Antimicrobial Resistance in The 21st Century (2018)
Antimicrobial Resistance in The 21st Century (2018)
Antimicrobial Resistance
in the 21st Century
Second Edition
Chapter 1
Introduction: Coordinated Global Action
Is Needed to Combat Antimicrobial
Resistance
I. W. Fong
Antimicrobial resistance is a global dilemma that threatens the health and safety of
populations in all countries of the world. Urgent actions are needed to be taken
before it reaches a critical stage, when large numbers of people in communities
cannot be treated for life-threatening infections due to lack of effective drugs.
Although the threat is most imminent from antibacterial resistance to commonly
used antibiotics for infections seen regularly in intensive care units and hospitals, it
is more prevalent and widespread and involves a wide spectrum of microbes. This
second edition of “Antimicrobial Resistance and Implications for the 21st Century”
provides not only updates and advances since the original edition but provides a
wider spectrum of topics on the issue. Although most chapters of this new edition
address issues of common bacterial resistance, others provide up-to-date reviews on
resistance trends with viruses, including human immunodeficiency virus [HIV] and
human herpes group of viruses.
To understand the evolution of microbial resistance, it is appropriate to review
historical aspects. Development of antimicrobial chemotherapy is usually attributed
to Paul Ehrlich [“father of chemotherapy”] based on his quest to find a cure for
parasitic infections, toward the latter part of the nineteenth century, with natural
dyes and heavy metals [mercury and arsenicals]. Penicillin was subsequently
discovered in 1928 and administered clinically in the 1940s; sulfonamides were
introduced clinically in 1937 [1]. Thus, the “antibiotic era” was under way by the
early 1940s. Penicillin and other antibiotics were initially derived from environmental
fungi and bacteria, often with improvements made by chemical synthesis. Hence,
the origin of antibiotics is through naturally derived substances produced to
antagonize or inhibit the growth of other microorganisms, probably due to an evo-
lutionary process that protects environmental niches of the producing organisms.
I. W. Fong (*)
Department of Medicine, University of Toronto, Toronto, ON, Canada
e-mail: [email protected]
Streptomycin and the precursor of cephalosporins were also obtained from soil
microbes in the 1940s. Penicillin was first used in 1941, and by 1944 penicillinase-
producing Staphylococcus aureus was described. Streptomycin was introduced
clinically in 1944 for treatment of tuberculosis, but resistance soon developed dur-
ing treatment [2]. By the mid-1950s, most of the major antibiotic families, including
aminoglycosides, chloramphenicol, tetracycline, and macrolides, had been devel-
oped [1]. Synthetic chemical agents with antibacterial activity were introduced in
the early 1950s with para-aminosalicylic acid and isoniazid as antituberculosis
agents; nitrofurantoin was found around the same time, followed by trimethoprim
in 1956. Nalidixic acid, discovered in the early 1960s, was the precursor of the fluo-
roquinolones, with norfloxacin and ciprofloxacin developed in the 1980s. Rifampin
was introduced for tuberculosis in 1968.
The spectacular success of antimicrobial therapy led to widespread use and
emergence of resistance. Pharmaceutical companies saw profit in new, more potent
derivatives, which gave rise to broad-spectrum antipseudomonal penicillins and
second-generation cephalosporins in the 1970s, with subsequent introduction in the
1980s of third-generation cephalosporins. Later, beta-lactamase inhibitors com-
bined with broad-spectrum penicillins and carbapenems, new glycopeptides, newer
macrolides, later-generation quinolones, linezolid [a new class of oxazolidinone],
and glycylcyclines [tigecycline] were introduced. Currently, there are more than
100 antimicrobial compounds available.
the 1950s, but toxicity kept them from widespread use for systemic infections. Now
they are gaining clinical use with infections caused by Pseudomonas aeruginosa,
Acinetobacter baumannii, and carbapenemase-producing Enterobacteriaceae that
are resistant to all other antibiotics [4]. Full vancomycin resistance among MRSA
strains is rare, probably because mutational changes cause impaired fitness for
MRSA as susceptibility decreases. As a result, vancomycin-intermediate resistance
[VISA] is slowly becoming a problem. Resistance to polymyxins appears to be
emerging readily with increased use.
Over the years, development of new antibiotics to counter resistant bacteria led
to appearance of novel resistant strains to these new drugs, which become wide-
spread with increasing use of the antibiotics. This pattern has been seen with every
new class of antibiotic developed over the years. There is strong correlation
between the frequency and quantity of antibiotics used in humans and animals and
the rate of development of antibacterial drug resistance. The logarithmic growth of
resistant bacteria since the 1970s is reflected by the number of β-lactamase
enzymes identified during the antibiotic era. Before 1970, there were only several
β-lactamase enzymes described, and now about 900 β-lactamase enzymes have
been identified [2].
Mobile genetic elements, extrachromosomal self-replicating structures [plas-
mids] and transposons, found in bacterial cells provide a resistance threat that
maybe unconquerable. Our understanding of horizontal transfer of bacterial
resistance was heralded by the discovery of antibiotic resistance plasmids that could
be disseminated by bacterial conjugation in the mid-1950s [2]. Since 1989 we have
gained much greater knowledge of the genetics of bacterial resistance following the
discovery of integrons. Integrons are versatile genetic elements, commonly found in
bacterial genomes, that allow efficient capture and expression of exogenous genes.
Plasmids and transposons are considered mobile integrons. Integrons play a major
role in the spread of antibiotic resistance, particularly in Gram-negative pathogens;
the majority of these pathogens carry integrons with resistant genes [5]. The process
of microbial resistance is complex, and besides plasmids and DNA mutations
[acquired or heritable and transferable], other mechanisms include biofilms, which
harbor hypermutator bacteria that select for resistance more frequently, and pheno-
typic tolerance, a situation in which bacteria are not killed by antimicrobials [6].
Antibiotic pressure predisposes to resistance and tolerance. Despite the call for
intensive research on mechanisms of microbial resistance and development of novel
compounds to counteract the spread in 2012 [6], no innovative agent is on the
horizon.
The origins, evolution, genetics, and biochemistry of antibiotic resistance have
been studied over the last 60 years. Figure 1.1 outlines the history of antibiotic dis-
covery and subsequent development of antibiotic resistance. The emergence of anti-
biotic resistance of pathogenic bacteria after antibiotic development and clinical
use, plus the absence of resistance in bacteria of the pre-antibiotic era [7], suggested
that resistance is a modern phenomenon. However, metagenomic analyses of
authenticated ancient DNA from 30,000-year-old Beringian permafrost sediments
identified genes encoding resistance to β-lactam, tetracycline, and glycopeptide
4 I. W. Fong
Fig. 1.1 Discovery of antibiotics and evolution of antimicrobial resistance. Abbreviations: MRSA
methicillin-resistant S. aureus, VRE vancomycin-resistant enterococci, ESBL extended
β-lactamase
antibiotics [8]. Thus, antibiotic resistance is a natural phenomenon that predates the
modern selective pressure associated with clinical and animal use. This explains the
rapid emergence of resistance to new antibiotics; resistance will continue to emerge
with drugs now in development. Hence, it is predictable that new antibiotics will
select for preexistent resistance determinants that have been present and circulating
in universal, environmental microbial pangenome for hundreds of years [8]. Thus,
the era of antibiotics has shifted naturally to the era of antibiotic resistance. Our
challenge is to minimize the problem.
It has been argued that the lack of access to life-saving antibiotics is as important an
issue as antibiotic resistance. It has been estimated that universal access to antibiotics
could prevent 445,000 deaths out of 590,000 deaths from pneumonia [75% reduc-
tion] in 101 countries [13]. However, increased use of pneumococcal and
Haemophilus influenza type B vaccines could prevent up to 11.4 million days on
antibiotics – a 47% reduction in 75 countries. Carriage of multiresistant bacteria is
not restricted to travelers or developing countries. In a study from Germany, of 4376
patients admitted to general wards, third-generation cephalosporin-resistant
Enterobacteriaceae were detected from rectal swabs in almost 10% of patients
immediately after admission [14]. Risk factors for presence of these multiresistant
bacteria included prior antimicrobial treatment, travel outside Europe, stay in a
long-term care facility, and use of proton pump inhibitors for gastroesophageal
reflux.
Many high-income countries, including the US and parts of Europe, have created
national plans as well as regulation to address antibiotic resistance issues. However,
the brunt of the problem will be borne by low-income and middle-income countries
that cannot afford the newer, expensive drugs.
Health care-associated infections are a major source of the problem, and inten-
sive care units are “generators” of resistant bacteria. The empiric institution of
broad-spectrum antimicrobials for infections in very ill patients is the force behind
this problem. There is a great need for inexpensive, rapid and reliable microbiologi-
cal/molecular diagnostic tests to alleviate some of the empiric overuse of antibiotics
in health-care settings. Conventional culture methods usually take >2 days for iden-
tification and susceptibility determination, but a rapid multiplex polymerase chain
reaction [rmPCR] can provide identification in 1.3 h [15]. Moreover, rapid point-of-
care tests are needed to distinguish viral and noninfectious inflammatory conditions
from bacterial infections. Antibiotic stewardship in hospitals in North America and
other countries reduces antibiotic use, improves patient outcome, decreases adverse
events such as superinfection with Clostridium difficile and antibiotic resistance [to
a modest degree], and is cost-effective [16]. Wider adoption of stringent steward-
ship programs is needed for all community hospitals globally, but it will be difficult
to implement in resource-poor countries.
Inappropriate antibiotic use is still widespread for acute upper respiratory tract
infections despite attempts to curb the abuse in outpatient primary-care practice by
education. Efforts had been made to improve prescription behavior and provide
guidelines for antibiotic use across the USA, but success has been limited. Overall
antibiotic use for acute respiratory infections has significantly declined in children
[17], but use in adults remains high, especially for broad-spectrum antibiotics and
macrolides [18], as confirmed by recent data from the Veterans Affairs health system
[19]. A review of outpatient antibiotic use in the USA reported that about 13% of all
visits [about 154 million per year] resulted in antibiotic prescriptions of which at
least 30% were considered unnecessary [20]. Thus, unnecessary use of antibiotic
1 Introduction: Coordinated Global Action Is Needed to Combat Antimicrobial… 7
for acute respiratory or minor infections remains high. One method worth exploring
is for public health officials and medical associations to send frequent e-mail
messages to primary-care physicians concerning the dangers of antibiotic
overprescribing [no proven benefit]. Another option is to provide financial incentives
for not prescribing antibiotics by medical insurance companies. Delayed prescribing
[delay between receiving the prescription and collecting the drugs] has shown some
success in reducing antibiotic use [21]. In Thailand, the Antibiotic Smart Use
program has shown that alternative treatment options were important in restricting
antibiotic outpatient use, such as oral rehydration and zinc for diarrheal diseases,
and herbal drugs packaged in antibiotic-like capsules for viral upper respiratory
infections [22].
What is being implemented to combat antimicrobial resistance? Several coun-
tries are now taking steps to improve antibiotic prescribing, and the World Health
Day in 2011 was dedicated to antimicrobial resistance. The Infectious Disease
Society of America in 2011 outlined a road map to counter antimicrobial resistance:
regular surveillance and data collection on resistance patterns and prevalence, uni-
versal antibiotic stewardship for hospitals, and provision of research and develop-
ment [R&D] incentives for drug companies to facilitate licensing of novel
antimicrobials [23]. But as pointed out, new antimicrobials will simply delay the
problem. More recently, the Presidential Advisory Council on Combating Antibiotic
Resistant Bacteria and Innovative Medicines Initiative suggested public-private
partnership to provide financial resources to assist R&D [24]. However, proposed
funding cuts by the Trump administration to the CDC's antimicrobial resistance
[AMR] fund by 14% and the NIAID by 23% threatens the progress in the fight
against antimicrobial resistance [Boucher et al. Proposed US funding cuts threaten
progress on antinicrobial resistance. Ann Int Med 2017; 167:738–9]
The WHO or United Nations could provide leadership to facilitate multinational
global collaboration in this effort. The WHO has just released priority pathogens
list for R&D of new antibiotics: priority 1 [critical] includes carbapenem-
resistant A. baumannii, P. aeruginosa, and multiresistant Enterobacteriaceae;
priority 2 [high] includes vancomycin-resistant Enterococcus faecium [VRE],
vancomycin-intermediate MRSA, clarithromycin-resistant Helicobacter pylori,
fluoroquinolone-resistant Campylobacter and Salmonella spp., and third-
generation-resistant Neisseria gonorrhoeae; and priority 3 [medium] includes
penicillin-non-susceptible Streptococcus pneumonia, ampicillin-resistant H.
influenzae, and fluoroquinolone-resistant Shigella spp. [25]. R&D for new drugs
for multiresistant tuberculosis and artemisinin-resistant malaria were previously
noted as high priority by the WHO. Development of new antibiotics to combat
resistant bacteria is a short-term solution to meet current needs. Resistance will
eventually develop to these agents as well. Innovative biological substances for
therapeutics where resistance is unlikely to develop are needed; research in this
area should be encouraged. This could include use of probiotics to counter and
prevent enteric colonization of resistant bacteria or bacteriophages to lyse
colonized resistant organisms such as MRSA and VRE.
8 I. W. Fong
On a global scale, there is much that can be done to reduce the risk of infection
and decrease the need for antibiotics, mainly in low-income countries. These
activities include wider and more universal use of vaccines, such as the pneumococcal
conjugate vaccine, H. influenzae type B vaccine, pertussis vaccine in pregnancy,
rotavirus vaccine, and measles vaccine, which could save lives and dramatically
reduce the use of pediatric antibiotics worldwide. Yearly universal influenza vacci-
nation of children and adults could also reduce the outpatient use of antibiotics for
respiratory infections in all countries. A major problem in developing and low-
income countries is poor sanitation and lack of clean water, which predisposes per-
sons to a variety of infectious diarrheas that leads to antibiotic overuse and increased
antimicrobial resistance. In general, better infection control practices in health-care
institutions could reduce the need for antibiotics and lead to reduced prevalence of
resistance. In the USA alone, it is predicted that within 5 years multiresistant bacte-
ria could cause 340,000 deaths per year, but immediate implementation of a national
intervention strategy involving all elements of the healthcare network [hospitals,
nursing homes, etc.] through infection control and universal antibiotic stewardship
could save 37,000 lives and avert 619,000 infections over the next 5 years [26].
Current estimates are that antibiotic-resistant bacteria cause 2 million illness and
23,000 deaths each year in the USA with a annual cost to the health care system of
over $20 billion [CDC. Antibiotic resistance threats in the United States, 2013.
www.cdc.gov/drugresistance/pdf/ar-trhreats-2013-508.pdf]
References
21. Little P, Moore M, Kelly J, et al. the PIPS Investigators. Delayed antibiotic prescribing strate-
gies for respiratory tract infections in primary care: pragmatic, factorial, randomized con-
trolled trial. BMJ. 2014;348:g1606.
22. Dar OA, Hasa R, Schlundt J, et al. Antimicrobials: access and sustainable effectiveness.
Exploring the evidence base for national and regional policy interventions to combat resis-
tance. Lancet. 2016;387:285–95.
23. Spellberg B, Blaser M, Guidos RJ, et al. Infectious Disease Society of America [IDSA].
Combating antimicrobial resistance: policy recommendations to save lives. Clin Infect Dis.
2011;52(Suppl 5):S397-428.
24. Presidential Advisory Council on Combating Antibiotic-Resistant Bacteria. 2016. Accessed at
www.hhs.gov/sites/default/fioles/paccarb-final-report-03312016.pdf
25. WHO. Global priority list of antibiotic-resistant bacteria to guide research, discovery, and
development of new antibiotics. Accessed at http://www.int/medicines/publications/global-
priority-list-antibiotic-resistant-bacteria/en/. On 31 Mar 2017.
26. Slayton RB, Toth D, Lee BY, et al. Estimated effects of a coordinated approach for action to
reduce antibiotic-resistant infections in health care facilities—United States. MMWR Morb
Mortal Wkly Rep. 2015;64:826.
27. Cheng AC, Turnbridge J, Collignon P, Looke D, Barton M, Gottlieb T. Control of fluoroquino-
lone resistance through successful regulation, Australia. Emerg Infect Dis. 2012;18:1453–60.
28. Zhu YG, Zhao Y, Li B, et al. Continental-scale pollution of estuaries with antibiotic resistant
genes. Nat Microbiol. 2017;2:16270.
Part I
Examples of Resistance
Chapter 2
Antimicrobial Resistance Among
Streptococcus pneumoniae
2.1 Introduction
Antibiotic resistance is a direct result of antibiotic consumption [1, 2]. In the United
States, it is estimated that antibiotic resistance is responsible for more than 2 million
infections and 23,000 deaths each year, with a direct cost of $20 billion and addi-
tional productivity losses of $35 billion [3, 4]. Data from Europe showed that
approximately 25,000 deaths are attributable to antibiotic-resistant infections, with
a related cost of $1.5 billion annually [5]. The use of antibiotics in primary care is
high; the most frequent indications for their use are respiratory tract infections [6].
Streptococcus pneumoniae (pneumococcus) is the leading cause of community-
acquired pneumonia and is considered to be a major cause of death of children
under 5 years old worldwide. In a recent report on global antibiotic resistance, pub-
lished by the World Health Organization (WHO) in 2014, pneumococcus was con-
sidered to be one of the nine bacteria of international concern [7]. Other infections
caused by pneumococcus include bacteremia, otitis media, and meningitis. In bacte-
rial meningitis, pneumococcus is associated with mortality rates ranging from 16%
to 37%. About 30–50% of adult survivors present permanent residual symptoms [8,
9]. The study by Van Boeckel et al. [10], regarding global antibiotic consumption
from 2000 to 2010, reported that it grew by more than 30%, from approximately 50
Erythromycin 1953
1943-45 1960 1962 1968 1980 1990 1996 1998 2000 2005 2010
Year of reported resistance
Levofloxacin resistant
Penicillin resistant S. pneumoniae
S. pneumoniae
Cephalosprin resistant
S. pneumoniae
MDR
Erythromycin resistant S. pneumoniae
S. pneumoniae
• Infections other than meningitis: susceptible < 2 μg/ml, intermediate < 4 μg/ml,
and resistant ≥ 8 μg/ml
• Meningitis: susceptible ≤ 0.06 μg/ml, intermediate ≥ 0.12 μg/ml, and resis-
tant ≥ 2 μg/ml
The breakpoints for penicillin susceptibility are based on three criteria: micro-
biological data, pharmacokinetic/pharmacodynamics of β-lactam antibiotics, and
clinical outcome of pneumococcal infections. In a patient treated with a dose of
intravenous penicillin, the levels achieved in the lung will be 100 times greater than
those reached in the brain. Thus, use of low concentrations of β-lactam for pneumo-
coccal infections, such as otitis media or meningitis, could lead to treatment failure.
In contrast, with pulmonary infections the levels of β-lactam reached are generally
sufficient to clear infection. Therefore, treating the same pathogen will require dif-
ferent doses of a given β-lactam depending on the site of infection. Likewise, we
must considerer pneumococcal resistance in different sites of infection differently,
and breakpoints for resistance will be different.
Macrolides inhibit bacterial protein synthesis by binding to the 23S rRNA compo-
nent of the 50S ribosomal subunit in bacteria. There are two main mechanisms of
macrolide resistance in pneumococcus. One involves target-site (ribosomal) altera-
tion by an enzyme that methylates 23S rRNA, an enzyme that is encoded by the
ermB (erythromycin-resistance methylase) gene. The resistance phenotype is called
MLSB (macrolide, lincosamide, streptogramin B type) and is responsible for a high
level of macrolide resistance. In a low proportion of cases, ermB gene variation that
modifies the binding site for macrolides and lincosamides confers complete cross-
resistance to clindamycin [22].
The second mechanism of resistance involves active efflux pumps encoded by
the mefE or mefA (macrolide efflux) genes. These mutations result in low-level
resistance to macrolides but not to the other two agents. The mefA gene is predomi-
nant in Europe, whereas mefB gene predominates in North America.
The relative frequency of the two macrolide resistance mechanisms varies by
geographic region [23–36] (Table 2.2): in European countries, approximately 90%
of the isolates of pneumococcus presented the MLSB phenotype, which is associ-
ated with high levels of macrolide resistance, whereas in North America between
50% and 65% of the resistant pneumococcus isolates contained efflux mutations
that were associated with lower levels of macrolide resistance [12]. In Asian coun-
tries, strains that showed both mechanism of resistance are a major concern, with
between 12% and 40% of the resistant isolates displaying both mechanisms [31,
37–40]. In South American countries, isolates reporting both mechanisms vary
between 4% and 20% [34, 41]. Worldwide resistance to macrolides in pneumococ-
18 C. Cillóniz et al.
cus has increased recently and is associated with the extensive global use of macro-
lides, principally for community-acquired respiratory tract infections.
The use of a β-lactam antibiotic in the previous 3–6 months is the main risk factor
associated with penicillin-resistant pneumococcal infection [12, 21, 58–60]. A
study by Ruhe et al. [61] regarding the duration of previous antibiotic treatment and
its association with penicillin-resistant bacteremic infection revealed that the risk
depends on the class of prior antibiotic exposure and the duration of therapy. The
study analyzed 303 patients with pneumococcal bacteremia. In 98 (32%) cases of
bacteremia caused by penicillin-non-susceptible S. pneumoniae, statistical analysis
showed that the use of β-lactams, sulfonamides, and macrolides within the last
1–6 months before presentation was associated with penicillin-non-susceptible S.
pneumoniae bacteremia (p < 0.05). In a second study with the same bacteremic
population, Ruhe et al. [62] identified 33 (11%) cases of bacteremia caused by high-
level resistant S. pneumoniae. In these cases, three risk factors for high-level
penicillin-resistant pneumococcal infection were identified: β-lactam antibiotic use
in the previous 6 months, previous residence in a risk area (defined as stays in day-
care facilities, prisons, homeless shelters, nursing homes, or other long-term care
facilities), and respiratory tract infection in the previous year.
Age extremes (<5 years or > 65 years) are a recognized risk factor for penicillin-
resistant pneumococcal infections [12, 17, 63]. As pointed out above, nasopharyn-
geal carriage of pneumococcus in healthy children ranges from 20% to 50%, and in
the healthy adult population, nasopharyngeal carriage rates range from 5% to 30%
[64, 65]. Consequently, it is not difficult to understand why several studies have
shown that day-care centers are a risk factor for colonization and infection of chil-
dren due to penicillin-resistant pneumococcus [66–68]. Similarly, institutionalized
adults, especially those older than 65 years of age, have increased risk for penicillin-
resistant pneumococcal infections [69]. Moreover, the presence of specific comor-
bidities, such as human immunodeficiency virus (HIV) and chronic pulmonary
disease, especially chronic obstructive pulmonary disease (COPD), is a recognized
risk factor for penicillin-resistant pneumococcal infection [58].
Several studies have addressed the association between antibiotic consumption
and resistance selection. A study by van Eldere et al. [70], concerning the impact of
antibiotic usage in ambulatory patients in Belgium, involved 14,448 Streptococcus
pneumoniae isolates collected between 1994 and 2004. This work showed a modest
relationship between consumption and resistance; additional factors were high pop-
ulation density and proximity to high-resistance regions, particularly for the devel-
opment of multiple resistances in pneumococcus. In this Belgian population, the
highest levels of resistance were to erythromycin, followed by resistance to tetracy-
cline and penicillin; the highest prevalence of co-resistance to two antibiotics was
for erythromycin-tetracycline.
In 2001 the prevalence of non-susceptibility to erythromycin in the Belgium
study peaked at 36.7% and stayed mostly stable until 2004. Prevalence of non-
susceptibility to tetracycline reached its highest level (31.7%) in 2000; penicillin
non-susceptibility hit 17.7% in 2000 and declined to 11.6% in 2004. The prevalence
22 C. Cillóniz et al.
Recent therapy by macrolides is the main risk factor for macrolide-resistant nasal
colonization and pneumococcal infection [1, 12, 73, 74]. The study by Dias et al.
[75], which evaluated the role of antimicrobial and vaccine use in the trends of
resistance to penicillin and erythromycin in Portugal from 1994 to 2004, found
that the use of macrolides was the main factor associated with an increase of
2 Antimicrobial Resistance Among Streptococcus pneumoniae 23
penicillin and erythromycin non-susceptible isolates among adults (p < 0.01) and
erythromycin non-susceptible isolates among children (p = 0.006). The study also
suggested that the heptavalent vaccine is failing to reduce antimicrobial resis-
tance, possibly due to the increased consumption of azithromycin (p = 0.04).
Other works showed that there is an increased risk of macrolide-resistant infec-
tion in cases related to certain pneumococcus serotypes, in particular 6A, 6B,
11A, 14, 23F, and 19F [76, 77].
Other important risk factors are age below 5 years [78–81], attendance in a day-
care center [82–84], middle ear infection [85–87], and nosocomial acquisition [26].
As with β-lactams, there is strong evidence correlating the prevalence of macrolide
resistance of pneumococcus and overall macrolide consumption within specific
geographic areas [70, 71, 88].
Previous exposure to fluoroquinolones is considered the main risk factor for fluo-
roquinolone resistance [89–92]. Other risk factors, reported worldwide, are
COPD, nosocomial acquisition, and residence in a nursing home [43, 93, 94]. A
retrospective review of cases of invasive pneumococcal infections in adults in
Spain reported that residence in public shelters (OR 26.13, p = 0.002), previous
hospitalization (OR 61.77, p < 0.001), human immunodeficiency virus (HIV)
infection (OR 28.14, p = 0.009), and heavy smoking (OR 14.41, p = 0.016) are
risk factors associated with acquiring an infection by levofloxacin-resistant pneu-
mococci [95–97].
Fig. 2.2 How antibiotic resistance arises and spreads in bacterial population. (Figure reproduced
by the permission of the author: Laura Piddock and Victoria Wells. Longevity Bulletin:
Antimicrobial Resistance, Chapter 3: How antimicrobial resistance emerges. Issue 8, May 2016)
2 Antimicrobial Resistance Among Streptococcus pneumoniae 25
Two types of pneumococcal vaccines are currently available: the polyvalent pneu-
mococcal polysaccharide vaccine (PPV) and the pneumococcal conjugate vaccine
(PCV).
The PPV23 vaccine includes 23 purified capsular polysaccharide antigens of
Streptococcus pneumoniae (1, 2, 3, 4, 5, 6B, 7F, 8, 9N, 9V, 10A, 11A, 12F, 14, 15B,
17F, 18C, 19A, 19F, 20, 22F, 23F, and 33F); it was licensed in the United States in
1983. PPV23 induces antibodies primarily through a T-cell-independent immune
response that enhances phagocytosis, thereby killing the bacterium [121]. The
immune system of young children does not produce an adequate response to the
polysaccharide capsule; consequently, the vaccine is not used in this age group.
The pneumococcal conjugate vaccine 7-valent (PCV7), which included seven
pneumococcal sertotypes (4, 6B, 9V, 14, 18C, 19F, and 23F), was introduced in the
United States in 2000. It is recommended for infants and young children. This vac-
cine is highly effective in preventing invasive disease, with percentages of efficacy
of about 90%. The routine use of PCV7 has resulted not only in a tremendous reduc-
tion in invasive pneumococcal infections in children but also decreased rates of
pneumococcal disease in adults.
Also, after 4 years of the introduction of PCV7 in the United States, the inci-
dence of invasive pneumococcal disease caused by penicillin-non-susceptible S.
pneumoniae and multidrug-resistant S. pneumoniae decreased. In 1999, the rate of
invasive disease caused by penicillin-non-susceptible strains was 6.3 cases per
100,000 – it decreased to 2.7 cases per 100,000 in 2004. Similarly, in 1999 the rate
of cases caused by strains not susceptible to multiple antibiotics was 4.1 cases per
100,000 and decreased to 1.7 cases per 100,000 in 2004 [122]. The study by Whitney
et al. [123] demonstrated that the PCV7 vaccine prevents invasive disease in both
healthy and chronically ill children. Despite the success of PCV7, studies have
noted an increase in the incidence of invasive pneumococcal disease (IPD) caused
by non-vaccine serotypes, such as 1, 3, 5, 6A, 6C, 7F, 12F, 19A, and 22F [124].
These serotypes are related to penicillin-non-susceptible clones. The emergence of
serotype 19A, which correlates with high-level penicillin and multidrug resistance,
is a main concern globally [125]. This serotype presents a dual macrolide-resistance
phenotype (erm B and mefA).
A new 13-valent pneumococcal polysaccharide-protein conjugate vaccine
(PCV13) was approved by the Food and Drug Administration in February 2010 for
the prevention of IPD in infants and young children. PCV13 contains capsular poly-
saccharides from serotypes 1, 3, 4,6A, 7F, 9 V, 14, 18C, 19A, 19F, and 23F. In
March 2010, the Advisory Committee for Immunization Practices (ACIP) recom-
mended that PCV13 replace PCV7 for the vaccination of children. New studies
show a similar reduction in IPD following the introduction of the PCV13 vaccine,
as seen previously with the PCV7 vaccine. The study by Moore et al. [126] analyzed
IPD cases (33,688 cases, of which 89% contained serotyping results) during July
2004–June 2013 and classified as being caused by the PCV13 serotypes against
28 C. Cillóniz et al.
which PCV7 has no effect (PCV13/nonPCV7). The work found a reduction in IPD
in adults associated with PCV13 introduction in children. In all adult age groups,
PCV13/nonPCV7-type IPD (especially serotypes 19A and 7F) declined by 58–72%,
which was comparable to that observed early after PCV7 introduction. The PCV13
led to overall reductions of IPD of 12–32% [126]. However, the phenomenon of
serotype replacement, which is thought to be caused by non-vaccine serotypes
(NVT) that occupy nasopharyngeal natural niches vacated after pneumococcal vac-
cination, is again observed with pneumococcal serotypes 11A, 15A, 23B, and 35B,
the most frequent serotypes. Serotypes 15A and 23B show a high proportion of
penicillin non-susceptibility [127].
Treatment failure has been observed with patients treated with fluoroquinolones
who had infections caused by fluoroquinolone-resistant strains [43, 132]. However,
the global rates of fluoroquinolone resistance remain low [32, 99, 114, 118, 133],
making correlation between resistance and outcome statistically marginal. In a 2013
study, Kang et al. [134] evaluated the impact of levofloxacin resistance on 136 adult
patients with invasive pneumococcal disease (IPD). In this work, pneumonia was
the most frequent disease (68%), followed by primary bacteremia (11%) and men-
ingitis (11%). The rate of levofloxacin resistance in invasive pneumococcal isolates
30 C. Cillóniz et al.
was 3.7% (5/136) of the isolates. The overall 30-day mortality rate was 26.5%
(36/136). In univariate analysis, the factors associated with 30-day mortality in
patients with IPD were corticosteroid use, presentation with septic shock, and
development of acute respiratory distress syndrome (ARDS). The authors found an
association between levofloxacin resistance and increased mortality, although sta-
tistical significance was not reached (p = 0.083). However, multivariate analysis
revealed that presentation with septic shock, corticosteroid use, development of
ARDS, and levofloxacin resistance were independent factors associated with 30-day
mortality.
Several worldwide reports about antimicrobial resistance in pneumococcus
noted that in countries where the rates of β-lactam resistance and macrolide resis-
tance are high, the prevalence of fluoroquinolone resistance is also high [70, 114].
It may be that in those situations the consumption of fluoroquinolones is also high.
mutation that confers high resistance to macrolides, whereas in the United States,
the dominant mechanism of resistance is active efflux, which confers low levels of
resistance to macrolides. These data suggest the importance of clinical studies in
different geographical areas before recommending particular antibiotics. A com-
pletely different question is how to slow transmission among young children and
elderly persons in long-term care facilities. Solutions may involve reducing antimi-
crobial consumption, the main driver of newly acquired resistance.
Continued surveillance to quantify pneumococcal resistance is also needed to
detect the emergence of new strains exhibiting high-level resistance to penicillin.
Moreover, we need to better understand the clinical relevance and impact of antibi-
otic resistance on pneumococcal infections, since there is not always a clear rela-
tionship between resistance and treatment failure.
Major Points
• Streptococcus pneumoniae remains an important pathogen worldwide.
Pneumococcal infections are related to high rates of morbidity and mortality
especially in young children, older adults, and immunocompromised persons.
• Worldwide pneumococcal infections remain a big challenge for physicians
because of its resistance to penicillin and increasing resistance to macrolides.
• Efforts to reduce antibiotic consumption should be encouraged by educational
programs and guidelines for healthcare professionals.
• The best way to prevent pneumococcal infection is by the implementation of
conjugate pneumococcal vaccinations.
• It is important to monitor the evolution of pneumococcal disease, focusing on
serotype replacement.
• Studies focusing on the development of new vaccine designs should be addressed
in order to avoid serotype replacement.
References
27. Wierzbowski AK, Nichol K, Laing N, et al. Macrolide resistance mechanisms among
Streptococcus pneumoniae isolated over 6 years of Canadian Respiratory Organism
Susceptibility Study (CROSS) (1998 2004). J Antimicrob Chemother. 2007;60(4):733–40.
28. Hawkins PA, Chochua S, Jackson D, Beall B, McGee L. Mobile elements and chromosomal
changes associated with MLS resistance phenotypes of invasive pneumococci recovered in
the United States. Microb Drug Resist. 2015;21(2):121–9.
29. El Ashkar S, Osman M, Rafei R, et al. Molecular detection of genes responsible for macro-
lide resistance among Streptococcus pneumoniae isolated in North Lebanon. J Infect Public
Health. 2017;10(6):745–8.
30. Kawaguchiya M, Urushibara N, Aung MS, et al. Emerging non-PCV13 serotypes of nonin-
vasive Streptococcus pneumoniae with macrolide resistance genes in northern Japan. New
Microbes New Infect. 2016;9:66–72.
31. Azadegan A, Ahmadi A, Lari AR, Talebi M. Detection of the efflux-mediated erythromycin
resistance transposon in Streptococcus pneumoniae. Ann Lab Med. 2015;35(1):57–61.
32. Kim SH, Song JH, Chung DR, et al. Changing trends in antimicrobial resistance and sero-
types of Streptococcus pneumoniae isolates in Asian countries: an Asian Network for
Surveillance of Resistant Pathogens (ANSORP) study. Antimicrob Agents Chemother.
2012;56(3):1418–26.
33. Ramos V, Duarte C, Diaz A, Moreno J. Mobile genetic elements associated with
erythromycin- resistant isolates of Streptococcus pneumoniae in Colombia. Biomedica.
2014;34(Suppl 1):209–16.
34. Reijtman V, Gagetti P, Faccone D, et al. Macrolide resistance in Streptococcus pneumoniae
isolated from Argentinian pediatric patients suffering from acute otitis media. Rev Argent
Microbiol. 2013;45(4):262–6.
35. Diawara I, Zerouali K, Katfy K, et al. Phenotypic and genotypic characterization of
Streptococcus pneumoniae resistant to macrolide in Casablanca, Morocco. Infect Genet Evol.
2016;40:200–4.
36. Rachdi M, Boutiba-Ben Boubaker I, Moalla S, et al. Phenotypic and genotypic character-
ization of macrolide resistant Streptococcus pneumoniae in Tunisia. Pathol Biol (Paris).
2008;56(3):125–9.
37. Reinert RR, Filimonova OY, Al-Lahham A, et al. Mechanisms of macrolide resistance
among Streptococcus pneumoniae isolates from Russia. Antimicrob Agents Chemother.
2008;52(6):2260–2.
38. Ma X, Yao KH, Xie GL, et al. Characterization of erythromycin-resistant Streptococcus
pneumoniae isolates causing invasive diseases in Chinese children. Chin Med J (Engl).
2013;126(8):1522–7.
39. Zhou L, Yu SJ, Gao W, Yao KH, Shen AD, Yang YH. Serotype distribution and antibiotic
resistance of 140 pneumococcal isolates from pediatric patients with upper respiratory infec-
tions in Beijing, 2010. Vaccine. 2011;29(44):7704–10.
40. Geng Q, Zhang T, Ding Y, et al. Molecular characterization and antimicrobial susceptibility
of Streptococcus pneumoniae isolated from children hospitalized with respiratory infections
in Suzhou, China. PLoS One. 2014;9(4):e93752.
41. Caierao J, Hawkins P, Sant’anna FH, et al. Serotypes and genotypes of invasive Streptococcus
pneumoniae before and after PCV10 implementation in southern Brazil. PLoS One.
2014;9(10):e111129.
42. Fuller JD, McGeer A, Low DE. Drug-resistant pneumococcal pneumonia: clinical relevance
and approach to management. Eur J Clin Microbiol Infect Dis. 2005;24(12):780–8.
43. Fuller JD, Low DE. A review of Streptococcus pneumoniae infection treatment failures asso-
ciated with fluoroquinolone resistance. Clin Infect Dis. 2005;41(1):118–21.
44. Urban C, Rahman N, Zhao X, et al. Fluoroquinolone-resistant Streptococcus pneumoniae
associated with levofloxacin therapy. J Infect Dis. 2001;184(6):794–8.
45. Woodhead M, Blasi F, Ewig S, et al. Guidelines for the management of adult lower respira-
tory tract infections--full version. Clin Microbiol Infect. 2011;17(Suppl 6):E1–59.
34 C. Cillóniz et al.
46. Mandell LA, Wunderink RG, Anzueto A, et al. Infectious Diseases Society of America/
American Thoracic Society Consensus guidelines on the management of community-
acquired pneumonia in adults. Clin Infect Dis. 2007;44(Suppl 2):S27–72.
47. Donhofer A, Franckenberg S, Wickles S, Berninghausen O, Beckmann R, Wilson
DN. Structural basis for TetM-mediated tetracycline resistance. Proc Natl Acad Sci U S A.
2012;109(42):16900–5.
48. Moscoso M, Domenech M, Garcia E. Vancomycin tolerance in Gram-positive cocci. Environ
Microbiol Rep. 2011;3(6):640–50.
49. Van BF, Reinert RR, Appelbaum PC, Tulkens PM, Peetermans WE. Multidrug-resistant
Streptococcus pneumoniae infections: current and future therapeutic options. Drugs.
2007;67(16):2355–82.
50. Sheppard C, Fry NK, Mushtaq S, et al. Rise of multidrug-resistant non-vaccine serotype
15A Streptococcus pneumoniae in the United Kingdom, 2001 to 2014. Euro Surveill.
2016;21(50):3042.
51. Richter SS, Diekema DJ, Heilmann KP, Dohrn CL, Riahi F, Doern GV. Changes in pneumo-
coccal serotypes and antimicrobial resistance after introduction of the 13-valent conjugate
vaccine in the United States. Antimicrob Agents Chemother. 2014;58(11):6484–9.
52. Golden AR, Rosenthal M, Fultz B, et al. Characterization of MDR and XDR Streptococcus
pneumoniae in Canada, 2007–13. J Antimicrob Chemother. 2015;70(8):2199–202.
53. Golden AR, Adam HJ, Zhanel GG. Invasive Streptococcus pneumoniae in Canada, 2011–
2014: Characterization of new candidate 15-valent pneumococcal conjugate vaccine sero-
types 22F and 33F. Vaccine. 2016;34(23):2527–30.
54. Olarte L, Kaplan SL, Barson WJ, et al. Emergence of Multidrug-resistant Pneumococcal
Serotype 35B among U.S. Children. J Clin Microbiol. 2017;55(3):724–34.
55. Nakano S, Fujisawa T, Ito Y, et al. Serotypes, antimicrobial susceptibility, and molecular
epidemiology of invasive and non-invasive Streptococcus pneumoniae isolates in paediatric
patients after the introduction of 13-valent conjugate vaccine in a nationwide surveillance
study conducted in Japan in 2012–2014. Vaccine. 2016;34(1):67–76.
56. Elshafie S, Taj-Aldeen SJ. Emerging resistant serotypes of invasive Streptococcus pneu-
moniae. Infect Drug Resist. 2016;9:153–60.
57. Bahy RH, Hamouda HM, Shahat AS, Yassin AS, Amin MA. Emergence of neoteric sero-
types among multidrug resistant strains of Streptococcus pneumoniae prevalent in Egypt.
Jundishapur J Microbiol. 2016;9(4):e30708.
58. Aspa J, Rajas O, Rodríguez de Castro F, et al. Drug-resistant pneumococcal pneumonia:
clinical relevance and related factors. Clin Infect Dis. 2004;38(6):787–98.
59. Pallares R, Gudiol F, Linares J, et al. Risk factors and response to antibiotic therapy in adults
with bacteremic pneumonia caused by penicillin-resistant pneumococci. N Engl J Med.
1987;317:18–22.
60. Tsai HY, Lauderdale TL, Wang JT, et al. Updated antibiotic resistance and clinical spectrum
of infections caused by Streptococcus pneumoniae in Taiwan: emphasis on risk factors for
penicillin nonsusceptibilities. J Microbiol Immunol Infect. 2013;46(5):345–51.
61. Ruhe JJ, Hasbun R. Streptococcus pneumoniae bacteremia: duration of previous antibiotic
use and association with penicillin resistance. Clin Infect Dis. 2003;36(9):1132–8.
62. Ruhe JJ, Myers L, Mushatt D, Hasbun R. High-level penicillin-nonsusceptible Streptococcus
pneumoniae bacteremia: identification of a low-risk subgroup. Clin Infect Dis.
2004;38(4):508–14.
63. Mollendorf CV, Cohen C, De Gouveia L, et al. Factors associated with ceftriaxone non-
susceptibility of Streptococcus pneumoniae: analysis of South African national surveillance
data, 2003 to 2010. Antimicrob Agents Chemother. 2014;58(6):3293–305.
64. Bogaert D, de Groot R, Hermans PW. Streptococcus pneumoniae colonisation: the key to
pneumococcal disease. Lancet Infect Dis. 2004;4(3):144–54.
65. Gritzfeld JF, Wright AD, Collins AM, et al. Experimental human pneumococcal carriage.
J Vis Exp. 2013;(72):50115.
2 Antimicrobial Resistance Among Streptococcus pneumoniae 35
66. Vasoo S, Singh K, Hsu LY, et al. Increasing antibiotic resistance in Streptococcus pneumoniae
colonizing children attending day-care centres in Singapore. Respirology. 2011;16(8):1241–8.
67. Jain A, Kumar P, Awasthi S. High nasopharyngeal carriage of drug resistant Streptococcus
pneumoniae and Haemophilus influenzae in North Indian schoolchildren. Trop Med Int Heal.
2015;10:234–9.
68. Dunais B, Bruno-Bazureault P, Carsenti-Dellamonica H, Touboul P, Pradier C. A decade-
long surveillance of nasopharyngeal colonisation with Streptococcus pneumoniae among
children attending day-care centres in south-eastern France: 1999–2008. Eur J Clin Microbiol
Infect Dis. 2011;30(7):837–43.
69. Nuorti JP, Butler JC, Crutcher JM, et al. An outbreak of multidrug-resistant pneumococ-
cal pneumonia and bacteremia among unvaccinated nursing home residents. N Engl J Med.
1998;338:1861–8.
70. Van Eldere J, Mera RM, Miller LA, Poupard JA, Amrine-Madsen H. Risk factors for devel-
opment of multiple-class resistance to Streptococcus pneumoniae Strains in Belgium over
a 10-year period: antimicrobial consumption, population density, and geographic location.
Antimicrob Agents Chemother. 2007;51(10):3491–7.
71. van de Sande-Bruinsma N, Grundmann H, Verloo D, et al. Antimicrobial drug use and resis-
tance in Europe. Emerg Infect Dis. 2008;14(11):1722–30.
72. Farshad AA, Enferadi M, Bakand S, Jamshidi OR, Mirkazemi R. Penicillin dust exposure
and penicillin resistance among pharmaceutical workers in Tehran, Iran. Int J Occup Environ
Health. 2016:22(3): 218–23.
73. Hare KM, Leach AJ, Morris PS, et al. Impact of recent antibiotics on nasopharyngeal car-
riage and lower airway infection in Indigenous Australian children with non-cystic fibrosis
bronchiectasis. Int J Antimicrob Agents. 2012;40(4):365–9.
74. Metlay JP, Fishman NO, Joffe MM, Kallan MJ, Chittams JL, Edelstein PH. Macrolide
resistance in adults with bacteremic pneumococcal pneumonia. Emerg Infect Dis.
2006;12(8):1223–30.
75. Dias R, Canica M. Trends in resistance to penicillin and erythromycin of invasive pneumo-
cocci in Portugal. Epidemiol Infect. 2008;136(7):928–39.
76. Baek JY, Kim SH, Kang CI, et al. Prevalence of antimicrobial resistant Streptococcus pneu-
moniae serotype 11A isolates in Korea, during 2004–2013, due to the increase of multidrug-
resistant clone, CC166. Infect Genet Evol. 2016;38:122–5.
77. Zhou L, Ma X, Gao W, et al. Molecular characteristics of erythromycin-resistant
Streptococcus pneumoniae from pediatric patients younger than five years in Beijing, 2010.
BMC Microbiol. 2012;12:228.
78. Pan F, Han L, Huang W, et al. Serotype distribution, antimicrobial susceptibility, and molecu-
lar epidemiology of Streptococcus pneumoniae Isolated from children in Shanghai, China.
PLoS One. 2015;10(11):e0142892.
79. Hadjipanayis A, Efstathiou E, Alexandrou M, et al. Nasopharyngeal Pneumococcal Carriage
among Healthy Children in Cyprus Post Widespread Simultaneous Implementation of
PCV10 and PCV13 Vaccines. PLoS One. 2016;11(10):e0163269.
80. Vanderkooi OG, McConnell A, Church DL, Kellner JD. Antimicrobial susceptibility of inva-
sive and lower respiratory tract isolates of Streptococcus pneumoniae, 1998 to 2007. Can
J Infect Dis Med Microbiol. 2009;20(4):e139–44.
81. Matsumoto A, Hosoya M, Kawasaki Y, Katayose M, Kato K, Suzuki H. The emergence of
drug-resistant Streptococcus pneumoniae and host risk factors for carriage of drug-resistant
genes in northeastern Japan. Jpn J Infect Dis. 2007;60(1):10–3.
82. Grivea IN, Priftis KN, Giotas A, et al. Dynamics of pneumococcal carriage among day-care
center attendees during the transition from the 7-valent to the higher-valent pneumococcal
conjugate vaccines in Greece. Vaccine. 2014;32(48):6513–20.
83. Zuccotti G, Mameli C, Daprai L, et al. Serotype distribution and antimicrobial susceptibili-
ties of nasopharyngeal isolates of Streptococcus pneumoniae from healthy children in the
13-valent pneumococcal conjugate vaccine era. Vaccine. 2014;32(5):527–34.
36 C. Cillóniz et al.
84. Korona-Glowniak I, Siwiec R, Malm A. Resistance determinants and their association with
different transposons in the antibiotic-resistant Streptococcus pneumoniae. Biomed Res Int.
2015;2015:836496.
85. Ding Y, Geng Q, Tao Y, et al. Etiology and epidemiology of children with acute otitis media
and spontaneous otorrhea in Suzhou, China. Pediatr Infect Dis J. 2015;34(5):e102–6.
86. Abdelnour A, Arguedas A, Dagan R, et al. Etiology and antimicrobial susceptibility of
middle ear fluid pathogens in Costa Rican children with otitis media before and after the
introduction of the 7-valent pneumococcal conjugate vaccine in the National Immunization
Program: acute otitis media microbiology in Costa Rican children. Medicine (Baltimore).
2015;94(2):e320.
87. Pumarola F, Mares J, Losada I, et al. Microbiology of bacteria causing recurrent acute oti-
tis media (AOM) and AOM treatment failure in young children in Spain: shifting patho-
gens in the post-pneumococcal conjugate vaccination era. Int J Pediatr Otorhinolaryngol.
2013;77(8):1231–6.
88. Arason VA, Sigurdsson JA, Erlendsdottir H, Gudmundsson S, Kristinsson KG. The role of
antimicrobial use in the epidemiology of resistant pneumococci: a 10-year follow up. Microb
Drug Resist. 2006;12(3):169–76.
89. Ho PL, Tse WS, Tsang KW, et al. Risk factors for acquisition of levofloxacin-resistant
Streptococcus pneumoniae: a case-control study. Clin Infect Dis. 2001;32(5):701–7.
90. Kang CI, Song JH, Kim SH, et al. Risk factors for levofloxacin-nonsusceptible Streptococcus
pneumoniae in community-acquired pneumococcal pneumonia: a nested case-control study.
Eur J Clin Microbiol Infect Dis. 2014;33(1):55–9.
91. Kim ES, Hooper DC. Clinical importance and epidemiology of quinolone resistance. Infect
Chemother. 2014;46(4):226–38.
92. Kuster SP, Rudnick W, Shigayeva A, et al. Previous antibiotic exposure and antimicrobial
resistance in invasive pneumococcal disease: results from prospective surveillance. Clin
Infect Dis. 2014;59(7):944–52.
93. Wolter N, Du Plessis M, Von Gottberg A, De Gouveia L, Klugman KP. Molecular charac-
terization of emerging non-levofloxacin-susceptible pneumococci isolated from children in
South Africa. J Clin Microbiol. 2009;47(5):1319–24.
94. Ben-David D, Schwaber MJ, Adler A, et al. Persistence and complex evolution
of fluoroquinolone- resistant Streptococcus pneumoniae clone. Emerg Infect Dis.
2014;20(5):799–805.
95. Isea-Pena MC, Sanz-Moreno JC, Esteban J, Fernandez-Roblas R, Fernandez-Guerrero
ML. Risk factors and clinical significance of invasive infections caused by levofloxacin-
resistant Streptococcus pneumoniae. Infection. 2013;41(5):935–9.
96. Kupronis BA, Richards CL, Whitney CG. Invasive pneumococcal disease in older adults resid-
ing in long-term care facilities and in the community. J Am Geriatr Soc. 2003;51(11):1520–5.
97. Vanderkooi OG, Low DE, Green K, Powis JE, McGeer A. Predicting antimicrobial resistance
in invasive pneumococcal infections. Clin Infect Dis. 2005;40(9):1288–97.
98. Thabit AK, Crandon JL, Nicolau DP. Antimicrobial resistance: impact on clinical and
economic outcomes and the need for new antimicrobials. Expert Opin Pharmacother.
2015;16(2):159–77.
99. Song JH, Dagan R, Klugman KP, Fritzell B. The relationship between pneumococcal sero-
types and antibiotic resistance. Vaccine. 2012;30(17):2728–37.
100. Thummeepak R, Leerach N, Kunthalert D, Tangchaisuriya U, Thanwisai A, Sitthisak S. High
prevalence of multi-drug resistant Streptococcus pneumoniae among healthy children in
Thailand. J Infect Public Health. 2015;8(3):274–81.
101. Jin P, Kong F, Xiao M, et al. First report of putative Streptococcus pneumoniae serotype 6D
among nasopharyngeal isolates from Fijian children. J Infect Dis. 2009;200(9):1375–80.
102. Oliver MB, van der Linden MP, Kuntzel SA, Saad JS, Nahm MH. Discovery of Streptococcus
pneumoniae serotype 6 variants with glycosyltransferases synthesizing two differing repeat-
ing units. J Biol Chem. 2015;290(44):26474–5.
2 Antimicrobial Resistance Among Streptococcus pneumoniae 37
103. Calix JJ, Porambo RJ, Brady AM, et al. Biochemical, genetic, and serological characteriza-
tion of two capsule subtypes among Streptococcus pneumoniae Serotype 20 strains: discov-
ery of a new pneumococcal serotype. J Biol Chem. 2012;287(33):27885–94.
104. Jauneikaite E, Tocheva AS, Jefferies JM, et al. Current methods for capsular typing of
Streptococcus pneumoniae. J Microbiol Methods. 2015;113:41–9.
105. Brueggemann AB, Peto TE, Crook DW, Butler JC, Kristinsson KG, Spratt BG. Temporal
and geographic stability of the serogroup-specific invasive disease potential of Streptococcus
pneumoniae in children. J Infect Dis. 2004;190(7):1203–11.
106. Browall S, Backhaus E, Naucler P, et al. Clinical manifestations of invasive pneumococcal
disease by vaccine and non-vaccine types. Eur Respir J. 2014;44(6):1646–57.
107. Sjostrom K, Spindler C, Ortqvist A, et al. Clonal and capsular types decide whether pneumo-
cocci will act as a primary or opportunistic pathogen. Clin Infect Dis. 2006;42(4):451–9.
108. Henriques-Normark B, Blomberg C, Dagerhamn J, Battig P, Normark S. The rise and fall of
bacterial clones: Streptococcus pneumoniae. Nat Rev Microbiol. 2008;6(11):827–37.
109. Meropol SB, Stange KC, Jacobs MR, Weiss JK, Bajaksouzian S, Bonomo RA. Bacterial
colonization and antibiotic resistance in a prospective cohort of newborn infants during the
first year of life. Open Forum Infect Dis. 2016;3(4):ofw221.
110. Boken DJ, Chartrand SA, Goering RV, Kruger R, Harrison CJ. Colonization with
penicillin-resistant Streptococcus pneumoniae in a child-care center. Pediatr Infect Dis
J. 1995;14(10):879–84.
111. Centers for Disease Control and Prevention (CDC). Antibiotic Resistance Threats in the
United States. 2013.
112. Navarro TA, Dias JG, Quinten C, et al. European enhanced surveillance of invasive pneumo-
coccal disease in 2010: data from 26 European countries in the post-heptavalent conjugate
vaccine era. Vaccine. 2014;32(29):3644–50.
113. European Centre for Disease Prevention and Control (ECDC). Antimicrobial resistance sur-
veillance in Europe 2012. 2012.
114. Jones RN, Sader HS, Mendes RE, Flamm RK. Update on antimicrobial susceptibility trends
among Streptococcus pneumoniae in the United States: report of ceftaroline activity from
the SENTRY Antimicrobial Surveillance Program (1998–2011). Diagn Microbiol Infect Dis.
2013;75(1):107–9.
115. Jenkins SG, Farrell DJ. Increase in pneumococcus macrolide resistance, United States.
Emerg Infect Dis. 2009;15(8):1260–4.
116. Xiao Y, Wei Z, Shen P, et al. Bacterial-resistance among outpatients of county hospitals
in China: significant geographic distinctions and minor differences between central cities.
Microbes Infect. 2015;17(6):417–25.
117. Cilloniz C, Albert RK, Liapikou A, et al. The effect of macrolide-resistance on the presenta-
tion and outcome of patients hospitalized for Streptococcus pneumoniae pneumonia. Am
J Respir Crit Care Med. 2015;191(11):1265–72.
118. Pletz MW, van der Linden M, von Baum H, Duesberg CB, Klugman KP, Welte T. Low preva-
lence of fluoroquinolone resistant strains and resistance precursor strains in Streptococcus
pneumoniae from patients with community-acquired pneumonia despite high fluoroquino-
lone usage. Int J Med Microbiol. 2011;301(1):53–7.
119. Jacobs E, Dalhoff A, Korfmann G. Susceptibility patterns of bacterial isolates from hospital-
ised patients with respiratory tract infections (MOXIAKTIV Study). Int J Antimicrob Agents.
2009;33(1):52–7.
120. Ho PL, Cheng VC, Chow KH. Decreasing prevalence of levofloxacin-resistant Streptococcus
pneumoniae in Hong Kong, 2001 to 2007. J Antimicrob Chemother. 2009;63(4):836–8.
121. Vila-Corcoles A, Ochoa-Gondar O. Preventing pneumococcal disease in the elderly: recent
advances in vaccines and implications for clinical practice. Drugs Aging. 2013;30(5):263–76.
122. Kyaw MH, Lynfield R, Schaffner W, et al. Effect of introduction of the pneumococcal conjugate
vaccine on drug-resistant Streptococcus pneumoniae. N Engl J Med. 2006;354(14):1455–63.
38 C. Cillóniz et al.
123. Whitney CG, Pilishvili T, Farley MM, et al. Effectiveness of seven-valent pneumococcal con-
jugate vaccine against invasive pneumococcal disease: a matched case-control study. Lancet.
2006;368(9546):1495–502.
124. Gertz RE Jr, Li Z, Pimenta FC, et al. Increased penicillin nonsusceptibility of nonvaccine-
serotype invasive pneumococci other than serotypes 19A and 6A in post-7-valent conjugate
vaccine era. J Infect Dis. 2010;201(5):770–5.
125. Reinert R, Jacobs MR, Kaplan SL. Pneumococcal disease caused by serotype 19A: review of
the literature and implications for future vaccine development. Vaccine. 2010;28(26):4249–59.
126. Moore MR, Link-Gelles R, Schaffner W, et al. Effect of use of 13-valent pneumococcal conju-
gate vaccine in children on invasive pneumococcal disease in children and adults in the USA:
analysis of multisite, population-based surveillance. Lancet Infect Dis. 2015;15(3):301–9.
127. van der Linden M, Perniciaro S, Imohl M. Increase of serotypes 15A and 23B in IPD in
Germany in the PCV13 vaccination era. BMC Infect Dis. 2015;15:207.
128. Garcia-Vidal C, Ardanuy C, Tubau F, et al. Pneumococcal pneumonia presenting with septic
shock: host- and pathogen-related factors and outcomes. Thorax. 2010;65(1):77–81.
129. Mongardon N, Max A, Bougle A, et al. Epidemiology and outcome of severe pneumococcal
pneumonia admitted to intensive care unit: a multicenter study. Crit Care. 2012;16(4):R155.
130. Kaplan SL, Mason EO Jr. Management of infections due to antibiotic-resistant Streptococcus
pneumoniae. Clin Microbiol Rev. 1998;11(4):628–44.
131. Schroeder MR, Stephens DS. Macrolide resistance in Streptococcus pneumoniae. Front Cell
Infect Microbiol. 2016;6:98.
132. Endimiani A, Brigante G, Bettaccini AA, Luzzaro F, Grossi P, Toniolo AQ. Failure of levo-
floxacin treatment in community-acquired pneumococcal pneumonia. BMC Infect Dis.
2005;5:106.
133. Kim L, McGee L, Tomczyk S, Beall B. Biological and epidemiological features of antibiotic-
resistant Streptococcus pneumoniae in pre- and post-conjugate vaccine eras: a United States
perspective. Clin Microbiol Rev. 2016;29(3):525–52.
134. Kang CI, Song JH, Kim SH, et al. Association of levofloxacin resistance with mortality in
adult patients with invasive pneumococcal diseases: a post hoc analysis of a prospective
cohort. Infection. 2013;41(1):151–7.
Chapter 3
Emergence of MRSA in the Community
Staphylococcus aureus has been recognized as a cause of human infection for over
100 years, and its role in causing clinical syndromes, such as sepsis and abscesses,
was first described by Ogston in the late nineteenth century [119]. S. aureus can
colonize human hosts without causing disease, but infections with S. aureus, espe-
cially antimicrobial-resistant varieties such as methicillin-resistant S. aureus
(MRSA), contribute significantly to the burden of infectious diseases in humans.
Penicillin was first introduced to treat patients with bacterial infections in 1941,
and resistance to penicillin was first reported in S. aureus within 1–2 years [87].
These resistant strains were first found in hospitals after the Second World War,
where patients were exposed to this new antimicrobial agent [7]. S. aureus had
quickly acquired the ability to produce penicillinase, an enzyme that inactivates
penicillin. An “epidemic strain” of penicillin-resistant S. aureus, which was charac-
teristically lysed by bacteriophage 80 and 81, was noted to cause hospital outbreaks
in Australia, Canada, and the United States in the 1950s, particularly in hospitalized
children and otherwise healthy young adults [50].
In Denmark in the late 1960s, the first large-scale study of penicillin-resistant
S. aureus discovered that not only was the majority of S. aureus found in hospitals
resistant to penicillin but also the resistance gene had spread to a majority of
S. aureus strains collected from patients in community settings [79]. Within a
decade, the majority of community S. aureus strains in the United States were
penicillin-resistant [132]. New drug development provided a solution to penicillin-
resistant strains with the release of the semisynthetic penicillins (e.g., methicillin,
oxacillin, nafcillin) that resisted penicillinases produced by the majority of S. aureus
strains.
The first S. aureus isolates resistant to methicillin were isolated from patients in
England within months of the introduction of methicillin in 1959 [80]. Reports of
MRSA in the United States soon followed [8]. As with penicillin-resistant S. aureus,
MRSA strains were first seen in hospitals, prompting concerns that MRSA would
spread outside the hospital. Over 50 years later, MRSA has established itself as a
common cause of infections in community settings.
In this chapter, we discuss the epidemiology and mechanisms of resistance of
community-associated MRSA (CA-MRSA), an approach to management of
CA-MRSA infections, recommendations for prevention of MRSA in the commu-
nity, and future directions for research, focusing mostly on CA-MRSA in the United
States.
In 1999, the Centers for Disease Control and Prevention (CDC) published a report
concerning four children from 12 months to 13 years of age who had died from
MRSA infections [17]. None of the children had risk factors for healthcare-
associated MRSA (HA-MRSA), which at that time included recent hospitalization
or surgery, residence in a long-term care facility, or a history of injection drug use.
Although CA-MRSA had been recently reported in children [1, 71], most previ-
ously described MRSA infections in the United States were associated with health-
care settings or injection drug use in adults [93, 135]. These four cases demonstrated
that not only could patients develop MRSA disease outside of the hospital but also
that CA-MRSA disease could be severe or fatal. In addition to the four pediatric
deaths, CA-MRSA infections were reported in other populations, such as prisoners
[18] and military personnel [81]. In response to these reports, CDC initiated active
surveillance to describe the epidemiology and drug resistance patterns of MRSA
isolates in the community in the United States [56].
This and other early studies demonstrated that there were CA-MRSA strains
with unique characteristics compared to methicillin-susceptible S. aureus (MSSA)
and traditional healthcare-associated MRSA strains. Infections from these
CA-MRSA strains often had different epidemiologic risk factors, clinical manifes-
tations, and microbiological characteristics than HA-MRSA strains.
In the literature, the designations CA-MRSA and HA-MRSA have been used to
describe both distinct MRSA strains and infections with different epidemiologic
3 Emergence of MRSA in the Community 41
Virulence factors enhance the ability of bacteria to cause infection by evading the
host’s defenses, increasing adherence to tissues, or spreading through tissues.
Examples of virulence factors in S. aureus include production of coagulase, toxins,
3 Emergence of MRSA in the Community 43
and proteins intrinsic to the cell wall. S. aureus produces coagulase, which interacts
with fibrinogen causing plasma to clot. This clumping creates a loose polysaccha-
ride capsule that can interfere with phagocytosis. The combination of these viru-
lence factors may cause localization of an infection, such as in an abscess, a common
clinical manifestation of CA-MRSA infection.
Panton-Valentine leukocidin (PVL) is a cytotoxin (coded by the lukS-PV and
lukF-PV genes) first identified in methicillin-susceptible S. aureus [121]. PVL kills
leukocytes by creating pores in the cell membrane of affected cells or by activating
apoptosis pathways. Pore formation leads to increased cell wall permeability and
leakage of protein from the cell causing cell death and tissue necrosis. PVL genes
have been associated with severe abscesses, necrotizing pneumonia, and increased
complications in osteomyelitis [96, 102]. PVL genes are found in most CA-MRSA
strains, such as with pulsed-field gel electrophoresis type USA300. PVL, however,
is not limited to CA-MRSA, as the toxin is also found in the majority of MSSA
strains isolated from patients with community-acquired skin and soft tissue infec-
tions (SSTIs) [64]. While PVL is rare in other S. aureus strain collections such as
colonization or clinical isolates from bloodstream infections, it is highly associated
with SSTIs [144].
In addition to PVL, other toxins may be produced by S. aureus: α-toxin, which
causes tissue necrosis and acts on cell membranes; exfoliative toxins ETA, ETB,
and ETD, which are encoded on different genetic elements and cause skin separa-
tion in diseases such as bullous impetigo and staphylococcal scalded skin syndrome;
enterotoxins A–E, G–J, and R–T (SEA-SEE, SEG-SEJ, SER-SET) which can cause
vomiting and diarrhea associated with food poisoning; and toxic shock syndrome
toxin I (TSST-1) which induces production of interleukin-1 and tumor necrosis fac-
tor leading to shock [64]. Peptidoglycans, which comprise 50% by weight of the
cell wall of staphylococci, can have endotoxin properties as well. Other cell wall
polymers (e.g., teichoic acid) and cell surface proteins (e.g., protein A and fibronec-
tin- and collagen-binding proteins) may also be virulence factors for S. aureus [64].
Recent DNA sequencing of the most common molecular type of CA-MRSA
(USA300) suggests that encoded gene products might enhance the ability of the
strain to live on the host’s skin [41].
Molecular typing of MRSA strains is used to link cases in a cluster, locate sources
of specific outbreaks, and conduct macroepidemiology and evolutionary studies.
Using the antimicrobial agent susceptibility profile to determine genetic related-
ness of strains of S. aureus is unreliable. Historically, pulsed-field gel electropho-
resis (PFGE) was one of the most commonly used methods for MRSA strain typing
in outbreak investigations. Pulsed-field types are still commonly recognized in the
United States and around the world. Sequence-based typing methods such as
multi-locus sequence typing (MLST) and spa (Staphylococcal protein A) typing
have been used in more recent years for the analysis of long-term epidemiology
3 Emergence of MRSA in the Community 45
and evolution of MRSA. MLST and spa typing have been found to be highly con-
cordant, and both typing methods are easier and less costly than performing PFGE,
providing unambiguous typing results that can be compared between laboratories
and over time [120]. An early application of whole genome sequencing was to
characterize MRSA outbreaks in healthcare settings [89].The requirement for bio-
informatics expertise is one factor that has limited adoption by healthcare institu-
tions. Nonetheless, advances in WGS and the accessibility of a large number of
assembled bacterial genomes have made possible new methods for strain-level
epidemiologic tracking of isolates. This includes the ability to apply MLST
schemes on a genome- wide scale [37]. Phylogenetic analysis, derived from
sequencing results, has been used to study the population structures of outbreaks
and to build transmission networks and help identify factors associated with
USA300 strains [127].
In the United States, a limited number of MRSA strains have been implicated in
most community outbreaks. A recent description of clinical MRSA isolates from 43
centers across the United States indicated that the USA300 pulsed-field type was the
most common type in all regions and from all specimen sources [40]. Other com-
munity MRSA genotypes (USA400, USA1000, and USA1100) also cause disease
[103]. Molecular typing has further classified the USA300 strain as ST8 and most
commonly spa type t008. The highly conserved USA300 strain (USA300–0114)
has been implicated in multiple outbreaks across the United States in diverse popu-
lations that are not epidemiologically related, such as athletes, prisoners, and chil-
dren [85].
As discussed earlier, some of the first reports of MRSA infection without traditional
healthcare-associated risk factors were in children [17, 71]. The age distribution for
noninvasive CA-MRSA syndromes, such as SSTI, is not well described. However,
from population-based surveillance data for MRSA in the United States, invasive
(i.e., isolated from a normally sterile body site) CA-MRSA infections are most
common in persons aged 50 or older and in children less than 1 year of age [26].
Age differences exist in nasal colonization rates as well. A national study of S.
aureus nasal colonization among noninstitutionalized persons at least 1 year of age
showed that persons age 60 years or older had the highest odds of MRSA nasal car-
riage; however, this does not specifically refer to carriage of USA300 MRSA or
other community MRSA strains. For example, the same study reported that among
persons colonized with MRSA, younger persons most often had USA300 or
USA400 community MRSA strains [62]. In addition, older age was associated with
decreased odds of USA300 MRSA nasal colonization in one study [55].
Trends for MRSA colonization by sex have varied. S. aureus nasal carriage has
been reported to be more common in men than women in most studies [62, 76].
Some studies have shown higher prevalence of MRSA nasal carriage in men [76],
but a large population study of nasal carriage in the United States in 2001–2004 did
not find significantly higher prevalence of MRSA carriage in men [62]. Most studies
3 Emergence of MRSA in the Community 47
Rates of CA-MRSA infection vary between racial and ethnic groups. Compared
with Australians of European descent, high rates of CA-MRSA infection have been
noted in Maori and Pacific Islanders in Australia [72]. Similarly, Pacific Islanders in
Hawaii have much higher rates of CA-MRSA infection than Hawaiians of Asian
descent. In one investigation of MRSA in Hawaii, 76% of patients with CA-MRSA
were Pacific Islanders, but only 35% of patients who received care in the facility
were Pacific Islanders [48]. Rates of CA-MRSA SSTIs are high among Alaskan
natives [128].
Population-based surveillance data from the continental United States have con-
sistently shown higher incidence rates of invasive CA-MRSA infection among
black versus white persons, in both adult and pediatric populations [65, 67, 78, 88,
129]. African-American race has also been described as an independent risk factor
for colonization with the USA300 strain [55]. Analysis of public health surveillance
has demonstrated that when socioeconomic factors such as income and crowding
are accounted for, no significant differences in invasive community-associated
MRSA rates by race remain [141]. However, the specific mechanisms by which dif-
ferences in socioeconomic status lead to racial disparities in invasive community-
associated MRSA have not yet been elucidated. It may occur through differences in
concurrent diseases such as diabetes that result from socioeconomic disparities or
from factors such as limited access to healthcare and crowded housing conditions,
which are more common in some racial groups [48]. For example, a study in urban
Chicago has described incarceration and public housing as risk factors for
CA-MRSA SSTIs [74]. The importance of socioeconomic factors as a contributor
to CA-MRSA rates is underscored by a study showing transient residence and sub-
stance abuse as factors associated with CA-MRSA compared to patients with
MSSA [159].
HIV infection, illicit drug use, temporary housing, and incarceration are associated
with colonization with USA300 MRSA [126]. In addition, antimicrobial use is also
a risk factor for MRSA colonization in the community [47, 105, 106].
48 L. P. Gleason et al.
Clinical treatment guidelines for MRSA were published by the Infectious Disease
Society of America (IDSA) in 2011 [97]. This section will review major recommen-
dations from those guidelines as well as new developments that have occurred since
then.
MRSA has increased in prevalence as a cause of purulent SSTIs since the early
2000s and is the most common cause in many communities [151]. Some of the
major considerations for treatment of these infections outlined in the IDSA guide-
lines, along with a brief discussion of the literature, are listed below:
3 Emergence of MRSA in the Community 49
1. Incision and drainage should be routine for skin lesions that can be drained.
Some clinicians suggest the use of ultrasonography to distinguish if there is a
drainable collection [148]. Incision and drainage has been the primary mode of
treatment for skin and soft tissue abscesses for many centuries and is essential.
Although the 2011 IDSA guidelines suggest that additional data are needed to
describe whether antimicrobials are needed for simple abscesses/boils, a ran-
domized trial at five US emergency departments showed that trimethoprim-
sulfamethoxazole (TMP-SMX) treatment led to higher cure rates among patients
with a drained cutaneous abscess compared with patients who received placebo
[152]. Antimicrobial treatment may also reduce recurrent infections [45, 139,
152] and the need for subsequent procedures [152]. Available literature suggests
that clinicians should weigh the costs and benefits of prescribing antimicrobial
agents: for example, considering an individual patient’s risk for recurrent
infections.
2. Collect diagnostic specimens for culture. Cultures should be obtained from
patients with both draining and non-draining purulent lesions. Obtaining isolates
helps guide treatment of individual patients and monitor the antimicrobial agent
susceptibility patterns in the community. Cultures are also recommended by
some experts when there is a severe local infection, signs of systemic illness,
inadequate response to initial treatment, or concerns about a cluster or outbreak
of cases [148].
3. Use of antimicrobial agents. For adult patients, the IDSA guidelines suggest
using clindamycin, TMP-SMX, a tetracycline, or linezolid for empiric coverage
of CA-MRSA in outpatients with SSTI when empiric coverage is needed, with
the additional suggestion that if coverage for β-hemolytic streptococci is also
needed, a beta-lactam be added if TMP-SMX or a tetracycline is prescribed. For
hospitalized patients with complicated SSTI, empiric therapy is suggested to
include vancomycin, linezolid, daptomycin, telavancin, or clindamycin. Local
antimicrobial susceptibility data for outpatient S. aureus SSTIs should guide
empiric treatment decisions. The 2011 IDSA guidelines specifically suggest
empiric therapy for CA-MRSA be prescribed for outpatients with purulent cel-
lulitis, hospitalized patients with complicated SSTI, and abscess associated with
certain conditions—severe or extensive disease or rapid progression with cellu-
litis, signs/symptoms of systemic illness, immunosuppression, extremes of age,
lack of response to initial treatment, septic phlebitis, or anatomically difficult-to-
drain location. The guidelines state that the role of CA-MRSA in outpatient non-
purulent cellulitis is unknown, and despite use of state-of-the-art techniques for
molecular identification, the microbiologic etiology of nonpurulent cellulitis in
the community and role of MRSA in particular remain unknown [32]. Notably a
recent study has shown no difference in outcomes with the use of cephalexin
alone compared with the use of cephalexin plus TMP-SMX in patients with
uncomplicated cellulitis [109].
4. Perform careful and thorough personal and environmental hygiene. MRSA can
be transmitted from person to person or to the environment through contact with
draining skin and soft tissue lesions. After incision and drainage, wounds should
50 L. P. Gleason et al.
Several antimicrobials have been developed more recently with in vitro activity
against MRSA, many of which also have indications for skin infections or other
syndromes that may be CA-MRSA infections (Table 3.2). These include dalba-
vancin, oritavancin, ceftaroline, and tedizolid [14]. Some clinicians have suggested
combination therapy (e.g., either vancomycin or daptomycin, in conjunction with a
β-lactam) for complicated MRSA bacteremia [39]. Clinical trials are underway to
further define the role of combination therapy [154]. In addition, a new clinical
treatment guideline for S. aureus bacteremia is planned by IDSA and may clarify
the role of newer agents or combination therapy for bacteremia for treatment of
CA-MRSA syndromes.
3 Emergence of MRSA in the Community 51
Table 3.2 Description of major antimicrobials or antimicrobial classes that have been used to treat
methicillin-resistant Staphylococcus aureus infections
Antimicrobial Mechanism of action Other comments
Ceftaroline Binding to penicillin-binding High affinity for penicillin-binding protein
proteins (fifth-generation 2A (PBP2A), leading to greater in vitro
cephalosporin) activity against MRSA than other
cephalosporin antibiotics
Clindamycin Inhibition of bacterial protein Inducible resistance can occur
synthesis
Dalbavancin Same as vancomycin Dosing for skin infections approved as
either single dose or two doses 1 week
apart
Daptomycin Cyclic lipopeptide; binds to Inactivated by surfactant and not
bacterial cell membranes recommended for treatment of MRSA
pneumonia
Linezolid Inhibition of ribosomal Myelosuppression reported; could cause
protein synthesis serotonin syndrome in conjunction with
some medications (e.g., antidepressants)
Oritavancin Inhibits cell wall synthesis Once/week dosing
(similar to vancomycin) and
disrupts cell membrane
barrier function
Quinupristin- Inhibits peptide bond Arthralgias/myalgias are common adverse
dalfopristin formation in ribosome events
Tedizolid Same as linezolid Approved for skin infections; advantage
over linezolid is once/day dosing
Telavancin Inhibits cell wall synthesis Synthetic derivative of vancomycin but
(similar to vancomycin) and once/day dosing
disrupts cell membrane
barrier function
Tetracyclines (class) Inhibition of bacterial protein Not recommended in children because of
synthesis effects on bones and teeth
Tigecycline Inhibits protein translation FDA boxed warning about increased risk
of death
Trimethoprim- Blocks production of folic
sulfamethoxazole acid
Vancomycin Inhibits cell wall synthesis Monitoring of serum levels recommended
by IDSA guidelines; dosing in obese
patients controversial; often drug of choice
for patients with bacteremia
mupirocin may limit the effectiveness of nasal decolonization [98]. Proposed regi-
mens for decolonization include intranasal mupirocin twice a day for 5–10 days and
antiseptic body wash, such as with chlorhexidine, for 5–14 days [97, 150]. Dilute
bleach baths have been suggested for patients with recurrent MRSA infections, but
children who underwent routine hygienic measures plus twice weekly bleach baths
for 3 months did not experience a significant reduction in recurrent SSTI requiring
medical attention within a year of treatment compared with those using routine
hygienic measures [83]. A randomized controlled trial testing the effect of skin
cleaning with chlorhexidine gluconate (CHG) three times per week for 6 months in
a jail showed no difference in MRSA carriage between groups cleaning with CHG
cloths compared with water-soaked cloths [35]. When household transmission is
suspected, implementation of personal and environmental hygiene measures and
evaluation of symptomatic contacts are recommended in IDSA guidelines, which
state that decolonization may be considered [97]. Decolonization with a 5-day regi-
men of hygiene, nasal mupirocin, and chlorohexidine body washes of all household
members was associated with a significant decrease in self-reported recurrent SSTI
at 12-month follow-up compared to a group in which only the index patient was
decolonized [57]. However, SSTI still recurred in the majority of cases [57]. The
use of oral antimicrobial therapy is generally not recommended for decolonization
in current guidelines [97].
Prevention strategies for CA-MRSA need to include public health officials, medical
providers and infection control practitioners, and patients and community members.
The following are considerations for these different groups for prevention of
CA-MRSA.
ity of illness, the presence of ongoing transmission, and the likelihood that an
intervention could be successfully implemented.
2. Enhance surveillance. Both prospective and retrospective surveillance are impor-
tant to identify cases of MRSA in the community and intervene in outbreak set-
tings. Consider notifying contacts of patients with MRSA infection to identify
new cases in outbreak settings and to ensure that they are receiving proper
treatment.
New mechanisms of preventing and treating S. aureus infection are being studied.
Vaccines against S. aureus are being developed. Although significant difficulties
have been encountered with previous studies of potential S. aureus vaccines, there
are clinical trials of vaccine candidates ongoing [10, 59, 100, 138]. In addition,
some studies of antibodies to treat or prevent S. aureus infections are also underway
[59]. Another research area that has attracted some interest recently is the use of
bacteriophages for prevention or treatment of S. aureus infections, though the appli-
cation of such an approach is likely distant [86].
might clarify the dynamics of, risk factors for, and importance of this potential hori-
zontal transmission between pets and humans.
CA-MRSA strains have been increasingly described as the cause of disease and
outbreaks in healthcare settings. USA300 MRSA was the most common MRSA
strain causing bloodstream infections in a Chicago hospital and has been reported to
make up an increasing proportion of hospital-onset MRSA infections [34, 130].
Surveillance from CDC’s EIP suggests that the increase in the proportion of
hospital-
onset MRSA infections caused by USA300 has occurred principally
because the incidence of MRSA infections caused by other strains in hospitals has
declined [140]. Data from the same surveillance system show that the incidence of
invasive USA300 infections has not decreased over the past decade, and additional
strategies for preventing USA300 MRSA infections in both healthcare and the com-
munity are needed.
3 Emergence of MRSA in the Community 59
Acknowledgments/Disclaimer The findings and conclusions in this report are those of the
authors and do not necessarily represent the official position of the Centers for Disease Control and
Prevention. Use of trade names and commercial sources is for identification only and does not
imply endorsement by the Centers for Disease Control and Prevention, the Public Health Service,
or the US Department of Health and Human Services. The authors would like to thank the authors
of the previous version of this chapter (Adam Cohen, Daniel Jernigan, and Rachel Gorwitz) for
permission to include some materials from the previous edition.
References
Shieh WJ, Zaki SR, Sejvar JJ, Shay DK, Harper SA, Cox NJ, Fukuda K, Uyeki TM, Influenza
Special Investigations Team. Influenza-associated deaths among children in the United
States, 2003-2004. N Engl J Med. 2005;353:2559–67.
12. Calfee DP, Salgado CD, Milstone AM, Harris AD, Kuhar DT, Moody J, Aureden K, Huang
SS, Maragakis LL, Yodoe DS, Society for Healthcare Epidemiology of America. Strategies to
prevent methicillin-resistant Staphylococcus aureus transmission and infection in acute care
hospitals: 2014 update. Infect Control Hosp Epidemiol. 2014;35:s108–32.
13. Campbell KM, Vaughn AF, Russell KL, Smith B, Jimenez DL, Barrozo CP, Minarcik JR, Crum
NF, Ryan MA. Risk factors for community-associated methicillin-resistant Staphylococcus
aureus infections in an outbreak of disease among military trainees in San Diego, California,
in 2002. J Clin Microbiol. 2004;42:4050–43.
14. Cardona AF, Wilson SE. Skin and soft-tissue infections: a critical review and the role of tela-
vancin in their treatment. Clin Infect Dis. 2015;61:S69–78.
15. Casey JA, Curriero FC, Cosgrove SE, Nachman KE, Schwartz BS. High-density livestock
operations, crop field application of manure, and risk of community-associated methicillin-
resistant Staphylococcus aureus infection in Pennsylvania. JAMA Int Med. 2013;173:1980.
16. Cefai C, Ashurst S, Owens C. Human carriage of methicillin-resistant Staphylococcus aureus
linked with pet dog. Lancet. 1994;344:539–40.
17. Centers for Disease Control and Prevention. Four pediatric deaths from community-acquired
methicillin-resistant Staphylococcus aureus—Minnesota and North Dakota, 1997-1999.
MMWR Morb Mortal Wkly Rep. 1999;48:707–10.
18. Centers for Disease Control and Prevention. Methicillin-resistant Staphylococcus aureus skin
or soft tissue infections in a state prison—Mississippi, 2000. MMWR Morb Mortal Wkly
Rep. 2001;50:919–22.
19. Centers for Disease Control and Prevention. Staphylococcus aureus Resistant to vancomy-
cin—United States, 2002. MMWR Morb Mortal Wkly Rep. 2002;51:565–7.
20. Centers for Disease Control and Prevention. Vancomycin-resistant Staphylococcus aureus—
Pennsylvania, 2002. MMWR Morb Mortal Wkly Rep. 2002;51:902.
21. Centers for Disease Control and Prevention. Methicillin-resistant Staphylococcus aureus
infections among competitive sports participants—Colorado, Indiana, Pennsylvania, and Los
Angeles County, 2000-2003. MMWR Morb Mortal Wkly Rep. 2003;52:793.
22. Centers for Disease Control and Prevention. Methicillin-resistant Staphylococcus aureus
infections in correctional facilities—Georgia, California, and Texas, 2001-2003. MMWR
Morb Mortal Wkly Rep. 2003;52:992–6.
23. Centers for Disease Control and Prevention. Infectious disease and dermatologic conditions
in evacuees and rescue workers after hurricane Katrina—multiple states, august-September,
2005. MMWR Morb Mortal Wkly Rep. 2005;54:961–4.
24. Centers for Disease Control and Prevention. Methicillin-resistant Staphylococcus aureus
skin infections among tattoo recipients—Ohio, Kentucky, and Vermont, 2004-2005. MMWR
Morb Mortal Wkly Rep. 2006;55:677–9.
25. Centers for Disease Control and Prevention. Severe co-infection with seasonal influenza a
(H3N2) virus and Staphylococcus aureus – Maryland, February-march 2012. MMWR Morb
Mortal Wkly Rep. 2012;61:289–91.
26. Centers for Disease Control and Prevention. Emerging infections program network report:
methicillin-resistant Staphylococcus aureus, 2015. Published February 28, 2017. Available
online at: https://www.cdc.gov/hai/eip/pdf/2015-MRSA-Report.pdf. 2017.
27. Chambers HF. The changing epidemiology of Staphylococcus aureus? Emerg Infect Dis.
2001;7:178–82.
28. Clinical and Laboratory Standards Institute. Performance standards for antimicrobial suscep-
tibility testing; 16th informational supplement. Wayne: The Institute; 2006. p. M100–S16.
29. Cohen AL, Shuler C, McAllister S, Fosheim GE, Brown MG, Abercrombie D, Anderson
K, McDougal LK, Drenzek C, Arnold K, Jernigan D, Gorwitz R. Methamphetamine
3 Emergence of MRSA in the Community 61
use and methicillin-resistant Staphylococcus aureus skin infections. Emerg Infect Dis.
2007;13:1707–13.
30. Cohen PR. Community-acquired methicillin-resistant Staphylococcus aureus skin infections:
implications for patients and practitioners. Am J Clin Dermatol. 2007;8:259–70.
31. Coronado F, Nicholas JA, Wallace BJ, Kohlerschmidt DJ, Musser K, SchoonmakerBopp DJ,
Zimmerman SM, Boller AR, Jernigan DB, Kacica MA. Community-associated methicillin-
resistant Staphylococcus aureus skin infections in a religious community. Epidemiol Infect.
2006;135:492–501.
32. Crisp JG, Takhar SS, Moran GJ, Krishnadasan A, Dowd SE, Finegold SM, Summanen PH,
Talan DA, EMERGency ID Net Study Group. Inability of polymerase chain reaction, pyro-
sequencing, and culture of infected and uninfected site skin biopsy specimens to identify the
cause of cellulitis. Clin Infect Dis. 2015;61:1679–87.
33. David MZ, Rudolph KM, Hennessey TW, Boyle-Vavra B, Daum RS. Molecular epidemiol-
ogy of methicillin-resistant Staphylococcus aureus, rural southwestern Alaska. Emerg Infect
Dis. 2008;14:1693–9.
34. David MZ, Cadilla A, Boyle-Vavra S, Daum RS. Replacement of HA-MRSA by CA-MRSA
infections at an academic medical center in the Midwestern United States, 2004-2005 to
2008. PLoS One. 2014;9:e92760.
35. David MZ, Siegel JD, Henderson J, Leos G, Lo K, Iwuora J, Porsa E, Schumm LP, Boyle-
Vavra S, Daum RS. A randomized, controlled trial of Chlorhexidine-soaked cloths to reduce
methicillin-resistant and methicillin-susceptible Staphylococcus aureus carriage prevalence
in an urban jail. Infect Control Hosp Epidemiol. 2014;35:1466–73.
36. Davis J, Jackson C, Fedorka-Cray P, Barrett J, Brousse J, Gustafson J, Kucher M. Carriage
of methicillin-resistant staphylococci by healthy companion animals in the US. Lett Appl
Microbiol. 2014;59:1–8.
37. Dekker JP, Frank KM. Next-generation epidemiology: using real-time core genome multilo-
cus sequence typing to support infection control policy. J Clin Microbiol. 2016;54:2850–3.
38. DeLeo FR, Otto M, Kreiswirth BN, Chambers HF. Community-associated methicillin-
resistant Staphylococcus aureus. Lancet. 2010;375:1557–68.
39. Dhand A, Sakoulas G. Daptomycin in combination with other antibiotics in the treat-
ment of complicated methicillin-resistant Staphylococcus aureus bacteremia. Clin Ther.
2014;36:1303–16.
40. Diekema DJ, Richter SS, Heilmann KP, Dohrn CL, Riahi F, Tendolkar S, JS MD, Doern
GV. Continued emergence of USA300 methicillin-resistant Staphylococcus aureus in
the United States: results from a nationwide study. Infect Control Hosp Epidemiol.
2014;35:285–92.
41. Diep BA, Gill SR, Chang RF, Phan TH, Chen JH, Davidson MG, Lin F, Lin J, Carleton
HA, Mongodin EF, Sensabaugh GF, Perdreau-Remington F. Complete genome sequence
of USA300, an epidemic clone of community-acquired meticillin-resistant Staphylococcus
aureus. Lancet. 2006;367:731–9.
42. Dominguez TJ. It’s not a spider bite, it’s community-acquired methicillin-resistant
Staphylococcus aureus. Am Board Fam Med. 2004;17:220–6.
43. Duijekeren EV, Wolfhagen MJ, Box AT, Heck ME, Wannet WJ, Fluit AC. Human-to-dog trans-
mission of methicillin-resistant Staphylococcus aureus. Emerg Infect Dis. 2004;10:2235–7.
44. Duijekeren E, Wolfhagen MJ, Heck ME, Wannet WJ. Transmission of a panton-valentine
leucocidin-positive, methicillin-resistant Staphylococcus aureus strain between humans and
a dog. J Clin Microbiol. 2005;43:6209–11.
45. Duong M, Markwell S, Peter J, Barenkamp S. Randomized, controlled trial of antibiotics
in the Management of Community-Acquired Skin Abscesses in the pediatric patient. Ann
Emerg Med. 2010;55:401–7.
46. Ekkelenkamp MB, Sekkat M, Carpaij N, Troelstra A, Bonten MJ. Endocarditis due to
methicillin-resistant Staphylococcus aureus originating from pigs. Ned Tijdschr Geneeskd.
2006;150:2442–7.
62 L. P. Gleason et al.
47. Ellis MW, Hospenthal DR, Dooley DP, Gray PJ, Murray CK. Natural history of community-
acquired methicillin-resistant Staphylococcus aureus colonization and infection in soldiers.
Clin Infect Dis. 2004;39:971–9.
48. Estivariz CF, Park SY, Hageman JC, Dvorin J, Melish MM, Arpon R, Coon P, Slavish S,
Kim M, McDougal LK, Jensen B, McAllister S, Lonsway D, Killgore G, Effler PE, Jernigan
DB. Emergence of community-associated methicillin-resistant Staphylococcus aureus in
Hawaii, 2001-2003. J Infect. 2007;54:349–57.
49. Faires MC, Tater KC, Weese JS. An investigation of methicillin-resistant Staphylococcus
aureus colonization in people and pets in the same household with an infected person or
infected pet. J Am Vet Med Assoc. 2009;235:540–3.
50. Fekety FR, Bennett IL Jr. The epidemiological virulence of staphylococci. Yale J Biol Med.
1959;32:23–32.
51. Ferreira JP, Anderson KL, Correa MT, Lyman R, Ruffin F, Reller LB, Fowler VG. Transmission
of MRSA between companion animals and infected human patients presenting to outpatient
medical care facilities. PLoS One. 2011;6:e26978.
52. Fiebelkorn KR, Crawford SA, McElmeel ML, Jorgensen JH. Practical disk diffusion method
for detection of inducible clindamycin resistance in Staphylococcus aureus and coagulase-
negative staphylococci. J Clin Microbiol. 2003;41:4740–4.
53. Fleisch F, Zbinden R, Vanoli C, Ruef C. Epidemic spread of a single clone of methicillin-
resistant Staphylococcus aureus among injection drug users in Zurich, Switzerland. Clin
Infect Dis. 2001;32:581–6.
54. Frank AL, Marcinak JF, Mangat PD, Tjhio JT, Kelkar S, Schreckenberger PC, Quinn
JP. Clindamycin treatment of methicillin-resistant Staphylococcus aureus infections in chil-
dren. Pediatr Infect Dis J. 2002;21:530–4.
55. Freitas EAF, Harris RM, Blake RK, Salgado CD. Prevalence of USA300 strain type of
methicillin-resistant Staphylococcus aureus among patients with nasal colonization identi-
fied with active surveillance. Infect Control Hosp Epidemiol. 2010;31:469–75.
56. Fridkin SK, Hageman JC, Morrison M, Sanza LT, Como-Sabetti K, Jernigan JA, Harriman K,
Harrison LH, Lynfield R, Farley MM. Methicillin-resistant Staphylococcus aureus disease in
three communities. N Engl J Med. 2005;352:1436–44.
57. Fritz SA, Hogan PG, Hayek G, Eisenstein KA, Rodriguez M, Epplin EK, Garbutt J, Fraser
VJ. Household versus individual approaches to eradication of community-associated
Staphylococcus aureus in children: a randomized trial. Clin Infect Dis. 2011;54:743–51.
58. Fritz SA, Hogan PG, Singh LN, Thompson RM, Wallace MA, Whitney K, Al-Zubeidi D,
Burnham CD, Fraser VJ. Contamination of environmental services with Staphylococcus
aureus in households with children infected with methicillin-resistant S aureus. JAMA Ped.
2014;168:1030–8.
59. Giersing BK, Dastgheyb SS, Modjarrad K, Moorthy V. Status of vaccine research and devel-
opment of vaccines for Staphylococcus aureus. Vaccine. 2016;34:2962–6.
60. Gingrich EN, Kurt T, Hyatt DR, Lappin MR, Ruch-Gallie R. Prevalence of methicillin-resistant
staphylococci in northern Colorado shelter animals. J Vet Diagn Investig. 2011;23:947–50.
61. Glaser P, Martins-Simoes P, Villain A, Barbier M, Tristan A, Bouchier C, Ma L, Bes M,
Laurent F, Guillemot D, Wirth T, Vandenesch F. Demography and Intercontinental spread of
the USA300 community acquired methicillin-resistant Staphylococcus aureus lineage. mBio.
2016;7:e02183–15.
62. Gorwitz RJ, Kruszon-Moran D, McAllister SK, McQuillan G, McDougal LK, Fosheim
GE, Jensen BJ, Killgore G, Tenover FC, Kuehnert MJ. Changes in the prevalence of nasal
colonization with Staphylococcus aureus in the United States, 2001-2004. J Infect Dis.
2008;197:1226–34.
63. Groom AV, Wolsey DH, NaimI TS, Smith K, Johnson S, Boxrud D, Moore KA, Cheek
JE. Community-acquired methicillin-resistant Staphylococcus aureus in a rural American
Indian community. JAMA. 2001;286:1201–5.
3 Emergence of MRSA in the Community 63
64. Grumann D, Nübel U, Bröker BM. Staphylococcus aureus Toxins--their functions and genet-
ics. Infect Genet Evol. 2014;21:583–92.
65. Gualandi NR, Mu Y, Dumyati G, Harrison L, Lesher L, Nadle L, Petit S, Ray SM, Schaffner W,
Townes J, Fridkin S, See I. Racial disparities in invasive methicillin-resistant Staphylococcus
aureus infections, 2005–2012. 2015 Council of State and Territorial Epidemiologists Annual
Conference, Boston, MA. Presented June 15, 2015. Abstract #4203; 2015.
66. Hageman JC, UyekI TM, Francis JS, Jernigan DB, Wheeler JG, Bridges CB, Barenkamp SJ,
Sievert DM, Srinivasan A, Doherty MC, McDougal LK, Killgore GE, Lopatin UA, Coffman
R, MacDonald JK, McAllister SK, Fosheim GE, Patel JB, McDonald C. Severe community-
acquired pneumonia caused by an emerging strain of Staphylococcus aureus during the 2003-
04 influenza season. Emerg Infect Dis. 2006;12:894–9.
67. Ham DC, Jackson K, Bamberg W, Dumyati G, Harrison L, Lynfield R, Nadle J, Petit S, Ray
SM, Schaffner W, Townes J, Kallen AJ, See I. Community associated invasive methicillin-
resistant Staphylococcus aureus (MRSA), 2005–2014. Spring 2017 Conference for the
Society for Healthcare Epidemiology of America, St Louis, MO. Presented March 31, 2017.
Abstract #8707; 2017.
68. Hatcher SM, Rhodes SM, Stewart JR, Silbergeld E, Pisanic N, Larsen J, Jiang S, Krosche A,
Hall D, Carroll KC, Heaney CD. The prevalence of antibiotic-resistant Staphylococcus aureus
nasal carriage among industrial hog operation workers, community residents, and children
living in their households: North Carolina, USA. Environ Health Perspect. 2017;125:560–9.
69. Heller J, Innocent G, Denwood M, Reid S, Kelly L, Mellor D. Assessing the probability of
acquisition of meticillin-resistant Staphylococcus aureus (MRSA) in a dog using a nested
stochastic simulation model and logistic regression sensitivity analysis. Prev Vet Med.
2011;99:211–24.
70. Heller J, Kelly L, Reid SW, Mellor DJ. Qualitative risk assessment of the acquisition of
Meticillin-resistant Staphylococcus aureus in pet dogs. Risk Anal. 2010;30:458–72.
71. Herold BC, lmmergluck LC, Maranan MC, Lauderdale DS, Gaskin RE, BoyleVavra S, Leitch
CD, Daum RS. Community-acquired methicillin-resistant Staphylococcus aureus in children
with no identified predisposing risk. JAMA. 1998;279:593–8.
72. Hill PC, Birch M, Chambers S, Drinkovic D, Ellis-Pegler RB, Everts R, Murdoch D,
Pottumarthy S, Roberts SA, Swager C, Taylor SL, Thomas MG, Wong CG, Morris
AJ. Prospective study of 424 cases of Staphylococcus aureus bacteraemia: determination of
factors affecting incidence and mortality. Intern Med J. 2001;31:97–103.
73. Hiramatsu K, Hanaki H, Ino T, Yabuta K, Oguri T, Tenover FC. Methicillinresistant
Staphylococcus aureus clinical strain with reduced vancomycin susceptibility. J Antimicrob
Chemother. 1997;40:135–6.
74. Hota B, Ellenbogen C, Hayden MK, Aroutcheva A, Rice TW, Weinstein RA. Community
associated methicillin-resistant Staphylococcus aureus skin and soft tissue infections at a
public hospital: do public housing and incarceration amplify transmission? Arch Int Med.
2007;167:1026–33.
75. Huang SS, Septimus E, Kleinman K, Moody J, Hickok J, Avery TR, Lankiewicz J, Bombosev
A, Terpstra L, Hartford F, Hayden MK, Jernigan JJ, Weinstein RA, Fraser VJ, Haffenreffer K,
Cui E, Kaganov RE, Lolans K, Perline JB, Platt R. Targeted versus universal decolonization
to prevent ICU infection. N Engl J Med. 2013;368:2255–65.
76. Humphreys H, Fitzpatrick F, Harvey BJ. Gender differences in rates of carriage and blood-
stream infection caused by methicillin-resistant Staphylococcus aureus: are they real, do they
matter and why? Clin Infect Dis. 2015;61:1708–14.
77. Hwang JH, Tsai HY, Liu TC. Community-acquired methicillin-resistant Staphylococcus
aureus infections in discharging ears. Acta Otolaryngol. 2002;122:827–30.
78. Iwamoto M, Mu Y, Lynfield R, Bulens SN, Nadle J, Aragon D, Petit S, Ray SM, Harrison LH,
Dumyati G, Townes JM, Schaffner W, Gorwitz RJ, Lessa FC. Trends in invasive methicillin-
resistant Staphylococcus aureus infections. Pediatrics. 2013;132:e817.
64 L. P. Gleason et al.
79. Jessen O, Rosendal K, Bulow P, Faber V, Eriksen KR. Changing staphylococci and staphy-
lococcal infections. A ten-year study of bacteria and cases of bacteremia. N Engl J Med.
1969;281:627–35.
80. Jevons MP. “Celbenin”-resistance staphylococci. BMJ. 1961;1:124.
81. Kallen AJ, Driscoll TJ, Thornton S, Olson PE, Wallace MR. Increase in community-acquired
methicillin-resistant Staphylococcus aureus at a Naval Medical Center. Infec Control Hosp
Epidemiol. 2000;21:223–6.
82. Kaplan SL. Implications of methicillin-resistant Staphylococcus aureus as a communityac-
quired pathogen in pediatric patients. lnfect Dis Clin North Am. 2005;19:747–57.
83. Kaplan SL, Forbes A, Hammerman WA, Lamberth L, Hulten KG, Minard CG, Mason
EO. Randomized trial of “bleach baths” plus routine hygienic measures vs routine hygienic
measures alone for prevention of recurrent infections. Clin Infect Dis. 2013;58:679–82.
84. Katayama Y, Ito T, Hiramatsu K. A new class of genetic element, staphylococcus cassette
chromosome mec, encodes methicillin-resistance in Staphylococcus aureus. Antimicrob
Agents Chemother. 2000;44:1549–55.
85. Kazakova SV, Hageman JC, Matava M, Srinivasan A, Phelan L, Garfinkel B, Boo T,
McAllister S, Anderson J, Jensen B, Dodson D, Lonsway D, McDougal LK, Arduino M,
Fraser VJ, Killgore G, Tenover FC, Cody S, Jernigan DB. A clone of methicillin-resistant
Staphylococcus aureus among professional football players. N Engl J Med. 2005;352:468–75.
86. Kazmierczak Z, Gorski A, Dabrowska K. Facing antibiotic resistance: Staphylococcus aureus
phages as a medical tool. Virus. 2014;6:2551–70.
87. Kirby WMM. Extraction of a highly potent penicillin inactivator from penicillin resistant
staphylococci. Science. 1944;99:452–3.
88. Klevens RM, Morrison MA, Nadle J, Petit S, Gershman K, Ray S, Harrison LH, Lynfield
R, Dumyati G, Townes JM, Craig AS, Zell ER, Fosheim GE, McDougal LK, Carey RB,
Fridkin SK. Invasive methicillin-resistant Staphylococcus aureus infections in the United
States. JAMA. 2007;298:1763–71.
89. Köser CU, Holden MT, Ellington MJ, Cartwright EJ, Brown NM, Ogilvy-Stuart AL, Hsu
LY, Chewapreecha C, Croucher NJ, Harris SR, Sanders M, Enright MC, Dougan G, Bentley
SD, Parkhill J, Fraser LJ, Betley JR, Schulz-Trieglaff OB, Smith GP, Peacock SJ. Rapid
whole-genome sequencing for investigation of a neonatal MRSA outbreak. N Engl J Med.
2012;366:2267–75.
90. Laupland KB, Conly JM. Treatment of Staphylococcus aureus colonization and prophylaxis
for infection with topical intranasal mupirocin: an evidence-based review. Clin lnfect Dis.
2003;37:933–8.
91. Lee NE, Taylor MM, Bancroft E, Ruane PJ, Morgan M, McCoy L, Simon PA. Risk factors
for community-associated methicillin-resistant Staphylococcus aureus skin infections among
HIV-positive men who have sex with men. Clin Infect Dis. 2005;40:1529–34.
92. Lessa FC, Mu Y, Ray SM, Dumyati G, Bulens S, Gorwitz RJ, Fosheim G, AS DV, Schaffner
W, Nadle J, Gershman K, Fridkin SK, Active Bacterial Core surveillance (ABCs); MRSA
Investigators of the Emerging Infections Program. Impact of USA300 methicillin-resistant
Staphylococcus aureus on clinical outcomes of patients with pneumonia or central line asso-
ciated bloodstream infections. Clin Infect Dis. 2012;55:232–41.
93. Levine DP, Cushing RD, Jui J, Brown WJ. Community-acquired methicillin-resistant
Staphylococcus aureus endocarditis in the Detroit Medical Center. Ann Intern Med.
1982;97:330–8.
94. Limbago B, Fosheim GE, Schoonover V, Crane CE, Nadle J, Petit S, Heltzel D, Ray SM,
Harrison LH, Lynfield R, Dumyati G, Townes JM, Schaffner W, Mu Y, Fridkin SK, Active
Bacterial Core Surveillance MRSA Investigators. Characterization of methicillin-resistant
Staphylococcus aureus isolates collected in 2005 and 2006 from patients with invasive dis-
ease: a population-based analysis. J Clin Microbiol. 2009;47:1344–51.
95. Limbago BM, Kallen AJ, Zhu W, Eggers P, McDougal LK, Albrecht VS. Report of the
13th vancomycin-resistant Staphylococcus aureus isolates from the United States. J Clin
Microbiol. 2014;52:998–1002.
3 Emergence of MRSA in the Community 65
96. Lina G, Piemont Y, Godail-Gamot F, Bes M, Peter MO, Gauduchon V, Vandenesch F, Etienne
J. Involvement of Panton-Valentine leukocidin-producing Staphylococcus aureus in primary
skin infections and pneumonia. Clin lnfect Dis. 1999;29:1128–32.
97. Liu C, Bayer A, Cosgrove SE, Daum RS, Fridkin SK, Gorwitz RJ, Kaplan SL, Karchmer
AW, Levine DP, Murray BE, J Rybak M, Talan DA, Chambers HF. Clinical practice guide-
lines by the infectious diseases society of America for the treatment of methicillin-resistant
Staphylococcus aureus infections in adults and children. Clin Infect Dis. 2011;52:e18–55.
98. Loeb M, Main C, Walker-Dilks C, Eady A. Antimicrobial drugs for treating methicillin-resistant
Staphylococcus aureus colonization. Cochrane Database Syst Rev. 2003;(4):CD003340.
99. Mammina C, Calà C, Plano MR, Bonura C, Vella A, Monastero R, Palma DM. Ventilator-
associated pneumonia and MRSA ST398, Italy. Emerg Infect Dis. 2010;16:730–1.
100. Mancini F, Monaci E, Lofano G, Torre A, Bacconi M, Tavarini S, Sammicheli C, Arcidiacono
L, Galletti B, Laera D, Pallaoro M, Tuscano G, Fontana MR, Bensi G, Grandi G, Rossi-
Paccani S, Nuti S, Rappuoli R, De Gregorio E, Bagnoli F, Soldaini E, Bertholet S. One dose
of Staphylococcus aureus 4C-staph vaccine formulated with a novel TLR7-dependent adju-
vant rapidly protects mice through antibodies, effector CD4+ T cells, and IL-17A. PLoS One.
2016;11:e0147767.
101. Manian FA. Asymptomatic nasal carriage of mupirocin-resistant, methicillin-resistant
Staphylococcus aureus (MRSA) in a pet dog associated with MRSA infection in household
contacts. Clin Infect Dis. 2003;36:e26–8.
102. Martinez-Aguilar G, Avalos-Mishaan A, Hulten K, Hammerman W, Mason EO Jr, Kaplan
SL. Community-acquired, methicillin-resistant and methicillin-susceptible Staphylococcus
aureus musculoskeletal infections in children. Pediatr lnfect Dis J. 2004;23:701–6.
103. McDougal LK, Steward CD, Killgore GE, Chaitram JM, McAllister SK, Tenover FC. Pulsed-
field gel electrophoresis typing of oxacillin-resistant Staphylococcus aureus isolates from the
United States: establishing a national database. J Clin Microbiol. 2003;41:5113–20.
104. Miller LG, Perdreau-Remington F, Rieg G, Mehdi S, Perlroth J, Bayer AS, Tang AW, Phung
TO, Spellberg B. Necrotizing fasciitis caused by communityassociated methicillin-resistant
Staphylococcus aureus in Los Angeles. N Engl J Med. 2005;352:1445–53.
105. Miller LG, Eells SJ, Taylor AR, David MZ, Ortiz N, Zychowski D, Kumar N, Cruz D,
Boyle-Vavra S, Daum RS. Staphylococcus aureus Colonization among household contacts
of patients with skin infections: risk factors, strain discordance, and complex ecology. Clin
Infect Dis. 2012;54:1523–35.
106. Miller MB, Weber DJ, Goodrich JS, Popowitch EB, Poe MD, Nguyen V, Shope TR,
Foster DT, Miller JR, Kotch J. Prevalence and risk factor analysis for methicillin-resistant
Staphylococcus aureus nasal colonization in children attending child care centers. J Clin
Microbiol. 2011;49:1041–7.
107. Moran GJ, Amii RN, Abrahamian FM, Talan DA. Methicillin-resistant Staphylococcus
aureus in community-acquired skin infections. Emerg Infect Dis. 2005;11:928–30.
108. Moran GJ, Krishnadasan A, Gorwitz RJ, Fosheim GE, Albrecht V, Limbago B, Talan DA,
EMERGEncy ID NET Study Group. Prevalence of methicillin-resistant Staphylococcus aureus
as an etiology of community-acquired pneumonia. Clin Infect Dis. 2012;54(8):1126–33.
109. Moran GJ, Krishnadasan A, Mower WR, Abrahamian FM, LoVecchio F, Steele MT, Rothman
RE, Karras DJ, Hoagland R, Pettibone S, Talan DA. Effect of cephalexin plus trimethoprim-
sulfamethoxazole vs. cephalexin alone on clinical cure of uncomplicated cellulitis: a random-
ized clinical trial. JAMA. 2017;317:2088–96.
110. Morgan DJ, Murthy R, Munoz-Price LS, Barnden M, Camins BC, Johnston BL, Rubin Z,
Sullivan KV, Shane AL, Dellinger EP, Rupp ME, Bearman G. Reconsidering contact pre-
cautions for endemic methicillin-resistant Staphylococcus aureus and vancomycin-resistant
Enterococcus. Infect Control Hosp Epidemiol. 2015;36:1163–72.
111. Morris DO, Lautenbach E, Zaoutis T, Leckerman K, Edelstein PH, Rankin SC. Potential
for pet animals to harbour methicillin-resistant Staphylococcus aureus when residing with
human MRSA patients. Zoonoses Public Health. 2012;59:286–93.
66 L. P. Gleason et al.
112. Naimi TS, Ledell KH, Boxrud DJ, Groom AV, Steward CD, Johnson SK, Besser JM, O'Boyle
C, Danila RN, Cheek JE, Osterholm MT, Moore KA, Smith KE. Epidemiology and clonal-
ity of community-acquired methicillin-resistant Staphylococcus aureus in Minnesota, 1996-
1998. Clin Infect Dis. 2001;33:990–6.
113. Nair R, Ammann E, Rysavy M, Schweizer ML. Mortality among patients with methicillin-
resistant Staphylococcus aureus USA300 versus non-USA300 invasive infections: a meta-
analysis. Infect Control Hosp Epidemiol. 2014;35:31–41.
114. Nerby JM, Gorwitz R, Lesher L, Juni B, Jawahir S, Lynfield R, Harriman K. Risk factors
for household transmission of community associated methicillin-resistant Staphylococcus
aureus. Pediatr Infect Dis J. 2011;30:927–32.
115. Neyra RC, Frisancho JA, Rinsky JL, Resnick C, Carroll KC, Rule AM, Ross T, You Y, Price
LB, Silbergeld EK. Multidrug-resistant and methicillin-resistant Staphylococcus aureus
(MRSA) in hog slaughter and processing plant workers and their community in North
Carolina (USA). Environ Health Perspect. 2014;122:471–7.
116. Ng W, Faheem A, McGeer A, Simor AE, Gelosia A, Willey BM, Watt C, Richardson
DC, Wong H, Ostrowska K, Vernich L, Muller MP, Gnanasuntharam P, Porter V, Katz
K. Community and healthcare associated methicillin-resistant Staphylococcus aureus strains:
an investigation into household transmission, risk factors, and environmental contamination.
Infect Control Hosp Epidemiol. 2017;38:61–7.
117. Nguyen DM, Mascola L, Brancoft E. Recurring methicillin-resistant Staphylococcus aureus
infections in a football team. Emerg Infect Dis. 2005;11:526–32.
118. Nimmo GR. USA300 abroad: global spread of a virulent strain of community associated
methicillin-resistant Staphylococcus aureus. Clin Microbiol Infect. 2012;18:725–34.
119. Ogston A. Micrococcus poisoning. J Anat. 1882;17:24–58.
120. O’Hara FP, Suaya JA, Ray GT, Baxter R, Brown ML, Mera RM, Close NM, Thomas E,
Amrine-Madsen H. Spa typing and multilocus sequence typing show comparable perfor-
mance in a macroepidemiologic study of Staphylococcus aureus in the United States. Microb
Drug Resist. 2016;22:88–96.
121. Panton PN, Valentine FC. Staphylococcal toxin. Lancet. 1932;222:506–8.
122. Perl TM, Cullen JJ, Wenzel RP, Zimmerman MB, Pfaller MA, Sheppard D, Twombley J,
French PP, Herwaldt LA. Intranasal mupirocin to prevent postoperative Staphylococcus
aureus infections. N Engl J Med. 2002;346:1871–7.
123. Planet PJ, LaRussa SJ, Dana A, Smith H, Xu A, Ryan C, Uhlemann AC, Boundy S, Goldberg J,
Narechania A, Kulkarni R, Ratner AJ, Geoghegan JA, Kolokotronis SO, Prince A. Emergence
of the epidemic of methicillin-resistant Staphylococcus aureus strain USA300 coincides with
horizontal transfer of the arginine catabolic mobile element and speG-mediated adaptations
for survival on skin. mBio. 2013;4:e00889–13.
124. Planet PJ, Diaz L, Kolokotronis S, Narechania A, Reyes J, Xing G, Rincon S, Smith H,
Panesso D, Ryan C, Smith DP, Guzman M, Zurita J, Sebra R, Deikus G, Nolan RL, Tenover
FC, Weinstock GM, Robinson DA, Arias CA. Parallel epidemics of community associated
methicillin-resistant Staphylococcus aureus USA300 infection in North and South America.
J Infect Dis. 2015;212:1874–82.
125. Podewils LJ, Liedtke LA, McDonald LC, Hageman JC, Strausbaugh LJ, Fischer TK, Jernigan
DB, Uyeki TM, Kuehnert MJ. A national survey of severe influenza-associated complications
among children and adults, 2003-2004. Clin Infect Dis. 2005;40:1693–6.
126. Popovich KJ, Hota B, Aroutcheva A, Kurien L, Patel J, Lyles-Banks R, Grasso AE, Spec
A, Beavis KG, Hayden MK, Weinstein RA. Community associated methicillin-resistant
Staphylococcus aureus colonization burden and HIV-infected patients. Clin Infect Dis.
2013;56:1067–74.
127. Popovich KJ, Snitkin ES, Hota B, Green SJ, Pirani A, Aroutcheva A, Weinstein RA. Genomic
and epidemiological evidence for community origins of hospital-onset methicillin-resistant
Staphylococcus aureus bloodstream infections. J Infect Dis. 2017;215:1640–7.
128. Raczniak GA, Gaines J, Bulkow LR, Kinzer MH, Hennessy TW, Klejka JA, Bruce MG. A
survey of knowledge, attitudes, and practices towards skin and soft-tissue infections in rural
Alaska. Int J Circumpolar Health. 2016;75:30603.
3 Emergence of MRSA in the Community 67
129. Ray GT, Suaya JA, Baxter R. Incidence, microbiology, and patient characteristics of skin and
soft-tissue infections in a US population: a retrospective population-based study. BMC Infect
Dis. 2013;13:252.
130. Rhee Y, Aroutcheva A, Hota B, Weinstein RA, Popovich KJ. Evolving epidemiology of
Staphylococcus aureus bacteremia. Infect Control Hosp Epidemiol. 2015;36:1417–22.
131. Rolo J, Miragaia M, Turlej-Rogacka A, Empel J, Bouchami O, Faria NA, Tavares A,
Hryniewicz W, Fluit AC, de Lencastre H, The CONCORD Working Group. High genetic
diversity among community associated Staphylococcus aureus in Europe: results from a mul-
ticenter study. PLoS One. 2012;7:e34768.
132. Ross S, Rodriguez W, Controni G, Khan W. Staphylococcal susceptibility to penicillin G. The
changing pattern among community strains. JAMA. 1974;229:1075–7.
133. Rossi F, Diaz L, Wollam A, Panesso D, Zhou Y, Rincon S, Narechania A, Xing G, Di Gioia
TS, Doi A, Tran TT, Reyes J, Munita JM, Carvajal LP, Hernandez-Roldan A, Brandão D, van
der Heijden IM, Murray BE, Planet PJ, Weinstock GM, Arias CA. Transferable vancomycin
resistance in a community associated MRSA lineage. N Engl J Med. 2014;370:1524–31.
134. Santos F, Mankarious LA, Eavey RD. Methicillin-resistant Staphylococcus aureus: pediatric
otitis. Arch Otolaryngol Head Neck Surg. 2000;126:1383–5.
135. Saravolatz LD, Markowitz N, Arking L, Pohlod D, Fisher E. Methicillin-resistant
Staphylococcus aureus. Epidemiologic observations during a community-acquired outbreak.
Ann lntern Med. 1982;96:11–6.
136. Sattler CA, Mason EO Jr, Kaplan SL. Prospective comparison of risk factors and demographic
and clinical characteristics of community-acquired, methicillin-resistant versus methicillin-
susceptible Staphylococcus aureus infection in children. Pediatr lnfect Dis J. 2002;21:910–7.
137. Schinasi L, Wing S, Augustino KL, Ramsey KM, Nobles DL, Richardson DB, Price LB,
Aziz M, PD MD, Stewart JR. A case control study of environmental and occupational expo-
sures associated with methicillin resistant Staphylococcus aureus nasal carriage in patients
admitted to a rural tertiary care hospital in a high density swine region. Environ Health.
2014;13:54.
138. Schmidt CS, White CJ, Ibrahim AS, Filler SG, Yeaman MR, Edwards JE Jr, Hennessey JP Jr.
NDV-3, a recombinant alum-adjuvanted vaccine for Candida and Staphylococcus aureus, is
safe and immunogenic in healthy adults. Vaccine. 2012;30:7594–600.
139. Schmitz GR, Bruner D, Pitotti R, Olderog C, Livengood T, Williams J, Huebner K,
Lightfoot J, Ritz B, Bates C, Schmitz M, Mete M, Deye G. Randomized controlled trial
of trimethoprim-sulfamethoxazole for uncomplicated skin abscesses in patients at risk for
community-associated methicillin-resistant Staphylococcus aureus infection. Ann Emerg
Med. 2010;56:283–7.
140. See I, Albrecht V, Mu Y, Dumyati G, Koeck M, Lynfield R, Nadle J, Ray SM, Schaffner
W, Limbago B, Kallen A. Changes in incidence and strains of methicillin-resistant
Staphylococcus aureus bloodstream infections, 2005–2013. IDweek 2016 conference, New
Orleans, LA. Presented October 29, 2016. Abstract #1750; 2016.
141. See I, Wesson P, Gualandi N, Dumyati G, Harrison LH, Lesher L, Nadle J, Petit S, Reisenauer
C, Schaffner W, Tunali A, Mu Y, Ahern J. Socioeconomic factors explain racial disparities
in invasive community-associated methicillin-resistant Staphylococcus aureus disease rates.
Clin Infect Dis. 2017;64:597–604.
142. Sehulster L, Chinn RY. Guidelines for environmental infection control in health-care facili-
ties. Recommendations of CDC and the healthcare infection control practices advisory com-
mittee (HICPAC). MMWR Recomm Rep. 2003;52:1–42.
143. Self WH, Wunderink RG, Williams DJ, Zhu Y, Anderson EJ, Balk RA, Fakhran SS, Chappell
JD, Casimir G, Courtney DM, Trabue C, Waterer GW, Bramley A, Magill S, Jain S, Edwards
KM, Grijalva CG. Staphylococcus aureus Community-acquired pneumonia: prevalence,
clinical characteristics, and outcomes. Clin Infect Dis. 2016;63(3):300–9.
144. Shallcross LJ, Fragaszy E, Johnson AM, Hayward AC. The role of the Panton-Valentine leu-
cocidin toxin in staphylococcal disease: a systematic review and meta-analysis. Lancet Infec
Dis. 2013;13:43–54.
68 L. P. Gleason et al.
145. Siberry GK, Tekle T, Carroll K, Dick J. Failure of clindamycin treatment of methicillin-
resistant Staphylococcus aureus expressing inducible clindamycin resistance in vitro. Clin
Infect Dis. 2003;37:1257–60.
146. Siegel JD, Rhinehart E, Jackson M, Chiarello L, Healthcare Infection Control Practices
Advisory Committee. Management of multidrug-resistant organisms in healthcare settings,
2006. Available at: https://www.cdc.gov/infectioncontrol/guidelines/mdro/. 2006.
147. Siegel JD, Rhinehart E, Jackson M, Chiarello L, Healthcare Infection Control Practices
Advisory Committee. Guideline for isolation precautions: preventing transmission of infec-
tious agents in healthcare settings 2007. Available at http://www.cdc.gov/ncidod/dhqp/pdf/
guidelines/Isolation2007.pdf. 2007.
148. Singer AJ, Talan DA. Management of Skin Abscesses in the era of methicillin-resistant
Staphylococcus aureus. N Engl J Med. 2014;370:1039–47.
149. Stacey AR, Endersby KE, Chan PC, Marples RR. An outbreak of methicillin resistant
Staphylococcus aureus infection in a rugby football team. Br J Sports Med. 1998;32:153–4.
150. Stevens DL, Bisno AL, Chambers HF, Everett ED, Dellinger P, Goldstein EJ, Gorbach SL,
Hirschmann JV, Kaplan EL, Montoya JG, Wade JC. Practice guidelines for the diagnosis and
management of skin and soft-tissue infections. Clin Infect Dis. 2005;41:1373–406.
151. Talan DA, Krishnadasan A, Gorwitz RJ, Fosheim GE, Limbago B, Albrecht V, Moran GJ,
EMERGEncy ID Net Study Group. Comparison of Staphylococcus aureus from skin and
soft-tissue infections in US emergency department patients, 2004 and 2008. Clin Infect Dis.
2011;15:144–9.
152. Talan DA, Mower WR, Krishnadasan A, Abrahamian FM, Lovecchio F, Karras DJ, Steele
MT, Rothman RE, Hoagland R, Moran GJ. Trimethoprim–Sulfamethoxazole versus placebo
for uncomplicated skin abscess. N Engl J Med. 2016;374:823–32.
153. Toleman MS, Reuter S, Coll F, Harrison EM, Blane B, Brown NM, Török ME, Parkhill
J, Peacock SJ. Systematic surveillance detects multiple silent introductions and household
transmission of methicillin-resistant Staphylococcus aureus USA300 in the east of England.
J Infect Dis. 2016;214:447–53.
154. Tong SYC, Nelson J, Paterson DL, Fowler VG, Howden BP, Cheng AC, Chatfield M, Lipman
J, Van Hal S, O’Sullivan M, Robinson JO, Yahav D, Lye D, Davis JS, CAMERA2 study
group and the Australasian Society for Infectious Diseases Clinical Research Network.
CAMERA2 – combination antibiotic therapy for methicillin-resistant Staphylococcus aureus
infection: study protocol for a randomized controlled trial. Trials. 2016;17:170.
155. Uhlemann AC, Dordel J, Knox JR, Raven KE, Parkhill J, Holden MTG, Peacock SJ, Lowy
FD. Molecular tracing of the emergence, diversification, and the transmission of S. aureus
sequence type 8 in a New York community. Proc Natl Acad Sci. 2014;111:6739–43.
156. Uhlemann AC, Otto M, Lowy FD, DeLeo FR. Evolution of community- and healthcare-
associated methicillin-resistant Staphylococcus aureus. Infect Genet Evol. 2014;21:563–74.
157. van Cleef BA, Verkade EJ, Wulf MW, Buiting AG, Voss A, Huijsdens XW, van Pelt W,
Mulders MN, Kluytmans JA. Prevalence of livestock-associated MRSA in communities with
high pig-densities in the Netherlands. PLoS One. 2010;5:e9385.
158. VandenBergh MF, Yzerman EP, Van Belkum A, Boelens HA, Sijmons M, Verbrugh
HA. Follow-up of Staphylococcus aureus nasal carriage after 8 years: redefining the persis-
tent carrier state. J Clin Microbiol. 1999;37:3133–40.
159. Vayalumkal JV, Suh KN, Toye B, Ramotar K, Saginur R, Roth VR. Skin and soft tissue infec-
tions caused by methicillin-resistant Staphylococcus aureus (MRSA): an affliction of the
underclass. CJEM. 2012;14:335–43.
160. Vetter RS, Bush SP. Reports of presumptive brown recluse spider bites reinforce improb-
able diagnosis in regions of North America where the spider is not endemic. Clin Infect Dis.
2002;35:442–5.
161. von Eiff C, Becker K, Machka K, Stammer H, Peters G. Nasal carriage as a source of
Staphylococcus aureus bacteremia. Study Group. N Engl J Med. 2001;344:11–6.
3 Emergence of MRSA in the Community 69
Much has transpired in the realm of antibiotic resistance in the 10 years since the first
edition of this text. This chapter began in 2007 with the line, “At the beginning of the
twenty first century, we now find ourselves experiencing a taste of what life was like
prior to the advent of the antibiotic age…,” and this reality is continuing to sink in. So
much so, in fact, that antibiotic resistance is now routinely broached in the popular
media and has the attention of government agencies and philanthropic groups and to
some extent may be prompting a return to antibiotic discovery within the pharmaceuti-
cal industry. In 2009, the Infectious Diseases Society of America (IDSA) released the
updated call to action for a coordinated effort to bring antibiotic development to the
forefront, specifically regarding the “ESKAPE” pathogens, Enterococcus faecium,
Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii,
Pseudomonas aeruginosa, and Enterobacter spp. [1]. These pathogens cause the major-
ity of US hospital infections, and resistance is a major issue. The Gram-negative bacilli
are well represented in this group and pose a very significant emerging problem, particu-
larly in the case of pan-antibiotic-resistant A. baumannii, multidrug-resistant (MDR) P.
aeruginosa, and carbapenem-resistant Enterobacteriaceae (CRE). More recently the
IDSA has begun “the 10x20 Initiative” (http://www.idsociety.org/10x20/). In February
of 2017, the World Health Organization established its priority list for drug-resistant
pathogens, and in the “critical” category are carbapenem-resistant A. baumannii and P.
aeruginosa and carbapenem-resistant, extended-spectrum β-lactamase (ESBL)-
producing Enterobacteriaceae (http://www.who.int/mediacentre/news/releases/2017/
bacteria-antibiotics-needed/en/). The notion of tackling antimicrobial resistance was
also addressed by the economist Jim O’Neill, who articulated the human costs and eco-
nomic and security threats associated with a failure to act (https://amr-review.org/sites/
default/files/160518_Final%20paper_with%20cover.pdf). Discussions about how to
incentivize antibiotic discovery have followed and the establishment of research funding
through agencies such as the Biomedical Advanced Research and Development
Authority (BARDA, https://www.phe.gov/about/BARDA/Pages/default.aspx) and
Combating Antibiotic-Resistant Bacteria Biopharmaceutical Accelerator (CARB-X,
http://www.carb-x.org); Wellcome Trust and Pew Charitable Trust have also come
online to promote antibiotic discovery. These activities, although encouraging, still
underscore the challenge upon us. Therefore, an understanding of resistance in Gram-
negative pathogens is informative and will be discussed in the following sections.
This chapter addresses intrinsic resistance and mutationally or horizontally
acquired resistance mechanisms. Intrinsic (or innate) resistance varies widely
among different pathogens and is determined by the general makeup of a cell, where
the overall complement of genes and their expression levels establish a baseline
susceptibility to an antibacterial. Here we focus on two broad elements important
for intrinsic resistance, the impermeability of the Gram-negative cell envelope,
which impedes a compound’s entry into the cell to exert its effect, and energy-
dependent active efflux which extrudes a compound back out of the cell before it
can engage its target. We begin there, since (i) these can be important hurdles to
overcome in efforts to discover new antibiotics for Gram-negative pathogens and
(ii) they can facilitate/exacerbate the emergence of mutationally or horizontally
acquired resistance. The organism-specific genetic blueprint for intrinsic resistance
provides the background within which mutations can be selected that further
decrease susceptibility to antibacterial compounds. As well, the horizontal acquisi-
tion of new genetic material is an important route of acquired resistance. The pro-
gression to resistance is often multifactorial, and several acquired mechanisms can
accumulate over time to cause clinically significant resistance and multidrug resis-
tance (MDR). In that regard, the meaning of “resistance” is context-specific. In
clinical antimicrobial susceptibility testing, resistance is based on a specific mini-
mal inhibitory concentration (MIC) of an antibiotic tested under standardized con-
ditions, and clinical resistance occurs if the MIC of the antibiotic is above an
established clinical resistance “breakpoint” [2]. Here, we use the term more gener-
ally to convey the idea that the mechanisms discussed will alter (increase) the level
of resistance (or decrease susceptibility), but not all resistance mechanisms will
cause the specifically defined clinical resistance (shift over the breakpoint). The first
edition of this chapter pertained mainly to antibiotics that had been in clinical use
for some time (e.g., fluoroquinolones). These sections are updated here, but two
additional aspects are now included. In 2007, tigecycline was just entering the
clinic, and we update on what has happened in the approximately 12 years it has
been in widespread clinical use (Sect. 4.2.4). Second, polymyxins were reintro-
duced into the clinic as a last line of defense against MDR Gram-negative p athogens.
In a relatively short time, resistance has emerged and has begun to erode the clinical
utility of these compounds, and this is discussed in Sect. 4.2.6.
4 Resistance of Gram-negative Bacilli to Antimicrobials 73
Bacteria have a broad range of efflux pumps that can actively extrude molecules
from the cell. Efflux pumps can serve natural physiological roles such as extrusion
of metabolites but also function in efflux of toxic molecules that enter the cells.
Efflux of toxic molecules serves to lower their intracellular accumulation to reduce
access to the intracellular target(s), thereby protecting the bacteria. The five broad
efflux pump superfamilies most important in bacteria are the ATP-binding cassette
(ABC) family, the major facilitator superfamily (MFS), the small multidrug resis-
tance (SMR) family, the multidrug and toxic compound extrusion (MATE) family,
and the resistance-nodulation-cell division (RND) family (reviewed in [3] (Fig. 4.1)).
The ABC family differs from the other families in that they derive energy to drive
active efflux from hydrolysis of ATP, whereas the other families derive energy from
the proton gradient maintained at the bacterial cytoplasmic membrane. The pump
proteins that mediate compound recognition and energy-dependent extrusion for all
families are situated in the bacterial cytoplasmic membrane. Members of all pump
families except RND pumps are found in both Gram-positive and Gram-negative
bacteria. The RND family pumps are unique to Gram-negative bacteria and have
additional components and an overall architecture necessary for efflux across the
Gram-negative outer membrane (OM) (Fig. 4.1). Depending on the context, all of
the families can contribute to resistance in Gram-negative pathogens, but non-RND
family pumps can only efflux compounds into the periplasmic space between the
cytoplasmic and OM but not to the outside of the cell. Furthermore, RND pumps are
Fig. 4.1 General architecture of efflux pump families and placement in the Gram-negative cell
envelope. The Gram-negative envelope has two membranes (the inner membrane, shown here as a
symmetrical bilayer in blue, and an outer membrane, which is asymmetrical) and has phospholipid
(blue) at the inner leaflet and lipopolysaccharide (gold) at the outer leaflet. RND family pumps
have an architecture that spans both membranes. Some compounds can enter the cells via water-
filled porins (yellow)
74 C. R. Dean et al.
notable for their large amorphous compound-binding pockets [4, 5] which confer
the ability to recognize and extrude a very broad range of structurally unrelated
molecules. For these reasons, RND family pumps are regarded as the most signifi-
cant efflux pumps overall in Gram-negative bacteria in terms of antibacterial resis-
tance. However, it is also clear that there is cooperation between networks of pumps
of different families when their substrates overlap [6]. In those cases, the single-
component pumps may efflux a substrate into the periplasm, and the RND pump
may then expel the compound from the periplasm to the outside of the cell. RND
pumps are tripartite structures, comprised of the inner membrane-located RND
pump component, an OM channel component (outer membrane factor (OMF)), and
a periplasmic membrane fusion protein (MFP) that links these components
(Fig. 4.1). This architecture spans the double membrane of the Gram-negative cell
to allow compound extrusion across the OM through the OMF, driven by the proton-
motive force (PMF) at the inner membrane. RND pumps are typically named in the
order MFP-pump-OMF, and the best studied RND pumps are AcrA-AcrB-TolC
(shortened to AcrAB-TolC) of E. coli and MexAB-OprM of P. aeruginosa. RND
family pumps have been found in all Gram-negative bacteria so far studied, and
most RND pumps have a broad substrate range, allowing them overall to accom-
modate most classes of antibiotics, biocides, dyes, organic solvents, detergents, bile
salts, β-lactamase inhibitors, and other molecules [3]. Moreover, some bacteria pos-
sess several different RND pumps with partially overlapping substrate specificities,
increasing their ability to deal with toxic compounds (Table 4.1). The complement
of efflux pumps in a particular Gram-negative species likely reflects the variability
of its environment. For example, the ubiquitous environmental organism P. aerugi-
nosa has a large and highly regulated genome that encodes 12 different putative
RND family efflux pumps [7, 8], presumably enhancing survival in the presence of
toxic molecules, including natural product antibacterials encountered in the envi-
ronment. In contrast, Haemophilus influenzae, which is adapted mainly to the
human respiratory tract, has only one RND pump.
Table 4.1 Example RND efflux pumps in Gram-negative pathogens and range of antibiotics
accommodated by each pump
Pump component
Organism MFP RND OMF Antibiotics pumped
A. baumannii AdeA AdeB AdeC AG, CM, FQ, TC (MC), TG
AdeIa AdeJa AdeKa BL, CM, EM, FQ, TC (MC), TG
AdeF AdeG AdeH FQ, TG
B. cepacia CeoA CeoB OpcM CM, FQ, TM
E. coli AcrAa AcrBa TolCa BL, CM, FQ, ML, NO, RF
AcrA AcrD Tolc AG, FU, NO
acrE Acrf TolC FQ
H. influenzae AcrAa AcrBa TolCa EM, NO
K. pneumoniae AcrA AcrB TolC BL, CM, EM, FQ, TG
OqxA OqxB TG
KpgA KpgB KpgC TG
P. aeruginosa MexAa MexBa OprMa AG, BL, CM, ML, NO, TC, TG, TM,CM,
CP, FQ, TC
MexC MexD OprJ CM, FQ
MexE MexF OprN EM, TC
MexJ MexK OprM/ CM, EM, FQ, TC
OprH
MexV MexW OprM AG, ML, TC, TG
MexM MexN OprM BL
MexX MexY OprM AG, ML, TC, TG
S. enterica serovar AcrAa AcrBa TolCa BL, CM, EM, FQ, NO, RF, TC
Typhimurium
S. maltophilia SmeA SmeB SmeC AG, BL, FQ
SmeDa SmeEa SmeFa EM, FQ, TC (MC), TG
Table 4.1 summarizes data extracted from Li et al. [3]. For additional pumps and details regarding
substrate ranges and pump regulation, consult this very comprehensive review
MFP membrane fusion protein, RND resistance-nodulation-division pump component, OMF outer
membrane factor
Antibiotics: AG aminoglycosides, BL β-lactams, CM chloramphenicol, EM erythromycin, FQ
fluoroquinolones, FU fusidic acid, ML macrolides, NO novobiocin, RF rifampicin, TC tetracy-
clines, MC minocycline, TG tigecycline, TM trimethoprim
a
Denotes a pump that is expressed constitutively (housekeeping pump), but regulatory mutations
can further upregulate expression
switch to an open state of TolC [9]. A direct interaction between AcrB and TolC has
also been shown in vitro and also in cells using chemical cross-linking [10], but
other models suggest an alternative mechanism of assembly where AcrB and TolC
do not interact [11]. A very recent study showing in vitro reconstitution of AcrAB-
TolC and MexAB-OprM using nanodisc technology and characterization by single-
particle electron microscopy revealed a structure whereby the pump and OMF were
linked by the MFP but did not directly interact [12]. Whether that structure repre-
sents the final functional pump assembly in a cellular context or if direct interaction
76 C. R. Dean et al.
Fig. 4.2 Rotating functional mechanism of efflux by RND family pumps (represented by AcrAB-
TolC); compounds enter at AcrB access conformation; AcrB undergoes a conformational change
to the binding mode and then to the extrusion mode where the compound is released into the outer
membrane channel. Side view of assembled pump with access and extrusion depicted (top); top
view cross section of functional AcrB rotamers (bottom)
between the pump and OMF is required for function is not currently resolved, but
these observations support the notion that the MFP component itself likely forms
part of the exit duct between the RND pump component and the OMF. Structural
studies with AcrB done by independent groups [13–15] and later simulation studies
[16, 17] revealed that drug efflux occurs by a functional rotation mechanism
(Fig. 4.2). Each of the three protomers of the assembled pump component (e.g.,
AcrB) can exist in one of three states referred to as “access” (or “loose”), “binding”
(or “tight”), and “extrusion” (or “open”) (Fig. 4.2). A complete functional cycle
occurs as follows: a compound enters the access conformation of AcrB from the
periplasm (likely from the outer leaflet of the cytoplasmic membrane) via a trans-
membrane domain called the vestibule. AcrB then changes conformation to the
binding conformation, which opens the large compound-binding pocket to accom-
modate the entry of the compound, and finally AcrB rotates to the open (extrusion)
conformation which releases the compound from the binding pocket into the funnel
region toward the OMF (TolC). As mentioned above, the interaction of the MFP
component with TolC keeps TolC in an open formation allowing compounds to be
expelled outside the cell. Energy for this process is derived from transport of pro-
tons from the periplasm to the cytoplasm, and it is suggested that a proton is released
to the cytoplasm when AcrB transitions from the tight to the open conformation
[14]. Consistent with RND pumps requiring proton-motive force to function, energy
4 Resistance of Gram-negative Bacilli to Antimicrobials 77
decouplers like CCCP inhibit efflux. The location of the vestibule in pump proteins
like AcrAB is such that compounds enter from the periplasmic leaflet of the cyto-
plasmic membrane, thereby suggesting generally that RND pumps recognize com-
pounds as they are entering the cell rather than after they ultimately reach the
cytosol. This is consistent with early observations that certain RND family pumps
reduced susceptibility to β-lactam antibiotics or β-lactamase inhibitors that target
penicillin-binding proteins or β-lactamase enzymes, respectively, which are located
in the periplasm [18–20], and with the reported importance of amino acid residues
in periplasmic loops of the inner membrane pump components in determining sub-
strate recognition [21–23]. As mentioned above, single-component pumps from
other families may also play a possibly underappreciated role in acting coopera-
tively with RND pumps when they have overlapping substrate specificities and the
cellular antibacterial target of a compound is cytosolic [6]. This has been fairly well
established in the case of tetracycline-specific MFS (TetA/C) pumps which specifi-
cally efflux tetracycline into the periplasm where broader specificity RND pumps
that recognize tetracycline, such as MexAB-OprM, can extrude the compound from
the periplasm [24]. It remains to be determined in detail where these multi-pump
interactions are important in terms of clinical resistance. Without the contribution of
an RND pump in this sequential efflux, the compound may accumulate in the peri-
plasm where it may readily diffuse back in across the cytoplasmic membrane. When
effluxed out of the periplasm by the RND pump, it can diffuse away or alternatively
must reenter the cell by again traversing the OM. This raises the concept of com-
pound influx and the role of the Gram-negative OM permeability barrier as it relates
to RND-mediated efflux, which is discussed in the next section.
Fig. 4.3 Chemical structure of lipopolysaccharide from E. coli 0157:H7. The lipid A forms the
outer leaflet of the asymmetrical outer membrane. Acyl chain number and lengths and level of lipid
A and core phosphorylation vary among different Gram-negative bacteria
The requirement for nutrient uptake across the OM bilayer is generally met by
water-filled protein β-barrel channels that span the bilayer, known as porins (see
Fig. 4.1). Porins allow for the passage of small hydrophilic molecules across the
OM, essentially establishing the overall OM as a molecular sieve. These channels
are also thought to allow influx of certain hydrophilic antibiotic molecules that are
small enough to traverse the porin channels [27]. Gram-negatives such as E. coli
have several relatively nonspecific large porins such as OmpC and OmpF with
molecular weight cutoffs of approximately 600 Da [28]. In contrast P. aeruginosa
harbors a number of more specialized or restrictive (smaller) porins to allow influx
of nutrients. This organism also has the general porin OprF, but this exists only
occasionally in the conformation that allows the channel to be open [29]. This
highlights that, along with variability in the lipid bilayer characteristics among
different Gram-negative bacteria, the number and characteristics of the OM porin
channels can also vary, causing large differences in the effectiveness of the OM
permeability barrier. Reflecting this, the OM of P. aeruginosa was estimated to be
more than tenfold less permeable than that of E. coli [30]. The best studied exam-
ples of antibiotic permeation via porins center on various β-lactams. For example,
carbapenems can enter E. coli via OmpC porins, and some carbapenems such as
imipenem enter P. aeruginosa via OprD. In the latter case, the natural function of
4 Resistance of Gram-negative Bacilli to Antimicrobials 79
OprD is transport of basic amino acids, which bear some structural resemblance
to certain carbapenems. Imipenem was also recently shown to enter P. aeruginosa
via OpdP [31], suggesting that some antibiotics may access cells via multiple
porins. Hydrophobic and/or larger compounds not able to enter by porins can
enter the bacterium by diffusion across the membrane bilayer, although these pro-
cesses are likely slower and are not well understood. As well, there are a limited
number of specialized energy-dependent active transporters in the OM that import
scarce nutrients such as iron (as siderophore or protein-bound complexes) [32] or
cobalamin [33].
Overall, the OM of Gram-negative bacteria is highly evolved to provide a strong
protective permeability barrier, but it still allows influx of important nutrients, either
by passive diffusion in the case of porins or active transport in other cases. For most
antibiotics though, the OM slows their influx considerably, and the combination of
reduced antibiotic influx due to the OM permeability barrier and active efflux
together typically determines the levels of susceptibility [34, 35]. RND family
efflux pumps do not always exhibit a high velocity of antibiotic efflux, so beyond
whether the pump recognizes a specific antibiotic, its effectiveness depends to a
large extent on how slowly a given substrate antibiotic is entering the cell. If the
influx rate is too fast, the pump may not “keep up,” and even if the compound is a
substrate, the pump may not confer meaningful resistance. In contrast, if the mem-
brane barrier is slowing influx, efflux pumps can then become a very significant
resistance factor. This was elegantly shown for oxacillin compared to ampicillin in
E. coli. These antibiotics were shown to be very similar as substrates of the AcrB
efflux pump, but oxacillin traversed porins more slowly. Deletion of the acrB gene
had a much larger impact on susceptibility to oxacillin than to ampicillin [36, 37].
Furthermore, disruption of either efflux or the OM permeability barrier in P. aeru-
ginosa strongly increased antibiotic susceptibility, with an even greater increase in
susceptibility when both were disrupted simultaneously, showing the interplay
between these two factors [38]. More recently it was reported that expression of an
OM iron-siderophore transporter that had been engineered to create very large
porin-like channels in the bacterial OM strongly increased susceptibility to a range
of antibacterial compounds, further showing the important role of the OM permea-
bility barrier and its interrelationship with active efflux in several Gram-negative
pathogens [39, 585]. Correspondingly, mutations that impact either the OM perme-
ability barrier or efflux in Gram-negative bacteria can decrease susceptibility to
antibiotics, and this is discussed in the next section.
The combination of the OM permeability barrier and efflux is important for dictat-
ing the spectrum of Gram-negative pathogens that a clinically used or novel antibi-
otic under development will be sufficiently active. In those cases where useful
antibacterial activity does occur (e.g., with currently used antibiotics), mutations
80 C. R. Dean et al.
that increase the expression of efflux pumps, alter substrate recognition, or impact
the OM permeability barrier (decreased compound influx) erode this activity over
time and limit a compound’s therapeutic longevity. Since many RND efflux pumps
have broad substrate ranges, selection of pump upregulation with one compound
will usually affect susceptibility to multiple antibiotics, contributing to multidrug
resistance. These factors are particularly problematic since they contribute to the
therapeutic demise of current antibiotics while also being the major impediment to
the discovery of novel replacement antibiotics. As an example illustrating both of
these points, tigecycline, a glycyl derivative of minocycline that evades the classic
tetracycline-specific resistance mechanisms of ribosomal protection and efflux by
tetracycline-specific single-component efflux pumps (e.g., TetA) (discussed in
detail in Sect. 4.2.4), was still subject to intrinsic RND-mediated efflux in P. aeru-
ginosa [40], and therefore its spectrum does not include this pathogen. Although
tigecycline achieves useful antibacterial activity against other Gram-negatives and
has been successfully implemented clinically, mutations leading to RND pump
upregulation can erode this activity in organisms such as Proteus mirabilis [41], K.
pneumoniae [42], E. coli [43], and A. baumannii [44].
The impact of cell impermeability on the discovery of new anti-Gram-negative
antibiotics is difficult to overstate, as the vast majority of compounds are subject to
some level of efflux and/or limited influx. In one study, the majority of antimicrobial
compounds identified from direct antibacterial screening in E. coli were AcrAB-
TolC pump substrates [45]. In a broader discussion of overall screening efforts con-
ducted at AstraZeneca, the inability of compounds to accumulate in Gram-negative
bacteria was cited as a significant impediment to novel antibiotic discovery using
corporate compound collections [46]. Additional examples of novel compounds
that inhibit specific bacterial targets but are subject to efflux include the peptide
deformylase inhibitor LBM415, which is subject to AcrAB-TolC-mediated efflux in
H. influenza [47], and CHIR-090 an LpxC inhibitor that has potent intrinsic activity
against P. aeruginosa but selects in vitro for mutations that upregulate expression of
several RND efflux pumps [48]. Similarly, standard antibiotics with good intrinsic
activity against Gram-negative pathogens, such as fluoroquinolones, many
β-lactams, and aminoglycosides, also select for mutations causing increased pump
expression which can decrease susceptibility substantially [3].
The selection of mutations leading to pump upregulation underscores the idea
that although RND pump expression is generally subject to intricate regulation,
pump expression is not typically induced by antibiotics of clinical importance. An
exception is the strong induction of the MexXY efflux pump of P. aeruginosa by
compounds that perturb protein synthesis. Even in that case, however, induction
occurs in response to a range of structurally and mechanistically unrelated protein
synthesis inhibitors including aminoglycosides and tetracyclines [40], novel ribo-
some inhibitors such as argyrin B [49] or to mutations that impair ribosome function
[50, 51]. Therefore, MexXY expression is responsive to ribosome impairment [52,
53] rather than to the specific antibacterial compounds, and in some cases such as
argyrin B, the inducing compound may not be a pump substrate. Novobiocin was
also shown to directly bind to the NalD repressor of the MexAB-OprM efflux pump
and induce pump expression [54].
4 Resistance of Gram-negative Bacilli to Antimicrobials 81
[75]. The latter affects the ability of polymyxin (cationic peptides) to interact with
LPS which is required for their entry into cells (discussed in Sect. 4.2.6). These
examples serve to illustrate the extensive ability of many Gram-negative bacteria to
survive exposure to toxic molecules by preventing their accumulation in the cell.
The ability of mechanisms such as efflux and porin loss to enhance survival under
exposure to antibiotics also supports the emergence of other resistance mechanisms
such as specific target mutations, ultimately facilitating the emergence of very high
levels of resistance.
One strategy to potentiate partner antibiotics that has garnered extensive interest
over the years is the design of efflux pump inhibitors (EPIs). In theory, a potent EPI
could improve the spectrum and potency of a range of currently used antibiotics
whose usefulness is compromised by efflux. EPIs could also serve to enhance and
extend the clinical usefulness of novel agents that are or may become affected by
efflux. Several EPIs have been described over the last two decades.
The first EPI described in detail as an inhibitor of multiple RND family pumps,
MC207,110 (phe-arg-β-napthylamide, PAβN) was originally identified by screen-
ing for compounds that potentiated the activity of pump substrate fluoroquinolone
4 Resistance of Gram-negative Bacilli to Antimicrobials 83
The mechanisms by which translocation of compounds occurs via porins have been
the subject of extensive investigation [126–128], and assays to evaluate porin trans-
location for use in drug design are being pursued [129, 130]. A recent study sug-
gested that porin traversal may be optimal for small polar compounds with charged
groups and a dipole moment having a component aligned perpendicular to its main
axis [128]. Since β-lactams have periplasmic targets, they may represent the class of
antibiotic that can best be optimized specifically for permeation through porins. It is
likely that this contributes in some cases to reducing the impact of efflux since rapid
influx can overwhelm the capacity of RND family efflux pumps even if the com-
pound is a pump substrate [36, 37]. The majority of novel antibacterial targets how-
ever are located in the cytosol, and understanding compound penetration (and
evasion of efflux) becomes more complex in that case than for compounds such as
β-lactams. This stems from the necessity to traverse the two distinctly different
membranes. The OM severely limits influx of larger or more hydrophobic com-
pounds, which must diffuse across the asymmetrical OM bilayer which has low
fluidity and presents a high barrier for lipophilic compounds, since they are excluded
from entry through porins [3]. Smaller more hydrophilic molecules can traverse the
OM through porins, but extensive optimization for polarity to maximize this can
hinder entry across the symmetrical phospholipid inner membrane bilayer, which
favors diffusion of hydrophobic molecules. Antibiotics directed at cytosolic targets
may therefore need some element of amphiphilicity to cross both membranes,
which also may increase recognition by RND efflux pumps [131]. A much better
understanding of the chemical property space required for the design of cell active
inhibitors of intracellular targets, and the representation of this chemical property
space typically found in corporate screening libraries, is likely required to overcome
the ongoing inability to deliver novel antibacterials in this area. Initial efforts toward
this understanding were described by O’Shea and Moser [132], who examined the
properties of a wide range of antibacterial compounds and, consistent with our
understanding of the cell envelope, correlated properties such as molecular weight
and polarity with Gram-negative antibacterial activity. A more recent look at data
from a wide range of screening efforts at AstraZeneca also suggested that polarity
and small size correlated with reduced efflux and cellular activity and increasing
compound hydrophobicity could drive biochemical target inhibition but possibly at
the expense of cellular activity [46]. However, increased polarity itself was not suf-
ficient to ensure antibacterial activity [46], again suggesting a fine balance of prop-
erties is likely necessary. This is consistent with the notion that the cell envelope and
efflux likely coevolved with intracellular targets to exclude molecules with the
properties to strongly bind and inhibit essential targets. These properties may also
be compound scaffold specific. Modulation of physicochemical properties (pKa and
logD) improved the antibacterial activity of novel bacterial type II topoisomerase
inhibitors [133]. A very recent study directly measured accumulation of compounds
in E. coli using mass spectrometry, and computational analysis indicated that rigid,
amphiphilic compounds with low globularity and containing an amine moiety accu-
mulated better. These rules were applied to convert a compound that was active only
against Gram-positive bacteria into one with E. coli activity [134]. This suggests it
4 Resistance of Gram-negative Bacilli to Antimicrobials 87
Fig. 4.4 Representative structures for each class of β-lactam. The core structure is depicted in
blue; the specific side chains are depicted in black
may be possible to derive some general rules to engineer compounds with Gram-
negative accumulation and cellular activity, but more research will be necessary to
validate this concept. Furthermore, compound accumulation must be achieved
together with low toxicity in order for resulting compounds to be therapeutically
useful. Efforts are also underway to explore new methods complementary to mass
spectrometry to measure cellular accumulation of compounds, which could further
assist in defining rules for cell penetration [135, 136]. Finally, the Trojan-horse
concept has also been applied to antibacterial design. In this scenario, an antibiotic
is linked to a compound that is actively transported into cells, thereby exploiting the
active uptake mechanism to drive intracellular accumulation of the chimeric antibi-
otic molecule. Among others, a recent example of this is compound cefiderocol
(S-649266), a catechol cephalosporin that is proposed to utilize energy-dependent
siderophore-iron uptake systems in Gram-negative bacteria for improved cellular
access [137].
4.2.2.1 β-Lactamases
The discovery of β-lactamases predated the clinical use of benzylpenicillin, but the
widespread use of these agents in the clinic has, over time, led to the emergence of
an astonishing number of β-lactamase variants [144, 145], which as a group can
degrade most or all β-lactam antibiotics. Indeed the development of new β-lactam
antibiotics is to some extent a continuing story of addressing the emergence of new
β-lactamases [146, 147], as is the ongoing development of β-lactamase inhibitors
(BLIs) for use in combination with β-lactams to restore their activity against
β-lactamase-expressing strains (see Sect. 4.2.2.3). β-lactamases hydrolyze the
β-lactam ring of all classes of β-lactam antibiotics by one of the two major mecha-
nisms. The first is mediated by an active-site serine (Ser), via a covalent enzyme
intermediate that is rapidly hydrolyzed causing inactivation of the antibiotic.
β-lactamases that operate by this mechanism are therefore referred to as serine
4 Resistance of Gram-negative Bacilli to Antimicrobials 89
functionality and divides β-lactamases into three major groups: group 1 cephalospo-
rinases (class C), group 2 serine β-lactamases (classes A and D), and group 3
metallo-β-lactamases. Each major group is then divided into several subgroups
based on specific attributes [145]. The first β-lactamase, TEM-1, identified in a clin-
ical isolate was reported in the early 1960s in an E. coli isolate from a patient in
Greece [149]. Since then, the number of β-lactamases identified has constantly
grown. A recent report estimated that over 2000 unique β-lactamases sequences
have been identified [150]. The major players in the clinic for infections caused by
Gram-negative pathogens are the extended-spectrum β-lactamases (ESBLs), the
AmpC cephalosporinases, and the serine and metallo-carbapenemases.
ESBL rates increased from 11.4% (2012) to 16.9% (2014). The predominant ESBLs
identified in this study were CTX-M-15 (59% of ESBLs) followed by SHV (19% of
ESBLs), both mainly in K. pneumoniae isolates, and CTX-M-14 (18% of ESBLs)
[157]. In Europe the ESBL rates vary considerably by country. The prevalence of
ESBLs in E. coli in the 2014 European surveillance varies from 3.3% in Iceland to
40.4% in Bulgaria. Even more alarming is the prevalence of ESBLs in K. pneu-
moniae with rates over 70% in Greece, Bulgaria, and Romania. On a positive note,
the rates of ESBLs did not increase from 2009 to 2014, attributed to the increased
use of carbapenems [158]. Also across Europe, the most prevalent ESBL types iden-
tified in clinical isolates were the CTX-M family β-lactamases, but the specific type
varies considerably among countries, with CTX-M-9 and CTX-M-14 enzymes
dominant in Spain and CTX-M-3 and CTX-M-15 dominant elsewhere [159].
Another growing family of ESBLs is the OXA-type enzymes that confer resis-
tance to ampicillin and cephalothin, are characterized by their high hydrolytic activ-
ity against oxacillin and cloxacillin, and are very poorly inhibited by clavulanic
acid. The OXA-type enzyme genes differ genetically from all other ESBLs. To date,
over 500 different OXA-type variants have been reported, but not all are ESBLs.
The OXA-type enzymes with activity against oxyimino-cephalosporins are OXA-
10 and its variants (OXA-11, OXA-14, OXA-16, and OXA-17), OXA-13 and its
variants (OXA-19 and OXA-32), and some other OXA enzymes (OXA-15, OXA-
18, and OXA-45). These enzymes have been identified mainly in P. aeruginosa
isolates [155, 160].
Even though ESBL incidence rates have not been increasing in the past few
years, they are still very high in some parts of the world and are a major health con-
cern. Further, ESBLs are often present on mobile genetic elements with other anti-
biotic-resistant determinants, including those for aminoglycosides and
fluoroquinolones. The use of carbapenems to treat infections caused by ESBL-
producing pathogens is increasing the emergence of carbapenem-resistant strains,
starting the debate on how to better treat those pathogens. Using a β-lactamase
inhibitor (see Sect. 4.2.2.3) with a β-lactam is in principle a targeted and effective
approach. A detailed analysis on the benefit of β-lactam/β-lactamase inhibitor com-
binations for the treatment of ESBL-producing pathogens can be found in a recent
review by Viale et al. [161].
AmpC β-lactamases belong to class C and functional group 1. They confer resis-
tance to cephamycins, such as cefoxitin and cefotetan, and cephalosporins, includ-
ing oxyimino-cephalosporins such as ceftazidime, cefotaxime, and ceftriaxone.
They are also able to hydrolyze to a lesser extent penicillins and aztreonam [162].
The majority of AmpC β-lactamases are not or are only weakly inhibited by inhibi-
tors of class A enzymes such as clavulanic acid, sulbactam, and tazobactam. Some
AmpC variants have been reported to be inhibited by tazobactam or sulbactam [162,
163]. Several AmpC β-lactamases are chromosomally encoded enzymes, found in
92 C. R. Dean et al.
Carbapenemases
Carbapenems are considered the most effective β-lactams for the treatment of seri-
ous infections caused by Gram-negative bacteria and present a broad spectrum of
antibacterial activity. Furthermore, carbapenems are relatively stable to most ESBLs
and class C enzymes and are deemed to be safer to use than any other last-resort
antibiotic. Therefore, the increasing number of reports of β-lactamases able to
hydrolyze carbapenems over the last few years is of major concern. Carbapenemases
are a heterogeneous group of β-lactamases, including members from classes A, B,
and D [178, 179]. Most carbapenemases are able to hydrolyze a very broad spec-
trum of β-lactams. Several class A enzymes, functional group 2f, with carbapenem-
inactivating activity, have been reported over the years. The first, SME-1, was
reported in 1990 in a Serratia marcescens isolate from the United Kingdom [180].
Subsequently, GES-1 in K. pneumoniae, SFC-1 in Serratia fonticola, and IBC-1/
IMI-1/NMC-A in Enterobacter cloacae have been reported [181]. One of the most
recent and widespread class A carbapenemases is the K. pneumoniae carbapene-
mases (KPCs). KPC-1 was the first carbapenemase identified for this family of
enzymes, reported in 1996 in North Carolina [182, 183]. KPCs, in contrast to the
other class A carbapenemases that are chromosomally encoded, are usually on
mobile genetic elements and since their discovery have spread to many other organ-
isms including most species of Enterobacteriaceae (including Enterobacter spp.,
Serratia marcescens, and Salmonella spp.), P. aeruginosa, and several other genera.
A recent study that analyzed 147 cases of infections due to carbapenem-resistant K.
pneumoniae from 2013 to 2014, in one hospital in Northern Italy, showed that the
major resistance determinant was KPC-3 (83.8%). The death rate was an alarming
24.0% in 2013 and 37.5% in 2014 [184]. This is of great concern especially in
Southern European countries where CREs expressing KPC are spreading rapidly.
Several enzymes of the class D family (OXA type) are able to degrade car-
bapenems [185]. The first identified enzyme in this class able to hydrolyze imi-
penem was OXA-23. It was isolated in the United Kingdom in 1985 from an A.
baumannii isolate and was originally characterized as ESBL [186]. Since this
initial report, several other OXA enzymes have been described in
Enterobacteriaceae, such as OXA-23-like, OXA-40-like, OXA-51-like, OXA-
58-like, and OXA-48-like enzymes [185, 187]. The most widespread enzymes in
this class are the OXA-48-like enzymes. OXA-48 was first isolated in a patient
with UTI in Turkey in 2001 from a strain of K. pneumoniae and is now widely
94 C. R. Dean et al.
The first of these to be identified and brought to the clinic was clavulanic acid, a
natural product isolated from Streptomyces clavuligerus [205], followed by the
semisynthetic penicillanic acid sulfone class of inhibitors (sulbactam and tazobac-
tam) [206, 207]. These BLIs all possess the basic core structure of a β-lactam which
allows for recognition and binding to β-lactamase. However, key structural differ-
ences from β-lactams eliminate most or all intrinsic antibacterial activity against
many bacteria and render them mechanism-based “suicide inhibitors” of sensitive
β-lactamases [208]. The mechanism of β-lactamase inactivation by these inhibitors
is complex, but in general, the active-site serine of the β-lactamase attacks the car-
bonyl group in the β-lactam ring of clavulanic acid leading to acylation of the
β-lactamase. This is then followed by a series of secondary reactions in the enzyme
active site that irreversibly inactivate the enzyme [209–211]. A main difference
between the BLI molecules and β-lactams that facilitates this mechanism is that
BLIs possess good leaving groups at the C-1 position of their five-membered rings.
This allows for secondary ring opening and subsequent β-lactamase enzyme modi-
fication. Important factors for BLI efficacy include high acylation and low deacyla-
tion rates, which localize them for a longer time period in the enzyme active site and
a low number of hydrolytic events (inhibitor molecules hydrolyzed per unit time)
necessary for complete enzyme inactivation (termed turnover number or tn).
Differences exist in these factors between clavulanic acid, sulbactam, and tazobac-
tam, and these are affected by differences in the active sites among β-lactamases.
Clavulanic acid and tazobactam cover most class A β-lactamases including ESBLs.
Sulbactam also covers these but is less potent against some enzymes. Tazobactam
and sulbactam are better inhibitors of class C carbapenemases than clavulanic acid
4 Resistance of Gram-negative Bacilli to Antimicrobials 97
and notably differ from clavulanic acid in that they do not induce expression of
AmpC in bacteria where this enzyme is inducible [212]. However, none of these can
cover strains producing metallo-β-lactamases such as NDM-1. Clavulanic acid is
partnered in the clinic with amoxicillin or ticarcillin, sulbactam with ampicillin, and
tazobactam with either piperacillin or ceftolozane.
Avibactam, the first non-β-lactam BLI approved for clinical use, is a broad inhib-
itor of class A (including KPCs), class C, and some class D β-lactamases. Avibactam
is a member of the diazabicyclooctane (DBO) chemical class [213], and as such it
has a different mechanism of inhibition from previous BLIs. This mechanism is not
yet fully understood, but it appears that avibactam functions as a slowly reversible
covalent inhibitor with release of intact avibactam for most class A and C
β-lactamases [214]. The enzymes appear to be slowly acylated and slowly deacyl-
ated, with no or only low-level hydrolysis of the inhibitor molecule. An exception to
this was inhibition of KPC-2 which was rapidly acylated but slowly deacylated with
hydrolysis of avibactam, so differences do exist. The release of intact avibactam in
most cases however is thought to allow for recycling of the inhibitor by β-lactamases
in the cell, leading to better inhibitory efficiency than BLIs like clavulanic acid or
tazobactam that are hydrolyzed. Avibactam is a substantially more effective inhibi-
tor of key β-lactamases like TEM-1, KPC-2, and AmpC from P. aeruginosa than
clavulanic acid, sulbactam, or tazobactam but has limited coverage of class D
enzymes, although it does cover OXA-48. Like previous BLIs, avibactam does not
cover metallo-β-lactamases like NDM-1. Avibactam was introduced into the clinic
very recently (2015), partnered with ceftazidime, and is in clinical trials for combi-
nation use with ceftaroline or aztreonam. The latter partnering with the monobac-
tam aztreonam is meant to capitalize on the idea that monobactams are inherently
stable to metallo-β-lactamases, and the combination should therefore cover strains
expressing metallo-β-lactamases and/or serine β-lactamases. Although ceftazidime/
avibactam has only been in clinical use a short while, resistance to this combination
due to porin loss and upregulation of β-lactamase expression has been reported
[215]. It has recently been suggested that ceftazidime/avibactam could be used in
combination with aztreonam for coverage of metallo-β-lactamase/serine-β-
lactamase-producing clinical isolates [216], but appropriate dosing would need to
be established.
The success of avibactam has inspired efforts to identify next-generation
DBO β-lactamase inhibitors. These include relebactam (MK7655), directed at
some class A, including KPCs, and class C β-lactamases (currently in Phase III
trials, in combination with imipenem [217]). Relebactam possesses a narrower
spectrum than avibactam since it does not include class D β-lactamases such as
OXA-48. Nacubactam (RG6080, OP0595), directed at class A and C β-lactamases,
is currently in Phase I trials and is intended for combination with meropenem
(Table 4.3). Some DBOs also possess intrinsic antibacterial activity and this war-
rants some discussion. They inhibit PBP2 in some Gram-negative pathogens,
similar to the β-lactam mecillinam [218]. PBP2 inhibition can be synergistic
with inhibition of other PBPs (i.e., with other β-lactams), as has been reported
for nacubactam [218, 219], and this has been referred to as an “enhancer effect”
98 C. R. Dean et al.
avibactam; this combination has the potential to cover strains expressing both
metallo-β-lactamase and serine β-lactamases. This combination is currently under-
going Phase III clinical trials. More recently, an innovative approach was under-
taken to design novel next-generation monobactams that are not significantly
impacted by most serine β-lactamases while retaining their intrinsic stability to the
metallo-β-lactamases. One of these, LYS228 [223], demonstrated excellent potency
against MDR Enterobacteriaceae, including CRE [224, 225], and has entered Phase
II clinical trials (Novartis). Significant effort has also been devoted to the discovery
of therapeutically useful inhibitors of the class B metalloenzymes. This has lagged
to some extent since class B enzymes have a different mechanism than serine
β-lactamases, and it appears that the design of inhibitors capable of covering mul-
tiple clinically important class B enzymes is technically challenging. Inhibitors of
class B enzymes would also need to be highly specific and avoid human metalloen-
zymes to avoid toxicity issues. While the prevalence of class B enzymes remained
relatively low in the past and their contribution to worldwide carbapenem resistance
was initially considered minimal, this viewpoint has changed in recent years with
increased spread of class B enzymes like NDM-1 and their linkage with other resis-
tance determinants. Recent efforts in the search for class B inhibitors include the
discovery of the natural product aspergillomarasmine, which is active against
NDM-1 and VIM-2 [226]. Novel bisthiazolidine (BTZ) inhibitors of class B
enzymes have also recently been described [227]. Of most interest are reports from
The Medicines Company on a new series of cyclic boronate compounds derived
from RPX7009 (Table 4.3) with broad-spectrum carbapenemase activity including
metallo-β-lactamases, in preclinical development [228].
inducible ampC [234, 235]. Sulbactam and tazobactam do not have this induction
effect and as such can be better options (e.g., tazobactam paired with piperacillin
against P. aeruginosa). Active efflux of BLIs [20, 236] or changes in influx, possibly
due to porin loss, may also serve to reduce their concentration relative to the
β-lactamases in the cells, decreasing their effectiveness, although influx of com-
pounds such as avibactam is not currently well understood [237]. Defects in porins
OmpK35 and/or OpmK36 were, however, associated with decreased effectiveness
of imipenem/relebactam and meropenem/RPX7009 in clinical isolates isolated
from hospitals in New York [219, 238]. Efflux was implicated as an important medi-
ator of resistance to ceftazidime/avibactam in P. aeruginosa, but this remains to be
further explored [239]. Porin mutations combined with upregulated expression of
plasmid-borne KPC-3 have been associated with clinical resistance to ceftazidime/
avibactam in K. pneumoniae [215], and porin mutations were associated with cef-
taroline/avibactam resistance in E. cloacae mutants selected in vitro [240].
Resistance to BLIs, mainly clavulanic acid, also resulted from the emergence of
new β-lactamases that are resistant to the BLI. Very soon after the introduction of
clavulanic acid into clinical use, such variants began to emerge, with the first
reported being variants of the class A TEM enzyme that were resistant to clavulanic
acid, found in E. coli clinical isolates [241, 242]. These TEM variants were altered
at their Arg244 residues to either Cys or Ser [243]. This position had been shown
earlier to be important for clavulanic acid inhibitory function, and so this clinical
outcome might have been expected. These variants were initially designated
inhibitor-resistant TEM (IRT-1 and IRT-2), but since then, 37 clavulanic acid-
resistant TEM variants have been identified (cataloged at http://www.lahey.org/
Studies/temtable.asp, functional group br), and the convention now is that these all
have TEM numerical designations. These variants are found mainly in E. coli iso-
lates but also occur in Klebsiella [244], Proteus [245], Shigella [246], and
Citrobacter [247]. Inhibitor resistance can also be combined with amino acid sub-
stitutions conferring β-lactamase activity against oxyimino-β-lactams (ESBL), and
these are referred to as complex mutant TEMs (CMT). Currently 11 of these vari-
ants have been described (http://www.lahey.org/Studies/temtable.asp, functional
group ber). As well, seven inhibitor-resistant variants of the class A enzyme SHV
have also been described (http://www.lahey.org/Studies/), with the most recent
being SHV-107 found in a K. pneumoniae clinical isolate [248]. It should be noted
that inhibitor-resistant β-lactamases generally refer to clavulanic acid, and these can
also be resistant to sulbactam, but generally they remain susceptible to tazobactam
[249–251]. Therefore these enzymes mainly affect amoxicillin/clavulanate, ticarcil-
lin/clavulanate, or ampicillin/clavulanate but not piperacillin/tazobactam. However,
in 2010, the emerging class A ESBL KPC-2 carbapenemase [252] was shown to
also be resistant to clavulanic acid, sulbactam, and tazobactam, raising serious con-
cerns [253].
Although ceftazidime/avibactam is active against KPC producers, providing
effective treatments in the near term, the emergence of β-lactamase variants resis-
tant to avibactam has already been reported [254]. In vitro selection studies demon-
strated that variants of KPC-3 [254] or certain inhibitor-resistant SHV enzymes
4 Resistance of Gram-negative Bacilli to Antimicrobials 101
4.2.3 Quinolones
Quinolones and the related fluoroquinolones (Fig. 4.5) were introduced into clinical
use in the 1960s and 1980s, respectively. First-generation quinolones (e.g., nalidixic
acid) were restricted generally to treating urinary tract infections, because of subop-
timal systemic distribution and somewhat limited activity. Second-generation fluo-
roquinolones (e.g., norfloxacin and ciprofloxacin) had improved tissue distribution
and a broadened antibacterial spectrum, allowing for expanded use and perhaps
overuse. Newer third- and fourth-generation fluoroquinolones (e.g., ofloxacin, lome-
floxacin, levofloxacin, trovafloxacin, gatifloxacin, moxifloxacin, sparfloxacin) were
focused mainly on improved Gram-positive and atypical (e.g., Mycoplasma,
Legionella) coverage. Quinolone antibiotics act by inhibiting DNA gyrase and
topoisomerase IV enzymes that control DNA topology and play essential roles in
DNA replication, transcription, and recombination [258]. The DNA gyrase holoen-
zyme tetramer consists of two subunits each of GyrA and GyrB, which act to intro-
duce negative superhelicity into DNA. This is required for initiation of replication,
replication fork movement, and transcription [259]. The domain responsible for
DNA strand passage resides on GyrA, whereas GyrB contains an ATPase domain.
The topoisomerase IV tetrameric holoenzyme similarly consists of two subunits,
each of ParC and ParE, and functions to relax both positive and negative supercoils
and to direct decatenation (unlinking) of replicated chromosome copies to allow for
chromosomal partitioning upon cell division. The DNA strand passing domain is
located on ParC, and the ATPase activity is mediated by ParE. Both holoenzymes are
102 C. R. Dean et al.
Fig. 4.5 Chemical structure of key quinolones. Nalidixic acid (first generation), ciprofloxacin
(second generation), levofloxacin (third generation), and trovafloxacin (fourth generation)
type II topoisomerases that introduce double-stranded breaks in DNA and pass DNA
strands/helices through each other via a transient “cleaved complex” where the
enzyme, covalently linked to the DNA, serves as a bridge between the DNA ends,
mediating strand breakage, strand passage, and resealing [259]. Although the exact
mechanisms by which different quinolones kill bacteria have not been fully unrav-
eled, their general mechanism involves forming reversible non-covalent complexes
with the topoisomerases bound to DNA. This forms a drug-enzyme-DNA complex
(ternary complex) that is trapped as the cleaved complex and ligation of the DNA
ends is prevented [258]. Subsequent destabilization of the complex without rejoin-
ing the ends introduces double-stranded DNA breakage, fragmenting the genome
and ultimately causing cell death [258, 260]. The trapped cleaved complexes also
interfere with progression of replication forks, blocking DNA synthesis [261] and
with transcription by blocking RNA polymerase [262] and disrupting the action of
DNA helicases [263]. Additional mechanisms may contribute to cell killing in cer-
tain Gram-negatives. For example, a recent report detailing the transcriptomic inter-
rogation of ciprofloxacin-treated P. aeruginosa implicated induction of a pyocin
system in cell-killing activity [264] (See also Chaps. 16 and 20).
Since there is significant amino acid sequence homology between the GyrA
and ParC and GyrB and ParE proteins, individual quinolone molecules can
inhibit the activities of both enzymes, and most quinolones will inhibit both tar-
gets to varying degrees. Either topoisomerase can constitute the primary or sec-
ondary target of quinolones in different bacteria. Given their broad-spectrum and
4 Resistance of Gram-negative Bacilli to Antimicrobials 103
The role of Gram-negative RND family efflux pumps in intrinsic and mutationally
acquired resistance to antibiotics is covered in detail in Sect. 4.2.1 and in recent
reviews [3]. The potent broad-spectrum activity of fluoroquinolones against even
intrinsically resistant Gram-negative bacteria such as P. aeruginosa indicates that
RND efflux does not mediate enough intrinsic resistance to some fluoroquinolones
to limit their spectrum. This may relate to some extent to the hydrophilicity [3] as
well as to the overall target potency and cidality of fluoroquinolones. However, fluo-
roquinolones are substrates of a wide range of RND family pumps (Table 4.1)
including the AcrAB-TolC pump of E. coli and Salmonella spp., the AcrEF pumps
of E. coli and Salmonella enterica, the CmeABC pump of Campylobacter jejuni,
and MexAB-OprM, MexXY-OprM, MexCD-OprJ, and MexEF-OprN of P. aerugi-
nosa [3]. Mutations in regulatory genes causing pump overexpression and decreased
susceptibility can be readily selected by in vitro exposure to quinolones. However,
in most cases overexpression of efflux pumps alone, in the absence of other mecha-
nisms, affords only modest reductions in fluoroquinolone susceptibility (Table 4.4).
Efflux pump overexpressing mutants are routinely found among clinical isolates
[276]. Upregulation of the AcrAB-TolC pump in fluoroquinolone-resistant E. coli
clinical isolates contributed to high-level fluoroquinolone resistance along with
QRDR mutations [67, 277]. Similarly, AcrAB-TolC upregulation played a role in
fluoroquinolone resistance in K. pneumoniae and K. oxytoca clinical isolates [278].
A P. aeruginosa clinical isolate with a mutation in gyrB and upregulated for MexAB-
OprM emerged during ciprofloxacin monotherapy [279]. Since RND pumps have
broad substrate ranges, selection of pump overexpression during previous treatment
with various antibiotics will result in selection of pump upregulation which will
affect fluoroquinolones and vice versa. RND family efflux pumps function in con-
cert with the OM permeability barrier, and therefore any reduction in a compound’s
ability to cross the OM will have a corresponding enhancing effect on
4 Resistance of Gram-negative Bacilli to Antimicrobials 105
efflux-
mediated resistance. Fluoroquinolones cross the outer membrane either
through water-filled porin channels or by diffusion through lipid domains in the
outer membrane depending on the hydrophobicity of the quinolone [280]. Reduced
porin levels have also been associated with fluoroquinolone resistance in E. coli
[277, 281] and S. enterica [282] clinical isolates. Reduced porin levels often occur
concomitantly with upregulation of efflux pumps [283] potentially linking reduced
influx and increased efflux via a single mutation.
several pumps could not be correlated with higher-level resistance seen for some
isolates that also harbored QRDR mutations, suggesting that other as yet unidenti-
fied factors may mediate higher-level resistance in some QRDR mutants [275].
Similar interplay between plasmid-borne resistance mechanisms and other mecha-
nisms is likely also occurring, both in terms of determining susceptibility and in
facilitating the emergence of high-level resistance. Qnr proteins only cause a mar-
ginal shift in quinolone susceptibility [284] (Table 4.4), but this is additive with
target-based or other mechanisms and will contribute to the emergence of higher-
level clinically relevant resistance [295]. Like qnr, the level of resistance conferred
by aac(6′)-1b-cr alone was modest; however, more significant levels of resistance
were observed when aac(6′)-1b-cr and qnrA were found together (Table 4.4).
Furthermore, the presence of plasmid-borne aac(6′)-1b-cr in E. coli resulted in a
greater recovery of resistant mutants during selection experiments with ciprofloxa-
cin [298], essentially by widening the mutant selection window. This again high-
lights the interplay of determinants such as aac(6′)-1-cr and qnrA in the stepwise
acquisition of clinically significant resistance [318]. An additional factor that may
have contributed to the emergence of high-level fluoroquinolone resistance is that
DNA damage and interference with DNA replication caused by quinolones induce
the SOS response, leading to upregulation of error-prone DNA polymerases.
Evolution of quinolone resistance in E. coli in vitro and in an animal model of infec-
tion was curtailed in mutants lacking the SOS response [319].
of action distinct from fluoroquinolones that may involve targeting GyrB (AZD0914,
now ETX0914, Entasis Therapeutics) are in clinical trials and may find utility in
treating infections due to Gram-positive and/or fastidious Gram-negative patho-
gens, such as N. gonorrhoeae [326–328]. Another class of novel inhibitors of GyrB,
which bind to the TOPRIM domain and are not affected by fluoroquinolone resis-
tance mutations, has recently been described [329]. A detailed discussion of non-
quinolone inhibitors of topoisomerases is presented in Chap. 19.
4.2.4 Tetracyclines
Tetracyclines are bacteriostatic and prevent bacterial growth by binding to the ribo-
some, thereby blocking protein synthesis. They bind the A-site of the ribosomal 30S
subunit which prevents the entrance of aminoacyl-tRNAs into the mRNA-ribosome
complex, ultimately preventing incorporation of amino acids into the newly emerg-
ing polypeptide [330–332]. The ribosomal target is relatively conserved in bacteria,
and tetracyclines can therefore have a broad spectrum of antibacterial activity, cov-
ering many Gram-positive, Gram-negative, anaerobic, and atypical pathogens. The
original tetracycline, chlortetracycline (also referred to as Aureomycin), is a natural
product produced by Streptomyces aureofaciens and was identified in the late 1940s
by Benjamin Duggar at Lederle Laboratories [333]. Over time other natural exam-
ples were discovered, and routes for making semisynthetic tetracyclines were devel-
oped. The latter allowed detailed exploration of this chemical scaffold, leading to
second-generation tetracyclines doxycycline and minocycline and culminating with
third-generation tetracyclines omadacycline and the glycylcycline tigecycline [334]
which has now been in clinical use for over 10 years. Tigecycline (Tygacil®, Pfizer
Inc.) is approved in the United States and Europe for the treatment of complicated
skin and intra-abdominal infections and in the United States for community-
acquired bacterial pneumonia. More recently eravacycline (TP-434), a fully syn-
thetic fluorocycline of the tetracycline class, has completed a Phase II study in
complicated intra-abdominal infection (cIAI) and is currently undergoing Phase III
studies in both cIAI and complicated urinary tract infection (cUTI) (www.clinical-
trials.gov) [335]. The latter two compounds are of particular interest in that they
have a broader spectrum of antibacterial activity and largely evade the tetracycline-
specific, acquired resistance mechanisms of MFS efflux and ribosomal protection
[336], described below in Sects. 4.2.4.1 and 4.2.4.2. Examples of tetracycline chem-
ical structures are shown in Fig. 4.6.
Tetracyclines have now been in use for several decades in human and veterinary
medicine as well as in agriculture. Correspondingly, resistance to earlier-generation
tetracyclines became fairly widespread some time ago [337–339]. There are two
main tetracycline-specific mechanisms of resistance in Gram-negative pathogens:
tetracycline-specific active efflux and ribosomal protection. Additional mechanisms
are active-site rRNA mutations and tetracycline-modifying enzymes. Efflux by
broad specificity RND family pumps (described in Sect. 4.2.1) also affects suscep-
110 C. R. Dean et al.
Fig. 4.6 Example of tetracyclines (left side) and glycylcyclines (right side). The key modification
at the 9 position (tert-butyl-glycylamido) differentiating the glycylcycline scaffold is depicted in
blue for tigecycline
4.2.4.1 Efflux
There are many tetracycline-specific efflux pumps described for both Gram-positive
and Gram-negative bacteria [340], and they function by actively extruding tetracy-
cline from the cell and preventing accumulation to a level sufficient to fully inhibit
the ribosome. Examples of pumps that are commonly found in Gram-negative
pathogens are listed in Table 4.5, and an updated table of the distribution of these
4 Resistance of Gram-negative Bacilli to Antimicrobials 111
ers that operate through the exchange of a proton for the tetracycline molecule
which drives transport of the tetracycline against a chemical concentration gradient,
in this case from the cytoplasm across the inner membrane into the periplasmic
space between the inner membrane and OM. Single-component pumps like TetA,
located in the inner membrane, are generally thought to be more effective at extrud-
ing compounds from the cytoplasm than are pumps of the more broadly active RND
family. The latter are generally thought to recognize compounds in the periplasm or
when diffusing into the inner membrane. This is important, since in Gram-negative
bacteria, the single-component pumps cannot extrude antibiotics completely out of
the cell into the surrounding milieu but will deposit the compound into the periplas-
mic space between the inner and outer membranes. This can concentrate the tetra-
cycline in the periplasm, from which it could diffuse back across the inner membrane
into the cell in the absence of additional efflux across the OM. Consistent with this,
higher levels of resistance mediated by pumps like TetA often require interplay with
efflux across the OM by RND family pumps such as MexAB-OprM in P. aerugi-
nosa or AcrAB-TolC in E. coli [6, 24]. These RND family pumps have a broad
substrate specificity which includes tetracyclines, and the combined effect of spe-
cific single pump efflux from the cytosol and subsequent RND-mediated efflux
from the periplasm to the outside of the bacterium can lead to high levels of resis-
tance in some Gram-negative pathogens.
Although the TetA MFS family efflux pumps are currently widespread among
clinical isolates, presumably driven by the extensive use of earlier-generation tetra-
cyclines, there have been significant advancements in circumventing the impact of
these pumps with each subsequent generation of tetracyclines. Understanding the
emergence of clinical resistance to early generations of tetracyclines, along with
improved understanding of tetracycline mechanism of action, was a key driving
force for renewed interest in developing new tetracyclines that were not subject to
this mechanism. Correspondingly, efforts leading to the identification of tigecycline
were specifically directed toward achieving cellular activity against tetracycline-
resistant bacteria [334, 349–351], including those expressing tetracycline efflux
pumps. That this was achieved with tigecycline is shown by the potent antibacterial
activity in broad susceptibility testing with resistant clinical isolates that supported
clinical development as an “expanded-spectrum” antibiotic for treatment of
multidrug-resistant Gram-negative infections (excluding P. aeruginosa) [349, 352–
359]. Specifically showing that tigecycline circumvents resistance mediated by
these pumps, expression of Tet(A), Tet(B), or Tet(X) in a susceptible E. coli strain
background conferred very high levels of resistance to tetracyclines (MIC ≥ 128 μg/
mL) but had a much smaller or no impact on the third-generation tigecycline (or
eravacycline (TP-434)), depending on the pump [360]. Furthermore, no correlation
was seen between the presence of tetracycline-specific efflux genes tet(A) to tet(E)
and insusceptibility to tigecycline in strains of Enterobacteriaceae [353]. It should
be noted though that some variation in amino acid residues important for recogni-
tion of tetracyclines has been reported for TetA proteins expressed from tet(A) genes
residing on different genetic elements and this can have a modest effect on how
much susceptibility is shifted when the pump is expressed [360, 361]. Whether this
4 Resistance of Gram-negative Bacilli to Antimicrobials 113
portends the selection of mutations in tetracycline pump genes over time that
increase recognition of tigecycline remains to be seen.
Tetracyclines are substrates of several RND family pumps [3] including AcrAB-
TolC in E. coli and K. pneumoniae and the MexAB-OprM, MexXY-OprM, and
MexCD-OprJ pumps in P. aeruginosa. RND efflux pumps are therefore important
for determining the Gram-negative spectrum of these compounds. Second-
generation compounds such as doxycycline and minocycline possess a broader anti-
bacterial spectrum than tetracycline, most notably against Acinetobacter,
Burkholderia, and Stenotrophomonas, but their activities against P. aeruginosa and
most species of Enterobacteriaceae are still limited, in large part due to RND-
mediated efflux [362]. Tigecycline has an expanded spectrum, covering a range of
Gram-negative pathogens including Enterobacteriaceae and non-fermenters such
as Acinetobacter, Stenotrophomonas, and Burkholderia [363]. Therefore basal-
level RND-mediated efflux alone does not exclude these organisms from the spec-
trum of tigecycline. However, RND-mediated efflux is a factor excluding P.
aeruginosa (MexAB-OprM, MexXY-OprM) [40] and Proteus mirabilis (AcrAB)
[41] from the spectrum of tigecycline. Since tigecycline is a substrate of AcrAB-
TolC present in many Gram-negative pathogens within the spectrum of tigecycline,
RND-mediated efflux also posed a threat as a resistance mechanism, either via
mutational upregulation of pump expression or indirectly by exacerbating other as
yet unknown mechanisms. Supporting this, RND-mediated efflux has been impli-
cated as a determinant of resistance in laboratory and clinical isolates of Morganella
morganii [364], K. pneumoniae [42, 365, 366], E. coli [43], Enterobacter cloacae
[367]/E. aerogenes [368], and Salmonella enterica [369]. As well, another RND
family pump, OqxAB, may play a role in decreasing susceptibility to tigecycline in
K. pneumoniae [366], and the AdeABC [61, 370, 371], AdeFGH [372, 373], and
AdeIJK [374] efflux pumps have been associated with decreased susceptibility to
tigecycline in A. baumannii.
Resistance to some tetracyclines (first and second generation) can be caused by the
action of ribosomal protection proteins (RPPs). Several of these proteins have been
identified (e.g., Tet(B), Tet(O), Tet(M), Tet (S), Tet(Q), Tet(W)) [375] (Table 4.5),
and a comprehensive update on the distribution of these determinants in Gram-
negative bacteria is maintained at http://faculty.washington.edu/marilynr/tetweb2.
pdf. The best studied of these proteins are Tet(O) and Tet(M) [375]. Plasmid-borne
tet(O) was first identified in a Campylobacter jejuni clinical isolate [376] and later
in Campylobacter coli. A similar gene, designated otr(A), was identified in the
114 C. R. Dean et al.
Current efforts in the search for next-generation tetracyclines are largely being done
by Tetraphase Pharmaceuticals, specifically centered on the fully synthetic fluoro-
cyclines eravacycline, TP-271, and TP-6076. Eravacycline (currently in Phase III)
4 Resistance of Gram-negative Bacilli to Antimicrobials 117
Fig. 4.7 Chemical structure of key aminoglycosides used in the clinic. The positions of covalent
chemical modification by aminoglycoside-modifying enzymes are shown in blue
is generally more potent overall and has a slightly better spectrum than tigecycline,
but also does not cover P. aeruginosa. TP-271 has a much more limited spectrum
and is directed at the bacterial pathogens responsible for community-acquired pneu-
monia [406]. TP-6076 has potent activity against a range of pathogens including A.
baumannii and carbapenem-resistant Enterobacteriaceae and has entered Phase I
trials as of this writing. It was also selected for funding support from CARB-X
(www.carb-x.org).
4.2.5 Aminoglycosides
Aminoglycosides are one of the major classes of antibiotics used in the clinic to
treat Gram-negative bacillary infections. The most widely used aminoglycosides
are tobramycin, gentamicin, and amikacin, mainly for the treatment of P. aerugi-
nosa meningitis and pneumonia. Tobramycin is also used in two inhaled formula-
tions (TOBI ® Podhaler ®, Novartis) for the treatment of chronic P. aeruginosa
infections in cystic fibrosis patients (Fig. 4.7). Streptomycin, neomycin, and kana-
mycin are used for the treatment of infections caused by E. coli, Proteus species,
Enterobacter aerogenes, K. pneumoniae, Serratia marcescens, and Acinetobacter
species [407–410]. Aminoglycosides act by binding to bacterial ribosomes and
therefore blocking bacterial protein synthesis. They mainly bind to the aminoacyl-
tRNA recognition site (A-site) of the 16S ribosomal RNA of the 30S ribosome [331,
411–414]. This binding causes codon misreading and the corresponding introduc-
tion of incorrect amino acids in the growing polypeptide. This amino acid mis-
incorporation causes rapid cell death [415]. All aminoglycosides are bactericidal
and have a prolonged postantibiotic effect due to the extended time needed to
118 C. R. Dean et al.
Table 4.6 Major aminoglycoside-modifying enzymes present in clinical isolates and their
resistance profile
Type Enzymes Resistance conferred References
Aminoglycoside AAC(6′)-I Tobramycin, amikacin, [441, 444]
acetyltransferases (AACs) netilmicin, dibekacin, sisomicin,
kanamycin, isepamicin
AAC(3)-IIa Tobramycin, gentamicin, [441, 445]
netilmicin, dibekacin, sisomicin
AAC(3)-I Gentamicin, sisomicin, [441, 446]
fortimicin
AAC(6′)-Ib-cr Kanamycin, amikacin and [447, 448]
tobramycin, ciprofloxacin, and
norfloxacin
Aminoglycoside APH (3′)-Ia Kanamycin, neomycin, [409, 431–436,
phosphotransferases (APHs) streptomycin, lividomycin, 441, 449, 450]
paromomycin, and ribostamycin
APH (3″)-III Kanamycin, neomycin, [441, 451]
lividomycin, paromomycin,
butirosin, and ribostamycin
Aminoglycoside ANT(2″) Tobramycin, gentamicin, [441, 452]
nucleotidyltransferases dibekacin, sisomicin, kanamycin
(ANTs) ANT(4′) Tobramycin, amikacin, [441, 453]
dibekacin, kanamycin,
isepamicin
4.2.5.4 Biofilms
Growth in the biofilm mode is another barrier for the entry of aminoglycosides in
bacteria, contributing to intrinsic and adaptive resistance to this class of antibiotics.
Biofilms are defined as an intertwined community of bacteria adhering on a surface
and surrounded by a self-produced matrix composed of extracellular DNA, pro-
teins, and polysaccharides. Biofilms play a key role in chronic P. aeruginosa infec-
tions and have been associated with pulmonary infections in patients with CF where
aminoglycosides and, in particular, tobramycin are routinely used [477, 478].
Therefore understanding the role of biofilms in relation to aminoglycoside resis-
tance is of high importance. Subinhibitory concentrations of aminoglycosides,
especially tobramycin, have been shown to induce biofilm formation in P. aerugi-
nosa by the induction of the aminoglycoside response regulator (arr) gene. This
gene is postulated to be involved in the regulation of cell surface adhesiveness and
therefore contributes to biofilm-specific aminoglycoside resistance. Some studies
122 C. R. Dean et al.
Fig. 4.8 Structure of novel aminoglycosides currently in clinical development (plazomicin and
arbekacin) and preclinical characterization (TS3112)
have hypothesized the ability of the biofilm matrix to bind aminoglycosides and
therefore drastically reduce their antibacterial activity [479, 480]. Several in vitro
studies have shown co-dosing of an aminoglycoside with a cationic steroid antibi-
otic, CSA-13, or a cationic peptide, DJK-5, is effective in overcoming biofilm-
mediated resistance [481, 482]. More study to understand the molecular mechanisms
of biofilm induction by aminoglycosides as well as device strategies to inhibit bio-
film formation are needed.
not be used [494, 495]. Arbekacin is stable to some of the most common APHs,
ANTs, and AACs present in clinical isolates. Plazomicin, a sisomicin analog with
potent activity against Enterobacteriaceae (MIC90 ≤ 2 μg/mL), has recently com-
pleted Phase III trials in cUTI and in patients with serious bacterial infections due
to CRE (http://www.achaogen.com/plazomicin/). Plazomicin, like its parent com-
pound sisomicin, is resistant to several AMEs such as APH(3′)-III, APH(3′)-VI, and
APH(3′)-VII and ANT(4′)). The addition of hydroxyl-aminobutyric acid substitu-
ent at the N-1 position and hydroxyethyl substituent at the 6′ position has rendered
plazomicin resistant to AAC [3], ANT(2″), APH(2″), and AAC(6′) enzymes [409].
Plazomicin activity is abrogated by RMTs that are frequently present on mobile
genetic elements that also carry β-lactamases like NDM-1 in Enterobacteriaceae
[431, 434, 496]. This may turn out to be a liability for the clinical longevity of
plazomicin against Enterobacteriaceae. Meiji Seika Pharma Co. recently reported a
semisynthetic apramycin, named TS3112 (Fig. 4.8), which is active against Gram-
positive and Gram-negative bacteria producing both AMEs and RMTs. TS3112
showed potent bactericidal activity in a murine thigh model of K. pneumoniae
expressing RTMs [497]. TS3112 is currently in early-stage characterization, show-
ing encouraging in vitro results.
New delivery strategies for aminoglycosides have been adopted in the past few
years that provide higher local concentration at the infection site with a lower total
amount of drug delivered, which reduces systemic exposure and safety liabilities of
this class of drug. The best example of new delivery method for inhaled aminogly-
coside is tobramycin inhalation powder (TOBI® Podhaler®), delivered via the
T-326 inhaler (Novartis). Long-term safety studies in patients with CF have shown
that it is well tolerated, there was no evidence of serum tobramycin accumulation
with successive cycles, and no unexpected adverse events were observed. Further,
the new powder delivery method improved compliance due to shorter administra-
tion time, convenience, and ease of use [498, 499]. Bayer Healthcare, in collabora-
tion with Nektar, is currently developing BAY41-6551, a drug-device combination
of a specially formulated amikacin. BAY41-6551 has recently completed Phase III
as an adjunctive treatment for intubated and mechanically ventilated patients with
Gram-negative pneumonia and showed bactericidal activity against most isolates
tested with amikacin MICs ≤ 256 μg/mL [500–502].
Aminoglycosides are a key class of antibiotic used by physicians to treat serious
infections caused by MDR Gram-negative and Gram-positive pathogens. Continued
characterization of aminoglycoside resistance mechanisms may enable the design
of resistance determinant inhibitors and/or new aminoglycosides capable of circum-
venting these mechanisms. Additionally, efforts aimed at optimization of dosing
regimens and discovery of new delivery strategies should help to maintain and
extend the clinical utility of this important antibiotic class.
124 C. R. Dean et al.
4.2.6 Polymyxins
Polymyxins are an older class of cationic cyclic lipopeptide antibiotics that were
introduced into clinical use in the 1950s (polymyxin B and colistin, also known
as polymyxin E). However, the use of polymyxins declined sharply around the
early 1970s due to concerns of toxicity [503, 504] and the availability of safer
antibiotics. The mechanism by which polymyxins kill bacteria is not fully under-
stood. One mechanistic step that is well established is an initial interaction of the
cationic peptide with negative charges on the lipopolysaccharide (LPS) that
forms the outer leaflet of the Gram-negative OM [505, 506]. This interaction
occurs mainly via negatively charged phosphate residues located on lipid A and
is required for the “self-promoted uptake” of polymyxins into the bacteria.
Binding of polymyxin is thought to displace divalent cations (Mg2+, Ca2+) that
cross-link adjacent LPS molecules, and this can disrupt to some extent the per-
meability barrier of the OM, which also increases uptake of the polymyxin. This
is unlikely to be responsible entirely for cell killing. As discussed in more detail
in Sect. 4.2.6.4 below, derivatives of polymyxin (e.g., polymyxin B nonapeptide)
exist with dramatically reduced antibacterial activity that retain the OM disrup-
tion activity. Polymyxins may ultimately kill via mechanisms that include lysis
of the inner membrane [505] generation of toxic hydroxyl radicals [507] and
inhibition of respiration via targets such as type II NADH-quinone oxidoreduc-
tases (NDH-2) [508]. The interaction of polymyxins with the Gram-negative OM
LPS has two main implications: first, that susceptibility to this class of com-
pound can be decreased by restructuring LPS to reduce its negative charge (dis-
cussed below) and, second, that it limits the spectrum of polymyxins to some but
not all Gram-negative pathogens. This includes the important Gram-negative
ESKAPE pathogens E. coli, K. pneumoniae, P. aeruginosa, and A. baumannii.
Polymyxins were reintroduced into the clinic in the early 2000s as a last-line
therapeutic option to address the emergence of MDR and XDR in these patho-
gens. Not long after this reintroduction into clinical practice, emergence of colis-
tin-resistant strains began to increase, prompting concern about their ongoing
therapeutic usefulness, especially considering that no other options may exist in
scenarios where colistin is being used. The main mechanism by which Gram-
negative pathogens that are susceptible to polymyxins develop resistance is via
alterations in their lipopolysaccharide that reduce its net negative charge, thereby
reducing uptake of polymyxins. This can occur through selection of chromo-
somal mutations in genes involved in regulating LPS remodeling or by horizontal
acquisition of plasmids harboring genes mediating this process. Some Gram-
negative pathogens (e.g., Burkholderia cepacia, Proteus mirabilis, Serratia
marcescens) are not susceptible to polymyxins (intrinsic resistance), because
their LPS always possesses such modifications. These mechanisms are discussed
in the sections below.
4 Resistance of Gram-negative Bacilli to Antimicrobials 125
Gram-negative bacteria have a broad ability to remodel their OM. The best under-
stood and most widespread mechanism that decreases susceptibility to polymyxins
utilizes this capacity by modification of LPS to reduce its negative charge and, con-
sequently, the initial binding step of cationic polymyxin to the cell surface. This is
highly complex and varies among different strains but in general occurs in
Enterobacteriaceae and P. aeruginosa via masking the negatively charged 4′-phos-
phates on lipid A, and to a lesser extent the 1 position phosphate or the 3-deoxy-d-
manno-octulosonic acid (KDO), by addition of 4-amino-4-deoxy-l-arabinose
(l-Ara4N) [509]. Synthesis and transfer of l-Ara4N are mediated by the products
of the arn locus (e.g., arnBCADTEFpmrE in P. aeruginosa, also known as pmrHFI-
JKLME). This is perhaps the most common mechanism of reducing the negative
charge of LPS described in these organisms. Enterobacteriaceae can also add phos-
phoethanolamine (pEtN) (mainly to the 1 position but also to other locations such
as 4′-lipid A, KDO, or core oligosaccharide) via transferases such as PmrC. The
contribution of pEtN to resistance appears to be smaller than that of l-Ara4N, but
both clearly play a role, and both decorations can occur together. Recently, P. aeru-
ginosa has also been shown to be able to modify its LPS with pEtN when zinc is
present, under the regulation of the ColRS two-component regulatory system [510].
A. baumannii lacks an arn locus and therefore cannot carry out L-Ara4N modifica-
tion but can undergo pEtN modification [526] or, as recently described, galactos-
amine modification [511]. Regulatory control of these modifications is highly
complex and is often mediated by interrelated networks of two-component regula-
tory systems (TCSs). The PmrAB regulator pair controls l-Ara4 and/or pEtN modi-
fication and is widespread in Gram-negative pathogens. Similarly the PhoPQ
system, also present in several organisms such as S. enterica, E. coli, K. pneu-
moniae, and P. aeruginosa (but absent in A. baumannii), is important for control of
LPS modification and can be interconnected with the PmrAB system. For example,
these two regulatory systems are interconnected in S. enterica and E. coli via the
PmrD protein [512]. These systems can upregulate LPS modification (e.g., upregu-
late expression of the arn locus) in response to certain conditions such as magne-
sium limitation or exposure to polymyxin or other cationic peptides [513–515].
Although there is likely some adaptive change in susceptibility to polymyxins medi-
ated by these systems upon drug exposure, or by exposure to cationic peptides in the
host, resistance is generally mutationally acquired, via selection of mutations in the
genes encoding these regulators, which leads to strong constitutive upregulation of
LPS modification. Some mutations in these regulatory genes may also lead to stron-
ger inducibility of the systems by the polymyxin [516]. Regulatory mutations can
be selected in vitro and have also been associated with resistance in the clinical set-
ting, including mutations found in pmrA/pmrB in K. pneumoniae [517–520], P.
aeruginosa [516, 521, 522], and A. baumannii [523–527] and phoP-phoQ in K.
pneumoniae [528–530] and P. aeruginosa [464, 522, 531]. A. baumannii lacks both
phoPQ and an arn locus, so pmrB mutations are frequently found in this pathogen,
126 C. R. Dean et al.
and these mutants will have pEtN modification through activation of the pmrC
transferase gene located in the pmrABC locus [526]. However, mutations in pmrAB
are not always found in colistin-resistant A. baumannii clinical isolates suggesting
that mutations elsewhere on the chromosome can upregulate pmrABC [532].
The importance of the PmrAB and PhoPQ systems in controlling OM remodel-
ing and resistance to polymyxins is well established; however a full understanding
of these phenomena is still forthcoming. Recently in the case of K. pneumoniae,
mutation of mgrB, which encodes a negative feedback regulator of the PhoPQ two-
component system, was revealed as an important mediator of LPS modification and
colistin resistance [533, 534]. Loss of MgrB function leads to constitutive activation
of PhoPQ and LPS modification. This mechanism appears to be relatively wide-
spread in K. pneumoniae clinical isolates [529, 535, 536] and can be mediated by
insertion of genetic elements that may also carry other resistance genes such as
β-lactamases [537]. The fact that colistin resistance can arise from any loss-of-
function mutation of mgrB likely explains its relatively high prevelance among
colistin-resistant K. pneumoniae isolates. Several additional TCSs involved in poly-
myxin resistance have been characterized more recently including the CrrAB (colis-
tin resistance regulon) in K. pneumoniae [530]. Changes in CrrB result in
upregulation of the PmrAB system via CrrR, which then upregulates arn genes and
pmrC, leading to LPS modification and colistin resistance [538]. However, not all
K. pneumoniae harbor the crrAB genes. In P. aeruginosa, three additional TCSs also
known to be involved in modulating susceptibility to polymyxins have been
described: ParRS [539], ColRS [540], and CprRS [541]. ParRS and CprRS partici-
pate in adaptive resistance to polymyxins by upregulating expression from the arn
locus upon sensing polymyxins or other cationic peptides. ColRS and CprRS are
required for high-level polymyxin resistance resulting from mutations in phoPQ
[540]. These interactions appear complex, and mutational analysis also suggested
that additional factors beyond l-Ara4N modification of lipid A could be involved in
resistance in P. aeruginosa, but this remains to be fully elucidated [540]. Mutations
in the parRS genes were subsequently shown to reduce susceptibility to multiple
classes of antibiotic due to coordinately upregulating expression from the mexXY
efflux pump genes and arn and downregulating expression of the oprD porin gene
[75]. For additional details on polymyxin resistance mechanisms, see Jeannot et al.
[542]. Another aspect of resistance relevant to colistin is the phenomenon of hetero-
resistance, which refers to the presence of a substantial stable resistant subpopula-
tion in cultures of isolates that may score as susceptible to an antibiotic by standard
susceptibility testing. Colistin heteroresistance has been described mainly in K.
pneumoniae [528, 543, 544] and A. baumannii [545], and resistant subpopulations
can harbor a range of resistance mutations [543]. Heteroresistant isolates can be
recovered from patients with no prior treatment with polymyxins, and it is expected
that the use of colistin could rapidly enrich for the resistant subpopulation leading
to clinical failures. Heteroresistance can be missed by standard susceptibility tests,
suggesting that the rates of, and potential for, selecting colistin resistance in the
clinic may be underestimated. For more information on heteroresistance, see Chap.
9 in this volume. Finally, as mentioned above, a number of Gram-negative bacteria
4 Resistance of Gram-negative Bacilli to Antimicrobials 127
Synthesis of LPS (in particular the lipid A portion) and assembly of the LPS-
containing OM are essential for the growth and/or viability of most Gram-negative
pathogens, but there are a few exceptions to this. Neisseria meningitidis, Moraxella
catarrhalis [586, 587] and a subset of A. baumannii have been shown to tolerate
loss of LPS biosynthesis. Indeed, this was uncovered in the case of A. baumannii
during in vitro studies of colistin resistance, where mutations in genes involved in
lipid A biosynthesis were directly selected from A. baumannii strain ATCC 19606
or other strains on polymyxin-containing medium [112, 546]. Point mutations were
initially identified in the lpxA, lpxC, or lpxD genes [112] that encode enzymes cata-
lyzing the first three steps of lipid A biosynthesis [25]. A follow-up experiment
identified mutants where lpxA or lpxC were inactivated by insertion sequence
ISAb11 [546], and subsequently an engineered mutant deleted for lpxC was reported
(described in [547]), confirming that lpxA and lpxC are dispensable in A. baumannii
ATCC 19606, at least under laboratory growth conditions. These mutants lack lipid
A, the target of the initial interaction with polymyxins, and as such are highly resis-
tant to polymyxins but are also highly susceptible to a range of other antibiotics due
to loss of the protective lipid A-containing OM [112]. To date, it appears that this
mechanism may be confined to mutations in genes encoding enzymes occurring
early in the lipid A biosynthesis pathway since inactivation of steps occurring later
in the lipid A biosynthetic pathway (e.g., LpxH [548] or LpxK [549]) causes toxic
accumulation of lipid A synthetic pathway intermediates and is not tolerated.
Furthermore, this mechanism does not apply across all A. baumannii, as only a
subset appears to tolerate loss of lipid A biosynthesis (e.g., ATCC 19606). The rea-
sons for this are not fully understood, but a recent study showed potentially com-
pensatory transcriptomic changes in response to loss of lpxA [550], whereas others
showed that expression of penicillin-binding protein (PBP)-1A rendered lipid A
loss lethal in strains that could otherwise tolerate lipid A loss and that cells lacking
both PBP 1A and lipid A had increased expression of lipoproteins on their surface
that may compensate for lipid A loss [114]. Although lipid A loss and colistin resis-
tance can be readily selected in vitro, it stimulates debate about its relevance in the
clinic, both in terms of colistin resistance and, as discussed above in Sect. 4.2.1.5,
with respect to the evaluation of novel antibacterial targets within the lipid A bio-
synthetic pathway (e.g., LpxC). This stems from the notion of whether A. bauman-
nii lacking lipid A (LPS) can survive during infection and therefore could be selected
during colistin treatment.
Since the Gram-negative OM provides protection from the host immune system,
it is generally thought that loss of lipid A (OM) would render the cells unfit in the
128 C. R. Dean et al.
Mutations mediating resistance to colistin can occur fairly rapidly in some Gram-
negative pathogens as described above, but there were initially no reports of hori-
zontal transfer of mobile elements carrying genes mediating LPS modification and
colistin resistance. This changed in 2015 with reporting of the plasmid-borne mcr-1
gene encoding a pEtN transferase in E. coli strains in China and its distribution in
strains isolated from raw meat, animals (pigs), and humans [556]. It was quickly
established that: plasmid-borne mcr-1 was widespread in many regions of the world;
occurred in isolates from food animals, meat and vegetables, the environment, and
humans; was found mainly in E. coli but also occurred in other bacteria; and was
detected in isolate collections dating back to the 1980s [557, 558]. The first identi-
fication of the mcr-1 gene in E. coli from a patient in the United States was reported
in 2016 [559]. Additional mcr-1.2 [560], mcr-2 [561], and mcr-3 [562] variants have
now also been identified. The specific impact of the mcr-1 gene in all four Gram-
negative ESKAPE pathogens has very recently been reported. Mcr-1 mediates pEtN
4 Resistance of Gram-negative Bacilli to Antimicrobials 129
Current efforts in the area of novel polymyxins are aimed at the design of non-
antibacterial polymyxin analogs for use as potentiators of currently used antibiotics
or the design of new antibacterial analogs with reduced toxicity allowing for a
higher therapeutic index or with increased antibacterial activity against emerging
polymyxin-resistant isolates. Several of these efforts exploit the earlier finding that
130 C. R. Dean et al.
Fig. 4.9 Chemical structure of polymyxin B. Acyl chain depicted in blue is a key determinant of
antibacterial activity but not outer membrane disruption activity
the N-terminal acyl chain of polymyxin B (Fig. 4.9) is involved in both antibacterial
activity and toxicity.
A derivative of polymyxin B, polymyxin B nonapetide (PMBN), lacks this moi-
ety and is less toxic and less potent as an antibacterial but retains the ability to
interact with the bacterial OM and permeabilize cells. PMBN itself has been the
subject of much interest and research over the years as a possible potentiating mol-
ecule for use in combination with other antibiotics, but the potential for unaccept-
able residual toxicity still exists. The number of positive charges on polymyxin has
also been associated with toxicity. Northern Antibiotics/Spero has exploited this to
design a PMBN derivative (SPR741) that contains an N-acetyl-threonine-d-serine
side chain, thereby reducing the number of positive charges from five to three rela-
tive to PMBN [121]. This molecule does not have significant antibacterial activity
and is reported to be less toxic but retains antibiotic potentiation activity in E. coli,
K. pneumoniae, and A. baumannii, although it does not potentiate in P. aeruginosa
[574–576]. As of this writing, SPR741 has entered Phase I clinical trials. Cubist
Pharmaceuticals (now Merck) have designed a polymyxin decapeptide derivative
containing a halo-aryl moiety at its N-terminus (CB-182804) to pursue a reduction
in toxicity [577]. CB-182804 exhibited slightly lower antibacterial potency relative
to polymyxin B but was efficacious in animal models of infection and is reported to
have reduced toxicity. CB-182804 entered Phase I clinical trials but appears to be
discontinued. Along the same lines, Pfizer reported a series of analogs replacing the
N-terminal acyl chain with biaryl moieties and substituting the diamino-butyrate
moiety at amino acid 3 with diamino-propionate. One of these, named 5 X, had
slightly improved antibacterial activity and indications of reduced toxicity, but
based on studies in dogs, the therapeutic index was not significantly better than
polymyxin B [578]. Researchers at Monash University are exploring novel poly-
myxin lipopeptides to define structure activity relationships for gaining activity
against colistin-resistant isolates [579] and have presented data on other less toxic
polymyxin derivatives in conjunction with the Medicines Company [580]. Cantab
4 Resistance of Gram-negative Bacilli to Antimicrobials 131
has also reported on the piperazine derivative that showed reduced cytotoxicity and
improved in vivo efficacy over polymyxin B in A. baumannii and P. aeruginosa
lung infection models. Additional details and chemical structures for these and
other novel polymyxins can be found in the review by Brown and Dawson [581].
Finally, there is renewed interest in the octapeptin natural products which also inter-
act with and traverse the Gram-negative OM but do so via a different mechanism
and therefore may be active against polymyxin-resistant strains [582–584].
The discovery of antibiotics, along with vaccines and improved concepts in hygiene,
could be considered the greatest achievement in healthcare-related science in his-
tory. Unfortunately, decades of antibiotic use and perhaps misuse, in both medicine
and agriculture, have enriched for resistant bacteria in the clinical setting, eroding
the effectiveness of the antibiotics upon which we have relied and setting the stage
for a potentially very different reality in medicine from what most of us had grown
accustomed to. This is especially unfortunate since so much of medical practice, for
example, surgery, has relied on antibiotics for success. As can be seen from the
above discussions, antibacterial resistance is complex and multifactorial. However
there are key mechanisms that affect susceptibility to certain classes, such as
β-lactamases for β-lactams and AMEs for aminoglycosides, which may provide
specific strategies for next-generation versions of these antibiotics that address
those mechanisms. It is clear that no effort should be spared on these approaches for
the near term and that new agents directed at previously unexploited novel targets
should be aggressively pursued where they show promise. Hopefully the new aware-
ness of the issue of antimicrobial resistance, and the various incentivizing efforts
spawned from this, will be successful in moving us in the right direction to address
this threat. Finally, even if new agents come along in the near term, it is imperative
that complacency in antibiotic discovery never again sets in. New agents will likely
be the last line of defense, and as such, when resistance to them emerges, the overall
issue of untreatable infections would again be upon us.
Major Points
• Gram-negative pathogens have a unique additional asymmetric outer membrane
(OM). This membrane establishes a significant permeability barrier (reduces
influx) to toxic molecules including antibiotics. Mutations decreasing compound
permeability can be selected under antibiotic exposure.
• Gram-negative pathogens have unique RND family efflux pumps that extrude
most antibiotics and other toxic molecules; these work together with the OM to
reduce intracellular compound accumulation. Upregulation of efflux pump
expression, or changes in compound specificity, can be selected under antibiotic
exposure.
132 C. R. Dean et al.
References
1. Boucher HW, Talbot GH, Bradley JS, Edwards JE, Gilbert D, Rice LB, et al. Bad bugs, no
drugs: no ESKAPE! An update from the Infectious Diseases Society of America. Clin Infect
Dis. 2009;48(1):1–12.
2. Turnidge J, Paterson DL. Setting and revising antibacterial susceptibility breakpoints. Clin
Microbiol Rev. 2007;20(3):391–408.
3. Li XZ, Plesiat P, Nikaido H. The challenge of efflux-mediated antibiotic resistance in Gram-
negative bacteria. Clin Microbiol Rev. 2015;28(2):337–418.
4. Yu EW, Aires JR, Nikaido H. AcrB multidrug efflux pump of Escherichia coli: compos-
ite substrate-binding cavity of exceptional flexibility generates its extremely wide substrate
specificity. J Bacteriol. 2003;185(19):5657–64.
5. Yu EW, McDermott G, Zgurskaya HI, Nikaido H, Koshland DE Jr. Structural basis of multiple
drug-binding capacity of the AcrB multidrug efflux pump. Science. 2003;300(5621):976–80.
6. Tal N, Schuldiner S. A coordinated network of transporters with overlapping specificities
provides a robust survival strategy. Proc Natl Acad Sci U S A. 2009;106(22):9051–6.
7. Kumar A, Schweizer HP. Bacterial resistance to antibiotics: active efflux and reduced uptake.
Adv Drug Deliv Rev. 2005;57(10):1486–513.
8. Winsor GL, Griffiths EJ, Lo R, Dhillon BK, Shay JA, Brinkman FS. Enhanced annotations
and features for comparing thousands of Pseudomonas genomes in the Pseudomonas genome
database. Nucleic Acids Res. 2016;44(D1):D646–53.
9. Du D, Wang Z, James NR, Voss JE, Klimont E, Ohene-Agyei T, et al. Structure of the AcrAB-
TolC multidrug efflux pump. Nature. 2014;509(7501):512–5.
10. Tamura N, Murakami S, Oyama Y, Ishiguro M, Yamaguchi A. Direct interaction of multidrug
efflux transporter AcrB and outer membrane channel TolC detected via site-directed disulfide
cross-linking. Biochemistry. 2005;44(33):11115–21.
11. Kim JS, Jeong H, Song S, Kim HY, Lee K, Hyun J, et al. Structure of the tripartite mul-
tidrug efflux pump AcrAB-TolC suggests an alternative assembly mode. Mol Cells.
2015;38(2):180–6.
12. Daury L, Orange F, Taveau JC, Verchere A, Monlezun L, Gounou C, et al. Tripartite assembly
of RND multidrug efflux pumps. Nat Commun. 2016;7:10731.
13. Murakami S, Nakashima R, Yamashita E, Matsumoto T, Yamaguchi A. Crystal struc-
tures of a multidrug transporter reveal a functionally rotating mechanism. Nature.
2006;443(7108):173–9.
14. Seeger MA, Schiefner A, Eicher T, Verrey F, Diederichs K, Pos KM. Structural asymmetry of
AcrB trimer suggests a peristaltic pump mechanism. Science. 2006;313(5791):1295–8.
15. Sennhauser G, Amstutz P, Briand C, Storchenegger O, Grutter MG. Drug export pathway of
multidrug exporter AcrB revealed by DARPin inhibitors. PLoS Biol. 2007;5(1):e7.
16. Ruggerone P, Vargiu AV, Collu F, Fischer N, Kandt C. Molecular dynamics computer simula-
tions of multidrug RND efflux pumps. Comput Struct Biotechnol J. 2013;5:e201302008.
4 Resistance of Gram-negative Bacilli to Antimicrobials 133
17. Schulz R, Vargiu AV, Collu F, Kleinekathofer U, Ruggerone P. Functional rotation of the
transporter AcrB: insights into drug extrusion from simulations. PLoS Comput Biol.
2010;6(6):e1000806.
18. Li XZ, Ma D, Livermore DM, Nikaido H. Role of efflux pump(s) in intrinsic resistance
of Pseudomonas aeruginosa: active efflux as a contributing factor to β-lactam resistance.
Antimicrob Agents Chemother. 1994;38(8):1742–52.
19. Nikaido H, Basina M, Nguyen V, Rosenberg EY. Multidrug efflux pump AcrAB of
Salmonella typhimurium excretes only those β-lactam antibiotics containing lipophilic side
chains. J Bacteriol. 1998;180(17):4686–92.
20. Li XZ, Zhang L, Srikumar R, Poole K. Β-lactamase inhibitors are substrates for the
multidrug efflux pumps of Pseudomonas aeruginosa. Antimicrob Agents Chemother.
1998;42(2):399–403.
21. Kobayashi N, Tamura N, van Veen HW, Yamaguchi A, Murakami S. β-Lactam selectivity of
multidrug transporters AcrB and AcrD resides in the proximal binding pocket. J Biol Chem.
2014;289(15):10680–90.
22. Elkins CA, Nikaido H. Substrate specificity of the RND-type multidrug efflux pumps AcrB
and AcrD of Escherichia coli is determined predominantly by two large periplasmic loops.
J Bacteriol. 2002;184(23):6490–8.
23. Eda S, Maseda H, Nakae T. An elegant means of self-protection in gram-negative bacte-
ria by recognizing and extruding xenobiotics from the periplasmic space. J Biol Chem.
2003;278(4):2085–8.
24. Lee A, Mao W, Warren MS, Mistry A, Hoshino K, Okumura R, et al. Interplay between efflux
pumps may provide either additive or multiplicative effects on drug resistance. J Bacteriol.
2000;182(11):3142–50.
25. Raetz CR, Whitfield C. Lipopolysaccharide endotoxins. Annu Rev Biochem.
2002;71:635–700.
26. Walsh AG, Matewish MJ, Burrows LL, Monteiro MA, Perry MB, Lam JS. Lipopolysaccharide
core phosphates are required for viability and intrinsic drug resistance in Pseudomonas
aeruginosa. Mol Microbiol. 2000;35(4):718–27.
27. Pages JM, James CE, Winterhalter M. The porin and the permeating antibiotic: a selective
diffusion barrier in Gram-negative bacteria. Nat Rev Microbiol. 2008;6(12):893–903.
28. Nikaido H. Porins and specific diffusion channels in bacterial outer membranes. J Biol Chem.
1994;269(6):3905–8.
29. Sugawara E, Nagano K, Nikaido H. Alternative folding pathways of the major porin OprF of
Pseudomonas aeruginosa. FEBS J. 2012;279(6):910–8.
30. Hancock RE, Woodruff WA. Roles of porin and β-lactamase in β-lactam resistance of
Pseudomonas aeruginosa. Rev Infect Dis. 1988;10(4):770–5.
31. Isabella VM, Campbell AJ, Manchester J, Sylvester M, Nayar AS, Ferguson KE, et al.
Toward the rational design of carbapenem uptake in Pseudomonas aeruginosa. Chem Biol.
2015;22(4):535–47.
32. Faraldo-Gomez JD, Sansom MS. Acquisition of siderophores in gram-negative bacteria. Nat
Rev Mol Cell Biol. 2003;4(2):105–16.
33. Kadner RJ. Vitamin B12 transport in Escherichia coli: energy coupling between membranes.
Mol Microbiol. 1990;4(12):2027–33.
34. Nikaido H. The role of outer membrane and efflux pumps in the resistance of gram-negative
bacteria. Can we improve drug access? Drug Resist Updat. 1998;1(2):93–8.
35. Nikaido H. Molecular basis of bacterial outer membrane permeability revisited. Microbiol
Mol Biol Rev. 2003;67(4):593–656.
36. Lim SP, Nikaido H. Kinetic parameters of efflux of penicillins by the multidrug efflux trans-
porter AcrAB-TolC of Escherichia coli. Antimicrob Agents Chemother. 2010;54(5):1800–6.
37. Mazzariol A, Cornaglia G, Nikaido H. Contributions of the AmpC β-lactamase and the
AcrAB multidrug efflux system in intrinsic resistance of Escherichia coli K-12 to β-lactams.
Antimicrob Agents Chemother. 2000;44(5):1387–90.
134 C. R. Dean et al.
38. Li XZ, Zhang L, Poole K. Interplay between the MexA-MexB-OprM multidrug efflux system
and the outer membrane barrier in the multiple antibiotic resistance of Pseudomonas aerugi-
nosa. J Antimicrob Chemother. 2000;45(4):433–6.
39. Krishnamoorthy G, Wolloscheck D, Weeks JW, Croft C, Rybenkov VV, Zgurskaya
HI. Breaking the permeability barrier of Escherichia coli by controlled hyperporination of
the outer membrane. Antimicrob Agents Chemother. 2016;60(12):7372–81.
40. Dean CR, Visalli MA, Projan SJ, Sum PE, Bradford PA. Efflux-mediated resistance to tige-
cycline (GAR-936) in Pseudomonas aeruginosa PAO1. Antimicrob Agents Chemother.
2003;47(3):972–8.
41. Visalli MA, Murphy E, Projan SJ, Bradford PA. AcrAB multidrug efflux pump is associ-
ated with reduced levels of susceptibility to tigecycline (GAR-936) in Proteus mirabilis.
Antimicrob Agents Chemother. 2003;47(2):665–9.
42. Ruzin A, Visalli MA, Keeney D, Bradford PA. Influence of transcriptional activator RamA
on expression of multidrug efflux pump AcrAB and tigecycline susceptibility in Klebsiella
pneumoniae. Antimicrob Agents Chemother. 2005;49(3):1017–22.
43. Keeney D, Ruzin A, McAleese F, Murphy E, Bradford PA. MarA-mediated overexpression of
the AcrAB efflux pump results in decreased susceptibility to tigecycline in Escherichia coli.
J Antimicrob Chemother. 2008;61(1):46–53.
44. Deng M, Zhu MH, Li JJ, Bi S, Sheng ZK, Hu FS, et al. Molecular epidemiology and mecha-
nisms of tigecycline resistance in clinical isolates of Acinetobacter baumannii from a Chinese
university hospital. Antimicrob Agents Chemother. 2014;58(1):297–303.
45. Li X, Zolli-Juran M, Cechetto JD, Daigle DM, Wright GD, Brown ED. Multicopy suppres-
sors for novel antibacterial compounds reveal targets and drug efflux susceptibility. Chem
Biol. 2004;11(10):1423–30.
46. Brown DG, May-Dracka TL, Gagnon MM, Tommasi R. Trends and exceptions of physical
properties on antibacterial activity for Gram-positive and Gram-negative pathogens. J Med
Chem. 2014;57(23):10144–61.
47. Dean CR, Narayan S, Daigle DM, Dzink-Fox JL, Puyang X, Bracken KR, et al. Role of
the AcrAB-TolC efflux pump in determining susceptibility of Haemophilus influen-
zae to the novel peptide deformylase inhibitor LBM415. Antimicrob Agents Chemother.
2005;49(8):3129–35.
48. Caughlan RE, Jones AK, Delucia AM, Woods AL, Xie L, Ma B, et al. Mechanisms decreas-
ing in vitro susceptibility to the LpxC inhibitor CHIR-090 in the gram-negative pathogen
Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2012;56(1):17–27.
49. Jones AK, Woods AL, Takeoka KT, Shen X, Wei JR, Caughlan RE, et al. Determinants of
antibacterial spectrum and resistance potential of the elongation factor G inhibitor argyrin B
in key gram-negative pathogens. Antimicrob Agents Chemother. 2017;61(4):e02400–16.
50. Caughlan RE, Sriram S, Daigle DM, Woods AL, Buco J, Peterson RL, et al. Fmt bypass in
Pseudomonas aeruginosa causes induction of MexXY efflux pump expression. Antimicrob
Agents Chemother. 2009;53(12):5015–21.
51. Balibar CJ, Iwanowicz D, Dean CR. Elongation factor P is dispensable in Escherichia coli
and Pseudomonas aeruginosa. Curr Microbiol. 2013;67(3):293–9.
52. Morita Y, Gilmour C, Metcalf D, Poole K. Translational control of the antibiotic inducibility
of the PA5471 gene required for mexXY multidrug efflux gene expression in Pseudomonas
aeruginosa. J Bacteriol. 2009;191(15):4966–75.
53. Morita Y, Sobel ML, Poole K. Antibiotic inducibility of the MexXY multidrug efflux system
of Pseudomonas aeruginosa: involvement of the antibiotic-inducible PA5471 gene product.
J Bacteriol. 2006;188(5):1847–55.
54. Chen W, Wang D, Zhou W, Sang H, Liu X, Ge Z, et al. Novobiocin binding to NalD induces
the expression of the MexAB-OprM pump in Pseudomonas aeruginosa. Mol Microbiol.
2016;100(5):749–58.
55. Srikumar R, Paul CJ, Poole K. Influence of mutations in the mexR repressor gene on expres-
sion of the MexA-MexB-oprM multidrug efflux system of Pseudomonas aeruginosa.
J Bacteriol. 2000;182(5):1410–4.
4 Resistance of Gram-negative Bacilli to Antimicrobials 135
56. Daigle DM, Cao L, Fraud S, Wilke MS, Pacey A, Klinoski R, et al. Protein modulator of multi-
drug efflux gene expression in Pseudomonas aeruginosa. J Bacteriol. 2007;189(15):5441–51.
57. Morita Y, Cao L, Gould VC, Avison MB, Poole K. nalD encodes a second repressor of
the mexAB-oprM multidrug efflux operon of Pseudomonas aeruginosa. J Bacteriol.
2006;188(24):8649–54.
58. Fetar H, Gilmour C, Klinoski R, Daigle DM, Dean CR, Poole K. mexEF-oprN multidrug
efflux operon of Pseudomonas aeruginosa: regulation by the MexT activator in response to
nitrosative stress and chloramphenicol. Antimicrob Agents Chemother. 2011;55(2):508–14.
59. Sobel ML, Neshat S, Poole K. Mutations in PA2491 (mexS) promote MexT-dependent
mexEF-oprN expression and multidrug resistance in a clinical strain of Pseudomonas aeru-
ginosa. J Bacteriol. 2005;187(4):1246–53.
60. Baranova N, Nikaido H. The baeSR two-component regulatory system activates transcription
of the yegMNOB (mdtABCD) transporter gene cluster in Escherichia coli and increases its
resistance to novobiocin and deoxycholate. J Bacteriol. 2002;184(15):4168–76.
61. Lin MF, Lin YY, Yeh HW, Lan CY. Role of the BaeSR two-component system in the regula-
tion of Acinetobacter baumannii adeAB genes and its correlation with tigecycline suscepti-
bility. BMC Microbiol. 2014;14:119.
62. Rosner JL, Martin RG. Reduction of cellular stress by TolC-dependent efflux pumps in
Escherichia coli indicated by BaeSR and CpxARP activation of spy in efflux mutants.
J Bacteriol. 2013;195(5):1042–50.
63. Tian ZX, Yi XX, Cho A, O”Gara F, Wang YP. CpxR activates MexAB-OprM efflux pump
expression and enhances antibiotic resistance in both laboratory and clinical nalB-type iso-
lates of Pseudomonas aeruginosa. PLoS Pathog. 2016;12(10):e1005932.
64. Adewoye L, Sutherland A, Srikumar R, Poole K. The mexR repressor of the mexAB-oprM
multidrug efflux operon in Pseudomonas aeruginosa: characterization of mutations compro-
mising activity. J Bacteriol. 2002;184(15):4308–12.
65. Sobel ML, Hocquet D, Cao L, Plesiat P, Poole K. Mutations in PA3574 (nalD) lead to
increased MexAB-OprM expression and multidrug resistance in laboratory and clinical iso-
lates of Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2005;49(5):1782–6.
66. Cao L, Srikumar R, Poole K. MexAB-OprM hyperexpression in NalC-type multidrug-
resistant Pseudomonas aeruginosa: identification and characterization of the nalC gene
encoding a repressor of PA3720-PA3719. Mol Microbiol. 2004;53(5):1423–36.
67. Wang H, Dzink-Fox JL, Chen M, Levy SB. Genetic characterization of highly fluoroquinolone-
resistant clinical Escherichia coli strains from China: role of acrR mutations. Antimicrob
Agents Chemother. 2001;45(5):1515–21.
68. Purssell A, Poole K. Functional characterization of the NfxB repressor of the mexCD-
oprJ multidrug efflux operon of Pseudomonas aeruginosa. Microbiology. 2013;159(Pt
10):2058–73.
69. Poole K. Efflux-mediated antimicrobial resistance. J Antimicrob Chemother.
2005;56(1):20–51.
70. Blair JM, Webber MA, Baylay AJ, Ogbolu DO, Piddock LJ. Molecular mechanisms of anti-
biotic resistance. Nat Rev Microbiol. 2015;13(1):42–51.
71. Poulou A, Voulgari E, Vrioni G, Koumaki V, Xidopoulos G, Chatzipantazi V, et al. Outbreak
caused by an ertapenem-resistant, CTX-M-15-producing Klebsiella pneumoniae sequence
type 101 clone carrying an OmpK36 porin variant. J Clin Microbiol. 2013;51(10):3176–82.
72. Wozniak A, Villagra NA, Undabarrena A, Gallardo N, Keller N, Moraga M, et al. Porin alter-
ations present in non-carbapenemase-producing Enterobacteriaceae with high and intermedi-
ate levels of carbapenem resistance in Chile. J Med Microbiol. 2012;61(Pt 9):1270–9.
73. Lavigne JP, Sotto A, Nicolas-Chanoine MH, Bouziges N, Pages JM, Davin-Regli A. An adap-
tive response of Enterobacter aerogenes to imipenem: regulation of porin balance in clinical
isolates. Int J Antimicrob Agents. 2013;41(2):130–6.
74. Bajaj H, Scorciapino MA, Moynie L, Page MG, Naismith JH, Ceccarelli M, et al. Molecular
basis of filtering carbapenems by porins from β-Lactam-resistant clinical strains of
Escherichia coli. J Biol Chem. 2016;291(6):2837–47.
136 C. R. Dean et al.
93. Balibar CJ, Grabowicz M. Mutant alleles of lptD increase the permeability of Pseudomonas
aeruginosa and define determinants of intrinsic resistance to antibiotics. Antimicrob Agents
Chemother. 2016;60(2):845–54.
94. Shen X, Johnson AK, Jones AK, Barnes SW, Walker JR, Ranjitkar S, Woods AL, Six DA,
Dean CR. Genetic characterization of the hypersusceptible Pseudomonas aeruginosa strain
Z61: identification of a defect in LptE. Abstr C-105 Abstr 54th Intersci Conf Antimicrob
Agents Chemother American Society for Microbiology, Washington, DC; 2014.
95. Piizzi G, Parker DT, Peng Y, Dobler M, Patnaik A, Wattanasin S, et al. Design, synthesis, and
properties of a potent inhibitor of Pseudomonas aeruginosa deacetylase LpxC. J Med Chem.
2017;60(12):5002–14.
96. Barb AW, Jiang L, Raetz CR, Zhou P. Structure of the deacetylase LpxC bound to the antibi-
otic CHIR-090: time-dependent inhibition and specificity in ligand binding. Proc Natl Acad
Sci U S A. 2007;104(47):18433–8.
97. Barb AW, McClerren AL, Snehelatha K, Reynolds CM, Zhou P, Raetz CR. Inhibition of lipid
A biosynthesis as the primary mechanism of CHIR-090 antibiotic activity in Escherichia
coli. Biochemist. 2007;46(12):3793–802.
98. Brown MF, Reilly U, Abramite JA, Arcari JT, Oliver R, Barham RA, et al. Potent inhibitors
of LpxC for the treatment of Gram-negative infections. J Med Chem. 2012;55(2):914–23.
99. Hale MR, Hill P, Lahiri S, Miller MD, Ross P, Alm R, et al. Exploring the UDP pocket of
LpxC through amino acid analogs. Bioorg Med Chem Lett. 2013;23(8):2362–7.
100. Liang X, Lee CJ, Chen X, Chung HS, Zeng D, Raetz CR, et al. Syntheses, structures and
antibiotic activities of LpxC inhibitors based on the diacetylene scaffold. Bioorg Med Chem.
2011;19(2):852–60.
101. Liang X, Lee CJ, Zhao J, Toone EJ, Zhou P. Synthesis, structure, and antibiotic activity of
aryl-substituted LpxC inhibitors. J Med Chem. 2013;56(17):6954–66.
102. Mansoor UF, Vitharana D, Reddy PA, Daubaras DL, McNicholas P, Orth P, et al. Design
and synthesis of potent Gram-negative specific LpxC inhibitors. Bioorg Med Chem Lett.
2011;21(4):1155–61.
103. McAllister LA, Montgomery JI, Abramite JA, Reilly U, Brown MF, Chen JM, et al.
Heterocyclic methylsulfone hydroxamic acid LpxC inhibitors as Gram-negative antibacterial
agents. Bioorg Med Chem Lett. 2012;22(22):6832–8.
104. Montgomery JI, Brown MF, Reilly U, Price LM, Abramite JA, Arcari J, et al. Pyridone meth-
ylsulfone hydroxamate LpxC inhibitors for the treatment of serious gram-negative infections.
J Med Chem. 2012;55(4):1662–70.
105. Murphy-Benenato KE, Olivier N, Choy A, Ross PL, Miller MD, Thresher J, et al. Synthesis,
structure, and SAR of tetrahydropyran-Based LpxC Inhibitors. ACS Med Chem Lett.
2014;5(11):1213–8.
106. Titecat M, Liang X, Lee CJ, Charlet A, Hocquet D, Lambert T, et al. High susceptibility
of MDR and XDR Gram-negative pathogens to biphenyl-diacetylene-based difluoromethyl-
allo-threonyl-hydroxamate LpxC inhibitors. J Antimicrob Chemother. 2016;71(10):2874–82.
107. Kasar R, Linsell MS, Aggen JB, Lu Q, Wang D, Church T, et al. Hydroxamic acid derivatives
and their use in the treatment of bacterial infections. Google Patents. 2012.
108. Cigana C, Bernardini F, Facchini M, Alcala-Franco B, Riva C, De Fino I, et al. Efficacy of
the novel antibiotic POL7001 in preclinical models of Pseudomonas aeruginosa pneumonia.
Antimicrob Agents Chemother. 2016;60(8):4991–5000.
109. Kalinin DV, Holl R. Insights into the zinc-dependent deacetylase LpxC: biochemical proper-
ties and inhibitor design. Curr Top Med Chem. 2016;16(21):2379–430.
110. Srinivas N, Jetter P, Ueberbacher BJ, Werneburg M, Zerbe K, Steinmann J, et al.
Peptidomimetic antibiotics target outer-membrane biogenesis in Pseudomonas aeruginosa.
Science. 2010;327(5968):1010–3.
111. Steeghs L, de Cock H, Evers E, Zomer B, Tommassen J, van der Ley P. Outer membrane
composition of a lipopolysaccharide-deficient Neisseria meningitidis mutant. EMBO
J. 2001;20(24):6937–45.
138 C. R. Dean et al.
112. Moffatt JH, Harper M, Harrison P, Hale JD, Vinogradov E, Seemann T, et al. Colistin resis-
tance in Acinetobacter baumannii is mediated by complete loss of lipopolysaccharide pro-
duction. Antimicrob Agents Chemother. 2010;54(12):4971–7.
113. Garcia-Quintanilla M, Carretero-Ledesma M, Moreno-Martinez P, Martin-Pena R, Pachon
J, McConnell MJ. Lipopolysaccharide loss produces partial colistin dependence and collat-
eral sensitivity to azithromycin, rifampicin and vancomycin in Acinetobacter baumannii. Int
J Antimicrob Agents. 2015;46(6):696–702.
114. Boll JM, Crofts AA, Peters K, Cattoir V, Vollmer W, Davies BW, et al. A penicillin-binding
protein inhibits selection of colistin-resistant, lipooligosaccharide-deficient Acinetobacter
baumannii. Proc Natl Acad Sci U S A. 2016;113(41):E6228–E37.
115. Bojkovic J, Richie DL, Six DA, Rath CM, Sawyer WS, Hu Q, et al. Characterization of an
Acinetobacter baumannii lptD deletion strain: permeability defects and response to inhibition
of lipopolysaccharide and fatty acid biosynthesis. J Bacteriol. 2016;198(4):731–41.
116. Garcia-Quintanilla M, Caro-Vega JM, Pulido MR, Moreno-Martinez P, Pachon J, McConnell
MJ. Inhibition of LpxC Increases antibiotic susceptibility in Acinetobacter baumannii.
Antimicrob Agents Chemother. 2016;60(8):5076–9.
117. Lin L, Tan B, Pantapalangkoor P, Ho T, Baquir B, Tomaras A, et al. Inhibition of LpxC pro-
tects mice from resistant Acinetobacter baumannii by modulating inflammation and enhanc-
ing phagocytosis. MBio. 2012;3(5):e00312.
118. Clements JM, Coignard F, Johnson I, Chandler S, Palan S, Waller A, et al. Antibacterial
activities and characterization of novel inhibitors of LpxC. Antimicrob Agents Chemother.
2002;46(6):1793–9.
119. Kimura Y, Matsunaga H, Vaara M. Polymyxin B octapeptide and polymyxin B hepta-
peptide are potent outer membrane permeability-increasing agents. J Antibiot (Tokyo).
1992;45(5):742–9.
120. Vaara M. Agents that increase the permeability of the outer membrane. Microbiol Rev.
1992;56(3):395–411.
121. Corbett D, Wise A, Langley T, Skinner K, Trimby E, Birchall S, et al. Potentiation of anti-
biotic activity by a novel cationic peptide: potency and spectrum of activity of SPR741.
Antimicrob Agents Chemother. 2017;61(8):e00200–17.
122. Murray B, Pillar C, Pucci M, Shinabarger D. Mechanism of action of SPR741, a potentiator
molecule for Gram-negative pathogens. Abstr Saturday-491. New Orleans: Microbe; 2017.
123. Urfer M, Bogdanovic J, Lo Monte F, Moehle K, Zerbe K, Omasits U, et al. A peptidomimetic
antibiotic targets outer membrane proteins and disrupts selectively the outer membrane in
Escherichia coli. J Biol Chem. 2016;291(4):1921–32.
124. DiGiandomenico A, Keller AE, Gao C, Rainey GJ, Warrener P, Camara MM, et al. A mul-
tifunctional bispecific antibody protects against Pseudomonas aeruginosa. Sci Transl Med.
2014;6(262):262ra155.
125. Thanabalasuriar A, Surewaard BG, Willson ME, Neupane AS, Stover CK, Warrener P, et al.
Bispecific antibody targets multiple Pseudomonas aeruginosa evasion mechanisms in the
lung vasculature. J Clin Invest. 2017;127(6):2249–61.
126. Tran QT, Williams S, Farid R, Erdemli G, Pearlstein R. The translocation kinetics of antibiot-
ics through porin OmpC: insights from structure-based solvation mapping using WaterMap.
Proteins. 2013;81(2):291–9.
127. Tran QT, Pearlstein RA, Williams S, Reilly J, Krucker T, Erdemli G. Structure-kinetic rela-
tionship of carbapenem antibacterials permeating through E. coli OmpC porin. Proteins.
2014;82(11):2998–3012.
128. Bajaj H, Acosta Gutierrez S, Bodrenko I, Malloci G, Scorciapino MA, Winterhalter M, et al.
Bacterial outer membrane porins as electrostatic nanosieves: exploring transport rules of
small polar molecules. ACS Nano. 2017;11(6):4598–606.
129. Ghai I, Winterhalter M, Wagner R. Probing transport of charged β-lactamase inhibi-
tors through OmpC, a membrane channel from E. coli. Biochem Biophys Res Commun.
2017;484(1):51–5.
4 Resistance of Gram-negative Bacilli to Antimicrobials 139
130. Giammanco A, Cala C, Fasciana T, Dowzicky MJ. Global assessment of the activity of tige-
cycline against multidrug-resistant Gram-negative pathogens between 2004 and 2014 as part
of the tigecycline evaluation and surveillance trial. mSphere. 2017;2(1):e00310.
131. Lomovskaya O, Watkins WJ. Efflux pumps: their role in antibacterial drug discovery. Curr
Med Chem. 2001;8(14):1699–711.
132. O’Shea R, Moser HE. Physicochemical properties of antibacterial compounds: implications
for drug discovery. J Med Chem. 2008;51(10):2871–8.
133. Reck F, Ehmann DE, Dougherty TJ, Newman JV, Hopkins S, Stone G, et al. Optimization
of physicochemical properties and safety profile of novel bacterial topoisomerase type II
inhibitors (NBTIs) with activity against Pseudomonas aeruginosa. Bioorg Med Chem.
2014;22(19):5392–409.
134. Richter MF, Drown BS, Riley AP, Garcia A, Shirai T, Svec RL, et al. Predictive compound
accumulation rules yield a broad-spectrum antibiotic. Nature. 2017;545(7654):299–304.
135. Hong S, Moritz TJ, Rath CM, Tamrakar P, Lee P, Krucker T, et al. Assessing antibiotic perme-
ability of Gram-negative bacteria via nanofluidics. ACS Nano. 2017;11(7):6959–67.
136. Tian H, Six DA, Krucker T, Leeds JA, Winograd N. Subcellular chemical imaging of
antibiotics in single bacteria using C60-secondary ion mass spectrometry. Anal Chem.
2017;89(9):5050–7.
137. Ito A, Kohira N, Bouchillon SK, West J, Rittenhouse S, Sader HS, et al. In vitro antimicrobial
activity of S-649266, a catechol-substituted siderophore cephalosporin, when tested against
non-fermenting Gram-negative bacteria. J Antimicrob Chemother. 2016;71(3):670–7.
138. Fleming A. On the antibacterial action of cultures of a penicillium, with special reference to
their use in the isolation of B. influenzae. Br J Exp Pathol. 1929;10:226–36.
139. Bush K, Bradford PA. β-Lactams and β-Lactamase Inhibitors: an overview. Cold Spring Harb
Perspect Med. 2016;6(8):a025247.
140. Spratt BG. Penicillin-binding proteins and the future of β-lactam antibiotics. The seventh
fleming lecture. J Gen Microbiol. 1983;129(5):1247–60.
141. Sauvage E, Kerff F, Terrak M, Ayala JA, Charlier P. The penicillin-binding proteins: structure
and role in peptidoglycan biosynthesis. FEMS Microbiol Rev. 2008;32(2):234–58.
142. van Heijenoort J. Peptidoglycan hydrolases of Escherichia coli. Microbiol Mol Biol Rev.
2011;75(4):636–63.
143. Cho H, Uehara T, Bernhardt TG. Β-lactam antibiotics induce a lethal malfunctioning of the
bacterial cell wall synthesis machinery. Cell. 2014;159(6):1300–11.
144. Massova I, Mobashery S. Kinship and diversification of bacterial penicillin-binding proteins
and β-lactamases. Antimicrob Agents Chemother. 1998;42(1):1–17.
145. Bush K, Jacoby GA. Updated functional classification of β-lactamases. Antimicrob Agents
Chemother. 2010;54(3):969–76.
146. Medeiros AA. Evolution and dissemination of β-lactamases accelerated by generations of
β-lactam antibiotics. Clin Infect Dis. 1997;24(Suppl 1):S19–45.
147. Rossolini GM, Docquier JD. New β-lactamases: a paradigm for the rapid response of bacte-
rial evolution in the clinical setting. Future Microbiol. 2006;1(3):295–308.
148. De Pascale G, Wright GD. Antibiotic resistance by enzyme inactivation: from mechanisms to
solutions. Chembiochem. 2010;11(10):1325–34.
149. Datta N, Kontomichalou P. Penicillinase synthesis controlled by infectious R factors in
Enterobacteriaceae. Nature. 1965;208(5007):239–41.
150. Bonomo RA. β-Lactamases: a focus on current challenges. Cold Spring Harb Perspect Med.
2017;7(1):a025239.
151. Knothe H, Shah P, Krcmery V, Antal M, Mitsuhashi S. Transferable resistance to cefotaxime,
cefoxitin, cefamandole and cefuroxime in clinical isolates of Klebsiella pneumoniae and
Serratia marcescens. Infection. 1983;11(6):315–7.
152. Ramphal R, Ambrose PG. Extended-spectrum β-lactamases and clinical outcomes: current
data. Clin Infect Dis. 2006;42(Suppl 4):S164–72.
140 C. R. Dean et al.
194. Zmarlicka MT, Nailor MD, Nicolau DP. Impact of the New Delhi metallo-β-lactamase on
β-lactam antibiotics. Infect Drug Resist. 2015;8:297–309.
195. Borgia S, Lastovetska O, Richardson D, Eshaghi A, Xiong J, Chung C, et al. Outbreak of
carbapenem- resistant Enterobacteriaceae containing blaNDM-1, Ontario, Canada. Clin
Infect Dis. 2012;55(11):e109–17.
196. Sugawara E, Kojima S, Nikaido H. Klebsiella pneumoniae major porins OmpK35 and
OmpK36 allow more efficient diffusion of β-Lactams than their Escherichia coli Homologs
OmpF and OmpC. J Bacteriol. 2016;198(23):3200–8.
197. Adams-Sapper S, Nolen S, Donzelli GF, Lal M, Chen K, Justo da Silva LH, et al. Rapid
induction of high-level carbapenem resistance in heteroresistant KPC-producing Klebsiella
pneumoniae. Antimicrob Agents Chemother. 2015;59(6):3281–9.
198. Castanheira M, Mills JC, Farrell DJ, Jones RN. Mutation-driven β-lactam resistance mecha-
nisms among contemporary ceftazidime-nonsusceptible Pseudomonas aeruginosa isolates
from U.S. hospitals. Antimicrob Agents Chemother. 2014;58(11):6844–50.
199. Moya B, Beceiro A, Cabot G, Juan C, Zamorano L, Alberti S, et al. Pan-β-lactam resis-
tance development in Pseudomonas aeruginosa clinical strains: molecular mechanisms,
penicillin-binding protein profiles, and binding affinities. Antimicrob Agents Chemother.
2012;56(9):4771–8.
200. Bradford PA, Urban C, Mariano N, Projan SJ, Rahal JJ, Bush K. Imipenem resistance in
Klebsiella pneumoniae is associated with the combination of ACT-1, a plasmid-mediated
AmpC β-lactamase, and the foss of an outer membrane protein. Antimicrob Agents
Chemother. 1997;41(3):563–9.
201. Yamachika S, Sugihara C, Kamai Y, Yamashita M. Correlation between penicillin-binding
protein 2 mutations and carbapenem resistance in Escherichia coli. J Med Microbiol.
2013;62(Pt 3):429–36.
202. Alm RA, Johnstone MR, Lahiri SD. Characterization of Escherichia coli NDM isolates
with decreased susceptibility to aztreonam/avibactam: role of a novel insertion in PBP3.
J Antimicrob Chemother. 2015;70(5):1420–8.
203. Zhang Y, Kashikar A, Brown CA, Denys G, Bush K. Unusual Escherichia coli PBP 3
insertion sequence identified from a collection of carbapenem-resistant Enterobacteriaceae
tested In vitro with a combination of ceftazidime-, ceftaroline-, or aztreonam-avibactam.
Antimicrob Agents Chemother. 2017;61(8):e00389.
204. Durand-Reville TF, Guler S, Comita-Prevoir J, Chen B, Bifulco N, Huynh H, et al. ETX2514
is a broad-spectrum β-lactamase inhibitor for the treatment of drug-resistant Gram-negative
bacteria including Acinetobacter baumannii. Nat Microbiol. 2017;2:17104.
205. Reading C, Cole M. Clavulanic acid: a β-lactamase-inhiting β-lactam from Streptomyces
clavuligerus. Antimicrob Agents Chemother. 1977;11(5):852–7.
206. English AR, Retsema JA, Girard AE, Lynch JE, Barth WE. CP-45,899, a β-lactamase inhibi-
tor that extends the antibacterial spectrum of β-lactams: initial bacteriological characteriza-
tion. Antimicrob Agents Chemother. 1978;14(3):414–9.
207. Fisher J, Belasco JG, Charnas RL, Khosla S, Knowles JR. Β-lactamase inactivation by
mechanism-based reagents. Philos Trans R Soc Lond Ser B Biol Sci. 1980;289(1036):309–19.
208. Cole M. Biochemistry and action of clavulanic acid. Scott Med J. 1982;27:S10–6.
209. Brown RP, Aplin RT, Schofield CJ. Inhibition of TEM-2 β-lactamase from Escherichia coli
by clavulanic acid: observation of intermediates by electrospray ionization mass spectrom-
etry. Biochemistry. 1996;35(38):12421–32.
210. Drawz SM, Bonomo RA. Three decades of β-lactamase inhibitors. Clin Microbiol Rev.
2010;23(1):160–201.
211. Sulton D, Pagan-Rodriguez D, Zhou X, Liu Y, Hujer AM, Bethel CR, et al. Clavulanic acid
inactivation of SHV-1 and the inhibitor-resistant S130G SHV-1 β-lactamase. Insights into the
mechanism of inhibition. J Biol Chem. 2005;280(42):35528–36.
212. Weber DA, Sanders CC. Diverse potential of β-lactamase inhibitors to induce class I enzymes.
Antimicrob Agents Chemother. 1990;34(1):156–8.
4 Resistance of Gram-negative Bacilli to Antimicrobials 143
230. Canton R, Morosini MI, de la Maza OM, de la Pedrosa EG. IRT and CMT β-lactamases and
inhibitor resistance. Clin Microbiol Infect. 2008;14(Suppl 1):53–62.
231. Martinez JL, Cercenado E, Rodriguez-Creixems M, Vincente-Perez MF, Delgado-Iribarren
A, Baquero F. Resistance to β-lactam/clavulanate. Lancet. 1987;2(8573):1473.
232. Martinez JL, Vicente MF, Delgado-Iribarren A, Perez-Diaz JC, Baquero F. Small plasmids
are involved in amoxicillin-clavulanate resistance in Escherichia coli. Antimicrob Agents
Chemother. 1989;33(4):595.
233. Lister PD, Gardner VM, Sanders CC. Clavulanate induces expression of the Pseudomonas
aeruginosa AmpC cephalosporinase at physiologically relevant concentrations and antago-
nizes the antibacterial activity of ticarcillin. Antimicrob Agents Chemother. 1999;43(4):882–9.
234. Derbyshire H, Kay G, Evans K, Vaughan C, Kavuri U, Winstanley T. A simple disc diffu-
sion method for detecting AmpC and extended-spectrum β-lactamases in clinical isolates of
Enterobacteriaceae. J Antimicrob Chemother. 2009;63(3):497–501.
235. Yan JJ, Ko WC, Jung YC, Chuang CL, Wu JJ. Emergence of Klebsiella pneumoniae Isolates
producing inducible DHA-1 -lactamase in a university hospital in Taiwan. J Clin Microbiol.
2002;40(9):3121–6.
236. Nikaido H, Pages JM. Broad-specificity efflux pumps and their role in multidrug resistance
of Gram-negative bacteria. FEMS Microbiol Rev. 2012;36(2):340–63.
237. Pages JM, Peslier S, Keating TA, Lavigne JP, Nichols WW. Role of the outer membrane
and porins in susceptibility of β-lactamase-oroducing Enterobacteriaceae to ceftazidime-
avibactam. Antimicrob Agents Chemother. 2015;60(3):1349–59.
238. Lapuebla A, Abdallah M, Olafisoye O, Cortes C, Urban C, Landman D, et al. Activity of imi-
penem with relebactam against Gram-Negative pathogens from New York City. Antimicrob
Agents Chemother. 2015;59(8):5029–31.
239. Winkler ML, Papp-Wallace KM, Hujer AM, Domitrovic TN, Hujer KM, Hurless KN, et al.
Unexpected challenges in treating multidrug-resistant Gram-negative bacteria: resistance to
ceftazidime-avibactam in archived isolates of Pseudomonas aeruginosa. Antimicrob Agents
Chemother. 2015;59(2):1020–9.
240. Livermore DM, Mushtaq S, Barker K, Hope R, Warner M, Woodford N. Characterization of
β-lactamase and porin mutants of Enterobacteriaceae selected with ceftaroline + avibactam
(NXL104). J Antimicrob Chemother. 2012;67(6):1354–8.
241. Blazquez J, Baquero MR, Canton R, Alos I, Baquero F. Characterization of a new TEM-
type β-lactamase resistant to clavulanate, sulbactam, and tazobactam in a clinical isolate of
Escherichia coli. Antimicrob Agents Chemother. 1993;37(10):2059–63.
242. Vedel G, Belaaouaj A, Gilly L, Labia R, Philippon A, Nevot P, et al. Clinical isolates of
Escherichia coli producing TRI β-lactamases: novel TEM-enzymes conferring resistance to
β-lactamase inhibitors. J Antimicrob Chemother. 1992;30(4):449–62.
243. Belaaouaj A, Lapoumeroulie C, Canica MM, Vedel G, Nevot P, Krishnamoorthy R, et al.
Nucleotide sequences of the genes coding for the TEM-like β-lactamases IRT-1 and IRT-2
(formerly called TRI-1 and TRI-2). FEMS Microbiol Lett. 1994;120(1-2):75–80.
244. Lemozy J, Sirot D, Chanal C, Huc C, Labia R, Dabernat H, et al. First characterization
of inhibitor-resistant TEM (IRT) β-lactamases in Klebsiella pneumoniae strains. Antimicrob
Agents Chemother. 1995;39(11):2580–2.
245. Bret L, Chanal C, Sirot D, Labia R, Sirot J. Characterization of an inhibitor-resistant enzyme
IRT-2 derived from TEM-2 β-lactamase produced by Proteus mirabilis strains. J Antimicrob
Chemother. 1996;38(2):183–91.
246. Sirot D, Chanal C, Bonnet R, De Champs C, Bret L. Inhibitor-resistant TEM-33 β-lactamase
in a Shigella sonnei isolate. Antimicrob Agents Chemother. 2001;45(7):2179–80.
247. Hunter JE, Corkill JE, McLennan AG, Fletcher JN, Hart CA. Plasmid encoded β-lactamases
resistant to inhibition by clavulanic acid produced by calf faecal coliforms. Res Vet Sci.
1993;55(3):367–70.
248. Manageiro V, Ferreira E, Cougnoux A, Albuquerque L, Canica M, Bonnet R. Characterization
of the inhibitor-resistant SHV β-lactamase SHV-107 in a clinical Klebsiella pneumoniae
strain coproducing GES-7 enzyme. Antimicrob Agents Chemother. 2012;56(2):1042–6.
4 Resistance of Gram-negative Bacilli to Antimicrobials 145
249. Bonomo RA, Rudin SA, Shlaes DM. Tazobactam is a potent inactivator of selected inhibitor-
resistant class A β-lactamases. FEMS Microbiol Lett. 1997;148(1):59–62.
250. Chaibi EB, Sirot D, Paul G, Labia R. Inhibitor-resistant TEM β-lactamases: phenotypic,
genetic and biochemical characteristics. J Antimicrob Chemother. 1999;43(4):447–58.
251. Pagan-Rodriguez D, Zhou X, Simmons R, Bethel CR, Hujer AM, Helfand MS, et al.
Tazobactam inactivation of SHV-1 and the inhibitor-resistant Ser130 -->Gly SHV-1
β-lactamase: insights into the mechanism of inhibition. J Biol Chem. 2004;279(19):19494–501.
252. Nordmann P, Cuzon G, Naas T. The real threat of Klebsiella pneumoniae carbapenemase-
producing bacteria. Lancet Infect Dis. 2009;9(4):228–36.
253. Papp-Wallace KM, Bethel CR, Distler AM, Kasuboski C, Taracila M, Bonomo RA. Inhibitor
resistance in the KPC-2 β-lactamase, a preeminent property of this class A β-lactamase.
Antimicrob Agents Chemother. 2010;54(2):890–7.
254. Livermore DM, Warner M, Jamrozy D, Mushtaq S, Nichols WW, Mustafa N, et al. In vitro
selection of ceftazidime-avibactam resistance in Enterobacteriaceae with KPC-3 carbapen-
emase. Antimicrob Agents Chemother. 2015;59(9):5324–30.
255. Winkler ML, Papp-Wallace KM, Taracila MA, Bonomo RA. Avibactam and inhibitor-
resistant SHV β-lactamases. Antimicrob Agents Chemother. 2015;59(7):3700–9.
256. Shields RK, Chen L, Cheng S, Chavda KD, Press EG, Snyder A, et al. Emergence of
ceftazidime-avibactam resistance due to plasmid-borne blaKPC-3 mutations during treatment
of carbapenem-resistant Klebsiella pneumoniae infections. Antimicrob Agents Chemother.
2017;61(3):e02097.
257. Lahiri SD, Bradford PA, Nichols WW, Alm RA. Structural and sequence analysis of class
A β-lactamases with respect to avibactam inhibition: impact of Omega-loop variations.
J Antimicrob Chemother. 2016;71(10):2848–55.
258. Drlica K, Malik M. Fluoroquinolones: action and resistance. Curr Top Med Chem.
2003;3(3):249–82.
259. Champoux JJ. DNA topoisomerases: structure, function, and mechanism. Annu Rev
Biochem. 2001;70:369–413.
260. Drlica K, Hiasa H, Kerns R, Malik M, Mustaev A, Zhao X. Quinolones: action and resistance
updated. Curr Top Med Chem. 2009;9(11):981–98.
261. Hiasa H, Yousef DO, Marians KJ. DNA strand cleavage is required for replication
fork arrest by a frozen topoisomerase-quinolone-DNA ternary complex. J Biol Chem.
1996;271(42):26424–9.
262. Willmott CJ, Critchlow SE, Eperon IC, Maxwell A. The complex of DNA gyrase and qui-
nolone drugs with DNA forms a barrier to transcription by RNA polymerase. J Mol Biol.
1994;242(4):351–63.
263. Shea ME, Hiasa H. Interactions between DNA helicases and frozen topoisomerase
IV-quinolone-DNA ternary complexes. J Biol Chem. 1999;274(32):22747–54.
264. Brazas MD, Hancock RE. Ciprofloxacin induction of a susceptibility determinant in
Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2005;49(8):3222–7.
265. Economou V, Gousia P. Agriculture and food animals as a source of antimicrobial-resistant
bacteria. Infect Drug Resist. 2015;8:49–61.
266. Hooper DC, Jacoby GA. Mechanisms of drug resistance: quinolone resistance. Ann N Y
Acad Sci. 2015;1354:12–31.
267. Morais Cabral JH, Jackson AP, Smith CV, Shikotra N, Maxwell A, Liddington RC. Crystal
structure of the breakage-reunion domain of DNA gyrase. Nature. 1997;388(6645):903–6.
268. Yoshida H, Bogaki M, Nakamura M, Nakamura S. Quinolone resistance-determining
region in the DNA gyrase gyrA gene of Escherichia coli. Antimicrob Agents Chemother.
1990;34(6):1271–2.
269. Laponogov I, Veselkov DA, Crevel IM, Pan XS, Fisher LM, Sanderson MR. Structure of an
‘open’ clamp type II topoisomerase-DNA complex provides a mechanism for DNA capture
and transport. Nucleic Acids Res. 2013;41(21):9911–23.
270. Wohlkonig A, Chan PF, Fosberry AP, Homes P, Huang J, Kranz M, et al. Structural basis of
quinolone inhibition of type IIA topoisomerases and target-mediated resistance. Nat Struct
Mol Biol. 2010;17(9):1152–3.
146 C. R. Dean et al.
271. Willmott CJ, Maxwell A. A single point mutation in the DNA gyrase A protein greatly reduces
binding of fluoroquinolones to the gyrase-DNA complex. Antimicrob Agents Chemother.
1993;37(1):126–7.
272. Aldred KJ, Kerns RJ, Osheroff N. Mechanism of quinolone action and resistance.
Biochemistry. 2014;53(10):1565–74.
273. Bagel S, Hullen V, Wiedemann B, Heisig P. Impact of gyrA and parC mutations on quinolone
resistance, doubling time, and supercoiling degree of Escherichia coli. Antimicrob Agents
Chemother. 1999;43(4):868–75.
274. Komp Lindgren P, Marcusson LL, Sandvang D, Frimodt-Moller N, Hughes D. Biological
cost of single and multiple norfloxacin resistance mutations in Escherichia coli implicated in
urinary tract infections. Antimicrob Agents Chemother. 2005;49(6):2343–51.
275. Bruchmann S, Dotsch A, Nouri B, Chaberny IF, Haussler S. Quantitative contributions of tar-
get alteration and decreased drug accumulation to Pseudomonas aeruginosa fluoroquinolone
resistance. Antimicrob Agents Chemother. 2013;57(3):1361–8.
276. Kriengkauykiat J, Porter E, Lomovskaya O, Wong-Beringer A. Use of an efflux pump
inhibitor to determine the prevalence of efflux pump-mediated fluoroquinolone resistance
and multidrug resistance in Pseudomonas aeruginosa. Antimicrob Agents Chemother.
2005;49(2):565–70.
277. Paltansing S, Tengeler AC, Kraakman ME, Claas EC, Bernards AT. Exploring the contribu-
tion of efflux on the resistance to fluoroquinolones in clinical isolates of Escherichia coli.
Microb Drug Resist. 2013;19(6):469–76.
278. Mazzariol A, Zuliani J, Cornaglia G, Rossolini GM, Fontana R. AcrAB Efflux system:
expression and contribution to fluoroquinolone resistance in Klebsiella spp. Antimicrob
Agents Chemother. 2002;46(12):3984–6.
279. Le Thomas I, Couetdic G, Clermont O, Brahimi N, Plesiat P, Bingen E. In vivo selection
of a target/efflux double mutant of Pseudomonas aeruginosa by ciprofloxacin therapy.
J Antimicrob Chemother. 2001;48(4):553–5.
280. Chapman JS, Georgopapadakou NH. Routes of quinolone permeation in Escherichia coli.
Antimicrob Agents Chemother. 1988;32(4):438–42.
281. Chenia HY, Pillay B, Pillay D. Analysis of the mechanisms of fluoroquinolone resistance in
urinary tract pathogens. J Antimicrob Chemother. 2006;58(6):1274–8.
282. Miro E, Verges C, Garcia I, Mirelis B, Navarro F, Coll P, et al. Resistance to quinolones and
β-lactams in Salmonella enterica due to mutations in topoisomerase-encoding genes, altered
cell permeability and expression of an active efflux system. Enferm Infecc Microbiol Clin.
2004;22(4):204–11.
283. Randall LP, Woodward MJ. The multiple antibiotic resistance (mar) locus and its signifi-
cance. Res Vet Sci. 2002;72(2):87–93.
284. Martínez-Martínez L, Pascual A, Jacoby GA. Quinolone resistance from a transferable plas-
mid. Lancet. 1998;351(9105):797–9.
285. Jacoby GA, Walsh KE, Mills DM, Walker VJ, Oh H, Robicsek A, et al. qnrB, another plasmid-
mediated gene for quinolone resistance. Antimicrob Agents Chemother. 2006;50(4):1178–82.
286. Hata M, Suzuki M, Matsumoto M, Takahashi M, Sato K, Ibe S, et al. Cloning of a novel
gene for quinolone resistance from a transferable plasmid in Shigella flexneri 2b. Antimicrob
Agents Chemother. 2005;49(2):801–3.
287. Vetting MW, Hegde SS, Fajardo JE, Fiser A, Roderick SL, Takiff HE, et al. Pentapeptide
repeat proteins. Biochemistry. 2006;45(1):1–10.
288. Tran JH, Jacoby GA. Mechanism of plasmid-mediated quinolone resistance. Proc Natl Acad
Sci U S A. 2002;99(8):5638–42.
289. Tran JH, Jacoby GA, Hooper DC. Interaction of the plasmid-encoded quinolone resistance
protein QnrA with Escherichia coli topoisomerase IV. Antimicrob Agents Chemother.
2005;49(7):3050–2.
290. Tran JH, Jacoby GA, Hooper DC. Interaction of the plasmid-encoded quinolone resis-
tance protein Qnr with Escherichia coli DNA gyrase. Antimicrob Agents Chemother.
2005;49(1):118–25.
4 Resistance of Gram-negative Bacilli to Antimicrobials 147
291. Vetting MW, Hegde SS, Wang M, Jacoby GA, Hooper DC, Blanchard JS. Structure of QnrB1,
a plasmid-mediated fluoroquinolone resistance factor. J Biol Chem. 2011;286(28):25265–73.
292. Jacoby GA, Hooper DC. Phylogenetic analysis of chromosomally determined qnr and related
proteins. Antimicrob Agents Chemother. 2013;57(4):1930–4.
293. Robicsek A, Strahilevitz J, Sahm DF, Jacoby GA, Hooper DC. qnr prevalence in ceftazidime-
resistant Enterobacteriaceae isolates from the United States. Antimicrob Agents Chemother.
2006;50(8):2872–4.
294. Briales A, Rodriguez-Martinez JM, Velasco C, de Alba PD, Rodriguez-Bano J, Martinez-
Martinez L, et al. Prevalence of plasmid-mediated quinolone resistance determinants qnr and
aac(6’)-Ib-cr in Escherichia coli and Klebsiella pneumoniae producing extended-spectrum
β-lactamases in Spain. Int J Antimicrob Agents. 2012;39(5):431–4.
295. Robicsek A, Jacoby GA, Hooper DC. The worldwide emergence of plasmid-mediated quino-
lone resistance. Lancet Infect Dis. 2006;6(10):629–40.
296. Jacoby GA, Strahilevitz J, Hooper DC. Plasmid-mediated quinolone resistance. Microbiol
Spectr. 2014;2(5). https://doi.org/10.1128/microbiolspec.
297. Guillard T, Grillon A, de Champs C, Cartier C, Madoux J, Bercot B, et al. Mobile insertion
cassette elements found in small non-transmissible plasmids in Proteeae may explain qnrD
mobilization. PLoS One. 2014;9(2):e87801.
298. Robicsek A, Strahilevitz J, Jacoby GA, Macielag M, Abbanat D, Park CH, et al.
Fluoroquinolone-modifying enzyme: a new adaptation of a common aminoglycoside acetyl-
transferase. Nat Med. 2006;12(1):83–8.
299. Hansen LH, Jensen LB, Sorensen HI, Sorensen SJ. Substrate specificity of the OqxAB
multidrug resistance pump in Escherichia coli and selected enteric bacteria. J Antimicrob
Chemother. 2007;60(1):145–7.
300. Hansen LH, Johannesen E, Burmolle M, Sorensen AH, Sorensen SJ. Plasmid-encoded multi-
drug efflux pump conferring resistance to olaquindox in Escherichia coli. Antimicrob Agents
Chemother. 2004;48(9):3332–7.
301. Sorensen AH, Hansen LH, Johannesen E, Sorensen SJ. Conjugative plasmid conferring resis-
tance to olaquindox. Antimicrob Agents Chemother. 2003;47(2):798–9.
302. Kim HB, Wang M, Park CH, Kim EC, Jacoby GA, Hooper DC. oqxAB encoding a multidrug
efflux pump in human clinical isolates of Enterobacteriaceae. Antimicrob Agents Chemother.
2009;53(8):3582–4.
303. Li L, Liao X, Yang Y, Sun J, Li L, Liu B, et al. Spread of oqxAB in Salmonella enterica
serotype Typhimurium predominantly by IncHI2 plasmids. J Antimicrob Chemother.
2013;68(10):2263–8.
304. Liu BT, Yang QE, Li L, Sun J, Liao XP, Fang LX, et al. Dissemination and characteriza-
tion of plasmids carrying oqxAB-bla CTX-M genes in Escherichia coli isolates from food-
producing animals. PLoS One. 2013;8(9):e73947.
305. Zhao J, Chen Z, Chen S, Deng Y, Liu Y, Tian W, et al. Prevalence and dissemination of oqxAB
in Escherichia coli isolates from animals, farmworkers, and the environment. Antimicrob
Agents Chemother. 2010;54(10):4219–24.
306. Perez F, Rudin SD, Marshall SH, Coakley P, Chen L, Kreiswirth BN, et al. OqxAB, a quino-
lone and olaquindox efflux pump, is widely distributed among multidrug-resistant Klebsiella
pneumoniae isolates of human origin. Antimicrob Agents Chemother. 2013;57(9):4602–3.
307. Wong MH, Chan EW, Chen S. Evolution and dissemination of OqxAB-like efflux pumps,
an emerging quinolone resistance determinant among members of Enterobacteriaceae.
Antimicrob Agents Chemother. 2015;59(6):3290–7.
308. Bialek-Davenet S, Lavigne JP, Guyot K, Mayer N, Tournebize R, Brisse S, et al. Differential
contribution of AcrAB and OqxAB efflux pumps to multidrug resistance and virulence in
Klebsiella pneumoniae. J Antimicrob Chemother. 2015;70(1):81–8.
309. Yamane K, Wachino J, Suzuki S, Kimura K, Shibata N, Kato H, et al. New plasmid-mediated
fluoroquinolone efflux pump, QepA, found in an Escherichia coli clinical isolate. Antimicrob
Agents Chemother. 2007;51(9):3354–60.
148 C. R. Dean et al.
330. Hierowski M. Inhibition of protein synthesis by Chlortetracycline in the E. Coli in vitro sys-
tem. Proc Natl Acad Sci U S A. 1965;53:594–9.
331. Brodersen DE, Clemons WM Jr, Carter AP, Morgan-Warren RJ, Wimberly BT, Ramakrishnan
V. The structural basis for the action of the antibiotics tetracycline, pactamycin, and hygro-
mycin B on the 30S ribosomal subunit. Cell. 2000;103(7):1143–54.
332. Pioletti M, Schlunzen F, Harms J, Zarivach R, Gluhmann M, Avila H, et al. Crystal struc-
tures of complexes of the small ribosomal subunit with tetracycline, edeine and IF3. EMBO
J. 2001;20(8):1829–39.
333. Duggar BM. Aureomycin; a product of the continuing search for new antibiotics. Ann N Y
Acad Sci. 1948;51(Art. 2):177–81.
334. Sum PE, Lee VJ, Testa RT, Hlavka JJ, Ellestad GA, Bloom JD, et al. Glycylcyclines. 1. A
new generation of potent antibacterial agents through modification of 9-aminotetracyclines.
J Med Chem. 1994;37(1):184–8.
335. Solomkin J, Evans D, Slepavicius A, Lee P, Marsh A, Tsai L, et al. Assessing the efficacy and
safety of Eravacycline vs Ertapenem in complicated intra-abdominal infections in the inves-
tigating Gram-Negative infections treated with Eravacycline (IGNITE 1) trial: a randomized
clinical trial. JAMA Surg. 2017;152(3):224–32.
336. Livermore DM, Mushtaq S, Warner M, Woodford N. In Vitro Activity of Eravacycline against
Carbapenem-Resistant Enterobacteriaceae and Acinetobacter baumannii. Antimicrob Agents
Chemother. 2016;60(6):3840–4.
337. Schnappinger D, Hillen W. Tetracyclines: antibiotic action, uptake, and resistance mecha-
nisms. Arch Microbiol. 1996;165(6):359–69.
338. Chopra I, Roberts M. Tetracycline antibiotics: mode of action, applications, molecular biol-
ogy, and epidemiology of bacterial resistance. Microbiol Mol Biol Rev. 2001;65(2):232–60.
339. Thaker M, Spanogiannopoulos P, Wright GD. The tetracycline resistome. Cell Mol Life Sci.
2010;67(3):419–31.
340. Guillaume G, Ledent V, Moens W, Collard JM. Phylogeny of efflux-mediated tetracycline
resistance genes and related proteins revisited. Microb Drug Resist. 2004;10(1):11–26.
341. Levy SB. Active efflux mechanisms for antimicrobial resistance. Antimicrob Agents
Chemother. 1992;36(4):695–703.
342. McMurry L, Petrucci RE Jr, Levy SB. Active efflux of tetracycline encoded by four geneti-
cally different tetracycline resistance determinants in Escherichia coli. Proc Natl Acad Sci U
S A. 1980;77(7):3974–7.
343. Ohnuki T, Katoh T, Imanaka T, Aiba S. Molecular cloning of tetracycline resistance genes
from Streptomyces rimosus in Streptomyces griseus and characterization of the cloned genes.
J Bacteriol. 1985;161(3):1010–6.
344. Hillen W, Berens C. Mechanisms underlying expression of Tn10 encoded tetracycline resis-
tance. Annu Rev Microbiol. 1994;48:345–69.
345. Saenger W, Orth P, Kisker C, Hillen W, Hinrichs W. The tetracycline repressor-a paradigm for
a biological switch. Angew Chem Int Ed Engl. 2000;39(12):2042–52.
346. Hinrichs W, Kisker C, Duvel M, Muller A, Tovar K, Hillen W, et al. Structure of the
Tet repressor-tetracycline complex and regulation of antibiotic resistance. Science.
1994;264(5157):418–20.
347. Kisker C, Hinrichs W, Tovar K, Hillen W, Saenger W. The complex formed between Tet
repressor and tetracycline-Mg2+ reveals mechanism of antibiotic resistance. J Mol Biol.
1995;247(2):260–80.
348. Orth P, Schnappinger D, Hillen W, Saenger W, Hinrichs W. Structural basis of gene regulation
by the tetracycline inducible Tet repressor-operator system. Nat Struct Biol. 2000;7(3):215–9.
349. Petersen PJ, Jacobus NV, Weiss WJ, Sum PE, Testa RT. In vitro and in vivo antibacterial
activities of a novel glycylcycline, the 9-t-butylglycylamido derivative of minocycline (GAR-
936). Antimicrob Agents Chemother. 1999;43(4):738–44.
350. Tally FT, Ellestad GA, Testa RT. Glycylcyclines: a new generation of tetracyclines.
J Antimicrob Chemother. 1995;35(4):449–52.
150 C. R. Dean et al.
351. Testa RT, Petersen PJ, Jacobus NV, Sum PE, Lee VJ, Tally FP. In vitro and in vivo antibacte-
rial activities of the glycylcyclines, a new class of semisynthetic tetracyclines. Antimicrob
Agents Chemother. 1993;37(11):2270–7.
352. Bradford PA, Weaver-Sands DT, Petersen PJ. In vitro activity of tigecycline against iso-
lates from patients enrolled in phase 3 clinical trials of treatment for complicated skin
and skin-structure infections and complicated intra-abdominal infections. Clin Infect Dis.
2005;41(Suppl 5):S315–32.
353. Fluit AC, Florijn A, Verhoef J, Milatovic D. Presence of tetracycline resistance determi-
nants and susceptibility to tigecycline and minocycline. Antimicrob Agents Chemother.
2005;49(4):1636–8.
354. Petersen PJ, Bradford PA. Effect of medium age and supplementation with the biocatalytic
oxygen-reducing reagent oxyrase on in vitro activities of tigecycline against recent clinical
isolates. Antimicrob Agents Chemother. 2005;49(9):3910–8.
355. Fritsche TR, Sader HS, Stilwell MG, Dowzicky MJ, Jones RN. Potency and spectrum of tige-
cycline tested against an international collection of bacterial pathogens associated with skin
and soft tissue infections (2000–2004). Diagn Microbiol Infect Dis. 2005;52(3):195–201.
356. Fritsche TR, Sader HS, Stilwell MG, Dowzicky MJ, Jones RN. Antimicrobial activity of
tigecycline tested against organisms causing community-acquired respiratory tract infection
and nosocomial pneumonia. Diagn Microbiol Infect Dis. 2005;52(3):187–93.
357. Fritsche TR, Strabala PA, Sader HS, Dowzicky MJ, Jones RN. Activity of tigecycline tested
against a global collection of Enterobacteriaceae, including tetracycline-resistant isolates.
Diagn Microbiol Infect Dis. 2005;52(3):209–13.
358. Sader HS, Jones RN, Dowzicky MJ, Fritsche TR. Antimicrobial activity of tigecycline tested
against nosocomial bacterial pathogens from patients hospitalized in the intensive care unit.
Diagn Microbiol Infect Dis. 2005;52(3):203–8.
359. Sader HS, Jones RN, Stilwell MG, Dowzicky MJ, Fritsche TR. Tigecycline activity tested
against 26,474 bloodstream infection isolates: a collection from 6 continents. Diagn Microbiol
Infect Dis. 2005;52(3):181–6.
360. Grossman TH, Starosta AL, Fyfe C, O’Brien W, Rothstein DM, Mikolajka A, et al. Target-
and resistance-based mechanistic studies with TP-434, a novel fluorocycline antibiotic.
Antimicrob Agents Chemother. 2012;56(5):2559–64.
361. Guay GG, Tuckman M, Rothstein DM. Mutations in the tetA(B) gene that cause a change
in substrate specificity of the tetracycline efflux pump. Antimicrob Agents Chemother.
1994;38(4):857–60.
362. Castanheira M, Mendes RE, Jones RN. Update on Acinetobacter species: mechanisms of
antimicrobial resistance and contemporary in vitro activity of minocycline and other treat-
ment options. Clin Infect Dis. 2014;59(Suppl 6):S367–73.
363. Stein GE, Craig WA. Tigecycline: a critical analysis. Clin Infect Dis. 2006;43(4):518–24.
364. Ruzin A, Keeney D, Bradford PA. AcrAB efflux pump plays a role in decreased susceptibility
to tigecycline in Morganella morganii. Antimicrob Agents Chemother. 2005;49(2):791–3.
365. Hentschke M, Wolters M, Sobottka I, Rohde H, Aepfelbacher M. ramR mutations in clini-
cal isolates of Klebsiella pneumoniae with reduced susceptibility to tigecycline. Antimicrob
Agents Chemother. 2010;54(6):2720–3.
366. Veleba M, Schneiders T. Tigecycline resistance can occur independently of the ramA gene in
Klebsiella pneumoniae. Antimicrob Agents Chemother. 2012;56(8):4466–7.
367. Keeney D, Ruzin A, Bradford PA. RamA, a transcriptional regulator, and AcrAB, an RND-
type efflux pump, are associated with decreased susceptibility to tigecycline in Enterobacter
cloacae. Microb Drug Resist. 2007;13(1):1–6.
368. Veleba M, De Majumdar S, Hornsey M, Woodford N, Schneiders T. Genetic characterization
of tigecycline resistance in clinical isolates of Enterobacter cloacae and Enterobacter aero-
genes. J Antimicrob Chemother. 2013;68(5):1011–8.
369. Hentschke M, Christner M, Sobottka I, Aepfelbacher M, Rohde H. Combined ramR mutation
and presence of a Tn1721-associated tet(A) variant in a clinical isolate of Salmonella enterica
serovar Hadar resistant to tigecycline. Antimicrob Agents Chemother. 2010;54(3):1319–22.
4 Resistance of Gram-negative Bacilli to Antimicrobials 151
370. Ruzin A, Keeney D, Bradford PA. AdeABC multidrug efflux pump is associated with
decreased susceptibility to tigecycline in Acinetobacter calcoaceticus-Acinetobacter bau-
mannii complex. J Antimicrob Chemother. 2007;59(5):1001–4.
371. Bratu S, Landman D, Martin DA, Georgescu C, Quale J. Correlation of antimicrobial
resistance with β-lactamases, the OmpA-like porin, and efflux pumps in clinical isolates
of Acinetobacter baumannii endemic to New York City. Antimicrob Agents Chemother.
2008;52(9):2999–3005.
372. Coyne S, Guigon G, Courvalin P, Perichon B. Screening and quantification of the expres-
sion of antibiotic resistance genes in Acinetobacter baumannii with a microarray. Antimicrob
Agents Chemother. 2010;54(1):333–40.
373. Coyne S, Rosenfeld N, Lambert T, Courvalin P, Perichon B. Overexpression of resistance-
nodulation-cell division pump AdeFGH confers multidrug resistance in Acinetobacter bau-
mannii. Antimicrob Agents Chemother. 2010;54(10):4389–93.
374. Rumbo C, Gato E, Lopez M, Ruiz de Alegria C, Fernandez-Cuenca F, Martinez-Martinez L,
et al. Contribution of efflux pumps, porins, and β-lactamases to multidrug resistance in clinical
isolates of Acinetobacter baumannii. Antimicrob Agents Chemother. 2013;57(11):5247–57.
375. Connell SR, Tracz DM, Nierhaus KH, Taylor DE. Ribosomal protection proteins and their
mechanism of tetracycline resistance. Antimicrob Agents Chemother. 2003;47(12):3675–81.
376. Taylor DE. Plasmid-mediated tetracycline resistance in Campylobacter jejuni: expression
in Escherichia coli and identification of homology with streptococcal class M determinant.
J Bacteriol. 1986;165(3):1037–9.
377. Doyle D, McDowall KJ, Butler MJ, Hunter IS. Characterization of an oxytetracycline-
resistance gene, otrA, of Streptomyces rimosus. Mol Microbiol. 1991;5(12):2923–33.
378. Sanchez-Pescador R, Brown JT, Roberts M, Urdea MS. Homology of the TetM with trans-
lational elongation factors: implications for potential modes of tetM-conferred tetracycline
resistance. Nucleic Acids Res. 1988;16(3):1218.
379. Burdett V. Tet(M)-promoted release of tetracycline from ribosomes is GTP dependent.
J Bacteriol. 1996;178(11):3246–51.
380. Spahn CM, Blaha G, Agrawal RK, Penczek P, Grassucci RA, Trieber CA, et al. Localization
of the ribosomal protection protein Tet(O) on the ribosome and the mechanism of tetracycline
resistance. Mol Cell. 2001;7(5):1037–45.
381. Connell SR, Trieber CA, Dinos GP, Einfeldt E, Taylor DE, Nierhaus KH. Mechanism of
Tet(O)-mediated tetracycline resistance. EMBO J. 2003;22(4):945–53.
382. Connell SR, Trieber CA, Stelzl U, Einfeldt E, Taylor DE, Nierhaus KH. The tetracycline
resistance protein Tet(o) perturbs the conformation of the ribosomal decoding centre. Mol
Microbiol. 2002;45(6):1463–72.
383. Donhofer A, Franckenberg S, Wickles S, Berninghausen O, Beckmann R, Wilson
DN. Structural basis for TetM-mediated tetracycline resistance. Proc Natl Acad Sci U S A.
2012;109(42):16900–5.
384. Li W, Atkinson GC, Thakor NS, Allas U, Lu CC, Chan KY, et al. Mechanism of tetracycline
resistance by ribosomal protection protein Tet(O). Nat Commun. 2013;4:1477.
385. Draper MP, Weir S, Macone A, Donatelli J, Trieber CA, Tanaka SK, et al. Mechanism
of action of the novel aminomethylcycline antibiotic omadacycline. Antimicrob Agents
Chemother. 2014;58(3):1279–83.
386. Macone AB, Caruso BK, Leahy RG, Donatelli J, Weir S, Draper MP, et al. In vitro and in vivo
antibacterial activities of omadacycline, a novel aminomethylcycline. Antimicrob Agents
Chemother. 2014;58(2):1127–35.
387. Ross JI, Eady EA, Cove JH, Cunliffe WJ. 16S rRNA mutation associated with tetracycline
resistance in a gram-positive bacterium. Antimicrob Agents Chemother. 1998;42(7):1702–5.
388. Gerrits MM, de Zoete MR, Arents NL, Kuipers EJ, Kusters JG. 16S rRNA mutation-
mediated tetracycline resistance in Helicobacter pylori. Antimicrob Agents Chemother.
2002;46(9):2996–3000.
389. Trieber CA, Taylor DE. Mutations in the 16S rRNA genes of Helicobacter pylori mediate
resistance to tetracycline. J Bacteriol. 2002;184(8):2131–40.
152 C. R. Dean et al.
429. Haidar G, Alkroud A, Cheng S, Churilla TM, Churilla BM, Shields RK, et al. Association
between the presence of aminoglycoside-modifying enzymes and in vitro activity of gentami-
cin, tobramycin, amikacin, and plazomicin against Klebsiella pneumoniae carbapenemase-
and extended-spectrum-β-lactamase-producing Enterobacter species. Antimicrob Agents
Chemother. 2016;60(9):5208–14.
430. Gonzalez-Padilla M, Torre-Cisneros J, Rivera-Espinar F, Pontes-Moreno A, Lopez-Cerero
L, Pascual A, et al. Gentamicin therapy for sepsis due to carbapenem-resistant and colistin-
resistant Klebsiella pneumoniae. J Antimicrob Chemother. 2015;70(3):905–13.
431. Bercot B, Poirel L, Nordmann P. Updated multiplex polymerase chain reaction for detection
of 16S rRNA methylases: high prevalence among NDM-1 producers. Diagn Microbiol Infect
Dis. 2011;71(4):442–5.
432. Miro E, Grunbaum F, Gomez L, Rivera A, Mirelis B, Coll P, et al. Characterization of
aminoglycoside-modifying enzymes in Enterobacteriaceae clinical strains and characteriza-
tion of the plasmids implicated in their diffusion. Microb Drug Resist. 2013;19(2):94–9.
433. Petrosillo N, Vranic-Ladavac M, Feudi C, Villa L, Fortini D, Barisic N, et al. Spread of
Enterobacter cloacae carrying blaNDM-1, blaCTX-M-15, blaSHV-12 and plasmid-mediated
quinolone resistance genes in a surgical intensive care unit in Croatia. J Glob Antimicrob
Resist. 2016;4:44–8.
434. Poirel L, Savov E, Nazli A, Trifonova A, Todorova I, Gergova I, et al. Outbreak caused by
NDM-1- and RmtB-producing Escherichia coli in Bulgaria. Antimicrob Agents Chemother.
2014;58(4):2472–4.
435. Poirel L, Yilmaz M, Istanbullu A, Arslan F, Mert A, Bernabeu S, et al. Spread of NDM-1-
producing Enterobacteriaceae in a neonatal intensive care unit in Istanbul, Turkey. Antimicrob
Agents Chemother. 2014;58(5):2929–33.
436. Rahman M, Prasad KN, Pathak A, Pati BK, Singh A, Ovejero CM, et al. RmtC and RmtF 16S
rRNA methyltransferase in NDM-1-producing Pseudomonas aeruginosa. Emerg Infect Dis.
2015;21(11):2059–62.
437. Bradford PA, Dean CR. Resistance of Gram-Negative Bacilli to antimicrobials. In: Fong
IW, Drlica K, editors. Antimicrobial aesistance and implications for the twenty-first century.
Boston, MA: Springer US; 2008. p. 97–159.
438. Honore N, Marchal G, Cole ST. Novel mutation in 16S rRNA associated with strep-
tomycin dependence in Mycobacterium tuberculosis. Antimicrob Agents Chemother.
1995;39(3):769–70.
439. Nessar R, Reyrat JM, Murray A, Gicquel B. Genetic analysis of new 16S rRNA mutations
conferring aminoglycoside resistance in Mycobacterium abscessus. J Antimicrob Chemother.
2011;66(8):1719–24.
440. Criswell D, Tobiason VL, Lodmell JS, Samuels DS. Mutations conferring aminoglyco-
side and spectinomycin resistance in Borrelia burgdorferi. Antimicrob Agents Chemother.
2006;50(2):445–52.
441. Vakulenko SB, Mobashery S. Versatility of aminoglycosides and prospects for their future.
Clin Microbiol Rev. 2003;16(3):430–50.
442. Ramirez MS, Tolmasky ME. Aminoglycoside modifying enzymes. Drug Resist Updat.
2010;13(6):151–71.
443. Wright GD. Aminoglycoside-modifying enzymes. Curr Opin Microbiol. 1999;2(5):499–503.
444. Kosmidis C, Giannopoulou M, Flountzi A, Markogiannakis A, Goukos D, Petrikkos G, et al.
Genetic basis of aminoglycoside resistance following changes in aminoglycoside prescrip-
tion patterns. J Chemother. 2013;25(4):217–21.
445. Rather PN, Munayyer H, Mann PA, Hare RS, Miller GH, Shaw KJ. Genetic analysis of bacte-
rial acetyltransferases: identification of amino acids determining the specificities of the ami-
noglycoside 6’-N-acetyltransferase Ib and IIa proteins. J Bacteriol. 1992;174(10):3196–203.
446. Wohlleben W, Arnold W, Bissonnette L, Pelletier A, Tanguay A, Roy PH, et al. On the evo-
lution of Tn21-like multiresistance transposons: sequence analysis of the gene (aacC1) for
4 Resistance of Gram-negative Bacilli to Antimicrobials 155
466. Aires JR, Nikaido H. Aminoglycosides are captured from both periplasm and cytoplasm by
the AcrD multidrug efflux transporter of Escherichia coli. J Bacteriol. 2005;187(6):1923–9.
467. Rosenberg EY, Ma D, Nikaido H. AcrD of Escherichia coli is an aminoglycoside efflux
pump. J Bacteriol. 2000;182(6):1754–6.
468. Chan YY, Tan TM, Ong YM, Chua KL. BpeAB-OprB, a multidrug efflux pump in
Burkholderia pseudomallei. Antimicrob Agents Chemother. 2004;48(4):1128–35.
469. Aires JR, Kohler T, Nikaido H, Plesiat P. Involvement of an active efflux system in the natural
resistance of Pseudomonas aeruginosa to aminoglycosides. Antimicrob Agents Chemother.
1999;43(11):2624–8.
470. Westbrock-Wadman S, Sherman DR, Hickey MJ, Coulter SN, Zhu YQ, Warrener P, et al.
Characterization of a Pseudomonas aeruginosa efflux pump contributing to aminoglycoside
impermeability. Antimicrob Agents Chemother. 1999;43(12):2975–83.
471. Jeannot K, Sobel ML, El Garch F, Poole K, Plesiat P. Induction of the MexXY efflux
pump in Pseudomonas aeruginosa is dependent on drug-ribosome interaction. J Bacteriol.
2005;187(15):5341–6.
472. Karlowsky JA, Hoban DJ, Zelenitsky SA, Zhanel GG. Altered denA and anr gene expression
in aminoglycoside adaptive resistance in Pseudomonas aeruginosa. J Antimicrob Chemother.
1997;40(3):371–6.
473. Yamamoto M, Ueda A, Kudo M, Matsuo Y, Fukushima J, Nakae T, et al. Role of MexZ and
PA5471 in transcriptional regulation of mexXY in Pseudomonas aeruginosa. Microbiology.
2009;155(Pt 10):3312–21.
474. Lau CH, Krahn T, Gilmour C, Mullen E, Poole K. AmgRS-mediated envelope stress-
inducible expression of the mexXY multidrug efflux operon of Pseudomonas aeruginosa.
Microbiologyopen. 2015;4(1):121–35.
475. Guenard S, Muller C, Monlezun L, Benas P, Broutin I, Jeannot K, et al. Multiple muta-
tions lead to MexXY-OprM-dependent aminoglycoside resistance in clinical strains of
Pseudomonas aeruginosa. Antimicrob Agents Chemother. 2014;58(1):221–8.
476. Pirnay JP, Bilocq F, Pot B, Cornelis P, Zizi M, Van Eldere J, et al. Pseudomonas aeruginosa
population structure revisited. PLoS One. 2009;4(11):e7740.
477. Taccetti G, Campana S, Neri AS, Boni V, Festini F. Antibiotic therapy against Pseudomonas
aeruginosa in cystic fibrosis. J Chemother. 2008;20(2):166–9.
478. Ciofu O, Mandsberg LF, Wang H, Hoiby N. Phenotypes selected during chronic lung infec-
tion in cystic fibrosis patients: implications for the treatment of Pseudomonas aeruginosa
biofilm infections. FEMS Immunol Med Microbiol. 2012;65(2):215–25.
479. Sadovskaya I, Vinogradov E, Li J, Hachani A, Kowalska K, Filloux A. High-level antibi-
otic resistance in Pseudomonas aeruginosa biofilm: the ndvB gene is involved in the pro-
duction of highly glycerol-phosphorylated β-(1->3)-glucans, which bind aminoglycosides.
Glycobiology. 2010;20(7):895–904.
480. Mah TF, Pitts B, Pellock B, Walker GC, Stewart PS, O’Toole GA. A genetic basis for
Pseudomonas aeruginosa biofilm antibiotic resistance. Nature. 2003;426(6964):306–10.
481. Crabbe A, Liu Y, Matthijs N, Rigole P, De La Fuente-Nunez C, Davis R, et al. Antimicrobial
efficacy against Pseudomonas aeruginosa biofilm formation in a three-dimensional lung epi-
thelial model and the influence of fetal bovine serum. Sci Rep. 2017;7:43321.
482. Nagant C, Tre-Hardy M, El-Ouaaliti M, Savage P, Devleeschouwer M, Dehaye JP. Interaction
between tobramycin and CSA-13 on clinical isolates of Pseudomonas aeruginosa in a model
of young and mature biofilms. Appl Microbiol Biotechnol. 2010;88(1):251–63.
483. Burk DL, Hon WC, Leung AK, Berghuis AM. Structural analyses of nucleotide binding to an
aminoglycoside phosphotransferase. Biochemistry. 2001;40(30):8756–64.
484. Hon WC, McKay GA, Thompson PR, Sweet RM, Yang DS, Wright GD, et al. Structure of
an enzyme required for aminoglycoside antibiotic resistance reveals homology to eukaryotic
protein kinases. Cell. 1997;89(6):887–95.
485. Sakon J, Liao HH, Kanikula AM, Benning MM, Rayment I, Holden HM. Molecular struc-
ture of kanamycin nucleotidyltransferase determined to 3.0-A resolution. Biochemistry.
1993;32(45):11977–84.
4 Resistance of Gram-negative Bacilli to Antimicrobials 157
486. Wolf E, Vassilev A, Makino Y, Sali A, Nakatani Y, Burley SK. Crystal structure of a GCN5-
related N-acetyltransferase: Serratia marcescens aminoglycoside 3-N-acetyltransferase. Cell.
1998;94(4):439–49.
487. Wybenga-Groot LE, Draker K, Wright GD, Berghuis AM. Crystal structure of an aminogly-
coside 6’-N-acetyltransferase: defining the GCN5-related N-acetyltransferase superfamily
fold. Structure. 1999;7(5):497–507.
488. Caldwell SJ, Berghuis AM. Small-angle X-ray scattering analysis of the bifunctional anti-
biotic resistance enzyme aminoglycoside (6’) acetyltransferase-ie/aminoglycoside (2”)
phosphotransferase-ia reveals a rigid solution structure. Antimicrob Agents Chemother.
2012;56(4):1899–906.
489. Burk DL, Ghuman N, Wybenga-Groot LE, Berghuis AM. X-ray structure of the AAC(6’)-Ii
antibiotic resistance enzyme at 1.8 A resolution; examination of oligomeric arrangements in
GNAT superfamily members. Protein Sci. 2003;12(3):426–37.
490. Burk DL, Xiong B, Breitbach C, Berghuis AM. Structures of aminoglycoside acetyltransfer-
ase AAC(6’)-Ii in a novel crystal form: structural and normal-mode analyses. Acta Crystallogr
D Biol Crystallogr. 2005;61(Pt 9):1273–9.
491. Nurizzo D, Shewry SC, Perlin MH, Brown SA, Dholakia JN, Fuchs RL, et al. The crystal
structure of aminoglycoside-3’-phosphotransferase-IIa, an enzyme responsible for antibiotic
resistance. J Mol Biol. 2003;327(2):491–506.
492. Young PG, Walanj R, Lakshmi V, Byrnes LJ, Metcalf P, Baker EN, et al. The crystal struc-
tures of substrate and nucleotide complexes of Enterococcus faecium aminoglycoside-2″-
phosphotransferase-IIa [APH(2″)-IIa] provide insights into substrate selectivity in the
APH(2″) subfamily. J Bacteriol. 2009;191(13):4133–43.
493. Labby KJ, Garneau-Tsodikova S. Strategies to overcome the action of aminoglycoside-
modifying enzymes for treating resistant bacterial infections. Future Med Chem.
2013;5(11):1285–309.
494. Lee JH, Lee CS. Clinical usefulness of arbekacin. Infect Chemother. 2016;48(1):1–11.
495. Sader HS, Rhomberg PR, Farrell DJ, Jones RN. Arbekacin activity against contemporary
clinical bacteria isolated from patients hospitalized with pneumonia. Antimicrob Agents
Chemother. 2015;59(6):3263–70.
496. Livermore DM, Mushtaq S, Warner M, Zhang JC, Maharjan S, Doumith M, et al. Activity
of aminoglycosides, including ACHN-490, against carbapenem-resistant Enterobacteriaceae
isolates. J Antimicrob Chemother. 2011;66(1):48–53.
497. Ago K, Umemura E, Takahashi Y, Igarashi M, Hayashi C, Shibasaki M, Yamada K, Ida T,
Yonezawa M. TS3112, a novel aminoglycoside antibiotic active against multidrug-resistant
pathogens producing 16S rRNA methyltransferases: synthesis and structure-activity relation-
ships ASM microbe 2017; New Orleans; 2017 Jun 4.
498. Hamed K, Debonnett L. Tobramycin inhalation powder for the treatment of pulmonary
Pseudomonas aeruginosa infection in patients with cystic fibrosis: a review based on clinical
evidence. Ther Adv Respir Dis. 2017;11(5):193–209.
499. McKeage K. Tobramycin inhalation powder: a review of its use in the treatment of
chronic Pseudomonas aeruginosa infection in patients with cystic fibrosis. Drugs.
2013;73(16):1815–27.
500. Niederman MS, Chastre J, Corkery K, Fink JB, Luyt CE, Garcia MS. BAY41-6551 achieves
bactericidal tracheal aspirate amikacin concentrations in mechanically ventilated patients
with Gram-negative pneumonia. Intensive Care Med. 2012;38(2):263–71.
501. Luyt CE, Eldon MA, Stass H, Gribben D, Corkery K, Chastre J. Pharmacokinetics and toler-
ability of amikacin administered as BAY41-6551 aerosol in mechanically ventilated patients
with gram-negative pneumonia and acute renal failure. J Aerosol Med Pulm Drug Deliv.
2011;24(4):183–90.
502. Adams PD, Grosse-Kunstleve RW, Hung LW, Ioerger TR, McCoy AJ, Moriarty NW, et al.
PHENIX: building new software for automated crystallographic structure determination.
Acta Crystallogr D. 2002;58:1948–54.
158 C. R. Dean et al.
503. Koch-Weser J, Sidel VW, Federman EB, Kanarek P, Finer DC, Eaton AE. Adverse effects
of sodium colistimethate. Manifestations and specific reaction rates during 317 courses of
therapy. Ann Intern Med. 1970;72(6):857–68.
504. Yow EM, Moyer JH. Toxicity of polymyxin B. II. Human studies with particular reference to
evaluation of renal function. AMA Arch Intern Med. 1953;92(2):248–57.
505. Velkov T, Thompson PE, Nation RL, Li J. Structure--activity relationships of polymyxin
antibiotics. J Med Chem. 2010;53(5):1898–916.
506. Pristovsek P, Kidric J. The search for molecular determinants of LPS inhibition by proteins
and peptides. Curr Top Med Chem. 2004;4(11):1185–201.
507. Sampson TR, Liu X, Schroeder MR, Kraft CS, Burd EM, Weiss DS. Rapid killing of
Acinetobacter baumannii by polymyxins is mediated by a hydroxyl radical death pathway.
Antimicrob Agents Chemother. 2012;56(11):5642–9.
508. Deris ZZ, Akter J, Sivanesan S, Roberts KD, Thompson PE, Nation RL, et al. A second-
ary mode of action of polymyxins against Gram-negative bacteria involves the inhibition of
NADH-quinone oxidoreductase activity. J Antibiot (Tokyo). 2014;67(2):147–51.
509. Olaitan AO, Morand S, Rolain JM. Mechanisms of polymyxin resistance: acquired and
intrinsic resistance in bacteria. Front Microbiol. 2014;5:643.
510. Nowicki EM, O’Brien JP, Brodbelt JS, Trent MS. Extracellular zinc induces phosphoethanol-
amine addition to Pseudomonas aeruginosa lipid A via the ColRS two-component system.
Mol Microbiol. 2015;97(1):166–78.
511. Chin CY, Gregg KA, Napier BA, Ernst RK, Weiss DS. A PmrB-regulated deacetylase
required for lipid A modification and polymyxin resistance in Acinetobacter baumannii.
Antimicrob Agents Chemother. 2015;59(12):7911–4.
512. Rubin EJ, Herrera CM, Crofts AA, Trent MS. PmrD is required for modifications to esch-
erichia coli endotoxin that promote antimicrobial resistance. Antimicrob Agents Chemother.
2015;59(4):2051–61.
513. Prost LR, Daley ME, Le Sage V, Bader MW, Le Moual H, Klevit RE, et al. Activation of the
bacterial sensor kinase PhoQ by acidic pH. Mol Cell. 2007;26(2):165–74.
514. Bader MW, Sanowar S, Daley ME, Schneider AR, Cho U, Xu W, et al. Recognition of anti-
microbial peptides by a bacterial sensor kinase. Cell. 2005;122(3):461–72.
515. Gunn JS, Richards SM. Recognition and integration of multiple environmental signals by the
bacterial sensor kinase PhoQ. Cell Host Microbe. 2007;1(3):163–5.
516. Schurek KN, Sampaio JL, Kiffer CR, Sinto S, Mendes CM, Hancock RE. Involvement
of pmrAB and phoPQ in polymyxin B adaptation and inducible resistance in non-cystic
fibrosis clinical isolates of Pseudomonas aeruginosa. Antimicrob Agents Chemother.
2009;53(10):4345–51.
517. Pragasam AK, Shankar C, Veeraraghavan B, Biswas I, Nabarro LE, Inbanathan FY, et al.
Molecular mechanisms of colistin resistance in Klebsiella pneumoniae causing bacteremia
from India-a first report. Front Microbiol. 2016;7:2135.
518. Cannatelli A, Di Pilato V, Giani T, Arena F, Ambretti S, Gaibani P, et al. In vivo evolution
to colistin resistance by PmrB sensor kinase mutation in KPC-producing Klebsiella pneu-
moniae is associated with low-dosage colistin treatment. Antimicrob Agents Chemother.
2014;58(8):4399–403.
519. Cheng HY, Chen YF, Peng HL. Molecular characterization of the PhoPQ-PmrD-PmrAB
mediated pathway regulating polymyxin B resistance in Klebsiella pneumoniae CG43.
J Biomed Sci. 2010;17:60.
520. Jayol A, Poirel L, Brink A, Villegas MV, Yilmaz M, Nordmann P. Resistance to colistin asso-
ciated with a single amino acid change in protein PmrB among Klebsiella pneumoniae iso-
lates of worldwide origin. Antimicrob Agents Chemother. 2014;58(8):4762–6.
521. Moskowitz SM, Brannon MK, Dasgupta N, Pier M, Sgambati N, Miller AK, et al. PmrB
mutations promote polymyxin resistance of Pseudomonas aeruginosa isolated from colistin-
treated cystic fibrosis patients. Antimicrob Agents Chemother. 2012;56(2):1019–30.
4 Resistance of Gram-negative Bacilli to Antimicrobials 159
522. Barrow K, Kwon DH. Alterations in two-component regulatory systems of phoPQ and
pmrAB are associated with polymyxin B resistance in clinical isolates of Pseudomonas aeru-
ginosa. Antimicrob Agents Chemother. 2009;53(12):5150–4.
523. Dahdouh E, Gomez-Gil R, Sanz S, Gonzalez-Zorn B, Daoud Z, Mingorance J, et al. A novel
mutation in pmrB mediates colistin resistance during therapy of Acinetobacter baumannii.
Int J Antimicrob Agents. 2017;49(6):727–33.
524. Lean SS, Suhaili Z, Ismail S, Rahman NI, Othman N, Abdullah FH, et al. Prevalence and
genetic characterization of carbapenem- and polymyxin-resistant Acinetobacter baumannii
isolated from a tertiary hospital in Terengganu, Malaysia. ISRN Microbiol. 2014;2014:953417.
525. Adams MD, Nickel GC, Bajaksouzian S, Lavender H, Murthy AR, Jacobs MR, et al.
Resistance to colistin in Acinetobacter baumannii associated with mutations in the PmrAB
two-component system. Antimicrob Agents Chemother. 2009;53(9):3628–34.
526. Arroyo LA, Herrera CM, Fernandez L, Hankins JV, Trent MS, Hancock RE. The pmrCAB
operon mediates polymyxin resistance in Acinetobacter baumannii ATCC 17978 and clini-
cal isolates through phosphoethanolamine modification of lipid A. Antimicrob Agents
Chemother. 2011;55(8):3743–51.
527. Lesho E, Yoon EJ, McGann P, Snesrud E, Kwak Y, Milillo M, et al. Emergence of colistin-
resistance in extremely drug-resistant Acinetobacter baumannii containing a novel pmrCAB
operon during colistin therapy of wound infections. J Infect Dis. 2013;208(7):1142–51.
528. Jayol A, Nordmann P, Brink A, Poirel L. Heteroresistance to colistin in Klebsiella pneu-
moniae associated with alterations in the PhoPQ regulatory system. Antimicrob Agents
Chemother. 2015;59(5):2780–4.
529. Cheng YH, Lin TL, Pan YJ, Wang YP, Lin YT, Wang JT. Colistin resistance mecha-
nisms in Klebsiella pneumoniae strains from Taiwan. Antimicrob Agents Chemother.
2015;59(5):2909–13.
530. Wright MS, Suzuki Y, Jones MB, Marshall SH, Rudin SD, van Duin D, et al. Genomic and
transcriptomic analyses of colistin-resistant clinical isolates of Klebsiella pneumoniae reveal
multiple pathways of resistance. Antimicrob Agents Chemother. 2015;59(1):536–43.
531. Miller AK, Brannon MK, Stevens L, Johansen HK, Selgrade SE, Miller SI, et al. PhoQ
mutations promote lipid A modification and polymyxin resistance of Pseudomonas aeru-
ginosa found in colistin-treated cystic fibrosis patients. Antimicrob Agents Chemother.
2011;55(12):5761–9.
532. Park YK, Choi JY, Shin D, Ko KS. Correlation between overexpression and amino acid
substitution of the PmrAB locus and colistin resistance in Acinetobacter baumannii. Int
J Antimicrob Agents. 2011;37(6):525–30.
533. Poirel L, Jayol A, Bontron S, Villegas MV, Ozdamar M, Turkoglu S, et al. The mgrB gene
as a key target for acquired resistance to colistin in Klebsiella pneumoniae. J Antimicrob
Chemother. 2015;70(1):75–80.
534. Cannatelli A, D’Andrea MM, Giani T, Di Pilato V, Arena F, Ambretti S, et al. In vivo emer-
gence of colistin resistance in Klebsiella pneumoniae producing KPC-type carbapenemases
mediated by insertional inactivation of the PhoQ/PhoP mgrB regulator. Antimicrob Agents
Chemother. 2013;57(11):5521–6.
535. Cannatelli A, Giani T, D’Andrea MM, Di Pilato V, Arena F, Conte V, et al. MgrB inactivation
is a common mechanism of colistin resistance in KPC-producing Klebsiella pneumoniae of
clinical origin. Antimicrob Agents Chemother. 2014;58(10):5696–703.
536. Olaitan AO, Diene SM, Kempf M, Berrazeg M, Bakour S, Gupta SK, et al. Worldwide emer-
gence of colistin resistance in Klebsiella pneumoniae from healthy humans and patients in Lao
PDR, Thailand, Israel, Nigeria and France owing to inactivation of the PhoP/PhoQ regulator
mgrB: an epidemiological and molecular study. Int J Antimicrob Agents. 2014;44(6):500–7.
537. Jayol A, Nordmann P, Desroches M, Decousser JW, Poirel L. Acquisition of broad-spectrum
cephalosporin resistance leading to colistin resistance in Klebsiella pneumoniae. Antimicrob
Agents Chemother. 2016;60(5):3199–201.
160 C. R. Dean et al.
538. Cheng YH, Lin TL, Lin YT, Wang JT. Amino acid substitutions of CrrB responsible for resis-
tance to colistin through CrrC in Klebsiella pneumoniae. Antimicrob Agents Chemother.
2016;60(6):3709–16.
539. Fernandez L, Gooderham WJ, Bains M, McPhee JB, Wiegand I, Hancock RE. Adaptive resis-
tance to the “last hope” antibiotics polymyxin B and colistin in Pseudomonas aeruginosa
is mediated by the novel two-component regulatory system ParR-ParS. Antimicrob Agents
Chemother. 2010;54(8):3372–82.
540. Gutu AD, Sgambati N, Strasbourger P, Brannon MK, Jacobs MA, Haugen E, et al. Polymyxin
resistance of Pseudomonas aeruginosa phoQ mutants is dependent on additional two-
component regulatory systems. Antimicrob Agents Chemother. 2013;57(5):2204–15.
541. Fernandez L, Jenssen H, Bains M, Wiegand I, Gooderham WJ, Hancock RE. The two-
component system CprRS senses cationic peptides and triggers adaptive resistance in
Pseudomonas aeruginosa independently of ParRS. Antimicrob Agents Chemother.
2012;56(12):6212–22.
542. Jeannot K, Bolard A, Plesiat P. Resistance to polymyxins in Gram-negative organisms. Int
J Antimicrob Agents. 2017;49(5):526–35.
543. Halaby T, Kucukkose E, Janssen AB, Rogers MR, Doorduijn DJ, van der Zanden AG, et al.
Genomic characterization of colistin heteroresistance in Klebsiella pneumoniae during a
nosocomial outbreak. Antimicrob Agents Chemother. 2016;60(11):6837–43.
544. Bardet L, Baron S, Leangapichart T, Okdah L, Diene SM, Rolain JM. Deciphering heterore-
sistance to Colistin in a Klebsiella pneumoniae isolate from Marseille, France. Antimicrob
Agents Chemother. 2017;61(6):e00356.
545. Li J, Rayner CR, Nation RL, Owen RJ, Spelman D, Tan KE, et al. Heteroresistance to
colistin in multidrug-resistant Acinetobacter baumannii. Antimicrob Agents Chemother.
2006;50(9):2946–50.
546. Moffatt JH, Harper M, Adler B, Nation RL, Li J, Boyce JD. Insertion sequence ISAba11 is
involved in colistin resistance and loss of lipopolysaccharide in Acinetobacter baumannii.
Antimicrob Agents Chemother. 2011;55(6):3022–4.
547. Bojkovic J, Richie DL, Six DA, Rath CM, Sawyer WS, Hu Q, et al. Characterization of an
Acinetobacter baumannii lptD deletion strain: permeability defects and response to inhibition
of lipopolysaccharide and fatty acid biosynthesis. J Bacteriol. 2015;198(4):731–41.
548. Richie DL, Takeoka KT, Bojkovic J, Metzger LE, Rath CM, Sawyer WS, et al. Toxic accu-
mulation of LPS pathway intermediates underlies the requirement of LpxH for growth of
Acinetobacter baumannii ATCC 19606. PLoS One. 2016;11(8):e0160918.
549. Wei J-R, Richie DL, Mostafavi M, Metzger LE IV, Rath CM, Sawyer WS, Takeoka KT, Dean
CR. LpxK is essential for growth of Acinetobacter baumannii ATCC 19606: relationship
to toxic accumulation of lipid A pathway intermediates. mSPHERE. 2017;4(2). https://doi.
org/10.1128/mSphere.00199-17.
550. Henry R, Vithanage N, Harrison P, Seemann T, Coutts S, Moffatt JH, et al. Colistin-resistant,
lipopolysaccharide-deficient Acinetobacter baumannii responds to lipopolysaccharide loss
through increased expression of genes involved in the synthesis and transport of lipopro-
teins, phospholipids, and poly-β-1,6-N-acetylglucosamine. Antimicrob Agents Chemother.
2012;56(1):59–69.
551. Beceiro A, Moreno A, Fernandez N, Vallejo JA, Aranda J, Adler B, et al. Biological cost of
different mechanisms of colistin resistance and their impact on virulence in Acinetobacter
baumannii. Antimicrob Agents Chemother. 2014;58(1):518–26.
552. Hood MI, Becker KW, Roux CM, Dunman PM, Skaar EP. genetic determinants of intrinsic
colistin tolerance in Acinetobacter baumannii. Infect Immun. 2013;81(2):542–51.
553. Lean SS, Yeo CC, Suhaili Z, Thong KL. Comparative genomics of two ST 195 carbapenem-
resistant Acinetobacter baumannii with different susceptibility to polymyxin revealed under-
lying resistance mechanism. Front Microbiol. 2015;6:1445.
554. Thi Khanh Nhu N, Riordan DW, Do Hoang Nhu T, Thanh DP, Thwaites G, Huong Lan NP,
et al. The induction and identification of novel Colistin resistance mutations in Acinetobacter
baumannii and their implications. Sci Rep. 2016;6:28291.
4 Resistance of Gram-negative Bacilli to Antimicrobials 161
555. Regenbogen B, Willmann M, Steglich M, Bunk B, Nubel U, Peter S, et al. Rapid and con-
sistent evolution of colistin resistance in XDR Pseudomonas aeruginosa during morbidostat
culture. Antimicrob Agents Chemother. 2017; 61(9):e00043–17.
556. Liu Y-Y, Wang Y, Walsh TR, Yi L-X, Zhang R, Spencer J, et al. Emergence of plasmid-
mediated colistin resistance mechanism MCR-1 in animals and human beings in China: a
microbiological and molecular biological study. Lancet Infect Dis. 2016;16(2):161–8.
557. Skov RL, Monnet DL. Plasmid-mediated colistin resistance (mcr-1 gene): three months later,
the story unfolds. Euro Surveill. 2016;21(9):30155.
558. Al-Tawfiq JA, Laxminarayan R, Mendelson M. How should we respond to the emergence of
plasmid-mediated colistin resistance in humans and animals? Int J Infect Dis. 2017;54:77–84.
559. McGann P, Snesrud E, Maybank R, Corey B, Ong AC, Clifford R, et al. Escherichia coli
Harboring mcr-1 and blaCTX-M on a novel IncF plasmid: first report of mcr-1 in the United
States. Antimicrob Agents Chemother. 2016;60(7):4420–1.
560. Di Pilato V, Arena F, Tascini C, Cannatelli A, Henrici De Angelis L, Fortunato S, et al.
mcr-1.2, a new mcr variant carried on a transferable plasmid from a colistin-resistant KPC
carbapenemase-producing Klebsiella pneumoniae strain of sequence type 512. Antimicrob
Agents Chemother. 2016;60(9):5612–5.
561. Xavier BB, Lammens C, Ruhal R, Kumar-Singh S, Butaye P, Goossens H, et al. Identification
of a novel plasmid-mediated colistin-resistance gene, mcr-2, in Escherichia coli,
Belgium, June 2016. Euro Surveillance. 2016;21(27). https://doi.org/10.2807/1560-7917.
ES.2016.21.27.30280.
562. Yin W, Li H, Shen Y, Liu Z, Wang S, Shen Z, et al. Novel plasmid-mediated colistin resis-
tance gene mcr-3 in Escherichia coli. MBio. 2017;8(3):e00543.
563. Liu YY, Chandler CE, Leung LM, McElheny CL, Mettus RT, Shanks RMQ, et al. Structural
modification of lipopolysaccharide conferred by mcr-1 in Gram-negative ESKAPE
Pathogens. Antimicrob Agents Chemother. 2017;61(6):e00580.
564. Du H, Chen L, Tang Y-W, Kreiswirth BN. Emergence of the mcr-1 colistin resistance gene in
carbapenem-resistant Enterobacteriaceae. Lancet Infect Dis. 2016;16(3):287–8.
565. Yao X, Doi Y, Zeng L, Lv L, Liu J-H. Carbapenem-resistant and colistin-resistant Escherichia
coli co-producing NDM-9 and MCR-1. Lancet Infect Dis. 2016;16(3):288–9.
566. Lai CC, Chuang YC, Chen CC, Tang HJ. Coexistence of MCR-1 and NDM-9 in a clinical
carbapenem-resistant Escherichia coli isolate. Int J Antimicrob Agents. 2017;49(4):517–8.
567. Mediavilla JR, Patrawalla A, Chen L, Chavda KD, Mathema B, Vinnard C, et al. Colistin- and
carbapenem-resistant Escherichia coli harboring mcr-1 and blaNDM-5, causing a compli-
cated urinary tract infection in a patient from the United States. MBio. 2016;7(4):e01191.
568. Guh AY, Bulens SN, Mu Y, Jacob JT, Reno J, Scott J, et al. Epidemiology of
carbapenem- resistant Enterobacteriaceae in 7 US communities, 2012–2013. JAMA.
2015;314(14):1479–87.
569. Castanheira M, Griffin MA, Deshpande LM, Mendes RE, Jones RN, Flamm RK. Detection
of mcr-1 among Escherichia coli clinical isolates collected worldwide as part of the SENTRY
antimicrobial surveillance program in 2014 and 2015. Antimicrob Agents Chemother.
2016;60(9):5623–4.
570. Caniaux I, van Belkum A, Zambardi G, Poirel L, Gros MF. MCR: modern colistin resistance.
Eur J Clin Microbiol Infect Dis. 2017;36(3):415–20.
571. Gao R, Hu Y, Li Z, Sun J, Wang Q, Lin J, et al. Dissemination and mechanism for the MCR-1
colistin resistance. PLoS Pathog. 2016;12(11):e1005957.
572. Sun J, Xu Y, Gao R, Lin J, Wei W, Srinivas S, et al. Deciphering MCR-2 colistin resistance.
MBio. 2017;8(3):e00625.
573. Ye H, Li Y, Li Z, Gao R, Zhang H, Wen R, et al. Diversified mcr-1-harbouring plasmid reser-
voirs confer resistance to colistin in human gut microbiota. MBio. 2016;7(2):e00177.
574. Vaara M, Fox J, Loidl G, Siikanen O, Apajalahti J, Hansen F, et al. Novel polymyxin deriva-
tives carrying only three positive charges are effective antibacterial agents. Antimicrob
Agents Chemother. 2008;52(9):3229–36.
162 C. R. Dean et al.
Neil W. Schluger
The problem of drug resistance in tuberculosis was apparent from the first experi-
ment involving antibiotic treatment of this disease. Streptomycin, the first antibiotic
with activity against M. tuberculosis, was discovered by Selman Waksman and
Albert Schatz of Rutgers University in the 1940s. The first rigorous investigation of
its use was conducted by the British Medical Research Council and reported in a
paper in the British Medical Journal published in 1948 [1]. In that experiment, gen-
erally also acknowledged as the first randomized controlled trial ever to be pub-
lished, 100 men were chosen to receive either bed rest, the standard of care at the
time for tuberculosis (TB), or bed rest plus injections of streptomycin. The results
were striking: they demonstrated that streptomycin was clearly effective in patients
with respect to improvements in symptoms, chest radiographic findings, and results
of sputum bacteriology (most of the study participants converted to negative sputum
cultures within a few months after beginning streptomycin injections). However,
after (often very soon after) converting to negative sputum cultures, essentially all
the patients in the trial relapsed and again developed positive sputum cultures. In all
these relapsed cases, cultures that had initially been susceptible to streptomycin had
become resistant. That very first trial, which clearly established that antibiotic treat-
ment of tuberculosis was effective, also demonstrated that drug resistance could
easily emerge when the disease was treated with a single antibiotic for more than a
few days. The need for multidrug regimens, in order to prevent the emergence of
resistance during treatment of tuberculosis, was suggested [2]. Combination
N. W. Schluger (*)
Departments of Medicine, Epidemiology and Environmental Health Sciences,
Columbia University Medical Center, New York, NY, USA
e-mail: [email protected]
treatment became possible in the early 1950s with the introduction of isoniazid, and
the development of other so-called “first-line” drugs (rifampin, ethambutol, pyra-
zinamide) followed.
As noted above, isoniazid (INH) was initially introduced in the early 1950s and
was recognized immediately as an effective antituberculosis drug [3–7]. Although
the BMRC streptomycin trial had certainly brought the emergence of drug resis-
tance to the fore, the highly active properties of INH led some to think that this drug
could be used as a single agent in the treatment of TB disease in poor countries with
limited resources. That would eliminate the need to obtain and use more expensive
and complex agents such as streptomycin, which had to be given by injection. Trials
in the late 1950s and early 1960s in Africa and India using INH monotherapy for TB
showed a high rate of favorable responses, but also a very high rate—as high as
53%—of the development of INH resistance. This certainly (and perhaps predict-
ably) underscored the findings in the initial streptomycin trial.
These early experiments clearly demonstrated the potential for drug-resistant
strains of M. tuberculosis to emerge in a relatively short period of time following
exposure to a single drug. Consequently, the use of single-drug regimens was dis-
couraged and fell out of practice on a programmatic basis. On a worldwide level,
relatively little attention was paid to the issue of drug resistance as a public health
issue for the first several decades of the chemotherapy era.
By the 1980s, a short-course regimen consisting of isoniazid, rifampin, pyrazin-
amide, and ethambutol for 2 months followed by a continuation phase of isoniazid
and rifampin for an additional 4 months (a 6-month course in total) had been shown
to be effective in achieving cure in nearly all cases, and this became and has
remained the standard regimen for treating tuberculosis around the world [8]. When
administered correctly, this regimen should easily guard against the emergence of
drug resistance.
rapid rise in overall TB cases in the United States that began in the mid-1980s and
peaked in 1992. This rise in cases has been attributed to several factors, including
the emergence of the HIV epidemic; the deterioration and neglect of the public
health tuberculosis control infrastructure; social conditions such as drug use, prison
overcrowding, and homelessness; and poor infection control in hospitals and other
congregate facilities.
In 1992, there were 3811 cases of tuberculosis in New York City (an incidence
rate of 50/100,000), and of those roughly 12%, an astonishing percentage, had
MDR-TB [15]. Patients infected with drug-resistant strains were more likely to have
been HIV-infected and to have been previously treated for TB than patients with
drug-susceptible strains. Several instances of nosocomial transmission of MDR-TB,
usually among patients with HIV infection, were also documented using molecular
epidemiology techniques, in several American cities. Extremely worrisome was the
very high mortality rate associated with MDR-TB. In an early report from New York
City, MDR was almost always a fatal disease, with reported mortality rates in excess
of 80% [16].
Following these reports concerning the emergence of drug resistance in New York
City, a landmark global survey was commissioned by the World Health Organization
(WHO) and published in 1998 [17]. In many countries, particularly those with the
highest burden of tuberculosis cases, resources were (and are still) not sufficient to
perform drug susceptibility testing on all M. tuberculosis isolates. The global
MDR-TB survey effort involved a network of 14 reference laboratories around the
world and the development of a standardized methodology for defining and estimat-
ing the number of patients with MDR. The results of the first global MDR-TB sur-
vey were disturbing. Nearly 10% of patients with TB were infected with strains that
had at least one type of drug resistance. Resistance to INH alone was found in 7.3%
of cases. Although only 1.4% of the total cases in the survey were MDR, this distri-
bution was far from uniform, as several hotspots were identified. In the original
survey, prevalence rates of MDR-TB in Estonia and Latvia exceeded 10% among
patients who had never been treated for TB. In patients previously treated, rates of
MDR exceeded 10% in Argentina, Cuba, the Dominican Republic, England and
Wales, Estonia, Latvia, Peru, Portugal, Puerto Rico, South Korea, Romania, Russia,
Sierra Leone, and Spain. Over half of Latvian re-treatment patients were infected
with MDR strains!
Subsequently, regular annual surveys of the prevalence of drug resistance have
been published, and the resulting information has been included in the World Health
Organization’s annual Global TB Report [18]. A more complete, and extremely
frightening, picture of drug-resistant TB around the world has emerged. Several
countries of the former Soviet Union, including Russia, Belarus, Ukraine,
Kazakhstan, and nearby Mongolia, all have rates of drug resistance that surpass
18% in previously untreated cases. In previously treated cases, rates exceed 50% in
several of these countries.
In terms of sheer numbers, China contributes a large percentage of the world’s
drug-resistant cases, even though the percentage of MDR cases there is much lower
than in Russia and the countries of the former Soviet Union.
166 N. W. Schluger
As has been detailed in the WHO’s annual Global TB Report, precise knowledge
about the prevalence of drug-resistant tuberculosis is limited by the inability of
many national TB control programs to conduct drug susceptibility testing for all
diagnosed cases of TB [19]. This limitation is due to a combination of factors,
including clinical diagnosis where culture methods are unavailable. Even when cul-
turing is performed, comprehensive drug susceptibility testing is too complex and
expensive in many resource-constrained settings. Testing has been a particular prob-
lem in Africa, where WHO maps indicate that no data is available for many of the
countries. It is hoped that introduction of genotypic drug susceptibility testing,
based either on nucleic acid amplification or whole-genome sequencing, will pro-
vide more robust and accessible drug susceptibility testing than traditional pheno-
typic testing using solid or liquid media.
At the time of this writing, isoniazid monoresistance is quite common around the
world, with a prevalence of above 10% in many regions [19]. Evidence of the clini-
cal consequences of INH monoresistance is somewhat mixed. Some data suggest
that patients with this pattern of resistance can be treated with the standard regimen
for drug-susceptible tuberculosis and achieve cure and relapse rates similar to
patients with fully susceptible TB. But other studies have suggested that the stan-
dard short-course regimen is associated with higher treatment failure and relapse
rates when used for INH-monoresistant disease [20–23].
Rifampin monoresistance tends to be lower than with other agents. Many experts
feel that this is because rifampin was introduced relatively late compared with the
other first-line drugs and has never been used as a single agent to treat
tuberculosis.
have received intermittent high doses of rifampin [28–31]. As rifampin is the most
active drug in the standard treatment regimen, treatment of rifampin monoresistance
generally falls under the umbrella of treatment for multidrug-resistant
tuberculosis.
Pyrazinamide (PZA) monoresistance has been somewhat difficult to detect
because of the stringent acidic conditions under which PZA phenotypic testing must
be performed. Thus, PZA monoresistance may be more common than previously
thought [32]. Patients with PZA monoresistance require 9 months of treatment with
INH and rifampin rather than the shorter course 6-month regimen.
The term polydrug resistance is used to describe patients who are infected with
isolates that are resistant to isoniazid and one of the other first-line drugs (pyrazin-
amide, ethambutol, streptomycin) but not rifampin. Some data indicate that persons
with polydrug resistance have worse outcomes than patients with fully drug-
susceptible tuberculosis when treated with standard regimens, though all these data
are retrospective. Optimal treatment regimens have yet to be defined for polyresis-
tant strains of M. tuberculosis.
Multidrug-resistant tuberculosis (MDR) is generally treated with regimens that
include one of the advanced fluoroquinolones (levofloxacin, moxifloxacin, or gati-
floxacin) and an injectable agent (streptomycin, amikacin, kanamycin, or capreo-
mycin). In carefully monitored settings, favorable outcomes for patients with
MDR-TB can be achieved in 80–85% of cases using drug therapy alone [33].
The term pre-XDR-TB has come into use recently to describe isolates of M.
tuberculosis that are resistant to INH, rifampin, and either a fluoroquinolone or one
of the injectable drugs but not both. XDR-TB (extensively drug-resistant TB) indi-
cates an isolate that is resistant to INH, rifampin, fluoroquinolones, and the inject-
ables [34]. Both pre-XDR and XDR-TB are extremely difficult to treat successfully
and require long duration of therapy with drugs that have a high rate of potentially
serious adverse effects. Newer agents, such as bedaquiline, delamanid, and line-
zolid, may be useful in treating these infections, although the optimal dosing regi-
mens with these newer drugs remain to be defined. In cases of XDR and pre-XDR-TB,
surgery may be useful as an adjunctive therapy if the disease is localized.
of Xpert that tests for resistance to a large number of drugs: isoniazid, rifampin,
fluoroquinolones, and aminoglycosides.
As molecular biology techniques have become more automated and less expen-
sive, the realistic prospect of using whole-genome sequencing (WGS) has emerged
as a means for detecting drug resistance [49]. WGS in theory should be the most
complete means of examining DNA sequences for mutations associated with drug
resistance, whereas PCR-based methods have not, up to now, been able to interro-
gate all mutations in all genes associated with antibiotic resistance. WGS can exam-
ine the entire genome and in theory can detect any mutation known to be associated
with drug resistance. Since this technique has an inherent error rate that could
impair its robustness, the method will have to be evaluated in field trials to move
forward.
It is worth exploring in some depth the role of genotypic detection of drug resistance
in M. tuberculosis in light of developments in molecular biology, gene sequencing,
and highly automated detection of mutations and single nucleotide polymorphisms
that hold the promise for inexpensive, rapid, and accurate detection. As noted above,
the Cepheid GeneXpert platform has now been used for several years to identify M.
tuberculosis in sputum samples and for the rapid detection of rifampin resistance.
Quite recently, a significant expansion of this platform holds the promise of rapid
determination of drug susceptibility for a large number of antibiotics.
Xie and colleagues recently evaluated a rapid molecular drug susceptibility plat-
form for detection of resistance to several anti-TB drugs [45]. This study provides
some of the best data, using the most up-to-date methodology, for evaluating
genotypic and phenotypic approaches to identifying drug-resistant strains of
M. tuberculosis. Using a platform derived from the Cepheid MTB/RIF instrument,
samples, obtained from patients in China and South Korea, were interrogated for
the presence of mutations in several genes associated with antibiotic resistance
(katG, inhA, gyrA, and rrs). In theory, this assay could detect resistance to isonia-
zid, fluoroquinolones, and the aminoglycosides kanamycin and amikacin. Results
from the molecular assay were compared with phenotypic testing performed with
the widely used BACTEC MGIT 960 system, using commonly employed break-
point MICs for the determination of resistance: 0.1 μgm/ml for INH, 0.5 and 2.0
μgm/ml for moxifloxacin, 2 μgm/ml for ofloxacin, 1 μgm/ml for amikacin, and 2.5
μgm/ml for kanamycin. As a further check, sequencing of the target genes was car-
ried out using the Sanger method on all the clinical isolates (aliquots of the same
samples that were used for the phenotypic testing).
Using phenotypic testing as the gold standard, the investigational nucleic acid
amplification platform had a sensitivity and specificity for detection of resistance to
170 N. W. Schluger
drugs as follows: for isoniazid, 83.3% and 99.2%, respectively; for ofloxacin, 88.4%
and 96.6%; for moxifloxacin at an MIC of 0.5 μgm/ml, 87.6% and 94.3%; for moxi-
floxacin at an MIC of 2.0 μgm/ml, 96.2% and 84%; for kanamycin 71.4% and
98.4%; and for amikacin, 70.7% and 99.6%. As compared to direct sequencing of
resistance genes, the nucleic acid amplification method had the following sensitivity
and specificity for individual drugs: for isoniazid, 98.1% and 100%, respectively;
for fluoroquinolones, 95.8% and 100%; for kanamycin 92.7% and 99.6%; and for
amikacin, 96.8% and 100%. There were 13 specimens out of 304 that had mutations
that were not detected by the investigational assay.
These results indicate that the highly automated investigational platform, mod-
eled after the current Cepheid MTB/RIF GeneXpert, is a highly accurate means of
identifying mutations in genes that are associated with resistance to the drugs that
define MDR-TB, pre-XDR-TB, and XDR-TB. As compared with phenotypic test-
ing, specificity for drug resistance was quite good. Thus, a result from the investiga-
tional platform indicating drug resistance could be reliably used to exclude that drug
from a therapeutic regimen in nearly every case. However, the sensitivity of the
investigational platform was not quite as good as the specificity. It was excellent for
high-level resistance to moxifloxacin, somewhat less reliable for low-level resis-
tance to moxifloxacin, for ofloxacin, and for isoniazid, and probably unsatisfactory
for amikacin and kanamycin. Looked at another way, the positive predictive value
for determining drug resistance using the investigational platform was generally
excellent, but the negative predictive value fell short of clinical desirability. If only
the molecular testing were used to guide selection of an antibiotic regimen, as many
as 20% of patients might be treated with drugs, such as aminoglycosides, that would
not be clinically useful and could have serious adverse consequences, such as hear-
ing loss.
The results described above underscore the discrepancies that are sometimes
observed between phenotypic and genotypic drug susceptibility testing. These dis-
crepancies point to at least two possible explanations. First, it is certainly possible
that as yet unidentified mutations in genes not examined are responsible for drug
resistance. Second, it is possible that there are limitations in the critical concentra-
tions used to test drug susceptibility and that the phenotypic results obtained in the
laboratory are in fact not entirely reflective of what might be achieved in clinical
practice using standard dosing of the drugs studied in this experiment. At present,
either of these two possibilities, or both, may be operative (see also Chap. 9 on
heteroresistance).
The investigational nucleic acid platform returned results faster than would have
been obtained using the WHO-recommended line probe assay, and it is more suit-
able for use in local labs, hospitals, and clinics. The line probe assay is generally
confined to use in reference centers in most high-prevalence, low-resource settings.
Since many of these settings already have experience using the MTB/RIF version of
the platform, the only change would be in the cartridge that is used; the remaining
hardware is identical. Also, for the purposes of the study, cultures using MGIT were
highly controlled for quality; that level of quality might not be seen under field con-
ditions, whereas the performance of the DNA amplification platform is fairly robust
under most conditions.
5 Drug Resistance in Tuberculosis 171
5.7.1 Isoniazid
Isoniazid, or isonicotinic acid hydrazide (INH), was one of the earliest antitubercu-
losis drugs to be developed, coming into clinical use in the 1950s. Eventually,
Winder and Collins demonstrated convincingly that INH worked by inhibiting syn-
thesis of mycolic acids, the long-chain α-alkyl β-hydroxy fatty acids that are essen-
tial components of the mycobacterial cell wall [50, 51]. INH is a prodrug that is
metabolized by the catalase-peroxidase enzyme KatG (encoded by the katG gene)
and then binds to NAD to form an INH-NAD adduct as an intermediate form [52].
The INH-NAD adduct binds to and inhibits a reductase called InhA. InhA plays an
important role in fatty acid synthesis leading to mycolic acid production; conse-
quently, when its action is blocked by the binding of the INH-NAD adduct, cell wall
synthesis is interrupted, thereby accounting for the anti-TB action of isoniazid.
Resistance to INH occurs more commonly than resistance for any other first-line
agent used for the treatment of tuberculosis, and it can develop in a number of ways.
First, initial activation of the prodrug can be inhibited by mutations in the catalase-
peroxidase enzymes that are required to create the active form of the molecule.
Thus, katG mutations cause INH resistance; they are found in a high percentage of
clinical M. tuberculosis isolates, ranging from 30% to 90% of INH-resistant strains
[53, 54]. Second, the inhibition of InhA by the binding of the INH-NAD adduct can
be inhibited by mutations in the InhA gene, which can also lead to resistance to
INH. The latter mechanism of resistance is observed less commonly than katG
mutations, although it is nonetheless clinically significant [55]. Overall, mutations
responsible for isoniazid resistance occur at a frequency of 10−5 to 10−6 bacilli,
probably the highest frequency of naturally occurring resistance to any of the first-
line antituberculosis drugs.
172 N. W. Schluger
Although complete deletion of katG has been known to account for resistance to
INH in clinical isolates, mutation within that gene is a much more common mode of
resistance. Over 300 different mutations in katG have been identified. Most muta-
tions, however, occur at codon S315, where each base (AGC) can be found mutated
(mutants have Thr, Asn, Arg, Ile, Gly, or Leu residues). Interestingly, certain muta-
tions are more likely associated with monoresistant strains, while others are more
likely to be found in multidrug-resistant strains [56]. InhA mutations are less often
the cause of INH resistance; they are often seen in cases of low- rather than high-
level resistance, but they are clinically significant nonetheless [53].
Mutations in several other genes that are involved in the action of INH have been
reported to cause resistance to isoniazid [57]. These include mutations in furA (a
gene whose product regulates katG expression), sigI (a sigma factor that regulates
katG expression), and glf (which encodes an NAD+ − and flavin adenine
dinucleotide-dependent UDP galactopyranose mutase). These seem to be less com-
mon or important clinically as mediators of isoniazid resistance. As will be dis-
cussed below, the relative contribution of these various mutations in clinical isolates
is important for evaluating diagnostic tests that rely on genotypic rather than pheno-
typic approaches to identifying drug-resistant strains.
Rifampin was first studied as a potential antituberculosis drug in the 1960s, and
resistant strains were quickly identified. Rifampin is a potent, sterilizing drug, with
MICs for susceptible strains generally less than 1.0 μgm/ml [58]. It is the corner-
stone of all modern short-course regimens for the treatment of tuberculosis, and its
introduction and use allowed treatment of tuberculosis to be reduced first to 9- and
then to the current standard 6-month short-course regimen. The introduction of
rifampin made the quite toxic injectable agents, notably streptomycin, obsolete in
the vast majority of cases of drug-susceptible tuberculosis.
Unlike isoniazid, which has several genes that can be associated with the devel-
opment of resistance, resistance to rifampin (and the other clinically used rifamy-
cins, rifabutin, and rifapentine) is controlled almost exclusively by a single gene,
rpoB [59, 60]. Rifampin interrupts RNA synthesis through binding to the beta sub-
unit of RNA polymerase, and mutations in the gene that encodes that subunit, rpoB,
prevent the binding of rifampin and allow RNA synthesis to continue unimpeded.
The binding site of rifampin onto RpoB occurs upstream from its catalytic center,
and elongation of the RNA chain cannot occur.
Mutations of rpoB occur at a natural frequency of 10−7–10−8, much less com-
monly than mutations associated with isoniazid resistance. Detection of rifampin
resistance by genotypic methods has been aided by the fact that nearly all the rpoB
mutations that are associated with resistance occur in a very small hotspot in the
gene, in an 81-base-pair segment [59, 60]. It is felt that mutations in this hotspot
region account for greater than 95% of clinical cases of rifampin resistance. Thus,
5 Drug Resistance in Tuberculosis 173
genotypic tests for rifampin resistance were among the first to come into widespread
clinical use. This is particularly important given rifampin’s importance in the treat-
ment of tuberculosis. Testing for rifampin resistance alone allows one to exclude the
possibility of MDR-TB (if rpoB is wild-type) or to alter therapy to include the use
of potent second-line agents, such as injectable aminoglycosides and fluoroquino-
lones, if rpoB mutations are detected. In general, mutations in rpoB produce high-
level resistance, requiring MICs of greater than 32 μgm/ml, a serum concentration
not generally achievable with safe and well-tolerated dosing of rifampin. As noted
above, all rifamycins have the same mechanism of action, so any detected mutation
in rpoB should be taken as evidence for resistance. However, it has been noted that
particular mutations, at codons 511, 516, 518, and 522, are associated with retention
of susceptibility to rifabutin [61]. These mutations are less common, however.
Rifampin was the last of the so-called first-line drugs to be developed and to
come into widespread clinical use. It was added to already effective regimens to
allow treatment shortening, but it was never used alone. Likely because of this rea-
son, rifampin monoresistance is substantially less common than multidrug resis-
tance. Recent surveys bear this out. In Peru, a country where there had previously
been a high prevalence of MDR-TB, only 2% of cases were found to have rifampin
monoresistance [62]. Among a recently reported California cohort of HIV-infected
patients (a group previously identified as being at higher risk of rifampin monore-
sistance), only 0.4% were found to have strains with this pattern of resistance [63].
MDR-TB was seen in only 1.5%. As if to underscore the effect of immunocompro-
mise on the development of rifampin monoresistance (RMR), rates of RMR and
MDR were sharply lower in the era of highly effective antiviral therapy than they
were in the pre-HAART era. Although uncommon, in both the Peruvian and
California cohorts, RMR strains were significantly associated with an increased risk
of death, at least double that of patients infected with fully drug-susceptible strains.
In the Western Cape province of South Africa, a region burdened by extraordinarily
high rates of both tuberculosis and HIV infection, the number of cases of RMR-TB
seems to be rising, as a recent report documents a tripling of the number of cases in
a relatively short time frame. This seems to be recapitulating the experience of
rifampin monoresistance in New York City [31, 64] from the early to mid-1990s,
where a sudden rise in cases was noted among patients with HIV infection. At the
time, the incidence of tuberculosis in the city had risen to 50/100,000, the equivalent
of a medium-burden country, and roughly one-third of all persons with tuberculosis
were also infected with HIV at a time when effective antiretroviral therapy was
unavailable for most patients.
5.7.3 Pyrazinamide
resistance, the use of genotypic testing has rapidly increased. pncA is 558 base pairs
long, and mutations can occur all along the gene. Thus determining PZA resistance
is a more difficult diagnostic problem than interrogating the 81-base pair hotspot
region of rpoB that is responsible for rifampin resistance.
A recent review of global PZA resistance pooled a large number of studies to
develop estimates of phenotypic and genotypic resistance [32]. Resistance was seen
in every region of the globe. Allowing for the fact that there is probable ascertain-
ment bias in deciding which strains to test for PZA resistance, results are nonethe-
less striking. By phenotypic testing, 16.2% of isolates from a worldwide collection
were determined to be resistant. Geographically, this ranged from 11.4% of M.
tuberculosis isolates in the European region to 21.9% in the Americas. Among
strains at high risk for MDR, PZA resistance was found in 41.3% of patients, and in
strains collected from people with MDR-TB, PZA resistance was found in a stag-
gering 60.5%. In reviewing genotypic testing results, Whitfield and colleagues
determined that 20.9% of a sample of 8651 reported cases showed resistance. As
seen in previous reports of smaller sample sizes, the pncA mutations were diverse
and widely scattered; 608 unique polymorphisms were noted at 397 positions in the
gene. Although some polymorphisms were more common than others, the 20 most
common accounted for only one-third of all isolates with phenotypic resistance.
These more recent estimates by Whitfield differ somewhat from a slightly older
review published by Chang and colleagues [69]. They found that PZA resistance
was present in 51% of multidrug-resistant strains of M. tuberculosis and in only 5%
of non-MDR strains. The differences between these two reviews can be at least
partially explained by sampling or ascertainment bias, as well as some differences
in methodology of resistance detection. It is likely that in the future, more precise
estimates of the prevalence of PZA resistance will be made as molecular testing
becomes used more widely around the world.
better and more consistent care, and by 2011, among smear-positive patients, only
3.3% had streptomycin monoresistance. INH resistance was also lower than in
many parts of the world, at 4.7%. Kanamycin resistance was very uncommon,
reported in only 0.2% of strains tested, and MDR-TB was seen in only 1.1% of
cases in the most recent survey. A recent report from Hunan province in China
found streptomycin resistance in 20.5% of cases [78, 79]. Consistent with the dis-
cussion of resistance mechanism, resistance to capreomycin, amikacin, and kana-
mycin was found in only 2.3%, 1.2%, and 1.8% of isolates, respectively. Mutations
in the rrs gene associated with streptomycin resistance were all in the 388–1084 bp
region, whereas mutations in the gene associated with capreomycin, amikacin, or
kanamycin resistance were in the 1158–1674 bp region, as expected.
5.7.5 Fluoroquinolones
Overall, the bulk of studies in this area indicates that gyrA mutations should almost
always be taken to mean that there is high-level resistance to fluoroquinolones.
It is worth noting that the fluoroquinolones are among the most widely used
antibiotics in the world for a host of bacterial infections, and they are available over
the counter and without prescription in many countries. It is probably for this reason
that the prevalence of fluoroquinolone resistance in M. tuberculosis strains is
extraordinarily high in certain locations. Invariably, some patients with tuberculosis
are being treated (either self-treated or under a physician’s care) with fluoroquino-
lones as single agents for misdiagnosed pneumonia or bronchitis, a practice that is
leading to the generation of FQ-resistant tuberculosis. A very recent multi-country
survey of fluoroquinolone resistance in over 5000 tuberculosis patients found that
the prevalence of resistance ranged from 0.5% to 12.4% for levofloxacin and 0.9%
to 14.6% for moxifloxacin when using a breakpoint definition of resistance of MIC
greater than 0.5 μgm/ml, with the highest rates of resistance occurring in Pakistan
[91]. Using an MIC of 2 μgm/ml as a resistance breakpoint for moxifloxacin was
uncommon, regardless of the country examined.
5.7.6 Bedaquiline
Studies of the molecular basis of resistance focus on two genes. One, atpE,
encodes the previously mentioned F1/F0 ATP synthase; mutations in this gene have
been reported in about 30% of M. tuberculosis strains that have phenotypic bedaqui-
line resistance. Resistance through this mechanism seems to be mediated by abnor-
malities in the C subunit of the ATP synthase.
In the majority of reports, resistance to bedaquiline appears to be mediated by a
gene called rv0678, which encodes a protein of the same name. All bedaquiline-
resistant isolates identified in South Africa to date have had mutations in this gene.
(Of note, mutations in this gene also confer resistance to clofazimine, a second-line
drug that had long been relegated to the end of the therapeutic line, but which
recently regained some prominence for the treatment of multidrug-resistant strains,
particularly as a component of the Bangladesh regimen [95, 96].) The rv0678 pro-
tein encodes a transcriptional repressor of the genes encoding the MmpS5-MmpL5
efflux pump, and there is evidence that resistance via this mechanism can be over-
come in part by the use of efflux pump inhibitors such as verapamil. However, the
feasibility of such an approach in clinical practice is unclear, since high, toxic levels
of verapamil may be needed to achieve this effect. The naturally occurring fre-
quency of resistance to bedaquiline is on the order of 10−8, similar to that of rifampin;
thus resistance should, in general, emerge slowly. However, bedaquiline also has a
very long tissue half-life, which may favor selection of resistant strains of M. tuber-
culosis in clinical use. There has not yet been enough experience with this drug to
make confident statements about the likelihood of resistance emerging as a signifi-
cant clinical issue.
5.7.7 Clofazimine
As noted above, clofazimine has long been relegated to the bottom of the list of
second-line agents, primarily because of its side effect profile, which includes skin
hyperpigmentation, which can be quite marked and bothersome to many patients,
although it is generally said to be a reversible condition. Clofazimine is now a com-
ponent of the 9-month Bangladesh regimen for the treatment of MDR-TB that has
recently been endorsed for use in selected cases by the WHO [96]. Additionally,
recent use of clofazimine in murine models of tuberculosis has encouraged recon-
sideration its potential as a treatment-shortening agent for patients with drug-
susceptible tuberculosis [97].
Even some 50 years after its identification as a drug useful in the treatment of
tuberculosis, the precise mechanism of clofazimine action remains obscure. There
is some evidence that it exerts its effect through disruption of cellular redox cycling,
and there is also evidence that the drug directly disrupts the cell membrane [95].
Clinical isolates that are resistant to clofazimine seem extremely rare, although
resistant strains have been developed in this laboratory. Most of those laboratory
strains have resistance mutations in the same rv0678 gene that was described above
as causing resistance to bedaquiline [95]. This gene is a transcriptional regulator
5 Drug Resistance in Tuberculosis 181
that represses expression of mmpS5-mmpL5, the gene that encodes an efflux pump
of the same name. Interestingly, the rv0678 locus mmpS5-mmpL5 is absent in M.
leprae, the causative agent of leprosy, and no strain of this mycobacterium has been
found that is resistant to clofazimine. There have been other genes (rv1979c and
rv2535c) that have been associated with clofazimine resistance, but the mechanisms
for their involvement have not been identified. As expected, Rv0678 mutations lead
to cross-resistance with bedaquiline.
Although the majority of cases of antituberculosis drug resistance occur singly for
each drug that is used to treat infection, in at least some cases, efflux pumps are
associated with resistance to multiple drugs, as seen with cancer chemotherapy and
which seems to occur at least in vitro in cases of cross-resistance to bedaquiline and
clofazimine.
Recently, there has been evidence that efflux pumps may be playing a role in
resistance to multiple drugs used to treat tuberculosis. As described by Almeida da
Silva, there are five superfamilies of efflux pumps: the ATP-binding cassette (ABC),
major facilitator superfamily (MFS), resistance nodulation division (RND), small
multidrug resistance (SMR), and multidrug and toxic-compound extrusion (MATE)
[98]. The ABC superfamily is of particular interest in M. tuberculosis. Roughly
2.5% of all genes in M. tuberculosis encode ABC transporters, and more than 12
efflux pump genes have been identified. Genes of other superfamilies have also
been implicated in tuberculosis drug resistance. Several experiments have shown
that MICs of antituberculosis drugs can be reduced by administration of efflux
pump inhibitors such as verapamil. This suggests strongly that these pumps can
play a role in at least some forms of drug resistance.
At the outset, it is important to state that there is no well-defined regimen for the
treatment of most forms of drug-resistant tuberculosis, whether monoresistant,
polyresistant, multidrug resistant, pre-extensively drug resistant, or extensively
drug resistant. Treatment recommendations for all of these forms of resistance rely
on expert opinion based on experience, observational cohorts, small case series, and
extrapolations from clinical trials early in the antibiotic era when drugs such as
rifampin were not yet available. Newer drugs that have become available in the last
few years, such as delamanid, bedaquiline, fluoroquinolones, and linezolid, have
also been studied mostly as single agents added to optimized background regimens
rather than as part of novel regimens tested in randomized controlled trials.
182 N. W. Schluger
5.9 Conclusions
Major Points
• Tuberculosis is a serious, often fatal, airborne disease if untreated or
untreatable.
• Drug resistance emerges so readily within individual patients that tuberculosis
serves as a paradigm for implementing combination therapy.
• Most of the antituberculosis drugs are preferentially active with M. tuberculosis;
thus long treatment times do not contribute to resistance in other pathogens (fluo-
roquinolones are an exception).
• Effective treatment regimens exist, even for MDR-TB; however, the necessary
infrastructure is often lacking in resource-challenged countries, which then
become “breeding grounds” for new drug-resistant strains.
• The increasing prevalence of resistance requires that new antituberculosis agents
be developed and administered with sufficient resources and evidence-based
knowledge to stem what is becoming a global healthcare crisis.
References
6. The prevention and treatment of isoniazid toxicity in the therapy of pulmonary tuberculosis.
Bull World Health Organ. 1963;29:457–81.
7. British Medical Research Council. Rate of inactivation of isoniazid in South Indian patients
with pulmonary tuberculosis*. Bull World Health Organ. 1961;25:765–77.
8. Combs DL, O’Brien RJ, Geiter LJ. USPHS Tuberculosis Short-Course Chemotherapy Trial
21: effectiveness, toxicity, and acceptability. The report of final results. Ann Intern Med.
1990;112:397–406.
9. Centers for Disease Control. Primary resistance to antituberculosis drugs—United States.
MMWR Morb Mortal Wkly Rep. 1983;32:521–3.
10. Doster B, Caras GJ, Snider DE. A continuing survey of primary drug resistance in tubercu-
losis, 1961 to 1968. A U.S. Public Health Service cooperative study. Am Rev Respir Dis.
1976;113:419–25.
11. Hobby GL, Johnson PM, Lenert TF, et al. A continuing study of primary drug resistance
in tuberculosis in a Veteran population within the United States. I. Am Rev Respir Dis.
1964;89:337–49.
12. Kopanoff DE, Kilburn JO, Glassroth JL, Snider DE Jr, Farer LS, Good RC. A continu-
ing survey of tuberculosis primary drug resistance in the United States: March 1975 to
November 1977. A United States Public Health Service cooperative study. Am Rev Respir
Dis. 1978;118:835–42.
13. Frieden TR, Sterling T, Pablos-Mendez A, Kilburn JO, Cauthen GM, Dooley SW. The emer-
gence of drug-resistant tuberculosis in New York City. N Engl J Med. 1993;328:521–6.
14. Frieden TR, Fujiwara PI, Washko RM, Hamburg MA. Tuberculosis in New York City—turn-
ing the tide. N Engl J Med. 1995;333:229–33.
15. Health NYCDo. New York City annual tuberculosis report 1992. 1992.
16. Park MM, Davis AL, Schluger NW, Cohen H, Rom WN. Outcome of MDR-TB patients,
1983–1993. Prolonged survival with appropriate therapy. Am J Respir Crit Care Med.
1996;153:317–24.
17. Pablos-Mendez A, Raviglione MC, Laszlo A, et al. Global surveillance for antituberculosis-
drug resistance, 1994–1997. World Health Organization-International Union against
Tuberculosis and Lung Disease Working Group on Anti-Tuberculosis Drug Resistance
Surveillance. N Engl J Med. 1998;338:1641–9.
18. Zignol M, Dean AS, Falzon D, et al. Twenty years of global surveillance of antituberculosis-
drug resistance. N Engl J Med. 2016;375:1081–9.
19. World Health Organization. Global tuberculosis report. Geneva: World Health Organization;
2016.
20. Baez-Saldana R, Delgado-Sanchez G, Garcia-Garcia L, et al. Isoniazid mono-resistant tuber-
culosis: impact on treatment outcome and survival of pulmonary tuberculosis patients in
Southern Mexico 1995–2010. PLoS One. 2016;11:e0168955.
21. Chien JY, Chen YT, Wu SG, Lee JJ, Wang JY, Yu CJ. Treatment outcome of patients with
isoniazid mono-resistant tuberculosis. Clin Microbiol Infect. 2015;21:59–68.
22. Gegia M, Winters N, Benedetti A, van Soolingen D, Menzies D. Treatment of isoniazid-
resistant tuberculosis with first-line drugs: a systematic review and meta-analysis. Lancet
Infect Dis. 2017;17:223–34.
23. Nolan CM, Goldberg SV. Treatment of isoniazid-resistant tuberculosis with isoniazid,
rifampin, ethambutol, and pyrazinamide for 6 months. Int J Tuberc Lung Dis. 2002;6:952–8.
24. Five-year follow-up of a controlled trial of five 6-month regimens of chemotherapy for pul-
monary tuberculosis. Hong Kong Chest Service/British Medical Research Council. Am Rev
Respir Dis. 1987;136:1339–42.
25. Controlled trial of 2, 4, and 6 months of pyrazinamide in 6-month, three-times-weekly regi-
mens for smear-positive pulmonary tuberculosis, including an assessment of a combined
preparation of isoniazid, rifampin, and pyrazinamide. Results at 30 months. Hong Kong
Chest Service/British Medical Research Council. Am Rev Respir Dis. 1991;143:700–6.
186 N. W. Schluger
69. Chang KC, Yew WW, Zhang Y. Pyrazinamide susceptibility testing in Mycobacterium
tuberculosis: a systematic review with meta-analyses. Antimicrob Agents Chemother.
2011;55:4499–505.
70. Davies J, Gorini L, Davis BD. Misreading of RNA codewords induced by aminoglycoside
antibiotics. Mol Pharmacol. 1965;1:93–106.
71. Finken M, Kirschner P, Meier A, Wrede A, Bottger EC. Molecular basis of streptomycin
resistance in Mycobacterium tuberculosis: alterations of the ribosomal protein S12 gene
and point mutations within a functional 16S ribosomal RNA pseudoknot. Mol Microbiol.
1993;9:1239–46.
72. Honore N, Cole ST. Streptomycin resistance in mycobacteria. Antimicrob Agents Chemother.
1994;38:238–42.
73. Nair J, Rouse DA, Bai GH, Morris SL. The rpsL gene and streptomycin resistance in
single and multiple drug-resistant strains of Mycobacterium tuberculosis. Mol Microbiol.
1993;10:521–7.
74. Spies FS, da Silva PE, Ribeiro MO, Rossetti ML, Zaha A. Identification of mutations related
to streptomycin resistance in clinical isolates of Mycobacterium tuberculosis and possible
involvement of efflux mechanism. Antimicrob Agents Chemother. 2008;52:2947–9.
75. Alangaden GJ, Kreiswirth BN, Aouad A, et al. Mechanism of resistance to amikacin and
kanamycin in Mycobacterium tuberculosis. Antimicrob Agents Chemother. 1998;42:1295–7.
76. Suzuki Y, Katsukawa C, Tamaru A, et al. Detection of kanamycin-resistant Mycobacterium
tuberculosis by identifying mutations in the 16S rRNA gene. J Clin Microbiol. 1998;36:1220–5.
77. Maus CE, Plikaytis BB, Shinnick TM. Mutation of tlyA confers capreomycin resistance in
Mycobacterium tuberculosis. Antimicrob Agents Chemother. 2005;49:571–7.
78. Zhao LL, Liu HC, Sun Q, et al. Identification of mutations conferring streptomycin resistance
in multidrug-resistant tuberculosis of China. Diagn Microbiol Infect Dis. 2015;83:150–3.
79. Zhao LL, Chen Y, Chen ZN, et al. Prevalence and molecular characteristics of drug-
resistant Mycobacterium tuberculosis in Hunan, China. Antimicrob Agents Chemother.
2014;58:3475–80.
80. WHO treatment guidelines for drug-resistant tuberculsis (2016 update). Geneva: World
Health Organization; 2016.
81. Gillespie SH, Crook AM, McHugh TD, et al. Four-month moxifloxacin-based regimens for
drug-sensitive tuberculosis. N Engl J Med. 2014;371:1577–87.
82. Jindani A, Harrison TS, Nunn AJ, et al. High-dose rifapentine with moxifloxacin for pulmo-
nary tuberculosis. N Engl J Med. 2014;371:1599–608.
83. Merle CS, Fielding K, Sow OB, et al. A four-month gatifloxacin-containing regimen for treat-
ing tuberculosis. N Engl J Med. 2014;371:1588–98.
84. Drlica K, Malik M. Fluoroquinolones: action and resistance. Curr Top Med Chem.
2003;3:249–82.
85. Alangaden GJ, Manavathu EK, Vakulenko SB, Zvonok NM, Lerner SA. Characterization
of fluoroquinolone-resistant mutant strains of Mycobacterium tuberculosis selected in the
laboratory and isolated from patients. Antimicrob Agents Chemother. 1995;39:1700–3.
86. Takiff HE, Salazar L, Guerrero C, et al. Cloning and nucleotide sequence of Mycobacterium
tuberculosis gyrA and gyrB genes and detection of quinolone resistance mutations.
Antimicrob Agents Chemother. 1994;38:773–80.
87. Zhou J, Dong Y, Zhao X, et al. Selection of antibiotic-resistant bacterial mutants: allelic
diversity among fluoroquinolone-resistant mutations. J Infect Dis. 2000;182:517–25.
88. Avalos E, Catanzaro D, Catanzaro A, et al. Frequency and geographic distribution of gyrA
and gyrB mutations associated with fluoroquinolone resistance in clinical Mycobacterium
tuberculosis isolates: a systematic review. PLoS One. 2015;10:e0120470.
89. Farhat MR, Jacobson KR, Franke MF, et al. Gyrase mutations are associated with vari-
able levels of fluoroquinolone resistance in Mycobacterium tuberculosis. J Clin Microbiol.
2016;54:727–33.
90. Farhat MR, Mitnick CD, Franke MF, et al. Concordance of Mycobacterium tuberculo-
sis fluoroquinolone resistance testing: implications for treatment. Int J Tuberc Lung Dis.
2015;19:339–41.
5 Drug Resistance in Tuberculosis 189
Audrey N. Schuetz
6.1 Introduction
A. N. Schuetz (*)
Mayo Clinic, Rochester, MN, USA
e-mail: [email protected]
Fig. 6.1 Minimal inhibitory concentration (MIC) distribution curve of meropenem for 976
Bacteroides fragilis isolates with Clinical and Laboratory Standards Institute (CLSI) breakpoints
noted
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 193
Fig. 6.2 Graphic display of the general process by which breakpoints are determined
the lowest drug concentration that blocks visible bacterial growth. In the agar dilu-
tion assay, the antimicrobial is placed in agar on which a dilute bacterial culture at a
standardized concentration is applied (a set of plates is prepared in which the anti-
microbial concentration is varied). The drug concentration that inhibits colony for-
mation after suitable incubation is taken as the MIC. Another method that is used by
clinical laboratories is called agar gradient diffusion. Examples of this commer-
cially available method include the Etest (bioMérieux, Durham, NC) and the MTS
strip (Liofilchem, Italy). Plastic or paper strips are impregnated on one face with
concentrations of an antimicrobial in a gradient fashion. When the strip is placed on
an agar plate that has been inoculated with a standardized number of bacteria, the
antimicrobial agent diffuses onto the agar and inhibits bacterial growth where con-
centrations are above the MIC. The MIC is read from a scale printed on the strip –
the point at which the ellipse of bacterial growth meets the strip is taken as
MIC. Results from agar gradient diffusion assays may not always correlate well
with the gold standards of broth microdilution and agar dilution, but these commer-
cial strips are widely used due to their ease of use [6]. Although appropriate for use
with aerobic organisms, disk diffusion testing (placement of antibiotic-impregnated
disks onto an agar plate in which a bacterial culture grows) is not suggested by the
CLSI as a test method for anaerobes due to inaccurate results and poor correlation
with the agar dilution method [3]. However, some research groups are currently try-
ing to develop EUCAST disk diffusion breakpoints for testing of the B. fragilis
group organisms [7]. Unfortunately, MIC results from these various AST methods
do not always correlate well with each other. Thus, knowledge of the testing method
used to obtain a particular data set is important when assessing the literature or
reviewing the MIC of a particular isolate. Moreover, the assays can be difficult to
adapt for routine testing by clinical microbiology laboratories. The important point
is that published reports using anaerobic AST may vary in resistance rates simply
from differences among the methods used to determine the MIC.
AST for anaerobes is still an unsettled situation. For example, CLSI recommends
agar dilution as the testing method for anaerobes, but few clinical laboratories use
this methodology [3]. Use of broth microdilution is limited to the B. fragilis group.
CLSI is now undertaking studies to reevaluate whether broth microdilution can be
performed for anaerobes other than the B. fragilis group, but results for other
anaerobes, such as C. difficile, do not look promising [8]. Moreover, broth
microdilution itself can be difficult to perform, and restricting it to only one group
of anaerobic organisms limits its utility for clinical laboratories where many types
of anaerobes other than solely members of the B. fragilis group are isolated.
When considering the published literature, careful attention must be placed on
the breakpoints applied to the datasets, because breakpoints can differ between
breakpoint-setting organizations (see Table 6.1 below for examples of differing
breakpoints for CLSI and EUCAST). Thus, when comparing resistance prevalence
between regions, one may expect a higher prevalence of resistance reported for
ertapenem if the EUCAST resistant breakpoint is applied rather than the CLSI
breakpoint, which for this drug is three doubling dilutions higher. The reason for the
196 A. N. Schuetz
Table 6.1 Examples of current resistant breakpoint differences for EUCAST and CLSI
EUCAST resistant CLSI resistant
Antimicrobial Organism(s) breakpoint breakpoint
Penicillin Gram-positive and Gram-negative >0.5 μg/mL ≥2 μg/mL
anaerobes
Piperacillin- Gram-positive and Gram-negative >16/4 μg/mL ≥128/4 μg/mL
tazobactam anaerobes
Ertapenem Gram-positive and Gram-negative >1 μg/mL ≥16 μg/mL
anaerobes
Metronidazole Gram-positive and Gram-negative >4 μg/mL ≥32 μg/mL
anaerobes, except C. difficile
different breakpoints is due in part to the different testing methodologies used in the
data gathered for setting breakpoints and to differences in interpretation of break-
point data.
Updates in bacterial taxonomy have also contributed to confusion in interpreta-
tion of previous reports. Some anaerobes that had previously been grouped within a
particular species are now grouped in a different genus (e.g., Peptoniphilus harei
was previously listed within the Peptostreptococcus genus). Other anaerobes have
been newly described (e.g., Murdochiella asaccharolytica), and it is uncertain
where they would have fallen in prior classification schemes [9, 10]. Taxonomic
refinements are ongoing for anaerobes and will continue to require close attention
until new data sets are accumulated.
Finally, the increasing use of nucleic acid and proteomic methods to identify
anaerobic bacteria has resulted in more accurate differentiation and characterization
of isolates. Thus, more anaerobes are identified to the genus or species level as
compared to previous practices that occasionally reported the organism according
to a morphologic group (e.g., anaerobic Gram-positive cocci). As a result, more
accurate AST results are published on anaerobes when attributed to particular
bacterial species.
In summary, surveys of resistance trends among anaerobes vary in their quality
and size. Resistance patterns differ by geographic region, and some resistance rates
have changed significantly over time. It is unclear how much of the differences
among regions is due to true geographic variation, to methodologic testing variation,
or to application of different breakpoints. Patterns of resistance that deserve close
attention due to rising rates include:
• Resistance to the β-lactam-β-lactamase inhibitor combinations among the B. fra-
gilis group.
• Increasing clindamycin resistance among all anaerobes.
• Metronidazole resistance is no longer limited to the B. fragilis group, as it now
includes Gram-positive cocci and bacilli.
• Resistance of Clostridium to vancomycin.
Moreover, many anaerobes are frequently isolated from polymicrobial infections;
sometimes as many as eight or more different organisms are cultured from the site
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 197
Taxonomic updates and the increased ability to identify organisms more accurately
over time must be taken into account when interpreting AST resistance trends and
patterns over different geographic regions.
of 3% have been noted in some large surveys [22]. Although penicillin susceptibility
rates for Peptoniphilus asaccharolyticus and Anaerococcus prevotii are similar to
those of other anaerobic Gram-positive cocci, high MIC90 (e.g., MIC of 90% of the
isolates) values for clindamycin of >32 μg/mL have been noted with these species
[17, 23].
Peptostreptococcus anaerobius and P. stomatis are taxonomically distinct from
other anaerobic Gram-positive cocci. These species are generally associated with
female genital tract and intra-abdominal infections [10]. Resistance of P. anaerobius
to penicillin was notably high (7%) as compared to <1% for the other anaerobic
Gram-positive cocci assessed in one resistance survey [16]. Antimicrobial
susceptibility comparisons between P. anaerobius and P. stomatis demonstrate
higher resistance rates of the former to amoxicillin-clavulanate and consistently
higher MIC50 and MIC90 values for many other antimicrobials [24]. Thus, species
identification can be an important guide for therapy.
Peptoniphilus harei is another Gram-positive anaerobic coccus isolated from
infections of skin and soft tissue. It generally demonstrates low MICs to a variety of
antimicrobials [25]. This species can be difficult to identify phenotypically, since it
resembles P. asaccharolyticus biochemically, and in the past it has likely been
misidentified as P. asaccharolyticus [25]. Thus, past reports of AST results of P.
asaccharolyticus may have been limited by misidentification.
In general, Veillonella spp. are highly susceptible to several of the anti-anaerobic
agents, such as penicillins, β-lactam-β-lactamase inhibitor combinations,
carbapenems, clindamycin, and metronidazole. Indeed, no resistance to amoxicillin,
amoxicillin-clavulanate, clindamycin, or metronidazole was found in a recent
survey of Veillonella in the Netherlands [26].
While mechanisms of antimicrobial resistance among the anaerobic cocci have not
been extensively studied, Reig and colleagues postulated that penicillin-binding
proteins may account for the elevated MICs to penicillin that are occasionally seen
in Veillonella spp., because β-lactamases were not detected in penicillin-resistant
Veillonella isolates [27]. Likewise, penicillin resistance in Gram-positive cocci is
thought to be due to penicillin-binding proteins [28]. Resistance of anaerobes to
clindamycin is due to methylation of 23S rRNA at the site of drug action in the 50S
subunit of ribosomes [29]. The nitroimidazole nimB gene has been detected in a
large number of anaerobic Gram-positive cocci, but it is not always associated with
metronidazole resistance [30]. In a study by Theron and colleagues, the nimB gene
was present in 19/21 metronidazole-susceptible strains of Gram-positive cocci.
Thus, the presence of the nimB gene is not sufficient in and of itself for expression
of metronidazole resistance.
200 A. N. Schuetz
The most common antimicrobials used to treat CDI are metronidazole and vanco-
mycin. These agents are effective for most cases of CDI, although isolates with
elevated MICs to these antimicrobials have been reported [48]. It is important to
note, however, that the existing clinical breakpoints for C. difficile are based on
systemic infections that utilize systemic rather than intra-luminal (intraintestinal)
pharmacokinetics and pharmacodynamics of the antimicrobials. For example, the
levels of vancomycin orally administered to treat CDI are very high intraluminally
in the intestine (e.g., higher than 1000 μg/mL in feces), which is well above the MIC
resistant breakpoint of >2 μg/mL [49].
Antimicrobial resistance of C. difficile is important to monitor, not only for
patient treatment but also for the epidemiologic associations with certain ribotypes.
For example, the emergence of the hypervirulent C. difficile strain 027/BI/NAP has
been associated with consumption of fluoroquinolones [50]. Even though
fluoroquinolones are not used clinically to treat C. difficile infection, antimicrobial
pressure from widespread use of these agents is believed to have allowed the spread
of this fluoroquinolone-resistant, hypervirulent strain.
Resistance of C. difficile to metronidazole is reported to be relatively infrequent
in most areas of the world, but this conclusion depends upon which breakpoints are
used to interpret the data [51]. CLSI metronidazole breakpoints for C. difficile are
susceptible ≤8 μg/mL, intermediate 16 μg/mL, and resistant ≥32 μg/mL. EUCAST
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 203
breakpoints are lower: susceptible ≤2 μg/mL and resistant >2 μg/mL. In a US-based
survey of 925 isolates obtained in 2011 and 2012, 2.6% displayed resistance to
metronidazole based on the EUCAST breakpoint, while no isolate was resistant
based on the higher CLSI breakpoint [52].
Elevated MICs of C. difficile to vancomycin have also been reported. Most stud-
ies have applied the EUCAST vancomycin breakpoint (susceptible ≤2 μg/mL;
resistant >2 μg/mL), since CLSI has not set a clinical breakpoint for vancomycin
with anaerobes. Surveys have generally reported vancomycin resistance below 5%,
but higher percentages have been reported from smaller studies [52, 53]. Elevated
MICs to rifampin (>16 μg/mL) have also been reported, ranging from 8% to over
50% in different regions [53, 54]. There are no EUCAST or CLSI breakpoints for
rifampin. Fidaxomicin is also a common alternative agent used to treat antibiotic-
refractory CDI or recurrent CDI. Decreased susceptibility to fidaxomicin is currently
rare, although one C. difficile isolate has been reported to have an MIC of 16 μg/
mL. This isolate was discovered in a clinical trial of fidaxomicin from a case of
recurrent CDI [55].
C. difficile has shown resistance to a variety of antimicrobials that include
clindamycin (8–100%), cephalosporins (51%), erythromycin (13–100%), and
fluoroquinolones (47%), based on 30 AST studies published between 2012 and
2015 [51]. Among the cephamycins, resistance to cefoxitin and cefotetan is
common, with at least 80% resistance in tested isolates [51]. In North America,
ribotype 027 is resistant to multiple antimicrobials, including rifampin, clindamycin,
and moxifloxacin [54]. Another hypervirulent strain, ribotype 078, also shows
resistance to ciprofloxacin, moxifloxacin, erythromycin, and imipenem [56]. The
most relevant statement is that elevated MICs to both metronidazole and vancomycin
have been reported among C. difficile, while elevated MICs to fidaxomicin are
relatively rare.
Many of the most common species of non-C. difficile clostridia isolated from human
infections are highly susceptible to antimicrobials having anti-anaerobic activity
(e.g., metronidazole, clindamycin, carbapenems, and piperacillin-tazobactam).
Among the non-C. difficile clostridia species, the “RIC” group – namely, C. ramo-
sum, C. innocuum, and C. clostridioforme – demonstrates the highest resistance
rates. C. clostridioforme produces β-lactamase and thus is resistant to several
β-lactam antibiotics but is susceptible to vancomycin [63, 64]. C. innocuum is resis-
tant to cefoxitin and cefotetan, and it displays high MICs (8–32 μg/mL) to vanco-
mycin [65]. Fortunately, C. innocuum is susceptible to metronidazole. Finally,
C. ramosum also demonstrates high vancomycin MICs. Many species of clostridia
have shown resistance to clindamycin, including C. perfringens, C. ramosum,
C. tertium, and C. sporogenes [66]. In general, clindamycin resistance among clos-
tridia ranges from 10% to 20%, but it varies by species [4]. Metronidazole,
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 205
abdominal), bone, and skin lesions [68]. Although usually recovered from abscesses
as mixtures with other anaerobes and aerobes, Bacteroides spp. are usually recovered
from blood as the only pathogen present.
Close to 97% of all Bacteroides isolates are resistant to penicillin and ampicillin,
principally due to β-lactamase production. However, penicillin-binding proteins can
also be expressed by strains [76]. Penicillin and cephalosporin resistance is mediated
primarily by cepA or cfxA [77]. A chromosomal cephalosporinase is encoded by
cepA, which leads to cephalosporin and aminopenicillin resistance. However,
isolates expressing cepA may remain susceptible to piperacillin and β-lactam-β-
lactamase inhibitor combinations. Porin losses can also lead to increases in β-lactam
MICs, including MICs to the β-lactam-β-lactamase inhibitor combinations, as
demonstrated by one group which reported amoxicillin-clavulanate resistance in a
B. thetaiotaomicron strain [78]. The cfxA gene encodes a broad-spectrum
β-lactamase that is responsible for loss of susceptibility to cefoxitin and several
other β-lactam drugs [79, 80]. Carbapenem resistance is often mediated by a zinc
metallo-β-lactamase, encoded by cfiA, that confers resistance to β-lactams and
β-lactam-β-lactamase inhibitor combinations [75]. cfiA in Bacteroides is silent
unless an insertion sequence element activates it [81]. A report from Turkey noted
10% resistance to carbapenems in B. fragilis group isolates, many of which were
confirmed to carry cfiA [82]. Resistance to clindamycin is mediated by erm, which
is located on transferable plasmids [83]. Metronidazole resistance is linked to nim
(nitroimidazole reductase). Expression of this gene leads to reduction of the nitrate
moiety of metronidazole to an amino derivative, which decreases effectiveness of
the antibiotic. However, the mere presence of nim is not predictive of metronidazole
resistance, as nim can be found in metronidazole-susceptible strains. Of 206 B.
fragilis isolates in one study, nim was detected in 24%, and metronidazole MICs of
the nim-positive isolates ranged from 1.5 (susceptible) to >256 (resistant) μg/mL
[84]. Although many nim-positive Bacteroides strains do not show a loss of
susceptibility, they can be induced to express high MICs to metronidazole by
subinhibitory concentrations of the antibiotic [85]. Resistance to moxifloxacin is
mediated by gyrA mutations and efflux pumps. In summary, there are many different
types of resistance genes that may be carried and expressed by B. fragilis group
isolates.
208 A. N. Schuetz
Most Prevotella spp. are resistant to penicillin and ampicillin due to β-lactamase
production. When penicillin resistance is present in Porphyromonas, Fusobacterium,
or B. wadsworthia, it is usually as a result of β-lactamases [90]. Most β-lactamases
in the Gram-negative anaerobic bacilli other than the B. fragilis group are penicil-
linases, although cephalosporinases have been described in Prevotella spp. [29].
Although anaerobic bacteria cause serious diseases, they have not received as much
attention as their aerobic counterparts, in part because they are more difficult to
cultivate. In addition, anaerobes are often present in polymicrobial aerobic/anaerobic
infections or mixed with other anaerobes; therefore, their exact role in contributing
toward infection can sometimes be difficult to ascertain. Some antimicrobials are
effective against both aerobic and anaerobic bacteria, but certain antimicrobials,
such as metronidazole, are only effective against anaerobes. Since many of the
anaerobes tend to grow slowly in culture, susceptibility testing is slow. For
anaerobes, pathogen identification is often obtained well before antimicrobial
susceptibility test results are available. Thus, the anaerobe antibiogram can be
particularly useful in guiding appropriate empirical choice of antimicrobial. It has
been shown that AST results vary according to the type of assay used; thus, future
anaerobic antimicrobial surveys should clearly present the methodologic method of
testing and the breakpoint cutoff applied. Clear delineation of data will aid data
interpretation. Ideally, authors should also present AST data on different species
separately, rather than lumping together groups of morphologically similar
organisms (e.g., anaerobic cocci).
In addition to the technical issues discussed above, aerobic and anaerobic patho-
gens may differ in virulence. Such activity is not reflected in susceptibility (MIC)
measurements due to the nature of the assays, but it is likely to be clinically impor-
210 A. N. Schuetz
tant for clearing infection. It has been proposed that with aerobic bacteria, reactive
oxygen species contribute to the lethal activity of aerobes (see Chap. 20). These
toxic molecules are not expected to be present in anaerobes. Thus, as resistance
emerges to carbapenems and metronidazole, for example, it may be necessary to
develop more agents that are specifically lethal for anaerobes. Such agents could
have additional utility against both aerobic and anaerobic organisms.
Major Points
• With the increasing use of MALDI-TOF MS to identify anaerobic bacteria in
clinical microbiology laboratories, more accurate species identification and anti-
microbial susceptibility results are becoming available. More accurate differen-
tiation and characterization of strains provide more meaningful AST results.
• When reviewing anaerobic AST data, it is important to be aware of the method
of AST utilized in generating the data and the specific breakpoints applied, since
various methods and breakpoints affect interpretation of data.
• A unified AST testing approach for clinical laboratories needs to be developed
for anaerobes. For example, broth microdilution testing is recommended only for
the B. fragilis group. Agar dilution testing can be performed on all anaerobes,
including the B. fragilis group.
• It is important to recognize that the data presented in the anaerobe antibiogram
from CLSI M100 document [4] are derived from isolates obtained globally and
may not be applicable to isolates within a given hospital or region.
• Taxonomic name changes must be taken into account when assessing antimicro-
bial susceptibility testing results of many anaerobes including the anaerobic
Gram-positive cocci, the propionibacteria, and the clostridia.
• Bacteroides spp. are some of the most frequently encountered anaerobes in the
clinical microbiology laboratory and are also among the most virulent. They are
typically more resistant than other anaerobes to antimicrobial agents.
• Certain patterns of resistance in anaerobes that are on the rise include resistance
to the β-lactam-β-lactamase inhibitor combinations among the B. fragilis group,
increasing clindamycin resistance among all anaerobes, and resistance of
Clostridium to vancomycin. In addition, metronidazole resistance is no longer
limited to the B. fragilis group, as it now includes Gram-positive cocci and
bacilli. Although the prevalence of resistance to certain antimicrobials is rising,
effective antimicrobials are generally available.
References
1. Brook I, Wexler HM, Goldstein EJ. Antianaerobic antimicrobials: spectrum and susceptibility
testing. Clin Microbiol Rev. 2013;26(3):526–46.
2. CLSI. Analysis and presentation of cumulative antimicrobial susceptibility test data; approved
guideline, 4th ed. CLSI M39-A4. Wayne, PA: Clinical and Laboratory Standards Institute;
2014.
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 211
3. CLSI. Methods for antimicrobial susceptibility testing of anaerobic bacteria; approved stan-
dard, 8th ed. CLSI M11-A8. Wayne, PA: Clinical and Laboratory Standards Institute; 2012.
4. CLSI. Performance standards for antimicrobial susceptibility testing. 27th ed. CLSI supple-
ment M100. Wayne, PA: Clinical and Laboratory Standards Institute; 2017.
5. Hastey CJ, Boyd H, Schuetz AN, Anderson K, Citron DM, Dzink-Fox J, et al. Changes in the
antibiotic susceptibility of anaerobic bacteria from 2007–2009 to 2010–2012 based on the
CLSI methodology. Anaerobe. 2016;42:27–30.
6. Poulet PP, Duffaut D, Lodter JP. Evaluation of the Etest for determining the in-vitro sus-
ceptibilities of Prevotella intermedia isolates to metronidazole. J Antimicrob Chemother.
1999;43(4):610–1.
7. Nagy E, Justesen US, Eitel Z, Urban E, Infection ESGoA. Development of EUCAST disk dif-
fusion method for susceptibility testing of the Bacteroides fragilis group isolates. Anaerobe.
2015;31:65–71.
8. Hastey CJ, Dale SE, Nary J, Citron D, Law JH, Roe-Carpenter DE, et al. Comparison of
Clostridium difficile minimum inhibitory concentrations obtained using agar dilution vs broth
microdilution methods. Anaerobe. 2017;44:73–7.
9. Munson E, Carroll KC. What’s in a name? New bacterial species and changes to taxonomic
status from 2012 through 2015. J Clin Microbiol. 2017;55(1):24–42.
10. Song Y, Finegold SM. Peptostreptococcus, Finegoldia, Anaerococcus, Peptoniphilus,
Veillonella, and other anaerobic cocci. In: Jorgensen JH, Pfaller MA, editors. Manual of clini-
cal microbiology. 11th ed. Washington, DC: ASM Press; 2015. p. 909–39.
11. Murdoch DA. Gram-positive anaerobic cocci. Clin Microbiol Rev. 1998;11(1):81–120.
12. Brook I. Recovery of anaerobic bacteria from clinical specimens in 12 years at two military
hospitals. J Clin Microbiol. 1988;26(6):1181–8.
13. Brazier J, Chmelar D, Dubreuil L, Feierl G, Hedberg M, Kalenic S, et al. European surveil-
lance study on antimicrobial susceptibility of Gram-positive anaerobic cocci. Int J Antimicrob
Agents. 2008;31(4):316–20.
14. Wren MW. Anaerobic cocci of clinical importance. Br J Biomed Sci. 1996;53(4):294–301.
15. Wildeboer-Veloo AC, Harmsen HJ, Welling GW, Degener JE. Development of 16S rRNA-
based probes for the identification of Gram-positive anaerobic cocci isolated from human
clinical specimens. Clin Microbiol Infect. 2007;13(10):985–92.
16. Brazier JS, Hall V, Morris TE, Gal M, Duerden BI. Antibiotic susceptibilities of Gram-positive
anaerobic cocci: results of a sentinel study in England and Wales. J Antimicrob Chemother.
2003;52(2):224–8.
17. Goldstein EJ, Citron DM, Warren YA, Tyrrell KL, Merriam CV, Fernandez HT. In vitro activities
of dalbavancin and 12 other agents against 329 aerobic and anaerobic Gram-positive isolates
recovered from diabetic foot infections. Antimicrob Agents Chemother. 2006;50(8):2875–9.
18. Mory F, Lozniewski A, Bland S, Sedallian A, Grollier G, Girard-Pipau F, et al. Survey of
anaerobic susceptibility patterns: a French multicentre study. Int J Antimicrob Agents.
1998;10(3):229–36.
19. Goto T, Yamashita A, Hirakawa H, Matsutani M, Todo K, Ohshima K, et al. Complete
genome sequence of Finegoldia magna, an anaerobic opportunistic pathogen. DNA Res.
2008;15(1):39–47.
20. Aldridge KE, Ashcraft D, Cambre K, Pierson CL, Jenkins SG, Rosenblatt JE. Multicenter
survey of the changing in vitro antimicrobial susceptibilities of clinical isolates of Bacteroides
fragilis group, Prevotella, Fusobacterium, Porphyromonas, and Peptostreptococcus species.
Antimicrob Agents Chemother. 2001;45(4):1238–43.
21. Hawser SP. Activity of tigecycline and comparators against recent clinical isolates of
Finegoldia magna from Europe. Eur J Clin Microbiol Infect Dis. 2010;29(8):1011–3.
22. Koeth LM, Good CE, Appelbaum PC, Goldstein EJ, Rodloff AC, Claros M, et al.
Surveillance of susceptibility patterns in 1297 European and US anaerobic and capnophilic
isolates to co-amoxiclav and five other antimicrobial agents. J Antimicrob Chemother.
2004;53(6):1039–44.
212 A. N. Schuetz
23. Citron DM, Kwok YY, Appleman MD. In vitro activity of oritavancin (LY333328), vanco-
mycin, clindamycin, and metronidazole against Clostridium perfringens, Propionibacterium
acnes, and anaerobic Gram-positive cocci. Anaerobe. 2005;11(1–2):93–5.
24. Kononen E, Bryk A, Niemi P, Kanervo-Nordstrom A. Antimicrobial susceptibilities of
Peptostreptococcus anaerobius and the newly described Peptostreptococcus stomatis isolated
from various human sources. Antimicrob Agents Chemother. 2007;51(6):2205–7.
25. Veloo AC, Welling GW, Degener JE. Antimicrobial susceptibility of clinically relevant Gram-
positive anaerobic cocci collected over a three-year period in the Netherlands. Antimicrob
Agents Chemother. 2011;55(3):1199–203.
26. Veloo AC, van Winkelhoff AJ. Antibiotic susceptibility profiles of anaerobic pathogens in the
Netherlands. Anaerobe. 2015;31:19–24.
27. Reig M, Mir N, Baquero F. Penicillin resistance in Veillonella. Antimicrob Agents Chemother.
1997;41(5):1210.
28. Reig M, Baquero F. Antibacterial activity of clavulanate and tazobactam on Peptostreptococcus
spp. J Antimicrob Chemother. 1994;33(2):358–9.
29. Garcia-Rodriguez JA, Garcia-Sanchez JE, Munoz-Bellido JL. Antimicrobial resistance in
anaerobic bacteria: current situation. Anaerobe. 1995;1(2):69–80.
30. Theron MM, Janse Van Rensburg MN, Chalkley LJ. Nitroimidazole resistance genes (nimB) in
anaerobic Gram-positive cocci (previously Peptostreptococcus spp.). J Antimicrob Chemother.
2004;54(1):240–2.
31. Hall V, Copsey SD. Propionibacterium, Lactobacillus, Actinomyces, and other non-spore-
forming anaerobic Gram-positive rods. In: Jorgensen JH, Pfaller MA, editors. Manual of clini-
cal microbiology. 11th ed. Washington, DC: ASM Press; 2015. p. 920–39.
32. Siqueira JF Jr, Rocas IN. Polymerase chain reaction detection of Propionibacterium propioni-
cus and Actinomyces radicidentis in primary and persistent endodontic infections. Oral Surg
Oral Med Oral Pathol Oral Radiol Endod. 2003;96(2):215–22.
33. Cannon JP, Lee TA, Bolanos JT, Danziger LH. Pathogenic relevance of Lactobacillus: a retro-
spective review of over 200 cases. Eur J Clin Microbiol Infect Dis. 2005;24(1):31–40.
34. Land MH, Rouster-Stevens K, Woods CR, Cannon ML, Cnota J, Shetty AK. Lactobacillus
sepsis associated with probiotic therapy. Pediatrics. 2005;115(1):178–81.
35. Rautio M, Saxen H, Siitonen A, Nikku R, Jousimies-Somer H. Bacteriology of histopathologi-
cally defined appendicitis in children. Pediatr Infect Dis J. 2000;19(11):1078–83.
36. Stackebrandt E, Cummins CS, Johnson JL. Family Propionibacteriaceae: the genus
Propionibacterium. In: Dworkin M, Falkow S, Rosenberg E, Schleifer K, Stackebrandt E, edi-
tors. The prokaryotes. New York: Springer; 2006. p. 400–18.
37. Butler-Wu SM, Sengupta DJ, Kittichotirat W, Matsen FA 3rd, Bumgarner RE. Genome
sequence of a novel species, Propionibacterium humerusii. J Bacteriol. 2011;193(14):3678.37.
38. Scholz CF, Kilian M. The natural history of cutaneous propionibacteria, and reclassifica-
tion of selected species within the genus Propionibacterium to the proposed novel genera
Acidipropionibacterium gen. nov., Cutibacterium gen. nov. and Pseudopropionibacterium gen.
nov. Int J Syst Evol Microbiol. 2016;66(11):4422–32.
39. Oprica C, Nord CE, Bacteria ESGoARiA. European surveillance study on the antibiotic sus-
ceptibility of Propionibacterium acnes. Clin Microbiol Infect. 2005;11(3):204–13.
40. Tyrrell KL, Citron DM, Warren YA, Fernandez HT, Merriam CV, Goldstein EJ. In vitro
activities of daptomycin, vancomycin, and penicillin against Clostridium difficile, C.
perfringens, Finegoldia magna, and Propionibacterium acnes. Antimicrob Agents Chemother.
2006;50(8):2728–31.
41. Smith AJ, Hall V, Thakker B, Gemmell CG. Antimicrobial susceptibility testing of Actinomyces
species with 12 antimicrobial agents. J Antimicrob Chemother. 2005;56(2):407–9.
42. Goldstein EJC, Citron DM. Resistance trends in antimicrobial susceptibility of anaerobic bac-
teria, part I. Clin Microbiol Newsl. 2011;33(1):1–8.
43. Goldstein EJ, Citron DM, Merriam CV, Warren YA, Tyrrell KL, Fernandez HT. In vitro
activities of the new semisynthetic glycopeptide telavancin (TD-6424), vancomycin, dapto-
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 213
mycin, linezolid, and four comparator agents against anaerobic Gram-positive species and
Corynebacterium spp. Antimicrob Agents Chemother. 2004;48(6):2149–52.
44. Edmiston CE, Krepel CJ, Seabrook GR, Somberg LR, Nakeeb A, Cambria RA, et al. In
vitro activities of moxifloxacin against 900 aerobic and anaerobic surgical isolates from
patients with intra-abdominal and diabetic foot infections. Antimicrob Agents Chemother.
2004;48(3):1012–6.44.
45. CLSI, editor. Methods for antimicrobial dilution and disk susceptibility testing of infre-
quently isolated or fastidious bacteria. 3rd ed. CLSI guideline M45. Wayne, PA: Clinical and
Laboratory Standards Institute; 2015.
46. Salminen MK, Rautelin H, Tynkkynen S, Poussa T, Saxelin M, Valtonen V, et al. Lactobacillus
bacteremia, species identification, and antimicrobial susceptibility of 85 blood isolates. Clin
Infect Dis. 2006;42(5):e35–44.
47. Lawson PA, Citron DM, Tyrrell KL, Finegold SM. Reclassification of Clostridium difficile as
Clostridioides difficile (Hall and O'Toole 1935) Prévot 1938. Anaerobe. 2016 Aug;40:95–9.
48. Goudarzi M, Goudarzi H, Alebouyeh M, Azimi Rad M, Shayegan Mehr FS, Zali MR, et al.
Antimicrobial susceptibility of clostridium difficile clinical isolates in Iran. Iran Red Crescent
Med J. 2013;15(8):704–11.
49. Baines SD, Wilcox MH. Antimicrobial resistance and reduced susceptibility in Clostridium
difficile: potential consequences for induction, treatment, and recurrence of C. difficile infec-
tion. Antibiotics (Basel). 2015;4(3):267–98.
50. He M, Miyajima F, Roberts P, Ellison L, Pickard DJ, Martin MJ, et al. Emergence and global
spread of epidemic healthcare-associated Clostridium difficile. Nat Genet. 2013;45(1):109–13.
51. Spigaglia P. Recent advances in the understanding of antibiotic resistance in Clostridium dif-
ficile infection. Ther Adv Infect Dis. 2016;3(1):23–42.
52. Snydman DR, McDermott LA, Jacobus NV, Thorpe C, Stone S, Jenkins SG, et al. U.S.-Based
National Sentinel Surveillance Study for the epidemiology of Clostridium difficile-associated
diarrheal isolates and their susceptibility to fidaxomicin. Antimicrob Agents Chemother.
2015;59(10):6437–43.
53. Freeman J, Vernon J, Morris K, Nicholson S, Todhunter S, Longshaw C, et al. Pan-European
longitudinal surveillance of antibiotic resistance among prevalent Clostridium difficile ribo-
types. Clin Microbiol Infect. 2015;21(3):248 e9–e16.
54. Tenover FC, Tickler IA, Persing DH. Antimicrobial-resistant strains of Clostridium difficile
from North America. Antimicrob Agents Chemother. 2012;56(6):2929–32.
55. Goldstein EJ, Babakhani F, Citron DM. Antimicrobial activities of fidaxomicin. Clin Infect
Dis. 2012;55(Suppl 2):S143–8.
56. Keessen EC, Hensgens MP, Spigaglia P, Barbanti F, Sanders IM, Kuijper EJ, et al. Antimicrobial
susceptibility profiles of human and piglet Clostridium difficile PCR-ribotype 078. Antimicrob
Resist Infect Control. 2013;2:14.
57. Peng Z, Jin D, Kim HB, Stratton CW, Wu B, Tang YW, et al. Update on antimicrobial resis-
tance in Clostridium difficile: resistance mechanisms and antimicrobial susceptibility testing.
J Clin Microbiol. 2017;55(7):1998–2008.
58. Shen J, Wang Y, Schwarz S. Presence and dissemination of the multiresistance gene cfr in
Gram-positive and Gram-negative bacteria. J Antimicrob Chemother. 2013;68(8):1697–706.
59. O'Connor JR, Galang MA, Sambol SP, Hecht DW, Vedantam G, Gerding DN, et al. Rifampin
and rifaximin resistance in clinical isolates of Clostridium difficile. Antimicrob Agents
Chemother. 2008;52(8):2813–7.
60. Leeds JA, Sachdeva M, Mullin S, Barnes SW, Ruzin A. In vitro selection, via serial passage,
of Clostridium difficile mutants with reduced susceptibility to fidaxomicin or vancomycin.
J Antimicrob Chemother. 2014;69(1):41–4.
61. Stevens DL, Bryant AE, Carroll KC. Clostridium. In: Jorgensen JH, Pfaller MA, editors.
Manual of clinical microbiology. 11th ed. Washington, DC: ASM Press; 2015. p. 940–66.
62. Leal J, Gregson DB, Ross T, Church DL, Laupland KB. Epidemiology of Clostridium species
bacteremia in Calgary, Canada, 2000-2006. J Infect. 2008;57(3):198–203.
214 A. N. Schuetz
63. Dubreuil L, Odou MF. Anaerobic bacteria and antibiotics: what kind of unexpected resistance
could I find in my laboratory tomorrow? Anaerobe. 2010;16(6):555–9.
64. Finegold SM, Song Y, Liu C, Hecht DW, Summanen P, Kononen E, et al. Clostridium clos-
tridioforme: a mixture of three clinically important species. Eur J Clin Microbiol Infect Dis.
2005;24(5):319–24.
65. Finegold SM, Molitoris D, Vaisanen ML. Study of the in vitro activities of rifaximin and com-
parator agents against 536 anaerobic intestinal bacteria from the perspective of potential utility
in pathology involving bowel flora. Antimicrob Agents Chemother. 2009;53(1):281–6.
66. Winn WC, Allen SD, Janda WM, Koneman EW, Procop G, Schrechenberger PC, et al.
Koneman’s color atlas and textbook of diagnostic microbiology. Philadelphia: Lippincott
Williiams & Wilkins; 2005.
67. Ackermann G, Schaumann R, Pless B, Claros MC, Goldstein EJ, Rodloff AC. Comparative
activity of moxifloxacin in vitro against obligately anaerobic bacteria. Eur J Clin Microbiol
Infect Dis. 2000;19(3):228–32.
68. Wexler HM. Bacteroides: the good, the bad, and the nitty-gritty. Clin Microbiol Rev.
2007;20(4):593–621.
69. Snydman DR, Jacobus NV, McDermott LA, Golan Y, Hecht DW, Goldstein EJ, et al. Lessons
learned from the anaerobe survey: historical perspective and review of the most recent data
(2005-2007). Clin Infect Dis. 2010;50(Suppl 1):S26–33.
70. Trevino M, Areses P, Penalver MD, Cortizo S, Pardo F, del Molino ML, et al. Susceptibility
trends of Bacteroides fragilis group and characterisation of carbapenemase-producing strains
by automated REP-PCR and MALDI TOF. Anaerobe. 2012;18(1):37–43.
71. Nagy E, Urban E, Nord CE, Bacteria ESGoARiA. Antimicrobial susceptibility of
Bacteroides fragilis group isolates in Europe: 20 years of experience. Clin Microbiol Infect.
2011;17(3):371–9.
72. Snydman DR, Jacobus NV, McDermott LA, Supran S, Cuchural GJ Jr, Finegold S, et al.
Multicenter study of in vitro susceptibility of the Bacteroides fragilis group, 1995 to 1996,
with comparison of resistance trends from 1990 to 1996. Antimicrob Agents Chemother.
1999;43(10):2417–22.
73. Seifert H, Dalhoff A, Group PS. German multicentre survey of the antibiotic susceptibility of
Bacteroides fragilis group and Prevotella species isolated from intra-abdominal infections:
results from the PRISMA study. J Antimicrob Chemother. 2010;65(11):2405–10.
74. Snydman DR, Jacobus NV, McDermott LA, Golan Y, Goldstein EJ, Harrell L, et al. Update on
resistance of Bacteroides fragilis group and related species with special attention to carbapen-
ems 2006-2009. Anaerobe. 2011;17(4):147–51.
75. Sherwood JE, Fraser S, Citron DM, Wexler H, Blakely G, Jobling K, et al. Multi-drug resistant
Bacteroides fragilis recovered from blood and severe leg wounds caused by an improvised
explosive device (IED) in Afghanistan. Anaerobe. 2011;17(4):152–5.
76. Piriz S, Vadillo S, Quesada A, Criado J, Cerrato R, Ayala J. Relationship between penicillin-
binding protein patterns and beta-lactamases in clinical isolates of Bacteroides fragilis with
different susceptibility to beta-lactam antibiotics. J Med Microbiol. 2004;53(Pt 3):213–21.
77. Aldridge KE, Ashcraft D, O'Brien M, Sanders CV. Bacteremia due to Bacteroides fragilis
group: distribution of species, beta-lactamase production, and antimicrobial susceptibility pat-
terns. Antimicrob Agents Chemother. 2003;47(1):148–53.
78. Behra-Miellet J, Calvet L, Dubreuil L. A Bacteroides thetaiotamicron porin that could take
part in resistance to beta-lactams. Int J Antimicrob Agents. 2004;24(2):135–43.
79. Boente RF, Ferreira LQ, Falcao LS, Miranda KR, Guimaraes PL, Santos-Filho J, et al.
Detection of resistance genes and susceptibility patterns in Bacteroides and Parabacteroides
strains. Anaerobe. 2010;16(3):190–4.
80. Garcia N, Gutierrez G, Lorenzo M, Garcia JE, Piriz S, Quesada A. Genetic determinants
for cfxA expression in Bacteroides strains isolated from human infections. J Antimicrob
Chemother. 2008;62(5):942–7.
6 Anaerobic Bacteria: Antimicrobial Susceptibility Testing and Resistance Patterns 215
81. Soki J, Edwards R, Hedberg M, Fang H, Nagy E, Nord CE, et al. Examination of cfiA-medi-
ated carbapenem resistance in Bacteroides fragilis strains from a European antibiotic suscep-
tibility survey. Int J Antimicrob Agents. 2006;28(6):497–502.
82. Toprak NU, Uzunkaya OD, Soki J, Soyletir G. Susceptibility profiles and resistance genes
for carbapenems (cfiA) and metronidazole (nim) among Bacteroides species in a Turkish
University Hospital. Anaerobe. 2012;18(1):169–71.
83. Roberts MC, Sutcliffe J, Courvalin P, Jensen LB, Rood J, Seppala H. Nomenclature for macro-
lide and macrolide-lincosamide-streptogramin B resistance determinants. Antimicrob Agents
Chemother. 1999;43(12):2823–30.
84. Gal M, Brazier JS. Metronidazole resistance in Bacteroides spp. carrying nim genes and
the selection of slow-growing metronidazole-resistant mutants. J Antimicrob Chemother.
2004;54(1):109–16.
85. Lofmark S, Fang H, Hedberg M, Edlund C. Inducible metronidazole resistance and nim
genes in clinical Bacteroides fragilis group isolates. Antimicrob Agents Chemother.
2005;49(3):1253–6.
86. Roberts SA, Shore KP, Paviour SD, Holland D, Morris AJ. Antimicrobial susceptibility of
anaerobic bacteria in New Zealand: 1999-2003. J Antimicrob Chemother. 2006;57(5):992–8.
87. Liu CY, Huang YT, Liao CH, Yen LC, Lin HY, Hsueh PR. Increasing trends in antimicro-
bial resistance among clinically important anaerobes and Bacteroides fragilis isolates causing
nosocomial infections: emerging resistance to carbapenems. Antimicrob Agents Chemother.
2008;52(9):3161–8.
88. Bahar H, Torun MM, Demirci M, Kocazeybek B. Antimicrobial resistance and beta-lacta-
mase production of clinical isolates of prevotella and porphyromonas species. Chemotherapy.
2005;51(1):9–14.
89. Molitoris E, Wexler HM, Finegold SM. Sources and antimicrobial susceptibilities of
Campylobacter gracilis and Sutterella wadsworthensis. Clin Infect Dis. 1997;25(Suppl
2):S264–5.
90. King A, Downes J, Nord CE, Phillips I, European Study G. Antimicrobial susceptibility of
non-Bacteroides fragilis group anaerobic Gram-negative bacilli in Europe. Clin Microbiol
Infect. 1999;5(7):404–16.
Chapter 7
Clinical Significance and Biologic Basis
of HIV Drug Resistance
Rodger D. MacArthur
7.1 Introduction
R. D. MacArthur (*)
Medical College of Augusta at Augusta University, Augusta, GA, USA
e-mail: [email protected]
Mutations in various structural HIV proteins are random events. They occur as long
as the virus is replicating. These two statements result in Key Concept 1: the only
way to prevent HIV viral resistance from occurring is to shut down viral replication
completely. This concept and other key concepts are summarized in Table 7.1. There
are two corollaries to this concept: (a) If the virus is not replicating, it is not mutat-
ing; and (b) if the virus is replicating, whether in the presence or absence of antiret-
roviral therapy (ART), it is mutating. The rate of mutation development in various
structural HIV proteins is impressive. For instance, it is estimated that in the absence
of antiretroviral therapy, every single possible mutation in every single position of,
say, the 99 amino acid protease, occurs once each day. Of course, many of these
mutations are “dead-end” mutations that are not compatible with viral existence.
Other mutations result in a viral “quasispecies” that is replication-competent but has
a lower replicative capacity (i.e., is less “fit”) than “wild-type” virus. Due to limita-
tions in typically used commercial resistance assays, these less-fit subpopulations
are not detected until their numbers exceed 10–20% of the total viral pool. However,
in the presence of drug, they can become the dominant population, as Fig. 7.1
shows. In addition, when using more sensitive assays that can detect these mutated
quasispecies down to a level of 1%, it has been shown that, when present, they
adversely affect treatment options [4]. Even very low levels of viral replication,
occurring in the presence of drug, will result in the accumulation of drug-limiting
resistance mutations.
Factors that affect the likelihood of selecting for drug-limiting resistance muta-
tions include (a) adherence to the antiretroviral regimen, (b) the potency of the anti-
retroviral drugs, (c) the drug’s intrinsic barrier to resistance, and (d) the duration of
time that antiretroviral drug levels exceed zero but are lower than the level needed
to achieve maximal viral suppression. Intermediate rates of adherence (e.g., 20–80%
of drug taken) typically are associated with the greatest risk for selecting drug-
limiting resistance mutations. Lower adherence rates (e.g., below 20%) usually
result in drug levels too low to exert any selective pressure on the virus, and higher
adherence rates usually result in drug levels high enough to maximally suppress
viral replication. On the other hand, factors such as poor absorption of the drugs, or
poor bioavailability, will change the dynamics such that higher levels of adherence,
resulting in higher but suboptimal drug levels, will lead to substantial selective pres-
sure on the virus to mutate. Similarly, lower-potency antiretrovirals, or lower-
potency combinations of antiretrovirals, have the potential to exert substantial
selective pressure on the virus if maximal suppression of replication does not occur.
In vitro, potency of an antiretroviral drug is expressed as percent inhibition of HIV
for any given drug concentration (Fig. 7.2). Clinically, potency of an antiretroviral
drug is expressed as the log reduction in HIV RNA level when that drug is used
220 R. D. MacArthur
alone. Different drugs and different classes have different intrinsic potencies. Log
reductions in HIV RNA level for each drug contribute to the potency of the entire
regimen in an additive fashion: the anticipated potency of any particular a ntiretroviral
regimen is the sum of the potencies (in log HIV RNA reduction) of the individual
components. Several key concepts follow: Key Concept 2 – if a drug or combination
of drugs is unable to completely shut down viral replication, the higher the drug
levels, the greater the selective pressure on the virus to mutate. Key Concept 3: if a
drug has no activity, there will be no selective pressure on the virus, and no further
mutations will occur. Key Concept 4: the longer drug levels exceed zero but are
lower than the level needed to completely shut down viral replication, the greater is
the selective pressure on the virus to mutate.
Mutations that occur as a result of selective pressure of a drug have the effect of
increasing the amount of drug necessary to achieve the same degree of killing
(Fig. 7.3). That is, graphically, they shift the “kill curve” to the right. If the shift
results, for instance, in ten times more drug necessary to get the same amount of
killing, this effect often is referred to as a tenfold increase in resistance to a particu-
lar drug. These mutations may have no effect on other antiretrovirals of the same
class, or they may have the same effect, a greater effect, or a lesser effect. In some
situations (discussed later), these mutations may even shift the kill curve to the left
for another drug in the same class, effectively resulting in increased activity of that
drug. This phenomenon is referred to as “hypersusceptibility” of HIV to the drug
with the left-shifted curve.
If the initial mutation that occurs as a result of selective pressure of a particular
drug is not sufficient to eliminate all activity of that drug, additional mutations
will accumulate in the presence of that drug until that drug (and typically many
others in the same antiretroviral class) no longer has any activity against the virus.
In some circumstances, these subsequent mutations will not decrease the activity
of the offending drug but rather restore fitness (replicative capacity) to the virus.
7 Clinical Significance and Biologic Basis of HIV Drug Resistance 221
There are three HIV enzymes that are the main targets for antiretrovirals: reverse
transcriptase, protease, and integrase. Reverse transcriptase is a 560 amino acid
(AA) heterodimer comprised of p51 and p66 subunits. The p51 subunit contains
the first 450 amino acids; the p66 subunit contains the rest [5]. Most resistance
assays stop sequencing around AA 250 or so, as most of the relevant mutations
occur from AA 41 through AA 219. Protease is composed of two non-covalently
associated, structurally identical monomers of 99 amino acids each [5]. Integrase is
a 288 AA dimer comprised of 3 domains: the N-terminal domain, the catalytic core
domain, and the C-terminal domain. The catalytic core contains the triad of AAs
comprising the D,D-35-E motif, made up of aspartic acid at positions 64 and 116
and glutamine at position 152. Mutations at any of these positions essentially abol-
ish all catalytic activity of integrase [6]. As such, these mutations occur uncom-
monly. Most of the relevant mutations selected by exposure to the integrase strand
transfer inhibitor class of antiretrovirals occur in the catalytic core domain, from
AA 74 through AA 230.
By convention, mutations at any position in these enzymes are listed as the one-
letter AA code that is expected at a particular numbered position, followed by the
one-letter AA code of the mutated quasispecies. So, for instance, the very com-
monly occurring valine-for-methionine mutation at position 184 of reverse tran-
scriptase is listed as M184V. Often, multiple mutations can be seen (detected) at
each position, including position 184 of reverse transcriptase. The convention is to
list these mutations separated by a slash (e.g., M184V/I). Not every detectable
mutation at a given position results in the same loss of activity to a particular drug.
In addition, because three nitrogenous bases encode each amino acid, and multiple
triads can code for the same amino acid, reverting (or back reverting) to wild type
from a mutated quasispecies often involves going through an intermediate form [7].
So, for instance, the zidovudine-selected mutation, T215Y, tends to mutate to T215S
prior to reverting to wild type. Typically, the back revertants (e.g., T215S) tend to be
fitter than the original mutated version. Finally, mutations such as T215Y that
222 R. D. MacArthur
Table 7.3 are never selected during combination therapy. For instance, the M184V
mutation is not selected when abacavir is used in combination with zidovudine,
likely due to the hypersusceptibility that M184V imparts to zidovudine. The extent
of resistance conferred to a particular drug by a particular mutation often depends
on both the drug and the mutation. So, for instance, the M184V mutation confers
high-level resistance to lamivudine and emtricitabine, hypersusceptibility to zid-
ovudine, tenofovir, and stavudine, and only partial resistance to abacavir. On the
other hand, the mutations that are selected by both zidovudine and stavudine,
referred to as “thymidine analogue mutations (TAMs),” confer some degree of resis-
tance to all of the NRTIs.
There are some mutations that are associated with properties not limited to one
or a few of the NRTIs. Foremost among these is the M184V mutation. This muta-
tion, while having the effect on individual drugs described above, also results in a
viral quasispecies that has substantial loss in fitness. This phenomenon, first
described by Mark Wainberg and colleagues in Montreal [8], results clinically in a
sustained 0.5 log reduction in HIV RNA; at the same time the virus has developed
up to 1000-fold resistance to either lamivudine or emtricitabine [9]. In addition, this
mutation results in a viral quasispecies that has greater “fidelity,” meaning that it is
more difficult for the virus to revert to wild type [10].
Other mutations of note that are selected by exposure to the NRTIs are (1)
Q151M (along with accessory mutations at positions 62, 75, 77, and 116) which
conveys high-level resistance to all NRTIs; (2) 69 insertion complex, which
refers to amino acid insertions between codons 67 and 70 and conveys resistance
to all NRTIs; (3) 70 deletion complex, which refers to amino acid deletions
between codons 67 and 70 and conveys resistance to all NRTIs; and (4) E40F,
E44D/A, and V118I are referred to as “accessory” mutations that are seen occa-
sionally in association with the TAMs and contribute to the extent of resistance
conferred by the TAMs.
Mutations, in general, impart some degree of structural change to the relevant
enzyme. In the case of NRTI-limiting mutations, the resulting structural changes to
224 R. D. MacArthur
reverse transcriptase have one of two effects: they prevent the nucleoside analogue
from binding, due to steric hindrance, or they facilitate excision of the already incor-
porated analogue (Fig. 7.4). The excision activity of HIV reverse transcriptase cata-
lyzes the 3′-terminal residue of the analogue to an acceptor substrate, most likely
ATP [11]. Some mutations, such as K65R, have the effect of decreasing binding/
incorporation for several of the NRTIs, while at the same time having a variable
effect on the stability of the incorporated drug. So, for instance, the effect of K65R
is to make the virus hypersusceptible to zidovudine, susceptible to stavudine and
abacavir, and more resistant to didanosine and tenofovir.
The NNRTI class of antiretrovirals also inhibit HIV reverse transcriptase,
although the mechanism is not as straightforward as with the NRTIs. Unlike the
NRTIs, the NNRTIs are structurally quite diverse. However, all of them bind allo-
sterically to reverse transcriptase in the same general region, albeit to different
amino acids. The NNRTI-binding pocket of reverse transcriptase is hydrophobic,
located about 10 angstroms from the catalytic site in the “palm” region of the p66
subunit [12]. Binding of the NNRTIs induces conformational changes that inhibit
the catalytic activities of reverse transcriptase. Drugs in this class are more potent
than the NRTIs, typically able to decrease HIV RNA levels by 2–4 logs or more.
However, single-point mutations at positions 103 and 181 (K103N, Y181C) devel-
oped quickly in the presence of the “first generation” of NNRTIs, nevirapine, dela-
virdine, and efavirenz [13]. These mutations alter the size, shape, and polarity of the
NNRTI-binding pocket, thereby reducing access to the pocket by those NNRTIs
[12]. Interestingly, although not clinically relevant due to the lack of current use of
zidovudine, it was shown that the addition of zidovudine to nevirapine changed the
mutation selection pattern from predominantly Y181C to K103N [14]. The “second-
generation” NNRTIs, etravirine and rilpivirine, retain good activity against HIV in
7 Clinical Significance and Biologic Basis of HIV Drug Resistance 225
the presence of the K103N mutation but have reduced activity in the presence of the
Y181C/I/V mutation. Other mutations that limit the activity of etravirine and rilpi-
virine include V90I, E138A/K, and M230L. The effect of these mutations on the
NNRTIs is summarized in Table 7.4.
The protease inhibitor class of antiretrovirals, when “boosted” with the cytochrome
p450 inhibitors ritonavir or cobicistat, is the most potent of all of the classes. The
drugs in this class inhibit a key enzyme (protease) in the replicative cycle of
HIV. This enzyme cleaves unnecessary pieces of viral protein just prior to maturity
of the virus, thereby allowing the virus to fold into its compact, infectious, shape/
form. The class is the least susceptible to the development of resistance mutations,
when combined with one of the boosters. Even monotherapy (with a booster) results
in sustained HIV RNA decreases of 4–6 logs. It is the only class for which mono-
therapy can be considered, although it is not recommended to do so routinely. It is
extraordinarily rare for any mutations to develop in previously antiretroviral-naïve
individuals when on a standard multidrug combination. Mutations have developed,
rarely, in persons on lopinavir and atazanavir, for instance, when each is boosted
with ritonavir. Much of our knowledge of the virologic consequences of the devel-
opment of mutations in this class comes from 20 years ago before the use of a
booster was common. For instance, unboosted indinavir selected for mutations
slowly and retained antiviral activity even in the presence of several mutations. On
the other hand, by the time additional mutations developed, HIV protease typically
was resistant to all of the drugs in the class (unboosted). From a historical perspec-
tive, mutations that developed at key binding locations in protease resulted in
lengthening of the bond between the enzyme and the inhibitor, as well as other
structural changes. For instance, the L33F multidrug resistance mutation [15] has
226 R. D. MacArthur
been shown to reduce flexibility of protease in the 30s and 80s loops [16]. As a
result of early mutational changes in protease, the enzyme typically loses a substan-
tial degree of fitness. Over time, additional mutations develop away from the bind-
ing site that have the effect of restoring fitness by “tightening” the structure of the
enzyme
Certain other unboosted protease inhibitor-selected mutations are now of passing
interest only, as these drugs are not used either at all or in the absence of a booster.
So, for instance, the PI nelfinavir selected for the D30N mutation, which substan-
tially limited the activity of that drug but had no effect on other PIs. The atazanavir-
selected I50L mutation resulted in high-level resistance to atazanavir but had no
effect on the activity of amprenavir. Similarly, the I50V amprenavir-selected muta-
tion resulted in high-level resistance to amprenavir but had no effect on the activity
of atazanavir. As with the NRTI and NNRTI classes, there were a number of other
“signature” mutations (e.g., G48V and L90M for saquinavir) that often would reli-
ably indicate the previous drug exposure history of the virus.
The integrase strand transfer inhibitor (INSTI) class also is extremely potent, result-
ing in sustained reductions in HIV RNA of 4–5 logs. However, the class is more
susceptible to the development of drug-limiting resistance mutations than is the PI
class. For instance, in the setting of prior virologic failure on an NRTI-based regi-
men, a switch from a boosted PI regimen (lopinavir-ritonavir) to a raltegravir-based
regimen was associated with a substantial virologic failure rate in the raltegravir
arm, despite participants having an HIV RNA level < 400 copies/ml at the time of
the switch [17].
In addition to raltegravir, there are two other currently available integrase strand
transfer inhibitors: elvitegravir, which must be boosted (and coformulated) with
cobicistat, and dolutegravir. All of these drugs work at the last step of integration,
by inhibiting insertion of viral DNA into host chromosomal DNA (strand transfer
step). There is substantial overlap in the mutations that are selected by, and limit the
activity of, raltegravir and elvitegravir. These include T66A/I/K, E92Q, E138K/A/T,
G140S/A/C, Y143R/C/H (raltegravir only), and, especially, Q148H/R/K and
N155H. These mutations are located near the catalytic core domain of integrase,
which is where the INSTIs bind [18]. These mutations have the effect of excluding
binding of the INSTIs but at the expense of reduced enzymatic fitness. As might be
anticipated, a number of accessory mutations often are seen in association with the
major mutations. These secondary mutations have the effect of restoring fitness to
the virus.
Dolutegravir, the third INSTI currently available, is the most resilient and least
susceptible to the development of resistance of the three. In fact, mutations which
substantially limit the activity of both raltegravir and elvitegravir typically have
little impact on the activity of dolutegravir. However, mutations do develop, espe-
7 Clinical Significance and Biologic Basis of HIV Drug Resistance 227
Enfuvirtide is the only drug in the fusion inhibitor class of antiretrovirals. It also is
the only antiretroviral which must be administered by injection. Relatively little is
known about the development of resistance to this drug and even less about the
impact of specific mutations. Furthermore, routine resistance testing is not avail-
able, and the use of the drug is limited to the heavily treatment-experienced group
of patients, those defined by mutations in multiple other classes, and comprising at
most 5% of the total HIV-infected population. Enfuvirtide binds to the HR1 domain
of GP41, after the envelope protein has bound to the CD4 receptor of the cell [19].
Binding of enfuvirtide prevents the conformational changes that are needed for
membrane fusion. While mutations in the envelope gene at positions 36–43 have
been detected in viruses from patients on enfuvirtide, it is not clear what impact any
single or multiple mutations has on drug activity. In fact, most clinical trials using
resistance as an entry criterion default to “ever use” of enfuvirtide to arbitrarily
declare that HIV-1 likely is resistant to that drug.
Maraviroc is the only antiretroviral that targets a human protein, the CCR5 co-
receptor. The drug has good activity in both B- and non-B-subtypes of HIV-1, as
long as the individual is CCR5-tropic and not dual- (CCR5, CXCR4) or mixed-
tropic. The main resistance mechanism involves utilization of maraviroc-bound
CCR5 co-receptors for entry, as a result of multiple mutations in the V3 loop of
GP120 [20]. Another way HIV can enter in the presence of maraviroc is by co-
receptor switching (i.e., utilizing CXCR4). Co-receptor switching occurs infre-
quently, as it results in substantially reduced fitness and efficiency of entry, as a
result of multiple mutations in GP160 [20]. There is no cross-resistance with enfu-
virtide. Like enfuvirtide, maraviroc is used infrequently at the present time. Early
assays used to detect minority variants that were dual- or mixed-tropic were not
sensitive enough to exclude this population of patients from entry into clinical
trials.
Mutation “pathways” have been well-described for the thymidine analogue drugs
zidovudine and stavudine, as well as for the INSTI raltegravir. Pathways refer to the
accumulation of specific mutations based on the previous development of other
mutations. Little is known about the factors that “force” the virus down one pathway
228 R. D. MacArthur
or another, but fitness likely plays a role. The two thymidine analogue pathways are
shown in Fig. 7.5. While the mutations seen at the full expression of each pathway
overlap, the process by which they accumulate differs between the pathways. In
addition, it has been observed that the mutation L210W is always found in associa-
tion with T215Y, a mutation that is not present in the “second” TAM pathway. Note
that the first TAM pathway, involving the mutations L210W and T215Y, is the more
common of the two and results in a virus more resistant to all of the NRTIs.
Even less is known about the 2–3 raltegravir pathways, the N155H pathway, the
Q148H/R/K pathway, and the Y143R/C/H pathway. The Q148H/R/K pathway is
associated with the least reduction in viral fitness. Minority variants that exist at the
time of drug initiation may play a role in determining the specific pathway the virus
takes to accumulate mutations.
There are two main types of HIV resistance tests: genotypic tests, based on sequenc-
ing and determining the specific amino acids present at each position in key
enzymes, and phenotypic tests, which test for the degree of killing of the isolated
virus in the presence of different antiretrovirals. Genotypic tests are an indirect mea-
sure of resistance, based primarily on assigning a “likelihood of failure” score to a
particular mutation or set of mutations. Phenotyping is a direct measure of a drug’s
activity in killing the mutated or unmutated virus. Genotyping is quicker, more
readily available (e.g., not limited to commercial laboratories), and less expensive.
7 Clinical Significance and Biologic Basis of HIV Drug Resistance 229
with over 90,000 paired phenotype specimens and assigning weights to each singlet
and pair of mutations associated with different antiretrovirals [26]. By doing so,
Virco™ was able to show that a single mutation (i.e., I84V) that substantially
reduced the activity of, for instance, tipranavir, behaved substantially differently in
combination with the L10F mutation. Today, as the cost of phenotyping has dropped
in North America and Europe, genotype and phenotype tests often are run together.
When genotype tests are run without a phenotype, the reported results typically are
substantiated by a large database of isolates.
The history of treating HIV-infected persons with antiretrovirals has benefitted from
three decades of research on resistance issues specific to these drugs. Treatment
guidelines [1] have evolved as we have acquired a better understanding of the devel-
opment and prevention of resistance to antiretrovirals. For instance, prior to 2005,
the DHHS guidelines recommended baseline resistance testing only in the acutely
or recently infected individual. This recommendation was based on the belief that
drug-selected resistance mutations would be overgrown by (revert to) wild-type
virus in the absence of continuous selective pressure of antiretroviral therapy. That
7 Clinical Significance and Biologic Basis of HIV Drug Resistance 231
Major Points
• Understand basic principles of antiretroviral resistance
• Learn key drug- and class-related resistance mutations
• Apply basic science-related principles of drug resitance to clinical applications
References
1. Panel on Antiretroviral Guidelines for Adults and Adolescents. Guidelines for the Use
of Antiretroviral Agents in Adults and Adolescents Living with HIV. Department of
Health and Human Services. Available at http://aidsinfo.nih.gov/contentfiles/lvguidelines/
AdultandAdolescentGL.pdf. Accessed 15 Nov 2017.
2. World Health Organization. HIV drug resistance report 2017. Available at www.who.int/hiv/
pub/drugresistance/hivdr-report-2017/en. Accessed 17 Nov 2017.
3. Beyrer C, Pozniak A. HIV drug resistance – an emerging threat to epidemic control. N Engl
J Med. 2017;377:1605–7.
4. Simen BB, Simons JF, Hullsiek KH, et al. Low-abundance drug-resistant viral variants in
chronically HIV-infected, antiretroviral treatment-naive patients significantly impact treatment
outcomes. J Infect Dis. 2009;199:693–701.
5. Shafer RW, Jung DR, Betts BJ, et al. Human immunodeficiency virus reverse transcriptase and
protease sequence database. Nucleic Acids Res. 2000;28:346–8.
6. Craigie R. HIV integrase, a brief overview from chemistry to therapeutics. J Biol Chem.
2001;276:23213–6.
7. Kuritzkes DR. A fossil record of zidovudine resistance in transmitted isolates of HIV-1. PNAS.
2001;98:13485–7.
8. Quan Y, Brenner BG, Oliveira M, Wainberg MA. Lamivudine can exert a modest antivi-
ral effect against human immunodeficiency virus type 1 containing the M184V mutation.
Antimicrob Agents Chemother. 2003;47:747–54.
9. Eron J, Benoit S, Jemsek J, et al. The safety and efficacy of lamivudine (3TC)/zidovudine com-
bination therapy vs. zidovudine monotherapy and lamivudine monotherapy in therapy-naive,
HIV-positive patients with CD4 cell counts of 200- 500/mm3: a randomized, double-blind,
multicenter, controlled trial. N Engl J Med. 1995;333:1662–9.
10. Wainberg MA, Drosopoulos WC, Salomon H. Enhanced fidelity of 3TC-selected mutant
HIV-1 reverse transcriptase. Science. 1996;271:1282–5.
11. Acosta-Hoyos A, Scott WA. The role of nucleotide excision by reverse transcriptase in HIV
drug resistance. Virus. 2010;2:372–94.
12. deBéthune MP. Non-nucleoside reverse transcriptase inhibitors (NNRTIs), their discovery,
development, and use in the treatment of HIV-1 infection: a review of the last 20 years (1989 –
2009). Antivir Res. 2010;85:75–90.
232 R. D. MacArthur
13. van den Berg-Wolf M, Hullsiek KH, Peng G, Kozal MJ, Novak RM, Chen L, Crane LR,
MacArthur RD, CPCRA 058 Study Team, the Terry Beirn Community Programs for Clinical
Research on AIDS (CPCRA), and The International Network for Strategic Initiative in Global
HIV Trials (INSIGHT). Virologic, immunologic, clinical, safety, and resistance outcomes
from a long-term comparison of efavirenz-based versus nevirapine- based antiretroviral regi-
mens as initial therapy in HIV-1-infected persons. HIV Clin Trials. 2008;9:324–36.
14. Richman DD, Havlir D, Corbeil J, et al. Nevirapine resistance mutations of human immunode-
ficiency virus type 1 selected during therapy. J Virol. 1994;68:1660–6.
15. Kozal MJ, Hullsiek KH, Leduc R, Novak RM, MacArthur RD, Lawrence J, Baxter JD. Terry
Beirn Community Programs for Clinical Research on AIDS (CPCRA). Prevalence and impact
of HIV-1 protease codon 33 mutations and polymorphisms in treatment-naive and treatment-
experienced patients. Antivir Ther. 2006;11:457–63.
16. Kuiper BD, Keusch BJ, Dewdney TG, et al. The L33F darunavir resistance mutation acts as a
molecular anchor reducing the flexibility of the HIV-1 protease 30s and 80s loops. Biochem
Biophys Rep. 2015;2:160–5.
17. Bianco JL, Gonzalez-Cordón A, Llibre JM, et al. Impact of prior virologic failure and nucelos(t)
ide genotypic resistance mutations on the efficacy of switching from ritonavir-boosted prote-
ase inhibitors to raltegravir. Antivir Ther. 2015;20:487–92.
18. Mbisa JL, Martin SA, Cane PA. Patterns of resistance development with integrase inhibitors in
HIV. Infect Drug Resist. 2011;4:65–76.
19. Reeves JD, Lee F-H, Miamidian JL, et al. Enfuvirtide resistance mutations: impact on human
immunodeficiency virus envelope function, entry inhibitor sensitivity, and virus neutrilazation.
J Virol. 2005;79:4991–9.
20. MacArthur RD, Novak RM. Reviews of anti-infective agents: maraviroc: the first of a new
class of antiretroviral agents. Clin Infect Dis. 2008;47:236–41.
21. Kempf DJ, Isaacson JD, King MS, et al. Identification of genotypic changes in human immu-
nodeficiency virus protease that correlate with reduced susceptibility to the protease inhibitor
lopinavir among protease inhibitor-experienced patients. J Virol. 2001;75:7462–9.
22. Parkin N, Chappey C, Petropoulos C. Improving lopinavir genotype algorithm through
phenotype correlations: novel mutation patterns and amprenavir cross-resistance. AIDS.
2003;17:955–61.
23. Grant P, Wong EC, Rode R, et al. Virologic response to lopinavir-ritonavir-based antiretroviral
regimens in a multicenter international cohort: comparison of genotypic interpretation scores.
Antimicrob Agents Chemother. 2008;52:4050–6.
24. King MS, Rode R, Cohen-Codar I, et al. Predictive genotypic algorithm for virologic
response to lopinavir-ritonavir in protease inhibitor-experienced patients. Antimicrob Agents
Chemother. 2007;51:3076–4.
25. Schapiro JM, Scherer J, Vinisko R, et al. Genotypic tipranavir scores as predictors of response.
11th European AIDS conference; Madrid, Spain; 24–27 October 2007; P3.4/16.
26. Van Houtte M, Picchio G, Van Der Borght K, et al. A comparison of HIV-1 drug susceptibility
as provided by conventional phenotyping and by a phenotype prediction tool based on viral
genotype. J Med Virol. 2009;81:1702–9.
27. Novak RM, Chen L, MacArthur RD, Baxter J, Huppler Hullsiek K, Peng G, Xiang Y, Henley C,
Schmetter B, Uy J, van den Berg-Wolf M, Kozal M, and the Terry Beirn Community Programs
for Clinical Research on AIDS 058 Study Team. Prevalence of antiretroviral drug resistance
mutations in chronically human immunodeficiency virus- infected, treatment-naïve patients:
implications for routine resistance screening before initiation of antiretroviral therapy. Clin
Infect Dis. 2005;40:468–74.
28. Sungkanuparth S, Manosuth W, Kiertiburanakul S, et al. Options for a second-line antiretrovi-
ral regimen for HIV type 1-infected patients whose intial regimen of a fixed-dose combination
of stavudine, lamivudine, and nevirapine fails. Clin Infect Dis. 2007;44:447–52.
29. Stanford University HIV Drug Resistance Database. Available at: https://hivdb.stanford.edu.
Last accessed 19 Nov 2017.
30. Wensing AM, Calvez V, Günthard HF, et al. 2017 update of the drug resistance mutations in
HIV-1. Top Antivir Med. 2017;24:132–41.
Chapter 8
Resistance of Herpesviruses to Antiviral
Agents
Three antiviral agents and a prodrug are currently available for the systemic treat-
ment of human cytomegalovirus (HCMV) infections. Ganciclovir (GCV, Cytovene®,
Roche) is a deoxyguanosine analogue and was the first drug to be approved in 1988.
Since then, it has remained the first-line treatment for HCMV infections in immu-
nocompromised patients. Upon entry in HCMV-infected cells, GCV is selectively
phosphorylated by a viral phosphotransferase (the product of the UL97 gene,
pUL97). Subsequently, cellular kinases convert GCV-monophosphate into GCV-
triphosphate, which acts as a potent inhibitor of the HCMV DNA polymerase (pol)
by competing with deoxyguanosine triphosphate on the enzyme binding site
(Fig. 8.1). Ganciclovir is also incorporated into the viral DNA where it slows down
and eventually stops chain elongation [15, 22, 213]. Ganciclovir formulations are
available for intravenous (IV) or oral administration as well as intravitreal injections
for the treatment of HCMV diseases in immunocompromised patients. Due to its
poor bioavailability (~6%), efforts were made to develop a prodrug of
GCV. Valganciclovir (VGCV, Valcyte®, Roche) is a L-valyl ester formulation of
GCV exhibiting about ten times the bioavailability of GCV following oral adminis-
tration [175].
The other two compounds approved for systemic treatment of HCMV infections
are also potent inhibitors of the viral DNA pol. However, due to their toxicity
W. L. Drew
Departments of Pathology and Laboratory Medicine, University of California,
San Francisco, CA, USA
J. Piret · G. Boivin (*)
Research Center in Infectious Diseases of the CHU of Quebec- Laval University,
Quebec City, Canada
e-mail: [email protected]
Fig. 8.1 Mechanisms of action of the different classes of antiviral agents. The nucleoside ana-
logues such as ganciclovir (GCV), acyclovir (ACV), and penciclovir (PCV) must be first phos-
phorylated by the UL97 protein kinase or viral thymidine kinase (TK) and then by cellular kinases
to be converted into their active forms. The cyclic nucleoside analogue brivudine (BVDU) is con-
verted into monophosphate and diphosphate forms by the viral TK followed by an additional
phosphorylation by cellular kinase. The acyclic nucleoside phosphonate derivatives such as cido-
fovir (CDV) must be phosphorylated by cellular kinases only to be active. The resulting triphos-
phate forms compete with deoxynucleotide triphosphates (dNTPs) to inhibit viral replication. The
pyrophosphate analogue foscarnet (FOS) directly inhibits the activity of the viral DNA poly-
merase. Key: Ⓟ represents one phosphate group
p rofiles and absence of oral formulation, they are usually reserved for patients fail-
ing or not tolerating GCV therapy. Cidofovir (CDV, Vistide®, Gilead Sciences) is a
nucleotide analogue of cytidine (also called acyclic nucleoside phosphonate) that
only requires activation (phosphorylation) by cellular but not viral enzymes to exert
its antiviral activity [61]. The diphosphate form of CDV is a competitive inhibitor of
HCMV DNA pol, and it may act as a DNA chain terminator when two consecutive
incorporations into DNA occur (Fig. 8.1). The IV formulation of CDV is indicated
for the treatment of HCMV retinitis in patients with AIDS and is also occasionally
used in transplant recipients. Foscarnet (FOS, Foscavir®, AstraZeneca), a pyrophos-
phate analogue, differs from the two previous antivirals both by its mechanism of
action and by the fact that it does not require any activation step to exert its antiviral
activity. Foscarnet binds to and blocks the pyrophosphate binding site on the viral
polymerase, thus preventing incorporation of incoming deoxynucleotide triphos-
phate (dNTP) into viral DNA (Fig. 8.1) [59]. The IV formulation of FOS is indi-
cated for the treatment of HCMV retinitis in individuals with AIDS and for
GCV-resistant HCMV infections in immunocompromised patients.
In addition to the treatment of established HCMV disease, antivirals have also
been used to prevent symptomatic episodes, especially in transplant recipients. The
8 Resistance of Herpesviruses to Antiviral Agents 235
Fig. 8.2 Confirmed mutations associated with resistance to ganciclovir in the UL97 gene and to
ganciclovir (GCVR), foscarnet (FOSR), and/or cidofovir (CDVR) in the UL54 gene of clinical
HCMV isolates. In UL97 gene, the ATP-binding site, the phosphate transfer (P-transfer) domain,
the nucleoside-binding site (NBS), and some regions conserved among the protein kinase family
(i.e., I, II, III, VIB, VII, VIII, and IX) are represented by the black boxes. In UL54 gene, conserved
regions among the Herpesviridae DNA polymerases are represented by the black boxes. The
roman numbers (I to VII) and δ-region C corresponding to each of these regions are indicated
above the boxes. Conserved motifs (Exo I, Exo II, and Exo III) in the exonuclease domain are also
indicated above the boxes. Scale represents nucleotide positions in each gene. Bars (│) indicate
amino acid substitutions associated with antiviral drug resistance. ashaded area corresponds to the
codon 590–603 region in UL97 gene where different amino acid deletions conferring ganciclovir
resistance were identified (i.e., deletions 591–594, 591–607, 595, 595–603, 600, and 601–603).
b
amino acid deletion 981–982 in UL54 gene that confers resistance to all three antivirals
GCV resistance was demonstrated after only 7–24 days (median 10 days) of cumu-
lative GCV therapy. Finally, the emergence of GCV-resistant strains has been
recently associated with previously uncommon central nervous system HCMV dis-
ease and retinitis occurring late after HSCT [111, 237].
been studied in marker transfer experiments for their effect on viral fitness. Among
those, mutations T700A and V715M (conserved region II) [11], K513N (δ-region
C) [63], and D301N (Exo I motif) [55] were shown to significantly reduce the yield
of progeny virus in cell culture supernatants, whereas some others (D413E, T503I,
L516R, and E756K/D) were only associated with a modest attenuation of viral rep-
lication [55]. In the case of HCMV DNA pol mutants selected during GCV therapy,
it should be noted that UL97 mutations have been generally shown to emerge first
and to confer a low level of resistance (EC50 < 30 μM), whereas subsequent emer-
gence of UL54 mutations usually leads to a high level of drug resistance
(EC50 > 30 μM) [86, 122, 205]. Most clinical isolates resistant to GCV have muta-
tions in UL97 only. However, isolated UL54 mutations have been reported occa-
sionally in clinical HCMV strains [31].
Fig. 8.3 Suggested algorithm for the management of suspected drug-resistant HCMV infections
in solid organ transplant recipients. Key: GCV ganciclovir, FOS foscarnet, CDV cidofovir, BID
twice a day, IV intravenous, EC50, concentration of antiviral that reduces HCMV replication in
cultured cells by 50% compared to control (without drug) determined in phenotypic assay.
(Adapted from [134])
As mentioned, resistance is more likely when stable or rising viral loads (espe-
cially DNAemia levels) or persistence of clinical symptoms are observed more than
2–3 weeks after initiating appropriate full-dose IV antiviral therapy. In this context,
clinical decisions on disease management should be based on genotypic analysis of
UL97 and UL54 genes, the patient’s immune status (e.g., high-risk D+/R– recipi-
ents, lung transplant recipients), and disease severity (i.e., sight- or life-threatening
conditions) [77, 145] (Fig. 8.3). Despite the limitation mentioned above, genotypic
resistance testing is more practical and rapid (results in 72–96 h) than phenotypic
assays. Thus, rescue therapy should be ideally based on results of the genotypic
assays. In centers where genotypic testing is unavailable or performed infrequently,
initial management should avoid the use of drugs with similar pathways of resis-
tance. For instance, patients failing GCV should be given FOS in the absence of any
sequencing data because of the possibility of UL54 mutations that could confer
resistance to both GCV and CDV. On the other hand, if UL97 and UL54 sequencing
data are available and indicate that only UL97 mutations are present, then CDV
therapy can be attempted. Other empiric options for patients failing GCV therapy
244 W. L. Drew et al.
could consist in re-inducing the patient with higher than normal doses of GCV (up
to 10 mg/kg IV BID) or combination therapy with reduced doses of GCV and FOS
[165, 208] although these strategies are associated with significant toxicity and can
be clinically risky in patients with life- or sight-threatening diseases. Leflunomide,
an anti-inflammatory compound, which appears to inhibit viral capsid assembly, has
not been systemically evaluated but is the subject of successful case reports [182].
When GCV resistance is encountered, discontinuation of the drug and the use of
foscarnet alone may hasten return to “wild,” i.e., sensitive virus [78]. Whenever pos-
sible, improvement of the patient’s immune status (i.e., reduction of immunosup-
pressive regimen in transplant patients or aggressive antiretroviral therapy in AIDS
patients) should also be considered. HCMV viral load should be carefully moni-
tored (once weekly) to quantitate a response to the change in therapy.
Four mechanisms are involved in HSV resistance to ACV and lead to different phe-
notypes: (i) a complete deficiency in viral TK activity (TK-deficient); (ii) a decreased
production of viral TK (TK low producer); (iii) a viral TK protein with altered sub-
strate specificity (TK altered), i.e., the enzyme is able to phosphorylate thymidine,
the natural substrate, but does not phosphorylate ACV; and finally (iv) a viral DNA
pol with altered substrate specificity (DNA pol altered). Alteration or absence of the
TK protein is the most frequent mechanism seen in the clinic, probably because TK
is not essential for viral replication in most tissues and cultured cells [99, 162].
Thymidine kinase mutants resistant to ACV usually exhibit a reduction in fitness
and neurovirulence. In animal models, TK low producers show some reduction in
pathogenicity compared with wild-type strains but are generally able to reactivate.
In contrast, TK-deficient mutants demonstrate impaired pathogenicity as well as a
lower efficiency to establish latency in sensory ganglia and a poor reactivation com-
pared with wild-type strains. However, it has been suggested that ultralow levels of
TK enzyme activity could be sufficient to allow reactivation [18]. Mutants with
altered DNA pol activity exhibit different degrees of neurovirulence attenuation in
animals [2, 67].
The TK phenotype can be determined by the selective incorporation of iodode-
oxycytidine and thymidine into infected cells using plaque autoradiography [156].
However, it can be difficult to evaluate residual TK activity in HSV isolates of
immunocompromised patients in whom heterogeneous populations (TK-competent/
TK-deficient) may coexist [99, 191, 223]. Nonisotopic methods have been devel-
oped using ADP-Glo kinase assay [39, 174] and DiviTum assay based on
8 Resistance of Herpesviruses to Antiviral Agents 245
Antiviral drugs against herpesviruses provide some of the best examples of effective
and selective antiviral therapy. However, drug-resistant viruses have been rapidly
selected in the laboratory and also identified in the clinic. Contrasting with HCMV
resistance data, no extensive survey has been performed to evaluate the rate of emer-
gence of drug-resistant HSV isolates according to the duration of antiviral therapy.
Such study would be a difficult task considering that oral and topical ACV formula-
tions are widely used.
In immunocompetent hosts, HSV resistance to ACV is not a clinically important
problem. Studies have shown that 0.1–0.6% of HSV isolates recovered from
untreated, prophylaxed, or treated immunocompetent subjects harbor a resistant
phenotype to ACV (EC50 ≥ 8.8 μM) as assessed by a plaque reduction assay, and
this seems to reflect the natural occurrence of TK-deficient mutants in a viral popu-
lation [8, 9, 33, 60, 69, 91, 159, 211, 232]. Except for a few notable cases [133, 136,
214], the occasional recovery of ACV-resistant HSV-2 from immunocompetent
hosts has not been associated with clinical failure and proved to be transient [110,
232]. A higher prevalence (6.4%) of ACV-resistant HSV-1 isolates has been reported
in immunocompetent patients with recurrent herpetic keratitis [80], and some of
these cases were clinically refractory to ACV therapy [41, 125, 170, 224]. The lower
immune surveillance in the cornea, which is an immune-privileged site, could
explain the rapid selection of drug-resistant viruses [6]. Herpes simplex virus strains
resistant to ACV are more often isolated in immunocompromised hosts, and such
isolates have been associated with persistent and/or disseminated diseases [26, 47,
60, 84, 89, 114, 162, 184, 228]. In the few clinical surveys reported, the rate of
ACV-resistant HSV isolates has varied from 4.3% to 14% among all immunocom-
promised groups [60, 69, 72, 84, 143, 169, 181, 211]. The prevalence of ACV resis-
tance has ranged from 3.5% to 7% in HIV-positive patients in several studies [60,
84, 143, 149, 181, 239]. It is estimated that 6.5% of HSV isolates obtained from
patients with cancer were resistant to ACV compared to 10% from heart or lung
transplant recipients [60] and 5–14% from other SOT recipients [84]. Of note, high
resistance rates have been reported in HSCT recipients, ranging from 4.1% to 10.9%
8 Resistance of Herpesviruses to Antiviral Agents 247
[47, 69, 85, 96, 162, 229, 233]. In another study, 8% of allogeneic cell transplant
recipients demonstrated persistent HSV excretion despite ACV therapy, whereas
5% of HSV isolates showed significant level of ACV resistance in vitro [42]. Morfin
and Thouvenot reported that patients receiving either autologous or allogenic bone
marrow have a similar incidence, i.e., 9%, of HSV infection, but resistance only
occurred in allogenic transplants, reaching a prevalence of 30% [163]. The severity
of immunosuppression and the prolonged use of ACV are considered two important
factors for the development of drug resistance. The importance of the severity of
immunosuppression is underscored by Langston et al. who studied adult patients
undergoing lymphocyte-depleted hematopoietic progenitor cell transplant from
HLA-matched family donors [140]. All seven evaluable HSV-1 or HSV-2 seroposi-
tive patients reactivated at a median of 40 days posttransplant, and the five strains
tested were all resistant to ACV. Furthermore, FOS resistance developed rapidly in
the three patients treated with this drug [140]. Importantly, the prevalence of ACV-
resistant HSV isolates has remained stable in immunocompromised patients over
the past decades [69, 211], and there has been no unequivocal evidence of transmis-
sion of a resistant HSV strain from person to person.
The emergence of VZV isolates resistant to ACV has not been described in
immunocompetent individuals with primary VZV infection or herpes zoster, except
for one case report of a patient with an ACV-resistant VZV keratitis [109]. Acyclovir-
resistant VZV isolates in the clinic have been mainly found in AIDS patients with
low CD4 cell counts who presented with atypical, disseminated, or relapsing zoster
lesions [27, 161, 189, 216, 226]. Cases of resistance to ACV have also been described
in SOT and HSCT recipients as well as in hemato-oncological patients with VZV
reactivations unresponsive to therapy. In these patients, VZV infections not respond-
ing to ACV therapy persist in the form of chronic skin lesions and are associated
with significant morbidity and mortality due to visceral dissemination. An unusual
verrucous form of VZV infection caused by ACV-resistant mutants has also been
described in some patients [38, 66]. Two cases of immunocompromised children
presenting herpes zoster due to the Oka vaccine strain and who developed chronic,
disseminated drug-resistant VZV infections following ACV therapy have been
reported [38, 142]. However, the prevalence of ACV-resistant cases in these differ-
ent populations is unknown because only case reports have been published so far. It
was reported that 27% of hemato-oncological patients, including HSCT recipients,
with persistent VZV infections had mutations probably associated with ACV resis-
tance [222].
Only a few FOS-resistant HSV (EC50 ≥ 330 μM or ≥ 100 μg/ml and at least a
threefold increase in EC50 value compared with the parental susceptible strain) have
been reported in the clinic [64, 119, 199]. Nine FOS-resistant HSV clinical isolates
from HIV-infected subjects in whom ACV and FOS therapy sequentially failed have
been described [19, 199]. A few reports have described the emergence of VZV
strains resistant to FOS in immunocompromised patients [16, 92, 186, 226, 227].
248 W. L. Drew et al.
Fig. 8.4 Confirmed mutations associated with acyclovir resistance in the UL23 and ORF36 gene
of clinical HSV-1, HSV-2, and VZV isolates. Conserved regions among the thymidine kinases of
Herpesviridae including the ATP-binding site (ATP) and the nucleoside-binding site (NBS) are
represented by the black boxes. Scale represents nucleotide positions in the gene. Bars (│) indicate
amino acid substitutions, whereas dots (●) represent nucleotide additions and/or deletions that
confer resistance to acyclovir
Mutations in the TK of HSV (encoded by the UL23 gene) leading to ACV resistance
consist of either additions or deletions in homopolymer runs of Gs and Cs associ-
ated with a premature stop codon or single nucleotide substitutions in conserved and
non-conserved regions of the gene (Fig. 8.4) [176, 177]. Each mechanism accounts
for approximately 50% of ACV-resistant phenotypes in the clinic [99, 162].
However, recent studies reported an increased proportion of additions/deletions
which accounted for 62% [97] or even 80% [40] of TK gene mutations. Nucleotide
substitutions are scattered within the TK gene including the three catalytic sites of
the enzyme (ATP-binding site, nucleoside-binding site, and a.a. 336) [70]. Albeit
rarely seen in clinic, most mutations in the DNA polymerase of HSV (encoded by
the UL30 gene) conferring drug resistance are located in the conserved regions of
the enzyme, most specifically in regions II, VI, III, and VII (Fig. 8.5) which are
directly or indirectly involved in the recognition and binding of nucleotides or pyro-
phosphate as well as in catalysis [176, 177]. The greatest clusters of mutations in the
DNA pol enzyme have been found in conserved regions II and III. Most FOS-
resistant clinical HSV isolates contain single-base substitutions in conserved regions
II, VI, III, or VII and in a non-conserved region (between I and VII) of the DNA pol.
Some of these isolates retain susceptibility or borderline levels of susceptibility to
ACV and CDV. However, some mutations, in particular in regions II (V715G and
S724N) and VII (Y941H) of the DNA pol, can confer resistance to both ACV and
8 Resistance of Herpesviruses to Antiviral Agents 249
Fig. 8.5 Confirmed mutations associated with resistance to acyclovir (ACVR), foscarnet (FOSR),
and/or cidofovir (CDVR) in the UL30 and ORF28 genes of clinical HSV-1, HSV-2, and VZV iso-
lates. Conserved regions among the Herpesviridae DNA polymerases are represented by the black
boxes. The roman numbers (I to VII) and δ-region C or region A corresponding to each of these
regions are indicated above the boxes. Scale represents nucleotide positions in the gene. Colored
bars (│) indicate amino acid substitutions
FOS. Mutations associated with CDV resistance mapped in DNA pol regions II
(R700M), VI (L773M), III (G841C and G850I), and VII (Y941H) and in δ-region C
located in the Exo III motif (V573M).
A recent study of Topalis et al. showed that mutations conferring drug resistance
in DNA pol of HCMV are mostly detected in the 3′-5′ exonuclease domain (60.6%)
and to a lower extent in the palm (18.2%), fingers (16.7%), and thumb (4.6%)
domains, whereas those identified in DNA pol of HSV are mainly located in the
palm (25.0%), fingers (25.0%), and thumb (21.5%) domains with a lower propor-
tion of mutations being found in the 3′-5′ exonuclease domain (27.3%) [220]. The
different distribution of mutations in DNA pol domains may reflect various mecha-
nisms of drug resistance. Mutations conferring resistance to nucleoside analogues
located within conserved regions of the pol domain might reduce the binding of the
inhibitor or the incorporation of the active drug into growing DNA [117]. It has
been suggested that mutations conferring resistance to nucleoside analogues located
in the exonuclease domain might enhance the rate of excision of the incorporated
drug [55]. However, a recent study demonstrated that mutant HCMV with reduced
exonuclease activity might efficiently synthetize DNA in the absence of drug exci-
sion [46]. It has been proposed that resistance of HSV-1 and HCMV to FOS may
250 W. L. Drew et al.
result from subtle conformational changes in the DNA pol that adopts a more open
conformation to which the drug binds with a lower affinity [178, 218, 238].
In VZV clinical isolates, resistance to ACV is mostly associated with mutations
in the viral TK (encoded by the ORF36 gene) and, less frequently, with mutations
in the viral DNA pol (encoded by the ORF28 gene) [177, 179]. The string of six
cytosines located at codons 493–498 of the TK gene emerged as a hot spot for the
insertion or deletion of nucleotides involved in ACV resistance (Fig. 8.4) [3, 27,
161, 222]. Deletions of nucleotides that result in frameshift reading leading to a stop
codon at position 231 are often detected in ACV-resistant VZV clinical isolates
[161]. In addition, nonsynonymous nucleotide substitutions conferring resistance to
ACV are widely dispersed in the TK gene. However, these amino acid changes
occur more frequently in the ATP-binding and nucleoside-binding sites of the TK
enzyme [27, 93, 161, 189, 192, 197, 216]. A few reports have described ACV-
resistant and/or FOS-resistant VZV clinical isolates with mutations in the DNA pol
gene (Fig. 8.5) [127, 192, 226]. The amino acid substitutions are mainly found in
the catalytic site and in the conserved regions of the DNA pol and may confer cross-
resistance to ACV and FOS [177, 179]. The TK and DNA pol genes of VZV are
highly conserved compared with those of HSVs, and only a few natural polymor-
phisms have been identified in these genes [192].
Fig. 8.6 Proposed algorithm for the management of suspected nucleoside analogue-resistant HSV
infections. Key: ACV acyclovir, VACV valacyclovir, FCV famciclovir, FOS foscarnet, CDV cido-
fovir, TID thrice a day, IV intravenous
considered. In parallel, isolates from the lesions should be submitted for susceptibil-
ity testing and/or genotypic analysis of the TK gene. If there is still no improvement
of HSV disease after 7–10 days, continuous infusion of high-dose ACV at a dosage
of 1.5–2.0 mg/kg/h for 6 weeks could be initiated, as it is a well-tolerated option for
severe ACV-resistant or multidrug-resistant HSV infections [83, 132]. Cidofovir
(5 mg/kg once a week for 3–4 weeks) could be considered, as it has shown some
efficacy in the treatment of progressive multidrug-resistant mucocutaneous HSV
infection in immunocompromised patients [150, 207], but it is not approved for this
indication. Although topical formulations of FOS [126] and CDV [138, 183] were
effective in the treatment of mucocutaneous infections not responding to ACV and
could avoid the adverse effects associated with their IV administration, they are not
commercially available. A topical formulation containing 5% imiquimod, an
252 W. L. Drew et al.
All currently available antiviral agents target the viral DNA pol. The development of
new anti-herpetic compounds with different mechanisms of action and with adequate
safety profile is urgently needed. Some promising compounds are currently in clinical
trials. The orally bioavailable lipid ester prodrug of CDV (hexadecyloxypropyl-cido-
fovir; brincidofovir) could avoid the dose-limiting toxicity of the parent drug and
8 Resistance of Herpesviruses to Antiviral Agents 253
Major Points
• Resistance of herpesviruses to antiviral drugs is mostly detected in immunocom-
promised patients but it is increasingly recognized in immunocompetent indi-
viduals with herpetic keratitis.
• Genotypic testing is more frequently used for the detection of antiviral drug
resistance in herpesvirus infections.
• Interpretation of genotypic testing requires a database linking amino acid changes
to mutations associated with natural polymorphisms or drug resistance.
• Algorithms are proposed for the management of infections caused by drug-resis-
tant herpesvirus strains.
• Novel antiviral agents acting on viral targets other than the viral DNA poly-
merase are in development for the treatment of herpesvirus infections.
254 W. L. Drew et al.
Acknowledgments This study was supported by a Foundation Grant from the Canadian Institutes
of Health Research (grant no. 148361 to G.B.). G.B. is the holder of the Canada research chair on
emerging viruses and antiviral resistance.
References
1. Ahmed AM, Brantley JS, Madkan V, Mendoza N, Tyring SK. Managing herpes zoster in
immunocompromised patients. Herpes. 2007;14:32–6.
2. Andrei G, Fiten P, Froeyen M, De Clercq E, Opdenakker G, Snoeck R. DNA polymerase
mutations in drug-resistant herpes simplex virus mutants determine in vivo neurovirulence
and drug-enzyme interactions. Antivir Ther. 2007;12:719–32.
3. Andrei G, Topalis D, Fiten P, McGuigan C, Balzarini J, Opdenakker G, et al. In vitro-selected
drug-resistant varicella-zoster virus mutants in the thymidine kinase and DNA polymerase
genes yield novel phenotype-genotype associations and highlight differences between anti-
herpesvirus drugs. J Virol. 2012;86:2641–52.
4. Andrei G, Georgala A, Topalis D, Fiten P, Aoun M, Opdenakker G, et al. Heterogeneity and
evolution of thymidine kinase and DNA polymerase mutants of herpes simplex virus type 1:
implications for antiviral therapy. J Infect Dis. 2013;207:1295–305.
5. Andrei G, Snoeck R. Advances in the treatment of varicella-zoster virus infections. Adv
Pharmacol. 2013;67:107–68.
6. Andrei G, Snoeck R. Herpes simplex virus drug-resistance: new mutations and insights. Curr
Opin Infect Dis. 2013;26:551–60.
7. Arthurs SK, Eid AJ, Pedersen RA, Kremers WK, Cosio FG, Patel R, et al. Delayed-onset
primary cytomegalovirus disease and the risk of allograft failure and mortality after kidney
transplantation. Clin Infect Dis. 2008;46:840.
8. Bacon TH, Boon RJ, Schultz M, Hodges-Savola C. Surveillance for antiviral-agent-resistant
herpes simplex virus in the general population with recurrent herpes labialis. Antimicrob
Agents Chemother. 2002;46:3042–4.
9. Bacon TH, Levin MJ, Leary JJ, Sarisky RT, Sutton D. Herpes simplex virus resistance
to acyclovir and penciclovir after two decades of antiviral therapy. Clin Microbiol Rev.
2003;16:114–28.
10. Baldanti F, Silini E, Sarasini A, Talarico CL, Stanat SC, Biron KK, et al. A three-nucleotide
deletion in the UL97 open reading frame is responsible for the ganciclovir resistance of a
human cytomegalovirus clinical isolate. J Virol. 1995;69:796–800.
11. Baldanti F, Underwood MR, Stanat SC, Biron KK, Chou S, Sarasini A, et al. Single amino
acid changes in the DNA polymerase confer foscarnet resistance and slow-growth phenotype,
while mutations in the UL97-encoded phosphotransferase confer ganciclovir resistance in
three double-resistant human cytomegalovirus strains recovered from patients with AIDS. J
Virol. 1996;70:1390–5.
12. Baldanti F, Underwood MR, Talarico CL, Simoncini L, Sarasini A, Biron KK, et al. The
Cys607-->Tyr change in the UL97 phosphotransferase confers ganciclovir resistance to
two human cytomegalovirus strains recovered from two immunocompromised patients.
Antimicrob Agents Chemother. 1998;42:444–6.
13. Baldanti F, Michel D, Simoncini L, Heuschmid M, Zimmermann A, Minisini R, et al.
Mutations in the UL97 ORF of ganciclovir-resistant clinical cytomegalovirus isolates differ-
entially affect GCV phosphorylation as determined in a recombinant vaccinia virus system.
Antivir Res. 2002;54:59–67.
14. Baldanti F, Lilleri D, Campanini G, Comolli G, Ridolfo AL, Rusconi S, et al. Human cyto-
megalovirus double resistance in a donor-positive/recipient-negative lung transplant patient
8 Resistance of Herpesviruses to Antiviral Agents 255
33. Boon RJ, Bacon TH, Robey HL, Coleman TJ, Connolly A, Crosson P, et al. Antiviral sus-
ceptibilities of herpes simplex virus from immunocompetent subjects with recurrent herpes
labialis: a UK-based survey. J Antimicrob Chemother. 2000;46:324–5.
34. Breton G, Fillet AM, Katlama C, Bricaire F, Caumes E. Acyclovir-resistant herpes zoster in
human immunodeficiency virus-infected patients: results of foscarnet therapy. Clin Infect
Dis. 1998;27:1525–7.
35. Brink AA, van Gelder M, Wolffs PF, Bruggeman CA, van Loo IH. Compartmentalization of
acyclovir-resistant varicella zoster virus: implications for sampling in molecular diagnostics.
Clin Infect Dis. 2011;52:982–7.
36. Brunnemann AK, Bohn-Wippert K, Zell R, Henke A, Walther M, Braum O, et al. Drug resis-
tance of clinical varicella-zoster virus strains confirmed by recombinant thymidine kinase
expression and by targeted resistance mutagenesis of a cloned wild-type isolate. Antimicrob
Agents Chemother. 2015;59:2726–34.
37. Brunnemann AK, Liermann K, Deinhardt-Emmer S, Maschkowitz G, Pohlmann A, Sodeik
B, et al. Recombinant herpes simplex virus type 1 strains with targeted mutations relevant for
aciclovir susceptibility. Sci Rep. 2016;6:29903.
38. Bryan CJ, Prichard MN, Daily S, Jefferson G, Hartline C, Cassady KA, et al. Acyclovir-
resistant chronic verrucous vaccine strain varicella in a patient with neuroblastoma. Pediatr
Infect Dis J. 2008;27:946–8.
39. Burrel S, Bonnafous P, Hubacek P, Agut H, Boutolleau D. Impact of novel mutations of her-
pes simplex virus 1 and 2 thymidine kinases on acyclovir phosphorylation activity. Antivir
Res. 2012;96:386–90.
40. Burrel S, Aime C, Hermet L, Ait-Arkoub Z, Agut H, Boutolleau D. Surveillance of herpes
simplex virus resistance to antivirals: a 4-year survey. Antivir Res. 2013;100:365–72.
41. Burrel S, Boutolleau D, Azar G, Doan S, Deback C, Cochereau I, et al. Phenotypic and geno-
typic characterization of acyclovir-resistant corneal HSV-1 isolates from immunocompetent
patients with recurrent herpetic keratitis. J Clin Virol. 2013;58:321–4.
42. Chakrabarti S, Pillay D, Ratcliffe D, Cane PA, Collingham KE, Milligan DW. Resistance
to antiviral drugs in herpes simplex virus infections among allogeneic stem cell transplant
recipients: risk factors and prognostic significance. J Infect Dis. 2000;181:2055–8.
43. Chatis PA, Miller CH, Schrager LE, Crumpacker CS. Successful treatment with foscarnet of
an acyclovir-resistant mucocutaneous infection with herpes simplex virus in a patient with
acquired immunodeficiency syndrome. N Engl J Med. 1989;320:297–300.
44. Chatis PA, Crumpacker CS. Analysis of the thymidine kinase gene from clinically isolated
acyclovir-resistant herpes simplex viruses. Virology. 1991;180:793–7.
45. Chemaly RF, Ullmann AJ, Stoelben S, Richard MP, Bornhauser M, Groth C, et al. Letermovir
for cytomegalovirus prophylaxis in hematopoietic-cell transplantation. N Engl J Med.
2014;370:1781–9.
46. Chen H, Beardsley GP, Coen DM. Mechanism of ganciclovir-induced chain termination
revealed by resistant viral polymerase mutants with reduced exonuclease activity. Proc Natl
Acad Sci U S A. 2014;111:17462–7.
47. Chen Y, Scieux C, Garrait V, Socie G, Rocha V, Molina JM, et al. Resistant herpes simplex
virus type 1 infection: an emerging concern after allogeneic stem cell transplantation. Clin
Infect Dis. 2000;31:927–35.
48. Chilukuri S, Rosen T. Management of acyclovir-resistant herpes simplex virus. Dermatol
Clin. 2003;21:311–20.
49. Chou S, Erice A, Jordan MC, Vercellotti GM, Michels KR, Talarico CL, et al. Analysis of the
UL97 phosphotransferase coding sequence in clinical cytomegalovirus isolates and identifi-
cation of mutations conferring ganciclovir resistance. J Infect Dis. 1995;171:576–83.
50. Chou S, Marousek G, Guentzel S, Follansbee SE, Poscher ME, Lalezari JP, et al. Evolution
of mutations conferring multidrug resistance during prophylaxis and therapy for cytomegalo-
virus disease. J Infect Dis. 1997;176:786–9.
8 Resistance of Herpesviruses to Antiviral Agents 257
51. Chou S, Marousek G, Parenti DM, Gordon SM, LaVoy AG, Ross JG, et al. Mutation in region
III of the DNA polymerase gene conferring foscarnet resistance in cytomegalovirus isolates
from 3 subjects receiving prolonged antiviral therapy. J Infect Dis. 1998;178:526–30.
52. Chou S, Lurain NS, Weinberg A, Cai GY, Sharma PL, Crumpacker CS. Interstrain varia-
tion in the human cytomegalovirus DNA polymerase sequence and its effect on genotypic
diagnosis of antiviral drug resistance. Adult AIDS Clinical Trials Group CMV Laboratories.
Antimicrob Agents Chemother. 1999;43:1500–2.
53. Chou S, Meichsner CL. A nine-codon deletion mutation in the cytomegalovirus UL97
phosphotransferase gene confers resistance to ganciclovir. Antimicrob Agents Chemother.
2000;44:183–5.
54. Chou S, Waldemer RH, Senters AE, Michels KS, Kemble GW, Miner RC, et al.
Cytomegalovirus UL97 phosphotransferase mutations that affect susceptibility to ganciclo-
vir. J Infect Dis. 2002;185:162–9.
55. Chou S, Lurain NS, Thompson KD, Miner RC, Drew WL. Viral DNA polymerase mutations
associated with drug resistance in human cytomegalovirus. J Infect Dis. 2003;188:32–9.
56. Chou S, Van Wechel LC, Lichy HM, Marousek GI. Phenotyping of cytomegalovirus drug
resistance mutations by using recombinant viruses incorporating a reporter gene. Antimicrob
Agents Chemother. 2005;49:2710–5.
57. Chou S, Hakki M, Villano S. Effects on maribavir susceptibility of cytomegalovirus UL97
kinase ATP binding region mutations detected after drug exposure in vitro and in vivo. Antivir
Res. 2012;95:88–92.
58. Chou S, Ercolani RJ, Marousek G, Bowlin TL. Cytomegalovirus UL97 kinase cata-
lytic domain mutations that confer multidrug resistance. Antimicrob Agents Chemother.
2013;57:3375–9.
59. Chrisp P, Clissold SP. Foscarnet. A review of its antiviral activity, pharmacokinetic properties
and therapeutic use in immunocompromised patients with cytomegalovirus retinitis. Drugs.
1991;41:104–29.
60. Christophers J, Clayton J, Craske J, Ward R, Collins P, Trowbridge M, et al. Survey of
resistance of herpes simplex virus to acyclovir in Northwest England. Antimicrob Agents
Chemother. 1998;42:868–72.
61. Cihlar T, Chen MS. Identification of enzymes catalyzing two-step phosphorylation of cidofo-
vir and the effect of cytomegalovirus infection on their activities in host cells. Mol Pharmacol.
1996;50:1502–10.
62. Cihlar T, Fuller MD, Cherrington JM. Characterization of drug resistance-associated muta-
tions in the human cytomegalovirus DNA polymerase gene by using recombinant mutant
viruses generated from overlapping DNA fragments. J Virol. 1998;72:5927–36.
63. Cihlar T, Fuller MD, Mulato AS, Cherrington JM. A point mutation in the human cytomega-
lovirus DNA polymerase gene selected in vitro by cidofovir confers a slow replication phe-
notype in cell culture. Virology. 1998;248:382–93.
64. Collins P, Larder BA, Oliver NM, Kemp S, Smith IW, Darby G. Characterization of a DNA
polymerase mutant of herpes simplex virus from a severely immunocompromised patient
receiving acyclovir. J Gen Virol. 1989;70:375–82.
65. Collins P, Ellis MN. Sensitivity monitoring of clinical isolates of herpes simplex virus to
acyclovir. J Med Virol. 1993;(Suppl 1):58–66.
66. Crassard N, Souillet AL, Morfin F, Thouvenot D, Claudy A, Bertrand Y. Acyclovir-resistant
varicella infection with atypical lesions in a non-HIV leukemic infant. Acta Paediatr.
2000;89:1497–9.
67. Dambrosi S, Martin M, Yim K, Miles B, Canas J, Sergerie Y, et al. Neurovirulence and
latency of drug-resistant clinical herpes simplex viruses in animal models. J Med Virol.
2010;82:1000.
68. Dankner WM, Scholl D, Stanat SC, Martin M, Sonke RL, Spector SA. Rapid antiviral DNA-
DNA hybridization assay for human cytomegalovirus. J Virol Methods. 1990;28:293–8.
258 W. L. Drew et al.
the acquired immunodeficiency syndrome (AIDS). An uncontrolled trial. Ann Intern Med.
1989;110:710–3.
89. Erlich KS, Mills J, Chatis P, Mertz GJ, Busch DF, Follansbee SE, et al. Acyclovir-resistant
herpes simplex virus infections in patients with the acquired immunodeficiency syndrome. N
Engl J Med. 1989;320:293–6.
90. Field HJ, Vere Hodge RA. Recent developments in anti-herpesvirus drugs. Br Med Bull.
2013;106:213–49.
91. Fife KH, Crumpacker CS, Mertz GJ, Hill EL, Boone GS. Recurrence and resistance patterns
of herpes simplex virus following cessation of > or = 6 years of chronic suppression with
acyclovir. Acyclovir study group. J Infect Dis. 1994;169:1338–41.
92. Fillet AM, Visse B, Caumes E, Dumont B, Gentilini M, Huraux JM. Foscarnet-resistant mul-
tidermatomal zoster in a patient with AIDS. Clin Infect Dis. 1995;21:1348–9.
93. Fillet AM, Dumont B, Caumes E, Visse B, Agut H, Bricaire F, et al. Acyclovir-resistant
varicella-zoster virus: phenotypic and genetic characterization. J Med Virol. 1998;55:250–4.
94. Fillet AM, Auray L, Alain S, Gourlain K, Imbert BM, Najioullah F, et al. Natural polymor-
phism of cytomegalovirus DNA polymerase lies in two nonconserved regions located between
domains delta-C and II and between domains III and I. Antimicrob Agents Chemother.
2004;48:1865–8.
95. Florescu DF, Qiu F, Schmidt CM, Kalil AC. A direct and indirect comparison meta-analysis
on the efficacy of cytomegalovirus preventive strategies in solid organ transplant. Clin Infect
Dis. 2014;58:785–803.
96. Frangoul H, Wills M, Crossno C, Engel M, Domm J. Acyclovir-resistant herpes simplex virus
pneumonia post-unrelated stem cell transplantation: a word of caution. Pediatr Transplant.
2007;11:942–4.
97. Frobert E, Burrel S, Ducastelle-Lepretre S, Billaud G, Ader F, Casalegno JS, et al. Resistance
of herpes simplex viruses to acyclovir: an update from a ten-year survey in France. Antivir
Res. 2014;111:36–41.
98. Fyfe JA, Keller PM, Furman PA, Miller RL, Elion GB. Thymidine kinase from herpes sim-
plex virus phosphorylates the new antiviral compound, 9-(2-hydroxyethoxymethyl)guanine.
J Biol Chem. 1978;253:8721–7.
99. Gaudreau A, Hill E, Balfour HH Jr, Erice A, Boivin G. Phenotypic and genotypic charac-
terization of acyclovir-resistant herpes simplex viruses from immunocompromised patients.
J Infect Dis. 1998;178:297–303.
100. Gerna G, Baldanti F, Zavattoni M, Sarasini A, Percivalle E, Revello MG. Monitoring of gan-
ciclovir sensitivity of multiple human cytomegalovirus strains coinfecting blood of an AIDS
patient by an immediate-early antigen plaque assay. Antivir Res. 1992;19:333–45.
101. Gerna G, Sarasini A, Lilleri D, Percivalle E, Torsellini M, Baldanti F, et al. In vitro model
for the study of the dissociation of increasing antigenemia and decreasing DNAemia and
viremia during treatment of human cytomegalovirus infection with ganciclovir in transplant
recipients. J Infect Dis. 2003;188:1639–47.
102. Gilbert C, Roy J, Belanger R, Delage R, Beliveau C, Demers C, et al. Lack of emergence
of cytomegalovirus UL97 mutations conferring ganciclovir (GCV) resistance following
preemptive GCV therapy in allogeneic stem cell transplant recipients. Antimicrob Agents
Chemother. 2001;45:3669–71.
103. Gilbert C, Bestman-Smith J, Boivin G. Resistance of herpesviruses to antiviral drugs: clinical
impacts and molecular mechanisms. Drug Resist Updat. 2002;5:88–114.
104. Gilbert C, Boivin G. Discordant phenotypes and genotypes of cytomegalovirus (CMV) in
patients with AIDS and relapsing CMV retinitis. AIDS. 2003;17:337–41.
105. Gilbert C, Boivin G. New reporter cell line to evaluate the sequential emergence of mul-
tiple human cytomegalovirus mutations during in vitro drug exposure. Antimicrob Agents
Chemother. 2005;49:4860–6.
106. Gilbert C, Boivin G. Human cytomegalovirus resistance to antiviral drugs. Antimicrob
Agents Chemother. 2005;49:873–83.
260 W. L. Drew et al.
125. James SH, Prichard MN. A possible pitfall in acyclovir prophylaxis for recurrent herpetic
keratitis? J Infect Dis. 2013;208:1353–5.
126. Javaly K, Wohlfeiler M, Kalayjian R, Klein T, Bryson Y, Grafford K, et al. Treatment of
mucocutaneous herpes simplex virus infections unresponsive to acyclovir with topical foscar-
net cream in AIDS patients: a phase I/II study. J Acquir Immune Defic Syndr. 1999;21:301–6.
127. Kamiyama T, Kurokawa M, Shiraki K. Characterization of the DNA polymerase gene of
varicella-zoster viruses resistant to acyclovir. J Gen Virol. 2001;82:2761–5.
128. Karamitros T, Harrison I, Piorkowska R, Katzourakis A, Magiorkinis G, Mbisa JL. De novo
assembly of human herpes virus type 1 (HHV-1) genome, mining of non-canonical structures
and detection of novel drug-resistance mutations using short- and long-read next generation
sequencing technologies. PLoS One. 2016;11:e0157600.
129. Kaul DR, Stoelben S, Cober E, Ojo T, Sandusky E, Lischka P, et al. First report of success-
ful treatment of multidrug-resistant cytomegalovirus disease with the novel anti-CMV com-
pound AIC246. Am J Transplant. 2011;11:1079–84.
130. Kessler HA, Hurwitz S, Farthing C, Benson CA, Feinberg J, Kuritzkes DR, et al. Pilot study
of topical trifluridine for the treatment of acyclovir-resistant mucocutaneous herpes simplex
disease in patients with AIDS (ACTG 172). AIDS Clinical Trials Group. J Acquir Immune
Defic Syndr Hum Retrovirol. 1996;12:147–52.
131. Kesson AM, Zeng F, Cunningham AL, Rawlinson WD. The use of flow cytometry to detect
antiviral resistance in human cytomegalovirus. J Virol Methods. 1998;71:177–86.
132. Kim JH, Schaenman JM, Ho DY, Brown JM. Treatment of acyclovir-resistant herpes sim-
plex virus with continuous infusion of high-dose acyclovir in hematopoietic cell transplant
patients. Biol Blood Marrow Transplant. 2011;17:259–64.
133. Kost RG, Hill EL, Tigges M, Straus SE. Brief report: recurrent acyclovir-resistant genital
herpes in an immunocompetent patient. N Engl J Med. 1993;329:1777–82.
134. Kotton CN, Kumar D, Caliendo AM, Asberg A, Chou S, Danziger-Isakov L, et al. Updated
international consensus guidelines on the management of cytomegalovirus in solid-organ
transplantation. Transplantation. 2013;96:333–60.
135. Krawczyk A, Arndt MA, Grosse-Hovest L, Weichert W, Giebel B, Dittmer U, et al.
Overcoming drug-resistant herpes simplex virus (HSV) infection by a humanized antibody.
Proc Natl Acad Sci U S A. 2013;110:6760–5.
136. Kriesel JD, Spruance SL, Prichard M, Parker JN, Kern ER. Recurrent antiviral-resistant geni-
tal herpes in an immunocompetent patient. J Infect Dis. 2005;192:156–61.
137. Kruger RM, Shannon WD, Arens MQ, Lynch JP, Storch GA, Trulock EP. The impact of
ganciclovir-resistant cytomegalovirus infection after lung transplantation. Transplantation.
1999;68:1272–9.
138. Lalezari J, Schacker T, Feinberg J, Gathe J, Lee S, Cheung T, et al. A randomized, double-
blind, placebo-controlled trial of cidofovir gel for the treatment of acyclovir-unresponsive
mucocutaneous herpes simplex virus infection in patients with AIDS. J Infect Dis.
1997;176:892–8.
139. Landry ML, Stanat S, Biron K, Brambilla D, Britt W, Jokela J, et al. A standardized plaque
reduction assay for determination of drug susceptibilities of cytomegalovirus clinical iso-
lates. Antimicrob Agents Chemother. 2000;44:688–92.
140. Langston AA, Redei I, Caliendo AM, Somani J, Hutcherson D, Lonial S, et al. Development
of drug-resistant herpes simplex virus infection after haploidentical hematopoietic progenitor
cell transplantation. Blood. 2002;99:1085–8.
141. Lascaux AS, Caumes E, Deback C, Melica G, Challine D, Agut H, et al. Successful treatment
of aciclovir and foscarnet resistant Herpes simplex virus lesions with topical imiquimod in
patients infected with human immunodeficiency virus type 1. J Med Virol. 2012;84:194–7.
142. Levin MJ, Dahl KM, Weinberg A, Giller R, Patel A, Krause PR. Development of resistance
to acyclovir during chronic infection with the Oka vaccine strain of varicella-zoster virus, in
an immunosuppressed child. J Infect Dis. 2003;188:954–9.
262 W. L. Drew et al.
143. Levin MJ, Bacon TH, Leary JJ. Resistance of herpes simplex virus infections to nucleoside
analogues in HIV-infected patients. Clin Infect Dis. 2004;39:S248–57.
144. Limaye AP, Corey L, Koelle DM, Davis CL, Boeckh M. Emergence of ganciclovir-
resistant cytomegalovirus disease among recipients of solid-organ transplants. Lancet.
2000;356:645–9.
145. Limaye AP. Ganciclovir-resistant cytomegalovirus in organ transplant recipients. Clin Infect
Dis. 2002;35:866–72.
146. Limaye AP, Bakthavatsalam R, Kim HW, Randolph SE, Halldorson JB, Healey PJ, et al.
Impact of cytomegalovirus in organ transplant recipients in the era of antiviral prophylaxis.
Transplantation. 2006;81:1645–52.
147. Lischka P, Hewlett G, Wunberg T, Baumeister J, Paulsen D, Goldner T, et al. In vitro and
in vivo activities of the novel anticytomegalovirus compound AIC246. Antimicrob Agents
Chemother. 2010;54:1290–7.
148. Liu W, Kuppermann BD, Martin DF, Wolitz RA, Margolis TP. Mutations in the cytomegalovi-
rus UL97 gene associated with ganciclovir-resistant retinitis. J Infect Dis. 1998;177:1176–81.
149. Lolis MS, Gonzalez L, Cohen PJ, Schwartz RA. Drug-resistant herpes simplex virus in HIV
infected patients. Acta Dermatovenerol Croat. 2008;16:204–8.
150. LoPresti AE, Levine JF, Munk GB, Tai CY, Mendel DB. Successful treatment of an acyclovir-
and foscarnet-resistant herpes simplex virus type 1 lesion with intravenous cidofovir. Clin
Infect Dis. 1998;26:512–3.
151. Lowance D, Neumayer HH, Legendre CM, Squifflet JP, Kovarik J, Brennan PJ, et al.
Valacyclovir for the prevention of cytomegalovirus disease after renal transplantation.
International Valacyclovir Cytomegalovirus Prophylaxis Transplantation Study Group. N
Engl J Med. 1999;340:1462–70.
152. Lurain NS, Weinberg A, Crumpacker CS, Chou S, Adult AIDS Clinical Trials Group-CMV
Laboratories. Sequencing of cytomegalovirus UL97 gene for genotypic antiviral resistance
testing. Antimicrob Agents Chemother. 2001;45:2775–80.
153. Lurain NS, Bhorade SM, Pursell KJ, Avery RK, Yeldandi VV, Isada CM, et al. Analysis and
characterization of antiviral drug-resistant cytomegalovirus isolates from solid organ trans-
plant recipients. J Infect Dis. 2002;186:760–8.
154. Lurain NS, Chou S. Antiviral drug resistance of human cytomegalovirus. Clin Microbiol Rev.
2010;23:689–712.
155. Malartre N, Boulieu R, Falah N, Cortay JC, Lina B, Morfin F, et al. Effects of mutations
on herpes simplex virus 1 thymidine kinase functionality: an in vitro assay based on detec-
tion of monophosphate forms of acyclovir and thymidine using HPLC/DAD. Antivir Res.
2012;95:224–8.
156. Martin JL, Ellis MN, Keller PM, Biron KK, Lehrman SN, Barry DW, et al. Plaque autora-
diography assay for the detection and quantitation of thymidine kinase-deficient and thymi-
dine kinase-altered mutants of herpes simplex virus in clinical isolates. Antimicrob Agents
Chemother. 1985;28:181–7.
157. Marty FM, Winston DJ, Rowley SD, Vance E, Papanicolaou GA, Mullane KM, et al.
CMX001 to prevent cytomegalovirus disease in hematopoietic-cell transplantation. N Engl
J Med. 2013;369:1227–36.
158. McSharry JM, Lurain NS, Drusano GL, Landay A, Manischewitz J, Nokta M, et al. Flow
cytometric determination of ganciclovir susceptibilities of human cytomegalovirus clinical
isolates. J Clin Microbiol. 1998;36:958–64.
159. Mertz GJ, Jones CC, Mills J, Fife KH, Lemon SM, Stapleton JT, et al. Long-term acyclo-
vir suppression of frequently recurring genital herpes simplex virus infection. A multicenter
double-blind trial. JAMA. 1988;260:201–6.
160. Michel D, Hohn S, Haller T, Jun D, Mertens T. Aciclovir selects for ganciclovir-cross-
resistance of human cytomegalovirus in vitro that is only in part explained by known muta-
tions in the UL97 protein. J Med Virol. 2001;65:70–6.
8 Resistance of Herpesviruses to Antiviral Agents 263
180. Piret J, Goyette N, Boivin G. Novel method based on real-time cell analysis for drug sus-
ceptibility testing of herpes simplex virus and human cytomegalovirus. J Clin Microbiol.
2016;54:2120–7.
181. Reyes M, Shaik NS, Graber JM, Nisenbaum R, Wetherall NT, Fukuda K, et al. Acyclovir-
resistant genital herpes among persons attending sexually transmitted disease and human
immunodeficiency virus clinics. Arch Intern Med. 2003;163:76–80.
182. Rifkin LM, Minkus CL, Pursell K, Jumroendararasame C, Goldstein DA. Utility of lefluno-
mide in the treatment of drug resistant cytomegalovirus retinitis. Ocul Immunol Inflamm.
2017;25:93–6.
183. Sacks SL, Shafran SD, Diaz-Mitoma F, Trottier S, Sibbald RG, Hughes A, et al. A multi-
center phase I/II dose escalation study of single-dose cidofovir gel for treatment of recurrent
genital herpes. Antimicrob Agents Chemother. 1998;42:2996–9.
184. Safrin S, Assaykeen T, Follansbee S, Mills J. Foscarnet therapy for acyclovir-resistant muco-
cutaneous herpes simplex virus infection in 26 AIDS patients: preliminary data. J Infect Dis.
1990;161:1078–84.
185. Safrin S, Ashley R, Houlihan C, Cusick PS, Mills J. Clinical and serologic features of herpes
simplex virus infection in patients with AIDS. AIDS. 1991;5:1107–10.
186. Safrin S, Berger TG, Gilson I, Wolfe PR, Wofsy CB, Mills J, et al. Foscarnet therapy in five
patients with AIDS and acyclovir-resistant varicella-zoster virus infection. Ann Intern Med.
1991;115:19–21.
187. Safrin S, Crumpacker C, Chatis P, Davis R, Hafner R, Rush J, et al. A controlled trial com-
paring foscarnet with vidarabine for acyclovir-resistant mucocutaneous herpes simplex in
the acquired immunodeficiency syndrome. The AIDS Clinical Trials Group. N Engl J Med.
1991;325:551–5.
188. Safrin S, Kemmerly S, Plotkin B, Smith T, Weissbach N, De Veranez D, et al. Foscarnet-
resistant herpes simplex virus infection in patients with AIDS. J Infect Dis. 1994;169:193–6.
189. Saint-Leger E, Caumes E, Breton G, Douard D, Saiag P, Huraux JM, et al. Clinical and viro-
logic characterization of acyclovir-resistant varicella-zoster viruses isolated from 11 patients
with acquired immunodeficiency syndrome. Clin Infect Dis. 2001;33:2061–7.
190. Sarisky RT, Crosson P, Cano R, Quail MR, Nguyen TT, Wittrock RJ, et al. Comparison of
methods for identifying resistant herpes simplex virus and measuring antiviral susceptibility.
J Clin Virol. 2002;23:191–200.
191. Sasadeusz JJ, Sacks SL. Spontaneous reactivation of thymidine kinase-deficient, acyclovir-
resistant type-2 herpes simplex virus: masked heterogeneity or reversion? J Infect Dis.
1996;174:476–82.
192. Sauerbrei A, Taut J, Zell R, Wutzler P. Resistance testing of clinical varicella-zoster virus
strains. Antivir Res. 2011;90:242–7.
193. Sauerbrei A, Liermann K, Bohn K, Henke A, Zell R, Gronowitz S, et al. Significance of
amino acid substitutions in the thymidine kinase gene of herpes simplex virus type 1 for
resistance. Antivir Res. 2012;96:105–7.
194. Sauerbrei A, Vodisch S, Bohn K, Schacke M, Gronowitz S. Screening of herpes simplex
virus type 1 isolates for acyclovir resistance using DiviTumR assay. J Virol Methods.
2013;188:70–2.
195. Sauerbrei A. Diagnosis, antiviral therapy, and prophylaxis of varicella-zoster virus infections.
Eur J Clin Microbiol Infect Dis. 2016;35:723–34.
196. Sauerbrei A, Bohn-Wippert K, Kaspar M, Krumbholz A, Karrasch M, Zell R. Database
on natural polymorphisms and resistance-related non-synonymous mutations in thymidine
kinase and DNA polymerase genes of herpes simplex virus types 1 and 2. J Antimicrob
Chemother. 2016;71:6–16.
197. Sawyer MH, Inchauspe G, Biron KK, Waters DJ, Straus SE, Ostrove JM. Molecular analy-
sis of the pyrimidine deoxyribonucleoside kinase gene of wild-type and acyclovir-resistant
strains of varicella-zoster virus. J Gen Virol. 1988;69:2585–93.
8 Resistance of Herpesviruses to Antiviral Agents 265
198. Schliefer K, Gumbel HO, Rockstroh JK, Spengler U. Management of progressive outer reti-
nal necrosis with cidofovir in a human immunodeficiency virus-infected patient. Clin Infect
Dis. 1999;29:684–5.
199. Schmit I, Boivin G. Characterization of the DNA polymerase and thymidine kinase genes of
herpes simplex virus isolates from AIDS patients in whom acyclovir and foscarnet therapy
sequentially failed. J Infect Dis. 1999;180:487–90.
200. Schnepf N, Boiteau N, Petit F, Alain S, Sanson-Le Pors MJ, Mazeron MC. Rapid determina-
tion of antiviral drug susceptibility of human cytomegalovirus by real-time PCR. Antivir Res.
2009;81:64–7.
201. Schubert A, Ehlert K, Schuler-Luettmann S, Gentner E, Mertens T, Michel D. Fast selection
of maribavir resistant cytomegalovirus in a bone marrow transplant recipient. BMC Infect
Dis. 2013;13:330.
202. Sergerie Y, Boivin G. Thymidine kinase mutations conferring acyclovir resistance in herpes
simplex type 1 recombinant viruses. Antimicrob Agents Chemother. 2006;50:3889–92.
203. Singh N. Antiviral drugs for cytomegalovirus in transplant recipients: advantages of preemp-
tive therapy. Rev Med Virol. 2006;16:281–7.
204. Smith IL, Shinkai M, Freeman WR, Spector SA. Polyradiculopathy associated with
ganciclovir-resistant cytomegalovirus in an AIDS patient: phenotypic and genotypic charac-
terization of sequential virus isolates. J Infect Dis. 1996;173:1481–4.
205. Smith IL, Cherrington JM, Jiles RE, Fuller MD, Freeman WR, Spector SA. High-level resis-
tance of cytomegalovirus to ganciclovir is associated with alterations in both the UL97 and
DNA polymerase genes. J Infect Dis. 1997;176:69–77.
206. Smith IL, Taskintuna I, Rahhal FM, Powell HC, Ai E, Mueller AJ, et al. Clinical failure
of CMV retinitis with intravitreal cidofovir is associated with antiviral resistance. Arch
Ophthalmol. 1998;116:178–85.
207. Snoeck R, Andrei G, Gerard M, Silverman A, Hedderman A, Balzarini J, et al. Successful treat-
ment of progressive mucocutaneous infection due to acyclovir- and foscarnet-resistant herpes
simplex virus with (S)-1-(3-hydroxy-2-phosphonylmethoxypropyl)cytosine (HPMPC). Clin
Infect Dis. 1994;18:570–8.
208. SOCA. Combination foscarnet and ganciclovir therapy vs monotherapy for the treatment of
relapsed cytomegalovirus retinitis in patients with AIDS. The Cytomegalovirus Retreatment
Trial. The studies of ocular complications of AIDS research group in collaboration with the
AIDS clinical trials group. Arch Ophthalmol. 1996;114:23–33.
209. Springer KL, Chou S, Li S, Giller RH, Quinones R, Shira JE, et al. How evolution of mutations
conferring drug resistance affects viral dynamics and clinical outcomes of cytomegalovirus-
infected hematopoietic cell transplant recipients. J Clin Microbiol. 2005;43:208–13.
210. Stoelben S, Arns W, Renders L, Hummel J, Muhlfeld A, Stangl M, et al. Preemptive treat-
ment of cytomegalovirus infection in kidney transplant recipients with letermovir: results of
a phase 2a study. Transpl Int. 2014;27:77–86.
211. Stranska R, Schuurman R, Nienhuis E, Goedegebuure IW, Polman M, Weel JF, et al. Survey
of acyclovir-resistant herpes simplex virus in the Netherlands: prevalence and characteriza-
tion. J Clin Virol. 2005;32:7–18.
212. Strasfeld L, Lee I, Tatarowicz W, Villano S, Chou S. Virologic characterization of multidrug-
resistant cytomegalovirus infection in 2 transplant recipients treated with maribavir. J Infect
Dis. 2010;202:104–8.
213. Sullivan V, Talarico CL, Stanat SC, Davis M, Coen DM, Biron KK. A protein kinase homo-
logue controls phosphorylation of ganciclovir in human cytomegalovirus-infected cells.
Nature. 1992;358:162–4.
214. Swetter SM, Hill EL, Kern ER, Koelle DM, Posavad CM, Lawrence W, et al. Chronic vul-
var ulceration in an immunocompetent woman due to acyclovir-resistant, thymidine kinase-
deficient herpes simplex virus. J Infect Dis. 1998;177:543–50.
266 W. L. Drew et al.
215. Swierkosz EM, Hodinka RL, Moore BM, Sacks S, Scholl DR, Wright DK. Antiviral suscep-
tibility testing: Herpes simplex virus by plaque reduction assay; approved standard. Wayne,
PA: Clinical and Laboratory Standards Institute; 2004. p. 1–39.
216. Talarico CL, Phelps WC, Biron KK. Analysis of the thymidine kinase genes from acyclovir-
resistant mutants of varicella-zoster virus isolated from patients with AIDS. J Virol.
1993;67:1024–33.
217. Tatarowicz WA, Lurain NS, Thompson KD. In situ ELISA for the evaluation of antiviral
compounds effective against human cytomegalovirus. J Virol Methods. 1991;35:207–15.
218. Tchesnokov EP, Gilbert C, Boivin G, Gotte M. Role of helix P of the human cytomegalovirus
DNA polymerase in resistance and hypersusceptibility to the antiviral drug foscarnet. J Virol.
2006;80:1440–50.
219. Telenti A, Smith TF. Screening with a shell vial assay for antiviral activity against cytomega-
lovirus. Diagn Microbiol Infect Dis. 1989;12:5–8.
220. Topalis D, Gillemot S, Snoeck R, Andrei G. Distribution and effects of amino acid changes
in drug-resistant alpha and beta herpesviruses DNA polymerase. Nucleic Acids Res.
2016;44:9530–54.
221. Tyring SK, Plunkett S, Scribner AR, Broker RE, Herrod JN, Handke LT, et al. Valomaciclovir
versus valacyclovir for the treatment of acute herpes zoster in immunocompetent adults: a
randomized, double-blind, active-controlled trial. J Med Virol. 2012;84:1224–32.
222. van der Beek MT, Vermont CL, Bredius RG, Marijt EW, van der Blij-de Brouwer CS, Kroes
AC, et al. Persistence and antiviral resistance of VZV in hematological patients. Clin Infect
Dis. 2013;56:335–43.
223. van Velzen M, van Loenen FB, Meesters RJ, de Graaf M, Remeijer L, Luider TM, et al.
Latent acyclovir-resistant herpes simplex virus type 1 in trigeminal ganglia of immunocom-
petent individuals. J Infect Dis. 2012;205:1539–43.
224. van Velzen M, Missotten T, van Loenen FB, Meesters RJ, Luider TM, Baarsma GS, et al.
Acyclovir-resistant herpes simplex virus type 1 in intra-ocular fluid samples of herpetic uve-
itis patients. J Clin Virol. 2013;57:215–21.
225. Vere Hodge RA, Field HJ. Antiviral agents for herpes simplex virus. Adv Pharmacol.
2013;67:1–38.
226. Visse B, Dumont B, Huraux JM, Fillet AM. Single amino acid change in DNA polymerase is
associated with foscarnet resistance in a varicella-zoster virus strain recovered from a patient
with AIDS. J Infect Dis. 1998;178:S55–7.
227. Visse B, Huraux JM, Fillet AM. Point mutations in the varicella-zoster virus DNA poly-
merase gene confers resistance to foscarnet and slow growth phenotype. J Med Virol.
1999;59:84–90.
228. Wade JC, Newton B, McLaren C, Flournoy N, Keeney RE, Meyers JD. Intravenous acyclo-
vir to treat mucocutaneous herpes simplex virus infection after marrow transplantation: a
double-blind trial. Ann Intern Med. 1982;96:265–9.
229. Wade JC, McLaren C, Meyers JD. Frequency and significance of acyclovir-resistant herpes
simplex virus isolated from marrow transplant patients receiving multiple courses of treat-
ment with acyclovir. J Infect Dis. 1983;148:1077–82.
230. Wald A, Corey L, Timmler B, Magaret A, Warren T, Tyring S, et al. Helicase-primase inhibi-
tor pritelivir for HSV-2 infection. N Engl J Med. 2014;370:201–10.
231. Weinberg A, Jabs DA, Chou S, Martin BK, Lurain NS, Forman MS, et al. Mutations confer-
ring foscarnet resistance in a cohort of patients with acquired immunodeficiency syndrome
and cytomegalovirus retinitis. J Infect Dis. 2003;187:777–84.
232. Whitley RJ, Gnann JW Jr. Acyclovir: a decade later. N Engl J Med. 1992;327:782–9.
233. Williamson EC, Millar MR, Steward CG, Cornish JM, Foot AB, Oakhill A, et al. Infections
in adults undergoing unrelated donor bone marrow transplantation. Br J Haematol.
1999;104:560–8.
8 Resistance of Herpesviruses to Antiviral Agents 267
234. Wolf DG, Lee DJ, Spector SA. Detection of human cytomegalovirus mutations associated
with ganciclovir resistance in cerebrospinal fluid of AIDS patients with central nervous sys-
tem disease. Antimicrob Agents Chemother. 1995;39:2552–4.
235. Wolf DG, Smith IL, Lee DJ, Freeman WR, Flores-Aguilar M, Spector SA. Mutations in
human cytomegalovirus UL97 gene confer clinical resistance to ganciclovir and can be
detected directly in patient plasma. J Clin Invest. 1995;95:257–63.
236. Wolf DG, Yaniv I, Honigman A, Kassis I, Schonfeld T, Ashkenazi S. Early emergence of
ganciclovir-resistant human cytomegalovirus strains in children with primary combined
immunodeficiency. J Infect Dis. 1998;178:535–8.
237. Wolf DG, Lurain NS, Zuckerman T, Hoffman R, Satinger J, Honigman A, et al. Emergence
of late cytomegalovirus central nervous system disease in hematopoietic stem cell transplant
recipients. Blood. 2003;101:463–5.
238. Zahn KE, Tchesnokov EP, Gotte M, Doublie S. Phosphonoformic acid inhibits viral replica-
tion by trapping the closed form of the DNA polymerase. J Biol Chem. 2011;286:25246–55.
239. Ziyaeyan M, Alborzi A, Japoni A, Kadivar M, Davarpanah MA, Pourabbas B, et al. Frequency
of acyclovir-resistant herpes simplex viruses isolated from the general immunocompe-
tent population and patients with acquired immunodeficiency syndrome. Int J Dermatol.
2007;46:1263–6.
Chapter 9
Heteroresistance: A Harbinger of Future
Resistance
9.1 Introduction
K. Drlica (*)
Public Health Research Institute, New Jersey Medical School,
Rutgers Biomedical and Health Sciences, Newark, NJ, USA
e-mail: [email protected]
B. Shopsin
Departments of Medicine and Microbiology, New York University School of Medicine,
New York, NY, USA
X. Zhao
Public Health Research Institute, New Jersey Medical School, Rutgers Biomedical
and Health Sciences, Newark, NJ, USA
Department of Microbiology, Biochemistry, & Molecular Genetics, New Jersey Medical
School, Rutgers Biomedical and Health Sciences, Newark, NJ, USA
State Key Laboratory of Molecular Vaccinology and Molecular Diagnostics, School of Public
Health, Xiamen University, Xiamen, Fujian Province, China
Heteroresistance manifests itself in several ways. The most graphic is the growth of
bacterial colonies within a zone of inhibition created when an antimicrobial diffuses
from a central source on agar that had been covered with bacteria prior to incubation
(for example, see Fig. 1 in Ref. [7]). If the colonies in the inhibition zone test posi-
tive for antimicrobial resistance using assays that measure minimal inhibitory con-
centration (MIC), the overall population is said to be heteroresistant. When those
resistant colonies continue to test resistant following multiple rounds of growth in
or on drug-free medium, the heteroresistance is said to be stable. Many examples
exist in which the resistance phenotype is lost during subculturing in the absence of
drug. Those situations are called unstable heteroresistance. The “colonies within the
inhibition zone” is the easiest method for detecting heteroresistance and is
9 Heteroresistance: A Harbinger of Future Resistance 271
Resistant
Colonies recovered
Heteroresistant
Wild type
Drug concentration
Fig. 9.1 Population analysis profile. A bacterial culture or specimen is applied to a series of agar
plates containing different concentrations of the test antimicrobial. After incubation to allow col-
ony formation, colonies are counted and plotted for each drug concentration. A resistant culture is
unaffected by the drug until concentrations are very high, while a fully susceptible culture (wild
type) exhibits a sharp drop in colony recovery at MIC. Results for a heteroresistant culture contain-
ing a variety of subpopulations having reduced susceptibility are depicted. (Data for hVISA can be
seen in Refs [5, 6])
commonly used as an initial screen with samples that would otherwise be scored as
susceptible by diagnostic laboratories.
With some pathogens, discordance in susceptibility testing indicates heteroresis-
tance. For example, with M. tuberculosis, DNA tests may indicate the presence of
mutations associated with drug resistance, while drug susceptibility testing (deter-
mination of MIC) indicates that the isolate is in the drug-susceptible category.
Discordance can also occur between liquid-growth and agar-plate tests. In both situ-
ations the discordance arises from assay sensitivity differences.
The gold standard for establishing heteroresistance is detection of “resistant”
subpopulations in a population analysis profile (PAP, Ref. [8]). For this assay, a
series of agar plates is prepared such that each plate contains a different concentra-
tion of drug. A large number of cells, generally >106, are applied to each agar plate,
and after incubation at the appropriate growth temperature (usually 37 °C), the
number of colonies is scored. A fully susceptible pathogen isolate will exhibit a
sharp drop in colony number when the drug concentration in the agar reaches
MIC. In contrast, a heteroresistant isolate will show colonies at concentrations
above MIC. The resulting plot of colony number versus drug concentration is the
population analysis profile (Fig. 9.1). The area under the curve (AUC) generated by
the PAP provides an integrated description of the heteroresistant subpopulations;
normalization to a reference strain lacking detectible heteroresistance provides a
single number for comparing the heteroresistance status of pathogen samples.
Although PAP can be readily applied to any pathogen that forms colonies on
solid medium, including mycobacteria [9, 10] and fungi [11], the method is very
labor intensive. Thus, it is generally used only for research purposes or to confirm
the presence of heteroresistance in a clinical setting. For research it is important to
recognize that incubation time can be a factor if the antimicrobial induces resis-
272 K. Drlica et al.
Fig. 9.2 Common types of clonal heteroresistance. Four general themes are shown schematically.
Type I represents the acquisition of a resistance mutation, either spontaneously or by horizontal
transfer, that is maintained in the population. Resistant subpopulations are enriched with each
antimicrobial challenge until resistant cells dominate the population. Examples are fluoroquino-
lone and rifampicin resistance in M. tuberculosis. In Type II heteroresistance, antimicrobial pres-
sure is needed to maintain the resistant phenotype as it is enriched. When drug pressure is removed
or relaxed, susceptible members of the population regain dominance. An example of this type of
heteroresistance is represented by gene amplification in S. enterica. Type III illustrates a situation
in which multiple pathways lower susceptibility and also reduce pathogen fitness. Fitness prob-
lems can limit the loss of susceptibility to a state called intermediate resistance. An example of
Type III heteroresistance is seen with vancomycin-intermediate S. aureus. In Type IV heteroresis-
tance, a resistance determinant enters the population by horizontal transfer and is rapidly enriched
due to continuing horizontal transfer. If the element is unstable, it can be lost when antimicrobial
pressure is reduced. Heteroresistance is seen as a balance of resistance acquisition, loss, and anti-
microbial pressure. An example of Type IV heteroresistance is methicillin-resistant S. aureus
Although a single cell can acquire resistance in a single step, pathogen populations
generally require amplification of the resistant cell. During selective amplification,
the population will be heteroresistant. Thus, heteroresistance may be a general
aspect of the emergence of resistance.
Among infections exhibiting heteroresistance are those caused by commensal
bacteria that occasionally act as pathogens. Important examples include MRSA,
vancomycin-resistant Enterococcus (VRE), Acinetobacter baumannii, Escherichia
coli, and Klebsiella pneumoniae. Repeated antimicrobial exposure, sometimes
aimed at other bacterial species, results in subpopulations of resistant mutants.
These mutant subpopulations can be enriched during treatment and thereby restrict
therapeutic options when the microbes cause infection. E. coli, a common inhabit-
ant of the human digestive tract, serves as an example. Fluoroquinolone treatment
for a variety of reasons unrelated to E. coli populations can selectively enrich
fluoroquinolone-resistant E. coli in the digestive tract. If these organisms contami-
nate the urinary tract, they can cause fluoroquinolone-resistant urinary infection,
which is now a global problem [14].
Heteroresistance is also associated with pathogens for which infection is required
for transmission. Among these obligate pathogens are M. tuberculosis and the
human immunodeficiency virus (HIV). Spontaneously resistant mutants emerge
readily, which makes every treated patient at risk for developing a resistant infec-
tion. As a result, the standard of care involves the use of multiple antimicrobials.
With tuberculosis, the antimicrobials are administered daily by healthcare workers
to assure adherence to treatment protocols. In the next section, we consider hetero-
resistance in M. tuberculosis as an example of emerging resistance with an obligate
pathogen.
With individual patients, monotherapy for tuberculosis often leads to the emergence
of resistance and treatment failure [15–17]. Host defense systems appear unable to
readily clear M. tuberculosis, and thus some tuberculosis patients harbor large num-
bers of pathogen cells (on the order of 109) [18, 19]. This feature, coupled with the
early finding that cell cultures contain resistant mutants at a high frequency (about
10−6 for isoniazid and 10−8 for rifampicin and streptomycin [17, 20–22]), led to the
idea that monotherapy simply enriches existing mutant subpopulations. More recent
measurement of mutation rate, which avoids the jackpot effects of frequency assays,
suggests that mutation rate for cultured M. tuberculosis is similar to that of other
9 Heteroresistance: A Harbinger of Future Resistance 275
bacteria (discussed by McGrath et al. [23]). Thus, lack of immune clearance, which
is especially obvious with patients coinfected with HIV, results in heavy bacterial
burden; the bacterial load is probably a key factor in the emergence of resistance
rather than an abnormally high mutation rate. With M. tuberculosis, mutagenesis
can be induced by DNA damage [24], which likely contributes to the mutagenic
effect of some antimicrobials. As expected, combination therapy largely overcomes
the rapid emergence of resistance [25].
Enrichment of mutant subpopulations is further favored by the long treatment
time required to achieve cure. During infection, M. tuberculosis is thought to
remodel its metabolism such that part of the bacterial population enters a drug-
tolerant, semi-quiescent state known as dormancy (persistence). When this state is
modeled in the laboratory, dormant bacteria are difficult to eradicate [26]. In addi-
tion, infection occurs in diverse compartments [27, 28], some of which may not be
readily accessible to active compounds [28]; moreover, some cells may be non-
culturable but viable. Consequently, antibiotic treatment must be maintained for
many months to be effective. When adherence to treatment is poor, drug exposure
becomes intermittent, which allows cycles of population expansion followed by
selective reductions. These cycles also occur with many other pathogens but usually
in different patients rather than in a single individual. The easily observed emer-
gence of resistance with M. tuberculosis has made tuberculosis a paradigm for
understanding the process.
Awareness of tuberculosis heteroresistance is high in part because PCR-based
detection of resistant subpopulations is straightforward: resistance arises from point
mutations in an otherwise highly conserved pathogen genome (reviewed in [23]).
Moreover, the need for rapid diagnosis has been a high priority, which has led to the
widespread application of DNA-based methods. These tests now show that 10–20%
of patients in localities of high tuberculosis incidence have infections containing
diverse subpopulations. Heterogeneity is seen with both HIV-positive and HIV-
negative patients [29], and it is detected with a variety of genes, including those that
encode resistance to ethambutol [30–33], isoniazid [32–34], rifampicin [34], fluoro-
quinolones [32, 35], streptomycin [33], pyrazinamide [32], and amikacin [36].
Thus, M. tuberculosis heteroresistance within individual patients is common to
many antimicrobials.
A study from South Africa illustrates the complex dynamics of clonal heteroresis-
tance [32]. The subjects of the study suffered from multidrug-resistant tuberculosis
(MDR-TB) that had persisted through more than 12 months of treatment. Since the
overall prevalence of MDR-TB in the community was low (0.3% in new patients,
1.7% in previously treated patients), clonal heterogeneity was more likely to occur
than mixed infections. Indeed, when sputum samples from 13 HIV-negative
MDR-TB patients were examined at 2-week intervals, no evidence was found for
superinfection: all carried bacteria having a single IS6110 RFLP type and spoligo-
type pattern. Nucleotide sequence analysis for eight resistance genes (katG,
inhA = isoniazid; pncA = pyrazinamide; embB = ethambutol; rrs = amikacin, kana-
mycin; rpsL = streptomycin; gyrA = fluoroquinolone; rpoB = rifampicin) showed
that several of the infections changed resistance patterns over the course of
sampling.
One patient in the South African study [32] was examined for mutations in katG,
embB, and gyrA during 56 weeks of therapy. At the beginning of sampling, all three
genes were wild type, but at weeks 4 and 6, the katG marker was scored as resistant.
Subsequent samples showed that it returned to wild type. The embB marker con-
verted to resistant by 6 weeks, and it remained resistant throughout the observation
period. The gyrA gene showed a mixture of alleles at week 6, and in subsequent
samples transient changes were observed among several gyrA resistance forms,
often mixed with wild-type alleles. Even after 48 weeks, gyrA was a mixture of
resistant (D94C) and wild-type alleles. By week 52 a different gyrA allele (D94G)
had emerged as the dominant form.
Specimens from two other patients [32] also contained different alleles of genes
involved in drug resistance. For example, one patient evolved a mixture of wild-type
and resistant alleles for pncA, changes in gyrA alleles over time, and a mixture of rrs
alleles at the beginning of sampling that later saw one allele emerge as dominant.
Another patient began with wild-type gyrA that after 36 weeks changed to resistant.
But after 48 weeks, gyrA returned to wild type. Wild-type pncA also persisted until
week 36, and then it shifted to resistant for the remainder of the study (week 52).
The katG gene began as wild type, but after week 6 it was resistant, except for one
sample at week 30 that was wild type. Overall, these fluctuations in drug-resistance
markers illustrate the dynamic and varied nature of clonal heteroresistance when
sputum samples are the source of information.
Examination of lung tissue provides an explanation for the allelic diversity: het-
eroresistance measured in sputum samples arises at least in part from independent
clonal evolution in different regions of the lung. When surgical samples were exam-
ined from three HIV-negative patients having undergone long-term therapy, DNA
IS6110 fingerprints were the same for bacteria from different regions of the lung;
thus, the isolates from individual patients appeared to be clonally related [27]. One
patient had a streptomycin-resistant strain in an open lesion, while wild-type cells
were detected in a closed granuloma. Wild-type cells were also recovered from
278 K. Drlica et al.
sputum. Bacteria from a second patient carried two different gyrA resistance alleles
when isolated from open lesions, while wild-type gyrA alleles were obtained from
sputum and from two closed lesions. The third patient produced three types of M.
tuberculosis: (1) cells from apparently normal lung tissue had wild-type genes for
katG, embB, and rrs, (2) cells from sputum and four pathological sites had katG and
embB resistance markers but wild-type rrs, and (3) another pathological site yielded
bacteria with resistance for all three genes. These observations, plus similar findings
in another study [44] and autopsies [45], led to the conclusion that evolution occurs
independently in different lung compartments and that wild-type cells can survive
treatment. The results of sputum-based analyses probably reflect opening of granu-
lomas and release of bacteria at different times during infection. Thus, analysis of a
single sputum sample may not accurately reflect the diversity of bacterial popula-
tions in the infection.
changing requires rapid and accurate diagnostic methods. Below we consider the
development of genetic (DNA-based) assays.
Although M. tuberculosis population heterogeneity had been known for many years
from phage-typing of M. tuberculosis subpopulations [47, 48], it was recognized as
an important phenomenon only after molecular diagnostic methods emerged. When
PCR was used to amplify specific regions of M. tuberculosis DNA encoding pro-
teins associated with resistance and the amplified fragments were separated by gel
electrophoresis, the size distribution characteristic of both susceptible and resistant
alleles was observed from a single bacterial specimen [33]. Heteroresistance was
then used to explain the occasional discordance between results from drug-
susceptibility testing and DNA-based methods: the DNA tests indicated resistance,
but only susceptibility was detected following the bacterial outgrowth required for
susceptibility testing. A fitness advantage among the susceptible bacteria was
thought to allow them to dominate during outgrowth [49].
The various DNA-based tools differ in sensitivity (Table 9.1). For example,
Sanger DNA sequencing of PCR products reported 15% of isolates as heteroresis-
tant, while with the same samples deep sequencing found almost 40% heteroresis-
tance [35]. When heteroresistance is greater than a few percent, current hybridization
methods are sufficiently sensitive. Unfortunately, PCR-based diagnostic methods
encounter a specificity problem when subpopulations are below 1%, because tem-
plates from the major bacterial population can generate false-positive, variant sig-
nals due to mis-priming, mis-incorporation, and mis-hybridization.
As pointed out above, sensitivity to 1% is unlikely to be adequate for monitoring
the emergence of resistance, because 1% is considered fully resistant for that marker
if the equivalent of monotherapy is employed. Moreover, a negative result cannot
rule out heteroresistance. In essence, current genetic diagnostics can give false-
negative results. We conclude that other methods are needed to detect heteroresis-
tance at levels low enough to allow successful intervention.
Work in cancer biology is driving new, DNA-based tests for heteroresistance – a
priority in the cancer field is detection of a small number of transformed cells within
a large background of normal cells. One approach is called digital PCR [57]. In this
method, the sample is diluted into a series of wells in a multi-well microfluidic plate
such that only a single molecule of mutant DNA is expected to be present in a given
well (most wells will contain only wild-type DNA). Amplification of DNA in the
wells produces a digital readout: either the presence or absence of mutant DNA. The
fraction of total wells scoring positive estimates the percent of the sample contain-
ing mutant DNA. In principle, the sensitivity of this method is limited only by the
number of wells assayed. Digital PCR has been applied to M. tuberculosis isolates
by mixing wild-type cells with M. tuberculosis containing resistance mutations in
katG, rpoB, gyrA, and rrs. The method reliably detects heteroresistance at a ratio of
280 K. Drlica et al.
1 mutant per 1000 wild-type cells [50], which is about 10-times more sensitive than
previous PCR-based methods. For digital PCR to achieve this sensitivity with spu-
tum samples, the samples must contain more than 1000 M. tuberculosis cells per ml
(bacillary content varies among sputum samples, but it can exceed one million cfu
[58–60]).
Another approach, single-nucleotide primer extension, is used to incorporate a
nucleotide having a distinctive mass modification that can be identified by mass
spectroscopy [61]. The method, called iPLEX Gold, has the advantage of detecting
multiple resistance alleles in the same reaction mixture. In one application of the
method, a reconstruction experiment detected one amikacin-resistant cell per 200
wild-type cells [36].
A third strategy is called pyrophosphorolysis-activated polymerization [62–67].
In this method, a primer containing a dideoxyribonucleotide at its 3′ terminus (noted
as P*) is hybridized to the test DNA at the preselected mutation site. Removal of the
dideoxyribonucleotide by pyrophosphorolysis, which is highly specific for perfect
hybridization of the primer, is required for extension of the primer by DNA
polymerase. Primer extension then amplifies the signal for real-time detection by
fluorescent probes. When a P* primer is used that contains the complement of the
mutant sequence, the polymerization assay is expected to detect mutant alleles at a
frequency as low as 10−8 of wild-type DNA, a level that approaches background
(spontaneous) mutation frequency. To our knowledge, the pyrophosphorolysis-
activated polymerization method has not been applied to detection of heteroresis-
tant M. tuberculosis.
9 Heteroresistance: A Harbinger of Future Resistance 281
A fourth strategy, which is also derived from cancer diagnosis, uses what are
called SuperSelective primers for real-time PCR assays [68]. In this system, a DNA
primer is designed in which one region hybridizes strongly to a portion of the target
DNA being queried. This anchor region is separated from a detector region, the
“foot”, by a long region expected to mispair with the target and thus form a loop.
The foot is designed to hybridize only with the mutant sequence in the target. By
adjusting the length of the loop and the foot, conditions can be obtained in which
hybridization only occurs with mutant DNA. The resulting hybrid then primes real-
time PCR. The system can detect multiple mutations in the same reaction tube by
using fluorophores having different colors to discriminate the amplification prod-
ucts. This method has not yet been applied to diagnosis of heteroresistance.
A fifth strategy is based on CRISPR, a bacterial system that recognizes and
destroys foreign nucleic acids. The underlying idea is as follows. A DNA sample
from the pathogen is transcribed in vitro and incubated with the Cas13a protein
system plus a quenched, fluorescently labeled reporter RNA. Recognition of the
target RNA by Cas13a, which is designed to occur only if the resistance mutation is
present, will cause collateral damage in the reporter RNA, eliminate the quenching,
and generate a fluorescent signal. This method, which has been dubbed SHERLOCK
[69], has single-molecule sensitivity, similar to droplet digital PCR and quantitative
PCR (qPCR). Moreover, it has point-of-care diagnostic features. To our knowledge
SHERLOCK has not been applied to detection of heteroresistant M. tuberculosis.
However, the CRISPR system has been modified to function in this pathogen [70].
A general problem associated with PCR-based diagnosis of resistant bacterial
subpopulations is laboratory contamination by amplicons present in the laboratory
from previous tests. Published estimates of laboratory cross-contamination using
open-tube methods are presently almost 4% [38, 40]. Although closed-tube methods
exist [53, 71, 72], current closed-tube methods require refinement to be sensitive
enough for heteroresistance detection.
Heteroresistance with S. aureus, which has been known for many years [73], is not
routinely detected by standard susceptibility testing (MIC determination). Such
determinations typically examine only 104 to 105 cells, and the frequency of resis-
tant subpopulations is generally below 10−5. However, when susceptibility testing
uses a large number of cells, on the order of 107 to 1010, subpopulations having
reduced susceptibility can be seen. For example, heterogeneity is a distinctive fea-
ture of methicillin resistance due to the presence of a mobile chromosomal element
called SCCmec [6, 74]. SCCmec elements, which vary in size, contain a gene, mecA,
that encodes a low-affinity penicillin-binding protein (PBP2’ or PBP2a). PBP2’ is a
transpeptidase [75] that allows S. aureus to form cell walls in the presence of
282 K. Drlica et al.
MRSA infection is commonly treated with the glycopeptide vancomycin. The result
has been the emergence of an intermediate level of resistance (VISA, which is dis-
tinct from the rare, vanA-mediated, fully vancomycin-resistant S. aureus). VISA is
associated with a poorly defined thickening of the bacterial cell wall that reduces the
uptake of vancomycin [79]. Other features associated with VISA are excess pepti-
doglycan production, low fitness manifested by reversion toward susceptibility dur-
ing growth in vitro [80], and attenuated virulence in animal models of infection
[81–84]. VISA probably represents the end stage of evolution from heteroresistant
strains (hVISA) in which subpopulations slightly elevate the overall MIC of an
MRSA isolate. Since clinical isolates of MRSA are heteroresistant due to instability
of the SCCmec elements [85], hVISA emerging during vancomycin treatment can
be co-heteroresistant (heteroresistant for two or more antimicrobials).
With both hVISA and VISA, S. aureus populations exhibit considerable hetero-
geneity in their susceptibility to vancomycin (see examples in Ref. [5]). To better
detect hVISA, the breakpoint for full susceptibility was lowered to an MIC of 2 μg/
ml [86]. In some samples, hVISA cells are abundant enough to raise vancomycin
9 Heteroresistance: A Harbinger of Future Resistance 283
MIC to the high end of the susceptible range (MIC = 0.5 to <2 μg/ml; with most
antimicrobial-pathogen combinations, heteroresistance has no effect on MIC
because the subpopulations are small). Thus, a slightly elevated MIC can indicate
hVISA, but a more sensitive test is needed.
PAP (population analysis profile) determination can definitively identify hVISA,
but the method is too labor intensive for routine clinical use. Consequently, efforts
have shifted toward other agar-plate screens [87]. Developing DNA tests is still
challenging, because multiple genetic pathways lead to VISA. For example, when
seven successive samples of MRSA were obtained from a single patient and ana-
lyzed by vancomycin-PAP, a clear evolution from susceptible to hVISA to VISA
was observed [88]. Whole-genome sequencing then identified six mutations that
generated five distinct genetic profiles that correlated with evolution along three
pathways involving cell wall biology.
In another whole-genome sequencing study, a VISA strain was compared to a
closely related, susceptible isolate. Several gene differences were found, and wild-
type alleles were introduced into the VISA strain to regain susceptibility and thereby
identify genes involved in VISA [79, 89]. Among these were the GraSR and VraSR
two-component systems that contribute to the evolution of VISA by upregulating
cell wall synthesis (for additional detail, see Chap. 15). Other associated alterations,
listed in Ref. [77], include a substitution in RNA polymerase (H481Y/L/N) that also
confers rifampicin resistance (whether rifampicin resistance is acquired before
VISA is not known). We conclude that unlike the situation with M. tuberculosis
heteroresistance, a simple DNA-based diagnostic for hVISA is not likely to be
available soon.
The clinical importance of VISA and hVISA can be assessed by surveillance
studies. VISA represents a few percent of MRSA recovered from serious infections,
and hVISA prevalence is 4 to 5 times higher [90, 91]. Apparently, the genetic
changes that create VISA have fitness costs that trap most S. aureus in the
intermediate-resistance state. Since antimicrobial resistance is inherently a local
phenomenon, variation in prevalence is expected (assay methods also differ, which
adds to variation). For example, surveys from Asia (excluding China) indicate a
prevalence for hVISA of a few percent [92], but a report from Taiwan places it at
about 10% (2012–2013; Ref. [91]). A hospital study from Turkey also reported a
high prevalence of hVISA in blood isolates (almost 14%, Ref. [93]). In a troubling
study from Michigan (cited in Ref. [94]), the prevalence of hVISA in blood isolates
increased from 2% (1987–1993) to 8% (2003–2007), and in a multicenter US study,
the prevalence increased from 0.4% in 2009 to 1.2% in 2011. Thus, hVISA dissemi-
nation within and between hospitals is taken seriously; it may require special con-
sideration by infection control departments to limit transmission from patients
colonized by or infected with hVISA.
hVISA surveillance data also suggest that treatment changes are needed.
Although VISA is associated with treatment failure, reports on hVISA are mixed.
For example, a Michigan study of MRSA bacteremia found that hVISA was not
clearly associated with treatment failure [90]. However, in another study, hVISA
284 K. Drlica et al.
for the two countries). While the reasons for differences in heteroresistance are com-
plex, these data show that clinicians in developing countries should be watchful for
multiple infections that might impact susceptibility testing.
Klebsiella pneumoniae causes serious diseases, such as pneumonia, meningitis,
and urinary infections. Since K. pneumoniae inhabits the human digestive tract, it
readily disseminates in hospitals by fecal contamination. Thus, when multidrug-
resistant K. pneumoniae strains develop resistance to carbapenems, they become a
major nosocomial problem. Low reproducibility of MIC tests for carbapenems, fol-
lowed by population analysis profiling, led to the conclusion that K. pneumoniae
heteroresistance is overlooked by automated susceptibility testing [110].
Heteroresistance appears to arise from drug-induced expression of carbapenemases,
since heteroresistance to meropenem is lost when drug pressure is withdrawn [111].
As the prevalence of resistance to the major antimicrobials mounts, colistin is
being used to treat K. pneumoniae infections. The result has been a sharp increase
in colistin resistance. For example, in one Greek hospital, resistance to colistin rose
from 0% in 2007, to 8% in 2008, and 24% in 2009 [112]. When PAP was performed
on a small set of patient isolates, heteroresistance to colistin was observed in 12/16
isolates that had been deemed susceptible by standard MIC assays [112]. With K.
pneumoniae, colistin heteroresistance is associated with the PhoPQ regulatory sys-
tem [119], as pointed out below for E. cloacae. The PhoPQ system alters the lipo-
polysaccharide of cell surfaces (the negative charge on lipid A is reduced, thereby
lowering the affinity for colistin, a cationic peptide). Colistin monotherapy is con-
traindicated for serious disease caused by K. pneumoniae.
Pseudomonas aeruginosa is an opportunistic pathogen that is particularly prob-
lematic for patients suffering from cystic fibrosis. Antimicrobial resistance with P.
aeruginosa is mediated by multiple efflux systems and production of β-lactamases.
In a study from Greece, 27% of presumably susceptible isolates exhibited stable
carbapenem heteroresistance [114]. This result may be common for P. aeruginosa,
as a similar finding was reported from China [115]. With P. aeruginosa, it may be
necessary to perform heteroresistance testing on many isolates, since automated
methods do not reliably detect heteroresistance.
Salmonella enterica serovar Typhimurium is noted for causing outbreaks of food
poisoning. Since isolates that exhibit multidrug resistance are associated with
increased mortality and morbidity, colistin is being considered for treatment of S.
enterica-associated diseases. A study of laboratory-generated colistin heteroresis-
tance with S. enterica revealed a correlation between heteroresistance and a
moderate gene dosage of pmrD, a gene that upregulates proteins that modify lipid A
and thereby lower susceptibility to colistin [116]. Successive passages in the pres-
ence of colistin increased amplification of pmrD, while the number of amplified
copies declined when cells were passaged on drug-free medium. A similar phenom-
enon may have contributed to tetracycline heteroresistance in a clinical isolate
[116]. Antimicrobial resistance arising from gene amplification has also been
observed with M. tuberculosis [120], suggesting that it may underlie heteroresis-
tance in a variety of pathogens.
288 K. Drlica et al.
Major Points
• Antimicrobial heteroresistance derives from a variety of phenomena ranging
from subpopulations of stable, fully resistant mutants to reversible, antimicrobial-
mediated induction or amplification of protective genes.
• Heteroresistance is common: it has been observed in many different pathogenic
bacterial species and found in almost 25% of patient isolates
• Heteroresistance can evolve to full drug resistance.
• The importance of heteroresistance has been underappreciated, because infec-
tions containing heteroresistant pathogen populations can often be treated
successfully.
290 K. Drlica et al.
Acknowledgments We thank the following for helpful discussions and critical comments:
Veronique Dartois, Dorothy Fallows, Marila Gennaro, Ben Gold, Barry Kreiswirth, Richard Pine,
and George Zhanel. The authors’ work was supported by NIH grants 1DP20D007423,
1R01AI073491, 1R21A03781, and 1R01AI87671.
References
15. East African Hospitals and British Medical Research Council. Comparative trial of isoniazid
alone in low and high dosage and isoniazid plus PAS in the treatment of acute pulmonary
tuberculosis in East Africa. Tubercle. 1960;40:83–102.
16. Crofton J, Mitchison D. Streptomycin resistance in pulmonary tuberculosis. Br Medical
J. 1948;2:1009–15.
17. Canetti G, leLerzin M, Porven G, Rist N, Grumbach F. Some comparative apects of rifampi-
cin and isoniazid. Tubercle. 1968;49:367–76.
18. Mitchison DA. Drug resistance in mycobacteria. Brit. Med Bull. 1984;40:84–90.
19. Canetti G. The J. Burns Amberson Lecture: present aspects of bacterial resistance in tubercu-
losis. Am Rev Resp Dis. 1965;92:687–703.
20. Lorian V, Finland M. In vitro effect of rifampicin on mycobacteria. App Microbiol.
1969;17:202–7.
21. Hobby G, Lenert T, Mater-Engallena J. In vitro activity of rifampin against the H37Rv strain
of Mycobacterium tuberculosis. Am Rev Respir Dis. 1968;99:453–6.
22. Canetti G, Froman S, Grosset J, Hauduroy P, Langerová M, Mahler H, et al. Mycobacteria:
laboratory methods for testing drug sensitivity and resistance. Bull World Health Organ.
1963;29:564–78.
23. McGrath M, GeyvanPittius N, vanHelden P, Warren R, Warner D. Mutation rate and the
emergence of drug resistance in Mycobacterium tuberculosis. J Antimicrob Chemother.
2014;69:292–302.
24. Warner D, Ndwandwe D, Abrahams G, Kana B, Machowski E, Venclovas C, et al. Essential
roles for imuA’- and imuB-encoded accessory factors in DnaE2-dependent mutagenesis in
Mycobacterium tuberculosis. Proc Natl Acad Sci U S A. 2010;107:13093–8.
25. Fox W, Elklard G, Mitchison D. Studies on the treatment of tuberculosis undertaken by the
British Medical Research Council Tuberculosis Units, 1946–1986, with relevant subsequent
publications. Int J Tuberc Lung Dis. 1999;3:S231–S79.
26. Wayne LG, Hayes LG. An in vitro model for sequential study of shiftdown of Mycobacterium
tuberculosis through two stages of nonreplicating persistence. Infect Immun. 1996;64:2062–9.
27. Kaplan G, Post F, Moreira A, Wainwright H, Kreiswirth B, Tanverdi M, et al. Mycobacterium
tuberculosis growth at the cavity surface: a microenvironment with failed immunity. Infect
Immun. 2003;71:7099–108.
28. Prideaux B, Via L, Zimmerman M, Eum S, Sarathy J, O'Brien P, et al. The association between
sterilizing activity and drug distribution into tuberculosis lesions. Nat Med. 2015;21:1223–7.
29. Zetola N, Modongo C, Moonan P, Ncube R, Matlhagela K, Sepako E, et al. Clinical out-
comes among persons with pulmonary tuberculosis caused by Mycobacterium tuberculosis
isolates with phenotypic heterogeneity in results of drug-susceptibility tests. J Infect Dis.
2014;209:1754–63.
30. Cullen M, Sam N, Kanduma E, McHugh T, Gillespie S. Direct detection of heteroresistance in
Mycobacterium tuberculosis using molecular techniques. J Med Microbiol. 2006;55(8):1157.
31. Plinke C, Cox H, Kalon S, Doshetov D, Rüsch-Gerdes S, Niemann S. Tuberculosis etham-
butol resistance: concordance between phenotypic and genotypic test results. Tuberculosis.
2009;89:448–52.
32. Post F, Willcox P, Mathema B, Steyn L, Shean K, Ramaswamy S, et al. Genetic polymor-
phism in Mycobacterium tuberculosis isolates from patients with chronic multidrug-resistant
tuberculosis. J Inf Dis. 2004;190:99–106.
33. Rinder H, Mieskes K, Loscher T. Heteroresistance in Mycobacterium tuberculosis. Int
J Tuberc Lung Dis. 2001;5:339–45.
34. Hofmann-Thiel S, van Ingen J, Feldmann K, Turaev L, Uzakova G, Murmusaeva G, et al.
Mechanisms of heteroresistance to isoniazid and rifampin of Mycobacterium tuberculosis in
Tashkent, Uzbekistan. Eur Respir J. 2009;33:368–74.
35. Eilertson B, Maruri F, Blackman A, Herrera M, Samuels D, Sterling T. High proportion of
heteroresistance in gyrA and gyrB in fluoroquinolone-resistant Mycobacterium tuberculosis
clinical isolates. Antimicrob Agents Chemother. 2014;58:3270–5.
292 K. Drlica et al.
36. Zhang X, Zhao B, Huang H, Zhu Y, Peng J, Dai G, et al. Co-occurrence of amikacin-resistant
and -susceptible Mycobacterium tuberculosis isolates in clinical samples from Beijing,
China. J Antimicrob Chemother. 2013;68:1537–42.
37. Shamputa I, Rigouts L, Eyongeta L, El Aila N, van Deun A, Salim A, et al. Genotypic and
phenotypic heterogeneity among Mycobacterium tuberculosis isolates from pulmonary
tuberculosis patients. J Clin Microbiol. 2004;42:5528–36.
38. van Rie A, Victor T, Richardson M, Johnson R, vanderSpuy G, Murray E, et al. Reinfection
and mixed infection cause changing Mycobacterium tuberculosis drug-resistance patterns.
Am J Respir Crit Care Med. 2005;172:636–42.
39. Blaas S, Mütterlein R, Weig J, Neher A, Salzberger B, Lehn N, et al. Extensively drug resis-
tant tuberculosis in a high income country: a report of four unrelated cases. BMC Infect Dis.
2008;8:60.
40. Warren R, Victor T, Streicher E, Richardson M, Beyers N, GeyvanPittius N, et al. Patients
with active tuberculosis often have different strains in the same sputum specimen. Am J Crit
Care Med. 2004;169:610–4.
41. Meacci F, Orrù G, Iona E, Giannoni F, Piersimoni C, Pozzi G, et al. Drug resistance evolu-
tion of a Mycobacterium tuberculosis strain from a noncompliant patient. J Clin Microbiol.
2005;43:3114–20.
42. Xu C, Kreiswirth B, Sreevatsan S, Musser J, Drlica K. Fluoroquinolone resistance associ-
ated with specific gyrase mutations in clinical isolates of multidrug-resistant Mycobacterium
tuberculosis. J. Inf. Dis. 1996;174:1127–30.
43. Shamputa I, Jugheli L, Sadradze N, Willery E, Portaels F, Supply P, et al. Mixed infection and
clonal representativeness of a single sputum sample in tuberculosis patients from a peniten-
tiary hospital in Georgia. Respir Res. 2006;7:99.
44. Vadwai V, Daver G, Udwadia Z, Sadani M, Shetty A, Rodrigues C. Clonal population
of Mycobacterium tuberculosis strains reside within multiple lung cavities. PLoS One.
2011;6:e24770.
45. Lieberman T, Wilson D, Misra R, Xiong L, Moodley P, Cohen T, et al. Genomic diversity in
autopsy samples reveals within-host dissemination of HIV-associated Mycobacterium tuber-
culosis. Nat Med. 2016;22:1470–4.
46. Behr M. Tuberculosis due to multiple strains: a concern for the patient? A concern for tuber-
culosis control? Am J Respir Crit Care Med. 2004;169(5):554.
47. Mankiewicz E, Liivak M. Phage types of Mycobacterium tuberculosis in cultures isolated
from Eskimo patients. Am Rev Respir Dis. 1975;111:307–12.
48. Bates J, Stead W, Rado T. Phage type of tubercle bacilli isolated from patients with two or
more sites of organ involvement. Am Rev Respir Dis. 1976;114:353–8.
49. Parsons L, Salfinger M, Clobridge A, Dormandy J, Mirabello L, Polletta V, et al. Phenotypic
and molecular characterization of Mycobacterium tuberculosis isolates resistant to both iso-
niazid and ethambutol. Antimicrob Agents Chemother. 2005;49:2218–25.
50. Pholwat S, Stroup S, Foongladda S, Houpt E. Digital PCR to detect and quantify heteroresis-
tance in drug resistant Mycobacterium tuberculosis. PLoS One. 2013;8:e57238.
51. Folkvardsen D, Svensson E, Thomsen V, Rasmussen E, Bang D, Werngren J, et al. Can
molecular methods detect 1% isoniazid resistance in Mycobacterium tuberculosis? J Clin
Microbiol. 2013a;51:1596–9.
52. Folkvardsen D, Thomsen V, Rigouts L, Rasmussen E, Bang D, Bernaerts G, et al. Rifampin
heteroresistance in Mycobacterium tuberculosis cultures as detected by phenotypic and geno-
typic drug susceptibility test methods. J Clin Microbiol. 2013b;51:4220–2.
53. Hu S, Li G, Li H, Liu X, Niu J, Quan S, et al. Rapid detection of isoniazid resistance in
Mycobacterium tuberculosis isolates by use of real-time-PCR-based melting curve analysis.
J Clin Microbiol. 2014;52:1644–52.
54. Chakravorty S, Kothari H, Aladegbami B, Cho E, Lee J, Roh S, et al. Rapid, high-throughput
detection of rifampin resistance and heteroresistance in Mycobacterium tuberculosis by use of
sloppy molecular beacon melting temperature coding. J Clin Microbiol. 2012;50:2194–202.
9 Heteroresistance: A Harbinger of Future Resistance 293
55. Chakravorty S, Aladegbami B, Thoms K, Lee J, Lee E, Rajan V, et al. Rapid detection of
fluoroquinolone-resistant and heteroresistant Mycobacterium tuberculosis by use of sloppy
molecular beacons and dual melting-temperature codes in a real-time PCR assay. J Clin
Microbiol. 2011;49:932–40.
56. de Oliveira M, da Silva-Rocha A, Cardoso-Oelemann M, Gomes H, Fonseca L, Werneck-
Barreto A, et al. Rapid detection of resistance against rifampicin in isolates of Mycobacterium
tuberculosis from Brazilian patients using a reverse-phase hybridization assay. J Microbiol
Methods. 2003;53:335–42.
57. Vogelstein B, Kinsler K. Digital PCR. Proc Natl Acad Sci U S A. 1999;96:9236–41.
58. Yajko D, Wagner C, Tevere V, Kocagöz T, Hadley W, Chambers H. Quantitative culture of
Mycobacterium tuberculosis from clinical sputum specimens and dilution endpoint of its
detection by the Amplicor PCR assay. J Clin Microbiol. 1995;33:1944–7.
59. Diacon A, Patientia R, Venter A, van Helden P, Smith P, McIlleron H, et al. Early bactericidal
activity of high-dose rifampin in patients with pulmonary tuberculosis evidenced by positive
sputum smears. Antimicrob Agents Chemother. 2007;51:2994–6.
60. Brindle R, Odhiambo J, Mitchison D. Serial counts of Mycobacterium tuberculosis in sputum
as surrogate markers of the sterilising activity of rifampicin and pyrazinamide in treating
pulmonary tuberculosis. BMC Pulm Med. 2001;1:2.
61. Bouakaze C, Keyser C, Gonzalez A, Sougakoff W, Veziris N, Dabernat H, et al. Matrix-assisted
laser desorption ionization-time of flight mass spectrometry-based single nucleotide polymor-
phism genotyping assay using iPLEX gold technology for identification of Mycobacterium
tuberculosis complex species and lineages. J Clin Microbiol. 2011;49(9):3292.
62. Liu Q, Sommer S. Pyrophosphorolysis-activated polymerization (PAP): application to allele-
specific amplification. BioTechniques. 2000;29:1072–83.
63. Liu Q, Nguyen VQ, Li X, Sommer SS. Multiplex dosage pyrophosphorolysis-activated
polymerization: application to the detection of heterozygous deletions. BioTechniques.
2006;40(5):661–8.
64. Liu Q, Sommer SS. Pyrophosphorolysis-activatable oligonucleotides may facilitate detection
of rare alleles, mutation scanning and analysis of chromatin structures. Nucleic Acids Res.
2002;30(2):598–604.
65. Liu Q, Sommer SS. PAP: detection of ultra rare mutations depends on P* oligonucle-
otides: “sleeping beauties” awakened by the kiss of pyrophosphorolysis. Hum Mutat.
2004a;23(5):426–36.
66. Liu Q, Sommer SS. Detection of extremely rare alleles by bidirectional pyrophosphorolysis-
activated polymerization allele-specific amplification (Bi-PAP-A): measurement of mutation
load in mammalian tissues. BioTechniques. 2004b;36(1):156–66.
67. Liu Q, Sommer SS. Pyrophosphorolysis by Type II DNA polymerases: implications for
pyrophosphorolysis-activated polymerization. Anal Biochem. 2004c;324(1):22–8.
68. Vargas D, Kramer F, Tyagi S, Marras S. Multiplex real-time PCR assays that measure the
abundance of extremely rare mutations associated with cancer. PLoS One. 2016;11:e0156546.
69. Gootenberg J, Abudayyeh O, Lee J, Essletzbichler P, Dy A, Joung J, et al. Nucleic acid detec-
tion with CRISPR-Cas13a/C2c2. Science. 2017;pii: eaam9321.
70. Rock J, Hopkins F, Chavez A, Diallo M, Chase M, Gerrick E, et al. Programmable transcrip-
tional repression in mycobacteria using an orthogonal CRISPR interference platform. Nat
Microbiol. 2017;2:16274.
71. Rice J, Reis A, Rice L, Carver-Brown R, Wangh L. Fluorescent signatures for variable DNA
sequences. Nucleic Acids Res. 2012;40:e164.
72. Huang Q, Liu Z, Liao Y, Chen X, Zhang Y, Li Q. Multiplex fluorescence melting curve analysis
for mutation detection with dual-labeled, self-quenched probes. PLoS One. 2011;6:e19206.
73. Seligman S. Methicillin-resistant staphylococci: genetics of the minority population. J Gen
Microbiol. 1966;42:315–22.
74. Ryffel C, Strässle A, Kayser F, Berger-Bächi B. Mechanisms of heteroresistance in
methicillin-resistant Staphylococcus aureus. Antimicrob Agents Chemother. 1994;38:724–8.
294 K. Drlica et al.
75. Lim D, Strynadka N. Structural basis for the beta lactam resistance of PBP2a from methicillin-
resistant Staphylococcus aureus. Nat Struct Biol. 2002;9:870–6.
76. Aiba Y, Katayama Y, Hishinuma T, Murakami-Kuroda H, Cui L, Hiramatsu K. Mutation
of RNA polymerase β-subunit gene promotes heterogeneous-to-homogeneous conversion
of β-lactam resistance in methicillin-resistant Staphylococcus aureus. Antimicrob Agents
Chemother. 2013;57:4861–71.
77. Alam M, Petit R, Crispell E, Thornton T, Conneely K, Jiang Y, et al. Dissecting vancomycin-
intermediate resistance in Staphylococcus aureus using genome-wide association. Genome
Biol Evol. 2014;6:1174–85.
78. Saravolatz SN, Martin H, Pawlak J, Johnson LB, Saravolatz LD. Ceftaroline-heteroresistant
Staphylococcus aureus. Antimicrob Agents Chemother. 2014;58(6):3133.
79. Cui L, Neoh H, Shoji M, Hiramatsu K. Contribution of vraSR and graSR point mutations
to vancomycin resistance in vancomycin-intermediate Staphylococcus aureus. Antimicrob
Agents Chemother. 2009;53:1231–4.
80. Boyle-Vavra S, Berke S, Lee J, Daum R. Reversion of the glycopeptide resistance phenotype
in Staphylococcus aureus clinical isolates. Antimicrob Agents Chemother. 2000;44:272–7.
81. Cameron D, Ward D, Kostoulias X, Howden B, Moellering R, Eliopoulos G, et al. Serine/
threonine phosphatase Stp1 contributes to reduced susceptibility to vancomycin and viru-
lence in Staphylococcus aureus. J Inf Dis. 2012;205:1677–87.
82. Gao W, Cameron D, Davies J, Kostoulias X, Stepnell J, Tuck K, et al. The RpoB H481Y
rifampicin resistance mutation and an active stringent response reduce virulence and increase
resistance to innate immune responses in Staphylococcus aureus. J Inf Dis. 2013;207:929–39.
83. Majcherczyk P, Barblan J, Moreillon P, Entenza J. Development of glycopeptide-intermediate
resistance by Staphylococcus aureus leads to attenuated infectivity in a rat model of endocar-
ditis. Microb Pathog. 2008;45:408–14.
84. Peleg A, Monga D, Pillai S, Mylonakis E, Moellering R, Eliopoulos G. Reduced susceptibil-
ity to vancomycin influences pathogenicity in Staphylococcus aureus infection. J Inf Dis.
2009;199:532–6.
85. Donnio P, Oliveira D, Faria N, Wilhelm N, LeCoustumier A, deLencastre H. Partial exci-
sion of the chromosomal cassette containing the methicillin resistance determinant results in
methicillin-susceptible Staphylococcus aureus. J Clin Microbiol. 2005;43:4191–3.
86. Tenover F, Moellering R. The rationale for revising the Clinical and Laboratory Standards
Institute vancomycin minimal inhibitory concentration interpretive criteria for Staphylococcus
aureus. Clin Inf Dis. 2007;44:1208–15.
87. Satola S, Farley M, Anderson K, Patel J. Comparison of detection methods for heteroresis-
tant vancomycin-intermediate Staphylococcus aureus, with the population analysis profile
method as the reference method. J Clin Microbiol. 2011;49:177–83.
88. Chen C, Huang Y, Chiu C. Multiple pathways of cross-resistance to glycopeptides and dapto-
mycin in persistent MRSA bacteraemia. J Antimicrob Chemother. 2015;70:2965–72.
89. Hafer C, Lin Y, Kornblum J, Lowy F, Uhlemann A. Contribution of selected gene muta-
tions to resistance in clinical isolates of vancomycin-intermediate Staphylococcus aureus.
Antimicrob Agents Chemother. 2012;56:5845–51.
90. Khatib R, Jose J, Musta A, Sharma M, Fakih M, Johnson L, et al. Relevance of vancomycin-
intermediate susceptibility and heteroresistance in methicillin-resistant Staphylococcus
aureus bacteraemia. J Antimicrob Chemother. 2011;66:1594–9.
91. Huang S, Chen Y, Chuang Y, Chiu S, Fung C, Lu P, et al. Prevalence of vancomycin-
intermediate Staphylococcus aureus (VISA) and heterogeneous VISA among methicillin-
resistant S. aureus with high vancomycin minimal inhibitory concentrations in Taiwan: a
multicenter surveillance study, 2012–2013. J Microbiol Immunol Infect. 2016;49:701–7.
92. Chung D, Lee C, Kang Y, Baek J, Kim S, Ha Y, et al. Genotype-specific prevalence of
heterogeneous vancomycin-intermediate Staphylococcus aureus in Asian countries. Int
J Antimicrob Agents. 2015;46:338–41.
9 Heteroresistance: A Harbinger of Future Resistance 295
93. Sancak B, Yagci S, Gür D, Gülay Z, Ogunc D, Söyletir G, et al. Vancomycin and dapto-
mycin minimum inhibitory concentration distribution and occurrence of heteroresistance
among methicillin-resistant Staphylococcus aureus blood isolates in Turkey. BMC Infect Dis.
2013;13:583.
94. Gomes D, Ward K, LaPlante K. Clinical implications of vancomycin heteroresistant and
intermediately susceptible Staphylococcus aureus. Pharmacotherapy. 2015;35:424–32.
95. Claeys K, Lagnf A, Hallesy J, Compton MT, Gravelin A, Davis S, et al. Pneumonia caused
by methicillin-resistant Staphylococcus aureus: does vancomycin heteroresistance matter?
Antimicrob Agents Chemother. 2016;60:1708–16.
96. Zhao X, Drlica K. Restricting the selection of antibiotic-resistant mutants: a general strategy
derived from fluoroquinolone studies. Clin Inf Dis. 2001;33(Suppl 3):S147–S56.
97. Rybak M, Lomaestro B, Rotschafer J, Moellering R, Craig W, Billeter M, et al. Therapeutic
monitoring of vancomycin in adults summary of consensus recommendations from the
American Society of Health-System Pharmacists, the Infectious Diseases Society of America,
and the Society of Infectious Diseases Pharmacists. Pharmacotherapy. 2009;29:1275–9.
98. Mei Q, Ye Y, Zhu Y, Cheng J, Yang HF, Liu Y, et al. Use of Monte Carlo simulation to evaluate
the development of vancomycin resistance in meticillin-resistant Staphylococcus aureus. Int
J Antimicrob Agents. 2015a;45(6):652.
99. Khatib R, Sharma M, Johnson L, Riederer K, Shemes S, Szpunar S. Decreasing prevalence of
isolates with vancomycin heteroresistance and vancomycin minimum inhibitory concentra-
tions ≥2 mg/L in methicillin-resistant Staphylococcus aureus over 11 years: potential impact
of vancomycin treatment guidelines. Diagn Microbiol Infect Dis. 2015;82:245–8.
100. Ikonomidis A, Neou E, Gogou V, Vrioni G, Tsakris A, Pournaras S. Heteroresistance
to meropenem in carbapenem-susceptible Acinetobacter baumannii. J Clin Microbiol.
2009;47:4055–9.
101. Hung K, Wang M, Huang A, Yan J, Wu J. Heteroresistance to cephalosporins and penicillins
in Acinetobacter baumannii. J Clin Microbiol. 2012;50:721–6.
102. Choi H, Kil M, Choi J, Kim S, Park K, Kim Y, et al. Characterisation of successive
Acinetobacter baumannii isolates from a deceased haemophagocytic lymphohistiocytosis
patient. Int J Antimicrob Agents. 2017;49:102–6.
103. Rodríguez C, Nastro M, Fiorilli G, Dabos L, Lopez-Calvo J, Fariña M, et al. Trends in the
resistance profiles of Acinetobacter baumannii endemic clones in a university hospital of
Argentina. J Chemother. 2016;28:25–7.
104. Álvarez-Pérez S, Blanco J, Harmanus C, Kuijper E, García M. Subtyping and antimicrobial
susceptibility of Clostridium difficile PCR ribotype 078/126 isolates of human and animal
origin. Vet Microbiol. 2017;199:15–22.
105. Tran T, Jaijakul S, Lewis C, Diaz L, Panesso D, Kaplan H, et al. Native valve endocar-
ditis caused by Corynebacterium striatum with heterogeneous high-level daptomycin
resistance: collateral damage from daptomycin therapy? Antimicrob Agents Chemother.
2012;56:3461–4.
106. Ma W, Sun J, Yang S, Zhang L. Epidemiological and clinical features for cefepime hetero-
resistant Escherichia coli infections in Southwest China. Eur J Clin Microbiol Infect Dis.
2016;35:571–8.
107. Sun J, Huang S, Yang S, Pu S, Zhang C, Zhang LZ. Impact of carbapenem heteroresistance
among clinical isolates of invasive Escherichia coli in Chongqing, southwestern China. Clin
Microbiol Infect. 2014;21:469.e1–10.
108. Cherkaoui A, Diene S, Renzoni A, Emonet S, Renzi G, François P, et al. Imipenem heterore-
sistance in nontypeable Haemophilus influenzae is linked to a combination of altered PBP3,
slow drug influx and direct efflux regulation. Clin Microbiol Infect. 2017;23:118.e9–e19.
109. Ben-Mansour K, Fendri C, Battikh H, Garnier M, Zribi M, Jlizi A, et al. Multiple and mixed
Helicobacter pylori infections: comparison of two epidemiological situations in Tunisia and
France. Infect Genet Evol. 2016;37:43–8.
296 K. Drlica et al.
Patricia A. Bradford
10.1 Introduction
P. A. Bradford (*)
Antimicrobial Development Specialists, LLC, Nyack, NY, USA
e-mail: [email protected]
Bacterial resistance to antibiotics can result via three main pathways: modification
of the bacterial target for the antibacterial, decreased intracellular concentrations
due to reduced permeability and efflux, or enzymatic inactivation of the drug. In
some cases, all members of a given species might be resistant to a particular antibi-
otic. For example, all isolates of the Gram-negative non-fermenter Stenotrophomonas
maltophilia express a chromosomally encoded metallo-β-lactamase. Therefore,
resistance to imipenem and other carbapenems is a diagnostic tool for identifying
this organism. Alternatively, resistance can develop in previously susceptible organ-
isms through genetic mutation or by acquisition of foreign DNA encoding resis-
tance genes. The specific mechanisms that affect various classes of antibiotics are
discussed in other chapters. The discussion here will focus on the selection and
spread of resistance when it occurs.
The use of antibacterial drugs disrupts the microbiome of the patient being treated.
In turn, the hospital unit or other groups of people in close proximity such as in
daycare centers or in long-term care facilities can be affected. As a consequence, an
entire species of bacterial pathogen might be selected with antibiotic pressure due
to natural resistance occurring in that species. For example, the increased role of
enterococci as opportunist pathogens in the 1980s and 1990s correlated with the
introduction and increased usage of fluoroquinolones and cephalosporins, as these
organisms are inherently resistant to those agents [2]. Similarly, the increasing inci-
dence of coagulase-negative staphylococci and α-hemolytic streptococci in hema-
tology patients, especially those who have indwelling central lines, correlated with
the increased use of fluoroquinolones in these patients [3]. Among Gram-negative
pathogens, Acinetobacter baumannii and S. maltophilia have become increasingly
prevalent in many intensive care units (ICUs) following the increased usage of car-
bapenems, especially among patients with mechanical ventilation [4, 5]. S. malto-
philia has a naturally occurring metallo-β-lactamase that renders it resistant to
carbapenems, and A. baumannii is often resistant to all antibacterials except
trimethoprim-sulfamethoxazole. The introduction of each of these new therapies
has led to the unexpected consequence of shifting the etiology of some of the com-
mon hospital-based infections to species that are naturally more resistant than the
pathogens they replaced.
10 Epidemiology of Bacterial Resistance 301
As bacteria grow, the DNA (both the chromosome and plasmids) is replicated
through a process that is highly prone to errors in base incorporation. These errors,
leading to base substitutions, occur randomly, at a frequency of approximately 10−9
per gene [6]. Even one amino acid substitution can greatly alter the functionality of
a gene. For example, the substitution of serine for glycine at residue 238 in the
SHV-1 β-lactamase led to the first extended-spectrum β-lactamase (ESBL), SHV-2,
that conferred resistance to expanded spectrum cephalosporins [7]. In addition to
these random point mutations, replication errors may lead to deletions or insertions
of small pieces of genes. Each of these mutations may result in the altered interac-
tion of antibacterial agents with the bacterium through changes of the drug target,
enzymatic inactivation of the drug inactivation, or changes in the efflux or uptake
inactivated by the importation of insertion sequences, such as the case with the ccrA
gene expressing a metallo-β-lactamase in Bacteroides fragilis that is only expressed
only if an insertion sequence has inserted upstream of this structural gene [8].
Exposure to antibiotics does not cause the mutations but rather selects for strains
that have pre-existing mutations that allow the bacterial cell to survive in the pres-
ence of the antibiotic.
Most mutations occurring in the drug target or in an antibiotic-modifying enzyme
affect only a single antibacterial class. However, mutations also occur in genes
encoding outer membrane porin proteins that allow penetration through the outer
membrane by passive diffusion, or efflux systems that expel out of the cell multiple
antibiotic classes as well as other cell toxins such as dyes can greatly impact the
susceptibility of a bacterial cell to the antibiotic [9]. The maintenance of a mutation
in a bacterial pathogen causing antibiotic resistance is completely dependent upon
whether or not that mutation affects the fitness or virulence of that organism. If
resistant mutants emerge at high frequency and are still able to replicate and cause
disease, they can gain a foothold in the bacterial population that is further selected
through continued use of the drug [10]. There have been several antimicrobials
introduced in the 1980s and 1990s that had reduced utility following mutational
resistance in certain species. Resistance to fluoroquinolones among staphylococci
rapidly emerged by the upregulation of NorA-mediated efflux [11]. Another exam-
ple was the use of imipenem that led to the selection of P. aeruginosa that have lost
the OprD porin, which provides carbapenem-specific pores through the outer mem-
brane [9]. Interestingly, the recent development of resistance to linezolid due modi-
fication of the domain V of 23S rRNA (the binding site for linezolid) in
Staphylococcus aureus and Enterococcus spp. has not led to widespread resistance
among clinical isolates [12, 13].
302 P. A. Bradford
DNA transfer among bacteria primarily occurs via plasmids, some of which are
self-transmissible, in that they carry genes to initiate the direct transfer to another
bacterium. Many plasmids are large and are able to accommodate multiple resis-
tance genes. These large transferrable plasmids are the ideal vector for the dissemi-
nation of resistance genes. Within plasmids, resistance genes are often carried by
transposons, which can transfer determinants between plasmids, or transport them
into and out of the chromosome [14]. In addition, resistant bacteria often contain
integrons that have the capability to acquire and express resistance determinants
behind a single promoter. They are widely distributed among Gram-negative bacte-
ria and are found within plasmids and transposons [14]. Very diverse resistance
determinants have been found in integrons, including genes conferring target-based
resistance to trimethoprim and fosfomycin, efflux-mediated quinolone resistance,
and metallo-β-lactamase-mediated carbapenem resistance [15–17]. Mechanisms of
transferrable resistance are presented in detail in Chap. 11.
The dissemination of plasmids, transposons, and integrons among bacterial
pathogens has resulted in “gene epidemics” [10]. The TEM-1 plasmid-mediated
β-lactamase was first described in 1965 in an Escherichia coli isolate from a patient
in Greece but has since spread globally to multiple species. It has been found in up
to 60% of clinical isolates of Enterobacteriaceae, to a few Pseudomonas aerugi-
nosa, and up to 50% of Haemophilus influenzae and Neisseria gonorrhoeae isolates
[18]. There are probably multiple factors that determine whether or not a mobilized
gene will spread widely, but these are not well understood. For example, TEM-2
β-lactamase differs from TEM-1 by only a single amino acid substitution and pro-
vides an identical spectrum of resistance. It is also found on similar kinds of plas-
mids and transposons. However, the β-lactamase TEM-2 is at least tenfold less
prevalent than TEM-1 in every region [18].
Many of the resistance determinants now found on plasmids, integrons, and
transposons are believed to have originated in the chromosomes of other bacterial
species, a phenomenon that has been well-documented in plasmid-mediated
β-lactamases. The SHV-type β-lactamases are derived from the chromosomal
β-lactamases of Klebsiella pneumoniae; plasmid-encoded AmpC enzymes expressed
in K. pneumoniae and E. coli are nearly identical to chromosomal AmpC genes
found in E. cloacae (ACT-1, MIR-1), Citrobacter freundii (CMY-type), Hafnia
alvei (ACC-1), Morganella morganii (DHA-1), and the very successful cefotaxime-
hydrolyzing CTX-M-type ESBLs from Kluyvera spp. [7, 19–21]. In addition, many
aminoglycoside-modifying enzymes found in pathogenic bacteria were determined
to have originated in environmental species of Acinetobacter [22, 23]. Many genes
that are responsible for resistance to antibiotics that are natural products have
migrated from the antibiotic-producing organisms (mostly streptomycetes), which
have developed and retained these genes in order protect themselves against their
own by-products. For example, the erm determinants that methylate 23S rRNA
block binding of macrolides, lincosamides, and group B streptogramins to the target
10 Epidemiology of Bacterial Resistance 303
Typing systems for the epidemiological study of bacterial pathogens are based on
the observation that, although different isolates of the same genus and species share
microbiological, biochemical, serological, and physiological characteristics that
distinguish them from other species, they also have detectable genetic differences
that make discrimination at the intraspecies level possible [26]. In many circum-
stances, intraspecies variability is very high among unrelated isolates, and therefore,
it is easily detectable. However, when dealing with human disease, several species
of bacterial pathogens share overlapping niches and are subjected to identical envi-
ronmental selective pressures. Molecular genetic studies of bacterial populations
have demonstrated that there is some degree of homogeneity between pathogenic
and environmental strains and making genetic differentiation relatively more com-
plicated [27]. Consequently, one must understand that different typing methods give
different, sometimes somewhat contradictory information that should be viewed as
a totality of information for an examination of the phylogenetic and epidemiologi-
cal relationships between pathogens. Molecular typing methods that utilize the
genetic structure of bacterial pathogens have been used to address many different
problems such as the study of genomic organization and evolution. In the context of
bacterial resistance, they are now being used for the identification of patterns of
infection and sources of transmission, the epidemiological surveillance of infec-
tious diseases, and outbreak investigations [28].
10.3.1.1 Antibiogram
the detection of acquired resistance determinants may not coincide with therapeutic
breakpoints used in the clinical microbiology laboratory. In addition, minimal inhib-
itory concentration (MIC) values are more informative than qualitative resistance
patterns. However, discrimination is dependent on the diversity and relative preva-
lence of detectable resistance in the isolates in question. One drawback to using the
antibiogram for epidemiology is that the stability of resistance pattern can be insuf-
ficient for use as a clonal marker, because resistance determinants may be encoded
on plasmids or resistance genes may be expressed under control of complex regula-
tory systems [29–31]. The antibiogram is often the most valuable first-line typing
methods in clinical laboratories that can quickly be used to assess the prevalence of
resistance or the appearance of an outbreak strain. However, the integrity of data
used to generate the antibiogram is crucial and is dependent upon the methods used
for determining susceptibility. Many automated systems use short dilution ranges
that surround the breakpoint for a given drug and may not provide enough informa-
tion to discriminate between strains [32]. Nevertheless, the generation of an antibio-
gram has the advantage of being technically easy to use and interpret, even in small
and resource-limited laboratories. It is relatively low-cost test suitable for testing
large numbers of isolates and relies on routine clinical practice. Good reproducibil-
ity allows its use for definitive typing if a standard method such as MIC or disk
diffusion as well as a standard set of marker antibiotics are utilized [28].
10.3.1.2 Serotyping
Different high resolution molecular-based procedures have been used to detect the
unique features of each individual organism. As a result, guidelines and some inter-
pretive criteria have been proposed in an attempt to standardize what constitutes the
“same strain” [27, 39].
10 Epidemiology of Bacterial Resistance 305
The profile of the number and size of bacterial plasmids, some of which carry anti-
microbial resistance determinants, can be used to determine the relatedness of
strains during an epidemiologic investigation especially when combined with the
utilization of restriction endonucleases to generate a restriction fragment length
polymorphism (RFLP) analysis [40, 41]. However, plasmids that can be transferred
even to other strains, including those of different bacterial species, are often unsta-
ble and may be lost or new ones acquired spontaneously. This makes plasmid fin-
gerprinting somewhat difficult to reproduce [26]. Because of this variability in
plasmid content, the use of plasmid profiling has been found to be insufficient for
use as a clonal marker in some studies [29, 31, 41]. It is best combined with other
genomic typing methods (at the chromosome level) to distinguish between spread
of a clone and that of a plasmid [28].
10.3.2.2 Ribotyping
Several amplification techniques using PCR have been proposed as bacterial typing
systems. The various PCR-based fingerprinting methods may involve the entire
genome by the use of either arbitrary primers or primer pairs directed at the short
sequences lying between repeat motifs in the bacterial genome [28, 51]. They are
universal typing methods that can be applied to most bacterial species and exhibit a
high level of discrimination between strains [51]. Major advantages of these tech-
niques include flexibility, technical simplicity, wide availability of equipment and
reagents, and same-day results [28].
RAPD/AP-PCR
Rep-PCR
VNTR/MLVA Analysis
The intergenic regions of many bacterial genomes contain repetitive DNA sequences
that are highly variable with regard to the number or structure of these repeat units.
These variable-number tandem repeats (VNTR) allow for the differentiation
between strains using a PCR-based method [62]. When multiple regions of these
repeat sequences are analyzed, VNTR analysis is sometimes also referred to as mul-
tiple locus variable analysis (MLVA). The speed at which these intergenic regions
change and evolve is also highly variable, making the various loci important for
determining the evolutionary clock for the strains. This adds to the discriminatory
power of this tool [33]. MLVA was famously applied during the 2001 bioterrorism-
associated anthrax outbreak in the USA [63]. Bacillus anthracis is notoriously dif-
ficult to type using most molecular strain typing methods because there is very little
variability among strains. We used MLVA to subtype 135 isolates from infected
patients, powders from mail sources, and environmental samples. Subtyping of B.
anthracis allowed anthrax cases to be linked to environmental specimens and pow-
ders and provided information about potential sources. MLVA was able to not only
determine that all of the samples involved in the terror attack were identical to each
other but that they were distinguishable from other Ames strains that had been iso-
lated from other sources [63]. VNTR/MLVA has been used for multiple other appli-
cations. VNTR analysis is considered to be the gold standard for molecular typing
of Mycobacterium tuberculosis [64]. It was also shown to be superior to Rep-PCR
for strain discrimination when typing Mycobacterium bovis [65]. In addition, VNTR
analysis was used to characterize an outbreak of linezolid-resistant isolates of
enterococci in Polish hospitals [66].
PFGE analysis (also known as macrorestriction) can simplify the fingerprint of bac-
terial pathogens by utilizing restriction endonucleases that infrequently cut the
chromosomal DNA into a relatively small number of large fragments (20 kb to
>1 Mb) [33]. PFGE uses a current that pulses in alternating directions to separate
and resolve significantly larger fragments of DNA than is possible using constant
field gel electrophoresis. In addition, shearing of these large fragments is avoided by
stabilizing the DNA by embedding samples into an agarose plug [67]. PFGE can be
applied to isolates of most species, although the restriction enzymes used may be
specific to a particular organism [29]. The profiles generated by PFGE are highly
reproducible [26]. One limitation for using PFGE to track resistance is that it is very
labor intensive and can take up to 4 days to complete. Therefore, it cannot be used
in a rapid response to an outbreak.
Interpretive criteria for chromosomal DNA macrorestriction patterns produced
by PFGE have been proposed and guidelines applied successfully for different bac-
terial organisms [39]. Strains are considered to be unrelated if there are three or
more bands (number and/or size of fragments) differing between two isolates. This
308 P. A. Bradford
standardization has allowed for not only the use of PFGE for comparing isolates at
a local level but also across many sites. The Centers for Disease Control (CDC)
developed PulseNet in 1996, which is a national laboratory network that uses PFGE
to detect thousands of local and multistate outbreaks of foodborne illnesses https://
www.cdc.gov/pulsenet/index.html [68]. PFGE has been applied in many different
scenarios for tracking resistance, including carbapenem-resistant K. pneumoniae in
China [69], following NDM-1 among A. baumannii in Israel [70], and determining
nosocomial transmission of MRSA in Malaysia [71].
In recent years, next generation sequencing technology has become an easy and
cost-effective method for performing molecular epidemiology by sequencing the
entire genome of pathogens of interest. The strain relatedness of VRE isolated dur-
ing three outbreaks in a hospital in Sweden was investigated to determine how WGS
would compare to PFGE and MLST. The whole genome sequencing (WGS) data
was analyzed using the average nucleotide identity analysis. PFGE analysis of the
isolates confirmed what was already known by the clinical epidemiological investi-
gation: that three outbreaks had occurred. However, there was no indication of fur-
ther strain relatedness, or if there was a larger cluster. In contrast, the WGS analysis
could clearly distinguish six clusters among the isolates [74]. WGS was also used to
investigate a prolonged outbreak of KPC-producing K. pneumoniae and E. cloacae
in a burn unit in the USA. WGS revealed that this outbreak, which seemed epide-
miologically unrelated, was in fact genetically linked. The outbreak was primarily
maintained by a clonal expansion of E. cloacae sequence ST114 that contained
multiple resistance determinants including blaKPC-3 that was transmitted via plas-
mids containing an identical Tn4401b [83].
Looking at the genome for any differences between strains can be overwhelming
with the amount of data that this generates. Therefore, it is often more useful to
focus on a subset of conserved genes in the bacterial species. Using this approach,
carbapenemase-producing K. pneumoniae isolates from two distinct outbreaks that
occurred in Switzerland in 2013 and 2015 were analyzed. The analysis correctly
identified the two clusters of strains from the two outbreaks and differentiated these
from K. pneumoniae that were unrelated to the outbreak [84]. Many of the previ-
ously described typing methods that utilize PCR and sequencing to detect differ-
ences in strains can now be done by WGS. The PulseNet International network
conducts global laboratory-based surveillance for foodborne illnesses. PulseNet
relies on MLST typing to track outbreaks of many pathogens. Previously, the MLST
was done by PCR and sequencing; however, they have now transitioned into the
standardized use of WGS to perform this subtyping [73, 85]. WGS was recently
used to track an outbreak of carbapenem-resistant K. pneumoniae expressing OXA-
232 to a contaminated duodenoscope in a California hospital [86].
same hospital. In the USA, a study found that the incidence of MRSA, VRE,
ceftazidime-resistant E. cloacae and P. aeruginosa, and imipenem-resistant P. aeru-
ginosa was two times higher in ICU patients than in patients in general wards or
outpatients at the same hospitals [87]. Likewise, in Europe, the prevalence of MRSA
was noted to be higher in ICUs than in the general patient population [88]. Most
outbreaks and cluster cases of resistant pathogens involve a few patients in a single
unit. The prevalence of resistance is often highest in units where the most debili-
tated patients are located. These units are also often where antibacterial usage is the
greatest, resulting in a constant selective pressure for resistant strains [10].
The epidemiology of resistance can also vary greatly depending on region. In
North America, resistance rates are generally higher in the USA than in Canada.
The prevalence of MRSA among hospitalized adults in Canada was 22–28% but
was 42–45% in the USA [89, 90]. In Europe, an extreme variation in the prevalence
of resistance between countries is noted in that there is a very low incidence in the
Scandinavian countries and very high percentages in the Mediterranean countries.
The incidence of MRSA among S. aureus ranges from <1% in Sweden to 5–9% in
Austria;10–28% in the UK and France; 25–49% in Portugal, Italy, and Greece; and
a high of 50–>75% in Ukraine [91]. Very high prevalence of resistance has been
noted in Asia and Latin America. Prevalence of penicillin-resistant S. pneumoniae
(PRSP) in pediatric patients was 91.3% in Taiwan, 85.8% in Korea, and 70.4% in
Vietnam, compared to <1%–5% in the UK and Scandinavian countries [91, 92].
Differences in the prevalence of specific pathogens are likely due to a combination
of factors including variations in medical care and antibiotic prescribing habits.
Strains with similar resistance phenotypes to organisms causing outbreaks and
epidemics have been found in the same hospitals or patient groups; however, they
fail to spread [10]. The reason for the success of a particular resistant clone is not
well understood. However, some potential factors may include (i) increased
adherence or virulence mechanisms; (ii) retention of the fitness of the bacterium in
the presence of the resistance genes; (iii) greater tolerance of desiccation, thus
remaining longer in the environment outside of the human body; and (iv) resistance
to disinfectants. A few of these outbreak strains have been studied in detail with
regard to virulence factors contributing to their proliferation. In one study, one of
the major subunits of a new fimbrial protein KPF-28 that aids adherence to the gut
mucosa was found to be co-localized on a plasmid that also encoded the ESBL
SHV-4 in serotype K25 K. pneumoniae strains that were circulating in France and
Belgium in the 1990s [93].
Resistant strains have now disseminated globally. In part, this is due to increased
international travel and open immigration. Several examples of the spread of resis-
tant strains are discussed in the sections below.
10 Epidemiology of Bacterial Resistance 311
10.5.1 MRSA
Shortly after methicillin (then called Celbenin) was introduced into clinical use in
the 1950s in the UK, there were reports of clinical isolates of S. aureus that were
resistant to methicillin [94]. S. aureus strains became methicillin-resistant by the
acquisition and expression of PBP 2a, a low-affinity PBP not native to S. aureus and
encoded by mecA [95]. The mecA gene confers resistance to all β-lactam antibiotics
except for the anti-MRSA cephalosporins ceftaroline and ceftobiprole that can bind
to PBP 2a [96]. The mecA gene is located within a mobile element in the S. aureus
chromosome known as the staphylococcal chromosomal cassette (SCCmec) region.
Numerous different mec regions (SCCmec I to XI) have been described [97–99].
However, transfer of the mec region between staphylococcal strains has never been
documented.
Humans are a natural reservoir for S. aureus, including MRSA, with asymptom-
atic colonization of the nasopharynx, perineum, or skin being commonplace. This
often occurs shortly after birth and may be transient throughout one’s lifetime [100].
Transmission usually occurs by direct contact to a colonized carrier. In adults, the
incidence of MRSA carriage is 25–50% for the general public. A higher incidence
is observed in injection drug users, persons with insulin-dependent diabetes, patients
with dermatologic conditions, and patients with long-term indwelling intravascular
catheters. It is thought that healthcare workers may be an important reservoir for
MRSA [101]. Young children tend to have higher colonization rates, most likely due
to frequent contact with bodily secretions from other children [102]. Epidemic
strains have spread through entire hospitals and even cities. The stability of the
mecA genetic environments differ between various clones of S. aureus clones,
suggesting a potential explanation for the limited lineages within which the resis-
tance determinant has been found [103]. In 2011, MRSA of both human and animal
origin isolates from Europe which were found to have a divergent mecA homologue
termed mecC were first reported [104]. Isolates harboring mecC can be detected by
susceptibility testing but not by commercial assays targeting mecA or PBP 2a.
Multilocus enzyme electrophoresis and other molecular population genetic tech-
niques were used to determine the extent of mec distribution among phylogenetic
lineages of the species and genetic relationships among MRSA strains isolated from
1961 to 1992 from various geographic regions [105]. This study found that the mec
gene was harbored by many divergent phylogenetic lineages representing wide
chromosomal diversity in the S. aureus. The conclusions were that multiple epi-
sodes of horizontal transfer and recombination contributed to the spread of this
resistance determinant in natural populations of bacteria [105]. However, there can
be regional difference in the spread. This study also identified a single multilocus
enzyme genotype among MRSA isolates recovered in the UK, Denmark,
Switzerland, Egypt, and Uganda, from the 1960s, which indicated that MRSA iso-
lates recovered from those countries at that time were progeny of a single ancestral
cell that had probably recently acquired the mec determinant.
312 P. A. Bradford
Within a few years after that first occurrence, hospital outbreaks caused by
MRSA occurred in Europe. By the mid-1970s, MRSA were recognized as an impor-
tant hospital infection control problem in the USA, and subsequently, these organ-
isms have now achieved global distribution [27]. In a study of bacteremic patients
infected with S. aureus in England and Wales, the incidence of MRSA remained low
(1–3%) for the first years of the study (1989–1993), but then the percentage of
MRSA rapidly increased, reaching 42% by the year 2000 [106]. This increase coin-
cided with the emergence of two new epidemic strains, (designated EMRSA) 15
and 16, which now comprise 95% of all S. aureus bacteremias in England and Wales
[107]. Presently, many industrialized countries report that MRSA comprises at least
25–50% of S. aureus isolated in hospitals from infected patients [89]. In Japan, it
was noted that high antibiotic consumption rates lead to increased MRSA burden
over time [108]. In contrast, some countries such as the Netherlands and the
Scandinavian countries have a low incidence of MRSA infections (around 1%)
[109]. This is likely due to rigid infection control and surveillance policies, as well
as restraint in antibiotic prescription in those countries.
At first, when cases of MRSA infection were identified in the community setting,
an investigation usually exposed a history of recent hospitalization; close contact
with a person who has been hospitalized; or previous antimicrobial-drug therapy.
However, there were some notable exceptions. During 1980–1981, there was an
outbreak of MRSA infections in Detroit among people with no history of hospital-
ization [110]. The majority of these patients were found to be injection drug users.
The source of the Detroit outbreak was never identified, but it is assumed that fre-
quent needle sharing was the mode of transmission in the community setting. The
unexpected deaths of four Native American children with no known risk factors for
acquiring MRSA in the late 1990s launched a new worry of community-associated
MRSA (CA-MRSA) [111]. This more recent development in the spread of MRSA
represents the migration of MRSA being completely healthcare associated
(HA-MRSA) to now often finding CA-MRSA. These strains of CA-MRSA tend to
be microbiologically distinct from HA-MRSA, typically more susceptible to com-
monly used antibiotics than are HA-MRSA. CA-MRSA is thought to have devel-
oped as an independent acquisition of mecA by a strain of methicillin-susceptible S.
aureus (MSSA) [100]. A common clone of CA-MRSA belonging to pulsed-field gel
electrophoresis (PFGE) type USA300 was originally widespread only in the USA;
however, it has now become widespread throughout North America, Latin America,
and the Caribbean [112–114]. CA-MRSA is discussed in detail in Chap. 3.
S. pneumoniae is a major cause of illness and death worldwide, causing otitis media,
acute sinusitis, community-acquired pneumonia, and meningitis [115]. Since the
introduction of penicillin in the mid-1940s, the treatment of pneumococcal infec-
tions has primarily been with penicillin and other β-lactam antibiotics. Only 20 years
10 Epidemiology of Bacterial Resistance 313
later, strains with decreased susceptibility to β-lactams were detected for the first
time in 1967 [116]. PRSP were found periodically during the late 1970s but then
rapidly emerged and disseminated soon after. In the USA, approximately 5% of
pneumococcal isolates recovered during 1979 to 1987 were reported as testing with
penicillin MICs of ≥1 μg/ml [117]. During the years 1993 to 1994, the percentage
of nonsusceptible isolates was 14.1% but had increased to 25% by 1999 [118].
Antibiotic resistance in S. pneumoniae is now a global public health problem,
although the incidence of PRSP differs by geographic region. 60–89% of S. pneu-
moniae are resistant to penicillin in parts of Latin America and Asia [27].
Penicillin resistance in S. pneumoniae is due to the expression of altered high-
molecular-weight penicillin-binding proteins (PBPs) that have reduced affinity for
binding and subsequent inhibition by β-lactam antibiotics [119]. It has been shown
that alterations in the five high molecular-weight PBPs 1A, 2X, 2A, and 2B corre-
lated with resistant patient isolates [25]. Acquisition of penicillin-binding protein
gene segments from foreign donors, such as oral streptococci, thereby creating
mosaic genes, is the primary mechanism for acquired resistance [120]. Following a
recombination event in a single bacterium, drug-resistant progeny can proliferate
with antibiotic pressure and subsequently can be transmitted both locally and glob-
ally by person-to-person spread [121]. The mechanisms of resistance in S. pneu-
moniae are discussed in detail in Chap. 2.
A good illustration of the rapid local clonal spread of antibiotic-resistant S. pneu-
moniae can be found by looking at the history of the development of resistance in
Iceland. During a program to monitor of antibiotic resistance in S. pneumoniae in
Iceland, no PRSP were isolated in years 1983 through most of 1988. However, in
December of 1988, the first penicillin-resistant strain was isolated. The percentage
of PRSP among S. pneumoniae then rose sharply 2.3% in 1989 to 17% by the first
part of 1992 [122]. Approximately 70% of the isolates of PRSP were also resistant
to multiple other antibiotics. Some of these isolates were characterized for related-
ness, including serotype, PBP pattern, PFGE, and multilocus enzyme electrophore-
sis typing [123]. All of the isolates were found to be serotype 6B and had similar or
identical patterns for each of the molecular markers examined. Interestingly, the
PRSP isolated in Iceland were indistinguishable from a subgroup of serotype 6B
PRSP that was frequently isolated in Spain. It was thought that the Spanish clone
was imported to Iceland, as Spain was a popular vacation spot for Icelandic families
with young children [124]. Continuing on with the study, it was noted that in 1995
PRSP comprised 24.2% of all pneumococci. There was then an interesting decline
of the incidence of PRSP to 13.6% in 2001, followed by another rapid increase to
38.6% in 2010. The study found that at some point after 2001, 19F replaced type 6B
in frequency, and by the end of the study, it accounted for 85.8% of all serogrouped
PRSP [124]. The factors responsible for the spread of these clones of PRSP in
Iceland remain unknown. However, it was noted that there was a high amount of
trimethoprim-sulfamethoxazole and tetracycline usage in Iceland compared to other
Nordic countries, which may have contributed to selecting the resistant clones. In
addition, a high proportion of Icelandic children attend day care centers, which may
have contributed to the rapid spread of PRSP in that country [27].
314 P. A. Bradford
The incidence and spread of PRSP took a dramatic downturn following the intro-
duction of the pneumococcal seven-valent conjugate vaccine (PCV7) in the year
2000 that included serotypes 4, 6B, 9V, 14, 18C, 19F, and 23F, which are the most
common serotypes found in invasive pneumococcal disease. As shown in Fig. 10.1,
the incidence of PRSP among pediatric cases of invasive pneumococcal disease was
21.9 per 100,000 cases in 1999 [125]. Approximately 80% of PRSP isolated in the
pre-PCV7 era in the USA were of serotypes commonly carried by healthy children
(serotypes 6B, 9V, 14, 19F, and 23F), and serotype 19A was rare [118]. Following
the introduction of the vaccine, the incidence dropped to 2.3 in just 10 years, fol-
lowed by a further reduction to <1 PRSP in 2013. However, there was also a noted
difference in the serotypes now responsible for β-lactam resistance in S. pneu-
moniae. Following the widespread implementation of the PCV7 vaccine, the highly
resistant serotype 19A became the predominant serotype among PRSP. Fortunately,
a 13 valent vaccine was introduced in 2010 that included serotype 19A; therefore,
the further dissemination of PRSP was limited [126].
Fig. 10.1 Approximate numbers of cases per 100,000 individuals caused by penicillin resistance for each year (left). These data encompass all clonal com-
315
plexes and serotypes associated with penicillin resistance (MIC of 2 μg/ml) in cases of pediatric invasive pneumococcal disease during 1999, 2009, and 2013.
The circle diameters reflect relative incidence of disease [125]
316 P. A. Bradford
10.5.4 KPC
first-line therapeutic options [153, 154]. The successful spread of KPC has been
primarily due to the spread of K. pneumoniae isolates belonging to the successful
clonal complex ST258 [153, 154]. For isolates belonging to ST258, blaKPC is most
often found inside of the Tn4401 transposon. Furthermore, the promiscuous plas-
mids harboring blaKPC also commonly carry genes encoding resistance to aminogly-
cosides and additional β-lactamases, including ESBLs and metallo-β-lactamases
[144, 145]. KPC has also been found in the same strain as metallo-β-lactamases.
blaKPC has also been identified the bacterial chromosome [155].
10.5.5 NDM
approximately 22.8% during 2012–2015 [140]. However, there was a wide varia-
tion in the percent resistance among different European countries. The fluoroqui-
nolone resistance for Europe ranged from 2.9% in Iceland up to 70% in Slovakia.
A recent survey of 10,000 E. coli isolates from UTI in US hospitals showed an
overall incidence of resistance to fluoroquinolones of 34.5% [166]. The epidemic
CTX-M-15 expressing clone E. coli ST131 is often also resistant to fluoroquino-
lones, and this is thought to have played a significant role in the spread and main-
tenance of this strain [79]. The potential for the spread of fluoroquinolone
resistance has increased with the development of plasmid-meditated fluoroquino-
lone resistance genes such as qnr [167]. In several longitudinal surveys, the inci-
dence of qnr among K. pneumoniae in Taiwan increased from zero in 2000 to
7.6% in 2005 [168]. Similarly, among clinical isolates of Enterobacter spp. iso-
lates from Israel, there were none positive for qnr during 1990–1993, but 6.8%
were positive in 2005 [169].
Until the recent years, the incidence of resistance to colistin was very low. However,
because of the rise of carbapenem resistance among Enterobacteriaceae, there has
been an increased usage of the drugs of polymyxin class, chiefly colistin [177].
Subsequently, reports of colistin resistance are being seen with increasing fre-
quency. A recent study of colistin resistance among a global collection of clinical
isolates of Enterobacteriaceae from 2012 to 2013 showed only 1.6% in the overall
population; 12% in carbapenemase positive isolates were resistant [178]. Colistin
susceptibility was higher among MBL-positive isolates (92.6%) than those positive
for a KPC (87.9%) or OXA-48 (84.2%). In this study, approximately 2.4% of all K.
pneumoniae isolates were colistin resistant compared to only 0.3% of E. coli.
Interestingly, the prevalence of colistin resistance was relatively high among
Enterobacter spp. but varied from 39% in E. asburiae to 0.4% in E. aerogenes. This
study also noted regional differences in the prevalence of colistin resistance. For K.
pneumoniae, the prevalence ranged from 5% in Greece, 4.7% in Italy, and 3.2% in
Romania to 1.3% in North America [178]. Other studies have also shown K. pneu-
moniae resistance rates ranging from 10.5–20% in resistance hotspots such as
Greece to 2.9% in Canada [179].
Previously, the spread of colistin resistance was limited to the expansion of resis-
tant clones because the development of resistance depended upon mutations in the
bacterial chromosome. However, the recent discovery of colistin resistance due to
the plasmid borne mcr-1 may have a very great impact in the coming years. The
10 Epidemiology of Bacterial Resistance 321
mcr-1 gene was first described in a porcine isolate of E. coli in China [180]. Since
that time, a multitude of reports from all over the globe have emerged. It has been
found in patients with Salmonella spp. in China, E. coli in Belgium and Oman, and
K. pneumoniae in South Africa [181–184]. The first detection of the mcr-1 gene in
E. coli from a patient in the USA was reported in 2016 [185]. In a survey of 390
colistin-resistant isolates collected in 2014–2015, none of the K. pneumoniae iso-
lates harbored mcr-1 (0 of 331 isolates), whereas 19 of 59 E. coli isolates had mcr-1
[186]. A study of fecal carriage of mcr-1 in the Netherlands showed detection of the
gene, although among very few patients (0.35%), but it was detected nonetheless
[187]. There continues to be a high correlation of animal to human spread of mcr-1.
This gene was found among patient and animal samples of Salmonella spp. in China
[188]. Similarly, a recent study detected mcr-1 in E. coli and Salmonella spp. iso-
lates of animal origin in Europe, in every year from 2004 to 2014 [189]. It was also
found among E. coli from a pig farm in Germany [190]. There have also been
reports of bacteria containing mcr-1 in sewage and wastewater in China, Germany,
and Spain [191–193]. It is difficult to know if mcr-1 is really spreading or if it is just
an increased awareness of this resistance determinant which is causing its increased
detection.
Daycare centers that look after infants and preschool-aged children have been
shown to harbor and spread resistant isolates. This is likely due to the fact that these
children frequently share bodily secretions and receive multiple courses of antibiot-
ics in their first few years of life. The high incidence of PRSP in Iceland (discussed
in Sect. 10.5.2 above) was partially attributed to the large number of children that
attend daycare in that country [27]. A study to document the increased carriage of
PRSP by children in daycare centers was undertaken in Japan. Researchers sampled
the nasopharyngeal passages of children from newborns to 6 years attending two
daycare centers [195]. From 363 children cultured, they found that the overall
322 P. A. Bradford
In the early 1990s, long-term care facilities (nursing homes), mostly comprised of
elderly patients, became recognized as an important reservoir for antibiotic-resistant
pathogens during one of the earliest outbreaks of ESBL producing (TEM-10) that
occurred in Chicago [198]. The source of the outbreak was tracked to an index
patient that had been admitted to the hospital from a long-term care facility.
Subsequently, the authors conducted a study to determine the extent of prevalence
of ESBL-producing strains among these patients [40]. With that goal, they identi-
fied 55 hospitalized patients who were colonized or infected with ESBL-producing
E. coli or K. pneumoniae, during an 18-month period in 1990–1992. Of these 31
patients that were admitted from 8 different nursing homes were positive for TEM-
10 expressing pathogens, all of which were resistant to ceftazidime, gentamycin,
and tobramycin. As a case-control study, 24 nursing home patients colonized with
resistant strains on hospital admission were compared with 16 patients admitted
from nursing homes that were not colonized on hospital admission. A strong cor-
relation ESBL carriage aligned with poor cognitive function, presence of a gastros-
tomy tube or decubitus ulcers, and prior treatment with antibiotics [40]. At the same
time across the country, another outbreak of ESBL-producing E. coli (SHV-7) was
described among elderly patients from nursing homes admitted to a New York hos-
pital [199]. More recently, the prevalence of ESBL-producing Enterobacteriaceae
10 Epidemiology of Bacterial Resistance 323
was studied among nursing home residents in Germany [200]. Using rectal swabs to
survey, they found that 14.7% of the residents were colonized with ESBL-producing
E. coli. All of the isolates expressed a variant of blaCTX-M, with the most common
isolated being CTX-M-15 in 65.2% and CTX-M-27 in 21.7%. Moreover, 69.6%
could be assigned by polymerase chain reaction (PCR)-typing to the epidemic
clonal lineage E. coli ST131 [200].
Other studies have examined the prevalence of multiple types of resistant patho-
gens among residents of long-term care facilities. Trick et al. conducted a study to
determine the frequency of and risk factors for colonization of nursing home resi-
dents by MRSA, vancomycin-resistant Enterococcus (VRE), and ESBL-producing
K. pneumoniae or E. coli [201]. Of 117 residents that were sampled, 43% were
culture positive in at least one antimicrobial-resistant pathogen as follows: MRSA
24%, VRE 3.5%, ESBL-producing K. pneumoniae 18%, and ESBL-producing E.
coli 15%. At the time of the study, only three of the residents were under contact
isolation precautions. Risk factors for colonization included a total dependence on
healthcare workers for activities of daily living and prior therapy with antibiotics.
Another study examined the molecular epidemiology and antimicrobial suscepti-
bilities of C. difficile strains in long-term care facilities in Israel. Toxigenic C. dif-
ficile isolates were recovered 23.6% of the nursing home patients that were sampled.
Many of these were of the predominant ribotype 027, which had increased MICs of
vancomycin and metronidazole [47]. The authors recommend increased contact iso-
lation and other infection control measures be implemented by the facilities [201].
Fig. 10.5 Epidemic curve graph (top) and field position diagram (bottom) of cases of MRSA
infection among St. Louis Rams professional football players in 2003. Each box on the epidemic
curve graph and field diagram represents an MRSA infection; different colors designate different
players; boxes of the same color thus represent recurrent infections. On the field diagram, X repre-
sents a defensive player position and O an offensive player position [211]
With the ease of international travel in today’s world, there has been an increase in
documented transplantation of previously regional resistant strains into a naïve pop-
ulation of people. This mainly occurs when people return home from abroad, either
infected or colonized with a resistant pathogen, and then transmit it to another indi-
vidual. Recently, there have also been cases of resistant bacterial pathogens being
imported to a destination country by refugees or those seeking asylum from their
country of origin. A very well-documented example of the importation of a resis-
tance mechanism was NDM-1 arriving in the UK in an individual who had recently
returned from India (discussed above in Sect. 10.5.9) [158]. A study in Finland
revealed that 46% of travelers returning from South Asia were colonized with
ESBL-producing Enterobacteriaceae [212].
It has been shown that popular travel destinations can be a localized reservoir for
resistant pathogens. In a systematic review of the literature from 2002–2017,
Leangapichart et al. looked at reports of antibiotic resistance among pilgrims and
326 P. A. Bradford
workers attending the Hajj in Mecca, Saudi Arabia [213]. MRSA carriage was
reported in 20% of pilgrims and food handlers. Carbapenem resistance was detected
in less than 10% of E. coli isolates, but up to 100% of K. pneumoniae and A. bau-
mannii isolates. Colistin-resistant Salmonella spp., E. coli, and K. pneumoniae,
including mcr-1-mediated colistin resistance, were detected among the pilgrims
[213]. A cluster of genetically related, azithromycin-resistant N. gonorrhoeae was
detected in Oahu, Hawaii, in a 1-month period in 2016. The majority of the isolates
were also resistant to ceftriaxone, which eliminated both of the currently recom-
mended therapies for treatment from consideration [214].
Antibiotic resistance has also been noted in cases of traveler’s diarrhea, some of
which caused larger outbreaks in the home country after return. Kim et al. studied
an outbreak of intestinal illness caused by Shigella sonnei in a daycare center in
South Korea. The outbreak strain was resistant to extended-spectrum cephalospo-
rins (via CTX-M-15) and fluoroquinolones. The index case was a child who had
recently returned from traveling to Vietnam [215]. Another study in Belgium exam-
ined antibiotic resistance patterns among isolates of Campylobacter spp. obtained
from returned international travelers that were symptomatic for traveler’s diarrhea
and analyzed the data based on travel destination [216]. For the group as a whole,
60.9% of the isolates of Campylobacter spp. were resistant to ciprofloxacin, with
the prevalence ranging from 50.8% in Africa to 75.0% in Asia. Resistance to
erythromycin was 4.6%, with the highest incidence (15.2%) seen in isolates from
individuals who had traveled to Southern Asia (six of seven patients returning from
India) and 48.3% were resistant to tetracycline [216]. A recent study examined 453
cases of enteric fever in London caused by Salmonella enterica subspecies Typhi
and Paratyphi. For patients with a history of travel, 88% of S. Paratyphi A isolates
were resistant to ciprofloxacin. For isolate of S. Typhi, 80% were resistant to cipro-
floxacin, 26% to ampicillin, and 27% to chloramphenicol [217].
Migration due to refugees and asylum seekers has been noted to be one of the
risk factors for the spread of multidrug-resistant organisms, which can challenge
local healthcare systems. Ravensbergen et al. analyzed cultures performed in Dutch
hospital on asylum seekers during 2014–2015 and compared the results to those
obtained from Dutch nationals. Of 118 asylum seekers with S. aureus in clinical
cultures, almost 19% were MRSA positive compared to only 1.3% in the general
public. In addition 20.3% were infected with Enterobacteriaceae that produced
ESBLs [109]. Similarly, 20% E. coli isolated from refugees admitted to a German
hospital were found to be ESBL producers [218].
10.7 Conclusions
References
1. CDC. Antibiotic resistance threats in the United States, 2013. 2014. Available from: https://
www.cdc.gov/drugresistance/threat-report-2013/index.html.
2. Woodford N. Glycopeptide-resistant enterococci: a decade of experience. J Med Microbiol.
1998;47(10):849–62.
3. Oppenheim BA. The changing pattern of infection in neutropenic patients. J Antimicrob
Chemother. 1998;41 Suppl D:7–11.
4. Gomez J, Simarro E, Banos V, Requena L, Ruiz J, Garcia F, et al. Six-year prospective study
of risk and prognostic factors in patients with nosocomial sepsis caused by Acinetobacter
baumannii. Eur J Clin Microbiol Infect Dis. 1999;18(5):358–61.
5. Elting LS, Khardori N, Bodey GP, Fainstein V. Nosocomial infection caused by Xanthomonas
maltophilia: a case-control study of predisposing factors. Infect Control Hosp Epidemiol.
1990;11(3):134–8.
6. Miller JH. Mutational specificity in Bacteria. Annu Rev Genet. 1983;17(1):215–38.
7. Bradford PA. Extended-spectrum β-lactamases in the 21st century: characterization,
epidemiology, and detection of this important resistance threat. Clin Microbiol Rev.
2001;14(4):933–51.
8. Podglajen I, Breuil J, Collatz E. Insertion of a novel DNA sequence, 1S1186, upstream of the
silent carbapenemase gene cfiA, promotes expression of carbapenem resistance in clinical
isolates of Bacteroides fragilis. Mol Microbiol. 1994;12(1):105–14.
9. Poole K. Outer membranes and efflux: the path to multidrug resistance in Gram-negative
bacteria. Curr Pharm Biotechnol. 2002;3(2):77–98.
10. Livermore DM. Bacterial resistance: origins, epidemiology, and impact. Clin Infect Dis.
2003;36(Suppl 1):S11–23.
11. Piddock LJ. Mechanisms of fluoroquinolone resistance: an update 1994-1998. Drugs.
1999;58(Suppl 2):11–8.
12. Prystowsky J, Siddiqui F, Chosay J, Shinabarger DL, Millichap J, Peterson LR, et al. Resistance
to linezolid: characterization of mutations in rRNA and comparison of their occurrences in
vancomycin-resistant enterococci. Antimicrob Agents Chemother. 2001;45(7):2154–6.
328 P. A. Bradford
33. Gerner-Smidt P, Hyytiä-Trees E, Rota PA. Molecular epidemiology. In: Versalovic J, editor.
Manual of clinical microbiology, vol. 1. 10th ed. Washington, D.C.: American Society for
Microbiology Press; 2011.
34. Yoshida C, Franklin K, Konczy P, McQuiston JR, Fields PI, Nash JH, et al. Methodologies
towards the development of an oligonucleotide microarray for determination of Salmonella
serotypes. J Microbiol Methods. 2007;70(2):261–71.
35. Fitzgerald C, Collins M, van Duyne S, Mikoleit M, Brown T, Fields P. Multiplex, bead-based
suspension array for molecular determination of common Salmonella serogroups. J Clin
Microbiol. 2007;45(10):3323–34.
36. Neal S, Beall B, Ekelund K, Henriques-Normark B, Jasir A, Johnson D, et al. International
quality assurance study for characterization of Streptococcus pyogenes. J Clin Microbiol.
2007;45(4):1175–9.
37. Facklam RF, Martin DR, Lovgren M, Johnson DR, Efstratiou A, Thompson TA, et al.
Extension of the Lancefield classification for group A streptococci by addition of 22 new
M protein gene sequence types from clinical isolates: emm103 to emm124. Clin Infect Dis.
2002;34(1):28–38.
38. Pai R, Gertz RE, Beall B. Sequential multiplex PCR approach for determining capsular sero-
types of Streptococcus pneumoniae isolates. J Clin Microbiol. 2006;44(1):124–31.
39. Tenover FC, Arbeit RD, Goering RV, Mickelsen PA, Murray BE, Persing DH, et al.
Interpreting chromosomal DNA restriction patterns produced by pulsed-field gel electropho-
resis: criteria for bacterial strain typing. J Clin Microbiol. 1995;33(9):2233–9.
40. Wiener J, Quinn JP, Bradford PA, Goering RV, Nathan C, Bush K, et al. Multiple antibiotic-
resistant Klebsiella and Escherichia coli in nursing homes. JAMA. 1999;281(6):517–23.
41. Mayer LW. Use of plasmid profiles in epidemiologic surveillance of disease outbreaks and in
tracing the transmission of antibiotic resistance. Clin Microbiol Rev. 1988;1(2):228–43.
42. Grimont F, Grimont PA. Ribosomal ribonucleic acid gene restriction patterns as potential
taxonomic tools. Ann Inst Pasteur Microbiol. 1986;137b(2):165–75.
43. Blumberg HM, Rimland D, Kiehlbauch JA, Terry PM, Wachsmuth IK. Epidemiologic typ-
ing of Staphylococcus aureus by DNA restriction fragment length polymorphisms of rRNA
genes: elucidation of the clonal nature of a group of bacteriophage-nontypeable, ciprofloxacin-
resistant, methicillin-susceptible S. aureus isolates. J Clin Microbiol. 1992;30(2):362–9.
44. Popovic T, Bopp CA, Olsvik O, Kiehlbauch JA. Ribotyping in molecular epidemiology.
In: Persing DH, editor. Diagnostic molecular microbiology, principles and applications.
Washington, D.C.: American Society for Microbiology; 1993. p. 573–83.
45. Bingen EH, Denamur E, Elion J. Use of ribotyping in epidemiological surveillance of noso-
comial outbreaks. Clin Microbiol Rev. 1994;7(3):311–27.
46. Jones CH, Tuckman M, Keeney D, Ruzin A, Bradford PA. Characterization and sequence
analysis of extended-spectrum-β-lactamase-encoding genes from Escherichia coli, Klebsiella
pneumoniae, and Proteus mirabilis isolates collected during tigecycline phase 3 clinical tri-
als. Antimicrob Agents Chemother. 2009;53(2):465–75.
47. Adler A, Miller-Roll T, Bradenstein R, Block C, Mendelson B, Parizade M, et al. A national
survey of the molecular epidemiology of Clostridium difficile in Israel: the dissemination of
the ribotype 027 strain with reduced susceptibility to vancomycin and metronidazole. Diagn
Microbiol Infect Dis. 2015;83(1):21–4.
48. McAleese F, Murphy E, Babinchak T, Singh G, Said-Salim B, Kreiswirth B, et al. Use of
ribotyping to retrospectively identify methicillin-resistant Staphylococcus aureus isolates
from phase 3 clinical trials for tigecycline that are genotypically related to community-
associated isolates. Antimicrob Agents Chemother. 2005;49(11):4521–9.
49. Bouchet V, Huot H, Goldstein R. Molecular genetic basis of ribotyping. Clin Microbiol Rev.
2008;21(2):262–73.
50. Fawley WN, Knetsch CW, MacCannell DR, Harmanus C, Du T, Mulvey MR, et al.
Development and validation of an internationally-standardized, high-resolution capillary
330 P. A. Bradford
69. Zheng B, Dai Y, Liu Y, Shi W, Dai E, Han Y, et al. Molecular epidemiology and risk factors of
Carbapenem-resistant Klebsiella pneumoniae infections in Eastern China. Front Microbiol.
2017;8:1061.
70. Adler A, Glick R, Lifshitz Z, Carmeli Y. Does Acinetobacter baumannii serve as a source
for blaNDM dissemination into Enterobacteriaceae in hospitalized patients? Microb Drug
Resist. 2017;
71. Sit PS, Teh CS, Idris N, Sam IC, Syed Omar SF, Sulaiman H, et al. Prevalence of methicillin-
resistant Staphylococcus aureus (MRSA) infection and the molecular characteristics of
MRSA bacteraemia over a two-year period in a tertiary teaching hospital in Malaysia. BMC
Infect Dis. 2017;17(1):274.
72. Feil EJ, Li BC, Aanensen DM, Hanage WP, Spratt BG. eBURST: inferring patterns of evo-
lutionary descent among clusters of related bacterial genotypes from multilocus sequence
typing data. J Bacteriol. 2004;186(5):1518–30.
73. Nadon C, Van Walle I, Gerner-Smidt P, Campos J, Chinen I, Concepcion-Acevedo J, et al.
PulseNet International: Vision for the implementation of whole genome sequencing (WGS)
for global food-borne disease surveillance. Euro Surveill. 2017, 22;(23):pii: 30544.
74. Lytsy B, Engstrand L, Gustafsson A, Kaden R. Time to review the gold standard for genotyp-
ing vancomycin-resistant enterococci in epidemiology: comparing whole-genome sequenc-
ing with PFGE and MLST in three suspected outbreaks in Sweden during 2013-2015. Infect
Genet Evol. 2017;54:74–80.
75. Hall TA, Sampath R, Blyn LB, Ranken R, Ivy C, Melton R, et al. Rapid molecular genotyp-
ing and clonal complex assignment of Staphylococcus aureus isolates by PCR coupled to
electrospray ionization-mass spectrometry. J Clin Microbiol. 2009;47(6):1733–41.
76. Nicolas-Chanoine M-H, Bertrand X, Madec J-Y. Escherichia coli ST131, an intriguing clonal
group. Clin Microbiol Rev. 2014;27(3):543–74.
77. Peirano G, Pitout JDD. Molecular epidemiology of Escherichia coli producing CTX-M
β-lactamases: the worldwide emergence of clone ST131 O25:H4. Int J Antimicrob Agents.
2010;35(4):316–21.
78. Peirano G, Bradford PA, Kazmierczak KM, Badal RE, Hackel M, Hoban DJ, et al.
Global incidence of carbapenemase-producing Escherichia coli ST131. Emerg Infect Dis.
2014;20(11):1928–31.
79. Banerjee R, Johnson JR. A new clone sweeps clean: the enigmatic emergence of Escherichia
coli sequence type 131. Antimicrob Agents Chemother. 2014;58(9):4997–5004.
80. Mathers AJ, Peirano G, Pitout JDD. The role of epidemic resistance plasmids and inter-
national high-risk clones in the spread of multidrug-resistant Enterobacteriaceae. Clin
Microbiol Rev. 2015;28(3):565–91.
81. Peirano G, Bradford PA, Kazmierczak KM, Chen L, Kreiswirth BN, Pitout JD. Importance
of clonal complex 258 and IncFK2-like plasmids among a global collection of Klebsiella
pneumoniae with blaKPC. Antimicrob Agents Chemother. 2017;61(4)
82. Baraniak A, Izdebski R, Zabicka D, Bojarska K, Gorska S, Literacka E, et al. Multiregional
dissemination of KPC-producing Klebsiella pneumoniae ST258/ST512 genotypes in Poland,
2010-14. J Antimicrob Chemother. 2017;72(6):1610–6.
83. Kanamori H, Parobek CM, Juliano JJ, van Duin D, Cairns BA, Weber DJ, et al. A prolonged
outbreak of KPC-3-producing Enterobacter cloacae and Klebsiella pneumoniae driven by
multiple mechanisms of resistance transmission at a large academic burn center. Antimicrob
Agents Chemother. 2017;61(2):pii: e01516-16.
84. Ruppe E, Olearo F, Pires D, Baud D, Renzi G, Cherkaoui A, et al. Clonal or not clonal?
Investigating hospital outbreaks of KPC-producing Klebsiella pneumoniae with whole-
genome sequencing. Clin Microbiol Infect. 2017;23(7):470–5.
85. Deng X, den Bakker HC, Hendriksen RS. Genomic epidemiology: whole-genome-
sequencing-powered surveillance and outbreak investigation of foodborne bacterial patho-
gens. Annu Rev Food Sci Technol. 2016;7:353–74.
332 P. A. Bradford
86. Humphries RM, Yang S, Kim S, Muthusamy VR, Russell D, Trout AM, et al. Duodenoscope-
related outbreak of a carbapenem resistant Klebsiella pneumoniae identified using advanced
molecular diagnostics. Clin Infect Dis. 2017;65(7):1159–66.
87. Archibald L, Phillips L, Monnet D, JE MG Jr, Tenover F, Gaynes R. Antimicrobial resistance
in isolates from inpatients and outpatients in the United States: increasing importance of the
intensive care unit. Clin Infect Dis. 1997;24(2):211–5.
88. Hetem DJ, Derde LP, Empel J, Mroczkowska A, Orczykowska-Kotyna M, Kozinska A, et al.
Molecular epidemiology of MRSA in 13 ICUs from eight European countries. J Antimicrob
Chemother. 2016;71(1):45–52.
89. Sader HS, Mendes RE, Streit JM, Flamm RK. Antimicrobial susceptibility trends among
Staphylococcus aureus from United States hospitals: results from 7 years of the Ceftaroline
(AWARE) Surveillance Program (2010-2016). Antimicrob Agents Chemother. 2017;61(9):pii:
e01043-17.
90. Adam HJ, Baxter MR, Davidson RJ, Rubinstein E, Fanella S, Karlowsky JA, et al.
Comparison of pathogens and their antimicrobial resistance patterns in paediatric, adult and
elderly patients in Canadian hospitals. J Antimicrob Chemother. 2013;68(Suppl 1):i31–7.
91. ECDC. Surveillance atlas of infectious disease 2017. Available from: http://atlas.ecdc.europa.
eu/public/index.aspx.
92. Lee NY, Song JH, Kim S, Peck KR, Ahn KM, Lee SI, et al. Carriage of antibiotic-resistant
pneumococci among Asian children: a multinational surveillance by the Asian Network for
Surveillance of Resistant Pathogens (ANSORP). Clin Infect Dis. 2001;32(10):1463–9.
93. Di Martino P, Livrelli V, Sirot D, Joly B, Darfeuille-Michaud A. A new fimbrial antigen har-
bored by CAZ-5/SHV-4-producing Klebsiella pneumoniae strains involved in nosocomial
infections. Infect Immun 1996;64(6):2266–2273.
94. Jevons MP. “Celbenin” – resistant staphylococci. Br Med J. 1961;1(5219):124–5.
95. Chambers HF. Methicillin resistance in staphylococci: molecular and biochemical basis and
clinical implications. Clin Microbiol Rev. 1997;10(4):781–91.
96. Kosowska-Shick K, McGhee PL, Appelbaum PC. Affinity of ceftaroline and other β-lactams
for penicillin-binding proteins from Staphylococcus aureus and Streptococcus pneumoniae.
Antimicrob Agents Chemother. 2010;54(5):1670–7.
97. Deurenberg RH, Stobberingh EE. The evolution of Staphylococcus aureus. Infect Genet
Evol. 2008;8(6):747–63.
98. Zhang K, McClure J-A, Elsayed S, Conly JM. Novel Staphylococcal Cassette Chromosome
mec type, tentatively designated type VIII, harboring Class A mec and type 4 ccr gene
complexes in a Canadian epidemicxtrain of methicillin-resistant Staphylococcus aureus.
Antimicrob Agents Chemother. 2009;53(2):531–40.
99. Elements IWGCSCC. Classification of Staphylococcal Cassette Chromosome mec
(SCCmec): guidelines for reporting novel SCCmec elements. Antimicrob Agents Chemother.
2009;53(12):4961–7.
100. Chambers HF. The changing epidemiology of Staphylococcus aureus? Emerg Infect Dis.
2001;7(2):178–82.
101. Albrich WC, Harbarth S. Health-care workers: source, vector, or victim of MRSA? Lancet
Infect Dis. 2008;8(5):289–301.
102. Adcock PM, Pastor P, Medley F, Patterson JE, Murphy TV. Methicillin-resistant
Staphylococcus aureus in two child care centers. J Infect Dis. 1998;178(2):577–80.
103. Katayama Y, Robinson DA, Enright MC, Chambers HF. Genetic background affects stability
of mecA in Staphylococcus aureus. J Clin Microbiol. 2005;43(5):2380–3.
104. Paterson GK, Harrison EM, Holmes MA. The emergence of mecC methicillin-resistant
Staphylococcus aureus. Trends Microbiol. 2014;22(1):42–7.
105. Musser JM, Kapur V. Clonal analysis of methicillin-resistant Staphylococcus aureus strains
from intercontinental sources: association of the mec gene with divergent phylogenetic lin-
eages implies dissemination by horizontal transfer and recombination. J Clin Microbiol.
1992;30(8):2058–63.
10 Epidemiology of Bacterial Resistance 333
106. Reacher MH, Shah A, Livermore DM, Wale MCJ, Graham C, Johnson AP, et al. Bacteraemia
and antibiotic resistance of its pathogens reported in England and Wales between 1990 and
1998: trend analysis. BMJ. 2000;320(7229):213–6.
107. Johnson AP, Aucken HM, Cavendish S, Ganner M, Wale MCJ, Warner M, et al. Dominance
of EMRSA-15 and -16 among MRSA causing nosocomial bacteraemia in the UK: analy-
sis of isolates from the European Antimicrobial Resistance Surveillance System (EARSS).
J Antimicrob Chemother. 2001;48(1):143–4.
108. Nakamura A, Miyake K, Misawa S, Kuno Y, Horii T, Hori S, et al. Association between anti-
microbial consumption and clinical isolates of methicillin-resistant Staphylococcus aureus: a
14-year study. J Infect Chemother. 2012;18(1):90–5.
109. Ravensbergen SJ, Berends M, Stienstra Y, Ott A. High prevalence of MRSA and ESBL
among asylum seekers in the Netherlands. PLoS One. 2017;12(4):e0176481.
110. Saravolatz LD, Markowitz N, Arking L, Pohlod D, Fisher E. Methicillin-resistant
Staphylococcus aureus. Epidemiologic observations during a community-acquired outbreak.
Ann Intern Med. 1982;96(1):11–6.
111. CDC. Four pediatric deaths from community-acquired methicillin-resistant Staphylococcus
aureus – Minnesota and North Dakota, 1997-1999. MMWR Morb Mortal Wkly Rep
1999;48(32):707–710.
112. Ugarte Torres A, Chu A, Read R, MacDonald J, Gregson D, Louie T, et al. The epidemiology
of Staphylococcus aureus carriage in patients attending inner city sexually transmitted infec-
tions and community clinics in Calgary, Canada. PLoS One. 2017;12(5):e0178557.
113. Akpaka PE, Roberts R, Monecke S. Molecular characterization of antimicrobial resistance
genes against Staphylococcus aureus isolates from Trinidad and Tobago. J Infect Public
Health. 2017;10(3):316–23.
114. Zuma AV, Lima DF, Assef AP, Marques EA, Leao RS. Molecular characterization of
methicillin-resistant Staphylococcus aureus isolated from blood in Rio de Janeiro displaying
susceptibility profiles to non-beta-lactam antibiotics. Braz J Microbiol. 2017;48(2):237–41.
115. Lynch JP, Zhanel GG. Streptococcus pneumoniae: epidemiology, risk factors, and strategies
for prevention. Semin Respir Crit Care. 2009;30(2):189–209.
116. Hansman D, Bullen M. A resistant pneumococcus. Lancet. 1967;290(7509):264–5.
117. Michel J, Dickman D, Greenberg Z, Bergner-Rabinowitz S. Serotype distribution of
penicillin-resistant pneumococci and their susceptibilities to seven antimicrobial agents.
Antimicrob Agents Chemother. 1983;23(3):397–401.
118. Kyaw MH, Lynfield R, Schaffner W, Craig AS, Hadler J, Reingold A, et al. Effect of introduc-
tion of the pneumococcal conjugate vaccine on drug-resistant Streptococcus pneumoniae. N
Engl J Med. 2006;354(14):1455–63.
119. Jabes D, Nachman S, Tomasz A. Penicillin-binding protein families: evidence for the
clonal nature of penicillin resistance in clinical isolates of pneumococci. J Infect Dis.
1989;159(1):16–25.
120. Dowson CG, Coffey TJ, Kell C, Whiley RA. Evolution of penicillin resistance in Streptococcus
pneumoniae; the role of Streptococcus mitis in the formation of a low affinity PBP2B in S.
pneumoniae. Mol Microbiol. 1993;9(3):635–43.
121. Sibold C, Wang J, Henrichsen J, Hakenbeck R. Genetic relationships of penicillin-susceptible
and -resistant Streptococcus pneumoniae strains isolated on different continents. Infect
Immun. 1992;60(10):4119–26.
122. Kristinsson KG, Hjalmarsdottir MA, Steingrimsson O. Increasing penicillin resistance in
pneumococci in Iceland. Lancet. 1992;339(8809):1606–7.
123. Soares S, Kristinsson KG, Musser JM, Tomasz A. Evidence for the introduction of a multi-
resistant clone of serotype 6B Streptococcus pneumoniae from Spain to Iceland in the late
1980s. J Infect Dis. 1993;168(1):158–63.
124. Hjalmarsdottir MA, Kristinsson KG. Epidemiology of penicillin-non-susceptible pneumo-
cocci in Iceland, 1995-2010. J Antimicrob Chemother. 2014;69(4):940–6.
334 P. A. Bradford
125. Kim L, McGee L, Tomczyk S, Beall B. Biological and epidemiological features of antibiotic-
resistant Streptococcus pneumoniae in pre- and post-conjugate vaccine eras: a United States
perspective. Clin Microbiol Rev. 2016;29(3):525–52.
126. Prevention CfDCa. Active Bacterial Core Surveillance (ABCs) Report Emerging Infections
Program Network Streptococcus pneumoniae, 2015. Available from: https://www.cdc.gov/
abcs/reports-findings/survreports/spneu-types.html.
127. Leclercq R, Derlot E, Duval J, Courvalin P. Plasmid-mediated resistance to vancomycin and
teicoplanin in Enterococcus faecium. N Engl J Med. 1988;319(3):157–61.
128. Sahm DF, Kissinger J, Gilmore MS, Murray PR, Mulder R, Solliday J, et al. In vitro suscepti-
bility studies of vancomycin-resistant Enterococcus faecalis. Antimicrob Agents Chemother.
1989;33(9):1588–91.
129. Acar J, Casewell M, Freeman J, Friis C, Goossens H. Avoparcin and virginiamycin as
animal growth promoters: a plea for science in decision-making. Clin Microbiol Infect.
2000;6(9):477–82.
130. Kirst HA, Thompson DG, Nicas TI. Historical yearly usage of vancomycin. Antimicrob
Agents Chemother. 1998;42(5):1303–4.
131. Frieden TR, Munsiff SS, Low DE, Willey BM, Williams G, Faur Y, et al. Emergence of
vancomycin-resistant enterococci in New York City. Lancet. 1993;342(8863):76–9.
132. Deshpande LM, Fritsche TR, Moet GJ, Biedenbach DJ, Jones RN. Antimicrobial resistance
and molecular epidemiology of vancomycin-resistant enterococci from North America and
Europe: a report from the SENTRY antimicrobial surveillance program. Diagn Microbiol
Infect Dis. 2007;58(2):163–70.
133. O'Driscoll T, Crank CW. Vancomycin-resistant enterococcal infections: epidemiology, clini-
cal manifestations, and optimal management. Infect Drug Resist. 2015;8:217–30.
134. Baden LR, Critchley IA, Sahm DF, So W, Gedde M, Porter S, et al. Molecular characterization
of vancomycin-resistant Enterococci repopulating the gastrointestinal tract following treat-
ment with a novel glycolipodepsipeptide, ramoplanin. J Clin Microbiol. 2002;40(4):1160–3.
135. Snyder GM, Thom KA, Furuno JP, Perencevich EN, Roghmann MC, Strauss SM, et al.
Detection of methicillin-resistant Staphylococcus aureus and vancomycin-resistant entero-
cocci on the gowns and gloves of healthcare workers. Infect Control Hosp Epidemiol.
2008;29(7):583–9.
136. Ghanem G, Hachem R, Jiang Y, Chemaly RF, Raad I. Outcomes for and risk factors associated
with vancomycin-resistant Enterococcus faecalis and vancomycin-resistant Enterococcus
faecium bacteremia in cancer patients. Infect Control Hosp Epidemiol. 2007;28(9):1054–9.
137. Hidron AI, Edwards JR, Patel J, Horan TC, Sievert DM, Pollock DA, et al. NHSN annual
update: antimicrobial-resistant pathogens associated with healthcare-associated infec-
tions: annual summary of data reported to the National Healthcare Safety Network at the
Centers for Disease Control and Prevention, 2006-2007. Infect Control Hosp Epidemiol.
2008;29(11):996–1011.
138. Sievert DM, Ricks P, Edwards JR, Schneider A, Patel J, Srinivasan A, et al. Antimicrobial-
resistant pathogens associated with healthcare-associated infections: summary of data
reported to the National Healthcare Safety Network at the Centers for Disease Control and
Prevention, 2009-2010. Infect Control Hosp Epidemiol. 2013;34:1):1–14.
139. Zhanel GG, Adam HJ, Baxter MR, Fuller J, Nichol KA, Denisuik AJ, et al. Antimicrobial
susceptibility of 22746 pathogens from Canadian hospitals: results of the CANWARD 2007-
11 study. J Antimicrob Chemother. 2013;68(Suppl 1):i7–22.
140. ECDC. European Centre for Disease Prevention and Control. Antimicrobial resistance
surveillance in Europe 2015. Annual report of the European Antimicrobial Resistance
Surveillance Network (EARS-Net). 2017.
141. Yigit H, Queenan AM, Anderson GJ, Domenech-Sanchez A, Biddle JW, Steward CD, et al.
Novel carbapenem-hydrolyzing β-lactamase, KPC-1, from a carbapenem-resistant strain of
Klebsiella pneumoniae. Antimicrob Agents Chemother. 2001;45(4):1151–61.
10 Epidemiology of Bacterial Resistance 335
142. Bradford PA, Bratu S, Urban C, Visalli M, Mariano N, Landman D, et al. Emergence of
carbapenem-resistant Klebsiella species possessing the class A carbapenem-hydrolyzing
KPC-2 and inhibitor-resistant TEM-30 β-lactamases in New York City. Clin Infect Dis.
2004;39(1):55–60.
143. Navon-Venezia S, Leavitt A, Schwaber MJ, Rasheed JK, Srinivasan A, Patel JB, et al. First
report on a hyperepidemic clone of KPC-3-producing Klebsiella pneumoniae in Israel geneti-
cally related to a strain causing outbreaks in the United States. Antimicrob Agents Chemother.
2009;53(2):818–20.
144. Kazmierczak KM, Biedenbach DJ, Hackel M, Rabine S, de Jonge BLM, Bouchillon SK,
et al. Global dissemination of blaKPC into bacterial species beyond Klebsiella pneumoniae
and in vitro susceptibility to ceftazidime-avibactam and Aztreonam-avibactam. Antimicrob
Agents Chemother. 2016;60(8):4490–500.
145. Y-y H, D-x G, J-c C, H-w Z, Zhang R. Emergence of KPC-2-producing Pseudomonas aeru-
ginosa sequence type 463 isolates in Hangzhou, China. Antimicrob Agents Chemother.
2015;59(5):2914–7.
146. Liang Y, Yin X, Zeng L, Chen S. Clonal replacement of epidemic KPC-producing Klebsiella
pneumoniae in a hospital in China. BMC Infect Dis. 2017;17(1):363.
147. Kim JO, Song SA, Yoon EJ, Shin JH, Lee H, Jeong SH, et al. Outbreak of KPC-2-producing
Enterobacteriaceae caused by clonal dissemination of Klebsiella pneumoniae ST307 car-
rying an IncX3-type plasmid harboring a truncated Tn4401a. Diagn Microbiol Infect Dis.
2017;87(4):343–8.
148. Vubil D, Figueiredo R, Reis T, Canha C, Boaventura L, GJ DAS. Outbreak of KPC-3-
producing ST15 and ST348 Klebsiella pneumoniae in a Portuguese hospital. Epidemiol
Infect. 2017;145(3):595–9.
149. Snitkin ES, Zelazny AM, Thomas PJ, Stock F, Henderson DK, Palmore TN, et al. Tracking
a hospital outbreak of carbapenem-resistant Klebsiella pneumoniae with whole-genome
sequencing. Sci Transl Med. 2012;4(148):148ra16-ra16.
150. Abdallah M, Olafisoye O, Cortes C, Urban C, Landman D, Ghitan M, et al. Rise and fall
of KPC-producing Klebsiella pneumoniae in New York City. J Antimicrob Chemother.
2016;71(10):2945–8.
151. Walther-Rasmussen J, Høiby N. Class A carbapenemases. J Antimicrob Chemother.
2007;60(3):470–82.
152. Hecker SJ, Reddy KR, Totrov M, Hirst GC, Lomovskaya O, Griffith DC, et al. Discovery
of a cyclic boronic acid beta-lactamase inhibitor (RPX7009) with utility vs class A serine
carbapenemases. J Med Chem. 2015;58(9):3682–92.
153. Munoz-Price LS, Poirel L, Bonomo RA, Schwaber MJ, Daikos GL, Cormican M, et al.
Clinical epidemiology of the global expansion of Klebsiella pneumoniae carbapenemases.
Lancet Infect Dis. 2013;13(9):785–96.
154. Pitout JDD, Nordmann P, Poirel L. Carbapenemase-producing Klebsiella pneumoniae,
a key pathogen set for global nosocomial dominance. Antimicrob Agents Chemother.
2015;59(10):5873–84.
155. Chen L, Chavda KD, DeLeo FR, Bryant KA, Jacobs MR, Bonomo RA, et al. Genome
sequence of a Klebsiella pneumoniae sequence type 258 isolate with prophage-encoded K.
pneumoniae carbapenemase. Genome Announc. 2015;3(3)
156. Yong D, Toleman MA, Giske CG, Cho HS, Sundman K, Lee K, et al. Characterization of
a new metallo-β-lactamase gene, blaNDM-1, and a novel erythromycin esterase gene car-
ried on a unique genetic structure in Klebsiella pneumoniae sequence type 14 from India.
Antimicrob Agents Chemother. 2009;53(12):5046–54.
157. Nordmann P, Poirel L. The difficult-to-control spread of carbapenemase producers among
Enterobacteriaceae worldwide. Clin Microbiol Infect. 2014;20(9):821–30.
158. Kumarasamy KK, Toleman MA, Walsh TR, Bagaria J, Butt F, Balakrishnan R, et al.
Emergence of a new antibiotic resistance mechanism in India, Pakistan, and the UK: a molec-
ular, biological, and epidemiological study. Lancet Infect Dis. 2010;10(9):597–602.
336 P. A. Bradford
159. Kazmierczak KM, Rabine S, Hackel M, McLaughlin RE, Biedenbach DJ, Bouchillon SK,
et al. Multiyear, multinational survey of the incidence and global distribution of metallo-β-
lactamase-producing Enterobacteriaceae and Pseudomonas aeruginosa. Antimicrob Agents
Chemother. 2016;60(2):1067–78.
160. Bocanegra-Ibarias P, Garza-González E, Morfín-Otero R, Barrios H, Villarreal-Treviño L,
Rodríguez-Noriega E, et al. Molecular and microbiological report of a hospital outbreak of
NDM-1-carrying Enterobacteriaceae in Mexico. PLoS One. 2017;12(6):e0179651.
161. Bosch T, Lutgens SPM, Hermans MHA, Wever PC, Schneeberger PM, Renders NHM, et al.
An outbreak of NDM-1 producing Klebsiella pneumoniae in a Dutch hospital with interspe-
cies transfer of the resistance plasmid and unexpected occurrence in unrelated healthcare
centers. J Clin Microbiol. 2017;
162. Yu J, Wang Y, Chen Z, Zhu X, Tian L, Li L, et al. Outbreak of nosocomial NDM-1-
producing Klebsiella pneumoniae ST1419 in a neonatal unit. J Global Antimicrobial Resist.
2017;8:135–9.
163. Lin J-N, Chang L-L, Lai C-H, Huang Y-H, Chen W-F, Yang C-H, et al. High prevalence of
fluoroquinolone-nonsusceptible Streptococcus pyogenes emm12 in Taiwan. Diagn Microbiol
Infect Dis. 2015;83(2):187–92.
164. McCurdy SP, Jones RN, Mendes RE, Puttagunta S, Dunne MW. In vitro activity of
Dalbavancin against drug-resistant Staphylococcus aureus isolates from a global surveillance
program. Antimicrob Agents Chemother. 2015;59(8):5007–9.
165. Weiner LM, Webb AK, Limbago B, Dudeck MA, Patel J, Kallen AJ, et al. Antimicrobial-
resistant pathogens associated with healthcare-associated infections: summary of data
reported to the National Healthcare Safety Network at the Centers for Disease Control and
Prevention, 2011–2014. Infect Control Hosp Epidemiol. 2016;37(11):1288–301.
166. Bidell MR, Palchak M, Mohr J, Lodise TP. Fluoroquinolone and third-generation-
cephalosporin resistance among hospitalized patients with urinary tract infections due to
Escherichia coli: do rates vary by hospital characteristics and geographic region? Antimicrob
Agents Chemother. 2016;60(5):3170–3.
167. Strahilevitz J, Jacoby GA, Hooper DC, Robicsek A. Plasmid-mediated quinolone resistance:
a multifaceted threat. Clin Microbiol Rev. 2009;22(4):664–89.
168. Wu JJ, Ko WC, Wu HM, Yan JJ. Prevalence of Qnr determinants among bloodstream iso-
lates of Escherichia coli and Klebsiella pneumoniae in a Taiwanese hospital, 1999-2005.
J Antimicrob Chemother. 2008;61(6):1234–9.
169. Strahilevitz J, Engelstein D, Adler A, Temper V, Moses AE, Block C, et al. Changes in
qnr prevalence and fluoroquinolone resistance in clinical isolates of Klebsiella pneu-
moniae and Enterobacter spp. collected from 1990 to 2005. Antimicrob Agents Chemother.
2007;51(8):3001–3.
170. Castanheira M, Costello SE, Jones RN, Mendes RE, editors. Prevalence of aminoglycoside
resistance genes among contemporary Gram-negative resistant isolates collected worldwide.
25th European Congress of Clinical Microbiology and Infectious Diseases (ECCMID); 2015
April 25–28; Copenhagen.
171. Miro E, Grunbaum F, Gomez L, Rivera A, Mirelis B, Coll P, et al. Characterization of
aminoglycoside-modifying enzymes in Enterobacteriaceae clinical strains and characteriza-
tion of the plasmids implicated in their diffusion. Microb Drug Resist. 2013;19(2):94–9.
172. Almaghrabi R, Clancy CJ, Doi Y, Hao B, Chen L, Shields RK, et al. Carbapenem-resistant
Klebsiella pneumoniae strains exhibit diversity in aminoglycoside-modifying enzymes,
which exert differing effects on Plazomicin and other agents. Antimicrob Agents Chemother.
2014;58(8):4443–51.
173. Bell J, Andersson P, Jones RN, Turnidge J, editors. 16S rRNA methylase containing
Enterobacteriaceae in the SENTRY Asia-Pacific region frequently harbour plasmid-mediated
quinolone resistance and CTX-M types. European Congress of Clinical Microbiology and
Infectious Diseases (ECCMID); 2010 April 10–13; Vienna.
10 Epidemiology of Bacterial Resistance 337
191. Zhou H-W, Zhang T, Ma J-H, Fang Y, Wang H-Y, Huang Z-X, et al. Occurrence of plas-
mid- and chromosome-carried mcr-1 in waterborne Enterobacteriaceae in China. Antimicrob
Agents Chemother. 2017;61(8)
192. Hembach N, Schmid F, Alexander J, Hiller C, Rogall ET, Schwartz T. Occurrence of the
mcr-1 colistin resistance gene and other clinically relevant antibiotic resistance genes in
microbial populations at different municipal wastewater treatment plants in Germany. Front
Microbiol. 2017;8:1282.
193. Ovejero CM, Delgado-Blas JF, Calero-Caceres W, Muniesa M, Gonzalez-Zorn B. Spread of
mcr-1-carrying Enterobacteriaceae in sewage water from Spain. J Antimicrob Chemother.
2017;72(4):1050–3.
194. Kizny Gordon AE, Mathers AJ, Cheong EYL, Gottlieb T, Kotay S, Walker AS, et al.
The hospital water environment as a reservoir for carbapenem-resistant organisms caus-
ing hospital-acquired infections-a systematic review of the literature. Clin Infect Dis.
2017;64(10):1435–44.
195. Hashida K, Shiomori T, Hohchi N, Ohkubo J, Ohbuchi T, Mori T, et al. Nasopharyngeal
Streptococcus pneumoniae carriage in Japanese children attending day-care centers. Int
J Pediatr Otorhinolaryngol. 2011;75(5):664–9.
196. Dueger EL, Asturias EJ, Matheu J, Gordillo R, Torres O, Halsey N. Increasing penicillin
and trimethoprim-sulfamethoxazole resistance in nasopharyngeal Streptococcus pneumoniae
isolates from Guatemalan children, 2001-2006. Int J Infect Dis. 2008;12(3):289–97.
197. Braga EDV, Aguiar-Alves F, de Freitas MFN, de e Silva MO, Correa TV, Snyder RE, et al.
High prevalence of Staphylococcus aureus and methicillin-resistant S. aureus colonization
among healthy children attending public daycare centers in informal settlements in a large
urban center in Brazil. BMC Infect Dis. 2014;14
198. Rasmussen BA, Bradford PA, Quinn JP, Wiener J, Weinstein RA, Bush K. Genetically
diverse ceftazidime-resistant isolates from a single center: biochemical and genetic charac-
terization of TEM-10 β-lactamases encoded by different nucleotide sequences. Antimicrob
Agents Chemother. 1993;37(9):1989–92.
199. Bradford PA, Urban C, Jaiswal A, Mariano N, Rasmussen BA, Projan SJ, et al. SHV-7, a
novel cefotaxime-hydrolyzing β-lactamase, identified in Escherichia coli isolates from hos-
pitalized nursing home patients. Antimicrob Agents Chemother. 1995;39(4):899–905.
200. Valenza G, Nickel S, Pfeifer Y, Pietsch M, Voigtlander E, Lehner-Reindl V, et al. Prevalence
and genetic diversity of extended-spectrum beta-lactamase (ESBL)-producing Escherichia
coli in nursing homes in Bavaria, Germany. Vet Microbiol. 2017;200:138–41.
201. Trick WE, Weinstein RA, DeMarais PL, Kuehnert MJ, Tomaska W, Nathan C, et al.
Colonization of skilled-care facility residents with antimicrobial-resistant pathogens. J Am
Geriatr Soc. 2001;49(3):270–6.
202. Jiménez-Truque N, Saye EJ, Soper N, Saville BR, Thomsen I, Edwards KM, et al.
Longitudinal assessment of colonization with Staphylococcus aureus in healthy collegiate
athletes. J Pediatric Infect Dis Soc. 2016;5(2):105–13.
203. Lindenmayer JM, Schoenfeld S, O'Grady R, Carney JK. Methicillin-resistant Staphylococcus
aureus in a high school wrestling team and the surrounding community. Arch Intern Med.
1998;158(8):895–9.
204. Begier EM, Frenette K, Barrett NL, Mshar P, Petit S, Boxrud DJ, et al. A high-morbidity out-
break of methicillin-resistant Staphylococcus aureus among players on a college football team,
facilitated by cosmetic body shaving and turf burns. Clin Infect Dis. 2004;39(10):1446–53.
205. CDC. Methicillin-resistant Staphylococcus aureus infections among competitive sports par-
ticipants–Colorado, Indiana, Pennsylvania, and Los Angeles County, 2000–2003. MMWR.
2003;52:793–5.
206. Young LM, Motz VA, Markey ER, Young SC, Beaschler RE. Recommendations for best
disinfectant practices to reduce the spread of infection via wrestling mats. J Athl Train.
2017;52(2):82–8.
10 Epidemiology of Bacterial Resistance 339
George A. Jacoby
11.1 Introduction
G. A. Jacoby (*)
Lahey Hospital and Medical Center, Burlington, MA, USA
More than a hundred plasmid-mediated enzymes are known that modify aminogly-
cosides by transferring acetyl, phospho, or adenyl groups to essential –OH or –NH2
groups on the aminoglycoside core (Fig 11.1). Such modifications reduce the bind-
ing of the drug to its 16S rRNA target in the 30S ribosome. Two nomenclatures are
in use. In one three letters identify the enzyme activity (AAC, APH, ANT) followed
by the site of modification in parenthesis, a roman numeral for the resistance profile,
and a lower case letter as an individual identifier [25, 26]. In the other system, the
genes are labeled aac, aph, and aad followed by a letter or a number that identifies
the site of modification and another number as a unique identifier [2]. Thus,
aac(6′)-Ib and aacA4 represent the same gene for an aminoglycoside
N-acetyltransferase modifying position 6′ and giving a Ib pattern of resistance, a
potentially confusing situation. A table of equivalent nomenclature is available [25].
The aminoglycoside acetyltransferase group is the largest. There are four sub-
classes named according to the position of the amino group that is modified. AAC(1)
and AAC(3) target groups at positions 1 and 3 of the 2-deoxystreptamine ring, while
AAC(2′) and AAC(6′) target groups found at the 2′ or 6′ position of the
2,6-dideoxy-2,6-diaminoglucose ring. The AAC(6′) group can be further subdi-
vided into type I giving resistance to amikacin but not gentamicin and type II con-
ferring resistance to gentamicin but not amikacin. Both AAC(6′) groups modify
tobramycin and netilmicin and have many subgroups. AAC(6′) also exists as a
bifunctional hybrid with APH(2″) in Gram-positive bacteria. The second largest is
the phosphotransferase group, which targets 3′, 4′, and 2″ hydroxyls and provides
resistance to kanamycin, to neomycin, and for some enzymes to additional amino-
glycosides including amikacin. APH(3″) and APH(6) that modify streptomycin also
belong to this group. The final group of nucleotidyltransferases attacks hydroxyls at
the 2″, 3″, 4′, 6, and 9 positions. Clinically important members include ANT(2″)
active on 4,6-disubstituted aminoglycosides with a deoxystreptamine ring including
gentamicins, kanamycins, sisomicin, and tobramycin, ANT (3″) modifying the 3″
position of streptomycin and the 9-hydroxyl group of spectinomycin, and ANT(4′)
344 G. A. Jacoby
ANT(2) APH(2)
11.2.2 Bacitracin
described in an online database [39]. TEM-1 and TEM-2 belong to class A or group
2b as does SHV-1, another common plasmid-determined β-lactamase, differing in
sequence from TEM-1 by 36%. Group 2b enzymes hydrolyze penicillins such as
ampicillin and early cephalosporins such as cephalothin or cefazolin and are inhib-
ited by clavulanic acid, sulbactam, or tazobactam (Table 11.3).
Pharmaceutical chemists devised oxyimino-β-lactams such as cefotaxime,
ceftazidime, ceftriaxone, and the monobactam aztreonam that are effective against
TEM-1-, TEM-2-, or SHV-1-producing bacteria. Within a few years, transmissible
resistance to oxyimino-β-lactams appeared and proved to be due to mutations in
TEM-1, TEM-2, and SHV-1 that changed the charge and configuration of the active
site allowing access to β-lactams with the oxyimino side chain. Many such extended-
spectrum β-lactamases or ESBLs in the TEM (TEM-3, TEM-10, TEM-12, TEM-26,
etc.) and SHV (SHV-2, SHV-4, SHV-5, etc.) families are known and belong to
group 2be. Group 2be also includes a large family of CTX-M enzymes and smaller
ones of PER and VEB β-lactamases.
Combinations of clavulanic acid, sulbactam, or tazobactam with β-lactamase
susceptible drugs were also developed. Inhibitor-resistant TEM and SHV varieties
followed (group 2br) such as TEM-30 or SHV-10 as well as a few enzymes that
combined the extended-spectrum and inhibitor-resistant phenotypes (group 2ber),
including TEM-50 and TEM-68. β-lactamases pay a price for broader spectrum or
inhibitor resistance in terms of efficiency, so that compensatory promoter mutations
increasing gene expression often accompany such mutations [40] as well as changes
in the bacterial host to decrease β-lactam accumulation.
Carbenicillin and ticarcillin were developed to treat infections caused by P. aeru-
ginosa. β-lactamases with high activity toward carbenicillin (group 2c) were found
and, since they were at first thought to be pseudomonas-specific, were given names
like PSE-1 (for Pseudomonas-specific enzyme) or, for carbenicillinase activity,
CARB-3. A related variety able to hydrolyze cefepime as well as carbenicillin
(group 2ce) has also evolved.
Cephamycins, such as cefoxitin and cefotetan, were also developed as agents
effective against class A β-lactamase-producing bacteria and others. Class C (group
1) enzymes can hydrolyze cephamycins, but these enzymes were determined by
chromosomal ampC genes and hence restricted to bacteria expressing them until
plasmid-mediated class 3 enzymes (CMY-1, MIR-1, ACT-1, DHA-1, FOX-1,
MOX-1, and more in each family) were discovered [41]. These β-lactamases could
also hydrolyze oxyimino-β-lactams and were naturally resistant to the first genera-
tion of β-lactamase inhibitors. Group 1e enzymes have short deletions or amino acid
substitutions in the R2 loop that enhance activity toward particular substrates, espe-
cially ceftazidime. With host porin and efflux pump changes to reduce drug accu-
mulation, AmpC enzymes can even provide carbapenem resistance [42].
Carbapenems such as imipenem, meropenem, doripenem, and ertapenem came
to be widely used against ESBL and AmpC β-lactamase-producing bacteria with the
all too predictable appearance of carbapenemases belonging to class B (group 3a:
VIM-1, IMP-1, NDM-1) and class A (group 2f: KPC-2, IMI-1) β-lactamases, which
have spread worldwide [43] (see Chaps. 10 and 12 for further details). In the United
348
Long after colistin (polymyxin E) was first used clinically but within a few years of
its extensive use in animal husbandry, plasmid-mediated colistin resistance was
reported in 2016 in an isolate from an intensive pig farm in China. The plasmid gene
mcr-1 encodes a membrane-bound phosphoethanolamine transferase, a zinc metal-
loprotein, that modifies lipid A decreasing the binding affinity of polymyxins and
thus conferring a colistin MIC of 4–8 μg/ml [9]. Within a year of its discovery, mcr-
1 was found on a wide variety of plasmids in a diversity of Enterobacteriaceae in
countries on five continents and in isolates from as early as the 1980s [50]. ISApl1
is often found upstream from the mcr-1 gene [51]. mcr-1 has been found in 10% of
350 G. A. Jacoby
Many naturally occurring antibiotics and synthetic agents target DNA gyrase
besides quinolones. Qnr protects against compounds with a somewhat quinolone-
like structure, for example, 2-pyridone, quinazoline-2,4-dione, or spiropyrimidine-
trione, so it is not strictly quinolone specific. Qnr, however, does not block agents
acting on the GyrB subunit [58].
Expression of qnrB, qnrD, and qnrE is regulated by components of the bacterial
SOS system in which DNA damage induced by quinolones activates RecA protease
to cleave LexA protein which otherwise blocks expression by binding to specific
DNA sequence upstream from these qnr genes. Expression of qnrS1 is also induced
by quinolone and requires DNA sequence upstream from the gene, but qnrS1 regu-
lation is independent of the SOS system. Expression of qnrA in the aquatic organ-
ism Shewanella algae, on the other hand, is induced not by quinolone but by cold
shock.
Qnr plasmids have been found around the world in a variety of Enterobacteriaceae,
especially E. coli, E. cloacae, K. pneumoniae, and S. enterica but rarely in non-
fermenting bacteria such as P. aeruginosa or A. baumannii. Plasmids carrying qnr
genes vary in size and incompatibility specificity, indicating that the spread of mul-
tiple plasmids has been responsible for their dissemination and that plasmid acquisi-
tion has occurred multiple times. A mobile or transposable element is almost
invariably associated with plasmid-mediated qnr genes, especially insertion
sequences ISCR1, ISExp1, and IS26. The complex is often inserted into a sul1-type
integron. qnrVC is so far the only qnr gene located in a cassette with a linked attC
site ready by itself for integron capture. Because of such linkage, qnr genes are
often found on plasmids with genes for other resistance determinants such as ESBLs
and carbapenemases. Qnr prevalence seems to be increasing and has reached as
high as 39% in a sample of E. cloacae isolates at one hospital in China [59].
The second plasmid-mediated quinolone resistance mechanism to be discovered
involves modification. AAC(6′)-Ib-cr is an acetyltransferase providing resistance to
kanamycin, tobramycin, and amikacin that by two amino acid substitutions has
broadened its spectrum to acetylate quinolones with an amino nitrogen on the piper-
azinyl ring, such as ciprofloxacin and norfloxacin [60]. Acetylation decreases the
antibacterial potency raising the ciprofloxacin MIC eightfold. Both the substitutions
Trp102Arg and Asp79Tyr are required for quinolone-acetylating activity. Seven
nucleotide variants of aac(6′)-Ib-cr are known, one with a 26 amino acid insertion
changing the position of the amino acid changes essential for anti-quinolone activ-
ity [61]. The aac(6’)-Ib-cr gene has been found worldwide in a variety of
Enterobacteriaceae (especially E. coli) and is often more common in surveys than
qnr alleles. It is usually found in a cassette as part of a class 1 integron in a multire-
sistance plasmid, which may contain other plasmid-mediated quinolone resistance
genes. Association with CTX-M-15 is particularly common.
The third mechanism of plasmid-mediated quinolone resistance involves efflux
pumps. QepA is a plasmid-acquired efflux pump in the major facilitator family that
decreases susceptibility to hydrophilic fluoroquinolones. qepA has often been found
on plasmids together with aminoglycoside ribosomal methyltransferase rmtB. At
least three qepA variants are known. OqxAB is an efflux pump in the
352 G. A. Jacoby
11.2.8 Glycopeptide
Ddl
a
L-Ala D-Ala futile D-Ala- D-Ala Penta Penta
cycle
racemase
VanX
adding VanY Tetra
Tri
enzyme
VanH VanA
Pyruvate D-Lac D-Ala-D-Lac Pentadepsi Pentadepsi
PR PH
VanR VanS VanH VanA VanX VanY VanZ
Fig. 11.2 VanA-type glycopeptide resistance. a: Organization of the vanA operon in Tn1546.
Open arrows represent coding sequences and indicate the direction of transcription. IRL and IRR,
inverted repeat left and right, respectively. b: Synthesis of peptidoglycan precursors in a VanA-
type-resistant strain after induction with glycopeptide; Ddl, D-Ala:D-Ala ligase [79]
only distantly (34% and 24%) related to those of VanA. A vanW gene of unknown
function is included but no vanZ.
The VanC cluster produces peptidoglycan precursors ending in D-Ala-D-Ser and
is chromosomal and not transferable. VanT is a racemase converting L-Ser to D-Ser
which is joined to D-Ala by VanC ligase and incorporated into peptidoglycan pre-
cursors in place of D-Ala-D-Ala. The two enzymes of VanA or VanB that eliminate
D-Ala-D-Ala targets are combined in VanC in a single VanXYC D,D-peptidase.
Expression of vanC, vanXYC, and vanT is again controlled by a two-component
regulatory system VanRC and VanSC. The VanE cluster has a similar organization
and 41–60% identity to VanC [81].
VanA and VanB are the most frequent causes of vancomycin resistance in
Enterococcus spp. and have occasionally been detected in coryneform bacteria and
other streptococci. VanA has also appeared in S. aureus causing high-level resis-
tance but fortunately only rarely.
In a sample of enterococci causing bloodstream infections between 2001 and
2014 from the United States, the frequency of vancomycin resistance in E. faecalis
was 6.5% in 2008 declining to 1.9% in 2014 but reached 80.7% in E. faecium in
2010 and 68.4% in 2014. In Europe, the corresponding vancomycin resistance fre-
quencies were 2.6% in E. faecalis and peaked at 27.3% in E. faecium in 2012. All
three lipoglycopeptides are active against VanB enterococci, while oritavancin and
telavancin have activity as well against VanA strains but of unproven clinical signifi-
cance [82].
11 Transmissible Antibiotic Resistance 355
genes have been distinguished based on >20% difference in sequence. Many are
determined by plasmids, transposons, or integrative and conjugative elements [88].
Erm proteins add one or two methyl groups to adenine 2058 in domain V of 23S
rRNA preventing MLSB antibiotic attachment. Resistance is produced to 14-, 15-,
and 16-membered macrolides, ketolides, lincosamides, and streptogramin B antibi-
otics. erm(A), erm(B), and erm(C) are typically found in staphylococci. erm(B) and
a subclass of erm(A) [erm(TR)] are widespread in enterococci and streptococci.
erm(F) has been found in anaerobes and H. influenzae. erm(A) is part of transposon
Tn554 or its close relative Tn6133, while erm(B) is part of transposons Tn917 and
Tn551. erm(C) is often located on small plasmids in staphylococci and erm(T) on
larger ones. erm(33) is the result of in vivo recombination between erm(A) and
erm(C). erm expression may be inducible or constitutive. Erythromycin and other
14- and 15-membered macrolides tend to be good inducers via a mechanism that
involves ribosome stalling while translating an upstream leader peptide with conse-
quent changes in the structure of erm mRNA that allows it to be translated. Ketolides
and lincosamides are usually not inducers but may become so by deletions, inser-
tions, and point mutations in this attenuator system [87, 89, 90]. Hence a staphylo-
coccal strain with erm(A) or erm(C) may appear erythromycin resistant but
clindamycin susceptible, but if exposed to clindamycin, it can mutate to resistance
to both agents [91].
A different methyl transferase is encoded by the cfr gene and confers resistance
to lincosamides, streptogramins A, phenicols, oxazolidinones, and pleuromutilins
and decreased susceptibility to such 16-membered macrolides as josamycin and
spiramycin. It adds a methyl group from S-adenosyl-L-methionine to the C8 posi-
tion of adenine 2503 at the peptidyltransferase center in domain V of 23 rRNA by a
two-step mechanism involving intermediate methylation of a Cys residue on the
enzyme [92]. The cfr gene has been found worldwide in Staphylococcus spp.,
Enterococcus spp., other Gram-positive organisms, and P. vulgaris and E. coli on
plasmids or together with insertion sequences. Cfr(B) with 74.9% amino acid iden-
tity to Cfr(A) has been described in E. faecium [93] and Cfr(C) with 55.1% identity
in Campylobacter spp. [94].
Plasmid-mediated efflux genes are also involved in MLSB resistance. Msr(A) in
the ABC transporter family confers resistance to 14- and 15-membered macrolides,
and streptogramin B antibiotics and low-level resistance to ketolides, while Mef(A)
in the major facilitator superfamily provides resistance to most 14- and 15-membered
macrolides but not 16-membered macrolides, lincosamides, or streptogramin B
[87]. Msr(A) has mainly been found in Staphylococcus spp. but also in Streptococci,
Corynebacterium, and Pseudomonas [95]. Mef(A) has been detected in strepto-
cocci including pneumococci and Group A and D organisms and also in Gram-
negative bacteria. Plasmid- or transposon-borne vga(A), vga(AV), vga(A)LC, vga(B),
vga(C), vga(E), and vga(E)V encode ABC transporters that export streptogramin A
antibiotics, while vga(A), vga(C), and vga(E) export lincosamides and pleuromuti-
lins as well. lsa(B) found on a plasmid in S. sciuri encodes an ABC transporter
active on clindamycin but probably not streptogramins [96]. lsa(E) on the other
hand confers resistance to lincosamides, streptogramins A, and pleuromutilins and
11 Transmissible Antibiotic Resistance 357
has been found in S. aureus and several species of Enterococcus [97]. OptrA in the
ABC transporter family confers resistance to oxazolidinones and phenicols and has
been found in E. faecium, E. faecalis, Staphylococcus sciuri, and Streptococcus suis
[98, 99]. Insertion sequence IS1216E has been implicated in the spread of optrA
among enterococcal plasmids and to the streptococcal chromosome [100, 101].
11.3.1 Mupirocin
11.3.2 Nitrofuran
11.3.3 Nitroimidazole
Plasmid and chromosomally located nim genes (A through J) [105, 106] encode
nitroimidazole reductases that convert 5-nitroimidazole to 5-aminoimidazole thus
blocking formation of toxic nitroso derivatives that are essential for bactericidal
activity by metronidazole and tinidazole [107]. The nim genes are usually tran-
scribed from promotors located within different insertion elements: IS1168 for
nimA-nimB, IS1169 for nimD, and IS1170 for nimC. nim plasmids characterized in
Bacteroides spp. have been nonconjugative (7.2 to 10–kb in size) but mobilizable by
larger plasmids or transferable by electroporation [10, 108]. Metronidazole resis-
tance in the B. fragilis group has been quite rare in the United States [109].
358 G. A. Jacoby
11.3.4 Oxazolidinone
Linezolid targets the peptidyltransferase center of the bacterial 50S ribosomal sub-
unit and, like other drugs with the same target, is blocked by the Cfr 23S rRNA
methyltransferase with an increase in the MIC of S. aureus from 0.5 to 8–16 μg/ml
[22]. The plasmid-mediated ABC transporter OptrA also exports linezolid with
typical MIC increases in S. aureus or enterococci of 2 to 8–16 μg/ml. Susceptibility
of tedizolid is affected to a lesser extent by OptrA [98] and not at all by Cfr [110].
11.3.5 Phenicol
11.3.6 Rifamycin
11.3.7 Sulfonamide
11.3.8 Tetracycline
More than 60 genes conferring resistance to tetracycline are known, most associated
with mobile elements that allow for gene exchange. Most common are genes encod-
ing energy-dependent efflux proteins. Others code for ribosomal protection proteins
or inactivating enzymes. Further details can be found in the web site maintained by
M. Roberts [88].
tet(X) encoding a NADP-dependent monooxygenase that requires oxygen to
degrade tetracycline was originally discovered as part of a conjugative transposon
in Bacteroides sp. where, lacking oxygen, it does not confer tetracycline resistance
on its host. Subsequently, it has been found in E. cloacae, K. pneumoniae, and other
360 G. A. Jacoby
11.3.9 Trimethoprim
Like that for sulfonamide, the strategy for plasmid-mediated trimethoprim resis-
tance is a resistant substitute for the trimethoprim target, dihydrofolate reductase. A
genetically diverse set of more than 30 dfr genes are known, mostly located in
Gram-negative organisms in integron cassettes or associated with ISCR elements.
The dfrA genes encode dimeric dihydrofolate reductases and include at least 26
alleles with dfrA1 and dfrA17 the most common [124]. dfrB genes encode smaller
trimeric enzymes of seven varieties [125, 126]. In addition, several dfr genes confer-
ring high-level trimethoprim resistance are known in Gram-positive organisms:
dfrA located on transposon Tn4003 in S. aureus and other Staphylococcus spp.,
dfrD found on small plasmids in staphylococci and Listeria monocytogenes [127],
11 Transmissible Antibiotic Resistance 361
dfrG which seems to be the most common variety in S. aureus [128], and dfrK found
on small plasmids in species of Staphylococcus [129].
The multidrug OqxAB pump effluxes trimethoprim [114].
Resistance genes existed long before the antibiotic era and have been found, for
example, in ancient permafrost, at the bottom of isolated caves, and in the gut
microbiome of a pre-Columbian mummy [130]. An origin in the organisms that
produce antibiotics with hence a need to be protected from their action is plausible
[131, 132]. Alternatively, organisms living in the same environment as antibiotic
producers, often soil, also need resistance genes to be able to compete. In other
cases, a housekeeping gene playing no apparent role in antibiotic production or
defense could have been adapted to this new use [133]. Sophisticated metagenomic
studies have found sequences in DNA from soil, oceans, and human feces 100%
identical to genes for resistance to aminoglycosides, β-lactams, glycopeptides,
phenicols, tetracycline, and other agents supporting an environmental source for
exchange of resistance genes, although the direction of this transfer is not always
obvious [134, 135].
Some examples of potential sources are listed in Table 11.4. Streptomyces gri-
seus, producer of streptomycin, makes phosphotransferases modifying the same-
OH groups as APH enzymes in pathogens. Streptomyces fradiae, producer of
neomycin, has acetyltransferases with the same specificity as the plasmid-mediated
AACs. Micromonospora purpurea, producer of gentamicin, has methylases that
modify 16S rRNA like ArmA and Rmt enzymes. Streptoalloteichus tenebrarius,
producer of tobramycin, has an rRNA methylase similar to the acquired NpmA
methylase. Bacillus licheniformis, producer of bacitracin, protects itself with a
BcrABC transporter similar to that determined by plasmids in E. faecalis.
Streptococcus (Saccharopolyspora) erythreus, producer of erythromycin, has 23S
rRNA methylases similar to acquired Erm methylases. Amycolatopsis orientalis
and Streptomyces toyocaensis, glycopeptide producers, have Van-like systems for
self-protection. Pseudomonas fluorescens, producer of mupirocin, has a resistant
isoleucyl-tRNA synthetase like the acquired Mup enzyme. Streptomyces rimosus,
an oxytetracycline producer, has ribosome protecting Tet(M) and Tet(O)-like pro-
teins. In each case, although the mechanisms are the same, the amino acid identity
between producer and plasmid-mediated resistance protein is too low to accommo-
date a direct transfer. Both could have a common ancestor, but convergent evolution
has not been ruled out.
Candidates with much closer sequence identity have been found for other resis-
tance genes. AAC(6′)-Ih, so far found only on plasmids in Acinetobacter spp., and
the more broadly distributed APH(3′)-VI are 99–100% identical to chromosomal
enzymes from particular species of Acinetobacter. QnrA, QnrC, and QnrS have
97–99% identical analogues in aquatic bacteria such as Shewanella and Vibrio spp.,
362 G. A. Jacoby
(continued)
11 Transmissible Antibiotic Resistance 363
while QnrB originates from Citrobacter and QnrE from Enterobacter spp. The
QepA efflux pump is related to ones in the order Burkholderiales such as species of
Pseudorhodoferax, while the OqxAB pump has close relatives in K. pneumoniae,
an organism that is also the likely source of fosA genes. Plasmid-mediated fosA3,
fosA5, and fosA6 are surrounded by truncated genes that also delimit the fosA gene
on the chromosome of strains of K. pneumoniae.
Many β-lactamases also have a clear pedigree. The origin of TEM-1 is not
known, but blaSHV-1 is a chromosomal as well as a plasmid gene in K. pneumoniae
and has been mobilized onto plasmids at least twice [165]. Close homologues of
blaCTX-M genes can be found on the chromosome of rarely pathogenic Kluyvera
species with blaCTX-M groups 1 and 2 related to genes of K. ascorbata; blaCTX-M
groups 8, 9, and 25 related to genes of K. georgiana; and blaCTX-M-37 related to
genes of K. cryocrescens [48]. Several plasmid-mediated OXA-type carbapene-
mases are close enough in sequence to chromosomal genes in Acinetobacter or
Shewanella spp. to make them likely progenitors. Plasmid-mediated AmpC-type
β-lactamases have close homologues in chromosomally determined enzymes of
various species. Enzymes in Chromobacterium spp. are as much as 76% identical
in amino acid sequence to KPC β-lactamase. NDM β-lactamase appears to be a
chimera formed, probably in A. baumannii [166], between the aminoglycoside
resistance gene aphA6 and a metallo-β-lactamase such as ElBla2 from Erythrobacter
litoralis [151].
Aminoglycoside nucleotidyltransferases are missing among aminoglycoside pro-
ducers. Several ANTs, however, share structural similarity and catalytic mechanism
with housekeeping enzymes such as DNA polymerase β, which has a similar rela-
tionship with lincosamide nucleotidyltransferases LnuA and LinB [167].
Chloramphenicol acetyl transferase has been found in species of Streptomyces, such
as S. albus, but not in Streptomyces venezuelae, the organism known to produce it
[152]. The cfr methyltransferase gene has homologues in Bacillus spp. Moraxella
spp. contain chromosomal mcr-like genes and also ISApl1 that is often associated
with them [154]. Species of Aeromonas have also been suggested for the origin of
mcr-like genes [52]. Mycobacterium sp. has a rifampin ribosyltransferase 63% iden-
tical to the plasmid Arr-2 enzyme. The origin of acquired sul and dfr genes is not
known.
364 G. A. Jacoby
Plasmids vary in size from a few to more than 500-kb. Core plasmid functions
include systems for maintenance (replication, stability, and copy control), for parti-
tioning between daughter cells at the time of bacterial division, and for mobility
(mobilizability and conjugal transfer). Small plasmids usually exist in multiple cop-
ies within the cell, while replication of larger ones is limited to a few copies.
Plasmids in Gram-negative organisms smaller than about 25-kb lack space for the
genetic machinery involved in mating pair formation but may be mobilized by a
conjugative helper plasmid. In Gram-positive organisms, many plasmids rely on
chemical signals mediated by oligopeptides for mating pair formation [168]. A third
group of plasmids, found across the size spectrum, is neither conjugative nor mobi-
lizable and is thought to rely on transformation or transduction for transfer [169].
Several plasmid classification schemes have been developed but can be compro-
mised by plasmid plasticity and recombination [170]. Historically, plasmids were
classified into Inc or incompatibility groups based on whether two plasmids were
unable to coexist stably in the same bacterial host, a property based on replication
specificity and copy control. Inc grouping is now tested with specific primers by
PCR-based replicon typing. In Enterobacteriaceae as of 2014, PCR-based replicon
typing could identify 24 distinct plasmid replicons with IncFII, IncA/C, IncL/M, and
IncI1 being the most common groups among typed resistance plasmids [171, 172].
Since the system was based on established Inc groups, it relates directly to the older
classifications. In Acinetobacter baumannii, plasmids have been subdivided into 19
GR types based on replicon sequences [173], and in Enterococcus and Staphylococcus
spp., more than 25 rep families have been defined [174, 175]. Alternatively, MOB
classification is based on variations in relaxase, an enzyme in the plasmid mobiliza-
tion system that nicks DNA at a specific site to produce a single-stranded substrate
for transfer. Use of degenerate primers recognizing the conserved N-terminal portion
of the relaxase gene allows five MOB types to be distinguished for plasmids of
γ-Proteobacteria [176]. Advantages of the MOB scheme include broad applicability
to plasmids of Acinetobacter and Pseudomonas spp. as well as Enterobacteriaceae,
and inclusion as well of integrative conjugative or mobilizable elements (ICE and
IME). A disadvantage is the limited resolution inherent in the current number of
MOB types. For some plasmid groups, a multilocus sequence typing (pMLST) sys-
tem based on 2–6 core plasmid genes is available for subtyping [172, 177]. With
neither replicon nor MOB typing, however, can all plasmids be classified at present,
so further evolution of plasmid taxonomy can be anticipated.
Plasmids vary in host range, due mainly to specificity of replication rather than
requirements of the conjugative system itself. In liquid culture, little if any mating
occurs between Gram-negative and Gram-positive bacteria or between strict anaer-
obes and facultative organisms. Plasmids found in Enterobacteriaceae are usually
transferable within that family, but only those belonging to a few Inc groups are
transferable to P. aeruginosa, which has its own set of plasmids transferable to other
Pseudomonas spp. [178]. Similarly, among plasmids in Gram-positive organisms,
11 Transmissible Antibiotic Resistance 365
some are species specific, while others have a broad host range and can be found in
both Enterococcus and Staphylococcus spp. [174].
Plasmids carry accessory functions besides antimicrobial resistance such as met-
abolic pathways, colonization and virulence factors, sex factor activity, bacteriocin
production and resistance, restriction/modification systems, biocide resistance, and
heavy metal ion resistance. Hughes and Datta examined over 400 enterobacterial
isolates collected in the “pre-antibiotic era” between 1917 and 1954 and found that
24% contained conjugative plasmids, many in the same Inc groups as contemporary
resistance plasmids, but none carried antibiotic resistance genes [179, 180]. How
then did naked plasmids and other mobile genetic elements acquire the genes for
resistance?
Figure 11.3 shows tools that bacteria have used to capture genes and incorporate
them into plasmids [181]. An insertion sequence (IS) is a 700 to 2500-bp DNA seg-
ment usually bounded by short, identical, sometimes imperfect inverted repeats (IRL
and IRR) and containing one or two transposase (tnp) genes that code for enzymes
that recognize the IRs and catalyze movement to another DNA site where integra-
tion generates direct repeats of 2 to14-bp depending on the IS [182]. As originally
defined, classical IS did not carry resistance genes but may locate to provide an
active promoter to activate an adjacent gene. Two copies of the same IS or related
ones can, however, surround a resistance gene creating a composite transposon that
can now move as a unit to another plasmid or chromosomal location. In particular,
820-bp IS26 is very common in multiresistance regions of plasmids and is a fre-
quent flanking element in composite transposons.
A few IS are unusual in that a single copy of the element can capture and move
an adjacent resistance gene. For example, 1656-bp ISEcp1 is bounded by 14-bp IRs
but on moving can utilize IRL and a new IRR distal to an adjacent gene which con-
sequently becomes part of the mobile unit. ISEcp1 has been implicated in the mobi-
lization of blaCTX-M, qnr, rmt, and other resistance genes [183]. ISCR elements differ
in moving by rolling circle replication and can incorporate larger segments of DNA
than ISEcp1. They are bounded not by IRs but by a downstream origin (oriIS) and
an upstream terminus (terIS) and do not create DR. Failure to recognize terIS allows
replication to continue into an adjacent gene which is thus mobilized. More than 20
ISCR elements have been distinguished based on the sequence of their transposases.
ISCR have been involved in the mobilization of virtually every class of antibiotic
resistance genes in Gram-negative organisms [184] .
An integron is an even more sophisticated system for capturing resistance genes
packaged in cassettes. A cassette contains the gene, often preceded by a ribosome
binding site but usually not a promoter, and an attC recombination site. The inte-
gron is made up of an intI gene encoding an integrase of the tyrosine recombinase
family, an attI recombination site, and a Pc promoter. The integrase catalyzes site-
specific recombination between the attI and attC sites capturing or releasing gene
cassettes which can be lined up in tandem, all under the control of the Pc promoter.
One hundred or so different cassettes are known carrying antibiotic resistance genes
and three main groups of integrons, classified on the basis of intI sequences. Class
1 integrons are the most common [185, 186].
366 G. A. Jacoby
IRL IRR
Insertion
(DR) tnp (DR) (DR) tnp tnp (DR)
sequence
Composite transposon
IRL IRR IRalt
DR DR
ISEcp1 (5) tnp (5) Terminal inverted repeat (IR)
ter lS ori lS
ISCR rcr
Gene
cassettes attC attC
attl
Integron
intl Cassette array
IRtnp IR
res
Tn3-subgroup DR DR
(5) tnpA tnpR (5)
transposon
IRtnp IR
res
Tn21-subgroup DR DR
tnpA tnpR mer or
transposon (5) (5)
IR IRt
DR
res
MIC in parallel DR
with ISCR (5) tniR tniQ tniB tniA (5)
intl1/attlt1
or mer
Fig. 11.3 Mobile elements involved in the capture and mobilization of antibiotic resistance genes
in Gram-negative bacteria. DR, direct repeat; tnp, tni, transposition functions; IRL and IRR, left and
right inverted repeats; IRalt, alternative IR; rcr, rolling circle replicase; oriIS, origin of ISCR ele-
ments; terIS, terminus of ISCR elements; attC, cassette recombination site; attI, integron recombi-
nation site; res, resolvase site. Elements that create DR are indicated and the DR length given,
except for IS, where the DR length varies for different elements. Tn21-subfamily transposons may
carry resistance genes as part of class 1 integrons inserted in or near the res site. MIC mobile inser-
tion cassette. (Adapted from [181])
The first moveable units on plasmids to be described were complex or unit trans-
posons, such as 4957-bp Tn3 encoding TEM-1 β-lactamase. Members of the Tn3
family are bigger than IS and include a transposase gene (tnpA), a resolvase gene
(tnpR), and a resolution (res) site as well as one or more resistance genes all bounded
by 38-bp IR. Movement is replicative and involves formation of a cointegrate
11 Transmissible Antibiotic Resistance 367
intermediate consisting of two copies of the transposon linking donor and recipient
molecules. Tn21 (19,671-bp) and Tn21-like transposons contain the same tnpA,
tnpR, and res genes in different orientations and often include mer genes for resis-
tance to Hg++. Integrons are often found within transposons and ISCR elements
within integrons.
Another transposable element termed mic for mobile insertion cassette is com-
posed of a resistance gene bracketed by IR but lacking an integrase/transposase,
which must be supplied in trans for the unit to move [187].
Integrative and conjugative elements (ICE) (also known as conjugative transpo-
sons) encode a phage-like integrase (int) that catalyzes recombination between an
attP site on the unit and the host chromosome. The chromosomal integration site is
typically specific for a particular ICE family. They are bounded by IRs, and most
ICE also encode an excisionase (xis) that removes the ICE from the chromosome as
a circular molecule. Transfer of the circular form to a new host requires plasmid-like
genes that control DNA transfer and genes that form a mating pair between donor
and recipient. Lack of the latter function produces an integrative mobilizable ele-
ment (IME) that requires the missing functions in trans for transfer by conjugation.
Both ICE and IME can contain transposons, ISs, and integrons. ICE and IME thus
share many of the functions of conjugative and nonconjugative plasmids except for
their preference for a chromosomal location. Surveys of prokaryotic genomes indi-
cate that ICE are more common than plasmids and mobilizable elements outnumber
self-conjugative ones [188]. They occur in Gram-positive, Gram-negative, and
strictly anaerobic organisms.
Genomic islands are gene clusters, some very large, fixed in the chromosome
with features that suggest a foreign origin. In one strain of A. baumannii, an 86-kb
resistance island containing a variety of ISs, transposons, integrases, and 45 resis-
tance genes has been identified and obviously allows for rapid development of pan-
resistance [189]. Genomic islands are also important in the evolution of
multiresistance in P. aeruginosa [190].
Phage particles carrying resistant genes (blaTEM, blaCTX-M, qnrA, qnrS, armA)
have been identified in wastewater or the human gut and constitute another class of
mobile elements [191, 192].
These elements can interact in various and complex ways. For example, a plas-
mid in K. pneumoniae carrying genes for both carbapenem, aminoglycoside, and
quinolone resistance was found to contain a complex transposon incorporating
blaKPC-3 inserted into a Tn3-family complex transposon with aminoglycoside resis-
tance genes and blaTEM-1 that also contained qnrB19 mobilized by ISEcp1 [193].
Because these mobile elements are built in modules, they can exchange, rearrange,
insert, delete, and recombine to generate remarkable diversity [194]. They have also
been doing this for a long time. Plasmid NR1 (also known as R100), one of the
original transmissible elements discovered in Japan in the 1950s, already contained
both integrons and transposons [195].
368 G. A. Jacoby
Major Points
Our adversaries turn out to be cleverer than we thought with an abundant reservoir of
resistance genes and a toolkit of efficient genetic devices to mobilize, incorporate,
and share them. Resistance is increasing, and one by one agents that we thought
could still be counted on have become less reliable. Knowledge of resistance mecha-
nisms has allowed the development of antibiotics and combinations effective for a
time against resistant pathogens, but bacteria will continue to evolve resistance.
Speedier diagnostic tests will facilitate choice of effective agents, but new antibiotics
and new ideas to combat resistance are urgently needed.
References
15. Martínez-Martínez L, Pascual A, Jacoby GA. Quinolone resistance from a transferable plas-
mid. Lancet. 1998;351(9105):797–9.
16. Mendoza C, Garcia JM, Llaneza J, Mendez FJ, Hardisson C, Ortiz JM. Plasmid-determined
resistance to fosfomycin in Serratia marcescens. Antimicrob Agents Chemother.
1980;18(2):215–9.
17. Bauernfeind A, Chong Y, Schweighart S. Extended broad spectrum ß-lactamase in Klebsiella
pneumoniae including resistance to cephamycins. Infection. 1989;17(5):316–21.
18. Knothe H, Shah P, Krcmery V, Antal M, Mitsuhashi S. Transferable resistance to cefotaxime,
cefoxitin, cefamandole and cefuroxime in clinical isolates of Klebsiella pneumoniae and
Serratia marcescens. Infection. 1983;11(6):315–7.
19. Watanabe M, Iyobe S, Inoue M, Mitsuhashi S. Transferable imipenem resistance in
Pseudomonas aeruginosa. Antimicrob Agents Chemother. 1991;35(1):147–51.
20. Rahman M, Noble WC, Cookson B. Mupirocin-resistant Staphylococcus aureus. Lancet.
1987;ii:387.
21. Le Goffic F, Capmau ML, Bonnet D, Cerceau C, Soussy C, Dublanchet A, Duval J. Plasmid-
mediated pristinamycin resistance. PAC IIA: a new enzyme which modifies pristinamycin
IIA. J Antibiot. 1977;30(8):665–9.
22. Long KS, Poehlsgaard J, Kehrenberg C, Schwarz S, Vester B. The Cfr rRNA methyltrans-
ferase confers resistance to phenicols, lincosamides, oxazolidinones, pleuromutilins, and
streptogramin A antibiotics. Antimicrob Agents Chemother. 2006;50(7):2500–5. https://doi.
org/10.1128/AAC.00131-06.
23. Xavier BB, Das AJ, Cochrane G, De Ganck S, Kumar-Singh S, Aarestrup FM, Goossens H,
Malhotra-Kumar S. Consolidating and exploring antibiotic resistance gene data resources.
J Clin Microbiol. 2016;54(4):851–9. https://doi.org/10.1128/JCM.02717-15.
24. Armstrong ES, Kostrub CF, Cass RT, Moser HE, Serio AW, Miller GH. Aminoglycosides.
In: Dougherty TJ, Pucci MJ, editors. Antibiotic discovery and development, vol. I. 1st ed.
New York/Dordrecht/Heidelberg/London: Springer; 2012. p. 229–69.
25. Shaw KJ, Rather PN, Hare RS, Miller GH. Molecular genetics of aminoglycoside resistance
genes and familial relationships of the aminoglycoside-modifying enzymes. Microbiol Rev.
1993;57(1):138–63.
26. Ramirez MS, Tolmasky ME. Aminoglycoside modifying enzymes. Drug Resist Updat.
2010;13(6):151–71. https://doi.org/10.1016/j.drup.2010.08.003.
27. Krause KM, Serio AW, Kane TR, Connolly LE. Aminoglycosides: an overview. Cold Spring
Harb Perspect Med. 2016;6(6). https://doi.org/10.1101/cshperspect.a027029.
28. Yokoyama K, Doi Y, Yamane K, Kurokawa H, Shibata N, Shibayama K, Yagi T, Kato H,
Arakawa Y. Acquisition of 16S rRNA methylase gene in Pseudomonas aeruginosa. Lancet.
2003;362(9399):1888–93. https://doi.org/10.1016/S0140-6736(03)14959-8.
29. Doi Y, Arakawa Y. 16S ribosomal RNA methylation: emerging resistance mechanism against
aminoglycosides. Clin Infect Dis. 2007;45(1):88–94. https://doi.org/10.1086/518605.
30. Wachino J, Shibayama K, Kurokawa H, Kimura K, Yamane K, Suzuki S, Shibata N, Ike
Y, Arakawa Y. Novel plasmid-mediated 16S rRNA m1A1408 methyltransferase, NpmA,
found in a clinically isolated Escherichia coli strain resistant to structurally diverse ami-
noglycosides. Antimicrob Agents Chemother. 2007;51(12):4401–9. https://doi.org/10.1128/
AAC.00926-07. AAC.00926-07 [pii]
31. Gauntlett JC, Gebhard S, Keis S, Manson JM, Pos KM, Cook GM. Molecular analysis of
BcrR, a membrane-bound bacitracin sensor and DNA-binding protein from Enterococcus
faecalis. J Biol Chem. 2008;283(13):8591–600. https://doi.org/10.1074/jbc.M709503200.
32. Matos R, Pinto VV, Ruivo M, Lopes Mde F. Study on the dissemination of the bcrABDR clus-
ter in Enterococcus spp. reveals that the BcrAB transporter is sufficient to confer high-level
bacitracin resistance. Int J Antimicrob Agents. 2009;34(2):142–7. https://doi.org/10.1016/j.
ijantimicag.2009.02.008.
33. Ambler RP, Coulson AFW, Frère J-M, Ghuysen J-M, Joris B, Forsman M, Levesque RC,
Tiraby G, Waley SG. A standard numbering scheme for the class A ß-lactamases. Biochem
J. 1991;276:269–70.
372 G. A. Jacoby
34. Bush K, Jacoby GA, Medeiros AA. A functional classification scheme for ß-lactamases and
its correlation with molecular structure. Antimicrob Agents Chemother. 1995;39(6):1211–33.
35. Bush K, Jacoby GA. Updated functional classification of ß-lactamases. Antimicrob Agents
Chemother. 2010;54(3):969–76. https://doi.org/10.1128/AAC.01009-09. AAC.01009-09
[pii]
36. Kernodle DS, Stratton CW, McMurray LW, Chipley JR, McGraw PA. Differentiation of
ß-lactamase variants of Staphylococcus aureus by substrate hydrolysis profiles. J Infect Dis.
1989;159(1):103–8.
37. East AK, Dyke KG. Cloning and sequence determination of six Staphylococcus aureus
β-lactamases and their expression in Escherichia coli and Staphylococcus aureus. J Gen
Microbiol. 1989;135(4):1001–15. https://doi.org/10.1099/00221287-135-4-1001.
38. Murray BE. β-lactamase-producing enterococci. Antimicrob Agents Chemother.
1992;36(11):2355–9.
39. ß-Lactamase classification and amino acid sequences for TEM, SHV and OXA extended-
spectrum and inhibitor resistant enzymes http:// www.lahey.org/studies.
40. Lartigue MF, Leflon-Guibout V, Poirel L, Nordmann P, Nicolas-Chanoine MH. Promoters
P3, Pa/Pb, P4, and P5 upstream from blaTEM genes and their relationship to ß-lactam resis-
tance. Antimicrob Agents Chemother. 2002;46(12):4035–7.
41. Philippon A, Arlet G, Jacoby GA. Plasmid-determined AmpC-type ß-lactamases. Antimicrob
Agents Chemother. 2002;46(1):1–11.
42. Jacoby GA. AmpC ß-lactamases. Clin Microbiol Rev. 2009;22(1):161–82. https://doi.
org/10.1128/CMR.00036-08. 22/1/161 [pii]
43. Nordmann P, Poirel L. The difficult-to-control spread of carbapenemase produc-
ers in Enterobacteriaceae worldwide. Clin Microbiol Infect. 2014. https://doi.
org/10.1111/1469-0691.12719.
44. Bontron S, Nordmann P, Poirel L. Transposition of Tn125 encoding the NDM-1 carbapen-
emase in Acinetobacter baumannii. Antimicrob Agents Chemother. 2016;60(12):7245–51.
https://doi.org/10.1128/AAC.01755-16.
45. Poirel L, Naas T, Nordmann P. Diversity, epidemiology, and genetics of class D ß-lactamases.
Antimicrob Agents Chemother. 2010;54(1):24–38. https://doi.org/10.1128/AAC.01512-08.
doi:AAC.01512-08 [pii]
46. Medeiros AA. Evolution and dissemination of ß-lactamases accelerated by generations of
ß-lactam antibiotics. Clin Infect Dis. 1997;24(Suppl 1):S19–45.
47. Castanheira M, Mendes RE, Jones RN, Sader HS. Changes in the frequencies of β-lactamase
genes among Enterobacteriaceae isolates in U.S. hospitals, 2012 to 2014: activity of
ceftazidime-avibactam tested against β-lactamase-producing isolates. Antimicrob Agents
Chemother. 2016;60(8):4770–7. https://doi.org/10.1128/AAC.00540-16.
48. D'Andrea MM, Arena F, Pallecchi L, Rossolini GM. CTX-M-type β-lactamases: a success-
ful story of antibiotic resistance. Int J Med Microbiol. 2013;303(6–7):305–17. https://doi.
org/10.1016/j.ijmm.2013.02.008.
49. Bevan ER, Jones AM, Hawkey PM. Global epidemiology of CTX-M β-lactamases: temporal
and geographical shifts in genotype. J Antimicrob Chemother. 2017:2145–55. https://doi.
org/10.1093/jac/dkx146.
50. Schwarz S, Johnson AP. Transferable resistance to colistin: a new but old threat. J Antimicrob
Chemother. 2016;71(8):2066–70. https://doi.org/10.1093/jac/dkw274.
51. Snesrud E, Ong AC, Corey B, Kwak YI, Clifford R, Gleeson T, Wood S, Whitman TJ,
Lesho EP, Hinkle M, McGann P. Analysis of serial isolates of mcr-1-positive Escherichia
coli reveals a highly active ISApl1 transposon. Antimicrob Agents Chemother. 2017;61(5).
https://doi.org/10.1128/AAC.00056-17.
52. Yin W, Li H, Shen Y, Liu Z, Wang S, Shen Z, Zhang R, Walsh TR, Shen J, Wang Y. Novel
plasmid-mediated colistin resistance gene mcr-3 in Escherichia coli. MBio. 2017;8(3).
https://doi.org/10.1128/mBio.00543-17.
11 Transmissible Antibiotic Resistance 373
71. Fernandes P. Fusidic acid: a bacterial elongation factor inhibitor for the oral treatment of acute
and chronic staphylococcal infections. Cold Spring Harb Perspect Med. 2016;6(1):a025437.
https://doi.org/10.1101/cshperspect.a025437.
72. O'Neill AJ, Larsen AR, Skov R, Henriksen AS, Chopra I. Characterization of the epidemic
European fusidic acid-resistant impetigo clone of Staphylococcus aureus. J Clin Microbiol.
2007;45(5):1505–10. https://doi.org/10.1128/JCM.01984-06.
73. Wang JL, Tang HJ, Hsieh PH, Chiu FY, Chen YH, Chang MC, Huang CT, Liu CP, Lau
YJ, Hwang KP, Ko WC, Wang CT, Liu CY, Liu CL, Hsueh PR. Fusidic acid for the treat-
ment of bone and joint infections caused by methicillin-resistant Staphylococcus aureus. Int
J Antimicrob Agents. 2012;40(2):103–7. https://doi.org/10.1016/j.ijantimicag.2012.03.010.
74. Farrell DJ, Mendes RE, Castanheira M, Jones RN. Activity of fusidic acid tested against
staphylococci isolated from patients in U.S. medical centers in 2014. Antimicrob Agents
Chemother. 2016;60(6):3827–31. https://doi.org/10.1128/AAC.00238-16.
75. Bennett AD, Shaw WV. Resistance to fusidic acid in Escherichia coli mediated by the type I
variant of chloramphenicol acetyltransferase. A plasmid-encoded mechanism involving anti-
biotic binding. Biochem J. 1983;215(1):29–38.
76. Murray IA, Cann PA, Day PJ, Derrick JP, Sutcliffe MJ, Shaw WV, Leslie AG. Steroid rec-
ognition by chloramphenicol acetyltransferase: engineering and structural analysis of a high
affinity fusidic acid binding site. J Mol Biol. 1995;254(5):993–1005. https://doi.org/10.1006/
jmbi.1995.0671.
77. Courvalin P. Vancomycin resistance in gram-positive cocci. Clin Infect Dis. 2006;42(Suppl
1):S25–34. https://doi.org/10.1086/491711.
78. Cattoir V, Leclercq R. Twenty-five years of shared life with vancomycin-resistant enterococci:
is it time to divorce? J Antimicrob Chemother. 2013;68(4):731–42. https://doi.org/10.1093/
jac/dks469.
79. Périchon B, Courvalin P. Glycopeptide resistance. In: Dougherty TJ, Pucci MJ, editors.
Antibiotic discovery and development, vol. 1. 1st ed. New York/Dordrecht/Heidelberg/
London: Springer; 2012. p. 515–42.
80. Holman TR, Wu Z, Wanner BL, Walsh CT. Identification of the DNA-binding site for the
phosphorylated VanR protein required for vancomycin resistance in Enterococcus faecium.
Biochemistry. 1994;33(15):4625–31.
81. Abadia Patino L, Courvalin P, Perichon B. vanE gene cluster of vancomycin-resistant
Enterococcus faecalis BM4405. J Bacteriol. 2002;184(23):6457–64.
82. Zeng D, Debabov D, Hartsell TL, Cano RJ, Adams S, Schuyler JA, McMillan R, Pace
JL. Approved glycopeptide antibacterial drugs: mechanism of action and resistance. Cold
Spring Harb Perspect Med. 2016. https://doi.org/10.1101/cshperspect.a026989.
83. Schwarz S, Shen J, Kadlec K, Wang Y, Michael GB, Fessler AT, Vester B. Lincosamides,
streptogramins, phenicols, and pleuromutilins: mode of action and mechanisms of resistance.
Cold Spring Harb Perspect Med. 2016. https://doi.org/10.1101/cshperspect.a027037.
84. Morar M, Pengelly K, Koteva K, Wright GD. Mechanism and diversity of the erythromy-
cin esterase family of enzymes. Biochemistry. 2012;51(8):1740–51. https://doi.org/10.1021/
bi201790u.
85. Korczynska M, Mukhtar TA, Wright GD, Berghuis AM. Structural basis for streptogramin
B resistance in Staphylococcus aureus by virginiamycin B lyase. Proc Natl Acad Sci U S A.
2007;104(25):10388–93. https://doi.org/10.1073/pnas.0701809104.
86. Bozdogan B, Berrezouga L, Kuo MS, Yurek DA, Farley KA, Stockman BJ, Leclercq R. A
new resistance gene, linB, conferring resistance to lincosamides by nucleotidylation in
Enterococcus faecium HM1025. Antimicrob Agents Chemother. 1999;43(4):925–9.
87. Fyfe C, Grossman TH, Kerstein K, Sutcliffe J. Resistance to macrolide antibiotics in public
health pathogens. Cold Spring Harb Perspect Med. 2016;6(10). https://doi.org/10.1101/csh-
perspect.a025395.
88. http://faculty.washington.edu/marilynr/.
89. Weisblum B. Insights into erythromycin action from studies of its activity as inducer of resis-
tance. Antimicrob Agents Chemother. 1995;39(4):797–805.
11 Transmissible Antibiotic Resistance 375
90. Dzyubak E, Yap MN. The expression of antibiotic resistance methyltransferase correlates
with mRNA stability independently of ribosome stalling. Antimicrob Agents Chemother.
2016;60(12):7178–88. https://doi.org/10.1128/AAC.01806-16.
91. Siberry GK, Tekle T, Carroll K, Dick J. Failure of clindamycin treatment of methicillin-
resistant Staphylococcus aureus expressing inducible clindamycin resistance in vitro. Clin
Infect Dis. 2003;37(9):1257–60. https://doi.org/10.1086/377501.
92. Boal AK, Grove TL, McLaughlin MI, Yennawar NH, Booker SJ, Rosenzweig AC. Structural
basis for methyl transfer by a radical SAM enzyme. Science. 2011;332(6033):1089–92.
https://doi.org/10.1126/science.1205358.
93. Deshpande LM, Ashcraft DS, Kahn HP, Pankey G, Jones RN, Farrell DJ, Mendes
RE. Detection of a new cfr-like gene, cfr(B), in Enterococcus faecium isolates recovered from
human specimens in the United States as part of the SENTRY antimicrobial surveillance
program. Antimicrob Agents Chemother. 2015;59(10):6256–61. https://doi.org/10.1128/
AAC.01473-15.
94. Tang Y, Dai L, Sahin O, Wu Z, Liu M, Zhang Q. Emergence of a plasmid-borne multidrug
resistance gene cfr(C) in foodborne pathogen Campylobacter. J Antimicrob Chemother.
2017;72(6):1581–8. https://doi.org/10.1093/jac/dkx023.
95. Ojo KK, Striplin MJ, Ulep CC, Close NS, Zittle J, Luis H, Bernardo M, Leitao J, Roberts
MC. Staphylococcus efflux msr(a) gene characterized in Streptococcus, Enterococcus,
Corynebacterium, and Pseudomonas isolates. Antimicrob Agents Chemother.
2006;50(3):1089–91. https://doi.org/10.1128/AAC.50.3.1089-1091.2006.
96. Kehrenberg C, Ojo KK, Schwarz S. Nucleotide sequence and organization of the mul-
tiresistance plasmid pSCFS1 from Staphylococcus sciuri. J Antimicrob Chemother.
2004;54(5):936–9. https://doi.org/10.1093/jac/dkh457.
97. Li XS, Dong WC, Wang XM, Hu GZ, Wang YB, Cai BY, Wu CM, Wang Y, Du XD. Presence
and genetic environment of pleuromutilin-lincosamide-streptogramin a resistance gene lsa(E)
in enterococci of human and swine origin. J Antimicrob Chemother. 2014;69(5):1424–6.
https://doi.org/10.1093/jac/dkt502.
98. Wang Y, Lv Y, Cai J, Schwarz S, Cui L, Hu Z, Zhang R, Li J, Zhao Q, He T, Wang D, Wang Z,
Shen Y, Li Y, Fessler AT, Wu C, Yu H, Deng X, Xia X, Shen J. A novel gene, optrA, that con-
fers transferable resistance to oxazolidinones and phenicols and its presence in Enterococcus
faecalis and Enterococcus faecium of human and animal origin. J Antimicrob Chemother.
2015;70(8):2182–90. https://doi.org/10.1093/jac/dkv116.
99. Fan R, Li D, Wang Y, He T, Fessler AT, Schwarz S, Wu C. Presence of the optrA gene in
methicillin-resistant Staphylococcus sciuri of porcine origin. Antimicrob Agents Chemother.
2016;60(12):7200–5. https://doi.org/10.1128/AAC.01591-16.
100. He T, Shen Y, Schwarz S, Cai J, Lv Y, Li J, Fessler AT, Zhang R, Wu C, Shen J, Wang
Y. Genetic environment of the transferable oxazolidinone/phenicol resistance gene optrA
in Enterococcus faecalis isolates of human and animal origin. J Antimicrob Chemother.
2016;71(6):1466–73. https://doi.org/10.1093/jac/dkw016.
101. Huang J, Chen L, Wu Z, Wang L. Retrospective analysis of genome sequences revealed
the wide dissemination of optrA in Gram-positive bacteria. J Antimicrob Chemother.
2017;72(2):614–6. https://doi.org/10.1093/jac/dkw488.
102. Gilbart J, Perry CR, Slocombe B. High-level mupirocin resistance in Staphylococcus aureus:
evidence for two distinct isoleucyl-tRNA synthetases. Antimicrob Agents Chemother.
1993;37(1):32–8.
103. Seah C, Alexander DC, Louie L, Simor A, Low DE, Longtin J, Melano RG. MupB, a new
high-level mupirocin resistance mechanism in Staphylococcus aureus. Antimicrob Agents
Chemother. 2012;56(4):1916–20. https://doi.org/10.1128/AAC.05325-11.
104. Antonov NK, Garzon MC, Morel KD, Whittier S, Planet PJ, Lauren CT. High prevalence
of mupirocin resistance in Staphylococcus aureus isolates from a pediatric population.
Antimicrob Agents Chemother. 2015;59(6):3350–6. https://doi.org/10.1128/AAC.00079-15.
105. Trinh S, Reysset G. Detection by PCR of the nim genes encoding 5-nitroimidazole resistance
in Bacteroides spp. J Clin Microbiol. 1996;34(9):2078–84.
376 G. A. Jacoby
106. Husain F, Veeranagouda Y, Hsi J, Meggersee R, Abratt V, Wexler HM. Two multidrug-
resistant clinical isolates of Bacteroides fragilis carry a novel metronidazole resistance nim
gene (nimJ). Antimicrob Agents Chemother. 2013;57(8):3767–74. https://doi.org/10.1128/
AAC.00386-13.
107. Carlier JP, Sellier N, Rager MN, Reysset G. Metabolism of a 5-nitroimidazole in suscep-
tible and resistant isogenic strains of Bacteroides fragilis. Antimicrob Agents Chemother.
1997;41(7):1495–9.
108. Reysset G, Haggoud A, Sebald M. Genetics of resistance of Bacteroides species to
5-nitroimidazole. Clin Infect Dis. 1993;16(Suppl 4):S401–3.
109. Snydman DR, Jacobus NV, McDermott LA, Ruthazer R, Golan Y, Goldstein EJ, Finegold SM,
Harrell LJ, Hecht DW, Jenkins SG, Pierson C, Venezia R, Yu V, Rihs J, Gorbach SL. National
survey on the susceptibility of Bacteroides fragilis group: report and analysis of trends in
the United States from 1997 to 2004. Antimicrob Agents Chemother. 2007;51(5):1649–55.
https://doi.org/10.1128/AAC.01435-06.
110. Sahm DF, Deane J, Bien PA, Locke JB, Zuill DE, Shaw KJ, Bartizal KF. Results of the sur-
veillance of tedizolid activity and resistance program: in vitro susceptibility of gram-positive
pathogens collected in 2011 and 2012 from the United States and Europe. Diagn Microbiol
Infect Dis. 2015;81(2):112–8. https://doi.org/10.1016/j.diagmicrobio.2014.08.011.
111. Schwarz S, Kehrenberg C, Doublet B, Cloeckaert A. Molecular basis of bacterial resistance
to chloramphenicol and florfenicol. FEMS Microbiol Rev. 2004;28(5):519–42. https://doi.
org/10.1016/j.femsre.2004.04.001.
112. Schwarz S, Werckenthin C, Kehrenberg C. Identification of a plasmid-borne chloramphenicol-
florfenicol resistance gene in Staphylococcus sciuri. Antimicrob Agents Chemother.
2000;44(9):2530–3.
113. Butaye P, Cloeckaert A, Schwarz S. Mobile genes coding for efflux-mediated antimicro-
bial resistance in Gram-positive and Gram-negative bacteria. Int J Antimicrob Agents.
2003;22(3):205–10.
114. Hansen LH, Jensen LB, Sorensen HI, Sorensen SJ. Substrate specificity of the OqxAB
multidrug resistance pump in Escherichia coli and selected enteric bacteria. J Antimicrob
Chemother. 2007;60(1):145–7.
115. Tribuddharat C, Fennewald M. Integron-mediated rifampin resistance in Pseudomonas aeru-
ginosa. Antimicrob Agents Chemother. 1999;43(4):960–2.
116. Almeida AC, Cavalcanti FL, Martins WM, Vilela MA, Gales AC, Morais Junior MA, Morais
MM. First description of KPC-2-producing Klebsiella oxytoca in Brazil. Antimicrob Agents
Chemother. 2013;57(8):4077–8. https://doi.org/10.1128/AAC.02376-12.
117. Shen P, Yi M, Fu Y, Ruan Z, Du X, Yu Y, Xie X. Detection of an Escherichia coli ST167 strain
with two tandem copies of blaNDM-1 encoded in the chromosome. J Clin Microbiol. 2016.
https://doi.org/10.1128/JCM.01581-16.
118. Bean DC, Livermore DM, Hall LM. Plasmids imparting sulfonamide resistance in Escherichia
coli: implications for persistence. Antimicrob Agents Chemother. 2009;53(3):1088–93.
https://doi.org/10.1128/AAC.00800-08.
119. Leski TA, Bangura U, Jimmy DH, Ansumana R, Lizewski SE, Stenger DA, Taitt CR, Vora
GJ. Multidrug-resistant tet(X)-containing hospital isolates in Sierra Leone. Int J Antimicrob
Agents. 2013;42(1):83–6. https://doi.org/10.1016/j.ijantimicag.2013.04.014.
120. Connell SR, Tracz DM, Nierhaus KH, Taylor DE. Ribosomal protection proteins and their
mechanism of tetracycline resistance. Antimicrob Agents Chemother. 2003;47(12):3675–81.
121. Warburton PJ, Amodeo N, Roberts AP. Mosaic tetracycline resistance genes encoding
ribosomal protection proteins. J Antimicrob Chemother. 2016;71(12):3333–9. https://doi.
org/10.1093/jac/dkw304.
122. Orth P, Schnappinger D, Hillen W, Saenger W, Hinrichs W. Structural basis of gene regulation
by the tetracycline inducible Tet repressor-operator system. Nat Struct Biol. 2000;7(3):215–
9. https://doi.org/10.1038/73324.
123. Fiedler S, Bender JK, Klare I, Halbedel S, Grohmann E, Szewzyk U, Werner G. Tigecycline
resistance in clinical isolates of Enterococcus faecium is mediated by an upregulation of
11 Transmissible Antibiotic Resistance 377
141. Brown S, Young HK, Amyes SG. Characterisation of OXA-51, a novel class D carbapen-
emase found in genetically unrelated clinical strains of Acinetobacter baumannii from
Argentina. Clin Microbiol Infect. 2005;11(1):15–23.
142. Rottman M, Benzerara Y, Hanau-Bercot B, Bizet C, Philippon A, Arlet G. Chromosomal
ampC genes in Enterobacter species other than Enterobacter cloacae, and ancestral asso-
ciation of the ACT-1 plasmid-encoded cephalosporinase to Enterobacter asburiae. FEMS
Microbiol Lett. 2002;210(1):87–92.
143. Bauernfeind A, Stemplinger I, Jungwirth R, Wilhelm R, Chong Y. Comparative characteriza-
tion of the cephamycinase blaCMY-1 gene and its relationship with other ß-lactamase genes.
Antimicrob Agents Chemother. 1996;40(8):1926–30.
144. Bauernfeind A, Stemplinger I, Jungwirth R, Ernst S, Casellas JM. Sequences of ß-lactamase
genes encoding CTX-M-1 (MEN-1) and CTX-M-2 and relationship of their amino acid
sequences with those of other ß-lactamases. Antimicrob Agents Chemother. 1996;40:509–13.
145. Papanicolaou GA, Medeiros AA, Jacoby GA. Antimicrob Agents Chemother. 1990;34:2200–9.
146. Horii T, Arakawa Y, Ohta M, Ichiyama S, Wacharotayankun R, Kato N. Plasmid-
mediated AmpC-type ß-lactamase isolated from Klebsiella pneumoniae confers resis-
tance to broad-spectrum ß-lactams, including moxalactam. Antimicrob Agents Chemother.
1993;37(5):984–90.
147. Fosse T, Giraud-Morin C, Madinier I, Labia R. Sequence analysis and biochemical char-
acterisation of chromosomal CAV-1 (Aeromonas caviae), the parental cephalosporinase of
plasmid-mediated AmpC 'FOX' cluster. FEMS Microbiol Lett. 2003;222(1):93–8.
148. Verdet C, Arlet G, Barnaud G, Lagrange PH, Philippon A. A novel integron in Salmonella
enterica serovar Enteritidis, carrying the blaDHA-1 gene and its regulator gene ampR, origi-
nated from Morganella morganii. Antimicrob Agents Chemother. 2000;44(1):222–5.
149. Girlich D, Karim A, Spicq C, Nordmann P. Plasmid-mediated cephalosporinase ACC-1 in
clinical isolates of Proteus mirabilis and Escherichia coli. Eur J Clin Microbiol Infect Dis.
2000;19:893–5.
150. Gudeta DD, Bortolaia V, Jayol A, Poirel L, Nordmann P, Guardabassi L. Chromobacterium
spp. harbour Ambler class A β-lactamases showing high identity with KPC. J Antimicrob
Chemother. 2016;71(6):1493–6. https://doi.org/10.1093/jac/dkw020.
151. Zheng B, Tan S, Gao J, Han H, Liu J, Lu G, Liu D, Yi Y, Zhu B, Gao GF. An unexpected simi-
larity between antibiotic-resistant NDM-1 and beta-lactamase II from Erythrobacter litoralis.
Protein Cell. 2011;2(3):250–8. https://doi.org/10.1007/s13238-011-1027-0.
152. Shaw WV, Hopwood DA. Chloramphenicol acetylation in Streptomyces. J Gen Microbiol.
1976;94(1):159–66. https://doi.org/10.1099/00221287-94-1-159.
153. Hansen LH, Planellas MH, Long KS, Vester B. The order Bacillales hosts functional
homologs of the worrisome cfr antibiotic resistance gene. Antimicrob Agents Chemother.
2012;56(7):3563–7. https://doi.org/10.1128/AAC.00673-12.
154. Kieffer N, Nordmann P, Poirel L. Moraxella species as potential sources of MCR-like poly-
myxin resistance determinants. Antimicrob Agents Chemother. 2017;61(6). https://doi.
org/10.1128/AAC.00129-17.
155. Poirel L, Rodriguez-Martinez JM, Mammeri H, Liard A, Nordmann P. Origin of plasmid-
mediated quinolone resistance determinant QnrA. Antimicrob Agents Chemother.
2005;49(8):3523–5.
156. Jacoby GA, Griffin CM, Hooper DC. Citrobacter spp. as a source of qnrB alleles. Antimicrob
Agents Chemother. 2011;55(11):4979–84. https://doi.org/10.1128/AAC.05187-11.
AAC.05187-11 [pii]
157. Hooper DC, Jacoby GA. Topoisomerase inhibitors: fluoroquinolone mechanisms of action
and resistance. Cold Spring Harb Perspect Med. 2016;6(9). https://doi.org/10.1101/cshper-
spect.a025320.
158. Albornoz E, Tijet N, De Belder D, Gomez S, Martino F, Corso A, Melano RG, Petroni A.
qnrE1, a member of a new family of plasmid-located quinolone resistance genes, originated
from the chromosome of Enterobacter species. Antimicrob Agents Chemother. 2017;61(5).
https://doi.org/10.1128/AAC.02555-16.
11 Transmissible Antibiotic Resistance 379
159. Yuan J, Xu X, Guo Q, Zhao X, Ye X, Guo Y, Wang M. Prevalence of the oqxAB gene complex
in Klebsiella pneumoniae and Escherichia coli clinical isolates. J Antimicrob Chemother.
2012;67(7):1655–9. https://doi.org/10.1093/jac/dks086.
160. Wachino J, Yamane K, Suzuki S, Kimura K, Arakawa Y. Prevalence of fosfomycin resistance
among CTX-M-producing Escherichia coli clinical isolates in Japan and identification of
novel plasmid-mediated fosfomycin-modifying enzymes. Antimicrob Agents Chemother.
2010;54(7):3061–4. https://doi.org/10.1128/AAC.01834-09.
161. Guo Q, Tomich AD, McElheny CL, Cooper VS, Stoesser N, Wang M, Sluis-Cremer N, Doi
Y. Glutathione-S-transferase FosA6 of Klebsiella pneumoniae origin conferring fosfomycin
resistance in ESBL-producing Escherichia coli. J Antimicrob Chemother. 2016;71(9):2460–
5. https://doi.org/10.1093/jac/dkw177.
162. Arthur M, Brisson-Noel A, Courvalin P. Origin and evolution of genes specifying resistance
to macrolide, lincosamide and streptogramin antibiotics: data and hypotheses. J Antimicrob
Chemother. 1987;20(6):783–802.
163. Marshall CG, Broadhead G, Leskiw BK, Wright GD. D-Ala-D-Ala ligases from glycopep-
tide antibiotic-producing organisms are highly homologous to the enterococcal vancomycin-
resistance ligases VanA and VanB. Proc Natl Acad Sci U S A. 1997;94(12):6480–3.
164. Hughes J, Mellows G, Soughton S. How does Pseudomonas fluorescens, the producing organ-
ism of the antibiotic pseudomonic acid A, avoid suicide? FEBS Lett. 1980;122(2):322–4.
165. Ford PJ, Avison MB. Evolutionary mapping of the SHV ß-lactamase and evidence for two
separate IS26-dependent blaSHV mobilization events from the Klebsiella pneumoniae chro-
mosome. J Antimicrob Chemother. 2004;54(1):69–75.
166. Toleman MA, Spencer J, Jones L, Walsh TR. blaNDM-1 is a chimera likely constructed in
Acinetobacter baumannii. Antimicrob Agents Chemother. 2012;56(5):2773–6. https://doi.
org/10.1128/AAC.06297-11.
167. Cox G, Stogios PJ, Savchenko A, Wright GD (2015) Structural and molecular basis for
resistance to aminoglycoside antibiotics by the adenylyltransferase ANT(2″)-Ia. mBio 6 (1).
doi:https://doi.org/10.1128/mBio.02180-14.
168. Dunny GM, Berntsson RP. Enterococcal sex pheromones: evolutionary pathways to com-
plex, two-signal systems. J Bacteriol. 2016;198(11):1556–62. https://doi.org/10.1128/
JB.00128-16.
169. Garcillan-Barcia MP, Alvarado A, de la Cruz F. Identification of bacterial plasmids based on
mobility and plasmid population biology. FEMS Microbiol Rev. 2011;35(5):936–56. https://
doi.org/10.1111/j.1574-6976.2011.00291.x.
170. Orlek A, Stoesser N, Anjum MF, Doumith M, Ellington MJ, Peto T, Crook D, Woodford
N, Walker AS, Phan H, Sheppard AE. Plasmid classification in an era of whole-genome
sequencing: application in studies of antibiotic resistance epidemiology. Front Microbiol.
2017;8:182. https://doi.org/10.3389/fmicb.2017.00182.
171. Carattoli A. Resistance plasmid families in Enterobacteriaceae. Antimicrob Agents
Chemother. 2009;53(6):2227–38. https://doi.org/10.1128/AAC.01707-08. AAC.01707-08
[pii]
172. Carattoli A, Zankari E, Garcia-Fernandez A, Voldby Larsen M, Lund O, Villa L, Moller
Aarestrup F, Hasman H. In silico detection and typing of plasmids using PlasmidFinder and
plasmid multilocus sequence typing. Antimicrob Agents Chemother. 2014;58(7):3895–903.
https://doi.org/10.1128/AAC.02412-14.
173. Bertini A, Poirel L, Mugnier PD, Villa L, Nordmann P, Carattoli A. Characterization and
PCR-based replicon typing of resistance plasmids in Acinetobacter baumannii. Antimicrob
Agents Chemother. 2010;54(10):4168–77. https://doi.org/10.1128/AAC.00542-10.
174. Jensen LB, Garcia-Migura L, Valenzuela AJ, Lohr M, Hasman H, Aarestrup FM. A classifi-
cation system for plasmids from enterococci and other Gram-positive bacteria. J Microbiol
Methods. 2010;80(1):25–43. https://doi.org/10.1016/j.mimet.2009.10.012.
175. Lozano C, Garcia-Migura L, Aspiroz C, Zarazaga M, Torres C, Aarestrup FM. Expansion of a
plasmid classification system for Gram-positive bacteria and determination of the diversity of
380 G. A. Jacoby
plasmids in Staphylococcus aureus strains of human, animal, and food origins. Appl Environ
Microbiol. 2012;78(16):5948–55. https://doi.org/10.1128/AEM.00870-12.
176. Garcillan-Barcia MP, de la Cruz F. Ordering the bestiary of genetic elements transmissible by
conjugation. Mob Genet Elements. 2013;3(1):e24263. https://doi.org/10.4161/mge.24263.
177. Hancock SJ, Phan MD, Peters KM, Forde BM, Chong TM, Yin WF, Chan KG, Paterson
DL, Walsh TR, Beatson SA, Schembri MA. Identification of IncA/C plasmid replication
and maintenance genes and development of a plasmid multilocus sequence typing scheme.
Antimicrob Agents Chemother. 2017;61(2). https://doi.org/10.1128/AAC.01740-16.
178. Jacoby GA. Resistance plasmids of Pseudomonas. In: Sokatch JR, editor. The bacteria, vol.
X. The biology of Pseudomonas. Orlando: Academic Press, Inc.; 1986. p. 265–93.
179. Hughes VM, Datta N. Conjugative plasmids in bacteria of the ‘pre-antibiotic’ era. Nature.
1983;302(5910):725–6.
180. Datta N, Hughes VM. Plasmids of the same Inc groups in Enterobacteria before and after the
medical use of antibiotics. Nature. 1983;306(5943):616–7.
181. Partridge SR. Analysis of antibiotic resistance regions in Gram-negative bacteria. FEMS
Microbiol Rev. 2011;35(5):820–55. https://doi.org/10.1111/j.1574-6976.2011.00277.x.
182. Siguier P, Gourbeyre E, Varani A, Ton-Hoang B, Chandler M. Everyman’s guide to bac-
terial insertion sequences. Microbiol Spectr. 2015;3(2):MDNA3-0030-2014. https://doi.
org/10.1128/microbiolspec.MDNA3-0030-2014.
183. Lartigue MF, Poirel L, Aubert D, Nordmann P. In vitro analysis of ISEcp1B-mediated mobi-
lization of naturally occurring β-lactamase gene blaCTX-M of Kluyvera ascorbata. Antimicrob
Agents Chemother. 2006;50(4):1282–6. https://doi.org/10.1128/AAC.50.4.1282-1286.2006.
184. Toleman MA, Bennett PM, Walsh TR. ISCR elements: novel gene-capturing systems of the
21st century? Microbiol Mol Biol Rev. 2006;70(2):296–316.
185. Labbate M, Case RJ, Stokes HW. The integron/gene cassette system: an active
player in bacterial adaptation. Methods Mol Biol. 2009;532:103–25. https://doi.
org/10.1007/978-1-60327-853-9_6.
186. Partridge SR, Tsafnat G, Coiera E, Iredell JR. Gene cassettes and cassette arrays in
mobile resistance integrons. FEMS Microbiol Rev. 2009;33(4):757–84. https://doi.
org/10.1111/j.1574-6976.2009.00175.x.
187. Guillard T, Grillon A, de Champs C, Cartier C, Madoux J, Bercot B, Lebreil AL, Lozniewski
A, Riahi J, Vernet-Garnier V, Cambau E. Mobile insertion cassette elements found in
small non-transmissible plasmids in Proteeae may explain qnrD mobilization. PLoS One.
2014;9(2):e87801. https://doi.org/10.1371/journal.pone.0087801.
188. Guglielmini J, Quintais L, Garcillan-Barcia MP, de la Cruz F, Rocha EP. The repertoire of
ICE in prokaryotes underscores the unity, diversity, and ubiquity of conjugation. PLoS Genet.
2011;7(8):e1002222. https://doi.org/10.1371/journal.pgen.1002222.
189. Fournier PE, Vallenet D, Barbe V, Audic S, Ogata H, Poirel L, Richet H, Robert C, Mangenot
S, Abergel C, Nordmann P, Weissenbach J, Raoult D, Claverie JM. Comparative genomics
of multidrug resistance in Acinetobacter baumannii. PLoS Genet. 2006;2(1):e7. https://doi.
org/10.1371/journal.pgen.0020007.
190. Roy Chowdhury P, Scott M, Worden P, Huntington P, Hudson B, Karagiannis T, Charles
IG, Djordjevic SP. Genomic islands 1 and 2 play key roles in the evolution of extensively
drug-resistant ST235 isolates of Pseudomonas aeruginosa. Open Biol. 2016;6(3). https://doi.
org/10.1098/rsob.150175.
191. Colomer-Lluch M, Jofre J, Muniesa M. Quinolone resistance genes (qnrA and qnrS) in bac-
teriophage particles from wastewater samples and the effect of inducing agents on pack-
aged antibiotic resistance genes. J Antimicrob Chemother. 2014;69(5):1265–74. https://doi.
org/10.1093/jac/dkt528.
192. Quiros P, Colomer-Lluch M, Martinez-Castillo A, Miro E, Argente M, Jofre J, Navarro
F, Muniesa M. Antibiotic resistance genes in the bacteriophage DNA fraction of human
fecal samples. Antimicrob Agents Chemother. 2014;58(1):606–9. https://doi.org/10.1128/
AAC.01684-13.
11 Transmissible Antibiotic Resistance 381
193. Rice LB, Carias LL, Hutton RA, Rudin SD, Endimiani A, Bonomo RA. The KQ element,
a complex genetic region conferring transferable resistance to carbapenems, aminoglyco-
sides, and fluoroquinolones in Klebsiella pneumoniae. Antimicrob Agents Chemother.
2008;52(9):3427–9. https://doi.org/10.1128/AAC.00493-08. AAC.00493-08 [pii]
194. Gillings MR, Paulsen IT, Tetu SG. Genomics and the evolution of antibiotic resistance. Ann
N Y Acad Sci. 2017;1388(1):92–107. https://doi.org/10.1111/nyas.13268.
195. Liebert CA, Hall RM, Summers AO. Transposon Tn21, flagship of the floating genome.
Microbiol Mol Biol Rev. 1999;63(3):507–22.
196. Jacoby GA, Munoz-Price LS. The new ß-lactamases. N Engl J Med. 2005;352(4):380–91.
197. Shaw KJ, Poppe S, Schaadt R, Brown-Driver V, Finn J, Pillar CM, Shinabarger D, Zurenko
G. In vitro activity of TR-700, the antibacterial moiety of the prodrug TR-701, against
linezolid-resistant strains. Antimicrob Agents Chemother. 2008;52(12):4442–7. https://doi.
org/10.1128/AAC.00859-08.
198. Arias CA, Mendes RE, Stilwell MG, Jones RN, Murray BE. Unmet needs and prospects for
oritavancin in the management of vancomycin-resistant enterococcal infections. Clin Infect
Dis. 2012;54(Suppl 3):S233–8. https://doi.org/10.1093/cid/cir924.
199. Shen Z, Ding B, Bi Y, Wu S, Xu S, Xu X, Guo Q, Wang M. CTX-M-190, a novel β-lactamase
resistant to tazobactam and sulbactam, identified in an Escherichia coli clinical isolate.
Antimicrob Agents Chemother. 2017;61(1). https://doi.org/10.1128/AAC.01848-16.
200. Livermore DM, Mushtaq S, Warner M, Vickers A, Woodford N. In vitro activity of cefepime/
zidebactam (WCK 5222) against gram-negative bacteria. J Antimicrob Chemother.
2017;72(5):1373–85. https://doi.org/10.1093/jac/dkw593.
201. Shields RK, Chen L, Cheng S, Chavda KD, Press EG, Snyder A, Pandey R, Doi Y,
Kreiswirth BN, Nguyen MH, Clancy CJ. Emergence of ceftazidime-avibactam resistance
due to plasmid-borne blaKPC-3 mutations during treatment of carbapenem-resistant Klebsiella
pneumoniae infections. Antimicrob Agents Chemother. 2017;61(3). https://doi.org/10.1128/
AAC.02097-16.
202. Davies J, Davies D. Origins and evolution of antibiotic resistance. Microbiol Mol Biol Rev.
2010;74(3):417–33. https://doi.org/10.1128/MMBR.00016-10. 74/3/417 [pii]
203. Bates J, Jordens JZ, Griffiths DT. Farm animals as a putative reservoir for vancomycin-
resistant enterococcal infection in man. J Antimicrob Chemother. 1994;34(4):507–14.
204. Kunin CM, Tupasi T, Craig WA. Use of antibiotics. A brief exposition of the problem and
some tentative solutions. Ann Intern Med. 1973;79(4):555–60.
205. Barlam TF, Cosgrove SE, Abbo LM, MacDougall C, Schuetz AN, Septimus EJ, Srinivasan A,
Dellit TH, Falck-Ytter YT, Fishman NO, Hamilton CW, Jenkins TC, Lipsett PA, Malani PN,
May LS, Moran GJ, Neuhauser MM, Newland JG, Ohl CA, Samore MH, Seo SK, Trivedi
KK. Implementing an antibiotic stewardship program: guidelines by the Infectious Diseases
Society of America and the Society for Healthcare Epidemiology of America. Clin Infect Dis.
2016;62(10):e51–77. https://doi.org/10.1093/cid/ciw118.
Chapter 12
Antibiotics and Resistance
in the Environment
Marilyn C. Roberts
12.1 Introduction
The discovery and use of antibiotics was one of the greatest public health achieve-
ments of the twentieth century. Antibiotics have saved millions of human and ani-
mal lives, reduced agricultural losses, and contributed to increased food production.
These agents have extended the lives of people with genetic conditions and have
become indispensible in modern medicine. The majority of antibiotics currently in
use were originally produced by living microbes that were then modified by man.
Antibiotics either inhibit growth of other microbes or kill them by interacting with
specific microbial targets. Most of the targets are unique to microbes, which has led
to the agents being safe enough to use with eukaryotic organisms.
In the mid-twentieth century, antibiotics became the foundation for treating bac-
terial infections in both humans and animals. Antibiotic-resistant bacteria [ARB]
and antibiotic resistance genes [ARGs] were recognized within a year after penicillin
was first used in humans, and soon after it was seen with agricultural use [1, 2].
ARB infections now contribute to thousands of deaths each year plus increased
morbidity and medical cost. Currently, it is estimated that ~10 million deaths due to
antibiotic-resistant infections occur each year; this number is expected to rise in
coming years [3]. In essence, antibiotic resistance has changed treatable infections
into untreatable diseases, thereby moving us closer to the “post-antibiotic era.”
Multidrug-resistant pathogens were first identified in the 1950s [4]. ARB were
initially limited to hospital settings and few outbreaks occurred; ARB were not seen
as a major concern for general community medicine. Today it is known that
antibiotic use in humans and agriculture results in increased antibiotic resistance in
M. C. Roberts (*)
Department of Environmental and Occupational Health Sciences, School of Public Health,
University of Washington, Seattle 98195, WA, USA
e-mail: [email protected]
countries. This is a major task given that in the USA there is no national surveillance
program for the most common pathogens across most states. Instead, the Centers for
Disease Control and Prevention (CDC) has used surveillance systems that focus on
nine representative states [17]. The European Union does a more comprehensive job
of covering their member states (http://ecdc.europa.eu/en/healthtopics/antimicro-
bial-resistance-and-consumption/antimicrobial_resistance/EARS-Net/Pages/
EARS-Net.aspx); other parts of the world have varying success with human surveil-
lance systems [17]. The problem is difficult, because ARGs are not randomly distrib-
uted among bacterial species. Data suggests that a clear link exists between bacterial
taxonomy and specific types of ARG [18, 19]. This phenomenon has been particu-
larly well documented with tetracycline resistance genes [20–22].
The environmental dissemination of ARGs and the development of ARGs are
thought to be primarily due to horizontal gene transfer. The most common way
bacteria exchange ARGs is by conjugation, which allows rapid transfer of ARGs
between species and genera within and between ecosystems [21]. However, our
knowledge is limited in regard to how the environment contributes to transmission
between the environment, wildlife, domesticated animals, plants, and humans. It is
critical when examining specific antibiotic resistance genes to know whether a
given gene is normally associated with a mobile element and whether that element
has a narrow or broad host range. Clearly a mobile element with a broad host range
will allow for wider transmission across multiple genera than a narrow host range
element [23, 24]. It is important to identify the specific ARGs associated with
specific bacterial species and/or genera within the environment. Durso et al. [18]
suggested that the same antibiotic resistance gene might have different risks for
environmental transmission that depends on the specific bacterial taxa within which
it is found. For example, if the bacteria are widely distributed among a variety of
environments, the ARGs associated with them are more likely to spread widely. If,
on the other hand, the bacteria have a limited environmental range, the ARGs will
tend to remain associated with them specifically. If they have a limited host range,
they may also not be widely distributed. It is equally important to know how ARGs
and ARB are distributed among human and animal populations and how these
ecosystems interact with various environments. Moreover, we need to know how
microbial distribution differs by region, nation, and worldwide [25]. Other issues
include the fact that most environmental studies look at a selected group of ARGs
by qPCR, which determines the presence or absence of particular genes [26], or
they use microbiome studies that usually do not look specifically for selected ARGs
[25]. Thus, environmental studies should include bacterial culturing, in addition to
molecular studies, to fully understand the distribution within the bacterial ecosystem
of various environments. The more comprehensive analysis is especially important
because many of the new ARGs are coming from the environment rather than from
either human or animal sources, which makes it difficult to know the bacterial
source of a given ARG (http://faculty.washington.edu/marilynr/).
Organized environmental surveillance of ARGs/ARB will hopefully allow iden-
tification of major gaps in our understanding of the forces that act on selection and
transmission of bacterial resistance. This effort in turn may lead scientists in direc-
386 M. C. Roberts
tions that could either slow or stop the march to a time when common infections and
minor injuries kill, as they did prior to the introduction and widespread use of anti-
biotics (this phenomenon is well illustrated by the recent spread of NDM-1
β-lactamase carrying bacteria [27]). It is clear that a global “One Health” approach
is needed in which animal and human usage and environmental contamination are
considered together, along with an understanding of how ARGs and ARB move
between the ecosystems.
Some antibiotics have non-bacterial effects on humans and animals and have been
used to treat non-bacterial conditions, especially skin diseases. A review of the non-
antibiotic properties of minocycline by Garrido-Mesa et al. [28] is a useful guide to
other properties that this antibiotic has and the non-bacterial conditions for which
minocycline is used as treatment. A 2013 paper [29] reported that minocycline
improves symptoms of fragile X syndrome when given to children and adolescents.
Another study explored the use of tetracyclines with cancer targets through a
randomized phase II trial [30]. In a third case, the macrolide azithromycin stimulated
immune and epithelial cell modulation of transcription factors AP-1 and NFκB with
subsequent delayed inhibitory effects on cell function and may cause lysosomal
accumulation of the macrolide with disruption of protein and lipid transport through
the Golgi apparatus and effects the surface receptor expression, including
macrophage phenotype changes and autophagy [31]. In addition, azithromycin
inhibits quorum sensing and biofilm formation by Pseudomonas aeruginosa, even
though the drug does not inhibit growth. Moreover, azithromycin, given
prophylactically, can reduce the incidence of ventilation-associated pneumonia
[31]. It is important to note that the use of antibiotics for non-bacterial conditions
increases the exposure of individuals’ microbiomes to the selective pressures
underlying the emergence of bacterial resistance (M. Roberts unpublished results).
It also increases the potential for environmental contamination by the antibiotic, its
residues, and ultimately selection of ARGs and ARB resistant to these antibiotics.
The One Health approach contrasts with the traditional practice of human and
animal medicine, which have been studied and practiced in isolation rather than as
part of an ecosystem. The environmental contribution to global health has not been
generally considered, or if studied, rarely, until recently in connection with the
health of man and/or animals. The world microbial ecosystem includes the micro-
biomes associated with each domain of life and the direct and indirect mixing of
12 Antibiotics and Resistance in the Environment 387
the different microbiomes that, in some cases, may lead to disease. The human-
animal interface is ancient, but it has expanded with the development of farming
animals and fish. It is a continuum of contacts and interactions that allow for bar-
rier breaches of pathogens to occur and an increased driver of infections. This is
illustrated by the estimation that ~75% of emerging infectious diseases in humans
over the last 20 years have been zoonotic, i.e., the pathogen spread from animals
or insects to people. In some cases, the pathogen becomes established and then
spreads within human populations. However, more commonly, there are recurrent
events of transmission from an animal/insect reservoir to humans, with limited
human-to-human transmission. An example of this situation is observed with the
Zika virus [32, 33]. Other examples include many foodborne bacterial infections,
such as those caused by E. coli O157:H7 and enterotoxigenic E. coli O114:H4. E.
coli O114:H4 caused a huge outbreak in 2011, which, besides causing death and
infections, created tension among EU members involving boycotts of vegetables
within the EU [34]. Dealing with emerging and reemerging infections that cross
species barriers not only impacts humans but also impacts livestock, pets, wildlife,
crops, and aquaculture. These pathogens can contaminate the environment, and in
worse cases, they may impact food resources and food security. The importance of
global ecological changes due to human impact on the environment and techno-
logical changes in society, along with important changes in how food is produced,
processed, and transported, combines to increase the potential risk of disease
transmission [32]. With environmental contamination as a major by-product of
these endeavors, changing the downward spiral of increasing global contamination
can only be addressed by improved communication, cooperation, and collabora-
tion across disciplines and the realization that there are multiple ways contamina-
tion can enter the food chain.
How a particular antibiotic can influence where and how antibiotic resistance
and ARB develop and spread from one domain to all three has been illustrated in the
literature. One good example is the development of vancomycin-resistant
Enterococcus faecium (VRE) in North America, the EU, and the rest of the world.
In Europe and other parts of the world, a vancomycin-related drug avoparcin was
used as a growth promoter in livestock. Over time, VRE developed in chickens and
swine to where it could be readily detected in processed meat [35]. Transmission of
VRE genes, or the intact bacterium, from animals to humans occurs in the EU
setting. Once VRE was established in livestock populations, farmers and those
slaughtering the VRE+ animals acquired VRE in their intestinal tracts. VRE
ultimately was isolated from hospitalized persons [36]. In contrast, avoparcin was
never used as a food additive in the USA. Early studies suggested that VRE was not
found in chickens in the USA, and there was little evidence to suggest transmission
of VRE in healthy adults prior to 2000 [35]. In contrast to the EU, which did not use
vancomycin heavily in the hospital setting, vancomycin was used extensively in US
hospital. The result was the emergence of VRE as a major nosocomial pathogen
within US hospitals [38]. This was due, in part, to the persistence of viable VRE on
contaminated surfaces within the hospital for weeks and even months. Rooms
housing patients colonized or infected with VRE were difficult to clean.
388 M. C. Roberts
Most studies on ARGs, over the last 70 years, have focused on clinically important
bacteria found in humans and animals. It is estimated that there are ~5 × 1030 bacteria
on earth, with only a small subset adapted to live either in or on humans and animals.
More striking is the estimate that <1% of the total number of bacteria in the world
have been cultured [42]. The natural world is rich in chemicals made by living
organisms and human activity – antibiotics are not the only compounds that have
influenced the evolution of microbiomes [43].
As stated above, knowing which type of bacterium carries a particular ARG can
be critical in designing studies of the environment. For example, many in the field
use E. coli as a model system for ARG carriage. Yet many ARGs found in E. coli are
unique to Gram-negative facultative aerobes and not found in anaerobes, other
Gram-negative bacteria, or Gram-positive species [22]. Thus, when E. coli is the
model, most acquired ARGs are associated with plasmids that independently
replicate and tend to have a host range limited to Gram-negative bacteria. In contrast,
many ARGs in Gram-positive and anaerobic species are on mobile elements that are
normally found in the chromosome; thus, they have a different host range dynamics
that can be much broader than classical Gram-negative plasmids. Therefore, by
looking for both classical Gram-negative and Gram-positive ARGs, researchers can
select ARGs that are likely to be most important for a particular ecosystem from a
large set that confers resistance to the same class of antibiotic. This is especially
important when molecular methods of detection are used, because only a limited
number of genes can be assayed, and if the rare genes are chosen, it will bias the
results leading to an unrealistic picture in that ecosystem.
Environmental studies are moving away from culturing bacteria, instead of
determining which ARGs are present by using either PCR or qPCR. These molecular
assays have now been used for direct detection of ARGs in food [44], animal feeding
facilities, and agricultural soils amended with manure [45] and as indicators for
water quality changes [46]. In these studies, known ARGs were used without
12 Antibiotics and Resistance in the Environment 389
determining their likely distribution in the particular sample source, which can lead
to biased results. For example, if only ARGs present in Gram-negative bacteria are
used for screening, then no information will be obtained regarding the Gram-
positive and anaerobic component of the sample source. Such studies have a very
limited ability to identify novel resistance genes.
To overcome the shortcomings of nucleotide sequence-dependent methods, soil
bacteria were screened for the ability to degrade or inactivate antibiotics. In one
study, strains were randomly isolated from 11 diverse rural and urban soils, and
they were then tested for the ability to utilize 18 different antibiotics as sole sources
of carbon and nitrogen [47]. Many of the bacteria were Burkholderia spp. and
Pseudomonas spp., which are naturally part of the soil microbiome and only rarely
cause disease. These bacteria could grow on antibiotic-supplemented media and
were resistant to multiple antibiotics at clinically relevant concentrations, suggest-
ing the presence of an unappreciated reservoir of antibiotic resistance genes in
these soils [47]. This work led to development of the functional metagenomic
approach (described below) in which the antibiotic resistome of an environment is
examined. This has led to identification of novel resistance genes in addition to
know ARGs [22].
Functional genomic studies have also been used to study a variety of microbial
environments [48]. This assay determines whether the cloned DNA can be expressed
and confer resistance when transferred into a host E. coli. When the cloned DNA
allows the host bacteria to grow in the presence of antibiotic-supplemented media,
the resistance-conferring DNA fragments can be sequenced and compared to known
ARGs. The Antibiotic Resistance Genes Database now lists ~20,000 potential
resistance genes( [49]; http://ardb.cbcb.umd.edu/), while the Comprehensive
Antibiotic Resistance Database [CARD] also has a large number of resistance genes
that can be used to screen sequences [50]. A variety of new potential ARGs have
been identified using this method ([51]; http://faculty.washington.edu/marilynr/).
One issue with these databases is they rely on GenBank information, which can be
confusing and inaccurate because the system allows authors to name their own
genes rather than going through a system. They also make it difficult to change
names. Thus many of the ARGs in GenBank do not have the correct nomenclature
for specific ARGs. Consulting other sources such as http://faculty.washington.edu/
marilynr/ and http://www.lahey.org/studies/ should be used to get the correct names
for tetracycline, for macrolide-lincosamide-streptogramin genes, and for β-lactamase
genes, respectively. Recent reviews are also good sources for the current
nomenclature [22].
The term antibiotic “resistome” is defined as the collection of all genes that can
either directly or indirectly contribute antibiotic resistance to its bacterial host [52].
Research groups have been examining the microbial resistome of natural and
clinical environments [51, 53–55]. Studies have looked for ARGs in samples linked
to human activity, such as food production [56, 57], polluted waterways, and
wastewater treatment plants [26]. Resistome studies suggest that environmental
bacteria may be antibiotic resistant by virtue of both previously characterized,
known genes and unknown resistance genes, mutations, and resistance genes on
390 M. C. Roberts
mobile elements [55, 56]. A number of new ARGs have been identified from these
studies; in most cases, the bacterial host is unknown (http://faculty.washington.edu/
marilynr/). One example of a ARG identified in a molecular study is the tet(43),
which encodes an efflux protein. It was isolated from metagenomic analysis of soil
taken from an apple orchard that had been repeatedly treated with streptomycin
[56]. It is unknown what type of bacterium actually carries tet(43), and little is
known about the distribution of this gene. Similarly, the nine genes [tet(47)-tet(55)],
identified more recently, code for enzymes that inactivate tetracycline. These were
identified, cloned, and sequenced from soil samples where functional metagenomic
analysis was done [58]. Again the bacterial sources of each of the genes are not
known. This later work increased the number of characterized enzymes that
inactivate tetracycline from 3 to 13 (http://faculty.washington.edu/marilynr/),
clearly showing that a variety of new ARGS may be present in environments.
Many bacteria, including environmental bacteria, encode β-lactamases, which
hydrolyze and inactivate β-lactam antibiotics (http://www.lahey.org/studies/). They
are the most widely distributed of all ARGs [3]. One example is ampC, which was
originally an inducible chromosomal cephalosporinase found in a variety of
Enterobacteriaceae. This gene has been found in opportunistic pathogens belong-
ing to normal intestinal floral of humans and animals, in bacterial species that nor-
mally live in either soil or water, and in both pathogenic and nonpathogenic bacteria
[59, 60]. It has been proposed that ampC originated in environmental bacteria. The
first AmpC-positive clinical strains were E. coli isolated in the 1940s, just as the
first antibiotics were being developed and used. In a host background that has porin
deficiencies, ampC, when expressed, confers carbapenem resistance due to
increased production of the AmpC β-lactamase. This increased production of the
AmpC β-lactamase is usually due to mutations that up-regulate expression of the
enzyme. Today the chromosomal AmpC β-lactamases are associated with plas-
mids, which was first noticed in the 1980s. These mobile plasmids often tend to be
large, and they carry multiple ARGs. Plasmid-mediated AmpC β-lactamases have
greatly expanded the host range of this group of enzymes that are now found in
epidemic human pathogens such as E. coli ST131. This E. coli strain has been iso-
lated from fresh vegetables, food-producing animals, fish farms, pets, and water
environments [61, 62].
Many ARGs are associated with soil antibiotic producers such as Streptomyces.
Some of these natural ARGs have the same mode of action as those found in
clinically resistant bacteria [3]. In the past, it was assumed that most environmental
bacteria were poorly adapted for life in humans and animals. However this idea is
changing, as progress in medical science allows severely immunocompromised
patients to live in the community where they can be infected with environmental
organisms. Other susceptible persons include those who have foreign objects
permanently implanted in their bodies and persons with various types of occupational
exposure [63]. Moreover, the distinction between environmental and non-
environmental bacteria has become difficult, because the mixing of the two sources
of bacteria has become increasingly common as human and agricultural
contamination of the environment has become widespread. Indeed, very few
12 Antibiotics and Resistance in the Environment 391
ecosystems around the world have not been touched by the activities of human
civilization – whether it is in polar regions or the Amazon jungle [64, 65]. The
continual mixing of environmental and non-environmental bacteria provides
opportunities for horizontal genetic exchange of ARGs between man, animal, and
environmental bacteria.
Antibiotics, antibiotic residues, ARBs, and ARGs move by water and wind [66],
wastewater treatment discharges [26, 67], biosolids, and manure applications [68],
isolated from recreational beaches [40]. They are also moved along with the
transportation of goods and people around the world [69, 70]. One result of this
movement has been the global spread of specific strains, such as Clostridium difficile
NAP1/027/BI [71]. C. difficile spores are robust, and they can survive in hospital
dust for extended time periods. C. difficile was originally classified as a nosocomial
pathogen. Today C. difficile is known to be a foodborne and community pathogen.
Similarly, 25 years ago, Acinetobacter baumannii was a rarely identified human
pathogen. At that time, Acinetobacter spp. were primarily found in the environment.
They are well adapted to grow at a wide range of temperature and pH values, can
use a variety of carbon and energy sources, and persist in both moist and dry
locations for extended time periods. Today multidrug-resistant A. baumannii is
considered an opportunistic pathogen that has become a major concern for military
trauma patients and causes infections that are very difficult to treat due to limited
therapeutic options [72].
12.5 Agriculture
Antibiotics are used for both human and agricultural activities for prevention and
treatment of infections. They are also used as food additives and growth promot-
ers in food production in the USA. However this widespread use is changing. In
June 2015, the US Food and Drug Administration published a final rule, known as
the veterinary feed directive (VFD), that extends the use of veterinary feed direc-
tives to an increased number of medically important antimicrobials used in food
animal production [73]. The rule became effective on October 1, 2015, and may
have impacted use and prescribing of medically important antibiotics in food ani-
mals in years prior to implementation because evidence supporting this idea is
derived from the experience of the EU. On July 1, 1989, an EU-wide ban on the
use of four growth-promoting antibiotics, spiramycin, tylosin, bacitracin zinc,
and virginiamycin, came into effect. The result of this ban was a dramatic drop in
the sales of antimicrobial growth-promoting agents. In 2006, the remaining anti-
biotics used as growth promoters (monensin, avilamycin, salinomycin, and flavo-
mycin) came under an EU-wide ban. It is projected that a further dramatic
decrease in sales of growth promoters will occur [74]. Therefore, it is hoped that
the US FDA ruling will reduce overall uses of antibiotics used annually in live-
stock raised in the USA.
392 M. C. Roberts
Human activity may directly influence the development of ARBs in built-up envi-
ronments. For example, several studies have recovered antibiotic-resistant E. coli
and S. aureus from air in homes that are enriched relative to samples from outside
of the home, even though the latter have higher bacterial levels. However, there was
variability both in study design and results [66]. Potentially, ARB may contaminate
the environment either directly, as occurs when manure is applied to enrich agricul-
tural fields, or indirectly due to sewage contamination of receiving waters where
the final effluent is deposited such as a river, lake, or ocean. The first description of
the tet(M) gene in Bacillus spp. and of TcrBacillus cereus strains carrying the
tet(M) gene, on a Tn916 mobile element, was found in animal manure and in fields
where the manure was spread. These results suggested that the presence of tet(M)-
carrying B. cereus in fields was a direct result of manure application to the soil.
12 Antibiotics and Resistance in the Environment 393
Whether tet(M)-carrying B. cereus will act as a donor and transfer the tet(M) gene
to either related B. anthracis or B. thuringiensis is unknown. However, some toxin-
encoding plasmids are shared among these three species [68].
An example of human wastes increasing ARB was illustrated by a 1980’s study
that observed three groups of wild baboons in Kenya. Two of the groups lived in
their natural habitat with either limited or no human contact; these groups had low
levels of antibiotic-resistant Gram-negative enteric bacteria. The third group lived
close to a tourist lodge that provided opportunities for daily contact with unprocessed
human refuse. From these animals, high levels of antibiotic-resistant Gram-negative
enteric bacteria were recovered with >90% tetracycline resistant [Tcr]. These results
suggested that contact with human refuse greatly increased the carriage of Tcr
bacteria in these wild primates [78]. Unfortunately, the surrounding environmental
bacteria were not sampled. One could speculate that the level of environmental
ARB was likely higher around the human refuse site than in areas where the two
other baboon groups lived in a more natural setting. Other studies have recovered
antibiotic-resistant E. coli from arctic and subarctic seals [79], wild boars [80], and
wild rabbits [81]. More recently, bacteria carrying extended-spectrum beta-
lactamases (ESBLs) have been isolated from water birds in remote locations [82].
Birds and wild animals can also be found feeding either in or around wastewater
treatment ponds, waste landfill sites, and septic tank discharges. Birds have the
potential for long-distance dissemination of ARB and ARGs to remote environments.
Such transmission sources may explain why ARB and ARGs can be found in
environments having little anthropogenic activity, such as the remote arctic [66].
In many studies it has been assumed that ARG flowed from humans and animals
to the environment. But in other cases, the use of antibiotics for food production has
created antibiotic-resistant bacteria in animals and farm environment that has spread
to man. One classic example of animal-to-human spread is the use of avoparcin in
farm animals in the EU [83]. Vancomycin-resistant enterococci [VRE] develop on
these farms, contaminating the farm ecosystem, including animal, environmental,
and human microbiomes. The VRE strains were passed to farm workers and families
living on the farm. In other cases, the plasmids carrying the vanA/vanB genes were
transmitted from animal to human enterococci [40]. In contrast, VRE development
in hospital settings in North America has occurred because vancomycin was
commonly used in hospitalized individuals but not in the general community
population. More recently, VRE strains have spread to the environment in the USA
where they are now isolated in a variety of settings, from recreational beaches to
birds to farms [39, 40, 84].
12.7 Aquaculture
As the taste for seafood and shellfish increases, the use of aquaculture around the
world, especially in Asia, has increased. Integrated aquaculture is a traditional
practice used by small-scale farmers in Asia. The fish are raised in ponds along
394 M. C. Roberts
with livestock. The livestock manure is used to feed the fish. This system allows
for mixing of ARGs and ARB, as well as for creating recombinant influenza
viruses [85]. Other parts of the world are less likely to practice integrated aquacul-
ture. Varying sizes of fish farms, both of the fresh water and marine type, grow
many types of fish for global export. Tilapia (Oreochromis niloticus) is among the
most cultured and internationally traded food fish, with an estimated 1.45 million
tons produced in China in 2013 [85].
ARGs are enriched in sediments below fish farms in Finland, even though selec-
tive pressure from antibiotics was low. A new study, which looked at 364 PCR
primer sets for detecting ARGs, mobile genetic elements, and 16S rRNA genes,
detected 28 genes in fish feces and fish farm sediments. The ARGs included
aminoglycoside (aadA1, aadA2), chloramphenicol (catA1), macrolide (mef(A),
msr(A)), sulfonamide (sul1), trimethoprim (dfrA1), and tetracycline ribosomal
protection genes [tet(32), tet(M), tet(O), tet(W)]. The same ARGs were found in fish
feces, suggesting that fish contribute to the ARG enrichment of the farm sediments
even though no antibiotic treatment of the fish in the farms was performed. Individual
farms had their own unique resistome compositions [86]. The Baltic Sea has no tide,
and water circulation is slow; thus, ARGs in the sediment underneath the fish pens
and up to 200 m from the fish farms were expected to reflect activity in the farm.
Muziasari et al. [86] concluded that their findings provide indirect evidence for the
hypothesis that selected ARGs are introduced into the sediment underneath fish
farms in the Northern Baltic Sea by farmed fish. The antibiotic concentrations in the
sediments were ~1–100 ng/g of sediment.
Tetracyclines have been used extensively in aquaculture, and Tcr bacteria have
been characterized from numerous sources, including fish pathogens and envi-
ronmental bacteria associated with finfish aquaculture from around the world
[87–91]. Tcr bacteria can be found in fish feed, in the sediment under the fish
pens, as well as in the water entering and leaving fresh water ponds [92]. Some
of the greatest diversity in Tcr genes has been identified in the aquaculture envi-
ronment. In one of our studies, ~40% of the Tcr bacteria isolated from Chilean
salmon fish farms carried previously unidentified Tcr genes, suggesting that the
diversity in the types of tet resistance genes is higher than routinely found in col-
lections from either man or other food animals [57]. Some of these bacteria were
later found to carry tet(39), while other genes are still unknown [93]. It is com-
mon to find previously characterized tet genes in new bacterial genera. Many of
these tet genes were not readily transferred under laboratory conditions, thereby
raising the question of how some of the genes are transferred to bacteria across
the world and from very different environments [57]. The diversity of type and
number of Tcr bacteria found in the aquaculture setting suggests that this may be
one environment where there is rapid evolution of Tcr bacteria and a hotspot for
ARG transmission.
12 Antibiotics and Resistance in the Environment 395
of ARB in outlets [102]. Hotspots of ARGs and ARB may be at WWTP outflows
where wastewater effluents are discharged into bodies of water. Thus, WWTP efflu-
ent may contribute to the dissemination of specific ARGs in the natural environment
[102, 103]. Similarly, other studies have shown that use of reclaimed water is a
reservoir for ARGs which increase in the soils after repeated irrigation with
reclaimed water. This has potential implications for human health [104].
Residual ARB/ARGs in the final effluent are normally deposited into bodies of
water where they can then be taken up by fresh water and marine wildlife and ulti-
mately cycle back to humans, land animals, and/or marine life [105]. Preliminary
data supports this hypothesis. High levels of ARGs were detected where WWTP
and CSO outflows discharged into Puget Sound WA USA (Dr. L. Rhodes personnel
communications). This release may be one reason why the southern resident killer
whales carry Gram-negative and Gram-positive resistant and multiresistant bacteria
in their respiratory tracts, as determined by cultures from exhaled breath samples
[106]. Similarly, antibiotic-resistant enterococci have been isolated from feces of
sea turtles, seabirds, and marine mammals from the southern coast of Brazil [105].
We conclude that the major waterways are sources and reservoirs of ARGs and ARB
worldwide.
Conventional wastewater treatment does reduce the total number of fecal bacte-
ria, but it does not necessarily reduce the fraction of ARGs/ARB present. Over
30 years ago, Walter and Vennes [107] showed that between 0.35% and 5% of the
coliforms from a domestic sewage system were resistant to ≥1 antibiotics, with
~75% of the multiple resistant strains capable of resistance gene transfer. Other
studies have isolated and characterized multidrug-resistant fecal coliforms and/or
enterococci from municipal water from multiple geographical areas [108, 109]. To
complicate the issue, wastewater effluent is now being used for urban landscaping
and to replenish urban aquifers. Thus what is in the effluent can make its way into
the drinking water ([110]; http://www.ocwd.com/what-we-do/water-reuse/).
The wastewater treatment process, besides increasing the abundance of ARGs
and the diversity of ARBs, may also provide selective pressure to increase the
diversity of antibiotic-resistant phenotypes and transmission of ARGs to new
bacterial species. These final WWTP products can ultimately contaminate a variety
of ecosystems, with particular impact on health through aquaculture, agriculture,
the human workers in these industries, and persons who consume these products
[104]. Occupational exposure risk to human and animal health is just now being
recognized [110]. ARGs and ARB have been found throughout the wastewater treat-
ment process, from raw influent, primary and secondary effluent, aeration tanks,
activated sludge, and residual biosolids [111, 112]. The biosolids represent the
majority of the biomass and thus the highest concentration of the ARGs and ARB
from the treatment process. This material is now widely used to enrich both urban
and agricultural environments. This can lead to environmental contamination of soil
and water and, most importantly, the potential to contaminate food consumed by the
general public [101]. This potential contamination needs to be considered when
12 Antibiotics and Resistance in the Environment 397
t rying to determine where the bacteria causing an outbreak were introduced into the
food product of interest. Moreover, knowing which specific ARG(s) are found in
which bacterial species and/or genera in WWTP products is critical when selecting
specific ARGs for regional, national, and international surveillance studies. It is
likely that there are common microbes in most WWTP systems (E. coli and entero-
cocci), but they may differ in the carriage of ARGs. Thus, unlike isolating bacteria,
which can also lead to biases, determining which ARGs are carried by specific bac-
teria is key to the success of future surveillance efforts using molecular methods.
The use of whole genome sequencing of WWTP products with emphasis on a large
number of different ARGs would be extremely useful in determining which suite of
ARGs should be examined when screening various components of the WWTP. This
needs to be done in different types of WWTP systems in both rural and urban setting
and both economically advantaged and disadvantaged nations.
Few studies have been conducted concerning metagenome analysis of plasmids
[113] or the microbiome of human sewage [114]. More research needs to be done to
determine whether there are variations by geographical location, seasons, and other
factors. Thus most studies in the literature that screen for specific ARGs and/or
resistant plasmids are inherently biased, because of the very large number of differ-
ent ARGs that are known. This bias should be taken into account when reviewing
the literature, including studies cited below.
A variety of studies have looked for specific ARGs in influent wastewater, after
primary settling, treated effluent, activated sludge, and treated biosolids. Most of
these studies select a small subset of the known antibiotic resistance genes
characterized by conferring resistance to a particular antibiotic class. For example,
one study looked at 10 different tet genes out of 59 that are known ([22]; http://
faculty.washington.edu/marilynr/). The genes included Gram-negative-specific
efflux genes tet(A), tet(E), and tet(G) and ribosomal protection genes tet(M), tet(O),
tet(Q), and tet(S) that are found in both Gram-negative and Gram-positive bacteria
[115] from the 18 samples over a 12-month period. The Gram-negative efflux tet(A)
and tet(C) genes were identified from all samples (n = 18). The other Gram-negative
efflux genes were isolated from 9–16 of the samples. The least common Gram-
negative efflux gene, tet(D), was identified in 9 of the 18 samples. The results are
not surprising, given the distribution of the different tet genes (http://faculty.wash-
ington.edu/marilynr/). It is interesting that most common efflux gene, tet(L), which
is isolated in similar numbers of Gram-negative (n = 19) and Gram-positive (n = 22)
bacteria, was not examined ([22]; http://faculty.washington.edu/marilynr/). This is a
common issue with many of the environmental sample studies published. The
authors selected tetracycline resistance genes to survey based on what previous
studies have used rather than base the work on abundance or on those most widely
distributed ARGs among different genera in the system they are studying. This
approach provides a significant bias to many of the environmental studies, including
those on WWTP products [101, 116].
398 M. C. Roberts
Bacteria carrying Tcr are widely distributed throughout the world. They have been
isolated from deep, subsurface trenches; in wastewater, surface water, and
groundwater, sediments, and soils; and in pristine environments untouched by
human civilization, such as penguins in Antarctica and seals in the Arctic [42, 56,
65, 79]. Seventeen (39%) of the 43 known tet genes including 12 (44%) of the
efflux, 3 (25%) of the ribosomal protection, and 2 (66%) of the enzymatic tet genes
are uniquely ascribed to environmental bacteria. Whether this is an accurate
representation, with some tet genes being truly “unique” to environmental bacteria,
or whether these genes have not been used in surveillance studies of either animal
or human bacteria is unclear. As of 2017, there are 59 tet genes with many of the
new genes not having been identified in specific bacteria (http://faculty.washington.
edu/marilynr/).
Five different resistance genes from Streptomyces, designated otr(A), otr(B),
otr(C), tcr3, and tet, have been identified in the chromosome of antibiotic-pro-
ducing strains. Today the otr(A) and otr(B) are now found in classical Bacillus
and Mycobacterium species that were primarily environmental bacteria but
recently have caused animal and human disease. It is possible that over time
other environmental “tet genes” will move into bacteria of clinical importance
and become associated with animals and man. For example, Clostridium spp. are
found in the environment, but they are also associated with the intestinal tract of
humans and animals. The tetA(P) and tetB(P) genes appear to be unique to
Clostridium spp. Other environmental genes included are the tet(V) gene that
has been found in Mycobacterium smegmatis, which is thought to be an environ-
mental bacterium; the tet(30) gene in Agrobacterium; the tet(33) that has been
found in environmental Arthrobacter and Corynebacterium spp.; the tet(35)
gene in Vibrio and Stenotrophomonas spp., which can cause human disease; and
the tet(41) gene in Serratia spp. which rarely causes human disease. The tet(42)
gene found in Bacillus, Microbacterium, Micrococcus, Paenibacillus, and
Pseudomonas spp. was isolated from a deep-sea trench. The tet(34) gene was
first described in Vibrio spp. and more recently identified in Pseudomonas spp.
and Serratia spp.(http://faculty.washington.edu/marilynr/). To determine if these
genes are truly environmental will require new surveillance studies in human
and animal bacteria to determine if some of genes currently assigned as “uniquely
environmental” are really only associated with bacteria isolated in the
environment.
Among the 97 genes that confer resistance to one or more macrolide, lincos-
amide, and streptogramin (MLS) antibiotics, there are a number of resistance genes
that are exclusively identified in the Streptomyces spp. including rRNA methylase
genes [erm(H), erm(I), erm(N), erm(O), erm(S), erm(U), erm(Z), erm(30), erm(31),
and erm(32)], ATP-binding transporters [car(A), ole(C), srm(B), tlr(C)], and a
major facilitator [lmr(A)] gene. Other rRNA methylases are found innately in vari-
12 Antibiotics and Resistance in the Environment 399
12.10 Conclusions
References
1. Feighner SD, Dashkevicz MP. Subtherapeutic levels of antibiotics in poultry feeds and their
effects on weight gain, feed efficiency, and bacterial cholyltaurine hydrolase activity. Appl
Environ Microbiol. 1987;53:331–6.
2. Marshall BM, Levy SB. Food animals and antimicrobials: impacts on human health. Clin
Microbiol Rev. 2011;24:718–33.
3. Davies J, Davies D. Origins and evolution of antibiotic resistance. Microbiol Mol Biol Rev.
2010;74:417–33.
4. Watanabe T. Fukasawa T. Episome-mediated transfer of drug resistance in Enterobacteriaceae.
1. Transfer of resistance factors by conjugation. J Bacteriol. 1961;81:669–78.
5. WHO Global Action Plan on Antimicrobial Resistance. 2015. [cited 2017 May 20.] Available
from: http://www.who.int/drugresistance/global_action_plan/en/
6. Collignon PC, Conly JM, Andremont A, McEwen SA, Aidara-Kane A. World Health
Organization ranking of antimicrobials according to their importance in human medicine:
a critical step for developing risk management strategies to control antimicrobial resistance
from food animal production. Clin Infect Dis. 2016;63(8):1087–93.
7. Boucher HW, Bakken JS, Murray BE. The United Nations and the urgent need for coordinated
global action in the fight against antimicrobial resistance. Ann Intern Med. [serial online].
2016. [cited 2017 May 4.] Available from: www.annals.org
8. Center for Diseases Dynamics, Economics & Policy. The State of the World’s Antibiotics
2015. [cited 2017 May 4.] Available from: http://www.cddep.org/publications/state_worlds_
antibiotics_2015#sthash.3wasnwH1.dpbs
402 M. C. Roberts
9. Wall, BA, Mateus A, Marshall L, Pheiffer DU. Drivers, dynamics and epidemiology of
antimicrobial resistance in animal production. Food and Agriculture Organization of the
United Nations. Rome, 2016. [cited 2017 May 4.] Available from: http://www.fao.org/3/a-
i6209e.pdf
10. Collignon P. The importance of a one health approach to preventing the development and
spread of antibiotic resistance. Cur Top Microbiol Immunol. 2013;366:19–36.
11. Center for Disease Control. Antibiotic Resistant Threats in the United States. 2013. [cited
2017 May 4.] Available from: http://www.cdc.gov/drugresistance/threat-report-2013/pdf/ar-
threats-2013-508.pdf
12. WHO Antibiotic Resistance Global Report of Surveillance 2014. [cited 2017 June 2].
Available from: http://apps.who.int/iris/bitstream/10665/112642/1/9789241564748_eng.
pdf?ua=1
13. Van Boeckel TP, Brower C, Gilbert M, et al. Global trends in antimicrobial use in food ani-
mals. PNAS. 2015;112:5649–54.
14. Berendonk TU, Manaia CM, Merlin C, et al. Tackling antibiotic resistance: the environmen-
tal framework. Nat Rev Microbiol. 2015;13:310–7.
15. National Strategy for Combating Antibiotic-Resistant Bacteria. 2014. [cited 2017 June 2.]
Available from: https://www.whitehouse.gov/sites/default/files/docs/carb_national_strategy.
pdf
16. Taylor, J, Hafner M, Yerushalmi E, et al. Estimating the economic costs of antimicrobial
resistance: Model and results. RAND. 2014 [cited 2017 May15] Available from: http://www.
rand.org/content/dam/rand/pubs/research_reports/RR900/RR911/RAND_ RR911.pdf
17. Tillotson GS. Where in the world? The role of geography in antibiotic resistance and the
potential impact in pulmonary infections. Postgrad Med. 2016;128(5):449–50.
18. Durso LM, Miller DN, Wienhold BJ. Distribution and quantification of antibiotic resis-
tant genes and bacteria across agricultural and non-agricultural metagenomes. PLoS ONE.
[serial online] 2012. [cited 2017 June 2] Available from: https://doi.org/10.1371/journal.
pone.0048325
19. Roberts M C, Schwarz S, Aarts H. Acquired antibiotic resistance genes: an overview. Front
Microbiol. [serial online] 2012 [cited. 2017 May 30] 3:384; 1–17. Available from: http://www.
frontiersin.org/Antimicrobials,_Resistance_and_Chemotherapy/10.3389/fmicb.2012.00384/full
20. Chopra I, Roberts MC. Tetracycline antibiotics: mode of action, applications, molecular biol-
ogy and epidemiology of bacterial resistance. Microbiol Mol Biol Rev. 2001;65:232–60.
21. Roberts MC. Mechanisms of bacterial antibiotic resistance and lessons learned from environ-
mental tetracycline resistant bacteria. In: Keen P, Montforts M, editors. Antibiotic resistance
in the environment: John Wiley & Sons, Inc.; 2012. p. 93–121.
22. Roberts MC, Schwarz S. Tetracycline and phenicol resistance genes and mechanisms: impor-
tance for agriculture, the environment and humans. J Environ Quality. 2016;45:576–92.
23. Roberts MC, Knapp JS. Host range of the conjugative 25.2 Mdal tetracycline resistance plas-
mid from Neisseria gonorrhoeae. Antimicrob Agents Chemother. 1988;32:488–91.
24. Roberts MC, Knapp JS. Transfer of β-lactamase plasmids from Neisseria gonorrhoeae to
Neisseria meningitidis and commensal Neisseria species by the 25.2-Megadalton conjugative
plasmid. Antimicrob Agents Chemother. 1988;32:1430–2.
25. Hu Y, Yang X, Lv N, et al. The bacterial mobile resistome transfer network connecting the
animal and human microbiomes. Appl Environ Microbiol. 2016;82(22):6672–81.
26. Rizzo L, Manaia C, Merlin C, et al. Sci Total Environ. 2013;447:345–60.
27. Nordmann P, Naas T, Poirel L. Global spread of carbapenemase-producing Enterobacteri
aceae. Emerg Infect Dis. 2011;17:1791–8.
28. Garrido-Messa N, Zarzuelo A, Galvez J. Minocycline: far beyond an antibiotic. Br
J Pharmacol. 2013;169:337–52.
29. Leigh MJS, Nguyen DV, My Y, et al. A randomized double-blind, placebo-controlled trial of
minocycline in children and adolescents with Fragile X syndrome. J Devel Behav Pediatr.
2014;34:147–55.
12 Antibiotics and Resistance in the Environment 403
30. Dezube BJ, Krown SE, Lee JY, et al. Randomized phase II trial of matrix metalloproteinase
inhibitor COL-3 in AIDS-related Kaposi’s sarcoma: an AIDS Malignancy Consortium Study.
J Clin Oncol. 2006;24:1389–94.
31. Parnham MJ, Haber VE, Giamarellos-Bourboulis EJ, Perletti G, Verleden GM, Vos
R. Azithromycin: mechanism of action and their relevance for clinical applications.
Pharmacol Ther. 2014;143:225–45.
32. Lindahl JF, Grace D. The consequences of human actions on risks for infectious disease:
a review. Infect Ecol Epidem. [serial online] 2015 [cited 2017 May 25]; Available from:
5:30048-https://doi.org/10.3402//lee.v5.30048
33. USDA Animal and Plant Health Inspection service Veterinary services 2015 One Health.
What is One Health? [cited 2017 May 20.] Available from: https://www.aphis.usda.gov/ani-
mal_health/one_health/downloads/one_health_info_sheet.pdf.
34. Heymann DL, Dixon M. The value of the one health approach: shifting from emergency
response to prevention of zoonotic disease threats at their source. In: Atlas RM, Maloy S, edi-
tors. One health people, animals, and the environment. Washington, DC: American Society
for Microbiology; 2014. p. 17–31.
35. Cetinkaya Y, Falk P, MayHall CG. Vancomycin-resistant enterococci. Clin Microbiol Rev.
2000;13:686–707.
36. van den Bogaard AE, Willems R, London N, Top J, Stobberingh EE. Antibioitc resistance
of faecal enterococci in poultry, poultry farmers and poultry slaughterers. J Antimicrob
Chemoth. 2002;49:497–505.
37. Eckstein BC, Adams DA, Eckstein EC, et al. Reduction of Clostridium difficile and vanco-
mycin-resistant Enterococcus contamination of environmental surfaces after an intervention
to improve cleaning methods. BMC Infect Dis. [serial online] 2007 [cited 2017 May 20];7:61
Available from:http://www.biomedcentral.com/1471-2334/7/61
38. Rice LB. The complex dynamics of antimicrobial activity in the human gastrointestinal tract.
Trans Am Clin Climatol Assoc. 2013;124:123–32.
39. Oravcova V, Zurek L, Townsend A, et al. American crows as carriers of vancomycin-resistant
enterococci with vanA gene. Environ Microbiol. 2014;16:939–49.
40. Roberts MC, Soge OO, Giardino MA, Mazengia E, Ma G, Meschke JS. Vancomycin- resis-
tant Enterococcus spp. in environments from the west coast of the USA. J Appl Microbiol.
2009;107:300–7.
41. Roberts MC, No DB, Marzluff JM, DeLap JH, Turner R. Vancomycin resistant Enterococcus
spp. from crows and their environment in Metropolitan Washington State, USA: is there a cor-
relation between VRE positive crows and the environment? Vet Microbiol. 2016;194:48–54.
42. Kümmerer K. Resistance in the environment. J Antimicrob Chemother. 2004;54:311–20.
43. Wright GD. The antibiotic resistome: the nexus of chemical and genetic diversity. Nat Rev
Microbiol. 2007;5:175–86.
44. Guarddon M, Miranda JM, Rodriquez JA, et al. Real-time polymerase chain reaction for
the quantitative detection of tetA and tetB bacterial tetracycline resistance genes in food. Int
J Food Microbiol. 2011;146:284–9.
45. Peng S, Wang Y, Zhou B, et al. Long-term application of fresh and composted manure
increase tetracycline resistance in the arable soil of eastern China. Sci Total Environ.
2015;506-507:279–86.
46. Harnisz M, Korzeniewska E, Ciesielski S, et al. tet genes as indicators of changes in the water
environment: relationships between culture-dependent and culture-independent approaches.
Sci Total Environ. 2015;505:704–11.
47. Dantas G, Sommer MOS, Oluwasegun RD, et al. Bacterial subsisting on antibiotics. Science.
2008;320:100–3.
48. Moore AM, Patel S, Forsberg KJ, et al. Pediatric fecal microbiota harbor diverse and novel
antibiotic resistance genes. PLoS One. [serial online] 2013 [cited 2017 June 1]; 8:e78822
Available from: http://www.plosone.org/article/fetchObject.action?uri=info:doi/10.1371/
journal.pone.0078822&representation=PDF
404 M. C. Roberts
49. Lin B, Pop M. ARDB-Antibiotic Resistance Genes Database. Nucleic Acids Res. 2009 [cited
2017 May 11]; Jan:37(Database issue):D443-7 Available from:http://ardb.cbcb.umd.edu/
50. Jia B, Raphenya AR, Alcock B, et al. CARD 2017: expansion and model-centric curation of
the comprehensive antibiotic resistance database. Nucl Acids Res. [serial online] 2017; [cited
June 1]; 45:D566–D573 Available from: doi:https://doi.org/10.1093/nar/gkw1004.
51. D’Costa V, McGrann KM, Hughes DW, et al. Sampling the antibiotic resistome. Science.
2006;311:374–7.
52. Perry JA, Wright GE. The antibiotic resistance “mobiolime”: Searching for the link between
environment and clinic. Front Microbiol. [serial online] 2013 [cited 2017 June 2];4:138
Available from: doi: https://doi.org/10.3389/fmicb.2013.00138.
53. Forsberg KJ, Reyes A, Wang B, et al. The shared antibiotic resistome of soil and human
pathogens. Science. 2012;337:1107–11.
54. Perry JA, Westman EL, Wright GE. The antibiotic resistome: what’s new? Curr Opin
Microbiol. 2014;21:45–50.
55. Thaker M, Spanogiannopoulos P, Wright GD. The tetracycline resistome. Cell Mol Life Sci.
2010;67:419–31.
56. Donato JJ, Moe LA, Converse BJ, et al. Metagenomic analysis of apple orchard soil reveals
antibiotic resistance genes encoding predicted bifunctional proteins. Appl Environ Microbiol.
2010;76:4396–401.
57. Miranda CD, Kehrenberg C, Ulep C, et al. Diversity of tetracycline resistance genes in bac-
teria from Chilean salmon farms. Antimicrob Agents Chemother. 2003;47:883–8.
58. Forsberg KJ, Patel S, Wencewicz TA, Dantas G. The tetracycline destructases: a novel family
of tetracycline-inactivation enzymes. Chem Biol. 2015;22:888–97.
59. de Oliveria DV, Nunes LS, Barth AL, Van Der Sand. Genetic background of β-lactamases in
Enterobacteriaceae isolates from environmental samples. Microb Ecol. [serial online] 2017
[cited 2017 May 16.] Available from:doi:https://doi.org/10.1007/s00248-017-0970-6
60. Capkin E, Terzi E, Altinok I. Occurrence of antibiotic resistance genes in culturable bacte-
ria isolated from Turkish trout farms and their local aquatic environment. Dis Aquat Org.
2015;114:127–37.
61. Guyomard-Rabenirina S, Darton C, Falord M, et al. Resistance to antimicrobial drugs in
different surface waters and wastewaters of Guadelopupe. PLoS One. [serial online] 2017.
[cited 2017 May 16]; Mar 2;12(3):e0173155. Available from: doi:https://doi.org/10.1371/
journal.pone.0173155.
62. van Hoek AHAM, Schouls L, van Santen MG, Florihn A, de Greeff SC, van Duijkeren
E. Molecular characteristics of extended-spectrum cephalosporin-resistant Enterobacteriaceae
from humans in the community. PLoS One. [serial online] 2015 [cited 2017 May 16]:
10(6):e0129085. Available from; doi:https://doi.org/10.1371/journal.pne.0129085
63. Rowlinson M-C, Bruckner DA, Hinnebusch C, et al. Clearance of Cellulosimicrobium cel-
lulans bacteremia in a child without central venous catheter removal. J Clin Microbiol.
2006;44:2605–54.
64. Pallecchi L, Bartonloni A, Riccobono E, et al. Quinolone resitance in absence of selective
pressure: The experience of a very remote community in the Amazon forest. PLoS Neg
Trop Dis. [serial online] 2012 [cited 2016 July 10]; 6:e1790 Available from: doi: https://doi.
org/10.1371/journal.pntd.0001790
65. Rahman MH, Sakamoto KQ, Nonaka L, et al. (2008). Occurrence and diversity of tetra-
cycline tet(M) in enteric bacteria of Antarctic Adelie penguins. J Antimicrob Chemother.
2008;62:627–8.
66. Allen HK, Donato J, Wang HH, et al. Call of the wild: antibiotic resistance gene in natural
environments. Nat Rev Microbiol. 2010;8:251–9.
67. Berglund B, Fick J, Lindgren PE. Urban wastewater effluent increases antibiotic resis-
tance genes concentrations in a receiving northern European river. Environ Toxicol Chem.
2015;34:192–6.
12 Antibiotics and Resistance in the Environment 405
68. Agersø Y, Jensen LB, Givskov M, et al. The identification of a tetracycline resistance gene
tet(M), on a Tn916-like transposon, in the Bacillus cereus group. FEMS Microbiol Lett.
2002;214:251–6.
69. Dobbs FC, Goodrich AL, Tomson FS III, et al. Pandemic serotypes of Vibrio cholerae iso-
lated from ships’ ballast tanks and coastal waters: assessment of antibiotic resistance and
virulence genes (tcpA and ctxA). Microb Ecol. 2013;65:969–74.
70. MacFadden DR, Bogoch II, Brownstein JS, et al. A passage from India: association between
air traffic and reported cases of New Delhi metallo-beta-lactase 1 from 2007 to 2012. Travel
Med Infect Dis. 2015;13:295–9.
71. Gould LH, Limbago B. Clostridium difficile in food and domestic animals: a new foodborne
pathogen? Clin Infect Dis. 2010;51(5):577–82.
72. Abbo A, Navon-Venezia S, Hamemer-Muntz O, et al. Multidrug-resistant Acinetobacter bau-
mannii. Emerg Infect Dis. 2005;11:22–9.
73. U.S. Food and Drug Administration (FDA) FACT SHEET: Veterinary Feed Directive
Final Rule and Next Steps. [cited 2017 May 30.] Available from: https://www.fda.gov/
AnimalVeterinary/DevelopmentApprovalProcess/ucm449019.htm
74. Lord Soulsby of Swaffham Prior. The 2008 Garrod Lecture: antimicrobial resistance--ani-
mals and the environment. J Antimicrob Chemother. 2008;62(2):229–33.
75. Price LB, Johnson E, Vailes R, Silbergeld E. Fluoroquinolone-resistant Campylobacter
isolates from conventional and antibiotic-free chicken products. Environ Health Perspect.
2005;113:557–60.
76. Hao R, Zhao R, Qiu S, et al. Antibiotics crisis in China. Science. 2015;348:1100–1.
77. Facinelli B, Roberts MC, Giovanetti E, et al. Genetic basis of tetracycline resistance in food
borne isolates of Listeria innocua. Appl Environ Microbiol. 1993;59:614–6.
78. Rolland RM, Hausfater G, Marshall B, et al. Antibiotic-resistant bacteria in wild pri-
mates: increased prevalence in baboons feeding on human refuse. Appl Environ Microbiol.
1985;49:791–4.
79. Glad T, Kristiansen VF, Nielsen KM, Brusetti L, Wright A-DG, Sundset MA. Ecological
characterisation of the colonic microbiota in arctic and sub-arctic seals. Microb Ecol.
2010;60:320–30.
80. Poeta P, Radhouani H, Pinto L, et al. Wild boars as reservoirs of extended-spectrum beta-
lactamase (ESBL) producing Escherichia coli of different phylogenetic groups. J Basic
Microbiol. 2009;49:584–8.
81. Marinho C, Igrehas G, Goncalves A, et al. Azorean wild rabbits as reservoirs of antimicrobial
resistant Escherichia coli. Anaerobe. 2014;30:116–9.
82. Ardiles-Villegas K, Gonzalez-Acuna D, Waldenstrom J, Olsen B, Hernandez J. Antibiotic
resistance patterns in fecal bacteria isolated from Christmas shearwater (Puffinus nativitatis)
and masked booby (Sula dactylatra) at remote Easter Island. Avian Dis. 2011;55:4896–489.
83. Nilsson O (2012) Vancomycin resistant enterococci in farm animals-occurrence and impor-
tance. Infect Ecol Epidemiol. [serial online] 2012 [cited 2017 June 2; 2:16959 Available
from: https://doi.org/10.3402/lee.v2i0.16959
84. Gordoncillo MJN, Donabedian S, Bartlett PC, et al. Isolation and molecular characterization
of vancomycin-resistance Enterococcus faecium from swine in Michigan, USA. Zoonoses
Pub Health. 2012. 2012;60:319–26.
85. Li K, Petersen G, Barco L, Hvidtfeldt K, Liu L, Dalsgaard A. Salmonella Weltevreden in inte-
grated and non-integrated tilapia aquaculture systems in Guangdong, China. Food Microbiol.
2017;65:19–24.
86. Muziasari WI, Pikanen LK, Sorum H, Stedtfeld RD, Tiedje JM, Virta M. The resistome of
farmed fish feces contributes to the enrichment of antibiotic resistance genes in sediments
below Baltic Sea fish farms. Front Microbiol. [serial online] 2017 [cited 2017 May 8]
Available from: doi:https://doi.org/10.3380/fmicb.2016.02137ss.
87. Akinbowale OL, Peng H, Barton MD. Diversity of tetracycline resistance genes in bacteria
from aquaculture sources in Australia. J Appl Microbiol. 2007;103:2016–25.
406 M. C. Roberts
88. DePaola A, Roberts MC. Class D and E tetracycline resistance determinants in gram-negative
catfish pond bacteria. Mol Cell Probes. 1995;9:311–3.
89. Furushita M, Shiba T, Maeda T, et al. Similarity of tetracycline resistance genes isolated from
fish farm bacteria to those from clinical isolates. Appl Environ Microbiol. 2003;69:5336–42.
90. Jacobs L, Chenia HY. Characterization of integrons and tetracycline resistance determinants
in Aeromonas spp. isolated from South African aquaculture systems. Int J Food Microbiol.
2007;114:295–306.
91. Nawaz M, Khan AA, Khan S, et al. Isolation and characterization of tetracycline-resistant
Citrobacter spp. from catfish. Food Microbiol. 2008;25:85–91.
92. Miranda CD, Zemelman R. Bacterial resistance to oxytetracycline in Chilean salmon farms.
Aquaculture. 2002;212:31–47.
93. Roberts MC, No D, Kuchmiy E, Miranda CD. The tetracycline resistant tet(39) gene identi-
fied in three new genera of bacteria isolated in 1999 from Chilean Salmon farms. J Antimicrob
Chemother. 2015;70:619–20.
94. Hayden EC. California faces arid future. Nature. 2015;526:14–5.
95. Reinthaler FF, Posch J, Feierl G, et al. Antibiotic resistance of E. coli in sewage and sludge.
Water Res. 2003;37:1685–90.
96. Hespanhol I. Wastewater a as resource, Chpt 4. In R Helmer, I Hespanhol, editors. Water
Pollution Control-A guide to the use of water quality management principles. 1997. Available
from: http://www.who.int/water_sanitation_health/resourcesquality/wpcchap4.pdf
97. Fuhrimann S, Winkler MS, Schneeberger PHH, et al. Health risk assessment along the waste-
water and faecal sludge management and reuse chain of Kampala, Uganda: a visualization.
Geospat Health. 2014;9:251–5.
98. Fenollare F, Marth T, Lagier JC, Angelakis E, Raoult D. Sewage workers with low antibody
response may be colonized successively by several Tropheryma whipplei strains. Int J Infect
Dis. 2015;35:51–5.
99. Rouch DA, Mondal T, Pai S, et al. Microbial safety of air-dried and rewetted biosolids.
J Water Health. 2011;9:403–14.
100. Xu C, Weese SJ, Flemming C, Odumeru J, Warriner K. Fate of Clostridium difficile dur-
ing wastewater treatment and incidence in Southern Ontario watersheds. J Appl Microbiol.
2014;117:891–904.
101. Burch TR, Sadowsky MJ, LaPara TM. Air-drying beds reduce the quantities of antibiotic
resistance genes and class 1 integrons in residual municipal wastewater solids. Environ Sci
Technol. 2013;47:9965–71.
102. Dropa M, Lincopan N, Balsalobre LC, et al. Genetic background of novel sequence types of
CTS-M-8- and CTX-M-15-producing Escherichia coli and Klebsiella pneumoniae from pub-
lic wastewater treatment plants in Sao Paul, Brazil. Environ Sci Pollut Res. 2016;23:4953–8.
103. Proia L, von Schiller D, Sanchez-Melsio A, et al. Occurrence and persistence of antibiotic
resistance genes in river biofilms after wastewater inputs in small rivers. Environ Pollut.
2016;210:121–8.
104. Fahrenfeld N, Ma Y, O’Brien M, Pruden A. Reclaimed water as a reservoir of antibiotic resis-
tance genes: distribution system and irrigation implications. Front Microbiol. [serial online]
2013 [cited 2017 May 30] Available from: doi:https://doi.org/10.3389/micb.2013.00130.
105. Richula J, Pereira RI, Wachholz GR, et al. Resistance to antimicrobial agents among entero-
cocci isolated from fecal samples of wild marine species in the southern coast of Brazil. Mar
Pollut Bull. 2016;105(1):51–7.
106. Raverty SA, Rhodes LD, Zabek E, et al. Respiratory microbiome of endangered Southern
resident killer whales and microbiota of surrounding sea surface microlayer in the Eastern
North Pacific. Sci Reports. [serial online] 2017 [cited 2017 June 1];7:394 Available from:
doi:https://doi.org/10.1038/s41598-017-00457-5.
107. Walter MV, Vennes JW. Occurrence of multiple-antibiotic-resistant enteric bacteria in domes-
tic sewage and oxidation lagoons. Appl Environ Microbiol. 1985;50:930–3.
12 Antibiotics and Resistance in the Environment 407
108. da Costa PM, Vaz-Pires P, Bernardo F. Antimicrobial resistance in Enterococcus spp. isolated
in inflow, effluent and sludge from municipal sewage waters treatment plants. Water Res.
2006;40:1735–40.
109. Luczkiewica A, Jankowska K, Fudala-Ksiazek S, Olanczuk-Neyman K. Antimicrobial
resistance of fecal indicators in municipal wastewater treatment plant. Water Res.
2010;44:5089–97.
110. Rosenberg Goldstein ER, Micallef SA, Gibbs SG, et al. Occupational exposure to
Staphylococcus aureus and Enterococcus spp. among spray irrigation workers using
reclaimed water. Int J Environ Res Public Health. 2014;11:4340–55.
111. Diallo AA, Brugere H, Kerouredan M, et al. Persistence and prevalence of pathogenic and
extended-spectrum beta-lactase-producing Escherichia coli in municipal wastewater treat-
ment plant receiving slaughterhouse wastewater. Water Res. 2013;47:4719–29.
112. Huijbers PMC, Blaak H, de Jong MCM, et al. Role of the environment in the transmission of
antimicrobial resistance to humans: a review. Environ Sci Technol. 2015;49:11993–2004.
113. Szczepanowski R, Linke B, Krahn I, et al. Detection of 140 clinically relevant antibiotic-
resistance genes in the plasmid metagenome of wastewater treatment plant bacteria showing
reduced susceptibility to selected antibiotics. Microbiology. 2009;155:2306–19.
114. Cai L, Ju F, Zhang T. Tracking human sewage microbiome in a municipal wastewater treat-
ment plant. Appl Microbiol Biotechnol. 2014;98:3317–26.
115. Auerbach EA, Seyfried EE, McMahon KD. Tetracycline resistance genes in activated sludge
wastewater treatment plants. Water Res. 2007;41:1143–51.
116. Naquin A, Shrestha A, Sherpa M, Nathaniel R, Boopathy R. Presence of antibiotic resis-
tance genes in a sewage treatment plant in Thibodaux, Louisiana, USA. Bioresour Technol.
2015;188:79–83.
117. Aminov RI. Horizontal gene exchange in environmental microbiota. Front Microbiol.
[serial online] 2011[cited 2017 May 22] Available from: doi:https://doi.org/10.3389/
fmmicb.2011.00158.
118. Ferro G, Guarino F, Castiglione S, Rizzo L. Antibioitc resistance spread potential in urban
wastewater effluents disinfected by UV/H2O2 process. Sci Total Environ. 2016;560-561:29–35.
Chapter 13
Phenotypic Tolerance and Bacterial
Persistence
Carl Nathan
13.1 Introduction
This chapter reproduces, adapts, and updates portions of a wider-ranging essay, “Fundamental
Immunodeficiency and Its Correction” by C. Nathan, published in the Journal of Experimental
Medicine 214: 2175-2191, 2017. Passages from that publication are reproduced here with permis-
sion of The Rockefeller University Press.
C. Nathan (*)
Department of Microbiology and Immunology, Weill Cornell Medicine, New York, NY, USA
e-mail: [email protected]
I discuss what is known about the mechanisms for each class and approaches to
overcome them.
Over the past six generations, humans have found or invented several thousand
medicines. Among them, the antimicrobial agents, discovered over the past four
generations, are unique in two aspects. First, until recently, antimicrobial agents
were the only medicines that cured large numbers of the sick, and they remain the
only medicines that do so routinely. Within the last two generations, some antineo-
plastic regimens have been curative, including some that are immunity-based, and
corticosteroids sometimes cure temporal arteritis. Second, antimicrobial agents are
the only medicines whose use hastens their loss of usefulness for people who have
not yet taken them.
The first claim hinges on using “cure” in the true sense. Administration of an
appropriately chosen antimicrobial agent has the routine capacity to restore an indi-
vidual to the state of wellness that prevailed before the onset of an illness that would
not otherwise have resolved, that would not otherwise have resolved as quickly, or
whose unaided resolution would not restore the individual to their prior state of
wellness. In contrast, when the administration of most other medicines stops, the
individual returns to the state of illness that invited intervention, unless the illness
had resolved spontaneously or from a change in contributory factors, such as diet.
Some other medicines help prevent the onset of illness rather than treating it.
The definition of “cure” given above is admittedly idealized. Clinical cure can be
ambiguous. “Cure” does not return the patient to the previous state of health if tissue
damage already caused by the pathogen or the host’s reaction to it is irreparable, as
is often the case in successfully treated TB. Finally, cure achieved with broad spec-
trum antimicrobial agents often comes at the cost of a long-lasting perturbation of
the microbiota, and in that sense an important component of the host’s overall
makeup has not returned to its preexistent state. Nonetheless, within the bounds of
these ambiguities and qualifications, antimicrobial agents stand out among medi-
cines for their ability to cure large numbers of people routinely.
However, the ability of antimicrobial agents to cure the majority of patients for
whom such drugs are appropriately prescribed is handicapped by the second unique
feature of this class of medicines: their use eventually selects for resistance. The
resistant pathogens are eventually shared among hosts, or the determinants of resis-
tance are eventually shared among pathogens. Thus we are all likely to need antimi-
crobial agents, yet the more a given agent is used, the nearer it comes to being
useless.
In sum, antimicrobial agents are at once among the most important and least
permanent achievements of medicine.
13 Phenotypic Tolerance and Bacterial Persistence 411
Beginning with the use of penicillin in civilian populations in the mid-1940s, physi-
cians, scientists, and much of the public quickly came to regard antimicrobial agents
as both indispensable and invincible [1]. Beginning just 20 years later, taking anti-
microbial agents for granted put us on a path to losing them.
Over the past few decades, a declining rate of success in discovering new antimi-
crobial agents discouraged much of the pharmaceutical industry from continuing
the search [2]. Meanwhile, levels of AMR continue to rise. These respectively fall-
ing and rising curves have crossed in recent years for one pathogen after another;
antimicrobial agents are now lacking to treat a significant proportion of formerly
curable infections caused by nearly a dozen different bacterial species. As the
remaining agents become less often useful, elective surgery and cancer chemother-
apy may become prohibitively risky, trauma care ineffective, premature babies non-
viable, and incidental wounds potentially lethal.
To imagine what it might be like to return to a pre-antibiotic era, consider the
reaction to the introduction of penicillin to public use after World War II. Alexander
Fleming “was showered with gifts of carnations… people whose lives had been
saved by penicillin … now knelt before him to kiss his hands” [3]. In 1964, the city
of Madrid installed statues of Fleming and of a bullfighter saluting him outside the
municipal bullring, because antibiotics had so greatly reduced the lethality of mata-
dors’ wounds.
One of the first postwar impacts of penicillin was the cure of gonorrhea with a
single injection. Yet Neisseria gonorrhoeae is one of the bacterial pathogens some
of whose clinical isolates are now resistant to most antibiotics. Others include
Enterococcus faecium; Staphylococcus aureus; Klebsiella pneumonia; Acinetobacter
baumannii; Pseudomonas aeruginosa; Enterobacter species; some Salmonella,
including invasive, non-typhoidal strains; some Shigella; and Mycobacterium tuber-
culosis. Leaving out the single most prevalent instance of AMR—drug-resistant
tuberculosis— it is estimated that drug-resistant bacterial pathogens now kill some
700,000 people a year, and if present trends continue, the toll will rise to 10 million
deaths per year by 2050 [4]. Authorities seem reluctant to factor drug-resistant
tuberculosis into this tally, perhaps fearing that its unfamiliarity to the citizenry of
economically advanced countries might blunt their concern. Nearly 500,000 people
a year develop drug-resistant tuberculosis; as matters now stand, over 50% of them
will die from it.
After decades of advocacy by scientists and physicians, beginning with Fleming
himself in his Nobel Prize acceptance speech in 1945, acknowledgment of the grav-
ity of AMR has finally come from leaders in business and government, as voiced by
the World Health Organization, the World Economic Forum, the G20, and the G7.
In 2015, President Obama issued a National Action Plan for Combating Antibiotic-
Resistant Bacteria [5]. In May 2016 a panel commissioned by the British govern-
ment issued cogent recommendations for coordinated global action [6]. In July
412 C. Nathan
That an immunologic perspective might help derives from four additional points:
(i) Mtb has no known naturally transmitting host but humans. (ii) As noted earlier,
for its transmission, Mtb needs a live human whose immune response is vigorous
enough to liquefy infected lung and erode into an airway. (This dependency proba-
bly accounts for the striking finding that the nucleotide sequences most highly con-
served among 1226 clinical isolates of Mtb were those encoding human T cell
epitopes, that is, the specific oligopeptides within a given protein that bind to anti-
gen receptors on T lymphocytes [16].) (iii) Untreated, the active disease has a fatal-
ity rate of 50% or more. (iv) Nonetheless, after an estimated 70,000 years of
parasitism, neither species—Mtb nor humans—has eliminated the other.
From these considerations we can reach four conclusions: Mtb has evolved the
ability to incite, titrate [17], survive, and exploit the human immune response.
To the degree that we understand the host-pathogen relationship in tuberculosis,
we should be able to apply strategies for drug development that accommodate or
even capitalize on those relationships rather than ignoring them and paying the price
for unappreciated antagonism.
The best understood form of AMR is heritable. There are bacterial genes that encode
resistance to antibiotics that were not invented or deployed at the time that the bac-
teria acquired the genes [18], and it is usually possible to isolate bacteria that have
become heritably resistant to any new antibiotic as soon as there is enough of the
antibiotic on hand to conduct a selection [19]. Apparent exceptions [20–22] are
likely to involve compounds with multiple targets or no specific target. Only a few
such agents are sufficiently selective to be clinically useful. In general, the issue
with heritable AMR is not whether but when the deployment of a given antibiotic
will select for the emergence of heritable resistance in clinical settings.
While correct use of antibiotics will usually lead in time to heritable AMR, other
forms of use hasten its emergence: misuse, overuse, and underuse.
Misuse is exemplified by feeding over half of the United States’ antibiotic ton-
nage to healthy food animals and plants to accelerate their growth; the proportion is
thought to be higher in China [23]. Another form of misuse is the routine failure to
account for individual variation in drug levels attained with standard dosing,
although it is possible to conduct therapeutic drug monitoring on finger-prick blood
spots [24]. Without dose adjustment, peak rifampin levels in the blood vary by
nearly two orders of magnitude in people treated for tuberculosis [25], with some
40–70% being undertreated [26]. Undertreatment fosters the emergence of
resistance.
Overuse results from lack of rapid, point-of-care diagnostics. An estimated 30%
of antibiotic prescriptions in the United States are written for the wrong indication,
typically a viral infection [27]. Overuse is also fostered in settings where the
414 C. Nathan
p rescribers are the purveyors or the consumers, that is, where doctors sell the drugs
or patients purchase them without recourse to doctors.
Underuse is a problem when the drugs are diluted by inexpert manufacture or
fraudulent intent or when patients discontinue them prematurely because they feel
better, feel worse, or cannot afford to buy more of them.
Mechanisms of heritable AMR are still being discovered. They include mutation
or posttranslational modification of the target so that it continues to support the
viability of the organism but no longer binds the antibiotic, increased expression of
the target so that it titrates the antibiotic, expression of a pathway that compensates
for the impairment caused by the antibiotic, inactivation of the antibiotic inside the
bacterium [28] or by a secreted bacterial product [29], decreased activation of a
prodrug form of the antibiotic, and decreased uptake or increased export of the
antibiotic.
Discovery of mechanisms of AMR has profoundly impacted both basic science
and clinical care. In basic science, studies of heritable AMR played a prominent role
in introducing the concept that small chemical compounds can have specific macro-
molecular targets in biological systems and can serve as tools to identify the targets’
functions [30]. Clinically, mechanistic understanding of heritable AMR allowed the
design of combination chemotherapy with agents that thwart resistance. For exam-
ple, the World Health Organization’s list of essential medicines includes the combi-
nation of amoxicillin, which is a β-lactam, with clavulanate, an inhibitor of some
bacterial β-lactamases. Moreover, mechanistic understanding of heritable AMR
allows combination chemotherapy with agents to which bacteria manifest resistance
by different mechanisms. Combination chemotherapy was introduced to the prac-
tice of medicine in the 1950s with the discovery that there was no other way to avoid
routine emergence of resistance in the treatment of TB [31]. The practice was later
adopted for the treatment of cancer and HIV/AIDS.
To set the stage for a discussion of phenotypic tolerance as a major form of AMR, it
helps to acknowledge the seemingly paradoxical negative impact of host immunity
on the action of anti-infectives that were developed without taking immunity into
account.
Because a primary function of the immune system is to protect the host from
infection and the purpose of administering antimicrobial agents is the same, then
immunity and antimicrobial chemotherapy can be expected to exert additive or syn-
ergistic effects, and no special effort should be necessary to take advantage of their
common actions.
Indeed, it is sometimes difficult to cure an infection with antibiotics in someone
whose encoded immune system is dysfunctional. For example, most patients with
13 Phenotypic Tolerance and Bacterial Persistence 415
In contrast to the situation with heritable AMR, we have very limited understanding
of nonheritable AMR, also called “phenotypic tolerance,” a term introduced by
Tuomanen [37]. Phenotypic tolerance can be defined as conditional drug resistance
that is not attributable to changes in the nucleic acid sequence of the pathogen’s
genome. Phenotypic tolerance gives rise to bacterial persistence: survival of bacte-
ria during treatment of a host with a drug to which the same strain of pathogen is
susceptible under standard laboratory conditions at concentrations achieved in the
host. Phenotypic tolerance predisposes to emergence of mutants with heritable
resistance [44].
The first two studies of phenotypic tolerance hold such important lessons for
today that they deserve detailed discussion. The purification of penicillin was
reported in 1942 [45]. That same year, Gladys Hobby and her colleagues reported
that at 37 °C, about 1 streptococcus remained viable after 48 hours of exposure to
penicillin for every 106 present in the control culture at the end of that period. The
authors did not comment on that but drew attention to the survival of nearly all the
penicillin-treated streptococci if the exposure took place at 4 °C, conditions in
which there was no increase in bacterial number in the untreated control culture.
The authors concluded, “It is apparent that penicillin is capable of destroying bacte-
ria only if multiplication takes place” [46].
In 1944, Joseph Bigger repeated and extended the experiments using staphylo-
cocci [47]. He introduced the term “persisters” to stress the observation that about
1 in 106 staphylococci survived the treatment of logarithmically replicating cultures
at body temperature. He inferred that persisters to penicillin must be “cocci …
which happen to be, when exposed to it, in a phase in which they are insusceptible
to its action,” because “If persisters had an abnormally high resistance, either natu-
ral [that is, heritable and existing prior to the experiment] or acquired [that is, heri-
table but acquired during the experiment], it is probable that their descendants
would also possess abnormally high resistance. The descendants of a number of
persisters which had survived contact with 1 unit per c.cm. penicillin for 3–5 days
were found to be killed by 1/8 unit per c.cm. within 46 hours and to have no greater
tendency than normal forms to produce persisters” [47].
Bigger went on to confirm the observation of Hobby et al. [46] that cooling the
bacteria elevated the frequency of persisters to nearly 100%, that is, by 6 orders of
magnitude. He demonstrated the same effect by acidifying the medium or lowering
its tonicity. He concluded that “persisters are cocci which survive contact with peni-
cillin because they are in dormant, non-dividing phase” [47].
In fact, within 2 years of the report of penicillin’s publication, the two groups
mentioned above, working on two continents with two different pathogens, had each
observed two different classes of phenotypic tolerance, but without distinguishing
them. It took another 70 years before the distinction was made, driven by the recog-
nition that the two classes have different implications for drug discovery [11].
13 Phenotypic Tolerance and Bacterial Persistence 417
Not all the anti-infectives that kill Mtb in some non-replicating states kill Mtb in
other non-replicating states. For example, rifampin generated rather than killed the
differentially detectable Mtb described above, while thioridazine did not generate
such cells but did kill them [55]. These antimicrobial agents serve as chemical
probes to teach us that class II phenotypic tolerance encompasses a spectrum of
states—at our present state of knowledge, at least two. Class IIa phenotypic toler-
ance is characteristic of bacteria that stop replicating in response to a given set of
stresses but form CFU when those stresses are relieved. Class IIb phenotypic toler-
ance is a feature of bacteria that stop replicating in response to different stresses and
remain viable when those stresses are removed, but do not form CFU [55]. This
complicates the task of finding anti-infectives that can kill bacteria displaying phe-
notypic tolerance.
To the extent that individual bacteria in an otherwise antibiotic-susceptible popu-
lation manifest class I phenotypic tolerance to two different antibiotics by different
mechanisms, then the cells that are phenotypically tolerant to the first antibiotic are
likely to be susceptible to the second. In such a case, to kill the whole population, it
should suffice to combine antibiotics in such a way that no one bacterium is pheno-
typically tolerant to all of them, provided that each of the drugs in the combination
reaches the bacteria in adequate concentrations at the same time. (In vitro, class I
phenotypically tolerant Mtb could be killed by forcing them to produce extra ROS
in the presence of rifampin or isoniazid by supplying them with small thiols [56].)
In contrast, if all the bacteria in a population are phenotypically tolerant to several
different antibiotics, then each individual bacterium must be tolerant to each of
13 Phenotypic Tolerance and Bacterial Persistence 419
Class I phenotypic tolerance can theoretically arise by any mechanism that confers
heritable AMR, from epigenetic regulation to posttranslational modification, as
long as the mechanism does not depend on a change in the pathogen’s coding
sequence. As noted earlier, the size of the tolerant subpopulation may be affected by
a change in coding sequence, as long as the tolerant subpopulation remains such a
minority that the overall population does not manifest an increase in the antibiotic’s
MIC.
Much of the research in this field has wrestled with a descriptive question,
whether class I phenotypic tolerance is as tightly linked with non-replication as
Hobby et al. [46] and Bigger [47] inferred. In short, the answer is “no.”
The first study to use time-lapse photomicroscopy of bacteria in microfluidic
chambers to study phenotypic tolerance at the single cell level [60] revealed that in
an otherwise replicating population of E. coli, most of the few cells that survived
ampicillin were non-replicating at the time of exposure to the drug. However, some
of the other surviving E. coli had been replicating. This study was rendered feasible
by using E. coli with compound mutations in hipA that raised the frequency of class
I phenotypically tolerant E. coli by several orders of magnitude without changing
the MIC of the overall population.
Nine years later, a study of similar design reached a different conclusion while
studying the action of isoniazid on Mtb [61]. Isoniazid is a prodrug whose activation
depends on the Mtb catalase-peroxidase KatG. The investigators showed that sto-
chastic extinction of KatG expression conferred resistance to isoniazid. Growth rate
had nothing to do with it [61].
The same year, Orman and Brynildsen showed that E. coli persisters to ampicil-
lin and fluoroquinolones are enriched among the non-replicating subpopulation, but
not confined to it nor highly prevalent in it [62]. Natural clinical and veterinary
isolates of E. coli each showed the same MICs to a given antibiotic, yet each showed
different levels of persistence to different sets of antibiotics [63]. This suggested
that different individual cells were phenotypically tolerant to different antibiotics,
meaning that non-replication of a given cell could not be a universal explanation for
phenotypic tolerance.
420 C. Nathan
One of the most important challenges for antibiotic research is to understand mech-
anisms of class II phenotypic tolerance, a state for which incompletely effective
immunity and sublethal antibiotic therapy bear much of the responsibility.
We have a long way to go. We do not know if a given bacterial species that enters
a non-replicating state in response to different host conditions manifests class II
13 Phenotypic Tolerance and Bacterial Persistence 421
hibernation factor and ribosome modulation factor, proteins that bind ribosomes
and inhibit translation [79]. It is clear how these actions could suppress replication,
but as noted above, suppression of replication does not suffice as a general explana-
tion of phenotypic tolerance.
stage of the assay to the stage of the assay where recovery is assessed under condi-
tions that support replication [93]. Rifampin has genuine bactericidal action on non-
replicating Mtb in vitro but at far higher concentrations than needed to kill replicating
Mtb, and even then, the maximum extent of killing in vitro is far less [93]. This is
not meant to disparage the proven clinical utility of these drugs but rather to suggest
that they do not represent an ideal solution to the problem of class II phenotypic
tolerance.
Fortunately, compounds can be found that extensively kill bacteria in a state that
confers class II phenotypic tolerance to conventional antibiotics. An early example
was a thioxothiazolidine that killed Mtb only when the Mtb was non-replicating,
without regard to diverse conditions tested that imposed non-replication [81].
Another target-based screen led to two chemically distinct classes of Mtb-selective
proteasome inhibitors [88, 92] that killed Mtb that was rendered non-replicating by
nitrosative stress [88, 92] or starvation [94]. A whole-cell screen designed to iden-
tify compounds that kill non-replicating Mtb identified oxyphenbutazone [35] and
other compounds [95]. Subsequently, over 100 compounds have been reported to
kill non-replicating Mtb selectively, including novel cephalosporins [96]. However,
in only a few cases did the investigators exclude the possibility that carry-over of
compound into the replicative phase of the assay may have led to a false impression
of activity in the preceding, non-replicative phase of the assay [30].
Why are some compounds only able to kill non-replicating bacteria, sparing the
same cells when they replicate? Barring compound modification under one of the
two sets of assay conditions, and assuming equivalent uptake under both, the ques-
tion becomes why some targets are nonessential under conditions that support rep-
lication but essential under conditions that do not. For example, at least four sets of
Mtb enzymes involved in central carbon metabolism—hydroxyoxoadipate syn-
thase, dihydrolipoamide acyltransferase, lipoamide dehydrogenase, and the isoci-
trate lyases—are dispensable for survival under nonstressed conditions but become
essential for Mtb to withstand oxidative or nitrosative stresses that impose non-
replication [68, 76, 81, 84]. This invites the speculation that some pathways that
would afford redundancy in a critical function targeted by the antibiotic are inacti-
vated under non-replicative conditions, or a singular essential pathway incompletely
inhibited by the antibiotic is further inhibited by the non-replicative conditions.
Even more encouraging are antibiotics that can kill bacteria extensively not only
when they are replicating but also when they are not replicating and are phenotypi-
cally tolerant to other antibiotics. With respect to tuberculosis, this has been reported
with 8-hydroxyquinolines [97, 98] and nitazoxanide, an antibiotic approved for
other indications [20]. In vitro, the nitroimidazole PA-824 (Pretomanid) kills both
replicating and non-replicating Mtb to comparable extents and at comparable con-
centrations [30, 99]. Under non-replicating conditions, the mechanism involves
generation of reactive nitrogen species [99], a striking example of a synthetic anti-
biotic mimicking host immunity [100].
424 C. Nathan
Major Points
• Phenotypic tolerance prevents an antimicrobial agent from eradicating a patho-
gen population; it likely accounts for relapse and contributes to the emergence of
heritable resistance.
• Type I phenotypic tolerance occurs when a minority (subpopulation) survives
antibiotic treatment in conditions permissive for growth of the majority popula-
tion and individual tolerant bacterial cells are each tolerant to a different
antibiotic.
• Type II phenotypic tolerance is a bacterial response to exogenous stress that
impairs growth and pertains to all of the bacteria whose growth is impaired and
individual bacterial cells are each tolerant to multiple antibiotics.
• Host cell immunity can foster phenotypic tolerance and thereby work at cross-
purposes with antimicrobials.
• Better mechanistic understanding of the different classes of phenotypic tolerance
will help improve antimicrobial chemotherapy and help reduce the emergence of
heritable antimicrobial resistance.
References
1. Nathan C. Cooperative development of antimicrobials: looking back to look ahead. Nat Rev
Microbiol. 2015;13(10):651–7.
2. Payne DJ, Gwynn MN, Holmes DJ, Pompliano DL. Drugs for bad bugs: confronting the chal-
lenges of antibacterial discovery. Nat Rev Drug Discov. 2007;6(1):29–40.
3. Brown K. Penicillin man: Alexander Fleming and the antibiotic revolution. Stroud: History
Press; UK. 2013.
4. O’Neill J. Tackling drug-resistant infections globally. Review on Antimicrobial Resistance.
London, England, UK. 2014.
5. White House. National action plan for combating antibiotic-resistant bacteria. Washington,
DC. 2015. 62 pp.
6. O’Neill J. Tackling drug resistant infections globally: final report and recommendations.
Review on Antimicrobial Resistance. London, England, UK. 2016.
7. Outterson K, Rex JH, Jinks T, Jackson P, Hallinan J, Karp S, et al. Accelerating global
innovation to address antibacterial resistance: introducing CARB-X. Nat Rev Drug Discov.
2016;15(9):589–90.
8. Bagley N, Outterson K. How to avoid a post-antibiotic world. New York Times. 2017:Op-Ed.
9. United Nations. Draft political declaration of the high-level meeting of the General Assembly
on antimicrobial resistance. New York City, New York. 2016.
10. Nathan C. Antibiotics at the crossroads. Nature. 2004;431(7011):899–902.
11. Nathan C. Fresh approaches to anti-infective therapies. Sci Transl Med. 2012;4(140):140sr2.
12. Nathan C, Cars O. Antibiotic resistance–problems, progress, and prospects. N Engl J Med.
2014;371(19):1761–3.
13 Phenotypic Tolerance and Bacterial Persistence 425
13. Nathan C. Making space for anti-infective drug discovery. Cell Host Microbe.
2011;9(5):343–8.
14. Nathan C. Taming tuberculosis: a challenge for science and society. Cell Host Microbe.
2009;5(3):220–4.
15. Pietersen E, Ignatius E, Streicher EM, Mastrapa B, Padanilam X, Pooran A, et al. Long-term
outcomes of patients with extensively drug-resistant tuberculosis in South Africa: a cohort
study. Lancet. 2014;383(9924):1230–9.
16. Coscolla M, Copin R, Sutherland J, Gehre F, de Jong B, Owolabi O, et al. M. tuberculosis
T cell epitope analysis reveals paucity of antigenic variation and identifies rare variable TB
antigens. Cell Host Microbe. 2015;18(5):538–48.
17. Marakalala MJ, Raju RM, Sharma K, Zhang YJ, Eugenin EA, Prideaux B, et al.
Inflammatory signaling in human tuberculosis granulomas is spatially organized. Nat Med.
2016;22(5):531–8.
18. Bhullar K, Waglechner N, Pawlowski A, Koteva K, Banks ED, Johnston MD, et al. Antibiotic
resistance is prevalent in an isolated cave microbiome. PLoS One. 2012;7(4):e34953.
19. Kling A, Lukat P, Almeida DV, Bauer A, Fontaine E, Sordello S, et al. Antibiotics. Targeting
DnaN for tuberculosis therapy using novel griselimycins. Science. 2015;348(6239):1106–12.
20. de Carvalho LP, Lin G, Jiang X, Nathan C. Nitazoxanide kills replicating and nonreplicating
Mycobacterium tuberculosis and evades resistance. J Med Chem. 2009;52(19):5789–92.
21. Ling LL, Schneider T, Peoples AJ, Spoering AL, Engels I, Conlon BP, et al. A new antibiotic
kills pathogens without detectable resistance. Nature. 2015;517(7535):455–9.
22. Moreira W, Aziz DB, Dick T. Boromycin kills mycobacterial persisters without detectable
resistance. Front Microbiol. 2016;7:199.
23. Van Boeckel TP, Brower C, Gilbert M, Grenfell BT, Levin SA, Robinson TP, et al. Global
trends in antimicrobial use in food animals. Proc Natl Acad Sci U S A. 2015;112(18):5649–54.
24. Alsultan A, Peloquin CA. Therapeutic drug monitoring in the treatment of tuberculosis: an
update. Drugs. 2014;74(8):839–54.
25. Wilkins JJ, Savic RM, Karlsson MO, Langdon G, McIlleron H, Pillai G, et al. Population phar-
macokinetics of rifampin in pulmonary tuberculosis patients, including a semimechanistic
model to describe variable absorption. Antimicrob Agents Chemother. 2008;52(6):2138–48.
26. Um SW, Lee SW, Kwon SY, Yoon HI, Park KU, Song J, et al. Low serum concentrations
of anti-tuberculosis drugs and determinants of their serum levels. Int J Tuberc Lung Dis.
2007;11(9):972–8.
27. Fleming-Dutra KE, Hersh AL, Shapiro DJ, Bartoces M, Enns EA, File TM Jr, et al. Prevalence
of inappropriate antibiotic prescriptions among US Ambulatory Care Visits, 2010–2011.
JAMA. 2016;315(17):1864–73.
28. Warrier T, Kapilashrami K, Argyrou A, Ioerger TR, Little D, Murphy KC, et al. N-methylation
of a bactericidal compound as a resistance mechanism in Mycobacterium tuberculosis. Proc
Natl Acad Sci U S A. 2016;113(31):E4523–30.
29. El-Halfawy OM, Klett J, Ingram RJ, Loutet SA, Murphy ME, Martin-Santamaria S, et al.
Antibiotic capture by bacterial Lipocalins uncovers an extracellular mechanism of intrinsic
antibiotic resistance. MBio. 2017;8(2)
30. Gold B, Nathan C. Targeting phenotypically tolerant Mycobacterium tuberculosis. In: Jacobs
Jr WR, McShane H, Mizrahi V, Orme I, editors. Tuberculosis and the tubercle Bacillus. 2nd
ed: American Society of Microbiolgy Press; 2017.
31. Fox W, Sutherland IA. Five-year assessment of patients in a controlled trial of streptomy-
cin, Para-aminosalicylic acid, and streptomycin plus Para-aminosalicylic acid, in pulmonary
tuberculosis. Q J Med. 1956;25(98):221–43.
32. Lin CH, Chi CY, Shih HP, Ding JY, Lo CC, Wang SY, et al. Identification of a major epitope
by anti-interferon-gamma autoantibodies in patients with mycobacterial disease. Nat Med.
2016;22(9):994–1001.
33. Xie QW, Cho HJ, Calaycay J, Mumford RA, Swiderek KM, Lee TD, et al. Cloning and
characterization of inducible nitric oxide synthase from mouse macrophages. Science.
1992;256(5054):225–8.
426 C. Nathan
34. McCune RM, Feldmann FM, Lambert HP, McDermott W. Microbial persistence.
I. The capacity of tubercle bacilli to survive sterilization in mouse tissues. J Exp Med.
1966;123(3):445–68.
35. Gold B, Pingle M, Brickner SJ, Shah N, Roberts J, Rundell M, et al. Nonsteroidal anti-
inflammatory drug sensitizes Mycobacterium tuberculosis to endogenous and exogenous
antimicrobials. Proc Natl Acad Sci U S A. 2012;109(40):16004–11.
36. Helaine S, Cheverton AM, Watson KG, Faure LM, Matthews SA, Holden DW. Internalization
of Salmonella by macrophages induces formation of nonreplicating persisters. Science.
2014;343(6167):204–8.
37. Tuomanen E. Phenotypic tolerance: the search for beta-lactam antibiotics that kill nongrow-
ing bacteria. Rev Infect Dis. 1986;8(Suppl 3):S279–91.
38. Liu Y, Tan S, Huang L, Abramovitch RB, Rohde KH, Zimmerman MD, et al. Immune activa-
tion of the host cell induces drug tolerance in Mycobacterium tuberculosis both in vitro and
in vivo. J Exp Med. 2016;213(5):809–25.
39. Zemke AC, Gladwin MT, Bomberger JM. Sodium nitrite blocks the activity of amino-
glycosides against Pseudomonas aeruginosa biofilms. Antimicrob Agents Chemother.
2015;59(6):3329–34.
40. Gusarov I, Shatalin K, Starodubtseva M, Nudler E. Endogenous nitric oxide protects bacteria
against a wide spectrum of antibiotics. Science. 2009;325(5946):1380–4.
41. Group ICGDCS. A controlled trial of interferon gamma to prevent infection in chronic gran-
ulomatous disease. The International Chronic Granulomatous Disease Cooperative Study
Group. N Engl J Med. 1991;324(8):509–16.
42. Sun K, Yajjala VK, Bauer C, Talmon GA, Fischer KJ, Kielian T, et al. Nox2-derived oxidative
stress results in inefficacy of antibiotics against post-influenza S. aureus pneumonia. J Exp
Med. 2016;213(9):1851–64.
43. Nathan C, Gold B, Lin G, Stegman M, de Carvalho LP, Vandal O, et al. A philosophy of
anti-infectives as a guide in the search for new drugs for tuberculosis. Tuberculosis (Edinb).
2008;88(Suppl 1):S25–33.
44. Levin-Reisman I, Ronin I, Gefen O, Braniss I, Shoresh N, Balaban NQ. Antibiotic tolerance
facilitates the evolution of resistance. Science. 2017;355(6327):826–30.
45. Abraham EP, Chain E. Purification of penicillin. Nature. 1942;149:328.
46. Hobby GL, Meyer K, Chaffee E. Observations on the mechanism of action of penicillin. Proc
Soc Exp Biol Med. 1942;50:281–5.
47. Bigger J. Treatment of staphylococcal infections with penicillin by intermittent sterilisation.
Lancet. 1944;244:497–500.
48. Nathan C, Barry CE 3rd. TB drug development: immunology at the table. Immunol Rev.
2015;264(1):308–18.
49. Kohanski MA, DePristo MA, Collins JJ. Sublethal antibiotic treatment leads to multidrug
resistance via radical-induced mutagenesis. Mol Cell. 2010;37(3):311–20.
50. Boshoff HI, Reed MB, Barry CE 3rd, Mizrahi V. DnaE2 polymerase contributes to in vivo
survival and the emergence of drug resistance in Mycobacterium tuberculosis. Cell.
2003;113(2):183–93.
51. Xu HS, Roberts N, Singleton FL, Attwell RW, Grimes DJ, Colwell RR. Survival and viability
of nonculturable Escherichia coli and Vibrio cholerae in the estuarine and marine environ-
ment. Microb Ecol. 1982;8(4):313–23.
52. Chengalroyen MD, Beukes GM, Gordhan BG, Streicher EM, Churchyard G, Hafner R, et al.
Detection and quantification of differentially culturable Tubercle Bacteria in sputum from
tuberculosis patients. Am J Respir Crit Care Med. 2016;194(12):1532–40.
53. Dartois V, Saito K, Warrier T, Nathan C. Editorial: new evidence for the complexity of the
population structure of Mycobacterium tuberculosis increases the diagnostic and biologic
challenges. Am J Resp Crit Care Med. 2016;194:1448–50.
54. Mukamolova GV, Turapov O, Malkin J, Woltmann G, Barer MR. Resuscitation-promoting
factors reveal an occult population of tubercle Bacilli in Sputum. Am J Respir Crit Care Med.
2010;181(2):174–80.
13 Phenotypic Tolerance and Bacterial Persistence 427
76. Nandakumar M, Nathan C, Rhee KY. Isocitrate lyase mediates broad antibiotic tolerance in
Mycobacterium tuberculosis. Nat Commun. 2014;5:4306.
77. Harms A, Maisonneuve E, Gerdes K. Mechanisms of bacterial persistence during stress and
antibiotic exposure. Science. 2016;354(6318)
78. Jankevicius G, Ariza A, Ahel M, Ahel I. The toxin-antitoxin system DarTG catalyzes revers-
ible ADP-ribosylation of DNA. Mol Cell. 2016;64(6):1109–16.
79. Beckert B, Abdelshahid M, Schafer H, Steinchen W, Arenz S, Berninghausen O, et al.
Structure of the Bacillus subtilis hibernating 100S ribosome reveals the basis for 70S dimer-
ization. EMBO J. 2017;
80. Rath P, Huang C, Wang T, Wang T, Li H, Prados-Rosales R, et al. Genetic regulation of
vesiculogenesis and immunomodulation in Mycobacterium tuberculosis. Proc Natl Acad Sci
U S A. 2013;110(49):E4790–7.
81. Bryk R, Gold B, Venugopal A, Singh J, Samy R, Pupek K, et al. Selective killing of nonrep-
licating mycobacteria. Cell Host Microbe. 2008;3(3):137–45.
82. Bryk R, Griffin P, Nathan C. Peroxynitrite reductase activity of bacterial peroxiredoxins.
Nature. 2000;407(6801):211–5.
83. Bryk R, Lima CD, Erdjument-Bromage H, Tempst P, Nathan C. Metabolic enzymes
of mycobacteria linked to antioxidant defense by a thioredoxin-like protein. Science.
2002;295(5557):1073–7.
84. Maksymiuk C, Balakrishnan A, Bryk R, Rhee KY, Nathan CF. E1 of alpha-ketoglutarate
dehydrogenase defends Mycobacterium tuberculosis against glutamate anaplerosis and
nitroxidative stress. Proc Natl Acad Sci U S A. 2015;112(43):E5834–43.
85. Vandal OH, Pierini LM, Schnappinger D, Nathan CF, Ehrt S. A membrane protein pre-
serves intrabacterial pH in intraphagosomal Mycobacterium tuberculosis. Nat Med.
2008;14(8):849–54.
86. Darwin KH, Nathan CF. Role for nucleotide excision repair in virulence of Mycobacterium
tuberculosis. Infect Immun. 2005;73(8):4581–7.
87. Darwin KH, Ehrt S, Gutierrez-Ramos JC, Weich N, Nathan CF. The proteasome
of Mycobacterium tuberculosis is required for resistance to nitric oxide. Science.
2003;302(5652):1963–6.
88. Lin G, Li D, de Carvalho LP, Deng H, Tao H, Vogt G, et al. Inhibitors selective for mycobac-
terial versus human proteasomes. Nature. 2009;461(7264):621–6.
89. Vaubourgeix J, Lin G, Dhar N, Chenouard N, Jiang X, Botella H, et al. Stressed mycobacteria
use the chaperone ClpB to sequester irreversibly oxidized proteins asymmetrically within and
between cells. Cell Host Microbe. 2015;17(2):178–90.
90. Bryk R, Arango N, Maksymiuk C, Balakrishnan A, Wu YT, Wong CH, et al. Lipoamide
channel-binding sulfonamides selectively inhibit mycobacterial lipoamide dehydrogenase.
Biochemistry. 2013;52(51):9375–84.
91. Bryk R, Arango N, Venugopal A, Warren JD, Park YH, Patel MS, et al.
Triazaspirodimethoxybenzoyls as selective inhibitors of mycobacterial lipoamide dehydro-
genase. Biochemistry. 2010;49(8):1616–27.
92. Lin G, Chidawanyika T, Tsu C, Warrier T, Vaubourgeix J, Blackburn C, et al. N,C-capped
dipeptides with selectivity for mycobacterial proteasome over human proteasomes: role of
S3 and S1 binding pockets. J Am Chem Soc. 2013;135(27):9968–71.
93. Gold B, Roberts J, Ling Y, Quezada LL, Glasheen J, Ballinger E, et al. Rapid, semiquantita-
tive assay to discriminate among compounds with activity against replicating or nonreplicat-
ing Mycobacterium tuberculosis. Antimicrob Agents Chemother. 2015;59(10):6521–38.
94. Russo F, Gising J, Akerbladh L, Roos AK, Naworyta A, Mowbray SL, et al. Optimization and
evaluation of 5-Styryl-Oxathiazol-2-one Mycobacterium tuberculosis proteasome inhibitors
as Potential Antitubercular Agents. ChemistryOpen. 2015;4(3):342–62.
95. Warrier T, Martinez-Hoyos M, Marin-Amieva M, Colmenarejo G, Porras-De Francisco E,
Alvarez-Pedraglio AI, et al. Identification of novel anti-mycobacterial compounds by screen-
ing a pharmaceutical small-molecule library against nonreplicating Mycobacterium tubercu-
losis. ACS Infect Dis. 2015;1(12):580–5.
13 Phenotypic Tolerance and Bacterial Persistence 429
96. Gold B, Smith R, Nguyen Q, Roberts J, Ling Y, Lopez Quezada L, et al. Novel cephalo-
sporins selectively active on nonreplicating Mycobacterium tuberculosis. J Med Chem.
2016;59(13):6027–44.
97. Darby CM, Nathan CF. Killing of non-replicating Mycobacterium tuberculosis by
8-hydroxyquinoline. J Antimicrob Chemother. 2010;65(7):1424–7.
98. Shah S, Dalecki AG, Malalasekera AP, Crawford CL, Michalek SM, Kutsch O, et al.
8-Hydroxyquinolines are boosting agents of copper-related toxicity in Mycobacterium tuber-
culosis. Antimicrob Agents Chemother. 2016;60(10):5765–76.
99. Singh R, Manjunatha U, Boshoff HI, Ha YH, Niyomrattanakit P, Ledwidge R, et al. PA-824
kills nonreplicating Mycobacterium tuberculosis by intracellular NO release. Science.
2008;322(5906):1392–5.
100. Nathan C. An antibiotic mimics immunity. Science. 2008;322(5906):1337–8.
Chapter 14
Staphylococcus aureus Adaptation During
Infection
14.1 Introduction
B. Shopsin (*)
Departments of Medicine and Microbiology, New York University School of Medicine,
New York, NY, USA
e-mail: [email protected]
R. Copin
Department of Medicine, New York University School of Medicine, New York, NY, USA
S. aureus is responsible for a large variety of diseases in both community and hospital
settings [5]. Despite advances in care, S. aureus infections remain associated with
considerable morbidity and mortality. In addition, treatment of methicillin-resistant
14 Staphylococcus aureus Adaptation During Infection 433
AIP
B
C
D
A
P2 P3
A C D B
Fig. 14.1 The agr quorum-sensing system. (A) The agr locus consists of two divergent transcrip-
tion units driven by promoters P2 and P3. The P2 operon encodes the signaling module, which
contains four genes – agrB, D, C, and A – each of which is required for transcriptional activation of
the agr regulon (reviewed in [9]). AgrC is the receptor-histidine kinase, and AgrA is the response
regulator. AgrD is the autoinducing, secreted peptide that is derived from a propeptide processed by
AgrB. The P3 transcript is a regulatory RNA (RNAIII) that also encodes the structural gene for
hemolysin. Regulation of target genes by agr occurs through two pathways: (1) an RNAIII-
dependent regulation of virulence genes and (2) an RNAIII-independent, AgrA-mediated regulation
of metabolic genes and small cytolytic toxins known as phenol-soluble modulins (modulins). The
regulatory connection between these processes links virulence to metabolism. Agr has a dual, time-
dependent regulatory role (in vitro) that is characterized by (1) increased post-exponential produc-
tion of toxins and exoenzymes (e.g., α-hemolysin) that facilitate dissemination of bacteria via tissue
invasion; (2) decreased production of cell surface proteins that facilitate adherence and attachment
(e.g., fibronectin-binding proteins); and (3) decreased production of factors that promote the eva-
sion of host defense (e.g., protein A). Thus, the agr locus coordinates a switch from an adherent
state to an invasive state dependent on bacterial population density. This important duality has been
exploited by the use of agr quorum-sensing inhibitors for the prevention and treatment of experi-
mental S. aureus infections, including catheter and vascular prosthetic graft infection [10–13]
of examining the role of agr in colonization. Screening assays to detect agr func-
tionality among isolates from healthy subjects indicate that although agr dysfunc-
tion is not an absolute barrier to colonization and transmission, carriage of
agr-defective strains is strongly associated with hospitalization rather than with
healthy patients [27]. Collectively, these observations suggested that: (1) agr-
defective mutants are fit for transmission (they are not a “dead-end” state), and (2)
the hospital environment is a reservoir of attenuated, agr-defective variants.
Presumably, disruption of protective barrier functions by disease and clinical inter-
vention (e.g., intravenous catheter use) permits S. aureus lacking full virulence to
cause infection. Analysis of paired S. aureus clones from blood infection and nasal
carriage sites in individual hospitalized patients presenting with bacteremia indicate
that recovery of an agr-defective mutant from blood is usually predicted by the agr
status of carriage isolates [28]. Thus, fieldwork supports the idea that the transition
from commensalism to opportunism in S. aureus does not require full virulence in
hospitalized patients.
The strong association of agr dysfunction with the hospital environment and
infection suggests an unappreciated role for agr: colonization by S. aureus is
responsible for maintaining agr function. Indeed, although fully agr-defective
mutant isolates colonize and transmit, they do not persist indefinitely in natural
populations of hospital-associated MRSA [29]. This suggests that, in the case of agr
mutation, attenuation of virulence is the product of short-sighted evolution within
hosts – although attenuation of agr-mediated virulence may help S. aureus adapt to
host tissues in the short term, it appears to put S. aureus at a disadvantage in the long
term.
The combination of ubiquity and relatively short lifespan suggests that the
occurrence of agr-defective mutants results from frequent within-host selection in
situations such as persistent bacteremia [18, 21, 22, 28]. However, an experimen-
tal system demonstrating transmission following invasive bacteremia was lacking,
and thus implications of within-host adaptation for between-host transmission –
and therefore for hospital epidemiology – were unknown. While a disease-pro-
moting agr mutation that occurs during the course of bacteremia could confer a
transient advantage to the bacterium, such an adaption would be a dead end for the
bacterium in the absence of transmission. Recently, S. aureus was found to dis-
seminate to the gastrointestinal tract of mice via the gall bladder following intra-
venous injection, and the bacterium readily transmits to cohoused naive mice
[30]. These findings established an animal model to investigate gastrointestinal
dissemination of S. aureus and the role of adaptive mutations in genes such as agr.
The work suggests that selective processes taking place over the course of blood
infection can go beyond a single host. Both intestinal dissemination and transmis-
sion were linked to the production of virulence factors based on gene deletion
studies of two-component virulence regulatory systems, including agr. Thus, the
animal data are consistent with data from hospital isolates that indicate that agr
inactivation can attenuate colonization-transmission but is selected during
bacteremia.
436 B. Shopsin and R. Copin
agr-positive strains could adhere to the agr-defective variants while producing their
toxic exoproteins. Additionally, inactivation of agr upregulates fibronectin-binding
proteins, which play an important role in the ability of S. aureus to colonize, persist
within, and damage cardiovascular tissue [49]. agr inactivation also increases resis-
tance to endogenous thrombin-induced microbicidal proteins [50], key mediators of
host defense that are secreted by platelets at sites of cardiovascular damage and
infection.
It is also possible that agr mutation promotes S. aureus survival inside host
cells, as opposed to within-host tissues. Although considered an extracellular
pathogen, S. aureus clearly thrives inside host phagocytic, epithelial, and endothe-
lial cells ([51–56]; reviewed in [57]). The importance of this intracellular lifestyle
is highlighted by the recent finding that ablation of intracellular S. aureus improves
outcome from experimental infection [54, 55]. While the large majority of work on
MRSA-phagocyte interactions, including work from our laboratories, has been
performed with neutrophils, recent findings indicate that disseminated S. aureus
infection is tied to survival inside macrophages [52–55]. For some cell types, agr-
defective mutants exhibit prolonged intracellular residence due to attenuated cyto-
toxicity and delay in initiation of host cell death [58, 59]. But the agr-defective
phenotype is not noted in isolates from primary skin and soft tissue infection (e.g.,
[60]), suggesting that such attenuated toxicity and intracellular survival may be
particularly important in infections in which persistence is an issue. Persistence is
a particular problem with S. aureus endocarditis, while it is not such an issue with
acute infection of skin and soft tissue. Thus, a better understanding of the interac-
tion between-host cells, agr-mutant, and wild-type strains will generate the knowl-
edge needed to confront the growing problem of complicated disease and poor
outcomes.
Other investigators showed that the use of antibiotics, such as fluoroquinolones
or beta-lactams, is a risk factor for loss of agr functionality in vitro and during treat-
ment of infection in the hospital [61, 62]. Furthermore, agr-defective strains are
associated with the development of vancomycin tolerance in vitro and during treat-
ment of patients with bacteremia in vivo, perhaps owing to defects in autolysis and
consequent changes in cell wall structure that mitigate fitness costs associated with
the evolution of vancomycin resistance (reviewed in [17]). Indeed, all known van-
comycin intermediate-resistant S. aureus (VISA) strains are agr-defective.
Moreover, recent work demonstrates that vraR, a member of the two-component
vraRS regulatory system that is upregulated in vancomycin-resistant strains, sup-
presses transcription of agr [63]. Thus, vancomycin resistance appears to be an
example of how a fitness trait that is initially dependent on attenuation of agr can
evolve transcriptional independence.
Enhanced fitness when agr function is compromised can enrich underlying poly-
morphisms in the locus such that a threshold is passed and the phenotype is expressed
in all members of the population. The effect of antimicrobials on selection for agr-
defective strains is discussed further in sections below in a different context – dimin-
ished antimicrobial lethality (tolerance) rather than inhibition of growth (resistance).
We conclude this section by noting that agr dysfunction offers potential advantages
438 B. Shopsin and R. Copin
to S. aureus not just during infection but also more generally by promoting protec-
tion against antimicrobials. Indeed, loss of agr expression has been described as a
potential “win-win” situation for a nosocomial pathogen [64].
a reservoir for relapse and the evolution to antibiotic resistance. Thus, understanding
tolerance is critical for addressing the decreasing efficacy of antibiotics. We note that,
although definitions have been debated, tolerance is related to but distinct from
another important contributor to pathogen survival – the phenomenon of persistence.
Persisters are considered to be slow growing, metabolically dormant cells that exhibit
tolerance and are less likely than growing cells to exhibit ROS-mediated killing by
antimicrobials (for additional discussion of tolerance, see Chap. 13).
One oxidative stress-protective mechanism that might promote antimicrobial tol-
erance in S. aureus involves mutation of agr, which has a built-in oxidation-sensing
mechanism through an intramolecular disulfide switch possessed by the DNA-
binding domain of the response regulator AgrA [69]. Oxidation of AgrA decreases
DNA-binding activity, which results in derepression of the bsaA gene, which
encodes the antioxidant glutathione peroxidase. As a result, agr-defective mutants
are less susceptible to oxidative stress.
The frequent occurrence of in vivo-selected agr-defective mutants during persis-
tent infection highlights a possible link between oxidative stress and antibiotic toler-
ance in this organism. The mechanism underlying agr dysfunction among strains
derived from serial passage in vitro and from clinical isolates is almost always traced
to inactivating mutations in agrC and agrA, the sensor component and response
regulator, respectively, of the agr system (Fig. 14.2). The intuition explaining this
observation is that selection for agr-defective strains occurs in mixtures containing
agr-positive parental strains. Accordingly, inactivation of agrD or agrB does not
silence agr owing to the production of autoinducing peptide in trans by the agr-
positive strain. However, this scenario does not explain why RNAIII, the effector of
the agr response, is not targeted by selection for loss of agr function. We hypothesize
that an agrA-dependent, bsaA-mediated antioxidant phenotype provides protection
against antibiotic-dependent oxidative damage, thereby resolving the dilemma.
A C D B RNAIII
Fig. 14.2 Localization of inactivating mutations in agr. The mechanism underlying agr dysfunc-
tion can usually be traced to mutations in agr that inactivate the locus [18, 22, 27, 29, 66, 70, 71].
Representative mutations identified by DNA sequencing of the agr operon in different clinical
isolates that were negative for δ-hemolysin. (Adapted from [18]). The numbers on the figure refer
to the location of agr mutations for different isolates. The isolates were derived from patients with
various clinical infections (the strain number is an arbitrary designation). Some strains had more
than one mutation in agr, but complementation with the relevant gene on a plasmid showed that
only one mutation per strain resulted in agr inactivation. Notably, agr defects in S. aureus and
Staphylococcus epidermidis usually result from quorum-sensing deficiency (due to a mutation in
agrA or agrC) rather than from quorum-signaling deficiency (due to a mutation in agrB or agrD).
Presumably, selection for agr-defective strains occurs in mixtures with agr-positive parental
strains. Thus, inactivation of agrD or agrB cannot silence agr owing to the production of AIP in
trans by the agr-positive strain
440 B. Shopsin and R. Copin
Collectively, the data support the hypothesis that inactivation of agrA can result
in degradation of both antimicrobials themselves and the lethal response to
antimicrobial-mediated stress. Given the data indicating that fitness gains are asso-
ciated with inactivation of RNAIII (mentioned above, [42]), we conclude that two
distinct subsets of agr antimicrobial fitness exist: an RNAIII-independent one that
impacts antimicrobial lethality and an RNAIII-dependent form that controls
antimicrobial-associated fitness for growth (Fig. 14.3).
Fig. 14.3 Overview of agr mutation and its consequences in complex host environments. Effect
of agr deficiency on sublethal and lethal action. (A) Sublethal stress. By switching to the inactive
form and modulating expression of appropriate factors, S. aureus cells gain enhanced replicative
fitness (RNAIII pathway). In this model, RNAIII deficiency enhances energy resources. Decreased
protein synthesis and ATP could be involved in this coupling; however, as with many factors in the
agrA pathway, the physiological relevance, as well as the mechanism by which these factors act, is
poorly understood. (B) Lethal stress. agr deficiency modulates survival against lethal stress (agrA
pathway). agrA-mediated effects are stress dependent, for example, derepression of the antioxidant
bsaA results in protection against ciprofloxacin, whereas upregulation of unknown factors control
survival in trans against gentamicin (not shown). For both RNAIII and agrA pathways, enhanced
survivability is achieved at the cost of virulence, which may or may not be compensated for by the
presence of coinfecting agr-positive strains
442 B. Shopsin and R. Copin
Given the metabolic burden associated with agr function, the question arises as
to whether it is advantageous for S. aureus populations to consist purely of
signaling-proficient cells or whether there might be situations in which “cheat-
ers” would be favored. Social cheaters reap the benefit of public goods while
contributing less than average to the cost [76, 77]. In S. aureus, many agr-regu-
lated products are released into the extracellular environment and benefit not
only the producing cell but also its neighbors. Mutants that do not respond to
agr autoinducing peptide signals do not incur the cost of producing these “pub-
lic goods,” but they may gain the benefit of production of goods shared by
neighbors. Cheating theory therefore predicts that agr-defective cells should be
at a disadvantage to their wild-type counterparts when grown in monoculture. In
support of this hypothesis, it is well known that agr-defective cells are rapidly
eliminated and are less virulent in animal models of acute infection [14, 23, 78,
79]. Indeed, it was unequivocally demonstrated that blocking of agr attenuates
staphylococcal virulence and that the administration of an agr-positive superna-
tant along with agr-defective organisms protects the bacteria in an abscess
model of infection [14]. This suggested that the spread of agr-defective strains
within populations in vivo is due in part to the exploitation of shared products
produced by their wild-type neighbors. Thus, social cheating may be relevant
during infection owing to the differential impact of host defenses on bacterial
survival.
Evolution by natural selection involves two main steps: the generation of heritable
variations (e.g., mutations) and the differential proliferation of the variants in the
environment. Hypermutability may function to create a diverse bacterial population,
increasing the likelihood of environmentally adapted variants. Thus, enhanced
mutability may provide the substrate for selection of attenuated agr-mediated viru-
lence in S. aureus infection. However, recent work by Plata et al. suggests that heri-
table elevations of mutation frequency are not likely the cause of agr diversification:
agr-defective mutants, and their parent strains showed similar mutation frequencies
in the range of what is commonly found in the species [80]. Nonetheless, the authors
reported that generation of heterogeneous resistance to oxacillin is enhanced by agr
mutation and antimicrobial-related stress [80], giving rise to the related ideas that
(1) agr suppresses genome plasticity and (2) agr dysfunction can result in bursts of
mutations. In this scenario, agr dysfunction potentially serves as a “driver” muta-
tion that promotes accumulation of additional genetic alterations and rapid evolu-
tion in the complex environmental milieu of invasive infection. agr inactivation and
genetic instability in turn may result in subclones that display emergent properties,
14 Staphylococcus aureus Adaptation During Infection 443
The observations described above provide an entry point for additional screening to
identify other bacterial regulators having activities that can be self-protective,
depending on the type and level of lethal stress. Elucidating the basis of such effects
can be clinically significant when they inform efforts to personalize management of
antimicrobials through pathogen strain-specific characteristics. For example, the
use of anti-agr agents or therapeutic vaccines [26] may be counter-productive for
applications in which the absence of agr reduces lethal activity of an antimicrobial.
Likewise, identification of adaptations that erode the lethal activities of antimicrobi-
als might inform the development of novel strategies to selectively bolster antimi-
crobial effectiveness [82–84].
60
118
40
263
20
0
High cytotoxicity Low cytotoxicity
Fig. 14.4 Neutrophil cytotoxity values for CA-MRSA (orange) and HA-MRSA (blue) strains. In
our recent survey of more than 380 MRSA isolates from patients enrolled in a randomized con-
trolled trial of nosocomial pneumonia [85, 86], CA-MRSA lineages were found to be almost uni-
formly highly cytotoxic to human neutrophils (82% vs 11%; p < 0.0001). In contrast, HA-MRSA
lineages were weakly cytotoxic, often below levels expected to correlate with mortality in animals.
Furthermore, many strains (22/381 [6%] HA-MRSA) produced no detectable cytotoxic activity.
Left, overall, 69% (263/381) of isolates were assigned to CA- or HA-MRSA subsets based on
SCCmec type. More than 75% of CA-MRSA demonstrated ≥90% killing of neutrophils within
2 h. In contrast, ≥90% of HA-MRSA strains demonstrated cytotoxicity levels below 90%
Hospital adaptation
Hospital-associated Communiy-associated
S. aureus S. aureus
Antimicrobial
resistance/ Virulence
tolerance
Inflammation
Response to treatment
Fig. 14.5 Variations in resistance and virulence, represented by the continuum symbol, influence
the balance between stimulation of proinflammatory host responses and antimicrobial-tolerant
states. High levels of virulence but low antimicrobial resistance results in immune and antimicro-
bial clearance among hospitalized patients with bacteremia. The same characteristics result in
exuberant bacterial growth and increased inflammation among hosts in the community [91–94]. In
the community, lack of prompt antimicrobial treatment leads to severe inflammation, tissue necro-
sis, and poor outcomes. Although lower levels of virulence result in decreased inflammation during
infection, high tolerance to antimicrobials and lack of effective immune responses may still result
in high bacterial burden and poor outcomes among hospitalized patients
We now know that bacterial isolates from invasive infection often differ pheno-
typically from naturally occurring strains that colonize subjects in the commu-
nity, exhibiting high frequencies of decreased cytotoxicity and reduced agr
activity. Moreover, they are associated with worse outcomes [24, 25, 38, 70,
87]. Colonizing isolates represent the source for disease [28], and isolates in
hospital-associated settings are the reservoir for low-virulence phenotypes [27].
A better understanding of low-virulence strains is critical not only to evaluate
therapeutic efficacy of vaccine and monoclonal antibody strategies under
14 Staphylococcus aureus Adaptation During Infection 447
development for prophylaxis and therapy but also for the important first step of
target identification. To date, efforts to define the polymorphism and expression
heterogeneity of antigens associated with targets of protection have been mostly
limited to individual agr-regulated toxin antigens, such as hla, that are expressed
poorly, if at all, among HA-MRSA lineages that are implicated in complicated
infections, such as bacteremia [24, 89, 90]. However, it is possible that adaptive
variants result from changes that do not alter agr sequences or expression. To
test for non-agr specific selection processes, additional mutations can be sought
using genome sequencing to define the complete set of changes that accompany
the transition to host and hospital-adapted phenotypes.
Recent studies have used genome sequencing to provide evidence for rapid
pathogen genome diversification, some of which could potentially affect the course
of disease (reviewed in [95]). Much of the variation described has been measured
between bacterial isolates of a single patient during infection, either at a single time
point or longitudinally. Although longitudinal studies in single patients have been
highly informative (e.g., [21]), they often focus on resistance mutations during anti-
biotic exposure; effects of selection are superimposed upon mutational patterns
generated by antimicrobial damage and repair processes. In contrast, the overlap
between the within-host and the between-host levels that defines adaptation to the
hospital environment results in heterogeneity of selective pressures that derive from
from functional trade-offs, wherein mutations in genes such as agr may promote
survival in certain host niches in the short term but represent a liability to clones
over the longer term. A consequence of the differential populational stability of
mutations is that pathogens evolve toward fitness on a global scale. Thus, to under-
stand adaptation to a complex environment, such as the hospital, we must examine
evolution in populations as well as in single patients.
Divergence of HA-MRSA strains from community isolates occurred decades
ago, complicating the identification of specific genetic changes associated with
the transition from community to hospital. CA-MRSA strains, such as USA300,
the predominant CA-MRSA clone in the United States, are only distantly related
to hospital-associated MRSA; until recently, they were not linked to the health-
care environment [96, 97]. To limit confounding effects of genetic variability,
genetic background can be controlled by leveraging the recent introduction and
dissemination of CA-MRSA strain USA300 into hospitals. The limited number
of functional mutations associated with this lineage helps identify specific vari-
ants that are associated with the evolution to low-virulence phenotypes in hospi-
tals. Using such an approach, recent genomic studies demonstrate that a number
of MRSA genes, including but not limited to agr, are targets for mutation during
hospital adaptation [70, 98]. Among the non-agr genes are sucD, clpC, tsK, and
rpsA [70].
Identification of genetic changes that potentially drive or are associated with
emergence of host- and hosptial-adapted strains highlight the need to rapidly
reduce bacterial burden and restrict the emergence of adapted mutants, as well
as vigorous surveillance and early public health intervention to limit further
adaptation. Additionally, the mutations that accumulate during adaptation pro-
448 B. Shopsin and R. Copin
Although genome sequencing has improved our understanding of the mutations and
genetic polymorphism associated with S. aureus strains isolated from patients dur-
ing invasive and hospital-associated infections [99–101], the contribution of mobile
genetic elements (MGE) remains under-investigated. MGE discovery and identifi-
cation are important goals for genome analysis because almost all S. aureus strains
harbor one or more MGE with potentially syndrome-specific and tissue-specific
functions. Staphylococcal MGEs include plasmids, transposons, integrons, genomic
islands, S. aureus pathogenicity islands (SaPIs), integrative conjugative elements,
staphylococcal chromosome cassettes, and phages [102–104]. MGEs are abundant,
representing 15–20% of the S. aureus genome. Staphylococcal MGEs are thus an
important source of variation in S. aureus strains and remain, for the most part,
functionally uncharacterized. Together, phages and plasmids are the main source of
MGE diversity among S. aureus strains [105].
Phages are bacterial viruses that can integrate into the S. aureus chromosome. As
the primary vehicles for horizontal transfer of MGEs between strains, phages are
major drivers of staphylococcal genome evolution [106]. All S. aureus genomes
sequenced to date contain at least one phage, and many strains carry up to four
[107]. In addition to numerous hypothetical proteins, phages carry antibiotic resis-
tance genes and virulence factors, such as the Panton-Valentine leukocidin (PVL)
toxin [108, 109]. Phages also mobilize other MGEs, such as plasmids and staphylo-
coccal pathogenicity islands (SaPIs) [103]. SaPIs, which integrate into specific
attachment sites in the chromosome, employ phage machinery for replication and
dissemination [102]. Thus, phages impact both the virulence and adaptability of S.
aureus.
Genomic comparisons between successful S. aureus clones and distantly
related strains have provided insights into MGE contribution to S. aureus adap-
tation [110–112]. For example, genomic studies have pinpointed the dramatic
impact of MGE on the success of MRSA clones, such as USA300, the predomi-
nant CA-MRSA clone in the United States [113]. Notably, they highlighted the
importance of the association between phage-derived PVL with purulent skin
and soft tissue infection and with necrotizing pneumonia and other syndrome-
specific MGE, such as the arginine catabolic mobile element [111, 112]. Other
14 Staphylococcus aureus Adaptation During Infection 449
Even modest genetic changes in a bacterial pathogen can have a dramatic impact on
virulence factor expression and the host-pathogen interaction [121–123]. Although
the presence or absence of a particular gene or polymorphism can be determined,
we do not yet have tools to readily extract information about gene expression in an
organism from the aggregate of its sequence data alone. Thus, given that multiple
mutations and MGEs are involved in the process of host and hospital adaptation, it
is not obvious that an understanding of any one mutation or change will be sufficient
to define what constitutes a hospital-adapted strain.
The research challenge posed by these observations is to reduce the heterogeneity
of hospital adaptation so that we can identify a proxy for hospital adaptation that can
explain epidemiological patterns and clinical outcomes. To augment mutation discov-
ery, we propose a novel approach focused on the phenotypic effects of mutation: anal-
ysis of gene expression, proteomic, and metabolic pathways using information-rich
datasets derived from low-virulence MRSA populations. Once identified, pathways
can be directly assessed for phenotypic and genetic variation in individual strains.
The metabolic underpinnings of bacterial virulence, host defense, fitness, and
antimicrobial lethality underscore metabolic capacity as an attractive candidate to
determine the ability of MRSA to adapt to the hospital environment. The frequent
occurrence of dysfunctional tricarboxylic acid (TCA) cycle activity in clinical iso-
lates of staphylococci supports the idea that genetic adaptation can redirect metabo-
lism [124], as does the observation that downregulation of TCA cycle activity is
common among small colony-variant phenotypes that are associated with chronic
infections and antimicrobial resistance [125–127].
These observations give rise to the hypothesis that MRSA use mutation to fine-
tune their metabolism to sustain fitness in response to selective pressures in the
hospital environment. As CA-MRSA become HA-MRSA, metabolic heterogeneity
ensues. Our working model is shown in Fig. 14.6.
Fig. 14.6 Differential susceptibility hypothesis of nosocomial adaptation. Current work focuses
on resistant and susceptible S. aureus; we hypothesize that S. aureus displays a spectrum of meta-
bolic states that causes strains to differ in the degree to which they tolerate host and antimicrobial
stress. Hospital-adaptation is accompanied by distinct stress tolerance states that are associated
with distinct metabolic states, and these metabolic states can be identified as signatures that define
specificity for different pathways to metabolize substrates, as well as differences in virulence phe-
notypes and functions. Differential susceptibility of an organism to stress is sometimes called the
orchid hypothesis. In the figure, the boundary between community and hospital-adapted states is
not distinct; however, CA-MRSA are like delicate orchids; and HA-MRSA are like dandelions.
The former are highly virulent, and if left untreated causing more severe disease, they quickly
wither if exposed to antimicrobial stress. The latter prove resilient to the effects of antimicrobials,
but at the same time, they are not particularly virulent. CA-MRSA represents a “high” metabolic
state, although high-metabolic states may be favored during certain hospital environments, which
can account for inter-strain variability. In this scenario, the Goldilocks metaphor is applicable;
extreme states impair fitness and are considered either deleterious or “short-sighted.” Metabolism
refers to growth rate and metabolic “complexity”
changes that have a profound impact on the S. aureus transcriptome and on viru-
lence has important implications for the study of gene regulation in the pathogenic-
ity of CA-MRSA infection. For example, differences in pathology observed in
infections caused by mutants constructed in the laboratory may be obscured or com-
pounded by a failure to undergo mutation of agr to one form or another: only careful
comparison of strains going into and coming out of a host can identify the extent to
which this is a problem. Thus, field work is essential. By unraveling the evolution-
ary steps occurring during infection and characterizing the genetic basis of such
changes, we expect to provide a framework for interpreting the phenotype of newly
emergent clones.
Major Points
• Functional compromise of pleiotropic regulators, which control many genes in
complex genetic networks, provides a broad framework in which alterations in
cellular circuitry promote the emergence of new traits.
• Mutation of agr is an explicit example of a mechanism by which hospital adapta-
tion can alter the relationship between virulence, infectivity, and host and syn-
thetic antimicrobial susceptibility.
• Mechanisms of adaptation-dependent changes in antimicrobial susceptibility
include direct and indirect effects on fitness and antimicrobial tolerance.
• Mutations in genes such as agr may be adaptive for survival in the infected host,
but they are counter-adaptive outside infected host tissues or in situtations that
involve high attack rates, such as outbreaks.
• Hospital adaptations are likely to derive from a range of heritable changes in
virulence attenuation, metabolic activity, and antimicrobial susceptibility, all of
which are linked to the oxidative stress network that participates in host and
synthetic antimicrobial-mediated killing.
• Identification of polymorphisms and targets of selection that affect the cellular
response of MRSA to antimicrobial drugs may increase our understanding of
drug tolerance mechanisms; such identification will also provide an initial step
toward predictive tailoring of drug treatments to individuals to maximize thera-
peutic benefit.
Acknowledgments We thank Karl Drlica for helpful discussions and critical comments. Our
work was supported by NIH grants AI103268 and N272201400019C.
References
1. Dwyer DJ, Belenky PA, Yang JH, et al. Antibiotics induce redox-related physiological altera-
tions as part of their lethality. Proc Natl Acad Sci U S A. 2014;111:E2100–9.
2. Dwyer DJ, Collins JJ, Walker GC. Unraveling the physiological complexities of antibiotic
lethality. Annu Rev Pharmacol Toxicol. 2015;55:313–32.
3. Zhao X, Drlica K. Reactive oxygen species and the bacterial response to lethal stress. Curr
Opin Microbiol. 2014;21:1–6.
14 Staphylococcus aureus Adaptation During Infection 453
4. Zhao X, Hong Y, Drlica K. Moving forward with reactive oxygen species involvement in
antimicrobial lethality. J Antimicrob Chemother. 2015;70:639–42.
5. Lowy FD. Staphylococcus aureus infections. N Engl J Med. 1998;339:520–32.
6. Smith TL, Pearson ML, Wilcox KR, et al. Emergence of vancomycin resistance in
Staphylococcus aureus. Glycopeptide-intermediate Staphylococcus aureus working group.
N Engl J Med. 1999;340:493–501.
7. From the Centers for Disease Control. Staphylococcus aureus resistant to vancomycin–
United States, 2002. JAMA. 2002;288:824–5.
8. From the Centers for Disease Control and Prevention. Vancomycin resistant Staphylococcus
aureus–Pennsylvania, 2002. JAMA. 2002;288:2116.
9. Novick RP. Autoinduction and signal transduction in the regulation of staphylococcal viru-
lence. Mol Microbiol. 2003;48:1429–49.
10. Cirioni O, Giacometti A, Ghiselli R, et al. Prophylactic efficacy of topical temporin A and
RNAIII-inhibiting peptide in a subcutaneous rat Pouch model of graft infection attributable to
staphylococci with intermediate resistance to glycopeptides. Circulation. 2003;108:767–71.
11. Domenico P, Gurzenda E, Giacometti A, et al. BisEDT and RIP act in synergy to prevent graft
infections by resistant staphylococci. Peptides. 2004;25:2047–53.
12. Dell'Acqua G, Giacometti A, Cirioni O, et al. Suppression of drug-resistant staphylococ-
cal infections by the quorum-sensing inhibitor RNAIII-inhibiting peptide. J Infect Dis.
2004;190:318–20.
13. Cirioni O, Giacometti A, Ghiselli R, et al. RNAIII-inhibiting peptide significantly reduces
bacterial load and enhances the effect of antibiotics in the treatment of central venous
catheter-associated Staphylococcus aureus infections. J Infect Dis. 2006;193:180–6.
14. Wright JS 3rd, Jin R, Novick RP. Transient interference with staphylococcal quorum sensing
blocks abscess formation. Proc Natl Acad Sci U S A. 2005;102:1691–6.
15. Goerke C, Matias y Papenberg S, Dasbach S, et al. Increased frequency of genomic altera-
tions in Staphylococcus aureus during chronic infection is in part due to phage mobilization.
J Infect Dis. 2004;189:724–34.
16. Painter KL, Krishna A, Wigneshweraraj S, Edwards AM. What role does the quorum-sensing
accessory gene regulator system play during Staphylococcus aureus bacteremia? Trends
Microbiol. 2014;22:676–85.
17. Sakoulas G, Moellering RC Jr, Eliopoulos GM. Adaptation of methicillin-resistant
Staphylococcus aureus in the face of vancomycin therapy. Clin Infect Dis. 2006;42(Suppl
1):S40–50.
18. Traber KE, Lee E, Benson S, et al. agr function in clinical Staphylococcus aureus isolates.
Microbiology. 2008;154:2265–74.
19. Vuong C, Kocianova S, Yao Y, Carmody AB, Otto M. Increased colonization of indwelling
medical devices by quorum-sensing mutants of Staphylococcus epidermidis in vivo. J Infect
Dis. 2004;190:1498–505.
20. Vuong C, Saenz HL, Gotz F, Otto M. Impact of the agr quorum-sensing system on adherence
to polystyrene in Staphylococcus aureus. J Infect Dis. 2000;182:1688–93.
21. Mwangi MM, Wu SW, Zhou Y, et al. Tracking the in vivo evolution of multidrug resis-
tance in Staphylococcus aureus by whole-genome sequencing. Proc Natl Acad Sci U S A.
2007;104:9451–6.
22. Traber K, Novick R. A slipped-mispairing mutation in AgrA of laboratory strains and clin-
ical isolates results in delayed activation of agr and failure to translate delta- and alpha-
haemolysins. Mol Microbiol. 2006;59:1519–30.
23. Schwan WR, Langhorne MH, Ritchie HD, Stover CK. Loss of hemolysin expression in
Staphylococcus aureus agr mutants correlates with selective survival during mixed infections
in murine abscesses and wounds. FEMS Immunol Med Microbiol. 2003;38:23–8.
24. Fowler VG Jr, Sakoulas G, McIntyre LM, et al. Persistent bacteremia due to methicillin-
resistant Staphylococcus aureus infection is associated with agr dysfunction and low-
level in vitro resistance to thrombin-induced platelet microbicidal protein. J Infect Dis.
2004;190:1140–9.
454 B. Shopsin and R. Copin
25. Schweizer ML, Furuno JP, Sakoulas G, et al. Increased mortality with accessory gene regu-
lator (agr) dysfunction in Staphylococcus aureus among bacteremic patients. Antimicrob
Agents Chemother. 2011;55:1082–7.
26. Khan BA, Yeh AJ, Cheung GY, Otto M. Investigational therapies targeting quorum-
sensing for the treatment of Staphylococcus aureus infections. Expert Opin Investig Drugs.
2015;24:689–704.
27. Shopsin B, Drlica-Wagner A, Mathema B, Adhikari RP, Kreiswirth BN, Novick
RP. Prevalence of agr dysfunction among colonizing Staphylococcus aureus strains. J Infect
Dis. 2008;198:1171–4.
28. Smyth DS, Kafer JM, Wasserman GA, et al. Nasal carriage as a source of agr-defective
Staphylococcus aureus bacteremia. J Infect Dis. 2012;206:1168–77.
29. Shopsin B, Eaton C, Wasserman GA, et al. Mutations in agr do not persist in natural popula-
tions of methicillin-resistant Staphylococcus aureus. J Infect Dis. 2010;202:1593–9.
30. Kernbauer E, Maurer K, Torres VJ, Shopsin B, Cadwell K. Gastrointestinal dissemination and
transmission of Staphylococcus aureus following bacteremia. Infect Immun. 2015;83:372–8.
31. Hunter P. The great leap forward. Major evolutionary jumps might be caused by changes in
gene regulation rather than the emergence of new genes. EMBO Rep. 2008;9:608–11.
32. Wittkopp PJ, Haerum BK, Clark AG. Evolutionary changes in cis and trans gene regulation.
Nature. 2004;430:85–8.
33. Barrier M, Robichaux RH, Purugganan MD. Accelerated regulatory gene evolution in an
adaptive radiation. Proc Natl Acad Sci U S A. 2001;98:10208–13.
34. Quan S, Ray JC, Kwota Z, et al. Adaptive evolution of the lactose utilization network in
experimentally evolved populations of Escherichia coli. PLoS Genet. 2012;8:e1002444.
35. Saxer G, Krepps MD, Merkley ED, et al. Mutations in global regulators lead to metabolic
selection during adaptation to complex environments. PLoS Genet. 2014;10:e1004872.
36. Spencer CC, Bertrand M, Travisano M, Doebeli M. Adaptive diversification in genes that
regulate resource use in Escherichia coli. PLoS Genet. 2007;3:e15.
37. Carter MQ, Parker CT, Louie JW, Huynh S, Fagerquist CK, Mandrell RE. RcsB contrib-
utes to the distinct stress fitness among Escherichia coli O157:H7 curli variants of the 1993
hamburger-associated outbreak strains. Appl Environ Microbiol. 2012;78:7706–19.
38. Das S, Lindemann C, Young BC, et al. Natural mutations in a Staphylococcus aureus viru-
lence regulator attenuate cytotoxicity but permit bacteremia and abscess formation. Proc Natl
Acad Sci U S A. 2016;113:E3101–10.
39. Kisiela DI, Radey M, Paul S, et al. Inactivation of transcriptional regulators during within-
household evolution of Escherichia coli. J Bacteriol. 2017;199:e00036-17.
40. Smith EE, Buckley DG, Wu Z, et al. Genetic adaptation by Pseudomonas aeruginosa to the
airways of cystic fibrosis patients. Proc Natl Acad Sci U S A. 2006;103:8487–92.
41. Balasubramanian D, Harper L, Shopsin B, Torres VJ. Staphylococcus aureus pathogenesis in
diverse host environments. Pathog Dis. 2017;75(1):ftx005.
42. Novick RP, Geisinger E. Quorum sensing in staphylococci. Annu Rev Genet. 2008;42:541–64.
43. D'Argenio DA, Wu M, Hoffman LR, et al. Growth phenotypes of Pseudomonas aeru-
ginosa lasR mutants adapted to the airways of cystic fibrosis patients. Mol Microbiol.
2007;64:512–33.
44. Feltner JB, Wolter DJ, Pope CE, et al. LasR variant cystic fibrosis isolates reveal an adaptable
quorum-sensing hierarchy in Pseudomonas aeruginosa. MBio. 2016;7:e01513.
45. Hoffman LR, Kulasekara HD, Emerson J, et al. Pseudomonas aeruginosa lasR mutants are
associated with cystic fibrosis lung disease progression. J Cyst Fibros. 2009;8:66–70.
46. Wilder CN, Allada G, Schuster M. Instantaneous within-patient diversity of Pseudomonas
aeruginosa quorum-sensing populations from cystic fibrosis lung infections. Infect Immun.
2009;77:5631–9.
47. Yarwood JM, Paquette KM, Tikh IB, Volper EM, Greenberg EP. Generation of virulence fac-
tor variants in Staphylococcus aureus biofilms. J Bacteriol. 2007;189:7961–7.
48. Pratten J, Foster SJ, Chan PF, Wilson M, Nair SP. Staphylococcus aureus accessory regula-
tors: expression within biofilms and effect on adhesion. Microbes Infect. 2001;3:633–7.
14 Staphylococcus aureus Adaptation During Infection 455
49. Xiong YQ, Bayer AS, Yeaman MR, Van Wamel W, Manna AC, Cheung AL. Impacts of sarA
and agr in Staphylococcus aureus strain Newman on fibronectin-binding protein A gene
expression and fibronectin adherence capacity in vitro and in experimental infective endocar-
ditis. Infect Immun. 2004;72:1832–6.
50. Sakoulas G, Eliopoulos GM, Fowler VG Jr, et al. Reduced susceptibility of Staphylococcus
aureus to vancomycin and platelet microbicidal protein correlates with defective autoly-
sis and loss of accessory gene regulator (agr) function. Antimicrob Agents Chemother.
2005;49:2687–92.
51. DuMont AL, Yoong P, Surewaard BG, et al. Staphylococcus aureus elaborates leukocidin AB
to mediate escape from within human neutrophils. Infect Immun. 2013;81:1830–41.
52. Koziel J, Maciag-Gudowska A, Mikolajczyk T, et al. Phagocytosis of Staphylococcus aureus
by macrophages exerts cytoprotective effects manifested by the upregulation of antiapoptotic
factors. PLoS One. 2009;4:e5210.
53. Kubica M, Guzik K, Koziel J, et al. A potential new pathway for Staphylococcus aureus
dissemination: the silent survival of S. aureus phagocytosed by human monocyte-derived
macrophages. PLoS One. 2008;3:e1409.
54. Lehar SM, Pillow T, Xu M, et al. Novel antibody-antibiotic conjugate eliminates intracellular
S. aureus. Nature. 2015;527:323–8.
55. Surewaard BG, Deniset JF, Zemp FJ, et al. Identification and treatment of the Staphylococcus
aureus reservoir in vivo. J Exp Med. 2016;213:1141–51.
56. Zeng Z, Surewaard BG, Wong CH, Geoghegan JA, Jenne CN, Kubes P. CRIg functions as
a macrophage pattern recognition receptor to directly bind and capture blood-borne gram-
positive bacteria. Cell Host Microbe. 2016;20:99–106.
57. Fraunholz M, Sinha B. Intracellular Staphylococcus aureus: live-in and let die. Front Cell
Infect Microbiol. 2012;2:43.
58. Soong G, Paulino F, Wachtel S, et al. Methicillin-resistant Staphylococcus aureus adaptation
to human keratinocytes. MBio. 2015;6:e00289-15.
59. Wickersham M, Wachtel S, Wong Fok Lung T, et al. Metabolic stress drives keratinocyte
defenses against Staphylococcus aureus infection. Cell Rep. 2017;18:2742–51.
60. Tsuji BT, MacLean RD, Dresser LD, McGavin MJ, Simor AE. Impact of accessory gene
regulator (agr) dysfunction on vancomycin pharmacodynamics among Canadian community
and health-care associated methicillin-resistant Staphylococcus aureus. Ann Clin Microbiol
Antimicrob. 2011;10:20.
61. Butterfield J, Tsuji B, Brown J, et al. Predictors of agr dysfunction in methicillin-resistant
Staphylococcus aureus (MRSA) isolates among patients with MRSA bloodstream infec-
tions in the “15-20 mg/L” target vancomycin trough era. Antimicrob Agents Chemother.
2011;55:5433.
62. Paulander W, Nissen Varming A, Baek KT, Haaber J, Frees D, Ingmer H. Antibiotic-mediated
selection of quorum-sensing-negative. MBio. 2013;3:e00459–12.
63. Dai Y, Chang W, Zhao C, et al. VraR binding to the promoter region of agr inhibits its
function in vancomycin-intermediate Staphylococcus aureus (VISA) and heterogeneous
VISA. Antimicrob Agents Chemother. 2017;61:e02740.
64. Sakoulas G, Moise PA, Rybak MJ. Accessory gene regulator dysfunction: an advantage for
Staphylococcus aureus in health-care settings? J Infect Dis. 2009;199:1558–9.
65. Mishra NN, Bayer AS, Moise PA, Yeaman MR, Sakoulas G. Reduced susceptibility to host-
defense cationic peptides and daptomycin coemerge in methicillin-resistant Staphylococcus
aureus from daptomycin-naive bacteremic patients. J Infect Dis. 2012;206:1160–7.
66. Somerville GA, Beres SB, Fitzgerald JR, et al. In vitro serial passage of Staphylococcus
aureus: changes in physiology, virulence factor production, and agr nucleotide sequence.
J Bacteriol. 2002;184:1430–7.
67. Dunman PM, Murphy E, Haney S, et al. Transcription profiling-based identifica-
tion of Staphylococcus aureus genes regulated by the agr and/or sarA loci. J Bacteriol.
2001;183:7341–53.
456 B. Shopsin and R. Copin
68. Zorzet A, Andersen JM, Nilsson AI, Moller NF, Andersson DI. Compensatory mutations in
agrC partly restore fitness in vitro to peptide deformylase inhibitor-resistant Staphylococcus
aureus. J Antimicrob Chemother. 2012;67:1835–42.
69. Sun F, Liang H, Kong X, et al. Quorum-sensing agr mediates bacterial oxidation response via
an intramolecular disulfide redox switch in the response regulator AgrA. Proc Natl Acad Sci
U S A. 2012;109:9095–100.
70. Laabei M, Uhlemann AC, Lowy FD, et al. Evolutionary trade-offs underlie the multi-faceted
virulence of Staphylococcus aureus. PLoS Biol. 2015;13:e1002229.
71. McNamara PJ, Milligan-Monroe KC, Khalili S, Proctor RA. Identification, cloning, and ini-
tial characterization of rot, a locus encoding a regulator of virulence factor expression in
Staphylococcus aureus. J Bacteriol. 2000;182:3197–203.
72. Kumar K, Chen J, Drlica K, Shopsin B. Tuning of the lethal response to multiple stress-
ors with a single-site mutation during clinical infection by Staphylococcus aureus. MBio.
2017;8:e01476.
73. Liu Y, Liu X, Qu Y, Wang X, Li L, Zhao X. Inhibitors of reactive oxygen species accumula-
tion delay and/or reduce the lethality of several antistaphylococcal agents. Antimicrob Agents
Chemother. 2012;56:6048–50.
74. Pader V, Hakim S, Painter KL, Wigneshweraraj S, Clarke TB, Edwards AM. Staphylococcus
aureus inactivates daptomycin by releasing membrane phospholipids. Nat Microbiol.
2016;2:16194.
75. Queck SY, Jameson-Lee M, Villaruz AE, et al. RNAIII-independent target gene control
by the agr quorum-sensing system: insight into the evolution of virulence regulation in
Staphylococcus aureus. Mol Cell. 2008;32:150–8.
76. Keller L, Surette MG. Communication in bacteria: an ecological and evolutionary perspec-
tive. Nat Rev Microbiol. 2006;4:249–58.
77. Sandoz KM, Mitzimberg SM, Schuster M. Social cheating in Pseudomonas aeruginosa quo-
rum sensing. Proc Natl Acad Sci U S A. 2007;104:15876–81.
78. Booth MC, Cheung AL, Hatter KL, Jett BD, Callegan MC, Gilmore MS. Staphylococcal
accessory regulator (sar) in conjunction with agr contributes to Staphylococcus aureus viru-
lence in endophthalmitis. Infect Immun. 1997;65:1550–6.
79. Gillaspy AF, Hickmon SG, Skinner RA, Thomas JR, Nelson CL, Smeltzer MS. Role of
the accessory gene regulator (agr) in pathogenesis of staphylococcal osteomyelitis. Infect
Immun. 1995;63:3373–80.
80. Plata KB, Rosato RR, Rosato AE. Fate of mutation rate depends on agr locus expression
during oxacillin-mediated heterogeneous-homogeneous selection in methicillin-resistant
Staphylococcus aureus clinical strains. Antimicrob Agents Chemother. 2011;55:3176–86.
81. Malik M, Hoatam G, Chavda K, Kerns RJ, Drlica K. Novel approach for comparing the abili-
ties of quinolones to restrict the emergence of resistant mutants during quinolone exposure.
Antimicrob Agents Chemother. 2010;54:149–56.
82. Gusarov I, Shatalin K, Starodubtseva M, Nudler E. Endogenous nitric oxide protects bacteria
against a wide spectrum of antibiotics. Science. 2009;325:1380–4.
83. Liu Y, Zhou J, Qu Y, et al. Resveratrol antagonizes antimicrobial lethality and stimulates
recovery of bacterial mutants. PLoS One. 2016;11:e0153023.
84. Shatalin K, Shatalina E, Mironov A, Nudler E. H2S: a universal defense against antibiotics in
bacteria. Science. 2011;334:986–90.
85. Rose HR, Holzman RS, Altman DR, et al. Cytotoxic virulence predicts mortality in nos-
ocomial pneumonia due to methicillin-resistant Staphylococcus aureus. J Infect Dis.
2015;211:1862–74.
86. Mendes RE, Deshpande LM, Smyth DS, Shopsin B, Farrell DJ, Jones RN. Characterization
of methicillin-resistant Staphylococcus aureus strains recovered from a phase IV clinical trial
for linezolid versus vancomycin for treatment of nosocomial pneumonia. J Clin Microbiol.
2012;50:3694–702.
87. Baines SL, Holt KE, Schultz MB, et al. Convergent adaptation in the dominant global hospi-
tal clone ST239 of methicillin-resistant. MBio. 2015;6:e00080.
14 Staphylococcus aureus Adaptation During Infection 457
88. Cheung GY, Kretschmer D, Duong AC, et al. Production of an attenuated phenol-soluble
modulin variant unique to the MRSA clonal complex 30 increases severity of bloodstream
infection. PLoS Pathog. 2014;10:e1004298.
89. Deleo FR, Kennedy AD, Chen L, et al. Molecular differentiation of historic phage-type
80/81 and contemporary epidemic Staphylococcus aureus. Proc Natl Acad Sci U S A.
2011;108:18091–6.
90. Fowler VG Jr, Nelson CL, McIntyre LM, et al. Potential associations between hematoge-
nous complications and bacterial genotype in Staphylococcus aureus infection. J Infect Dis.
2007;196:738–47.
91. Alegre ML, Chen L, David MZ, et al. Impact of Staphylococcus aureus USA300 colo-
nization and skin infections on systemic immune responses in humans. J Immunol.
2016;197(4):1118–26
92. Malachowa N, Kobayashi SD, Braughton KR, et al. Staphylococcus aureus leukotoxin GH
promotes inflammation. J Infect Dis. 2012;206:1185–93.
93. Malachowa N, Kobayashi SD, Freedman B, Dorward DW, DeLeo FR. Staphylococcus
aureus leukotoxin GH promotes formation of neutrophil extracellular traps. J Immunol.
2013;191:6022–9.
94. Yoong P, Pier GB. Immune-activating properties of Panton-valentine leukocidin improve
the outcome in a model of methicillin-resistant Staphylococcus aureus pneumonia. Infect
Immun. 2012;80:2894–904.
95. Didelot X, Walker AS, Peto TE, Crook DW, Wilson DJ. Within-host evolution of bacterial
pathogens. Nat Rev Microbiol. 2016;14:150–62.
96. Popovich KJ, Weinstein RA, Hota B. Are community-associated methicillin-resistant
Staphylococcus aureus (MRSA) strains replacing traditional nosocomial MRSA strains? Clin
Infect Dis. 2008;46:787–94.
97. Seybold U, Kourbatova EV, Johnson JG, et al. Emergence of community-associated
methicillin-resistant Staphylococcus aureus USA300 genotype as a major cause of health
care-associated blood stream infections. Clin Infect Dis. 2006;42:647–56.
98. Laabei M, Recker M, Rudkin JK, et al. Predicting the virulence of MRSA from its genome
sequence. Genome Res. 2014;24:839–49.
99. Harris SR, Cartwright EJ, Torok ME, et al. Whole-genome sequencing for analysis of an
outbreak of meticillin-resistant Staphylococcus aureus: a descriptive study. Lancet Infect Dis.
2013;13:130–6.
100. Planet PJ, Diaz L, Kolokotronis SO, et al. Parallel epidemics of community-associated
methicillin-resistant Staphylococcus aureus USA300 infection in North and South America.
J Infect Dis. 2015;212:1874–82.
101. Tong SY, Holden MT, Nickerson EK, et al. Genome sequencing defines phylogeny and
spread of methicillin-resistant Staphylococcus aureus in a high transmission setting. Genome
Res. 2015;25:111–8.
102. Penades JR, Christie GE. The phage-inducible chromosomal islands: a family of highly
evolved molecular parasites. Annu Rev Virol. 2015;2:181–201.
103. Novick RP, Christie GE, Penades JR. The phage-related chromosomal islands of gram-
positive bacteria. Nat Rev Microbiol. 2010;8:541–51.
104. Malachowa N, DeLeo FR. Mobile genetic elements of Staphylococcus aureus. Cell Mol Life
Sci. 2010;67:3057–71.
105. Copin R, Shopsin B, Torres JV. After de deluge: mining Staphylococcus aureus genomic
data for clinical associations and host pathogen interactions. Curr Opin Microbiol.
2018;41C:43–50.
106. Xia G, Wolz C. Phages of Staphylococcus aureus and their impact on host evolution. Infect
Genet Evol. 2014;21:593–601.
107. Lindsay JA. Genomic variation and evolution of Staphylococcus aureus. Int J Med Microbiol.
2010;300:98–103.
108. Haaber J, Leisner JJ, Cohn MT, et al. Bacterial viruses enable their host to acquire antibiotic
resistance genes from neighbouring cells. Nat Commun. 2016;7:13333.
458 B. Shopsin and R. Copin
109. Nguyen LT, Vogel HJ. Staphylokinase has distinct modes of interaction with antimicrobial
peptides, modulating its plasminogen-activation properties. Sci Rep. 2016;6:31817.
110. Everitt RG, Didelot X, Batty EM, et al. Mobile elements drive recombination hotspots in the
core genome of Staphylococcus aureus. Nat Commun. 2014;5:3956.
111. Gillet Y, Issartel B, Vanhems P, et al. Association between Staphylococcus aureus strains car-
rying gene for Panton-valentine leukocidin and highly lethal necrotising pneumonia in young
immunocompetent patients. Lancet. 2002;359:753–9.
112. Goering RV, McDougal LK, Fosheim GE, Bonnstetter KK, Wolter DJ, Tenover
FC. Epidemiologic distribution of the arginine catabolic mobile element among selected
methicillin-resistant and methicillin-susceptible Staphylococcus aureus isolates. J Clin
Microbiol. 2007;45:1981–4.
113. Diep BA, Gill SR, Chang RF, et al. Complete genome sequence of USA300, an epi-
demic clone of community-acquired meticillin-resistant Staphylococcus aureus. Lancet.
2006;367:731–9.
114. Rolo J, Worning P, Boye Nielsen J, et al. Evidence for the evolutionary steps leading to
mecA-mediated beta-lactam resistance in staphylococci. PLoS Genet. 2017;e1006674:13.
115. Salgado-Pabon W, Herrera A, Vu BG, et al. Staphylococcus aureus beta-toxin production
is common in strains with the beta-toxin gene inactivated by bacteriophage. J Infect Dis.
2014;210:784–92.
116. Herrera A, Kulhankova K, Sonkar VK, et al. Staphylococcal beta-toxin modulates human
aortic endothelial cell and platelet function through sphingomyelinase and biofilm ligase
activities. MBio. 2017;8:e00273.
117. van Wamel WJ, Rooijakkers SH, Ruyken M, van Kessel KP, van Strijp JA. The innate
immune modulators staphylococcal complement inhibitor and chemotaxis inhibitory pro-
tein of Staphylococcus aureus are located on beta-hemolysin-converting bacteriophages.
J Bacteriol. 2006;188:1310–5.
118. Goerke C, Gressinger M, Endler K, et al. High phenotypic diversity in infecting but not in
colonizing Staphylococcus aureus populations. Environ Microbiol. 2007;9:3134–42.
119. Benson MA, Ohneck EA, Ryan C, et al. Evolution of hypervirulence by a MRSA clone
through acquisition of a transposable element. Mol Microbiol. 2014;93:664–81.
120. Altman DR, Sebra R, Hand J, et al. Transmission of methicillin-resistant Staphylococcus
aureus via deceased donor liver transplantation confirmed by whole genome sequencing. Am
J Transplant Off J Am Soc Transplant Am Soc Transplant Surg. 2014;14:2640–4.
121. Beres SB, Richter EW, Nagiec MJ, et al. Molecular genetic anatomy of inter- and intrasero-
type variation in the human bacterial pathogen group A streptococcus. Proc Natl Acad Sci U
S A. 2006;103:7059–64.
122. Kennedy AD, Otto M, Braughton KR, et al. Epidemic community-associated methicillin-
resistant Staphylococcus aureus: recent clonal expansion and diversification. Proc Natl Acad
Sci U S A. 2008;105:1327–32.
123. Sumby P, Whitney AR, Graviss EA, DeLeo FR, Musser JM. Genome-wide analysis of group
a streptococci reveals a mutation that modulates global phenotype and disease specificity.
PLoS Pathog. 2006;2:e5.
124. Thomas VC, Kinkead LC, Janssen A, et al. A dysfunctional tricarboxylic acid cycle enhances
fitness of Staphylococcus epidermidis during beta-lactam stress. MBio. 2013;4:e00437-13.
125. Kriegeskorte A, Block D, Drescher M, et al. Inactivation of thyA in Staphylococcus aureus
attenuates virulence and has a strong impact on metabolism and virulence gene expression.
MBio. 2014;5:e01447-14.
126. Kriegeskorte A, Grubmuller S, Huber C, et al. Staphylococcus aureus small colony vari-
ants show common metabolic features in central metabolism irrespective of the underlying
auxotrophism. Front Cell Infect Microbiol. 2014;4:141.
127. Kriegeskorte A, Konig S, Sander G, et al. Small colony variants of Staphylococcus aureus
reveal distinct protein profiles. Proteomics. 2011;11:2476–90.
14 Staphylococcus aureus Adaptation During Infection 459
128. Allison KR, Brynildsen MP, Collins JJ. Metabolite-enabled eradication of bacterial persisters
by aminoglycosides. Nature. 2011;473:216–20.
129. Gardner SG, Marshall DD, Daum RS, Powers R, Somerville GA. Metabolic mitigation of
Staphylococcus aureus vancomycin intermediate-level susceptibility. Antimicrob Agents
Chemother. 2018;62:e01608-17.
130. Gutierrez A, Jain S, Bhargava P, Hamblin M, Lobritz MA, Collins JJ. Understanding
and sensitizing density-dependent persistence to quinolone antibiotics. Mol Cell.
2017;68:1147–54. e3.
Chapter 15
Bacterial Signal Transduction Systems
in Antimicrobial Resistance
15.1 Introduction
Bacteria must continually sample and assess their environment to survive. Once
pathogens are inside a host, they must detect important external stimuli, including
nutrient availability, oxygen levels, osmolarity, temperature, and even light quality
[1]. Pathogens must also be able to respond to assaults from other competing
microbes within the host, as well as antibacterial activity from the host itself. Indeed,
bacteria produce an entourage of antimicrobial compounds to assure that a particu-
lar niche is maintained in the presence of other bacterial competitors or, in the case
of the host, factors that are produced with the sole intent of clearing the invasive
microbe. It is important that the microbial pathogen responds to such environmental
signals only when necessary. This phenomenon is often referred to as inducible
responses. Bacteria (and most life as we know it) use inducible systems to respond
to a present threat rather than to anticipate one, as the response itself is usually
entergetically costly (e.g., transcription and translation require ATP, etc.).
Consequently, such energy, if its use is not required, is better redirected to another
task or stored for a later time. From these considerations it is thus reasonable to
conclude that antimicrobial resistance and/or tolerance can have a severe fitness
cost, and thus they must be tightly regulated to ensure the long-term survival of an
invading microbe.
In order to properly respond to their changing environment in the most efficient
manner, bacteria have evolved intricate detection mechanisms called signal trans-
duction systems. These systems are based largely on the transfer or “relay” of a
phosphate, as a posttranslational modification, from one protein moiety to another
Environmental Environmental
Stimuli Stimuli
Tyr
Kinase
HK H eSTK ST ST Y
antibiotics
D eSTP ST Y
REC ST SD
ED ED
Transcriptional response
Fig. 15.1 Bacterial signaling basics. Classical two-component signaling (TCS) involves an initial
sensory input, usually detected from outside the bacterial cell by a membrane-bound histidine
kinase (HK). This signal results in autophosphorylation of the HK on a conserved histidine (H)
residue, whose phosphate group is then transferred to a conserved aspartate residue (D) on a
receiver (REC) domain. This posttranslational modification (PTM) then results in conserved struc-
tural rearrangements that ultimately activate the effector domain (ED) of the receiver, which can
take many forms including DNA binding (most common). Eukaryotic-like serine-threonine
kinases (eSTKs), their cognate phosphatases (eSTPs), and tyrosine (Tyr) kinases can also phos-
pho-regulate TCS systems or one-component systems, the latter which have been shown to directly
bind antibiotics in many cases through a sensory domain (SD). The end result is the sensory input
initiates a transcriptional response to the external threat
fold that is almost identical to their eukaryotic homologs [6, 15], suggesting that the
mechanism by which they act is evolutionarily conserved between kingdoms.
Especially prominent, and relevant to our discussion, are the so-called PASTA (pen-
icillin-binding protein and serine/threonine kinase associated) eSTKs found in
Gram-positive pathogens. PASTA eSTKs contain repeat PASTA domains that are
suspected to directly interact with the peptidoglycan and/or free peptidoglycan frag-
ments. They serve as crucial regulatory players in bacterial cell division, cellular
stress responses and infection [6, 15]. In general, Gram-positive bacteria have only
one eSTK/eSTP pair, or a single tyrosine (Tyr) kinase, although some species have
many more. For example, Mycobacterium tuberculosis possesses 11 eSTKs (but
only a single eSTP [6]). Another difference from the two-component systems is the
pleiotropic nature of the targets for eSTKs and Tyr kinases, where many substrates
have been identified, including transcription factors ([6, 14, 15]; Fig. 15.1).
Surprisingly, these eukaryotic-like signaling proteins in bacteria also directly regu-
late TCS proteins [16]. In addition to the multitude of other signaling proteins
involved in bacterial AMR sensory systems, this observation suggests a complexity
464 A. T. Ulijasz et al.
The classic example of inducible vancomycin resistance is observed with the entero-
cocci (e.g., E. faecium and E. faecalis), where it was originally observed in the late
1980s [18]. Among the enterococci, VanA-type strains display high levels of
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 465
Capsule
muropeptides?
PASTA
PASTA
PASTA
Cell wall
PASTA
PASTA
PASTA
PASTA
PASTA
PASTA
vancomycin lipid II LL-37 and indolicidin ?
PASTA
PASTA
PASTA
Membrane
D D D D
CAMP efflux
Sythesis and cell wall VraF VraG cell wall simulon
Chromosome
-Reduced autolysis
incorporation of D-ala-D-lac -Increase in cell wall thickness
D-alanylation of cell wall teichoic acid murein hydrolases
vanH vanA vanX cell wall regulatory operon dlt operon
Lys additon to PG
MprF
central metabolism
enterococci/
Actinomycetes staphylococci
Fig. 15.2 Vancomycin resistance and tolerance signaling mechanisms in the enterococci and S.
aureus. In the enterococci, a single two-component signaling (TCS) system conveys vancomycin
resistance through the vanHAX operon. This histidine kinase VanS has been shown to likely bind
vancomycin directly to initiate a D-Ala-D-lac modification of the cell wall pentapeptide. In S.
aureus there are at least three TCS systems responsible for the VISA phenotype, WalKR, GraSR,
and VraSR, all of which are controlled through the S. aureus serine-threonine kinase Stk1 (an
eSTK). The serine-threonine phosphatase (eSTP) acts on Stk1 and its substrates for counter-regu-
lation. The histidine kinase WalK responds through contact with lipid II, and GraS binds and
responds to the human CAMP LL-37. The VraS ligand remains unknown. The ligand for Stk1
is still under debate, but is suspected to be cell wall muropeptides. Dotted lines represent direct
regulation by phosphorylation; dashed lines represent regulatory interactions with GraX. ST, ser-
ine-threonine phosphorylation; D, aspartate phosphorylation; H, histidine phosphorylation; REC,
receiver domain; DBD, DNA-binding domain; PASA, penicillin-binding and serine-threonine
kinase associated domain
Adjacent to the vanHAXYZ operon is the TCS pair VanRS, which controls the
synthesis of vanHAXYZ and therefore renders vancomycin resistance inducible
[19]. Upon addition of vancomycin, the histidine kinase VanS phosphorylates the
response regulator VanR. VanR then binds the vanHAXYZ promoter, inducing the
operon, thereby conferring resistance to the drug [20]. Deletion of VanS results in a
constitutively active phenotype [19]. To date, the precise mechanism for how van-
comycin interacts with the Gram-positive versions of VanS to induce resistance has
remained enigmatic. However, in recent years progress has been made with the
study of the VanS histidine kinase-driven system from Actinomycetes, the bacteria
that naturally produce vancomycin and therefore require their own inducible resis-
tance mechanism when manufacturing the antibiotic. Wright and colleagues were
the first to show that the Actinomycete version of VanS directly binds vancomycin
[21]. The antibiotic binds within the first 41 N-terminal residues of the protein,
which comprise the first transmembrane region plus a predicted short extracellular
sensory peptide containing the sequence DQGW. Vancomycin is predicted to inter-
act with this peptide based on biochemical data [21] (Fig. 15.3). These data hold
true for Actinomycete versions of VanS, which directly bind vancomycin and confer
intrinsic resistance (i.e., resistance to vancomycin by these native producers [21]).
On the other hand, the mechanism by which the enterococci, staphylococci, and
Gram-positive pathogens in general sense the presence of vancomycin through the
VanS histidine kinase receptor has yet to be determined. The primary amino acid
sequences of the Actinomycete VanS are divergent enough from the enterococcal
and staphylococcal versions to suggest a possible alternative mechanism (Fig. 15.3)
that might include more indirect signaling mechanisms, such as interactions with
cell envelope components. However, a recent study indicates that vancomycin and
teicoplanin might bind directly to the receptor at aromatic residues, similar to what
was observed with Actinomycetes [22]. Indeed, an alignment of VanS receptors
shows that some aromatic residues, such as the tryptophan residue proposed to
directly interact with vancomycin [21], are conserved among VanS versions
(Fig. 15.3). An interesting topic of future research would be to resolve exactly how
VanS from pathogenic Gram-positive bacteria sense the presence of vancomycin.
TM2
VanS(sc) PGTVFLRSFAPTAAWVMAFLLVFGLVGGWFLAGRMLAPLDRITEATRTAATGSLS...HR
VanS(ef ) HLDAMKLYQYSIRNNIDIFIYVAIVISILILCRVMLSKFAKYFDEINTGIDVLIQNEDKQ
consensus
VanS(sc) IRLPGRRDEYRELADAFDEMLAR..LEAHVAEQRR..FAANASHELRTPLAVSK...AIL
VanS(ef ) IELSAEMDVMEQKLNTLKRTLEKREQDAKLAEQRKNDVVMYLAHDIKTPLTSIIGYLSLL
consensus
VanS(sc) DVARTDPHQDPGEIIDRLHAVNTRAIDLTEALLLLSRAGQR..SFTREQVDMSLLAEEAT
VanS(ef ) DEAPDMPVDQKAKYVHITLDKAYRLEQLIDEFFEITRYNLQTITLTKTHIDLYYMLVQMT
consensus
VanS(sc) ETLLPFAEKHGVTLETRGHVTLALG.SPALLLQLTTNLVHNAIVHNLPGRGRVWIHTAAG
VanS(ef ) DEFYPQLSAHGKQAVIHAPEDLTVSGDPDKLARVFNNILKNAAAYSE.DNSIIDITAGLS
consensus
HATPase domain
VanS(sc) PHTTRLVVENTGDLISPHQASTLPEPFQRGTERIHTDHPGVGLGLAIVNTITQAHDGTLT
VanS(ef ) GDVVSIEFKNTG.SIPKDKLAAIFEKFYRLDNARSSDTGGAGLGLAIAKEIIVQHGGQIY
consensus
VanS(sc) LTPRHSGGLRVTVELPAAAPHTGR...
VanS(ef ) AESNDN.YTTFRVELPAMPDLVDKRRS
consensus
Fig. 15.3 Alignment of VanS from Streptomyces coelicolor (VanS(sc)) and VanS from
Enterococcus faecium (VanS(ef)). Yellow bars indicate transmembrane (TM) regions, the green
bar the vancomycin sensory loop, and the red bar the histidine kinase ATPase domain. The arrow
indicates the phosphorylated histidine. The purple highlighted area indicates the residues (DQGW)
proposed to bind vancomycin directly in VanS(sc). Black-colored residues are identical and gray
similar. For consensus,! is identical residues and * similar
ceptibility [16]. This novel discovery demonstrates that eSTKs and TCS can work
together to co-regulate essential cellular processes and antibiotic susceptibility, and
opens new opportunities and strategies for bacterial signaling perturbation to facili-
tate novel antibiotic treatments.
Beta-lactams, which are among the most widely prescribed antibiotics, were the
first to be discovered (in the 1920s by Alexander Fleming at St. Mary’s Hospital,
now part of Imperial College London) [38]. Members of this antibiotic class target
470 A. T. Ulijasz et al.
the aptly named penicillin-binding proteins (PBPs), enzymes that are responsible
for catalyzing the later steps in the assembly of the bacterial cell wall. When applied,
the beta-lactams can cause malformation of the cell wall, cell lysis, and death. Due
to the early introduction and overuse of beta-lactams, especially with Gram-positive
pathogens, resistance is now widespread. As with many cell envelope-acting antibi-
otics, beta-lactam susceptibility is controlled through bacterial signaling proteins.
S. aureus presents one of the most serious problems with respect to beta-lactam-
resistant hospital infections [10]. After the introduction of penicillin and the ensuing
development of resistance provided by penicillinase, a semisynthetic penicillinase-
resistant beta-lactam called methicillin was introduced in 1959. Within a year resis-
tant strains emerged, and MRSA was born as a problematic nosocomial pathogen
[41]. Although methicillin is no longer in clinical use, MRSA strains of S. aureus
continue to become resistant to new beta-lactams. Resistance is conferred by intro-
duction of an alternative PBP target (called PBP2a), which lowers the affinity of the
antibiotic. PBP2a is encoded by the mecA gene, which is in a mobile genetic ele-
ment that is referred to as SCCmec. MecA is under control of an inducible signaling
system, as depicted in Fig. 15.4.
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 471
BL
BL
BL
Cell wall
BL
MecA BL
BL
BL
MecA BL BL
BL
MecA BL
BL
PBD
MecR1
Membrane
MecI
Cytoplasm
DBD
MecR2
MecI MecI
Chromosome
DBD DBD
Fig. 15.4 MecA (PBP2a)-mediated regulation in S. aureus. MecI is a transcription factor that
binds to the promoter of mecA (PBP2a) and represses its transcription until it is degraded via
MecR1-mediated proteolysis. MecR1 contains a penicillin-binding domain (PBD) that senses the
beta-lactam (BL, green octagons). The binding of cell wall fragments (specifically g-D-Glu-L-Lys)
and MecR2 to MecI destabilizes its binding to DNA and enhances MecR1-mediated degradation.
Once mecA repression is remediated, MecA is produced and binds the beta-lactam to lessen drug
susceptibility
cally γ-D-Glu-L-Lys, bind the C-terminal domain of the MecI repressor and act as
a cofactor that mediates MecI cleavage and inactivation by MecR1 (Fig. 15.4).
Once produced, the mecA (PBP2a) gene product interacts with beta-lactams.
Other reports show that MecA is more complicated, as it is apparently also con-
trolled allosterically through binding the D-Ala-D-Ala terminus of the pentapep-
tide stem, likely generated by the inhibition of cell wall synthesis by beta-lactams.
The allosteric site requires occupation to then unveil the active binding pocket for
the beta-lactam [42, 43].
and CisRS systems and the molecular mechanism underlying antibiotic resistance
remain to be deciphered. Based on the differing targets and structures of the antibi-
otics that CroRS/CisRS respond to, it appears that the mechanism may involve gen-
eral sensing of cell envelope stresses.
BlrA response regulator. Standard bioinformatics analysis reveals that the closest
homolog to BlrAB is the CreBC TCS system. Interestingly, the Aeromonas CreBC
TCS, when introduced into in a heterologous host (E. coli), regulates the E. coli
beta-lactamases. This regulation occurs through the E. coli CreC response regulator
binding a conserved CreC-binding motif (TTCACnnnnnnTTCAC), which activates
gene expression. A recent report confirmed this same binding motif for CreC in
Aeromonas, and it importantly demonstrated that the CreBC system is specifically
responsive to inhibition of PBP4 by beta-lactams [53]. The authors hypothesize that
the BlrAB and CreBC TCS systems respond to peptidoglycan recycling and there-
fore levels of AmpC-inducing muropeptides that control beta-lactam susceptibility
[53]. Future work will need to elucidate the precise mechanisms underlying the
BlrAB and CreBC systems and determine whether their roles are ubiquitous among
Gram-negative pathogens.
Host CAMPs,
Polymyxins, pEtN,
other antibiotics L-Ara4N,
galactosamine
Lipid A Core O-Antigen
LPS
Cell wall Membrane
Outer
low pH
CcrB PhoQ PmrB ColS ParS CprS indolicidin
Inner
MgrB
H H H H H H
Cytoplasm
pEtN
ccrC pmrD eptA oprD
Chromosome
mgrR L-Ara4N
pEtN (E. coli) arn operon
mexX mexY
eptB
galactosamine
(A. baumannii) pEtN (E. coli)
eptA
naxD
pEtN (K. pneumoniae, A. baumannii)
mgrB pmrC pmrA pmrB
Fig. 15.5 The complex signaling pathways of Gram-negative polymyxin and CAMP resistance.
TCS systems shown in gray (ColR, ParR, and CprR) are found in P. aeruginosa. The CcrABC
system and MgrB are found in K. pneumoniae. Small regulatory RNAs micA and mgrR are found
in E. coli. Sensory inputs vary depending on the pathogen. Here we show some common histidine
kinase stimulants (e.g., Mg2+, Fe3+) among the Gram-negative pathogens. For a complete list, refer
to the text and references therein. Phosphoethanolamine (pEtN), galactosamine, and 4-amino-
4-deoxy-L-arabinose (L-Ara-4N) molecules are all added to Lipid A of the LPS to modify the
outer membrane charge. A. baumannii lacks a L-Ara-4N-modifying system and instead has galac-
tosamine and pEtN enzymes. The mexXY/oprD multidrug resistance efflux system is found in P.
aeruginosa
Since resistance costs valuable energy, signal transduction networks have been
adapted by bacteria to control the antibiotic response on an as-needed basis.
15.2.3.1 Enterobacteriaceae
Salmonella, the pathogen species in which they were originally discovered [56–58].
Although much is still not understood about the complexities of how polymyxin
resistance arises, resistance appears to be induced nonspecifically by the presence
of cationic compounds through the action of TCS networks. Inducing agents include
polymyxins, low magnesium (Mg2+) concentrations, high ferric iron (Fe3+) concen-
trations, and acidic pH [54, 56]. In some cases metals such as aluminum and zinc
induce the polymyxin signaling cascade. Specifically, magnesium or cationic drugs
stimulate phosphorylation of the histidine kinase PhoQ by direct interaction with an
acidic patch within its sensory domain. PhoQ then dimerizes and phosphorylates
the intracellular response regulator PhoP, which then modulates transcriptional
activity [59]. In K. pneumoniae, PhoP directly activates the arn operon to initiate a
chemical modification that changes the net charge of LPS. In E. coli and Salmonella
species, signaling is accomplished through an intermediary relay protein PmrD,
which signals between the PhoPQ and a second TCS cognate pair, PmrAB [58]. The
PhoPQ/PmrAB systems have common activators and are also able to respond to
their own individualized signals. Both signaling systems respond to cationic pep-
tides (i.e., polymyxins) and low pH; however, they differ in that the PhoQ histidine
kinase responds to low magnesium, while the PmrB histidine kinase responds to
high ferric iron (Fig. 15.5). Interestingly, in E. coli PmrB responds to other metals,
such as zinc and aluminum [54].
Downstream signaling can differ among Enterobacteriaceae species. However,
it is important to note that independent of species, the end result of PhoPQ/PmrAB
induction is the same phenotypic change: the enzymatic restructuring of the LPS as
a charge switch through upregulation of 4-amino-4-deoxy-L-arabinose (L-Ara-4N),
phosphoethanolamine (pEtN), or galactosamine additions. Nevertheless, there are
slight variations among Enterobacteriaceae that are exploited to achieve this com-
mon goal. For example, in K. pneumoniae, aside from activating LPS-modifying
operons, PhoP also activates synthesis of a membrane protein called MgrB, which
inhibits the PhoP kinase to complete a negative feedback regulatory loop (Fig. 15.5).
To further complicate matters, in K. pneumoniae a third TCS, CrrAB, can respond
to polymyxins by activating the transcription of an intermediary signaling protein,
CrrC, which then activates the PmrB response regulator to induce the PmrC LPS-
modifying enzyme (Fig. 15.5). E. coli PhoPQ/PmrAB signaling differs from K.
pneumoniae regulation by adding another layer of regulatory complexity in which
two small regulatory RNAs, mgrR and micA, are involved in inhibition of the phos-
phoethanolamine LPS-modifying operon and PhoP, respectively (Fig. 15.5).
The genomes of many polymyxin-resistant clinical isolates have been sequenced,
thereby verifying that mutations within the Enterobacteriaceae PhoPQ/PmrAB
TCS relays and associated signaling systems are sufficient to result in polymyxin
resistance. In particular, MgrB has been the subject of several reports, as a mutation
within this gene is sufficient to result in colistin resistance by strongly activating the
PhoPQ signaling system [60]. These single mutations within the K. pneumoniae
mgrB gene have been identified in clinical isolates from globally sampled patients,
indicating that this form of resistance is common and possibly arises from indepen-
dent mutagenic events. In fact, one study reported that over 40% of colistin-resistant
isolates, collected from several countries worldwide, had an mgrB mutation [60]. In
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 477
addition to MgrB, independent mutations within the PmrB kinase also produce
resistance to colistin. This observation was noted in a study of globally collected
and sequenced K. pneumoniae colistin-resistant genomes in which a single con-
served amino acid change in the PmrB regulator (threonine-157 to proline) resulted
in (up to) a 170-fold induction in the PmrC LPS-modifying enzyme [61].
bacteremia (2006), it has been reserved for the most serious infections caused by
methicillin-resistant S. aureus (MRSA) and vancomycin-resistant E. faecium (VRE)
[63]. The mechanism by which DAP exerts its bactericidal activity is not fully
understood; however, it is thought to bind and insert itself into the Gram-positive
bacterial membrane via a process that is enhanced by the presence of calcium and is
dependent upon interaction with the anionic membrane phospholipid phosphatidyl-
glycerol (PG) [64–67]. How this results in cell death is still unclear, but it may
involve the formation of oligomeric DAP pore-like structures that results in ion
leakage and disruption of membrane potential [66]. Another possibility is the
recently proposed lipid extracting effect in which the accumulation of DAP in the
cell membrane leads to the aggregation and subsequent release of lipid from the cell
membrane [67], both of which would affect permeabilization, metabolism, and cell
division.
Due to its distinct mechanism of action, resistance to DAP is rare. Nevertheless,
treatment failure is a serious concern [68, 69]. Some of the mechanisms that under-
lie DAP resistance (DAP-R) involve enzymes participating in phospholipid metab-
olism and membrane homeostasis (Fig. 15.2) [70]. Regarding signal transduction
genes, the previously mentioned VraSR-regulated cell wall synthesis stimulon and
the essential TCS, WalKR, have been shown to contribute to the DAP-R phenotype
in clinical and laboratory isolates [71, 72] (see section above regarding Vancomycin
resistance/Staphylococcus aureus). VraSR is involved in the regulation of cell wall
biosynthesis via transcription of a number of genes including pbp2 (penicillin-bind-
ing protein 2) [73] and is upregulated in the presence of DAP. Differential gene
expression analysis between DAP-R and DAP-susceptible (DAP-S) clinical iso-
lates reveals an upregulation of the VraSR TCS system. In addition, deletion of
VraSR from a DAP-R isolate conferred a DAP-S phenotype, pointing to the impor-
tance of these signaling systems in producing daptomycin AMR. In S. aureus, the
WalKR TCS system is involved in the control of peptidoglycan biosynthesis and
can influence peptidoglycan turnover, cross-linking, and chain length by sensing
membrane fluidity, likely through lipid II (Fig. 15.2) [74, 75]. Due to the similari-
ties between a DAP-R and a WalKR-deficient phenotype that includes thickened
cell walls, increased membrane fluidity, and resistance to membrane disruption, it
has been proposed that mutations in WalKR may lead to the downregulation of cell
wall homeostasis, which leads to an increase in bacterial survival in the presence of
DAP [70].
The success of S. aureus as a pathogen can be attributed largely to its ability to
produce a wide range of virulence factors and accessory genes, many of which are
under the control of the Agr quorum-sensing system, a classical TCS module (see
also Chap. 14). Importantly, dysfunction of this system has recently been implicated
in a transient defense mechanism that protects against daptomycin activity [76].
Quorum sensing is a form of intercellular communication that enables bacteria to
initiate density-dependent changes in gene expression, allowing populations of bac-
teria to restrict the expression of genes whose resulting phenotypes are most benefi-
cial at high cell densities. Examples are biofilm production, bioluminescence, and
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 479
virulence factor secretion (for review see Ref. [77]). Quorum sensing involves the
production and secretion of small molecules that are sensed by neighboring cells.
Once the concentration of these small molecules, termed “autoinducers,” exceeds a
certain threshold, population-wide responses are initiated. Broadly speaking, S.
aureus uses the Agr system to coordinate the upregulation of exotoxin expression
and downregulation of surface proteins, such as adhesion molecules, at high cell
densities [78].
The Agr quorum-sensing system is encoded by a four-gene operon (agrBDCA)
and a regulatory RNA, RNAIII, which are expressed from two divergent promoters,
P2 and P3, respectively (Fig. 15.6). AgrA and AgrC comprise a classical TCS sys-
tem. AgrD is the activating ligand (or autoinducer) which is N- and C-terminally
processed in the bacterial cell and then secreted via AgrB into the extracellular
milieu to produce the final autoinducing peptide (AIP) (Fig. 15.6). AIP is sensed by
the transmembrane histidine kinase AgrC, which induces phosphorylation of the
cytoplasmic HPK domain upon AIP binding. This phosphate is transferred to AgrA,
the response regulator TCS transcription factor, thereby activating AgrA and tran-
scription of the two agr promoters, P2 and P3. The Agr system is an example of an
autoactivating system; agr autoactivation leads to an exponential increase in expres-
sion of the two agr promoters. At high cell densities, RNAIII, whose transcription
is under the control of the P3 promoter, is a regulatory molecule that is responsible
for the differential expression of many genes. These large changes in gene expres-
sion inflict a significant metabolic burden upon the cell and have been hypothesized
to partly explain the selective enrichment of Agr-defective mutants among seriously
ill, hospitalized patients [78] (see also Chap. 14).
The development of Agr-defective mutants during invasive infection and the
hypothesis that these mutations may incur a survival advantage in the presence of
antibiotics led the Edwards lab to investigate the role of the Agr system in daptomy-
cin susceptibility [76]. Somewhat counterintuitively, they discovered that the loss
of AgrA or AgrC, the TCS module, allowed for survival of S. aureus in the presence
of daptomycin. Investigation into the mechanism by which this occurred revealed
that Agr-defective mutants can survive antibiotic exposure by actively releasing
membrane phospholipids that bind to and inactivate daptomycin. This process also
occurs in wild-type bacteria; however, a set of Agr-regulated genes that are
expressed in wild-type bacteria mitigate this observed inactivating effect. The
genes encode molecules called phenol-soluble modulins (PSMs), small cytolytic
toxins that promote binding of daptomycin to the bacteria by sequestering the shed
membrane phospholipid, likely via their surfactant (or detergent-like) properties
[76]. Pader et al. also found that the enhanced survival of Agr-defective mutants in
the presence of daptomycin could be mitigated by the addition of the β-lactam
antibiotic oxacillin, which reduced the rate of lipid release from the bacterial mem-
brane and therefore the inactivation of daptomycin. This result suggests addition of
oxacillin as an immediate clinical remedy for daptomycin-tolerant S. aureus. This
mechanism has been extended to include the enterococci and streptococci in a
recent report [79].
480 A. T. Ulijasz et al.
DAP DAP
DAP
Extracellular
DAP
DAP
DAP
PSMa
DAP DAP
PSMa PSMa
AIP C
S PSMa
AgrC
Membrane
AgrB
D AgrA
REC
Cytoplasm
DBD
Virulence factors
ArgD
Chromosome
RNAIII
P2 P3
argA argC argD argB RNAIII PSMα1-4
Fig. 15.6 The Agr signaling pathway and daptomycin resistance. AgrC and AgrA comprise a TCS
pair that senses the AIP quorum-sensing autoinducer peptide (AIP) signal. AIP is made from ArgD
peptide, which is processed and secreted by ArgB. The ArgCA TCS system also regulates (alpha
1–4 type) phenol-soluble modulin (PSM) production, which are excreted and bind daptomycin
(DAP) to inactivate it. agrBDCA is transcribed from the P2 promoter, and the virulence regulatory
RNAIII is transcribed from the P3 promoter. Both are controlled by AgrA
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 481
The drug fosfomycin, discovered in the late 1960s, has lately reemerged as a sec-
ond-line treatment for urinary tract infections due to the increasing prevalence of
resistance to commonly prescribed antibiotics, such as trimethoprim-sulfamethoxa-
zole and ciprofloxacin [80]. Fosfomycin is primarily taken up by the glycerol-
3-phosphate transport system (GlpT) but may also be internalized by the hexose
phosphate uptake transport system (UhpT) when glucose-6-phosphate is present.
Once inside the bacterium, fosfomycin inactivates cytosolic N-acetylglucosamine
enolpyruvyl transferase (MurA). This impairs bacterial cell wall formation by
inhibiting the first step in peptidoglycan synthesis, the formation of N-acetylmuramic
acid from N-acetylglucosamine and phosphoenolpyruvate [81–83]. Since its clini-
cal launch, several mechanisms for fosfomycin resistance have emerged, including
inducible modulation of the GlpT and UhpT transporters by the Cpx TCS system
[84]. Although this section will focus on the Cpx system as it is understood in E. coli
and as it relates to fosfomycin, the Cpx system has also been implicated in resis-
tance to other antibiotics in E. coli, such as aminoglycosides and beta-lactams, as
well as forms of resistance in Salmonella and P. aeruginosa [85–87].
The Cpx system is comprised by the genes cpxA and cpxR, which encode the
inner membrane sensor histidine kinase CpxA and the response regulator transcrip-
tion factor CpxR [88]. CpxA responds to membrane stressors such as unfolded or
misfolded proteins, as well as to changes in pH. In the absence of stressors, CpxA
functions as a phosphatase, keeping CpxR in its unphosphorylated, inactive form.
The activation of CpxA is modulated by the periplasmic protein CpxP, which inter-
acts with the sensing domain of CpxA; CpxP is displaced by misfolded proteins [89,
90]. Free CpxA is activated by additional signals that are currently not well under-
stood. In typical TCS fashion, once activated, CpxA autophosphorylates and then
transfers the phosphate to CpxR [89, 91]. P-CpxR then acts as a transcription factor
for a variety of genes, including the cpx regulon, degP, glpT, and uhpT (Fig. 15.7)
[84, 88–91].
Interestingly, the Cpx response is shut off by feedback inhibition modulated by
the amount of unfolded protein present in the bacterial cell. P-CpxR increases the
transcription levels of cpxP, and CpxP subsequently inhibits the CpxA kinase, while
also carrying out its function in binding unfolded proteins. Increases in P-CpxR also
increase transcription and translation of the periplasmic protease DegP [88], whose
function is to relieve membrane stress by degrading misfolded or unfolded proteins
in the periplasmic space. DegP recognizes unfolded proteins bound to CpxP, and the
CpxP complex is then degraded by DegP. Importantly, the digestion of CpxP does
not occur in the absence of unfolded proteins. This careful balance suggests a mech-
anism whereby unfolded proteins displace CpxP from CpxA, allowing for auto-
phosphorylation and the subsequent phosphorylation of CpxR, which, in turn, leads
to increased levels of both CpxP and DegP through transcriptional activation
(Fig. 15.7). While unfolded protein remains in the periplasmic space, CpxP will
bind to it and target it for degradation by DegP. As unfolded protein levels fall, more
482 A. T. Ulijasz et al.
Cell wall
DegP
CpxP
Periplasm
(active)
CpxP
(free)
CpxP misfolded
protein G3P F F G6P
quinone
CpxA ArcB pool GlpT UhpT
Membrane
Inner
H H
F F G6P
Cytoplasm
G3P
G3P G6P
G3P
G3P G6P
REC REC
DBD DBD
F MurA
uhpT
Fig. 15.7 Fosfomycin resistance and the Cpx regulatory system. Fosfomycin (F) acts on the early
stages of cell wall synthesis by binding MurA and inhibiting its function. The drug enters the cell
via GlpT and UhpT transporters that normally transport glycerol-3-phosphate (G3P) and glycerol-
6-phosphate (G6P) sugars, respectively. The CpxAR TCS system represses glpT and uhpT tran-
scription and activates degP transcription. Free CpxP inhibits CpxA activity. CpxP functions to
bind unfolded protein and deliver it to the DegP protease in the periplasm. The TCS system ArcAB
senses cellular quinone pools and, in turn, regulates aerobic metabolism and resultant reactive
oxygen species (ROS) that is proposed to induce fosfomycin antibiotic lethality
CpxP is free to bind CxpA; this change in the CpxP/CpxA stoichiometry shifts
activity from kinase to phosphatase function in CpxA, increasing the levels of CpxR
relative to P-CpxR (i.e., phosphorylated and active CpxR (Fig. 15.7)). The altering
of the CpxR phosphorylation equilibrium ultimately results in a decrease in the
transcription of membrane stress-response genes [92].
The Cpx system promotes fosfomycin resistance through P-CpxR by negatively
regulating the transcription of glpT and uhpT. This limits fosfomycin entry into the
cell via the GlpT and UhpT transporters (current data suggests that the GlpT trans-
porter is the more physiologically relevant [84]). It is worth noting that the role of
the Cpx system in mediating a response to antibiotics is still under debate. Some
work suggests that the Cpx system activation is involved in antibiotic-mediated
accumulation of reactive oxygen species (ROS), which then can lead to cell death
[93]. This conclusion was drawn by observing that a null cpxA strain of E. coli
reduced the lethality of several antibiotic classes and accumulated less ROS. Other
studies show that mutations turning CpxA into a constitutively active kinase lacking
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 483
bacteria are engulfed. However, CAMPs can also be released into the medium after
phagosomal digestion to kill pathogens that have a more extracellular lifestyle, such
as streptococci, staphylococci, and Pseudomonas species. In the case of epithelial-
type cells, CAMPs can be secreted and act as an intrinsic barrier to pathogens enter-
ing organs or tissues where they are unwelcome. These activities place CAMPs at
the forefront of maintaining the normal gut flora and general microbiome composi-
tion of the host [100].
In order to successfully colonize and infect their host, pathogenic bacteria have
evolved a multitude of CAMP-resistance mechanisms. Among these mechanisms
are (1) repulsion via recharging their cell envelope, (2) sequestration, (3) export via
transporters, and (4) direct enzymatic-driven breakdown of the peptides [101].
Since these collective processes require energy, bacteria must have ways to detect
and respond quickly to host CAMP assaults without compromising precious energy
needed for colonization and pathogenicity. In this section we describe several spe-
cific examples of how bacterial pathogens sense and respond to CAMPs.
15.3.1 Salmonellae
Bacterial CAMP-resistance mechanisms often use the same general signaling and
resistance determinants that they use for bacteria-produced antimicrobials. Probably
the most well-studied CAMP-resistance signaling cascade is PhoPQ TCS pair
[102]. Aside from being an integral signaling component of polymyxin resistance
(see the above section on polymyxins), after engulfment of the bacteria by the host,
PhoPQ is required for sensing the acidic shift in pH and an accompanying drop in
manganese concentrations within the phagosome (see Fig. 15.5). This detection is
crucial to the Salmonella species life cycle, since they are intracellular pathogens
and must therefore escape and/or subvert innate immunity to remain viable and
replicate within host cells, including innate immune cells such as macrophages.
Since phagocytes produce CAMPs to lyse the engulfed Salmonella, once PhoPQ is
activated it initiates expression of several genes that modify the cell envelope charge
and repel CAMPs. To accomplish this, PhoPQ controls a large regulon containing
more than twenty genes, which together change several chemical moieties within
lipopolysaccharides, glycerophospholipids, and outer membrane proteins that alter
the net charge of the cell envelope from negative to positive (Fig. 15.5). As described
in the previous section on polymyxins, these changes are also influenced by the
PhoPQ induction of a second TCS, PmrAB. For a more extensive description of the
chemical basis for PhoPQ-induced CAMP resistance see Ref. [102] and references
therein.
The collective action of PhoPQ/PmrAB induction results in CAMP resistance
and a concomitant decrease in innate immune recognition that ensures the survival
of Salmonella within macrophages and the host environment. One interesting aspect
of the many outer membrane modifications is that PhoPQ/PmrAB may make spe-
cific changes that depend on the particular infected host tissue and environment. For
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 485
15.3.2 Streptococci
Gram-positive bacteria are also subject to CAMPs from phagocytes and other host
cells. Differing from Gram-negative microbes, Gram-positive bacteria lack an outer
membrane and instead possess an expanded cell wall for fortification, often sur-
rounded by a protective polysaccharide capsule (see Fig. 15.2). These layers col-
lectively act as barriers to phagocytosis and unwanted host molecules. Taking these
differences into consideration, it is not surprising that Gram-positive bacteria have
evolved systems of CAMP resistance that may differ considerably from those of
Gram-negatives, in both signaling and mechanistic outcomes. For example, strepto-
cocci (e.g., S. pneumoniae, S. agalactiae, and S. pyogenes) and staphylococci spe-
cies often rely on an L-lysinylation-protective strategy facilitated by MprF (for
multiple peptide resistance factor). MprF conveys its phenotype by adding posi-
tively charged L-lysine groups to the membrane lipid phosphatidylglycerol, thereby
enabling repulsion of host CAMPs [105–108]. Other strategies include the
D-alanylation of cell wall teichoic acid by the dlt operon, a mechanism conserved
among several streptococcal and other Gram-positive species. This chemical modi-
fication results in increased density and surface charge of the Gram-positive cell
wall, which then acts to absorb and repel CAMPs. D-alanylation generated by the
Dlt proteins also confers resistance to host-produced acid, lysozyme, neutrophil-
produced neutrophil extracellular traps (NETs), and antimicrobial peptides known
as bacteriocins [106]. Such broad resistance capabilities suggest that the dlt operon
and other antimicrobial systems act as versatile signaling and host-responsive
mechanisms, rather than specifically targeting a single antibiotic. This reoccurring
theme is emphasized below.
486 A. T. Ulijasz et al.
15.3.3 Staphylococci
The main CAMP sensory system in S. aureus and related species is the GraRS/
VraFG signaling cascade, which modifies the overall charge of the cell envelope
(described above with reference to its contribution to vancomycin resistance [25];
Fig. 15.2). Thus, this signaling cascade is a general, nonspecific means to facilitate
resistance based on electrostatic repulsion of host-produced CAMPs. However, the
resistance conveyed by GraRS/VraFG can differ greatly depending on the staphylo-
coccus species. For example, GraRS/VraFG recognizes and responds to host LL-37
and indolicidin in both S. aureus and S. epidermidis, but only S. epidermidis GraRS/
VraFG responds to human beta-defensin-3, which protects against skin infections.
S. epidermidis, as its name suggests, is part of the normal flora of human skin and
causes opportunistic skin infections. The species-specific recognition is facilitated
by a short extracellular loop within the GraS histidine kinase sensory domain [115].
Induction of the GraRS signaling cascade (histidine kinase phosphorylation of the
response regulator; Fig. 15.1) then initiates expression of the dlt operon, which is
488 A. T. Ulijasz et al.
responsible for enzymatic incorporation of D-alanine into the cell wall teichoic
acid. Simultaneously, GraRS upregulates MprF, which adds a lysine moiety onto
phosphatidylglycerol (Fig. 15.2). These modifications collectively alter the overall
cell wall and membrane charge of the bacterial envelope, thereby electrostatically
repelling host CAMPs. In addition, GraRS activates the VraFG TCS system, which
has been proposed to govern the efflux of CAMPs from the bacterial cell [107]. The
requirement of the GraRS and VraFG signaling systems for CAMP resistance there-
fore explains experimental evidence showing that these TCS pairs are required to
survive neutrophil attacks and are critical to the success of staphylococcal infections
[107, 116].
In addition to the all-important and very ubiquitous GraRS/VraFG systems, the
staphylococci also use other signaling cascades to counteract CAMP activity. The
global regulators Agr (see section above on daptomycin resistance) and SarA, as
well as the TCS system SaeRS, have been implicated in CAMP resistance through
activation of the controlled expression and release of CAMP-degrading proteases,
such as the exoprotease SepA (Ref. [116]; for review see Ref. [107]). Finally, a
report from 2013 describes another TCS pair, LytSR, which confers staphylococcal
CAMP resistance. Interestingly, the authors proposed that LytSR senses subtle
changes in membrane potential, alerting the bacterial cell of early exposure to host
CAMPs that perturb the normal electrical gradient [117]. An appropriate response
can then be elicited in time. These studies collectively suggest that S. aureus and
other staphylococci have evolved complex regulatory networks to survive host
onslaughts from innate immune cells, and specifically from CAMPs.
films rapidly, as is the case for several other antimicrobial classes such as the
polyketide tetracycline (see Ref. [120] and references therein).
One commonality to all biofilms is the upregulation of genes producing a protec-
tive glycocalyx or capsular sugary matrix once the bacteria are attached to a host
surface. This physiological change upon surface attachment is often accompanied
by the initiation of a number of AMR resistance mechanisms, including enzymati-
cally driven antibiotic degradation (e.g., by beta-lactamases) and the production of
multidrug resistance efflux pumps. Indeed, both are hypothesized to contribute sig-
nificantly to the staying power of the bacterial biofilm [119, 120].
Biofilms also contain a large DNA component, aptly named “extracellular DNA”
or eDNA, which was found to act as a protectant against innate immunity and some
antibiotics by adsorption [121]. A large fraction of the bacteria within the biofilm
community are also metabolically less active, especially within the deeper layers of
the substratum. This phenotype has now been associated with bacterial stress
responses that include the ppGpp-driven stringent response, or in other cases the
SOS response. Both are involved in controlling bacterial cell death and longevity.
This slowing of metabolism is a hallmark of the so-called “persister” cell pheno-
type, or the subset of cells that “persist” in the wake of antimicrobial or other envi-
ronmental stress [17, 120]. Data now show that persister cells comprise a
considerable proportion of the biofilm and thus might contribute significantly to the
longstanding question of why biofilms render antibiotics less effective (for an
expanded description on persisting microbes see the “Persisters” section below).
The transition from a planktonic state to initial adherence to gradual, stepwise
increments toward a mature and resistant biofilm is a highly complex process that
requires bacterial cells to undergo dramatic physiological, metabolic, and pheno-
typic changes [118]. This process is reversibly controlled through a variety of sig-
naling mechanisms that include quorum-sensing systems (i.e., cell-cell
communication) and signal transduction cascades. In many cases, biofilms are
induced by subinhibitory concentrations of the antibiotic itself, strongly suggesting
that signal transduction systems are involved at sensing and responding to the threat
and play an integral role in tolerance and resistance [122]. Here we will discuss
specific examples from both Gram-positive and Gram-negative bacteria concerning
how pathogens undergo biofilm formation and, importantly, the relevant signaling
proteins involved in biofilm-mediated resistance and tolerance to antibiotics.
many P. aeruginosa clinical isolates from cystic fibrosis patients’ sputum contained
mutations within BrlR, suggesting that in chronic infections BrlR might be inacti-
vated to enhance tolerance to host-produced CAMPs [126]. Indeed, recent clinically
relevant cystic fibrosis models developed in swine have shown that P. aeruginosa
and other bacteria are better able to colonize the diseased cystic fibrosis lung due to
the drastic change in lung pH. It was hypothesized that the pH change alters the
charge of host-produced CAMPs, making the lung far less effective at clearing
microbes from the usually more or less “sterile lung” environment [127]. On the
other hand, when present in its wild-type form, BrlR is responsible for activation of
biofilms and, as a result, tolerance to the aminoglycoside tobramycin. Collectively,
these studies suggest that tobramycin is more effective with isolates that have lost
BrlR function in the more chronic, later stages of the cystic fibrosis disease and,
conversely, might be less effective if given early on.
Almost a decade before the discovery of BrlR the association of a signaling path-
way connecting aminoglycoside resistance and biofilm formation was discovered.
Miller and colleagues described the aminoglycoside response regulator Arr as being
responsible for induction of biofilm formation at subinhibitory concentrations of
aminoglycosides, such as tobramycin, in both P. aeruginosa and E. coli [128]. Arr
contains a c-di-GMP phosphodiesterase (or EAL) domain that directly regulates
biofilm formation and tobramycin resistance. This regulation is accomplished by
responding to subinhibitory levels of antibiotic via an as-of-yet unknown mecha-
nism and, in turn, regulating c-di-GMP levels through breakdown by the Arr EAL
domain [128].
Other regulators that affect the levels of c-di-GMP have also been implicated in
antibiotic resistance, such as the PvrR response regulator, originally described as
controlling P. aeruginosa phenotypic variance and antibiotic resistance. With a pvrR
mutant, the aminoglycoside kanamycin induced small-colony variants, which are
hyper-adhesive and better resist antimicrobials [129]. Similar to Arr, PvrR contains
an EAL domain and is therefore involved in responding to and controlling cellular
c-di-GMP levels. Results from these studies collectively point to bacterial cell-cell
communication systems responding to often subinhibitory concentrations of antibi-
otics and then changing the cell phenotype to enable resistance/tolerance. In this
way the bacteria can communicate the threat to their community and respond in a
timely fashion. The concept of subinhibitory levels of antibiotic resulting in biofilm
formation has now been demonstrated with many diverse pathogens, including
Gram-positive bacteria such as S. aureus. The general response to antimicrobial
treatment and induction of the protective biofilm state is a result of the ability of
signal transduction systems to detect and quickly eliminate the threat. For a compre-
hensive review on antimicrobial induction of biofilms see Ref. [122].
As mentioned in the introduction of this section, biofilms also contain an exten-
sive amount of eDNA. These web-like structures are either secreted by dedicated
export systems in the bacteria or, more indirectly, can be a consequence of cell lysis
[121]. Once outside the cell, eDNA can act as neutrophil extracellular traps (or
NETs) to enable immune system evasion. Interestingly, within the context of the
Pseudomonas biofilm, eDNA was recently shown to play a role in signaling CAMP
492 A. T. Ulijasz et al.
and other AMR mechanisms [121]. Due to the highly anionic nature of eDNA, it
was shown to act as an absorber of cationic metal ions, such as magnesium. As a
consequence, eDNA depletes extracellular magnesium and activates the PhoPQ/
PmrAB TCS, which chemically modifies the P. aeruginosa polysaccharide to repel
host CAMPs (see Fig. 15.5). The same effect holds true with other Gram-negative
bacteria, such as Salmonella species [121]. A surprising discovery was that addition
of eDNA to planktonic cultures of P. aeruginosa cells induced the expression of a
three-gene cluster adjacent to PmrAB that is responsible for the production and
export to the outer membrane of the cationic molecule spermidine. It was hypothe-
sized that spermidine then acts as a positively charged surrogate to magnesium,
repelling CAMPs and positively charged antibiotics, including polymyxins and
aminoglycosides (e.g., tobramycin and gentamicin) [121].
15.4.2 Staphylococci
that direct biofilm formation in S. aureus, namely, the agr quorum-sensing signaling
cascade ([107]; see the section on daptomycin resistance above). Maturation of the
protective S. aureus biofilm structure is facilitated by phenol-soluble modulins
(PSMs), which has been confirmed to be a signaling requirement for catheter-related
biofilm infections in animal infection models ([132, 134]). Despite a plethora of
examples of antibiotic-induced biofilm formation in staphylococcal species, most of
the regulatory connections remain tenuous at best (for review see Ref. [122]).
The direct association between induction of staphylococcal biofilm formation
and antibiotic treatment, especially at subinhibitory concentrations, was first
observed in 1940 by Arthur Gardner with the Gram-positive Clostridium perfrin-
gens (for review Ref. [122]). This effect was then established in several other Gram-
positive and Gram-negative pathogens with many additional antibiotics. However,
the signaling responsible for induction remains unknown for most Gram-positive
cases, with a few recent exceptions with staphylococci. One of these exceptions
comes from a report from Michael Otto and colleagues, which demonstrates that
in vivo PSMs are key contributors to the S. aureus biofilm maturation process [132].
The authors found that PSM placement within the matrix dictates the local struc-
tures of the biofilm. Otto and colleagues propose a model by which this local varia-
tion is controlled by targeted activity of the agr quorum-sensing system (described
in the daptomycin resistance section and displayed in Fig. 15.6) [132]. Agr and
phenol-soluble modulins (PSMs) also control biofilm detachment. Another report
from Schilcher et al. builds upon previous knowledge that at subinhibitory concen-
trations clindamycin induces higher eDNA content in the S. aureus biofilm matrix
[135]. This effect was then determined to be triggered by the alternative sigma fac-
tor B (σB) and its upregulation of known biofilm-associated factors. This report is
important because it provided the missing link in staphylococcal signaling between
inducible biofilm formation and biofilm-driven resistance. The extent to which σB is
integrated within other Gram-positive antibiotic-induced biofilm resistance signal-
ing networks remains to be determined.
15.5 Persisters
Bacterial toxin-antitoxin (TA) systems were the first signaling systems discovered
to induce persistence through environmental ques. TA systems are ubiquitous in
bacteria, with many species possessing multiple versions. TA systems are usually
encoded within an operon that consists of two genes, one encoding a toxin that regu-
lates cell growth in some manner and the other encoding a cognate antitoxin that
regulates the levels of the toxin (for review see Ref. [139]). In the most simplistic
system, the antitoxin, which is normally a DNA-binding domain-containing tran-
scription factor, directly binds to the toxin. In doing so the toxin then acts as a core-
pressor with the antitoxin to bind DNA and self-regulate repression of the TA
operon [138]. Because the antitoxin is usually produced in excess of toxin, the toxin
then controls the TA relationship through the toxin-to-antitoxin ratio. In this manner
the TA ratio can fine-tune and control the cellular growth and therefore the persis-
tence phenotype. It is interesting to note that the “toxin” in most cases is not actually
a toxin in the classical sense, but rather a regulatory protein or RNA whose action
indirectly results in cellular toxicity (e.g., through posttranslationally modifying
other proteins that actually confer the toxicity). To date there is only one example in
which the toxin component of the TA pair directly influences cellular toxicity (Ref.
[140] and see below). To add another layer of complexity, it is known that with
many TA systems a protease (Lon protease) degrades the antitoxins, which is, in
turn, controlled by cellular phosphate levels [136, 138, 139, 141]. One hypothesis
for how TA systems are able to direct reversible phenotypic heterogeneity involves
the balance the TA module provides between cell growth and arrest. In doing so, it
enables a responsive subpopulation to occur only when required. The mechanism
for how this balance is controlled is still the subject of much debate [141].
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 495
Most studies investigating the mechanisms of persister formation have been accom-
plished using laboratory strains of E. coli in test tubes (i.e., in vitro). Although these
studies have yielded a plethora of information [136], it is also important, perhaps
even more so, to study persister formation in actual pathogens and in the context of
their native host environment. Using S. Typhimurium single-cell analysis with a
murine model of infection, Helaine and colleagues have done just this with surpris-
ing results. They found that within 30 minutes after macrophages engulf S.
Typhimurium in an animal, as much as 20% of the population changes to the per-
sister state, becoming tolerant to antibiotic treatments [144]. This surprising result
was in stark contrast to all previous reports from in vitro experiments, which showed
that the fraction of persisters is normally no greater than 1% of the total population
[138]. The persistence was shown to be aided by Lon protease/ppGpp-dependent TA
modules and, interestingly, triggered by the drastic change in pH and nutrient depra-
vation when the bacteria enter the macrophage vacuole [144]. A second paper by the
same group describes the mechanism by which this occurs through a new class of
uncharacterized TA modules harboring Gcn5 N-acetyltransferase (i.e., acetylation)
activity [145]. It was determined that the acetyl transferase, dubbed TacT for “tRNA
acetylating toxin,” signals for initiation of the persister state by acetylating tRNA,
thereby inhibiting tRNA function and simultaneously inducing lon-ppGpp-medi-
ated cell growth arrest [145]. Furthermore, this acetylation and therefore growth
could be reversed by the S. Typhimurium deacetylase CobB [145]. These studies are
important, as they show a more comprehensive picture of how TA system signaling
can facilitate reversible persister formation that is independent of genotype.
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 497
In this chapter we discussed how bacterial cells sense antimicrobials and respond in
a timely manner to tolerate/resist them using signal transduction systems. Although
many systems are specific to a particular species or genus, there are also examples
of conservation that might elicit the targeting of these pathways for new broad-
spectrum antimicrobials. One of these research areas that has recently gained atten-
tion involves drugs that inhibit quorum-sensing pathways, which are shared by both
Gram-positive and Gram-negative organisms and generally control biofilm forma-
tion in many pathogens [146]. In this regard, the most progress has been made with
P. aeruginosa and related pathogens [147]. Other attempts to subvert antibiotic
resistance by inhibition of signaling systems have been made in the area of two-
component signaling. Although two-component signaling systems have been
deemed too diverse and prone to mutagenesis for the design of serious broad-spec-
trum inhibitors, recent efforts to target the conserved WalRK system have resusci-
tated this area of research [148]. An important consideration favoring these
approaches is that both quorum-sensing and two-component signaling are com-
pletely absent from the mammalian genome. Promising data have also been gener-
ated with a recent screening effort for antibiotics that specifically inhibit persister
formation [149]. Thus, this work could have broad implications for preventing bio-
films and ensuing antibiotic resistance/tolerance. This idea seems rational, as persis-
tence signaling seems to eventually culminate in a finite number of stress-response
pathways that are common to many bacterial pathogens. Recent studies have shown
that some antibiotics derived from previously “unculturable” bacteria demonstrate
great potential to provide a new class of persister-targeting antibiotics [149, 150].
Perhaps the future holds a more multifaceted approach to the problem of resistance
and tolerance, such that one antibiotic is given to clear an infection and another to
clear the anticipated tolerant subpopulation.
Major Points
• Tolerance and resistance mechanisms can be energetically costly; thus bacteria
must have a means to sense a threat, induce a response, and terminate a response.
• Bacterial signal transduction systems provide a means to sense and respond to
both their extracellular and intracellular environments.
• Bacterial signaling systems are largely comprised of two-component signaling
(TCS) systems, but they can also be serine-threonine kinase (eSTK) and phos-
phatase (eSTP) systems, along with more specialized signaling such as the S.
aureus Agr cascade.
• A general response to many cell-envelope targeting antimicrobials is the modifi-
cation of the envelope charge to electrostatically repel the drug. Some signaling
systems that detect and respond to antibiotics in this way, such as the PhoQP and
PmrAB TCS systems, are conserved among many bacterial species.
• Bacterial signaling systems, especially cell-cell communication systems (or quo-
rum sensing), are required to regulate biofilm formation and antibiotic tolerance.
498 A. T. Ulijasz et al.
• Within the biofilm, a subpopulation of persister cells exist that are metabolically
less active than their majority counterparts and whose state in many cases is
controlled by toxin-antitoxin (TA) modules.
• Bacterial signaling systems accumulate mutations that contribute to resistance
and tolerance; therefore, they are an important component of antimicrobial
resistance.
References
1. Swartz TE, Tseng TS, Frederickson MA, Paris G, Comerci DJ, Rajashekara G, et al.
Blue-light-activated histidine kinases: two-component sensors in bacteria. Science.
2007;317(5841):1090–3.
2. Kobir A, Shi L, Boskovic A, Grangeasse C, Franjevic D, Mijakovic I. Protein phosphoryla-
tion in bacterial signal transduction. Biochim Biophys Acta. 2011;1810(10):989–94.
3. Hu LI, Lima BP, Wolfe AJ. Bacterial protein acetylation: the dawning of a new age. Mol
Microbiol. 2010;77(1):15–21.
4. Antelmann H, Helmann JD. Thiol-based redox switches and gene regulation. Antioxid Redox
Signal. 2011;14(6):1049–63.
5. Hoch JA, Silhavy TJ, editors. Two-component signal transduction. Washington, D.C.:
American Society for Microbiology; 1995.
6. Wright DP, Ulijasz AT. Regulation of transcription by eukaryotic-like serine-threonine
kinases and phosphatases in gram-positive bacterial pathogens. Virulence. 2014;5:863–85.
7. Burnside K, Rajagopal L. Regulation of prokaryotic gene expression by eukaryotic-like
enzymes. Curr Opin Microbiol. 2012;15(2):125–31.
8. Howden BP, Davies JK, Johnson PD, Stinear TP, Grayson ML. Reduced vancomycin sus-
ceptibility in Staphylococcus aureus, including vancomycin-intermediate and heterogeneous
vancomycin-intermediate strains: resistance mechanisms, laboratory detection, and clinical
implications. Clin Microbiol Rev. 2010;23(1):99–139.
9. D'Costa VM, King CE, Kalan L, Morar M, Sung WW, Schwarz C, et al. Antibiotic resistance
is ancient. Nature. 2011;477(7365):457–61.
10. CDC. Antibiotic resistance threats in the United States. US Department of Health and Human
Services; 2013.
11. Ulrich LE, Koonin EV, Zhulin IB. One-component systems dominate signal transduction in
prokaryotes. Trends Microbiol. 2005;13(2):52–6.
12. Maule AF, Wright DP, Weiner JJ, Han L, Peterson FC, Volkman BF, et al. The aspar-
tate-less receiver (ALR) domains: distribution, structure and function. PLoS Pathog.
2015;11(4):e1004795.
13. Grangeasse C, Cozzone AJ, Deutscher J, Mijakovic I. Tyrosine phosphorylation: an emerging
regulatory device of bacterial physiology. Trends Biochem Sci. 2007;32(2):86–94.
14. Grangeasse C, Nessler S, Mijakovic I. Bacterial tyrosine kinases: evolution, biological func-
tion and structural insights. Philos Trans R Soc Lond Ser B Biol Sci. 2012;367(1602):2640–55.
15. Pereira SF, Goss L, Dworkin J. Eukaryote-like serine/threonine kinases and phosphatases in
bacteria. Microbiol Mol Biol Rev. 2011;75(1):192–212.
16. Libby EA, Goss LA, Dworkin J. The eukaryotic-like Ser/Thr kinase PrkC regulates the essen-
tial WalRK two-component system in Bacillus subtilis. PLoS Genet. 2015;11(6):e1005275.
17. Conlon BP, Rowe SE, Lewis K. Persister cells in biofilm associated infections. Adv Exp Med
Biol. 2015;831:1–9.
18. Levine DP. Vancomycin: a history. Clin Infect Dis. 2006;42(Suppl 1):S5–12.
19. Courvalin P. Vancomycin resistance in gram-positive cocci. Clin Infect Dis. 2006;42(Suppl
1):S25–34.
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 499
20. Holman TR, Wu Z, Wanner BL, Walsh CT. Identification of the DNA-binding site for the
phosphorylated VanR protein required for vancomycin resistance in Enterococcus faecium.
Biochemistry. 1994;33(15):4625–31.
21. Koteva K, Hong HJ, Wang XD, Nazi I, Hughes D, Naldrett MJ, et al. A vancomycin pho-
toprobe identifies the histidine kinase VanSsc as a vancomycin receptor. Nat Chem Biol.
2010;6(5):327–9.
22. Hughes CS, Longo E, Phillips-Jones MK, Hussain R. Characterisation of the selective bind-
ing of antibiotics vancomycin and teicoplanin by the VanS receptor regulating type A vanco-
mycin resistance in the enterococci. Biochim Biophys Acta. 2017;1861(8):1951–9.
23. Hu Q, Peng H, Rao X. Molecular events for promotion of vancomycin resistance in vanco-
mycin intermediate Staphylococcus aureus. Front Microbiol. 2016;7:1601.
24. Cui L, Neoh HM, Shoji M, Hiramatsu K. Contribution of vraSR and graSR point mutations
to vancomycin resistance in vancomycin-intermediate Staphylococcus aureus. Antimicrob
Agents Chemother. 2009;53(3):1231–4.
25. Falord M, Karimova G, Hiron A, Msadek T. GraXSR proteins interact with the VraFG
ABC transporter to form a five-component system required for cationic antimicrobial
peptide sensing and resistance in Staphylococcus aureus. Antimicrob Agents Chemother.
2012;56(2):1047–58.
26. Dubrac S, Bisicchia P, Devine KM, Msadek T. A matter of life and death: cell wall homeo-
stasis and the WalKR (YycGF) essential signal transduction pathway. Mol Microbiol.
2008;70(6):1307–22.
27. Fridman M, Williams GD, Muzamal U, Hunter H, Siu KW, Golemi-Kotra D. Two unique
phosphorylation-driven signaling pathways crosstalk in Staphylococcus aureus to modulate
the cell-wall charge: Stk1/Stp1 meets GraSR. Biochemistry. 2013;52(45):7975–86.
28. Sun F, Ding Y, Ji Q, Liang Z, Deng X, Wong CC, et al. Protein cysteine phosphorylation
of SarA/MgrA family transcriptional regulators mediates bacterial virulence and antibiotic
resistance. Proc Natl Acad Sci U S A. 2012;109(38):15461–6.
29. Beltramini AM, Mukhopadhyay CD, Pancholi V. Modulation of cell wall structure and anti-
microbial susceptibility by a Staphylococcus aureus eukaryote-like serine/threonine kinase
and phosphatase. Infect Immun. 2009;77(4):1406–16.
30. Kadioglu A, Weiser JN, Paton JC, Andrew PW. The role of Streptococcus pneumoniae viru-
lence factors in host respiratory colonization and disease. Nat Rev Microbiol. 2008;6(4):
288–301.
31. Ulijasz AT, Andes DR, Glasner JD, Weisblum B. Regulation of iron transport in Streptococcus
pneumoniae by RitR, an orphan response regulator. J Bacteriol. 2004;186(23):8123–36.
32. Novak R, Charpentier E, Braun JS, Tuomanen E. Signal transduction by a death signal pep-
tide: uncovering the mechanism of bacterial killing by penicillin. Mol Cell. 2000;5(1):49–57.
33. Haas W, Sublett J, Kaushal D, Tuomanen EI. Revising the role of the pneumococcal vex-
vncRS locus in vancomycin tolerance. J Bacteriol. 2004;186(24):8463–71.
34. Robertson GT, Zhao J, Desai BV, Coleman WH, Nicas TI, Gilmour R, et al. Vancomycin
tolerance induced by erythromycin but not by loss of vncRS, vex3, or pep27 function in
Streptococcus pneumoniae. J Bacteriol. 2002;184(24):6987–7000.
35. Hakenbeck R, Grebe T, Zahner D, Stock JB. beta-lactam resistance in Streptococcus pneu-
moniae: penicillin-binding proteins and non-penicillin-binding proteins. Mol Microbiol.
1999;33(4):673–8.
36. Moscoso M, Domenech M, Garcia E. Vancomycin tolerance in clinical and laboratory
Streptococcus pneumoniae isolates depends on reduced enzyme activity of the major LytA
autolysin or cooperation between CiaH histidine kinase and capsular polysaccharide. Mol
Microbiol. 2010;77(4):1052–64.
37. Liu X, Li JW, Feng Z, Luo Y, Veening JW, Zhang JR. Transcriptional repressor PtvR
regulates phenotypic tolerance to vancomycin in Streptococcus pneumoniae. J Bacteriol.
2017;199(14):e00054–17.
38. Fleming A. Classics in infectious diseases: on the antibacterial action of cultures of a peni-
cillium, with special reference to their use in the isolation of B. influenzae by Alexander
500 A. T. Ulijasz et al.
Fleming, Reprinted from the British Journal of Experimental Pathology 10:226-236, 1929.
Rev Infect Dis. 1980;2(1):129–39.
39. Muller M, Marx P, Hakenbeck R, Bruckner R. Effect of new alleles of the histidine kinase
gene ciaH on the activity of the response regulator CiaR in Streptococcus pneumoniae R6.
Microbiology. 2011;157(Pt 11):3104–12.
40. Chewapreecha C, Marttinen P, Croucher NJ, Salter SJ, Harris SR, Mather AE, et al.
Comprehensive identification of single nucleotide polymorphisms associated with beta-lac-
tam resistance within pneumococcal mosaic genes. PLoS Genet. 2014;10(8):e1004547.
41. Peacock SJ, Paterson GK. Mechanisms of methicillin resistance in Staphylococcus aureus.
Annu Rev Biochem. 2015;84:577–601.
42. Lim D, Strynadka NC. Structural basis for the beta lactam resistance of PBP2a from methi-
cillin-resistant Staphylococcus aureus. Nat Struct Biol. 2002;9(11):870–6.
43. Otero LH, Rojas-Altuve A, Llarrull LI, Carrasco-Lopez C, Kumarasiri M, Lastochkin
E, et al. How allosteric control of Staphylococcus aureus penicillin binding protein 2a
enables methicillin resistance and physiological function. Proc Natl Acad Sci U S A.
2013;110(42):16808–13.
44. Balish E, Warner T. Enterococcus faecalis induces inflammatory bowel disease in interleu-
kin-10 knockout mice. Am J Pathol. 2002;160(6):2253–7.
45. Amarnani R, Rapose A. Colon cancer and enterococcus bacteremia co-affection: a dangerous
alliance. J Infect Public Health. 2017;10:681.
46. Kristich CJ, Wells CL, Dunny GM. A eukaryotic-type Ser/Thr kinase in enterococcus fae-
calis mediates antimicrobial resistance and intestinal persistence. Proc Natl Acad Sci U S A.
2007;104(9):3508–13.
47. Squeglia F, Marchetti R, Ruggiero A, Lanzetta R, Marasco D, Dworkin J, et al. Chemical
basis of peptidoglycan discrimination by PrkC, a key kinase involved in bacterial resuscita-
tion from dormancy. J Am Chem Soc. 2011;133(51):20676–9.
48. Hall CL, Tschannen M, Worthey EA, Kristich CJ. IreB, a Ser/Thr kinase substrate, influ-
ences antimicrobial resistance in Enterococcus faecalis. Antimicrob Agents Chemother.
2013;57(12):6179–86.
49. Comenge Y, Quintiliani R Jr, Li L, Dubost L, Brouard JP, Hugonnet JE, et al. The CroRS two-
component regulatory system is required for intrinsic beta-lactam resistance in Enterococcus
faecalis. J Bacteriol. 2003;185(24):7184–92.
50. Kellogg SL, Kristich CJ. Functional dissection of the CroRS two-component system required
for resistance to cell wall stressors in enterococcus faecalis. J Bacteriol. 2016;198(8):1326–36.
51. Jaurin B, Grundstrom T. ampC cephalosporinase of Escherichia coli K-12 has a different
evolutionary origin from that of beta-lactamases of the penicillinase type. Proc Natl Acad Sci
U S A. 1981;78(8):4897–901.
52. Zeng X, Lin J. Beta-lactamase induction and cell wall metabolism in Gram-negative bacteria.
Front Microbiol. 2013;4:128.
53. Zamorano L, Moya B, Juan C, Mulet X, Blazquez J, Oliver A. The Pseudomonas aeruginosa
CreBC two-component system plays a major role in the response to beta-lactams, fitness,
biofilm growth, and global regulation. Antimicrob Agents Chemother. 2014;58(9):5084–95.
54. Jeannot K, Bolard A, Plesiat P. Resistance to polymyxins in Gram-negative organisms. Int
J Antimicrob Agents. 2017;49(5):526–35.
55. Tsubery H, Ofek I, Cohen S, Fridkin M. Structure activity relationship study of polymyxin B
nonapeptide. Adv Exp Med Biol. 2000;479:219–22.
56. Trimble MJ, Mlynarcik P, Kolar M, Hancock RE. Polymyxin: alternative mechanisms of
action and resistance. Cold Spring Harb Perspect Med. 2016;6(10):pii: a025288.
57. Wosten MM, Kox LF, Chamnongpol S, Soncini FC, Groisman EA. A signal transduction
system that responds to extracellular iron. Cell. 2000;103(1):113–25.
58. Kato A, Latifi T, Groisman EA. Closing the loop: the PmrA/PmrB two-component system
negatively controls expression of its posttranscriptional activator PmrD. Proc Natl Acad Sci
U S A. 2003;100(8):4706–11.
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 501
59. Bader MW, Sanowar S, Daley ME, Schneider AR, Cho U, Xu W, et al. Recognition of anti-
microbial peptides by a bacterial sensor kinase. Cell. 2005;122(3):461–72.
60. Olaitan AO, Diene SM, Kempf M, Berrazeg M, Bakour S, Gupta SK, et al. Worldwide emer-
gence of colistin resistance in Klebsiella pneumoniae from healthy humans and patients in Lao
PDR, Thailand, Israel, Nigeria and France owing to inactivation of the PhoP/PhoQ regulator
mgrB: an epidemiological and molecular study. Int J Antimicrob Agents. 2014;44(6):500–7.
61. Jayol A, Poirel L, Brink A, Villegas MV, Yilmaz M, Nordmann P. Resistance to colistin asso-
ciated with a single amino acid change in protein PmrB among Klebsiella pneumoniae iso-
lates of worldwide origin. Antimicrob Agents Chemother. 2014;58(8):4762–6.
62. Gutu AD, Sgambati N, Strasbourger P, Brannon MK, Jacobs MA, Haugen E, et al. Polymyxin
resistance of Pseudomonas aeruginosa phoQ mutants is dependent on additional two-compo-
nent regulatory systems. Antimicrob Agents Chemother. 2013;57(5):2204–15.
63. Steenbergen JN, Alder J, Thorne GM, Tally FP. Daptomycin: a lipopeptide antibiotic for the
treatment of serious Gram-positive infections. J Antimicrob Chemother. 2005;55(3):283–8.
64. Jung D, Rozek A, Okon M, Hancock RE. Structural transitions as determinants of the action
of the calcium-dependent antibiotic daptomycin. Chem Biol. 2004;11(7):949–57.
65. Muraih JK, Pearson A, Silverman J, Palmer M. Oligomerization of daptomycin on mem-
branes. Biochim Biophys Acta. 2011;1808(4):1154–60.
66. Silverman JA, Perlmutter NG, Shapiro HM. Correlation of daptomycin bactericidal activ-
ity and membrane depolarization in Staphylococcus aureus. Antimicrob Agents Chemother.
2003;47(8):2538–44.
67. Chen YF, Sun TL, Sun Y, Huang HW. Interaction of daptomycin with lipid bilayers: a lipid
extracting effect. Biochemistry. 2014;53(33):5384–92.
68. Stefani S, Campanile F, Santagati M, Mezzatesta ML, Cafiso V, Pacini G. Insights and clini-
cal perspectives of daptomycin resistance in Staphylococcus aureus: a review of the available
evidence. Int J Antimicrob Agents. 2015;46(3):278–89.
69. Seaton RA, Menichetti F, Dalekos G, Beiras-Fernandez A, Nacinovich F, Pathan R, et al.
Evaluation of effectiveness and safety of high-dose daptomycin: results from patients
included in the European Cubicin((R)) outcomes registry and experience. Adv Ther.
2015;32(12):1192–205.
70. Miller WR, Bayer AS, Arias CA. Mechanism of action and resistance to daptomycin in
Staphylococcus aureus and Enterococci. Cold Spring Harb Perspect Med. 2016;6(11):pii:
a026997.
71. Muthaiyan A, Silverman JA, Jayaswal RK, Wilkinson BJ. Transcriptional profiling reveals
that daptomycin induces the Staphylococcus aureus cell wall stress stimulon and genes
responsive to membrane depolarization. Antimicrob Agents Chemother. 2008;52(3):980–90.
72. Mehta S, Cuirolo AX, Plata KB, Riosa S, Silverman JA, Rubio A, et al. VraSR two-compo-
nent regulatory system contributes to mprF-mediated decreased susceptibility to daptomycin
in in vivo-selected clinical strains of methicillin-resistant Staphylococcus aureus. Antimicrob
Agents Chemother. 2012;56(1):92–102.
73. Kuroda M, Kuroda H, Oshima T, Takeuchi F, Mori H, Hiramatsu K. Two-component
system VraSR positively modulates the regulation of cell-wall biosynthesis pathway in
Staphylococcus aureus. Mol Microbiol. 2004;49(3):807–21.
74. Dubrac S, Msadek T. Identification of genes controlled by the essential YycG/YycF two-
component system of Staphylococcus aureus. J Bacteriol. 2004;186(4):1175–81.
75. Turck M, Bierbaum G. Purification and activity testing of the full-length YycFGHI proteins
of Staphylococcus aureus. PLoS One. 2012;7(1):e30403.
76. Pader V, Hakim S, Painter KL, Wigneshweraraj S, Clarke TB, Edwards AM. Staphylococcus
aureus inactivates daptomycin by releasing membrane phospholipids. Nat Microbiol.
2016;2:16194.
77. Rutherford ST, Bassler BL. Bacterial quorum sensing: its role in virulence and possibilities
for its control. Cold Spring Harb Perspect Med. 2012;2(11):a012427.
78. Novick RP, Geisinger E. Quorum sensing in staphylococci. Annu Rev Genet. 2008;42:541–64.
502 A. T. Ulijasz et al.
79. Ledger EVK, Pader V, Edwards AM. Enterococcus faecalis and pathogenic streptococci inac-
tivate daptomycin by releasing phospholipids. Microbiology. 2017;163(10):1502–8.
80. Sastry S, Doi Y. Fosfomycin: resurgence of an old companion. J Infect Chemother.
2016;22(5):273–80.
81. Kahan FM, Kahan JS, Cassidy PJ, Kropp H. The mechanism of action of fosfomycin (phos-
phonomycin). Ann N Y Acad Sci. 1974;235(0):364–86.
82. Silhavy TJ, Hartig-Beecken I, Boos W. Periplasmic protein related to the sn-glycerol-3-phos-
phate transport system of Escherichia coli. J Bacteriol. 1976;126(2):951–8.
83. Kadner RJ, Winkler HH. Isolation and characterization of mutations affecting the transport of
hexose phosphates in Escherichia coli. J Bacteriol. 1973;113(2):895–900.
84. Kurabayashi K, Hirakawa Y, Tanimoto K, Tomita H, Hirakawa H. Role of the CpxAR two-
component signal transduction system in control of fosfomycin resistance and carbon sub-
strate uptake. J Bacteriol. 2014;196(2):248–56.
85. Mahoney TF, Silhavy TJ. The Cpx stress response confers resistance to some, but not all,
bactericidal antibiotics. J Bacteriol. 2013;195(9):1869–74.
86. Huang H, Sun Y, Yuan L, Pan Y, Gao Y, Ma C, et al. Regulation of the two-component regu-
lator CpxR on aminoglycosides and beta-lactams resistance in Salmonella enterica serovar
typhimurium. Front Microbiol. 2016;7:604.
87. Tian ZX, Yi XX, Cho A, O'Gara F, Wang YP. CpxR activates MexAB-OprM efflux pump
expression and enhances antibiotic resistance in both laboratory and clinical nalB-type iso-
lates of Pseudomonas aeruginosa. PLoS Pathog. 2016;12(10):e1005932.
88. Danese PN, Snyder WB, Cosma CL, Davis LJ, Silhavy TJ. The Cpx two-component signal
transduction pathway of Escherichia coli regulates transcription of the gene specifying the
stress-inducible periplasmic protease, DegP. Genes Dev. 1995;9(4):387–98.
89. Raivio TL, Popkin DL, Silhavy TJ. The Cpx envelope stress response is controlled by ampli-
fication and feedback inhibition. J Bacteriol. 1999;181(17):5263–72.
90. Tschauner K, Hornschemeyer P, Muller VS, Hunke S. Dynamic interaction between the
CpxA sensor kinase and the periplasmic accessory protein CpxP mediates signal recognition
in E. coli. PLoS One. 2014;9(9):e107383.
91. Raivio TL, Silhavy TJ. Transduction of envelope stress in Escherichia coli by the Cpx two-
component system. J Bacteriol. 1997;179(24):7724–33.
92. Isaac DD, Pinkner JS, Hultgren SJ, Silhavy TJ. The extracytoplasmic adaptor protein CpxP
is degraded with substrate by DegP. Proc Natl Acad Sci U S A. 2005;102(49):17775–9.
93. Kohanski MA, Dwyer DJ, Wierzbowski J, Cottarel G, Collins JJ. Mistranslation of mem-
brane proteins and two-component system activation trigger antibiotic-mediated cell death.
Cell. 2008;135(4):679–90.
94. Zhao X, Hong Y, Drlica K. Moving forward with reactive oxygen species involvement in
antimicrobial lethality. J Antimicrob Chemother. 2015;70(3):639–42.
95. Takahata S, Ida T, Hiraishi T, Sakakibara S, Maebashi K, Terada S, et al. Molecular mecha-
nisms of fosfomycin resistance in clinical isolates of Escherichia coli. Int J Antimicrob
Agents. 2010;35(4):333–7.
96. Nilsson AI, Berg OG, Aspevall O, Kahlmeter G, Andersson DI. Biological costs and
mechanisms of fosfomycin resistance in Escherichia coli. Antimicrob Agents Chemother.
2003;47(9):2850–8.
97. Mookherjee N, Hancock RE. Cationic host defence peptides: innate immune regulatory pep-
tides as a novel approach for treating infections. Cell Mol Life Sci. 2007;64(7–8):922–33.
98. Mansour SC, Pena OM, Hancock RE. Host defense peptides: front-line immunomodulators.
Trends Immunol. 2014;35(9):443–50.
99. Bauer ME, Shafer WM. On the in vivo significance of bacterial resistance to antimicrobial
peptides. Biochim Biophys Acta. 2015;1848(11 Pt B):3101–11.
100. Muniz LR, Knosp C, Yeretssian G. Intestinal antimicrobial peptides during homeostasis,
infection, and disease. Front Immunol. 2012;3:310.
15 Bacterial Signal Transduction Systems in Antimicrobial Resistance 503
101. Richards SM, Strandberg KL, Conroy M, Gunn JS. Cationic antimicrobial peptides serve as
activation signals for the Salmonella typhimurium PhoPQ and PmrAB regulons in vitro and
in vivo. Front Cell Infect Microbiol. 2012;2:102.
102. Dalebroux ZD, Miller SI. Salmonellae PhoPQ regulation of the outer membrane to resist
innate immunity. Curr Opin Microbiol. 2014;17:106–13.
103. Needham BD, Trent MS. Fortifying the barrier: the impact of lipid A remodelling on bacterial
pathogenesis. Nat Rev Microbiol. 2013;11(7):467–81.
104. Ernst RK, Yi EC, Guo L, Lim KB, Burns JL, Hackett M, et al. Specific lipopolysaccharide
found in cystic fibrosis airway Pseudomonas aeruginosa. Science. 1999;286(5444):1561–5.
105. Ernst CM, Staubitz P, Mishra NN, Yang SJ, Hornig G, Kalbacher H, et al. The bacterial
defensin resistance protein MprF consists of separable domains for lipid lysinylation and
antimicrobial peptide repulsion. PLoS Pathog. 2009;5(11):e1000660.
106. LaRock CN, Nizet V. Cationic antimicrobial peptide resistance mechanisms of streptococcal
pathogens. Biochim Biophys Acta. 2015;1848(11 Pt B):3047–54.
107. Joo HS, Otto M. Mechanisms of resistance to antimicrobial peptides in staphylococci.
Biochim Biophys Acta. 2015;1848(11 Pt B):3055–61.
108. Peschel A, Jack RW, Otto M, Collins LV, Staubitz P, Nicholson G, et al. Staphylococcus
aureus resistance to human defensins and evasion of neutrophil killing via the novel viru-
lence factor MprF is based on modification of membrane lipids with l-lysine. J Exp Med.
2001;193(9):1067–76.
109. Graham MR, Smoot LM, Migliaccio CA, Virtaneva K, Sturdevant DE, Porcella SF, et al.
Virulence control in group A Streptococcus by a two-component gene regulatory system:
global expression profiling and in vivo infection modeling. Proc Natl Acad Sci U S A.
2002;99(21):13855–60.
110. Burnside K, Rajagopal L. Aspects of eukaryotic-like signaling in Gram-positive cocci: a
focus on virulence. Future Microbiol. 2011;6(7):747–61.
111. Pensinger DA, Aliota MT, Schaenzer AJ, Boldon KM, Ansari IU, Vincent WJ, et al. Selective
pharmacologic inhibition of a PASTA kinase increases Listeria monocytogenes susceptibility
to beta-lactam antibiotics. Antimicrob Agents Chemother. 2014;58(8):4486–94.
112. Klinzing DC, Ishmael N, Dunning Hotopp JC, Tettelin H, Shields KR, Madoff LC, et al. The
two-component response regulator LiaR regulates cell wall stress responses, pili expression
and virulence in group B Streptococcus. Microbiology. 2013;159(Pt 7):1521–34.
113. Cotter PD, Guinane CM, Hill C. The LisRK signal transduction system determines the sensi-
tivity of Listeria monocytogenes to nisin and cephalosporins. Antimicrob Agents Chemother.
2002;46(9):2784–90.
114. Nizet V, Ohtake T, Lauth X, Trowbridge J, Rudisill J, Dorschner RA, et al. Innate antimicrobial
peptide protects the skin from invasive bacterial infection. Nature. 2001;414(6862):454–7.
115. Cheung AL, Bayer AS, Yeaman MR, Xiong YQ, Waring AJ, Memmi G, et al. Site-specific
mutation of the sensor kinase GraS in Staphylococcus aureus alters the adaptive response to
distinct cationic antimicrobial peptides. Infect Immun. 2014;82(12):5336–45.
116. Cheung GY, Rigby K, Wang R, Queck SY, Braughton KR, Whitney AR, et al.
Staphylococcus epidermidis strategies to avoid killing by human neutrophils. PLoS Pathog.
2010;6(10):e1001133.
117. Yang SJ, Xiong YQ, Yeaman MR, Bayles KW, Abdelhady W, Bayer AS. Role of the LytSR
two-component regulatory system in adaptation to cationic antimicrobial peptides in
Staphylococcus aureus. Antimicrob Agents Chemother. 2013;57(8):3875–82.
118. Valentini M, Filloux A. Biofilms and cyclic di-GMP (c-di-GMP) signaling: lessons from
Pseudomonas aeruginosa and other bacteria. J Biol Chem. 2016;291(24):12547–55.
119. Lebeaux D, Ghigo JM, Beloin C. Biofilm-related infections: bridging the gap between clini-
cal management and fundamental aspects of recalcitrance toward antibiotics. Microbiol Mol
Biol Rev. 2014;78(3):510–43.
120. Mulcahy LR, Isabella VM, Lewis K. Pseudomonas aeruginosa biofilms in disease. Microb
Ecol. 2014;68(1):1–12.
504 A. T. Ulijasz et al.
144. Helaine S, Cheverton AM, Watson KG, Faure LM, Matthews SA, Holden DW. Internalization
of Salmonella by macrophages induces formation of nonreplicating persisters. Science.
2014;343(6167):204–8.
145. Cheverton AM, Gollan B, Przydacz M, Wong CT, Mylona A, Hare SA, et al. A Salmonella
toxin promotes persister formation through acetylation of tRNA. Mol Cell. 2016;63(1):86–96.
146. Grandclement C, Tannieres M, Morera S, Dessaux Y, Faure D. Quorum quenching: role in
nature and applied developments. FEMS Microbiol Rev. 2016;40(1):86–116.
147. O'Loughlin CT, Miller LC, Siryaporn A, Drescher K, Semmelhack MF, Bassler BL. A quo-
rum-sensing inhibitor blocks Pseudomonas aeruginosa virulence and biofilm formation. Proc
Natl Acad Sci U S A. 2013;110(44):17981–6.
148. Worthington RJ, Blackledge MS, Melander C. Small-molecule inhibition of bacterial
two-component systems to combat antibiotic resistance and virulence. Future Med Chem.
2013;5(11):1265–84.
149. Ling LL, Schneider T, Peoples AJ, Spoering AL, Engels I, Conlon BP, et al. A new antibiotic
kills pathogens without detectable resistance. Nature. 2015;517(7535):455–9.
150. Lewis K. New approaches to antimicrobial discovery. Biochem Pharmacol. 2017;134:87–98.
Chapter 16
Bacterial Type II Topoisomerases
and Target-Mediated Drug Resistance
16.1 Introduction
Fluoroquinolones are among the most efficacious and broad-spectrum oral antibac-
terials currently in clinical use [1–4]. They are used as frontline treatments for a
wide variety of infections caused by Gram-negative and Gram-positive bacteria [5].
Among the diseases treated with fluoroquinolones are urinary tract infections and
pyelonephritis, sexually transmitted diseases, prostatitis, skin and tissue infections,
chronic bronchitis, community-acquired and nosocomial pneumonia, and intra-
abdominal and pelvic infections [5]. Fluoroquinolones are also the first line of pro-
phylactic treatment for anthrax, the “biological agent most likely to be used” in a
bioterrorist attack, according to the Centers for Disease Control and Prevention
(CDC) [6]. Furthermore, they are commonly used to treat tuberculosis in cases of
resistance or patient intolerance to established regimens [7]. Tuberculosis recently
overtook HIV/AIDS as the deadliest disease in the world caused by a single infec-
tive agent [8].
Fluoroquinolones kill bacteria by increasing levels of double-stranded DNA
breaks generated by enzymes known as type II topoisomerases [2, 9–12]. The vast
E. G. Gibson
Vanderbilt University School of Medicine, Department of Pharmacology,
Nashville, TN, USA
R. E. Ashley
Vanderbilt University School of Medicine, Department of Biochemistry, Nashville, TN, USA
R. J. Kerns
University of Iowa College of Pharmacy, Department of Pharmaceutical Sciences and
Experimental Therapeutics, Iowa City, IA, USA
N. Osheroff (*)
Vanderbilt University School of Medicine, Departments of Biochemistry and Medicine, VA
Tennessee Valley Healthcare System, Nashville, TN, USA
e-mail: [email protected]
majority of bacteria encode two type II enzymes, gyrase and topoisomerase IV [10,
13–15]. These enzymes are essential for cell survival, and both appear to be physi-
ological targets for fluoroquinolones [2, 9, 10, 14]. In contrast, a handful of species
encode only gyrase. This group includes a number of disease-causing organisms,
including Treponema pallidum (syphilis) [16], Helicobacter pylori (stomach and
intestinal ulcers) [17], Campylobacter jejuni (gastroenteritis) [18], Mycobacterium
leprae (leprosy) [19], and Mycobacterium tuberculosis (tuberculosis) [20]. In these
species, gyrase takes on dual characteristics and can fulfill its own functions as well
as those of topoisomerase IV [21].
Unfortunately, fluoroquinolone usage is being threatened by an increasing preva-
lence of resistance, which extends to every bacterial infection treated by this drug
class [1, 2]. The most common and clinically relevant form of resistance is target-
mediated, which is caused by specific mutations in gyrase and topoisomerase IV
[22, 23]. Therefore, it is critically important to understand how this drug class inter-
acts with and alters the activity of its enzyme targets to better guide drug develop-
ment and to overcome resistance [11]. In this chapter, we will discuss fluoroquinolone
action and targeting, resistance mechanisms, and efforts to overcome this
resistance.
16.2 Fluoroquinolones
The history of the fluoroquinolones began in 1962, when Lesher et al. made the
accidental discovery of nalidixic acid (Fig. 16.1) as a by-product of the synthesis of
the antimalarial compound chloroquine [24]. This first-generation quinolone dis-
played limited efficacy and was used mainly for the treatment of uncomplicated
urinary tract infections caused by Gram-negative enteric bacteria [25]. In the 1980s,
the second generation of quinolones was established when norfloxacin (Fig. 16.1)
was synthesized [1, 25, 26]. This drug featured a fluorine at the C6 position, making
it the first true fluoroquinolone, and a cyclic diamine piperazine at the C7 position.
The fluorine at the C6 position increased tissue penetration and has been included
in every subsequent clinically relevant member of this drug class [1, 25–27].
Even with improved tissue penetration, norfloxacin was still confined to the uri-
nary tract and displayed low serum concentrations [1, 25–27]. However, it broad-
ened the use of quinolones to include sexually transmitted diseases [1, 25–27].
Ciprofloxacin (Fig. 16.1) was the first fluoroquinolone to display efficacy toward
both Gram-positive and Gram-negative bacterial species and was the first with suf-
ficiently high tissue penetration and serum concentration to be used outside the
urinary tract [1, 25, 26]. The clinical success of ciprofloxacin spawned the develop-
ment of third-generation fluoroquinolones that include moxifloxacin, gatifloxacin,
and levofloxacin (Fig. 16.1) [1, 25–27]. These drugs all exhibit improved half-lives
compared to ciprofloxacin [28]. Moreover, they have extended the spectrum of
fluoroquinolone activity to include a broader array of Gram-positive bacteria
(including a number of respiratory infections), as well as atypical pathogens such as
Legionella pneumophila and Chlamydia pneumoniae [2–5].
16 Bacterial Type II Topoisomerases and Target-Mediated... 509
Fig. 16.1 Fluoroquinolone structures. Nalidixic acid, a first-generation quinolone, is the founding
member of this drug class. This drug displayed limited efficacy for systemic infections and had a
narrow antibacterial spectrum. Norfloxacin and ciprofloxacin, second-generation fluoroquino-
lones, had improved efficacy, with ciprofloxacin being the more efficacious of the two.
Ciprofloxacin displayed an improved antibacterial profile that included additional Gram-positive
bacterial infections and improved Gram-negative coverage. Moxifloxacin, gatifloxacin, and levo-
floxacin, third-generation fluoroquinolones, are the most efficacious and broad-spectrum fluoro-
quinolones in clinical use today
510 E. G. Gibson et al.
Type II topoisomerases control the topological state of the DNA in the cell [9–11,
29–31]. These enzymes modulate DNA under- and overwinding (i.e., negative and
positive supercoiling, respectively) and remove tangles and knots from the genome
[11, 12, 31–35]. As discussed earlier, most bacterial species encode two type II
topoisomerases, gyrase and topoisomerase IV [9–11, 29–31]. Gyrase was the first
type II topoisomerase to be described in any species and was originally reported in
1976 [36]. It is an A2B2 heterotetramer in which the two subunit types are GyrA and
GyrB [9–11, 29–31]. The A subunits contain the active site tyrosine residues that
cleave the DNA (shown in blue in Fig. 16.2). The B subunits form the N-terminal
gate of the enzyme and contain the sites of ATP binding and hydrolysis (shown in
green in Fig. 16.2) [9–11, 29–31].
The subunits of topoisomerase IV were first identified in Gram-negative species
as being required for chromosome partitioning and were named ParC and ParE (blue
and green in Fig. 16.2, respectively) [10–13, 31, 37–39]. Sequence analysis revealed
that these proteins were homologous to GyrA and GyrB, respectively. In 1990, it
was determined that the ParC/ParE complex was a heterotetramer that functioned as
a distinct type II topoisomerase [10–13, 31, 37–39]. The enzyme was subsequently
named topoisomerase IV. Whereas the subunits of topoisomerase IV are denoted as
ParC and ParE in Gram-negative species because of their historic roles in chromo-
some partitioning, they are called GrlA and GrlB, respectively, (which comes from
their initial name gyrase-like proteins) in Gram-positive species.
Gyrase and topoisomerase IV regulate DNA topology by using a double-stranded
DNA passage mechanism [10–13, 31, 34]: the enzymes generate a double-stranded
break in the gate or G-segment (green in Fig. 16.2) and pass the transport or
T-segment (red) through the open DNA gate. The transport helix eventually exits the
enzymes when the two subunits at the bottom of the enzyme open to form the exit
gate. This reaction takes place at the expense of ATP binding, which opens the DNA
gate and induces a conformational change that moves the T-segment through the
open gate, and ATP hydrolysis, which drives enzyme turnover.
As a prerequisite for opening the DNA gate, gyrase and topoisomerase IV gener-
ate a double-stranded break in the G-segment [10–13, 31, 34]. The scissile bonds on
the two strands of the double helix are located across the major groove from one
another. Cleavage results in 5′-overhanging termini with a four-base cohesive stag-
ger. In order to maintain genomic integrity during the DNA cleavage event, the
enzymes form a covalent phosphotyrosine linkage between active site residues and
the newly generated 5′ termini. This covalent enzyme-cleaved DNA complex is criti-
cal for the actions of quinolones and is called the cleavage complex [10–13, 31, 34].
Despite the sequence and structural similarities between gyrase and topoisomer-
ase IV, differences in the C-terminal domains of GrlA/ParC and GyrA confer these
enzymes with unique catalytic activities [10–13, 31, 34]. The C-terminal domain of
GrlA/ParC allows topoisomerase IV to interact with two distal DNA segments.
Thus, the enzyme uses a “canonical” strand-passage mechanism in which it cap-
16 Bacterial Type II Topoisomerases and Target-Mediated... 511
Fig. 16.2 Cellular functions of bacterial type II topoisomerases. Topoisomerase IV (bottom left)
uses a canonical double-stranded DNA passage mechanism. The enzyme can remove positive
supercoils but acts primarily behind the replication fork (middle) to remove precatenanes and
unlink daughter chromosomes. Gyrase (bottom right) uses a DNA-wrapping mechanism that is
superimposed upon the double-stranded DNA passage reaction. The enzyme removes positive
DNA supercoils ahead of transcription (top) and replication (middle) complexes and maintains the
negative superhelicity of the genome. (Artwork by Ethan Tyler, NIH Medical Arts)
tures existing intra- or intermolecular DNA crossovers (Fig. 16.2, left) [10–13, 31,
34]. This allows the enzyme to relax (i.e., remove) positive or negative DNA super-
coils and to remove DNA tangles and knots in a highly efficient manner [10–13, 31,
34]. Although topoisomerase IV is able to alleviate torsional stress ahead of DNA
tracking systems and appears to play a role in regulating genomic superhelicity, its
major function is to remove the precatenanes that form behind DNA replication
forks (Fig. 16.2 middle), separate daughter chromosomes following replication, and
512 E. G. Gibson et al.
Fluoroquinolones
Fig. 16.3 Bacterial type II topoisomerases: the critical balance of DNA cleavage and ligation. The
DNA cleavage/ligation activity of bacterial type II topoisomerases must be regulated in the cell.
When an appropriate level of cleavage complexes is maintained, topological problems within DNA
are resolved, and the cell can grow normally. If the levels of cleavage complexes decrease, slow
growth rates and mitotic failure can cause cell death. Conversely, if the levels of cleavage com-
plexes are too high, the resulting strand breaks can block essential nucleic acid functions and
induce the SOS response, generate mutations, and lead to cell death. Compounds that increase
levels of gyrase or topoisomerase IV cleavage complexes, such as fluoroquinolones, act as topoi-
somerase “poisons” and convert the proteins to cellular toxins that have the potential to fragment
the genome. Compounds that inhibit the catalytic activity of gyrase or topoisomerase IV without
increasing levels of DNA cleavage work by robbing the cell of essential enzyme function leading
to slow growth rates, mitotic failure, and cell death. (Adapted from Pendleton et al. [46])
remove DNA knots that form during DNA recombination [13, 40–45]. If topoisom-
erase IV activity drops below threshold levels, cells die of mitotic failure (Fig. 16.3)
[10–13, 31, 34].
In contrast to the canonical mechanism used by topoisomerase IV, gyrase uses a
mechanism in which the C-terminal domain of the GyrA subunit (gray, Fig. 16.2)
wraps DNA, inducing a positive crossover between the G- and T-segments that
mimics a positive supercoil [13, 47–49]. Because of this “wrapping” mechanism,
the captured G- and T-segments are proximal to one another [50]. As a result, gyrase
greatly favors the catalysis of intra- over intermolecular strand-passage reactions.
Consequently, the enzyme can efficiently alter superhelical density but is very poor
at removing tangles and knots [50, 51]. In addition, because gyrase always acts on
the induced positive crossover, it works in a unidirectional manner [13, 52]; in the
presence of ATP, the enzyme can remove positive, but not negative supercoils.
Furthermore, gyrase is able to induce negative supercoils into relaxed DNA [36,
53]. The major cellular roles of gyrase stem from its DNA-wrapping mechanism.
Gyrase functions ahead of replication forks and transcription complexes to alleviate
the torsional stress induced by DNA overwinding (Fig. 16.2, middle and bottom)
[31, 54]. Furthermore, in conjunction with the ω protein (commonly called topoi-
somerase I), a type I topoisomerase, gyrase modulates the superhelicity of the bac-
16 Bacterial Type II Topoisomerases and Target-Mediated... 513
terial chromosome and allows the organism to maintain its genetic material in an
underwound state [55]. If gyrase activity in the cell drops, rates of replication/tran-
scription are severely impacted (Fig. 16.3) [9–13, 53]. Furthermore, a number of
pleiotropic effects on gene expression are observed due to changes in superhelicity
of the bacterial chromosome [56].
Although gyrase and topoisomerase IV are essential enzymes, they also pose a
threat to the bacterial cell. Indeed, if a replication fork, transcription complex, or
DNA tracking system encounters and attempts to pass through a gyrase- or topoi-
somerase IV-mediated DNA cleavage complex, it can disrupt the complex and ren-
der the enzyme unable to ligate the DNA [40]. This event generates double-stranded
DNA breaks that require recombination pathways to repair. Thus, these breaks
block essential nucleic acid functions, induce the SOS response, generate muta-
tions, and trigger processes that ultimately impair cell survival [2, 3, 12, 26, 57, 58].
Compounds that increase levels of gyrase or topoisomerase IV cleavage com-
plexes are referred to as “poisons” [59], because they are said to poison these pro-
teins, converting them to cellular toxins that have the potential to fragment the
bacterial chromosome [2, 9, 26, 57, 58]. The term “poison” distinguishes these
compounds from “catalytic inhibitors,” which act primarily by robbing the cell of
the catalytic functions of these enzymes [34].
Fluoroquinolones are potent gyrase/topoisomerase IV poisons [9, 11, 12, 23, 60–
62]. These drugs interact with both the protein and DNA within a cleavage complex
and intercalate into the DNA backbone at the cleaved scissile bonds [23, 62].
Consequently, two fluoroquinolone molecules are required to stabilize double-
stranded breaks induced by the bacterial type II enzymes (Fig. 16.4). The interca-
lated fluoroquinolones likely produce some distortions within the enzyme active
site; however, these drugs act primarily as “molecular doorstops” that form a physi-
cal barrier to DNA ligation [23]. Thus, the presence of fluoroquinolones inhibits the
rate of gyrase- and topoisomerase IV-mediated DNA ligation. Furthermore, drugs
that induce higher levels of enzyme-mediated DNA strand breaks appear to form
more stable interactions within the cleavage complex and allow these complexes to
persist for longer periods of time [63, 64].
In addition to generating DNA strand breaks in the cell, fluoroquinolones also
inhibit the overall catalytic strand-passage activities of gyrase and topoisomerase IV
[3, 11, 12]. As a result, there is debate as to whether the inhibition of strand passage
contributes to drug efficacy in cells. Although this issue has yet to be definitively
decided, a recent study suggests that the deleterious actions of drugs result primarily
from the enhancement of DNA cleavage [65]. In this study, the effects of ciprofloxa-
cin on three different fluoroquinolone-resistant mutations of Escherichia coli
topoisomerase IV that are associated with clinical resistance were examined in vitro.
With all three enzymes, ciprofloxacin displayed virtually no ability to enhance DNA
514 E. G. Gibson et al.
Fig. 16.4 Crystal structure of a topoisomerase IV-DNA cleavage complex formed with A. bau-
mannii enzyme in the presence of moxifloxacin. This structure is a top view of the cleavage com-
plex with two fluoroquinolone molecules intercalating four base pairs apart at the cleaved scissile
bonds. The presence of the intercalated fluoroquinolones likely produces distortions within the
enzyme active site; however, these drugs act primarily as “molecular doorstops” that form a physi-
cal barrier to ligation. The catalytic core of the enzyme (blue and green for the A and B subunit,
respectively), moxifloxacin (red), and DNA (yellow) are shown. (Adapted from Aldred et al. [11])
cleavage but showed wild-type ability to inhibit DNA relaxation catalyzed by the
type II enzymes. Therefore, it appears that the ability to induce DNA cleavage is the
primary factor that determines quinolone-induced cytotoxicity.
The World Health Organization (WHO) ranks fluoroquinolones as one of the five
“highest priority” and “critically important” classes of antimicrobials [66]. However,
due to their widespread use and overuse, resistance has been on the rise since the
1990s [1, 2, 11]. As an extreme example, the CDC has classified Neisseria gonor-
rhoeae, the causative agent of gonorrhea, as one of its top three “urgent level” drug-
resistant threats to the United States [67], primarily due to fluoroquinolone
resistance. Along with the WHO, it has issued dire warnings that gonorrhea is on the
precipice of joining HIV/AIDS and herpes as the third “incurable” sexually trans-
mitted disease [68].
Fluoroquinolones were used routinely to treat gonorrhea starting in 1993 and
were used in more than 40% of the cases by the year 2003 [69–71]. However, the
use of fluoroquinolones as frontline therapy against this disease was discontinued in
2006 due to the high incidence of resistance; 22.4% of cases reported in the United
16 Bacterial Type II Topoisomerases and Target-Mediated... 515
States in 2015 were resistant to fluoroquinolones (this value rose to 32.1% among
men who have sex with men) [69, 72, 73]. In parts of Asia, fluoroquinolone resis-
tance exceeds 80% [73]. Other infectious bacteria that have raised concerns due to
their high level of fluoroquinolone resistance include Campylobacter spp.,
Salmonella spp., and E. coli [74].
Thus far, three mechanisms of fluoroquinolone resistance have been described
[11, 58]. The first is “target-mediated resistance,” which results from specific muta-
tions in gyrase or topoisomerase IV [75–77]. The second is “plasmid-mediated
resistance,” which is caused by the presence of extrachromosomal DNA fragments
that encode three different classes of proteins [11, 78, 79]. Some plasmids encode
acetylases, which modify and inactivate quinolones and other drugs. Others encode
Qnr proteins, which block type II topoisomerases from binding to their DNA sub-
strates or to fluoroquinolones. Still others encode efflux pumps, which decrease the
fluoroquinolone concentration in cells. The third mechanism of fluoroquinolone
resistance is “chromosome mediated,” in which the expression of efflux pumps is
elevated or the expression of porins, which play a role in fluoroquinolone uptake, is
downregulated [11, 26, 78, 80].
Although the latter two mechanisms contribute significantly to fluoroquinolone
resistance, the target-mediated mechanism is generally the form most often associ-
ated with clinical resistance [11, 81, 82]. Because target-mediated resistance repre-
sents the most common and clinically relevant form of resistance, the remainder of
this chapter will focus on this mechanism.
Initial quinolone resistance is almost always associated with specific mutations
in gyrase, topoisomerase IV, or both. For example, in a recent clinical study on drug
resistance [83], 97% of 60 quinolone-resistant isolates of E. coli carried mutations
in gyrase, and 90% of these isolates also carried mutations in topoisomerase IV.
In general, the most commonly observed (up to ~90%) fluoroquinolone resis-
tance mutation is in a highly conserved serine residue that was first described as
Ser83 in the A subunit of E. coli gyrase [84–88]. This residue resides in helix-IV of
GyrA. The majority of other resistance mutations usually map to a conserved glu-
tamic/aspartic acid residue that is four amino acids downstream from the serine and
also resides in helix-IV. Mutations at these positions often provide a tenfold or
higher reduction in susceptibility to clinically relevant fluoroquinolones.
Corresponding mutations in E. coli topoisomerase IV also result in fluoroquinolone
resistance in vitro [85–88].
The prevalence of resistance mutations at the serine residue may reflect the fact
that this residue is highly conserved but nonessential. To this point, the common
mutations at this residue display no known phenotype, in cells or in vitro, with the
exception of fluoroquinolone resistance. It is not clear why this residue is conserved;
however, the presence of the serine appears to provide protection against nybomy-
cin, a naturally occurring antibiotic [89]. Thus, it has been proposed there has been
natural selection to maintain the serine in the bacterial genome. It is notable that
mutations at the glutamic/aspartic residue often decrease the overall catalytic
activity of gyrase and topoisomerase IV [77, 90]. This may explain why a higher
proportion of resistance mutations are observed at the serine residue.
516 E. G. Gibson et al.
Although the association of the serine and glutamic/aspartic residues with fluoro-
quinolone resistance was established in the late 1980s [99–102], the mechanism by
which they lead to resistance was described only recently. Ultimately, the mechanis-
tic basis for fluoroquinolone action and resistance turned out to be inextricably
linked [11, 63, 65, 103]. Thus, these two important aspects of fluoroquinolone-
enzyme interaction will be discussed together.
The initial insight into the roles of the serine and glutamic/aspartic acid residues
of fluoroquinolone actions and resistance came from structural studies of cleavage
complexes formed with topoisomerase IV or gyrase in the presence of fluoroquino-
lones [23, 60–62]. Although these studies all localized fluoroquinolones in the same
binding pocket, which was proximal to the conserved amino acid residues, there
was disagreement regarding drug orientation within the pocket. Furthermore, none
of the studies found that the bound fluoroquinolone was close enough to either
amino acid to form a direct interaction.
However, one of the structures (which examined the cleavage complex of
Acinetobacter baumannii topoisomerase IV formed in the presence of m oxifloxacin)
provided a potential mechanism by which mutations at the serine or glutamic/aspar-
tic residue could lead to fluoroquinolone resistance [23]. It had long been known
16 Bacterial Type II Topoisomerases and Target-Mediated... 517
Glu/Asp
O
O
HO Ser
Mg+2
O O
F
5 4
6 3 O
7
8 1 2
R1 N
R2
Fig. 16.5 A water-metal ion bridge mediates critical interactions between fluoroquinolones and
bacterial type II topoisomerases. A generic fluoroquinolone structure is depicted in black, water
molecules are in blue, Mg2+ is in orange, and the coordinating serine and glutamic/aspartic acid
residues are in green. Blue dashed lines indicate the interaction between the divalent metal ion,
four water molecules, and the C3/C4 keto acid of the fluoroquinolone. The green dashed lines
represent hydrogen bonds between the serine and glutamic/aspartic acid side-chain hydroxyl
groups and the water molecules
that the C3/C4 keto acid of fluoroquinolones chelates divalent metal ions, but the
physiological role of these bound metal ions, if any, was unknown. The structure of
A. baumannii topoisomerase IV was the first to capture this fluoroquinolone-metal
ion interaction within a cleavage complex. In this structure, the C3/C4 keto acid of
moxifloxacin chelated a non-catalytic magnesium ion that appeared to be coordi-
nated to four water molecules. Two of these water molecules were in sufficiently
close proximity to Ser84 and Glu88 (equivalent to E. coli GyrA Ser83 and Glu87)
to form hydrogen bonds. Thus, the authors suggested that this water-metal ion coor-
dination might play a role in mediating interactions between fluoroquinolones and
bacterial type II topoisomerases. A subsequent study that determined the structures
of cleavage complexes formed with M. tuberculosis gyrase in the presence of moxi-
floxacin, 8-methyl-moxifloxacin, ciprofloxacin, levofloxacin, or gatifloxacin also
observed the chelated metal ion, the associated water molecules, and the protein
contacts [64]. A generalized diagram of the proposed water-metal ion “bridge” that
facilitates fluoroquinolone interactions with the conserved serine and glutamic/
aspartic residues is shown in Fig. 16.5 [22, 65, 90].
The initial functional evidence for the existence and role for the water-metal ion
bridge in mediating fluoroquinolone activity and resistance came from biochemical
studies on B. anthracis topoisomerase IV [90]. These studies utilized wild-type and
drug-resistant enzymes that carried mutations in the serine (Ser81) and/or glutamic
acid (Glu85) residues. The authors demonstrated that (1) the ability of
fluoroquinolones to poison topoisomerase IV relied on the presence of a non-cata-
lytic divalent metal ion; (2) mutations in either the serine or glutamic acid restricted
518 E. G. Gibson et al.
the metal ions that could be used to support drug activity; and (3) mutations in either
amino acid decreased the affinity of the metal ion. Later studies extended these
conclusions to topoisomerase IV from E. coli and gyrase from B. anthracis and M.
tuberculosis [11, 63, 65]. Thus, it appears that the water-metal ion bridge is used to
mediate fluoroquinolone-enzyme interactions in a variety of bacterial species.
Furthermore, the loss of one or both of the amino acids that anchor the bridge is
sufficient to disrupt these interactions and cause drug resistance [22, 63, 65].
Despite the importance and apparent “universality” of the water-metal ion bridge,
it seems to be used differently by enzymes from different bacterial species.
Whereas the bridge is critical for the binding of clinically relevant fluoroquinolones
to B. anthracis gyrase and topoisomerase IV and M. tuberculosis gyrase, it is used
primarily to align fluoroquinolones in the active site of E. coli topoisomerase IV
[11, 63, 65, 90].
The divalent metal ion of the water-metal ion bridge interacts with fluoroquino-
lones through the C3/C4 keto acid of the drug scaffold [63, 65, 90, 97]. This may
explain why clinically relevant fluoroquinolones can accommodate such a wide
variety of substituents at the N1, C7, and C8 positions. Whereas substituents at the
latter positions are unlikely to form critical gyrase or topoisomerase IV interactions,
they may contribute minor or species-specific interactions. Furthermore, they may
influence the pharmacokinetics of the drugs.
Finally, the water-metal ion bridge appears to be the feature of drug-enzyme
interactions that allows discrimination between the bacterial and human type II
topoisomerases. Indeed, the amino acids in human topoisomerase IIα that corre-
spond to the serine and acidic residues of the bacterial helix-IV are methionine resi-
dues. This likely explains why clinically relevant fluoroquinolones display such poor
activity against the human type II enzymes. If these methionine residues in topoi-
somerase IIα are converted to serine and glutamic acid residues, the activity of cip-
rofloxacin and moxifloxacin against the human enzyme rises four- to fivefold [90].
Fig. 16.6 Fluoroquinolone-like compounds with unique properties. The quinazolinedione scaf-
fold is similar to the fluoroquinolone core; however, the loss of the C3/C4 keto acid disrupts the
ability to chelate the divalent metal ion used in the water-metal ion bridge. 8-Methyl-2,4-dione, a
quinazolinedione, and 8-methyl-moxifloxacin, a fluoroquinolone, overcome resistance by mediat-
ing interactions with bacterial type II enzymes through their C7 and C8 substituents, respectively.
CP 115,955, a fluoroquinolone, displays high activity against human (through the C7 substituent)
and bacterial (through the water-metal ion bridge) type II topoisomerases
These substitutions disrupt the C3/C4 keto acid required to chelate the divalent
metal ion used in the water-metal ion bridge. Thus, it is not surprising that interac-
tions between quinazolinediones and bacterial type II topoisomerases are indepen-
dent of bridge function and are unaffected by resistance mutations in the serine or
acidic amino acid residues that serve as bridge anchors. Consequently, it was
believed that quinazolinediones represented “fluoroquinolones” with a scaffold that
was impervious to classic resistance mutations [87].
However, three lines of evidence lead to the conclusion that the ability of quin-
azolinediones to overcome resistance results from the substituent at C7 rather than
the C3/C4 portion [22, 75]. First, the quinazolinediones reported in the literature
most often contained a 3′-(aminomethyl)-1-pyrrolidinyl [3′-(AM)P] (or related)
substituent at the C7 position. The 3′-(AM)P moiety is not represented in any clini-
cally relevant fluoroquinolone, opening the possibility that it has properties not pre-
520 E. G. Gibson et al.
Recent studies strongly suggest that substituents at the C8 position can have dra-
matic effects on fluoroquinolone resistance [63]. At the present time, structure-
activity relationship studies that examined the effects of C8 substituents on resistance
have been confined to relatively minor changes at this position: hydrogen, methyl,
or methoxy groups. However, major effects on resistance have been observed. In
general, compounds that include a methyl or methoxy group at C8 display higher
16 Bacterial Type II Topoisomerases and Target-Mediated... 521
activity against enzymes that carry mutations in the bridge-anchoring serine or glu-
tamic/aspartic acid. In some cases, dramatic differences in sensitivity have been
reported. For example, converting the C8 methoxy of moxifloxacin to a methyl
group results in a fluoroquinolone that poisons M. tuberculosis gyrase with twice
the potency and efficacy of moxifloxacin and completely overcomes clinically rel-
evant resistance mutations in a purified enzyme system [63, 64]. The fact that such
a minor alteration in fluoroquinolone structure can produce such a dramatic differ-
ence in the resistance profile of the drug suggests that the C8 position is a ripe target
for future drug discovery.
Despite the potential of C8 substituents for overcoming drug resistance, the basis
for the high activity of “8-methyl-moxifloxacin” against fluoroquinolone-resistant
M. tuberculosis gyrase is unknown. Although this compound induces gyrase-
mediated DNA strand breaks that are more stable than observed with moxifloxacin,
structural studies indicate that 8-methyl-moxifloxacin occupies a space within the
cleavage complex that is identical to that of the parent drug [63, 64]. Furthermore,
no specific protein or DNA contacts were observed with the C8 substituent. Thus,
future chemical studies will need to be combined with strong efforts in mechanistic
enzymology and structure in order to fully exploit the C8 substituent as a means to
overcome fluoroquinolone resistance.
Currently, fluoroquinolones are the only antibacterials in clinical use that target
gyrase or topoisomerase IV [9, 11, 23, 60–62]. However, recent drug discovery
efforts have resulted in new classes (two of which are in clinical trials) with clinical
potential. All of these compounds lack the keto acid that fluoroquinolones use in
conjunction with the water-metal ion bridge to interact with their bacterial targets.
Consequently, they all display activity against fluoroquinolone-resistant bacterial
strains.
Fig. 16.7 Novel gyrase/topoisomerase IV-targeted compounds. NBTIs, such as gepotidacin, act
as gyrase/topoisomerase IV poisons. However, in contrast to fluoroquinolones, they induce only
single-stranded DNA breaks. MGIs, such as GSK000, were derived from NBTIs in an effort to
optimize activity against M. tuberculosis. The founding member of the spiropyrimidinetrione class
of antibacterials is zoliflodacin (ETX0914/AZD0914)
MGIs (Fig. 16.7) were derived from NBTIs in an effort to optimize activity against
M. tuberculosis [111]. These compounds display high activity against wild-type and
fluoroquinolone-resistant strains. On the basis of mutagenesis studies, MGIs are
16 Bacterial Type II Topoisomerases and Target-Mediated... 523
16.8.3 Spiropyrimidinetriones
16.9 Conclusions
Fluoroquinolones are one of the most important and widely prescribed classes of
antibacterials in clinical use. However, their usefulness is being eroded by the rise
of drug resistance. Of the mechanisms that impair fluoroquinolone actions, those
that result from mutations in gyrase and topoisomerase IV are the most common
and detrimental. The mechanistic studies described above have led to a more com-
plete understanding of how fluoroquinolones interact with their enzyme targets and
how mutations alter these interactions. Furthermore, these studies have suggested
new strategies for overcoming resistance. Among these strategies are the design of
novel fluoroquinolones and the development of new drug classes that do not rely on
the water-metal ion bridge for their actions. Hopefully, these approaches will allow
gyrase and topoisomerase IV to remain important antibacterial targets in the decades
to come.
Major Points
• Although fluoroquinolones are the most efficacious and broad-spectrum oral
antibacterials in the clinic, their use is being eroded by resistance.
• Fluoroquinolone binding to gyrase and topoisomerase IV, the cellular targets of
these drugs, involves a water-metal ion bridge that is anchored by a keto acid on
the drug and a highly conserved serine and glutamic/aspartic acid residue in the
enzyme.
• Substitutions in the serine and glutamic/aspartic amino acid residues are respon-
sible for most of the target-mediated fluoroquinolone resistance.
• The development of compounds that interact with gyrase and topoisomerase IV
through mechanisms that are independent of the water-metal ion bridge may
provide an approach to bypassing existing target-mediated resistance.
524 E. G. Gibson et al.
Acknowledgments The authors of this article were supported by the US Veterans Administration
(Merit Review Award I01 Bx002198 to N.O.), the National Institutes of Health (grants AI87671 to
R.J.K., GM33944 and GM126363 to N.O., and GM007628 to E.G.G.), the National Science
Foundation (grant DGE-0909667 to R.E.A.), the Pharmaceutical Research and Manufacturers of
America Foundation (to E.G.G.), and the American Association of Pharmaceutical Scientists
Foundation (to E.G.G.).
References
1. Andriole VT. The quinolones: past, present, and future. Clin Infect Dis. 2005;41(Suppl.
2):S113–9.
2. Drlica K, Hiasa H, Kerns R, Malik M, Mustaev A, Zhao X. Quinolones: action and resistance
updated. Curr Top Med Chem. 2009;9(11):981–98.
3. Hooper DC. Mechanisms of action of antimicrobials: focus on fluoroquinolones. Clin Infect
Dis. 2001;32(Suppl. 1):S9–S15.
4. Linder JA, Huang ES, Steinman MA, Gonzales R, Stafford RS. Fluoroquinolone prescribing
in the United States: 1995 to 2002. Am J Med. 2005;118(3):259–68.
5. Hooper DC. Clinical applications of quinolones. Biochim Biophys Acta. 1998;1400:45–61.
6. Ressel G. Centers for Disease Control and Prevention. CDC updates interim guide-
lines for anthrax exposure management and antimicrobial therapy. Am Fam Physician.
2001;64(11):1901–2. 4
7. Jeon D. WHO treatment guidelines for drug-resistant tuberculosis, 2016 update: applicability
in South Korea. Tuberc Respir Dis (Seoul). 2017;80(4):336–43.
8. WHO. Global tuberculosis report 2016. 2016.
9. Anderson VE, Osheroff N. Type II topoisomerases as targets for quinolone antibacterials:
turning Dr. Jekyll into Mr. Hyde. Curr Pharm Des. 2001;7(5):337–53.
10. Champoux JJ. DNA topoisomerases: structure, function, and mechanism. Annu Rev
Biochem. 2001;70:369–413.
11. Aldred KJ, Kerns RJ, Osheroff N. Mechanism of quinolone action and resistance.
Biochemistry. 2014;53(10):1565–74.
12. Bush NG, Evans-Roberts K, Maxwell A. DNA topoisomerases. EcoSal Plus. 2015;6(2)
13. Levine C, Hiasa H, Marians KJ. DNA gyrase and topoisomerase IV: biochemical activities,
physiological roles during chromosome replication, and drug sensitivities. Biochim Biophys
Acta. 1998;1400:29–43.
14. Deweese JE, Burch AM, Burgin AB, Osheroff N. Use of divalent metal ions in the DNA
cleavage reaction of human type II topoisomerases. Biochemistry. 2009;48(9):1862–9.
15. Gentry AC. DNA topoisomerases: type II. In: Lennarz WJ, Daniel Lane M, editors. The
encyclopedia of biological chemistry. Waltham: Academic; 2013. p. 163–8.
16. Fraser CM, Norris SJ, Weinstock GM, White O, Sutton GG, Dodson R, et al. Complete genome
sequence of Treponema pallidum, the syphilis spirochete. Science. 1998;281(5375):375–88.
17. Tomb JF, White O, Kerlavage AR, Clayton RA, Sutton GG, Fleischmann RD, et al.
The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature.
1997;388(6642):539–47.
18. Parkhill J, Wren BW, Mungall K, Ketley JM, Churcher C, Basham D, et al. The genome
sequence of the food-borne pathogen Campylobacter jejuni reveals hypervariable sequences.
Nature. 2000;403(6770):665–8.
19. Cole ST, Eiglmeier K, Parkhill J, James KD, Thomson NR, Wheeler PR, et al. Massive gene
decay in the leprosy bacillus. Nature. 2001;409(6823):1007–11.
20. Cole ST, Barrell BG. Analysis of the genome of Mycobacterium tuberculosis H37Rv.
Novartis Found Symp. 1998;217:160–72. discussion 72–7
16 Bacterial Type II Topoisomerases and Target-Mediated... 525
21. Ashley RE, Blower TR, Berger JM, Osheroff N. Recognition of DNA supercoil geometry by
Mycobacterium tuberculosis gyrase. Biochemistry. 2017;56:5440.
22. Aldred KJ, Schwanz HA, Li G, McPherson SA, Turnbough CL, Kerns RJ, et al. Overcoming
target-mediated quinolone resistance in topoisomerase IV by introducing metal-ion-
independent drug–enzyme interactions. ACS Chem Biol. 2013;8(12):2660–8.
23. Wohlkonig A, Chan PF, Fosberry AP, Homes P, Huang J, Kranz M, et al. Structural basis of
quinolone inhibition of type IIA topoisomerases and target-mediated resistance. Nat Struct
Mol Biol. 2010;17(9):1152–3.
24. Lesher GY, Froelich EJ, Gruett MD, Bailey JH, Brundage RP. 1,8-Naphthyridine derivatives.
A new class of chemotherapeutic agents. J Med Pharm Chem. 1962;91:1063–5.
25. Emmerson AM, Jones AM. The quinolones: decades of development and use. J Antimicrob
Chemother. 2003;51(Suppl 1):13–20.
26. Mitscher LA. Bacterial topoisomerase inhibitors: quinolone and pyridone antibacterial
agents. Chem Rev. 2005;105(2):559–92.
27. Stein GE. The 4-quinolone antibiotics: past, present, and future. Pharmacotherapy.
1988;8(6):301–14.
28. Drusano G, Labro MT, Cars O, Mendes P, Shah P, Sorgel F, et al. Pharmacokinetics and
pharmacodynamics of fluoroquinolones. Clin Microbiol Infect. 1998;4(Suppl 2):S27–41.
29. Nitiss JL. Targeting DNA topoisomerase II in cancer chemotherapy. Nat Rev Cancer.
2009;9(5):338–50.
30. Deweese JE, Osheroff MA, Osheroff N. DNA topology and topoisomerases: teaching a
“knotty” subject. Biochem Mol Biol Educ. 2008;37(1):2–10.
31. Vos SM, Tretter EM, Schmidt BH, Berger JM. All tangled up: how cells direct, manage and
exploit topoisomerase function. Nat Rev Mol Cell Biol. 2011;12(12):827–41.
32. Wang JC. Cellular roles of DNA topoisomerases: a molecular perspective. Nat Rev Mol Cell
Biol. 2002;3(6):430–40.
33. Liu Z, Deibler RW, Chan HS, Zechiedrich L. The why and how of DNA unlinking. Nucleic
Acids Res. 2009;37(3):661–71.
34. Deweese JE, Osheroff N. The DNA cleavage reaction of topoisomerase II: wolf in sheep’s
clothing. Nucleic Acids Res. 2009;37(3):738–48.
35. Pommier Y, Leo E, Zhang H, Marchand C. DNA topoisomerases and their poisoning by anti-
cancer and antibacterial drugs. Chem Biol. 2010;17(5):421–33.
36. Gellert M, Mizuuchi K, O'Dea MH, Nash HA. DNA gyrase: an enzyme that introduces super-
helical turns into DNA. Proc Natl Acad Sci U S A. 1976;73:3872–6.
37. Bates AD, Berger JM, Maxwell A. The ancestral role of ATP hydrolysis in type II topoisom-
erases: prevention of DNA double-strand breaks. Nucleic Acids Res. 2011;39(15):6327–39.
38. Gentry AC, Osheroff N. DNA topoisomerases: type II. In:Encyclopedia of biological chem-
istry. Amsterdam: Elsevier Inc.; 2013. p. 163–8.
39. Corbett KD, Berger JM. Structure, molecular mechanisms, and evolutionary relationships in
DNA topoisomerases. Annu Rev Biophys Biomol Struct. 2004;33:95–118.
40. Hiasa H, Marians KJ. Topoisomerase IV can support oriC DNA replication in vitro. J Biol
Chem. 1994;269(23):16371–5.
41. Zechiedrich EL, Khodursky AB, Bachellier S, Schneider R, Chen D, Lilley DM, et al. Roles
of topoisomerases in maintaining steady-state DNA supercoiling in Escherichia coli. J Biol
Chem. 2000;275(11):8103–13.
42. Crisona NJ, Strick TR, Bensimon D, Croquette V, Cozzarelli NR. Preferential relaxation of
positively supercoiled DNA by E. coli topoisomerase IV in single-molecule and ensemble
measurements. Genes Dev. 2000;14(22):2881–92.
43. Wang X, Reyes-Lamothe R, Sherratt DJ. Modulation of Escherichia coli sister chromosome
cohesion by topoisomerase IV. Genes Dev. 2008;22(17):2426–33.
44. Joshi MC, Magnan D, Montminy TP, Lies M, Stepankiw N, Bates D. Regulation of sis-
ter chromosome cohesion by the replication fork tracking protein SeqA. PLoS Genet.
2013;9(8):e1003673.
526 E. G. Gibson et al.
45. Zawadzki P, Stracy M, Ginda K, Zawadzka K, Lesterlin C, Kapanidis AN, et al. The localiza-
tion and action of topoisomerase IV in Escherichia coli chromosome segregation is coordi-
nated by the SMC complex, MukBEF. Cell Rep. 2015;13(11):2587–96.
46. Pendleton M, Lindsey RH Jr, Felix CA, Grimwade D, Osheroff N. Topoisomerase II and
leukemia. Ann N Y Acad Sci. 2014;1310:98–110.
47. Liu LF, Wang JC. DNA-DNA gyrase complex: the wrapping of the DNA duplex outside the
enzyme. Cell. 1978;15(3):979–84.
48. Kampranis SC, Maxwell A. Conversion of DNA gyrase into a conventional type II topoisom-
erase. Proc Natl Acad Sci U S A. 1996;93(25):14416–21.
49. Kramlinger VM, Hiasa H. The “GyrA-box” is required for the ability of DNA gyrase to wrap
DNA and catalyze the supercoiling reaction. J Biol Chem. 2006;281(6):3738–42.
50. Ullsperger C, Cozzarelli NR. Contrasting enzymatic acitivites of topoisomerase IV and DNA
gyrase from Escherichia coli. J Biol Chem. 1996;271:31549–55.
51. Marians KJ. DNA gyrase-catalyzed decatenation of multiply linked DNA dimers. J Biol
Chem. 1987;262(21):10362–8.
52. Brown PO, Cozzarelli NR. A sign inversion mechanism for enzymatic supercoiling of
DNA. Science. 1979;206(4422):1081–3.
53. Drlica K, Snyder M. Superhelical Escherichia coli DNA: relaxation by coumermycin. J Mol
Biol. 1978;120(2):145–54.
54. Koster DA, Crut A, Shuman S, Bjornsti MA, Dekker NH. Cellular strategies for regulating
DNA supercoiling: a single-molecule perspective. Cell. 2010;142(4):519–30.
55. Pruss GJ, Manes SH, Drlica K. Escherichia coli DNA topoisomerase I mutants: increased
supercoiling is corrected by mutations near gyrase genes. Cell. 1982;31(1):35–42.
56. Menzel R, Gellert M. Regulation of the genes for E. coli DNA gyrase: homeostatic control of
DNA supercoiling. Cell. 1983;34(1):105–13.
57. Drlica K, Malik M, Kerns RJ, Zhao X. Quinolone-mediated bacterial death. Antimicrob
Agents Chemother. 2008;52(2):385–92.
58. Hooper DC. Mechanisms of fluoroquinolone resistance. Drug Resist Updat. 1999;2(1):38–55.
59. Kreuzer KN, Cozzarelli NR. Escherichia coli mutants thermosensitive for deoxyribonucleic
acid gyrase subunit A: effects on deoxyribonucleic acid replication, transcription, and bacte-
riophage growth. J Bacteriol. 1979;140(2):424–35.
60. Bax BD, Chan PF, Eggleston DS, Fosberry A, Gentry DR, Gorrec F, et al. Type IIA topoisom-
erase inhibition by a new class of antibacterial agents. Nature. 2010;466(7309):935–40.
61. Laponogov I, Pan XS, Veselkov DA, McAuley KE, Fisher LM, Sanderson MR. Structural
basis of gate-DNA breakage and resealing by type II topoisomerases. PLoS One.
2010;5(6):e11338.
62. Laponogov I, Sohi MK, Veselkov DA, Pan XS, Sawhney R, Thompson AW, et al. Structural
insight into the quinolone-DNA cleavage complex of type IIA topoisomerases. Nat Struct
Mol Biol. 2009;16(6):667–9.
63. Aldred KJ, Blower TR, Kerns RJ, Berger JM, Osheroff N. Fluoroquinolone interactions with
Mycobacterium tuberculosis gyrase: enhancing drug activity against wild-type and resistant
gyrase. Proc Natl Acad Sci U S A. 2016;113(7):E839–46.
64. Blower TR, Williamson BH, Kerns RJ, Berger JM. Crystal structure and stability of gyrase-
fluoroquinolone cleaved complexes from Mycobacterium tuberculosis. Proc Natl Acad Sci U
S A. 2016;113(7):1706–13.
65. Aldred KJ, Breland EJ, Vlčková V, Strub MP, Neuman KC, Kerns RJ, et al. Role of the
water-metal ion bridge in mediating interactions between quinolones and Escherichia coli
topoisomerase IV. Biochemistry. 2014;53(34):5558–67.
66. Collignon PC, Conly JM, Andremont A, McEwen SA, Aidara-Kane A, World Health
Organization Advisory Group BMoISoAR, et al. World Health Organization ranking of anti-
microbials according to their importance in human medicine: a critical step for developing
risk management strategies to control antimicrobial resistance from food animal production.
Clin Infect Dis. 2016;63(8):1087–93.
16 Bacterial Type II Topoisomerases and Target-Mediated... 527
Nature is rich in bioactive molecules that can be used as medicines. The earliest
records describing natural medicines are found on clay tablets dating from
2600 BCE in Mesopotamia; these records contain over 1000 plant derived-sub-
stances [1, 2]. The most well-known ancient medicinal record is the Egyptian Ebers
Papyrus, which dates from 1500 BCE and contains over 700 natural remedies, most
of plant origin [3]. Natural/herbal treatments are found throughout history and from
all over the world. Examples are the Chinese Materia Medica (Shennong Bencao
Jing) (1100 BCE – 659 CE) [3, 4] and the Indian Ayurvedic system (1000 BCE)
[3, 5]. These collections of ancient remedies were directed at a range of ailments
that included infections; some are still in use today. Such traditional medicines
generally consist of complex extracts and mixtures of agents whose bioactive
component(s) went unidentified for hundreds of years. Improvements came largely
from trial and error efforts that were hindered by confirmational bias and placebo
effects.
In the early 1800s, morphine became the first bioactive natural product isolated
from a medicinal plant [6]. This milestone led the Western medical field away from
Authors Fern R. McSorley and Jarrod W. Johnson contributed equally to this work.
F. R. McSorley · J. W. Johnson · G. D. Wright (*)
Department of Biochemistry and Biomedical Sciences, McMaster University, Michael
G. DeGroote Institute for Infectious Disease Research, McMaster University,
Hamilton, ON, Canada
e-mail: [email protected]
complex mixtures toward the pharmacology of pure compounds [7]. Old methods
based on impure mixtures were unreliable: the variability of growth conditions and
extractions from plant materials and microorganisms impacted the concentrations
of both beneficial and toxic bioactive compounds present in the mixtures. Isolating
the beneficial bioactive compound away from other material found in the natural
source, some of which could cause negative effects, allows in-depth analysis of
mode of action and efficacy. It also enables physicians to more accurately regulate
the dosage of the bioactive compound.
In 1928, just over 100 years after the isolation of morphine, Fleming serendipi-
tously discovered that the fungus Penicillium notatum secretes an agent that pre-
vents the growth of bacteria. A decade and a half later, penicillin became the first
natural-product antimicrobial to reach widespread clinical use [8]. Selman Waksman
coined the term antibiotics to refer to microbially produced compounds that are
“against life”; they either halt the growth of (bacteriostatic) or kill (bactericidal)
microbes. Among these are small peptides having antibiotic activity [9]; they typi-
cally disrupt bacterial membranes in a nonselective manner. Indeed, these peptides
are the first line of defense against bacterial infections. A larger suite of highly
selective antibiotics is produced by microbes. These compounds have been the main
source of our antibiotic drugs thus far [10]. The physiological roles of microbially
produced antibiotics are still debated; these molecules evolved either for signaling
functions or as chemical warfare agents to ward off neighboring microbes [11]. The
density and diversity of resistance elements in microbes are consistent with an
evolved detoxifying role to protect against the growth-impeding effects of antibiot-
ics. Regardless of the evolutionary basis for antibiotics, the introduction of penicil-
lin into the clinic in the early 1940s led to the “antibiotic era” of medicine. Natural
antibiotics have aided in the treatment and control of infections for the last 80 years
(Fig. 17.1). By controlling infections, antibiotics have revolutionized medicine,
allowing physicians to perform lifesaving organ transplants and invasive surgeries
and to treat cancer using disruptive immune system chemotherapy. The emergence
of antibiotic resistance now threatens these breakthroughs, our quality of life, and
our life expectancy.
Fig. 17.1 Timeline for the introduction of each class of antibiotic into clinical use. (Adapted from
ref. [12])
17 Natural Products in Antibiotic Discovery 535
The first antibacterial compounds to enter the clinic were synthetic molecules dis-
covered in the laboratory of Paul Ehrlich [13, 14]. After noticing that some microbes
stained differently than others when exposed to synthetic aniline and azo dyes,
Ehrlich postulated that a chemical compound could selectively target pathogenic
cells and not host cells. His “magic bullet” theory led him to screen hundreds of
organoarsenic derivatives for activity. One of these, arsphenamine (Salvarsan), was
successfully used to treat syphilis. Salvarsan, although difficult to administer, was
the most prescribed antimicrobial drug until it was replaced by penicillin [15].
Four classes of synthetic antibiotics1 remain successful as drugs in the clinic
today: sulfonamides, diaminopyrimidines, quinolones, and oxazolidinones
(Table 17.1, Fig. 17.2). The initial class of systemic synthetic antimicrobials was the
sulfonamide sulfa drugs [14, 17]. Prontosil, the first sulfa drug for human use, is a
prodrug; after administration, it is metabolized into the bioactive agent sulfanil-
amide. Sulfonamides work by inhibiting dihydropteroate synthase, a critical com-
ponent in folate synthesis. Humans acquire folate from their diet, while bacteria
must biosynthesize this essential nutrient. Consequently, sulfa drugs are selectively
active against microbes. Sulfonamides are generally co-administered with the
diaminopyrimidine trimethoprim. This synthetic antibiotic also targets folate bio-
synthesis, inhibiting dihydrofolate reductase. The synergistic combination of trim-
ethoprim and sulfamethoxazole (co-trimoxazole) is sold under a variety of trade
names (e.g., Septra, Bactrim).
The quinolone drugs target type II DNA topoisomerases and replication. These
agents are potent against both Gram-negative and Gram-positive bacteria. A widely
used example is ciprofloxacin (Fig. 17.2), a second-generation fluoroquinolone
that is orally available and used to treat urinary tract infections, sinusitis, and many
other infections. Several newer generations of quinolones have found a place in the
market, making the fluoroquinolones the most successful class of synthetic drugs
to date.
The oxazolidinones comprise the fourth class of synthetic antibiotic. These com-
pounds, which block the ribosomal peptidyl transferase center essential for protein
synthesis, are effective primarily against Gram-positive bacteria. Linezolid
(Fig. 17.2) represents the first generation of this class; it was approved by the US
Food and Drug Administration in 2000 and became the first novel chemical scaffold
to enter the clinic in several decades.
Fleming’s discovery that P. notatum secretes a bactericidal substance helped
launch the intensive mining of microbes as sources of antibiotics rather than exam-
ining synthetic chemical libraries [8]. Florey, Chain, and Heatley’s efforts to develop
a protocol for isolation of penicillin, followed by in vivo efficacy studies in animals
and clinical tests in humans, showed that penicillin, a natural product, was a viable
1
Here we are deviating from Waksman’s definition of antibiotics restricted narrowly to compounds
synthesized by microbes to include synthetic and semisynthetic human-made compounds as well.
Table 17.1 Timeline of discovery and introduction of antibiotic classes
Resistance
Antibiotic class; examples Discovery Introduction observed Mechanism of action Activity; target species
Organoarsenics; salvarsan 1909 1910 1924 Unknown Bactericidal; antisyphilitic
Sulfadrugsa; prontosil, sulfanilamide 1932 1936 1942 Inhibits dihydropteroate synthetase in folate synthesis Bacteriostatic; Gram-positive bacteria
β-Lactams; penicillins, 1928 1938 1945 Inhibits penicillin-binding proteins in cell-wall Bactericidal; broad-spectrum
cephalosporins, carbapenems biosynthesis
Aminoglycosides; streptomycin, 1943 1946 1946 Binds 30S ribosomal subunit inhibiting protein synthesis Bactericidal; broad-spectrum
spectinomycin, kanamycin,
neomycin
Chloramphenicols; 1946 1948 1950 Binds 50S ribosomal subunit inhibiting protein synthesis Bacteriostatic; broad-spectrum
chloramphenicol
Macrolides; erythromycin, 1948 1951 1955 Binds peptidyl transferase center of 50S ribosomal Bacteriostatic; broad-spectrum
clarithromycin subunit inhibiting protein synthesis
Tetracyclines; chlortetracycline, 1944 1952 1950 Binds 30S ribosomal subunit inhibiting protein synthesis Bacteriostatic; broad-spectrum
doxycycline
Ansamycins; rifamycins rifampicin 1957 1958 1962 Binds RNA polymerase β-subunit inhibiting RNA synthesis Bactericidal; Gram-positive bacteria
Glycopeptides; vancomycin, 1953 1958 1960 Transpeptidase blockade inhibiting cell-wall biosynthesis Bactericidal; Gram-positive bacteria
teicoplanin
Quinolonesa; ciprofloxacin 1961 1968 1968 Binds DNA gyrase inhibiting DNA synthesis Bactericidal; broad-spectrum
Streptogramins; pristinamycins 1963 1998 1964 Binds 50S ribosomal subunit inhibiting protein synthesis Bactericidal, Gram-positive bacteria
Oxazolidinones; linezolid 1955 2000 2001 Binds peptidyl transferase center of 50S ribosomal Bacteriostatic; Gram-positive
subunit inhibiting protein synthesis bacteria
Lipopeptides; daptomycin 1986 2003 1987 Depolarization of cell membrane Bactericidal; Gram-positive bacteria
Mutulins; pleuromutilin, 1951 2007 1999 Binds peptidyl transferase center of 50S ribosomal Bacteriostatic; Gram-positive
retapamulin subunit inhibiting protein synthesis bacteria
Fidaxomicin (targeting Clostridium 1975 2011 1977 Inhibition of RNA polymerase Bactericidal; Gram-positive bacteria
difficile)
Diarylquinolinesa; bedaquiline 1997 2012 2006 Inhibition of F1FO-ATPase Narrow-spectrum activity,
Mycobacterium tuberculosis
a
Synthetic antibiotic classes. Table assembled from references [11, 12, 16]
17 Natural Products in Antibiotic Discovery 537
O
O O NH2
S F CO2H O
OMe O
NH2 NH2 N N
N N N N O
N H 2N N OMe
HN F
OMe NHAc
H 2N
prontosil trimethoprim ciprofloxacin linezolid
(sulfonamide) (diaminopyrimidine) (fluoroquinolone) (oxazolidinone)
antibiotic drug candidate. The discovery of penicillin gave rise to the “golden age”
of antibiotic discovery (Fig. 17.1). For the next 20 years, extensive screening of
microbes, particularly soil-dwelling actinomycetes, was conducted to identify natu-
ral antibiotic compounds. In that time many natural-product scaffolds (aminoglyco-
sides, chloramphenicol, macrolides, tetracyclines, ansamycins, glycopeptides, and
streptogramins) were identified and often rapidly developed for clinical use
(Table 17.1). By the middle of the 1960s, this simple but effective screening plat-
form [18] appeared to have exhausted its sources, as new antibiotic scaffolds suit-
able for drug development became harder and harder to find.
The decline of success in isolating new chemical scaffolds from natural sources
suitable for drug development, the generally poor pharmacological and toxicologi-
cal properties of natural products as drugs, and the problematic emergence of
resistance ushered in the next phase of antibiotic innovation. The focus shifted to
medicinal chemistry efforts as we entered the “medicinal chemistry era.” Synthetic
chemists prepared and derivatized core antibiotic scaffolds already used in the clinic
and screened them for improvements. Chemists were successful in creating so-
called “generations” of enhanced synthetic variants that improved pharmacological
properties, expanded antibiotic spectra, and evaded resistance. Although many
improved drugs and new generations of known drugs emerged from these efforts, no
truly novel chemical scaffold entered the clinic from the 1960s to 2000s.
The lack of innovation in antibiotic discovery over the past two decades and the
general failure of in vitro target-based drug discovery methods have prompted a
renewed interest in natural products [16, 19]. This return to the natural-product
compounds that previously dominated antibiotic drug discovery reflects the historic
success of these drugs, a growing understanding of the physiochemical properties of
small molecules for efficacy against bacterial targets, and new thinking resulting
from advances in bacterial genomics, synthetic biology, and the properties of anti-
microbial targets.
Microbial natural products are the source of most of our antibiotic scaffolds in cur-
rent clinical use. Brief descriptions of the most prominent, clinically used drugs and
their modes of action are outlined below.
538 F. R. McSorley et al.
β-Lactams
Penicillin falls into the β-lactam category of natural products. Five classes of
β-lactams are important in the antibiotic field: penams, cephems, carbapenems, cla-
vams, and monobactams (Fig. 17.3). All β-lactams contain a strained 4-membered
β-lactam ring system that in the majority of clinically relevant compounds is fused
to a 5- or 6-membered ring system.
Penams, cephems, and carbapenems covalently bind to penicillin-binding proteins
(PBPs), essential enzymes that process the D-Ala-D-Ala termini of the pentapeptide
portion of the peptidoglycan intermediates in cell-wall metabolism. The electro-
philic β-lactam antibiotics mimic the D-Ala-D-Ala substrate (Fig. 17.4) and cova-
lently bind to the PBP [20], preventing the enzyme from facilitating transpeptidation
in the final step of peptidoglycan synthesis. This results in a complex series of
molecular events that include inhibition of cell division and eventually cell
rupture.
HO R
A H H R' H H H OMe H H H R' R' H
RHN S R S RHN S RHN S RHN R'' R O
R'
X X R''
N R'' N N N N N N
O O O X O X O O X O
CO2H CO2H CO2H CO2H CO2H CO2H
penam penem cephem cephamycin carbapenem monobactam clavam
H H H penicillin H H
B N S deacylase H 2N S acylation
6 5 2 semisynthetic
7
O N N 3
penicillins
O O
CO2H CO2H
penicillin G 6-APA
Me
O NH2 OMe
N H H H H H H O O Ph
N N H H H H H H
S S N O N
S N N S
O N O N H
O R O MeO O N N O N
O Et O
CO2H CO2H
CO2H CO2H
oxacillin ampicillin: R = H methicillin piperacillin
amoxicillin: R = OH
C
H H H enzymatic H H
HO2C N S H 2N S acylation
or 7 6 2 semisynthetic
8 3
NH2 O N OAc chemical N 4 3' OAc cephalosporins
O O
cleavage
CO2H CO2H
cephalosporin C 7-ACA
CO2H Me
N
O OMe OEt
N N N
H H H H H H H H H H H H
S N S N N S N N S N N S
H2N H2N H 2N N
O N OAc S O N N S O N N S N O N
S S
O O O Me O
CO2H CO2H CO2H CO2H
cefalothin ceftazidime cefepime ceftaroline
CO2H
D E
HO HO HO Me O
H H H H H H O
H N
NH2 N H NMe2 H H Me
5 1 N
S S S N H
N N N NH H 2N
O O NH O
S O N
CO2H CO2H CO2H O SO3H
thienamycin imipenem meropenem aztreonam
Fig. 17.3 β-Lactam antibiotics. (A) Major structural classes of β-lactam antibiotics. (B) The deac-
ylation of penicillin G to generate 6-APA opened the door for the synthesis of new semisynthetic
penicillins having improved antibiotic potency and spectrum of activity. (C) Similarly, 7-ACA has
been used as a key intermediate for the preparation of countless semisynthetic cephalosporins hav-
ing potent, broad-spectrum activity against clinically important Gram-negative bacteria. Ceftaroline
has activity against methicillin-resistant Staphylococcus aureus (MRSA). (D) Natural and syn-
thetic carbapenem antibiotics. (E) Aztreonam, a representative monobactam
17 Natural Products in Antibiotic Discovery 539
A Me
H Me Me H
CO2 CO2
O O
N H N
H S
H
penicillin N N D-Ala-D-Ala
H H
R H R Me
O O
C O
H HO O HO O H
O OH S S H 2N S N
N
N N N
N N N O B
O N CO2H
O O HO O
CO2H CO2H CO2H O OSO3Na
clavulanic acid sulbactam tazobactam avibactam vaborbactam
Fig. 17.4 β-Lactams mimic the D-Ala-D-Ala terminus of the peptidoglycan peptide strands and
block the cross-linking step in cell-wall biosynthesis. (A) Structural analogy of penicillin with the
D-Ala-D-Ala portion of peptidoglycan, as proposed by Tipper and Strominger in 1965 [20]. (B)
Interactions of a penicillin with penicillin-binding proteins (PBPs) and β-lactamases. The acyla-
tion of PBP active-site serine prevents cross-linking of the bacterial cell wall and leads to cell lysis.
In contrast, β-Lactamases hydrolyze β-lactams rapidly and confer high-level resistance. (C)
β-Lactamase inhibitors that are used in combination with β-lactams to overcome resistance due to
β-lactamases
Aminoglycosides
In 1944, Waksman’s laboratory discovered streptomycin, the first aminoglycoside
antibiotic, as a product of a strain of Streptomyces griseus [27]. Over the next two
decades, several members of the class were discovered that included kanamycin in
1956 from Streptomyces kanamyceticus and gentamicin C in 1963 from
Micromonospora purpurea [28, 29]. These were the first antibiotics to show effi-
cacy in the treatment of tuberculosis and in infections caused by Gram-negative
bacteria. Aminoglycosides consist of a core aminocyclitol ring modified to varying
extents by saccharides. Multiple amino groups provide a positive charge at physio-
logical pH and impart high water solubility. Currently, three aminoglycosides are
commonly used in the clinic: the natural products, gentamicin C and tobramycin,
and the semisynthetic agent amikacin, each of which contains the 2-deoxystreptamine
aminocyclitol core (Fig. 17.5a) [30]. Streptomycin, which is unique in this drug
class for having a streptamine core aminocyclitol ring, continues to find some clini-
cal use in the treatment of tuberculosis and in combination with penicillins for
enterococcal infections that are difficult to treat.
The amino and hydroxy groups of the aminoglycosides interact with the 16S
rRNA in the 30S ribosomal unit through a network of hydrogen bonds (Fig. 17.5b).
The binding of aminoglycosides to the 16S rRNA results in a conformational change
that impedes cognate codon–anticodon validation by the ribosome. The result is the
corruption of the genetic code and the synthesis of proteins with incorrect amino
acids. This corruption contributes to cell death.
Resistance to aminoglycosides can take many forms. Active efflux can be a sig-
nificant contributor to resistance in bacteria such as P. aeruginosa; however, the
main mechanisms of resistance are chemical modification of the drugs or the target
[30]. A large number of aminoglycoside N-acetyltransferases, O-phosphotransferases,
and O-adenylyltransferases are present in both Gram-positive and Gram-negative
pathogens (Fig. 17.5c). Over the past decade, ribosomal methyltransferases that
modify the 16S rRNA (e.g., G-1405 and A-1408) and confer high-level aminogly-
coside resistance in Gram-negative pathogens have also emerged as significant
clinical challenges (Fig. 17.5b).
Macrolides
The term macrolide is a portmanteau that combines macrolactone, a lactone ring
containing eight or more atoms, and polyketide. Erythromycin is a first-generation
macrolide that was isolated from a soil actinomycete Saccharopolyspora erythraea.
Erythromycin contains a 14-membered macrolactone framework and a 2-amino
sugar (Fig. 17.6). The presence of a ketone at position 9 of the macrolactone can
result in formation of a hemiketal with the hydroxyl group at position 6 under acidic
conditions (e.g., exposure to gastric acids), thereby decreasing the levels of bio-
available drug [31]. Semisynthetic conversion of erythromycin to clarithromycin or
azithromycin removes this possibility and results in improved efficacy and bioavail-
ability. Macrolide antibiotics are most effective against Gram-positive bacteria, but
they also have efficacy against some common Gram-negative, upper respiratory
tract pathogens.
17
HN
A NH2 OH OH OH
HN OH R
HO O R OH HO O H2N R Me O H2N
Me O O OH O O O
HO NH H2N H2N MeHN
HO OH NH2 HO OH NH2 HO OH NH2
OHC OH NH 2 O O O O
O O
HN H2N NH2 H2N NH2 H2N NH2
HO O
HO
O streptomycin kanamycin A: R = OH tobramycin: R = OH gentamicin C2: R = Me
HO NHMe
HO kanamycin B: R = NH2 dibekacin: R = H gentamicin B: R = H
NH2 OH OH
NH2 OH
HO O O O
HO HO OH Me H2N
HO O H2N O MeHN H
H2N NH2 OH OH O N
O HO NH2 HO
OH O O O
O O NH HO O NH2 HO HN NH2 OH
HN
HO OH
H2N O H2N O
HO OH butirosin amikacin plazomicin
Natural Products in Antibiotic Discovery
Fig. 17.5 Aminoglycosides. (A) Structures of natural and semisynthetic aminoglycosides. (B) Neomycin B and its binding mode in the H44 region of the E.
coli ribosome (PDB: 4V9C). Asterisks at G-1405 and A-1408 indicate sites of methylation by methyltransferases that confer aminoglycoside resistance (e.g.,
ArmA and NpmA, respectively). (C) Major target sites of aminoglycoside-modifying enzymes (AMEs), including O-phosphotransferases (APHs),
541
O Me
A Me Me Me N Me
N N
O
9 N Me
1 Me
HO OR HO OMe O
N
Me OH 6 Me NMe2 Me OH Me NMe2 OMe
Me HO Me HO O Me
Et O O O Me Et O O O Me Me NMe2
Me HO
NH2 O O O Me
O O OR2 O O OMe
Me
Me Me Me Me O O
O OH O OH F Me
Me Me
erythromycin A: R1 = H, R2 = H azithromycin solithromycin
erythromycin C: R1 = H, R2 = Me
clarithromycin: R1 = Me, R2 = Me
C
Me Me Me Me Me Me Me Me Me Me Me Me
AcO AcO AcO AcO
O O O O O O N O O N
O O HO N N
O O R O O
Me Me Me NMe Me
Me
rifamycin S rifamycin SV: R = H rifampin rifaximin
rifamycin B: R = CH2CO2H
Fig. 17.6 Selected examples of macrolides (A), tetracyclines (B), and ansamycin antibiotics (C)
F. R. McSorley et al.
17 Natural Products in Antibiotic Discovery 543
Macrolides interact with the peptidyl transferase center of the 50S ribosomal
subunit, blocking peptide-chain elongation. Specifically, the hydrophobic surface of
the macrolide binds to the sidewall of the exit tunnel and causes premature release
of short peptidyl-tRNAs [31].
The main causes of clinical macrolide resistance are the 23S rRNA methyltrans-
ferases (Erm) [32]. However, a number of macrolide kinases, which modify the
desosamine sugar, are increasing in prevalence and diversity. Ring-opening ester-
ases and desosamine glycosyltransferases are also known, but they are less fre-
quently encountered. Active macrolide efflux is common in many Gram-positive
pathogens.
Tetracyclines
Chlortetracycline (Aureomycin) and oxytetracycline (Terramycin), discovered in
1948 and 1950, respectively, were the first members of the tetracycline class
(Fig. 17.6b). Tetracyclines contain a tetracyclic polyketide core of four fused
6-membered rings; they were the first broad-spectrum antibiotics (i.e., effective
against both Gram-positive and Gram-negative pathogens) to enter clinical use.
After it was found that the C6-hydroxy group could be reductively removed to form
a more stable 6-deoxytetracline, further modification ensued. Multiple generations
of tetracyclines have now emerged that include doxycycline, minocycline, and tige-
cycline [33]. Tigecycline (approved for use in the USA in 2005) has broad-spectrum
activity for both Gram-positive and Gram-negative bacteria; it is also effective
against methicillin-resistant Staphylococcus aureus (MRSA).
Before the emergence of resistance limited their use, the tetracyclines were used
for decades to treat infections of the respiratory tract, middle ear, and urinary tract.
Like aminoglycosides, tetracyclines bind to the 30S ribosomal subunit. However,
they do not cause the production of aberrant proteins. Instead they bind to the
aminoacyl-tRNA binding site of the ribosome, thereby competitively preventing
translation of mRNA.
Tetracycline resistance in Gram-negative bacteria is most often the result of
active efflux. In Gram-positive bacteria, the expression of ribosomal protection pro-
teins lowers the affinity of tetracycline for the bacterial ribosome. Enzyme-mediated
inactivation has been reported through the action of TetX, a flavin-dependent mono-
oxygenase that hydroxylates the antibiotic, thereby precipitating a nonenzymatic
breakdown of the compounds [34].
Ansamycins
As with the macrolides, ansamycins are polyketide macrocycles; however, ansa-
mycins cyclize to form a macrolactam instead of a macrolactone (Fig. 17.6c). The
ansa-bridged macrolactam is formed with an intramolecular amine nucleophile
derived from the biosynthetic starter unit, 3-amino-5-hydroxybenzoyl-CoA. The
natural ansamycin, rifamycin, is an 18-membered macrolactam that is converted
through semisynthesis to the commonly used rifampin. Addition of the pipera-
zinyl hydrazide to the rifamycin naphthyl core in rifampin increases its oral bio-
availability and broadens antimicrobial activity. Rifampin is a WHO essential
544 F. R. McSorley et al.
Glycopeptides
As their name implies, glycopeptides are peptides that are decorated with sugar
moieties. Vancomycin and teicoplanin (Fig. 17.7) exemplify this class; both have
been developed as Gram-positive-directed antibiotics. Vancomycin, which was dis-
covered in the early 1950s as a product of the actinomycete Amycolatopsis orienta-
lis, was used only sporadically for several decades, largely due to difficulties in
obtaining pure compound. However, use increased in the 1980s following the wide-
spread emergence of MRSA in hospitals [38]. Emergence of resistance in entero-
cocci (VRE) and then intermediate resistance in S. aureus (VISA) spurred the
development of second-generation glycopeptides such as telavancin, dalbavancin,
and oritavancin that are less susceptible to resistance [39].
Vancomycin and teicoplanin are highly cross-linked pentapeptides that have a
high affinity for D-Ala-D-Ala termini of uncrosslinked peptidoglycan chains.
Vancomycin forms five hydrogen bonds with the D-Ala-D-Ala terminus of lipid II
and prevents the formation of interpeptidyl cross-links by PBPs. That reduces the
integrity of the cell wall and leads to cell death. Although glycopeptides and
β-lactams both inhibit cell-wall biosynthesis, the glycopeptides sequester the sub-
strate of transpeptidation rather than directly interacting with the PBP catalyst.
Resistance to glycopeptide antibiotics can take two forms. In Enterococci, repro-
gramming of cell-wall biosynthesis to terminate in either D-Ala-D-Lac or D-Ala-D-
Ser reduces affinity of the antibiotic. In Staphylococcus aureus, acquisition of the
HO HO OH Cl
NH2
H
Me O N OH
Me OH
O NH
OH HO O O
vancomycin O O OH HO
OH HO O Me O
Cl Me OH
O O O O O
HO NH2 O OH
O O
HO Cl OH AcHN Cl Me OH
O O O Cl
O H H O O O O O
6 5 3 1 H H Me
N N NHMe N N O
O N 4 N 2 N O N N O Cl OH
H H N
H H H H O O O
NH O O O O H H
O HN NH2 N N NHMe
7 O N N
HO2C N
NH2 HO2C O H H H
OH O O HO HN O O
Me O O
OH HO
HO X HO HO HO2C NH2
N O O OH
H HO
peptide stem O Me OH OH
HO HO
OH teicoplanin A2-2 oritavancin
oflipidII -D-Ala-D-Ala: X = NH
-D-Ala-D-Lac: X = O
Fig. 17.7 Glycopeptide antibiotics and the interaction between vancomycin and the D-Ala-D-Ala
portion of lipid II. In vancomycin-resistant enterococci (VRE), the peptide stem contains a D-Ala-
D-Lac terminal region, and the affinity for the glycopeptide is decreased 1000-fold [37]
17 Natural Products in Antibiotic Discovery 545
Streptogramins
Dalfopristin and quinupristin (Fig. 17.8a) are semisynthetic derivatives of virginia-
mycins/pristinamycins that belong to the A-type and B-type streptogramin families.
Type B streptogramins are cyclic hexapeptides, while type A streptogramins are
cyclic polyketide–peptide hybrids. This illustrates that a pair of molecules, which
have only bacteriostatic activity and are not effective treatments alone, can be com-
bined to work synergistically and form a potent bactericidal drug. The molecular
mechanism of synergy is based on the affinity of the compounds for different but
adjacent regions of the bacterial ribosome [40]. Dalfopristin binds to the peptidyl
transferase center where it reduces the affinity of aminoacyl-tRNAs for the amino-
acyl site, which lowers subsequent peptide bond formation and chain elongation in
the peptidyl site. In contrast, quinupristin binds in a similar manner to erythromycin
at the proximal end of the tunnel, thereby accelerating the release of small
oligopeptidyl-tRNAs [41]. When administered together, the two agents form a com-
bination drug known as Synercid. It is used to treat staphylococcal infections [42].
Both type A and B streptogramins are susceptible to efflux-mediated resistance;
indeed, the efflux protein Lsa intrinsic to Enterococcus faecalis confers resistance
to type B streptogramins, limiting Synercid use [40]. A group of O-acetyltransferases
confer high-level resistance to type A streptogramins, while Vgb is a ring-opening
C–O lyase that provides resistance to type B antibiotics.
Lipopeptides
As their name suggests, these compounds are peptides that contain a lipid moiety.
Both linear and cyclic, macrolactone and macrolactam lipopeptides exist. Due to the
large variations in structure, these molecules have few well-characterized cellular
targets. Daptomycin (Fig. 17.8b), initially discovered in the 1980s and discarded at
Phase II clinical trials by Eli Lilly due to toxicity, was revisited with a new dosing
regimen; it was approved for clinical use in 2003 [43]. Daptomycin has pleiotropic
effects on the membrane of Gram-positive bacteria that result in depolarization and
physical alteration of the cell membrane that leads to cell death [44]. Daptomycin is
effective against most Gram-positive pathogens, including drug-resistant forms
such as VRE and MRSA.
Colistin (polymyxin E) is a lipopeptide of the polymyxin class. It is one of the
few antibiotics in clinical use that was derived from a non-actinomycete bacterium,
Paenibacillus polymyxa. Discovered in 1949, colistin has been used sparingly for
the treatment of serious infections caused by Gram-negative bacteria due to toxic-
ity issues [45]. As a result of the rise of carbapenem-resistant Gram-negative
pathogens, clinicians have been left with few therapeutic options other than colis-
tin. Consequently, its use has increased significantly. The mode of action of poly-
myxins involves disruption of the outer and inner membranes of Gram-negative
bacteria [46].
546
O O
A N
O O S
H H
OH N OH N
N N N N
H OH H OH
Me O O Me O O
O O O O O O O O O O O O
N O N N O N
O O Me R O O Me NMe2
N N HN H N N HN H
O N O O N
N O S N
O O
H2N H2N
B C NH2
NH
O O
H Me HN
N NH
NH N N HO O O
H H O
CONH2 O NH O HN O O
O O HO2C HN
H OH O NH
N Me HO2C O O O O
N N H H H
H H N N N
NH O O O HO NH N N N
CO2H O H H H
H O O O
O N
NH2 N N O
H H NH2 NH2 NH2
O
O Me
colistin
daptomycin CO2H (= polymyxin E)
Cl Et
O
Me OMe
HO O O
HO O
Cl OH
Me Me
OH OH O OH HO O
S O OH
MeN H H
OH RO O O
Me Me HO Me Me Me Me Me
HN CHCl2 Me Me P
O2N Me Me HO O Me
O O HO
O O Me O O
chloramphenicol mutilin: R = H retapamulin fosfomycin O fidaxomicin
pleuromutilin: R = COCH2OH HO Me
Fosfomycin
The phosphonate fosfomycin is a small (138 Da) antibiotic produced by Streptomyces
fradiae. It has been in clinical use since the early 1970s for the treatment of urinary
tract infections. The rapid emergence of resistance has limited its use, but its high
water solubility and low toxicity enables a single dose (3 g of drug/day) or very
short course therapy. Fosfomycin targets bacterial cell-wall biosynthesis by inhibi-
tion of MurA, an enzyme involved in the first step in the biosynthesis of
N-acetylmuramic acid [52].
Fosfomycin resistance is readily selected during therapy in the form of mutants
in the glycerol-3-phosphate transporter, which is needed for fosfomycin entry into
the cell but not essential for bacterial cell growth [52]. A series of enzymes, includ-
ing glutathione transferases and epoxide hydrolases, are known to inactivate the
antibiotic via opening the essential epoxide ring.
Fidaxomicin
Fidaxomicin is the first member of the newest class of natural products to enter the
market (2011). It consists of an 18-membered macrolactone polyketide that was
discovered independently in Italy (lipiarrmycin), Japan (clostomicin), and the USA
(tiacumicin) in the early 1970s [53]. The macrolactone is decorated with two acyl-
ated rhamnoses (Fig. 17.9). Fidaxomicin inhibits RNA polymerase by binding to a
site distinct from the rifamycin binding site. There, fidaxomicin blocks the conver-
sion of bound promoter DNA to the open single-strand complex that forms the
transcription bubble. It has been approved for clinical use for treatment of
Clostridium difficile-associated diarrhea.
The vast majority of antibacterials used clinically are natural products, semisyn-
thetic derivatives, or analogues thereof. As mentioned above, the only synthetic
classes that are not derived from natural scaffolds are the sulfonamides, diaminopy-
rimidines, oxazolidinones, and quinolones. Natural products have always been a
major source of human medicines and continue to be especially important as leads
for antimicrobials – 74 of the 98 small molecules approved as antibacterials from
1981 to 2006 are natural products, semisynthetic derivatives, or natural-product
mimics [10].
In the 1980s the arrival of combinatorial chemistry allowed the rapid synthesis of
large numbers of synthetic compounds. That transformed the pharmaceutical indus-
try, and companies began to favor screening vast libraries of synthetic compounds
over natural-product extracts. With major advances in high-throughput screening
(HTS) technologies through the 1990s and 2000s, companies obtained the ability to
17 Natural Products in Antibiotic Discovery 549
quickly screen libraries of millions of compounds. This strategy has been enor-
mously successful in identifying lead compounds having targets in human cells, but
it has not been successful for identifying new antibacterials. Extensive high-
throughput screening campaigns at GlaxoSmithKline (GSK) [54] and AstraZeneca
[55] both failed to identify any candidate structure worthy of further clinical devel-
opment. The lack of success is due to a combination of several factors, with retro-
spective analyses of both campaigns pointing to a lack of chemical diversity in the
compound libraries.
The chemical libraries of pharmaceutical companies have largely been con-
structed with guidance from Lipinski’s Rules [56], which aim to improve the likeli-
hood of oral bioavailability by keeping molecular weights (MW) under 500,
measures of hydrophobicity (logP or logD) less than 5, and the number of hydrogen-
bond donors and acceptors in the molecule less than 10. However, antibiotics have
long been known to occupy a different “chemical space” than other drugs, and often
they exhibit multiple violations of Lipinski’s guidelines. An analysis of physio-
chemical properties by O’Shea and Moser in 2008 [57] compared a reference set of
human drugs against compounds active against Gram-positive and Gram-negative
bacteria. Average molecular weights were 338, 813, and 414, and average clogD7.4
values were 1.6, −0.2, and −2.8, respectively. Anti-Gram-positive compounds are
more polar than reference drugs and can be much larger, especially if their targets
are on the cell exterior (e.g., glycopeptides, lipopeptides). Compounds active against
Gram-negative bacteria, which must cross the outer membrane, are much more
polar and have a strict molecular weight cutoff at 600 Da, likely due to the limita-
tions of transport through porins. Overington pointed out, however, that the bacterial
target should be taken into account, since the physiochemical properties of antibiot-
ics targeting the ribosome fall further outside Lipinski’s rules than antibacterials
that have bacterial protein targets [58].
An analysis of 23 HTS campaigns at AstraZeneca, reported by Brown et al. [59],
showed that active antibacterial project compounds were significantly more polar
than the screening collection average. Improving biochemical potency through
chemical modification of active leads often came with an increase in hydrophobicity
and an increased probability of problematic plasma protein binding or cytotoxicity.
In cases in which biochemical potency was maintained by increasing polarity,
whole-cell activity remained elusive; designing polar compounds was not sufficient
for antibacterial activity. Overall, the study highlights the complexities of bacterial
cell penetration and efflux systems, especially in Gram-negative bacteria. The
authors note that one possibility for improving the antibiotic chemical space of
screening libraries would be to return to natural-product screening.
While there is little overlap in the chemical space of compounds in screening
libraries with that of antibacterials, there is far more overlap in the physiochemical
properties between antibiotics and natural products [60, 61]. In addition to hydro-
philicity (i.e., log P), other properties, such as the number of rotatable bonds
(molecular flexibility), polar surface area, H-bond donors and acceptors, molecular
complexity, and 3-dimensionality [62, 63], are better represented in natural-product
chemical space.
550 F. R. McSorley et al.
The time is right for a reevaluation of natural products in antibiotic drug discovery.
Their historical success as drugs, the comparative shortcomings of screens of syn-
thetic compound libraries, and the serious need for innovation in securing new anti-
biotics demand a fresh look at this source of bioactive chemistry. The rediscovery of
well-known chemical scaffolds, which prompted a move away from natural prod-
ucts, can be mitigated in several ways. First, previously unsuccessful scaffolds can
be reevaluated; second, successful antibiotic drugs can be reinvigorated by combin-
ing them with inhibitors of resistance and other antibiotic adjuvants; third, new
scaffolds can be sourced from previously neglected genera or through mining of
microbial genomes and metagenomes; and finally, “new-to-nature” compounds can
be generated through synthetic biology strategies.
The three most recent natural-product antibiotics to enter the clinic, daptomycin,
fidaxomicin, and the pleuromutilins, were all discovered and discarded decades
before their successful clinical launch. In the case of daptomycin, off-target human
toxicity was deemed a sufficient concern by Eli Lilly to halt clinical development.
A decade later, with more careful drug dosing to avoid undesired effects, daptomy-
cin was championed by Cubist Pharmaceuticals, which successfully brought the
compound to market [43]. Fidaxomicin was discarded in the 1970s due to poor solu-
bility and narrow spectrum, properties that are advantages in its new incarnation as
an orally dosed drug to combat C. difficile [53]. These examples offer hope, perhaps
even certainty, that new antibiotic drugs can be sourced from known compounds.
The estimate that ~28,000 natural-product antibiotics have been reported, while
fewer than 500 have entered into clinical use, is encouraging that we can revisit
these compounds as antimicrobial sources.
There are challenges to this route, however. A practical consideration is that there
is no ready way to obtain these compounds for reevaluation. Most were reported in
the scientific or patent literature decades ago, but some remain in the yellowing lab
books in the vaults of pharma. Unless the compounds progressed in the development
process, the bacterial strains that produce them may not be available in public cul-
ture collections. The fate of the extensive libraries of producing organisms held by
many companies active during the 1950s–1980s is not widely known. Some have
been captured by new entities. For example, the historical Merck collections are now
foundational resources of Fundación MEDINA and the Natural Products Discovery
Institute. Most strain libraries, however, are not easy to access. This means that
interesting chemical scaffolds may be lost until rediscovered by traditional screens.
Another challenge is securing intellectual property on known natural compounds
and their activities [67]. Nevertheless, a deep reservoir of knowledge and chemistry
exists that can be tapped for twenty-first-century antibiotic drug discovery.
552 F. R. McSorley et al.
H
A B C N NH2
O
O
O OH CONH2 N
N
HO2C HN O H
H O
HO2C N O NH O O O O HN
N CO2H H H H H
H HO OH MeHN N N N N
NH2 CO2H N N N N O
Me OH H H H H
O O O O
aspergillomarasmine A (AMA) bicyclomycin OH OH
teixobactin
OH OH Me Me OH Cl
HO O O
Me CO2H Me
Me O Me O O Cl OH O OH O NH2
HO2C
O
OH O
mupirocin enacyloxin IIa
(= pseudomonic acid A) Me
Me H
OH O Me O
Cl O NH
Cl
HO OH
Me O
H N NH2 O
MeO N Me H2N
O O N O Me O
Me O N CO2H Cl
O Me O O H
myxopyronin armeniaspirole C pantocin A salinosporamide A
The bulk of the actinomycetes screened by the pharmaceutical industry for antibi-
otic activities originate from soil environments. These sources are easy to access
and offer a wide variety of conditions for enriching various genera. The question of
whether such sampling reflects a reasonable representative distribution of microbial
natural-product diversity is unresolved. The same common scaffolds can be readily
found in samples from around the world, supporting the axiom that “everything is
everywhere, the environment selects.” However, careful genomic sampling of sig-
nature natural-product biosynthetic elements, such as ketosynthase domains from
polyketide synthases and adenylation domains of non-ribosomal peptide synthases
from a variety of soil environments, suggests that indeed there are significant envi-
ronmental differences in natural-product potential: there is significant chemical
diversity still to be identified [76]. Indeed, marine actinomycetes are sources of
several new compounds, many having biological activity (Fig. 17.11).
The fact that most of the focus in antibiotic natural-product discovery has been
on the actinomycetes has prompted a search in other orders of bacteria. The Gram-
negative betaproteobacteria, such as members of the genus Burkholderia, are prodi-
gious producers of antibiotics [77]. The gamma-proteobacteria, pseudomonads, and
the deltaproteobacteria, such as the Myxococci [78], also produce numerous natural-
product antibiotics (Fig. 17.11). These have only just begun to be mined to discover
new chemical scaffolds.
554 F. R. McSorley et al.
These sources though, are limited to strains that we can readily grow in the labo-
ratory. The “great plate count anomaly” refers to the fact that we are generally lim-
ited to growing <5% of the detectable microbes in a soil sample. Strategies to mine
this “microbial dark matter” offer ways to access new microbial genetic and chemi-
cal diversity [79]. An example of this approach is the iChip, a simple 96-compartment
device to capture microbes and grow them in situ, with access to nutrients and
growth factors of their natural environment [80]. Using this device and strategy, a
new antibiotic, teixobactin, which represents a new scaffold, was identified from a
previously uncultured bacterium [75]. Teixobactin, produced by a Gram-negative
bacterium, has a Gram-positive-only profile. Its mode of action involves binding to
lipid II, which is required for cell wall biosynthesis. Mining other difficult-to-grow
bacteria for new chemistry should be possible and therefore offers hope that addi-
tional antibiotic scaffolds can be identified.
OH
O
H
OH N N
NH2 O HN N OH
H H
N O O NH
HO2C N O HN O
H
O OH O
O O
H
O N NH
N
H HN NH HN NH
HO O HN
telomycin HN turbomycin A turbomycin B
While genome mining has greatly expanded our access to known and new anti-
biotic scaffolds, the majority of environmental microbes remain difficult to culture.
Here, metagenomic strategies in which total DNA is collected from a source (e.g., a
soil sample, animal, or plant microbiomes) are being employed. Such strategies are
yielding new antimicrobial compounds, such as the turbomycins [87], variants of
glycopeptides [94], and colicins [95] (Fig.17.12).
The long-term future for obtaining antibiotic diversity may be through the genera-
tion of nonnatural or synthetic natural products. This oxymoron refers to the engi-
neering of biosynthetic gene clusters to produce novel compounds, not yet known
to nature, using synthetic biology concepts [96–98]. The modularity of biosynthetic
gene clusters lends itself well to systematic synthetic biology. Biosynthetic gene
clusters include a predictable parts list: genes encoding scaffold assembly, tailoring
enzymes, supply of components not easily scavenged from primary metabolism
(amino acids, sugars, etc.), self-resistance, regulation, and transport (Fig. 17.13). In
principle, these elements can be mixed and matched to generate new compounds
having novel activities. For example, we have used this approach to generate new-
to-nature glycopeptide antibiotics that evade certain forms of resistance in VRE [99]
(Fig. 17.13).
As the costs of DNA synthesis continue to drop, one can envision synthesis of
large numbers of biosynthetic gene cluster parts, the combinatorial generation of
libraries of scaffolds, tailoring enzymes, regulatory elements, etc., and their
expression in a suitable heterologous host. The result would be millions of previ-
ously untested combinations of biosynthetic genes (Fig. 17.13). With suitable
selection, such libraries could deliver hits and lead for new antibiotic drug
development.
A Teicoplanin (Actinoplanes teichomyceticus)
556
Fig. 17.13 A synthetic biology approach for increasing chemical diversity in glycopeptide antibiotics. (A) Biosynthetic clusters for selected glycopeptides
with genes colored according to function. Significant portions of many clusters are comprised of tailoring genes responsible for decorating the glycopeptide
backbone (e.g., methyltransferases, sulfotransferases, and glycosyltransferases). (B) Novel glycopeptides have been generated by mixing biosynthetic genes.
In a recent study by Yim et al. [99], the biosynthetic clusters for A47934 and desulfo-A47934 were expressed in S. coelicolor together with a variety of tailoring
F. R. McSorley et al.
genes from the biosynthetic clusters of other glycopeptides. Several new A47934- and desulfo-A47934 derivatives were produced and are more potent than
vancomycin against E. faecalis and VRE B (C)
17 Natural Products in Antibiotic Discovery 557
Natural products, in particular those generated by bacteria and fungi, are the source
of the majority of our successful antibiotic drugs. These agents have changed the
world of medicine. For the first time in human history, we have good control over
infection. With that control has come much of modern medicine. Our natural-
product antibiotics have also helped us feed the world by changing the way we raise
and care for food animals. It is not hyperbole to suggest that natural-product antibi-
otics may be the most important scientific discovery of the twentieth century.
Unfortunately, the evolution of antibiotic resistance and its selection in once-
susceptible pathogens gravely threatens these advances. We need new antibiotics
and alternatives to maintain our control over infectious disease. The advances in our
knowledge of how natural products are made by microbes, new and unparalleled
access to the genetic determinants of natural-product biosynthesis by NGS of
microbial genomes and metagenomes, and the ability to harness this information to
identify and exploit this information are growing exponentially.
The proven efficacy of natural products as antibiotics, plus the disappointing
results of the past two decades of focus on synthetic compounds, means that we
must pivot back to these ancient compounds for leads and inspiration. There is good
reason to believe that the era of resistance depicted in Fig. 17.1 will be followed by
an era of anti-infective innovation.
Major Points
• Microbial natural products are the source of most of our successful antibiotic drugs.
• These compounds are the result of evolutionary processes that select for optimal
penetration and retention in target bacterial cells.
• The chemical diversity and physiochemical properties of microbial natural prod-
ucts cannot yet be effectively matched in most synthetic libraries.
• The re-isolation of known natural-product scaffolds diminished enthusiasm for
the natural-product approach in antibiotic discovery.
• Efforts to identify new antibiotic chemical diversity through revisiting discarded
compounds, mining of bacterial genomes, isolation of hitherto rare or unsampled
microbes, and increasing chemical diversity using synthetic biology strategies
offer new routes to identifying antibiotic leads.
• The use of inhibitors of resistance or other adjuvants can also extend the clinical
effectiveness of existing antibiotic scaffolds.
References
1. Borchardt JK. The beginnings of drug therapy: ancient Mesopotamian medicine. Drug News
Perspect. 2002;15(3):187–92.
2. Scurlock J. Sourcebook for ancient Mesopotamian medicine: Society of Biblical Literature;
2014.
558 F. R. McSorley et al.
3. Cragg GM, Newman DJ. Natural products: a continuing source of novel drug leads. Biochim
Biophys Acta. 2013;1830(6):3670–95.
4. Huang KC. The pharmacology of chinese herbs. 2nd ed. Boca Raton: CRC Press; 1999.
5. Kapoor LD. CRC handbook of ayurvedic medicinal plants. Boca Raton: CRC Press; 1990.
6. Sertuerner F. Ueber das Morphium, eine neue salzfähige Grundlage, und die Mekonsäure, als
Hauptbestandtheile des Opiums. Ann Phys (Berl). 1817;55(1):56–89.
7. Newman DJ, Cragg GM. Chapter 1 Natural products as drugs and leads to drugs: the historical
perspective. Natural Product Chemistry for Drug Discovery: The Royal Society of Chemistry;
2009. p. 3–27.
8. Fleming A. On the antibacterial action of cultures of a Penicillium, with special reference to
their use in the isolation of B. influenzæ. Br J Exp Pathol. 1929;10(3):226–36.
9. Zasloff M. Antimicrobial peptides of multicellular organisms. Nature. 2002;415(6870):389–95.
10. Newman DJ, Cragg GM. Natural products as sources of new drugs from 1981 to 2014. J Nat
Prod. 2016;79(3):629–61.
11. Walsh CT, Wencewicz T. Antibiotics: challenges mechanisms opportunities. Washington, DC:
ASM Press; 2016.
12. Brown ED, Wright GD. Antibacterial drug discovery in the resistance era. Nature.
2016;529(7586):336–43.
13. Ehrlich P, Hata S. Die experimentelle Chemotherapie der Spirillosen:(Syphilis, Rückfallfieber,
Hühnerspirillose, Frambösie). Wiesbaden: Springer; 1910.
14. Aminov RI. A brief history of the antibiotic era: lessons learned and challenges for the future.
Front Microbiol. 2010;1:134.
15. Mahoney JF, Arnold RC, Harris A. Penicillin treatment of early syphilis-a preliminary report.
Am J Public Health Nations Health. 1943;33(12):1387–91.
16. Lewis K. Platforms for antibiotic discovery. Nat Rev Drug Discov. 2013;12(5):371–87.
17. Domagk G. Ein Beitrag zur Chemotherapie der bakteriellen Infektionen. Dtsch Med
Wochenschr. 1935;61(7):250–3.
18. Lewis K. Antibiotics: recover the lost art of drug discovery. Nature. 2012;485(7399):439–40.
19. Wright GD. Opportunities for natural products in 21st century antibiotic discovery. Nat Prod
Rep. 2017;34(7):694–701.
20. Tipper DJ, Strominger JL. Mechanism of action of penicillins: a proposal based on their struc-
tural similarity to acyl-D-alanyl-D-alanine. Proc Natl Acad Sci U S A. 1965;54(4):1133–41.
21. Rolinson GD. Forty years of beta-lactam research. J Antimicrob Chemother.
1998;41(6):589–603.
22. Drawz SM, Papp-Wallace KM, Bonomo RA. New β-lactamase inhibitors: a therapeutic renais-
sance in an MDR world. Antimicrob Agents Chemother. 2014;58(4):1835–46.
23. Wang DY, Abboud MI, Markoulides MS, Brem J, Schofield CJ. The road to avibactam: the
first clinically useful non-β-lactam working somewhat like a β-lactam. Future Med Chem.
2016;8(10):1063–84.
24. Castanheira M, Rhomberg PR, Flamm RK, Jones RN. Effect of the β-lactamase inhibi-
tor Vaborbactam combined with Meropenem against serine carbapenemase-producing
Enterbacteriaceae. Antimicrob Agents Chemother. 2016;60(9):5454–8.
25. Sykes RB, Cimarusti CM, Bonner DP, Bush K, Floyd DM, Georgopapadakou NH, Koster
WM, Liu WC, Parker WL, Principe PA, Rathnum ML, Slusarchyk WA, Trejo WH, Wells
JS. Monocyclic beta-lactam antibiotics produced by bacteria. Nature. 1981;291(5815):489–91.
26. Sykes RB, Bonner DP, Bush K, Georgopapadakou NH. Azthreonam (SQ26,776), a synthetic
monobactam specifically active against aerobic gram-negative bacteria. Antimicrob Agents
Chemother. 1982;21(1):85–92.
27. Waksman SA, Reilly HC, Johnston DB. Isolation of streptomycin-producing strains of
Streptomyces griseus. J Bacteriol. 1946;52(3):393–7.
28. Umezawa H. Kanamycin: its discovery. Ann N Y Acad Sci. 1958;76(2):20–6.
17 Natural Products in Antibiotic Discovery 559
29. Weinstein MJ, Luedemann GM, Oden EM, Wagman GH, Rosselet JA, Coniglio CT, Charney
W, Herzog HL, Black J. Gentamicin, a new antibiotic complex from micromonospora. J Med
Chem. 1963;6(4):463.
30. Garneau-Tsodikova S, Labby KJ. Mechanisms of resistance to aminoglycoside antibiotics:
overview and perspectives. Medchemcomm. 2016;7(1):11–27.
31. Katz L, Ashley GW. Translation and protein synthesis: macrolides. Chem Rev.
2005;105(2):499–528.
32. Gomes C, Martinez-Puchol S, Palma N, Horna G, Ruiz-Roldán L, Pons MJ, Ruiz J. Macrolide
resistance mechanisms in Enterobacteriaceae: focus on azithromycin. Crit Rev Microbiol.
2017;43(1):1–30.
33. Wright PM, Seiple IB, Myers AG. The evolving role of chemical synthesis in antibacterial
drug discovery. Angew Chem Int Ed Engl. 2014;53(34):8840–69.
34. Yang W, Moore IF, Koteva KP, Bareich DC, Hughes DW, Wright GD. TetX is a flavin-
dependent monooxygenase conferring resistance to tetracycline antibiotics. J Biol Chem.
2004;279(50):52346–52.
35. Floss HG, Yu TW. Rifamycin-mode of action, resistance, and biosynthesis. Chem Rev.
2005;105(2):621–32.
36. Goldstein BP. Resistance to rifampicin: a review. J Antibiot (Tokyo). 2014;67(9):625–30.
37. Bugg TD, Wright GD, Dutka-Malen, S, Arthur, M, Courvalin, P, Walsh CT. Molecular basis
for vancomycin resistance in Enterococcus faecium BM4147: biosynthesis of a depsipeptide
peptidoglycan precursor by vancomycin resistance proteins VanH and VanA. Biochem. 1991;
30(43):10408–10415.
38. Kirst HA, Thompson DG, Nicas TI. Historical yearly usage of vancomycin. Antimicrob
Agents Chemother. 1998;42(5):1303–4.
39. Kahne D, Leimkuhler C, Lu W, Walsh C. Glycopeptide and lipoglycopeptide antibiotics.
Chem Rev. 2005;105(2):425–48.
40. Mukhtar TA, Wright GD. Streptogramins, oxazolidnones, and other inhibitors of bacgerial
protein synthesis. Chem Rev. 2005;105(2):529–42.
41. Wilson DN. The A-Z of bacterial translation inhibitors. Crit Rev Biochem Mol Biol.
2009;44(6):393–433.
42. Allington DR, Rivey MP. Quinupristin/dalfopristin: a therapeutic review. Clin Ther.
2001;23(1):24–44.
43. Eisenstein BI, Oleson FB Jr, Baltz RH. Daptomycin: from the mountain to the clinic, with
essential help from Francis Tally, MD. Clin Infect Dis. 2010;50(1):S10–5.
44. Taylor SD, Palmer M. The action mechanism of daptomycin. Bioorg Med Chem.
2016;24(24):6253–68.
45. Ordooei Javan A, Shokouhi S, Sahraei Z. A review on colisten nephrotoxicity. Eur J Clin
Pharmacol. 2015;71(7):801–10.
46. Velkov T, Thompson PE, Nation RL, Li J. Structure-activity relationships of polymyxin anti-
biotics. J Med Chem. 2010;53(5):1898–916.
47. Yu Z, Quin W, Lin J, Fang S, Qiu J. Antibacterial mechanisms of polymyxin and bacterial
resistance. Biomed Res Int. 2015;2015:679109.
48. Srinivas P, Rivard K. Polymyxin resistance in gram-negative pathogens. Curr Infect Dis Rep.
2017;19(11):38.
49. Miller WR, Bayer AS, Arias CA. Mechanism of action and resistance to daptomycin in
Staphylococcus aureus and Enterocci. Cold Spring Harb Perspect Med. 2016;6(11):a026997.
50. Novak R. Are pleuromutilin antibiotics finally fit for human use? Ann N Y Acad Sci.
2011;1241:71–81.
51. Dinos GP, Athanassopoulos CM, Missiri DA, Giannopoulou PC, Vlachogiannis IA,
Papadopoulos GE, Papaioannou D, Kalpaxis DL. Chloramphenicol derivatives as antibac-
terial and anticancer agents: historic problems and current solutions. Antibiotics (Basel).
2016;5(2):20.
560 F. R. McSorley et al.
52. Silver LL. Fosfomycin: mechanism and resistance. Cold Spring Harb Perpect Med.
2017;7(2):a025262.
53. Fidaxomicin. A novel agent for the treatment of Clostridium difficile infection. Can J Infect
Dis Med Microbiol. 2015;26(6):305–12.
54. Payne DJ, Gwynn MN, Holmes DJ, Pompliano DL. Drugs for bad bugs: confronting the chal-
lenges of antibacterial discovery. Nat Rev Drug Discov. 2007;6(1):29–40.
55. Tommasi R, Brown DG, Walkup GK, Manchester JI, Miller AA. ESKAPEing the labyrinth of
antibacterial discovery. Nat Rev Drug Discov. 2015;14(8):529–42.
56. Lipinski CA, Lombardo F, Dominy BW, Feeney PJ. Experimental and computational
approaches to estimate solubility and permeability in drug discovery and development set-
tings. Adv Drug Deliv Rec. 2001;46(1–3):3–26.
57. O'Shea R, Moser HE. Physicochemical properties of antibacterial compounds: implications
for drug discovery. J Med Chem. 2008;51(10):2871–8.
58. Mugumbate G, Overington JP. The relationship between target-class and the physiochemical
properties of antibacterial drugs. Bioorg Med Chem. 2015;23(16):5218–24.
59. Brown DG, Tl M-D, Gagnon MM, Tommasi R. Trends and exceptions of physical proper-
ties on antibacterial activity for Gram-positive and Gram-negative pathogens. J Med Chem.
2014;57(23):10144–61.
60. Ganesan A. The impacet of natural product upon modern drug discovery. Curr Opin Chem
Biol. 2008;12(3):306–17.
61. Harvey AL, Edrada-Ebel R, Quinn RJ. The re-emergene of natural products for drug discovery
in the genomics era. Nat Rev Drug Discov. 2015;14(2):111–29.
62. Lovering F, Bikker J, Humblet C. Escape from flatland: increasing saturation as an approach
to improving clinial success. J Med Chem. 2009;52(21):6752–6.
63. Lovering F. Escape from Flatland 2: complexity and promiscuity. Med Chem Commun.
2013;4(3):515–9.
64. Buggs CW. Ten years after streptomycin; past and current practice in antibiotic therapy. J Natl
Med Assoc. 1957;49(3):142–9.
65. Baltz RH. Antimicrobials from actinomycetes: back to the future. Microbe. 2007;2(3):125.
66. Katz L, Baltz RH. Natural product discovery: past, present, and future. J Ind Microbiol
Biotechnol. 2016;43(2–3):155–76.
67. Chang B-C, Wang S-J. The impact of patent eligibility on biotech patents: a flow chart for
determining patent eligibility and an immune therapy case study. Hum Vaccin Immunother.
2015;11(3):789–94.
68. Challis GL, Hopwood DA. Synergy and contingency as driving forces for the evolution of
multiple secondary metabolite production by Streptomyces species. Proc Natl Acad Sci U S A.
2003;100(2):14555–61.
69. Wright GD. Antibiotic adjuvants: rescuing antibiotics from resistance. Trends Microbiol.
2016;24(11):862–71.
70. Melander RJ, Melander C. The challenge of overcoming antibiotic resistance: an adjuvant
approach? ACS Infect Dis. 2017;3(8):559–63.
71. Cox G, Siron A, King AM, De Pascale G, Pawlowski AC, Koteva K, Wright GD. Cell Chem
Biol. 2017;24(1):98–109.
72. King AM, Reid-Yu SA, Wang W, King DT, De Pascale G, Strynadka NC, Walsh TR, Coombes
BK, Wright GD. Aspergillomarasmine A overcomes metallo-β-lacamse antibiotic resistance.
Nature. 2014;510(7506):503–6.
73. Stermitz FR, Lorenz P, Tawara JN, Zenewicz LA, Lewis K. Synergy in a medicinal plant:
antimicrobial action of berberine potentiated by 5′-methoxyhydnocarpin, a multidrug pump
inhibitor. Proc Natl Acad Sci U S A. 2000;97(4):1433–7.
74. Malik M, Li L, Zhao X, Kerns RJ, Berger JM, Drlica K. Lethal synergy involving bicyclomy-
cin: an approach for reviving old antibiotics. J Antimicrob Chemother. 2014;69(12):3227–35.
75. Ling LL, Schneider T, Peoples AJ, Spoering AL, Engels I, Conlon BP, Mueller A, Schäberle
TF, Hughes DE, Epstein S, Jones M, Lazarides L, Steadman VA, Cohen DR, Felix CR,
17 Natural Products in Antibiotic Discovery 561
Fetterman KA, Millet WP, Nitti AG, Zullo AM, Chen X, Lewis K. A new antibiotic kills
pathogens without detectable resistance. Nature. 2015;517(7535):455–9.
76. Charlop-Powers Z, Owen JG, Reddy BV, Ternei MA, Guimarães DO, de Frias UA, Pupo MT,
Seepe P, Feng Z, Brady SF. Global biogeographical sampling of bacterial secondary metabo-
lism. elife. 2015;4:e05048.
77. Masschelein J, Jenner M, Challis GL. Antibiotics from Gram-negative bacteria: a comprehen-
sive overview and selected biosynthetic highlights. Nat Prod Rep. 2017;34(7):712–83.
78. Hermann J, Fayad AA, Müller R. Natural products from myxobacteria: novel metabolites and
bioactivities. Nat Prod Rep. 2017;34(2):135–60.
79. Lok C. Mining the microbial dark matter. Nature. 2015;522(7556):270–3.
80. Nichols D, Cahoon N, Trakhtenber EM, Pham L, Mehta A, Belanger A, Kanigan T, Lewis K,
Epstein SS. Use of ichip for high-throughput in situ cultivation of “uncultivable” microbial
species. Appl Environ Microbiol. 2010;76(8):2445–50.
81. Blin K, Wolf T, Chevrette MG, Lu X, Schwalen CJ, Kautsar SA, Suarez Duran HG, de Los
Santos ELC, Kim H, Nave M, Dickschat JS, Mitchell DA, Shelest E, Breitling R, Takano E,
Sy L, Webe T, Medema MH. antiSMASH 4.0-improvements in chemistry predictions and
gene cluster boundary identification. Nucleic Acids Res. 2017;45(W1):W36–41. https://doi.
org/10.1093/nar/gkx319.
82. Ibrahim A, Yang L, Johnston C, Liu X, Ma B, Magarvey NA. Dereplicating nonribosomal
peptides using an informatic search algorithm for natural products (iSNAP) discovery. Proc
Natl Acad Sci U S A. 2012;109(47):19196–201.
83. Johnston CW, Skinnider MA, Wyatt MA, Li X, Ranieri MR, Yang L, Zecehl DL, Ma B,
Magarvey NA. An automated Gencomes-to-Natural Products platform (GNP) for the discov-
ery of modular natural products. Nat Commun. 2015;8:8421.
84. Medema MH, Osbourn A. Computational genomic identification and functional reconstitution
of plant natural product biosynthetic pathways. Nat Prod Rep. 2016;33(8):951–62.
85. Skinnifer MA, Merwin NJ, Johnston CW, Magarvey NA. PRISM 3: expanded predic-
tion of natural product chemical structures from microbial genomes. Nucleic Acids Res.
2017;45(W1):W49–54.
86. Johnston CW, Skinnider MA, Ca D, Rees PN, Chen GM, Walker CG, French S, Brown ED,
Bérdy J, Liu DY, Magarvey NA. Assembly and clustering of natural antibiotics guides target
identification. Nat Chem Biol. 2016;12(4):233–9.
87. Gillespie DE, Brady SF, Bettermann AD, Cianciotto NP, Liles MR, Rondon MR, Clardy J,
Goodman RM, Handelsman J. Isolation of antibiotics turbomycin A and B from a metage-
nomic library of soil microbial DNA. Appl Environ Microbiol. 2002;68(9):4301–6.
88. Tanaka Y, Tokuyama S, Ochi K. Activation of secondary metabolite-biosynthetic gene
clusters by generating rsmG mutations in Streptomyces griseus. J Antibiot (Tokyo).
2009;62(12):669–73.
89. Tanaka Y, Kasahara K, Hirose Y, Murakami K, Kugimiya R, Ochi K. Activation and products
of the cryptic secondary metabolite biosynthetic gene clusters by rifampin resistance (rpoB)
mutations in actinomycetes. J Bacteriol. 2012;195(10):2959–70.
90. Yoon V, Nodwell JR. Activating secondary metabolism with stress and chemicals. J Ind
Microbiol Biotechnol. 2014;41(2):415–24.
91. Onaka H. Novel antibiotic screening methods to awaken silent or cryptic secondary metabolic
pathways in actinomycetes. J Antibiot (Tokyo). 2017;70(8):865–70.
92. Yamanaka K, Reynolds KA, Kersten RD, Tyan KS, Gonzalez DJ, Nizet V, Dorrestein PC,
Moore BS. Direct cloning and refactoring of a silent lipopeptide biosynthetic gene cluster
yields the antibiotic taromycin A. Proc Natl Acad Sci U S A. 2014;111(5):1957–62.
93. Li Y, Li Z, Yamanaka K, Xy Y, Zhang W, Wlamakis H, Kolter R, Moore BS, Qian PY. Directed
natural product biosynthesis gene cluster capture and expression in the model bacterium
Bacilus subtilis. Sci Rep. 2015;5(9383).
94. Banik JJ, Brady SF. Cloning and characterization of new glycopeptide gene clusters found in
environmental DNA megalibrary. Proc Natl Acad Sci U S A. 2008;105(45):17273–7.
562 F. R. McSorley et al.
95. Cohen LJ, Han S, Huang YH, Brady SF. Identification of the colicin V bacteriocin gene cluster
by functional screening of a human microbiome metagenomic library. ACS Infec Dis; 2017.
4(1):27–32.
96. Wright GD. Perspective: synthetic biology revives antibiotics. Nature. 2014;509(7498):S13.
97. Thaker MN, Wright GD. Opportunities for synthetic biology in antibiotics:expanding glyco-
peptide diversity. ACS Synth Biol. 2015;4(3):195–206.
98. Braff D, Shis D, Collins JJ. Synthetic biology platform technologies for antimicrobial applica-
tions. Adv Drug Deliv Rev. 2016;105(Pt A):35–43.
99. Yim G, Wang W, Thaker MN, Tan S, Wriight GD. How to make a glycopeptide: asynthetic
biology approach to expand glycopeptide antibiotic chemcial diversity. ACS Infect. Dis.
2016;2(9):642–50.
Chapter 18
The New Versus Old Target Debate
for Drug Discovery
Alice L. Erwin
If I had been invited to write on this topic a couple of decades ago, I would not have
understood what there was to debate. By mid-2000, complete genome sequences
had been published for a dozen or so bacterial pathogens, providing a wealth of
potential targets. For those of us whose careers in antibiotic discovery started just
before the turn of the century, target evaluation seemed extremely simple. At that
time it seemed obvious to me and to my coworkers that new targets were better than
old. We argued that for drugs with new mechanisms there would be no pre-existing
resistance. A second argument was that while the empiric methods of the past had
identified only a small number of antibiotic classes and even fewer targets, new
technologies would allow us to cast our net much more widely and be much more
productive.
At that time, I saw the need for new antibiotics to replace drugs for which resis-
tance had become common. I had no idea of the limitations of existing antibiotics
other than resistance. Moreover, it did not occur to me to wonder whether inhibitors
of the new targets (mostly enzymes) would be as effective as existing antibiotics
(most of which target the machinery of macromolecular synthesis).
Today, I consider that one of the most important advantages of new targets is the
possibility of finding drugs that are not only new but in some way better than current
antibiotics. Features that might be considered desirable for new anti-infective drugs
include reduced likelihood of resistance emergence, improved activity against per-
sistent infections, or better safety, including less disruption of normal flora.
The sections below will present my view of the advantages and risks of old and
new targets. I will illustrate my discussion with examples of antibacterial drugs that
A. L. Erwin (*)
Erwin Consulting, Seattle, WA, USA
were recently approved or are currently in clinical trials, as well as some interesting
new compounds with efficacy in animals. These can be considered in three groups:
• Inhibitors of three good old targets, including both new members of old chemical
classes and first-in-class antibiotics with similar mechanisms to old drugs.
Contrary to the idea that old targets were not amenable to modern methods,
medicinal chemistry was guided by structural and biochemical data in addition
to antibacterial assays for many of these.
• New antibiotics with other, well-defined targets, for which knowledge of the
target was used in evaluating analogs. This list is short, as target-directed pro-
grams have not yet been very successful.
• Antibacterial drugs with new, often complex mechanisms. Some of these were
discovered empirically, with the mechanism of action determined later (if at all).
Others are not antibiotics per se but increase the susceptibility of bacteria to host
defenses or to other drugs. Some were designed to address aspects of infection
not well handled by existing antibiotics.
This review is not intended to provide a complete list of antibacterial drugs in the
clinical development pipeline. The reader is referred to recent reviews [12], to the
NIH database https://clinicaltrials.gov/, and to the pipeline web pages maintained
by the Pew Charitable Trusts Antibiotic Resistance Project (http://www.pewtrusts
.org/en/multimedia/data-visualizations/2014/antibiotics-currently-in-clinical-
development; http://www.pewtrusts.org/en/multimedia/data-visualizations/2017/
nontraditional-products-for-bacterial-infections-in-clinical-development).
The largest class of synthetic antibiotics is the fluoroquinolones, which inhibit two
targets, the GyrA subunit of DNA gyrase and the ParC subunit of topoisomerase
IV. Many fluoroquinolones have been approved, the most recent being Baxdela
(delafloxacin) from Melinta Therapeutics; and there are others currently in various
stages of clinical development. Moreover, clinical candidates have been identified
for several new scaffolds of Type II topoisomerase inhibitors. Gepotidacin
(GlaxoSmithKline) and Zoliflodacin (Entasis Therapeutics), both in phase 2, both
inhibit gyrase A with mechanisms different from fluoroquinolones [12].
The lethality of fluoroquinolones is the result of the fact that they not only inhibit
the ATPase activity of gyrase but also stabilize the covalent enzyme–DNA complex.
Balanced inhibition of both gyrase and topoisomerase IV is important for the char-
acteristically low frequency of spontaneous mutation to high-level fluoroquinolone
resistance. Novobiocin has a much higher frequency of resistance, and this has been
attributed to its disproportionate inhibition of gyrase (GyrB) vs topoisomerase IV
(ParE). This insight inspired programs at both Vertex Pharmaceuticals and Trius
Therapeutics aimed at discovery of compounds with balanced inhibition of GyrB
and ParE. Both groups employed structure-based drug design and successfully iden-
tified dual-active compounds with efficacy in animal models that had the predicted
low frequency of resistance. Tricyclic GyrB/ParE (TriBE) inhibitors active against
both Gram-negative and Gram-positive bacteria were discovered by Trius before
their acquisition by Cubist and then by Merck. Vertex gyrase inhibitors are primar-
ily active against Gram-positive pathogens; their technology was recently licensed
by Spero Therapeutics. Using an antisense whole-cell screen, Merck scientists iden-
tified a natural product, kibdelomycin, that is also a dual inhibitor of GyrB and
ParE. Kibdelomycin is active primarily against Gram-positive bacteria and showed
efficacy in a hamster model of Clostridium difficile infection [7, 32, 37, 62, 97].
As noted, there are multiple classes of antibiotics targeting each of the mechanisms
described above. Each has been shown to be amenable to rational drug design as
well as to the empiric methods that led to the first members of each scaffold.
Fluoroquinolones were probably the first class of antibiotics for which data on
potency toward the target enzyme as well as antibacterial activity were used to
inform design of new analogs [23]. Structural information on the interaction of com-
pounds with their targets was critical for discovery of the dual GyrB/ParE inhibitors
and was likely to have been important for the new classes of β-lactamase inhibitors
also. Ribosomal crystal structures now allow understanding of binding specificity of
new inhibitors but have not yet led to completely novel scaffolds [111].
568 A. L. Erwin
As the impetus for discovery of new antibiotics has usually been the emergence of
resistance to existing antibiotics, it may seem paradoxical to describe drugs with
these old targets as having a low frequency of resistance. The key distinction is
between endogenous resistance, resulting from spontaneous mutations that produce
a substantial loss in susceptibility to an antibiotic, and exogenous resistance, result-
ing from acquisition of genes encoding resistance determinants. Mutations that con-
fer high-level resistance in a single step can lead to rapid selection for resistance,
sometimes within a single course of treatment. Drugs for which such mutations are
frequent cannot be used as single antibiotics because treatment failure is common.
This phenomenon was recognized in the 1950s, leading to the decision that strepto-
mycin and rifampin could safely be used in combination (typically in tuberculosis)
but neither could be used as monotherapy. In an influential paper, Lynn Silver
described the concept that the reason monotherapy is possible for many of the com-
mon broad-spectrum antibiotics is that they effectively have multiple targets [87].
For the classes of antibiotics described above, nearly all resistance is exogenous.
Acquired mechanisms of resistance include aminoglycoside-modifying enzymes,
β-lactamases, efflux pumps specific for macrolides or tetracyclines, rRNA methyl-
transferases, alternative PBPs, and other target-modifying enzymes. These resis-
tance determinants nearly always have origins in bacterial populations other than
the pathogens being treated by the antibiotic. The existence of transmissible resis-
tance does not prevent a new antibiotic from being useful, sometimes for many
decades after the first appearance of a resistance determinant. Understanding the
molecular basis of acquired resistance to an existing drug has been extremely valu-
able, allowing the discovery of new drugs with the same or similar mechanism of
action but which avoid resistance.
The success of antibiotics with these old targets has given us confidence that we
know how antibiotics work, how to use them, and how to find new ones. Antibiotics
are probably the most successful and best understood type of drug. Broadly speak-
ing, all antibiotics have the same, extremely simple mechanism: stop growth of the
infecting organism and let the patient recover. Bacterial infections can be consid-
ered as either the invasion of the patient by a foreign pathogen or the intrusion of a
commensal organism into a normally sterile site. In either scenario, the vast major-
ity of such invasions are easily controlled by the patient’s inflammatory system
without any problem. If the bacteria somehow get away from the normal host
defenses, the result is symptomatic disease. A drug that slows growth of the invader
will often allow the patient to regain the upper hand and clear the infection. Although
it might seem desirable for antibiotics to kill bacteria outright, in actuality many
successful antibiotics are simply bacteriostatic.
18 The New Versus Old Target Debate for Drug Discovery 569
Study of the antibiotics discovered during the twentieth century led to standard
in vitro methods for determining the antimicrobial susceptibility of bacteria isolated
from patients and to understanding how antibacterial activity and pharmacokinetics
of an antibiotic contribute to its efficacy in an animal model. The developing field
of PK/PD not only allows accurate prediction of effective dosing in animal studies
but allows extension of those data to human trials.
For most existing antibiotics, we have a fairly good understanding of mechanism
of action, usually inhibition of an essential bacterial process. Collectively, the study
of existing antibiotics has given us the feeling that a chemical compound will pre-
vent bacterial growth if it is able to enter bacterial cells and inhibit an essential
bacterial process with sufficient potency. If in vitro antibacterial activity is suffi-
cient, the compound will probably have efficacy in mouse models of infection as
long as it is not toxic and has appropriate pharmacokinetics to provide sufficient
exposure to bacteria at the site of infection.
These features should make antibiotics far easier to discover and develop than
drugs for other therapeutic areas. Consider drugs for hypertension, cardiac disease,
Alzheimer’s, rheumatoid arthritis, etc. Discovery of a new drug often means simul-
taneously developing new understanding of the molecular basis of the disease and
finding compounds that affect that aspect of human biology while leaving the rest of
human physiology intact. For many human diseases, animal models are not very
useful for testing efficacy of new drugs. If a new drug is approved, it is necessary to
teach physicians which patients will benefit and how to use the new drug. In con-
trast, the principles of antibiotics are well understood by scientists, by physicians,
and by regulatory agencies. We know how to evaluate candidate drugs in the lab and
in patients, and physicians understand how to use the new drug if it is approved.
At the turn of the century, it was recognized that the so-called Golden Age of
Antibiotics had come to an end. Although new members of existing antibiotic
classes were being developed, the only new class that had been discovered for
decades was the oxazolidinones. Discovery of promising natural products had
stalled. The search for new targets was driven by a feeling that as antibiotic resis-
tance was continuing to increase, new experimental approaches would be required.
The explosion of target-directed projects in the early twenty-first century, par-
ticularly the widespread use of in vitro enzyme assays for high-throughput screen-
ing of synthetic libraries, is often described as the “genome era” of antibiotic
research. In addition to the bacterial genome sequences that were appearing,
industrial research programs at this time made use of other components of target-
directed drug discovery that were already in place in biotechnology and pharmaceu-
tical companies, having been developed for other therapeutic areas. There was
increasing interest in enzymes as targets, as it would be possible to screen large
chemical libraries with in vitro assays using robotics. If the enzyme could be crys-
570 A. L. Erwin
tallized, structural data would be available to guide chemistry. Rational drug design
seemed feasible and more attractive than the empiric methods by which most anti-
biotic classes had previously been developed.
The critical value of the genome sequences available by 2000 or so was not
the thousands of previously unknown genes but the incredible level of detail on
the genes for which functions were known or could be proposed. A gene previ-
ously known only in E. coli or Bacillus subtilis could be amplified from P. aeru-
ginosa or S. aureus in order to produce recombinant protein. Transposon
mutagenesis and other molecular biology techniques allowed genome-wide
assessment of genes required for in vitro growth or for full virulence in animals.
Research programs focused on targeting enzymes for which in vitro assays could
easily be developed would compile a list of 100–150 potential targets that were
considered to be essential in vitro, absent from mammalian cells, and conserved
across the pathogens of interest. A complementary approach used antisense tech-
nology to downregulate individual genes, generating strains that were hypersen-
sitive to inhibitors of the target enzyme (or pathway). Such strains could be used
in initial screening or in evaluating compounds that were known to be potent
inhibitors of the target [24, 59, 63].
The use of new, genome-wide studies together with robotics and combinational
chemistry seemed very exciting at the time but did not lead to the rapid discovery of
new antibiotics, as had been hoped. Some of the limitations of these programs will
be discussed in a later section.
18.5.1 rotein Synthesis: Targets Other than the 30S and 50S
P
Ribosomal Subunits
Several inhibitors of protein synthesis with mechanisms different from those dis-
cussed above have reached clinical trials.
18 The New Versus Old Target Debate for Drug Discovery 571
Peptide deformylase (PDF) PDF was a very popular target because it was thought
there is no equivalent mammalian enzyme and that initiation of bacterial proteins
with formyl-methionine was a universal feature of bacterial proteins. An antibacte-
rial natural product, actinonin, was found to inhibit PDF, providing chemical valida-
tion. However, actinonin-resistant mutants of S. aureus were easily isolated and
found to have inactivated fmt, the gene encoding formyl-methionine transferase.
Finding that fmt-null mutants can be viable suggested that PDF inhibitors would be
effective only for species in which fmt is essential. These include Streptococcus
pneumoniae and Haemophilus influenzae. There is also some evidence that fmt-null
mutants of S. aureus are reduced in virulence. PDF is a metalloenzyme, so it was
expected that finding potent inhibitors would be straightforward. Indeed, at least
three PDF inhibitors reached clinical trials [15, 48, 110]. Two of these, BB-83698
and LBM-415, were discontinued after phase 1. GSK1322322 entered phase 2 and
has since been discontinued.
FabI The fatty acid biosynthetic enzyme FabI (enoyl-ACP reductase) has been
targeted in several programs. This enzyme had been expected to be conserved across
bacterial species. However, the availability of genome sequences made it apparent
that some species contain alternative enzymes, designated FabK, FabL, or
FabV. These catalyze the same reaction as FabI but are so different in structure that
inhibition of all by a single compound appears impossible. Moreover, some patho-
gens are able to use exogenous fatty acids during infection and are therefore
expected to be intrinsically resistant to inhibitors of endogenous fatty acid synthesis
[108]. Two FabI inhibitors are currently in clinical trials, CG400549 from
CrystalGenomics and Debio 1450 from Debiopharm [12].
572 A. L. Erwin
18.5.3 Antifolates
The antifolates are the oldest synthetic antibiotics still in use today. Sulfonamides
were discovered empirically in the 1930s by Gerhard Domagk, with antibacterial
activity determinations used to drive medicinal chemistry of compounds derived
from azo dyes. After introduction of the first sulfa drugs, it was realized that they
mimic para-aminobenzoic acid (PABA), thereby inhibiting dihydropterate synthase
(DHPS) [92]. Trimethoprim was discovered more rationally. As described by van
Miert, George Hitchings hypothesized that analogs of purine and pyrimidine bases
might serve as antimetabolite drugs. He discovered several 2,4-diaminopyrimidine
inhibitors of the folate pathway that differed in specificity, including trimethoprim
for bacteria, methotrexate for cancer, and pyrimethamine for protozoal infections
(e.g., malaria); these are all inhibitors of dihydrofolate reductase (DHFR) [102].
The DHFR inhibitor iclaprim was designed to address trimethoprim resistance.
Development was stalled in 2008, but it is again in phase 3 clinical trials, sponsored
by Motif Bio [12]. Both DHPS and DHFR are being pursued in academic research
labs, but no clinical candidates have yet been identified [28, 35].
Discovered in the 1950s, rifamycins are the major antibiotic class targeting RNA
polymerase. Fidaxomicin, another inhibitor of RNA polymerase, is a natural prod-
uct described in 1975 that was approved in 2011 for C. difficile infection. A limita-
tion of both these classes is their very high frequency of resistance due to mutations
18 The New Versus Old Target Debate for Drug Discovery 573
in rpoB. Rifamycins are used primarily in combination with other antibiotics and
occasionally as monotherapy in situations where the number of bacteria to be treated
is very low, such as for prophylaxis. A recent report described the discovery of pseu-
douridimycin by screening a library of natural product extracts for inhibitors of
RNA polymerase. Pseudouridimycin is active against several Gram-positive species
and fastidious Gram-negative bacteria and was reported to be efficacious in a mouse
model of Streptococcus pyogenes infection. Of note, the frequency of resistance to
pseudouridimycin was reported to be tenfold lower than that of rifampin, apparently
because it has a different binding site that is less tolerant of mutations [47, 52, 78].
One of the risks of any new target is a high likelihood of failure, but the same can be
said of any drug discovery effort. Even with well-established antibiotic classes, it is
rare for a new clinical candidate to emerge, and the time between discovery and
approval is many years. It is thus not surprising that few of the antibiotics approved
in the past 5 years came from projects that were initiated after 2000 or so. However,
it is surprising that there are so few scaffolds in the pipeline from the target-directed
projects started in the genome era.
An influential review from scientists at GlaxoSmithKline described the dismal
failure of 67 high-throughput screens of bacterial targets, with very few producing
hits that were worth exploratory chemistry. A more recent review from the antibac-
terial program at AstraZeneca was somewhat more positive, in that they had identi-
fied hit scaffolds for 57 of the 65 targets they screened. Lead series were identified
for 19 targets. It seems likely that differences in the synthetic libraries at the two
pharmaceutical companies accounted for the differing hit rates. Despite initial
promising results, the AZ team dropped nearly all their targets when they were
unable to find compounds with broad-spectrum antibacterial activity. A contributing
factor was that at the time, Gram-negative activity was seen as essential. If they had
been primarily interested in drugs for staphylococcal and enterococcal infections,
they might have considered a smaller proportion of projects to be failures. Both
reviews emphasized, as others have, that pursuing such high numbers of targets
might have contributed to failure of the overall programs [74, 99].
The prevailing culture held that all genetically validated targets were equally
good and that screening would identify those that were worth pursuing. There was
high pressure to run as many screens as possible, leading to an emphasis on targets
for which assays would be easy to develop. In practice, these were nearly all
enzymes. Knowledge of the biochemical reaction and structure of the active site
aided prioritization of “druggable” targets. Despite the idea of pursuing novel tar-
gets identified in genome sequences, most work was on genes whose function had
already been identified in the previous decades of academic research. Protein–pro-
tein interactions were generally not considered druggable.
574 A. L. Erwin
Hit evaluation and early chemistry focused on improving affinity for the tar-
get, sometimes without evidence that any antibacterial activity observed was
mechanism-based. The mantra “fail early and often” meant that one project after
another was terminated without a chance to learn how to improve the chances of
success.
A common experience is that biochemical and structural data could allow
the chemists to improve potency of a scaffold substantially but without a cor-
responding improvement in antibacterial activity. This is in large part because
we lack information on how to improve the accumulation of compounds within
bacterial cells – an important subject beyond the scope of this review. It is
often suggested that chemical libraries in most companies are not well suited
for antibacterial discovery and that we would be more successful with different
starting points.
Because it is difficult to confer antibacterial activity on chemical scaffolds that
lack it, some programs have run target-focused screens that use bacterial cells. A
number of approaches using reporter genes, differential growth, morphology, or
labeling by a dye or radioisotope have been described [63, 90]. Even in programs
that placed a strong emphasis on cell-based assays rather than in vitro enzyme
assays, the overall success rate has been low, considered in terms of drugs entering
clinical trials.
For some targets, we have some information about why development of a spe-
cific compound was halted. Inhibitors of two metalloenzymes (PDF and LpxC)
have failed to progress after phase 1, and we can guess that these compounds had
toxicity that might have been mechanism-based. Whether toxicity will keep other
LpxC inhibitors from reaching the clinic is difficult to determine. The rapid selec-
tion of mutants resistant to the LeuRS inhibitor GSK2251052 is consistent with our
general understanding of antibiotics with single targets. However, preclinical data
suggest that resistance to the MetRS inhibitor CRS3123 may be less of an issue
[17]. Perhaps surprisingly, the frequency of resistance to LpxC inhibitors is so far
very low [27].
For projects that never reach the clinic, there is rarely any public information
about the status of any project or the reason it was terminated. It is therefore
impossible to determine why any target has not yielded new drugs, or even to say
conclusively that it has failed. A scaffold may be active vs a few isolates of S.
aureus or E. coli and perhaps show efficacy in a mouse model but never reach the
desired antibacterial spectrum; or it may have poor pharmacokinetics or show
toxicity in preclinical studies. Most of these liabilities are characteristics of the
compound or the chemical scaffold but shed no light on the validity of the
molecular target. Often resources are diverted from one project to another for
reasons that are strategic rather than scientific. Should we conclude in such a
case that the target has now been validated, not only genetically but chemically
and pharmacologically? If no clinical candidate appears, should we conclude the
target is a failure?
18 The New Versus Old Target Debate for Drug Discovery 575
Today there is increasing recognition that there are two kinds of unmet need for
bacterial infections.
Traditional antibiotics We need new traditional antibiotics to replace those for
which resistance has become common. The discovery programs described in the
previous sections are intended to meet this need. As has been discussed in numerous
conferences and white papers, new antibiotics are not emerging at the rate that is
required. However, even if we had a full pipeline of broad-spectrum antibiotics for
Gram-positive and Gram-negative infections, there would still be a need for other
types of drug.
One issue is that the desired target-product profiles change over time. At the turn
of the century, a common criticism of LpxC projects was that the resulting drugs
would not be useful for treating MRSA. It was difficult to convince senior manage-
ment that Gram-negative resistance was on the rise. Infectious disease physicians
generally did see value in a new antibiotic that would be limited to Gram-negative
bacteria – as long as it was active against P. aeruginosa as well as enteric bacteria.
Today we need drugs for A. baumannii as well as for P. aeruginosa. At the same
time, antibiotic resistance in Klebsiella and Enterobacter spp. and E. coli has
increased to the point where an enteric-only drug would be valuable. C. difficile
infection, long recognized as a complication of treatment with broad-spectrum anti-
biotics, became a much greater problem as antibiotic-resistant strains became prev-
alent. It has become common for new Gram-positive agents with poor PK and oral
bioavailability to be developed as C. difficile drugs. For some of these, microbiome
studies are incorporated into clinical trials.
The idea of narrow-spectrum antibiotics is becoming much more attractive. It is
thought that resistance to narrow-spectrum drugs may emerge more slowly, as the
normal flora will not provide a reservoir for resistance determinants. Further,
narrow-spectrum antibiotics are not expected to induce the dysbiosis that is associ-
ated with broad-spectrum antibiotics. The development and use of narrow-spectrum
antibiotics are far from straightforward but may become more feasible in the next
several years [9]. It is therefore reasonable to reconsider some of the targets that
were rejected during the genome era because they were not broadly conserved
across pathogens.
Second unmet need We also need drugs for prevention and treatment of bacterial
infections for which antibiotics have never been fully effective. These include recur-
rent, latent, or persistent infections, often associated with biofilms. Such infections
have plagued humans for millennia but are much more prominent today. Examples
are ventilator-associated pneumonia, infections in immunocompromised patients
with organ transplants or cancer chemotherapy, infected joint replacements or other
orthopedic devices, diabetic foot ulcers, respiratory infections in patients with cys-
tic fibrosis, and recurrent urinary tract infections.
576 A. L. Erwin
The antibiotics we have today act mainly on growing bacteria. They are probably
most effective in treating acute infections in previously healthy individuals. The
same is likely to be true for new antibiotics discovered with either the old targets or
the enzyme targets of the genome era. We need drugs that are active against non-
growing bacteria and against bacteria that have become tolerant to antibiotics.
Some aspects of infection are not affected by antibiotics at all. These include the
direct toxic effects of secreted cytotoxins, some of which are treatable with antibod-
ies, and the uncontrolled inflammatory response to lipopolysaccharides or toxic-
shock syndrome toxin. Research on sepsis has not yet produced effective drugs, in
part because animal models are not very predictive.
Overall, new antibiotics with standard kinds of targets are likely to have the same
limitations as existing antibiotics. Different targets, or different experimental
approaches, will be needed in order to discover drugs that are not only new but bet-
ter anti-infective therapeutic agents. Some of these are described in the next
section.
driver was antibacterial activity rather than potency vs target. At the time these
compounds were discovered, the target was not known (and may still be unknown).
Teixobactin A platform at NovoBiotic for seeking antibacterial compounds pro-
duced by previously uncultured microorganisms has revived hope in empiric natural
products discovery. In 2015, they described a new Gram-positive antibiotic with
efficacy in mouse models of S. aureus and pneumococcal infection. Teixobactin
inhibits cell wall biosynthesis, binding glycolipids lipid II and lipid III (precursors
of peptidoglycan and teichoic acids, respectively). This mechanism of action is
similar to that of vancomycin, which binds a different site on lipid II. Because it is
not the product of a single gene, it is difficult for spontaneous mutation to alter the
structure of lipid II. The frequency of endogenous resistance to vancomycin is thus
extremely low, and the same appears to be true for teixobactin. Resistance to vanco-
mycin became common only after a resistance cassette derived from the natural
producer was transferred into enterococci, encoding enzymes that produced an
altered lipid II. Teixobactin binds to both forms of lipid II and is thus active vs both
vancomycin-sensitive and vancomycin-resistant bacteria. Several other antibacte-
rial compounds that bind lipid II are known, including the peptides plectasin and
nisin. Apart from the vancomycin derivatives telavancin, dalbavancin, and orita-
vancin, none of these other lipid II-binding compounds is used clinically [49, 81].
Bedaquiline This review does not otherwise discuss tuberculosis drugs, but two
classes of antimycobacterial antibiotics are mentioned here as successful recent
examples of using antibacterial activity to guide medicinal chemistry, without
knowledge of the molecular target. The recently approved tuberculosis drug beda-
quiline targets ATP synthase and is bactericidal for both growing and nongrowing
mycobacteria. Bedaquiline was discovered at Janssen by screening for compounds
that inhibit growth of Mycobacterium smegmatis and then optimizing by medicinal
chemistry [1]. It is of interest because of the success of this empiric approach and
also because of the unprecedented nature of the target. No other antibacterial drug
has a well-defined direct effect on the respiratory pathway.
Phage lysins The idea of using lytic peptides derived from phage or bacteria as
antibacterial drugs is not new. A key issue is finding lysins with appropriate speci-
ficity. Many such peptides are active against only certain isolates of a bacterial spe-
cies; others, like nisin and gramicidin, are so broad in spectrum that they are
generally toxic. Two phage lysins are being developed for staphylococcal infection.
N-Rephasin (SAL200) from iNtRON is in phase 2 studies. It was derived from bac-
teriophage SAP-1 and has broad activity against staphylococci but is not active
against other genera [42]. CF-301 (previously PlySs2) from Contrafect is currently
in phase 1. CF-301 was derived from a Streptococcus suis phage and is broadly
cidal for staphylococci and streptococci, including biofilms [82].
Minor groove binders Drugs that bind the minor groove of DNA, such as the
antiparasitic drug pentamidine, have a long history. While there is some specificity
in the binding sites, these are not generally designed to affect transcription of a
particular gene (as for the antisense approach described below). Analogs are evalu-
ated by their antimicrobial activity and by selectivity for the desired target organ-
isms. Minor groove binders have the advantage that they kill rapidly and have low
frequency of resistance. However, the nonspecific nature of their activity makes
toxicity a concern. MGB-BP-3, currently in phase 1, is derived from the natural
product distamycin and is being developed for C. difficile infection. A series of bis-
indole compounds with activity in mouse models of Gram-positive infection dis-
covered at Microbiotix was recently reported to have a mechanism that involves
DNA binding [5, 72, 95].
The idea that studying host–pathogen interaction will lead to discovery of improved
therapeutic agents is intellectually attractive but has not yet been very successful
with regard to standard antibiotics like those discussed above. In contrast, nearly all
the agents discussed below resulted from research on how bacteria cause disease or
respond to therapy.
patients with P. aeruginosa HAP or VAP. Aerucin binds alginate, and Aerumab
(previously AR-101 or Panobacumab) binds LPS of serotype O11 [76].
Two antibody-based agents are being studied in patients with S. aureus bactere-
mia: monoclonal antibody 514G3 from XBiotech and an antibody-antibiotic conju-
gate (DSTA4637S, RG7861) from Genentech.
Antibiotic potentiators Given the difficulty in discovering new antibiotics, an
obvious potential alternative is to find drugs that can either restore the susceptibility
of resistant strains or extend the spectrum of Gram-positive antibiotics to Gram-
negative bacteria. Apart from the β-lactamase inhibitors discussed in a previous
section, no such drugs have reached the market. The only clinical candidate in this
category is SPR741.
SPR741(previously NAB741) is a polymyxin derivative being developed by
Spero Therapeutics as an antibiotic-potentiating agent, currently in phase 1 [101].
Because its mechanism is permeabilization of the outer membrane, it is possible
that SPR741 might also increase the susceptibility of bacteria to host defenses such
as defensins and complement.
Many academic and industrial labs have screened for compounds that increase
the activity of β-lactams for MRSA. Where mechanisms of active compounds have
been identified, they have been surprisingly diverse. A more systematic approach
has been taken by Merck scientists, seeking inhibitors of wall teichoic acid synthe-
sis [39, 66, 98, 107].
Efflux pump inhibition is another area that has been extremely frustrating,
despite advances in the assembly and function of multidrug efflux pumps and in the
interaction of pump components with their substrates [36, 51, 53, 91, 96].
Targeting pathogenesis is often suggested but rarely has been very successful. One
issue is that potential targets are identified by demonstrating the inability of mutants
to initiate infection. It is by no means certain that drugs inhibiting the production or
function of those virulence determinants would be able to reverse an established
function. In addition to these concerns, an additional argument against anti-viru-
lence drugs is that they would be narrow in spectrum. That is now seen as less of a
disadvantage than it was several years ago. It must be admitted that there has been
very little serious effort by the pharmaceutical industry to discover anti-virulence
drugs.
Antitoxin mAbs Like the surface-binding antibodies discussed in a previous sec-
tion, passively administered toxin-neutralizing antibodies have an extremely long
history of successful use. Shigamab, being developed by Taro, consists of two anti-
bodies against Shiga toxin [60]. ASN100 from Arsanis is also a mixture of two
antibodies, one active against five cytotoxins and the other active against LukGH
leukocidin [2, 21]. ASN100 is being tested as a prophylactic agent in patients colo-
582 A. L. Erwin
Some aspects of infectious disease result not from the bacteria per se but from an
overwhelming inflammatory response to bacterial components. This can persist
even if the infection is treated with appropriate antibiotics. Attempts to rescue septic
patients with interleukins or other immunomodulators have been largely unsuccess-
ful, even though some such agents appeared to be effective in mice. Two immuno-
modulatory agents are now in clinical trials.
AB103 from Atox Bio, now in phase 3, is an octapeptide that attenuates the sig-
naling through CD28 that is involved in induction of many proinflammatory cyto-
kines [77]. It is being tested in patients with necrotizing soft tissue infections.
CAL02, now in phase 2 from Combioxin, is a liposomal drug that neutralizes
bacterial toxins. It is being tested in addition to standard of care in intensive care
unit patients with severe pneumococcal pneumonia [6].
The Kranz lab, University of Illinois, engineered T-cell receptor domains to bind
staphylococcal enterotoxins B and C with high affinity and reported efficacy in rab-
bit models of necrotizing pneumonia and infective endocarditis [56, 86].
The previous section described a number of experimental approaches that have the
possibility of improving treatment of infectious disease. Several of the antibacterial
agents are bactericidal for nongrowing bacteria, making it possible that they will be
more effective against persistent infection than current antibiotics are. Many of
them were discovered by empiric methods and/or have mechanisms that would not
have been considered appropriate for target-directed drug discovery. With the
18 The New Versus Old Target Debate for Drug Discovery 583
exception of the tuberculosis drugs, these are all very early in development. It can-
not be concluded that these nonclassical approaches will always be more successful
than classical target-directed antibiotic discovery.
The target-directed programs of the genome era focused on enzymes. As dis-
cussed above, there are a number of reasons why any given project might have
failed. There is no obvious reason to think that enzymes as a group should be less
suitable as antibiotic targets than the macromolecular synthesis machines that most
existing antibiotics inhibit. Indeed, the first broad-spectrum synthetic antibiotics,
the antifolates, are enzyme inhibitors. It is puzzling that even for a pathway as well
validated as peptidoglycan synthesis, it has been impossible to find good inhibitors
of the soluble enzymes (GlmU, MurA-MurF) apart from fosfomycin. One limita-
tion of enzymes as targets is the risk that mutants with endogenous resistance occur
at high frequency [87]. This risk is not simply theoretical, as indicated by the failure
of GSK2251052 during phase 2 [70]. It is unlikely, however, that high frequency of
resistance was the characteristic that killed most of the genome-era projects before
a clinical candidate was identified.
It is more likely that after initial identification of a hit series with mechanism-
based antibacterial activity, most projects progressed for a while but failed to come
close to a solid lead within an allotted period of time. A common experience is that
chemistry can produce very potent inhibitors, as assessed in an in vitro biochemical
assay, but that potency does not translate into antibacterial activity. The standard
explanation is that these compounds “don't get in.” This is certainly true to a great
extent. Compounds active vs staphylococci and streptococci but not vs Gram-
negative bacteria are often active against an efflux-deficient or hyperpermeable
mutant of E. coli or P. aeruginosa. In that case, it is appropriate to conclude that the
intrinsic defenses of Gram-negative bacteria are a major limitation to antibacterial
activity. Poor antibacterial activity against Gram-positive bacteria is harder to
explain, although again failure to cross the cytoplasmic membrane and accumulate
against the concentration gradient probably contributes to the problem [89].
We must consider the possibility that we simply don’t know enough about bacte-
ria to choose good targets. The assumption of the new-targets programs of the
genome era was that we understand antibiotics and infectious disease well enough
to be able to do this rationally. Discovery programs at that time were often driven
not by new insight into how antibiotics work but by new technology, combined with
a feeling that no new insight was needed. I would argue that not all essential genes
are equally good targets. Moreover, research into the mechanisms by which bacteria
avoid the effects of antibiotics has suggested that drugs with apparently similar
mechanisms can have very different effects on bacterial physiology.
One potential problem is that inhibition of the target may have relatively little
immediate effect on bacterial physiology. It is very difficult to determine “how
essential” a bacterial process is. Will growth stop if catalytic activity is reduced by
25%? Will it be necessary to inhibit 99.9% of activity? Very few studies have
attempted to address this, and indeed there are no general methods. Tuberculosis
researchers have developed methods for targeted degradation of specific proteins in
order to assess the impact on growth and survival of the cell [43, 105]. Extending
584 A. L. Erwin
these methods beyond mycobacteria would be useful, although reducing the amount
of a protein within the cell may have a different effect from chemical inhibition of
that protein’s activity. For a few enzymes, mutants with partial activity can give us
some information as to the level of inhibition that is tolerated. The envA1 mutant of
E. coli has an 18-fold reduction in LpxC activity, compared to wild-type strains
[109]. Similarly, the E. coli mutant ligA251 has a point mutation that reduces activ-
ity of DNA ligase by 20- to 60-fold [46].
Systems biology may allow improved prioritization of targets, with the caveat
that we may not yet have the information needed to generate predictive models. One
example is the modeling of the lipid A synthetic pathway. A combination of compu-
tational and experimental methods led to the suggestion that LpxK would be a better
target for inhibition than LpxC if previous knowledge of the pathway’s regulation is
incorporated into the model. If regulation was not considered, then LpxC appeared
to be a better target, as suggested by previous enzymology [26].
While target-directed discovery programs tend to focus on the effect of chemical
compounds on the target, it may be more useful to focus on the effect of compounds
on the bacterial cell. The most obvious such effects are bacterial stasis or death,
morphological changes such as filamentation or spheroplasting, or changes in pro-
cesses like macromolecular synthesis that can be monitored easily. Bacterial
responses to differing antibiotic stresses may also be important. It may be useful to
look for targets that, when inhibited, induce the same stress responses as one or
more of the well-established existing classes of antibiotics. The SOS response to
DNA damage is one example. Fluoroquinolines induce the SOS response, and it is
also involved in the thymineless death induced by trimethoprim [30].
Reporters of the cell wall stress response have been used by many groups to
screen chemical libraries, finding hits with diverse targets in envelope biogenesis –
not only peptidoglycan synthesis but also LPS synthesis and lipoprotein export [19,
68, 100]. The machines involved in export of proteins and lipopolysaccharide to the
outer membrane and in maintenance of its permeability barrier could be considered
as the only truly novel bacterial pathways that have been discovered in recent
decades [58, 79, 80].
The stringent response was first studied decades ago in the context of amino acid
starvation. Much more recently it was identified as a critical issue in antibiotic toler-
ance and biofilm formation. A better understanding of persistence and tolerance
may allow us to design more effective antibiotics. One might expect all bacteriostatic
inhibitors of protein synthesis to induce similar responses in bacteria. However, a
recent study of the effect of bacteriostatic agents on induction of tolerance to
β-lactams found that the Met-tRNA synthetase inhibitor mupiricin activates RelA,
while the ribosome binders tetracycline and chloramphenicol inhibit induction of
the stringent response [45, 54, 85].
Finally, several lines of evidence suggest that the mechanisms of existing antibi-
otics are more complicated than the simple picture presented above. In 2007, the
Collins lab at Boston University proposed that the bactericidal activities of β-lactams,
aminoglycosides, and fluoroquinolones have a common mechanism involving pro-
duction of hydroxyl radicals, as a consequence of the immediate effects of these
18 The New Versus Old Target Debate for Drug Discovery 585
Old targets, benefits and risks The field has been successful in identifying new
antibiotics for the old targets, utilizing current structural biology, and making use of
new understanding of mechanisms and resistance determinants as they emerge.
Even for the old antibiotic classes, new synthetic methods are allowing increased
chemical diversity. However, the side effects and other liabilities of these antibiotic
classes must be considered in developing new members. A further problem is that
inhibitors of the old targets tend to be broad-spectrum antibiotics, with inherent
effects on normal flora. Resistance may be slow to emerge but is likely to spread
rapidly, just as for existing antibiotics. Moreover, new inhibitors of the old targets
are unlikely to be better than the old drugs at treating persistent infections or toxic
effects.
New targets, benefits and risks New targets have the possibility of yielding better
drugs – more diverse in chemical structure and mechanism, possibly narrower in
spectrum, and/or providing improved treatment of conditions not well handled by
existing antibiotics. However, focusing narrowly on targets we think we understand
has not been very successful. We need to take a variety of approaches, allowing use
of technology and incorporating new knowledge as it becomes available.
We don’t know enough about compound uptake One of the major problems in
antibiotic discovery is our inability to design compounds with good access to their
targets. Aminoglycosides and tetracyclines owe their success in part to their ability
to achieve concentrations within a bacterial cell that are higher than external con-
centrations. β-lactams and other inhibitors of late stages of peptidoglycan synthesis
do not need to cross the cytoplasmic membrane, but for Gram-negative bacteria the
outer membrane is a barrier to these drugs. Academic research has been very suc-
cessful in elucidating the structure and synthesis of the outer membrane and of the
multidrug efflux pumps that constitute the intrinsic resistance mechanisms of Gram-
586 A. L. Erwin
negative bacteria. This knowledge has not yet provided concepts or tools to guide
medicinal chemists in improving the antibacterial activity of potent compounds.
Major Points
• Most target-directed drug discovery projects focus on seeking inhibitors of indi-
vidual enzymes
• No antibiotic that has ever been approved was discovered through this approach
• Only a few target-directed projects have yielded clinical candidates
• A critical problem that is not well appreciated is that not all essential bacterial
functions are equally good drug targets
• A second important problem is that optimization of a chemical scaffold for
potency in a cell-free assay often does not improve its antibacterial activity or
drug characteristics
• The antibacterial drug development pipeline is sparse, but includes small mole-
cules and biologics with a wide variety of mechanisms
• Some of these were discovered empirically, others through rational methods
including structure-based drug design
References
tors of DNA gyrase and topoisomerase IV. ACS Infect Dis. 2015;1(1):4–41. https://doi.
org/10.1021/id500013t.
8. Blount KF, Megyola C, Plummer M, Osterman D, O'Connell T, Aristoff P, et al. Novel
riboswitch-binding flavin analog that protects mice against Clostridium difficile infection
without inhibiting cecal flora. Antimicrob Agents Chemother. 2015;59(9):5736–46. https://
doi.org/10.1128/AAC.01282-15.
9. Boucher HW, Ambrose PG, Chambers HF, Ebright RH, Jezek A, Murray BE, et al. White
paper: developing antimicrobial drugs for resistant pathogens, narrow-spectrum indications,
and unmet needs. J Infect Dis. 2017. https://doi.org/10.1093/infdis/jix211.
10. Bourne CR. Utility of the biosynthetic folate pathway for targets in antimicrobial discovery.
Antibiotics (Basel). 2014;3(1):1–28. https://doi.org/10.3390/antibiotics3010001.
11. Brötz-Oesterhelt H, Brunner NA. How many modes of action should an antibiotic have? Curr
Opin Pharmacol. 2008;8(5):564–73. https://doi.org/10.1016/j.coph.2008.06.008.
12. Bush K, Page MGP. What we may expect from novel antibacterial agents in the pipeline
with respect to resistance and pharmacodynamic principles. J Pharmacokinet Pharmacodyn.
2017;44(2):113–32. https://doi.org/10.1007/s10928-017-9506-4.
13. Casadevall A. The third age of antimicrobial therapy. Clin Infect Dis. 2006;42(10):1414–6.
https://doi.org/10.1086/503431.
14. Coleman K. Diazabicyclooctanes (DBOs): a potent new class of non-β-lactam
β-lactamase inhibitors. Curr Opin Microbiol. 2011;14(5):550–5. https://doi.org/10.1016/j.
mib.2011.07.026.
15. Corey R, Naderer OJ, O‘Riordan WD, Dumont E, Jones LS, Kurtinecz M, et al. Safety, toler-
ability, and efficacy of GSK1322322 in the treatment of acute bacterial skin and skin structure
infections. Antimicrob Agents Chemother. 2014;58(11):6518–27. https://doi.org/10.1128/
AAC.03360-14.
16. Côté J-P, French S, Gehrke SS, MacNair CR, Mangat CS, Bharat A, et al. The genome-wide
interaction network of nutrient stress genes in Escherichia coli. MBio. 2016;7(6):e01714–6.
https://doi.org/10.1128/mBio.01714-16.
17. Critchley IA, Green LS, Young CL, Bullard JM, Evans RJ, Price M, et al. Spectrum of activ-
ity and mode of action of REP3123, a new antibiotic to treat Clostridium difficile infections.
J Antimicrob Chemother. 2009;63(5):954–63. https://doi.org/10.1093/jac/dkp041.
18. Davies J, Spiegelman GB, Yim G. The world of subinhibitory antibiotic concentrations. Curr
Opin Microbiol. 2006;9(5):445–53. https://doi.org/10.1016/j.mib.2006.08.006.
19. DeCenzo M, Kuranda M, Cohen S, Babiak J, Jiang Z-D, Su D, et al. Identification of
compounds that inhibit late steps of peptidoglycan synthesis in bacteria. J Antibiot.
2002;55(3):288–95.
20. Deng Y, Sun C, Hunt DK, Fyfe C, Chen C-L, Grossman TH, et al. Heterocyclyl tetracyclines.
1. 7-Trifluoromethyl-8-pyrrolidinyltetracyclines: potent, broad spectrum antibacterial agents
with enhanced activity against Pseudomonas aeruginosa. J Med Chem. 2017;60(6):2498–
512. https://doi.org/10.1021/acs.jmedchem.6b01903.
21. Diep BA, Le VTM, Visram ZC, Rouha H, Stulik L, Dip EC, et al. Improved protection in
a rabbit model of CA-MRSA necrotizing pneumonia upon neutralization of leukocidins in
addition to alpha-hemolysin. Antimicrob Agents Chemother. 2016:AAC.01213-16. https://
doi.org/10.1128/AAC.01213-16.
22. DiGiandomenico A, Keller AE, Gao C, Rainey GJ, Warrener P, Camara MM, et al. A mul-
tifunctional bispecific antibody protects against Pseudomonas aeruginosa. Sci Transl Med.
2014;6(262):262ra155. https://doi.org/10.1126/scitranslmed.3009655.
23. Domagala JM, Hanna LD, Heifetz CL, Hutt MP, Mich TF, Sanchez JP, et al. New structure-
activity relationships of the quinolone antibacterials using the target enzyme. The develop-
ment and application of a DNA gyrase assay. J Med Chem. 1986;29(3):394–404.
24. Donald RGK, Skwish S, Forsyth RA, Anderson JW, Zhong T, Burns C, et al. A Staphylococcus
aureus fitness test platform for mechanism-based profiling of antibacterial compounds. Chem
Biol. 2009;16(8):826–36. https://doi.org/10.1016/j.chembiol.2009.07.004.
588 A. L. Erwin
25. Ehmann DE, Jahić H, Ross PL, Gu R-F, Hu J, Kern G, et al. Avibactam is a covalent, revers-
ible, non–β-lactam β-lactamase inhibitor. Proc Natl Acad Sci U S A. 2012;109(29):11663–8.
https://doi.org/10.1073/pnas.1205073109.
26. Emiola A, George J, Andrews SS. A complete pathway model for lipid A biosynthesis in
Escherichia coli. PLoS One. 2014. https://doi.org/10.1371/journal.pone.0121216.
27. Erwin AL. Antibacterial drug discovery targeting the lipopolysaccharide biosynthetic enzyme
LpxC. Cold Spring Harb Perspect Med. 2016. https://doi.org/10.1101/cshperspect.a025304.
28. Estrada A, Wright DL, Anderson AC. Antibacterial antifolates: from development through
resistance to the next generation. Cold Spring Harb Perspect Med. 2016;6(8). https://doi.
org/10.1101/cshperspect.a028324.
29. Fernandes P, Martens E, Pereira D. Nature nurtures the design of new semi-synthetic macro-
lide antibiotics. J Antibiot. 2017;70(5):527–33. https://doi.org/10.1038/ja.2016.137.
30. Fonville NC, Bates D, Hastings PJ, Hanawalt PC, Rosenberg SM. Role of RecA and the SOS
response in thymineless death in Escherichia coli. PLoS Genet. 2010;6(3):e1000865. https://
doi.org/10.1371/journal.pgen.1000865.
31. Giersing BK, Dastgheyb SS, Modjarrad K, Moorthy V. Status of vaccine research and devel-
opment of vaccines for Staphylococcus aureus. Vaccine. 2016;34(26):2962–6. https://doi.
org/10.1016/j.vaccine.2016.03.110.
32. Grillot A-L, Le Tiran A, Shannon D, Krueger E, Liao Y, O’Dowd H, et al. Second-generation
antibacterial benzimidazole ureas: discovery of a preclinical candidate with reduced meta-
bolic liability. J Med Chem. 2014;57(21):8792–816. https://doi.org/10.1021/jm500563g.
33. Grossman TH. Tetracycline antibiotics and resistance. Cold Spring Harb Perspect Med.
2016;6(4):a025387. https://doi.org/10.1101/cshperspect.a025387.
34. Gullberg E, Cao S, Berg OG, Ilbäck C, Sandegren L, Hughes D, et al. Selection of resis-
tant bacteria at very low antibiotic concentrations. PLoS Pathog. 2011;7(7). https://doi.
org/10.1371/journal.ppat.1002158.
35. Hammoudeh DI, Zhao Y, White SW, Lee RE. Replacing sulfa drugs with novel DHPS inhibi-
tors. Future Med Chem. 2013;5(11):1331–40. https://doi.org/10.4155/fmc.13.97.
36. Haynes KM, Abdali N, Jhawar V, Zgurskaya HI, Parks JM, Green AT, et al. Identification and
structure-activity relationships of novel compounds that potentiate the activities of antibiotics
in Escherichia coli. J Med Chem. 2017. https://doi.org/10.1021/acs.jmedchem.7b00453.
37. Hooper DC. Emerging mechanisms of fluoroquinolone resistance. Emerg Infect Dis.
2001;7(2):337–41. https://doi.org/10.3201/eid0702.700337.
38. Howard JJ, Sturge CR, Moustafa DA, Daly SM, Marshall-Batty KR, Felder CF, et al.
Inhibition of Pseudomonas aeruginosa by peptide-conjugated phosphorodiamidate mor-
pholino oligomers. Antimicrob Agents Chemother. 2017:AAC.01938-16. https://doi.
org/10.1128/AAC.01938-16.
39. Huber J, Donald RGK, Lee SH, Jarantow LW, Salvatore MJ, Meng X, et al. Chemical
genetic identification of peptidoglycan inhibitors potentiating carbapenem activity against
methicillin-resistant Staphylococcus aureus. Chem Biol. 2009;16(8):837–48. https://doi.
org/10.1016/j.chembiol.2009.05.012.
40. Hurdle JG, O’Neill AJ, Chopra I. Prospects for aminoacyl-tRNA synthetase inhibitors as
new antimicrobial agents. Antimicrob Agents Chemother. 2005;49(12):4821–33. https://doi.
org/10.1128/AAC.49.12.4821-4833.2005.
41. Janardhanan J, Chang M, Mobashery S. The oxadiazole antibacterials. Curr Opin Microbiol.
2016;33:13–7. https://doi.org/10.1016/j.mib.2016.05.009.
42. Jun SY, Jang IJ, Yoon S, Jang K, Yu K-S, Cho JY, et al. Pharmacokinetics and tolerance
of the phage endolysin-based candidate drug SAL200 after a single intravenous adminis-
tration among healthy volunteers. Antimicrob Agents Chemother. 2017;61(6). https://doi.
org/10.1128/AAC.02629-16.
43. Kim J-H, O’Brien KM, Sharma R, Boshoff HIM, Rehren G, Chakraborty S, et al. A genetic
strategy to identify targets for the development of drugs that prevent bacterial persistence.
Proc Natl Acad Sci. 2013;110(47):19095–100. https://doi.org/10.1073/pnas.1315860110.
18 The New Versus Old Target Debate for Drug Discovery 589
44. Kohanski MA, Dwyer DJ, Hayete B, Lawrence CA, Collins JJ. A common mechanism of
cellular death induced by bactericidal antibiotics. Cell. 2007;130(5):797–810. https://doi.
org/10.1016/j.cell.2007.06.049.
45. Kudrin P, Varik V, Oliveira SRA, Beljantseva J, Santos TDP, Dzhygyr I, et al. Subinhibitory
concentrations of bacteriostatic antibiotics induce relA-dependent and relA-independent
tolerance to β-lactams. Antimicrob Agents Chemother. 2017;61(4):e02173–16. https://doi.
org/10.1128/AAC.02173-16.
46. Lavesa-Curto M, Sayer H, Bullard D, MacDonald A, Wilkinson A, Smith A, et al.
Characterization of a temperature-sensitive DNA ligase from Escherichia coli. Microbiology.
2004;150(Pt 12):4171–80. https://doi.org/10.1099/mic.0.27287-0.
47. Leeds JA. Antibacterials developed to target a single organism: mechanisms and frequen-
cies of reduced susceptibility to the novel anti-Clostridium difficile compounds fidaxomicin
and LFF571. Cold Spring Harb Perspect Med. 2016;6(2):a025445. https://doi.org/10.1101/
cshperspect.a025445.
48. Leeds JA, Dean CR. Peptide deformylase as an antibacterial target: a critical assessment.
Curr Opin Pharmacol. 2006;6(5):445–52. https://doi.org/10.1016/j.coph.2006.06.003.
49. Ling LL, Schneider T, Peoples AJ, Spoering AL, Engels I, Conlon BP, et al. A new antibi-
otic kills pathogens without detectable resistance. Nature. 2015;advance online publication.
https://doi.org/10.1038/nature14098.
50. Liu Y, Imlay JA. Cell death from antibiotics without the involvement of reactive oxygen spe-
cies. Science. 2013;339(6124):1210–3. https://doi.org/10.1126/science.1232751, https://doi.
org/10.1126/science.1232688.
51. Lomovskaya O, Bostian KA. Practical applications and feasibility of efflux pump inhibitors
in the clinic–a vision for applied use. Biochem Pharmacol. 2006;71(7):910–8. https://doi.
org/10.1016/j.bcp.2005.12.008.
52. Maffioli SI, Zhang Y, Degen D, Carzaniga T, Gatto GD, Serina S, et al. Antibacterial
nucleoside-analog inhibitor of bacterial RNA polymerase. Cell. 2017;169(7):1240–8.e23.
https://doi.org/10.1016/j.cell.2017.05.042.
53. Mahmood HY, Jamshidi S, Sutton JM, Rahman KM. Current advances in developing inhibi-
tors of bacterial multidrug efflux pumps. Curr Med Chem. 2016;23(10):1062–81.
54. Maisonneuve E, Castro-Camargo M, Gerdes K. (p) ppGpp controls bacterial persistence
by stochastic induction of toxin-antitoxin activity. Cell. 2013;154(5):1140–50. https://doi.
org/10.1016/j.cell.2013.07.048, https://doi.org/10.1128/mBio.02267-16.
55. Mann J, Taylor PW, Dorgan CR, Johnson PD, Wilson FX, Vickers R, et al. The discovery
of a novel antibiotic for the treatment of Clostridium difficile infections: a story of an effec-
tive academic-industrial partnership. Med Chem Commun. 2015;6(8):1420–6. https://doi.
org/10.1039/c5md00238a.
56. Mattis DM, Spaulding AR, Chuang-Smith ON, Sundberg EJ, Schlievert PM, Kranz
DM. Engineering a soluble high-affinity receptor domain that neutralizes staphylococcal
enterotoxin C in rabbit models of disease. Protein Eng Des Sel. 2013;26(2):133–42. https://
doi.org/10.1093/protein/gzs094.
57. Maura D, Ballok AE, Rahme LG. Considerations and caveats in anti-virulence drug develop-
ment. Curr Opin Microbiol. 2016;33:41–6. https://doi.org/10.1016/j.mib.2016.06.001.
58. May KL, Silhavy TJ. Making a membrane on the other side of the wall. Biochim Biophys
Acta. 2016. https://doi.org/10.1016/j.bbalip.2016.10.004.
59. McLeod SM, Dougherty TJ, Pucci MJ. Novel antibacterial targets/identification of new tar-
gets by comparative genomics. In: Dougherty TJ, Pucci MJ, editors. Antibiotic discovery and
development. Boston: Springer US; 2012. p. 881–900.
60. Melton-Celsa AR, OʼBrien AD. New therapeutic developments against shiga toxin-producing
Escherichia coli. Microbiol Spectr. 2014;2(5). https://doi.org/10.1128/microbiolspec.
EHEC-0013-2013.
61. Metz JT, Hajduk PJ. Rational approaches to targeted polypharmacology: creating and navi-
gating protein-ligand interaction networks. Curr Opin Chem Biol. 2010;14(4):498–504.
https://doi.org/10.1016/j.cbpa.2010.06.166.
590 A. L. Erwin
62. Miesel L, Hecht DW, Osmolski JR, Gerding D, Flattery A, Li F, et al. Kibdelomycin is a potent
and selective agent against toxigenic Clostridium difficile. Antimicrob Agents Chemother.
2014;58(4):2387–92. https://doi.org/10.1128/AAC.00021-14.
63. Mills SD, Dougherty TJ. Cell-based screening in antibacterial discovery. In: Dougherty
TJ, Pucci MJ, editors. Antibiotic discovery and development. Boston: Springer US; 2012.
p. 901–29.
64. Mukherjee T, Boshoff H. Nitroimidazoles for the treatment of TB: past, present and future.
Future Med Chem. 2011;3(11):1427–54. https://doi.org/10.4155/fmc.11.90.
65. Mullane K, Lee C, Bressler A, Buitrago M, Weiss K, Dabovic K, et al. A multi-center, ran-
domized clinical trial to compare the safety and efficacy of LFF571 and vancomycin for
Clostridium difficile infections. Antimicrob Agents Chemother. 2014. https://doi.org/10.1128/
AAC.04251-14.
66. Nair DR, Monteiro JM, Memmi G, Thanassi J, Pucci M, Schwartzman J, et al. Characterization
of a novel small molecule that potentiates β-lactam activity against Gram positive and
Gram negative pathogens. Antimicrob Agents Chemother. 2015. https://doi.org/10.1128/
AAC.04164-14.
67. Nayak SU, Griffiss JM, Blumer J, O’Riordan MA, Gray W, McKenzie R, et al. Safety, toler-
ability, systemic exposure and metabolism of CRS3123, a methionyl-tRNA synthetase inhib-
itor developed for treatment of Clostridium difficile infections, in a Phase I study. Antimicrob
Agents Chemother. 2017. https://doi.org/10.1128/AAC.02760-16.
68. Nayar AS, Dougherty TJ, Ferguson KE, Granger BA, McWilliams L, Stacey C, et al. Novel
antibacterial targets and compounds revealed by a high-throughput cell wall reporter assay.
J Bacteriol. 2015;197(10):1726–34. https://doi.org/10.1128/JB.02552-14.
69. Novak R. Are pleuromutilin antibiotics finally fit for human use? Ann N Y Acad Sci.
2011;1241(1):71–81. https://doi.org/10.1111/j.1749-6632.2011.06219.x.
70. O’Dwyer K, Spivak AT, Ingraham K, Min S, Holmes DJ, Jakielaszek C, et al. Bacterial resis-
tance to leucyl-tRNA synthetase inhibitor GSK2251052 develops during treatment of com-
plicated urinary tract infections. Antimicrob Agents Chemother. 2015;59(1):289–98. https://
doi.org/10.1128/AAC.03774-14.
71. O’Neill AJ, Chopra I. Preclinical evaluation of novel antibacterial agents by microbiological
and molecular techniques. Expert Opin Investig Drugs. 2004;13(8):1045–63.
72. Opperman TJ, Kwasny SM, Li JB, Lewis MA, Aiello D, Williams JD, et al. DNA target-
ing as a likely mechanism underlying the antibacterial activity of synthetic bis-indole anti-
biotics. Antimicrob Agents Chemother. 2016;60(12):7067–76. https://doi.org/10.1128/
AAC.00309-16.
73. Page MGP. Beta-lactam antibiotics. In: Dougherty TJ, Pucci MJ, editors. Antibiotic discovery
and development. Boston: Springer US; 2012. p. 79–117.
74. Payne DJ, Gwynn MN, Holmes DJ, Pompliano DL. Drugs for bad bugs: confronting the
challenges of antibacterial discovery. Nat Rev Drug Discov. 2007;6(1):29–40. https://doi.
org/10.1038/nrd2201.
75. Qiu X-Q, Wang H, Lu X-F, Zhang J, Li S-F, Cheng G, et al. An engineered multidomain bac-
tericidal peptide as a model for targeted antibiotics against specific bacteria. Nat Biotechnol.
2003;21(12):1480–5. https://doi.org/10.1038/nbt913.
76. Que YA, Lazar H, Wolff M, François B, Laterre PF, Mercier E, et al. Assessment of pan-
obacumab as adjunctive immunotherapy for the treatment of nosocomial Pseudomonas
aeruginosa pneumonia. Eur J Clin Microbiol Infect Dis. 2014;33(10):1861–7. https://doi.
org/10.1007/s10096-014-2156-1.
77. Ramachandran G, Kaempfer R, Chung C-S, Shirvan A, Chahin AB, Palardy JE, et al. Cd28
homodimer interface mimetic peptide acts as a preventive and therapeutic agent in models of
severe bacterial sepsis and Gram-negative bacterial peritonitis. J Infect Dis. 2015;211(6):995–
1003. https://doi.org/10.1093/infdis/jiu556.
78. Rothstein DM. Rifamycins, alone and in combination. Cold Spring Harb Perspect Med.
2016;6(7). https://doi.org/10.1101/cshperspect.a027011.
18 The New Versus Old Target Debate for Drug Discovery 591
79. Ruiz N, Kahne D, Silhavy TJ. Advances in understanding bacterial outer-membrane biogen-
esis. Nat Rev Microbiol. 2006;4(1):57–66. https://doi.org/10.1038/nrmicro1322.
80. Ruiz N, Kahne D, Silhavy TJ. Transport of lipopolysaccharide across the cell envelope:
the long road of discovery. Nat Rev Microbiol. 2009;7(9):677–83. https://doi.org/10.1038/
nrmicro2184.
81. Schneider T, Sahl H-G. An oldie but a goodie – cell wall biosynthesis as antibiotic target path-
way. Int J Med Microbiol. 2010;300(2–3):161–9. https://doi.org/10.1016/j.ijmm.2009.10.005.
82. Schuch R, Khan BK, Raz A, Rotolo JA, Wittekind M. Bacteriophage lysin CF-301: a
potent anti-staphylococcal biofilm agent. Antimicrob Agents Chemother. 2017. https://doi.
org/10.1128/AAC.02666-16.
83. Scott RW, Tew GN. Mimics of host defense proteins; strategies for translation to therapeutic
applications. Curr Top Med Chem. 2017;17(5):576–89.
84. Seiple IB, Zhang Z, Jakubec P, Langlois-Mercier A, Wright PM, Hog DT, et al. A platform
for the discovery of new macrolide antibiotics. Nature. 2016;533(7603):338–45. https://doi.
org/10.1038/nature17967.
85. Shan Y, Gandt AB, Rowe SE, Deisinger JP, Conlon BP, Lewis K. ATP-dependent per-
sister formation in Escherichia coli. MBio. 2017;8(1):e02267–16. https://doi.org/10.1128/
mBio.02267-16.
86. Sharma P, Wang N, Kranz DM. Soluble T cell receptor Vβ domains engineered for high-affinity
binding to staphylococcal or streptococcal superantigens. Toxins (Basel). 2014;6(2):556–74.
https://doi.org/10.3390/toxins6020556.
87. Silver LL. Multi-targeting by monotherapeutic antibacterials. Nat Rev Drug Discov.
2007;6(1):41–55. https://doi.org/10.1038/nrd2202.
88. Silver LL. Challenges of antibacterial discovery. Clin Microbiol Rev. 2011;24(1):71–109.
https://doi.org/10.1128/CMR.00030-10.
89. Silver LL. A Gestalt approach to Gram-negative entry. Bioorg Med Chem. 2016. https://doi.
org/10.1016/j.bmc.2016.06.044.
90. Singh SB, Young K, Miesel L. Screening strategies for discovery of antibacterial natural
products. Expert Rev Anti-Infect Ther. 2011;9(8):589–613. https://doi.org/10.1586/eri.11.81.
91. Sjuts H, Vargiu AV, Kwasny SM, Nguyen ST, Kim H-S, Ding X, et al. Molecular basis for
inhibition of AcrB multidrug efflux pump by novel and powerful pyranopyridine derivatives.
Proc Natl Acad Sci U S A. 2016;113(13):3509–14. https://doi.org/10.1073/pnas.1602472113.
92. Sköld O. Sulfonamides and trimethoprim. Expert Rev Anti-Infect Ther. 2010;8(1):1–6.
https://doi.org/10.1586/eri.09.117.
93. Srinivas N, Jetter P, Ueberbacher BJ, Werneburg M, Zerbe K, Steinmann J, et al.
Peptidomimetic antibiotics target outer-membrane biogenesis in Pseudomonas aeruginosa.
Science. 2010;327(5968):1010–3. https://doi.org/10.1126/science.1182749.
94. Starkey M, Lepine F, Maura D, Bandyopadhaya A, Lesic B, He J, et al. Identification of anti-
virulence compounds that disrupt quorum-sensing regulated acute and persistent pathogenic-
ity. PLoS Pathog. 2014;10(8). https://doi.org/10.1371/journal.ppat.1004321.
95. Suckling CJ. The antibacterial drug MGB-BP3 : from discovery to clinical trial. Chem Biol
Interface. 2015;5(3):166–74.
96. Takatsuka Y, Chen C, Nikaido H. Mechanism of recognition of compounds of diverse struc-
tures by the multidrug efflux pump AcrB of Escherichia coli. Proc Natl Acad Sci U S A.
2010;107(15):6559–65. https://doi.org/10.1073/pnas.1001460107.
97. Tari LW, Li X, Trzoss M, Bensen DC, Chen Z, Lam T, et al. Tricyclic GyrB/ParE (TriBE)
inhibitors: a new class of broad-spectrum dual-targeting antibacterial agents. PLoS One.
2013;8(12):e84409. https://doi.org/10.1371/journal.pone.0084409.
98. Therien AG, Huber JL, Wilson KE, Beaulieu P, Caron A, Claveau D, et al. Broadening the
spectrum of β-lactam antibiotics through inhibition of signal peptidase type I. Antimicrob
Agents Chemother. 2012;56(9):4662–70. https://doi.org/10.1128/AAC.00726-12.
99. Tommasi R, Brown DG, Walkup GK, Manchester JI, Miller AA. ESKAPEing the labyrinth of
antibacterial discovery. Nat Rev Drug Discov. 2015;14(8):529–42. https://doi.org/10.1038/
nrd4572.
592 A. L. Erwin
100. Utaida S, Dunman PM, Macapagal D, Murphy E, Projan SJ, Singh VK, et al. Genome-wide
transcriptional profiling of the response of Staphylococcus aureus to cell-wall-active antibiot-
ics reveals a cell-wall-stress stimulon. Microbiology. 2003;149(Pt 10):2719–32. https://doi.
org/10.1099/mic.0.26426-0.
101. Vaara M, Siikanen O, Apajalahti J, Fox J, Frimodt-Møller N, He H, et al. A novel polymyxin
derivative that lacks the fatty acid tail and carries only three positive charges has strong syn-
ergism with agents excluded by the intact outer membrane. Antimicrob Agents Chemother.
2010;54(8):3341–6. https://doi.org/10.1128/AAC.01439-09.
102. van Miert AS. The sulfonamide-diaminopyrimidine story. J Vet Pharmacol Ther.
1994;17(4):309–16.
103. Wang DY, Abboud MI, Markoulides MS, Brem J, Schofield CJ. The road to avibactam: the
first clinically useful non-β-lactam working somewhat like a β-lactam. Future Med Chem.
2016;8(10):1063–84. https://doi.org/10.4155/fmc-2016-0078.
104. Wang H, Mann PA, Xiao L, Gill C, Galgoci AM, Howe JA, et al. Dual-targeting small-
molecule inhibitors of the Staphylococcus aureus FMN riboswitch disrupt riboflavin
homeostasis in an infectious setting. Cell Chem Biol. 2017;24(5):576–88.e6. https://doi.
org/10.1016/j.chembiol.2017.03.014.
105. Wei J-R, Krishnamoorthy V, Murphy K, Kim J-H, Schnappinger D, Alber T, et al.
Depletion of antibiotic targets has widely varying effects on growth. Proc Natl Acad Sci.
2011;108(10):4176–81. https://doi.org/10.1073/pnas.1018301108.
106. Wright PM, Seiple IB, Myers AG. The evolving role of chemical synthesis in antibacterial
drug discovery. Angew Chem Int Ed Engl. 2014;53(34):8840–69. https://doi.org/10.1002/
anie.201310843.
107. Xu W, Wang W, Wang X. Gold-catalyzed cyclization leads to a bridged tetracyclic indolenine
that represses β-lactam resistance. Angew Chem Int Ed. 2015;54(33):9546–9. https://doi.
org/10.1002/anie.201503736.
108. Yao J, Rock CO. Bacterial fatty acid metabolism in modern antibiotic discovery. Biochim
Biophys Acta. 2016. https://doi.org/10.1016/j.bbalip.2016.09.014.
109. Young K, Silver LL, Bramhill D, Cameron P, Eveland SS, Raetz CR, et al. The envA perme-
ability/cell division gene of Escherichia coli encodes the second enzyme of lipid A biosyn-
thesis. UDP-3-O-(R-3-hydroxymyristoyl)-N-acetylglucosamine deacetylase. J Biol Chem.
1995;270(51):30384–91.
110. Yuan Z, White RJ. The evolution of peptide deformylase as a target: contribution of bio-
chemistry, genetics and genomics. Biochem Pharmacol. 2006;71(7):1042–7. https://doi.
org/10.1016/j.bcp.2005.10.015.
111. Zhou J, Bhattacharjee A, Chen S, Chen Y, Duffy E, Farmer J, et al. Design at the atomic level:
generation of novel hybrid biaryloxazolidinones as promising new antibiotics. Bioorg Med
Chem Lett. 2008;18(23):6179–83. https://doi.org/10.1016/j.bmcl.2008.10.014.
Chapter 19
Non-quinolone Topoisomerase Inhibitors
19.1 Introduction
There is no question that the development of antibiotics has been among the most
important advances of the twentieth century, saving a countless number of lives. Not
only are antibiotics used in the direct treatment of bacterial infections, but they are
also important in surgical situations (e.g., transplant surgery, joint replacement) by
preventing infections. However, the twenty-first century has seen increasing con-
cern due to the rise in antimicrobial-resistant bacterial infections [1]. Antibiotic
resistance is a growing global threat, with resistance to all classes of antibiotics now
reported worldwide. The resistance problem is compounded by a lack of innovation
and few new structural classes of antibiotics being brought to the clinic [2–4]. To
tackle antibiotic resistance, we need to review our stewardship of existing antibiot-
ics and expand efforts to discover new agents that are not susceptible to known
resistance mechanisms.
Quinolones (specifically fluoroquinolones (FQs); Fig. 19.1) are a potent class of
synthetic antibiotics that target DNA gyrase and DNA topoisomerase IV, essential
enzymes that are ubiquitous among bacterial species. The quinolones were discov-
ered in the early 1960s, and they are now the most successful class of topoisomerase
inhibitors; fluoroquinolones are widely prescribed in the USA, Europe, and most
regions of the world [5, 6]. This heavy consumption of fluoroquinolones has led to
an increase in resistance that derives from a variety of processes including upregula-
tion of efflux pumps, reduced ability to take up the drug, plasmid-based resistance,
or mutations in the gyrase and/or topo IV genes [7, 8]. This widespread resistance
has resulted in revised stewardship guidelines for quinolones [6] as well as the
Fig. 19.1 Fluoroquinolone compounds. The basic FQ skeleton is shown along with three FQ
compounds in clinical use
DNA topoisomerases are enzymes that can interconvert different topological forms
of DNA; their reactions include relaxation of supercoils, decatenation, and unknot-
ting [10, 11]. They are present and essential in all organisms and are involved in
DNA replication and transcription, preventing the buildup of unwanted supercoils
and resolving catenated products [12]. Topoisomerases are classified into two types,
I and II, depending on whether they catalyze reactions involving single (I)- or dou-
ble (II)-stranded breaks in DNA; they can also be further divided into subtypes IA,
IB, and IC and IIA and IIB, dependent on mechanistic and evolutionary consider-
ations [13, 14]. The subjects of this chapter, DNA gyrase (Fig. 19.2) and DNA
topoisomerase (topo) IV, are type IIA enzymes. Type I enzymes catalyze their reac-
tions by transiently breaking one strand of DNA and forming a covalent bond
between either the 5′ (IA) or 3′ (IB) phosphate at the break site and the active-site
tyrosine of the enzyme. The topoisomerase reaction occurs via a swivel mechanism
(type IB) or by strand passage, where a single- or double-stranded segment of DNA
is passed through the break (type IA). Type II topoisomerases make a transient
double-strand break in DNA, forming covalent bonds with the 5′-phosphates at the
break site and passing a double-stranded segment of DNA (the so-called “T” or
“transported” segment) through the break [10, 11] in the “G” or “gate” segment.
Although this is the case for both IIA and IIB enzymes, the two subtypes are differ-
ent in several other structural and mechanistic aspects [15]. The principal member
of the type IIB subtype is topo VI, originally discovered in archaea and now shown
to occur in plants and plasmodial parasites [16, 17]. The members of the IIA sub-
type, gyrase, topo IV, and eukaryotic topo II, are better known and have been more
extensively studied. All these enzymes can relax supercoiled DNA; gyrase is dis-
tinct in also being able to catalyze the introduction of negative supercoils into DNA.
Structural and mechanistic aspects of these enzymes have been extensively reviewed
Fig. 19.2 Schematic representation of DNA gyrase (A2B2) complexed with G-segment
DNA. NTD, N-terminal domain; CTD, C-terminal domain; TOPRIM, topoisomerase-primase
domain; WHD, winged-helix domain
596 A. Maxwell et al.
elsewhere [10, 11, 18]. The main topic of this chapter concerns their role as targets
for antibacterial agents; other recent reviews also address this topic [10, 19–21].
Because of their essentiality and the fact that their reactions proceed via transient
breaks in DNA, topoisomerases have become targets for both anticancer and anti-
bacterial chemotherapy [10, 21, 22]. Many topoisomerase-targeted drugs act by sta-
bilizing the DNA-protein covalent complexes that form between the enzymes and
DNA during the topoisomerase reaction cycle. However, there are other agents (see
below) that act by interrupting the reaction cycle in other ways; these have been
dubbed “catalytic inhibitors.” Quinolone compounds, and more specifically fluoro-
quinolones (FQs), such as ciprofloxacin, moxifloxacin, and levofloxacin (Fig. 19.1),
are highly successful antibiotics that are used for a wide range of clinical conditions
[23, 24]. However, due to resistance to these compounds, new agents that match
their clinical characteristics, but are not affected by quinolone-resistance mutations,
are urgently needed. This chapter reviews such compounds.
Although, in principle, DNA topo I is a valid target for antibacterial compounds
and a number have been investigated [25], there are currently no topo I inhibitors in
clinical use, although recent work suggests that there is scope to develop antibiotics
targeted to this enzyme [26, 27]; topo I will not be further discussed in this chapter.
The scope of DNA gyrase and DNA topo I as targets for tuberculosis therapy has
recently been reviewed [20].
There are two main mechanisms of inhibition of topoisomerases: catalytic
inhibition and topoisomerase poisons. Catalytic inhibitors arrest enzyme activity
and need to bind to their targets with reasonably high affinity to be effective.
Poisons are compounds that stabilize the topoisomerase-DNA cleavage complex
and tend to be more effective, as relatively low occupancy of the inhibitor bound
to its target can lead to cell death, which involves chromosome fragmentation,
induction of the SOS response, and possibly the induction of reactive oxygen
species [28, 29] (see Chap. 20).
death; it is also the basis for the mode of action of a number of successful antitumor
drugs that target human topo II [30]. Finding agents that act in a similar way that can
substitute for FQs has been a major challenge.
As we now know, from X-ray crystallography, how FQs work at the molecular
level [31, 32], we can use this information to develop other molecules that act in a
similar way but that may avoid the problems of quinolone resistance. As will be
seen from some of the examples below, it is possible to find compounds that bind at
sites distinct from the quinolone-binding site and that can also stabilize cleavage
complexes. This type of compound that can avoid quinolone cross-resistance per-
haps represents the best opportunities for finding novel gyrase−/topo IV-targeted
antibiotics with clinical potential. The term NBTI (novel (non-fluoroquinolone)
bacterial type II topoisomerase inhibitor) has been introduced [31] to encompass
compounds that demonstrate these properties.
FQs work through intercalation of two compounds, one at each cleavage site,
into the DNA (Fig. 19.3) [31, 32]; the compounds also interact with the enzyme
through a water-metal-ion bridge [33]. Mutations affecting the residues involved in
contacting the metal ion are found in clinically-isolated strains of FQ-resistant
pathogenic bacteria. It is likely that disrupting the water-metal ion bridge destabi-
lizes the ternary complex, mitigating cleavage-complex formation (see Chap. 16).
Efforts have therefore been made to develop compounds that do not rely on this
water-metal ion bridge. Several compounds share a scaffold with the FQs, for exam-
ple, the quinazolinediones (QZDs) [34–36] and the imidazopyrazinones (IPYs)
[37, 38]. Compounds of these classes are able to intercalate at the FQ-binding site
and stabilize cleavage complexes without relying on the water-metal ion bridge [37,
39, 40]. The water-metal ion bridge is a major contributor to the stable binding of
bacterial topoisomerase by the FQs. Removing it requires other contacts to be estab-
lished by putative intercalating compounds of similar scaffold to afford the same
affinity and, by extension, equivalent potency (although the relationship between
efficiency of poisoning in vitro and potency against bacterial pathogens is far from
straightforward). Indeed in the case of QZDs, efficiency of poisoning tends to be
significantly lower than the FQs [41–43], and some of the medicinal chemistry
effort has focused on developing alternative contacts with the protein, for instance,
by modifying the C7 substituent [44]. In the case of the IPYs, the absence of a water-
metal ion bridge is compensated for by direct contact with one of the residues
involved in the bridge, which unfortunately affords some degree of cross-resistance
between quinolones and these compounds [37]. The IPYs also establish contact
with the arginine situated next to the catalytic tyrosine, which is presumably essen-
tial for catalytic activity [37]. In the case of the QZDs, attempts have been made to
develop such contacts in order to improve the activity of the compounds while mini-
mizing resistance [45].
Structural information on the protein-DNA-drug complexes has shown that they
all share a characteristic DNA extension, increasing the distance between the two
scissile phosphates compared to an uncleaved “binary” enzyme-DNA complex [31,
33, 37, 46]. In gyrase, this extension is associated with conformational movement
involving sliding of the two GyrA subunits against each other and tilting of the
598 A. Maxwell et al.
Fig. 19.3 Binding sites of cleavage complex-stabilizing compounds. Two structures are super-
posed to show binding by moxifloxacin (PDB 5CDQ), GSK945237, and thiophene2 (PDB 5NPP)
to the S. aureus gyrase core (C-terminal region of GyrB fused to the N-terminal region of GyrA).
The TOPRIM domains from 5CDQ are superposed and displayed instead of the ones from 5NPP;
the DNA and GyrA domain displayed are from 5NPP. An orientation cartoon shows the color-
coded compounds binding at distinct sites on the enzyme. While moxifloxacin intercalates directly
at the cleavage site (as with the IPYs and QZDs), the triazaacenaphthylene and thiophene com-
pounds bind at different allosteric sites and stabilize the cleaved intermediate. Triazaacenaphthylenes
block the sliding of the GyrA subunits against one another, thereby blocking resealing, whereas the
thiophene blocks the TOPRIM domain in a tilted conformation against the DNA
TOPRIM domains toward the dyad axis, suggesting that the enzyme can manipulate
DNA geometry to favor cleavage. The dynamic nature of the DNA gate also allows
for a variety of structurally unrelated compounds to bind at the FQ site without
requiring a water-metal ion bridge and thereby bypassing most FQ-resistance muta-
tions. For instance, etoposide, a eukaryotic topoisomerase II inhibitor [46], is also
able to stabilize gyrase-DNA cleavage complexes through intercalation into DNA
and interaction with residues conserved in the related human topoisomerase
II. Likewise, the spiropyrimidinetriones (such as QPT-1) can stabilize cleavage
complexes by intercalation much like the FQs despite being structurally different
[46]. However, unlike etoposide, QPT-1 is not active on human topoisomerase II
despite interacting with conserved residues. It was thought that the water-metal ion
bridge conferred bacterial specificity to the FQs, but the case of QPT-1 shows that
more is at play. This is an important issue, as development of bacterial topoisomerase
19 Non-quinolone Topoisomerase Inhibitors 599
thiazole heterocycles [50]. MccB17 targets bacterial gyrase and can stabilize the
cleavage complex but in a manner distinct from quinolones [51, 52]. The only
known mutation in gyrase that confers resistance to MccB17 is at the C-terminal
end of GyrB (Trp751 to Arg) [52, 53]; no other drug-resistance mutations map here.
Although MccB17 is a potentially attractive option as an antibacterial compound,
its poor physicochemical properties have hampered its development as a drug can-
didate. Despite significant work on this toxin, its binding site and mode of action on
gyrase are not known. However, the toxin and fragments thereof have been chemi-
cally synthesized, and fragments have also been made using molecular biology/
biochemical methods [54–57]. Some of these fragments, with molecular weights
<2 kDa, show activity suggesting that it may be possible to generate smaller ver-
sions of MccB17 with more attractive physicochemical properties that might have
potential as antibacterial agents in the future.
Other non-small molecule inhibitors of gyrase include the phytotoxin albicidin,
the CcdB protein toxin, the E. coli GyrI protein, and the pentapeptide-repeat pro-
teins, such as Qnr and MfpA; these agents have been reviewed elsewhere [10, 19].
Although the toxin protein CcdB (MW ~12 kDa) is outside the scope of this review,
it is interesting to note that peptides based on CcdB as short as 18 amino acids can
retain inhibitory activity on gyrase [58].
The term “catalytic inhibitors” refers to agents that do not inhibit gyrase/topo IV by
stabilizing the DNA cleavage complex but affect another aspect of the catalytic
cycle. The majority of these are ATPase inhibitors, e.g., aminocoumarins and cyclo-
thialidines, but other types, such as simocyclinones, which inhibit DNA binding,
also exist. Although arguably catalytic inhibitors are less likely than cleavage
complex-stabilizing compounds to be effective antibiotics, such catalytic inhibition
is effective with other antibiotic targets (e.g., rifamycin on RNA polymerase, trim-
ethoprim on dihydrofolate reductase [59]); thus, there is no reason a priori that they
should not succeed as antibiotics. Indeed novobiocin, see below, has been utilized
as a clinical antibiotic.
Aminocoumarin antibiotics (Fig. 19.4) that target DNA gyrase were discovered
as Streptomyces natural products in the 1950s; these “classical” agents are novobio-
cin, clorobiocin, and coumermycin A1 [60, 61]. Early on it was established that
these compounds are competitive inhibitors of the gyrase ATPase reaction [62]; they
also have activity against topo IV [63]. In this sense, they are classic catalytic inhibi-
tors that affect ATP-dependent topoisomerase reactions without stabilizing the
cleavage complex. Specifically, they bind to the ATPase (N-terminal) domain of
GyrB and block the binding of ATP, a process that was definitively established using
X-ray crystallography [64, 65]. This conclusion was somewhat surprising given that
aminocoumarins do not obviously resemble ATP (Fig. 19.4). In fact, it is the sugar
ring of the aminocoumarins that overlaps the adenine-binding site in the ATPase
19 Non-quinolone Topoisomerase Inhibitors 601
Fig. 19.4 Catalytic inhibitors of gyrase and topo IV. (a) Novobiocin, clorobiocin, novclobiocin
401, and coumermycin A1, (b) cyclothialidine, (c) simocyclinone D8. ATPase inhibitors are boxed
in orange, DNA-binding inhibitors in green
pocket, thereby preventing the binding of ATP. Several crystal structures of gyrase
B fragments bound to aminocoumarin compounds now exist (Fig. 19.5 [19, 61]);
these fragments have potentiated the design of alternative compounds that can bind
at the same site (see below).
Although aminocoumarins are very effective inhibitors of gyrase and topo IV,
with Kd values in the 1–20 nM range [66], they have not had a high degree of suc-
cess as clinical antibiotics. Although novobiocin has been used on its own and in
combination as an antibiotic, safety concerns have led to discontinuation of its
usage [19]. Its toxicity issues may stem in part from its binding site, the ATPase
domain of GyrB/ParE, which is part of the GHKL ATPase/kinase superfamily [67],
and secondary eukaryotic targets are therefore likely. Indeed it has been possible to
“redesign” novobiocin to target Hsp90 (see below) [68]. Furthermore, the amino-
coumarins suffer from solubility issues, making it difficult to develop them as drugs.
To circumvent these problems, attempts have been made to prepare aminocou-
marin derivatives with superior properties. These efforts have been made possible
by the identification, sequencing, and annotation of the gene clusters for the three
“classical” aminocoumarins [69]. This has led to a detailed understanding of the
biosynthetic pathways for these compounds [60]. With this information it has been
possible to use various technologies: genetic engineering, combinatorial biosynthe-
sis, and mutasynthesis, to generate novel aminocoumarins, which can then be tested
for antibacterial activity and for their effects on the target enzymes [70]. A signifi-
cant amount of this effort has been carried out by Heide and coworkers, who coined
602 A. Maxwell et al.
Fig. 19.5 Binding site for aminocoumarins and other ATPase inhibitors. The superposition of
structures of the ATP-binding domain of DNA gyrase bound to ADPNP (5′-adenylyl-β,γ-
imidodiphosphate), novobiocin, and cyclothialidine. The nucleotide binds inside a furrow and pro-
motes dimerization through binding of a projection from the cognate monomer covering the
opening of the furrow. Both novobiocin and cyclothialidine impinge on nucleotide binding as their
binding sites overlap with the nucleotide-binding site
the name “novclobiocins” for molecules that are hybrids between clorobiocin and
novobiocin [71]. Many novclobiocins have been produced and their activities
assessed [69, 72]. One specific example is novclobiocin 401 (Fig. 19.4), which con-
tains the catechol moiety 3,4-dihydroxybenzoic acid, which assists import across
the bacterial cell envelope [73]. This modification improved the penetration into E.
coli, and the analog also retained good activity against gyrases from E. coli and
Staphylococcus aureus. Whether further improvements to the aminocoumarins can
generate compounds with clinical potential remains to be seen.
Cyclothialidines (Fig. 19.4) are Streptomyces-derived natural products and, like
aminocoumarins, are also competitive inhibitors of the gyrase ATPase reaction and
bind at essentially the same site in GyrB [65]. Although their antibacterial potency
is generally poor, their novel structures and good in vitro activity against gyrase
warranted further investigation [74–76]. For example, a chemistry program at
19 Non-quinolone Topoisomerase Inhibitors 603
Fig. 19.6 Binding site of simocyclinone. The compound binds in a “saddle” formed by the two
GyrA subunits that normally accommodates the G-DNA; two molecules of simocyclinone D8 bind
per GyrA dimer. GyrA55 is a truncated version of the GyrA-NTD
However, apart from the limited success of novobiocin, these have yet to be devel-
oped as clinically useful compounds. But, given their high degree of chemical diver-
sity and the great success of natural products directed at other targets, it is realistic
to expect that gyrase/topo IV-targeted compounds will be successful in the future.
There are a whole host of compounds that have been discovered that target gyrase/
topo IV and that may have clinical potential going forward. On account of space
limitations, only a few examples are given here to illustrate this type of work.
In 2007, a novel inhibitor, identified in an in vivo E. coli screen, demonstrated
inhibitory effects on cellular division [95]: N-benzyl-3-sulfonamidopyrrolidine,
then referred to as “534F6” and now as gyramide A (Fig. 19.7). The target of gyra-
mide A was shown to be DNA gyrase, reportedly with a unique binding site and
mechanism of inhibition [96]. Gyramides B and C were synthesized through modi-
fication of gyramide A. Using an E. coli strain having an inactive AcrAB-TolC mul-
tidrug efflux pump, gyramide A gave an MIC of 10 μM. The sequencing of
spontaneous gyramide A-resistant mutants revealed amino acid substitutions in both
GyrA and GyrB that clustered adjacent to the cleavage-religation site and were dis-
tinct from those that produce quinolone resistance [96]. Cross-resistance to quino-
lones was also absent in these gyramide A-resistant mutants. When paired with an
efflux pump inhibitor, gyramides A, B, and C demonstrated effective antibacterial
action against Gram-negative bacteria, as well as activity against some Gram-
positive species [96].
In 2014, gyramide A was suggested to be a competitive inhibitor of ATP hydro-
lysis [97], in contrast to the earlier report, although this was later withdrawn [98]. It
is possible that gyramide A reduces the rate of ATP hydrolysis indirectly by perturb-
ing the binding to DNA. The 2014 paper also showed that gyramide A is a specific
inhibitor of gyrase, showing no activity against topo IV in vitro.
606 A. Maxwell et al.
IV decatenation (IC50 0.009 and 0.5 μM, respectively), but it only weakly inhibits
E. coli topo IV decatenation (IC50 29 μM). Also, producing an effect very similar to
novobiocin, kibdelomycin is a potent inhibitor of the ATPase activity of both E. coli
gyrase and topo IV (IC50 0.011 and 0.9 μM). When crystallized bound to both GyrB
and ParE, kibdelomycin exhibited a unique U-shaped binding mode having exten-
sive hydrophobic and polar interactions with surface residues of both proteins,
while the pyrrolamide moiety extended deep into the ATP-binding pocket. This
behavior is distinct from the binding of aminocoumarins and is consistent in the
lack of cross-resistance between the two [107].
Closthioamide, a member of the polythioamide class of DNA gyrase inhibitors
(Fig. 19.7), was the first secondary metabolite isolated from the obligate anaerobe
Clostridium cellulolyticum. While genome mining identified genes involved in
polyketide and peptide synthesis, no bioactive compounds were found until the
addition of aqueous soil extract to the culture led to the production of closthioamide
[108]. Closthioamide is symmetrical in structure with six thioamide groups flanked
on either side by a phenol group. It has demonstrated potent activity against Gram-
positive strains including MRSA and VRE, giving MICs of 0.14 and 0.4 mg/L,
respectively. Only moderate inhibitory effects were found using wild-type E. coli
strains (MICs of 2.5–3.5 mg/L); however, activity could be increased dramatically
using the membrane permeability enhancer polymyxin B nonapeptide (PMBN) or
using a drug efflux pump-deficient strain with and without PMBN, to MICs of
0.625, 0.035, and 0.31 mg/L, respectively. These data clearly suggest that the outer
membrane and efflux pumps of Gram-negative bacteria are responsible for the
reduction of closthioamide’s efficacy [109].
Closthioamide’s mode of action is unlikely to involve cleavage-complex stabili-
zation, as very little linear DNA is detected when the agent is present in gyrase-
DNA reaction mixtures. However, closthioamide did reduce the ATPase activity of
gyrase and topo IV by 80% and 60%, respectively. Although the compound also
inhibits the ATP-independent relaxation activity of gyrase, it is doubtful that it is a
competitive ATPase inhibitor. Rather, it is more likely to allosterically interfere with
ATP hydrolysis, inhibiting the enzyme using a novel binding mode, one that has
been likened to the mode of action diospyrin [109].
Using the soil-dwelling actinomycete, Amycolatopsis sp., amycolamicin
(Fig. 19.7) was isolated and found to have potent, broad-spectrum antibiotic activ-
ity. Its structure, determined using a combination of NMR spectroscopy, chemical
degradation, X-ray analysis, and functional group modification, is described in
detail in ref. [110]. The compound has shown promise against the Gram-positive
MRSA, VRE, and penicillin-resistant S. pneumoniae (all with MICs in the range of
0.25–1 μg/mL) as well as against the Gram-negative ampicillin-resistant and beta-
lactamase-positive amoxicillin-clavulanate-resistant strains of H. influenzae (MIC
0.5 and 2 μg/mL, respectively). The target was determined to be bacterial type II
topoisomerases, with amycolamicin inhibiting gyrase and topo IV with IC50s of
0.024 and 6.2 μg/mL, respectively.
608 A. Maxwell et al.
A clear message that emerges from this review is that gyrase and topo IV are excel-
lent targets for antibacterial chemotherapy. The success of the fluoroquinolones
attests to the value of these targets. However, what is clearly needed are new agents,
ideally cleavage complex-stabilizing compounds, that can replace the quinolones.
The inhibitors described in this chapter have been discovered and developed using
a variety of approaches, including screening chemical libraries, following natural
product leads, and fragment-based and in silico approaches; they include both
target-led and phenotypic-led methodologies. Although arguments can be advanced
that favor one or another of these approaches, it is likely that we need to retain a
diverse range of approaches to discover the types of agent we seek. Increased chem-
ical diversity, perhaps through investigating novel sources of natural products, is
likely to be a key component to success going forward.
610 A. Maxwell et al.
While the approaches to new compound discovery are an important issue for
discussion, a more challenging question is who will carry out this work? The vast
majority of antibiotics available for clinical use have been developed and produced
by large pharma companies. While academics and small companies may have dis-
covered and researched compounds, it is only large pharma that has the know-how
and resources to bring them to market. However, in the current economic and politi-
cal climate, large pharma is pulling out of antibiotic R & D, mainly on account of
profitability issues, with a consequent reduction in effort in some cases and com-
plete withdrawal in others [135, 136]. We are now faced with potential significant
shortcomings in the discovery pipeline [137]. This is leading to the perilous situa-
tion of increasing antimicrobial-resistant bacterial infections and a paucity of new
agents to treat them [138]. It is probably essential that governments, ideally working
in cooperation, confront this challenge and provide the necessary resources and
incentives to sustain the antibiotic discovery effort. It is likely that this will be
increasingly carried out by the academic and SME sectors, resourced through public
financing [139, 140]; it is clear that governments need to take action to address this
crisis.
Government action would be starting with a solid base, as some pharma compa-
nies are still developing novel quinolones, particularly for niche markets (e.g., dela-
floxacin produced by Melinta Therapeutics [141]). Moreover, there are many
examples of non-quinolone agents being developed, some of which have been
described in this chapter. Among the active companies are AstraZeneca (e.g., spiro-
pyrimidinetriones [142]), Cubist (pyrazolopyridones [125]), GSK (NBTIs and thio-
phenes [31, 48]), Pfizer (quinazolinediones [35]), Sanofi (IPYs [38]), and Vertex
(benzimidazoles [143]). It is to be hoped that mechanisms will be found to sustain
these efforts and ensure that the considerable expertise in this area of investigation
is not lost.
Major Points
• Quinolones are highly successful antibiotics, but resistance is a serious
problem.
• DNA topoisomerases (particularly gyrase and topo IV) are important targets for
antimicrobial chemotherapy that should continue to be exploited.
• Cleavage-complex stabilization is an excellent mode of action for antibiotics; it
is possible to find new, non-quinolone, compounds that work via this
mechanism.
• Catalytic inhibitors of topoisomerases can, in principle, be developed as antibiot-
ics of the future.
• It is important to sustain a variety of approaches for discovering new
antibiotics.
• Big pharma companies cannot necessarily be relied upon to develop new antibi-
otics going forward; other ways of developing new drugs need to be explored.
19 Non-quinolone Topoisomerase Inhibitors 611
References
1. Boucher HW, Talbot GH, Bradley JS, Edwards JE, Gilbert D, Rice LB, et al. Bad bugs, no
drugs: no ESKAPE! An update from the Infectious Diseases Society of America. Clin Infect
Dis. 2009;48(1):1–12.
2. Bush K, Courvalin P, Dantas G, Davies J, Eisenstein B, Huovinen P, et al. Tackling antibiotic
resistance. Nat Rev Microbiol. 2011;9(12):894–6.
3. Lewis K. Antibiotics: Recover the lost art of drug discovery. Nature. 2012;485(7399):439–40.
4. Walsh CT, Wencewicz TA. Prospects for new antibiotics: a molecule-centered perspective.
J Antibiot (Tokyo). 2014;67(1):7–22.
5. Linder JA, Huang ES, Steinman MA, Gonzales R, Stafford RS. Fluoroquinolone prescribing
in the United States: 1995 to 2002. Am J Med. 2005;118(3):259–68.
6. Pitiriga V, Vrioni G, Saroglou G, Tsakris A. The impact of antibiotic stewardship programs in
combating quinolone resistance: a systematic review and recommendations for more efficient
interventions. Adv Ther. 2017;34(4):854–65.
7. Hooper DC. Mechanisms of quinolone resistance. In: Hooper DC, Rubinstein E, editors.
Quinolone antimicrobial agents. 3rd ed. Washington, DC: ASM Press; 2003. p. 41–67.
8. Redgrave LS, Sutton SB, Webber MA, Piddock LJ. Fluoroquinolone resistance: mechanisms,
impact on bacteria, and role in evolutionary success. Trends Microbiol. 2014;22(8):438–45.
9. WHO. Critically important antimicrobials for human medicine – 5 rev. June 2017 Ed.
Geneva: World Health Organisation – WHO Advisory Group on Integrated Surveillance of
Antimicrobial Resistance (AGISAR); 2017.
10. Bush NG, Evans-Roberts K, Maxwell A. DNA topoisomerases. EcoSal Plus. 2015;6(2).
11. Vos SM, Tretter EM, Schmidt BH, Berger JM. All tangled up: how cells direct, manage and
exploit topoisomerase function. Nat Rev Mol Cell Biol. 2011;12(12):827–41.
12. Wang JC. Cellular roles of DNA topoisomerases: a molecular perspective. Nat Rev Mol Cell
Biol. 2002;3(6):430–40.
13. Forterre P, Gribaldo S, Gadelle D, Serre MC. Origin and evolution of DNA topoisomerases.
Biochimie. 2007;89(4):427–46.
14. Wang JC. DNA topoisomerases. Annu Rev Biochem. 1996;65:635–92.
15. Gadelle D, Filee J, Buhler C, Forterre P. Phylogenomics of type II DNA topoisomerases.
BioEssays. 2003;25(3):232–42.
16. Aravind L, Iyer LM, Wellems TE, Miller LH. Plasmodium biology: genomic gleanings. Cell.
2003;115(7):771–85.
17. Sugimoto-Shirasu K, Stacey NJ, Corsar J, Roberts K, McCann MC. DNA topoisomerase VI
is essential for endoreduplication in Arabidopsis. Curr Biol. 2002;12(20):1782–6.
18. Wang JC. Moving one DNA double helix through another by a type II DNA topoisomerase:
the story of a simple molecular machine. Q Rev Biophys. 1998;31(2):107–44.
19. Mayer C, Janin YL. Non-quinolone inhibitors of bacterial type IIA topoisomerases: a feat of
bioisosterism. Chem Rev. 2014;114(4):2313–42.
20. Nagaraja V, Godbole AA, Henderson SR, Maxwell A. DNA topoisomerase I and DNA gyrase
as targets for TB therapy. Drug Discov Today. 2017;22(3):510–8.
21. Collin F, Karkare S, Maxwell A. Exploiting bacterial DNA gyrase as a drug target: current
state and perspectives. Appl Microbiol Biotechnol. 2011;92(3):479–97.
22. Pommier P, Capranico G, Orr A, Kohn KW. Distribution of topoisomerase II cleavage sites in
simian virus 40 DNA and the effects of drugs. J Mol Biol. 1991;222:909–24.
23. De Souza MV. New fluoroquinolones: a class of potent antibiotics. Mini Rev Med Chem.
2005;5(11):1009–17.
24. Oliphant CM, Green GM. Quinolones: a comprehensive review. Am Fam Physician.
2002;65(3):455–64.
612 A. Maxwell et al.
25. Tse-Dinh YC. Bacterial topoisomerase I as a target for discovery of antibacterial compounds.
Nucleic Acids Res. 2009;37(3):731–7.
26. Sandhaus S, Annamalai T, Welmaker G, Houghten RA, Paz C, Garcia PK, et al. Small-
molecule inhibitors targeting topoisomerase I as novel antituberculosis agents. Antimicrob
Agents Chemother. 2016;60(7):4028–36.
27. Cheng B, Cao S, Vasquez V, Annamalai T, Tamayo-Castillo G, Clardy J, et al. Identification
of anziaic acid, a lichen depside from Hypotrachyna sp., as a new topoisomerase poison
inhibitor. PLoS One. 2013;8(4):e60770.
28. Aldred KJ, Kerns RJ, Osheroff N. Mechanism of quinolone action and resistance.
Biochemistry. 2014;53(10):1565–74.
29. Drlica K, Hiasa H, Kerns R, Malik M, Mustaev A, Zhao X. Quinolones: action and resistance
updated. Curr Top Med Chem. 2009;9(11):981–98.
30. Pommier Y, Leo E, Zhang H, Marchand C. DNA topoisomerases and their poisoning by anti-
cancer and antibacterial drugs. Chem Biol. 2010;17(5):421–33.
31. Bax BD, Chan PF, Eggleston DS, Fosberry A, Gentry DR, Gorrec F, et al. Type IIA topoisom-
erase inhibition by a new class of antibacterial agents. Nature. 2010;466(7309):935–40.
32. Laponogov I, Sohi MK, Veselkov DA, Pan X-S, Sawhney R, Thompson AW, et al. Structural
insight into the quinolone-DNA cleavage complex of type IIA topoisomerases. Nat Struct
Mol Biol. 2009;16(6):667–9.
33. Wohlkonig A, Chan PF, Fosberry AP, Homes P, Huang J, Kranz M, et al. Structural basis of
quinolone inhibition of type IIA topoisomerases and target-mediated resistance. Nat Struct
Mol Biol. 2010;17(9):1152–3.
34. Ellsworth EL, Tran TP, Showalter HD, Sanchez JP, Watson BM, Stier MA, et al.
3-aminoquinazolinediones as a new class of antibacterial agents demonstrating excellent
antibacterial activity against wild-type and multidrug resistant organisms. J Med Chem.
2006;49(22):6435–8.
35. Huband MD, Cohen MA, Zurack M, Hanna DL, Skerlos LA, Sulavik MC, et al. In vitro
and in vivo activities of PD 0305970 and PD 0326448, new bacterial gyrase/topoisomerase
inhibitors with potent antibacterial activities versus multidrug-resistant gram-positive and
fastidious organism groups. Antimicrob Agents Chemother. 2007;51(4):1191–201.
36. Tran TP, Ellsworth EL, Sanchez JP, Watson BM, Stier MA, Showalter HD, et al. Structure-
activity relationships of 3-aminoquinazolinediones, a new class of bacterial type-2 topoisom-
erase (DNA gyrase and topo IV) inhibitors. Bioorg Med Chem Lett. 2007;17(5):1312–20.
37. Germe T, Voros J, Jeannot F, Tailler T, Stavenger RA, Bacqué E, et al. A new class of antibac-
terials, the imidizopyrazinones, reveal structural transitions involved in DNA gyrase poison-
ing and mechanisms of resistance. Nucleic Acids Res. (in press).
38. Jeannot F, Taillier T, Despeyroux P, Renard S, Rey A, Mourez M, et al. Imidazopyrazinones
(IPYs): novel non-quinolone bacterial topoisomerase inhibitors showing partial cross-
resistance with quinolones. (in press).
39. Aldred KJ, McPherson SA, Wang P, Kerns RJ, Graves DE, Turnbough CL Jr, et al. Drug
interactions with Bacillus anthracis topoisomerase IV: biochemical basis for quinolone action
and resistance. Biochemistry. 2012;51(1):370–81.
40. Laponogov I, Pan XS, Veselkov DA, McAuley KE, Fisher LM, Sanderson MR. Structural
basis of gate-DNA breakage and resealing by type II topoisomerases. PLoS One.
2010;5(6):e11338.
41. Aldred KJ, Schwanz HA, Li G, McPherson SA, Turnbough CL Jr, Kerns RJ, et al.
Overcoming target-mediated quinolone resistance in topoisomerase IV by introducing metal-
ion-independent drug-enzyme interactions. ACS Chem Biol. 2013;8(12):2660–8.
42. German N, Malik M, Rosen JD, Drlica K, Kerns RJ. Use of gyrase resistance mutants to
guide selection of 8-methoxy-quinazoline-2,4-diones. Antimicrob Agents Chemother.
2008;52(11):3915–21.
19 Non-quinolone Topoisomerase Inhibitors 613
43. Oppegard LM, Streck KR, Rosen JD, Schwanz HA, Drlica K, Kerns RJ, et al. Comparison of
in vitro activities of fluoroquinolone-like 2,4- and 1,3-diones. Antimicrob Agents Chemother.
2010;54(7):3011–4.
44. Mustaev A, Malik M, Zhao X, Kurepina N, Luan G, Oppegard LM, et al. Fluoroquinolone-
gyrase-DNA complexes: two modes of drug binding. J Biol Chem. 2014.
45. Drlica K, Mustaev A, Towle TR, Luan G, Kerns RJ, Berger JM. Bypassing fluoroquinolone
resistance with quinazolinediones: studies of drug-gyrase-DNA complexes having implica-
tions for drug design. ACS Chem Biol. 2014;9(12):2895–904.
46. Chan PF, Srikannathasan V, Huang J, Cui H, Fosberry AP, Gu M, et al. Structural basis of
DNA gyrase inhibition by antibacterial QPT-1, anticancer drug etoposide and moxifloxacin.
Nat Commun. 2015;6:10048.
47. Biedenbach DJ, Bouchillon SK, Hackel M, Miller LA, Scangarella-Oman NE, Jakielaszek
C, et al. In vitro activity of gepotidacin, a novel triazaacenaphthylene bacterial topoisomerase
inhibitor, against a broad spectrum of bacterial pathogens. Antimicrob Agents Chemother.
2016;60(3):1918–23.
48. Chan PF, Germe T, Bax BD, Huang J, Thalji RK, Bacque E, et al. Thiophene antibacterials
that allosterically stabilize DNA-cleavage complexes with DNA gyrase. Proc Natl Acad Sci
U S A. 2017;114(22):E4492–E500.
49. Aldred KJ, Breland EJ, McPherson SA, Turnbough CL Jr, Kerns RJ, Osheroff N. Bacillus
anthracis GrlAV96A topoisomerase IV, a quinolone resistance mutation that does not affect
the water-metal ion bridge. Antimicrob Agents Chemother. 2014;58(12):7182–7.
50. Parks WM, Bottrill AR, Pierrat OA, Durrant MC, Maxwell A. The action of the bacterial
toxin, microcin B17, on DNA gyrase. Biochimie. 2007;89:500–7.
51. Heddle JG, Blance SJ, Zamble DB, Hollfelder F, Miller DA, Wentzell LM, et al. The antibi-
otic microcin B17 is a DNA gyrase poison: characterisation of the mode of inhibition. J Mol
Biol. 2001;307(5):1223–34.
52. Vizan JL, Hernandez-Chico C, del Castillo I, Moreno F. The peptide antibiotic microcin
B17 induces double-strand cleavage of DNA mediated by E. coli DNA gyrase. EMBO
J. 1991;10(2):467–76.
53. del Castillo FJ, del Castillo I, Moreno F. Construction and characterization of mutations at
codon 751 of the Escherichia coli gyrB gene that confer resistance to the antimicrobial pep-
tide microcin B17 and alter the activity of DNA gyrase. J Bacteriol. 2001;183(6):2137–40.
54. Thompson RE, Collin F, Maxwell A, Jolliffe KA, Payne RJ. Synthesis of full length
and truncated microcin B17 analogues as DNA gyrase poisons. Org Biomol Chem.
2014;12(10):1570–8.
55. Collin F, Thompson RE, Jolliffe KA, Payne RJ, Maxwell A. Fragments of the bacterial toxin
microcin b17 as gyrase poisons. PLoS One. 2013;8(4):e61459.
56. Thompson RE, Jolliffe KA, Payne RJ. Total synthesis of Microcin B17 via a fragment con-
densation approach. Org Lett. 2011;13(4):680–3.
57. Videnov G, Kaiser D, Brooks M, Jung G. Synthesis of the DNA gyrase inhibitor microcin
B17, a 43-peptie antibiotic with eight heterocycles in its backbone. Agnew Chem Int Ed Engl.
1996;35(13/14):1506–8.
58. Trovatti E, Cotrim CA, Garrido SS, Barros RS, Marchetto R. Peptides based on CcdB protein
as novel inhibitors of bacterial topoisomerases. Bioorg Med Chem Lett. 2008;18(23):6161–4.
59. Walsh C, Wencewicz T. Antibiotics challenges mechanisms opportunites. Washington, DC:
ASM Press; 2016. 477 p.
60. Heide L. The aminocoumarins: biosynthesis and biology. Nat Prod Rep. 2009;26(10):1241–50.
61. Maxwell A, Lawson DM. The ATP-binding site of type II topoisomerases as a target for
antibacterial drugs. Curr Top Med Chem. 2003;3(1):283–303.
62. Mizuuchi K, O’Dea MH, Gellert M. DNA gyrase: subunit structure and ATPase activity of
the purified enzyme. Proc Natl Acad Sci U S A. 1978;75:5960–3.
614 A. Maxwell et al.
102. Karkare S, Chung TT, Collin F, Mitchenall LA, McKay AR, Greive SJ, et al. The naphthoqui-
none diospyrin is an inhibitor of DNA gyrase with a novel mechanism of action. J Biol Chem.
2013;288(7):5149–56.
103. Phillips JW, Goetz MA, Smith SK, Zink DL, Polishook J, Onishi R, et al. Discovery of kibde-
lomycin, a potent new class of bacterial type II topoisomerase inhibitor by chemical-genetic
profiling in Staphylococcus aureus. Chem Biol. 2011;18(8):955–65.
104. Singh SB. Discovery and development of kibdelomycin, a new class of broad-spectrum anti-
biotics targeting the clinically proven bacterial type II topoisomerase. Bioorg Med Chem.
2016;24(24):6291–7.
105. Singh SB, Dayananth P, Balibar CJ, Garlisi CG, Lu J, Kishii R, et al. Kibdelomycin is a
bactericidal broad-spectrum aerobic antibacterial agent. Antimicrob Agents Chemother.
2015;59(6):3474–81.
106. Miesel L, Hecht DW, Osmolski JR, Gerding D, Flattery A, Li F, et al. Kibdelomycin is a
potent and selective agent against toxigenic Clostridium difficile. Antimicrob Agents
Chemother. 2014;58(4):2387–92.
107. Lu J, Patel S, Sharma N, Soisson SM, Kishii R, Takei M, et al. Structures of kibdelomycin
bound to Staphylococcus aureus GyrB and ParE showed a novel U-shaped binding mode.
ACS Chem Biol. 2014;9(9):2023–31.
108. Lincke T, Behnken S, Ishida K, Roth M, Hertweck C. Closthioamide: an unprecedented
polythioamide antibiotic from the strictly anaerobic bacterium Clostridium cellulolyticum.
Angew Chem Int Ed Engl. 2010;49(11):2011–3.
109. Chiriac AI, Kloss F, Kramer J, Vuong C, Hertweck C, Sahl HG. Mode of action of clos-
thioamide: the first member of the polythioamide class of bacterial DNA gyrase inhibitors.
J Antimicrob Chemother. 2015;70(9):2576–88.
110. Sawa R, Takahashi Y, Hashizume H, Sasaki K, Ishizaki Y, Umekita M, et al. Amycolamicin:
a novel broad-spectrum antibiotic inhibiting bacterial topoisomerase. Chemistry.
2012;18(49):15772–81.
111. Agrawal A, Roue M, Spitzfaden C, Petrella S, Aubry A, Hann M, et al. Mycobacterium
tuberculosis DNA gyrase ATPase domain structures suggest a dissociative mechanism that
explains how ATP hydrolysis is coupled to domain motion. Biochem J. 2013;456(2):263–73.
112. Hearnshaw SJ, Chung TT, Stevenson CE, Maxwell A, Lawson DM. The role of monova-
lent cations in the ATPase reaction of DNA gyrase. Acta Crystallogr D Biol Crystallogr.
2015;71(Pt 4):996–1005.
113. Stanger FV, Dehio C, Schirmer T. Structure of the N-terminal gyrase B fragment in com-
plex with ADPPi reveals rigid-body motion induced by ATP hydrolysis. PLoS One.
2014;9(9):e107289.
114. Lewis RJ, Singh OM, Smith CV, Maxwell A, Skarzynski T, Wonacott AJ, et al. Crystallization
of inhibitor complexes of an N-terminal 24 kDa fragment of the DNA gyrase B protein. J Mol
Biol. 1994;241(1):128–30.
115. Tsai FT, Singh OM, Skarzynski T, Wonacott AJ, Weston S, Tucker A, et al. The high-
resolution crystal structure of a 24-kDa gyrase B fragment from E. coli complexed with one
of the most potent coumarin inhibitors, clorobiocin. Proteins. 1997;28(1):41–52.
116. Bellon S, Parsons JD, Wei Y, Hayakawa K, Swenson LL, Charifson PS, et al. Crystal struc-
tures of Escherichia coli topoisomerase IV ParE subunit (24 and 43 kilodaltons): a single resi-
due dictates differences in novobiocin potency against topoisomerase IV and DNA gyrase.
Antimicrob Agents Chemother. 2004;48(5):1856–64.
117. Oblak M, Kotnik M, Solmajer T. Discovery and development of ATPase inhibitors of DNA
gyrase as antibacterial agents. Curr Med Chem. 2007;14(19):2033–47.
118. Garg G, Zhao H, Blagg BSJ. Design, synthesis, and biological evaluation of ring-constrained
Novobiocin analogues as Hsp90 C-terminal inhibitors. ACS Med Chem Lett. 2014.
119. Hall JA, Seedarala S, Zhao H, Garg G, Ghosh S, Blagg BS. Novobiocin analogues that inhibit
the MAPK pathway. J Med Chem. 2016;59(3):925–33.
19 Non-quinolone Topoisomerase Inhibitors 617
120. Kampranis SC, Bates AD, Maxwell A. A model for the mechanism of strand passage by DNA
gyrase. Proc Natl Acad Sci U S A. 1999;96(15):8414–9.
121. Morris SK, Baird CL, Lindsley JE. Steady-state and rapid kinetic analysis of topoisomerase
II trapped as the closed-clamp intermediate by ICRF-193. J Biol Chem. 2000;275(4):2613–8.
122. Ehmann DE, Lahiri SD. Novel compounds targeting bacterial DNA topoisomerase/DNA
gyrase. Curr Opin Pharmacol. 2014;18:76–83.
123. Boehm HJ, Boehringer M, Bur D, Gmuender H, Huber W, Klaus W, et al. Novel inhibitors
of DNA gyrase: 3D structure based biased needle screening, hit validation by biophysical
methods, and 3D guided optimization. A promising alternative to random screening. J Med
Chem. 2000;43(14):2664–74.
124. Mesleh MF, Cross JB, Zhang J, Kahmann J, Andersen OA, Barker J, et al. Fragment-based
discovery of DNA gyrase inhibitors targeting the ATPase subunit of GyrB. Bioorg Med Chem
Lett. 2016;26(4):1314–8.
125. Cross JB, Zhang J, Yang Q, Mesleh MF, Romero JA, Wang B, et al. Discovery of pyrazo-
lopyridones as a novel class of gyrase B inhibitors using structure guided design. ACS Med
Chem Lett. 2016;7(4):374–8.
126. Zhang J, Yang Q, Cross JB, Romero JA, Poutsiaka KM, Epie F, et al. Discovery of Azaindole
Ureas as a novel class of bacterial gyrase B inhibitors. J Med Chem. 2015;58(21):8503–12.
127. Eakin AE, Green O, Hales N, Walkup GK, Bist S, Singh A, et al. Pyrrolamide DNA gyrase
inhibitors: fragment-based nuclear magnetic resonance screening to identify antibacterial
agents. Antimicrob Agents Chemother. 2012;56(3):1240–6.
128. Jeankumar VU, Renuka J, Kotagiri S, Saxena S, Kakan SS, Sridevi JP, et al. Gyrase ATPase
domain as an antitubercular drug discovery platform: structure-based design and lead optimi-
zation of nitrothiazolyl carboxamide analogues. ChemMedChem. 2014;9(8):1850–9.
129. Jeankumar VU, Saxena S, Vats R, Reshma RS, Janupally R, Kulkarni P, et al. Structure-guided
discovery of antitubercular agents that target the gyrase ATPase domain. ChemMedChem.
2016;11(5):539–48.
130. Jeankumar VU, Reshma RS, Vats R, Janupally R, Saxena S, Yogeeswari P, et al. Engineering
another class of anti-tubercular lead: hit to lead optimization of an intriguing class of gyrase
ATPase inhibitors. Eur J Med Chem. 2016;122:216–31.
131. Saxena S, Samala G, Renuka J, Sridevi JP, Yogeeswari P, Sriram D. Development of 2-amino-
5-phenylthiophene-3-carboxamide derivatives as novel inhibitors of Mycobacterium tubercu-
losis DNA GyrB domain. Bioorg Med Chem. 2015;23(7):1402–12.
132. Gjorgjieva M, Tomasic T, Barancokova M, Katsamakas S, Ilas J, Tammela P, et al. Discovery of
benzothiazole scaffold-based DNA gyrase B inhibitors. J Med Chem. 2016;59(19):8941–54.
133. Jakopin Z, Ilas J, Barancokova M, Brvar M, Tammela P, Sollner Dolenc M, et al. Discovery
of substituted oxadiazoles as a novel scaffold for DNA gyrase inhibitors. Eur J Med Chem.
2017;130:171–84.
134. Sun J, Lv PC, Yin Y, Yuan RJ, Ma J, Zhu HL. Synthesis, structure and antibacterial activ-
ity of potent DNA gyrase inhibitors: N'-benzoyl-3-(4-bromophenyl)-1H-pyrazole-5-
carbohydrazide derivatives. PLoS One. 2013;8(7):e69751.
135. O'Brien S. Meeting the societal need for new antibiotics: the challenges for the pharmaceuti-
cal industry. Br J Clin Pharmacol. 2015;79(2):168–72.
136. Bax R, Green S. Antibiotics: the changing regulatory and pharmaceutical industry paradigm.
J Antimicrob Chemother. 2015;70(5):1281–4.
137. Silver LL. Challenges of antibacterial discovery. Clin Microbiol Rev. 2011;24(1):71–109.
138. Cooper MA, Shlaes D. Fix the antibiotics pipeline. Nature. 2011;472(7341):32.
139. Payne DJ, Miller LF, Findlay D, Anderson J, Marks L. Time for a change: addressing R&D
and commercialization challenges for antibacterials. Philos Trans R Soc Lond Ser B Biol Sci.
2015;370(1670):20140086.
140. Kostyanev T, Bonten MJ, O'Brien S, Steel H, Ross S, Francois B, et al. The Innovative
Medicines Initiative's New Drugs for Bad Bugs programme: European public-private part-
618 A. Maxwell et al.
nerships for the development of new strategies to tackle antibiotic resistance. J Antimicrob
Chemother. 2016;71(2):290–5.
141. Candel FJ, Penuelas M. Delafloxacin: design, development and potential place in therapy.
Drug Des Devel Ther. 2017;11:881–91.
142. Basarab GS, Brassil P, Doig P, Galullo V, Haimes HB, Kern G, et al. Novel DNA gyrase
inhibiting spiropyrimidinetriones with a benzisoxazole scaffold: SAR and in vivo character-
ization. J Med Chem. 2014;57(21):9078–95.
143. Charifson PS, Grillot AL, Grossman TH, Parsons JD, Badia M, Bellon S, et al. Novel
dual-targeting benzimidazole urea inhibitors of DNA gyrase and topoisomerase IV pos-
sessing potent antibacterial activity: intelligent design and evolution through the judi-
cious use of structure-guided design and structure-activity relationships. J Med Chem.
2008;51(17):5243–63.
Chapter 20
Antimicrobial-Mediated Bacterial Suicide
20.1 Introduction
Controlling antibiotic resistance can be considered from two perspectives: limit the
emergence of new resistance and halt the dissemination/horizontal transfer of exist-
ing resistance. For both applications, we expect better results from compounds that
rapidly reduce bacterial burden. A lower pathogen burden will then reduce our reli-
ance on host immune responses, which are likely to decline as populations age, to
clear infection. Lower burden will also help control pathogens as we increase our
use of immunosuppressants. Thus, rapid killing by antimicrobials will become
increasingly important. The present chapter addresses a common mechanism of
rapid killing by focusing on the hypothesis that bacteria respond to severe stress by
accumulating toxic reactive oxygen species (ROS) and thereby self-destruct.
Y. Hong · K. Drlica
Public Health Research Institute, New Jersey Medical School, Rutgers Biomedical
and Health Sciences, Newark, NJ, USA
X. Zhao (*)
Public Health Research Institute, New Jersey Medical School, Rutgers Biomedical
and Health Sciences, Newark, NJ, USA
Department of Microbiology, Biochemistry, & Molecular Genetics, New Jersey Medical
School, Rutgers Biomedical and Health Sciences, Newark, NJ, USA
State Key Laboratory of Molecular Vaccinology and Molecular Diagnostics,
School of Public Health, Xiamen University, Xiamen, Fujian Province, China
e-mail: [email protected]
Understanding this process may help us better control bacterial populations. Readers
interested in earlier reviews on ROS and on programmed cell death are referred to
references [1–4].
We begin by distinguishing between antimicrobial lethality and blocking growth,
the usual measure of antimicrobial activity, because the terms are frequently mis-
used. Experimentally, blocking growth is measured with antimicrobial in the test
medium. The output is expressed as either minimal inhibitory concentration (MIC)
or as efficiency of plating. MIC is currently the basis for most antimicrobial consid-
erations, including diagnosis, resistance surveillance, new compound development,
and dosing. In contrast, killing is measured as survival after treatment with antimi-
crobial, usually by scoring colony formation on drug-free agar. Input cells that can-
not form colonies after removal of drug are commonly considered to be dead.
Several variations exist for measuring killing. One is to determine the concentra-
tion of drug that reduces survival by a particular amount following a long incuba-
tion, often overnight for rapidly growing bacteria. This measurement, when
performed under standardized conditions, is called minimal bactericidal concentra-
tion (MBC). Many secondary events can occur during the long incubation required
for MBC measurement, which makes it difficult to characterize direct, specific
lethal mechanisms using MBC. In contrast, rapid killing, which determines rate and
extent of killing occurring within a few hours of drug exposure, is measured to study
mechanism and find ways to improve the direct lethal activity of antimicrobials.
Work discussed in the present chapter focuses on rapid killing.
A key concept is that rapid antimicrobial-mediated killing occurs in two general
ways: (1) the primary lesion is sufficient to directly kill the cell, and (2) the primary
lesion induces a lethal, self-destructive stress response (see Fig. 20.1). The contribu-
tion of ROS to lethal activity falls in the second category [3]. Since the relevance of
ROS action relies on cell death, we begin with a brief discussion of how the impor-
tance of lethality is growing as the prevalence of antimicrobial resistance increases.
Traditionally, lethal action has been thought to be important mostly for curing indi-
vidual infections, particularly for problematic diseases such as endocarditis and
meningitis. Dosing decisions have been based on relationships between incubation
conditions and killing, largely using measurements of MBC [5]. Compounds fall
into two general groups. Members of one group, represented by fluoroquinolones
and aminoglycosides, are said to be concentration-dependent killers because
increases in drug concentration result in increased killing. Members of the other,
which includes β-lactams and macrolides, are time-dependent killers – as long as
antimicrobial concentration exceeds MIC by a multiple of 2–5, further increases in
drug concentration have little effect on killing [6]. According to this approach, some
antimicrobials, in particular fluoroquinolones, are thought to be lethal enough to
cure most infections. That perspective questions the need for improving lethality.
20 Antimicrobial-Mediated Bacterial Suicide 621
Fig. 20.1 Relationships among resistance, tolerance, and killing. (a) Antimicrobials create pri-
mary damage that is specific to the compound class. The resulting lesions block growth; the effect
is quantified by the minimal inhibitory concentration (MIC). Resistance interferes with the forma-
tion of the primary lesion. (b) When the initial lesion is sufficiently damaging, a series of events
occur that lead to accumulation of ROS; those events can be perturbed in a variety of ways. (c)
When an ROS threshold is passed, the reactive species cause secondary damage that can then elicit
the accumulation of even more ROS (d) and cell death (e). (f) Some antimicrobials produce lesions
that are lethal independent of ROS production. Bacterial tolerance (g) occurs when lethal action is
inhibited even though the primary lesion occurs and growth is still blocked. Tolerant cells may
suppress lethal pathway (e) by inhibiting respiration and thereby ROS production/accumulation.
Whether tolerance also arises from blockage of lethal pathway (f) is likely compound-specific
least susceptible mutant subpopulation, resistance is less likely to emerge than if the
goal is only to keep concentrations above the MIC of the major portion of the popu-
lation [10–12]. Unfortunately, with most pathogen-antimicrobial combinations,
keeping antimicrobial concentrations high enough and exposure time long enough
to restrict the emergence of resistance is likely to have adverse effects on patients.
Thus, it is encouraging that studies with animal models show that some highly lethal
compounds restrict the emergence of resistance without needing to keep concentra-
tions above the MIC of mutant subpopulations throughout treatment [13–16].
Lethal activity is also an important consideration with drug-tolerant cells (see
Chap. 13 for discussion of tolerance). Such bacteria, when exposed to highly lethal
antimicrobials, fail to grow, but they are not killed (in contrast, resistant cells grow
in the presence of drug; see Fig. 20.1 for relationships). Tolerant cells (persisters),
which are usually a small fraction of the population, are a problem because they
may survive treatment and subsequently cause relapse. Since many antimicrobials
are active primarily with replicating cells, non-growing cells can display a form of
tolerance, as seen with ampicillin and first-generation quinolones. Antibiotic toler-
ance can also arise from mutation, and it can facilitate the emergence of resistance
[17]. Thus, finding ways to overcome tolerance, i.e., weakened lethal activity, is of
major importance for restricting the emergence of resistance.
Key evidence for the contribution of ROS to lethal antimicrobial action derives from
manipulating ROS levels and correlating those changes with bacterial survival. One
type of genetic approach involves increasing or decreasing the expression of cata-
lase/peroxidase. For example, deletion of katG increases ROS accumulation and
lethality arising from treatment of E. coli with quinolones [18–20], from thymine
starvation [21], and from exposure to UV irradiation (Y. Hong & X. Zhao, unpub-
lished observation). Similarly, deletion of ahpCF, which encodes a peroxidase,
increases the lethal activity of kanamycin and ampicillin [18]. In a reciprocal exper-
iment, overexpression of katG suppresses antimicrobial lethality [20].
Many other genes that protect from oxidative stress have more complex effects.
Among these are sodA and sodB, genes that encode superoxide dismutases (enzymes
that convert superoxide to peroxide). A deficiency of either gene has little effect on
the lethal activity of norfloxacin [18]; however, a sodA sodB double mutant protects
from the lethal stress arising from norfloxacin, ampicillin, and kanamycin, as if
elevation of superoxide concentration is protective (wild-type genes would reduce
superoxide levels). A similar conclusion emerged from a study of the DNA-
damaging agent bleomycin [22]. In this case, a sodA sodB double mutant exhibited
reduced killing from bleomycin, while overexpression of superoxide dismutase
20 Antimicrobial-Mediated Bacterial Suicide 623
One chemical approach involves treating bacterial cells with antioxidants, such as
thiourea, dimethyl sulfoxide, ascorbic acid, glutathione, or resveratrol. Such com-
pounds are expected to scavenge hydroxyl radical. With chemical perturbation, an
important consideration is the concentration of the perturbing agent, since that con-
centration needs to be adjusted to avoid interfering with bacterial growth – growth
inhibition can create a type of tolerance. Another consideration is whether antioxi-
dants actually scavenge hydroxyl radical, since the rapid reaction of hydroxyl radi-
cal with other biomolecules might obscure any reaction with thiourea [24]. That
624 Y. Hong et al.
would make the observed protection from killing by thiourea [25] an off-target
effect. To our knowledge, no off-target effect of antioxidants has been identified.
Another chemical perturbation uses iron chelators, such as 2,2′ bipyridyl. These
agents are expected to act in two ways. One is by inhibiting the iron-requiring
Fenton reaction in which hydrogen peroxide produces hydroxyl radical. A second
involves iron-requiring proteins involved in respiration, a source of superoxide and
ultimately hydroxyl radical. Experiments employing bipyridyl have consistently
supported the conclusion that lethal activity arising from a variety of stressors
involves ROS [18, 25].
Ruling out off-target effects due to chelators and antioxidants is difficult. One
argument is that antioxidants that interfere with antimicrobial lethality are chemi-
cally diverse and unlikely to have the same off-target effect. Another argument
derives from consideration of catalase as an antioxidant. We found that E. coli cells,
thought to be killed by the lethal antimicrobial trimethoprim and then plated on
drug-free agar, are revived by addition of catalase to the plated cells (Y. Hong and
X. Zhao, unpublished observations). Since hydrogen peroxide readily enters and
exits from cells, extracellular catalase would lower the overall peroxide concentra-
tion and provide a specific protective effect that would not be attributed to either
off-target effects or growth inhibition. Overall, effects of chelators and antioxidants
fit well with their suppression of ROS accumulation.
Credit for the general nature of the ROS-lethality hypothesis is usually attributed to
a 2007 report from the Collins laboratory [25], although involvement of ROS with
lethal activity had been proposed earlier. For example, activation of the SoxRS regu-
lon conferred resistance to multiple antimicrobial classes [27–29]; moreover, anti-
oxidants, such as vitamin C and glutathione, raised minimal inhibitory concentration
(MIC) and efficiency of plating for quinolones and aminoglycosides [30, 31]. In
addition, elevated levels of oxidative stress signals were detected in cells treated
with antimicrobials [32, 33]. However, such observations do not establish a connec-
tion between ROS and cell death, because the measurements reported inhibition of
bacterial growth, not cell death. Thus, the work by Kohanski et al. [25] was a quali-
tative advance because killing was measured. We later showed that the killing was
separate from bacteriostatic effects by normalizing killing to MIC [18].
Among the central observations from the Kohanski et al. study [25] was that kill-
ing caused by norfloxacin, ampicillin, or kanamycin was suppressed by thiourea
and 2,2′ bipyridyl. Although the experimental conditions also suppressed bacterial
growth, which itself was known to interfere with killing, subsequent work [18]
showed that subinhibitory concentrations of thiourea and bipyridyl protect from
20 Antimicrobial-Mediated Bacterial Suicide 625
antimicrobial lethality. The latter study also described genetic perturbations that
supported the hypothesis [18].
It soon became clear that not all derivatives of an antimicrobial class depend on
ROS to kill cells [34]. For example, the quinolones were known to kill by two gen-
eral pathways, one that was blocked by chloramphenicol, an inhibitor of protein
synthesis, and one that was not [35–37]. This dichotomy also applied to the contri-
bution of ROS: 2,2′ bipyridyl treatment [34] and anaerobic conditions [38] paral-
leled the behavior of chloramphenicol. Overall, the quinolone experiments
established that first-generation compounds, such as nalidixic and oxolinic acids,
kill E. coli by a mechanism that relies heavily on ROS; the fluoroquinolones addi-
tionally kill bacteria by a pathway that appears to rely more on chromosome frag-
mentation than on ROS, although fluoroquinolones still trigger ROS accumulation
[34]. Within this scheme, norfloxacin is an outlier having an intermediate,
concentration-dependent activity [38]. As pointed out below in Sect. 5, subsequent
challenges to the ROS-lethality hypothesis relied in part on work with norfloxacin.
Steps leading from stress to accumulation of ROS have been investigated by identi-
fying genes that, when defective, alter ROS levels and stress-mediated cell death.
The action of the gene products can be fit into a scheme (Fig. 20.2). One of the fac-
tors is the MazEF toxin-antitoxin pair. The MazF protein is an endoribonuclease
that during stress cleaves mRNA and thereby blocks protein synthesis. At low, bac-
teriostatic concentrations of antimicrobial, blocking gene expression and bacterial
growth by MazF would allow time for cells to efflux noxious molecules and repair
damage. However, at high, lethal levels of stress, MazF would produce toxic levels
of truncated mRNA and misfolded proteins that perturb cell membrane function and
elevate ROS levels. Thus, MazF is expected to be bifunctional with respect to stress.
Indeed, bifunctionality has been observed with Bacillus subtilis. For example, at
low doses of UV irradiation or low concentrations of moxifloxacin, a ΔndoA (mazF)
mutant is more readily killed than wild-type cells, but at high doses the opposite is
seen [39]. Bifunctionality is a key feature of a stress response that either allows
repair of minor damage or causes self-destruction when damage is severe.
EF4 is another factor that contributes to ROS accumulation. This ribosomal elon-
gation protein is normally sequestered in the cell membrane, but during stress, the
protein enters the cytosol, binds to stress-stalled ribosomes, and stimulates ribo-
somal back-translation. These events allow protein synthesis to recover from mod-
erate stress. But EF4 also blocks tagging of truncated proteins for degradation by
tmRNA – this activity would facilitate the accumulation of toxic peptides, presum-
ably arising from MazF action. Indeed, the absence of EF4 protects E. coli from
being killed by quinolones (wild-type protein would be destructive) [40]. Thus, EF4
and MazEF appear to be bifunctional proteins that protect from moderate stress but
promote death when stress is high. How the action of MazF and EF4 is connected
to the initial lesion is unknown.
626 Y. Hong et al.
Fig. 20.2 Scheme describing bifunctional nature of factors involved in the live-or-die stress-
response pathway. Low to moderate levels of stress result in a protective stress response. The stress-
ors, such as quinolones, generate specific primary lesions that in an unknown way stimulate the
MazF toxin to cleave mRNA, thereby halting translation and allowing cells time to repair damage.
The protein fragments resulting from translation or MazF-mediated mRNA cleavage are tagged for
degradation by tmRNA and EF4, a ribosomal protein that facililtates restart of stalled ribosomes.
Truncated, misfolded peptides that enter cell membranes stimulate the Cpx two-component system
to induce genes involved in membrane protein repair. Safety valves YihE kinase and KatG catalase
negatively regulate MazF and detoxify peroxide, respectively. At high stress levels, mRNA cleavage
by MazF is extensive, and EF4 blocks the tagging of truncated peptides by tmRNA. Thus, high
levels of protein fragments accumulate and enter cell membranes. That causes the Cpx system to
induce the Arc two-component system, which in turn leads to high-level production of superoxide.
Superoxide dismutates to peroxide; the Fenton reaction then converts peroxide to hydroxyl radical.
Hydroxyl radical damages nucleotides and many macromolecule types, causing mutations and cell
death. To assure death, MazF cleaves katG mRNA, which lowers the level of KatG, a protein that
would otherwise reduce peroxide levels. Other protective functions, such as induction of membrane
repair by Cpx/Arc and induction of the SoxRS/OxyRS regulon by superoxide/peroxide, are
overwhelmed
Insertion of truncated, misfolded proteins into the cell membrane activates a two-
component membrane stress-response system called Cpx [41, 42]. When stress is
high, activation of Cpx stimulates another two-component system (Arc) that then
contributes to the generation of elevated levels of ROS [43] as discussed below in
Sect. 6. The destructive feature of Cpx is revealed by a CpxR deficiency protecting
20 Antimicrobial-Mediated Bacterial Suicide 627
from nalidixic acid-mediated cell death [19]. The Cpx system also serves to repair
membrane protein damage [44] and to mitigate MazF toxicity [19] when stress is
moderate. Thus, Cpx has both destructive and protective functions; it represents a
third bifunctional system involved in lethal antimicrobial action. Arc also exhibits
protective and destructive properties (arc is discussed in more detail in Sect. 6.2).
Superoxide occupies a central position in the lethal stress response. At moderate
stress levels, superoxide is thought to accumulate and stimulate protective gene
responses, such as induction of the SoxRS regulon. During periods of harsh stress,
superoxide may accumulate to high levels and rapidly dismutate, thereby creating
elevated levels of hydrogen peroxide (superoxide dismutases are 1000-fold more
efficient than catalase/peroxidase, enzymes that detoxify hydrogen peroxide by con-
verting it to water). The result is accumulation of hydrogen peroxide, which can
then be converted to hydroxyl radical, a compound whose toxic effects may over-
whelm the protective functions stimulated by superoxide. As pointed out above,
such a bifunctional nature of superoxide would explain why low concentrations of
metabolic generators of superoxide, such as plumbagin and paraquat, reduce the
lethal action of bleomycin and the quinolones [22, 23, 45], even though high con-
centrations of plumbagin kill bacteria.
Several safety valves operate within this scheme of bifunctional factors. One
involves a protein kinase called YihE. The absence of YihE elevates the lethal action
of nalidixic acid by 100-fold but only if the MazEF toxin is present [19]. Thus, YihE
appears to be a negative regulator of MazF. Another safety valve is the katG cata-
lase/peroxidase, which, as mentioned above, detoxifies peroxide by converting it to
water. Deletion of katG increases norfloxacin lethality by 20-fold without affecting
MIC [18]. Similarly, deletion of ahpCF increases lethality of both ampicillin and
kanamycin, although this deletion has no effect on quinolone-mediated killing.
Among the many protective genes (safety valves) induced by oxidative stress are
efflux pumps that remove noxious stressors [23, 45]; presumably the pumps reduce
the signal leading to ROS accumulation.
Variations on the theme developed above are likely to emerge as more is learned
about individual stressors. For example, each stressor generates a unique lesion that
is likely to be recognized in a unique way, and many pathways may connect the
lesion to the accumulation of ROS. Thus, the scheme in Fig. 20.2 should be consid-
ered only as a framework for future testing.
The most toxic of the reactive oxygen species is hydroxyl radical. It readily dam-
ages DNA by creating single-strand breaks and by converting single-strand DNA
lesions into double-strand breaks [46]. Hydroxyl radical also carbonylates proteins
[47] and peroxidates membrane lipids [48]. In addition to directly breaking DNA,
hydroxyl radical oxidizes the guanine nucleotide pool, thereby producing 8-oxo-
guanine. When 8-oxo-guanine is incorporated into DNA, it is expected to lead to
double-strand breaks. The DNA breaks are thought to derive from excision of the
628 Y. Hong et al.
8-oxo-dG, since deficiencies in the excision enzymes protect (10–20 fold) from the
lethal action of norfloxacin, ampicillin, and kanamycin [20, 49]. Overexpression of
mutT, which “sanitizes” 8-oxo-guanine from the nucleotide pool, reduces the rate of
killing for the three antimicrobial classes [49]. Thus, lethal DNA damage is a com-
mon theme associated with killing, even by antimicrobials that do not have DNA as
their primary target. That explains why ampicillin induces the SOS response [7].
DNA damage leads to the interesting possibility that ROS is self-amplifying.
When a threshold concentration of ROS is reached, DNA damage could be suffi-
cient to stimulate the production of even more ROS. Like a nuclear reaction, ROS
accumulation would be unstoppable, and death would be assured. In this sense,
bacterial cells self-destruct when faced with severe lethal stress [19]. To our knowl-
edge, stress-mediated ROS accumulation due to self-amplification has not been
demonstrated. One prediction from the self-amplification idea is that primary stress
of one type, such as DNA damage by a quinolone, will cause ROS-mediated dam-
age of a second type, such as protein carbonylation. Repair of protein damage, if
specifically blocked by a lon or hslV deficiency, would exacerbate damage of a third
type, such as lipid peroxidation. Our unpublished work supports this scenario
(X. Wang & X. Zhao).
The initial proposal for ROS-mediated antimicrobial lethality [25] required modifi-
cation with respect to sod effects (Sect. 3.1), reconsideration of mutations affecting
iron-sulfur proteins (Sect. 3.1), and complexities with respect to fluoroquinolones
having at least two lethal mechanisms [34, 38]. However, the overarching idea was
independently solidified [18, 19, 23]. At about the same time, two laboratories chal-
lenged the idea that ROS contribute to the lethal action of multiple antimicrobial
classes. Below we discuss the resulting controversy.
One of the challenging studies [24] reported failure to confirm ROS involvement
in the lethal action of norfloxacin, kanamycin, and ampicillin under conditions that
were similar to those reported by Kohanski et al. [25]. Those conditions, which
involved use of a single drug concentration, appear to have been too narrowly
defined to take into account differences in conditions between laboratories. For
example, norfloxacin concentration was known to be important for observing lethal-
ity [38], so it needed to be varied. In the other work, Keren et al. [50] focused on
whether killing depends on ROS (requires ROS). By examining high levels of stress,
they confirmed that fluoroquinolones have an ROS-independent mode of killing [34,
38] that is insensitive to anaerobiosis [38]. The issue of whether ROS contribute to
killing, as proposed by Collins [25] and extended by our work [18], was actually
supported by Keren et al. [50] at low levels of stress. The high levels examined were
outside the range of discrimination: only residual “persisters” survived (such cells
20 Antimicrobial-Mediated Bacterial Suicide 629
The respiratory chain is a major source of superoxide and hydrogen peroxide [54,
55], as these ROS form when molecular oxygen oxidizes redox enzymes, such as
fumarate reductase (Frd), succinate dehydrogenase (Sdh), and aspartate oxidase
(NadB) [56]. Indeed, about a quarter of the cytoplasmic H2O2 derives from NadB
[57]. The electrons involved in formation of superoxide and hydrogen peroxide are
thought to derive largely from the tricarboxylic acid (TCA) cycle, which we discuss
below. In principle, dismutation of superoxide can also serve as a source of H2O2,
which can then form highly toxic hydroxyl radical via the Fenton reaction.
TCA cycle. Several lines of evidence connect the TCA cycle with the lethal action
of antibiotics. One is that treatment of E. coli cultures with bactericidal antibiotics
leads to upregulation of genes involved in central metabolism, including the TCA
cycle and respiration [25, 43, 58]. A second line links the efficacy of bactericidal
antibiotic therapy to carbon flux through the TCA cycle [59, 60]. A third line showed
that mutations in genes involved in the TCA cycle reduce stress-mediated lethality.
Surprisingly, antibiotic classes, represented by norfloxacin, ampicillin, and kana-
mycin, differ in the TCA cycle genes involved. For example, norfloxacin-mediated
lethality is reduced by mutation of icdA and acnB, while killing by ampicillin is
reduced by mutation of these two genes (icdA and acnB) and sucB. Killing by kana-
mycin is reduced by mutation of four genes, icdA, acnB, sucB, and mdh [25].
Additional work is needed to determine whether these differences are characteristic
of drug mechanism or whether they result from differences in the concentration of
the three drugs examined (normalized to MIC).
Respiration. E. coli encodes three cytochrome terminal oxidases: bd-I (CydAB),
bd-II (AppAB), and bo (CyoABCDE). Reduction of stress-mediated lethality by a
deficiency of cydB, which is associated with decreased ROS levels during hydroxy-
urea treatment [61] and thymineless death [21], establishes the importance of the
respiratory chain. Moreover, the rate of oxygen consumption serves as a measure of
respiration that can be compared with antimicrobial-mediated cell death [20, 62].
For example, a deficiency of atpA increases oxygen consumption, ROS level, and
killing during treatment of cells with ampicillin or norfloxacin [62]. Conversely, a
cytochrome oxidase null mutant fails to show an acceleration of respiration or cell
death when treated with norfloxacin, ampicillin, or gentamicin [62]. Treatment of
wild-type cells with the bacteriostatic agent chloramphenicol rapidly attenuates cel-
lular respiration and the lethality associated with lethal doses of norfloxacin or
ampicillin [62]. In yet another example, NADH-coupled electron transport (NADH
dehydrogenase I) is a common upregulated pathway for all three bactericidal drugs
[25]. Such data fit with upregulation of genes involved in central metabolism and
respiration being associated with exposure to lethal antimicrobials [25, 43].
20 Antimicrobial-Mediated Bacterial Suicide 631
Complexities of drug uptake and killing. We have proposed that ROS contributes to
the lethal activity of aminoglycosides (kanamycin), because lethal action, when
normalized to MIC to rule out effects of drug uptake, efflux, and target interactions,
was decreased by mutation of superoxide dismutase genes and increased by a defi-
ciency in a catalase/peroxidase [18]. Other works implicated the TCA cycle in ami-
noglycoside lethality [25, 43], and recently fumarate or glyoxylate, used to activate
or inhibit the TCA cycle, potentiated or suppressed, respectively, the lethal activity
of tobramycin with Pseudomonas aeruginosa [63]. Complexity arises because
proton-motive force, which derives from respiration and is necessary for aminogly-
coside uptake [26], is a required precursor to lethal activity. Moreover, uptake and
killing are stimulated by the same factor (e.g., respiration). Thus, normalization to
MIC is required to remove uptake from consideration before the effect of respiration
on ROS and killing can be assessed.
out in an earlier section, Cpx protects from low levels of lethal stress and contributes
to death when stress is high [19]. Indeed, Cpx may contribute to the accumulation
of ROS by stimulating Arc to accelerate the TCA cycle [43]. We reiterate the gen-
eral theme in which a set of proteins (MazEF, EF-4, Cpx, and ArcAB) protect from
moderate levels of lethal stress, but when stress is severe, the proteins contribute to
the accumulation of ROS and cell death.
lation would benefit other members [69]. Nevertheless, the existence of bacterial
PCD is supported by studies of sporulation and by detection of apoptosis biomark-
ers previously reported for eukaryotic systems (with antibiotic stress, E. coli cells
exhibit DNA fragmentation, chromosome condensation, and membrane depolariza-
tion [70]). Among the examples of bacterial development are mother cell lysis dur-
ing spore formation, which is seen with Bacillus subtilis and Myxococcus xanthus
(as reviewed in [4]). With Xanthomonas campestris, apoptosis-like phenotypes, as
mentioned above, are seen during death caused by incubation in rich Luria-Bertani
(LB) medium [71, 72]. However, in none of these examples have cells been shown
to continue along a documented death pathway after withdrawal of the primary
stressor – in each case, the primary stressor was present throughout the experiment.
Thus, data showing that lethal stress propels bacteria along an ROS-requiring death
pathway, described above, add key support to the idea that bacteria undergo a self-
destructive process in response to stress.
A study of thymineless death provides insight into one way in which ROS kill bac-
teria. Thymineless death refers to the rapid loss of bacterial viability that occurs
during starvation for thymine or thymidine. The phenomenon is of considerable
interest, because the underlying molecular events also apply to the action of several
antibacterial (trimethoprim, sulfamethoxazole), antimalarial (pyrimethamine, sul-
fonamide), anticancer (methotrexate, fluorouracil), and immune-modulating (meth-
otrexate) agents. Many explanations have been proposed to explain thymineless
death. Among these are unbalanced growth, toxin-antitoxin module action, nucleo-
tide misincorporation, induction of the SOS regulon, destruction of replication
forks, and partial degradation of DNA at oriC [75, 76]. None of these explanations
have satisfied all of the experimental observations. Recent attention has focused on
proteins involved in recombinational repair, as some appear to contribute to death
while others facilitate survival [77–80].
Since DNA replication is severely slowed by withdrawal of thymidine and since
other means of inhibiting DNA replication (fluoroquinolone or hydroxyurea treat-
ment) cause the accumulation of ROS, we examined the possibility that ROS con-
tribute to thymineless death. Two processes appear to be required for rapid death:
generation of persistent single-strand DNA regions and accumulation of ROS [21].
The attack of single-strand DNA regions by ROS then leads to lethal d ouble-stranded
DNA breaks. Interference with either production of single-strand DNA or accumu-
lation of ROS inhibits thymineless death. Our current view is that proteins involved
in recombinational DNA repair, such as RecF and RecQ, expand the single-strand
DNA substrate for ROS attack (the absence of these proteins reduces thymineless
death). Other DNA repair proteins, such as RecBC, are involved in surviving the
damage, and the SOS response increases expression of SulA, which serves as a
checkpoint that halts cell division until the damage to DNA is repaired in cells that
do not die [77].
20 Antimicrobial-Mediated Bacterial Suicide 635
Part of the evidence supporting a role for ROS in antimicrobial lethality stems from
the protective effects of antioxidants. For example, thiourea, glutathione, and vita-
min C reduce fluoroquinolone lethality by orders of magnitude [18, 25, 51, 81];
thiourea and glutathione reduce lethality of daptomycin and oxacillin by 10–100-
fold [51]. Since human consumption of antioxidants is large [82], the potential for
interference with antimicrobial action exists. One aspect may involve food prod-
ucts. In a recent study, Marathe et al. [83] examined effects of the antioxidant cur-
cumin on ciprofloxacin-mediated lethality with Salmonella. Curcumin is a common
food ingredient in Southeast Asia, and it is often used medicinally. Curcumin
reduces the lethal activity of ciprofloxacin with cultured bacteria, with bacteria
infecting macrophages, and with bacteria infecting mice. Since curcumin also sup-
presses the antibacterial activity of the immune response, it may act in two ways to
increase bacterial survival [83]. Thus, a cautionary note has been raised concerning
antibiotic therapy and consumption of foods having antioxidant activity [83].
Antioxidants are also consumed as nutritional supplements. One of the popular
agents is resveratrol, a natural polyphenol antioxidant [81]. Resveratrol is thought
to have beneficial effects for ailments such as cardiovascular disease [84], neurode-
generative disorders [84, 85], and some forms of cancer [84, 86]. To address whether
antioxidant nutritional supplements are likely to interfere with the ability of antimi-
crobials to kill bacteria, we added resveratrol to cultures of E. coli and S. aureus,
treated them with antimicrobial, and assayed for bacterial survival [87]. Resveratrol,
at concentrations likely to be present during human consumption, reduced killing by
two- to threefold during a 2-h exposure to moxifloxacin or kanamycin. At higher but
still subinhibitory concentrations, resveratrol lowered antimicrobial lethality by
more than 1000-fold. Resveratrol also reduced the accumulation of ROS
characteristic of treatment with oxolinic acid, a first-generation quinolone.
Collectively these observations support the general idea that the lethal activity of
some antimicrobials involves ROS.
Subinhibitory concentrations of resveratrol also promoted (two- to sixfold) the
recovery of rifampicin-resistant mutants arising from the action of ciprofloxacin,
kanamycin, or daptomycin. This finding can be explained by resveratrol lowering
ROS to sublethal levels that are still mutagenic, while the absence of resveratrol
allows ROS levels to be high enough to kill mutagenized cells. Suppression of kill-
636 Y. Hong et al.
During infection, bacteria experience stress from host defense systems. One conse-
quence may be preparation of the pathogen for antimicrobial-mediated stress. For
example, host-derived reactive oxygen or reactive nitrogen species may induce bac-
terial protective systems that would counter bacterial production of ROS induced by
antimicrobials. According to the scheme presented in Fig. 20.1, suppression of
ROS-mediated killing would constitute a form of bacterial tolerance.
Heritable, genetic changes that increase antimicrobial tolerance can be selected
both in vitro and in vivo. Indeed, genome-wide maps of tolerance-associated genes,
the “tolerome,” have been described [88] that include numerous genes and pathways
20 Antimicrobial-Mediated Bacterial Suicide 637
that slow bacterial growth. Particularly interesting are mutations in global transcrip-
tional regulators, because they can produce a rapid, pleiotropic phenotypic effect:
they may constitute a prominent mechanism underlying tolerance. A clinically rel-
evant example of a global regulator of tolerance is seen with S. aureus mutants that
are deficient in the quorum-sensing agr regulon (see also Chap. 14). Paradoxically,
defects in this virulence factor are selected during serious hospital infection and are
associated with worse outcome [89, 90]. Recent work indicates that Δagr mutants
are less readily killed by gentamicin, ciprofloxacin, and several other lethal stressors
without affecting MIC [91], thereby conforming to a tolerance phenotype. Wild-
type agr normally downregulates a peroxidase that would otherwise protect from
ROS-mediated quinolone lethality. In the absence of agr and downregulation, cip-
rofloxacin loses some lethality. With gentamicin and daptomycin, modulation of S.
aureus tolerance by Δagr involves leakage of bacterial components that affect anti-
microbial lethality [91, 92]. Thus, mutation of global regulators during clinical
infection can result in multiple phenotypes associated with tolerance that are not
detected as resistance.
The initial assertion that ROS contribute to the lethal action of multiple antimicro-
bial classes [25] has expanded to include a contribution of ROS to several severe
stress conditions. The idea that bacteria produce toxic ROS supports the more gen-
eral concept that bacteria respond to lethal stress by self-destructing. Existing data,
both direct and indirect, lead to the following scenario: when bacteria experience
severe stress, they make a live-or-die decision based on the response of several
bifunctional genes that are protective at low levels of stress but destructive at high
levels (Fig. 20.2). Destruction is achieved by ROS accumulation exceeding a thresh-
old beyond which genes designed to protect from oxidative stress fail to halt the
self-amplification of ROS. How bacterial suicide may confer a selective advantage
upon bacterial populations is currently unknown (see [4]).
The bifunctional nature of the lethal stress-response system makes application of
the principles challenging, because lethal treatment may be either enhanced or
diminished by increasing oxidative stress. For example, elevation of intracellular
superoxide levels prior to drug exposure may be protective, because the cells are less
susceptible to antimicrobials due to induction of protective genes. However, if stress
is severe, ROS accumulation will lead to cell death. Collectively, these observations
suggest that superoxide concentration, localization (intracellular, extracellular), and
regulation are complex. They may even change during the course of drug exposure
and infection. Nevertheless, several practical applications emerge from ROS contrib-
uting to antibacterial lethality. One is to avoid consuming antioxidants during ther-
apy with lethal antimicrobials, as antioxidants quench ROS. Another is that ROS
accumulation may represent a new form of interfering cross-reaction for tolerance
638 Y. Hong et al.
among seemingly unrelated drugs: drugs that reduce metabolism could inhibit the
production of ROS needed for full lethal activity of other antimicrobials.
Overall, consideration of ROS enriches our understanding of how bacteria
respond to severe stress and opens new ways to control bacterial populations.
Major Points
• Although lethal activity of current antimicrobials may be sufficient to clear most
infections, antimicrobial lethality needs to be even more rapid and extensive to
restrict the induction of resistance.
• Bacteria exhibit a lethal, ROS-mediated stress response that contributes to the
lethal activity of multiple antimicrobial classes; bacteriostatic assays of antimi-
crobial activity are often insensitive to perturbation of lethal stress responses.
• The effects of ROS are transient, which makes measurements of MBC uninfor-
mative with respect to rapid killing and suppression of induction of new resistant
mutants.
• Suppression of the ROS-mediated lethal stress response with antioxidants may
contribute to the emergence of resistance.
• Modulation of the ROS-mediated lethal stress response offers a new way to con-
trol bacterial populations.
Acknowledgments We thank the following for critical comments on the manuscript: Marila
Gennaro and Bo Shopsin.
References
11. Drlica K, Zhao X. Mutant selection window hypothesis updated. Clin Infect Dis. 2007;44:681–8.
12. Baym M, Lieberman T, Kelsic E, Chait R, Gross R, Yelin I, et al. Spatiotemporal microbial
evolution on antibiotic landscapes. Science. 2016;353:1147–51.
13. Cui J, Liu Y, Wang R, Tong W, Drlica K, Zhao X. The mutant selection window demonstrated
in rabbits infected with Staphylococcus aureus. J Infect Dis. 2006;194:1601–8.
14. Ni W, Song X, Cui J. Testing the mutant selection window hypothesis with Escherichia coli
exposed to levofloxacin in a rabbit tissue cage infection model. Eur J Clin Microbiol Infect
Dis. 2014;33:385–9.
15. Zhang B, Gu X, Li Y, Li X, Gu M, Zhang N, et al. In vivo evaluation of mutant selection
window of cefquinome against Escherichia coli in piglet tissue-cage model. BMC Vet Res.
2014;10:297.
16. Xiong M, Wu X, Ye X, Zhang L, Zeng S, Huang Z, et al. Relationship between cefquinome
PK/PD parameters and emergence of resistance of Staphylococcus aureus in rabbit tissue-cage
infection model. Front Microbiol. 2016;7:874.
17. Levin-Reisman I, Ronin I, Gefen O, Braniss I, Shoresh N, Balaban N. Antibiotic tolerance
facilitates the evolution of resistance. Science. 2017;355:826–30.
18. Wang X, Zhao X. Contribution of oxidative damage to antimicrobial lethality. Antimicrob
Agents Chemother. 2009;53:1395–402.
19. Dorsey-Oresto A, Lu T, Mosel M, Wang X, Salz T, Drlica K, et al. YihE kinase is a central
regulator of programmed cell death in bacteria. Cell Rep. 2013;3:528–37.
20. Dwyer D, Belenky P, Yang J, MacDonald I, Martell J, Takahashi N, et al. Antibiotics induce
redox-related physiological alterations as part of their lethality. Proc Natl Acad Sci U S A.
2014;111:E2100–E9.
21. Hong Y, Li L, Luan G, Drlica K, Zhao X. Contribution of reactive oxygen species to thymine-
less death in Escherichia coli. Nat Microbiol. 2017;(in press).
22. Burger R, Drlica K. Superoxide protects Escherichia coli from bleomycin mediated lethality.
J Inorg Biochem. 2009;109:1273–7.
23. Mosel M, Li L, Drlica K, Zhao X. Superoxide-mediated protection of Escherichia coli from
antimicrobials. Antimicrob Agents Chemother. 2013;57:5755–9.
24. Liu Y, Imlay J. Cell death from antibiotics without the involvement of reactive oxygen species.
Science. 2013;339:1210–3.
25. Kohanski M, Dwyer D, Hayete B, Lawrence C, Collins J. A common mechanism of cellular
death induced by bactericidal antibiotics. Cell. 2007;130:797–810.
26. Ezraty B, Vergnes A, Banzhaf M, Duverger Y, Huguenot A, Brochado A, et al. Fe-S clus-
ter biosynthesis controls uptake of aminoglycosides in a ROS-less death pathway. Science.
2013;340:1583–7.
27. Greenberg JT, Monach P, Chou JH, Josephy PD, Demple B. Positive control of a global anti-
oxidant defense regulon activated by superoxide-generating agents in Escherichia coli. Proc
Natl Acad Sci U S A. 1990;87(16):6181–5.
28. Oethinger M, Podglajen I, Kern WV, Levy SB. Overexpression of the marA or soxS regulatory
gene in clinical topoisomerase mutants of Escherichia coli. Antimicrob Agents Chemother.
1998;42(8):2089–94.
29. Koutsolioutsou A, Martins EA, White DG, Levy SB, Demple B. A soxRS-constitutive muta-
tion contributing to antibiotic resistance in a clinical isolate of Salmonella enterica (Serovar
typhimurium). Antimicrob Agents Chemother. 2001;45(1):38–43.
30. Goswami M, Mangoli SH, Jawali N. Involvement of reactive oxygen species in the action of
ciprofloxacin against Escherichia coli. Antimicrob Agents Chemother. 2006;50(3):949–54.
31. Goswami M, Mangoli SH, Jawali N. Effects of glutathione and ascorbic acid on streptomycin
sensitivity of Escherichia coli. Antimicrob Agents Chemother. 2007;51(3):1119–22.
32. Albesa I, Becerra MC, Battan PC, Paez PL. Oxidative stress involved in the antibacterial action
of different antibiotics. Biochem Biophys Res Commun. 2004;317(2):605–9.
33. Becerra MC, Albesa I. Oxidative stress induced by ciprofloxacin in Staphylococcus aureus.
Biochem Biophys Res Commun. 2002;297(4):1003–7.
640 Y. Hong et al.
34. Wang X, Zhao X, Malik M, Drlica K. Contribution of reactive oxygen species to pathways of
quinolone-mediated bacterial cell death. J Antimicrob Chemother. 2010;65:520–4.
35. Howard BM, Pinney RJ, Smith JT. 4-quinolone bactericidal mechanisms. Arzneimittelforschung/
Drug Res. 1993;43:1125–9.
36. Chen C-R, Malik M, Snyder M, Drlica K. DNA gyrase and topoisomerase IV on the bacterial
chromosome: quinolone-induced DNA cleavage. J Mol Biol. 1996;258:627–37.
37. Zhao X, Malik M, Chan N, Drlica-Wagner A, Wang J-Y, Li X, et al. Lethal action of quino-
lones with a temperature-sensitive dnaB replication mutant of Escherichia coli. Antimicrob
Agents Chemother. 2006;50:362–4.
38. Malik M, Hussain S, Drlica K. Effect of anaerobic growth on quinolone lethality with
Escherichia coli. Antimicrob Agents Chemother. 2007;51:28–34.
39. Wu X, Wang X, Drlica K, Zhao X. A toxin-antitoxin module in Bacillus subtiltis can both
mitigate and amplify effects of lethal stress. PLoS One. 2011;6:e23909.
40. Li L, Hong Y, Luan G, Mosel M, Malik M, Drlica K, et al. Ribosomal elongation factor 4
promotes cell death associated with lethal stress. MBio. 2014;5:e01708.
41. Pogliano J, Lynch A, Belin D, Lin E, Beckwith J. Regulation of Escherichia coli cell envelope
proteins involved in protein folding anddegradation by the Cpx two-component system. Genes
Dev. 1997;11:1169–82.
42. Raivio T, Silhavy T. Transduction of envelope stress in Escherichia coli by the Cpx two-
component system. J Bacteriol. 1997;179:7724–33.
43. Kohanski M, Dwyer D, Wierzbowski J, Cottarel G, Collins J. Mistranslation of membrane
proteins and two-component system activation trigger antibiotic-mediated cell death. Cell.
2008;135:679–90.
44. Raivio T, Silhavy T. Periplasmic stress and ECF sigma factors. Annu Rev Microbiol.
2001;55:591–624.
45. Wu Y, Vulic M, Keren I, Lewis K. Role of oxidative stress in persister tolerance. Antimicrob
Agents Chemother. 2012;56:4922–6.
46. Kobayashi S, Ueda K, Komano T. The effects of metal ions on the DNA damage induced by
hydrogen peroxide. Agric Biol Chem. 1990;54:69–76.
47. Grimsrud P, Xie H, Griffin T, Bernlohr D. Oxidative stress and covalent modification of protein
with bioactive aldehydes. J Biol Chem. 2008;283:21837–41.
48. Girotti A. Lipid hydroperoxide generation, turnover, and effector action in biological systems.
J Lipid Res. 1998;39:1529–42.
49. Foti J, Devadoss B, Winkler J, Collins J, Walker G. Oxidation of the guanine nucleotide pool
underlies cell death by bactericidal antibiotics. Science. 2012;336:315–9.
50. Keren I, Wu Y, Inocencio J, Mulcahy L, Lewis K. Killing by bactericidal antibiotics does not
depend on reactive oxygen species. Science. 2013;339:1213–6.
51. Liu Y, Liu X, Qu Y, Wang X, Li L, Zhao X. Inhibitors of reactive oxygen accumulation delay
and/or reduce the lethality of several antistaphylococcal agents. Antimicrob Agents Chemother.
2012;56:6048–58.
52. Imlay J. Diagnosing oxidative stress in bacteria: not as easy as you might think. Curr Opin
Microbiol. 2015;24:124–31.
53. Mahoney TF, Silhavy TJ. The Cpx stress response confers resistance to some, but not all, bac-
tericidal antibiotics. J Bacteriol. 2013;195:1869–74.
54. Korshunov S, Imlay J. Detection and quantification of superoxide formed within the periplasm
of Escherichia coli. J Bacteriol. 2006;188:6326–34.
55. Gonzalez-Flecha B, Demple B. Metabolic sources of hydrogen peroxide in aerobically grow-
ing Escherichia coli. J Biol Chem. 1995;270(23):13681–7.
56. Messner K, Imlay J. Mechanism of superoxide and hydrogen peroxide formation by fuma-
rate reductase, succinate dehydrogenase, and aspartate oxidase. J Biol Chem. 2002;277:
42563–71.
57. Korshunov S, Imlay J. Two sources of endogenous hydrogen peroxide in Escherichia coli. Mol
Microbiol. 2010;75:1389–401.
20 Antimicrobial-Mediated Bacterial Suicide 641
58. Dwyer D, Kohanski M, Hayete B, Collins J. Gyrase inhibitors induce an oxidative damage
cellular death pathway in Escherichia coli. Mol Syst Biol. 2007;3:91. Epub.
59. Baek S, Li A, Sassetti C. Metabolic regulation of mycobacterial growth and antibiotic sensitiv-
ity. PLoS Biol. 2011;(5):e1001065.
60. Thomas V, Kinkead L, Janssen A, Schaeffer C, Woods K, Lindgren J, et al. A dysfunctional tri-
carboxylic acid cycle enhances fitness of Staphylococcus epidermidis during β-lactam stress.
MBio. 2013;4:e00437–13.
61. Davies B, Kohanski M, Simmons L, Winkler J, Collins J, Walker G. Hydroxyurea induces
hydroxyl radical-mediated cell death in Escherichia coli. Mol Cell. 2009;36:845–60.
62. Lobritz M, Belenky P, Porter C, Gutierrez A, Yang J, Schwarz E, et al. Antibiotic efficacy is
linked to bacterial cellular respiration. Proc Natl Acad Sci U S A. 2015;112:8173–80.
63. Meylan S, Porter C, Yang J, Belenky P, Gutierrez A, Lobritz M, et al. Carbon sources tune
antibiotic susceptibility in Pseudomonas aeruginosa via tricarboxylic acid cycle control. Cell
Chem Biol. 2017;24:195–206.
64. Park D, Akhtar M, Ansari A, Landick R, Kiley P. The bacterial response regulator ArcA
uses a diverse binding site architecture to regulate carbon oxidation globally. PLoS Genet.
2013;9:e1003839.
65. Loui C, Chang A, Lu S. Role of the ArcAB two-component system in the resistance of
Escherichia coli to reactive oxygen stress. BMC Microbiol. 2009;9:183.
66. Lu S, Killoran P, Fang F, Riley L. The global regulator ArcA controls resistance to reactive
nitrogen and oxygen intermediates in Salmonella enterica serovar Enteritidis. Infect Immun.
2002;70:451–61.
67. Deponte M. Programmed cell death in protists. Biochim Biophys Acta. 2008;1783(7):1396–405.
68. Jimenez C, Capasso JM, Edelstein CL, Rivard CJ, Lucia S, Breusegem S, et al. Different
ways to die: cell death modes of the unicellular chlorophyte Dunaliella viridis exposed to
various environmental stresses are mediated by the caspase-like activity DEVDase. J Exp Bot.
2009;60(3):815–28.
69. Bayles KW. Bacterial programmed cell death: making sense of a paradox. Nat Rev Microbiol.
2014;12(1):63–9.
70. Dwyer DJ, Camacho DM, Kohanski MA, Callura JM, Collins JJ. Antibiotic-induced bac-
terial cell death exhibits physiological and biochemical hallmarks of apoptosis. Mol Cell.
2012;46(5):561–72.
71. Gautam S, Sharma A. Involvement of caspase-3-like protein in rapid cell death of Xanthomonas.
Mol Microbiol. 2002;44(2):393–401.
72. Raju KK, Gautam S, Sharma A. Molecules involved in the modulation of rapid cell death in
Xanthomonas. J Bacteriol. 2006;188(15):5408–16.
73. Crumplin GC, Smith JT. Nalidixic acid: an antibacterial paradox. Antimicrob Agents
Chemother. 1975;8:251–61.
74. Malik M, Zhao X, Drlica K. Lethal fragmentation of bacterial chromosomes mediated by
DNA gyrase and quinolones. Mol Microbiol. 2006;61:810–25.
75. Hanawalt P. A balanced perspective on unbalanced growth and thymineless death. Front
Microbiol. 2015;6:504.
76. Khodursky AGE, Hanawalt PC. Thymineless death lives on: new insights into a classic phe-
nomenon. Annu Rev Microbiol. 2015;69:247–63.
77. Fonville N, Bates D, Hastings P, Hanawalt P, Rosenberg S. Role of RecA and the SOS response
in thymineless death in Escherichia coli. PLoS Genet. 2010;6:e1000865.
78. Fonville N, Vaksman Z, DeNapoli J, Hastings P, Rosenberg S. Pathways of resistance to thy-
mineless death in Escherichia coli and the function of UvrD. Genetics. 2011;189:23–36.
79. Sangurdekar D, Hamann B, Smirnov D, Srienc F, Hanawalt P, Khodursky A. Thymineless
death is associated with loss of essential genetic information from the replication origin. Mol
Microbiol. 2010;75:1455–67.
642 Y. Hong et al.
80. Kuong K, Kuzminov A. Disintegration of nascent replication bubbles during thymine star-
vation triggers RecA- and RecBCD-dependent replication origin destruction. J Biol Chem.
2012;287:23958–70.
81. Stivala L, Savio M, Carafoli F, Perucca P, Bianchi L, Maga G, et al. Specific structural determi-
nants are responsible for the antioxidant activity and the cell cycle effects of resveratrol. J Biol
Chem. 2001;276:22586–94.
82. Radimer K, Bindewald B, Hughes J, Ervin B, Swanson C, Picciano M. Dietary supplement use
by US adults: data from the National Health and nutrition examination survey, 1999–2000. Am
J Epidemiol. 2004;160:339–49.
83. Marathe S, Kumar R, Ajitkumar P, Nagaraja V. DC. Curcumin reduces the antimicrobial
activity of ciprofloxacin against Salmonella typhimurium and Salmonella typhi. J Antimicrob
Chemother. 2013;68:139–52.
84. Aires V, Delmas D. Common pathways in health benefit properties of RSV in cardiovascular
diseases, cancers and degenerative pathologies. Curr Pharmaceut Biotech. 2015;16:219–44.
85. Granzotto A, Zatta P. Resveratrol and Alzheimer's disease: message in a bottle on red wine and
cognition. Front Aging Neurosci. 2014;6:95.
86. Yang X, Li X, Ren J. From French paradox to cancer treatment: anti-cancer activities and
mechanisms of resveratrol. Anti Cancer Agents Med Chem. 2014;14:806–25.
87. Liu Y, Zhou J, Qu Y, Yang X, Shi G, Wang X, et al. Resveratrol antagonizes antimicrobial
lethality and stimulates ecovery of bacterial mutants. PLoS One. 2016;11:e0153023.
88. Brauner A, Fridman O, Gefen O, Balaban N. Distinguishing between resistance, tolerance and
persistence to antibiotic treatment. Nat Rev Microbiol. 2016;14:320–30.
89. Fowler V, Sakoulas G, McIntyre L, Meka V, Arbeit R, Cabell C, et al. Persistent bacteremia due
to methicillin-resistant Staphylococcus aureus infection is associated with agr dysfunction and
low-level in vitro resistance to thrombin-induced platelet microbicidal protein. J Infect Dis.
2004;190:1140–9.
90. Schweizer M, Furuno J, Sakoulas G, Johnson J, Harris A, Shardell M, et al. Increased mortality
with accessory gene regulator (agr) dysfunction in Staphylococcus aureus among bacteremic
patients. Antimicrob Agents Chemother. 2011;55:1082–7.
91. Kumar K, J Chen, Drlica K, Shopsin B. Dysfunction of the agr virulence regulator modulates
antimicrobial-mediated killing of Staphylococcus aureus. MBio. 2017;in press.
92. Pader V, Hakim S, Painter K, Wigneshweraraj S, Clarke T, Edwards A. Staphylococcus aureus
Inactivates daptomycin by releasing membrane phospholipids. Nat Microbiol. 2016;2:16194.
Chapter 21
PK/PD-Based Prediction of “Anti-Mutant”
Antibiotic Exposures Using In Vitro
Dynamic Models
21.1 Introduction
hypothesis [3, 4]. In this regard, the present chapter is an extension of an earlier
publication [5].
Targeted bacterial resistance studies using in vitro dynamic models [6–16] were
launched at the turn of the new millennium. In most of the resistance studies, loss in
susceptibility of antibiotic-exposed pathogens and/or the enrichment of resistant
mutant subpopulations was related to the ratio of 24-h area under the concentration-
time curve (AUC24) to the MIC. The fact that AUC24/MIC ratio rather than AUC24
per se related to susceptibility and/or population analysis data implies that AUC24/
MIC is an inter-strain resistance predictor that allows generalization of findings
obtained with different bacterial strains.
Contrary to expectations, most of these early works did not reveal clear relation-
ships between antibiotic exposure, expressed by the AUC24/MIC ratio, and the loss
in susceptibility of antibiotic-exposed pathogen cultures and/or the enrichment of
resistant mutants. For example, similar low resistance frequencies were reported in
an in vitro staphylococcal study at AUC24/MICs for ciprofloxacin that varied by a
factor of 16 [7]. Moreover, in the same study, frequency of resistance in
Staphylococcus aureus exposed to norfloxacin was paradoxically more pronounced
at a relatively high AUC24/MIC ratio (55 h) than at a low value of AUC24/MIC (3 h).
Obviously, in the absence of relationships between AUC24/MIC and loss in suscep-
tibility of antibiotic-exposed pathogens and/or the enrichment of resistant mutant
subpopulations, the AUC24/MIC thresholds reported to protect against the enrich-
ment of resistant mutants [8–15] are questionable.
These failures result from shortcomings in study design summarized elsewhere
[5], among which are the absence of resistant mutants in the starting inoculum,
insufficient duration of simulated treatments, simulation of AUC24/MIC ratios that
provide either sub-optimal or super-optimal effects of antibiotics on resistant
subpopulation(s) but not intermediate AUC24/MICs ratios, and inappropriate data
analysis. In particular, simple AUC24/MIC relationships with the emergence of
resistance – the greater the AUC24/MIC ratio, the less pronounced the enrichment of
resistant mutants or loss in susceptibility of antibiotic-exposed bacteria – were
reported in some studies, although more complex relationships between antibiotic
concentration and bacterial resistance are expected based on the “mutant selection
window” (MSW) hypothesis [3, 4]. According to this idea, resistant mutants are
selectively enriched at antibiotic concentrations between the MIC and the mutant
prevention concentration (MPC) but not at concentrations below the MIC or above
the MPC.
Strong support for the window hypothesis was first provided by an in vitro
dynamic model study in which S. aureus was exposed to 3-day dosing with
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 645
Although the AUC24/MIC ratio has been related to the enrichment of resistant bacte-
rial mutants, it is not the only predictor for the emergence of antibiotic resistance.
Other predictors that emerge from the mutant window hypothesis are the ratio of
AUC24 to the MPC, the time during which antibiotic concentration is inside the
MSW (TMSW), and the time when drug concentration is above the MPC (T>MPC).
Each is briefly discussed below.
21.3.1 AUC24/MPC
From the standpoint of the MSW hypothesis, the AUC24/MPC ratio rather than the
AUC24/MIC might better predict the enrichment of resistant mutants and/or loss in
susceptibility of antibiotic-exposed bacteria [36–38] because MPC directly mea-
sures mutant susceptibility, which in many cases does not correlate well with bulk
culture MIC. Nevertheless, AUC24/MIC- and AUC24/MPC-resistance relationships
for a given organism can differ only in quantitative, not in qualitative terms (bell-
shaped curves shifted along the x-axis). Consequently, the predictive potentials of
AUC24/MIC and AUC24/MPC ratios can be distinguished most clearly by their abil-
ity to serve as inter-strain predictors of resistance.
Obviously, bacterial resistance studies that utilize at least two bacterial strains
with different MPC/MIC ratios are needed to distinguish between AUC24/MPC and
AUC24/MIC as potential predictors of resistance. To compare the abilities of AUC24/
MPC and AUC24/MIC as inter-strain predictors of resistance, two strains of S.
aureus with distinctly different MPC/MIC (4 versus 16) were used in a study that
simulated twice-daily dosing of ciprofloxacin for 3 days [29]. When comparing the
descending portions of the AUC24/MPC and AUC24/MIC relationships with AUBCM
over a wide range of drug exposure, the AUC24/MIC plots were more stratified than
the respective AUC24/MPC plots. For example, with mutants resistant to 4 × MIC
ciprofloxacin, the square correlation coefficient for the AUBCM against log AUC24/
MPC relationship was 1.6-fold greater (r2 0.70) than for the AUBCM against log
AUC24/MIC relationship (r2 0.43). Even greater differences between AUC24/MPC
and AUC24/MIC relationships were reported with mutants resistant to 8 × MIC of
antibiotic (r2 0.72 versus 0.35). Figure 21.1 shows a systematic increase in the pre-
dictive power of AUC24/MPC and a concomitant decrease in the predictive power of
AUC24/MIC with an increase in culture MIC. These findings suggest that the AUC24/
MPC ratio is a more potent inter-strain predictor for staphylococcal resistance to the
fluoroquinolone than the AUC24/MIC ratio. This implies lower strain-to-strain vari-
ability in AUC24/MPC thresholds that prevent mutant enrichment. Less variation in
the “anti-mutant” thresholds is desired because clinical recommendations need to
be suitable for many strains.
648 A. A. Firsov et al.
Fig. 21.1 MPC- and MIC-related pharmacokinetic variables as predictors of the enrichment of
ciprofloxacin-resistant mutants of S. aureus. (Reconstructed from Ref. [29])
The distinct advantages of the AUC24/MPC over the AUC24/MIC ratio were dem-
onstrated subsequently in a similarly designed study with levofloxacin-exposed S.
aureus [19]. In this work three strains with the same MIC for levofloxacin but with
distinctly different MPCs (MPC/MIC from 8 to 64) were treated with once-daily
fluoroquinolone for 3 days. According to our analysis, plotting MICfinal/MICinitial
against either AUC24/MPC or AUC24/MIC did not allow combination of data
obtained with individual S. aureus strains: both AUC24/MPC and AUC24/MIC rela-
tionships with MICfinal/MICinitial were too stratified to be combined. However, quali-
tative characteristics of resistance, i.e., the loss in susceptibility (posttreatment MIC
elevation) or the absence of such a loss, were better correlated to AUC24/MPC than
to AUC24/MIC in a strain-independent manner. Reconstructed from reported data
[19], Fig. 21.2 demonstrates bacterial strain specificity of the AUC24/MIC-resistance
relationships in contrast to the strain-independent AUC24/MPC-resistance
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 649
r elationship. As seen in the figure, unlike the stratified AUC24/MIC plots observed
with individual strains, the respective AUC24/MPC are virtually superimposed.
In contrast to fluoroquinolone-exposed S. aureus, with Gram-negative bacteria,
correlations between AUBCM, which reflects the enrichment of ciprofloxacin-
resistant mutants, and simulated AUC24/MPC were not as strong as between AUBCM
and AUC24/MIC. The respective r2s with ciprofloxacin-exposed Escherichia coli
were 0.69 versus 0.86 [24], with Klebsiella pneumoniae they were 0.72 versus 0.76
[27], and with P. aeruginosa they were 0.65 versus 0.75 for the AUC24/MPC and
AUC24/MIC ratios [28]. This difference between Gram-negative and Gram-positive
bacteria could reflect strain-to-strain variability in the “anti-mutant” thresholds. As
seen in Fig. 21.3, with each Gram-negative situation, the scattering of the “anti-
mutant” AUC24/MPC was more pronounced than with AUC24/MIC: with E. coli it
was 4-fold versus 2-fold, with K. pneumoniae it was 25-fold versus 2-fold, and with
P. aeruginosa it was 26-fold versus 5-fold differences. Unlike the Gram-negative
bacteria tested, S. aureus strains exhibit less variable “anti-mutant” AUC24/MPC
ratios than the respective AUC24/MIC ratios (2.4-fold versus 3.8-fold differences).
In a recent study that exposed S. aureus strains to 5-day treatment with linezolid
(MPC/MIC from 2.5 to 5), a lower r2 (0.79) was also reported for the AUBCM rela-
tionship with AUC24/MPC than with AUC24/MIC (r2 0.91) [34]. However, strain-to-
strain variability for the “anti-mutant” AUC24/MPC and AUC24/MIC ratios was
identical: twofold differences (from 48 to 96 h and from 120 to 240 h, respectively).
In another study that simulated 5-day treatments of two strains of S. aureus with
daptomycin (MPC/MIC from 3 to 5) and vancomycin (MPC/MIC from 3 to 8) [31],
the AUC24/MPC ratio was less predictive for S. aureus resistance than the AUC24/
MIC ratio. In contrast to reasonable AUC24/MIC relationships with Nmax/Ninitial (r2
0.68 for mutants resistant to 2 × MIC and r2 0.66 for mutants resistant to 4 × MIC
of the tested antibiotics) and MICfinal/MICinitial (r2 0.64), there was no correlation
between AUC24/MPC and either population analysis or susceptibility data. Based on
650 A. A. Firsov et al.
data obtained with two S. aureus strains exposed to daptomycin and vancomycin,
MICfinal/MICinitial plots against AUC24/MPC were characterized by widely scattered
points. Thus, preference for a particular parameter may vary according to the
antibiotic-pathogen pairs studied.
In this section we confined ourselves to studies containing clear evidence on the
advantages or disadvantages of AUC24/MPC over AUC24/MIC. There are other stud-
ies in which the conclusion that one of the predictors of bacterial resistance (usually
AUC24/MPC ratio) is preferable was not supported by the experimental findings.
One example is a resistance study with two strains of S. aureus exposed to five fluo-
roquinolones (three AUC24/MPC ratios per one antibiotic-pathogen pair, MPC/MIC
ratios from 2 to 15) [39]. According to the author’s statement, AUC24/MPC was
recognized as “the only parameter to correlate with the development of resistance”
although this correlation was extremely weak (r2 0.2). As seen in Fig. 21.4, when
plotting MICfinal/MICinitial against simulated AUC24/MPC or AUC24/MIC ratios, a
cloud of scattered points is observed using both potential predictors for emergence
of staphylococcal resistance. It is possible that this scatter could be avoided by using
a wider range of simulated AUC24/MPC or AUC24/MIC ratios and treatments longer
than 48 h. At least 72-hour treatments were recommended in our resistance studies
with fluoroquinolones [32]. Short observation times (24 h) could have affected the
results even more for a single-dose study with ciprofloxacin-exposed E. coli (three
strains with MPC/MIC from 4 to 16) [40]. For this reason, the authors’ conclusion
that “AUC/MPC ratio was the single pharmacodynamic index that predicted preven-
tion of resistant mutant development” should be taken with caution.
Thus, contrary to expectations, use of AUC24/MPC as an inter-strain predictor of
resistance has an advantage over AUC24/MIC only with some antibiotic-pathogen
pairs (fluoroquinolone-exposed S. aureus) but not with other pairs (fluoroquinolone-
exposed E. coli, K. pneumoniae, P. aeruginosa, daptomycin- and vancomycin-
exposed S. aureus). Further studies are needed to understand apparent differences
among bacterial species and antimicrobials.
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 651
21.3.2 TMSW
[41], this conclusion was the result of inappropriate combination of data obtained in
simulations of isoniazid pharmacokinetics in fast and slow acetylators (half-lives
1.8 and 4.2 h, respectively). Quite possibly fast and slow acetylator data might relate
to different TMSW-resistance relationships. On the other hand, there were insufficient
data to establish a specific relationship for each type of pharmacokinetic profile
(two points per one profile only). Moreover, in both cases, TMSW varied over very
small ranges: from 30% to 54% (fast acetylator simulations) and from 80% to 100%
(slow accelerator simulations) of the dosing interval. Based on these limited data,
delineation of a relationship or lack of a relationship is questionable.
A less clear situation is found with a study [44] in which TMSW failed to be predic-
tive for moxifloxacin and levofloxacin resistance with S. pneumoniae (four strains
exposed to 3-day treatments with the fluoroquinolones). In this case the unsuccess-
ful attempts to relate susceptibility of antibiotic-exposed bacteria with TMSW might
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 653
have resulted from the overestimation of MPCs, at least for a S. pneumoniae strain
that exhibited a biphasic pattern in the frequency-concentration curve. As shown in
our study with ciprofloxacin-exposed E. coli [24], the higher MPC derived from the
second phase of such curves describes the second-step mutations that might or
might not be present in pharmacokinetic simulations. As a result, the true value of
TMSW could be overestimated, and the value of T>MPC could be underestimated.
Moreover, in most simulations (e.g., in five out of six experiments with moxifloxa-
cin), there was no loss in susceptibility of S. pneumoniae. Because of the unbal-
anced study design, TMSW was used to explain the lack of resistance rather than the
emergence of resistance. We conclude that arguments against the predictive value of
TMSW are at best weak.
However, even in studies in which relationships between TMSW and emergence of
resistance were demonstrated, the predictive power of TMSW was always lower than
the AUC24/MIC ratio. For example, in the studies described above, with S. aureus
exposed to four fluoroquinolones, the respective r2s were 0.72 versus 0.90 [17], with
ciprofloxacin-exposed E. coli the r2s were 0.61 versus 0.84 [25], and with
ciprofloxacin-exposed P. aeruginosa they were 0.56 versus 0.80 [45]. With
doripenem-exposed P. aeruginosa r2s were 0.69 versus 0.80 [45], and with
glycopeptide-exposed S. aureus they were 0.60 versus 0.68 (MICfinal/MICinitial data)
or 0.50 versus 0.64 (Nmax/Nmin data) [31].
The differences in predictive power described above may be due to relating
AUC24/MIC to TMSW while ignoring information about the position of simulated
antibiotic concentrations inside the MSW, a feature that is likely to be very impor-
tant with respect to mutant amplification. To test this hypothesis, the enrichment of
ciprofloxacin-resistant S. aureus was examined at drug concentrations that oscil-
lated near the MPC, i.e., close to the top of the MSW (“upper case”), or close to the
MIC, i.e., at the lower limit of the MSW (“lower case”). In both cases the TMSW was
the same [46]. In this study, two methicillin-resistant strains of S. aureus (MPC/
MIC 4 and 16) were exposed to twice-daily ciprofloxacin for 3 consecutive days.
The simulated AUC24/MIC were 50 h (“lower case”) and 260 h (“upper case”) to
provide TMSW of 75% of the dosing interval with one strain and 30 h (“lower case”)
and 100 h (“upper case”) to provide TMSW of 56% with another strain. With each
strain, AUBCM (a measure of mutant enrichment) observed in the “lower case” was
much greater than in the “upper case,” thereby showing less pronounced enrichment
of ciprofloxacin-resistant staphylococci at antibiotic concentrations oscillating near
the MPC than near the MIC, even though for each strain TMSW was the same.
Heterogeneity of the MSW was further examined in a study that exposed four
Escherichia coli strains to twice-daily ciprofloxacin dosing for 3 days [25]. To
explore the different predictive powers of TMSW and AUC24/MIC, the enrichment of
ciprofloxacin-resistant E. coli mutants was studied at wide ranges of TMSWs and
AUC24/MICs (up to eight points per strain). Peak antibiotic concentrations were
simulated to be close to the MIC, between the MIC and MPC, and above the MPC;
TMSW varied from 0% to 100% of the dosing interval. The amplification (enrich-
ment) of resistant mutants was monitored by plating on media with 8 × MIC of the
antibiotic. With each organism, TMSW plots of the AUBCM split into two portions,
654 A. A. Firsov et al.
Fig. 21.7 TMSW-dependent changes in the AUBCM reflecting the enrichment of E. coli mutants
resistant to 4 × MIC of ciprofloxacin fitted by Eq. 1, separately for points that belong to the ascend-
ing portion (Y0 = 0, x0 = 20.18, a = 459.1, b = 15.86) and descending portion (Y0 = 0, x0 = 83.40,
a = 500.1, b = 6.969) of the AUBCM-AUC24/MIC curve. AUC24/MIC values are shown in callouts.
(Reconstructed from Ref. [25])
one for antibiotic concentrations below the MPC (T>MPC = 0) and the other for
c oncentrations consistently above the MPC (T>MPC > 0). The result was a hysteresis
loop. Figure 21.7 illustrates a TMSW relationship with AUBCM observed with one of
the E. coli strains examined. As seen in the figure, when antibiotic concentrations
were below the MPC (points corresponding to the ascending portion of the bell-
shaped AUBCM-AUC24/MIC curve – AUC24/MIC ratios of 15, 30 and 60 h), the
AUBCM at a given TMSW was greater than at the same TMSW relevant to the descend-
ing portion of the AUBCM-AUC24/MIC curve (AUC24/MIC ratios of 360 and 720 h
gave the same TMSW).
The distinct T>MPC-dependent splitting of the AUBCM-TMSW curves (Fig. 21.7)
prevents consideration of data obtained at T>MPC = 0 and at T>MPC > 0 as a single data
set. When the data with the four E. coli strains were combined, a sigmoid function
fits well with AUBCM versus TMSW data sets taken separately at T>MPC = 0 and
T>MPC > 0 (r2s 0.81 and 0.92, respectively). In both cases, correlation of TMSW with
resistance appeared to be of the same power as observed with the AUC24/MIC ratio
(r2 0.84). In contrast to the separated analysis of the TMSW data referring to the condi-
tions of T>MPC = 0 or T>MPC > 0, fitting the whole data pool while ignoring T>MPC
exhibited a weaker correlation between TMSW and mutant enrichment (r2 0.61).
Hysteresis loops have also been reported for TMSW relationships with S. aureus
resistance to linezolid [47]. Using inocula of three methicillin-resistant S. aureus
strains (MIC of linezolid = 2 mg/L), spiked with low concentrations of previously
selected resistant mutants (MIC, 8 mg/L), AUC24/MIC- and TMSW-dependent mutant
enrichment was observed in 5-day treatments with twice-daily linezolid. With each
strain, TMSW relationships with the AUBCM (for mutants resistant to 4 × MIC) exhib-
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 655
ited a hysteresis loop, with the upper sigmoid corresponding to T>MPC = 0, and the
lower one to the T>MPC > 0. Based on combined data obtained with the three bacterial
strains, AUBCM correlated better with TMSW data taken separately when T>MPC was
zero or exceeded zero (r2 0.99) than with pooled data ignoring T>MPC (r2 0.24).
We conclude that a hysteresis loop is inherent in the TMSW relationships with
mutant enrichment. It is very likely that the incorrect combination of data obtained
at T>MPC = 0 and at T>MPC > 0 is among the reasons for an underestimation of the true
role of TMSW as a predictor of the emergence of bacterial resistance. For example, in
a resistance study with meropenem-exposed Acinetobacter baumannii [48], the
conclusion that TMSW is not a suitable parameter relating to mutant enrichment might
result from inappropriate combining TMSWs belonging to the upper (T>MPC = 0) and
lower (T>MPC > 0) portions of the hysteresis loop. With each A. baumannii strain, the
TMSWs observed at the minimal antibiotic exposure met the condition T>MPC = 0,
whereas the TMSWs at the maximal exposure met the situation in which T>MPC > 0.
Overall, although TMSW is mutually related to the MSW, the appropriate use of
this parameter requires consideration of the T>MPC data.
21.3.3 T>MPC
The available reports on the use of T>MPC as a predictor of bacterial resistance are
much less frequent than those that report AUC24/MIC, AUC24/MPC, and TMSW, in
part because antibiotic concentrations simulated in these studies exceeded the MPCs
for only a short time or did not reach the MPCs. Even in cases in which T>MPCs were
positive, the reported data [39, 40] are too limited to delineate quantitative T>MPC
relationships with the enrichment of resistant mutants. However, unlike the staphy-
lococcal study with five fluoroquinolones [39], a reasonable link between the emer-
gence of bacterial resistance (qualitative characteristics only) and T>MPC can be seen
from the E. coli study using ciprofloxacin [40]. With each of three E. coli strains, at
least in simulations of ciprofloxacin pharmacokinetics having a half-life of 4 h, the
emergence of resistance was consistently associated with lower T>MPC. Apparently,
the authors’ conclusion that the emergence of bacterial resistance cannot be pre-
dicted by the T>MPC reflects the inability to combine data obtained with different E.
coli strains: at the same T>MPC, resistance to ciprofloxacin developed with one strain
but not with another.
In another study, suppression of A. baumannii resistance to meropenem (again,
qualitative characteristics – the presence or absence of resistant mutants of antibiotic-
exposed bacteria) was achieved for two strains with MPC/MIC ratios of approxi-
mately 15 and 60 [48] at similar T>MPCs. It is noteworthy that strain-independent
T>MPC-resistance relationships could be established for each mode of antibiotic
administration (0.5- and 3-h infusions). These relationships are specific for the type
of simulated pharmacokinetic profile: the protective T>MPC was lower in the longer
656 A. A. Firsov et al.
than in shorter meropenem infusions (Fig. 21.8). Such data are further evidence for
pharmacokinetic profile-dependent emergence of bacterial resistance.
A quantitative resistance index, AUBCM, was first related to T>MPC in the above-
mentioned ciprofloxacin study with E. coli [25]. When AUBCM versus T>MPC data
sets for four strains of E. coli were combined, a mono-exponential decay function
fits these data with a relatively high r2 (0.71). Using the points that met the condition
of T>MPC > 0, similar correlations between AUBCM with AUC24/MPC (r2 0.74) and
with AUC24/MIC (r2 0.81) were observed. Thus, the predictive power of T>MPC was
not inferior to AUC24/MPC or to AUC24/MIC ratios.
Even stronger correlations were reported recently between AUBCM and T>MPC
with linezolid-exposed S. aureus [47]. A sigmoid function fits combined data for
three S. aureus strains with a high r2 (0.99). For the points that meet the condition
T>MPC > 0, the sum of TMSW and T>MPC equals 100% of the dosing interval, and the
T>MPC plot of AUBCM is a mirror image of the TMSW plot at T>MPC > 0 with the same
r2. In this study, both T>MPC and TMSW at T>MPC > 0 exhibited stronger correlations
with AUBCM than did AUC24/MPC (r2 0.80) and AUC24/MIC (r2 0.85).
Thus, together with AUC24/MIC, AUC24/MPC, and TMSW, T>MPC can be consid-
ered as a strain-independent predictor for the emergence of bacterial resistance.
21.4.1 Monotherapy
The “anti-mutant” AUC24/MIC ratios were established when S. aureus was exposed
to fluoroquinolones [17]. Based on the bell-shaped AUC24/MIC relationships with
MICfinal/MICinitial, predicted “protective” AUC24/MICs appeared to be similar for
levofloxacin (201 h), moxifloxacin (222 h), gatifloxacin (241 h), and ciprofloxacin
(244 h). However, these thresholds are clinically attainable only with moxifloxacin.
With a 400 mg dose of moxifloxacin, the “anti-mutant” AUC24/MIC ratio is 66% of
the clinically attainable value, whereas with two 500 mg doses of ciprofloxacin, a
500 mg dose of levofloxacin, or a 400 mg dose of gatifloxacin, the respective anti-
mutant AUC24/MICs are 420%, 220%, and 190% of the clinically attainable values.
Thus, at least against S. aureus, moxifloxacin is expected to protect against resis-
tance development in a clinical setting, whereas the three other fluoroquinolones
will likely enrich mutant subpopulations.
Resistance thresholds reported in vitro studies from different research groups
exhibit considerable variability. For example, with grepafloxacin-exposed S. pneu-
moniae, “protective” AUC24/MIC ratios varied from 32 h [14] to 80 h [10] while
those of levofloxacin were from 9 h [14] to 26 h [9] and 35 h [11]. Furthermore,
although moxifloxacin-resistant S. pneumoniae were not found at AUC24/MIC ratios
of 60 h [11] and 107 h [10], significant losses in susceptibility were seen at AUC24/
MICs as high as 43,500 h [14]. Analysis of these findings [5] indicates that different
estimates of the “anti-mutant” AUC24/MIC ratio can be attributed to differences in
study design and data processing. For this reason, it is of particular interest to com-
pare “anti-mutant” AUC24/MICs obtained under the same experimental conditions.
Based on data reported in ciprofloxacin resistance studies that determine “anti-
mutant” AUC24/MIC ratios using the descending portion of the MICfinal/MICinitial or
AUBCM versus AUC24/MIC curve [17, 29, 50], lower resistance thresholds were
established with Gram-positive than with Gram-negative bacteria. The “anti-
mutant” AUC24/MIC ratios were 125–244 h with S. aureus (three strains) [17, 29],
700–1100 h with E. coli (four strains), 1300–2600 h with K. pneumoniae (three
strains), and 300–1400 h with P. aeruginosa (four strains) [50]. However, when
related to clinically attainable AUC24/MIC ratios for each individual strain, different
distributions emerge for ciprofloxacin “anti-mutant” potentials for different species.
658 A. A. Firsov et al.
Fig. 21.9 “Anti-mutant” thresholds of AUC24/MIC related to the clinically attainable AUC24/MIC
ratios: ciprofloxacin against Gram-positive and Gram-negative bacteria. (Reconstructed from Ref.
[17, 29, 50])
As shown in Fig. 21.9, with E. coli but not with S. aureus, P. aeruginosa, and K.
pneumoniae, the predicted “anti-mutant” AUC24/MICs are achieved in a clinical
setting (ratio of the resistance threshold to the clinically achievable AUC24/MIC <1).
Our resistance study with daptomycin- and vancomycin-exposed S. aureus [31]
provides an example of a more favorable situation, at least for daptomycin. Based
on combined data obtained with two S. aureus strains, the predicted “anti-mutant”
AUC24/MIC ratio (200 h) was smaller than the clinically attainable AUC24/MIC90s:
380 h for a 4 mg/kg dose of daptomycin and 570 h for a 6 mg/kg dose of the antibi-
otic. Unlike daptomycin, the “anti-mutant” target for vancomycin virtually coin-
cided with the clinically attainable AUC24/MIC90 ratio (200 h for 1 gm twice-daily
dosing). Reasonably optimistic predictions were made in a linezolid study with S.
aureus [34]: with two of three studied strains, the “anti-mutant” AUC24/MIC ratios
were equal to the clinically attainable value (120 h for a 600 mg dose twice a day).
However, with the third S. aureus strain, the “anti-mutant” threshold was twofold
greater.
Overall, these examples show that clinically achievable AUC24/MIC ratios may
not always overlap the “anti-mutant” AUC24/MIC thresholds. In conjunction with
extensive MPC testing of various antibiotic-pathogen pairs [51], these findings pre-
dict emergence of bacterial resistance with the use of many existing antibiotics as
usually prescribed. As dose escalation is rarely possible due to limited patient toler-
ability, combined antibiotic therapy provides an alternative to the replacement of a
less “protective” agent by a more “protective” agent.
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 659
21.5 Conclusions
these relationships and their clinical relevance remain limited because of the scar-
city of dynamic resistance studies with many antibiotic classes and diverse patho-
gens. Indeed, only a few bacterial species have been examined; quantitative findings
reported with a limited number of pathogens will remain applicable only to those
antimicrobial-pathogen pairs and not to other strains of the same species until the
data are generalized. Moreover, further studies that compare inter-strain predictions
of mutant enrichment using AUC24/MIC and AUC24/MPC are particularly needed,
due to apparent differences between fluoroquinolone-exposed Gram-positive and
Gram-negative bacteria. Nevertheless, bacterial resistance studies using dynamic
models provide notable progress in understanding of the mutant selection window
as a framework for predicting the selective enrichment of resistant mutants.
Major Points
• Relationships between PK/PD (pharmacokinetic/pharmacodynamic) indices and
the emergence of bacterial resistance are a basis for designing “anti-mutant”
antibiotic dosing regimens, i.e., regimens that are expected to prevent or restrict
the enrichment of resistant mutant subpopulations.
662 A. A. Firsov et al.
References
1. Linder KE, Nicolau DP, Nailor MD. Predicting and preventing antimicrobial resistance utiliz-
ing pharmacodynamics: part I gram positive bacteria. Expert Opin Drug Metabol Toxicol.
2016;12:267–80.
2. Abdelraouf K, Linder KE, Nailor MD, Nicolau DP. Predicting and preventing antimicrobial
resistance utilizing pharmacodynamics: part II gram-negative bacteria. Expert Opin Drug
Metabol Toxicol, 2017; doi: https://doi.org/10.1080/17425255.2017.1329417.
3. Zhao X, Drlica K. Restricting the selection of antibiotic-resistant mutants: a general strategy
derived from fluoroquinolone studies. Clin Infect Dis. 2001;33:147–56.
4. Blondeau JM, Zhao X, Hansen G, Drlica K. Mutant prevention concentrations of fluoroqui-
nolones for clinical isolates of Streptococcus pneumoniae. Antimicrob Agents Chemother.
2001;45:433–8.
5. Firsov AA, Zinner SH. Lubenko IY. In: Nightingale CH, Ambrose PG, Drusano GL,
Murakawa T, editors. vitro dynamic models as tools to predict antibiotic pharmacodynamics,
Antimicrobial pharmacodynamics in theory and clinical practice. 2nd ed. New York: Informa
Healthcare USA, Inc.; 2007. p. 45–78.
6. Lacy MK, Lu W, Xu X, Tessier PR, Nicolau DP, Quintiliani R, Nightingale
CH. Pharmacodynamic comparisons of levofloxacin, ciprofloxacin, and ampicillin against
Streptococcus pneumoniae in an in vitro model of infection. Antimicrob Agents Chemother.
1999;43:672–7.
7. Aeschlimann JR, Kaatz GW, Rybak MJ. The effects of NorA inhibition on the activities of levo-
floxacin, ciprofloxacin and norfloxacin against two genetically related strains of Staphylococcus
aureus in an in-vitro infection model. J Antimicrob Chemother. 1999;44:343–9.
8. Peterson ML, Hovde LB, Wright DH, Hoang AD, Raddatz JK, Boysen PJ, Rotschafer
JC. Fluoroquinolone resistance in Bacteroides fragilis following sparfloxacin exposure.
Antimicrob Agents Chemother. 1999;43:2251–5.
9. Madaras-Kelly KJ, Demasters TA. In vitro characterization of fluoroquinolone concentra-
tion/MIC antimicrobial activity and resistance while simulating clinical pharmacokinetics of
levofloxacin, ofloxacin, or ciprofloxacin against Streptococcus pneumoniae. Diagn Microbiol
Infect Dis. 2000;37:253–60.
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 663
10. Coyle EA, Kaatz GW, Rybak MJ. Activities of newer fluoroquinolones against ciprofloxacin-
resistant Streptococcus pneumoniae. Antimicrob Agents Chemother. 2001;45:1654–9.
11. Zhanel GG, Walters M, Laing N, Hoban DJ. In vitro pharmacodynamic modelling simulat-
ing free serum concentrations of fluoroquinolones against multidrug-resistant Streptococcus
pneumoniae. J Antimicrob Chemother. 2001;47:435–40.
12. Thorburn CE, Edwards DI. The effect of pharmacokinetics on the bactericidal activity of cip-
rofloxacin and sparfloxacin against Streptococcus pneumoniae and the emergence of resis-
tance. J Antimicrob Chemother. 2001;48:15–22.
13. Ross GH, Wright DH, Hovde LB, Peterson ML, Rotschafer JC. Fluoroquinolone resistance in
anaerobic bacteria following exposure to levofloxacin, trovafloxacin, and sparfloxacin in an in
vitro pharmacodynamic model. Antimicrob Agents Chemother. 2001;45:2136–40.
14. Klepser ME, Ernst EJ, Petzold CR, Rhomberg P, Doern GV. Comparative bactericidal activi-
ties of ciprofloxacin, clinafloxacin, grepafloxacin, levofloxacin, moxifloxacin, and trova-
floxacin against Streptococcus pneumoniae in a dynamic in vitro model. Antimicrob Agents
Chemother. 2001;45:673–8.
15. Peterson ML, Hovde LB, Wright DH, Brown GH, Hoang AD, Rotschafer
JC. Pharmacodynamics of trovafloxacin and levofloxacin against Bacteroides fragilis in an in
vitro pharmacodynamic model. Antimicrob Agents Chemother. 2002;46:203–10.
16. Wright DH, Gunderson SM, Hovde LB, Ross GH, Ibrahim AS, Rotschafer JC. Comparative
pharmacodynamics of three newer fluoroquinolones versus six strains of staphylococci in
an in vitro model under aerobic and anaerobic conditions. Antimicrob Agents Chemother.
2002;46:1561–3.
17. Firsov AA, Vostrov SN, Lubenko IY, Drlica K, Portnoy YA, Zinner SH. In vitro pharmaco-
dynamic evaluation of the mutant selection window hypothesis using four fluoroquinolones
against Staphylococcus aureus. Antimicrob Agents Chemother. 2003;47:1604–13.
18. Firsov AA, Vostrov SN, Lubenko IY, Arzamastsev AP, Portnoy YA, Zinner SH. ABT492 and
levofloxacin: comparison of their pharmacodynamics and their abilities to prevent the selec-
tion of resistant Staphylococcus aureus in an in vitro dynamic model. J Antimicrob Chemother.
2004;54:178–86.
19. Liang B, Bai N, Cai Y, Wang R, Drlica K, Zhao X. Mutant prevention concentration-based
pharmacokinetic/pharmacodynamic indices as dosing targets for suppressing the enrich-
ment of levofloxacin-resistant subpopulations of Staphylococcus aureus. Antimicrob Agents
Chemother. 2011;55:2409–12.
20. Tam VH, Louie A, Deziel MR, Liu W, Leary R, Drusano GL. The relationship between quino-
lone exposures and resistance amplification is characterized by an inverted U: a new paradigm
for optimizing pharmacodynamics to counter select resistance. Antimicrob Agents Chemother.
2007;51:744–7.
21. MacGowan AP, Rogers CA, Holt HA, Bowker KE. Activities of moxifloxacin against, and
emergence of resistance in, Streptococcus pneumoniae and Pseudomonas aeruginosa in an in
vitro pharmacokinetic model. Antimicrob Agents Chemother. 2003;47:1088–95.
22. Zinner SH, Lubenko IY, Gilbert D, Simmons K, Zhao X, Drlica K, Firsov AA. Emergence of
resistant Streptococcus pneumoniae in an in vitro dynamic model that simulates moxifloxacin
concentrations inside and outside the mutant selection window: related changes in susceptibil-
ity, resistance frequency and bacterial killing. J Antimicrob Chemother. 2003;52:616–22.
23. Gebru E, Choi M-J, Lee S-J, Damte D, Park SC. Mutant prevention concentration and mech-
anism of resistance in clinical isolates and enrofloxacin/marbofloxacin-selected mutants of
Escherichia coli of canine origin. J Med Microbiol. 2011;60:1512–22.
24. Firsov AA, Strukova EN, Shlykova DS, Portnoy YA, Kozyreva VK, Edelstein MV, Dovzhenko
SA, Kobrin MB, Zinner SH. Bacterial resistance studies using in vitro dynamic models: the
predictive power of the mutant prevention and minimum inhibitory antibiotic concentrations.
Antimicrob Agents Chemother. 2013;57:4956–62.
25. Firsov AA, Portnoy YA, Strukova EN, Shlykova DS, Zinner SH. Predicting bacterial resis-
tance using the time inside the mutant selection window: possibilities and limitations. Int
J Antimicrob Agents. 2014;44:301–5.
664 A. A. Firsov et al.
26. Firsov AA, Portnoy YA, Strukova EN, Shlykova DS, Zinner SH. Bacterial antibiotic resistance
studies using in vitro dynamic models: population analysis vs. susceptibility testing as end-
points of mutant enrichment. Int J Antimicrob Agents. 2015;46:313–8.
27. Strukova EN, Portnoy YA, Romanov AV, Edelstein MV, Zinner SH Firsov AA. Searching for
the optimal predictor of ciprofloxacin resistance in Klebsiella pneumoniae by using in vitro
dynamic models. Antimicrob Agents Chemother. 2016;60:1208–15.
28. Strukova EN, Portnoy YA, Zinner SH, Firsov AA. Predictors of bacterial resistance using
in vitro dynamic models: area under the concentration-time curve related to either the mini-
mum inhibitory or mutant prevention antibiotic concentration. J Antimicrob Chemother.
2016;71:678–84.
29. Firsov AA, Smirnova MV, Strukova EN, Vostrov SN, Portnoy YA, Zinner SH. Enrichment of
resistant Staphylococcus aureus at ciprofloxacin concentrations simulated within the mutant
selection window: bolus versus continuous infusion. Int J Antimicrob Agents. 2008;32:488–93.
30. Golikova MV, Strukova EN, Portnoy YA, Firsov AA. PK/PD modeling as a tool to predict
bacterial resistance to antibiotics: alternative analyses of the experimental data. Antibiot
Chimiother. 2015;60:11–6. rus
31. Firsov AA, Smirnova MV, Lubenko IY, Vostrov SN, Portnoy YA, Zinner SH. Testing the
mutant selection window hypothesis with Staphylococcus aureus exposed to daptomycin and
vancomycin in an in vitro dynamic model. J Antimicrob Chemother. 2006;58:1185–92.
32. Smirnova MV, Vostrov SN, Strukova EN, Dovzhenko SA, Kobrin MB, Portnoy YA, Zinner
SH, Firsov AA. The impact of duration of antibiotic exposure on bacterial resistance predic-
tions using in vitro dynamic models. J Antimicrob Chemother. 2009;64:815–20.
33. VanScoy BD, McCauley J, Ellis-Grosse EJ, Okusanya OO, Bhavnani SM, Forrest A, Ambrose
PG. Exploration of the pharmacokinetic-pharmacodynamic relationships for fosfomycin effi-
cacy using an in vitro infection model. Antimicrob Agents Chemother. 2015;59:7170–7.
34. Firsov AA, Alieva KN, Strukova EN, Golikova MV, Portnoy YA, Dovzhenko SA, Kobrin MB,
Romanov AV, Edelstein MV, Zinner SH. Testing the mutant selection window hypothesis
with Staphylococcus aureus exposed to linezolid in an in vitro dynamic model. J Antimicrob
Chemother. 2017;72:3100–7.
35. Firsov AA, Golikova MV, Strukova EN, Portnoy YA, Romanov AV, Edelstein MV, Zinner SH.
In vitro resistance studies with bacteria that exhibit low mutation frequencies: the prediction
of “antimutant” linezolid concentrations using a mixed inoculum containing both susceptible
and resistant Staphylococcus aureus. Antimicrob Agents Chemother. 2015;59:1014–9.
36. Drlica K, Zhao X. Mutant selection window hypothesis updated. Clin Infect Dis.
2007;44:681–68.
37. Zhao X, Drlica K. A unified anti-mutant dosing strategy. J Antimicrob Chemother.
2008;62:434–6.
38. Blondeau JM. New concepts in antimicrobial susceptibility testing: the mutant prevention con-
centration and mutant selection window approach. Vet Dermatol. 2009;20:383–96.
39. Allen GP, Kaatz GW, Rybak MJ. In vitro activities of mutant prevention concentration-targeted
concentrations of fluoroquinolones against Staphylococcus aureus in a pharmacodynamic
model. Int J Antimicrob Agents. 2004;24:150–60.
40. Olofsson SK, Marcusson LL, Lindgren PK, Hughes D, Cars O. Selection of ciprofloxacin
resistance in Escherichia coli in an in vitro kinetic model: relation between drug exposure and
mutant prevention concentration. J Antimicrob Chemother. 2006;57:1116–21.
41. Firsov AA, Vostrov SN, Lubenko IY, Zinner SH, Portnoy YA. Concentration-dependent
changes in the susceptibility and killing of Staphylococcus aureus in an in vitro dynamic
model that simulates normal and impaired gatifloxacin elimination. Int J Antimicrob Agents.
2004;23:60–6.
42. Campion JJ, McNamara PJ, Evans ME. Evolution of ciprofloxacin-resistant Staphylococcus
aureus in in vitro pharmacokinetic environments. Antimicrob Agents Chemother.
2004;48:4733–44.
43. Gumbo T, Louie A, Liu W, Brown D, Ambrose PG, Bhavnani SM, Drusano GL. Isoniazid
bactericidal activity and resistance emergence: integrating pharmacodynamics and pharma-
21 PK/PD-Based Prediction of “Anti-Mutant” Antibiotic Exposures Using In Vitro… 665
with cephalosporins or gentamicin against Streptococcus mitis group strains in an in vitro model
of simulated endocardial vegetations (SEVs). J Antimicrob Chemother. 2017;72:2290–6.
59. Drusano GL, Neely M, Van Guilder M, Schumitzky A, Brown D, Fikes S, Peloquin C, Louie
A. Analysis of combination drug therapy to develop regimens with shortened duration of treat-
ment for tuberculosis. PLoS One. 2014;9:1–10.
60. Firsov AA, Golikova MV, Strukova EN, Portnoy YA, Dovzhenko SA, Kobrin MB, Zinner
SH. Pharmacokinetically based prediction of the effects of antibiotic combinations on resis-
tant Staphylococcus aureus mutants: in vitro model studies with linezolid and rifampicin.
J Chemother. 2017;29:220–6.
61. Zhanel GG, Mayer M, Laing N, Adam HJ. Mutant prevention concentrations of levofloxacin
alone and in combination with azithromycin, ceftazidime, colistin (polymyxin E), merope-
nem, piperacillin-tazobactam, and tobramycin against Pseudomonas aeruginosa. Antimicrob
Agents Chemother. 2006;50:2228–30.
62. Cai Y, Yang J, Kan Q, Nie X, Wang R, Liang B, Bai N. Mutant prevention concentration of
colistin alone and in combination with levofloxacin or tobramycin against multidrug-resistant
Acinetobacter baumannii. Int J Antimicrob Agents. 2012;40:477–8.
63. Liu LG, Zhu YL, Hu LF, Cheng J, Ye Y, Li JB. Comparative study of the mutant prevention
concentrations of vancomycin alone and in combination with levofloxacin, rifampicin and fos-
fomycin against methicillin-resistant Staphylococcus epidermidis. J Antibiot. 2013;66:709–12.
64. Wei W, Yang H, Hu L, Ye Y, Li J. Activity of levofloxacin in combination with colistin against
Acinetobacter baumannii: in vitro and in a Galleria mellonella model. J Microbiol Immunol
Infect. 2015; doi: https://doi.org/10.1016/j.jmii.2015.10.010.
65. Wu J, Jiang TT, Su JR, Li L. Antimicrobial activity of linezolid combined with minocycline
against vancomycin-resistant Enterococci. Chin Med J. 2013;126:2670–5.
Part IV
Bringing Compounds to Market
Chapter 22
The Role of Pharmacometrics
in the Development of Antimicrobial
Agents
pathogens, enrollment of patients in clinical trials for these agents can be slow and
often requires several years to accrue a modest level of enrollment. This severely
limits the amount of information available to conduct traditional statistical analyses
of clinical data. Moreover, it is unethical to enroll patients into a randomized clini-
cal trial of any design for which one treatment arm is not reliably active against
MDR or XDR pathogens. Consequently, if we are to develop antimicrobial agents
for the treatment of seriously ill patients infected with MDR and XDR pathogens,
the normal paradigm of basing antimicrobial approval on the results of multiple
randomized clinical studies is difficult to impossible. This is ultimately due to the
lack of comparators with suitable efficacy and low numbers of patients with these
infections. In this context, we must therefore consider other data to supply the evi-
dence necessary for antimicrobial drug approval. The focus of this chapter is not
only on basic pharmacometric concepts in the setting of pathogens with usual drug
resistance (UDR) but also on how pharmacometric analyses can be used to leverage
limited clinical data packages in order to support antimicrobial drug approval for
the treatment of patients with infections due to specified MDR or XDR pathogens.
reviewed. The value of these models and breadth of questions they are able to
answer will depend largely upon the richness of the data upon which they are built
and refined. This section will present the consideration needed for ensuring that one
is collecting data which are relevant for answering pivotal questions.
Section 22.4 describes the iterative process of dosing regimen selection. Analyses
to support early-stage dosing regimen selection integrate the aforementioned pre-
clinical information with healthy volunteer PK data using Monte Carlo simulation
in the context of the minimum inhibitory concentration (MIC) distribution(s) for the
target pathogen(s). These analyses should explicitly account for between-species
differences in PK, protein binding, and effect site exposures. The underlying popu-
lation PK model used for the simulations should be developed using a robust clini-
cal dataset, which initially includes data from healthy subjects after receiving single
and multiple doses, and is ultimately refined using data from special populations
and target patient populations treated with dosing regimens intended for labeling.
Finally, Sect. 22.5 will discuss the value of PK-PD analyses based on clinical data in
the context of clinical data packages in the setting of UDR or MDR and/or XDR. The
use of these data to confirm that adequate drug exposures relative to nonclinical PK-PD
targets for efficacy are achieved in the context of both robust and limited clinical data
packages is reviewed. The opportunities for evaluating PK-PD relationships for safety
endpoints and use these data to guide labeling and/or clinical practice guidelines will be
addressed. Finally, the concept of Bayesian analyses, which integrate preclinical and
clinical PK-PD information to inform clinical trial design questions, will be discussed.
672 J. C. Bader et al.
The preclinical PK-PD package for a new drug application (NDA) serves as the
foundation for selecting and supporting dosing regimens for clinical study. These
data are vital to ensuring the success of any drug development program but should
be held in higher regard when developing antimicrobial agents to treat patients with
infections due to MDR and XDR organisms. As will be discussed in greater detail
in Sect. 22.5, clinical data are likely to be limited in such programs; thus, we must
put greater weight on preclinical data to increase regulatory certainty. Consequently,
as described herein, the selection, design, and execution of preclinical studies must
be thoughtfully planned to ensure a robust PK-PD data package is obtained. Such
data will then allow for more informative preclinical inputs for dose selection analy-
ses as described in Sect. 22.3.
To begin formulating a preclinical PK-PD data package, the first question which
must be asked and answered is in regard to the PK-PD index which is most associ-
ated with efficacy for a given antimicrobial. Antimicrobials are typically said to
exhibit concentration- or time-dependent patterns of killing activity [12]. In the case
of antimicrobials with concentration-dependent activity, the rate and extent of kill-
ing increase in tandem with drug concentrations. This pattern of activity is best
described using the ratios of the area under the concentration-time curve (AUC) or
maximum concentrations (Cmax) over the MIC (AUC:MIC and Cmax:MIC ratios,
respectively). The objective when dosing concentration-dependent antimicrobials is
to achieve exposures in patients which maximize the killing of pathogens while
minimizing the likelihood of witnessing drug-induced toxicities.
Alternatively, the objective when administering antimicrobials which exhibit
time-dependent killing is not to maximize drug exposures but rather to optimize
dosing to maintain drug concentrations above a target threshold such as an
MIC. Accordingly, this pattern of activity can be characterized by the percentage
of time drug concentrations remain above an MIC or other threshold (%T > MIC
and %T > threshold, respectively). Jointly, the AUC:MIC ratio, Cmax:AUC ratio,
and %T > MIC comprise the three most commonly utilized PK-PD indices to
describe antimicrobial activity (Fig. 22.2). Determining the PK-PD index which
best describes efficacy for a given antimicrobial can be challenging if one does
not take extra precautions. Given that the magnitude of drug introduced into a
system impacts all of the aforementioned PK-PD indices, significant collinearity
is observed when attempting to differentiate these indices on the basis of dose
alone. That is to say AUC:MIC ratio, Cmax:AUC ratio, and %T > MIC all increase
with dose. Dose-fractionation studies are used to mitigate the impact of this col-
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 673
Fig. 22.3 Relationships between change in log10 CFU from baseline at 24 h and total-drug
AUC:MIC ratio, Cmax:MIC ratio, and %T > MIC for anidulafungin against C. glabrata based on
data from a neutropenic murine candidiasis model [13]. (Reproduced from Ref. [14] with permis-
sion from J Antimicrob Chemother. Copyright © 2017 British Society for Antimicrobial
Chemotherapy, [Journal of Antimicrobial Chemotherapy, 2018; 73 (suppl 1):i44-50.])
linearity through the administration of dosing regimens which utilize the same
total dose of an antimicrobial agent but which are differentiated in their frequency
of dosing (e.g., 600 mg once daily, 300 mg twice daily, 150 mg four times daily,
etc.). A range of exposures is obtained by administering regimens in a similar
manner over multiple dose levels.
Using data obtained from a dose-fractionation study that was conducted using
a neutropenic murine candidiasis model [13], Fig. 22.3 shows relationships
between change in log10 CFU from baseline at 24 h and total-drug AUC:MIC
ratio, Cmax:MIC ratio, and %T > MIC for anidulafungin against C. glabrata [14].
674 J. C. Bader et al.
In this study, neutropenic mice were infected with Candida glabrata and admin-
istered 1 of 20 anidulafungin dosing regimens. Total doses of 1.25, 5, 20, 80, or
320 mg/kg were administered over a 96-h period in the form of one, two, four, or
six divided doses (i.e., doses were given every 96, 48, 24, or 16 h, respectively).
Hill models were used to characterize the relationships between changes in fungal
density (i.e., colony-forming units [CFU]) in homogenized kidney tissue at 96 h
relative to baseline and the three aforementioned PK-PD indices. The data pre-
sented demonstrated that changes in fungal density were most closely associated
with anidulafungin AUC:MIC and Cmax:MIC ratios, indicating that this agent
exhibits a concentration-dependent pattern of fungicidal activity. However, unlike
AUC values, the Cmax is achieved at a transient time point, making it difficult to
accurately capture in studies and apply to support dose selection. Therefore, in
such situations, AUC:MIC ratio serves as a more reliable and predictable PK-PD
index than Cmax:MIC ratio.
Once the PK-PD index most associated with efficacy is known, the next step is to
determine the magnitudes of this index which are associated with various levels of
pathogen killing. These thresholds are commonly known as PK-PD targets for effi-
cacy, and they provide crucial information to assist in estimating the likelihood of
achieving efficacious drug concentrations in patients following the administration
of a given antimicrobial dosing regimen. Dose-ranging studies are used to derive
these PK-PD targets, wherein changes in microbial density are evaluated across a
wide range of antimicrobial doses. Given that the PK-PD index most associated
with efficacy is known by this point in time, there is no longer a need to account for
potential collinearities. Consequently, all doses are administered over identical dos-
ing intervals (e.g., every 24 h). The interval evaluated in these studies will be that
which best describes the relationship between change in bacterial burden and the
PK-PD index most associated with efficacy, as established by the results obtained
from prior dose-fractionation studies. Common thresholds assessed in these studies
include net stasis (i.e., no change in the density of bacteria or fungi from that
observed at baseline) and 1- and 2-log10 reductions in the counts of CFUs relative to
baseline observations.
Figure 22.4, which shows data from VanScoy et al. [15], illustrates the type
of data that can be derived from an in vitro dose-ranging study. In these studies,
a one-compartment in vitro infection model was used to simulate total-drug epi-
thelial lining fluid (ELF) AUC values ranging from 33.3 to 7942 mg•h/L follow-
ing administration of arbekacin, an investigational aminoglycoside. These drug
exposures were evaluated against four Pseudomonas aeruginosa isolates, the
MIC values for which ranged from 2 to 8 mg/L. The relationship between
change in log10 CFU from baseline at 24 h and total-drug ELF AUC:MIC ratio,
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 675
Fig. 22.4 Arbekacin total-drug ELF AUC:MIC0–24 ratio targets associated with net bacterial stasis
and 1- and 2-log10 CFU reductions from baseline for P. aeruginosa based on data from a one-
compartment in vitro infection model [15]
the PK-PD index associated with arbekacin efficacy, was evaluated using a Hill
model. Using this model, the magnitudes of total-drug ELF AUC:MIC ratio
associated with net b acterial stasis and 1- and 2-log10 CFU reductions from
baseline, which were 56.9, 142, and 393, respectively, were identified as shown
in Fig. 22.4.
These data exhibit several characteristics which indicate the robustness of the
above-described AUC:MIC ratio targets. We can be assured that the PK-PD rela-
tionship was well captured as evidenced by the nearly complete sigmoidal curve
obtained by the fitted Hill model. This is a product of designing the dose-ranging
study to obtain a wide range of AUC:MIC ratios by evaluating a large range of doses
and various isolates with differing MIC values. Moreover, the coefficient of deter-
mination (r2) of 0.856 for this relationship was high, which tells us that the relation-
ship between change in log10 CFU and the AUC:MIC ratio is strong. Finally, the
data pertaining to each of the various isolates evaluated are well dispersed along the
fitted relationship with no apparent trends, indicating that no substantial differences
in efficacy were observed across these isolates.
676 J. C. Bader et al.
Devoting time and resources to the design of studies that account for variability in
PK-PD relationships for efficacy is crucial to the development of a robust preclini-
cal PK-PD package. Bacteria and fungi are extremely complex and adaptive organ-
isms that can develop a myriad of antimicrobial resistance mechanisms and undergo
changes in their inherent fitness. Consequently, the variability among isolates for a
given pathogen needs to be considered when designing preclinical studies. The con-
sideration of such variability provides an opportunity to better characterize the
PK-PD of a given antimicrobial agent. The following will detail best practices for
designing PK-PD studies in order to maximize the information that can be gained in
light of the above-described variability.
To begin, let us review the design of dose-fractionation studies and consider how
best to select an isolate for evaluation. Given that the primary objective when con-
ducting these studies is to discriminate among the various PK-PD indices and deter-
mine which is most associated with efficacy for an antimicrobial, the intention
should be to minimize the potential of generating variable and inexplicable results.
Therefore, when evaluating an antimicrobial agent, it is best to select a well-defined
isolate that is known to grow well in the in vitro or in vivo system intended for study
and for which consistent and predictable PK-PD data have been generated previ-
ously (e.g., as observed in prior time-kill studies).
Regarding dose-ranging studies, the objective when selecting isolates should be
to study a diverse collection of isolates such that inter-isolate variability can be
adequately characterized. The challenge panel of isolates should have MIC values
that encompass a clinically relevant range and that express applicable resistant
determinants. Given that these studies are employed to derive PK-PD targets associ-
ated with efficacy which are then used to forecast dosing regimens for patients in the
UDR setting or even the setting of MDR and XRD, it is important to account for the
population of pathogens expected in either of these clinical settings. Examples of
isolate collections used for PK-PD analyses that meet the above-described criteria
are described below.
Figure 22.5 shows the relationship between change in log10 CFU from baseline
and free-drug plasma AUC0–24:MIC ratio for an investigational anti-staphylococcal
agent, afabicin, against seven Staphylococcus aureus isolates with MIC values rang-
ing from 0.004 to 0.06 mg/L, the data for which was obtained from studies utilizing
a murine-thigh infection model [16]. When assessed relative to the range of MIC
values (<0.001–0.25 mg/L, MIC90 = 0.008 mg/L) evaluated in a recent in vitro sur-
veillance study of 660 S. aureus isolates collected from European, North American,
Latin American, and Asian hospitals from 2013 to 2014 [17], the range of MIC
values for the isolate collection studied was considered robust. In addition, the latter
consideration is presented in Fig. 22.6 which shows data obtained from one-
compartment in vitro infection model studies of eravacycline, an investigational
tetracycline, against five Escherichia coli isolates stratified by those that were tetra-
cycline susceptible and non-susceptible based on either the Clinical and Laboratory
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 677
Fig. 22.5 Relationship between the change in log10 CFU from baseline at 24 h and free-drug
plasma AUC0–24:MIC ratio for afabicin based on data for seven S. aureus isolates studied in a neu-
tropenic murine-thigh infection model [16]
tam, ceftaroline concentrations were much lower than those expected. This is due to
ceftaroline being hydrolyzed by the endogenous β-lactamase enzymes secreted by
the above-described K. pneumoniae isolate. These data underscore the importance
of collecting PK as part of all in vitro and in vivo infection model studies. Had these
investigators not collected PK samples during their study, they would have greatly
overestimated ceftaroline exposures achieved in the in vitro infection model.
The evaluation of in vivo PK is critical to conducting PK-PD analyses based on
data derived from in vivo infection models. Thus, careful consideration needs to be
given to the design of such studies to ensure that useful data are generated. To this
end, measures can be taken to mitigate sources of variability in PK of the antimicro-
bial agent by controlling for factors such as route of administration, species, sex,
and weight. The use of genetically modified animals of the same species, sex, and
weight has allowed for control of the latter set of factors. Along these same lines, the
route of drug administration is another important consideration. While intraperito-
neal injections are routinely utilized due to the relatively large potential space for
injection and the ease and rapidity with which drug administration can be carried
out, training of research staff would be required to ensure proper technique in order
to prevent drug administration into abdominal organs or adipose tissue. Intravenous
(IV) injections are also desirable for the rapid delivery of drug directly into the
bloodstream. However IV injections can be extremely difficult to administer to
smaller species such as BalbC as compared to CD-1 mice. Subcutaneous injections
are often the most preferred and relatively straightforward injection site, resulting in
minimal variability among injections and technicians. Other factors to consider that
can affect PK include the potential for drug interactions between the antimicrobial
agent and analgesic or anesthetic agent.
680 J. C. Bader et al.
Fig. 22.8 Targeted (solid line) concentration-time profiles with observed concentrations (sym-
bols) overlaid for ceftaroline (triangles) and avibactam (NXL104, squares) among in vitro treat-
ment arms including both agents (C to G) or ceftaroline or avibactam alone (B and H, respectively).
(Reproduced from Ref. [22] with permission from Antimicrob Agents Chemother. Copyright ©
American Society for Microbiology)
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 681
Other study design elements to consider include infection status and PK sam-
pling. Although in vivo PK studies are often carried out in noninfected animals, it
can be advantageous to evaluate PK in infected animals. Moreover, when appli-
cable, PK samples from the effect site of interest should be collected in addition
to blood samples (e.g., ELF in the case of murine-lung infection models). Intensive
PK sampling enables non-compartmental analyses (commonly referred to as
“SHAM analyses”) of the PK data. Non-compartmental analyses allow for lines
across the individual or pooled concentrations obtained from animals to be con-
nected, thus allowing for the shape of the concentration-time profile(s) to be char-
acterized. Based on these data, exposure measures such as the AUC and Cmax can
be calculated. While PK sampling is typically intensive, it is also possible to
design less intensive PK sampling strategies and use compartmental analyses to
derive drug exposures. Compartmental analyses offer the benefit of allowing for
more precise estimates of drug exposure with less reliance on intensively sampled
PK from study animals. If non-compartmental analyses are used to evaluate PK
data, intensive sampling around multiple doses may be required to ensure reliable
estimates of drug exposure. Consider the case presented in Fig. 22.9 in which
animals were dosed every 12 h over a 48-h period, but PK were only intensively
sampled following the first and third doses. If a non-compartmental analysis were
used to evaluate these data, exposures would be greatly underestimated given that
the concentrations following the second and fourth doses would be largely
ignored. Thus, it is important to ensure that the PK sampling strategy is adequate,
682 J. C. Bader et al.
Fig. 22.10 Changes in the total and resistant bacterial populations over time for single ceftolozane, piperacillin-tazobactam, and control dosing regimens (a)
683
and a range of ceftolozane-tazobactam dosing regimens (b). (Reproduced from Ref. [23] with permission from Antimicrob Agents Chemother. Copyright ©
American Society for Microbiology)
684 J. C. Bader et al.
Selection of a dosing regimen for Phase 2 and 3 trials is a critical decision in the
development of new antimicrobials. The dose selected needs to be low enough to
prevent severe drug-related toxicities, yet high enough to achieve efficacy in the
majority of patients. Ideally, this dose should also be of sufficient magnitude to
prevent the emergence of on-therapy resistance. The first step in selecting a clinical
dosing regimen is to understand the PK of the drug in humans, including PK vari-
ability and factors contributing to this variability. Once this is known and thor-
oughly understood, simulations of various proposed dosing regimens can be
conducted as will be discussed in Sect. 22.4.
Population PK modeling is the current gold standard for performing PK analyses
in this context. Although simpler tools are available (such as non-compartmental,
naïve pooling, and standard two-stage approaches), a population PK model offers a
few key advantages. The first of these is that the population approach allows for the
quantification of between-subject variability. This is particularly useful when simu-
lating various dosing regimens as one can estimate the width and shape of the
expected exposure distribution. The second advantage is the ability to quantify and
explain variability in PK through factors such as body size, age, gender, or clearing
organ function (i.e., patient-specific covariates). Again, this is useful when perform-
ing simulations as one can determine if the same dose can be administered to all
patients or if dose should be adjusted based on patient characteristics (e.g., weight-
based dosing or adjustments based on patient renal function). The final key advan-
tage of population PK models is that they can be informed by sparsely collected PK
data based on an optimized sampling scheme. This is especially important when
utilizing data from Phase 2 and 3 trials as it is rarely feasible to collect rich PK data
in all patients in these studies due to ethical and logistical constraints.
In essence, population PK models use differential equations to describe the time
course of drug concentrations across a population of subjects. Traditionally, com-
partmental models are utilized, wherein the model compartments represent hypo-
thetical spaces to which drug may distribute (e.g., vasculature, tissue, macrophages,
etc.). The central compartment most often represents the vascular space and all
areas of the body where drug rapidly equilibrates. Various peripheral compart-
ments may be added to the base model to represent areas of the body with slower
distribution characteristics. In addition, absorption compartments can be added to
characterize the time course of drug disposition following various routes of admin-
istration such as absorption through the gastrointestinal tract after oral administra-
tion. Each compartment is represented with a differential equation using various
PK parameters (e.g., clearance or volume of distribution) to describe the move-
ment of drug into and out of the compartment. The values of these parameters can
vary across subjects. For instance, take an oral antibiotic with a concentration-time
profile which can be described with a one-compartment model. The model can be
described with two differential equations – one for the gut/gastrointestinal tract
and one for the plasma and other places in the body where the drug rapidly distrib-
686 J. C. Bader et al.
utes. The differential equations often use a “ka” parameter to describe absorption
from the gut/gastrointestinal tract, a volume of distribution term to relate drug
amounts to measured drug concentrations, and a clearance term to describe removal
of the drug from the body. Using a population PK model, one can determine a
mean parameter value across all subjects (known as the “population mean value,”
as shown in Fig. 22.11). In addition, the model can also estimate an individual
parameter estimate for each subject, as shown in Fig. 22.11. The distribution
around the population mean value is termed “interindividual variability.”
Quantification of this variability is crucial for dose selection analyses, which are
discussed further in Sect. 22.4.
The development of the population PK model should be an iterative process,
beginning during Phase 1 development. If the studies are designed appropriately,
analysis of Phase 1 data should allow for the identification of anomalies in PK, such
as nonlinearity or nonstationary, early in drug development. Nonlinearity refers to a
change in exposure which is not proportional to the change in dose. For example, a
doubling of dose would typically be expected to cause a doubling of the AUC, but
in the case of a drug which exhibits nonlinear PK, the AUC might only increase
slightly following this doubling of dose. Oftentimes, this is due to saturation of an
absorption process in the gastrointestinal tract following oral administration. More
troubling are instances in which the dose doubles, but the resulting AUC is more
than doubled. This is often due to the saturation of an elimination pathway.
Nonstationary, on the other hand, refers to PK parameters that change with time
irrespective of dose. For example, a drug with auto-induction of clearance can result
in lower exposures on Day 5 of therapy relative to Day 1, even if the same dose is
administered in both cases. Other common PK issues which can be observed in
Phase 1 studies are food and diurnal effects. All of the abovementioned PK com-
plexities can be built into the population PK model structure.
Throughout the iterative process of developing and refining a structural popula-
tion model, covariate analyses can be conducted. Covariate analyses allow for iden-
tification and evaluation of factors that explain variability in the PK parameters.
Typical covariates evaluated include body size measures [e.g., weight, height, body
surface area, body mass index (BMI), lean body weight], sex, race, age, clearing
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 687
organ function (e.g., creatinine clearance or hepatic function tests), and genotypes.
Trends between covariates and individual PK parameters are explored. Any
covariate-parameter pairs with a considerable relationship are tested in the model,
and those which are found to have statistically significant relationships remain in
the model. Oftentimes, the value of these analyses may be limited during early-
stage development when data stem largely from Phase 1 studies comprised of
largely homogenous healthy volunteers. However, covariate analyses become more
informative as they are enriched with data obtained from special populations such
as patients with renal impairment or obesity and Phase 2 and 3 studies, which enroll
large numbers of patients with greater covariate variability.
As evidenced by results of analyses that have demonstrated differences in clear-
ance, volume of distribution, and tissue penetration in patients compared to healthy
volunteers [27–29], differences in PK for patients relative to healthy subjects may be
observed. This is not surprising given the physiologic changes observed in patients,
particularly those who are critically ill. In infected and acutely ill patients, greater PK
variability has been observed [30–32]. Regardless of the scenario, these differences
can be quantified by refining the population PK model. Thus, after development of
the Phase 1 population PK model, it is important that the model is refined using data
collected from Phase 2 and/or Phase 3 studies. Phase 2 is typically the first time the
investigational agent is administered to patients rather than healthy volunteers.
However, in the case of the development of antimicrobial agents for MDR and XDR
pathogens, the conduct of Phase 1b studies conducted in patients prior to small Phase
3 clinical studies is increasingly becoming commonplace.
Lastly, it is important to update the results of the covariate analysis after refining
the population PK model as there is typically a wider range of covariates observed
in Phase 2 and 3 studies. For example, the protocol for Phase 1 studies usually
excludes patients with a BMI above 30 kg/m2. However, in Phase 2 and 3, often-
times there is no exclusion criteria based on BMI. Important relationships between
BMI and PK parameters such as clearance may not be evident until a wider range of
BMI data are included in the analysis datasets. In addition to the wider range of
covariates for patients in Phase 2 and 3 studies relative to subjects enrolled in Phase
1 studies, additional covariates which could not be assessed in healthy volunteers
(e.g., renal function, APACHE II score, and/or infection type) can be evaluated.
of dosing regimens and develop in vitro susceptibility testing criteria (i.e., suscepti-
bility breakpoints) [35]. In this section, we will briefly outline the use of Monte
Carlo simulation to carry out PK-PD target attainment analyses and discuss how the
results of these analyses may be used to evaluate dosing regimens and inform the
selection of interpretative criteria for in vitro susceptibility testing.
The optimal application of Monte Carlo simulation for dose selection evalua-
tions requires the use of compartmental models as previously discussed in Sect.
22.3. A population PK model for the antimicrobial under investigation can be used
to generate drug exposures of interest for simulated patients. As discussed previ-
ously, interindividual variability can exist on many of the parameters within popula-
tion PK models. As illustrated in Fig. 22.11, by randomly assigning parameter
estimates for simulated patients based on the distributions for these parameters,
interindividual variability can be considered. As shown in Fig. 22.12, PK parameter
estimates assigned to simulated patients can be used to generate concentration-time
profiles after administration of the dosing regimen of interest, which in this example
was an intravenously administered antibiotic infused over 2 h twice daily. The dis-
tribution of these concentration-time profiles, as represented by the median and 5th
and 95th percentiles, allows for an understanding of the variability expected in the
actual patient population.
These above-described concentration-time profiles can be used in conjunction
with fixed MIC values to calculate the PK-PD index associated with efficacy for a
given pathogen by MIC for individual dosing regimens administered to each sim-
ulated patient. The percentage of simulated patients achieving a given PK-PD
target by MIC is then determined. As shown by data assessing the percent proba-
bilities of PK-PD target attainment by MIC for ceftaroline relative to nonclinical
%T > MIC targets for S. aureus shown in Fig. 22.13 [35], these data are com-
monly interpreted in the context of observed MIC values for isolates based on
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 689
Fig. 22.13 Percentage of simulated patients with normal renal function (80≤ creatinine clearance
≤170 mL/min/1.73 m2) achieving free-drug (f) %T > MIC targets by MIC following administra-
tion of ceftaroline fosamil 600 mg q12h, overlaid on a histogram showing the MIC distributions
for MRSA and MSSA. (Reproduced from Ref. [35] with permission from Antimicrob Agents
Chemother. Copyright © American Society for Microbiology)
PK-PD target attainment. Endpoints for such PK-PD targets range from net bacte-
rial stasis to a 2-log10 CFU reduction from baseline. Typically, these data are derived
from neutropenic murine-thigh or murine-lung infection models or when warranted,
in vitro infection models. Net bacterial stasis has been suggested to be an appropri-
ate endpoint for a PK-PD target when selecting dosing regimens to treat patients
with infections associated with lower bacterial inoculums and/or for which source
control, including surgical intervention, is an option. This endpoint may also be
reasonable to assess for inferences about patient populations that are expected to be
immunocompetent and for whom the response rate associated with no treatment is
expected to be relatively high (e.g., ≥60%). Examples of indications that meet these
criteria include acute bacterial skin and skin structure infections (ABSSSI), cIAI,
and cUTI. Reduction of 1-log10 CFU from baseline has been suggested to be an
appropriate endpoint for a PK-PD target when selecting dosing regimens to treat
patients with infections associated with higher bacterial inoculums such as pneumo-
nia, endocarditis or bacteremia, and/or for infected patients who are
immunocompromised. In such populations, the response rate associated with no
treatment may be low (e.g., ≤40%) [11, 37, 38].
Support for each of the above-described endpoints is based on successful
translations between the results of previous PK-PD analyses based on nonclini-
cal and clinical data [3, 39–42]. Results of analyses of these data have demon-
strated that the same magnitude of the PK-PD target associated with net bacterial
stasis from neutropenic murine-thigh infection models for a given antimicrobial
agent was associated with a high percentage of successful outcomes among
patients with cIAI or ABSSSI [3, 39–42]. The choice of a 1-log10 CFU reduction
from baseline for the treatment of patients with infections with a higher no-
treatment response is based on an assessment of PK-PD target attainment analy-
sis results for antimicrobial agents that were evaluated for pneumonia. As the
percent probability of achieving a PK-PD target associated with a 1-log10 CFU
reduction from baseline increased, so too did the probability of a successful
regulatory outcome [2]. The latter was considered an indicator of meeting non-
inferiority in pivotal clinical trials.
While a 2-log10 CFU reduction from baseline has been suggested as an endpoint
for indications such as HABP/VABP [11], attainment of a PK-PD target associated
with such a level of bacterial reduction may not be possible for many antimicrobial
agents, including those currently available and commonly used for these indica-
tions. As previously shown for a meropenem dosing regimen of 2 g q8h infused over
1 h [4], while it is possible to achieve the %T > MIC target associated with a 2-log10
CFU reduction from baseline, large interpatient variability can hinder the likelihood
of achieving this PK-PD target in many patients. However, one strategy to overcome
this is to administer the same dose as a prolonged infusion over 3 h [43]. Such a
strategy was employed for development of meropenem-vaborbactam, a β-lactam/β--
lactamase inhibitor combination recently approved by the US FDA [44]. From a
drug development perspective, the margin of safety should be weighed against goals
for efficacy when considering an endpoint of a 1- vs 2-log10 CFU reduction from
baseline for indications such as HABP/VABP [38]. If an antimicrobial agent has a
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 691
wide safety margin, developers have a greater opportunity to utilize a 2-log10 CFU
reduction endpoint.
In addition to the PK-PD target, it is important to consider exposures at the effect
site when applicable. For example, if an antimicrobial agent is being developed to
treat patients with pneumonia, it is important to evaluate the likelihood of achieving
efficacious drug concentrations in ELF. To enable the consideration of ELF expo-
sures, PK data from healthy volunteers and if available, from patients, should be
considered when developing the population PK model. This model would then be
used together with ELF PK-PD targets associated with efficacy derived from a
murine-lung infection model and Monte Carlo simulation to assess PK-PD target
attainment for dosing regimens.
As described above, the assessment of dosing regimens to evaluate in Phase 2
or 3 studies requires the use of a population PK model constructed using PK data
from healthy volunteers enrolled in Phase 1 studies. As such, the interindividual
variability in PK will be limited and may not be reflective of the target patient
population. In such cases, inflating variance in PK parameters (e.g., increasing
the interindividual variability terms on PK parameters such as clearance and vol-
ume) as a part of a sensitivity analysis may be a useful approach to further dis-
criminate among candidate dosing regimens [38]. Another limitation of a
population PK model developed using Phase 1 PK data is that covariate distribu-
tions are relatively narrow. Since studies for special populations, including sub-
jects with renal or hepatic impairment, are typically not completed early in a
clinical development program, the evaluation of covariates is not usually avail-
able until late-stage development. Thus, early-stage development decisions for
dose selection should be confirmed after a population PK model has been refined
using data from the target patient population and special populations. Additionally,
an understanding of covariates that are highly influential on PK allows for the
assessment of dosing regimens in simulated patients stratified by ranges of such
covariates to support dosing recommendations for special populations. Data
from simulated patients can be used to support dosing recommendations even if
such dosing regimens were not assessed in clinical trials. This strategy was used
for delafloxacin, a fluoroquinolone that was recently approved by the US FDA
for the treatment of patients with ABSSSI. The delafloxacin dosing regimen
approved for patients with severe renal impairment [45], 450 mg by mouth twice
daily, was not studied in clinical trials [46, 47] but was supported by the results
of population PK and PK-PD target attainment analyses [48].
The above-described strategy to use preclinical PK-PD data, population PK
models, and Monte Carlo simulation both for early- and late-stage development
decisions about dose selection allows developers to mitigate risk and increase the
likelihood of regulatory success. The results of such analyses can also be used to
inform recommendations for interpretative criteria for in vitro susceptibility testing
criteria for the antimicrobial agent of interest against target pathogens. The data
obtained from these simulations can be used in conjunction with clinical outcome
data by MIC and pathogen susceptibility distributions to support susceptibility
breakpoint decisions. Results of PK-PD target attainment analyses to support sus-
692 J. C. Bader et al.
Irrespective of whether the clinical data package is robust or limited, the data derived
from PK-PD analyses are valuable. However, as described below, objectives of such
analyses will vary depending on the data package available. Important prerequisites
for both types of clinical data packages include the collection of PK data from all
patients and the evaluation of informative endpoints. As described in Sect. 22.3, the
benefit of developing a population PK model based on Phase 1 data is that the model
can be used to determine sparse PK sampling strategies for implementation in clini-
cal trials. Such strategies are designed to ensure that optimal information to estimate
drug exposure in each patient is obtained using a minimal number of blood samples
for drug assay as possible. Using these sparse PK data, the goal of additional popu-
lation PK analyses is to refine the existing model developed using Phase 1 data in
order to enable precise and unbiased estimation of drug exposure in individual
patients, including the applicable PK-PD index for efficacy (e.g., AUC:MIC ratio,
Cmax:MIC ratio, or %T > MIC).
In addition to reliable estimates of drug exposure, well-defined and reproducible
efficacy and safety endpoints are needed to evaluate PK-PD relationships for such
endpoints. Objective criteria, determined by observations collected at informative
time points, are required to assess drug effect. Clinical trial endpoints for efficacy
are typically categorical variables, such as clinical response to therapy (success or
failure) assessed at the test-of-cure visit (i.e., a window of time after the end of study
drug; TOC) and/or at the end of therapy. However, recent US FDA guidance for a
number of indications has described the assessment of efficacy endpoints evaluated
earlier in therapy [5, 6]. For patients with ABSSSI and CABP, clinical response is
assessed on Days 2 to 3 and 3 to 5, respectively. PK-PD relationships for efficacy
have been largely described using dichotomous efficacy endpoints assessed at TOC
[3, 31, 41, 51–54]. In contrast, there is comparatively less experience evaluating
efficacy endpoints assessed earlier in therapy [55]. Despite the lack of experience
with the latter, given the natural course of infection, which involves eradication of
the pathogen followed by macrophage and inflammatory modulator activity, which
is then followed by resolution of signs and symptoms, it may be difficult to identify
PK-PD relationships for efficacy early after therapy has been initiated [56–58].
Consequently, the time at which efficacy is assessed can influence the likelihood of
identifying PK-PD relationships for efficacy.
While dichotomous efficacy endpoints are typically evaluated in clinical trials
for antimicrobial agents and serve as primary endpoints upon which sample size is
determined, the evaluation of continuous or time-to-event efficacy endpoints can
also be informative. Examples of continuous endpoints include change in bacterial
density or lesion size, while examples of time-to-event endpoints include time to
resolution of signs and symptoms, lesion size reduction, or bacteriologic eradica-
tion. Continuous or time-to-event endpoints have the benefit of being more sensitive
than categorical endpoints for capturing drug effect. When measures of efficacy are
assessed serially, this provides the opportunity to identify the time period during
694 J. C. Bader et al.
which treatment effect is greatest [55, 59]. Evaluation of such endpoints for PK-PD
analyses for efficacy has the potential to inform decisions about dose and duration
using a relative smaller sample size than that for a dichotomous efficacy endpoint
[59]. For example, while evaluations of clinical or microbiological response for 38
tigecycline-treated patients with CABP failed to reveal PK-PD relationships for
efficacy, a relationship between free-drug AUC:MIC ratio and time to fever resolu-
tion was identified [60]. The median time to fever resolution was 12 and 24 h for
patients with a free-drug AUC:MIC ratio >12.8 and ≤12.8, respectively. Thus,
despite not representing the primary clinical trial endpoint for efficacy, relationships
for such endpoints can be used to support dose selection and even provide potential
insights about the duration of therapy.
For the assessment of PK-PD relationships for safety endpoints, the same prin-
ciples as described above are applicable. While safety endpoints, such as the pres-
ence or absence of a given safety event or a dichotomous threshold for a continuous
laboratory measure, are dichotomous in nature, the assessment of continuous
endpoints including laboratory measures or physiologic measurements such as
blood pressure collected serially provides the opportunity to develop informative
multivariable models [61, 62]. Such models, which can be constructed to describe
the effect of varying drug exposures on laboratory measures over the course of
therapy in the context of other patient factors, can then be applied to simulated data
to discriminate among potential dosing regimens to be studied in Phase 3 trials. For
example, the percentage of simulated patients with laboratory measures above clini-
cally relevant folds of the upper limit of normal (ULN) (e.g., 3, 5 or 10 × ULN) or
in the case of systolic blood pressure, the percentage of patients with readings
≥160 mmHg can be determined for individual dosing regimens. This information,
together with assessments of the percent probabilities of achieving each efficacy
endpoints (based on clinical PK-PD relationships for efficacy) and/or nonclinical
PK-PD targets, can be used to balance considerations for safety and efficacy. Or
using multivariable models developed using Phase 2 and/or 3 data, percent proba-
bilities of elevation of these safety endpoints can be evaluated among all simulated
patients and subgroups at increased risk who receive intended dosing regimens. The
identification of patient populations at increased risk and the characterization of the
elevations for such safety endpoints can be used to inform use for labeling and/or
clinical practice guidelines.
As described above, the robustness of the clinical data package guides the objectives
of the PK-PD analyses for efficacy. For antimicrobial agents for patients with infec-
tions arising from pathogens in the setting of UDR, the sample size of evaluable
populations is expected to be sufficient to support the assessment of PK-PD rela-
tionships for efficacy. Thus, the objective is to determine if PK-PD relationships for
efficacy endpoints can be identified. However, despite the robustness of the sample
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 695
size of analysis populations, other factors may influence the ability to characterize
PK-PD relationships, including the duration of therapy. It is important to evaluate
patients from the microbiologically evaluable population in order to consider
patients who received a sufficient number of doses and who had pathogen(s) iso-
lated at baseline. The former ensures that the lack of clinical response is not attrib-
uted to insufficient duration of drug exposure, and the latter allows for drug
exposures to be indexed to pathogen MIC values in order to enable PK-PD indices
to be determined. For infections for which there are baseline pathogens with antici-
pated or known (as determined by preclinical data) PK-PD characteristics that dif-
fer, evaluation of subpopulations may be necessary to characterize PK-PD
relationships for individual pathogens. Or in the setting of infections with polymi-
crobial pathogens, careful consideration needs to be given to how the primary
pathogen used for calculating the PK-PD index is identified. Additionally, consider-
ation needs to be given to the definitions for clinical failure. If the reasons for declar-
ing a clinical failure include those not related to study drug (e.g., an adverse events),
data for patients failing for these reasons should be excluded given the potential for
these data to impede the ability to identify PK-PD relationships.
Finally, and perhaps most importantly, while it is important to consider all of the
above-described factors to ensure that every opportunity has been provided to allow
for PK-PD relationships for efficacy based on a robust clinical data package to be
identified, the lack of identification of PK-PD relationships for efficacy is a predict-
able outcome when patients have received PK-PD optimized dosing regimens. In such
cases, it is still valuable to demonstrate that drug exposures from patients indexed to
MIC values from pathogens identified at baseline exceed nonclinical PK-PD targets
for efficacy to confirm the basis for dose selection. When PK-PD relationships for
efficacy are identified based on clinical data from patients who received PK-PD-
optimized dosing regimens, these relationships are usually based on a dichotomized
variable for the PK-PD index of interest, the threshold for which is optimally deter-
mined using a number of statistical approaches. These approaches can include using
the threshold of the PK-PD index representing the first split of a classification or
regression tree, a receiver operating characteristic curve, or using a model fit to esti-
mate a threshold for achieving a target efficacy outcome or probability. PK-PD rela-
tionships identified in this manner resemble step functions and allow for patients with
both lower PK-PD indices and percentages of successful response to be contrasted
from those with higher PK-PD indices and percentages of successful response [51,
59]. Table 22.1 summarizes the results of two separate PK-PD analyses of Phase 3
data for patients who received PK-PD optimized dalbavancin or oritavancin regimens.
In both cases, PK-PD relationships identified were based on two-group variables for
AUC:MIC ratio [63, 64]. The differences in the percentage of successful clinical
responses between patients in the lower and higher exposure groups were 10.9 and
13.6%, respectively, with percentage of patients with successful clinical response in
the lower AUC:MIC ratio groups of 89.1 and 82.6% for dalbavancin and oritavancin,
respectively. Thus, when PK-PD-optimized dosing regimens are studied and PK-PD
relationships based on two-group variables are identified, the differences between the
lower and higher exposures groups are unlikely to be impressive.
696 J. C. Bader et al.
Table 22.1 Summary of PK-PD relationships for efficacy for dalbavancin and oritavancin based
on dichotomous two-group AUC:MIC ratio variables
Percentage of patients <
or ≥ threshold achieving
Threshold the efficacy endpoint
value of (n/N)
Antimicrobial Efficacy PK-PD ≥
agent endpoint PK-PD index index < threshold threshold P-value
Dalbavancin Clinical AUCavg:MIC 21,267 89.1 (98/110) 100 0.01
[63] success at the ratiob (52/52)
test-of-cure
visita
Oritavancin Clinical AUC0– 11,982 82.6 (19/23) 96.2 0.029
[64] success at the 72:MIC ratiod (126/131)
post-therapy
evaluationc
a
The test-of-cure visit occurred 14 days [± 2 days] after the end of therapy
b
AUCavg:MIC ratio was calculated by dividing the average 24-h AUC from 0 to 120 h by the base-
line MIC of the infecting pathogen
c
The post-therapy evaluation occurred 7–14 days after the end of therapy
d
AUC0–72:MIC ratio was calculated by dividing the AUC from 0 to 72 h by the baseline MIC of the
infecting pathogen
Given the current paradigm for obtaining robust preclinical PK-PD data and using
these data with Phase 1 PK data and Monte Carlo simulation to predict doses for
Phase 2 and 3 clinical trials, the likelihood for failed clinical trials has been reduced.
Evaluation of data based on contemporary clinical trials that did not make full use
of these approaches to select dose, together with innovative statistical approaches,
provides the opportunity to answer questions about the no-treatment effect. Such
data represent valuable inputs for power and sample size calculations for future
clinical trials in the setting of UDR. As described below, data for tigecycline from
61 patients with HABP/VABP who were microbiologically evaluable and who had
sufficient PK data, the clinical trial that failed to demonstrate non-inferiority
22 The Role of Pharmacometrics in the Development of Antimicrobial Agents 697
Fig. 22.14 Frequentist (A) and Bayesian (B) logistic regression-estimated relationships between
clinical response and the tigecycline free-drug AUC:MIC ratio based on data from 61 patients with
HABP/VABP. The solid lines represent the fitted functions based on frequentist and Bayesian
logistic regression, while the dashed lines represent the upper and lower 95% pointwise confidence
and credible bounds, respectively. The green histogram represents the distribution of observed
values for free-drug AUC:MIC ratio. (Reproduced from Ref. [66] with permission from Antimicrob
Agents Chemother. Copyright © American Society for Microbiology)