Mak Shina
Mak Shina
Increasing demand for renewable feedstock-based chemicals is driving the interest of both academic and
industrial research to substitute petrochemicals with renewable chemicals from biomass-derived
resources. The search towards novel platform chemicals is challenging and rewarding, but the main
research activities are concentrated on finding efficient pathways to produce familiar drop-in chemicals
and polymer building blocks. A diversity of industrially important monomers like alkenes, conjugated
dienes, unsaturated carboxylic acids and aromatic compounds are thus targeted from renewable
feedstock. In this context, on-purpose production of 1,3-butadiene from biomass-derived feedstock is an
interesting example as its production is under pressure by uncertainty of the conventional fossil feedstock.
Ethanol, obtained via fermentation or (biomass-generated) syngas, can be converted to butadiene,
although there is no large commercial activity today. Though practised on a large scale in the beginning
of the 20th century, there is a growing worldwide renewed interest in the butadiene-from-ethanol route.
An alternative route to produce butadiene from biomass is through direct carbohydrate and gas
fermentation or indirectly via the dehydration of butanediols. This review starts with a brief discussion on
Received 6th March 2014 the different feedstock possibilities to produce butadiene, followed by a comprehensive summary of the
DOI: 10.1039/c4cs00105b current state of knowledge regarding advances and achievements in the field of the chemocatalytic
conversion of ethanol and butanediols to butadiene, including thermodynamics and kinetic aspects of the
www.rsc.org/csr reactions with discussions on the reaction pathways and the type of catalysts developed.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
1
emerging production of BD from petroleum sources. The ethanol Table 1 Butadiene content (kg kg ethylene) for various feedstocks2,3
route only survived up to now in China and India. In the US,
Feedstock Butadiene/ethylene
a large percentage of BD production is obtained by dehydro-
genation of n-butane and n-butenes.2 Direct extraction from C4 Ethane 0.02
Propane 0.07
streams arising from naphtha crackers (ethylene plants) is Butane 0.07–0.11
substantially practised in Europe since the early seventies. Naphtha 0.13
The separation of BD from C4 streams though requires an Gas oil 0.26
expensive extractive distillation with a selectivity of only 4–5%
due to its azeotrope formation with butane. Recent trends in steam cracking show a lightened feedstock
The growth of ethylene production outpaced that of BD as a result of lower costs and larger availability of ethane from
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
demand resulting in an oversupply of BD. Although this situa- natural gas like shale gas, this change in feedstock affecting the
tion has led to a shut-down of many on-purpose BD production supply and price of BD. Table 1 illustrates the lower BD yield
units, it also stimulated the development of chemical processes from lighter feedstock relative to naphtha and light atmospheric
using BD as the building block. BD is thus considered as a high gas oil in the steam cracker.2,3 Whereas an ethane-only cracker
value by-product of ethylene production from steam crackers with- has the lowest capital cost and the highest selectivity towards
out clear production economics. Since many years, the BD supply ethylene, it shows the lowest BD yield. As a consequence, restric-
and price have been influenced by the demand for ethylene. tions on the availability of BD and high BD prices are expected in
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
Pierre A. Jacobs has been since Bert F. Sels obtained his PhD in
2008 professor emeritus with 2000 at KU Leuven in the field
special mandate at KU Leuven. of oxidation chemistry with
During his active career, he has heterogeneous catalysis. He was
been professor of Physical- awarded the DSM Chemistry
Chemistry and Catalysis at KU Award in 2000, and in 2005 the
Leuven and authored over 650 Incentive Award by the Societè
reviewed papers. The industrial Chimique Belge for his research
impact of his research is visible achievements. In 2005 he was
in numerous patents and culmi- appointed assistant professor at
nated in the recent nomination for KU Leuven, and full professor in
the 2013 Alwin Mittasch special 2008. He now teaches courses on
Pierre A. Jacobs prize marking the centennial com- Bert F. Sels spectroscopic tools in organic
missioning of the first ammonia chemistry, sustainability of
plant. Among the important awards received for his research are the chemical processes, spectroscopic tools in surface chemistry and
Paul Emmett Award in 1981, for his contribution to fundamental heterogeneous catalysis. He heads a research group in the Center
understanding of zeolite catalysts, the Donald Breck Award in 1998 for for Surface Chemistry and Catalysis (KU Leuven). His current field
his work on zeolite-based biomimetic catalysis. In recognition of his of research interest focuses on synthesis, characterization and use
contributions to bridge homogeneous, heterogeneous and bio-catalysis, of heterogeneous catalysis, mainly in transformation of renewable
he received the Synetix Award and the Kreitman Award in 2001. resources to chemicals. He was recently elected the co-chair of the
He was Karl Ziegler Guest Professor in 2004. IZA Catalysis Commission.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
low-cost and non-food/feed biomass feedstocks.10 Ethanol from under development. Syngas may originate from reforming of
cellulosic and algal sources, the so-called 2nd and 3rd genera- natural gas such as shale gas, although gasification of biomass
tion ethanol, is promising in this respect.11–17 or waste gases from the steel industry are valid alternatives.
While the industrial feasibility of the ethanol-to-BD route One route to BD runs via the production of 2,3-BDO followed by
has been proven long time ago, a number of improvements its dehydration. It was announced for technical implementation
both from an engineering and catalytic point of view is still vital by 2016 in the frame of new technology for adiponitrile (ADP) and
to attain high BD yield and high BD productivity. In order to nylon-6,6 production by LanzaTech and Invista.26 The butanediol
compete with existing oil- and gas-based technologies like C4 is produced from syngas via fermentation, for instance from the
dehydrogenation, issues related to carbon efficiency, catalyst waste gases of a steal producer or from natural gas steam
cost, toxicity, catalyst performance, feedstock tolerance, back- reforming.27 In the long term the direct gas fermentation of CO
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
end optimization, catalyst lifetime and stability require further to BD is envisaged. In a partnership with Orochem Technologies,
optimization. So, it seems that extensive research is necessary both companies claim an economically viable thermocatalytic
to improve the viability of the ethanol route to BD vis-à-vis the 2,3-BDO dehydration process to BD with high selectivity. While
conventional routes. In 2013, Axens, IFPEN and Michelin have the CO gas fermentation lifts the process out of the sugar value
launched their joint research program to develop an economically chain (food – chemicals), the batch sugar production is replaced
competitive process to produce bio-sourced synthetic rubber from by continuous CO gas production.
bioethanol.18,19 Other routes to BD from natural gas were investigated. They
Next to ethanol, other bio-derived chemicals like butanols mainly focused on the direct and indirect production of chemical
(BuOH), viz. n-butanol and Gevo’s isobutanol, 2,3-butanediol and fermentation ethanol, which is converted to BD. Several
(2,3-BDO) and 1,4-butanediol (1,4-BDO) may also be useful to papers and reviews have been published on the catalytic conver-
produce BD. Most of the C4 alcohols are now available from sion of syngas to ethanol and higher alcohols using chemo-
renewable feedstocks via fermentation of biomass or syngas catalysis.28–31 Three chemocatalytic routes are known in the
(e.g. biomass derived).20,21 For instance, Genomatica – a company literature to convert syngas to ethanol: (i) the direct conversion
designing microorganisms for the conversion of sugars and via selective CO hydrogenation on a solid catalyst, (ii) methanol
cellulosic biomass to sustainable chemicals – is currently pro- homologation, which involves reductive carbonylation of methanol
ducing 1,4-BDO at a demonstration plant. The first commercial- over a redox catalyst, and (iii) a multistep process, wherein classical
scale plant is scheduled, with Novamont as a partner.22,23 methanol production from syngas, is followed by carbonylation to
Together with Versalis, Eni’s chemicals subsidiary leader in acetic acid, and subsequent hydrogenation to ethanol. None of
the production of elastomers, they recently announced the these routes have been commercially practised. The carbonylation
establishment of a strategic partnership to enable the produc- of methanol to ethanol is the most promising route, while the
tion of BD from renewable feedstocks. Main incentive of the direct formation of ethanol from syngas is probably commercially
collaboration is the successful direct production of million less interesting due to a low ethanol yield and selectivity.32 A recent
pound quantities of bio-based 1,4-BDO from biomass resources example of the methanol–acetic acid route (TCX technology) is
in 2011 by Genomatica.24 reported by Celanese.33
Acid catalysed dehydration of (bio-based) butanols leads to a The fermentative production of ethanol from syngas has also
mixture of n-butenes, requiring further dehydrogenation to gained considerable attention. The syngas could be an industrial
produce BD with existing technology.2 Next to dehydration of waste gas or it may be obtained by natural gas reforming or
monohydric alcohols direct butanediol (BDO) dehydration to BD gasification of biomass. The utilization of the whole biomass of
was also performed. This strategy has the advantage of avoiding low quality and the elimination of biomass fractionation are major
the expensive dehydrogenation step. The reaction involves a advantages here. The state-of-the art and the most important
double dehydration, preferably employing a catalytic process, challenges recently have been critically reviewed.34–38
the catalyst and the reaction conditions employed differing with Hydrocarbon and natural gas routes. Synthetic ethanol,
the position of the alcohol groups. Dehydration of 1,4-BDO is produced by the direct or indirect hydration of ethylene from
well-known from the acetylene-based Reppe route.2 It is carried the cracker or dehydrogenation of ethane in natural gas resources,
out at 280 1C by a rather complex method using an excess of may be converted to BD. However, the capacity of worldwide
steam and THF in the presence of a sodium phosphate catalyst. production of synthetic ethanol is rather limited and represents
The dehydration chemistry of 1,3-BDO is slightly different. High less than 10% of the global production.
BD yields above 80% were obtained at 400 1C using the same Dehydrogenation of butane from natural gas is a better
catalyst in the presence of steam. Dehydration of the other diols economical solution to BD production. This well-established
is less selective. So, 2,3-BDO easily dehydrates to methyl ethyl technology has been practiced until the 1990’s in Russia and
ketone, while its diacetate can be pyrolysed in high yield to BD the US. The main disadvantages of the process are the high
upon heating to 475–600 1C, the acetic acid co-product being endothermicity and the high process temperature (600–700 1C)
recycled.25 New catalytic advances in the dehydration of diols to achieve an economical equilibrium conversion. As the reac-
and polyols in detail will be described below. tion runs at incomplete conversion, separation of feed and
Gasification routes to butadiene. New chemical and bio- products is mandatory. Due to coke formation, catalysts usually
chemical technologies producing BD from syngas are currently run for only 5 to 15 minutes. As heat from catalyst regeneration
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
fuels the endothermic dehydrogenation, the overall lifetime of is attempted in the first part. The contribution includes details
the catalysts is much longer. The Houdry-Catadiene process of the reaction network, mechanistic and thermodynamic con-
using a mixture of Al oxide and Cr oxide typically shows a siderations and compares catalytic data (in terms of BD yield,
reactor outlet with 15 vol% butadiene, and about 50% carbon ethanol conversion, carbon yield and volume productivity) of
yield.39 Somewhat higher yields (65%) are reported for the two- various catalyst types. Catalytic properties and structure/activity–
step Phillips Petroleum technology, involving butane dehydro- selectivity relationships were made as much as possible, if appro-
genation to n-butenes with Cr/Na/Al oxide, followed by their priate, in order to guide and inspire the reader.
extractive distillation and conversion to butadiene over Fe/K/Al Routes to BD from 4-carbon diols (BDO) have attracted
oxide.39 Application of vacuum or addition of water are used to much recent attention and this evolution is mainly explained
reduce the partial pressure (Le Chatelier) and to overcome coke by the availability of bio-derived BDOs from renewable and
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
formation by stimulating the water gas shift reaction, respec- waste feedstocks. Though the chemistry seems less complicated
tively (see also technologies from Dow and Shell). The conver- and has already been described in part of the old BD processes
sion and selectivity of the process are also greatly improved by from acetylene and ethanol, the selective double dehydration of
hydrogen removal. In the presence of dioxygen, water is BDOs to BD under milder conditions remains challenging
formed, the exothermicity partly being used to compensate requiring unique catalysis. The second part of the review is
the energy required for the dehydrogenation. Whereas the therefore devoted to summarize and describe the most impor-
presence of water and O2 increases catalyst lifetime, O2 also tant catalytic achievements to dehydrate such diols including
helps the difficult abstraction of allylic hydrogen. With Phillips the mechanistic proposals and thermodynamics, the type of
O–X–D at 480–600 1C, 80% conversion and 90% BD selectivity catalyst developed and the current status.
were obtained. The Oxo-D process of PetroTex shows 93% BD
selectivity at 65% conversion.2 Both processes run only on
n-butenes since the addition of oxygen to butane would result 2. Ethanol-based butadiene
in products oxidation at the temperature required for butane 2.1. The technological options: one vs. two-step processes
dehydrogenation. Although use of iodine to remove H2 as HI In 1903 the catalytic synthesis of BD from ethanol was first
(Idas process, Shell France) has been proposed, serious corro- performed by Ipatiev in Russia.62 Low BD yield of about 1.5%
sion has been encountered. The yield of BD on butane was said was obtained by passing ethanol at 550–600 1C vapour over Al
to be 70%. Butenes could be also resourced from dehydrating powder. Filippov obtained a BD yield of 5% from diethyl ether
biobutanols to produce bio-based BD, though the price of the at 400–500 1C on the same catalyst.63 Ostromyslensky proposed
bioalcohols is currently too high for commercial use. a catalytic route from a mixture of ethanol and acetaldehyde
The dehydrogenation processes are competitive with the over alumina or clay catalysts at 440–460 1C, yielding 18% BD.64
cracking process when BD price is high and raw C4 is available. Maximoff claimed the formation of BD from a mixture of
Honeywell UOP in partnership with TPC group and BASF in ethanol with acetaldol (or crotonaldehyde) over aluminium
collaboration with Linde recently announced the development hydroxide.65 Later, Lebedev66–68 proposed his first process on
and licensing of such a process for the on-purpose BD produc- a mixture of zinc oxide and alumina at 400 1C. He achieved a BD
tion from butane.40–42 selectivity of 18%, directly from ethanol according to following
1.2. Scope of the review reaction (1):
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
2.2. Routes and chemistry from ethanol to BD authors considered that the reactions, including the Lebedev
and Ostromyslensky synthesis, all proceeded through the
Ethanol as a building block can be converted into BD. The intermediate formation of crotonaldehyde.69,72,73,77,78 The
reaction network involves a series of reactions, leading via mechanism includes the formation of acetaldol via the aldol
multiple pathways to a variety of products that are controlled condensation of two acetaldehyde molecules, followed by
differently by kinetics and thermodynamic restrictions. Before dehydration of the aldol to crotonaldehyde, in accordance with
describing the thermodynamic and kinetic aspects, the different reaction (2):
reactions involved and the complex reaction network in the
synthesis of BD from ethanol will be clarified first.
Butadiene formation. Since the successful industrial pro- (2)
duction of BD from ethanol (or a mixture of ethanol and
acetaldehyde), the study of the reaction mechanism has been
performed extensively. Many reaction mechanisms proposed Further transformation of crotonaldehyde to BD has been
before 1945 have been discussed in depth in the review by the subject of many discussions for a long time. According to
Egloff and Hulla.44 This review briefly mentions the most Quattlebaum et al., crotonaldehyde is deoxygenated with ethanol,
important statements and conclusions of that period, but we as presented in Scheme 5(i).73 Moreover, the authors for the first
will focus more on the recently anticipated reaction schemes. time attempted a description of the mechanism taking into
Initially, Lebedev et al. proposed a radical mechanism to account the participation of surface atoms of the catalyst. They
describe the direct formation of BD from ethanol, in accordance assumed a synthesis route via hydrogen transfer from ethanol to
with the reactions in Scheme 3.68 A few years later, Balandin the enol-form of crotonaldehyde.73 It has been suggested that the
argued that this radical reaction mechanism was energetically reactive bridged siloxane of the silica-based catalyst participated in
possible.71 All possible combinations of molecules in the reaction the deoxygenation process by reacting with ethanol and the enol,
mixture were analysed based on the multiplet theory. Whereas the thereby supporting an intermolecular hydride shift (Scheme 5ii–iv),
formation of 49 different possible reaction products was predicted, while the oxide promoter (e.g., tantalum, zirconium or niobium
only 21 of them could be detected. Acetaldehyde was identified as oxides) performs the aldol condensation. The latter reaction
a primary product, while BD was assumed to be a third generation was identified as the rate-controlling step for the Ostromyslensky
product. Baladin concluded that the reaction proceeded via 1,3- process.73
BDO according to the following reaction sequence (Scheme 4).
This scheme was essentially similar to the mechanism
proposed earlier by Ostromyslensky for his two-step process
using a mixture of ethanol and acetaldehyde.64 Egloff and Hulla
supported the intermediate formation of 1,3-BDO.44
The above mechanistic proposals were all rejected a few
years later by different research groups.69,72–76 By studying the
Scheme 3 BD formation mechanism according to Lebedev.68 Scheme 5 BD formation mechanism according to Quattlebaum et al.73
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
Jones et al.76 studied the reaction mechanism of the two-step participation of acid and basic sites. High BD yields necessitate
process over the same catalyst, viz. 2%Ta2O5–SiO2. They sug- a balanced ratio between the acidity and basicity.80
gested an opposite effect from the catalyst’s components, the Kvisle et al. also using the MgO–SiO2 catalyst81 concluded
unmodified silica being more effective for acetaldehyde conden- that the rate-limiting step consists in the condensation of
sation and the 2%Ta2O5–SiO2 material being more active for feed acetaldehyde occurring prior to the hydrogen transfer step,
mixtures requiring deoxygenation of the aldehyde. Thus, the rather than the dehydrogenation of ethanol to acetaldehyde.
proposed mechanism of the process (in Scheme 5i) was supported, In the absence of quantitative details, they also emphasized the
while reaction sequence ii–iv was rejected (Scheme 5), because synergic effect of the catalyst’s components for active and
participation of the Ta promoter was not included. selective catalysis.
At the same time, a slightly different pathway for crotonalde- Separately, the possible formation of 1,3-BDO as an active
hyde transformation was proposed (Scheme 6).69,72 Crotonalde- intermediate was studied in more detail.73,74,82 Based on the
hyde was first reduced to crotyl alcohol (Scheme 6i and ii), transformation of pure butanediol or its mixture with ethanol,
followed by its dehydration to BD (Scheme 6iii). Ethanol (or any it was concluded that the former did not participate in BD
other alcohol present in the system) served as a hydrogen donor. formation from ethanol.73,82 The analysis by Natta and Rigamonti
Kagan et al. over Lebedev type catalysts showed BD production by supported the thermodynamically impossible formation of 1,3-BDO
the direct reduction of crotonaldehyde with hydrogen.72 In view of from ethanol and acetaldehyde.74
the overall scheme of the reaction mechanism, it was concluded that Although the route of BD formation (Scheme 7) is generally
this hydrogenation process is less likely than the hydrogen transfer accepted, alternative schemes were proposed as well. In
from the alcohol to the aldehyde (Scheme 6ii). Natta and Scheme 8 it is suggested that BD formation results from an
Rigamonti74 and later Bhattacharyya and Sanyal,75 based on thermo- interaction between acetaldehyde and ethylene via the Prins
dynamic calculations of the proposed reaction scheme (Scheme 6), reaction.83,84 Thermodynamically, the Prins mechanism is
have confirmed that the formation of BD proceeded via the selective energetically possible but less favourable than the generally
reduction of crotonaldehyde by ethanol to crotyl alcohol, the direct accepted mechanism via aldol condensation.51,74 However,
reduction of crotonaldehyde by hydrogen being thermodynamically no experimental studies confirming the involvement of the
less favourable. In the overall reaction network (Scheme 7), croton- Prins reaction have been performed.
aldehyde formation from acetaldehyde, i.e. the aldol condensation In the literature, all mechanistic and reaction network studies
step, was indicated as the rate-controlling step.72,75 were based on the experimental results of catalytic tests using
The reaction network in Scheme 7 was also confirmed by various feed combinations, temperatures and contact times. There
Vinogradova et al. using C14 labelled acetaldehyde.79 They showed has been only one attempt studying BD formation from ethanol by
IR spectroscopy.86 Ethanol adsorption and its chemical transforma-
tion were studied over an alumina–zinc catalyst in the temperature
range of 20–400 1C. Above 300 1C a decrease of surface alkoxy-
species in the region 1080–1120 cm 1 with increasing reaction
temperature was observed at the expense of several new vibration
bands. The most intense signal was assigned to the characteristic
vibrations of CQO (1750 cm 1) and CQC–O (1035 and 1680 cm 1),
while the origin of other, less intense bands was not specified.
However, a conclusive mechanism was not proposed.
Scheme 7 Overall scheme of ethanol conversion to BD. Scheme 8 Prins reaction scheme for BD formation from ethanol.83–85
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
Identification and origin of by-products. Reaction of ethanol product of deoxygenated acetaldol with subsequent rearrange-
to BD generates several by-products. Their formation has been ment (reaction (6))73 or it may result from dehydration of
studied intensively.73,87,88 Quattlebaum et al. proposed a classi- 1,3-BDO (see below, Section 3.2), or it could result from crotyl
fication of the by-products according to their formation mecha- alcohol isomerization.90 Butanal can be an isomerization pro-
nism,73 one group including the products formed by simple duct of crotyl alcohol (reaction (7)), while butenes and butanol
dehydration or ester-forming disproportionation, another one may be produced from butanal via deoxygenation (reaction (8))
containing the products formed similarly as in the reaction and reduction (reaction (9)), respectively.73 1,3-BDO was con-
cascade towards BD. The flow chart of the most important sidered as a product of acetaldol reduction (reaction (10)).
by-products is presented in Scheme 9. Furthermore, C6-hydrocarbons and C6-oxygenated products were
According to this classification, the first group includes formed by the condensation of C4-aldehydes such as butanal
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
ethylene and diethyl ether, being products of the direct ethanol and crotonaldehyde with acetaldehyde via similar mechanisms
dehydration (reaction (1) and (2), Scheme 9). Ethyl acetal is a (reaction sequence, ii–vi).73
condensation product in the reaction between ethanol and The presence of molecules containing three and five carbon
acetaldehyde (reaction (3)). Ethyl acetate is a product of acet- atoms was rationalized by the cleavage of 1,3-BDO into
aldehyde disproportionation, also known as the Tishchenko propylene and formaldehyde (reaction (11), Scheme 9). The
reaction (reaction (4), Scheme 9). Finally, acetic acid is a occurrence of the last reaction explains the formation of
product of ethyl acetate hydrolysis (reaction (5), Scheme 9).73 C3-hydrocarbons and C3-oxygenated compounds as a result of
The second group of by-products consists of crotonaldehyde, an interaction of formaldehyde with ethanol and acetaldehyde,
a product of acetaldehyde condensation to acetaldol (reactions as well as C5-hydrocarbons and C5-oxygenated products as a
(iii), Scheme 9) and its subsequent dehydration (reaction (iv)); result of an interaction of formaldehyde with C4-oxygenated
crotyl alcohol, a product of the reduction of crotonaldehyde compounds.73 Also the reduction of formaldehyde could be a
(reaction (v)). Methyl ethyl ketone (butan-2-one) could be a source of methanol (reaction (12)), which further reacts with
other alcohols resulting in ether formation and ultimately in
aromatics according to the Methanol-to-Gasoline concept.91
An almost similar scheme of by-product formation was pro-
posed by Gorin et al.87,88 Despite the experimental evidence of
reaction (11) (Scheme 9),82 an alternative mechanism was
proposed to explain the formation of by-products with an odd
number of carbon atoms.88 The steps proposed are the keton-
ization of acetic acid, formed via reaction (5), resulting in
acetone with release of CO2 (reaction (13)). Acetone is further
reduced to isopropyl alcohol (reaction (14)), which is likely the
origin of propylene via dehydration (reaction (15)). Acetone
and isopropanol could participate in the formation of the
C5-hydrocarbon pool92,93 via mechanisms similar to those
proposed for BD formation (Scheme 9, reactions (ii)–(vi)).
Alternatively, it has been suggested that n-butanol is formed
directly via hydrogenation of crotonaldehyde (reaction (16),
Scheme 9)53,59,89 by surface hydrogen atoms resulting from
ethanol dehydrogenation (reaction (ii)), whereas the butenes can
be obtained by dehydration of n-butanol (reaction (17)).59 Further-
more, it has been also proposed that acetone could be produced
from acetaldol (reaction (18)).94–96 Reaction proceeds via formation
of surface carboxylate intermediate CH3CH(OH)CH2COO(s)
followed by dehydrogenation and decarboxylation.
All compounds identified as by-products in both Lebedev’s
and Ostromyslensky’s processes are summarized in Table 2.
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
2 Ethylene Acetaldehyde
Ethane Acetic acid
Propane Acetone
Propanol and isopropanol
7 Toluene
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
BD formation. One reaction, viz. the bimolecular MPV ((v) in also stimulates reaction (iii) + (iv) to continuously produce
Fig. 1 and Schemes 7, 9), forming crotyl alcohol (CrOH), crotonaldehyde.
shows very low DG values over the entire temperature range. It may thus be concluded that the choice of the temperature
However, as long as CrOH is consumed by dehydration in the range for BD formation from ethanol will be defined by kinetic
consecutive reaction to BD (vi), the MPV reaction (v) will be factors, and thus by the ability of the catalyst components to
consuming crotonaldehyde (CrH) to form new CrOH. Further- accelerate the individual reactions.43
more, as the equilibrium of reaction (ii) is in favour of acet- Fig. 2 shows the equilibrium compositions (mole fractions)
aldehyde formation, an excess of acetaldehyde in the system in the direct conversion of ethanol to BD. These data were
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
Fig. 2 Equilibrium compositions calculated for: (a) the overall reaction of ethanol to BD, reaction (i); (b) the reaction of ethanol to BD taking into account
the reaction sequence, reactions (ii) to (vi); (c) ethanol dehydrogenation to acetaldehyde, reaction (ii); (d) aldol condensation of acetaldehyde to
crotonaldehyde, reaction (iii) + (iv); (e) MPV reaction of ethanol with crotonaldehyde to synthesize crotyl alcohol, reaction (v), with a mixture of ethanol
and crotonaldehyde (molar ratio is 1 : 1); (f) dehydration of crotyl alcohol to BD, reaction (vi). The pressure in the calculations is assumed 1 atmosphere.
The simulation is performed using Equilibrium based reactor (REquil) in Aspen Pluss software. The reactions (i) to (vi) are in line with the reactions in Fig. 1
and in Scheme 7. (CrOH = crotyl alcohol; CrH = crotonaldehyde; AcH = acetaldehyde; BD = 1,3-butadiene).
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
Table 3 Catalytic activity of doped alumina materials in the direct butadiene synthesis from ethanol
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
and sulphates, because the anion could be completely removed the most acidic sites by K2O. Addition of hydrogen peroxide
by washing and calcination. In addition, substitution of aqueous (0.8–1.5 wt%) in the feed significantly improved the catalytic
ammonia as a precipitating agent by sodium or potassium results. A BD yield of 48% was found over the same catalyst
hydroxides resulted in adverse effects.107 Optimal contact times under similar conditions showing 90% selectivity for BD, corre-
attaining maximum BD yields were reported for the binary sponding to a productivity of 550 gBD lcat 1 h 1. The promoting
systems. A WHSV of 1.5 h 1 showed the highest BD yields for effect of peroxide was explained, on one hand, by hydroxylation
most catalysts, except for the SiO2–Al2O3 catalyst which showed of the catalyst surface during the reaction with peroxide and,
the highest BD yields at a WHSV of 1 h 1. Note that many authors on the other hand, by cleaning of the surface from coke by
were using much higher contact times.74,101,102 Bhattacharyya radicals resulting from peroxide decomposition.
et al. have clearly demonstrated the remarkable importance of In conclusion, numerous catalysts of the Zn–alumina type
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
an optimal WHSV. In the case of ZnO–Al2O3, BD yield dropped were not retained because of low activity or selectivity. High
from 56 to 34% with a WHSV decreasing from 1.5 to 1 h 1, or to activity was reached after modification with promoters, facili-
12% with an increase of WHSV from 1.5 h 1 to 2 h 1.104 It is tating simultaneously the five elemental reactions for fast BD
worth mentioning that high BD selectivity and yields at a high formation, avoiding formation of the thermodynamically more
EtOH feed rate is beneficial because high space-time yields can be favourable side-products such as ethylene and butenes. High
reached (Table 3, compare entries 11–13 with entries 2–9). With temperature and low contact times were beneficial for the
regard to the temperature, the highest BD yield was observed at BD yield. Synthesis procedures and choice of precursors were
425 1C for most binary systems. Over the ZnO–Al2O3 catalyst a essential to reproduce the best performing catalysts.
decrease of the temperature from 425 to 375 1C showed a BD yield
decrease from 56 to 14%.104 Furthermore, it has been shown Magnesia–silica catalysts. Magnesia on silica has long been
that use of a fluidized bed reactor allowed us to significantly studied for converting ethanol in BD. Many catalytic studies
increase BD yield up to 73% compared to 56% obtained with a have been reported to reproduce the catalytic experiments and
fixed bed reactor.108 to better understand the active sites on the magnesia–silica
Tret’yakov et al. reported a similar zinc–alumina catalyst catalyst type. Table 4 compares the most important catalytic
promoted with a small percentage of potassium oxide, silica and data for un-promoted systems.
magnesia, viz. 30%ZnO–1%K2O–4%SiO2–4%MgO–Al2O3.105 They Szukiewicz109–111 has reported between 350 and 450 1C the
reported 34% BD yield with 80% selectivity at 420 1C and low production of BD from ethanol over a MgO–SiO2 catalyst with a
contact time, e.g. LHSV of 2.5 h 1 (Table 3, entry 14). Catalysts BD selectivity of 30–40%, using a LHSV of 0.3 h 1 (Table 4,
were synthesised by mixing of zinc, aluminium and potassium entry 1). Unfortunately, ethanol conversion data were absent.
oxides with suspension of magnesia and silica. The catalyst The catalyst had a MgO : SiO2 weight ratio of 60 : 40 and was
composition contained all the necessary active elements to prepared by wet-kneading of a physical mixture of the dry oxide
catalyse the sequence of transformations such as dehydrogena- powders in water, the amount of water added only specified as
tion (reaction (ii), Scheme 7) with Zn, the base-catalysed con- 1.5–4 parts of water to one part of powder mixture by weight.
densation (iii and iv) with MgO, the MPV reaction (v) with Zn or The aging time (2 to 12 h) was not essential. At 410 1C and with
MgO, and the dehydration (vi) with SiO2. The dehydration a LHSV of 0.4 h 1, Corson et al.,101 with a similar MgO–SiO2
of ethanol to ethylene is avoided possibly due to poisoning of catalyst, having a molar Mg/Si ratio of 2.2, were able to
Table 4 Catalytic activity of MgO–SiO2 materials in the direct butadiene synthesis from ethanol
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
wet-mixing (Table 4, entry 13).117 An optimal calcination tempera- time keeping the remaining MgO highly dispersed. However, for
ture of 500 1C was suggested, which is in good agreement with the design of improved catalysts, more systematic characterization
data reported earlier. The best performance was obtained with a of the surface properties of the active catalysts is a prerequisite.
MgO content of 80 wt% (or a molar Mg to Si ratio of 6).
In addition to the catalyst composition and synthesis pro- Magnesia–silica materials doped with a dehydrogenation promoter.
cedure, reaction conditions also affect catalytic activity and BD The commercial magnesia–silica catalysts industrially used in USSR
selectivity. With the commercial catalyst (Table 4, entry 8) contained various dopants affording high and stable BD yields.
increasing BD yields are obtained at reaction temperatures up to These dopants were initially thought to improve the dehydrogena-
460 1C in agreement with thermodynamics, while the BD selectivity tion capacity of the commercial catalyst, resulting in faster
reaches a maximum of 48% at 430–440 1C and then decreases.114 conversion of ethanol to acetaldehyde. According to the patents
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
Kvisle et al. have investigated the effect of the ethanol of Szukiewicz,109–111 modification with chromium oxide up to
concentration on its conversion to BD at 350 1C with their 10 wt% led to a significant increase in BD selectivity of about 5%.
magnesia–silica catalyst (entry 11) at a constant carrier gas flow The Cr modification reduced the formation of some by-products
rate of 18 ml min 1. The authors observed higher BD yields at like ethylene, while others, like butenes, were increasing. The
lower ethanol partial pressures (3000 ppm), in agreement with effect of different dopants on the catalytic performance has been
the increase in the number of product molecules in the studied in detail.
chemical reaction of ethanol to BD (eqn (1)).81 Lowering the Natta and Rigamonti74,112 have employed the addition of
partial pressure obviously leads to lower volume productivity, chromic acid as a promoter of a MgO–SiO2 catalyst with an
while an increase of the carrier gas flow rate (or decrease of the atomic Mg : Si ratio of 2.2. At 415 1C and with a LHSV of 0.2 h 1
contact time) at constant ethanol concentration in the feed (entry 1, Table 5) ethanol was converted to a high yield and
leads to a lower butadiene yield.80 selectivity of BD of 43 and 52%, respectively. With the Cr-free
Summarizing the results mentioned, it follows that high BD catalyst only the respective values of 35 and 41% were obtained.
yields (up to 30–40%) and selectivity (40 to 60%) are possible The positive effect of the Cr-promotion was connected to the
over the un-promoted binary magnesia–silica materials of formation of magnesium chromate preventing excessive forma-
Table 4 at temperatures ranging from 350 to 400 1C, using tion of magnesium silicate. It appears that modification with
space velocities from 0.3 to 0.5 h 1. For low partial pressures of chromium oxide has an effect similar to that of ammonia or
ethanol, a value of 10 vol% is a compromise between BD yield acetic acid (see before).
(selectivity) and volume productivity. Higher partial pressures Corson et al.101 showed that among the more than 500 tested
require higher reaction temperatures resulting in the formation of materials, the 2%Cr2O3–59%MgO–39%SiO2 catalyst was superior.
more side-products. Although contradictory data are reported on In a fixed-bed reactor at 400 1C with a WHSV of 0.4 h 1, it exhibited
the optimal Mg to Si molar ratio and the preferred synthesis a BD yield of 38% with 56% selectivity (Table 5, entry 2),
recipe, it appears that the catalytic performance of the MgO–SiO2 corresponding to a volume productivity of 65 gBD lcat 1 h 1.
catalysts is very sensitive to the acid–base properties of the Fig. 7 shows an overview of the promotion with 2 wt% Cr
synthesized material. There is agreement that the synthesis pro- of MgO–SiO2 catalysts with varying atomic ratios of Mg : Si,
cedure should promote formation of Mg–O–Si species, at the same as prepared and tested by different authors.101,102,112,116,118
Table 5 Catalytic activity of modified MgO–SiO2 materials and related clay minerals in the direct butadiene synthesis from ethanol
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
selectivity to BD. Co-impregnation often forms additional silica by other layered silicates such as talc and kaolinite showed
silicate phases or causes blockage or creation of acid sites. similarly reduced BD yields.129 Generally, both MgO and SiO2
There is no consensus on the optimal atomic ratio of Mg : Si in components being essential in the catalyst composition should
the doped catalysts, probably resulting from differences in the not be replaced for more than 20 wt% by the layered silica to
preparation methods and in the nature of the dopant. It is attain BD yields of 35 to 40%.129
therefore recommended to compare the different catalysts The situation is somewhat different when MgO3SiO23H2O,
based on their surface properties like acid–base relationship an amorphous magnesium hydrosilicate, is used as a catalyst
(in terms of amounts, strength and location) rather than on the precursor.129 At 400 1C and with a LHSV of 0.3 h 1 it gives up to
basis of the elemental composition only. Today, systematic 80% yield of ethylene from ethanol. After precipitation of
studies with advanced surface characterization tools in combi- magnesium hydroxide with aqueous ammonia or sodium
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
nation with catalytic measurements are missing for the ethanol- hydroxide (MgO 4 60 wt%) and subsequent calcination at 550 1C,
to-BD reaction. BD yields of over 50% were noticed. Based on X-ray diffraction, a
partially crystalline phase, viz. composition, 3MgO4SiO22H2O,
Magnesia–silica catalysts promoted by alkaline metal cations could be identified next to MgO. Whereas the presence of free
and clay minerals. As ethylene is thermodynamically favoured in MgO has been suggested to be essential to ensure fast dehydro-
the reaction of ethanol to BD, promotion of magnesia-silica genation of ethanol to acetaldehyde, the precise catalytic role of
with alkaline metal oxides has been studied. Accordingly, the Mg silicate phase has not been clarified.
Butterbaugh and Spence125,126 have reported active MgO–SiO2 Aleixandre et al.123 have reported the effect of clay-modified
catalysts, yielding BD from ethanol with a selectivity up to 48% MgO on the BD selectivity. At 440 1C and using a LHSV of
at 435 1C, by contacting the precursor mixture of magnesium 0.4 h 1, four different clay minerals, viz. bentonite A and B,
hydroxide and silica with an alkaline solution at 90–100 1C for halloysite, and kaolin, were physically mixed with a variable
1–2 h. The alkaline treatment improved the BD selectivity at the amount of MgO (Fig. 11). Bentonite B was obtained from
expense of the ethylene selectivity. The most promising results bentonite A by extracting free silica. The MgO–clay catalyst
were obtained with alkali-digested MgO–SiO2 materials con- containing 57 wt% of bentonite A exhibited the highest BD
taining 20 to 50 wt% magnesia, corresponding to an atomic selectivity (31%). Unfortunately, no ethanol conversions were
Mg : Si ratio of 0.4 to 1.5. reported (entry 15, Table 5). Catalytic differences were rationa-
Later, Ohnishi et al. used post-impregnation with sodium or lised by taking into account the number of accessible active
potassium hydroxide followed by calcination at 500 1C, result- acid–base sites on the clay mineral. Compared to halloysite and
ing in the formation of highly active MgO–SiO2 catalysts. kaolin, montmorillonite – the main component of bentonites – is
At 350 1C and with a WHSV of 0.2 h 1, Na2O/MgO–SiO2 and known for its large interlamellar voids and high cation-exchange
K2O/MgO–SiO2 yielded 87 and 70% BD, respectively.115 Though capacity. However, the presence of different impurities in
the post-synthetic selective poisoning of acid sites by addition of bentonites like iron oxides which likely could promote ethanol
alkali is similar to that proposed by Butterbaugh and Spence,125,126 dehydrogenation was not taken into consideration.
the catalytic performance reported by Ohnishi et al. was almost In conclusion, the catalytic performance of magnesia–silica
twice as high. The different time-on-stream behaviour of the catalysts can be promoted with alkaline metal cations, by
catalyst possibly is at the origin of the different catalytic data. In suppressing (too) high acidity, thus reducing/blocking the
our experience, early sampling under non-steady state conditions ethanol to ethylene dehydration. Whereas modification of the
often leads to higher conversions. Within one to two hours BD magnesia–silica catalyst with layered silicates or clay minerals
yield and selectivity converge to lower steady state values, followed also results in an increase of BD selectivity, the catalytic role of
by slow deactivation at longer time-on-stream.
László et al.102 have studied ethanol to BD conversion with
magnesia–silica catalysts mixed with clays or clay-related
minerals. Replacement in a 2 wt% chromium oxide promoted
catalyst of half of the silica by kaolin, at 435 1C and with a LHSV
of 0.3 h 1 showed a BD yield and selectivity increase from 40 to
48% and from 54 to 63%, respectively, while the ethylene
selectivity decreased (compare entries 3 and 6, Table 5). Kovařı́k
et al.118 using a similar catalyst reported a similar increase
of the BD yield with partial substitution of silica for kaolin.
At 400 1C and with a LHSV of 0.3 h 1, 37 and 47% BD yields
were obtained with 2%Cr2O3–MgO–SiO2 and 2%Cr2O3–MgO–
SiO2–kaolin, respectively (compare entries 4 and 8, Table 5).
A complete substitution of silica for kaolin resulted in a reduced
BD yield and enhanced ethylene yield. Under the same condi-
tions, a MgO–kaolin catalyst yielded 22.5% BD, while the MgO– Fig. 11 Influence of the nature of various clay minerals on the selectivity
SiO2 catalyst produced 35% BD yield.74 Complete substitution of of the MgO/silicate catalysts (T = 440 1C, LHSV = 0.4 h 1).123
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
the silicates is less clearly defined. Despite these improve- oxides132,134,136 or with mixtures of vanadium oxide with tung-
ments, addition of dehydrogenation modifiers is more effective sten or molybdenum oxides.137
to reach high BD yields. Gruver et al.83 have reported that a 6 wt% Ag promoted
sepiolite is able to produce BD from ethanol. At 280 1C and a
Magnesium silicate clay minerals. A series of unmodified Mg WHSV of 0.8 h 1, they obtained about 9% BD yield at 63%
and Si containing clay minerals have also been used in the ethanol conversion. As the major product was acetaldehyde,
ethanol to BD reaction. Dandy et al. demonstrated BD formation obtained at about 20% yield, the catalyst probably lacks appro-
over sepiolite for temperatures ranging from 200 to 300 1C.130 priate basicity needed to compensate for the fast dehydrogena-
Higher BD yields were generally obtained after modification of tion caused by the presence of metallic silver (entry 17, Table 5).
the clay with dopants, affecting the acid–base and dehydrogena- In summary, clay minerals like magnesia–silica catalysts
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
tion properties of the catalyst. Suzuki et al.124 have employed require careful promotion of the dehydrogenation activity and
synthetic Li modified fluorohectorite as a catalyst in the ethanol fine-tuning of the acid–base properties in order to achieve high
to BD reaction at 375 1C using a LHSV of 0.3 h 1. The catalyst BD yields from ethanol. While the aforementioned modified
exhibited an excellent BD selectivity of 48% at 33% ethanol magnesia–silica catalysts generally show better BD yields and
conversion (Table 5, entry 16). An IR study with adsorbed pyridine selectivity than the promoted clays, individuals like ZnO pro-
and CO2 showed that the presence of Li+ and F in the catalyst moted sepiolite also exhibited promising catalytic properties.
modified the Lewis and Brønsted acid sites, thus optimising the
dehydration properties of the catalyst. Oxygen atoms bound to Other catalysts. Catalysts free of Mg2+ and Al3+ have been
Mg2+ at the edge of the layers were responsible for the dehydro- reported as well. Spence and Butterbaugh138 have reported the
genation activity. use of 20 wt% zinc oxide supported on silica (like diatomaceous
To further improve the dehydrogenation ability of the clay earth) for butadiene formation from ethanol, showing a selec-
catalysts, various metal oxide dopants have been explored. Cd, tivity of 58% and a yield of 29% at 450 1C using a LHSV of 1 h 1
Zn and Cu oxide promoters have been tested to improve the (Table 6, entry 1). The reaction temperature for BD formation
BD yields with smectite and bentonite clays.131 Smectite clay was lowered by modifying the Zn catalyst with phosphate or
modified with Cd showed better results than other catalysts tungstate. ZnO on silica and ZnO/Zn(II)phosphate on silica showed
yielding BD with 17% selectivity. Kitayama et al.132–137 used BD yields of 29 and 22% and a BD selectivity of 58 and 63%, at
sepiolites modified by impregnation with various transition 450 and 415 1C, respectively. Corson et al.101 have found a BD
metal oxides as ethanol conversion catalysts in a recirculation yield of 20% at 425 1C using 0.6 h 1 LHSV with a 10 wt% ZnO
reactor at 320 1C. For the unmodified sepiolite ethylene selec- on a silica catalyst (Table 6, entry 3).
tivity was very high, while after doping BD was one of the major Thorium and zirconium oxides on silica exhibit catalytic activity
products, the nature and amount of the precursor salt of the for BD formation from ethanol as well, at 445 1C amounting to BD
promoter determining the BD yield. With manganese content selectivities of 35 and 44%, respectively (Table 6, entries 4 and 5).
increasing from 39 to 79 mol% in the MnO2–sepiolite, the BD Unfortunately, no details about ethanol conversion were
yield increased spectacularly from 8.4 to 33.4%. The catalyst reported.127,139 Corson et al. reported 23% of BD yield using a
prepared from manganese acetate showed better activity than ZrO2–SiO2 catalyst (Table 6, entry 6).101
that from manganese chloride.135 Superior results were obtained Spence et al.127,139 have claimed copper and molybdenum
with ZnO modified sepiolite containing 4.4 wt% Zn, yielding oxides as potential dopants for the ZrO2–SiO2 catalyst. Unfortu-
63.5% of BD at full conversion after 2 h of reaction at 260 1C.133 nately, most of the catalytic data refer to catalytic compositions
High BD yields ranging from 33 to 58% were obtained over for which detailed information on the synthesis procedure and
sepiolites modified with copper, nickel, cobalt and vanadium the physicochemical properties of the catalyst surface is missing.
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
Conversion of ethanol to BD has recently been reported by catalyst (Table 6, entry 12). Although lead and cadmium oxide
Jones et al.140 over silica doubly promoted with metal oxides of silica promoters have been proposed, no systematic approach to
zinc, zirconium, copper, cobalt, manganese, cerium and hafnium. variations in catalyst composition was present nor was detailed
Emphasis was on the effect of the nature of the promoters and the information concerning the preparation methods available.
texture of the silica on the catalytic activity and BD selectivity. The Finally, although it has been postulated that modification of
best results were obtained over a ZnO–ZrO2/SiO2 catalyst contain- silica gel with 0.5–10 wt% titanium oxide should give suitable
ing 1.5 and 0.5 wt% ZrO2 and ZnO, respectively. Higher ethylene catalysts for ethanol transformation into BD, no catalytic data
selectivity occurred in parallel with higher promoter contents. were presented to support this.142
At 375 1C and a space velocity of 1.5 h 1, the catalyst showed Next to the reported silica supported oxide materials, silica-
18 mol% BD yield with a BD selectivity of 39 mol%, corre- free catalysts have been found to be active in the direct
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
sponding to a high volume productivity of 124 gBD lcat 1 h 1 conversion of ethanol to BD. Bhattacharyya et al. studied binary
(entry 7, Table 6). Using this catalyst as reference, it could be systems containing zirconium or thorium oxides.104 Among the
shown that impregnation of the promoters rather than sol–gel tested materials, ZrO2–Fe2O3, having a weight ratio of 40 to 60,
synthesis is the preferred synthesis method to achieve high BD showed a high BD yield of 38% with 45% selectivity at 450 1C and
yields. High ethylene yields were observed as well. a space velocity of 1.5 h 1 (Table 6, entry 13). Arata et al.143 have
Selective poisoning of acid sites with alkali metal cations, as shown that promotion of zirconia with weakly basic TiO2 cata-
proposed by Ohnishi115 for the MgO–SiO2 system, was not very lysed the ethanol to BD reaction, yielding 12.4% BD at 330 1C.
successful for the Zn/Zr oxide silica system. In contrast, adding a Tsuchida et al.53 have employed hydroxyapatite to produce
third element like Cu oxide to the Zn–Zr doped silica significantly BD from ethanol at 350 1C and 1 h 1. The acid–base properties
improved the BD selectivity, yielding 19 mol% BD with a BD of the catalyst surface were controlled by changing the Ca to P
selectivity of 50 mol% (compare entries 7 and 9 in Table 6). Finally, ratio in the hydroxyapatite composition. Catalysts with a molar
an increase in BD yield and selectivity with pore size of the silica Ca : P ratio larger than 1.62 possessed the appropriate acid–base
support increasing from 40 to 150 Å was noticed for the Zn–Zr balance, yielding 24% BD at an ethanol conversion of 20%.
modified silica. Doped large pore silica yielded 23% BD with 48% 2.4.2 Two-step process. The two-step process of ethanol to
BD selectivity under the same reaction conditions (compare entries BD involves the dehydrogenation of ethanol to acetaldehyde in
7 and 8, Table 6). According to authors140 the large pore silica a separate step, followed by BD production in the second stage
catalyst contained lower proportion of acidic silanols, as measured by co-feeding ethanol and acetaldehyde. The latter reaction was
by the fractions of Q4 vs. Q2 and Q3 in 29Si MAS NMR that results in a discovery of Ostromyslensky at the beginning of the 20th
the decrease of ethylene yield. However diffusional effects and century.64 In this way, dehydrogenation of ethanol is decoupled
dispersion of supported metal oxides particles were not taken into from aldol condensation, MPV reduction and dehydration.
account to explain the changes in selectivity. Catalytic tests at Commercialized by Union Carbide, the process has been
different space velocities showed that a decrease of the space supplying the major amount of synthetic BD in the US during
velocity from 1.5 to 0.75 h 1 results in a slight increase in BD yield, 1943 to 1944.1 Although many studies on the two-step process
albeit with lower BD productivity. No explanation of the contact were not showing much pertinent information, most of the
time effect was advanced. It should be stressed that the preferred available bibliographic data refer to reactions of co-fed ethanol and
LHSV of 1.5 h 1 used is similar to that used by Bhattacharyya acetaldehyde in the presence of catalysts, previously developed for
et al.104,107,108 for the superior modified alumina catalysts. This the one-step process. It was attempted mostly to elucidate the
contact time is several times higher than that usually applied for reaction mechanism or to evaluate the effect of the presence of the
the magnesia–silica and other related catalytic systems (compare different components on the catalyst composition. In this respect,
Table 6, entries 7–9 with the entries in Tables 4 and 5). various authors noticed that addition of acetaldehyde results in an
Recently, Ordomskiy et al.141 have claimed that silver, cerium, tin increase of BD yield. However, it remains undecided whether the
and antimony oxide promoted ZrO2–SiO2 are highly selective for BD. two-step process should be preferred over the one-step process.
At 325 1C and with a WHSV of 0.3 h 1, 72% BD selectivity at an Corson et al. have reported that BD formation from ethanol
ethanol conversion of 34% was obtained over a 0.4 wt% Ag modified via the two-step process is preferable.101 This preference is
catalyst, corresponding to a productivity of 40 g kgcat 1 h 1 based on the enhanced flexibility for selection of appropriate
(entry 10, Table 6). A remarkably low temperature was used conditions for each reaction separately and on the decreased
compared to the usual temperature window for ethanol con- catalyst complexity with respect to composition and properties
version to BD. of the two catalyst types used in each stage. The two-step
Furthermore, a high BD yield has been observed over a process allowed a reduction of the reaction temperature from
tantalum oxide doped silica – the commercial catalyst developed 400–425 1C to 350 1C. The purity of BD extracted from the
for the two-step process.101 Good yields of BD up to 16% have 4-carbon cut amounting to 98% BD in the two-step process was
been observed over the catalysts with 2 wt% Ta2O5 on silica at only 80% in the one-step process.
425 1C and 0.6 h 1 (Table 6, entry 11).101 Under similar reaction Conversely, Bhattacharyya et al. have claimed that the one-
conditions, an increase in BD yield up to 25% with 30% BD step process is preferable.108 Under appropriate conditions,
selectivity was found upon doping the Ta oxide catalyst with Cu the amount of formed BD is high enough so that it is easily
oxide, likely due to an improved dehydrogenation capacity of the separated from other gases.108 Jones et al. showed that the
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
addition of acetaldehyde next to ethanol does not result in any commercial silica gel, while the BD selectivity increased from
improvement of BD yield for catalysts that are able to produce 72–75 to 80%. Moreover, Ta2O5 supported on ordered silica
significant amounts of acetaldehyde (AcH) from ethanol.140 exhibited a better coke tolerance and stability with time-on-stream.
Ostromyslensky has reported that a higher yield of BD was The best catalytic results in terms of activity and selectivity were
obtained over aluminium oxide at 440–460 1C using an equi- obtained using silica support material with nano-sized morpho-
molar mixture of EtOH and AcH.64,144 Later, Maximoff and logy and large pore sizes, indicating the importance of efficient
Canonici145,146 have shown that aluminium sulphate, and basic pore diffusion. 2D or 3D organization of the mesopores like in
aluminium sulphate alone or deposited on pumice-stone or SBA-15 or KIT-6 was found to be less important. The advantage of
Kieselguhr, were capable of catalysing the formation of BD hierarchical mesoporous silicas is related to the ability of reaching
from EtOH–AcH mixtures at a reaction temperature ranging higher dispersion of tantalum making more active sites available
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
from 280 to 450 1C yielding a BD selectivity of 33% with a molar and preventing agglomeration of Ta oxide during synthesis and
EtOH to AcH ratio of 2. catalytic reaction.
The two-step process of BD formation has been studied Romanovsky and Jordan148–150 have reported that the addition
intensively at Union Carbide and Carbon Chemical Corp. as of phosphate modifiers to magnesia–rich MgO–silica catalysts
early as 1947 with tantalum, zirconium or niobium oxide with an atomic Mg to Si ratio of 5 resulted in a higher conversion
promoted silica.70,78 The first step proceeded over a standard of the ethanol–acetaldehyde mixture and a higher BD selectivity.
copper dehydrogenation catalyst. The best results of the second Key promoters for silica–magnesia were 5 to 25 wt% calcium or
step were obtained over a catalyst containing 2 wt% tantalum calcium–nickel phosphate complexes having molar Ca to Ni ratio
oxide on silica. At 325–350 1C and a space velocity of 0.4–0.6 h 1, of 7.5–9.2. The best results were obtained at 400 1C using a
the catalyst produced 35% BD yield with a selectivity of 67% using LHSV of 0.4–1.0 h 1, yielding 21 and 23% BD with 56 and 58%
a mixture of 69% by weight ethanol, 24% acetaldehyde and 7% selectivity, respectively.
water. Higher Ta2O5 contents in the catalyst resulted in a further In addition to the catalyst composition, the reaction operat-
increase of the BD yield but at the expense of BD selectivity. It has ing conditions like temperature, contact time and feed molar
been stated that the activity of catalysts containing less than ratio of ethanol to AcH also played an important role in the
1 wt% of tantalum oxide was not reproducible.101 catalytic activity and BD selectivity. The old commercial process
The catalytic activity of zirconium and niobium oxides on operating between 325 and 350 1C used a mixture of 7 wt%
silica was lower than the Ta2O5–SiO2 material. 1.6 wt% ZrO2 on water and 93 wt% feed with a molar ethanol to AcH ratio of 2.7.
SiO2 showed a BD selectivity of 59% under comparable reaction Toussaint et al.78 investigated the influence of the reaction
conditions.78 Corson et al. have reported that hafnia promoted temperature, the molar ratio of EtOH to AcH and the space
silica gel was selective and active in the two-step production velocity on the catalyst activity and BD selectivity of a commer-
of BD. At 300 1C a BD yield of 30–40% and a BD selectivity of cial 2 wt% Ta2O5 on silica. High BD selectivity was obtained at
50–60% could made from an ethanol–acetaldehyde mixture.101 325 1C for a feed mixture with a EtOH to AcH ratio of 3, whereas
Ordomskiy et al. have claimed that zirconium, titanium, better results were found at 350 1C with a feed mixture ratio of
tantalum, niobium or magnesium oxides supported on silica EtOH to AcH of 2–2.5. Catalyst deactivation was less pronounced
promoted by silver, gold, copper, cerium, tin or antimony oxides at lower temperature. Though an increase in BD yield was
are highly active and selective catalysts for the production of BD observed at 350 1C with a EtOH to AcH feed molar ratio of 2, a
from a mixed acetaldehyde–ethanol feed at 325 1C.141 Among the considerable loss in activity was detected after 160 h on stream,
Ag promoted MOx–SiO2 catalysts, mildly basic titanium oxide when compared to a reaction at 325 1C using a feed molar ratio
was found to be the most selective catalyst showing a BD yield of of 3, the latter conditions being considered as optimal. Corson
25% and a BD selectivity of 72%. Ag promoted MgO-containing et al. have reported that for a 2%Ta2O5–SiO2 the optimal condi-
catalyst produced more BD, yielding 29%, but with a BD selec- tions for BD production were achieved at 350 1C, using a LHSV of
tivity of only of 64%. Among the zirconium containing materials, 0.6 h 1 and a molar EtOH to AcH molar feed ratio of 2.75.151 The
M–ZrO2–SiO2 doped with different metals, gold was claimed to amount of coke deposited on the catalyst was significantly higher
be the best choice producing BD with a yield of 25% and a when the EtOH to AcH molar ratio was decreased below 2.75,
selectivity of 82%. Promotion of a Ag modified ZrO2–SiO2 with resulting in faster catalyst deactivation.
cerium oxide further increased the BD yield when compared to The old commercial process with 2 wt% Ta2O5 on silica used
tin, antimony and sodium dopants, yielding 33% BD with 81% 7 wt% water in the feed. Romanovsky and Jordan employed 40
selectivity. Silica was the best support, and use of promoted to 60 wt% water with the phosphate supported magnesia–silica
alumina and aluminosilicate resulted in lower BD yields, 18 and catalyst to reach the highest BD yield of 23% at 410 1C using a
25%, respectively, and selectivity was 57 and 74%, respectively. LHSV of 0.4 h 1. However, the precise role of water has not
Chae et al.147 have studied the effect of the pore size, pore been investigated.148–150
volume and pore organisation on the catalytic conversion of an In conclusion, 2 wt% Ta2O5 on silica seems to be the pre-
ethanol–AcH mixture at 350 1C and a LHSV of 1 h 1 by using ferred catalyst for converting a mixture of ethanol and acet-
2 wt% Ta2O5 on several mesoporous silica like SBA-15, KIT-6 aldehyde to BD contains. High ethanol to acetaldehyde molar
and MCM-41 type materials. The BD yield increased from 22–24 ratios are required to obtain high BD selectivity and to avoid
till 37% when a SBA-15 silica support was used instead of a coke formation. Likely, water in the feed mixture has a
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
beneficial though not further determined role in the catalysis. different BDOs is addressed first. 1,4-BDO is a versatile inter-
High catalyst stability and high BD yield are obtained for reaction mediate in the chemical industry with an annual production in
temperatures between 325 and 350 1C, the preferred contact times the range of 1 million metric tons.2 The main synthesis route
corresponding to a LHSV ranging from 0.3 to 1 h 1. Use of large is based on the Reppe process starting with the reaction of
pore size mesoporous silica is beneficial for the BD yield and acetylene with formaldehyde to form 2-butyne-1,4-diol, followed
selectivity due to high Ta oxide dispersion and efficient pore by hydrogenation to 1,4-BDO. The latter is not only a feedstock
diffusion. Best BD yield and selectivity were reported to be about for tetrahydrofuran (THF), but it is also used for polyurethanes
25 and 80% respectively, corresponding to a volume productivity and poly(butylene) terephthalate. Very recently, Genomatica
of 350 gBD l 1 h 1. unveiled an inventive bio-based route towards 1,4-BDO, using
a metabolically engineered Escherichia coli, which is capable of
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
Most reports on the conversion of 2,3-BDO in the gas or liquid carbon backbone.189 Other zeolite topologies have also been
phase show a predominant pinacol type dehydration. The pinacol investigated. Lee et al. have investigated a series of zeolites with
reaction is accompanied by BD formation in the liquid phase only different Si/Al ratios within MFI, MOR, BEA and FAU topologies.
under non-classical conditions. For instance, Waldmann et al. They noted that the small pore zeolites favoured butan-2-one
have studied the dehydration of 2,3-BDO to BD in excess of pthalic formation,190 whereas the large pore zeolites mostly yielded
anhydride with benzenesulfonic acid.178,180 The pyrolysis of 2,3- acetals and ketals. However, no significant formation of BD was
acetoxy derivatives of 2,3-BDO into BD is also known.181 The use of reported. Very recently, Zhang et al. have studied the influence
amines also helps in guiding the dehydration to BD, while avoid- of boric acid modification of H-ZSM-5 zeolites191 in a study
ing butan-2-one formation from 2,3-BDO.182 Hereto, a vapour concerned with the effect of acid site density and strength. In
phase reaction over a catalyst, containing silica and tungsten oxide all cases however, BD yields were in the range of 1%.190–192
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
on alumina, succeeded yielding 48% of BD from 2,3-BDO in a Olah et al. in a study on the use of H-Nafion, a perfluorinated
single pass at 330 1C. However, this high yield was only possible by resin with sulfonic acid groups, reported that this superacid
co-feeding butan-2-one, water and triethylamine. Later, a patent mainly catalysed the pinacol route.179 In a reactive distillation
was filed on a similar catalyst composition regarding the forma- setup, 1,3-dioxolane derivatives were predominantly formed via
tion of 2-methyl-butadiene (isoprene) in high yields (490%) from acetalisation of butan-2-one with 2,3-BDO, while butan-2-one
2-methyl-2,3-butanediol in the vapour phase without co-feeding was the main product in a continuous flow reactor setup. At
additional reagents. Unfortunately, the patent did not demonstrate 150 1C, the conversion of 2,3-BDO was about 78%, while the
the conversion of 2,3-BDO to BD.178,183 selectivity to BD only reached 4%. Butan-2-one and its acetals
The first real attempts on the single-step 2,3-BDO-to-BD accounted thus for about 94% of the recovered products. At
conversion without sacrificial or stoichiometric reagents date 175 1C, the yield of BD increased to 8% at full BDO conversion.
back to the 1940s. Bourns et al. have assessed a Morden Other strong acids like heteropolyacids examined by Török
bentonite clay as a catalyst and they noticed full conversion of et al. showed a dominant pinacol pathway.193 The preference
2,3-BDO into butan-2-one below 350 1C. Between 450 and 700 1C, for pinacol formation with acid catalysts is attributed to the
more gaseous products were formed containing small quantities strength of the acid sites. Strong acidity better stabilises the
of BD. Dilution of the feed stream with steam increased the BD oxonium ion transition state (thus the carbocation formed from
yield to 25%.184 Winfield was the first who truly aimed at making 2,3-BDO), hence favouring the 1,2-hydride (and also some 1,2-
BD form 2,3-BDO in as single step and thoroughly screened a methyl) shift. Ironically, when 2,3-dimethyl-2,3-butanediol
range of catalysts including several metal oxides and salts such (actual pinacol) is the substrate instead of 2,3-BDO, pinacol
as SiO2, B2O3, ThO2, BeO, CaHPO4, CaCO3 and CaSO4 in a rearrangement is less dominant and diene formation is more
vapour phase reactor setup in 1945.185 The author aimed at a pronounced. Ironically, when 2,3-dimethyl-2,3-butanediol (actual
reaction step, proceeding via 1,2-elimination towards 3-buten-2- pinacol) is the substrate instead of 2,3-BDO, pinacol rearrange-
ol, and further dehydration to BD (Scheme 11). While most ment is less dominant and diene formation is more pronounced.
catalysts promoted pinacol dehydration with almost exclusive This change in the reaction pathway is likely related to the relative
formation of butan-2-one, ThO2, prepared according to a specific ease of pinacol rearrangement with 2,3-BDO, since the migration
synthesis protocol, produced selectively BD. The best result was of the hydride as in 2,3-BDO is easier than the shift of a methyl
obtained over the ThO2 catalyst at 350 1C in a fixed-bed reactor group as in 2,3-dimethyl-2,3-butanediol.188,189
under reduced pressure, yielding 60% BD and 20% 3-buten-2-ol. Sato et al. have reported the use of basic CeO2 to dehydrate
According to the authors, the rate of the second dehydration 2,3-BDO at 425 1C, noting the formation of butan-2-one as the
step, converting 3-buten-2-ol to BD, appears inversely propor- major product, while no BD was formed (Table 7, entry 7B).
tional to the applied pressure. Winfield then thoroughly inves- Next to the unique ThO2 catalyst of Winfield, a recent patent
tigated the adsorption equilibria of his catalytic system to showed promising dehydration of 2,3-BDO to BD in the pre-
elucidate the profound effect of water adsorption and the con- sence of an hydroxyapatite–alumina catalyst at 320 1C, yielding
comitant retardation of the second dehydration.186 BD yields above 45%.194
The work of Winfield on ThO2 may be considered as the Finally, an alternative and original strategy was recently
breakthrough in converting 2,3-BDO to BD, since most other proposed, which tries to bridge the fermentative production
catalytic materials produced butan-2-one.178,187 Bartok et al. of 2,3-BDO with its conversion to BD. The strategy is composed
have reported that the pinacol route is the preferred dehydra- of (i) the esterification of the diol with formic or acetic acid –
tion pathway in the presence of zeolites like Na+ and Na+/H+ which are naturally present in glucose-to-BDO fermentation
exchanged X and Y type faujasites at 250 1C.188,189 The authors liquors-, followed by (ii) the pyrolysis of the di-esters formed
noticed that a lower acidity within X-type zeolites not only led to with production of BD at 500 1C. The yields of step 1 are in the
more pronounced 1,2-elimination and thus BD, but also to a range of 70–85% depending on the acid, whereas the pyrolysis
decreased conversion rate.188 For instance, at 250 1C, use of the of the pure di-esters yields BD in the range of 82–94%. This
NaH–X zeolite showed 59% conversion, distributed over 78% appears to be an effective way of avoiding pinacol type reactions,
butan-2-one, 4% 2-methyl-propanal and 16% BD.179,188 Zeolites but requires pyrolysis at elevated temperatures.195
generally require high reaction temperatures for dehydration In conclusion, direct dehydration of 2,3-BDO into BD in high
and therefore they also tend to catalyse fragmentation of the yields is very challenging due to the competitive butan-2-one
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
formation according to the pinacol rearrangement. While the high concentrations of unsaturated alcohols like 3-buten-1-ol
latter is catalysed preferably by most acid catalysts, Winfield’s and 2-buten-1-ol. Further increase of the temperature above
work of 1945 using ThO2 and recent work with hydroxyapatite 400 1C led to the predominant formation of BD.178 According to
prove that the competitive dehydration of 2,3-BDO to BD is Freidlin et al., with phosphate type catalysts THF was said not
possible, provided that the appropriate catalyst is used. There- to isomerise into the butenols, but to convert directly into BD at
fore, research urgently needs further and deeper exploration to temperatures above 360 1C.197
clarify the required catalytic properties and surface reactions. BD is thus formed both from THF and 3-buten-1-ol, as
Some companies159 very recently have announced development visually presented in Scheme 11. The primary pathway is said
of such new catalytic technology, in light of the recent break- to occur via 3-buten-1-ol, which is the product of dehydration
throughs in the production of 2,3-BDO from syngas and bio- via the catalysed 1,2-elimination mechanism.178 The work by
mass fermentation.158 Reppe et al. has confirmed that the 1,4-cyclodehydration is
Dehydration of 1,4-BDO. The acidic dehydration of 1,4-diols mildly exothermic and rather easy, while the conversion of THF
(g-diols) such as 1,4-BDO almost always leads to the formation to BD reaction is highly exothermic and is much more difficult.198
of cyclic ethers under medium temperature conditions. As sum- Nevertheless, they have described a process converting 1,4-BDO
marized in a comprehensive chapter by Bartok and Molnar,178 vapours mixed with steam, over a supported H3PO4 catalyst bed,
the 1,4-cyclodehydration reaction is nearly independent of the containing sodium phosphates. The residence time could be
chemical structure and the degree of substitution of the BDO. optimized to form almost quantitative yields of THF, including
It has been demonstrated with countless acidic catalysts,196 small percentages of BD. By separating BD and recycling THF
ranging from organic to mineral acids, metal salts like ZnCl2 over the catalyst bed, BD yields of over 90% were shown.198,199 The
and oxides like alumina and phosphates in both liquid and gas reverse process, viz. 1,4-BDO production from BD, is currently
phase conditions.178 Cyclodehydration of 1,4-BDO leads to the commercially practiced. This process proceeds via 1,4-diacetoxy-
formation of tetrahydrofuran (THF, in Scheme 11). As this butene formation from BD and acetic acid and subsequent
dehydration process is highly selective to the cyclic ether, it is hydrogenation followed by a final acid hydrolysis. Acetic acid can
obviously one of three commercially practiced processes to be recycled and, next to 1,4-BDO, THF is formed.2,200,201
produce THF. In the vapour phase, the dehydration of 1,4-BDO Sato et al. have studied the dehydration of 1,4-BDO to
over chromium oxide, alumina or calcium phosphate at 250–320 1C unsaturated alcohols. A wide range of oxides such as Al2O3,
almost quantitatively leads to THF.2 SiO2–Al2O3, ZrO2, MgO were screened as well as plethora of
Production of BD from 1,4-BDO demands a higher reaction lanthanide oxides, with a special focus on CeO2.202,203 Logically,
temperature. The formation of THF over Ca3(PO4)2 is selective the different catalytic behaviour of SiO2–Al2O3 and Al2O3
in the temperature window of 250 to 320 1C. At higher tem- showed that acidity influences the dehydration activity. The
peratures, the formation of some BD was observed along with dehydration reaction over alumina at 200 1C (entry 1 in Table 7)
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
shows a conversion of 1,4-BDO of 17%, whereas full conversion long time has been an industrially practiced route.2,39,178 The
is obtained with SiO2–Al2O3 at the same temperature (entry 3, reaction, which proceeds mainly via the 1,2-elimination mecha-
Table 7). While at low conversions the selectivity to BD was nism, is catalysed by a plethora of acids, acid anhydrides and
high, viz. 70% for alumina, THF is the dominant product at full metal salts.
BDO conversion (entries 2 and 3), while no unsaturated alco- Different products such as BD, intermediately dehydrated
hols were formed. enol products, and other side-products have been encountered
Yamamoto et al.204,205 have investigated the use of ZrO2 for in the dehydration of 1,3-BDO. Their pathways are depicted in
the selective dehydration of 1,4-BDO to produce THF and Scheme 11. The intermediate products are the result of a
3-buten-1-ol at 275 1C and a WHSV of 9 h 1 (entry 4, Table 7). different regioselectivity in the first dehydration step. Accord-
The competitive formation of both products is likely due to the ing to the classic acid-catalysed 1,2-dehydration mechanism,
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
presence of both acid and basic sites on ZrO2.204 This surface the diol should transform to 2-buten-1-ol, the most stable
composition is able to compete for 1,4-cyclodehydration to THF alkene in agreement with Zaitsev’s rules. The anti-Zaitsev
and 1,2-elimination to unsaturated alcohols. In order to avoid product, 3-buten-1-ol, is formed abundantly with 1,3-BDO. In
the formation of THF, the authors decided to investigate purely addition, while the elimination of water from a secondary alcohol
basic cerium oxides. Usage of CeO2 completely avoided the is usually easier than from a primary one due to the higher
formation of THF. At 275 1C and a WHSV of 6.0 h 1, only 6% stability of the more substituted carbocation, the terminal hydro-
1,4-BDO conversion was noticed with a high BD selectivity close xyl in the case of 1,3-BDO also undergoes serious dehydration,
to 90%, while a 73% conversion was obtained at 425 1C albeit rendering the formation of 3-buten-2-ol, as a third enol product
with a selectivity shift from BD to 3-buten-1-ol (compare entries in the mixture. BD formation via dehydration of the three enols
5 and 6 of Table 7). Upon increasing the contact time with a occurs with different reactivity for each enol type. Competitive
factor of three, BD selectivity increased from 5 to 24% at the isomerization with the formation of butan-2-one is an unwanted
expense of 3-buten-1-ol (entry 7A, Table 7), while a 91% con- side-reaction in view of BD production. Another mode of
version of 1,4-BDO was obtained. These results highlight the dehydration, namely 1,3-cyclodehydration resulting in the forma-
importance of tuning the contact time to promote single or tion of the cyclic 2-methyl oxirane, is usually not encountered for
double dehydration. The excellent redox properties of CeO2 is 1,3-BDO. Next to dehydration, fragmentations may occur via
responsible for a selective stimulation of dehydration to the retro-aldolisation if one of the hydroxyls is able to undergo
unsaturated alcohol.202,206 A similar behaviour was noticed with dehydrogenation to yield a carbonyl group.
In2O3207 and weakly basic, heavy rare earth oxides202,203,206,208–210 Patent literature on the synthesis of BD from 1,3-BDO is
such as Yb2O3 (entry 8, Table 7). The latter catalyst was very active ample. The patents describe the conditions and catalyst com-
and selective for dehydrating 1,4-BDO selectively to 3-buten-1-ol positions, which favour BD formation such as supported and
at 375 1C and a WHSV of 20.3 h 1 with a selectivity of 82% at a unsupported acidic and neutral phosphates like Na-(poly)-
conversion of 40%. The very high WHSV (and thus low contact phosphates (at 270 1C)2 and CePO4 (at 320 1C).212 Supported
time) reveals a high activity of Yb2O3. Although rather basic in H3PO4 and more complex compositions have also been used.213
nature, the catalytic action of the rare-earth oxides likely pro- BD yields over 90% are not uncommon in these patents.
ceeded via a concerted acid–base mechanism, as hinted from gas Sato et al.211 have thoroughly investigated the conversion of
poisoning experiments.211 Sato et al. recently reviewed their 1,3-BDO in the vapour phase with focus on the production of the
efforts in the dehydration of diols with rare-earth oxides.211 unsaturated alcohol, 3-buten-2-ol, rather than on BD (Scheme 11).
In conclusion, the most favourable dehydration route of Such enols are important raw materials for the synthesis of
1,4-BDO is cyclodehydration, rendering the formation of THF. custom and added-value specialty chemicals. Although not aimed
Although THF may be further dehydrated to BD, this reaction at producing BD, these studies are relevant for gaining mecha-
requires severe thermal conditions. The more straightforward nistic insight into the dehydration of 1,3-BDO over solid acid and
route from 1,4-BDO follows the double 1,2-elimination with base catalysts. A selection of the catalytic data is displayed in
intermediate formation of 3-buten-1-ol. This pathway is less Table 7 (entries 9 to 18). Strongly acidic catalysts such as
dominant over acidic catalysts. Neutral to basic catalysts like SiO2–Al2O3 were more active and yielded BD next to the enols
phosphates, ZrO2, CeO2 and Yb2O3 favour the formation of the (entries 9 to 13, Table 7). 19% of 1,3-BDO could be converted
unsaturated alcohol, BD being formed at high enough contact into 20% BD, 44% 3-buten-1-ol and 14% 2-buten-1-ol over SiO2–
time and increased reaction temperature. However, since the Al2O3 at 200 1C and WHSV of 11.4 h 1. Increasing the reaction
formation of BD was never the purpose of the aforementioned temperature to 250 1C, both the conversion and the BD selectivity
catalytic studies, we anticipate that much better catalytic results increased to 74% and 36%, respectively, (compare entries 9
are expected soon once the contact time, catalyst composition and 10, Table 7). Interestingly, and in contrast to dehydration
and other engineering parameters are fine-tuned with focus on of 1,4-BDO, dehydration of 1,3-BDO is more selective to BD in
1,4-BDO dehydration to BD. the presence of acid catalysts.
Dehydration of 1,3-BDO to BD. Dehydration of 1,3-BDO is Less acidic oxides like alumina tend to catalyze other reaction
the textbook example of 1,3-diols dehydration. As mentioned channels. In the presence of alumina, formation of formalde-
above, 1,3-BDO is the intermediate in the Reppe acetylene- hyde and 4-methyl-1,3-dioxane with 28 and 18% selectivity was
based route to BD. In fact, the double acidic dehydration for a noticed next to some 3-buten-1-ol (entry 11). Formaldehyde likely
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
In the thermodynamic simulations of 1,4-BDO dehydration Table 8 Catalytic results for the DODH approach applied to erythritol
to BD, the presence of THF in the equilibrium composition was
E Catalyst Solvent/red. Feed BDa 1a 2a 3a 4a 5a Ref.
unimportant (data not shown). Thus THF is not thermo-
dynamically favoured, while its abundant formation in the catalytic 1b Cp*ReO3 C6H5Cl/PPh3 ery 80 obs.b obs.b — — — 217
2c Re2(CO)10 3-Octanol f ery — — — — — 62 221
experiments demonstrates the fast kinetics of the 1,4-cyclo- 3d Bu4NReO4 benzene/SO3 ery 27 0 3 0 0 6 222
dehydration. In the long run, BD formation is favoured, albeit with 4e MeReO3 3-Octanol f ery 89 0 0 0 0 11 234
a high energy barrier. Lowering the energy barrier will be a great 5e MeReO3 3-Octanol f threig 81 0 0 0 13 0 234
6h HReO3 3-Octanol f ery 73 0 0 0 0 7 234
challenge for catalytic scientists if production of BD from 1,4-BDO is 7e MeReO3 3-Octanol f 2i 70 0 0 0 6 0 235
aimed at. 8e MeReO3 3-Octanol f 3i 70 0 0 0 0 0 235
9j XReO3k PhCl/PPh3 ery 18 0 6 2 — — 229
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
3.3. Mechanistic and catalytic chemistry of deoxydehydration 10l XReO3k Pyridine/PPh3 ery 30 4 3 5 — — 229
11e XReO3k 3-Octanol f ery 67 0 0 0 — 7 229
Besides the dehydration route for converting BDOs into BD, there a
Yields of BD and compounds 1–5, as seen in Scheme 12i, are expressed in
also exists a catalytic deoxydehydration (DODH) methodology, mol%. b 0.044 M erythritol (ery), 1.5 mol% Re, 0.044 M PPh3, 135 1C, 28 h
involving dehydration with concomitant oxygen transfer to a obs = observed. c 1 M ery, 1.25 mol% Re, 1.7 mol% TsOH, 160 1C, 12 h, air.
d
reductant (Red.). This approach is generalized for vicinal diols in 0.2 M ery, 10 mol% Re, 0.3 M Na2SO3, 155 1C, 100 h. e 0.3 M ery,
2.5 mol% Re, 170 1C, 1.5 h, N2. f 3-Octanol plays the role of solvent and
eqn (4), L and M being ligand and (transition) metal, respectively. reductant. g Reaction with DL-threitol instead of erythritol, main side
product is a diastereoisomer of compound 4. h Conditions as e but at
155 1C for 5.5 h. i Resp. intermediates from Scheme 12i as substrate, 0.5 h of
(4) reaction. j 0.1 M erythritol, 2 mol% Re, 0.22 M PPh3, 180 1C, 24 h, N2. k X =
1,2,4-tri(tert-butyl)cyclopentadienyl. l Conditions as j but 15 h. E = Entry.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
yielding 1,4-anhydroerythritol (4 in Scheme 12), followed by a vicinal main product. This points to the high preference of catalytic
DODH reaction leading to the formation of 2,5-dihydrofuran DODH rather than cyclodehydration under the conditions used.
(5 in Scheme 12). The presence of the acid additive was sug- Yi et al. also used MeReO3, but mainly reported on a solvent-
gested to be the driver for cyclisation to 1,4-anhydroerythritol. less DODH approach, practiced with glycerol.227 To remove and
Such a mechanism corroborates with the well-established collect the gaseous products, a continuous distillation setup was
substitution-independent acidic cyclisation of g-diols (see the used, mainly producing allyl alcohol from glycerol. The latter is
1,4-BDO section).178 not only the reactant, it also participates in transfer hydrogenation
In 2011, Ahmad et al. have investigated a sulphite driven with co-production of dihydroxyacetone. With 2 mol% of MeReO3
oxorhenium catalysed approach.222 Bu4NReO4 (perrhenate salts) and 1.5 molar equivalents of 1-heptanol as a transfer reagent,
was tested with a Na2SO3 reductant in benzene for the DODH of erythritol and threitol were converted mainly into gaseous
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
erythritol (entry 3, Table 8) at 155 1C. BD, cis-2-butene-1,4-diol products.214 Using the continuous distillation procedure, no
and 2,5-dihydrofuran were encountered, with BD being the BD was condensed in the collector. Erythritol mainly yielded
major product (24%). MeReO3 was also used in combination 2,5-dihydrofuran and cis-but-2-enal, while meso-threitol was
with sulphites, but this Re catalyst performed less well. The converted into (3S,4S)-tetra-hydrofuran-3,4-diol, an isomer of
reaction products of these DODH reactions are fully in line with compound 4 in Scheme 12i.227 Differences with the results of
the proposed Re reaction system of Cook et al. with PPh3 (entry 1), Shiramizu and Toste234 (entries 4 and 5, Table 2) may originate
and helped corroborating the reaction path in Scheme 12i. either from the excess of sacrificial alcohol used or, more likely,
However, Nicholas et al.209 attributed the formation of from the differences in setup; viz. an open distillative system
2,5-dihydrofuran via fast 2,3-DODH, followed by a 1,4-cyclo- versus a closed approach. It appears that the open reactor
dehydration. Given the long reaction times, the sulphite225 promotes 1,4-cyclodehydration, next to DODH chemistry, pre-
system seemingly presents a less active DODH system for ferably leading to the cyclic species of Scheme 12 rather than to BD.
tetritols. Mechanistically, the first 1,2-diol DODH transforming The resemblance with the product spectrum of the Bergman
erythritol into 3-butene-1,2-diol is generally considered faster system is obvious.221 However, rather than being the effect of
than a 2,3-DODH (Scheme 12i). The second DODH, of 3-butene- Brønsted acidity (Bergman), here a plausible explanation could
1,2-diol into BD, is considered to be even faster, because this be a shift in reaction equilibria, caused by the open distillation,
diol was never noticed in the initial reaction stage. Although hereby promoting dehydration.
considerably slower, and thus not a major pathway, the initial To expand the scope of the oxorhenium catalyzed DODH,
2,3-diol DODH leading to the cis rather than to the trans form of Shiramizu and Toste have investigated different substrates.235
2-butene-1,4-diol, further confirms the preference of oxorhenium Oxorhenium compounds not only catalyse DODH chemistry,
for a syn-oriented elimination. but in parallel accelerate 1,3-OH shifts in allylic alcohols. The
Both Shiramizu and Toste234 and Yi et al.227 independently latter reaction thus formally presents 1,4-DODH and 1,6-DODH
have further explored the DODH reaction on erythritol, both with (eqn (5) and (6), respectively), rather than classic cis-vicinal or
a less complex and commercially available MeReO3 catalyst. The 1,2-DODH (eqn (4)).
former authors have provided a breakthrough in the frame of BD
production, by adopting a closed system in excess of a sacrificial
alcohol reductant – as in the work of Bergman221 – with the (5)
significantly simpler MeReO3 catalyst to assist the DODH. This
transfer hydrogenation, here demonstrated with 3-octanol to
3-octanone, is useful in view of the production of bio-alcoholic
solvents like isobutanol. Moreover, the formed ketone is easily
reduced and recycled, if necessary, or may be used in other (6)
reactions like aldol condensation. The authors have reported
very efficient double DODH on erythritol and DL-threitol at
170 1C, yielding BD in 89% yield at high conversion (entries 4 For intermediates like cis- and trans-2-butene-1,4-diol
and 5, Table 8). The major side-product from erythritol was (2 and 3, Scheme 12i) tested under conditions similar to those
2,5-dihydrofuran in line with the Bergman’s report, suggesting a of the original tetritol, very smooth progression of 1,4-DODH
dominant 1,4-cyclodehydration activity (see Scheme 12i). In the chemistry has been noticed for both isomers.222 The catalytic
case of DL-threitol, the major side-product was a trans isomer of results (entries 7 and 8, Table 8) indicate that the formation of a
1,4-anhydroerythritol, a cyclic internal trans-diol. Its formation 7-membered rhenium-diolate from the catalyst and the 1,4-diol
demonstrates the need for cis-diol substrates to get an efficient is excluded as both isomers showed identical reactivity. On this
vicinal-directed DODH. To check whether the introduction basis, the 1,3-OH shift causing isomerisation of allylic alcohols
of Brønsted acidity would result in an increase of the 2,5- may be included in Scheme 12i (dotted line). Note that Cook
dihydrofuran yield via the acidic 1,4-cyclodehydration path et al.217 already noticed the formation of 3-butene-1,2-diol from
(Scheme 12i), a reaction with HReO4 as a catalyst was performed cis-2-butene-1,4-diol, though at a much slower rate than the
under similar conditions (entry 6). However, no changes in DODH reaction. Shiramizu and Toste have proposed a catalytic
the product distribution were noted, while BD remained the cycle for their tandem isomerisation/DODH protocol, which is
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
in fact the combination of a 1,4 and 1,2-DODH (Scheme 12ii). systems are worthy of exploring for BD production from tetri-
In the tentative catalytic cycle it is proposed that the 1,4-DODH tols, provided 2,5-dihydrofuran formation can be suppressed.
reaction proceeds through a 5-ring rheniumV-diolate before
expelling the isomerised olefin, i.e. 3-butene-1,2-diol (1) in the
case of erythritol. This intermediate is identical to the one in 4. Conclusions and perspectives
the last step of the final 1,2-DODH reaction, just before BD
expulsion (Scheme 12ii). This merged isomerisation-DODH On-purpose synthesis of BD has received a recent revival of
mechanism might explain the high overall DODH efficiency interest because its availability from naphtha and gas crackers
of the MeReO3–3-octanol system, in conjunction with the high is diminishing. Worldwide investment in BD extraction capa-
substrate solubility. Theoretical density functional studies city at the crackers and dehydrogenation of the butane fraction,
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
by Qu et al. have demonstrated the possibility of a pathway either as a component of natural gas and shale gas or as a waste
involving a MeReO(OH)2 species formed from MeReO3 and fraction from petrorefinery, are two obvious ways to satisfy the
the reducing alcohol,231 coordinating diols through double needs of a firm BD market. Although the butane dehydrogena-
dehydration. Thus, the alcohol may act as a shuttle facilitating tion route to BD is an energy consuming process because of its
various hydrogen-transfer steps. Other mechanistic proposals, endothermicity, state-of-the-art technology has proven its relia-
elaborated for the synthesis of trans-stilbene, involve a bility in the past and new initiatives were recently announced.
rheniumIII-diolate, formed from 3-octanol reduction of methyl- However, driven by the shortage of light alkanes due to light-
dioxorheniumV.228 Finally, Raju et al. reported the conversion ening of cracker feeds and the increasing global demand for
of erythritol via DODH with a bulky 1,2,4-tri(tert-butyl)cyclo- bio-derived chemicals, biomass-based conversion might offer
pentadienyl-ReO3 catalyst.229 Entries 9 and 10 in Table 8 show new avenues for on-purpose BD production from renewable
these results, using PPh3 as a reductant in chlorobenzene or feedstock. The major technical challenge here is to produce BD
pyridine. Low BD yields were noticed and several intermediates, from biomass and/or biomass-derived syngas that will be cost
including trans-2-butene-1,4-diol, were obtained, pointing to an competitive with fossil oil-based BD. Several research and
inefficient DODH protocol. Under the conditions of Shiramizu development needs are required to answer critical questions
and Toste, 67% BD yield was noticed with the aforementioned regarding the technical, viz. process (operational) costs, and
bulky catalyst (entry 11), underlining the importance of solvent economic, viz. feedstock price and BD price, feasibility to
and the reductant in DODH reactions, as well as the appropri- integrate a bio-derived BD in the chemical industry.
ateness of using a sacrificial alcohol such as 3-octanol.221,223,234 Ethanol can be converted to BD. This route has been
As mentioned before, the four-carbon sugar-based feedstock practised on the industrial scale long before fossil oils and gases
is currently not produced on a large scale and certainly not on found general use. The industrial production of bio-ethanol is under
the scale requested for BD production. Therefore, it seems extensive development, enabling production of large amounts of
unlikely that it will become shortly a viable feedstock for BD ethanol in the near future. Other encouraging developments show
via DODH chemistry. Alternatively, the development has been ethanol production from syngas, either derived from natural gas,
reported for new routes to tetritols via decarbonylation,236 or to waste gas or biomass feedstock. Also methanol carbonylation,
tetroses via decarboxylation237 of larger carbohydrates such as followed by acetic acid hydrogenation might turn out promis-
pentoses and hexoses and via aldol condensation of glycolalde- ing for ethanol production. Meanwhile, fermentation became a
hyde.173 Pentose and hexoses are abundantly present in (hemi-)- very competitive process to ethanol.
cellulosic biomass,238–240 while glycol aldehyde is a dominant The conversion of ethanol to BD is thermodynamically
compound in bio-oils from biomass pyrolysis.173,241,242 Interest- feasible and proceeds via a sequence of reactions, receiving
ingly, the elegant DODH approach was also performed with general acceptance, viz. ethanol dehydrogenation, aldol con-
glucose-derived sugar alcohols such as mannitol and sorbitol, densation, MPV reduction, and dehydration. Clearly, a subtle
leading to a hexatriene as the major product.234 Such hexitol balance of the four reactions is required to achieve high
feedstocks can be directly derived from cellulosics.175,243,244 BD yield.
A second issue of current DODH next to feedstock avail- The literature overview indicates that a strong base like MgO
ability and cost lies in the use of the expensive complex Re or ZnO is essential to catalyse the aldol coupling, while Ag and
catalysts. Taking into account their turnover rate, their use on a CuO are good candidates to assist the dehydrogenation step.
large scale appears uneconomical for the production of BD. The catalytic system tolerates the presence of weak Brønsted
Note however that NH4ReO4 has a significantly lower cost.223 acidity, e.g. surface hydroxyls of silica, alumina or clays, cata-
For an industrial breakthrough of the DODH approach, lysing dehydration, though its amount and strength should be
research should be directed to the development of stable and limited, to avoid the thermodynamically favourable ethylene as
practical heterogeneous catalysts. Denning et al. have recently the side-product. The activity of the MPV hydrogen transfer
ventured into this area, with the synthesis of a carbon sup- reaction is likely most difficult to understand at the molecular
ported perrhenate DODH catalyst.224 In addition, the solvent- level, but it is related to the presence of Lewis acid sites,
less distillation of Ji et al.,214 which was mainly applied to on incompletely coordinated Mg2+ in MgO and some clays or on
convert glycerol, could be envisioned occurring in a full con- Al3+ in alumina and clays. The addition of ZrO2 is an interesting
tinuous mode. It seems that such open and continuous DODH option to assist aldol condensation and MPV reduction.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
Current BD selectivity with such catalysts ranges from 50 to (redox, acid–base) properties, is needed, ultimately helping
80%, the latter being obtained at modest conversion, thus improving overall catalytic performance.
showing room for improvement and development. Hydrocarbons Next to ethanol, there have been recent developments in
like ethylene and butenes and oxygenates like acetaldehyde are producing bio-derived 4-carbon alcohols, like n-butanol, iso-
produced, thereby reducing total BD selectivity. While ethylene butanol and various butanediols such as 1,3-BDO, 1,4-BDO and
may be valorised, the formation of butenes is unfortunate 2,3-BDO using fermentation of sugars. Also fermentation of
because of separation issues with BD. Another challenge of the syngas from natural gas or biomass has been demonstrated to
ethanol-to-BD reaction consists in its low volume productivity provide some of the diols. Dehydration of the monohydric
and often low BD concentration in the product stream. Current alcohols provides bio-derived butenes, which could be dehydro-
values range from 50 to 400 gBD h 1 per litre catalyst volume, with genated to BD, preferably in the presence of oxidants to
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
values below 100 being reported most frequently. Thus one faces compensate for the endothermicity. Availability of cheap alco-
the challenging task to develop more active and selective catalysts hols will allow commercialization of the ODH technology.
at higher ethanol concentration. A low ethanol conversion should Dehydration of butanediols to BD has received increasing
not always be an inconvenience per se as efficient recycling of interest given the availability of bio- and gas-derived BDO’s.
unconverted ethanol is optional, as well as a product recycle since Each diol has its own unique features and challenges. Dehydra-
many by-products can be formed from the various intermediates tion of 2,3-BDO faces a facile pathway to butan-2-one, following
in the reaction sequence. Besides recycling reactors, reactor types a pinacol-type rearrangement. Whereas higher temperatures
like a fluidized bed could favour the cascade of consecutive will stimulate BD formation thermodynamically, kinetic control
reactions to BD due to better heat supply and back-mixing;108 is key to achieve high BD yields. This remains the key challenge
there has been no systematic study of the influence of the reactor for the catalytic chemist, especially with the renewable produc-
design and reactor type on the production of BD from ethanol. tion of 2,3-BDO just around the corner. For as such unknown
Most of the research and catalyst developments were devoted to reasons, a ThO2 catalyst currently delivers the best catalytic
experiments in plug flow reactors. performances.
There is also little information available about the importance Dehydration of 1,4-BDO suffers from the production of THF.
of the purity of the ethanol for BD production. Sporadically, a Again, use of higher temperatures is part of the thermodynamic
beneficial effect of water is reported, possibly as a dilute of ethanol solution, although kinetics to a major extent still determine the
or as a reactant that poisons strong acid–base sites. Others selectivity. THF can be further processed into BD, albeit at high
reported on the usage of diethyl ether in the feed stream as a temperatures. There is another challenging route proceeding
better reagent for BD production. The presence of impurities of according to a double dehydration with intermediate formation
larger alcohols like isoamyl alcohol in the bio-ethanol feed has not of 3-buten-1-ol. Suitable catalysts are neutral to basic like CeO2
been investigated in the context of on-purpose BD production. and Yb2O3. Combination of effective catalysis with the recent
Catalysts of particular interest for further improvement breakthrough in fermentative 1,4-BDO production is a route to
would include a combination of the abovementioned elements. BD that probably will be picked up in the very near future.
Catalyst preparation and pretreatment techniques are very Dehydration of 1,3-BDO is a classical route, as this diol was
important in order to maintain high dispersion and ideal the key intermediate in the acetylene-based industrial BD
mixtures/phases of the different functions. Structural modifiers production long time ago. Therein, cheap Na–polyphosphate
are likely required to stabilize the high dispersion against sinter- catalysis allowed good catalytic dehydration results.2 Alternatively, a
ing. Scalability and cost of catalyst preparation/modification are two-step reaction with CeO2 may be envisioned producing selec-
important factors to consider as well with these multi-component tively 2-buten-1-ol and 3-buten-2-ol, followed by their dehydration
catalysts. to BD using common Brønsted acid catalysts like silica–alumina.
Whereas whole series of catalyst compositions have been However, the bottleneck in this approach resides in the develop-
tested, many reports unfortunately lack thorough physicochemical ment of a renewable production of 1,3-BDO.
characterization and kinetic approaches defining the ideal Finally, the pioneering work of deoxydehydration of butane-
catalyst in terms of acid–base and redox properties. Studies of diols and tetritols is an elegant chemistry to produce BD.
individual reaction types in the conditions of BD formation are Although this route is probably far from commercialisation,
highly welcome,90,245,246 but this is not always possible due to it shows that creative thinking could lead to original catalytic
the thermodynamic constraints. They will form a first step in pathways towards useful drop-in chemicals from bio-derived
the deeper understanding of the correlation between the active chemicals. Development of cheap catalysts is essential though,
sites. A better understanding of the adsorption phenomena and and perhaps heterogeneous catalysis may offer the solution.
surface chemistry is crucial, especially of the critical reaction
steps determining selectivity and rate limiting events. Advanced
(in situ and operando) spectroscopic tools are available today to Acknowledgements
better monitor the reaction processes at the microscopic and
molecular level and the catalyst behaviour with nanoscale E.V.M. acknowledges the KU Leuven (Spiro fellowship) for
resolution under reaction conditions. Crucial information on financial support. M.D. thanks the Research Foundation
the overall catalyst behaviour, in terms of textural and chemical Flanders (FWO Vlaanderen) for a postdoctoral fellowship.
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
W.J. thanks KU Leuven for PhD financing (FLOF). IAP (Belspo) 21 Z.-L. Xiu and A.-P. Zeng, Appl. Microbiol. Biotechnol., 2008,
and the Belgium government are acknowledged for general 78, 917–926.
financial support of the research of B.F.S. 22 M. J. Burk, A. P. Burgard, R. E. Osterhout, J. Sun and
P. Pharkya, WO 2012/177710, 2012.
23 R. E. Osterhout, A. P. Burgard and M. J. Burk, WO 2013/
Notes and references 028519, 2013.
24 Genomatica, Successful commercial-scale production of
1 R. F. Goldstein and A. L. Waddams, The petroleum chemi- BDO, 2013, http://www.genomatica.com/products/bdo/.
cals industry, E.&F.N. Spon LTD, London, 1967. 25 R. Hill and E. Isaacs, GB 483989, 1938.
2 H.-J. Arpe, Industrial Organic Chemistry, WILEY-VCH Verlag 26 M. Köpke, C. Mihalcea, F. Liew, J. H. Tizard, M. S. Ali,
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
GmbH & Co. KGaA, 2010. J. J. Conolly, B. Al-Sinawi and S. D. Simpson, Appl. Environ.
3 H. N. Sun and J. P. Wristers, Kirk-Othmer Encyclopedia of Microbiol., 2011, 77, 5467–5475.
Chemical Technology, John Wiley & Sons, Inc., 2000, 27 M. Mohammadi, G. D. Najafpour, H. Younesi, P. Lahijani,
pp. 365–392. M. H. Uzir and A. R. Mohamed, Renew. Sustainable Energy
4 K. Laird, Bio-based chemicals receive boost from shale gas Rev., 2011, 15, 4255–4273.
boom in North America, 2014, http://www.plasticstoday.com/ 28 G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006,
articles/bio-based-chemicals-receive-boost-from-shale-gas- 106, 4044–4098.
boom-in-north-america-140523?cid=nl.plas08. 29 Y. Liu, K. Murata, M. Inaba, I. Takahara and K. Okabe,
5 M. Hackett, Chemical Building Blocks from Renewables, Catal. Today, 2011, 164, 308–314.
2014, IHS Chemical Special Report, http://www.ihs. 30 J. R. Rostrup-Nielsen, Catal. Rev. Sci. Eng., 2004, 46,
com/products/chemical/planning/special-reports/chemical- 247–270.
renewables.aspx. 31 J. J. Spivey and A. Egbebi, Chem. Soc. Rev., 2007, 36,
6 G. Centi and R. A. van Santen, Catalysis for Renewables, 1514–1528.
Wiley-VCH Verlag GmbH & Co. KGaA, 2007, pp. 387–411. 32 V. Subramani and S. K. Gangwal, Energy Fuels, 2008, 22,
7 RFA’s Ethanol Industry Outlook, 2013, http://www.ethanolrfa. 814–839.
org/pages/annual-industry-outlook. 33 M. Scates, R. D. Shaver, J. Zink, R. Zinobile and
8 C. A. Cardona and Ó. J. Sánchez, Bioresour. Technol., 2007, O. Oyerinde, WO 2013/070216, 2013.
98, 2415–2457. 34 F. R. Bengelsdorf, M. Straub and P. Duerre, Environ.
9 L. Panella, Sugar Tech, 2010, 12, 288–293. Technol., 2013, 34, 1639–1651.
10 B. Kamm, P. R. Gruber and M. Kamm, Ullmann’s Encyclo- 35 D. W. Griffin and M. A. Schultz, Environ. Prog. Sustainable
pedia of Industrial Chemistry, Wiley-VCH Verlag GmbH & Energy, 2012, 31, 219–224.
Co. KGaA, 2000. 36 A. M. Henstra, J. Sipma, A. Rinzema and A. J. M. Stams,
11 C. S. Goh and K. T. Lee, Renew. Sustainable Energy Rev., Curr. Opin. Biotechnol., 2007, 18, 200–206.
2010, 14, 842–848. 37 M. Köpke, C. Mihalcea, J. C. Bromley and S. D. Simpson,
12 R. P. John, G. S. Anisha, K. M. Nampoothiri and A. Pandey, Curr. Opin. Biotechnol., 2011, 22, 320–325.
Bioresour. Technol., 2011, 102, 186–193. 38 P. C. Munasinghe and S. K. Khanal, Bioresour. Technol.,
13 M. Y. Menetrez, Environ. Sci. Technol., 2012, 46, 7073–7085. 2010, 101, 5013–5022.
14 V. Menon and M. Rao, Prog. Energy Combust. Sci., 2012, 38, 39 J. Grub and E. Löser, Ullmann’s Encyclopedia of Industrial
522–550. Chemistry, Wiley-VCH Verlag GmbH & Co. KGaA, 2000.
15 R. C. Saxena, D. K. Adhikari and H. B. Goyal, Renew. 40 Linde and BASF intend to jointly develop on-purpose
Sustainable Energy Rev., 2009, 13, 167–178. butadiene technology, 2014, http://www.basf.com/group/
16 A. Singh, P. S. Nigam and J. D. Murphy, Bioresour. Technol., pressrelease/P-14-249.
2011, 102, 10–16. 41 Honeywell UOP Licensing TPC Group’s Advanced Process
17 J. O. B. Carioca, Biotechnol. J., 2010, 5, 260–273. Technology to Produce Key Ingredient for Synthetic Rubber,
18 Axens, IFPEN and Michelin Join Forces to Create a Syn- 2014, http://www.uop.com/?press_release=honeywell-uop-
thetic Rubber Production Channel Using Biomass, 2013, licensing-tpc-groups-advanced-process-technology-to-produce-
http://www.axens.net/news-and-events/news/297/axens-ifpen- key-ingredient-for-synthetic-rubber.
and-michelin-join-forces-to-create-a-synthetic-rubber-production- 42 T. Black and J. Kaskey, Honeywell, BASF tout rival solutions
channel-using-biomass.html. for global butadiene shortage, 2014, http://www.hydrocarbon
19 Axens partners with Michelin, IFP in French research initiative processing.com/Article/3351368/Channel/194955/Honeywell-
for biobased butadiene, 2013, IHS Chemical Week, http://www. BASF-tout-rival-solutions-for-global-butadiene-shortage.html.
chemweek.com/markets/basic_chemicals/petrochemicals/ 43 J. M. Berak, R. Guczalski, J. Wójcik and S. Zalwert, Przem.
butadiene/Axens-partners-with-Michelin-IFP-in-French- Chem., 1962, 41, 80–84.
research-initiative-for-biobased-butadiene_56616.html. 44 G. Egloff and G. Hulla, Chem. Rev., 1945, 36, 63–141.
20 E. Celinska and W. Grajek, Biotechnol. Adv., 2009, 27, 45 J. A. Gamma and T. Inouye, Chem. Metall. Eng., 1942, 49,
715–725. 97–100.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
46 J. A. Gamma and T. Inouye, Chem. Metall. Eng., 1943, 50, 79 O. M. Vinogradova, N. P. Keier and S. Z. Roginskii, Dokl.
94–97. Akad. Nauk SSSR, 1957, 112, 1075–1078.
47 A. Talalay and M. Magal, Synthetic Rubber from Alcohol. 80 H. Niiyama, S. Morii and E. Echigoya, Bull. Chem. Soc. Jpn.,
A Survey Based on the Russian Literature, Interscience 1972, 45, 655–659.
Publishers, Inc., New York, 1945. 81 S. Kvisle, A. Aguero and R. P. A. Sneeden, Appl. Catal., 1988,
48 A. Talalay and L. Talalay, Rubber Chem. Technol., 1942, 15, 43, 117–131.
403–429. 82 Y. A. Gorin, K. N. Charskaya and A. V. Bochkareva, Sb.
49 L. Tong, Z.-Z. Liu and M.-h. Zhang, Chem. Ind. Eng., 2012, Statei Obshch. Khim., 1953, 2, 818–822.
29, 38–44. 83 V. Gruver, A. Sun and J. J. Fripiat, Catal. Lett., 1995, 34,
50 Y. Wang and S. Liu, J. Bioprocess Eng. Biorefinery, 2012, 1, 359–364.
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
116 E. V. Makshina, W. Janssens, B. F. Sels and P. A. Jacobs, 144 I. I. Ostromyslensky, J. Russ. Phys.-Chem. Soc., 1915, 47,
Catal. Today, 2012, 198, 338–344. 1509–1529.
117 L. Tong and Z.-Z. Liu, Xiandai Huagong, 2012, 32, 39–44. 145 A. Maximoff and O. Canonici, GB 535,678, 1941.
118 B. Kovařı́k, J. Benı́šek and J. Zavřel, Chem. Prum., 1960, 10, 146 A. Maximoff and O. Canonici, US 2,297,424, 1941.
81–83. 147 H.-J. Chae, T.-W. Kim, Y.-K. Moon, H.-K. Kim, K.-E. Jeong, C.-U.
119 R. I. Silva, R. Frety and Y. L. Lam, Actas do 9.o simpósio Kim and S.-Y. Jeong, Appl. Catal., B, 2014, 150–151, 596–604.
iberoamericano de catálise, Lisbon, Portugal, 1984, 148 C. Romanovsky and T. E. Jordan, US 2,800,517, 1957.
pp. 604–613. 149 C. Romanovsky and T. E. Jordan, US 2,819,327, 1958.
120 O. B. Krylov, Heterogeneous Catalysis, Akademkniga, 150 C. Romanovsky and T. E. Jordan, US 2,819,327, 1958.
Moscow, 2004. 151 B. B. Corson, E. E. Stahly, H. E. Jones and H. D. Bishop,
121 J. M. Berak, R. Guczalski and J. Wójcik, Acta Chir. Acad. Sci. Ind. Eng. Chem., 1949, 41, 1012–1017.
Hung., 1966, 50, 163–166. 152 M. J. Burk, D. S. J. Van, A. Burgard and W. Niu, WO 2008/
122 Y. Kitayama, M. Satoh and T. Kodama, Catal. Lett., 1996, 115840, 2008.
36, 95–97. 153 W. Clark, M. Japs and M. J. Burk, WO 2010/141780, 2010.
123 V. Aleixandre and T. Fernandez, An. Edafol. Fisiol. Veg., 154 H. Yim, R. Haselbeck, W. Niu, C. Pujol-Baxley, A. Burgard,
1958, 17, 805–832. J. Boldt, J. Khandurina, J. D. Trawick, R. E. Osterhout,
124 E. Suzuki, S. Idemura and Y. Ono, Appl. Clay Sci., 1988, 3, R. Stephen, J. Estadilla, S. Teisan, H. B. Schreyer,
123–134. S. Andrae, T. H. Yang, S. Y. Lee, M. J. Burk and S. Van
125 D. J. Butterbaugh and L. R. U. Spence, US 2,447,181, 1948. Dien, Nat. Chem. Biol., 2011, 7, 445–452.
126 D. J. Butterbaugh and L. R. U. Spence, US 2,423,681, 1948. 155 H. Gräfje, W. Körnig, H.-M. Weitz, W. Reiß, G. Steffan,
127 L. R. U. Spence, D. J. Butterbaugh and D. G. Kundiger, US H. Diehl, H. Bosche, K. Schneider and H. Kieczka,
2,436,125, 1948. Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH
128 M. Lewandowski, G. S. Babu, M. Vezzoli, M. D. Jones, Verlag GmbH & Co. KGaA, 2000.
R. E. Owen, D. Mattia, P. Plucinski, E. Mikolajska, 156 X.-J. Ji, H. Huang and P.-K. Ouyang, Biotechnol. Adv., 2011,
A. Ochenduszko and D. C. Apperley, Catal. Commun., 29, 351–364.
2014, 49, 25–28, DOI: 10.1016/j.catcom.2014.02.003. 157 M. J. Syu, Appl. Microbiol. Biotechnol., 2001, 55, 10–18.
129 B. Kovařı́k, Collect. Czech. Chem. Commun., 1961, 26, 158 J. Daniell, M. Köpke and S. Simpson, Energies, 2012, 5,
1918–1924. 5372–5417.
130 A. J. Dandy and M. S. Nadiye-Tabbiruka, Clays Clay Miner., 159 LANZATECH, INVISTA and LanzaTech Sign Joint Develop-
1982, 30, 347–352. ment Agreement for Bio-Based Butadiene http://www.
131 V. Aleixandre and T. Fernandez, An. Edafol. Fisiol. Veg., lanzatech.com/sites/default/files/imce_uploads/news_release_-_
1958, 17, 305–319. invista_announces_lanzatech_partnership_-_embargoed_
132 T. Kitayama and T. Wada, JP 57 102822, 1982. final_0.pdf.
133 Y. Kitayama and A. Abe, J. Chem. Soc. Jpn., Chem. Ind. 160 A. P. Burgard, M. J. Burk, R. E. Osterhout, P. Pharkya and
Chem., 1989, 11, 1824–1829. J. Li, WO 2013/036764, 2013.
134 Y. Kitayama, T. Hoshina, T. Kimura and T. Asakawa, 161 R. E. Osterhout, A. P. Burgard and M. J. Burk, WO 2013/
Proceedings of the 8th International congress on catalysis, 028519, 2013.
Verlag Chemie, Weinheim, FRG, Berlin (West), 1984, 162 A. P. Burgard, M. J. Burk, R. E. Osterhout and P. Pharkya,
pp. V647–V657. WO 2010/127319, 2010.
135 Y. Kitayama and A. Michishita, J. Chem. Soc., Chem. 163 P. Brandão, A. Philippou, J. Rocha and M. Anderson, Catal.
Commun., 1981, 401–402. Lett., 2002, 80, 99–102.
136 Y. Kitayama, K. Shimizu, T. Kodama, S. Murai, 164 M. A. Makarova, E. A. Paukshtis, J. M. Thomas, C. Williams
T. Mizusima, M. Hayakawa and M. Muraoka, Stud. Surf. and K. I. Zamaraev, J. Catal., 1994, 149, 36–51.
Sci. Catal., 2002, 142, 675–682. 165 M. Mascal, Biofuels, Bioprod. Biorefin., 2012, 6, 483–493.
137 Y. Kitayama and T. Wada, JP 58 059928, 1983. 166 R. M. West, D. J. Braden and J. A. Dumesic, J. Catal., 2009,
138 L. R. U. Spence and D. J. Butterbaugh, US 2,423,951, 1947. 262, 134–143.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.
View Article Online
167 P. Anbarasan, Z. C. Baer, S. Sreekumar, E. Gross, 196 I. Bucsi, Á. Molnár, M. Bartók and G. A. Olah, Tetrahedron,
J. B. Binder, H. W. Blanch, D. S. Clark and F. D. Toste, 1995, 51, 3319–3326.
Nature, 2012, 491, 235–239. 197 L. K. Freidlin and V. Z. Sharf, Izv. Akad. Nauk SSSR, Ser.
168 V. Garcı́a, J. Päkkilä, H. Ojamo, E. Muurinen and R. L. Khim., 1960, 1700–1702.
Keiski, Renew. Sustainable Energy Rev., 2011, 15, 964–980. 198 W. Reppe, Justus Liebigs Ann. Chem., 1955, 596, 80–158.
169 P. Dürre, Curr. Opin. Biotechnol., 2011, 22, 331–336. 199 W. Reppe, A. Steinhofer and G. Daumiller, DE 899350,
170 E. M. Green, Curr. Opin. Biotechnol., 2011, 22, 337–343. 1953.
171 P. C. A. Bruijnincx and B. M. Weckhuysen, Angew. Chem., 200 H. Katori and N. Murai, JP 07082191, 1995.
Int. Ed., 2013, 52, 11980–11987. 201 T. Haas, B. Jaeger, R. Weber, S. F. Mitchell and C. F. King,
172 M. Dusselier, P. Van Wouwe, F. de Clippel, J. Dijkmans, Appl. Catal., A, 2005, 280, 83–88.
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
D. W. Gammon and B. F. Sels, ChemCatChem, 2013, 5, 202 S. Sato, R. Takahashi, T. Sodesawa and N. Yamamoto,
569–575. Catal. Commun., 2004, 5, 397–400.
173 M. Dusselier, P. Van Wouwe, S. De Smet, R. De Clercq, 203 S. Sato, R. Takahashi, T. Sodesawa and N. Honda, J. Mol.
L. Verbelen, P. Van Puyvelde, F. E. Du Prez and B. F. Sels, Catal. A: Chem., 2004, 221, 177–183.
ACS Catal., 2013, 3, 1786–1800. 204 N. Yamamoto, S. Sato, R. Takahashi and K. Inui, Catal.
174 B. M. Kabyemela, T. Adschiri, R. M. Malaluan, K. Arai and Commun., 2005, 6, 480–484.
H. Ohzeki, Ind. Eng. Chem. Res., 1997, 36, 5063–5067. 205 N. Yamamoto, S. Sato, R. Takahashi and K. Inui, J. Mol.
175 S. Van de Vyver, J. Geboers, W. Schutyser, M. Dusselier, Catal. A: Chem., 2006, 243, 52–59.
P. Eloy, E. Dornez, J. W. Seo, C. M. Courtin, E. M. 206 A. Igarashi, N. Ichikawa, S. Sato, R. Takahashi and
Gaigneaux, P. A. Jacobs and B. F. Sels, ChemSusChem, T. Sodesawa, Appl. Catal., A, 2006, 300, 50–57.
2012, 5, 1549–1558. 207 M. Segawa, S. Sato, M. Kobune, T. Sodesawa, T. Kojima,
176 R. Fittig, Justus Liebigs Ann. Chem., 1860, 114, 54–63. S. Nishiyama and N. Ishizawa, J. Mol. Catal. A: Chem., 2009,
177 M. B. Smith and J. March, March’s Advanced Organic 310, 166–173.
Chemistry: Reactions, Mechanisms, and Structure, Wiley, 208 A. Igarashi, S. Sato, R. Takahashi, T. Sodesawa and
2007. M. Kobune, Catal. Commun., 2007, 8, 807–810.
178 M. Bartók and Á. Molnár, Ethers, Crown Ethers, Hydroxyl 209 S. Sato, R. Takahashi, M. Kobune, H. Inoue, Y. Izawa, H. Ohno
Groups and their Sulphur Analogues (1981), John Wiley & and K. Takahashi, Appl. Catal., A, 2009, 356, 64–71.
Sons, Ltd., 2010, pp. 721–760. 210 S. Sato, R. Takahashi, T. Sodesawa, N. Honda and
179 I. Bucsi, Á. Molnár, M. Bartók and G. A. Olah, Tetrahedron, H. Shimizu, Catal. Commun., 2003, 4, 77–81.
1994, 50, 8195–8202. 211 S. Sato, F. Sato, H. Gotoh and Y. Yamada, ACS Catal., 2013,
:
180 H. Waldmann and F. Petru, Chem. Ber., 1950, 83, 287–291. 3, 721–734.
181 J. van Haveren, E. L. Scott and J. Sanders, Biofuels, Bioprod. 212 W. Reppe and U. Hoffmann, DE 578994, 1933.
Biorefin., 2008, 2, 41–57. 213 L. P. Kyriakides, J. Am. Chem. Soc., 1914, 36, 987–1005.
182 W. J. Hale and H. Miller, US 2400409, 1946. 214 N. Ichikawa, S. Sato, R. Takahashi and T. Sodesawa, J. Mol.
183 W. J. Hale, US 2441966, 1948. Catal. A: Chem., 2006, 256, 106–112.
184 A. N. Bourns and R. V. V. Nicholls, Can. J. Res., Sect. B, 215 S. Sato, R. Takahashi, T. Sodesawa, A. Igarashi and
1947, 25B, 80–89. H. Inoue, Appl. Catal., A, 2007, 328, 109–116.
185 M. E. Winfield, J. C. S. I. R. (Aust.), 1945, 18, 412–423. 216 H. Gotoh, Y. Yamada and S. Sato, Appl. Catal., A, 2010, 377,
186 M. Winfield, Aust. J. Chem., 1950, 3, 290–305. 92–98.
187 S. V. Kannan and C. N. Pillai, Indian J. Chem., 1969, 7, 217 G. K. Cook and M. A. Andrews, J. Am. Chem. Soc., 1996, 118,
1164–1166. 9448–9449.
188 Á. Molnár, I. Bucsi and M. Bartók, Stud. Surf. Sci. Catal., 218 S. Stanowski, K. M. Nicholas and R. S. Srivastava, Organo-
1988, 41, 203–210. metallics, 2012, 31, 515–518.
189 A. Molnar, I. Bucsi and M. Bartok, Acta Phys. Chem., 1985, 219 L. Hills, R. Moyano, F. Montilla, A. Pastor, A. Galindo,
31, 571–579. E. Álvarez, F. Marchetti and C. Pettinari, Eur. J. Inorg.
190 J. Lee, J. B. Grutzner, W. E. Walters and W. N. Delgass, Chem., 2013, 3352–3361.
Stud. Surf. Sci. Catal., 2000, 130C, 2603–2608. 220 G. Chapman and K. M. Nicholas, Chem. Commun., 2013,
191 W. Zhang, D. Yu, X. Ji and H. Huang, Green Chem., 2012, 49, 8199–8201.
14, 3441–3450. 221 E. Arceo, J. A. Ellman and R. G. Bergman, J. Am. Chem. Soc.,
192 A. Multer, N. McGraw, K. Hohn and P. Vadlani, Ind. Eng. 2010, 132, 11408–11409.
Chem. Res., 2013, 52, 56–60. 222 I. Ahmad, G. Chapman and K. M. Nicholas, Organometallics,
193 B. Török, I. Bucsi, T. Beregszászi, I. Kapocsi and Á. Molnár, 2011, 30, 2810–2818.
J. Mol. Catal. A: Chem., 1996, 107, 305–311. 223 C. Boucher-Jacobs and K. M. Nicholas, ChemSusChem,
194 Y. H. Han and H. R. Kim, KR 2012107353, 2012. 2013, 6, 597–599.
195 J. Baek, T. Y. Kim, W. Kim, H. J. Lee and J. Yi, Green Chem., 224 A. L. Denning, H. Dang, Z. Liu, K. M. Nicholas and
2014, DOI: 10.1039/C4GC00485J. F. C. Jentoft, ChemCatChem, 2013, 5, 3567–3570.
Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2014
View Article Online
225 P. Liu and K. M. Nicholas, Organometallics, 2013, 32, 236 R. N. Monrad and R. Madsen, J. Org. Chem., 2007, 72,
1821–1831. 9782–9785.
226 J. E. Ziegler, M. J. Zdilla, A. J. Evans and M. M. Abu-Omar, 237 G. Hourdin, A. Germain, C. Moreau and F. Fajula, J. Catal.,
Inorg. Chem., 2009, 48, 9998–10000. 2002, 209, 217–224.
227 J. Yi, S. Liu and M. M. Abu-Omar, ChemSusChem, 2012, 5, 238 S. Van de Vyver, J. Geboers, P. A. Jacobs and B. F. Sels,
1401–1404. ChemCatChem, 2011, 3, 82–94.
228 S. Liu, A. Senocak, J. L. Smeltz, L. Yang, B. Wegenhart, J. Yi, 239 R. Rinaldi and F. Schüth, ChemSusChem, 2009, 2, 1096–1107.
H. I. Kenttämaa, E. A. Ison and M. M. Abu-Omar, Organo- 240 J. A. Geboers, S. Van de Vyver, R. Ooms, B. Op de Beeck, P. A.
metallics, 2013, 32, 3210–3219. Jacobs and B. F. Sels, Catal. Sci. Technol., 2011, 1, 714–726.
229 S. Raju, J. T. B. H. Jastrzebski, M. Lutz and R. J. M. Klein 241 M. S. Mettler, D. G. Vlachos and P. J. Dauenhauer, Energy
Published on 04 July 2014. Downloaded by Old Dominion University on 08/07/2014 07:03:52.
This journal is © The Royal Society of Chemistry 2014 Chem. Soc. Rev.