0% found this document useful (0 votes)
96 views21 pages

Coordination Chemistry Reviews: Juan Chen, Wesley R. Browne

Uploaded by

Siti Zhaafira
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
96 views21 pages

Coordination Chemistry Reviews: Juan Chen, Wesley R. Browne

Uploaded by

Siti Zhaafira
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 21

Coordination Chemistry Reviews 374 (2018) 15–35

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Photochemistry of iron complexes


Juan Chen, Wesley R. Browne ⇑
Molecular Inorganic Chemistry, Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747AG Groningen, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: Although iron is among the most abundant of the bio-essential transition metals and its coordination
Received 10 May 2018 chemistry is of central importance to bio-inorganic and bioinspired chemistry, its photochemistry has
Accepted 12 June 2018 been overshadowed by ruthenium polypyridyl complexes since the 1970s. The photochemistry of iron
complexes is nevertheless rich and presents a multitude of opportunities in a wide range of fields.
Here, we review the state of the art and especially recent progress in the photochemistry of iron com-
Keywords: plexes, focusing on aspects of relevance to environmental, biological and photocatalytic chemistry.
Iron
Ó 2018 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND
Photochemistry
Photoreduction
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Photooxidation
Photocatalysis

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2. Electronic structures and photophysics in iron complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3. Photochemistry of iron complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1. Photo-assisted Fenton Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2. Photo-induced ligand degradation – decarboxylation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3. Photo-induced release of small molecules. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3.1. Photo-induced N-N cleavage – N2 release. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3.2. Photo-induced displacement of labile ligands – CO release . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3.3. Photo-induced reductive elimination Fe-hydride – H2 evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4. Potential anticancer metallodrugs – photocytotoxicity of iron complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4.1. Photocytotoxicity of FeII complexes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4.2. Photocytotoxicity of FeIII complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4.3. Photocytotoxicity of FeIII-oxo bridged complexes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.5. Photochemistry of iron complexes in catalytic oxidations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.5.1. Photo-induced catalytic reaction—the use of a photosensitizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.5.2. Photo-catalytic reactions through direct photo-excitation of iron complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4. Conclusion and overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Appendix A. Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

1. Introduction honored by the 2016 Nobel prize for chemistry [1,2], and is central
to most life on this planet. At its most basic level, photochemistry
Photochemistry is a small but essential branch of chemistry, is the conversion of electromagnetic radiation to chemical energy
recognized for example, as the basis for the ‘‘molecular machines” [3] and enables induction of chemical transformations with spatial
and temporal control [1]. The chemical transformations and
⇑ Corresponding author. changes in reactivity form the basis of ‘‘photodynamic therapy”
E-mail address: [email protected] (W.R. Browne). treatments in medicine [4] and in photo(redox)catalysis [5], and

https://doi.org/10.1016/j.ccr.2018.06.008
0010-8545/Ó 2018 The Authors. Published by Elsevier B.V.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
16 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

hence exploring new photochemical processes in both organic and  The release of N2, for example, from porphyrin-ligated FeIII azide
inorganic molecular systems opens opportunities in medicine, complexes used to generate high valent iron complexes.
materials and chemical reactivity.  H2 evolution, for example, through reductive elimination from
The range of organic and inorganic compounds of interest to Fe-hydride complexes, which holds potential in energy storage.
photochemistry is limited due primarily to the requirement that  CO-release, mainly from Fe-CO complexes, which is of impor-
an accessible electronically excited state is either dissociative or tance in CO-related cytoprotection, anti-inflammation, and
is sufficiently long-lived to engage in energy or electron transfer, vasodilatory therapeutic treatments.
or to react with other compounds [6]. From an inorganic perspec-
tive, this demand has mostly limited attention to transition metal Iron is a bio-essential element and its complexes are well recog-
complexes such as those of chromium, ruthenium, and iridium nized as candidates in photometallodrugs in cancer treatment,
[7,8]. In the case of iron complexes, the lowest excited states are specifically DNA cleavage and photocytotoxicity as shown by the
metal centered (e.g., eg t2g) and are displaced with respect to series of iron complexes discussed in Section 3.4. Last but not least,
the ground state facilitating rapid radiationless deactivation, and photocatalytic reactions using iron complexes are seeing increas-
hence quenching photochemical reactivity. Nevertheless there ing attention, with both heme and non-heme iron complexes as
are iron complexes that show significant photochemistry, but the photo-catalysts in the oxidation of organic substrates. This area
design of new photoreactive complexes presents the challenge of is discussed in the final section of this review.
identifying and understanding the approaches available to achiev-
ing photoreactivity. In this review, we will discuss the known pho- 2. Electronic structures and photophysics in iron complexes
tochemistry of iron complexes and categorize the various reaction
classes to build a picture of the state of the art. The known oxidation states of iron range from Fe0 to FeVI, all of
The reported iron complexes are mostly coordinated with which have been observed experimentally. Its cations have domi-
organic ligands in an octahedral or, less often, tetrahedral coordi- nated the field of transition metal oxidation chemistry, due to its
nation environment and the photochemistry reported to date cor- great importance in both bioinorganic and synthetic chemistry.
relates with formal oxidation state, spin state, as well as the ligand The chemistry of iron is enriched by the number of accessible spin
structure. Hence, we will begin our discussion by introducing elec- states; including high-, intermediate-, and low-spin iron com-
tronic structures and the electronic configurations of iron com- plexes. In bioinorganic and biomimetic chemistry, the majority of
plexes in their various oxidation states. iron complexes are in an octahedral or pseudo-octahedral environ-
The earliest, and perhaps the best-documented, photochemical ment. Scheme 1 illustrates the energetic ordering of the d orbitals
reactions are the photo-assisted Fenton and photo-induced decar- and the electronic configurations of the oxidation states from FeII
boxylation reactions, both of which are of relevance to environ- to FeIV in octahedral environments. Fe0, FeV, and FeVI are not
mental and materials science (see Section 3.1 and Section 3.2). included due to the lack of reports on photoactivity of their com-
The photochemically induced release of small molecules, a field plexes. Low-spin FeII complexes are the only diamagnetic members
of growing importance, is dominated by: of this series of possible oxidation and spin states with the rest

Scheme 1. (left) Oxidation and spin states of iron complexes in an octahedral geometry and (right) the ground state occupation of d-orbitals of oxo-iron(IV) complexes in
pseudo-octahedral geometry for non-heme and heme complexes (structures shown are the representative for each class).
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 17

being paramagnetic. Amongst the paramagnetic states, there is a Pavel and co-workers recently reported a low-spin Fe(III) complex
special case: the antiferromagnetic coupled arrangement of the with a relatively long lived (100 ps) doublet ligand to metal
FeIII dimer, in which there are five alpha d electrons on one FeIII charge transfer state (2LMCT) at room temperature [16], achieved
ion and five beta d electrons on the other. In a pseudo-octahedral using a strong r-donor and p-acceptor NHC ligand (Scheme 2).
complex, the degeneracy of the t2g and eg orbitals (Scheme 1 left) The 100 ps lifetime, although short, is promising as it is sufficient
is lifted further (Scheme 1 right). For example, in oxo-iron(IV) com- to engage in photochemical processes together with the spin-
plexes the five d orbitals are different in energy, and the ligands allowed radiative decay to the ground state.
can change the energy ordering of the orbitals; the dx2 y2 orbitals It is worth noting the earlier work of Toftlund, McGarvey and
are higher in energy than the dz2 in a heme ligand environment co-workers on the photochemistry of iron(II) polypyridyl com-
(e.g., a tetracarbene – iron(IV)oxo complex in Scheme 1) [9] and plexes in which population of the 3MC excited states lead to ligand
vice versa in most of the non-heme ligand environments (e.g., dissociation with recovery on the late nanosecond timescale. The
N4Py, TQA) [10]. In non-heme oxo-iron(IV) complexes, the ligand photochemistry described in the last section of this review shows
also has dramatic influence on the electron configuration (spin that long lived excited states are not necessarily essential to
sates). Although most of the reported non-heme oxo-iron(IV) com- achieve useful photochemical reactivity and instead photo-
plexes are in low-spin state (S = 1, e.g., [(N4Py)FeIV=O]2+) [11], induced generation of reactive species is of relevance also [22].
there are a few examples that are in a high-spin state (S = 2) with
a trigonal-bipyramidal geometry [12,13]. [(TQA)FeIV@O]2+ is the
only reported pseudo-octahedral oxo-iron(IV) complex with 3. Photochemistry of iron complexes
high-spin (S = 2) ground state [14].
The photochemistry of iron complexes was studied extensively Although exciting progress is being made in controlling the
prior to the 1970s, but has been overshadowed by the photo- photophysical properties of iron complexes, their excited state
physics and chemistry of ruthenium(II) polypridyl complexes for properties remain largely underexplored in comparison to their
which hundreds of variants are known [15]. The long lived excited ruthenium and osmium analogs. This, however, is not to say that
states observed in the latter provide ample opportunity to engage the photochemistry of iron complexes is limited. Indeed, iron com-
in photo-redox and other uni- and bi-molecular chemical reac- plexes show a rich and diverse range of photochemically induced
tions. Scheme 2a shows possible transitions in an octahedral ligand reactions, including photo-assisted Fenton Reactions, photo-
environment, including metal-centered (MC), ligand centered (LC), induced ligand degradation-decarboxylation and release of small
ligand to metal (LMCT), and metal to ligand charge transfer (MLCT). molecules.
Compared to their ruthenium analogs, the energy gap between the
t2g and eg orbitals is much less for iron complexes bringing the 3MC
3.1. Photo-assisted Fenton Reactions
state lower in energy than the other possible excited states. Hence,
the lifetime of the charge transfer states are in the sub picosecond
The photo-assisted Fenton reaction is one of earliest examples
domain in iron complexes [16] compared to the microsecond life-
of the photochemistry of iron complexes reported [23–25]. The
times observed in ruthenium complexes, e.g., the 3MLCT excited
Fenton (H2O2/FeII) [24,26] or Fenton like (H2O2/FeIII) reaction, are
state lifetime of [Ru(bpy)3]2+ is 1 ls [17,18] and for [Fe(bpy)3]2+
well known reactions, which are used mainly to generate reactive
it is <150 fs [19].
oxygen species (HO and HOO) with H2O as by-product and are
Recent efforts [20,21] to increase the lifetime of the 3MLCT
used widely in the treatment of waste water [27–30]. The produc-
states of iron complexes has focused on increasing the ligand field
tion of hydroxyl radicals (HO) is strongly accelerated under UV
strength (and hence raising the 3MC states’ energies) through the
irradiation due to the rapid regeneration of Fe2+ through photo-
use of strong r-donor ligands such as N-heterocyclic carbenes
reduction of Fe(OH)2+. It should be noted that reactions 1-(1, 2,
(NHC). This approach has lead to an increase in 3MLCT lifetime to
3), shown below, are not the only reactions involved, and that
a few tens of picoseconds in several iron(II) complexes. Notably,
the mechanisms by which these reactions proceed are still under

Scheme 2. (a) Jablonski diagram showing disposition of metal and ligand orbitals and possible electronic transitions for an octahedral ligand field complex, (b) Electronic
transitions for low-spin d5 iron(III) complexes [16].
18 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

debate, due in part to the possible direct photolysis of H2O2, as well tation. Pozdnyakov et al. proposed that intramolecular charge
as, the indistinguishability of Fe4+ and {Fe3+(HO)} [26]. transfer from oxalate ligand to the FeIII center results in reduction
to FeII, which is in contrast to that proposed earlier by Rentzepis
and co-workers in which the FeIII–O bond cleaves before electron
Fe2þ H2 O2 ! Fe3þ HO þ OH ð1-1Þ transfer. The recent advances in ultrafast high resolution transient
spectroscopy have enabled reexamination of this reaction. In 2017,
Fe3þ H2 O2 ! Fe2þ HOO þ Hþ ð1-2Þ the Gilbert and co-workers [38] proposed a mechanism based on
direct spectroscopic evidence for the first step in this photolysis
FeðOHÞ2þ hm Fe2þ HO ð1-3Þ reaction (Fig. 1). Electron transfer occurs within 0.1 ps of photoex-
!
citation and results in the formation of an intermediate ferrioxalate
radical anion, which then dissociates rapidly to form thermally
3.2. Photo-induced ligand degradation – decarboxylation excited CO2 and CO II
2 . The CO2 relaxes and then leaves the Fe cen-
ter whereas the CO 2 radical anion remains coordinated for 10 ns.
The photochemical activity of FeIII carboxylato complexes, is The photochemistry of FeIII-carboxylato complexes in aqueous
generally more pronounced than the solvated FeIII ion, due to the solutions is highly dependent on the ligand environment [43]. As
possibility of metal to ligand charge transfer (LMCT) excitations for oxalate (L1), di-carboxylato ligands, such as succinate (L5),
in the former. LMCT excitation results in the transient reduction citrate (L8), and glutarate (L10), coordinate strongly to FeIII centers,
of the iron center and can lead to oxidative degradation of the and their complexes show strong LMCT absorption bands. Photol-
organic ligand [31–34]. Fig. 1 shows one of the most famous exam- ysis mostly follows the mechanism shown in Fig. 1. In contrast,
ples of such photochemistry, the ferrioxalato complex. The photo- complexes of mono-carboxylate ligands, 2-propanoate (L2), 2-
induced reduction of ferrioxalate was first reported by Parker in oxoacetate (L3), and gluconate (L9), show a dependence of the
1953 [35], followed by a report of similar photochemical activity photo reaction on ligand concentration, because the carboxylato-
in iron complexes bearing a carboxylato group in their ligand complexed FeIII complex is in equilibrium with FeIII(OH) species.
structures (Fig. 2) [32,34,36]. Among the carboxylato ligands At low ligand concentration, the photolysis of FeIII(OH) dominates,
reported, the ferrioxalato complex is used widely as an actinome- and produces OH primarily (Eq. 1-(1–3)).
ter in quantum yield determinations [37]. The mechanism of this The photo-induced decarboxylation of FeIII complexes bearing
photoreduction, which is accompanied by degradation of the car- carboxylato groups (L1-L10) is of substantial importance in the
boxylato ligand, was studied with pump/probe transient absorp- treatment of environmental pollutants. Indeed small carboxylate
tion spectroscopy and quantum chemical simulations by several compounds are abundant on earth and frequently invoked in bio-
groups [38–42]. There are two mechanistic aspects still under dis- geochemical cycles [43–46]. Attention has been directed to ligands
cussion; especially the steps immediately following the photoexci- of the type L11 over the last decades due to their potential applica-
tion in catalytic oxidations and structural variability [47]. The poly-
pyridine amine based tripodal amine chelated FeIII complex, [(L11)
FeIII-X], with a carboxylato moiety undergoes ligand decarboxyla-
tion under UV irradiation concomitant with reduction of FeIII ion
to FeII (Fig. 3) [47].
Siderophores are a wide range of compounds produced by bac-
teria. They bear the a-hydroxy carboxylic acid functionality and
coordinate readily to FeIII ions enabling the passive uptake of iron.
Recently, Butler and co-workers [48,49] showed that Fe(III)-
siderophore complexes containing an a-hydroxy acid group can
undergo photo-induced decarboxylation. Ligand to metal charge
transfer excitation upon UV irradiation results in decarboxylation
and oxidation of the petrobactin ligand to form a new ligand (loss
of the central carboxylic acid group with a 3-ketoglutarate group or
an enol group remaining on the original citrate backbone, Fig. 3).
Fig. 1. Proposed pathways for photo-induced ligand degradation of ferrioxalate. For
clarity, two of the oxalate ligands are omitted [36,38]. Indeed the photochemistry of siderophore-like FeIII complexes

Fig. 2. Structures of carboxylate ligands mentioned in the text.


J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 19

Fig. 3. Examples of ligand decarboxylation in FeIII complexes under irradiation [47,48,50].

(L12, L13) is important in the transportation of iron in vivo for phy- low toxicity of iron complexes is attractive in the development of
toplanktonic communities [49]. therapeutics in which these small molecules are released upon
Melman and co-workers applied photo-induced decarboxyla- irradiation, and hence can be released locally.
tion in FeIII complexes to achieve gel-sol transitions of a hydrogel
with UV or visible light [51]. The hydrogels consist of alginate
cross-linked with iron(III) cations. In the presence of sacrificial 3.3.1. Photo-induced N-N cleavage – N2 release
hydroxy carboxylates, irradiation leads to photoreduction of the Wagner and Nakamoto noted, already in 1989, the photo-
Fe(III) ions to Fe(II). This change in redox state results in dissocia- induced release of N2 from a porphyrin-ligated (L16-17) FeIII azide
tion from the alginate, and hence a loss of crosslinking that induces complex (thin film) under irradiation with UV–visible (406.7–
a gel-sol transition of the hydrogel. Later Ostrowski and co- 514.5 nm) light in frozen dichloromethane (30 K) [56]. Photo-
workers [52] developed this approach without the use of sacrificial induced heterolytic cleavage of the NAN bond was accompanied
components, using hydrogels consisting of Fe(III) ions and the by oxidation of the iron center to form an iron nitrido complex
polysaccharides poly[guluronan-co-mannuronan] (alginate, L14) (L)FeV„N with concomitant release of N2 (Fig. 5). The formation
or poly[galacturonan] (pectate, L15), (Fig. 4). Notably, they of the (L)FeV„N complex was confirmed by resonance Raman
observed that the photoreactivity was dependent on the configura- spectroscopy with the m(FeAN) band at 876 cm1 and 873 cm1
tion of the chiral center bound to the iron ions, which provides for (L16, L18)FeV„N and (L17)FeV„N, respectively [56]. Due to
additional control over the stability and photoresponse of metal- the substantial electron deficiency of (L)FeV„N, the photochem-
coordinated materials, and is of importance for broader application istry could only be studied in a cryogenic inert matrix or thin film.
to biological and tissue engineering [53–55]. As with the non-innocence of the porphyrin ligand in oxo-iron(IV)
enzymes (P450 compound I) [57], the d3-iron(V) center with a
3.3. Photo-induced release of small molecules closed shell dianionic ligand is a resonance structure of a d4-
configured iron(IV) center with the ligand radical monoanion by
The release of small molecules, especially NO and CO, are of virtue of magnetic coupling. Hence, the possibility of similar pho-
contemporary interest due to the biological activity of such com- tolysis in redox-innocent non-heme ligand environment is
pounds and especially their role in cellular signaling. The generally challenging.

Fig. 4. Examples of gel-solution transition utilizing a photo-induced FeIII-decarboxylation (adapted from [52]. Copyright ACS).
20 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

Fig. 5. Examples of photo-induced oxidation of porphyrin-ligated FeIII azide complex induced by the release of N2.

The photo-induced cleavage of a NAN bond in a non-heme iron adiabatic cooling (internal conversion, vibrational relaxation),
complex was first reported by Wieghardt and co-workers (L19, resulting in full conversion of electronic energy to thermal energy.
Fig. 6) [58], in which the FeIII center was located in a pseudo- The ‘‘productive” channel of N-N cleavage and buildup of the iron
octahedral coordination environment with azide ligands at its axial (V) product are due to intramolecular vibrational energy redistri-
positions and four equatorial sites occupied by a redox-innocent bution which corresponds to the azide-associated low frequency
macrocyclic ligand. High valent FeV intermediates were observed modes, leading to N-N cleavage. Furthermore, due to the relatively
at 4 and 77 K by EPR and Mӧssbauer spectroscopy. In contrast to high barrier for rebound between FeV/N2 to FeIII/N3, the formed
heme systems, a five-coordinate ferrous species has also been dinitrogen escapes the reaction cage readily [59].
observed in the same reaction, which originates from homolytic The photolysis pathways are highly dependent on reaction con-
cleavage of the FeAN bond. Although the details of the photochem- ditions and ligand environment. Vӧhringer and co-workers [60]
ical mechanism are not established, femtosecond mid-infrared found, in contrast to the studies in cryogenic inert matrices [64],
spectroscopy provides insight into the overall dynamics of the that irradiation of an FeIII azide precursor (L20), with one nitrogen
photo-induced release of N2 and formation of FeV [59–62]. Excita- atom replaced by an acetate group in an axial position trans to the
tion of the FeIII azide precursor at 266 nm results in mainly non- N3 group, almost exclusively forms the solvent-stabilized ferrous

Fig. 6. Examples of photo-induced oxidation (or reduction) of non-heme FeIII azide complexes with a release of N2 [58–60,63].
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 21

complex by Fe-N cleavage at 266 nm in acetonitrile at room tem- resolved infrared spectroscopy by George and co-workers [79].
perature [60]. Ferryl complex L21, generated electrochemically CO release proceeds in a step-wise manner, accompanied by a
from ferric states, undergoes similar N-N cleavage with evolution change in the symmetry of the complex. For example, at low tem-
of N2 under irradiation at 650 nm at 77 K with formation of (L21) perature (<20 K), one CO is released to form a C2v symmetric Fe
FeVI„N, which is the only identified FeVI complex to date [63]. (CO)4, with prolonged irradiation resulting in the release of second
Only a few illustrative examples of the photolysis of Fe-N3 com- CO to form Fe(CO)3. The formed Fe(CO)3 and Fe(CO)4 can recom-
plexes were discussed in this section but there are an increasing bine with CO or the matrix molecules, (e.g., CH4, Xe) depending
number of ligands developed, incorporating, for example, pyridine on the irradiation conditions [79–83]. Other ligands have been
moieties into the amine-based ligand backbone by Costas and co- incorporated into Fe-CO complexes to control the steric and elec-
workers [65,66], which show similar photoreactivity to the com- tronic properties at the iron centre. Lynam and co-workers [84]
plexes discussed above. It should be noted that, to date, photo- developed a series of tricarbonyl complexes containing 2-pyrone
induced N-N cleavage is still the most common route to convert ligands (L22CO-2). Tuning the substitution on the backbone signif-
FeIII-azido precursors to high-valent FeV or FeVI complexes, with icantly affects CO release [84]. In 2010, Westerhausen and co-
only a few reports of thermal reactions yielding these species [61]. workers reported a biogenic dicarbonyl bis(aminoethylthiolato)
iron(II) complex(L22CO-3) (Fig. 7a), which shows advantages in
regard to therapeutic applications. It is soluble in water, and the
3.3.2. Photo-induced displacement of labile ligands – CO release CO release is triggered by irradiation with visible light (k > 400
Carbon monoxide (CO) is a key molecule in biochemistry; nm). The complex shows minor adverse effects in physiological
in vivo, it is a natural metabolite and produced mainly by heme tests [85]. Later, inspired by hydrogenases [86], which release H2,
oxygenase-1 [67,68]. Certain levels of CO have a positive biochem- complexes L22CO-4 and L22CO-5 [87–91] containing diiron centres
ical effect; including cytoprotection, anti-inflammation, and vasodi- bearing thiolate ligands were prepared. These complexes release
lation, and it is often used in therapeutic treatments [69]. However, CO under irradiation and generate a solvent coordinated complex
due to the high affinity for the iron center of hemoglobin, excess CO (Fig. 7b). The CO-release rate is sensitive to the thiolates’ struc-
can shut down oxygen transportation [70]. Hence, the controlled tures. Complexes with two monothiolates (‘‘open” form) show
release of CO from carbon monoxide releasing molecules (CORMS) greater photoreactivity than a dithiolate coordinated complex
is a promising strategy in the targeted delivery of CO to tissues. (‘‘closed” form) [92]. The dimercaptopropanoate-bridged diiron
Metal-CO complexes have seen the most attention due to hexacarbonyl complex, reported by Fan and co-workers [90] shows
propensity for cleavage of the M–CO bond with release of CO under rapid CO-release with six CO ligands disassociating within 30 min
irradiation. There are several recent reviews published on this to 1 h, and forms a water soluble iron thiolate salt eventually [90].
topic [71–76] and here the field will be mentioned only briefly. Epithelial cell tests did not show obvious cytotoxicity. A common
Iron complexes are particularly attractive in CORM studies due to feature of these Fe0 and Fe1 complexes is that they tend to undergo
the low toxicity of iron, and in this section we focus on the full decomposition under irradiation, i.e. release all bound CO
reported Fe-CO photoCORMs. ligands. The non-heme [(N4Py)FeII(CO)] complex (L22CO-6,
The first Fe-CO photoCORMs for biological applications was Fig. 7c), reported by Kodanko and co-workers [93], shows similar
reported by Motterlini and co-workers in 2002 [75]. However, iron photo-induced CO release but with extraordinary thermal stability
pentacarbonyl, [Fe0(CO)5], which also shows photo-induced in aqueous solution. The polypyridyl ligand environment opens
release of CO, and its related complexes were reported prior to possibilities for ligand modification, taking advantage of the range
the 1970s [77,78]. Exposure of [Fe0(CO)5] to a cold light source of such ligands developed for oxidation catalysis over the last dec-
results in the release of CO, quantified using the conversion of ades. Modification of the ligand with a peptide provides a handle
deoxymyoglobin to carbonmonoxymyoglobin. The mechanism of by which photoCORM transportation to a target tissue can be
photolysis was studied using picosecond and nanosecond time- envisaged in therapeutic applications [93].

Fig. 7. Examples of photo-induced carbon monoxide releasing molecules (photoCORMs).


22 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

3.3.3. Photo-induced reductive elimination Fe-hydride – H2 evolution platinum-based metal complexes as metallodrugs. However, drug
In contrast to CO-release, in which a ligand is dissociated fully, the resistance and severe side effects stimulate the search for new
release of H2 is generally achieved by photo-induced reductive elim- non-platinum alternatives [101]. The successful application of
ination from an Fe-hydride complex; an approach reviewed by Per- Bleomycin [102], an iron-chelating antibiotic for cancer treatment
utz and Procacci recently [94]. The two main classes of complexes, drew attention to iron complexes and a series of biomimetic com-
for which photo-induced H2 release are observed, are monohydride plexes with a variety of ligand environments have been reported
and dihydride. The biomimetic hydrogenase complex (L23H2-1), a with pre-clinical testing towards cytotoxicity in many cases
representative example for Fe-monohydrides, was reported by [103]. Photo-activated or photodynamic treatment (PDT), in which
Rauchfuss and co-workers [95], showed four turnovers for H2 evolu- the drug is only active under irradiation and not cytotoxic in the
tion under irradiation in the presence of triflic acid (Fig. 8a) [95]. dark, appears to be a promising approach due to the high spatial
Fe-dihydride complexes are mainly based on Fe-carbonyl selectivity for targeting tumors [104,105].
(Fig. 8b) and Fe-phosphine structures (Fig. 8c). Sweany [96] noted Achieving photo-induced DNA cleaving activity and hence pho-
that irradiation of H2Fe(CO)4 with a Hg lamp resulted in the tocytotoxicity with iron complexes necessitates consideration of
appearance of a characteristic CAO stretching band (the totally several aspects: (a) binding of the complex to DNA, interaction
symmetric mode of Fe(CO)4) in matrix-isolation studies with FTIR with targets or related proteins for transportation, (b) transport
spectroscopy. This reaction is reversed upon visible radiation into cells, which is closely related to the lipophilicity of the com-
(using a Nernst glower). The recombination of Fe(CO)4 with the plexes, (c) high molar absorptivity in the PDT window, (d) oxida-
released H2 was inferred from the reappearance of IR bands for tion states of the metal center that are capable of generating
H2Fe(CO)4. In contrast to Fe-carbonyl complexes, Fe-phosphine reactive oxygen species under irradiation. The first three properties
dihydrides show greater photo-reactivity and have been studied can be addressed by ligand modification, while the last is key to the
extensively towards the activation of strong CAH bonds under irra- generation of reactive oxygen species and must consider the condi-
diation. The first step under irradiation is still the photo-reductive tions encountered in vivo. As there are a number of reviews cover-
elimination of molecular H2 (Fig. 8c), which leaves a vacant site on ing the cytotoxicity of iron complexes and their application in
the iron centre for small molecule oxidative addition and hence cancer treatment [103,106–108], here we focus only on recent
CAH or CAS activation of substrates [97–99]. reports of the photocytotoxicity of reported iron complexes giving
In addition to the above mentioned complexes, the so called
‘‘Janus intermediate”, which has a redox active 4[FeAS] core with Table 1
two bridging Fe hydrides (E4(4H)), was studied by Hoffman and Abbreviations of cell types mentioned in this section.
co-workers by in situ EPR and ENDOR spectroscopy. This complex Abbreviation Cell type Abbreviation Cell type
shows photo-induced reductive elimination of H2 at 20 K, and
HeLa Human cervical
reverts to (E4(4H)) by oxidation of H2 at 175 K.
carcinoma,
HaCaT Human keratinocytes, MCF-7 Human breast
cancer
3.4. Potential anticancer metallodrugs – photocytotoxicity of iron Hep G2 Human hepatocellular MRC-5 Human fibroblast
complexes carcinoma
HPL1D Nontransformed A549 Human non-small
human epithelial lung cell lung carcinoma
The report of cisplatin as an anticancer chemotherapeutic by
cells
Rosenberg [100] stimulated the development of several

Fig. 8. Examples of photo-induced H2 evolution by Fe-containing complexes.


J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 23

representative examples rather than a comprehensive discussion quinoxaline (dpq) FeII complex reported by Chakravarty and co-
of each system. Furthermore, we categorized reported iron com- workers is photo-inactive but its FeIII complex shows significant
plexes by oxidation state of the iron center here and for clarity DNA cleavage activity under visible light irradiation [114,115].
Table 1 lists the cell lines examined in photocytotoxicity studies Fig. 10 shows several recently reported iron(III) complexes
which are mentioned in this review. (Fig. 10) that engage in photo-induced DNA cleavage in different
cell lines under irradiation with visible light [116–118]. The con-
3.4.1. Photocytotoxicity of FeII complexes siderable number of complexes reported present certain ‘design
The earliest reports of photocytoxic properties of iron com- rules’. In these complexes (L31-L51), the phenolato moieties
plexes were predominantly focused on FeII complexes (Fig. 9). induce intense LMCT bands, which facilitates photo-irradiation
Roelfes and co-workers [109] reported that the DNA cleavage and broadens the PDT window to the 620–850 nm region (L46
activity of a polypyridyl amine-based pentadentate FeII complex – L49); the bulky tert-butyl (in L31-L36) and hydrophilic group
(L24), which is a biomimetic model of the widely used anti- (-SO3H, in L35) increase and decrease lipophilicity, respectively;
cancer drug Fe-Bleomycin. The DNA cleavage activity in vitro was planar aromatic groups increase their DNA binding affinity and
later shown to be enhanced by irradiation with visible light. Mod- act as additional photosensitizers (in L31-L36, L38, L41-L51);
ification of the ligand with covalently attached chromophores [119] biomarkers, e.g., biotin (vitamin H or vitamin B7, in L33
(L25-L28) resulted a 56-fold enhancement of the photoactivity and L34) and sugars (in L37-L39) increase cytotoxic selectivity
for single-stand DNA cleavage (L25). The mechanism of DNA cleav- for certain cancer cell lines; [120] Schiff base pyridoxal ligands,
age was originally attributed to the generation of reactive oxygen instead of tetradentate phenolate-based ligands allow other cellular
species (ROS), HO, O 1
2 , O2, under irradiation (355, 400.8, 473
components to be targeted (e.g., the endoplasmic reticulum) (L40-
nm) [109]. In contrast to covalent attachment, the presence of L44); [121,122] As in previous studies, the cytotoxicity discussed
chromophores (9-aminoacridine, and of 1,8-naphthalimide) did above is due to apoptosis induced by production of reactive oxygen
not show a synergistic effect in the photo-enhancement of the species upon irradiation of DNA bound complexes [123,119].
DNA cleavage activity of L24. ROS are generally accepted as the
responsible agents in DNA cleavage. However, surprisingly, addi- 3.4.3. Photocytotoxicity of FeIII-oxo bridged complexes
tion of ROS scavengers (Na3N, DMSO and superoxide dismutase), As a photo-metallodrug, an iron complex should have as low as
significantly increased DNA cleavage activity under irradiation. possible dark toxicity to healthy cells. In contrast to the FeII and
This unexpected result revealed the complexity of the ROS mecha- FeIII oxidation states, complexes in the oxidation states Fe0 and
nisms and the importance of maintaining steady-state concentra- Fe1 are generally unstable at ambient or physiological conditions.
tions of ROS in DNA cleavage [110]. Beyond the DNA cleavage, Complexes in the oxidation states FeIV or higher although also
the photocytotoxcity of these complexes towards living cells was invoked as reactive intermediates in oxidation reactions, are not
also studied and compared with the natural antibiotic Bleomycin; considered for PDT treatments. Diiron(III) complexes, however,
complexes L24 and L25 both show comparable efficiency in are possible candidates. Fig. 11 lists a collection of diiron com-
nuclear DNA cleavage with that of Bleomycin. However, the mech- plexes showing photocytotoxcity.
anisms are different. Both iron complexes induced apoptosis and The first report of photo-induced DNA cleavage by oxo-bridged
not cell cycle arrest induced by mitotic catastrophe as observed diiron(III) complexes was by Chakravarty in 2008 [124,125], in
with Bleomycin [111]. Apart from the pentadentate ligand bound which the almost linear Fe-O-Fe centers were coordinated by
complexes, two complexes reported recently, one bearing two pla- L-histidine and phenanthroline based ligands (L53-L54) (Fig. 11).
nar DNA binding phenanthroline groups(L29) [112], and the other These complexes display double-strand DNA cleavage under visi-
a boron dipyrromethene group attached to a NCN pincer (L30) ble light irradiation. The dipyrido quinoxaline (dpq) group (L54)
[113], both show significant photocytotoxcity to living cells (HaCaT was proposed to bind the DNA through groove binding. Notably,
and MCF-7 for L29, and HeLa and MCR-5 for L30) under visible irra- the complexes were flipped when binding with DNA; the two
diation with minor dark toxicity. dpq planar groups rotate via the Fe-O-Fe center to a trans configu-
ration in order to reduce the steric effect of binding. In addition, in
3.4.2. Photocytotoxicity of FeIII complexes contrast to the dpq group, the phenanthroline (L53) diiron complex
Several examples in which the iron(III) complexes of particular showed only single-strand DNA cleavage activity under visible
ligands show greater photo-reactivity than their corresponding light irradiation. Phenanthroline ligands in complexes L53 and
iron(II) complexes have been reported. For example, the dipyrido- L54 are not only the DNA binder but also the photosensitizer.

Fig. 9. Examples of FeII complexes showing photocytotoxicity.


24 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

Fig. 10. Examples of FeIII complexes showing photocytotoxicity.

In contrast, the complex L52, which lacks the photosensitizer, is two complexes are active only at shorter wavelengths, and com-
photo-inactive towards the DNA cleavage [124]. Diiron complexes plex L55 is photo-inactive. The mechanism of DNA cleavage in all
that do not have the L-histidine group but instead another dpq cases (L53, L54, L57, L58) is attributed partially to the Fe-
(L56 and L58) or phenanthroline (L55 and L57) ligand [126], bind carboxylato group present, which, see Section 3.2, under irradia-
DNA more strongly [124] than complexes L52-L54. The near lin- tion undergoes decarboxylation to produce reactive radical species
early oxo-bridged diiron complexes, L55 and L56 are more active that cleave DNA.
than the acetate bridged complexes L57 and L58. Photo-induced The cytotoxicity of these complexes, under specific conditions
DNA cleavage activity is wavelength dependent for these com- (excess H2O2), has been established to be greater in the case of
plexes: complex L58 shows activity under red light while the other cancer cells than in healthy cells. This difference is amplified by
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 25

Fig. 11. Examples of diiron(III) complex showing photocytotoxicity.

irradiation with visible light [127]. The photocytotoxic diiron(III) ent in the subsequent sections, it is likely that the sensitized
complex (L60), which bears a curcumin unit and not a FeIII- systems need to be reevaluated given the intrinsic photoreactivity
carboxylato group, shows enhanced stability and photocytotoxicity of the iron complexes used that was not considered in the original
towards HeLa and MCF-7 cells under irradiation. Curcumin is pro- studies.
posed to be the photo-active group since the analogous complex
L59 is photo-inactive [128]. 3.5.1. Photo-induced catalytic reaction—the use of a photosensitizer
As discussed above, iron is a bio-essential metal and its com- The use of a photosensitizer to introduce extra energy to a reac-
plexes generally show low toxicity in vivo, but can be activated tion is well established, not least in the field of photoredox cataly-
to induce a variety of cellular processes leading to cell death and sis. In regard to iron catalysis, the most widely used
thereby hold potential as photo-metallodrugs, not least in the photosensitizers are RuII complexes (e.g., [Ru(bpy)3]2+) due to their
treatment of cancer. Ligand structure plays a key role in determin- outstanding photophysical and photochemical properties. The
ing photocytotoxicity, and the abundance of small molecules, excited state of [Ru(bpy)3]2+⁄ generated under visible light irradia-
which are already well-documented in clinic studies for anticancer tion can engage in electron transfer to form [Ru(bpy)3]3+, which is a
treatment or reagent transportation, could be readily incorporated strong oxidant (1.26 V vs SCE, Fig. 12) and can oxidize most iron
into ligand structures. This increases the possibilities for the use of complexes to their higher oxidation states. This approach is a so-
iron complexes as photometallodrugs in PDT. However, the mech- called ‘‘multi-catalyst strategy” (Fig. 12).
anisms are still unclear with the lack of the information on the
photo-activity of the iron complexes themselves limiting under- 3.5.1.1. The application of photosensitized heme-based catalytic
standing of the cytotoxic mechanisms. Hence, close examination systems. Gray and co-workers [133] reported the photo-induced
of the photoreactivity of iron complexes is essential in order to generation of high-valent metalloenzyme intermediates in a heme
advance this field further. complex using the photosensitizer [Ru(bpy)3]2+ as electron donor
(Fig. 13). Nanosecond transient absorption showed that excitation
3.5. Photochemistry of iron complexes in catalytic oxidations of [Ru(bpy)3]2+ to its excited state [Ru(bpy)3]2+⁄ is followed by elec-
tron transfer to [Ru(NH3)6]3+ (electron acceptor, EA) to form the
The increase in the number and variety of organic ligands has strong oxidant [Ru(bpy)3]3+. It can directly oxidize ferric
opened many opportunities in controlling the reactivity of iron- microperoxidase-8 (MP8) [(P)FeIII-OH2] (L59) to its cation radical
ferric form [(P )FeIII-OH2], which is in equilibrium with the ferryl
+
based homogenous catalysts [129], and especially redox coopera-
IV
tivity between metal and ligand promises replacement of noble MP8 [(P)Fe @O] (Compound II). Lowering pH (pH <6) shifts the
metals by iron. Furthermore, modern spectroscopy provides exten- equilibrium away from the ferryl MP8 and vice versa [133]. Later,
sive information about catalyst structure and mechanism, which the same concept was applied to the heme Horseradish Peroxidase
has enabled biomimetic approaches to ligand design. High-valent (HRP) [134], with [Co(NH3)5Cl]2+ as the electron acceptor. The HRP
iron-oxo species are invoked as the reactive species in the oxida- ferryl species (Compound II) is formed via a ferric p-cation por-
phyrin radical species [(P )FeIII-OH2] intermediate at alkaline pH.
+
tion of organic substrates in both enzyme and synthetic catalytic
reactions frequently. There are several approaches to accessing Notably, the ferryl porphyrin radical species [(P+))FeIV@O] (Com-
high oxidation states, such as: sacrificial oxidants (e.g., H2O2, PhIO, pound I) formed by oxidation of Compound II was also observed
m-CPBA etc.) [130], electrochemical oxidation [131], and photo [134].
excitation [132]. Photochemistry is advantageous due to its non- As discussed above, the rate determining step is the porphyrin-
invasive and atom economic nature and is drawing more attention based ligand oxidation and hence this approach is unsuitable for
in recent years. In this section, we focus on the application of pho- the thiolate-ligated heme cytochromes P450 since the heme is bur-
tochemistry in catalytic oxidations by iron complexes. The photo- ied deep inside the enzyme. Cheruzel and co-workers [135]
induced generation of higher oxidation states or reactive oxygen reported a new strategy, in which the photosensitizer [(IA-phen)
species can be achieved by either of two ways: the direct excitation Ru(bpy)2]2+ (IA-phen = 5-iodoacetamino-1,10-phenanthroline)
and the indirect excitation with the use of a photosensitizer. We was covalently bound to cytochrome P450 BM3 (RuIIK97C–FeIII P450)
will discuss briefly the second approach, despite that on first (Fig. 14). Under irradiation three species with well-defined transient
glance it appears as a trivial example of photochemistry with iron absorption spectra were observed, first ⁄RuIIK97C–FeIII P450, second
complexes since the actual photochemistry is carried out by a sep- RuIII III II III
K97C–FeP450 and returning back to the initial RuK97C–FeP450. Kinetic
arate species (e.g., a ruthenium(II) complex). As will become appar- studies reveal three transient species assigned as RuIIK97C–(P+)
26 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

Fig. 12. Photochemistry of [Ru(bpy)]2+ (a) and a multi-catalyst strategy in catalytic oxidation with iron complexes (b).

Fig. 13. Examples of generation of high-valent metalloenzymes by a multi catalyst strategy.

+
FeIII II III II
P450(A)(OH2), RuK97C–[(P )FeP450(B)(OH2) and RuK97C–[(P)FeP450
IV
reduces the excited [Ru(bpy)3]2+⁄ to [Ru(bpy)3]+, which is a strong
II + III
(OH) (Compound II). The conversion of RuK97C–(P )FeP450(A)(OH2) reductant (1.26 V vs SCE). As with the modified photooxidative
to RuIIK97C–[(P+)FeIII
P450(B) (OH2) was rationalized by solvent or strategy, a polypyridyl Ru(II) moiety was attached covalently to a
polypeptide conformational changes, as with MP8 and HRP. The heme enzyme, however, the fate of the intermediate Ru(I)-Fe(III)
conversion from ferric radical cation RuII–[(P+)FeIII(OH2) to ferryl species is not as clear as in the photo oxidative strategy. Neverthe-
RuIIK97C–[(P)FeIV
P450(OH) is pH dependent, both of which are present less, CAH functionalization studies show higher total turnover
transiently and a fast recovery to the initial complex RuIIK97C–FeIII P450 numbers under irradiation with visible light than control reactions
is observed [135]. Similarly, Farmer and co-workers [136] reported [138], which further emphasizes that reactivity can be controlled
another example in which the photosensitizer [Ru(bpy)3]2+ was both by the varying the nature of photosensitizer and modification
attached covalently to a heme enzyme via a –(CH2)7– linker (L62). of the heme structure [139]. In short, several successful examples
This linking strategy was already reported by Oishi and co-workers to use photochemistry to trigger the reactivity of heme metalloen-
as earlier as 1999, who demonstrated that it facilitated intramolec- zymes have been described in this field can expect further develop-
ular electron transfer with observation of the ferric–porphyrin cation ment in the near future [140,141]. However, as will be discussed
radical spectroscopically [137]. In contrast to RuIIK97C–FeIII
P450, the dis- below, these studies may need to be revisited to take into consid-
tance between photosensitizer and heme unit was large enough to eration the intrinsic photochemistry of the iron complexes
prevent recovery from the ferric porphyrin radical species themselves.
RuII–(P+))FeIII to the initial RuII–FeIII state. The fate of this ferric
radical is either to oxidize the iron center to form the ferryl form 3.5.1.2. Photosensitized non-heme based catalytic systems. The num-
(Compound II) or oxidize a surrounding amino acid residue ber of heme and non-heme iron dependent enzymes involved in
(Fig. 13). The oxidation of the protein also suggested that the protein oxidations make mimicking such systems highly attractive in bio-
environment significant influenced the reaction [136]. mimetic catalyst design. Several high-valent complexes of bioinor-
In addition to the photoxidation strategy in Fig. 12b, there is ganic relevance have been reported generated with oxidants and
also a photoreductive strategy, in which the electron acceptor there are recently several reports of their photochemical genera-
([Ru(NH3)6]3+, [Co(NH3)5Cl]2+) is replaced by electron donor (e.g., tion. Fukuzumi and co-workers [132] reported the first photo-
diethyldithiocarbamate = DTC). As an excitation quencher DTC, chemical generation of a non-heme high-valent iron-oxo
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 27

Fig. 14. Examples of generation of high-valent metalloenzymes by a modified multi catalyst strategy.

complex with the pentadentate polypyridyl ligand N4Py (L63) in In addition to the photosensitized oxidative formation of high-
2010 using a photo-induced oxidative pathway (Fig. 12b). The pho- valent iron complexes, a photosensitized reductive pathway to
tosensitizer [Ru(bpy)3]2+ when excited with visible light is oxidized form a high-valent iron complex can be used in the catalytic oxida-
by the electron acceptor [Co(NH3)5Cl]2+ to form [Ru(bpy)3]3+, tion of PPh3 with several turnovers [145]. Instead of an electron
which in turn oxidizes the non-heme FeII complex (L63) to the FeIV- acceptor, Et3N was used as an electron donor. The excited state
@O complex (with water as the oxygen source) in a step-wise man- [Ru(bpy)3]2+⁄ was quenched to form Ru(I) complex [Ru(bpy)3]+
ner (Fig. 15) [132]. which is a strong reductant (1.26 V vs SCE) and reduces the
Later in 2014, Dhar and co-worker [142] reported the first pho- diiron(III) complex to mononuclear Fe(II). Subsequent reaction
tochemical generation of iron(V)-oxo with tetra-amidoma- with molecular oxygen to form l-peroxo diiron(III) complex and
crocyclic TAML ligands (L64 and L65). In this case, they started then higher oxidation states to form two iron(IV)-oxo moieties that
with FeIII complex as in heme system, and S2O2 8 was used as elec- are responsible for substrate oxidation (Fig. 17). Notably, the cova-
tron acceptor. The FeIII state was oxidized to FeIV@O state first as lently linked complex L67 shows lower photo-efficiency than the
[Ru(bpy)3]3+ is not strong enough to oxidize FeIV state to FeV state. non-covalently connected Ru/iron system, presumably due to
This step contrasts with the previous case. The formation of FeV deactivation of the ruthenium complexes excited state by deproto-
was attributed to the SO4 radical oxidation (Fig. 15) [142] and this nation of the imidazole linker [145]. In this studies dioxygen acti-
reactive intermediate is responsible for water oxidation. In 2017, vation and formation of l-peroxo diiron(III) complex were
the complex L64 was also studied for its photocatalytic hydroxyla- demonstrated, however, the formation of iron(IV)@O species was
tion and epoxidation reaction by Sen Gupta and co-workers [143]. not confirmed.
Notably, in this photocatalytic reaction the SO 4 radical species is In summary, in both heme and non-heme systems, the use of
not observed due to the use of [Co(NH3)5Cl]2+ as electron acceptor photosensitizers dramatically increases the reactivity of the iron
instead of S2O2 8 . The (L64)Fe
IV
monomer formed immediately complexes/enzyme. Covalent linking of the photosensitizer and
forms the dimer [{(L64)FeIV}2(l-O)]2+, which is the active oxidant iron complex further contributes to efficiency through intramolec-
[143]. ular electron transfer. However, as for heme based photosensitized
As with heme systems, the covalent linking strategy binding systems, in the non-heme systems there is still an open question as
photosensitizer to the iron complex was also used in non-heme to the impact the photoreactivity of the iron complexes them-
systems. Banse and co-workers [144] reported a chromophore- selves. As is discussed below, these latter systems open opportuni-
catalyst complex L66 (Fig. 16), in which the non-heme iron ties for photo-driven oxidation with a single catalyst.
complex was attached covalently to the photosensitizer ([Ru
(bpy)3]2+) as in the heme systems (L61 and L62). The complex 3.5.2. Photo-catalytic reactions through direct photo-excitation of iron
[RuII-FeII(OH2)]2+ was promoted to the excited state complexes
[⁄RuII-FeII(OH2)]2+ under irradiation (k = 450 nm), which then 3.5.2.1. Direct photo activation of mononuclear heme iron com-
formed [RuIII-FeII(OH2)]2+ by oxidation with [Ru(NH3)6]3+, followed plexes. Direct photoactivation of a heme complex was reported by
by intramolecular electron transfer from iron(II) center to ruthe- Newcomb and co-workers in 2005 [146]. The photooxidation of
nium(III) center to form [RuII-FeIII(OH)]2+. The high-valent of Compound II (a neutral iron(IV)-oxo porphyrin compound) to
[RuII-FeIV(O)]2+ complex was formed by a second cycle (Fig. 16) Compound I (a radical iron(IV)-oxo porphyrin compounds) occurs
[144]. under UV irradiation (k = 355 nm, Fig. 18b), for complexes L68,
28 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

Fig. 15. Examples of the photosensitized catalytic reactions in non-heme systems.

Fig. 16. Examples of the modified photosensitized catalytic reaction in non-heme systems.

horseradish peroxidase (HRP) (L69) and horse skeletal Myoglobin. (Fig. 18c) [147]. The porphyrin(IV)@O complex has a axial sub-
Usually, compound I models are formed by addition of terminal stituent (NO2, ClO2), which under UV irradiation (k = 350 nm)
oxidants (H2O2, PhIO, m-CPBA) to the porphyrin-iron(III) precursor, undergoes heterolytic cleavage to form iron(V)-oxo compounds,
followed by reaction with substrates to form the relatively stable which react over 100 times more rapidly with substrates than
compound II. Under irradiation, compound II forms compound I, the corresponding iron(IV) compound [147].
manifested in a change in UV–vis absorption and compound I per-
sists for several seconds in the absence of substrates [146]. 3.5.2.2. The direct photo activation of dinuclear heme iron com-
Later, the same group reported photochemical generation of plexes. The light-driven catalytic oxidation of substrates using a sin-
even higher oxidation states, i.e. an iron(V)-oxo porphyrin complex gle iron-containing catalyst that can undergo direct photoexcitation
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 29

Fig. 17. Examples of photo-induced reductive formation high-valent iron complex.

Fig. 18. Examples of photolysis reaction on heme system.

is highly attractive in simplifying the design of catalytic systems. In spring-loaded complex was further modified in the porphyrin ring
this regard, photo-induced disproportionation of porphyrin diiron with three pentafluorophenyl groups (L73), which showed higher
(III) complexes is one of most promising pathways. Richman and turnover numbers towards sulfide [153], olefin [153], and hydro-
co-workers [148] reported the first example of photo-induced carbon oxidation [154] under visible light irradiation with molec-
hetero-cleavage of the oxo-bridge of a l-oxo porphyrin diiron(III) ular oxygen as terminal oxidation and without use of a co-
[(FeTPP)2O]4+ complex (L70) with formation of 2 equiv. of FeIITPP reductant. An ethane linked cofacial diiron(III) l-oxo porphyrin
as the final product under UV-irradiation (O ? Fe LMCT band), reported by Rath and co-workers [155] showed photocatalytic oxi-
accompanied by the oxidation of the substrate PPh3. The Fe(III)- dation of P(OR3) (R: Me, Et) via a photo-induced disproportionation
O-Fe(III) complex was regenerated by oxidation with molecular reaction mechanism. The pillar linked cofacial diiron(III) l-oxo
oxygen. Formation of high-valent iron(IV)-oxo intermediates (FeIV- porphyrin complexes show a common feature in that they have
OTPP) via disproportionation was proposed based on substrate oxi- much smaller Fe-O-Fe angles (150–160°) compared to the 170–
dation patterns, as well as, quantum yields [148,149]. The water 178° of non-covalently linked complexes and favor attack on sub-
soluble complex (L71) showed similar photo-induced dispropor- strates in a side-on manner [155].
tionation reactions [150]. The first system with inequivalent ligands (i.e. heme/nonheme)
Nocera and co-workers [151] reported a strategy for selective was the l-oxo diiron(III) [(L)FeIII-O-FeIII(L’), L77] complex reported
oxidation of substrates using l-oxo porphyrin diiron(III) com- by Karlin and co-workers in 2004 (Fig. 19c) [156], which shows
plexes under photo catalytic conditions. Cofacial bisporphyrine photo-induced catalytic oxidation of a series of substrates, PPh3
l-oxo diiron(III) complexes bearing dibenzofuran (DPD, L74) and to OPPh3, tetrahydrofuran to c-butyrolactone, and toluene to ben-
xanthene (DPX, L75) (Fig. 19), in which the two ‘Pacman’ moieties zaldehyde. Transient absorption spectroscopy indicates that the
were used as a pillar to build up a molecular spring architecture, photo-induced disproportionation of the l-oxo diiron(III) to form
confined the attack on the substrates to favor a side-on geometry. an FeIV@O/FeII pair occurs, in which the FeIV@O is the reactive
The size of the two spacers controls the pocket size for the sub- towards substrate oxidation [156]. Notably, photo-induced dispro-
strates and inhibits recombination to form the l-oxo diiron(III) portionation of diiron(III) is not the only case reported, with an
states. The photocatalytic oxidation of substrates (dimethyl sulfox- even higher oxidation state, iron(V) generated by photo-induced
ide) was studied and compared with the complex without such a disproportionation of a bis-corrole-diiron(IV)-m-oxo dimer
spacer (L72). The DPD-bridged complex L74 shows similar quan- together with one equiv. iron(III). The reactivity of the iron(V)
tum efficiency to the non-bridged complex L72, but much higher intermediates is greater than that of the corresponding iron(IV)
efficiency towards substrate oxidation [151,152]. Later, this complex [157].
30 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

Fig. 19. Examples of photo-induced disproportionation of diiron(III) porphyrin complexes.

Form the discussion above, we can conclude that the photo- highly favorable coordination of acetonitrile to the FeII center in
induced disproportionation of l-oxo diiron(III)/diiron(IV) is a acetonitrile. Later, Bartlett and co-workers reported a non-heme
promising strategy in photocatalytic oxidations with iron com- iron(II) complex bearing a tetradentate (bpmcn) ligand (L80 in
plexes. The intermediate high-valent iron complexes formed are Fig. 20b), which undergoes similar photo-induced oxidation from
key reactive species. However, limitations remain; the quantum the iron(II) to iron(III) states by activation of dioxygen. In this case,
yield in these heme systems is quite low due to the large driving there are two labile coordination sites on the iron center compared
force for recombination of Fe(IV)@O and Fe(II) units, which shuts with the one site available in the L78 based complex and hence O2
down productive oxidation pathways. Compared to the heme sys- coordination is expected to be more facile, and photo-induced oxi-
tem, the non-heme iron complex provides greater flexibility in dation occurs in acetonitrile also [160].
terms of ligand modification. Furthermore the thermal reactivity Recently, the effect of near UV-excitation on the reactivity of a
toward a variety of substrates with non-heme high-valent iron- series of non-heme iron(IV)-oxo complexes towards CAH activa-
oxo complexes has been studied extensively. Hence it is worth- tion was reported (Fig. 21) [161]. Hydrogen atom abstraction
while to explore driving the oxidation of organic substrates by (HAT) of a CAH bond shows a large kinetic isotope effect (KIE) in
non-heme iron complexes photochemically. the thermal reaction between the Fe(IV)@O species and alkanes,
consistent with HAT as the rate-determining step, however,
3.5.2.3. The direct photo activation of non heme iron complexes. The although the reaction is accelerated by photoexcitation the KIE is
potential application of non-heme iron complexes under photo- much reduced (kH  kD). The wavelength dependence of the acti-
chemical conditions requires that they are stable under catalytic vation (only near-UV light accelerates the thermal reaction) sug-
conditions (e.g., FeII, FeIII, and FeIV in certain cases) and hence FeI gested the excitation into the near-UV ligand(oxo)-to-metal
and FeV complexes, which are highly reactive at ambient condi- charge-transfer (LMCT). This charger redistribution results in a
tions, are considered as intermediates only. Although non-heme weakening and hence elongation of the Fe(IV)AO bond and an
iron photochemistry is dominated by photo-induced decarboxyla- increase in its oxyl radical character, making it a more powerful
tion of iron(III) complexes, discussed in Sections 3.1, 3.2, and more CAH bond abstracting agent.
recent reports of the photochemistry of iron(II) complexes in rela- More recently, the photocatalytic oxidation of methanol was
tion to the activation of dioxygen (see below), direct activation (as reported using a single photo-catalyst, the non-heme iron(III) com-
opposed to systems in which a photosensitizer is used) of non- plexes [(L)FeIII-X]2+ (L = N4Py (L78), MeN4Py (L79); X = Cl, OCH3,
heme iron complexes is not readily apparent from the literature. Fig. 22)) [162]. Under anaerobic irradiation in methanol, the non-
The photo-induced oxidation of a non-heme iron(II) complex in heme iron(III) complex undergoes photo reduction to form the cor-
the presence of molecular oxygen as the terminal oxidant (L78 and responding [(L)FeII-X]2+ without ligand degradation, and, more
L79 in Fig. 20a) was reported first in 2009 [158]. The non-heme importantly, is accompanied by oxidation of sub stoichiometric
iron(II) complex (L78), designed as a functional mimic of iron bleo- methanol to formaldehyde. Mechanistic studies indicate the most
mycin and studied extensively in its reactivity with oxidants such likely reactive species is its corresponding l-oxo-dimer, an
as H2O2, forms reactive high-valent iron-oxo species with certain FeIII-l-O-FeIII complex, which is formed and in equilibrium with
terminal oxidants [159]. Irradiation of [(L78)Fe(II)OCH3]2+ in aero- monomer ([(L)FeIII-X]2+) complex upon dissolution in methanol,
bic methanol or [(L78)Fe(II)OH]2+ in H2O results in oxidation to the the dinuclear complex undergoes photo-induced disproportiona-
corresponding solvent-coordinated iron(III) complex. Under anaer- tion to form [(L)FeII-X]2+ and [(L)FeIV@O]2+ species, and it is the
obic conditions in acetonitrile oxidation does not occur due to the high-valent iron(IV)-oxo that is responsible for methanol
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 31

Fig. 20. Examples of photo-induced oxidation of non-heme iron(II) complexes with molecular oxygen as terminal oxidant.

4. Conclusion and overview

The photochemistry of iron complexes continues to surprise


and it is clear that the paradigm that iron complexes show limited
photochemistry due to the rapid excited state relaxation that
occurs via low-lying metal centered states of iron complexes is
increasingly challenged. The possibility of activating iron com-
plexes in a range of oxidation states towards oxidative transforma-
tions opened opportunities in photoredox catalysis. However,
whereas the heme systems are amenable to study by time resolved
spectroscopies, probing the excited states responsible for the pho-
Fig. 21. Direct photochemical activation of non-heme Fe(IV)@O complexes. tochemical reactions in non-heme iron complexes remains a chal-
lenge to be met.

Acknowledgments

Financial support was provided by The Netherlands Ministry of


Education, Culture and Science (Gravity Program 024.001.035 to
W.R.B.) and Chinese Scholarship Council (J.C.). COST action
CM1305 ECOSTBio is acknowledged for discussion.

Appendix A. Supplementary data

Supplementary data associated with this article can be found, in


the online version, at https://doi.org/10.1016/j.ccr.2018.06.008.

References

[1] T. van Leeuwen, A.S. Lubbe, P. Štacko, S.J. Wezenberg, B.L. Feringa, Dynamic
control of function by light-driven molecular motors, Nat. Rev. Chem. 1
Fig. 22. A non-heme iron photo-catalyst for light driven aerobic oxidation of (2017) 96, https://doi.org/10.1038/s41570-017-0096.
methanol. [2] E.C. Harvey, B.L. Feringa, J.G. Vos, W.R. Browne, M.T. Pryce, Transition metal
functionalized photo- and redox-switchable diarylethene based molecular
switches, Coord. Chem. Rev. 282–283 (2015) 77–86, https://doi.org/10.1016/j.
oxidation. The reactivity is further enhanced by excitation of the ccr.2014.06.008.
iron(IV)@O at same wavelength (vide supra). This mechanism [3] B. Vincenzo, C. Alberto, V. Margherita, Photochemical conversion of solar
was supported by the transient observation of the high-valent energy, ChemSusChem 1 (2008) 26–58, https://doi.org/10.1002/
cssc.200700087.
iron(IV)-oxo species upon irradiation into the LMCT band of the [4] P. Agostinis, K. Berg, K.A. Cengel, T.H. Foster, A.W. Girotti, S.O. Gollnick, S.M.
FeIII-l-O-FeIII complex at high concentrations. Together with the Hahn, M.R. Hamblin, A. Juzeniene, D. Kessel, M. Korbelik, J. Moan, P. Mroz, D.
previous report of photooxidation (from [(L)FeII-X]+ to [(L)FeIII- Nowis, J. Piette, B.C. Wilson, J. Golab, Photodynamic therapy of cancer: an
update, CA, Cancer J. Clin. 61 (2011) 250–281, https://doi.org/
X]2+) with molecular oxygen as terminal oxidant, the photocat- 10.3322/caac.20114.
alytic cycle (methanol oxidation to methanal with O2) is closed, [5] R.M.N. Yerga, M.C.Á. Galván, F. del Valle, J.A. Villoria de la Mano, J.L.G. Fierro,
and high turnover numbers were obtained by irradiation of [(L) Water splitting on semiconductor catalysts under visible-light irradiation,
ChemSusChem 2 (2009) 471–485, https://doi.org/10.1002/cssc.200900018.
FeIII-X]2+) as a single (photo)catalyst in aerobic methanol with only [6] V. Balzani, G. Bergamini, S. Campagna, F. Puntoriero, Photochemistry and
minor catalyst deactivation over time [163]. photophysics of coordination compounds: overview and general concepts BT
32 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

– photochemistry and photophysics of coordination compounds I, Springer [26] S.H. Bossmann, E. Oliveros, S. Göb, S. Siegwart, E.P. Dahlen, L. Payawan, M.
Berlin Heidelberg, Berlin, Heidelberg, 2007, pp. 1–36, https://doi.org/10.1007/ Straub, M. Wörner, A.M. Braun, New evidence against hydroxyl radicals as
128_2007_132. reactive intermediates in the thermal and photochemically enhanced fenton
[7] S. Campagna, F. Puntoriero, F. Nastasi, G. Bergamini, V. Balzani, reactions, J. Phys. Chem. A 102 (1998) 5542–5550, https://doi.org/10.1021/
Photochemistry and photophysics of coordination compounds: ruthenium jp980129j.
BT – photochemistry and photophysics of coordination compounds I, Springer [27] F. Harber, J. Weiss, The catalytic decomposition of hydrogen peroxide by iron
Berlin Heidelberg, Berlin, Heidelberg, 2007, pp. 117–214, https://doi.org/ salts, Proc. R. Soc. London. Ser. A – Math. Phys. Sci. 147 (1934) 332–351,
10.1007/128_2007_133. http://rspa.royalsocietypublishing.org/content/147/861/332.abstract.
[8] N.A.P. Kane-Maguire, Photochemistry and photophysics of coordination [28] M. Barbeni, C. Minero, E. Pelizzetti, E. Borgarello, N. Serpone, Chemical
compounds: chromium BT – photochemistry and photophysics of degradation of chlorophenols with Fenton’s reagent (Fe2+ + H2O2),
coordination compounds I, Springer Berlin Heidelberg, Berlin, Heidelberg, Chemosphere 16 (1987) 2225–2237, https://doi.org/10.1016/0045-6535(87)
2007, pp. 37–67, https://doi.org/10.1007/128_2007_141. 90281-5.
[9] S. Ye, C. Kupper, S. Meyer, E. Andris, R. Navrátil, O. Krahe, B. Mondal, M. [29] L. Cermenati, P. Pichat, C. Guillard, A. Albini, Probing the TiO2 photocatalytic
Atanasov, E. Bill, J. Roithová, F. Meyer, F. Neese, Magnetic circular dichroism mechanisms in water purification by use of quinoline, photo-fenton
evidence for an unusual electronic structure of a tetracarbene–oxoiron(IV) generated OH radicals and superoxide dismutase, J. Phys. Chem. B 101
complex, J. Am. Chem. Soc. 138 (2016) 14312–14325, https://doi.org/ (1997) 2650–2658, https://doi.org/10.1021/jp962700p.
10.1021/jacs.6b07708. [30] C. Domı ´ nguez, J. Garcı́a, M.A. Pedraz, A. Torres, M.A. Galán, Photocatalytic
[10] K.-B. Cho, H. Hirao, S. Shaik, W. Nam, To rebound or dissociate? This is the oxidation of organic pollutants in water, Catal. Today 40 (1998) 85–101,
mechanistic question in C-H hydroxylation by heme and nonheme metal-oxo https://doi.org/10.1016/S0920-5861(97)00125-9.
complexes, Chem. Soc. Rev. 45 (2016) 1197–1210, https://doi.org/10.1039/ [31] P. Borer, S.J. Hug, Photo-redox reactions of dicarboxylates and a-
C5CS00566C. hydroxydicarboxylates at the surface of Fe(III)(hydr)oxides followed with
[11] A.R. McDonald, L. Que, High-valent nonheme iron-oxo complexes: synthesis, in situ ATR-FTIR spectroscopy, J. Colloid Interface Sci. 416 (2014) 44–53,
structure, and spectroscopy, Coord. Chem. Rev. 257 (2013) 414–428, https:// https://doi.org/10.1016/j.jcis.2013.10.030.
doi.org/10.1016/j.ccr.2012.08.002. [32] C. Weller, S. Horn, H. Herrmann, Photolysis of Fe(III) carboxylato complexes:
[12] J. England, Y. Guo, K.M. Van Heuvelen, M.A. Cranswick, G.T. Rohde, E.L. Fe(II) quantum yields and reaction mechanisms, J. Photochem. Photobiol. A
Bominaar, E. Münck, L. Que, A More Reactive Trigonal-Bipyramidal High-Spin Chem. 268 (2013) 24–36, https://doi.org/10.1016/j.jphotochem.2013.06.022.
Oxoiron(IV) Complex with a cis-Labile Site, J. Am. Chem. Soc. 133 (2011) [33] J. Šima, J. Makáňová, Photochemistry of iron (III) complexes, Coord. Chem.
11880–11883, https://doi.org/10.1021/ja2040909. Rev. 160 (1997) 161–189, https://doi.org/10.1016/S0010-8545(96)01321-5.
[13] J. England, M. Martinho, E.R. Farquhar, J.R. Frisch, E.L. Bominaar, E. Münck, L. [34] I.P. Pozdnyakov, F. Wu, A.A. Melnikov, V.P. Grivin, N.M. Bazhin, S.V. Chekalin,
Que, A synthetic high-spin oxoiron(IV) complex: generation, spectroscopic V.F. Plyusnin, Photochemistry of iron(iii)-lactic acid complex in aqueous
characterization, and reactivity, Angew. Chem. Int. Ed. 48 (2009) 3622–3626, solutions, Russ. Chem. Bull. 62 (2013) 1579–1585, https://doi.org/10.1007/
https://doi.org/10.1002/anie.200900863. s11172-013-0227-6.
[14] A.N. Biswas, M. Puri, K.K. Meier, W.N. Oloo, G.T. Rohde, E.L. Bominaar, E. [35] A new sensitive chemical actinometer. I. Some trials with potassium
Münck, L. Que, Modeling TauD-J: A High-Spin Nonheme Oxoiron(IV) Complex ferrioxalate, Proc. R. Soc. London. Ser. A. Math. Phys. Sci. 220 (1953) 104
with High Reactivity toward C-H Bonds, J. Am. Chem. Soc. 137 (2015) 2428– LP-116. http://rspa.royalsocietypublishing.org/content/220/1140/104.
2431, https://doi.org/10.1021/ja511757j. abstract.
[15] A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser, A. von Zelewsky, Ru [36] Z. Wang, X. Chen, H. Ji, W. Ma, C. Chen, J. Zhao, Photochemical cycling of iron
(II) polypyridine complexes: photophysics, photochemistry, eletrochemistry, mediated by dicarboxylates: special effect of malonate, Environ. Sci. Technol.
and chemiluminescence, Coord. Chem. Rev. 84 (1988) 85–277, https://doi. 44 (2010) 263–268, https://doi.org/10.1021/es901956x.
org/10.1016/0010-8545(88)80032-8. [37] A new sensitive chemical actinometer – II. Potassium ferrioxalate as a
[16] P. Chábera, Y. Liu, O. Prakash, E. Thyrhaug, A. El Nahhas, A. Honarfar, S. Essén, standard chemical actinometer, Proc. R. Soc. London. Ser. A. Math. Phys. Sci.
L.A. Fredin, T.C.B. Harlang, K.S. Kjær, K. Handrup, F. Ericson, H. Tatsuno, K. 235 (1956) 518–536. http://rspa.royalsocietypublishing.org/content/235/
Morgan, J. Schnadt, L. Häggström, T. Ericsson, A. Sobkowiak, S. Lidin, P. Huang, 1203/518.abstract.
S. Styring, J. Uhlig, J. Bendix, R. Lomoth, V. Sundström, P. Persson, K. [38] D.M. Mangiante, R.D. Schaller, P. Zarzycki, J.F. Banfield, B. Gilbert, Mechanism
Wärnmark, A low-spin Fe(iii) complex with 100-ps ligand-to-metal charge of ferric oxalate photolysis, ACS Earth Sp. Chem. 1 (2017) 270–276, https://
transfer photoluminescence, Nature 543 (2017) 695, https://doi.org/ doi.org/10.1021/acsearthspacechem.7b00026.
10.1038/nature21430. [39] J. Chen, H. Zhang, I.V. Tomov, M. Wolfsberg, X. Ding, P.M. Rentzepis, Transient
[17] D. Patra, A.H. Chaaban, S. Darwish, H.A. Saad, A.S. Nehme, T.H. Ghaddar, Time structures and kinetics of the ferrioxalate redox reaction studied by time-
resolved study of three ruthenium(II) complexes at micellar surfaces: A new resolved EXAFS, optical spectroscopy, and DFT, J. Phys. Chem. A 111 (2007)
long excited state lifetime probe for determining critical micelle 9326–9335, https://doi.org/10.1021/jp0733466.
concentration of surfactant nano-aggregates, Colloids Surfaces B [40] J. Chen, H. Zhang, I.V. Tomov, P.M. Rentzepis, Electron transfer mechanism
Biointerfaces 138 (2016) 32–40, https://doi.org/10.1016/ and photochemistry of ferrioxalate induced by excitation in the charge
j.colsurfb.2015.11.037. transfer band, Inorg. Chem. 47 (2008) 2024–2032, https://doi.org/10.1021/
[18] J. Van Houten, R.J. Watts, Effect of ligand and solvent deuteration on the ic7016566.
excited state properties of the tris(2,20 -bipyridyl)ruthenium(II) ion in [41] J. Chen, A.S. Dvornikov, P.M. Rentzepis, Comment on ‘‘New insight into
aqueous solution. Evidence for electron transfer to solvent, J. Am. Chem. photochemistry of ferrioxalate”, J. Phys. Chem. A 113 (2009) 8818–8819,
Soc. 97 (1975) 3843–3844, https://doi.org/10.1021/ja00846a062. https://doi.org/10.1021/jp809535q.
[19] W. Gawelda, A. Cannizzo, V.-T. Pham, F. van Mourik, C. Bressler, M. Chergui, [42] I.P. Pozdnyakov, O.V. Kel, V.F. Plyusnin, V.P. Grivin, N.M. Bazhin, Reply to
Ultrafast nonadiabatic dynamics of [FeII(bpy)3]2+ in solution, J. Am. Chem. ‘‘Comment on ‘New insight into Photochemistry of Ferrioxalate’”, J. Phys.
Soc. 129 (2007) 8199–8206, https://doi.org/10.1021/ja070454x. Chem. A 113 (2009) 8820–8822, https://doi.org/10.1021/jp810301g.
[20] Y. Liu, T. Harlang, S.E. Canton, P. Chabera, K. Suarez-Alcantara, A. Fleckhaus, D. [43] Z. Wang, C. Chen, W. Ma, J. Zhao, Photochemical coupling of iron redox
A. Vithanage, E. Goransson, A. Corani, R. Lomoth, V. Sundstrom, K. Warnmark, reactions and transformation of low-molecular-weight organic matter, J.
Towards longer-lived metal-to-ligand charge transfer states of iron(ii) Phys. Chem. Lett. 3 (2012) 2044–2051, https://doi.org/10.1021/jz3005333.
complexes: an N-heterocyclic carbene approach, Chem. Commun. 49 (2013) [44] D.A. Thomas, M.M. Coggon, H. Lignell, K.A. Schilling, X. Zhang, R.H. Schwantes,
6412–6414, https://doi.org/10.1039/C3CC43833C. R.C. Flagan, J.H. Seinfeld, J.L. Beauchamp, Real-time studies of iron oxalate-
[21] Y. Liu, P. Persson, V. Sundström, K. Wärnmark, Fe N-heterocyclic carbene mediated oxidation of glycolaldehyde as a model for photochemical aging of
complexes as promising photosensitizers, Acc. Chem. Res. 49 (2016) 1477– aqueous tropospheric aerosols, Environ. Sci. Technol. 50 (2016) 12241–
1485, https://doi.org/10.1021/acs.accounts.6b00186. 12249, https://doi.org/10.1021/acs.est.6b03588.
[22] C. Brady, P.L. Callaghan, Z. Ciunik, C.G. Coates, A. Døssing, A. Hazell, J.J. [45] M. Passananti, V. Vinatier, A.-M. Delort, G. Mailhot, M. Brigante, Siderophores
McGarvey, S. Schenker, H. Toftlund, A.X. Trautwein, H. Winkler, J.A. Wolny, in cloud waters and potential impact on atmospheric chemistry:
Molecular structure and vibrational spectra of spin-crossover complexes in photoreactivity of iron complexes under sun-simulated conditions, Environ.
solution and colloidal media: resonance Raman and time-resolved resonance Sci. Technol. 50 (2016) 9324–9332, https://doi.org/10.1021/acs.est.6b02338.
Raman studies, Inorg. Chem. 43 (2004) 4289–4299, https://doi.org/10.1021/ [46] D. Nansheng, W. Feng, L. Fan, X. Mei, Ferric citrate-induced photodegradation
ic049809t. of dyes in aqueous solutions, Chemosphere 36 (1998) 3101–3112, https://doi.
[23] J. Kiwi, C. Pulgarin, P. Peringer, M. Grätzel, Beneficial effects of homogeneous org/10.1016/S0045-6535(98)00014-9.
photo-Fenton pretreatment upon the biodegradation of anthraquinone [47] J.E. Vernia, M.R. Warmin, J.A. Krause, D.L. Tierney, M.J. Baldwin,
sulfonate in waste water treatment, Appl. Catal. B Environ. 3 (1993) 85–99, Photochemistry and anion-controlled structure of Fe(III) complexes with an
https://doi.org/10.1016/0926-3373(93)80070-T. a-hydroxy acid-containing tripodal amine chelate, Inorg. Chem. 56 (2017)
[24] H.J.H. Fenton, LXXIII.-Oxidation of tartaric acid in presence of iron, J. Chem. 13029–13034, https://doi.org/10.1021/acs.inorgchem.7b01799.
Soc. Trans. 65 (1894) 899–910, https://doi.org/10.1039/CT8946500899. [48] K. Barbeau, G. Zhang, D.H. Live, A. Butler, Petrobactin, a photoreactive
[25] J.J. Pignatello, Dark and photoassisted iron(3+)-catalyzed degradation of siderophore produced by the oil-degrading marine bacterium marinobacter
chlorophenoxy herbicides by hydrogen peroxide, Environ. Sci. Technol. 26 hydrocarbonoclasticus, J. Am. Chem. Soc. 124 (2002) 378–379, https://doi.
(1992) 944–951, https://doi.org/10.1021/es00029a012. org/10.1021/ja0119088.
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 33

[49] F.C. Küpper, C.J. Carrano, J.-U. Kuhn, A. Butler, Photoreactivity of iron(III)– [73] C.C. Romão, H.L.A. Vieira, Metal carbonyl prodrugs: CO delivery and beyond,
aerobactin: photoproduct structure and iron(III) coordination, Inorg. Chem. in: Bioorganometallic Chem., Wiley-VCH Verlag GmbH & Co. KGaA, 2014, pp.
45 (2006) 6028–6033, https://doi.org/10.1021/ic0604967. 165–202, https://doi.org/10.1002/9783527673438.ch06.
[50] J.E. Grabo, M.A. Chrisman, L.M. Webb, M.J. Baldwin, Photochemical reactivity [74] P.V. Simpson, U. Schatzschneider, Small signaling molecules and CO-releasing
of the iron(III) complex of a mixed-donor, a-hydroxy acid-containing chelate molecules (CORMs) for the modulation of the cellular redox metabolism BT –
and its biological relevance to photoactive marine siderophores, Inorg. Chem. redox-active therapeutics, Springer International Publishing, Cham, 2016, pp.
53 (2014) 5781–5787, https://doi.org/10.1021/ic500635q. 311–334, https://doi.org/10.1007/978-3-319-30705-3_13.
[51] R.P. Narayanan, G. Melman, N.J. Letourneau, N.L. Mendelson, A. Melman, [75] R. Motterlini, J.E. Clark, R. Foresti, P. Sarathchandra, B.E. Mann, C.J. Green,
Photodegradable iron(III) cross-linked alginate gels, Biomacromolecules 13 Carbon monoxide-releasing molecules, Circ. Res. 90 (2002), e17 LP-e24.
(2012) 2465–2471, https://doi.org/10.1021/bm300707a. http://circres.ahajournals.org/content/90/2/e17.abstract.
[52] G.E. Giammanco, C.T. Sosnofsky, A.D. Ostrowski, Light-responsive iron(III)– [76] E. Kottelat, Z. Fabio, Visible light-activated PhotoCORMs, Inorganics 5 (2017),
polysaccharide coordination hydrogels for controlled delivery, ACS Appl. https://doi.org/10.3390/inorganics5020024.
Mater. Interfaces 7 (2015) 3068–3076, https://doi.org/10.1021/am506772x. [77] R. Fields, M.M. Germain, R.N. Haszeldine, P.W. Wiggans, Metal carbonyl
[53] V.A. Kumar, N.L. Taylor, A.A. Jalan, L.K. Hwang, B.K. Wang, J.D. Hartgerink, A chemistry. Part IX. Improved syntheses and some reactions of
nanostructured synthetic collagen mimic for hemostasis, Biomacromolecules tetracarbonylcyclo-octafluorotetramethyleneiron, J. Chem. Soc. A Inorg.
15 (2014) 1484–1490, https://doi.org/10.1021/bm500091e. Phys. Theor. (1970) 1964–1969, https://doi.org/10.1039/J19700001964.
[54] G.E. Giammanco, B. Carrion, R.M. Coleman, A.D. Ostrowski, Photoresponsive [78] R. Victor, R. Ben-Shoshan, S. Sarel, Ferraindene-iron tricarbonyl complexes by
polysaccharide-based hydrogels with tunable mechanical properties for a novel dehydrobromination of o-bromostyrene on photolysis with Fe(CO)5,
cartilage tissue engineering, ACS Appl. Mater. Interfaces 8 (2016) 14423– J. Chem. Soc. D Chem. Commun. (1971) 1241–1242, https://doi.org/10.1039/
14429, https://doi.org/10.1021/acsami.6b03834. C29710001241.
[55] X. Yang, Y. Guo, X. Luo, N. Zheng, T. Ma, J. Tan, C. Li, Q. Zhang, J. Gu, Self- [79] P. Portius, J. Yang, X.-Z. Sun, D.C. Grills, P. Matousek, A.W. Parker, M. Towrie,
healing, recoverable epoxy elastomers and their composites with desirable M.W. George, Unraveling the photochemistry of Fe(CO)5 in solution:
thermal conductivities by incorporating BN fillers via in-situ polymerization, observation of Fe(CO)3 and the conversion between 3Fe(CO)4 and 1Fe(CO)4
Compos. Sci. Technol. 164 (2018) 59–64, https://doi.org/10.1016/ (Solvent), J. Am. Chem. Soc. 126 (2004) 10713–10720, https://doi.org/
j.compscitech.2018.05.038. 10.1021/ja048411t.
[56] W.D. Wagner, K. Nakamoto, Resonance Raman spectra of nitridoiron(V) [80] M. Poliakoff, J.J. Turner, Infrared spectra and photochemistry of the complex
porphyrin intermediates produced by laser photolysis, J. Am. Chem. Soc. 111 pentacarbonyliron in solid matrices at 4 and 20 K: evidence for formation of
(1989) 1590–1598, https://doi.org/10.1021/ja00187a010. the complex tetracarbonyliron, J. Chem. Soc. Dalton Trans. (1973) 1351–1357,
[57] J. Rittle, M.T. Green, Cytochrome P450 Compound I: Capture, https://doi.org/10.1039/DT9730001351.
Characterization, and C-H Bond Activation Kinetics, Science 330 (2010) [81] M. Poliakoff, J.J. Turner, Structure and reactions of matrix-isolated
933–937. http://science.sciencemag.org/content/330/6006/933.abstract. tetracarbonyliron(0), J. Chem. Soc. Dalton Trans. (1974) 2276–2285, https://
[58] K. Meyer, E. Bill, B. Mienert, T. Weyhermüller, K. Wieghardt, Photolysis of cis- doi.org/10.1039/DT9740002276.
and trans-[FeIII(cyclam)(N3)2]+ complexes: spectroscopic characterization of [82] M. Poliakoff, Infrared spectrum of matrix isolated tricarbonyliron, J. Chem.
a nitridoiron(V) species, J. Am. Chem. Soc. 121 (1999) 4859–4876, https://doi. Soc. Dalton Trans. (1974) 210–212, https://doi.org/10.1039/DT9740000210.
org/10.1021/ja983454t. [83] T.J. Barton, R. Grinter, A.J. Thomson, B. Davies, M. Poliakoff, Magnetic circular
[59] J. Torres-Alacan, J. Lindner, P. Vöhringer, Probing the primary photochemical dichroism evidence for the paramagnetism of tetracarbonyliron(0): low-
processes of octahedral iron(V) formation with femtosecond mid-infrared temperature matrix studies, J. Chem. Soc. Chem. Commun. (1977) 841–842,
spectroscopy, ChemPhysChem 16 (2015) 2289–2293, https://doi.org/ https://doi.org/10.1039/C39770000841.
10.1002/cphc.201500370. [84] A.J. Atkin, I.J.S. Fairlamb, J.S. Ward, J.M. Lynam, CO release from
[60] J. Torres-Alacan, O. Krahe, A.C. Filippou, F. Neese, D. Schwarzer, P. Vöhringer, norbornadiene iron(0) tricarbonyl complexes: importance of ligand
The Photochemistry of [FeIIIN3(cyclam-ac)]PF6 at 266 nm, Chem. – A Eur. J. dissociation, Organometallics 31 (2012) 5894–5902, https://doi.org/
18 (2012) 3043–3055, https://doi.org/10.1002/chem.201103294. 10.1021/om300419w.
[61] J. Torres-Alacan, P. Vöhringer, Generating high-valent iron with light: [85] R. Kretschmer, G. Gessner, H. Görls, S.H. Heinemann, M. Westerhausen,
photochemical dynamics from femtoseconds to seconds, Int. Rev. Phys. Dicarbonyl-bis(cysteamine)iron(II): a light induced carbon monoxide
Chem. 33 (2014) 521–553, https://doi.org/10.1080/0144235X.2014.973197. releasing molecule based on iron (CORM-S1), J. Inorg. Biochem. 105 (2011)
[62] J. Torres-Alacan, U. Das, A.C. Filippou, P. Vöhringer, Observing the formation 6–9, https://doi.org/10.1016/j.jinorgbio.2010.10.006.
and the reactivity of an octahedral iron(V) nitrido complex in real time, [86] M.W.W. Adams, The structure and mechanism of iron-hydrogenases,
Angew. Chem. Int. Ed. 52 (2013) 12833–12837, https://doi.org/10.1002/ Biochim. Biophys. Acta – Bioenerg. 1020 (1990) 115–145, https://doi.org/
anie.201306621. 10.1016/0005-2728(90)90044-5.
[63] J.F. Berry, E. Bill, E. Bothe, S.D. George, B. Mienert, F. Neese, K. Wieghardt, An [87] J.W. Tye, M.Y. Darensbourg, M.B. Hall, De novo design of synthetic di-iron(I)
octahedral coordination complex of iron(VI), Science 312 (1937) (2006), LP- complexes as structural models of the reduced form of iron–iron
1941. http://science.sciencemag.org/content/312/5782/1937.abstract. hydrogenase, Inorg. Chem. 45 (2006) 1552–1559, https://doi.org/10.1021/
[64] T. Petrenko, S. DeBeer George, N. Aliaga-Alcalde, E. Bill, B. Mienert, Y. Xiao, Y. ic051231f.
Guo, W. Sturhahn, S.P. Cramer, K. Wieghardt, F. Neese, Characterization of a [88] J. Brown-McDonald, S. Berg, M. Peralto, C. Works, Photochemical studies of
genuine iron(V)–nitrido species by nuclear resonant vibrational spectroscopy iron-only hydrogenase model compounds, Inorganica Chim. Acta 362 (2009)
coupled to density functional calculations, J. Am. Chem. Soc. 129 (2007) 318–324, https://doi.org/10.1016/j.ica.2008.03.110.
11053–11060, https://doi.org/10.1021/ja070792y. [89] J. Marhenke, A.E. Pierri, M. Lomotan, P.L. Damon, P.C. Ford, C. Works, Flash
[65] G. Sabenya, L. Lázaro, I. Gamba, V. Martin-Diaconescu, E. Andris, T. and continuous photolysis kinetic studies of the iron-iron hydrogenase model
Weyhermüller, F. Neese, J. Roithova, E. Bill, J. Lloret-Fillol, M. Costas, (l-pdt)[Fe(CO)3]2 in different solvents, Inorg. Chem. 50 (2011) 11850–
Generation, spectroscopic, and chemical characterization of an octahedral 11852, https://doi.org/10.1021/ic201523r.
iron(V)-nitrido species with a neutral ligand platform, J. Am. Chem. Soc. 139 [90] H.T. Poh, B.T. Sim, T.S. Chwee, W.K. Leong, W.Y. Fan, The dithiolate-bridged
(2017) 9168–9177, https://doi.org/10.1021/jacs.7b00429. diiron hexacarbonyl complex Na2[(l-SCH2CH2COO)Fe(CO)3]2 as a water-
[66] E. Andris, R. Navrátil, J. Jašík, G. Sabenya, M. Costas, M. Srnec, J. Roithová, soluble PhotoCORM, Organometallics 33 (2014) 959–963, https://doi.org/
Detection of indistinct Fe–N stretching bands in iron(V) nitrides by 10.1021/om401013a.
photodissociation spectroscopy, Chem. – A Eur. J. (2018), https://doi.org/ [91] W.A. Thornley, T.E. Bitterwolf, Intramolecular C. H activation and
10.1002/chem.201705307. metallacycle aromaticity in the photochemistry of [FeFe]-hydrogenase
[67] T. Sjostrand, Endogenous formation of carbon monoxide in man, Nature 164 model compounds in low-temperature frozen matrices, Chem. – A Eur. J. 21
(1949) 580, https://doi.org/10.1038/164580a0. (2015) 18218–18229, https://doi.org/10.1002/chem.201503826.
[68] E.M. Sikorski, T. Hock, N. Hill-Kapturczak, A. Agarwal, The story so far: [92] X. Jiang, L. Long, H. Wang, L. Chen, X. Liu, Diiron hexacarbonyl complexes as
molecular regulation of the heme oxygenase-1 gene in renal injury, Am. J. potential CO-RMs: CO-releasing initiated by a substitution reaction with
Physiol. Physiol. 286 (2004) F425–F441, https://doi.org/10.1152/ cysteamine and structural correlation to the bridging linkage, Dalton Trans.
ajprenal.00297.2003. 43 (2014) 9968–9975, https://doi.org/10.1039/C3DT53620C.
[69] S.W. Ryter, J. Alam, A.M.K. Choi, Heme oxygenase-1/carbon monoxide: from [93] C.S. Jackson, S. Schmitt, Q.P. Dou, J.J. Kodanko, Synthesis, characterization, and
basic science to therapeutic applications, Physiol. Rev. 86 (2006) 583–650, reactivity of the stable iron carbonyl complex [Fe(CO)(N4Py)](ClO4)2:
https://doi.org/10.1152/physrev.00011.2005. photoactivated carbon monoxide release, growth inhibitory activity, and
[70] B. Widdop, Analysis of carbon monoxide, Ann. Clin. Biochem. 39 (2002) 378– peptide ligation, Inorg. Chem. 50 (2011) 5336–5338, https://doi.org/10.1021/
391, https://doi.org/10.1258/000456302760042146. ic200676s.
[71] Y. Gong, T. Zhang, H. Liu, Y. Zheng, N. Li, Q. Zhao, Y. Chen, B. Liu, Synthesis, [94] R.N. Perutz, B. Procacci, Photochemistry of transition metal hydrides, Chem.
toxicities and cell proliferation inhibition of CO-releasing molecules Rev. 116 (2016) 8506–8544, https://doi.org/10.1021/acs.chemrev.6b00204.
containing cobalt, Transit. Met. Chem. 40 (2015) 413–426, https://doi.org/ [95] W. Wang, T.B. Rauchfuss, L. Bertini, G. Zampella, Unsensitized photochemical
10.1007/s11243-015-9931-4. hydrogen production catalyzed by diiron hydrides, J. Am. Chem. Soc. 134
[72] S.H. Heinemann, T. Hoshi, M. Westerhausen, A. Schiller, Carbon monoxide - (2012) 4525–4528, https://doi.org/10.1021/ja211778j.
physiology, detection and controlled release, Chem. Commun. 50 (2014) [96] R.L. Sweany, Matrix photolysis of tetracarbonyldihydridoiron. Evidence for
3644–3660, https://doi.org/10.1039/C3CC49196J. oxidative addition of dihydrogen on tetracarbonyliron, J. Am. Chem. Soc. 103
(1981) 2410–2412, https://doi.org/10.1021/ja00399a047.
34 J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35

[97] I.E. Buys, L.D. Field, T.W. Hambley, A.E.D. McQueen, Photochemical reactions remarkable photocytotoxicity, Inorg. Chem. 54 (2015) 3748–3758, https://
of [cis-Fe(H)2(Me2PCH2CH2PMe2)2] with thiophenes: insertion into C-H and doi.org/10.1021/ic5027625.
C-S bonds, J. Chem. Soc. Chem. Commun. (1994) 557–558, https://doi.org/ [122] T. Sarkar, S. Banerjee, A. Hussain, Significant photocytotoxic effect of an iron
10.1039/C39940000557. (iii) complex of a Schiff base ligand derived from vitamin B6 and
[98] M.V. Baker, L.D. Field, Reaction of ethylene with a coordinatively unsaturated thiosemicarbazide in visible light, RSC Adv. 5 (2015) 29276–29284, https://
iron complex Fe(DEPE)2: sp2 carbon-hydrogen bond activation without prior doi.org/10.1039/C5RA04207K.
formation of a.pi.-complex, J. Am. Chem. Soc. 108 (1986) 7436–7438, https:// [123] U. Basu, I. Pant, I. Khan, A. Hussain, P. Kondaiah, A.R. Chakravarty, Iron(III)
doi.org/10.1021/ja00283a065. catecholates for cellular imaging and photocytotoxicity in red light, Chem. –
[99] L.D. Field, A.V. George, B.A. Messerle, Methane activation by an iron An Asian J. 9 (2014) 2494–2504, https://doi.org/10.1002/asia.201402207.
phosphine complex in liquid xenon solution, J. Chem. Soc., Chem. Commun. [124] M. Roy, T. Bhowmick, S. Ramakumar, M. Nethaji, A.R. Chakravarty, Double-
(1991) 1339–1341, https://doi.org/10.1039/C39910001339. strand DNA cleavage from photodecarboxylation of ([small mu ]-oxo)diiron
[100] In remembrance of Barnett Rosenberg, Dalton Trans. (2009) 10648–10650, (iii) l-histidine complex in visible light, Dalton Trans. (2008) 3542–3545,
https://doi.org/10.1039/B918993A. https://doi.org/10.1039/B802533A.
[101] N.J. Wheate, S. Walker, G.E. Craig, R. Oun, The status of platinum anticancer [125] M. Roy, T. Bhowmick, R. Santhanagopal, S. Ramakumar, A.R. Chakravarty,
drugs in the clinic and in clinical trials, Dalton Trans. 39 (2010) 8113–8127, Photo-induced double-strand DNA and site-specific protein cleavage activity
https://doi.org/10.1039/C0DT00292E. of l-histidine ([small mu ]-oxo)diiron(iii) complexes of heterocyclic bases,
[102] J. Chen, J. Stubbe, Bleomycins: towards better therapeutics, Nat. Rev. Cancer 5 Dalton Trans. (2009) 4671–4682, https://doi.org/10.1039/B901337G.
(2005) 102, https://doi.org/10.1038/nrc1547. [126] M. Roy, R. Santhanagopal, A.R. Chakravarty, DNA binding and oxidative DNA
[103] W.A. Wani, U. Baig, S. Shreaz, R.A. Shiekh, P.F. Iqbal, E. Jameel, A. Ahmad, S.H. cleavage activity of ([small mu ]-oxo)diiron(iii) complexes in visible light,
Mohd-Setapar, M. Mushtaque, L. Ting Hun, Recent advances in iron Dalton Trans. (2009) 1024–1033, https://doi.org/10.1039/B815215B.
complexes as potential anticancer agents, New J. Chem. 40 (2016) 1063– [127] S.B. Chanu, S. Banerjee, M. Roy, Potent anticancer activity of photo-activated
1090, https://doi.org/10.1039/C5NJ01449B. oxo-bridged diiron(III) complexes, Eur. J. Med. Chem. 125 (2017) 816–824,
[104] K. Woodburn, Chemical Aspects of Photodynamic Therapy By Raymond https://doi.org/10.1016/j.ejmech.2016.09.090.
Bonnett (University of London). Gordon and Breach Science Publishers: [128] T. Sarkar, R.J. Butcher, S. Banerjee, S. Mukherjee, A. Hussain, Visible light-
London and Newark. 2000. xi + 305 pp. $48.00. ISBN 90-5699-248-1, J. Am. induced cytotoxicity of a dinuclear iron(III) complex of curcumin with low-
Chem. Soc. 123 (2001) 3622. doi:10.1021/ja0048228. micromolar IC50 value in cancer cells, Inorg. Chim. Acta 439 (2016) 8–17,
[105] M. Ethirajan, Y. Chen, P. Joshi, R.K. Pandey, The role of porphyrin chemistry in https://doi.org/10.1016/j.ica.2015.09.026.
tumor imaging and photodynamic therapy, Chem. Soc. Rev. 40 (2011) 340– [129] A. Fürstner, Iron catalysis in organic synthesis: a critical assessment of what
362, https://doi.org/10.1039/B915149B. it takes to make this base metal a multitasking champion, ACS Cent. Sci. 2
[106] I. Ott, R. Gust, Non platinum metal complexes as anti-cancer drugs, Arch. (2016) 778–789, https://doi.org/10.1021/acscentsci.6b00272.
Pharm. (Weinheim) 340 (2007) 117–126, https://doi.org/10.1002/ [130] W. Nam, Synthetic mononuclear nonheme iron-oxygen intermediates, Acc.
ardp.200600151. Chem. Res. 48 (2015) 2415–2423, https://doi.org/10.1021/acs.
[107] P. Koepf-Maier, H. Koepf, Non-platinum group metal antitumor agents. accounts.5b00218.
History, current status, and perspectives, Chem. Rev. 87 (1987) 1137–1152, [131] M.J. Collins, K. Ray, L. Que, Electrochemical generation of a nonheme oxoiron
https://doi.org/10.1021/cr00081a012. (IV) complex, Inorg. Chem. 45 (2006) 8009–8011, https://doi.org/10.1021/
[108] M.J. Clarke, F. Zhu, D.R. Frasca, Non-platinum chemotherapeutic ic061263i.
metallopharmaceuticals, Chem. Rev. 99 (1999) 2511–2534, https://doi.org/ [132] H. Kotani, T. Suenobu, Y.-M. Lee, W. Nam, S. Fukuzumi, Photocatalytic
10.1021/cr9804238. generation of a non-heme oxoiron(IV) complex with water as an oxygen
[109] Q. Li, W.R. Browne, G. Roelfes, Photoenhanced oxidative DNA cleavage with source, J. Am. Chem. Soc. 133 (2011) 3249–3251.
non-heme iron(II) complexes, Inorg. Chem. 49 (2010) 11009–11017, https:// [133] D.W. Low, J.R. Winkler, H.B. Gray, Photoinduced oxidation of
doi.org/10.1021/ic1014785. microperoxidase-8: generation of ferryl and cation-radical porphyrins, J.
[110] Q. Li, W.R. Browne, G. Roelfes, DNA cleavage activity of Fe(II)N4Py under Am. Chem. Soc. 118 (1996) 117–120, https://doi.org/10.1021/ja9530477.
photo irradiation in the presence of 1,8-naphthalimide and 9-aminoacridine: [134] J. Berglund, T. Pascher, J.R. Winkler, H.B. Gray, Photoinduced oxidation of
unexpected effects of reactive oxygen species scavengers, Inorg. Chem. 50 horseradish peroxidase, J. Am. Chem. Soc. 119 (1997) 2464–2469, https://doi.
(2011) 8318–8325, https://doi.org/10.1021/ic2008478. org/10.1021/ja961026m.
[111] Q. Li, M.G.P. van der Wijst, H.G. Kazemier, M.G. Rots, G. Roelfes, Efficient [135] M.E. Ener, Y.-T. Lee, J.R. Winkler, H.B. Gray, L. Cheruzel, Photooxidation of
nuclear DNA cleavage in human cancer cells by synthetic bleomycin mimics, cytochrome P450-BM3, Proc. Natl. Acad. Sci. 107 (2010), 18783 LP-18786.
ACS Chem. Biol. 9 (2014) 1044–1051, https://doi.org/10.1021/cb500057n. http://www.pnas.org/content/107/44/18783.abstract.
[112] A. Garai, U. Basu, I.L.A. Pant, P. Kondaiah, A.R. Chakravarty, Polypyridyl iron [136] C.E. Immoos, A.J. Di Bilio, M.S. Cohen, W. Van der Veer, H.B. Gray, P.J. Farmer,
(II) complexes showing remarkable photocytotoxicity in visible light, J. Chem. Electron-transfer chemistry of Ru–Linker–(Heme)-modified myoglobin:
Sci. 127 (2015) 609–618, https://doi.org/10.1007/s12039-015-0815-0. rapid intraprotein reduction of a photogenerated porphyrin cation radical,
[113] L. Tabrizi, Novel cyclometalated Fe(II) complex with NCN pincer and BODIPY- Inorg. Chem. 43 (2004) 3593–3596, https://doi.org/10.1021/ic049741h.
appended 40 -Ethynyl-2,20 :60 ,200 -terpyridine as mitochondria-targeted [137] I. Hamachi, S. Tsukiji, S. Shinkai, S. Oishi, direct observation of the ferric-
photodynamic anticancer agents, Appl. Organomet. Chem. (2017) e4161, porphyrin cation radical as an intermediate in the phototriggered oxidation
https://doi.org/10.1002/aoc.4161. of ferric- to ferryl-heme tethered to Ru(bpy)3 in reconstituted myoglobin, J.
[114] M. Roy, B. Pathak, A.K. Patra, E.D. Jemmis, M. Nethaji, A.R. Chakravarty, New Am. Chem. Soc. 121 (1999) 5500–5506, https://doi.org/10.1021/ja984199f.
insights into the visible-light-induced DNA cleavage activity of [138] N.-H. Tran, N. Huynh, T. Bui, Y. Nguyen, P. Huynh, M.E. Cooper, L.E. Cheruzel,
dipyridoquinoxaline complexes of bivalent 3d-metal ions, Inorg. Chem. 46 Light-initiated hydroxylation of lauric acid using hybrid P450 BM3 enzymes,
(2007) 11122–11132, https://doi.org/10.1021/ic701450a. Chem. Commun. 47 (2011) 11936–11938, https://doi.org/10.1039/
[115] M.S. Shongwe, C.H. Kaschula, M.S. Adsetts, E.W. Ainscough, A.M. Brodie, M.J. C1CC15124J.
Morris, A phenolate-induced trans influence: crystallographic evidence for [139] N.-H. Tran, D. Nguyen, S. Dwaraknath, S. Mahadevan, G. Chavez, A. Nguyen, T.
unusual asymmetric coordination of an a-diimine in ternary complexes of iron Dao, S. Mullen, T.-A. Nguyen, L.E. Cheruzel, An efficient light-driven P450
(III) possessing biologically relevant hetero-donor N-centered tripodal ligands, BM3 biocatalyst, J. Am. Chem. Soc. 135 (2013) 14484–14487, https://doi.org/
Inorg. Chem. 44 (2005) 3070–3079, https://doi.org/10.1021/ic048835o. 10.1021/ja409337v.
[116] S. Saha, R. Majumdar, M. Roy, R.R. Dighe, A.R. Chakravarty, An iron complex of [140] M. Kato, Q. Lam, M. Bhandarkar, T. Banh, J. Heredia, A.U.L. Cheruzel, Selective
dipyridophenazine as a potent photocytotoxic agent in visible light, Inorg. C–H bond functionalization with light-driven P450 biocatalysts, Comptes
Chem. 48 (2009) 2652–2663, https://doi.org/10.1021/ic8022612. Rendus Chim. 20 (2017) 237–242, https://doi.org/10.1016/j.crci.2015.10.005.
[117] S. Saha, D. Mallick, R. Majumdar, M. Roy, R.R. Dighe, E.D. Jemmis, A.R. [141] S. Fukuzumi, W. Nam, Thermal and photoinduced electron-transfer catalysis
Chakravarty, Structure–activity relationship of photocytotoxic iron(III) of high-valent metal-oxo porphyrins in oxidation of substrates, J. Porphyr.
complexes of modified dipyridophenazine ligands, Inorg. Chem. 50 (2011) Phthalocyan. 20 (2016) 35–44, https://doi.org/10.1142/S1088424616300032.
2975–2987, https://doi.org/10.1021/ic1024229. [142] M. Ghosh, K.K. Singh, C. Panda, A. Weitz, M.P. Hendrich, T.J. Collins, B.B. Dhar,
[118] S. Saha, R. Majumdar, A. Hussain, R.R. Dighe, A.R. Chakravarty, Biotin- S. Sen Gupta, Formation of a Room Temperature Stable Fe V (O) Complex :
conjugated tumour-targeting photocytotoxic iron(III) complexes, Philos. Reactivity, 2014, 40–43.
Trans. R. Soc. A Math. Phys. Eng. Sci. 371 (2013), https://doi.org/10.1098/ [143] B. Chandra, K.K. Singh, S. Sen, Gupta, Selective photocatalytic hydroxylation
rsta.2012.0190. and epoxidation reactions by an iron complex using water as the oxygen
[119] A. Garai, I. Pant, P. Kondaiah, A.R. Chakravarty, Iron(III) salicylates of source, Chem. Sci. 8 (2017) 7545–7551, https://doi.org/10.1039/C7SC02780J.
dipicolylamine bases showing photo-induced anticancer activity and [144] C. Herrero, A. Quaranta, M. Sircoglou, K. Senechal-David, A. Baron, I.M. Marin,
cytosolic localization, Polyhedron 102 (2015) 668–676, https://doi.org/ C. Buron, J.-P. Baltaze, W. Leibl, A. Aukauloo, F. Banse, Successive light-
10.1016/j.poly.2015.10.026. induced two electron transfers in a Ru-Fe supramolecular assembly: from Ru-
[120] U. Basu, I. Khan, A. Hussain, B. Gole, P. Kondaiah, A.R. Chakravarty, Fe(ii)-OH2 to Ru-Fe(iv)-oxo, Chem. Sci. 6 (2015) 2323–2327, https://doi.org/
Carbohydrate-appended tumor targeting iron(III) complexes showing 10.1039/C5SC00024F.
photocytotoxicity in red light, Inorg. Chem. 53 (2014) 2152–2162, https:// [145] F. Avenier, C. Herrero, W. Leibl, A. Desbois, R. Guillot, J. Mahy, A. Aukauloo,
doi.org/10.1021/ic4028173. Photoassisted generation of a dinuclear iron(III) peroxo species and oxygen-
[121] U. Basu, I. Pant, A. Hussain, P. Kondaiah, A.R. Chakravarty, Iron(III) complexes atom transfer, Angew. Chem. Int. Ed. 52 (2013) 3634–3637, https://doi.org/
of a pyridoxal schiff base for enhanced cellular uptake with selectivity and 10.1002/anie.201210020.
J. Chen, W.R. Browne / Coordination Chemistry Reviews 374 (2018) 15–35 35

[146] R. Zhang, R.E.P. Chandrasena, E. Martinez, J.H. Horner, M. Newcomb, Inorg. Chim. Acta 363 (2010) 2791–2799, https://doi.org/10.1016/j.
Formation of compound I by photo-oxidation of compound II, Org. Lett. 7 ica.2010.03.005.
(2005) 1193–1195, https://doi.org/10.1021/ol050296j. [156] I.M. Wasser, H.C. Fry, P.G. Hoertz, G.J. Meyer, K.D. Karlin, Photochemical
[147] D.N. Harischandra, R. Zhang, M. Newcomb, Photochemical generation of a organic oxidations and dechlorinations with a l-oxo bridged heme/non-
highly reactive iron–oxo intermediate. A true iron(V)–Oxo species?, J Am. heme diiron complex, Inorg. Chem. 43 (2004) 8272–8281, https://doi.org/
Chem. Soc. 127 (2005) 13776–13777, https://doi.org/10.1021/ja0542439. 10.1021/ic0490932.
[148] R.M. Richman, M.W. Peterson, Photodisproportionation of.mu.-oxo-bis [157] D.N. Harischandra, G. Lowery, R. Zhang, M. Newcomb, Production of a
[(tetraphenylporphinato)iron(III)], J. Am. Chem. Soc. 104 (1982) 5795–5796, putative iron(V)–oxocorrole species by photo-disproportionation of a bis-
https://doi.org/10.1021/ja00385a045. corrole–diiron(IV)–l-oxo dimer: implication for a green oxidation catalyst,
[149] M.W. Peterson, D.S. Rivers, R.M. Richman, Mechanistic considerations in the Org. Lett. 11 (2009) 2089–2092, https://doi.org/10.1021/ol900480p.
photodisproportionation of.mu.-oxo-bis((tetraphenylporphinato)iron(III)), J. [158] A. Draksharapu, Q. Li, G. Roelfes, W.R. Browne, Photo-induced oxidation of
Am. Chem. Soc. 107 (1985) 2907–2915, https://doi.org/10.1021/ja00296a013. [FeII(N4Py)CH3CN] and related complexes, Dalton Trans. 41 (2012) 13180–
[150] M.W. Peterson, R.M. Richman, Photodisproportionation of (.mu.-oxo)bis 13190, https://doi.org/10.1039/c2dt30392b.
((tetrakis(4-carboxyphenyl)porphinato)iron(III), Inorg. Chem. 24 (1985) [159] G. Roelfes, M. Lubben, R. Hage, J. Que, B.L Feringa Lawrence, Catalytic
722–725, https://doi.org/10.1021/ic00199a018. oxidation with a non-heme iron complex that generates a low-spin FeIIIOOH
[151] B.J. Pistorio, C.J. Chang, D.G. Nocera, A phototriggered molecular spring for intermediate, Chem. – A Eur. J. 6 (2000) 2152–2159, https://doi.org/10.1002/
aerobic catalytic oxidation reactions, J. Am. Chem. Soc. 124 (2002) 7884– 1521-3765(20000616)6:12<2152::AID-CHEM2152>3.0.CO;2-O.
7885, https://doi.org/10.1021/ja026017u. [160] S.L. Esarey, J.C. Holland, B.M. Bartlett, Determining the fate of a non-heme
[152] J.M. Hodgkiss, C.J. Chang, B.J. Pistorio, D.G. Nocera, Transient absorption iron oxidation catalyst under illumination, oxygen, and acid, Inorg. Chem. 55
studies of the pacman effect in spring-loaded diiron(III) l-Oxo bisporphyrins, (2016) 11040–11049, https://doi.org/10.1021/acs.inorgchem.6b01538.
Inorg. Chem. 42 (2003) 8270–8277, https://doi.org/10.1021/ic034751o. [161] J. Chen, A. Draksharapu, E. Harvey, W. Rasheed, L. Que, W.R. Browne, Direct
[153] J. Rosenthal, B.J. Pistorio, L.L. Chng, D.G. Nocera, Aerobic catalytic photochemical activation of non-heme Fe(IV)O complexes, Chem. Commun.
photooxidation of olefins by an electron-deficient pacman bisiron(III) l-oxo 53 (2017) 12357–12360, https://doi.org/10.1039/c7cc07452b.
porphyrin, J. Org. Chem. 70 (2005) 1885–1888, https://doi.org/10.1021/ [162] J. Chen, S. Stepanovic, A. Draksharapu, M. Gruden, W.R. Browne, A non-heme
jo048570v. iron photocatalyst for light-driven aerobic oxidation of methanol, Angew.
[154] J. Rosenthal, T.D. Luckett, J.M. Hodgkiss, D.G. Nocera, Photocatalytic oxidation Chem. Int. Ed. 57 (2018) 3207–3211, https://doi.org/10.1002/
of hydrocarbons by a bis-iron(III)-l-oxo pacman porphyrin using O2 and anie.201712678.
visible light, J. Am. Chem. Soc. 128 (2006) 6546–6547, https://doi.org/ [163] J. Chen, D. Unjaroen, S. Stepanovic, A. van Dam, M. Gruden, W.R. Browne,
10.1021/ja058731s. Selective photo-induced oxidation with O 2 of a non-heme iron(III) complex
[155] S.K. Ghosh, R. Patra, S.P. Rath, Synthesis, structure and photocatalytic activity to a bis(imine-pyridyl)iron(II) complex, Inorg. Chem. 57 (2018) 4510–4515,
of a remarkably bent, cofacial ethene-linked diiron (III) l-oxobisporphyrin, https://doi.org/10.1021/acs.inorgchem.8b00187.

You might also like