You are on page 1of 57

pubs.acs.

org/CR Review

Second Sphere Effects on Oxygen Reduction and Peroxide


Activation by Mononuclear Iron Porphyrins and Related Systems
Sarmistha Bhunia,† Arnab Ghatak,† and Abhishek Dey*

Cite This: Chem. Rev. 2022, 122, 12370−12426 Read Online

ACCESS Metrics & More Article Recommendations


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Activation and reduction of O2 and H2O2 by synthetic and biosynthetic iron
porphyrin models have proved to be a versatile platform for evaluating second-sphere effects
Downloaded via CARNEGIE MELLON UNIV on May 31, 2023 at 14:50:20 (UTC).

deemed important in naturally occurring heme active sites. Advances in synthetic techniques
have made it possible to install different functional groups around the porphyrin ligand,
recreating artificial analogues of the proximal and distal sites encountered in the heme
proteins. Using judicious choices of these substituents, several of the elegant second-sphere
effects that are proposed to be important in the reactivity of key heme proteins have been
evaluated under controlled environments, adding fundamental insight into the roles played by these weak interactions in nature. This
review presents a detailed description of these efforts and how these have not only demystified these second-sphere effects but also
how the knowledge obtained resulted in functional mimics of these heme enzymes.

CONTENTS Acknowledgments 12417


Abbreviations 12417
1. Introduction 12370 References 12417
1.1. Heme Proteins in Nature 12371
1.2. Definition of Second Sphere 12372
2. Role of Second Sphere in Synthetic Models of
CcO 12372 1. INTRODUCTION
2.1. Role of the Distal Cu Center 12372 Reduction of O2 to H2O during respiration provides the energy
2.2. Role of the Distal Phenol Group 12384 needed to sustain all aerobic life forms� from unicellular
3. Mononuclear Metallo-porphyrins as ORR Cata- bacteria to higher mammals.1,2 The solar energy stored in O2
lysts 12386 during photosynthesis is utilized to generate ATP, the energy
3.1. Role of Hydrophobicity and Steric Effect in carrier molecule, in the mitochondria. Thus, apart from being a
the Distal Pocket 12387 process fundamentally important to life on earth, the process
3.2. Role of Distal Water-Mediated H-Bonding 12389 has direct implication to the chemical storage of energy.
3.3. Role of Distal Redox Active Center 12391 Conventionally, O2 has served as oxidant in most abundant
3.4. Role of Distal H-Bonding Interactions 12395 combustion processes that release energy.3−5 The release of
3.5. Role of Electrostatic Interactions 12403 CO2 and other unamicable gases in the atmosphere associated
4. Biosynthetic Models of CcO 12405 with combustion of carbon-based fuels has led to an emprise
4.1. Role of Distal Cu 12405 for cleaner energy sources where sustainable and noncarbona-
4.2. Role of Tyrosine Residue 12406 ceous fuels like H2, which can be oxidized at the anode and
4.3. Electrocatalytic ORR by Biosynthetic Models utilized for O2 reduction at the cathode in a fuel-cell to
of CcO 12408 generate electricity, is being hailed as a potent replace-
4.4. Role of Distal Arginine Residue in Amylin ment.6−11 Consequently, the understanding of this fundamen-
Peptide 12410 tally important 4e−/4H+ reduction of O2 to H2O has now
5. Role of Second-Sphere Distal Residues in generated considerable interest in the research community as it
Peroxide Activation 12411 may aid in access to clean and sustainable energy.12−15
6. Conclusion 12416
Author Information 12416
Special Issue: Catalysis beyond the First Coordination
Corresponding Author 12416
Sphere
Authors 12416
Author Contributions 12417 Received: December 9, 2021
Notes 12417 Published: April 11, 2022
Biographies 12417

© 2022 American Chemical Society https://doi.org/10.1021/acs.chemrev.1c01021


12370 Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 1. Active site structures of (A) myoglobin (pdb id: 1mbo),39 (B) cytochrome P450 (pdb id: 2a1m),40 (C) horse radish peroxidase (pdb id:
1hh5),41 cytochrome c oxidase (pdb id: 2y69),42 Sir (pdb id: 3 mm5),43 and Nir (pdb id: 3d1i).48

To begin conceiving or designing a catalyst for O2 reduction while the axial thiolate ligand is key to the monooxygenase
a logical place to start is the active sites of metallo-enzymes activity, the threonine residue in the distal site (Figure 1B)
that are known to do it in nature.16 These would be either the transfers the protons needed to cleave the O−O bond to
heme-based cytochrome c oxidase or copper-based multi- generate the reactive oxidant such that the mutation of this
copper oxidases.17−21 There has been a substantial volume of residue results in loss of monooxygenase activity.38,52 The
work in the development of O2 reduction catalysts in the past distal site of horseradish peroxidases (HRP) has an arginine
several decades, which include different cathode materials, residue and a histidine residue (Figure 1C) which are
molecular catalysts, often adsorbed on cathodes, and enzymatic responsible for the efficient binding of the oxidant H2O2 and
catalysts.15,22−26 Many of these, at least the molecular ones, the substrates and heterolytic cleavage of the O−O bond to
were often prescribed following the design of the active sites of generate the highly oxidizing compound I species. Mutation of
these natural metalloenzymes. Such efforts have been reviewed either of the two species is deleterious to the function of
extensively very recently.10,27−31 In this paper, the intension is peroxidases.47,53−55 The distal site of cytochrome c oxidase
to review the recent developments in understanding the effect (CcO) contains a CuB (Figure 1D) which is held by three
of weaker, often noncovalent, interactions that are known to histidine residues, one of which is post-translationally cross-
play a major role in controlling the reactivity of heme proteins linked with a tyrosine residue. The CuB and tyrosine residues
and could be integrated in the design of artificial catalysts to are indispensable for the 4e−/4H+ reduction of O2 catalyzed by
enhance their reactivity. this active site. Additionally, the coupling between the O−O
1.1. Heme Proteins in Nature bond cleavage and proton transfer from the tyrosine residue is
One of the astounding features of the heme-based proteins is likely responsible for creating vectorial proton transfer during
the ability to catalyze diverse sets of chemical reactions with O2 reduction which is responsible for generating the proton
apparently small modifications of the active site�such gradient across the mitochondrial membrane to be later
diversity is perhaps only matched by iron−sulfur proteins.32−38 utilized for the synthesis of ATP from ADP and phosphate
The heme active sites vary in the nature of the heme cofactor, during oxidative phosphorylation.56−59
in the ligation to the protein, and in the environment around The organization of the protein residues around the
the heme active sites.39−43 While the different porphyrin metalloporphyrin active site in sulfite reductases is very
macrocycles and axial ligand to the iron are ostentatious impressive to say the least. This siroheme active site catalyzes
changes in the active site and may be expected to affect the the 6e−/6H+ reduction of SO2/SO32− to H2S and plays a major
chemical reactivity of the active site, the environment around role in the geochemical sulfur cycle.60 The roles of the distal
the heme is a major enabler of the reactivity of these active lysine and arginine residues (Figure 1E) include binding and
sites and is the inspiration behind the topic of this review.44−47 orientation of the substrate and transfer of protons to the
In the water-soluble O2 binding/storage proteins like catalytic site.61 In addition to the distal residues there is a Fe4S4
myoglobin, the distal environment, the active site in the cubane bridged to the cluster which provides electrons to the
vicinity of the exchangeable axial ligand, contains hydrophobic active site, which may achieve the transformation in three 2e−
residues and a histidine residue (Figure 1A) which decelerate steps.61−65 The same siroheme active site is found in
the hydrolysis of the FeIII−O2− adduct by both limiting access assimilatory nitrite reductases which reduce NO2− to
of water and hydrogen bonding to the bent Fe−O2 unit.49−51 NH4+.66 The cytochrome cd1 nitrite reductase, on the other
In the active site of cytochrome P450 superfamily of enzymes, hand, does not possess the Fe4S4 cluster and uses heme d1
12371 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 2. Active site of sulfite reductase (SiR) used to demonstrate the second sphere interactions: hydrogen bonding (gray lines), electrostatic
potential (columbic charge surfaces; generated using Chimera), and a second Fe4S4 cluster.

cofactor and can reduce NO2− to NO and not further.67 The lysine have pKa > 9.60,84,85 Some of these directly hydrogen
active site of Cd1 NiR contains arginine and histidine residues bond with the substrate (shown with gray lines, Figure 2),
(Figure 1F) which are believed to serve the same purpose as while some, that are further away, create a positive electrostatic
the ones present in the active site of SiRs.68−71 field (blue surfaces, Figure 2) in the vicinity of the substrate;
1.2. Definition of Second Sphere both of these effects will aid binding, orientation, and
reduction of the bound substrate.62 At the same time, the
A metallo-protein active site design can thus be divided into Fe4S4 cubane, albeit bridged, stores electrons and shuttles it to
two distinct components, the first being the immediate co- the heme cofactor during catalysis.85−87 These components of
ordination sphere of the metal. In the case of heme proteins the metallo-enzyme active sites can broadly be classified as the
this will include the porphyrin ring that forms the organic part second sphere, and their effect on the reactivity of the active
of the heme cofactor and the axial ligand used to bind the site can be summarized as second-sphere effects on catalysis.88
heme cofactor to the proteins.72,73 There are several different Second-sphere effects comprise effects induced by amino
varieties of the porphyrin ring known to exist in different heme acids such as hydrophobicity, hydrogen bonding, and electro-
active sites.74,75 Similarly, axial ligands reported to date for statics as well as the effect of introducing another redox active
heme include histidine (imidazole headgroup), cysteine metal in the vicinity of the iron porphyrin active site. Over the
(thiolate headgroup), tyrosine (phenolate headgroup), me- past decade or so, bioinspired modeling of heme active sites
thionine (thioether headgroup), and lysine (primary amine has advanced to include these second-sphere effects on
headgroup).76−78 Variations in the primary co-ordination spectroscopy and reactivity of artificial analogues. A
sphere affect the electronic structure of the heme active site comprehensive review of such efforts toward the understanding
which leads to very discernible changes in spectroscopic and development of ORR is presented in this paper.
features of heme proteins and plays a dominant role in
stabilizing, and thereby determining, the nature of the key 2. ROLE OF SECOND SPHERE IN SYNTHETIC MODELS
reactive intermediate in these proteins.79−83 OF CCO
The second component to the active-site design is the
noncoordinated amino acids which may or may not affect the 2.1. Role of the Distal Cu Center
electronic structure of the heme active site in its resting state The role of a distal Cu ion in the reduction of O2 by iron
but manifest themselves in the reactivity. The distal site of the porphyrins has been extensively investigated by several groups
SiR protein, for example, has several amino acids like arginine with the aim of generating functional mimics of naturally
and lysine which are likely to be protonated under occurring enzyme CcO.29,35 CcO, the terminal enzyme of the
physiological conditions as their head groups guanidine and human respiratory chain, catalyzes the 4e−/4H+ reduction of
12372 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 3. Synthetic CcO model complexes (A−E) reported by Collman and co-workers.109−111

dioxygen to water in mitochondria.89,90 A proton gradient independently addressed and either a constant or sweeping
along the mitochondrial membrane resulting from this process potential can be applied.106−108
is used to synthesize ATP.4,5,91 This process of oxidative Collman and co-workers synthesized some earlier model
phosphorylation in mitochondria stores the energy obtained by complexes of CcO such as a cobalt(II) porphyrin which
O2 reduction by NADH/NADPH released from different consists of a copper triazacyclononane macrocycle fastened to
metabolic pathways in a eukaryotic cell.92−94 The exact the distal side along with a linker bearing an imidazole
mechanism of the binuclear hemea3-CuB enzyme active site headgroup at the proximal side (Figure 3A). This model could
that performs the O2 reduction without any leakage of partially bind O2 irreversibly in 1:1 ratio to form a stable adduct. This
reduced oxygen species (PROS) like peroxide and superoxide, oxygen-derived adduct is a bridging peroxide which was
which cause oxidative stress, has been a fundamentally characterized by electrospray ionization mass (ESI-MS)
important question.95,96 Apart from biochemists investigating spectra and IR spectra [ν(O−O) at 804 cm−1, Δ16/18O2 =
the membrane bound protein itself, synthetic inorganic 46 cm−1], and this adduct regenerated the deoxygenated form
chemists have forwarded some noteworthy artificial mimics when titrated with 4 equiv of a reducing agent like cobaltocene,
that allowed investigating the inorganic active site outside of i.e., it catalyzes the 4e−/4H+ reduction. On the contrary, the
the protein environment.28,97,98 Cu free form of the complex 3A showed reversible O2 binding
The initial attempts focused on mimicking the O2 reduction but not O2 reduction. RRDE data of complex 3A at
activity electrochemically under heterogeneous conditions physiological pH showed 4e−/4H+ O2 reduction to H2O.
where the synthetic mimics were adsorbed on electrodes The selectivity is proposed to be derived from the presence of
Cu in the distal structure which aids the O−O bond
(generally graphitic) such that the electrons were directly
cleavage.28,109
provided from the electrode and protons and oxygen could be
The electrochemical behavior of two well-defined synthetic
obtained from aqueous solutions where these electrodes were
FeII−CuI model complexes of CcO (Figure 3B,C) were
dipped in.99,100 Rotating disc electrochemistry (RDE) and explored. These two complexes differ only in the nature of the
rotating ring disc electrochemistry (RRDE) are two convenient distal ligand structure to which the Cu ion binds. Compound
methods that have been extensively used to estimate the rate 3B has a 1,4,7-triazacyclononane (TACN) cap which can bind
and selectivity for the 4e−/4H+ oxygen reduction proc- the Cu ion with three nitrogen atoms, while complex 3C has a
ess.101,102 In RDE, the plot of 1/icat (icat is the catalytic N,N′,N”-tribenzyltris(aminomethyl)amine (TBTren) cap con-
current) and 1/ω1/2 (ω is the angular rotation velocity of the taining four nitrogen atoms. When both of the complexes were
electrode) at a potential in the region where the catalytic deposited over the edge-plane graphite (EPG) electrode and
current is mass transfer limited (i.e., current independent of RRDE experiments were performed, 3B was found to catalyze
potential) yields a straight line where the Y-intercept yields the 2e− reduction of O2 to generate H2O2, while 3C catalyzed 4e−
rate of O 2 reduction (kcat ) at infinite ω where the O2 reduction at pH 7. The difference in selectivity for O2
concentration of the substrate at the electrode surface is the reduction was ascribed to the different CuII/I potential induced
same as the concentration in bulk solution, while the slope by the TACN cap in 3B and TBTren cap in 3C. The 3C
yields the number of electrons supplied to the substrate (n). complex with TBTren cap in the distal structure was proposed
Ideally, the n should be 4 for a 4e+/4H+ ORR and the highest to stabilize the side-on FeIII(O22−)-CuII peroxo intermediate,
value of kcat can be the diffusion rate of O2 or H+ in solution at and as a result, the conversion to H2O2 is endergonic and only
a given pH.103−105 In RRDE, the working electrode of RDE is H2O formation is exergonic, which favors the 4e− O2
encircled by a Pt ring secondary electrode that can be reduction. On the contrary, for the 3B complex, both H2O2
12373 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 4. Synthetic CcO model complexes (A−F) studied by Boitrel and co-workers.112−115

formation and H2O formation become exergonic. Thus, the Around the same time, Boitrel and co-workers synthesized a
selectivity of oxygen reduction depends on the rate of series of biomimetic models of CcO in order to exact the role
hydrolysis of the peroxo intermediate, which is less stable for of the distal Cu center and Cu ion binding environment in the
3B due to TACN cap in the distal structure, and hence, it electrocatalytic O2 reduction. The quinolinyl picket porphyr-
hydrolyses and releases H2O2, thereby catalyzing the 2e− O2 ins, containing external nitrogen base (4A) and tailed nitrogen
reduction. Notably, this was one of the earliest reports where base (4B) were specifically designed (1) to evaluate their
PROS formation during electrochemical O2 reduction was efficiency of electrocatalytic O2 reduction to water, (2)
attributed to hydrolysis of peroxide intermediates. The model introduce structural similarity with the natural enzyme by
complex 3C, which selectively reduces O 2 to H 2 O, incorporating nitrogen atoms that is part of aromatic system to
demonstrates the role of CuII/I potential in controlling the stabilize Cu and off-centring of two metal atoms which is often
stability of the peroxide intermediate and in turn controlling neglected, and (3) prevention of any intermolecular reaction
the fate of O2 reduction.10,110 on the distal side of the porphyrin between the copper atoms
Collman and co-workers, in their continued pursuit, of two such constructs (Figure 4). Both of the Fe−Cu-
synthesized even closer structural analogues of Fea3/CuB active containing porphyrins were found to bind oxygen irreversibly,
site of CcO where the Cu atom is coordinated by three while the Fe-only one (not shown) could bind O2 reversibly.
imidazole ligands.111 Both compound 3D and 3E contain three Surprisingly, when these compounds were physiadsorbed over
distal imidazole ligands tethered to the porphyrin core via the EPG electrode and RRDE experiments were performed at
pH 6.86, they found that the Fe only model complexes are
acetamide linkages, but they differ in the nature of the
better 4e− catalysts than the Fe−Cu analogues (4A and 4B),
covalently attached axial base. Molecular modeling suggested
which catalyze both the 2e− and 4e− reductions. Boitrel and
that due to the presence of acetamide linkages in both the
co-workers explained this deviation considering two possible
complexes the Cu···Fe distance was in the ∼4.5−5 Å range but scenarios: (1) the copper center does not interfere with the O2
does not coordinate with the iron atom. RRDE experiments, molecule bound to the iron during the reduction, it only
with these catalysts adsorbed on EPG electrode, revealed that interferes during O2 binding and (2) the labile copper ion of
both these complexes showed 4e− reduction of oxygen across a the tris(N-base) complex is no longer coordinated when the
pH range of 3.5−8.5. The complex 3E exhibited electro- catalyst adsorbed onto the graphite electrode is in contact with
catalytic reduction of O2 at very low overpotential (ηORR) the aqueous solution.10,112
compared to other reported analogues. At pH 7 buffer The same group had almost similar observations with their
solution, addition of anions like chloride, bromide, and porphyrins capped with a tris(2-aminoethyl)amine (tren)
carboxylate did not affect the O2 reduction, suggesting that it motif as functional models for CcO. The Fe−Cu porphyrin
is the water-derived ligands (e.g., aquo-hydroxo or hydroxo 4C was obtained by grafting the tren moiety on a
ligands) which determine the redox potential of the Fe/Cu pair tris(chloroacetamide) picket porphyrin, leading to a ligand in
during the catalytic cycle.28,111 In these series, the selectivity which the tren is held by three arms of the porphyrin and is
for 4e−/4H+ reduction of O2 was rationalized mostly through more off-centered and closer to the porphyrin than in model
the stability and O−O activation of FeIII−(O22−)−CuII 4D (where all four of the porphyrin arms hold the tren ligand).
intermediates, although its presence was never established The Cu free forms of 4C and 4D were found to be more
directly. selective toward 4e−/4H+ O2 reduction as well when the
12374 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 5. CcO active site structure (A). Synthetic CcO model complexes (B−D) studied by Collman and co-workers.125,126

RRDE experiment was performed in a pH 6.6 buffer solution decades later by the Dey’s group and is discussed in the later
over EPG electrode. These observations could be related to the sections of the manuscript.
fact that the μ-peroxo intermediate between iron and copper In native CcO enzyme, the electrons required for O2
atoms (FeIII(O22−)−CuII) might not be a prerequisite for O2 reduction arrive via cytochrome C (cyt C) at very slow rates
reduction, whereas the involvement of a protonated end-on (∼4 s−1). While the fully reduced CcO does not require
ferric peroxo unit was likely to be more compatible with the electron transfer (ET) from Cyt C as the ET from reduced
reactivity. These results seem to be in contradiction with that CuA and heme b are rapid (104−105 s−1), Wickstrom and
of Collman’s results where the nature of the FeIII(O22−)−CuII others had postulated that during turnover the mixed valent
species was deemed crucial. In these cases, the ferric state where the heme−Cu site is reduced, and not the CuA and
hydroperoxide complexes might have been stabilized through heme b, is more likely to be catalytically active.1,57,116,117 This
H-bonding from the secondary amino groups present in the was supported by the isolation of a oxoferryl-tyr• radical
distal structure, which likely remain protonated at pH 6.6 in intermediate PM (FeIV�O, CuII OH/H2O, and Tyr•) where
aqueous solution.113 out of the 4e− needed for O2 reduction three are obtained from
In subsequent work, they showed that the substituents on the heme−Cu site and one is obtained from the Tyr244
residue.56 The X-ray crystallographic structure of the CcO
the secondary amino groups significantly influence the
active site is shown in Figure 5A for clarification (it includes a
electrocatalytic O2 reduction activity over the electrode
binuclear site composed of hemea3 with a proximal histidine
surface. The alkylated secondary amine in the distal structure
along with a histidine bound copper (CuB) in the distal site).
of 4C with −CO2Et and benzyl groups, respectively (4E and One of these three ligating histidine residues is covalently
4F), showed higher ηORR in reducing O2 compared to 4C. In cross-linked to a nearby tyrosine residue which acts as a source
particular, the alkylation by benzyl group increased the ηORR by of electron and proton in the reduction of O2 and aids vectorial
0.5 V. In addition, the Fe-only forms of 4E and 4F showed proton translocation.57,118,119 To address this, Collman et al.
partial O2 reduction reaction, while 4C reduces O2 selectively synthesized a CcO model that mimic both CuB and Tyr244 at
via the 4e− pathway only. The authors concluded that the the distal site. To mimic the slow ET rate of Cyt C, self-
secondary amines in the distal structures were much more assembled monolayer (SAM) on Au electrodes were
efficient than tertiary amines in catalyzing the selective 4e−/ used.120−122 The electron tunnelling rates across SAM
4H+ reduction of O2. The results from Boitrel’s work suggested exponentially decays with its length and ET rates across
that a mononuclear iron porphyrin adsorbed on an electrode is SAM can be attenuated between 105 and 10 s−1 by controlling
an intrinsically efficient catalyst for the reduction provided the the chain length.123,124 Immobilization of CcO models using
rate of flow of electrons and protons are very rapid, which is CuI-catalyzed azide−alkyne cycloaddition (click chemistry)
exactly the case over EPG electrode in aqueous neutral pH atop SAM allowed control of the electron flux, which enabled
buffer solution.114,115 This conclusion was vindicated two understanding the exact role of redox active centers under
12375 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

varying the rates of electron delivery. In the new structural and Scheme 1. (A) Schematic Representation of Two Different
functional CcO model complex synthesized, a proximal SAM (Slow SAM S1 and Fast SAM S2) Modified Au
imidazole, a distal copper bound to a tris-imidazole ligand, a Electrode Used by Collman and Co-workers for Variation in
phenol group covalently attached to one of the copper-ligating Electron Flux, Where the Model Complex Was Covalently
imidazoles and an alkyne functional group for immobilization Attached. (B) PROS Bar Diagram of the Three CcO Model
were assembled in what can only be described as one of the Complexes over Two Different SAM-Modified Au
most elaborate, albeit arduous, feats of synthetic inorganic Electrode125
chemistry in the known literature (Figure 5B).125 Distal copper
was installed over the heme plane in such a way that the Fe···
Cu distance remained 5 Å, which is very close to the natural
mammalian CcO enzyme, and the covalently attached phenol
group which mimics the tyrosine residue of CcO, was
introduced close to the O2 binding site of Fe porphyrin. The
proximal imidazole ligand was covalently introduced as a trans
axial ligand to Fe, mimicking His376 of the natural enzyme and
the purpose of introducing terminal alkyne was to selectively
attach these systems over azide terminated SAM solution
through azide−alkyne cycloaddition. The control model 5C
lacks the external phenol which is masked with a −OMe group
and their electrochemical behavior was also compared with the
iron only model 22A (discussed in section 3.2) which lacks
both Cu and Tyr244.
The role of the phenol groups introduced in O2 reduction
was evaluated by comparing the selectivity for O2 reduction
using RRDE to detect any PROS released under slow ET
conditions. Model complexes were covalently attached onto
mixed SAM combining of 1-azidohexadecanethiol and
hexadecanethiol (slow SAM S1) where the ET from electrode
is rate limiting having rate constant kET = 6.0 ± 0.1 s−1 and, for
comparison, a fast SAM S2, comprising azidophenylene-
ethynylenebenzylthiol and octanethiol, was also used which
has an ET rate constant (kET) as high as 104 s−1 (Scheme
1A).123,127 Although different SAMs do not change the
potential at which ORR takes place (0.3 V vs NHE in both
SAMs), the kET affects the kinetics of the catalysis, i.e., while in
slow SAM S1 ET is rate limiting and in fast SAM S2 O2
diffusion to the electrode is rate limiting. When RRDE was
performed over an S2-modified Au electrode, it was found that active site. As a result, more than 96% selectivity toward 4e−/
the Fe-only model 22A showed 89 ± 2% 4e−/4H+− ORR 4H+ O2 reduction was recorded over a slow SAM S1-modified
selectivity. While incorporation of Cu in 5C increased the electrode (Scheme 1B). Thus, these model complexes help to
selectivity to 97 ± 1%, incorporation of both Cu and phenol propose that both CuB and Tyr244 are required in the distal
(5B) did not alter the selectivity further (Scheme 1B). The site as electron donors for selective four-electron reduction of
situation on fast SAM S2 was similar to that of fully reduced O2 under rate-limiting ET conditions.125 The role of the
CcO in a single turnover experiment, and the involvement of phenol group alone could not be evaluated in these
distal redox-active tyrosine was not crucial as long as electrons investigations. Different groups have later investigated this in
could be readily delivered from the electrode which surrogates details, and these results are discussed later (sections 2.2 and
at heme b and CuA. Alternatively, the Slow SAM S1 might have 3.3).
better mimicked the physiological conditions, where CcO is Apart from the electrocatalytic investigations, stoichiometric
not normally at its fully reduced state during turnover due to reaction with O2 in organic solvents at low temperature was
very slow electron delivery from Cytochrome C. Over slow monitored using different spectroscopic techniques. The
SAM S1 modified Au electrode, complex 22A lacking both Cu heme/Cu complex 5D, which has a built-in histidine-tyrosine
and phenol produced ∼29% PROS in RRDE experiments and cross-link, reacted with O2 resulting in the formation of a heme
resulted in rapid degradation during O2 reduction. The superoxide species which was generally stable in organic
analogue which includes Cu only in absence of phenol, 5C solutions at low temperatures and could be characterized using
was more stable than 22A but still showed marked degree of resonance Raman (rR) and 1H nuclear magnetic resonance
PROS leakage. Complex 5C could store up to three electrons (1H NMR) spectroscopy. When the phenol group was
at the active site (two from iron and one from copper) and, included, the reaction with O2 proceeded via the heme−
thus, one additional electron is required from the electrode for superoxide intermediate to eventually result in a species which
the 4e− reduction of oxygen. Since the ET rate from the had electron paramagnetic resonance (EPR) spectroscopic
electrode to the catalyst was slow in SAM S1, PROS was being features like that of PM in CcO. It is important to note that this
generated before the electron reaches the active site. In case of series of heme/Cu complexes did not present a strong case of a
5B, an additional redox active phenol is present along with Cu FeIII−(O22−)−CuII intermediate in contrast to previous reports
in the distal site and four electrons could be supplied from the from this group. Rather, a direct 3e− reduction of the bound
12376 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 6. Structural representation of the TMPA−Cu complex and the porphyrin (A) used by Karlin and co-workers for stoichiometric reaction
with O2. The formation of the corresponding peroxo and oxo complexes are also shown in (B) and (C), respectively.131

Figure 7. (A) Tethered Cu−TMPA with Fe−TPP complex (A) and its constitutional isomer (C) used by Karlin and co-workers. The
corresponding peroxo complex after reaction of (A) with O2 is shown in (B).132

Fe−O2 species was proposed which is similar to the earlier In their attempts to mimic the O2 reduction process in CcO,
conclusions from Boitrel and co-workers.128 Recently, Solo- Karlin and co-workers utilized Cu-TMPA complex (where
mon et al. determined the geometric and electronic structural TMPA = tris(2-pyridylmethyl)amine) and FeII tetrakis (2,6-
contribution in heme Cu oxidase (HCO) active site for the difluorophenyl)porphyrin FeII(F8TPP) (Figure 6A).131 When
O−O bond cleavage and how this is coupled to proton O2 was made to react with the above two in a 1:1 ratio in
pumping by spectroscopically defining the PM intermediate.112 acetonitrile at −40 °C, heterobinuclear FeIII−(O22−)−CuII
In the previous literature reports although evidence for the peroxo complex (Figure 6B) was formed. The peroxo complex
Tyr• radical in PM intermediate had been extensively discussed, was not stable and underwent thermal degradation to produce
the presence of both CuII and Tyr• in HCOs was still elusive, the μ-oxo (FeIII−O−CuII) complex subsequently (6C). The
and this was the motivation for this particular work.129,130 formation of peroxo species was established using rR
12377 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 8. Oxygen-bound species (A−C) formed by the reaction of tridentate Cu ligand LMe2N with 6A and O2 as studied by Karlin and co-
workers.133

spectroscopy, where the stretching frequency, ν(O−O), was [Fe−(O2)···CuII−(O2−)]+. These two species were formed
observed at 808 cm−1 (ν16/18O2 = 46 cm−1). Solution magnetic independently within 1 ms, and the ratio of these two products
moment measurement and 1H NMR indicated that 6B has an depended on the reaction temperature and O2 concentration.
S = 2 ground state having spin only magnetic moment of 5.1 At low temperature, intermediate having λmax = 538 nm was
μB at −40 °C, which arose because of antiferromagnetic favored, and it was converted to the peroxo complex 7B with a
coupling between the (S = 5/2) FeIII and (S = 1/2) CuII first-order kinetics at −94 to −60 °C where ΔH⧧ = 37.4 ± 0.4
centers. Formation of a peroxo complex was followed using kJ mol−1 and ΔS⧧ = −28.7 ± 2.3 J mol−1K−1. Surprisingly,
stopped-flow kinetics in the temperature range of −94 to −75 while performing the same O2 reactivity with the constitutional
°C in acetone solvent. The data revealed that prior to the isomer of 7A, i.e., 7C, it was found that the resulting peroxo
formation of 6B (λmax = 556 nm) from 6A, a transient heme− complex (not shown) was diamagnetic as compared to the
superoxide species was generated having λmax= 537 nm, and paramagnetic 7B. Also, spectrophotometric titration revealed
this transformation was a first-order process with ΔH⧧ = 45 ± that stoichiometry of the reaction is 7C/O2 = 2:1 and the
1 kJ mol−1, ΔS⧧ = −19 ± 6 J mol−1K−1, and k = 0.007 s−1 at ν(O−O) of 809 cm−1 was higher compared to 7B. The
−90 °C. Formation of this transient ferric superoxo species contrasting nature of these two peroxo complexes might have
instead of the corresponding Cu(II)−superoxo species been due to the ligand architecture. Nevertheless, this study
indicated that heme was preferred for O2 binding. The peroxo showed that if a Cu unit can be tethered and appended
complex (t1/2= 1015 ± 20 s, at room temperature) finally properly in the distal site of an iron porphyrin moiety, a μ-
transformed to the μ-oxo complex 6C by releasing 0.5 equiv of peroxo species might be formed upon the reaction of the fully
oxygen.35,131 reduced complex with O229,35,132
The same group synthesized a binuclear complex 6L−FeCu Tridentate ligands for the Cu instead of tetradentate TMPA
(7A) in which a Cu−TMPA moiety was tethered to a ligands such as PY2, where PY2 = N,N-bis[2- (2-pyridyl)-
(TPP)Fe II−porphyrin derivative and reported their O 2 ethyl]amine, when mixed in 1:1 ratio with FeII(F8TPP), 6A, in
reactivity instead of separately reacting TMPA, FeII(F8TPP), the presence of dioxygen, resulted in a heme superoxo species
and O2 in a 1:1:1 ratio.132 When 7A reacted with O2 in 1:1 [(S)(F8TPP)Fe-(O2)] (8A, S = acetone or THF) which
ratio in acetonitrile solvent, the peroxo adduct (7B) (FeIII− subsequently reacted with the Cu to produce the bridging μ-
O22−−CuII) was formed which was characterized by spectro- peroxo species (Figure 8). However, unlike the tetradentate
scopic data and MALDI-TOF-MS, this time with much higher Cu chelate, in this case the peroxo species was unstable even at
stability having t1/2 = 60 min at room temperature (Figure 7). low temperature and transformed to an isolable and character-
The rR spectra revealed that ν(O−O) of 7B was at 787 cm−1 izable μ-oxo species [(F8TPP)FeIII−(O2−)−CuII(MePY2)]+
(ν16/18O2 = 43 cm−1). The 1H, 2H NMR and EPR spectra immediately. When a more electron-donating Cu binding
revealed that 7B was a paramagnetic high spin (Stotal = 2) LMe2N was used, a short-lived heme superoxo species
peroxo complex which was formed by the antiferromagnetic [(S)(F8TPP)Fe-(O2)] (8A) was formed further, reacting
coupling between high spin Fe (S = 5/2) and Cu (S = 1/2). with the CuI to form the bridging peroxo species [(F8TPP)-
Upon reaction of 7A in a 1:1 ratio with O2, a short-lived FeIII−(O22−)−CuII(LMe2N)]+(8B) having slightly more stability
intermediate was formed (λmax = 538 nm) apart from 7B (λmax at low temperature. The 2H and 1H NMR spectroscopic data
= 561 nm) which Karlin and co-workers postulated as a indicated that this species was also an antiferromagnetic
superoxo species [Fe−(O2)···CuI]+ and/or a bis-O2 adduct coupled S = 2 system, arising from high spin FeIII and CuII
12378 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 9. (A) PImHFeII complex with its O2 reactivity in absence of Cu (B) and in the presence of Cu (C−E) reported by Karlin and co-workers.134

centers. However, rR spectra showed two distinct peroxo cm −1 ((ν 16/18 O 2 = 61 cm −1 ) and ν(Fe−O) at 575
species (arising from different isomer or conformer) with cm−1(ν16/18O2 = 24 cm−1) and the reaction was reversible;
ν(O−O) vibrations at 767 and 752 cm−1 (ν16/18O2 = 41 and i.e., warming 9B at room temperature gave back 9A. They
45 cm−1), respectively, and these values were significantly obtained the fully reduced heterobinuclear heme copper
lower than those observed for tetradentate systems. A μ-η2:η2 complex [(PImH)FeIICuI]+(9C) at low temperature by addition
side-on bound heme−peroxo−copper structure with a bent of 1 equiv of [CuI(CH3CN)4](B(C6F5)4] to the reduced
Cu−O2−Fe core was proposed in contrast to the (μ-1,2) end- complex 9A. When dry O2 gas in THF was passed through 9C
on peroxo structure for the tetradentate Cu ligands. The at low temperature, a new species [(PImH)FeIII−(O22−)−
compound 8B converted thermally to the μ-oxo complex CuII]+(9D) was formed, which was characterized using
[(F8TPP)FeIII−O2−−CuII(LMe2N)]+, 8C, as seen for other absorption spectral features at 420, 545, and 565 nm.
related systems.35,133 Surprisingly, for this high-spin peroxo complex 9D, ν(O−O)
In a recent study, Karlin et al. reported a CcO model system stretching was found at 799 cm−1 (ν16/18O2 = 48 cm−1), which
where a histamine moiety is appended to the periphery of a was higher compared to other tridentate ligands with CuII ion,
fluorinated tetrapheylporphyrin to produce a new binuclear studied by Karlin et al. where side-on μ−η2:η2−O22− binding
tridentate ligand system PImH (Figure 9).134 They investigated to both FeIII and CuII had been postulated.133,135,136 The
the dioxygen reactivity of the PImHFeIIcomplex (9A) both in vibrations [ν(Fe−O) = 524 cm−1; ν(O−O) = 799 cm−1] of
the presence and absence of copper ion. Reaction of 9A with 9D were closer to the previously reported HS−TMPA
O2 in THF at −80 °C produced the PImHFe−O2 (9B) species, complexes. Based on the X-ray absorption fine structure
which was characterized by UV−vis absorption spectroscopy. (EXAFS) data and density functional theory (DFT)
Complex 9B was EPR silent and showed upfield-shifted pyrrole calculations, it was found that PImH tridentate chelate induced
resonance (δ = 9.12 ppm) in 2H NMR, indicating a six- considerable geometric distortion in the CuII coordination
coordinated low spin (S = 0) species where the tethered sphere, which forced the peroxo to binding to the Cu ion in an
imidazoyl arm serves as the sixth coordination. The rR spectra essentially η1 fashion instead of η2 fashion. So, the formulation
of 9B showed the characteristic ν(O−O) at 1171 of 9D peroxo species was proposed to be [(PImH)FeIII−
12379 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 10. Parent TMPA heme−peroxo−copper complex (A) along with eight other low-spin heme−peroxo−copper complexes bearing
derivatized TMPA-based copper chelates (B−I) indicating their roles as reported by Karlin and co-workers.142

(μ−η2:η1−O22−)−CuII]+. The species 9D was also EPR silent S either lowering the energy barrier and/or selectively enhancing
= 2 resulting from the antiferromagnetic coupling between FeIII heterolytic (over homolytic) O−O bond cleavage. Previously,
(S = 5/2) and CuII (S = 1/2) centers.134 On the contrary, a reaction of biomimetic heme−peroxo−copper complexes with
low-spin analogue of [(PImH)FeIII−(O22−)−CuII]+(9E) was exogeneous substrates like phenol or 4-nitrophenol was
obtained when an axial base DCHIm in THF was introduced investigated to understand the role of H-bonding in “O−O”
at −80 °C. In absorption spectroscopy, incorporation of bond cleavage.137,138 Mimicking H-bonding with exogeneous
DCHIm changed the Soret and Q-band to 425 and 538 nm substrates in synthetic systems is complicated by the loss of
from 9D, and relatively intense low energy features were entropy in positioning a H-bond participant with the reactive
observed in the 800−900 nm region originating from the unit which is absent in the enzyme systems.137−139 Therefore,
_
peroxo to metal LMCT band indicating an LS μ-1,2 peroxo Karlin et al. prepared 16 novel synthetic heme−(μ-O22 )− d

species. In addition, the 2H NMR spectra confirmed the spin Cu(XTMPA) complexes using the parent TMPA ligand
change from 9D to 9E with the change in chemical shift value (mentioned earlier) where they incorporated intramolecular
from 93.0 to 10.2 ppm. The S = 0 EPR silent ground state of H-bonding moieties to stabilize and spectroscopically detect
the 9E species resulted from antiferromagnetic coupling of the the proposed Ip intermediate where a putative metal-bound
LS FeIII (S = 1/2) and CuII centers. rR spectra indicated that hydroperoxide species in the active site is H-bonded with
ν(Fe−O) dramatically increased by 86−610 cm−1 due to the tyrosine residue through a water molecule to facilitate the
formation of LS peroxo species 9E. necessary proton coupled electron transfer (PCET) from
H-bonding plays a crucial role in the heme−Cu oxidase tyrosine residue29,140,141 and to understand the steric and
chemistry, and it is likely that the reductive cleavage of the electronic effects of H-bonding on O2-reduction. A series of
“O−O” bond in the active site is often controlled by it by TMPA-based Cu ligands which differ in the substituents in the
12380 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

6
pyridyl position like CH3TMPA, (CH3)2TMPA, OCH3TMPA, peroxo−Cu core. Therefore, optimal H-bonding interactions
PV
TMPA, A 2 TMPA, N H 2 TMPA, ( N H 2 ) 2 TMPA, and were required to the peroxo core for activating the O−O bond,
NH2,CH3
TMPA (Figure 10) were paired with tetrakis(2,6- which was exactly served by the variant with two amine groups
difluorophenyl)porphyrinate (F8) heme for generation of low- (10I) revealing a key orientational requirement of H-bonding
spin heme−peroxo−copper synthetic model complexes for O−O bond cleavage.
[(DCHIm)(F8)FeIII−(O22−)−CuII(TMPA)]+, (DCHIm = Naruta and co-workers prepared a binuclear complex,
1,5-dicyclohexylimidazole) and were characterized using [(TMP)Fe I I -(5MeTPA)Cu I ]BPh 4 (TMP-5MeTPA =
absorption and rR spectroscopy.142 It was found that 10,15,20-tris(2,4,6-trimethylphenyl)-5-(2′-bis((5′′-methyl-2′′-
derivatives of the parent ligands which contain one or two pyridylmethyl)aminomethyl)pyridine-5′-
amine ligands preferentially formed strong H-bonding with the carboxyamidophenyl)porphyrin (11A) which undertook re-
Cu bound oxygen atom of the heme−(μ-O22−)−Cu(XTMPA) action with O2 in CH3CN at −30 °C (Figure 11).143
complex (10G, 10H, and 10I), which was evidenced by the Absorption spectra showed that bands at 428 and 534 nm
blue shift of the peroxo to Fe charge-transfer band in the 800− were replaced by new bands at 420, 557, and 612 nm
900 nm region compared to the parent LS−(DCHIm)− indicating the formation of a peroxo-bridged FeIII−O22−−CuII
(F8)FeIII−(O22−)−CuII(XTMPA) complex (10A) and the species (11B). This μ-peroxo complex could be crystallized
control samples LS−CH3TMPA (10E), LS−(CH3)2TMPA from its CH3CN solution and yielded dark purple crystals on
(10F). This was due to the fact that the H-bonding interaction prolonged standing at −30 °C, without decomposition. This
lowers the energy of peroxo π* donor orbitals more compared was the first reported crystal structure for a heme−Cu μ-
to Fe acceptor orbitals. Remarkably, these set of complexes peroxo complex, whose single-crystal X-ray analysis showed
allowed the detection of the ν(Cu−O) which was typically that the FeIII−O22−−CuII moiety has a μ−η2:η1 coordination;
unidentified in similar LS CuII−(μ-O22−)−FeIII complexes. both the oxygen atoms (O(2) and O(3)) of the peroxo ligand
The rR data indicated that with the increase in the steric factor bind to Fe(1), while only O(2) binds to Cu(1). The observed
(presence of ligand methyl substituents) the stretching O−O bond length was found to be 1.460 Å, with a Fe(1)−
frequency of the Fe−O and Cu−O bonds gradually decreased O(2)−Cu(1) bond angle of 166. In additon, the crystal data
(Table 1). However, in the case of H-bonding amine showed that the iron atom lies 0.595 Å above the porphyrin
plane. Solution EXAFS data and rR spectroscopy were
Table 1. Spectroscopic Properties of LS Heme−Peroxo consistent with the solid-state structure, the latter exhibiting
Complex with TMPA-Derived Cu-chelate142 an isotope-sensitive ν(O−O) at 790 (16O2)/746 (18O2) cm−1.
Raman 16O2 (cm−1) The Mossbauer spectrum of 11B at 77K (zero field) showed a
sharp quadrupole doublet with parameters (ΔEQ = 1.17 mm
(DCHIm)
F8FeIII−O22−− ν ν ν s−1, δ = 0.56 mm s−1) that are typical for an HS FeIII
CuII(L) UV−vis (nm) (O−O) (Fe−O) (Cu−O) ν2, ν4 compound and suggested that this was a peroxo bound
10A 426, 535, 850, 812 623 535 1572, iron(III) complexes. The ν(O−O) and high spin FeIII ground
950, 1020 1364 state of the heme iron observed in 11B adds credence to the
(br)
assignment of 10B discussed earlier based on its reported
10G 426, 536, 790, 775 625 533 1570,
900 (br) 1364 ν(O−O) and high spin FeIII ground state.143
10I 425, 538, 785, 760 614 506 1571, Apart from investigating O2 reactivity in homogeneous
897 (br) 1367 solutions, Karlin and co-workers also studied the electro-
10H 425, 533, 789, 765 610 510 1571, catalytic behavior of the Fe−Cu complexes that were
897 (br) 1367
synthesized with tetradentate Cu ligands (TMPA).109 Electro-
10E 422, 534, 855, 789 612 525 1570,
980 (br) 1363 catalysis was carried out over the EPG electrode in 0.1 M
10F 420, 538, 845, 776 593 501 1570, CF3COOH by physiadsorbing three different complexes:
940 (br) 1363 mononuclear porphyrin complex 6A along with the tethered
10B 429, 536, 840 806 573 489 1571, binuclear Fe−Cu complex, 6L−FeCu (7A), and the Cu free
(br) 1365 form of 7A, i.e., 6L−Fe with the Cu free tether. RDE and
10C 426, 535, 823 802 580 484 1571, subsequent K−L analysis showed that both 7A and its Cu free
(br) 1366
variant catalyzed oxygen reduction by 4e− to produce water,
10D 424, 539, 853, 805 575 483 1569, while 6A, the simple F8TPP−FeII-reduced oxygen primarily in
946 (br) 1365 the 2e− pathway produced H2O2. RRDE data quantitatively
showed that efficiency of 4e− reduction of 7A was 74%, while
substituents in 10H and 10I, there seemed to be no significant its Cu free analogue had 59% efficiency and was only 20% for
effect in the Fe−O and Cu−O bonds and only the O−O bond 6A. These data supported the prior results of Boitrel and co-
became significantly weak (Table 1). This weakening of the workers showing that the Cu ion might not be necessary for
O−O bond by the amine mediated H-bonding could help in the 4e− ORR; rather, it was the stability of the superoxide/
the heterolytic cleavage facilitating selective protonation of the peroxide intermediates formed during O2 reduction that
oxygen atom bound to Cu in these CuII−(μ-O22−)−FeIII mattered. In the Cu free form of 7A, the pyridyl groups of
complexes, which was also supported by DFT calculations. the empty distal tether were likely protonated in acidic 0.1 M
When amide functionalities (10B, 10C) were used instead of CF3COOH solution which stabilizes the negatively charged
their amine counter-parts, it distorted the local geometry to heme-superoxide species [(S)(F8TPP)Fe-(O2)], and hence, it
such an extent that H-bonding to the peroxo core could only showed higher 4e−/4H+ ORR selectivity. The role of Cu in 7A
impart a weak electronic effect despite the fact that amide was to do the same by stabilizing the heme superoxide leading
groups could provide much stronger H-bonds to the heme− to the formation of 7B and the corresponding μ-oxo species.144
12381 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 11. (A) CcO model complex studied by Naruta and co-workers. The side-on peroxo complex (B) formed by the reaction of (A) with O2
has a single-crystal XRD structure.143

Scheme 2. Proposed Mechanism of Homogeneous O2 Reduction by 6L−FeCu and Its Cu Free Form, 6L−Fe in Organic
solution (A) at rt and (B) at −60 °C27,a

a
Reproduced with permission from ref 27. Copyright 2011 PNAS.

The authors further established the role of distal Cu in ORR complex at −80 °C. The change in Soret band as a function of
mechanism by demonstrating selective and efficient (turnovers the reaction temperature represented a change in the identity
1000) 4e− reduction of O2 to water using 7A and its Cu free of the steady state intermediate. The 422 nm band of the
form as catalyst, where decamethylferrocene (Fc*) was used as reduced heme (FeII) in the steady state at room temperature
electron source and trifluoroacetic acid (TFA) as proton shifted to the 415 nm band representing the FeIII−OOH
donor. A detailed kinetic investigation revealed that at room species at low temperature. The O−O bond cleavage of this
temperature (25 °C) O2 binding to heme (FeII) was the rate- newly observed FeIII−OOH species was assigned to be the
determining step for both 7A and its Cu-free form as was rate-determining step at −60 °C (Scheme 2B). At room
evident from accumulation of the reduced heme−FeII species temperature, the rate of O2 binding to 6L−FeCu as measured
with soret at 422 nm (Scheme 2A). The spectral changes by the turnover frequency (TOF = 41 s−1) was almost twice
observed at −60 °C, were quite different from that at room compared to 6L−Fe (TOF = 24 s−1), whereas the rate of O−O
temperature. At −60 °C, the species observed under steady bond cleavage of the FeIII−OOH species at −60 °C was same
state was assigned to be a FeIII−OOH species for both the for both 7A and its Cu free form. Therefore, Cu in the distal
complexes. This (FeIII−OOH, CuII) species was characterized site was proposed to enhance the rate of O2 binding during the
using UV−vis absorption (415 and 538 nm), ESI-MS, and catalytic cycle, but it did not seem to contribute to the O−O
EPR spectroscopy and could also be independently generated bond cleavage process for these model systems during
by adding excess TFA to the EPR silent (FeIII−(O22−)−CuII) homogeneous catalytic ORR.27
12382 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

The mechanistic details of ORR by CcO model complexes Collman.125 These complexes showed the general trend of
were mostly gained under single turnover conditions in organic increasing the amount of PROS generation with a decrease in
solvents and in the presence of a strong acid. The mechanistic ET rate as observed for other CcO models. However, the 6L−
details in aqueous solvent (more relevant to the working Fe(Im)Cu complex showed higher 4e−/4H+ selectivity (4 ±
conditions of the natural enzyme) remained unexplored due to 1% PROS only) under slow electron flux (C16−SAM)
a lack of direct spectroscopic evidence for any intermediate compared to Fe−Cu (13.5 ± 1%) and Fe-only (23 ± 3%)
involved. Dey et al. developed a setup where dynamic complexes which showed almost 94% selectivity for the 4e−
electrochemistry is combined with a spectroscopic technique, reduction of O2 to H2O under fast ET conditions over the
surface-enhanced resonance Raman (SERRS), which allows EPG electrode. Complex 12B followed the opposite trend, and
direct in situ observation of intermediates under steady-state the amount of PROS was 25% over EPG, i.e., showed only
conditions, and from that mechanistic insight of ORR 75% selectivity for the complete 4e−/4H+ reduction of O2 to
catalyzed by iron porphyrin can be obtained in aqueous water. SERRS-RDE data of all the three complexes over the
medium.145 In this technique, all those species whose rate of C8−SAM-modified Ag electrode at cathodic ORR potential
formation is greater than the rate of decay get accumulated revealed that both 6L−FeCu (7A) and 6L−Fe(Im)Cu (12B)
over the electrode during ORR. The synthetic CcO model showed a LS−FeIII component as the major species under
complexes of Karlin and co-workers, 6L−FeCu (7A), its cu free steady-state conditions with ν4 and ν2 values (oxidation and
form (12A), and the imidazole adduct of 6L−FeCu (12B), spin state marker band) at 1366 and 1567 cm−1 and at 1368
were used by Dey et al. for electrocatalytic ORR under and 1569 cm−1 (red traces in Figure 13A,B along with their
physiological conditions to understand the role of Cu and Lorentzian fitted data in the right side of the spectra),
imidazole and to detect the intermediates formed during ORR respectively. Apart from the LS−FeIII species, two minor
(Figure 12).80 components HS−FeII denoted by 1349 cm−1 (ν4) and 1542
cm−1 (ν2) for 7A (Figure 13A, Lorentzian fit, green trace) and
1345 cm−1 (n4) and 1544 cm−1 (n2) for 12B (Figure 13B,
Lorentzian fit, green trace) and HS−FeIII denoted by 1359
cm−1(ν4) and 1555 cm−1 (ν2) for 7A (Figure 13A, Lorentzian
fit, purple trace) and 1357 cm−1 (ν4) and 1559 cm−1 (ν2) for
12B (Figure 13B, Lorentzian fit, purple trace) were also found
under the steady state. In contrast, complex 12A, the Fe-only
form, showed an HS−FeIII species as the major component
with ν4 and ν2 values at 1361 and 1556 cm−1 (Figure 13C, red
trace in the Lorentzian fit), while the minor components were
HS−FeII (green trace in 13C), LS−FeIII (sky blue trace in
13C), and FeIV�O (purple trace in 13C). SERRS-RDE data in
Figure 12. Synthetic CcO model complexes of Karlin et al. used for the low-frequency region indicated the formation of a bridging
electrocatalytic ORR in aqueous medium.80 peroxide intermediate [no H/D shift on ν(O−O)] during O2-
reduction by both complexes 7A and 12B under steady-state
reaction conditions, suggesting that O−O bond heterolysis was
RDE data of these three complexes and their subsequent K- likely to be the rate-determining step (rds) at the mass transfer
L analysis revealed that the selectivity and efficiency for
limited region. The O−O vibrational frequencies were at 819
complete 4e− reduction of O2 over the EPG electrode
cm−1 (759 cm−1 in 18O2 saturated pH 7 buffer) for the 6L−
exhibited by 6L−Fe(12A) and binuclear 6L−FeCu (7A) were
FeCu complex and at 847 cm−1 (786 cm−1 in 18O2 saturated
higher (n = 4) than its imidiazole adduct (12B) (n = 3.4). The
pH 7 buffer) for the 6L−Fe(Im)Cu complex, indicating the
second-order rate constants of ORR (kcat) were obtained as
formation of side-on and end-on bridging peroxo intermedi-
(3.65 ± 1.1) × 106 M−1 s−1, (4.49 ± 0.9) × 106 M−1 s−1, and
(1.28 ± 0.11) × 106 M−1 s−1 for complexes 12A, 7A, and 12B, ates, respectively (Figure 14A,B). Alternatively, in the case of
6
respectively. It should be noted that the kcat value of 6L−FeCu L−Fe, no such bridging peroxide was formed in the absence
(7A) is 1 order of magnitude greater than synthetic heme− of a Cu center. The peroxo adduct in the organic solution was
copper complexes reported by Collman previously.28 Addi- a HS FeIII−O22−−CuII compound having a ν(O−O) value 30
tionally, RRDE experiments provided quantitative detection of cm−1 less compared to the one in aqueous medium. The
PROS during ORR as well as its dependence with the variation binding mode of the bridging side-on peroxo to the Fe and Cu
of ET rates from the electrodes. In order to control the ET rate centers was likely six coordinated LS μ−η1:η2 (η2 at Cu and η1
to the catalyst, self-assembled monolayers (SAM) of thiol- at Fe) in aqueous medium. These data suggested that side-on
modified Au electrodes were used as working electrode (as bridging peroxide intermediates could indeed be involved in
previously demonstrated by Collman and Chidsey) by Dey and fast and selective O2 reduction in these synthetic complexes.
co-workers, where the modification in the chain length of the Accumulation of end-on bridging peroxide during the ORR
thiols can vary the ET rate from 105 s−1 to 10 s−1.123,127 RRDE mechanism by 12B indicated that at first O2 bound to the Cu
data of 7A and 12B over the EPG electrode, C8−SAM center, which went through 1e−/1H+ O2 reduction to produce
(equivalent to fast SAM S2) modified Au electrode, and C16− higher PROS under fast ET conditions (Scheme 3B). The
SAM (equivalent to slow SAM S1, used by Collman) modified plausible mechanism of O2 reduction by complexes 7A and
Au electrode suggested that the presence of an additional 12B is shown in Scheme 3A,B.80 These results clearly indicated
redox center like Cu reduced the amount of PROS generation the proposals from Boitrel, Karlin, and Naruta on the likely
as compared to the Fe-only analogue, and the effect was more importance and involvement of a μ-peroxo species in O2
prominent under slow electron flux as has been reported by reduction by heme/Cu systems.
12383 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 13. SERRS-RDE data of (A) 6L−FeCu (complex 7A), (B) imidazole adduct of 6L−FeCu (complex 12B), and (C) iron only complex (12A)
in the high-frequency region.28

Figure 14. SERRS-RDE data in the low-frequency region under steady-state conditions in the presence of air and 18O2-saturated pH 7 buffer for
(A) 6L−FeCu (complex 7A) and (B) the imidazole adduct of a 6L−FeCu (complex 12B) on C8SH-modified Ag surfaces.28

2.2. Role of the Distal Phenol Group externally added phenolic substrates.137,138,146 The reactivity of
In order to thoroughly investigate the role of the redox-active two different low-spin bridging peroxo complexes, LS-
tyrosine residue (Tyr244) in CcO for the selective and efficient 4DCHIm (15A) and LS-3DCHIm, toward O2 were compared
4e−/4H+ reduction of O2 to water, Karlin and Solomon used a using different proton sources and Fc* as a reductant in an
combination of experimental and theoretical methods on the organic solvent.138 In 15A, the weakly acidic 4-NO2−phenol
heme−peroxo−copper model complexes in the presence of H-bonded with the Cu-bound Operoxo atom and with the
12384 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Scheme 3. Proposed Mechanism of O2 Reduction in was formed by a rate-limiting ET process to [LS-4DCMIm-


Aqueous Medium by the Complex 7A (A)28 and by the (ArOH)] which was then followed by another 1e−/1H+
Complex 12B (B),28 Respectivelya reduction step at a rate 3 times faster than the earlier rate
(Figure 15). The activation of the O−O bond in the (15A +
PhOH) adduct, [LS-4DCHIm(ArOH)], was supported by the
rR data where ν(O−O) decreased from ∼870 to 827 cm−1
(>40 cm−1) on H-bonding and the corresponding ν(Fe−O)
increased slightly (∼10 cm−1) from ∼590 to 598 cm−1; i.e., the
O−O bond gets weakened and Fe−O bond strengthens. The
H-bonding interaction was also indicated in the DFT-
optimized structure of [LS-4DCHIm(ArOH)], and on the
basis of a combination of the experimental and theoretical
results, it was concluded that the observed shift (∼10 cm−1) of
the ν(Fe−O) mode was due to H-bonding of the PhOH with
the distal O atom of the FeIII−O2−CuII species as reported
previously for the same in heme metalloenzymes.147,148
Deuterated PhOD indicated a kinetic isotope effect (KIE) of
1.6 in the PhOH association step and 1.9 in the rds that is
quite lower than what is expected for phenolic H-atom transfer
(HAT), but it lies in the range of a PCET reaction149,150
confirming the involvement of phenolic O−H/O−D bond in
both H-bond association and rds. The analogous heme−
peroxo−copper complex having three monodentate DCHIm
groups around the copper, LS-3DCHIm, did not react with the
weakly acidic phenol due to steric hindrance as phenol would
not have access to the relatively compressed peroxo core for
the H-bonded adduct formation unlike 15A, which mainly
contributes to the difference in reactivity between these
complexes. In contrast, a relatively stronger acid like [DMF-
H+](OTf−) resulted in M−O bond cleavage via protonation to
the proximal oxygen atom with respect to the iron-releasing
H2O2 while reacting with both of these complexes. Thus, the
O−O bond homolysis was affected by the hydrogen-bonding
network of the properly oriented phenol group close to the
peroxo, which facilitated the PCET reaction (here, from 4-
NO2−phenol and Fc*, and in CcO the cross-linked Tyr
residue) in a fashion similar to the generally proposed CcO
enzymatic mechanism.57
a
In a related work, Solomon and Karlin groups established
Reproduced from ref 28. Copyright 2015 American Chemical the mechanism of O−O bond cleavage by the weakly acidic
Society. redox-active phenol in a reaction with biomimetic low-spin
heme−peroxo−-copper complex, {[(DCHIm)(F8)−FeIII]−
addition of 2 equiv of Fc*, and the O−O bond was cleaved (O22−)−[CuII(AN)]}+ (16A, AN = bis[(3-(dimethylamino)-
yielding Fe(III) and Cu(II) as the final products. The kinetics propyl)amine).137 This involved an H-atom transfer from
suggested that a transient ferryl species [(DCHIm)F8FeIV�O] external electron-rich phenol (4-OMe-phenol) to the distal

Figure 15. Overall proposed mechanism for the reaction of LS-4DCHIm (15A) with 4-NO2−phenol and Fc*.138

12385 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 16. Two possible reaction mechanisms for phenol-induced O−O cleavage of 16A.137

oxygen atom of the bridging peroxide leading to the formation mechanism and the factors controlling the overall rate and
of FeIV�O, CuII−OH, and phenoxyl radicals analogous to the selectivity are crucial. In the case of complexes dissolved in
reaction mechanism of HCOs. Two different possible solutions, the intermediates involved in a reaction can be
mechanisms were considered: one involved the proton- investigated by trapping them at low temperatures and
initiated pathway (PI), where the proton transfer from the interrogating them with spectroscopic methods. However,
phenol occur before the barrier, and the other involved H- mechanistic investigation of reaction catalyzed by surface-
bond initiated pathway (HB) where the O−O bond hemolysis adsorbed thin layer (often monolayers) of catalysts is
could be the predominant reaction with the phenol remaining challenging as conventional spectroscopic techniques are not
H-bonded to the Cu-bound Operoxo atom in the TS and PT amenable to such systems and the amount of sample is too low
primarily occur after the barrier (Figure 16). In both to generate good signal. For this purpose, Dey et al. developed
mechanisms, ET from phenol occurred after the PT (and SERRS-RDE technique which probes a system under steady
after the barrier) and the barrier of O−O bond cleavage was state while performing electrocatalytic ORR in aqueous
only influenced by the interaction with the phenolic proton. medium (Figure 17).145,153 To start with, simple mononuclear
Furthermore, a KIE value of 1.7 was obtained, and the barrier
for this process (ΔG⧧expt) was estimated to be ∼15 kcal/mol
(at −70 °C), which is very similar to the theoretically
predicted value of HB mechanism exhibiting smaller secondary
KIE (<2 for the HB mechanism and >5 for the PI mechanism
independent of the functional used), therefore suggesting the
reaction of 16A with 4-OMePhOH proceeded via the H-bond-
assisted O−O homolysis mechanism.137 To understand the
relevance of such reaction (16A + PhOH), a computational
model of CcO active site was considered in which a cross- Figure 17. Mononuclear iron porphyrins FeTPP(A) and FeEs4(B),
linked tyrosine residue is widely known to participate in the studied by Dey and co-workers in SERRS-RDE.145
O−O cleavage step. It was determined that the active-site
tyrosine residue likely acts as the proton donor generating
tyrosinate which again provides an electron to favor the O−O iron porphyrin, FeTPP (17A), which does not have any distal
bond cleavage. It was proposed that in CcO the water residue, was immobilized over a C8−SAM-modified Ag
molecules present around the active site could interact with the electrode to carry out the mechanistic investigation along
peroxide by means of PT or H-bonding interactions lowering with another mononuclear iron porphyrin, FeEs4 (17B), which
the barrier for O−O bond cleavage in a manner analogous to had a distal H-bonding cavity supported via triazole
the phenol in this model complex. The effect of PhOH in O2 substituents on the meso-phenyl rings of the TPP framework.
reduction was also investigated later in detail in mononuclear The high-frequency region of the SERRS-RDE data indicated a
Iron porphyrins where phenol was covalently attached to the high population of FeIII-LS species during the ORR (Figure
porphyrin ring and is discussed in the next section.148 18A, red trace) compared to its data at 0 V (Figure 18A, blue
trace). The Lorentzian fit of the ν4 region obtained during
catalytic ORR showed a strong peak at 1369 cm−1 at an
3. MONONUCLEAR METALLO-PORPHYRINS AS ORR applied potential of −0.50 V (Figure 18B), and the difference
CATALYSTS spectra (reduced at −0.50 V - oxidized resting state) showed
In order to design efficient molecular ORR catalysts as that the resting FeIII−HS species (1360 cm−1) (Figure 18C,
alternatives of elegant but synthetically challenging binuclear yellow trace) was almost fully converted to FeIII−LS species at
CcO mimics, which are known to retain high selectivity and steady state along with very small population of FeIV�O
faster kinetics, the understanding of the oxygen reduction species denoted by a vibration at 1371 cm−1 (Figure 18C,
12386 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 18. (A) Overlay of oxidized (blue at 0 V) and aerobically reduced (red at −0.5 V) SERRS data in the high-frequency region for FeEs4. (B)
Lorentzian ν4 fit of the aerobically reduced spectra (at −0.5 V) of FeEs4. (C) Difference spectra of the same (−0.5 V) to resting oxidized state along
with Lorentzian ν4 fit. All potentials are expressed with respect to Ag/AgCl (saturated KCl) reference electrode.145

brown trace). To obtain metal ligand stretches, labeling which were then used to investigate O2 and H2O2 activation
experiments with 18O2 saturated pH 7 as well as pD 7 and subsequent reactivity and are discussed below.
phosphate buffer solution was done. It was found in the low- 3.1. Role of Hydrophobicity and Steric Effect in the Distal
frequency region of the data for both 17A and 17B, a 830 cm−1 Pocket
vibration gained intensity as the potential was gradually Collman and co-workers first demonstrated reversible oxygen-
lowered to cathodic potential where ORR took place (Figure ation of myoglobin (Mb) and hemoglobin (Hb) by
19A,C for 17B, data not shown for 17A). This 830 cm−1 band introducing sterically hindered hydrophobic matrix over
was shifted to 782 cm−1 in 18O2 buffer (Figure 19B,D), which porphyrin in the 1970s. They synthesized “Picket fence”
was denoted as ν(O−O) of a LS FeIII−OOH species based on porphyrin (20A) which prevented the irreversible bimolecular
previous literature reports.147,154 The corresponding ν(Fe−O) condensation of iron(II) porphyrins to form μ-oxo-bridged
of FeIII−OOH species were obtained at 631 cm−1 (ν16/18O2 = dimer upon reacting with O2 (Figure 20).156 This was the first
37 cm−1) (see the difference spectrum at different potentials instance where Fe−O2 adduct was trapped and characterized
from 16O2 to 18O2 in Figure 19E) and at 634 cm−1 (ν16/18O2 = to be a diamagnetic species. It was shown that oxygen binds
48 cm−1) (not shown) for 17B and 17A, respectively. The the Fe in an angular end-on fashion. This model complex, 20A,
ν(Fe−O) of high valent FeIV�O species were observed at 780 has higher O2 binding affinity close to that of Hb and Mb. The
cm−1 (Figure 19A,C) (ν16/18O2 = 28 cm−1) and at 778 cm−1 hydrophobic distal pocket in picket fence makes the redox-
for 17B and 17A, respectively. This spectro-electrochemical active metal center accessible for the nonpolar diatomic oxygen
data represented the first report of intermediates formed molecule.
during ORR by iron porphyrins on electrodes and revealed H-bonding interaction in a hydrophobic environment is the
that the accumulation of LS FeIII−OOH species dominated key to dioxygen activation in the natural enzyme Cyt P450, as
under steady state, allowing more hydrolysis and larger PROS has been reflected in synthetic model systems by Naruta and
production for porphyrins having no distal residues (FeTPP) co-workers.157 The authors synthesized a unique ligand system
or porphyrins having hydrogen-bonding triazole residues which contains a bulky binaphthyl-bridged porphyrin ligand
(FeEs4).145,155 Note that the accumulation of FeIV�O species around the central Fe-atom which creates a hydrophobic
under steady state at cathodic potential is surprising as the environment and pendant hydroxyl groups from the binaphthyl
driving force for its reduction >1 V. An FeIV�O species was moieties ensure secondary interaction.157 This system was
also observed during O2 reduction by heme/Cu complexes called single coronet porphyrin (20B), and using this, they
when Fc* was used as a reductant. This suggests that the examined reactions with both dioxygen and carbon monoxide
reduction of FeIV�O to FeIII−OH (a PCET step) is slow, by the help of flash photolysis. Surprisingly, the single coronet
likely due to a large reorganization energy. The presence of had greater affinity toward oxygen than CO, whereas normal
FeIV�O was evidenced later by using it to oxidize an external heme centers have ∼20000 times more affinity toward CO.
substrate. This remarkable property is likely due to the hydrogen
Different second-sphere interactions were investigated using bonding between the terminal oxygen atom of the dioxygen
mononuclear metallo-porphyrins. These platforms are gen- molecule and the hydroxy groups attached to the binaphthyl
erally synthetically amenable, allowing the integration of bridges in 20B. Ligand-iron bond strength and H-bonding was
different types of second-sphere residues in the molecule investigated using IR and rR spectroscopy. Factors like polarity
12387 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

porphyrins.99,158−161 The main drawback is that these Co-


containing catalysts generally tend to favor the two-electron
reduction of O2 to H2O2 except for the bimolecular
path.160,162−165 Therefore, several attempts have been made
successfully to date for the design of efficient and selective
mononuclear Co catalysts introducing hydrogen bonding,166
proton and electron transfer,167,168 electrostatic interactions,169
as these effects are found to be crucial for iron porphyrins to
exhibit 4e−/4H+ ORR.
In a recent work, Cao and co-workers reported electro-
chemical ORR with the four atropisomers of a cobalt−
porphyrin complex, each containing a bulky ο-amide
substituent on the meso-phenyl groups (Figure 21).170 They

Figure 19. SERRS-RDE data of FeEs4 obtained at various potentials


under 16O2-containing (A) and 18O2-containing (B) pH 7 buffers. The
spectra obtained from the difference of various potentials from 0 V for Figure 21. Molecular structures of the four atropisomers of Co
16
O2- and 18O2-containing buffers are shown in (C) and (D), porphyrins: αααα (A), αααβ (B), αβαβ (C), and ααββ (D).170
respectively. (E) Difference spectra of 18O2 from 16O2 obtained at
respective potentials. (F) Shifts obtained, along with their fits.145 demonstrated a new approach for control over product
selectivity, i.e., 2e−/2H+ vs 4e−/4H+ ORR, by tuning the
could also be responsible for this discrimination between O2 steric effect of the different atropisomers. These complexes
and CO binding in enzymes as indicated by Naruta and co- were physiadsorbed over carbon black or CNTs and examined
workers. The role of the hydroxyl group was further verified by for electrocatalytic O2 reduction in 0.1 M KOH solution under
replacing hydroxy groups of the single coronet binaphthyl aerobic condition. All the four atropisomers exhibited similar
bridge with methoxy group where CO affinity of the porphyrin ORR activity with the E1/2 values of around 0.73 V vs NHE but
was found to be strong relative to O2 as expected. diverse selectivity due to their dissimilar steric effect around
For the last several decades, Co porphyrins have been largely the redox-active Co center. The selectivity was determined
investigated as ORR electrocatalysts as these complexes usually from RRDE and K−L analysis. The αβαβ isomer, 21C showed
function at low ηORR in comparison with their analogous iron high selectivity for 2e− ORR (n = 2.10) in the absence of any

Figure 20. Synthetic model complex FePf (A) and single coronet system (B) synthesized by Collman and Naruta et al., respectively.156,157

12388 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

bimolecular path because of the large steric encumbrance on


both sides of the porphyrin ring. Without such a pronounced
steric effect in ααββ (21D) and αααβ isomers (21B), poor
selectivity was obtained for either 2e− or 4e− ORR with n =
2.89−3.10, which is commonly observed for most reputed Co
porphyrins. In contrast, the αααα isomer (21A) catalyzed the
4e−/4H+ ORR with high selectivity (n = 3.75) by improving
O2 binding inside the pocket as observed in UV and EPR data
even in the absence of any axial ligand to block the open side
of the porphyrin ring. Moreover, there is strong spectroscopic
evidence including IR and NEXAFS (K-edge near-edge X-ray
absorption fine structure) for the formation of stable Fe−O2
adduct in the αααα isomer, 21A. In FTIR, the changes in
ν(N−H) and ν(C�O) of amide functional group indicated Figure 22. Synthetic model complexes (A−C) studied by Collman
potential H-bonding interactions between the Fe−O2 adduct and co-workers.126
and amide groups in its distal pocket resulting in stronger O2
binding affinity in 21A than the other three isomers. Therefore, the H-bonded array present in its superstructure (through H-
it was concluded that the hydrophobicity as well as steric effect bonding with the nitrogen centers in the imidazole pendants)
in the distal site employed by bulky tert-butyl groups covalently increased the basicity of the water, which is required for the
attached to the porphyrin ring by the amide functionality can stability of superoxide. This H-bonded array of the water
strongly impact the higher O2 binding affinity to the metal cluster was absent for picket fence 22B. For the complex 22C
center (both Fe and Co) and result in stable Fe−O2 adduct where there is no distal residue, solvent access promoted the
formation, which could enhance selectivity of ORR. Very hydrolysis of Fe−O2 instead of stabilizing it like 22A.
recently, Cao et al. further demonstrated that similar cobalt Therefore, PROS increased in these complexes from 22A to
porphyrins can be grafted over metal organic frameworks 22B to finally 22C, lowering the selectivity of 4e−/4H+
(MOF) through ligand exchange. This porphyrin−MOF ORR.126 The H-bonded water cluster present in 22A which
hybrid synthetic assembly was shown to be a very reactive increases the basicity of water as well as kinetic stability of the
and robust electrocatalyst for ORR having low overpotential superoxide or Fe−O2 species may also be operational in the
compared to simple Co porphyrins. This hybrid set up was gas binding pocket of natural enzyme CcO.
found to be more selective toward 4e−/4H+ ORR owing to Dey et al. synthesized a family of porphyrins with triazole-
further reduction of H2O2 by active MOF surfaces, which containing distal superstructures having electron-withdrawing
enabled it to be used successfully in Zn−air batteries.171 and electron-donating groups. The tetra-triazole distal pocket
3.2. Role of Distal Water-Mediated H-Bonding entrapped H2O molecules as evidenced by X-ray and FTIR
analysis. They showed the effect of H-bonding in both FeIII−
Investigation using biomimetic metallo-porphyrins with distal hydroxide and FeIII−superoxide complexes using electro-
superstructure has aided in a better understanding of the role chemical and spectroscopic data.172,173 In the low-frequency
of distal structure in O2 binding. Collman et al. showed the region of rR spectra, an isotope-sensitive band in the region of
role of water present in the distal pocket of Fe-only CcO 540−580 cm−1 was identified as the Fe−OH stretching
model complexes during catalytic O2 reduction.126 Heteroge- frequency for these set of complexes, which was found to be
neous electrochemical analysis revealed that the PROS 574 cm−1 (544 cm−1 in deuterium) for FeFc4 (26A).When
production during O2 reduction sometimes depends on the electron-donating substituents were introduced in the triazole
presence or absence of water cluster in the distal site of the ring as in the case of Fe(tBu)4 (23B) and Fe(PhOMe)4 (23C),
catalyst. Using the azide SAM-coated Au−Pt interdigitated the Fe−OH stretching frequency of the hydroxide complex
array electrodes (IDAs), this study demonstrated how the decreased to 571 and 568 cm−1, respectively. In the case of
factors that affect the rate of oxygen binding changed the porphyrins having electron-withdrawing substituents in the
selectivity of ORR. Among the three complexes investigated triazole ring like FeEs4 (17B) and Fe(PhCN)4 (23A), the
(Figure 22), 22B and 22C bound O2 quite rapidly with rates as ν(Fe−O) was observed at somewhat higher values at 576
high as 107−108 M−1 s−1. The PROS produced for 22C was cm−1. For example, solution FTIR quantification indicated that
much higher compared to 22B. On the contrary, the tris- two water molecules were encapsulated inside the triazole
imidazole complex 22A bound O2 rather slowly, but it cavity in the distal structure which could form H-bonding with
produced almost half as much PROS compared to that of the triazole N-atoms allowing the external ligand to bind Fe in
the picket fence porphyrin, 22B. Collman et al. showed from the sterically hindered side through H-bonding with the water
titration experiments that six water molecules were present molecules (Figure 23, right). When electron-withdrawing
inside the triazole cavity of 22A among which two to three substituents were introduced in the triazole ring, the electron
water molecules were displaced during O2 binding. The density of the ring decreased resulting in weaker H-bonding
remaining water molecules form a stable H-bonding network interactions with water molecules, which in turn increased the
with the Fe−O2 species which was confirmed by FTIR H-bonding of the trapped water molecules with the distal
spectroscopy. Free superoxide is known to disproportionate at bound OH group; hence, the O−H bond strength decreased
high concentration in water, but at low concentrations they and the Fe−O bond strength increased. This was further
could be spectroscopically identified as superoxide di-, tetra-, verified with FePf (20A), where no H-bonding is present and
or hexahydrate water clusters. Ab initio calculations also therefore the Fe−OH stretching was very weak and the
supported this observation and indicated that the concen- corresponding Fe−OH vibration was observed at 556 cm−1. As
tration of superoxide-bound porphyrin was low for 22A and the H-bonding to the bound axial hydroxide decreased, the π
12389 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 23. Synthetic model complexes (A−C) studied by Dey and co-workers.172

bonding between the dyz orbital of Fe and the 2p orbital of OH


ligand increased with the increase in the Fe−O bond strength,
which imparted more covalency in the Fe−O bond. As a result,
the hydroxide complex of FeEs4 with a stronger Fe−O bond
had a lower pKa value of 3.8, while for Fe(PhOMe)4 the pKa
value increased to 5.5 among the complexes investigated by
Dey at al. They also demonstrated the same effect of water-
mediated H-bonding in the Fe−O bond strength of the ferric
superoxide complexes of these porphyrins, measured in the rR
spectra.172−174 Here too. the ν(Fe−O) of the Fe−O2 species
shifted from 585 cm−1 in FeFc4 to 577 cm−1 in Fe(PhOMe)4
and the effect was mediated by competitive hydrogen bonding Figure 24. Representative structure for [Fe{porphyrin(pyH)4}(O2)·
between the Fe−O2 unit and the distal triazole for the H2O (H2O)2]4+ species with a water cluster as studied by Mayer et al.175
trapped in the cavity.
Mayer et al. reported soluble iron meso-tetra(pyridyl)- Table 2. Electrochemical ORR Data of 25A at Different
porphyrins as efficient electrocatalysts for ORR in acidic pH176
aqueous solution with the goal of affecting proton addition to
pH kcat (M−1 s−1) × 105 PROS (%)
the O2-derived ligand in iron center.175 They synthesized
FeIII−meso-tetra(2-pyridyl)porphyrin (29A), with the pyridyl 4 (59.8 ± 0.7) (2.52 ± 0.3)
nitrogens pointing inward, albeit far away from the iron center, 4.56 (19.9 ± 0.5) (1.82 ± 0.15)
4.67 (65.9 ± 1.2) (0.39 ± 0.4)
and compared its electrocatalytic behavior with 4-pyridyl
5 (27.4 ± 0.4) (2.09 ± 0.25)
isomer (29B) with the pyridines pointing outward. Higher
6 (6.09 ± 0.09) (3.86 ± 0.2)
selectivity toward 4H+/4e− ORR for the 2-pyridyl derivative
7 (1.85 ± 0.02) (5.01 ± 0.15)
compared to the 4-pyridyl isomer was proposed to originate
8 (29.0 ± 0.6) (3.03 ± 0.1)
mainly due to the differences in the proton delivery mechanism
9 (123 ± 11.0) (1.43 ± 0.2)
to the O2-derived intermediates formed during the catalytic
10 (121 ± 10.0) (0.03 ± 0.1)
cycle likely via the protonated pyridines in acidic aqueous
11 (169 ± 15.0) (0.03 ± 0.05)
medium. Computational models suggested that indeed the
protonated pyridine nitrogens were quite far from the [Fe−
O2] complex (pyH+···O = 3.80 Å). In the iron(II) complex increase in second order ORR rate (kcat) compared to neutral
[Fe{porphyrin(pyH)4}]4+, the pKa for the pyridinium groups pH, while almost 40-fold increase in kcat values was observed at
was computed to be about 4.4. Nonetheless, when more than low pH (Table 2). A plot of log kcat vs pH indicated a pKa
one H2O molecule was included between pyridinium ions and value of (5.6 ± 1) at low pH where rate enhancement occurred
bound O2, pyridinium ions seemed to organize them into a which matched with the pKa of phenanthroline (∼5) in water.
cluster. From DFT calculations, this water cluster mediated Therefore, at low pH the phenanthroline moieties were
system was supposed to help in proton delivery through H- protonated and behaved like protonated histidine at the active
bonding to the FeIII−OOH derivative (Figure 24). This was site of HRP which exerts a “pull effect” to the O−O bond of
likely the reason behind the higher 4H+/4e− ORR selectivity of the FeIII−OOH intermediate during ORR, increasing the ORR
these electrocatalysts, especially the 2-pyridyl isomer (29A). rate (Figure 25C). On the other hand, at alkaline pH the
Recently, a synthetic complex Fe−bisPhen (25A), where the putative (H2O)−FeIII−OOH intermediate was likely to be
proton relay along with water mediated H-bonding from distal deprotonated to become a trans OH− bound species. The pKa
structure to the Fe center of porphyrins facilitating ORR value of this change was estimated by them to be 7.9 ± 1. The
activity, was reported by Dey and co-workers.176 Taking OH− has a stronger push effect compared to neutral H2O
inspiration from the two distal residues (Arg, His) of HRP, molecule, which was proposed to enhance O−O heterolysis
complex 25A was synthesized where two 1,10-phenanthroline rates. Apart from the push effect from trans OH− ligand,
moieties above the porphyrin plane were envisaged. This was pyridine groups of phenanthroline directly H-bonds to the
shown to be a very effective and selective 4H+/4e− ORR axial ligands of metal center and through water-mediated H-
electrocatalyst when physiadsorbed over EPG electrode in bonding in aqueous medium. This simultaneous pull effect
aqueous medium throughout a pH range of 4−11. RRDE data from the distal site and push effect from the trans axial OH−
revealed that catalyst maintained >95% selectivity for water ligand might increase the polarity of O−O bond in FeIII−OOH
formation throughout a wide pH range (Table 2). RDE and species, which could explain the 100-fold ORR rate increase at
the subsequent K−L analysis in phosphate buffer solution alkaline pH. The “push−pull” mechanism resulted in a bell-
indicated that at high pH, 25A showed almost 100-fold shaped pH dependence against log kcat which is generally
12390 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 25. Single-crystal X-ray structure of Fe-bisPhen (A), plot of log (kcat) vs pH (B), and the proposed schematic representation of “O−O”
bond activation via “push” and “pull” effects at different pH by Fe−bisPhen (C) as reported by Dey and co-workers.176

typical of enzymatic active sites (Figure 25B). The direct porphyrin, namely, α4-FeFc4 (26A) (Figure 26).173 They
delivery of proton through water-mediated H-bonding or showed that α4-FeFc4 could reduce O2 under homogeneous
directly through phenanthroline nitrogen atoms to the distal
oxygen atom of FeIII−OOH species likely helped to increase
the 4H+/4e− ORR selectivity (Figure 25C).176

Table 3. Cobalt Porphyrins as ORR Catalysts under


Heterogeneous Conditions
catalyst medium Ecat/2 (mV) vs NHE PROS (%) ref
28A 0.5 M H2SO4a 455 29 165
28B 440 67 165
28C 365 51 165
28D 444 65 165
28E 400 61 165
Figure 26. Pictorial representation of the triazole moiety in the
28F 425 52 165 second sphere of porphyrin-based catalysts (A) α4-FeFc4 and (B) α4-
32A 1 M H2SO4b 440 73 163 FeFc1Es3.185
32B 495 30 163
32C 510 55 163
conditions in an organic solvent in the presence of acid and
32D 450 39−47 166
further could catalyze oxygen reduction with equal efficiency in
32E 480 30−35 166
aqueous solution under heterogeneous conditions over a pH
32F 430 5−13 166
a
range of 1−9. In heterogeneous aqueous medium when the ET
Catalysts physiadsorbed over MWCNT films dispersed on GC rate was very fast, i.e., over the EPG electrode, 26A reduced O2
electrode. bCatalysts physiadsorbed over EPG electrode. with >95% 4H+/4e− selectivity. Under moderate SAM (C8−
SAM with ET rate 103 s−1) and slow SAM (C16−SAM with ET
rate 10 s−1), α4-FeFc4 produced 2 ± 1 and 10 ± 1% PROS,
3.3. Role of Distal Redox Active Center respectively, during ORR to H2O (Table 4).123 They also
As discussed in the earlier sections, several metallo-enzymes in decreased the number of Fc groups in the distal structure of
nature which catalyze fundamental reactions like O2 activation, the porphyrin and synthesized α4-FeFc1Es3 (26B) and α4-
reduction to H2O (CcO and multicopper oxidases), O2 FeEs4 (17B) to evaluate this possibility. It was found that when
reduction, and organic substrate oxidation (Cyt. P450) or the number of Fc groups was reduced from 4 to 1 in the distal
nitrogen reduction (nitrogenases) require multiple protons and site, FeFc1Es3 produced 16 ± 1% and 42 ± 2% PROS,
electrons.17,114,177,178 The multiple electrons required for these respectively, over C8 and C16−SAM, while in the absence of all
reactions are generally not supplied by the substrate. Rather, the Fc groups, FeEs4 followed the 2H+/2e− pathway of ORR in
these are derived from ET sites; e.g., in CcO the electrons are the case of slow C16−SAM. Thus, under slow ET rates from
obtained from the redox-active centers present in distal site of the electrode, the hydrolysis of Fe−O2 or FeIII−OOH was
the heme.38,89,179−184 Using a click chemistry approach, the faster than the electron transfer rate from the electrode. When
Dey group incorporated four ferrocene (Fc) groups over the the Fc groups were present, the oxidation of Fc to Fc+ (E0 of
H-bonding triazole residues in the distal structure of Fc+/Fc0 couple in aqueous medium is 0.38 V vs Ag/AgCl)
12391 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Table 4. Iron Porphyrins Used as ORR Catalysts under Heterogeneous Conditions


PROS analysis (%)
catalyst medium Ecat/2 (mV) vs Ag/AgCl kcat(×107) (M−1s−1) EPG C8-SAM C16-SAM Imz-SAM ref
22A pH 7 −30 0.51 × 10−2 10 >20 122
5C pH 7 −30 0.12 × 10−1 7.7 11 122
5B pH 7 5 122
12A pH 7 −50 0.36 6±1 9 ± 0.5 23 ± 3 78
7A pH 7 −80 0.45 6.3 ± 0.1 7 ± 0.1 13.5 ± 1 78
12B pH 7 −50 0.13 25 ± 2 10 ± 1 4±1 78
20A pH 7 10 ± 1 rapid degradation 170
26A pH 7 2 ± 1 10 170
17B pH 7 −145 0.50 × 10−2 15.6 ± 0.2 28.0 ± 4 100 10.0 ± 2.0 170
33B pH 7 −155 0.73 2.30 ± 0.20 6.15 ± 0.3 10.45 ± 0.9 193
33C pH 7 −103 1.80 1.25 ± 0.05 5.20 ± 0.3 4.90 ± 0.5 193
33D pH 7 −105 1.35 2.00 ± 0.10 5.10 ± 0.3 4.00 ± 0.5 2.9 ± 0.1 193
17A pH 7 −320 10 29 12.0 ± 1 193
33E pH 7 −196 0.42 2.54 ± 0.10 8.85 ± 0.1 18.0 ± 0.2 4.95 ± 0.4 194
33E-(N-Me)Imd pH 7 0.18 1.43 ± 0.2 3.10 ± 0.5 7.00 ± 0.1 2.90 ± 0.1 194
33E-(Me) pH 7 2.40 ± 0.3 5.43 ± 0.5 7.90 ± 0.1 2.64 ± 0.1 194
33E-(Ph) pH 7 1.80 ± 0.1 4.73 ± 0.5 9.20 ± 0.4 1.32 ± 0.1 194
31A 1 M H2SO4 60 1.4 22 ± 2 163
25A pH 7 −100 1.85 × 10−2 5.01 ± 0.2 172
27A pH 7 −160 0.19 3.3 ± 0.2 5.5 ± 0.6 3.3 ± 0.4 149
27B pH 7 −220 0.71 4.0 ± 0.4 3.2 ± 0.3 3.6 ± 0.4 149
27C pH 7 −240 5.31 1.8 ± 0.5 4.6 ± 0.2 10.2 ± 0.5 149

Scheme 4. (a) Proposed ORR Mechanistic Cycle in Heterogeneous Medium by α4-FeFc4 in the pH Range 1−9 Indicated by
Pathway A, at pH = 0, when the Triazole Rings Get Protonated by Pathway B Is Observed, while Pathway C Denotes an
Alternative 2e−/1H+ Pathway via Superoxide Protonation. (b) Proposed ORR Mechanistic Cycle in Homogeneous Organic
Medium by α4-FeFc4, in the Presence of 2-3 equiv of Triflic Acid Following Pathway A and Following Pathway B in the
Presence of Excess Acid, Where Triazoles Are Protonated185

during ORR maintained a rapid electron flow to the On the contrary, in spite of the presence of four redox-active
intermediate Fe−O2 species formed during ORR, which Fc groups in nonpolar organic solvent, in the absence of acid
decreased their hydrolysis rate, promoting 4H+/4e− selectivity. source, 26A formed a Fe−O2 intermediate that led to the 1e−
A plot of EPORR (peak potential of ORR catalytic current) vs reduction of O2 to O2− upon hydrolysis, which in turn
pH in aqueous medium showed a 28 mV slope for per unit pH disproportionated to yield 48 ± 3% H2O2, detected by xylenol
unit in the case of α4-FeFc4-catalyzed 4H+/4e− reduction of orange assay of the resultant solution. This difference in
O2, which indicated a 2e−/1H+ PCET step involvement in reactivity between aqueous and organic medium was due to
the shift in reduction potential of Fc+/Fc0 redox couple in
ORR cycle. The proposed reaction pathway in heterogeneous
nonpolar organic medium to 0.73 V from 0.38 V in aqueous
aqueous medium both in the pH 1−9 range and at pH = 0 is medium. Note that in aqueous medium there is enhanced
shown in Scheme 4a. The triazole residues in the distal solvation of the oxidized Fc+ ion, which enables the oxidation
environment having pKa ∼ 1.5 acted as a local buffer in the of Fc centers at lower potential, allowing it to participate in a
vicinity in the O2 reducing iron porphyrin center, promoting PCET process during ORR, which was not possible in organic
4H+/4e− selectivity, while at pH <1, triazoles were protonated, medium. Alternatively, addition of 2−3 equiv of triflic acid to
promoting the hydrolysis pathway of decreasing selectivity of the organic medium ensured the 4H+/4e− ORR to H2O by
4H+/4e− ORR. 26A. Three out of four Fc were oxidized to Fc+ (monitored by
12392 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

absorption spectra) in the process, and the other electron from the semiquinone ring and decayed to form a FeIV�O
needed to reduce O2 to H2O was derived from the iron in the species with the concomitant formation of an oxidized
porphyrin ring. Thus, protonation of the ferric superoxide Quinone(Q). The appearance of ν(Fe�O) at 754 cm−1 in
species could cause oxidation of the Fc moieties resulting in the rR spectra, which was replaced by the Fermi doublet 729
the 4e−/4H+ reduction of O2. The fact that the authors were and 698 cm−1 on using 18O2 and simultaneous decrease of the
not able to trap any FeIII−OOH species or high-valent FeIV� above-mentioned two EPR signals clearly indicated the
O species even at −80 °C suggested that the decay of FeIII− formation of FeIV�O[Q]. RT FTIR data suggested that
OOH via oxidation of Fc was very fast. Only the final product after O2 addition to FeQH2 a quinone was formed as indicated
of ORR cycle FeIII−OH was observed in the EPR spectra at by a vibration at 1520 cm−1 instead of quinolinic band at 1316
−80 °C as indicated by a rhombic signal at g = 6 after addition cm−1 before O2 addition. The intermediate FeIV�O[Q] finally
of 3 equiv of triflic acid to the Fe−O2 species. Thus, in organic decayed to FeIII−OH[Q] with a rhombically split axial g = 6
medium the Fe−O2 adduct underwent a proton transfer EPR signal (Scheme 5). Thus, out of the 4H+/4e− required for
followed by electro transfer (PTET) (Scheme 4b). However, ORR to produce H2O, 2e−/2H+ were obtained from iron (FeII
as shown by Dey et al., an excess equivalent of acid caused getting oxidized to FeIV) and remaining 2e−/2H+ were
hydrolysis of iron superoxide and increased the PROS supplied by the QH2 group appended to the porphyrin
(Scheme 4). The proton delivery pathway to the Fe-bound ring.151 The [Fe−O2] species in synthetic analogues were
O2 species during the course of ORR by these porphyrins was kinetically competent in abstracting HAT from QH2 with bond
made possible by the fact that O2 bound to the site where dissociation free energy (BDFE) of 80 kcal/mol. However,
triazole moieties were present and not to the open side. This DFT calculations indicated that the H-bonding interaction
was supported in the rR spectra at −80 °C of the Fe−O2 between ferric superoxide and hydroperoxide with −OH
adduct of 26A showing a diamagnetic low-spin, six-coordinated groups of quinol played a vital role in this facile reactivity.
(CH3OH as trans ligand) Fe−O2 adduct with ν(Fe−O) at 585 The effect of redox active quinol group (FeQH2, 27A) in the
cm−1 which shifted to 561 cm−1 in 18O2. If oxygen bound to second sphere was further compared with that of pendant
the open side, a S = 0, μ-oxo dimer having different phenol (FePh, 27B) and 2,5-dimethoxyquinol derivative
spectroscopic features would have formed.173,174,185 (FeQMe2, 27C), and their O2 reduction activity was
Recently, a covalently attached redox-active hydroxyquinone investigated both in aqueous medium under heterogeneous
group was introduced in the second sphere of iron porphyrin aqueous condition as well as in organic solution.152 RDE data
(FeQH2, 27A) by Dey et al. to evaluate the reactivity of heme of these complexes over EPG electrode and their subsequent
ferric superoxide species toward H atom donors151 which may K−L analysis revealed that all of these complexes 27A, 27B,
also help to understand the mechanism of action of heme and 27C showed selective 4e−/4H+ ORR with second-order
dioxygenases like indoleamine 2,3 dioxygenase (IDO) and rate constants of (7.09 ± 1) × 106 M−1 s−1, (1.97 ± 1) × 106
tryptophan 2,3 dioxygenases (TDO) (Figure 27).186−188 The M−1 s−1, and (5.31 ± 2) × 107 M−1 s−1, respectively (Table 4).
Their corresponding cyclic voltammetry data over the EPG
electrode under anaerobic conditions as well as in the presence
of air-saturated pH 7 buffer solutions showed that for 27A and
27B E1/2 values of FeIII/II redox couples (−250 mV and −310
mV vs Ag/AgCl, respectively) matched with the EpORR
(defined as the peak potential of the ORR current) values,
indicating that reduction of FeIII to FeII was the most
thermodynamically uphill process in the catalytic cycle of
ORR for these two complexes. On the contrary, 27C showed
EPORR at 130 mV which was cathodically shifted from the
FeIII/II redox process, signifying that more driving force than
Figure 27. Synthetic iron porphyrin complexes having hydroquinone E1/2 was required for at least one step during the ORR. E1/2 of
(FeQH2, A), phenol (FePh, B), and 2,5-dimethoxyphenyl rings the FeIII/II process and EpORR were further investigated at
(FeQMe2, C) studied by Dey and co-workers.151 different pH, and the pH dependence of E0 and EPORR had the
slope of 60 mV/pH for both 27A and 27C indicating a 1e−/
dioxygen adduct of reduced FeQH2 at first underwent an 1H+ PCET for both. However, in the case of 27B, EPORR vs pH
intramolecular hydrogen atom transfer (HAT) reaction (kH/kD had a slope of 30 mV/pH which was indicative of a 2e−/1H+
= 4) from the attached hydroxyquinone to form FeIII−OOH, PCET step as the potential determining step of ORR.
and the hydroxyquinone was oxidized to form semiquinone RRDE data of these three complexes over the EPG electrode
(SQ). The FeIII−OOH species was characterized by the (ET rate 106 s−1) agreed well with the K−L analysis of the
formation of 574 and 617 nm band in the UV absorption RDE data. These complexes followed the 4e−/4H+ ORR
spectra along with isotope sensitive ν(Fe−O) and ν(O−O) at pathway with only 4.0 ± 0.5, 3.3 ± 0.2, and 1.8 ± 0.5% PROS
531 cm−1 (ν16/18O = 21) and at 783 cm−1 (ν16/18O = 35), for 27B, 27A, and 27C respectively. When the electron-transfer
respectively, in the rR spectra. On the other hand, the rate was slowed down by using a SAM-attached Au electrode,
formation of SQ species was confirmed by the SQ vibrations at the PROS value remained almost similar to >95% selectivity
834 and 816 cm−1 in the rR spectra, which shifted to 830 and for 4e−/4H+ ORR in 27A and 27B. However, the PROS
811 cm−1, respectively, on deuteration of the quinolinic group. increased gradually from 1.8% in EPG to 4.6% in C8−SAM,
The LS FeIII−OOH[SQ] intermediate species, which resulted and finally up to 10% (5-fold increase) over C16−SAM for 27C
in the simultaneous growth in the g = 2.38, 2.20, 1.90 (for (Table 4). Although these three complexes with second-sphere
FeIII−OOH) and at g = 2.004 (for SQ) signal in the X-band distal residues showed preference for 4H+/4e− ORR and
EPR underwent another intramolecular HAT (kH/kD = 2.2) almost similar rates, their difference in ORR onset potential
12393 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Scheme 5. Proposed Mechanism of O2 Reduction by FeQH2 (27A) Mediated by Hydroquinone, Probed at −80 °C by Dey and
Co-workers151

Scheme 6. (A) Schematic Representation of a Mechanism of an ORR by FePh Involving a Disproportionation Reaction of
Ferric Superoxide Complex. (B) Schematic Representation of an ORR by FeQMe2 in Both Organic and Aqueous Medium152

and its pH dependence, selectivity (% PROS value) under slow electrons to be supplied from electrode. The low reduction
ET rate indicated that these three complexes followed different potential of hydroquinone (∼0.46 V vs NHE) in neutral pH
mechanistic pathways to ORR. In the case of 27C, the allowed ORR by FeQH2 (27A) to progress via two consecutive
reduction of FeIII−O2, i.e., iron superoxide to hydroperoxide H atom transfer pathways (HAT) of the Fe−O2 species. Thus,
step (FeIII−OOH), required an ET followed by a PT pathway oxygen could be fully reduced to water once it bound to the
(ETPT) in aqueous medium. This was consistent with the ferrous state to form Fe−O2 without any additional electron
ORR onset potential being lower than the E1/2 of FeIII/II from the external source as two electrons were supplied by the
process in aqueous medium (Scheme 6B). The pendant ferrous porphyrin and another two electrons were supplied by
methylated quinol group could restrain the hydrolysis of FeIII− the distal hydroxyquinol group being converted to the FeIV�
OOH species producing high 4H+/4e− selectivity under fast O and quinone species, respectively, which mitigated the four
ET flux (EPG). Rather, slow ET conditions (C16−SAM) electrons required for selective 4e−/4H+ ORR (Scheme 5).
resulted in a higher %PROS for FeQMe2 (27C) as it required The FeIV�O and quinone formed during the electrochemical
12394 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

ORR might be reduced back to the ferrous quinol state on the


electrode at the potential where ORR was observed. Therefore,
27A should be intrinsically more selective toward 4e−/4H+
ORR even if ET rate from electrode was slowed. In organic
medium, consecutive HAT pathways during ORR involved
species like Fe−O2−QH2, FeIII−OOH−SQ, and FeIV−oxo−Q
intermediates which were all characterized using UV−vis, EPR,
and rR spectroscopy in organic medium, already discussed
previously (Scheme 5).151 On the other hand, phenol has a
higher reduction potential (∼0.8 V vs NHE) compared to
hydroquinone but has low pKa (pKa = 9.98), i.e., greater
acidity. This phenomenon helped FePh (27B) to make the
reduction of Fe−O2 to a FeIII−OOH species more facile via an
intramolecular PCET pathway, where proton was delivered Figure 28. (Left) Hangman carboxylic acid containing Co porphyrins
from the pendant phenol and the electron came from another (CoHPX) with substituted meso-phenyl groups (A−E). (Right)
Fe−O2 species present in solution (Scheme 6A). The DFT- Pentafluorotetraphenyl Co porphyrin [Co(C6H5)4] (F) studied as an
optimized structures of the reactant Fe−O2 and product FeIII− ORR catalyst by Nocera and co-workers.168
OOH of FeTPP, FePh, and FeQH2 showed that the O−O
bond length of Fe−O2 was shorter while that for FeIII−OOH
catalytic ORR activities were measured by RDE and RRDE
was longer for FeTPP relative to both FePh and FeQH2, which
techniques. The onset potentials of oxygen reduction varied
were responsible for the smaller displacement in O−O bond
between 550 and 600 mV vs NHE for these complexes, and
during the PCET process for the FePh/FeQH2 than FeTPP.
the peak maxima of the catalytic waves were found between
The calculated inner sphere reorganization energy (λinner‑sphere)
300 and 400 mV at low rotation rates. The number of
for a PCET process, which varied with the square of the
electrons (n) transferred to the oxygen reduction process was
displacement, became 3 kcal/mol lower in both FePh and
calculated by a K−L equation from the RDE data. The higher
FeQH2 than that of FeTPP making the process more favorable
amount of H2O2 production and number of electrons (n)
for the complexes containing pendant phenol/quinol groups involved in ORR for the series of Co porphyrins indicated that
capable of hydrogen-bonding interactions with the intermedi- with the shift in E1/2 of ORR toward a more cathodic potential
ate species. In addition, the prearranged favorable orientations (for the electron-rich substituents at the meso positions B, D,
of the second-sphere residues helped in H-bonding stabiliza- and E) the number of electrons (as obtained from K−L plot)
tion of the FeIII−OOH species to retain >95% selectivity for delivered to O2 decreased (Table 3). The results indicated that
the 4H+/4e− ORR even under slow ET flux from the electrode a competing 2e− reduction pathway was preferred at more
in neutral buffer medium. cathodic potentials (high ηORR). The CoHPX complex 28A
3.4. Role of Distal H-Bonding Interactions showed the highest selectivity (∼70%) for H2O production in
Natural heme enzymes utilize noncovalent secondary inter- the presence of electron-withdrawing pentafluorophenyl
actions like hydrogen bonding in the outer coordination sphere groups (C6F5). Moreover, the reactivities were compared
to stabilize O2 adducts as well as proton-transfer residues to with the electronically similar nonhangman analogue Co-
facilitate protonation of the intermediates involved in the (C6F5)4, 28F, which produced only 48% H2O. These
catalytic cycles of O2 activation and reduction. To understand experimental data suggested that the presence of hanging
the role of these hydrogen-bonding interactions and proton groups along with the electron-deficient porphyrin ligand
shuttles during the dioxygen activation in heme monoox- backbone might be important for the higher prevalence of H2O
ygenases, the challenging task of designing artificial metal- as the product of ORR.168 In particular, the intramolecular
loporphyrins with different hydrogen-bonding and proton- proton-transfer groups installed at the distal site were
transfer groups in the second coordination sphere had been demonstrated to assist facile O−O bond cleavage toward the
pursued and their reactivities were investigated for several 4e−/4H+ O2 reduction to form H2O.
years. Nocera and co-workers constructed a series of Another class of porphyrin, iron(III) meso-
(tetracarboxyphenyl)porphyrin chloride complexes (29C,
mononuclear cobalt hangman porphyrins functionalized with
29D) with ortho- and para-substituted pendant carboxylic
pendant carboxylic acid on a xanthene scaffold attached to the
acid groups, were examined for ORR by Mayer’s group (Figure
meso position of the porphyrin rings (Figure 28).189 The
29).12 The electrochemical oxygen reduction was carried out in
oxygen reduction properties of various meso substitution upon
organic medium (acetonitrile/DMF) in the presence of [DMF-
cobalt hangman porphyrins (CoHPX, A−E) was evaluated.168
H+][OTf−] acid source, and the 2-carboxyphenyl isomer 29C
The effect of the hanging group was evaluated in comparison
with a nonhangman analogue for the proton coupled
multielectron process in ORR, and the electronic properties
of the porphyrin macrocycle was also established with the
meso-phenyl substituents.
Electrochemical O 2 reduction was performed under
heterogeneous conditions in oxygen-saturated 0.5 M H2SO4
by immobilizing CoHPX (Figure 28A−E) on multiwall carbon
nanotubes (MWCNTs), and thin films were produced by
drop-casting CoHPX-MWCNTs solution on a glassy carbon Figure 29. Iron-porphyrin complexes (A−D) investigated as ORR
(GC) electrode which was used as the working electrode. The catalysts by Mayer and co-workers.12,175

12395 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Table 5. Iron Porphyrins Used as ORR Catalysts under Homogeneous Conditions


catalyst solvent proton source E1/2 (mV) vs Fc+/Fc0 TOF (s−1) % H2O2 ref
+ −
17A DMF (DMF-H )Otf −530 2.7 × 101 <15 187
30A ACN (DMF-H+)Otf− −40 2.0 × 102 9 187
30A DMF (DMF-H+)Otf− −630 2.0 × 103 <15 187
30F DMF (DMF-H+)Otf− −486 1.5 × 101 <15 187
30B ACN (DMF-H+)Otf− −390 2.2 × 106 <15 187
30B DMF (DMF-H+)Otf− −611 2.5 × 103 <15 187
30C DMF (DMF-H+)Otf− −547 1.6 × 102 <15 187
30D DMF (DMF-H+)Otf- −536 1.8 × 102 <15 187
30E DMF (DMF-H+)Otf− −491 5.0 × 10° <15 187
30J ACN (DMF-H+)Otf− −326 6.5 × 104 <15 187
30I ACN (DMF-H+)Otf− −296 2.2 × 104 <15 187
30K ACN (DMF-H+)Otf− −280 2.2 × 102 <15 187
30G DMF (DMF-H+)Otf− −362 3.0 × 10° <15 187
29A water HOtf −250 vs NHE 6.0 × 102 5 171
29B water HOtf −150 NA 11−15 171
26A THF TFA NA NA 0 or 50 181
7A acetone TFA Me10Fc 4.0 × 101 minor pdt 27
12A acetone TFA Me10Fc 2.4 × 101 minor pdt 27
33B DMF (DMF-H+)Otf− −545 1.6 × 103 NA 193
33C DMF (DMF-H+)Otf− −502 3.96 × 103 NA 193
33D DMF (DMF-H+)Otf− −531 2.97 × 103 NA 193

Figure 30. (Left) Iron-porphyrin complexes used as ORR catalysts and (right) correlation of log(TOF) and effective overpotential (defined at
catalyst E1/2) for these catalysts.191

was found to be more selective toward 4e−/4H+ reduction of In addition to the iron porphyrin complexes 29C and 29D,
O2 to water than the 4-carboxyphenyl isomer 29D. A Mayer and co-workers compared the selectivity of oxygen
combination of techniques like RRDE and spectroelectro- reduction with 2-pyridyl and 4-pyridyl derivatives (29A and
chemistry along with reactivity with H2O2 indicated the 29B) of iron(III) meso-tetra(pyridyl)porphyrins as ORR
formation of insignificant amounts of H2O2, and H2O was catalysts in acidic aqueous medium (0.25 M HCl/0.5 M
established as the major product of oxygen reduction. KCl) under homogeneous conditions.175 The RRDE data for
However, the 4-carboxyphenyl isomer did not show any these two isomers revealed that 2-pyridyl-substituted chloride
evidence for proton relay, likely because the carboxylic acid complex produced no PROS (H2O2) and the triflate analogue
pointing away from the iron center was a much less selective was 95% selective for 4e−/4H+ reduction of O2. The 4-pyridyl
catalyst. On the basis of a direct comparison between these two chloride and triflate complexes produced almost 15% and 11%
isomers, the role of proton relay in the second coordination H2O2, respectively, which was much higher than the 2-pyridyl
sphere was proposed to be crucial to retain high selectivity complex (Table 5). These water-soluble complexes showed
during ORR. Cyclic voltammetry (CV) measurements under electrocatalytic oxygen reduction at reasonably high rates
purely kinetic conditions at a high substrate/catalyst ratio and (TOF ∼ 600 s−1). The onset potential of the 2-pyridyl
low scan rate allowed the determination of TOF to be 200 s−1 derivative, 29A, was 400 mV vs NHE, i.e., 800 mV more
for a soluble ORR electrocatalyst, 29C, for the first time. negative than the thermodynamic potential of ORR. These
12396 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

pyridine moieties were envisaged to have a proton relay effect, the 2-pyridyl substituents upon addition of [DMF-H+] that
but the second sphere residues were far away from the metal also shifted the E1/2 of FeIII/II to more positive values. As a
center as suggested by the pyH+−O bond distances (∼3.80 Å) consequence, TOF for 30G decreased with an increase in the
in the DFT-optimized structures of the Fe−O2 adduct. amount of [DMF-H+]. The presence of potential proton relays
To investigate O2 reduction with all of these complexes did not displace the pKO2 and the pKa of the catalyst from the
under similar conditions (0.1 M HClO4), the iron(III) (meso- straight lines substantially, although that might suggest poor
tetraarylporphyrin) chloride complexes (29, A−D) were proton translocation.
immobilized on the electrode surfaces using three conventional A series of iron(III) tetra(aryl)porphyrins (31, A−C) where
methods�(i) solution of catalyst in 0.5% nafion, drop casted three meso positions bear a phenyl group and the remaining
onto a GC working electrode (nafion membrane), (ii) a meso position was modified by either a 2-pyridyl, 2-benzoic
suspension of catalyst solution in 0.5% Nafion with 2.5 mg/mL acid, or 2-hydroxyphenyl group and were examined for oxygen
vulcan carbon additive drop casted on GC (nafion/carbon), reduction by Warren and co-workers.192 The oxygen reduction
and (iii) catalysts physiosorbed on edge-plane graphite reaction was performed by drop-casting the catalysts onto BPG
electrodes (EPG conditions)�and RRDE was performed in and EPG electrodes under heterogeneous condition in acidic
order to determine the % H2O2 produced and selectivity of O2 aqueous medium (O2-saturated 1 M H2SO4). The incorpo-
reduction.190 The difference in reactivity upon changing the ration of a single acid/base group at the meso position caused
medium indicated that the selectivity for O2 reduction was significant improvement in electrocatalytic ORR activity
influenced by the nature of the catalyst film on electrode, relative to the parent iron tetraphenylporphyrin (FeTPP).
although this conclusion needs spectroscopic validation of the The acidic pH was needed to confirm the protonation of the
same kind. Nonetheless, the comparison among the four functional groups that is essential for exhibiting a proton relay
catalysts in two isomeric forms clearly suggested that a effect.
properly positioned proton relay (2-substituents) showed All three complexes (Figure 31) showed oxygen reduction
better selectivity for the 4e−/4H+ pathway under similar with onset around 400 mV vs NHE, like Mayer’s
conditions. Since there were four potential relays in this set of
complexes, the effect of individual or single such group might
be indiscernible.
Later, Mayer and co-workers demonstrated that the
installation of such acid/base functional groups on meso-
phenyl rings of 30A and 30G had a negligible effect on the
turnover frequencies (TOFs) of the ORR. Rather, a correlation
was obtained for the TOF vs ηORR under homogeneous
conditions which could be regulated with the E1/2 of the Figure 31. Asymmetric iron(III) tetra(aryl)porphyrins studied by
catalysts in a series of iron porphyrins by changing the solvent Warren and co-workers.192
(DMF, ACN) and acid [DMF-H+] concentrations as well
(Table 5).191 TOF was dependent on the O2 binding
equilibrium, rate-limiting protonation of ferric superoxide
intermediate (Fe−O2) instead of any electrochemical step.
These results were in agreement with the mechanism proposed
for FeFc4 by the Dey group. The reported TOFs vary over a
wide range (100−106) s−1 for these complexes, and the
log(TOF) vs ηORR correlations obtained under different
experimental conditions were linear (Figure 30, right). The
linear scaling relationship was analyzed on the basis of Figure 32. Asymmetric cobalt(II) tetra(aryl)porphyrins studied by
computational studies where the free energy of O2 binding, Warren and co-workers.166,169
pKO2, and protonation of (Fe−O2) species were found to be
directly related to the reduction potential of the catalyst. The
electron-donating substituent shifted the E1/2 of the FeIII/II complexes,12,175 and these were comparatively stable even
redox couple to a more negative value by increasing the O2 after repeated voltammetry scans. The ηORR values for all these
binding affinity and the pKa of ferric superoxide that eventually complexes lie between 800 mV and 1 V, which were quite high
led to the linear increase in TOF, and the presence of an similar to the tetrapyridyl-substituted derivatives.175 Longer
electron-withdrawing substituent showed the opposite trend. steady current densities of these iron porphyrins, relative to
This might likely be the origin of overpotential correlations. FeTPP (17A), indicated higher catalyst stability. Of these, the
The 2-CO2H isomer (30A) in DMF exhibited 100 times current response for FeTPPy (31A) degraded by only 50%
higher TOF than the 4-CO2H isomer (30F) at an extent of during prolonged electrocatalysis and retained a stable current
∼140 mV higher ηORR. Again, catalyst 30A and its 2-CO2Me density of ≥1 mA/cm2, suggesting that the catalyst was not
derivative 30B showed the highest TOFs reported so far, which only more active but also more stable. In aqueous acidic
were also very close in both ACN and DMF medium due to solutions, the pyridine group having lower pKa (pKa = 5.2)
their high ηORR having almost similar redox potentials (E1/2), might be a more suitable acid than phenol (pKa = 10)193,194 in
despite the fact that only 30A appeared to have proton relays. the absorbed catalyst layer on the electrode surface for
However, the 2-pyridine-substituted catalyst 30G might exist promoting proton relay. These results suggested that the iron
in some unique feature other than any other complex in the porphyrins functionalized with pyridine/carboxylic acids (31A
series in DMF and followed the inverse trend of TOF vs acid and 31C) could shuttle protons to the iron−oxygen
concentration. This was likely because of the protonation of intermediates and enhanced stability of the molecular catalysts
12397 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 33. Synthetic iron porphyrins containing pendant bases in the second coordination sphere [(A) FeMPh, (B) FeL1, (C) FeL2, (D) FeL3,
and (E) Fe-MARG] studied by Dey and co-workers.197,198

immobilized onto the electrode surface by noncovalent intercepts from K−L plots constructed from plateau currents
interactions. of the RDE data. The second-order catalytic rate constants
The electrochemical properties of analogous CoTPPy, 32B, (kcat) for CoTPPy (32B) and FeTPPy (31A) were calculated
immobilized on EPG were explored in 1 M H2SO4 under 1 to be 2.9 × 107 and 1.4 × 107 M−1 s−1, respectively, implying
atm O2 using a combination of techniques like CV, RDE, and that the former achieved two times higher kcat at 230 mV lower
RRDE experiments.166 The presence of a single 2-pyridyl ηORR at pH 0 and these kcat values were much higher than
group in 32B improved the O2 reduction kinetics and changed simple cobalt porphyrins196 and also very similar to those
the selectivity to 4e−/4H+ reduction (H2O) like those reported obtained for iron porphyrins having basic functional groups,
by Anson in the 1990s,23,158,159 as opposed to 2e− reduction reported earlier.197
(H2O2) commonly associated with unfunctionalized cobalt To facilitate the rate-determining O−O bond heterolysis of
porphyrins. The ORR onset potential was in between 600−700 a FeIII−OOH species, as determined by in situ SERRS-RDE
mV which was about 50 mV higher than Nocera’s Co- formed during heterogeneous ORR, Dey and co-workers
hangman porphyrins.168 No decay was observed in the incorporated pendant bases of different pKa values in a series of
catalytic current during the first 1.5 h of electrolysis implying simple mononuclear iron porphyrin complexes (Figure 33)
that the catalyst to be as stable as 31A.192 The value of n which were rationally designed, inspired by HRP, to achieve
calculated from the RRDE experiment for the CoTPPy (32B) efficient O−O bond activation and site-selective proton
adsorbed on EPG was 3.51 ± 0.02 (∼70 ± 3% H2O) at 0.4 V transfer to effect facile and selective electrochemical reduction
applied potential (ηORR = 0.83 V), whereas unfunctionalized of O2 to water.197,198 The complexes FeL1, FeL2, FeL3, and
CoTPP (32A) showed n = 2.6 ± 0.1 (∼27 ± 3% H2O) under Fe-MARG (Figure 33B−E) containing N,N-dibenzylaniline,
identical conditions (Table 3). The 70% selectivity for ORR to pyridine, and aliphatic primary and guanidine groups,
H2O is similar to the selectivity exhibited by CoHPX (28A) respectively, were physiadsorbed on EPG electrode and
from Nocera. However, RRDE analysis of analogous iron analyzed for electrochemical ORR in neutral aqueous buffer
porphyrins provided n = 3.6 (∼80% H2O) for both 31A and (pH = 7) medium.
17A at 0.1 V applied potential (ηORR = 1.1 V). For the Fe All these catalysts showed electrocatalytic O2 reduction in
complexes investigated, the 2-pyridyl group did not change neutral pH 7 buffer with a substrate diffusion limited current at
selectivity but greatly improved stability. The observed an onset of −100 mV vs Ag/AgCl for FeL2 (33C), FeL3
increase in selectivity for 4e−/4H+ ORR in 32B versus 32A (33D), and Fe-MARG (33E). The kinetics and selectivity of
was specifically ascribed to the 2-pyridyl group and more O2 reduction were determined using K−L analysis from the
generally due to the presence of a proton donor or H-bonding RDE data, and the kcat values obtained on the EPG electrode
group. The presence of the 2-hydroxyphenyl group in complex surface for 33B, 33C, 33D, and 33E were (7.28 ± 0.02) × 106
32C resulted in greater selectivity for the 4e− reduction of O2 M−1 s−1, (1.80 ± 0.07) × 107 M−1 s−1, (1.35 ± 0.09) × 107
to H2O relative to 32A but not as much as that of the pyridyl M−1 s−1, and (4.2 ± 0.05) × 106 M−1 s−1, respectively (Table
group, 32B. At 0.50 V applied potential (ηORR = 0.73 V), the 4). The measured kcat values were >106 M−1 s−1, 1 to 2 orders
calculated yield of H2O for 32C was almost half compared to of magnitude higher than the values reported for unfunction-
that for 32B. As is the case with the Mayer complexes,175 the 2- alized iron−porphyrin complexes,174 heme/Cu model com-
pyridine group was unlikely to interact directly as a proton plexes of Cc,O80,199,200 and similar to those reported for
relay with the active intermediates, although the pH-dependent sophisticated biosynthetic scaffolds.201 However, the role of
RRDE experiment showed that the protonated pyridyl group the pendant amines in enhancing the stability of FeIII−OOH
could significantly affect the selectivity for H2O vs H2O2. species against hydrolysis and subsequent PROS generation
However, it should be noted that the Co porphyrins was best evaluated under slow ET rates as established by
functionalized with para-substituted analogues always pro- Collman and Chidsey.125,202
duced higher H2O2 under similar experimental condi- In an RRDE experiment, the amount of PROS produced was
tions.23,195 The kinetics of ORR in aqueous medium (1 M quantified on the EPG working electrode as well as the Au
H2SO4) by 32B and 31A were also evaluated using the electrode surface attached to SAM of thiols having different
12398 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 34. (A) SERRS-RDE data of complex FeL3, 33D, over SAM-coated Ag electrode in the high-frequency region at oxidized (blue) and
steady-state conditions (red at −0.4 V, green at −0.5 V) in air-saturated pH 7 buffer under aerobic condition. The difference spectra from −0.5 to 0
V potential along with Lorentzian ν4 fits and Lorentzian ν2 fits showing different components are given in (B) and (C), respectively. All potentials
are expressed with respect to Ag/AgCl (saturated KCl) reference electrode.204

chain lengths resulting in differing the rate of ET from the (33C, 33D) having pendant pyridine and primary amine
electrode to the catalyst.123,125,203 As the kET was lowered by moieties. Two structural variants of 33E, o-
104−6 times from EPG (fast ET transfer) to C16−SAM ((slow monomethylguanidinotetraphenyliron(III)−porphyrin (Fe-
ET transfer flux), the percentage of ROS generated also MeMARG), and o-monophenylguanidinotetraphenyliron-
increased from 2% to 15% in these complexes (Table 4) which (III)−porphyrin (Fe-PhMARG) were synthesized, and their
indicated that the selectivity for 4e−/4H+ O2 reduction was reactivities were compared with 33E.198 The 1H NMR spectra
mostly retained even under slow ET conditions. Thus, the of the free ligand Me-MARG and Ph-MARG showed that the
selectivity exhibited by this series of complexes was in stark resonances for the methyl CH’s of the guanidine were at 1.35
contrast to all of the previous reports of mononuclear iron ppm and that for the aromatic CH’s of guanidine were at 5.8,
porphyrin complexes (e.g., FeTPP, FePf, FeEs4) without any 6.2, and 6.4 ppm respectively. This data demonstrated that
additional redox centers (e.g., distal Cu, Fc, or phenol), which both methyl and phenyl protons were shielded by the aromatic
produced about 50−100% PROS in C16−SAM (Table 4), porphyrin groups, which could only happen if the methyl and
which often led to rapid catalyst degradation.174,202 These iron phenyl groups were poised just on top of the porphyrin ring.
porphyrins were also chemically attached to the ImdC11SH- This phenomenon made the distal site much more hydro-
functionalized Au electrode using a mixed thiol SAM phobic, having an impact in the selectivity of 4H+/4e− oxygen
combining a linker ImdC11SH and a diluent C8SH (Imz reduction. The % PROS value for Fe-MeMARG and Fe-
SAM), where the imidazole terminal of the SAM acted as an PhMARG dropped to 7.9 ± 0.1% and 9.2 ± 0.4% over the
axial ligand to produce “push” effect to the attached catalyst C16−SAM-modified Au electrode, respectively, which further
during O2 activation.198 The RRDE data indicated that Imz dropped to 2.6 ± 0.1% and 1.3 ± 0.1% over the Imz SAM-
SAM bound 33D and 33E recorded only 2.9 ± 0.1 and 4.9 ± modified Au electrode, respectively (Table 4). Nonetheless,
0.1% PROS, respectively (Table 4), which were almost half of the in situ mechanistic investigation seemed necessary to shed
the PROS produced in their respective C8−SAM though the light into the differences in selectivity observed for the
ET rate (103 s−1) was the same in both cases. These data complexes 33D and 33E.
clearly suggested that the combination of “push” of imidazole In subsequent work, Dey et al. employed the previously
and “pull” of amine/guanidine functional group were very developed SERRS-RDE approach to find out the mechanism of
effective in allowing very selective ORR.204 These four ORR at neutral pH, catalyzed by the iron porphyrins
mononuclear porphyrin complexes were over 95% selective immobilized over the SAM-coated electrode.204 In the case
for 4e−/4H+ O2 reduction to H2O like the FeCu CcO models. of FeL3 (33D), i.e., the porphyrin with a distal amine group,
Despite having a suitably positioned H-bonding guanidine two intermediate species were accumulated during the catalytic
moiety in the distal superstructure as in HRP, Fe-MARG turnover while the resting state of the SERRS data over C8−
(33E) produced 18% PROS over C16SH-modified Au SAM showed the presence of a HS FeIII species (Figure 34A).
electrode which was higher relative to other porphyrins The difference spectra (data at −0.5 V potential, resting
12399 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 35. SERRS-RDE data of complex FeL3, 33D, over SAM-coated Ag electrode (A) in the low-frequency region at oxidized (blue) and at
steady-state conditions (red, at −0.4 V) in air-saturated pH 7 buffer. (B) Overlay of the difference spectra from −0.4 to 0 V in 16O2 and 18O2. (C)
Overlay of the difference spectra from −0.4 to 0 V in air-saturated pH 7 and pD 7 buffer, respectively. * represents the characteristic phosphate
buffer peak.204

Figure 36. (A) SERRS-RDE data of complex Fe-MARG, 33E, over C8-SAM in the high-frequency region at oxidized (blue at 0 V) and steady-state
conditions (red at −0.3 V, green at −0.5 V) in air-saturated pH 7 buffer under aerobic conditions. The difference spectra from −0.5 to 0 V
potential along with Lorentzian ν4 fits and ν2 fits showing different components are given in (B) and (C), respectively. * represents the plasma
line.204

oxidized) confirmed that the major species under steady state with ν4 and ν2 bands at 1366 and 1563 cm−1 (Figure 34B,C,
was a LS FeII species having a ν4 band at 1356 cm−1 and a ν2 brown line), respectively, was also observed. In the low
band (characteristic of the spin state marker) at 1547 cm−1 frequency region, a new band gained intensity at 824 cm−1 in
(Figure 34B,C, green line) and another minor LS FeIII species 16
O2 saturated pH 7 buffer during ORR (Figure 35A), and this
12400 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 37. SERRS-RDE data of complex Fe-MARG, 33E, over C8-SAM (A) in the low-frequency region at oxidized (blue) and at steady-state
conditions (red at −0.3 V, green at −0.5 V) in air-saturated pH 7 buffer. (B) Overlay of the data in 16O2- and 18O2-saturated buffer and (C) overlay
of the data in air-saturated pH 7 and pD 7 buffer, respectively. * represents the characteristic phosphate buffer peak.204

band was reproducibly shifted to 765 cm−1 in 18O2 saturated shift of 22 cm−1 was slightly less for 33E compared to that of
buffer and to 815 cm−1 in pD7 buffer (Figure 35B,C). The 824 the theoretical value (36 cm−1) owing to the fact that Fe−O
cm−1 vibrations, with ν16/18O2 = 59, were thus identified as the vibration in FeIV�O species is coupled to other out-of-plane
ν(O−O), originating from a LS FeIII−OOH species in motions of the porphyrin ring. Again, an 3 cm−1 H/D shift
agreement with the ν4 and ν2 bands from the high-frequency from 820 to 817 cm−1 was also observed for this ν(Fe−O) of
region. Unfortunately, a further low-frequency region of the 33E in pD 7 phosphate buffer (Figure 37C), which suggested
spectrum did not show the corresponding Fe−O vibration. that the FeIV�O unit might be hydrogen bonded to the
Judging by the intensities of the observed vibrations, it was pendant protonated guanidine of Fe-MARG. Therefore,
concluded that although LS−FeIII−OOH species were spectro-electrochemical analysis revealed that for 33E on
accumulated over the electrode under catalytic turnover for C8−SAM the LS FeIII−OOH intermediate was not observed at
33D, the population of the species was substantially lower all and only high valent FeIV�O and unreacted FeII species
compared to that of 17A and 17B. The low population of the were accumulated under steady state.
FeIII−OOH species likely jeopardized the identification of the The results obtained using the SERRS-RDE technique
weak ν(Fe−O) stretch in SERRS-RDE under steady-state clearly revealed that simple mononuclear iron porphyrins
conditions for 33D. showed the accumulation of an LS FeIII−OOH species, which
Alternatively, Fe-MARG (33E) showed substantially differ- in turn was prone to hydrolysis and subsequently produced
ent spectra containing a mixture of species in the SERRS-RDE more PROS during the ORR catalytic cycle.145,204 The
data under steady state when the resting state comprised only production of PROS swayed the selectivity toward 2e−/2H+
HS FeIII species (ν4 at 1362 cm−1 and ν2 at 1554 cm−1) ORR. On the contrary, the presence of pendant hydrogen-
(Figure 36A). The ν4 and ν2 marker bands in the difference bonding groups in iron porphyrins which stayed protonated at
spectra showed the development of a HS FeII species as the neutral pHs significantly reduced the population of the rate-
major component havinga ν4 band at 1346 cm−1 and ν2 band determining LS FeIII−OOH species under steady state for 33D
at 1542 cm−1 along with a minor population of high valent and eliminated the accumulation of LS FeIII−OOH species for
FeIV�O species with ν4 and ν2 bands at 1372 and 1569 cm−1, 33E. This phenomenon could be better explained based on the
respectively, under catalytic turnover conditions (Figure faster heterolytic cleavage of the O−O bond of FeIII−OOH
36B,C). In the low-frequency region, a gradual increase of an species which was directly correlated with the enhanced ORR
820 cm−1 vibration was observed when the cathodic potential rates and 4H+/4e− selectivity in these complexes.
was applied in an air-saturated pH 7 buffer (Figure 37A). Data Homogeneous ORR electrocatalysis were also investigated
collected in the 18O2-saturated pH 7 buffer indicated that the for 33B, 33C, and 33D in the presence of strong acid, [DMF-
820 cm−1 band in 16O2 was shifted to 798 cm−1 in 18O2 (Figure H+]OTf−, in O2-saturated DMF solution.197 The first-order
37B). This vibration was assigned as the ν(Fe−O) of a FeIV� rate constants of O2 reduction (TOF) were quantified using
O species, which was also suggested by the ν4 and ν2 bands at the foot-of-the-wave analysis, and those kcat values for 33C
1372 and 1569 cm−1. The experimentally observed ν16/18O2 (FeL2) and 33D (FeL3) were significantly higher at the same
12401 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 38. Optimized structures of six-coordinate low-spin (A) [FeIIITPP-OOH]; (B) [FeIIIL1-OOH]H+; (C) [FeIIIL2-OOH]H+; (D) [FeIIIL3-
OOH]H+; (E) [FeIIIMARG-OOH]H+ with hydroperoxide and water as axial ligands and protonated distal basic residues in PCM model
considering water as solvent. Color codes: carbon, gray; hydrogen, white; nitrogen, blue; oxygen, red; and iron, bluish gray.197

Figure 39. DFT-optimized structures of water-bound LS ferric hydroperoxide (FeIII−OOH) species for both FeL3 and Fe-MARG. (A) H-bonded
geometry with a distal oxygen atom is energetically more stable by 2.35 kcal/mol than the structure where H-bonding is primarily with the proximal
oxygen atom for FeL3. (B) H-bonded geometry where H-bonding is primarily with the proximal oxygen atom is energetically more stable by 1.70
kcal/mol than the structure where H-bonding is with the distal oxygen atom for Fe-MARG.204

ηORR with respect to the other catalysts (Table 5). Thus, the close to or higher than 7 and were expected to remain
iron porphyrins having distal basic groups (33C, 33D) slightly protonated at pH 7.0 buffer or in acidic DMF.194 The
deviated from the linear correlation of log(TOF) vs ηORR as theoretically determined absolute pKa values of these pendant
demonstrated by Mayer. Hence, the enhancement of rates at a bases of ferric hydroperoxide species in water were found to be
particular ηORR observed here reflected other factors than solely higher by 4−5 units than that of free bases because of strong
the primary coordination sphere. hydrogen-bonding interactions with the distal oxygen atom of
DFT calculations were performed to understand the effect of the bound superoxide or hydroperoxide anion.197 The
hydrogen-bonding interactions with the basic groups for the optimized structures of the Fe−O2 species with protonated
spectroscopically observed models of low spin ferric super- distal bases showed only slight elongation of the O−O bond
oxides (FeIII−O2) and ferric hydroperoxides (FeIII−OOH) lengths (>1.30 Å) due to H-bonding interactions with the
with water as axial ligand. These amine groups have pKa very distal oxygen atom. A significant elongation of O−O bond
12402 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 40. Molecular structures of the four atropisomers (A−D) of the [FeIII(o-TMA)]5+ cation.209

length (1.52 Å) was obtained for [FeIIIL1-OOH]H+ which is of the iron porphyrin complexes mimicking the distal basic
much lower compared to other complexes as the protonated residues in the enzyme active sites (Histidine, arginine and
tertiary amine residue is further away from the bound lysine) emulated the “pull effect”, which is the primary
hydroperoxide species relative to FeL2, FeL3, and Fe-MARG determinant in modulating the facile rates and high selectivity
(Figure 38). The hydrogen-bonding interactions from of ORR. However, the direct observation of O−O bond
protonated nitrogen bases to the distal oxygen atoms were heterolysis of a putative FeIII−OOH species was not directly
substantially strengthened for 38C, 38D, and 38E in the series, demonstrated (see section 5).
and O−O bond lengths were elongated to 1.88 and 1.86 Å in 3.5. Role of Electrostatic Interactions
[FeIIIL2-OOH]H+ and [FeIIIL2-OOH]H+, respectively. The
bond length elongation was maximum (1.90 Å) in the case of There has been a growing interest in including properly
Fe-MARG (38E) that appeared to highly activate the O−O oriented charged functional groups as electrostatic motifs in
bond of ferric hydroperoxide species the most for further the second coordination sphere to improve molecular
cleavage in the next step of ORR. This predicted faster O−O reactivity, catalysis, and electrocatalysis. Based on an earlier
bond heterolysis in 33E was consistent with the SERRS-RDE report of Saveant and Robert on CO2 reduction with highly
data where accumulation of FeIII−OOH species was not efficient electrocatalytic activity using positively charged iron
observed in the steady state unlike FeL3 (33D), during the porphyrins,205 Mayer et al. described electrocatalytic ORR
catalytic turnover.204 Two different possible H-bonding process using the same polycationic iron porphyrin, [FeIII(o-
orientations to the proximal and distal oxygen atoms of TMA)]5+ bearing four positively charged o-trimethylanilinium
bound hydroperoxide ligand were considered for the low-spin [N(CH3)3+] groups with buffered weak acids as an example of
FeIII−OOH structures in 38E. These calculations predicted a exceptional homogeneous electrocatalyst (Figure 40).206 The
ΔG of 1.7 kcal/mol at rt corresponding to an equilibrium ORR was investigated in CH3CN solvent using the iron
constant of around ∼17 favoring the orientation where the H- porphyrin catalyst in a series of buffers. The strongest acid
bonding was with the proximal oxygen of FeIII−OOH; i.e., the source, ([DMF-H+]OTf−; pKa = 6.1) produced TOFmax = 8.5
ratio of proximal H-bonding conformation and distal H- s−1 and ηORR = 1.16 V (Table 6). In contrast, acetic acid
bonding confirmation was ∼17 (Figure 39B). Alternatively, for (AcOH, pKa = 23.5) gave distinctly improved catalysis: a faster
38D, the energy difference was 2.35 kcal/mol, favoring the TOFmax (170 s−1) at less than half ηORR (0.54 V) which is ∼104
distal H-bonding conformation which transformed to an
equilibrium constant of ∼1/50 for the same ratio (Figure Table 6. Electrocatalytic ORR Parameters of Fe(o-TMA) in
39A). This was consistent with the hydrolysis of FeIII−OOH, Different Buffers206
leading to the formation of PROS under slow ET conditions
E1/2 (mV) vs ηORR TOFmax
for 33E even though it had a favorably disposed proton transfer buffer pKa Fc+/Fc0 (V) (s−1) log(TOFmax)
residue. On the other hand, 38D geometry stabilized the
none −0.295
orientation with H-bonding to the distal oxygen atom and thus
[DMF-H+]/ 6.1 −0.25 1.16 8.5 0.91
primarily allowed O−O heterolysis. Overall, the optimized DMF
geometries of all these iron porphyrin complexes showed that TFAH/TFA− 12.6 −0.349 0.88 3.2 0.51
the proton donor groups are oriented in a way which allowed [Lut-H+]/Lut 14.1 −0.23 0.68 0.07 −1.17
selective delivery of the protons needed to the distal oxygen SalOH/ 16.7 −0.536 0.82 12 1.08
atom of the hydroperoxide species, activating it for heterolytic SalO−
O−O bond cleavage (Figure 38). All the experimental results BzOH/BzO− 21.5 −0.653 0.67 63 1.80
confirmed that these pendant bases in the second coordination AcOH/AcO− 23.5 −0.651 0.54 170 2.23

12403 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

times faster at similar ηORR than any earlier reported molecular with and without an acetate ligand. The αβαβ isomer (40D)
catalyst for the homogeneous ORR.25,207 While using other was most reactive with the highest TOFmax, while the αααα
acids with the decreasing order of acid strength from (40A) had the lowest TOFmax and the other two isomers lay in
trifluoroacetic acid (TFAH), salicylic acid (SalOH), to benzoic between them.
acid (BzOH), similar improvements in both TOFmax and ηORR The four atropisomers showed a roughly linear relationship
were obtained (Table 6). This result contradicted previously between log(TOFmax) and ηORR. The most important factor
derived log(TOFmax)/ηeff relationships, which always predicted controlling the relative values of both TOFmax and ηORR for
that a lower ηORR might give a slower rate, as observed for these systems was the catalyst E1/2 of FeIII/II under electro-
another series of iron tetraphenylporphyrins.191 In due catalytic conditions. These relationships are likely to be
consideration of these results, it was concluded that a single directly linked to the O2 binding and the barrier for rate-
scaling relationship was not adequate for predicting the limiting proton transfer to the ferric superoxide species which
combined changes in log(TOFmax) vs ηORR plot and a vector have dependence on the catalyst E1/2. In general, catalysts with
sum of two different scaling relationship was developed.206 more negative reduction potentials (E1/2 values) are more
Mayer and his co-workers examined how covalently nucleophilic and have higher binding affinity toward O2 and
positioned charged groups affected substrate binding and result in more basic Fe−O2 intermediates, facilitating their
other key pre-equilibrium during ORR in another work by protonation (rds) and rapid turnover. The slope of the
comparing the αβαβ isomer of polycationic iron porphyrin log(TOFmax)/E1/2 relationship for the Fe(o-TMA) atro-
complex, [FeIII(o-TMA)]5+ ,with the analogous unsubstituted pisomers (34 ± 7 mV per decade) was nearly two times
complex, FeTPP.208 Developing electrostatic effects to control steeper than the previously reported slope (55 ± 2 mV per
chemical reactivity and catalysis might depend on a decade) for a series of substituted neutral iron porphyrin
fundamental understanding of how electrostatics impact the catalysts.191 The steeper slope indicated a more sensitive
thermodynamics of individual chemical steps in a reaction. relationship between the catalyst E1/2, the free energy of O2
Despite the fact that [FeIII(o-TMA)]5+ was found to be a binding, the distal oxygen atom basicity of the corresponding
remarkable molecular electrocatalyst for O2 reduction with superoxide intermediate, and the barrier to superoxide
high rates at low ηORR, the key electrostatic effect has no role protonation. The positioning of the charged groups had
for the O2 binding to the ferrous porphyrin but rather the pre- multiple effects and extended beyond one specific step of the
equilibrium binding of acetate (the anionic conjugate base of catalytic cycle and further characterization of the nature of
the buffer), is enhanced.206 The charged groups facilitated the these intermediates should shed further light on this. These
binding of an acetate ligand, which caused a cathodic shift in data highlighted the significance of net change in charge
the E1/2 of the acetate form, [FeII(o-TMA)(AcO)]3+. Acetate density upon ligand binding rather than orientation, on the
binding to the FeII(o-TMA)4+ was directly enhanced by thermodynamics and kinetics of multistep molecular electro-
electrostatic interactions with the cationic-porphyrin ligand, chemical transformation.
with its acetate binding constant (KAcO) being 65 times higher Warren and co-workers explored the strategies of using the
than that of the neutral FeIITPP under low ionic strength in pendant groups with secondary interactions that have been
absence of electrolyte. These results provided valuable utilized for iron porphyrin ORR electrocatalysts to improve the
references in the literature to identify and quantify the effects performance of the cobalt porphyrins. In this work, the authors
of intramolecular electrostatic interactions for ligand binding investigated electrochemical ORR of a series of Co porphyrins
and small molecule activation. with modifications in the 2-position of one of the phenyl
In a following work, Mayer and his co-workers studied groups containing −NH2, −N(CH3)2, and −N(CH3)3+ groups
electrocatalytic ORR using all the four possible atropisomers of (Figure 32D−F) to probe systematically the relative
Fe(o-TMA) (Figure 40) and determined how the positioning importance of proton relay effect and electrostatic interactions
of the tetra-cationic groups affected catalysis by identifying the of ancillary groups in catalysis.169 All these complexes were
similarities and differences that exist among the isomers.209 drop casted over the EPG electrode in order to perform pH
These charged [N(CH3)3+] groups were effective in facilitating dependence heterogeneous ORR, and they showed a definite
multiproton multielectron reactions like ORR that involved increase in current near the CoIII/II couple. The CV data of the
charge redistribution or high-energy charged intermediates or CoIII/II couple for Co porphyrins (32D−32F) showed pH
transition states in molecular electrocatalysis. The E1/2 values dependence that might be attributed to the protonation/
for each of the FeIII/FeII redox couples of Fe(o-TMA) deprotonation of axially bound water. In addition, the
atropisomers were significantly positive than the same reported protonation state of CoTPPNH2 (32D) and CoTPPNMe2
for neutral iron tetraphenylporphyrins.210 The electrochemical (32E) depended on the pH at which the experiments were
oxygen reduction by the four atropisomers were examined conducted. Considering the pKa values of aniline (pKa = 4.6)
under the similar homogeneous conditions (CH3CN contain- and N,N-dimethylaniline (pKa = 5.2) in water as models, 32D
ing 0.1 M H2O, 1:1 buffered acetic acid/acetate (AcOH/ and 32E were expected to remain fully protonated at pH 0 and
AcO−) that was employed for the mixture of the isomers deprotonated at pH 7. However, the trimethylanilinium group
producing the best catalysis data. All of these four isomers in CoTPPNMe3+ (32F) being positively charged, was
exhibited reversible FeIII/II redox processes under anaerobic independent of the pH in the medium. Investigations using
conditions with the cathodically shifted E1/2 values between CV and RRDE showed that the presence of a cationic −
−0.595 V (αααα) and −0.644 V (αβαβ) in the presence of N(CH3)3+ group proximal to the Co center in 32F, improved
AcOH/AcO− buffer, which were replaced by a large selectivity for the 4H+/4e− reduction of O2 was observed
irreversible current in the presence of O2 indicating catalytic across a wide pH range from pH= 0 to 7 (Table 3). In contrast,
turnover. The isomers catalyzed O2 reduction with faster rates 32D and 32F were more selective for reduction of O2 to H2O
at low ηORR, and their catalytic rate constants (TOFmax) range at low pH when the pendant nitrogenous bases were
from 1 to 60 s−1 with the ηeff values differed only by 50 mV protonated, but the production of H2O2 was increased at
12404 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 41. Crystal structures and computer models of biosynthetic CcO mimics. (A) Overlay of the crystal structures of native bovine CcO (black)
and designed E-CuBMb (cyan); (B) overlay of the crystal structures of bovine CcO (black) and E-F33Y-CuBMb (cyan); (C) overlay of the
computer model (yellow) and crystal structure (cyan) of E-F33Y-CuBMb; (D) overlay of the E-CuBMb crystal structure (cyan) and the computer
model of E-G65Y-CuBMb (yellow).218,219 Reproduced with permission from ref 219. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KGaA.

higher pH. In particular, at pH 0, the selectivity of CoTPPy binding site in the distal pocket that closely mimics the
(32B, pKa of pyridine = 5.3) was quite similar to those of binuclear center in CcO, and this was the first reported
CoTPPNH2 (32D) and CoTPPNMe2 (32E) because of the protein-based model for CcO (Figure 41).216 The reversible
comparable pKa values (∼5) of all the proton relays, despite oxygen-binding property of WTswMb is well-known, and the
the different distances between the ionizable nitrogen and the incorporation of the CuB binding site in swMb provided ample
metal. The Co porphyrin analogue without any secondary scope to investigate the effect of copper ion on O 2
effects, CoTPP (32A), catalyzed primarily the 2e− reduction of reduction.213,217 Exposure of the CuII-free reduced CuBMb to
O2 at all pH, consistent with the other reports.23,195 The O2 resulted in partial conversion to the Fe−O2 adduct. The
second-order rate constants of O2 reduction for CoTPPNMe3+ mutations introducing two histidine residues in the distal
(32F) were determined to be pH-independent and also pocket with no bound metal ions, reduced the O2 binding
significantly high (∼106 M−1 s−1) throughout the entire pH affinity of the protein compared to the native WTswMb.
range at low ηORR. It was proposed that instead of hydrogen Binding of AgI and ZnII to the three histidine distal sites,
bonding, electrostatic stabilization of anionic intermediates by increased the O2 binding affinity. These results strongly
conjugate acids of nitrogen bases, or trimethylanilinium suggested that the presence of CuB center is important for O2
(cationic) groups was important in determining selectivity binding forming a stable Fe−O2 adduct and increases the O2
for the Co porphyrin ORR catalysts. binding affinity of the protein, albeit with a lower affinity than
the wild-type protein.213
4. BIOSYNTHETIC MODELS OF CCO WTMb is known to reduce O2 upon exposure to air in the
presence of a reductant and induce heme degradation to
Using a biosynthetic approach, stable naturally occurring
proteins were used as scaffolds for creating mimics of several verdoheme and then FeIII−biliverdin-like heme oxygenase. It
metalloenzymes. There are few examples where biosynthetic was shown previously that WTMb is incapable of O2 reduction
models that structurally and functionally mimic CcO have to H2O in the presence of subequivalent amounts of catalase
been reported.29,211 despite the addition of excess CuII ions and reductants like
ascorbate and TMDA (N,N,N′,N′-tetramethyl-p-phenylenedi-
4.1. Role of Distal Cu amine), implying exogenous peroxides must be involved in
It has been difficult to obtain and handle large membrane- coupled oxidation unlike heme oxygenase (HO). To further
bound proteins, HCO for practical O2 reduction. On the other probe the role of CuB center during O2 reduction, the reaction
hand biosynthetic models of HCO’s created via site directed of reduced CuBMb was followed with UV−vis absorption
mutagenesis in myoglobin proteins are easier to handle. kinetics. The spectral changes suggested that the CuB ion in
Myoglobin being a small globular protein that serves as an O2 CuBMb facilitated the reduction and then converted the heme
storage in nature, is easy to obtain in large quanitities.212−214 Yi in CuBMb to verdoheme whereas oxy-WTMb did not show
Lu and his co-workers designed and engineered copper any spectral changes. No O2 reduction was observed in other
binding sites in the active site of cytochrome c peroxidase215 control experiments, e.g., either the CuB metal was absent in
and sperm whale myoglobin (swMb).216 The CuBMb mutant CuBMb mutant or the position was occupied by redox-inactive
of swMb was created by introducing two non-native histidine ZnII or AgI metal. The CuBMb would likely to go through the
residues through Leu29His and Phe43His mutations which peroxy-heme intermediate during O2 reduction similar to HO
along with the native distal His64 residue formed the Cu- as suggested by similarities in the reaction products.213 WTMb
12405 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

only reversibly binds to dioxygen without reducing it and site of CuBMb to understand the role of the Tyr244 residue for
P450, HO shows 2e− reduction of O2, whereas HCO reduces O2 reduction in the CcO-active site and examined their
oxygen by 4e− to water. It was proposed that the first step in reactivities toward O2 reduction.219 The absorption spectros-
the catalytic cycle of HCO, P450, OH, and CuBMb mutant of copy was measured for both E-F33Y-CuBMb and E-
swMb involves the formation of two-electron reduction of G65YCuBMb in the presence of excess reductant [ascorbate
oxygen to peroxide intermediate (Scheme 7).46,212,220,221 Thus, with TMPD as a mediator] in air saturated buffer solutions,
and the results suggested reduction of O2. The rates of O2
Scheme 7. Proposed Reaction Mechanism of CuBMb with reduction were measured quantitatively using an O2 electrode
O2213 in the presence of reductant to monitor the concentration of
O2 over time, following similar methods reported for native
HCO.223 To identify the reduced product (O2−, O22−, or
H2O), superoxide dismutase (SOD) and catalase were added
which selectively reacted with superoxide and peroxide,
respectively, producing O2 as one of their products. By
comparing the rates of O2 consumption in the absence and
presence of SOD and catalase, the portion of O2 reduction that
is due to water formation versus superoxide or peroxide
formation could be calculated (see Figure 42).
Not surprisingly, most of the O2 consumption by WTswMb
was due to PROS formation, consistent with autoxidation, and
these rates remained unaffected in the presence of Zn2+, Ag+,
or Cu2+ ions. Again, introducing two histidine residues
(CuBMb) into the distal pocket of WTswMb resulted in
substantial inhibition of superoxide/peroxide formation
relative to WTswMb but did not significantly contribute to
interestingly, these enzymes perform different reactions even water formation and resulted in a decrease in the overall rate of
after going through the same putative peroxy-heme inter- O2 consumption. In contrast, introducing a tyrosine next to the
mediate. Introducing the Cu-binding site into myoglobin one of the histidine ligands at either position 33 or 65 in F33Y-
transformed it from a simple oxygen carrier into a copper CuBMb and G65Y-CuBMb, respectively, resulted in an increase
dependent heme oxygenase at pH 8,213,214,222 which degraded in water formation and overall rate of O2 consumption. The
the heme cofactor to verdoheme. Since one of the key features rate of O2 reduction to water by E-G65Y-CuBMb was
in the catalytic active site of native CcO is the presence of measured to be 28 min−1, which is about 150-fold of native
tyrosine residue covalently linked to the histidine ligand bound HCOs.224 The mutants E-F33Y-CuBMb and E-G65Y-CuBMb
to the CuB center,57 the lack of it might have been detrimental attained more than 505 and 1056 turnovers, respectively. The
to the reactivity of these biosynthetic models. Furthermore, crystal structures of E-CuBMb, E-F33Y-CuBMb, and WTswMb
proton delivery coupled to the O2 reduction is necessary to indicated that the distal pocket containing two histidine
avoid a HO reaction pathway that would lead to heme residues and an additional tyrosine residue had the capability
degradation and loss of HCO activity. of stabilizing two and three water molecules in the distal site,
4.2. Role of Tyrosine Residue respectively (Figure 43), in comparison to one water molecule
Lu et. al reported two mutants, F33Y-CuBMb and G65Y- in the distal pocket of WTswMb. Hence, WTswMb has only
CuBMb (Figure 42), introducing a tyrosine residue at two one hydrogen bond donor site (H64) available to interact with
different positions close to the histidines of the copper-binding the bound oxygen, and two and three more hydrogen-bond-
donating residues are available in E-CuBMb and E-F33Y-
CuBMb, respectively.219 On the basis of the results, the authors
proposed that the tyrosine and its associated hydrogen
bonding network activated the ferric-superoxo species to
accept another electron forming ferric peroxide which
underwent the PCET process to produce a ferryl intermediate
for the complete reduction.
The formation of a Tyr• was directly observed in the EPR
spectroscopy during the reactions of 1 equiv of H2O2- and O2-
saturated buffer with the oxidized and reduced protein,
respectively, providing a strong evidence for the active role
of a tyrosine residue in the enzymatic reaction mechanism.225
The EPR data, both X and Q-bands and hyperfine splitting on
an anisotropic radical signal at g ∼ 2, matched with those of the
Tyr244 radical in HCO.129,226 This Tyr•, being transient and
unstable, disappeared completely within ∼100 ms, which is
Figure 42. Oxygen consumption rates in the presence of EDTA and why it difficult to obtain under turnover conditions.
varying amounts of copper in the absence (blue) or presence (red) of In a subsequent work, the detailed EPR and X-ray
catalase and SOD. Error bars indicate standard deviation. Reproduced crystallographic data revealed that an extensive H-bonding
with permission from ref 219. Copyright 2012 Wiley-VCH Verlag network existed in the active site of oxy-F33YCuBMb that
GmbH & Co. KGaA. helped in promoting proton delivery to the active site and
12406 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 43. Water networks in the distal pockets of (A) E-F33Y-CuBMb and (B) E-CuBMb crystal structures. Reproduced with permission from ref
219. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KGaA.

Figure 44. (A) Proposed mechanism of O2 reduction by F33YCuBMb in comparison with WTMb. (B) Crystal structure of the oxy-F33Y-CuBMb
(1.27 Å resolution; PDB ID: 5HAV) with two water molecules in the active site pocket (left). Cryo-reduction method for formation and trapping
of (hydro)peroxo intermediates (X = active site His or Tyr residue) (center). The EPR spectra of oxy-F33Y-CuBMb cryoreduced at 77 K showing a
ferric-hydroperoxo (H-bonded to an active site water, green) which decays upon stepwise annealing (right). Reproduced from ref 227. Copyright
2016 American Chemical Society.

facilitated the reduction process, thus inducing oxidase activity resolution (1.27 Å) X-ray crystallographic data were obtained
in a biochemical model of HCO.227 It had been already for the oxy-F33Y-CuBMb showing a H-bond network of two
established that cryo-reduction of the parent Fe−O2 complex water molecules linking the two engineered residues (Tyr33
at 77 K followed by stepwise annealing revealed the and His29) to the terminal O atom of the Fe−O2 species
intermediates formed during O2 reduction pathway in heme (Figure 44B, left). These were not present in the crystal
proteins. 228−232 Oxy-F33YCu B Mb upon cryo-reduction structure of oxy-WTMb234 which was incapable of O2
showed two rhombic EPR signals corresponding to the ferric reduction. Altogether, the spectroscopic and crystallographic
peroxo (g = 2.24, 2.13, and 1.95) and ferric hydroperoxo (g = evidence suggested the importance of H-bonding interactions
2.29, 2.17, and 1.96) intermediates, and these species were that would transform a simple O2 carrier protein, WTMb to
again confirmed by the 1H ENDOR measurements. Then the F33Y-CuBMb which could activate O2 and reduced it to H2O
decomposition of the hydroperoxo species at high temperature (Figure 44A), and this was the first successful design of
resulted in O−O bond cleavage and an EPR-silent ferryl artificial enzyme system known so far.
species (compound II), and this intermediate underwent Similarly, Karlin and Solomon’s combined work based on
further cryoreduction at 77 K generating the prominent low- the synthetic heme−peroxo−copper models also supported the
spin rhombic EPR signal (g = 2.52, 2.15, and 1.90), assigned as role of a tyrosine residue.137,138 In these systems, introduction
ferric hydroxo species (Figure 44B, right). In contrast, the of external phenol formed a stable adduct with the bridging
cryoreduced oxy-WTMb was unable to deliver a proton and peroxo complex which mimicked the active site structure of
showed a dominant rhombic signal (g = 2.22, 2.13 and 1.96) HCOs. From the kinetic study and theoretical calculation, it
which was identified to be a ferric-peroxo species.232,233 High- was evident that PhOH played an important role for the H-
12407 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Scheme 8. Construction of Electrode-Bearing Biosynthetic Modelsa

a
Reproduced with permission from ref 201. Copyright 2018 American Chemical Society.

bond assisted homolytic O−O bond cleavage via a PCET could be directly transferred to the haem from the electrode to
mechanism. The energy barrier for the reductive O−O bond facilitate O2 reduction.237
cleavage decreased in the presence of either H-bonding In air-saturated buffer solutions, the bioinspired electrode
interactions or proton transfer from the phenolic O−H bond. loaded with the Mb mutant G65YCuBMb showed a large
In a recent study, Dey et. al demonstrated the role of a electrocatalytic O 2 reduction current at pH 7. The
redox-active quinol group in ORR mechanism using synthetic G65YCuBMb protein predominantly catalyzed a 4e−/4H+
model porphyrin.151 They reported that even a simple ferrous reduction of O2 to H2O as RRDE data indicated the formation
porphyrin complex (27A) containing a pendant quinol group of only 6% PROS during ORR. The G65YCuBMb biosynthetic
showed complete O2 reduction to water in terms of oxidation Mb scaffold-based bioelectrode was found to be remarkably
of quinol to quinone in the absence of any other additional stable for the ORR.236 Any SAM-bearing covalently attached
electron and proton source. The ferric superoxide species of O2-reducing electrocatalysts reported so far have not been
27A underwent the first HAT resulting in a ferric peroxide stable enough to allow these dynamic electrochemical
which could perform another HAT from the semiquinone experiments to determine the kinetic parameters. The kcat
group and generated ferryl species along with the residual was evaluated to be 1.98 × 107 M−1 s−1, yielding a TON in the
quinone group. Therefore, these results might contribute to order of 104 from RDE data and their K-L analysis (Table 7).
establish the importance of redox-active H-bond donor sites in
model systems like the residues present in enzyme active sites. Table 7. Biosynthetic CcO Models as ORR Catalysts199,201
In the F33Y-CuBMb model, the presence of tyrosine and its
kcat (× 107) TON PROS
specific positioning had been shown to play an important role catalysts (M−1 s−1) (104) (%) SKIE
in directing the product formation from the complete
CuBMb ND ND 6 ± 1 ND
reduction of O2 with the minimum release of PROS similar G65YCuBMb 1.98 × 107 >0.65 6 ± 1 16.00 ± 1.0
to the activity of terminal oxidases, with more than 1000 V68ECuBMb 2.2 × 107 >2.8 4 ± 1 4.30 ± 0.5
turnovers. Although the designed proteins were less active than V68E/I107E CuBMb 0.86 × 107 >0.5 7 ± 1 2.43 ± 0.07
terminal oxidases, it is noteworthy that oxygen reduction to synthetic CcO model 1.2 × 10−2 ND ∼8 ND
water was achieved by a much smaller protein, myoglobin, with (FeCu analogue)
only three mutations of the distal pocket.219,235 The rate of
electron transfer to the active site remained a major issue in
achieving efficient catalysis in these systems. The catalytic ORR rate by G65YCuBMb exceeded those
reported for the best artificial synthetic analogues as well as
4.3. Electrocatalytic ORR by Biosynthetic Models of CcO
native CcO itself. Efforts on attaching native CcO on electrode
To address the slow ET rates, the biosynthetic model of CcO similar to bioelectrodes resulted in very sluggish O2 reduction
containing the distal CuB and tyrosine residue was immobilized due to improper alignment of this membrane-bound protein
on a bioinspired electrode developed by Dey and co-workers on the electrode, which prevented efficient electron transfer to
following their earlier method employed for wTMb.236 The the active site.238,239 As a result, a CcO-functionalized SAM-
electrode bearing the biosynthetic model, G65YCuBMb, was covered Au electrode showed <1 μA electrochemical O2
constructed in situ where the native haem cofactor in Mb was reduction current at −300 mV, whereas this bioelectrode
replaced by a modified hemin cofactor bearing an alkyne group produced ∼100 μA current at similar potentials.240
(hemin-yne, Scheme 8) and reconstitution of apoprotein was To understand the mechanism of facile and selective O2
done in situ with the hemin-yne group tethered to the reduction catalyzed by the G65YCuBMb biochemical model,
functionalized SAM of thiols on Au electrode so that electrons the SERRS-RDE technique was applied. With the help of this
12408 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Scheme 9. Comparison between the Mechanism of Native CcO in Solution and the Biosynthetic G65YCuBMb Model on the
Electrode236

technique, in situ spectroscopic detection of reactive resting high spin ferric state and was likely to be the reduction
intermediates can be done over the electrode surface during of the FeIII−OH species. Thus, immobilized of an electrode,
electrocatalytic ORR under steady state. When a cathodic one could circumvent the kinetic barrier associated with the
potential was applied in an oxygenated buffer, the SERRS-RDE dissociation of the hydroxide by directly reducing it to ferrous
data clearly showed the characteristic features of low spin heme at more negative potentials. This direct ET to the ferric
and ferryl species. The lack of signals of high spin ferrous and hydroxide species, which is an intermediate in the catalytic
high spin ferric species indicated that the O2 binding was more cycle of CcO, is an ET shunt analogous to peroxide shunt in
facile for this mutant having the design principle of Mb which cytochrome P450. Thus, the rds of ORR by these biosynthetic
had 10 times faster binding rate than native CcO and the ET to models was expected to be the protonation of the FeIII−O22−
FeIII resting state was also very facile at these potentials as may species with a first-order rate of 5000 s−1 (Scheme 9). Here,
be expected due to direct attachment of the Hemin-yne to the too, the tyrosine residue present in the biosynthetic model of
electrode. The clear increase in intensities of bands having ν3 CcO (Y65 in G65YCuBMb mutant) could help in both proton
and ν2 at 1508 and 1591 cm−1, respectively, and 18O2-sensitive and electron transfer during ORR and substantially reduced
bands in the low-frequency region corresponding to the ferryl the amount of PROS relative to CuBMb, which is analogous to
species (FeIV�O) entailed the O−O bond cleavage faster for G65YCuBMb without the Y65 residue and resulted in higher
the formation than its decay under steady state. A Gedanken TONs.
steady-state turnover experiment with CcO, where the electron The oxygen-reducing enzyme CcO and nitric oxide
transfer to the active site was very fast (that is, in the reductase (NOR) both belong to the HCO superfam-
hypothetical situation where ET from Cyt c to CcO is not the ily.4,244,245 In spite of having similarities in the active-site
rds), the species that would accumulate during turnover, based structures, they have distinct reactivities likely due to the
on the Babkock−Wikstrő m mechanism were the FeII, FeII−O2, differences in distal metal and/or the key second sphere
FeIII−O22−, FeIII−OOH, FeIV�O, and FeIII−OH for which the residues.216,219,246,247 For example, the CcO active site
rates of formation of these species were greater than their rates contains a CuB and a conserved and cross-linked Tyr244
of decay (Scheme 9).241,242 Out of these, the FeII−O2, FeIII− residue while the active site of NOR has FeB and two
O22−, and FeIII−OOH species would have rR signatures of low conserved glutamate residues. Lu and co-workers successfully
spin heme, FeIII−OH would have rR signature of high spin designed and engineered the active site of NOR where FeB was
heme and the FeIV�O would have signature unique to heme introduced in the distal site instead of CuB and the Val68
ferryl species.4,242,243 The in situ SERRS-RDE data were residue of Mb was mutated with Glutamate residue resulting in
consistent with such an expectation. the mutant V68ECuBMb.247 Introduction of another glutamate
The overall rate-limiting step of native CcO in solution is the residue mutating the I1e107 residue in the distal side of heme
dissociation of hydroxide from FeIII−OH, which is the end in Mb resulted in a mutant I107ECuBMb.246 In order to create
product of O2 reduction reaction, to get back to the active myoglobin based perturbed models of CcO, the FeB metals in
resting ferric state with a first-order rate constant of 500 the distal site of V68ECuBMb and V68E/I107ECuBMb
s−1.91,242,243 The potential determining step of ORR by biosynthetic models of NOR were replaced with Cu and
G65YCuBMb was 166 mV more negative than E1/2 for the proton transfer glutamate residues were introduced instead of
12409 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

the redox active tyrosine residues. The glutamate residue Scheme 10. Comparison between the Mechanisms of
containing mutants provided a unique opportunity to probe Oxygen Reduction of the Two Biosynthetic Models while
the potential role of proton transfer in the O2 reduction Attached on Top of the Electrodea
process in general and specifically in CcO.
The electrochemical O2 reduction was investigated for these
two mutants by Dey et al. with a combination of H2O/D2O
solvent kinetic isotope effect and in situ SERRS-RDE to unveil
the main role played by the local acidic residues in determining
the selectivity and kinetics of O2 reduction as well as in
modulating the rds of 4e−/4H+ ORR.201 The heterogeneous
ORR were performed by reconstitution of apoproteins of the
mutants to the hemin−yne-modified functionalized SAM
surfaces like G65YCuBMb. The LSV data of these electrodes
bearing mutants showed large irreversible electrocatalytic O2
reduction current in air saturated pH 7 buffer. The kinetics of
ORR and selectivity were determined with RDE and K-L
analysis and thereby, the biosynthetic models, (V68ECuBMb
and V68E/I107ECuBMb) exhibited second-order rate con-
stants of >107 M−1 s−1 as high as for the biosynthetic CcO
model G65YCuBMb (Table 7). For V68E/I107ECuBMb, the
slightly lower value of the kcat was due to the presence of 37%
LS FeIII component which was unable to take part in ORR.
The presence of the additional proton donor residues near the
active site did not compromise the selectivity of 4e−/4H+ ORR
as these V68ECuBMb and V68E/I107ECuBMb mutants
produced minimum amount (4−7%) of PROS. The kcat
obtained in deuterated buffer allowed estimation of H2O/
D2O solvent kinetic isotope effect (SKIE) which, in turn,
reflected the involvement of proton transfer in the rds. The
SKIE of O2 reduction for G65YCuBMb was 16.0 which was
reduced to 4.3 in V68ECuBMb (Table 7). Thus, the presence
of proton transfer glutamate residue in V68ECuBMb reduced
a
the SKIE. The introduction of an additional glutamate residue Reproduced from ref 201. Copyright 2018 American Chemical
in V68E/I107ECuBMb mutant further reduced the SKIE to Society.
∼2.4. The loss of preorganized proton transfer channel due to
mutation of proton transfer residues was also proposed to be
responsible for high SKIE in heme enzymes like Cytochrome protonation steps in V68ECuBMb relative to G65YCuBMb.
P450 and peroxidases.248,249 Hence, the large H/D isotope Thus, the distal glutamate residues in the biosynthetic models,
effect ∼16 associated with G65YCuBMb indicated a rate- V68ECuBMb and V68EI107ECuBMb, behaved as proton
determining protonation step and the lack of an efficient transfer residues similar to those present in natural enzyme
proton transfer channel to the active site. resulting in facile and selective 4e−/4H+ ORR and also lowered
For G65YCuBMb mutant, the protonation of LS ferric the H2O/D2O SKIE value of kcat.201
peroxide was likely to be the rds as indicated by the 4.4. Role of Distal Arginine Residue in Amylin Peptide
accumulation of LS ferric peroxide species in the SERRS- Dey et al. assembled naturally occurring amylin (Ay) peptide
RDE data and was also supported by the large SKIE. The on top of Au electrodes with thiol groups of two cysteine
detailed electrochemical study of ORR for all of the mutants residues (Cys2 and Cys7) present in its 1−19 domain (total 37
showed that they reduce oxygen at comparable rates with the amino acid residues are present) to form a stable AySAM
only variation in their H2O/D2O SKIE values (2.4−16) (Table assembly.250 The assembly was bound to heme with the
7). However, the same rds for G65YCuBMb, V68ECuBMb, and His18 residue present in the 1−19 domain of amylin to
V68E/I107ECuBMb having zero, one, and two preorganized construct a potential biochemical scaffold which was shown to
proton transfer residues, respectively, would result in a be an excellent O2 reduction catalyst by Dey and co-workers.
variation in the kcat which was not the case suggesting that Wild-type (WT) heme-AySAM and its Arg11ASN mutant
the rate-determining steps were different depending on the (AySAMR11N) bound heme assembly were at first fully
nature of the second-sphere residues of these mutants. On the characterized using CV, SERRS, and atomic force microscopy
basis of the electrochemical and SERRS-RDE data, the (AFM) spectroscopy. Formation of a monolayer of site-
mechanism of ORR by the mutants other than G65YCuBMb isolated heme−AySAM active sites was concluded from low
also had been proposed. The SERRS-RRDE data of surface coverage value (9 × 1012 molecules/cm2) in CV data,
V68ECuBMb, the construct with the highest rate, showed and heme binding was ensured from the SERRS data,
only HS FeIII species under steady state, and neither a LS FeIII indicating the presence of a six coordinate FeIII−HS species
peroxide nor a FeIV�O species were observed unlike in on the surface. RDE data with the above-mentioned
G65YCuBMb (Scheme 10). Since, decay of both species bioinspired electrode systems and their subsequent K−L
required protons and the fact that these species did not analysis revealed that both heme−AySAMWT and AySAMR11N
accumulate during steady state was consistent with faster followed selective 4H+/4e− O2 reduction pathway with kcat
12410 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

values (1.8 ± 0.6) × 107 M−1 s−1 and (2.7 ± 0.4) × 105 M−1 [MnIII(TMP)OH] reacted the same way with peroxyacids to
s−1, respectively. RRDE was applied to detect PROS formation generate a (acylperoxo)MnIII TMP adduct followed by the
for these two systems, which is about 9 ± 0.5% for heme− formation of the oxo-manganese(V) intermediate which was
AySAMWT and 6.6 ± 1% for the R11N mutant. This particular well characterized.277,278 It was suggested that the nature as
construct with Ay attached to the Au electrode via two Cys− well as rate of the O−O bond cleavage would have been
Au linkages formed a very stable monolayer; otherwise, the kcat affected upon changing the substituents in peroxyacids. In
value could not be calculated for monolayers of catalysts. A order to understand the “push effect” on the O−O bond
facile H+ transfer pathway from the Arg11 residue was cleavage reaction of the FeIII−OOH species, Watanabe et al.
operative, which was the key factor behind higher kcat in examined the reactivity of a series of (acylperoxo)iron(III)−
ORR by the WT Ay compared to R11N mutant. The porphyrins by changing substituents at the meso positions of
difference in PROS was also due to the presence of the acidic the porphyrin ring (Figure 45) as well as introducing an
Arg11 residue, which although not directly involved in bonding
with heme provided an H-bonding interaction from the second
sphere. In this report, naturally occurring stable peptide
scaffold with heme binding was shown to be 100 times more
facile in ORR compared to simple mononuclear synthetic
mimics.

5. ROLE OF SECOND-SPHERE DISTAL RESIDUES IN


PEROXIDE ACTIVATION
The oxygen reduction cycle involves the formation of a ferric
hydroperoxide species (FeIII−OOH), also known as com-
pound 0, which is generally a very short-lived intermedi-
ate.251,252 It is formed by two-electron reduction of molecular
O2 followed by protonation in many heme metalloenzymes like Figure 45. Structures of meso-phenyl-substituted iron porphyrin
complexes (A−E) employed in this study.280
Cyt P450 and nitric oxide synthase. The same intermediate can
be generated by adding H2O2 directly to the resting ferric state
of heme enzymes.253 The compound 0 intermediate plays a imidiazole axial ligand.279,280 It was demonstrated that the O−
crucial role for the formation of high valent iron(IV)−oxo O bond cleavage mechanism can be homolytic or heterolytic
porphyrin cation radical [(FeIV�O)Por*+], commonly known and it could be controlled by changing the solvent polarity
as compound I, intermediate which is the active oxidant in from CH2Cl2 (heterolysis) to toluene (homolysis).272,276,281
monooxygenase, peroxygenase and peroxidases.253−255 The Generally the O−O bond cleavage step is the rds in these
formation of compound I from compound 0 involves the reactions and the first-order rate constants (khetero) for the
heterolytic cleavage of the O−O bond in this FeIII−OOH heterolytic O−O bond cleavage of these substituted porphyr-
species. Factors that can affect heterolytic O−O bond cleavage ins were evaluated (Table 9). The complexes having electron-
of metal hydroperoxide species have been, thus, of great donating substituents at the meso positions of a porphyrin ring
interest.256−259 Investigations using monofunctional perox- were found to enhance the rates of O−O bond cleavage
idases, catalases, cyt P450 monooxygenases, and peroxygenases consistent with the “push effect” exerted by the porphyrin
and their site-directed mutants revealed the importance of both ligand, whereas electron-withdrawing groups decelerated the
first- and second-coordination sphere residues in the active-site formation of compound I. In addition to that the nature of
structures that are essential for the generation of high-valent substituents on the peroxy acids affected the rate of heterolysis
catalytic intermediates via heterolysis of the O−O bond and the (acylperoxo)iron(III)−porphyrin complexes decom-
present in compound 0.47,55,260−264 The generation of posed faster for the p-nitroperbenzoate derivatives than the m-
compound 0 and the subsequent O−O bond cleavage to chloro derivatives. The most electron rich iron(III)-porphyrin
form compound I in peroxidases are facilitated by residues at (45A) exhibited the highest khetero values among all other
the distal pocket acting as acid−base catalysts, as proposed first substituted analogous porphyrins using both the peroxy acids
by Poulos and Kraut based on structural and mechanistic under similar experimental conditions. These investigations
investigations.256,265−267 Similarly, the O−O bond heterolysis laid the foundation of later work in the area where second-
of compound 0 in cytP450 is aided by a threonine residue in sphere residues were introduced.
the distal site directing the proton necessary to the FeIII−OOH In order to understand the O−O bond activation, Nocera
unit which is further activated by the “push” effect of the and his co-workers constructed a class of metal-porphyrin
electron rich axial thiolate ligation.37,268−270 Over the last complexes, hangman porphyrins, where −COOH functional
several decades, several synthetic metal-porphyrin complexes groups were positioned above the redox-active metal center of
were designed for the investigation of the “push-pull” effect iron porphyrins for the facile proton delivery to the metal-
during the O−O bond cleavage followed by the formation of bound ligand in a face-to-face arrangement within a single
compound I using various oxidants.271−275 molecular architecture.184,282 The hanging carboxylic acid or
Groves and Watanabe first reported the reaction of ester group was included on a rigid xanthene (HPX) or
chloro(5,10,15,20-tetramesitylporphyrinato)iron(III) dibenzofuran (HPD) scaffold at the meso position of the
[FeIII(TMP)(Cl)] with peroxyacids in organic medium trimesitylporphyrin platform to develop a set of HPX and HPD
(CH2Cl2) at low temperature and monitored O−O bond complexes containing transition metals (Fe, Mn) (Figure
cleavage step by direct observation of the conversion of 46A−K).283
acylperoxoiron(III) porphyrins to form high-valent spe- The likely role of protonation assisted O−O bond activation
cies. 272,276 The corresponding manganese analogue was examined with a set of complexes where both the proton
12411 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

donor ability (carboxylic acid versus ester) and orientation


(xanthene versus dibenzoufuran) of the pendant distal
functionality was varied. These iron hangman porphyrins
were evaluated for, catalase like, disproportionation of H2O2
under biphasic buffered condition (CH2Cl2/H2O, pH 7) in the
presence of 1,5-dicyclohexylimidazole, axial ligand to the metal,
at 25 °C. A simple porphyrin analogue, FeIII(TMP)(Cl) (45C)
having similar steric effect and electronic structure due to the
presence of mesityl groups, was used as a control. The complex
46B, containing a xanthene spacer and carboxylic acid group,
showed the highest activity toward H2O2 disproportionation
with 10−100 times higher TON obtained for O2 production
than any other catalysts examined (46C, 46J, and 45C). The
catalytic activity decreased significantly by replacing xanthene
with dibenzofuran spacer with the pendant carboxylic acid
functionality remaining the same or by replacing only the
carboxylic acid with a methyl ester keeping the xanthene
scaffold intact. The addition of 1 equiv of either benzoic acid
or methyl benzoate externally to 45C, as a source of
intermolecular proton, failed to induce catalytic catalase
Figure 46. Hangman porphyrins based on xanthene (A−I) and activity. The catalytic epoxidation of olefins e.g., styrene, cis-
dibenzofuran (J, K) scaffolds studied by Nocera and co-workers.283,284 cyclooctene, using hydrogen peroxide as a terminal oxidant,
was carried out using manganese HPX and HPD derivatives as

Scheme 11. (Top) Formation of Compound I (b) and Compound II (c) Intermediates from Acylperoxo−FeIII(HPX-CO2R)
(a) by the Reaction of 46C with m-CPBA and PPAA, Respectively. (Bottom) Proposed Reaction Mechanism of H2O2
Activation by (HPX-CO2H)FeIII(OH)274

12412 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

catalysts (46H, 46I, and 46K) and MnIII(TMP)Cl as were not significantly perturbed by changing the acid
control.283 The results for epoxidation reactivity of the functional group as the rates of O−O heterolysis in the
manganese derivatives followed the same general trend hangman scaffold, bearing either the −COOH or the
observed for catalase reactions catalyzed using the correspond- −COOMe substituents, were similar and comparable to the
ing iron analogues. The results showed that Complex 46H parent FeTMP analogue (45C). However, changing the
bearing a protic carboxylic acid group on a xanthene scaffold organic medium to an even more nonpolar solvent, toluene,
was most active, affording styrene epoxide (yield 70%) and cis- the reaction of 45C and m-CPBA at −20 °C resulted in the
cyclooctene oxide (yield 92%) with TONs almost twice as high formation of ferric N-oxide species via a homolytic O−O bond
as those obtained for the ester derivative, 46I. Under similar cleavage where high-valent compound II (TMP)FeIV�O was
conditions, both the HPD complex 46K and MnIII(TMP)Cl initially formed and subsequent attack on the porphyrin
produced significantly less epoxide products upon reaction pyrrole N atom.272 The formation of the ferric N-oxide
with either of the two substrates. Altogether, activity product formation was significantly decreased in the presence
enhancement observed for the xanthene-bridged platform of hangman ester scaffold of 46C under the same reaction
with a pendant carboxylic acid group (HPX-CO2H) indicated conditions, and no ferric N-oxide product was observed at all
a clear advantage of a geometrically more favorable proton for the hangman −COOH analogue 46B. The generation of
source in reactivity. Increasing the distance between the proton compound II from peracid entailed homolysis of the O−O
donor −COOH group and the metal in the porphyrin cavity bond of the peroxide intermediate. While compound II could
by using a dibenzofuran scaffold instead of the xanthene be directly obtained from the reactions of 46C with
scaffold resulted in reactivity comparable to the control TMP phenylperoxyacetic acid (PPAA) at −40 °C in toluene
despite the presence of a pendant protic hydrogen bonding (Scheme 11, top), no compound II could be observed for
group. Structure−function relationships of the HPX and HPD 46B having pendant −COOH group suggesting that hangman
complexes in both catalase-like H2O2 disproportionation and architecture with properly oriented acid functional group
peroxide-mediated olefin epoxidations revealed that both the favored the heterolytic O−O bond cleavage to generate
nature of the hydrogen-bonding group and its spatial compound I over competing homolytic cleavage (1e− path-
orientation are crucial for the efficient activation of O−O way) (Scheme 11, bottom).
bond of putative metal peroxides. In addition to the hangman porphyrins from Nocera, several
In a subsequent work, a series of hangman iron porphyrins Mn porphyrins with positively changed meso substituents were
based on xanthene scaffold (HPX) containing pendant groups reported to have encouraging reactivities.285 The first-order
with a wide range of different acid−base functionalities (pKa rate constant of O−O bond heterolysis was obtained as 66 ±
ranging from ∼2 to 25 in water) was synthesized (Figure 12 s−1 with activation parameters estimated as ΔH⧧ = 4.2 ±
46).284 The kinetics of H2O2 disproportionation suggested that 0.2 kcal mol−1 and ΔS⧧ = −36 ± 1 cal mol−1 K−1. The positive
the initial rate constants for catalase activity followed the charge on the substituent on the meso carbons of porphyrin
acidity of the pendant group such that the complex 46A with ring exerted strong electron-withdrawing effect which resulted
strongly acidic phosphonic acid functional group, showed the in ∼200 mV higher MnIII/II reduction potential relative to
highest rate constant. Thus, HPX complexes (46A and 46B) other Mn porphyrins.286,287 The positively charged porphyrin
bearing the most acidic pendants (phosphonic acid and meso substituent was also proposed to aid in O−O heterolysis
carboxylic acid) exhibited similar initial rate constants that via electrostatic effect.285
were 3 orders of magnitude higher than that of the complex In a recent work, Dey et al. developed a structural as well as
45C. The complex 46A decomposed rapidly during the a functional model of peroxidase (Fe-MARG), containing a
disproportionation reaction, but complex 46B retained its covalently attached guanidine moiety in the second sphere of
activity. It was proposed that hydrogen-bonding groups with porphyrin ligand (Figure 33E), which mimicked the Arg38
higher pKa values were less competent in donating the protons residue present in the active site of HRP.288 The role of Arg38
required for the heterolytic O−O bond cleavage, while the in HRP had been established since the early 1980s, but
conjugate base of the more acidic phosphonic acid guanidine functional groups were not modeled in synthetic
functionality (pKa < 2) was incapable for accepting protons systems. This was the first instance where the guanidine
during the catalytic cycle. However, the overall catalase activity residue was incorporated in a synthetic model porphyrin that
was maximum for 46B, iron hangman porphyrin bearing a showed peroxidase-like activity affecting the O−O bond
carboxylic acid group (pKa ∼ 4.2) which exhibited the highest heterolysis of FeIII−OOH intermediate compound 0 by the
TON where an optimal balance might have been achieved for reaction of H2O2, being used as a green oxidant. To investigate
acid−base catalysis needed for O−O bond heterolysis. the peroxidase assay in a molecular catalyst, 3,3′,5,5′-
Nocera et al. also directly investigated the O−O bond tetramethylbenzidine (TMB) was used as a model sub-
cleavage of peroxyacids resulting in high-valent oxo inter- strate.289,290 The reaction kinetics for the catalytic oxidation
mediates to decipher the influence of a proton donor group of TMB to diimine formation by Fe-MARG (33E) using H2O2
using iron hangman porphyrins (P)FeIII(OH) having carbox- oxidant and a trace amount of acetic acid in a 5% v/v H2O−
ylic acid (P = HPX-CO2H, 46B) and methyl ester (P = HPX- acetonitrile mixture, was monitored using absorption spectros-
CO2Me, 46C) by following the kinetics of formation of copy. The rate of TMB oxidation was much faster in 33E than
compound I species.274 Both of these complexes upon reaction other analogous porphyrins like FeTPP (17A) and FePf (20A),
with m-CPBA in CH2Cl2 underwent heterolytic O−O bond which clearly indicated the role of pendant guanidinium group
cleavage generating compound I intermediate at cryogenic in accelerating catalytic substrate oxidation. In addition, 33E
temperature (−40 °C) having characteristic absorption catalyzed a variety of substrates like [tetramethylbenzidine
features (Scheme 11, top). The rate constants for the O−O (TMB), o-phenylenediamine (OPD), and 2,4,6-tri-tert-butyl-
bond heterolysis were determined from the stopped-flow phenol (TBPH)] oxidations291,292 using other oxidant like m-
kinetics data and the rates of compound I formation which CPBA. Thus, the suitably poised guanidine moiety in the distal
12413 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Table 8. Comparison of Kinetic Parameters of Oxidation of Organic Substrates Using H2O2


catalyst KMH2O2 (mM) kcat (s−1) ηH2O2 = kcat/KMH2O2 (M−1 s−1) ref
Fe(III)-MC6 44 ± 2 370 ± 10 8.4 ± 0.6 × 103 290
E2L(TD)-MC6 31 ± 2 780 ± 60 25 ± 3 × 103
Protein-86 n.d ∼280 500 296
Protein-K5 ∼217 125
CTMs-C45 94 1200 1.3 × 104 294
μ-peroxidase 21 680 297
reconstituted-Mb 7.4 ± 0.5 1.1 ± 0.1 1.4× 102 293
antibody−heme complex 24 6.56 274 296
peroxidase (HRP, MPO) 3.7 0.01−2100* 4.6 × 106 291, 292
FeIIICl-MARG (33E) 4.76 ± 0.5 25 5.2 × 103 284

site of simple TPP analogue made Fe-MARG (33E) an Scheme 12. Plausible mechanism of FeIIICl-MARG catalysed
indiscriminate peroxidase mimic. Additionally, complex 33E substrate oxidation. Note that the Cl− ion is likely to be
showed a Michaelis−Menten-type saturation-kinetics over a exchanged by OH− under the reaction conditions (5%v/v
reasonable range of TMB and H2O2 concentration, which is H2O/CH3CN).288
very rarely observed for any artificial catalyst as a catalyst−
substrate [CS] complex285,293,294 analogous to enzyme−
substrate complex [ES]. The two most relevant kinetic
parameters KMH2O2 and the catalytic efficiency (ηH2O2) were
determined to be 4.76 ± 0.5 and 3.4 ± 0.4 mM, respectively.
The KM values of both TMB and H2O2 were also similar to
that of native HRP (Table 8).295,296 The catalytic rate toward
TMB oxidation of Fe-MARG is 25 ± 5 s−1 and the catalytic
efficiency (ηH2O2) is 5.2 × 103 M−1 s−1, which are lower than
that of the natural enzyme but comparable to or even higher
than that of most of the artificial proteins and engineered
myoglobin reported (Table 8).294,297−301 In addition, the
nonsequential multiple substrate binding was observed for this
complex, having low KM for both TMB and H2O2 just like
natural enzymes, suggesting a “ping-pong” mechanism in a
molecular catalyst.295 The reactive intermediate formed on the
reaction of 33E with peroxides was compound I as evident
from typical absorption features. The second sphere guanidine
group present in 33E remained protonated at neutral pHs and
assisted in facile heterolytic cleavage of FeIII−OOH species amine functional group, respectively, were rationally designed
leading to the compound I formation. The guanidinium group to emulate the distal histidine residue in peroxidases. A
being a strong hydrogen bond donor and could also provide structurally analogous iron porphyrin without basic moiety,
auxiliary binding site to the substrates used here, e.g., TMB, FeMPh (33A), containing a pendant benzyl group in the distal
OPD, and TBPH (Scheme 12). Overall, this model complex, site had been employed as control. The formation of a high
Fe-MARG, utilized the “ping pong” mechanism for substrate valent iron−oxo intermediate was observed in the reaction of
oxidation with KMH2O2 and kMsubstrate values similar to that of these iron(III)−porphyrin complexes with m-CPBA at low
native enzyme and exhibited competitive inhibition kinetics temperature using stopped-flow kinetics. This intermediate had
(with azide), thus emulating enzymatic activity on inclusion of been trapped and unequivocally identified at a compound I
second sphere guanidine residue.288,302 These results clearly species by 9-GHz EPR spectroscopy at 4K. The broad (2000
suggested that the presence of distal guanidinium group G) and axial EPR spectrum of an exchange-coupled
transformed the simple iron porphyrin into a miniature oxoferrylporphyrin radical species, the [(FeIV�O)Por*+]
enzyme both qualitatively as well as quantitatively. with gO5eff = 3.80 and g||eff = 1.99, was observed upon reaction
Dey et al. used a series of ferric iron porphyrin complexes of 33D (FeL3) with m-CPBA. The formation of this reactive
(FeL2, FeL3, and FeMPh) having pendant basic groups intermediate was much faster for 33C and 33D than that of the
(pyridine, primary amine, and control) covalently attached to benzyl analogue, 33A. The khetero values of O−O bond
the amine group of the meso-phenyl ring in the TPP heterolysis were determined to be more than 2 orders of
architecture (Figure 33), in order to investigate the rates of magnitude faster in 33C and 33D with respect to 33A which
O−O bond cleavage and to characterize the ensuing provided definitive proof of the advantage of having pendant
intermediates formed.303 In the active site of peroxidases, nitrogenous bases during O−O heterolytic cleavage. The khetero
histidine residue plays a vital role with the “pull effect” values for compound I formation for these two complexes were
transferring proton to the distal oxygen atom of compound 0 much higher than some other previously reported model
for the heterolytic O−O bond cleavage. To date, the role of the complexes (Table 9).274,280 In contrast, no compound I
histidine residue remains to be successfully modeled in any intermediate was observed during the reaction of parent
synthetic heme-based model system. These model complexes FeTPP (17A) with m-CPBA under identical conditions rather
FeL2 (33C) and FeL3 (33D) containing pendant pyridine and it showed a slow formation of isoporphyrin, a species
12414 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Table 9. Rate of O−O Heterolysis and the Corresponding Scheme 13. Proposed Mechanism of the Catalytic Cycle for
Activation Parametersa the Formation of High-Valent Intermediates Starting from
Iron(III) Porphyrins in the Presence of m-
temp ΔG⧧
catalyst kheterolysis (s−1) (°C) (kcal/mol−1) ref Chloroperbenzoic Acid (Oxidant) and One-Electron
45A 16.6 × 10 −3
−80 12.78 276
Donor/Substrate (S)303
45B 4 × 10−3 −80 13.33 276
45D 3.94 × 10−3 −40 16.19 276
45E 0.5 × 10−4 −80 14.14 276
45C (6.5 ± 1)×10−3 −40 15.96 ± 0.1 270
46B (6 ± 3)×10−3 −40 15.99 ± 0.2 270
46C (1 ± 0.2)×10−2 −40 15.76 ± 0.1 270
33C 21.67 ± 1 −30 12.72 ± 0.02 299
33D 69.6 ± 2 −30 12.15 ± 0.01 299
33A (8.5 ± 1)×10−2 −30 15.08 ± 0.05 299
HRP 200−500 25 13−14 295, 302
H42A/H42 V (5−50)×10−4 25 21 ± 1 53, 303
a
m-CPBA was used as the oxidant for all of the catalysts except for
those used in refs 53, 295, and 303 (where hydrogen peroxide is the
oxidant).

isoelectronic with compound I and an intermediate in heme


degradation of heme oxygenase cycle, that has characteristic
bands at 810 and 900 nm.304,305 The formation of isoporphyrin
was not only very minor but also sluggish in 33C and 33D,
indicating that the pendent basic groups avoid degradation of
the metallopophyrins upon their reaction with peroxyacids. of FeIII−OOH species of these complexes (33C and 33D)
The energy barrier for O−O heterolysis was found to be were involved in strong H-bonding interactions with the
very low (12−13 kcal/mol) for 33C and 33D (Table 9) protonated pendant bases and the O−O bond distances are
relative to previously investigated iron porphyrin complexes elongated to 1.88 and 1.86 Å in their corresponding DFT-
including the hangman porphyrins. The activation parameters optimized structures of ferric hydroperoxide species of 33C
for compound I formation of FeL2 were determined to have and 33D, respectively (Figure 38).197 A molecular orbital
ΔH⧧ = 2.34 ± 0.2 kcal/mol) and ΔS⧧ = −44.60 ± 1 cal mol−1 (MO) diagram was constructed to elucidate the bonding in a
K−1. The remarkably low enthalpic barrier and large negative putative compound 0 intermediate of 33C and 33D having
entropy of activation indicated a very facile O−O bond pendant pyridine and amine functional groups. The substantial
cleavage step. The corresponding activation energy barrier activation of the O−O bond of a FeIII−OOH species with H-
ΔG⧧ for HRP (∼13−14 kcal/mol) was higher than those bonding by a protonated base was the root of the “pull effect”
achieved in 33C and 33D (Table 9). Additionally, the three proposed in peroxidase active sites.46,253,308 The enhanced
times faster rate and 0.5 kcal/mol lower ΔG⧧ of 33D with an back-bonding from the Fe 3d orbitals was facilitated by the
aliphatic amine (pKa ∼ 16.9 in CH3CN) than 33C with a lowering of the O−O σ* orbital energies due to hydrogen
pyridine group (pKa ∼ 12.33 in CH3CN) suggested that the bonding interactions with the pendent basic groups of [FeIIIL2-
O−O bond heterolysis was more favored by a pendent amine OOH]H+ (38C) and [FeIIIL3-OOH]H+] species (38D). This
with higher basicity than the peroxyacid. Therefore, these two strong H-bonding polarized the electron density of the O−O
residues, in principle, easily acted as efficient acid−base catalyst bond toward the distal OH group reducing the overlap
capable of translocating proton from proximal oxygen atom of between the two O2p orbitals as a result of which the O−O
bound m-CPBA to the distal oxygen atom that eventually bond was weakened and the O−O σ* orbital energy was
facilitated the heterolytic O−O bond cleavage (Scheme 13). lowered. The DFT calculated wave functions showed that the
Note that mutation of the His42 residue in native HRP led to Fe dπ character in the O−O σ* orbitals significantly increased
several orders of magnitude decrease in the rate of compound I from 4.86% in the non H-bonded system to 36.89% and
formation via O−O bond heterolysis,55,299,306,307 which was 38.69% in FeL3 and FeL2, respectively (Figure 47). Thus, the
reasonably captured by 2 orders of magnitude decrease in the “pull effect” might be interpreted as pulling the energy of the
rate of compound I formation for 33A relative to 33C and O−O σ* orbital down substantially by hydrogen bonding to a
33D. Thus, the “pull effect” operating by the distal histidine protonated basic residue to allow better back-bonding into the
residue in peroxidases was successfully mimicked in synthetic O−O σ* orbital from filled Fe dπ orbital of the metal. The O2p
models including pendent basic groups (analogous to histidine coefficient increased in the Fe dπ orbital from 3.62% in
residue) in the second coordination sphere of the iron FeIIITPP-OOH to 18.05% in [FeIIIL3-OOH]H+ and 17.55% in
porphyrin. The proposed enhancement in O−O bond [FeIIIL2-OOH]H+, signifying increased covalent donation
heterolysis in FeL2, FeL3 and Fe-MARG based on in situ from the proximal oxygen to the vacant Fe dπ orbital, i.e.,
SERRS-RDE data obtained during electrocatalytic ORR, strong π bonding (Figure 47). A similar increase in σ covalency
indicated the same. was observed in the Fe dz2 orbital. These were consistent with
DFT calculations provided further insights into the effect of the shortening of Fe−O bond upon elongation of the O−O
pendent basic groups covalently attached to iron porphyrins in bond converting the reactant FeIII−OOH to the product
relation to such O−O bond cleavage. The distal oxygen atoms [(FeIV�O)Por*+].
12415 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Figure 47. Calculated (BP86) ground-state MO diagram of FeIIITPP-OOH (no H-bonding residue), [FeIIIL3-OOH]H+ and [FeIIIL2-OOH]H+
(protonated H-bonding residue). The relative energies are normalized with respect to the nonbonding dxy orbital. Only the β orbitals are shown.
The orbital contributions of the occupied dxz orbital and the unoccupied O−O σ* orbital are shown. The contributions of dz2 and the dxy orbitals of
the Fe that σ and π bonds, respectively, to the hydroperoxide are also shown. The nonbonding dxy and the dx2‑y2 are not shown for clarity.303

Inclusion of pendent basic groups in iron porphyrins and biosynthetic models provide convenient platforms for
mimicking the distal basic residues in the active site structure understanding the finer attributes of metalloenzyme activity
of peroxidases, enabled facile O−O bond heterolysis of outside the protein matrix. In addition to understanding the
peroxyacid lowering the activation energy barrier for the reactivity, some recent examples have mimicked the catalytic
generation of high-valent compound I intermediate and in function of the enzymes as well which has been a long-term
doing so allow the “pull effect” exhibited by the secondary goal of the bioinspired community. It is perhaps not too
coordination environment of the heme cofactor of heme
hydroperoxidase enzymes to be modeled. premature to state that judicious inclusion of second sphere
interactions in synthetic models likely hold the key for the
transition from structural modeling to functional modeling, a
6. CONCLUSION
much needed change in characterization of the bioinspired
Inclusion of second-sphere residues in the design of small- community.
molecule artificial bioinspired complexes has produced some
remarkable advancements in our understanding of how these
weak interactions play vital roles in attenuating the reactivity of AUTHOR INFORMATION
metallo-enzymes and has accelerated the community along Corresponding Author
toward generating functional mimics of these naturally Abhishek Dey − School of Chemical Science, Indian
occurring elegant catalysts. The nature of second-sphere Association for the Cultivation of Science, Kolkata 700032,
residues varies substantially and so does the effect it has on India; orcid.org/0000-0002-9166-3349;
the reactivity of the active site. A detailed understanding of the Email: abbeyde@gmail.com
nuances introduced by the second-sphere residues in the
geometric and electronic structure of the reactive species in the Authors
active site is prerequisite to successful implementation of these
design principles in small molecules. The dramatic improve- Sarmistha Bhunia − School of Chemical Science, Indian
ments in O−O bond cleavage in small molecule catalysts on Association for the Cultivation of Science, Kolkata 700032,
switching pendant carboxylic acids with pendant ammines India
(mimics histidine and arginine in HRP) as well as the clear Arnab Ghatak − School of Chemical Science, Indian
advantage of a properly positioned phenol in affecting O−O Association for the Cultivation of Science, Kolkata 700032,
bond cleavage are good examples of that. The examples India
discussed here, while limited to O2 and H2O2 activation by Complete contact information is available at:
heme active sites, succinctly demonstrate how these synthetic https://pubs.acs.org/10.1021/acs.chemrev.1c01021
12416 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Author Contributions PT proton transfer



S.B. and A.G. contributed equally. Q quinone
RDE rotating disc electrochemistry
Notes
RDS rate-determining step
The authors declare no competing financial interest. rR resonance Raman
Biographies RRDE rotating ring disc electrochemistry
SAM self-assembled monolayer
Sarmistha Bhunia received her B.Sc. (2013) from Midnapore college SERRS surface-enhanced resonance Raman spectroscopy
and M.Sc. (2015) from the Indian Association for the Cultivation of SERRS-RDE surface-enhanced resonance Raman spectros-
Science. She joined as a Ph.D. scholar in the same institute under the copy coupled to rotating disc electrochemistry
supervision of Prof. A. Dey in 2015. Her current research focuses on SKIE solvent kinetic isotope effect
synthesis and reactivity of iron porphyrin complexes having second SQ semi quinone
sphere distal residues. TBPH 2,4,6-tritert-butylphenol
Arnab Ghatak received his B.Sc. (2014) from Presidency University, TMB tetramethylbenzidine
Kolkata, and M.Sc. (2016) from Indian Institute of Technology TOF turn over frequency
Kharagpur. He joined the Indian Association for the Cultivation of TON turn over number
Science as a Ph.D. scholar under the supervision of Prof. A. Dey in Tyr• tyrosyl radical
2016. His current research focuses on the spectroscopic investigation Fc* decamethyl ferrocene
of the mechanism of oxygen reduction reaction catalyzed by iron
porphyrin complexes. REFERENCES
Abhishek Dey received his B.Sc. (1999) from Presidency College, (1) Babcock, G. T. How Oxygen Is Activated and Reduced in
Kolkata, M.Sc. (2001) from the Indian Institute of Technology, Respiration. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 12971−12973.
Kanpur, and Ph.D. (2007) from Stanford University, CA. After a (2) Kovaleva, E. G.; Lipscomb, J. D. Versatility of Biological Non-
postdoctoral stay at Stanford University, he moved to the Indian Heme Fe(II) Centers in Oxygen Activation Reactions. Nat. Chem.
Association for the Cultivation of Science (IACS), Kolkata, as an Biol. 2008, 4, 186−193.
Assistant Professor (2009). Currently, he is a professor at the same (3) Proshlyakov, D. A.; Pressler, M. A.; Babcock, G. T. Dioxygen
department. His primary research interests include inorganic reaction Activation and Bond Cleavage by Mixed-Valence Cytochrome c
mechanisms relevant to renewable energy and a clean environment. Oxidase. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 8020−8025.
(4) Ferguson-Miller, S.; Babcock, G. T. Heme/Copper Terminal
Oxidases. Chem. Rev. 1996, 96, 2889−2908.
ACKNOWLEDGMENTS (5) Ludwig, B.; Bender, E.; Arnold, S.; Hüttemann, M.; Lee, I.;
The work is supported by a Department of Science and Kadenbach, B. Cytochrome c Oxidase and the Regulation of Oxidative
Technology grant (SERB/STR/2019/000081). S.B. acknowl- Phosphorylation. ChemBioChem. 2001, 2, 392−403.
edges the IACS integrated Ph.D. program, and A.G. acknowl- (6) Savéant, J. M. Molecular Catalysis of Electrochemical Reactions.
edges UGC SRF and IACS for fellowships. Mechanistic Aspects. Chem. Rev. 2008, 108, 2348−2378.
(7) Jaouen, F.; Proietti, E.; Lefèvre, M.; Chenitz, R.; Dodelet, J. P.;
ABBREVIATIONS Wu, G.; Chung, H. T.; Johnston, C. M.; Zelenay, P. Recent Advances
in Non-Precious Metal Catalysis for Oxygen-Reduction Reaction in
CcO cytochrome c oxidase Polymer Electrolyte Fuel Cells. Energy Environ. Sci. 2011, 4, 114−130.
DFT density functional theory (8) Wang, Y. H.; Goldsmith, Z. K.; Schneider, P. E.; Anson, C. W.;
EPG edge-plane graphite Gerken, J. B.; Ghosh, S.; Hammes-Schiffer, S.; Stahl, S. S. Kinetic and
EPR electron paramagnetic resonance Mechanistic Characterization of Low-Overpotential, H2O2-Selective
ET electron transfer Reduction of O2 Catalyzed by N2O2-Ligated Cobalt Complexes. J.
EXAFS X-ray absorption fine structure Am. Chem. Soc. 2018, 140, 10890−10899.
Fc ferrocene (9) Wu, G.; Zelenay, P. Nanostructured Nonprecious Metal
GC glassy carbon Catalysts for Oxygen Reduction Reaction. Acc. Chem. Res. 2013, 46,
HAT hydrogen atom abstraction 1878−1889.
Hb hemoglobin (10) Zhang, W.; Lai, W.; Cao, R. Energy-Related Small Molecule
HCO heme copper oxidase Activation Reactions: Oxygen Reduction and Hydrogen and Oxygen
HRP horseradish peroxidase Evolution Reactions Catalyzed by Porphyrin- and Corrole-Based
kET ET rate constant Systems. Chem. Rev. 2017, 117, 3717−3797.
kcat ORR rate constant (11) Chatterjee, S.; Sengupta, K. Direct Methanol Fuel Cell
KIE kinetic isotope effect Technology ∥ 6-Carbon-Based Electrodes for Direct Methanol Fuel
Cells; Elsevier, 2020; pp 135−176.
K-L Koutecky−Levich
(12) Carver, C. T.; Matson, B. D.; Mayer, J. M. Electrocatalytic
Mb myoglobin Oxygen Reduction by Iron Tetra-Arylporphyrins Bearing Pendant
m-CPBA m-chloroperbenzoic acid Proton Relays. J. Am. Chem. Soc. 2012, 134, 5444−5447.
MO molecular orbital (13) Fukuzumi, S.; Kotani, H.; Lucas, H. R.; Doi, K.; Suenobu, T.;
ηORR overpotential of ORR Peterson, R. L.; Karlin, K. D. Mononuclear Copper Complex-
NMR nuclear magnetic resonance Catalyzed Four-Electron Reduction of Oxygen. J. Am. Chem. Soc.
OPD o-phenylenediamine 2010, 132, 6874−6875.
ORR oxygen reduction reaction (14) Thorseth, M. A.; Letko, C. S.; Tse, E. C. M.; Rauchfuss, T. B.;
PCET proton-coupled electron transfer Gewirth, A. A. Ligand Effects on the Overpotential for Dioxygen
PET proton transfer followed by electron transfer Reduction by Tris(2-Pyridylmethyl)Amine Derivatives. Inorg. Chem.
PROS partially reduced oxygen species 2013, 52, 628−634.

12417 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

(15) Dey, S.; Mondal, B.; Chatterjee, S.; Rana, A.; Amanullah, S.; (36) Denisov, I. G.; Makris, T. M.; Sligar, S. G.; Schlichting, I.
Dey, A. Molecular Electrocatalysts for the Oxygen Reduction Structure and Chemistry of Cytochrome P450. Chem. Rev. 2005, 105,
Reaction. Nat. Rev. Chem. 2017, 1, 98. 2253−2278.
(16) Cracknell, J. A.; Vincent, K. A.; Armstrong, F. A. Enzymes as (37) Sono, M.; Roach, M. P.; Coulter, E. D.; Dawson, J. H. Heme-
Working or Inspirational Electrocatalysts for Fuel Cells and Containing Oxygenases. Chem. Rev. 1996, 96, 2841−2888.
Electrolysis. Chem. Rev. 2008, 108, 2439−2461. (38) Meunier, B.; de Visser, S. P.; Shaik, S. Mechanism of Oxidation
(17) Palmer, A. E.; Lee, S. K.; Solomon, E. I. Decay of the Peroxide Reactions Catalyzed by Cytochrome P450 Enzymes. Chem. Rev. 2004,
Intermediate in Laccase: Reductive Cleavage of the O-O Bond. J. Am. 104, 3947−3980.
Chem. Soc. 2001, 123, 6591−6599. (39) Phillips, S. E. V. Structure and Refinement of Oxymyoglobin at
(18) Solomon, E. I.; Sundaram, U. M.; Machonkin, T. E. 1·6 Å Resolution. J. Mol. Biol. 1980, 142, 531−554.
Multicopper Oxidases and Oxygenases. Chem. Rev. 1996, 96, 2563− (40) Nagano, S.; Poulos, T. L. Crystallographic Study on the
2606. Dioxygen Complex of Wild-Type and Mutant Cytochrome P450cam:
(19) Messerchmidt, A.; Huber, R. The Blue Oxidases, Ascorbate IMPLICATIONS FOR THE DIOXYGEN ACTIVATION MECH-
Oxidase, Laccase and Ceruloplasmin Modelling and Structural ANISM*⧫. J. Biol. Chem. 2005, 280, 31659−31663.
Relationships. Eur. J. Biochem. 1990, 187, 341−352. (41) Berglund, G. I.; Carlsson, G. H.; Smith, A. T.; Szöke, H.;
(20) Dooley, D. M.; Rawlings, J.; Dawson, J. H.; Stephens, P. J.; Henriksen, A.; Hajdu, J. The Catalytic Pathway of Horseradish
Andreasson, L. E.; Malmstrom, B. G.; Gray, H. B. Spectroscopic Peroxidase at High Resolution. Nature 2002, 417, 463−468.
Studies of Rhus Vernicifera and Polyporus Versicolor Laccase. (42) Kaila, V. R. I.; Oksanen, E.; Goldman, A.; Bloch, D. A.;
Electronic Structures of the Copper Sites. J. Am. Chem. Soc. 1979, Verkhovsky, M. I.; Sundholm, D.; Wikström, M. A Combined
101, 5038−5046. Quantum Chemical and Crystallographic Study on the Oxidized
(21) Ducros, V.; Brzozowski, A. M.; Wilson, K. S.; Brown, S. H.; Binuclear Center of Cytochrome c Oxidase. Biochim. Biophys. Acta -
Østergaard, P.; Schneider, P.; Yaver, D. S.; Pedersen, A. H.; Davies, G. Bioenerg. 2011, 1807, 769−778.
J. Crystal Structure of the Type-2 Cu Depleted Laccase from (43) Parey, K.; Warkentin, E.; Kroneck, P. M. H.; Ermler, U.
Coprinus Cinereus at 2.2 Å Resolution. Nat. Struct. Biol. 1998, 5, Reaction Cycle of the Dissimilatory Sulfite Reductase from
310−316. Archaeoglobus Fulgidus,. Biochemistry 2010, 49, 8912−8921.
(22) Shi, C.; Anson, F. C. Comparison of the Catalytic Reduction of (44) Dawson, J. H.; Sono, M. Cytochrome P-450 and Chloroper-
Oxygen by [5,10,15,20-Tetrakis((Pentaammineruthenio(II)-4- oxidase: Thiolate-Ligated Heme Enzymes. Spectroscopic Determi-
Pyridyl)Porphyrinato]Cobalt(II) in Solution and on Graphite nation of Their Active-Site Structures and Mechanistic Implications of
Thiolate Ligation. Chem. Rev. 1987, 87, 1255−1276.
Electrode Surfaces. Inorg. Chem. 1995, 34, 4554−4561.
(45) Karlin, S.; Zhu, Z.-Y.; Karlin, K. D. The Extended Environment
(23) Song, E.; Shi, C.; Anson, F. C. Comparison of the Behavior of
of Mononuclear Metal Centers in Protein Structures. Proc. Natl. Acad.
Several Cobalt Porphyrins as Electrocatalysts for the Reduction of O2
Sci. U. S. A. 1997, 94, 14225−14230.
at Graphite Electrodes. Langmuir 1998, 14, 4315−4321.
(46) Ozaki, S.; Roach, M. P.; Matsui, T.; Watanabe, Y. Investigations
(24) Gewirth, A. A.; Varnell, J. A.; DiAscro, A. M. Nonprecious
of the Roles of the Distal Heme Environment and the Proximal Heme
Metal Catalysts for Oxygen Reduction in Heterogeneous Aqueous
Iron Ligand in Peroxide Activation by Heme Enzymes via Molecular
Systems. Chem. Rev. 2018, 118, 2313−2339.
Engineering of Myoglobin. Acc. Chem. Res. 2001, 34, 818−825.
(25) Pegis, M. L.; Wise, C. F.; Martin, D. J.; Mayer, J. M. Oxygen
(47) Poulos, T. L. Heme Enzyme Structure and Function. Chem.
Reduction by Homogeneous Molecular Catalysts and Electrocatalysts. Rev. 2014, 114, 3919−3962.
Chem. Rev. 2018, 118, 2340−2391. (48) Polyakov, K. M.; Boyko, K. M.; Tikhonova, T. V.; Slutsky, A.;
(26) Wikström, M.; Krab, K.; Sharma, V. Oxygen Activation and Antipov, A. N.; Zvyagilskaya, R. A.; Popov, A. N.; Bourenkov, G. P.;
Energy Conservation by Cytochrome c Oxidase. Chem. Rev. 2018, Lamzin, V. S.; Popov, V. O. High-Resolution Structural Analysis of a
118, 2469−2490. Novel Octaheme Cytochrome c Nitrite Reductase from the
(27) Halime, Z.; Kotani, H.; Li, Y.; Fukuzumi, S.; Karlin, K. D. Haloalkaliphilic Bacterium Thioalkalivibrio Nitratireducens. J. Mol.
Homogeneous Catalytic O2 Reduction to Water by a Cytochrome c Biol. 2009, 389, 846−862.
Oxidase Model with Trapping of Intermediates and Mechanistic (49) Olson, J. S.; Mathews, A. J.; Rohlfs, R. J.; Springer, B. A.;
Insights. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 13990−13994. Egeberg, K. D.; Sligar, S. G.; Tame, J.; Renaud, J. P.; Nagai, K. The
(28) Collman, J. P.; Boulatov, R.; Sunderland, C. J.; Fu, L. Role of the Distal Histidine in Myoglobin and Haemoglobin. Nature
Functional Analogues of Cytochrome c Oxidase, Myoglobin, and 1988, 336, 265−266.
Hemoglobin. Chem. Rev. 2004, 104, 561−588. (50) Shiro, Y.; Morishima, I. Modification of the Heme Distal Side
(29) Adam, S. M.; Wijeratne, G. B.; Rogler, P. J.; Diaz, D. E.; Quist, in Myoglobin by Cyanogen Bromide. Heme Environmental Structures
D. A.; Liu, J. J.; Karlin, K. D. Synthetic Fe/Cu Complexes: Toward and Ligand Binding Properties of the Modified Myoglobin.
Understanding Heme-Copper Oxidase Structure and Function. Chem. Biochemistry 1984, 23, 4879−4884.
Rev. 2018, 118, 10840−11022. (51) Chen, H.; Ikeda-Saito, M.; Shaik, S. Nature of the Fe-O2
(30) Zhang, X. P.; Chandra, A.; Lee, Y. M.; Cao, R.; Ray, K.; Nam, Bonding in Oxy-Myoglobin: Effect of the Protein. J. Am. Chem. Soc.
W. Transition Metal-Mediated O-O Bond Formation and Activation 2008, 130, 14778−14790.
in Chemistry and Biology. Chem. Soc. Rev. 2021, 50, 4804−4811. (52) Ortiz de Montellano, P. R. Hydrocarbon Hydroxylation by
(31) Li, Y.; Wang, N.; Lei, H.; Li, X.; Zheng, H.; Wang, H.; Zhang, Cytochrome P450 Enzymes. Chem. Rev. 2010, 110, 932−948.
W.; Cao, R. Bioinspired N4-Metallomacrocycles for Electrocatalytic (53) Veitch, N. C. Horseradish Peroxidase: A Modern View of a
Oxygen Reduction Reaction. Coord. Chem. Rev. 2021, 442, 213996. Classic Enzyme. Phytochemistry 2004, 65, 249−259.
(32) Riggs, A. F. Hemoglobins. Curr. Opin. Struct. Biol. 1991, 1 (6), (54) Tatoli, S.; Zazza, C.; Sanna, N.; Palma, A.; Aschi, M. The Role
915−921. of Arginine 38 in Horseradish Peroxidase Enzyme Revisited: A
(33) Møller, J. K. S.; Skibsted, L. H. Nitric Oxide and Myoglobins. Computational Investigation. Biophys. Chem. 2009, 141, 87−93.
Chem. Rev. 2002, 102, 1167−1178. (55) Newmyer, S. L.; de Montellano, P. R. O. Horseradish
(34) Flögel, U.; Merx, M. W.; Gödecke, A.; Decking, U. K. M.; Peroxidase His-42 → Ala, His-42 → Val, and Phe-41 → Ala Mutants:
Schrader, J. Myoglobin: A Scavenger of Bioactive NO. Proc. Natl. HISTIDINE CATALYSIS AND CONTROL OF SUBSTRATE
Acad. Sci. U. S. A. 2001, 98, 735−740. ACCESS TO THE HEME IRON*. J. Biol. Chem. 1995, 270,
(35) Kim, E.; Chufán, E. E.; Kamaraj, K.; Karlin, K. D. Synthetic 19430−19438.
Models for Heme-Copper Oxidases. Chem. Rev. 2004, 104, 1077− (56) Proshlyakov, D. A.; Pressler, M. A.; DeMaso, C.; Leykam, J. F.;
1134. DeWitt, D. L.; Babcock, G. T. Oxygen Activation and Reduction in

12418 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Respiration: Involvement of Redox-Active Tyrosine 244. Science 2000, (76) Samanta, S.; Das, P. K.; Chatterjee, S.; Sengupta, K.; Mondal,
290, 1588−1591. B.; Dey, A. O2 Reduction Reaction by Biologically Relevant Anionic
(57) Kaila, V. R. I.; Verkhovsky, M. I.; Wikström, M. Proton- Ligand Bound Iron Porphyrin Complexes. Inorg. Chem. 2013, 52,
Coupled Electron Transfer in Cytochrome Oxidase. Chem. Rev. 2010, 12963−12971.
110, 7062−7081. (77) Rydberg, P.; Sigfridsson, E.; Ryde, U. On the Role of the Axial
(58) McCauley, K. M.; Vrtis, J. M.; Dupont, J.; van der Donk, W. A. Ligand in Heme Proteins: A Theoretical Study. JBIC J. Biol. Inorg.
Insights into the Functional Role of the Tyrosine-Histidine Linkage in Chem. 2004, 9, 203−223.
Cytochrome c Oxidase. J. Am. Chem. Soc. 2000, 122, 2403−2404. (78) Li, T.; Bonkovsky, H. L.; Guo, J. Structural Analysis of Heme
(59) Kaila, V. R. I.; Johansson, M. P.; Sundholm, D.; Laakkonen, L.; Proteins: Implications for Design and Prediction. BMC Struct. Biol.
Wikström, M. The Chemistry of the CuB Site in Cytochrome c 2011, 11, 13.
Oxidase and the Importance of Its Unique His-Tyr Bond. Biochim. (79) Das, P. K.; Chatterjee, S.; Samanta, S.; Dey, A. EPR, Resonance
Biophys. Acta - Bioenerg. 2009, 1787, 221−233. Raman, and DFT Calculations on Thiolate- and Imidazole-Bound
(60) Swygert, S. G.; Senapati, S.; Bolukbasi, M. F.; Wolfe, S. A.; Iron(III) Porphyrin Complexes: Role of the Axial Ligand in Tuning
Lindsay, S.; Peterson, C. L. SIR Proteins Create Compact the Electronic Structure. Inorg. Chem. 2012, 51, 10704−10714.
Heterochromatin Fibers. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, (80) Chatterjee, S.; Sengupta, K.; Hematian, S.; Karlin, K. D.; Dey,
12447−12452. A. Electrocatalytic O2-Reduction by Synthetic Cytochrome c Oxidase
(61) Crane, B. R.; Siegel, L. M.; Getzoff, E. D. Probing the Catalytic Mimics: Identification of a “Bridging Peroxo” Intermediate Involved
Mechanism of Sulfite Reductase by X-Ray Crystallography: Structures in Facile 4e−‑/4H+ O2-Reduction. J. Am. Chem. Soc. 2015, 137,
of the Escherichia Coli Hemoprotein in Complex with Substrates, 12897−12905.
Inhibitors, Intermediates, and Products,. Biochemistry 1997, 36, (81) Bandyopadhyay, S.; Rana, A.; Mittra, K.; Samanta, S.; Sengupta,
12120−12137. K.; Dey, A. Effect of Axial Ligand, Spin State, and Hydrogen Bonding
(62) Murphy, M. J.; Siegel, L. M.; Tove, S. R.; Kamin, H. Siroheme: on the Inner-Sphere Reorganization Energies of Functional Models of
A New Prosthetic Group Participating in Six-Electron Reduction Cytochrome P450. Inorg. Chem. 2014, 53, 10150−10158.
Reactions Catalyzed by Both Sulfite and Nitrite Reductases. Proc. (82) Sigfridsson, E.; Olsson, M. H. M.; Ryde, U. A Comparison of
Natl. Acad. Sci. U. S. A. 1974, 71, 612−616. the Inner-Sphere Reorganization Energies of Cytochromes, Iron-
(63) Demenocal, P. B. Plio-Pleistocene African Climate. Science Sulfur Clusters, and Blue Copper Proteins. J. Phys. Chem. B 2001, 105,
1995, 270, 53−59. 5546−5552.
(64) Brânzanic, A. M. V; Ryde, U.; Silaghi-Dumitrescu, R. Why Does (83) Xie, L.; Zhang, X.-P.; Zhao, B.; Li, P.; Qi, J.; Guo, X.; Wang, B.;
Sulfite Reductase Employ Siroheme? Chem. Commun. 2019, 55, Lei, H.; Zhang, W.; Apfel, U. P.; Cao, R. Enzyme-Inspired Iron
14047−14049. Porphyrins for Improved Electrocatalytic Oxygen Reduction and
(65) Smith, K. W.; Stroupe, M. E. Mutational Analysis of Sulfite Evolution Reactions. Angew. Chemie Int. Ed. 2021, 60, 7576−7581.
(84) Avalos, J. L.; Boeke, J. D.; Wolberger, C. Structural Basis for the
Reductase Hemoprotein Reveals the Mechanism for Coordinated
Mechanism and Regulation of Sir2 Enzymes. Mol. Cell 2004, 13,
Electron and Proton Transfer. Biochemistry 2012, 51, 9857−9868.
639−648.
(66) Einsle, O.; Messerschmidt, A.; Huber, R.; Kroneck, P. M. H.;
(85) Crane, B. R.; Siegel, L. M.; Getzoff, E. D. Structures of the
Neese, F. Mechanism of the Six-Electron Reduction of Nitrite to
Siroheme- and Fe4S4-Containing Active Center of Sulfite Reductase
Ammonia by Cytochrome c Nitrite Reductase. J. Am. Chem. Soc.
in Different States of Oxidation: Heme Activation via Reduction-
2002, 124, 11737−11745.
Gated Exogenous Ligand Exchange,. Biochemistry 1997, 36, 12101−
(67) Rinaldo, S.; Giardina, G.; Castiglione, N.; Stelitano, V.;
12119.
Cutruzzolà, F. The Catalytic Mechanism of Pseudomonas Aeruginosa
(86) Liu, Y.; Wang, F.; Li, P.; Tan, X. Insights into the Mechanistic
Cd1 Nitrite Reductase. Biochem. Soc. Trans. 2011, 39, 195−200. Role of the [Fe4S4] Cubane in the A-Cluster {[Fe4S4]-(SR)-
(68) Maia, L. B.; Moura, J. J. G. How Biology Handles Nitrite. Chem. [NipNid]} of Acetyl-Coenzyme A Synthase. ChemBioChem. 2011, 12,
Rev. 2014, 114, 5273−5357. 1417−1421.
(69) Shiro, Y. Structure and Function of Bacterial Nitric Oxide (87) Stiebritz, M. T. A Role for [Fe4S4] Clusters in TRNA
Reductases: Nitric Oxide Reductase, Anaerobic Enzymes. Biochim. Recognition�a Theoretical Study. Nucleic Acids Res. 2014, 42, 5426−
Biophys. Acta - Bioenerg. 2012, 1817, 1907−1913. 5435.
(70) Brown, K.; Djinovic-Carugo, K.; Haltia, T.; Cabrito, I.; Saraste, (88) Mittra, K.; Chatterjee, S.; Samanta, S.; Sengupta, K.;
M.; Moura, J. G.; Moura, I.; Tegoni, M.; Cambillau, C. Revisiting the Bhattacharjee, H.; Dey, A. A Hydrogen Bond Scaffold Supported
Catalytic CuZ Cluster of Nitrous Oxide (N2O) Reductase: Synthetic Heme FeIII-O2− Adduct. Chem. Commun. 2012, 48, 10535−
EVIDENCE OF A BRIDGING INORGANIC SULFUR*. J. Biol. 10537.
Chem. 2000, 275, 41133−41136. (89) Yoshikawa, S.; Shimada, A. Reaction Mechanism of
(71) Amanullah, S.; Dey, A. The Role of Porphyrin Peripheral Cytochrome c Oxidase. Chem. Rev. 2015, 115, 1936−1989.
Substituents in Determining the Reactivities of Ferrous Nitrosyl (90) Babcock, G. T.; Wikström, M. Oxygen Activation and the
Species. Chem. Sci. 2020, 11, 5909−5921. Conservation of Energy in Cell Respiration. Nature 1992, 356, 301−
(72) Sengupta, K.; Chatterjee, S.; Samanta, S.; Bandyopadhyay, S.; 309.
Dey, A. Resonance Raman and Electrocatalytic Behavior of Thiolate (91) Rousseau, D. L. Two Phases of Proton Translocation. Nature
and Imidazole Bound Iron Porphyrin Complexes on Self Assembled 1999, 400, 412−413.
Monolayers: Functional Modeling of Cytochrome P450. Inorg. Chem. (92) Wikström, M.; Sharma, V. Proton Pumping by Cytochrome c
2013, 52, 2000−2014. Oxidase - A 40 year Anniversary. Biochim. Biophys. Acta - Bioenerg.
(73) Samanta, S.; Das, P. K.; Chatterjee, S.; Dey, A. Effect of Axial 2018, 1859, 692−698.
Ligands on Electronic Structure and O2 Reduction by Iron Porphyrin (93) Michel, H. Cytochrome c Oxidase: Catalytic Cycle and
Complexes: Towards a Quantitative Understanding of the “Push Mechanisms of Proton Pumping-A Discussion. Biochemistry 1999, 38,
Effect. J. Porphyr. Phthalocyanines 2015, 19, 92−108. 15129−15140.
(74) Battersby, A. R.; McDonald, E. Origin of the Pigments of Life: (94) Zaslavsky, D.; Gennis, R. B. Proton Pumping by Cytochrome
The Type-III Problem in Porphyrin Biosynthesis. Acc. Chem. Res. Oxidase: Progress, Problems and Postulates. Biochim. Biophys. Acta -
1979, 12, 14−22. Bioenerg. 2000, 1458, 164−179.
(75) Jensen, K. P.; Ryde, U. Comparison of the Chemical Properties (95) Petrosillo, G.; Ruggiero, F. M.; Pistolese, M.; Paradies, G.
of Iron and Cobalt Porphyrins and Corrins. ChemBioChem. 2003, 4, Reactive Oxygen Species Generated from the Mitochondrial Electron
413−424. Transport Chain Induce Cytochrome c Dissociation from Beef-Heart

12419 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Submitochondrial Particles via Cardiolipin Peroxidation. Possible Water: Mimicking the Active Site of Cytochrome c Oxidase.
Role in the Apoptosis. FEBS Lett. 2001, 509, 435−438. ChemBioChem. 2001, 2, 144−148.
(96) Yoshikawa, S.; Muramoto, K.; Shinzawa-Itoh, K. The O2 (113) Ricard, D.; Andrioletti, B.; Boitrel, B.; L’Her, M. Electro-
Reduction and Proton Pumping Gate Mechanism of Bovine Heart catalytic Reduction of Dioxygen to Water by Tren-Capped
Cytochrome c Oxidase. Biochim. Biophys. Acta - Bioenerg. 2011, 1807, Porphyrins, Functional Models of Cytochrome c Oxidase†. Chem.
1279−1286. Commun. 1999, 16, 1523−1524.
(97) Hematian, S.; Garcia-Bosch, I.; Karlin, K. D. Synthetic Heme/ (114) Ricard, D.; L’Her, M.; Richard, P.; Boitrel, B. Iron Porphyrins
Copper Assemblies: Toward an Understanding of Cytochrome c as Models of Cytochrome c Oxidase. Chem.�Eur. J. 2001, 7, 3291−
Oxidase Interactions with Dioxygen and Nitrogen Oxides. Acc. Chem. 3297.
Res. 2015, 48, 2462−2474. (115) Didier, A.; L’Her, M.; Boitrel, B. Substituted Tren-Capped
(98) Collman, J. P.; Wagenknecht, P. S.; Hutchison, J. E. Molecular Porphyrins: Probing the Influence of Copper in Synthetic Models of
Catalysts for Multielectron Redox Reactions of Small Molecules: The Cytochrome c Oxidase. Org. Biomol. Chem. 2003, 1, 1274−1276.
“Cofacial Metallodiporphyrin” Approach. Angew. Chem., Int. Ed. Engl. (116) Morgan, J. E.; Verkhovsky, M. I.; Palmer, G.; Wikström, M.
1994, 33, 1537−1554. Role of the PR Intermediate in the Reaction of Cytochrome c Oxidase
(99) Chang, C. J.; Loh, Z.-H.; Shi, C.; Anson, F. C.; Nocera, D. G. with O2. Biochemistry 2001, 40, 6882−6892.
Targeted Proton Delivery in the Catalyzed Reduction of Oxygen to (117) Michel, H.; Behr, J.; Harrenga, A.; Kannt, A. CYTO-
Water by Bimetallic Pacman Porphyrins. J. Am. Chem. Soc. 2004, 126, CHROME C OXIDASE: Structure and Spectroscopy. Annu. Rev.
10013−10020. Biophys. Biomol. Struct. 1998, 27, 329−356.
(100) Kadish, K. M.; Frémond, L.; Ou, Z.; Shao, J.; Shi, C.; Anson, (118) Svensson-Ek, M.; Abramson, J.; Larsson, G.; Törnroth, S.;
F. C.; Burdet, F.; Gros, C. P.; Barbe, J.-M.; Guilard, R. Cobalt(III) Brzezinski, P.; Iwata, S. The X-Ray Crystal Structures of Wild-Type
Corroles as Electrocatalysts for the Reduction of Dioxygen: Reactivity and EQ(I-286) Mutant Cytochrome c Oxidases from Rhodobacter
of a Monocorrole, Biscorroles, and Porphyrin-Corrole Dyads. J. Am. Sphaeroides. J. Mol. Biol. 2002, 321, 329−339.
Chem. Soc. 2005, 127, 5625−5631. (119) Tsukihara, T.; Aoyama, H.; Yamashita, E.; Tomizaki, T.;
(101) Chatterjee, S.; Sengupta, K.; Samanta, S.; Das, P. K.; Dey, A. Yamaguchi, H.; Shinzawa-Itoh, K.; Nakashima, R.; Yaono, R.;
Concerted Proton-Electron Transfer in Electrocatalytic O2 Reduction Yoshikawa, S. The Whole Structure of the 13-Subunit Oxidized
by Iron Porphyrin Complexes: Axial Ligands Tuning H/D Isotope Cytochrome c Oxidase at 2.8 Å. Science 1996, 272, 1136−1144.
Effect. Inorg. Chem. 2015, 54, 2383−2392. (120) Poirier, G. E.; Pylant, E. D. The Self-Assembly Mechanism of
(102) Chattopadhyay, S.; Mukherjee, M.; Kandemir, B.; Bowman, S. Alkanethiols on Au(111). Science 1996, 272, 1145−1148.
E. J.; Bren, K. L.; Dey, A. Contributions to Cytochrome c Inner- and (121) Nuzzo, R. G.; Allara, D. L. Adsorption of Bifunctional Organic
Outer-Sphere Reorganization Energy. Chem. Sci. 2021, 12, 11894− Disulfides on Gold Surfaces. J. Am. Chem. Soc. 1983, 105, 4481−4483.
11913. (122) Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D.
(103) Bard, A. J. Elements of Molecular and Biomolecular Spontaneously Organized Molecular Assemblies. 4. Structural
Electrochemistry: An Electrochemical Approach to Electron Transfer Characterization of n-Alkyl Thiol Monolayers on Gold by Optical
Chemistry By Jean-Michel Savéant (Université de Paris 7, Denis
Ellipsometry, Infrared Spectroscopy, and Electrochemistry. J. Am.
Diderot). J. Wiley & Sons, Inc.: Hoboken, NJ. 2006. Xviii + 486 Pp.
Chem. Soc. 1987, 109, 3559−3568.
$135. ISBN 0−471. J. Am. Chem. Soc. 2007, 129, 242.
(123) Chidsey, C. E. D. Free Energy and Temperature Dependence
(104) Wilhelm, E.; Battino, R.; Wilcock, R. J. Low-Pressure
of Electron Transfer at the Metal-Electrolyte Interface. Science 1991,
Solubility of Gases in Liquid Water. Chem. Rev. 1977, 77, 219−262.
251, 919−922.
(105) McCrory, C. C. L.; Ottenwaelder, X.; Stack, T. D. P.; Chidsey,
(124) Bain, C. D.; Evall, J.; Whitesides, G. M. Formation of
C. E. D. Kinetic and Mechanistic Studies of the Electrocatalytic
Monolayers by the Coadsorption of Thiols on Gold: Variation in the
Reduction of O2 to H2O with Mononuclear Cu Complexes of
Head Group, Tail Group, and Solvent. J. Am. Chem. Soc. 1989, 111,
Substituted 1,10-Phenanthrolines. J. Phys. Chem. A 2007, 111,
12641−12650. 7155−7164.
(106) Zhang, Y.; Wilson, G. S. Electrochemical Oxidation of H2O2 (125) Collman, J. P.; Devaraj, N. K.; Decréau, R. A.; Yang, Y.; Yan,
on Pt and Pt + Ir Electrodes in Physiological Buffer and Its Yi-L.; Ebina, W.; Eberspacher, T. A.; Chidsey, C. E. D.; Cytochrome c
Applicability to H2O2-Based Biosensors. J. Electroanal. Chem. 1993, Oxidase Model Catalyzes Oxygen to Water Reduction Under Rate-
345, 253−271. Limiting Electron Flux. Science 2007, 315, 1565−1568.
(107) Lee, S. J.; Pyun, S. I.; Lee, S. K.; Kang, S. J. L. Fundamentals of (126) Collman, J. P.; Decréau, R. A.; Lin, H.; Hosseini, A.; Yang, Y.;
Rotating Disc and Ring-Disc Electrode Techniques and Their Dey, A.; Eberspacher, T. A. Role of a Distal Pocket in the Catalytic O2
Applications to Study of the Oxygen Reduction Mechanism at Pt/C Reduction by Cytochrome c Oxidase Models Immobilized on
Electrode for Fuel Cells. Isr. J. Chem. 2008, 48, 215−228. Interdigitated Array Electrodes. Proc. Natl. Acad. Sci. U. S. A. 2009,
(108) Santhanagopalan, S.; White, R. E. Estimating Parameters from 106, 7320−7323.
Rotating Ring Disc Electrode Measurements. Russ. J. Electrochem. (127) Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.;
2017, 53, 1087−1099. Whitesides, G. M. Self-Assembled Monolayers of Thiolates on Metals
(109) Collman, J. P.; Fu, L.; Herrmann, P. C.; Zhang, X. A as a Form of Nanotechnology. Chem. Rev. 2005, 105, 1103−1170.
Functional Model Related to Cytochrome c Oxidase and Its (128) Collman, J. P.; Decréau, R. A.; Yan, Y.; Yoon, J.; Solomon, E.
Electrocatalytic Four-Electron Reduction of O2. Science 1997, 275, I. Intramolecular Single-Turnover Reaction in a Cytochrome c
949−951. Oxidase Model Bearing a Tyr244 Mimic. J. Am. Chem. Soc. 2007,
(110) Collman, J. P.; Fu, L.; Herrmann, P. C.; Wang, Z.; Rapta, M.; 129, 5794−5795.
Bröring, M.; Schwenninger, R.; Boitrel, B. A Functional Model of (129) Yu, M. A.; Egawa, T.; Shinzawa-Itoh, K.; Yoshikawa, S.;
Cytochrome c Oxidase: Thermodynamic Implications. Angew. Chemie Guallar, V.; Yeh, S. R.; Rousseau, D. L.; Gerfen, G. J. Two Tyrosyl
Int. Ed. 1998, 37, 3397−3400. Radicals Stabilize High Oxidation States in Cytochrome c Oxidase for
(111) Collman, J. P.; Rapta, M.; Bröring, M.; Raptova, L.; Efficient Energy Conservation and Proton Translocation. J. Am. Chem.
Schwenninger, R.; Boitrel, B.; Fu, L.; L’Her, M. Close Structural Soc. 2012, 134, 4753−4761.
Analogues of the Cytochrome c Oxidase Fea3/CuB Center Show (130) Budiman, K.; Kannt, A.; Lyubenova, S.; Richter, O. M. H.;
Clean 4e− Electroreduction of O2 to H2O at Physiological PH. J. Am. Ludwig, B.; Michel, H.; MacMillan, F. Tyrosine 167: The Origin of
Chem. Soc. 1999, 121, 1387−1388. the Radical Species Observed in the Reaction of Cytochrome c
(112) Ricard, D.; Didier, A.; L’Her, M.; Boitrel, B. Application of 3- Oxidase with Hydrogen Peroxide in Paracoccus Denitrificans.
Quinolinoyl Picket Porphyrins to the Electroreduction of Dioxygen to Biochemistry 2004, 43, 11709−11716.

12420 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

(131) Ghiladi, R. A.; Hatwell, K. R.; Karlin, K. D.; Huang, H.; catalytic O2 Reduction by Iron Porphyrins. Proc. Natl. Acad. Sci. U. S.
Moënne-Loccoz, P.; Krebs, C.; Huynh, B. H.; Marzilli, L. A.; Cotter, A. 2013, 110, 8431−8436.
R. J.; Kaderli, S.; Zuberbühler, A. D. Dioxygen Reactivity of (146) Kim, E.; Kamaraj, K.; Galliker, B.; Rubie, N. D.; Moënne-
Mononuclear Heme and Copper Components Yielding A High- Loccoz, P.; Kaderli, S.; Zuberbühler, A. D.; Karlin, K. D. Dioxygen
Spin Heme-Peroxo-Cu Complex. J. Am. Chem. Soc. 2001, 123, 6183− Reactivity of Copper and Heme-Copper Complexes Possessing an
6184. Imidazole-Phenol Cross-Link. Inorg. Chem. 2005, 44, 1238−1247.
(132) Ghiladi, R. A.; Kretzer, R. M.; Guzei, I.; Rheingold, A. L.; (147) Denisov, I. G.; Mak, P. J.; Makris, T. M.; Sligar, S. G.; Kincaid,
Neuhold, Y.-M.; Hatwell, K. R.; Zuberbühler, A. D.; Karlin, K. D. J. R. Resonance Raman Characterization of the Peroxo and
(F 8 TPP)Fe II /O2 Reactivity Studies {F 8 TPP = Tetrakis(2,6- Hydroperoxo Intermediates in Cytochrome P450. J. Phys. Chem. A
Difluorophenyl)Porphyrinate(2-)}: Spectroscopic (UV-Visible and 2008, 112, 13172−13179.
NMR) and Kinetic Study of Solvent-Dependent (Fe/O2 = 1:1 or 2:1) (148) Gregory, M.; Mak, P. J.; Sligar, S. G.; Kincaid, J. R. Differential
Reversible O2-Reduction and Ferryl Formation. Inorg. Chem. 2001, Hydrogen Bonding in Human CYP17 Dictates Hydroxylation versus
40, 5754−5767. Lyase Chemistry. Angew. Chemie Int. Ed. 2013, 52, 5342−5345.
(133) Kim, E.; Helton, M. E.; Wasser, I. M.; Karlin, K. D.; Lu, S.; (149) Osako, T.; Ohkubo, K.; Taki, M.; Tachi, Y.; Fukuzumi, S.;
Huang, H.; Moënne-Loccoz, P.; Incarvito, C. D.; Rheingold, A. L.; Itoh, S. Oxidation Mechanism of Phenols by Dicopper-Dioxygen
Honecker, M.; Kaderli, S.; Zuberbühler, A. D. Superoxo, μ-Peroxo, (Cu2/O2) Complexes. J. Am. Chem. Soc. 2003, 125, 11027−11033.
and μ-Oxo Complexes from Heme/O2 and Heme-Cu/O2 Reactivity: (150) Warren, J. J.; Tronic, T. A.; Mayer, J. M. Thermochemistry of
Copper Ligand Influences in Cytochrome c Oxidase Models. Proc. Proton-Coupled Electron Transfer Reagents and Its Implications.
Natl. Acad. Sci. U. S. A. 2003, 100, 3623−3628. Chem. Rev. 2010, 110, 6961−7001.
(134) Kim, H.; Sharma, S. K.; Schaefer, A. W.; Solomon, E. I.; (151) Singha, A.; Dey, A. Hydrogen Atom Abstraction by Synthetic
Karlin, K. D. Heme-Cu Binucleating Ligand Supports Heme/O2 and Heme Ferric Superoxide and Hydroperoxide Species. Chem. Commun.
FeII-CuI/O2 Reactivity Providing High- and Low-Spin FeIII-Peroxo- 2019, 55, 5591−5594.
CuII Complexes. Inorg. Chem. 2019, 58, 15423−15432. (152) Singha, A.; Mondal, A.; Nayek, A.; Dey, S. G.; Dey, A. Oxygen
(135) Chufán, E. E.; Puiu, S. C.; Karlin, K. D. Heme-Copper/ Reduction by Iron Porphyrins with Covalently Attached Pendent
Dioxygen Adduct Formation, Properties, and Reactivity. Acc. Chem. Phenol and Quinol. J. Am. Chem. Soc. 2020, 142, 21810−21828.
Res. 2007, 40, 563−572. (153) Chatterjee, S.; Sengupta, K.; Mondal, B.; Dey, S.; Dey, A.
(136) Kim, E.; Shearer, J.; Lu, S.; Moënne-Loccoz, P.; Helton, M. E.; Factors Determining the Rate and Selectivity of 4e−‑/4H+ Electro-
Kaderli, S.; Zuberbü hler, A. D.; Karlin, K. D. Heme/Cu/O2 catalytic Reduction of Dioxygen by Iron Porphyrin Complexes. Acc.
Reactivity: Change in FeIII-(O22‑)-CuII Unit Peroxo Binding Chem. Res. 2017, 50, 1744−1753.
Geometry Effected by Tridentate Copper Chelation. J. Am. Chem. (154) Liu, J.-G.; Ohta, T.; Yamaguchi, S.; Ogura, T.; Sakamoto, S.;
Soc. 2004, 126, 12716−12717. Maeda, Y.; Naruta, Y. Spectroscopic Characterization of a Hydro-
(137) Schaefer, A. W.; Kieber-Emmons, M. T.; Adam, S. M.; Karlin, peroxo-Heme Intermediate: Conversion of a Side-On Peroxo to an
K. D.; Solomon, E. I. Phenol-Induced O-O Bond Cleavage in a Low-
End-On Hydroperoxo Complex. Angew. Chemie Int. Ed. 2009, 48,
Spin Heme-Peroxo-Copper Complex: Implications for O2 Reduction
9262−9267.
in Heme-Copper Oxidases. J. Am. Chem. Soc. 2017, 139, 7958−7973.
(155) Sengupta, K.; Chatterjee, S.; Dey, A. In Situ Mechanistic
(138) Adam, S. M.; Garcia-Bosch, I.; Schaefer, A. W.; Sharma, S. K.;
Investigation of O2 Reduction by Iron Porphyrin Electrocatalysts
Siegler, M. A.; Solomon, E. I.; Karlin, K. D. Critical Aspects of Heme-
Using Surface-Enhanced Resonance Raman Spectroscopy Coupled to
Peroxo-Cu Complex Structure and Nature of Proton Source Dictate
Rotating Disk Electrode (SERRS-RDE) Setup. ACS Catal. 2016, 6,
Metal-Operoxo Breakage versus Reductive O-O Cleavage Chemistry.
J. Am. Chem. Soc. 2017, 139, 472−481. 6838−6852.
(139) Halime, Z.; Kieber-Emmons, M. T.; Qayyum, M. F.; Mondal, (156) Collman, J. P.; Gagne, R. R.; Reed, C.; Halbert, T. R.; Lang,
B.; Gandhi, T.; Puiu, S. C.; Chufán, E. E.; Sarjeant, A. A. N.; G.; Robinson, W. T. Picket Fence Porphyrins. Synthetic Models for
Hodgson, K. O.; Hedman, B.; Solomon, E. I.; Karlin, K. D. Heme- Oxygen Binding Hemoproteins. J. Am. Chem. Soc. 1975, 97, 1427−
Copper-Dioxygen Complexes: Toward Understanding Ligand-Envi- 1439.
ronmental Effects on the Coordination Geometry, Electronic (157) Kossanyi, A.; Tani, F.; Nakamura, N.; Naruta, Y. Properties of
Structure, and Reactivity. Inorg. Chem. 2010, 49, 3629−3645. a Binaphthyl-Bridged Porphyrin-Iron Complex Bearing Hydroxy
(140) Noodleman, L.; Han Du, W.-G.; Fee, J. A.; Götz, A. W.; Groups inside Its Cavity. Chem.�Eur. J. 2001, 7, 2862−2872.
Walker, R. C. Linking Chemical Electron-Proton Transfer to Proton (158) Shi, C.; Anson, F. C. (5,10,15,20-Tetramethylporphyrinato)-
Pumping in Cytochrome c Oxidase: Broken-Symmetry DFT Cobalt(II): A Remarkably Active Catalyst for the Electroreduction of
Exploration of Intermediates along the Catalytic Reaction Pathway O2 to H2O. Inorg. Chem. 1998, 37, 1037−1043.
of the Iron-Copper Dinuclear Complex. Inorg. Chem. 2014, 53, 6458− (159) Shi, C.; Steiger, B.; Yuasa, M.; Anson, F. C. Electroreduction
6472. of O2 to H2O at Unusually Positive Potentials Catalyzed by the
(141) Poiana, F.; von Ballmoos, C.; Gonska, N.; Blomberg, M. R. A.; Simplest of the Cobalt Porphyrins. Inorg. Chem. 1997, 36, 4294−
Adelroth, P.; Brzezinski, P. Splitting of the O-O Bond at the Heme- 4295.
Copper Catalytic Site of Respiratory Oxidases. Sci. Adv. 2017, 3, 1−9. (160) Chang, C. J.; Deng, Y.; Shi, C.; Chang, C. K.; Anson, F. C.;
(142) Ehudin, M. A.; Schaefer, A. W.; Adam, S. M.; Quist, D. A.; Nocera, D. G. Electrocatalytic Four-Electron Reduction of Oxygen to
Diaz, D. E.; Tang, J. A.; Solomon, E. I.; Karlin, K. D. Influence of Water by a Highly Flexible Cofacial Cobalt Bisporphyrin. Chem.
Intramolecular Secondary Sphere Hydrogen-Bonding Interactions on Commun. 2000, 15, 1355−1356.
Cytochrome c Oxidase Inspired Low-Spin Heme-Peroxo-Copper (161) Tang, H.; Yin, H.; Wang, J.; Yang, N.; Wang, D.; Tang, Z.
Complexes. Chem. Sci. 2019, 10, 2893−2905. Molecular Architecture of Cobalt Porphyrin Multilayers on Reduced
(143) Chishiro, T.; Shimazaki, Y.; Tani, F.; Tachi, Y.; Naruta, Y.; Graphene Oxide Sheets for High-Performance Oxygen Reduction
Karasawa, S.; Hayami, S.; Maeda, Y. Isolation and Crystal Structure of Reaction. Angew. Chemie Int. Ed. 2013, 52, 5585−5589.
a Peroxo-Bridged Heme-Copper Complex. Angew. Chemie Int. Ed. (162) Fukuzumi, S.; Mandal, S.; Mase, K.; Ohkubo, K.; Park, H.;
2003, 42, 2788−2791. Benet-Buchholz, J.; Nam, W.; Llobet, A. Catalytic Four-Electron
(144) Shin, H.; Lee, D. H.; Kang, C.; Karlin, K. D. Electrocatalytic Reduction of O2 via Rate-Determining Proton-Coupled Electron
Four-Electron Reductions of O2 to H2O with Cytochrome c Oxidase Transfer to a Dinuclear Cobalt-μ-1,2-Peroxo Complex. J. Am. Chem.
Model Compounds. Electrochim. Acta 2003, 48, 4077−4082. Soc. 2012, 134, 9906−9909.
(145) Sengupta, K.; Chatterjee, S.; Samanta, S.; Dey, A. Direct (163) Liu, Y.; Zhou, G.; Zhang, Z.; Lei, H.; Yao, Z.; Li, J.; Lin, J.;
Observation of Intermediates Formed during Steady-State Electro- Cao, R. Significantly Improved Electrocatalytic Oxygen Reduction by

12421 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

an Asymmetrical Pacman Dinuclear Cobalt(II) Porphyrin-Porphyrin (181) Shaik, S.; Kumar, D.; de Visser, S. P.; Altun, A.; Thiel, W.
Dyad. Chem. Sci. 2020, 11, 87−96. Theoretical Perspective on the Structure and Mechanism of
(164) Fukuzumi, S.; Okamoto, K.; Tokuda, Y.; Gros, C. P.; Guilard, Cytochrome P450 Enzymes. Chem. Rev. 2005, 105, 2279−2328.
R. Dehydrogenation versus Oxygenation in Two-Electron and Four- (182) Rittle, J.; Green, M. T. Cytochrome P450 Compound I:
Electron Reduction of Dioxygen by 9-Alkyl-10-Methyl-9,10-Dihy- Capture, Characterization, and C-H Bond Activation Kinetics. Science
droacridines Catalyzed by Monomeric Cobalt Porphyrins and 2010, 330, 933−937.
Cofacial Dicobalt Porphyrins in the Presence of Perchloric Acid. J. (183) Weinberg, D. R.; Gagliardi, C. J.; Hull, J. F.; Murphy, C. F.;
Am. Chem. Soc. 2004, 126, 17059−17066. Kent, C. A.; Westlake, B. C.; Paul, A.; Ess, D. H.; McCafferty, D. G.;
(165) Fukuzumi, S.; Okamoto, K.; Gros, C. P.; Guilard, R. Meyer, T. J. Proton-Coupled Electron Transfer. Chem. Rev. 2012, 112,
Mechanism of Four-Electron Reduction of Dioxygen to Water by 4016−4093.
Ferrocene Derivatives in the Presence of Perchloric Acid in (184) Rosenthal, J.; Nocera, D. G. Role of Proton-Coupled Electron
Benzonitrile, Catalyzed by Cofacial Dicobalt Porphyrins. J. Am. Transfer in O-O Bond Activation. Acc. Chem. Res. 2007, 40, 543−553.
Chem. Soc. 2004, 126, 10441−10449. (185) Mittra, K.; Chatterjee, S.; Samanta, S.; Dey, A. Selective 4e−‑/
(166) Sinha, S.; Ghosh, M.; Warren, J. J. Changing the Selectivity of 4H+ O2 Reduction by an Iron(Tetraferrocenyl)Porphyrin Complex:
O2 Reduction Catalysis with One Ligand Heteroatom. ACS Catal. From Proton Transfer Followed by Electron Transfer in Organic
2019, 9, 2685−2691. Solvent to Proton Coupled Electron Transfer in Aqueous Medium.
(167) Anson, C. W.; Stahl, S. S. Cooperative Electrocatalytic O2 Inorg. Chem. 2013, 52, 14317−14325.
Reduction Involving Co(Salophen) with p-Hydroquinone as an (186) Chung, L. W.; Li, X.; Sugimoto, H.; Shiro, Y.; Morokuma, K.
Electron-Proton Transfer Mediator. J. Am. Chem. Soc. 2017, 139, Density Functional Theory Study on a Missing Piece in Under-
18472−18475. standing of Heme Chemistry: The Reaction Mechanism for
(168) McGuire, R., Jr; Dogutan, D. K.; Teets, T. S.; Suntivich, J.; Indoleamine 2,3-Dioxygenase and Tryptophan 2,3-Dioxygenase. J.
Shao-Horn, Y.; Nocera, D. G. Oxygen Reduction Reactivity of Am. Chem. Soc. 2008, 130, 12299−12309.
Cobalt(II) Hangman Porphyrins. Chem. Sci. 2010, 1, 411−414. (187) Basran, J.; Booth, E. S.; Lee, M.; Handa, S.; Raven, E. L.
(169) Zhang, R.; Warren, J. J. Controlling the Oxygen Reduction Analysis of Reaction Intermediates in Tryptophan 2,3-Dioxygenase: A
Selectivity of Asymmetric Cobalt Porphyrins by Using Local Comparison with Indoleamine 2,3-Dioxygenase. Biochemistry 2016,
Electrostatic Interactions. J. Am. Chem. Soc. 2020, 142, 13426−13434. 55, 6743−6750.
(170) Lv, B.; Li, X.; Guo, K.; Ma, J.; Wang, Y.; Lei, H.; Wang, F.; Jin, (188) Makino, R.; Obayashi, E.; Hori, H.; Iizuka, T.; Mashima, K.;
X.; Zhang, Q.; Zhang, W.; Long, R.; Xiong, Y.; Apfel, U. P.; Cao, R. Shiro, Y.; Ishimura, Y. Initial O2 Insertion Step of the Tryptophan
Controlling Oxygen Reduction Selectivity through Steric Effects: Dioxygenase Reaction Proposed by a Heme-Modification Study.
Electrocatalytic Two-Electron and Four-Electron Oxygen Reduction
Biochemistry 2015, 54, 3604−3616.
with Cobalt Porphyrin Atropisomers. Angew. Chemie Int. Ed. 2021, 60, (189) Dogutan, D. K.; Bediako, D. K.; Teets, T. S.; Schwalbe, M.;
12742−12746.
Nocera, D. G. Efficient Synthesis of Hangman Porphyrins. Org. Lett.
(171) Liang, Z.; Guo, H.; Zhou, G.; Guo, K.; Wang, B.; Lei, H.;
2010, 12, 1036−1039.
Zhang, W.; Zheng, H.; Apfel, U. P.; Cao, R. Metal-Organic-
(190) Rigsby, M. L.; Wasylenko, D. J.; Pegis, M. L.; Mayer, J. M.
Framework-Supported Molecular Electrocatalysis for the Oxygen
Medium Effects Are as Important as Catalyst Design for Selectivity in
Reduction Reaction. Angew. Chemie Int. Ed. 2021, 60, 8472−8476.
Electrocatalytic Oxygen Reduction by Iron-Porphyrin Complexes. J.
(172) Singha, A.; Mittra, K.; Dey, A. Effect of Hydrogen Bonding on
Innocent and Non-Innocent Axial Ligands Bound to Iron Porphyrins. Am. Chem. Soc. 2015, 137, 4296−4299.
(191) Pegis, M. L.; McKeown, B. A.; Kumar, N.; Lang, K.;
Dalt. Trans. 2019, 48, 7179−7186.
(173) Samanta, S.; Sengupta, K.; Mittra, K.; Bandyopadhyay, S.; Wasylenko, D. J.; Zhang, X. P.; Raugei, S.; Mayer, J. M. Homogenous
Dey, A. Selective Four Electron Reduction of O2 by an Iron Porphyrin Electrocatalytic Oxygen Reduction Rates Correlate with Reaction
Electrocatalyst under Fast and Slow Electron Fluxes. Chem. Commun. Overpotential in Acidic Organic Solutions. ACS Cent. Sci. 2016, 2,
2012, 48, 7631−7633. 850−856.
(174) Samanta, S.; Mittra, K.; Sengupta, K.; Chatterjee, S.; Dey, A. (192) Sinha, S.; Aaron, M. S.; Blagojevic, J.; Warren, J. J.
Second Sphere Control of Redox Catalysis: Selective Reduction of O2 Electrocatalytic Dioxygen Reduction by Carbon Electrodes Non-
to O2−‑ or H2O by an Iron Porphyrin Catalyst. Inorg. Chem. 2013, 52, covalently Modified with Iron Porphyrin Complexes: Enhancements
1443−1453. from a Single Proton Relay. Chem.�Eur. J. 2015, 21, 18072−18075.
(175) Matson, B. D.; Carver, C. T.; Von Ruden, A.; Yang, J. Y.; (193) Bordwell, F. G. Equilibrium Acidities in Dimethyl Sulfoxide
Raugei, S.; Mayer, J. M. Distant Protonated Pyridine Groups in Solution. Acc. Chem. Res. 1988, 21, 456−463.
Water-Soluble Iron Porphyrin Electrocatalysts Promote Selective (194) Tshepelevitsh, S.; Kütt, A.; Lõkov, M.; Kaljurand, I.; Saame, J.;
Oxygen Reduction to Water. Chem. Commun. 2012, 48, 11100− Heering, A.; Plieger, P. G.; Vianello, R.; Leito, I. On the Basicity of
11102. Organic Bases in Different Media. Eur. J. Org. Chem. 2019, 2019,
(176) Mukherjee, S.; Nayek, A.; Bhunia, S.; Dey, S. G.; Dey, A. A 6735−6748.
Single Iron Porphyrin Shows PH Dependent Switch between “Push” (195) YUASA, M.; STEIGER, B.; ANSON, F. C. Hydroxy-
and “Pull” Effects in Electrochemical Oxygen Reduction. Inorg. Chem. Substituted Cobalt Tetraphenylporphyrins as Electrocatalysts for the
2020, 59, 14564−14576. Reduction of O2. J. Porphyr. Phthalocyanines 1997, 1, 181−188.
(177) Hasemann, C. A.; Kurumbail, R. G.; Boddupalli, S. S.; (196) Durand, R. R.; Anson, F. C. Catalysis of Dioxygen Reduction
Peterson, J. A.; Deisenhofer, J. Structure and Function of at Graphite Electrodes by an Adsorbed Cobalt(II) Porphyrin. J.
Cytochromes P450:A Comparative Analysis of Three Crystal Electroanal. Chem. Interfacial Electrochem. 1982, 134, 273−289.
Structures. Structure 1995, 3, 41−62. (197) Bhunia, S.; Rana, A.; Roy, P.; Martin, D. J.; Pegis, M. L.; Roy,
(178) Peters, J. W.; Fisher, K.; Dean, D. R. NITROGENASE B.; Dey, A. Rational Design of Mononuclear Iron Porphyrins for
STRUCTURE AND FUNCTION: A Biochemical-Genetic Perspec- Facile and Selective 4e−‑/4H+ O2 Reduction: Activation of O-O Bond
tive. Annu. Rev. Microbiol. 1995, 49, 335−366. by 2nd Sphere Hydrogen Bonding. J. Am. Chem. Soc. 2018, 140,
(179) Hoffman, B. M.; Lukoyanov, D.; Yang, Z.-Y.; Dean, D. R.; 9444−9457.
Seefeldt, L. C. Mechanism of Nitrogen Fixation by Nitrogenase: The (198) Ghatak, A.; Bhakta, S.; Bhunia, S.; Dey, A. Influence of the
Next Stage. Chem. Rev. 2014, 114, 4041−4062. Distal Guanidine Group on the Rate and Selectivity of O2 Reduction
(180) Peters, J. W.; Szilagyi, R. K. Exploring New Frontiers of by Iron Porphyrin. Chem. Sci. 2019, 10, 9692−9698.
Nitrogenase Structure and Mechanism. Curr. Opin. Chem. Biol. 2006, (199) Boulatov, R.; Collman, J. P.; Shiryaeva, I. M.; Sunderland, C. J.
10, 101−108. Functional Analogues of the Dioxygen Reduction Site in Cytochrome

12422 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Oxidase: Mechanistic Aspects and Possible Effects of CuB. J. Am. Cytochrome c Oxidase Model. J. Am. Chem. Soc. 2010, 132, 1598−
Chem. Soc. 2002, 124, 11923−11935. 1605.
(200) Collman, J. P.; Decréau, R. A. Functional Biomimetic Models (218) Tsukihara, T.; Shimokata, K.; Katayama, Y.; Shimada, H.;
for the Active Site in the Respiratory Enzyme Cytochrome c Oxidase. Muramoto, K.; Aoyama, H.; Mochizuki, M.; Shinzawa-Itoh, K.;
Chem. Commun. 2008, 41, 5065−5076. Yamashita, E.; Yao, M.; Ishimura, Y.; Yoshikawa, S. The Low-Spin
(201) Mukherjee, S.; Mukherjee, M.; Mukherjee, A.; Bhagi- Heme of Cytochrome c Oxidase as the Driving Element of the
Damodaran, A.; Lu, Y.; Dey, A. O2 Reduction by Biosynthetic Proton-Pumping Process. Proc. Natl. Acad. Sci. U. S. A. 2003, 100,
Models of Cytochrome c Oxidase: Insights into Role of Proton 15304−15309.
Transfer Residues from Perturbed Active Sites Models of CcO. ACS (219) Miner, K. D.; Mukherjee, A.; Gao, Y. G.; Null, E. L.; Petrik, I.
Catal. 2018, 8, 8915−8924. D.; Zhao, X.; Yeung, N.; Robinson, H.; Lu, Y. A Designed Functional
(202) Decréau, R. A.; Collman, J. P.; Hosseini, A. Electrochemical Metalloenzyme That Reduces O2 to H2O with Over One Thousand
Applications. How Click Chemistry Brought Biomimetic Models to Turnovers. Angew. Chemie Int. Ed. 2012, 51, 5589−5592.
the next Level: Electrocatalysis under Controlled Rate of Electron (220) Sun, Y.; Wang, L.; Liu, H. Myoglobin Functioning as
Transfer. Chem. Soc. Rev. 2010, 39, 1291−1301. Cytochrome P450 for Biosensing of 2,4-Dichlorophenol. Anal.
(203) Finklea, H. O.; Hanshew, D. D. Electron-Transfer Kinetics in Methods 2012, 4, 3358−3363.
Organized Thiol Monolayers with Attached Pentaammine(Pyridine) (221) Raven, E. L.; Mauk, A. G. Chemical Reactivity of the Active
Ruthenium Redox Centers. J. Am. Chem. Soc. 1992, 114, 3173−3181. Site of Myoglobin. Academic Press 2000, 51, 1−50.
(204) Ghatak, A.; Bhunia, S.; Dey, A. Effect of Pendant Distal (222) Wang, N.; Zhao, X.; Lu, Y. Role of Heme Types in Heme-
Residues on the Rate and Selectivity of Electrochemical Oxygen Copper Oxidases: Effects of Replacing a Heme b with a Heme o
Reduction Reaction Catalyzed by Iron Porphyrin Complexes. ACS Mimic in an Engineered Heme-Copper Center in Myoglobin§. J. Am.
Catal. 2020, 10, 13136−13148. Chem. Soc. 2005, 127, 16541−16547.
(205) Azcarate, I.; Costentin, C.; Robert, M.; Savéant, J. M. (223) Pawate, A. S.; Morgan, J.; Namslauer, A.; Mills, D.; Brzezinski,
Through-Space Charge Interaction Substituent Effects in Molecular P.; Ferguson-Miller, S.; Gennis, R. B. A Mutation in Subunit I of
Catalysis Leading to the Design of the Most Efficient Catalyst of CO2- Cytochrome Oxidase from Rhodobacter Sphaeroides Results in an
to-CO Electrochemical Conversion. J. Am. Chem. Soc. 2016, 138, Increase in Steady-State Activity but Completely Eliminates Proton
16639−16644. Pumping. Biochemistry 2002, 41, 13417−13423.
(206) Martin, D. J.; Mercado, B. Q.; Mayer, J. M. Combining Scaling (224) Chang, H. Y.; Hemp, J.; Chen, Y.; Fee, J. A.; Gennis, R. B. The
Relationships Overcomes Rate versus Overpotential Trade-Offs in O2 Cytochrome ba3 Oxygen Reductase from Thermus Thermophilus
Molecular Electrocatalysis. Sci. Adv. 2020, 6, 3318−3325. Uses a Single Input Channel for Proton Delivery to the Active Site
(207) Wang, Y. H.; Schneider, P. E.; Goldsmith, Z. K.; Mondal, B.; and for Proton Pumping. Proc. Natl. Acad. Sci. U. S. A. 2009, 106,
Hammes-Schiffer, S.; Stahl, S. S. Brønsted Acid Scaling Relationships 16169−16173.
Enable Control Over Product Selectivity from O2 Reduction with a (225) Yu, Y.; Mukherjee, A.; Nilges, M. J.; Hosseinzadeh, P.; Miner,
Mononuclear Cobalt Porphyrin Catalyst. ACS Cent. Sci. 2019, 5, K. D.; Lu, Y. Direct EPR Observation of a Tyrosyl Radical in a
1024−1034. Functional Oxidase Model in Myoglobin during Both H2O2 and O2
(208) Martin, D. J.; Johnson, S. I.; Mercado, B. Q.; Raugei, S.; Reactions. J. Am. Chem. Soc. 2014, 136, 1174−1177.
Mayer, J. M. Intramolecular Electrostatic Effects on O2, CO2, and (226) MacMillan, F.; Kannt, A.; Behr, J.; Prisner, T.; Michel, H.
Acetate Binding to a Cationic Iron Porphyrin. Inorg. Chem. 2020, 59, Direct Evidence for a Tyrosine Radical in the Reaction of
17402−17414. Cytochrome c Oxidase with Hydrogen Peroxide. Biochemistry 1999,
(209) Martin, D. J.; Mayer, J. M. Oriented Electrostatic Effects on 38, 9179−9184.
O2 and CO2 Reduction by a Polycationic Iron Porphyrin. J. Am. (227) Petrik, I. D.; Davydov, R.; Ross, M.; Zhao, X.; Hoffman, B.;
Chem. Soc. 2021, 143, 11423−11434. Lu, Y. Spectroscopic and Crystallographic Evidence for the Role of a
(210) Martin, D. J.; Mercado, B. Q.; Mayer, J. M. All Four Water-Containing H-Bond Network in Oxidase Activity of an
Atropisomers of Iron Tetra(o-N,N,N-Trimethylanilinium)Porphyrin Engineered Myoglobin. J. Am. Chem. Soc. 2016, 138, 1134−1137.
in Both the Ferric and Ferrous States. Inorg. Chem. 2021, 60, 5240− (228) Davydov, R.; Macdonald, I. D. G.; Makris, T. M.; Sligar, S. G.;
5251. Hoffman, B. M. EPR and ENDOR of Catalytic Intermediates in
(211) Nastri, F.; Chino, M.; Maglio, O.; Bhagi-Damodaran, A.; Lu, Cryoreduced Native and Mutant Oxy-Cytochromes P450cam:
Y.; Lombardi, A. Design and Engineering of Artificial Oxygen- Mutation-Induced Changes in the Proton Delivery System. J. Am.
Activating Metalloenzymes. Chem. Soc. Rev. 2016, 45, 5020−5054. Chem. Soc. 1999, 121, 10654−10655.
(212) Lu, Y.; Berry, S. M.; Pfister, T. D. Engineering Novel (229) Davydov, R.; Makris, T. M.; Kofman, V.; Werst, D. E.; Sligar,
Metalloproteins: Design of Metal-Binding Sites into Native Protein S. G.; Hoffman, B. M. Hydroxylation of Camphor by Reduced Oxy-
Scaffolds. Chem. Rev. 2001, 101, 3047−3080. Cytochrome P450cam: Mechanistic Implications of EPR and
(213) Sigman, J. A.; Kim, H. K.; Zhao, X.; Carey, J. R.; Lu, Y. The ENDOR Studies of Catalytic Intermediates in Native and Mutant
Role of Copper and Protons in Heme-Copper Oxidases: Kinetic Enzymes. J. Am. Chem. Soc. 2001, 123, 1403−1415.
Study of an Engineered Heme-Copper Center in Myoglobin. Proc. (230) Davydov, R.; Razeghifard, R.; Im, S.-C.; Waskell, L.; Hoffman,
Natl. Acad. Sci. U. S. A. 2003, 100, 3629−3634. B. M. Characterization of the Microsomal Cytochrome P450 2B4 O2
(214) Zhao, X.; Yeung, N.; Wang, Z.; Guo, Z.; Lu, Y. Effects of Activation Intermediates by Cryoreduction and Electron Para-
Metal Ions in the CuB Center on the Redox Properties of Heme in magnetic Resonance. Biochemistry 2008, 47, 9661−9666.
Heme-Copper Oxidases: Spectroelectrochemical Studies of an (231) Davydov, R.; Kofman, V.; Fujii, H.; Yoshida, T.; Ikeda-Saito,
Engineered Heme-Copper Center in Myoglobin. Biochemistry 2005, M.; Hoffman, B. M. Catalytic Mechanism of Heme Oxygenase
44, 1210−1214. through EPR and ENDOR of Cryoreduced Oxy-Heme Oxygenase
(215) Sigman, J. A.; Kwok, B. C.; Gengenbach, A.; Lu, Y. Design and and Its Asp 140 Mutants. J. Am. Chem. Soc. 2002, 124, 1798−1808.
Creation of a Cu(II)-Binding Site in Cytochrome c Peroxidase That (232) Davydov, R.; Kofman, V.; Nocek, J. M.; Noble, R. W.; Hui,
Mimics the CuB-Heme Center in Terminal Oxidases. J. Am. Chem. H.; Hoffman, B. M. Conformational Substates of the Oxyheme
Soc. 1999, 121, 8949−8950. Centers in α and β Subunits of Hemoglobin As Disclosed by EPR and
(216) Sigman, J. A.; Kwok, B. C.; Lu, Y. From Myoglobin to Heme- ENDOR Studies of Cryoreduced Protein. Biochemistry 2004, 43,
Copper Oxidase: Design and Engineering of a CuB Center into 6330−6338.
Sperm Whale Myoglobin. J. Am. Chem. Soc. 2000, 122, 8192−8196. (233) Kappl, R.; Höhn-Berlage, M.; Hüttermann, J.; Bartlett, N.;
(217) Lu, C.; Zhao, X.; Lu, Y.; Rousseau, D. L.; Yeh, S. R. Role of Symons, M. C. R. Electron Spin and Electron Nuclear Double
Copper Ion in Regulating Ligand Binding in a Myoglobin-Based Resonance of the [FeO2]- Centre from Irradiated Oxyhemo- and

12423 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Oxymyoglobin. Biochim. Biophys. Acta - Protein Struct. Mol. Enzymol. Solutions at Low Temperatures. Cytochrome C. FEBS Lett. 1974, 45,
1985, 827, 327−343. 256−258.
(234) Unno, M.; Chen, H.; Kusama, S.; Shaik, S.; Ikeda-Saito, M. (252) Symons, M. C. R.; Petersen, R. L. Electron Capture at the
Structural Characterization of the Fleeting Ferric Peroxo Species in Iron-Oxygen Centre in Single Crystals of Oxymyoglobin Studied by
Myoglobin: Experiment and Theory. J. Am. Chem. Soc. 2007, 129, Electron Spin Resonance Spectroscopy. Biochim. Biophys. Acta -
13394−13395. Protein Struct. 1978, 535, 241−246.
(235) Liu, X.; Yu, Y.; Hu, C.; Zhang, W.; Lu, Y.; Wang, J. Significant (253) Baek, H. K.; Van Wart, H. E. Elementary Steps in the
Increase of Oxidase Activity through the Genetic Incorporation of a Formation of Horseradish Peroxidase Compound I: Direct Observa-
Tyrosine-Histidine Cross-Link in a Myoglobin Model of Heme- tion of Compound 0, a New Intermediate with a Hyperporphyrin
Copper Oxidase. Angew. Chemie Int. Ed. 2012, 51, 4312−4316. Spectrum. Biochemistry 1989, 28, 5714−5719.
(236) Mukherjee, S.; Mukherjee, A.; Bhagi-Damodaran, A.; (254) Dunford, H. B.; Stillman, J. S. On the Function and
Mukherjee, M.; Lu, Y.; Dey, A. A Biosynthetic Model of Cytochrome Mechanism of Action of Peroxidases. Coord. Chem. Rev. 1976, 19,
c Oxidase as an Electrocatalyst for Oxygen Reduction. Nat. Commun. 187−251.
2015, 6, 8467. (255) Dolphin, D.; Forman, A.; Borg, D. C.; Fajer, J.; Felton, R. H.
(237) Mukherjee, S.; Sengupta, K.; Das, M. R.; Jana, S. S.; Dey, A. Compounds I of Catalase and Horse Radish Peroxidase: π-Cation
Site-Specific Covalent Attachment of Heme Proteins on Self- Radicals. Proc. Natl. Acad. Sci. U. S. A. 1971, 68, 614−618.
Assembled Monolayers. J. Biol. Inorg. Chem. 2012, 17, 1009−1023. (256) Poulos, T. L.; Kraut, J. The Stereochemistry of Peroxidase
(238) Haas, A. S.; Pilloud, D. L.; Reddy, K. S.; Babcock, G. T.; Catalysis. J. Biol. Chem. 1980, 255, 8199−8205.
Moser, C. C.; Blasie, J. K.; Dutton, P. L. Cytochrome c and (257) Jones, P.; Dunford, H. B. The Mechanism of Compound I
Cytochrome c Oxidase: Monolayer Assemblies and Catalysis. J. Phys. Formation Revisited. J. Inorg. Biochem. 2005, 99, 2292−2298.
Chem. B 2001, 105, 11351−11362. (258) Groves, J. T.; Boaz, N. C.; Fishing for Peroxidase Protons.
(239) Friedrich, M. G.; Plum, M. A.; Santonicola, M. G.; Kirste, V. Science 2014, 345, 142−143.
U.; Knoll, W.; Ludwig, B.; Naumann, R. L. C. In Situ Monitoring of (259) Groenhof, A. R.; Ehlers, A. W.; Lammertsma, K. Proton
the Catalytic Activity of Cytochrome c Oxidase in a Biomimetic Assisted Oxygen-Oxygen Bond Splitting in Cytochrome P450. J. Am.
Architecture. Biophys. J. 2008, 95, 1500−1510. Chem. Soc. 2007, 129, 6204−6209.
(240) Su, L.; Kelly, J.; Hawkridge, F. M. Electroreduction of O2 on (260) Sundaramoorthy, M.; Terner, J.; Poulos, T. L. The Crystal
Cytochrome c Oxidase Modified Electrode for Biofuel Cell. ECS Structure of Chloroperoxidase: A Heme Peroxidase-Cytochrome
Trans. 2007, 2, 1−6. P450 Functional Hybrid. Structure 1995, 3, 1367−1378.
(241) Varotsis, C.; Zhang, Y.; Appelman, E. H.; Babcock, G. T. (261) Yi, X.; Conesa, A.; Punt, P. J.; Hager, L. P. Examining the Role
Resolution of the Reaction Sequence during the Reduction of O2 by of Glutamic Acid 183 in Chloroperoxidase Catalysis*. J. Biol. Chem.
Cytochrome Oxidase. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 237− 2003, 278, 13855−13859.
241. (262) Lee, D.-S.; Yamada, A.; Sugimoto, H.; Matsunaga, I.; Ogura,
(242) Han, S.; Takahashi, S.; Rousseau, D. L. Time Dependence of H.; Ichihara, K.; Adachi, S.; Park, S.-Y.; Shiro, Y. Substrate
the Catalytic Intermediates in Cytochromec Oxidase*. J. Biol. Chem. Recognition and Molecular Mechanism of Fatty Acid Hydroxylation
2000, 275, 1910−1919. by Cytochrome P450 from Bacillus Subtilis: CRYSTALLO-
(243) Han, S.; Ching, Y.; Rousseau, D. L. Ferryl and Hydroxy GRAPHIC, SPECTROSCOPIC, AND MUTATIONAL STUDIES*.
Intermediates in the Reaction of Oxygen with Reduced Cytochrome c J. Biol. Chem. 2003, 278, 9761−9767.
Oxidase. Nature 1990, 348, 89−90. (263) Dawson, J. H. Probing Structure-Function Relations in Heme-
(244) Zumft, W. G. Nitric Oxide Reductases of Prokaryotes with Containing Oxygenases and Peroxidases. Science 1988, 240, 433−439.
Emphasis on the Respiratory, Heme-Copper Oxidase Type. J. Inorg. (264) Adachi, S.; Nagano, S.; Ishimori, K.; Watanabe, Y.; Morishima,
Biochem. 2005, 99, 194−215. I.; Egawa, T.; Kitagawa, T.; Makino, R. Roles of Proximal Ligand in
(245) van der Oost, J.; De Boer, A. P. N.; de Gier, J.-W. L.; Zumft, Heme Proteins: Replacement of Proximal Histidine of Human
W. G.; Stouthamer, A. H.; Van Spanning, R. J. M. The Heme-Copper Myoglobin with Cysteine and Tyrosine by Site-Directed Mutagenesis
Oxidase Family Consists of Three Distinct Types of Terminal as Models for P-450, Chloroperoxidase, and Catalase. Biochemistry
Oxidases and Is Related to Nitric Oxide Reductase. FEMS Microbiol. 1993, 32, 241−252.
Lett. 1994, 121, 1−9. (265) Poulos, T. L.; Freer, S. T.; Alden, R. A.; Edwards, S. L.;
(246) Lin, Y.-W.; Yeung, N.; Gao, Y.-G.; Miner, K. D.; Tian, S.; Skogland, U.; Takio, K.; Eriksson, B.; Xuong, N.; Yonetani, T.; Kraut,
Robinson, H.; Lu, Y. Roles of Glutamates and Metal Ions in a J. The Crystal Structure of Cytochrome c Peroxidase. J. Biol. Chem.
Rationally Designed Nitric Oxide Reductase Based on Myoglobin. 1980, 255, 575−580.
Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 8581−8586. (266) Erman, J. E.; Vitello, L. B.; Miller, M. A.; Shaw, A.; Brown, K.
(247) Yeung, N.; Lin, Y.-W.; Gao, Y.-G.; Zhao, X.; Russell, B. S.; Lei, A.; Kraut, J. Histidine 52 Is a Critical Residue for Rapid Formation of
L.; Miner, K. D.; Robinson, H.; Lu, Y. Rational Design of a Structural Cytochrome c Peroxidase Compound I. Biochemistry 1993, 32, 9798−
and Functional Nitric Oxide Reductase. Nature 2009, 462, 1079− 9806.
1082. (267) Vitello, L. B.; Erman, J. E.; Miller, M. A.; Wang, J.; Kraut, J.
(248) Vidakovic, M.; Sligar, S. G.; Li, H.; Poulos, T. L. Effect of Arginine-48 Replacement on the Reaction between
Understanding the Role of the Essential Asp251 in Cytochrome Cytochrome c Peroxidase and Hydrogen Peroxide. Biochemistry
P450cam Using Site-Directed Mutagenesis, Crystallography, and 1993, 32, 9807−9818.
Kinetic Solvent Isotope Effect. Biochemistry 1998, 37, 9211−9219. (268) Gerber, N. C.; Sligar, S. G. Catalytic Mechanism of
(249) Efimov, I.; Badyal, S. K.; Metcalfe, C. L.; Macdonald, I.; Cytochrome P-450: Evidence for a Distal Charge Relay. J. Am.
Gumiero, A.; Raven, E. L.; Moody, P. C. E. Proton Delivery to Ferryl Chem. Soc. 1992, 114, 8742−8743.
Heme in a Heme Peroxidase: Enzymatic Use of the Grotthuss (269) Harris, D. L.; Loew, G. H. A Role for Thr 252 in Cytochrome
Mechanism. J. Am. Chem. Soc. 2011, 133, 15376−15383. P450cam Oxygen Activation. J. Am. Chem. Soc. 1994, 116, 11671−
(250) Sengupta, K.; Chatterjee, S.; Mukherjee, S.; Dey, S. G.; Dey, 11674.
A. Heme Bound Amylin Self-Assembled Monolayers on an Au (270) Schlichting, I.; Berendzen, J.; Chu, K.; Stock, A. M.; Maves, S.
Electrode: An Efficient Bio-Electrode for O2 Reduction to H2O. A.; Benson, D. E.; Sweet, R. M.; Ringe, D.; Petsko, G. A.; Sligar, S. G.
Chem. Commun. 2014, 50, 3806−3809. The Catalytic Pathway of Cytochrome P450cam at Atomic
(251) Blumenfeld, L. A.; Davydov, R. M.; Fel’, N. S.; Magonov, S. Resolution. Science 2000, 287, 1615−1622.
N.; Vilu, R. O. Studies on the Conformational Changes of (271) Higuchi, T.; Shimada, K.; Maruyama, N.; Hirobe, M.
Metalloproteins Induced by Electrons in Water�Ethylene Glycol Heterolytic Oxygen-Oxygen Bond Cleavage of Peroxy Acid and

12424 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

Effective Alkane Hydroxylation in Hydrophobic Solvent Mediated by Transient- and Steady-State Kinetics. Biochemistry 1997, 36, 9349−
an Iron Porphyrin Coordinated by Thiolate Anion as a Model for 9355.
Cytochrome P-450. J. Am. Chem. Soc. 1993, 115, 7551−7552. (291) Bovaird, J. H.; Ngo, T. T.; Lenhoff, H. M. Optimizing the O-
(272) Groves, J. T.; Watanabe, Y. Reactive Iron Porphyrin Phenylenediamine Assay for Horseradish Peroxidase: Effects of
Derivatives Related to the Catalytic Cycles of Cytochrome P-450 Phosphate and pH, Substrate and Enzyme Concentrations, and
and Peroxidase. Studies of the Mechanism of Oxygen Activation. J. Stopping Reagents. Clin. Chem. 1982, 28, 2423−2426.
Am. Chem. Soc. 1988, 110, 8443−8452. (292) Traylor, T. G.; Lee, W. A.; Stynes, D. V. Model Compound
(273) Bell, S. R.; Groves, J. T. A Highly Reactive P450 Model Studies Related to Peroxidases. Mechanisms of Reactions of Hemins
Compound I. J. Am. Chem. Soc. 2009, 131, 9640−9641. with Peracids. J. Am. Chem. Soc. 1984, 106, 755−764.
(274) Soper, J. D.; Kryatov, S. V.; Rybak-Akimova, E. V.; Nocera, D. (293) Das, S. K.; Chaudhury, P. K.; Biswas, D.; Sarkar, S. Modeling
G. Proton-Directed Redox Control of O-O Bond Activation by Heme for the Active Site of Sulfite Oxidase: Synthesis, Characterization, and
Hydroperoxidase Models. J. Am. Chem. Soc. 2007, 129, 5069−5075. Reactivity of [MoVIO2(Mnt)2]2- (Mnt2- = 1,2-Dicyanoethylenedi-
(275) Huang, X.; Groves, J. T. Oxygen Activation and Radical thiolate). J. Am. Chem. Soc. 1994, 116, 9061−9070.
Transformations in Heme Proteins and Metalloporphyrins. Chem. (294) Vitale, R.; Lista, L.; Cerrone, C.; Caserta, G.; Chino, M.;
Rev. 2018, 118, 2491−2553. Maglio, O.; Nastri, F.; Pavone, V.; Lombardi, A. An Artificial Heme-
(276) Groves, J. T.; Watanabe, Y. Oxygen Activation by Metal- Enzyme with Enhanced Catalytic Activity: Evolution, Functional
loporphyrins Related to Peroxidase and Cytochrome P-450. Direct Screening and Structural Characterization. Org. Biomol. Chem. 2015,
Observation of the Oxygen-Oxygen Bond Cleavage Step. J. Am. Chem. 13, 4859−4868.
Soc. 1986, 108, 7834−7836. (295) Rodríguez-López, J. N.; Gilabert, M. A.; Tudela, J.; Thorneley,
(277) Groves, J. T.; Watanabe, Y.; McMurry, T. J. Oxygen R. N. F.; García-Cánovas, F. Reactivity of Horseradish Peroxidase
Activation by Metalloporphyrins. Formation and Decomposition of Compound II toward Substrates: Kinetic Evidence for a Two-Step
an Acylperoxymanganese(III) Complex. J. Am. Chem. Soc. 1983, 105, Mechanism. Biochemistry 2000, 39, 13201−13209.
4489−4490. (296) Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.;
(278) Groves, J. T.; Watanabe, Y. Heterolytic and Homolytic Wang, T.; Feng, J.; Yang, D.; Perrett, S.; Yan, X. Intrinsic Peroxidase-
Oxygen-Oxygen Bond Cleavage Reactions of Acylperoxomanganese- like Activity of Ferromagnetic Nanoparticles. Nat. Nanotechnol. 2007,
(III) Porphyrins. Inorg. Chem. 1986, 25, 4808−4810. 2, 577−583.
(279) Yamaguchi, K.; Watanabe, Y.; Morishima, I. Push Effect on (297) Hayashi, T.; Hitomi, Y.; Ando, T.; Mizutani, T.; Hisaeda, Y.;
the Heterolytic Oxygen-Oxygen Bond Cleavage of Peroxoiron(III) Kitagawa, S.; Ogoshi, H. Peroxidase Activity of Myoglobin Is
Porphyrin Adducts. Inorg. Chem. 1992, 31, 156−157. Enhanced by Chemical Mutation of Heme-Propionates. J. Am.
(280) Yamaguchi, K.; Watanabe, Y.; Morishima, I. Direct Chem. Soc. 1999, 121, 7747−7750.
Observation of the Push Effect on the Oxygen-Oxygen Bond (298) Watkins, D. W.; Jenkins, J. M. X.; Grayson, K. J.; Wood, N.;
Steventon, J. W.; Le Vay, K. K.; Goodwin, M. I.; Mullen, A. S.; Bailey,
Cleavage of Acylperoxoiron(III) Porphyrin Complexes. J. Am.
H. J.; Crump, M. P.; MacMillan, F.; Mulholland, A. J.; Cameron, G.;
Chem. Soc. 1993, 115, 4058−4065.
Sessions, R. B.; Mann, S.; Anderson, J. L. R. Construction and in Vivo
(281) Watanabe, Y.; Yamaguchi, K.; Morishima, I.; Takehira, K.;
Assembly of a Catalytically Proficient and Hyperthermostable de
Shimizu, M.; Hayakawa, T.; Orita, H. Remarkable Solvent Effect on
Novo Enzyme. Nat. Commun. 2017, 8, 358.
the Shape-Selective Oxidation of Olefins Catalyzed by Iron(III)
(299) Savenkova, M. I.; Kuo, J. M.; Ortiz de Montellano, P. R.
Porphyrins. Inorg. Chem. 1991, 30, 2581−2582.
Improvement of Peroxygenase Activity by Relocation of a Catalytic
(282) Yeh, C. Y.; Chang, C. J.; Nocera, D. G. “Hangman”
Histidine within the Active Site of Horseradish Peroxidase.
Porphyrins for the Assembly of a Model Heme Water Channel. J.
Biochemistry 1998, 37, 10828−10836.
Am. Chem. Soc. 2001, 123, 1513−1514. (300) Cochran, A. G.; Schultz, P. G. Peroxidase Activity of an
(283) Chang, C. J.; Chng, L. L.; Nocera, D. G. Proton-Coupled O-O Antibody-Heme Complex. J. Am. Chem. Soc. 1990, 112, 9414−9415.
Activation on a Redox Platform Bearing a Hydrogen-Bonding (301) Dallacosta, C.; Monzani, E.; Casella, L. Reactivity Study on
Scaffold. J. Am. Chem. Soc. 2003, 125, 1866−1876. Microperoxidase-8. JBIC J. Biol. Inorg. Chem. 2003, 8, 770−776.
(284) Chng, L. L.; Chang, C. J.; Nocera, D. G. Catalytic O-O (302) Ortiz de Montellano, P. R.; David, S. K.; Ator, M. A.; Tew, D.
Activation Chemistry Mediated by Iron Hangman Porphyrins with a Mechanism-Based Inactivation of Horseradish Peroxidase by Sodium
Wide Range of Proton-Donating Abilities. Org. Lett. 2003, 5, 2421− Azide. Formation of Meso-Azidoprotoporphyrin IX. Biochemistry
2424. 1988, 27, 5470−5476.
(285) Jin, N.; Lahaye, D. E.; Groves, J. T. A “Push-Pull” Mechanism (303) Bhunia, S.; Rana, A.; Dey, S. G.; Ivancich, A.; Dey, A. A
for Heterolytic O-O Bond Cleavage in Hydroperoxo Manganese Designed Second-Sphere Hydrogen-Bond Interaction That Critically
Porphyrins. Inorg. Chem. 2010, 49, 11516−11524. Influences the O-O Bond Activation for Heterolytic Cleavage in
(286) Batinić-Haberle, I.; Spasojević, I.; Hambright, P.; Benov, L.; Ferric Iron-Porphyrin Complexes. Chem. Sci. 2020, 11, 2681−2695.
Crumbliss, A. L.; Fridovich, I. Relationship among Redox Potentials, (304) Garcia-Bosch, I.; Sharma, S. K.; Karlin, K. D. A Selective
Proton Dissociation Constants of Pyrrolic Nitrogens, and in Vivo and Stepwise Heme Oxygenase Model System: An Iron(IV)-Oxo
in Vitro Superoxide Dismutating Activities of Manganese(III) and Porphyrin π-Cation Radical Leads to a Verdoheme-Type Compound
Iron(III) Water-Soluble Porphyrins. Inorg. Chem. 1999, 38, 4011− via an Isoporphyrin Intermediate. J. Am. Chem. Soc. 2013, 135,
4022. 16248−16251.
(287) Lahaye, D.; Groves, J. T. Modeling the Haloperoxidases: (305) Cong, Z.; Kurahashi, T.; Fujii, H. Formation of Iron(III)
Reversible Oxygen Atom Transfer between Bromide Ion and an Oxo- Meso-Chloro-Isoporphyrin as a Reactive Chlorinating Agent from
Mn(V) Porphyrin. J. Inorg. Biochem. 2007, 101, 1786−1797. Oxoiron(IV) Porphyrin π-Cation Radical. J. Am. Chem. Soc. 2012,
(288) Bhakta, S.; Nayek, A.; Roy, B.; Dey, A. Induction of Enzyme- 134, 4469−4472.
like Peroxidase Activity in an Iron Porphyrin Complex Using Second (306) Davies, D. M.; Jones, P.; Mantle, D. The Kinetics of
Sphere Interactions. Inorg. Chem. 2019, 58, 2954−2964. Formation of Horseradish Peroxidase Compound I by Reaction with
(289) Josephy, P. D.; Eling, T.; Mason, R. P. The Horseradish Peroxobenzoic Acids. PH and Peroxo Acid Substituent Effects.
Peroxidase-Catalyzed Oxidation of 3,5,3′,5′-Tetramethylbenzidine. Biochem. J. 1976, 157, 247−253.
Free Radical and Charge-Transfer Complex Intermediates. J. Biol. (307) Savenkova, M. I.; Newmyer, S. L.; Ortiz de Montellano, P. R.
Chem. 1982, 257, 3669−3675. Rescue of His-42 → Ala Horseradish Peroxidase by a Phe-41 → His
(290) Marquez, L. A.; Dunford, H. B. Mechanism of the Oxidation Mutation: ENGINEERING OF A SURROGATE CATALYTIC
of 3,5,3‘,5‘-Tetramethylbenzidine by Myeloperoxidase Determined by HISTIDINE*. J. Biol. Chem. 1996, 271, 24598−24603.

12425 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426
Chemical Reviews pubs.acs.org/CR Review

(308) Derat, E.; Shaik, S.; Rovira, C.; Vidossich, P.; Alfonso-Prieto,
M. The Effect of a Water Molecule on the Mechanism of Formation
of Compound 0 in Horseradish Peroxidase. J. Am. Chem. Soc. 2007,
129, 6346−6347.

Recommended by ACS
Amine Groups in the Second Sphere of Iron Porphyrins
Allow for Higher and Selective 4e–/4H+ Oxygen Reduction
Rates at Lower Overpotentials
Sarmistha Bhunia, Abhishek Dey, et al.
FEBRUARY 06, 2023
JOURNAL OF THE AMERICAN CHEMICAL SOCIETY READ

Origin of the Distinctive Electronic Structure of Co- and Fe-


Porphyrin-Nitrene and Its Effect on Their Nitrene Transfer
Reactivity
Mayank Mahajan and Bhaskar Mondal
MARCH 28, 2023
INORGANIC CHEMISTRY READ

Second-Sphere Hydrogen-Bond Donors and Acceptors Affect


the Rate and Selectivity of Electrochemical Oxygen
Reduction by Iron Porphyrins Differently
Arnab Ghatak, Abhishek Dey, et al.
AUGUST 08, 2022
INORGANIC CHEMISTRY READ

Thermally Stable Redox Noninnocent Bathocuproine-Iron


Complex for Cycloaddition Reactions
Mae Féo, Guillaume Lefèvre, et al.
MARCH 27, 2023
ACS CATALYSIS READ
Get More Suggestions >

12426 https://doi.org/10.1021/acs.chemrev.1c01021
Chem. Rev. 2022, 122, 12370−12426

You might also like