Copper Bioinorganic Chemistry From Health To Bioinspired Catalysis (Simaan J.A., Réglier M. (Ed.) )
Copper Bioinorganic Chemistry From Health To Bioinspired Catalysis (Simaan J.A., Réglier M. (Ed.) )
Copper Bioinorganic Chemistry From Health To Bioinspired Catalysis (Simaan J.A., Réglier M. (Ed.) )
Chemistry
From Health to Bioinspired Catalysis
Copper Bioinorganic
Chemistry
From Health to Bioinspired Catalysis
Editors
A Jalila Simaan • Marius Réglier
Centre Nationale de la Recherche Scientifique, France &
Aix Marseille Université, France
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_fmatter
Preface
v
vi Preface
A Jalila Simaan
Marius Réglier
July 2023
viii Preface
References
1. Kitajima, N., Fujisawa, K., Morooka, Y., Toriumi, K. m-η2:η2-Peroxo binuclear copper
complex, [Cu(HB(3,5-(Me2CH)2pz)3)]2(O2). J. Am. Chem. Soc. 111, 8975–8976
(1989).
2. Solomon, E. I., Sundaram, U. M., Machonkin, T. E. Multicopper Oxidases and
Oxygenases. Chem. Rev. 96, 2563–2606 (1996).
3. Karlin, K. D., Kaderli, S., Zuberbühler, A. D. Kinetics and Thermodynamics of
Copper(I)/Dioxygen Interaction. Acc. Chem. Res. 30, 139–147 (1997).
4. Halfen, J. A. et al. Reversible cleavage and formation of the dioxygen O-O bond within
a dicopper complex. Science 271, 1397–1400 (1996).
5. Koo, C. W., Tucci, F. J., He, Y., Rosenzweig, A. C. Recovery of particulate methane
monooxygenase structure and activity in a lipid bilayer. Science 375, 1287–1291
(2022).
6. Mirica, L. M., Ottenwaelder, X., Stack, T. D. P. Structure and spectroscopy of copper-
dioxygen complexes. Chem. Rev. 104, 1013–1045 (2004).
Contents
Preface v
I. Introduction 2
II. Copper, a Biocidal Compound 3
III. Copper Homeostasis Systems in Bacteria 5
IV. Copper Toxicity: A Phenomenon Dependent on
the Bioavailability of Copper and on Bacteria 7
V. Specificity of Copper Chemistry 8
A. Cu Coordination Chemistry 8
B. Cu Reactivity: The Case of ROS Production 9
C. In Vivo Complexation of Cu 10
VI. Mechanisms of Cu Toxicity: Macromolecules Targeted by
Copper10
A. Membranes 11
B. DNA 12
C. Proteins 13
VII. Chemistry of Copper Complexes 14
A. Reactivity of Cu Complexes 15
B. Stability of the Cu-L Complex 16
ix
x Contents
2 T
ransition State Analogue Molecules as Mechanistic
Tools and Inhibitors for Tyrosinase 45
Clarisse Faure, Amaury du Moulinet d’Hardemare,
Hélène Jamet, Catherine Belle, Elisabetta Bergantino,
Luigi Bubacco, Maurizio Benfatto, A. Jalila Simaan,
and Marius Réglier
List of Abbreviations 46
I. Introduction 47
II. Biological Functions of Melanins 47
A. Melanin as Color Pigment 48
B. Melanin as a Defensive Barrier 50
Contents xi
III. Tyrosinase 51
A. Structures of the Tyrosinase Active Sites 51
B. Catalytic Mechanism of Tyrosinase 53
C. Tyrosinase-Related Proteins TRP1 and TRP2 55
IV. Dysfunction in Tyrosinase Activity 58
A. Pathologies Linked to TYR Dysfunction 58
B. Skin Whitening 59
V. Tyrosinase Inhibition 59
A. Variation in TYR Sources 60
B. Transition-State Analogue (TSA) Inhibitors 60
1. L-Mimosine 63
2. Kojic acid and derivatives 63
3. Tropolone and derivatives 67
4. HOPNO inhibitor 68
5. HOPNO derivatives 73
VI. Conclusions 74
Acknowledgment76
VII. References 76
I. Introduction 81
A. General Considerations 81
B. Scope of the Review 83
II. Dicopper Proteins — Brief Overview 84
A. Hemocyanins 85
B. Tyrosinases 85
C. Catechol Oxidase 86
III. Three Cu2O2 Core Structures 87
IV. Biomimetic Studies on Tyrosinase 88
A. Intramolecular m-Xylyl and Aromatic Ring
Hydroxylation88
xii Contents
4 M
onooxygenation of Phenols by Small-molecule Models
of Tyrosinase: Correlations Between Structure and
Catalytic Activity 123
Alexander Koch, Tobias A. Engesser, Ramona Jurgeleit,
and Felix Tuczek
5 E
lectrochemistry and Spectroelectrochemistry of
Copper-Oxygen-Relevant Species 153
Nicolas Le Poul
235
235
241
242
246
247
251
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0001
Abstract
Copper is an essential micronutrient for most living beings including
bacteria. This is mainly due to essential roles of catalytic Cu-centers in
enzymes. However, an excess of Cu is toxic and thereby used over cen-
turies as a powerful antimicrobial agent. Bacteria developed several
systems to detect Cu and protect themselves from high intracellular
Cu concentration, namely via Cu-extrusion and/or Cu-storage. Several
Cu-detoxification systems have been described and new ones are still
1
2 Zuily et al.
I. Introduction
Before the Great Oxygenation Event, copper (Cu) was found as copper
sulfide minerals such as chalcocite (Cu2S) and chalcopyrite (CuFeS2),
impairing its bioavailability. In this anoxic environment, earliest anaerobic
prokaryotes were not using copper as a metal center for their enzymes to
operate.1 Apparition of oxygen on earth led to the oxidation of copper,
increasing its solubility, hence turning Cu into a bioavailable metal.2,3
Living organisms had to adapt to this drastic modification of their environ-
ment and they subsequently modify their metallome. For instance, in cur-
rent aerobic organisms, key enzymatic reactions like the reduction of
dioxygen into water are performed with copper-containing proteins.4
However, such benefits came with its part of inconveniency. Indeed,
while being a key element, Cu-chemical properties render this metal
highly toxic for living organisms. Living cells had to face this paradox of
an essential yet toxic element. General strategies have been developed by
bacteria to remove any excess of copper. Nonetheless, saturation of these
Ligands as a Tool to Tune the Toxicity of Cu on Bacteria 3
Potable water
Wound
dressings Filters
Agriculture Clothes
Paints Intra-body
compounds
(contraception,
Hospitals
dentistry)
surfaces
Cu(I)
CTR1
ATOX1
NAD+ NADH
ATP7A
Macrophage
Cu(II) Cu(I)
O2 O2
Phagolysosome
H2O2 OH
CueO
Outer membrane
CusC
CueO Cu(I) CusF
CusB
CusB
Cu(II)
Periplasm
CusA
CopA CusS
CopA
Inner membrane
P Cytoplasm
ATP ADP+Pi
CusR
CopZ
CueR
P
CusR
CueR CueR
A. Cu Coordination Chemistry
Cu(I) and Cu(II) are cations soluble in polar solvents. In water, Cu(II) is
the main form encountered, and is coordinated with six water molecules.
At higher pH, copper hydroxide Cu(OH)2 is formed, with a very low solu-
bility in water. Cu(I) is not stable in pure water: either it is readily oxi-
dized to Cu(II) by dioxygen in aerobic conditions, or it disproportionates
into Cu(II) and Cu(0) under anaerobic conditions. Hence, to exist as Cu(I)
in aqueous solvent, it needs to be stabilized by appropriate ligands.
Interestingly, Cu(I) and Cu(II) prefer different coordination geometries, as
referred in Table 1, which means that each of these two forms can be sta-
bilized by different sets of ligands.37,38 Worthy of note, Cu(I) is a soft
cation (the softest of the essential metal ions) and hence is very thiophilic.
In addition, due to the difference in coordination chemistry of Cu(II) and
Cu(I), the redox potential can be tuned over a wide range.39,40
Ligands as a Tool to Tune the Toxicity of Cu on Bacteria 9
Table 1. Key coordination properties of Cu(I) and Cu(II) to understand their behav-
ior in biological environment.
Cu(I) Cu(II)
Preferred geometry Digonal, trigonal, tetrahedral, Square planar, with one or
or penta-coordinated two weaker axial ligands
Preferred ligands Thiolates > thioether, amines > Amines, sulfur > oxygen
oxygen
Stability in aerobic Easily oxidized Stable
condition
Stability in water Not stable Stable
(pH 7) disproportionation
Solubility in water KsCuOH = 1.2 10–6(a) (KsCu(OH)2 = 2.2 × 10–20)b
Interaction with Can bind strongly to thiols Is reduced to Cu(I) by
thiols thiols
a
K s CuOH =
[Cu+ ] × [OH − ]
[CuOH]
[Cu2+ ] × [OH − ]
2
b
Ks Cu ( OH )2
=
[Cu ( OH )2 ]
Cu ( I ) + H 2O2 → Cu ( II ) + HO • + HO −
Cu ( II ) + red → Cu ( I ) + ox
Cu ( I ) + O2 → Cu ( II ) + O2• −
Cu ( I ) + O2• − + 2 H + → H 2O2 + Cu( II )
C. In Vivo Complexation of Cu
In cell, both Cu(I) and Cu(II) are present, in tight interaction with BMs
(mainly proteins), which can either stabilize Cu(I) or Cu(II) or cycle
Cu(I)/Cu(II) in redox active enzymes. Cu(II) and Cu(I) are found in the
periplasmic environment of Gram-negative bacteria, whereas Cu(I)
predominates in the cytosol. The presence of thiols (as GSH) in mM
concentration explains why copper can be maintained in its reduced
form.42 Under growth condition with low level of Cu in the media,
intracellular free Cu(I) is estimated to be extremely low. Around 10–21
M Cu(I) is found in the bacteria cytoplasm, compared to 10–15 M
Zn.43,44 Actually, keeping intracellular free copper concentration as low
as possible constitutes one of the main strategies to prevent intracellular
Cu toxicity.
Cu(II)
Cu(II) Cu(I)
Fenton reaction
DNA
+H2O2
S
Cu(I) +OH. Proteins
Fe Fe Fenton reaction
S Lipides
Metalloproteins
inactivation GSH Cu(I)
S
Cu Cu GSH-Cu+ complex
S
increase Misfolding
M Aggregation
Figure 4. Intracellular Cu impacts. Copper enters the bacterial cell via an unknown
pathway. In aerobic conditions, copper can be found in the periplasmic compartment
as Cu(I) or Cu(II). However, in the reducing environment of the cytoplasm, Cu(II) will
be rapidly reduced into Cu(I). Cu(I) in both compartments can be involved in Fenton-
type reactions to produce hydroxyl radicals that are highly reactive and toxic to mac-
romolecules. Copper can also have a direct impact on lipids, proteins, and nucleic
acid by altering their structure. Free copper or probably the accumulation of glu-
tathione-copper complexes will interfere with metalloproteins by the displacement of
the physiological metal and the perturbation of the metallocenter assembly.
A. Membranes
Cellular membrane is, in fact, the first component Cu encounters, and
lipids have been described as one of the Cu targets. Nevertheless, the
action of Cu on membranes appears to be related to indirect effect via
ROS production or via cysteine-binding to proteins, hence, leading to the
modification of proteins and/or lipids that will end in the destabilization
of the membrane integrity.
Membrane phospholipids such as phosphatidyl ethanolamine (PE) or
phosphatidyl serine (PS) have been shown to coordinate Cu(II) ions via
reactive groups such as the primary amine of PE found at its head group.46,47
Hence, the production of ROS is locally increased explaining the high level
of lipid oxidation in presence of Cu(II).47 Other reports confirm that Cu ions
bound to the membrane will induce the production of hydroxyl radicals,
which will react with lipids and induce lipid peroxidation of the double
12 Zuily et al.
B. DNA
Cu toxicity has long been attributed to DNA damages. Indeed, several
experimental evidences demonstrate in vitro a clear denaturation of DNA
fragments in presence of increasing amount of Cu(II).51 By binding to
phosphate and nitrogen ligands, Cu prevents the correct folding of
DNA.52,53 Once bound, Cu further reacts and induces DNA break via the
production of ROS.54 However, all these earlier studies were performed in
a non-cellular environment. In an environment containing millimolar
GSH concentrations, the affinity of DNA to the relevant redox state Cu(I)
is low compared to Cu(I) binding to GSH and to other thiolate or sulfide
containing proteins (such as those containing FeS clusters). The stronger
complex DNA-Cu(II) would never form in vivo as Cu(II) will be reduced
immediately by GSH. In addition, in E. coli, most of Cu ions are found in
the periplasm. Very little amount may enter in the cytoplasm where the
formation of hydroxyl radicals is unlikely given the high concentration of
GSH in this compartment and its potency to stabilize Cu(I). In accord-
ance, Macomber et al. demonstrated the absence of severe DNA damages
in E. coli after Cu treatment.35 In fact, the Cu binding capacity of DNA
might be used by bacteria to scavenge extracellular Cu ions and prevent
Cu intracellular damages. Dalecki et al.55 proposed that such strategy
could be used by bacteria pathogens to protect themselves against copper
by the release of extracellular DNA once inside eukaryotic cells56 or once
they formed biofilm matrix notably composed of extracellular DNA.57
Further studies might be of interest to better define the role played by
DNA among bacterial strategies to survive copper excess.
Ligands as a Tool to Tune the Toxicity of Cu on Bacteria 13
C. Proteins
Proteins are one of the most sensitive targets to Cu overload in bacteria
cells. Cu ions bind to O-, N-, and S- donors, elements that are found in the
protein backbone as well as in several amino acid side chains (such as
cysteine, methionine, histidine, etc.). Such variability of interactions leads
to a multitude of impact on proteins and, in fine, on cells.
In aerobic conditions, intracellular Cu excess leads to severe oxidative
damage linked to hydroxyl radical production via the Fenton reaction, as
well as to the oxidation of cysteine residues that may form unwanted
disulfide bonds. Deletion of periplasmic repair system known to fix abnor-
mal disulfide bonds, decreases drastically E. coli survival upon Cu stress in
aerobic conditions.58 This was not observed under anaerobic conditions,
underlying the different mechanism of action of Cu under anaerobic versus
aerobic conditions.59 Cu might also inhibit the reduction of disulfide bonds
in the periplasm, a process required to maturate cytochrome c and this has
been shown to perturb key pathways impairing cell survival.28 Recently,
methionine oxidation of the periplasmic protein CusF has been reported.60
These results confirm the production of ROS in the periplasmic environ-
ment of E. coli. Under these conditions, periplasmic methionine sulfoxide
reductase repairs the oxidized methionines of CusF, demonstrating the
importance for cells to keep their copper homeostasis systems active by
repairing the oxidation of their residues.
In addition, Cu(II) induces protein misfolding in vitro and this was
recently demonstrated in vivo.59,61,62 As described earlier, Cu(II) metal ion
possesses a strong affinity for proteins compared to other metals, reflecting
its place in the Irving–Williams series. Furthermore, Cu(I) shows a high
affinity for thiols, such as cysteines. Hence, any intracellular concentration
increase of Cu could lead to a strong perturbation of Zn-, Fe-, or other
metalloproteins, as intracellular concentration of metals have been shown
to dictate some proteins metalation.63,64 Under anaerobic conditions, direct
inactivation of metalloproteins by Cu has been demonstrated in vivo by
Imlay et al.45 The authors clearly showed that the iron–sulfur enzyme (iso-
propylmalate dehydratase) is one of the first Cu-target in E. coli. Cu-induced
protein inactivation leads to cell death unless amino acids were added to the
media. Inactivation of iron–sulfur enzymes or other metalloproteins, as
well as perturbation in iron–sulfur biosynthesis or heme biosynthesis, has
been further confirmed by other authors in other microorganisms.65–68
14 Zuily et al.
A. Reactivity of Cu Complexes
Since Cu-cytotoxicity can arise from different chemical mechanisms (see
Section 1.6), ligands (L) are a good tool to tune it and/or control its speci-
ficity. Indeed, ligands will promote or prevent some of the mechanisms
listed below.
(a) (b)
(c) (d)
Besides, one must care of the stoichiometry of the complexes before com-
paring their Kd. When appreciating apparent dissociation constant of
complexes, important concepts have to be considered, such as chelate
effect, Pearson’s principle (hard and soft acids and bases [HSAB]), coor-
dination geometry, and charges (including partial and electronic densi-
ties). Of course, another important concern is the physiological context in
which the metal-complex will be used. All of this needs to be taken into
account in in vitro measurements to get a better idea of Cu-L complex
intracellular stability and consequently intracellular behavior.
D. Redox Potential
Cu ions in a Cu-L complex can have very different redox potentials. The
main biological redox couple is Cu(I)/Cu(II). Depending on the properties
of the ligand (geometry, number, and type of coordination atoms (O, N, S,
etc.) and charge or electron density of the ligand (see Table 3), the redox
potential can vary from below –0.4 up to +0.8 V versus NHE. This means
that a range of potentials can be reached in which Cu can switch from
quasi-inert behavior in biological environment to a high redox activity.
development of new ligands toward the tuning and control of the killing
efficiency of copper-based drugs: (i) improving membrane permeability,
(ii) improving the specificity of the reaction, (iii) tuning the compartment
targeted, or (iv) tuning the mechanism of action. In this chapter, we will
discuss the different intracellular routes Cu-complexes could undertake.
A. Localization
Cu-L cellular localization is largely influenced by its charge: (i) hydro-
phobic Cu-L might localize into the membranes, (ii) less hydrophobic (yet
not too hydrophilic) ones could cross the membranes, and (iii) highly
charged ones are expected not to passively penetrate membranes.
Noteworthy, a hydrophobic complex (neutral global charge, hydrophobic
ligand) might have low solubility in aqueous media. Therefore, experi-
menters must care about precipitation.
Cu (I/II) - Complexes
Ligand alone
Substrate Product
Substrate Product
Substrate Product
Cu
toxicity
Reductive
dissociation Metal interference
Figure 6. Possible routes and activities of Cu bound to ligands. The left part of the
panel describes the activity of a Cu(I/II)-L encountering the bacteria with three con-
ceptual routes. On the right part of the panel, the Cu-free ligand (L) encounters the
bacteria, but once inside the cells, it will interact with endogenous Cu enzymes or
interfere with intracellular metal homeostasis. For Cu-L, route 1 describes the iono-
phore activity (also called “mechanism I” in this review). The role of the ligand is to
help the Cu to pass the membranes hence increasing the intracellular Cu concentra-
tion. Very often the Cu dissociates from the ligand triggered by Cu(II) reduction. The
released ligand could bind other metals or interfere with other targets as well the Cu.
Route 2 consists of a new function of the Cu-L. Classically, this would be a Cu-L
catalysing a reaction, most likely the production of ROS (also called “mechanism II”
in this review). In route 3, the Cu-L inhibits an essential function (loss of function)
such as binding to an essential cysteine in a cytoplasmic or periplasmic enzyme.
transcription factor CueR (Kd < zM) and the consequent expression of the
Cu extrusion machinery. Such complex might not have strong intracellular
toxic effect. Otherwise, Cu(I)-L can dissociate in the cytosol, exerting an
ionophore activity.
would bind endogenous Cu in the bacteria (Figure 6). This may have bio-
logical activity via two possible mechanisms:
Thermodynamic stability
Formation of ternary
complexes (e.g., with
proteins)
Reduction by physiological
reducing agent
In green LE correspond to external ligand. External ligands are classically water (or other sol-
vents) or anions during the preparation of the complexes (e.g., halides). Upon addition in cell
culture medium or application to living organisms, they are supposed to be exchanged easily
with small molecules (water, amino acid, and anions) or with proteins and hence form the
ternary complex (L-Cu-BM).
a - .74
24 Zuily et al.
presented above have not yet been fully exploited and further efforts are
required to improve our knowledge on the impact of diverse Cu-L on
bacteria.
A. Dithiocarbamates Compounds
Dithiocarbamates (DTC) are a class of metal chelators that can be formed
from thiuram disulfides by a two-electron reduction (Figure 7). They bind
Cu(II) bidentately and can form Cu(II)-DTC and Cu(II)-DTC2 complexes.
Cu(II)-DTC2 is neutral and the redox potential is quite low (Table 3).
Thiuram disulfides have a much lower affinity for copper ions than DTCs.
Thus, in terms of copper binding, thiuram disulfides are prodrugs, and
need to be activated by reduction, which is supposed to occur mainly
intracellularly by thiols. The most prominent example is the antabuse drug
disulfiram (a neutral thiuram disulfide) that upon reduction yields the
active DTC drug, diethyldithiocarbamate (DEDTC).
Thiuram disulfides, or most likely the intracellular compound DTC,
have been shown to have antibacterial activity on Staphylococcus aureus,75
on Mycobacterium tuberculosis,76 and on P. aeruginosa.77 A link
between the antimicrobial activity and DTC ability to bind copper was
proposed but not clearly demonstrated. As a thiol agent, DTC might react
with cysteine from key enzymes to form a thioester.77 DTC might also be
able to bind Cu(II) in the oxidizing periplasmic compartment of Gram-
negative bacteria and per se perturbs Cu trafficking toward Cu-enzymes.
Last, DTC might transport Cu(II) into the cytoplasm where it is expected
to be released as Cu(I) due to the high thiol content.72 Coherently, in this
reducing environment of bacteria cytoplasm, DTC might not form a com-
plex with Cu(I) as it will not compete with GSH found in mM
concentration.72
Instead of adding only the ligand DTC, direct addition of Cu(II)-
DTC2 complex in the media has a strong antimicrobial effect. This activity
can be explained by intracellular accumulation of Cu due to ionophoric
activity (mechanism I). Accordingly, strong Cu-dependent antimicrobial
activity of disulfiram/DTC was reported against M. tuberculosis, with a
minimum inhibitory concentration (MIC) around 0.3 mM,78 and against
S. pneumoniae.79 The authors further showed the ability of Cu(II)-DTC2
complex to cross bacteria membranes and inactivate cytoplasmic
enzymes.78 In such condition, intracellular Cu stress response is induced.
This confirms that Cu(II)-DTC2 entering the cytoplasm might release
high Cu(I) concentration which perturbs metabolic pathways. Such
Cu-dependent growth inhibition of disulfiram/DTC has also been observed
recently on mollicutes.80 The addition of a Cu(II) chelator in the media
completely modifies the MIC values of disulfiram from nanomolar to
micromolar range. This result further confirms the critical role of Cu for
disulfiram antimicrobial activity.
Besides, Cu(II)-DTC2 has been shown to interact with zinc finger
proteins in cancer cells, and to oxidize Zn(II)-thiolates to form disulfide.81,82
Even though this mechanism has not been reported in bacterial cells, its
occurrence cannot be ruled out.
C. Phenanthrolines
Phenanthrolines (Phen) are bidentate ligands forming both 1:1 and 1:2
complexes with Cu(I) and Cu(II) (see Figure 9). 2,9-unsubstituted
D. Pyrithione
Pyrithione (PT) is a neutral ligand which can bind Cu(II) as a Cu(II)-PT2
complex, which is neutral as well, due to the loss of the two hydroxy-
lamine protons (Figure 10). Thus, in terms of charge, PT has properties
suited for ionophore activity (mechanism I), as Cu(II)-PT2 can enter the
cytoplasm and PT exit it. The affinity of Cu(I) to PT is unknown, but
according to Pearson theory, Cu(I) has a low affinity for N-OH. This
means that once reduced in the cytoplasm, Cu(I) will dissociate from PT.
This was further confirmed by a recent study,62 where the authors dem-
onstrated the intracellular increase of copper in E. coli after treatment
with a mix of Cu and PT and in comparison, with Cu alone. Cu-induced
32
Affinity (pH 7.4)
in log Kd for
E° Cu(II)/Cu(I) Intracellular reactivity
Compound class Representative Cu(II) Cu(I) (V/NHE) (cytoplasm) Denticity
Cu alone 0.16 – Fast reduction of Cu(II) —
– Cu(I) binds to GSH
Dithiocarbamate DETC – Fast reduction 2
(DTC) Cu + DETC 11.7c – Complete dissociation
Cu-DETC + DETC 12.2c –0.14k – Cu(I) binds to GSH
8-hydroxyquinoline Clioquinol (CQ) – Fast reduction 2
(HQ) Cu + CQ 10.8f – Dissociation
Zuily et al.
Cu + HQ 9.75g –0.36h – Cu(I) binds to GSH
Cu-CQ + CQ 9.2f
Cu-HQ + HQ 8.65g
Phenanthrolines Cu + Phen 9.1l 10.3l – Fast reduction 2
Cu-Phen + Phen 6.8l 5.5l – Partial dissociation
Cu + 2 Phen 15.9l 15.8l 0.17i – Cu(I) binds to GSH
BCS
Cu + 2BCS 12.5d 20.8d 0.62j
Pyrithione Cu + PT > 8.5e – Fast reduction 2
Cu(II)-PT + PT 5.9e – Dissociation
– Cu(I) binds to GSH
Bis- Cu + Atsm ~13a –0.40a – Stable 4
thiosemicarbazone Cu + Gtsm ~13a –0.24a – Slow
Cu + Kts ~19b –0.12b (pH 6.6) – Dissociation
a 95 b. 94 c
; ; in methanol;97; d 98 e 99 f
; ; in 80% Methanol and 20% H2O, 100 g 101 h 102 i 103 j 104 k 105 l 106
; ; ; ; ; ;
Ligands as a Tool to Tune the Toxicity of Cu on Bacteria 33
copper dependent manner. Still, the use of Cu(II) chelator in the media
decreased gtsm antibacterial activity, in line with the fact that small
amount of Cu in the media is sufficient to boost the growth inhibitory
effect of gtsm.
The higher toxicity of Cu-gtsm compare to Cu-atsm applies to other
bacteria (Gram-positive and -negative).91,96 However, the susceptibility of
the bacteria is very different, and is directly related to the bacterial physi-
ology. Bacteria with a high developed Cu-detoxification system like
E. coli are quite resistant, other with a poorer Cu extrusion system, like
N. gonorrhoeae are much more susceptible (IC 50 < µM). This can be
explained by the fact that N. gonorrhoeae have only one exporter, CopA,
whose expression is not inducible by Cu. Moreover, in N. gonorrhoeae an
additional toxicity mechanism was described related to the inhibition of
dehydrogenases in the respiratory chain.96 The results suggested that
Cu-bTSC complexes bind to membrane proteins, likely in hydrophobic
pockets where they interfere with ubiquinol. This mechanism is independ-
ent of Cu release or redox activity, in line with experiments showing that
Zn-bTSC showed similar activity.96 However, other mechanisms may
explain the toxicity of Cu-gtsm, as bacteria growing under fermentation
(S. pneumoniae and E. coli) are also highly sensitive to Cu-gtsm.91
X. Peptides-Based Cu-Chelators
Antimicrobial peptides (AMPs) are described as peptides of about 10 to
100 amino acids that can be either naturally produced by the immune
system to combat invading pathogens or chemically synthesized with non-
natural sequences. For certain natural AMPs, the binding of copper or
other metal ions was proposed to be required for their activity. The main
idea was that Cu-binding to AMPs may improve the antimicrobial activity
either through Cu-catalyzed ROS production or Cu-induced conforma-
tional changes enhancing the interaction of the AMP with its target. Two
main approaches are reported: (i) identification of native AMPs contain-
ing a high affinity-Cu-binding site, (ii) engineering such a site in a native
or artificial peptide (recent review on AMPs107).
Short peptides are often dynamic molecules lacking a well-defined
3D structure. Hence, Cu-binding to peptides is generally weaker than to
34 Zuily et al.
versus mammalian cells. Some of these compounds are even used to treat
cancer and are known to negatively impact eukaryotic cells.118,119 To
increase selectivity, several strategies can be envisioned.
Festa et al.83 used a compound that gets activated once inside mac-
rophages. They took advantage of the high level of ROS as well as of
copper in macrophages to transform a non-toxic compound into a highly
toxic one. In particular, oxidation of a pro-drug released the 8-HQ ligand
that complexed Cu and then acted as ionophore to aid the immune
system to kill the pathogens. They were able to show that such condi-
tional activation worked in vitro and in mouse model on the pathogen
C. neoformans.83
Another strategy followed by Wolschendorf and coworkers was to
screen different compounds in combination with copper against
S. aureus.120 They were able to highlight the potential of new copper-
dependent molecules such as thiourea inhibitors which were able to
inhibit efficiently S. aureus growth and seemed tolerated in cell culture.
They also demonstrated the toxicity of PZP-915 (5-Benzyl-3-(4-
chlorophenyl) -2-methyl-4H,7H-pyrazolo [1,5-a] pyrimidin-7-one) toward
S. aureus resistant strains with a certain selectivity toward bacteria over
eukaryotic cells.121
The use of peptide with Cu-binding motifs is another way to obtain
higher selectivity toward bacteria as it is exemplified by the natural AMPs.
Peptides could also be used for targeting a Cu-site to compartments or
proteins. Accordingly, cytoplasm-targeted peptides could act as iono-
phores by adding a Cu(II)-site to the peptide. This strategy would need
optimization of the peptide sequence toward cell-penetration, so far, an
area of research poorly explored. However, it is not clear whether peptides
can be as efficient as small organic ligands. For the antimicrobial strategy
consisting in employing a Cu-site on an AMP to catalyze ROS production,
reports are mainly limited to Xxx-Zzz-His motifs. This motif is quite
redox inert as it stabilizes Cu(II) and hence ROS production is low. Other
high affinity Cu-binding sites are needed to increase ROS production.
This seems to be very challenging for short peptides, and protein-like
folding and/or organic ligands are needed to reach the balance between
stability and redox reactivity. Moreover, peptides form a scaffold which
38 Zuily et al.
XII. References
1. Decaria, L.; Bertini, I.; Williams, R. J. P. Metallomics. 2011, 3(1), 56–60. doi:
10.1039/C0MT00045K.
2. Dupont, C. L.; Grass, G.; etRensing, C. Metallomics. 2011, 3(11), 1109. doi:
10.1039/c1mt00107h.
3. Hong Enriquez R. P.; Do, T. N. Life. 2012, 2(4), 274–285. doi: 10.3390/life2040274.
4. Bertini, I.; Cavallaro, G.; McGreevy, K. S. Coord. Chem. Rev. 2010, 254(5–6),
506–524. doi: 10.1016/j.ccr.2009.07.024.
Ligands as a Tool to Tune the Toxicity of Cu on Bacteria 39
5. Borkow, G.; Gabbay, J. CMC 2005, 12(18), 2163–2175. doi: 10.2174/
0929867054637617.
6. Borkow, G.; Gabbay, J. Curr. Chem. Biol. 2009, 3(3), 272–278. doi:
10.2174/187231309789054887.
7. Grass, G.; Rensing, C.; Solioz, M. Appl. Environ. Mictobiol. 2011, 77(5), 5. doi:
10.1128/AEM.02766-10.
8. Arendsen, L. P.; Thakar, R.; Sultan, A. H. Clin. Microbiol. Rev. 2019, 32(4),
e00125–18. /cmr/32/4/CMR.00125-18.atom. doi: 10.1128/CMR.00125-18.
9. Hodgkinson, V.; Petris, M. J. J. Biol. Chem. 2012, 287(17), 13549–13555. doi:
10.1074/jbc.R111.316406.
10. Hood, M. I.; Skaar, E. P. Nat. Rev. Microbiol. 2012, 10(8), 525–537. doi: 10.1038/
nrmicro2836.
11. Wagner, D. et al. J. Immunol. 2005, 174(3), 1491–1500. doi: 10.4049/
jimmunol.174.3.1491.
12. Djoko, K. Y.; Ong, C. Y.; Walker, M. J.; McEwan, A. G. J. Biol. Chem. 2015,
290(31), 31. doi: 10.1074/jbc.R115.647099.
13. White, C.; Lee, J.; Kambe, T.; Fritsche, K.; Petris, M. J. J. Biol. Chem. 2009,
284(49), 49. doi: 10.1074/jbc.M109.070201.
14. Giachino, A.; Waldron, K. J. Mol. Microbiol. 2020, Sep; 114(3):377–390. doi:
10.1111/mmi.14522.
15. G. Ghssein et al., Science, 2016, 352(6289) 1105–1109.
16. H. J. Kim et al., Science, 2004, 305(5690), 1612–1615, doi: 10.1126/science.
1098322.
17. Koh, E.-I.; Robinson, A. E.; Bandara, N.; Rogers, B. E.; Henderson, J. P. Nat. Chem.
Biol. 2017, 13(9), 1016–1021. doi: 10.1038/nchembio.2441.
18. Andrei, A. et al. Membranes. 2020, 10(9), 242. doi: 10.3390/membranes10090242.
19. Rosenzweig, A. C. Chem. Biol. 2002, 9(6), 673–677. doi: 10.1016/S1074-
5521(02)00156-4.
20. Mealman, T. D.; Blackburn, N. J.; McEvoy, M. M. in Current Topics in Membranes.
2012, 69, Elsevier, pp. 163–196. doi: 10.1016/B978-0-12-394390-3.00007-0.
21. Munson, G. P.; Lam, D. L.; Outten, F. W.; O’Halloran, T. V. J. Bacteriol. 2000,
182(20), 5864–5871. doi: 10.1128/JB.182.20.5864-5871.2000.
22. Outten, F. W.; Huffman, D. L.; Hale, J. A.; O’Halloran, T. V. J. Biol. Chem. 2001,
276(33),33. doi: 10.1074/jbc.M104122200.
23. Singh, S. K.; Grass, G.; Rensing, C.; Montfort, W. R. J. Bacteriol. 2004, 186(22),
22. doi: 10.1128/JB.186.22.7815-7817.2004.
24. Vita, N. et al. Nature. 2015, 525(7567), 140–143. doi: 10.1038/nature14854.
25. Vita, N. et al. Sci. Rep. 2016, 6(1), 39065. doi: 10.1038/srep39065.
26. Stewart, L. J. et al. mBio. 2020, 11(6), e02804–20. /mbio/11/6/mBio.02804-20.
atom. doi: 10.1128/mBio.02804-20.
27. Durand, A. et al. Metallomics. 2021, 13(12), mfab067. doi: 10.1093/mtomcs/
mfab067.
40 Zuily et al.
51. Eichhorn, G. L.; Shin, Y. A. J. Am. Chem. Soc. 1968, 90(26), 7323–7328. doi:
10.1021/ja01028a024.
52. Geierstanger, B. H.; Kagawa, T. F.; Chen, S. L.; Quigley, G. J.; Ho, P. S. J. Biol.
Chem. 1991, 266(30), 20185–20191. doi: 10.2210/pdb1d40/pdb.
53. Rifkind, J. M.; Shin, Y. A.; Heim, J. M.; Eichhorn, G. L. Biopolymers 1976, 15(10),
1879–1902. doi: 10.1002/bip.1976.360151002.
54. Sagripanti, J. L.; Kraemer, K. H. J. Biol. Chem. 1989, 264(3), 1729–1734.
55. Dalecki, A. G.; Crawford, C. L.; Wolschendorf, F. in Advances in Microbial
Physiology. 2017, 70, Elsevier, pp. 193–260. doi: 10.1016/bs.ampbs.2017.01.007.
56. Manzanillo, P. S.; Shiloh, M. U.; Portnoy, D. A.; Cox, J. S. Cell Host & Microbe
2012, 11(5), 469–480. doi: 10.1016/j.chom.2012.03.007.
57. Whitchurch, C. B.; Science 2002, 295(5559), 1487–1487. doi: 10.1126/
science.295.5559.1487.
58. Hiniker, A.; Collet, J.-F.; Bardwell, J. C. A. J. Biol. Chem. 2005, 280(40), 33785–
33791. doi: 10.1074/jbc.M505742200.
59. Zuily, L. et al. mBio 2022, 13(2), e03251–21. doi: 10.1128/mbio.03251-21.
60. Vergnes A. et al., PLoS Genet 2022, 18(7), e1010180, doi: 10.1371/journal.
pgen.1010180.
61. Saporito-Magriñá, C. M. et al. Metallomics 2018, 10(12), 1743–1754. doi: 10.1039/
C8MT00182K.
62. Wiebelhaus, N.; Zaengle-Barone, J. M.; Hwang, K. K.; Franz, K. J.; Fitzgerald, M.
C. ACS Chem. Biol. 2021, 16(1), 214–224. doi: 10.1021/acschembio.0c00900.
63. Foster, A. W.; Osman, D.; Robinson, N. J. J. Biol. Chem. 2014, 289(41), 28095–
28103. doi: 10.1074/jbc.R114.588145.
64. Tottey, S. et al. Nature 2008, 455(7216), 1138–1142. doi: 10.1038/nature07340.
65. Chillappagari, S.; Seubert, A.; Trip, H.; Kuipers, O. P.; Marahiel, M. A.; Miethke,
M. J. Bacteriol. 2010, 192(10), 2512–2524. doi: 10.1128/JB.00058-10.
66. Djoko, K. Y.; McEwan, A. G. ACS Chem. Biol. 2013, 8(10), 2217–2223. doi:
10.1021/cb4002443.
67. Johnson, M. D. L.; Kehl-Fie, T. E.; Rosch, J. W. Metallomics 2015, 7(5), 786–794.
doi: 10.1039/C5MT00011D.
68. Tan, G. et al. Appl. Environ. Microbiol. 2017, 83(16), 16. doi: 10.1128/AEM.
00867-17.
69. Gugala, N.; Salazar-Alemán, D. A.; Chua, G.; Turner, R. J. Metallomics 2022, 14(1),
mfab071. doi: 10.1093/mtomcs/mfab071.
70. O'Hern C.; Djoko, K. Y. MBio, 2022, 13(3), e00434-22, doi:10.1128/mbio.00434-22
71. Bossak-Ahmad, K.; Frączyk, T.; Bal, W.; Drew, S. C. ChemBioChem 2020, 21(3),
331–334. doi: 10.1002/cbic.201900435.
72. Santoro, A.; Calvo, J. S.; Peris-Díaz, M. D.; Krężel, A.; Meloni, G.; Faller, P. Angew.
Chem. Int. Ed. 2020, 59(20), 7830–7835. doi: 10.1002/anie.201916316.
73. Gonzalez, P. et al. Chem. Eur. J. 2018, 24(32), 8029–8041. doi: 10.1002/
chem.201705398.
42 Zuily et al.
74. Rajalakshmi, S.; Fathima, A.; Rao, J. R.; Nair, B. U. RSC Adv. 2014, 4(60), 32004–
32012. doi: 10.1039/C4RA03241A.
75. Phillips, M.; Malloy, G.; Nedunchezian, D.; Lukrec, A.; Howard, R. G. Antimicrob.
Agents Chemother. 1991, 35(4), 785–787. doi: 10.1128/aac.35.4.785.
76. Byrne, S. T. et al. Antimicrob. Agents Chemother. 2007, 51(12), 4495–4497. doi:
10.1128/AAC.00753-07.
77. Zaldívar-Machorro, V. J.; López-Ortiz, M.; Demare, P.; Regla, I.; Muñoz-Clares, R.
A. Biochimie 2011, 93(2), 286–295. doi: 10.1016/j.biochi.2010.09.022.
78. Dalecki, A. G. et al. Antimicrob Agents Chemother. 2015, 59(8), 8. doi: 10.1128/
AAC.00692-15.
79. Menghani, S. V. et al. Microbiol Spectr. 2021. doi: 10.1128/Spectrum.00778-21.
80. Totten, A. H.; Crawford, C. L.; Dalecki, A. G.; Xiao, L.; Wolschendorf, F.; Atkinson,
T. P. Front. Microbiol. 2019, 10, 1720. doi: 10.3389/fmicb.2019.01720.
81. Skrott, Z. et al. Nature 2017, 552(7684), 194–199. doi: 10.1038/nature25016.
82. Xu, L. et al. Angew. Chem. Int. Ed. 2019, 58(18), 6070–6073. doi: 10.1002/
anie.201814519.
83. Festa, R. A.; Helsel, M. E.; Franz, K. J.; Thiele, D. J. Chem. Biol. 2014, 21(8),8. doi:
10.1016/j.chembiol.2014.06.009.
84. Tardito, S. et al. J. Am. Chem. Soc. 2011, 133(16), 6235–6242. doi: 10.1021/
ja109413c.
85. Shah, S. et al. Antimicrob. Agents Chemother. 2016, 60(10), 5765–5776. doi:
10.1128/AAC.00325-16.
86. Cahoon, L. Nat Med. avr. 2009, 15(4), 356–359. doi: 10.1038/nm0409-356.
87. Ananthan, S. et al. Tuberculosis. 2009, 89(5), 334–353. doi: 10.1016/j.tube.
2009.05.008.
88. Gilbert, B. C.; Silvester, S.; Walton, P. H. J. Chem. Soc., Perkin Trans. 1999, 2(6),
1115–1122. doi: 10.1039/a901179j.
89. Calvo, J. S.; Lopez, V. M.; Meloni, G. Metallomics 2018, 10(12), 1777–1791. doi:
10.1039/C8MT00264A.
90. Speer, A. et al. Antimicrob. Agents Chemother. 2013, 57(2), 1089–1091. doi:
10.1128/AAC.01781-12.
91. Djoko, K. Y.; Goytia, M. M.; Donnelly, P. S.; Schembri, M. A.; Shafer, W. M.;
McEwan, A. G. Antimicrob. Agents Chemother. 2015, 59(10),10. doi: 10.1128/
AAC.01289-15.
92. Chiem, K. et al. Antimicrob. Agents Chemother. 2015, 59(9), 5851–5853. doi:
10.1128/AAC.01106-15.
93. Patteson, J. B. et al. Science 2021, 374(6570), 1005–1009. doi: 10.1126/science.
abj6749.
94. Petering, D. H. Bioinorg. Chem. 1972, 1(4), 273–288. doi: 10.1016/S0006-
3061(00)81002-9.
95. Xiao, Z.; Donnelly, P. S.; Zimmermann, M.; Wedd, A. G. Inorg. Chem. 2008,
47(10), 4338–4347. doi: 10.1021/ic702440e.
Ligands as a Tool to Tune the Toxicity of Cu on Bacteria 43
96. Djoko, K. Y.; Paterson, B. M.; Donnelly, P. S.; McEwan, A. G. Metallomics 2014,
6(4), 854–863. doi: 10.1039/C3MT00348E.
97. Labuda, J.; Skatulokova, M.; Nemeth, M.; Gergely, S. Chem. Zvesti. 1984, 597—605.
98. Bagchi, P.; Morgan, M. T.; Bacsa, J.; Fahrni, C. J. J. Am. Chem. Soc. 2013, 135(49),
18549–18559. doi: 10.1021/ja408827d.
99. Sun, P. J. Quintus, Fernando, Henry, Freiser, Anal. Chem. 1964, 36(13), 2485–2488.
doi: 10.1021/ac60219a034.
100. Budimir, A.; Humbert, N.; Elhabiri, M.; Osinska, I.; Biruš, M.; Albrecht-
Gary, A.-M. J. Inorg. Biochem. 2011, 105(3), 490–496. doi: 10.1016/j.jinorgbio.
2010.08.014.
101. Smith, R. M.; Martell, A. E. in Critical Stability Constants. Boston, MA: Springer
US, 1989. doi: 10.1007/978-1-4615-6764-6.
102. Wehbe, M. et al. Drug Deliv. Transl. Res. 2018, 8(1), 239–251. doi: 10.1007/
s13346-017-0455-7.
103. Postnikova, G. B.; Shekhovtsova, E. A. Biochemistry Moscow 2016, 81(13), 1735–
1753. doi: 10.1134/S0006297916130101.
104. Chen, D.; Darabedian, N.; Li, Z.; Kai, T.; Jiang, D.; Zhou, F. Anal. Biochem. 2016,
497, 27–35. doi: 10.1016/j.ab.2015.12.014.
105. Hendrickson, A. R.; Martin, R. L.; Rohde, N. M. Inorg. Chem. 1976, 15(9), 2115–
2119. doi: 10.1021/ic50163a021.
106. McBryde, W. A. E. Can. J. Biochem. 1967, 2093–2100.
107. Portelinha, J. et al. Chem. Rev. 2021, acs.chemrev.0c00921. doi: 10.1021/acs.
chemrev.0c00921.
108. Alexander, J. L.; Thompson, Z.; Cowan, J. A. ACS Chem. Biol. 2018, 13(4), 844–
853. doi: 10.1021/acschembio.7b00989.
109. Santoro, A.; Walke, G.; Vileno, B.; Kulkarni, P. P.; Raibaut, L.; Faller, P. Chem.
Commun. 2018, 54(84), 11945–11948. doi: 10.1039/C8CC06040A.
110. Conklin, S. E.; Bridgman, E. C.; Su, Q.; Riggs-Gelasco, P.; Haas, K. L.; Franz, K.
J. Biochemistry 2017, 56(32), 4244–4255. doi: 10.1021/acs.biochem.7b00348.
111. Melino, S.; Santone, C.; Di Nardo, P.; Sarkar, B. FEBS J. 2014, 281(3),
657–672. doi: 10.1111/febs.12612.
112. Esmieu, C.; Ferrand, G.; Borghesani, V.; Hureau, C. Chem. Eur. J. 2021, 27(5),
1777–1786. doi: 10.1002/chem.202003949.
113. Schwab, S.; Shearer, J.; Conklin, S. E.; Alies, B.; Haas, K. L. J. Inorg. Biochem.
2016, 158, 70–76. doi: 10.1016/j.jinorgbio.2015.12.021.
114. Santoro, A.; Ewa Wezynfeld, N.; Vašák, M.; Bal, W.; Faller, P. Chem. Commun.
2017, 53(85), 11634–11637. doi: 10.1039/C7CC06802F.
115. Stefaniak, E.; Płonka, D.; Szczerba, P.; Wezynfeld, N. E.; Bal, W. Inorg. Chem.
2020, 59(7), 4186–4190. doi: 10.1021/acs.inorgchem.0c00427.
116. Bouraguba, M. et al. J. Inorg. Biochem. 2020, 213, 111255. doi: 10.1016/j.jinorg-
bio.2020.111255.
117. Salina, E. G. et al. Metallomics 2018, 10(7), 992–1002. doi: 10.1039/C8MT00067K.
Zuily et al.
118. Djoko, K. Y.; Donnelly, P. S.; McEwan, A. G. Metallomics 2014, 6(12), 2250–2259.
doi: 10.1039/C4MT00226A.
119. Oliveri, V. Coord. Chem. Rev. 2020, 422, 213474. doi: 10.1016/j.ccr.2020.213474.
120. Dalecki, A. G.; Malalasekera, A. P.; Schaaf, K.; Kutsch, O.; Bossmann, S. H.;
Wolschendorf, F. Metallomics 8(4), 412–421. 2016. doi: 10.1039/C6MT00003G.
121. Crawford, C. L. et al. Metallomics 2019, 11(4), 784–798. doi: 10.1039/C8MT00316E.
122. Crawford, C. L.; Dalecki, A. G.; Perez, M. D.; Schaaf, K.; Wolschendorf, F.; Kutsch,
O. Sci Rep. 2020, 10(1), 8955. doi: 10.1038/s41598-020-65978-y.
123. Fang, L. et al. Sci Rep. 2016, 6(1), 25312. doi: 10.1038/srep25312.
124. Arai, N. et al. Antimicrob. Agents Chemother. 2019, 63(9), e00429–19. /aac/63/9/
AAC.00429-19.atom. doi: 10.1128/AAC.00429-19.
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0002
45
46 Faure et al.
List of Abbreviations
Ab Agaricus bisporus
Ao Aspergillus oryzae
AUS Aureusidin synthase
Bm Bacillus megaterium
CO Catechol oxidase
DHI Dihydroxyindole
DHICA Dihydroxyindole carboxylic acid
DNA Desoxyribonucleic acid
ESEEM Electron spin echo envelope modulation
EXAFS Extended X-Ray Absorption Fine Structure
H-BPEP 2,6-Bis[(bis(2-pyridylethyl)amino)
methyl]-4-methylphenol
H-BPMEP 2-[(Bis(2-pyridylmethyl)amino)methyl]-6-[(bis
(2-pyridylethyl)amino)methyl]-4-methylphenol
H-BPMP 2,6-Bis[(bis(2-pyridylmethyl)amino)
methyl]-4-methylphenol
Hc Hemocyanins
HOPNO 2-Hydroxypyridine N-oxide
Hs Homo sapiens
HSPNO 2-Pyridinethiol N-oxide
HYSCORE Hyperfine sublevel correlation
Ib Ipomoea batatas
Jr Juglans regia
KA Kojic acid
L-DOPA L-3,4-dihydroxyphenylalanine
Mim L-mimosine
Mj Marsupenaeus japonicas
MM Molecular mechanics
Ms Manduca sexta
Transition State Analogue Molecules as Mechanistic Tools 47
I. Introduction
Tyrosinases (TYRs) are copper-containing redox enzymes present in
prokaryotes and eukaryotes (plants, arthropods, fungi, and mammals)
where they are involved in tissue pigmentation, wound healing, radiation
protection, and primary immune response. Acting in phenol oxidation,
TYRs belong to the class of polyphenol oxidases (PPOs). The TYR-
catalyzed oxygenation consists of two consecutive reactions, namely the
ortho-hydroxylation of phenol to catechol (cresolase or monophenolase
activity, EC 1.14.18.1) and its subsequent oxidation into ortho-quinone
(catecholase or diphenolase activity, EC 1.10.3.1) (Figure 1). In organ-
isms, TYR is involved in the biosynthesis of melanin pigments.
O 2 + 2 H+ H 2O 1/2 O2 H2O
HO HO O
R R R
TYR HO TYR O
organisms. The common feature of all melanins is that they are formed by
the polymerization of phenolic and indolic compounds catalyzed by TYR
or TYR-like enzymes.
III. Tyrosinase
A. Structures of the Tyrosinase Active Sites
TYR belong to type 3 copper-containing proteins, which encompass
hemocyanins (Hc), catechol oxidase (CO, EC 1.10.3.1), and aureusidin
synthase (AUS, EC 1.21.3.6).9–11 The active site of TYR had been identi-
fied very early as a structural analogue12,13 of that of CO14,15 and Hc16 for
which 3D X-ray structures were available. It was not until 2006 that
Matoba et al. confirmed this analogy with the resolution of the 3D struc-
ture of the recombinant TYR from Streptomyces castaneoglobisporus
bacteria (ScTYR).17 In type 3 copper-containing proteins, the common
feature is an active site composed of two copper ions at a distance of about
3.6 Å both coordinated by three N-imidazoles of histidine residues.18 In
fungi19–22 and plant23 TYRs known to date (Table 1), one of the six histi-
dine residues forms an unusual thioether bond with an adjacent cysteine
residue (Figure 3a). The role of this post-translational modification that
does not significantly modify the structure of the active site is not clear.
Several hypotheses have been made such as the stabilization of the binu-
clear copper site to facilitate the electron transfer during the catalytic
reactions24 or a role in the copper incorporation process into the apo-
TYR.19 However, despite this post-translational modification, the active
sites of all TYRs are similar and nearly superimposable (Figure 3b).
No structure of human TYR is yet available with a good resolution.28
However, two homology models based on ScTYR and Ipomoea batatas
(a) (b)
Figure 4. Overlay of the homology model of HsTYR (green) based on TRP130 and
the one obtains from AlphaFold AI approach (AF-P14679-F1 model_v2) (pink).
Copper ions are in blue.
(a) (b)
(c) (d)
HO
2 H+ OH
His His
II O II HO
Cu Cu
His His
O O O
H+
II O His His HisI
His II oxy
Cu Cu O2
His O His
His His
His His
Catecholase Cresolase O
I His
activity CuI Cu activity O
3 H+ His His His II
II
Cu Cu
His His His
His O
deoxy His
H2O
H2O His
O
O
H+
His H His
II O O O-atom transfer
II O His
Cu Cu O
II O II
His His His
Cu Cu
His His His His
I O
I
met H His
His
2 H+
HO
OH
Figure 6. General mechanism for the cresolase and catecholase TYR activities.
Transition State Analogue Molecules as Mechanistic Tools 55
kinetic study of the TYR catalyzed phenol hydroxylation, Itoh et al. have
shown that the rate-determining step is the O-atom transfer occurring
through an electrophilic aromatic substitution mechanism.38
With data from different groups, the TYR mechanism is quite well
established39 (Figure 7). The first step is the formation of a phenolate
bond on the CuA site of the oxy-TYR. Two mechanisms have been pro-
posed for the formation of the phenolate bond on the CuA site. The first
was proposed by Decker and Tuczek on the base of crystallographic data
of ScTYR.40 The authors proposed a pre-orientation of the phenolate by
p-staking interaction with the H194 of the CuB site followed by a pheno-
late sliding toward CuA (Figure 7A). With AbTYR, a similar proposal
involving a pre-orientation of the phenolate by a p-staking interaction
with a histidine residue of the CuB site was proposed by Dijkstra.41 Based
on crystallographic data on AoTYR, Itoh et al. proposed another mecha-
nism for the formation of the CuA-phenolate. While Decker and Tuczek
propose a shift from phenolate toward CuA, Itoh proposes an opposite
movement, that of CuA toward phenolate located in a stabilizing enzy-
matic pocket42 (Figure 7B). This latter proposal is supported by several
strong evidences obtained with AoTYR mutants, copper-depleted or zinc-
replaced AoTYR.
Both mechanisms lead to phenolate binding to CuA site in trans posi-
tion relative to H63 for ScTYR with a concomitant elongation of the
H63-CuA distance (H103 for AoTYR). Phenolate binding impulses a 90°
rotation of the O–O plan of the peroxo group toward the aromatic ring
that allows the peroxide σ* orbital to overlap with the p orbital of the
phenolate ortho-position thus facilitating the electrophilic attack of the
peroxide by the phenolate.39,40 This process leads to the formation of a
catecholate bridging the two copper centers. The structure of this cat-
echolate that is not well defined could be a κ2-catecholate on CuA (a), or
a m-1,4-catecholate (c) or the combination of both, that is (η2:η1)-
catecholate (b) (Figure 7C).
Figure 7. Mechanisms proposed by (A) Decker and Tuczek39 on ScTYR and (B) Itoh
on AoTYR42 with putative structures (a)–(c) for the catecholate-TYR form.
Transition State Analogue Molecules as Mechanistic Tools 57
common ancestral gene. While TRP2 was very early identified as zinc-
containing analogue involved in DHIC formation, TRP1 was firstly identi-
fied as copper-containing analog having a catecholase activity and
involved in the oxidation of DHI and DHICA into the eumelanin pigments
precursors.43 However, the recent 3D structure of TRP1 solved by Soler-
López and Dijkstra et al.44 revealed, against all expectations, the presence
of Zn(II) instead of Cu(II) ions in the active site thus ruling out the pos-
sible role of TRP1 in DHICA oxidation. Since this publication, TRP1
function has become enigmatic, and several hypotheses have been formu-
lated such as a role in stabilization of HsTYR as proposed by Dolinska
et al.45 or its contribution in increasing the ratio between eumelanin and
pheomelanin by promoting the synthesis of eumelanin as proposed by
Ishikawa et al.46 But an unresolved question still remains: “could TRP1
integrate Cu(II) ions in vivo?” Indeed, ex vivo experiments have shown
that catecholase activity (DHIC oxidation) can be partially conferred on
the human TRP1 when its Zn(II) ions are replaced by Cu(II) ions.44
Another possibility is that like TRP2 (Figure 8b), TRP1 may have an
(a)
(b)
Figure 8. Reaction catalyzed by TRP2 (a) and reaction which may be catalyzed by
TRP1 (b).
58 Faure et al.
activity linked to the Lewis acid property of the Zn(II) ions such as the
decarboxylation of Dopachrome into DHI that is supposed to be non-
catalyzed (Figure 8b).
B. Skin Whitening
Depigmenting agents are commonly used in dermatology in the treatment
of hyperpigmentary disorders.60 Kojic acid in association with hydroqui-
none is prescribed in Europe under medical supervision, in the treatment
of post-inflammatory hyperpigmentation61 and in melasma.62
Unfortunately, the effectiveness of these agents has led to their use being
diverted, toward non-therapeutic purposes, in voluntary depigmentation
with the objective of lightening the natural complexion of the skin.63,64
This practice, which is widespread in sub-Saharan Africa, has become a
real public health problem as it leads to serious dermatological and sys-
temic complications. The ingredients historically encountered in cosmetic
products intended to lighten the skin are hydroquinone, mercury deriva-
tives, and corticosteroids. Recent studies have demonstrated the harmful
effects of these active ingredients for human health with the development
of ochronosis, skin complications, addiction to steroids, periorbital hyper-
chromia, facial lesions, or even kidney failure.65,66 Although better con-
trolled, this practice is also rampant in Asia (China, Japan, and India)
where it is estimated that these products represent nearly 10% of the
cosmetics market.
In this context, measures must be taken to curb this tendency. Facing
the difficulty of fighting these dangerous practices through laws, the
development of harmless and inexpensive bleaching agents accessible to
all could be a solution to this problem.
V. Tyrosinase Inhibition
Because inhibition of TYR is a well-established approach to control
in vivo melanin production, the development of inhibitors has a huge eco-
nomic and industrial impact in medicine and cosmetics (vide supra) but
60 Faure et al.
(a)
(b)
(c)
(a)
(b)
Figure 10. TYR Inhibitors that target the copper dinuclear active site: (a) metal
chelating by the thiourea moieties in blue and (b) TSAs.
(a) (b)
Figure 11. TRP1 active site with (a) L-mimosine (PDB ID: 5m8n) and (b) kojic acid
(PDB ID: 5m8m).44 The coloring of the cycles in yellow shows the stabilizing
p-stacking interactions.
HsTYR, shows comparable inhibitory properties for the fungal and the
human TYRs.
1. L-Mimosine
A TRP1 structure with L-mimosine as ligand has been recently solved by
Soler-López and Dijkstra et al.44 Surprisingly, this structure reveals that
L-mimosine does not directly interact with the Zn(II) ions but that the two
O-atoms of L-mimosine are H-bonded to the water molecule bridging the
two Zn(II) ions (Figure 11a). The structure also points out additional inter-
actions such as p-stacking with the His381 and H-bonding of one of the
O-atom with Ser394.
Figure 12. Kojic acid bound to a binuclear copper center, (a) SaTYR-KA adduct
according to X-ray absorption spectroscopy studies34 (ScTYR numbering); (b) X-ray
crystal structure with a copper model complex,108 and (c) QM/MM dynamics simu-
lations for ScTYR-KA adduct.108 Stable position is obtained after 2.5 ps of QM/MM
dynamics simulations. The coloring of the cycles in yellow shows the stabilizing
p-stacking interaction and the dotted red line the H-bond interaction between S206
and KA.
(a) (b)
Figure 13. BmTYR active site with kojic acid (a) oriented to the active site (PDB ID:
5i38)102 and (b) in entrance to the active site (PDB ID: 3nq1).25 The coloring of the
cycles in yellow shows the stabilizing p-stacking interaction.
66 Faure et al.
(a) (b)
Figure 15. Tropolone in interaction with (a) AbTYR active site (PDB ID: 3y9x)20 and
(b) showing the hydrophobic pocket above the active site where tropolone is
stacked.
Figure 16. Tropolone in interaction with the TRP1 active site (PDB ID: 5m80).44
The coloring of the cycles in yellow shows the stabilizing p-stacking interaction.
4. HOPNO inhibitor
In the absence of 3D structures of TYR with the HOPNO, XAS spectra
were recorded on SaTYR in complex with HOPNO (Figure 18).
Significant modifications were observed in the XANES part of the XAS
spectra upon addition of HOPNO to the SaTYR indicating an interaction
of HOPNO involving a rearrangement around the coordination sphere of
Transition State Analogue Molecules as Mechanistic Tools 69
Figure 18. XAS of SaTYR alone (green) and with KA (blue) and HOPNO (red).
the Cu(II) ions. The same behavior was also observed upon addition of
kojic acid, but significant differences indicate that OPNOa and kojic acid
bind differently the copper center.
The biomimetic approach in developing small molecular complexes
reproducing the main structural characteristics of the active site and the
reactivity of metalloenzymes allows to better understand their structure–
function relationships [Chapters 3 and 4].116 With this strategy, Belle
et al. has reported the synthesis and characterization of a copper complex
based on H-BPMP ligand that reproduces: (i) the copper coordination
with 3 N-atoms and a m-phenolate and m-hydroxy groups allowing a copper–
copper distance of 2.966 Å and (ii) the catecholase activity of TYR and
CO109 (Figure 19).
On the contrary of the kojic acid adduct to BPMP(Cu2)OH complex
that features a η2:η1-KA coordination mode107 (vide supra), it was
observed a HOPNO adduct to BPMP(Cu2)OH complex featuring a
κ2-OPNO coordination mode on only one of two copper ions.97 This high-
lights the differences between inhibitors interaction at a dinuclear active
site and the difficulty to draw general conclusions for all TSA inhibitors.
a
Since the pKa for the N–OH group of HOPNO is 6.07 [98], in the conditions used
(pH = 7) HOPNO is mainly in the deprotonated form OPNO.
70 Faure et al.
(a)
Figure 20. OPNO ligand binding modes on dicopper(II) center. (a) met2-ScTYR
active site (PDB ID: 2zmy17) with the t factor117 for the two copper centers; (b,c) the
three different coordination mode obtained with model complexes.
(a) (b)
(c) (d)
Figure 21. The two minimized structures (a) the m-1,4-OPNO adduct and (b) the
κ1-OPNO adduct. (b) XAS simulation of the m-1,4-OPNO adduct and (d) of the
κ1-OPNO adduct. XAS of SaTYR (black plain line) and simulation (red dotted line).
dotted line vs. black plain line). These results tend to confirm a dynamic
behavior of the OPNO ligand in the TYR active site, with probably sev-
eral species in equilibrium such as the κ1-OPNO and m-1,4-OPNO
adducts.
5. HOPNO derivatives
Naturally occurring aurone compounds have been reported in the litera-
ture as inhibitors in melanin biosynthesis in human melanocytes.118 The
true effect of aurones is controversial since some compounds of this fam-
ily act as alternative substrates rather than inhibitors.119,120 Nevertheless,
aurones display strong affinities for TYR, which encourages their integra-
tions in TSA moieties such as HOPNO. Therefore, several HOPNO-
embedded compounds have been reported in the literature for TYR
inhibition.104,119,121
The HOPNO-embedded aurones I–III were evaluated for the inhibi-
tion of three forms of TYR, one mammalian (HsTYR), one bacterial
(SaTYR) and one fungal (AbTYR) (Table 3). HOPNO-embedded aurones
I–III exhibit better inhibition properties than the parent compound
HOPNO, although the effect is less pronounced for AbTYR. In the case
of SaTYR, the substitution of the aurone moieties seems does not signifi-
cantly modify the inhibition properties since the three compounds I–III
exhibit KI in the range 0.1 µM. On the contrary for the HsTYR, this effect
is more important, with a KI = 0.35 µM for compound I and KI = 1.20 µM.
for compound III.
The HOPNO-embedded aurones I–III were also tested in the inhibi-
tion of melanin biosynthesis in a human-integrated cellular model (MNT-1
74 Faure et al.
cells). Results from MNT-1 cell lysates confirmed the tendency observed
with purified HsTYR (I > II ~ III >> HOPNO). However, they displayed
a low capacity to reduce melanogenesis in MNT-1 whole cells. Except
compound II, the others exhibit a low cytotoxicity (Table 4).
HOPNO-embedded aurones have a great inhibition potency. However,
the lower efficiency in suppressing melanogenesis in MNT-1 human
whole cells remains to be addressed.
VI. Conclusions
From the medicinal point of view, the TYR TSA compounds are promis-
ing inhibitors because they target the binuclear copper active site, which
Transition State Analogue Molecules as Mechanistic Tools 75
Acknowledgment
The authors gratefully acknowledge the Cosmethics project, an
“Investissements d’Avenir” program (ANR-15-IDEX-02) and the French
national synchrotron facility Soleil (proposal number 20120202).
VII. References
1. Cordero, Casadevall, R. J. B.; Melanin, A. Curr. Biol. 2020, R135–R158.
2. Solano, Melanins, F. New J Sci. 2014, 1–28.
3. d’Ischia, M. et al. Pigment Cell Melanoma Res. 2013, 26, 616–633.
4. d’Ischia, M. et al. Pigment Cell Melanoma Res. 2015, 28, 520–544.
5. Hong, L.; Simon, J. D. J. Phys. Chem. 2007, 111, 7938–7947.
6. Chatelain, M.; Gasparini, J.; Jacquin, L.; Frantz, A. Biol. Lett. 2014, 10, 20140164.
7. McGraw, K. J. Oikos. 2003, 102, 402–406.
8. Butler, M. J.; Day, A. W. Can. J. Microbiol. 1998, 44, 1115–1136.
9. Kaintz, C.; Mauracher, S. G.; Rompel, A. Adv. Protein Chem. Str. 2014, 97, 1–35.
10. Belle C. Catechol Oxidase and Tyrosinase. In: Encyclopedia of Metalloproteins,
Kretsinger, Robert H.; Uversky, Vladimir N.; Permyakov, Eugene A. (Eds.).
Springer-Verlag Berlin Heidelberg 2013, 574–579.
11. Bijelic A., Rompel A. and Belle C. Tyrosinases: enzymes, models and related appli-
cations, In Series on Chemistry, Energy and the environment: Bioinspired chemistry,
from enzymes to synthetic models, Volume 5, M. Réglier (Ed), World Scientific.
2019, 5, 155–183.
12. Tepper, A.W.J.W., Lonardi, E., Bubacco, L. and Canters, G.W. Structure,
Spectroscopy, and Function of Tyrosinase; Comparison with Hemocyanin and
Catechol Oxidase. In Encyclopedia of Inorganic and Bioinorganic Chemistry, R.A.
Scott (Ed.) 2011 (https://doi.org/10.1002/9781119951438.eibc0683)
13. Solomon, E. I.; Sundaram, U. M.; Machonkin, T. E. Chem. Rev. 1996, 96, 2563–
2606.
14. Klabunde, T.; Eicken, C.; Sacchettini, J. C.; Krebs, B. Nat. Struct. Biol. 1998, 5,
1084–1090.
15. Prexler, S. M.; Frassek, M.; Moerschbacher, B. M.; Dirks-Hofmeister, M. E. Angew.
Chem. Int. Ed. 2019, 58, 8757–8761.
16. Volbeda, A.; Hol, W. G. J. J. Mol. Biol. 1989, 209, 249–279.
17. Matoba, Y.; Kumagai, T.; Yamamoto, A.; Yoshitsu, H.; Sugiyama, M. J. Biol.
Chem. 2006, 281, 8981–8990.
18. Kaintz, C.; Mauracher, S. G.; Rompel, A. Adv. Protein Chem. Str. 2014, 97, 1–35.
19. Fujieda, N. et al. J. Biol. Chem. 2013, 288, 22128–22140.
20. Ismaya, W. T. et al. Biochemistry. 2011, 50, 5477–5486.
Transition State Analogue Molecules as Mechanistic Tools 77
21. Mauracher, S. G.; Molitor, C.; Al-Oweini, R.; Kortz, U.; Rompel, A. Acta Cryst.
2014, 70, 2301–2315 ().
22. Pretzler, M.; Bijelic, A.; Rompel, A. Acta Cryst. 2017, 7:1810, 1–10.
23. Bijelic, A.; Pretzler, M.; Molitor, C.; Zekiri, F.; Rompel, A. Angew. Chem. Int. Ed.
Engl. 2015, 54, 14677–14680.
24. Klabunde, T.; Eicken, C.; Sacchettini, J. C.; Krebs, B. Nat. Struct. Biol. 1998, 5,
1084–1090.
25. Sendovski, M.; Kanteev, M.; Ben-Yosef, V. S.; Adir, N.; Fishman, A. J. Mol. Biol.
2011, 405, 227–237.
26. Li, Y.; Wang, Y.; Jiang, H.; Deng, J. Proc. Nat. Acad. Sci. 2009, 106, 17002–17006.
27. Masuda, T.; Momoji, K.; Hirata, T.; Mikami, B. FEBS J. 2014, 281, 2659–2673.
28. Lai, X.; Soler-López, M.; Wichers, H. J.; Dijkstra, B. W. Plos One. 2016, 11,
e0161697.
29. Favre, E.; Daina, A.; Carrupt, P.-A.; Nurisso, A. Chem. Biol. Drug Design. 2014, 84,
206–215.
30. Lai, X.; Wichers, H. J.; Soler-López, M.; Dijkstra, B. W. Chem. Eur. J. 2017, 24,
47–55.
31. Jumper, J. et al. Nature 2021, 596, 583–589.
32. Zou, C. et al. Molecules. 2017, 22, 1836–1847.
33. Matoba, Y.; Yoshitsu, H.; Jeon, H.-J.; Oda, K.; Noda, M.; Kumagai, T.; Sugiyama, M.
2008, PDB. doi: 10.2210/pdb2ZMY/pdb.
34. Bubacco, L.; Spinazze, R.; Longa, S.; Della, Benfatto, M. Arch. Biochem. Biophys.
2007, 465, 320–327.
35. Mirica, L. M. et al. J. Am. Chem. Soc. 2006, 128, 2654–2665.
36. Mirica, L. M.; Ottenwaelder, X.; Stack, T. D. P. Chem. Rev. 2004, 104, 1013–1045.
37. Nasir, M. S.; Cohen, B. I.; Karlin, K. D. J. Am. Chem. Soc. 1992, 114, 2482–2494.
38. Yamazaki, S.-I.; Itoh, S. J. Am. Chem. Soc. 2003, 125, 13034–13035.
39. Rolff, M.; Decker, H.; Tuczek, F. Chem. Soc. Rev. 2011, 40, 4077–22.
40. Decker, H.; Schweikardt, T.; Tuczek, F. Angew. Chem. Int. Ed. Eng. 2006, 45,
4546–4550.
41. Ismaya, W. T. et al. Acta Cryst. 2011, 67, 575–578.
42. Fujieda, N. et al. Angew. Chem. Int. Ed. Eng. 2020, 59, 13385–13390.
43. Gautron, A. et al. Pigm. Cell Melanoma R. 2021, 34, 836–852.
44. Lai, X.; Wichers, H. J.; Soler-López, M.; Dijkstra, B. W. Angew. Chem. Int. Ed. Eng.
2017, 56, 9812–9815.
45. Dolinska, M. B.; Wingfield, P. T.; Young, K. L.; Sergeev, Y. V. Pigment Cell
Melanoma Res. 2019, 32, 753–765.
46. Ishikawa, M.; Kawase, I.; Ishii, F. Biol. Pharm. Bull. 2007, 30, 677–681.
47. Bastonini, E.; Kovacs, D.; Picardo, M. Ann. Dermatol. 2016, 28, 279–289.
48. Boyle, A. L.; Boyle, J. L.; Haupt, H. M.; Stern, J. B.; Multhaupt, H. A. B. Arch.
Pathol. Lab. Med. 2002, 126, 816–822.
78 Faure et al.
49. Weinstein, D.; Leininger, J.; Hamby, C.; Safai, B. J. Clin. Aesthet. Dermatol. 2014,
7, 13–24.
50. World Cancer Research Fund International <https://www.wcrf.org/cancer-trends/
skin-cancer-statistics/>
51. Brożyna, A. A.; Jóźwicki, W.; Carlson, J. A.; Slominski, A. T. Hum. Pathol. 2013,
44, 2071–2074.
52. Rezaei, T. et al. Pigment Cell Melanoma Res. 2021, 34, 869–891.
53. Jawaid, S.; Khan, T. H.; Osborn, H. M. I.; Williams, N. A. O. Anticancer Agents
Med. Chem. 2009, 9, 717–727.
54. Vargas, A. J. et al. Integr. Cancer Ther. 2011, 10, 328–340.
55. Nishimura, M. I.; Al-Khami, A. A.; Mehrotra, S.; Wolfel, T. Cancer Therapeutic
Targets, Ed. J. L.; Marshall J. L.; Springer, New York, 2016, 1–8.
56. Cabaço, L. C.; Tomás, A.; Pojo, M.; Barral, D. C. Front. Oncol. 2022, 12, 887366.
57. Buitrago, E. et al. Curr. Top. Med. Chem. 2016, 16, 3033–3047.
58. Roulier, B.; Pérès, B.; Haudecoeur, R. J. Med. Chem. 2020, 63, 13428–13443.
59. Slominski, R. M. et al. Front. Oncol. 2022, 12, 842496.
60. Rendon, M.; Horwitz, S. Ann. Dermatol. Vénéréol. 2012, 139, S153–S158.
61. Desai, S. Treatment of Skin Disease Comprehensive. Therapeutic Strategies (5th
edition), Eds. Lebwohl, M. G.; Heymann, W. R.; Berth-Jones, J.; Elsevier, 2017,
658–661.
62. Ogden, S.; Griffiths, C. E. M. Treatment of Skin Disease Comprehensive. Therapeutic
Strategies (5th edition), Eds. Lebwohl, M. G.; Heymann, W. R.; Berth-Jones, J.;
2017, Elsevier, 493–495.
63. Qian, W. et al. Exp. Ther. Med. 2020, 20, 173–185.
64. Pillaiyar, T.; Manickam, M.; Namasivayam, V. J. Enzym. Inhib. Med. Chem. 2020,
32, 403–425.
65. Burki, T. Lancet Diabetes Endocrinol. 2020, 9, 10.
66. Desmedt, B. et al. J. Eur. Acad. Dermatol. 2016, 30, 943–950.
67. Loizzo, M. R.; Tundis, R.; Menichini, F. Compr. Rev. Food Sci. Food Saf. 2012, 11,
378–398.
68. Gębalski, J.; Graczyk, F.; Załuski, D. J. Enzym. Inhib. Med. Chem. 2022, 37, 1120–
1195.
69. Vaezi, M. J. Biomol. Struct. Dyn. 2022, 1–13.
70. Peng, Z. et al. Crit. Rev. Food Sci. 2022, 62, 1–42.
71. Obaid, R. J. et al. RSC Adv. 2021, 11, 22159–22198.
72. Hariri, R.; Saeedi, M.; Akbarzadeh, T. J. Pept. Sci. 2021, 27, e3329.
73. Riaz, R. et al. Mini Rev. Org. Chem. 2021, 18, 808–828.
74. Zhang, X. et al. J. Enzym. Inhib. Med. Chem. 2021, 36, 2104–2117.
75. Bonesi, M. et al. Curr. Med. Chem. 2019, 26, 3279–3299.
76. Mendes, E.; Perry, M.; de J.; Francisco, A. P. Expert Opin. Drug Discov. 2014, 9,
533–554.
Transition State Analogue Molecules as Mechanistic Tools 79
77. Kampatsikas, I.; Bijelic, A.; Pretzler, M.; Rompel, A. Sci. Rep. 2017, 7, 8860.
78. Criton, M.; Mellay-Hamon, V. L. Bioorg. Med. Chem. Letters 2008, 18, 3607–3610.
79. Yi, W. et al. Chem. Pharm. Bull. 2010, 58, 752–754.
80. Yi, W. et al. Chem. Pharm. Bull. 2009, 57, 1273–1277.
81. Yi, W. et al. Eur. J. Med. Chem. 2011, 46, 4330–4335.
82. Hałdys, K.; Latajka, R. Med. Chem. Comm. 2019, 10, 378–389.
83. Liu, J.; Yi, W.; Wan, Y.; Ma, L.; Song, H. Bioorg. Med. Chem. 2008, 16, 1096–1102.
84. Liu, J. et al. Eur. J. Med. Chem. 2009. 44, 1773–1778.
85. Li, Z.-C. et al. J. Agri. Food Chem. 2010, 2–5.
86. Xue, C.-B. et al. Bioorg. Med. Chem. 2007, 15, 2006–2015.
87. Thanigaimalai, P.; Hoang, T. A. L.; Lee, K.-C.; Kim, Y.; Jung, S.-H. Bioorg. Med.
Chem. Lett. 2010, 20, 2991–2993.
88. Pan, Z.-Z. et al. J. Agri. Food Chem. 2012, 60, 10784–10788.
89. Buitrago, E. et al. Inorg. Chem. 2014, 53, 12848–12858.
90. Tudela, J.; Lozano, J. A.; Garcia-Canovas, F. Phytochemistry. 1987, 26, 917–919.
91. Kyriakou, S. et al. Invest. New Drug. 2020, 38, 621–633.
92. Chen, J. S. et al. J. Agr. Food Chem. 1991, 39, 1396–1401.
93. Chen, J. S.; Wei, C. I.; Marshall, M. R. J. Agr. Food Chem. 1991, 39, 1897–1901.
94. Espín, J. C.; Wichers, H. J. J. Agr. Food. Chem. 1999, 47, 2638–2644.
95. Takahashi, S. et al. Bioorg. Med. Chem. 2010, 18, 8112–8118.
96. Yoshimori, A. et al. Bioorg. Med. Chem. 2014, 22, 6193–6200.
97. Peyroux, E. et al. Inorg. Chem. 2009, 48, 10874–10876.
98. Orio, M. et al. Chem. Eur. J. 2011, 17, 13482–13494.
99. Goliĉnik, M.; Stojan, J. Biochem. Mol. Biol. Edu. 2004, 32, 228–235.
100. Fogal, S. et al. Mol. Biotechnol. 2014, 57, 45–57.
101. Nesterov, A. et al. Chem. Pharm. Bull. 2008, 56, 1292–1296.
102. Deri, B. et al. Sci. Rep. 2016, 6, 34993–10.
103. Bochot, C. et al. Chem. Eur. J. 2013, 19, 3655–3664.
104. Haudecoeur, R. et al. ACS Med. Chem. Lett. 2017, 8, 55–60.
105. Bubacco, L.; Gastel, M.; van, Groenen, E. J. J.; Vijgenboom, E.; Canters, G. W. J.
Biol. Chem. 2003, 278, 7381–7389.
106. Gastel, M.; van, Bubacco, L.; Groenen, E.; Vijgenboom, E.; Canters, G. W. FEBS
Lett. 2000, 474, 228–232.
107. Benfatto, M.; Della Longa, S.; Pace, E.; Chillemi, G.; Padrin, C.; Natoli, C. R.;
Sanna, N. Comput. Phys. Commun. 2021, 265, 107992.
108. Bochot, C. et al. Chem. Comm. 2014, 50, 308–310.
109. Torelli, S. et al. Inorg. Chem. 2000, 39, 3526–3536.
110. Buitrago, E. et al. Chem. Eur. J. 2021, 27, 4384–4393.
111. Lee, Y. S.; Park, J. H.; Kim, M. H.; Seo, S. H.; Kim, H. J. Arch. Pharm. 2006, 339,
111–114.
112. Kahn, V.; Andrawis, A. Phytochemistry. 1985, 24, 905–908.
80 Faure et al.
113. Espin, J. C.; Wichers, H. J. J. Agri. Food Chem. 1999, 47, 2638–2644.
114. Yoshimori, A. et al. Bioorg. Med. Chem. 2014, 22, 6193–6200.
115. Takahashi, S. et al. Bioorg. Med. Chem. 2010, 18, 8112–8118.
116. “Bioinspired Chemistry: From Enzymes to Synthetic Models.”, Ed. M. Réglier, In
“Series in Chemistry, Energy and Environment”, Eds. K. Kadish and R. Guilard,
World Scientific. Vol. 5: 2019.
117. Addison, A. W.; Rao, T. N.; Reedijk, J.; van Rijn, J.; Verschoor, G. C. J. Chem. Soc.
Dalton Trans. 1984, 1349–1356.
118. Okombi, S. et al. J. Med. Chem. 2006, 49, 329–333.
119. Dubois, C. et al. Chem. Bio. Chem. 2012, 13, 559–565.
120. Marková, E. et al. J. Agri. Food Chem. 2016, 64, 2925–2931.
121. Haudecoeur, R. et al. Chem. Bio. Chem. 2014, 15, 1325–1333.
122. Dubois, C. La tyrosinase: étude de nouveaux effecteurs, Aix-Marseille Université,
Oct. 2012.
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0003
¶
Department of Chemistry, Indian Institute of Technology Roorkee,
Roorkee, Uttarakhand 247 667, India
†
Department of Chemistry, Indian Institute of Technology Kanpur,
Kanpur, Uttar Pradesh 208 016, India
‡
Present address: Department of Chemistry and Chemical Biology,
Indian Institute of Technology (Indian School of Mines) Dhanbad,
Dhanbad, Jharkhand 826 004, India
§
[email protected]
I. Introduction
A. General Considerations
Dioxygen is crucial for aerobic organisms, serving as a primary source of
energy with its thermodynamically favorable reduction to water. It is used
as an oxidant in nature. However, the use of dioxygen as an oxidant is not
straightforward. Dioxygen is a diradical species and hence its triplet
ground state electronic structure makes it unreactive, under ambient con-
ditions, toward most organic substrates, which are generally diamagnetic
81
82 Gupta and Mukherjee
and hence have singlet ground state. This is due to spin-forbidden nature
of the reaction.1,2 In nature, metalloenzymes activate O2 and its further
processing is done by reduction for oxidation reactions.3 The reduced
forms of O2 such as O22− (peroxide) or O2− (oxide) exist in singlet ground
states, therefore they can easily react with singlet organic molecules.
In addition, O atom incorporation into biological substrates occurs by
functionalization through mono- or dioxygenation processes mediated by
metalloenzymes. The corresponding metalloenzymes are called monoox-
ygenases and dioxygenases, respectively. The former one incorporates
one O atom to the substrate and the other O atom of dioxygen is reduced
to H2O. The latter one incorporates both the O atoms to the substrate.
Molecular-level understanding of basic chemical principles and
mechanisms of O2 activation at metal sites and its subsequent processing
for oxygenations and oxidations of organic molecules have been of inter-
est for a few decades now. The continued interests are because of their
potential relevance to biological processes.4–16 Due to its generally acces-
sible Cu(II) and Cu(I) states and bioavailability, copper plays a wide
variety of roles in nature.1a,4
In the domain of synthetic modeling approach involving biomimetic
studies,17 many factors affect the reactivity of low-molecular-weight
metal complexes with tailor-made organic ligands toward dioxygen.1–16,17b
Since the reaction causes one-electron oxidation of the metal center with
reduction of dioxygen, the reduction potential of MII/MI (for M = Cu) or
MIII/MII (for M = Fe) redox couples is one of the important factors gov-
erning the reaction between the metal complex and dioxygen.9c The redox
potentials of metal complexes vary dramatically depending upon the
ligand environment. The solvent exerts a strong effect as well. One-
electron reduction of dioxygen is a thermodynamically unfavorable pro-
cess, since for O2 + 1e– = O2– redox process the Eo = −0.35 V versus NHE
at pH 7. However, for the two-electron, O2 + 2H+ + 2e– = H2O2 reduction
process (Eo = 0.28 V vs. NHE at pH 7) and for four-electron, O2 + 4H+ +
4e− = 2H2O reduction process (Eo = 0.82 V vs. NHE at pH 7) are favorable
reactions. Although the reduction potential is a good predictor of the reac-
tivity of the metal complexes toward dioxygen, it is not directly correlated
with the feasibility for the formation of the dioxygen adduct, since the
Modeling Tyrosinase Activity Using m-Xylyl-Based Ligands 83
A. Hemocyanins
Hemocyanins (Hc) are the oxygen-carrier protein of molluscs and arthro-
pods. Hemocyanin, unlike hemoglobin, has no heme group; the copper is
bound directly to the protein side chain. Reversible binding of dioxygen
is depicted in Figure 1. X-ray structures of both deoxy- and oxy-forms of
Hc have been determined.4,14b,18
B. Tyrosinases
Polyphenol oxidase is a generic term for the group of enzymes that cata-
lyze the oxidation of phenolic compounds to produce brown color on cut
surfaces of fruits and vegetables.4,14c,19 The enzyme tyrosinase (Tyr)
catalyzes the hydroxylation of monophenols (tyrosine) to o-diphenols
(L-3,4-dihydroxyphenylalanine, L-DOPA; monophenolase activity) and
subsequent two-electron oxidation to o-quinones (L-DOPA-quinone;
diphenolase activity), which constitutes the first step of melanin biosynthe-
sis through a series of spontaneous, non-enzymatic reactions (Figure 2).
Tyrosinase appears to be ubiquitous in living organisms. It is widely
distributed in plants, fungi, bacteria, mammalians, and animals. When
potatoes, apples, bananas, sweet potatoes, or mushrooms are injured they
turn brown.4,19 This is due to the conversion of tyrosine to the pigment
melanin, by the sequence of reactions shown in Figure 2. The same process
causes skin to turn brown, following exposure to ultraviolet radiation. The
enzymatic reactions are catalyzed by tyrosinase. The enzyme is present in
86 Gupta and Mukherjee
the interior of the plant material and, since the reaction requires molecular
oxygen, the pigmentation does not occur until the interior is exposed.
Since melanin is a key pigment of some phenomena such as suntan, skin
disorder, and bruising of fruits, tyrosinase has practical and economic
importance, and thus attracting much attention from cosmetology, medi-
cine, and agriculture to control the synthesis of the melanin pigment.19
Tyrosinase is an O2-activating enzyme. The active site of tyrosinase
has a dinuclear copper center with each copper coordinated by three his-
tidines. The oxy-state of the enzyme contains a O22− ion bound to two
copper(II) centers in a η2-fashion (Figure 3).19a The X-ray structure of
oxy-Tyr is very similar to oxy-Hc (Figure 1), the only difference being the
active site of Tyr is more exposed to external substrate than in Hc.
C. Catechol Oxidase
Catechol oxidase (Cat Ox) is an enzyme that catalyzes the two-electron
oxidation of catechols to o-quinones (diphenolase activity/catecholase
Modeling Tyrosinase Activity Using m-Xylyl-Based Ligands 87
Distances
(Å)
{Cu2PS} {Cu2O2} {Cu2PE}
d(Cu…Cu) 3.51 2.80 4.36
d(O–O) 1.42 2.32 1.43
d(Cu–O) 1.92 1.82 1.85
Figure 5. Three isoelectronic Cu2O2 binding modes, their shorthand notations, and
geometrical parameters. L: supporting ligands.
N N
N N
were also reported with symmetrical Schiff base macrocyclic ligands. The
intermediates 29–32 are schematically displayed in Figure 9. The cores 30
and 31 have been shown to mediate tyrosinase-like phenolate ortho-
hydroxylation reactions (Figure 9). Apart from abovementioned intramo-
lecular aromatic ring hydroxylation reactions, a couple of biomimetic
Modeling Tyrosinase Activity Using m-Xylyl-Based Ligands 97
R 2+ 2+
N N N N
N N O
II
N Cu CuII N O
N O N N CuIII CuIII N
N N O
N N
R = H, Me
29 30
2+
2+
N O N
N N CuIII CuIII
N Cu II O O
N HN N
N O H
N N
CuII
N N N N
31 32
2+
N O O N
CuII CuII
N
N
N NN
N
33
binuclear copper complexes were synthesized and tested for their capabil-
ity to convert external monophenolic substrates to o-diphenols or
o-quinones.
Réglier et al. reported31a that a mixture of a ligand, containing pyri-
dylethylimine sidearms bridged by a biphenyl spacer (Figure 10) placed
98 Gupta and Mukherjee
Figure 10. (a) Réglier’s ligand with biphenyl spacer, (b) tridentate analogue of the
Casella’s ligand used in 29 (Figure 9), and (c) bidentate analogue of Réglier’s biphe-
nyl spacer ligand.
Figure 11. Intermediates (a) containing two coordinated phenolates and (b) asym-
metrically coordinated m-catecholato m-hydroxo-bridging.
HO OH
HO CO2Me
CO2Me
HO OH O O
HO
HO OH
(a) NaO CO2Me
CO2Me
HO OH
(b) NaO X
X
(i) acetone; 183 K; X = Cl, F, CO2Me, CN
Figure 13. Tyrosinase-like activity by 29–31 intermediates: (a) with 29, (b) with 30,
and (c) with 31.
rans-μμ-1,2-peroxo-dicopper(II): {CuII2(μμ-ηη1:ηη1-O2)}2++
2. T
{Cu2PE}
Similar to {Cu2PS} (Section 1), {Cu2PE} cores also contain two Cu(II)
centers bound to a O22− ligand; however, unlike a η2-binding mode as in
{Cu2PS}, {Cu2PE} cores have a η1-binding mode between the two Cu(II)
and O22− ions (Figure 5). In {Cu2PE}, the in-plane pip*(O−O) orbital of
O22− overlaps with the Cu(dz2) orbitals to form a bonding (HOMO-2) and
an antibonding (LUMO) FMOs (Figure 15). The Cu(dz2) orbitals attain a
head-to-head overlap with pip*(O−O). In {Cu2PS} cores, the Cu(dx2–y2)
and pip*(O−O) orbitals do not have head-to-head overlap like in {Cu2PE}
cores. Thus, the Cu−O distances in {Cu2PS} (~1.92 Å) are longer com-
pared to {Cu2PE} (~1.85 Å).
The O−O distance in {Cu2PS} (~1.42 Å) and {Cu2PE} (~1.43 Å) are
comparable as in both the cores only the O2 p-bond is broken and the
σ-bond is still intact. The {Cu2PS} core is more compact than {Cu2PE} as
the distance between the two Cu in {Cu2PS} (3.51 Å) is smaller than in
{Cu2PE} (4.36 Å). Thus, {Cu2PE} species are preferred in large-size
ligands, where due to the steric crowding of the ligands the two Cu cannot
achieve a shorter distance. The inorganic complexes containing {Cu2PE}
104 Gupta and Mukherjee
cores do not show EPR peaks, suggesting that the two Cu(II) centers are
antiferromagnetically coupled. This antiferromagnetic interaction is
achieved by following a super-exchange pathway in the bonding FMO
(HOMO-2).
Figure 19. Intramolecular aromatic ligand hydroxylation via {Cu2PS} core to the
product (24).
2-position of the first ring to the bridged oxygen and eventually yield a
µ-phenoxo µ-hydroxo product. Hence, Solà and coworkers showed the use
of a second m-xylyl ring for hydrogen transfer, such a mechanism could
be envisioned in tyrosinase since histidine moieties close to tyrosine may
assist the hydrogen transfer in aromatic C−H bond oxidation.
Figure 20. Formation of the σ-complex 3 from the {Cu2O2} intermediate in Tolman
model.
Figures 21. Two pathways leading to the hydroxylated product from the σ-complex
3 in the model complex of Tolman.
benzene ring decrease the barrier. Hence, the authors proposed that
{Cu2O2}’s attack on the aromatic ring is an electrophilic substitution
reaction.26b
The groups of Holthausen, Schindler, and Tuczek showed an intramo-
lecular aromatic ligand hydroxylation via {Cu2O2} species. Blomberg
and coworkers employed DFT to investigate aromatic hydroxylation of an
external phenolate via a {Cu2O2} species (Figures 22).51 For this purpose,
the Blomberg group chose an experimental reaction established by the
Stack group. In this reaction, Stack and coworkers used a DBED (DBED =
N,N′-di-tert-butyl-ethylenediamine) supported copper/O2 complex.27
DFT-computed energetics of DBED-supported {Cu2O2} and {Cu2PS}
systems confirmed that the two isomers are in equilibrium. However, after
the addition of the external phenolate substrate in the reaction mixture, the
equilibrium is shifted toward the phenolate-bound {Cu2O2} isomer. In
the next step, {Cu2O2} core at attacks 2-position of the ring carbon (of
phenolate) to yield a σ complex. Subsequently, the bridged m-oxido
abstracts H atom from 2-position of the ring to yield the hydroxylated
114 Gupta and Mukherjee
Figure 24. Equilibrium between {Cu2PS} and {Cu2O2} cores (before and after phe-
nolate binding).
({Cu2PE}, {Cu2PS}, {Cu2O2}), and it was found that the {Cu2PE} is the
most stable species out of the three. So, the authors concluded that the
{Cu2PE} core is involved in aromatic hydroxylation. Not many DFT
studies are available on {Cu2PE} as compared to {Cu2PS} and {Cu2O2}
cores.
VI. Conclusions
The use of dioxygen as an oxidant in organic transformations would very
much benefit from fundamental knowledge on the mechanisms of its acti-
vation. Copper enzymes have been illustrated to process dioxygen in an
efficient and purposeful manner. Modeling copper–dioxygen chemistry
occurring in enzymes has attracted the attention of synthetic inorganic
chemists. A powerful tool to extract mechanistic information on copper-
catalyzed biochemical transformations, such as the tyrosinase reaction, is
the study of small-molecule complexes, which mimic the active site struc-
ture of their biological counterparts. An additional challenge in copper–
dioxygen chemistry has been the synthesis of catalytic model systems of
118 Gupta and Mukherjee
VII. References
1. (a) Trammell, R.; Rajabimoghadam, K.; Garcia-Bosch, I. Chem. Rev. 2019, 119,
2954−3031. (b) Solomon, E. I.; Stahl, S. S. Chem. Rev. 2018, 118, 2299−2301.
2. Serrano-Plana, J.; Garcia-Bosch, I.; Company, A.; Costas, M. Acc. Chem. Res. 2015,
48, 2397−2406.
3. Que, L., Jr.; Tolman, W. B. Nature 2008, 455, 333−340.
4. (a) Solomon, E. I.; Heppner, D. E.; Johnston, E. M.; Ginsbach, J. W.; Cirera, J.;
Qayyum, M.; Kieber-Emmons, M. T.; Kjaergaard, C. H.; Hadt, R. G.; Tian, L. Chem.
Rev. 2014, 114, 3659−3853. (b) Solomon, E. I.; Chen, P.; Metz, M.; Lee, S.-K.;
Modeling Tyrosinase Activity Using m-Xylyl-Based Ligands 119
Palmer, A. E. Angew. Chem. Int. Ed. 2001, 40, 4570−4590. (c) Solomon, E. I.;
Sundaram, U. M.; Machonkin, T. E. Chem. Rev. 1996, 96, 2563–2605.
5. (a) Quist, D. A.; Diaz, D. E.; Liu, J. J.; Karlin, K. D. J. Biol. Inorg. Chem. 2017, 22,
253–288. (b) Garcia-Bosch, I.; Cowley, R. E.; Díaz, D. E.; Peterson, R. L.; Solomon,
E. I.; Karlin, K. D. J. Am. Chem. Soc. 2017, 139, 3186−3195. (c) Hatcher, L. Q.;
Karlin, K. D. J. Biol. Inorg. Chem. 2004, 9, 669–683. (d) Tyeklár, Z.; D. Karlin,
K. D. Acc. Chem. Res. 1989, 22, 241–248. (e) Karlin, K. D.; Gultneh, Y. Prog. Inorg.
Chem. 1987, 35, 219–327.
6. (a) Keown, W.; Gary, J. B.; Stack, T. D. P. J. Biol. Inorg. Chem. 2017, 22, 289–305.
(b) Citek, C.; Herres-Pawlis, S.; Stack, T. D. P. Acc. Chem. Res. 2015, 48, 2424–2433.
(c) Stack, T. D. P. Dalton Trans. 2003, 1881–1889.
7. (a) Elwell, C. E.; Gagnon, N. L.; Neisen, B. D.; Dhar, D. Spaeth, A. D.; Yee, G. M.;
Tolman, W. B. Chem. Rev. 2017, 117, 2059−2107. (b) Tolman, W. B. J. Biol. Inorg.
Chem. 2006, 11, 261–271. (c) Holland, P. L.; Tolman, W. B. Coord. Chem. Rev. 1999,
190–192, 855–869. (d) Tolman, W. B. Acc. Chem. Res. 1997, 30, 227–237.
8. (a) Mirica, L. M.; Ottenwaelder, X.; Stack, T. D. P. Chem. Rev. 2004, 104, 1013–
1045. (b) Lewis, E. A.; Tolman, W. B. Chem. Rev. 2004, 104, 1047–1076.
9. (a) Schindler, S. Eur. J. Inorg. Chem. 2000, 2311–2326. (b) Tolman, W. B. Struct.
Bonding (Berlin, Ger.) 2000, 97, 179−211. (c) Kitajima, N.; Moro-oka, Y. Chem.
Rev. 1994, 94, 737−757. (d) Kitajima, N. Adv. Inorg. Chem. 1992, 39, 1−77.
10. Itoh, S. Dicopper Enzymes. In Comprehensive Coordination Chemistry – II: From
Biology To Nanotechnology; McCleverty, J. A.; Meyer, T. J., Eds. (Que Jr., L.;
Tolman, W. B., Vol. Eds.); Elsevier/Pergamon: Amsterdam, 2004, Vol. 8, pp 369–393.
11. (a) Hatcher, L. Q.; Karlin, K. D. Adv. Inorg. Chem. 2006, 58, 131−184. (b) Karlin, K.
D.; Tyeklár, Z. Adv. Inorg. Biochem. 1994, 9, 123–172.
12. (a) Itoh, S.; Fukuzumi, S. Acc. Chem. Res. 2007, 40, 592–600. (b) Itoh, S.; Tachi, Y.
Dalton Trans. 2006, 4531–4538. (c) Kunishita, A.; Osako, T.; Tachi, Y.; Teraoka, J.;
Itoh, S. Bull. Chem. Soc. Jpn. 2006, 79, 1729–1741. (d) Komiyama, K.; Furutachi,
H.; Nagatomo, S.; Hashimoto, A.; Hayashi, H.; Fujinami, S.; Suzuki, M.; Kitagawa,
T. Bull. Chem. Soc. Jpn. 2004, 77, 59–72. (e) Itoh, S.; Fukuzumi, S. Bull. Chem. Soc.
Jpn. 2002, 75, 2081–2095.
13. Kodera, M.; Kano, K. Bull. Chem. Soc. Jpn. 2007, 80, 662–676.
14. (a) Decker, H.; Schweikardt, T.; Tuczek, F. Angew. Chem. Int. Ed. 2006, 45, 4546–
4550. (b) H. Decker, H.; Dillinger, R.; Tuczek, F. Angew. Chem., Int. Ed. 2000, 39,
1591–1595. (c) Decker, H.; Tuczek, F. Trends Biochem. Sci. 2000, 25, 392−397.
15. Battaini, G.; Granata, A.; Monzani, E.; Gullotti, M.; Casella, L. Adv. Inorg. Chem.
2006, 58, 185−233.
16. Que, Jr., L.; Tolman, W. B. Angew. Chem. Int. Ed. 2002, 41, 1114–1137.
17. (a) Solomon, E. I.; Lowery, M. D. Science 1993, 259, 1575−1581. (b) Karlin, K. D.
Science 1993, 261, 701−708. (c) Ibers, J. A.; Holm, R. H. Science 1980, 209,
223−235.
120 Gupta and Mukherjee
18. (a) Magnus, K. A.; Hazes, B.; Ton-That, H.; Bonaventura, C.; Bonaventura, J.; Ho1,
W. G. J. Proteins: Structure, Function and Genetics 1994, 19, 302−309. (b) Hazes,
B.; Magnus, K. A.; Bonaventura, C.; Bonaventure, J.; Dauter, Z.; Kalk, K. H.; Hol.
W. G. J. Protein Sci. 1993, 2, 597−619.
19. (a) Matoba, Y.; Kumagai, T.; Yamamoto, A.; Yoshitsu, H.; Sugiyama, M. J. Biol.
Chem. 2006, 281, 8981–8990. (b) Pillaiyar, T.; Manickam, M.; Namasivayam, V. J.
Enz. Inhib. Med. Chem. 2007, 32, 403–425. (c) Parvez, S.; Kang, M.; Chung, H.-S.;
Bae, H. Phytother. Res. 2007, 21, 805–816. (d) Zolghadri, S.; Bahrami, A.; Khan, M.
T. H.; Munoz-Munoz, J.; Garcia-Molina, F.; Garcia-Canovas, F.; Saboury, A. A. J.
Enz. Inhib. Med. Chem. 2019, 34, 279–309. (e) Shi, F.; Xie, L.; Lin, O.; Tong, C.; Fu,
Q.; Xu, J.; Xiao, J.; Shi, S. Food Chem. 2020, 312, 126042.
20. (a) Klabunde, T.; Eicken, C.; Sacchettini, J. C.; Krebs, B. Nat. Struct. Biol. 1998, 5,
1084-1090. (b) Gerdemann, C.; Eicken, C.; Krebs, B. Acc. Chem. Res. 2002, 35,
183−191. (b) Dey, S. K.; Mukherjee, A. Coord. Chem. Rev. 2016, 310, 80–115. (c)
Ackermann, J.; Buchler, S.; Meyer, F. C. R. Chimie 2007, 10, 421–432. (d) Koval, I.
A.; Gamez, P.; Belle, C.; Selmeczi, K.; Reedijk, J. Chem. Soc. Rev. 2006, 35, 814–
840. (e) Selmeczi, K.; Réglier, M.; Giorgi, M.; Speier, G. Coord. Chem. Rev. 2003,
245, 191–201.
21. (a) Karlin, K. D.; Hayes, J. C.; Gultneh, Y.; Cruise, R. W.; McKown, J. W.;
Hutchinson, J. H.; Zubieta, J. J. Am. Chem. Soc. 1984, 106, 2121−2128. (b) Karlin,
K. D.; Nasir, M. S.; Cohen, B. I.; Cruse, R. W.; Kaderii, S.; Andreas D. Zuberbühler,
A. D. J. Am. Chem. Soc. 1994, 116, 1324–1336. (c) Pidcock, E.; Obias, H. V.; Zhang,
C. X.; Karlin, K. D.; Solomon, E. I. J. Am. Chem. Soc. 1998, 120, 7841−7847.
22. (a) De, A.; Mandal, S.; Mukherjee, R. J. Inorg. Biochem. 2008, 102, 1170–1189.
(b) Spodine, E.; Manzur, J. Coord. Chem. Rev. 1992, 119, 171–198. (c) Sorrell, T.N.
Tetrahedron 1989, 45, 3–68.
23. Mukherjee, R. Copper. In Comprehensive Coordination Chemistry II: From Biology
to Nanotechnology; McCleverty, J. A.; Meyer, T. J., Eds. (Fenton, D. E., Vol. Ed.);
Elsevier/Pergamon: Amsterdam, 2004; Vol. 6, pp 747−910.
24. Rolff, M.; Schottenheim, J.; Decker, H.; Tuczek, F. Chem. Soc. Rev. 2011, 40,
4077–4098.
25. (a) Mahapatra, S.; Kaderli, S.; Llobet, A.; Neuhold, Y.-M.; Palanché, T.; Halfen, J.
A.; Young, V. G., Jr.; Kaden, T. A.; Que, L., Jr.; Zuberbühler, A. D.; Tolman, W. B.
Inorg. Chem. 1997, 36, 6343−6356. (b) Matsumoto, T,; Furutachi, H.; Kobino, M.;
Tomii, M.; Nagatomo, S.; Tosha, T.; Osako, T.; Fujinami, S.; Itoh, S.; Kitagawa, T.;
Suzuki, M. J. Am. Chem. Soc. 2006, 128, 3874−3875. (c) Foxon, S. P.; Utz, D.;
Astner, J.; Schindler, S.; Thaler, F.; Heinemann, F. W.; Liehr, G.; Mukherjee, J.;
Balamurugan, V.; Ghosh, D.; Mukherjee, R. Dalton Trans. 2004, 2321−2328. (d)
Murthy, N. N.; Mahroof-Tahir, M.; Karlin, K. D. Inorg. Chem. 2001, 40, 628−635.
(e) Ghosh, D.; Lal, T. K.; Ghosh, S.; Mukherjee, R. Chem. Commun. 1996, 13−14.
(f) Ghosh, D.; Mukherjee, R. Inorg. Chem. 1998, 37, 6597−6605. (g) Gelling, O. J.;
van Bolhuis, F.; Meetsma, A.; Feringa, B. L. Chem. Commun. 1988, 552−554. (h)
Modeling Tyrosinase Activity Using m-Xylyl-Based Ligands 121
Casella, L.; Rigoni, L. J. Chem. Soc., Comm. Commun. 1985, 1668−1669. (i) Casella,
L.; Gullotti, M.; Pallanza, G.; Rigoni, L. J. Am. Chem. Soc. 1988, 110, 4221-4227. (j)
Casella, L.; Gullotti, M.; Bartosek, M.; Pallanza, G.; Laurenti, E. J. Chem. Soc.,
Comm. Commun. 1991, 1235−1237. (k) Gloria Alzuet, G.; Casella, L.; Villa, M. L.;
Carugo, O.; Gullotti, M. J. Chem. Soc., Dalton Trans. 1997, 4789–4794. (l) Sorrell,
T. N.; Garrity, M. L. Inorg. Chem. 1991, 30, 210−215. (m) Mandal, S.; Mukherjee,
R. Inorg. Chim. Acta 2006, 359, 4019−4026. (n) Drew, M. G. B.; Trocha-Grimshaw,
J.; McKillop, K. P. Polyhedron 1989, 8, 2513−2515. (o) Sander, O.; Henß, A.; Näther,
C.; Würtele, C.; Holthausen, M. C.; Schindler, S.; Tuczek, F. Chem. Eur. J. 2008, 14,
9714–9729. (p) Mandal, S.; Mukherjee, J.; Lloret, F.; Mukherjee, R. Inorg. Chem.
2012, 51, 13148−13161. (q) Gupta, R.; Mukherjee, R. Inorg. Chim. Acta 1997, 263,
133−137. (r) Menif, R.; Mattell, A. E. J. Chem. Soc., Comm. Commun. 1989,
1521−1523. (s) Menif, R.; Martell, A. E.; Squattrito, P. J.; Clearfield, A. inorg. Chem.
1990, 29, 4723-4729. (t) Costas, M.; Ribas, X.; Poater, A.; Valbuena, J. M. L.; Xifra,
R.; Company, A.; Duran, M.; Sola, M.; Llobet, A.; Corbella, M.; Usón, M. A.; Mahía,
J.; Solans, X.; Shan, X.; Benet-Buchholz, J. Inorg. Chem. 2006, 45, 3569−3581.
(u) Poater, A.; Ribas, X.; Llobet, A.; Cavallo, L.; Solà, M. J. Am. Chem. Soc. 2008,
130, 17710−17717.
26. (a) Patrick L. Holland, P. L.; Rodgers, K. R.; Tolman, W. B. Angew. Chem. Int. Ed.
1999, 38, 1139–1142. (b) Becker, J.; Gupta, P.; Angersbach, F.; Tuczek, F.; Näther,
C.; Holthausen, M. C.; Schindler, S. Chem. Eur. J. 2015, 21, 11735–11744.
27. Mirica, L. M.; Vance, M.; Rudd, D. J.; Hedman, B.; Hodgson, K. O.; Solomon, E. I.;
Stack, T. D. P. Science 2005, 308, 1890–1892.
28. Fusi, V.; Llobet, A.; Mahía, J.; Micheloni, M.; Paoli, P.; Ribas, X.; Rossi, P. Eur. J.
Inorg. Chem. 2002, 987–990.
29. Hamann, J. N.; Rolff, M.; Tuczek, F. Dalton Trans. 2015, 44, 3251–3258.
30. (a) Palavicini, S.; Granata, A.; Monzani, E.; Casella, L. J. Am. Chem. Soc. 2005, 127,
18031–18036. (b) Company, A. ; Lamata, D.; Poater, A.; Solà, M.; Rybak-Akimova,
E. V.; Que, Jr., L.; Fontodrona, X.; Parella, T; Llobet, A.; Costas, M. Inorg. Chem.
2006, 45, 5239–5241. (c) Company, A.; Palavicini, S.; Garcia-Bosch, I.; Mas-
Ballesté, R.; Que, Jr., L.; Rybak-Akimova, E. V.; Casella, L.; Ribas, X.; Costas, M.
Chem. Eur. J. 2008, 14, 3535–3538. (d) Garcia-Bosch, I.; Company, A.; Frisch, J. R.;
Torrent-Sucarrat, M.; Cardellach, M.; Gamba, I.; Mireia Güell, M.; Casella, L.; Que,
Jr., L.; Ribas, X.; Luis, J. M.; Costas, M. Angew. Chem. Int. Ed. 2010, 49, 2406–2409.
(e) Company, A.; GÓmez, L.; Mas-Ballesté, R.; Korendovych, I. V.; Ribas, X.; Poater,
A.; Parella, T. ; Fontodrona, X.; Benet-Buchholz, J.; Solà, M.; Que, Jr., L.; Rybak-
Akimova, E. V.; Costas, M. Inorg. Chem. 2007, 46, 4997–5012.
31. (a) Réglier, M.; Jorand, C.; Waegell, B. J. Chem. Soc. Chem. Commun. 1990, 1752–
1755. (b) Battaini, G.; De Carolis, M.; Monzani, E.; Tuczek, F.; Casella, L. Chem.
Commun. 2003, 726–727. (c) Rolff, M.; Schottenheim, J.; Peters, G.; Tuczek, F.
Angew. Chem. Int. Ed. 2010, 49, 6438–6442.
32. Gupta, R.; Mukherjee, R. Tet. Lett. 2000, 41, 7763–7767.
122 Gupta and Mukherjee
33. (a) Casella, L.; Gullotti, M.; Radaelli, R.; Di Gennaro, P. J. Chem. Soc., Chem.
Commun. 1991, 1611–1612. (b) Casella, L.; Monzani, E.; Gullotti, M.; Cavagnino,
D.; Cerina, G.; Santagostini, L.; Ugo, R. Inorg. Chem. 1996, 35, 7516–7525.
34. Halfen, J. A.; Mahapatra, S.; Wilkinson, E. C. ; Kaderli, S.; Young, V. G.; Que, L.;
Zuberbühler, A. D.; Tolman, W. B. Science 1996, 271, 1397–1400.
35. Gherman, B. F.; Cramer, C. J. Coord. Chem. Rev. 2009, 253, 723–753.
36. Siegbahn, P. E. M.; Blomberg, M. R. A.; Chen, S.-L. J. Chem. Theory Comput. 2010,
6, 2040–2044.
37. Liakos, D. G.; Neese, F. J. Chem. Theory Comput. 2011, 7, 1511–1523.
38. Gupta, P.; Diefenbach, M.; Holthausen, M. C.; Förster, M.; Chem. Eur. J. 2017, 23,
1427–1435.
39. Cramer, C. J.; Wloch, M.; Piecuch, P.; Puzzarini, C.; Gagliardi, L. J. Phys. Chem. A
2006, 110, 1991–2004.
40. Rohrmüller, M.; Hoffmann, A.; Thierfelder, C.; Herres-Pawlis, S.; Schmidt, W. G. J.
Compt. Chem. 2015, 36, 1672–1685.
41. (a) Werner, H.-J.; Knowles, P. J. J. Chem. Phys. 1988, 89, 5803–5814; (b) Knowles,
P. J.; Werner, H. J. Chem. Phys. Lett. 1988, 145, 514–522.
42. (a) Kowalski, K.; Piecuch, P. J. Chem. Phys. 2000, 113, 18–35. (b) Piecuch, P.;
Kowalski, K.; Pimienta, I. S. O.; Fan, P. D.; Lodriguito, M.; McGuire, M. J.;
Kucharski, S. A.; Kuś, T.; Musiał, M. Theor. Chem. Acc. 2004, 112, 349–393. (c)
Piecuch, P.; Wloch, M.; Gour, J. R.; Kinal, A. Chem. Phys. Lett. 2006, 418, 467–474.
43. (a) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1988, 37, 785–789. (b) Becke, A. D.
Phys. Rev. A 1988, 38, 3098–3100.
44. Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Chem. Phys. Lett. 1989, 157,
200–206.
45. Reiher, M.; Salomon, O.; Artur Hess, B. Theor. Chem. Acc. 2001, 107, 48–55.
46. (a) Neese, F.; Hansen, A.; Liakos, D. G. J. Chem. Phys. 2009, 131, 064103–064117.
(b) Liakos, D. G.; Hansen, A.; Neese, F. J. Chem. Theory Comput. 2010, 7, 76–87.
47. (a) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. J. Phys. Chem.
1994, 98, 11623–11627. (b) Becke, A. D. J. Chem. Phys. 1993, 98, 5648–5652.
48. Nasir, M. S.; Cohen, B. I.; Karlin, K. D. J. Am. Chem. Soc. 1992, 114, 2482–2494.
49. Liu, Y. F.; Shen, J.; Chen, S.-L.; Qiao, W.; Zhou, S.; Hong, K. Dalton Trans. 2019,
48, 16882–16893.
50. (a) Karlin, K. D.; Dahlstrom, P. L.; Cozzette, S. N.; Scensny, P. M.; Zubieta, J. J.
Chem. Soc. Chem. Commun. 1981, 881–882. (b) Karlin, K. D.; Gultneh, Y.;
Hutchinson, J. P.; Zubieta, J. J. Am. Chem. Soc. 1982, 104, 5240–5242.
51. Liu, Y. F.; Yu, J. G.; Siegbahn, P. E. M.; Blomberg, M. R. A. Chem. Eur. J. 2013,
19, 1942–1954.
52. Besalú-Sala, P.; Magallσn, C.; Costas, M.; Company, A.; Luis, J. M. Inorg. Chem.
2020, 59, 17018–17027.
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0004
4 Monooxygenation of Phenols
by Small-molecule Models of
Tyrosinase: Correlations
Between Structure and
Catalytic Activity
I. Introduction
Copper enzymes are involved in important biochemical processes in ani-
mals or plants such as dioxygen transport and metabolism or electron
transfer and are present in the three domains of life.1,2 They are often
categorized by classes (type I–IV, CuA, CuB, CuZ, and Cu0), which derive
from their structural environment or the spectroscopic properties of the
metal center(s).1,3 Type III refers to a (oxygen-binding) binuclear copper
protein family comprising tyrosinase (Ty), tyrosinase-related proteins,
catechol oxidase (CO), and hemocyanin.1,3–5
Tyrosinase is a phenoloxidase and catalyzes the conversion and oxidation
of mono- and diphenols into the corresponding ortho-quinones whereas the
catechol oxidase only exhibits the latter reactivity.1,6–8 These enzymes mediate
123
124 Koch et al.
phenolic substrates, have been published in the past. This has provided
insight into relevant steps of the monophenolase cycle, that is, the forma-
tion of the peroxo complex, the interaction of the substrate with the peroxo
complex leading to a coordinated catecholate (including the role of base in
this reaction), the two-electron oxidation of the bound catecholate and the
release of the ortho-quinone product. As this chapter focuses on tyrosinase
activity, small-molecule models of catechol oxidase will not be discussed
here. Excellent reviews on the latter topic can be found elsewhere.3,12–16
Figure 1. The XYL system by Karlin et al., which is able to perform the hydroxyla-
tion of the aromatic ring in the ligand backbone and the L4-H system that performs
hydroxylation of the phenolic residue.19,28
126 Koch et al.
Figure 2. Early model systems with the ability to catalytically convert monophenols
to ortho-quinones.35,36,55
(d)
Figure 3. The first mononucleating model system with tyrosinase activity, devel-
oped by Stack et al. and (a) the m-η2:η2-peroxo dicopper(II) complex, which is formed
upon coordination to copper(I) and oxygenation at low temperatures, (b) the ternary
adduct between a bis-µ-oxo dicopper(III) complex with bound phenolate, (c) the
catecholato adduct, and (d) the resulting copper(II)-semiquinone complex.34,47
130 Koch et al.
(a) (b)
Figure 4. Model systems with an imine moiety and an N-heterocycle and the
obtained TONs under Bulkowski–Réglier conditions with 2,4-DTBP-H as substrate
and a catalyst concentration of 500 µmol/L. (a) The TONs were obtained using 0.1
mmol catalyst, 10 eq. of 4-tert-butylphenol and 5 eq. of NEt3.37,39,40,44,48
Table 1. pKa values of imines and different N-heterocycles (in their protonated
form).65,66,67
imine imidazole benzimidazole pyridine pyrazole triazole
pKa 7–8 7 5.6 5.2 2.5 1.17
134 Koch et al.
the other hand, exchange of pyridine by imidazole should lower the cata-
lytic activity as s-donation is increased, again in agreement with the
experimental observation.
Of particular interest is the large activity increase observed upon
replacing pyridine by benzimidazole. Based on the pKa values, this sub-
stitution should have little effect. However, benzannulation lowers both
the HOMO and the LUMO of imidazole.67,68 While the former explains
the decrease of s-donation (and concomitant decrease of pKa) upon going
from imidazole to benzimidazole (cf. Table 1), the latter acts to increase
the p-backbonding capability, that is, benzimidazole should be more back-
bonding than imidazole. Probably, the combination of both effects makes
benzimidazole superior to pyridine when employed as a donor group in
copper monooxygenation catalysts. This has already been exploited by
Casella whose L66- and L6-supported tyrosinase models, for example,
contained benzimidazoles as terminal N-donors (and not pyridines; see
above).55,56
Figure 5. Symmetric model systems for tyrosinase and their catalytic activity under
Bulkowski-Réglier conditions vs. 2,4-DTBP-H and a catalyst concentration of 500
µmol/L.37,42,44,64
complex supported by the BIMZ ligand, again in line with the results
obtained for the asymmetric ligands.
Overall, however, the catalytic activities of the Cu(I) complexes sup-
ported by the symmetric ligands are conspicuously lower than those of
their counterparts with asymmetric ligands. This result was surprising as
it is not compatible with the bonding properties discussed so far: based on
the corresponding pKa values, all of the employed heterocycles are
weaker s-donors than imine (cf. Table 1). From a purely electronic point
of view, replacement of imine by these moieties thus should increase (and
not decrease) the catalytic activity. The complete catalytic inactivity of the
DPM complex was even more striking; based on the activity sequence of
the asymmetric systems it should lead to a TON between that of the imi-
dazole and the pyrazole system.
In an attempt to understand the described findings, the formation and
possible constitution of the ternary intermediate 3 (cf. Scheme 3) are con-
sidered in more detail. Note that in a stoichiometric reaction mode (upon
which the mechanism depicted in Scheme 3 is based) the substrate is
bound to the Cu(I) precursor before oxygenation. In a catalytic run, how-
ever, the substrate is added to the Cu(I) precursor [Cu(L)(CH3CN)2]+
(L = bidentate ligand) and will bind to the µ-η2:η2-peroxo complex formed
upon oxygenation of the latter, generating ternary intermediate 3. In order
to get hydroxylated, the substrate has to bind within the Cu2O2 plane
(Scheme 4).6
Such a configuration, in which the peroxide s* orbital overlaps with
the p-system of the arene ring6 can in principle be reached by two path-
ways which correspond to an associative and a dissociative mechanism. In
the associative scheme initial binding of the substrate proceeds axially to
one of the Cu(II) centers of the planar µ-η2:η2-dicopper complex; e.g. CuB
N
N O N
Cu Cu
O N
O
Scheme 4. Overlap of a phenolate p-type orbital and the σ* orbital of the peroxo
ligand.6
Monooxygenation of Phenols by Small-molecule Models of Tyrosinase 137
Table 2. Observed TOFs (min–1) after 15 min for different model systems upon
catalytic conversion of 2,4-DTBP-H.37,39,40,42,44
Lhpz1 BPM
Lpy1 Lbzm1 (Lhpz2) Limz1 (mBPM/dmBPM) BIMZ
TOF 0.41 0.84 0.85 0.38 0.27 0.17
(0–15 min) (0.62) (0.24/ 0.22)
intermediate would be less stabilized and total loss of the bidentate ligand
may occur. It is thus conceivable that mixed ligands with one imine group
(as a comparatively strong donor) and one (weaker coordinating)
N-heterocycle follow a dissociative pathway for the binding of the sub-
strate whereas symmetric ligands with two comparatively weak donors
follow an associative pathway. The latter pathway is a priori kinetically
disfavored compared to the former due to the fact that equatorial coordi-
nation of the substrate only occurs after an axial → equatorial rearrange-
ment whereas this is not necessary within the dissociative scheme. This
mechanistic difference may explain the overall lower activity of the sym-
metric ligands, which is also evident from lower observed TOFs as com-
pared to their asymmetric, imine-containing analogs
(cf. Table 2). The total lack of catalytic activity observed for the DPM
system, finally, may be associated with the fact that this is the only sym-
metric ligand that is composed of two six-membered rings as opposed to
the other symmetric ligands that contain two N-heterocycles with five-
membered rings. This renders the bite angle of DPM smaller than that of
the latter systems, which in turn could become relevant, e.g., in the course
of the axial → equatorial rearrangement.
Figure 6. The LOLX model systems with the ability to mediate the tyrosinase reac-
tion in a stoichiometric fashion under Bulkowski-Réglier conditions vs. 2,4-DTBP-H
and a catalyst concentration of 500 µmol/L.44
Figure 7. Ligands containing different N-heterocycles and the TON of the corre-
sponding model systems under Bulkowski-Réglier conditions with 2,4-DTBP-H as
substrate and a catalyst concentration of 500 µmol/L.45,63
Table 4. Overview of the obtained TONs with different bidentate model systemsa
and different substrates.
BPM TMP1
Lhpz1 (mBPM/ PMP (TMP2/
Lpy1 (Lhpz2) dmBPM) LIMZ1 BIMZ LOL1–3 (dmPMP) TMP3)
2,4-DTBP-H 18 29 21 16 9 <1 14 20
(23) (19 / 11) (11) (22 / 24)
3-TBP-H 22 20 6 25 20
(15 / 12) (13) (19 / 24)
4-MeOP-H 35 34 6 34 42
(10 / 15) (33) (45 / 48)
4-MeP-H 14
(12 / 10)
P-H 9
(9 / 9)
NATEE 18 25 0
(15 / 11)
a
DPM and L2 are catalytically inactive and therefore not listed
IV. Summary
Tyrosinase is a dinuclear copper protein that converts L-tyrosine to
L-dopaquinone. Besides structural and mechanistic investigations of the
natural enzyme relevant information about the underlying chemistry can
be obtained using small-molecule model systems. In the last decade sev-
eral catalytic tyrosinase models with bidentate N-donor ligands have been
established. Through detailed investigations of these systems, correlations
between the properties of the ligand and the substrate on the one side and
the observed catalytic activity on the other could be obtained. An impor-
tant aspect of these correlations is the s-donor/p-acceptor strength of the
employed N-donor groups. Moreover, asymmetric ligands are more effec-
tive than their symmetric counterparts, which can be attributed to the
kinetics of substrate binding.
Apart from the electronic influence of the N-donor groups the pres-
ence of bulky residues in the ligand backbone also influences the catalytic
activity of the system. While stabilizing the peroxo complex, sterically
demanding substituents may inhibit coordination of the substrate and
electrophilic attack by the peroxo group. On the other hand, small steric
hindrance in the vicinity of the ortho-position to the hydroxyl group and
an increased electron density within the aromatic system (e.g., through a
+M-effect of a methoxy group in 4-position) facilitate bonding of a phe-
nolic substrate to the copper-peroxo intermediate and subsequent electro-
philic attack by the peroxo group. The reactivity of the presented
tyrosinase model systems towards external monophenols thus is well
understood. Remaining challenges to be adressed refer to increasing the
turnover numbers, which are still low, and reducing the amount of base
needed for a catalytic activity of the investigated model systems.
V. Acknowledgment
The authors would like to thank M. Rolff, J. Schottenheim, J. Hamann, F.
Wendt, B. Herzigkeit, R. Schneider and other present and former mem-
bers of the copper workgroup for their contributions to the research
described in this review.
Monooxygenation of Phenols by Small-molecule Models of Tyrosinase 149
VI. References
1. Solomon, E. I.; Heppner, D. E.; Johnston, E. M.; Ginsbach, J. W.; Cirera, J.; Qayyum,
M.; Kieber-Emmons, M. T.; Kjaergaard, C. H.; Hadt, R. G.; Tian, L. Chem. Rev. 2014,
114, 3659.
2. Festa, R. A.; Thiele, D. J. Curr. Biol. 2011, 21, R877–83.
3. Koval, I. A.; Gamez, P.; Belle, C.; Selmeczi, K.; Reedijk, J. Chem. Soc. Rev. 2006, 35,
814.
4. Decker, H.; Schweikardt, T.; Tuczek, F. Angew. Chem. Int. Ed. 2006, 45, 4546.
5. van Holde, K. E.; Miller, K. I.; Decker, H. J. Biol. Chem. 2001, 276, 15563.
6. Rolff, M.; Schottenheim, J.; Decker, H.; Tuczek, F. Chem. Soc. Rev. 2011, 40, 4077.
7. Hamann, J. N.; Herzigkeit, B.; Jurgeleit, R.; Tuczek, F. Coord. Chem. Rev. 2017, 334, 54.
8. Solem, E.; Tuczek, F.; Decker, H. Angew. Chem. Int. Ed. 2016, 55, 2884.
9. Battaini, G.; Granata, A.; Monzani, E.; Gullotti, M.; Casella, L. Adv. Inorg. Chem.
2006, 58, 185.
10. a) Mirica, L. M.; Ottenwaelder, X.; Stack, T. D. P. Chem. Rev. 2004, 104, 1013. b)
Lewis, E. A.; Tolman, W. B. Chem. Rev. 2004, 104, 1047. c) Que, L.; Tolman, W. B.
Nature. 2008, 455, 333. d) Hatcher, L. Q.; Karlin, K. D. J. Biol. Inorg. Chem. 2004,
9, 669. e) Karlin, K. D.; Itoh, S.; Rokita, S. Copper-oxygen chemistry. Wiley-VCH,
Hoboken, N. J., 2011, f) Quist, D. A.; Diaz, D. E.; Liu, J. J.; Karlin, K. D. J. Biol.
Inorg. Chem. 2017, 22, 253.
11. Elwell, C. E.; Gagnon, N. L.; Neisen, B. D.; Dhar, D.; Spaeth, A. D.; Yee, G. M.;
Tolman, W. B. Chem. Rev. 2017, 117, 2059.
12. Than, R.; Feldmann, A. A.; Krebs, B. Coord. Chem. Rev. 1999, 182, 211.
13. Simándi, L. I.; Simándi, T. M.; May, Z.; Besenyei, G. Coord. Chem. Rev. 2003, 245, 85.
14. Gerdemann, C.; Eicken, C.; Krebs, B. Acc. Chem. Res. 2002, 35, 183.
15. Selmeczi, K.; Réglier, M.; Giorgi, M.; Speier, G. Coord. Chem. Rev. 2003, 245, 191.
16. Dey, S. K.; Mukherjee, A. Coord. Chem. Rev. 2016, 310, 80.
17. a) Osório, Renata, E. H. M. B.; Neves, A.; Camargo, T. P.; Mireski, S. L.; Bortoluzzi,
A. J.; Castellano, E. E.; Haase, W.; Tomkowicz, Z. Inorg. Chim. Acta 2015, 435, 153.
b) Banu, K. S.; Chattopadhyay, T.; Banerjee, A.; Bhattacharya, S.; Suresh, E.; Nethaji,
M.; Zangrando, E.; Das, D. Inorg. Chem. 2008, 47, 7083. c) Banu, K. S.;
Chattopadhyay, T.; Banerjee, A.; Bhattacharya, S.; Zangrando, E.; Das, D. J. Mol.
Catal. A: Chem. 2009, 310, 34. d) Wegner, R.; Gottschaldt, M.; Görls, H.; Jäger, E.-G.;
Klemm, D. Chem. Eur. J. 2001, 7, 2143. e) Smith, S. J.; Noble, C. J.; Palmer, R. C.;
Hanson, G. R.; Schenk, G.; Gahan, L. R.; Riley, M. J. J. Biol. Inorg. Chem. 2008, 13,
499. f) Monzani, E.; Quinti, L.; Perotti, A.; Casella, L.; Gullotti, M.; Randaccio, L.;
Geremia, S.; Nardin, G.; Faleschini, P.; Tabbì, G. Inorg. Chem. 1998, 37, 553. g)
Garcia-Bosch, I.; Karlin, K. D. Copper Peroxide Bioinorganic Chemistry: From
Metalloenzymes to Bioinspired Synthetic Systems, Wiley-VCH, Hoboken, N. J., 2014.
18. Lo Presti, E.; Monzani, E.; Santagostini, L.; Casella, L. Inorg. Chim. Acta 2018, 481, 47.
150 Koch et al.
19. Karlin, K. D.; Hayes, J. C.; Gultneh, Y.; Cruse, R. W.; McKown, J. W.; Hutchinson,
J. P.; Zubieta, J. J. Am. Chem. Soc. 1984, 106, 2121.
20. Pidcock, E.; Obias, H. V.; Zhang, C. X.; Karlin, K. D.; Solomon, E. I. J. Am. Chem.
Soc. 1998, 120, 7841.
21. Pidcock, E.; Obias, H. V.; Abe, M.; Liang, H.-C.; Karlin, K. D.; Solomon, E. I. J. Am.
Chem. Soc. 1999, 121, 1299.
22. Sanyal, I.; Mahroof-Tahir, M.; Nasir, M. S.; Ghosh, P.; Cohen, B. I.; Gultneh, Y.;
Cruse, R. W.; Farooq, A.; Karlin, K. D.; Liu, S. et al. Inorg. Chem. 1992, 31, 4322.
23. Mahapatra, S.; Halfen, J. A.; Wilkinson, E. C.; Pan, G.; Cramer, C. J.; Que, L.;
Tolman, JR. W. B. J. Am. Chem. Soc. 1995, 117, 8865.
24. López, I.; Cao, R.; Quist, D. A.; Karlin, K. D.; Le Poul, N. Chem. Eur. J. 2017, 23,
18314.
25. Gennarini, F.; David, R.; López, I.; Le Mest, Y.; Réglier, M.; Belle, C.; Thibon-
Pourret, A.; Jamet, H.; Le Poul, N. Inorg. Chem. 2017, 56, 7707.
26. Holland, P. L.; Rodgers, K. R.; Tolman, W. B. Angew. Chem. 1999, 111, 1210.
27. Becker, J.; Gupta, P.; Angersbach, F.; Tuczek, F.; Näther, C.; Holthausen, M. C.;
Schindler, S. Chem. Eur. J. 2015, 21, 11735.
28. Hamann, J. N.; Rolff, M.; Tuczek, F. Dalton trans. 2015, 44, 3251.
29. Karlin, K. D.; Cohen, B. I.; Farooq, A.; Liu, S.; Zubieta, J. Inorg. Chim. Acta 1988,
153, 9.
30. Tabuchi, K.; Ertem, M. Z.; Sugimoto, H.; Kunishita, A.; Tano, T.; Fujieda, N.;
Cramer, C. J.; Itoh, S. Inorg. Chem. 2011, 50, 1633.
31. Itoh, S.; Abe, T.; Morimoto, Y.; Sugimoto, H. Inorg. Chim. Acta 2018, 481, 38.
32. Kunishita, A.; Kubo, M.; Sugimoto, H.; Ogura, T.; Sato, K.; Takui, T.; Itoh, S. J. Am.
Chem. Soc. 2009, 131, 2788.
33. Rolff, M.; Hamann, J. N.; Tuczek, F. Angew. Chem. Int. Ed. 2011, 50, 6924.
34. Mirica, L. M.; Vance, M.; Rudd, D. J.; Hedman, B.; Hodgson, K. O.; Solomon, E. I.;
Stack, T. D. P. Science 2005, 308, 1890.
35. Bulkowski, J. E. 1985, 4545937,
36. Réglier, M.; Jorand, C.; Waegell, B. J. Chem. Soc. Chem. Commun. 1990, 107, 1752.
37. Rolff, M.; Schottenheim, J.; Peters, G.; Tuczek, F. Angew. Chem. Int. Ed. 2010, 49, 6438.
38. Hoffmann, A.; Citek, C.; Binder, S.; Goos, A.; Rübhausen, M.; Troeppner, O.;
Ivanović-Burmazović, I.; Wasinger, E. C.; Stack, T. D. P.; Herres-Pawlis, S. Angew.
Chem. Int. Ed. 2013, 52, 5398.
39. Schottenheim, J.; Fateeva, N.; Thimm, W.; Krahmer, J.; Tuczek, F.; Z. anorg. allg.
Chem. 2013, 639, 1491.
40. Hamann, J. N.; Tuczek, F. Chem. Commun. 2014, 50, 2298.
41. Schottenheim, J.; Gernert, C.; Herzigkeit, B.; Krahmer, J.; Tuczek, F.; Eur. J. Inorg.
Chem. 2015, 3501.
42. Hamann, J. N.; Schneider, R.; Tuczek, F. J. Coord. Chem. 2015, 68, 3259.
Monooxygenation of Phenols by Small-molecule Models of Tyrosinase 151
43. Wilfer, C.; Liebhäuser, P.; Erdmann, H.; Hoffmann, A.; Herres-Pawlis, S. Eur. J.
Inorg. Chem. 2015, 2015, 494.
44. Wendt, F.; Näther, C.; Tuczek, F. J. Biol. Inorg. Chem. 2016, 21, 777.
45. Herzigkeit, B.; Flöser, B. M.; Engesser, T. A.; Näther, C.; Tuczek, F. Eur. J. Inorg.
Chem. 2018, 2018, 3058.
46. Trammell, R.; See, Y. Y.; Herrmann, A. T.; Xie, N.; Díaz, D. E.; Siegler, M. A.; Baran,
P. S.; Garcia-Bosch, I. J. Org. Chem. 2017, 82, 7887.
47. Op’t Holt, B. T.; Vance, M. A.; Mirica, L. M.; Heppner, D. E.; Stack, T. D. P.;
Solomon, E. I. J. Am. Chem. Soc. 2009, 131, 6421.
48. Esguerra, K. V. N.; Fall, Y.; Lumb, J.-P. Angew. Chem. Int. Ed. 2014, 53, 5877.
49. Esguerra, K. V. N.; Fall, Y.; Petitjean, L.; Lumb, J.-P. J. Am. Chem. Soc. 2014, 136,
7662.
50. Askari, M. S.; Rodríguez-Solano, L. A.; Proppe, A.; McAllister, B.; Lumb, J.-P.;
Ottenwaelder, X. Dalton trans. 2015, 44, 12094.
51. Askari, M. S.; Esguerra, K. V. N.; Lumb, J.-P.; Ottenwaelder, X. Inorg. Chem. 2015,
54, 8665.
52. Xu, B.; Lumb, J.-P.; Arndtsen, B. A. Angew. Chem. Int. Ed. 2015, 54, 4208.
53. Huang, Z.; Kwon, O.; Esguerra, K. V. N.; Lumb, J.-P. Tetrahedron 2015, 71, 5871.
54. a) Capdevielle, P.; Maumy, M. Tetrahedron Lett. 1982, 23, 1573. b) Capdevielle, P.;
Maumy, M. Tetrahedron Lett. 1982, 23, 1577.
55. Casella, L.; Gullotti, M.; Bartosek, M.; Pallanza, G.; Laurenti, E. J. Chem. Soc. Chem.
Commun. 1991, 1235.
56. Santagostini, L.; Gullotti, M.; Monzani, E.; Casella, L.; Dillinger, R.; Tuczek, F.
Chem. Eur. J. 2000, 6, 519.
57. Battaini, G.; de Carolis, M.; Monzani, E.; Tuczek, F.; Casella, L.; Chem. Commun.
2003, 726.
58. Garcia-Bosch, I.; Cowley, R. E.; Díaz, D. E.; Peterson, R. L.; Solomon, E. I.; Karlin,
K. D. J. Am. Chem. Soc. 2017, 139, 3186.
59. Garcia-Bosch, I.; Cowley, R. E.; Díaz, D. E.; Siegler, M. A.; Nam, W.; Solomon, E.
I.; Karlin, K. D. Chem. Eur. J. 2016, 22, 5133.
60. Park, G. Y.; Qayyum, M. F.; Woertink, J.; Hodgson, K. O.; Hedman, B.; Narducci
Sarjeant, A. A.; Solomon, E. I.; Karlin, K. D. J. Am. Chem. Soc. 2012, 134, 8513.
61. Paul, M.; Teubner, M.; Grimm-Lebsanft, B.; Golchert, C.; Meiners, Y.; Senft, L.;
Keisers, K.; Liebhäuser, P.; Rösener, T.; Biebl, F. et al. Chem. Eur. J. 2020, 26, 7556.
62. Schneider, R.; Engesser, T. A.; Näther, C.; Krossing, I.; Tuczek, F. Angew. Chem. Int.
Ed. 2022, 61, e202202562.
63. Herzigkeit, B.; Flöser, B. M.; Meißner, N. E.; Engesser, T. A.; Tuczek, F.
ChemCatChem 2018, 10, 5402.
64. Canty, A. J.; Minchin, N. J. Aust. J. Chem. 1986, 39, 1063.
152 Koch et al.
65. Thies, S.; Bornholdt, C.; Köhler, F.; Sönnichsen, F. D.; Näther, C.; Tuczek, F.; Herges,
R. Chem. Eur. J. 2010, 16, 10074.
66. a) Abboud, J.-L. M.; Foces-Foces, C.; Notario, R.; Trifonov, R. E.; Volovodenko, A.
P.; Ostrovskii, V. A.; Alkorta, I.; Elguero, J. Eur. J. Org. Chem. 2001, 2001, 3013. b)
Lõkov, M.; Tshepelevitsh, S.; Heering, A.; Plieger, P. G.; Vianello, R.; Leito, I. Eur.
J. Org. Chem. 2017, 2017, 4475. c) Frey, P. A.; Hegeman, A. D. Enzymatic reaction
mechanisms, Oxford University Press, Oxford, 2007.
67. Buncel, E.; Joly, H. A.; Jones, J. R. Can. J. Chem. 1986, 64, 1240.
68. Gradert, C.; Krahmer, J.; Sönnichsen, F. D.; Näther, C.; Tuczek, F. J. Organomet.
Chem. 2014, 770, 61.
69. Solomon, E. I.; Sundaram, U. M.; Machonkin, T. E. Chem. Rev. 1996, 96, 2563.
70. Herzigkeit, B.; Jurgeleit, R.; Flöser, B. M.; Meißner, N. E.; Engesser, T. A.; Näther,
C.; Tuczek, F. Eur. J. Inorg. Chem. 2019, 2019, 2258.
71. Kwon, O.; Esguerra, K.; Glazerman, M.; Petitjean, L.; Xu, Y.; Ottenwaelder, X.;
Lumb, J.-P. Synlett 2017, 28, 1548.
72. Nilges, M. J.; Swartz, H. M.; Riley, P. A. J. Biol. Chem. 1984, 2446.
73. Bailey, S. I.; Ritchie, I. M.; Hong-Guang, Z. Bioelectrochem. Bioenerg. 1988, 19,
521.
74. Waite, J. H. Anal. Biochem. 1976, 75, 211.
75. Rouet-Mayer, M.-A.; Ralambosoa, J.; Philippon, J. Phytochemistry 1990, 29, 435.
76. Taylor, S. W.; Molinski, T. F.; Rzepecki, L. M.; Waite, J. H. J. Nat. Prod. 1991, 54,
918.
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0005
5 Electrochemistry and
Spectroelectrochemistry of
Copper-Oxygen-Relevant
Species
Nicolas Le Poul
I. Introduction
Redox reactions involving molecular oxygen (O2) are ubiquitous in chem-
istry and biology.1 For instance, the hydroxylation of dopamine into nor-
epinephrine in the catecholamine synthetic pathway is catalyzed by the
copper-based DbM glycoprotein (DbM = Dopamine b Monooxygenase).
This redox enzyme, among others such as the particulate methane
monooxygenase (pMMO)2 or peptidylglycine α-hydroxylating monooxy-
genase (PHM),3 is able to cleave the O–O bond of O2 and to perform the
incorporation of one oxygen atom into a C–H bond at a specific position.
Alternatively, the four-electron reduction of dioxygen into water is carried
153
154 Le Poul
abstraction (HAA) of strong C–H bonds of alkanes (BDFE > 90 kcal mol–1,
BDFE = Bond Dissociation Free Energy).8 Significant progress has been
recently made on the elucidation of the different mechanistic pathways,
which can occur for HAA reactions involving Cun:O2 species. In particu-
lar, the concerted or stepwise features of the coupled electron–proton
reactions have been thoroughly investigated in both thermodynamics and
kinetics terms.9 This has often, but not always, required the determination
of redox potential values of the copper–oxygen adducts by either chemical
or electrochemical means.
Independently, many copper systems have also been studied as poten-
tial electrocatalysts for the oxygen reduction reaction (ORR) in fuel
cells10,11 and oxygen evolution reaction (OER) from water oxidation.12
Most of reported studies have been carried out by using voltammetric
methods, with the aim of quantifying the electrode kinetics from electro-
chemical analysis (Tafel slope, E/pH variation). More rarely, key active
copper–oxygen species have been characterized by decoupling electron
and proton transfers in presence of organic solvents.13
As a matter of fact, electrochemical and spectroelectrochemical char-
acterizations of Cun:O2 species have been seldom reported, although they
have been well investigated by UV–vis or resonance Raman spectrosco-
pies.5,14 One main reason is that these species are often unstable in the
conditions of experimental measurement (room temperature). As it will be
shown in the following sections throughout chosen examples, several
strategies have been proposed to overcome this issue: (i) modification of
the ligand topology by using a pre-organized framework to trap O2, or by
inserting H-donors to interact with O atoms; (ii) in situ electrochemical
reduction of stable Cu(II) species, which can further react with O2, or
electrochemical oxidation of stable Cu(II) precusors such as such as
copper(II) hydroxo, alkoxo, or carboxylato cores (Figure 1) into reactive
oxidized species; and (iii) decrease of the temperature of measurement to
increase the lifetime of the copper–oxygen adduct. These three main strat-
egies have been successfully applied (sometimes combined) and have
allowed electrochemical characterization of several species. An alterna-
tive approach based on the reaction of copper–oxygen adducts with
chemical reductants/oxidants has also been developed these last years.15
156 Le Poul
Figure 2. Electron and proton–electron transfer reactions for mono and dinuclear
superoxo and peroxo copper species. Plain-line rectangles designate the species of
interest, here superoxo and peroxo Cu(II) cores. Dotted-line rectangles indicate reac-
tions that have not yet been evidenced by electrochemical methods.
species have been postulated as being strong oxidant for HAA reactions.16
Figure 2 also displays the redox chemistry starting from end-on or side-on
dinuclear peroxo species. In particular, monoelectronic reduction of side-
on dicopper(II) peroxides may lead to O–O bond cleavage and formation
of mixed-valent Cu(II)–Cu(III) bis(µ-oxo) adducts. The latter have been
suggested to be key species in the oxidation of methane in pMMOs.17
First investigations on redox properties of peroxide/superoxide
copper–oxygen species were initiated in 1992 by Karlin and co-workers
for three different side-on peroxo dicopper(II) complexes 1–3 bearing two
linked bis-pyridyl(ethylamine) moieties (Figure 3).18 These adducts were
prepared at low temperature (193 K) in dichloromethane from the reaction
of bis-Cu(I) complexes with dioxygen. UV–visible spectroscopic analyses
supported by EXAFS studies of these adducts evidenced the formation of
bent-butterfly peroxo dicopper(II) species resulting from an
Table 1. Electrochemical data for side-on peroxo (SP), end-on peroxo (EP), bis(µ-
oxo) (O), and end-on superoxo (ES) complexes 1–9 (red: reduction; ox: oxidation).
Complex E1/2 / V vs. Fc+/Fc Conditions Adduct Ref.
a,b S
1 −1.77 (red) 193 K, CH2Cl2/NBu4PF6 P [18]
a,b S
2 −1.57 (red) 193 K, CH2Cl2/NBu4PF6 P [18]
a,b S
3 −1.87 (red) 193 K, CH2Cl2/NBu4PF6 P [18]
4 0.08 (red)c 195 K, CH2Cl2/NBu4(B(C6F5)4) S
P/O [20]
c S
5 −0.03 (red) 195 K, CH2Cl2/NBu4(B(C6F5)4) P/O [20]
E
6 −0.36 (ox) 193 K, CH2Cl2/NBu4ClO4 P [21]
7 −0.75 (red)a 293 K, CH2Cl2/NBu4PF6 S
P [19]
d E
8 −0.59 (ox) 273 K, CH3CN/NBu4PF6 P [9c]
E
9 < −0.90 (red) 213 K, Acetone/NBu4PF6 S [22]
a
Irreversible peak; bRecalculated versus Fc+/Fc by taking E0(Fc+/Fc) = 0.67 V versus Ag/AgCl
in this solvent; cRecalculated versus Fc+/Fc by taking E0(Fc+/Fc) = 0.47 V versus SCE in this
solvent; dRecalculated versus Fc+/Fc by taking E0(Fc+/Fc) = 0.08 V versus Ag/AgNO3 in this
solvent.
(a) (b)
In 2017, the groups of Karlin and Le Poul on one side, as well as that
of Meyer on the other side reported the electrochemical properties of end-
on peroxo complexes 6 and 8.9c,21 For the complex 8 (Figure 3),9c cyclic
voltammetry analysis in acetonitrile at 0°C was sufficient to demonstrate
the reversibility of the peroxide/superoxide reaction at −0.59 V versus
Fc+/Fc (Table 1). This half-wave value was further utilized to calculate the
BDFE(O-H) (72 kcal mol–1) of the corresponding hydroperoxo species
knowing the pKa for the peroxo/hydroperoxo couple (22.2 in MeCN). For
the peroxo complex 6 (Figure 3), Lopez et al. investigated several aspects
of the peroxide/superoxide reaction upon oxidation of the complex at 193
K.21. A reversible monoelectronic system (Figure 5) was detected at a
slightly higher potential (−0.36 V vs. Fc+/Fc) than for complex 8 (Table 1).
Electrochemical impedance spectroscopy allowed the determination of
the electron transfer kinetics for this process. The relatively low inner-
sphere reorganizational energy (0.11 eV) was ascribed to an electron-
transfer process occurring exclusively on the O2 core. Noteworthy, in situ
UV–vis characterization of the superoxide was carried out at 193 K by a
cryo-spectroelectrochemical setup (Figure 5). Reduction of the superox-
ide allowed the back formation of the complex 6.
Electrochemistry and Spectroelectrochemistry 161
(a) (b)
Figure 5. (a) CVs of complex 6 in CH2Cl2/NBu4ClO4 for 183 < T < 203 K. (b) Low-
temperature (193 K) UV–vis in situ spectroelectrochemical monitoring of the perox-
ide oxidation (absorbance at 530 nm) leading to the formation of the superoxide
complex (absorbance at 406 nm). Adapted from Ref. [21] with permission from John
Wiley and Sons.
Figure 6. Electron and proton–electron transfer reactions for mono and dinuclear
hydroxo, alkoxo, and carbonato copper species. Plain-line rectangles designate the
species of interest, here hydroxo, alkoxo, and carbonato Cu(II) cores. The dotted-line
rectangle indicates a reaction that has not yet been evidenced by electrochemical
methods.
Table 2. Electrochemical data for hydroxo (OH), alkoxo (OA), carboxylato (OC), and
µ-oxo (Cu2O) complexes 10–31 (red: reduction; ox: oxidation).
Complex E1/2 / V vs. Fc+/Fc Conditions Adduct Ref.
H
10 0.18 (ox1); 0.47 (ox2) 293 K, DMF/NBu4PF6 O [23]
H
11 −0.17 (ox) 293 K, Acetone or THF/ O [8a,24]
NBu4PF6
12 0.12 (ox) 293 K, DFB/NBu4PF6 OH [15b]
13 0.14 (ox) 293 K, DFB/NBu4BArF4 O H
[25]
H
14 −0.13 (ox) 293 K, DFB/NBu4PF6 O [25]
15 −0.26 (ox) 293 K, DFB/NBu4PF6 OH [15b]
16 −0.20 (ox) 293 K, THF/NBu4PF6 OH [24b]
A
17 0.04 (ox) 293 K, THF/NBu4PF6 O [24b]
18 −0.02 (ox) 293 K, THF/NBu4PF6 OA [24b]
C
19 0.23 (ox) 293 K, THF/NBu4PF6 O [26]
20 0.24 (ox) 293 K, THF/NBu4PF6 OC [26]
C
21 0.17 (ox) 293 K, THF/NBu4PF6 O [26]
C
22 0.15 (ox) 293 K, THF/NBu4PF6 O [26]
23 0.30 (ox) 293 K, THF/NBu4PF6 OC [26]
C
24 0.15 (ox) 293 K, THF/NBu4PF6 O [26]
a C
25 0.63 (ox) 293 K, DMF/NBu4PF6 O [27]
26 0.12 (ox1); 0.64 (ox2)a 213 K, DMF/NBu4ClO4 OH [28]
H
27 −0.48 (red);1.87 (ox) 243 K, MeCN/NBu4ClO4 O [9d,e]
H
28 0.54 (ox) 243 K, MeCN/NBu4ClO4 O [9d]
29 −1.05 (ox1);−0.28 293 K, DMF/NBu4PF6 OH [29]
(ox2)
30 1.26 (ox) 293 K, MeCN/NBu4PF6 OH [30]
Cu2
31 0.76 (ox1); 1.35 (ox2) 243 K, MeCN/NBu4ClO4 O [9d]
a
Irreversible peak.
Electrochemistry and Spectroelectrochemistry 165
(a) (b)
Figure 8. (a) Mechanistic pathway for THF oxidation mediated by the mono-
oxidized Cu(III)-OH form of complex 11. (b) Room-temperature CV at 0.2 V s–1 of
complex 11 in THF/NBu4PF6 (black) and THF-d8/NBu4PF6 (red). Adapted with per-
mission from Ref. [8a]. Copyright (2015) American Chemical Society.
166 Le Poul
(a) (b)
(a)
(b)
Figure 10. (a) X-ray structure of complex 30. (b) Electrocatalytic oxidation of SH by
the complex 30+ in presence of 2,6-lutidine (Lut). Adapted from Ref. [31] with per-
mission of the Royal Society of Chemistry.
was found to be equal to 74.9 kcal mol–1. This value indicated that the
µ-oxo dicopper(II) complex 31 was a relatively poor oxidant for H-atom
abstraction. On the other hand, irreversible oxidation of complex 27 was
detected at high potential in dry and cooled (243 K) acetonitrile by square
wave voltammetry (SWV) (1.87 V vs. Fc+/Fc). Addition of a strong base
(in excess) allowed the generation of the unprotonated µ-oxo complex 31,
which displayed an oxidation peak at 0.76 V versus Fc+/Fc. According to
the authors, the system was reversible in the SWV timescale hence sug-
gesting the formation of the mono-oxidized oxo species, formally a CuIII-
O-CuII adduct. Further oxidation of this transient species was observed at
1.35 V versus Fc+/Fc by SWV, which was assigned to the dicopper(III)
µ(oxo) adduct. With the support of DFT calculations, and on the basis of
determined pKa values, the authors concluded that both mono and bis-
oxidized oxo adducts were strong oxidants for HAA (103.4 and 91.7 kcal
mol–1, respectively). Noteworthy, the better oxidative properties for the
mono-oxidized were ascribed to the linear feature of the Cu-O-Cu core,
which favored strong superexchange between Cu centers through the cen-
tral oxygen and was merely considered as a dicopper(II)-oxyl adduct.
Alternatively, Garcia-Bosch and co-workers carried out room-
temperature voltammetric studies of the mononuclear copper(II)-hydroxide
complex 29 bearing a redox-active ligand comprising a tridentateamine and
two ureanyl H-bonds donors (Figure 7).29 Two reversible monoelectronic
oxidation waves were detected at –1.05 V and −0.28 V versus Fc+/Fc
(Table 2). Noteworthy, UV–vis, EPR, and XAS spectroscopic methods
using chemical oxidants evidenced that the electron transfer processes were
essentially located on the ligand backbone and not on the metal. While the
generated semi-quinone Cu(II) hydroxide species was found to be a moder-
ate H-abstractor, the bis-oxidized quinone Cu(II) hydroxo complex exhib-
ited a significant HAA 2e−/2H+ activity according to two sequential steps.
Figure 11. Electron and proton–electron transfer reactions for mono and dinuclear
hydroperoxo or alkoxoperoxo copper species. Plain-line rectangles designate the
species of interest, here hydroperoxo and alkylperoxo Cu(II) cores. Dotted-line rec-
tangles indicate reactions that have not yet been evidenced by electrochemical
methods.
(a) (b)
Figure 14. Electron and proton–electron transfer reactions for mono and dinuclear
copper phenoxide species. Plain-line rectangles designate the species of interest,
here phenolato Cu(II) cores.
Table 4. Electrochemical data for phenoxo (OPh) copper(II) complexes 35–55 (red:
reduction; ox: oxidation).
Complex E1/2 / V vs. Fc+/Fc Conditions Adduct Ref.
a,b Ph
35 0.24 (ox) 293 K, CH2Cl2/NBu4PF6 O [34a]
b Ph
36 0.12 (ox) 293 K, CH2Cl2/NBu4PF6 O [34a]
b Ph
37 0.25 (ox) 293 K, CH2Cl2/NBu4PF6 O [34a]
b Ph
38 0.32 (ox) 293 K, CH2Cl2/NBu4PF6 O [34b]
39 0.03 (ox1)b; 0.31 (ox2)b 293 K, CH2Cl2/NBu4PF6 OPh [34b]
c c Ph
40 0.41 (ox1) ; 0.58 (ox2) 233 K, CH3CN/NBu4ClO4 O [35]
Ph
41 0.55 (ox) 233 K, CH2Cl2/NBu4PF6 O [36]
Ph
42 0.28 (ox) 233 K, CH2Cl2/NBu4PF6 O [36]
Ph
43 0.46 (ox) 293 K, CH2Cl2/NBu4ClO4 O [37]
Ph
44 0.45 (ox1); 0.65 (ox2) 293 K, CH2Cl2/NBu4ClO4 O [38]
Ph
45 0.15 (ox1); 0.54 (ox2) 293 K, CH2Cl2/NBu4ClO4 O [38a]
Ph
46 0.08 (ox1); 0.21 (ox2) 293 K, CH2Cl2/NBu4ClO4 O [38]
Ph
47 0.44 (ox1); 0.76 (ox2) 293 K, CH2Cl2/NBu4ClO4 O [39]
Ph
48 0.22 (ox1); 0.50 (ox2) 293 K, CH2Cl2/NBu4ClO4 O [39]
Ph
49 0.56 (ox1); 0.75 (ox2) 293 K, CH2Cl2/NBu4ClO4 O [39]
50 0.34 (ox1); 0.53 (ox2) 293 K, CH2Cl2/NBu4ClO4 OPh [39]
Ph
51 0.51 (ox) 293 K, CH3CN/NBu4PF6 O [33d,40]
a Ph
52 0.71 (ox) 293 K, CH3CN/NBu4PF6 O [40]
a Ph
53 1.20 (ox) 293 K, CH3CN/NBu4PF6 O [40]
54 1.04 (ox)a 293 K, CH3CN/NBu4PF6 OPh [40]
a Ph
55 0.90 (ox) 293 K, CH3CN/NBu4PF6 O [40]
a b + 0 +
Irreversible peak; Recalculated versus Fc /Fc by taking E (Fc /Fc) = 0.47 V versus SCE in this
solvent; cRecalculated versus Fc+/Fc by taking E0(Fc+/Fc) = 0.40 V versus SCE in this solvent.
(a) (b)
Figure 16. (a) Room-temperature CVs at 0.1 V s−1 of complex 44 (black), 45 (red),
and 46 (blue) in CH2Cl2/NBu4ClO4 (R = tBu). (b) Schematic representation of the
mono-oxidized species 44+, 45+, and 46+. Adapted with permission from Refs. [38a]
and [38b]. Copyright (2018) Elsevier. Copyright (2008) American Chemical Society.
Figure 18. Electron transfer reactions for mono and dinuclear nitrosoarene and (bi)
sulfido copper species. Plain-line rectangles designate the species of interest, here
nitrosoarene, sufido, and bisulfido Cu(II) cores. The dotted-line rectangle indicates a
reaction that has not yet been evidenced by electrochemical methods.
between −0.36 and 0.28 V versus Fc+/Fc, depending on the nature of sub-
stituting group. These complexes were shown to be capable to perform
H-atom abstraction on C–H bonds of dihydroanthracene (BDFE = 76 kcal
mol–1) but at low reaction rates.
Another electrochemical study of interest was carried out by Tolman
and co-workers in 2008 on side-on disulfido dicopper complexes,43 which
are structural models of side-on peroxide copper adducts. Voltammetry at
room temperature in dichloromethane displayed an irreversible reduction
peak at −0.30 V and −0.77 V versus Fc+/Fc for complexes 60 and 61,
respectively (Figure 19 and Table 5). The difference of potential was
ascribed to the donor ability of the neutral ligand (peralkylated bi- or tri-
amine). In contrast to complex 61, complex 60 was shown to be a two-
electron oxidant for substituted phenolates. More recently, the same group
conducted researches on sulfur-containing analogues of monocopper-
hydroxo species.44 Two Cu(II) complexes bearing a macrocyclic ligand
and a hydrosulfido and thiophenolato anion were synthesized (Figure 19,
complexes 62 and 63, respectively). Room-temperature voltammetric
studies of these complexes in THF exhibited a quasi-reversible 1e– system
in oxidation at −0.21 V and −0.25 V versus Fc+/Fc (Table 5). The slightly
more negative E1/2 value for the complex 63 versus 62 was consistent with
the better donor properties of the phenyl moiety compared to the hydrogen
atom. Analogously, complex 62 showed a significantly lower oxidation
Electrochemistry and Spectroelectrochemistry 181
potential value (−50 mV) than that obtained for its hydroxo counterpart
(complex 11), as a result of the higher electronegativity of S versus O.
Reactivity of complexes 62 and 63 with TEMPOH was much slower than
for the OH analogue (complex 11) in agreement with the electrochemical
data and the lower basicity of the SH- and SPh- ligands relative the
hydroxide.
VII. Summary
In conclusion, this chapter has reviewed the electrochemical and spectro-
electrochemical studies of relevant copper species, including unstable
copper oxygen adducts and their precursors. While the reported examples
remain still scarce, one can observe that the number of electrochemical
studies of these species has dramatically increased for the last five years.
One main reason is that electrochemical data (thermodynamics) can be
directly correlated to HAA reactivity, hence allowing the calculation of
BDFEs through the well-known Bordwell analysis knowing pKa values.
Moreover, under certain circumstances, kinetics of electron transfer have
been directly quantified from electrochemical measurements (CV or
EIS), giving the possibility to determine inner and outer reorganizational
energies with the help of heterogeneous electron transfer theories.
Another source of progress has been achieved with in situ UV–visible
cryo-spectroelectrochemical measurements that have afforded direct
characterization of transient species generated at the electrode surface at
low temperature21,28,45 hence avoiding exhaustive electrolysis and uncer-
tainty about the real nature of the generated transient species.
From all data given in this chapter, it seems difficult to provide defini-
tive conclusions and/or to predict any redox behavior for any or other
specific family of copper–oxygen adducts. The principal reason is that
subtle variation of the geometric or electronic properties of the ligand
itself can induce significant modification of the redox properties, such as
the half-wave potential value, or even move the location of the electron
transfer. This is best exemplified by the different UV–vis spectroelectro-
chemical responses that were obtained upon oxidation of the phenoxo-
based complexes 52 and 55, which only differ by the spacer chain length
on one side of the dinucleating ligand.
182 Le Poul
VIII. References
1. (a) Solomon, E. I.; Heppner, D. E.; Johnston, E. M.; Ginsbach, J. W.; Cirera, J.;
Qayyum, M.; Kieber-Emmons, M. T.; Kjaergaard, C. H.; Hadt, R. G.; Tian, L. Chem.
Rev. 2014, 114, 3659–3853. (b) Quist, D. A.; Diaz, D. E.; Liu, J. J.; Karlin, K. D.
J. Biol. Inorg. Chem. 2017, 22, 253–288. (c) Lee, J. Y.; Karlin, K. D. Curr. Opin.
Chem. Biol. 2015, 25, 184–193.
2. Meier, K. K.; Jones, S. M.; Kaper, T.; Hansson, H.; Koetsier, M. J.; Karkehabadi, S.;
Solomon, E. I.; Sandgren, M.; Kelemen, B. Chem. Rev. 2018, 118, 2593–2635.
3. Cowley, R. E.; Tian, L.; Solomon, E. I. Proc. Natl. Acad. Sci. U.S.A. 2016, 113,
12035–12040.
4. Solomon, E. I. Inorg. Chem. 2016, 55, 6364–6375.
5. Elwell, C. E.; Gagnon, N. L.; Neisen, B. D.; Dhar, D.; Spaeth, A. D.; Yee, G. M.;
Tolman, W. B. Chem. Rev. 2017, 117, 2059–2107.
6. (a) Lyu, Z.; Zhou, Y.; Dai, W.; Cui, X.; Lai, M.; Wang, L.; Huo, F.; Huang, W.; Hu,
Z.; Chen, W. Chem. Soc. Rev. 2017. (b) Kwabi, D. G.; Bryantsev, V. S.; Batcho, T. P.;
Itkis, D. M.; Thompson, C. V.; Shao-Horn, Y. Angew. Chem. Int. Ed. 2016, 55,
3129–3134. (c) Bruce, P. G.; Freunberger, S. A.; Hardwick, L. J.; Tarascon, J. M. Nat.
Mater. 2011, 11, 19–29.
7. (a) Poulos, T. L. Chem. Rev. 2014, 114, 3919–3962. (b) Sheng, Y.; Abreu, I. A.;
Cabelli, D. E.; Maroney, M. J.; Miller, A. F.; Teixeira, M.; Valentine, J. S. Chem. Rev.
2014, 114, 3854–3918.
8. (a) Dhar, D.; Tolman, W. B. J. Am. Chem. Soc. 2015, 137, 1322–1329. (b) Warren, J.
J.; Tronic, T. A.; Mayer, J. M. Chem. Rev. 2010, 110, 6961–7001.
9. (a) Mandal, M.; Elwell, C. E.; Bouchey, C. J.; Zerk, T. J.; Tolman, W. B.; Cramer, C.
J. J. Am. Chem. Soc. 2019, 141, 17236–17244. (b) Bailey, W. D.; Dhar, D.; Cramblitt,
Electrochemistry and Spectroelectrochemistry 183
A. C.; Tolman, W. B. J. Am. Chem. Soc. 2019, 141, 5470–5480. (c) Kindermann, N.;
Gunes, C. J.; Dechert, S.; Meyer, F. J. Am. Chem. Soc. 2017, 139, 9831–9834. (d)
VanNatta, P. E.; Ramirez, D. A.; Velarde, A. R.; Ali, G.; Kieber-Emmons, M. T. J. Am.
Chem. Soc. 2020, 142, 16292–16312. (e) Ali, G.; VanNatta, P. E.; Ramirez, D. A.;
Light, K. M.; Kieber-Emmons, M. T. J. Am. Chem. Soc. 2017, 139, 18448–18451.
10. Mano, N.; de Poulpiquet, A. Chem. Rev. 2018, 118, 2392–2468.
11. (a) Thorseth, M. A.; Tornow, C. E.; Tse, E. C. M.; Gewirth, A. A. Coord. Chem. Rev.
2013, 257, 130–139. (b) Langerman, M.; Hetterscheid, D. G. H. Angew. Chem. Int.
Ed. 2019, 58, 12974–12978. (c) Gentil, S.; Serre, D.; Philouze, C.; Holzinger, M.;
Thomas, F.; Le Goff, A. Angew. Chem. Int. Ed. Engl. 2016, 55, 2517–2520.
(d) Thorseth, M. A.; Letko, C. S.; Rauchfuss, T. B.; Gewirth, A. A. Inorg. Chem.
2011, 50, 6158–6162. (e) Kato, M.; Yagi, I. e-J. Surf. Sci. Nanotech. 2020, 18, 81–93.
12. (a) Lukács, D.; Szyrwiel, Ł.; Pap, J. Catalysts 2019, 9, 83. (b) Koepke, S. J.; Light,
K. M.; VanNatta, P. E.; Wiley, K. M.; Kieber-Emmons, M. T. J. Am. Chem. Soc. 2017,
139, 8586–8600. (c) Kafentzi, M. C.; Papadakis, R.; Gennarini, F.; Kochem, A.;
Iranzo, O.; Le Mest, Y.; Le Poul, N.; Tron, T.; Faure, B.; Simaan, A. J.; Reglier, M.
Chem. Eur. J. 2018, 24, 5213–5224. (d) Barnett, S. M.; Goldberg, K. I.; Mayer, J. M.
Nat. Chem. 2012, 4, 498–502. (e) Su, X. J.; Gao, M.; Jiao, L.; Liao, R. Z.; Siegbahn,
P. E.; Cheng, J. P.; Zhang, M. T. Angew. Chem. Int. Ed. 2015, 54, 4909–4914.
(f) Nestke, S.; Ronge, E.; Siewert, I. Dalton Trans. 2018, 47, 10737–10741.
13. Tahsini, L.; Kotani, H.; Lee, Y.-M.; Cho, J.; Nam, W.; Karlin, K. D.; Fukuzumi, S.
Chem. Eur. J. 2012, 18, 1084–1093.
14. Mirica, L. M.; Ottenwaelder, X.; Stack, T. D. P. Chem. Rev. 2004, 104, 1013–1046.
15. (a) Cao, R.; Saracini, C.; Ginsbach, J. W.; Kieber-Emmons, M. T.; Siegler, M. A.;
Solomon, E. I.; Fukuzumi, S.; Karlin, K. D. J. Am. Chem. Soc. 2016, 138, 7055–7066.
(b) Dhar, D.; Yee, G. M.; Spaeth, A. D.; Boyce, D. W.; Zhang, H.; Dereli, B. Cramer,
C. J.; Tolman, W. B. J. Am. Chem. Soc. 2016, 138, 356–368.
16. Gagnon, N.; Tolman, W. B. Acc. Chem. Res. 2015, 48, 2126–2131.
17. Shiota, Y.; Yoshizawa, K. Inorg. Chem. 2009, 48, 838–845.
18. Karlin, K. D.; Tyeklar, Z.; Farooq, A.; Haka, M. S.; Ghosh, P.; Cruse, R. W.; Gultneh,
Y.; Hayes, J. C.; Toscano, P. J.; Zubieta, J. Inorg. Chem. 1992, 31, 1436–1451.
19. Lopez, I.; Porras-Gutierrez, A. G.; Douziech, B.; Wojcik, L.; Le Mest, Y.; Kodera, M.;
Le Poul, N. Chem. Commun. 2018, 54, 4931–4934.
20. Shearer, J.; Zhang, C. X.; Zakharov, L. N.; Rheingold, A. L.; Karlin, K. D. J. Am.
Chem. Soc. 2005, 127, 5469–5483.
21. Lopez, I.; Cao, R.; Quist, D. A.; Karlin, K. D.; Le Poul, N. Chem. Eur. J. 2017, 23,
18314–18319.
22. De Leener, G.; Over, D.; Smet, C.; Cornut, D.; Porras-Gutierrez, A. G.; Lopez, I.;
Douziech, B.; Le Poul, N.; Topic, F.; Rissanen, K.; Le Mest, Y.; Jabin, I.; Reinaud, O.
Inorg. Chem. 2017, 56, 10971–10983.
184 Le Poul
23. Halvagar, M. R.; Solntsev, P. V.; Lim, H.; Hedman, B.; Hodgson, K. O.; Solomon, E.
I.; Cramer, C. J.; Tolman, W. B. J. Am. Chem. Soc. 2014, 136, 7269–7272.
24. (a) Donoghue, P. J.; Tehranchi, J.; Cramer, C. J.; Sarangi, R.; Solomon, E. I.; Tolman,
W. B. J. Am. Chem. Soc. 2011, 133, 17602–17605. (b) Krishnan, V. M.; Shopov, D.
Y.; Bouchey, C. J.; Bailey, W. D.; Parveen, R.; Vlaisavljevich, B.; Tolman, W. B.
J. Am. Chem. Soc. 2021, 143, 3295–3299.
25. Dhar, D.; Yee, G. M.; Tolman, W. B. Inorg. Chem. 2018, 57, 9794–9806.
26. Elwell, C. E.; Mandal, M.; Bouchey, C. J.; Que, L.; Cramer, Jr. C. J.; Tolman, W. B.
Inorg. Chem. 2019, 58, 15872–15879.
27. Kochem, A.; Gennarini, F.; Yemloul, M.; Orio, M.; Le Poul, N.; Riviere, E.; Giorgi,
M.; Faure, B.; Le Mest, Y.; Reglier, M.; Simaan, A. J. Chem. Plus. Chem. 2017, 82,
615–624.
28. Thibon-Pourret, A.; Gennarini, F.; David, R.; Isaac, J. A.; Lopez, I.; Gellon, G.;
Molton, F.; Wojcik, L.; Philouze, C.; Flot, D.; Le Mest, Y.; Reglier, M.; Le Poul, N.;
Jamet, H.; Belle, C. Inorg. Chem. 2018, 57, 12364–12375.
29. Wu, T.; MacMillan, S. N.; Rajabimoghadam, K.; Siegler, M. A.; Lancaster, K. M.;
Garcia-Bosch, I. J. Am. Chem. Soc. 2020, 142, 12265–12276.
30. Isaac, J. A.; Gennarini, F.; Lopez, I.; Thibon-Pourret, A.; David, R.; Gellon, G.;
Gennaro, B.; Philouze, C.; Meyer, F.; Demeshko, S.; Le Mest, Y.; Reglier, M.; Jamet,
H.; Le Poul, N.; Belle, C. Inorg. Chem. 2016, 55, 8263–8266.
31. Isaac, J. A.; Thibon-Pourret, A.; Durand, A.; Philouze, C.; Le Poul, N.; Belle, C.
Chem. Commun. 2019, 55, 12711–12714.
32. Neisen, B. D.; Gagnon, N. L.; Dhar, D.; Spaeth, A. D.; Tolman, W. B. J. Am. Chem.
Soc. 2017, 139, 10220–10223.
33. (a) Thomas, F. Eur. J. Inorg. Chem. 2007, 2007, 2379–2404. (b) Benisvy, L.; Blake,
A. J.; Davies, E. S.; Garner, C. D.; McMaster, J.; Wilson, C.; Collison, D.; McInnes,
E. J. L.; Whittaker, G. Chem. Commun. 2001, 1824–1825. (c) Uma Maheswari, P.;
Hartl, F.; Quesada, M.; Buda, F.; Lutz, M.; Spek, A. L.; Gamez, P.; Reedijk, J. Inorg.
Chim. Acta 2011, 374, 406–414. (d) Michel, F.; Torelli, S.; Thomas, F.; Duboc, C.;
Philouze, C.; Belle, C.; Hamman, S.; Saint-Aman, E.; Pierre, J. L. Angew. Chem. Int.
Ed. 2005, 44, 438–441.
34. (a) Halfen, J. A.; Young, V. G.; Tolman, W. B. Angew. Chem. Int. Ed. 1996, 35,
1687–1690. (b) Halfen, J. A.; Jazdzewski, B. A.; Mahapatra, S.; Berreau, L. M.;
Wilkinson, E. C.; Que, L.; Tolman, W. B. J. Am. Chem. Soc. 1997, 119, 8217–8227.
35. Wang, Y.; Stack, T. D. P. J. Am. Chem. Soc. 1996, 118, 13097–13098.
36. Balaghi, S. E.; Safaei, E.; Chiang, L.; Wong, E. W.; Savard, D.; Clarke, R. M.; Storr,
T. Dalton Trans. 2013, 42, 6829–6839.
37. Alaji, Z.; Safaei, E.; Chiang, L.; Clarke, R. M.; Mu, C.; Storr, T. Eur. J. Inorg. Chem.
2014, 2014, 6066–6074.
Electrochemistry and Spectroelectrochemistry 185
38. (a) Chiang, L.; Wasinger, E. C.; Shimazaki, Y.; Young, V.; Storr, T.; Stack, T. D. P.
Inorg. Chim. Acta 2018, 481, 151–158. (b) Storr, T.; Verma, P.; Pratt, R. C.; Wasinger,
E. C.; Shimazaki, Y.; Stack, T. D. J. Am. Chem. Soc. 2008, 130, 15448–15459.
39. Kunert, R.; Philouze, C.; Berthiol, F.; Jarjayes, O.; Storr, T.; Thomas, F. Dalton Trans.
2020, 49, 12990–13002.
40. Gennarini, F.; David, R.; Lopez, I.; Le Mest, Y.; Reglier, M.; Belle, C.; Thibon-
Pourret, A.; Jamet, H.; Le Poul, N. Inorg. Chem. 2017, 56, 7707–7719.
41. Pratt, R. C.; Stack, T. D. J. Am. Chem. Soc. 2003, 125, 8716–8717.
42. Askari, M. S.; Effaty, F.; Gennarini, F.; Orio, M.; Le Poul, N.; Ottenwaelder, X. Inorg.
Chem. 2020, 59, 8678–8689.
43. Bar-Nahum, I.; York, J. T.; Young, V. G.; Tolman, Jr. W. B. Angew. Chem. Int. Ed.
2008, 47, 533–536.
44. Wu, W.; De Hont, J. T.; Parveen, R.; Vlaisavljevich, B.; Tolman, W. B. Inorg. Chem.
2021, 60, 5217–5223.
45. López, I.; Le Poul, N. Coord. Chem. Rev. 2021, 436, 213823.
This page intentionally left blank
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0006
6 Inorganic Models of
Lytic Polysaccharide
Monooxygenases
Ivan Castillo*
I. Introduction
Global sources of energy and chemicals have depended heavily on petro-
leum-based platforms, which have been affordable thus far, at least from
an economic standpoint. It is evident, however, that sustainable practices
need to be prioritized to avoid serious environmental consequences,
beyond the ones we experience nowadays.1 To this end, efficient use of
renewable carbon feedstocks is paramount for the sustainable production
of fuels and chemical supplies.2,3
The aim of this chapter is to present a brief background of the
discovery of copper-dependent lytic polysaccharide monooxygenases
(LPMOs),4–6 within the context of inorganic and coordination chemistry.
Their discovery and characterization have led to the design of synthetic
systems that possess some of the properties of the enzymatic active sites.
Copper complexes thus developed will be discussed in terms of their capa-
bility to act as LPMO mimics, and to provide insight on the mechanism
187
188 Castillo
B. Copper Complexes
Studies regarding copper complexes as hydrolase mimics involve the
commonly employed p-nitrophenylpyranosides model substrates, with
cleavage of the p-nitrophenolate moiety conveniently monitored by
A. Copper(II)/Bis(Benzimidazolyl)Amine System
Structural analogs of the active site of LPMO enzymes were reported as
early as 2014, with the tridentate bis(benzimidazolyl)amine ligand 2BB
in Figure 5 providing the T-shaped N3 set that includes the amine as cen-
tral donor.22 An important feature of the ligands employed for LPMO
mimics is the size of the chelate defined by the cis-N2 moieties: the two
Inorganic Models of Lytic Polysaccharide Monooxygenases 193
Figure 5. Left: 2BB; center: schematic representation of the T-shape active site of
AA9-11 LPMOs; right: Mercury diagram of [(2BB)Cu(H2O)2](OTf)2 at the 50% prob-
ability level; H atoms and triflate counterions omitted for clarity. Color code: C, gray;
N, blue; O, red; Cu, turquoise.
B. Copper(III)/Bis(Carboxamido)Pyridine System
Tolman and coworkers have exploited the dianionic N3 scaffold provided
by the ligand in Figure 8, featuring a central pyridine flanked by two car-
boxamido donors and sterically-encumbering N-aryl substituents, which
enforces meridional coordination.34 Its potential to mimic the active site
of LPMO was identified in 2015 in the form of a CuIII–OH complex (1)
as oxidant,35 which is capable of oxidizing strong C–H bonds by hydro-
gen atom abstraction (HAT). Although the dianionic ligand differs signifi-
cantly from the neutral donor set provided by the enzymes, its relevance
arises from the potential involvement of the invoked deprotonated primary
Inorganic Models of Lytic Polysaccharide Monooxygenases 197
(a) (b)
(c)
Figure 8. Left: complex 1−, X = OH. Right: oxidative glycosidic cleavage catalyzed
by LPMO and proposed structures for active site oxidants (a–c), including a CuIII–OH
complex (c), featuring a deprotonated amine.
C. Copper(II)/Copper(I)/(Pyridyl,Imidazolyl)Amine System
The complexes reported by Concia and coworkers feature closely related
ligands, where the central nitrogen donor takes the shape of an amine
(LAM) or an imine (LIM).38 In both cases, a 2-pyridyl fragment is con-
nected to the central N-atom by an ethylene linker that results in six-
membered chelates upon binding of copper. On the other hand, the
2-imidazolyl moiety is connected by a single carbon atom bridge to the
central nitrogen, resulting in five-membered chelates in both LAM and
LIM (Figure 10).
Structural characterization of the corresponding copper complexes
resulted in pseudo-octahedral geometries, with both chelating ligands act-
ing as equatorial N3-donors complemented by a molecule of water, and
the perchlorate anions as weak axial O-donors. The Cu–N distances range
from 1.97 to 2.03 Å, also within the range of reported LPMO active
sites. Aqueous EPR spectra are consistent with axial geometry in solution
and a dx2–y2 ground state, g⊥ = 2.059, g|| = 2.260 (A|| = 530 MHz) for
[LAMCu(H2O)2](ClO4)2 (1), and g⊥ = 2.060, g|| = 2.265 (A|| = 530 MHz)
for [LIMCu(H2O)2](ClO4)2 (2). All parameters are comparable to those
reported for LPMOs.4,9–11,29 An important physicochemical property that
appears to be directly related to the reactivity of LPMOs for these types
of systems is the redox potential, which was determined for both com-
plexes 1 and 2 by CV and redox titrations. The values determined range
from 5 to 50 mV versus SHE, and contrast with those reported for the
metalloenzymes.4,10,11,29
Complexes 1 and 2 represent functional models of LPMO on the
model substrate p-nitrophenyl-β-D-glucopyranoside that features the
p-nitrophenolate anion as leaving group. Thus, incubation of the com-
plexes with hydrogen peroxide as oxidant at pH 10.5 in carbonate buffer
resulted in p-nitrophenolate cleavage from the model substrate. The oxi-
dative nature of the reaction was confirmed by the low conversion in the
absence of H2O2, and by identification of gluconic acid as the product of
the initially formed gluconolactone by ESI-MS, see Figure 11. For both 1
and 2, analysis of the reaction with H2O2 in the presence of Et3N as base
in aqueous solution by UV–vis spectroscopy resulted in the detection of
an intense band at 305 nm, with a shoulder around 375 nm (Figure 12).
The intermediates were postulated as CuII–OOH species based on previ-
ous reports,39 the d–d band at 650 nm, and X-band EPR spectra that
resemble those of the parent monometallic complexes.
D. Copper(II)/(Pyridyl)Diazepane Complexes
Mayilmurugan and coworkers devised the pyridyl-appended ligands L1
through L3 from the 1,4-diazepane backbone (1-methyl homopiperazine)
for mimicking of the N3 coordination environment in LPMOs.40 While the
diazepane moiety provides a rigid framework for CuII binding, there are
differences among the ligands regarding the nature of the chelates formed
with the pyridyl substituents. L1 features a methylene linker between the
unsubstituted 2-pyridyl group and the diazepane N-atom, L2 has an
Figure 12. Room temperature absorption spectra of 0.25 mM aqueous buffer solu-
tion of [LAMCu(H2O)2](ClO4)2 (dotted line) and after addition of 25 mM H2O2 (solid
lines).
cathodic peaks DE from 145 to 163 mV. The redox potentials reported
have increasing values at E1/2 = 8, 41, and 112 mV relative to the normal
hydrogen electrode (NHE). However, this is inconsistent with the values
reported versus Ag/AgCl as reference, which correspond to the descend-
ing values −213, −246, and −317 mV for 1–3, respectively. Thus, further
analysis based on the structure of the complexes and ligands is not pos-
sible, as is a comparison of the values reported here with those of other
model systems and LPMOs.
With these systems, p-nitrophenyl-β-D-glucopyranoside was also
used as a model substrate, with 30% aqueous hydrogen peroxide and Et3N
as oxidant mixture. Complex 3 featuring electron-donating methyl and
methoxy groups on the pyridine donor exhibited the highest rate constant
for p-nitrophenolate cleavage, as evidenced by UV–vis spectroscopy. The
authors attribute this behavior to stabilization of the putative CuII–OOH
intermediate, but this is counterintuitive as it would result in a less elec-
trophilic species. An oxidative process is assumed to be operative,
although in preparative scale reactions the co-product was identified as
D-allose, see Figure 14. This result contrasts with the observation of glu-
conic acid as an oxidation product,25,38 whereas D-allose may be expected
in the case of hydrolysis. The reasonable doubt regarding the proposed
oxidative nature of the cleavage reaction is further supported by the lack
of activity of the most active catalyst 3 in the attempted oxidations of
benzyl alcohol and 1-phenylethanol.
Attempts to identify the cupric hydroperoxo intermediate with
10 equivs. of H2O2 in the presence of Et3N in aqueous solution by
UV–vis spectroscopy resulted in the detection of LMCT bands around
375 nm, accompanied by d–d transitions at ca. 630 nm; unfortunately, the
extinction coefficient for the LMCT bands was not reported. Addition of
tert-butyl hydroperoxide to solutions of 3 resulted in a band assigned as
LMCT at 380 nm, although the extinction coefficient reported is low for
such assignment (ε = 182 M−1cm−1). In the latter experiment, the authors
Inorganic Models of Lytic Polysaccharide Monooxygenases 203
Figure 14. Proposed reaction scheme for the cleavage of p-nitrophenol from
p-nitrophenyl-b-D-glucopyranoside with complexes 1–3 as catalysts, with D-allose
as co-product.
E. Copper(II)/Bis(Imidazolyl)Amine System
Modeling of the histidine brace of LPMO by the group of Itoh involved a
tridentate N3 donor that incorporates two imidazole groups and a central
amine.41 In addition to the T-shaped geometry, the authors deemed neces-
sary to have a large “twist” angle between the two imidazole planes, as is
the case in the active site of the enzymes, and acidic N–H protons at
the same N-heterocycles. These considerations notwithstanding, t-N-
methylation of imidazole moieties in N-terminal histidines of fungal
LPMOs is a common occurrence.5,42 The ligand (N-(2-(1H-imidazol-4-
yl)-benzyl)histamine) (LH3) and its cupric complex [Cu(LH3)(TFA)2]
(TFA = trifluoroacetate) were thus prepared. The authors mention that
under neutral or basic conditions, insoluble blue crystals were obtained
when Cu(ClO4)2 was added to solutions of the ligand, which they attribute
to deprotonation of the N–H group of the imidazolyl moieties and subse-
quent formation of coordination polymers, but potentiometric measure-
ments were not undertaken to determine their pKa values.
Structural characterization of [Cu(LH3)(TFA)2] revealed a SP geom-
etry (t5 = 0.06),43 with the basal plane defined by the N3-donor set and a
monodentate TFA counterion; the axial position is occupied by the second
204 Castillo
(a) (b)
Figure 15. Comparison of (a) bond lengths (Å) and (b) bond angles (°) of [Cu(LH3)
(TFA)2] (top) with those of oxidized LPMOs (bottom).
potential of 76 mV, while the actual reported value was 226 mV versus
SHE despite the angle being 14.5° in the cupric complex.25
Catalytic tests of [Cu(LH3)(TFA)2] with hydrogen peroxide (20 mM)
as an oxidant in carbonate buffer at pH 10 were also carried out with
p-nitrophenyl-β-D-glucopyranoside as substrate. As is the case with
[Cu(Ln)(H2O)ClO4](ClO4) (Ln = L1–L3, complexes 1–3) in Section IV.D,
an oxidative cleavage was reported, but the products observed correspond
to p-nitrophenolate and D-allose. Once again, the identification of
D-allose would correspond to hydrolysis rather than oxidation, raising
questions about the mechanism of model substrate cleavage.
F. Copper(II)/Bis(Picolyl)Amine System
The most recent example of a simple model system that mimics some
aspects of LPMO activity is the one reported by Cowan and coworkers.44
Thus, bis(picolyl)amine copper(II) chloride (1a) was employed along
with di- and tetranuclear analogs that will not be discussed, because the
enzymatic active sites are mononuclear. The preparation of 1a was previ-
ously reported,45 and its solid-state structure established that the Cu(II)
center has a distorted SP coordination environment, with all N-donors in
basal positions.46 Its redox potential was reported relative to the reversible
hydrogen electrode (RHE) at a value of E1/2 = 230 mV,47 within the values
reported for LPMOs.
The report focuses on glycosidase activity by oxidative cleavage of
p-nitrophenol conjugated glucose, galactose, and related disaccharides as
substrates in the presence of ascorbate and H2O2 under physiological con-
ditions.44 Similar to the earlier examples, the cleavage of p-nitrophenolate
was monitored by UV–vis spectroscopy, starting with ascorbate and
dioxygen, which resulted in ~10% activity when compared to hydrogen
peroxide as oxidant. The authors indicate that Fenton-like chemistry
might be responsible for the observed reactivity, with hydroxy radicals
promoting H-abstraction from the saccharides in the initial step of
p-nitrophenolate cleavage. Unfortunately, the identity of the saccharide
fragments after degradation was not established, leaving the open question
of whether the observed glycoside cleavage is indeed oxidative in nature.
206 Castillo
V. Concluding Remarks
The meridional donor set provided by the N3 scaffold in the form of the
histidine brace and an additional histidine imidazole in LPMOs is a req-
uisite that needs to be fulfilled to access functional model systems. The
effect of the angle between the imidazole rings remains to be understood
although it does not appear to be crucial to access the range of redox
potentials of the CuII/CuI couple in the enzymes. Moreover, the half wave
potentials may be a first approximation to the electronic environment
around the copper centers in the active sites of LPMOs, but this physico-
chemical parameter does not appear to be a strict requisite to develop
mimics, at least when model substrates are considered. Another observa-
tion that emerges is that hydrogen peroxide affords far superior results in
the degradation of model substrates and cellulose, which is in line with
recent studies that point to H2O2 as the oxidant in the enzymatic systems,
rather than O2. It may be time to consider renaming the enzymes as Lytic
Polysaccharide Peroxygenases.
While the minimum requirements appear to have been laid out for the
development of oxidative glycosidase model systems, key points remain
to be addressed: can model systems actually degrade the natural substrates
cellulose and/or chitin? This is as challenging to explore with model sys-
tems due to the insoluble and recalcitrant nature of such polysaccharides,
as it is in nature. Thus far, only the bis(benzimidazole)amine-based copper
complexes have demonstrated oxidative cleavage activity on cellulose.32
Another question that remains to be answered is whether the true oxidant
in the enzymatic systems is dioxygen or hydrogen peroxide, and simple
model complexes may provide relevant information. In this regard, LPMO
mimics may be active at oxidatively degrading model and natural sub-
strates, but they may operate through different reaction mechanisms. In
some cases, even the oxidative nature of the model substrate degradation
needs to be firmly established. Finally, perhaps the most fundamental
question that remains open to speculation is the nature of the copper–
oxygen intermediate responsible for the activation of the strong C–H bond
of insoluble substrates in LPMOs. Initial proposals considered cupric-
superoxide [CuII–O2]+ species, but recent experimental and theoretical
studies tend to favor a copper-oxyl [CuII–O]+ intermediate for H-abstraction
Inorganic Models of Lytic Polysaccharide Monooxygenases 207
from strong C–H bonds. Model systems have added to the discussion the
potential involvement of a high-valent copper(III)-hydroxide [CuIII–
OH]2+. Fenton-type reactivity may also be invoked, whether it occurs in
the proximity of the copper complexes that may generate hydroxy radi-
cals, or in bulk solution after their release. The combined data obtained
from LPMOs themselves, and model systems inspired by the enzymes,
should provide answers to these fundamental questions in due course.
VI. References
1. Chu, S.; Majumdar, A. Nature. 2012, 488, 294–303.
2. Sun, Z.; Fridrich, B.; de Santi, A.; Elangovan, S.; Barta, K. Chem. Rev. 2018, 118,
614–678.
3. Gómez Millán, G.; Hellsten, S.; Llorca, J.; Luque, R.; Sixta, H.; Balu, A. M.
ChemCatChem. 2019, 11, 2022–2042.
4. Vaaje-Kolstad, G.; Westereng, B.; Horn, S. J.; Liu, Z.; Zhai, H.; Sorlie, M.; Eijsink,
V. G. H. Science. 2010, 330, 219–222.
5. Quinlan, R. J.; Sweeney, M. D.; Lo Leggio, L.; Otten, H.; Poulsen, J.-C. N.; Johansen,
K. S.; Krogh, K. B. R. M.; Jorgensen, C. I.; Tovborg, M.; Anthonsen, A.; Tryfona, T.;
Walter, C. P.; Dupree, P.; Xu, F.; Davies, G. J.; Walton, P. H. Proc. Natl. Acad. Sci.
USA. 2011, 108, 15079–15084.
6. Phillips, C. M.; Beeson IV, W. T.; Cate, J. H.; Marletta, M. A. ACS Chem. Biol. 2011,
6, 1399–1406.
7. Sheldon, R. A. ACS Sustainable Chem. Eng. 2018, 6, 4464–4480.
8. Østby, H.; Hansen, L. D.; Horn, S. J.; Eijsink, V. G. H.; Várnai, A. J. Ind. Microbiol.
Biotechnol. 2020, 47, 623–657.
9. Frandsen, K. E. H.; Simmons, T. J.; Dupree, P.; Poulsen, J.-C. N.; Hemsworth, G. R.;
Ciano, L.; Johnston, E. M.; Tovborg, M.; Johansen, K. S.; von Freiesleben, P.; Marmuse,
L.; Fort, S. E. B.; Cottaz, S.; Driguez, H.; Henrissat, B.; Lenfant, N.; Tuna, F.; Baldansuren,
A.; Davies, G. J.; Lo Leggio, L.; Walton, P. H. Nat. Chem. Biol. 2016, 12, 298–303.
10. Beeson, W. T.; Vu, V. V.; Span, E. A.; Phillips, C. M.; Marletta, M. A. Annu. Rev.
Biochem. 2015, 84, 923–946.
11. Hemsworth, G. R.; Johnston, E. M.; Davies, G. J.; Walton, P. H. Trends Biotechnol.
2015, 33, 747–761.
12. Lee, J. Y.; Karlin, K. D. Curr. Opin. Chem. Biol. 2015, 25, 184–193.
13. Kjaergaard, C. H.; Qayyum, M. F.; Wong, S. D.; Xu, F.; Hemsworth, G. R.; Walton,
D. J.; Young, N. A.; Davies, G. J.; Walton, P. H.; Johansen, K. S.; Hodgson, K. O.;
Hedman, B.; Solomon, E. I. Proc. Natl. Acad. Sci. USA 2014, 111, 8797–8802.
14. Bissaro, B.; Rohr, A. K.; Muller, G.; Chylenski, P.; Skaugen, M.; Forsberg, Z.; Horn,
S. J., Vaaje-Kolstad, G.; Eijsink, V. G. H. Nat. Chem. Biol. 2017, 13, 1123–1128.
208 Castillo
15. Elwell, C. E.; Gagnon, N. L.; Neisen, B. D.; Dhar, D.; Spaeth, A. D.; Yee, G. M.;
Tolman, W. B. Chem. Rev. 2017, 117, 2059–2107.
16. Striegler, S.; Dunaway, N. A.; Gichinga, M. G.; Barnett, J. D.; Nelson, A.-G. D. Inorg.
Chem. 2010, 49, 2639–2648.
17. Striegler, S.; Barnett, J. D.; Dunaway, N. A. ACS Catal. 2012, 2, 50–55.
18. Haldar, S.; Patra, A.; Bera, M. RSC Adv. 2014, 4, 62851–62861.
19. Striegler, S.; Fan, Q.-H.; Rath, N. P. J. Catal. 2016, 338, 349–364.
20. Comba, P.; Eisenschmidt, A.; Kipper, N.; Schiebl, J. J. Inorg. Boiochem. 2016, 159,
70–75.
21. Li, Z.; Chen, C. H.; Liu, T.; Mathrubootham, V.; Hegg, E. L., Hodge, D. B.
Biotechnol. Bioeng. 2012, 110, 1078–1086.
22. Castillo, I.; Neira, A. C.; Nordlander, E.; Zeglio, E. Inorg. Chim. Acta 2014, 422,
152–157.
23. Ambundo, E. A.; Deydier, M.-V.; Grall, A. J.; Aguera-Vega, N.; Dressel, L. T.;
Cooper, T. H.; Heeg, M. J.; Ochrymowycz, L. A.; Rorabacher, D. B. Inorg. Chem.
1999, 38, 4233–4242.
24. Hemsworth, G. R.; Henrissat, B.; Davies, G. J.; Walton, P. H. Nat. Chem. Biol. 2014,
10, 122–126.
25. Neira, A. C.; Martínez-Alanis, P. R.; Aullón, G.; Flores-Alamo, M.; Zerón, P.;
Company, A.; Chen, J.; Kasper, J. B.; Browne, W. R.; Nordlander, E.; Castillo, I. ACS
Omega. 2019, 4, 10729–10740.
26. Bacik, J.-P.; Mekasha, S.; Forsberg, Z.; Kovalevsky, A. Y.; Vaaje-Kolstad, G.; Eijsink,
V. G. H.; Nix, J. C.; Coates, L.; Cuneo, M. J.; Unkefer, C. J.; Chen, J. C. Biochemistry.
2017, 56, 2529–2532.
27. Caldararu, O.; Oksanen, E.; Ryde, U.; Hedegård, E. D. Chem. Sci. 2019, 10, 576–586.
28. Kandemir, B.; Kubie, L.; Guo, Y.; Sheldon, B.; Bren, K. L. Inorg. Chem. 2016, 55,
1355–1357.
29. Garajova, S.; Mathieu, Y.; Beccia, M. R.; Bennati-Granier, C.; Biaso, F.; Fanuel, M.;
Ropartz, D.; Guigliarelli, B.; Record, E.; Rogniaux, H.; Henrissat, B.; Berrin, J.-G.
Sci. Rep. 2016, 6, 28276.
30. Casella, L.; Gullotti, M.; Radaelli, R.; Di Gennaro, P. J. Chem. Soc., Chem. Commun.
1991, 1161–1612.
31. Baldwin, M. J.; Root, D. E.; Pate, J. E.; Fujisawa, K.; Kitajima, N.; Solomon, E. I. J.
Am. Chem. Soc. 1992, 114, 10421–10431.
32. Castillo, I.; Torres-Flores, A. P.; Abad-Aguilar, D. F.; Berlanga-Vázquez, A.; Orio,
M.; Martínez-Otero, D. ChemCatChem. 2021, 13, 4700–4704.
33. Rodríguez Solano, L. A.; Aguiñiga, I.; López Ortiz, M.; Tiburcio, R.; Luviano, A.;
Regla, I.; Santiago-Osorio, E.; Ugalde-Saldívar, V. M.; Toscano, R. A.; Castillo, I.
Eur. J. Inorg. Chem. 2011, 3454–3460.
34. Donoghue, P. J.; Tehranchi, J.; Cramer, C. J.; Sarangi, R.; Solomon, E. I.; Tolman, W.
B. J. Am. Chem. Soc. 2011, 133, 17602–17605.
Inorganic Models of Lytic Polysaccharide Monooxygenases 209
35. Dhar, D.; Tolman, W. B. J. Am. Chem. Soc. 2015, 137, 1322–1329.
36. Krishnan, V. M.; Shopov, D. Y.; Bouchey, C. J.; Bailey, W. D.; Parveen, R.;
Vlaisavljevich, B.; Tolman, W. B. J. Am. Chem. Soc. 2021, 143, 3295–3299.
37. Luo, Y. Handbook of bond dissociation energies, 1st Ed. CRC Press, USA, 2002.
38. Concia, A. L.; Beccia, M. R.; Orio, M.; Ferre, F. T.; Scarpellini, M.; Biaso, F.;
Guigliarelli, B.; Réglier, M.; Simaan, A. J. Inorg. Chem. 2017, 56, 1023–1026.
39. Wada, A.; Harata, M.; Hasegawa, K.; Jitsukawa, K.; Masuda, H.; Mukai, M.;
Kitagawa, T.; Einaga, H. Angew. Chem. Int. Ed. 1998, 37, 798–799.
40. Muthuramalingam, S.; Maheshwaran, D.; Velusamy, M.; Mayilmurugan, R. J. Catal.
2019, 372, 352–361.
41. Fukatsu, A.; Morimoto, Y.; Sugimoto, H.; Itoh, S. Chem. Commun. 2020, 56, 5123–
5126.
42. Hemsworth, G. R.; Taylor, E. J.; Kim, R. Q.; Gregory, R. C.; Lewis, S. J.; Turkenburg,
J. P., Parkin, A., Davies, G. J., Walton, P. H. J. Am. Chem. Soc. 2013, 135, 6069–6077.
43. Addison, A. W.; Rao, T. N.; Reedijk, J.; van Rijn, J.; Verschoor, G. C. J. Chem. Soc.
Dalton Trans. 1984, 1349–1356.
44. Yu, Z.; Thompson, Z.; Behnke, S. L.; Fenk, K. D.; Huang, D.; Shafaat, H. S.; Cowan,
J. A. Inorg. Chem. 2020, 59, 11218–11222.
45. Humphryes, K. J.; Karlin, K. D.; Rokita, S. E. J. Am. Chem. Soc. 2002, 124, 8055–
8066.
46. Choi, K.-Y.; Ryu, H.; Sung, N.-D.; Suh, M. J. Chem. Crystallogr. 2003, 33, 947–950.
47. Tse, E. C. M.; Schilter, D.; Gray, D. L.; Rauchfuss, T. B.; Gewirth, A. A. Inorg. Chem.
2014, 53, 8505–8516.
This page intentionally left blank
© 2023 World Scientific Publishing Company
https://doi.org/10.1142/9789811269493_0007
211
212 Behar et al.
I. Introduction
Biomimetic oligomers are synthetic molecules akin to natural biopoly-
mers, namely peptides, proteins, and oligonucleotides.8–10 Both natural
biopolymers and their artificial mimics are sequence-specific oligomers
capable of folding into well-defined three-dimensional structures in solu-
tion. One difference between them, however, is the identity of their back-
bone: while peptides and proteins are composed of a-amino acids,
biomimetic oligomers are assembled from non-natural monomers.11,12
Among the peptide mimics, peptoids are unique, as they combine proper-
ties of both biological materials and synthetic polymers; they are inert
toward many catalytic transformation13 and oxidative conditions14–16; and
possess better physiochemical and pharmacokinetic properties relative
to a-peptides, including stability in oxidative conditions13 and high
Structure and Function of Cu–Peptoid Complexes 213
A. Peptoid Synthesis
Peptoid oligomers have identical backbone to that of polypeptides, but the
side chains are appended to the nitrogen rather than to the a-carbon
(Figure 1a). Consequently, peptoids lack backbone chirality and cannot
form hydrogen bonds. In addition, the amide bonds in peptoids are tertiary
and this allows isomerization between the cis and trans conformations
(Figure 2a),21 in contrast to the a-peptides, where the trans conformation
is preferable. Furthermore, the lack of amide protons restricts the stabili-
zation of secondary structures by backbone hydrogen bonding, as
observed in a-peptides. Nevertheless, peptoids have several properties that
make them superior over peptides, and one of the main advantages is the
ease of the peptoids synthesis. Peptoids can be synthesized efficiently on
solid support, using the submonomer approach22 (Figure 1b), which
involves two main steps: N, N-diisopropylcarbodiimide (DIC) mediated
(a)
(b)
chiral side chains that were shown to induce helicity within peptoids, a
few of them have been determined as the side chains that shift the cis/
trans isomerization toward a majorly of cis amide bonds, for example,
N-(S)-1-phenylethylglycine (Nspe), N-(R)-1-phenylethylglycine (Nrpe)
(S)-N-(1-naphthylethyl)glycine (Ns1npe)17,24,26 and some were even able
to lead to “all cis” helices, for example, (R)-(−)-3,3-dimethyl-2-butylamine
(Nr1tbe), (S)-(+)-3,3-dimethyl-2-butylamine (Ns1tbe)21,27 (Figure 2b).
Several sequences requirements for obtaining peptoid helices were estab-
lished. First, the peptoid helix can be stabilized when the majority of the
peptoid’s side chains (at least two-third) are bulky and chiral.28 Second,
placing a chiral bulky monomer at the C-terminal of the peptoids sequence
can induce folding of the peptoid into a helical structure, so-called “ser-
geant-soldier” effect; in this case, the overall number of chiral subunits
within peptoid scaffold could be reduced.28 Third, the longer the peptoid
sequence is, the more stable is the helical structure (peptoid decamers and
beyond).28
The secondary structure of the peptoids can be initially estimated by
circular dichroism (CD) spectroscopy. For example, the typical spectra of
Nspe-bearing peptoids show typical double minima indicating a negative
(a)
(b) (c)
ellipticity near 218 nm and 202 nm, and a maximum indicating a positive
ellipticity near 190 nm (Figure 2c).17,28 These bands correspond to
n → p* transition of the amide chromophore and high- and low-wave-
length components of the exciton split p → p* transition. The similar,
mirror-like pattern of opposite chirality (double-maxima and minima,
correspondingly) is observed for Nrpe-bearing peptoids (Figure 2c).17 In
accordance with the helical peptoids having a pitch of about three residues
per turn,17 pendant groups located at positions i and i + 3 along the peptoid
backbone are facing the same side of the helix, and this can serve as a
handle for the design of peptoids as selective chelators for metal ions,
including Cu.
2. D
etermining the association constants and evaluating the
selectivity of the peptoid chelators to Cu(II)
One of the important parameters for evaluating the Cu(II) binding to pep-
toids is the association/dissociation constant of the Cu(II)–peptoid com-
plex. Association constants of metallopeptoids can be determined by one
of the following methods depending on the solubility of the peptoid and
the MB ligands. For peptoids with low water solubility UV–vis titrations
in aqueous solutions containing 10–20% methanol (the peptoid is dis-
solved in methanol prior to water addition). Low peptoid and metal ion(s)
concentrations are used in order to ensure low ionic strength. Association
constants are estimated by fitting the obtained data via non-linear regres-
sion using the appropriate kinetic equations.15 For water-soluble peptoids,
isothermal titration calorimetry (ITC) measurements can be used in buffer
solutions.42,43 Association constants and other physical parameters are
extrapolated directly, using the ITC software. In addition, in order to
verify the results (obtained in either method), competition experiments
can be conducted, either between the metal ions competing over the pep-
toid ligand, or between the peptoid ligand and a strong Cu(II) chelator
with a known association constant. Using the first competition method,
two different metal ions are titrated with a peptoid solution, followed by
UV–vis spectroscopy and MS. From these experiments, quantitative data
about the selectivity of one metal ion over the other is deduced. Using the
second competition method, the dissociation constant for Cu(II)–peptoid
complex could be determined by metal ions titrations with both chelators
followed by UV–vis spectroscopy. From the obtained UV–vis data, using
the appropriate equations and pH correction factors, dissociation constant
is found from the calculated slope.44
To determine the selectivity of the peptoids toward Cu(II), several
approaches could be applied. Typically, each peptoid is treated with a
mixture of different metal ions in different concentration and the solution
mixture is investigated. First, the solution mixture is analyzed by UV–vis
222 Behar et al.
(a)
(b)
(c)
Figure 6. Schematic illustration of the peptoids 7mer-HQ2 (a) and 12mer-HQ4 and
(b) forming intramolecular complex (7mer-HQ2)Cu and (12mer-HQ4)Cu2 upon
binding to Cu ions. (c) UV–vis titrations of 7mer-HQ2 and 12mer-HQ4 with Cu(II)
ions. Insets: molar-to-ratio plots, suggesting the stoichiometric peptoid:Cu(II) ratio.
Structure and Function of Cu–Peptoid Complexes 227
Figure 7. Schematic illustration of peptoid 7P6 forming a PP-I-like helix upon bind-
ing to Cu(II) ions. Changes in the CD spectra of free peptoid 7P6 (left) and (7P6)Cu
(right).
228 Behar et al.
aromatic ring stacking (Figure 9e, f), which leads to a unique assembly
resulting in the first metallopeptoid helicates (Figure 9h–i).46 Notably,
both experimental spectroscopic data and DFT calculations suggested that
the Cu–peptoid macrocycles assembled from the peptoids having ethanol
or methoxy groups are not present in solution but rather exist as mono-
meric complexes of the type PCu in acetonitrile, presenting a solid/solu-
tion state equilibrium. In contrast, the macrocycle assembled from the
peptoid having the amine group does exist in solution.46
Although the macrocycles described above represent the first example
of metal ions–mediated self-assembly of peptoids, they can still be
defined as distinct structures as they did not pack nor stack to form more
complex (and thus more biomimetic) supramolecular architectures.
Indeed, targeted/controlled self-assembly of subunits toward the forma-
tion of distinct/desired supramolecular architectures is one of the most
significant phenomena that happens in natural system and thus is highly
significant in chemistry and biology.
In natural processes, peptide-based secondary structures form versa-
tile turns, different helices, or multiple types of sheets by directed associa-
tion of the peptide to produce 3D well-defined frameworks that work as
membranes and/or channels.54 These supramolecular architectures play
significant role in several biochemical processes spanning recognition
and/or transportation of biologically important ions and molecules.55
Although there are several examples showing self-assembly of artificial
peptides upon metal binding,56–58 controlling this process toward a spe-
cific architecture is still challenging. Therefore, a worthy goal is control-
ling/directing the Cu(II)-mediated self-assembly of peptoids toward the
controlled formation of various architectures.
To this aim, a set of peptoid trimers having Bipy in the second posi-
tion within the sequence together with a pyridine (Pam) group in the
N-terminal that acts as coordinating group, and a non-coordinating group
at the C-terminus, were designed (Figure 10).47 The Pam group in the
N-terminal was included as it has the potential to interact with a benzyl or
Pam group from another P2Cu2 unit via p-p interactions, such that it might
enable further self-assembly to form supramolecular structures. For the
monomer in C-terminus, various aromatic and aliphatic groups were
232 Behar et al.
Figure 10. Peptoid sequences with varying substitution in C-terminus with Bipy in
the second and pyridine in the third substitution.
(a) (b)
(c)
Figure 11. (a) Folded single metallopeptoid duplex of Cu2+ with peptoid having
benzyl amine substitution. (b) Supramolecular rod-like helical strand formed by self-
assembly of several metallopeptoid duplexes. (c) Cryo-TEM analysis of a sample
crystal of Cu2+ with peptoid having benzyl amine substitution (100 mM).
(a) (b)
(c)
Figure 12. Spatial orientation of Cu(II) with peptoid having cyclohexyl amine sub-
stitution supramolecular cavity along the crystallographic axis in the absence (a) and
presence (b) of water (Color code — red: oxygen; blue: nitrogen; cyan: Cu(II); and
grey: carbon. Hydrogen atoms and perchlorate ions are removed for clarity). (c)
Cryo-TEM of the crystals (100 mM in water) with expanded view marked framed in
purple.
Structure and Function of Cu–Peptoid Complexes 235
Thus these were suggested to be the channel pores. Overall, for the first
time, controlling the sequence by simply modifying the non-coordinating
group at the C-terminus lead to control over the supramolecular architec-
ture of metallopeptoids.
detected upon the addition of H2PO4−. These results indicated that nano-
channels assembled from Cu(II)-PCy, having larger pores are a non-
selective host for Cl−, whereas the nano-channels assembled from
Cu(II)-PTBu interact with Cl− only.
The host–guest interactions between Cu–peptoid nano-channels and
anions were initially monitored by gradually increasing the concentration
of glycerol in the solution containing Cu(II)-PCy resulted in a gradual
shift of the vibrational stretching near 1605 cm−1 to a higher wavenumber
near 1616 cm−1 while gradual addition of glucose, arabinose, or mannose
did not lead to any shift in the FT-IR spectra. These results suggested that
there is a selective interaction, and possible selective binding, between
Cu(II)-PCy and glycerol, while the larger, sugar molecules, do not inter-
act with the metallopeptoid host. Similar titrations to the solution of
Cu(II)-PTBu did not lead to changes in the vibrational stretching of the
metallopeptoids, suggesting that the pore size of the corresponding nano-
channels is too small to host such biomolecules.
The interaction between the Cu–peptoid nano-channels and glycerol
was further verified by 1H-NMR study, in which Cu(II)-PCy or Cu(II)-
PTBu were added to D2O solutions of either glycerol or glucose. The
1
H-NMR spectra of glycerol and glucose were measured and monitored
before and after the addition of the Cu–peptoids. The spectra showed that
while there was no change in the chemical shift of glucose upon the addi-
tion of each Cu–peptoid solution or in the chemical shift of glycerol after
the addition of Cu(II)-PTBu (Figure 13a, b), a clear downfield shift was
observed after the addition of Cu(II)-PCy to glycerol (Figure 13c, d), sup-
porting the results obtained from the FT-IR analysis. Furthermore, the
sample solution containing the Cu(II)-PCy loaded with glycerol was lyo-
philized and the microcrystalline solid obtained was washed thoroughly
with methanol to remove residues of unbound dissolved glycerol (if any),
and dried in vacuum. The dry solid was solubilized in diluted hydrochloric
acid and treated with bicarbonate, and a sample from this solution was
analyzed by ESI-MS. The ESI-MS analysis showed a signal indicating the
mass of glycerol, indicating again the interaction of Cu(II)-PCy with
glycerol. Similar workup and analysis performed on solution sample of
Cu(II)-PTBu loaded with glucose did not reveal any trace of glucose in
Structure and Function of Cu–Peptoid Complexes 237
(a) (c)
(b) (d)
Figure 13. 1H-NMR of glycerol in D2O (a) before and (b) after addition of Cu(II)-
PTBu and (c) before and (d) after addition of Cu(II)-PCy (host and guest concentration
is 10 mM).
the ESI-MS spectrum, supporting the results obtained from the FT-IR and
1
H-NMR experiments indicating no glucose recognition.
Overall, these results demonstrate that the self-assembled nano-chan-
nels can interact with biologically relevant anions or molecules and that
by controlling their pore size we can regulate interactions with specific
anions and small biomolecules toward the use of these channels as
238 Behar et al.
selective receptors. The selective binding of the anions and small biomol-
ecules to Cu(II)-PCy and Cu(II)-PTBu also validate the self-assembly of
the Cu–peptoids in solution.
Figure 14. A schematic summary of Cu(II) binding to the helical peptoid Helix i + 3,
which is capable of intramolecular binding of Cu(II) ions or intermolecular binding
of two different metal ions in a selective manner.
(a)
(b)
(c)
Figure 16. (a) Peptoid sequences of 12mer peptoids 12P5 and 12P6. (b) A cartoon
demonstrating positive allosteric cooperative binding to (12P5)Cu, compared to (c)
no allosteric cooperativity of (12P6)Cu.
D. C
u(II)–Peptoid Complexes as Versatile, Efficient, and
Selective Bio-Inspired Catalysts
Enzymatic catalysis is one of the most important functions carried out by
natural metalloproteins. The catalytic activity of enzymes is largely
dependent on the cooperativity between binding sites, creating catalytic
Structure and Function of Cu–Peptoid Complexes 243
(a)
(b)
Figure 18. The three different cooperativity modes suggested for the catalytic activ-
ity of TG-DI (intra-resin and inter-resin) and of TG-BT (intra-peptoid).
Structure and Function of Cu–Peptoid Complexes 245
(a) (b)
ligand for Cu(II), an –OH group, as a tyrosine mimic, and a benzyl group;
the two non-catalytic groups were designed to act as a second coordina-
tion mimic aiming to stabilize the Cu(II)/Cu(III) center and facilitate its
activity. Upon treating the peptoid with Cu(II) in alkaline buffer, the
complex Cu(II)BPT(OH)2 was formed and both experimental and compu-
tational data revealed that Cu(II) is bound to BPT via Bipy and two
hydroxyl ions originating from the basic solution. At pH 11, the catalyst
was stable over at least 15 hours of electrolysis and could be reused for at
least nine times in 40-minute runs, resulting in an overall TON of ~56
within 6 hours, if the pH was readjusted after each run. Electrochemical
experiments revealed that the reversible CuIII/II oxidation wave, which is
a key for water oxidation, occurred at an unusually low E1/2 of +0.30 V
versus NHE in contrast to a peptoid analog, in which the ethanolic –OH
group was replaced by –OCH3 to give a potential of +0.50 V for the CuIII/
II
transition. Based on these as well as other electrochemical experiments,
spectroscopic data, and DFT calculations, a stable key peroxide interme-
diate was identified and an intramolecular cooperative catalytic pathway
was proposed, suggesting that the proximal –OH group and the etheric
oxygen atom attached to the Bipy moiety form strong hydrogen bonding
with the coordinated hydroxide groups, thus has a major role in the high
stability of the complex.
V. Summary
The incorporation of MB ligands within peptoids, first demonstrated in
2009 by Maayan et al. has opened up the new field of metallopeptoids.
Extensive research on Cu(II) coordination has expanded the field of pep-
toids and of peptidomimetics in general, beyond structure and structural
requirements and considerations, toward function. The broad understand-
ing regarding the role of Cu(II) in peptoid folding and self-assembly led
to the development of Cu(II)–peptoids as functional materials that are
inspired by natural Cu(II)-binding peptides and proteins. This includes the
ability to mimic complicated biological activities such as selective recog-
nition, allosteric cooperativity, and intramolecular cooperative catalysis
similar to the one performed by enzymes. Although exciting, these
achievements represent only the initial steps toward the many possibilities
Structure and Function of Cu–Peptoid Complexes 247
VI. References
1. Zastrow, M. L.; Pecoraro, V. L. Coord. Chem. Rev. 2013, 257(17–18), 2565–2588.
2. Crichton, R. R. Biological Inorganic Chemistry, 2nd Ed., Ed. Crichton, R. R.,
Chapter 14 “Copper — Coping with Dioxygen.” Elsevier, 2012, 279–296.
3. Bush, A. I. Curr. Opin. Chem. Biol. 2000, 4, 91–184.
4. Ala, A.; Walker, A. P.; Ashkan, K.; Dooley, J. S.; Schilsky, M. L. Lancet 2007, 369,
397–408.
5. Lovejoy, D. B.; Jansson, P. J.; Brunk, U. T.; Wong, J.; Ponka, P.; Richardson, D. R.
Cancer Res. 2011, 71, 5871–5880.
6. Andersen, O. Chem. Rev. 1999, 99, 2683–2710.
7. Kalia, K.; Flora, S. J. S. J. Occup. Health. 2005, 47, 1–21.
8. Gellman, S. H. Acc. Chem. Res. 1998, 31(4), 173–180.
9. Huc, I. Eur. J. Org. Chem. 2004, 17–29.
10. Hill, D. J.; Mio, M. J.; Prince, R. B.; Hughes, T. S. and Moore, J. S. Chem. Rev. 2001,
101 (12), 3893–4012.
11. Seebach, D. and Gardiner, J. Acc. Chem. Res. 2008, 41(10), 1366–1375.
12. Cheng, R. P.; Gellman, S. H.; DeGrado, W. F. Chem. Rev. 2001, 101(10), 3219–3232.
13. Ghosh, T.; Ghosh, P.; Maayan, G., ACS Catal. 2018, 8(11), 10631–10640.
14. Mohan, D. C.; Ghosh, P.; Ghosh, T.; Maayan, G., Chem. Eur. J. 2020, 26, 9573–9579.
15. Maayan, G.; Ward, M. D.; Kirshenbaum, K. Proc. Natl. Acad. Sci. U.S.A. 2009, 106
(33), 13679–13684.
16. Schettini, R.; De Riccardis, F.; Della Sala, G.; Izzo, I. J. Org. Chem. 2016, 81,
2494−2505.
17. Kirshenbaum, K.; Barron, A. E.; Goldsmith, R. A.; Armand, P.; Bradley E. K.;
Truong, K. T. V.; Dill, K. A.; Cohen, F. E.; Zuckermann, R. N. Proc. Natl. Acad. Sci.
U.S.A. 1998, 95(8), 4303–4308.
18. Ruan, G.; Engelberg, L.; Ghosh, P.; Maayan, G. Chem. Commun. 2021, 57, 939–942.
19. Miller, S. M.; Simon, R. J.; Ng, S.; Zuckermann, R. N.; Kerr, J. M.; Moos, W. H.
Drug Dev. Res. 1995, 35, 20–32.
20. Kwon, Y.; Kodadek, T. J. Am. Chem. Soc. 2007, 129(6), 1508–1509.
21. Roy, O.; Dumonteil, G.; Faure, S.; Jouffret, L.; Kriznik, A.; Taillefumier, C. J. Am.
Chem. Soc. 2017, 139(38), 13533–13540.
248 Behar et al.
22. Zuckermann, R. N.; Kerr, J. M.; Kent, S. B. H.; Moos W. H. J. Am. Chem. Soc. 1992,
114(26), 10646–10647.
23. Culf, A. S.; Ouellette, R. J. Molecules. 2010, 15(8), 5282–5335.
24. Stringer, J. R.; Crapster, J. A.; Guzei, I. A.; Blackwell, H. E. J. Am. Chem. Soc. 2011,
133(39), 15559–15567.
25. Pokorski, J. K.; Miller Jenkins, L. M.; Feng, H.; Durell, S. R.; Bai, Y.; Appella, D. H.
Org. Lett. 2007, 9(12), 2381–2383.
26. Wu, C. W.; Sanborn, T. J.; Zuckermann, R. N.; Barron A. E. J. Am. Chem. Soc. 2001,
123(13), 2958–2963.
27. Baskin, M.; Panz, L.; Maayan, G. Chem. Commun. 2016, 52, 10350–10353.
28. Wu, C. W.; Sanborn, T. J.; Huang, K.; Zuckermann, R. N.; Barron, A. E. J. Am. Chem.
Soc. 2001, 123(28), 6778–6784.
29. Maayan, G.; Ward, M. D.; Kirshenbaum, K. Chem. Commun. 2009, 2009, 56–58.
30. Baskin, M.; Maayan, G. Biopolymers. 2015, 104, 577–584.
31. Baskin, M.; Maayan, G. Chem. Sci. 2016, 7, 2809–2820.
32. Baskin, M.; Zhu, H.; Qu, Z. W.; Chill, J. H.; Grimme, S.; Maayan, G. Chem. Sci.
2019, 10, 620–632.
33. Pearson, R. G. J. Am. Chem. Soc.1963, 85(22), 3533–3539.
34. Rubino, J. T.; Franz, K. J. J. Inorg. Biochem. 2012, 107, 129–143.
35. Griffith, J. S.; Orgel, L. E. Q. Rev. Chem. Soc. 11, 381–393.
36. Pflaum, R. T.; Brandt, W. W. J. Am. Chem. Soc. 1954, 76(24), 6215–6219.
37. Osamu, Y.; Hiroshi, B.; Akitsugu, N. Bull. Chem. Soc. Jpn. 1973, 46(11), 3458–3462.
38. Johnston, W. D.; Freiser, H. J. Am. Chem. Soc. 1952, 74(21), 5239–5242.
39. Fleischel, O.; Wu, N.; Petitjean, A. Chem. Commun. 2010, 46, 8454–8456.
40. Galezowska, J.; Boratynski, P. J.; Kowalczyk, R.; Lipke, K.; and Czapor-Irzabek, H.
Polyhedron. 2017, 121, 1–8.
41. Maayan, G.; Yoo, B.; Kirshenbaum, K. Tetrahedron Lett. 2008, 49(2), 335–338.
42. Zabrodski, T.; Baskin, M.; Kaniraj, P. J.; Maayan, G. Synlett. 2015, 26(04), 461–466.
43. Ford, B. K.; Hamza, M.; Rabenstein, D. L. Biochemistry. 2013, 52(21), 3773–3780.
44. Ghosh, P.; Maayan, G. Chem. Sci. 2020, 11, 10127–10134.
45. Zborovsky, L.; Smolyakova, A.; Baskin, M.; Maayan, G. Chem. Eur. J. 2018, 24,
1159–1167.
46. Ghosh, T.; Fridman, N.; Kosa, M.; Maayan, G. Angew. Chem. Int. Ed. 2018, 57(26),
7703–7708.
47. Ghosh, P.; Fridman, N.; Maayan, G. Chem. Eur. J. 2021, 27, 634–640.
48. Ghosh, P.; Maayan, G. Chem. Eur. J. 2021, 27, 1383–1389.
49. Ghosh, P.; Rozenberg, I.; Maayan, G. J. Inorg. Biochem, 2021, 217, 111388.
50. Sakaguchi, U.; Addison, A. W. J. Chem. Soc. Dalton Trans. 1979, 600–608.
51. Garriba, E. and Micera, G. J. Chem. Educ. 2006, 83(8), 1229.
52. Macedi, E.; Meli, A.; De Riccardis, F.; Rossi, P.; Smith, V. J.; Barbour, L. J.; Izzo, I.;
Tedesco, C. Cryst. Eng. Comm. 2017, 19, 4704–4708.
Structure and Function of Cu–Peptoid Complexes 249
53. Shin, S. B. Y.; Yoo, B.; Todaro, L. J.; Kirshenbaum, K. J. Am. Chem. Soc. 2007, 129,
3218–3225.
54. Hu, C.; Chan, S. I.; Sawyer, E. B.; Yu, Y.; Wang, J. Chem. Soc. Rev. 2014, 43, 6498.
55. Li, J. Supramolecular Chemistry of Biomimetic Systems. Singapore, Springer, 2017.
56. Sawada, T.; Yamagami, M.; Ohara, K.; Yamaguchi, K.; Fujita, M. Angew. Chem. Int.
Ed. 2016, 55, 4519.
57. Sawada, T.; Yamagami, M.; Akinaga, S.; Miyaji, T.; Fujita, M. Chem. Asian J. 2017,
12, 1715.
58. Kwon, S.; Shin, H. S.; Gong, J.; Eom, J. H.; Jeon, A.; Yoo, S. H.; Chung, I. S.; Cho,
S. J.; Lee, H. S. J. Am. Chem. Soc. 2011, 133, 17618.
59. Moore, S. J.; Wenzel, M.; Light, M. E.; Morley, R.; Bradberry, S. J.; Gómez-Iglesias,
P.; Soto-Cerrato, V.; Pérez-Tomas, R.; Gale, P. A., Chem. Sci. 2012, 3(8), 2501–2509.
60. Valkenier, H.; Judd, L. W.; Li, H.; Hussain, S.; Sheppard, D. N.; Davis, A. P. J. Am.
Chem. Soc. 2014, 136(35), 12507–12512.
61. Hordyjewska, A.; Popiołek, Ł.; Kocot, J. Bio. Metals. 2014, 27, 611–621.
62. Lee, S.; Barin, G.; Ackerman, C. M.; Muchenditsi, A.; Xu, J.; Reimer, J. A.;
Lutsenko, S.; Long, J. R.; Chang, C. J. J. Am. Chem. Soc. 2016, 138, 7603–7609.
63. Savelieff, M. G.; Nam, G.; Kang, J.; Lee, H. J.; Lee, M.; Lim, M. H. Chem. Rev. 2019,
119(2), 1221–1322.
64. Esmieu, C.; Guettas, D.; Conte-daban, A.; Sabater, L.; Faller, P.; Hureau, C. Inorg.
Chem. 2019, 58(20), 13509–13527.
65. Mohan, C. D.; Kaniraj, P. J.; Maayan, G. Org. Biomol. Chem. 2018, 16, 1480–1488.
66. Hoover, J. M.; Stahl, S. S. J. Am. Chem. Soc. 2011, 33, 16901–16911.
67. Prathap, K. J.; Maayan, G. Chem. Comm. 2015, 51(55), 11096–11099.
68. Mohan, D. C.; Sadhukha, A.; Maayan, G. J. Catal. 2017, 355, 139–144.
69. Tian, H.; Yu, X.; Li, Q.; Wang, J.; Xu, Q. Adv. Synth. Catal. 2012, 354, 2671−2677.
70. Stamatin, Y.; Maayan, G. Eur. J. Org. Chem. 2020, 3147–3152.
71. Hunter, B. M.; Gray, H. B.; Müller, A. M. Chem. Rev. 2016, 116, 14120–14136.
72. Zhang, T.; Wang, C.; Liu, S.; Wang, J. L.; Lin, W. J. Am. Chem. Soc. 2014, 136,
273–281.
73. Barnett, S. M.; Goldberg, K. I.; Mayer, J. M. Nat. Chem. 2012, 4, 498–502.
Index
affinity to Cu(II), 216, 222, 238, 239 bis-oxidized quinone Cu(II) hydroxo,
albinism, 58 168
alkoxo, 162, 166 bis-phenoxide, 175
alkoxoperoxide, 168 bis-phenoxide salen, 176
allostery, 241 bis-thiosemicarbazones (bTSC), 21,
antimicrobial peptides (AMPs), 22, 30
33 bis(μ-oxido)dicopper(III) /
applications, 212, 217, 235 {CuIII2(μ-O)2}2+ / {Cu2O2}, 104
aromatic C−H hydroxylation, 108, Bond Dissociation Free Energy
111, 116 (BDFE), 155, 160, 165,180, 181
aromatic ring hydroxylation, 84, 88,
89, 91, 94, 95, 96, 110 C−F bond, 114, 115, 116
association constants, 221, 222 C−H oxidation, 107, 108
aureusidin synthase (AUS), 51 calix[6]trenamide Cu(I) complex, 161
carboxylato, 166
backbone conformation, 214 carboxylato complexes, 162
bakcbone isomerisation inducers, catechol oxidase (CO), 51, 83, 84,
214, 215 86, 87
bathocuproine (BC), 28 cellobiose, 195, 196
bathocuproine disulfonate (BCS), 28 cellulose, 188, 190, 196, 206
benzimidazole, 194, 206 chalcocite (Cu2S), 2
benzylic alcohol, 175 chalcopyrite (CuFeS2), 2
bidentate ligand, 129, 130, 134, characterisation methods, 220
136–139, 142 chelators, 213, 216, 221, 238, 240
biomimetic, 70, 82–84, 88, 89, 96, chiral, 214–216, 224, 225, 227,
102, 107, 118, 212 240
biomolecules (BMs), 5 chitin, 188, 190, 206
Bipy, 218, 230–232, 240, 243, 245, circular dichroism (CD), 215,
246 223–225, 227, 228, 240, 241
251
252 Index