You are on page 1of 10

Coordination Chemistry Reviews 257 (2013) 130–139

Contents lists available at SciVerse ScienceDirect

Coordination Chemistry Reviews


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

Review

Cu complexes that catalyze the oxygen reduction reaction


Matthew A. Thorseth, Claire E. Tornow, Edmund C.M. Tse, Andrew A. Gewirth ∗
Department of Chemistry, University of Illinois at Urbana-Champaign, 600 S. Matthews Avenue, Urbana, IL 61801, United States

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
2. Copper complexes with porphyrins and phthalocyanines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
3. Copper complexes with amino-alkyl ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4. Copper complexes with substituted 1,10-phenanthrolines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
5. Copper complexes with other aromatic N-donor ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6. Copper complexes with tris(2-pyridylmethyl)amine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
7. Copper complexes with substituted triazoles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

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

Article history: The oxygen reduction reaction (ORR) is employed in a large number of systems such as fuel cells and air
Received 26 December 2011 batteries. Currently, the catalyst with the lowest overpotential for the ORR is the enzyme laccase. Laccase
Received in revised form 2 March 2012 only functions at a very narrow pH range, and its large size prevents high current densities. Using copper
Accepted 30 March 2012
based catalysts to mimic the ORR activity is an area with many spectroscopic results, but relatively few
Available online 5 April 2012
electrochemical studies. This review catalogs the various copper based ORR catalysts and their activities.

Keywords:
© 2012 Elsevier B.V. All rights reserved.
Oxygen reduction reaction (ORR)
Cu catalysts
Electrochemistry

1. Introduction

The oxygen reduction reaction (ORR) is a four-electron reduc-


tion of dioxygen to water (Eq. (1)). The standard reduction potential
Abbreviations: ads, adsorbed; AFC, alkaline fuel cell; apy- for the reaction is 1.23 V, meaning that dioxygen is a powerful oxi-
hist, (4-imidazolyl)ethylene-2-amino-1-ethylpyridine; baEtO, dizing agent [1]. The triplet ground state of dioxygen often prevents
2-(bis(2-aminoethyl)amino)ethanol; bistripic, 1,2-bis(6-(bis(6-methylpyridin-
2-yl)methyl)pyridin-2-yl)ethane); C–E, chemical–electrochemical; 5-Cl-phen,
direct reaction with many molecules, which typically have a sin-
5-chloro-1,10-phenanthroline; DMP, 2,9-dimethyl-1,10-phenanthroline; DPP, 2,4- glet ground state, and thus prevents rapid reactivity of dioxygen
bis(2-pyridyl)pyrimidine; 4,7-dppds, 4,7-diphenyl-l,l0-phenanthrolinedisulfonate; despite the high oxidation potential. In acid, the ORR can proceed
ESR, electron spin resonance; GC, glassy carbon; Hdatrz, 3,5-diamino-1,2,4-triazole; by a direct four-electron and four-proton reaction with O2 to yield
Im, imidazole; Me3 TACN, 1,4,7-trimethyl-1,4,7-triazacyclononane; Me6 tren, N1 ,N1 -
H2 O, or can proceed by consecutive two-electron and two-proton
bis(2-(dimethylamino)ethyl)-N2 ,N2 -dimethylethane-1,2-diamine; NHE, normal
hydrogen electrode; NPM, non-platinum metal; ORR, oxygen reduction reaction; steps to yield H2 O2 and then H2 O. In base, the four-electron direct
PBS, phosphate buffered saline; PDT, 3-(2-pyridyl)-5,6-diphenyl-1,2,4-triazine; reduction results in the formation of four equivalents of hydrox-
phen, 1,10-phenanthroline; Piv, pivalamido; PMAP, (2-(pyridine-2-yl)- ide, while the two-electron reduced product is one equivalent of
N-(2-(pyridine-2-yl)ethyl)-N-(pyridine-2-ylmethyl)ethanamine); PMEA, hydroxide and one hydroperoxyl anion. The hydroperoxyl anion
(2-(pyridine-2-yl)-N,N-bis(pyridine-2-ylmethyl)ethanamine); poly-his, poly-
l-histidine; POM, polyoxometalate; RHE, reversible hydrogen electrode;
can then be further reduced by two electrons to three hydroxide
RRDE, rotating ring-disk electrode; SCE, saturated calomel electrode; TEPA, ions.
tris(2-(pyridine-2-yl)ethyl)amine; TPA, tris(2-pyridylmethyl)amine; TPT,
2,4,6-tris(2-pyridyl)-1,3,5-triazine; tren, tris(2-aminoethyl)amine; trpn, tris(3-
aminopropyl)amine. Acid form
∗ Corresponding author. Tel.: +1 217 333 8329; fax: +1 217 244 3186.
E-mail address: agewirth@illinois.edu (A.A. Gewirth). O2 + 4e− + 4H+  2H2 O E 0 = 1.229 V (1)

0010-8545/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ccr.2012.03.033
M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139 131

O2 + 2e− + 2H+  H2 O2 E 0 = 0.695 V (2)

H2 O2 + 2e− + 2H+  2H2 O E 0 = 1.763 V (3)

Base form

O2 + 2H2 O + 4e−  4OH− E 0 = 0.401 V (4)

O2 + H2 O + 2e−  HO2 − + OH− E 0 = −0.076 V (5)

HO2 − + H2 O + 2e−  3OH− E 0 = 0.878 V (6)

The ORR is important in systems ranging from fuel cells, to air-


batteries, to corrosion. In fuel cells, the ORR forms one of the half
reactions responsible for extracting energy from a fuel; the other
is the oxidation half reaction. Many fuels are utilized for fuel cells,
but the most common one for low temperature (<100 ◦ C) is hydro-
gen [2–4]. Fuel cells are interesting because they have the potential
to extract most of the energy available in a fuel, without the Carnot
cycle limitations that attend combustion as would occur, for exam-
ple, in an internal combustion engine or a gas turbine. Thus, the Fig. 1. Polarization diagram of dioxygen electroreduced at a carbon fiber electrode
thermodynamic efficiency of a fuel cell could be quite high. modified with laccase wired with polymers (curves 1 and 2) vs. a 6-␮m diameter
platinum electrode (curve 3) in pH 5 citrate buffer and Pt fiber electrode in 0.5 M
H2 SO4 in air.
Although fuel cells were invented some 170 years ago, and mod- Reprinted with permission from the American Chemical Society [15].
ern implementations have been in use since the 1960s, there is still
a variety of factors which inhibits widespread use of low temper-
ature fuel cells. One of the most compelling challenges is the slow 2. Copper complexes with porphyrins and phthalocyanines
kinetics and substantial overpotentials which attend the ORR. The
consequence of this overpotential means that fuel cells are unable Jasinski’s early discovery, that cobalt phthalocyanine could cat-
to deliver the full 1.23 V that a H2 /O2 fuel cell should deliver. Rather, alyze the ORR, has laid the foundation for subsequent research
the fuel cell potential is only ca. 40–60% of the thermodynamic on non-platinum metal (NPM)-based ORR catalysts [20]. Shortly
value when practical current densities are reached. This drop in following this discovery, Collman and Anson reported a more
potential means that the full thermodynamic efficiency of the sys- active dicobalt porphyrin dimer system [21–23]. Although a great
tem is not achieved, which obviates the major advantage of the fuel amount of work has since contributed to the initial successes
cell in the first place. of cobalt complexes, limited success using copper as the metal
The origin of the overpotential for the ORR is the high bond center has been achieved using both the phthalocyanine and por-
strength of the dioxygen double bond (498 kJ) [5]. Pt or its alloys phyrin sub-structures. Studies using metal phthalocyanines for
are the most commonly used cathode catalysts, and these exhibit O2 reduction report an activity trend following the order of:
an overpotential of ∼300 mV for the ORR [5–7]. Unfortunately, the Fe(II) > Co(II) > Ni(II) > Cu(II) [24,25]. These differences in electro-
large amount of Pt found in typical implementations makes the fuel catalytic activity have been clarified using MO theory. Bonding
cell cost prohibitive. There are also issues with catalyst stability and of O2 to the metal orbitals depends largely on vacancies in the d
susceptibility toward poisoning. Catalyst cost is prompting a search orbitals of the metal center [26]. Fe(II), Co(II), Ni(II), and Cu(II) each
for non-precious metal containing materials that might function as have filled dxz and dyz orbitals; however, Fe(II) and Co(II) possess
ORR catalysts. either an empty or half-filled dz 2 orbital, which can accept elec-
Many different non-Pt containing materials have been exam- trons from O2 . Following these observed trends, a study by Itaya
ined for their ORR activity [8–12]. One area of renewed focus et al. demonstrated that while Co phthalocyanine adsorbed on a
is copper-containing proteins, including laccase. Many different Au(1 1 1) surface shows significant activity toward the ORR, a Cu
methods have been utilized to connect these proteins to electrode phthalocyanine-modified electrode exhibits decreased O2 reduc-
surfaces for interrogation, including adsorption, covalent attach- tion current [27]. More recently, Bekaroğlu et al. reported the O2
ment, and by using mediators, both attached to the electrode reduction activity of ball-type supramolecular metallophthalocya-
surface, or otherwise available [5,13,14]. Laccase has an almost nines using both Co and Cu [28]. In acidic solution, O2 reduction
nonexistent overpotential (20 mV), a turn-over rate of 2.1O2 per with the Co complex occurs at a significantly lower overpotential
laccase per second, as well as reaching diffusion limited behavior than the Cu complex, a feature that was again attributed to the
by 70 mV of overpotential (Fig. 1) [15,16]. Unfortunately, laccase enhanced binding ability of O2 by the Co structure.
only functions in a narrow pH range, and its size makes it diffi- The copper–porphyrin systems also exhibit limited activ-
cult to achieve the high packing densities that would lead to high ity toward O2 reduction as well as poor catalyst stability. Qui
current densities with good mass transport. Additionally, the sta- et al. reported an insoluble copper(II) triphenylporphyrin complex
bility of these systems can be limited, particularly in the aggressive deposited on a glassy carbon (GC) electrode that exhibits no cat-
environment of a fuel cell [17–19]. alytic activity for the reduction of O2 in acidic conditions [29]. By
One way to achieve the activity of a copper enzyme might be modifying the porphyrin with a hexadecyl-pyridiniumyl group, a
through the agency of inorganic complexes. While many copper modest onset of 0.13 V vs. RHE was observed, yet the electrocata-
compounds exhibit reactivity with dioxygen and have been stud- lyst appears to be stable for only a few scans. Similarly, Takehira
ied spectroscopically, there have been relatively few attempts to and coworkers discovered that a water-soluble complex of cop-
understand their reactivity for the ORR electrochemically (Fig. 2). per and a sulfonated tetraphenylporphyrin does not reduce O2 ,
This review summarizes key results from copper ORR electrocata- unlike the analogous Co, Fe, and Mn porphyrin complexes [30].
lysts and develops future directions for research. Zhuang et al., however, demonstrated an onset of ORR at 0.44 V
132 M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139

Fig. 2. Diagram of some of the reported Cu ORR catalysts.

vs. RHE with a copper(II) tetraphenylporphyrin complex at pH 4.0 Cu complex. The complex exhibited a reversible CuI/II couple at
using a paraffin-impregnated graphite electrode [31]. Swavey et al. −0.23 V vs. SCE in pH 6.4 Britton–Robinson buffer with 0.1 mM
attempted to coat edge-plane pyrolytic graphite with a copper(II) imidazole. Removal of the imidazole resulted in a decrease in
tetraphenylporphyrin, however, due to limited stability of the cop- the CuI/II couple current upon subsequent scans, but this current
per(II) porphyrin in acidic conditions, it was not possible to evaluate could be revived with the addition of imidazole to the electrolyte.
the electrocatalytic reduction of O2 [32]. Upon addition of O2 to the solution, increasing cathodic current
was observed with an onset of 0.58 V vs. RHE with n = 4e− by
3. Copper complexes with amino-alkyl ligands Koutecky–Levich analysis. Reduction of H2 O2 occurs at the same
potential with an n = 2e− , and at similar rates to O2 reduction.
A copper complex with tris(3-aminopropyl)amine (trpn) A similar complex, [Cu(baEtO)(Im)](ClO4 )2 [baEtO = 2-(bis(2-
and imidazole (Im) was examined by Cai et al. [33]. aminoethyl)amino)ethanol], has an E1/2 (CuI/II ) = −0.248 V vs. SCE
[Cu(trpn)(Im)](ClO4 )2 was adsorbed onto a pyrolytic graphite in pH 7 phosphate buffer also adsorbed to pyrolytic graphite [34].
electrode by soaking the electrode in a solution containing the The CuI/II couple is observed even in the absence of imidazole in
M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139 133

the electrolyte and appears to be stable for 48 h of potential cycling.


The onset of O2 reduction occurs at 0.57 V vs. RHE with an n = 4e− as
determined from the Cottrell equation, while H2 O2 reduction was
not studied. The complex exhibits two linear regions in a plot of
peak current for O2 reduction vs. pH: the region of pH 4–6 where
the potential is independent of the pH, and pH 6–10, where the
peak current varies by −56 mV/pH.
Our group tested a few copper(II) complexes
with tris(2-aminoethyl)amine (tren), N1 ,N1 -bis(2-
(dimethylamino)ethyl)-N2 ,N2 -dimethylethane-1,2-diamine
(Me6 tren), and 1,4,7-trimethyl-1,4,7-triazacyclononane
(Me3 TACN) [35]. [Cu(tren)(H2 O)](ClO4 )2 and
[Cu(Me6 tren)(H2 O)](ClO4 )2 showed similar voltamme-
try to [Cu(trpn)(Im)](ClO4 )2 under Ar; no CuI/II couple
was observed. Onsets of O2 reduction were observed
for the [Cu(tren)(H2 O)](ClO4 )2 at 0.43 V vs. RHE and
[Cu(Me6 tren)(H2 O)](ClO4 )2 at 0.33 V vs. RHE, but both only exhib-
ited the 2e− reduction of O2 to H2 O2 . [Cu(Me3 TACN)(H2 O)](ClO4 )2
does have a CuI/II couple at 0.18 V vs. RHE and an onset of ORR
at 0.30 V vs. RHE, much more negative than other reported com-
plexes. We attributed the poor activity of the alkyl complexes to
slow electron transfer to the Cu center. The overlap between the
␴* orbitals on the alkyl ligand and the ␲* orbitals of the graphitic
carbon is small, leading to poor electron conduction from the
support to the catalyst.

4. Copper complexes with substituted


1,10-phenanthrolines

A series of copper(II) complexes with 1,10-phenanthroline


(phen), 5-chloro-1,10-phenanthroline (5-Cl-phen), 4,7-
diphenyl-l,l0-phenanthrolinedisulfonate (4,7-dppds),
2,9-dimethyl-1,10-phenanthroline (DMP), which may form a
mixture of the [CuL]2+ and [CuL2 ]2+ complexes, adsorbed onto
edge plane graphite electrodes were prepared by Anson and
coworkers [36–40]. Attaching chloro groups to the 5 position
resulted in a positive shift in E1/2 (CuI/II ) by 75 mV compared to the
unsubstituted phenanthroline. The CuI/II couple of the Cu complex
with DMP shifts further positive by 300 mV vs. copper phen. The
onset potentials for ORR for the phen, 5-Cl-phen, and 4,7-dppds
complexes are similar at ∼0.5 V vs. RHE. The addition of sterically
hindering methyl groups resulted in a much larger positive shift in
the onset of ORR to 0.69 V vs. RHE. The shift in ORR onset potential
is attributed to a simple outer sphere electron transfer to O2 being
the rate-determining step. All the catalysts examined also showed Fig. 3. Cyclic voltammograms of Cu complexes with substituted 1,10-
phenanthrolines adsorbed onto edge plane graphite in N2 (dotted) and air
activity for H2 O2 reduction, although at much slower rates than
(solid) saturated solutions at pH 4.8.
for O2 reduction.
Reprinted with permission copyright American Chemical Society [41].
In 2007, a more comprehensive library of copper complexes
with substituted 1,10-phenanthrolines was reported by Chidsey,
Stack, and coworkers [41]. The potential of the CuI/II couple can be on the quantity of immobilized copper complexes was observed
shifted more positive by 125 mV, as compared to phen, by substi- and the electroreduction of O2 proceeded via a binuclear [Cu2 O2 ]
tution with carboethoxy and nitro electron withdrawing groups. intermediate. In Anson’s case, since the copper complexes were
A more significant increase (∼300 mV) of the CuI/II couple was physisorbed onto the carbon electrodes, after binding one equiva-
observed by substitution of sterically hindering methyl and ethyl lent of O2 with [CuI (phen)]ads , a second [CuI (phen)]ads could rapidly
groups at the 2,9 positions (Fig. 3). Increasing the CuI/II couple migrate to form the [Cu2 O2 ] adduct, which resulted in the observed
increased the peak potential for ORR, as seen in Anson’s work, up to first order dependence on Cu.
0.59 V vs. RHE for Cu complex with 2,9-diethyl-phen. The proposed
mechanism of the ORR, as proposed by Anson, was modified. The 5. Copper complexes with other aromatic N-donor ligands
CuII center is first reduced to CuI , followed shortly by the binding of
oxygen. The Cu–O2 complex is then reduced and protonated in sub- [LCu2 (CH3 CO2 )2 ](ClO4 )2 ·5H2 O [L = 3,7,11,18,22,26-
sequent steps, an inner-sphere mechanism. The authors assumed hexaazatricyclo-[26.2.2.213,16 ]tetratriaconta-
a mono-Cu reaction mechanism. 1(31),13(33),14,16(34),28(32),29-hexaene)] was prepared and
Chidsey later covalently attached 3-ethynyl-phenanthroline to tested for its ORR activity [43,44]. The di-Cu hexaaza-xylyl macro-
azide-modified GC surfaces via the click reaction [42]. By varying cyclic complex exhibits a reversible 2e− reduction wave at −0.1 V
the copper coverage on the electrode, a second-order dependence vs. SCE under Ar. ESR spectroscopy does not exhibit exchange
134 M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139

Fig. 4. (a) Current–potential curves for reduction of O2 electrocatalyzed by a Cu complex with TPT adsorbed onto graphite electrode, recorded at various electrode rotating
rates. (b) Koutecky–Levich plots for reduction of O2 . O2 concentration: 2.4 × 10−4 mol/dm3 . Britton–Robinson buffer: pH 5.3. Scan rate: 100 mV/s. TPT surface concentration:
9.97 × 10−10 mol/cm2 .
Reprinted with permission copyright Elsevier [46].

coupling between the two Cu atoms, likely due to the large spacing electrode in basic solution. RRDE measurements reveal the pro-
(8.40 Å) between them. Introduction of O2 increases the cathodic duction of a small amount of H2 O2 , demonstrating that a parallel
current at potentials similar to the CuI/II couple, with an onset of 2e− pathway is followed, where O2 is reduced to H2 O2 rather than
ORR at −0.1 V vs. RHE. At low loading of the Cu complex, increased directly to H2 O. Studies in both basic and acidic media show that
H2 O2 production was detected. Increasing the loading eliminated the potentials at which the reduction of O2 takes place are similar,
the peroxide oxidation peak due to further reduction of H2 O2 suggesting that the rate-determining step is the pH independent
by the excess Cu complex. The mechanism of O2 reduction was formation of superoxide.
proposed to be a 2e− reduction of O2 to H2 O2 and then subsequent Na14 [SiW9 O34 Cu3 (N3 )2 (OH)(H2 O)]2 ·24H2 O, the first multidi-
reduction of the H2 O2 to H2 O, at a slower rate. mensional, antiferromagnetic, hexacopper complex based on azido
Treating [Cu2 (apyhist)2 Cl2 ]2+ [apyhist = (4- polyoxometalate (POM) units, was synthesized by Nadjo and
imidazolyl)ethylene-2-amino-1-ethylpyridine] with pH 9 buffer coworkers [48]. Cyclic voltammetry exhibits two Cu reduction
resulted in the formation of a tetracopper complex from the peaks at 0.38 V and 0.25 V vs. RHE, indicating two reduction pro-
deprotonation of the imidazole groups [45]. The Cu4 complex was cesses, CuI/II , and Cu0/I . The multiple one-electron reduction steps
cast directly onto a GC electrode, where poor electron transfer were suggested to be the key to trigger the reduction of O2 . The
rates (1.8 × 10−11 cm2 /s) to the Cu centers were observed. Intro- compound reduces O2 at a high rate via an overall four-electron
duction of O2 resulted in a modest increase of the onset potential process; however, the onset for the ORR is fairly negative at 0.33 V
of 350 mV over the GC electrode. The onset was 0.53 V vs. RHE in vs. RHE.
pH 9 solutions and chronoamperometric experiments revealed a In effort to synthesize an enzyme biomimetic, based on the
4e− transfer process. framework of laccase, Bard and coworkers reported the use of poly-
Cu complexes with 2,4-bis(2-pyridyl)pyrimidine (DPP), 3- l-histidine as a matrix and ligand to complex CuII [49]. Using a Cu
(2-pyridyl)-5,6-diphenyl-1,2,4-triazine (PDT), and 2,4,6-tris(2- complex-coated thin film GC electrode, it was observed that O2
pyridyl)-1,3,5-triazine (TPT) exhibit similar redox waves at reduction started at ∼0.07 V vs. Ag/AgCl, a slightly more positive
potentials more negative than −0.5 V vs. SCE [39,46]. Addition of onset than that achieved using a bare electrode (Fig. 5). At −0.3 V
Cu to the ligand adsorbed onto pyrolytic graphite electrodes exhibit vs. Ag/AgCl, the current for the CuII -poly-his-modified electrode
CuI/II couples at 0.0 V vs. SCE. The Cu complex with PDT does not was 2.3-fold larger as compared to the bare GC electrode, demon-
exhibit any reactivity with O2 , which is attributed to the tetrahe- strating electrocatalytic success with this proof of principle system.
dral coordination of the CuI complex. Both Cu(I) complexes with However, further performance enhancements are expected using
DPP and TPT as supporting ligands do catalyze the ORR with onsets other polypeptide mixtures in hopes of better mimicking laccase.
at 0.56 V and 0.51 V vs. RHE respectively with n = 4e− (Fig. 4). Both Pap and coworkers synthesized [CuII (LH−4 )]−2 ,
Cu(I) complexes also catalyze H2 O2 reduction to H2 O at similar (L = 8,17-dioxa-1,2,5,6,10,11,14,15-octaaza-tricyclo[13.3.1]
potentials, again at slower overall rates than O2 reduction. While eicosane-3,4,12,13-tetrone), which catalyzed the 4e− reduc-
the ligands exhibit reversible couples, their potentials are too neg- tion of oxygen and the simultaneous 2e− oxidation of ascorbic acid
ative to play a large role in the ORR catalysis. in basic condition (pH 8) [50]. Due to the rigidness of the planar,
Beyer et al. reported the use of polymeric film-modified tetradentate ligand, a square pyramidal intermediate is anticipated
Au electrodes of copper(II) Schiff-base complexes possessing when an additional ligand is coordinated to the apical position.
pyrrole groups for use as O2 reduction catalysts [47]. Specifi- In contrast to the biological system, which usually involves the
cally, [CuII (LH−1 )2 (L )] (L = 2-(3-pyrrol-1-yl-propylimino-methyl)- CuI/II redox couple, Cu(III) and Cu(II) are the active species in Pap’s
phenol, L = H2 O or a neutral 2e− donor) achieves electrochemical system, with a E1/2 of 0.64 V vs. RHE. According to the observed
O2 reduction at a potential 400 mV more positive than the bare rate expression, the Pap group proposed a peroxodicopper(III)
M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139 135

Fig. 5. Cyclic voltammograms of CuII -poly-his-modified GC electrode in 0.2 M PBS


(pH 7). The condition of modified films: 8 mM histidine residues, 2 mM Cu2+ . Scan
rate = 50 mV/s. (1) Bare GC, no O2 ; (2) bare GC, with O2 ; (3) modified GC, no O2 ; (4) Fig. 6. Graph of the onset of O2 (red squares) and H2 O2 (black triangles) for
modified GC, with O2 . [CuTPA](ClO4 )2 in Britton–Robinson buffers at various pHs.
Reprinted with permission copyright American Chemical Society [49]. Reprinted with permission copyright American Chemical Society [51].

intermediate, which dissociated into two copper(III)-oxyl radical (2-(pyridine-2-yl)-N,N-bis(pyridine-2-ylmethyl)ethanamine)


anions in the rate-determining step. Despite a high rate of O2 (PMEA), (2-(pyridine-2-yl)-N-(2-(pyridine-2-yl)ethyl)-
reduction, the onset for the ORR is 0.55 V vs. RHE. To date, this is N-(pyridine-2-ylmethyl)ethanamine) (PMAP), and
one of the rare Cu(III) complexes that carries out ORR. (tris(2-(pyridine-2-yl)ethyl)amine) (TEPA) [35]. The CuI/II cou-
ple potential increases as the lengths of the linkers between the
amine and the pyridine groups are increased, however, the onset
6. Copper complexes with tris(2-pyridylmethyl)amine
for the ORR is the same for all four complexes at 0.69 V vs. RHE
at pH 7. All four complexes also exhibit similar Tafel slopes. This
The mode of O2 binding to a Cu complex plays a large role in the
implies that the rate-determining step for TPA based complexes is
catalyst’s activity. O2 binds to di-Cu centers in two fashions, an end-
not the CuI/II couple reduction.
on trans ␮-1,2 coordination, or a side-on ␮-␩2 :␩2 mode. The initial
Addition of H-bonding groups has been shown in the past to
coordination is an oxidative addition that reduces O2 to either the
greatly affect the stability and reactivity of O2 complexes. The first
peroxo or oxo state, depending on the complex. [CuTPA]2+ binds
crystal structure of a Cu–OOH species was obtained with a bispi-
O2 in the end-on mode while [Cu2 bistripic(L )]2+ (L = a neutral 2e−
valamido complex that H-bonded to a H2 O2 reactant and was stable
donor) binds O2 side-on. In acidic solutions, [CuTPA]2+ has an over-
for over a month at a time, unlike previous reports [54]. Adding
potential for the ORR of 700 mV (pH 1), while [Cu2 bistripic(L )]2+
H-bonding groups to [CuTPA]2+ does not lower the overpotential,
has an overpotential of 890 mV (pH 2) [51]. The two complexes
however. [Cu((NH2 )2 -TPA)](NO3 )2 has the same ORR onset poten-
also exhibited different mechanisms based on the Tafel slopes:
tial of 0.53 V vs. RHE as TPA at pH 1, while [Cu(Piv2 -TPA)](NO3 )2
[CuTPA]2+ exhibits a 2e− reduction step (slope = 70 mV/dec) while
has a more negative onset of 0.40 V vs. RHE [35]. The lower onset
[Cu2 bistripic(L )]2+ likely goes through a chemical–electrochemical
for the bispivalamido complex is due to the steric repulsions of the
(C–E) step (slope = 167–680 mV/dec). The 2e− rate-determining
tert-butyl groups, preventing a di-Cu–O2 interaction. The similar-
step for [CuTPA]2+ is corroborated by the 30 mV/pH slope of the
ity of the voltammetry between [CuTPA]2+ and [Cu(Piv2 -TPA)]2+
onset potential between pH 4 and 10 (Fig. 6). We speculated
leads to the conclusion that the mechanism is the same between
that the differences between the two complexes were due to the
the two complexes and that protonation is likely not a part of the
nucleophilicity of the end-on complex (the TPA system) which
rate-determining step, which is in agreement with the mechanism
leads to easy protonation of the peroxo intermediate, allowing fur-
suggested by the Tafel slope.
ther reduction of a neutral hydroperoxo intermediate. The side-on
peroxo species (the bistripic system) is electrophilic, and the pro-
tonation of the species is slow and leads to greater overpotentials. 7. Copper complexes with substituted triazoles
By changing the loading of [CuTPA]2+ , we determined that the
ORR mechanism must be at least a di-Cu mechanism [51]. Decreas- Inspired by the activity of the multicopper oxidase active site
ing the partial pressure of O2 over the solution results in a linear of laccase, Thorum et al. have introduced one of the most effi-
decrease of the limiting current density [35]. The current density is cient synthetic copper electrocatalysts for the ORR to date [55].
proportional to the rate of reaction, so the reaction is first order in The carbon-supported, insoluble copper coordination complex of
O2 . The rate-determining reaction then must consist of two Cu and copper with 3,5-diamino-1,2,4-triazole (Hdatrz) has a reported
one O2 , as is seen in solution [52,53]. onset of 0.73 V vs. RHE at pH 7, and the highest overall onset of
The electroreduction of O2 involves a number of electron 0.86 V vs. RHE at pH 13 (Fig. 7). The onset potential vs. pH has a
and proton transfer events. As was seen with Cu complexes slope of 30 mV/pH. It is also reported to be stable at pH 7 over a
with substituted 1,10-phenanthrolines, the CuI/II couple usually testing period of 24 h. The crystal structure of [Cu2 (Hdatrz)2 (␮-
plays a large role in the ORR overpotential. Complexes based off of OH2 )(H2 O)4 (SO4 )](SO4 )·3.5H2 O reveals a binuclear complex with
[CuTPA]2+ do not exhibit a change in overpotential with a change in two copper centers, bridged by neutral guanazole ligands and one
the CuI/II couple. This is seen by examining the Cu complexes with water molecule, with Cu· · ·Cu spacing comparable to Cu spacing
136 M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139

Fig. 7. Graph of the onset potentials for the ORR vs. pH for [Cu(Hdatrz)(H2 O)2 ]2+ in
Britton–Robinson buffers. Fig. 8. Graph of the onset potentials for the ORR for the reported Cu catalysts nor-
Reprinted with permission copyright Angewandte [55]. malized to the RHE as listed in Table 1 vs. the reported pH.

in laccase, approximately 3.5 Å [56]. Magnetic susceptibility mea- scans, specifically in the mixed kinetic diffusion-control region,
surements of the carbon-supported complex demonstrate that spin reaching current densities of approximately −5.4 mA/cm2 .
pairing between Cu centers occurs, illustrating the presence of on
electrode synthetic multi-copper sites. 8. Conclusions
Studies of the O2 reduction activity of [Cu(Hdatrz)(H2 O)2 ]2+ in
the presence of several anions and poisons probed whether or not While laccase remains the single best ORR catalyst to date, repli-
these multi-copper sites are the active site for the ORR [57]. Using cating its activity with synthetic compounds remains elusive. To
poisons that have demonstrated coordination to Cu complexes, date, only a very limited number of Cu catalysts have been exam-
a significant decrease in O2 reduction activity (up to ∼200 mV ined for their ORR activity when compared to the number of Cu
decrease in the onset potential) is observed in the presence of compounds that exhibit O2 reactivity. To summarize results based
sodium fluoride, potassium thiocyanate, and ethanethiol. A lack on the catalysis of the ORR with Cu complexes, Table 1 and Fig. 8
of poisoning by sodium azide suggests the existence of an active show the onset potentials of the catalysts discussed in this publica-
site containing a neutral copper(II) complex with Hdatrz and SO4 2− tion. In Table 1, the reported onset potentials are converted to RHE,
ligands. and the pH at which the measurement was made is noted.
[Cu(Hdatrz)(H2 O)2 ]2+ is also a promising candidate for future Of the compounds studied, a few trends can be established.
application as a cathode catalyst in alkaline fuel cells (AFC). This Firstly, most of the compounds examined likely react in a di-Cu
is the first reported synthetic multi-copper complex for use in an fashion, as elegantly shown by Chidsey’s click chemistry study.
AFC environment. Employing an alkaline microfluidic H2 /O2 fuel Most of the observed Cu complexes catalyze both O2 and H2 O2
cell platform, Brushett et al. demonstrated that on a per metal basis, reduction, but the peroxide reduction occurs at a much slower rate.
the copper complex with triazole ligand outperforms both Pt/C and The ORR consists of 4H+ transfers, and as a result, a large depen-
Ag/C cathode catalysts [58]. However, the power density achieved dence on the pH is observed for many Cu complexes. The overall
using the copper triazole complex still falls below that achieved trend is a 30 mV/pH increase in the onset potential for ORR, seen
with Pt/C electrocatalysts. in the copper Hdatrz and copper TPA systems. The 30 mV/pH vs.
Based on the initial success of copper coordinated with an RHE is a −30 mV/pH dependence vs. NHE, which implies that the
amine-substituted triazole, it was expected that further increases rate-determining step for most catalysts is a 2e− transfer for every
of ORR onset potentials may come as a result of tailoring the tri- proton transferred, half the −60 mV/pH predicted for a 1e− transfer
azole ligand with different functionalities. Since the triazole motif by the Nernst equation. Anson also observed a similar −30 mV/pH
is expected to be essential for the ligand to coordinate the active trend in the CuI/II couple potential in copper complexes with sub-
copper site, various substitutions at the triazole’s 3,4, and 5 posi- stituted 1,10-phenanthrolines. Finding a catalyst with a smaller
tions have exposed possible correlations between structure and pH dependence on the ORR activity would be a large step toward
activity [59]. Further substitution of the 3,5-diamino-1,2,4-triazole developing a better ORR catalyst.
ligand at the 4 position with an amino group produces a linear The ligand choice plays a very large role in the efficacy of the
trinuclear unit with a sulfate group bridging the copper centers. ORR. Alkyl ligands seem to perform quite poorly, possibly due to
This complex has an onset of oxygen reduction at 0.67 V vs. RHE poor electronic conductivity. Most of the porphyrin and phthalo-
and reaches a steady-state diffusion-limited current of approx- cyanine complexes also exhibit low onset potentials for the ORR.
imately −5.0 mA/cm2 . In contrast, the copper(II) complex with The cause of the poor reactivity may be due to the forced square pla-
3,5-dimethyl-4-amino-1,2,4-triazole produces a tricopper triangu- nar geometry around the Cu center. Since open coordination sites
lar cluster framework, possessing a ␮-3-OH group bridging the are only available above and below the square planar plane, only the
trinuclear center. The onset of O2 reduction for this complex occurs filled dz 2 orbital is available to interact with the oxygen rather than
at 0.58 V vs. RHE and reaches diffusion-limited current densities the half-occupied dx 2 –y 2 orbital. Thus, O2 is not interacting with
only slightly better than the Vulcan carbon support. Removal of all the orbital which forms the ground state of the complex. Ligands
substitutions produces a copper(II) complex of 1,2,4-triazole, which with pyridine, pyrrole, imidazole, and triazole structures appear to
results in a triangular Cu(II) cluster, with an OH− group bridging the have the best overall activity.
tricopper center. In this case, O2 reduction commences at 0.70 V vs. Future studies of Cu based ORR catalysts should focus on further
RHE, but displays irreproducibility between anodic and cathodic understanding the fundamental reaction mechanisms, since it has
M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139 137

Table 1
Listing of the Cu catalysts studied for their ORR activity.

Catalyst Cond (pH/electrode) E1/2 CuI/II E1/2 CuI/II vs. RHEa EORR vs. Ref EORR vs. RHEa Ref.
2+
[Cu(phthalocyanine)] Adlayer on Au (111), hanging – – 0 V vs. RHE 0V [27]
meniscus, 0.1 M HClO4
[Cu(heptadecafluorodecyl GC disk, complex supported on Vulcan, – – −0.1 V vs. SCE 0.14 V [28]
substituted ball-type 0.5 M H2 SO4 ,
metallophthalocyanine)]2+
(BTMPcs)
[Cu(5-(4-pyridyl)-10,15,20- Complex adsorbed on GC disk, 0.05 M – – – – [29]
triphenylporphyrin)]2+ H2 SO4
[Cu(5-(4-N-hexadecyl- Complex adsorbed on GC disk, 0.05 M – – −0.15 V vs. Ag/AgCl 0.13 V [29]
pyridiniumyl)-10,15,20- H2 SO4
triphenylporphyrin
bromide)]2+
[Cu(meso-tetrakis(p-sulfon- Adsorbed on Ag electrode, 0.05 M – – −0.3 V vs. SCE 0.02 V [30]
atophenyl)porphyrins)]2+ H2 SO4
[Cu(5,10,15,20-tetraphenyl- Paraffin-impregnated graphite – – ∼ 0 V vs. Ag/AgCl 0.44 V [31]
21H,23H-porphyrin)]2+ electrode (PIGE), 0.1 M KCl (pH 4)
[Cu(5,10,15,20-tetrakis(4- Edge-plane pyrolytic graphite 0.10 V vs. Ag/AgCl 0.28 V – – [32]
hydroxy-3- electrode, 0.5 M H2 SO4
methoxyphenyl)porphyrin)]2+
[Cu(3,5-diamino-1,2,4- GC disk, supported on Vulcan, pH 7 – – 0.73 V vs. RHE 0.73 V [55]
triazole)]2+ Britton–Robinson
[Cu(1,2,4-triazole)]2+ GC disk, supported on Vulcan, pH 7 – – 0.70 V vs. RHE 0.70 V [59]
Britton–Robinson
[Cu(3,5-dimethyl-4-amino- GC disk, supported on Vulcan, pH 7 – – 0.58 V vs. RHE 0.58 V [59]
1,2,4-triazole)]2+ Britton–Robinson
[Cu(3,4,5-triamino-1,2,4- GC disk, supported on Vulcan, pH 7 – – 0.67 V vs. RHE 0.67 V [59]
triazole)]2+ Britton–Robinson
[Cu(phen)]2+ Edge-plane pyrolytic graphite −0.19 V vs. SCE 0.37 V −0.1 V vs. SCE 0.46 V [37]
electrode, pH 5.2, Britton–Robinson
2+
[Cu(phen)] pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0. 025 V vs. NHE 0.313 V 0.01 V vs. NHE 0.30 V [41]
20 mM AcOH
[Cu(5-Cl-phen)]2+ Edge-plane pyrolytic graphite −0.17 V vs. SCE 0.39 V −0.07 V vs. SCE 0.49 V [37]
electrode, pH 5.2 Britton–Robinson
buffer
[Cu(5-Cl-phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.05 V vs. NHE 0.34 V 0.04 V vs. NHE 0.33 V [41]
20 mM AcOH
2+
[Cu(2,9-Me2 -phen)] Edge-plane pyrolytic graphite 0.05 V vs. SCE 0.59 V 0.05 V vs. SCE 0.59 V [36]
electrode, pH 5 Britton–Robinson
buffer
[Cu(2,9-Me2 -phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.31 mV vs. NHE 0.60 V 0.29 V vs. NHE 0.58 V [41]
20 mM AcOH
[Cu(4,7-diphenyl-phen- pH 5.3, Britton Robinson, −0.18 V vs. SCE 0.38 V −0.15 V vs. SCE 0.41 V [40]
disulfonate)]2+
[Cu(5-NH2 -phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.02 V vs. NHE 0.31 V 0.01 V vs. NHE 0.30 V [41]
20 mM AcOH
[Cu(5-NO2 -phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.075 V vs. NHE 0.37 V 0.040 V vs. NHE 0.33 V [41]
20 mM AcOH
[Cu(3-CO2 Et-4-Cl-phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.090 V vs. NHE 0.38 V 0.065 V vs. NHE 0.36 V [41]
20 mM AcOH
[Cu(3,8-(CO2 Et)2 -4,7-Cl2 - pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.15 V vs. NHE 0.44 V 0.130 V vs. NHE 0.42 V [41]
phen)]2+ 20 mM AcOH
[Cu(2-Me-phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.215 V vs. NHE 0.51 V 0.205 V vs. NHE 0.50 V [41]
20 mM AcOH
[Cu(5-NH2 -2,9-Me2 -phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.285 V vs. NHE 0.58 V 0.275 V vs. NHE 0.57 V [41]
20 mM AcOH
[Cu(2,9-Et2 -phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.335 V vs. NHE 0.63 V 0.305 V vs. NHE 0.59 V [41]
20 mM AcOH
2+
[Cu(2,9-nBu2 -phen)] pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.340 V vs. NHE 0.63 V 0.260 V vs. NHE 0.55 V [41]
20 mM AcOH
[Cu(5-NO2 -2,9-Me2 -phen)]2+ pH 4.8, 100 mM NaClO4 , 20 mM NaAcO, 0.390 V vs. NHE 0.68 V 0.080 V vs. NHE 0.37 V [41]
20 mM AcOH
[Cu(3-ethynyl- pH 4.8, 50 mM NaAcO, 50 mM AcOH, 0.28 V vs. NHE 0.56 V 0.10 V vs. NHE 0.39 V [42]
phenanthroline)]2+ 1 M NaClO4
[Cu(trpn)(Im)](ClO4 )2 pH 6.4 Britton–Robinson buffer −0.23 V vs. SCE 0.40 V −0.05 V vs. SCE 0.58 V [33]
[Cu(baEtO)(Im)](ClO4 )2 Pyrolytic graphite electrode, pH 7 −0.248 V vs. SCE 0.42 V −0.10 V vs. SCE 0.57 V [34]
phosphate buffer
[Cu(tren)](ClO4 )2 pH 1, 0.1 M HClO4 , on Vulcan C – – 0.43 V vs. RHE 0.43 V [35]
[Cu(Me6 tren)](ClO4 )2 pH 1, 0.1 M HClO4 , on Vulcan C – – 0.33 V vs. RHE 0.33 V [35]
[Cu(Me3 TACN)](ClO4 )2 pH 1, 0.1 M HClO4 , on Vulcan C 0.18 V vs. RHE 0.18 V vs. RHE 0.30 V vs. RHE 0.30 V [35]
[Cu(TPA)](ClO4 )2 pH 1, 0.1 M HClO4 , on Vulcan C – – 0.53 V vs. RHE 0.53 V [51]
[Cu(TPA)](ClO4 )2 pH 7 Britton–Robinson buffer, on 0.23 V vs. RHE 0.23 V 0.69 V vs. RHE 0.69 V [51]
Vulcan C
[Cu(tripic)(NCMe)]PF6 pH 2 Britton–Robinson buffer, on – – 0.34 V vs. RHE 0.34 V [51]
Vulcan C
[Cu2 (bistripic)(NCMe)2 ](PF6 )2 pH 2 Britton–Robinson buffer, on – – 0.40 V vs. RHE 0.40 V [51]
Vulcan C
138 M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139

Table 1 (Continued)

Catalyst Cond (pH/electrode) E1/2 CuI/II E1/2 CuI/II vs. RHEa EORR vs. Ref EORR vs. RHEa Ref.

[Cu(PMEA)](ClO4 )2 pH 7 Britton–Robinson buffer, on 0.37 V vs. RHE 0.37 V 0.69 V vs. RHE 0.69 V [35]
Vulcan C
[Cu(PMAP)](ClO4 )2 pH 7 Britton–Robinson buffer, on 0.42 V vs. RHE 0.42 V 0.69 V vs. RHE 0.69 V [51]
Vulcan C
[Cu(TEPA)](ClO4 )2 pH 7 Britton–Robinson buffer, on 0.52 V vs. RHE 0.52 V 0.69 V vs. RHE 0.69 V [35]
Vulcan C
[Cu((NH2 )2 -TPA)](NO3 )2 pH 1 0.1 M HClO4 , on Vulcan C 0.42 V vs. RHE 0.42 V 0.53 V vs. RHE 0.53 V [35]
[Cu(Piv2 -TPA)](NO3 )2 pH 1 0.1 M HClO4 , on Vulcan C 0.22 V vs. RHE 0.22 V 0.40 V vs. RHE 0.40 V [35]
Cu with hexaaza-xylyl pH 7.3 borate buffer in solution at GC −0.1 V vs. SCE 0.58 V 0.0 V vs. SCE 0.68 V [44]
macrocycle electrode
[Cu(DPP)(L )]2+ pH 5.3 Britton–Robinson buffer, PG 0.0 V vs. SCE 0.56 V 0.0 V vs. SCE 0.56 V [39]
electrode
[Cu(PDT)(L )]2+ pH 5.3 Britton–Robinson buffer, PG 0.0 V vs. SCE 0.56 V – – [39]
electrode
[Cu(TPT)(L )]2+ pH 5.3 Britton–Robinson buffer, PG −0.17 V vs. SCE 0.39 V −0.05 V vs. SCE 0.51 V [46]
electrode
[Cu2 (apyhist)2 Cl2 ]2+
pH 9 phosphate buffer, GC electrode −0.35 V vs. SCE 0.43 V −0.25 V vs. SCE 0.53 V [45]
Cu with poly-l-histidine matrix Complex-coated thin film on GC – – −0.07 V vs. Ag/AgCl 0.54 V vs. RHE [49]
electrode, 0.2 M PBS (pH 7)
[Cu(2-(3-pyrrol-1-yl- Complex-coated Au electrode, 0.1 M – – ∼ 0 V vs. SCE 0.30 V vs. RHE [47]
propylimino-methyl)- H2 SO4
phenol)]
·24H
Na14 [SiW9 O34 Cu3 (N3 )2 (OH)(H2 O)]2pH 5,21OM (CH3 COOLi + CH3 COOH) −0.17 V vs. SCE 0.38 V −0.22 V vs. SCE 0.33 V [48]
[CuIII (LH−4 )]− (L = 8,17-dioxa- pH 8, 0.1 M NaClO4 −0.09 V vs. 3 M CE 0.64 V −0.18 V vs. 3 M CE 0.55 V [50]
1,2,5,6,10,11,14,15-octaaza-
tricyclo[13.3.1]
eicosane-3,4,12,13-tetrone)

E1/2 of CuI/II couple and EORR were estimated from the data provided in references. When no specific number was quoted in the text, potentials were estimated from the
graphs provided in the reference. Potentials were then referenced to the RHE by conversion of the reported reference potential to the NHE (0.197 V for Ag/AgCl, and 0.244 V
for SCE). NHE potentials were then converted to RHE by assuming a 60 mV/pH positive shift by the Nernst equation.

been observed that the O2 binding plays a critical role in how the [18] A. Zloczewska, M. Joensson-Niedziolka, J. Rogalski, M. Opallo, Electrochim. Acta
catalyst behaves. Progress toward this goal can be seen in recent 56 (2011) 3947.
[19] S. Shleev, J. Tkac, A. Christenson, T. Ruzgas, A.I. Yaropolov, J.W. Whittaker, L.
work by Tahsini et al. [60]. In attempts to elucidate the reactive Gorton, Biosens. Bioelectron. 20 (2005) 2517.
intermediate of O2 reduction, which could exist as a peroxo or bis- [20] R.J. Jasinski, Nature 201 (1964) 1212.
␮-oxo species, both binuclear and mononuclear copper(II) systems [21] J.P. Collman, M. Marrocco, P. Denisevich, C. Koval, F.C. Anson, J. Electroanal.
Chem. 101 (1979) 117.
were used to catalyze the reduction of O2 by decamethylferrocene. [22] J.P. Collman, C.S. Bencosme, R.R. Durand Jr., R.P. Kreh, F.C. Anson, J. Am. Chem.
By comparison of the activation entropies of electron transfer from Soc. 105 (1983) 2699.
ferrocene derivatives to either the binuclear copper complex or the [23] R.R. Durand Jr., C.S. Bencosme, J.P. Collman, F.C. Anson, J. Am. Chem. Soc. 105
(1983) 2710.
peroxo dicopper(II) intermediate, it was possible, for the first time,
[24] M. Savy, P. Andro, C. Bernard, G. Magner, Electrochim. Acta 18 (1973)
to determine what intermediates were produced and what roles 191.
they played. [25] P. Vasudevan, Santosh, N. Mann, S. Tyagi, Transit. Met. Chem. 15 (1990) 81.
[26] H. Alt, H. Binder, G. Sandstede, J. Catal. 28 (1973) 8.
With greater knowledge of the mechanisms, Cu–ligand systems
[27] S. Yoshimoto, A. Tada, K. Suto, K. Itaya, J. Phys. Chem. B 107 (2003) 5836.
can be tailor made to mimic the environment in the active site of [28] M. Ozer, A. Altndal, A.R. Ozkaya, O. Bekaroğlu, Dalton Trans. (2009) 3175.
laccase. As seen in Fig. 8, all of the reported Cu complexes to date, [29] R. Guilard, J.M. Barbe, S. Dong, Q. Qiu, Chin. J. Chem. 10 (1992) 309.
which are likely di-Cu, have very poor activity compared to laccase. [30] H. Orita, M. Shimizu, C. Nishihara, T. Hayakawa, K. Takehira, Can. J. Chem. 68
(1990) 787.
The first priority then is to obtain a functional tricopper model with [31] Q.-K. Zhuang, F. Scholz, J. Porphyrins Phthalocyanines 4 (2000) 202.
activity comparable to laccase. [32] G. Richards, S. Swavey, Eur. J. Inorg. Chem. (2009) 5367.
[33] C.X. Cai, K.H. Xue, X.Y. Xu, Q.H. Luo, J. Appl. Electrochem. 27 (1997) 793.
[34] M. Wang, X. Xu, J. Gao, N. Jia, Y. Cheng, Russ. J. Electrochem. 42 (2006) 878.
[35] M.A. Thorseth, C.S. Letko, T.B. Rauchfuss, A.A. Gewirth, in preparation.
References [36] J. Zhang, F.C. Anson, Electrochim. Acta 38 (1993) 2423.
[37] Y. Lei, F.C. Anson, Inorg. Chem. 33 (1994) 5003.
[1] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applica- [38] Y. Lei, F.C. Anson, Inorg. Chem. 34 (1995) 1083.
tions, 2nd edition, 2000. [39] J. Zhang, F.C. Anson, J. Electroanal. Chem. 348 (1993) 81.
[2] K.B. Prater, J. Power Sources 51 (1994) 129. [40] J. Zhang, F.C. Anson, J. Electroanal. Chem. 341 (1992) 323.
[3] L. Carrette, K.A. Friedrich, U. Stimming, ChemPhysChem 1 (2000) 162. [41] C.C.L. McCrory, X. Ottenwaelder, T.D.P. Stack, C.E.D. Chidsey, J. Phys. Chem. A
[4] S. Litster, G. McLean, J. Power Sources 130 (2004) 61. 111 (2007) 12641.
[5] A.A. Gewirth, M.S. Thorum, Inorg. Chem. 49 (2010) 3557. [42] C.C.L. McCrory, A. Devadoss, X. Ottenwaelder, R.D. Lowe, T.D.P. Stack, C.E.D.
[6] H.A. Gasteiger, S.S. Kocha, B. Sompalli, F.T. Wagner, Appl. Catal. B 56 (2005) 9. Chidsey, J. Am. Chem. Soc. 133 (2011) 3696.
[7] H.A. Gasteiger, J.E. Panels, S.G. Yan, J. Power Sources 127 (2004) 162. [43] C.J. McKenzie, H. Toftlund, M. Pietraszkiewics, Z. Stojek, K. Slowinski, Inorg.
[8] B. Wang, J. Power Sources 152 (2005) 1. Chim. Acta 210 (1993) 143.
[9] R. Bashyam, P. Zelenay, Nature 443 (2006) 63. [44] K. Slowinski, Z. Kublik, R. Bilewicz, M. Pietraszkiewicz, J. Chem. Soc. Chem.
[10] P.H. Matter, U.S. Ozkan, Catal. Lett. 109 (2006) 115. Commun. (1994) 1087.
[11] F. Jaouen, M. Lefevre, J.P. Dodelet, M. Cai, J. Phys. Chem. B 110 (2006) 5553. [45] I.O. Matos, T.L. Ferreira, T.R.L.C. Paixão, A.S. Lima, M. Bertotti, W.A. Alves, Elec-
[12] F.d.r. Jaouen, J.-P. Dodelet, J. Phys. Chem. C 113 (2009) 15422. trochim. Acta 55 (2010) 5223.
[13] S.C. Barton, H.-H. Kim, G. Binyamin, Y. Zhang, A. Heller, J. Phys. Chem. B 105 [46] V.L.N. Dias, E.N. Fernandes, L.M.S. da Silva, E.P. Marques, J.J. Zhang, A.L.B. Mar-
(2001) 11917. ques, J. Power Sources 142 (2005) 10.
[14] S.C. Barton, J. Gallaway, P. Atanassov, Chem. Rev. 104 (2004) 4867. [47] J. Losada, I. del Peso, L. Beyer, Inorg. Chim. Acta 321 (2001) 107.
[15] N. Mano, V. Soukharev, A. Heller, J. Phys. Chem. B 110 (2006) 11180. [48] C. Pichon, P. Mialane, A. Dolbecq, J. Marrot, E. Rivière, B. Keita, L. Nadjo, F.
[16] H. Schweiger, E. Vayner, A.B. Anderson, Electrochem. Solid-State Lett. 8 (2005) Sécheresse, Inorg. Chem. 46 (2007) 5292.
A585–A587. [49] Y.C. Weng, F.-R.F. Fan, A.J. Bard, J. Am. Chem. Soc. 127 (2005) 17576.
[17] L. Hussein, S. Rubenwolf, F. von Stetten, G. Urban, R. Zengerle, M. Krueger, S. [50] J.S. Pap, Ł. Szywriel, M. Rowińska-Żyrek, K. Nikitin, I.O. Fritsky, H. Kozłowski, J.
Kerzenmacher, Biosens. Bioelectron. 26 (2011) 4133. Mol. Catal. A: Chem. 334 (2011) 77.
M.A. Thorseth et al. / Coordination Chemistry Reviews 257 (2013) 130–139 139

[51] M.A. Thorseth, C.S. Letko, T.B. Rauchfuss, A.A. Gewirth, Inorg. Chem. 50 (2011) [56] E. Aznar, S. Ferrer, J. Borras, F. Lloret, M. Liu-Gonzalez, H. Rodriguez-Prieto, S.
6158. Garcia-Granda, Eur. J. Inorg. Chem. (2006) 5115.
[52] S. Fukuzumi, H. Kotani, H.R. Lucas, K. Doi, T. Suenobu, R.L. Peterson, K.D. Karlin, [57] M.S. Thorum, J.M. Hankett, A.A. Gewirth, J. Phys. Chem. Lett. 2 (2011) 295.
J. Am. Chem. Soc. 132 (2010) 6874. [58] F.R. Brushett, M.S. Thorum, N.S. Lioutas, M.S. Naughton, C. Tornow, H.-R.M.
[53] R.R. Jacobson, Z. Tyeklar, A. Farooq, K.D. Karlin, S. Liu, J. Zubieta, J. Am. Chem. Jhong, A.A. Gewirth, P.J.A. Kenis, J. Am. Chem. Soc. 132 (2010) 12185.
Soc. 110 (1988) 3690. [59] C.E. Tornow, A.A. Gewirth, in preparation.
[54] R.L. Shook, A.S. Borovik, Chem. Commun. (2008) 6095. [60] L. Tahsini, H. Kotani, Y.-M. Lee, J. Cho, W. Nam, K.D. Karlin, S. Fukuzumi, Chem.
[55] M.S. Thorum, J. Yadav, A.A. Gewirth, Angew. Chem. Int. Ed. 48 (2009) 165. Eur. J. 18 (2012) 1084.

You might also like