You are on page 1of 16

Perspective

pubs.acs.org/acscatalysis

Reaction Pathways of Hydrogen-Evolving Electrocatalysts:


Electrochemical and Spectroscopic Studies of Proton-Coupled
Electron Transfer Processes
Noémie Elgrishi,‡ Brian D. McCarthy,‡ Eric S. Rountree, and Jillian L. Dempsey*
Department of Chemistry, University of North Carolina, Chapel Hill, North Carolina 27599-3290, United States

ABSTRACT: Proton-coupled electron transfer (PCET) re-


actions are at the heart of the catalytic processes involved in
hydrogen evolution. In this Perspective, the state-of-the-art
spectroscopic and electrochemical methods available to
elucidate the mechanisms of PCET reactions of fuel-forming
catalysts are presented. Through examples of our recent work,
the applications of these methods are discussed with a focus on
the type of information and the accuracy that can be obtained
from each. Three case studies are presented to illustrate
different possible origins for peak shifts observed in cyclic
voltammograms.

KEYWORDS: PCET, electrocatalysis, cyclic voltammetry, stopped-flow, solar fuels

1. INTRODUCTION Scheme 1. Square Scheme Representation of a Homogeneous


PCET Reaction Where Species M Receives One Electron and
Development of cost-effective solar energy storage technology is
One Proton To Form MHa
paramount for realizing the transition away from fossil fuel
energy.1−3 In terms of chemical fuels, chemical bonds in energy-
poor molecules must be rearranged to yield energy-rich fuels.
Frequently, these rearrangements involve the movement of
multiple protons and electrons such as in H2 evolution, CO2
reduction, and water oxidation.4−6 Minimizing energy loss
during these rearrangements is imperative for practical large-
scale fuel production. As energy losses are intricately related to a
the kinetics of elementary chemical steps, careful modulation of H−A represents an acid molecule, and ET, PT, and CPET indicate
the pathway taken from energy-poor reactants to energy-rich electron transfer, proton transfer, and concerted proton−electron
products is critical. It is particularly attractive to mediate these transfer, respectively.
complex transformations using molecular transition metal
catalysts. As such, opportunities exist to modulate reaction necessarily proceed through high-energy intermediates. Con-
mechanism and perturb reaction kinetics through systematic sequently, mechanistic study of PCET reactions involving
changes to catalyst structure and electronics, as well as reaction transition-metal complexes is an important step toward design-
conditions. ing more efficient catalysts for hydrogen evolution reactions.
A common motivation for understanding reaction mechanisms Further, as PCET reactions are additionally relevant outside the
of catalytic transformations is the hope that this insight can lead subfield of fuel formation reactions, the quest to understand
to modification and improvement of a catalyst. For example, if PCET in hydrogen evolution catalysts simultaneously contrib-
the hydrogen evolution mechanism for a specific catalyst is utes more broadly useful fundamental knowledge.
understood, a second generation catalyst can be rationally Our group has focused on understanding PCET processes in
designed. An elegant example is the evolution and improvement matters relevant to energy conversion and storage processes;
of the [Ni(PR2 NR2 ′)2]2+ family of hydrogen evolution catalysts.7 specifically, in the context of homogeneous electrochemical
Notably, within the context of hydrogen evolution, the catalyst catalysts for the hydrogen evolution reaction (HER). As noted
orchestrates the union of two electrons and two protons, and above, it has become increasingly clear that an understanding of
elementary steps are frequently proton-coupled electron transfer
(PCET) reactions. PCET is traditionally viewed in terms of a Received: March 16, 2016
square scheme (Scheme 1) where concerted and stepwise routes Revised: April 21, 2016
are possible. Reactions that proceed via the stepwise routes

© XXXX American Chemical Society 3644 DOI: 10.1021/acscatal.6b00778


ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

electrochemical PCET processes are crucial for advancing HER reaction for a CPET process (Figure 1A). Although the
and other catalytic processes. Studies of PCET reaction formation of high-energy, charged intermediates are circum-
mechanisms to date have primarily involved organic substrates vented, this termolecular reaction can give rise to a high kinetic
relevant to biological charge transport pathways, with only a few barrier. As such, stepwise ET-PT or PT-ET mechanisms are
examples of transition metal-based systems.8−11 Further, while a often favored kinetically.37,38 Conversely, for electrochemical
rich literature exists for the study of homogeneous-based PCET, PCET, the reaction between the PCET substrate and the proton
where electron and proton transfers occur between discrete donor/acceptor occurs at an essentially infinite electrode surface
molecules (Figure 1A),9,12−16 there are fewer reports which (Figure 1B). As such, the kinetic barriers for an electrochemical
CPET process will be different than the termolecular
homogeneous example.
Second, the electron transfer driving force changes during a
potential-sweep electrochemical experiment. It has been
demonstrated that via careful selection of chemical oxidants,
the PCET oxidation of a tungsten-hydride bond can be induced
to undergo a CPET mechanism.10 In an electrochemical
experiment such as cyclic voltammetry, the driving force for
electron transfer changes as a function of time. This suggests that
the PCET mechanism may change during a single experiment as
the potential is swept (Figure 2). It has already been

Figure 1. Cartoon depictions of (A) homogeneous termolecular PCET


(case for two donors and one acceptor shown; more partners are
possible) and (B) electrochemical PCET.

detail individual electrochemical PCET steps (specifically where


the electrode is a partner in the electron transfer reaction, Figure Figure 2. As applied potential changes throughout a variable-potential
1B).11,17−24 This is not due to a lack of methodology or theory experiment, the PCET mechanism accessed may change as the driving
extensive information and experimental examples exist for the force changes.
study of electrochemical mechanism.21,22,25−32 In addition,
detailed theoretical analyses for mechanism-dependent electro- demonstrated that mechanisms for hydrogen evolution can in
chemical PCET rate expressions have been reported.33−35 some cases switch from ECEC to EECC as the applied potential
However, only recently have these methods received more use is scanned cathodically.24,39 This strongly suggests that a change
in the study of PCET processes relevant to energy of potential may promote a change in the mechanism of an
storage.21,23,24,27,31,36 individual PCET event.
In the study of electrochemical mechanisms, elementary steps In this Perspective, the methods available to chemists for the
are generally divided into two categories: electrochemical (E) study of PCET reactions in organic solvents will be introduced,
steps and chemical (C) steps. An E step involves the movement with a focus on recent developments in electrochemical methods
of an electron to/from the electrode, whereas any other and their applications in acetonitrile. Our efforts in applying
homogeneous chemical transformation corresponds to a C these methods to the elucidation of PCET mechanisms of
step. Additional subscripts may designate reversibility (subscript hydrogen-evolving catalysts will then be discussed, highlighting
r) or irreversibility (subscript i) of individual steps, while the term the role of each method in building a broad mechanistic picture.
(EC) indicates a concerted E and C step. This classic Finally, three case studies of model systems relevant to hydrogen
electrochemical nomenclature is readily paired with traditional production will be presented to reflect upon different origins of
homogeneous PCET terminology. For example, an electron peak shifts observed in cyclic voltammograms and their
transfer followed by a proton transfer, referred to as ET-PT in mechanistic implications.
PCET literature, is designated more generally as an EC
electrochemical mechanism. Meanwhile, the concerted-proton 2. METHODS FOR ELUCIDATING REACTION
electron transfer (i.e., CPET) pathway would be designated MECHANISMS AND EXPERIMENTAL
(EC). CONSIDERATIONS
Proton transfer generally occurs between discrete molecular In this section, methodology for the elucidation of PCET
proton donors and proton acceptors; however, electron transfer mechanisms, particularly PCET reactions in catalysis, will be
may occur between solution species or, as considered more discussed. The primary methods utilized in our laboratory and
frequently, to and from the electrode directly. As electron others will be described, along with the information that can be
transfer occurs between solution and an electrode, two gained from their application. Care will be taken to highlight
interesting and, as of yet, incompletely explored influences on factors influencing the quality of the data and analysis, as well as
the PCET mechanism arise. potential pitfalls.
First, the kinetic barrier of the CPET pathway is impacted. In 2.1. Time-Resolved Spectroscopy. Two additional spec-
purely homogeneous systems, PCET may involve three discrete troscopic techniques have found utility in the study of PCET
molecules which must encounter one another in a ternary reactionstopped-flow spectroscopy and transient absorption
3645 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

spectroscopy. Both techniques take advantage of changes in the py, which has a time resolution of microseconds due to the
absorption profiles during a reaction to gain mechanistic and necessary mixing time, transient absorption spectroscopy allows
kinetic information; specific details are discussed below. investigation of kinetics in the pico- to nanosecond regime.
2.1.1. Stopped-Flow Spectroscopy. Stopped-flow spectros- 2.2. Electrochemistry. Electrochemical methods have vast
copy is a powerful tool for extracting kinetic information and has precedent in the study of reaction mechanisms,25,26 and
been used to examine PCET in various systems.10,40−43 Stopped- consequently, electrochemistry has been increasingly utilized in
flow has been applied to the study of both catalytic cycles and the study of PCET reactions. One key benefit of electrochemistry
elementary reaction steps where the reaction is accompanied by is the dynamic control over the energy of electrons. Of the
clear spectral changes. In a typical stopped-flow experiment, the multitude of electrochemical techniques, cyclic voltammetry has
contents of two syringes are rapidly mixed and the absorbance of emerged as the favorite tool for uncovering reaction mechanisms
the system recorded over time, either as full spectra at designated and determining rate constants due to the rich information
time delays or as single-wavelength kinetics traces (absorbance vs available from current−potential responses and the extensive
time). A general-purpose UV−vis absorption spectrometer may analytical methods established for interpreting resulting data.
be used in cases where reaction kinetics are slow.44 Additionally, cyclic voltammetry is a nondestructive technique;
Stopped-flow methods are amendable to monitoring the only the minute volume of reactants immediately next to the
reaction between many reagent types or even combinations of working electrode (the reaction layer) is involved in a
reagents. For PCET studies, these may be chemical reductants, measurement. Finally, over the last 5 years, great strides have
oxidants, hydride donors, acid, or base, for example. Once the been made in the development of mathematical models to extract
two solutions have been rapidly mixed, the absorbance is quantitative kinetic information from cyclic voltammograms of
recorded as a function of time. When monitoring a single multielectron, multisubstrate reactions. Presented here is a brief
elementary step, pseudo-first-order conditions are often summary of the electrochemical tools most commonly used in
employed, enabling the reaction kinetics to be fit to a single our laboratory; more extensive discussions are available.26,32
exponential to extract an observed rate constant kobs. For 2.2.1. Kinetic Analysis Using Catalytic Plateau Current and
complex reactions, the kinetics traces can be fit using a kinetics Half-Wave Potential. In the case of electrochemical catalysis
model based on a series of differential equations describing the under ideal conditions with no limiting side phenomena or
concentrations of each reactant, intermediate, and product as a substrate depletion, an S-shaped response with superimposable
function of time. Obtaining rate constants through stopped-flow forward and reverse traces is obtained by cyclic voltammetry.32
measurements has been shown to corroborate electrochemically An observed rate constant kobs can be extracted using the
derived rate constants.24,40,41 experimental catalytic plateau current (ipl, which is independent
2.1.2. Transient Absorption Spectroscopy. Time-resolved of scan rate) and the potential at which half the catalytic current is
optical monitoring techniques can also be coupled to photo- reached (Ecat/2). For a two-electron catalytic process and
triggering methods to initiate PCET reactions. In traditional laser assuming that both electron transfers occur at the electrode
flash-photolysis methods, excited-state reductants (including with the second reduction being easier than the first, the catalytic
ruthenium, rhenium, iridium, and copper polypyridyl com- plateau current is given by eq 1 where [Cat]0 is the bulk solution
plexes) are employed to initiate the one-electron reduction of a catalyst concentration, D is the catalyst diffusion coefficient, F the
redox-active molecule upon excitation by a pulsed laser.8,45−53 In Faraday constant, and A is the electrochemically accessible
the presence of a proton source, reduction will often be electrode surface area.
accompanied by a protonation reaction.12,54,55 As in stopped-
flow spectroscopy, the distinct optical signatures of the different i pl = 2FA[Cat]0 D kobs (1)
redox forms of the catalysts, as well as the protonated species,
allow the reaction intermediates and products to be spectroscopi- eq 1 is often divided by the one-electron peak current in the
cally detected and monitored. However, in these phototriggered absence of substrate, i0p, given by the Randles−Sevcik equation
transient absorption experiments, it is the change in the sample (eq 2, where F is the Faraday constant, T is temperature, and υ is
absorbance (as opposed to the absolute absorbance) that is scan rate) to yield eq 3, which eliminates the need for the
monitored as a function of time. Data can be recorded as diffusion coefficient and electrode surface area to be
difference spectra at designated time delays or as single- independently determined:
wavelength kinetics traces, depending upon experimental setup. DυF
While phototriggered reactions typically involve the initiation i p0 = 0.446FA[Cat]0
of electron transfer, proton transfer reactions may be photo- RT (2)
initiated as well using photoacids.12 Photoacids are species that
become powerful proton donors in their electronic excited states, i pl kobsRT
allowing in situ generation of a proton donor upon pulsed laser = 0.448
i p0 υF (3)
excitation.56,57 Beginning with the reduced metal species,
photoexcitation of the photoacid will promote proton transfer
Expressions for kobs as a function of the rate constants for the
to yield a PCET reaction intermediate. Like photoinduced
electron transfer reactions, proton transfer reactions can also be first and second chemical steps, k1 and k2 respectively, can be
monitored optically if protonation is accompanied by a distinct derived for specific mechanisms.28 For the mechanisms ECEC
change in absorption. While identifying the conditions to and EECC, assuming S-shaped catalytic responses, expressions
phototrigger an electron transfer or proton transfer reaction for kobs as well as Ecat/2, the potential at which half of the plateau
for a laser flash-photolysis experiment may be challenging, current is reached, are given in Table 1 as a function of k1, k2, and
especially when other reactive substrates are present in solution, E1/2, the potential of the electrochemical event triggering
this time-resolved spectroscopy allows much faster reactions to catalysis. In many cases, either k1 or k2 is rate limiting, resulting
be optically monitored. Compared to stopped-flow spectrosco- in a collapse of the equations in Table 1 simpler expressions.
3646 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

Table 1. Values of kobs and Ecat/2 as a Function of the Rate figures of merits of different homogeneous catalysts by
Constants of the Two Chemical Steps Considered28,39 illustrating the trade-off between activity and overpotential.
The turnover frequency is defined as the moles of product
ECEC EECC generated per unit of time divided by the moles of the active form
k1 of the catalyst present in the reaction-diffusion layer, and the
k1
kobs k1 1+
k1 expression of the maximal turnover frequency TOFmax is
⎛ k2 ⎞
1+
k2 k2 ⎜⎜1 +

⎟⎟
k1 ⎠
dependent upon the mechanism considered.28,32
⎛ ⎞ TOFmax
RT ⎛ k1⎞ RT ⎜ k1 ⎟ TOF =
ln⎜1 ⎟ ⎡ − η − Ecat/2)⎤⎦
Ecat/2 E1/2 + ln⎜1 + ⎟ E1/2 + +
F ⎝ k2 ⎠ F


k 2 ⎜1 +
k2 ⎞
⎟⎟
F Solv
1 + exp⎣ RT (E HA/H
⎝ ⎝ k1 ⎠ ⎠ 2 (5)

The overpotential η is defined as the difference between the


kobs collapses to k rate limiting for both cases, while Ecat/2 applied potential at the electrode (E) and the apparent standard
collapses to E1/2 only when k1 is rate limiting. potential of the reduction of protons to hydrogen in the
conditions of acid and solvent studied ESolv HA/H2: η = EHA/H2 − E.
Solv
2.2.2. Foot-of-the-Wave Analysis (FOWA) for Kinetic
Evaluation. Experimentally, S-shaped cyclic voltammograms The catalytic Tafel plots are generated using TOFmax, Ecat/2,
and consequently accurate plateau currents may not be accessible HA/H2 at a fixed stated acid concentration. Determination of
and ESolv
due to substrate consumption, catalyst degradation, or other side the value of ESolv
HA/H2 is discussed in section 2.3.1. Savéant and
phenomena.27,32 This prohibits the analysis of the plateau
current to extract kinetic information. As the influence of these Artero suggest using the standard concentration of 1 M to
side reactions increase throughout a cyclic voltammetry generate the catalytic Tafel plots. Practically, researchers
experiment, the plateau current becomes inaccessible. Costentin experimentally determine the rate constants for the trans-
and Savéant recently recognized that in these cases, the current formations studied using the methods described above, and they
(i) vs applied potential (E) plot of nonideal catalytic cyclic generate catalytic Tafel plots by calculating at each overpotential
voltammograms overlays with the idealized cyclic voltammogram the expected turnover frequency. The aim of this tool is to
near the beginning of the catalytic scan.27 Consequently, kinetic provide a simple visual representation of the activity of different
information can still be gleaned using data recorded near the foot catalysts at different overpotentials.
of the wave where the influence of such phenomena is minimized Recent concerns have emerged over the usefulness of this tool
in a method coined foot-of-the-wave analysis (FOWA).27 The as the activity reported is necessarily idealized. One such concern
generic expression of the current near the foot of the wave, was over the actual acid concentration in the medium, as well as
assuming fast electron transfers occurring only at the electrode, is turning on catalyst-dependent side reactions and deactivation
pathways at such high hypothetical acid concentrations.39
i i pl /i p0 2.3. Considerations for Choosing an Acid Source. The
=
1 + exp⎡⎣ RT (E − Ecat/2)⎤⎦
i p0 F nature and choice of acid source is of paramount importance in
(4) PCET processes. Several key parameters should be considered
where the specific definition of ipl and Ecat/2 varies with when choosing an acid, including acid pKa, structure, standard
mechanism (see eq 1 and Table 1).28,32 FOWA has rapidly reduction potential for H2 production, potential of background
proven extremely valuable with successful application to the reduction at the electrode, and solution thermal equilibrium
study of a number of electrocatalytic PCET reactions including processes. This section will briefly touch on the influence of each
H2 evolution, CO2 reduction, and water oxidation.24,27,39,58−63 parameter, focusing on practical guidelines to be mindful of
Recently, the influence of the heterogeneous electron transfer during PCET studies. A case study of the impact of these
rate and the percentage of the current response used in FOWA parameters on reactivity is presented in section 3.3.
on the kinetic results were reported.39 It was found that utilizing 2.3.1. Acid pKa, Thermodynamic Potential for Reduction,
larger percentages of the catalytic wave for FOWA resulted in and Structure. Acid strength, quantified through pKa, is the
larger errors; additionally, the linear portion of the wave parameter most commonly used in the choice of acid for the
decreased with smaller (and more realistic) values of the study of PCET reactions. Extensive tabulation of acid pKa values
heterogeneous electron transfer rate constant. This study in acetonitrile is available; this existing data spans a wide range of
emphasizes the importance of only using the linear portion of acid types, including phenols, carboxylic acids, anilinium and
the data for FOWA. pyridinium derivatives, among others.64−67
Finally, obtaining accurate values from FOWA is only possible The pKa of the chosen acid directly relates to the overpotential
when an accurate value is known for the potential of the catalysis- for the specific electrochemical PCET reaction. The required
initiating redox process (E1/2). The potential of an electro- overpotential for a catalytic reaction has implications for the
chemical event in the absence of substrate is often used, but care comparison of catalysts as discussed in the context of catalytic
must be taken to ensure that this redox process is indeed the Tafel plots and for analyzing catalytic thermochemical cycles. For
initiating event for catalysis. This can be difficult to obtain in the example of hydrogen evolution, evaluation of the over-
some cases, especially when the solvent is the substrate. potential η as a function of the applied potential E, necessitates
2.2.3. Catalytic Tafel Plots and Standard States. To knowledge of two parameters: the standard reduction potential
benchmark catalysts studied in different conditions, in the of H+ to H2 in the solvent (E0,Solv
H+/H2) and the pKa of the acid (HA)
presence of different acids and at different overpotentials, in the solvent (Solv):68
Savéant and co-workers proposed the use of catalytic Tafel plots.
These plots, of the log of the turnover frequency (TOF) as a Solv
η = E HA/H − E = E H0,Solv − 0.059 × pK a Solv − E
+
function of overpotential (η) (the log plot of eq 5), showcase the 2 /H 2 (6)

3647 DOI: 10.1021/acscatal.6b00778


ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

Recently, a methodology was outlined for the direct depending on the acid structure, concentration, and the nature of
electrochemical determination of ESolv HA/H2 via measurement of the solvent.
the open circuit potential of an electrochemical cell containing homoconjugation: BH+ + B → BHB+ (7)
the respective acid and 1 atm of H2 gas at equilibrium in the
solvent studied.69 To date, the experimental complexity of the dimerization: 2BH → (BH)2 (8)
measurement has limited its use.
A recent report investigating the kinetics of the CPET reaction heteroconjugation: BH+ + H 2O → BH3O+ (9)
between acids and gold electrodes in acetonitrile shed light on
the influence of acid structure in terms of steric hindrance.70 Homoconjugation is particularly strong for phenols and
Faster kinetics of the CPET process to the electrode were carboxylic acids. For example, phenolate is reported to have a
measured for triethylammonium compared to diisopropylethy- homoconjugation constant close to 16 000 M−1, while benzoate
lammonium. This effect was attributed to the increased Tolman has a constant of almost 4000 M−1.64 By contrast, aniline has a
cone angle of 200° for the diisopropylethylammonium cation much lower homoconjugation constants of 4 M−1;64 it is often
compared to 150° for triethylammonium. assumed that most aniline derivatives have low homoconjugation
The possibility of interactions between catalysts and conjugate constants.72 As was recently demonstrated in the study of
bases is another parameter to be considered in the choice of acid/ [Ni(PR2 NR2 ′)2]2+ hydrogen evolution catalysts, failure to account
base pairs for mechanistic studies. The influence of base structure for homoconjugation when determining rate constants from
has recently been highlighted in our study of catalytic proton cyclic voltammetry data can lead to misinterpretations of the
reduction by cobaloxime complexes with substituted aniliniums influence of added base as well as errors in the determination of
as an acid source (see below). It has been shown that during the rate constants. However, careful consideration of homoconjuga-
course of catalysis, as conjugate base aniline is generated, the tion can permit accurate determination of kobs and mechanistic
aniline molecules can bind to the cobalt center, influencing what assignment.39
the active catalyst is and so altering the cyclic voltammetry Acid dimerization is another process to consider, especially
responses.71 This effect is readily apparent in the cyclic when the acid dimer has a lower pKa than the parent acid. The
voltammetry of Co(dmgBF2)2(CH3CN)2 (dmgBF2 = difluor- presence of an acidic species with a different pKa than the parent
oboryldimethylglyoxime) with and without added 4-MeO- acid may result in multiple PCET mechanisms, complicating
aniline (Figure 3). Upon base addition, the CoIII/II wave kinetic and mechanistic analysis. Dimerization can occur in
aprotic organic solvents when interaction between two acid
molecules yields a stabilization greater than solvation by the
aprotic organic solvent. Dimerization has been found to
influence the reactivity of carboxylic acids in acetonitrile.73
Heteroconjugation arises between acids and bases of a
different species; for example, the interaction of a carboxylic
acid with water. In the context of electrochemical studies in
acetonitrile, the effect of heteroconjugation has only been
considered when water is added to the organic solvent.73
2.3.3. Reduction of Acids at Electrode Surfaces. For
solutions containing protic species, an important control
experiment to keep in mind is at what potential these protic
species are directly reduced at the electrode. In other words,
control experiments for each individual component of an
electrochemical system are crucial. Glassy carbon electrodes
have long been used in the study of fuel-forming catalysts,
especially hydrogen evolution, due to their low background
activity for direct, heterogeneous acid reduction. While the
background activity is less significant than on other electrode
materials such as platinum, direct acid reduction at glassy carbon
electrodes still occurs. Our group recently published a systematic
Figure 3. Cyclic voltammograms recorded at 100 mV/s in 0.25 M study of the direct reduction of 20 acids in acetonitrile on glassy
[Bu4N][PF6] of 0.38 mM of Co(dmgBF2)2(CH3CN)2 in the absence carbon electrodes.74 These acids spanned over 35 pKa units in
(a) and presence (b) of 2 mM 4-methoxyaniline. Reprinted with acetonitrile, with the experimental direct acid reduction
permission from reference 71. Copyright 2016 American Chemical potential, reported as the current inflection potential (Einf),
Society. covering a range of over 2 V.
These results are visually summarized in Figure 4, where the
reversibility and potential is modified, without apparent changes color intensity reflects the current intensity observed as a
to the CoII/I wave. Separate UV−vis titration experiments percentage of the current observed at Einf. This study
demonstrated that one and two aniline molecules can reversibly underscored four factors influencing the observed voltammo-
coordinate to the catalyst.71 grams: electrode fouling, irreproducibility from scan to scan,
2.3.2. Acid Homoconjugation, Heteroconjugation, and curve crossing, and acids with erratic backgrounds. The influence
Dimerization. Acid and base molecules are well-known to of adding water to each acid was also studied. While no significant
interact in solution via the processes of homoconjugation, peak shifts were observed with water addition, the current
heteroconjugation, and dimerization. These three processes, increased for a number of acids known to have a high
depicted in eqs 7 to 9, have a varying degree of importance homoconjugation constants. The current increase was proposed
3648 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

As the number of reports grows, some key structural motifs


have emerged as susceptible to decomposition. Cobalt and nickel
compounds containing CN and N−O bonds were reported by
a number of groups80,81,83,84,89 to decompose in acidic and
reducing conditions to Co- and Ni-containing nanoparticles,
respectively. Firm experimental evidence that CN bonds can
predispose a complex to decomposition was provided by our
group via direct comparison of two complexes either with or
without CN bonds.87
The presence of Ni−S bonds also appears to render
compounds lacking CN or N−O bonds susceptible to
decompositions, as discussed by Roberts85 and by our group.11
In these cases, it appears that cleavage of adjacent S−C bonds
with concomitant ligand destructionis favored by Ni−S
bonds.11,85 Finally, we recently demonstrated with an Fe
complex that protic conditions are not required for decom-
Figure 4. Acid reduction potentials in acetonitrile (25 mM acid, 100 position.88
mV/s) depicting the range between the direct reduction potential Einf. 2.5. Digital Simulations. The modeling of electroanalytical
and the approximate thermodynamic potential for hydrogen evolution. experiments has advanced to a level of great accuracy over the last
Caveats for three acids are listed. Adapted with permission from ref 74.
Copyright 2014 American Chemical Society.
half century. Two general approaches have been used:
(1) The semianalytical solutions to the diffusion equations
have been obtained through the integral equation method
in order to describe the voltammetric response of specific
mechanisms (e.g., equations presented in section 2.2.1).
These solutions are often applied directly or used to
generate working curves based on dimensionless param-
eters.26
(2) Finite difference or finite element numerical methods have
been used to solve the diffusion equations in conjunction
with the kinetic equations describing the chemical reaction
of interest (known as digital simulation).
Both methods are very valuable, but digital simulation is
generally found to be more accessible to most researchers.
Digital simulation software can be used to generate simulated
voltammograms that can be directly compared with experimental
voltammograms. However, successful simulation of experimental
data requires accurate assessment of a large number of
parameters, and the sensitivity of electroanalytical techniques
to environmental factors that cannot be emulated in simulation
often prevents satisfactory fitting. In our experience, we have
found digital simulation to be most valuable in addressing “what
if” questions and to test if the already derived equations can be
extended to a proposed mechanism. Both of these cases are
highlighted below in our mechanistic studies of cobaloxime.71
Figure 5. Kinetic zone diagram and simulated CV waveforms for the 2.6. Traversing the Zone Diagram To Access S-Shaped
catalytic EC′ mechanism of the one electron reduction of a substrate via Voltammograms. Zone diagrams pictorially depict the
a redox catalyst mediator. Reprinted with permission from ref 32.
voltammetric waveform shapes expected as a function of
Copyright 2014 American Chemical Society.
dimensionless parameters;91 the zone diagram for an EC′
mechanism is shown in Figure 5. For an EC′ reaction, the kinetic
(λ) and excess factors (γ) describe how waveforms will evolve as
to be the result of the added water stabilizing the in situ generated a function of parameters such as ke (the rate constant for the
conjugate bases and so freeing additional acid molecules tied up reaction of the reduced catalyst and the substrate), υ (the scan
in homoconjugation for direct electrode reduction. Overall, this rate), and [Cat]0 and [CA]0 (the bulk concentration of catalyst
study allows for easier and mindful selection of a suitable acid for and substrate A, respectively):26,91
electrochemical PCET reactions.
2.4. Homogeneous to Heterogeneous Transformation ⎛ RT ⎞⎛ ke[Cat]0 ⎞
λ= ⎜ ⎟⎜ ⎟
of Catalysts. The past decade has witnessed increased ⎝ F ⎠⎝ υ ⎠ (10)
awareness of the possibility of transformation from a
homogeneous precatalyst to heterogeneous catalyst.75−78 [CA ]0
There has been a recent surge in the identification of molecular γ=
[Cat]0 (11)
species which decompose to heterogeneous material deposited
on the electrode surface; known examples are depicted in Table By changing these experimental parameters, waveforms
2.11,79−88 conforming to different zones can be accessed.32
3649 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

Table 2. Metal Complexes Known to Electrochemically Degrade in Organic Solventsa under Catalytic Conditions To Form
Electrode-Adsorbed Heterogeneous Materialsb

a
All examples were in either acetonitrile or dimethylformamide. bThis table does not include examples where the homogeneous catalyst only
degrades after harsh prolonged catalysis; e.g., ref 90.

Until recently, it was not clear if the EC′ Zone Diagram could Diagram could be qualitatively extended to multielectron,
be directly extended to the more complex multielectron, multisubstrate reactions.92 In addition to υ, [CA]0, and [Cat]0,
multisubstrate reactions associated with fuel production.92 the pKa of added acids could also be used to traverse the zone
However, voltammograms have now been reported for the diagram as it was found that the observed rate constant is linearly
catalytic reduction of protons from a series of eight para- dependent on acid strength.71,92
substituted aniliniums to hydrogen by the catalyst Co- Using the methods detailed above in section 2.2.1, global rate
(dmgBF2)2(CH3CN)2. Notably, the voltammetric responses constants can be extracted from the plateau region of S-shaped
recorded for varying kinetic and excess factors apparently voltammograms of Zone KS (and the related Zone KD) for both
spanned all of the waveform shapes associated with the EC′ Zone EC′ and various multielectron, multisubstrate reactions, such as
Diagram. Detailed analysis confirmed that the EC′ Zone the ECEC pathway.26,28 Further, it has long been known that
3650 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

kinetic information can also be extracted from waveforms in the the [Co(dmgBF2)2(CH3CN)2]0/− couple. Notably, the Ecat/2
“total catalysis” region (Zone KT2) for a simple EC′ (the half-wave potential of the sigmoidal response) shifts to
reaction.32,91 The catalytic peak potential of a total catalysis values positive of E°′([Co(dmgBF2)2(CH3CN)2]0/−). For acids
voltammogram varies as a function of ke, υ, [CA]0, and [Cat]0 (see of pKa > ca. 9.5, the catalytic response currents approach a
section 4.1).32 More recently, peak shift analysis has been plateau shape as the acid concentration is increased; these
extended to the ECEC mechanism and demonstrated, under catalytic plateau currents (ipl) were found to be first-order in both
limiting conditions, to provide information about the rate acid concentration and catalyst concentration, consistent with
constant k1.71 Ultimately, the primary utility of a multielectron, prior results.97
multisubstrate zone diagram is its use as a navigational tool to These three observations provide evidence to support
modify experimental conditions in order to enter either the zones assignment of the reaction mechanism (Figure 6A). Specifically,
in which kinetic information can be extracted. the shifting Ecat/2 indicates that the first chemical step (C1,
protonation of [Co(dmgBF2)2(CH3CN)2]− to form a putative
3. CASE STUDIES: REACTION MECHANISMS AND CoIII-hydride intermediate) is rapid, and the subsequent
KINETICS OF H2-EVOLVING CATALYSTS chemical step(s) is(are) rate-limiting. Previously, two independ-
The electrochemical and spectroscopic methods discussed in ent theoretical124,126 and one experimental125 study have shown
section 2 have been applied in concert to elucidate the reaction that the CoIII-hydride intermediate is more readily reduced than
pathways of catalysts for hydrogen evolution. Recent work has the parent CoII species, indicating that this intermediate will be
focused on two highly active catalysts, Co(dmgBF2)2(CH3CN)2 spontaneously reduced. Further, based on the influence of acid
(dmgBF 2 = difluoroboryldimethylglyoxime) and [Ni- and catalyst concentration on the catalytic plateau current, the
(P2PhN2Ph)2]2+ (P2PhN2Ph = 1,3,5,7-tetraphenyl-1,5-diaza-3,7-di- rate-limiting step for catalysis is also first-order in both acid and
phosphacyclooctane) (Scheme 2). In addition to identifying catalyst, indicating a second (rate-limiting) protonation step
(C2) and an overall heterolytic route for hydrogen production.
Scheme 2. Co(dmgBF2)2(CH3CN)2 (left) and Together, these data support a heterolytic ECEC pathway for H2
[Ni(PPh Ph 2+
2 N2 )2] (right) evolution from anilinium acids in acetonitrile. Notably, this
assignment is consistent with that from theoretical work54,124 as
well as conclusions from photocatalysis and laser flash photolysis
experiments.112,125
Of note, a homolytic route involving the bimetallic reaction of
two CoIII-hydride species has previously been recognized as a
viable pathway for H2 production by both computational
studies124,126 and thermochemical analysis.54 While clear
evidence diagnosing a homolytic pathway for hydrogen
production has not yet been reported, 29 recent work
dominant PCET reaction pathways, these studies have yielded investigating the cleavage of H2 by Co(dmgBF2)2(CH3CN)2
new insight into the kinetics of elementary reaction steps and under high H2 pressures suggests this reaction proceeds via a
revealed linear free energy relationships between rate constants
homolytic cleavage route.127−129 As such, a homolytic reaction
for these individual steps and acid pKa. The complexities of
pathway for hydrogen production may be accessible under
proton source reactivity in acetonitrile have also been established
alternative experimental conditions, or could be occurring as a
through the investigation of reaction kinetics as a function of
minor parallel pathway.
various acid parameters.
With this mechanistic assignment, kinetic details for
3.1. Mechanistic Study of Electrocatalytic H2 Evolution
by a Cobaloxime. Cobaloximes and their derivatives have been elementary steps can be elucidated through analysis of the
incorporated as catalysts in both electrochemical and photo- voltammograms. The second-order global rate constants for
chemical H2 production cycles.93−123 While their operating catalysis were determined from the analysis of the acid
mechanisms have been explored in depth under certain concentration-dependent plateau currents for acids with pKa >
conditions using both experimental and theoretical meth- ca. 9.5 (Figure 6B). As the plateau current reflects the observed
ods,54,93,94,96−98,124−126 electroanalytical methods were recently rate constant, kinetic details of the rate-limiting step can be
applied to explore the reaction pathways and kinetics of gleaned directly. As described above, the second protonation is
Co(dmgBF2)2(CH3CN)2 (dmgBF2 = difluoroboryldimethyl- the rate-limiting step under these conditions and described by
glyoxime) as a function of acid strength.71,92 Through the use rate constant k2. Across the series of aniliniums with pKa > ca. 9.5,
of a series of analogous acids, specifically para-substituted catalysis was found to accelerate as the strength of acid increased
aniliniums, complicating factors resulting from different and a linear correlation is observed between log(k2) and acid pKa
homoconjugation constants across dissimilar acids were reduced, (slope = −0.77). For acids with pKa < ca. 9.5, catalytic
allowing studies to focus on the direct effect of acid pKa on voltammograms remained peaked and the catalytic responses
reaction kinetics and mechanism. It should be noted that were determined to be controlled by diffusion of the acid to the
Co(dmgBF2)2(CH3CN)2 decomposes in the presence of strong reaction layer. As such, second-order rate constants could not be
acids; in the extreme case of strong acid and reducing conditions determined for these acids; however, extrapolation of the linear
cobalt-containing nanoparticles form, as noted in Table 2 above. relationship determined for weaker acids allows these rate
In our studies of Co(dmgBF2)2(CH3CN)2, great care was taken constants to be estimated.
to limit the extent of decomposition.71 Rate constants for the first protonation step (k1) were
Cyclic voltammograms of Co(dmgBF2)2(CH3CN)2 recorded determined from foot-of-the-wave (FOWA) analysis. Like k2, a
in acetonitrile in the presence of para-substituted aniliniums linear correlation between log(k1) and pKa was observed (slope =
exhibit an increase in cathodic current and loss of reversibility at −0.97). Values for k1 are ca. 3 orders of magnitude greater than
3651 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

determined from this limiting current value (ca. 125 s−1) likely
reflects either an H−H bond formation step or H2 release.
Kinetic and mechanistic insight gained from comprehensive
analysis of the catalytic voltammograms contributes to a more
detailed picture of hydrogen evolution from cobaloximes.
Specifically, it was learned that upon reduction of the CoII
species, rapid protonation (k1) yields a CoIII-hydride species
(that may undergo tautomerization128−130). This intermediate is
readily reduced to form a CoII-hydride (k2). The subsequent
second protonation step is 3 orders of magnitude slower than the
first step. The resulting CoII−H2 species releases H2 to close the
catalytic cycle via an acid-independent step with a first-order rate
constant of 125 s−1.
3.2. Mechanistic Study of Electrocatalytic H2 Evolution
by [Ni(P2PhN2Ph)2]2+. [Ni(P2PhN2Ph)2]2+ (P2PhN2Ph = 1,3,5,7-
tetraphenyl-1,5-diaza-3,7-diphosphacyclooctane) and its family
of hydrogen production/oxidation catalysts have been thor-
oughly studied over the past decade, with the notable result of
clearly demonstrating the importance of the secondary
coordination sphere in catalytic reactions.131−136 While original
reports assumed a single catalytic mechanism, the possibility of
multiple catalytic routes has been recently examined.24,39,137
Two generic pathways have been identified: an ECEC pathway,
whereby the first protonation occurs on the NiI species after a
single reduction, and an EECC pathway, where double reduction
to form a Ni0 species occurs prior to both protonations.
Two recent experimental studies probing these parallel
reaction pathways approached the problem very differently.
Wiedner and co-workers examined the reactivity of [Ni-
(PPh Ph
2 N2 )2]
2+
using dimethylformamidium (DMFH+) as the
proton source.39 DMFH+ is a relatively strong acid in acetonitrile
with a pKa of 6.1, resulting in cyclic voltammograms with both
the ECEC and EECC pathways readily apparent. Using a mixture
of electrochemical techniques, primarily based around FOWA,
the overall contributions of the two different pathways were
deconvoluted. In work in our lab, we utilized a weaker acid
(anilinium, pKa = 10.6) to isolate the EECC mechanism and then
used several of the previously discussed methods to determine
Figure 6. (A) Reaction mechanism for hydrogen evolution catalyzed by the individual rate constants of the elementary reaction steps.24
Co(dmgBF2)2(CH3CN)2 as supported by detailed analysis of catalytic 3.2.1. Electrochemical Studies To Elucidate Kinetics of Ni0
CVs. (B) Rate constants k1 and k2 determined from FOWA and plateau Protonation and Subsequent Reactivity. While analysis of
current analysis, respectively, of acid concentration-dependent catalytic
voltammograms of Co(dmgBF2)2(CH3CN)2 and a series of para-
cyclic voltammograms under pseudo-first-order conditions is
substituted aniliniums. k2 values for acids 6, 7, and 8 were extrapolated useful for extracting kinetic information, experiments carried out
from the linear relationship established at higher pKa values. kΩ under more stoichiometric catalyst:substrate conditions can
represents an acid-independent step and was determined from catalytic provide additional information on how the reaction proceeds. In
voltammograms recorded at high acid concentrations in which the the analysis of [Ni(PPh Ph 2+
2 N2 )2] , stoichiometric addition of
voltammetric responses were acid concentration and pKa-independent. anilinium confirmed rapid reactivity of acid with the Ni0 species
Acids identity: (1) 4-methoxyanilinium, (2) 4-t-butylanilinium, (3) by a signature kinetic peak shift (discussed further in section 4.2)
anilinium, (4) 4-chloroanilinium, (5) 4-trifluoromethoxyanilinium, (6) and the appearance of a new oxidation feature on the return trace.
4-(methylbenzoate)anilinium, (7) 4-trifluoromethylanilinium, and (8) Given that two electrons and a single proton had been
4-cyanoanilinium. Adapted with permission from ref 71. Copyright 2016
transferred to the catalyst, it was natural to assume hydride
American Chemical Society.
formation. Independent synthesis and cyclic voltammetric
analysis of a NiII-hydride confirmed that the new oxidation
those of k2, consistent with the interpretation of an ECEC feature observed did correspond to the hydride species (Figure
pathway with a rate-limiting second chemical step. 7).24
At high acid concentrations, the plateau currents reach a With the first chemical step established as hydride formation,
limiting value that is acid concentration and pKa-independent. we next examined the details of this elementary step under
This observation extends to both the weaker acids and the catalytic conditions. First, a rate constant was determined by
stronger acids described above which exhibit only peak-shaped analyzing the kinetic peak shift of the NiI/0 wave in the presence
voltammograms in their acid-dependent region. Across the range of acid. Conveniently, catalytic turnover could be prevented on
of acids explored, this limiting plateau current is constant, though the electrochemical time scale, even under pseudo-first-order
the concentration of the acid required to reach this limiting value conditions, by the addition of excess base. This rate constant was
increases as acid strength decreases. The first-order rate constant corroborated with FOWA. FOWA uniquely provides the rate
3652 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

chemical step of H2 productionshould control the overall


reaction kinetics. If this is the case, the plateau currents should
directly reflect this reactivity and spectroscopic monitoring of the
reaction of NiII-hydride with anilinium should result in a similar
observed rate constant.
The NiII-hydride is a stable complex with a distinct absorption
spectrum from [Ni(PPh Ph 2+
2 N2 )2] , allowing us to observe its
reactivity with anilinium directly, using stopped-flow spectros-
copy. As stopped-flow spectroscopy is a time-resolved technique
compared to the steady-state nature of electrochemical methods,
complementary chemical reactivity information can be gathered
by employing this experimental technique. Reaction of the NiII-
Figure 7. Cyclic voltammograms of 1.5 mM [Ni(PPh Ph 2+ hydride with anilinium affords a kinetic trace with two distinct
2 N2 ) 2] and
approximately 1 equiv of anilinium (blue) compared with the cyclic regions: a fast component on the order of seconds, and a slow
voltammogram of 0.6 mM of the independently synthesized NiII- component on the order of minutes (Figure 9). Under pseudo-
hydride (black) in CH3CN. The new oxidation observed at −0.4 V for
[Ni(P2PhN2Ph)2]2+ in the presence of acid matches that of the
independently synthesized hydride. Voltammograms recorded at 100
mV/s in 200 mM [NBu4][PF6] CH3CN solutions. Reprinted with
permission from ref 24. Copyright 2015 American Chemical Society.

constant for only the first step in the catalytic reaction.


Reasonable agreement (1.2 × 106 M−1 s−1 for peak shift analysis,
and 6.5 × 106 M−1 s−1 for FOWA) was found between the
resulting values.24
Subsequent reactivity of the NiII-hydride was then inves-
tigated. Comparison of the value estimated for kobs from the
catalytic plateau currents and the value determined for rate
constant k1 indicates that the overall rate of catalysis is not
governed by hydride formation. As such, kinetic information
obtained from the plateau current represents steps subsequent to
the first protonation. At low acid concentrations the reaction is Figure 9. Stopped-flow kinetics trace for a solution 0.39 mM NiII-
hydride and 100 mM anilinium. 88% of the change in absorbance is
approximately first-order in acid; however, at higher acid
accounted for in the “fast” kinetics (shown in inset), which occurs over 2
concentrations, an acid concentration-independent region is s followed by a “slow” process occurring over the course of 5 min.
reached (with anilinium the maximum rate constant was ∼20 s−1
as measured by the ECEC plateau equation, Figure 8).24
3.2.2. Investigation of Hydride Reactivity with Acid via first-order conditions, the fast component was fit to single
Stopped-Flow Spectroscopy. The cyclic voltammetry studies exponential kinetics. The rate constant obtained was found to be
described above establish that the formation of the hydride is first-order with respect to the anilinium concentration.24
rapid and these kinetics do not affect the overall catalytic rate At first glance, the first-order response with anilinium
described by kobs. With this assumption, it was asserted that the appeared to be in conflict with the acid-independent response
reactivity of the hydride with the proton sourcethe second obtained from cyclic voltammetry at higher acid concentrations.

Figure 8. Observed rate constants obtained from the current plateau (blue circles) for the reaction of [Ni(PPh Ph 2+
2 N2 )2] and anilinium, and from stopped-
flow kinetics traces (red circles) for the reaction of NiII-hydride and anilinium. Black circles represent the fit for the mechanism shown in the right panel,
and the appropriate kinetics models are shown for the time-resolved (red) and steady-state (blue) kinetics. Data from ref 24.

3653 DOI: 10.1021/acscatal.6b00778


ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

However, the discrepancy was reconciled by considering the Changing the proton source structure significantly affects its
possibility of an off-cycle intermediate, which has been reactivity (as will be shown in the following sections).
extensively explored in several previous studies.39,136,138−142 The effect of added base on reaction kinetics was explored.
These works reveal that the [Ni(PR2 NR2 ′)2]2+ molecules can be When base is added to solution, the reaction kinetics were
protonated in an “exo” position on the pendant amine which observed to slow. A trend between amount of base necessary to
prohibits catalytic turnover. This effectively inactivates the slow the reaction and acid pKa was observed. The higher the pKa,
catalyst prior to catalytic turnover. The “exo” position proton the less base required to lower the observed reaction rate (when
must be removed, followed by protonation in the “endo” position pKa is greater than 8.6, reactivity of acids below a pKa of 8.6 do
for productive catalysis to occur. Under the conditions accessed not have a pKa dependence as illustrated in Figure 10).
3.3.2. Molecular Association. Two different forms of
in cyclic voltammetry, an increase in the acid concentration
molecular association were examined, homoconjugation and
results in an increase in the concentration of the inactive catalyst,
dimerization (aggregation and heteroconjugation were also
reducing the apparent rate constant and giving the appearance of
considered for studies involving water). The list of proton
acid concentration independence. The time-resolved (stopped-
sources with recorded homoconjugation constants in the
flow) kinetics do not exhibit acid independence because no
working range of [Ni(PPh Ph
2 N2 )2]
2+
(pKa < 12) is limited, but
equilibrium can be established between the inactive and active
the effect could be examined through the comparison of 4-
forms of the catalyst. However, the generation of the inactive
cyanoanilinium (pKa = 7, log(KHC) = 0.1),73 p-toluenesulfonic
catalyst affects the kinetics analysis of the fast component and the
acid (pKa = 8.6, log(KHC) = 3),74 and dimethylformamidium
slow component observed corresponds conversion of the
(pKa = 8.6, log(KHC) = 1.7).74 While this series of acids had only
inactive species to the active species followed by hydrogen
small differences in their pKa values, the reaction rate constants
release.24 were found to be highly dependent on the homoconjugation
3.3. Reactivity of NiIIH with Various Proton Sources. As constants, with fast reaction rate constants observed for the acids
the reaction of the NiII-hydride with acid was established to be with larger homoconjugation constants (p-toluenesulfonic acid
the rate-limiting step for hydrogen production, this elementary (130 000 M−1 s−1) > dimethylformamidium (12 750 M−1 s−1) >
reaction step provided a convenient platform to take a deeper p-cyanoanilinium (750 M−1 s−1 )).73 In past work, the
look at the influence of the proton source on reaction acceleration of acid−base reactivity for acids with large
mechanisms in acetonitrile. Reactivity of the NiII-hydride with homoconjugation constants has been attributed to a lowering
a series of acids were examined using the same stopped-flow of the “effective” pKa as a result of stabilization of generated
procedure described for the reaction with anilinium (section conjugate base via homoconjugation with acid.69,72,136,143−145
3.2.2). The effect of various parameters, including pKa, presence However, rate constants for this series span more than 2 orders of
of base, molecular association, and interaction with water, were magnitude, and this span is too large to be accounted for simply
explored.73 by homoconjugation. Likely, larger homoconjugation constant
3.3.1. Acid pKa. As discussed above, pKa is the most commonly correlate with increased hydrogen bonding abilities, and these
considered parameter for acids in PCET studies. In the reaction acids may better stabilize the proton transfer transition state.
of NiII−hydride with acid, the anticipated linear free energy To our knowledge, proton source dimerization has yet to be
relationship is obtained between the log of the reaction rate and considered in the PCET fuel forming literature. While there are
the acid pKa (Figure 10). With a sufficiently strong driving force no recorded dimerization constants for proton sources with a pKa
for proton transfer the reaction rate constant becomes substrate less than 12, carboxylic acids are known to dimerize in
pKa independent (pKa < 8.6). The data in Figure 10 is limited to a acetonitrile and this generally results in a pKa for the dimerized
series of structurally related para-substituted aniline derivatives. species that is ∼3.5 units lower than the monomer.64 No
dimerization constants are recorded for trifluoroacetic acid (pKa
= 12.6) and trichloroacetic acid (pKa = 10.8). However, the
reaction of trifluoroacetic and trichloroacetic acid with the NiII-
hydride is second-order in acid, as would be expected for a highly
reactive acid dimer with a very low dimerization constant. This is
an important finding as trifluoroacetic acid is often used in
electrocatalytic studies; indeed, many studies report second-
order kinetics with trifluoroacetic acid.136,146−151
3.3.3. Addition of Water to Acetonitrile Solutions. Because
the target for most H2 evolution electrocatalytic systems is to
operate in water, water is often intentionally added to acetonitrile
solutions during catalysis to probe catalyst water stability and
performance.140,152 Consequently, it is important to understand
how the addition of water affects the proton source being
employed. Two cases are considered here: (1) water becomes
protonated and thereby hydronium is available to act as a proton
source, or (2) water heteroconjugates with the proton source,
affecting its reactivity. Hydronium has a very low pKa in
acetonitrile (2.2), but the ability of water to form large aggregates
Figure 10. Second-order rate constants vs acid pKa for the reaction of with concurrent stabilization of the hydronium ion significantly
NiII-hydride with para-substituted aniliniums. The linear region between increases its effective pKa with increasing water concentration.
pKa 12 to 8.6 has a slope of −0.7. Reprinted with permission from ref 73. Figure 11 demonstrates the influence of water on the rate
Copyright 2016 American Chemical Society. constant of reaction with NiII-hydride and acid over a range of
3654 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

4. CASE STUDIES: MODEL SYSTEMS TO INVESTIGATE


ELEMENTARY PCET REACTION STEPS
As detailed above, catalysis involves elementary steps whose
kinetics may be investigated by a variety of spectroscopic and
electrochemical methods. A key question for driving the design
of more energy-efficient catalysts is whether the elementary steps
may be controlled through rational design of catalysts or
experimental conditions. Toward this, efforts in our group have
focused on furthering cyclic voltammetry techniques for the
study of PCET systems.
Shifts in the applied potential at which reactivity is observed
are common in PCET systems (or all catalytic systems for that
matter), yet the reason for potential shifts can differ greatly.
Three separate factors, all dependent upon proton addition, are
presented herein: First, the appearance of total-catalysis
responses;71,92 second, a kinetic shift in the reduction potential
and loss of reversibility;24 and third, a thermodynamic shift in the
reduction potential.11 Distinguishing between these cases is not
Figure 11. Concentration of water versus observed rate constant for always straightforward; we hope that presentation of these three
1.25 mM triflic acid (pKa = 2.6), 1.3 mM dimethylformamidium
case studies can help aid researchers in elucidating the factors
(DMFH+, pKa = 6.1), and 2.5 mM 4-cyanoanilinium (4-CN, pKa = 7).
Observed rate constants obtained from reaction with NiII-hydride to that perturb peak location. Special emphasis is placed on the
form hydrogen. Reprinted with permission from ref 73. Copyright 2016 thermodynamic potential shift, which in this case was found to
American Chemical Society. correspond to a concerted proton electron transfer
4.1. Total Catalysis Gives Rise to a Second Peak. As
noted above, cobaloxime has been reported to yield total catalysis
responses under specific conditions (Figure 12). This occurs
water concentrations for three acids. triflic acid (pKa = 2.6),
dimethylformamidium (pKa = 6.1), and 4-cyanonanilinium (pKa
= 7) were compared; in each case, addition of water resulted in
hydronium formation, which showed a higher reaction rate than
the original proton source. Eventually, further increases in water
concentration began to shut down the reaction, as the excess
water reacts as a base in solution to hinder turnover, as discussed
above.73
Heteroconjugation of water to proton sources with higher pKa
values was investigated through the addition of water to solutions
of trifluoroacetic acid. The reaction rate decreased exponentially
with the concentration of water, suggesting the acid hetero-
Figure 12. Example of a total catalysis waveform for the electrochemical
conjugates to water. This chaotropic activity breaks up the hydrogen evolution with 5 mM catalyst Co(dmgBF2)2(CH3CN)2 and 5
dimerization of trifluoroacetic acid. mM 4-cyanoanilinium in acetonitrile at 100 mV/s with 0.25 M
3.4. Linear Free Energy Relationships in H2 Evolution. [Bu4N][PF6]. All of the acid is consumed near the reaction layer,
For both the Co(dmgBF2)2(CH3CN)2 and [Ni(PPh Ph
2 N2 ) 2 ]
2+
resulting in the appearance of initial catalytic current followed by the
catalysts, linear free energy relationships have been observed CoII/I redox wave.
between rate constants for elementary proton transfer steps and
acid pKa in acetonitrile, with larger rate constants observed for when the kinetics of catalysis are fast enough such that prior to
stronger acids. Plots of log(k) vs pKa for these catalysts have the primary redox wave the small amount of reduced catalyst (as
slopes that span −0.7 to −0.97. The physical interpretation of governed by the Nernst equation) consumes all of the available
these correlations is still under investigation, but the differences substrate to produce product and regenerate the catalyst.
likely reflect the nature of the proton transfer transition state. As Consequently, the normal redox wave of the catalyst is observed
such, differences between the nucleophilic species reacting in after the catalytic peak. Extracting accurate kinetic information
from the catalytic peak position in this case is possible but is
these proton transfer steps are reflected in these relationships.
significantly restricted to a small parameter range; therefore,
As the thermodynamic potential for reduction of an acid to
more versatile methods like foot-of-the-wave analysis are
hydrogen in acetonitrile depends linearly on the pKa of the acid generally preferred.24,71 Assignment of a feature as total catalysis
for structurally comparable acids,69,153 the larger rate constants can be confirmed via observation of the suspected catalytic peak
observed for stronger acids also correspond to larger over- location as a function of acid concentration, catalyst concen-
potentials for catalytic hydrogen production. As such, the linear tration, and scan rate.71
free energy relationships observed provide an experimentally 4.2. Kinetic Control over Peak Location. The mechanism
based visual representation of the catalytic activity in different for the first elementary steps for the [Ni(PPh Ph
2 N2 )2]
2+
catalyst
conditions. These relationships could prove useful in bench- discussed above was probed by cyclic voltammetry. While
marking catalysts, and illustrate the power of detailed kinetic and [Ni(PPh Ph
2 N2 )2]
2+
is a catalyst for hydrogen evolution, catalytic
mechanistic studies. turnover can be shut off by the addition of excess conjugate base,
3655 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

permitting investigation of the first proton transfer event.24 As Scheme 3. NiP2S2


shown in Figure 13, increasing acid concentration results in a

Figure 13. (A) Titration of a solution of 1 mM [Ni(PPh Ph 2+


2 N2 )2] and 1 M
aniline with increasing concentrations of anilinium results in a positive
kinetic shift of the reduction wave. (B) Plotting the shift versus
log[anilinium] gives a line with a slope of 33 mV/decade, consistent with
the theoretical slope of 30 mV/decade for an EC mechanism. Data taken
from ref 24.

positive shift of the NiI/0 reduction event due to the subsequent


chemical step: protonation of the Ni0 species. Plotting log[BH+]
vs the difference in peak potential yields a slope of 33 mV/
decade. This observation supports a stepwise bimolecular EC
reaction with initial Nernstian electron transfer to the Ni species
followed by irreversible reaction with solution acid BH+; the peak
location potential as a function of [BH+] can be described as25,154
RT ⎛ k1[BH ]RT ⎞
+
RT
Epeak = EI0/0 − (0.78) + ln⎜ ⎟
F 2F ⎝ Fυ ⎠ (12)
where E0I/0is the potential of electron transfer in the absence of
BH+, R is the gas constant, F is the Faraday constant, T
Figure 14. Analysis of reactivity of NiP2S2 with triethylammonium in
temperature, k1 is the rate of reaction of BH+ with reduced acetonitrile; data taken from reference.11 (A) Cyclic voltammograms at
catalyst, and υ is the scan rate. Analysis of the resulting shift vs 100 mV/s of solutions of NiP2S2 with and without one equivalent of
[anilinium] permitted evaluation of the protonation rate triethylammonium. (B) Theoretical peak shift analysis based on a
constant as 1.2 × 106 M−1 s−1.24 Importantly, the slope of plots stepwise ECE mechanism compared with real peak shift; current
of log([BH+]) vs the difference in peak potential is predicted by integrations demonstrating two electron and one proton reactivity. (C)
eq 12 to be 30 mV/decadea close match for the experimentally Observation of a kinetic isotope effect upon addition of either 0.24 M
observed slope of 33 mV/decade. Analysis of experimental plots H2O or D2O to a solution of NiP2S2 with substoichiometric
of log([acid]) vs the difference in peak potential provides an triethylammonium.
experimental means of testing if an EC mechanism is operative.
As the rate of the chemical step directly corresponds to the stepwise EEC mechanism was ruled out. Three possible
magnitude of the peak shift, this type of potential shift is often mechanisms were then consideredECE, CEE, or the
referred to as a kinetic shift. concerted (EC)E. No protonation or association between the
4.3. Thermodynamic Control of Peak Location. The last nickel complex and acid was detected by either UV−vis or 1H
example discussed here is that of thermodynamically controlled NMR spectroscopy, suggesting that a CEE mechanism was not
PCET. For a solution of a nickel bisphosphine dithiolate complex operative.
(NiP2S2, Scheme 3) with stoichiometric triethylammonium, A ECE mechanism with an irreversible chemical step could
reduction occurs at potentials much more positive and the result in a large kinetic shift of the reduction event, especially with
magnitude of current passed doubles (Figure 14A,B).11 a fast kinetic step. A kinetic shift was observed for the second case
Assuming no nickel−nickel interactions, two electrons and one study above with [Ni(PPh Ph 2+
2 N2 )2] . Digital simulations of an ECE
proton are added to each nickel molecule in this redox event. As mechanism demonstrated that the magnitude of a reasonable
two-electron transfer is not observed in the absence of acid, the kinetic shift was lower than that observed experimentally (Figure
3656 DOI: 10.1021/acscatal.6b00778
ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

14B). This left the (EC)E mechanism as a plausible candidate (7) Dubois, D. L. Inorg. Chem. 2014, 53 (8), 3935−3960.
where the first proton and electron transfer together at a unique (8) Dempsey, J. L.; Winkler, J. R.; Gray, H. B. Chem. Rev. 2010, 110
potentialin other words, a thermodynamic effect. In the (12), 7024−7039.
literature, evidence for electrochemical CPET has been (9) Warren, J. J.; Tronic, T. A.; Mayer, J. M. Chem. Rev. 2010, 110 (12),
6961−7001.
supported by observation of a kinetic isotope effect where a
(10) Bourrez, M.; Steinmetz, R.; Ott, S.; Gloaguen, F.; Hammarström,
small amount of a protic solvent in the proteo or deutero form L. Nat. Chem. 2015, 7 (2), 140−145.
was added to the solution.20,155−159 Comparison of the cyclic (11) McCarthy, B. D.; Donley, C. L.; Dempsey, J. L. Chem. Sci. 2015, 6
voltammograms of solutions of NiP2S2 with either H2O or D2O (5), 2827−2834.
showed a small but clear KIE (Figure 14C), supporting a (EC)E (12) Dempsey, J. L.; Winkler, J. R.; Gray, H. B. J. Am. Chem. Soc. 2010,
mechanism. It should be noted that while observation of an 132 (47), 16774−16776.
electrochemical KIE supports a concerted mechanism, the lack of (13) Reece, S. Y.; Hodgkiss, J. M.; Stubbe, J.; Nocera, D. G. Philos.
a KIE does not rule out a concerted mechanism. Trans. R. Soc., B 2006, 361 (1472), 1351−1364.
(14) Mayer, J. M. Annu. Rev. Phys. Chem. 2004, 55 (1), 363−390.
5. OUTLOOK: FUTURE DIRECTIONS FOR THE (15) Mayer, J. M.; Rhile, I. J.; Larsen, F. B.; Mader, E. A.; Markle, T. F.;
ELUCIDATION OF PCET PROCESSES IN MOLECULAR DiPasquale, A. G. Photosynth. Res. 2006, 87 (1), 3−20.
(16) Waidmann, C. R.; Miller, A. J. M.; Ng, C.-W. A.; Scheuermann, M.
CATALYSTS
L.; Porter, T. R.; Tronic, T. A.; Mayer, J. M. Energy Environ. Sci. 2012, 5
It is our hope that with increased accessibility to tools for (7), 7771−7780.
studying PCET processes, a greater research effort will be (17) Costentin, C.; Robert, M.; Savéant, J.-M. J. Am. Chem. Soc. 2006,
directed at investigating the PCET pathways involved in the 128, 4552−4553.
catalytic transformations associated with fuel production. Both (18) Costentin, C.; Louault, C.; Robert, M.; Savéant, J.-M. J. Am. Chem.
electrochemical and spectroscopic methods were presented, with Soc. 2008, 130 (47), 15817−15819.
a focus on applications of these methods and potential pitfalls. (19) Costentin, C.; Louault, C.; Robert, M.; Savéant, J.-M. Proc. Natl.
The case studies briefly discussed above illustrate the range of Acad. Sci. U. S. A. 2009, 106 (43), 18143−18148.
(20) Costentin, C.; Robert, M.; Savéant, J.-M.; Tard, C. Angew. Chem.,
PCET mechanisms thus far explored by our group. Observation
Int. Ed. 2010, 49 (22), 3803−3806.
of apparent CPET in the reactivity of NiP2S2 has provided (21) Costentin, C.; Drouet, S.; Robert, M.; Savéant, J.-M. Science 2012,
interesting directions for our future researchcan CPET be 338 (6103), 90−94.
intentionally promoted by varying catalyst electronics or acid (22) Savéant, J.-M.; Tard, C. J. Am. Chem. Soc. 2014, 136 (25), 8907−
strength? As the CPET pathway circumvents high energy 8910.
intermediates, inducing CPET may result in catalysts with (23) Bediako, D. K.; Solis, B. H.; Dogutan, D. K.; Roubelakis, M. M.;
decreased required operating potentials and thus higher Maher, A. G.; Lee, C. H.; Chambers, M. B.; Hammes-Schiffer, S.;
efficiencies. Current work seeks to gain control over PCET Nocera, D. G. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (42), 15001−
reaction mechanisms through experimental parameters. 15006.


(24) Rountree, E. S.; Dempsey, J. L. J. Am. Chem. Soc. 2015, 137 (41),
AUTHOR INFORMATION 13371−13380.
(25) Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Fundamentals
Corresponding Author and Applications, 2nd ed.; John Wiley & Sons, Inc.: Hoboken, 2001.
*E-mail: dempseyj@email.unc.edu. (26) Savéant, J.-M. Elements of Molecular and Biomolecular Electro-
Author Contributions chemistry; John Wiley & Sons, Inc.: Hoboken, 2006.
‡ (27) Costentin, C.; Drouet, S.; Robert, M.; Savéant, J. M. J. Am. Chem.
These authors contributed equally (N.E. and B.D.M).
Funding Soc. 2012, 134 (27), 11235−11242.
(28) Costentin, C.; Savéant, J.-M. ChemElectroChem 2014, 1 (7),
This work was supported by the University of North Carolina at 1226−1236.
Chapel Hill and the David and Lucile Packard Foundation. (29) Costentin, C.; Dridi, H.; Savéant, J.-M. J. Am. Chem. Soc. 2014,
Notes 136 (39), 13727−13734.
The authors declare no competing financial interest. (30) Savéant, J.-M. Chem. Rev. 2008, 108 (7), 2348−2378.

■ ACKNOWLEDGMENTS
This work was supported by the University of North Carolina at
(31) Roubelakis, M. M.; Bediako, D. K.; Dogutan, D. K.; Nocera, D. G.
Energy Environ. Sci. 2012, 5 (7), 7737−7740.
(32) Rountree, E. S.; McCarthy, B. D.; Eisenhart, T. T.; Dempsey, J. L.
Inorg. Chem. 2014, 53 (19), 9983−10002.
Chapel Hill. J.L.D. acknowledges support from a Packard (33) Venkataraman, C.; Soudackov, A. V.; Hammes-Schiffer, S. J. Phys.
Fellowship for Science and Engineering.


Chem. C 2008, 112 (32), 12386−12397.
(34) Ludlow, M. K.; Soudackov, A. V.; Hammes-Schiffer, S. J. Am.
REFERENCES Chem. Soc. 2010, 132, 1234−1235.
(1) Lewis, N. S.; Nocera, D. G. Proc. Natl. Acad. Sci. U. S. A. 2006, 103 (35) Horvath, S.; Fernandez, L. E.; Soudackov, A. V.; Hammes-
(43), 15729−15735. Schiffer, S. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (39), 15663−15668.
(2) McKone, J. R.; Lewis, N. S.; Gray, H. B. Chem. Mater. 2014, 26 (1), (36) Costentin, C.; Robert, M.; Savéant, J.-M. Chem. Soc. Rev. 2013, 42
407−414. (6), 2423−2436.
(3) Cook, T. R.; Dogutan, D. K.; Reece, S. Y.; Surendranath, Y.; Teets, (37) Hinshelwood, C. N.; Green, T. E. J. Chem. Soc. 1926, 129 (9),
T. S.; Nocera, D. G. Chem. Rev. 2010, 110 (11), 6474−6502. 730−739.
(4) Huynh, M. H. V; Meyer, T. J. Chem. Rev. 2007, 107 (11), 5004− (38) Gileadi, E. J. Electroanal. Chem. 2002, 532 (1−2), 181−189.
5064. (39) Wiedner, E. S.; Brown, H. J. S.; Helm, M. L. J. Am. Chem. Soc.
(5) Solis, B. H.; Hammes-Schiffer, S. Inorg. Chem. 2014, 53 (13), 2016, 138, 604−616.
6427−6443. (40) Gagliardi, C. J.; Murphy, C. F.; Binstead, R. A.; Thorp, H. H.;
(6) Weinberg, D. R.; Gagliardi, C. J.; Hull, J. F.; Murphy, C. F.; Kent, C. Meyer, T. J. J. Phys. Chem. C 2015, 119 (13), 7028−7038.
A.; Westlake, B. C.; Paul, A.; Ess, D. H.; McCafferty, D. G.; Meyer, T. J. (41) Wasylenko, D. J.; Rodríguez, C.; Pegis, M. L.; Mayer, J. M. J. Am.
Chem. Rev. 2012, 112 (7), 4016−4093. Chem. Soc. 2014, 136 (36), 12544−12547.

3657 DOI: 10.1021/acscatal.6b00778


ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

(42) Symes, M. D.; Surendranath, Y.; Lutterman, D. A.; Nocera, D. G. (74) McCarthy, B. D.; Martin, D. J.; Rountree, E. S.; Ullman, A. C.;
J. Am. Chem. Soc. 2011, 133 (14), 5174−5177. Dempsey, J. L. Inorg. Chem. 2014, 53 (16), 8350−8361.
(43) Braten, M. N.; Gamelin, D. R.; Mayer, J. M. ACS Nano 2015, 9 (75) Widegren, J. A.; Finke, R. G. J. Mol. Catal. A: Chem. 2003, 198 (1−
(10), 10258−10267. 2), 317−341.
(44) Mandal, S.; Shikano, S.; Yamada, Y.; Lee, Y.-M.; Nam, W.; Llobet, (76) Schley, N. D.; Blakemore, J. D.; Subbaiyan, N. K.; Incarvito, C. D.;
A.; Fukuzumi, S. J. Am. Chem. Soc. 2013, 135 (41), 15294−15297. D'Souza, F.; Crabtree, R. H.; Brudvig, G. W. J. Am. Chem. Soc. 2011, 133
(45) Irebo, T.; Johansson, O.; Hammarström, L. J. Am. Chem. Soc. (27), 10473−10481.
2008, 130 (29), 9194−9195. (77) Artero, V.; Fontecave, M. Chem. Soc. Rev. 2013, 42 (6), 2338−
(46) Sjödin, M.; Ghanem, R.; Polivka, T.; Pan, J.; Styring, S.; Sun, L.; 2356.
Sundström, V.; Hammarström, L. Phys. Chem. Chem. Phys. 2004, 6 (20), (78) Lin, Y.; Finke, R. G. Inorg. Chem. 1994, 33 (22), 4891−4910.
4851−4858. (79) El Ghachtouli, S.; Fournier, M.; Cherdo, S.; Guillot, R.; Charlot,
(47) Zhang, M.-T.; Nilsson, J.; Hammarström, L. Energy Environ. Sci. M.; Anxolabéhère-Mallart, E.; Robert, M.; Aukauloo, A. J. Phys. Chem. C
2012, 5, 7732−7736. 2013, 117 (33), 17073−17077.
(48) Irebo, T.; Zhang, M.-T.; Markle, T. F.; Scott, A. M.; (80) Anxolabéhère-Mallart, E.; Costentin, C.; Fournier, M.; Nowak, S.;
Hammarström, L. J. Am. Chem. Soc. 2012, 134 (39), 16247−16254. Robert, M.; Savéant, J. M. J. Am. Chem. Soc. 2012, 134 (14), 6104−6107.
(49) Zhang, M.-T.; Hammarström, L. J. Am. Chem. Soc. 2011, 133 (23), (81) El Ghachtouli, S.; Guillot, R.; Brisset, F.; Aukauloo, A.
8806−8809. ChemSusChem 2013, 6 (12), 2226−2230.
(50) Irebo, T.; Reece, S. Y.; Sjödin, M.; Nocera, D. G.; Hammarstrom, (82) Wombwell, C.; Reisner, E. Dalton Trans. 2014, 43 (11), 4483−
L. J. Am. Chem. Soc. 2007, 129 (50), 15462−15464. 4493.
(51) Reece, S. Y.; Seyedsayamdost, M. R.; Stubbe, J.; Nocera, D. G. J. (83) Anxolabéhère-Mallart, E.; Costentin, C.; Fournier, M.; Robert, M.
Am. Chem. Soc. 2007, 129 (8), 8500−8509. J. Phys. Chem. C 2014, 118 (25), 13377−13381.
(52) Reece, S. Y.; Seyedsayamdost, M. R.; Stubbe, J.; Nocera, D. G. J. (84) Cherdo, S.; El Ghachtouli, S.; Sircoglou, M.; Brisset, F.; Orio, M.;
Am. Chem. Soc. 2007, 129 (45), 13828−13830. Aukauloo, A. Chem. Commun. 2014, 50 (88), 13514−13516.
(53) Eisenhart, T. T.; Dempsey, J. L. J. Am. Chem. Soc. 2014, 136 (35), (85) Fang, M.; Engelhard, M. H.; Zhu, Z.; Helm, M. L.; Roberts, J. A. S.
12221−12224. ACS Catal. 2014, 4 (1), 90−98.
(54) Dempsey, J. L.; Winkler, J. R.; Gray, H. B. J. Am. Chem. Soc. 2010, (86) Wombwell, C.; Reisner, E. Chem. - Eur. J. 2015, 21 (22), 8096−
132 (3), 1060−1065. 8104.
(55) Rodenberg, A.; Orazietti, M.; Probst, B.; Bachmann, C.; Alberto, (87) Martin, D. J.; McCarthy, B. D.; Donley, C. L.; Dempsey, J. L.
R.; Baldridge, K. K.; Hamm, P. Inorg. Chem. 2015, 54 (2), 646−657. Chem. Commun. 2015, 51, 5290−5293.
(56) Pretali, L.; Doria, F.; Verga, D.; Profumo, A.; Freccero, M. J. Org. (88) Martin, D. J.; McCarthy, B. D.; Piro, N. A.; Dempsey, J. L.
Chem. 2009, 74 (3), 1034−1041. Polyhedron 2016, In Press. DOI: 10.1016/j.poly.2015.12.003.
(57) Tolbert, L. M.; Solntsev, K. M. Acc. Chem. Res. 2002, 35 (1), 19− (89) El Ghachtouli, S. El; Fournier, M.; Cherdo, S.; Guillot, R.;
Charlot, M.-F.; Anxolabéhère-Mallart, E.; Robert, M.; Aukauloo, A. J.
27.
Phys. Chem. C 2013, 117 (33), 17073−17077.
(58) Elgrishi, N.; Chambers, M. B.; Fontecave, M. Chem. Sci. 2015, 6
(90) Chen, L.; Chen, G.; Leung, C. F.; Yiu, S. M.; Ko, C. C.;
(4), 2522−2531.
Anxolabéhère-Mallart, E.; Robert, M.; Lau, T. C. ACS Catal. 2015, 5 (1),
(59) Garrido-Barros, P.; Funes-Ardoiz, I.; Drouet, S.; Benet-Buchholz,
356−364.
J.; Maseras, F.; Llobet, A. J. Am. Chem. Soc. 2015, 137 (21), 6758−6761.
(91) Savéant, J. M.; Su, K. B. J. Electroanal. Chem. Interfacial
(60) Graham, D. J.; Nocera, D. G. Organometallics 2014, 33 (18),
Electrochem. 1984, 171 (1−2), 341−349.
4994−5001.
(92) Martin, D. J.; McCarthy, B. D.; Rountree, E. S.; Dempsey, J. L.
(61) Matheu, R.; Ertem, M. Z.; Benet-Buchholz, J.; Coronado, E.;
Dalton Trans. 2016, Accepted, DOI: 10.1039/C6DT00302H.
Batista, V. S.; Sala, X.; Llobet, A. J. Am. Chem. Soc. 2015, 137 (33), (93) Connolly, P.; Espenson, J. H. Inorg. Chem. 1986, 25 (6), 2684−
10786−10795. 2688.
(62) Costentin, C.; Robert, M.; Savéant, J. M. Acc. Chem. Res. 2015, 48 (94) Dempsey, J. L.; Brunschwig, B. S.; Winkler, J. R.; Gray, H. B. Acc.
(12), 2996−3006. Chem. Res. 2009, 42 (12), 1995−2004.
(63) Mognon, L.; Benet-Buchholz, J.; Llobet, A. Inorg. Chem. 2015, 54 (95) Szajna-Fuller, E.; Bakac, A. Eur. J. Inorg. Chem. 2010, 2488−2494.
(24), 11948−11957. (96) Hu, X.; Cossairt, B. M.; Brunschwig, B. S.; Lewis, N. S.; Peters, J.
(64) Izutsu, K. Acid-Base Dissociation Constants in Dipolar Aprotic C. Chem. Commun. 2005, 1 (37), 4723−4725.
Solvents; IUPAC Chemical Data Series; Blackwell Science: Oxford, U.K., (97) Hu, X.; Brunschwig, B. S.; Peters, J. C. J. Am. Chem. Soc. 2007, 129
1990. (29), 8988−8998.
(65) Kaljurand, I.; Kütt, A.; Sooväli, L.; Rodima, T.; Mäemets, V.; (98) Razavet, M.; Artero, V.; Fontecave, M. Inorg. Chem. 2005, 44 (13),
Leito, I.; Koppel, I. A. J. Org. Chem. 2005, 70 (3), 1019−1028. 4786−4795.
(66) Kütt, A.; Leito, I.; Kaljurand, I.; Sooväli, L.; Vlasov, V. M.; (99) Baffert, C.; Artero, V.; Fontecave, M. Inorg. Chem. 2007, 46 (5),
Yagupolskii, L. M.; Koppel, I. A. J. Org. Chem. 2006, 71 (7), 2829−2838. 1817−1824.
(67) Kütt, A.; Rodima, T.; Saame, J.; Raamat, E.; Mäemets, V.; (100) Valdez, C. N.; Dempsey, J. L.; Brunschwig, B. S.; Winkler, J. R.;
Kaljurand, I.; Koppel, I. A.; Garlyauskayte, R. Y.; Yagupolskii, Y. L.; Gray, H. B. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (39), 15589−15593.
Yagupolskii, L. M.; Bernhardt, E.; Willner, H.; Leito, I. J. Org. Chem. (101) Laga, S. M.; Blakemore, J. D.; Henling, L. M.; Brunschwig, B. S.;
2011, 76 (2), 391−395. Gray, H. B. Inorg. Chem. 2014, 53 (24), 12668−12670.
(68) Artero, V.; Saveant, J.-M. Energy Environ. Sci. 2014, 7, 3808−3814. (102) Andreiadis, E. S.; Jacques, P.-A.; Tran, P. D.; Leyris, A.;
(69) Roberts, J. A. S.; Bullock, R. M. Inorg. Chem. 2013, 52 (7), 3823− Chavarot-Kerlidou, M.; Jousselme, B.; Matheron, M.; Pécaut, J.; Palacin,
3835. S.; Fontecave, M.; Artero, V. Nat. Chem. 2013, 5 (1), 48−53.
(70) Jackson, M. N.; Surendranath, Y. J. Am. Chem. Soc. 2016, 138 (9), (103) Muresan, N. M.; Willkomm, J.; Mersch, D.; Vaynzof, Y.; Reisner,
3228−3234. E. Angew. Chem., Int. Ed. 2012, 51 (51), 12749−12753.
(71) Rountree, E. S.; Martin, D. J.; McCarthy, B. D.; Dempsey, J. L. (104) McCrory, C. C. L.; Uyeda, C.; Peters, J. C. J. Am. Chem. Soc.
ACS Catal. 2016, 6, 3326−3335. 2012, 134 (6), 3164−3170.
(72) Fourmond, V.; Jacques, P.-A.; Fontecave, M.; Artero, V. Inorg. (105) Fihri, A.; Artero, V.; Razavet, M.; Baffert, C.; Leibl, W.;
Chem. 2010, 49 (22), 10338−10347. Fontecave, M. Angew. Chem., Int. Ed. 2008, 47 (3), 564−567.
(73) Rountree, E. S.; Dempsey, J. L. Inorg. Chem 2016, DOI: 10.1021/ (106) Fihri, A.; Artero, V.; Pereira, A.; Fontecave, M. Dalton Trans.
acs.inorgchem.6b00885. 2008, 5567−5569.

3658 DOI: 10.1021/acscatal.6b00778


ACS Catal. 2016, 6, 3644−3659
ACS Catalysis Perspective

(107) Du, P.; Knowles, K.; Eisenberg, R. J. Am. Chem. Soc. 2008, 130 (139) Hoffert, W. A.; Roberts, J. A. S.; Morris Bullock, R.; Helm, M. L.
(38), 12576−12577. Chem. Commun. 2013, 49 (71), 7767−7769.
(108) Wang, M.; Na, Y.; Gorlov, M.; Sun, L. Dalton Trans. 2009, (140) Appel, A. M.; Pool, D. H.; O’Hagan, M.; Shaw, W. J.; Yang, J. Y.;
6458−6467. Rakowski Dubois, M.; Dubois, D. L.; Bullock, R. M. ACS Catal. 2011, 1
(109) Probst, B.; Kolano, C.; Hamm, P.; Alberto, R. Inorg. Chem. 2009, (7), 777−785.
48 (5), 1836−1843. (141) Wiese, S.; Kilgore, U. J.; DuBois, D. L.; Bullock, R. M. ACS Catal.
(110) Li, C.; Wang, M.; Pan, J.; Zhang, P.; Zhang, R.; Sun, L. J. 2012, 2 (5), 720−727.
Organomet. Chem. 2009, 694 (17), 2814−2819. (142) Smith, S. E.; Yang, J. Y.; Dubois, D. L.; Bullock, R. M. Angew.
(111) Lazarides, T.; McCormick, T.; Du, P.; Luo, G.; Lindley, B.; Chem., Int. Ed. 2012, 51 (13), 3152−3155.
Eisenberg, R. J. Am. Chem. Soc. 2009, 131 (26), 9192−9194. (143) Kaupmees, K.; Kaljurand, I.; Leito, I. J. Phys. Chem. A 2010, 114
(112) Du, P.; Schneider, J.; Luo, G.; Brennessel, W. W.; Eisenberg, R. (43), 11788−11793.
Inorg. Chem. 2009, 48 (11), 4952−4962. (144) Wang, M.; Chen, L.; Sun, L. Energy Environ. Sci. 2012, 5, 6763−
(113) Artero, V.; Chavarot-Kerlidou, M.; Fontecave, M. Angew. Chem., 6778.
Int. Ed. 2011, 50 (32), 7238−7266. (145) Donovan, E. S.; Felton, G. A. N. J. Organomet. Chem. 2012, 711,
(114) Krawicz, A.; Cedeno, D.; Moore, G. F. Phys. Chem. Chem. Phys. 25−34.
2014, 16 (30), 15818−15824. (146) Haddad, A. Z.; Kumar, D.; Ouch Sampson, K.; Matzner, A. M.;
(115) Han, Z.; Eisenberg, R. Acc. Chem. Res. 2014, 47 (8), 2537−2544. Mashuta, M. S.; Grapperhaus, C. A. J. Am. Chem. Soc. 2015, 137 (29),
(116) McCormick, T. M.; Calitree, B. D.; Orchard, A.; Kraut, N. D.; 9238−9241.
Bright, F. V.; Detty, M. R.; Eisenberg, R. J. Am. Chem. Soc. 2010, 132 (147) Cavell, A. C.; Hartley, C. L.; Liu, D.; Tribble, C. S.; McNamara,
(44), 15480−15483. W. R. Inorg. Chem. 2015, 54 (7), 3325−3330.
(117) Krawicz, A.; Yang, J.; Anzenberg, E.; Yano, J.; Sharp, I. D.; (148) Kal, S.; Filatov, A. S.; Dinolfo, P. H. Inorg. Chem. 2014, 53 (14),
7137−7145.
Moore, G. F. J. Am. Chem. Soc. 2013, 135 (32), 11861−11868.
(149) Hartley, C. L.; DiRisio, R. J.; Chang, T. Y.; Zhang, W.;
(118) Hawecker, J.; Lehn, J. M.; Ziessel, R. Nouv. J. Chemie 1983, 7 (5),
McNamara, W. R. Polyhedron 2015, In Press, DOI: 10.1016/
271−277.
j.poly.2015.11.023.
(119) Veldkamp, B. S.; Han, W.-S.; Dyar, S. M.; Eaton, S. W.; Ratner,
(150) Wise, C. F.; Liu, D.; Mayer, K. J.; Crossland, P. M.; Hartley, C. L.;
M. A.; Wasielewski, M. R. Energy Environ. Sci. 2013, 6 (6), 1917−1928. McNamara, W. R. Dalton Trans. 2015, 44, 14265−14271.
(120) Li, L.; Duan, L.; Wen, F.; Li, C.; Wang, M.; Hagfeldt, A.; Sun, L. (151) Connor, G. P.; Mayer, K. J.; Tribble, C. S.; McNamara, W. R.
Chem. Commun. 2012, 48 (7), 988−990. Inorg. Chem. 2014, 53 (11), 5408−5410.
(121) Lakadamyali, F.; Reynal, A.; Kato, M.; Durrant, J. R.; Reisner, E. (152) Brown, H. J. S.; Wiese, S.; Roberts, J. A. S.; Bullock, R. M.; Helm,
Chem. - Eur. J. 2012, 18 (48), 15464−15475. M. L. ACS Catal. 2015, 5 (4), 2116−2123.
(122) Wen, F.; Yang, J.; Zong, X.; Ma, B.; Wang, D.; Li, C. J. Catal. (153) Felton, G. A. N.; Glass, R. S.; Lichtenberger, D. L.; Evans, D. H.
2011, 281 (2), 318−324. Inorg. Chem. 2006, 45 (23), 9181−9184.
(123) Zhang, P.; Wang, M.; Li, C.; Li, X.; Dong, J.; Sun, L. Chem. (154) Nicholson, R. S.; Shain, I. Anal. Chem. 1964, 36 (4), 706−723.
Commun. 2009, 46 (46), 8806−8808. (155) Costentin, C.; Robert, M.; Savéant, J.-M. Phys. Chem. Chem.
(124) Solis, B. H.; Hammes-Schiffer, S. Inorg. Chem. 2011, 50 (21), Phys. 2010, 12 (37), 11179−11190.
11252−11262. (156) Bonin, J.; Costentin, C.; Louault, C.; Robert, M.; Routier, M.;
(125) Dempsey, J. L.; Winkler, J. R.; Gray, H. B. J. Am. Chem. Soc. 2010, Savéant, J.-M. Proc. Natl. Acad. Sci. U. S. A. 2010, 107 (8), 3367−3372.
132 (47), 16774−16776. (157) Snir, O.; Wang, Y.; Tuckerman, M. E.; Geletii, Y. V.; Weinstock,
(126) Muckerman, J. T.; Fujita, E. Chem. Commun. 2011, 47 (46), I. A. J. Am. Chem. Soc. 2010, 132 (33), 11678−11691.
12456−12458. (158) Costentin, C.; Robert, M.; Savéant, J. M. Acc. Chem. Res. 2010, 43
(127) Li, G.; Estes, D. P.; Norton, J. R.; Ruccolo, S.; Sattler, A.; Sattler, (7), 1019−1029.
W. Inorg. Chem. 2014, 53 (19), 10743−10747. (159) Costentin, C.; Robert, M.; Savéant, J.-M. Chem. Rev. 2010, 110
(128) Li, G.; Han, A.; Pulling, M. E.; Estes, D. P.; Norton, J. R. J. Am. (12), PR1−PR40.
Chem. Soc. 2012, 134 (36), 14662−14665.
(129) Estes, D. P.; Grills, D. C.; Norton, J. R. J. Am. Chem. Soc. 2014,
136 (50), 17362−17365.
(130) Lacy, D. C.; Roberts, G. M.; Peters, J. C. J. Am. Chem. Soc. 2015,
137 (14), 4860−4864.
(131) Bullock, R. M.; Appel, A. M.; Helm, M. L. Chem. Commun. 2014,
50 (24), 3125−3143.
(132) Wilson, A. D.; Shoemaker, R. K.; Miedaner, A.; Muckerman, J.
T.; DuBois, D. L.; DuBois, M. R. Proc. Natl. Acad. Sci. U. S. A. 2007, 104
(17), 6951−6956.
(133) Rakowski DuBois, M.; DuBois, D. L. Chem. Soc. Rev. 2009, 38
(1), 62−72.
(134) Helm, M. L.; Stewart, M. P.; Bullock, R. M.; DuBois, M. R.;
DuBois, D. L. Science 2011, 333 (6044), 863−866.
(135) Kilgore, U. J.; Stewart, M. P.; Helm, M. L.; Dougherty, W. G.;
Kassel, W. S.; Dubois, M. R.; Dubois, D. L.; Bullock, R. M. Inorg. Chem.
2011, 50 (21), 10908−10918.
(136) Wiedner, E. S.; Helm, M. L. Organometallics 2014, 33 (18),
4617−4620.
(137) Ho, M.-H.; Rousseau, R.; Roberts, J. A. S.; Wiedner, E. S.;
Dupuis, M.; DuBois, D. L.; Bullock, R. M.; Raugei, S. ACS Catal. 2015, 5
(9), 5436−5452.
(138) O’Hagan, M.; Shaw, W. J.; Raugei, S.; Chen, S.; Yang, J. Y.;
Kilgore, U. J.; DuBois, D. L.; Bullock, R. M. J. Am. Chem. Soc. 2011, 133
(36), 14301−14312.

3659 DOI: 10.1021/acscatal.6b00778


ACS Catal. 2016, 6, 3644−3659

You might also like