Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal Pre-proof

Electron transfer at different electrode materials: metals, semiconductors and


graphene

Elizabeth Santos, Renat Nazmutdinov, Wolfgang Schmickler

PII: S2451-9103(19)30167-X
DOI: https://doi.org/10.1016/j.coelec.2019.11.003
Reference: COELEC 479

To appear in: Current Opinion in Electrochemistry

Received Date: 9 September 2019


Revised Date: 18 November 2019
Accepted Date: 19 November 2019

Please cite this article as: Santos E, Nazmutdinov R, Schmickler W, Electron transfer at different
electrode materials: metals, semiconductors and graphene, Current Opinion in Electrochemistry, https://
doi.org/10.1016/j.coelec.2019.11.003.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Elsevier B.V. All rights reserved.


Electron transfer at different electrode materials: metals,
semiconductors and graphene
Elizabeth Santos1,*, Renat Nazmutdinov2 and Wolfgang Schmickler1

Abstract
Electron transfer reactions are the most important processes at electrochemical
interfaces. They are determined by the interplay between the interaction of the
reactant with the solvent and the electronic levels of the electrode surface. Theoretical
treatments only based on Density Functional Theory calculations are not sufficient.
This review emphasizes mainly the effect of the electronic structure of the electrode
material on electron transfer under different kinetic regimes. Our goal is to understand
experimental results in the framework of a theory valid for arbitrary strengths of
electronic coupling.

Adresses
1
Institute of Theoretical Chemistry, Ulm University, D-89069 Ulm, Germany.
2
Kazan National Research Technological University, 420015 Kazan, Russian
Federation.
*
Corresponding author: Santos, Elizabeth, esantos@uni-ulm.de

1. Introduction
Electron transfer reactions in an electrochemical environment are governed by both
the electronic interactions of the reactant with the electrode material and with the
solvent. The first process is determined by the coupling parameter Veff or
chemisorption function ∆ (ε), as it is called in the model of Anderson and Newns
[1,2]. This quantity reflects the electronic structure of the electrode material and is
proportional to its density of states:
2
∆(ε ) ≈ π Veff ρelec (ε ) (1)
When an electron transfer takes place at the interface, it is accompanied by changes in
intramolecular degrees of freedom and the reorganization of the solvent. In the case of
chemisorption (strong electronic interactions), the reactant can totally lose its
solvation shell, and bond-breaking can also happen (inner sphere reaction). If the
interaction with the electrode is comparatively weak, the reactant preserves its whole
solvation shell. This is an outer sphere electron transfer reaction; the reactant is not in
direct contact with the electrode surface. At least one layer of solvent or some other
ligand separates reactant and electrode. Although the reactant preserves its inner shell,
the solvent in the vicinity of the reactant must reorient during the reaction because
reactant and product carry different charges. The reaction rate is mainly determined
by this reorganization of the solvation shell characterized by the concomitant energy
λ, and the theoretical basis is given by extensions of the Marcus – Hush model [3,4].
Figure 1a illustrates these two types of electron transfer reactions [5-8]. In the case of
inner sphere reactions, a strong dependence on the electrode material is observed,
while in the case of outer sphere reactions the electrode material does not play such an
important role. In order to emphasize the complexity of inner sphere reactions, the
paradigmatic hydrogen evolution reaction is shown as an example. This reaction
occurs at least through two elementary steps, and the limiting step on platinum group
metals and on mercury-like metals is different. A typical mistake for numerous
publications of "volcano plots" and other correlations is to consider the exchange
current without taking into account the different mechanisms, as has been stressed in
other publications [9-10]. It is not easy to obtain the components of the total current
corresponding to the different steps of the reaction.
Nevertheless, according to the strength of the coupling, an electron transfer can occur
within an adiabatic or nonadiabatic regime. The interactions between the reactant and
the metal by adiabatic reactions are sufficiently strong such that the electron

(a)
outer sphere reactions

e Z
Metal
Metal e

inner sphere reactions

e A

Metal
B
e Metal

(b)

Figure 1. (a) experimental values of the standard exchange current for an outer
sphere reaction (filled symbols); data obtained from references [5,6] and for an inner
sphere reaction, such as the hydrogen evolution reaction; data obtained from
references [7,8]. At right: schematic representations of outer and inner sphere
reactions. (b) dependence of the rate constant on the interaction strength ∆ based on a
computer simulation [15] [13]. The black line shows the prediction from first order
perturbation theory.

exchange takes place every time the system reaches the transition state. Thus the
system is in electronic equilibrium for all solvent configurations. In the case that the
electronic interactions are weaker, the system can pass the saddle point of the reaction
coordinate without an electron transfer, so that it subsequently returns to its initial
state. These reactions are called nonadiabatic, and the interaction strength will enter
into the pre-exponential factor of the expression for the velocity.

In summary, there are three different kinetic regimes depending on the strength of the
electronic interactions between the reactant and the electrode material as illustrated in
Figure 1b:
I. If the interactions are sufficiently weak, the reaction proceeds non-adiabatiacally
and first order perturbation theory can be applied [11-14][9-12]. This corresponds to
the linear behavior observed in Figure 1b, where the pre-exponential factor is
proportional to ∆.
II. If the interactions are of intermediate strength, solvent dynamics starts to influence
the rate through the friction factor, which can be described in terms of Kramer's
theory [15] [13]. Therefore, in a certain region the rate constant is almost independent
of ∆. This regime is obeyed by outer sphere reactions, and the behavior can be
described in the framework of Marcus and Hush model [3,4], where the reaction
barrier is not affected by the electronic coupling.
III. If the interactions are very large, a catalytic effect is observed, and the rate
constant increases again, due to the lowering of the activation energy by the electronic
coupling (more details about our theory of electrocatalysis can be found in [16-17]
[14-15] and previous review articles [18-21] [16-19]).
In the case of metallic electrodes, electron transfer reactions proceed mainly
adiabatically, while on semiconducting and carbonaceous electrodes is still
controversial if they take place in an adiabatic or non-adiabatic regime.

2. Electronic structure of electrodes from different materials


The electronic structure of different electrode materials can be quite complex.
Therefore, our strategy is to use idealized band shapes for calculations, in order to
understand the influence of the electronic structure on the mechanism of electron
transfer at different materials. Figure 2 illustrates different models assumed for the
density of states (DOS) for diverse electrode materials.
The simplest possible model is the wide band approximation with constant density of
states [22] [20]. In the case of metals, the Fermi level lies within this band. In the case
of semiconductors, a step function can be used assuming that the DOS is constant
within two bands and vanishes in the gap with the Fermi level within this band gap
[23,24**] [21,22**].
However, this approximation is certainly not realistic. A better model considers
semielliptic bands [1,2]. In Figure 2 it was assumed that the center of the band
coincides with the Fermi level for the metal, while two semiellipses represent the
valence and conductance bands for the semiconductor with the Fermi level at center
of the band gap [24**][22**]. Obviously, the center, the width and the separation
(band gap) of the bands can be shifted according to the specific system to be
investigated. Ideal intrinsic semiconductors show sharp band edges. However, doping
and defects introduce localized electronic states in the band gap. In order to analyze
this effect, we consider an amorphous semiconductor with a density of states in the
band gap that takes the form of a negative semiellipsis (see green line in Figure 2).
The electronic structure of graphene is particularly interesting. We have calculated the
density of states using the tight binding model [25] [23]. The presence of topological
defects or doping introduces mid gap states on the density of states [26-28] [24-26].
We have accounted for such distortions in the electronic structure introducing a
Lorentzian function at three different positions relative to the Fermi level (see Figure
2) [29**] [27**].

3. Electrochemical response for the different electrode materials


In the case of a non-adiabatic regime, the electronic interactions between electrode
and reactant are very weak, in consequence the electron transfer occurs from (or into)
a very sharp electronic level of the reactant. The expression for the current-
overpotential dependency j(η) for the reduction of a species A is given by [13,16]
[11,14]:
2
+∞
Veff
( 4πλ kBT ) ∫ ρ (ε ) f (ε,T ) W (ε, λ, η) dε
−1/2 (2)
jred = P elec FD elec ox
h −∞

Figure 2. Normalized density of electronic states (DOS) of all the model systems
considered in this work (more details in the text).
where:
Wox (ε, λ, η ) = exp − ( λ − ε − η ) / ( 4 λ kBT )
2

fFD (ε,Telec ) = 1/ 1+ exp (ε / kBTelec )


fFD(ε,Telec) is the Fermi-Dirac distribution, i.e. the probability to find an occupied state
in the electrode and η is the overpotential. In equation (2) the Fermi level is taken as
zero. P is an extra factor, which depends on the reactant work term (i.e. the
Boltzmann factor that converts the bulk to the surface concentration) and on the
reaction volume resulting from its integration over the distance (of the order of 10-10 –
10-9 m) [13,16] [11,14]:.
According to Gerischer’s interpretation [30,31] [28,29], Wox(ε,λ,η) is the probability
that an electron from state ε in the metal is transferred to an unoccupied (oxidized)
state in the solution. It accounts for the fluctuations in the solvent shell, and
therefore it is independent of the electrode material. This must not be confused with
the electronic density of states of the reactant, these are sharply localized at a given
energy.
An equivalent expression for the oxidation can be easily found.
In the case of an adiabatic regime, where the electronic interactions between reactant
and electrode are stronger, a broadening of the electronic level of the reactant takes
place. Therefore, the total energy of the system, also taking into account the
interactions with the solvent is given by:
EF =0
Etot (λ, q, η ) = Esolv (λ, q) + Eelec (λ, q, η ) = λ q + 2zqλ +
2
∫ ε ρ A (ε, q, η ) dε (3)
−∞
Here, q is the normalized solvent coordinate, which describes the reaction path [3,4,
16 14], and z is the charge number. The latter term in equation (3) is the electronic
energy, and it is this term that contributes to the decrease of the activation barrier
when strong interactions with the electrode causes the broadening of the density of
states of the reactant ρA.

4. Results of model calculations


Figure 3 shows the results of numerical calculations for a non-adiabatic regime of the
integral in equation (2) J(ρelec, λ, η) as a function of the overpotential η, using the
density of states ρelec for the different electrode materials of Figure 2.
The absolute values of the current are determined by the pre-factor containing the
coupling constant |Veff|2 and we shall return to this later. Now, it is interesting to
analyze the features of these curves, which are determined by the features of the
density of states of the electrode material. The usual near-activationless regime is
attained at η of about 0.7 eV for the metal, while it is still not observed for the other
materials even at higher overpotentials. It can be explained taking into account that
while the product fFD . ρelec for the metal is almost constant below the Fermi level, that
product for the other materials increases monotonically when the electronic energy
becomes more negative. An interesting behaviour is observed when the graphene
contains electronic states in the mid-gap. In some cases, a maximum is observed,
which could be interpreted as a vestige of the inverse region in Marcus theory.
However, a clear explanation based on the Gerischer’s approach can be found. The
contribution of the mid-gap states is stronger at low overpotentials, because the
overlap of the probability Wox(ε,λ,η) with these states is larger in this case. At larger
overpotentials this overlap occurs with the deeper electronic states, which are similar
for all graphene materials. Therefore a maximum appears at intermediate intervals.
Figure 3. Effect of the electronic structure of the electrode material on the current -
overpotential dependencies for the non adiabatic limit.

In contrast to the non-adiabatic regime in the adiabatic case, the reactant and the
electrode share their electrons and a broadening of the electronic level of the reactant
is produced. Herein lies the key effect of the electronic structure on the kinetics of the
reaction. These effects can be better understood by analyzing the density of states of
the reactant (for a detailed discussion see reference [29**27**]).

Pre-exponential factor of the rate constant


Finally, we shall briefly discuss the pre-exponential factor, which has recently
received renewed attention [32** 30**]. For adiabatic outer-sphere electron transfer
there is a fairly well developed theory, which in agreement with experiment gives pre-
exponential factors of the order of 104 cm s-1; tunneling effects of inner-sphere modes
can reduce this typically by one order of magnitude. For inner sphere reactions, both
theory and some experimental data suggest a lower pre-exponential factor than for
outer-sphere transfer, but the theory is not well developed, and experimental data are
sometimes contradictory and affected by anion adsorption. For details we refer to the
cited article.

For non-adiabatic reactions, this factor depends strongly on the overlap between the
reactants and thus on their electronic properties. Hence there can be no general rule
for this class of reactions. For specific cases, the pre-exponential factor can be
estimated from quantum-chemical calculations. These can be based on first order
perturbations theory following eq. (2) by evaluating the overlap matrix element |Veff |2
and placing it directly into this equation. Alternatively, a Landau-Zener factor γe
given by [10, 11, 33 31]:
κ = 1− exp− (2πγ ) (4)
e e
which has the meaning of a transmission coefficient in the sense of activated complex
or Kramer’s theory, i.e. the probability of electron transfer after the redox system
attains the saddle point. The quantity γe depends on the matrix element, and for small
interactions the two approaches are equivalent. Obviously the overlap decreases
strongly, typically exponentially, with the distance of the reactant from the electrode.
In the non adiabatic regime the electronic transmission coefficient is directly
proportional to the electrode DOS near the Fermi level, while in the adiabatic limit
this quantity does not depend on the DOS. The DOS dependent κe in the intermediate
regime is considered in detail in reference [34** 32*]. As an example, Figure 4 shows
the transmission coefficient for Fc/Fc+ couple as a function of the distance from a
graphene electrode, Two different orientations of the reactant have been considered –
the axis vertical or parallel to the surface – and two different positions relative to the
graphene lattice: either above a bridge or above a hollow site. For comparison, the
results for a Au(111) surface are also shown. For graphene, the vertical orientation is
more favorable, and only weakly depends on the site, while for the horizontal
orientation the bridge position is more favorable. For Au(111) κe decays much more
slowly and is independent of the orientation and position. Nevertheless, even for
graphene the reaction is predicted to be almost adiabatic κ e ≈ 1 for distances x<4.5 Å.

Figure 4. Electronic transmission coefficient (from reference [29**]) calculated for a


Fc/Fc+ redox couple at a graphene surface for several orientations of Fc+. The results
for the Au(111) surface were obtained in Ref.[33] using the same model and
averaging over four different orientations of Fc+ .

Conclusions
We have discussed electron transfer reactions in an electrochemical environment for
different electrode materials. We have focused on the effect of the electronic structure
of the substrate under diverse kinetics regimes: outer- and inner-sphere, adiabatic and
non adiabatic reactions. Idealized electronic structures have been used to predict the
electrochemical behavior depending on the strength of the interactions between
electrode and reactant. Our goal was to shed light on the controversy found in the
literature about the electron transfer mechanism. The related problems and the appeal
to theoreticians to develop new tools to understand the fundamental electrochemistry
of graphene has been emphasized by Patten et al. [35 33]. Electron transfer reactions
measurements at graphene electrochemical interfaces, particularly of outer sphere
redox processes have been reported in the literature. Experiments using scanning
electrochemical cell microscopy (SECCM) are interesting since they allow local
measurements at the nanoscale [36-38]. Frequently, oversimplified models are
applied. In the case of metals, density functional theory calculations are extended to
electrochemical interfaces. However, charged species and solvent effects are very
difficult to model within this computational approach. In the case of semiconductors
and graphene, The Marcus-Hush and Gerischer frameworks are the most popular in
explaining the experimental data of electron transfer. We have shown that much
attention must be paid to the strength of the interactions and the electronic structures
of the electrode material in order to correctly analyze experimental results.

Our theoretical work is complementary to recent extended reviews [34*39-40*],


which focus on electron transfer at graphene from the experimental point of view. The
results of our model calculations for pristine graphene agree qualitatively with recent
experimental findings [41 35, 36*]. The experimental data from [36*] cannot be fit
with extant Marcus−Hush theory. Therefore, the authors assume an empiric linear
variation with energy of the DOS of graphene and consequently obtain an improved
fit. However, they stress the necessity for a deeper theoretical investigation. Our
model can explain these results in a consistent and realistic way.
Another promising point is the modification of the electronic structure of
semiconductors by intercalation or doping, which might result in a noticeable increase
of the rate of electron transfer across such electrodes [4237]. In the case of graphene,
an interesting aspect is the role of graphene edges in the electron transfer kinetics [43]
Finally, an important advantage of our work is the fact that it is valid for arbitrary
interactions strength. This makes it possible to apply it for investigating
photocatalytic reactions on semiconductors. An interesting effect is the generation of
the so called hot electrons by means of a laser pulse [44* 38*]. This phenomenon can
be also well described within our model, as we have shown for graphene [29**27**].

References
[1] P.W. Anderson, Localized magnetic states in metals, Phys. Rev. 124 (1961) 41-53.
[2] D.M. Newns, Self-consistent model of hydrogen chemisorption, Phys. Rev. 178
(1969) 1123.
[3] R.A. Marcus, On the theory of Oxidation-Reduction reactions involving electron
transfer, J. Chem. Phys. 24 (1956) 966-978.
[4] N.S. Hush, Adiabatic rate processes at electrodes. I. Energy-charge relationships,
J. Chem. Phys. 28 (1958) 962-972.
[5] T. Iwasita, W. Schmickler, J.W. Schultze, The Influence of the Metal on the
Kinetics of Outer Sphere Redox Reactions, Ber. Bunsen-Ges. 89 (1985) 138.
[6] E. Santos, T. Iwasita, W. Vielstich, On the use of the coulostatic method for the
investigation of fast redox reactions, Electrochim. Acta 31 (1986) 431.
[7] J.K. Nørskov, T. Bligaard, A. Logadottir, J.R. Kitchin, J. G. Chen, S. Pandelov, U.
Stimming, Trends in the Exchange Current for Hydrogen Evolution, J. Electrochem.
Soc. 152 (2005) J23.
[8] B.E. Conway, E.M. Beatty, P.A.D. De Maine, Electrochemical kinetics of
hydrogen evolution at copper-nickel alloys. Relation to electronic properties of the
electrodes, Electrochim. Acta 7 (1962) 39.
[9] O. A. Petrii and G. A. Tsirlina, Electrocatalytic activity prediction for hydrogen
electrode reaction: intuition, art, science, Electrochim. Acta 39 (1994) 1739-1747.
[10] W. Schmickler and S Trasatti, Comment on “Trends in the Exchange Current for
Hydrogen Evolution” [J. Electrochem. Soc., 152, J23 (2005)], J. Electrochem. Soc.
153 (2006) L31-L32.
[11][9] V.G. Levich, Kinetics of reactions with charge transfer, in Physical
Chemistry, and Advanced Treatise, Vol. Xb, ed. H. Eyring, D. Henderson and W.
Jost, Academic Press, N. York, 1970.
[12][10] A. M. Kuznetsov, Charge Transfer in Physics, Chemistry and Biology.
Mechanisms of Elementary Processes and Introduction to the Theory; Gordon and
Breach Science Publ., Berkshire, 1995.
[13][11] R.R. Nazmutdinov, M.D. Bronshtein, T.T. Zinkicheva, D.V. Glukhov,
Modeling of Electron Transfer Across Electrochemical Interfaces; State-of-the-Art
and Challenges for Quantum and Computational Chemistry. Int. J. Quantum Chem.
116 (2016) 201.
[14][12] R. R. Dogonadze, in Reactions of Electrons at Electrodes, edited by N. S.
Hush, Interscience, London, 1971.
[15][13] W. Schmickler, J. Mohr, The rate of electrochemical electron-transfer
reactions, J. Chem. Phys. 117 (2002) 2867.
[16][14] E. Santos, W. Schmickler in Interfacial Electrochemistry. Springer Verlag,
2011.
[17][15] E. Santos, A. Lundin, K. Pötting, P. Quaino and W. Schmickler, A theory for
the electrocatalysis of hydrogen evolution, Phys. Rev. B 79 (2009) 235436.
[18][16] E. Santos and W. Schmickler, Electrochemical electron transfer: from
Marcus theory to electrocatalysis, in Fuel Cell Catalysis: a Surface Science
Approach”. Edited by M. Koper. Wiley Editorial. 2009.
[19][17] E. Santos and W. Schmickler, Recent Advances in Theoretical aspects of
electrocatalysis, in Modern Aspects of the Electrochemistry, Number 50: “Theory
and Experiment in Electrocatalysis". Edited by Ralph E. White, Perla Balbuena,
Venkat Subramanian. Springer Verlag. 2010.
[20][18] E. Santos and W. Schmickler, Catalysis of electron transfer at metal surfaces,
in Catalysis in Electrochemistry: from Fundamental Aspects to Strategies for Fuel
Cells
Development. Edited by E. Santos and W. Schmickler. Wiley 2011.
[21][19] G. Soldano, W. Schmickler, M.F. Juarez, P. Quaino, E. Santos, Electron
transfer in nanoelectrochemical systems, Nanoelectrochemistry monograph_K15931.
CRC – Press. 2014.
[22][20] W. Schmickler, A theory of adiabatic electron transfer reactions, J.
Electroanal. Chem. 204 (1986) 31-43.
[23][21] F. D. M. Haldane and P. W. Anderson, Simple model of multiple charge
states of transition-metal impurities in semiconductors, Phys. Rev. B 13 (1976) 2553.
[24**][22**] W. Schmickler, E. Santos, M. Bronshtein, R. Nazmutdinov, Adiabatic
Electron‐Transfer Reactions on Semiconducting Electrodes, Chem. Phys. Chem. 18
(2017) 111-116.
An interesting approach of electron transfer reactions on semiconductor electrodes
considering different ideal features of the conduction and valence band
[25][23] A.H. Castro Neto, F. Guinea, N.M.R. Peres, K.S. Novoselov, A.K. Geim,
The electronic properties of graphene, Rev. Mod. Phys. 81 (2009) 109−162.
[26][24] O.V. Yazyev, S.G. Louie, Topological defects in graphene: Dislocations and
grain boundaries, Phys. Rev. B 81 (2010) 195420.
[27][25] Y.C. Chang, S. Haas, S. Defect-induced resonances and magnetic patterns in
graphene, Phys. Rev. B 83 (2011) 085406.
[28][26] Z. Ding, L. Zhao, L. Suo, Y. Jiao, S. Meng, Y.-S Hu, Z. Wang, L. Chen,
Towards understanding the effects of carbon and nitrogen-doped carbon coating on
the electrochemical performance of Li4Ti5O12 in lithium ion batteries: a combined
experimental and theoretical study, Phys. Chem. Chem. Phys. 13 (2011)
15127−15133.
[29**][27**] R. R. Nazmutdinov, M.D. Bronshtein, and E. Santos, Electron Transfer
across the Graphene Electrode/Solution Interface: Interplay between Different Kinetic
Regimes, J. Phys. Chem. C 123 (2019) 12346-12354.
Systematic analysis of electron transfer reactions on graphene (pristine and with
topological defects) considering ideal electronic structures. Discussion of
controversial experimental results in the literature in the framework of the proposed
model.
[30][28] H. Gerischer, Über den Ablauf von Redoxreaktionen an Metallen und an
Halbleitern. Z. Phys. Chem. NF 1960, 26, 223−247.
[31][29] H. Gerischer, Über den Ablauf von Redoxreaktionen an Metallen und an
Halbleitern. III. Halbleiterelektroden. Z. Phys. Chem. 1961, 27, 48−79.
[32**][30**] Zheng-Da He, Yan-Xia Chen, E. Santos, and W. Schmickler, The Pre-
exponential Factor in Electrochemistry, Angew.Chem.Int.Ed. 57 (2018) 2-11.
Exhaustive discussions of the physical meaning of the pre-exponential factor of the
rate constant in electrochemical systems.
[33][31] V.A. Nikitina, S.A. Kislenko, R.R. Nazmutdinov, M.D. Bronshtein, G.A.
Tsirlina, Ferrocene/ferrocenium redox couple at Au(111)/ionic liquid and
Au(111)/acetonitrile interfaces: A molecular-level view at the elementary act. J. Phys.
Chem. C 118 (2014) 6151−6164.
[34*][32*] A.S. Berezin, R.R. Nazmutdinov, Monte Carlo simulations of
heterogeneous electron transfer: new challanges, Russ. J. Electrochem. 53 (2017)
1232-1238.
Detailed explanations of the dependence of the transmission coefficient in the pre-
exponential term of the rate constant of electron transfer on the density of states of the
substrate.
[35][33] H.V. Patten, M. Velický, R.W Dryfe, Electrochemistry of Graphene. In
Electrochemistry of Carbon Electrodes; Alkire, R. C.; Barlett, P. N.; Lipkowski, J.,
Eds.; Adv. in Electrochem. Sci. and Eng.; Wiley-VCH Verlag: Weinheim, 2015; Vol
16, pp 121−161.
[36] A.G. Güell, A.S. Cuharuc, Y.R. Kim, G. Zhang, S.Y. Tan, N. Ebejer, P.R.
Unwin, Redox-dependent spatially resolved electrochemistry at graphene and graphite
step edges. ACS Nano 9 (2015) 3558−3571.
[37] J. Hui, X. Zhou, R. Bhargava, A. Chinderle, J. Zhang, J. Rodríguez-López,
Kinetic Modulation of Outer-Sphere Electron Transfer Reactions on Graphene
Electrode with a Sub-Surface Metal Substrate, Electrochim. Acta 211 (2016)
1016−1023.
[38] J.-H. Zhong, J. Zhang, X. Jin, J.-Y. Liu, Q. Li, M.-H. Li, W. Cai, D.-Y. Wu, D.
Zhan, B. Ren, Quantitative Correlation between Defect Densiity and Heterogenous
Electron Transfer Rate of Single Layer Graphene, J. Am. Chem. Soc. 136 (2014)
16609–16617
[39*34*] A. Kaplan, Zhe Yuan, J. D. Benck, A. G. Rajan, Ximo S. Chu, Qing Hua
Wang and M. S. Strano, Current and future directions in electron transfer chemistry of
graphene, Chem. Soc. Rev. 46 (2017) 4530-4571.
Extensive review of experimental work on electron transfer reactions at graphene
interfaces with a large number of references such that the state of art in the field can
be easily followed.
[40**] L. J. A. Macedo, R. M. Iost, A. Hassan, K. Balasubramanian, and F. N.
Crespilho, Bioelectronics and Interfaces Using Monolayer Graphene,
ChemElectroChem 6 (2019) 31–59.
Both references [39-40] are extensive reviews of experimental work on electron
transfer reactions at graphene interfaces with a large number of references such that
the state of art in the field can be easily followed. While [38] focus on 3D complex
porous materials, [40] emphasize the application of 2D graphene to biolelectronics.
Particularly [40] shows an interesting table about electron transfer kinetics determined
for various graphene electrodes, with indication of the rate estimation models and
measurement methods for different electrochemistry redox probes.
[35] K.-W Chang, I.A. Santos, Y. Nguyen, Y.-H Su, C.C. Hsu, Y.P. Hsieh, M.
Hofmann, Electrostatic Control over the Electrochemical Reactivity of Graphene.
Chem. Mater. 30 (2018) 7178−7182.
[4136*] R. Narayanan, H. Yamada, B.C. Marin, A. Zaretski, P.R. Bandaru,
Dimensionality-Dependent Electrochemical Kinetics at the Single-Layer Graphene-
Electrolyte Interface. J. Phys. Chem. Lett. 8 (2017) 4004−4008.
Experimental data for the [Fe(CN)6]3-/4- redox couple at the graphene/ electrolyte
interface are shown as the ln k versus η plot in the overpotential interval of 0.7
eV.16. The plot looks very similar to our current− overpotential curve (compare
Figure 3B in this reference with the green curve in Figure 2 ).
[4237] S.A. Kislenko, M.F. Juárez, F. Dominguez-Flores, W. Schmickler, R.R.
Nazmutdinov, Tuning the rate of an outer sphere electron transfer by changing the
electronic structure of carbon nanotubes, J. Electroanal. Chem. 847 (2019) Article
number 113186.
[43] S. V. Pavlov, R. R. Nazmutdinov, M. V. Fedorov, and S. A. Kislenko, Role of
Graphene Edges in the Electron Transfer Kinetics: Insight from Theory and
Molecular Modeling, J. Phys. Chem. C 123 (2019) 6627−6634.
[44* 38*] Y. Lu, B. Ma, Y. Yang, E. Huang, Zh. Ge, T. Zhang, S. Zhang, L. Li, N.
Guan, Y. Ma, Y. Chen, High activity of hot electrons from bulk 3D graphene
materials for efficient photocatalytic hydrogen production, Nano Res. 10 (2017)
1662−1672.
This work should offer a new strategy for overcoming the high energy barriers of
many important reactions. Electron transfer and mechanism of photocatalytic
hydrogen production is discussed and agree with our predictions.
Declarations of conflict of interest
none

You might also like