Modeling Electron Transfer in Diffusive Multidimensional

4 hours ago - Watson shifts from drug discovery to the clinic. Responding to reports that it will no longer sell its Watson artificial intelligence sy...
0 downloads 0 Views 895KB Size
Subscriber access provided by UNIV OF LOUISIANA

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Modeling Electron Transfer in Diffusive Multidimensional Electrochemical Systems Alec J. Coffman, Aparna Karippara Harshan, Sharon Hammes-Schiffer, and Joseph E. Subotnik J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.9b02068 • Publication Date (Web): 01 May 2019 Downloaded from http://pubs.acs.org on May 1, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Modeling Electron Transfer in Diffusive Multidimensional Electrochemical Systems Alec J. Coffman,∗,† Aparna Karippara Harshan,‡ Sharon Hammes-Schiffer,¶ and Joseph E. Subotnik† Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, USA, Department of Chemistry, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA, and Department of Chemistry, Yale University, New Haven, Connecticut 06520, USA E-mail: [email protected]



To whom correspondence should be addressed Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, USA ‡ Department of Chemistry, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA ¶ Department of Chemistry, Yale University, New Haven, Connecticut 06520, USA †

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

May 1, 2019 Abstract We analyze different stochastic approaches for simulating electron transfer and potential sweep experiments in diffusive multidimensional electrochemical systems. In particular, we focus on a simple two-dimensional system, where one dimension is a traditional mass diffusion coordinate moving reactants from bulk solution to the electrode, and a second dimension represents a reorganization coordinate capturing solvent motion. We find that this multidimensional system can indeed be reduced to a simpler one-dimensional model, provided certain circumstances pertaining to separability between the two coordinates are met. Our results should begin to bridge the gap between continuum models of electrochemical dynamics/transport and ab initio models of surface electronic structure.

1

Introduction

Within the field of electrochemistry, one of the open questions is: what is the simplest computational model for capturing the details of Tafel plots and cyclic voltammograms for realistic experiments? Historically, when modeling such electron transfer (ET) events at the electrode surface, the usual approach has been to enforce either static or local empirical boundary conditions, especially Nernstian equilibrium or Butler-Volmer kinetics. 1 These boundary conditions are simple to implement and applicable in many common applications, even in the presence of diffuse charges on surfaces. 2 Nevertheless, the above approaches do not address the actual mechanism of a complicated ET event; 3 for instance, the Butler-Volmer approach hides a great deal of physics (all barrier heights and shapes, reorganization free energies, solvent recrossing probabilities) in one empirical parameter, namely the transfer coefficient. Many efforts to improve upon the canonical techniques of electrochemistry have been constructed, most commonly using Marcus-Hush (MH) theory which usually assumes sym2

ACS Paragon Plus Environment

Page 2 of 45

Page 3 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

metric parabolas. 4,5 However, there has also been some development for asymmetric parabolas 6 and non-parabolic surfaces (where local low-frequency vibrations can alter the effective force constant of the two oxidative states differently 7 ). More generally, one can estimate rate constants for electron transfer using Fermi’s golden rule and evaluating a one-dimensional integral (sometimes called a Marcus-Hush-Chidsey integral 8,9 ). And quite often, by fitting experimental I-V curves to Fermi’s golden rule (FGR) rates, one can extract reasonable thermodynamic parameters, e.g. reorganization energies. 10 However, even this approach is not without limitations, 11 and FGR integrals have been shown to be insufficient for replicating some voltammetry experiments quantitatively; 6 in fact, sometimes more involved FGR treatments of I-V curves perform worse than the simpler Butler-Volmer equations. 12 Current electrochemical research is exploring alternatives to Bulter-Volmer, including phase-field models, 13 density functional theory (DFT) aided molecular dynamics simulations, 14,15 and other approaches. In the end, the potential energy surfaces as relevant to electrochemistry are large and quite complicated, and the statistical mechanics of sampling such configurations is daunting. As far as we are aware, there is still no complete and practical methodology for simulating electrochemical phenomena that (i) can describe complex chemical transformations (including multi-electron transfer events) using (ii) only a few parameters (ideally extracted from electronic structure theory). We also note that modern simulations using the Nernstian and Butler-Volmer boundary conditions should be valid only when electron transfer at the surface is relatively fast and largely decoupled from other surface processes. Unlike these approaches, the ideal electrochemical simulation (iii) should be applicable whether electron transfer, mass diffusion, or some other transformation is rate limiting. From our point of view, even though there is also a long history of studying ET coupled to diffusion in solution, 16–19 as well as photoinduced ET and ion recombination in different enviornments, 20–22 there has still been very little development of the theory of electrochemical dynamics compared to the vast theoretical developments in photochemistry and photodynamics. This gap

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

between photochemistry and electrochemistry poses both a challenge and opportunity for chemical theory. Recently, there has has been recognition of the importance of modeling electrochemical ET through a more complete Marcus picture, 23 but previous works in this vein have usually assumed restrictive conditions, for example, focusing primarily on equilibrium statistical fluctuations. 24 Thus far, little attention has been given to methods that explicitly model the dynamics of electrochemical ET with molecular motion in bulk and at the surface. In this paper we analyze the simplest such multidimensional approach for simulating electrochemical dynamics: we study electron transfer with molecular motion possible along two different nuclear coordinates, one a solvent reorganization coordinate and the other a diffusion coordinate, under an applied potential. By modeling ET with a stochastic, nonadiabatic surface hopping 25 (SH) ansatz, we will build new intuition for how to understand I-V curves dynamically. An outline of this paper is as follows. In Sec. II we compare the standard approaches for electrochemical simulations, highlighting how these different approaches relate to one another. In Sec. III we present results for each approach, comparing and contrasting when a two-dimensional simulation can be reduced to a more simpler one-dimensional case. Finally, in Sec. IV, we discuss how to connect mass transport models with ab initio models of electrochemical dynamics and we point out the obvious improvements needed in order to make progress. We conclude in Sec. V.

2 2.1

Methods Basic electrochemical methods in one dimension

As is standard, we begin our discussion with an analysis of electrochemical dynamics in one dimension, where mass transport and diffusion are paramount.

4

ACS Paragon Plus Environment

Page 4 of 45

Page 5 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2.1.1

Mass transport with Butler-Volmer and Nernstian boundary conditions

Historically, simulation of electrochemical systems has been carried out by modeling solute mass transport and incorporating electrode boundary conditions that obey either Nernstian or Butler-Volmer boundary conditions. For the most part, this approach requires only a simple grid in one-dimension (to simulate diffusive motion of the electroactive species in solution) and the implementation of species exchange at the surface (to simulate charge transfer). To illustrate this approach, consider the following equation: k

f A + e– − )− −* −B

kb

Here, species A is reduced at the electrode to form B, and the reaction is reversible. The kinetics obey x ∂cA xˆ J~A = −DA ∂x x ∂cB J~B = −DB xˆ ∂x

(1a)

2 ∂cA x ∂ cA ~ · J~A = DA = −∇ ∂t ∂x2 2 ∂cB x ∂ cB ~ · J~B = DB = −∇ , ∂t ∂x2

(1b)

where xˆ is the unit vector for the diffusive coordinate, JA (JB ) is the current of species A (B) x x (DB ) is the diffusion coefficient in the x direction (which we take as equal for both and DA

species A and B). Equation 1 is an almost complete protocol for simulating electrochemical systems, but does require boundary conditions. The system is governed by the following initial conditions,

cA (t = 0, x) = cbulk =1 A

(2a)

cB (t = 0, x) = cbulk =0 B

(2b)

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 45

and bath boundary conditions

cA (t, x = ∞) = cbulk =1 A

(3a)

cB (t, x = ∞) = cbulk =0 B

(3b)

The initial conditions above state that only A is present at time zero and the boundary conditions are defined so that the concentrations of A and B are equal to their bulk concentrations infinitely far from the electrode. What remains to be determined, however, are the dynamical boundary conditions at the electrode surface, where two options are standard.

Nernstian: On the one hand, Nernstian boundary conditions enforce the proper equilibrium ratio for the surface concentration of species A divided by the surface concentration of species B at all times. We assume the surface concentrations at the boundary are related to the ratio of the forward and backward rate constants, kf and kb : 1,26 kf cB ◦ |x=0 = = e−β(∆G(t)−∆G ) . cA kb

(4)

Butler-Volmer: On the other hand, Butler-Volmer boundary conditions stipulate that the concentration gradient at the electrode surface must be equal to the net flux:

x DA

∂cA |x=0 = kf cA |x=0 − kb cB |x=0 , ∂x

(5)

The forward and backward rates are given by

kf = k ◦ e−αβ(∆G(t)−∆G

◦)

kb = k ◦ e(1−α)β(∆G(t)−∆G

◦)

(6a) (6b)

where k ◦ is the standard rate constant, α is the transfer coefficient, β = (kT )−1 , ∆G(t) is the external potential, and ∆G◦ is the redox potential of the A/B redox couple. 1,26 In the 6

ACS Paragon Plus Environment

Page 7 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

standard electrochemical literature, the value of the transfer coefficient is usually related to the symmetry of the energy barrier in going from A to B, and is equal to 0.5 in the case where the barrier from A to B is completely symmetric. For both Nernstian and ButlerVolmer dynamics we also assume that the flux of species A and B are equal and opposite at the electrode, JA = −JB . When supplemented with one more equation to ensure mass conservation, the above equations define either Nernstian (equation 4) or Butler-Volmer (equations 5 and 6) electrochemical dynamics. For more details about the implementation of these conditions, see Appendix C. Solving the above boundary value problem has a long history, attempted first by Randles 27 and Sevcik. 28 In the case of Nernstian dynamics for a linear sweep voltammetry experiment, there exists analytical values for the peak current and peak potential, Ip and ∆Gp , found by Laplace transforming the diffusion equations in 1(b) and employing the initial and boundary conditions in equations 2-4. 29 Assuming units of current in amperes, electrode area A in cm2 , diffusion constant D in cm2 /s, the scan rate ν in V/s and concentration cA in mol/cm3 , one arrives at 1/2

Ip = (2.69 × 105 )n3/2 ADA ν 1/2 cA (t = 0)

(7)

for the peak current and ∆Gp = ∆Gp/2 − 2.20

RT nF

(8)

for the peak potential (in V). Here n is the number of electrons transferred in the ET event, F is the Faraday constant, and ∆Gp/2 is the half-peak potential, 1 i.e. the driving force for which the current is half of the maximum current in equation 7. Equation 7 shows that, for Nernstian systems where ET kinetics are reversible, Ip is proportional to ν 1/2 and does not depend on k ◦ . For systems that exhibit more complicated dynamics than those of the reversible case, the values of Ip and ∆Gp become more complicated functions of a unitless ◦

parameter Λ ≡ √kDνF , 30 which can be seen as a measure of the relative rates of ET kinetics RT

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 45

q (k ◦ ) and mass transfer kinetics ( DνF ). The statements above can be extended beyond RT linear sweeps to the case of cyclic voltammetry, although again, exact analytical expressions exist only for the Nernstian case. 31,32

2.1.2

A generalized SH approach with flexible boundary conditions

The Butler-Volmer dynamics described above show similarities to a stochastic master equation approach. In fact, with a little work, it can be easily shown that equations 1 and 5 are identical to classical surface hopping solving for the dynamics of coupled nuclear and electronic motion near a metal surface. To be precise, consider the following generalized Anderson-Holstein (AH) Hamiltonian, where an electronic impurity (i.e. a redox-active ˆ α and molecule) is coupled to a collection of nuclei (with position qˆα and momentum Π ˆ α ] = i~) and a bath of metallic fermions (indexed by k): [ˆ qα , Π

H = Hs + Hb + Hc

(9a)

p ~·~ p Hs = E(~r)d† d + UA (~r) + 2m + Hp h i P 1 Π2α (gα ~ r )2 2 2 ) + (g q ~ r + q Hp = ( + m ω ) α α α α α 2 2 mα 2mα ωα α P Hb = (k − µ)c†k ck k P Hc = Wk (~r)(c†k d + d† ck )

(9b) (9c) (9d) (9e)

k

Here Hs represents a Hamiltonian for a molecule with molecular orbital d near a metal surface; Hp represents a collection of nuclear coordinates that lead to friction for the molecule; Hb represents a metal surface with multiple electronic states, indexed by k; and Hc represents the coupling between electrons in the metal and molecule. ~r is the position operator for the electronic impurity; in principle, ~r in equation 9 represents a multidimensional nuclear coordinate, however, in one dimension ~r becomes a scalar variable x. gα is the coupling between the molecule and the collection of nuclei indexed by coordinate α, d/d† represent creation and annihilation operators for a molecular orbital near a surface, µ is the chemical 8

ACS Paragon Plus Environment

Page 9 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

potential of the metal, ck /c†k are the annihilation and creation operators for the metallic bath of electronic states, and Wk is the coupling between the electronic bath and the molecular orbital d. In practice, we will use the function Γ,

Γ(~r, ) = 2π

X

|Wk (~r)|2 δ(k − ),

(10)

k

to characterize the strength of the electron-metal coupling, where

~ Γ

is the lifetime of an

electron placed on the molecule. In the wide band limit, one can assume Γ(~r, ) depends only on ~r (and not on ). The model in equation 9 incorporates two potential energy surfaces (PESs): we define UA to be the diabatic PES for the case where the impurity is unoccupied and we define UB to be the diabatic PES for the case where the impurity is occupied. Thus, the function E in equation 9(b) can be defined as the difference between the diabatic states, E(~r) ≡ UB (~r) − UA (~r). Furthermore, in one dimension, it is standard to choose flat potential energy surfaces with no x dependence1 , such that

UB ≡ UA + ∆G

(11a)

E = ∆G.

(11b)

A Classical Master Equation (CME) for solving the AH model: As shown in Ref. 34, if we consider the limit of weak coupling Γ and high temperature (where all nuclear motion is classical), and the limit of strong friction so that all translational motion is diffusive and governed by the Smoluchowski equation, one can propagate dynamics for the Hamiltonian 1

Interestingly, although this manuscript will not address dynamics in the adiabatic representation, for E defined in equation 9, if one commits to the adiabatic representation, the electronic friction tensor here arises exclusively from non-Condon effects (i.e. the electronic friction tensor is proportional to ∂Γ ∂~ r ). For details, see Ref. 33.

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 45

in equation 9 by solving the relevant master equation of the form: ∂cA ~ · J~A − (kf cA − kb cB )δ(x) =−∇ ∂t 2 x ∂ cA =DA − (kf cA − kb cB )δ(x) ∂x2

(12a)

∂cB ~ · J~B + (kf cA − kb cB )δ(x) =−∇ ∂t 2 x ∂ cB + (kf cA − kb cB )δ(x). =DB ∂x2

(12b)

Here, we have assumed Γ(x) = Γ0 δ(x) so that all hopping occurs at only x = 0. The relevant electrode and bath boundary conditions are

cA (t, x = −1) = cA (t, x = 0)

(13a)

cB (t, x = −1) = cB (t, x = 0)

(13b)

cA (t, x = ∞) = 1

(13c)

cB (t, x = ∞) = 0.

(13d)

Here, we have invoked a fictitious boundary point at x = −12 . In principle, equation 12 can be solved by sampling over a swarm of trajectories hopping back and forth between surfaces and undergoing stochastic motion with Langvein forces. However, to save time, we will integrate 12 directly, solving for the time dependence of the probability densities as they are propagated with the coupled Smoluchowski equations. Now, to prove the equivalence between equations 1, 5, and 12 we simply integrate both 2

In other words, for SH, the relevant boundary conditions are

10

∂cA − ∂x |x=0

ACS Paragon Plus Environment

=

∂cB − ∂x |x=0 .

Page 11 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

sides of equation 12 over a small length , where  → 0 : Z



−

Z



−

∂cA dx = ∂t

Z

 x (DA

−

∂ 2 cA − (kf cA + kb cB )δ(x)) dx ∂x2

∂cA 0 = DA |x=0 − kf cA |x=0 + kb cB |x=0 Z  ∂x 2 ∂ cB ∂cB (DB dx = + (kf cA − kb cB )δ(x)) dx ∂t ∂x2 − ∂cB 0 = DB |x=0 + kf cA |x=0 − kb cB |x=0 . ∂x

(14a)

(14b)

These are the Butler-Volmer boundary conditions. Apparently, adjusting the boundary conditions at x = 0 (the electrode surface) can reproduce Butler-Volmer dynamics entirely and with more generality, giving us another view of electrochemical dynamics. Note that, with the diffusion equations in equation 12, we find that the units of kf and kb are

cm , s

not

s−1 .

Including reactions far from the surface: Before proceeding further, one more point is now appropriate. The differential equation modeling electron transfer (equation 12) can be modified easily to include pre- and post- reactions. For the series of chemical reactions k

f A + e– − )− −* −B

kb

kr

B −−→ Z

,

where Z is an electronically inactive species, the differential equation in equation 12 simply becomes

∂cA ∂t ∂cB ∂t

2

x ∂ cA = DA − (kf cA − kb cB )δ(x) ∂x2 2

x ∂ cB = DB + (kf cA − kb cB )δ(x) − kr cB . ∂x2

(15a) (15b)

In this case, kf and kb have different units from kr ; the rate constants kf and kb have units of

cm s

and kr has units of s−1 . Note that the delta function δ(x) from equations 12 and 15

has units of cm-1 .

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2.1.3

Page 12 of 45

Ramp speed and measuring current

In order to model a linear sweep voltammetry or a cyclic voltammetry experiment, one must propagate the diffusion equations forward in time, while also changing the potential (∆G(t) in equations 6 and 11). For our purposes, we will imagine that the potential is ramped according to the following equation

∆G(t) = ∆Gi + νt

(16)

where ν is the scan rate and ∆Gi is the starting potential. For the cyclic voltammetry experiments, once ∆G(t) = ∆Gf = −∆Gi the scan is reversed and carried out until ∆G(t) = ∆Gi . Current in these experiments can be calculated in at least two equivalent ways: x I(t) = nFADA

∂c(t) |x=0 ∂x

(17a)

or

I(t) = nFA(kf (t)cA (t) − kb (t)cB (t))|x=0 .

(17b)

Here F is the Faraday constant, A is the area of the electrode surface, and n is the number of electrons transferred in the electron transfer. For simplicity, we set nFA = 1.

2.2 2.2.1

Electrochemical dynamics in two dimensions SH in 2D: explicitly modeling ET through a solvent reorganization coordinate

As described above, simple Butler-Volmer kinetics model electrochemical ET effectively through one parameter α; thus, the kinetics implicitly assume that ET dynamics are dictated by slow, rate-limiting motion along a generalized diffusion coordinate from bulk to the

12

ACS Paragon Plus Environment

Page 13 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

surface. While this approach is often reasonable in solution (for outer sphere ET), it is also true that dynamics are often limited by water reorganization and other solvent effects (as argued originally by Marcus 35 ). Thus, one can easily imagine at least two possible reaction coordinates. With this motivation in mind, and having already reviewed the basic theory of one-dimensional electrochemical dynamics for modeling CV experiments, we will now construct a multidimensional problem by including an additional degree of freedom into our Hamiltonian: an effective solvent reorganization coordinate ζ. Mathematically, we choose a Hamiltonian of the form from equation 9, with ~r = (x, ζ). For simplicity, we suppose that the PESs for species A and B are quadratic wells, with nuclear frequency ω, mass m, and centered at ζA (ζB ),

UA (x, ζ) = 12 mω 2 (ζA (x) − ζ)2

(18a)

UB (ζ) = 12 mω 2 (ζB − ζ)2 + ∆G

(18b)

E(x, ζ) = UB − UA = 12 mω 2 (ζB2 − ζA2 (x) − 2(ζB − ζA (x))ζ) + ∆G

(18c)

where UA can have an x dependence through the diabat minimum position ζA (x), and ∆G is determined by the external voltage. Note that UB is always independent of x. To model the electrochemical dynamics, we discretize our electrochemical system on a 2D grid, and we solve the following differential equations: Fx Fζ ζ ∂cA x ∂cA J~A = (−DA + A cA )ˆ x + (−DA + A cA )ζˆ ∂x γx ∂ζ γζ FBx FBζ ζ ∂cB x ∂cB ~ JB = (−DB + cB )ˆ x + (−DB + cB )ζˆ ∂x γx ∂ζ γζ

13

ACS Paragon Plus Environment

(19a)

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 45

2 2 ∂cA FAx ∂cA ∂FAx cA ζ ∂ cA x ∂ cA =DA − + D − A ∂t ∂x2 ∂ζ 2 γx ∂x ∂x γx ζ ζ F ∂cA ∂FA cA − A − − (kf cA − kb cB ) γζ ∂ζ ∂ζ γζ 2 2 ∂cB FBx ∂cB ∂FBx cB ζ ∂ cB x ∂ cB =DB − + D − B ∂t ∂x2 ∂ζ 2 γx ∂x ∂x γx ζ ζ F ∂cB ∂FB cB − + (kf cA − kb cB ) − B γζ ∂ζ ∂ζ γζ

(19b)

cA (t, x = −1, ζ) = cA (t, x = 0, ζ) (19c) cB (t, x = −1, ζ) = cB (t, x = 0, ζ). Dx , Dζ are diffusion coefficients in the x and ζ directions, respectively, γx , γζ are friction coefficients in the x and ζ directions, respectively (γx =

kT , Dx

γζ =

kT ), Dζ

and F x , F ζ are the

mean force acting on a given species in the x and ζ directions, respectively. From equation 18, the relevant forces are:

FAζ (x, ζ) = mω 2 (ζA (x) − ζ)

(20a)

FAx (x, ζ) = −mω 2 (ζA (x) − ζ) dζdxA

(20b)

FBζ (ζ) = mω 2 (ζB − ζ)

(20c)

FBx (ζ) = 0

(20d)

To visualize the effect of ∆G on the PESs, see figure 1. The addition of the ζ coordinate allows us to model electron transfer in the form of “hops” between the species, represented by hopping rates kf and kb . These hopping rates depend on the coordinate ζ, and are given by

kf (x, ζ) = kb (x, ζ) =

Γ(x) f (E(x, ζ)) ~

Γ(x) (1 ~

− f (E(x, ζ)))

Γ(x) = Γ0 δ(x)

14

ACS Paragon Plus Environment

(21a) (21b) (21c)

Page 15 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

where Γ(x) is the function that represents the strength of the electrode-molecule coupling and f (E) =

1 1+eβE

is the fermi function.

UB( G = 0.01) UB( G = 0) UA

UB( G = 0.01)

Figure 1: Illustration of the vertical shift in energy that occurs when changing the external voltage, and subsequently ∆G, on the PES UB from equation 18. The system is governed by the following initial conditions,

cA (t = 0, x, ζ) =

−βUA (x,ζ) R e −βU (x,ζ) A dζe

cB (t = 0, x, ζ) = 0

15

ACS Paragon Plus Environment

(22a) (22b)

The Journal of Physical Chemistry

0.005

UA UB UPMF UAbroad UBbroad

0.004 energy

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 45

0.003

cross

0.002 0.001 0.000

15

10

5

0

5

10

15

Figure 2: Potential energy surfaces as a function of the solvent reorganization coordinate ζ, unbroadened (UA and UB ) and broadened (UAbroad and UBbroad ), and potentials of mean force (UP M F ), for diabats in equation 18 and 29, with m = 2000, ω = 0.0002, ζA = −10, ζB = 10, Γ = 0.002, and ∆G = 0. All parameters are in atomic units (a.u.).

16

ACS Paragon Plus Environment

Page 17 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3: Potential energy surfaces as a function of the diffusion coordinate x and solvent reorganization coordinate ζ, with m = 2000, ω = 0.00025, ζA = −10, ζB = 10, Γ = 10−6 , and ∆G = 0. All parameters are in atomic units (a.u.).

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 45

and boundary conditions

cA (t, x = ∞, ζ) =

−βUA (x,ζ) R e −βU (x,ζ) A dζe

(23a)

cB (t, x = ∞, ζ) = 0

(23b)

cA (t, x, ζ = ±∞) = cB (t, x, ζ = ±∞) = 0,

(23c)

The initial conditions in equation 22 enforce the fact that only A is present initially, and at a uniform concentration everywhere in the x direction; furthermore, there is a Boltzmann distribution in the ζ direction. The boundary conditions in equation 23 enforce the fact that there is no concentration of either species at the boundaries of the ζ coordinate (ζ → ±∞) , and that the concentration of A and B are equal to their initial concentrations infinitely far from the electrode in the x direction (x → ∞). For simplicity, we fix the diffusion constant ζ ζ x x = DA = DB ). = DB to be constant and equal for both species A and B (DA

2.2.2

Nuances of broadening: a broadened classical master equation (BCME)

In cases where the metal molecule coupling Γ is larger than kT , level broadening becomes important: for molecules chemisorbed or physisorbed on surfaces, electrons are shared in an adiabatic state quantum mechanically between molecule and surface, and electronic motion is not simply hopping back and forth. To capture the more complex hybridized dynamics that arise from the continuum of electronic states in the metal, a reasonable fix is to modify our diabatic forces as follows: 36

FAx,broad (x, ζ) = −mω 2 (ζA (x) − ζ) dζdxA − FAζ,broad (x, ζ) = mω 2 (ζA (x) − ζ) −

dE (n(E(x, ζ)) dx

dE (n(E(x, ζ)) dζ

− f (E(x, ζ)))

− f (E(x, ζ)))

FBx,broad (x, ζ) = − dE (n(E(x, ζ)) − f (E(x, ζ))) dx FBζ,broad (x, ζ) = mω 2 (ζB − ζ) −

18

dE (n(E(x, ζ)) dζ

ACS Paragon Plus Environment

− f (E(x, ζ))).

(24a) (24b) (24c) (24d)

Page 19 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Here n(E) is the broadened population and W is the electronic bandwidth, Z

W

n(E) = −W

d Γ 2π ( − E)2 +

Γ2 4

f (),

(25)

chosen such that W  Γ. To visualize the surfaces from equation 18 and 24, see figure 2, and to visualize the 2D surfaces, see figure 3.

2.2.3

Measuring current in 2D

For all calculations where we explicitly model ζ, one can calculate current in two different ways: x =nFADA

I(t)

ζf

Z

dζ ζi

x ≈nFADA

I(t)

Nζ X i

∂cA (t, ζi ) |x=0 ∂x

∂cA (t, ζi ) dζ |x=0 ∂x

(26a)

or ζf

Z

dζ(kf (t, ζi )cA (t, ζi ) − kb (t, ζi )cB (t, ζi ))|x=0

I(t) =nFA ζi

I(t) ≈nFA

Nζ X

(26b) dζ(kf (t, ζi )cA (t, ζi ) − kb (t, ζi )cB (t, ζi ))|x=0 .

i

2.2.4

Coarse graining and marginalizing out ζ: The 1D Marcus-Hush (MH) model

Equation 19 represents a general approach to modeling electrochemical ET and yet one often seeks the simplest possible approach. To that end, one often seeks to marginalize out the reorganization coordinate direction, leading to the following marginal probability

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 45

distribution c(x, t): x ∂cA xˆ J~A = −DA ∂x x ∂cB J~B = −DB xˆ ∂x

(27a)

∂cA ~ · J~A − (kef cA − keb cB )δ(x) =−∇ ∂t 2 x ∂ cA =DA − (kef cA − keb cB )δ(x) ∂x2 ∂cB ~ · J~B + (kef cA − keb cB )δ(x) =−∇ ∂t 2 x ∂ cB =DB + (kef cA − keb cB )δ(x). ∂x2

(27b)

Obviously, equation 27 is equivalent to equation 12, although in this case the motion in the ζ direction gives a new average hopping rate, e kf /b , R∞ −βUA (0,ζ) f (E(0,ζ)) Γ0 −∞ Rdζ e ∞ −βUA (0,ζ) ~ −∞ dζ e R∞ −βU (ζ) B (1−f (E(0,ζ))) Γ0 −∞ dζR e . ∞ −βUB (ζ) ~ dζ e −∞

kef = keb =

(28a) (28b)

The rates kef , keb are just the original hopping rates weighted by the Boltzmann average of being at a given position in ζ at x = 0. Equation 28 arises from the assumption that equilibration in ζ is fast relative to other time scales in the simulation. For the Hamiltonian in equation 9 with roughly equal diffusion in x and ζ, the relevant condition for fast nuclear equilibration relative to the timescale of driving force is

mω 2 γ

=

Dmω 2 kT

>

ν . kT

This condition

states that the time required to change the driving force by kT should be larger than the relaxation time of a damped harmonic oscillator (which in the overdamped limit is

mω 2 ). γ

This condition should be satisfied for almost all experiments. Implementation of k˜f and k˜b as the rate constants for Butler-Volmer dynamics in equation 5 in lieu of the rate constants in equation 6 is known in the electrochemical literature as the Marcus-Hush (MH) boundary conditions. 37 Obviously, equations 28a and 28b are equivalent to the Marcus electrochemical rate constants; in this paper, even though we are in the slightly 20

ACS Paragon Plus Environment

Page 21 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

overdamped regime, the Zusman corrections (for solvent overdamping effects on the rate constant) should be small and will not be considered. Below, we will refer to the utilization of the rate constants k˜f and k˜b in the SH framework as a 1D MH model. Note that, if Γ is larger than kT , then the potentials UA and UB (within the BCME approximation) in the above equations need to be replaced with the broadened PESs,

UAbroad (0, ζ)

UBbroad (0, ζ)

2.3

1 = mω 2 (ζA (0) − ζ)2 + 2

Z

1 = mω 2 (ζB − ζ)2 + ∆G + 2

ζ

dζ 0 (n(E(0, ζ 0 )) − f (E(0, ζ 0 )))

ζ0

Z

dE dζ 0

ζ

dζ 0 (n(E(0, ζ 0 )) − f (E(0, ζ 0 )))

ζ0

(29a)

dE . dζ 0

(29b)

Alternative to solving ODE: diagonalizing the Fokker-Planck operator

The simplest means to propagate the electrochemical dynamics in equations 12, 19, and 27 is to use an implicit ordinary differential equation (ODE) solver. Note that an implicit ODE solver is necessary because a more conventional explicit ODE solver (e.g. RK4 38 ) is unstable for this purpose. The root of the problem is the stiff behavior of these ODEs; 39 for the 1D data in the results and discussion, we have chosen to use the Real-valued Variable-coefficient Ordinary Differential Equation (’vode’) solver with the backward difference formula (BDF) method. 40 As an alternative to ‘vode’, there is another approach: namely, one can cast

∂cA ∂t

( ∂c∂tB )

as a matrix equation and diagonalize the corresponding matrix operator to obtain an exact answer for the time evolution of cA (cB ) 3 . Diagonalization of this ‘Fokker-Planck’ operator has been previously applied to study ET dynamics in a two state system with diffusion. 41 3

Note that the mathematics in this section can be safely skipped by the reader who is not concerned about the stability of diagonalizing a non-Hermitian matrix. For the practical reader who wants to repeat our calculations, however, these details will be essential.

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 45

For our purposes, equation 12 can be written in the following way: ∂cA = LxA cA − (kf cA − kb cB )δ(x) ∂t

(30a)

∂cB = LxB cB + (kf cA − kb cB )δ(x). ∂t

(30b)

Here, we define superoperators:

and we allow for a drift term

Fx γx

LxA =

∂ Fx x ∂ (DA − A) ∂x ∂x γx

(31a)

LxB =

Fx ∂ x ∂ (DB − B ), ∂x ∂x γx

(31b)

for generality; note that, for the case of equation 12, F x ≡ 0.

Similarly, equation 19 can be written as: ∂cA = (LxA + LζA )cA − (kf cA − kb cB )δ(x) ∂t

(32a)

∂cB = (LxB + LζB )cB + (kf cA − kb cB )δ(x) ∂t

(32b)

∂ Fζ ζ ∂ (DA − A) ∂ζ ∂ζ γζ

(32c)

FBζ ∂ ζ ∂ (D − ). = ∂ζ B ∂ζ γζ

(32d)

LζA = LζB

By combining LxA , LxB , LζA , LζB , kf , and kb we obtain the general matrix equation for both cA and cB , 









kb δ(x)  cA  c˙A  LA − kf δ(x)  =    + ~y cB c˙B kf δ(x) LB − kb δ(x) ~c˙ = M~c + ~y

22

ACS Paragon Plus Environment

(33)

Page 23 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

where LA = LxA (LB = LxB ) in the 1D model and LA = LxA + LζA (LB = LxB + LζB ) in the 2D model. Here, the term ~y is used to include the bath conditions from equations 13 and 23 which maintain a static condition at the bath boundary, i.e. ~c˙(x = ∞) = ~0. In other words, ~y (x 6= ∞) = 0 and ~y (x = ∞) = −(M~c)(x = ∞). Obviously, M is in general non-Hermitian, given that (i) LA (LB ) is non-Hermitian if there is a potential UA (UB ), and (ii) the off-diagonal elements of M are not equal (kf 6= kb ) unless E = 0. Since diagonalization of non-Hermitian matrices is often unstable, one would like to transform M into a Hermitian matrix. To do so, there is a common method 42 for e A (L e B ), converting the non-Hermitian operator LA (LB ) to a Hermitian operator L

eA = e L

eB = e L

βUA 2

βUB 2

LA e−

βUA 2

LB e−

βUB 2

(34a)

.

(34b)

With the above transformation of LA (LB ) in mind, we can define a similarity transformation matrix S as

 e S=



βUA 2

0  ,

(35)

  βE − e LA − kf δ(x) kb δ(x)e 2  =  βE e B − kb δ(x) kf δ(x)e 2 L

(36)

0

e

βUB 2

f such that the new operator M

f = SMS −1 M

f will differ from is Hermitian and can be easily diagonalized. While the eigenvectors of M f are related by a similarity transformation, so that the eigenvalues are those of M, M and M identical for both and one can analyze the dynamics of the system at any arbitrary time. Note that, in equation 36 for a 2D simulation, kf , kb , and E are functions of the coordinates x and ζ.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 45

Simulating a linear sweep voltammetry experiment can be done with matrix algebra by f at each time step (since M f is a function of ∆G(t)), yielding M f = diagonalization of M P ΛP T . The eigenvalues Λ are always nonpositive, corresponding to decay rates along modes represented by the eigenvectors of P , and for such a dissipative system, there is always one zero eigenvalue corresponding to the equilibrium, or steady state4 . At this point, we can change variables and recast the differential equation for the time evolution of cA and cB into renormalized coordinates: ~b ≡ P T S~c

(37a)

~z ≡ P T S~y .

(37b)

~b˙ = Λ~b + ~z.

(37c)

The final equation of motion becomes:

Equation 37c can be solved exactly, with solution for any arbitrary time step given by

~b(t) = ~beq + (~b − ~beq )eΛt

(38a)

~beq = −Λ−1~z.

(38b) 



 cA  Finally, using equation 37a, we can convert ~b(t) back to ~c(t) =   using P and S: cB ~c = S −1 P~b. Hence, one can easily calculate the current at each time step with complete numerical stability. 4

In this paper, if we don’t use enough grid points, we must sometimes force the lowest eigenvalue to be equal to zero for numerical stability.

24

ACS Paragon Plus Environment

Page 25 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2.4

Parameterization

We will work in standard units throughout this paper and, unless stated otherwise, set ∆G◦ = 0 V, m = 1.82 · 10−27 kg (2000 a.u.), T = 300K (0.00095 a.u.), ζA (x) = ζA0 = −5.29 ˚ A (-10 a.u.), ζB = 5.29 ˚ A (10 a.u.), W = 103 · Γ, Nx = 50, and Nζ = 101. The size of the grid √ in x is given by Lx = 6 Dx tT , where tT is the total time of the simulation, defined by the starting and ending ∆G and ν, tT =

|∆Gf −∆Gi | ν

for linear sweeps and t =

2|∆Gf −∆Gi | ν

for cyclic

sweeps. The size of the grid in ζ is given by Lζ = 4(ζB − ζA ), with endpoints [4ζA , 4ζB ], and the grid spacing in both x and ζ is uniform, such that dx =

Lx Nx

and dζ =

Lζ . Nζ

The current

values in figures 5, 6, 7, 8, 10 and 11 are rescaled by a factor of 109 .

3

Results

We have thus far presented several distinct approaches for simulating electrochemical systems. To better understand these approaches, we will now consider a host of different situations and quantitatively analyze the prediction of each approach. When will different approaches agree? And if they disagree, which approach is appropriate?

3.1

Verification of 1D model

Our first goal is to numerically demonstrate the equivalence of surface hopping with boundary conditions (equations 12 and 13) against the Nernstian (equations 1 and 4) and Butler Volmer (equations 1 and 5) approaches. In figure 4 we plot simulated cyclic voltammograms for two chemical reaction types, with both methods for calculating current (equations 17(a)(b)) and three different dynamics methods; Nernstian (equations 1 and 4), Butler-Volmer (equations 1 and 5), and surface hopping (equations 12 and 13). In all cases, we see excellent agreement between the surface hopping dynamics and the standard Nernstian and ButlerVolmer approaches (however current is calculated). In addition to the case of a symmetric barrier between oxidation and reduction, the 25

ACS Paragon Plus Environment

The Journal of Physical Chemistry

SH, Eq. 17(a) BV, Eq. 17(a) Nernst, Eq. 17(a)

I(t)

SH, Eq. 17(b) BV, Eq. 17(b)

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2 1 (a) 0 1 2 1 3

0

1

2 1 (b) 0 1 2 1 3

2 (c)

2 (d)

1

1

0

1

0 G(t)

Page 26 of 45

1

0

1

0

1

0 G(t)

1 k

f −− Figure 4: 1D Cyclic voltammograms (CV) for the reaction A + e– ) −* − B, both with ((c)

kb

k

and (d)) and without ((a) and (b)) the post reaction B −−→ Z. (a) CV measuring current k by the difference in forward and backward rates, without the post reaction B −−→ Z. (b) k CV measuring current by the flux of A at the electrode, without the post reaction B −−→ Z. (c) CV measuring current by the difference in forward and backward rates, with the post k reaction B −−→ Z. (d) CV measuring current by the flux of A at the electrode, with the k post reaction B −−→ Z. Note that kf and kb are not explicitly modeled when using Nernst boundary conditions, preventing the use of the difference in forward and backward rates as a measure of current. Parameters are set to Dx = 1 cm2 /sec, ν = 1 V/sec, ∆Gi = 1 V, ∆Gf = −1 V, Nx = 200, k = 100 sec-1 , k ◦ = 1000 cm/sec, α = 0.5. We see quantitative agreement between the surface hopping dynamics and both Nernstian and Butler-Volmer dynamics, suggesting that all the approaches are applicable for simulating these systems. surface hopping method can also handle the case of an asymmetric barrier, α 6= 0.55 . In figure 5 we plot results for the same models and parameterizations as in figure 4, but with α = 0.2. 5

The transfer coefficient α represents the slope of the log current with respect to the external potential, 26 ◦ ∆G, at zero external potential, i.e. α = −kT ∂log(I) ∂∆G |∆G=∆G

26

ACS Paragon Plus Environment

Page 27 of 45

SH, Eq. 17(a) BV, Eq. 17(a) Nernst, Eq. 17(a)

I(t)

SH, Eq. 17(b) BV, Eq. 17(b)

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

2 1 (a) 0 1 2 1 3

0

1

2 1 (b) 0 1 2 1 3

2 (c)

2 (d)

1

1

0

1

0 G(t)

1

0

1

0

1

0 G(t)

1

Figure 5: 1D Cyclic voltammograms (CV) for the same conditions as in figure 1, but with an asymmetric barrier between oxidation and reduction, given by α = 0.2. Parameters are set to Dx = 1 cm2 /sec, ν = 1 V/sec, ∆Gi = 1 V, ∆Gf = −1 V, Nx = 200, k = 100 sec-1 , k ◦ = 1 cm/sec, α = 0.2. Note that k ◦ is decreased relative to the symmetric case; if the kinetics are very facile, as evinced by a large k ◦ , then the asymmetry of the barrier will have no effect on the dynamics. Similarly to the α = 0.5 case, we see quantitative agreement between the hopping approach and the Butler-Volmer approach. Note that Nernstian dynamics cannot recover the correct answer, since we have introduced an asymmetric barrier and relatively slow ET kinetics in this case. Obviously, we do not expect Nernstian dynamics to recover the correct current due to the asymmetric barrier, but we do see quantitative agreement between the hopping approach and Butler-Volmer kinetics, providing additional validation for the hopping approach.

27

ACS Paragon Plus Environment

The Journal of Physical Chemistry

3

I(t)

2

= 10 6a. u. 3

2D, Eq. 19 1D, Eq. 12 1D, Eq. 27

2

(a)

1 0 3 2

2

0.5

0.0

(c)

1 0

= 10 5a. u. (b)

1 0 0.5 = 10 4a. u. 3

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 45

0.5

0.0

0.5 = 10 3a. u.

(d)

1 0.5

0.0 G(t)

0

0.5

0.5

0.0 G(t)

0.5

Figure 6: Linear sweep voltammograms for simulations where the 1D MH model (equation 27) and full 2D model (equation 19) agree, with an intermediate reorganization energy. ∆Gi = 0.816 V, ∆Gf = −0.816 V, D = 5.5·10−4 cm2 /sec, ω = 1.03·1013 sec-1 , λ0 = 26.32kT , and ν = 112 V/sec for each subplot. (a) Γ~ = 1.40 · 107 cm/sec, (b) Γ~ = 1.40 · 108 cm/sec, (c) Γ~ = 1.40 · 109 cm/sec, (d) Γ~ = 1.40 · 1010 cm/sec. These plots suggest that utilizing the 1D MH model in lieu of the 2D model is appropriate regardless of coupling strength for moderate reorganization energies, although at lower coupling strengths the regular 1D model (equation 12) fails and does not agree with the 2D model. Note that the green and blue traces representing the 1D MH and 2D models, respectively, seem indistinguishable due to the quantitative agreement between the two methods.

3.2

Cases where 2D and 1D MH model agree

Next, we examine situations under which the 1D MH model can recover the same result as that of the full 2D model. Plotted in figures 6 and 7 are linear sweep voltammograms for cases where the 1D MH model agrees with the 2D model, for four different coupling strengths and two different reorganization energies. In the case of moderate reorganization

28

ACS Paragon Plus Environment

Page 29 of 45

3

I(t)

2

= 10 6a. u. 3

2D, Eq. 19 1D, Eq. 12 1D, Eq. 27

2

(a)

1 0 3 2

2

0.5

0.0

(c)

1 0

= 10 5a. u. (b)

1 0 0.5 = 10 4a. u. 3

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.5

0.0

0.5 = 10 3a. u.

(d)

1 0.5

0.0 G(t)

0

0.5

0.5

0.0 G(t)

0.5

Figure 7: Linear sweep voltammograms for simulations where the 1D MH model (equation 27) and full 2D model (equation 19) agree, with a moderately large reorganization energy. ∆Gi = 0.816 V, ∆Gf = −0.816 V, D = 5.5·10−4 cm2 /sec, ω = 1.24·1013 sec-1 , λ0 = 37.89kT , and ν = 112 V/sec for each subplot. (a) Γ~ = 1.40 · 107 cm/sec, (b) Γ~ = 1.40 · 108 cm/sec, (c) Γ = 1.40 · 109 cm/sec, (d) Γ~ = 1.40 · 1010 cm/sec. These plots suggest that utilizing the 1D ~ MH model in lieu of the 2D model is appropriate regardless of coupling strength for large reorganization energies. Again, as in figure 6, the regular 1D model is valid only when the coupling strength is large ( Γ~ > ω). energies (figure 6), all three methods (1D, 1D MH, and 2D) give the same results for large enough Γ, but as λ0 increases the agreement between the 1D model and the 1D MH/2D model decreases (figure 7). The regime under which the 1D, the 1D MH and 2D models all agree can be considered the case where the dynamics are mass transfer limited, i.e. the dynamics along the x coordinate are rate limiting. By contrast, the high reorganization energy/low coupling regime can be understood as the case where the dynamics are limited by motion along the solvent 29

ACS Paragon Plus Environment

The Journal of Physical Chemistry

reorganization coordinate ζ. Nevertheless, from the fact that the 1D MH model always recaptures the 2D results, it appears the 1D MH model may be adequate for capturing the dynamics of electrochemical systems in all cases. And yet, there should be instances where the average hopping rate of the 1D MH model should not be adequate for modeling the full dynamics of the 2D system. Let us now address this hypothesis.

3.3

Cases where 2D and 1D MH don’t agree

3

I(t)

2

= 10 6a. u. 3

2D, Eq. 19 1D, Eq. 12 1D, Eq. 27

2

(a)

1 0

2

0.5

0.0

0 0.5 = 10 4a. u. 3 2

(c)

1 0

= 10 5a. u. (b)

1

3

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 45

0.5

0.0

0.5 = 10 3a. u.

(d)

1 0.5

0.0 G(t)

0

0.5

0.5

0.0 G(t)

0.5

Figure 8: Linear sweep voltammograms for simulations where the 1D MH model (equation 27) and full 2D model (equation 19) do not agree, with ζA0 = −5.82 ˚ A and ζA∞ = −5.29 10 ˚ A in equation 39. ∆Gi = 0.816 V, ∆Gf = −0.816 V, σ = dx2 , D = 5.5 · 10−4 cm2 /sec, ω = 1.03 · 1013 sec-1 , λ0 = 29.01kT , and ν = 112 V/sec for each subplot. (a) Γ~ = 1.40 · 107 cm/sec, (b) Γ~ = 1.40 · 108 cm/sec, (c) Γ~ = 1.40 · 109 cm/sec, (d) Γ~ = 1.40 · 1010 cm/sec. By shifting ζA in the 2D model, and breaking separability of x and ζ in the Hamiltonian, we see that the 1D MH and 2D models can yield somewhat different results.

30

ACS Paragon Plus Environment

Page 31 of 45

3

I(t)

2

= 10 6a. u. 3

2D, Eq. 19 1D, Eq. 12 1D, Eq. 27

2

(a)

1 0

2

0.5

0.0

0 0.5 = 10 4a. u. 3 2

(c)

1 0

= 10 5a. u. (b)

1

3

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.5

0.0

0.5 = 10 3a. u.

(d)

1 0.5

0.0 G(t)

0

0.5

0.5

0.0 G(t)

0.5

Figure 9: Linear sweep voltammograms for simulations where the 1D MH model (equation A and ζA∞ = −5.29 27) and full 2D model (equation 19) do not agree at all, with ζA0 = −6.35 ˚ 10 −4 ˚ A in equation 39. ∆Gi = 0.816 V, ∆Gf = −0.816 V, σ = dx cm2 /sec, 2 , D = 5.5 · 10 Γ 13 -1 ω = 1.03 · 10 sec , λ0 = 31.84kT , and ν = 112 V/sec for each subplot. (a) ~ = 1.40 · 107 cm/sec, (b) Γ~ = 1.40 · 108 cm/sec, (c) Γ~ = 1.40 · 109 cm/sec, (d) Γ~ = 1.40 · 1010 cm/sec. By shifting ζA even more at x = 0, we see a complete breakdown of the 1D MH model as far as replicating the 2D model results. In order to construct a situation where the 1D MH and 2D models differ, let us choose nonseparable PESs, where ζ and x are entangled. Since the 1D MH model simulates dynamics only in x, we suspect that entanglement of x and ζ should give rise to different results compared to the 2D model. To do this, instead of using ζA (x) = ζA0 (in equation 18) as a constant value, we choose the functional form 2

ζA (x) = (ζA0 − ζA∞ )e−σx + ζA∞ ,

31

σ

ACS Paragon Plus Environment

1 . (dx)2

(39)

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

This creates a situation where the diabat well minimum for species A takes on the value ζA0 at x = 0 and the value ζA∞ for all other x. The switch off occurs over a length scale much smaller than the grid spacing dx. Consider now a situation where ζA∞ = −5.29 ˚ A for all x values except at the electrode surface. When species A approaches the metal from x > 0, necessarily there must be additional relaxation as the solute concentration cA adjusts to the new surface at x = 0. Moreover, this adjustment cannot be measured by a simple difference in reorganization energy at x = 0. In such cases, the solute concentrations cA and cB need not relax faster than the timescales kf and kb . Plotted in figures 8 and 9 are linear sweep voltammograms for different ζA positions at x = 0 for four different coupling strengths. While the difference between the 1D MH and 2D models is modest in figure 8 (ζA (x = 0) = −5.82 ˚ A), the difference becomes readily apparent in figure 9 (ζA (x = 0) = −6.35 ˚ A). Effectively, we find that the 1D MH model is adequate for situations where the two dimensions are separable, but the 1D MH model fails when separability is no longer guaranteed. Now, one may well question whether the data in figures 8 and 9 is anomalous and caused by a discontinuous PES. To that end, in Appendix B we study smaller σ values for ζA (x) and investigate model Hamiltonians that smoothly bridge the gap between separable and inseparable. We find that, even with smooth PESs, the 1D MH model still fails without separability. In the future, understanding the nature of the PESs and the separability of the diffusion and solvent reorganization coordinates will be essential: after all, the reorganization energy for an ion must depend strongly on proximity to the metal surface.

4 4.1

Discussion Electrostatic interactions

At this point, we have shown that, in many cases, 1D simulations of electrochemical voltammetry experiments are appropriate – unless there is significant non-separability between the 32

ACS Paragon Plus Environment

Page 32 of 45

Page 33 of 45

(a)

L = 12 Dt 0 L = 6 Dt 0 0.2 L = 3 Dt 0 Ramp off 0.1

0.1 0.0

0

1.00

200 t L = 12 Dt 0 L = 6 Dt 0 L = 3 Dt 0

0.75 0.50 0.25 0.00

0

20

x

40

0.0 19600 19800 20000

400

1.00

(b)

C(t = 300)

I(t)

0.2

CA(t = 20000)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.75

CA CB

0.50 0.25 0.00

60

(c)

0

20

x

40

60

Figure 10: Linear sweep voltammogram and concentration profile for a simulation that is allowed to run for long times after completing the sweep, for different length grids (xf − x0 ). D = 1, ν = 0.01, and t0 = 300. t0 is the time at which the potential ramp is turned off. √ (a) Dt0 , Linear sweep voltammograms for the case where the simulation grid length is set to 12 √ √ 0 0 6 Dt and 3 Dt . The dotted line is the point at which the potential is no longer ramped and held at that value for the remainder of the simulation. (b) Concentration profiles for √ A at the end of the simulation, from x = 0 to x = 3 Dt0 . (c) Concentration profiles √ for A and B at the time at which the potential ramp is turned off, from x = 0 to x = 3 Dt0 . At short times all three grid spacings provide the same concentration and current profile, but at long times begin to diverge. The effect of grid length on the outcome of these steady-state simulations suggests that accurate modeling of electrostatic interactions between particles is necessary for a complete understanding of the dynamics of these systems at long times. diffusion and solvent reorganization coordinates. That being said, there is still one additional important consideration that all the simulations in this work do not account for: electrostatic interactions between charged particles (redox species and/or electrolyte ions). These effects can have a measurable effect on the dynamics and subsequent I-V curve for a given

33

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 45

simulation. The importance of accounting for these interactions can be ascertained by investigating the concentrations and current profile in a simulation that is allowed to run for long times after reaching the final voltage in a linear sweep experiment. To that end, consider figure 10 where dx is fixed but we choose a different number of grid points. We present 1D linear sweep voltammograms using the surface hopping approach. We find that, at long times, the concentration profiles of cA and cB differ greatly as a function of the length of the 1D-grid, which is also detectable through a slight difference in the steadystate current between the simulations. In fact, a simple analysis shows that the steady state current using equation 17(a) is ∂css A |x=0 ∂x 0 ss x cA (x = dx) − cA Iss = nFADA , dx

x Iss = nFADA

(40)

where c0A is the steady state concentration of A at the electrode surface, whose ratio is almost (but not exactly) equal to the ratio of backward and forward hopping rates, c0A kb = + κ = eβ∆G + κ, 0 cB kf

(41)

where κ is a small number representing the deviation of the steady state concentrations at the electrode surface from the Nernstian value. To develop an analytical expression for Iss , we note that, since the diffusion constants of both A and B are equal, the flux of A and B are equal and that cA + cB = cbulk + cbulk A B everywhere in the simulation grid. In addition, we know that the long time concentration profile of cA and cB must be linear (satisfying the condition

∂css ∂t

2 ss

= Dx ∂∂xc2 = 0),

css A (x) =

cbulk − c0A A x + c0A Lx

(42a)

css B (x) =

cbulk − c0B B x + c0B . Lx

(42b)

34

ACS Paragon Plus Environment

Page 35 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Finally, substituting equation 42a into equation 40 yields a steady state concentration of

Iss =

x bulk x bulk nFADA (cA − c0A ) nFADA cA e−β∆G ≈ , Lx Lx 1 + e−β∆G

(43)

1 bulk since c0A ≈ cbulk = 0. A 1+e−β∆G when cB

A similar analysis can be done using the difference in forward and backward rates, equation 17(b), for the current: ss Iss =nFA(kf css A − kb cB )|x=0

(44)

nFAΓ −αβ∆G 0 e (cA − c0B eβ∆G ). Iss = ~ −β∆G

e Insertion of equation 41 for c0A into 44 (and assuming c0B ≈ cbulk A 1+e−β∆G ) yields

nFAΓ −αβ∆G 0 β∆G e (cB (e + κ) − c0B eβ∆G ) ~ nFAΓ −αβ∆G 0 nFAΓ −αβ∆G bulk e−β∆G Iss = e κcB ≈ e κcA . ~ ~ 1 + e−β∆G

Iss =

Comparison of equations 43 and 45 shows that κ =

~Dx eαβ∆G ΓLx

(45)

for the case where cbulk = 0, B

such that c0B = cbulk − c0A . A Figure 10 and equation 43 demonstrate that, in the long time limit, where mass diffusion is rate-limiting, the steady-state current is actually independent of coupling strength6 . That being said, figure 10 and equation 43 also show a fundamental artifact of all electrochemical simulations using equations 17 and 26; steady state current should not depend on the choice of grid. The problem of grid length has been previously recognized for unequally spaced grids, 43–45 and the general boundary flux problem is known to have a simulation box size dependence. 46–48 Looking forward, our belief is that an accurate treatment of electrostatic 6

Note that this long time simulation is not equivalent to a Tafel measurement. For the present paper, we allow the concentrations of A and B to change at the electrode surface subject to diffusion, as to study the effect of mass transfer on the voltammetry curves. By contrast, for a Tafel plot, one works under the assumption that the concentrations of A and B are largely fixed at the electrode surface and are not affected by diffusion

35

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 45

interactions in simulations should mitigate the aberrant behavior shown in figure 10, such that the steady state solution for cA (x) and cB (x) will not necessarily be linear between 0 and L (equation 42), and our modeling should be more accurate at long times. Including these electrostatic interactions within the surface hopping framework will be the focus of ongoing work.

4.2

Coarse-graining ab initio rate constants

The final issue we would like to address in this paper is the question of how we might parameterize the rate constants kf and kb in equations 6, 21, and 28. Throughout this paper, we have assumed that Γ is equal to a certain value at x = 0 and zero for all other x. Of course, in reality, Γ should vary smoothly in the x direction. Thus, how should we construct Γ(x) in practice? In figure 11, we carried out simulations using equation 12 but on a much denser grid (Nx = 500) where Γ is a smooth function of x: x

Γ(x) = Γ00 e− λ .

(46)

In figure 11 we also plot electrochemical simulations with Nx = 50 using Γ(x) = Γ0 δ(x), where Γ0 ≡ Γ00 λ. From the results shown in figure 11, simulations with dense or sparse grids yield equivalent answers, as long as λ is small enough such that there is no appreciable ability to hop beyond the first grid point in the more sparse simulation. Thus, altogether, the data in figure 11 suggests that if one can compute an ab initio reaction rate kf (x) for a species near a metal surface, one needs only to calculate Z kf =



kf (x)dx

(47)

0

in order to parameterize a 1D model for electrochemical dynamics (x = 0 represents the metal surface). We remind the reader that the final kf value will in fact have units of 36

ACS Paragon Plus Environment

cm s

Page 37 of 45

I(t)

3 2 1 0

I(t)

3

= 1.06 10 4cm

= 0 (x) = 00e x

(a) 0.5

0.0

1 3

(b) 0.5

0.0

0

0.5

= 1.06 10 6cm

2 1

0.5

= 1.06 10 5cm

2 0

I(t)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(c) 0.5

0.0 G

0.5

Figure 11: Linear sweep voltammograms for simulations with two different types of x dependence for Γ, Γ(x) = Γ00 e−λx (equation 46), Γ(x) = Γ0 δ(x) (equation 12), and we fix Γ0 Γ0 = λ0 . ∆Gi = 0.816 V, ∆Gf = −0.816 V, D = 5.5 · 10−4 cm2 /sec, Γ~0 = 1.40 · 108 cm/sec, x and ν = 112 V/sec for each subplot. Nx = 500 for the Γ(x) = Γ00 e− λ model and Nx = 50 for the Γ(x) = Γ0 δ(x) model. (a) λ = 1.06·10−4 cm, (b) λ = 1.06·10−5 cm (c) λ = 1.06·10−6 cm. The results in (a) show the case where appreciable hopping occurs past the first grid point in the more sparse simulation, illustrating the breakdown of the δ-function approximation. Overall, this data confirms how one should construct Γ0 for a sparse grid so as to match ab initio rate constants which drop off with length λ1 from the surface. compared to the ab initio reaction rate kf (x) which has units of s−1 . This simple relationship in equation 47 opens up many avenues for future explorations.

37

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

5

Conclusion

In this paper we have 1) shown the equivalence of the surface hopping approach with traditional electrochemical methods and 2) quantified instances under which a 2D surface hopping model can or cannot be marginalized to a simpler 1D MH model. Regarding the latter question, we find the key consideration is one of separability: If the solvent reorganization coordinate (ζ) and the diffusion coordinate (x) are nearly separable, a 1D MH model is good enough, and when compared with a multidimensional approach, the 1D MH is far more reliable than Butler-Volmer or Nernstian dynamics. Nevertheless, if there is no separability between ζ and x, a 2D treatment is necessary and can be easily accomplished with a multidimensional SH approach. Looking forward, if one seeks to model I-V curves under external potentials, the new approaches presented here should be of great value. After all, current voltammetry simulation packages, such as DigiElch, 49 usually assume 1D systems with predetermined rate constants, chosen as linearized Marcus rates (as one would usually associate with Butler-Volmer dynamics) or full Marcus rates (which, if calculated properly, should come from averaging over solvent fluctuations, as in equation 28). By contrast, our guiding principle is that the time has come to utilize modern nonadiabatic dynamics algorithms to study electrochemical ET for arbitrary PESs. Using the simple analysis in the discussion section, our hope is that, in the near future, we will be able to simulate unknown mechanistic pathways for unexplained chemical systems, with realistic parameters taken from ab initio DFT studies.

6

Acknowledgments

We thank Abe Nitzan, Alexander Soudackov, and Adam Willard for very helpful conversations. This material is based upon work supported by the Air Force Office of Scientific Research (AFOSR) under award number FA9550-18-1-0420.

38

ACS Paragon Plus Environment

Page 38 of 45

Page 39 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

7

Supporting Information

Information pertaining to numerical simulation details for the diagonalization of the FokkerPlanck operator, further analysis of nonseparable 2D Hamiltonians, and how to determine surface concentrations in traditional 1D Nerstian and Butler-Volmer simulations can be found in the Supporting Information. This material is available free of charge via the Internet at http://pubs.acs.org.

References (1) Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Fundamentals and Applications; Wiley: New York, 2001. (2) Yan, D.; Bazant, M. Z.; Biesheuvel, P. M.; Pugh, M. C.; Dawson, F. P. Theory of linear sweep voltammetry with diffuse charge: Unsupported electrolytes, thin films, and leaky membranes. Phys. Rev. E 2017, 95 . (3) Smalley, J. F.; Chidsey, C. E. D.; Creager, S. E.; Ferraris, J. P.; Finklea, H. O.; Chalfant, K.; Zawodzinsk, T.; Newton, M. D.; Feldberg, S. W.; Linford, M. R. Heterogeneous Electron-Transfer Kinetics for Ruthenium and Ferrocene Redox Moieties through Alkanethiol Monolayers on Gold. J. Am. Chem. Soc. 2003, 125, 2004–2013. (4) Feldberg, S. W. Implications of marcus-hush theory for steady-state heterogeneous electron transfer at an inlaid disk electrode. Anal. Chem. 2010, 82, 5176–5183. (5) Henstridge, M. C.; Laborda, E.; Rees, N. V.; Compton, R. G. Marcus-Hush-Chidsey theory of electron transfer applied to voltammetry: A review. Electrochim. Acta 2012, 84, 12–20. (6) Laborda, E.; Henstridge, M. C.; Batchelor-McAuley, C.; Compton, R. G. Asymmetric MarcusHush theory for voltammetry. Chem. Soc. Rev. 2013, 42, 4894.

39

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(7) Matyushov, D. V.; Newton, M. D. Q-model of electrode reactions: altering force constants of intramolecular vibrations. Phys. Chem. Chem. Phys. 2018, 20, 24176–24185. (8) Oldham, K. B.; Myland, J. C. On the evaluation and analysis of the Marcus-HushChidsey integral. J. Electroanal. Chem. 2011, 655, 65–72. (9) Migliore, A.; Nitzan, A. On the evaluation of the Marcus-Hush-Chidsey integral. J. Electroanal. Chem. 2012, 671, 99–101. ´ (10) Henstridge, M. C.; Laborda, E.; Wang, Y.; Suwatchara, D.; Rees, N.; Molina, A.; Mart´ınez-Ortiz, F.; Compton, R. G. Giving physical insight into the ButlerVolmer model of electrode kinetics: Application of asymmetric MarcusHush theory to the study of the electroreductions of 2-methyl-2-nitropropane, cyclooctatetraene and europium(III) on mercury microelectrodes. J. Electroanal. Chem. 2012, 672, 45–52. (11) Henstridge, M. C.; Laborda, E.; Dickinson, E. J. F.; Compton, R. G. Redox systems ˇ c´ık equaobeying Marcus-Hush-Chidsey electrode kinetics do not obey the Randles-Sevˇ tion for linear sweep voltammetry. J. Electroanal. Chem. 2012, 664, 73–79. (12) Tanner, E. E.; Xiong, L.; Barnes, E. O.; Compton, R. G. One electron oxygen reduction in room temperature ionic liquids: A comparative study of Butler-Volmer and Symmetric Marcus-Hush theories using microdisc electrodes. J. Electroanal. Chem. 2014, 727, 59–68. (13) Okajima, Y.; Shibuta, Y.; Suzuki, T. A phase-field model for electrode reactions with Butler-Volmer kinetics. Comput. Mater. Sci. 2010, 50, 118–124. (14) Sk´ ulason, E.; Tripkovic, V.; Bj¨orketun, M. E.; Gudmundsd´ottir, S.; Karlberg, G.; Rossmeisl, J.; Bligaard, T.; J´onsson, H.; Nørskov, J. K. Modeling the electrochemical hydrogen oxidation and evolution reactions on the basis of density functional theory calculations. J. Phys. Chem. C 2010, 114, 18182–18197.

40

ACS Paragon Plus Environment

Page 40 of 45

Page 41 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(15) Roldan, A. Frontiers in first principles modelling of electrochemical simulations. Curr. Opin. Electrochem. 2018, 10, 1–6. (16) Zusman, L. D. Outer-sphere electron transfer in polar solvents. Chem. Phys. 1980, 49, 295–304. (17) Rips, I.; Jortner, J. Dynamic solvent effects on outer-sphere electron transfer. J. Chem. Phys. 1987, 87, 2090–2104. (18) Straub, J. E.; Berne, B. J. A statistical theory for the effect of nonadiabatic transitions on activated processes. J. Chem. Phys. 1987, 87, 6111–6116. (19) Matyushov, D. V.; Schmid, R. A molecular treatment of solvent effects on intervalence electron transfer. J. Phys. Chem. 1994, 98, 5152–5159. (20) Dorfman, R. C.; Fayer, M. D. The influence of diffusion on photoinduced electron transfer and geminate recombination. J. Chem. Phys. 1992, 96, 7410–7422. (21) Gladkikh, V. S.; Burshtein, A. I.; Tavernier, H. L.; Fayer, M. D. Influence of diffusion on the kinetics of donor-acceptor electron transfer monitored by the quenching of donor fluorescence. J. Phys. Chem. A 2002, 106, 6982–6990. (22) Gladkikh, V.; Burshtein, A. I.; Feskov, S. V.; Ivanov, A. I.; Vauthey, E. Hot recombination of photogenerated ion pairs. J. Chem. Phys. 2005, 123, 4894. (23) Bai, P.; Bazant, M. Z. Charge transfer kinetics at the solidsolid interface in porous electrodes. Nat. Commun. 2014, 5, 3585. (24) Reed, S. K.; Madden, P. A.; Papadopoulos, A. Electrochemical charge transfer at a metallic electrode: A simulation study. J. Chem. Phys. 2008, 128, 124701–184102. (25) Tully, J. C. Molecular dynamics with electronic transitions. J. Chem. Phys. 1990, 93, 1061–1071. 41

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(26) Schmickler, W.; Santos, E. Interfacial Electrochem.; Springer-Verlag Berlin Heidelberg, 2010. (27) Randles, J. E. A cathode ray polarograph. Part II. - The current-voltage curves. Trans. Faraday Soc. 1948, 44, 327–338. ˇ c´ık, A. Oscillographic polarography with periodical triangular voltage. Collect. (28) Sevˇ Czechoslov. Chem. Commun. 1948, 13, 349–377. (29) Nicholson, R. S.; Shain, I. Theory of Stationary Electrode Polarography: Single Scan and Cyclic Methods Applied to Reversible, Irreversible, and Kinetic Systems. Anal. Chem. 1964, 36, 706–723. (30) Matsuda, H.; Ayabe, Y. Zur Theorie der RandlesSevˇcikschen KathodenstrahlPolarographie. Z. Electrochem 1955, 59, 494–503. (31) Nicholson, R. S. Theory and Application of Cyclic Voltammetry for Measurement of Electrode Reaction Kinetics. Anal. Chem. 1965, 37, 1351–1355. (32) Myland, J. C.; Oldham, K. B. An analytical expression for the current-voltage relationship during reversible cyclic voltammetry. J. Electroanal. Chem. 1983, 153, 43–54. (33) Dou, W.; Subotnik, J. E. Electronic friction near metal surfaces: A case where moleculemetal couplings depend on nuclear coordinates. J. Chem. Phys. 2017, 146, 092304. (34) Dou, W.; Nitzan, A.; Subotnik, J. E. Surface hopping with a manifold of electronic states. II. Application to the many-body Anderson-Holstein model. J. Chem. Phys. 2015, 142, 084110. (35) Marcus, R. A. On the Theory of Electron-Transfer Reactions. VI. Unified Treatment for Homogeneous and Electrode Reactions. J. Chem. Phys. 1965, 43, 679–701.

42

ACS Paragon Plus Environment

Page 42 of 45

Page 43 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(36) Dou, W.; Subotnik, J. E. A broadened classical master equation approach for nonadiabatic dynamics at metal surfaces: Beyond the weak molecule-metal coupling limit. J. Chem. Phys. 2016, 144, 024116. (37) Compton, R. G.; Laborda, E.; Ward, K. R. Understanding Voltammetry; Imperial College Press, 2014. (38) Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P. Numerical Recipes in C : The Art of Scientific Computing; Cambridge University Press, 1992. (39) Hull, T. E.; Enright, W. H.; Fellen, B. M.; Sedgwick, A. E. Comparing Numerical Methods for Ordinary Differential Equations. SIAM J. Numer. Anal. 1972, 9, 603– 637. (40) Brown, P. N.; Byrne, G. D.; Hindmarsh, A. C. VODE: A Variable-Coefficient ODE Solver. SIAM J. Sci. Stat. Comput. 1989, 10, 1038–1051. (41) Jung, Y. J.; Cao, J. Spectral analysis of electron transfer kinetics. II. J. Chem. Phys. 2002, 117, 3822–3836. (42) Risken, H. The Fokker-Planck Equations, Method of Solution and Applications, 2nd ed.; Springer-Verlag, 1989; p 103. (43) Rudolph, M. Digital simulations on unequally spaced grids . Part 1 . Critical remarks on using the point method by discretisation on a transformed grid. J. Electroanal. Chem. 2002, 529, 97–108. (44) Rudolph, M. Digital simulations on unequally spaced grids. Part 2. Using the box method by discretisation on a transformed equally spaced grid. J. Electroanal. Chem. 2003, 543, 23–39.

43

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(45) Rudolph, M. Digital simulations on unequally spaced grids. Part 3. Attaining exponential convergence for the discretisation error of the flux as a new strategy in digital simulations of electrochemical experiments. J. Electroanal. Chem. 2004, 571, 289–307. (46) Rudolph, M. Attaining exponential convergence for the flux error with second and fourthorder accurate finitedifference equations. I. Presentation of the basic concept and application to a pure diffusion system. J. Comput. Chem. 2005, 26, 619–632. (47) Rudolph, M. Attaining exponential convergence for the flux error with second and fourthorder accurate finitedifference equations. II. Application to systems comprising firstorder chemical reactions. J. Comput. Chem. 2005, 26, 633–641. (48) Rudolph, M. Attaining exponential convergence for the flux error with second- and fourth-order accurate finite-difference equations. Part 3. Application to electrochemical systems comprising second-order chemical reactions. J. Comput. Chem. 2005, 26, 1193– 1204. (49) Rudolph, M.; Reddy, D. P.; Feldberg, S. W. A Simulator for Cyclic Voltammetric Respooses. Anal. Chem. 1994, 66, 589–600.

44

ACS Paragon Plus Environment

Page 44 of 45

Journal of of Physical Chemi Page 45 45 3

= 10

2D, Eq. 19

6

a. u.

3

= 10

5

a. u.

1D, Eq. 12 1D, Eq. 27

I(t)

2

2

(a)

(b)

1 2 au au 3 4 5 6 S Paragon Plus Environme Gt Gt 7 8 1

1

0

0

0.5

0.0

3

I(t)

2

0.5

= 10

4

.

0.5

.

2

(c)

1

0.0

3

0.5

= 10

3

.

(d)

1

0

0

0.5

0.0

( )

0.5

0.5

0.0

( )

0.5

.