QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed...

39
Quantifying Confidence in DFT Predicted Surface Pourbaix Diagrams of Transition Metal Electrode-Electrolyte Interfaces Olga Vinogradova a , Dilip Krishnamurthy b , Vikram Pande b , Venkatasubramanian Viswanathan a,b,* a Department of Chemical Engineering, Carnegie Mellon University, Pittsburgh, Pennsylva- nia, 15213, USA b Department of Mechanical Engineering, Carnegie Mellon University, Pittsburgh, Pennsyl- vania, 15213, USA * Corresponding author, Email: [email protected] Abstract Density Functional Theory (DFT) calculations have been widely used to predict the activity of catalysts based on the free energies of reaction intermediates. The incorpora- tion of the state of the catalyst surface under the electrochemical operating conditions while constructing the free energy diagram is crucial, without which even trends in activity predictions could be imprecisely captured. Surface Pourbaix diagrams indicate the surface state as a function of the pH and the potential. In this work, we utilize error-estimation capabilities within the BEEF-vdW exchange correlation functional as an ensemble approach to propagate the uncertainty associated with the adsorption energetics in the construction of Pourbaix diagrams. Within this approach, surface- transition phase boundaries are no longer sharp and are therefore associated with a fi- nite width. We determine the surface phase diagram for several transition metals under reaction conditions and electrode potentials relevant for the Oxygen Reduction Reaction (ORR). We observe that our surface phase predictions for most predominant species are in good agreement with cyclic voltammetry experiments and prior DFT studies. We use the OH * intermediate for comparing adsorption characteristics on Pt(111), Pt(100), Pd(111), Ir(111), Rh(111), and Ru(0001) since it has been shown to have a higher pre- diction efficiency relative to O * , and find the trend Ru>Rh>Ir>Pt>Pd for (111) metal facets, where Ru binds OH * the strongest. We robustly predict the likely surface phase as a function of reaction conditions by associating c-values to quantifying the confidence in predictions within the Pourbaix diagram. We define a confidence quantifying metric using which certain experimentally observed surface phases and peak assignments can be better rationalized. The probabilistic approach enables a more accurate determina- tion of the surface structure, and can readily be incorporated in computational studies for better understanding the catalyst surface under operating conditions. 1 arXiv:1710.08407v2 [cond-mat.mtrl-sci] 5 Jul 2018

Transcript of QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed...

Page 1: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Quantifying Confidence in DFT Predicted SurfacePourbaix Diagrams of Transition Metal

Electrode-Electrolyte InterfacesOlga Vinogradovaa, Dilip Krishnamurthyb, Vikram Pandeb,

Venkatasubramanian Viswanathana,b,∗

a Department of Chemical Engineering, Carnegie Mellon University, Pittsburgh, Pennsylva-nia, 15213, USAb Department of Mechanical Engineering, Carnegie Mellon University, Pittsburgh, Pennsyl-vania, 15213, USA∗ Corresponding author, Email: [email protected]

Abstract

Density Functional Theory (DFT) calculations have been widely used to predict theactivity of catalysts based on the free energies of reaction intermediates. The incorpora-tion of the state of the catalyst surface under the electrochemical operating conditionswhile constructing the free energy diagram is crucial, without which even trends inactivity predictions could be imprecisely captured. Surface Pourbaix diagrams indicatethe surface state as a function of the pH and the potential. In this work, we utilizeerror-estimation capabilities within the BEEF-vdW exchange correlation functional asan ensemble approach to propagate the uncertainty associated with the adsorptionenergetics in the construction of Pourbaix diagrams. Within this approach, surface-transition phase boundaries are no longer sharp and are therefore associated with a fi-nite width. We determine the surface phase diagram for several transition metals underreaction conditions and electrode potentials relevant for the Oxygen Reduction Reaction(ORR). We observe that our surface phase predictions for most predominant species arein good agreement with cyclic voltammetry experiments and prior DFT studies. Weuse the OH∗ intermediate for comparing adsorption characteristics on Pt(111), Pt(100),Pd(111), Ir(111), Rh(111), and Ru(0001) since it has been shown to have a higher pre-diction efficiency relative to O∗, and find the trend Ru>Rh>Ir>Pt>Pd for (111) metalfacets, where Ru binds OH∗ the strongest. We robustly predict the likely surface phaseas a function of reaction conditions by associating c-values to quantifying the confidencein predictions within the Pourbaix diagram. We define a confidence quantifying metricusing which certain experimentally observed surface phases and peak assignments canbe better rationalized. The probabilistic approach enables a more accurate determina-tion of the surface structure, and can readily be incorporated in computational studiesfor better understanding the catalyst surface under operating conditions.

1

arX

iv:1

710.

0840

7v2

[co

nd-m

at.m

trl-

sci]

5 J

ul 2

018

Page 2: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

1 Introduction

Electrochemical processes are at the heart of many routes towards sustainable energy storagein the form of chemical bonds.1 The storage routes involve hydrogen and oxygen electrochem-istry occurring at solid-liquid interfaces.2 Determining the activity of electrode materialsactive for these processes requires a thorough understanding of the surface dynamics at theelectrode-electrolyte interfaces.3

Probing the electrode-electrolyte interface directly is an extremely challenging problem.4There has been substantial development towards the coupling of X-ray spectroscopy directlywith electrochemical cells.5,6 This can be complemented by electrochemical impedance spec-troscopy (EIS) to deconvolute the signals and assign them to surface adsorbed species.7Theoretically, density functional theory calculations have been used to construct surfacephase diagrams based on probing many different possible surface configurations.8 However,a major challenge associated with this approach is that the choice of the exchange-correlationfunctional often plays a key role in determining the dominant surface phases.9,10

Recently, the use of Bayesian error estimation techniques has brought capabilities to esti-mate uncertainty associated with DFT simulations.11 In this work, we develop a systematicmethod to assign a confidence value (c-value) associated with the stable surface phases at agiven electrode potential and pH. Using this method, we probe the surface Pourbaix diagramof several transition metals: Pt, Pd, Ir, Rh, and Ru. We then map c-values onto the surfacePourbaix determined with the best-fit functional to gain insight into surface phase transi-tions and compare them with those determined by cyclic voltammetry experiments. Basedon our method, we identify that on Pt(111) at an electrode potential of 0.65 V (vs. RHE) thesurface could be covered with a combination of OH∗ or O∗ with relatively equal confidence.Our method reproduces the well-known uncertainty associated with the DFT-predicted OH∗to O∗ phase transition on Pt(111).12 On Pt(100), we note that the predicted stable phases of13OH∗ and 1

2OH∗ are consistent with the suggested structural transitions by Han et al.13 Our

results indicate that on Pd(111) at potentials starting at ∼0.75 V (vs. RHE), the surfaceis O∗ covered, consistent with the experimentally determined current peak corresponding tooxidation.14 On Ir(111), Rh(111), and Ru(0001) we find voltage regimes where our predictionconfidence is limited relative to other surfaces explored in this work. The order of relativeadsorption strengths of 1

3OH∗ on (111) facet metals is Ru>Rh>Ir>Pt>Pd (in decreasing

adsorption strength where Ru is most oxophilic). Except for the order between Pt and Pd,this ordering is comparable to the trend established in a previous computational study byKarlberg et al.15 However, we observe that the adsorption strengths of 1

3OH∗ on (111) facets

of Pt and Pd are very close to each other (within 0.02 eV) demonstrating a need to considerprediction uncertainty. We show that we can use c-value calculation to consider a wide arrayof functionals within DFT to assign an error to each adsorption energy. This is particu-larly helpful in distinguishing between surface phases close to each other in energy. Widerdistributions of c-values correspond to greater uncertainties and vice versa. Comparing thewidths of c-value energy distributions with cyclic voltammetry studies, we also consider asituation where the Rh(111) surface could reconstruct into metal-oxides to explain certainfeatures in cyclic voltammograms. Given the importance of the stable state of the surface indetermining electrochemical reaction mechanisms, we believe this method will be extremely

2

Page 3: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

crucial in constructing free energy diagrams on the appropriate surface and will play animportant role in determining the subsequent limiting potential.16,17

2 Methods

2.1 Construction of Surface Pourbaix Diagrams

The surface Pourbaix diagram represents the stable state of the catalyst surface in electro-chemical systems as a function of pH and the electrode potential U . Consider an arbitrarymetal surface S with adsorption site ∗ in its pristine state. A generalized representation ofthe adsorption of oxygenated intermediates, denoted as OmH∗n, can be written as

S-OmH∗n + (2m− n)(e− + H+) ⇀↽ S∗ + mH2O (1)

where m and n are the number of oxygen and hydrogen atoms in the adsorbate respec-tively. Note that this representation corresponds to a concerted proton-coupled electrontransfer (PCET) reaction. The associated free energy change can be computed as:

∆G(U, pH) = GS∗ + mGH2O −GS-OmH∗n− (2m− n)(Ge− + GH+) (2)

Rewriting Eqn. 2, the free energy of electrons Ge− is defined as −eU where e is theelectron charge. The free energy of protons GH+ is determined from the computationalhydrogen electrode using the equilibrium relation H++e− ⇀↽ H2. This gives at non-standardconditions GH++Ge− = 1

2GH2−eUSHE+kBT (ln[aH+ ]) where USHE is the potential relative to

the standard hydrogen electrode (SHE). We then write the free energy change for adsorptionof OmH∗n intermediates as

∆G(U, pH) = GS∗ + mGH2O −GS-OmH∗n− (2m− n)(

1

2GH2 − USHE − 2.303kBTpH) (3)

Therefore, we can derive a relation between potential and pH for a wide variety of adsorbateson a surface S in reference to standard conditions when ∆G(U, pH) = 0.

Similarly, generalized pH independent electrochemical reactions such as the dissolutionof metal surfaces can be represented as:

Si+ ⇀↽ Sj+ + (j− i)e− (4)

where i and j represent the charge on the ion, which can be zero for metal surfacesand uncharged species. The free energy would therefore be pH-independent and appear as ahorizontal lines on the surface Pourbaix and the standard reduction potentials are taken fromthe CRC handbook.18 We note that unconcerted electrochemical reactions can be representedas linear combinations of the two generalized forms presented above.

Since the most stable state at a given condition is that which minimizes free energy, wedetermine the surface Pourbaix diagram by choosing the appropriate state at each set ofconditions of U and pH. The phase transitions between adsorbates appear as lines with a

3

Page 4: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

slope of -59mV/pH at standard conditions. Free energies are determined from calculatedenergies of adsorption ES∗ corrected for entropy, S, and zero point energies, ZPE, which weassume do not change significantly with temperature, by ∆G = ∆Eref,water − T∆S + ∆ZPE.Values for entropy and zero point energy correction are given in the Supporting Informationsection 5. In this work, we assume constant zero point energies across the ensemble offunctionals for simplicity, since it is expected to have negligible variation (of order kT , wherek is the Boltzmann constant) relative to the order of magnitude (10− 20 kT ) of the surfaceenergetics.

2.2 Bayesian Error Estimation Framework

With the recent development of the Bayesian Ensemble Error Functional with van der Waalscorrelations (BEEF-vdW),11 there exists a systemat approach to estimate the uncertaintyin energetics of a DFT calculation. BEEF-vdW is a semi-empirical exchange correlation(XC) functional including non-local contributions, developed by using training data setsfor molecular formation energies, molecular reaction energies, molecular reaction barriers,non-covalent interactions, solid state properties, and chemisorption on solid surfaces.

The exchange correlation energy in BEEF-vdW is expressed as a sum of the GGA ex-change energy expanded using Legendre polynomials, the LDA and PBE19 correlation ener-gies and the non local correlation energy from vdW-DF2.20

Exc =∑m

amEGGA−xm + αcE

LDA−c + (1− αc)EPBE−c + Enl−c (5)

The parameters am and αc are optimized with respect to the above mentioned data sets.The error estimation functionality is enabled by deriving an ensemble of energies from anensemble of exchange correlation functionals non-self-consistently following a self-consistentDFT calculation. The ensemble of exchange correlation functionals is generated using aprobability distribution function for the parameters am and αc such that the standard devia-tion of the ensemble of energies reproduces the standard deviation for the training propertiescalculated using BEEF-vdW self-consistently.

2.3 Prediction Confidence of Surface Pourbaix Diagrams

Pourbaix diagrams (figure 1) typically depict only the most stable state of the surface (min-imum Gibbs free energy) over a range of operating potentials (U) and pH values with sharpphase boundaries. We utilize error estimation capabilities within the BEEF-vdW exchangecorrelation functional to determine the ensemble of free energies for all considered states as afunction of U and pH. Therefore, for every functional there exists a unique set of stable sur-face states in the ranges of pH and U of interest, determined by the corresponding minimumfree energy surface state. This results in an ensemble of Pourbaix diagrams, allowing us toapply statistical tools to obtain a measure of the confidence in a predicted surface state byquantifying the agreement between functionals. We are interested in determining the levelof agreement between functionals at the GGA-level as to the most energetically favorablesurface state as a function of potential and pH. We use the confidence-value21,22 (c-value),

4

Page 5: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

which in this context is defined as the fraction of the ensemble that is in agreement with thehypothesis of the optimal BEEF-vdW (best-fit) functional, and is given by

c(U, pH) =1

Nens

Nens∑n=1

∏si 6=sopt

Θ(∆Gnsi

(U, pH)−∆Gnsopt(U, pH)) (6)

where, Nens is the number of functionals in the ensemble, which is 2000 in this work. Θ(x)denotes the Heaviside step function. At any given U and pH, si ∈ S, the set of all consideredsurface states, and sopt is the thermodynamically stable surface state predicted by the BEEF-vdW optimal functional. ∆Gn

sirefers to the free energy of the ith surface state given by the

nth member of the ensemble of functionals. This approach allows assigning confidence valuesto calculated surface Pourbaix diagrams.

In regimes of electrode potential and pH where the confidence value is lower than 1,it is important to determine whether the stable surface phase predicted by the majorityof functionals agree with the hypothesis of the best-fit functional. We extend the conceptof c-value to determine a measure of confidence in any considered surface phase being thelowest in free energy, looking beyond just agreement with respect to the most stable phase inaccordance with the optimal functional hypothesis. In order to define this measure, we usesp(U, pH) as the function that maps electrochemical conditions to the corresponding moststable surface phase ("sp" for "stable phase") from the set of possible/considered phasesdenoted by {0,1,2,...,i,...,n}, where i denotes the ith surface phase and i = 0 indicates the cleansurface. For example, for the ith considered surface phase, we define below the respectivestability confidence measure csp=1(U, pH) quantitatively as the fraction of functionals thatpredict that the phase labelled as i = 1 is the lowest in free energy:

csp=i(U, pH) =1

Nens

Nens∑n=1

δ(spn(U, pH)− i). (7)

where δ(x) denotes the Dirac delta function and the superscript n denotes the predictionfrom the nth functional.

3 Results and Discussion

We use the example case of the surface phase diagrams relevant for oxygen reduction reaction(ORR) on Pt-group transition metals: Pt, Pd, Ir, Rh, and Ru, to demonstrate an approachto quantify the ability of DFT in determining the state surface state. We present our findingsand compare them to experimental linear sweep voltammetry measurements, and discuss thedegree of confidence in our DFT predictions in relation to the assignment of voltammetricpeaks.

5

Page 6: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure 1: For Pt(111) the surface Pourbaix diagram shows regions of stability for surfacecoverages of OH∗ and O∗ at conditions of pH and potential relative to the standard hydrogenelectrode (SHE). Dissolution of Pt is shown at 1.18 V and is independent of pH.

Figure 2: Comparing free energies of all coverage fraction surfaces sampled for Pt(111) atpH=0 we observe from the best-fit solutions (thick lines) that 1/4 O∗ is <0.05 eV more stablethan 1/3 OH∗ at 0.63 V. The thin lines represent solutions for free energy for a sample setof 50 BEEF-vdW ensemble functionals depicting a spread of energies for each surface statedetermined using the best-fit functional.

6

Page 7: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure 3: Figure (a) shows the prediction confidence of each stable surface of a Pourbaixdiagram for Pt(111). Greatest prediction uncertainty occurs at phase transition boundaries.In (b) we plot the prediction confidence in relation to potential at pH of 0. Surface stateswith the largest c-values correspond to the solution determined with the best-fit BEEF-vdWXC-functional in figure 1. We also note that this representation of state surfaces may beextended to higher pH values (refer to Supporting Information section 8.

3.1 Pt(111)

The Pourbaix diagram for Pt(111), previously investigated8,12, is calculated using the BEEF-vdW exchange correlation functional and is shown in figure 1. We consider the followingoxygenated adsorbate surface states of Pt(111): 1/3 OH∗, 1/4 O∗, 1/3 O∗, 1/2 O∗, 2/3 O∗,3/4 O∗, and 1 O∗. While modeling OH∗ with water stabilization effects, we ignore potentialdependent water orientation effects23, and we neglect water stabilization for O∗ states24 asboth of these effects have been shown to have a negligible influence on the energetics. Locallymixed surface phases are expected to be energetically unfavorable and are not consideredsince surface O∗ repels water unlike adsorbed OH∗, where a hexagonal water stabilizinglayer occurs due to favorable hydrogen bonding.9,24 In addition, mixed phases are complexto simulate12 since it involves the incorporation of long range disorder, which is undoubtedlypresent.25 We also neglect configurational entropy of adsorbed oxygen intermediates, sincethe effect is expected to be small.26

The state of the surface is governed by the following water discharge reactions where *refers to an adsorption site on a metal catalyst surface:

H2O + ∗⇀↽ OH∗ + e− + H+

H2O + ∗⇀↽ O∗ + 2e− + 2H+

We compute the free energy changes of these reactions, which are functions of the pH

7

Page 8: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

and the electrode potential, based on Eqn. 3 (refer section 2.1).

∆G = (GOH∗ +1

2GH2 −GH2O) + kBTln[aH+ ]− eU (8)

∆G =1

2(GO∗ + GH2 −GH2O) + kBTln[aH+ ]− eU (9)

The surface Pourbaix diagram for Pt(111) shown in figure 1 is constructed using Eqns.(8) and (9) . The free-energy as a function of the electrode potential (figure 2) determines thesurface phase boundaries from the best-fit BEEF-vdW XC functional and therefore identifiesthe most stable state of the surface at given reaction conditions of potential and pH. Atproton activity of aH+=1, we predict that below ≈0.63 V (vs. RHE), clean Pt(111) surfaceis the most stable. Water oxidation occurs at higher potentials starting with 1/4 monolayercoverage of O∗ and increasing monotonically to coverage 1/2 O∗ at ≈0.91 V (vs. RHE). Near1.18 V, we indicate Pt dissociation18 through Pt ⇀↽ Pt2+ + 2e−, which is pH independentand appears as a horizontal line on the surface Pourbaix diagram. Although there existsome differences in the positions of the phase boundaries, the observed set of stable surfacestates are consistent with those predicted by Hansen et al8 using the RPBE XC-functionalwith the exception of the 1/3 OH∗ surface state which Hansen et al. predicts at 0.78 V (vs.RHE). However, we observe that in the range 0.63–0.83 V (vs. RHE) the 1/4 O∗ surfaceis marginally (by < 0.05 eV) more stable than the 1/3 OH∗ surface state, an observationalso noted by Hansen et al. in the region 0.78–0.84 V. Close energetics can be resolved morerobustly through the quantification of confidence in the relative stability of surface phases bythe use of error estimation capabilities based on agreement between functionals at the GGAlevel. Using the BEEF-vdW best-fit functional, we compute the equilibrium potential of the1/3 OH∗ surface state to be 0.68 V (vs. RHE). Since the ensemble of functionals withinthe BEEF-vdW family of functionals give rise to an ensemble of energies, we quantify theuncertainty by the standard deviation of the distribution, computed to be σOH=0.23, whichbounds within the 68% confidence interval the calculation of 0.81 V by Rossmeisel et al12using the RPBE XC-functional. Other examples of calculations bound within one standarddeviation of the ensemble approach include the OH∗ adsorption energy computed by Tayloret al27 to be 0.59 V for a lower coverage of 1/9 OH∗ using the PW91 XC-functional, andthe adsorption potential for 1/3 OH∗ to be 0.71 V (PBE XC-functional) and 0.69 V (RPBEXC-functional) by Jinnouchi et al.10 Similarly, calculations using RPBE XC-functional byAnderson et al. measured adsorption of a variety of combinations of OH∗ and H2O fractions,observing that adsorbed OH∗ begins to form at ∼0.6 V.9

The ensemble of functionals within the BEEF-vdW XC functional results in an ensembleof free energies for each state (figure 2) which gives rise to an ensemble of Pourbaix dia-grams.21 Using c-value to quantify confidence in the predicted surface surface phases basedon the fraction of the ensemble that is in agreement with the hypothesis of the best-fit (oroptimal BEEF-vdW) functional, we construct a modified surface Pourbaix diagram followingthe methodology described in section 2.3. The confidence-derived surface Pourbaix diagramof Pt(111) with the associated c-values is shown in figure 3a. We note that regions spanningsurface phase transitions are those with the lowest c-values. To better understand the rela-tive confidence values, we show the c-values of each surface phase as a function of potential

8

Page 9: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

at fixed pH value of 0 in figure 3b. In this figure, we plot csp=i(pH,U) (refer to eqn. 7 ofsection 2.3) for all i corresponding to the various considered surface phases to capture thedegree of agreement between functionals as to the lowest free energy phase.

The phase transitions in figure 3b predicted by the max(csp=i(U)) curve at fixed pHvalues lie within 0.02 eV of those predicted by the optimal BEEF-vdW XC-functional. Weobserve that the distribution of confidence values for OH∗ is wider than that for the O∗phases, and attribute this to a larger standard deviation in the distribution of energies forwater-stabilized OH∗ (σOH=0.23 eV) compared to O∗ (σO ≈ 0.1 eV) in reference to a waterlayer. The distributions of the free energies of adsorption have similar widths, however, incalculating the equilibrium potential, O∗ is associated with the concerted transfer of twoprotons and electrons while the formation of OH∗ involves only one proton-coupled electrontransfer (Eqns. 8 and 9). The additional factor of 2 provides rationale for the tighteningof the equilibrium potential distribution for O∗ relative to OH∗. We discuss the relativeconfidence values of the various phases at pH=0, however at different pH values the effectcan be captured by a constant chemical potential shift of the proton activity (see SupportingInformation section 8). At URHE < 0.55 V, we predict the clean Pt(111) surface with thehighest c-value (> 0.5). Around ≈0.63 V (RHE), we find that the confidence values of theclean Pt(111), 1/3 OH∗ and 1/4 OH∗ are nearly equal (≈ 0.33) indicating that within DFTat the GGA-level these surface phases are equally likely to be the lowest in free energy. Theuncertainty in the OH∗ to O∗ phase transition in the range from ∼0.6 V to ∼0.8 V (RHE) isconsistent with that reported earlier.6 We note that in linear sweep voltammetry experimentsWang et al28, Koper29, and Garcia-Araez et al30,31 attribute the reversible butterfly regionbetween 0.6 V and 0.85 V to OH∗, which is also supported by Wakisaka et al5 in ex situ XPSexperiments. In line with these findings, although at U ≈0.63 V (RHE) predictions from ouranalysis (figure 3b) have very close energetics, we note that the highest prediction confidencefor OH∗ adsorption occurs at this potential, a key observation that is missed from the surfacePourbaix diagram in figure 1. This emphasizes the usefulness of the defined quantity csp=i(U)in more robustly assigning electrode potential regimes for the stability of a given surfacestate, when close energetics can be resolved between surface phases through experimentalvalidation or higher-order DFT. At URHE > 0.63 V (RHE), O∗ surface adsorption coveragegradually increases to higher oxygen coverage and higher c-values for 1/4 O∗ and 1/2 O∗until dissolution at a potential of 1.18 V. The 2/3 O ∗ surface phase is not stable at potentialsbelow the dissolution potential. Wakisaka et al.5 have reported that OH∗ formation beginsat E > 0.6 V continuing to increase until 0.8 V which is captured within uncertainty boundsof our predictions. Our findings of a monotonic increase in the oxygen coverage comparewell with blank voltammetry7 of Pt(111) in 0.1 M HClO4, avoiding the specific adsorptionof sulfate that occurs in sulfuric acid electrolyte. We believe that the correlation betweenour predictions and state-of-the-art experiments strengthens the effectiveness of confidencevalues and Bayesian error estimation tools in quantifying uncertainty of predicted stabilityof surface phases.

3.2 Pt(100)

We apply a similar methodology to Pt(100), which has been well studied both experimen-tally32–34 and computationally13,35 since commercial state-of-the-art Pt catalysts are com-

9

Page 10: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure 4: Figure (a) shows the surface Pourbaix diagram for Pt(100) where the onset of 1/3OH∗ is at 0.64 V (vs. RHE). A free-energy vs potential figure is presented in the SupportingInformation section 9.1. Figure (b) depicts the c-value for each surface state at pH=0. Weobserve a gradual transition from 1/3 OH∗ to 1/2 OH∗ from ∼0.6 V to ∼0.9 V (RHE). AtURHE > 0.9 V there is greatest likelihood of observing 3/4 O∗. The region on the x-axisshaded in gray represents the dissolution potential occurring at 1.18 V for Pt.

prised typically of low-index (111) and (100) facets36,37. We consider the following surfacestates of Pt(100): 1/3 OH∗, 1/2 OH∗, 1/4 O∗, 1/2 O∗, 3/4 O∗ and 1 O∗. The surface phaseboundaries in figure 4a are determined using the best-fit BEEF-vdW functional. On thesurface Pourbaix diagram we overlay the c-values as a function of pH and potential, whichprovide a measure of agreement between GGA-level functionals in the BEEF-vdW familyof functional. We find that at URHE < 0.64 V, the clean Pt(100) surface is thermodynami-cally most favorable. At higher potentials we predict 1/3 OH∗ until 0.87 V (RHE) at whichthere is a narrow potential window where the 1/2 OH∗ phase is the lowest in free energy.At URHE ≈> 0.9 V the OH∗ phase transitions to the 3/4 O∗ surface. We observe thatthe prediction confidence (c-value) of the OH∗ surface states is relatively low (≈<0.4). Thenarrow region for stable 1/2 OH∗ suggests that precise determination of the phase bound-aries for this state is computationally challenging with the best-fit BEEF-vdW functional.We show the degree of agreement within the ensemble of functionals as to the most stablesurface phase for each individual surfaces state in figure 4b. We see that in the range 0.6V<URHE<0.7 V 1/3 OH∗ has a marginally greater c-value than 1/2 OH∗. However, at ≈0.8V (RHE) the 1/2 OH∗ surface state has a c-value of < 0.25 suggesting that the 1/3 OH∗phase transitions to the more thermodynamically stable 1/2 OH∗ with increasing electrodepotential. Previously it was suggested through ab-initio DFT calculations and Monte Carlosimulations13 that 1/3 OH∗ transforms into 1/2 OH∗ which is supported by our analysis; wefind that 1/2 OH∗ has the greatest confidence between 0.75 V and 0.9 V (vs. RHE). Of alloxygen covered surfaces considered, at URHE > 0.9 V we predict the 3/4 O∗ adsorption stateto be the most favorable until the dissolution potential. Based on cyclic voltammetry inhydrochloric acid performed by several groups33,34 a wide reversible peak centered at 0.36 V(vs. RHE) has been reported, and attributed to a strong overlap between hydrogen and OH∗

10

Page 11: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

adsorption regions which makes precise determination of adsorbed states difficult. Gomezet al34 used a deconvolution analysis to decouple OH coverage from H coverage, and in asubsequent work a maximum coverage value of 0.37 OH species per Pt surface atom was re-ported33, which corresponds to OH∗ adsorption assigned to a shoulder of the reversible peakappearing between 0.5 V and 0.7 V on the cyclic voltammograms. We highlight that theprediction from our analysis for the clean Pt(100) surface transition to OH∗ (1/3 OH∗ at 0.65V (vs. RHE)) falls within this region although the relatively low (< 0.4) confidence valuesindicate the uncertainty associated with this transition. Similar to the case of Pt(111), thisagain emphasizes the efficacy of the defined quantity csp=i(U) in assigning potential regimesfor the stability of a given surface state, when close energetics can be resolved between sur-face phases through insights from both theory and experiments. We do not consider H∗adsorption since it is beyond the scope of the current work owing to the necessary coveragedependent analysis.38

3.3 Pd(111)

Figure 5: Figure (a) shows the surface Pourbaix diagram for Pd(111), where at pH=0oxidation is predicted to begin at 0.64 V (vs. RHE). Figure (b) shows the c-value for eachsurface state at pH=0, where transition to oxygen covered surfaces begins at U > 0.7 V. Theshaded gray region represents the onset of dissolution of crystalline Pd at 0.95 V.

Palladium has been identified to be a promising alternative to Pt for the ORR39–42.We construct the Pourbaix diagram for Pd(111) by considering the following states of thesurface43 similar to those considered for Pt(111): clean surface, 1/3 OH∗, 1/4 O∗, 1/3 O∗,1/2 O∗, 2/3 O∗, 3/4 O∗, and 1 O∗. We predict from the constructed surface Pourbaixdiagram, shown in figure 5a, that the surface is clear of adsorbates until a potential of0.64 V (vs. RHE) which marks the onset of 1/4 O∗ coverage surface state. At higherelectrode potentials, O∗ adsorption increases monotonically to 1/3 O∗ followed by 1/2 O∗,2/3 O∗, and finally by 3/4 O∗. Pd ⇀↽ Pd2+ + 2e− dissolution is labeled at 0.95 V which isindependent of pH. Relatively low (< 0.4) prediction confidence regions in figure 5a showthat transitions between 1/4 O∗ and 1/3 O∗ in particular have a high degree of uncertainty

11

Page 12: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

along with the 2/3 O∗ surface state transitions between 1/2 O∗ and 3/4 O∗ regions, whichwe inspect further using csp=i(U, pH = 0) (refer to section 2.3) for all i corresponding tothe considered surface phases. Based on this uncertainty quantification measure, for eachsurface phase we depict the agreement within the BEEF-vdW ensemble of functionals infigure 5b. We notice a low degree of prediction confidence as to the lowest free energysurface phase at ≈0.72 V (vs. RHE) (pH=0) where the maximum c-value is < 0.3. The 1/3OH∗ surface has a relatively higher confidence value than the 1/4 O∗ surface suggesting thatthe clean surface transitions directly to the 1/3 OH∗ surface phase, an observation enabledby the defined quantity csp=i(U, pH). Experimentally, Hara et al.14 ascribe broad anodicand cathodoic current peaks to OH∗ species around 0.7 V based on cyclic voltammetry inperchloric acid solution. We quantify the maximum observed c-value for 1/3 OH∗ between0.6-0.7 V (vs. RHE) which lies close to the experimentally observed region within theprediction uncertainty. We observe comparatively large prediction confidence values for 1/3O∗ and 1/2 O∗ covered surfaces between 0.8-1.3 V (vs. RHE), which is supported by Haraet al. who report an oxidation peak corresponding to oxygen adsorption at 1.02 V.

3.4 Ir(111)

Figure 6: Figure (a) shows the surface Pourbaix diagram for Ir(111), where we observeoxidation to begin with 1/3 OH∗ in the range from 0.2 V to 0.55 V. We then see increasein oxidation until the dissolution potential at 1.16 V. Figure (b) shows each stable regionassigned prediction confidence on the surface Pourbaix at pH=0 where we see a region withlow confidence between 0.55 V and 0.7 V. In this region the surface may covered by 1/3 OH∗,1/3 O∗, 1/2 O∗, or 2/3 O∗. The shaded gray region on the x-axis represents dissolution intoIr3+.

Iridium is a highly corrosion resistant platinum group metal and is similar to Pt al-though relatively less extensively studied44. Stable surface phases determined with the best-fit BEEF-vdW functional are shown with the bold black lines in the constructed surfacePourbaix diagram in figure 6a. Oxygen coverage begins with 1/3 OH∗ to potentials of ≈0.56V (vs. RHE) at pH=0, consistent with prior DFT predictions that Ir is more oxophilic than

12

Page 13: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Pt(111).15,45 The surface transitions to 1/3 O∗ at ≈0.56 V, followed by 1/2 O∗ at ≈0.64V, 2/3 O∗ at ≈0.74 V, and 3/4 O∗ at ≈1.1 V (all vs. RHE), with dissolution at ≈1.16V.18 However, prediction confidence values corresponding to the 1/3 O∗, 1/2 O∗, and 3/4O∗ surface phases are low (c-value < 0.4) relative to 1/3 OH∗ phase. To further probe thedegree of agreement within the ensemble of functionals as to the lowest free energy phaseespecially close to the phase boundaries, we assess the uncertainty quantification metric csp=i

for all i corresponding to the considered surface phases as shown in figure 6b. We predictwith high confidence the 1/3 OH∗ adsorption phase approximately in the 0.2 < U < 0.55 V(vs. RHE) range where the c-value is > 0.5 with a maximum at ≈0.4 V. We observe thatbetween ≈ 0.55 V and ≈ 0.75 V (vs. RHE) a smooth transition from the 1/3 O∗ phase tothe 1/2 O∗ phase. At higher potentials until dissolution, with a large degree of agreementbetween the functionals we predict the 2/3 O∗ surface phase to be the lowest in free energy.The 3/4 O∗ surface has a maximum c-value of nearly 0.4 at ≈1.0 V (vs. RHE), therefore,within DFT uncertainty we expect a short voltage window where this phase is stable justbefore dissolution. We note some uncertainty concerning the interpretation of peaks in cyclicvoltammograms while comparing the findings from our analysis. Cyclic voltammograms forIr(111) in perchloric acid shows a broad peak between approximately 0.05 V to 0.35 V (vs.RHE).46–48 In this range, Marc Koper suggests that the surface is covered with hydrogen butalso tentatively suggests that the peak could represent a mixed transition phase from beingOH-covered to being H-covered,29 while Wan et al. have attributed this region to hydrogenadsorption.48 We find that our prediction of a stronger binding OH∗ surface as comparedto Pt(111) is closer in agreement with previous DFT calculations.15 Similarly, regarding theinterpretation of a pair of sharp cyclic voltammetry peaks at 0.92 V and 0.95 V in perchloricacid, Wan et al.48 attributes it to initial stages of oxidation representing hydroxide adsorp-tion, but this would depend on the interpretation of the peak between 0.05 V and 0.35 Vsince that would affect the length of the double-layer region.29 With high confidence (c-value>0.7), we observe that the surface is O∗ covered at 0.9 V (vs. RHE) which is in agreementwith DFT calculations15. We highlight that the defined confidence values provide a way toincorporate DFT uncertainty to interpret smooth surface phase transitions and aid a moreintegrated theory-experiment effort that is necessary for conclusive peak assignments.

3.5 Rh(111)

The Pt-group metal rhodium is less frequently studied since it is scarce and expensive.However, it is a versatile catalyst most known for use in the three-way-catalyst (TWC) forNOx reduction, and oxidation of CO and unburned hydrocarbons in automotive exhaust.49We show the constructed surface Pourbaix diagram for the Rh(111) surface in figure 7a.At pH=0, we observe the 1/3 OH∗ surface phase to be stable between ≈0.2-0.6 V (vs.RHE). The surface Pourbaix diagram also shows that Rh(111) binds oxygen intermediatesstronger than Pt(111) which is supported by prior DFT studies.15,50 At pH=0, we showthat dissolution is favored at 0.6 V (vs. RHE) of the crystalline Rh metal governed bythe equilibrium relation Rh+ + e− ⇀↽ Rh. Based on the best-fit BEEF-vdW functional, at aconstant electrode potential with increasing pH, we observe that the 1/2 O∗ surface appearsas the most stable phase followed by the 3/4 O∗ phase and subsequently the O∗ surfacephase. However, we notice that the 1/2 O∗ phase is predicted to have a narrow stability

13

Page 14: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure 7: Figure (a) shows the surface Pourbaix diagram for Rh(111). Figure (b) describesthe confidence for each surface state at pH=0 where dissolution from solid Rh to Rh1+ isshown at 0.60 V by the shaded gray region.

region associated with a comparatively low (c-value < 0.4) degree of prediction confidence.We understand the relative agreement between the functionals as to the lowest free energy

phase by constructing figure 7b based on csp=i(U, pH) for all i corresponding to the surfacesphases considered. We observe that 1/4 O∗, 1/3 O∗, and 2/3 O∗ surface phases are associatedwith low confidence values (< 0.3) in accordance with the fact that these phases do noappear in figure 7a based on the best-fit BEEF-vdW functional. From figure 7b at pH=0 wepredict the 1

3OH∗ surface to be stable between ≈0.2 and ≈0.6 V (vs. RHE) with relatively

high prediction confidence. Several experimental groups have studied cyclic voltammetry onRh(111) in perchloric solutions to understand the sharp reversible peak visible at 0.64 V51–53.The reversible nature of this peak has led it to be tentatively ascribed to OH adsorption29,51

although we predict the onset of 1/3 OH∗ phase at a lower potential. However, we cautiouslyspeculate that a surface reconstruction to RhO2 occurs based on the our prediction of theonset of OH∗ adsorption on RhO2 at 0.66 V and the observed reversible butterfly peakat 0.64 V (refer to section 4 of the Supporting Information for additional details). Thisreconstruction could correspond to the broad peak at ≈0.3 V and also explain why thedissolution potential for Rh to Rh1+ is experimentally observed18 at 0.6 V while the cyclicvoltammogram extends until ≈ 0.8 V.29,51 However, further insight into the stability of theRh oxide through a surface Pourbaix diagram is necessary to validate the hypothesis.

3.6 Ru(0001)

Ruthenium is used as a catalyst for nitrogen reduction54,55 and is the basis for several bimetal-lic catalyst systems in electrocatalysts.56,57 We consider the widely studied Ru(0001) surfacefor our analysis. The surface Pourbaix diagram for the Hexagonal close packed Ru(0001)surface is shown in figure 8a with the surface transition from 1/3 OH∗ to 2/3 O∗ surfaceconfiguration at 0.24 V (vs. RHE) at a pH=0. At a constant pH, with increasing potential,we observe a transition from the 2/3 O∗ surface phase to the 1 O∗ phase. Ru metal binds

14

Page 15: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure 8: Figure (a) shows the surface Pourbaix diagram for Ru(0001) showing strong oxi-dation at low potentials. Figure (b) shows the confidence prediction for each surface state atpH of 0. The region on the x-axis shaded in gray shows the dissolution potential occurringat 0.46 V.

oxygen strongly with lower onset potentials for adsorption of oxygenated species relative toother transition metals, which is in agreement with a previous DFT study by Taylor et al27.We highlight from figure 8a that the stable surface states appearing in the Pourbaix diagramare predicted with a high level of confidence (≈>0.5). We observe only the 1/3 OH∗ to2/3 O∗ transition to have a comparatively low prediction confidence (< 0.37). We assessthe relative prediction confidence of all surface phases using csp=i by constructing figure 8bat pH=0. The maximum prediction confidence (c-value of 0.85) associated with the 1/3OH∗ phase occurs near 0.03 V (vs. RHE). The dissolution potential of 0.46 V representsdissociation of crystalline Ru to Ru2+ under standard conditions.18 Our prediction of theOH∗ to O∗ phase transition at 0.24 V (vs. RHE) is close to the that reported by M. Koper(0.18 V) on considering the DFT prediction error (σOH=0.27, σO=0.09). While earlier com-putational work suggests a narrow double layer region,27, we do not observe a double layerregion (> −0.2 V vs. RHE) with a high degree of agreement between functionals consis-tent with the interpretation of cyclic voltammograms58–60 by M. Koper, which highlights theusefulness of uncertainty incorporation. We rationalize the fact that cyclic voltammogramextends beyond 0.8 V while at standard conditions the dissolution of Ru to Ru2+ is markedat 0.46 V based on a possible metal surface reconstruction to RuO2, which is supportedby a better agreement between our predicted onset of OH∗ adsorption on RuO2 at 0.23 Vand the observed reversible peak at 0.18 V. However, it remains to be confirmed that theOH∗ surface phase is the most stable in this potential regime through a Pourbaix diagramconstruction on RuO2.

Further improvements to the uncertainty incorporation method presented here can bemade by accounting for configurational and vibrational entropy and accounting for uncer-tainties in zero point energy although we expect these effects to much smaller than theuncertainties accounted for in this paper. For increased prediction accuracy of surface re-activity of catalysts, systematically accounting for a self-consistent loop between the stable

15

Page 16: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

state of the surface determined by surface Pourbaix diagrams and the reactivity determinedby the limiting potential derived from free energy diagrams is essential. In addition, thiswork enables a precise determination of the stable catalyst surface state, which has importantimplications for identifying dominant reaction mechanisms and selectivity.

4 Conclusions

Surface Pourbaix diagrams are crucial in determining electrocatalytic reaction mechanismsand in corrosion processes. In this work, we demonstrate a method to quantify uncertaintyin density functional theory calculated surface Pourbaix diagrams. Based on this method,we define a confidence value (c-value) associated with each possible surface configuration ata given electrode potential, U, and pH. The approach involving the BEEF-vdW exchangecorrelation functional with built-in error estimation capabilities is much more computation-ally efficient relative to performing many calculations based on different functionals sincethis involves non-self-consistent calculations based on the converged density from one self-consistent calculation. Using this method, we have constructed surface Pourbaix diagramsfor each of Pt(111), Pt(100), Pd(111), Ir(111), Rh(111), and Ru(0001) surface and the as-sociated confidence values for these surface phases. On Pt(111), our method captures thewell-known uncertainty in the phase transition from OH∗ to O∗. We find good agreementbetween our predicted phase diagrams on Pt(100) and Pd(111) compared to cyclic voltam-metry experiments for the onset of OH∗ and O∗ covered surfaces. Ir(111), Rh(111), andRu(0001) all exhibit strong oxidation of 1/3 OH∗ which is supports prior theoretical stud-ies. We also briefly consider oxidation on rutile metal-oxides as a possible explanation toconnect DFT and experimental observations. We suggest that reconstruction of the surface,particularly for Rh(111) and Ru(0001), could be important. An important implication fromthis work is the need to incorporate multiple surface phases in determining electrochemicalreactions mechanisms and the associated activity due to finite uncertainty associated withpredicting the stable state of the surface.

Acknowledgement

O.V. gratefully acknowledges funding support in part from Volkswagen AG. D.K. and V.Vgratefully acknowledge funding support from the National Science Foundation under awardCBET-1554273.

References(1) Lewis, N. S.; Nocera, D. G. Powering the planet: Chemical challenges in solar energy

utilization. PNAS 2006, 103, 15729–15735.

(2) Montoya, J. H.; Seitz, L. C.; Chakthranont, P.; Vojvodic, A.; Jaramillo, T. F.;Nørskov, J. K. Materials for solar fuels and chemicals. Nat. Mater. 2017, 16, 70–81.

16

Page 17: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

(3) Bockris, J. O.; Khan, S. U. Surface electrochemistry: a molecular level approach;Springer Science & Business Media, 2013.

(4) Bard, A. J.; Abruna, H. D.; Chidsey, C. E.; Faulkner, L. R.; Feldberg, S. W.; Itaya, K.;Majda, M.; Melroy, O.; Murray, R. W. The electrode/electrolyte interface-a statusreport. J. Phys. Chem. 1993, 97, 7147–7173.

(5) Wakisaka, M.; Suzuki, H.; Mitsui, S.; Uchida, H.; Watanabe, M. Identification andquantification of oxygen species adsorbed on Pt (111) single-crystal and polycrystallinePt electrodes by photoelectron spectroscopy. Langmuir 2009, 25, 1897–1900.

(6) Casalongue, H. S.; Kaya, S.; Viswanathan, V.; Miller, D. J.; Friebel, D.; Hansen, H. A.;Nørskov, J. K.; Nilsson, A.; Ogasawara, H. Direct observation of the oxygenated speciesduring oxygen reduction on a platinum fuel cell cathode. Nat. Commun. 2013, 4, 2817.

(7) Bondarenko, A. S.; Stephens, I. E.; Hansen, H. A.; Pérez-Alonso, F. J.; Trip-kovic, V.; Johansson, T. P.; Rossmeisl, J.; Nørskov, J. K.; Chorkendorff, I. The Pt(111)/electrolyte interface under oxygen reduction reaction conditions: an electrochem-ical impedance spectroscopy study. Langmuir 2011, 27, 2058–2066.

(8) Hansen, H. A.; Rossmeisl, J.; Nørskov, J. K. Surface Pourbaix diagrams and oxygenreduction activity of Pt, Ag and Ni (111) surfaces studied by DFT. Phys. Chem. Chem.Phys. 2008, 10, 3722–3730.

(9) Tian, F.; Jinnouchi, R.; Anderson, A. B. How potentials of zero charge and potentialsfor water oxidation to OH (ads) on Pt (111) electrodes vary with coverage. J. Phys.Chem. C 2009, 113, 17484–17492.

(10) Jinnouchi, R.; Anderson, A. B. Aqueous and surface redox potentials from self-consistently determined Gibbs energies. J. Phys. Chem. C 2008, 112, 8747–8750.

(11) Wellendorff, J.; Lundgaard, K. T.; Møgelhøj, A.; Petzold, V.; Landis, D. D.;Nørskov, J. K.; Bligaard, T.; Jacobsen, K. W. Density functionals for surface science:Exchange-correlation model development with Bayesian error estimation. Phys. Rev. B2012, 85, 235149.

(12) Rossmeisl, J.; Nørskov, J. K.; Taylor, C. D.; Janik, M. J.; Neurock, M. Calculatedphase diagrams for the electrochemical oxidation and reduction of water over Pt (111).J. Phys. Chem. B 2006, 110, 21833–21839.

(13) Han, B.; Viswanathan, V.; Pitsch, H. First-principles based analysis of the electrocat-alytic activity of the unreconstructed pt (100) surface for oxygen reduction reaction. J.Phys. Chem. C 2012, 116, 6174–6183.

(14) Hara, M.; Linke, U.; Wandlowski, T. Preparation and electrochemical characterizationof palladium single crystal electrodes in 0.1 MH 2 SO 4 and HClO 4: Part I. Low-indexphases. Electrochim. Acta 2007, 52, 5733–5748.

17

Page 18: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

(15) Karlberg, G. Adsorption trends for water, hydroxyl, oxygen, and hydrogen ontransition-metal and platinum-skin surfaces. Phys. Rev. B 2006, 74, 153414.

(16) Deshpande, S.; Kitchin, J. R.; Viswanathan, V. Quantifying Uncertainty in ActivityVolcano Relationships for Oxygen Reduction Reaction. ACS Catal. 2016, 6, 5251–5259.

(17) Krishnamurthy, D.; Sumaria, V.; Viswanathan, V. Maximal predictability approachfor identifying the right descriptors for electrocatalytic reactions. J. Phys. Chem. Lett.2018, 9, 588–595.

(18) Lide, D. R.; Kehiaian, H. V. CRC handbook of thermophysical and thermochemical data;Crc Press, 1994; Vol. 1.

(19) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation madesimple. Phys. Rev. Lett. 1996, 77, 3865.

(20) Lee, K.; Murray, É. D.; Kong, L.; Lundqvist, B. I.; Langreth, D. C. Higher-accuracyvan der Waals density functional. Phys. Rev. B 2010, 82, 081101.

(21) Sumaria, V.; Krishnamurthy, D.; Viswanathan, V. Quantifying Confidence in DFTPredicted Surface Pourbaix Diagrams and Associated Reaction Pathways for ChlorineEvolution. arXiv:1804.02766 [cond-mat.mtrl-sci] 2018,

(22) Houchins, G.; Viswanathan, V. Quantifying confidence in density functional theorypredictions of magnetic ground states. Phys. Rev. B 2017, 96, 134426.

(23) Garcia-Araez, N.; Climent, V.; Feliu, J. Potential-dependent water orientation on Pt(111), Pt (100), and Pt (110), as inferred from laser-pulsed experiments. Electrostaticand chemical effects. J. Phys. Chem. C 2009, 113, 9290–9304.

(24) Ogasawara, H.; Brena, B.; Nordlund, D.; Nyberg, M.; Pelmenschikov, A.; Pettersson, L.;Nilsson, A. Structure and bonding of water on Pt (111). Phys. Rev. Lett. 2002, 89,276102.

(25) Viswanathan, V.; Hansen, H. A.; Rossmeisl, J.; Jaramillo, T. F.; Pitsch, H.;Nørskov, J. K. Simulating linear sweep voltammetry from first-principles: applicationto electrochemical oxidation of water on Pt (111) and Pt3Ni (111). J. Phys. Chem. C2012, 116, 4698–4704.

(26) Rossmeisl, J.; Karlberg, G. S.; Jaramillo, T.; Nørskov, J. K. Steady state oxygen re-duction and cyclic voltammetry. Faraday Discuss. 2009, 140, 337–346.

(27) Taylor, C. D.; Kelly, R. G.; Neurock, M. First-principles prediction of equilibriumpotentials for water activation by a series of metals. J. Electrochem. Soc. 2007, 154,F217–F221.

(28) Wang, J.; Markovic, N.; Adzic, R. Kinetic analysis of oxygen reduction on Pt (111)in acid solutions: intrinsic kinetic parameters and anion adsorption effects. J. Phys.Chem. B 2004, 108, 4127–4133.

18

Page 19: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

(29) Koper, M. Blank voltammetry of hexagonal surfaces of Pt-group metal electrodes: com-parison to density functional theory calculations and ultra-high vacuum experimentson water dissociation. Electrochim. Acta 2011, 56, 10645–10651.

(30) Garcia-Araez, N.; Climent, V.; Feliu, J. M. Determination of the entropy of formationof the Pt (111)|perchloric acid solution interface. Estimation of the entropy of adsorbedhydrogen and OH species. J. Solid State Electrochem. 2008, 12, 387–398.

(31) Garcia-Araez, N.; Climent, V.; Herrero, E.; Feliu, J. M.; Lipkowski, J. Thermodynamicapproach to the double layer capacity of a Pt (111) electrode in perchloric acid solutions.Electrochim. Acta 2006, 51, 3787–3793.

(32) Stamenkovic, V.; Markovic, N.; Ross Jr, P. Structure-relationships in electrocatalysis:oxygen reduction and hydrogen oxidation reactions on Pt (111) and Pt (100) in solutionscontaining chloride ions. J. Electroanal. Chem. 2001, 500, 44–51.

(33) Climent, V.; Gómez, R.; Orts, J. M.; Feliu, J. M. Thermodynamic analysis of thetemperature dependence of OH adsorption on Pt (111) and Pt (100) electrodes inacidic media in the absence of specific anion adsorption. J. Phys. Chem. B 2006, 110,11344–11351.

(34) Gómez, R.; Orts, J. M.; Álvarez-Ruiz, B.; Feliu, J. M. Effect of temperature on hydrogenadsorption on Pt (111), Pt (110), and Pt (100) electrodes in 0.1 M HClO4. J. Phys.Chem. B 2004, 108, 228–238.

(35) Duan, Z.; Wang, G. Comparison of reaction energetics for oxygen reduction reactionson Pt (100), Pt (111), Pt/Ni (100), and Pt/Ni (111) surfaces: a first-principles study.J. Phys. Chem. C 2013, 117, 6284–6292.

(36) Song, H.; Kim, F.; Connor, S.; Somorjai, G. A.; Yang, P. Pt nanocrystals: shape controland langmuir- blodgett monolayer formation. J. Phys. Chem. B 2005, 109, 188–193.

(37) Wang, C.; Daimon, H.; Onodera, T.; Koda, T.; Sun, S. A General Approach to theSize-and Shape-Controlled Synthesis of Platinum Nanoparticles and Their CatalyticReduction of Oxygen. Angew. Chem. Int. Ed. 2008, 47, 3588–3591.

(38) Karlberg, G.; Jaramillo, T.; Skulason, E.; Rossmeisl, J.; Bligaard, T.; Nørskov, J. K.Cyclic voltammograms for H on Pt (111) and Pt (100) from first principles. Phys. Rev.Lett. 2007, 99, 126101.

(39) Shao, M.-H.; Sasaki, K.; Adzic, R. R. Pd- Fe nanoparticles as electrocatalysts for oxygenreduction. J. Am. Chem. Soc. 2006, 128, 3526–3527.

(40) Fernández, J. L.; Raghuveer, V.; Manthiram, A.; Bard, A. J. Pd- Ti and Pd- Co- Auelectrocatalysts as a replacement for platinum for oxygen reduction in proton exchangemembrane fuel cells. J. Am. Chem. Soc. 2005, 127, 13100–13101.

19

Page 20: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

(41) Fernández, J. L.; Walsh, D. A.; Bard, A. J. Thermodynamic guidelines for the designof bimetallic catalysts for oxygen electroreduction and rapid screening by scanningelectrochemical microscopy. M- Co (M: Pd, Ag, Au). J. Am. Chem. Soc. 2005, 127,357–365.

(42) Savadogo, O.; Lee, K.; Oishi, K.; Mitsushima, S.; Kamiya, N.; Ota, K.-I. New palla-dium alloys catalyst for the oxygen reduction reaction in an acid medium. Electrochem.Commun. 2004, 6, 105–109.

(43) Mitsui, T.; Rose, M.; Fomin, E.; Ogletree, D. F.; Salmeron, M. Water diffusion andclustering on Pd (111). Science 2002, 297, 1850–1852.

(44) Cherevko, S.; Geiger, S.; Kasian, O.; Kulyk, N.; Grote, J.-P.; Savan, A.; Shrestha, B. R.;Merzlikin, S.; Breitbach, B.; Ludwig, A. Oxygen and hydrogen evolution reactions onRu, RuO 2, Ir, and IrO 2 thin film electrodes in acidic and alkaline electrolytes: acomparative study on activity and stability. Catal. Today 2016, 262, 170–180.

(45) Krekelberg, W. P.; Greeley, J.; Mavrikakis, M. Atomic and molecular adsorption on Ir(111). J. Phys. Chem. B 2004, 108, 987–994.

(46) Ganassin, A.; Sebastián, P.; Climent, V.; Schuhmann, W.; Bandarenka, A. S.; Feliu, J.On the pH Dependence of the Potential of Maximum Entropy of Ir (111) Electrodes.Sci. Rep. 2017, 7 .

(47) Pajkossy, T.; Kibler, L.; Kolb, D. Voltammetry and impedance measurements of Ir(111) electrodes in aqueous solutions. J. Electroanal. Chem. 2005, 582, 69–75.

(48) Wan, L.-J.; Hara, M.; Inukai, J.; Itaya, K.; et al, In situ scanning tunneling microscopyof well-defined Ir (111) surface: high-resolution imaging of adsorbed sulfate. J. Phys.Chem. B 1999, 103, 6978–6983.

(49) Heck, R. M.; Farrauto, R. J. Automobile exhaust catalysts. Appl. Catal., A 2001, 221,443–457.

(50) Marković, N.; Ross, P. N. Surface science studies of model fuel cell electrocatalysts.Surf. Sci. Rep. 2002, 45, 117–229.

(51) Housmans, T.; Koper, M. CO oxidation on stepped Rh [n (1 1 1)×(1 1 1)] singlecrystal electrodes: Anion effects on CO surface mobility. Electrochem. Commun. 2005,7, 581–588.

(52) Clavilier, J.; Wasberg, M.; Petit, M.; Klein, L. Detailed analysis of the voltammetry ofRh (111) in perchloric acid solution. J. Electroanal. Chem. 1994, 374, 123–131.

(53) Wan, L.-J.; Yau, S.-L.; Swain, G. M.; Itaya, K. In-situ scanning tunneling microscopyof well-ordered Rh (111) electrodes. J. Electroanal. Chem. 1995, 381, 105–111.

(54) Jacobsen, C. J.; Dahl, S.; Hansen, P. L.; Törnqvist, E.; Jensen, L.; Topsøe, H.;Prip, D. V.; Møenshaug, P. B.; Chorkendorff, I. Structure sensitivity of supportedruthenium catalysts for ammonia synthesis. J. Mol. Catal. A: Chem. 2000, 163, 19–26.

20

Page 21: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

(55) Aika, K.-i.; Hori, H.; Ozaki, A. Activation of nitrogen by alkali metal promoted tran-sition metal I. Ammonia synthesis over ruthenium promoted by alkali metal. J. Catal.1972, 27, 424–431.

(56) Hansen, T. W.; Hansen, P. L.; Dahl, S.; Jacobsen, C. J. Support effect and active siteson promoted ruthenium catalysts for ammonia synthesis. Catal. Lett. 2002, 84, 7–12.

(57) Jackson, A.; Viswanathan, V.; Forman, A. J.; Larsen, A. H.; Nørskov, J. K.;Jaramillo, T. F. Climbing the activity volcano: core–shell Ru at Pt electrocatalystsfor oxygen reduction. ChemElectroChem 2014, 1, 67–71.

(58) Zhou, W.-P.; Lewera, A.; Bagus, P. S.; Wieckowski, A. Electrochemical and electronicproperties of platinum deposits on Ru (0001): Combined XPS and cyclic voltammetricstudy. J. Phys. Chem. C 2007, 111, 13490–13496.

(59) Hoster, H.; Richter, B.; Behm, R. Catalytic influence of Pt monolayer islands on the hy-drogen electrochemistry of Ru (0001) studied by ultrahigh vacuum scanning tunnelingmicroscopy and cyclic voltammetry. J. Phys. Chem. B 2004, 108, 14780–14788.

(60) El-Aziz, A.; Kibler, L. New information about the electrochemical behaviour of Ru (00 0 1) in perchloric acid solutions. Electrochem. Commun. 2002, 4, 866–870.

21

Page 22: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Supporting Information:

Quantifying Confidence in DFT Predicted

Surface Pourbaix Diagrams of Transition

Metal Electrode-Electrolyte Interfaces

Olga Vinogradova,† Dilip Krishnamurthy,‡ Vikram Pande,‡ and

Venkatasubramanian Viswanathan∗,‡,†

†Department of Chemical Engineering, Carnegie Mellon University, Pittsburgh, PA 15213,

USA

‡Department of Mechanical Engineering, Carnegie Mellon University, Pittsburgh, PA

15213, USA

1 Calculation Details

All calculations were performed using the projector augmented-wave (PAW)1,2 method as

implemented in GPAW3,4 with the ASE interface5. The Bayesian Error Estimation Func-

tional with Van der Waals corrections (BEEF-vdW)6 exchange-correlation functional was

employed, which has built-in error estimation capabilities and is calibrated to reproduce the

error as mapped to experimental training data. This error estimation capability has been

used to quantify uncertainty in heterogenous catalysis7, electrocatalytic actiivty8,9, magnetic

ground states10 and mechanical properties of solid electrolytes11.

For (111) facet surfaces, OH* adsorption is modeled on a (√

3 ×√

3)R30◦ unit cell and

1

Page 23: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

O* is modeled on a (2×2) unit cell each consisting of 4 layers of metal atoms. The k-point

grids are 4×4×1 and 6×6×1 respectively. For (100) facet, 4 layer unit cells of (2×2) and

(2×3) were used with 4×4×1 and 6×4×1 k-point grids. A real space grid discretization of

0.18Å was used for all calculations. All DFT calculations were converged with respect to

k-point and grid spacing. Each unit cell was periodic in the x and y directions only with

5Å layer of vacuum in z direction above and below the layers. The bottom two layers were

fixed and the top two layers including any adsorbates were relaxed to a net force less than

0.05eV/Å. A Fermi smearing of 0.05 eV was used to ensure self consistency in calculation of

electron occupations.

2 Lattice Parameters

The lattice sizes for each metal were extracted from bulk optimized crystal structures using

the BEEF-vdW exchange-correlation functional. The a lattice parameters used in all subse-

quent simulations are 4.022 Å for Pt, 4.015 Å for Pd, 3.897 Å for Ir Å 3.890 Å for Rh, and

2.738 Å and 4.332 Å for a and c respectively for Ru.

3 Calculating Adsorption Energies

The adsorption energies of these intermediates calculated using the BEEF-vdW best-fit

functional are reported in Table S1. The standard deviations of all adsorbed surface states

are also reported as σO∗ or σOH∗ as the fraction of BEEF-vdW ensemble functionals in

agreement with the best-fit functional.

2

Page 24: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Table S1: Adsorption Energies of the various intermediates considered on the surface oftransition metal surfaces. All energies are reported in eV. All O* intermediates adsorbed atthe hollow site unless italicized, which signal intermediates that relaxed to the bridge site.All errors are reported as the standard deviation of the BEEF ensemble.

Pt(111) Pt(100) Pd(111) Ir(111) Rh(111) Ru(0001)1/3 OH H2O 0.68 ± 0.23 0.65 ± 0.32 0.69 ± 0.32 0.21 ± 0.23 0.20 ± 0.30 -0.30 ± 0.271/2 OH H2O - 0.73 ± 0.30 - - - -1/4 O 0.64 ± 0.11 0.92 ± 0.13 0.65 ± 0.14 0.40 ± 0.10 0.34 ± 0.13 0.07 ± 0.081/3 O 0.92 ± 0.09 1.22 ± 0.12 0.67 ± 0.13 0.39 ± 0.05 0.40 ± 0.11 -0.02± 0.081/2 O 0.78 ± 0.12 0.79 ± 0.17 0.79 ± 0.14 0.47 ± 0.05 0.47 ± 0.13 0.11 ± 0.112/3 O 0.88 ± 0.11 1.27 ± 0.13 0.91 ± 0.13 0.54 ± 0.06 0.55 ± 0.11 0.11 ± 0.093/4 O 0.97 ± 0.11 0.87 ± 0.13 0.97 ± 0.13 0.60 ± 0.07 0.58 ± 0.13 0.18 ± 0.101 O 1.13 ± 0.11 1.35 ± 0.10 1.17 ± 0.12 0.75 ± 0.07 0.70 ± 0.12 0.25 ± 0.10

4 Adsorption on Bulk Oxide Slabs

The OH* intermediate was absorbed at the cus sites of 4 layer rutile (110) metal-oxide slabs

where the bottom two layers are constrained. The following figure shows the structure for

any metal-oxides considered.

Figure S1: Structure of clean and OH* adsorbed surface for RhO2.

The adsorption energy is calculated by the following equation consistent with the gener-

alized form in the main paper:

∆EOH∗ = EOH∗ − E∗ − EH2O + 1/2EH2 (1)

3

Page 25: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

5 Zero Point and Entropy Corrections

We report the ∆ZPE and ∆S◦ for all reactions at standard temperature T=298 K in Table

S2.12,13

Table S2: Zero point and entropic corrections to the free energies

TS T∆S ZPE ∆ZPE ∆ZPE − T∆SH2O 0.67 0 0.56 0 0OH∗ + 1

2H2 0.20 -0.47 0.44 -0.12 0.35

O ∗+H2 0.41 -0.27 0.34 -0.22 0.0512O2 + H2 0.73 0.05 0.32 -0.24 -0.29

H2 0.41 0.2712O2 0.32 0.05

O∗ 0 0.07OH∗ 0 0.30H∗ 0 0.17

4

Page 26: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

6 Quantifying Confidence in Adsorption Energies

Using the BEEF-vdW best-fit functional, the ensemble of adsorption energies for each surface

phase were calculated. The energies were referenced to a clean Pt(111) surfaces since the

error is much smaller due to similarities in metal-adsorbate bonding characteristics. We

calculate the error of each adsorption energy as the standard deviation of the ensemble

spread of energies. The errors are tabulated in table S1. Smaller errors imply similarity in

surface-adsorbate interactions of metal surfaces.

Figure S2: Normalized frequency as a function of adsorption energy of 1/3 OH∗ on Pt(111)metal surface. The standard deviation of the spread of energies is σOH=0.23 eV.

5

Page 27: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

7 Free Energy Ensemble for Pt(111) Surface

To illustrate the ensemble of energies derived from the BEEF-vdW ensemble of functionals,

in figure 2 in the main paper we reported a consistent set of 50 energies bounding each of

the energies calculated using the best-fit functional. The following figures show the ensemble

spread of energies around each individual solution derived with the best-fit functional. All

surface state energies are reported in reference to a clean Pt(111) surface.

Figure S3: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 1/3 OH∗ surface phase.

6

Page 28: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure S4: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 1/4 O∗ surface phase.

Figure S5: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 1/3 O∗ surface phase.

7

Page 29: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure S6: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 1/2 O∗ surface phase.

Figure S7: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 2/3 O∗ surface phase.

8

Page 30: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure S8: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 3/4 O∗ surface phase.

Figure S9: Distribution of energies derived from 50 BEEF-vdW ensemble functionals aboutthe 1 O∗ surface phase.

9

Page 31: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

8 Prediction Confidence at Nonzero pH Levels

In section 3.1 of the main paper, we presented figure 3b showing the prediction confidence of

each surface state at a pH of 0. Here, we show that we can extend the prediction confidence of

each surface state to other pH values by shifting the x axis by -59 mV/pH. To demonstrate

this analysis, we show prediction confidence for each surface phase examined for Pt(111)

surface in figure S11 at pH of 7. At USHE of approximately 0.2 V, we observe a region of low

confidence (c-value<0.4) where the surface may be either 1/3 OH∗, 1/4 O∗, or clean Pt(111).

The 1/4 O∗ surface shows high confidence (≈0.6) at about 0.4 V which transitions to the

1/2 O∗ surface at ≈0.4 V and to the 2/3 O∗ surface phase at ≈0.8 V before the dissolution

potential of 1.18 V for Pt(111). We highlight that this shows the necessity of utilizing the

prediction confidence metric to more accurately describe surface phase transitions.

Figure S10: Prediction confidence of surface states considered on Pt(111) at pH of 7.

10

Page 32: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

9 Other Supplementary Figures

9.1 Pt(100)

Figure S11: Surface Pourbaix diagram of Pt(100) derived from the best-fit BEEF-vdW XC-functional.

11

Page 33: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Figure S12: Stability of surface phases considered on Pt(100) in terms of free energy atpH=0. Bold lines are calculated from the best-fit BEEF-vdW functional.

12

Page 34: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

9.2 Pd(111)

Figure S13: Surface Pourbaix diagram of Pd(111) derived from the best-fit BEEF-vdWXC-functional.

Figure S14: Stability of surface phases considered on Pd(111) in terms of free energy atpH=0. Bold lines are calculated from the best-fit BEEF-vdW functional.

13

Page 35: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

9.3 Ir(111)

Figure S15: Surface Pourbaix diagram of Ir(111) derived from the best-fit BEEF-vdW XC-functional.

Figure S16: Stability of surface phases considered on Ir(111) in terms of free energy at pH=0.Bold lines are calculated from the best-fit BEEF-vdW functional.

14

Page 36: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

9.4 Rh(111)

Figure S17: Surface Pourbaix diagram of Rh(111) derived from the best-fit BEEF-vdWXC-functional.

Figure S18: Stability of surface phases considered on Rh(111) in terms of free energy atpH=0. Bold lines are calculated from the best-fit BEEF-vdW functional.

15

Page 37: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

9.5 Ru(0001)

Figure S19: Surface Pourbaix diagram of Ru(0001) derived from the best-fit BEEF-vdWXC-functional.

Figure S20: Stability of surface phases considered on Ru(0001) in terms of free energy atpH=0. Bold lines are calculated from the best-fit BEEF-vdW functional.

16

Page 38: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

References

(1) Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953.

(2) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-

wave method. Phys. Rev. B 1999, 59, 1758–1775.

(3) Mortensen, J. J.; Hansen, L. B.; Jacobsen, K. W. Real-space grid implementation of

the projector augmented wave method. Phys. Rev. B 2005, 71, 035109.

(4) Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev.

B 1976, 13, 5188.

(5) Larsen, A. H. et al. The atomic simulation environment a Python library for working

with atoms. J. Phys. Condens. Matter 2017, 29, 273002.

(6) Wellendorff, J.; Lundgaard, K. T.; Møgelhøj, A.; Petzold, V.; Landis, D. D.;

Nørskov, J. K.; Bligaard, T.; Jacobsen, K. W. Density functionals for surface science:

Exchange-correlation model development with Bayesian error estimation. Phys. Rev. B

2012, 85, 235149.

(7) Medford, A. J.; Wellendorff, J.; Vojvodic, A.; Studt, F.; Abild-Pedersen, F.; Jacob-

sen, K. W.; Bligaard, T.; Nørskov, J. K. Assessing the reliability of calculated catalytic

ammonia synthesis rates. Science 2014, 345, 197–200.

(8) Deshpande, S.; Kitchin, J. R.; Viswanathan, V. Quantifying Uncertainty in Activity

Volcano Relationships for Oxygen Reduction Reaction. ACS Catal. 2016, 6, 5251–5259.

(9) Krishnamurthy, D.; Sumaria, V.; Viswanathan, V. Maximal predictability approach for

identifying the right descriptors for electrocatalytic reactions. J. Phys. Chem. Lett.

(10) Houchins, G.; Viswanathan, V. Quantifying confidence in density functional theory

predictions of magnetic ground states. Phys. Rev. B 2017, 96, 134426.

17

Page 39: QuantifyingConfidenceinDensityFunctionalTheory ... · O* and 1 O*. Through the constructed Pourbaix diagram, shown in Fig.3a, we observe that at aH+ = 1 the surface exists as clean

Graphical TOC Entry

39