<<

CROATICA CHEMICA ACTA CCACAA, ISSN 0011-1643, e-ISSN 1334-417X Croat. Chem. Acta 83 (4) (2010) 457–474. CCA-3441

Feature Article

Theory, Experiment and Computer Simulation of the Electrostatic Potential at Crystal/Electrolyte Interfaces

Piotr Zarzycki,a,* Kevin M. Rosso,a Shawn Chatman,a Tajana Preočanin,b Nikola Kallay,b and Wojciech Piaseckic

aChemical and Materials Sciences Division, Pacific Northwest National Laboratory, Richland, WA 99354, USA bLaboratory of Physical Chemistry, Department of Chemistry, Faculty of Science, University of Zagreb, HR 10000 Zagreb, Croatia cDepartment of Biochemistry, The Jozef Pilsudski Academy of Physical Education, Warsaw, Poland

RECEIVED JUNE 16, 2010; REVISED OCTOBER 1, 2010; ACCEPTED OCTOBER 5, 2010

Abstract. In this feature article we discuss recent advances and challenges in measuring, analyzing and in- terpreting the electrostatic potential development at crystal/electrolyte interfaces. We highlight progress toward fundamental understanding of historically difficult aspects, including point of zero potential esti- mation for single faces of single crystals, the non-equilibrium pH hysteresis loop, and the origin of nonlinearities in the titration response. It has been already reported that the electrostatic potential is strongly affected by many second order type phenomena such as: surface heterogeneity, (sub)surface transformations, charge transfer reactions, and additional potential jumps at crystal face edges and/or Schottky barriers. Single-crystal electrode potentials seem particularly sensitive to these phenomena, which makes interpretation of experimental observations complicated. We hope that recent theory devel- opments in our research group including an analytical model of titration hysteresis, a perturbative surface potential expansion, and a new surface complexation model that incorporates charge transfer processes will help experimental data analysis, and provide unique insights into the electrostatic response of nonpo- larizable single-crystal electrodes. Keywords: single-crystal electrode, surface potential, Nernst potential, computer simulations, electron transfer, metal oxide

INTRODUCTION These phenomena that govern the development of the electrical interfacial layer are thought to be general Metal (hydr)oxides acquire a surface electric charge across a wide range of metal (hydr)oxides, including when brought into contact with an aqueous solution.1,2 more broadly oxides of metalloids such as silicon.4 Molecular and/or dissociative adsorption of water mole- Representative microscopic and macroscopic models of cules is a probable mechanism of obtaining surface electrical interfacial layers formed at metal oxide/elec- charge at metal-terminated surface sites, whereas proton trolyte interfaces are shown in Figure 1, and include uptake is a more probable scenario for oxygen- those that can be constructed using molecular simula- terminated surface sites.3 Because actual metal oxide tion (Figure 1a), surface complexation modeling (Figure surfaces are typically structurally nonuniform, both 1b), and analytical macroscopic theories (Figure 1c). charging mechanisms play an important role. Surface Experimental characterization of metal (hydr)oxide sites exhibit an amphoteric property, that is they are able acid-base properties involves determination of the sur- to uptake or release protons and thus form negatively, face charge (σ0), surface potential (ψ0), or ζ-potential as neutral, or positively charged surface groups.1,2 The net a function of pH. is the most frequently balance of protons at the metal oxide/aqueous interface reported characteristic, and it is typically obtained from is a primary charging mechanism, and for this reason H+ potentiometric titration of a microparticulate metal and OH– ions play a key role as potential determining oxide suspension.1,5,6 The electrokinetic potential (ζ) is ions (PDI). Electrolyte ion association with surface sites usually obtained in an electrophoretic experiment.7 The is considered a secondary process, which results in surface potential is the most difficult property to deter- formation of inner- or outer-sphere surface complexes.1 mine experimentally. The difficulties lie in both the

* Author to whom correspondence should be addressed. (E-mail: [email protected] or [email protected]) 458 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

Figure 1. (a) Molecular representation of the electrical interfacial layer at the goethite (101)/electrolyte interface (using a snapshot of the computation cell used in the molecular dynamics simulations of the goethite/electrolyte interface from Reference 86); (b) the metal oxide/electrolyte interface with inner- and outer-sphere surface complexes (e.g., of surface associated electrolyte ions); (c) simplified double-layer picture of an electrical interfacial layer at the metal oxide/electrolyte interface provided by the

Surface Complexation Model. The surface, β-plane and diffuse potentials (ψ0, ψβ, ψd) are correlated through the capacity constants (c1, c2). The ζ-potential is the potential at the slip plane located at the distance l from the beginning of a diffuse layer. construction of a sensor electrode and in finding a rela- not depend on the externally applied potential. Howev- tionship between measured cell voltage and the surface er, electrical interfacial layer formation causes a flow of potential. charging current through the cell.8 This current is an After introducing the methods and tools used for example of non-Faradaic current (i.e., not related to studying the surface potential at single crystal elec- reduction-oxidation processes)8 and it decays as the trodes, we briefly discuss recent associated challenges interfacial charge distribution approaches the equili- along with some proposed solutions, which were devel- brium condition of zero net current. When the potential oped in a combined two-team effort: Kevin M. Rosso's is measured between the metal oxide electrode and a research group at the Pacific Northwest National Labor- standard reference electrode (usually an ideal nonpola- atory, Washington, U.S.A., and Nikola Kallay's research rizable electrode), then we observe a voltage drop group at the University of Zagreb in Croatia. representing the charging current, which is described by Ohm's law:8 ViR (1) Electrical Interfacial Layer Near the Nonpolarizable

Electrode where i is the charging current passing through the cell Metal (hydr)oxide electrodes are an example of nonpo- and R describes the collective resistance of the setup larizable (depolarized) electrodes, whose response does and instrument. The absolute electrode potential is un-

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 459 available, only the potential difference between the metal oxide electrode and reference electrode can be directly measured.8,9 This voltage measurement, some- times referred to as an open circuit potential or rest potential measurement, can be conducted in either a two or three electrode configuration. The two-electrode setup is justified when the current that is passed through the circuit is small, such as the case with a high instru- ment resistance (or impedance) and/or low measured 8 potentials (V < 1–2 mV). In the case of larger currents (a) (b) the voltage measurements should be conducted using a three-electrode setup, in which an auxiliary electrode Figure 2. (a) one possible design of an open circuit potential (counter electrode) is used as the current sink for the setup for surface potential measurements using a single-crystal cell, and to ensure that the potential difference is meas- electrode; (b) a possible method for constructing the single- ured as close as possible to the zero net current (equi- crystal electrode. librium) condition. It is worth mentioning that the aux- iliary electrode cannot completely eliminate the poten- oxides, the first analytical models of ψ0 were built based tial drop due to electric interfacial layer equilibration. on ISFET electrode measurements. There is still a fraction of solution resistance which remains uncompensated.8 The resistance of nonpolariz- Coated Wire Electrodes able-electrode/electrolyte interfaces also strongly de- pends on how the working electrode is constructed. For The surface potential of the metal oxide/electrolyte determination of metal oxide surface potentials, three interface can also be examined by using a coated wire electrode, that is, a metallic wire covered by the metal kinds of working electrodes have been used so far: (i) 13–16 13 oxide. Measurements using Pt/hematite(α-Fe2O3), one based on the ion sensitive field-effect transistor 14 15,16 (ISFET),10–12 (ii) one using a metallic wire covered by Ti/TiO2 or Pd/PdO electrodes serve as a good the metal oxide in a fine particulate form13–16 and (iii) illustration of this method. Due to the intrinsic porosity one using a single-crystal electrode, typically in which a of the metal oxide layer in this approach, electrolyte single face of the crystal is exposed to solution ions are in simultaneous contact with both the metal (SCrE).17–21 oxide and the metallic core. Redox reactions occurring at the metallic core therefore contribute substantially to Ion-Sensitive Field Effect Transistor Electrode the measured potential, consequently the electrode be- haves as a so-called electrode of the second-kind.22 An electrode based on the ISFET consists of a metal Despite this aspect, if the contribution from the metallic attached to a semiconductor (e.g., Si) whose surface is core is sufficiently well understood or predictable, this covered by the phase of interest, an insulator (e.g., SiO2) configuration does provide useful information about the which is in contact with an electrolyte.10,11 A typical metal oxide surface potential. However, such problems electrochemical cell with an ISFET electrode has the arising from metal oxide porosity are completely elimi- following construction:10,11 nated by the single-crystal electrode method. metal |semiconductor|insulator|electrolyte| Single-crystal Electrode 1 (2) reference electrode|metal2 The concept of using a single-crystal as an electrode goes back many years (e.g., Quinn et al.23). However, An experimental observable in this configuration is the this approach has recently begun to be used again as a flat-band potential of the semiconductor, that is, the means to access detailed information about the behavior voltage applied to the metal (metal1) which makes the at individual crystallographic terminations for face- semiconductor surface potential zero (the zero band specific electrical interfacial layer properties. Detailed bending condition).10,11 The measured flat-band voltage studies have been introduced for ice crystals,24,25 and of the semiconductor is particularly sensitive to the then adapted to single-crystals of hematite.17,18 The potential drop at the insulator/electrolyte interface basic idea is to selectively expose only a single crystal which, in turn, is developed due to metal oxide surface face and measure the potential developed at its interface protonation. This is the reason that the ISFET electrode with electrolyte, an experiment typically performed can be useful as a chemically sensitive electronic device using the three-electrode setup. The cell and one method for H+ ion detection.10,11 Although ISFET electrodes are for constructing such an electrode are schematically rarely used to determine the surface potential of metal presented in Figure 2. In order to assure electrochemical

Croat. Chem. Acta 83 (2010) 457. 460 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces equilibrium, the cell potential is measured in the open ed by surface charges and adsorbed counter ions is still circuit fashion. The current through the circuit should be present. In analytical modeling, this remaining surface kept small, requiring that the resistance of the potentio- potential contribution is assumed to be constant during meter setup is several orders of magnitude higher than variation in surface charge (and potential).29 In addition, that of the single-crystal electrode. The surface potential it is expected that only the single-crystal electrode is is measured as the difference between reference and sensitive to the change in the bulk pH; all other terms SCrE electrodes. Simultaneously, the pH is recorded using are considered as a constant shift of ψ0 (denoted by 26,27 a glass electrode with respect to the same reference elec- ET): SCrE/sol trode. More detailed discussion on construction and prepa- ψ0 = E – ET = ∆φ – C – ET (5) ration of a single-crystal electrode (SCrE) can be found in papers by Kallay and coworkers.17–21,26,27 Moreover, the surface potential is related to the change The electromotive force of the cell is simply the in the free energy of surface protonation. For instance, potential difference between the single-crystal electrode by assuming a single protonation step we can write: (the working electrode) and the reference electrode, that 28 is: GeNψ0 where  G =  G NkB T ln Q , SCrE ref (6) EE E (3) and in equilibrium  GNkTK ln B  which can be expressed as the inner potential differenc- es at both interfaces (Galvani potentials)9,28 (considering where e is the elementary charge, N is the Avogadro only monovalent ions): constant (F = eN is the Faraday constant), kB is the Boltzmann constant (R = NkB is the gas constant), Q is the reaction quotient for protonation reaction, which in E = ∆ SCrE/sol – ∆ ref/sol – φ φ equilibrium (∆G = 0) is equivalent to the equilibrium 1 ()μSCrE/solμ ref/sol C (4) constant (K+). From Equation 6 we can obtain the F  ii Nernst equation:

x/sol where ∆ expresses the Galvani potential at the kT kT 2.303 kT φ ψ BBlnK ln[H ]  B (PZC  pH) (7) given electrode (x)/electrolyte interface, the last term 0 ee e expresses the difference in chemical potential of charge carriers near the SCrE and reference electrodes, and C is where pH = –log[H+] and PZC is a point of zero charge a constant representing all remaining potential jumps in defined by the intrinsic protonation constant (K+ or a the circuit unrelated to the surface reactions (presuma- combination of equilibrium constants if more than one bly independent of PDI concentration). The Galvani protonation reaction is postulated, see Analytical Mod- potential is a difference in the electrostatic potential felt els section). The Nernst equation (Equation 7) can be by a given ion in the interior of both phases, which is also derived by assuming equality of electrochemical exactly the electric work required to transfer this ion potentials of H+ ions at the surface and in the bulk (the from the interior of one phase to the interior of the oth- equilibrium condition).29 er.9,28 The potential arising from the reference electrode At pH = PZC, the surface charge and potential are is by definition only a voltage shift with respect to the zero. This requirement is used in the experimental pro- standard hydrogen electrode, and it can be put into the con- cedure to transform the measured cell potential to the SCrE/sol stant C, giving finally the expression: E = ∆φ – C. surface potential (by choosing that ET is equal to the In addition, simultaneous measurement of the potential measured cell voltage at pH = PZC). More detailed difference between the auxiliary and reference elec- discussion on transforming E = f(pH) into ψ0 = f(pH) trodes allows one to monitor the variation in the bulk can be found below in the Challenges section. electrolyte pH. To this point we have briefly introduced the fun- The surface charge changes due to net proton re- damentals of surface measurement using the metal lease or uptake, and it usually does not involve the dif- oxide/electrolyte interface as an example. Extension of fusion of H+ ions inside the solid metal oxide phase. For these concepts to the other nonpolarizable surfaces, in this reason, the electrostatic work of transferring protons particular to the silver halides as a key model system for adsorbed at the metal oxide surface into the interior of example, is straightforward, requiring replacement of metal oxide (surface potentials, χSCrE) can be neglected H+ ion concentration (the principal PDI for metal oxide) in the Galvani potential. In other words, the Galvani by the silver or halide ion concentration. Silver halide is potential (∆φx/sol) in Equation (4) can be replaced by a well studied model nonpolarizable electrode material the by Volta potential (∆ψx/sol). However, some part of with a relatively simple composition and structure the surface contribution arising from reorientation of (rocksalt).26,29,30 At the interface between a silver halide water molecules in the strong electrostatic field generat- surface and electrolyte, the source of surface charge

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 461

development is thought to be a combination of an un- 0.95 for Ta2O5/electrolyte interfaces (see Reference 31 even adsorption of charged ions and incongruent surface and references wherein). dissolution. In this case, ions common to the electrode In the case of direct experimental ψ0-meas- composition (Ag+ or halide anions) play the role of urement, deviation from Nernstian behavior (parameter PDIs. For a long time it has been generally believed that α) is estimated by linear regression of experimental data silver halides are the Nernstian electrodes, that is, the or if possible by the following numerical derivative:30 slope of ψ0 is approximately –59 mV/pPDI unit at room 30 temperature (pPDI = –log[PDI]). However, Kallay, e dψ0 26,27 α  (11) Preočanin and Šupljika have recently reported that 2.303kTd pH the actual potential measured at the single crystal AgBr B and AgI electrodes has a slightly lower slope than the From here one can see that α is directly related to the Nernstian one (by about 12 %). variation in protonation free energy due to changes in

the bulk pH by combining Equations (6) and (11) to Analytical Models of Surface Potential obtain: One of the first analytical expressions for the surface 1 dG α  (12) potential at nonpolarizable electrodes such as metal 2.303kTd pH oxides was derived by Bousse et al.,10,11 and it is based B on measurements using ISFET electrodes. By consider- For simplicity of notation, we assume that the ing only the primary source of surface charge, namely point of zero potential is independent of electrolyte surface protonation, Bousse et al.10,11 arrived at the concentration, that is, that the PZC referred to here is following expression for surface potential as a function also the point of zero salt effect.32 This is equivalent to of pH: making the assumption that there is no preferential spe- eψ eψ cific ion adsorption (i.e., both electrolyte ions have 2.303(PZC pH)00 arcsin (8)  similar surface affinity constants). In this idealized case, kTBBβkT the point of zero charge (the pH at which σ0 = 0) is where β is a system-dependent parameter which de- equal to the point of zero potential (the pH at which ψ0 = 0) and to the (the pH at which scribes the deviation of ψ0 from Nernstian behavior. The function arcsin can be replaced by its power expan- ζ = 0). In the more complex case of preferential adsorp- sion:30 tion of one electrolyte ion relative to another, these equalities can diverge and additional nomenclature has eψ 2.303(PZC pH) 0 to be introduced (see papers by Kallay and coworkers 17–21,26 kTB for details ). Effects arising from preferential spe- 21j (9) cific ion adsorption are discussed in the Challenges  (1)(2)! j j eψ 0 section.  2(!)(21)22j jj βkT j0 B Surface Potential in Surface Complexation Models and for small values of ψ , we can neglect all terms 0 The surface potential can be also expressed in the except j = 0, which leads to the so-called quasi- framework of surface complexation modeling, that is, as Nernstian formula:10,30 a function of the surface concentrations of protonated and deprotonated metal oxide surface groups. In the 33 β 2.303kTB case of the so-called 1-pK model, the surface proton ψ0  (PZC pH) β 1 e balance is described by a single reaction:  (10)

2.303kTB α (PZC pH) 1/2  1/2 e SOH H SOH2 (13)

The prefactor β/(β + 1) is usually expressed as a single with an intrinsic equilibrium constant defined as: parameter (α), which describes the decrease in the elec- trode response with respect to the ideal Nernstian beha- 1/ 2 SOH2  K  (14) vior. Values for this parameter are typically obtained by  1/2 int indirect fitting of a surface complexation model entail- SOH [H ] ing a presumed function for the relationship ψ0 = f(pH). For instance, α has been found to be about 0.8 for where {} brackets stand for the equilibrium surface alumina, 0.9 for TiO2, 0.75 for SiO2 and approximately concentration of protonated, deprotonated sites, whereas

Croat. Chem. Acta 83 (2010) 457. 462 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

int [] stands for the concentration of the unbound surface 1/2 2.303kT kT SOH2  + ψ BB(PZC pH) ln protons. The concentration of H ions at the surface can 0 1/2 + ee2 SOH be connected to the bulk H ion concentration by using (20) the Boltzmann distribution: 2.303kTBB kT θ (PZC pH) ln ee2 θ int bulk eψ [H ] [H ] exp0 (15) kTB where PZC = –(pK1+pK2)/2. Again, we can try to ex- press the parameter α from Equation (10) as a function By inserting Equation (15) into Equation (14) we of surface coverage: can define the surface potential as follows:17–19,34 1 θ α 1ln (21) 1/2 4.606(PZC pH) θ 2.303kT kT SOH2  ψ BB(PZC pH) ln 0 eeSOH1/2  (16) Finally, in the case of the multisite complexation model (MUSIC),37,38 wherein surface sites are treated as pos- 2.303kTBB kT θ (PZC pH) ln sessing fractional charge whose magnitude (z) depends eeθ  on the effective metal ion coordination, the general protonation reaction can be written as follows (using a where PZC = –pK+, and θ+,θ– stand for the fractional single protonation reaction as an analogue of the 1-pK occupation for positively and negatively charged surface model):34 z  1 z groups at a given pH (i.e., from use of an adsorption SOH H SOH2 (22) isotherm curve (θi = [i]/NS , where is the surface site density). Equation (16) predicts deviation from the with the equilibrium constant defined as: Nernst equation as a function of surface protona- tion/deprotonation processes; accordingly the previously 1z SOH2  30 K  (23) introduced parameter α can be redefined as: zint SOH [H ] 1 θ α 1ln (17) Applying a similar procedure of replacing surface H+ 2.303(PZC pH) θ concentration by the bulk one (Equation (15)) allows us

to express the surface potential as:34 which is well defined only for pH ≠ PZC and θi ≠ 0 (but can be estimated in the limit PZC – pH → 0 by employ- 1z 2.303kT kT (1 z ) SOH 2 ing the Bernoulli-l'Hospital rule, see Reference 30). In ψ BB(PZC pH) ln the case of linear surface potential α = const. and conse- 0 eez SOHz (24) quently (ln(θ /θ ))/(PZC – pH) = const., otherwise α + – 2.303kT kT θ becomes a function of pH. BB(PZC pH) ln  35,36 eeθ In the case of the 2-pK model, proton up-  take/release is described by two surface reactions: Now the deviation from the Nernst response (α) can be expressed as:    SO H SOH and SOH H SOH2 (18) 11 z θ α 1ln (25) 2.303(PZC pH) z θ with equilibrium constants defined as: 

In the two-reaction version of the MUSIC model,37,38  SOH SOH2  surface protonation is described as: K and K (19) 12SO [H ]int SOH [H ]int SOHzz11 H  SOH1 and 2 (26) + zz22  1 By replacing surface H concentration by the bulk SOH H SOH2 one (Equation 15), and rewriting both equations in order to eliminate [SOH], we can define the surface and the corresponding intrinsic equilibrium constants potential as:17–19 are given by:

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 463

11zz12 SOH22  SOH  K12 and K (27) SO(1)zz12 [H ]int SO (  1) [H ] int and the corresponding expression for the surface poten- tial is:

kT z12θzz z θ ψ B ln 12 0 ez(1 )θ Kz (1 )θ K 11zz12 2 2 (28) 2.303kT B pH e where θθ, represent the fractional surface occupa- z12z tions of the empty (negatively charged) sites with Figure 3. Lattice model of electrical interfacial layer formed charges z1 and z2, respectively. at the metal oxide/electrolyte interface used for Grand- Because the surface concentrations of individual Canonical Monte Carlo and Kinetic Monte Carlo simulations ions are rarely available, these definitions (Equations (see References 30, 31, 34, 44–50). The two continuum dielec- (16,20,27)) are of limited experimental use in practice. tric regions are used to describe the interface and bulk phase, However, they allow us explain the deviation of ψ from 0 with dielectric constants εS and εW, respectively. ideal Nernstian behavior as the consequence of surface complexation and protonations reactions. of surface charge.42,43 This approach includes use of a In principle, the surface electrostatic response of pair-wise electrostatic interaction potential between nonpolarizable electrodes can be defined as the ideal charged surface groups, with the effect of intervening Nernstian behavior (α = 1) decreased by the term ψ0(θ) electrolyte treated in a mean-field way by using Debye (containing all surface protonation/complexation screening. We have expanded this model by including 30 processes): effects of surface energetic44 or topographic nonunifor- mity, as well as by expanding to the cellular Potts model quasi-Nernst Nernst Nernst ψαψψψθ0000(pH) (pH) (pH) ( ) (29) analog to study the effect of electrolyte ion surface association.45,46 In our studies, equilibrium properties Computer Simulation of the Surface Potential: were obtained using Grand Canonical Monte Carlo Monte Carlo Lattice Model algorithm,44–46 whereas kinetic ones were obtained by using the Grand Canonical Kinetic Monte Carlo me- One of the most popular computational methods in thod.47,48 A basic model lattice we have used in combi- adsorption science is numerical simulation of a lattice nation with the Borkovec interaction potential to simu- surface model, which is an adaptation of the Ising model late the metal oxide/electrolyte interface is schematical- of a ferromagnetic material for solid/(liquid/gas) inter- 39,40 ly presented in Figure 3. Although this approach seems faces. For years it has been used as a numerical tool very simple, it has proven to be a valuable tool for using in testing analytical theories (e.g., to simulate adsorption 39,40 microscopic principles to test the validity of popular isotherms and surface phase transitions). To a cer- surface analytical models, such as ones that are based on tain extent, this simulation method can be viewed as a 44,49 41 the integral adsorption isotherm approach. Further- computer experiment. This approach has nonetheless more, because individual adsorption isotherms are direct provided very useful and unique insight into microscop- observables in lattice Monte Carlo simulation, we can ic chemical and physical phenomena behind macroscop- evaluate the equilibrium surface potential by employing ic behavior. In particular, lattice models can be tailored appropriate formula (e.g., from Equations (16,20, to a desired configuration to enable simultaneous devel- 24,27)).30,34 More detailed description of these simula- opment of analytical theoretical models. And in particu- tions and their utility for metal oxide/electrolyte inter- lar, strengths of this approach include computational faces can be found in our previous papers.30,34,47–50 efficiency, and the ability to switch elementary processes in the system on/off depending on the model Computer simulation of the surface potential: requirements, a feature that enables study of, for exam- Molecular dynamics ple, the effects of individual microscopic processes on macroscopic thermodynamic observables. Molecular dynamics, a widely used atomistic simulation A numeric lattice model description of the metal method that depends heavily on the accuracy of empiri- oxide/electrolyte interface was introduced by Borkovec cally derived interatomic interaction potentials, allows and coworkers.42,43 In its original formulation, the Bor- study of both equilibrium and kinetic properties of a kovec model includes only the uptake/release of surface metal oxide/electrolyte interfacial system. However, this H+ ions on a simple surface lattice as the primary source approach is in general more computationally intensive

Croat. Chem. Acta 83 (2010) 457. 464 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

inert solvent, and the halide or silver ions play the role of PDIs. In Figure 4 we show a molecular dynamics simulation model recently used to calculate electrostatic 3 properties of the AgCl(100)/KCl(aq)(0.24 mol/dm ) inter- face, along with the calculated electrostatic potential profile.56,57 As shown in Figure 4, the simulation cell in this instance is composed of the crystal slab and the electrolyte phase periodically replicated in three dimen- sions. The electrostatic potential profile is obtained by a (a) double integration of the ion and water charge density curves:60–65

1 zz ψ()z  σζ ()dζ (30) ε0 00

where ε0 is the dielectric permittivity of free space, z is the distance from the surface (z = 0) into the bulk elec- (b) trolyte phase and ζ represents the integration variable (distance). Figure 4. (a) simulation cell used in estimating electrostatic properties of AgCl/electrolyte interfaces using molecular CHALLENGES dynamics simulations; (b) The electrostatic potential between both solid planes calculated for the AgCl/KCl electrolyte (aq) Determination of Surface Potential Using a Single- system periodically replicated in each direction.The details of Crystal Electrode the molecular dynamics calculations can be found in Refer- ences 56 and 57. As mentioned above, in order to convert the measured electromotive force of the cell based on a single-crystal than in the Monte Carlo lattice model approach and electrode to the surface potential, the voltage shift (ET, therefore can suffer from inability to reach into long see Equation (5)) has to be known in advance. The sim- time scales (e.g., greater than a few tens or hundreds of plest way to assess ET is based on the requirement that nanoseconds) or very large system sizes (e.g., to simu- surface potential at pH = PZC is equal zero. If the PZC late low electrolyte concentration). Furthermore, incor- of a given crystal face exposed to the electrolyte was poration of a chemical reaction, which entails bond known, in principle one could simply set ET = E(PZC). making and bond breaking events, is much more diffi- Unfortunately, the PZC of individual crystal faces are cult than in a Monte Carlo lattice approach. Aspects of largely unavailable. It is much more common to have the difficulty lie in the ambiguity of defining universal only either the PZC or IEP from batch potentiometric criteria for when specific bonds are formed or broken, titration or electrophoretic measurements on powder and avoiding discontinuities in the potential energy metal oxide particulates, typically of unknown or un- surface when such events are specified to occur. There- specified crystallite morphologies.66 In both experimen- fore, in a molecular dynamics simulation of the inter- tal procedures, because a metal oxide suspension is face, chemical entities are usually restricted to remain used,67 the determined PZC/IEP represents is an average unchanged as the system dynamics evolve. To our best of numerous surfaces and types of surface sites (differ- knowledge, for metal oxide surfaces and their interfaces ent crystal faces, terminations, edges, kinks, defects). with aqueous solution, there is currently only one mole- Hence, use of this average characteristic for shifting the cular dynamics model that has successfully incorporated cell voltage for a SCrE setup is dubious. Furthermore, water dissociation.51,52 For these reasons, use of molecu- an estimation of the potential slope (α) is available di- lar dynamics simulation to analyze metal oxide surface rectly from SCrE measurements, and it does not require electrostatic properties as a function of pH is very li- prior knowledge of PZC. mited. There are, however, numerous studies of hy- Although in general there is no a priori knowledge drated metal oxides using non-dissociable water models of the PZC of a single crystal face, one can try to esti- (e.g., References (53–55)), which can provide some mate it based on the presumed or measured stoichiome- insight into the distribution of charge/ions across the try of the exposed crystal face using the MUSIC or CD- interface. MUSIC surface complexation models.37,38 However, in Fortunately, in the case of the silver halide/ practice this still requires an unsatisfying number of electrolyte interfaces, water molecules can be treated as assumptions to be made about the surface termination

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 465 and its level of compositional and structural perfection, and the activity of different surface sites. The estimated PZC will therefore be strongly model-dependent. In some of the first uses of an SCrE, the value of ET was approximated using the PZC or IEP obtained for the powdered metal oxides. However, recently we in- troduced a novel approach for estimating the PZC of individual crystal faces based on the observation that in SCrE measurements involving sequential forward and reverse pH titration typically one finds reproducible hysteresis in the measured cell potential as a function of pH. This phenomenon was first reported for the case of a single-crystal hematite electrode by Preočanin and Kallay.18 A similar hysteresis loop has also been ob- served in batch potentiometric titration68 and calorime- tric measurements.69 The kinetic character of hysteresis was immediately recognized: by extending the titration time or applying ultrasound mixing the hysteresis sepa- rating the acidimetric and alkalimetric titration branches could be reduced.18 If no steps are taken to reduce the Figure 5. A measured titration hysteresis loop obtained using hysteresis, one is left with the conclusion that in the single-crystal electrode measurements (markers) (hematite 3 region of separated branches the SCrE surface is under- 021/0.001 mold/dm NaNO3). By adapting the concept of the going overcharging in the alkalimetric titration (overpo- Maxwell construction to the problem of titration hysteresis, we tential- slightly higher potential than we should observe can estimate the surface PZC value and shift in cell potential at the equilibrium), and undercharging during the acidi- (Epzp=ET) as equal to 7.2, and 35 mV, respectively. The ex- metric titration (underpotential). Surface overcharging perimental and numerical details can be found in Reference 70. (overpotential) is a manifestation of the fact that not all surface protons were able to leave the surface before ponding to the PZC of the fully equilibrated system. another aliquot of base was added (with respect to equi- Based on the theory of the ferromagnetic hysteresis, we librium). Conversely, surface undercharging (under- found that this zero potential condition (pH = PZP) can potential) is evidence that not all protons were able to be be found using the Maxwell construction of an equal bound to the surface before another aliquot of acid was area. The pH = PZP corresponds to the vertical plane added.70 (pH = ζ) which divides hysteresis into two equal areas Given this behavior, the difference in measured (see Reference 70 for more details): potential between the acidimetric (ba, basic to acidic ζ pHend pH) and alkalimetric (ab, acidic to basic pH) titration I (ζ )Ed (pH) pH  Ed (pH) pH branches gives the free energy difference between both  (32) pHstart ζ titration curves: where EEab  E ba

1 EEab  E ba ()  G ab  G ba ψψ ab  ba (31) eN 00 Application of this procedure is illustrated for experi- mental data obtained for hematite(021)/0.001 mol/dm3 Therefore, under this condition, the cell, in which the NaClO4 electrolyte solution in Figure 5 (data from Ref- overall cell reaction has not yet reached equilibrium, erence 70). This procedure therefore allows us to deter- can do electrical work.8,9 Furthermore, as is the case in mine the PZP of individual crystal faces directly from magnetic hysteresis phenomena, the area enclosed by the titration experiment. Such an analysis provides the hysteresis loop is the work done on the system.71 unique insight into the acid-base response of metal The difference (∆E, Equation (31)) approaches zero as oxide surfaces not afforded by other surface spectros- the titration time is extended and the system is allowed copic probes. In principle, with knowledge of the sur- sufficient time to reach equilibrium at each pH value. In face structure, it enables one to estimate protonation principle, one should be able to find a pH value at which constants of individual crystal faces from experiment over/undercharging of the positively charged surface for the first time. For a more detailed description of this has the same free energy as over/undercharging of the approach, see Reference 70. negatively charged surface. This pH value should Another way of estimating the surface PZP is roughly estimate the pH of electrostatic metastability of based on analogy to Gran plots used in potentiometric the cell, and should be identical to the pH value corres- selective electrode titration.9 If high resolution cell po-

Croat. Chem. Acta 83 (2010) 457. 466 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

Figure 6. Theoretical surface charge density curve

(σ0* = f(pH)) and its first (–dσ0* /dpH) and second derivatives 2 2 Figure 7. Types of surface potential response to the change in (–d σ0* /dpH ) with respect to pH. The pH = PZC is an inflec- tion point of the charging curve. Results were obtained using potential determining ions (here H+ ions): (i) Nernstian re- Grand Canonical Monte Carlo simulations of a lattice model sponse, expected for the ideally nonpolarizable electrode (e.g., (see Reference 70 for details). Ag|AgCl), (ii) quasi-Nernst, expected for metal (hydr)oxide/ electrolyte interfaces (linear function of pH, but lower than tential measurement can be achieved (i.e., the ability to Nernstian slope), (iii) non-Nernstian (or non-linear) behavior, collect a large number of E = f(pH) points), the PZP can where the slope of the response is not constant and varies with be estimated by using numerical derivative (dE/dpH). It the bulk pH. has been experimentally proved using high resolution potentiometric titration,72,73 and Monte Carlo simulation tial as a function of pH, and also proved that the actual (Figure 6) that the surface charging curve possess an electrode response is somewhat of lower slope than inflection point at the pH = PZC. An analogous property predicted by the Nernst equation (Equation (7)). As can be expected from the surface potential curve, that is, mentioned before, this lower slope in the relationship is the maximum of the dE/dpH curve should coincide with quantitatively described by the α parameter in the quasi- the PZP (and d2E/dpH2 = 0 at pH = PZP). There are two Nernst equation (Equation (10)). However, recently the major obstacles to this approach in practice. First, as expectation of linear electrostatic response for metal already mentioned, the typical potential measurement oxide electrodes has started to be questioned on grounds shows a hysteresis loop which obscures the correct that at a fundamental level these surfaces are typically 30 inflection point. It is also possible at times to obtain an structurally and compositionally irregular. This gives essentially linear profile, which results in a constant rise to many possible sources of nonlinearity; one of the value of dE/dpH (each point will be an inflection point). most striking is the coupling of electrostatic effects and Secondly, current experimental methods typically do surface morphology. A distribution of possible not allow one to collect sufficiently high-resolution data ion/proton affinity constants (i.e., a spectrum of adsorp- to obtain a reliable derivative of the cell potential with tion energies) arises from the fact that at the atomic respect to pH. Hence this method is somewhat relegated scale, metal oxide surfaces are generally defective, with to systems in which hysteresis can be eliminated and the potential to deviate substantially from ideal crystal- where experimental capabilities are uniquely capable for lographic terminations. Although it has also been found this application. that, to some extent, long-range electrostatic interaction between surface sites couples with structural reconstruc- Nonlinear Single-crystal Electrode Response tion/relaxation to smooth an initially rough surface (e.g., one possible mechanism of surface free energy minimi- In the surface complexation community, there is a zation), an energetic spectrum of surface adsorption strong belief that surface potential should be a linear energies nonetheless remains from one termination to function of potential determining ion concentration at the next. Hence, at any particular metal oxide surface, the surface (ψ0 ≈ pH). This is understandable given that the ensemble of surface sites should be considered to be the metal oxides represent one kind of nonpolarizable in a state of dynamic equilibrium with each other. In surface, which in general is expected to obey the Nernst some cases this dynamic equilibrium appears to extend equation (e.g., similar to the Ag|AgCl reference elec- even further to intersurface interaction. For example, trode). Early measurements based on the ISFET elec- Yanina and Rosso74 showed near-macroscopic scale trode confirmed the expected linearity of surface poten- evidence that net structural distinction between surfaces

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 467 of a hematite crystal is sufficient to couple oxidative deposition of ferrous iron at one surface type to reduc- tive dissolution of ferric iron at another surface type, mediated by transfer of electrical current through the crystal bulk. The thermally activated self-diffusion of electrons (i.e., as small polarons) through the crystal is sufficient to allow multiple exposed crystal faces to remain in electrostatic contact and as a result induce dissolution of one and growth of another.74 Such inter- surface interaction warrants further investigation with respect to its role in influencing the surface potential at Figure 8. Schematic illustration of electrostatic screening at any particular surface of a whole crystal exposed to sterically constrained surface sites on energetical- electrolyte solution. ly/topographically rough surfaces. Surface sites exposed on For SCrE measurements, it turns out that the typi- the edges and kinks are in energetically excited states that cal ψ0 = f(pH) profile is only rarely a linear function. relax through proton adsorption, ion complexation, and local Representative Nernstian, quasi-Nernstian and non- geometric relaxation. However, long-range electrostatic repul- Nernstian (i.e., nonlinear) surface potential responses sion between adsorbed ions/protons can still be present and are illustrated in Figure 7. In past practice, the linearity can prevent some surface sites from relaxing via adsorption. of surface potential has required that the parameter α is independent of pH, that is: order (i). In principle, they depend on the source of nonlinearity, for example, arising from surface hetero- dα d 2ψ dG2 geneity or surface phase transformations and/or accom- 0, 0 0, or  0 (33) dpH ddpH22 pH panying reactions. However, application of Equation (34) to experimental data awaits the availability of high- er resolution SCrE surface potential measurements. The linearity can be analyzed by plotting α = f(pH), as 10 The relationship between nonlinear electrode re- already suggested by Bousse et al. In addition, earlier sponse and surface heterogeneity deserves further dis- we compared SCM expressions for the surface potential cussion.30 Surface heterogeneity so far has been dis- (Equations (16,20,24,27)) with a quasi-Nernst one (Eq- cussed in terms of coupled structural and chemical sur- uation (10)) and were able to re-express α as a function face complexity; a natural convolution of topographic of fractional surface coverage (Equations (17,21,25)). variability yielding a broad range of adsorption energies An outcome is that these expressions show that the for PDIs. Here for the purposes of illustrating separate slope of depends on the adsorption progress, and can be effects via computer simulation, we can discuss an expected to vary with pH (Equation (29)). A limitation operative distinction between pure topographic surface is the fact that α is composed somewhat differently in heterogeneity involving only surface roughness and each expression and the choice of which form to apply structurally distinct surface sites having differences remains somewhat arbitrary. limited in scope to geometric variability in physical To deal with this issue, we recently proposed a access and intersite electrostatic interaction, and pure new way of analyzing surface potential nonlinearity, energetic surface heterogeneity involving a discrete or based on pertubative expansion of the free energy with 30 continuous spectrum of PDI adsorption energies alone respect to reduced pH units (ζ = PZC – pH). By consi- without ties to surface topography (i.e., on a flat sur- dering an idealized Nernstian electrode as a reference face). It has been already shown that surface energetic system (unperturbed), we introduce terms for both devi- heterogeneity has a tremendous effect on acid-base ation from Nernstian behavior as well as the deviation 44–50 30 properties of metal oxide/electrolyte interfaces. from linearity as perturbations: Energetic heterogeneity is considered in lattice models by assuming a discrete/continuous spectrum of PDI 2.303kT ψαζ B i (34) adsorption energies (proton and ion affinity con- 0  i 44–50 e i1 stants). Perhaps more complicated is the separate role of topographic heterogeneity. We recently showed If the expansion is truncated after i = 1, than we that topographic heterogeneity, and any associated local obtain the quasi-Nernst equation (Equation (10)). In variation in intrinsic affinity, can be effectively screened principle, all higher terms in the expansion (i.e., by electrostatic interaction.50 As illustrated by Figure 8, i (kBT/(ne))αiζ for i > 1) have the meaning of perturbative even in the event that neighboring sites both possess corrections to the linear quasi-Nernst equation. Based strong proton affinity, the occupation of one site can on computer simulation results, we found that the mag- prevent the neighboring one from being occupied due to nitudes of the expansion coefficients (|αi|) decrease with electrostatic repulsion. Hence because site-to-site elec-

Croat. Chem. Acta 83 (2010) 457. 468 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces trostatic forces are of a longer range than chemical ones Specific Adsorption they can dominate the local site reactivity. As already As mentioned above, in order to simplify the notation mentioned, because it naturally is convoluted with site we assumed a lack of specific ion adsorption (i.e., reactivity, this effect is strong enough to help reduce similar affinities of electrolyte ions), which allowed us surface roughness through dynamic restructuring of the to write all equations using only the point of zero surface.45 In particular, this phenomenon needs to be charge. However, specific ion adsorption has been taken into account in any attempt to project the topo- frequently invoked to explain the data from SCrE mea- graphic heterogeneity studied in the gas-phase to surements,21,66 and it therefore is advisable to use the charged interfaces in aqueous solution (e.g., the surface term point of zero potential (PZP) while discussing pH area determined by BET adsorption isotherm analysis at which ψ = 0.21,66 By specific adsorption we mean does necessarily correspond to the actual surface area 0 that in the accumulation of a particular ion at the sur- when in contact with an aqueous phase). face the formation of chemical bonds is involved. The Although energetic heterogeneity is a well justi- formation of an electrical interfacial layer in the ab- fied and accepted characteristic of powder metal oxide sence of specific adsorption is driven by the electros- particulates, its applicability is less immediately clear in tatic interactions, and thus the variation in ionic the case of an SCrE electrode. However, even though a strength should not affect the point of zero charge. In single-crystal electrode is typically designed to expose this case, at pH < PZC the surface is positively only a single crystal face, it is often then required to charged and anion adsorption dominates, whereas at endure prolonged contact with aqueous solution. The pH > PZC the surface is negatively charged and cation high reactivity of freshly cleaved or otherwise dry pre- adsorption dominates. If specific adsorption occurs, pared oxide surfaces toward water (e.g., to water chemi- chemical bonding forces are also involved and ion- sorption at very low partial pressure) has already re- 75 surface interactions can be strong enough to overcome ported. Progressive surface relaxation/reconstruction electrostatic repulsion (e.g., anions can accumulate at a processes can thus occur to reduce the surface free negatively charged surface: pH > PZP, ψ < 0, or ca- energy, yielding a fundamentally different termination 0 tions at pH < PZP, ψ0 > 0). Specific adsorption is of the selected crystal face. It is also possible that an usually detected by looking at the dependence of the energetically more stable face forms by deposition on PZC and/or IEP on the ionic strength: the top of the original one. This dynamic phenomena can result in surface roughening or amorphization, par- ddPZC IEP ticularly if the electrode is extensively exposed to pH  0 and  0 (35) extremes. The likelihood of all such phenomena suggest dI dI that the most reliable ψ0-measurement possible requires that the measurements be accompanied by careful elec- Because specific adsorption varies with electrolyte trode surface monitoring (e.g., by using scanning probe concentration, the positions of the PZC and IEP depend microscopic techniques). on ionic strength. This is a known as the Esin-Markov effect and it is commonly encountered at met- We end this section by noting a key area for future 9 work. In most cases for metal oxide/electrolyte interfac- al/electrolyte interfaces. Quantitative description of the PZC dependence on specific adsorption is provided by es the interfacial charge is developed exclusively on the 9 metal oxide surface and any redistribution of charge the Esin-Markov coefficient (βEM): within the solid itself neglected. However, in the case of semiconducting metal oxides, the space-charge region 1PZCd βEM  (36) can be neglected only at certain value of the applied kTdB ln c potential, at which the number of electrons removed and supplied to electrode is equal.9 If this is not the case, In the case of symmetric ion adsorption (e.g., elec- there is additional potential drop within the space- trostatically driven complexation) we observe an equal charge region, with a finite capacity (i.e., the Mott- adsorption of electrolyte ions at pH = PZC (e.g., within 9 1/ 2 1/ 2  Schottky capacity). This capacity is usually a linear the 1-pK model: [SOH22 A ] [SOH C ]) and con- function of applied potential, but nonlinearity appears if sequently self-balancing of protonated/deprotonated 9 1/2 there are adsorbed species on semiconductor surface. sites (e.g., within the 1-pK model: [SOH2 ]  1/ 2 Moreover, the presence of the electrical interfacial layer [SOH2 ]). In this instance, the concentration of sur- results in another Schottky barrier formation at the sem- face complexed ions increases with ionic strength, but iconductor/electrolyte interface. These potential jumps due to the equality in their relative amounts at the PZC, are presently neglected in the Surface Complexation the position of the point of zero charge remains inde- Modeling approach but warrant further investigation. pendent of the electrolyte concentration.

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 469

(a) (b)

Figure 9. The dependence of surface potential on ionic strength in the case of cation (a) and anion (b) preferential adsorption. The shift in PZC with ionic strength and the relative position of the common intersection point (CIP) also known as the point of zero salt effect are schematically presented.

In the case of preferential cation adsorption we observed at pH > PZC (at ψ0 < 0). This situation is observe a higher surface concentration of adsorbed depicted in Figure 9b. cations than anions at pH=PZC (within the 1-pK mod- Specific ion adsorption is observed only rarely in 1/2 1/2  el: [SOH22 A ] [SOH C ]). This discrepancy in a potentiometric titration experiment, and it is usually ion surface concentration increases with increasing treated as an anomaly of the more commonly observed ionic strength. Because the cations are bound to the zero salt effect (see Reference 76 and references there- deprotonated groups (SOH1/2– in 1-pK) they effectively in). For instance, Cs+ ions are found to be preferential- contribute negative charge to the surface ly adsorbed on anatase,76,77 with the CIP at pH < PZC 2 + charge/potential. The protonation balance therefore has (σ0 = 0.2 µC/cm ). Similarly, complexation of Li on to be shifted toward increasing concentration of hematite78 is driven by specific adsorption. Preferential 1/ 2  SOH2 groups to maintain the overall surface neutral- adsorption of NO3 ions on Al2O3, with the CIP at 2 79 ity, and consequently the PZC is shifted towards lower pH > PZC (σ0  –3.5 µC/cm ) serves as an illustra- pH values (shift increases with an increase in ionic tion of anion specific adsorption. strength). At the point in which surface concentrations Contrary to potentiometric titration experiments of electrolyte ions are equal (i.e., in 1-pK model: of metal oxide suspensions, specific ion adsorption 1/ 2 1/ 2  [SOH22 A ] [SOH C ]) the titration curves inter- seems to be frequently observed in SCrE measure- sect. This pH value is called the common intersection ments.80 One can deduce from this that this is evi- point (CIP) or point of zero salt effect (PZSE).32 In the dence of the compensation effect of simultaneous case of preferential cation adsorption, the CIP can be exposure of many crystal faces in the former ap- achieved if anion adsorption is facilitated by electros- proach. The single-crystal electrode can detect specif- tatic attraction (i.e., the CIP is located at ic adsorption undetected by surface charge/ζ-potential pH < PZC). This case is illustrated in Figure 9a. measurements on suspensions, which also calls into The opposite situation is observed if anions are question the use of the IEP or PZC to transform meas- preferentially adsorbed. In this case, at pH = PZC we ured cell potential into surface potential. In summary, observe a higher surface concentration of adsorbed a single-crystal electrode method is very sensitive in anions than cations (i.e., in the 1-pK model: tracing the acid-base properties of metal oxide surfac- 1/2 1/2  [SOH22 A ] [SOH C ]). Because, the adsorbed es, and all things considered it seems to emerge supe- anions effectively contribute positive charge to the rior to electrokinetic/potentiometric titration experi- surface potential, to reach surface electroneutrality ments on suspensions. deprotonated sites have to balance the complexed anions and protonated sites. The PZC therefore shifts Electron Transfer towards higher pH values, and this shift is also directly proportional to the ionic strength. The equality of sur- Redox processes at the metal oxide surface can be also face concentration of complexed electrolyte ions can considered as the source of a nonlinear metal oxide be obtained if cation adsorption is enhanced by the electrode response. Electron transfer is a Franck- electrostatic attraction, and consequently the CIP is Condon process, which means that charge transfer

Croat. Chem. Acta 83 (2010) 457. 470 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

simulation methods, in particular the umbrella sam- pling technique developed by Warshel and cowork- ers.85 In Figure 11, we show simulation results for Fe(II) interfacial electron transfer with Fe(III) in the (101) surface of goethite in contact with aqueous solu- tion. In Panel A, the calculation of potential of mean force for the approach of Fe(II) to the surface is used to identify stable inner/outer surface complexes of adsorbed Fe(II), whereas in panel B the energetics of electron transfer in an inner-sphere configuration and within the mineral subsurface region are presented. More detailed discussion can be found in our recent paper.86

CONCLUSIONS

In this feature article, we summarized current experi- mental, theoretical, computational methods for analyz- ing and understanding the electrostatic potential at single-crystal electrodes of metal (hydr)oxides. Sub- stantial progress has been made in use of single-crystal electrodes as a tool for understanding fundamental Figure 10. Surface complexation model combined with an chemical and physical processes governing the elec- 2+ adiabatic interfacial electron transfer event (example for Fe trostatic potential distribution across metal ion adsorption at the goethite/electrolyte interface (see Refer- (hydr)oxide/electrolyte interfaces, but challenges re- ence 83). main. Improvements in the design of single-crystal electrodes are still under development. Here we have occurs fast compared to the timescale of nuclear fluc- overviewed some of the present difficulties and and tuations, and therefore occurs without affecting the have suggested solutions. One of them is to make use 81 nuclear degrees of freedom. However, prior to the of the hysteresis observed in cyclic titration experi- actual charge transfer, the nuclear configuration of ments to directly estimate the point of zero potential donor and acceptor have to reach a transition state for a well-defined single-crystal face. Analysis of geometry at which the electronic states of the donor evaluated ψ -profiles is also a difficult task, and many 82 0 and acceptor are energetically degenerate. In the second order processes can contribute to non- case of metal oxides, the electron transfer is typically Nernstian behavior. In particular, the surface rough- an adiabatic process, that is, there is strong electronic ness/recrystalization, redox reactions, specific adsorp- coupling between donor and acceptor wavefunctions tion and unrecognized potential jumps across the inter- at the transition state geometry, which substantially face have to be better understood, quantified, and tak- decreases the electron transfer energetic barrier (Fig- en into account. The experimental procedure indeed ure 10). The rate of adiabatic electron transfer can be requires more extensive study, and the various experi- described using the classical Arrhenius formalism, mental/theoretical procedures outlined herein are cur- hence electron transfer can be considered in the sim- rently being refined and tested between Kevin M. plest sense as an additional reaction in Surface Com- Rosso's and Nikola Kallay's research teams. plexation Models (see Figure 10).83 We are currently developing an analytical version of such a model Acknowledgements. This work was supported by the grant which includes the possibility of adiabatic electron from the US Department of Energy, Office of Basic Energy transfer events.83 The accompanying change in the Sciences, Geosciences Program. The research was performed surface site oxidation state affects the surface charge using EMSL, a national scientific user facility, sponsored by and potential, and in principle this is can be detected the Department of Energy’s Office of Biological and Envi- by a single crystal electrode measurement (see Refer- ronmental Research located at Pacific Northwest National ences 74 and 87). Laboratory. Tajana Preočanin and Nikola Kallay were sup- Molecular-level insight into such electron transfer ported by the Ministry of Science, Education and Sports of process can be obtained using computational molecular the Republic of Croatia (Project No. 119-1191342-2961).

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 471

(a)

(b)

Figure 11. Molecular dynamics study of electron transfer across the goethite 101/electrolyte interface and into the solid. The potential of mean force (a) enables identification of possible inner and outer-sphere complexes. Free energy profiles (b) were obtained using the umbrella sampling technique (for details see Reference 85).

Glossary E potential of the cell ESCrE the absolute potential of the Single-Crystal electrode Eref the absolute potential of the reference electrode V additional voltage drop due to formation of electrical interfacial layer i electric current caused by the formation of electrical interfacial layer R Ohmic resistance due to formation of the electrical interfacial layer ∆φSCr/sol the Galvani potential at the single crystal/electrolyte interface (continued)

Croat. Chem. Acta 83 (2010) 457. 472 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

∆φref/sol the Galvani potential at the reference/electrolyte interface SCr/sol μi chemical potential of potential determining ions near the Single Crystal Electrode ref/sol μi chemical potential of potential determining ions near the reference electrode constant shift in measured cell potential. It represents all potential jumps within the cell which do not depend on the C concentration of potential determining ions χSCrE electrostatic work required to transfer an ion across the electrode/electrolyte interface (part of the Galvani potential)

ET the difference between the measured cell potential (E) and surface potential (ψ0) ∆G the free energy difference for the cell ∆G° The free energy difference of the cell in the equilibrium state (∆G = 0) Q Reaction coefficient (equivalent of the equilibrium constant for non-equilibrium condition)

K+, K1, K2 Surface protonation equilibrium constant α, β Parameter describing deviation for the slope of the surface potential from Nernstian model Eab the measured cell potential in alkalimetric titration (ab: from acidic pH to basic pH) Eba the measured cell potential in acidimetric titration (ba: from basic pH to acidic pH) ∆Gab the free energy of the cell in alkalimetric titration ∆Gba The free energy surface in acidimetric titration ab ψ0 Surface potential obtained in the alkalimetric titration ba ψ0 Surface potential obtained in the acidimetric titration The area difference between left and right-hand sided of titration loop with respect to the position of dividing I(ζ) surface (ζ)

αi Surface potential nonlinearity coefficients in perturbative expansion ζ Integration variable (pH) in the titration hysteresis analysis method IEP Isoelectric point, defined as the value of pH at which electrokinetic potential (ζ)is equal 0, pHIEP

PZP Point of zero potential, defined as the pH value at which (ψ0 = 0), pHPZP Point of zero charge, defined as the pH value at which (σ = 0), pHPZC (extended classification can be found in PZC 0 Reference 32) Point of Zero Salt Effect, pH at which potentiometric titration curves obtained for various electrolyte concentration PZSE intersect, pHPZSE

REFERENCES 12. R. E. G. van Hal, J. C. T. Eijkel, and P. Bergveld, Adv. Interface Sci. 69 (1996) 31–32. 13. N. H. G. Panners, L.K. Koopal, and J. Lyklema, and 1. W. Stumm and J. J. Morgan, Aquatic Chemistry, Wiley, New Surfaces 21 (1986) 457–468. York, 1996. 14. M. J. Avena, O. R. Camara, and C. P. de Pauli, Colloids and 2. D. A. Dzombak and F. M. M. Morel, Surface Complexation Surfaces 69 (1993) 217–228. Modeling: Hydrous Ferric Oxide, Wiley, 1990. 3. P. Zarzycki, J. Phys. Chem. C 111 (2007) 7692–7703. 15. E. Kinoshita, F. Ingman, and G. Edwall, Electrochim. Acta 31 4. P. W. Schindler, B. Fürst, R. Dick, and P. U. Wolf, J. Colloid (1986) 29–38. Interface Sci. 55 (1976) 469–475. 16. W. T. Grubb and L. H. King, Anal. Chem. 52 (1980) 270–273. 5. S. Žalac and N. Kallay, J. Colloid Interface Sci. 149 (1992) 17. N. Kallay, Z. Dojnovič, and A. Čop, J. Colloid Interface Sci. 286 233–240. (2005) 610–614. 6. T. Preočanin and N. Kallay, Croat. Chem. Acta 71 (1998) 649. 18. T. Preočanin, A. Čop, and N. Kallay, J. Colloid Interface Sci. 7. A. V. Delgado, Interfacial Electrokinetics and , 299 (2006) 772–776. M. Dekker, Basel, 2002. 19. N. Kallay and T. Preočanin, J. Colloid Interface Sci. 318 (2008) 8. M. Ciobanu, J. P. Wilburn, M. L. Krim, and D. E. Cliffel in: C. 290–295. G. Zoski (Ed.) Handbook of Electrochemistry, Elsevier, 2007. 20. N. Kallay, T. Preočanin, and T. Ivšić, J. Colloid Interface Sci. 9. C. M. A. Brett and A. M. O. Brett, Electrochemistry: Principles, 309 (2007) 21–27. Methods, and Applications, Oxford University Press, Oxford, 21. N. Kallay, T. Preočanin, A. Selmani, F. Šupljika, and I. Leniček 1993. Croat. Chem. Acta 82 (2009) 323–327 10. L. Bousse, N. F. de Rooij, and P. Bergveld, IEEE Trans. Elec- 22. N. Kallay and S. Žalac, J. Colloid Interface Sci. 230 (2000) tron Devices 30 (1983) 1263–1270. 1–11. 11. L. Bousse, N. F. de Rooij, and P. Bergveld, Surf. Science 135 23. R. K. Quinn, R.D. Nasby, and R.J. Baughman, Mater. Res. Bull. (1983) 479–496. 11 (1976) 1011–1017.

Croat. Chem. Acta 83 (2010) 457. P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces 473

24. N. Kallay and D. Čakara, J. Colloid Interface Sci. 232 (2000) mann, E. Gileadi and M. Urbakh (Eds), Encyclopedia of Electro- 81–85. chemistry, Wiley-VCH: New York, 2002; pp. 33–132. 25. N. Kallay, A. Čop, E. Chibowski, and L. Holysz, J. Colloid In- 60. E. Spohr, Computer simulation of the structure and dynamics of terface Sci. 259 (2003) 89–96. water near metal surfaces, in: G. Jerkiewich, M. P. Soriaga, K. 26. N. Kallay, T. Preočanin, and F. Šupljika, J. Colloid Interface Sci. Uosaki, and A. Wieckowski (Eds), Solid-Liquid Electrochemical 327 (2008) 384–387. Interfaces, ACS Symposium Series 656; American Chemical So- 27. T. Preočanin, F. Šupljika, and N. Kallay, J. Colloid Interface Sci. ciety; Washington, DC, 1996; pp. 31–44. 346 (2010) 222–225. 61. M. L. Berkowitz, I.-C. Yeh, and E. Spohr, Structure of water at 28. S. Trasatti (Ed.), Pure Appl. Chem. 58 (1986) 956–966. the water-metal interface: molecular dynamics computer simula- 29. H.-H. Butt, K. Graf, and M. Kappl, Physics and Chemistry of tions, in: A. Wieckowski (Ed.), Interfacial Electrochemistry: Interfaces, Wiley, New York, 2003, chapter 5, pp 63–65. Theory, experiment, and applications, Dekker: New York, 1999, 30. P. Zarzycki and K. M. Rosso, J. Colloid Interface Sci. 341 pp. 33–45. (2010) 143–152. 62. E. Spohr, Electrochim. Acta 44 (1999) 1697–1705. 31. P. Zarzycki, R. Charmas, and P. Szabelski, J. Comput. Chem. 25 63. M. R. Philpott and J. N. Glosi, Molecular Dynamics Simulation (2004) 704–711. of Interfacial Electrochemical Processes: Electrical Double 32. G. Sposito, Environ. Sci. Technol. 32 (1998) 2815–2819. Layer Screening, in: G. Jerkiewich, M. P. Soriaga, K. Uosaki, 33. W. H. van Riemsdijk, J. C. M. de Wit, L. K. Koopal, and G. H. and A. Wieckowski (Eds), Solid-Liquid Electrochemical Inter- Bolt, J. Colloid Interface Sci. 116 (1987) 511–522. faces, ACS Symposium Series 656; American Chemical Society; 34. P. Zarzycki, Appl. Surface Sci. 253 (2007) 7604–7612. Washington, DC, 1996; pp. 12–30. 35. D. E. Yates and S. L. T. W. Healy, J. Chem. Soc., Faraday 64. J. N. Glosi and M. R. Philpott, Electrochim. Acta 41 (1996) Trans. 1 70 (1974) 1807–1818. 2145–2158. 36. D. Chan, J. W. Perram, L. R. White, and T. W. Healy, J. Chem. 65. M. R. Philpott and J. N. Glosi, J. Electroanal. Chem. 409 (1996) Soc., Faraday Trans. 1 71 (1975) 1046 – 1057. 65–72. 37. T. Hiemstra, W. H. Van Riemsdijk, and G.H. Bolt, J. Colloid 66. N. Kallay, T. Preočanin, D. Kovačević, J. Lützenkirchen, and E. Interface Sci. 133(1989) 91–104. Chibowski, Croat. Chem. Acta (2010) in press. 38. T. Hiemstra, J. C. M. de Wit, and W. H. van Riemsdijk, J. Col- 67. W. Janusz, Electrical Double-Layer at Oxide-Solution Interfac- loid Interface Sci. 133 (1989) 105–117. es, in: A. T. Hubbard (Ed.), Encyclopedia of Surface and Colloid 39. D. P. Landau and K. Binder, A guide to Monte Carlo simulations Science, volume II, Marcel Deker, New York, 2000. nd in statistical physics, 2 edition, Cambridge University Press, 68. M.-C. Jodin, F. Gaboriaud, and B. J. Humbert, Colloid Interface Cambridge, 2000. Sci. 287 (2005) 581–591. 40. D. Nicholson and N. G. Personage, Computer Simulation and the 69. S. R. Mehr, D. J. Eatough, L. D. Hansen, and E. A. Lewis, Statistical Mechanics of Adsorption, Academic Press, London Thermochimica Acta 154 (1989) 129–143. 1982. 70. P. Zarzycki, T. Preočanin, K. M. Rosso, and S. Chatman, Science 41. D. W. Heermann, Computer Simulation Methods in Theoretical (2010) submitted. Physics, Springer, Berlin, 1990. 71. G. Bertotti, Hysteresis in Magnetism, Academic Press, San Di- 42. M. Borkovec, Langmuir 13 (1997) 2608–2613. ego, 1998. 43. M. Borkovec, J. Daicic, G. J. M. Koper, Physica A 298 (2001) 1–23. 72. B. Prelot, R. Charmas, P. Zarzycki, F. Thomas, F. Villieras, W. 44. P. Zarzycki, P. Szabelski, and R. Charmas, J. Comput. Chem. 26 Piasecki, and W. Rudzinski, J. Phys. Chem. B 106 (2002) 13280– (2005) 1079–1088. 13286. 45. P. Zarzycki, J. Colloid Interface Sci. 306 (2007) 328–336. 73. R. Charmas, P. Zarzycki, F. Villieras, F. Thomas, B. Prelot, and 46. P. Zarzycki, P. Szabelski, and R. Charmas, Appl. Surface Sci. W. Piasecki, Colloids Surf. A 244 (2004) 9–17. 252 (2005) 752–758. 74. S. V. Yanina and K. M. Rosso, Science 320 (2008) 218–222. 47. P. Zarzycki, J. Colloid Interface Sci. 315 (2007) 54–62. 75. M. A. Blesa, P. J. Morando, and A. E. Regazzoni, Chemical Dis- 48. P. Zarzycki and K.M. Rosso, Langmuir 25 (2009) 6841–6848. solution of Metal Oxides, CRC, Boca Raton,1994. 49. P. Zarzycki, J. Colloid Interface Sci. 311 (2007) 622–627. 76. P. Zarzycki, R. Charmas, and W. Piasecki, Adsorption 10 (2004) 50. P. Zarzycki, Langmuir 22 (2006) 11234–11240. 139–149. 51. J. R. Rustad, A. R. Felmy, and B. P. Hay, Geochim. Cosmochim. 77. W. Janusz, I. Kobal, A. Sworska, and J. Szczypa, J. Colloid Acta 60 (1996) 1553–1562. Interface Sci. 187 (1997) 381–387. 52. J. R. Rustad, A. R. Felmy, and E. J. Bylaska, Geochim. Cosmo- 78. A. Breeuwsma and J. Lyklema, Discuss. Faraday Soc. 52 (1971) chim. Acta 67 (2003) 1001–1016. 324 – 333 53. M. Předota,P. T. Cummings, and D. J. Wesolowski, J. Phys. 79. S. Mustafa, B. Dilera, Z. Neelofer, A. Naeem, and S. Tasleem, J. Chem. C 111, (2007) 3071–3079. Colloid Interface Sci. 204 (1998) 284–293. 54. M. Předota, Z. Zhang, P. Fenter, D.J . Wesolowski, and P.T. 80. N. Kallay, T. Preočanin, and T. Ivšić, J. Colloid Interface Sci. Cummings, J. Phys. Chem. B 108 (2004) 12061–12072. 309 (2007) 21–27. 55. M. Předota, A. V. Bandura, P. T. Cummings, J. D. Kubicki, D. J. 81. V. May and O. Kuhn, Charge and Energy Transfer Dynamics in Wesolowski, A. Chialvo, and M. L. Machesky, J. Phys. Chem. B Molecular Systems, 2nd edition, Wiley, New York, 2004. 108 (2004) 12049–12060. 82. A. Nitzan, Chemical Dynamics in Condensed Physics, Oxford 56. P. Zarzycki, S. Kerisit, and K. M. Rosso, J. Phys. Chem. C 114 University Press, Oxford, 2006. (2010) 8905–8916. 83. P. Zarzycki, W. Piasecki, and K.M. Rosso, in preparation. 57. P. Zarzycki and K. M. Rosso, J. Phys. Chem. C 114 (2010) 84. K. M. Rosso, S. V. Yanina, C. A. Gorski, P. Larese-Casanova, 10019–10026. and M. M. Scherer, Environ. Sci. Technol. 44 (2010) 61–67. 58. E. Spohr, Electrochim. Acta 49 (2003), 23–27. 85. G. King and A. Warshel, J. Chem. Phys. 93 (1990) 8682–8692. 59. A. A. Kornyshev, E. Spohr, and M. A. Vorotyntsev, Electro- 86. P. Zarzycki, S. Kerisit, and K. M. Rosso, in preparation. chemical Interfaces: At the Border Lin, in: A. J. Bard, M. Strat-

Croat. Chem. Acta 83 (2010) 457. 474 P. Zarzycki et al., Electrostatic Potential at Crystal/Electrolyte Interfaces

SAŽETAK Teorija, eksperiment i kompjuterska simulacija elektrostatskog potencijala na međupovršini kristal/elektrolit

Piotr Zarzycki,a Kevin M. Rosso,a Shawn Chatman,a Tajana Preočanin,b Nikola Kallayb i Wojciech Piaseckic

aChemical and Materials Sciences Division, Pacific Northwest National Laboratory, Richland, WA 99354, USA bZavod za fizikalnu kemiju, Kemijski odsjek, Prirodoslovno-matematički fakultet, Sveučilište u Zagrebu, HR 10000 Zagreb, Hrvatska cDepartment of Biochemistry, The Jozef Pilsudski Academy of Physical Education, Warsaw, Poland

U ovom članku razmatrani su napredak i izazovi kod mjerenja, analiziranja i tumačenja nastanka elektrostatskog potencijala na međupovršini kristal/elektrolit. Analizirani su različiti modeli površinskog kompleksiranja, prikaza- na je metoda za procjenu točke nul-potencijala za plohu monokristala uz pomoć histereze do koje dolazi tijekom pH titracije, te porijeklo nelinearne ovisnosti površinskog potencijala o pH. Poznato je da elektrostatski potencijal znatno ovisi o mnogim pojavama poput heterogenosti površine, transformacijama na površini, reakcijama prijeno- sa naboja i dodatnog skoka potencijala na rubovima kristalne plohe i/ili Schottkyjevim barijerama. Potencijal monokristalneelektrode je također osjetljiv na te pojave što otežava tumačenje rezultata mjerenja. Predloženi analitički model titracijske histereze i novi Model površinskog kompleksiranja koji uzima u obzir procese prijeno- sa naboja doprinose tumačenju eksperimentalnih podataka i osiguravaju jednoznačno tumačenje elektrostatskog odziva nepolarizabilnih monokristalnih elektroda.

Croat. Chem. Acta 83 (2010) 457.