Modelling water adsorption on Au(2 1 0) surfaces: II. Monte Carlo simulations

8
This article was published in an Elsevier journal. The attached copy is furnished to the author for non-commercial research and education use, including for instruction at the author’s institution, sharing with colleagues and providing to institution administration. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Transcript of Modelling water adsorption on Au(2 1 0) surfaces: II. Monte Carlo simulations

This article was published in an Elsevier journal. The attached copyis furnished to the author for non-commercial research and

education use, including for instruction at the author’s institution,sharing with colleagues and providing to institution administration.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Author's personal copy

Modelling water adsorption on Au(210) surfaces: II. MonteCarlo simulations

Rodrigo S. Neves a, Artur J. Motheo a, Rui P.S. Fartaria b,Fernando M.S. Silva Fernandes b,*

a Laboratory of Interfacial Electrochemistry, Department of Physical Chemistry, Institute of Chemistry of Sao Carlos, University of Sao Paulo,

Avenida do Trabalhador Sancarlense, CP 780, 13560-970 Sao Carlos-SP, Brazilb Molecular Simulation Group, CCMM, Department of Chemistry and Biochemistry, Faculty of Science, University of Lisboa Campo Grande,

Bloco C8, 1749-016 Lisboa, Portugal

Received 8 November 2006; received in revised form 4 September 2007; accepted 19 September 2007Available online 6 October 2007

Abstract

Canonical Monte Carlo simulations for the Au(210)/H2O interface, using a force field recently proposed by us, are reported. Theresults exhibit the main features normally observed in simulations of water molecules in contact with different noble metal surfaces.The calculations also assess the influence of the surface topography on the structural aspects of the adsorbed water and on the distribu-tion of the water molecules in the direction normal to the metal surface plane. The adsorption process is preferential at sites in the firstlayer of the metal. The analysis of the density profiles and dipole moment distributions points to two predominant orientations. Most ofthe molecules are adsorbed with the molecular plane parallel to surface, while others adsorb with one of the O–H bonds parallel to thesurface and the other bond pointing towards the bulk liquid phase. There is also evidence of hydrogen bond formation between the firstand second solvent layers at the interface.� 2007 Elsevier B.V. All rights reserved.

Keywords: Monte Carlo simulation; Au(210); Water adsorption; Bilayers

1. Introduction

In the last decades, molecular simulations, namelyMonte Carlo (MC) and molecular dynamics (MD), havebecome a powerful tool in the branch of interfacial electro-chemistry. Several works on the subject have contributedwith important information about the molecular behaviourat electrochemical interfaces.

Simulations of water in contact with noble metal single-crystal surfaces have probed the distribution of the watermolecules in the interfacial region and its dependence onthe nature and structure of the surface [1–8]. These studieshave systematically demonstrated the preferential adsorp-

tion of the molecules on specific sites of the surface, show-ing that the structure of the adsorbed water layer stronglydepends on the surface topography.

Simulations of ions [7,9–13] and of organic molecules[14–17] adsorption on electrode surfaces, have also showna similar preferential site adsorption behaviour. Addition-ally, calculations of the potential of mean force (PMF)[13,15,16,18,19] have enabled the identification of the sol-vent contributions to the ionic and molecular adsorptionsat different sites of the surface and, in the case of organicmolecules, have led to a better understanding of the molec-ular reorientations along the adsorption process.

The first step for the realization of this kind of simula-tions is the construction of consistent force fields to modelthe water-surface interactions, which account for the influ-ence of the structural properties of the surface on the sol-vent layers, and on the preferential sites for the

0022-0728/$ - see front matter � 2007 Elsevier B.V. All rights reserved.

doi:10.1016/j.jelechem.2007.09.032

* Corresponding author. Tel.: +351 21 7500133; fax: +351 21 7500088.E-mail address: [email protected] (F.M.S. Silva Fernandes).

www.elsevier.com/locate/jelechem

Available online at www.sciencedirect.com

Journal of Electroanalytical Chemistry 612 (2008) 179–185

Journal ofElectroanalytical

Chemistry

Author's personal copy

adsorption. In a previous paper we have proposed ananalytical potential function to describe the interactionbetween water molecules and the Au(210) surface [20],based on DFT calculations. In the present article, wereport canonical Monte Carlo simulations of water adsorp-tion on Au(2 10) surfaces, at potential of zero charge (pzc),using the referred to force field.

The interest in simulating water on Au(2 10) surfaces isconcerned with some experimental features. This surfacedoes not undergo the charge induced reconstruction phe-nomena, observed in lower index gold single-crystal sur-faces [21]. It is therefore, in experimental terms, a stablesurface for adsorption studies in the total range of the elec-trical double layer potential region. As a consequence, theAu(2 10) surface is a suitable choice, e.g., for the study ofdissipative processes accompanying adsorption phenomena[22] and reorientations in the adsorbed solvent layer[23,24], since the reconstruction process itself is also asource of dissipation and may couple with adsorption dis-sipation, as in the case of halides adsorption onto Au(11 1)[25].

Once the scope of dissipation associated to phase transi-tions in the interface is still an open subject, molecular sim-ulations of the Au(2 10)/water interface regarding theadsorption process can contribute to a better understand-ing of experimental results obtained by two of us [22–25].The other motivation of the present paper is just to testthe consistency of the results obtained by using the forcefield for that interface recently proposed by us [20]. Indeed,the test of interaction potentials is one of the goals of com-puter simulation.

Section 2 contains the methodology and models used inthe Monte Carlo simulations. Section 3 discusses the simu-lation results. Section 4 presents the concluding remarks ofthe work.

2. Methodology and models

The canonical Monte Carlo simulations were performedon a system composed by 1000 molecules of water encapsu-lated between two Au(21 0) surfaces. The simulation boxwas rectangular in shape, with the dimensions of22.43 · 26.40 · 50.47 A in the directions x, y and z, respec-tively. Standard periodic boundary conditions (PBC) wereapplied on the x and y directions. The dimensions of thebox in the directions x and y were chosen in order to ensurethe periodicity of the crystallographic surface under thePBC. The total volume of the box reproduces the densityof water at temperature and pressure of 298.25 K and1 bar, respectively.

Equilibration runs of 5000 Monte Carlo cycles were fol-lowed by production runs of 25,000 cycles. One MC cycle iscomposed of random translations and rotations applied toeach water molecule of the sample, with an acceptance rateof �50%. The computational program used the time savingenergy-repository algorithm, recently discussed by Fartariaet al. [26].

The TIP4P [27] model for the interaction between watermolecules was used, considering its success in the simula-tion of some water–metal interfaces [15,18,19,28,29]. It isa rigid model describing the water molecule by four inter-action centres, three of them coinciding with the moleculeatoms and the fourth related to the electronic density asso-ciated to the pair of electrons of the oxygen atom. This lastcentre is located at 0.15 A from the oxygen, between thetwo hydrogen atoms. The calculations of the interactionsbetween the water molecules were conducted with a cut-off radius of 8.0 A and the usual long-range correctionsfor the dispersion terms of the potential.

The long-range corrections for the electrostatic compo-nent of the water potential were calculated by the Ewaldsum method adapted to a two-dimensional periodic system[30,31] in the x and y directions.

The interactions between water molecules and theAu(210) surfaces were worked out using the force field pre-sented in our previous work [20], with a cut-off radius of25 A.

3. Results and discussion

The normalised density profiles for the distribution ofthe water molecule atoms as a function of the distance fromthe surface, considering the average value over both sur-faces of the model, are shown on Fig. 1.

The results clearly show the formation of two welldefined water layers on the metallic surface. The first layercorresponds to the region from �1.2 to �3.8 A. In thisregion, the location of the two most intense peaks, forthe O and H atoms, at nearly the same distance of�2.5 A, shows that most of the water molecules are, onaverage, adsorbed with the molecular plane parallel tothe electrode surface. The peaks are narrow, indicating thatthe molecules have well defined positions and orientations.However, the occurrence of the small peak a, at �3.3 Afrom the surface, suggests that some molecules have a dif-ferent orientation with one or the two hydrogen atoms

2.5 5.0 7.5 10.0 12.5

1

2

3

4

5

(b)(a)

II

I

ρ n

r / Å

Fig. 1. Normalised density profiles for water molecules at the interface.Oxygen (——) and hydrogen (—-—). The inset shows a conformationconsistent with hydrogen bond formation.

180 R.S. Neves et al. / Journal of Electroanalytical Chemistry 612 (2008) 179–185

Author's personal copy

pointing to the second adsorption layer. The comparisonbetween the areas of the profiles related to the differentatoms in region I, indicates that �80% of the water mole-cules present the molecular plane parallel to the surface.This behaviour is observed in other simulations of waterin contact with metallic surfaces [5,6].

The second solvent layer corresponds to the region from�3.8 to �6.9 A, denoted as region II. The less intense andbroader peaks of this region, when compared with region I,show that this is a more diffuse and less organized layerwith all the atoms presenting a higher degree of mobility,therefore implying a broader distribution for the molecularorientations.

The structure of this layer presents the same overall ori-entations as in region I though not in the same degree, asindicated by the displacement of the peak referent to thehydrogen atoms distribution, in relation to the peak corre-sponding to the oxygen atoms.

The peak b, observed on region II at �4.3 A from thesurface, shows that some hydrogen atoms, located at thesecond solvent layer, are pointing towards the first solventlayer. It is seen that the distances from each of the hydro-gen peaks, a and b, to the oxygen peaks in regions II and I,respectively, is �1.9 A. The proximity of this value to thetypical bond length of 1.97 A for hydrogen bonds, stronglysuggests the existence of such bonds between the first andsecond adsorption layers. The inset in Fig. 1 shows anexample of a conformation consistent with hydrogen bondformation.

For distances greater than 12.5 A from the surface, thedensity profiles of both atoms converge to unity, meaningthat the influence of the metal is negligible for thosedistances.

Further information on the structure of the solvent lay-ers, in contact with the surface, has been obtained by calcu-lating the dipole moment and H–H vectors distributions atdifferent distances from the metal. The distribution of thecosines of the angles, between those vectors and the axisnormal to the surface, was calculated for three consecutivesolvent laminae measured from the metallic surface, withthickness of 4 A, which is, approximately, the thicknessof the adsorption layer.

The distributions of the dipole moment vectors orienta-tions are shown in Fig. 2. In the first solvent lamina,between 0 and 4 A from the surface (Fig. 2a), a great partof the dipole moments are approximately parallel to thesurface, as shown by the broad peak at �0 (region I).Moreover, a second well-defined peak is observed at�0.75 (region II). These results are consistent with the pre-vious analysis of the density profiles. That is, most of themolecules (�80%) are adsorbed with the molecular planeapproximately parallel to the surface, while other mole-cules adsorb with one of the O–H bonds approximatelyparallel to the surface, and the other pointing towardsthe bulk liquid phase, with the molecular plane at �90� rel-atively to the surface. In this case, the hydrogen atompointing outwards the electrode is at �3.4 A from the sur-

face, in accordance with peak a on Fig. 1, at �3.3 A. Thisbehaviour was also observed in MC simulations of wateron Cu(100) [32].

Such structural properties of the first adsorbed waterlayer resembles the features of water bilayers commonlyformed in the adsorption of water onto closed packednoble metals surfaces [33–35]. In the bilayers, the moleculescloser to the electrode are adsorbed via the electronic pairof the oxygen atom, whilst the molecules more distant fromthe electrode, but inside the lamina, are adsorbed viahydrogen bonds, with one hydrogen atom pointingtowards the solution side of the interface (H-up bilayer),or towards the surface (H-down bilayer). The resultantstructure of the bilayer formation is similar to the ice likestructure of water, with the formation of approximatelyhexagonal six member rings.

The present results for the water adsorption onAu(21 0), at pzc, suggest the formation a H-up bilayer inthe interfacial region comprehended by the solvent laminabetween 0 and 4 A from the electrode.

In the solvent lamina between 4 and 8 A from the sur-face, shown in Fig. 2b, it is noticeable a weak tendencyfor preferential orientation of the dipole moment vectorsat ��0.75 (region III), which implies that the hydrogenatoms of the water molecules are pointing towards the bulkliquid phase, consistently with the density profiles. Also, inthis layer it is possible to see a weak wave at �0.6 (regionIV), which is related to peak b of Fig. 1. This is a furtherindication of hydrogen bonds formation between the firstand second solvent layers.

0.2

0.4

0.6

0.8

-0.75 -0.50 -0.25 0.00 0.25 0.50 0.75

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8

p[co

s(θ)

]p[

cos(

θ)]

p[co

s(θ)

]cos (θ)

(c) 8 < z < 12 Å

(b) 4 < z < 8 Å

(a) 0 < z < 4 Å

IVIII

I II

Fig. 2. Distribution of the cosine of the angle (h) between the dipolemoment vector of the water molecules and the axis normal to the surfacefor three solvent lamina parallel to the surface. (a) First solvent layer, (b)second solvent layer and (c) region with weak influence from the metal.

R.S. Neves et al. / Journal of Electroanalytical Chemistry 612 (2008) 179–185 181

Author's personal copy

For the solvent lamina between 8 and 12 A, shown inFig. 2c, no specific tendency is detected for the dipolemoment vector distribution and, as expected, the liquidapproaches bulk behaviour.

The distributions of the H–H vectors orientations areshown in Fig. 3. For the first solvent lamina (Fig. 3a), agreat part of the H–H vectors are approximately parallelto the surface (region I, cos/ � 0), consistently with theprevious information about the parallel orientation of theadsorbed water molecules. The small peak observed (atcos/ � 0.6) supports, as suggested before, the adsorptionof water molecules with one O–H bond parallel to the sur-face and the other pointing to the liquid side of theinterface.

The solvent lamina between 4 and 8 A (Fig. 3b) shows amuch weaker tendency for preferential orientation of theH–H vectors, although resembling the distribution for thefirst lamina. This is consistent with hydrogen bonds forma-tion between the first and second solvent layers.

In the solvent lamina between 8 and 12 A (see Fig. 3c)no specific tendency is observed for the H–H vector distri-bution, as noted in the case of the dipole moment vectors.

The structural behaviour of the water adsorbed layerleads to a particular distribution of charges in the interfa-cial region, producing a characteristic electric field profile.

The charge density profile, calculated from the oxygenand hydrogen density profiles shown in Fig. 1, can beobserved in Fig. 4a. As expected, strong fluctuations inthe charge density are observed at distances closer than

6 A from the electrode. The order of magnitude of thecharge fluctuations at the first (at �2.5 A) and second (at�5 A) solvent layers is similar to the observed for thewater/platinum interfaces in the absence of external fields[36].

The dependence of the electric field on the distance fromthe metallic surface (E(z)) can be calculated from the den-sity profile by

EðzÞ ¼ 1

e0

Z z

0

qCðziÞdzi; ð1Þ

where qc(zi) is the charge density at a distance (zi) from thesurface and e0 is the vacuum permittivity. It is representedon Fig 4b.

As expected from the charge density profiles, typicalfield fluctuations are present in the transition from the firstto the second solvent layers. The magnitude and shape ofthe profile are similar to the ones observed for water in con-tact with platinum surfaces [36]. Another feature noted inRef. [36], and also observed in the results presented here,is the fact that the electrical field is not zero at the centreof the simulation box. Instead, it reaches a constant valueslightly different from 0. In those studies, the authors dem-onstrated that this feature is not related to the size of thesystem. In metal/solution interface simulations where ionicspecies are present, and consequently the drop potential atthe interface is not exclusively related to solvent dipoles,such behaviour is not observed or is too small relativelyto fluctuations [7,37]. Therefore, it seems to be related tolong-range relaxation of water dipoles at the interface inthe absence of ionic species.

The equilibrium between the water–metal and water–water interactions gives rise to particular molecular orien-tations at the interface. It also affects the molecular energydistributions, which are represented in Fig. 5 for the threedifferent solvent laminae at the interfacial region and thecentre of the simulation box.

0.5

1.0

1.5

2.0

2.5

3.0

0.5

1.0

1.5

2.0

2.5

3.0

0.25 0.50 0.75

0.5

1.0

1.5

2.0

2.5

3.0

I

(b) 4 < z < 8 Å

(a) 0 < z < 4 Å

II

p[co

s(φ)

]p[

cos(

φ)]

p[co

s(φ)

]

(c) 8 < z < 12 Å

|cos (φ)|

Fig. 3. Distribution of the jcos(/)j, where / is the angle between the H–Hvector of the water molecules and the axis normal to the surface: (a) firstsolvent layer, (b) second solvent layer and (c) region with weak influencefrom the metal.

-1

0

1

2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0 22.5

-6

-4

-2

0

2

ρ c / 1

09 C m

-3

(b)

(a)

z / Å

E /

109 V

m-1

Fig. 4. Charge density obtained from the density profiles (a) and electricfield obtained from the charge density profile (b).

182 R.S. Neves et al. / Journal of Electroanalytical Chemistry 612 (2008) 179–185

Author's personal copy

For the first solvent lamina the molecular energies aredistributed at ��109 kJ mol�1, while at the centre of thebox are distributed at ��84 kJ mol�1, which is consistentwith the value determined for the model by Jorgensenet al. in simulations of pure water [27,38]. The averagemolecular energy for the molecules in the second and thirdlaminae is near to the values at the centre of the simulationbox, indicating the low influence of the metal at those dis-tances from the surface.

Another important sub product of the simulations is theoxygen surface density distributions of the water moleculesin the first solvent layer (Fig. 6).

The surface density distribution over the total and unitcell surfaces (in Fig. 6a and b, respectively) shows that arelatively high number of oxygen atoms are found in thevicinities, or over, the gold T sites (top sites, see Ref.[20]). However, a significant number of oxygen atoms arealso observed over the H1 and H2 sites (hollow1 and hol-low2 sites, see Ref. [20]). It is suggested that such distribu-tions are a consequence of: (i) the water–Au(210)interaction mainly drives the molecules to top sites, whichfor this surface are set apart by �4–5 A; (ii) the O. . .O dis-tance in the most favourable hydrogen bond conformationis �3 A. These effects compete actively against each other,giving rise to a very noticeable distribution over sites otherthan the top ones. This behaviour presumably is not sopronounced in more closed packed surfaces, e.g. Au(111), where the top sites are set apart by �3 A. In suchcases (i) and (ii) should cooperate instead.

Other important aspect about the structure of theadsorbed water layer comes from the comparison betweenthe presented results and those from molecular simulationand DFT calculations of water bilayers on closed packedtransition metal surfaces [5,6,34,39]. In closed packed sur-faces, it is common to observe very regular hexagonalstructures propagating parallel to the surface. However,in the case of the Au(2 10) surface, the observed hexago-nal centred structures are irregular, as in the exampleshighlighted in Fig. 7, which give a qualitative view ofthe first solvent layer. This is coherent with the fact thatwater bilayers tend to adapt to the surface lattice in orderto retain the ice like structure [33]. In some cases, like theones in this work, that fact implies the compression of thebilayer, and its structure resembles the one of a singlelayer with two kind of adsorbed water molecules. Thisfeature arises from the equilibrium between the water–water and water–metal interactions, which stronglydepends on the surface topography and hydrophilicity[40,41].

-150 -125 -100 -75 -50 -25

0.000

0.005

0.010

0.015

0.020

0.025

0.030

p

E / kJ mol -1

Fig. 5. Molecular configurational energy distributions in the three solventlamina parallel to the surface and on the centre of the simulation box.0 < z < 4 A (n), 4 < z < 8 A (s), 8 < z < 12 A (m), centre of the simulationbox (e).

Fig. 6. Oxygen surface density distribution in the first solvent layer: (a) total surface and (b) surface unit cell.

R.S. Neves et al. / Journal of Electroanalytical Chemistry 612 (2008) 179–185 183

Author's personal copy

Most of the water molecules are adsorbed with themolecular plane parallel to the surface. However, some ofthem have one of the O–H bonds parallel to the surfaceand the molecular plane approximately perpendicular toit, as can be observed in the snapshots shown in Fig. 7.These molecules are those that participate in hydrogenbonds between the first and second layers, associated withpeak a in Fig. 1, and their distribution over the surface canbe visualized in the hydrogen surface density distributionspresented in Fig. 8. The analysis of Fig. 8a and b showsthat the water molecules which participate on hydrogenbonds between first and second solvent layers adsorb pref-erentially over T sites.

The adsorption of water molecules in these configura-tions is supported by experimental evidence [42,43]. Thereorganisation of the first solvent layer in order to formhydrogen bonds with the second one, was also observedin MC simulations of water adsorbed on Cu(100) [32].

4. Conclusions

We have reported canonical Monte Carlo simulationsfor the Au(2 10)/water interface, using a recently proposedforce field for modelling the Au(21 0)/water interaction.

The results show that water adsorbs trough the oxygenatom, preferentially over Au atoms of the first layer (topsites, T). The density profiles, the dipole moment and H–H vectors distributions also show the occurrence of twopredominant orientations for water molecules adsorbedat the metallic surface. Most of the molecules are adsorbedwith the molecular plane parallel to the surface. Some ofthem, however, are adsorbed with an angle of �42�between the dipole moment vector and the axis normal tothe metal plane, with the molecular plane in an orientationapproximately perpendicular to the surface. The moleculesadsorbed in the last orientation are found almost all overor in the vicinities of the sites T, and participate in the for-

Fig. 7. Snapshots of the water molecules at the Au(210) surface. The darkest oxygen atoms are of water molecules interacting with the first solvent layervia hydrogen bonds. Hexagonal centred structures are highlighted.

Fig. 8. Hydrogen surface density distribution of the hydrogen atoms in the first solvent layer, which form hydrogen bonds with the second solvent layer:(a) total surface and (b) surface unit cell.

184 R.S. Neves et al. / Journal of Electroanalytical Chemistry 612 (2008) 179–185

Author's personal copy

mation of hydrogen bonds between the first and the secondsolvent layers.

The developed potential function reproduces the mainfeatures found in molecular simulations of water adsorp-tion on various metallic surfaces. However, differently fromlower crystallographic index surfaces, the mobility of theadsorbed water layer seems to be more pronounced inthe case of the Au(2 10) surface.

The MC simulations, carried out with the proposedinteraction model, show consistent results, indicating thatthe model can be used for the study of adsorption processesof different molecules, in aqueous media, on Au(2 10) sur-faces, including the determination of potentials of meanforce (PMF). The formation of hydrogen bonds betweenthe adsorbed layers, predicted by the interaction model,certainly plays an important role on the adsorption phe-nomena. The simulations have also assessed the structureof the system water/Au(210) in the interfacial region.The molecular level picture of the interface may contributeto a better understanding of the dissipative behaviour, nor-mally associated to phase transitions on adsorbed layers,and of the adsorption process itself. Further work alongthese lines is in progress and will be reported soon.

Acknowledgements

F. Fernandes and R. Fartaria are grateful to Fundacaopara a Ciencia e Tecnologia (FCT) within the project POC-TI/QUI/36457/2000. R. Fartaria also gratefully acknowl-edges the support of FCT within the Ph.D. grant SFRH/BD/10405/2002.

R.S. Neves and A.J. Motheo are grateful to CAPES(BEX0532/00-2), CNPq (140250/00-0) and FAPESP (06/04122).

References

[1] G. Nagy, K. Heinzinger, J. Electroanal. Chem. 296 (1990) 549.[2] K. Heinzinger, Pure Appl. Chem. 63 (1991) 1733.[3] K. Raghavan, K. Foster, K. Motakabbir, M. Berkowitz, J. Chem.

Phys. 94 (1991) 2110.[4] X. Xia, L. Perera, U. Essmann, M.L. Berkowitz, Surf. Sci. 335 (1995)

401.[5] I.C. Yeh, M.L. Berkowitz, J. Electroanal. Chem. 450 (1998) 313.[6] E. Spohr, in: R.C. Alkire, D.M. Kolb (Eds.), Advances in Electro-

chemical Science and Engineering, vol. 6, Wiley-VHC, Nova Iorque,1999.

[7] E. Spohr, Solid State Ionics 150 (2002) 1.[8] E. Spohr, Electrochim. Acta 49 (2003) 23.[9] J.N. Glosli, M.R. Philpot, J. Chem. Phys. 96 (1992) 6962.

[10] J.N. Glosli, M.R. Philpot, J. Chem. Phys. 98 (1993) 9995.[11] D.A. Rose, I. Benjamin, J. Chem. Phys. 98 (1993) 2283.[12] M.R. Philpot, J.N. Glosli, S.-B. Zhu, Surf. Sci. 335 (1995) 422.[13] A. Ignaczak, J.A.N.F. Gomes, S. Romanowski, J. Electroanal. Chem.

450 (1998) 175.[14] M.R. Philpot, J.N. Glosli, Chem. Phys. 198 (1995) 53.[15] R.S. Neves, A.J. Motheo, F.M.S.S. Fernandes, R.P.S. Fartaria, J.

Braz. Chem. Soc. 15 (2004) 224.[16] R.P.S. Fartaria, F.F.M. Freitas, F.M.S.S. Fernandes, J. Electroanal.

Chem. 574 (2005) 321.[17] U. Reimer, M. Wahab, P. Schiller, H.-J. Mogel, Langmuir 21 (2005)

1640.[18] E. Spohr, Acta Chem. Scand. 49 (1995) 189.[19] J.A.N.F. Gomes, A. Ignaczak, J. Mol. Struct. (THEOCHEM) 463

(1999) 113.[20] R.S. Neves, A.J. Motheo, R.P.S. Fartaria, F.M.S.S. Fernandes, J.

Electroanal. Chem., in press, doi:10.1016/j.jelechem.2007.06.020.[21] D.M. Kolb, Prog. Surf. Sci. 51 (1996) 109.[22] R.S. Neves, E. De Robertis, A.J. Motheo, Electrochim. Acta 51

(2006) 1215.[23] A.J. Motheo, A. Sadkowski, R.S. Neves, J. Electroanal. Chem. 430

(1997) 253.[24] A. Sadkowski, A.J. Motheo, R.S. Neves, J. Electroanal. Chem. 455

(1998) 107.[25] R.S. Neves, E. De Robertis, A.J. Motheo, Appl. Surf. Sci. 253 (2006)

1379.[26] R.P.S. Fartaria, R.S. Neves, P.C.R. Rodrigues, F.F.M. Freitas,

F.M.S.S. Fernandes, Comput. Phys. Commun. 175 (2006) 116.[27] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, J. Chem. Phys. 79

(1983) 926.[28] J. Bocker, R.R. Nazmutdinov, E. Spohr, K. Heinzinger, Surf. Sci. 335

(1995) 372.[29] E. Spohr, G. Toth, K. Heinzinger, Electrochim. Acta 41 (1996)

2131.[30] E. Spohr, J. Chem. Phys. 107 (1997) 6342.[31] I.-C. Yeh, M.L. Berkiwitz, J. Chem. Phys. 111 (1999) 3155.[32] A. Ignaczak, in: Ph.D. thesis, Department of Chemistry, University of

Porto, Portugal, 1997, p. 186.[33] M.A. Henderson, Surf. Sci. Rep. 46 (2002) 1.[34] C.D. Taylor, M. Neurock, Curr. Opin. Solid State Mater. Sci. 9

(2005) 49.[35] J.-S. Filhol, M. Neurock, Angew. Chem. Int. Ed. 45 (2006) 402.[36] I.C. Yeh, M.L. Berkowitz, J. Chem. Phys. 110 (1999) 7935.[37] E. Spohr, Electrochim. Acta 44 (1999) 1697.[38] M.W. Mahoney, W.L. Jorgensen, J. Chem. Phys. 112 (2000) 8910.[39] A. Roudgar, A. Grob, Chem. Phys. Lett. 409 (2005) 157.[40] S. Meng, E.G. Wang, S. Gao, J. Chem. Phys. 119 (2003) 7617.[41] S. Meng, E.G. Wang, S. Gao, Phys. Rev. B 69 (2004) 195404.[42] J.D. Porter, A.S. Zinn, J. Phys. Chem. 97 (1993) 1190.[43] K.G. Lloyd, B.A. Banse, J.C. Hemminger, Phys. Rev. B 33 (1986)

2858.

R.S. Neves et al. / Journal of Electroanalytical Chemistry 612 (2008) 179–185 185