a reactive force field (reaxff) for zinc oxide · uncorrected proof 1 2 a reactive force field...

12
UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a , Adri C.T. van Duin b , Micael Baudin a , Kersti Hermansson a, * 4 a Department of Materials Chemistry, The A ˚ ngstro ¨ m Laboratory, Box 538, 752 21 Uppsala, Sweden 5 b Materials and Process Simulation Center (MSC), California Institute of Technology, Pasadena, CA 91125, USA 6 Received 28 August 2007; accepted for publication 13 December 2007 7 8 Abstract 9 We have developed a reactive force field (FF) within the ReaxFF framework, for use in molecular dynamics (MD) simulations to 10 investigate structures and reaction dynamics for ZnO catalysts. The force field parameters were fitted to a training set of data points 11 (energies, geometries, charges) derived from quantum-mechanical (QM) calculations. The data points were chosen to give adequate 12 descriptions of the equations of state for a number of zinc metal and zinc oxide phases, a number of low-index ZnO surfaces and 13 gas-phase zinc hydroxide clusters. Special emphasis was put on obtaining a good surface description. We have applied the force field 14 to the calculation of atomic vibrational mean square amplitudes for bulk wurtzite–ZnO at 20 K, 300 K and 600 K and we find good 15 agreement with experimental values from the literature. The force field was also applied in a study of the surface growth mechanism 16 for the wurtzite(0 0 0 1) surface. We find that the growth behavior depends on the presence of surface steps. 17 Ó 2007 Elsevier B.V. All rights reserved. 18 Keywords: Zinc oxide; Construction and use of effective interatomic interactions; Molecular dynamics; Density-functional calculations; Surface relaxation 19 and reconstruction; Surface energy 20 21 1. Introduction 22 The aim of this study is to create a high-level, transfer- 23 able interaction model for the condensed phases of zinc 24 oxide. ZnO is a highly interesting material with a broad 25 range of technical applications. It is known to catalyze a 26 number of chemical reactions, such as the decomposition 27 of formic acid and the formation of methanol from CO 28 and H 2 [1–3]. Recent studies of ZnO nanostructures imply 29 a connection between catalytic activity and surface mor- 30 phology [1]. ZnO also attracts interest because of its elec- 31 tronic properties, with a semi-conducting band-gap of 32 3.4 eV. It has potential applications in opto-electronic de- 33 vices, directly or as a substrate for the growth of other 34 semiconductors such as GaN and SiC; here the interest in 35 ZnO is partly fueled by the availability of high-quality sub- 36 strates [4]. 37 Because of this large versatility of ZnO, it is not surpris- 38 ing that a very large number of experimental and theoreti- 39 cal studies of ZnO exist in the literature. Thus, for the 40 various polymorphs of bulk ZnO only, numerous theoret- 41 ical studies have been published, using models based either 42 on quantum-mechanical (QM) methods (ab initio theory or 43 density-functional theory (DFT)) or analytical force fields 44 (FF). Some recent examples of such QM studies can be 45 found in Refs. [5–7] (and references therein). In Ref. [5], 46 Catti et al. studied the piezo-electric properties of ZnO 47 using Hartree–Fock calculations, and in Refs. [6,7], the 48 optimized geometry and electronic structures as well as 49 the effect of pressure on the vibrations in several ZnO poly- 50 morphs were studied using DFT lattice dynamics. 51 The development of DFT techniques, or hybrid techniques 52 such as the B3LYP functional [8,9], have resulted in methods 53 that are significantly more efficient than ab initio methods of 54 comparable accuracy. However, no DFT method available 55 today is fast enough to describe chemical and physical prop- 56 erties accurately in large, dynamic systems with a few thou- 57 sand atoms. Many systems with defects of various kinds 0039-6028/$ - see front matter Ó 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.susc.2007.12.023 * Corresponding author. Tel.: +46 18 47 137 25; fax: +46 18 513 548. Q1 E-mail address: [email protected] (K. Hermansson). www.elsevier.com/locate/susc Available online at www.sciencedirect.com Surface Science xxx (2008) xxx–xxx SUSC 17820 No. of Pages 12, Model 5+ 11 January 2008 Disk Used ARTICLE IN PRESS Please cite this article in press as: D. Raymand et al., Surf. Sci. (2008), doi:10.1016/j.susc.2007.12.023

Upload: hatuyen

Post on 13-Sep-2018

227 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

1

2

3

45

67

8

9

10

11

12

13

14

15

16

17

181920

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

Q1

Available online at www.sciencedirect.com

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

www.elsevier.com/locate/susc

Surface Science xxx (2008) xxx–xxx

OF

A reactive force field (ReaxFF) for zinc oxide

David Raymand a, Adri C.T. van Duin b, Micael Baudin a, Kersti Hermansson a,*

a Department of Materials Chemistry, The Angstrom Laboratory, Box 538, 752 21 Uppsala, Swedenb Materials and Process Simulation Center (MSC), California Institute of Technology, Pasadena, CA 91125, USA

Received 28 August 2007; accepted for publication 13 December 2007

TED

PR

O

Abstract

We have developed a reactive force field (FF) within the ReaxFF framework, for use in molecular dynamics (MD) simulations toinvestigate structures and reaction dynamics for ZnO catalysts. The force field parameters were fitted to a training set of data points(energies, geometries, charges) derived from quantum-mechanical (QM) calculations. The data points were chosen to give adequatedescriptions of the equations of state for a number of zinc metal and zinc oxide phases, a number of low-index ZnO surfaces andgas-phase zinc hydroxide clusters. Special emphasis was put on obtaining a good surface description. We have applied the force fieldto the calculation of atomic vibrational mean square amplitudes for bulk wurtzite–ZnO at 20 K, 300 K and 600 K and we find goodagreement with experimental values from the literature. The force field was also applied in a study of the surface growth mechanismfor the wurtzite(0001) surface. We find that the growth behavior depends on the presence of surface steps.� 2007 Elsevier B.V. All rights reserved.

Keywords: Zinc oxide; Construction and use of effective interatomic interactions; Molecular dynamics; Density-functional calculations; Surface relaxationand reconstruction; Surface energy

C

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

UN

CO

RR

E1. Introduction

The aim of this study is to create a high-level, transfer-able interaction model for the condensed phases of zincoxide. ZnO is a highly interesting material with a broadrange of technical applications. It is known to catalyze anumber of chemical reactions, such as the decompositionof formic acid and the formation of methanol from COand H2 [1–3]. Recent studies of ZnO nanostructures implya connection between catalytic activity and surface mor-phology [1]. ZnO also attracts interest because of its elec-tronic properties, with a semi-conducting band-gap of3.4 eV. It has potential applications in opto-electronic de-vices, directly or as a substrate for the growth of othersemiconductors such as GaN and SiC; here the interest inZnO is partly fueled by the availability of high-quality sub-strates [4].

54

55

56

57

0039-6028/$ - see front matter � 2007 Elsevier B.V. All rights reserved.

doi:10.1016/j.susc.2007.12.023

* Corresponding author. Tel.: +46 18 47 137 25; fax: +46 18 513 548.E-mail address: [email protected] (K. Hermansson).

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

Because of this large versatility of ZnO, it is not surpris-ing that a very large number of experimental and theoreti-cal studies of ZnO exist in the literature. Thus, for thevarious polymorphs of bulk ZnO only, numerous theoret-ical studies have been published, using models based eitheron quantum-mechanical (QM) methods (ab initio theory ordensity-functional theory (DFT)) or analytical force fields(FF). Some recent examples of such QM studies can befound in Refs. [5–7] (and references therein). In Ref. [5],Catti et al. studied the piezo-electric properties of ZnOusing Hartree–Fock calculations, and in Refs. [6,7], theoptimized geometry and electronic structures as well asthe effect of pressure on the vibrations in several ZnO poly-morphs were studied using DFT lattice dynamics.

The development of DFT techniques, or hybrid techniquessuch as the B3LYP functional [8,9], have resulted in methodsthat are significantly more efficient than ab initio methods ofcomparable accuracy. However, no DFT method availabletoday is fast enough to describe chemical and physical prop-erties accurately in large, dynamic systems with a few thou-sand atoms. Many systems with defects of various kinds

08), doi:10.1016/j.susc.2007.12.023

Page 2: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

T

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

2 D. Raymand et al. / Surface Science xxx (2008) xxx–xxx

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

UN

CO

RR

EC

and broken symmetries would ideally require such large sys-tems. Therefore force field based methods are necessary.

ZnO is challenging to model by means of force fieldmethods because of its partly covalent, partly ionic charac-ter. Traditionally, metal oxides have been modeled success-fully using pair interactions consisting of a short-range partand long-range Coulombic terms, the latter usually mod-eled by fixing the charges of the different ions at their for-mal values. Such a procedure of course limits the possibilityto describe a charge redistribution around a defect or at asurface. By adding an extra (charged) site to each ion con-nected via a spring, polarization can be handled; this is theidea behind the shell-model, [10] which can thus capturesome of the many-body interactions, even in an otherwisepair-additive model. The short-range interactions in theshell-model are usually described by an analytical expres-sion such as the Buckingham interaction model. Severalsuccessful attempts at using these schemes for ZnO canbe found in Refs. [11–14]. In Ref. [11], a shell-model poten-tial was used to predict the spatial arrangement of the dop-ant species sodium, lithium, and chlorine within the zincoxide lattice in varistors using molecular dynamics (MD)simulations. The shell-model of Ref. [11] was also appliedin the MD study of elastic properties of zinc oxide ‘‘nano-belts” in Ref. [14]. Whitmore et al. [12] used lattice dynam-ics calculations to parametrize a shell-model reproducingbulk vibrational properties of ZnO. This model has subse-quently been applied to embed QM-clusters [13] in theinvestigation of electron trapping at the oxygen terminatedpolar (0001) surface of ZnO and related surface F centers.

The pure two-body potentials as well as the various shellmodels generally have difficulties in describing the covalentcharacter of ZnO properly, as is evident from the fact thatthe four-coordinated wurtzite does not come out as thelowest-energy crystal structure in Refs. [11,12]. Instead,these models tend to favor the more isotropic, four-coordi-nated zincblende (sphalerite) structure. The description ofcovalent bonding can be aided by the use of explicit high-er-order terms in the many-body expansion. One suchexample of the addition of a three-body interaction termcan be found in the MD study of the growth mechanismfor the ZnO(0 00 1) surface in Ref. [15]. Another way tohandle the covalent character of ZnO has been to neglectthe ionic character altogether and use a bond-order poten-

tial, i.e. a potential where the short-range many-body ef-fects depend on the local environment of each atom (thenumber of neighbors and the distances to them). This hasrecently been done in a successful way by Erhart et al.[16]. Such an approach might, however, have difficultieswith modeling interfaces, or in other cases where Coulom-bic fields might be important. Moreover, if only nearestneighbors are considered, the model cannot describe theenergy difference between the two four-coordinated ZnOphases. Incidentally, Ref. [16] gives an excellent review ofdifferent analytical potential schemes used for ZnO in theliterature. In the present study, we will use the reactiveforce field (ReaxFF) to model ZnO.

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

OO

F

The ReaxFF model [17–19] is also a bond-order interac-tion model. It is bond-order consistence for the two-body,three-body and four-body short-range interaction terms.The model also includes Coulombic terms, where thecharges are calculated based on connectivity and geometryusing the Electronegativity Equalization Method (EEM)[20], thereby allowing redistribution of charges. In princi-ple these features allow the ReaxFF model to describemetallic, covalent and ionic bonds. The ReaxFF model isable to simulate the breaking and reforming phenomenaof bonds during dynamics. It can also reproduce the struc-ture and mechanical properties of condensed phases [17–19]. The ReaxFF framework was initially developed forhydrocarbons [17,18], and has since then been applied tomany such systems. It has also been employed in the studyof several oxide systems, namely MoOx;VxOy and BixOy

[21] oxides and Si=SiO2 [18] and Al=Al2O3 [19] oxideinterfaces.

All force field parameters within the ReaxFF frame-work are developed using results of quantum-chemical cal-culations. During the parametrization, measures are takento provide for transferability, e.g. the parameters describ-ing the oxygen–oxygen interaction are the same in allcompounds and molecules. Therefore the set of datapoints to which the model is fitted needs to contain infor-mation about a wide range of compounds. In the currentstudy this means having data not only for ZnO(s) and itssurfaces, but also for Zn(s) and Zn and O containing gas-phase species. The ZnO-parameters may thus in principlebe used not only for the metal oxide but also for the me-tal/oxide interface and for systems where Zn or ZnO inter-act/react with hydrogen, hydrocarbons, and otherinteresting molecules (provided that a parametrizationfor the molecule is available and that the ZnO moleculecross-terms can be made).

However, parameterizing a model for a broad range ofchemical environments might necessitate trade-offs. To en-sure a good description of the most important parts of thedata set, a weighting scheme is employed during theparametrization. In our case describing ZnO(s) and itssurfaces were prioritized over the description of gas-phasehydroxide clusters. If future studies require greater accu-racy for the gas-phase species, the weighting scheme couldbe redesigned to allow a closer reproduction of these datapoints.

The current parametrization of ZnO in the ReaxFFmodel consisted of the following steps: (i) The creation ofan initial training set of data points (energies, chargesand geometries) from QM calculations, (ii) selection ofthe appropriate number of terms in the ReaxFF energyexpression, (iii) fitting the parameters to the training set,(iv) evaluating if the data set is satisfactorily reproducedor if energy expression needs to be expanded with addi-tional terms which may in turn require the training set tobe expanded with additional data points, (v) repeating(i)–(iv) until the fit is satisfactory, and (vi) validation ofthe potential parameters by comparing properties extracted

08), doi:10.1016/j.susc.2007.12.023

Page 3: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

T

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

208208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

233233

D. Raymand et al. / Surface Science xxx (2008) xxx–xxx 3

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

C

from the model to experimental and quantum-chemicaldata from other studies. The goal of the present study isto create a good description of ZnO within the ReaxFFframework of transferable potential parameters.

The layout of the paper is as follows. Section 2 describesthe methods used, and gives about the ReaxFF model(Section 2.1), the force field optimization procedure (Sec-tion 2.2) and the quantum-mechanical calculations usedto generate the data set (Section 2.3). Section 3 is the re-sults and discussion section where we present the modeland make an assessment of its quality by discussing themodel’s ability to reproduce the relative energies for thedifferent crystalline phases and for certain low-indexsurfaces. In Section 4, we apply the new model in MD sim-ulations to calculate the vibrational mean square ampli-tudes or displacements of the atoms in bulk wurtziteZnO and to study wurtzite ZnO(0 001) surface growth.In the growth simulations we use a strategy as similar aspossible to the published force field study of Kubo et al.[15] and compare with their results. The paper ends witha summary section.

2. Method

Zinc oxide is a transition metal oxide with the wurtzitetype structure (space group P63mc; number 186 in the Inter-

national tables for crystallography [22]) with lattice con-stants 3.24992(5) and 5.20658(8) A at ambient conditions[23]. It has a bulk modulus of 140 GPa [24]. Another threepolymorphs are stable and are listed in Table 1 togetherwith experimental cell parameters and bulk moduli, wheresuch data exist. Calculated data in the table will be dis-cussed in Sections 2.3 (B3LYP) and 3 (ReaxFF).

UN

CO

RR

E

Table 1QM results and ReaxFF results (at 0 K) compared to experimental data from(Df H), bulk moduli and elastic constants of the four polymorphs (B1–B4) of Z

Structure Property B3L

Wurtzite(B4) a/A 3.28P63mc c/A 5.28

Df HB4/(kcal/mol)Bulk modulus/GPa 136c11/GPac12/GPac13/GPac33/GPac44/GPa

Zincblende(B1) Bulk modulus/GPa 162F �43m a/A 4.60

Df HB1 � Df HB4/(kcal/mol) 0.50

Rocksalt(B3) Bulk modulus/GPa 202Fm�3m a/A 4.30

Df HB3 � Df HB4/(kcal/mol) 8.57

Caesium chloride(B2) Bulk modulus/GPa 183Pm�3m a/A 2.68

Df HB2 � Df HB4/(kcal/mol) 37.8

a Explicitly entered in the dataset.

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

OO

F

2.1. The reactive force field

The total interaction energy expression of the ReaxFF ispartitioned into several energy terms as given in Eq. (1).

Esystem ¼ Ebond þ Eover þ Eunder þ Elp þ Eval þ Epen

þ Etors þ Econj þ EvdWaals þ ECoulomb: ð1Þ

These partial contributions include bond energies ðEbondÞ,under-coordination penalty energies ðEunderÞ, lone-pairsenergies ðElpÞ, over-coordination penalty energies ðEoverÞ,valence angles energies ðEvalÞ, energy penalty for handlingatoms with two double bonds ðEpenÞ, torsion angles ener-gies ðEtorsÞ, conjugated bonds energies ðEconjÞ and termsto handle non-bonded interactions, namely van der WaalsðEvdWaalsÞ and Coulomb ðECoulombÞ interactions. All termsexcept the last two are bond-order dependent, i.e. will con-tribute more or less depending on the local environment ofeach atom.

In the expression above, EvdWaals and ECoulomb are calcu-lated between every atom pair, irrespective of connectivityand include a shielding parameter to avoid excessive repul-sion at short distances. This treatment of non-bondedinteractions aids the ReaxFF model in describing covalent,metallic, ionic, and intermediate materials, thus, greatlyenhancing its transferability. A detailed description of theindividual terms can be found in Refs. [17,18]. However,in the present study, not all of the terms in Eq. (1) wereconsidered necessary and some were therefore set to zero,reducing the energy expression to Eq. (2).

Esystem ¼ Ebond þ Eover þ Eval þ EvdWaals þ ECoulomb: ð2Þ

the literature (at room temperature) for the cell axes, heat of formationnO

YP ReaxFF Experiment

a 3.29 3.25 [23]a 5.30 5.21 [23]

�91.2 �83.3a [37]144 141, 143 [24,44]222.9 209.7 [45]116.3 121.1 [45]103.5 105.1 [45]212.8 210.9 [45]57.1 42.47 [45]

130 Polymorph does not occur in naturea 4.62a 0.97

283 203,228 [44,46]a 4.44 4.27 [46]a 7.72

407 Polymorph does not occur in naturea 2.649a 37.58

08), doi:10.1016/j.susc.2007.12.023

Page 4: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

T

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

4 D. Raymand et al. / Surface Science xxx (2008) xxx–xxx

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

UN

CO

RR

EC

2.2. Force field optimization procedure

The ReaxFF was optimized using a successive one-parameter search technique as described by van Duinet al. [25]. Based on QM calculations for Zn(s), ZnO(s)and Zn hydroxide clusters, ReaxFF parameters were gener-ated for Zn–O and Zn–Zn bond energies and for Zn–O–Zn, O–Zn–O, O–Zn–Zn and Zn–O–H valence angleenergies.

The current parametrization of ZnO in the ReaxFFmodel consisted of the following steps, as outlined in theintroduction (the same roman numerals are used).

(i) An initial training set of QM data points was cre-ated for clusters and crystals. For the crystals,data for the energy dependence on volume of fourZn metal phases and four ZnO polymorphs wereincluded: Wurtzite(4), zincblende(4), rocksalt(6)and caesium chloride(8) for ZnO and fcc(12),hcp(12), bcc(8) and sc(6) metal phases (coordina-tion number in parentheses). Mulliken charges fothe ZnO bulk systems were added to the dataset.A bond dissociation profile of one Zn–OH bondin a ZnðOHÞ2-molecule was included. Data forthe energy dependence on valence angles inZnðOHÞ2 and OðZnOHÞ2 was added to includevalence angle variation in clusters. The ReaxFFparameters were fitted to the lowest-energy state(singlet or triplet depending on geometry).

(ii) A minimum number of terms in the ReaxFFenergy expression were selected (starting withEbond;Eover;EvdWaals;ECoulomb).

(iii) The parameters were fitted to the whole trainingset using the built-in one-parameter routine.

(iv)–(v) The valence angle term ðEvalÞ was added to thetotal energy expression to obtain a good fit tothe training set. Surface energies and geometriesfor certain low-index surfaces of three of theZnO(s) phases were added to improve the descrip-tion of ZnO surfaces. These surfaces were wurtziteð10�10Þ and ð11�20Þ, zincblende (100), rocksalt(110) and (100). To further expand the data set,Mulliken charges from the surface calculationswere added.

(vi) The potential parameters were validated by com-paring properties extracted from the model toexperimental and quantum-chemical data; see Sec-tion 3.

In total 44 parameters were fitted to a dataset containing347 data points.

2.3. The quantum-mechanical calculations and the systems

studied

The quantum-mechanical calculations were performedwith the GAUSSIAN03 program [26] for the hydroxide

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

clusters and the CRYSTAL03 program [27] for the ZnOcrystals. The calculations for Zinc metal were performedwith the SeqQuest software package [28].

ED

PR

OO

F

2.3.1. System descriptions

QM calculations were performed for the four crystalpolymorphs of the wurtzite, zincblende, rocksalt and cae-sium chloride structures (the structures are also referredto as h-ZnS, c-ZnS, NaCl and CsCl, respectively). Here3D periodic calculations were performed using the CRYS-TAL03 program [27], which makes use of space group sym-metry. In all cases only the asymmetric unit was included.For a range of volumes the cell parameters of the four crys-tal structures were optimized within the given space groupsin such a way that the minimum energy for a given volumewas obtained (e.g. the c=a-ratio was optimized for thewurtzite structure). Potential energy curves were calculatedfor use in the training set for the ReaxFF parametrization.The bulk moduli were calculated using the Murnaghanequation; see Ref. [24] a for detailed description of the pro-cedure for ZnO. The bulk moduli were used for ReaxFFmodel validation purposes (and were not included in thetraining set). The elastic constants of the wurtzite phasewere calculated following the procedure outlined in Ref.[29].

Five ZnO surface systems were studied, namely ð10�10Þand ð11�20Þ for wurtzite, (110) for zincblende and (110)and (100) for the rocksalt structure. Each surface systemwas modeled by an isolated slab, periodic in two dimen-sions (x and y), with free surfaces in the third (+z and�z). In all cases, 10-layer slabs were generated by thebuilt-in routine of CRYSTAL03, thus preserving symmetryoperations of the space group where possible. The numberof layers was chosen based on the slab-thickness investiga-tion for wurtzite by Wander and Harrison [30]. The peri-odic cell axes of the slabs were kept fixed at therespective optimized bulk values (see Table 1). The atomicpositions were optimized for all slabs (with fixed cellparameters).

The surface energy (defined as ðEsurf � nEbulkÞ=2Asurf ,where A is the surface area and n is the number of ZnO-for-mula units in the slab), the surface relaxation (defined asthe atomic displacements in z-coordinate) and the surfacerumpling (defined as the difference in relaxation betweenthe cations and the anions in one layer, positive rumplingmeaning that the cations displace more towards the centerof the slab compared to the anions) were calculated andused in the training set.

Equations of state were calculated for four polymorphsof zinc metal: hexagonally close packed (hcp), face centeredcubic (fcc), body centered cubic (bcc) and simple cubic (sc).In all cases the full crystallographic unit cell was included.Here 3D periodic calculations were performed using theSeqQuest program [28]. The cell parameters of the threecrystal structures were optimized within the given spacegroups, following the same procedure as for the ZnO crys-

08), doi:10.1016/j.susc.2007.12.023

Page 5: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

D. Raymand et al. / Surface Science xxx (2008) xxx–xxx 5

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

tals. Potential energy curves were calculated for use in thetraining set for the ReaxFF parametrization.

Geometry optimizations and potential energy curve cal-culations were performed for the two zinc hydroxide clus-ters, ZnðOHÞ2 and OðZnOHÞ2, using the GAUSSIAN03program [26].

357

358

359

360

361

2.3.2. Computational details

All calculations for the ZnO crystals and surfaces wereperformed using 3- and 2-dimensionally periodic systems

UN

CO

RR

EC

T

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396Q3

Table 2QM results and ReaxFF results (at 0 K) compared to experimental datafrom the literature (at room temperature) for the cell axes, cohesiveenergy, bulk moduli and elastic constants of Zn metal

Structure Property PBE ReaxFF Experiment

hcp a/A 2.63a 2.73b 2.67 [47]P63=mmc c/A 5.06a 4.46b 4.95 [47]

Ecohesive(hcp)/(kcal/mol)

�31.6 �31.1a [36]

Bulk modulus/GPa 66.7 87.7 64.5–75.1c

c11/GPa 193 163c

c12/GPa 41 31c

c13/GPa 42 48c

c33/GPa 188d 60c

c44/GPa 71d 39c

fcc Bulk modulus/GPa 81.8 87.7 Polymorph does notoccur in nature

F �43m a/A 3.86a 3.86Ecohesive(fcc)-Ecohesive(hcp)/(kcal/mol)

1.31 0.00b

bcc Bulk modulus/GPa 84.6 73.5 Polymorph does notoccur in nature

Im�3m a/A 3.06a 3.06Ecohesive(bcc)-Ecohesive(hcp)/(kcal/mol)

2.71 2.71

sc Bulk modulus/GPa 64.2 30.2 Polymorph does notoccur in nature

Pm�3m a/A 2.71a 2.71Ecohesive(sc)-Ecohesive(hcp)/(kcal/mol)

6.00 6.62

a Explicitly entered in the dataset.b Valence angle terms for pure Zn were not included, therefore ReaxFF

will preserve the ideal c/a ratio ðð8=3Þ1=2Þ of a close packed crystal. Alsothe volume energy relationship for the fcc and hcp phases will be identical.

c Experimental values taken from the review article by Ledbetter [48].d These hcp-specific stresses are not correctly reproduced for the reason

stated in b.

Table 3The final atom parameters for Zn

r0 Coulomb parameters

g (eV) v (eV) c (A) po

Zn 1.9014 5.6773 0.5149 0.8009 �O 1.2477 8.9989 8.5000 1.0503 �For O, parameters from Ref. [49] was used. Definitions of the individual Reax

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

OO

F

(as mentioned) and the hybrid DFT functional B3LYP.Gaussian type basis sets were used for both the zinc atomand the oxygen atom, as optimized by Jaffe and Hess[24]. This basis set has been applied with good results inseveral earlier QM-studies of ZnO polymorphs and sur-faces [31,30,32]. In the 3-dimensionally periodic calcula-tions, the Brillouin zone was sampled with a 10� 10� 10k-point sampling mesh of the Monkhorst–Pack type [33].In the 2-dimensionally periodic calculations, the Brillouinzone was sampled with 4� 4 k-point sampling mesh ofthe Monkhorst–Pack type. Table 1 gives a comparison ofthe optimized QM results with experiment for ZnO.

The calculations for the zinc metal were performed using3-dimensionally periodic systems and the DFT functionalPBE [34]. A library SeqQuest atom file for Zn was used,describing a gaussian type basis set with double zeta qualityfor s and d plus polarization for p. The atom file includes apseudo potential generated using the Fritz–Haber Institutefhi98PP code [35]. However, the wave function coefficientsin the SeqQuest atom file were optimized to get good Zncell parameters and energies. The Brillouin zone was sam-pled with a 8� 8� 8 k-point sampling mesh. Table 2 givesa comparison of the optimized QM results with experimentfor Zn. The experimental cohesive energy for Zn was usedin the training set [36]. The zinc metal data and ZnO crystaldata were combined in the training set by putting the ener-gies on a heat of formation scale, using the experimentalheat of formation [37].

For the clusters, potential energy curves were calculatedas functions of the HOZn–OH bond distance and the HO–Zn–OH, Zn–O–H and (HOZn)–O–(ZnOH) valence angles,using the hybrid DFT functional B3LYP [8,9] and the 6-311 + G* basis set [38–40] as implemented in the GAUSS-IAN03 program.

3. Results and discussion

Our final fitted ReaxFF is given in Tables 3–6. The po-tential form was given schematically in Eq. (2). Unlessotherwise stated, all ReaxFF results in the following text,tables or figures refer to our final optimized ReaxFF, alsowhen we discuss selected cuts through the multi-dimen-sional potential energy surface (PESs), such as bond disso-ciation curves, etc., below. The quantities included in thefinal FF fit were listed in Section 2.2. In the figures pre-sented in the following, selected cuts through, or pointson, the final ReaxFF PESs will be compared with the cor-responding cuts from the B3LYP PESs (see Table 7).

van der Waals parameters

v=un rvdW (A) � (kcal/mol) a cvdW (A)

2.5000 1.9272 0.3000 11.4526 12.51073.6141 1.9236 0.0904 10.2127 7.7719

FF parameters in this table and Tables 4–6 can be found in Refs. [17,18].

08), doi:10.1016/j.susc.2007.12.023

Page 6: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

REC

T

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

Table 4The final van der Waals and bond radius parameters for the Zn–O bond

Bond rr (A) rvdW (A) � (kcal/mol) cvdW (A)

Zn–O 2.0933 2.1496 0.2992 9.7788

Table 5The final bond energy and bond-order parameters for the Zn–Zn and Zn–O bonds

Dre /(kcal/mol) pbe;1 pbe;2 pbo;1 pbo;2 pkov

Zn–Zn 47.2436 �0.3583 2.7689 0.0962 6.3816 0.0000Zn–O 124.8335 �0.8979 1.1391 �0.3406 6.7964 0.0000O–O 60.1436 �0.2802 0.2441 �0.1302 6.2919 0.9114

The O–O parameters from Ref. [49] were used.

Table 6The final valence angle parameters

Valenceangle

H0;0

(degree)ka (kcal/mol)

kb

ð1=radÞ2pv;1 pv;2

H–O–Zn 50.7176 11.0788 0.3108 0.9582 1.1056O–Zn–O 9.4094 26.4927 0.3247 2.0000 3.1031Zn–O–Zn 50.1486 6.7868 0.5367 0.1384 1.1000O–Zn–Zn 64.3179 6.5137 0.2554 0.5219 1.7398

Table 7QM results versus ReaxFF model results for surface energies of the fiveZnO surfaces present in the ReaxFF training set

B3LYP ReaxFFsurface energy/Jm�2 surface energy/Jm�2

Wurtziteð10�10Þ 1.32 0.96Wurtziteð11�20Þ 1.39 1.06Zincblende(100) 1.30 1.00Rocksalt(100) 1.16 0.73Rocksalt(110) 2.00 1.10

Zincblende

Caesium Chloride

Wurtzite

Rocksalt

ZincblendeCaesium Chloride

Wurtzite

-90

-80

-70

-60

-50

14 16 18 20 22 24 26 28 30

-80

-70

-60

-50

-40

Rocksalt

ΔΗf /

(kca

l/mol

)

B3LYP

ReaxFF

Volume (Å3/formula unit)

Equations of state for ZnO(s) phases

Fig. 1. Equation of state (compression and tension) for four crystalstructures (wurtzite(B4), zincblende(B3), caesium chloride(B2) and rock-salt(B1)) of ZnO calculated using the (a) B3LYP and (b) ReaxFFmethods.

6 D. Raymand et al. / Surface Science xxx (2008) xxx–xxx

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

UN

CO

R3.1. Equations of state for the Zn and ZnO crystals

As mentioned, the relative stabilities of four Zn crystalstructures (fcc, hcp, bcc and sc types) and four ZnO crystalstructures (wurtzite, zincblende, rocksalt and caesium chlo-ride) were used as data in the ReaxFF training set. The var-iation of coordination numbers and coordination figuresrepresented in these crystals give important informationconcerning how bond energies and valence angles energiesrelate to bond order. The heats of formation and cohesiveenergies for the various crystals were used in the trainingset and give information about bond strength as a functionof bond order. In this study, we find that the inclusion ofthe valence angle term has a large influence on the relativephase stability of the crystals. The cell axis lengths of thedifferent crystals give important information about thepositions of the minima of the bond dissociation profiles.No information about the elastic properties of the crystalswas entered in the data set other than the volume-energyrelationship (equations of state). Therefore calculating

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

OO

F

elastic constants and bulk moduli of the crystal structuresmay serve to evaluate the parametrization, which has beendone here and is presented below. During the force fieldoptimization, the description of the crystal phases was gi-ven priority over the clusters, since the crystals were con-sidered more important for future studies (see below).

Fig. 1 shows that the four ZnO polymorphs are repro-duced in the correct relative order of stability by the Rea-xFF model, and also the curve shapes agree with the QMdata. It is known from the literature that ReaxFF has a sys-tematic tendency to overestimate the stability of metaloxide phases; [21] this is also the case here. Table 1 givesa comparison between bulk moduli and cell axes of the var-ious polymorphs of ZnO obtained with the B3LYP meth-od, the ReaxFF model and experiment, respectively. Wefind that the ReaxFF model gives a good description ofthe elastic properties of wurtzite structured ZnO(s) com-pared to experimental values. This helps validate the qual-ity of the current parametrization since information aboutthe elastic properties of the crystals was only implicitly en-tered. A good description of the low-energy wurtzite phasewas prioritized in our fitting; this is visible from the overes-timation of the bulk moduli of the high-energy phases.

Fig. 2 shows the equations of state for zinc metal. Va-lence angle terms for pure Zn metal–metal interactionswere not included; therefore the ReaxFF model will pre-serve the ideal c/a ratio ((8/3)1/2) of a close-packed crystal(i.e. there will be no energy difference between hcp and fcc).This also means that the hcp-specific elastic constants, i.e.c33 and c44; corresponding to stresses in the c-direction ofthe hexagonal lattice, may not be correctly described. How-ever, these valence angle terms can be added to the model

08), doi:10.1016/j.susc.2007.12.023

Page 7: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

T448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

8 9 10 11 12 13 14 15 16 17 18

Volume (ų/atom)

Coh

esiv

e en

ergy

(kc

al/m

ol)

sc

bcccp

sc

bcccp

-20

0

20

40

-20

0

20

40 ReaxFF

DFT

Equations of state for Zn(s) phases

Fig. 2. Equation of state (compression and tension) for three crystalstructures (simple cubic (sc), body centered cubic (bcc) and close-packed(cp)) of Zn calculated using the (a) PBE and (b) ReaxFF methods.

D. Raymand et al. / Surface Science xxx (2008) xxx–xxx 7

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

at a later time if one wants to study properties relying onthe elastic properties of hcp-Zinc. In the current study thisfeature has not been prioritized. Table 2 gives a compari-son between bulk moduli and cell axes of the various poly-morphs of Zn metal calculated with the PBE method andthe ReaxFF model and from experiment.

488

489

490

491

492

493

REC3.2. ZnO surfaces – energies and geometries

The surface structures of five different ZnO slabs were fit-ted. The correct relative order of the surface energies isimportant for studies of e.g. crystal growth, where high sur-face-energy facets tend to grow the fastest. Only non-polarsurfaces were used in the training set because of the inherent

UN

CO

R

Dip

lace

men

t rel

ativ

e bu

lk p

ositi

ons

Δ(Å

)

0 1 2 3 4 5 6

O2-

0 1 2 3 4 5 6 7 0

-0.3

-0.2

-0.1

0

0.1

-0.3

-0.2

-0.1

0

0.1

0.2Zn 2+ Zn 2+

O2-

Distance from c 1 2 3 4

Zn2+

O 2-

B3LYPReaxFF

c-ZnS

Relaxationh-ZnS(1010) - h-ZnS(1120) -

Fig. 3. Atomic displacements (in A) from the bulk-terminated positions in Zcalculated with the B3LYP and ReaxFF methods.

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

OO

F

difficulties of making a good QM-model of polar surfaces.Table 6 gives a summary of surface energies. Reproducingthe correct rumpling trends is important for a properdescription of surface chemistry, and in our ReaxFF fitting,efforts were therefore made to reproduce the rumplingtrends observed in the B3LYP calculations. Several inter-io-nic distances in the slabs were added explicitly to the train-ing set, to put extra weight on the surface geometry.

Fig. 3 shows that ReaxFF tends to underestimate thesurface relaxation (this underestimation is also visible inthe surface energies) but reproduces the relative magni-tudes of the relaxation of the different surface orientationscompared to the QM-data.

3.3. Atomic charges in ZnO bulk and slabs

Since the EEM-method uses geometry and connectivityto calculate the magnitude of the ionic charges it is impor-tant to include such QM data in the training set. Thereforethe charges of the ions in the ZnO bulk crystals and slabswere included in the training set. The coordination of theions changes from the surface to the bulk of the slab andso do the charges. Here Mulliken charges were used to re-main consistent with previous parameterizations in theReaxFF model. Atomic charges in the four bulk crystalstructures and the five slabs were fitted (see Table 6 andFig. 4). In the slabs, the ReaxFF charge dependence onz-coordinate is seen to be slightly exaggerated comparedto the B3LYP results. Likewise in the bulk ZnO crystals,the charge difference between ions with different coordina-tion numbers is seen to be slightly exaggerated. Overall, theReaxFF gives a good representation of the partial atomiccharges of zinc and oxygen.

3.4. Potential energy curves for zinc hydroxide clusters

In line with the ReaxFF strategy of parametrization, wehave also included data for selected zinc and oxygen con-

enter of slab (Å) 5 6 7 8

(110)

1 2 3 4 5 6 0 1 2 3 4 5 6 7 8 9 0

Zn 2+ Zn 2+

O 2- O2-

NaCl(110) NaCl(100)

nO slabs as a function of depth (going from the middle to the surface),

08), doi:10.1016/j.susc.2007.12.023

Page 8: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

8 D. Raymand et al. / Surface Science xxx (2008) xxx–xxx

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

taining gas-phase clusters. To include cluster data for theZn–O bond, a dissociation profile was determined fromQM calculations for the ZnðOHÞ2 cluster. The ground statewas determined through full geometry optimization of thesinglet state; this optimized structure was later used as a ref-erence. Subsequently, the total potential energy dissociationprofile was constructed through calculations at modifiedgeometries by changing the bond length, allowing otherstructural parameters to relax, to avoid influence of otherstraining forces (possibly different in the QM and ReaxFFcases, respectively) on the potential energy curve. The disso-ciation curve was calculated for both the singlet and tripletstates. Finally, the total potential energy of the fully opti-mized reference structure was subtracted to obtain the inter-action energy profile. The QM and ReaxFF curves areshown in Fig. 5a. To the best of our knowledge experimentaldata such as bond lengths, bond strengths and valence an-gles for these gas-phase species do not exist, and are there-fore not presented in a table for comparison (See Fig. 6).

To include cluster data for the valence angles, two differ-ent clusters were used: ZnðOHÞ2 and OðZnOHÞ2. Explicit

UN

CO

RR

EC

T

B3L

YP

Rea

xFF

1.059

1.049

1.040

1.072

0.895 0.897

1.073

1.044

1.051

1.062

-0.917

-1.045

-1.052

-1.054

-1.053

-1.055

-1.054

-1.046

-1.052

-0.919

1.051

1.064

1.044

1.074

0.894 0.896

1.071

1.055

1.060

1.041

-0.919

-1.047

-1.053

-1.049

-1.055

-1.053

-1.053

-1.045

-1.051

-0.917-0.933 0.884

-1.049 1.048

-1.053 1.054

-1.038

1.045

1.103

-1.058

1.102-1.039

1.042-1.059

1.058-1.050

-1.051

0.880

1.047

-0.933

-0.909

-1.026

-1.009

0.946

1.026

-1.025

-1.021

0.988

1.024

1.007

-1.021

-1.025

1.007

-1.009

-1.026

1.024

-0.909

0.988

1.026

0.946

1.025

1.026

1.019

1.006

0.935 0.935

1.006

1.019

1.025

1.026

-0.916

-1.018

-1.025

-1.026

-1.026

-0.916

-1.018

-1.026

-1.026

-1.025

1.025

1.026

1.019

1.006

0.935

1.006

1.019

1.025

1.026

0.935-0.916

-1.018

-1.025

-1.026

-1.026

-0.916

-1.018

-1.026

-1.026

-1.025

h-ZnS(1120)h-ZnS(1010)

0.882

1.038

1.049

1.044

1.056

-1.034

-0.901

-1.043

-1.042

-1.042

0.924

1.007

1.014

0.987

1.014

-1.002

-1.014

-1.014

-1.012

-0.904

Atomic charges in ZnO slabs

Fig. 4. Atomic charges as a function of depth in the various ZnO slabs

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

RO

OF

cluster data for three out of the four different valence an-gles (H–O–Zn, O–Zn–O and Zn–O–Zn included) wasadded. Cluster data for the O–Zn–Zn angle was not addedbecause of the difficulty in getting good quality QM-datafor a small cluster with a O–Zn–Zn bond sequence. Follow-ing the same procedure as for the bond dissociation profile,the different clusters were relaxed through geometry opti-mization, creating two reference states. Afterwards theinteresting valence angle (see Fig. 5b–d) was locked inincrements, while the other structural parameters wereoptimized, again to avoid other straining forces in the clus-ter. This procedure was repeated for each of the two va-lence angles in the ZnðOHÞ2 cluster. The resultingpotential energy curves are shown in Fig. 5b–d.

In Fig. 5a, it is visible that the crystals in general wereprioritized over the clusters since the agreement betweenthe ReaxFF and the B3LYP minima for ZnðOHÞ2 inFig. 5a is not perfect. In fact, the ReaxFF minimum isshifted towards 2 A, which is the equilibrium bond lengthin the wurtzite and zincblende crystals. Also the depth ofthe ReaxFF potential well in Fig. 5a is seen to be too shal-

ED

PNaCl(100) NaCl(110)

1.084

1.066

-1.043

-1.051 1.134

0.810

-1.060

-1.053

1.011

-0.898

-0.898

1.010

-1.054

-1.060

0.810

1.135 -1.051

-1.043

1.066

1.085

-0.900

-1.062

-1.079

-1.083

-1.078

0.940

1.084

1.082

1.074

1.023

1.023

1.074

1.082

1.084

0.940

-1.078

-1.083

-1.079

-1.062

-0.900

c-ZnS(110)

1.057

1.042

1.048

1.035

0.880

-0.902

-1.046

-1.044

-1.041

-1.034

-1.083

-1.084

-1.083

-0.986

-1.083

1.060

1.082

1.083

1.086

1.007

1.007

-1.083

-1.083

-1.083

-1.084

-0.986

1.086

1.083

1.082

1.060

1.064

1.067

1.034

1.074

0.947

-0.960

-1.058

-1.049

-1.060

-1.060

1.073

1.065

1.066

1.032

0.951

-0.961

-1.050

-1.061

-1.060

-1.055

0.924

0.987

1.014

1.014

1.007

-1.012

-1.014

-1.014

-0.904

-1.002

. Mulliken charges from B3LYP and EEM-charges from ReaxFF.

08), doi:10.1016/j.susc.2007.12.023

Page 9: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

OO

F

536

537

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

ReaxFF

Albertsson et al.

Schultz et al.

Kihara et al.

0

0.005

0.01

0.015

0.02

0

0.005

0.01

0.015

0.02

0.025

0 200 400 600 800 0 200 400 600 800

Zn

T

c axis Zn || c axis

O || c axis

T

c axisO

T/K

Vib

ratio

nal m

ean

squa

re a

mpl

itude

s <

u >

/ Å

22

ReaxFF

ReaxFF ReaxFF

Albertsson et al.

Albertsson et al. Albertsson et al.

Schultz et al.Schultz et al.

Schultz et al.

Kihara et al.

Kihara et al. Kihara et al.

Atomic Vibrations in wurtzite ZnO(s)

Fig. 6. Atomic vibrational mean square amplitudes for wurtzite ZnO(s).ReaxFF and experimental values from neutron and X-ray diffraction.

D. Raymand et al. / Surface Science xxx (2008) xxx–xxx 9

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

low, to reduce the overestimation of the heat of formationfor ZnO(s) (cf. Fig. 1).

3.5. Errors in the fitted force field

We would like to add a comment concerning the errorsin our fitted force field parameters. It is not straightforwardto try to assess the quality of the fit by quoting an averageor maximum error for the quantities used as observables inthe fit. One reason for this is that weighted data points wereused in the fitting procedure, and the weighted error willdepend on the weighting scheme. The weighting schemewas designed to make sure that the features of the quan-tum-mechanical data that we considered to be the mostimportant were best reproduced by our fitted FF. Amongsuch features are, as mentioned, the curvature of the vol-ume–energy relationship close to the equilibrium forWurtzite ZnO, and the order of stability of the variousZnO phases and surfaces. Less emphasis was deliberatelyput on the gas-phase species, and that is where we considerthat the force field has the largest errors.

During the course of the fitting we have used several ini-tial values for the parameters and several weightingschemes. The ‘‘final parameters” represent a local mini-mum for the final weighting scheme and produces theZnO properties presented in the paper.

UN

CO

RR

EC

T

rZnO

(Å)

ZnO

HO

H

-100

-50

0

50

100

150

0 1 2 3 4 5 6

Ene

rgy

(kca

l/mol

)

rZnO

(Å)

B3LYP - triplet

B3LYP - singlet

ReaxFF

Bond dissociation

0

2

4

6

8

10

12

14

90 100 110 120 130 140 150 160 170

Ene

rgy

(kca

l/mol

)

(degree)Θ

Θ

Zn Zn

O

O

OH

H

B3LYP

ReaxFF

Angle distortion

a

c

Fig. 5. B3LYP and ReaxFF potential energy curves for: (a) dissociation of a ZZnðOHÞ2 molecule, (c) angle distortion of Zn–O–Zn in the OðZnOHÞ2 molecu

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

DPR4. Application: MD simulation of bulk and surfaces of

wurtzite ZnO

In this section we test our new ZnO force field by apply-ing it in two different types of simulation, namely one non-reactive and one reactive. We calculate vibrational meansquare amplitudes for the atoms in bulk ZnO and compare

E

0

10

20

30

40

50

60

70

60 80 100 120 140 160 180

Ene

rgy

(kca

l/mol

)

(degree)Θ

Θ

Zn

O

OH

H

B3LYP

ReaxFF

Angle distortion

0

5

10

15

20

25

30

35

60 80 100 120 140 160

Ene

rgy

(kca

l/mol

)

(degree)Θ

ΘO

ZnH

H

O

B3LYP

ReaxFF

Angle distortion

b

d

n–O bond in the ZnðOHÞ2 molecule, (b) angle distortion of O–Zn–O in thele, (d) angle distortion of Zn–O–H in the ZnðOHÞ2 molecule.

08), doi:10.1016/j.susc.2007.12.023

Page 10: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

O

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

Table 8Mulliken charges (B3LYP) and EEM charges (ReaxFF) for bulk ZnO (seetext for details)

B3LYP ReaxFF

Wurtzite(B4) Zn 1.0260 1.0464O �1.0260 �1.0464

Zincblende(B1) Zn 1.0130 1.0402O �1.0130 �1.0402

Rocksalt(B3) Zn 1.0950 1.0504O �1.0950 �1.0504

CsCl(B2) Zn 0.9780 0.9187O �0.9780 �0.9187

10 D. Raymand et al. / Surface Science xxx (2008) xxx–xxx

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

them to experimental values taken from neutron and X-raydiffraction, to evaluate the ability of ReaxFF to reproducebulk vibrational properties as a function of temperature.Moreover, we apply it to study crystal growth on thewurtzite ZnO(0 001) surface, following the procedure ofKubo et al. [15].

T597

598

599

600

601

602

603

604

605

606

607

4.1. Vibrational mean square amplitudes for bulk ZnO

In all simulations, the equations of motion were solvedwith the Verlet algorithm, using a time step of 0.25 fs.The temperature was kept constant using the Berendsenthermostat [41]. The volume of the box was equilibratedusing the NPT (constant temperature/constant pressure)MD method, where the Berendsen barostat [41] was usedto keep the internal pressure of the box at 0 GPa during

UN

CO

RR

EC

Table 9Atomic vibrational mean square amplitudes in wurtzite ZnO, hu2i, parallel ðU 33

at 20 K, 300 K and 600 K

20 K 300

? c axis k c axis ? c

hu2iZn=ðA2Þ 0.00098(8) 0.00041(2) 0.00

hu2iO=ðA2Þ 0.00081(7) 0.00036(2) 0.00

The number in parentheses is the standard error of the mean.

Crystal growth

a bFig. 7. Snapshots taken from one of the MD simulations of crystal growth on aafter the end of equilibration. ZnO molecules were added during the first 30 p

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

OF

the subsequent NVT simulations. The system consisted of360 ions in a simulation box, initially of the dimensions16:5 A� 16:6 A� 17:2 A. The system was equilibratedfor 200 ps simulation time prior to the production runswere trajectories where recorded during 800 ps.

Fig. 5 shows a comparison between atomic mean squareamplitudes from the ReaxFF model and anisotropic tem-perature coefficients from diffraction experiments. The tem-perature coefficients give a measure of the averagevibrational motion of the ions around their equilibriumpositions in the crystal. Neutron data was taken fromAlbertsson et al. [23] and X-ray data from Schultz andThiemann [42] and Kihara and Donnay [43]. The numeri-cal values from the ReaxFF model are given in Table 8.We observe good agreement between the ReaxFF modeland experimental data (see Table 9).

ED

PR4.2. ZnO surfaces and crystal growth

Here we employed conditions as close as possible to theconditions used in the simulations of ZnO(0 001) crystalgrowth by Kubo et al. in Ref. [15]. In this model, the topof the simulation box corresponds to a source which emitsZnO monomer molecules towards a ZnO slab correspond-ing to a substrate. Two slab models were used under three-dimensional periodic boundary conditions, namely onesmooth and one stepped surface. The stepped surface wascreated by removing half a surface layer on each side ofthe slab, resulting in two kinks on each side of the brokensurface layer, on both sides of the slab. The surface area

Þ and perpendicular ðU 11Þ to the c axis, calculated from the ReaxFF model

K 600 K

axis k c axis ? c axis k c axis

66(0) 0.0064(2) 0.0120(2) 0.0124(2)59(0) 0.0058(1) 0.0108(1) 0.0113(1)

csmooth wurtzite ZnO(0001)-surface, at (a) 30 ps, (b) 130 ps and (c) 300 pss of simulation.

08), doi:10.1016/j.susc.2007.12.023

Page 11: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

T

OF

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

Crystal growth

a b cFig. 8. Snapshots taken from one of the MD simulations of crystal growth on a stepped wurtzite ZnO(0001)-surface, at (a) 30 ps, (b) 130 ps and (c) 300 psafter the end of equilibration. ZnO molecules were added during the first 30 ps of simulation.

D. Raymand et al. / Surface Science xxx (2008) xxx–xxx 11

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

UN

CO

RR

EC

was in both cases 16:45� 17:10 A2, and the thickness

approximately 13.8 A. The vacuum gap separating theoxide slabs was 50 A. ZnO molecules were made to emergewith random velocities (distributed on a half sphere withradius 900 m/s) from the emitting source located 20 Aabove the surface. The molecules were emitted one byone at regular time intervals of 2 ps. The system was equil-ibrated for 40 ps, prior to the epitaxial growth simulations.The calculations were performed for 30 ps, and a total of15 ZnO molecules were deposited on the ZnO(0 001) sur-face; this corresponds to half a monolayer. The tempera-ture was kept constant at 700 K, which is similar toexperimental growth conditions.

In addition to the procedure outlined in Ref. [15], wecarried out additionally nine simulations runs for each ofthe two surface models, to improve the statistics. More-over, to allow the surfaces to properly relax and recon-struct, the simulations were extended with 300 ps, duringwhich time the number of ions was kept fixed. In all simu-lations, the equations of motion were solved with the Verletalgorithm using a time step of 0.25 fs. The temperature waskept constant using the Berendsen thermostat [41] and thevolume of the box was kept constant.

For both the smooth surface (see Fig. 7) and the steppedsurface (see Fig. 8) we see crystal growth. In this respectour model disagrees with the results in Ref. [15], wheregrowth was only observed for the stepped surface. In oursimulations of the smooth surface we find no preferentialgrowth position and the incoming dimers distribute them-selves randomly over the surface. During the subsequent,extended MD simulation, an increasingly well orderedZnO(0 001) layer was formed. We see this behavior in 8out of the 10 simulation runs.

In simulations of the stepped surface, both Ref. [15] andourselves find that the molecules preferentially adsorb closeto the kinks. We find that the ZnO molecules adsorb bothon the upper and the lower sides of the kinks. Our resultfor the stepped surface is that the growth, rather thanmending the broken layer by filling the trench, forms twonew facets on either side of the broken layer. We see this

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

O

behavior in 9 out of the 10 simulation runs. We believe thatif the growth had been allowed to continue, these two newfacets would eventually meet each other in the trench, andsubsequently order into a ZnO(0 001) monolayer. How-ever, the surface would then be permanently altered to-wards a stepped surface with terraces of lower surface-energy facets, most probably with large ZnOð10�10Þ char-acter. A more rigorous treatment is needed to extract themaximum amount of information from the force fieldabout mechanisms and rates for the present system. How-ever, we can already from our model conclude that differ-ent growth behavior is observed depending on thepresence of surface steps.

5. Summary

The present work presents a reactive force field for ZnO,developed to allow an accurate description of the chemicaland physical properties of condensed-phase ZnO. Goodagreement with DFT calculations is observed for variouscondensed systems, bulk and surfaces. Emphasis has beenput on obtaining a good surface description, to allow fu-ture studies of surface chemical reactions important forcatalysis. The force field successfully reproduces the atomicvibrational mean square amplitudes of bulk ZnO. More-over, the force field was used to model crystal growth.We observe a distinct growth pattern dependence on sur-face structure; on a smooth surface we find a randomgrowth pattern, while a stepped surface causes a distinctgrowth pattern related to preferential reactions near sur-face kink-sites. This latter observation is in agreement withearlier simulations on ZnO crystal growth from theliterature.

Acknowledgements

This work was supported by the Swedish ResearchCouncil (VR). Valuable assistance with the quantum-mechanical bulk calculations was given by Dr. Qing Zhang

08), doi:10.1016/j.susc.2007.12.023

Page 12: A reactive force field (ReaxFF) for zinc oxide · UNCORRECTED PROOF 1 2 A reactive force field (ReaxFF) for zinc oxide 3 David Raymand a, Adri C.T. van Duinb, Micael Baudina, …

T

683

684

685

686687688689690691692693694695696697698699700701702703704705706707708709710711712713714715716717718719720721722723724725726727728729730731

732733734735736737738739740741742743744745746747748749750751752753754755756757758759760761762763764765766767768769770771772773774775776777778779780Q2

781

12 D. Raymand et al. / Surface Science xxx (2008) xxx–xxx

SUSC 17820 No. of Pages 12, Model 5+

11 January 2008 Disk UsedARTICLE IN PRESS

OR

REC

and Mr. Henrik Eriksson. Many useful discussions withDr. Pavlin Mitev are acknowledged.

References

[1] Z.R. Tian, J.A. Voigt, J. Liu, B. McKenzie, M.J. McDermott, M.A.Rodriguez, H. Konishi, H. Xu, Nat. Mater. 2 (2003) 821.

[2] G.A. Somorjai, Introduction to Surface Chemistry and Catalysis,John Wiley & Sons, Inc., New York, USA, 1994.

[3] P. Persson, L. Ojamae, Chem. Phys. Lett. 321 (3–4) (2000) 302.[4] U. Ozgur, Y.I. Alivov, C. Liu, M.A.R.A. Teke, S. Dogan, V. Avrutin,

S.J. Cho, H. Morkoc, J. Appl. Phys. 98 (2005) 041301.[5] M. Catti, Y. Noel, R. Dovesi, J. Phys. Chem. Solids 64 (2003) 2183.[6] J. Serrano, A. Romero, F. Lauck, M. Cardona, A. Rubio, Phys. Rev.

B 69 (2004) 094306.[7] Z.G. Yu, H. Gong, P. Wu, J. Cryst. Growth 287 (2006) 199.[8] C. Lee, W. Yang, R. Parr, Phys. Rev. B 37 (1988) 785.[9] A.D. Becke, J. Chem. Phys. 98 (1993) 5968.

[10] B. Dick, A. Overhauser, Phys. Rev. 112 (1958) 90.[11] D.J. Binks, R.W. Grimes, J. Am. Ceram. Soc. 76 (1993) 2370.[12] L. Whitmore, A.A. Sokol, C.R.A. Catlow, Surf. Sci. 498 (2001) 135.[13] A.A. Sokol, S.T. Bromley, S.A. French, C.R.A. Catlow, P. Sher-

wood, Int. J. Quantum Chem. 99 (2004) 695.[14] A.J. Kulkarni, M. Zhou, F.J. Ke, Nanotechnology 16 (2005) 2749.[15] M. Kubo, Y. Oumi, H. Takaba, A. Chatterjee, A. Miyamoto, M.

Kawasaki, M. Yoshimoto, H. Koinuma, Phys. Rev. B 61 (23) (2000)16187.

[16] P. Erhart, N. Juslin, O. Goy, K. Nordlund, R. Muller, K. Albe, J.Phys. Cond. Matter 18 (2006) 6585.

[17] A.C.T. van Duin, S. Dasgupta, F. Lorant, W.A. Goddard III, J.Phys. Chem. A 105 (2001) 9396.

[18] A.C.T. van Duin, A. Strachan, S. Stewman, Q. Zhang, X. Xu, W.A.Goddard III, J. Phys. Chem. A 107 (2003) 3803.

[19] Q. Zhang, T. Cagin, A.C.T. van Duin, W.A. Goddard III, Y. Qi, G.Hector, Phys. Rev. B 69 (2004) 045423.

[20] G.O.A. Janssens, B.G. Baekelandt, H. Toufar, W.J. Martier, R.A.Shoonheydt, J. Phys. Chem. B 99 (1995) 3251.

[21] W.A. Goddard III, A.C.T. van Duin, K. Chenoweth, M. Cheng, S.Pudar, J. Oxgaard, B. Merinov, Y.H. Jang, P. Persson, Topics Catal.38 (2006) 1.

[22] T. Hahn, International Tables for Crystallography Volume A: Space-group symmetry, IUCr, Chester, UK, 2002.

[23] J. Albertsson, S.C. Abrahams, A. Kvick, Acta Crystall. B 45 (1)(1989) 34.

[24] J.E. Jaffe, A.C. Hess, Phys. Rev. B 48 (1993) 7903.[25] A.C.T. van Duin, J.M.A. Baas, B.J. van de Graaf, J. Chem. Soc.,

Faraday Trans. 90 (1994) 2881.[26] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb,

J.R. Cheeseman, J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C.Burant, J.M. Millam, S.S. Iyengar, J. Tomasi, V. Barone, B.Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson, H.

UN

C

Please cite this article in press as: D. Raymand et al., Surf. Sci. (20

ED

PR

OO

F

Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa,M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X.Li, J.E. Knox, H.P. Hratchian, J.B. Cross, V. Bakken, C. Adamo, J.Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R.Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A.Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich,A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck,K. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S.Clifford, J. Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P.Piskorz, I. Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill,B. Johnson, W. Chen, M.W. Wong, C. Gonzalez, J.A. Pople,Gaussian 03, Revision C.02, Gaussian, Inc., Wallingford, CT, 2004.

[27] V.R. Saunders, R. Dovesi, C. Roetti, R. Orlando, C.M. Zicovich-Wilson, N. Harrison, K. Doll, B. Civalleri, I. Bush, P.D. Arco, M.Llunell, CRYSTAL2003 User’s Manual (2003).

[28] P.J. Schultz, A description of the method is in P.J. Feibelman, Phys.Rev. B 35 (1987) 2626.

[29] L. Fast, J. Wills, B. Johansson, O. Eriksson, J. Inst. Metals 83 (1995)49.

[30] A. Wander, N. Harrison, Surf. Sci. 457 (2000) 342.[31] J. Jaffe, A. Hess, J. Chem. Phys. 104 (1996) 3348.[32] Y. Noel, C.M. Zicovich-Wilson, B. Civalleri, P. D’Arco, R. Dovesi,

Phys. Rev. B 65 (2001) 014111.[33] H. Monkhorst, J. Pack, Phys. Rev. B 13 (1976) 5188.[34] J. Perdew, K. Burke, M. Ernzehof, Phys. Rev. Lett. 77 (1996) 3865.[35] M. Fuchs, M. Scheffler, Comput. Phys. Commun. 119 (1999) 67.[36] G. Aylward, T. Findlay, S.I. Chemical Data, John Wiley & Sons,

Ltd., Milton, Australia, 1987.[37] W.L. Masterton, E.J. Slowinski, C.L. Stanitski, Chemical Principles,

CBS, New York, USA, 1987.[38] R. Krishnan, J.S. Binkley, R. Seeger, J.A. Pople, J. Chem. Phys. 72

(1980) 650.[39] T. Clark, J. Chandrasekhar, G.W. Spitznagel, P.v.R. Schleyer, J.

Comp. Chem. 4 (1983) 294.[40] K. Raghavachari, G.W. Trucks, J. Chem. Phys. 91 (1989) 1062.[41] H.J.C. Berendsen, J.P.M. Postma, W.F.V. Gunsteren, A. DiNola,

J.R. Haak, J. Chem. Phys. 81 (1984) 3684.[42] H. Schultz, K.H. Thiemann, Solid State Commun. 32 (1979) 783.[43] K. Kihara, G. Donnay, Can. Miner. 23 (1985) 647.[44] S. Desgreniers, Phys. Rev. B 58 (1998) 14102.[45] T. Bateman, J. Appl. Phys. 33 (1962) 3309.[46] W.P.H. Karzel, M. Kofferlein, W. Schiessl, M. Steiner, U. Hiller,

G.M. Kalvius, D.W. Mitchell, T.P. Das, P. Blaha, K. Schwarz, M.P.Pasternak, Phys. Rev. B 58 (1998) 14102.

[47] J. Brown, J. Inst. Metals 83 (1955) 49.[48] H. Ledbetter, J. Phys. Chem. Ref. Data 6 (1977) 1181.[49] K. Chenoweth, A.C.T. van Duin, W.A. Goddard III, J. Phys. Chem.

A, in press.

08), doi:10.1016/j.susc.2007.12.023