Monte Carlo Simulation of Ferroelectric Domain Structure: Electrostatic and Elastic Strain Energy Contributions

Similar documents
Monte Carlo Simulation of Ferroelectric Domain Structure: Electrostatic and Elastic Strain Energy Contributions

PHASE-FIELD SIMULATION OF DOMAIN STRUCTURE EVOLUTION IN FERROELECTRIC THIN FILMS

Domain-size dependence of piezoelectric properties of ferroelectrics

Journal. Three-Dimensional Computer Simulation of Ferroelectric Domain Formation. Theory and Modeling of Glasses and Ceramics

Monte Carlo simulation on dielectric and ferroelectric behaviors of relaxor ferroelectrics

Generalized continuum theory for ferroelectric thin films

Effect of interfacial dislocations on ferroelectric phase stability and domain morphology in a thin film a phase-field model

arxiv:cond-mat/ v1 10 Jun 1994 K. M. Rabe

Landau-Ginzburg model for antiferroelectric phase transitions based on microscopic symmetry

Classification of Dielectrics & Applications

Introduction to solid state physics

Micromechanical modeling and simulation of piezoceramic materials

Electro-shape-memory effect in Mn-doped BaTiO 3 single crystals and in situ observation of the reversible domain switching

Supplementary Figures:

Interfacial Coherency and Ferroelectricity of BaTiO 3 /SrTiO 3 Superlattice Films

MSE 201A Thermodynamics and Phase Transformations Fall, 2008 Problem Set No. 7

Phase-Field Simulation of the Effect of Elastic Inhomogeneity on Microstructure Evolution in Ni-Based Superalloys

Lab 70 in TFFM08. Curie & Ising

Mustafa Uludogan 1, Tahir Cagin, William A. Goddard, III Materials and Process Simulation Center, Caltech, Pasadena, CA 91125, U.S.A.

Effect of substrate-induced strains on the spontaneous polarization of epitaxial BiFeO 3 thin films

Roger Johnson Structure and Dynamics: Displacive phase transition Lecture 9

Effect of substrate constraint on the stability and evolution of ferroelectric domain structures in thin films

Spontaneous lateral composition modulation in InAlAs and InGaAs short-period superlattices

The Power of Feedback: From Dielectric to Ferroelectric Systems

Deepam Maurya 1*, Yuan Zhou 1, Yaojin Wang 2, Yongke Yan 1, Jiefang Li 2, Dwight Viehland 2, and Shashank Priya 1*

The Magnetic Properties of Superparamagnetic Particles by a Monte Carlo Method

Supplementary Information

Chap. 7. Dielectric Materials and Insulation

Relaxor characteristics of ferroelectric BaZr 0.2 Ti 0.8 O 3 ceramics

Electric field dependent sound velocity change in Ba 1 x Ca x TiO 3 ferroelectric perovskites

SUPPLEMENTARY MATERIAL

Ferroelectric Materials

Adaptive ferroelectric states in systems with low domain wall energy: Tetragonal microdomains

Cellular solid structures with unbounded thermal expansion. Roderic Lakes. Journal of Materials Science Letters, 15, (1996).

properties Michele Catti Dipartimento di Scienza dei Materiali Università di Milano Bicocca, Italy

Monte Carlo Simulation of the Ising Model. Abstract

Phenomenological Theory of a Single Domain Wall in Uniaxial Trigonal Ferroelectrics: Lithium Niobate and Lithium Tantalate

Aging effect evolution during ferroelectricferroelectric phase transition: A mechanism study

Synthesis and Ferroelectric Properties of KNO 3 films

Direct measurement of giant electrocaloric effect in BaTiO 3 multilayer thick film structure beyond theoretical prediction

File Name: Supplementary Information Description: Supplementary Figures, Supplementary Table, Supplementary Notes and Supplementary References.

Simulation of Hysteresis In Permalloy Films

Toughening due to domain switching in single crystal ferroelectric materials

INFLUENCE OF AN APPLIED STRAIN FIELD ON MICROSTRUCTURAL EVOLUTION DURING THE α 2 O- PHASE TRANSFORMATION IN Ti Al Nb SYSTEM

Constructing Landau-Ginzburg-Devonshire Type Models for Ferroelectric Systems Based on Symmetry

Ferroelectrics and ferroelectric domains

Thermodynamics Study on the Decay Properties of Reversed Domains in LiNbO 3. Single Crystals

In situ observation of reversible domain switching in aged Mn-doped BaTiO 3 single crystals

Depolarization Field Effect on Dielectric and Piezoelectric Properties of Particulate. Ferroelectric Ceramic-Polymer Composites

Intermolecular Forces and Monte-Carlo Integration 열역학특수연구

Two simple lattice models of the equilibrium shape and the surface morphology of supported 3D crystallites

Elastic constants and the effect of strain on monovacancy concentration in fcc hard-sphere crystals

Phase Transitions in Relaxor Ferroelectrics

Materials 218/UCSB: Phase transitions and polar materials

Magnetic properties of spherical fcc clusters with radial surface anisotropy

Surface stress and relaxation in metals

Trapping of oxygen vacancies on twin walls of CaTiO 3 :acomputer simulation study

Ferroelectricity. Phase transition. Material properties. 4/12/2011 Physics 403 Spring

Phase-field simulation of polarization switching and domain evolution in ferroelectric polycrystals

Metropolis Monte Carlo simulation of the Ising Model

3.320 Lecture 18 (4/12/05)

Dielectric, Ferroelectric, and Optical Properties

Domain switching and electromechanical properties of pulse poled Pb Zn 1Õ3 Nb 2Õ3 O 3 PbTiO 3 crystals

MODELING AND SIMULATION OF PIEZOCERAMIC MATERIALS USING MICROMECHANICAL APPROACH

Addition 1. Shear Stack Piezoelectric Elements and Shear Effect Basics

Phase-field modeling of the coupled microstructure and fracture evolution in ferroelectric single crystals

Transduction Based on Changes in the Energy Stored in an Electrical Field

Supplementary Figures

Nucleation and evaporation of domains due to electric field at room temperature in BaTiO 3 single crystals

Reconstruction and intermixing in thin Ge layers on Si 001

Computer simulation methods (2) Dr. Vania Calandrini

Chapter 19: Elastic Inclusions

Elements of Continuum Elasticity. David M. Parks Mechanics and Materials II February 25, 2004

Monoclinic phases arising across thermal inter-ferroelectric phase transitions

PIEZOELECTRIC TECHNOLOGY PRIMER

Ab initio study of ferroelectric domain walls in PbTiO 3

Quantitative analysis of the first-principles effective Hamiltonian approach to ferroelectric perovskites

Exploring the energy landscape

Numerical methods to compute optical errors due to stress birefringence

The electric field induced strain behavior of single. PZT piezoelectric ceramic fiber

SECOND PUBLIC EXAMINATION. Honour School of Physics Part C: 4 Year Course. Honour School of Physics and Philosophy Part C C3: CONDENSED MATTER PHYSICS

Kinetic Monte Carlo: from transition probabilities to transition rates

Potts And XY, Together At Last

Domain Structures in Epitaxial Ferroelectric Thin Films

Physical and Biological Properties of Agricultural Products Acoustic, Electrical and Optical Properties and Biochemical Property

Haicheng Guo Feng Li Nirand Pisutha-Arnond Danqing Wang. Introduction

A Monte Carlo Implementation of the Ising Model in Python

Curie-Weiss law in thin-film ferroelectrics

Improved model of nonaffine strain measure

Liquid Crystals IAM-CHOON 1(1100 .,4 WILEY 2007 WILEY-INTERSCIENCE A JOHN WILEY & SONS, INC., PUBLICATION. 'i; Second Edition. n z

High tunable dielectric response of Pb 0.87 Ba 0.1 La 0.02 (Zr 0.6 Sn 0.33 Ti 0.07 ) O 3 thin film

The Pennsylvania State University. The Graduate School. College of Engineering ON THE IMPACT OF ELECTRODE GEOMETRY ON THE PIEZOELECTRIC

Numerical Analysis on Magnetic-induced Shear Modulus of Magnetorheological Elastomers Based on Multi-chain Model

Basic Equations of Elasticity

I. B. Misirlioglu*, M. Yildiz. Faculty of Engineering and Natural Sciences, Sabanci University, Tuzla/Orhanli, Istanbul, Turkey.

Supplementary Figure 1. Optical and magneto-optical responses for 80 nm diameter particles

arxiv:mtrl-th/ v1 13 Feb 1995

Chapter 2 Magnetic Properties

Let's look at the force on a current loop. In a uniform field it is zero: F = I I (dl B) =I I dl B =0 (4) since B is constant and comes outside the in

Phonon Dispersion, Interatomic Force Constants Thermodynamic Quantities

Transcription:

Monte Carlo Simulation of Ferroelectric Domain Structure: Electrostatic and Elastic Strain Energy Contributions B.G. Potter, Jr., B.A. Tuttle, and V. Tikare Sandia National Laboratories Albuquerque, NM 87185-1405 Abstract. A lattice-monte Carlo approach was developed to simulate ferroelectric domain behavior. The model utilizes a Hamiltonian for the total energy that includes electrostatic terms (involving dipole-dipole interactions, local polarization gradients, and applied electric field), and elastic strain energy. The contributions of these energy components to the domain structure and to the overall applied field response of the system were examined. In general, the model exhibited domain structure characteristics consistent with those observed in a tetragonally distorted ferroelectric. Good qualitative agreement between the appearance of simulated electrical hysteresis loops and those characteristic of real ferroelectric materials was found. INTRODUCTION The electrical, mechanical and optical behavior of a ferroelectric is dictated by the morphology and reorientation dynamics of domains, regions consisting of likeoriented spontaneous polarization. The ferroelectric attempts to attain a domain configuration that minimizes the total energy while satisfying both electrostatic and mechanical boundary conditions. The interplay between domain configuration, microstructure, and service conditions (e.g. temperature, applied field, pressure) dictate the engineering performance of a ferroelectric. Development of a computational approach to simulate the domain structure of a ferroelectric would enhance a broader model to predict the mechanical and electrical response of a ferroelectric. This capability could be used to guide the processing and utilization of ferroelectric ceramics in various applications to optimize performance and evaluate aging effects. Numerical simulations of ferroelectrics and their properties have been the focus of numerous research efforts. Techniques including ab initio calculations [1-3], the solution of the Landau-Ginzburg energy relations [4-6], and Potts-based Monte Carlo simulations [7,8] have been pursued. The present work develops a lattice-based Monte Carlo simulation technique for the prediction of ferroelectric domain configurations. The model is two-dimensional and, rather than relying on the typical Pottstype Hamiltonian to evaluate neighboring lattice site interactions, the present approach utilizes potentials based on electrostatic and mechanical phenomena that occur in fer-

roelectrics. As we will discuss below, these longer range interactions produce domain morphologies and relative orientations of dipoles across domain boundaries that are consistent with those observed experimentally. The present work will describe the model and the influence of the electrostatic and elastic strain energy terms used on the evolution of domain assemblage. The response of the dipole ensemble to an oscillating applied field will also be examined through the simulation of electrical hysteresis loops. MODEL DESCRIPTION Crystal and Domain Representation In the current lattice-based Monte Carlo (MC) simulation, space is discretized into a two-dimensional, square array, or grid, of lattice points. Each lattice point is assigned the orientation and magnitude of the local, spontaneous dipole moment associated with a single ferroelectric unit cell. A tetragonal lattice distortion is assumed to describe the formation of permanent dipoles in the ferroelectric material. Two orthogonal basis vectors (twodimensions) are therefore used to describe the local dipole orientation. Four dipole orientations are allowed within the plane defined by the basis vectors. The simulation is further limited to a single crystal orientation with dipole moment basis vectors aligned with the principal axes of the square lattice used to discretize space. Allowed dipole orientations are therefore up, down, left and right. While there is no inherent restriction on the relative orientation and symmetry used for the simulation lattice, the alignment of the simulation lattice basis vectors with those used to represent the dipole moment is consistent with a one-to-one correspondence between a simulation lattice site and a single ferroelectric unit cell and the assumption that crystallographic lattice coherence is maintained through the cubic to tetragonal phase transformation. This choice of lattice and dipole orientation symmetries thus enables the simulation to clearly exhibit the general domain wall orientations and the relative orientations of dipoles in adjacent domains that are anticipated in a tetragonal ferroelectric material (e.g. 180 and 90 domain walls). The dipole moment magnitudes are assumed to be constant during the simulation. In most cases, the local dipole orientations were randomly assigned to lattice points to furnish a starting configuration. While the elastic strain energy is utilized in the model, the current implementation does not allow for motion of adjacent lattice sites with respect to one another, thus neglecting the influence of local displacements that describe the elastic strain field.

Interaction Potentials The present work attempts to introduce physically realistic interaction potentials based on the known electrostatic and elastic mechanical nature of the ferroelectric material associated with each lattice point. Since the lattice sites are associated with permanent dipole moments within the ferroelectric, a dipole-dipole potential energy expression has been introduced. Thus, the pairwise interaction potential energy between lattice points is given by the vector expression: ( 3( )( ) ) E K P L P L P L n P L n i j i j ij = 3 r where P is the dipole moment vector at a given lattice site, n is a unit vector connecting points i and j, r = magnitude of the separation vector between i and j, and K is a scaling factor (contains the material permittivity). The dipole-dipole interaction is truncated at a distance where the potential has dropped to <% of its magnitude at r = 1 lattice unit. In addition, the truncation distance must also be less than one-half the size of the simulation box to avoid a dipole interacting with its image when using periodic boundary conditions (minimum image convention). A truncation distance of 6 lattice unit distances was utilized in the present study; larger truncation distances did not significantly alter the steady-state dipole configurations obtained or their dipole-dipole potential energy [9]. A potential term linked to the presence of polarization gradients is used to include an electrostatically based domain wall energy. The expression is adapted from that used by Hu and Chen [6]. The potential energy density takes the form: (1) D i = P P P x y Px y + + + x y x y J P P P x y Px y + + y x y x () where x and y denote spatial variables, P is the local electric polarization, and J is a scaling factor for the domain wall interaction potential and it is positive, making the overall energy term positive. Unless otherwise noted, the magnitude of J in the present work was chosen to energetically favor a single domain configuration under periodic boundary conditions (infinite single crystal model) and a multi-domain structure under finite boundary conditions when using only electrostatic contributions to the total configuration energy. An applied bias field was also included in the simulation. The potential energy of the dipole in the field is expressed as the dot product between the field vector and

the dipole moment, as dictated by electrostatics. In the present study, the applied field is taken to be constant over the entire lattice without regard for electrostatic screening effects known to produce local modifications in the field strength. The field-related potential energy is therefore given as: E = α i ( P F ) (3) i where P is the dipole moment, F is the applied field, and α is a scaling factor. Elastic strain energy at each step in the Monte Carlo simulation is calculated based on an analytical expression developed by Khachaturyan [10] for a multiphase, coherent mixture assuming a uniform, anisotropic elastic modulus in each phase. In this case, each dipole is associated with a portion of material that has undergone a phase transformation from cubic (paraelectric) to tetragonal (ferroelectric). The orientation of the dipole at a location in the simulation then gives the orientation of the local transformation strain field, i.e. the c-axis distortion direction is aligned along the indicated dipole orientation. In the present study, the simulation is therefore treated as a composite of two different material phases with the spatial distribution of these phases given by the dipole configuration. The total elastic strain energy for the ferroelectric body is given in terms of the elastic strain energy associated with the local transformations (term 1 in equation (4) below) and the concomitant relaxation of the surrounding material (terms and 3). 1 o o 1 o o E elastic = Vλijklεij ( p) ε kl ( p) wp Vλijklεij ( p) ε kl ( q) w p p, q 1 3 d k o ˆ o nσˆ ( p) Ω( n) ˆ ( q) n Θ ( k) * 3 p Θ q p, q (π ) σ ( k) p w q (4) where V is the total volume, ε o (p) is the transformation strain of phase p, σ o (p) is the transformation stress of phase p, λ is the elastic modulus tensor, w p is the volume fraction of phase p, Θ p (k) is the Fourier transform of the shape function, Θ p (r) (a spatial mapping of the pth phase), and Ω is a tensor whose inverse contains λ. The relaxation energy is expressed in terms of homogeneous and heterogeneous strain components (terms and 3 in equation (4), respectively) where the heterogeneous strain term is defined such that the total real-space displacement at the body edges is zero. All displacements associated with the body edges are contained in the homogeneous strain term. Again, information regarding real-space displacements of the simulation lattice is not used in the current model. The strain energy calculated above is combined with the electrostatic energy to determine the total energy change associated with a given Monte Carlo move.

Simulation Units Electrostatic energies were calculated in reduced units with length defined by a simulation lattice constant (called a lattice unit (LU) ) and the dipole moment defined to be 1 charge-lu. Given the current assignment of a ferroelectric unit cell to each simulation lattice point, the LU is therefore directly related to the crystallographic lattice constant. An energy unit (EU) is then equal to the potential energy obtained with two parallel dipole moments of unit magnitude separated by a vector of length one lattice unit oriented perpendicular to the dipole moments. Temperature is reported in terms of energy, i.e. kt, in EU s. Elastic strain energies were calculated using elastic constants and transformation strains representative of BaTiO 3 : C 11 = 75 x 10 9 N/m, C 1 = 179 x 10 9 N/m, C 44 = 54.3 x 10 9 N/m, ε 11 = ε = -3.74 x 10-3, ε 33 = 6.3 x 10-3. The results were converted to energy units (EU) to allow combination with electrostatic energy results to derive a total energy for the simulation. The temporal unit is a Monte Carlo timestep (MCS). The MC timestep represents m x m MC move attempts, where m is the size of the simulation box. 50 x 50 or 100 x 100 LU simulation sizes were typically utilized in this study. Where smaller ensembles are noted, the results are representative of larger simulations. Simulation Procedure A Monte Carlo move attempt involves the random selection of a lattice site followed by random reorientation of the dipole at this site. Move acceptance is determined using the Metropolis importance sampling algorithm [11] based on the energy change associated with the dipole reorientation and the simulation temperature. Simulation temperatures were chosen to provide sufficient thermal energy to allow Monte Carlo move acceptance rates in the range 0. to 0.4, thus favoring evolution of a steady-state dipole configuration within a reasonable computation time without significant randomization of the configuration due to thermal energy. A steadystate condition was defined when little or no further minimization in the ensemble s average total energy was observed with additional MC moves. Simulations were performed using either the electrostatic potential only or using both electrostatic and elastic strain energy contributions to the total energy. This allowed the effect of each energy term on the evolution of the dipole ensemble configuration to be studied. A more detailed evaluation of the specific effects of the individual electrostatic terms was undertaken in previous work [9]. After gaining insight into the relative impact of the different energy terms and simulation conditions on the evolution of the dipole configuration, the model was extended to allow the simulation of ferroelectric-type electrical hysteresis loops. In this case, the applied field value was changed by a known increment and the simulation was allowed to evolve from the configuration state developed at the previous field value. The net polarization component aligned with the applied field was calculated from the dipole configuration at each field magnitude. The effective field frequency was varied by changing the equilibration time allowed the system at a given field

value. Electrical hysteresis was simulated for a purely electrostatic system and for the combined Hamiltonian including both the electrostatic and elastic strain energy terms. RESULTS AND DISCUSSION Figure 1a shows a representative dipole configuration for a purely electrostatic simulation at kt = 1.5 EU under periodic boundary conditions. (The domain wall scaling factor, J, has been adjusted to allow the observation of multiple domains.) The presence of 180 domain walls is evident. Head-to-tail alignment of dipoles across the boundaries is dictated by the dipole-dipole potential used. Addition of elastic strain energy significantly affects the evolution and appearance of the simulation, resulting in the formation of 90 domain walls (see Figure 1b). Again, a head-to-tail alignment across the walls is observed due to the dipole-dipole electrostatic interaction. a. b. FIGURE 1. Representative dipole configurations after 750 MCS for simulations including electrostatic potential terms only (a.) and both electrostatic and elastic strain energy terms (b.) The impact of the elastic strain energy on the switching dynamics of the dipole ensemble is indicated in Figure which contains representative hysteresis loops simulated at a frequency of.5 x 10-5 MCS -1 and kt = 4 EU for both pure electrostatic and combined electrostatic and elastic strain energy models, respectively. Both simulations utilized equal maximum field strengths (sufficient to insure saturation of the polarization alignment for the purely electrostatic model at this frequency). Significant noise is observed in the loops due to the small simulation size used (0 x 0).

1. electrostatic and elastic strain energy 1.0 electrostatic energy only 0.8 E/E max 0.6 0.4 0. -1.0-0.5 0.5 1.0-0. -0.4-0.6-0.8-1.0-1. P/P max FIGURE. Representative electrical hysteresis loops simulated using the lattice-mc approach. The solid curve was obtained using only electrostatic terms in the model Hamiltonian, the dashed curve utilized both electrostatic and elastic strain energy terms. It is clear that the polarization at maximum field is reduced and that the coercive field is increased with the introduction of the elastic strain energy term. Figure 3 shows the dipole configurations at the positive (upward) and negative (downward) maximum field values in the electrostatic and elastic strain energy model. a. b. FIGURE 3. Dipole configurations for the maximum negative (a.) and positive (b.) applied electric fields for the simulated hysteresis curve shown in Figure for the combined electrostatic and elastic strain energy model. Figure 4a depicts the 90 dipole reorientation rate (90 successes/90 attempts) with applied field for the combined electrostatic and strain energy run. Results for

180 switching are given in Figure 4b. Analogous data for the electrostatic only case are given in Figures 5a and 5b. In all cases, the reorientation acceptance rates are initially high as the simulation begins with a randomly oriented dipole ensemble. 90 degree reorientation success rate 0.018 0.016 0.014 0.01 0.010 0.008 0.006 0.004 0.00 0.000 electrostatic and elastic strain energy -1.0-0.5 0.0 0.5 1.0 Applied Field (normalized) 180 degree reorientation success rate 0.545 0.540 0.535 0.530 0.55 0.50 0.515 0.510 0.505 0.500 electrostatic and elastic strain energy -1.0-0.5 0.0 0.5 1.0 Applied Field (normalized) a. b. FIGURE 4. Dipole reorientation rates for 90 (a.) and 180 (b.) reorientations corresponding to the hysteresis curve given in Figure for the combined electrostatic and elastic strain energy model. 90 degree reorientation success rate 0. 0.0 0.18 0.16 0.14 0.1 0.10 0.08 0.06 0.04 0.0 electrostatic energy only 180 degree reorientation success rate 0.550 0.545 0.540 0.535 0.530 0.55 0.50 0.515 0.510 electrostatic energy only -1.0-0.5 0.0 0.5 1.0 Applied Field (normalized) -1.0-0.5 0.0 0.5 1.0 Applied Field (normalized) a. b. FIGURE 5. Dipole reorientation rates for 90 (a.) and 180 (b.) reorientations corresponding to the hysteresis curve given in Figure for the model containing only electrostatic energy. In Figure 4, the 90 reorientation rate becomes nearly constant at a value only 0% of its original value and 0.6% of the corresponding 180 reorientation rate at zero applied field after initial ordering of the dipoles with field early in the simulation. These consistent move success rates observed for 90 reorientations are largely due to

thermally allowed reorientations rather than a continuous switching process in the present simulation. In contrast, the 180 reorientation rate is observed to gradually increase in magnitude as the positive (negative) coercive field is approached from below (above), indicating that the field-driven dipole reorientation process is occurring over a range of field magnitudes and that it is due mainly to 180 switching at these field strengths. Peaks observed in both 90 and 180 reorientation rate curves correspond to the coercive field of the electrical hysteresis loop, a point where the maximum switching rate is expected. A peak in the switching rates at the coercive field is also observed in the data of Figures 5a and 5b for the electrostatic only model. In this case, however, both the 90 and 180 reorientation rate curves increase monotonically as the positive (negative) coercive field is approached from below (above). These observations indicate that both processes contribute to switching over a broad range of applied fields. The difference between the 90 reorientation rate behaviors of the combined electrostatic and elastic strain energy model (Figure 4a) and the electrostatic only model (Figure 5a) is due to the enhanced energy penalty assigned to 90 dipole reorientation when elastic strain energy is included. Significant 90 reorientation success in this case is only activated when the applied field magnitude is sufficient to offset this energy penalty and increase the Monte Carlo move acceptance probability. It is interesting to note that, while the magnitude of the 90 reorientation rate is generally greater in simulations based only on electrostatic interactions (see Figure 5a), the overall success rate ratio of 90 vs. 180 reorientations ranges from 0.1 to 0.4 (depending on applied field) for the electrostatic-only model. This indicates that, even in the absence of strain energy, 90 reorientation is not the favored process for realignment with an applied field in this model. The reduction in the polarization at maximum field in the hysteresis loop of Figure (using both electrostatic and elastic strain terms), then, is the result of a limited success rate for 90 dipole reorientation under even the maximum applied electric field. Examination of the dipole configurations of Figure 3, together with the reorientation rate data of Figure 4, indicate that reorientation of the dipole ensemble with the applied field is accomplished predominantly by 180 switching rather than 90 domain wall motion at these field strengths. It is anticipated that increased maximum field strengths will increase the 90 dipole reorientation rate as an increased electrostatic driving force offsets elastic strain energy considerations for this process. This will promote the movement of 90 domain walls to result in an increased maximum polarization value for the system. These issues are currently under study. CONCLUSION A two-dimensional, lattice-monte Carlo model for the simulation of ferroelectric domain configurations has been presented. The model contains both electrostatic and elastic strain energy contributions to a total potential energy that determines the evolution of steady-state domain assemblages and the dynamics associated with reori-

entation of the dipole ensemble under applied electric fields. Evaluation of steadystate domain configurations obtained using purely electrostatic energy and both electrostatic and elastic strain energy terms confirm that the introduction of elastic strain considerations results in the formation of 90 domain structures. Electrostatic contributions to the model favor the head-to-tail alignment of dipoles across these domain boundaries. These observations are consistent with experimental evidence in real ferroelectric materials. The influence of elastic strain was further found to significantly impact the switching behavior observed in simulated electrical hysteresis, producing a decreased polarization observed at the maximum applied field and an increased coercive field over the model that did not include elastic strain energy. The lower polarization is attributed to a reduced 90 dipole reorientation rate that precluded the long-range motion of 90 domain walls under the field strengths applied. ACKNOWLEDGMENTS The authors thank J. Aidun, S. Hwang, and L.-Q. Chen for valuable discussions concerning the model. Sandia National Laboratories is a multiprogram laboratory operated by Sandia Corporation, a Lockheed Martin Company. This work was supported by the U.S. DOE under contract DE-AC04-94AL85000. REFERENCES 1. W. Zhong, D. Vanderbilt, and K. M. Rabe, Phys. Rev. Lett. 73, 1861 (1994).. R.E. Cohen and H. Krakauer, Phys. Rev B 4, 6416 (1990). 3. A. Garcia and D. Vanderbilt, Appl. Phys. Lett. 7, 981 (1998). 4. W. Cao and L.E. Cross, Phys. Rev. B 44, 5 (1991). 5. S. Nambu and D.A. Sagala, Phys. Rev. B 50, 5838 (1994). 6. H.-L. Hu and L.-Q. Chen, J. Am. Ceram. Soc. 81, 49 (1998). 7. M. Abel and R. Siems, Ferroelectrics 16, 75 (199). 8. J.-M. Liu and Z.G. Liu, Mater. Lett. 36, 17 (1998). 9. B.G. Potter, Jr., V. Tikare, and B.A. Tuttle, J. Appl. Phys. 87, 4415 (000). 10. A.G. Khachaturyan, Theory of Structural Transformations in Solids, John Wiley and Sons, Inc., New York, NY, 1983, pp. 198-1. 11. N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.N. Teller and E. Teller, J. Chem. Phys., 1, p. 1087 (1953).