Pair distribution function analysis driven by atomistic simulations Application to microwave radiation synthesized TiO 2and ZrO 2 Shuyan Zhang1 Jie Gong1 Daniel Xiao2 B. Reeja Jayan1 and Alan J. H. McGaughey1

2025-05-06 0 0 5.46MB 33 页 10玖币
侵权投诉
Pair distribution function analysis driven by atomistic simulations: Application to
microwave radiation synthesized TiO2and ZrO2
Shuyan Zhang1, Jie Gong1, Daniel Xiao2, B. Reeja Jayan1, and Alan J. H. McGaughey1
1Department of Mechanical Engineering, Carnegie Mellon University,
Pittsburgh, Pennsylvania 15213, USA and
2Department of Materials Science and Engineering, Carnegie Mellon University,
Pittsburgh, Pennsylvania 15213, USA
A workflow is presented for performing pair distribution function (PDF) analysis of defected ma-
terials using structures generated from atomistic simulations. A large collection of structures, which
differ in the types and concentrations of defects present, are obtained through energy minimization
with an empirical interatomic potential. Each of the structures is refined against an experimental
PDF. The structures with the lowest goodness of fit Rwvalues are taken as being representative
of the experimental structure. The workflow is applied to anatase titanium dioxide (a-TiO2) and
tetragonal zirconium dioxide (t-ZrO2) synthesized in the presence of microwave radiation, a low
temperature process that generates disorder. The results suggest that titanium vacancies and inter-
stitials are the dominant defects in a-TiO2, while oxygen vacancies dominate in t-ZrO2. Analysis
of the atomic displacement parameters extracted from the PDF refinement and mean squared dis-
placements calculated from molecular dynamics simulations indicate that while these two quantities
are closely related, it is challenging to make quantitative comparisons between them. The workflow
can be applied to other materials systems, including nanoparticles.
I. INTRODUCTION
The atomic structure of good quality crystalline materials can be obtained from X-ray crystallography, which
only measures the Bragg peaks that result from the atomic periodicity. Crystallography alone is not sufficient for
specifying the atomic structure of nanocrystalline and/or highly-defected samples [1–4]. In pair distribution function
(PDF) experiments, on the other hand, X-rays or neutrons are used and a wider angular range is detected so that
both Bragg scattering and diffuse scattering are collected, the latter of which results from disorder, allowing for
structural characterization without assuming periodicity [4–6, 8]. The PDF is thus a powerful tool for quantitatively
characterizing short-range and long-range atomic structure [6].
The PDF, denoted by G(r), provides the scaled probability of finding two atoms a distance of rapart [2]. An
experimentally-measured PDF can be analyzed by adjusting the parameters of an assumed structure model, such as
the lattice constants, atomic positions, and grain/particle size. The structure is refined in real-space by minimizing
the difference between its PDF and the experimental PDF [1, 2]. The refinement process to perform this analysis has
been implemented in PDFgui [9], DiffPy-cmi [10], and TOPAS [11].
A major challenge in PDF modeling is the selection of the starting atomic structure [1, 2, 12]. Significant information
about the sample (e.g., the crystal phase) is required to achieve satisfactory results. Typically, PDF modeling involves
manual trial-and-error refinement of multiple structure models [12, 13]. There have been attempts to automate and
accelerate this process, where a large number of structures are either pulled from a materials database or generated
automatically [2, 13–15]. Such approaches can be efficient for the identification of the crystal structure of an unknown
material. They may not be sufficient, however, when the material and phase are already known and detailed structural
information is required. For example, if the sample is disordered and/or a nanoparticle, where there can be atomic
displacements away from the perfect bulk structure due to defects and/or surfaces [7]. Since the crystal structures
available in most databases are for perfect bulk phases [16–18], a method for generating candidate structures that
takes into account the atomic displacements induced by disorder and/or surfaces is needed.
Atomistic simulations provide a solution to this challenge. While density functional theory (DFT) can perform first
principles-based energy calculations, it is limited to small systems with minimal complexity by its large computational
cost. Empirical interatomic potentials, on the other hand, can efficiently provide the total energy of a large, complex
system as an explicit function of its atomic coordinates [19]. A well-parameterized potential can maintain high
accuracy when compared to a DFT calculation.
In minimizing the energy of a system, the atomic positions will change to account for perturbations to the perfect
structure (e.g., defects). Structure models generated via energy minimization with an interatomic potential can
mcgaughey@cmu.edu.
arXiv:2210.05890v1 [cond-mat.mtrl-sci] 12 Oct 2022
2
therefore be used to represent complex systems that are not available in materials databases. Furthermore, an
energy minimization-based approach is advantageous compared to the reverse Monte Carlo (RMC) method. In RMC,
atoms are allowed to move under certain constraints (e.g., bond angles, bond lengths) until the difference between the
calculated and experimental structural characteristic of interest (e.g, the PDF) is minimized [20]. Incorrect structures,
however, can be generated if the appropriate combinations of such constraints are not applied [21].
Defect-induced changes in the local atomic environment can be described by atomistic calculations if the crystal
phase and defect information (e.g., types, concentrations) are specified. Herein, we describe a PDF analysis workflow
for modeling defected crystals. Structures with different types and concentrations of defects are first created and then
relaxed using energy minimization. The PDF of each relaxed structure is then refined against the experimental PDF.
The defect types and concentrations are then estimated from the structures that yield refined PDFs with the smallest
differences compared to the experimental PDF.
Measured PDFs of anatase titanium dioxide (TiO2) and tetragonal zirconium dioxide (ZrO2) grown using low
temperature solution synthesis in the presence of microwave radiation (MWR) are used to demonstrate this method.
External fields can induce defects that alter the diffusion of space charges across grain boundaries and ultimately lead
to efficient processing [22–25]. Our goal is to identify the types and concentrations of point defects that are present
in MWR-grown anatase TiO2and tetragonal ZrO2. While we focus on structures with local disorder, this workflow
can also benefit PDF analysis of nanoparticles [26] and perovskites [27], which are challenging to characterize using
conventional methods.
The rest of the paper is organized as follows. In Secs. II A and II B, the atomistic calculations, structure generation,
and PDF analysis are described. In Sec. III A, the interatomic potential selection process is presented, which is
guided by predictions of lattice constants and elastic constants. In Sec. III B, defect formation energy predictions are
compared with literature values. The PDF analysis is presented in Sec. IV. We find that Ti vacancy and Ti interstitial
are the dominant defects in MWR-grown anatase TiO2and that O vacancy is the dominant defect in MWR-grown
tetragonal ZrO2.
II. METHODS
A. Atomistic calculations and structure model generation
The nature of the bonding in TiO2and ZrO2, which have relatively strong covalent characteristics, makes it
challenging to model the atomic interactions using an interatomic potential. We also require that the potential be able
to handle heterogeneous, defected systems. Atomistic calculations of materials with electrostatic interactions typically
maintain fixed charges on the ions. Charge transfer may happen around a defect site, however, causing deviations
of the electrostatic energies and particle positions [28, 29]. A variable-charge potential allows for the redistribution
of charges based on the local environment [30–32]. We identified two candidate variable charge potentials: second
moment tight-binding charge equilibrium (SMTB-Q) [33, 34] and reactive force field (ReaxFF) [35–37], which are
evaluated in Sec. III A.
All atomistic calculations are performed using the Large-scale Atomic/Molecular Massively Parallel Simulator
(LAMMPS) [38]. The space groups, supercell sizes, and number of atoms in the simulation boxes of anatase and
rutile TiO2, and tetragonal, cubic, and monoclinic ZrO2are listed in Table I. While we are focused on anatase TiO2
and tetragonal ZrO2, the other phases were included to best assess the potentials. Periodic boundary conditions
are applied in all three directions. The long-range electrostatic interaction is calculated using the Wolf summation
method [33, 39]. The time step for the molecular dynamics (MD) simulations is 0.2 fs. The charge on each atom is
updated every time step using the charge equilibrium (Qeq) scheme developed by Rapp´e and Goddard [40].
We built structures with different types and concentrations of randomly-placed point defects and relaxed them at
zero temperature by energy minimization. To investigate the sensitivity of the refined PDF to the defect locations,
ten structures were created for each configuration. We consider common point defects for both the cation (Ti and
Zr) and the anion (O): vacancies, interstitials, Frenkel pairs (a cation vacancy and interstitial pair), and anti-Frenkel
pairs (an anion vacancy and interstitial pair), and combinations of these defect types.
B. Pair distribution function (PDF) analysis
We use measured PDFs for MWR-grown anatase TiO2and tetragonal ZrO2thin films from previous studies [6, 25].
The data acquisition was performed at the X-ray Powder Diffraction Beamline, 28-ID-2, at the National Synchrotron
Light Source II at Brookhaven National Laboratory.
3
TABLE I. Space groups, supercell sizes, and number of atoms in the TiO2and ZrO2simulation boxes. a,b, and care the
lattice constants.
Phase Space Group Supercell Number of Atoms
Anatase TiO2I41/amd 7a×7a×3c1764
Rutile TiO2P42/mnm 6a×6a×9c1944
Tetragonal ZrO2P42/nmc 7a×7a×5c1470
Cubic ZrO2F m3m5a×5a×5a1500
Monoclinic ZrO2P21/c 5a×5b×5c1500
The PDF refinement is performed using the DiffPy-cmi package [10]. The structure models are fit against the
measured PDFs by refining global and phase-specific scaling factors, lattice parameters, isotropic atomic displacement
parameters (ADPs, denoted by Uiso), a low-rpeak sharpening coefficient for correlated motion of nearby atoms, and
a PDF peak envelope function, which dampens the signal as a function of separation to account for grain/particle
size. The quality of the fit is quantified by a goodness-of-fit value Rw, which is calculated as [4]
Rw=sPn(Gobs,n Gcalc,n)2
PnG2
obs,n
,(1)
where Gobs,n is the nth point on the experimentally measured PDF and Gcalc,n is the nth point on the refined PDF.
The PDF refinement is done from 1.5 ˚
A to 30 ˚
A in an increment of 0.01 ˚
A. Periodic boundary conditions are applied
to the structure models [42]. This step is important because if the supercell is not periodic, the PDF peaks will
dampen as the atomic separation approaches the simulation box size.
III. POTENTIAL SELECTION AND ASSESSMENT
A. Lattice constants and elastic constants
We performed atomistic calculations using SMTB-Q and ReaxFF for both TiO2and ZrO2. The SMTB-Q po-
tential [34] and two ReaxFF parametrizations, i.e., ReaxFF – TiO2/H2O [36] and ReaxFF – defect [37] for TiO2
all produce stable structures in MD simulations and will be further considered in the lattice constant and elastic
constant calculations. The ReaxFF parametrized for yttria-stabilized zirconia [35], however, does not generate stable
structures in MD simualtions for the tetragonal, cubic, and monoclinic phases of ZrO2. The SMTB-Q potential for
ZrO2[33] does generate stable structures for perfect and defected phases and will be used in the subsequent analysis.
The lattice constants and elastic constants of anatase TiO2calculated using the SMTB-Q and ReaxFF are presented
in Table II along with experimental values from the literature [43, 44]. The calculation details are provided in
Appendix A. The lattice parameters predicted by all three potentials are close to the experimental values with a
largest deviation of 1.7%. The accuracy of the elastic constants is quantified by the root-mean-square error (RMSE)
of the unique elements of the elastic constant tensor compared to the experimental values. The elastic constant RMSE
from the SMTB-Q potential is 17.9 GPa, thirteen and five times smaller than those predicted by ReaxFF – TiO2/H2O
and ReaxFF – defect. In addition, the elastic constants calculated by SMTB-Q maintain the crystal symmetry (i.e.,
C11 =C22,C13 =C23, and C44 =C55), which is not the case with those calculated by ReaxFF – TiO2/H2O and
ReaxFF – defect. Based on the lattice constant and elastic constant results, we decided to use the SMTB-Q potential
in our further calculations of TiO2. The lattice constants and elastic constants of rutile TiO2are provided in Table
S1 of the Supplemental Material [45].
The lattice constants and elastic constants of tetragonal ZrO2calculated using the SMTB-Q potential are also
reported in Table II. Although the ZrO2elastic constants have a larger RMSE compared to that for TiO2, the lattice
constants are well predicted with an average error of 2.4% compared to the experimental values [46, 47]. The elastic
constants that represent the reaction to a normal stress (i.e., C11,C22, and C33) are reasonably accurate, with an
average error of 18%. The lattice constants and elastic constants of monoclinic and cubic ZrO2are provided in Table
S2 of the Supplemental Material [45].
4
TABLE II. Lattice constants (a,c) (˚
A), elastic constants (cij ) (GPa), bulk modulus (BV) (GPa), and shear modulus (GV)
(GPa) calculated using the SMTB-Q and ReaxFF potentials, and literature experimental values.
Anatase TiO2Tetragonal ZrO2
SMTB-Q ReaxFF-TiO2/H2O ReaxFF-defect Exp. [43, 44] SMTB-Q Exp. [46, 47]
a3.809 3.798 3.832 3.784 3.681 3.591
c9.597 9.550 9.673 9.515 5.285 5.169
c11 341 401 402 337 294 327
c22 341 401 412 337 294 327
c33 198 211 245 192 178 264
c12 104 87 101 139 191 100
c13 103 115 145 136 16.5 62
c23 103 113 155 136 16.5 62
c44 56 -675 -238 54 41.7 59
c55 56 -58 90 54 41.7 59
c66 63 35 53 60 181 64
cij RMSE 17.9 249 106 - 190 -
BV
a167 182 207 178 135 152
GV
b58 -32 -32 57 89 83
aEq. (A3)
bEq. (A4)
B. Defect formation energy
The relaxed point defect structures predicted by SMTB-Q in anatase TiO2and tetragonal ZrO2, the phases that are
grown under MWR, are shown in Figs. 1(a)–1(e) and 2(a)–2(e). The calculation details are presented in Appendix A.
The resulting atomic displacements around the point defects in TiO2(denoted by arrows) are similar to those observed
in previous DFT calculations [48] and range from 0.1 ˚
A (O interstitial) to 0.4 ˚
A (Ti vacancy). For ZrO2, the atomic
displacements around a point defect are less than 0.0001 ˚
A and are not shown.
FIG. 1. Relaxed atomic structures of anatase TiO2from the SMTB-Q potential: (a) perfect, (b) Ti interstitial, (c) Ti
vacancy, (d) O interstitial, and (e) O vacancy. The dashed circles indicate the defect sites. The arrows indicate the direction
of displacements after relaxation. The structures are qualitatively similar to DFT-predicted structures [48].
The anatase TiO2defect formation energies predicted by SMTB-Q and from previous DFT calculations are provided
in Table III. Also provided are the tetragonal ZrO2defect formation energies predicted by SMTB-Q, previous DFT
calculations, and previous calculations for the polarizable Born-Mayer (BM) core-shell potential [49]. In the DFT
studies, the defect formation energies depend on the charge states, the elemental chemical potentials, the Fermi level,
and the valence-band maximum of the perfect crystal. As such, several discrete values are often reported for a given
5
FIG. 2. Relaxed atomic structures of tetragonal ZrO2from the SMTB-Q potential: (a) perfect, (b) Zr interstitial, (c) Zr
vacancy, (d) O interstitial, and (e) O vacancy. The dashed circles indicate the defect sites. The atomic displacements around
a point defect are less than 0.0001 ˚
A and are not shown.
TABLE III. Defect formation energies in eV for anatase TiO2and tetragonal ZrO2from the SMTB-Q potential, and previous
DFT and BM potential calculations.
Anatase TiO2Tetragonal ZrO2
SMTB-Q DFT SMTB-Q DFT [50] BM [49]
O Vacancy 1.85 0.70–4.23 [51] 4.42 -0.76 – 6.10 15.62
Cation Vacancy 2.93 -0.7 – 5.80 [48] 9.63 6.11 85.04
O Interstitial 4.37 0.60 – 4.30 [48] 7.74 1.79 -10.42
Cation Interstitial 1.07 0.70 – 7.74 [51] 1.85 1.96 -67.41
O Frenkel Pair 3.86 11.39 8.07 – 18.57 17.63
Cation Frenkel Pair 5.38 12.12 4.11 7.41 5.20
defect type. The full range is provided in Table III. The defect formation energies for other phases of TiO2and ZrO2
are provided in Tables S3 and S4 of the Supplemental Material [45].
For anatase TiO2, the SMTB-Q defect formation energies fall within the DFT energy ranges except for the O
interstitial (0.07 eV above the maximum DFT value). SMTB-Q predicts that the smallest formation energy is for a
Zr interstitial. For tetragonal ZrO2, the SMTB-Q defect formation energies fall close to the DFT results except for the
O interstitial. All SMTB-Q results are in better agreement with the DFT results than the BM potential. SMTB-Q
predicts that the smallest formation energy is for a Ti interstitial.
IV. PAIR DISTRIBUTION FUNCTION (PDF) ANALYSIS
A. Anatase TiO2
We now perform PDF analysis with starting structures obtained as described in Sec. II A to determine the defect
types (vacancies, interstitials, Frenkel pairs, anti-Frenkel pairs) and concentrations present in MWR-grown anatase
TiO2(this section) and tetragonal ZrO2(Sec. IV B). The variation of Rwwith defect concentration for different defects
in anatase TiO2is plotted in Fig. 3. The concentration of a defect type is specified by dividing the number of that
type of defect by the total number of atoms before defects are inserted. The mean values and standard deviations (as
shown by the error bars) are obtained from ten randomly generated structures for each defect type and concentration.
For all six defect types, Rwfirst decreases starting from 0.325 (the perfect structure, plotted as a dashed horizontal
line), reaches a minimum, and then increases as the defect concentration varies from 0 to 4.5%. Introducing Ti
defects improves the PDF refinement more than O defects as indicated by the lower Rwacross all concentrations.
The small error bars indicate that Rwis insensitive to the defect configuration except at the higher concentrations,
when clustering is possible. The lowest Rw(0.274) comes from the structure with 2.3% Ti Frenkel pairs.
Based on this result, we conducted a more comprehensive study. The concentrations of the four types of point
defect (Ti vacancy, Ti interstitial, O vacancy, and O interstitial) were simultaneously varied and the PDFs of the
摘要:

Pairdistributionfunctionanalysisdrivenbyatomisticsimulations:ApplicationtomicrowaveradiationsynthesizedTiO2andZrO2ShuyanZhang1,JieGong1,DanielXiao2,B.ReejaJayan1,andAlanJ.H.McGaughey11DepartmentofMechanicalEngineering,CarnegieMellonUniversity,Pittsburgh,Pennsylvania15213,USAand2DepartmentofMaterial...

展开>> 收起<<
Pair distribution function analysis driven by atomistic simulations Application to microwave radiation synthesized TiO 2and ZrO 2 Shuyan Zhang1 Jie Gong1 Daniel Xiao2 B. Reeja Jayan1 and Alan J. H. McGaughey1.pdf

共33页,预览5页

还剩页未读, 继续阅读

声明:本站为文档C2C交易模式,即用户上传的文档直接被用户下载,本站只是中间服务平台,本站所有文档下载所得的收益归上传人(含作者)所有。玖贝云文库仅提供信息存储空间,仅对用户上传内容的表现方式做保护处理,对上载内容本身不做任何修改或编辑。若文档所含内容侵犯了您的版权或隐私,请立即通知玖贝云文库,我们立即给予删除!
分类:图书资源 价格:10玖币 属性:33 页 大小:5.46MB 格式:PDF 时间:2025-05-06

开通VIP享超值会员特权

  • 多端同步记录
  • 高速下载文档
  • 免费文档工具
  • 分享文档赚钱
  • 每日登录抽奖
  • 优质衍生服务
/ 33
客服
关注