Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

Subscriber access provided by the University of Exeter

Article
Local Distortions and Dynamics in Hydrated Y-Doped BaZrO
3

Amangeldi Torayev, Luke Sperrin, Maria A. Gomez, John A. Kattirtzi, Céline Merlet, and Clare P. Grey
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.0c04594 • Publication Date (Web): 23 Jun 2020
Downloaded from pubs.acs.org on July 1, 2020

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 37 The Journal of Physical Chemistry

1
2
3
4
5
6
7 Local Distortions and Dynamics in Hydrated Y-doped BaZrO3
8
9
10
11 Amangeldi Torayev 1, Luke Sperrin 1, Maria A. Gomez 2, John A. Kattirtzi 1, Céline Merlet* 1,3,4,
12
13 Clare P. Grey* 1
14
15 1
Department of Chemistry, University of Cambridge, Lensfield Road, Cambridge CB2 1EW,
16
17 United Kingdom
18
19
2
20 Department of Chemistry, Mount Holyoke College, South Hadley, MA 01075, United States
21
22 3
23
CIRIMAT, Université de Toulouse, CNRS, Université Toulouse 3 - Paul Sabatier, 118 Route
24 de Narbonne, 31062 Toulouse cedex 9 - France
25
26
4
27 Réseau sur le Stockage Electrochimique de l'Energie (RS2E), FR CNRS 3459, HUB de
28
29 l’Energie, Rue Baudelocque, 80039 Amiens, France
30
31
Corresponding Authors: merlet@chimie.ups-tlse.fr, cpg27@cam.ac.uk
32
33
34
35 ABSTRACT
36
37
38
39 Y-doped BaZrO3 is a promising proton conductor for intermediate temperature solid oxide fuel
40
41 cells. In this work, a combination of static DFT calculations and DFT based molecular dynamics
42
43
44 (DFT-MD) was used to study proton conduction in such a material. Geometry optimisations of
45
46 100 structures with a 12.5% dopant concentration allowed us to identify a clear correlation
47
48 between the bending of the metal-oxygen-metal angle and the energies of the simulated cells.
49
50
51
Depending on the type of bending, two configurations, designated as inwards bending and
52
53 outwards bending, were defined. The results demonstrate that a larger bending decreases the
54
55 energy and that the lowest energies are observed for structures combining inwards bending with
56
57
58
59
60 ACS Paragon Plus Environment
1
The Journal of Physical Chemistry Page 2 of 37

1
2
3 protons being close to the dopant atoms. These lowest energy structures are the ones with the
4
5
6 strongest hydrogen bonds. DFT-MD simulations in cells with different yttrium distributions
7
8 provide complementary microscopic information on proton diffusion as they capture the
9
10 dynamic distortions of the lattice caused by thermal motion. A careful analysis of the proton
11
12
13
jumps between different environments confirmed that the inwards and outwards bending states
14
15 are relevant for the understanding of proton diffusion. Indeed, intra-octahedral jumps were
16
17 shown to only occur starting from an outwards configuration while the inwards configuration
18
19
seems to favor rotations around the oxygen. On average, in the DFT-MD simulations, the
20
21
22 hydrogen bond lengths are shorter for the outwards configuration which facilitates the intra-
23
24 octahedral jumps. Diffusion coefficients and activation energies were also determined and
25
26 compared to previous theoretical and experimental data showing a good agreement with previous
27
28
29 data corresponding to local proton motion.
30
31
32
TOC GRAPHICS
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52 KEYWORDS
53
54
55 proton dynamics, solid oxide fuel cells, DFT-MD, lattice distortions, perovskite
56
57
58
59
60 ACS Paragon Plus Environment
2
Page 3 of 37 The Journal of Physical Chemistry

1
2
3 1. Introduction
4
5
6
7 Perovskites have emerged as attractive electrolytes for intermediate-temperature Solid
8
9 Oxide Fuel Cells (SOFCs) owing to their good ionic transport properties as well as adequate
10
11 mechanical and chemical stabilities.1 In particular, Y-doped BaZrO3 has been widely studied as
12
13
14
this material shows very high protonic conductivities of up to 1x10-2 S cm-1 at 450°C when
15
16 exposed to water vapor.2,3 When BaZrO3 is doped with yttrium, the charge imbalance between
17
18 the yttrium and zirconium ions is accommodated by the generation of oxygen vacancies leading
19
20
to the composition BaZr1-xYxO3-x/2. These oxygen vacancies are responsible for the moderate
21
22
23 oxygen-anion conductivity at high temperatures.3 Under wet conditions, protons can be
24
25 incorporated in the material through absorption of water molecules that dissociate forming
26
27 hydroxide ions and filling the oxygen vacancies. The material can be fully hydrated, represented
28
29
30 by the chemical formula BaZr1-xYxO3-x(OH)x, or partially hydrated represented more generally as
31
32 BaZr1-xYxO3-x/2-y/2(OH)y (y < x) with oxygen vacancies remaining. The high ionic conductivities
33
34 achieved in wet conditions are promising for applications of this material in intermediate-
35
36
37
temperature SOFCs.3
38
39
40 Y-substituted BaZrO3 can be prepared with high concentrations of yttrium (up to 60%4)
41
42 and a number of studies4,5 have reported a dependence of the ionic conductivity on the yttrium
43
44
content. As a first guess, one would expect that the ionic conductivity increases with the doping
45
46
47 concentration following the increase in the number of charge carriers. However, it was shown
48
49 that an optimum conductivity is observed for a dopant concentration of around 10%-20%.4,5 In
50
51 fact, many experimental and theoretical observations suggest that the protons are trapped close to
52
53
54 the yttrium dopant sites increasing the activation barrier for diffusion and reducing the overall
55
56
57
58
59
60 ACS Paragon Plus Environment
3
The Journal of Physical Chemistry Page 4 of 37

1
2
3 ionic conductivity.6–8 This proposed proton trapping has attracted considerable attention in recent
4
5
6 years both from an experimental and from a theoretical point of view.8–17
7
8
9 In hydrated Y-doped BaZrO3, proton conduction occurs through series of transfers
10
11 (between neighboring oxygens) and rotations (around an oxygen). Static density functional
12
13
14
theory calculations (DFT) and molecular dynamics simulations (based on DFT or reactive force
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Figure 1. (a) Four possible proton-yttrium (H-Y) configurations in 2x2x2 supercells with a
37
38 12.5% dopant concentration as defined by Blanc et al.14 Barium atoms in the A sites of the
39
40
perovskite structure (at the centre of the Zr/Y-O cubes) are not shown for clarity. (b) Relative
41
42
43 energies of these cells with respect to the lowest energy structure after geometry optimization,
44
45 extracted from the data of Blanc et al.14
46
47
48 fields) have been proposed as powerful methods to estimate energy barriers for the intra- and
49
50
51 inter-octahedral transfers as well as for rotations. Static DFT calculations have shown that local
52
53 distortions arise close to the dopant atom, usually leading to stronger hydrogen bonds, which
54
55 could explain the trapping effect.10–13 However, local distortions also induce shorter O-O
56
57
58
59
60 ACS Paragon Plus Environment
4
Page 5 of 37 The Journal of Physical Chemistry

1
2
3 distances, which would tend to decrease the activation energies for proton jumping. It is worth
4
5
6 noting that trapping does not necessarily imply that jumps do not happen but rather that protons
7
8 spend a significant amount of time in the vicinity of the dopant and that the jumps are local
9
10 rather than involving long-range transport. In many studies, local activation energies are
11
12
13
combined with a jump-diffusion model to predict long-range energy barriers and diffusion
14
15 coefficients. The resulting activation energies are usually in good agreement with experiments
16
17 and trends observed for different dopant atoms could be reproduced by Björketun et al.11 In this
18
19
context, Blanc et al. have defined specific proton environments in the perovskite material
20
21
22 depending on the position of the proton with respect to the dopant atom (see Figure 1)14. In their
23
24 DFT calculations, the configurations with protons close to the yttrium atoms indeed correspond
25
26 to the lowest energy structures.
27
28
29
30 While static DFT calculations are very useful to identify proton diffusion pathways and
31
32 the effect of specific local distortions, thermal fluctuations are always present at the temperatures
33
34 used in the experiments (200°C - 700°C). These dynamic distortions may alter the picture
35
36
37
obtained from a static point of view. DFT-MD simulations have the advantage of being able to
38
39 capture both proton hopping and thermal fluctuations that occur on short time scales as
40
41 demonstrated by many studies for water in the bulk liquid phase and at interfaces.19–21 More
42
43
recently, Fronzi et al. applied DFT-MD simulations on proton conduction in undoped BaZrO3
44
45
46 and reported that proton diffusion is enhanced when compressive strain is applied on the
47
48 structure.22 The authors interpret this increase in diffusion to the shorter O-O distances observed
49
50 in the compressed material. The effect of strain on proton diffusion has also been demonstrated
51
52
53 in Y-doped BaZrO3 for two dopant concentrations (1% and 12.5%) using MD simulations with a
54
55 reactive force field based on an Empirical Valence Bond description.23 In these dynamical
56
57
58
59
60 ACS Paragon Plus Environment
5
The Journal of Physical Chemistry Page 6 of 37

1
2
3 studies and other related ones,15–17 no strong trapping is however observed. Raiteri et al. suggest
4
5
6 that for low dopant amounts (less than 15%), the trapping probability is low and thus not often
7
8 observed, or hard to characterize, from MD simulations.16 As such, there still remains a need to
9
10 develop a clear understanding of the trapping effect and the effect of lattice dynamics on
11
12
13
mobility at intermediate-to-high temperatures – the temperatures relevant to the operation of the
14
15 device.
16
17
18 In this work, we combine static DFT calculations and ab initio molecular dynamics to
19
20
obtain insights into proton diffusion and trapping in fully hydrated Y-doped BaZrO3. Static DFT
21
22
23 calculations are performed on structures with two different Y-dopant concentrations (12.5% and
24
25 25%) and are used to determine the energies of a large number of configurations and correlate
26
27 energies with distortions of the lattice. The effect of the relative position of yttrium atoms is
28
29
30 explored by using a variety of structures in both static and DFT-MD simulations; we assume that
31
32 no oxygen vacancies are present to simplify the analysis. Ab initio molecular dynamics
33
34 simulations are then conducted on structures with a 12.5% dopant concentration at three
35
36
37
temperatures. From the generated trajectories, the proton diffusion was analyzed through mean
38
39 square displacements as well as via the numbers of proton jumps between different
40
41 environments. The results show a relationship between specific local geometries and the
42
43
occurrence of proton jumps. In particular, intra-octahedral jumps are only seen when short
44
45
46 hydrogen bond lengths are formed, more common at high temperatures. Activation energies
47
48 calculated are in good agreement with previously reported data corresponding to local proton
49
50 motion.
51
52
53
54 2. Computational methods
55
56
57
58
59
60 ACS Paragon Plus Environment
6
Page 7 of 37 The Journal of Physical Chemistry

1
2
3 2.1 Static DFT calculations
4
5
6
7 For the static DFT study, results reported in the literature12,14 for a dopant concentration of
8
9 12.5% were combined with results obtained in this work for a dopant concentration of 25%. All
10
11 geometry optimizations were achieved using the PBE functional24,25 and were done with a 2x2x2
12
13
14
supercell having the generic formula BaZr1-xYxO3-x(OH)x. The supercells contain 41 atoms for
15
16 the 12.5% dopant concentration and 42 atoms for the 25% dopant concentration, i.e., the material
17
18 is considered fully protonated and there are no oxygen vacancies. Four of the structures with a
19
20
12.5% dopant concentration and all 26 structures with a 25% dopant concentration were
21
22
23 optimized using CASTEP26, a basis set cut-off of 680 eV, an energy convergence criterion of
24
25 10-6 eV (as in Ref.14) and a 3x3x3 Monkhorst-Pack27 k-point mesh. The 96 other structures with
26
27 a 12.5% dopant concentration have been reported in a previous work12 and were generated, using
28
29
30 VASP,28,29 as follows. The positions of all atoms in an initial 2x2x2 supercell with an yttrium
31
32 dopant, without any protons, are relaxed. To compensate the charge imbalance, VASP
33
34 automatically uses a uniform positive background gas. A proton is then placed in all possible
35
36
37
accommodating sites (24 oxygens x 4 possible orientations),12 and its position is relaxed while
38
39 keeping all the other atoms fixed. For these 96 proton binding structures, started from the lowest
40
41 energy Glazer distortion, the geometry optimizations were done using the Generalized Gradient
42
43
Approximation (GGA) PBE functional. A basis set cut-off of 600 eV, an electronic energy
44
45
46 convergence criterion of 10-4 eV and a 2x2x2 Monkhorst–Pack k-point mesh were used. The
47
48 geometry optimization energy convergence criterion was set to 10-3 eV. In the end, a total of 100
49
50 structures with a 12.5% dopant concentration and 26 structures with a 25% dopant concentration
51
52
53 was obtained. VESTA software was used for plotting all the structures shown in this article.
54
55
56 2.2 DFT-MD calculations
57
58
59
60 ACS Paragon Plus Environment
7
The Journal of Physical Chemistry Page 8 of 37

1
2
3 For the DFT-MD study, larger 4x4x4 supercells containing a total of 328 atoms
4
5
6 (including 8 protons) and having a single dopant concentration of 12.5% were simulated. All the
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 Figure 2. Starting configurations used for the DFT-MD simulations performed on cells
40
41
42
containing 12.5% yttrium. (a) YU: Yttrium dopant atoms are distributed uniformly in the 4x4x4
43
44 supercell. (b) YC: Some of the yttrium dopant atoms are nearest neighbors (clustered). (c) YP:
45
46 Some of the yttrium dopant atoms are in the planar configuration. In all cases, there are 8 protons
47
48 with 2 of them arranged in each H-Y configuration (near, planar, perpendicular, distant).
49
50
51
52 calculations were done with the CP2K software30 using the PBE functional and run in the NPT
53
54 ensemble under atmospheric pressure. The cell is cubic with an initial lattice size of 17.04 Å.
55
56
57
58
59
60 ACS Paragon Plus Environment
8
Page 9 of 37 The Journal of Physical Chemistry

1
2
3 The Gaussian Plane Wave (GPW) implementation uses a Gaussian DZVP basis set with GTH
4
5
6 pseudopotentials,31–33 a 400 eV plane wave cut-off, with only the Gamma point, and a 6x10-7 eV
7
8 electronic energy convergence criteria. Three different distributions of the yttrium dopant atoms
9
10 were considered: a uniform distribution (noted YU), a clustered configuration where two pairs of
11
12
13
yttrium atoms are nearest neighbors (noted YC) and a planar configuration where two pairs of
14
15 yttrium atoms are next nearest neighbors (noted YP). This is illustrated in Figure 2, which shows
16
17 the initial configurations for the DFT-MD simulations. To obtain the starting configurations,
18
19
8 protons were distributed in the system so that there are 2 protons in each H-Y configuration as
20
21
22 defined in Figure 1 (near, planar, perpendicular, distant)14. Three temperatures (500 K, 750 K
23
24 and 1000 K) were investigated in order to estimate the activation energies related to proton
25
26 diffusion. All the DFT-MD simulations were run for 20 ps with a 0.5 fs timestep.
27
28
29
30 3. Results and discussion
31
32
33 3.1 Static DFT calculations
34
35
36 As discussed in the introduction, the relative positions of the protons and yttrium atoms,
37
38
39
i.e., the H-Y configuration, have a large influence on the energy of the 12.5% dopant
40
41 concentration cells. To investigate the impact of the relative positions of the yttrium atoms, i.e.,
42
43 the Y-Y configuration, the configurational energies and cell geometries of 26 structures with a
44
45
dopant concentration of 25% are first analyzed. Figure 3 shows the relative energies of these
46
47
48 structures with respect to the lowest energy structure and illustrates the Y-Y configurations.
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
9
The Journal of Physical Chemistry Page 10 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure 3. (a) Relative configurational energies with respect to the lowest energy structure for
26
27
26 cells having a dopant concentration of 25% and different Y-Y configurations. (b) Schematic
28
29
30 representation of the different Y-Y coordination shells (these structures are not optimised and
31
32 simply illustrate the relative positions of the yttrium atoms).
33
34
35 From Figure 3, it is clear that a larger range of energies is observed for 1st and 2nd shell than for
36
37
38 3rd shell configurations. We also observe that the lowest energies are obtained for protons nearby
39
40 Y-Y in 1st and 2nd shell configurations. We note, however, that the Y-Y arrangements are
41
42 determined during the high temperature synthesis of the material to form BaZr1-xYxO3-x:
43
44
45
hydration is unlikely to change the Y-Y configurations at least at moderate temperatures. Due to
46
47 the high temperatures used to prepare these materials, yttrium atoms are likely to be randomly
48
49 distributed in the lattice as suggested by previous researchers.1,34 Different syntheses routes may
50
51
lead to different materials properties which could perhaps be related to different dopant spatial
52
53
54 distributions.35,36
55
56
57
58
59
60 ACS Paragon Plus Environment
10
Page 11 of 37 The Journal of Physical Chemistry

1
2
3 To characterize the lattice distortions due to the presence of an yttrium atom, the metal-
4
5
6 oxygen (Zr-O, Y-O) and metal-metal (Zr-B, Y-B) interatomic distances for different dopant
7
8 concentrations (B = Zr, Y) are examined. Figures 4.a and 4.b show a comparison of the average
9
10 interatomic distances calculated from the static DFT calculations and the experimental values32.
11
12
13
The metal-oxygen and metal-metal distances obtained from DFT are in good agreement with
14
15 experiments although slightly overestimated (by around 1-5%). These results are consistent with
16
17 other simulation results published on similar materials37 and show that the PBE functional is
18
19
suitable for such a study. From Figures 4.a and 4.b, it is clear that while the average metal-
20
21
22 oxygen distance is larger for yttrium-oxygen than for zirconium-oxygen, the average metal-metal
23
24 interatomic distances are relatively unaffected by the presence of yttrium. This suggests that
25
26 there is a bending of the B-O-B units, consistent with the octahedral tilting reported
27
28
29 previously.37,38 Figure 4.c shows the interatomic distances for different environments around the
30
31 oxygen in the presence or absence of protons. While the presence of a single yttrium atom only
32
33 leads to a limited variation in the metal-metal distance, the presence of two yttrium atoms results
34
35
36
in a noticeable decrease in distance. We also note that the metal-metal distances for atoms
37
38 surrounding an OH unit are systematically larger than in the absence of proton. This could be
39
40 due to steric or electrostatic repulsions between the proton and the adjacent Zr/Y atoms.
41
42
43
To investigate the effect of local distortions close to the H and Y atoms on the energy of
44
45
46 the cells, we focus on the 100 structures with a 12.5% dopant concentration that contain only one
47
48 Y and proton per supercell. (In the 25% doped structures, there are two protons and two yttrium
49
50 atoms, and it is thus difficult to separate the effects of the two H-Y configurations; furthermore
51
52
53 the relative energies of the different Y-Y configurations would need to be accounted for).
54
55
56
57
58
59
60 ACS Paragon Plus Environment
11
The Journal of Physical Chemistry Page 12 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Figure 4. (a) and (b) Metal-oxygen and metal-metal interatomic distances as a function of the
34
35 Y-dopant concentration calculated in this work from static DFT (12.5 and 25%) and determined
36
37
38 experimentally from X-ray diffraction and extended X-ray fine structure (EXAFS) analysis (0, 6
39
40 and 15%) by Giannici et al.37 (c) Metal-metal interatomic distances for specific metal-oxygen-
41
42 metal configurations obtained in this work.
43
44
45
46 To characterize the local distortions shown in Figure 5.a, we used a signed “B-OH-B bending
47
48 distance” which is the non-zero component of the vector from the center of the solid black line
49
50 connecting the two transition metals (Zr or Y) to the oxygen covalently bonded to the proton.
51
52
53
The 100 structures are classified into two categories depending on the sign of the B-OH-B
54
55 bending distance. If the B-OH-B bending distance is negative, the structure is assigned as “bent
56
57
58
59
60 ACS Paragon Plus Environment
12
Page 13 of 37 The Journal of Physical Chemistry

1
2
3 outwards” since the hydroxyl is outside of the interior BOB angle, the structure is otherwise
4
5
6 assigned as “bent inwards” when the hydroxyl is on the inside of the interior BOB angle.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Figure 5. (a) Illustration of inwards and outwards bending, showing how the bending distance is
37
38
39 quantified, a B-OH-B bending distance of 0 corresponding to colinear bonds (i.e., no bend).
40
41 (b) Relative energies of the 100 structures with a 12.5% dopant concentration classified
42
43 following their H-Y configuration and B-OH-B bending type. (c) Relative energies as a function
44
45
46 of the B-OH-B bending distance.
47
48
49 Figure 5.b shows a plot of the relative cell energies as a function of the H-Y
50
51 configuration. The different structures are colored according to their “bent inwards” or “bent
52
53
54
outwards” bending status. It is very clear from this figure that, for a given H-Y configuration, all
55
56 the structures which are “bent inwards” have lower energies than the “bent outwards”
57
58
59
60 ACS Paragon Plus Environment
13
The Journal of Physical Chemistry Page 14 of 37

1
2
3 configurations and are thus more favorable. To confirm this assumption, we plot the relative cell
4
5
6 energies as a function of the B-OH-B bending distance. This is shown in Figure 5.c where the
7
8 different structures are now colored according to their H-Y configuration. While the trend is not
9
10 perfect and while the energies of some structures overlap, there is a clear correlation between the
11
12
13
cell energy and the type (inwards/outwards) and amplitude of the bending. Independent of the
14
15 bending type, larger absolute distances (i.e., larger bendings) are correlated with lower energies.
16
17 We note that all the points corresponding to specific H-Y configurations are grouped and that the
18
19
lowest energies are observed for combined near/planar and bent inwards configurations. As far
20
21
22 as we know this correlation between specific distortions and cell energies has not been
23
24 highlighted before and confirms that the presence of yttrium dopant close to the protons can lead
25
26 to low energy configurations not accessible otherwise.
27
28
29
30 We now investigate what features makes the lowest energy structures more stable. While,
31
32 the proximity of protons, oxygen and yttrium atoms can lead to complex hydrogen bonding
33
34 situations, which can influence both the local structure and proton dynamics,12,39,40 here we focus
35
36
37
on the characterization of these hydrogen bonds to determine how important this is in
38
39 determining the calculated configurational energies. To describe the different environments, a
40
41 number of angles and distances, all defined in Figure 6.a, are used. dHO1 and dHO2 are the
42
43
distances between the proton and either the second or third nearest oxygen, respectively. dHO1 is
44
45
46 thus the traditional hydrogen bond length while dHO is the length of the covalent bond (distance
47
48 to the nearest oxygen). OHO1 corresponds to the hydrogen bond angle while OBO1 and OBO2
49
50 characterize the distortions around the metal atoms. Figures 6.b, 6.c, and 6d clearly show that
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
14
Page 15 of 37 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 6. (a) Schematic representation of the local geometry around a proton and labels adopted
39
40
to assess hydrogen bonding related geometries. (b) Hydrogen bond angle as a function of
41
42
43 hydrogen bond length (i.e., dHO1). (c) Distance between the proton and the 3rd nearest oxygen
44
45 (i.e., dHO2) as a function of the hydrogen bond length. (d) Metal-oxygen-metal angles around the
46
47 hydrogen bond. For (b), (c), (d) the structures are colored according to their relative energy with
48
49
50 respect to the lowest energy structure in black (energies are in eV cell-1).
51
52
53 the lowest energy configurations correspond to specific local environments where the hydrogen
54
55 bonds are the strongest. Indeed, Figure 6.b shows that all the lowest energy structures correspond
56
57
58
59
60 ACS Paragon Plus Environment
15
The Journal of Physical Chemistry Page 16 of 37

1
2
3 to hydrogen bond lengths smaller than 2.15Å and angles larger than 130°. Figure 6.c illustrates
4
5
6 the fact that all the lowest energy structures correspond to situations where one of the hydrogen
7
8 bond (HO1) is noticeably stronger than the other one (HO2). In these low energy environments,
9
10 dHO1 is clearly smaller than dHO2. Interestingly, these stronger hydrogen bonds are correlated
11
12
13
with less distorted OBO1 angles, i.e., with angles close to a value of 90 degrees which would be
14
15 seen in an ideal cubic structure (see Figure 6.d). It is worth noting that, for the remaining
16
17 configurations, a number of data points are close to the “y = x” line. In these cases, the proton is
18
19
engaged in two very similar hydrogen bonds. In the case of inwards bending, the formation of
20
21
22 hydrogen bonds is intuitively favorable as it is consistent with the octahedral tilting while in the
23
24 case of outwards bending, the formation of hydrogen bonds goes against the expected tilting.
25
26 This phenomenon is illustrated in Figure S1.
27
28
29
30 Overall, the static DFT calculations indicate that, in agreement with previous studies, the
31
32 presence of yttrium as a dopant can lead to the existence of low energy configurations which
33
34 could act as proton traps. These low energy structures are apparently stabilized by a strong
35
36
37
hydrogen bond in a locally distorted environment. We now turn to DFT-MD simulations to
38
39 investigate the existence and fate of such configurations in a dynamic system.
40
41
42 3.2 DFT-MD calculations
43
44
45
In order to obtain insight into the proton dynamics in Y-doped BaZrO3 materials, a
46
47
48 number of DFT-MD simulations were performed on 4x4x4 supercells having a single dopant
49
50 concentration of 12.5% at three temperatures (500 K, 750 K, and 1000 K). In the smaller 2x2x2
51
52 supercells used above for the DFT analysis of the static systems, there was only one Y-Y
53
54
55 arrangement, the different configurations arising from the different positions of the proton in the
56
57
58
59
60 ACS Paragon Plus Environment
16
Page 17 of 37 The Journal of Physical Chemistry

1
2
3 cell (Figure 1), however, the larger 4x4x4 cells require that we now consider the relative Y
4
5
6 positions (as was performed for the smaller cells containing 25% Y, Figure 3). Three
7
8 distributions of yttrium atoms were investigated (Figure 2). In one of the systems, the Y atoms
9
10 are distributed uniformly and quite far apart from each other (YU); while in the clustered system
11
12
13
(YC), some of the Y atoms are nearest neighbors; and in the last configuration, some of the Y
14
15 atoms are next nearest neighbors, i.e., in the planar configuration (YP). The initial structures are
16
17 generated so that the 8 protons are distributed equally in all H-Y configurations (2 protons in
18
19
near, 2 in planar, 2 in perpendicular and 2 in distant). The molecular dynamics simulations are
20
21
22 then conducted and the protons trajectories are recorded.
23
24
25 Figure 7.a provides an illustration of the environments visited by one of the protons along
26
27 a trajectory for the YU system at 1000 K. While the proton is originally in a planar configuration
28
29
30 at t = 0, it jumps between planar and near configurations ultimately visiting all possible H-Y
31
32 configurations along the 20 ps trajectory. The actual trajectories of the 8 protons present in the
33
34 same YU system at 1000 K are shown in Figure 7.b. The total residence time fractions of the
35
36
37
protons in the different H-Y configurations, i.e., the total time protons spend in a given
38
39 configuration divided by the total simulation time, for all systems are shown in Figure S2 (these
40
41 time fractions are normalized by the available sites in each system in Figure S3). They confirm
42
43
that all H-Y configurations are explored by protons. While the perpendicular and distant
44
45
46 configurations are generally slightly less favorable than the near and planar ones (Figure S2), the
47
48 differences between residences times between different sites are not as large as might be
49
50 expected based on the static DFT calculations (Figure 1). However, direct comparison with
51
52
53 predictions from static DFT energies is out of reach since these total residence time fractions are
54
55 not fully converged for the 20 ps trajectories conducted here (Figure S4). Note that the initial
56
57
58
59
60 ACS Paragon Plus Environment
17
The Journal of Physical Chemistry Page 18 of 37

1
2
3 configurations of the protons were generated to explore as many microstates as possible and
4
5
6 since all of the configurations are sampled on the timescales used here, the short times employed
7
8 in the MD runs should still allow comparisons between structural properties obtained in the static
9
10 calculations and by DFT-MD.
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Figure 7. (a) Illustration of the various environments visited by a single proton along a 20 ps
34
35 trajectory for the YU system simulated at 1000 K. The near configuration is split into two
36
37 environments depending on the proton having 1 or 2 yttrium atoms as nearest neighbors (“Near
38
39
40 Y-OH-Z” and “Near Y-OH-Y”). Note that in YU and YP, there are no Y-OH-Y configurations
41
42 so this environment cannot be visited in the trajectory shown here. (b) DFT-MD trajectories of
43
44 the 8 protons present, visualized using the OVITO program41. The proton whose trajectory is
45
46
47
analysed in (a) is indicated.
48
49
50 Since it is possible to distinguish the different proton environments with respect to the
51
52 yttrium atoms (near, planar, perpendicular, distant), we can calculate various properties
53
54 according to each H-Y configuration. Figure 8 shows the average residence time for each H-Y
55
56
57
58
59
60 ACS Paragon Plus Environment
18
Page 19 of 37 The Journal of Physical Chemistry

1
2
3 configuration, i.e., the average time that a proton stays in a given site before jumping to another
4
5
6 site. For most of the configurations, the average residence time decreases as expected when the
7
8 temperature increases. For the perpendicular configuration however, the average residence time
9
10 stays short even at 500 K. In fact, the average residence time in the perpendicular configuration
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 8. Average residence time of a proton in the various H-Y configurations.
42
43
44 is the shortest for all the systems considered. In the static DFT calculations, the configurational
45
46
47
energies for near and planar are low and close to each other, while the perpendicular and distant
48
49 configurations have similar and high energies. Following these results, longer residence times
50
51 could be expected for near and planar configurations, but this is not what is observed here. A low
52
53
energy barrier for proton transfers between near and planar might explain the relatively small
54
55
56 residence times determined for these H-Y configurations. In addition, fluctuations of the oxygen-
57
58
59
60 ACS Paragon Plus Environment
19
The Journal of Physical Chemistry Page 20 of 37

1
2
3 oxygen distances are probably an important component in defining the time between jumps, as
4
5
6 discussed below.
7
8
9 Proton transfers and rotations between environments are examined by counting how
10
11 many proton jumps occur from a given type of site to another. The total number of rotations and
12
13
14
intra-octahedral transfers are given in Table S1. It is worth noting that no inter-octahedral jump
15
16 is observed in any of the simulated systems. This is consistent with an activation barrier being
17
18 higher for this event compared to rotations and intra-octahedral jumps.10,38,42 There are
19
20
approximately 20 times more rotations than intra-octahedral transfers which is consistent with
21
22
23 activation barriers reported previously.10 When two yttrium atoms are nearest neighbors (YC),
24
25 there are much more near to near proton jumps. This could be because the local distortions
26
27 resulting from the proximity of two yttrium atoms reduces the oxygen-oxygen distances thereby
28
29
30 facilitating proton rotations and jumps as discussed in Ref.10,38 In addition, most of the jumps are
31
32 gathered on the diagonals, meaning that they do not change the type of H-Y configuration. The
33
34 exceptions are the rotations between perpendicular and planar and the intra-octahedral jumps
35
36
37
between planar and near.
38
39
40 To relate these jump events to local distortions, the jumps are now divided into jumps
41
42 from inwards and outwards configurations. The number of rotations classified in this way are
43
44
given in Table S2 and plotted in Figure 9 along with the fraction of inwards and outwards states
45
46
47 (note that while the trajectories are short, these fractions are quite well converged as can be seen
48
49 in Figure S5). This analysis shows that all intra-octahedral transfers occur when the protons are
50
51 in the outwards configuration. Figure 9 also shows that, in most cases, even though on average
52
53
54 there are more protons in the outwards states, more rotations originate from the bent inwards
55
56 configuration. This indicates that inwards configurations are more favorable for rotations. These
57
58
59
60 ACS Paragon Plus Environment
20
Page 21 of 37 The Journal of Physical Chemistry

1
2
3 considerations, combined with the lowest energies observed for near and planar H-Y
4
5
6 configurations with inwards bending indicate that the presence of yttrium atoms favor inwards
7
8 bending, and thus rotations, thereby reducing the occurrence of intra-octahedral jumps and thus
9
10 long-range diffusion. This could, at least partly, explain the proton trapping observed in Y-doped
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Figure 9: (a) Average fraction of inwards bending states observed in the different systems.
34
35
36 (b) Total number of rotations originating from inwards/outwards configurations.
37
38
39 hydrated BaZrO3. It is worth nothing that while there are fewer protons in inwards configurations
40
41 than in outwards configurations at the temperatures simulated here, this proportion decreases
42
43
44 with increasing temperature, in agreement with a lower energy for inwards configurations. A
45
46 rough estimate of the energy difference between inwards and outwards from their populations at
47
48 500 K and 1000 K indicates that the inwards configurations energies are around 8-10 kJ mol-1
49
50
51
lower than the outwards (corresponding to 0.10 eV cell-1 in units comparable to the static DFT
52
53 calculations reported in this work) which is consistent with the range of values observed in
54
55 Figure 5.c.
56
57
58
59
60 ACS Paragon Plus Environment
21
The Journal of Physical Chemistry Page 22 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 Figure 10. Several distances for different proton positions analyzed according to
40
41 inwards/outwards bending states in the YU system at 750 K: (a) distance between the proton and
42
43
its closest oxygen atom, (b) hydrogen bond length, (c) distance between two oxygen atoms
44
45
46 forming a covalent and a hydrogen bond to the same proton, (d) bending of B-O1-B angles.
47
48
49 We now try to correlate these results with a description of the hydrogen bonding in the
50
51 material. To begin with, the same geometric descriptors as in Figure 6 are used to see if major
52
53
54 differences between inwards and outwards configurations can be identified. The plots obtained
55
56 are shown in Figure S6. As expected, due to thermal motion, the clouds of points span a much
57
58
59
60 ACS Paragon Plus Environment
22
Page 23 of 37 The Journal of Physical Chemistry

1
2
3 wider area than in the case of the static DFT calculations. Moreover, the hydrogen bond lengths
4
5
6 and the OBO1 and OBO2 angles are on average a bit larger for the inwards configurations. To
7
8 obtain further insights into the difference between inwards and outwards configurations,
9
10 histograms of several descriptors of the hydrogen bonding environment are plotted; the
11
12
13
histograms were all checked to ensure that they are well converged over the simulated trajectory
14
15 for all systems. The histograms for the YU system simulated at 750 K are given in Figure 10.
16
17 Similar results are observed for the YC and YP systems at 750 K (Figure S7 and S8). Figure 10
18
19
shows that the covalent bond lengths are similar for inwards and outwards configurations and as
20
21
22 such cannot be used to interpret the proton jumps results. In contrast, inwards and outwards
23
24 configurations show some difference with regards to the hydrogen bond length and HO--O1
25
26 distance. On average the outwards configurations show shorter hydrogen bond lengths – in
27
28
29 contrast to what is observed in the static calculations – and shorter HO--O1 distances which
30
31 would tend to reduce the activation barriers for the jump to the H-bonded O1 site and thus
32
33 increase the jump rate. As such, intra-octahedral transfers are possible in the case of protons in
34
35
36
outwards configurations but only rotations, in which the covalent OH bond is not broken, are
37
38 observed for protons in the inwards configurations. It is worth highlighting here that local
39
40 distortions, motion of oxygen that is covalently bonded to the proton, and rotation of proton can
41
42
lead to inwards/outwards exchange and so the conclusions above do not imply that protons
43
44
45 which are in inwards configurations are blocked for the entire trajectory.
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
23
The Journal of Physical Chemistry Page 24 of 37

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Figure 11. (a) Effect of temperature on the hydrogen bond lengths for the YU system.
33
34 (b) Comparison of the hydrogen bond lengths for all the simulated systems at 1000 K. (c) and (d)
35
36 Histograms of hydrogen bond lengths and HO--O1 distances for the YU system at 1000 K. Black
37
38
39 lines correspond to all the hydrogen bonds existing in the simulation box over time. Red and blue
40
41 lines correspond to lengths at the time of a rotation or intra-octahedral jump, respectively. The
42
43 number of distances in each category is very different (320,000 for the entire set of lengths, up to
44
45
46
a few thousands for rotations and up to a few tens for intra-octahedral jumps) so, to facilitate the
47
48 comparison, histograms are given as a percentage in a given category.
49
50
51 As proton jumps are highly correlated with hydrogen bonds in the structure, it is
52
53 interesting to look more closely at the effect of temperature on the hydrogen bond length.
54
55
56 Figure 11.a shows the histograms for all the hydrogen bond lengths existing in the YU system at
57
58
59
60 ACS Paragon Plus Environment
24
Page 25 of 37 The Journal of Physical Chemistry

1
2
3 the three temperatures studied here. Similar plots for YC and YP are given in Figure S9 and
4
5
6 show once again that the Y-Y distribution does not affect significantly the results (as confirmed
7
8 by the direct comparison of histograms in Figure 11.b). At higher temperatures, wider
9
10 distributions of lengths are found, corresponding to larger fluctuations in the hydrogen bond
11
12
13
lengths. Both shorter and larger lengths are explored, the former favoring intra-octahedral jumps,
14
15 which occur preferentially when the hydrogen bonds, and HO--O1 distances, are the shortest. To
16
17 confirm this last statement, histograms of hydrogen bond lengths and HO--O1 distances when
18
19
jumps occur were plotted. Such histograms are shown in Figures 11.c and 11.d for the YU
20
21
22 system and in Figure S9 for YC and YP (other temperatures give similar results). Here,
23
24 histograms are normalised by the total number of distances found in each category (total,
25
26 rotations, intra-octahedral jumps) to ease the comparison. From these histograms, it is very clear
27
28
29 that intra-octahedral jumps only occur for short hydrogen bonds, a situation that happens more
30
31 often when the temperature increases. For rotations, the maximum of the hydrogen bond length
32
33 distribution is only slightly shifted to larger distances compared to the average bond length. The
34
35
36
results clearly indicate that low temperature static calculations do not necessarily capture
37
38 relevant motional processes at higher temperatures. It is worth pointing out that while the
39
40 possible existence of oxygen vacancies is omitted in this study, such vacancies could be present,
41
42
especially at high temperatures, and introduce compositional dependent chemical expansion
43
44
45 effects which would potentially impact the proton dynamics.43,44
46
47
48 We now examine the proton diffusion in a more global fashion to see if the local
49
50 dynamics observed lead to notable difference on long-range diffusion. From the DFT-MD
51
52
53 trajectories, the proton diffusion coefficients are determined using the Einstein relation:
54
55
56 𝑀𝑆𝐷 = ⟨(𝑟𝑖𝑡 − 𝑟𝑖0 )2 ⟩ = 6𝐷𝑡
57
58
59
60 ACS Paragon Plus Environment
25
The Journal of Physical Chemistry Page 26 of 37

1
2
3
where the MSD is the mean square displacements, 𝑟𝑖𝑡 is the position of proton 𝑖 at time 𝑡, D is the
4
5
6 diffusion coefficient and <> denotes an average over all protons. The MSDs calculated for the
7
8 different trajectories are given in Figure S10 and the diffusion coefficients obtained from the
9
10 MSDs and the corresponding activation energies, calculated using Arrhenius law and the values
11
12
13 at different temperatures, are gathered in Table 1 along with values from the literature obtained
14
15 experimentally or from classical molecular dynamics. Our calculations seem to overestimate the
16
17 diffusion coefficients compared to previously reported values at similar temperatures. It is
18
19
20
important to note that due to the computational limitations of DFT-MD, these simulations are
21
22 done on a small number of protons and the timescales and system sizes may be too short to
23
24 assess diffusion coefficients, which could explain this overestimation, as discussed in more detail
25
26
below.
27
28
29
30 The activation energies calculated (20.4, 16.0 and 21.3 kJ mol-1) are of the same order of
31
32 magnitude for the three Y distributions studied. This is consistent with the relatively similar
33
34 properties observed for these systems so far and with the fact that all structures contain a similar
35
36
37 distribution of H-Y configurations (Figure S3) so that the global motion is always an average
38
39 over protons in different configurations. The values calculated here are in the same range as the
40
41 activation energy of 24.1 kJ mol-1 reported in the computational work of Gomez et al. in which
42
43
44
no yttrium is present and thus no trapping can occur.38 Kinetic Monte Carlo (KMC) simulations,
45
46 which simulated proton dynamics using 96 binding sites with a 12.5% dopant concentration,
47
48 have predicted activation energies of 38 kJ mol-1 for a single proton45 and 43 kJ mol-1 when
49
50
8 protons were included in the simulation.46 In these simulations the protons spend a significant
51
52
53 amount of time in Y-OH-Z sites but travel between these sites through perpendicular and planar
54
55 sites. These activation energies are closer to the values of 40.5 kJ mol-1 and 45.0 kJ mol-1
56
57
58
59
60 ACS Paragon Plus Environment
26
Page 27 of 37 The Journal of Physical Chemistry

1
2
3 determined experimentally.1,8 KMC simulations are more suitable to characterize long-range
4
5
6 diffusion because of their ability to sample rare events more efficiently. It is worth noting though
7
8 that in both KMC and DFT-MD simulations, the distant sites were sampled only occasionally. In
9
10 addition, the KMC and the classical molecular simulations considered in table 116,17 simulate
11
12
13
proton motion on a potential energy surface derived from static DFT calculations, typically
14
15 through geometry optimisations and Nudge Elastic Band calculations. This could also explain
16
17 partly the discrepancies described here.
18
19
20
21 Diffusion Diffusion Diffusion
22 System Activation energies (kJ mol-1)
23
@500K @750K @1000K
24
25 YU
2.60 6.51 35.96 20.4
26
27
28
YC
29 3.79 24.10 22.67 16.0
30
31
32 YP
33 2.19 16.26 26.29 21.3
34
35
36 Ref 1
0.11 3.19 16.81 41.3
37
38
39
16 (> approx. 1000K)
40
41 Ref 8
0.07 1.16 3.30 45 (< 1000K)
42
43 13.0 (T1 measurements)
44
45
46 Ref 16
0.05 1.31 6.63 40.5
47
48
49 Ref 17
50 0.01 0.26 1.49 43.4
51
52
53 Table 1. Diffusion coefficients (/ 10-10 m2 s-1) and activation energies calculated for the 9 systems
54
55 studied in this work having a dopant concentration of 12.5% and from the literature
56
57
58
59
60 ACS Paragon Plus Environment
27
The Journal of Physical Chemistry Page 28 of 37

1
2
3 corresponding to experimental (dopant concentrations of 10% and 20%)1,8 and simulated values
4
5
6 (dopant concentrations of 0.46% and 12.5%)16,17.
7
8
9 Figure 7 shows that in the DFT-MD simulations, in some cases, a proton can jump
10
11 between perpendicular and planar sites at the beginning of a trajectory before spending a longer
12
13
14
time in Y-OH-Z sites with occasional excursions through planar sites. While not all protons
15
16 show a similar trajectory and there are also occurrences of protons escaping from near or planar
17
18 to distant sites, such a behavior, consistent with KMC simulations, has been observed for several
19
20
protons in the DFT-MD trajectories. This is again a hint that the current DFT-MD trajectories
21
22
23 may be too short to look at long range diffusion but that trapping could be observed more clearly
24
25 in longer simulations, unfortunately hard to achieve due to computational cost.
26
27
28 Looking more closely at the experimental data, the calculated activation energies are
29
30
31 intermediate between the activation energies determined at high (16.0 kJ mol-1) and low
32
33 temperatures (45.0 kJ mol-1) by Yamazaki et al.8 using impedance spectroscopy and H/D
34
35 exchange experiments. The difference between the high and low temperature impedance data of
36
37
38
29 kJ mol-1 was then ascribed to an association energy the protons need to overcome in order to
39
40 experience long-range diffusion. In the same work, the authors also characterized proton motion
41
42 by T1 relaxation measurements using NMR experiments and reported a single activation energy
43
44
of 13 kJ mol-1 across the whole temperature range monitored (room temperature to 873 K). The
45
46
47 T1 measurements are a probe of local motion while the impedance spectroscopy gives
48
49 information on long-range diffusion. This picture is consistent with the fact that, due to short
50
51 time and length scales, our simulations probe mainly the local motion and only rare events of
52
53
54 “trapped” to “trap-free” positions are sampled, which are responsible for the long-range
55
56 transport. As a consequence, while some trapping is captured in the simulations, the rare events
57
58
59
60 ACS Paragon Plus Environment
28
Page 29 of 37 The Journal of Physical Chemistry

1
2
3 corresponding to protons going from “trapped” to “trap-free” states are not sampled with
4
5
6 sufficient frequency to affect the activation energy in a significant way.
7
8
9 4. Conclusion
10
11
12 In this work, a combination of static DFT calculations and DFT-MD was used to study
13
14
15
the structural properties and proton dynamics in Y-doped BaZrO3 with two dopant
16
17 concentrations. The presence of yttrium leads to significant lattice distortions, which, as
18
19 quantified by, for example, the calculated metal-metal and metal-oxygen interatomic distances,
20
21
are in agreement with reported experimental results. Going beyond a simple description of the
22
23
24 average interatomic distances, the local distortions of the lattice have been characterized. This
25
26 allowed us to identify two types of configurations related to the bending of the metal-oxygen-
27
28 metal angle. The nature of the two configurations, labelled inwards bending and outwards
29
30
31 bending, were shown to affect the energies of the systems. In particular, i) the larger the bending,
32
33 the lower the energy and ii) the lowest energies were observed for structures combining near and
34
35 planar configurations with an inwards bending. The lowest energy structures were the ones with
36
37
38
the strongest hydrogen bonds, which is only possible in the inwards configuration in the static
39
40 calculations.
41
42
43 DFT-MD studies allowed us to explore the effect of temperature on the local structures
44
45
and proton trajectories. A detailed analysis of the proton jumps according to the different H-Y
46
47
48 configurations showed that inwards configurations favor rotations while intra-octahedral jumps
49
50 are only observed from outwards configurations. In fact, in contrast to static DFT calculations,
51
52 the DFT-MD simulations indicate that outwards bending leads to shorter hydrogen bond lengths
53
54
55 and HO--O1 distances on average. This is consistent with favoring intra-octahedral jumps as a
56
57
58
59
60 ACS Paragon Plus Environment
29
The Journal of Physical Chemistry Page 30 of 37

1
2
3 decrease in the HO--O1 distance should decrease the energy barrier for the jump. The DFT-MD
4
5
6 simulations confirm very clearly that intra-octahedral jumps only occur when the hydrogen bond
7
8 lengths and H--O1 distances are the shortest. It is important to note that local distortions and
9
10 rotations of protons can lead to inwards/outwards exchange. As a consequence, the statements
11
12
13
above do not imply that protons in inwards configurations at a given time are blocked for the
14
15 entire trajectory. The proton diffusion coefficients determined at three different temperatures
16
17 allowed for the extraction of activation energies, giving values in the same range as some
18
19
experimental results that probe short range motion but in general lower than previously reported
20
21
22 simulation results using KMC or classical molecular simulations based on potential energy
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 12. Scheme summarizing the main conclusions of this work regarding proton transfers.
48
49 Static DFT calculations have permitted the identification of two configurations, namely inwards
50
51 and outwards bending, which favor different proton transfers. The inwards bending is usually
52
53
54
lower in energy, especially close to dopant atoms, and favors rotations with no intra-octahedral
55
56 jumps observed in the DFT-MD simulations conducted here. The outwards bending on the
57
58
59
60 ACS Paragon Plus Environment
30
Page 31 of 37 The Journal of Physical Chemistry

1
2
3 contrary allows for intra-octahedral proton transfers although still less frequent than rotations.
4
5
6 These observations are consistent with proton trapping close to the yttrium dopant.
7
8
9 surfaces determined using static DFT methods. While trapping is believed to happen, it only
10
11 affects the long-range diffusion and the local motion of protons is largely unaffected. As a
12
13
14
consequence, the simulations reported in this work, probably too short to study long-range
15
16 diffusion, do not capture the larger activation energies for long range proton diffusion. Overall,
17
18 our results demonstrate the importance of the bending for proton dynamics and as local
19
20
distortions are related with the yttrium content, this work is a new step towards understanding the
21
22
23 existence of an optimal dopant concentration for the proton conduction. The results highlight the
24
25 role of temperature and lattice motion on controlling the mechanisms for proton motion.
26
27
28 ASSOCIATED CONTENT
29
30
31
32 Supporting Information. An illustration of the octahedral tilting, total residence time fractions,
33
34 number of H-Y configuration sites in the systems, numbers of proton jumps, fractions of inwards
35
36 and outwards bending as a function of time, characterizations of the interatomic distances related
37
38
39
to the hydrogen bonding, mean square displacements.
40
41
42 AUTHOR INFORMATION
43
44 The authors declare no competing financial interests.
45
46
47
ACKNOWLEDGMENT
48
49
50 We thank Dr Bartomeu Monserrat, Dr Rachel Kerber and Dr Frédéric Blanc for fruitful
51
52 discussions. C. M. acknowledges an Oppenheimer Research Fellowship from the School of
53
54 Physical Sciences from the University of Cambridge. This project has received funding from the
55
56
57
58
59
60 ACS Paragon Plus Environment
31
The Journal of Physical Chemistry Page 32 of 37

1
2
3 European Research Council (ERC) under the European Union’s Horizon 2020 research and
4
5
6 innovation programme (grant agreement no. 714581). Via our membership of the UK's HEC
7
8 Materials Chemistry Consortium, which is funded by EPSRC (EP/L000202), this work used the
9
10 ARCHER UK National Supercomputing Service. MAG’ work was supported by the National
11
12
13
Science Foundation under grant DMR-1709975 and the Mount Holyoke College Department of
14
15 Chemistry. Computational resources were provided in part by the MERCURY consortium under
16
17 NSF grant CHE-1626238. Data access statement: original data supporting this publication are
18
19
available as Supporting Information.
20
21
22
23
24 REFERENCES
25
26 (1) Kreuer, K. D. Proton - Conducting Oxides. Annu. Rev. Mater. Res. 2003, 33, 333–359.
27
28 (2) Yamazaki, Y.; Hernandez-Sanchez, R.; Haile, S. M. High Total Proton Conductivity in
29
30
31 Large-Grained Yttrium-Doped Barium Zirconate. Chem. Mater. 2009, 21, 2755–2762.
32
33 (3) Bohn, H. G.; Schober, T. Electrical Conductivity of the High-Temperature Proton
34
35 Conductor BaZr0.9Y0.1O2.95. J. Am. Ceram. Soc. 2000, 83, 768–772.
36
37
38
(4) Fabbri, E.; Pergolesi, D.; Licoccia, S.; Traversa, E. Does the Increase in Y-Dopant
39
40 Concentration Improve the Proton Conductivity of BaZr1−xYxO3−δ Fuel Cell Electrolytes?
41
42 Solid State Ionics 2010, 181, 1043–1051.
43
44
(5) Iguchi, F.; Tsurui, T.; Sata, N.; Nagao, Y.; Yugami, H. The Relationship between
45
46
47 Chemical Composition Distributions and Specific Grain Boundary Conductivity in
48
49 Y-Doped BaZrO3 Proton Conductors. Solid State Ionics 2009, 180, 563–568.
50
51 (6) Larring, Y. Protons in Rare Earth Oxides. Solid State Ionics 1995, 77, 147–151.
52
53
54 (7) Kreuer, K. Aspects of the Formation and Mobility of Protonic Charge Carriers and the
55
56 Stability of Perovskite-Type Oxides. Solid State Ionics 1999, 125, 285–302.
57
58
59
60 ACS Paragon Plus Environment
32
Page 33 of 37 The Journal of Physical Chemistry

1
2
3 (8) Yamazaki, Y.; Blanc, F.; Okuyama, Y.; Buannic, L.; Lucio-Vega, J. C.; Grey, C. P.;
4
5
6 Haile, S. M. Proton Trapping in Yttrium-Doped Barium Zirconate. Nat. Mater. 2013, 12,
7
8 647–651.
9
10 (9) Pornprasertsuk, R.; Kosasang, O.; Somroop, K.; Jinawath, S.; Prinz, F. B. Proton
11
12
13
Conductivity Studies of Y-Doped Barium Zirconate: Theoretical and Experimental
14
15 Approaches. ECS Transactions 2010, 25, 367–381.
16
17 (10) Merinov, B.; Goddard, W. Proton Diffusion Pathways and Rates in Y-Doped BaZrO3
18
19
Solid Oxide Electrolyte from Quantum Mechanics. J. Chem. Phys. 2009, 130, 194707.
20
21
22 (11) Björketun, M. E.; Sundell, P. G.; Wahnström, G. Effect of Acceptor Dopants on the
23
24 Proton Mobility in BaZrO3: A Density Functional Investigation. Phys. Rev. B 2007, 76,
25
26 054307.
27
28
29 (12) Gomez, M. A.; Chunduru, M.; Chigweshe, L.; Foster, L.; Fensin, S. J.; Fletcher, K. M.;
30
31 Fernandez, L. E. The Effect of Yttrium Dopant on the Proton Conduction Pathways of
32
33 BaZrO3, a Cubic Perovskite. J. Chem. Phys. 2010, 132, 214709.
34
35
36
(13) Draber, F. M.; Ader, C.; Arnold, J. P.; Eisele, S.; Grieshammer, S.; Yamaguchi, S.;
37
38 Martin, M. Nanoscale Percolation in Doped BaZrO3 for High Proton Mobility. Nat.
39
40 Mater. 2020, 19, 338–346.
41
42
(14) Blanc, F.; Sperrin, L.; Lee, D.; Dervişoğlu, R.; Yamazaki, Y.; Haile, S. M.; De Paëpe, G.;
43
44
45 Grey, C. P. Dynamic Nuclear Polarization NMR of Low-γ Nuclei: Structural Insights into
46
47 Hydrated Yttrium-Doped BaZrO3. J. Phys. Chem. Lett. 2014, 5, 2431–2436.
48
49 (15) Friman, J. Molecular Dynamics Simulations of Proton Diffusion in Yttrium Doped
50
51
52 Barium Zirconate, Master Dissertation, Chalmers University of Technology, 2013.
53
54 (16) Raiteri, P.; Gale, J. D.; Bussi, G. Reactive Force Field Simulation of Proton Diffusion in
55
56
57
58
59
60 ACS Paragon Plus Environment
33
The Journal of Physical Chemistry Page 34 of 37

1
2
3 BaZrO3 Using an Empirical Valence Bond Approach. J. Phys. Condens. Matter 2011, 23,
4
5
6 334213.
7
8 (17) van Duin, A. C. T.; Merinov, B. V.; Han, S. S.; Dorso, C. O.; Goddard, W. A. ReaxFF
9
10 Reactive Force Field for the Y-Doped BaZrO3 Proton Conductor with Applications to
11
12
13
Diffusion Rates for Multigranular Systems. J. Phys. Chem. A 2008, 112, 11414–11422.
14
15 (18) Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal,
16
17 Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276.
18
19
(19) Marx, D.; Tuckerman, M. E.; Hutter, J.; Parrinello, M. The Nature of the Hydrated Excess
20
21
22 Proton in Water. Nature 1999, 397, 601–604.
23
24 (20) Hassanali, A.; Prakash, M. K.; Eshet, H.; Parrinello, M. On the Recombination of
25
26 Hydronium and Hydroxide Ions in Water. Proc. Natl. Acad. Sci. 2011, 108, 20410–20415.
27
28
29 (21) Kattirtzi, J. A.; Limmer, D. T.; Willard, A. P. Microscopic Dynamics of Charge
30
31 Separation at the Aqueous Electrochemical Interface. Proc. Natl. Acad. Sci. 2017, 114,
32
33 13374–13379.
34
35
36
(22) Fronzi, M.; Tateyama, Y.; Marzari, N.; Nolan, M.; Traversa, E. First-Principles Molecular
37
38 Dynamics Simulations of Proton Diffusion in Cubic BaZrO3 Perovskite under Strain
39
40 Conditions. Mater. Renew. Sustain. Energy 2016, 5, 14.
41
42
(23) Ottochian, A.; Dezanneau, G.; Gilles, C.; Raiteri, P.; Knight, C.; Gale, J. D. Influence of
43
44
45 Isotropic and Biaxial Strain on Proton Conduction in Y-Doped BaZrO3: A Reactive
46
47 Molecular Dynamics Study. J. Mater. Chem. A 2014, 2, 3127.
48
49 (24) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made
50
51
52 Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.
53
54 (25) Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple.”
55
56
57
58
59
60 ACS Paragon Plus Environment
34
Page 35 of 37 The Journal of Physical Chemistry

1
2
3 Phys. Rev. Lett. 1998, 80, 890–890.
4
5
6 (26) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert, M. I. J.; Refson, K.;
7
8 Payne, M. C. First Principles Methods Using CASTEP. Zeitschrift für Krist. 2005, 220,
9
10 567–570.
11
12
13
(27) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B
14
15 1976, 13, 5188–5192.
16
17 (28) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy
18
19
Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186.
20
21
22 (29) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-
23
24 Wave Method. Phys. Rev. B 1999, 59, 1758–1775.
25
26 (30) Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. CP2K: Atomistic Simulations
27
28
29 of Condensed Matter Systems. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2014, 4, 15–25.
30
31 (31) Krack, M. Pseudopotentials for H to Kr Optimized for Gradient-Corrected Exchange-
32
33 Correlation Functionals. Theor. Chem. Acc. 2005, 114, 145–152.
34
35
36
(32) Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials.
37
38 Phys. Rev. B 1996, 54, 1703–1710.
39
40 (33) Hartwigsen, C.; Goedecker, S.; Hutter, J. Relativistic Separable Dual-Space Gaussian
41
42
Pseudopotentials from H to Rn. Phys. Rev. B 1998, 58, 3641–3662.
43
44
45 (34) Bjorketun, M.; Sundell, P.; Wahnstrom, G.; Engberg, D. A Kinetic Monte Carlo Study of
46
47 Proton Diffusion in Disordered Perovskite Structured Lattices Based on First-Principles
48
49 Calculations. Solid State Ionics 2005, 176, 3035–3040.
50
51
52 (35) Bi, L.; Traversa, E. Synthesis Strategies for Improving the Performance of Doped-BaZrO3
53
54 Materials in Solid Oxide Fuel Cell Applications. J. Mater. Res. 2013, 29, 1–15.
55
56
57
58
59
60 ACS Paragon Plus Environment
35
The Journal of Physical Chemistry Page 36 of 37

1
2
3 (36) Pergolesi, D.; Fabbri, E.; D’Epifanio, A.; Di Bartolomeo, E.; Tebano, A.; Sanna, S.;
4
5
6 Licoccia, S.; Balestrino, G.; Traversa, E. High Proton Conduction in Grain-Boundary-Free
7
8 Yttrium-Doped Barium Zirconate Films Grown by Pulsed Laser Deposition. Nat. Mater.
9
10 2010, 9, 846–852.
11
12
13
(37) Giannici, F.; Shirpour, M.; Longo, A.; Martorana, A.; Merkle, R.; Maier, J. Long-Range
14
15 and Short-Range Structure of Proton-Conducting Y:BaZrO3. Chem. Mater. 2011, 23,
16
17 2994–3002.
18
19
(38) Gomez, M. A.; Griffin, M. A.; Jindal, S.; Rule, K. D.; Cooper, V. R. The Effect of
20
21
22 Octahedral Tilting on Proton Binding Sites and Transition States in Pseudo-Cubic
23
24 Perovskite Oxides. J. Chem. Phys. 2005, 123, 094703.
25
26 (39) Sahraoui, D. Z.; Mineva, T. Structural Properties of Y-Doped BaZrO3 as a Function of
27
28
29 Dopant Concentration and Position: A Density Functional Study. Solid State Ionics 2013,
30
31 232, 1–12.
32
33 (40) Davies, R.; Islam, M. S.; Gale, J. D. Dopant and Proton Incorporation in Perovskite-Type
34
35
36
Zirconates. Solid State Ionics 1999, 126, 323–335.
37
38 (41) Stukowski, A. Visualization and Analysis of Atomistic Simulation Data with OVITO–the
39
40 Open Visualization Tool. Model. Simul. Mater. Sci. Eng. 2010, 18, 015012.
41
42
(42) Münch, W.; Kreuer, K. D.; Seifert, G.; Maier, J. Proton Diffusion in Perovskites:
43
44
45 Comparison between BaCeO3, BaZrO3, SrTiO3, and CaTiO3 Using Quantum Molecular
46
47 Dynamics. Solid State Ionics 2000, 136–137, 183–189.
48
49 (43) Bjørheim, T. S.; Løken, A.; Haugsrud, R. On the relationship between chemical expansion
50
51
52 and hydration thermodynamics of proton conducting perovskites. J. Mater. A. 2016, 4,
53
54 5917.
55
56
57
58
59
60 ACS Paragon Plus Environment
36
Page 37 of 37 The Journal of Physical Chemistry

1
2
3 (44) Kjølseth, C.; Wang, L.-Y.; Haugsrud, R.; Norby, T. Determination of the enthalpy of
4
5
6 hydration of oxygen vacancies in Y-doped BaZrO3 and BaCeO3 by TG-DSC. Solid State
7
8 Ion. 2010, 181, 1740.
9
10 (45) Krueger, R. A.; Haibach, F. G.; Fry, D. L.; Gomez, M. A. Centrality Measures Highlight
11
12
13
Proton Traps and Access Points to Proton Highways in Kinetic Monte Carlo Trajectories.
14
15 J. Chem. Phys. 2015, 142, 154110.
16
17 (46) Gomez, M. A.; Fry, D. L.; Sweet, M. E. Effects on the Proton Conduction Limiting
18
19
Barriers and Trajectories in BaZr0.875Y0.125O3 Due to the Presence of Other Protons. J.
20
21
22 Korean Ceram. Soc. 2016, 53, 521–528.
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
37

You might also like