Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

New High Pressure Phases of Titanite-type

CaTiSiO5 Predicted from First-Principles

Subhamoy Char,† P. S. Ghosh,∗,†,‡ Kawsar Ali,† and A. Arya†,‡

†Glass & Advanced Materials Division, Bhabha Atomic Research Center, Mumbai 400 085,
India
‡Homi Bhabha National Institute, Anushaktinagar, Mumbai 400 094, India

E-mail: psghosh@barc.gov.in

Abstract

Titanite-type CaTiSiO5 is one of the common accessory minerals in igneous, meta-

morphic, sedimentary, and ultrahigh-pressure rocks. Determination of the high-pressure

crystal-chemical behavior of silicates is crucial to comprehend the crystalline materials

in deep earth geospheres. The present study implements a density functional theory

calculation assisted structure search method using evolutionary algorithms and reveals

two new structural phase transitions: displacive-type C 2/c → P1̄ and reconstructive-

type P1̄ → Pnma at 10 GPa and 18 GPa, respectively. The topology of the high-

pressure Pnma structure can be identified as a combination of corner-shared distorted

zigzag SiO6 octahedra along [100] direction and edge-shared TiO6 prism (with SiO6 )

along [001] direction. This novel topological arrangement has not been observed in any

of the high-pressure structural prototypes of silicates reported in the literature.

1
Introduction

The most significant components of the Earth’s crust and mantle are silicates. Mineralogists
and crystallographers have long been interested in conducting systematic studies on these
minerals. 1 The behavior of silicates under high pressures (HP) and high temperatures (HT)
has received particular interest since it is crucial to our comprehension of crystalline materials
in deep-earth geospheres. The crystal-chemical and physical characteristics of silicates under
varying thermodynamic conditions are dictated by the local coordination of atoms. The HP
structures of the silicates are subdivided into eleven major groups according to their chemical
compositions 2 and further sub-divided based on Si-O coordination and topology of linkage
between Si coordination polyhedra. Among these classifications, titanite-related structures
are an interesting class of silicates where the crystal structures contain a mixture of four-
, five- or/and six-coordinated Si atoms 3 at ambient pressure. Under increasing pressure,
structural transformations change coordination to SiO6 octahedron or a mixture of SiO4 and
SiO6 leading to the formation of structures with higher 3,4 or lower symmetry, 5,6 respectively.
Hence, titanite-structure type minerals offer a rich crystal chemical behavior at HP.
The CaTiSiO5 with titanite-structure type is a ubiquitous accessory mineral in igneous,
metamorphic, sedimentary, and ultrahigh-pressure rocks. 7 It is ostensibly the most common
mineral containing titanium outside of rutile and ilmenite. 8 Geologists and material scientists
have been largely interested in its stability and crystal chemistry because of its importance
as a petrogenic indicator, 9 glass/ceramics, 10 versatile host for rare earth elements (REE),
and U-Pb geochronometer for dating geological events. 8,11,12 Even though it has geological
and material science related importance, the HP behavior of CaTiSiO5 is scarcely explored
in a limited pressure range. The structural phase transition of the constituent minerals in
magmatic and metamorphic rocks is one of the major concerns at extreme temperatures and
pressures. Moreover, several silicate minerals undergo a phase transition between 5 GPa (for
some framework silicates) to about 20 GPa, which corresponds to the pressure at the top
of the Earth’s lower mantle. 6,13,14 Therefore, the study of HP behavior in CaTiSiO5 offers

2
an avenue to understand the underlying physics and crystal chemistry leading to enriched
structural variety and chemical bonding.
The CaTiSiO5 exists in monoclinic structure (space group P 21 /c) at ambient conditions;
whereas at HT and/or HP, it undergoes a P 21 /c → C 2/c structural phase transition. 15–18
Around 487 K and ambient pressure, CaTiSiO5 transforms to C 2/c space group which is
also an antiferroelectric to paraelectric phase transition. 19 In P 21 /c structure, 20 the TiO6
octahedra is distorted due to the slight displacement of Ti atoms towards the vertex shared
O-atom whereas this well-known off-center displacement of Ti atoms disappears in C 2/c
structure. The Ti off-center displacement is parallel in a TiO6 chain but antiparallel to
the neighboring TiO6 chain and this leads to an antiferroelectric ground state. Another
structural phase transition is observed around 825 K at ambient pressure 21 and HT structure
is also characterized as C 2/c structure. In-situ HT-XRD study reveals this low temperature
(LT) C 2/c to high temperature (HT) C 2/c structural transition at 825 K govern by the
symmetrical reorganization of Ca-atoms around the Ti-centers. 22
Apart from the experimental investigations on the structural phase transition of CaTiSiO5
between temperature range from 300 K to 1600 K and pressure range from 1 atm to 6.9 GPa,
a few first-principles studies have been reported. 23–25 Most of the theoretical calculations ex-
plore P 21 /c → C 2/c temperature-induced phase transition focusing on the contribution
of Ti atoms displacement within the TiO6 octahedra. A recent stress-strain calculation of
CaTiSiO5 shows that with increasing pressure from 0–5 GPa, the P 21 /c structure transforms
to C 2/c structure due to the shear softening behavior. 25 Moreover, Density Functional Per-
turbation Theory (DFPT) calculations reveal that the appearance of unstable phonon modes
in the high temperature C 2/c structure corresponds to the displacement of Ti atoms in the
TiO6 octahedra leading to the experimentally observed C 2/c → P 21 /c phase transition with
reducing temperature. 23
Insufficient insight into the crystal-chemical behavior of CaTiSiO5 at HP has motivated
this work to focus on pressure-dependent structural phase transition in the 0-50 GPa pres-

3
sure range within the framework of first-principles calculations. The present study discovers
two new structural polymorphs from an evolutionary algorithm. Further, analysis of struc-
tural, elastic, and dynamical stabilities leads to the identification of stability regions of new
polymorphs as a function of pressure.

Computational Methods

The electronic structure calculations are performed using Density Functional Theory (DFT)
as implemented in the Vienna Ab-initio Simulation Package (VASP) with Projected Aug-
mented Wave (PAW) potentials. 26,27 Two levels of approximations for exchange-correlation
functional, namely, local gradient approximations (LDA) in Ceperley-Alder (CA) parame-
terization 28 and generalized gradient approximation (GGA) with Perdew-Burke-Ernzerhof
(PBE) parameterization 29 are considered. The explicitly treated valence electron configura-
tions of Si and O are (3s2 3p2 ) and (2s2 2p4 ), respectively. Two combinations of Ca and Ti
PAW potentials having valence electron configurations Ca pv: 3p6 4s2 and Ti: 3d3 4s1 (ab-
breviated as GGA1/LDA1) and Ca sv: 3s2 3p6 4s2 and Ti sv: 3s2 3p6 3d3 4s1 (abbreviated
as GGA2/LDA2), respectively, are considered. In all calculations, the cut-off energy of the
plane wave is set to 600 eV. Brillouin Zone summations are performed on 8 × 6 × 8 k-grid
centered at the Γ point which is sufficient to ensure the energy convergence of less than 1
meV/atom. During structural relaxations, the energy and forces are optimized up to 10−7
eV and 10−3 eV/Å, respectively.
To predict crystal structures at high pressures within the 0-50 GPa pressure range with
a 10 GPa interval, an evolutionary algorithm is used as implemented in USPEX code. 30–32
The USPEX code is widely used to predict stable high-pressure crystal structures without
requiring any experimental informations 33 and the traditional fixed-composition searches are
considered in the present study. Structure prediction is performed initially in conjunction
with the classical molecular dynamics package GULP 34 to reduce the computational burden

4
followed by the ab-initio structural optimization using VASP to achieve reliable predictions.
The short-range interactions in GULP are described using Buckingham potentials 35,36 while
the long-range interactions are considered to be dominated by Coulombic interactions with
fractional charges on the atoms. As CaTiSiO5 possesses 4 formula units per unit cell (Z=4),
the structure search is restricted to potential structural candidates with Z=4 with the simu-
lation box containing 4 Ca, 4 Ti, 4 Si, and 20 O atoms, and the volume is limited to 500 Å3 .
In the process of structure search, initially, 50 structures are randomly generated in the first
generation. These random structures are ranked by a well-defined fitness function and from
the next generations, different evolutionary operators are applied to the initial structures.
In the structure search, in each generation, a maximum of 50 structures are populated. The
contribution of different variation operators in the subsequent generations is as follows: 0.5
fractions of the generation produced by heredity, 0.20 fraction of generation produced ran-
domly from the lattice space group, 0.1 fractions of generation produced by permutation, 0.1
fraction of generation produced by soft mutation, and 0.1 fraction of generation produced by
lattice mutation. This process continued until the lattice energy remained the same for 20
generations in a row. The choice of Buckingham potential parameters, effective charges and
details of USPEX implementation are elaborated in the Supporting Information. The best
structures predicted from the USPEX-GULP combination are further optimized using VASP
with 2π × 0.2 Å−1 k -point spacing to ensure that the enthalpy calculations are converged to
better than 1 meV/atom.
The phonon dispersion curve of the high-pressure structures is calculated using the
PHONOPY package 37 and finite displacement method 38 to ensure that predicted structures
are dynamically stable in the desired pressure range. The symmetry analysis of all optimized
structures is performed using the FINDSYM utility of the ISOTROPY package. 39 The elastic
constants of the predicted structures are calculated using the stress-strain method as imple-
mented in the VASP code to ensure the mechanical stability of the predicted structures. The
projected Crystal Orbital Hamilton Population (pCOHP) is analyzed in LOBSTER pack-

5
age 40,41 to find the nature of bonding in the new structures. All the crystal structures are
visualized in the VESTA package. 42

Results and Discussion

0 GPa Structure

The ground state structure of CaTiSiO5 (P 21 /c) is subjected to structural optimization


using GGA1/GGA2 and LDA1/LDA2 at zero pressure and optimized lattice parameters
and volumes are compared with room temperature experiments as shown in Table 1. The
GGA1/GGA2 calculated lattice parameters are in good agreement (±1%) with experiments;
while LDA1/LDA2 calculated lattice parameters are underestimated maximum by 3%. The
equilibrium volume is 2.7% overestimated (maximum) and 3.8% underestimated (maximum),
respectively, for GGA1/GGA2 and LDA1/LDA2 compared to the experiment as expected.
At high temperature and/or pressure, P 21 /c phase undergoes a structural phase transition
to C 2/c phase. 15–18 The experimental structural information of C 2/c phase reported at
1.70 GPa and 627 K 43 is subjected to structural optimizations at zero pressure using both
GGA1/GGA2 and LDA1/LDA2. The optimized structural parameters are also tabulated in
Table 1. Structural similarity of P 21 /c and C 2/c phase leads to similar GGA1/GGA2 and
LDA1/LDA2 calculated lattice parameters and volumes. As a result, the energy difference
between P 21 /c and C 2/c phase is small which is of the order of 0.7 (0.2) meV/unit cell
for GGA1 (LDA1) and 2.0 (0.3) for GGA2 (LDA2), respectively. In all the cases, P 21 /c
structure is the ground state.
The Octahedrally coordinated Ti-atoms in P 21 /c structure are located at off-center posi-
tions within the TiO6 octahedra. The off-centering type distortion is parallel to the individual
TiO6 chain along [001] direction, but anti-parallel with neighboring TiO6 chain along [001̄]
direction. 16 The off-centering is absent in the C 2/c structure. The well-known off-center
distortion of Ti atoms in TiO6 octahedra is observed only in GGA2 optimized P 21 /c struc-

6
ture where Ti-O bond lengths are 1.83, 1.93, 1.99, 1.99, 2.03, and 2.04 Å. These Ti-O bond
lengths agree with experimental bond lengths of 1.77, 1.97, 1.98, 2.01, 2.01, and 2.04 Å. 16
The difference in the Ti-O bond distances between 1st and 2nd nearest neighbors is referred
to as anti-polar distortion. This distortion pattern is absent in the C 2/c structure where the
Ti-O bond lengths are 1.88, 1.88, 2.00, 2.00, 2.04, and 2.04 Å. However, such kind of Ti-atom
distortion is absent in the LDA1/LDA2 and GGA1 optimized P 21 /c structure at 0 GPa. It
can be concluded that even though LDA1/LDA2 or GGA1 predicts lattice parameters ac-
curately, only GGA2 exchange-correlation functional correctly distinguishes small structural
differences between P 21 /c and C 2/c.

Table 1: Calculated lattice parameters, volumes for both P 21 /c and C 2/c structures. For
comparison, experimental values are also included. These experimental lattice parameters
are reported at room temperature and atmospheric pressure.

Structure type Lattice parameter (a,b,c, β) Volume (Å3 )


GGA1 7.025 8.704 6.528 113.6 365.68
P 21 /c LDA1 6.866 8.546 6.362 113.3 342.95
GGA2 7.130 8.766 6.658 114.1 379.92
LDA2 6.962 8.611 6.481 113.7 355.80
Exp 15 7.068 8.714 6.562 113.8 369.73
GGA1 7.025 8.704 6.528 113.6 365.68
C 2/c LDA1 6.865 8.546 6.362 113.3 342.95
GGA2 7.122 8.767 6.659 114.0 379.63
LDA2 6.962 8.611 6.480 113.7 355.80

Phase transition in 0-10 GPa range

To determine the pressure response of CaTiSiO5 , the ground state structure P 21 /c is sub-
jected to hydrostatic pressure up to 50 GPa and full structural relaxation is performed. The
applied pressure is sequentially increased in steps of 2 GPa from the zero value with each
new simulation initialized by the output of the previous one. The variation of volumes,
lattice parameters, and Ti-O bond distances corresponding to the antipolar distortions as
a function of pressures calculated using LDA1/LDA2 and GGA1/GGA2 in the 0-10 GPa

7
12 10
a b c

Lattice parameter (Å)


9
Volume (Å3/atom ) 11
8
10
7
9
6
a) DFT−GGA Expt. b)
8 5
0 2 4 6 8 10 0 2 4 6 8 10

130 1.95
α β γ
120

110 1.90

dTi−O (Å)
Angle (°)

100

90 1.85

80
c) d)
70 1.80
0 2 4 6 8 10 0 2 4 6 8 10
Pressure (GPa) Pressure (GPa)

Figure 1: DFT-GGA2 calculated structural parameters of CaTiSiO5 at pressures up to 10


GPa. The filled square represents the experimental data. 17 Beyond 6 GPa, the antipolar Ti-
O distortion vanishes as the axial Ti-O interatomic separation (red and blue bonds) becomes
identical. In the inset, the cyan and red colors represent the Ti and O atoms, respectively.

pressure range is shown in Fig. S1 and S2, respectively. Experimentally available pressure
variations of lattice parameters 17 are also included for comparison. The LDA1/LDA2 calcu-
lated volumes and lattice parameters are underestimated throughout the pressure range and
GGA1/GGA2 calculated lattice parameters provide a better match with the experiments.
Figure 1 reports the evolution of the GGA2 calculated volumes, lattice parameters and Ti-O
bond distances corresponding to the antipolar distortions as a function of pressures (open
circles). The pressure-volume data is fitted to Birch–Murnaghan equation of state (solid
lines in Fig. 1(a)) and the calculated bulk modulus (B0 ) is 145.4 GPa. The GGA2 calcu-
lated B0 is overestimated compared to the experimental value of 113.4 GPa calculated at 298
K. 43 The lattice parameters decrease in response to the pressure and the associated linear
compressibilities Ka = − a1 ∂P
∂a 1 ∂b 1 ∂c
  
T
, K b = − b ∂P T
and K c = − c ∂P T
are calculated for a,

8
b and c-lattice parameters, respectively. The GGA2 calculated values are Ka = 23.3, Kb =
12.7, and Kc = 21.8 TPa−1 which signifies that the b axis is stiffer than the a and c axis in
agreement with experimental observation in that pressure range. 43 The monoclinic angle (β)
is also decreasing with increasing pressures.
The pressure evolution of the 1st and 2nd nearest neighbors Ti-O bond distances are
shown in Fig. 1(d). With increasing pressure, the differences between 1st and 2nd neighbors
decrease and at 8 GPa it becomes equal. This leads to P 21 /c → C 2/c phase transition.
According to the experimental observations, CaTiSiO5 undergoes a second-order structural
phase transition from P 21 /c → C 2/c at 6.9 GPa pressure and room temperature. 16 The
structural similarity between P 21 /c and C 2/c enforces almost identical GGA2 calculated
lattice parameters and volumes at zero and high pressures. Evidently P 21 /c → C 2/c tran-
sition is not identifiable from the pressure-volume or pressure-lattice parameter plots. It is
also important to note that we could identify P 21 /c → C 2/c phase transition only from the
GGA2 optimized structures. Hence, only GGA2 is considered further to study structural
evolution as a function of pressure.
Table 2 shows DFT-GGA2 calculated single crystal elastic constants of P 21 /c structure
at 0 GPa and C 2/c structure at 10 GPa. For P 21 /c and C 2/c structure, all the diagonal
elements (Cii ) are positive. The P 21 /c follows the required criteria of mechanical stability 44
whereas C 2/c structure violates those criteria indicating mechanical instability at 10 GPa.
The B0 calculated from single crystal elastic constants for P 21 /c structure is 114.1 GPa
matches preferably to the experimental value of 113.4 GPa (at 298 K). 43

Phase transition beyond 10 GPa

We have further determined the structural evolution of CaTiSiO5 beyond 10 GPa and up
to 50 GPa using GGA2 as shown in Fig. 2. The variation of volumes, lattice parameters
and angles as a function of pressure remain continuous up to 32 GPa (Fig. 2). As expected
when pressure increases, all the inter-atomic separations like Ca-O, Ti-O, and Si-O reduce.

9
12 10
a b c

Lattice parameter (Å)


9
Volume (Å3/atom )

11
8

10 7

6
9
5
a) b)
8 4
10 15 20 25 30 35 40 45 50 10 15 20 25 30 35 40 45 50

130 150
α β γ
120 140

Ti−O−Ti Angle (°)


110
Angle (°)

130
100
120
90

80 110
c) d)
70 100
10 15 20 25 30 35 40 45 50 10 15 20 25 30 35 40 45 50
Pressure (GPa) Pressure (GPa)

Figure 2: DFT-PBE calculated structural parameters of CaTiSiO5 after pressure 10 GPa.


Phase transitions are observed around 34 GPa and 42 GPa, respectively. The blue triangle
in the pressure-volume plot indicates the change of volume in the back pressure calculations
starting from 34 GPa and 42 GPa, respectively. There is a sharp drop of Ti-O-Ti angle at
the onset of transitions.

However, the C 2/c symmetry of the optimized structure is retained up to 32 GPa. We


have noted down the change of volume for different coordinated polyhedra with increasing
pressure and found the volume of CaO7 polyhedra is maximally reduced in comparison with
TiO6 and SiO4 polyhedra indicating CaO7 polyhedra to be the softest structural moiety in
CaTiSiO5 . At 32 GPa, the CaO7 polyhedral volume is reduced by 19 % with respect to
ground state P 21 /c volume; whereas this reduction is 11 % and 8 % for TiO6 and SiO4
polyhedra, respectively.
A discontinuous change in lattice angle is observed between 32 and 34 GPa leading to
inflection in volume-pressure and lattice parameter-pressure plots. In P 21 /c structure, each

10
Figure 3: DFT-GGA2 optimized structures of 32 GPa (a) and 34 GPa (b), respectively.
In 32 GPa structure, all the Ti-Ti interatomic separations are identical however, these are
different in 34 GPa structure which shows a similar structural analogy with the high-pressure
phase transition of malayaite. 5

Table 2: Calculated Elastic constants (in GPa) for P 21 /c at 0 GPa, C 2/c at 10 GPa, P1̄ at
10 GPa, and Pnma at 20 GPa.

C11 C22 C33 C44 C55 C66 C12 C13 C14


C15 C16 C23 C24 C25 C26 C34 C35 C36
C45 C46 C56
P 21 /c 186.0 294.4 239.0 80.2 55.2 41.7 66.2 33.0 -
10.1 - 82.5 - 8.5 - - 26.2 -
- 6.9 -
C 2/c 246.8 341.6 260.0 32.3 105.1 44.5 93.3 75.5 -
57.5 - 100.3 - 8.5 - - -18.0 -
- 47.7 -
P1̄ 241.0 339.5 260.9 46.5 104.1 58.3 91.1 79.5 -14.8
52.1 20.5 100.8 -9.9 7.1 9.0 -7.6 -18.8 0.5
-5.9 34.3 3.0
Pnma 535.8 379.6 585.2 66.8 87.2 77.9 64.8 159.3 -
- - 102.5 - - - - - -
- - -

11
TiO6 polyhedra is interconnected with other TiO6 by vertex shared O-atom with Ti-O-Ti
angle of 142.52° at 0 GPa (as shown in Fig. S3). With increasing pressure, this angle reduces
as TiO6 polyhedra are tilted to a greater extent. At 34 GPa, a sharp reduction in Ti-O-
Ti angle from 132.83° to 123.91° is identified leading to the change in lattice angle at that
pressure. The symmetry of the transformed structure at 34 GPa is identified as a triclinic
structure with a center of inversion (P1̄). Experimentally, 45 a similar C 2/c → P1̄ structural
transformation is indicated at 10.2 GPa pressure.
A comparison of TiO6 polyhedral arrangement of structures at 32 GPa (C 2/c) and 34
GPa (P1̄) reveals identical Ti-Ti distance along [110] and [11̄0] of 5.21 Å at 32 GPa however,
these distances are different in P1̄ structure (as shown in Fig. 3). It is observed that in
P1̄, Ti-Ti distances are different along the [110] direction and the [11̄0] direction leading
to two different Ti-Ti interatomic separations (5.12 Å and 5.40 Å) and loss of glide planes.
As a result, in P1̄ structure Ti atoms occupy two distinct Wyckoff sites rather than one
in C 2/c structure. This indicates P 21 /c → P1̄ phase transition is displacive-type and a
similar pressure-induced structural transition mechanism has been identified in CaSnSiO5
(malayaite) at 5.8 GPa pressure. 5
Another phase transition at 42 GPa with discontinuous change in volume, lattice pa-
rameters, and lattice angles is identified (Fig. 2). A unit cell volume change of 4.25% is
observed at 42 GPa compared to its 40 GPa structure. A sharp discontinuous change in
lattice parameters is also observed at 42 GPa. The space group of 42 GPa crystal structure
is determined to be C 2/c from the symmetry analysis, however, the chemical environment
is different from the previous low-pressure C 2/c structures predicted below 34 GPa. At 42
GPa, there is a change in the coordination around Si atoms from SiO4 tetrahedra to SiO6
octahedra, although, the coordination environment remains the same around Ti atoms. The
coordination number of Ca also changes and CaO7 polyhedral rearrangement leads to the
formation of CaO9 and these signatures refer to the reconstructive nature of the phase tran-
sition. The SiO6 octahedrons are interconnected with each other by sharing edges with a

12
Si-Si bond at a distance of 2.64 Å (at 42 GPa) which is unlikely in TiO6 octahedral chains
propagated by sharing vertex O atoms. In the deep mantle of Earth’s interior, higher coor-
dinated Si is found for molten SiO2 . 46 Experimentally, edged-sharing SiO6 octahedra is also
observed at 29.1 GPa for high-pressure SiO2 stishovite structure in which Si-Si internuclear
separation is 2.619 Å which is incidentally close to the calculated Si-Si bond distances of 42
GPa optimized structure.
To further inspect the nature of phase transitions, two additional sets of calculations are
performed where structural optimizations are conducted starting from 42 GPa and 34 GPa
structures with a reduction of pressures shown as a blue triangle in the pressure-volume
plot of Fig. 2. As the pressure reduces from 42 GPa to 32 GPa, no substantial changes in
the structure are observed. All the bond distances (Ca-O, Ti-O, and Si-O) increase as the
pressure releases from 42 GPa and there is no change in the symmetry of the structure. Our
calculations suggest that C 2/c structure does not fall back to P1̄ structure due to the hys-
teresis effect which indicates a reconstructive-type phase transition at 42 GPa. 47 In the case
of C 2/c → P1̄ phase transition at 34 GPa, with decreasing pressure the volume at 28 GPa
(blue open triangle in Fig. 2) almost overlaps with the initial optimized structure at 28 GPa
(red open circle in Fig. 2) and might be a signature of displacive-type phase transition. 48 It
is also important to note that structural optimization calculations with continuous increases
in pressure always suffer from hysteresis leading to over-prediction of transition pressures
and limited exploration of the potential energy landscape. However, the phase transition
identified in this process can be regarded as an indication of possible phase transitions and
favorable chemical environments at high pressures. Hence, an exhaustive potential energy
landscape search at high pressure is customary to find thermodynamically stable phases.
The next section shows the results of new HP structures and their thermodynamic stability
predicted from stochastic and evolutionary structure search algorithms.

13
High pressure crystal structure prediction (CSP)

To establish the methodology of evolutionary structure search algorithms using USPEX


in conjunction with GULP, initially, structure predictions have been made at 0 GPa. To
find the suitability of interatomic potential parameters considered in this calculation, the
GULP-calculated lattice parameters, volumes, Ca-O, Si-O, and Ti-O bond distances are
compared with DFT-GGA2 calculated values and experiments (Table S2). A reasonable
agreement is found between the calculated values and experiments. Fig. 4(a) shows enthalpies
of the USPEX-GULP predicted structures at 0 GPa arranged with respect to the most
stable structure. The space group of the ground state structure matches with experimentally
reported P 21 /c structure with lattice parameters shown in Table S2. The second lowest
enthalpy structure is 460 meV/fu higher in enthalpy than the ground state. The large
separation of enthalpy at 0 GPa between the ground state structure and the second lowest
enthalpy structure provides additional confirmation for the suitability of the Buckingham
potential parameters and the implementation of the USPEX-GULP conjunction for the
structure search.
The USPEX-GULP implementation is further expanded to predict high-pressure crystal
structures at 10, 20, 30, 40, and 50 GPa. The GULP calculated enthalpies of the USPEX
predicted crystal structures are shown in Fig. 4(a) and enthalpies of the structures lying
within 1 eV/unitcell range (with respect to the ground state at each pressure) are also shown
in Fig. 4(b). At each pressure, structural optimization is conducted on the structures lying
within the 1 eV/ unit-cell enthalpy range using DFT-GGA2. Fig. 4(b) compares enthalpies
of the GULP and DFT-GGA2 optimized structures and the top three structures are marked
in each case. In the case of 10 and 30 GPa, the relative ranking for the best three structures
remains the same in GULP and DFT, however, this ranking differs at other pressures. It is
important to note that the DFT-GGA2 calculated enthalpy differences between the second
rank structure and the lowest enthalpy structure is several hundreds of meV/unitcell higher
at any pressure. The structural parameters of the most stable structures optimized using

14
DFT-GGA2 are shown in the Supporting Information (Table S4).
20 1 5

1
0.8 2 3 4
15 1 3
2 1

∆HGULP [eV/cell]

∆HDFT [eV/cell]
0.6 3 3
∆H [eV/cell]

10 3
3
2
3
0.4 3 2
2 2
3
2
5
0.2 1
2
3
2 2
3
21 1 1 1 1 1 1
0 0 0
a) 0 10 20 30 40 50 b) 10 20 30 40 50
Pressure (GPa) Pressure (GPa)

Figure 4: (a) USPEX predicted structures at different pressures are ranked according to the
GULP calculated formation enthalpy. The enthalpies within 1 eV/unitcell range are shown
in a red box. (b) The relative ranking of those structures (shown in red box) is compared
by GULP (in red) and DFT (in blue) calculated enthalpies. The first three lowest enthalpy
structures are sequentially numbered.

At 10 GPa, a triclinic structure with space group P1̄ (CSP-10GPa) is predicted as the
ground state and this structure has a remarkable similarity with the structure generated in
C 2/c → P1̄ structural transformation at 34 GPa predicted from our pressure-volume study.
To determine mechanical stability, single-crystal elastic constants are calculated as shown in
Table 2 and found that the predicted triclinic structure is mechanically stable at 10 GPa. It
is worth mentioning that a similar phase transformation has been observed in a single high-
pressure study and referred to occur at 10.2 GPa. 45 In Supporting Information (Table S3)
a comparison of structural parameters of USPEX predicted P1̄ structure at 10 GPa and ex-
perimentally reported structure at 10.2 GPa is provided. Remarkable matching in structural
parameters again emphasizes the capability of the USPEX-GULP-DFT implementation to
predict accurate structures not only at 0 GPa but also at higher pressures. The triclinic P1̄
structure is topologically similar to the ground state P 21 /c structure. The SiO4 tetrahedra
are completely isolated from each other but connected to the chains of TiO6 octahedra (see
Fig. 5 (a) and (b)).

15
Figure 5: The Ti-O and Si-O polyhedral rearrangement of USPEX predicted and DFT
optimized structure at 0 GPa (a), 10/20 GPa (b), 30 GPa (c), 40 GPa (d), and 50 GPa (e).
There is a change of Si coordination environment from SiO4 to SiO6 at high pressures. For
visualization, Ca-atoms are not shown.

At 20 GPa a triclinic P1̄ structure (CSP-20GPa) is predicted which is identical to the 10


GPa predicted structure (see Table S4). The predicted structure at 30 GPa (CSP-30GPa,
SG - Pbam) and 50 GPa (CSP-50GPa, SG - Pnma) are orthorhombic, however, a monoclinic
structure (CSP-40GPa, P 21 /c) is predicted at 40 GPa. A rearrangement of SiO4 tetrahedra
to SiO6 octahedra is observed in USPEX-predicted structures at a pressure of 30 GPa and
higher. The topology of SiO6 octahedra is different in the USPEX predicted high-pressure
structures. In the CSP-30GPa structure, SiO6 octahedra makes an extended network by
sharing polyhedra edges along [001] direction. In CSP-40GPa structure, there is no long-
range connectivity as two combined SiO6 octahedra are isolated. The topology of CSP-50GPa
structure is a combination of corner shared distorted and zigzag SiO6 octahedral chains
(along [100] direction) and edge shared TiO6 prism (perpendicular to the SiO6 octahedral
chain) as shown in Fig. 5(e). The SiO6 octahedra and TiO6 prism are tilted so that Ca
atoms can fit into the interstitial space. Similar topological arrangements of zigzag SiO6
octahedra network are also found in orthorhombic MgSiO3 phase (SG - Pbnm) at 10.6 GPa
pressure. 49 However, zigzag SiO6 chains in MgSiO3 are cross-linked with each other which is
missing in the CSP-50GPa structure and makes it a topologically distinct prototype among
high-pressure silicates.

16
Thermodynamic and electronic properties of P21 /c, P1̄, and Pnma

structures

P21/c
0.25
CSP−10GPa
CSP−20GPa
0.15 CSP−30GPa
CSP−40GPa
Relative Enthalpy (eV/atom)

CSP−50GPa
0.05

−0.05

−0.15
P21/c
C2/c

−0.25 P1
Pnma

−0.35 0 4 8 12 16 20 24 28 32 36 40 44 48

0 4 8 12 16 20 24 28 32 36 40 44 48 52
Pressure (GPa)

Figure 6: Comparison of relative enthalpy of USPEX-predicted structures at high pressures.


The difference in enthalpy between ground state P 21 /c and P1̄ (CSP-10GPa) phases is neg-
ligibly small. However, one of the HP structures, Pnma (CSP-50GPa) is thermodynamically
most favorable beyond 18 GPa as shown in the figure inset.

Fig. 6 compares the enthalpies of the USPEX predicted structures (CSP-XGPa, with
X= 10, 20. 30, 40 and 50) with respect to ground state P 21 /c structure over the 0-52
GPa pressure range. The enthalpy difference between P 21 /c, C2/c and P1̄ structures is
small (2 meV/unit cell) and not resolvable in the enthalpy range shown in Fig. 6. The
inset figure shows the stability ranges of the CaTiSiO5 structures as a function of pressure
obtained in this study. The Pnma structure is 170 meV/atom higher in enthalpy than the
P 21 /c structure at 0 GPa. However, as pressure increases, Pnma becomes enthalpically
more favorable than all other structures, and beyond 18 GPa, it is thermodynamically the
most stbale structure.
Fig. 7 compares the total and atom-projected Density of States (DOS) of P 21 /c (at 0

17
8 8 8

6 6 6
a) b) c)
4 4 4

2 2 2
Energy [eV]

Energy [eV]

Energy [eV]
0 0 0

−2 −2 −2

−4 −4 −4
Total
−6 −6 −6
Ca
Ti
−8 −8 Si −8
O
−10 −10 −10
0 0.05 0.1 0.15 0 0.05 0.1 0.15 0 0.05 0.1 0.15

DOS [ states/eV/e ]
8 8 8

6 6 6
d) e) f)
4 4 4

2 2 2
Energy [eV]

Energy [eV]

Energy [eV]
0 0 0

−2 −2 −2

−4 −4 −4
Ca−O
Ti−O
−6 −6 Si−O −6

−8 −8 −8

−10 −10 −10


−1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5
−pCOHP

Figure 7: At top, DOS of P 21 /c at 0 GPa (a), P1̄ at 10 GPa (b), and Pnma at 20 GPa (c).
The horizontal black dashed line at 0 eV denotes the Fermi level and the energy of bands are
scaled according to the Fermi level. GGA2 calculated band gap for P 21 /c, P1̄, and Pnma
structures is 3.0, 3.1, and 2.2 eV respectively. At bottom, projected COHP (-pCOHP) of
P 21 /c at 0 GPa (d), P1̄ at 10 GPa (e), and Pnma at 20 GPa (f) are shown.

GPa), P1̄ (at 10 GPa), and Pnma (at 20 GPa). In the analysis of the DOS the valence band
(VB) from -10 eV upto the Fermi level is considered and 3s3p orbital of Ca and Ti atoms are
not shown. In the vicinity of the Fermi level, the VB is dominated by O 2p-states as there
are charge transfers from metal ions to O-orbitals. The characteristic feature of DOS for
both P 21 /c and P1̄ is identical as these structures have similar structural topology. In Pnma
structure, projected Si-O DOS is dispersed than P 21 /c and P1̄ which is expected as Si-O
bonds are elongated within the distorted SiO6 octahedra (1.76 Å) in Pnma (at 20 GPa) in
comparison with P 21 /c and P1̄ structures (1.64 Å). According to our calculations, the band
gap of Pnma structure is smaller than P 21 /c and P1̄ which indicates that Pnma structure

18
is less insulating than other two structures. The electron localization function (ELF) of
three structures is shown in the Supporting Information (see Fig. S4). As a result of charge
transfer from metal ions to O atoms, electrons are completely localized on the O-atoms.
In the case of Pnma structure, the electron density gets perturbed around those O-atoms
which are shared by two distorted SiO6 octahedra (along [100] direction) as mentioned in
the previous section. A positive value of projected Crystal Orbital Hamilton Population
(pCOHP) indicates a bonding state whereas the negative value denotes an anti-bonding
state (Fig. 7 (d)-(f)). Like DOS, the characteristic feature of pCOHP is comparable for
monoclinic P 21 /c and triclinic P1̄ structures which justifies a close crystal chemical behavior
among these two structures. A major difference in pCOHP is found when we compare the
degree of covalency of Si-O interactions. The integrated value of pCOHP [IpCOHP] measures
the extent of covalency in bonds. The calculated IpCOHP for Pnma is 5.6 however, this
value is significantly larger in P 21 /c (IpCOHP = 7.4) and P1̄ (IpCOHP = 7.6) structures.
It is expected as Si-O bonds become weaker due to the elongation of bonds in the Pnma
structure.
The Bader charge distribution 50,51 analysis of P 21 /c, P1̄, and Pnma shows net positive
charge on Ca, Ti, and Si atoms varies within 1.54 – 1.60, 2.15 – 2.28, and 3.12 – 3.25,
respectively. In Pnma, there are three chemically distinct O-atoms. The charge on these
O atoms is -1.31, -1.36, and -1.51 respectively. We have compared the charge distribution
on metal ions of these structures with those on ground states of CaO (SG-Fm3̄m), SiO2
(SG-P31 21 ), and TiO2 (SG-I41 /amd ) structures. The charge of Ca (in Ca2+ O2− ), Ti (in
Ti4+ O2 2− ) and Si (in Si4+ O2 2− ) are +1.48, +2.22, and +3.20, respectively. There is a close
resemblance between the charges of elements of these individual oxides and these predicted
structures. From this comparison, we can infer that the oxidation state of Ca, Ti, Si, and O
atoms of the predicted structures are +2, +4, +4, and -2 respectively.
We have calculated phonons for both P 1̄ and Pnma structures at 10 and 20 GPa with
2×1×2 and 2×2×1 supercell containing 128 atoms. The phonon dispersion plots of P 1̄

19
1200 1200
a) b)

1000 1000

800 800
Frequency (cm−1)

Frequency (cm−1)
600 600

400 400

200 200

0 0
Γ X|Z Γ T|U Γ Γ X S Y Γ Z U R T Z

Figure 8: Phonon band dispersion of P 1̄ at 10 GPa (a) and Pnma at 20 GPa (b).

and Pnma structure are shown in Fig. 8 along specific high symmetric q-points of the
first Brillouin zone. As Pnma is a stable structure beyond 18 GPa, the phonon dispersion is
calculated at 20 GPa. The absence of any imaginary phonon modes in the phonon dispersion
plot indicates the dynamical stability of these structures at the corresponding pressures. In
Table 2, we have shown the calculated elastic constants (Cij ) for this structure. There are
nine elastic constants (six diagonal and three off-diagonal) and all of them are positive. The
Pnma structure satisfies all the necessary criteria of mechanical stability 44 and hence it is
mechanically stable as well. The elastic constants C33 > C11 > C22 , indicate that compression
along b-axis is easier than a- and c-axes. It is also justified by the lattice parameter-pressure
plot of Pnma (see Fig. S5) as lattice parameter b decreases faster than lattice parameters a
and c with increasing pressure. Our calculations indicate that the predicted Pnma structure
is not only thermodynamically but also dynamically and mechanically stable. The predicted
Pnma structure is a suitable candidate for experimental validation and can be classified as
a new structural prototype among HP silicates.

20
Conclusions

High-pressure study of titanite-type CaTiSiO5 using an evolutionary algorithm and den-


sity functional theory calculations in the 0-50 GPa range is conducted to explore structural
phase transformations at HP. The analysis of DFT calculated pressure variation of Ti-O
bond lengths identifies P 21 /c → C 2/c phase transition at 8 GPa which is in agreement with
experimental observation of a second-order structural phase transition at 6.9 GPa pressure
and room temperature. 16 It is also identified that the GGA can distinguish small structural
differences between P 21 /c and C 2/c. Implementation of the structure search method using
evolutionary algorithms and DFT calculations reveals two new structural phase transitions:
displacive-type C 2/c → P1̄ and reconstructive-type P1̄ → Pnma at 10 GPa and 18 GPa,
respectively. The mechanical and dynamical stability of the new structures has also been
substantiated to find stability ranges. The topology of the Pnma structure can be identified
as a combination of corner shared distorted and zigzag SiO6 octahedra along [100] direction
and edge shared TiO6 prism (with SiO6 ) along [001] direction. This novel topological ar-
rangement has not been observed in any of the high-pressure structural prototypes reported
in the literature. We believe that our work provides a unique insight into the high-pressure
structural polymorphs of titanite-type important silicate minerals and will exhort further ex-
perimental efforts to validate our predictions to shed more light on the structural intricacies
at high pressures.

Acknowledgments

All the calculations were performed in the High Performance Computing Center of BARC,
India. S.C acknowledges to the Department of Atomic Energy (DAE), India for providing
postdoctoral research associateship.

21
References

(1) Heaney, P. J.; Prewitt, C. T.; Gibbs, G. V. Silica: Physical behavior, geochemistry, and
materials applications; Walter de Gruyter GmbH & Co KG, 2018; Vol. 29.

(2) Hazen, R. M.; Downs, R. T.; Finger, L. W. High-Pressure Framework Silicates. Science
1996, 272, 1769–1771.

(3) Angel, R. J.; Ross, N. L.; Seifert, F.; Fliervoet, T. F. Structural characterization of
pentacoordinate silicon in a calcium silicate. Nature 1996, 384, 441–444.

(4) Malcherek, T.; Bosenick, A. Structure and phase transition of CaGe2 O5 revisited. Phys.
Chem. Miner. 2004, 31, 224–231.

(5) Rath, S.; Kunz, M.; Miletich, R. Pressure-induced phase transition in malayaite,
CaSnOSiO4 . Am. Mineral. 2003, 88, 293–300.

(6) Finger, L.; Hazen, R. High-Temperature and High Pressure Crystal Chemistry; The
Mineralogical Society of America, 2000; Vol. 41; pp 123–156.

(7) Deer, W.; Deer, W.; Howie, R.; Zussman, J. Rock-Forming Minerals: Orthosilicates,
Volume 1A; Rock-forming minerals; Geological Society, 1982.

(8) Okrusch, M.; Frimmel, H. E. Mineralogy: An introduction to minerals, rocks, and


mineral deposits; Springer Nature, 2020.

(9) Kohn, M. J. Titanite Petrochronology. Rev. Mineral. Geochem. 2017, 83, 419–441.

(10) Hayward, P. J.; Cann, C. D.; Mitchell, S. L.; Vance, E. R. Crystallization of sphene-
based glass-ceramics for immobilization of high-level nuclear fuel reprocessing wastes;
American Ceramic Society: United States, 1984.

22
(11) Frost, B.; Chamberlain, K. R.; Schumacher, J. C. Sphene (titanite): phase relations
and role as a geochronometer. Chem. Geol. 2001, 172, 131–148, What are we dating?
Understanding the Crystallogernesis of U-Pb.

(12) Scibiorski, E.; Kirkland, C.; Kemp, A.; Tohver, E.; Evans, N. Trace elements in titanite:
A potential tool to constrain polygenetic growth processes and timing. Chem. Geol.
2019, 509, 1–19.

(13) Bauer, J. D.; Labs, S.; Weiss, S.; Bayarjargal, L.; Morgenroth, W.; Milman, V.;
Perlov, A.; Curtius, H.; Bosbach, D.; Zänker, H.; Winkler, B. High-Pressure Phase
Transition of Coffinite, USiO4 . J. Phys. Chem. C 2014, 118, 25141–25149.

(14) Strzelecki, A. C.; Zhao, X.; Baker, J. L.; Estevenon, P.; Barral, T.; Mesbah, A.;
Popov, D.; Chariton, S.; Prakapenka, V.; Ahmed, S.; Yoo, C.-S.; Dacheux, N.; Xu, H.;
Guo, X. High-Pressure Structural and Thermodynamic Properties of Cerium Orthosil-
icates (CeSiO4 ). J. Phys. Chem. C 2023, 127, 4225–4238.

(15) Taylor, M.; Brown, G. E. High-temperature structural study of the P21/a → A2/a
phase transition in synthetic titanite, CaTiSiO5 . Am. Mineral. 1976, 61, 435–447.

(16) Kunz, M.; Xirouchakis, D.; Lindsley, D. H.; Hausermann, D. High-pressure phase tran-
sition in titanite (CaTiOSiO4 ). Am. Mineral. 1996, 81, 1527–1530.

(17) Angel, R. J.; Kunz, M.; Miletich, R.; Woodland, A. B.; Koch, M.; Xirouchakis, D.
High-pressure phase transition in CaTiOSiO4 titanite. Phase Transitions 1999, 68,
533–543.

(18) Malcherek, T. Spontaneous strain in synthetic titanite, CaTiOSiO4 . Mineral. Mag.


2001, 65, 709–715.

(19) Ghose, S.; Ito, Y.; Hatch, D. M. Paraelectric-antiferroelectric phase transition in titan-
ite, CaTiSiO5 . Phys. Chem. Miner. 1991, 17, 591–603.

23
(20) Speer, J. A.; Gibbs, G. V. The crystal structure of synthetic titanite, CaTiOSiO4 , and
the domain textures of natural titanites. Am. Mineral. 1976, 61, 238–247.

(21) Zhang, M.; Salje, E.; Bismayer, U. Structural phase transition near 825 K in titanite:
Evidence from infrared spectroscopic observations. Am. Mineral. 1997, 82, 30–35.

(22) Malcherek, T.; Domeneghetti, C. M.; Tazzoli, V.; Salje, E. K. H.; Bismayer, U. A
high temperature diffraction study of synthetic titanite CaTiOSiO4 . Phase Transitions
1999, 69, 119–131.

(23) Malcherek, T.; Fischer, M. Phase transitions of titanite CaTiSiO5 from density func-
tional perturbation theory. Phys. Rev. Mater. 2018, 2, 023602.

(24) Gutmann, M. J.; Refson, K.; Zimmermann, M.; Swainson, I. P.; Dabkowski, A.;
Dabkowska, H. Room temperature single-crystal diffuse scattering and ab initio lat-
tice dynamics in CaTiSiO5 . J. Phys. Condens. Matter 2013, 25, 315402.

(25) Das, P. K.; Mondal, S. K.; Mandal, N. First principles prediction of exceptional me-
chanical and electronic behaviour of Titanite (CaTiSiO5 ). Materialia 2021, 15, 100964.

(26) Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calcu-
lations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186.

(27) Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B
1993, 47, 558–561.

(28) Ceperley, D. M.; Alder, B. J. Ground State of the Electron Gas by a Stochastic Method.
Phys. Rev. Lett. 1980, 45, 566–569.

(29) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made
Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.

(30) Oganov, A. R.; Glass, C. W. Crystal structure prediction using ab initio evolutionary
techniques: Principles and applications. J. Chem. Phys. 2006, 124, 244704.

24
(31) Zhu, Q.; Oganov, A. R.; Glass, C. W.; Stokes, H. T. Constrained evolutionary algo-
rithm for structure prediction of molecular crystals: methodology and applications.
Acta Crystallogr. B. 2012, 68, 215–226.

(32) Glass, C. W.; Oganov, A. R.; Hansen, N. USPEX—Evolutionary crystal structure


prediction. Comput. Phys. Commun. 2006, 175, 713–720.

(33) Oganov, A. R.; Lyakhov, A. O.; Valle, M. How Evolutionary Crystal Structure Predic-
tion Works—and Why. Acc. Chem. Res. 2011, 44, 227–237.

(34) Gale, J. D.; Rohl, A. L. The General Utility Lattice Program (GULP). Mol. Simul.
2003, 29, 291–341.

(35) Guillot, B.; Sator, N. A computer simulation study of natural silicate melts. Part I:
Low pressure properties. Geochim. Cosmochim. Acta 2007, 71, 1249–1265.

(36) Wang, M.; Krishnan, N. A.; Wang, B.; Smedskjaer, M. M.; Mauro, J. C.; Bauchy, M. A
new transferable interatomic potential for molecular dynamics simulations of borosili-
cate glasses. J. Non-Cryst. Solids 2018, 498, 294–304.

(37) Togo, A.; Oba, F.; Tanaka, I. First-principles calculations of the ferroelastic transition
between rutile-type and CaCl2 -type SiO2 at high pressures. Phys. Rev. B 2008, 78,
134106.

(38) Parlinski, K.; Li, Z. Q.; Kawazoe, Y. First-Principles Determination of the Soft Mode
in Cubic ZrO2 . Phys. Rev. Lett. 1997, 78, 4063–4066.

(39) Stokes, H. T.; Hatch, D. M. FINDSYM: program for identifying the space-group sym-
metry of a crystal. J. Appl. Crystallogr. 2005, 38, 237–238.

(40) Deringer, V. L.; Tchougréeff, A. L.; Dronskowski, R. Crystal orbital Hamilton popula-
tion (COHP) analysis as projected from plane-wave basis sets. J. Phys. Chem. A 2011,
115, 5461–5466.

25
(41) Nelson, R.; Ertural, C.; George, J.; Deringer, V. L.; Hautier, G.; Dronskowski, R. LOB-
STER: Local orbital projections, atomic charges, and chemical-bonding analysis from
projector-augmented-wave-based density-functional theory. J. Comput. Chem. 2020,
41, 1931–1940.

(42) Momma, K.; Izumi, F. VESTA3 for three-dimensional visualization of crystal, volumet-
ric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276.

(43) Kunz, M.; Arlt, T.; Stolz, J. In situ powder diffraction study of titanite (CaTiOSiO4 )
at high pressure and high temperature. Am. Mineral. 2000, 85, 1465–1473.

(44) Mouhat, F.; Coudert, F.-X. Necessary and sufficient elastic stability conditions in var-
ious crystal systems. Phys. Rev. B 2014, 90, 224104.

(45) Rath, S. Comparative pressure-dependent structural studies of titanite (CaTiOSiO4 )


and malayaite (CaSnOSiO4 ). Ph.D. thesis, Swiss Federal Institute of Technology Zurich,
2002.

(46) Bykova, E.; Bykov, M.; Černok, A.; Tidholm, J.; Simak, S. I.; Hellman, O.; Belov, M.;
Abrikosov, I. A.; Liermann, H.-P.; Hanfland, M., et al. Metastable silica high pressure
polymorphs as structural proxies of deep Earth silicate melts. Nat. Commun. 2018, 9,
4789.

(47) Ghosh, P. S.; Ali, K.; Arya, A. A computational study of high pressure polymorphic
transformations in monazite-type LaPO4 . Phys. Chem. Chem. Phys. 2018, 20, 7621–
7634.

(48) Ghosh, P. S.; Arya, A.; Tewari, R.; Dey, G. Alpha to omega martensitic phase trans-
formation pathways in pure Zr. J. Alloys Compd. 2014, 586, 693–698.

(49) Ross, N. L.; Hazen, R. M. High-pressure crystal chemistry of MgSiO3 perovskite. Phys.
Chem. Miner. 1990, 17, 228–237.

26
(50) Bader, R. F. Atoms in molecules. Acc. Chem. Res. 1985, 18, 9–15.

(51) Tang, W.; Sanville, E.; Henkelman, G. A grid-based Bader analysis algorithm without
lattice bias. J. Phys. Condens. Matter 2009, 21, 084204.

27
Supporting Information: New High Pressure
Phases of Titanite-type CaTiSiO5 Predicted from
First-Principles

Subhamoy Char,† P. S. Ghosh,∗,†,‡ Kawsar Ali,† and A. Arya†,‡

†Glass & Advanced Materials Division, Bhabha Atomic Research Center, Mumbai 400 085,
India
‡Homi Bhabha National Institute, Anushaktinagar, Mumbai 400 094, India

E-mail: psghosh@barc.gov.in

Crystal structure prediction by USPEX and GULP

In this article, we have predicted crystal structures for CaTiSiO5 from random and evolu-
tionary algorithm which is implemented in the USPEX package. As the unit cell contains 32
atoms, initial screening of structure search with first-principles based method is time demand-
ing and to surpass the high time-cost, we have computed the energy of predicted structures
by GULP package which is interfaced with USPEX. The USPEX-GULP package was jointly
adopted by several authors either to explore bulk crystal structures for PuCrO3 /PuAlO3 1 or
to predict effective surface orientations for Ga2 O3 /Al2 O3 compositions. 2 In this article, we
considered Buckingham potential for short range interaction and the required parameters are
tabulated in Table S1. The parameters were developed by Guillot et al 3 and these were ad-
justed to reproduce the experimental densities of several silicate minerals which are normally
found in magmatic melts. The liquid state of these silicate minerals is sufficiently enriched by
short range order and hence, these parameters are efficient to deal nearest neighbor chemical

S-1
interactions between different elements. We have computed lattice parameters and volume of
CaTiSiO5 with Buckingham potential parameters and compared with experimental values.
(as shown in Table S2)

Table S1: Buckingham potential parameters for different atomic interactions. The parame-
ters are used in different silicate glass. 3,4 The atomic charges are mentioned at the superscript
of elements. The cutoff of interaction was set to 10 Å.

Atom-pair Aij (eV) ρij (Å) cij (eV.Å6 )


Ca+0.945 -O−0.945 155667.70 0.178 42.2597
Ti+1.890 -O−0.945 50126.64 0.178 46.2978
Si+1.890 -O−0.945 50306.10 0.161 46.2978
O−0.945 -O−0.945 9022.79 0.265 85.0921

Table S2: Lattice parameter and internuclear separation of P 21 /c structure calculated from
GULP as well as DFT and compared with available experimental data.

Volume
Lattice parameter (a, b, c, β) dSi−O (Å) dT i−O (Å) dCa−O (Å)
(Å3 )
1.916, 1.916, 2.324, 2.409, 2.409,
1.628, 1.628,
GULP 7.213, 8.803, 6.554, 113.89 380.52 2.009, 2.009, 2.451, 2.451, 2.776,
1.630, 1.630
2.021, 2.021 2.776
1.833, 1.931, 2.274, 2.425, 2.434,
1.648, 1.648,
DFT 7.130, 8.766, 6.658, 114.09 379.92 1.995, 1.997, 2.463, 2.470, 2.599,
1.654, 1.654
2.039, 2.045 2.634
1.772, 1.971, 2.270, 2.389, 2.406,
1.624, 1.634,
exp. 5 7.068, 8.714, 6.562, 113.82 369.73 1.974, 1.986, 2.427, 2.433, 2.585,
1.654, 1.659
2.010, 2.049 2.678

S-2
Table S3: Comparison of the structural parameters between USPEX predicted structure
(CSP-10GPa) at 10 GPa and experimentally reported triclinic structure at 10.2 GPa.

USPEX predicted (CSP-10GPa) Experiment 6


Space Group P 1̄ P 1̄
a (Å) 6.415 6.362
b (Å) 8.639 8.604
c (Å) 6.910 6.841
α(◦ ) 91.25 90.02
β(◦ ) 113.06 112.81
γ(◦ ) 89.13 89.98
Volume (Å3 ) 352.30 345.20
Ti1-O (Å) 1.848, 1.968, 2.013 1.835, 1.968, 2.001
Ti2-O (Å) 1.842, 1.993, 2.002 1.839, 1.984, 1.999
Si-O (Å) 1.631, 1.631, 1.636, 1.644 1.595, 1.617, 1.625, 1.634
2.237, 2.333, 2.349, 2.364, 2.424, 2.216, 2.317, 2.337, 2.366, 2.374,
Ca-O (Å)
2.562, 2.633 2.582, 2.608
Ti1-O1-Ti2 (◦ ) 138.87 137.15
CaO7 polyhedra
18.55 18.39
volume (Å3 )
TiO6 polyhedra
9.78 9.69
volume (Å3 )
SiO4 polyhedra
2.23 2.16
volume (Å3 )

S-3
Table S4: Optimized structural parameters for high pressure predicted structures of
CaTiSiO5 composition with Z=4 formula units.

Wyckoff position x y z

CSP-10GPa (SG - P 1̄)


a = 6.415, b = 8.639, c = 6.911
α = 91.2, β = 113.1, γ = 89.1
Ca1 4i 0.74358 0.08399 0.76610
Ti1 2g 0.25000 0.25000 0.50000
Ti2 2c 0.25000 0.25000 0.00000
Si1 4i 0.75287 0.06770 0.25441
O1 4i 0.24962 0.17496 0.24732
O2 4i 0.56771 0.17684 0.07471
O3 4i 0.43137 -0.30989 0.41638
O4 4i 0.35228 0.03662 0.61199
O5 4i 0.13259 0.04226 0.87363

CSP-20GPa (SG - P 1̄)


a = 6.255, b = 8.519, c = 6.766
α = 93.2, β = 112.8, γ = 87.9
Ca1 4i 0.73689 0.08777 0.78417
Ti1 2c 0.25000 0.25000 0.00000
Ti2 2g 0.25000 0.25000 0.50000
Si1 4i 0.75662 0.06799 0.26558
O1 4i 0.23649 0.17055 0.23886
O2 4i 0.42805 -0.30177 0.42652
O3 4i 0.57069 0.16841 0.06970
O4 4i 0.11891 0.04704 0.85857
O5 4i 0.35046 0.03389 0.60152

S-4
CSP-30GPa (SG - Pbam)
a = 6.115, b = 16.726, c = 2.795
α = 90.0, β = 90.0, γ = 90.0
Ca1 4h 0.00153 0.74966 0.50000
Ti1 4h 0.74961 0.08505 0.50000
Si1 4g 0.75061 -0.09554 0.00000
O1 4g 0.75235 0.80474 0.00000
O2 4h 0.06110 0.09306 0.50000
O3 4h 0.06203 0.59426 0.50000
O4 4g 0.75011 0.16279 0.00000
O5 4g 0.74815 0.00767 0.00000

CSP-40GPa (SG - P 21 /c)


a = 6.733, b = 4.852, c = 10.293
α = 90.0, β = 127.3, γ = 90.0
Ca1 4e 0.51830 0.54908 0.14258
Ti1 4e 0.09244 -0.02830 0.69343
Si1 4e 0.79549 -0.04549 -0.00670
O1 4e 0.51530 0.45456 0.65255
O2 4e 0.18521 0.74630 0.40447
O3 4e 0.80073 0.75393 0.13419
O4 4e 0.18112 0.36060 0.07737
O5 4e 0.86297 0.10199 0.86892

CSP-50GPa (SG - Pnma)


a = 6.332, b = 4.828, c = 8.604
α = 90.0, β = 90.0, γ = 90.0
Ca1 4c 0.86333 0.25000 0.63430

S-5
Ti1 4c 0.85815 0.25000 -0.02653
Si1 4c 0.39666 0.25000 0.22337
O1 8d -0.04332 0.50720 0.13809
O2 8d 0.33882 0.00307 0.09434
O3 4c 0.15924 0.25000 0.32710

12 10

Lattice parameter (Å)


9
Volume (Å3/atom )

11
8
10
7
9 LDA1 6 a(LDA1) c(LDA1) b(LDA2)
LDA2 b(LDA1) a(LDA2) c(LDA2)
8 5
0 2 4 6 8 10 0 2 4 6 8 10

130 1.90

120

110 1.86
dTi−O (Å)
Angle (°)

100

90 1.82
α(LDA1) γ(LDA1) β(LDA2) Ti−O1(LDA1) Ti−O1(LDA2)
80
β(LDA1) α(LDA2) γ(LDA2) Ti−O2(LDA1) Ti−O2(LDA2)
70 1.78
0 2 4 6 8 10 0 2 4 6 8 10
Pressure (GPa) Pressure (GPa)

Figure S1: Phase transition up to 10 GPa with two different pseudo-potentials (LDA1 and
LDA2) with LDA functional. LDA1 indicates Ca pv (3p6 4s2 ) and Ti (4s1 3d3 ) pseudopo-
tentials whereas LDA2 indicates Ca sv (3s2 3p6 4s2 ) and Ti sv (3s2 3p6 4s1 3d3 ). The phase
transition below 10 GPa is not correctly reproduced with LDA functional. Solid square de-
notes experimental volume and lattice parameter. 7

S-6
12 10

Lattice parameter (Å)


9
Volume (Å3/atom )

11
8
10
7
9 GGA1 6 a(GGA1) c(GGA1) b(GGA2)
GGA2 b(GGA1) a(GGA2) c(GGA2)
8 5
0 2 4 6 8 10 0 2 4 6 8 10

130 1.95

120

110 1.90
Angle (°)

dTi−O
100

90 1.85
α(GGA1) γ(GGA1) β(GGA2) Ti−O1(GGA1) Ti−O1(GGA2)
80 Ti−O2(GGA1) Ti−O2(GGA2)
β(GGA1) α(GGA2) γ(GGA2)
70 1.80
0 2 4 6 8 10 0 2 4 6 8 10
Pressure (GPa) Pressure (GPa)

Figure S2: Phase transition up to 10 GPa with two different sets of pseudo-potentials (GGA1
and GGA2) with GGA functional. GGA1 indicates Ca pv (3p6 4s2 ) and Ti (4s1 3d3 ) pseu-
dopotentials whereas GGA2 indicates Ca sv (3s2 3p6 4s2 ) and Ti sv (3s2 3p6 4s1 3d3 ). GGA2
set provides better representation for P 21 /c → C 2/c phase transition below 10 GPa. Solid
square denotes experimental volume and lattice parameter. 7

Figure S3: The selected Ti-O-Ti angle between two TiO6 octahedra which follows a sharp
change at the onset of C 2/c → P 1̄ and P 1̄ → C 2/c phase transitions.

S-7
Figure S4: ELF for (a) P 21 /c at 0 GPa (b) P 1̄ at 10 GPa, and (c) Pnma structure at 20
GPa with isosurface value 0.75.

10
a b c

9
Lattice parameter (Å)

4
0 4 8 12 16 20 24 28 32 36 40 44 48 52
Pressure (GPa)

Figure S5: Change of lattice parameters of Pnma structure (CSP-50GPa) with pressures.
The response of b-axis is more significant with pressure than other two axes.

S-8
References

(1) Fullarton, M. L.; Qin, M. J.; Robinson, M.; Marks, N. A.; King, D. J. M.; Kuo, E. Y.;
Lumpkin, G. R.; Middleburgh, S. C. Structure, properties and formation of PuCrO3 and
PuAlO3 of relevance to doped nuclear fuels. J. Mater. Chem. A 2013, 1, 14633–14640.

(2) Hinuma, Y.; Kamachi, T.; Hamamoto, N.; Takao, M.; Toyao, T.; Shimizu, K.-i. Surface
Oxygen Vacancy Formation Energy Calculations in 34 Orientations of -Ga2 O3 and -
Al2 O3 . J. Phys. Chem. C 2020, 124, 10509–10522.

(3) Guillot, B.; Sator, N. A computer simulation study of natural silicate melts. Part I: Low
pressure properties. Geochim. Cosmochim. Acta 2007, 71, 1249–1265.

(4) Wang, M.; Krishnan, N. A.; Wang, B.; Smedskjaer, M. M.; Mauro, J. C.; Bauchy, M. A
new transferable interatomic potential for molecular dynamics simulations of borosilicate
glasses. J. Non-Cryst. Solids 2018, 498, 294–304.

(5) Taylor, M.; Brown, G. E. High-temperature structural study of the P21/a → A2/a phase
transition in synthetic titanite, CaTiSiO5 . Am. Mineral. 1976, 61, 435–447.

(6) Rath, S. Comparative pressure-dependent structural studies of titanite (CaTiOSiO4 )


and malayaite (CaSnOSiO4 ). Ph.D. thesis, Swiss Federal Institute of Technology Zurich,
2002.

(7) Angel, R. J.; Kunz, M.; Miletich, R.; Woodland, A. B.; Koch, M.; Xirouchakis, D. High-
pressure phase transition in CaTiOSiO4 titanite. Phase Transitions 1999, 68, 533–543.

S-9

You might also like