research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206

Phase transitions and (p–T–X) behaviour of centrosymmetric perovskites: modelling with transformed crystallographic data

crossmark logo

aWerkstofftechnik Glas and Keramik, Hochschule Koblenz, Rheinstrasse 56, 56203 Hoehr-Grenzhausen, Germany
*Correspondence e-mail: thomas@hs-koblenz.de

Edited by R. Černý, University of Geneva, Switzerland (Received 5 August 2021; accepted 30 November 2021; online 20 January 2022)

Dedicated to the memory of Christopher Jason Thomas (30.12.1988–21.10.2021).

A reversible transformation of the unit-cell parameters and atomic coordinates of centrosymmetric perovskites ABX3 into a Cartesian space is defined. Analytical expressions for the three vectors for the pseudocubic cell and three vectors for a BX6 octahedron are derived for space groups Pbmn, Cmcm, Ibmm, P4/mbm, P4/nmc, I4/mcm and R3c. The following structural parameters may be derived from these vectors: up to six pseudocubic parameters defining octahedral geometry; length- and angle-based octahedral distortion parameters λ and σ; inclination angles of tilted octahedra, θ1, θ2 and θ3; angles of tilt of octahedra; AX12:BX6 polyhedral volume ratio, VA/VB; parameters ηA and ηB defining the relative contraction of inner AX8 polyhedra and expansion of BX6 octahedra due to octahedral tilting. The application of these parameters is demonstrated by reference to published crystal structures. The variation of ηA and ηB with temperature in the compositional series SrxBa1–xSnO3 and SrxBa1–xHfO3, as well as the temperature series of BaPbO3 and CaTiO3, is related to the sequence of phases PbmnIbmmPm3m. Stabilization of the Cmcm phase is likewise interpreted in terms of these two parameters for NaTaO3 and NaNbO3. The pressure evolution of the structures of MgSiO3, YAlO3, (La1–xNdx)GaO3 (0 ≤ x ≤ 1) and YAl0.25Cr0.75O3 is modelled with the appropriate structural parameters, thereby also addressing the characteristics of the PbmnR3c transition. Simulation of MgSiO3 up to 125 GPa and of YAlO3 up to 52 GPa in space group Pbnm is carried out by using the Birch–Murnaghan equation of state. In both cases, full sets of oxygen coordinates assuming regular octahedra are generated. Octahedral distortion is also modelled in the latter system and predicted to have a key influence on structural evolution and the sequence of phase transitions. The core modelling procedures are made available as a Microsoft Excel file.

1. Introduction: the modelling of octahedral tilting in perovskites

Synthetic perovskite-related compounds continue to attract the attention of many scientists and technologists, irrespective of whether they are working, for example, on the development of lead-free piezoelectric ceramics (Shrout & Zhang, 2007[Shrout, T. R. & Zhang, S. J. (2007). J. Electroceram. 19, 113-126.]) or on organolead halide ABX3 nanocrystals for photochemical cells (Jena et al., 2019[Jena, A. K., Kulkarni, A. & Miyasaka, T. (2019). Chem. Rev. 119, 3036-3103.]). In general, the current availability of high-quality structural data evokes the need for all the nuances of structural change under varying conditions of pressure, temperature and composition] (p–T–X) to be precisely modelled. This expectation is particularly poignant in the case of perovskite-related compounds, ABX3, since, according to Mitchell (2002a[Mitchell, R. H. (2002a). Perovskites: Modern and Ancient, Preface. Thunder Bay, Ontario: Almaz Press.]), all elements, apart from the noble gases, can be found in a variant of these. In the pioneering work of Megaw (1973[Megaw, H. D. (1973). Crystal Structures - A Working Approach, pp. 285-302. London: Saunders.]), their structures were regarded as comprising three essential features: (a) tilting of the anion octahedra; (b) displacements of the cations; (c) octahedral distortions. Definitive work by Glazer (1972[Glazer, A. M. (1972). Acta Cryst. B28, 3384-3392.]) led to a system of classification based on the sense of tilt of the three tetrad axes of corner-linked regular octahedra about three perpendicular, or nearly perpendicular crystal axes.

A regular octahedron is characterized by three pairs of opposite vertices that are linked by stalks of equal length passing through its mid-point at right angles to one another. Two lines of regular octahedra are shown in Figs. 1[link](a) and 1[link](b), as they arise in space group Pbnm.

[Figure 1]
Figure 1
Tilting of regular octahedra about the xPC and yPC axes, as in space group Pbnm. Pseudocubic axes xPC, yPC and zPC are directed parallel to orthorhombic vectors [\left [{1\bar 10} \right]], [110 ] and [001], respectively. (a) Tilting around the yPC axis (in clinographic projection) with θy the angles of inclination of stalks of length s to the axis; (b) Tilting around the zPC axis (head-on view showing mirror planes); (c) yPC-axis tilting [as in (a)] viewed along the yPC axis towards the origin; (d) zPC-axis tilting [as in (b)] viewed along the zPC axis towards the origin.

Zigzag chains of adjacent octahedral stalks of uniform length s are formed, these lying in planes PQRS and TUVW. When viewed along the yPC axis-of-tilting, the red and blue octahedra are seen to be rotated in opposite senses [Fig. 1[link](c)], leading to Glazer notation [{b^ - }]. By comparison, the red and green octahedra are rotated in the same sense [Fig. 1[link](d)]. This leads to the notation c+ , since the tilt angle is different. As tilting around the xPC and yPC axes is equivalent, the three-dimensional tilt system in Pbnm is denoted by [{a^ - }{a^ - }{c^ + }] (Glazer, 1972[Glazer, A. M. (1972). Acta Cryst. B28, 3384-3392.]).

In this now well established a#b#c# nomenclature, superscripts # can be +, − or 0, denoting the angles of tilt of neighbouring octahedra along the three pseudocubic axes as in-phase (+), anti-phase (−) or zero. Importantly, Glazer (1972[Glazer, A. M. (1972). Acta Cryst. B28, 3384-3392.]) also showed that the 23 tilt systems for regular octahedra correlated with 15 alternative space groups. As well as aiding the correct interpretation of perovskite diffraction patterns (Glazer, 1975[Glazer, A. M. (1975). Acta Cryst. A31, 756-762.]), these correlations stimulated group-theoretical analysis of perovskites, their being confirmed significantly later by Howard & Stokes (1998[Howard, C. J. & Stokes, H. T. (1998). Acta Cryst. B54, 782-789.], 2002[Howard, C. J. & Stokes, H. T. (2002). Acta Cryst. B58, 565-565.]). The latter work led to minor modifications to the space groups originally assigned by Glazer and established a top-down, hierarchical system of group–subgroup relationships starting from the cubic [Pm \overline{3}m] aristotype for perovskite structures with tilted, regular octahedra (Fig. 2[link]).

[Figure 2]
Figure 2
Group–subgroup relationships among the 15 perovskite space groups arising from the tilting of regular octahedra. A dashed line signifies that the corresponding phase transition must be first order (Howard & Stokes, 2002[Howard, C. J. & Stokes, H. T. (2002). Acta Cryst. B58, 565-565.]). Reproduced with permission of the International Union of Crystallography.

Two independent contributions concerning octahedral tilting in perovskites were made by Woodward (1997a[Woodward, P. M. (1997a). Acta Cryst. B53, 32-43.],b[Woodward, P. M. (1997b). Acta Cryst. B53, 44-66.]) and Thomas (1996[Thomas, N. W. (1996). Acta Cryst. B52, 16-31.]). Woodward (1997a[Woodward, P. M. (1997a). Acta Cryst. B53, 32-43.]) concluded that perfectly regular octahedra could not be linked together in some tilt systems. The contentious space groups were subsequently narrowed down to Cmcm (tilt system [{a^0}{b^ + }{c^ - }]) and P42/nmc ([{a^ + }{a^ + }{c^ - }]) by Howard & Stokes (1998[Howard, C. J. & Stokes, H. T. (1998). Acta Cryst. B54, 782-789.]), with only the latter requiring irregular octahedra. The work of Thomas (1996[Thomas, N. W. (1996). Acta Cryst. B52, 16-31.]), by comparison, was primarily concerned with the dependence of AX12:BX6 polyhedral volume ratio, VA/VB, on octahedral tilting in orthorhombic and tetragonal perovskites. Inclination angles θy and θz in Figs. 1[link](a) and 1[link](b), and by extension θx, were used to relate octahedral stalk lengths s and tilting to the lengths of pseudocubic cell axes 1, 2 and 3 and therefore cell volume [equation (1)[link]].

[{V_A}/{V_B} \approx 6\cos^{2}\big[{({{\theta _1} + {\theta _2}} )/2}\big]\cos {\theta _3} - 1 \eqno(1)]

This form results from the coupling of inclination angles θ1 and θ2 when pseudocubic axes 1 and 2 are oriented at approximately 45° to the crystal axes. In untilted structures (with θ1 = θ2 = θ3 = 0), VA/VB is exactly equal to five (Thomas, 1989[Thomas, N. W. (1989). Acta Cryst. B45, 337-344.]).

Tamazyan & van Smaalen (2007[Tamazyan, R. & van Smaalen, S. (2007). Acta Cryst. B63, 190-200.]) subsequently argued that inclination angles θ1, θ2 and θ3 defined by Thomas (1996[Thomas, N. W. (1996). Acta Cryst. B52, 16-31.]) were unnecessarily influenced by octahedral distortion. They therefore proposed an alternative method of calculating tilt angles, although this was at the expense of losing the simple link to polyhedral volume ratio expressed by equation (1)[link]. Later work by Wang & Angel (2011[Wang, D. & Angel, R. J. (2011). Acta Cryst. B67, 302-314.]) restored the link between octahedral tilting and ratio VA/VB, employing a group-theoretical, rather than a geometrical approach to separate octahedral tilting and distortion. These authors expressed the ratio VA/VB as a function of the amplitudes of the normal modes in a cubic perovskite, rather than by direct calculation from anion coordinates in experimentally determined crystal structures.

The conflict between geometrical, i.e. crystal-chemical and group-theoretical methods in describing perovskite structures is somewhat contrived, since both have legitimate fields of application and ultimately have similar aims. This was made clear in seminal work by Knight (2009[Knight, K. S. (2009). Can. Mineral. 47, 381-400.]), who utilized group-theoretical methods to develop a general parameterization of centrosymmetric perovskites based on symmetry-adapted basis vectors of the [Pm\bar 3m] phase. He also pointed to difficulties with a method earlier proposed by this author for a general crystal-chemical parameterization of centrosymmetric perovskites (Thomas, 1998[Thomas, N. W. (1998). Acta Cryst. B54, 585-599.]). Two specific problems were identified, (a) that the method was `geometrically complex'; and (b) that it relied `totally on an empirical analysis of known crystal structures'. He also made the important point, (c), that the A site has a coordination number less than 12 due to a geometrically complex coordination polyhedron (Knight, 2009[Knight, K. S. (2009). Can. Mineral. 47, 381-400.]). In the current work, point (a) is addressed by focusing on the distortion of centrosymmetric octahedra as independent geometrical forms. Point (b) is addressed by deriving analytical expressions dependent on space group for the three stalk vectors defining octahedral geometry. Point (c) is addressed by developing a simple parameterization based on AX8 sub-polyhedra that correlates with the tilt classification system of Glazer (1972[Glazer, A. M. (1972). Acta Cryst. B28, 3384-3392.]). Details of these methodological improvements are given in the following section.

2. Elements of the revised crystal-chemical method

2.1. Pseudocubic representations of octahedra

In order that an object can be described as an octahedron, it requires three recognisable pairs of opposite vertices that may be linked to each another by vectors. If the ends of the stalks of a regular octahedron [Fig. 3[link](a)] are randomly displaced by limited amounts [Fig. 3[link](b)], the resulting form is still recognisable as an octahedron [Fig. 3[link](c)]. The six independent vertices lead to six sets of [x,y,z ] triplets in Cartesian space, i.e. 18 parameters. If only the form of the octahedron is relevant, and not its absolute position or orientation, these 18 parameters are reduced to 12 by subtracting six parameters: three to fix one vertex in space at [0,0,0] and three to define the orientation of the octahedron. In Fig. 3[link](c), one vertex has been fixed at [0,0,0] and the octahedron so rotated that the opposite vertex is fixed at [0 ,0, z1]. A further rotation of the octahedron about the z axis has been carried out to fix a vertex at [0, y2, z2]. The total of six coordinate-components equal to zero here signifies a reduction in independent parameters from 18 to 12, i.e. (z1; y2z2; xnynzn; n = 3,5). If, however, the translations at opposite ends of the three stalks are equal and opposite [Fig. 3[link](d)], a centrosymmetric octahedron results [Fig. 3[link](e)], in which the three stalks bis­ect one other at the centre of symmetry. Only one end of a stalk needs to be fixed in space in order to define the position of the opposite end, so that the number of independent parameters is halved to six [Fig. 3[link](e)]: (z1; y2z2; x3y3z3). This situation applies to all perovskites with B ions located at centres of symmetry.

[Figure 3]
Figure 3
The number of independent parameters of a generalized, isolated centrosymmetric octahedron is equal to six. Views (a) to (e) discussed in §2.1[link].

The six independent parameters of a centrosymmetric octahedron may be assigned to three stalk lengths a1, a2, a3 and three angles of intersection of the stalks, θ12, θ23 and θ31. In order to visualize the extent of its distortion, a parallelepiped enclosing the octahedron may be constructed by displacing the three stalk vectors to a common origin (Fig. 4[link]).

[Figure 4]
Figure 4
Visualization of the distortion of octahedra by enclosure in a pseudocube: (a), (b): centrosymmetric; (c), (d) non-centrosymmetric.

As the three independent edge lengths of the parallelipiped are approximately equal, it may also be termed a pseudocube. Such a pseudocube is defined by the same six parameters as the octahedron, i.e. a1, [ {a}_{2}, { a}_{3}, {\theta }_{12},{\theta }_{23}] and [{\theta _{31}}]. It follows that regular octahedra would lead to pseudocubic representations of octahedra (PCRO) of cubic form.

In the case of the non-centrosymmetric octahedron of Fig. 3[link](c), the six octahedral vertices do not touch their parallelogram faces at the meeting points of the four quadrants [Fig. 4[link](d)]. The 12 independent parameters may be accommodated by noting the 2D polar coordinates [\left [{r,\phi } \right]] of the three emergent octahedral vertices, giving rise to additional parameters r1, [ {r}_{2}, { r}_{3}, {\phi }_{1},{\phi }_{2}] and [{\phi _3}]. Thus the PCRO construction is of general validity for visualizing distorted octahedra. Its parameterization may prove to be useful for characterizing non-centrosymmetric and polar perovskites in future work. The transformation between octahedron and PCRO is also reversible.

An advantage of the PCRO visualization method is that the aggregate distortion parameters representing normal and shear distortion developed for pseudocubic representations of tetrahedra (PCRT), [ \lambda] and [\sigma], (Reifenberg & Thomas, 2018[Reifenberg, M. & Thomas, N. W. (2018). Acta Cryst. B74, 165-181.]; Fricke & Thomas, 2021[Frontline Systems Inc. (2021). https://www.solver.com/.]) may be taken over without modification [equations (2)[link], (3)[link]].

[{{\lambda}} = {{| {{a_1}\! -\! {L_0}}| + | {{a_2}\! -\! {L_0}} | + | {{a_3}\! - \!{L_0}} |} \over {3{L_0}}}\,\,\, {\rm with}\,\,\, {L_0} = {{({a_1}\! + \!{a_2} \!+\! {a_3})} \over 3} \eqno(2)]

[\sigma {(^\circ }) = {{\left| {{\theta _{12}}\left(^\circ \right) - 90^\circ } \right| + \left| {{\theta _{23}}\left(^\circ \right) - 90^\circ } \right| + \left| {{\theta _{31}}\left(^\circ \right) - 90^\circ } \right|} \over 3} \eqno(3)]

In this connection, the following values are obtained for the centrosymmetric and non-centrosymmetric octahedra of Figs. 3[link](e) and 3[link](c), respectively: [λ,σ] = [0.0367, 11.90°]; [λ,σ] = [0.0370, 7.10°]. It follows that the centrosymmetric octahedron displays a higher degree of angular distortion here. Furthermore, if the angular distortion parameters are expressed in radians instead of degrees, the relative degree of shear versus normal distortion can be quantified. Since 11.90° correspond to 0.21 and 7.10° to 0.12 radian, it follows that the shear distortion is greater than the normal distortion in both cases.

2.2. Parameterization of PCRO in terms of three vectors

The octahedra in perovskites are not isolated but form a three-dimensional corner-sharing network. Their crystal structural parameters, i.e. space group, unit-cell parameters and atomic coordinates, deliver full information on octahedral distortion, tilting and connectivity. Importantly, space group symmetry ensures connectivity. Once the distortion and tilting of a single octahedron in a structure with only one symmetry-independent octahedron has been defined, the distortions and tilting of all the other octahedra in the unit cell follow. To convey full information on the tilting of this single octahedron in a structure, its PCRO may be described by the three Cartesian stalk vectors, a1, a2 and a3 along its pseudocube edges. Generalized, analytical expressions for these vectors in different space groups may be derived inductively from known crystal structures. This process is described in §2.2.1[link] for space group Pbnm with B ions located at 4b special positions. The three vectors defining the axes of the pseudocubic unit cell are likewise defined in this Cartesian space, so that inclination angles of the stalks to the three pseudocubic axes, θ1, θ2 and θ3 may straightforwardly be calculated [see Fig. 1[link], equation (1)[link] and §2.3[link]]. In addition, the method also allows calculation of tilt angles ϕa, ϕb and ϕc of the octahedra around the three pseudocubic axes, even when the octahedra are distorted.

2.2.1. Space group Pbnm with B ions at 4b positions

The octahedral cage coordinating the B ion at 0½0 in the unit cell of orthorhombic space group Pbnm may be taken, this being the cab setting of space group 62 with standard symbol Pnma. 0½0 is one of the 4b special positions, with X1 ions in 4c special positions and X2 ions in 8d general positions (see Table S1 in the supporting information).1

Derivation of the analytical form of stalk, or equivalently PCRO vectors a1, a2 and a3 starts by taking an example structure, such as CaTiO3 at 296 K [Yashima & Ali (2009[Yashima, M. & Ali, R. (2009). Solid State Ionics, 180, 120-126.]); ICSD code 162908]: a = 5.3709 Å, b = 5.4280 Å, c = 7.6268 Å; x(O1) = 0.0708, y(O1) = 0.4830, x(O2) = 0.7113, y(O2) = 0.2891, z(O2) = 0.0375; z(O1) has the fixed value of ¼.

The essential step is to convert from numerical fractional coordinates of the titanium ion and its six oxygen neighbours (Table 1[link], columns 2–4) to analytical fractional coordinates (Table 1[link], columns 8–10). One way to generate the numerical fractional coordinates would be to use a simple computer program to generate the Cartesian coordinates of the titanium and oxygen ions. (Table 1[link], columns 5–7). The numerical fractional coordinates (Table 1[link], columns 2–4) would then follow by multiplying these by the inverse orthogonalization matrix, in this trivial case

[\left({\matrix{ {1/a} & 0 & 0 \cr 0 & {1/b} & 0 \cr 0 & 0 & {1/c} \cr } } \right).]

Letters a to f of the oxygen ions correspond to the atom labels in Fig. 1[link](a). They are based on the principle that the octahedral stalk from a to b is oriented closest to the xPC axis in the positive direction, with the stalks from c to d and from e to f oriented closest to the positive yPC and zPC directions, respectively.

Table 1
Calculation of analytical fractional coordinates of the six oxygen anions coordinating the titanium ion in CaTiO3 located at 0,½,0

Atom Numerical fractional coordinates Cartesian coordinates Analytical fractional coordinates
x y z xC yC zC x y z
Ti (4b) 0 0.5 0 0.0000 2.7140 0.0000 0 ½ 0
O a (8f) −0.2113 0.7891 0.0375 −1.1349 4.2832 0.2860 x(O2) + ½ y(O2) + ½ z(O2)
O b (8f) 0.2113 0.2109 −0.0375 1.1349 1.1448 −0.2860 x(O2) − ½ y(O2) + ½ z(O2)
O c (8f) −0.2887 0.2891 0.0375 −1.5506 1.5692 0.2860 x(O2) − 1 y(O2) z(O2)
O d (8f) 0.2887 0.7109 −0.0375 1.5506 3.8588 −0.2860 x(O2) + 1 y(O2) + 1 z(O2)
O e (4c) −0.0708 0.5170 −0.25 −0.3803 2.8063 −1.9067 x(O1) y(O1) + 1 −¼
O f (4c) 0.0708 0.4830 0.25 0.3803 2.6217 1.9067 x(O1) y(O1) ¼

The analytical expressions in columns 8 to 10 of Table 1[link] are obtained by inspection: the values in columns 2 to 4 are compared with the starting values of x(O1), y(O1), x(O2), y(O2) and z(O2). It is important to note that the results depend on the convention used for expressing these parameters. For example, the alternative of x(O1) = 0.4292 and y(O1) = −0.0170, although symmetrically equivalent, would lead instead to the results x(O e) = x(O1) − ½ and y(O e) = −y(O1) + ½.

The three vectors for the PCRO are now formed in Table 2[link] by taking the differences in the analytical fractional coordinates for atom pairs (a,b), (c,d) and (e,f) given in Table 1[link]. Cartesian vectors are formed by multiplying these differences by the orthogonalization matrix

[\left({\matrix{ a & 0 & 0 \cr 0 & b & 0 \cr 0 & 0 & c \cr } } \right).]

The Cartesian axes are parallel to the axes of the ortho­rhombic unit cell in this case.

Table 2
The three vectors defining the PCRO with mid-point 0,½,0 in space group Pbnm

    Cartesian components Nearest pseudocubic axes
Stalk PCRO vector X Y Z Pseudocubic Orthorhombic
O b ← O a a1 a(2x(O2) − 1) − 2by(O2) − 2cz(O2) xPC []0{\bar 1}0]]
O d ← O c a2 −2a(x(O2) − 1) b(2y(O2) − 1) −2cz(O2) yPC [110]
O f ← O e a3 2ax(O1) b(2y(O1) − 1) c/2 zPC [001]

By adopting this analytical representation, vectors a1, a2 and a3 are no longer tied to the example structure of CaTiO3: they apply to all perovskites in space group Pbnm, provided that the B ions occupy 4b sites and that the correct convention in choosing the values parameters x(O1), y(O1), x(O2), y(O2) and z(O2) has been followed. All eight crystallographic parameters are involved in defining the geometry of the PCRO in space group Pbnm: a, b, c, x(O1), y(O1), x(O2), y(O2), z(O2). For regular octahedra, six of these eight degrees of freedom (d.o.f.) would be used up in forming octahedra of a particular volume, with five defining the regular form and the sixth the volume. The remaining two d.o.f. would be used to define the unit-cell parameters and octahedral tilting. If the octahedra were only approximately regular, as is generally the case, more d.o.f. would be available for optimizing the octahedral tilting and unit-cell volume, in response to different A and B ion radii or to changing temperature or hydro­static pressure. Thus the idea of structural compromise in forming connected octahedral networks in perovskites can be modelled in response to varying (p–T–X) conditions.

2.2.2. Other space groups

The other space groups analysed in this work are those that commonly arise in experimentally determined crystal structures. They span the following space groups: Pbmn (B ions in 4a sites), Cmcm, Ibmm, P4/mbm, P42/nmc, I4/mcm, [R{\bar 3}c], [Pm{\bar 3}m]. Tables similar to Table 2[link] are generated for them in §S2 of the supporting information.

2.3. Implementation of the crystallographic to structural transformation in the Microsoft Excel Solver environment

Structural analysis requires a one-way transformation from crystallographic to structural parameters. For example, crystallographic parameters a, b, c, xi, yi, zi are transformed to structural parameters such as PCRO parameters and tilt angles. By comparison, structural prediction requires a reversible transformation between the two parameter sets. The modeller will seek to establish systematic variations in the structural parameters with (p–T–X). If successful, interpolations and extrapolations to other (p–T–X) values can be made before reverse-transforming to crystallographic parameters. This technique has been demonstrated for olivines (Thomas, 2017[Thomas, N. W. (2017). Acta Cryst. B73, 74-86.]), coesite (Reifenberg & Thomas, 2018[Reifenberg, M. & Thomas, N. W. (2018). Acta Cryst. B74, 165-181.]) and quartz (Fricke & Thomas, 2021[Frontline Systems Inc. (2021). https://www.solver.com/.]).

The requirements of ease and flexibility of use together with a reversible transformation suggest that the Microsoft Excel Solver environment is appropriate for applying the method. In this connection, an Excel datafile already programmed is provided in the supporting information. This consists of eight worksheets for the different space groups. A screenshot of the worksheet for Pbnm with B ions in 4b positions is given in Fig. 5[link].

[Figure 5]
Figure 5
Screenshot of the Pbnm worksheet, which contains the reference data for CaTiO3 at 296 K (Yashima & Ali, 2009[Yashima, M. & Ali, R. (2009). Solid State Ionics, 180, 120-126.]). No refinement has taken place, since the entries in the `Refined data' and `Reference data' boxes are identical.

The user enters the crystallographic data in the `Refined data' box in blocks K3:K5 (light-green background), N4:N8 and N10:N11 (mustard background), whereupon all the structural parameters are automatically calculated by Excel. The computational core of the spreadsheet is in blocks R9:T11 (deep-blue background) and U9:W11 (dark-green background). The formulas in these cells correspond to the entries in Table 2[link] and the pseudocubic axes in Cartesian coordinates, respectively. Values of all the structural parameters relating to octahedral distortion and tilting (light-blue background) can be seen as ways of describing the numerical values in these two blocks in a structurally meaningful way.

The purpose of the `Reference data' box is threefold. First, it acts as a repository for the source, reference data (cells with orange background). Secondly, the dependent structural parameters can be calculated for this reference data by clicking on the button in cell C15.2 Thirdly, both reference data and dependent structural parameters can be used as constraints in structural refinements, which apply to the cells in the `Refined data' box.

The different types of structural parameters, all with light-blue background, are summarized as follows.

a) PCRO parameters a1, a2, a3, θ23, θ31, θ12 are calculated as the lengths and intersectional angles of the vectors a1, a2, a3 in block R9:T11. For example, [{\theta }_{23} = \arccos\left({{\bf a}}_{2}\cdot{{\bf a}}_{3}/{a}_{2}{a}_{3}\right)]. Aggregate distortion parameters λ and σ [equations (2)[link] and (3)[link]] are calculated in block P13:P16, the latter quoted in both degrees and radians.

b) Pseudocubic angle γPC is calculated as the angle between pseudocubic axes aPC and bPC in cell P11: [{\gamma }_{\rm PC} = \arccos({{\bf a}}_{\rm PC}\cdot{{\bf b}}_{\rm PC}/{a}_{\rm PC}{b}_{\rm PC})].

c) Inclination angles θ1, θ2, θ3 of the octahedral stalks are quoted in block X9:X11. These describe the relationship of the two sets of vectors in blocks R9:T11 and U9:W11 to one another. With respect to vector a1 in block R9:T9 and its nearest pseudocubic axis vector aPC in block U9:W9, the following applies:

[{{\bf a}_1} = {\pmatrix{ {{a_{1X}}} \cr {{a_{1Y}}} \cr {{a_{1Z}}} \cr } };]

[{\bf a}_{\rm PC} = \pmatrix{ a_{{\rm PC},X} \cr a_{{\rm PC},Y} \cr a_{{\rm PC},Z}}.]

The subscripts X,Y,Z here denote the Cartesian X,Y and Z components, respectively. It follows that

[\eqalignno{{\theta }_{1} = &\arccos({{\bf a}}_{1}\cdot {{\bf a}}_{\rm PC}/{a}_{1}{a}_{\rm PC})\cr = &\arccos\big(\big[{a}_{1X}{a}_{{\rm PC},X}+{a}_{1Y}{a}_{{\rm PC},Y}+{a}_{1Z}{a}_{{\rm PC},Z}\big]/{a}_{1}{a}_{\rm PC}\big) &(4)} ]

Substitution of

[{a_{\rm PC}} = \pmatrix{ \,\,\,\,\,2.2697 \cr - 3.1385 \cr -0.5720}]

and

[{a_{\rm PC}} = \pmatrix{ \,\,\,\,5.3709 \cr - 5.4280 \cr \,\,\,\,0.0000}]

leads to θ1 = 12.16°. Angles θ2 and θ3 are derived from the corresponding pairs (a2,bPC) as well as (a3,cPC), respectively.

d) The method of calculation of angles of tilt ϕa and ϕc in the [{a^ - }{a^ - }{c^ + }] tilt system of space group Pbnm is shown in Fig. 6[link]. These angles constitute an alternative to the inclination angles for relating the two sets of vectors in blocks R9:T11 and U9:W11 to each other. Since tilt angles are to be calculated for distorted octahedra, the method evaluates the projections of oxygen atoms O a, O b, O c and O d in the planes perpendicular to pseudocubic axes aPC, bPC and cPC. These four atoms also define PCRO vectors a1 and a2: vector a3 does not influence the calculated tilt angles.

[Figure 6]
Figure 6
Tilted octahedra and basal parallelogram planes, e.g. PQRS, of AX12 polyhedra (in yellow) viewed along the orthorhombic z axis in space group Pbnm. The + and – signs in black denote z heights of the octahedral vertices relative to the z = 0 plane of the diagram. These are due to [{a^ - }] tilting. Green dashed lines show the directions of pseudocubic axes xPC and yPC, the reference directions for the [{a^ - }] tilting. Red dotted lines represent the perpendicular planes to the green dashed lines, which are projected as lines in the xy plane. Red circles lie above the red dotted lines at the same z height as the octahedral vertices O a to O d. They are denoted by O a′ to O d′ and result from the four projections (O a → O a′), (O b → O b′), (O c → O c′) and (O d → O d′). c+ tilting causes the octahedra to be rotated about the orthorhombic z axis (or equivalently, the zPC axis). The relevant tilt angles, ϕc,1 and ϕc,2, are shaded orange. (Vectors p and q are defined in §2.6[link].)

A tilt of type [{a^ - }] around aPC is the angle between vector (O c′-O d′) and its projection in the xy plane. Similarly, an [{a^ - }] tilt around bPC is the angle between vector (O a′-O b′) and its projection in the xy plane. Calculated values are shown in cells Y9 and Y10 in Fig. 5[link], with the mean value in cell Y11 (8.53°, 8.51° and 8.52°, respectively). The observed difference of 0.02° in calculated values results from an interplay of the deviation of γPC from 90° and the octahedral distortions.

The in-phase c+ tilting causes the projections of the a1 and a2 octahedral vectors in the xy plane to be rotated away from pseudocubic axes xPC and yPC (Fig. 6[link]). The two resulting tilt angles are given in cells Y12 and Y13 (8.82° and 8.87°), with the mean value in Y14 (8.84°).

The algorithms for calculating these tilt angles are described in §S3 of the supporting information, where reference is also made to their implementation in the Excel file in the supporting information.

e) Unit-cell volume VUC, polyhedral volumes VB, VA and volume ratio VA/VB are calculated in blocks P3:P4 and P6:P7. VUC = [{\bf a}\cdot {\bf b}\times {\bf c}] and

[{V_B} = {1 \over 6}\left| \matrix{ a_{1x} & a_{1y} & a_{1z} \cr a_{2x} & a_{2y} & a_{2z} \cr a_{3x} & a_{3y} & a_{3z}} \right|.]

The latter calculation method is facilitated by the centrosymmetry of the octahedron. Volume ratio VA/VB follows from these parameters as [(VUC/Z) − VB]/VB and VA as VB(VA/VB ).

f) Parameters ηA and ηB are likewise extracted from the core computational data (blocks R9:T11 and U9:W11) and defined in §2.6[link].

g) Parameters L and ϕ (block N18:19 with brown background) relate to the A ion positions and are independent of the anionic network. They are defined in §2.7[link].

2.4. Example Solver refinements

Initial insight into use of the Solver is gained here by addressing an issue relevant to the historical development of perovskite structural chemistry: the ability of different space groups to accommodate regular octahedra. A comparison is also made between tilt angles calculated as in point d) of §2.3[link] and the values yielded by commonly used approximations.

The five degrees of freedom required to define the regular octahedral form (see §2.2.1[link]) result in the constraints a) a2 = a1; b) a3 = a2; c) θ23 = 90°; d) θ31 = θ23; and e) θ12 = θ31. Since there are eight d.o.f. in total, two further constraints may be applied in the refinement. For example, it may be stipulated that the unit cell and the octahedral volumes remain unchanged: f) VUC = VUC,reference and g) VB = VB,reference. The coding of these constraints in the Solver in shown in Fig. 7[link](a), whereby constraints f) and g) also refer to cells $H$3 and $H$4 in Fig. 5[link]. The end-point of this tightly defined refinement is shown in Fig. 7[link](b): a cubic PCRO of side length 3.9014 Å is generated, corresponding to a regular octahedron with perpendicular stalks. The values of the other structural parameters described in §2.3[link] appear in the cells with a light-blue background.

[Figure 7]
Figure 7
(a) Solver settings for generating regular octahedra in space group Pbnm with unchanged unit cell and octahedral volumes. (b) Results of the refinement.

The crystal structures obtained by applying the same refinement conditions to reference structures in all the space groups are given in Table 3[link]. The Solver constraints for the respective space groups are pre-programmed in the Excel file in the supporting information.

Table 3
Summary of Solver refinements by space group corresponding to structures with regular octahedra

The octahedral and unit-cell volumes are identical to those in the reference structures.

Space group Pbnm (B: 4b) Pbnm (B: 4a) Cmcm Ibmm P4/mbm P42/nmc I4/mcm [R{\bar 3}c]
Reference system CaTiO3 LaCr0.7Ni0.3O3 NaNbO3 at 848 K BaPbO3 at 300 K NaNbO3 at 888 K CaMnO3 CaTiO3 at 1523 K La(Cr0.2Ni0.8)O3
Reference ICSD 162908 173471 192404 154038 280100 670342 162919 173475
a1(refined) (Å) 3.9014 3.9535 3.9495 4.3480 3.9500 3.8074 3.9025 3.8952
a (Å) 5.33816 5.46101 7.85139 6.08315 5.56480 7.52285 5.47004 5.42241
b (Å) 5.44997 5.57887 7.85672 6.14899        
c (Å) 7.64264 7.73997 7.89355 8.60287 3.94998 7.43089 7.80495 13.49346
x(O1) 0.07274 0.57385 0.27596 0.05216 0 0.03909 0 0.55170
y(O1) 0.48876 0.49521 0 0 0 x(O1) 0 0
z(O1) ¼ ¼ 0 ¼ ½ ¼ ¼ ¼
x(O2) 0.70974 0.23308 0 ¼ 0.22810 ¼ 0.21650  
y(O2) 0.28863 0.26621 0.22407 ¼ 0.72810 0.00306 0.71650  
z(O2) 0.03637 0.03692 0.00921 −0.02608 0 0.96043 0  
x(O3)     0.25921     ¼    
y(O3)     0.25096     y(O2)    
z(O3)     ¼     z(O2) + [{1 \over 2}]    
Method §2.3[link] §2.3[link] a) §2.3[link] b) §2.3[link] d)   §2.3[link] §2.3[link] c) §2.3[link]
ϕa (°) 8.52 8.47 11.91     6.37 8.72   7.13     5.90
ϕa,refined (°) 8.29 8.33 11.80     5.92 8.39   8.89     5.90
ϕb (°)       2.11 4.15       6.02      
ϕb,refined (°)       2.12 2.11       8.89      
ϕc (°) 8.84 4.90 4.90 5.38 5.38     5.01 9.34 7.63 7.63  
ϕc,refined (°) 8.96 3.79 3.79 5.92 5.92     5.01 0.70 7.63 7.63  
θ3 (°)   10.97   7.87      
θ3,refined (°)   11.80   8.39      

For space group P42/nmc, the obtaining of regular octahedra in tilt system a+a+c is at variance with the conclusion of Howard & Stokes (1998[Howard, C. J. & Stokes, H. T. (1998). Acta Cryst. B54, 782-789.]). A resolution of this discrepancy is thought likely to depend on the observed interdependence of the coordinates of atoms O2 and O3 in the refined structure (Table 3[link]), which implies a refinement into higher symmetry or pseudo-symmetry than P42/nmc.

The values of tilt angles below the dividing line in Table 3[link] serve to compare the values calculated by the method given in §2.3[link] and approximations for tilt angles used by other authors. Following an initial analytical treatment of perovskite tilt angles by Megaw (1973[Megaw, H. D. (1973). Crystal Structures - A Working Approach, pp. 285-302. London: Saunders.]), a more developed treatment according to space group was given by Kennedy, Prodjosantoso et al. (1999[Kennedy, B. J., Prodjosantoso, A. K. & Howard, C. J. (1999). J. Phys. Condens. Matter, 11, 6319-6327.]) as follows. The four parameterizations a) to d) here correspond to headers a) to d) in the row `Method' in Table 3[link].

a) Pbnm (B ions in 4a): O2:[\big({{1 \over 4} - u,{1 \over 4} + v,w}\big)]. [\phi _{\rm a} = \arctan\big(4\cdot {2}^{{{1}\over{2}}}\cdot w\big)]; [\phi _{\rm c} = \arctan \big({2[{u + v} ]}\big)].

b) Cmcm: O1: [\big({{1 \over 4} + {u_1},0,0}\big)]; O2: [\left({0,{1 \over 4} - {v_2},{w_2}} \right)]; O3: [\left({{1 \over 4} + {u_3},{1 \over 4} + {v_3},{1 \over 4}} \right)].

[\phi _{\rm b} = \arctan \left({2\left [{{u_3} + {w_2}} \right]} \right)]; [\phi _{\rm c} = \arctan \big({2[{{u_1} + {v_2}}]}\big)].

c) I4/mcm: O: [\big({1 \over 4} + u,{3 \over 4} + u,0 \big)]. [\phi _{c} = \arctan(4u)].

Mountstevens et al. (2003[Mountstevens, E. H., Attfield, J. P. & Redfern, S. A. T. (2003). J. Phys. Condens. Matter, 15, 8315-8326.]) augmented this set with a result for space group Imma, which upon resetting in Ibmm, is as follows:

d) Ibmm: O2: [({1 \over 4},{1 \over 4},w)]. [\phi _{\rm a} = \arctan\big(4\cdot {2}^{{{1}\over{2}}}\cdot w\big)].

This expression is equivalent to Pbnm with u = v = 0.

In general, the agreement between values generated in this work and approximations a) to d) above is satisfactory for the in-phase tilts, ϕc. Antiphase angles ϕa for space groups Pbnm and Ibmm show poorer agreement. In order to investigate whether this is a systematic discrepancy, the structural data of Kennedy, Prodjosantoso et al. (1999[Kennedy, B. J., Prodjosantoso, A. K. & Howard, C. J. (1999). J. Phys. Condens. Matter, 11, 6319-6327.]) for CaTiO3 (Pbnm) at 1273 K were fed into the Excel program, yielding the following two sets of tilt values and inclination angles θ3: ϕa = 6.53°; ϕa,refined = 6.59°; θ3 = 8.92°; θ3,refined = 9.24°. These are to be compared with the tilt values given by the Kennedy approximation: ϕa = 9.19°; ϕa,refined = 9.24°. These results and those in the Pbnm (B: 4a) and Ibmm columns of Table 3[link] lead to the following two conclusions: (1) The ϕa angles calculated by the method of §2.3[link] are systematically lower than those generated by the Kennedy approximation; (2) The Kennedy approximation generates exact values of inclination angle for refined structures with regular octahedra, i.e. θ3,refined, and not tilt angle ϕa as defined in this work.

In the case of Cmcm, good agreement for ϕa between §2.3[link] and Kennedy, Prodjosantoso et al. (1999[Kennedy, B. J., Prodjosantoso, A. K. & Howard, C. J. (1999). J. Phys. Condens. Matter, 11, 6319-6327.]) is obtained for regular octahedra in the refined structure, but not for the distorted octahedra in the unrefined structure.

A rationalization for the smaller values of ϕa generated by the method of §2.3[link] is that the inclination angle θ3 of a regular octahedron will always be larger than an angle within a plane of projection perpendicular to a pseudocubic axis (Fig. 6[link]). The former angle is generated by the approximations of Kennedy, Prodjosantoso et al. (1999[Kennedy, B. J., Prodjosantoso, A. K. & Howard, C. J. (1999). J. Phys. Condens. Matter, 11, 6319-6327.]) and the latter angle by the method of §2.3[link].

2.5. PCRO parameters, inclination angles and tilt systems with numbers of degrees-of-freedom by space group

Whether regular octahedra can be formed in a particular space group depends on whether the octahedra have sufficient degrees of freedom (d.o.f.). The number of d.o.f. defining the octahedral anionic network is N(UC) + N(X), of which N(tilt) are used to define independent tilt angles. In P42/nmc with irregular octahedra, the tilt system is [{a^ + }{b^ + }{c^ - }], i.e. N(tilt) = 3. For regular octahedra, this is reduced to [{a}^{+}{a}^{+}{c}^{-}] with N(tilt) = 2. The parameter denoting the remaining d.o.f. available to construct the octahedra, N(PCRO), is equal to N(UC) + N(X) − N(tilt) (Table 4[link]). In space groups Pbnm and Cmcm, N(PCRO) values equal to six signify a capacity to form regular octahedra independently of the values of the two tilt angles. Although N(PCRO) is less than 6 in Ibmm, I4/mcm, P4/mbm and [R{\bar 3}c], this does not preclude the formation of regular octahedra, since the space group symmetry itself provides partial regularity: in Ibmm: a1 = a2 and θ23 = θ31; in I4/mcm and P4/mbm: a1 = a2 with θ23 = θ31 = θ12 = 90°; in [R{\bar 3}c]: a1 = a2 = a3 with θ23 = θ31 = θ12. In space group P42/nmc, the N(PCRO) value of less than 6 signifies that the two tilt angles in the refined structure will be interdependent.

Table 4
Summary of inclination angles, PCRO parameters and degrees of freedom by space group

Space group Pbnm Cmcm Ibmm I4/mcm and P4/mbm P42/nmc [R{\bar 3}c] [Pm{\bar 3}m]
3-axis in equation (1)[link] z z z z
Inclination angles [{{{\theta}}_1} \ne 0] [{{{\theta}}_1} \ne 0] [{{{\theta}}_1} = {{{\theta}}_2} \ne 0] [{{{\theta}}_1} \ne 0] [{{{\theta}}_1} \ne 0] [{{{\theta}}_1} = {{{\theta}}_2} = {{{\theta}}_3} \ne 0] [{{{\theta}}_1} = {{{\theta}}_2} = {{{\theta}}_3} = 0]
  [{{{\theta}}_2} \ne 0] [{{{\theta}}_2} \ne 0] [{{{\theta}}_2} \ne 0] [{{{\theta}}_2} \ne 0]
  [{{{\theta}}_3} \ne 0] [{{{\theta}}_3} \ne 0] [{{{\theta}}_3} \ne 0] [{{{\theta}}_3} = 0] [{{{\theta}}_3} \ne 0]
PCRO parameter equalities     a1 = a3 a1 = a2   a1 = a2 = a3 a1 = a2 = a3
      [{{{\theta}}_{23}} = {{{\theta}}_{12}}] [{{{\theta}}_{23}} = {{{\theta}}_{31}} = {{{\theta}}_{12}} = 90^\circ]   [{{{\theta}}_{23}} = {{{\theta}}_{31}} = {{{\theta}}_{12}}] [{{{\theta}}_{23}} = {{{\theta}}_{31}} = {{{\theta}}_{12}} = 90^\circ]
Tilt system [{a^ - }{a^ - }{c^ + }] [{a^0}{b^ - }{c^ + }] [{a^ - }{a^ - }{c^0}] [{a^0}{a^0}{c^ - }] and a0a0c+ [{a^ + }{b^ + }{c^ - }] or [{a^ + }{a^ + }{c^ - }] [{a^ - }{a^ - }{a^ - }] a0a0a0
               
N(UC) 3 3 3 2 2 2 1
N(X) 5 5 2 1 5 1 0
N(tilt) 2 2 1 1 2 1 0
N(PCRO) 6 6 4 2 5 2 1
               
N(A) 2 2 1 0 0 0 0
N(B) 0 0 0 0 0 0 0

The final two rows in Table 4[link] relate to the d.o.f. assigned to A and B cations. The structures associated with the A cations are discussed in §2.7[link], after the primary structures associated with the parameter N(tilt) have been considered in the following sub-section.

2.6. Additional anionic network parameters related to octahedral tilting

Since octahedral tilting per se is a secondary structural feature relative to the BX6 octahedra, the primary structures resulting from this tilting necessarily concern the coordination of the A ions. It has long been customary to assume AX8 polyhedra in GdFeO3-type perovskites such as CaTiO3 in space group Pbnm (Liu & Liebermann, 1993[Liu, X. & Liebermann, R. C. (1993). Phys. Chem. Miner. 20, 171-175.]). Mitchell & Liferovich (2004[Mitchell, R. H. & Liferovich, R. P. (2004). J. Solid State Chem. 177, 4420-4427.]), adopting the perspective of coordination chemistry, reported an increase in coordination number from 8 to 9 with x in the solid solution series Ca1–xNaxTi1–xTaxO3. Mitchell (2002b[Mitchell, R. H. (2002b). Perovskites: Modern and Ancient, pp. 27-29. Thunder Bay, Ontario: Almaz Press.]) states generally that coordination numbers of 8, 9, 10 and 12 are possible, and calculates as an example the volume of uncoordinated space in SrZrO3 (I4/mcm) as 8.7%. This is the difference between VA, the volume assigned to AX12 polyhedra from volume filling with BX6 octahedra, and the volume of a more appropriate AX8 polyhedron, V(AX8 ).

From a functional viewpoint, the AX8 polyhedra in tilted perovskites provide scaffolding to support the eight octahedral faces of the full AX12 polyhedra. In Pbnm, the inner AX8 polyhedron is formed by the top and bottom parallelogram AX12 faces and four vertical struts (Fig. 8[link]).

[Figure 8]
Figure 8
AX8 and AX12 polyhedra in space group Pbnm. The two opposite, yellow parallelogram faces are joined by vertical blue struts to form AX8 polyhedra. The eight green faces are shared with BX6 octahedra.

It is now appropriate to interpret Fig. 6[link] in terms of the effect of octahedral tilting on A-ion coordination. The c+ tilting causes parallelograms PQRS to be formed instead of rectangles, with a concomitant reduction in area A(PQRS). The a tilting, by comparison, leads to equal and opposite displacements of vertices P, Q, R and S along z (denoted by + and −). The volume contributions made by plane PQRS and the other five faces of the AX8 polyhedron to V(AX8) are not affected by this tilting, since changes brought about by the equal and opposite displacements relative to the centre-of-symmetry at [0,0,0] cancel one another out.3 The mirror plane through A, B, C and D in all space groups with c+ tilting means that the horizontal positions, in this case the x and y coordinates of the bottom and top parallelograms in Fig. 8[link], are identical. It follows that the formula V(AX8) = A(PQRS)(c/2) could be used, since the centres of both parallelograms are separated by c/2 in all space groups with in-phase c-axis tilting apart from P4/mbm, where the separation is c. In space groups P42/nmc and I4/mcm, to which anti-phase, c tilting applies, the same formula is valid, since the mean cross-section [\perp z] is A(PQRS ).

Since the [{c}^{\pm}] component of the tilting is solely responsible for the reduction in A(PQRS) and V(AX8), the converse applies that A(PQRS) quantifies the extent of this tilting. Without [{c}^{\pm}] tilting, A(PQRS) would be equal to one quarter of the unit cell cross-sectional area, i.e. (ab)/4 in crystal systems with perpendicular x and y axes. A dimensionless coefficient [{\eta _A}] to quantify the amount of [{c}^{\pm}] tilting may therefore be defined as follows.

[{\eta _A} = {{A(PQRS)} \over {(ab)/4}} \eqno(5)]

Although the [{b^ - }] tilting about the xPC and yPC axes in Pbnm, more generally the [{a^ \pm }{b^ \pm }] tilting component when applied to all space groups, does not affect V(AX8 ), it does affect VB. For a given perovskite compound with unit-cell cross-section ab, this tilting allows larger VB values than would otherwise have been the case. The greater the degree of tilting, the larger the VB value. A dimensionless coefficient [{\eta _B}] to quantify this is given by relating VB to the volume of an upright, untilted octahedron of equal basal area in xy projection, [V_{B,{\rm ref}}]. The in-plane components of vectors p and q in Fig. 6[link] define the waist of this octahedron of perpendicular height c/2. These depend in turn on PCRO vectors a1 and a2.

[{\bf p} = {{1}\over{2}}\left(\matrix{{a}_{2x}+{a}_{1x}\cr {a}_{2y}+{a}_{1y}\cr 0}\right)]

and

[{\bf q} = {1 \over 2}\left({\matrix{ {{a_{2x}} - {a_{1x}}} \cr {{a_{2y}} - {a_{1y}}} \cr 0 \cr } } \right).]

[{\eta _B} = {{{V_B}} \over {{V_{B,{\rm ref}}}}} \eqno(6)]

In the presence of [{a^ \pm }{b^ \pm }] tilting, ηB > 1. VB is calculated as one-sixth of the volume of the PCRO and VB,ref as twice the volume of a right pyramid of height c/4 with parallelogram base of area

[\left| {\matrix{ {{p_x}} & {{p_y}} \cr {{q_x}} & {{q_y}} \cr } } \right|,]

i.e.

[{{1}\over{3}}\cdot {{c}\over{2}} \left|\matrix{{p}_{x}& {p}_{y}\cr {q}_{x}& {q}_{y}}\right|. ]

Analytical expressions for [{\eta _A}] and [{\eta _B}] are quoted in Table 5[link]. The form of [{\eta _B}] is similar for all space groups, and [{\eta _A}] can generally be traced back to simple expressions involving fractional coordinates. Particularly significant are the fixed, limiting values of [{\eta }_{A} = 1] and [{\eta _B} = 1] in space groups Ibmm and I4/mcm, respectively. These govern the sequence of phase transitions in series with increasing VA volume (see §3.1[link]). The corresponding derivations are in §S4 of the supporting information, along with diagrams of the [{c^ \pm }] and [{a^ \pm }{b^ \pm }] tilt patterns determining [{\eta _A}] and [{\eta _B}], respectively. These parameters are also calculated in the Excel datafile in the supporting information.

Table 5
Analytical expressions for structural parameters ηA and ηB to quantify [{c^ \pm }] and [{a^ \pm }{b^ \pm }] tilting

Space group N(tilt) ηA ηB
Pbnm (B in 4b) 2 (1 − [3 − 4x(O2)][4y(O2) − 1]) [4({\bf a}_{3} \cdot {\bf a}_{1} \times {\bf a}_{2})\times \Big( c \Big|\matrix{ a_{2x} + a_{1x} & a_{2y} + a_{1y}\cr a_{2x} - a_{1x} & a_{2y} - a_{1y} }\Big|\Big)^{-1}]
Pbnm (B in 4a) 2 ([1 + [4x(O2) − 1][4y(O2) − 1]) As for Pbnm (B in 4b)
Cmcm 2 16x(O1)y(O2) As for Pbnm (B in 4b)
 
Ibmm 1 1 As for Pbnm (B in 4b)
P4/mbmI4/mcm 1 8x(O2)(1− 2x(O2)) 1
P42/nmc 3 2a,2b: 4(−[{3\over 2}] + 2y(O3))([{1 \over 2}]+ 2y(O2)) As for Pbnm (B in 4b)
4d: 4([{{5}\over{2}}] − 2y(O3))([{{1}\over{2}}] − 2y(O2))
[R{\bar 3}c] 1 4x(O)(1 − x(O))

Since [{c^ \pm }] tilting reduces V(AX8 ) and therefore VA, whereas [{a^ \pm }{b^ \pm }] tilting increases VB, use of VA/VB as an indicator of the overall degree of tilting is vindicated. By using parameters [{\eta }_{A}] and [{\eta _B}] to quantify the tilting, which are simply calculated, the difficulties of defining and calculating tilt angles unequivocally can be circumvented. As may be inferred from the complexity of Fig. 6[link] and the comparison of calculated tilt angles in the lower part of Table 3[link], this could be seen as an advantage.

2.7. A cation positions

In earlier work on the parameterization of centrosymmetric perovskites (Thomas, 1998[Thomas, N. W. (1998). Acta Cryst. B54, 585-599.]), difficulties were experienced in fixing the positions of the A ions by reference to their coordinating X ions in the AX12 polyhedra. However, subsequent work by Magyari-Köpe et al. (2001[Magyari-Köpe, B., Vitos, L., Johansson, B. & Kollár, J. (2001). Acta Cryst. B57, 491-496.], 2002[Magyari-Köpe, B., Vitos, L., Johansson, B. & Kollár, J. (2002). Phys. Rev. B, 66, 092103-1.]) demonstrated that their positions could be treated as a function of the VA/VB ratio.

The essential objective is to quantify the small displacements from geometrically regular positions [0, 0] and [½, ½] in the relevant plane of projection (Fig. 9[link]). The parameterization applied here is limited to a transformation to polar coordinates L and ϕ in Pbnm [Fig. 9[link](a)], whereby L is twice the displacement of an individual ion. In space group Cmcm, two independent lengths, L1 and L2, apply [Fig. 9[link](b)]. These displacements are either along one axis or non-existent in the other space groups.

[Figure 9]
Figure 9
A cation positions in space groups (a) Pbnm and (b) Cmcm in 2D projection. Green circles: height +¼; brown circles: height −¼. A network with dashed lines is shown in (a).

3. Analysis of structures at variable temperature and chemical composition

The structural parameters defined in §2[link] are used here to characterize the sequences of phase transitions observed with increasing temperature and varying chemical composition in centrosymmetric perovskites. Li et al. (2004[Li, L., Kennedy, B. J., Kubota, Y., Kato, K. & Garrett, R. F. (2004). J. Mater. Chem. 14, 263-273.]) identified the following commonly occurring sequence with increasing temperature or mean A ion radius: PbnmIbmmI4/mcm[Pm{\bar 3}m]. This applies to Sr1−xBaxHfO3 and Sr1–xBaxZrO3 perovskites (Kennedy et al., 2001[Kennedy, B. J., Howard, C. J., Thorogood, G. J. & Hester, J. R. (2001). J. Solid State Chem. 161, 106-112.]), as well as to the systems Sr1–xBaxSnO3 and Ca1–xSrxSnO3 (Mountstevens et al., 2003[Mountstevens, E. H., Attfield, J. P. & Redfern, S. A. T. (2003). J. Phys. Condens. Matter, 15, 8315-8326.]). Identification of the intermediate Ibmm phase was significant, as it had previously been regarded as rare (Kennedy et al., 2001[Kennedy, B. J., Howard, C. J., Thorogood, G. J. & Hester, J. R. (2001). J. Solid State Chem. 161, 106-112.]). The temperature range of stabilization of this phase is highly variable, with values greater than 570 K in BaPbO3 (Fu et al., 2007[Fu, W. T., Visser, D., Knight, K. S. & IJdo, D. J. W. (2007). J. Solid State Chem. 180, 1559-1565.]), of 140 K in SrRuO3 (Kennedy et al., 2002[Kennedy, B. J., Hunter, B. A. & Hester, J. R. (2002). Phys. Rev. B, 65, 224103-1.]), 90 K in SrHfO3 (Li et al., 2004[Li, L., Kennedy, B. J., Kubota, Y., Kato, K. & Garrett, R. F. (2004). J. Mater. Chem. 14, 263-273.]) and less than 20 K in SrRhO3 (Kennedy et al., 2004[Kennedy, B. J., Yamaura, K. & Takayama-Muromachi, E. (2004). J. Phys. Chem. Solids, 65, 1065-1069.]). It is characterized by a fixed value of parameter [{\eta _A}] equal to one (Table 5[link]).

A second sequence was identified by Ahtee & Darlington (1980[Ahtee, M. & Darlington, C. N. W. (1980). Acta Cryst. B36, 1007-1014.]) for NaTaO3: [Pbmn \to Cmcm \to P4/mbm \to Pm\bar 3m], with further structural work by Kennedy, Prodjosantoso et al. (1999[Kennedy, B. J., Howard, C. J. & Chakoumakos, B. C. (1999). J. Phys. Condens. Matter, 11, 1479-1488.]) reporting transition temperature ranges of 738–753 K, 823–863 K and 883–913 K. More recently, Mitchell et al. (2014[Mitchell, R. H., Burns, P. C., Knight, K. S., Howard, C. J. & Chakhmouradian, A. R. (2014). Phys. Chem. Miner. 41, 393-401.]) reported the same series of phase symmetries for the mineral lueshite, NaNbO3.

Both sequences are characterized by an increasing VA/VB ratio with T or x, which may be attributed either to a larger volume expansion coefficient of VA relative to VB or to a compositionally induced expansion of the A-site volume relative to the B site. The term `A-ion perturbation' may be applied in both cases, since a change in the A-site geometry is the dominant driving force causing the whole structure to respond.

A solid solution series with a rising VA/VB ratio brought about by reducing VB has been established by Yang (2008[Yang, J. (2008). Acta Cryst. B64, 281-286.]) for LaCr1–xNixO3 at room temperature. Upon increasing x from 0 to 0.7, a Pbnm[R{\bar 3}c] phase transition is observed. In this case, a B-ion perturbation is being applied.

3.1. x- and T-series with A-ion perturbations

3.1.1. SrxBa1–xSnO3 and SrxBa1–xHfO3

Table 6[link] contains structural parameters calculated from the refinements of Mountstevens et al. (2003[Mountstevens, E. H., Attfield, J. P. & Redfern, S. A. T. (2003). J. Phys. Condens. Matter, 15, 8315-8326.]) for the solid solution series SrxBa1–xSnO3 and of Li et al. (2004[Li, L., Kennedy, B. J., Kubota, Y., Kato, K. & Garrett, R. F. (2004). J. Mater. Chem. 14, 263-273.]) for the SrxBa1–xHfO3 series.

Table 6
Structural parameters for two solid solutions spanning space group Ibmm at room temperature

Composition ICSD Space group ηA ηB VA3) VB3) VA/VB λ σ
SrSnO3 190602 Pbnm 0.9847 (1) 1.0313 (3) 54.181 (5) 11.454 (4) 4.730 (2) 0.0023 (4) 0.0104 (8)
Sr0.8Ba0.2SnO3 190611 Pbnm 0.9935 (2) 1.0236 (3) 55.10 (6) 11.42 (1) 4.824 (2) 0.0052 (7) 0.0195 (9)
Sr0.6Ba0.4SnO3 190610 Ibmm 1 1.0147 (2) 56.02 (7) 11.40 (2) 4.913 (1) 0.0004 (3) 0.0091 (8)
Sr0.4Ba0.6SnO3 190609 I4/mcm 0.9960 (1) 1 56.9 (1) 11.43 (2) 4.9758 (6) 0.0000 (4) 0
Sr0.2Ba0.8SnO3 190608 [Pm{\bar 3}m] 1 1 57.52 (8) 11.50 (2) 5 0 0
BaSnO3 190601 [Pm{\bar 3}m] 1 1 58.071 (4) 11.6142 (8) 5 0 0
SrHfO3 89383 Pbnm 0.9856 (6) 1.034 (1) 55.64 (2) 11.79 (2) 4.719 (8) 0.003 (2) 0.013 (1)
Sr0.8Ba0.2HfO3 55746 Pbnm 0.996 (2) 1.029 (4) 57.16 (6) 11.89 (6) 4.81 (3) 0.033 (6) 0.058 (9)
Sr0.6Ba0.4HfO3 55747 Ibmm 1 1.026 (1) 57.97 (1) 11.96 (1) 4.848 (7) 0.0058 (5) 0.039 (3)
Sr0.4Ba0.6HfO3 55748 I4/mcm 0.98755 1 58.88 (1) 11.95 (1) 4.926 (5) 0.0013 (2) 0
Sr0.2Ba0.8HfO3 55749 [Pm{\bar 3}m] 1 1 59.7948 (9) 11.9590 (2) 5 0 0
BaHfO3 [Pm{\bar 3}m] 1 1 60.2909 (4) 12.0582 (1) 5 0 0

On reading Table 6[link] from top to bottom, values of VA confirm that the increasing mean A-ion radius is the principal driving force for the structural changes in both systems. VA rises by 7.2% between x = 0 and x = 1 in the tin-containing and by 8.4% in the hafnium-containing series. Furthermore, the larger ionic radius of Hf4+ compared to Sn4+ in sixfold coordination [0.71 cf. 0.69 Å; Shannon (1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.])] causes the VA value at a given Sr:Ba ratio to be systematically larger in the system with hafnium as B ion. By comparison, the volume of the B site increases by 1.4% between x = 0 and x = 1 in the tin-containing series compared system compared to 2.3% in the hafnium series. The following ranges of the VA/VB parameter span the Ibmm phase in the two systems: 4.824 < VA/VB < 4.976 and 4.81 < VA/VB < 4.926.

The variation of parameters ηA and ηB with x reveals the mechanisms behind the PbnmIbmm and IbmmI4/mcm phase transitions in these series [Figs. 10[link](a) and 10[link](b)].

[Figure 10]
Figure 10
Characterization of phase transitions in terms of tilt-related parameters ηA and ηB: (a) SrxBa1–xSnO3, (b) SrxBa1–xHfO3, (c) BaPbO3, (d) CaTiO3. Circles denote experimentally determined points, which are joined by straight lines. Arrows signify limited flexibility in the positions of the phase transitions.

Reading these two diagrams from left to right, the PbnmIbmm transition is triggered by ηA reaching the limiting value of 1 whilst ηB is still > 1. The range of stability of the Ibmm phase is determined by the x value at which ηB becomes equal to 1, whereupon a phase transition to I4/mcm takes place. The diagrams suggest that the approach of ηA to 1 takes place continuously at the PbnmIbmm boundary. By comparison, the IbmmI4/mcm boundary is characterized by a discontinuous step in ηB down to 1. Also characteristic of the latter boundary is a fall in ηA to a value less than 1. This may be understood by considering that ηA and ηB can only both be equal to one in the aristotype [Pm\bar 3m] phase. However, at the start of the range of stability of the I4/mcm phase, VA/VB has not reached the limiting value of 5. The transition from I4/mcm to [Pm\bar 3m] takes place as soon as VA has increased sufficiently for this VA/VB limit to be reached.

3.1.2. BaPbO3

The above rationalization can be transferred without modification to series where temperature is responsible for an increasing VA volume. For example, the Ibmm phase is stabilized over a wide temperature range in BaPbO3 perovskites, these being of technological relevance due to superconductivity in the Ba-Pb-Bi-O system. Following earlier structural refinements by Moussa et al. (2001[Moussa, S. M., Kennedy, B. J. & Vogt, T. (2001). Solid State Commun. 119, 549-552.]) and Ivanov et al. (2001[Ivanov, S. A., Eriksson, S.-G., Tellgren, R. & Rundlöf, H. (2001). Mater. Sci. Forum, 378-381, 511-516.]) in the monoclinic space group C2/m on cooling and at room temperature, respectively, Fu et al. (2005[Fu, W. T., Visser, D. & IJdo, D. J. W. (2005). Solid State Commun. 134, 647-652.]) interpreted this monoclinic distortion as probably due to twinning. They also provided four structural refinements in Ibmm at room temperature and at 4.2 K. Upon heating, Fu et al. (2007[Fu, W. T., Visser, D., Knight, K. S. & IJdo, D. J. W. (2007). J. Solid State Chem. 180, 1559-1565.]) reported phase changes to tetragonal I4/mcm and thereafter to [Pm \overline{3}m] at approximate temperatures of 573 and 673 K respectively. Fig. 10[link](c) is generated from experimental points calculated from the structural refinements of Fu et al. (2005[Fu, W. T., Visser, D. & IJdo, D. J. W. (2005). Solid State Commun. 134, 647-652.], 2007[Fu, W. T., Visser, D., Knight, K. S. & IJdo, D. J. W. (2007). J. Solid State Chem. 180, 1559-1565.]). The absence of Pbnm symmetry in BaPbO3 can be ascribed to the VA/VB ratio remaining above 4.8320 (8) over the temperature range from 4.2 K to 773 K, this being higher than the threshold values of 4.81 and 4.824 noted in §3.1.1[link] for Ibmm symmetry. This interpretation is supported by the stabilization of the strontium compound SrPbO3 in Pbnm (Fu & Ijdo, 1995[Fu, W. T. & Ijdo, D. J. W. (1995). Solid State Commun. 95, 581-585.]). The calculated VA/VB value here is significantly lower, at 4.418 (2).

Despite the absence of a Pbnm phase in BaPbO3, Fig. 10[link](c) shows that the IbmmI4/mcm transition is also governed by discontinuous jumps in ηB to 1 and ηA to less than 1, as in Figs. 10[link](a) and 10[link](b). The approach to the aristotype phase follows the same principles as noted earlier.

3.1.3. CaTiO3

The mineral perovskite, CaTiO3, has not been observed in space group Ibmm, although it exists in Pbnm at room temperature (Sasaki et al., 1987[Sasaki, S., Prewitt, C. T., Bass, J. D. & Schulze, W. A. (1987). Acta Cryst. C43, 1668-1674.]) and at temperatures up to 1373 K (Liu & Liebermann, 1993[Liu, X. & Liebermann, R. C. (1993). Phys. Chem. Miner. 20, 171-175.]). Redfern (1996[Redfern, S. A. T. (1996). J. Phys. Condens. Matter, 8, 8267-8275.]) paved out the sequence of phase transitions at yet higher temperatures through I4/mcm to and [Pm\bar 3m] aristotype, reporting transition temperatures of 1373–1423 K and at ∼1523 K. Although Kennedy, Howard & Chakoumakos (1999[Kennedy, B. J., Howard, C. J. & Chakoumakos, B. C. (1999). J. Phys. Condens. Matter, 11, 1479-1488.]) subsequently raised the possibility of an orthorhombic Cmcm structure being stabilized at around 1380 K, this was refuted by Ali & Yashima (2005[Ali, R. & Yashima, M. (2005). J. Solid State Chem. 178, 2867-2872.]), who later generated the nineteen structural refinements upon which the experimental points of Fig. 10[link](d) are based (Yashima & Ali, 2009[Yashima, M. & Ali, R. (2009). Solid State Ionics, 180, 120-126.]). It is logical that no intermediate phase of symmetry Ibmm is formed with increasing temperature, since ηA remains significantly below one over the whole Pbnm range. This means that the degree of AO8 expansion induced thermally is smaller than the compositionally induced AO8-expansion analysed in §3.1.1[link]. Stabilization of the Pbnm phase is curtailed by ηB falling to 1 at the PbnmI4/mcm boundary. After the temperature of this phase transition has been reached, the overall expansion of the unit cell allows the volumetric requirements of the TiO6 octahedron to be satisfied without [{a^ - }] tilting in space group I4/mcm. The remaining tilting allows further comparatively greater AO8 expansion with increasing temperature up to the temperature of the phase transition to [Pm\bar 3m].

In all four series upon which Fig. 10[link] is based, the changes in tilt systems taking place at the phase transitions, these being critical events, are largely determined by the systematic changes in tilt angles taking place over wide compositional or temperature ranges. Parameters ηA and ηB are ideally suited for encapsulating these changes.

3.1.4. NaTaO3 and NaNbO3

These compounds provide an experimental basis for assembling the factors governing stabilization of the Cmcm phase, and more widely, the progression from Pbnm to [Pm\bar 3m] via Cmcm and P4/mbm instead of Ibmm and I4/mcm. I4/mcm and P4/mbm structures are very similar, the only difference being anti-phase or in-phase c axis tilting, respectively. Furthermore, structural parameters VA/VB, ηA and ηB cannot discriminate between these two structures. It is therefore appropriate to regard Cmcm and Imma, in the first instance, as alternative precursor phases to a generic phase of symmetry I4/mcm or P4/mbm.

An analysis of the phases of Cmcm symmetry in terms of VA/VB, ηA and ηB within these two compounds is given in Table 7[link].

Table 7
Focus on the Cmcm phases of NaTaO3 and NaNbO3 with comparative data for neighbouring Pbnm and P4/mbm phases

Compound ICSD Temperature (K) VA/VB ηA1 ηA2 ηB Reference
Pbnm              
NaTaO3 150430 293 4.74 (2) 0.988 (2)   1.032 (3) Mitchell & Liferovich (2004[Mitchell, R. H. & Liferovich, R. P. (2004). J. Solid State Chem. 177, 4420-4427.])
NaNbO3 192400 293 4.802 (5) 0.9868 (6)   1.0206 (6) Mitchell et al. (2014[Mitchell, R. H., Burns, P. C., Knight, K. S., Howard, C. J. & Chakhmouradian, A. R. (2014). Phys. Chem. Miner. 41, 393-401.])
Cmcm              
NaTaO3 280099 803 4.899 (2) 0.987 (3) 0.994 (3) 1.0076 (2) Darlington & Knight (1999[Darlington, C. N. W. & Knight, K. S. (1999). Acta Cryst. B55, 24-30.])
NaTaO3 241445 778 4.891 (2) 1.002 (2) 0.977 (2) 1.0076 (2) Knight & Kennedy (2015[Knight, K. S. & Kennedy, B. J. (2015). Solid State Sci. 43, 15-21.])
NaTaO3 239691 783 4.889 (3) 0.991 (5) 0.989 (5) 1.0087 (3) Arulnesan et al. (2016[Arulnesan, S. W., Kayser, P., Kennedy, B. J. & Knight, K. S. (2016). J. Solid State Chem. 238, 109-112.])
NaNbO3 192404 848 4.928 (3) 1.051 (3) 0.933 (3) 1.0041 (4) Mitchell et al. (2014[Mitchell, R. H., Burns, P. C., Knight, K. S., Howard, C. J. & Chakhmouradian, A. R. (2014). Phys. Chem. Miner. 41, 393-401.])
NaNbO3 192405 873 4.940 (2) 1.063 (3) 0.924 (3) 1.0033 (3) Mitchell et al. (2014[Mitchell, R. H., Burns, P. C., Knight, K. S., Howard, C. J. & Chakhmouradian, A. R. (2014). Phys. Chem. Miner. 41, 393-401.])
P4/mbm              
NaTaO3 88377 843 4.945 (1) 0.9907 (2)     Kennedy, Prodjosantoso et al. (1999[Kennedy, B. J., Howard, C. J. & Chakoumakos, B. C. (1999). J. Phys. Condens. Matter, 11, 1479-1488.])
NaNbO3 192406 898 4.9687 (7) 0.9948 (1)   1 Mitchell et al. (2014[Mitchell, R. H., Burns, P. C., Knight, K. S., Howard, C. J. & Chakhmouradian, A. R. (2014). Phys. Chem. Miner. 41, 393-401.])
†Doped with 1 mol% K.

The definitive characteristic of the Cmcm phase is the proximity of ηB values to one in Table 7[link]. This parameter has much larger values in the Ibmm phase, with ηB values greater than 1.03 observed [see Table 6[link] and Figs. 10[link](a), 10[link](b) and 10[link](c)]. This difference is due to there being only one tilt of type [{a^ \pm }{b^ \pm }] in Cmcm, compared to two in Ibmm (see §S4 of the supporting information). Values of ηB close to one confirm Cmcm as a precursor to an I4/mcm or P4/mbm phase, in which ηB is exactly equal to one. The existence of Cmcm symmetry in CaTiO3 at ∼1380 K, as proposed by Kennedy, Howard & Chakoumakos (1999[Kennedy, B. J., Howard, C. J. & Chakoumakos, B. C. (1999). J. Phys. Condens. Matter, 11, 1479-1488.]), is indeed possible, since ηB is close to one at this temperature.

Also to be noted in Table 7[link] is the splitting of ηA values made possible by there being two symmetry-independent A sites in Cmcm. Geometry simply requires a mean value 〈ηA〉 = (ηA1 + ηA2)/2 of less than one. This splitting is considerably more marked in NaNbO3. The ability for the lower ηA value to be significantly less than one could go some way towards rationalizing the phase coexistence of Pbnm and Cmcm phases at room temperature observed in NaTaO3 by Knight & Kennedy (2015[Knight, K. S. & Kennedy, B. J. (2015). Solid State Sci. 43, 15-21.]). A further contributory factor towards stabilization of Cmcm in 1 mol% K-doped samples of NaTaO3 at room temperature (Arulnesan et al., 2016[Arulnesan, S. W., Kayser, P., Kennedy, B. J. & Knight, K. S. (2016). J. Solid State Chem. 238, 109-112.]) could be preferential occupation of the sites with higher ηA value by the larger potassium ions, i.e. a partial ordering.

It is speculated that the proximity of ηB to one in Cmcm is conducive to a to a higher temperature phase transition to P4/mbm, whereas the sudden fall in ηB within Ibmm or Pbnm phases observed in Figs. 10[link](a)–10[link](d) favours a transition to I4/mcm.

3.2. X-series with B-ion perturbations

Calculated structural parameters for the LaCr1–xNixO3 solid solution series (Yang, 2008[Yang, J. (2008). Acta Cryst. B64, 281-286.]) are given in Table 8[link].

Table 8
Structural data relevant to the Pbnm[R{\bar 3}c] transition at x ∼ 0.7 in LaCr1–xNixO3

Composition ICSD Space group ηA ηB ϕc〉 (°) ϕa〉 (°) (〈ϕa〉 + 〈ϕc〉)/2 (°) VA3) VB3) VA/VB
x = 0 173469 Pbnm 0.9925 (1) 1.0355 (7) 5.02 (3) 8.5 (2) 6.78 (8) 48.38 (1) 10.182 (7) 4.751 (4)
x = 0.3 173471 Pbnm 0.9836 (4) 1.0278 (9) 8.67 (5) 5.4 (2) 7.05 (8) 48.28 (1) 10.178 (9) 4.744 (5)
x = 0.6 173472 Pbnm 0.9900 (1) 1.0301 (8) 6.13 (4) 6.5 (2) 6.32 (10) 48.388 (9) 10.152 (8) 4.766 (5)
x = 0.7 173473 Pbnm 0.9983 (3) 1.0254 (5) 6.36 (3) 4.64 (9) 5.50 (5) 48.204 (9) 9.956 (6) 4.841 (3)
x = 0.7 173474 [R{\bar 3}c] 0.9891 (1)     5.97 (3)   47.453 (5) 9.867 (4) 4.809 (2)
x = 0.8 173475 [R{\bar 3}c] 0.9893 (2)     5.90 (5)   47.414 (7) 9.850 (5) 4.814 (3)
x = 0.9 173476 [R{\bar 3}c] 0.9897 (1)     5.79 (3)   47.29 (1) 9.810 (4) 4.820 (2)
x = 1 173477 [R{\bar 3}c] 0.9900 (1)     5.72 (3)   46.778 (3) 9.696 (2) 4.825 (1)

Unlike the structures perturbed by A ions, the variation in ηA and ηB parameters is small and also unsystematic for structures in space group Pbnm. There is no tendency of either parameter to approach one. The expected cross-correlations between parameter pairs ηAϕc and ηBϕa are observed within the Pbnm phase field: the greater the magnitude of the deviations of ηA and ηB from one, the larger the 〈ϕc〉 and 〈ϕa〉 angles. The ϕa tilt angle falls to 4.64° and the mean tilt angle (〈ϕa〉 + 〈ϕc〉)/2 to 5.50° in Pbnm just before the phase transition, with ϕa rising to 5.97° in the [R\bar 3c] phase at x = 0.7. A gradual reduction in ϕa to 5.72° is observed up to x = 1, this still being a stable rhombohedral phase far from a transition to the cubic aristotype. The gradual fall in VB with x is consistent with the smaller ionic radius of Ni3+ compared to Cr3+ (Shannon, 1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]) and is responsible for the systematic increase in VA/VB with x, which peaks at 4.841 in Pbnm before the phase transition to [R{\bar 3}c] and falls to 4.809 thereafter. The [Pbnm \to R\bar 3c] transition is examined further in §4.4[link].

4. Modelling the structural variation of Pbmn perovskites with increasing pressure

The structures adopted by perovskites under increasing hydro­static pressure rest on the different compressibilities of the AX12 and BX6 polyhedra (Angel et al., 2005[Angel, R. J., Zhao, J. & Ross, N. L. (2005). Phys. Rev. Lett. 95, 025503-1.]), such that the volume ratio VA/VB may either increase or decrease. In general, experimental difficulties limit the pressure range over which full structure refinements can be obtained, a common practice being to report the variation of unit-cell parameters over a wider pressure range by means of the third-order Birch–Murnaghan equation of state [equation (7)[link]].

[P = { 3K_{T} \over 2}\Big [\Big({V_{0} \over {V} } \Big)^{7/3} \!-\! \Big({ V_{0} \over {V}} \Big)^{5/3}\Big] \!\Big\{ 1 + {3 \over 4}\Big(K_{0}^{\prime} - 4 \Big)\Big [\Big({V_{0} \over {V}} \Big)^{2/3} \!-\! 1 \Big] \!\Big\} \eqno(7)]

Here, coefficient KT is the bulk modulus and [K_0^{\prime}] is its derivative with respect to pressure. V0 represents the reference unit-cell volume. The appropriate notation to use for the variation of cell parameters a, b, c with pressure would be Ka0, [K_{a0}^{\prime}], a0 etc., whereby cubes a3,b3,c3 will have been used for the Birch–Murnaghan fitting. It is also common to quote linear compressibilities [{\beta _a}], [{\beta _b}], [ {\beta }_{c}], whereby [{\beta _a} = (\partial / \partial P)(a / a_{0} )_{T}], etc., these generally holding over a limited pressure range. By setting [K_0^{\prime}] values equal to 4, the term in curly brackets is equal to one, giving rise to the second-order Birch–Murnaghan equation.

The primary motivation for studying perovskites under pressure has been to simulate MgSiO3 perovskite in the lower mantle, for which a phase transition to a post-perovskite phase at pressures > 125 GPa and a temperature of 2700 K has been proposed (Murakami et al., 2004[Murakami, M., Hirose, K., Kawamura, K., Sata, N. & Ohishi, Y. (2004). Science, 304, 855-858.]). This system is simulated by assuming regular octahedra in the following sub-section.

4.1. Simulation of MgSiO3 under pressures of up to 125 GPa

It may be assumed that MgSiO3 remains in space group Pbnm at all pressures up to ∼125 GPa. Accordingly, the 12 sets of a, b, c unit-cell parameters reported by Vanpeteghem et al. (2006[Vanpeteghem, C. B., Zhao, J., Angel, R. J., Ross, N. L. & Bolfan-Casanova, N. (2006). Geophys. Res. Lett. 33, L03306-1.]) at pressures up to 10 GPa were used to derive the following Birch–Murnaghan constants to second order for the unit-cell volume and a and c unit-cell parameters4: KT = 252.90 GPa; V0 = 162.52 Å3; KT,a = 232.11 GPa; V0,a = 109.08 Å3; KT,c = 240.78 GPa; V0,c = 328.33 Å3.

The assumption was made that these constants could extended to 125 GPa, thereby yielding theoretical a, b and c unit-cell parameters over this whole range. This process was validated by comparing the values obtained (`Extrapolated  B-M') with values quoted by Murakami et al. (2004[Murakami, M., Hirose, K., Kawamura, K., Sata, N. & Ohishi, Y. (2004). Science, 304, 855-858.]) and Fiquet et al. (2000[Fiquet, G., Dewaele, A., Andrault, D., Kunz, M. & Le Bihan, T. (2000). Geophys. Res. Lett. 27, 21-24.]). Acceptable agreement is found in Table 9[link].

Table 9
Validation of Birch–Murnaghan (B-M) constants for use at pressures of up to to ∼125 GPa

Method/calculation type p (GPa) a (Å) b (Å) c (Å) Vuc3)
Murakami XRD 109 4.325 4.579 6.308 124.9
Murakami MD 109 4.403 4.574 6.410 129.1
Extrapolated B-M 109 4.365 4.559 6.315 125.7
Fiquet synchrotron XRD 40.66 143.26
Extrapolated B-M 40.71 4.571 4.743 6.606 143.22
Vuc: unit-cell volume.

200 equally spaced pressures were taken within the range up to 125 GPa and a, b, c values calculated from the above Birch–Murnaghan constants. Values of x(O1), y(O1), x(O2), y(O2) and z(O2) were allowed to vary within the core Excel Solver functionality for space group Pbnm (Fig. 7[link]), so as to generate regular octahedra. The resulting variation of parameters ηA and ηB with pressure is shown in Fig. 11[link](a).

[Figure 11]
Figure 11
Simulations of MgSiO3 and YAlO3 in terms of ηA and ηB. (a) MgSiO3 with regular octahedra up to p = 125 GPa; (b) comparison of YAlO3 with regular and distorted octahedra up to p = 8 GPa; (c) Extrapolation of YAlO3 with regular and distorted octahedra up to p = 52 GPa. Dashed lines indicate pressures for which structural data are quoted in Table 10[link].

Parameter ηA decreases from 0.9576 at 0 GPa to 0.9529 at 125 GPa, which indicates an increase in c+ tilting. By comparison, ηB increases from 1.0649 at 0 GPa to 1.0939 at 125 GPa, indicating an increase in [{a^ - }] tilting. The effect of pressure is to drive the structure ever further into the Pbnm phase field. It is likely that the imminent phase transition to post-perovskite at 125 GPa is triggered by the strong O⋯O repulsions that will be associated with this degree of octahedral tilting.

4.2. Simulation of YAlO3 at pressures up to 52 GPa

Ross et al. (2004[Ross, N. L., Zhao, J. & Angel, R. J. (2004). J. Solid State Chem. 177, 1276-1284.]) elucidated the response of the perovskite compound YAlO3 at pressures up to 8 GPa by providing eight structural refinements of a synthetic single crystal. A decrease in octahedral tilting with pressure was found, which was ascribed to the AlO6 octahedral compressibility being greater than that of the YO12 site. It follows that an approach towards cubic symmetry will occur, similar to the T-series with A ion perturbation in §3.1[link]. The Birch–Murnaghan constants to third order quoted by Ross et al. (2004[Ross, N. L., Zhao, J. & Angel, R. J. (2004). J. Solid State Chem. 177, 1276-1284.]) were used to generate unit-cell parameters at different pressures, as in §4.1[link]. Two alternative simulations were carried out for pressures up to 8 GPa, the first assuming regular octahedra as in §4.1[link], and the second exploiting the octahedral distortions reported by Ross et al. (2004[Ross, N. L., Zhao, J. & Angel, R. J. (2004). J. Solid State Chem. 177, 1276-1284.]) from an analysis of their structural refinements. For this purpose, the linear, decreasing trends in the three different Al—O bond lengths with increasing pressure observed by them were translated into linear relationships for PCRO parameters a1 to a3. In addition, linear relationships for PCRO angle parameters θ23, θ31 and θ12 were derived [equations (8)[link]].

[\eqalign{{a}_{1}\,(\AA) = 3.8421-0.007137\,[p({\rm GPa})] \cr {a}_{2}\,(\AA) = 3.8161-0.006407\,[p({\rm GPa})] \cr {a}_{3}\,(\AA) = 3.7915-0.005836\,[p({\rm GPa})]\cr \theta _{23}\,(^\circ ) = 90.893 - 0.034905\,[p({\rm GPa})]\cr \theta _{31}\,(^\circ) = 89.382 + 0.010050\,[p({\rm GPa})]\cr \theta _{12}\,(^\circ ) = 89.878 - 0.019462 \,[p({\rm GPa})]} \eqno(8)]

Solver-based refinements in Pbnm were carried out at equally spaced pressure values, the spacing determined by covering the pressure range from 0 to 60 GPa with 200 values. Parameters x(O1), y(O1), x(O2), y(O2) and z(O2) were allowed to vary in refinements with end-point determined by the minimum deviation from conditions (8). The deviation was observed to range between 0.001 and 0.050%.

The results of both simulations are shown in Fig. 11[link](b) and in that part of Fig. 11[link](c) with the blue-shaded background. The slope of both ηB curves is negative, compared to the positive slope of ηB for MgSiO3. The slope of ηA for the simulation with regular octahedra is positive, compared to the negative slope for MgSiO3. The opposite behaviour to MgSiO3 is shown by YAlO3, which shows a gradual progression away from Pbnm towards higher symmetry. The curve for ηA for distorted octahedra has a shallow minimum at ∼3.4 GPa, which is more clearly seen in Fig. 11[link](c). Thereafter ηA rises, as for undistorted octahedra.

In a computational experiment, the pressure was allowed to rise towards 60 GPa, with unit-cell constants calculated from the Birch–Murnaghan constants of Ross et al. (2004[Ross, N. L., Zhao, J. & Angel, R. J. (2004). J. Solid State Chem. 177, 1276-1284.]) for the pressure range up to 8 GPa. Fig. 11[link](c) shows that the curvature of both pairs of curves changes sign as the octahedra change from regular to distorted. Since the constraints in equations (8) only apply up to 8 GPa, it became increasingly difficult to reach low deviations at the end-points of the refinements. The maximum deviation amounted to 1.08%, which was obtained at the highest pressure.

The ascending curve for ηA with distorted octahedra in Fig. 11[link](c) reaches zero at a pressure of ∼41.5 GPa. This would be consistent with a phase transition to Ibmm [cf. Figs. 10[link](a) and 10[link](b)]. By comparison, the ηB curve is the first to reach zero in the simulation with regular octahedra. The corresponding pressure is ∼51.6 GPa. This would be consistent with a phase transition to I4/mcm or P4/mbm [cf. Fig. 10[link](d)]. The asymptotic approach of the ηB curve to 1 for regular octahedra gives rise to a broad pressure range in which Cmcm could be stabilized as well as Pbnm prior to this phase transition. This phenomenon has been observed in NaTaO3 by Knight & Kennedy (2015[Knight, K. S. & Kennedy, B. J. (2015). Solid State Sci. 43, 15-21.]) over a broad temperature range, as discussed in §3.1.4[link]. Assuming pattern similarity, a phase change to P4/mbm would be anticipated at 51.6 GPa. More generally, the simulations of YAlO3, both with and without octahedral distortion, indicate how this distortion is expected to have a direct effect on the sequence of phase transitions and the pressures at which they occur.

In summary, the analysis of (p–T–X)-induced phase transitions in terms of ηA and ηB parameters is a stimulus to further, targeted experimental work on the systems that have been analysed and simulated in §3.1[link], §4.1[link] and §4.2[link].

4.3. Crystal structures generated in the high-pressure simulations

A by-product of the simulations in §3.1[link] is the generation of full sets of oxygen ion coordinates. Table 10[link] contains these data corresponding to the conditions denoted by dashed lines in Figs. 11[link](a) and 11[link](c). The space group is Pbmn with the B ions in 4b positions. The full set of associated structural parameters is also quoted below the line in the table.

Table 10
Crystallographic and structural parameters generated in the simulations of MgSiO3 and YAlO3

Compound MgSiO3 MgSiO3 YAlO3 YAlO3 YAlO3 YAlO3
Pressure (GPa) 109 125 8 8 31.5 31.5
Octahedra Regular Regular Regular Distorted Regular Distorted
a (Å) 4.3649 4.3278 5.1285 5.1285 5.0308 5.0308
b (Å) 4.5592 4.5263 5.2443 5.2443 5.0751 5.0751
c (Å) 6.3151 6.2628 7.2899 7.2899 7.1349 7.1349
x(O1) 0.1067 0.1083 0.0756 0.0810 0.0470 0.0604
y(O1) 0.4780 0.4775 0.4925 0.4810 0.4965 0.4934
x(O2) 0.6936 0.6933 0.7241 0.7054 0.7309 0.7241
y(O2) 0.3017 0.3019 0.2748 0.2940 0.2687 0.2758
z(O2) 0.0533 0.0542 0.0378 0.0420 0.0235 0.0295
γPC (°) 92.49 92.57 91.28 91.28 90.50 90.50
VA3) 25.439 24.816 40.386 40.127 37.772 37.651
VB3) 5.979 5.855 8.630 8.889 7.769 7.890
VA/VB 4.255 4.239 4.680 4.514 4.862 4.772
θx (°) 16.89 17.05 10.27 13.64 6.88 8.90
θy (°) 16.89 17.05 10.27 13.71 6.88 8.91
θz (°) 16.79 17.03 12.06 13.19 7.58 9.72
ϕc〉 (°) 12.19 12.24 5.79 10.05 4.32 5.90
ϕa〉 (°) 12.05 12.25 8.55 9.48 5.36 6.72
ηA 0.9534 0.9529 0.9898 0.9686 0.9943 0.9893
ηB 1.0910 1.0939 1.0457 1.0549 1.0177 1.0285
a1 (Å) 3.2981 3.2750 3.7272 3.7845 3.5989 3.6193
a2 (Å) 3.2981 3.2750 3.7272 3.7647 3.5989 3.6138
a3 (Å) 3.2981 3.2750 3.7272 3.7437 3.5989 3.6194
θ23 (°) 90 90 90 90.61 90 89.85
θ31 (°) 90 90 90 89.46 90 89.75
θ12 (°) 90 90 90 89.69 90 89.71
λPC 0 0 0 0.0037 0 0.0007
σPC 0 0 0 0.012 0 0.004

In all cases, the effect of increasing pressure is to reduce VA and VB. However, the ratio VA/VB is reduced in MgSiO3 and increased in YAlO3. This signifies movement further into the Pbnm phase field in the former case [Fig. 11[link](a)] and movement away towards higher symmetry in the latter [Fig. 11[link](c)]. This is borne out by the changes in inclination angles θx, θx, θx, tilt angles ϕa, ϕc and tilt-related parameters ηA, ηB.

4.4. Analysis of (La1–xNdx)GaO3 structures at pressures of up to 12 GPa

Apart from the above simulations, the structural parameters used in this work are applied to the structural refinements of Angel et al. (2007[Angel, R. J., Zhao, J., Ross, N. L., Jakeways, C. V., Redfern, S. A. T. & Berkowski, M. (2007). J. Solid State Chem. 180, 3408-3424.]) for this 3:3 perovskite solid solution. In carrying out A-ion substitutional perturbations, it was found that a phase transition from Pbnm to [R\bar 3c] takes place under pressure for x values up to 0.20, but not for x = 0.62 or x = 1. The question arises as to whether a particular pattern of structural evolution within the Pbnm phase is associated with a phase transition to [R\bar 3c]. Data for the Pbmn and [R\bar 3c] phases at their lowest and highest investigated pressures for a given x value are given in Table 11[link].

Table 11
Structural parameters for (La1–xNdx)GaO3 in space groups Pbnm and [R\bar 3c]

x Space group p (GPa) ICSD ηA ηB ϕc〉 (°) ϕa〉 (°) (〈ϕa〉 + 〈ϕc〉)/2 (°) VA3) VA3) VA/VB
0 Pbnm 0.0001 160233 0.9926 (2) 1.0406 (7) 4.90 (8) 8.5 (1) 6.69 (7) 48.653 (8) 10.299 (8) 4.724 (4)
  Pbnm 2.038 (10) 160235 0.9939 (2) 1.0382 (7) 4.46 (7) 8.3 (1) 6.39 (7) 48.136 (7) 10.146 (7) 4.744 (4)
  [R\bar 3c] 2.356 (7) 160265 0.9868 (3)     6.56 (7)   48.024 (9) 10.066 (8) 4.771 (5)
  [R\bar 3c] 8.127 (23) 160270 0.9872 (3)     6.45 (7)   46.745 (8) 9.782 (8) 4.779 (4)
0.06 Pbnm 0.0001 160236 0.9902 (4) 1.044 (2) 5.7 (1) 9.1 (3) 7.4 (1) 48.50 (2) 10.33 (2) 4.693 (9)
  Pbnm 5.30 (5) 160240 0.9947 (4) 1.042 (2) 4.2 (2) 9.1 (3) 6.6 (2) 47.18 (2) 9.98 (2) 4.73 (1)
0.12 Pbnm 0.0001 160241 0.9907 (4) 1.040 (1) 5.5 (1) 8.5 (2) 7.0 (1) 48.55 (1) 10.29 (1) 4.716 (7)
  Pbnm 6.437 (7) 160243 0.9939 (3) 1.039 (1) 4.4 (1) 8.5 (2) 6.5 (1) 46.96 (1) 9.91 (1) 4.741 (6)
  [R{\bar 3}c] 8.020 (2) 160271 0.9875 (5)     6.4 (1)   46.65 (2) 9.75 (1) 4.783 (9)
  [R{\bar 3}c] 9.496 (13) 160272 0.9872 (4)     6.46 (9)   46.37 (1) 9.71 (1) 4.778 (6)
0.20 Pbnm 0.0001 160244 0.9884 (4) 1.0425 (7) 6.2 (1) 8.7 (1) 7.45 (7) 48.404 (9) 10.323 (8) 4.689 (5)
  Pbnm 8.671 (7) 160248 0.9931 (3) 1.042 (1) 4.7 (1) 8.8 (2) 6.79 (9) 46.38 (1) 9.83 (1) 4.720 (6)
0.62 Pbnm 0.0001 160250 0.9807 (5) 1.049 (1) 7.9 (1) 9.1 (1) 8.52 (8) 47.72 (1) 10.35 (1) 4.611 (6)
  Pbnm 9.432 160256 0.9842 (7) 1.046 (1) 7.2 (2) 8.9 (2) 8.0 (1) 45.60 (1) 9.81 (1) 4.648 (7)
1.00 Pbnm 0.0001 160257 0.9749 (5) 1.0537 (9) 9.00 (9) 9.6 (1) 9.28 (8) 47.18 (1) 10.36 (1) 4.555 (6)
  Pbnm 8.292 (9) 160264 0.9762 (4) 1.0514 (8) 8.78 (7) 9.4 (1) 9.11 (7) 45.324 (9) 9.909 (9) 4.574 (5)

The direct influence of the A-ion perturbations is seen in the values of VA, which decrease downwards in the table with increasing x for the p = 0.0001 GPa values. Since VB remains approximately constant, a parallel trend of decreasing VA/VB ratio with increasing x is generally observed. This is consistent with greater stabilization within the Pbnm phase field. Pressure induces the opposite trend to raising x, since the VA/VB ratios for maximum pressures within the Pbnm phase field are uniformly higher than at atmospheric pressure. A parallel increase in ηA and decrease in ηB is observed. At lower x values up to 0.20, which are associated with reduced Pbnm stabilization, the increased pressure induces a phase transition to [R\bar 3c]. Angel et al. (2007[Angel, R. J., Zhao, J., Ross, N. L., Jakeways, C. V., Redfern, S. A. T. & Berkowski, M. (2007). J. Solid State Chem. 180, 3408-3424.]) report the following approximate pressures for this phase transition: x = 0: 2.2 GPa; x = 0.06: 5.5 GPa; x = 0.12: 7.8 GPa; x = 0.20: 12 GPa. The expected cross-correlations between parameter pairs ηAϕc and ηBϕa are observed within the Pbnm phase field: the greater the magnitude of the deviations of ηA and ηB from one, the larger the 〈ϕc〉 and 〈ϕa〉 angles. It is proposed that the observed fall in 〈ϕc〉 to ∼4.2–4.4° with increasing pressure is the principal driving force for the phase transitions to [R{\bar 3}c]. After the transitions for x = 0 and x = 0.12, i.e. within the [R{\bar 3}c] phase field, the ϕa tilt angle is larger and approximately equal to the mean tilt angle (〈ϕa〉 + 〈ϕc〉)/2 in the Pbnm field beforehand. For x = 0.62 and x = 1.00, the critical range of ϕc between 4.2 and 4.4° is not reached at the pressures investigated, so that these compounds remain stabilized in space group Pbnm.

4.5. Structural parameters for YAl0.25Cr0.75O3 with locked octahedral tilting

Ardit et al. (2017[Ardit, M., Dondi, M. & Cruciani, G. (2017). Phys. Rev. B, 95, 024110-1.]) subsequently took up the theme of locked octahedral tilting in orthorhombic 3:3 perovskites (A ion +3; B ion +3) by reference to the solid solution YAl0.25Cr0.75O3 in space group Pbnm. Although in general agreement with Zhao et al. (2004[Zhao, J., Ross, N. L. & Angel, R. J. (2004). Acta Cryst. B60, 263-271.]) and Angel et al. (2005[Angel, R. J., Zhao, J. & Ross, N. L. (2005). Phys. Rev. Lett. 95, 025503-1.]) in assuming compressibility ratios β(AO12)/β(BO6) < 1 for a 3:3 perovskite and > 1 for a 2:4 perovskite, their compound showed a compressibility ratio approximately equal to one (Table 12[link]).

Table 12
Structural parameters for eleven structural refinements of Ardit et al. (2017[Ardit, M., Dondi, M. & Cruciani, G. (2017). Phys. Rev. B, 95, 024110-1.]) on YAl0.25Cr0.75O3

p (GPa) ηA ηB ϕc〉 (°) (§2.3[link]) ϕa〉 (°) (§2.3[link]) ϕc〉 (°) (Ardit) ϕa〉 (°) (Ardit) VA3) VB3) VA/VB
0.05 (5) 0.9557 (9) 1.082 (1) 11.9 (1) 11.7 (2) 12.0 (5) 16.6 (5) 43.52 (2) 10.10 (2) 4.309 (8)
1.44 (9) 0.9604 (4) 1.0868 (7) 11.24 (6) 12.32 (9) 12.7 (5) 16.6 (5) 43.238 (9) 10.030 (8) 4.311 (4)
3.32 (9) 0.9609 (5) 1.0862 (7) 11.18 (7) 12.32 (9) 12.7 (5) 16.5 (5) 42.892 (9) 9.938 (8) 4.316 (4)
4.93 (7) 0.9603 (5) 1.0871 (7) 11.25 (7) 12.36 (9) 12.8 (5) 16.6 (5) 42.561 (9) 9.878 (8) 4.309 (4)
6.36 (8) 0.9630 (5) 1.0890 (7) 10.88 (7) 12.67 (9) 13.1 (5) 16.5 (5) 42.321 (9) 9.812 (8) 4.313 (4)
8.48 (9) 0.9630 (5) 1.0890 (8) 10.88 (7) 12.67 (9) 13.1 (5) 16.6 (5) 41.954 (9) 9.727 (8) 4.313 (5)
10.54 (9) 0.9635 (6) 1.090 (1) 10.81 (9) 12.8 (1) 13.2 (5) 16.6 (5) 41.60 (1) 9.65 (1) 4.312 (6)
12.34 (7) 0.9580 (6) 1.086 (1) 11.57 (8) 12.2 (1) 12.6 (5) 16.5 (5) 41.28 (1) 9.59 (1) 4.303 (6)
14.40 (9) 0.9569 (8) 1.085 (1) 11.7 (1) 12.1 (2) 12.4 (5) 16.6 (5) 40.97 (1) 9.52 (1) 4.302 (7)
16.45 (9) 0.956 (1) 1.084 (2) 11.9 (2) 12.0 (2) 12.4 (5) 16.5 (5) 40.65 (2) 9.45 (2) 4.30 (1)
18.86 (9) 0.959 (2) 1.087 (3) 11.4 (3) 12.6 (3) 13.0 (5) 16.5 (5) 40.30 (3) 9.36 (3) 4.30 (2)

In spite of consistently falling VA and VB values with increasing pressure, the VA/VB ratio remains approximately constant, as do parameters [{\eta _A}], [{\eta _B}], 〈ϕc〉 and 〈ϕa〉. Discrepancies are observed between the values of the tilt angles calculated according to §2.3[link] and the values quoted by Ardit et al. (2017[Ardit, M., Dondi, M. & Cruciani, G. (2017). Phys. Rev. B, 95, 024110-1.]) without declaring the method of calculation. This is not unexpected, as discussed in §2.4[link]. Since values of 〈ϕc〉 and 〈ϕa〉 calculated according to §2.3[link] show the expected cross-correlations between parameter pairs ηAϕc and ηBϕa, indirect support is given for the correctness of these calculations.

5. Discussion

The ability of the method to analyse experimental structural data and the sequences of phase transitions between space groups has been demonstrated in §3 and §4. By comparison with the group-theoretical approach, it has not been necessary to relate these explicitly to the cubic aristotype, since deviations of VA/VB from the limiting value of 5 in the aristotype provide a direct indication of how far away a given structure is from the aristotype. Parameters VA, VB and VA/VB continue to yield valuable insight, for example in Table 11[link]. However, the direct calculation of Glazer tilt angles ϕc and [ϕa, ϕb] and associated derived parameters ηA and ηB has allowed a deeper analysis of the structural factors leading to phase transitions than a consideration of these volumes alone. The significance of the generalized algorithms introduced here for the Glazer tilt angles may be assessed by reference to Wang & Angel (2011[Wang, D. & Angel, R. J. (2011). Acta Cryst. B67, 302-314.]): `…the decomposition of a perovskite structure including tilted and distorted octahedra by geometric analysis does not result in an unambiguous definition of the Glazer (1972[Glazer, A. M. (1972). Acta Cryst. B28, 3384-3392.]) tilts and the problem is more acute in perovskites with lower space-group symmetries'. These authors noted further that `unambiguous expressions for both the Glazer tilts and their relationship to the VA/VB ratio are still to be determined explicitly for each space group, and in a general form'. These observations led Wang & Angel (2011[Wang, D. & Angel, R. J. (2011). Acta Cryst. B67, 302-314.]) to resort to group-theoretical methods in order to relate the amplitudes of symmetry-adapted modes to VA/VB ratio. The current work, by comparison, has remained strictly based on unit-cell parameters and atomic coordinates. Although it has not provided generalized analytical expressions for these tilt angles, it has led to two generalized algorithms (a) for all centrosymmetric space groups apart from [R{\bar 3}c] and (b) for space group [R{\bar 3}c] (and more generally, triclinic space groups). An analytical link between ϕc, [ϕa, ϕb] and VA/VB has not been provided, since inclination angles [{\theta }_{1}] to [{\theta _3}] are more suitable for this purpose [equation (1)[link]]. However, the linked-cell implementation within Excel allows the empirical derivation of numerical relationships between Glazer tilt angles and VA/VB.

The methodological innovations of the current work and their benefits are summarized in Table 13[link].

Table 13
Summary of innovations in the current work

Innovation Benefits
Quantification of distortion of centro­symmetric octahedra via PCRO. (a) Direct, concise visualization of octahedral distortion.
(b) Variation of PCRO parameters rooted in space group symmetry.
(c) New possibilities for simulating perovskite structures at extrapolated (p–T–X) conditions.
(d) Many more visualizable structural parameters cf. Thomas (1998[Thomas, N. W. (1998). Acta Cryst. B54, 585-599.]).
(e) Upgradable to non-centrosymmetric octahedra, and so applicable to all perovskites and more widely to all crystal structures containing octahedra.
(f) Provides summary distortion parameters λ and σ.
   
Generalized algorithms for calculating tilt angles [ϕa, ϕb] and ϕc. (a) Extension to space groups Cmcm, P42/nmc and [R{\bar 3}c] not previously covered by analytical approximations.
(b) Ability to calculate tilt angles in structures with distorted octahedra without approximation.
   
Structural parameters ηA and ηB. (a) ηA emphasizes the importance of AX8 inner polyhedra and focuses on octahedral tilting around the z (3) axis.
(b) ηB focuses on tilting around the x (1) and y (2) axes.
(c) Tracking structural evolution via ηA and ηB allows phase transitions between Pbnm and Ibmm, Cmcm, I4/mcm or P4/mbm to be rationalized and anticipated.
(d) Improved integration with the tilt classification of Glazer (1972[Glazer, A. M. (1972). Acta Cryst. B28, 3384-3392.]), and, by implication, group-theoretical methods.
   
Implementation in Excel Solver software environment. (a) Reversibility of transformation between crystallographic and structural parameters.
(b) The Excel file in the supporting information is a useful resource for calculating tilt angles and other parameters, as well as for refining crystallographic parameters under structural constraints.
(c) Upgradable to other programming languages (Frontline Systems Inc., 2021[Frontline Systems Inc. (2021). https://www.solver.com/.]).
   
Simulation of structural development at increasing pressure. The assumption of regular or idealized distorted octahedra allows the prediction of crystal structure and associated parameters, e.g. ηA, ηB from unit-cell parameters alone. The potential for modelling structures and phase transitions at high pressure is thereby increased.

The question of the relative stability of alternative perovskite phases has been addressed by several workers employing quantum-mechanical methods (Zagorac et al., 2014[Zagorac, J., Zagorac, D., Zarubica, A., Schön, J. C., Djuris, K. & Matovic, B. (2014). Acta Cryst. B70, 809-819.]) as well as the semi-empirical bond-valence method. Woodward (1997b[Woodward, P. M. (1997b). Acta Cryst. B53, 44-66.]) initiated this discussion by evaluating the ionic and covalent bonding in perovskites and analysing the conditions of stabilization of favoured tilt systems [{a^ - }{a^ - }{c^ + }], [{a^ - }{a^ - }{a^ - }] and a0a0a0. In later work (Lufaso & Woodward, 2001[Lufaso, M. W. & Woodward, P. M. (2001). Acta Cryst. B57, 725-738.]), an algorithm was developed to minimize the so-called global instability index (GII) in alternative tilt systems, this being the r.m.s. deviation between calculated bond valences and ideal cationic valences. The bond-valence method was also the favoured approach of Zhao et al. (2004[Zhao, J., Ross, N. L. & Angel, R. J. (2004). Acta Cryst. B60, 263-271.]) in rationalizing the relative compressibilities of the AO12 and BO6 polyhedra in perovskites. It led to a clear differentiation in behaviour between 2:4 and 3:3 perovskites, which has remained a feature of experimentally led investigations of perovskites under pressure, for example by Ardit et al. (2017[Ardit, M., Dondi, M. & Cruciani, G. (2017). Phys. Rev. B, 95, 024110-1.]).

The potential of direct transformation of crystallographic data into structural parameters for the analysis of sequences of phase transitions has been demonstrated in this work. The set of structural parameters has been extended and a baseline provided for future investigations of non-centrosymmetric perovskites. The intention is to promote more detailed interaction between experimental crystallography and modelling in developing new materials by atomic and molecular design.

6. Related literature

The following references are cited in the supporting information: Hahn (1995[Hahn, T. (1995). Editor. International Tables for Crystallography, Vol. A, Space-Group Symmetry. Dordrecht: Kluwer.]), Williams (1971[Williams, D. E. (1971). Acta Cryst. A27, 452-455.]).

Supporting information


Footnotes

1Since oxides are exclusively considered in this work, the notation `O1,O2…' is used interchangeably with `X1,X2…'.

2This activates a VBA macro, which calculates the values of structural parameters for the cells with a pink background. Although not included in the supporting information, the author is willing to supply an xlsm file containing VBA codes on request.

3Polyhedral volume = [{1 \over 3}\textstyle\sum \limits_i {A_i}{h_i}]. Ai is area of face; hi is perpendicular distance from face to common apex inside polyhedron.

4Values of KT and V0 are in good agreement with the values quoted by Vanpeteghem et al. (2006[Vanpeteghem, C. B., Zhao, J., Angel, R. J., Ross, N. L. & Bolfan-Casanova, N. (2006). Geophys. Res. Lett. 33, L03306-1.]): 253 (1) GPa and 162.51 (2) Å3.

Acknowledgements

I thank student Yannick Breuer and an unknown referee of the article by Fricke & Thomas (2021[Fricke, M. & Thomas, N. W. (2021). Acta Cryst. B77, 427-440.]) for rekindling my interest in perovskites. Thanks are likewise due to an anonymous referee of this article for some good suggestions. Finally, the work of staff of the Rheinische Landesbibliothek in Koblenz in retrieving numerous scientific articles is gratefully acknowledged. Open access funding enabled and organized by Projekt DEAL.

References

First citationAhtee, M. & Darlington, C. N. W. (1980). Acta Cryst. B36, 1007–1014.  CrossRef ICSD CAS IUCr Journals Web of Science Google Scholar
First citationAli, R. & Yashima, M. (2005). J. Solid State Chem. 178, 2867–2872.  Web of Science CrossRef ICSD CAS Google Scholar
First citationAngel, R. J., Zhao, J. & Ross, N. L. (2005). Phys. Rev. Lett. 95, 025503-1.  CrossRef PubMed CAS Google Scholar
First citationAngel, R. J., Zhao, J., Ross, N. L., Jakeways, C. V., Redfern, S. A. T. & Berkowski, M. (2007). J. Solid State Chem. 180, 3408–3424.  Web of Science CrossRef ICSD CAS Google Scholar
First citationArdit, M., Dondi, M. & Cruciani, G. (2017). Phys. Rev. B, 95, 024110-1.  CrossRef Google Scholar
First citationArulnesan, S. W., Kayser, P., Kennedy, B. J. & Knight, K. S. (2016). J. Solid State Chem. 238, 109–112.  CrossRef ICSD CAS Google Scholar
First citationDarlington, C. N. W. & Knight, K. S. (1999). Acta Cryst. B55, 24–30.  Web of Science CrossRef ICSD CAS IUCr Journals Google Scholar
First citationFiquet, G., Dewaele, A., Andrault, D., Kunz, M. & Le Bihan, T. (2000). Geophys. Res. Lett. 27, 21–24.  Web of Science CrossRef ICSD CAS Google Scholar
First citationFricke, M. & Thomas, N. W. (2021). Acta Cryst. B77, 427–440.  CrossRef IUCr Journals Google Scholar
First citationFrontline Systems Inc. (2021). https://www.solver.com/Google Scholar
First citationFu, W. T. & Ijdo, D. J. W. (1995). Solid State Commun. 95, 581–585.  CrossRef ICSD CAS Google Scholar
First citationFu, W. T., Visser, D. & IJdo, D. J. W. (2005). Solid State Commun. 134, 647–652.  CrossRef ICSD CAS Google Scholar
First citationFu, W. T., Visser, D., Knight, K. S. & IJdo, D. J. W. (2007). J. Solid State Chem. 180, 1559–1565.  CrossRef ICSD CAS Google Scholar
First citationGlazer, A. M. (1972). Acta Cryst. B28, 3384–3392.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationGlazer, A. M. (1975). Acta Cryst. A31, 756–762.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationHahn, T. (1995). Editor. International Tables for Crystallography, Vol. A, Space-Group Symmetry. Dordrecht: Kluwer.  Google Scholar
First citationHoward, C. J. & Stokes, H. T. (1998). Acta Cryst. B54, 782–789.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationHoward, C. J. & Stokes, H. T. (2002). Acta Cryst. B58, 565–565.  CrossRef CAS IUCr Journals Google Scholar
First citationIvanov, S. A., Eriksson, S.-G., Tellgren, R. & Rundlöf, H. (2001). Mater. Sci. Forum, 378–381, 511–516.  CrossRef ICSD CAS Google Scholar
First citationJena, A. K., Kulkarni, A. & Miyasaka, T. (2019). Chem. Rev. 119, 3036–3103.  CrossRef CAS PubMed Google Scholar
First citationKennedy, B. J., Howard, C. J. & Chakoumakos, B. C. (1999). J. Phys. Condens. Matter, 11, 1479–1488.  Web of Science CrossRef CAS Google Scholar
First citationKennedy, B. J., Howard, C. J., Thorogood, G. J. & Hester, J. R. (2001). J. Solid State Chem. 161, 106–112.  Web of Science CrossRef CAS Google Scholar
First citationKennedy, B. J., Hunter, B. A. & Hester, J. R. (2002). Phys. Rev. B, 65, 224103-1.  CrossRef Google Scholar
First citationKennedy, B. J., Prodjosantoso, A. K. & Howard, C. J. (1999). J. Phys. Condens. Matter, 11, 6319–6327.  Web of Science CrossRef ICSD CAS Google Scholar
First citationKennedy, B. J., Yamaura, K. & Takayama-Muromachi, E. (2004). J. Phys. Chem. Solids, 65, 1065–1069.  CrossRef CAS Google Scholar
First citationKnight, K. S. (2009). Can. Mineral. 47, 381–400.  Web of Science CrossRef ICSD CAS Google Scholar
First citationKnight, K. S. & Kennedy, B. J. (2015). Solid State Sci. 43, 15–21.  CrossRef ICSD CAS Google Scholar
First citationLi, L., Kennedy, B. J., Kubota, Y., Kato, K. & Garrett, R. F. (2004). J. Mater. Chem. 14, 263–273.  Web of Science CrossRef ICSD CAS Google Scholar
First citationLiu, X. & Liebermann, R. C. (1993). Phys. Chem. Miner. 20, 171–175.  CrossRef ICSD CAS Google Scholar
First citationLufaso, M. W. & Woodward, P. M. (2001). Acta Cryst. B57, 725–738.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMagyari-Köpe, B., Vitos, L., Johansson, B. & Kollár, J. (2001). Acta Cryst. B57, 491–496.  Web of Science CrossRef IUCr Journals Google Scholar
First citationMagyari-Köpe, B., Vitos, L., Johansson, B. & Kollár, J. (2002). Phys. Rev. B, 66, 092103–1.  Google Scholar
First citationMegaw, H. D. (1973). Crystal Structures – A Working Approach, pp. 285–302. London: Saunders.  Google Scholar
First citationMitchell, R. H. (2002a). Perovskites: Modern and Ancient, Preface. Thunder Bay, Ontario: Almaz Press.  Google Scholar
First citationMitchell, R. H. (2002b). Perovskites: Modern and Ancient, pp. 27–29. Thunder Bay, Ontario: Almaz Press.  Google Scholar
First citationMitchell, R. H., Burns, P. C., Knight, K. S., Howard, C. J. & Chakhmouradian, A. R. (2014). Phys. Chem. Miner. 41, 393–401.  Web of Science CrossRef ICSD CAS Google Scholar
First citationMitchell, R. H. & Liferovich, R. P. (2004). J. Solid State Chem. 177, 4420–4427.  CrossRef ICSD CAS Google Scholar
First citationMountstevens, E. H., Attfield, J. P. & Redfern, S. A. T. (2003). J. Phys. Condens. Matter, 15, 8315–8326.  Web of Science CrossRef ICSD CAS Google Scholar
First citationMoussa, S. M., Kennedy, B. J. & Vogt, T. (2001). Solid State Commun. 119, 549–552.  Web of Science CrossRef ICSD CAS Google Scholar
First citationMurakami, M., Hirose, K., Kawamura, K., Sata, N. & Ohishi, Y. (2004). Science, 304, 855–858.  Web of Science CrossRef ICSD PubMed CAS Google Scholar
First citationRedfern, S. A. T. (1996). J. Phys. Condens. Matter, 8, 8267–8275.  CrossRef CAS Web of Science Google Scholar
First citationReifenberg, M. & Thomas, N. W. (2018). Acta Cryst. B74, 165–181.  Web of Science CrossRef ICSD IUCr Journals Google Scholar
First citationRoss, N. L., Zhao, J. & Angel, R. J. (2004). J. Solid State Chem. 177, 1276–1284.  Web of Science CrossRef ICSD CAS Google Scholar
First citationSasaki, S., Prewitt, C. T., Bass, J. D. & Schulze, W. A. (1987). Acta Cryst. C43, 1668–1674.  CrossRef ICSD CAS Web of Science IUCr Journals Google Scholar
First citationShannon, R. D. (1976). Acta Cryst. A32, 751–767.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationShrout, T. R. & Zhang, S. J. (2007). J. Electroceram. 19, 113–126.  Web of Science CrossRef Google Scholar
First citationTamazyan, R. & van Smaalen, S. (2007). Acta Cryst. B63, 190–200.  Web of Science CrossRef IUCr Journals Google Scholar
First citationThomas, N. W. (1989). Acta Cryst. B45, 337–344.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationThomas, N. W. (1996). Acta Cryst. B52, 16–31.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationThomas, N. W. (1998). Acta Cryst. B54, 585–599.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationThomas, N. W. (2017). Acta Cryst. B73, 74–86.  Web of Science CrossRef IUCr Journals Google Scholar
First citationVanpeteghem, C. B., Zhao, J., Angel, R. J., Ross, N. L. & Bolfan-Casanova, N. (2006). Geophys. Res. Lett. 33, L03306-1.  CrossRef Google Scholar
First citationWang, D. & Angel, R. J. (2011). Acta Cryst. B67, 302–314.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWilliams, D. E. (1971). Acta Cryst. A27, 452–455.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationWoodward, P. M. (1997a). Acta Cryst. B53, 32–43.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationWoodward, P. M. (1997b). Acta Cryst. B53, 44–66.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationYang, J. (2008). Acta Cryst. B64, 281–286.  CrossRef ICSD IUCr Journals Google Scholar
First citationYashima, M. & Ali, R. (2009). Solid State Ionics, 180, 120–126.  Web of Science CrossRef ICSD CAS Google Scholar
First citationZagorac, J., Zagorac, D., Zarubica, A., Schön, J. C., Djuris, K. & Matovic, B. (2014). Acta Cryst. B70, 809–819.  Web of Science CrossRef ICSD IUCr Journals Google Scholar
First citationZhao, J., Ross, N. L. & Angel, R. J. (2004). Acta Cryst. B60, 263–271.  Web of Science CrossRef CAS IUCr Journals Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Follow Acta Cryst. B
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds