International
Tables for
Crystallography
Volume D
Physical properties of crystals
Edited by A. Authier

International Tables for Crystallography (2006). Vol. D. ch. 3.4, pp. 449-505
https://doi.org/10.1107/97809553602060000645

Chapter 3.4. Domain structures

V. Janoveca* and J. Přívratskáb

a Department of Physics, Technical University of Liberec, Hálkova 6, 461 17 Liberec 1, Czech Republic, and bDepartment of Mathematics and Didactics of Mathematics, Technical University of Liberec, Hálkova 6, 461 17 Liberec 1, Czech Republic
Correspondence e-mail:  janovec@fzu.cz

This chapter is devoted to the crystallographic aspects of static ferroic domain structures. The exposition is based on well defined concepts and rigorous relations that follow from the symmetry lowering at the ferroic phase transition. Necessary mathematical tools are explained in Section 3.2.3[link] and important points are illustrated with simple examples. Synoptic tables provide useful ready-to-use data accessible even without knowledge of deeper theory. Three main concepts needed in a rigorous analysis (both in a continuum and a microscopic description) of any domain structure are thoroughly discussed. (1) Domain states (orientation states or structural variants) representing inner structures of domains are classified according to their characteristic properties (ferroelastic, ferroelectric etc.) and their hierarchy (primary, secondary, principal, basic etc.). A synoptic table is given with all possible symmetry lowerings at ferroic transitions and contains the numbers of ferroic, ferroelectric and ferroelastic domain states, Aizu's classification and the representation characterizing the principal domain states. (2) Relations between domain states (twin laws) determine domain distinction, switching of domain states in external fields and properties of interfaces (domain walls) between coexisting domains. Tables give for each possible transition all independent twin laws and for each twin law the number of equal and distinct tensor components of material tensors up to rank 4 in two coexisting domains of a domain twin. (3) The basic properties of domain twins and domain walls are determined by their symmetry, which is expressed by crystallographic layer groups. The fundamental significance of this description is explained and illustrated. Synoptic tables give for each twin law the possible orientations of compatible domain walls and their symmetries.

Keywords: Aizu classification; Dauphiné twins; coherent domain walls; dichromatic complexes; domain pairs; domain states; domain structures; domain twins; domain walls; ferroelastic domain pairs; ferroelastic domain states; ferroelastic domain structures; ferroelastic domain twins; ferroelastic domain walls; ferroelastic single-domain states; ferroelectric domain states; ferroelectric domain structures; ferroic domain states; ferroic domain structures; ferroic transitions; layer groups; morphic tensor components; non-ferroelastic domain pairs; non-ferroelastic domain states; non-ferroelastic domain structures; non-ferroelastic domain twins; non-ferroelastic domain walls; non-ferroelastic phases; non-ferroelectric domain states; non-ferroelectric phases; parent clamping approximation; physical property tensors; stabilizers; switching; symmetry descent; twin laws; twinning group.

3.4.1. Introduction

| top | pdf |

3.4.1.1. Basic concepts

| top | pdf |

It was demonstrated in Section 3.1.2[link] that a characteristic feature of structural phase transitions connected with a lowering of crystal symmetry is an anomalous behaviour near the transition, namely unusually large values of certain physical properties that vary strongly with temperature. In this chapter, we shall deal with another fundamental feature of structural phase transitions: the formation of a non-homogeneous, textured low-symmetry phase called a domain structure.

When a crystal homogeneous in the parent (prototypic) phase undergoes a phase transition into a ferroic phase with lower point-group symmetry, then this ferroic phase is almost always formed as a non-homogeneous structure consisting of homogeneous regions called domains and contact regions between domains called domain walls. All domains have the same or the enantiomorphous crystal structure of the ferroic phase, but this structure has in different domains a different orientation, and sometimes also a different position in space. When a domain structure is observed by a measuring instrument, different domains can exhibit different tensor properties, different diffraction patterns and can differ in other physical properties. The domain structure can be visualized optically (see Fig. 3.4.1.1[link]) or by other experimental techniques. Powerful high-resolution electron microscopy (HREM) techniques have made it possible to visualize atomic arrangements in domain structures (see Fig. 3.4.1.2[link]). The appearance of a domain structure, detected by any reliable technique, provides the simplest unambiguous experimental proof of a structural phase transition.

[Figure 3.4.1.1]

Figure 3.4.1.1 | top | pdf |

Domain structure of tetragonal barium titanate (BaTiO3). A thin section of barium titanate ceramic observed at room temperature in a polarized-light microscope (transmitted light, crossed polarizers). Courtesy of U. Täffner, Max-Planck-Institut für Metallforschung, Stuttgart. Different colours correspond to different ferroelastic domain states, connected areas of the same colour are ferroelastic domains and sharp boundaries between these areas are domain walls. Areas of continuously changing colour correspond to gradually changing thickness of wedge-shaped domains. An average distance between parallel ferroelastic domain walls is of the order of 1–10 µm.

[Figure 3.4.1.2]

Figure 3.4.1.2 | top | pdf |

Domain structure of a BaGa2O4 crystal seen by high-resolution transmission electron microscopy. Parallel rows are atomic layers. Different directions correspond to different ferroelastic domain states of domains, connected areas with parallel layers are different ferroelastic domains and boundaries between these areas are ferroelastic domain walls. Courtesy of H. Lemmens, EMAT, University of Antwerp.

Under the influence of external fields (mechanical stress, electric or magnetic fields, or combinations thereof), the domain structure can change; usually some domains grow while others decrease in size or eventually vanish. This process is called domain switching. After removing or decreasing the field a domain structure might not change considerably, i.e. the form of a domain pattern depends upon the field history: the domain structure exhibits hysteresis (see Fig. 3.4.1.3[link]). In large enough fields, switching results in a reduction of the number of domains. Such a procedure is called detwinning. In rare cases, the crystal may consist of one domain only. Then we speak of a single-domain crystal.

[Figure 3.4.1.3]

Figure 3.4.1.3 | top | pdf |

Elastic hysteresis of ferroelastic lead phosphate Pb3(PO4)2 (Salje, 1990[link]). Courtesy of E. K. H. Salje, University of Cambridge. The dependence of strain on applied stress has the form of a loop. The states at the extreme left and right correspond to two ferroelastic domain states, steep parts of the loop represent switching of one state into the other by applied stress. The strain at zero stress corresponds to the last single-domain state formed in a field larger than the coercive stress defined by the stress at zero strain (the intersection of the loop with the axis of the applied stress). Similar dielectric hysteresis loops of polarization versus applied electric field are observed in ferroelectric phases (see e.g. Jona & Shirane, 1962[link]).

There are two basic types of domain structures:

  • (i) Domain structures with one or several systems of parallel plane domain walls that can be observed in an optical or electron microscope. Two systems of perpendicular domain walls are often visible (see Fig. 3.4.1.4[link]). In polarized light domains exhibit different colours (see Fig. 3.4.1.1[link]) and in diffraction experiments splitting of reflections can be observed (see Fig. 3.4.3.9[link]). Domains can be switched by external mechanical stress. These features are typical for a ferroelastic domain structure in which neighbouring domains differ in mechanical strain (deformation). Ferroelastic domain structures can appear only in ferroelastic phases, i.e. as a result of a phase transition characterized by a spontaneous shear distortion of the crystal.

    [Figure 3.4.1.4]

    Figure 3.4.1.4 | top | pdf |

    Transmission electron microscopy image of the ferroelastic domain structure in a YBa2Cu3O7−y crystal (Rosová, 1999[link]). Courtesy of A. Rosová, Institute of Electrical Engineering, SAS, Bratislava. There are two systems (`complexes'), each of which is formed by almost parallel ferroelastic domain walls with needle-like tips. The domain walls in one complex are nearly perpendicular to the domain walls in the other complex.

  • (ii) Domain structures that are not visible using a polarized-light microscope and in whose diffraction patterns no splitting of reflections is observed. Special methods [e.g. etching, deposition of liquid crystals (see Fig. 3.4.1.5[link]), electron or atomic force microscopy, or higher-rank optical effects (see Fig. 3.4.3.3[link])] are needed to visualize domains. Domains have the same strain and cannot usually be switched by an external mechanical stress. Such domain structures are called non-ferroelastic domain structures. They appear in all non-ferroelastic phases resulting from symmetry lowering that preserves the crystal family, and in partially ferroelastic phases.

    [Figure 3.4.1.5]

    Figure 3.4.1.5 | top | pdf |

    Non-ferroelastic ferroelectric domains in triglycine sulfate (TGS) revealed by a liquid-crystal method. A thin layer of a nematic liquid crystal deposited on a crystal surface perpendicular to the spontaneous polarization is observed in a polarized-light microscope. Black and white areas correspond to ferroelectric domains with antiparallel spontaneous polarization. The typical size of the domains is of order of 1–10µm. Courtesy of M. Połomska, Institute of Molecular Physics, PAN, Poznań. Although one preferential direction of domain walls prevails, the rounded shapes of the domains indicate that all orientations of non-ferroelastic walls are possible.

Another important kind of domain structure is a ferroelectric domain structure, in which domains differ in the direction of the spontaneous polarization. Such a domain structure is formed at ferroelectric phase transitions that are characterized by the appearance of a new polar direction in the ferroic phase. Ferroelectric domains can usually be switched by external electric fields. Two ferroelectric domains with different directions of spontaneous polarization can have different spontaneous strain [e.g. in dihydrogen phosphate (KDP) crystals, two ferroelectric domains with opposite directions of the spontaneous polarization have different spontaneous shear strain], or two ferroelectric domains with antiparallel spontaneous polarization can possess the same strain [e.g. in triglycine sulfate (TGS) crystals].

The physical properties of polydomain crystals are significantly influenced by their domain structure. The values of important material property tensor components, e.g. permittivity, piezoelectric and elastic constants, may be enhanced or diminished by the presence of a domain structure. Owing to switching and detwinning phenomena, polydomain materials exhibit hysteresis of material properties. These features have important practical implications, e.g. the production of anisotropic ceramic materials or ferroelectric memories.

The domain structure resulting from a structural phase transition belongs to a special type of twinning referred to as transformation twinning (see Section 3.3.7.2[link] ). Despite this, the current terminology used in domain-structure studies is different. The main terms were coined during the first investigations of ferroelectric materials, where striking similarities with the behaviour of ferromagnetic materials led researchers to introduce terms analogous to those used in studies of ferromagnetic domain structures that had been examined well at that time.

Bicrystallography (see Section 3.2.2[link] ) provides another possible frame for discussing domain structures. Bicrystallography and domain structure analysis have developed independently and almost simultaneously but different language has again precluded deeper confrontation. Nevertheless, there are common features in the methodology of both approaches, in particular, the principle of symmetry compensation (see Section 3.2.2[link] ), which plays a fundamental role in both theories.

In Chapter 3.1[link] , it is shown that the anomalous behaviour near phase transitions can be explained in the framework of the Landau theory. In this theory, the formation of the domain structures follows from the existence of several equivalent solutions for the order parameter. This result is a direct consequence of a symmetry reduction at a ferroic phase transition. It is this dissymmetrization which is the genuine origin of the domain structure formation and which determines the basic static features of all domain structures.

3.4.1.2. Scope of this chapter

| top | pdf |

This chapter is devoted to the crystallographic aspects of static domain structures, especially to the symmetry analysis of these structures. The main aim is to explain basic concepts, derive relations that govern the formation of domain structures and provide tables with useful ready-to-use data on domain structures of ferroic phases. The exposition uses algebraic tools that are explained in Section 3.2.3[link] , but the important points are illustrated with simple examples comprehensible even without mathematical details. The synoptic tables in Sections 3.4.2[link] and 3.4.3[link] present the main results of the analysis for all possible ferroic domain structures. More detailed information on certain points can be found in the software GI[\star]KoBo-1.

All these results are definite – their validity does not depend on any particular model or approximation – and form thus a firm basis for further more detailed quantitative treatments. `For the most part, the only exact statements which can be made about a solid state system are those which arise as a direct consequence of symmetry alone.' (Knox & Gold, 1967[link].)

The exposition starts with domain states, continues with pairs of domain states and domain distinction, and terminates with domain twins and walls. This is also the sequence of steps in domain-structure analysis, which proceeds from the simplest to more complicated objects.

In Section 3.4.2[link], we explain the concept of domain states (also called variants or orientational states), define different types of domain states (principal, ferroelastic, ferroelectric, basic), find simple formulae for their number, and disclose their hierarchy and relation with symmetry lowering and with order parameters of the transition. Particular results for all possible ferroic phase transitions can be found in synoptic Table 3.4.2.7[link], which lists all possible crystallographically non-equivalent point-group symmetry descents that may appear at a ferroic phase transition. For each descent, all independent twinning groups (characterizing the relation between two domain states) are given together with the number of principal, ferroelastic and ferroelectric domain states and other data needed in further analysis.

Section 3.4.3[link] deals with pairs of domain states and with the relationship between two domain states in a pair. This relationship, in mineralogy called a `twin law', determines the distinction between domain states, specifies switching processes between two domain states and forms a starting point for discussing domain walls and twins. We show different ways of expressing the relation between two domain states of a domain pair, derive a classification of domain pairs, find non-equivalent domain pairs and determine which tensor properties are different and which are the same in two domain states of a domain pair.

The presentation of non-equivalent domain pairs is divided into two parts. Synoptic Table 3.4.3.4[link] lists all representative non-equivalent non-ferroelastic domain pairs, and for each pair gives the twinning groups, and the number of tensor components that are different and that are the same in two domain states. These numbers are given for all important property tensors up to rank four. We also show how these data can be used to determine switching forces between two non-ferroelastic domain states.

Then we explain specific features of ferroelastic domain pairs: compatible (permissible) domain walls and disorientation of domain states in ferroelastic domain twins. A list of all non-equivalent ferroelastic domain pairs is presented in two tables. Synoptic Table 3.4.3.6[link] contains all non-equivalent ferroelastic domain pairs with compatible (coherent) domain walls. This table gives the orientation of compatible walls and their symmetry properties. Table 3.4.3.7[link] lists all non-equivalent ferroelastic domain pairs with no compatible ferroelastic domain walls.

Column K1j in Table 3.4.2.7[link] specifies all representative non-equivalent domain pairs that can appear in each particular phase transition; in combination with Tables 3.4.3.4[link] and 3.4.3.6[link], it allows one to determine the main features of any ferroic domain structure.

Section 3.4.4[link] is devoted to domain twins and domain walls. We demonstrate that the symmetry of domain twins and domain walls is described by layer groups, give a classification of domain twins and walls based on their symmetry, and present possible layer groups of non-ferroelastic and ferroelastic domain twins and walls. Then we discuss the properties of finite-thickness domain walls. In an example, we illustrate the symmetry analysis of microscopic domain walls and present conclusions that can be drawn from this analysis about the microscopic structure of domain walls.

The exposition is given in the continuum description with crystallographic point groups and property tensors. In this approach, all possible cases are often treatable and where possible are covered in synoptic tables or – in a more detailed form – in the software GI[\star]KoBo-1. Although the group-theoretical tools are almost readily transferable to the microscopic description (using the space groups and atomic positions), the treatment of an inexhaustible variety of microscopic situations can only be illustrated by particular examples.

Our attempt to work with well defined notions calls for introducing several new, and generalizing some accepted, concepts. Also an extended notation for the symmetry operations and groups has turned out to be indispensable. Since there is no generally accepted terminology on domain structures yet, we often have to choose a term from several existing more-or-less equivalent variants.

The specialized scope of this chapter does not cover several important aspects of domain structures. More information can be found in the following references. There are only two monographs on domain structures (both in Russian): Fesenko et al. (1990[link]) and Sidorkin (2002[link]). The main concepts of domain structures of ferroic materials are explained in the book by Wadhawan (2000[link]) and in a review by Schranz (1995[link]). Ferroelastic domain structures are reviewed in Boulesteix (1984[link]) and Wadhawan (1991[link]), and are treated in detail by Salje (1990[link], 1991[link], 2000a[link],b[link]). Different aspects of ferroelectric domain structures are covered in books or reviews on ferroelectric crystals: Känzig (1957[link]), Jona & Shirane (1962[link]), Fatuzzo & Merz (1967[link]), Mitsui et al. (1976[link]), Lines & Glass (1977[link]), Smolenskii et al. (1984[link]), Zheludev (1988[link]) and Strukov & Levanyuk (1998[link]). Applications of ferroelectrics are described in the books by Xu (1991[link]) and Uchino (2000[link]). Principles and technical aspects of ferroelectric memories are reviewed by Scott (1998[link], 2000[link]).

3.4.2. Domain states

| top | pdf |

3.4.2.1. Principal and basic domain states

| top | pdf |

As for all crystalline materials, domain structures can be approached in two ways: In the microscopic description, a crystal is treated as a regular arrangement of atoms. Domains differ in tiny differences of atomic positions which can be determined only indirectly, e.g. by diffraction techniques. In what follows, we shall pay main attention to the continuum description, in which a crystal is treated as an anisotropic continuum. Then the crystal properties are described by property tensors (see Section 1.1.1[link] ) and the crystal symmetry is expressed by crystallographic point groups. In this approach, domains exhibit different tensor properties that enable one to visualize domains by optical or other methods.

The domain structure observed in a microscope appears to be a patchwork of homogeneous regions – domains – that have various colours and shapes (see Fig. 3.4.1.1[link]). Indeed, the usual description considers a domain structure as a collection of domains and contact regions of domains called domain walls. Strictly speaking, by a domain [{\bf D}_i] one understands a connected part of the crystal, called the domain region, which is filled with a homogeneous low-symmetry crystal structure. Domain walls can be associated with the boundaries of domain regions. The interior homogeneous bulk structure within a domain region will be called a domain state. Equivalent terms are variant or structural variant (Van Tendeloo & Amelinckx, 1974[link]). We shall use different adjectives to specify domain states. In the microscopic description, domain states associated with the primary order parameter will be referred to as primary (microscopic, basic) domain states. In the macroscopic description, the primary domain states will be called principal domain states, which correspond to Aizu's orientation states. (An exact definition of principal domain states is given below.)

Further useful division of domain states is possible (though not generally accepted): Domain states that are specified by a constant value of the spontaneous strain are called ferroelastic domain states; similarly, ferroelectric domain states exhibit constant spontaneous polarization etc. Domain states that differ in some tensor properties are called ferroic or tensorial domain states etc. If no specification is given, the statements will apply to any of these domain states.

A domain [{\bf D}_i] is specified by a domain state [{\bf S}_j ] and by domain region [B_k]: [{\bf D}_i =] [{\bf D}_i({\bf S}_j,B_k)]. Different domains may possess the same domain state but always differ in the domain region that specifies their shape and position in space.

The term `domain' has also often been used for a domain state. Clear distinction of these two notions is essential in further considerations and is illustrated in Fig. 3.4.2.1[link]. A ferroelectric domain structure (Fig. 3.4.2.1a[link]) consists of six ferroelectric domains [{\bf D}_1], [{\bf D}_2], [\ldots], [{\bf D}_6] but contains only two domain states [{\bf S}_1], [{\bf S}_2] characterized by opposite directions of the spontaneous polarization depicted in Fig. 3.4.2.1(d[link]). Neighbouring domains have different domain states but non-neighbouring domains may possess the same domain state. Thus domains with odd serial number have the domain state [{\bf S}_1] (spontaneous polarization `down'), whereas domains with even number have domain state [{\bf S}_2] (spontaneous polarization `up').

[Figure 3.4.2.1]

Figure 3.4.2.1 | top | pdf |

Hierarchy in domain-structure analysis. (a) Domain structure consisting of domains [{\bf D}_1, {\bf D}_2,\ldots, {\bf D}_6] and domain walls [{\bf W}_{12}] and [{\bf W}_{21}]; (b) domain twin and reversed twin (with reversed order of domain states); (c) domain pair consisting of two domain states [{\bf S}_1] and [{\bf S}_2]; (d) domain states [{\bf S}_1] and [{\bf S}_2 ].

A great diversity of observed domain structures are connected mainly with various dimensions and shapes of domain regions, whose shapes depend sensitively on many factors (kinetics of the phase transition, local stresses, defects etc.). It is, therefore, usually very difficult to interpret in detail a particular observed domain pattern. Domain states of domains are, on the other hand, governed by simple laws, as we shall now demonstrate.

We shall consider a ferroic phase transition with a symmetry lowering from a parent (prototypic, high-symmetry) phase with symmetry described by a point group G to a ferroic phase with the point-group symmetry [F_1 ], which is a subgroup of G. We shall denote this dissymmetrization by a group–subgroup symbol [G \supset F_1] (or [G\Downarrow F_1 ] in Section 3.1.3[link] ) and call it a symmetry descent or dissymmetrization. Aizu (1970a[link]) calls these symmetry descents species and uses the letter F instead of the symbol [\subset].

As an illustrative example, we choose a phase transition with parent symmetry [G =4_z/m_zm_xm_{xy}] and ferroic symmetry [F_1=2_xm_ym_z] (see Fig. 3.4.2.2[link]). Strontium bismuth tantalate (SBT) crystals, for instance, exhibit a phase transition with this symmetry descent (Chen et al., 2000[link]). Symmetry elements in the symbols of G and [F_1] are supplied with subscripts specifying the orientation of the symmetry elements with respect to the reference coordinate system. The necessity of this extended notation is exemplified by the fact that the group [G = 4_z/m_zm_xm_{xy}] has six subgroups with the same `non-oriented' symbol [mm2]: [m_xm_y2_z], [2_xm_ym_z], [m_x2_ym_z ], [m_{x{\bar y}}m_{xy}2_z], [2_{x{\bar y}}m_{xy}m_z], [m_{x{\bar y}}2_{xy}m_z]. Lower indices thus specify these subgroups unequivocally and the example illustrates an important rule of domain-structure analysis: All symmetry operations, groups and tensor components must be related to a common reference coordinate system and their orientation in space must be clearly specified.

[Figure 3.4.2.2]

Figure 3.4.2.2 | top | pdf |

Exploded view of single-domain states [{\bf S}_1], [{\bf S}_2], [{\bf S}_3] and [{\bf S}_4 ] (solid rectangles with arrows of spontaneous polarization) formed at a phase transition from a parent phase with symmetry [G=4_z/m_zm_xm_{xy} ] to a ferroic phase with symmetry [F_1=2_xm_ym_z]. The parent phase is represented by a dashed square in the centre with the symmetry elements of the parent group [G=4_z/m_zm_xm_{xy}] shown.

The physical properties of crystals in the continuum description are expressed by property tensors. As explained in Section 1.1.4[link] , the crystal symmetry reduces the number of independent components of these tensors. Consequently, for each property tensor the number of independent components in the low-symmetry ferroic phase is the same or higher than in the high-symmetry parent phase. Those tensor components or their linear combinations that are zero in the high-symmetry phase and nonzero in the low-symmetry phase are called morphic tensor components or tensor parameters and the quantities that appear only in the low-symmetry phase are called spontaneous quantities (see Section 3.1.3.2[link] ). The morphic tensor components and spontaneous quantities thus reveal the difference between the high- and low-symmetry phases. In our example, the symmetry [F_1 =2_xm_ym_z ] allows a nonzero spontaneous polarization [{\rm P}^{(1)}_0] [=(P,0,0) ], which must be zero in the high-symmetry phase with [G=4_z/m_zm_xm_{xy} ].

We shall now demonstrate in our example that the symmetry lowering at the phase transition leads to the existence of several equivalent variants (domain states) of the low-symmetry phase. In Fig. 3.4.2.2[link], the parent high-symmetry phase is represented in the middle by a dashed square that is a projection of a square prism with symmetry [4_z/m_zm_xm_{xy} ]. A possible variant of the low-symmetry phase can be represented by an oblong prism with a vector representing the spontaneous polarization. In Fig. 3.4.2.2[link], the projection of this oblong prism is drawn as a rectangle which is shifted out of the centre for better recognition. We denote by [{\bf S}_1] a homogeneous low-symmetry phase with spontaneous polarization [{\rm P}^{(1)}_0=(P,0,0)] and with symmetry F1 = [2_xm_ym_z]. Let us, mentally, increase the temperature to above the transition temperature and then apply to the high-symmetry phase an operation [2_z], which is a symmetry operation of this high-symmetry phase but not of the low-symmetry phase. Then decrease the temperature to below the transition temperature. The appearance of another variant of the low-symmetry phase [{\bf S}_2] with spontaneous polarization [{\rm P}^{(2)}_0=(-P,0,0) ] obviously has the same probability of appearing as had the variant [{\bf S}_1]. Thus the two variants of the low-symmetry phase [{\bf S}_1 ] and [{\bf S}_2] can appear with the same probability if they are related by a symmetry operation suppressed (lost) at the transition, i.e. an operation that was a symmetry operation of the high-symmetry phase but is not a symmetry operation of the low-symmetry phase [{\bf S}_1 ]. In the same way, the lost symmetry operations [4_z] and [4^3_z] generate from [{\bf S}_1] two other variants, [{\bf S}_3 ] and [{\bf S}_4], with spontaneous polarizations [(0,P,0)] and [(0,-P,0)], respectively. Variants of the low-symmetry phase that are related by an operation of the high-symmetry group G are called crystallographically equivalent (in G) variants. Thus we conclude that crystallographically equivalent (in G) variants of the low-symmetry phase have the same chance of appearing.

We shall now make similar considerations for a general ferroic phase transition with a symmetry descent [G\supset F_1]. By the state S of a crystal we shall understand, in the continuum description, the set of all its properties expressed by property (matter) tensors in the reference Cartesian crystallophysical coordinate system of the parent phase (see Example 3.2.3.9[link] in Section 3.2.3.3.1[link] ). A state defined in this way may change not only with temperature and external fields but also with the orientation of the crystal in space.

We denote by [{\bf S}_1] a state of a homogeneous ferroic phase. If we apply to [{\bf S}_1] a symmetry operation [g_i] of the group G, then the ferroic phase in a new orientation will have the state [{\bf S}_j], which may be identical with [{\bf S}_1] or different. Using the concept of group action (explained in detail in Section 3.2.3.3.1[link] ) we express this operation by a simple relation: [g_j{\bf S}_1 = {\bf S}_j, \quad g_j \in G. \eqno(3.4.2.1) ]

Let us first turn our attention to operations [f_j \in G] that do not change the state [{\bf S}_1]: [f_j{\bf S}_1 = {\bf S}_j,\quad f_j \in G. \eqno(3.4.2.2) ]The set of all operations of G that leave [{\bf S}_1] invariant form a group called a stabilizer (or isotropy group) of a state [{\bf S}_1] in the group G. This stabilizer, denoted by [I_G({\bf S}_1)], can be expressed explicitly in the following way: [I_G({\bf S}_1) \equiv \{g \in G|g{\bf S}_1 = {\bf S}_1\}, \eqno(3.4.2.3) ]where the right-hand part of the equation should be read as `a set of all operations of G that do not change the state [{\bf S}_1]' (see Section 3.2.3.3.2[link] ).

Here we have to explain the difference between the concept of a stabilizer of an object and the symmetry of that object. By the symmetry group F of an object one understands the set of all operations (isometries) that leave this object [{\bf S}] invariant. The symmetry group F of an object is considered to be an inherent property that does not depend on the orientation and position of the object in space. (The term eigensymmetry is used in Chapter 3.3[link] for symmetry groups defined in this way.) In this case, the symmetry elements of F are `attached' to the object.

A stabilizer describes the symmetry properties of an object in another way, in which the object and the group of isometries are decoupled. One is given a group G, the symmetry elements of which have a defined orientation in a fixed reference system. The object can have any orientation in this reference system. Those operations of G that map the object in a given orientation onto itself form the stabilizer [I_G({\bf S}_1)] of [{\bf S}_i] in the group G. In this case, the stabilizer depends on the orientation of the object in space and is expressed by an `oriented' group symbol [F_1] with subscripts defining the orientation of the symmetry elements of [F_1]. Only for certain `prominent' orientations will the stabilizer acquire a symmetry group of the same crystal class (crystallographic point group) as the eigensymmetry of the object.

We shall define a single-domain orientation as a prominent orientation of the crystal in which the stabilizer [I_G({\bf S}_1)] of its state [{\bf S}_1] is equal to the symmetry group [F_1] which is, after removing subscripts specifying the orientation, identical with the eigensymmetry of the ferroic phase: [I_G({\bf S}_1) = F_1. \eqno(3.4.2.4) ]This equation thus declares that the crystal in the state [{\bf S}_1 ] has a prominent single-domain orientation.

The concept of the stabilizer allows us to identify the `eigensymmetry' of a domain state (or an object in general) [{\bf S}_i] with the crystallographic class (non-oriented point group) of the stabilizer of this state in the group of all rotations O(3), [I_{O(3)}({\bf S}_i) ].

Since we shall further deal mainly with states of the ferroic phase in single-domain orientations, we shall use the term `state' for a `state of the crystal in a single-domain orientation', unless mentioned otherwise. Then the stabilizer [I_G({\bf S}_1)] will often be replaced by the group [F_1], although all statements have been derived and hold for stabilizers.

The difference between symmetry groups of a crystal and stabilizers will become more obvious in the treatment of secondary domain states in Section 3.4.2.2[link] and in discussing disoriented ferroelastic domain states (see Section 3.4.3.6.3[link]).

As we have seen in our illustrative example, the suppressed operations generate from the first state [{\bf S}_1] other states. Let [g_j ] be such a suppressed operation, i.e. [g_j \in G ] but [g_j \not\in F_1]. Since all operations that retain [{\bf S}_1] are collected in [F_1], the operation [g_j] must transform [{\bf S}_1] into another state [{\bf S}_j], [g_j{\bf S}_1 = {\bf S}_j \not= {\bf S}_1,\quad g_j\in G,\quad g_j \not\in F_1, \eqno(3.4.2.5) ]and we say that the state [{\bf S}_j] is crystallographically equivalent (in G) with the state [{\bf S}_1], [{\bf S}_j \buildrel {G} \over \sim {\bf S}_1 ].

We define principal domain states as crystallographically equivalent (in G) variants of the low-symmetry phase in single-domain orientations that can appear with the same probability in the ferroic phase. They represent possible macroscopic bulk structures of (1) ferroic single-domain crystals, (2) ferroic domains in non-ferroelastic domain structures (see Section 3.4.3.5[link]), or (3) ferroic domains in any ferroic domain structure, if all spontaneous strains are suppressed [this is the so-called parent clamping approximation (PCA), see Section 3.4.2.5[link]]. In what follows, any statement formulated for principal domain states or for single-domain states applies to any of these three situations. Principal domain states are identical with orientation states (Aizu, 1969[link]) or orientation variants (Van Tendeloo & Amelinckx, 1974[link]). The adjective `principal' distinguishes these domain states from primary (microscopic, basic – see Section 3.4.2.5[link]) domain states and secondary domain states, defined in Section 3.4.2.2[link], and implies that any two of these domain states differ in principal tensor parameters (these are linear combinations of morphic tensor components that transform as the primary order parameter of an equitranslational phase transition with a point-group symmetry descent [G\supset F_1], see Sections 3.1.3.2[link] and 3.4.2.3[link]). A simple criterion for a principal domain state [{\bf S}_1] is that its stabilizer in G is equal to the symmetry [F_1] of the ferroic phase [see equation (3.4.2.4[link])].

When one applies to a principal domain state [{\bf S}_1] all operations of the group G, one gets all principal domain states that are crystallographically equivalent with [{\bf S}_1]. The set of all these states is denoted [G{\bf S}_1] and is called an G-orbit of [{\bf S}_1] (see also Section 3.2.3.3.3[link] ), [G{\bf S}_1 = \{{\bf S}_1, {\bf S}_2,\ldots,{\bf S}_n\}. \eqno(3.4.2.6) ]In our example, the G-orbit is [4_z/m_zm_xm_{xy}{\bf S}_1=\{{\bf S}_1,{\bf S}_2,{\bf S}_3,{\bf S}_4\} ].

Note that any operation g from the parent group G leaves the orbit [G{\bf S}_1] invariant since its action results only in a permutation of all principal domain states. This change does not alter the orbit, since the orbit is a set in which the sequence (order) of objects is irrelevant. Therefore, the orbit [G{\bf S}_1] is invariant under the action of the parent group G, [GG{\bf S}_1 = G{\bf S}_1].

A ferroic phase transition is thus a paradigmatic example of the law of symmetry compensation (see Section 3.2.2[link] ): The dissymmetrization of a high-symmetry parent phase into a low-symmetry ferroic phase produces variants of the low-symmetry ferroic phase (single-domain states). Any two single-domain states are related by some suppressed operations of the parent symmetry that are missing in the ferroic symmetry and the set of all single-domain states (G-orbit of domain states) recovers the symmetry of the parent phase. If the domain structure contains all domain states with equal partial volumes then the average symmetry of this polydomain structure is, in the first approximation, identical to the symmetry of the parent phase.

Now we find a simple formula for the number n of principal domain states in the orbit [G{\bf S}_1] and a recipe for an efficient generation of all principal domain states in this orbit.

The fact that all operations of the group [I_G({\bf S}_1)=F_1] leave [{\bf S}_1] invariant can be expressed in an abbreviated form in the following way [see equation (3.2.3.70[link] )]: [F_1{\bf S}_1 = {\bf S}_1. \eqno(3.4.2.7)]We shall use this relation to derive all operations that transform [{\bf S}_1] into [{\bf S}_j = g_j{\bf S}_1]: [g_j{\bf S}_1 = g_j(F_1{\bf S}_1) = (g_jF_1){\bf S}_1 = {\bf S}_j, \quad g_j\in G. \eqno(3.4.2.8) ]The second part of equation (3.4.2.8[link]) shows that all lost operations that transform [{\bf S}_1] into [{\bf S}_j] are contained in the left coset [g_jF_1] (for left cosets see Section 3.2.3.2.3[link] ).

It is shown in group theory that two left cosets have no operation in common. Therefore, another left coset [g_kF_1] generates another principal domain state [{\bf S}_k] that is different from principal domain states [{\bf S}_1] and [{\bf S}_j]. Equation (3.4.2.8)[link] defines, therefore, a one-to-one relation between principal domain states of the orbit [G{\bf S}_1] and left cosets of [F_1] [see equation (3.2.3.69[link] )], [{\bf S}_j \leftrightarrow g_{j}F_1, \quad F_1= I_G({\bf S}_1), \quad j = 1,2,\ldots, n. \eqno(3.4.2.9) ]From this relation follow two conclusions:

  • (1) The number n of principal domain states equals the number of left cosets of [F_1]. All different left cosets of [F_1] constitute the decomposition of the group G into left cosets of [F_1] [see equation (3.2.3.19[link] )], [G = g_1F_{1} \cup g_2F_{1} \cup\ldots\cup g_jF_{1}\cup\ldots \cup g_nF_{1}, \eqno(3.4.2.10) ]where the symbol [\cup] is a union of sets and the number n of left cosets is called the index of G in [F_1] and is denoted by the symbol [[G:F_1]]. Usually, one chooses for [g_1 ] the identity operation e; then the first left coset equals [F_1]. Since each left coset contains [|F_1|] operations, where [|F_1|] is number of operations of [F_1] (order of [F_1]), the number of left cosets in the decomposition (3.4.2.10[link]) is [n=[G:F_1]=|G|:|F_1|, \eqno(3.4.2.11)]where [|G|, |F_1|] are orders of the point groups [G, F_1], respectively. The index n is a quantitative measure of the degree of dissymmetrization [G \supset F_1]. Thus the number of principal domain states in orbit [G{\bf S}_1] is equal to the index of [F_1] in G, i.e. to the number of operations of the high-symmetry group G divided by the number of operations of the low-symmetry phase [F_1]. In our illustrative example we get [n =|4_z/m_zm_xm_{xy}|:|2_xm_ym_z|= 16:4= 4].

    The basic formula (3.4.2.11[link]) expresses a remarkable result: the number n of principal domain states is determined by how many times the number of symmetry operations increases at the transition from the low-symmetry group [F_1] to the high-symmetry group G, or, the other way around, the fraction [{{1}\over{n}}] is a quantitative measure of the symmetry decrease from G to [F_1], [|F_1| = {{1}\over{n}}|G|]. Thus it is not the concrete structural change, nor even the particular symmetries of both phases, but only the extent of dissymmetrization that determines the number of principal domain states. This conclusion illustrates the fundamental role of symmetry in domain structures.

  • (2) Relation (3.4.2.9[link]) yields a recipe for calculating all principal domain states of the orbit [G{\bf S}_1]: One applies successively to the first principal domain states [{\bf S}_1] the representatives of all left cosets of [F_1]: [G{\bf S}_1 = \{{\bf S}_1, g_2{\bf S}_1,\ldots, g_j{\bf S}_1,\ldots, g_n{\bf S}_1\}, \eqno(3.4.2.12) ]where the operations [g_1 = e,g_2,\ldots,g_j,\ldots,g_n] are the representatives of left cosets in the decomposition (3.4.2.10[link]) and e is an identity operation. We add that any operation of a left coset can be chosen as its representative, hence the operation [g_j] can be chosen arbitrarily from the left coset [g_jF_1], [j=1,2,\ldots,n ].

This result can be illustrated in our example. Table 3.4.2.1[link] presents in the first column the four left cosets [g_j\{2_xm_ym_z\}] of the group [F_1 = 2_xm_ym_z]. The corresponding principal domain states [{\bf S}_j], [j = 1,2,3,4,] and the values of spontaneous polarization in these principal domain states are given in the second and the third columns, respectively. It is easy to verify in Fig. 3.4.2.2[link] that all operations of each left coset transform the first principal domain state [{\bf S}_1] into one principal domain state [{\bf S}_j], [j = 2, 3, 4.]

Table 3.4.2.1 | top | pdf |
Left and double cosets, principal and secondary domain states and their tensor parameters for the phase transition with [G=4_z/m_zm_xm_{xy} ] and [F_1=2_xm_ym_z]

Left cosets [g_j{\bf S}_1] Principal domain states Secondary domain states
[1] [2_x] [m_y] [m_z] [{\bf S}_1] [(P00) ] [(000g00) ] [{\bf R}_1] [u_{1}-u_{2}] [Q_{11}-Q_{22}]
[\bar{1} ] [m_x] [2_y] [2_z] [{\bf S}_2 ] [(-P00) ] [(000{-g}00) ]
[2_{xy} ] [4_z ] [\bar{4}^3_{z} ] [m_{x\bar{y}}] [{\bf S}_3 ] [(0P0) ] [(0000{-g}0) ] [{\bf R}_2 ] [u_{2}-u_{1} ] [Q_{22}-Q_{11} ]
[2_{x\bar{y}} ] [4^3_z] [\bar{4}_{z}] [m_{xy}] [{\bf S}_4 ] [(0{-P}0) ] [(0000g0)]

The left coset decompositions of all crystallographic point groups and their subgroup symmetry are available in the software GI[\star ]KoBo-1, path: Subgroups\View\Twinning Group.

Let us turn briefly to the symmetries of the principal domain states. From Fig. 3.4.2.2[link] we deduce that two domain states [{\bf S}_1 ] and [{\bf S}_2] in our illustrative example have the same symmetry, [F_1 =] [F_2 =2_xm_ym_z], whereas two others [{\bf S}_3] and [{\bf S}_4] have another symmetry, [F_3 =F_4 =m_x2_ym_z]. We see that symmetry does not specify the principal domain state in a unique way, although a principal domain state [{\bf S}_j] has a unique symmetry [F_i =I_G({\bf S}_j)].

It turns out that if [g_j] transforms [{\bf S}_1] into [{\bf S}_j], then the symmetry group [F_j] of [{\bf S}_j] is conjugate by [g_j] to the symmetry group [F_1] of [{\bf S}_1 ] [see Section 3.2.3.3[link] , Proposition 3.2.3.13[link] and equation (3.2.3.55[link] )]: [\hbox {if }{\bf S}_j = g_j{\bf S}_1, \hbox { then } F_j = g_jF_1g_j^{-1}. \eqno(3.4.2.13) ]One can easily check that in our example each operation of the second left coset of [F_1=2_xm_ym_z] (second row in Table 3.4.2.1[link]) transforms [F_1=2_xm_ym_z] into itself, whereas operations from the third and fourth left cosets yield [F_3=F_4=m_x2_ym_z]. We shall return to this issue again at the end of Section 3.4.2.2.3[link].

3.4.2.2. Secondary domain states, partition of domain states

| top | pdf |

In this section we demonstrate that any morphic (spontaneous) property appears in the low-symmetry phase in several equivalent variants and find what determines their number and basic properties.

As we saw in Fig. 3.4.2.2[link], the spontaneous polarization – a principal tensor parameter of the [4_z/m_zm_xm_{x y}\supset 2_xm_ym_z ] phase transition – can appear in four different directions that define four principal domain states. Another morphic property is a spontaneous strain describing the change of unit-cell shape; it is depicted in Fig. 3.4.2.2[link] as a transformation of a square into a rectangle. This change can be expressed by a difference between two strain components [u_{11}-u_{22}={\lambda}^{(1)} ], which is a morphic tensor parameter since it is zero in the parent phase and nonzero in the ferroic phase. The quantity [{\lambda}^{(1)}=u_{11}-u_{22} ] is a secondary order parameter of the transition [4_z/m_zm_xm_{xy} \supset 2_xm_ym_z ] (for secondary order parameters see Section 3.1.3.2[link] ).

From Fig. 3.4.2.2[link], we see that two domain states [{\bf S}_1] and [{\bf S}_2] have the same spontaneous strain, whereas [{\bf S}_3] and [{\bf S}_4] exhibit another spontaneous strain [{\lambda}^{(2)}=u_{22}-u_{11}=-{\lambda}^{(1)}]. Thus we can infer that a property `to have the same value of spontaneous strain' divides the four principal domain states [{\bf S}_1], [{\bf S}_2], [{\bf S}_3 ] and [{\bf S}_4] into two classes: [{\bf S}_1] and [{\bf S}_2] with the same spontaneous strain [{\lambda}^{(1)}] and [{\bf S}_3] and [{\bf S}_4] with the same spontaneous strain [{\lambda}^{(2)}=-{\lambda}^{(1)}]. Spontaneous strain appears in two `variants': [{\lambda}^{(1)} ] and [{\lambda}^{(2)}=-{\lambda}^{(1)}].

We can define a ferroelastic domain state as a state of the crystal with a certain value of spontaneous strain [\lambda], irrespective of the value of the principal order parameter. Values [\lambda={\lambda}^{(1)} ] and [{\lambda}^{(2)}=-{\lambda}^{(1)}] thus specify two ferroelastic domain states [{\bf R}_1] and [{\bf R}_2], respectively. The spontaneous strain in this example is a secondary order parameter and the ferroelastic domain states can therefore be called secondary domain states.

An algebraic version of the above consideration can be deduced from Table 3.4.2.1[link], where to each principal domain state (given in the second column) there corresponds a left coset of [F_1=2_xm_ym_z ] (presented in the first column). Thus to the partition of principal domain states into two subsets [\{{\bf S}_1,{\bf S}_2,{\bf S}_3,{\bf S}_4\} = \{{\bf S}_1,{\bf S}_2\}_{{\lambda}^{(1)}}\cup \{{\bf S}_3, {\bf S}_4\}_{{\lambda}^{(2)}}, \eqno(3.4.2.14) ]there corresponds, according to relation (3.4.2.9[link]), a partition of left cosets [\eqalignno{&4_z/m_zm_xm_{xy}&\cr&\quad =\{\{2_xm_ym_z\} \cup \bar1\{2_xm_ym_z\}\} \cup \{2_{xy}\{2_xm_ym_z\} \cup 2_{x y}\{2_xm_ym_z\}\}&\cr &\quad =m_xm_ym_z \cup 2_{xy}\{m_xm_ym_z\}, &(3.4.2.15)} ]where we use the fact that the union of the first two left cosets of [2_xm_ym_z] is equal to the group [m_xm_ym_z]. This group is the stabilizer of the first ferroelastic domain state [{\bf R}_1], [I_G({\bf R}_1) = m_xm_ym_z]. Two left cosets of [m_xm_ym_z] correspond to two ferroelastic domain states, [{\bf R}_1] and [{\bf R}_2], respectively. Therefore, the number [n_a] of ferroelastic domain states is equal to the number of left cosets of [m_xm_ym_z] in [4_z/m_zm_xm_{xy} ], i.e. to the index of [m_xm_ym_z] in [4_z/m_zm_xm_{xy} ], [n_a =] [[4_z/m_zm_xm_{xy}] : [m_xm_ym_z]=|4_z/m_zm_xm_{xy}| ] : [|m_xm_ym_z|=16:8 = 2], and the number [d_a] of principal domain states in one ferroelastic domain state is equal to the index of [2_xm_ym_z] in [m_xm_ym_z], i.e. [d_a =[m_xm_ym_z]: [2_xm_ym_z] =|m_xm_ym_z|] : [|2_xm_ym_z|=8:4=2].

A generalization of these considerations, performed in Section 3.2.3.3.5[link] (see especially Proposition 3.2.3.30[link] and Examples 3.2.3.10[link] and 3.2.3.33[link] ), yields the following main results.

Assume that [{\lambda}^{(1)}] is a secondary order parameter of a transition with symmetry descent [G \supset F_1]. Then the stabilizer [L_1] of this parameter [I_G({\lambda}^{(1)})\equiv L_1] is an intermediate group, [F_1 \subseteq I_G({\lambda}^{(1)})\equiv L_1 \subseteq G. \eqno(3.4.2.16) ]Lattices of subgroups in Figs. 3.1.3.1[link] and 3.1.3.2[link] are helpful in checking this condition.

The set of n principal domain states (the orbit [G{\bf S }_1 ]) splits into [n_{\lambda}] subsets[n_{\lambda}=[G:L_{1}]=|G|:|L_{1}|. \eqno(3.4.2.17)]

Each of these subsets consists of [d_{\lambda}] principal domain states,[d_{\lambda}=[L_1:F_{1}]=|L_1|:|F_{1}|. \eqno(3.4.2.18) ]The number [d_{\lambda}] is called a degeneracy of secondary domain states.

The product of numbers [n_{\lambda}] and [d_{\lambda}] is equal to the number n of principal domain states [see equation (3.2.3.26[link] )]: [n_{\lambda}d_{\lambda}=n. \eqno(3.4.2.19) ]

Principal domain states from each subset have the same value of the secondary order parameter [{\lambda}^{(j)}, j=1,2,\ldots,n_{\lambda}] and any two principal domain states from different subsets have different values of [{\lambda}^{(j)}]. A state of the crystal with a given value of the secondary order parameter [{\lambda}^{(j)}] will be called a secondary domain state [{\bf R}_j, j=1,2,\ldots,n_{\lambda}]. Equivalent terms are degenerate or compound domain state.

In a limiting case [L_1=F_1], the parameter [\lambda^{(1)}] is identical with the principal tensor parameter and there is no degeneracy, [d_{\lambda}=1].

Secondary domain states [{\bf R}_1,{\bf R}_2,\ldots,{\bf R}_j,\ldots,{\bf R}_{n_{\lambda}} ] are in a one-to-one correspondence with left cosets of [L_1] in the decomposition[G=h_1L_{1} \cup h_2L_{1} \cup\ldots\cup h_jL_{1} \cup\ldots\cup h_{n_{\lambda}}L_{1}, \eqno(3.4.2.20) ]therefore[{\bf R}_j=h_j{\bf R}_1,\quad j=1,2,\ldots,n_{\lambda}. \eqno(3.4.2.21) ]

Principal domain states of the first secondary domain state [{\bf R}_1 ] can be determined from the first principal domain state [{\bf S}_1 ]: [{\bf S}_k = p_k{\bf S}_1,\quad k=1,2,\ldots,d_{\lambda}, \eqno(3.4.2.22) ]where [p_k] is the representative of the kth left coset of [F_1] of the decomposition[L_{1}=p_1F_{1} \cup p_2F_{1} \cup \ldots \cup p_kF_{1} \cup \ldots \cup p_{d_{\lambda}}F_{1}. \eqno(3.4.2.23) ]

The partition of principal domain states according to a secondary order parameter offers a convenient labelling of principal domain states by two indices [j, k], where the first index j denotes the sequential number of the secondary domain state and the second index k gives the sequential number of the principal domain state within the jth secondary domain state [see equation (3.2.3.79[link] )]:[{\bf S}_{jk} = h_jp_k{\bf S}_{11}, \quad {\bf S}_{11}={\bf S}_1, \quad j=1,2,\ldots,n_{\lambda}, \quad k = 1,2,\ldots,d_{\lambda}, \eqno(3.4.2.24) ]where [h_j] and [p_k] are representatives of the decompositions (3.4.2.20[link]) and (3.4.2.23[link]), respectively.

The secondary order parameter [\lambda] can be identified with a principal order parameter of a phase transition with symmetry descent [G \subset L_1] (see Section 3.4.2.3[link]). The concept of secondary domain states enables one to define domain states that are characterized by a certain spontaneous property. We present the three most significant cases of such ferroic domain states.

3.4.2.2.1. Ferroelastic domain state

| top | pdf |

The distinction ferroelastic–non-ferroelastic is a basic division in domain structures. Ferroelastic transitions are ferroic transitions involving a spontaneous distortion of the crystal lattice that entails a change of shape of the crystallographic or conventional unit cell (Wadhawan, 2000[link]). Such a transformation is accompanied by a change in the number of independent nonzero components of a symmetric second-rank tensor [u ] that describes spontaneous strain.

In discussing ferroelastic and non-ferroelastic domain structures, the concepts of crystal family and holohedry of a point group are useful (IT A , 2005[link]). Crystallographic point groups (and space groups as well) can be divided into seven crystal systems and six crystal families (see Table 3.4.2.2[link]). A symmetry descent within a crystal family does not entail a qualitative change of the spontaneous strain – the number of independent nonzero tensor components of the strain tensor u remains unchanged.

Table 3.4.2.2 | top | pdf |
Crystal systems, holohedries, crystal families and number of spontaneous strain components

Point group M Crystal system Holohedry HolM Spontaneous strain components Crystal family FamM
Independent Nonzero
[23], [m\bar3], [432], [\bar43m], [m\bar3m] Cubic [m\bar3m] 1 3 Cubic
[6], [\bar6], [6/m], [622], [6mm], [\bar62m], [6/mmmm] Hexagonal [6/mmm] 2 3 Hexagonal
[3], [\bar3], [32], [3m], [\bar3m] Trigonal [\bar3m] 2 3
[4], [\bar4], [4/m], [422], [4mm], [\bar42m], [4/mmm] Tetragonal [4/mmm] 2 3 Tetragonal
[222], [mm2], [mmm] Orthorhombic [mmm] 3 3 Orthorhombic
[2], m, [2/m] Monoclinic [2/m] 4 4 Monoclinic
[1], [\bar1] Triclinic [\bar1] 6 6 Triclinic

We shall denote the crystal family of a group M by the symbol FamM. Then a simple criterion for a ferroic phase transition with symmetry descent [G \subset F] to be a non-ferroelastic phase transition is[F \subset G, \quad{\rm Fam}F ={\rm Fam}G. \eqno(3.4.2.25) ]

A necessary and sufficient condition for a ferroelastic phase transition is[F \subset G, \quad{\rm Fam}F \neq {\rm Fam}G. \eqno(3.4.2.26) ]

A ferroelastic domain state [{\bf R}_i] is defined as a state with a homogeneous spontaneous strain [u^{(i)}]. [We drop the suffix `s' or `(s)' if the serial number of the domain state is given as the superscript [(i)]. The definition of spontaneous strain is given in Section 3.4.3.6.1[link].] Different ferroelastic domain states differ in spontaneous strain. The symmetry of a ferroelastic domain state Ri is specified by the stabilizer [I_G(u^{(i)}) ] of the spontaneous strain [u^{(i)}] of the principal domain state [{\bf S}_i] [see (3.4.2.16[link])]. This stabilizer, which we shall denote by [A_i], can be expressed as an intersection of the parent group G and the holohedry of group [F_i], which we shall denote Hol[F_i] (see Table 3.4.2.2[link]):[A_i \equiv I_G(u^{(i)})=G\cap {\rm Hol}F_i. \eqno(3.4.2.27) ]This equation indicates that the ferroelastic domain state Ri has a prominent single-domain orientation. Further on, the term `ferroelastic domain state' will mean a `ferroelastic domain state in single-domain orientation'.

In our illustrative example, [\eqalign{A_1 &= I_{4_z/m_zm_xm_{xy}}(u_{11}-u_{22})\cr &= {\rm Hol}(2_xm_ym_z)\cap m4_z/m_zm_xm_{xy}\cr &=m_xm_ym_z \cap 4_z/m_zm_xm_{xy}= m_xm_ym_z.\cr} ]

The number [n_a] of ferroelastic domain states is given by[n_a = [G:A_1] = |G|:|A_1|. \eqno(3.4.2.28)]In our example, [n_a=|4_z/m_zm_xm_{xy}|:|m_xm_ym_z|=16:8=2]. In Table 3.4.2.7[link], last column, the number [n_a] of ferroelastic domain states is given for all possible ferroic phase transitions.

The number [d_a] of principal domain states compatible with one ferroelastic domain state (degeneracy of ferroelastic domain states) is given by[d_a=[A_1:F_1]=|A_1|:|F_1|. \eqno(3.4.2.29) ]In our example, [d_a=|m_xm_ym_z|:|2_xm_ym_z|=8:4=2], i.e. two non-ferroelastic principal domain states are compatible with each of the two ferroelastic domain states (cf. Fig. 3.4.2.2[link]).

The product of [n_a] and [d_a] is equal to the number n of all principal domain states [see equation (3.4.2.19[link])],[n_ad_a=[G:A_1][A_1:F_1]=[G:F_1]=n. \eqno(3.4.2.30) ]The number [d_a] of principal domain states in one ferroelastic domain state can be calculated for all ferroic phase transitions from the ratio of numbers n and [n_a] that are given in Table 3.4.2.7[link].

According to Aizu (1969[link]), we can recognize three possible cases:

  • (i) Full ferroelastics: All principal domain states differ in spontaneous strain. In this case, [n_a=n], i.e. [A_1=F_1], ferroelastic domain states are identical with principal domain states.

  • (ii) Partial ferroelastics: Some but not all principal domain states differ in spontaneous strain. A necessary and sufficient condition is [1 \,\lt\, n_a \,\lt\, n], or, equivalently, [F_1 \subset A_1 \subset G]. In this case, ferroelastic domain states are degenerate secondary domain states with degeneracy [n>d_a=|A_1|:|F_1|>1 ]. In this case, the phase transition [G\supset F_1] can also be classified as an improper ferro­elastic one (see Section 3.1.3.2[link] ).

  • (iii) Non-ferroelastics: All principal domain states have the same spontaneous strain. The criterion is [n_a=1 ], i.e. [A_1=G].

A similar classification for ferroelectric domain states is given below. Both classifications are summarized in Table 3.4.2.3[link].

Table 3.4.2.3 | top | pdf |
Aizu's classification of ferroic phases

[n_a] is the number of ferroelastic domain states, [n_e] is the number of ferroelectric domain states and [n_f] is the number of ferroic domain states.

Ferroelastic Ferroelectric
Fully Partially Non-ferroelastic Fully Partially Non-ferroelectric
[n_a=n] [1 \,\lt\, n_a \,\lt\, n] [n_a=1 ] [n_e=n] [1 \,\lt\, n_e \,\lt\, n] [n_e=0], 1

Example 3.4.2.1. Domain states in leucite.  Leucite (KAlSi2O6) (see e.g. Hatch et al., 1990[link]) undergoes at about 938 K a ferro­elastic phase transition from cubic symmetry [G=m\bar3m] to tetragonal symmetry [L=4/mmm]. This phase can appear in [|G=m\bar3m|:|4/mmm|=3] single-domain states, which we denote [{\bf R}_1], [{\bf R}_2], [{\bf R}_3 ]. The symmetry group of the first domain state [{\bf R}_1] is [L_1=4_x/m_xm_ym_z]. This group equals the stabilizer [I_G(u^{(1)}) ] of the spontaneous strain [u^{(1)}] of [{\bf R}_1] since Hol([4_x/m_xm_ym_z)] [=4_x/m_xm_ym_z] (see Table 3.4.2.2[link]), hence this phase is a full ferroelastic one.

At about 903 K, another phase transition reduces the symmetry [4/mmm] to [F= 4/m]. Let us suppose that this transition has taken place in a domain state [{\bf R}_1] with symmetry [L_1=4_x/m_xm_ym_z]; then the room-temperature ferroic phase has symmetry [F_1=4_x/m_x]. The [4_x/m_xm_ym_z \supset 4_x/m_x] phase transition is a non-ferroelastic one [[{\rm Hol}(4_x/m_x) =] [{\rm Hol}(4_x/m_xm_ym_z) =] [4_x/m_xm_ym_z]] with [|4_x/m_xm_ym_z|:|4_x/m_x|=8:4=2] non-ferro­elastic domain states, which we denote [{\bf S}_1] and [{\bf S}_2]. Similar considerations performed with initial domain states R2 and R3 generate another two couples of principal domain states [{\bf S}_3 ], [{\bf S}_4] and [{\bf S}_5], [{\bf S}_6], respectively. Thus the room-temperature phase is a partially ferroelastic phase with three degenerate ferroelastic domain states, each of which can contain two principal domain states. Both ferroelastic domains and non-ferroelastic domains within each ferroelastic domain have been observed [see Fig. 3.3.10.13[link] in Chapter 3.3[link] , Palmer et al. (1988[link]) and Putnis (1992)[link]].

3.4.2.2.2. Ferroelectric domain states

| top | pdf |

Ferroelectric domain states are defined as states with a homogeneous spontaneous polarization; different ferroelectric domain states differ in the direction of the spontaneous polarization. Ferroelectric domain states are specified by the stabilizer [I_G({\rm P}_s^{(1)})] of the spontaneous polarization [{\rm P}_s^{(1)}] in the first principal domain state [{\bf S}_1] [see equation (3.4.2.16[link])]:[F_1\subseteq C_1 \equiv I_G({\rm P}_s^{(1)}) \subseteq G. \eqno(3.4.2.31) ]The stabilizer [C_1] is one of ten polar groups: 1, 2, 3, 4, 6, m, [mm2], [3m], [4mm ], [6mm]. Since [F_1] must be a polar group too, it is simple to find the stabilizer [C_1] fulfilling relation (3.4.2.31[link]).

The number [n_e] of ferroelectric domain states is given by[n_e=[G:C_1]=|G|:|C_1|. \eqno(3.4.2.32)]If the polar group [C_1] does not exist, we put [n_e=0]. The number [n_e] of ferroelectric domain states is given for all ferroic phase transitions in the eighth column of Table 3.4.2.7[link].

The number [d_a] of principal domain states compatible with one ferroelectric domain state (degeneracy of ferroelectric domain states) is given by[d_e=[C_1:F_1]=|C_1|:|F_1|. \eqno(3.4.2.33) ]

The product of [n_e] and [d_e] is equal to the number n of all principal domain states [see equation (3.4.2.19[link])],[n_ed_e=n. \eqno(3.4.2.34)]The degeneracy [d_e] of ferroelectric domain states can be calculated for all ferroic phase transitions from the ratio of the numbers n and [n_e] that are given in Table 3.4.2.7[link].

According to Aizu (1969[link], 1970a[link]), we can again recognize three possible cases (see also Table 3.4.2.3[link]):

  • (i) Full ferroelectrics: All principal domain states differ in spontaneous polarization. In this case, [n_e=n], i.e. [C_1=F_1], ferroelectric domain states are identical with principal domain states.

  • (ii) Partial ferroelectrics: Some but not all principal domain states differ in spontaneous polarization. A necessary and sufficient condition is [1\,\lt\, n_e\,\lt \,n], or equivalently, [F_1 \subset C_1 \subset G]. Ferroelectric domain states are degenerate secondary domain states with degeneracy [n\,\gt \,d_e\,\gt\, 1]. In this case, the phase transition [G\supset F_1] can be classified as an improper ferroelectric one (see Section 3.1.3.2[link] ).

  • (iii) Non-ferroelectrics: No principal domain states differ in spontaneous polarization. There are two possible cases: (a) The parent phase is polar; then [C_1=G] and [n_e=1 ]. (b) The parent phase is non-polar; in this case a polar stabilizer [C_1] does not exist, then we put [n_e=0].

The classification of full-, partial- and non-ferroelectrics and ferroelastics is given for all Aizu's species in Aizu (1970a[link]).

This classification for all symmetry descents is readily available from the numbers n, [n_a], [n_e] in Table 3.4.2.7[link]. One can conclude that partial ferroelectrics are rather rare.

Example 3.4.2.3. Domain structure in tetragonal perovskites.  Some perovskites (e.g. barium titanate, BaTiO3) undergo a phase transition from the cubic parent phase with [G=m\bar3m] to a tetragonal ferroelectric phase with symmetry [F_1=4_xm_ym_z]. The stabilizer [A_1 =] Hol[(4_xm_ym_z)\cap m3m =] [m_xm_ym_z]. There are [n_a =] [ |m3m|: |m_xm_ym_z|  =] 3 ferroelastic domain states each compatible with [d_a =] [|m_xm_ym_z|:|4_xm_ym_z|  =] 2 principal ferroelectric domain states that are related e.g. by inversion [\bar1], i.e. spontaneous polarization is antiparallel in two principal domain states within one ferroelastic domain state.

A similar situation, i.e. two non-ferroelastic domain states with antiparallel spontaneous polarization compatible with one ferroelastic domain state, occurs in perovskites in the trigonal ferroic phase with symmetry [F=3m] and in the orthorhombic ferroic phase with symmetry [F_1=m_{x\bar y}2_{xy}m_z ].

Many other examples are discussed by Newnham (1974[link], 1975[link]), Newnham & Cross (1974a[link],b[link]), and Newnham & Skinner (1976[link]).

3.4.2.2.3. Domain states with the same stabilizer

| top | pdf |

In our illustrative example (see Fig. 3.4.2.2[link]), we have seen that two domain states [{\bf S}_1] and [{\bf S}_2] have the same symmetry group (stabilizer) [2_xm_ym_z]. In general, the condition `to have the same stabilizer (symmetry group)' divides the set of n principal domain states into equivalence classes. As shown in Section 3.2.3.3[link] , the role of an intermediate group [L_1] is played in this case by the normalizer [N_G(F_1)] of the symmetry group [F_1] of the first domain state [{\bf S}_1]. The number [d_F] of domain states with the same symmetry group is given by [see Example 3.2.3.34[link] in Section 3.2.3.3.5[link] and equation (3.2.3.95[link] )], [d_F=[N_G(F_1):F_1]=|N_G(F_1)|:|F_1|.\eqno(3.4.2.35) ]The number [n_F] of subgroups that are conjugate under G to [F_1] can be calculated from the formula [see equation (3.2.3.96[link] )][n_F=[G:N_G(F_1)]=|G|:|N_G(F_1)|.\eqno(3.4.2.36) ]The product of [n_F] and [d_F] is equal to the number n of ferroic domain states, [n=n_Fd_F.\eqno(3.4.2.37) ]

The normalizer [N_G(F_1)] enables one not only to determine which domain states have the symmetry [F_1] but also to calculate all subgroups that are conjugate under G to [F_1] (see Examples 3.2.3.22[link] , 3.2.3.29[link] and 3.2.3.34[link] in Section 3.2.3.3[link] ).

Normalizers [N_G(F_1)] and the number [d_F] of principal domain states with the same symmetry are given in Table 3.4.2.7[link] for all symmetry descents [G \supset F_1]. The number [n_F] of subgroups conjugate to [F_1] is given by [n_F=n:d_F ].

All these results obtained for point-group symmetry descents can be easily generalized to microscopic domain states and space-group symmetry descents (see Section 3.4.2.5[link]).

3.4.2.3. Property tensors associated with ferroic domain states

| top | pdf |

In the preceding section we derived relations for domain states without considering their specific physical properties. Basic formulae for the number of principal and secondary domain states [see equations (3.4.2.11[link]) and (3.4.2.17[link]), respectively] and the transformation properties of these domain states [equations (3.4.2.12[link]) and (3.4.2.21[link]), respectively] follow immediately from the symmetry groups G, [F_1] of the parent and ferroic phases, respectively. Now we shall examine which components of property tensors specify principal and secondary domain states and how these tensor components change in different domain states.

A property tensor [\tau] is specified by its components. The number [m_i(\tau)] of independent tensor components of a certain tensor [\tau ] depends on the point-group symmetry G of the crystal (see Chapter 1.1[link] ). The number [m_c(\tau)] of nonzero Cartesian (rectangular) components depends on the orientation of the crystal in the reference Cartesian coordinate system and is equal to, or greater than, the number [m_i(\tau) ] of independent tensor components; this number [m_i(\tau)] is independent of orientation. Then there are [m_c(\tau)-m_i(\tau)] linear relations between Cartesian tensor components. The difference [m_c(\tau)-m_i(\tau) ] is minimal for a `standard' orientation, in which symmetry axes of the crystal are, if possible, parallel to the axes of the reference coordinate system [for more on this choice, see Nye (1985[link]) Appendix B, Sirotin & Shaskolskaya (1982[link]), Shuvalov (1988[link]) and IEEE Standards on Piezoelectricity (1987[link])]. Even in this standard orientation, only for point groups of triclinic, monoclinic and orthorhombic crystal systems is the number [m_c(\tau)] of nonzero Cartesian components of each property tensor equal to the number [m_i(\tau) ] of independent tensor components, i.e. all Cartesian tensor components are independent. For all other point groups [m_c(\tau)-m_i(\tau)\,\gt\,0 ], i.e. there are always relations between some Cartesian tensor components. One can verify this statement for the strain tensor in Table 3.4.2.2[link].

The relations between Cartesian tensor components can be removed when one uses covariant tensor components. [Kopský (1979[link]); see also the manual of the the software GI[\star]KoBo-1 and Kopský (2001[link]). An analogous decomposition of Cartesian tensors into irreducible parts has been performed by Jerphagnon et al. (1978[link]).] Covariant tensor components are linear combinations of Cartesian tensor components that transform according to irreducible matrix representations [D^{(\alpha)}(G)] of the group G of the crystal (i.e. they form a basis of irreducible representations of G; for irreducible representations see Chapter 1.2[link] ). The number of covariant tensor components equals the number of independent components of the tensor [\tau].

The advantage of expressing property tensors by covariant tensor components becomes obvious when one considers a change of a property tensor at a ferroic phase transition. A symmetry descent [G\supset F_1] is accompanied by the preservation of, or an increase of, the number of independent Cartesian tensor components. The latter possibility can manifest itself either by the appearance of morphic Cartesian tensor components in the low-symmetry phase or by such changes of nonzero Cartesian components that break some relations between tensor components in the high-symmetry phase. This is seen in our illustrative example of the strain tensor u. In the high-symmetry phase with [G=4_z/m_zm_xm_{xy}], the strain tensor has two independent components and three nonzero components: [u_{11}\neq u_{22}=u_{33}]. In the low-symmetry phase with [F_1=2_xm_ym_z], there are three independent and three nonzero components: [u_{11}\neq u_{22}\neq u_{33}], i.e. the equation [u_{22}=u_{33}] does not hold in the parent phase. This change cannot be expressed by a single Cartesian morphic component.

Since there are no relations between covariant tensor components, any change of tensor components at a symmetry descent can be expressed by morphic covariant tensor components, which are zero in the parent phase and nonzero in the ferroic phase. In our example, the covariant tensor component of the spontaneous strain is [u_{11}-u_{22}], which is a morphic component since [u_{11}-u_{22}=0 ] for the symmetry [4_z/m_zm_xm_{xy}] but [u_{11}-u_{22}\neq 0 ] for symmetry [2_xm_ym_z].

Tensorial covariants are defined in an exact way in the manual of the software GI[\star]KoBo-1 and in Kopský (2001[link]). Here we give only a brief account of this notion. Consider a crystal with symmetry G and a property tensor [\tau] with [n_{\tau} ] independent tensor components. Let [D^{(\alpha)}(G)] be a [d_{\alpha}]-dimensional physically irreducible matrix representation of G. The [D^{(\alpha)}_{a}(G)] covariant of [\tau] consists of the following [d_{\alpha}] covariant tensor components: [\tau^{\alpha}_a =] [(\tau^{\alpha}_{ a,1}, \tau^{\alpha}_{a,2},\ldots,\tau^{\alpha}_{a,d_{\alpha}}) ], where a = [1, 2, \ldots] and [m=n_{\tau}/d_{\alpha}] numbers different [d_{\alpha}]-tuples formed from [n_{\tau}] components of [\tau]. These covariant tensor components are linear combinations of Cartesian components of [\tau] that transform as so-called typical variables of the matrix representation [D^{(\alpha)}(G)], i.e. the transformation properties under operations [g\in G] of covariant tensor components are expressed by matrices [D^{\alpha}(g)].

The relation between two presentations of the tensor [\tau] is provided by conversion equations, which express Cartesian tensor components as linear combinations of covariant tensor components and vice versa [for details see the manual and Appendix E of the software GI[\star]KoBo-1 and Kopský (2001[link])].

Tensorial covariants for all non-equivalent physically irreducible matrix representations of crystallographic point groups and all important property tensors up to rank four are listed in the software GI[\star]KoBo-1 and in Kopský (2001[link]). Thus, for example, in Table D of the software GI[\star]KoBo-1, or in Kopský (2001[link]) p. 5, one finds for the two-dimensional irreducible representation E of group 422 the following tensorial covariants: [(P_1,P_2)], [(d_{11},d_{22})], [(d_{12},d_{21}) ], [(d_{13}, d_{23})], [(d_{26},d_{16})], [(d_{35},d_{34}) ].

Let us denote by [\tau^{(\alpha)(1)}_a] a tensorial covariant of [\tau] in the first single-domain state [{\bf S}_1]. A crucial role in the analysis is played by the stabilizer [I_G(\tau^{(\alpha)(1)}_a) ] of these covariants, i.e. all operations of the parent group G that leave [\tau^{(\alpha)(1)}_a] invariant. There are three possible cases:

  • (1) If[I_G(\tau^{(\alpha)(1)}_a)=G, \eqno(3.4.2.38) ]then all components of [\tau^{(\alpha)(1)}_a] that are nonzero in the parent phase are also nonzero in the ferroic phase. All these components are the same in all principal domain states. For important property tensors and for all point groups G, these covariant tensor components are listed in the main tables of the software GI[\star]KoBo-1 and in Kopský (2001[link]). The corresponding Cartesian tensor components are available in Section 1.1.4[link] and in standard textbooks (e.g. Nye, 1985[link]; Sirotin & Shaskolskaya, 1982[link]).

  • (2) If [I_G(\tau^{(\alpha)(1)}_a)=F_1, \eqno(3.4.2.39) ]then any of [m=n_{\tau}/d_{\alpha}] tensorial covariants [\tau^{(\alpha)}_a], [a=1,2,\ldots,m], is a possible principal tensor parameter [\varphi^{(1)}] of the transition [G\supset F_1]. Any two of [n_f=|G|:|F_1|] principal domain states differ in some, or all, components of these covariants. The principal tensor parameter [\varphi] plays a similar symmetric (but generally not thermodynamic) role as the order parameter [\eta] does in the Landau theory. Only for equitranslational phase transitions is one of the principal tensor parameters (that with the temperature-dependent coefficient) identical with the primary order parameter of the Landau theory (see Section 3.1.3[link] ).

  • (3) If [I_G(\tau^{(\alpha)(1)}_a)=L_1, \ \ F_1\subset L_1\subset G, \eqno(3.4.2.40) ]then [\tau^{(\alpha)(1)}_a] represents the secondary tensor parameter [\lambda] (see Section 3.1.3.2[link] ). There exist [n_{\lambda}=|G|:|L_1|] secondary domain states [{\bf R}_1,] [{\bf R}_2,] [\ldots,] [{\bf R}_{n_{\lambda}}] that differ in [\lambda]. Unlike in the two preceding cases (1[link]) and (2[link]), several intermediate groups [L_1,M_2,\ldots ] (with secondary tensor parameters [\lambda, \mu,\ldots]) that fulfil condition (3.4.2.40[link]) can exist.

Now we shall indicate how one can find particular property tensors that fulfil conditions (3.4.2.39[link]) or (3.4.2.40[link]). The solution of this group-theoretical task consists of three steps:

  • (i) For a given point-group symmetry descent [G\supset F_1], or [G\supset L_1], one finds the representation [{\Gamma}_{\eta}] that specifies the transformation properties of the principal, or secondary, tensor parameter, which plays the role of the order parameters in a continuum description. This task is called an inverse Landau problem (see Section 3.1.3[link] for more details). The solution of this problem is available in Tables 3.4.2.7[link] and 3.1.3.1[link] , in the software GI[\star]KoBo-1 and in Kopský (2001[link]), where the letters A, B signify one-dimensional irreducible representations, and letters E and T two- and three-dimensional ones. The dimensionality [d_{\eta}], or [d_{\lambda}], of the representation [{\Gamma}_{\eta} ], or [{\Gamma}_{\lambda}], specifies the maximal number of independent components of the principal, or secondary, tensor parameter [\varphi], or [\lambda], respectively. `Reducible' indicates that [{\Gamma}_{\eta} ] is a reducible representation.

  • (ii) In Table 3.1.3.1[link] one finds in the second column, for a given G and [{\Gamma}_{\eta}], or [\Gamma_{\lambda}] (first column), the standard variables designating in a standardized way the covariant tensor components of the principal, or secondary, tensor parameters (for more details see Section 3.1.3.1[link] and the manual of the software GI[\star]KoBo-1). For two- and three-dimensional irreducible representations, this column contains relations that restrict the values of the components and thus reduce the number of independent components.

  • (iii) The association of covariant tensor components of property tensors with standard variables is tabulated for all irreducible representations in an abridged version in Table 3.1.3.1[link] , in the column headed Principal tensor parameters, and in full in the main table of the software GI[\star]KoBo-1 and of Kopský (2001[link]).

Phase transitions associated with reducible representations are treated in detail only in the software GI[\star]KoBo-1 and in Kopský (2001[link]). Fortunately, these phase transitions occur rarely in nature.

A rich variety of observed structural phase transitions can be found in Tomaszewski (1992[link]). This database lists 3446 phase transitions in 2242 crystalline materials.

Example 3.4.2.4. Morphic tensor components associated with [4_z/m_zm_xm_{xy}\supset 2_xm_ym_z ] symmetry descent  

  • (1) Principal tensor parameters [\varphi^{(1)}]. The representation [{\Gamma}_{\eta} ] that specifies the transformation properties of the principal tensor parameter [\varphi^{(1)}] (and for equitranslational phase transitions also the primary order parameter [\eta^{(1)}]) can be found in the first column of Table 3.1.3.1[link] for [G=4_z/m_zm_xm_{xy} ] and [F_1=2_xm_ym_z]; the R-irreducible representation (R-irep) [\Gamma_{\eta}=E_u]. Therefore, the principal tensor parameter [\varphi^{(1)}] (or the primary order parameter [\eta^{(1)} ]) has two components [(\varphi_1^{(1)},\varphi_2^{(1)})] [or [(\eta_1^{(1)},\eta_2^{(1)}])]. The standard variables are in the second column: [({\sf x}_1^{-}, 0)]. This means that only the first component [\varphi^{(1)}_1] (or [\eta^{(1)}_1]) is nonzero. In the column Principal tensor parameters, one finds that [\varphi^{(1)}_1=P_1] (or [\eta^{(1)}_1=P_1]), i.e. one principal tensor parameter is spontaneous polarization and the spontaneous polarization in the first domain state [{\bf S}_1] is [P_{(s)}=(P,00)]. Other principal tensor parameters can be found in the software GI[\star]KoBo-1 or in Kopský (2001[link]), p. 185: [(g_4,0), ] [(d_{11},0),] [(d_{12},0),] [(d_{13},0),] [(d_{26},0),] [(d_{35},0)] (the physical meaning of the components is explained in Table 3.4.3.5[link]).

  • (2) Secondary tensor parameters [\lambda^{(1)}, \mu^{(1)},\ldots ].

    In the group lattice (group–subgroup chains) in Fig. 3.1.3.1[link] , one finds that the only intermediate group between [4_z/m_zm_xm_{xy}] and [2_xm_ym_z] is [L_1=m_xm_ym_z]. In the same table of the software GI[\star]KoBo-1 or in Kopský (2001[link]), one finds [\Gamma_{\lambda}=B_{1g} ] and the following one-dimensional secondary tensor parameters: [u_1-u_2;] [A_{14}+A_{25},] [A_{36};] [s_{11}-s_{22}, ] [s_{13}-s_{23},] [s_{44}-s_{55};] [Q_{11}-Q_{22},] [Q_{12}-Q_{21},] [Q_{13}-Q_{23},] [Q_{31}-Q_{32},] [Q_{44}-Q_{55} ].

The use of covariant tensor components has two practical advantages:

Firstly, the change of tensor components at a ferroic phase transition is completely described by the appearance of new nonzero covariant tensor components. If needed, Cartesian tensor components corresponding to covariant components can be calculated by means of conversion equations, which express Cartesian tensor components as linear combinations of covariant tensor components [for details on tensor covariants and conversion equations see the manual and Appendix E of the software GI[\star]KoBo-1 and Kopský (2001[link])].

Secondly, calculation of property tensors in various domain states is substantially simplified: transformations of Cartesian tensor components, which are rather involved for higher-rank tensors, are replaced by a simpler transformation of covariant tensor components by matrices [D^{(\eta)}] of the matrix representation of [\Gamma_{\eta}], or of [\Gamma_{\lambda}] [see again the software GI[\star]KoBo-1 and Kopský (2001[link])]. The determination of the tensor properties of all domain states is discussed in full in the book by Kopský (1982[link]).

The relations between morphic properties, tensor parameters, order parameters and names of domain states are compared in Table 3.4.2.4[link], from which it is seen that what matters in distinguishing different domain states is the stabilizer of the spontaneous (morphic) property, where physically different parameters may possess a common stabilizer. The latter thermodynamic division, based on conditions of the stability, is finer than the former division, which is based on symmetry only. This difference manifests itself, for example, in the fact that two physically different tensor parameters, such as the principal order parameter [\varphi] and a `similar' order parameter [\sigma ], transform according to different representations [\Gamma_{\varphi} ] and [\Gamma_{\sigma}] but have the same stabilizer [F_1] (such symmetry descents are listed in Table 3.1.3.2[link] ) and possess common domain states. This `degeneracy' of domain states can be even more pronounced in the microscopic description, where the same stabilizer [{\cal F}_1] and therefore a common basic domain state can be shared by three physically different order parameters: a primary order parameter [\eta] (the order parameter, components of which form a quadratic invariant with a temperature-dependent coefficient in the free energy), a pseudoproper order parameter [\zeta] that transforms according to the same representation [\Gamma_{\eta}] as the primary order parameter but has a temperature coefficient that is not as strongly temperature-dependent as the primary order parameter, and a `similar' order parameter [\kappa] with a representation [\Gamma_{\kappa}] different from [\Gamma_{\eta}].

Table 3.4.2.4 | top | pdf |
Morphic properties, tensor parameters, order parameters, stabilizers and domain states

Morphic property Tensor or order parameter [\Gamma ] Stabilizer of morphic property Domain states
Principal tensor parameter [\varphi^{\,(1)}] [{\Gamma}_{\varphi} ] [F_1] Principal
`Similar principal' tensor parameter [\sigma^{\,(1)}] [{\Gamma}_{\sigma} ]
Secondary tensor parameter [\lambda^{(1)}] [{\Gamma}_{\lambda} ] [L_1, \ F_1\subset L_1\subset G] Secondary ferroic
Spontaneous polarization [{\bf P}_{(s)}] [{\Gamma}_{{\bf P}_{(s)}} ] [C_1=I_G(P_{(s)}^{\,(1)})] Ferroelectric
Spontaneous strain [u_{(s)}] [{\Gamma}_{u_{(s)}}] [A_1={\rm Hol}F_1\cap G] Ferroelastic
Primary order parameter [\eta^{(1)}] [\Gamma_{\eta}] [{\cal F}_1] Primary, basic, microscopic
Pseudoproper order parameter [\zeta^{\,(1)}]
`Similar' order parameter [\kappa^{\,(1)}] [{\Gamma}_{\kappa} ]
Secondary order parameter [{\tau}^{(1)}] [{\Gamma}_{\tau} ] [{\cal M}_1, {\cal F}_1\subset {\cal M}_1\subset {\cal G} ] Secondary microscopic

3.4.2.4. Synoptic table of ferroic transitions and domain states

| top | pdf |

The considerations of this and all following sections can be applied to any phase transition with point-group symmetry descent [G\supset F]. All such non-magnetic crystallographically non-equivalent symmetry descents are listed in Table 3.4.2.7[link] together with some other data associated with symmetry reduction at a ferroic phase transition. These symmetry descents can also be traced in lattices of point groups, which are displayed in Figs. 3.1.3.1[link] and 3.1.3.2[link] .

The symmetry descents [G\supset F_1] listed in Table 3.4.2.7[link] are analogous to Aizu's `species' (Aizu, 1970a[link]), in which the symbol F stands for the symbol [\supset ] in our symmetry descent, and the orientation of symmetry elements of the group [F_1] with respect to G is specified by letters p, s, ps, pp etc.

As we have already stated, any systematic analysis of domain structures requires an unambiguous specification of the orientation and location of symmetry elements in space. Moreover, in a continuum approach, the description of crystal properties is performed in a rectangular (Cartesian) coordinate system, which differs in hexagonal and trigonal crystals from the crystallographic coordinate system common in crystallography. Last but not least, a ready-to-use and user-friendly presentation calls for symbols that are explicit and concise.

To meet these requirements, we use in this chapter, in Section 3.1.3[link] and in the software GI[\star]KoBo-1 a symbolism in which the orientations of crystallographic elements and operations are expressed by means of suffixes related to a reference Cartesian coordinate system. The relation of this reference Cartesian coordinate system – called a crystallophysical coordinate system – to the usual crystallographic coordinate system is a matter of convention. We adhere to the generally accepted rules [see Nye (1985[link]) Appendix B, Sirotin & Shaskolskaya (1982[link]), Shuvalov (1988[link]), and IEEE Standards on Piezoelectricity, 1987[link]].

We list all symbols of crystallographic symmetry operations and a comparison of these symbols with other notations in Tables 3.4.2.5[link] and 3.4.2.6[link] and in Figs. 3.4.2.3[link] and 3.4.2.4[link].

Table 3.4.2.5 | top | pdf |
Symbols of symmetry operations of the point group [m\bar 3m ]

Standard: symbols used in Section 3.1.3[link] , in the present chapter and in the software; all symbols refer to the cubic crystallographic (Cartesian) basis, [p\equiv[111]] (all [{\underline{p}}]ositive), [q\equiv[\bar1\bar11], \ r\equiv [1\bar1\bar1], \ s\equiv [\bar11\bar1] ]. BC: Bradley & Cracknell (1972[link]). AH: Altmann & Herzig (1994[link]). IT A: IT A (2005[link]). Jones: Jones' faithful representation symbols express the action of a symmetry operation on a vector [(xyz)] (see e.g. Bradley & Cracknell, 1972[link]).

Standard BC AH IT A Jones Standard BC AH IT A Jones
1 or e E E 1 [x,y,z] [\bar{1}] or i I i [{\bar 1}]   [0,0,0] [\bar{x},\bar{y},\bar{z}]
[2_{z}] [C_{2z}] [C_{2z}] 2   [0,0,z] [{\bar x},{\bar y},z] [m_{z}] [\sigma_{z}] [\sigma_{z}] m    [x,y,0 ] [x,y,{\bar z}]
[2_{x}] [C_{2x}] [C_{2x}] 2   [x,0,0] [x,{\bar {y},{\bar z}}] [m_{x}] [\sigma_{x}] [\sigma_{x}] m    [0,y,z ] [{\bar x},y,z]
[2_{y}] [C_{2y}] [C_{2y}] 2   [0,y,0] [{\bar x},y,{\bar z}] [m_{y}] [\sigma_{y}] [\sigma_{y}] m    [x,0,z ] [x,{\bar y},z]
[2_{xy}] [C_{2a}] [C_{2a}^{\prime}] 2   [x,x,0] [y,x,{\bar z}] [m_{xy}] [\sigma_{da}] [\sigma_{d1}] m    [x,{\bar x},z ] [{\bar y},{\bar x},z]
[2_{x{\bar y}}] [C_{2b}] [C_{2b}^{\prime}] 2   [x,{\bar x},0 ] [{\bar y},{\bar x},{\bar z}] [m_{x{\bar y}}] [\sigma_{db}] [\sigma_{d2}] m    [x,x,z ] [y,x,z]
[2_{zx}] [C_{2c}] [C_{2c}^{\prime}] 2   [x,0,x,] [z,{\bar y},x] [m_{zx}] [\sigma_{dc}] [\sigma_{d3}] m    [{\bar x},y,x, ] [{\bar z},y,{\bar x}]
[2_{z{\bar x}}] [C_{2e}] [C_{2e}^{\prime}] 2   [{\bar x},0,x ] [{\bar z},{\bar y},{\bar x}] [m_{z{\bar x}}] [\sigma_{de}] [\sigma_{d5}] m    [x,y,x ] [z,y,x]
[2_{yz}] [C_{2d}] [C_{2d}^{\prime}] 2   [0,y,y] [{\bar x},z,y] [m_{yz}] [\sigma_{dd}] [\sigma_{d4}] m    [x,y,{\bar y} ] [x,{\bar z},{\bar y}]
[2_{y{\bar z}}] [C_{2f}] [C_{2f}^{\prime}] [2]   [0,y,{\bar y} ] [{\bar x},{\bar z},{\bar y}] [m_{y{\bar z}}] [\sigma_{df}] [\sigma_{d6}] m    [x,y,y ] [x,z,y]
[3_{p}] [C_{31}^{+}] [C_{31}^{+}] [3^{+}]   [x,x,x] [z,x,y] [{\bar 3}_{p}] [S_{61}^{-}] [S_{61}^{-}] [{\bar 3}^{+}]   [x,x,x] [{\bar z},{\bar x},{\bar y}]
[3_{q}] [C_{32}^{+}] [C_{32}^{+}] [3^{+}]   [{\bar x},{\bar x},x] [{\bar z},x,{\bar y}] [{\bar 3}_{q}] [S_{62}^{-}] [S_{62}^{-}] [{\bar 3}^{+}]   [{\bar x},{\bar x},x] [z,{\bar x},y]
[3_{r}] [C_{33}^{+}] [C_{33}^{+}] [3^{+}]   [x,{\bar x},{\bar x}] [{\bar z},{\bar x},y] [{\bar 3}_{r}] [S_{63}^{-}] [S_{63}^{-}] [{\bar 3}^{+}]   [x,{\bar x},{\bar x}] [z,x,{\bar y}]
[3_{s}] [C_{34}^{+}] [C_{34}^{+}] [3^{+}]   [{\bar x},x,{\bar x}] [z,{\bar x},{\bar y}] [{\bar 3}_{s}] [S_{64}^{-}] [S_{64}^{-}] [{\bar 3}^{+}]   [{\bar x},x,{\bar x}] [{\bar z},x,y]
[3_{p}^{2}] [C_{31}^{-}] [C_{31}^{-}] [3^{-}]   [x,x,x] [y,z,x] [{\bar 3}_{p}^{5}] [S_{61}^{+}] [S_{61}^{+}] [{\bar 3}^{-}]   [x,x,x] [{\bar y},{\bar z},{\bar x}]
[3_{q}^{2}] [C_{32}^{-}] [C_{32}^{-}] [3^{-}]   [{\bar x},{\bar x},x] [y,{\bar z},{\bar x}] [{\bar 3}_{q}^{5}] [S_{62}^{+}] [S_{62}^{+}] [{\bar 3}^{-}]   [{\bar x},{\bar x},x] [{\bar y},z,x]
[3_{r}^{2}] [C_{33}^{-}] [C_{33}^{-}] [3^{-}]   [x,{\bar x},{\bar x}] [{\bar y},z,{\bar x}] [{\bar 3}_{r}^{5}] [S_{63}^{+}] [S_{63}^{+}] [{\bar 3}^{-}]   [x,{\bar x},{\bar x}] [y,{\bar z},x]
[3_{s}^{2}] [C_{34}^{-}] [C_{34}^{-}] [3^{-}]   [{\bar x},x,{\bar x}] [{\bar y},{\bar z},x] [{\bar 3}_{s}^{5}] [S_{64}^{+}] [S_{64}^{+}] [{\bar 3}^{-}]   [{\bar x},x,{\bar x}] [y,z,{\bar x}]
[4_{z}] [C_{4z}^{+}] [C_{4z}^{+}] [4^{+}]   [0,0,z] [{\bar y},x,z] [{\bar 4}_{z}] [S_{4z}^{-}] [S_{4z}^{-}] [{\bar 4}^{+}]   [0,0,z] [y,{\bar x},{\bar z}]
[4_{x}] [C_{4x}^{+}] [C_{4x}^{+}] [4^{+}]   [x,0,0] [x,{\bar z},y] [{\bar 4}_{x}] [S_{4x}^{-}] [S_{4x}^{-}] [{\bar 4}^{+}]   [x,0,0] [{\bar x},z,{\bar y}]
[4_{y}] [C_{4y}^{+}] [C_{4y}^{+}] [4^{+}]   [0,y,0] [z,y,{\bar x}] [{\bar 4}_{y}] [S_{4y}^{-}] [S_{4y}^{-}] [{\bar 4}^{+}]   [0,y,0] [{\bar z},{\bar y},x]
[4_{z}^{3}] [C_{4z}^{-}] [C_{4z}^{-}] [4^{-}]   [0,0,z] [y,{\bar x},z] [{\bar 4}_{z}^{3}] [S_{4z}^{+}] [S_{4z}^{+}] [{\bar 4}^{-}]   [0,0,z] [{\bar y},x,{\bar z}]
[4_{x}^{3}] [C_{4x}^{-}] [C_{4x}^{-}] [4^{-}]   [x,0,0] [x,z,{\bar y}] [{\bar 4}_{x}^{3}] [S_{4x}^{+}] [S_{4x}^{+}] [{\bar 4}^{-}]   [x,0,0] [{\bar x},{\bar z},y]
[4_{y}^{3}] [C_{4y}^{-}] [C_{4y}^{-}] [4^{-}]   [0,y,0] [{\bar z},y,x] [{\bar 4}_{y}^{3}] [S_{4y}^{+}] [S_{4y}^{+}] [{\bar 4}^{-}]   [0,y,0] [z,{\bar y},{\bar x}]

Table 3.4.2.6 | top | pdf |
Symbols of symmetry operations of the point group [6/mmm]

Standard: symbols used in Section 3.1.3[link] , in the present chapter and in the software; suffixes (in italic) refer to the Cartesian crystallophysical coordinate system. BC: Bradley & Cracknell (1972[link]). AH: Altmann & Herzig (1994[link]). IT A: IT A (2005[link]), coordinates (in Sans Serif) are expressed in a crystallographic hexagonal basis. Jones: Jones' faithful representation symbols express the action of a symmetry operation of a vector [({\sf xyz})] in a crystallographic basis (see e.g. Bradley & Cracknell, 1972[link]).

Standard BC AH IT A Jones Standard BC AH IT A Jones
1 or e E E [{\sf 1}] [{\sf x,y,z}] [\bar 1] or i I I [{\bar{\sf 1}}]   [{\sf 0,0,0}] [\bar{\sf x},\bar{\sf y},\bar{\sf z}]
[6_{ z}] [C_{6}^{+}] [C_{6}^{+}] [{\sf 6^{+}}]   [{\sf 0,0,z}] [{\sf x-y,x,z}] [{\bar 6}_{ z}] [S_{3}^{-}] [S_{3}^{-}] [{\bar{\sf 6}^{+}}]   [{\sf 0,0,z}] [{\sf y-x},\bar{\sf x},\bar{\sf z}]
[3_{ z}] [C_{3}^{+}] [C_{3}^{+}] [{\sf 3^{+}}]   [{\sf 0,0,z}] [{\bar {\sf y}},{\sf x-y},{\sf z}] [{\bar 3}_{ z}] [S_{6}^{-}] [S_{6}^{-}] [{\bar {\sf 3}^{+}}]   [{\sf 0,0,z}] [{\sf y,y-x,}\bar{\sf z}]
[2_{ z}] [C_{2}] [C_{2}] [{\sf 2}]   [{\sf 0,0,z}] [\bar{\sf x},\bar{\sf y},{\sf z}] [m_{ z}] [\sigma_{h}] [\sigma_{h}] [{\sf m}]   [{\sf x,y,0}] [{\sf x},{\sf y},\bar{\sf z}]
[3_{ z}^{2}] [C_{3}^{-}] [C_{3}^{-}] [{\sf 3^{-}}]   [{\sf 0,0,z}] [{\sf y-x},\bar{\sf x},{\sf z}] [{\bar 3}_{ z}^{5}] [S_{6}^{+}] [S_{6}^{+}] [{\bar {\sf 3}^{-}}]   [{\sf 0,0,z}] [{\sf x-y},{\sf x},\bar{\sf z}]
[6_{ z}^{5}] [C_{6}^{-}] [C_{6}^{-}] [{\sf 6^{-}}]   [{\sf 0,0,z}] [{\sf y,y-x,z}] [{\bar 6}_{ z}^{5}] [S_{3}^{+}] [S_{3}^{+}] [\bar{\sf 6}^{-}]   [{\sf 0,0,z}] [\bar{\sf y},{\sf x-y},\bar{\sf z}]
[2_{ x}] [C_{21}{^\prime}{^\prime}] [C_{21}{^\prime}{^\prime}] [{\sf 2}]   [{\sf x,0,0}] [{\sf x-y},\bar{\sf y},\bar{\sf z}] [m_{ x}] [\sigma_{v1}] [\sigma_{v1}] [{\sf m}]   [{\sf x,2x,z}] [{\sf y-x,y,z}]
[2_{x^\prime}] [C_{22}{^\prime}{^\prime}] [C_{22}{^\prime}{^\prime}] [{\sf 2}]   [{\sf 0,y,0}] [\bar{\sf x},{\sf y-x},\bar{\sf z}] [m_{x^\prime}] [\sigma_{v2}] [\sigma_{v2}] [{\sf m}]   [{\sf 2x,x,z}] [{\sf x,x-y,z}]
[2_{x{^\prime}{^\prime}}] [C_{23}{^\prime}{^\prime}] [C_{23}{^\prime}{^\prime}] [{\sf 2}]   [{\sf x,x,0}] [{\sf y},{\sf x},\bar{\sf z}] [m_{x{^\prime}{^\prime}}] [\sigma_{v3}] [\sigma_{v3}] [{\sf m}]   [{\sf x},\bar{\sf x},{\sf z}] [\bar{\sf y},\bar{\sf x},{\sf z}]
[2_{y}] [C_{21}{^\prime}] [C_{21}{^\prime}] [{\sf 2}]   [{\sf x,2x,0}] [{\sf y-x},{\sf y},\bar{\sf z}] [m_{y}] [\sigma_{d1}] [\sigma_{d1}] [{\sf m}]   [{\sf x,0,z}] [{\sf x-y},\bar{\sf y},{\sf z}]
[2_{y{^\prime}}] [C_{22}{^\prime}] [C_{22}{^\prime}] [{\sf 2}]   [{\sf 2x,x,0}] [{\sf x,x-y},\bar{\sf z}] [m_{y{^\prime}}] [\sigma_{d2}] [\sigma_{d2}] [{\sf m}]   [{\sf 0,y,z}] [\bar{\sf x},{\sf y-x,z}]
[2_{y{^\prime}{^\prime}}] [C_{23}{^\prime}] [C_{23}{^\prime}] [{\sf 2}]   [{\sf x},\bar{\sf x},{\sf 0}] [\bar{\sf y},\bar{\sf x},\bar{\sf z}] [m_{y{^\prime}{^\prime}}] [\sigma_{d3}] [\sigma_{d3}] [{\sf m}]   [{\sf x,x,z}] [{\sf y,x,z}]
[Figure 3.4.2.3]

Figure 3.4.2.3 | top | pdf |

Oriented symmetry operations of the cubic group [m\bar3m] and of its subgroups. The Cartesian (rectangular) coordinate system [x, y, z ] is identical with the crystallographic and crystallophysical coordinate systems. Correlation with other notations is given in Table 3.4.2.5[link].

[Figure 3.4.2.4]

Figure 3.4.2.4 | top | pdf |

Oriented symmetry operations of the hexagonal group [6/mmm ] and of its hexagonal and trigonal subgroups. The coordinate system [x, y, z] corresponds to the Cartesian crystallophysical coordinate system, the axes [{\sf x, y, z}] of the crystallographic coordinate system are parallel to the twofold rotation axes [2_x, 2_{x^\prime}] and to the sixfold rotation axis [6_z]. Correlation with other notations is given in Table 3.4.2.6[link].

Now we can present the synoptic Table 3.4.2.7[link].

3.4.2.4.1. Explanation of Table 3.4.2.7[link]

| top | pdf |

  • G : point group expressing the symmetry of the parent (prototypic) phase. Subscripts of generators in the group symbol specify their orientation in the Cartesian (rectangular) crystallophysical coordinate system of the group G (see Tables 3.4.2.5[link] and 3.4.2.6[link], and Figs. 3.4.2.3 [link]and 3.4.2.4[link]).

    Table 3.4.2.7 | top | pdf |
    Group–subgroup symmetry descents [G \supset F_1]

    G : point-group symmetry of parent phase; [F_1]: point-group symmetry of single-domain state [{\bf S}_1]; [\Gamma_{\eta} ]: representation of G; [N_{G}(F_1)]: normalizer of [F_1] in G; [K_{1j}]: twinning groups; n: number of principal single-domain states; [d_F]: number of principal domain states with the same symmetry; [n_e]: number of ferroelectric single-domain states; [n_a]: number of ferroelastic single-domain states.

    G [F_1] [\Gamma_{\eta}] [N_{G}(F_1)] [K_{1j}] n [d_F ] [n_e ] [n_a ]
    [ {\bar {\bf 1}}] 1 [A_u] [\bar1] [\bar1^{\star}] 2 2 2 1
    [{\bf 2}_{\bi u}] 1 B [2_u] [2_u^\star] 2 2 2 2
    [{\bi m_u}] 1 [A{^\prime}{^\prime}] [m_u] [m_u^\star] 2 2 2 2
    [{\bf 2}_{\bi u}/{\bi m_u} ] [m_u] [B_u] [2_u/m_u] [2_u^\star/m_u] 2 2 2 1
    [2_u] [A_u] [2_u/m_u] [2_u/m_u^\star] 2 2 2 1
    [\bar 1 ] [B_g ] [2_u/m_u] [2_u^\star/m_u^\star] 2 2 0 2
    1 Reducible [2_u/m_u] [m_u^{\star}], [ 2_u^{\star}], [\bar 1^{\star}] 4 4 4 2
    [{\bf 2}_{\bi x}{\bf 2}_{\bi y}{\bf 2}_{\bi z} ] [2_z] [B_{1g} ] [2_x2_y2_z ] [2_x^\star2_y^\star2_z ] 2 2 2 2
    [2_x] [B_{3g} ] [2_x2_y2_z] [2_x2_y^\star2_z^\star ] 2 2 2 2
    [2_y] [B_{2g} ] [2_x2_y2_z] [2_x^\star2_y2_z^\star] 2 2 2 2
    1 Reducible [2_x2_y2_z] [2_z^{\star}], [2_x^{\star}], [2_y^{\star}] 4 4 4 4
    [{\bi m_xm_y}{\bf 2}_{\bi z} ] [m_x] [B_2] [m_xm_y2_z] [m_xm_y^{\star}2_z^{\star}] 2 2 2 2
    [m_y ] [B_1] [m_xm_y2_z] [m_x^{\star}m_y2_z^{\star}] 2 2 2 2
    [2_z] [A_2] [m_xm_y2_z] [m_x^{\star}m_y^{\star}2_z] 2 2 1 2
    1 Reducible [m_xm_y2_z] [m_x^{\star}], [m_y^{\star}], [2_z^{\star} ] 4 2 4 4
    [{\bi m_xm_ym_z}] [m_xm_y2_z ] [B_{1u} ] [m_xm_ym_z] [m_xm_ym_z^\star] 2 2 2 1
    [2_xm_ym_z ] [B_{3u} ] [m_xm_ym_z] [m_x^{\star}m_ym_z] 2 2 2 1
    [m_x2_ym_z ] [B_{2u} ] [m_xm_ym_z] [m_xm_y^{\star}m_z] 2 2 2 1
    [2_x2_y2_z ] [A_{1u} ] [m_xm_ym_z] [m_x^{\star}m_y^{\star}m_z^{\star}] 2 2 0 1
    [2_z/m_z] [B_{1g} ] [m_xm_ym_z] [m_x^{\star}m_y^{\star}m_z ] 2 2 0 2
    [2_x/m_x] [B_{3g} ] [m_xm_ym_z] [m_xm_y^{\star}m_z^{\star}] 2 2 0 2
    [2_y/m_y] [B_{2g} ] [m_xm_ym_z] [m_x^{\star}m_ym_z^{\star}] 2 2 0 2
    [m_z ] Reducible [m_xm_ym_z] [2_x^{\star}m_y^{\star}m_z], [m_x^{\star}2_y^{\star}m_z], [2_z^{\star}/m_z] 4 4 4 2
    [m_x ] Reducible [m_xm_ym_z] [m_xm_y^{\star}2_z^{\star}], [m_x2_y^{\star}m_z^{\star}], [2_x^{\star}/m_x] 4 4 4 2
    [m_y ] Reducible [m_xm_ym_z] [m_x^{\star}m_y2_z^{\star}], [2_x^{\star}m_ym_z^{\star}], [2_y^{\star}/m_y] 4 4 4 2
    [2_z ] Reducible [m_xm_ym_z] [m_x^{\star}m_y^{\star}2_z], [2_x^{\star}2_y^{\star}2_z], [2_z/m_z^{\star}] 4 4 2 2
    [2_x ] Reducible [m_xm_ym_z] [2_xm_y^{\star}m_z^{\star}], [2_x2_y^{\star}2_z^{\star}], [2_x/m_x^{\star}] 4 4 2 2
    [2_y ] Reducible [m_xm_ym_z] [m_x^{\star}2_ym_z^{\star}], [2_x^{\star}2_y2_z^{\star}], [2_y/m_y^{\star}] 4 4 2 2
    [\bar 1] Reducible [m_xm_ym_z] [2_z^{\star}/m_z^{\star}], [2_x^{\star}/m_x^{\star} ], [2_y^{\star}/m_y^{\star}] 4 4 0 4
    1 Reducible [m_xm_ym_z] [m_z^{\star}], [m_x^{\star}], [m_y^{\star}], [2_z^{\star}], [2_x^{\star}], [2_y^{\star} ], [\bar 1^{\star}] 8 8 8 4
    [{\bf 4}_{\bi z} ] [2_z ] B [4_z] [4_z^{\star} ] 2 2 1 2
    1 [^1E\oplus ^2E] [4_z] [4_z, 2_z^{\star}] 4 4 4 4
    [{\bar {\bf 4}}_{\bi z} ] [2_z ] B [\bar4_z ] [\bar 4_z^{\star}] 2 2 2 2
    1 [^1E\oplus ^2E ] [\bar4_z ] [\bar 4_z,] [2_z^{\star}] 4 2 4 4
    [{\bf 4}_{\bi z}/{\bi m_z} ] [\bar4_z] [B_u] [4_z/m_z ] [4_z^{\star}/m_z^{\star} ] 2 2 0 1
    [4_z ] [A_u] [4_z/m_z ] [4_z/m_z^{\star} ] 2 2 2 1
    [2_z/m_z ] [B_g] [4_z/m_z ] [4_z^{\star}/m_z ] 2 2 0 2
    [m_z] [^1E_u \oplus ^2E_u] [4_z/m_z] [4_z/m_z], [2_z^{\star}/m_z] 4 4 4 2
    [2_z] Reducible [4_z/m_z] [\bar 4_z^{\star}], [4_z^{\star} ], [2_z/m_z^{\star}] 4 4 2 2
    [\bar 1] [^1E_g \oplus ^2E_g ] [4_z/m_z] [4_z/m_z], [2_z^{\star}/m_z^{\star} ] 4 4 0 4
    1 Reducible [4_z/m_z] [\bar4_z], [4_z], [m_z^{\star} ], [2_z^{\star}], [\bar1^{\star}] 8 8 8 4
    [{\bf 4}_{\bi z}{\bf 2}_{\bi x}{\bf 2}_{\bi xy} ] [4_z] [A_2] [4_z2_x2_{xy}] [4_z2_x^{\star}2_{xy}^{\star}] 2 2 2 1
    [2_{x\bar{y}}2_{xy}2_z] [B_2] [4_z2_x2_{xy}] [4_z^{\star}2_x^{\star}2_{xy}] 2 2 0 2
    [2_x2_y2_z ] [B_1] [4_z2_x2_{xy} ] [4_z^{\star}2_x2_{xy}^{\star}] 2 2 0 2
    [2_{xy}] [(2_{x\bar{y}})] E [2_{x\bar{y}}2_{xy}2_z ] [4_z2_x2_{xy}], [2_{x\bar{y}}^{\star}2_{xy}2_z^{\star} ] 4 2 2 2
    [2_z ] Reducible [4_z2_x2_{xy}] [4_z^{\star}], [2_x^{\star}2_y^{\star}2_z ], [2_{x\bar{y}}^{\star}2_{xy}^{\star}2_z] 4 4 2 2
    [2_x] [(2_y)] E [2_{x\bar{y}}2_{xy}2_z] [4_z2_x2_{xy}], [2_x2_y^{\star}2_z^{\star} ] 4 2 2 2
    1 E [4_z2_x2_{xy}] [4_z], [2_z^{\star}], [2_x^{\star}(2)], [2_{xy}^{\star}(2)] 8 8 8 8
    [{\bf 4}_{\bi z}{\bi m}_{\bi x}{\bi m}_{\bi xy} ] [4_z ] [A_2] [4_zm_xm_{xy}] [4_zm_x^{\star}m_{xy}^{\star} ] 2 2 1 1
    [m_{x\bar{y}}m_{xy}2_z] [B_2] [4_zm_xm_{xy} ] [4_z^{\star}m_x^{\star}m_{xy} ] 2 2 1 2
    [m_xm_y2_z ] [B_1] [4_zm_xm_{xy} ] [4_z^{\star}m_xm_{xy}^{\star} ] 2 2 1 2
    [m_{xy}] [(m_{x\bar{y}})] E [m_{x\bar{y}}m_{xy}2_z] [4_zm_xm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}2_z^{\star} ] 4 2 4 4
    [m_x] [(m_y)] E [m_xm_y2_z ] [4_zm_xm_{xy}], [m_xm_y^{\star}2_z^{\star} ] 4 2 4 4
    [2_z] Reducible [4_zm_xm_{xy} ] [4_z^{\star}], [m_x^{\star}m_y^{\star}2_z ], [m_{x\bar{y}}^{\star}m_{xy}^{\star}2_z] 4 4 2 2
    1 E [4_zm_xm_{xy} ] [4_z], [m_x^{\star}(2)], [m_{xy}^{\star}(2)], [2_z^{\star}] 8 8 8 8
    [{\bar {\bf 4}}_{\bi z}{\bf 2}_{\bi x}{\bi m_{xy}} ] [\bar4_z] [A_2] [\bar4_z2_xm_{xy}] [\bar4_z2_x^{\star}m_{xy}^{\star}] 2 2 0 1
    [m_{x\bar{y}}m_{xy}2_z] [B_2] [\bar4_z2_xm_{xy}] [\bar4_z^{\star}2_x^{\star}m_{xy}] 2 2 2 2
    [2_x2_y2_z ] [B_1] [\bar4_z2_xm_{xy}] [\bar4_z^{\star}2_xm_{xy}^{\star}] 2 2 0 2
    [m_{xy}] [(m_{x\bar{y}})] E [m_{x\bar{y}}m_{xy}2_z] [\bar4_z2_xm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}2_z^{\star} ] 4 2 4 4
    [2_z ] Reducible [\bar4_z2_xm_{xy}] [\bar4_z^{\star}], [m_{x\bar{y}}^{\star}m_{xy}^{\star}2_z ], [2_x^{\star}2_y^{\star}2_z] 4 4 2 2
    [2_x ] [(2_y) ] E [2_x2_y2_z ] [\bar4_z2_xm_{xy}], [2_x2_y^{\star}2_z^{\star} ] 4 2 4 4
    1 E [\bar4_z2_xm_{xy}] [\bar4_z], [m_{xy}^{\star}(2)], [2_z^{\star}], [2_x^{\star}(2)] 8 8 8 8
    [\bar4_zm_x2_{xy}] [\bar4_z] [A_2] [\bar4_zm_x2_{xy}] [\bar4_zm_x^{\star}2_{xy}^{\star}] 2 2 0 1
    [m_xm_y2_z ] [B_2] [\bar4_zm_x2_{xy}] [\bar4_z^{\star}m_x2_{xy}^{\star}] 2 2 2 2
    [2_{x\bar{y}}2_{xy}2_z] [B_1] [\bar4_zm_x2_{xy}] [\bar4_z^{\star}m_x^{\star}2_{xy}] 2 2 0 2
    [m_x] [(m_y)] E [m_xm_y2_z ] [\bar4_zm_x2_{xy}], [m_xm_y^{\star}2_z^{\star} ] 4 2 4 4
    [2_{xy}] [(2_{x\bar{y}})] E [2_{x\bar{y}}2_{xy}2_z] [\bar4_zm_x2_{xy}], [2_{x\bar{y}}^{\star}2_{xy}2_z^{\star} ] 4 2 4 4
    [2_z ] Reducible [\bar4_zm_x2_{xy}] [\bar4_z^{\star} ], [m_x^{\star}m_y^{\star}2_z ], [2_{x\bar{y}}^{\star}2_{xy}^{\star}2_z] 4 4 2 2
    1 E [\bar4_zm_x2_{xy}] [\bar4_z], [m_x^{\star}(2)], [2_{xy}^{\star}(2)], [2_z^{\star}] 8 8 8 8
    [{\bf 4}_{\bi z}/{\bi m_zm_xm_{xy}} ] [\bar4_zm_x2_{xy}] [B_{2u}] [4_z/m_zm_xm_{xy} ] [4_z^{\star}/m_z^{\star}m_xm_{xy}^{\star} ] 2 2 0 1
    [\bar4_z2_xm_{xy} ] [B_{1u} ] [4_z/m_zm_xm_{xy} ] [4_z^{\star}/m_z^{\star}m_x^{\star}m_{xy} ] 2 2 0 1
    [4_zm_xm_{xy} ] [A_{2u} ] [4_z/m_zm_xm_{xy} ] [4_z^{\star}/m_z^{\star}m_xm_{xy}] 2 2 2 1
    [4_z2_x2_{xy} ] [A_{1u} ] [4_z/m_zm_xm_{xy} ] [4_z^{\star}/m_z^{\star}m_x^{\star}m_{xy} ] 2 2 0 1
    [4_z/m_z ] [A_{2g} ] [4_z/m_zm_xm_{xy} ] [4_z^{\star}/m_zm_x^{\star}m_{xy}^{\star} ] 2 2 0 1
    [\bar4_z ] Reducible [4_z/m_zm_xm_{xy} ] [\bar4_z2_x^{\star}m_{xy}^{\star}], [\bar4_zm_x^{\star}2_{xy}^{\star}], [4_z^{\star}/m_z^{\star}] 4 4 0 1
    [4_z] Reducible [4_z/m_zm_xm_{xy} ] [4_zm_x^{\star}m_{xy}^{\star} ], [4_z2_x^{\star}2_{xy}^{\star}], [4_z/m_z^{\star}] 4 4 2 1
    [m_{x\bar{y}}m_{xy}m_z ] [B_{2g} ] [4_z/m_zm_xm_{xy}] [4_z^{\star}/m_zm_x^{\star}m_{xy}] 2 2 0 2
    [m_xm_ym_z] [B_{1g} ] [m_xm_ym_z ] [4_z^{\star}/m_zm_xm_{xy}^{\star}] 2 2 0 2
    [2_{x\bar{y}}m_{xy}m_z] [(m_{x\bar{y}}2_{xy}m_z)] [E_u] [m_{x\bar{y}}m_{xy}m_z ] [4_z/m_zm_xm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}m_z ] 4 2 4 2
    [2_xm_ym_z] [(m_x2_ym_z)] [E_u] [m_xm_ym_z] [4_z/m_zm_xm_{xy}], [m_x^{\star}m_ym_z ] 4 2 4 2
    [m_{x\bar{y}}m_{xy}2_z ] Reducible [4_z/m_zm_xm_{xy}] [\bar4_z^{\star}2_x^{\star}m_{xy}], [4_z^{\star}m_x^{\star}m_{xy}], [m_{x\bar{y}}m_{xy}m_z^{\star} ] 4 4 2 2
    [m_xm_y2_z] Reducible [4_z/m_zm_xm_{xy}] [\bar4_z^{\star}m_x2_{xy}^{\star}], [4_z^{\star}m_xm_{xy}^{\star}], [m_xm_ym_z^{\star} ] 4 4 2 2
    [2_{x\bar{y}}2_{xy}2_z ] Reducible [4_z/m_zm_xm_{xy}] [\bar4_z^{\star}m_x^{\star}2_{xy}], [4_z^{\star}2_x^{\star}2_{xy}], [m_{x\bar{y}}^{\star}m_{xy}^{\star}m_z^{\star} ] 4 4 0 2
    [2_x2_y2_z] Reducible [4_z/m_zm_xm_{xy} ] [\bar4_z^{\star}2_xm_{xy}^{\star}], [4_z^{\star}2_x2_{xy}^{\star}], [m_x^{\star}m_y^{\star}m_z^{\star}] 4 4 0 2
    [2_{xy}/m_{xy}] [(2_{x\bar{y}}/m_{x\bar{y}})] [E_g] [m_{x\bar{y}}m_{xy}m_z ] [4_z/m_zm_xm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}m_z^{\star} ] 4 2 0 4
    [2_z/m_z ] Reducible [4_z/m_zm_xm_{xy}] [4_z^{\star}/m_z], [m_{x\bar{y}}^{\star}m_{xy}^{\star}m_z ], [m_x^{\star}m_y^{\star}m_z ] 4 4 0 4
    [2_x/m_x] [(2_y/m_y)] [E_g] [m_xm_ym_z] [4_z/m_zm_xm_{xy}], [m_xm_y^{\star}m_z^{\star} ] 4 2 0 4
    [m_{xy} (m_{x \bar y})] Reducible [m_{x\bar{y}}m_{xy}m_z ] [\bar4_z2_xm_{xy}], [4_zm_xm_{xy} ], [2_{x\bar{y}}^{\star}m_{xy}m_z^{\star}], [m_{x\bar{y}}^{\star}m_{xy}2_z^{\star} ], [2_{xy}^{\star}/m_{xy}] 8 4 8 4
    [m_z ] Reducible [4_z/m_zm_xm_{xy} ] [4_z/m_z], [2_{x\bar{y}}^{\star}m_{xy}^{\star}m_z(2) ], [2_x^{\star}m_y^{\star}m_z(2)], [2_z^{\star}/m_z ] 8 8 8 4
    [m_x] [(m_y) ] Reducible [m_xm_ym_z] [\bar4_zm_x2_{xy}], [4_zm_xm_{xy} ], [m_xm_y^{\star}2_z^{\star}], [m_x2_y^{\star}m_z^{\star}], [2_x^{\star}/m_x] 8 4 8 4
    [2_{xy}] [(2_{x\bar{y}})] Reducible [m_{x\bar{y}}m_{xy}m_z ] [\bar4_zm_x2_{xy}], [4_z2_x2_{xy} ], [m_{x\bar{y}}^{\star}2_{xy}m_z^{\star}], [2_{x\bar{y}}^{\star}2_{xy}2_z^{\star} ], [2_{xy}/m_{xy}^{\star}] 8 4 8 4
    [2_z] Reducible [4_z/m_zm_xm_{xy}] [\bar4_z^{\star}], [4_z^{\star} ], [m_x^{\star}m_y^{\star}2_z], [m_{x\bar{y}}^{\star}m_{xy}^{\star}2_z ], [2_x^{\star}2_y^{\star}2_z], [2_{x\bar{y}}^{\star}2_{xy}^{\star}2_z ], [2_z/m_z^{\star}] 8 8 2 4
    [2_x] [(2_y)] Reducible [m_xm_ym_z] [\bar4_z2_xm_{xy}], [4_z2_x2_{xy} ], [2_xm_y^{\star}m_z^{\star}], [2_x2_y^{\star}2_z^{\star}], [2_x/m_x^{\star}] 8 4 4 4
    [\bar1] [E_g] [4_z/m_zm_xm_{xy} ] [4_z/m_z], [2_{xy}^{\star}/m_{xy}^{\star}(2) ], [2_z^{\star}/m_z^{\star}], [2_x^{\star}/m_x^{\star}(2)] 8 8 0 8
    1 Reducible [4_z/m_zm_xm_{xy}] [\bar4_z], [4_z], [m_{xy}^{\star}(2) ], [m_z^{\star}], [m_x^{\star}(2)], [2_{xy}^{\star}(2) ], [2_z^{\star}], [2_x^{\star}(2)], [\bar1^{\star}] 16 16 16 8
    [{\bf 3}_{\bi z} ] 1 E [3_z ] [3_z ] 3 3 3 3
    [{\bar {\bf 3}}_{\bi z} ] [3_z ] [A_u] [\bar3_z ] [\bar 3_z^{\star} ] 2 2 2 1
    [\bar 1 ] [E_g] [\bar3_z ] [\bar 3_z] 3 3 0 3
    1 [E_u] [\bar3_z ] [\bar 3_z], [3_z], [\bar 1^{\star} ] 6 6 6 3
    [{\bf 3}_{\bi z}{\bf 2}_{\bi x} ] [3_z] [A_2] [3_z2_x] [3_z2_x^{\star}] 2 2 2 1
    [2_x] [(2_{x^{\prime}}], [2_{x^{{\prime}{\prime}}})] E [2_x ] [3_z2_x ] 3 1 3 3
    1 E [3_z2_x] [3_z, 2_x^{\star}(3) ] 6 6 6 6
    [3_z2_y ] [3_z] [A_2] [3_z2_y] [3_z2_y^{\star}] 2 2 2 1
    [2_y] [(2_{y^{\prime}}], [2_{y^{{\prime}{\prime}}})] E [2_y] [3_z2_y ] 3 1 3 3
    1 E [3_z2_y] [3_z], [2_y^{\star}(3)] 6 6 6 6
    [{\bf 3}_{\bi z}{\bi m_x} ] [3_z] [A_2] [3_zm_x] [3_zm_x^{\star}] 2 2 1 1
    [m_x] [(m_{x^\prime}], [m_{x{^\prime}{^\prime}})] E [m_x ] [3_zm_x ] 3 1 3 3
    1 E [3_zm_x] [3_z], [m_x^{\star}(3)] 6 6 6 6
    [3_zm_y] [3_z] [A_2] [3_zm_y] [3_zm_y^{\star}] 2 2 1 1
    [m_y ] [(m_{y{^\prime}}], [m_{y{^\prime}{^\prime}})] E [m_y] [3_zm_y ] 3 1 3 3
    1 E [3_zm_y] [3_z], [m_y^{\star}(3)] 6 6 6 6
    [{\bar {\bf 3}}_{\bi z}{\bi m_x} ] [3_zm_x ] [A_{2u} ] [\bar3_zm_x ] [\bar3_z^{\star}m_x ] 2 2 2 1
    [3_z2_x ] [A_{1u} ] [\bar3_zm_x ] [\bar3_z^{\star}m_x^{\star}] 2 2 0 2
    [\bar 3_z ] [A_{2g} ] [\bar3_zm_x ] [\bar3_zm_x^{\star} ] 2 2 0 1
    [3_z ] Reducible [\bar3_zm_x] [3_zm_x^{\star}], [3_z2_x^{\star} ], [\bar 3_z^{\star} ] 4 4 2 1
    [2_x/m_x] [(2_{x^\prime}/m_{x^\prime} ], [2_{x{^\prime}{^\prime}}/m_{x{^\prime}{^\prime}})] [E_g] [2_x/m_x] [\bar3_zm_x ] 3 1 0 3
    [m_x] [(m_{x{^\prime}}], [m_{x{^\prime}{^\prime}})] [E_u] [2_x/m_x] [\bar3_zm_x], [3_zm_x], [2_x^{\star}/m_x(3)] 6 2 6 3
    [2_x ] [(2_{x{^\prime}}], [2_{x{^\prime}{^\prime}})] [E_u] [2_x/m_x] [\bar3_zm_x], [3_z2_x], [2_x/m_x^{\star}(3) ] 6 2 6 3
    [\bar 1] [E_g] [\bar3_zm_x] [\bar 3_z], [2_x^{\star}/m_x^{\star}(3) ] 6 6 0 6
    1 [E_u] [\bar3_zm_x] [\bar 3_z], [3_z], [m_x^{\star}(3) ], [2_x^{\star}(3)], [\bar 1^{\star} ] 12 12 12 6
    [\bar3_zm_y] [3_zm_y] [A_{2u}] [\bar3_zm_y] [\bar3_z^{\star}m_y] 2 2 2 1
    [3_z2_y] [A_{1u}] [\bar3_zm_y] [\bar3_z^{\star}m_y^{\star}] 2 2 0 1
    [\bar 3_z] [A_{2g}] [\bar3_zm_y] [\bar3_zm_y^{\star}] 2 2 0 1
    [3_z] Reducible [\bar3_zm_y] [3_zm_y^{\star}], [3_z2_y^{\star} ], [\bar 3_z^{\star}] 4 4 0 1
    [2_y/m_y] [(2_{y{^\prime}}/m_{y{^\prime}} ], [2_{y{^\prime}{^\prime}}/m_{y{^\prime}{^\prime}})] [E_g] [2_y/m_y] [\bar3_zm_y ] 3 1 2 1
    [m_y] [(m_{y{^\prime}}], [m_{y{^\prime}{^\prime}})] [E_u] [2_y/m_y] [\bar3_zm_y], [3_zm_y], [2_y^{\star}/m_y(3)] 6 2 0 3
    [2_y ] [(2_{y{^\prime}}], [2_{y{^\prime}{^\prime}}) ] [E_u] [2_y/m_y] [\bar3_zm_y], [3_z2_y], [2_y/m_y^{\star}(3) ] 6 2 6 3
    [\bar 1] [E_g] [\bar3_zm_y] [\bar 3_z], [2_y^{\star}/m_y^{\star}(3) ] 6 6 0 3
    1 [E_u] [\bar3_zm_y] [\bar 3_z], [3_z], [m_y^{\star}(3) ], [2_y^{\star}(3)], [\bar 1^{\star}] 12 12 12 6
    [{\bf 6}_{\bi z} ] [3_z] B [6_z ] [6_z^{\star}] 2 2 1 1
    [2_z] [E_2] [6_z] [6_z ] 3 3 1 3
    1 [E_1] [6_z] [6_z], [3_z], [2_z^{\star} ] 6 6 6 6
    [{\bar{\bf 6}}_{\bi z} ] [3_z] [A{^\prime}{^\prime}] [\bar6_z] [\bar 6_z^{\star}] 2 2 2 1
    [m_z] [E{^\prime}] [\bar6_z] [\bar 6_z] 3 2 3 3
    1 [E{^\prime}{^\prime}] [\bar6_z] [\bar 6_z ], [3_z], [m_z^{\star} ] 6 6 6 6
    [{\bf 6}_{\bi z}/{\bi m_z} ] [\bar 6_z] [B_u ] [6_z/m_z] [6_z^{\star}/m_z] 2 2 0 1
    [6_z ] [A_u ] [6_z/m_z] [6_z/m_z^{\star}] 2 2 2 1
    [\bar 3_z ] [B_g ] [6_z/m_z] [6_z^{\star}/m_z^{\star} ] 2 2 0 1
    [3_z ] Reducible [6_z/m_z] [\bar 6_z^{\star}], [6_z^{\star} ], [\bar 3_z^{\star}] 4 4 2 1
    [2_z/m_z] [E_{2g} ] [6_z/m_z] [6_z/m_z] 3 3 0 3
    [m_z ] [E_{1u} ] [6_z/m_z] [6_z/m_z], [\bar 6_z], [2_z^{\star}/m_z ] 6 6 6 3
    [2_z ] [E_{2u} ] [6_z/m_z] [6_z/m_z], [6_z], [2_z/m_z^{\star} ] 6 6 2 3
    [\bar 1 ] [E_{1g} ] [6_z/m_z] [6_z/m_z], [\bar 3_z], [2_z^{\star}/m_z^{\star} ] 6 6 0 6
    1 Reducible [6_z/m_z] [\bar 6_z], [6_z], [\bar 3_z ], [3_z], [m_z^{\star}], [2_z^{\star}], [\bar 1^{\star} ] 12 12 12 6
    [{\bf 6}_{\bi z}{\bf 2}_{\bi x}{\bf 2}_{\bi y}] [6_z] [A_2 ] [6_z2_x2_y] [6_z2_x^{\star}2_y^{\star}] 2 2 2 1
    [3_z2_x] [B_1 ] [6_z2_x2_y ] [6_z^{\star}2_x2_y^{\star}] 2 2 0 1
    [3_z2_y] [B_2 ] [6_z2_x2_y ] [6_z^{\star}2_x^{\star}2_y] 2 2 0 1
    [3_z ] Reducible [6_z2_x2_y] [6_z^{\star}], [3_z2_x^{\star} ], [3_z2_y^{\star} ] 4 4 2 1
    [2_x2_y2_z] [(2_{x{^\prime}}2_{y{^\prime}}2_z ], [2_{x{^\prime}{^\prime}}2_{y{^\prime}{^\prime}}2_z)] [E_2] [2_x2_y2_z] [6_z2_x2_y ] 3 1 0 3
    [2_z ] [E_2] [6_z 2_x2_y] [6_z], [2_x^{\star}2_y^{\star}2_z(3) ] 6 6 2 6
    [2_x] [(2_{x{^\prime}}], [2_{x{^\prime}{^\prime}})] [E_1] [2_x2_y2_z] [6_z2_x2_y], [3_z2_x], [2_x2_y^{\star}2_z^{\star}] 6 2 6 6
    [2_y] [(2_{y{^\prime}}], [2_{y{^\prime}{^\prime}})] [E_1] [2_x2_y2_z] [6_z2_x2_y], [3_z2_y], [2_x^{\star}2_y2_z^{\star} ] 6 2 6 6
    1 [E_1] [6_z2_x2_y] [6_z], [3_z], [2_z^{\star} ], [2_x^{\star}(3)], [2_y^{\star}(3)] 12 12 12 12
    [{\bf 6}_{\bi z}{\bi m}_{\bi x}{\bi m}_{\bi y} ] [6_z] [A_2] [6_zm_xm_y] [6_zm_x^{\star}m_y^{\star} ] 2 2 1 1
    [3_zm_x] [B_2] [6_zm_xm_y] [6_z^{\star}m_xm_y^{\star}] 2 2 1 1
    [3_zm_y] [B_1] [6_zm_xm_y] [6_z^{\star}m_x^{\star}m_y] 2 2 1 1
    [3_z ] Reducible [6_zm_xm_y] [6_z^{\star}], [3_zm_x^{\star} ], [3_zm_y^{\star}] 4 4 1 1
    [m_xm_y2_z] [(m_{x{^\prime}}m_{y{^\prime}}2_z ], [m_{x{^\prime}{^\prime}}m_{y{^\prime}{^\prime}}2_z)] [E_2] [m_xm_y2_z] [6_zm_xm_y] 3 1 1 3
    [m_x] [(m_{x{^\prime}}], [m_{x{^\prime}{^\prime}})] [E_1] [m_xm_y2_z] [6_zm_xm_y], [3_zm_x], [m_xm_y^{\star}2_z^{\star}] 6 2 6 6
    [m_y ] [(m_{y{^\prime}}], [m_{y{^\prime}{^\prime}})] [E_1] [m_xm_y2_z] [6_zm_xm_y], [3_zm_y], [m_x^{\star}m_y2_z^{\star}] 6 2 6 6
    [2_z ] [E_2] [6_zm_xm_y] [6_z], [m_x^{\star}m_y^{\star}2_z(3) ] 6 6 1 6
    1 [E_1] [6_zm_xm_y] [6_z], [3_z], [2_z^{\star} ], [m_x^{\star}(3)], [m_y^{\star}(3)] 12 12 12 12
    [{\bar {\bf 6}}_{\bi z}{\bi m_x}{\bf 2}_{\bi y} ] [\bar 6_z ] [A_2{^\prime} ] [\bar6_zm_x2_y] [\bar6_zm_x^{\star}2_y^{\star}] 2 2 0 1
    [3_zm_x ] [A_2{^\prime}{^\prime}] [\bar6_zm_x2_y] [\bar6_z^{\star}m_x2_y^{\star} ] 2 2 2 1
    [3_z2_y ] [A_1{^\prime}] [\bar6_zm_x2_y] [\bar6_z^{\star}m_x^{\star}2_y ] 2 2 0 1
    [3_z ] Reducible [\bar6_zm_x2_y] [\bar 6_z^{\star} ], [3_zm_x^{\star} ], [3_z2_y^{\star}] 4 4 2 1
    [m_x2_ym_z] [(m_{x{^\prime}}2_{y{^\prime}}m_z ], [m_{x{^\prime}{^\prime}}2_{y{^\prime}{^\prime}}m_z)] [E{^\prime}] [m_x2_ym_z] [\bar6_zm_x2_y] 3 1 3 3
    [m_z ] [E{^\prime}] [\bar6_zm_x2_y] [\bar6_z, m_x^{\star}2_y^{\star}m_z(3) ] 6 6 6 6
    [m_x] [(m_{x{^\prime}}], [m_{x{^\prime}{^\prime}})] [E{^\prime}{^\prime}] [m_y2_ym_z] [\bar6_zm_x2_y], [3_zm_x], [m_x2_y^{\star}m_z^{\star}] 6 2 6 6
    [2_y] [(2_{y{^\prime}}], [2_{y{^\prime}{^\prime}})] [E{^\prime}{^\prime}] [m_x2_ym_z] [\bar6_zm_x2_y], [3_z2_y], [m_x^{\star}2_ym_z^{\star}] 6 2 3 6
    1 [E{^\prime}{^\prime}] [\bar6_zm_x2_y] [\bar 6_z], [3_z], [m_z^{\star} ], [m_x^{\star}(3)], [2_y^{\star}(3)] 12 12 12 12
    [\bar 6_z2_xm_y ] [\bar 6_z ] [A_2{^\prime} ] [\bar6_z2_xm_y] [\bar6_z2_x^{\star}m_y^{\star} ] 2 2 0 1
    [3_zm_y] [A_2{^\prime} ] [\bar6_z2_xm_y] [\bar6_z^{\star}2_x^{\star}m_y] 2 2 2 1
    [3_z2_x] [A_1{^\prime}{^\prime}] [\bar6_z2_xm_y ] [\bar6_z^{\star}2_xm_y^{\star} ] 2 2 0 1
    [3_z] Reducible [\bar6_z2_xm_y ] [\bar 6_z^{\star}], [3_zm_y^{\star} ], [3_z2_x^{\star}] 4 4 2 1
    [2_xm_ym_z] [(2_{x{^\prime}}m_{y{^\prime}}m_z ], [2_{x{^\prime}{^\prime}}m_{y{^\prime}{^\prime}}m_z)] [E{^\prime}] [m_x2_ym_z] [\bar6_z2_xm_y] 3 1 3 3
    [m_z] [E{^\prime}] [\bar6_z2_xm_y ] [\bar6_z], [2_x^{\star}m_y^{\star}m_z(3) ] 6 6 6 6
    [m_y] [(m_{y{^\prime}}], [m_{y{^\prime}{^\prime}})] [E{^\prime}{^\prime}] [m_x2_ym_z] [\bar 6_z2_xm_y], [3_zm_y], [2_x^{\star}m_ym_z^{\star}] 6 2 6 6
    [2_x] [(2_{x{^\prime}}], [2_{x{^\prime}{^\prime}})] [E{^\prime}{^\prime}] [m_x2_ym_z] [\bar 6_z2_xm_y], [3_z2_x], [2_xm_y^{\star}m_z^{\star}] 6 2 3 6
    1 [E{^\prime}{^\prime}] [\bar6_z2_xm_y ] [\bar 6_z], [3_z], [m_z^{\star} ], [m_y^{\star}(3)], [2_x^{\star}(3)] 12 12 12 12
    [{\bf 6}_{\bi z}/{\bi m_zm_xm_y} ] [\bar 6_zm_x2_y] [B_{2u}] [6_z/m_zm_xm_y] [6_z^{\star}/m_zm_xm_y^{\star}] 2 2 0 1
    [\bar 6_z2_xm_y] [B_{1u} ] [6_z/m_zm_xm_y] [6_z^{\star}/m_zm_x^{\star}m_y] 2 2 0 1
    [6_zm_xm_y] [A_{2u} ] [6_z/m_zm_xm_y] [6_z/m_z^{\star}m_xm_y ] 2 2 2 1
    [6_z2_x2_y] [A_{1u} ] [6_z/m_zm_xm_y] [6_z/m_z^{\star}m_x^{\star}m_y^{\star} ] 2 2 0 1
    [6_z/m_z ] [A_{2g} ] [6_z/m_zm_xm_y] [6_z/m_zm_x^{\star}m_y^{\star}] 2 2 0 1
    [\bar 6_z] Reducible [6_z/m_zm_xm_y] [\bar 6_zm_x^{\star}2_y^{\star}], [\bar6_z2_x^{\star}m_y^{\star} ], [6_z^{\star}/m_z ] 4 4 0 1
    [6_z] Reducible [6_z/m_zm_xm_y] [6_zm_x^{\star}m_y^{\star}], [6_z2_x^{\star}2_y^{\star}], [6_z/m_z^{\star}] 4 4 2 1
    [\bar 3_zm_x] [B_{1g} ] [6_z/m_zm_xm_y] [6_z^{\star}/m_z^{\star}m_xm_y^{\star} ] 2 2 0 1
    [\bar 3_zm_y] [B_{2g}] [6_z/m_zm_xm_y] [6_z^{\star}/m_z^{\star}m_x^{\star}m_y ] 2 2 0 1
    [3_zm_x] Reducible [6_z/m_zm_xm_y] [\bar 6_z^{\star}m_x2_y^{\star}], [6_z^{\star}m_xm_y^{\star}], [\bar 3_z^{\star}m_x ] 4 4 2 1
    [3_zm_y] Reducible [6_z/m_zm_xm_y] [\bar 6_z^{\star}2_x^{\star}m_y], [6_z^{\star}m_x^{\star}m_y], [\bar 3_z^{\star}m_y] 4 4 2 1
    [3_z2_x] Reducible [6_z/m_zm_xm_y] [\bar 6_z^{\star}2_xm_y^{\star}], [6_z^{\star}2_x2_y^{\star}], [\bar 3_z^{\star}m_x^{\star} ] 4 4 0 1
    [3_z2_y] Reducible [6_z/m_zm_xm_y] [\bar 6_z^{\star}m_x^{\star}2_y], [6_z^{\star}2_x^{\star}2_y], [\bar 3_z^{\star}m_y^{\star}] 4 4 0 1
    [\bar 3_z] Reducible [6_z/m_zm_xm_y] [6_z^{\star}/m_z^{\star}], [\bar 3_zm_x^{\star} ], [\bar 3_zm_y^{\star}] 4 4 0 1
    [3_z] Reducible [6_z/m_zm_xm_y] [\bar 6_z^{\star}], [6_z^{\star} ], [3_zm_x^{\star}], [3_zm_y^{\star}], [3_z2_x^{\star} ], [3_z2_y^{\star}], [\bar 3_z^{\star}] 8 8 2 1
    [m_xm_ym_z] [(m_{x{^\prime}}m_{y{^\prime}}m_z ], [m_{x{^\prime}{^\prime}}m_{y{^\prime}{^\prime}}m_z)] [E_{2g} ] [m_xm_ym_z ] [6_z/m_zm_xm_y] 3 1 0 3
    [m_xm_y2_z ] [(m_{x{^\prime}}m_{y{^\prime}}2_z ], [m_{x{^\prime}{^\prime}}m_{y{^\prime}{^\prime}}2_z)] [E_{2u}] [m_xm_ym_z] [6_z/m_zm_xm_y], [6_zm_xm_y], [m_xm_ym_z^{\star}] 6 2 2 3
    [2_xm_ym_z] [(2_{x{^\prime}}m_{y{^\prime}}m_z ], [2_{x{^\prime}{^\prime}}m_{y{^\prime}{^\prime}}m_z)] [E_{1u} ] [m_xm_ym_z] [6_z/m_z m_xm_z], [\bar 6_z2_xm_y ], [m_x^{\star}m_ym_z] 6 2 6 3
    [m_x2_ym_z] [(m_{x{^\prime}}2_{y{^\prime}}m_z ], [m_{x{^\prime}{^\prime}}2_{y{^\prime}{^\prime}}m_z)] [E_{1u}] [m_xm_ym_z] [6_z/m_z m_xm_z], [\bar 6_zm_x2_y ], [m_xm_y^{\star}m_z] 6 2 6 3
    [2_x2_y2_z] [(2_{x{^\prime}}2_{y{^\prime}}2_z ], [2_{x{^\prime}{^\prime}}2_{y{^\prime}{^\prime}}2_z)] [E_{2u}] [m_xm_ym_z ] [6_z/m_zm_xm_y], [6_z2_x2_y], [m_x^{\star}m_y^{\star}m_z^{\star}] 6 6 0 3
    [2_z/m_z] [E_{2g}] [6_z/m_zm_xm_y] [6_z/m_z], [m_x^{\star}m_y^{\star}m_z(3) ] 6 6 0 6
    [2_x/m_x] [(2_{x{^\prime}}/m_{x{^\prime}} ], [2_{x{^\prime}{^\prime}}/m_{x{^\prime}{^\prime}})] [E_{1g}] [m_xm_ym_z] [6_z/m_zm_xm_y], [\bar 3_zm_x], [m_xm_y^{\star}m_z^{\star} ] 6 2 0 6
    [2_y/m_y] [(2_{y{^\prime}}/m_{y{^\prime}} ], [2_{y{^\prime}{^\prime}}/m_{y{^\prime}{^\prime}})] [E_{1g}] [m_xm_ym_z ] [6_z/m_zm_xm_y], [\bar 3m_y], [m_x^{\star}m_ym_z^{\star} ] 6 2 0 6
    [m_z] [E_{1u}] [6_z/m_zm_xm_y] [6_z/m_z], [\bar 6_z], [2_x^{\star}m_y^{\star}m_z], [m_x^{\star}2_y^{\star}m_z], [2_z^{\star}/m_z ] 12 12 12 6
    [m_x] [(m_{x{^\prime}}], [m_{x{^\prime}{^\prime}})] Reducible [m_xm_ym_z ] [\bar 6_zm_x2_y], [6_zm_xm_y], [\bar 3_zm_x], [3_zm_x], [m_xm_y^{\star}2_z^{\star}], [m_x2_y^{\star}m_z^{\star}], [ 2_x^{\star}/m_x] 12 4 12 6
    [m_y] [(m_{y{^\prime}}], [m_{y{^\prime}{^\prime}})] Reducible [m_xm_ym_z] [\bar 6_z2_xm_y], [6_zm_xm_y], [\bar 3_zm_y], [3_zm_y], [m_x^{\star}m_y2_z^{\star}], [2_x^{\star}m_ym_z^{\star}], [2_y^{\star}/m_y] 12 4 12 6
    [2_z] [E_{2u}] [6_z/m_zm_xm_y ] [6_z/m_z], [6_z], [m_x^{\star}m_y^{\star}2_z (3) ], [2_x^{\star}2_y^{\star}2_z(3)], [2_z/m_z^{\star}] 12 12 2 6
    [2_x ] [(2_{x{^\prime}}], [2_{x{^\prime}{^\prime}})] Reducible [m_xm_ym_z] [\bar 6_z2_xm_y], [6_z2_x2_y], [\bar 3_zm_x], [3_z2_x], [2_xm_y^{\star}m_z^{\star}], [2_x2_y^{\star}2_z^{\star}], [2_x/m_x^{\star}] 12 4 6 6
    [2_y] [(2_{y{^\prime}}], [2_{y{^\prime}{^\prime}})] Reducible [m_xm_ym_z] [\bar 6_zm_x2_y], [6_z2_x2_y], [\bar 3_zm_y], [3_z2_y], [m_x^{\star}2_ym_z^{\star}], [2_x^{\star}2_y2_z^{\star}], [2_y/m_y^{\star}] 12 4 6 6
    [\bar 1 ] [E_{1g}] [6_z/m_zm_xm_y ] [6_z/m_z], [\bar 3_z], [2_z^{\star}/m_z^{\star}], [2_x^{\star}/m_x^{\star}(3)], [2_y^{\star}/m_y^{\star}(3) ] 12 12 0 12
    1 Reducible [6_z/m_zm_xm_y ] [\bar 6_z], [6_z], [\bar 3_z ], [3_z], [m_z^{\star}], [m_x^{\star}(3)], [m_y^{\star}(3) ], [2_z^{\star}], [2_x^{\star}(3)], [2_y^{\star}(3)], [\bar 1^{\star}] 24 24 24 12
    [{\bf 23}] [3_p] [(3_q], [3_r], [3_s)] T [3_p] 23 4 1 4 4
    [2_x2_y2_z] E 23 23 3 3 0 3
    [2_z] [(2_x], [2_y)] T [2_x2_y2_z] [23], [2_x^{\star}2_y^{\star}2_z ] 6 2 6 6
    1 T [23 ] [3_p(4)], [2_z^{\star}(3)] 12 12 12 12
    [{\bi m}{\bar {\bf 3}} ] 23 [A_u ] [m\bar3] [m^{\star}\bar3^{\star}] 2 2 0 1
    [\bar 3_p] [(\bar 3_q], [\bar3_r], [\bar 3_s) ] [T_g] [\bar3_p] [m\bar3] 4 1 0 4
    [3_p] [(3_q], [3_r], [3_s)] [T_u] [\bar3_p] [m\bar3, 23 ] 8 2 8 4
    [m_xm_ym_z] [E_g ] [m\bar3 ] [m\bar3] 3 3 0 3
    [m_xm_y2_z ] [(2_xm_ym_z], [m_x2_ym_z)] [T_u ] [m_xm_ym_z] [m\bar3, m_xm_ym_z^{\star}] 6 2 6 3
    [2_x2_y2_z] [E_u ] [m\bar3] [m\bar3], [23], [m_x^{\star}m_y^{\star}m_z^{\star} ] 6 6 0 3
    [2_z/m_z] [(2_x/m_x], [2_y/m_y)] [T_g] [m_xm_ym_z] [m\bar3], [m_x^{\star}m_y^{\star}m_z ] 6 2 0 6
    [m_z] [(m_x], [m_y)] Reducible [m_xm_ym_z ] [m\bar3], [2_x^{\star}m_y^{\star}m_z ], [m_x^{\star}2_y^{\star}m_z], [2_z^{\star}/m_z] 12 4 12 6
    [2_z] [(2_x], [2_y)] Reducible [m_xm_ym_z] [m\bar3], [23], [m_x^{\star}m_y^{\star}2_z ], [2_x^{\star}2_y^{\star}2_z], [2_z/m_z^{\star}] 12 4 6 6
    [\bar 1] [T_g ] [m\bar3] [\bar3_p(4)], [2_z^{\star}/m_z^{\star}(3) ] 12 12 0 12
    1 [T_u ] [m\bar3] [\bar3_p(4)], [3_p(4)], [m_z^{\star}(3)], [2_z^{\star}(3)], [\bar1^{\star}] 24 24 24 12
    [{\bf 432}] 23 [A_2] 432 [4^{\star}32^{\star}] 2 2 0 1
    [3_p2_{x\bar{y}}] [(3_q2_{x\bar{y}} ], [ 3_r2_{xy}], [3_s2_{xy})] [T_2] [3_p2_{x\bar{y}} ] 432 4 1 0 4
    [3_p] [(3_q], [3_r], [3_s)] [T_1] [3_p2_{x\bar{y}} ] [23], [3_p2_{x\bar{y}}^{\star} ] 8 2 8 4
    [4_z2_x2_{xy}] [(4_x2_y2_{yz}], [4_y2_z2_{xz})] E [4_z2_x2_{xy}] 432 3 1 0 3
    [4_z] [(4_x], [4_y)] [T_1] [4_z2_x2_{xy}] [432], [4_z2_x^{\star}2_{xy}^{\star} ] 6 2 6 3
    [2_x2_y2_z] E 432 [23], [4_z^{\star}2_x2_{xy}^{\star} ] 6 6 0 6
    [2_{x\bar{y}}2_{xy}2_z ] [(2_{y\bar{z}}2_{yz}2_x ], [2_{z\bar{x}}2_{zx}2_y)] [T_2] [4_z2_x2_{xy}] [432], [4_z^{\star}2_x^{\star}2_{xy} ] 6 2 0 6
    [2_z] [(2_x], [2_y)] Reducible [4_z2_x2_{xy}] 23, [4_y2_z2_{xy}], [4_z^{\star} ], [2_{x\bar{y}}^{\star}2_{xy}^{\star}2_z], [2_x^{\star}2_y^{\star}2_z ] 12 4 6 12
    [2_{xy}] [(2_{yz}], [2_{zx} ], [2_{x\bar{y}}], [2_{y\bar{z}}], [2_{z\bar{x}})] [T_1], [T_2 ] [2_{x\bar{y}}2_{xy}2_z ] [432], [3_r2_{xy}], [3_s2_{xy} ], [4_z2_x2_{xy}], [2_{x\bar{y}}2_{xy}^{\star}2_z^{\star}] 12 2 12 12
    1 [T_1], [T_2 ] 432 [3_p(4) ], [4_z(3)], [2_z^{\star}(3) ], [2_{xy}^{\star}(6)] 24 24 24 24
    [{\bar {\bf 4}}{\bf 3}{\bi m}] 23 [A_2 ] [\bar43m] [\bar4^{\star}3m^{\star} ] 2 2 0 1
    [3_pm_{x\bar{y}}] [(3_qm_{x\bar{y}} ], [3_rm_{xy}], [3_sm_{xy})] [T_2] [3_pm_{x\bar{y}} ] [\bar43m ] 4 1 4 4
    [3_p] [(3_q], [3_r], [3_s)] [T_1 ] [3_pm_{x\bar{y}} ] [\bar43m], [23], [3_pm_{x\bar{y}}^{\star} ] 8 2 4 4
    [\bar4_z2_xm_{xy}] [(\bar4_x2_ym_{yz} ], [\bar4_y2_zm_{zx})] E [\bar4_z2_xm_{x\bar{y}}] [\bar43m] 3 1 0 3
    [\bar4_z] [(\bar4_x, \bar4_y)] [T_1 ] [\bar4_z2_xm_{x\bar{y}}] [\bar43m], [\bar4_z2_x^{\star}m_{xy}^{\star} ] 6 2 0 3
    [m_{x\bar{y}}m_{xy}2_z] [(m_{y\bar{z}}m_{yz}2_x ], [m_{z\bar{x}}m_{zx}2_y)] [T_2 ] [\bar4_z2_xm_{x\bar{y}}] [\bar43m], [\bar4_z^{\star}2_x^{\star}m_{xy} ] 6 2 6 6
    [2_x2_y2_z] E [\bar43m ] [23, \bar4_z^{\star}2_xm_{xy}^{\star} ] 6 6 0 6
    [m_{xy}] [(m_{yz}], [m_{zx} ], [m_{x\bar{y}}], [m_{y\bar{z}}], [m_{z\bar{x}})] [T_1, T_2] [m_{x\bar{y}}m_{xy}2_z] [\bar43m], [3_rm_{xy}], [3_sm_{xy}], [\bar4_z2_xm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}2_z^{\star} ] 12 2 12 12
    [2_z] [(2_x], [2_y)] Reducible [\bar4_z2_xm_{xy} ] 23, [\bar4_z^{\star}], [4_z^{\star} ], [m_{x\bar{y}}^{\star}m_{xy}^{\star}2_z], [2_x^{\star}2_y^{\star}2_z ] 12 4 6 12
    1 [T_1], [T_2] [\bar43m ] [3_p(4) ], [\bar4_z(3)], [m_{xy}^{\star}(6)], [2_z^{\star}(3)] 24 24 24 24
    [{\bi m}{\bar {\bf 3}}{\bi m} ] [\bar43m] [A_{2u}] [m\bar3m] [m^{\star}\bar3^{\star}m] 2 2 0 1
    432 [A_{1u}] [m\bar3m] [m^{\star}\bar3^{\star}m^{\star}] 2 2 0 1
    [m\bar3 ] [A_{2g}] [m\bar3m] [m\bar3m^{\star}] 2 2 0 1
    23 Reducible [m\bar3m] [\bar4^{\star}3m^{\star}], [4^{\star}32^{\star} ], [m_z^{\star}\bar3_p ] 4 4 0 1
    [\bar 3_pm_{x\bar{y}}] [(\bar 3_qm_{x\bar{y}} ], [\bar3_rm_{xy}], [\bar 3_sm_{xy})] [T_{2g}] [\bar3_pm_{x\bar{y}}] [m\bar3m] 4 1 0 4
    [3_pm_{x\bar{y}}] [(3_qm_{x\bar{y}} ], [3_rm_{xy}], [3_sm_{xy})] [T_{1u}] [\bar3_pm_{x\bar{y}} ] [m\bar3m ], [\bar43m], [\bar3_p^{\star}m_{x\bar{y}}] 8 2 8 4
    [3_p2_{x\bar{y}}] [(3_q2_{x\bar{y}} ], [3_r2_{xy}], [3_s2_{xy})] [T_{2u}] [\bar3_pm_{x\bar{y}} ] [m\bar3m], [432], [\bar3_p^{\star}m_{x\bar{y}} ] 8 2 0 4
    [\bar 3_p] [(\bar 3_q], [\bar3_r], [\bar 3_s)] [T_{1g}] [\bar3_pm_{x\bar{y}}] [m\bar3m], [m\bar3], [\bar3_pm_{x\bar{y}}^{\star} ] 8 2 0 4
    [3_p] [(3_q], [3_r], [3_s)] Reducible [\bar3_pm_{x\bar{y}}] [\bar43m], [432], [m\bar3 ], [23], [3_pm_{x\bar{y}}^{\star}], [3_p2_{x\bar{y}}^{\star} ], [\bar3_p^{\star}] 16 4 8 4
    [4_z/m_zm_xm_{xy}] [(4_x/m_xm_ym_{yz} ], [4_y/m_ym_zm_{zx})] [E_g ] [4_z/m_zm_xm_{xy}] [m\bar3m] 3 1 0 3
    [\bar4_z2_xm_{xy}] [(\bar4_x2_ym_{yz} ], [\bar4_y2_zm_{zx})] [E_u ] [4_z/m_zm_xm_{xy}] [m\bar3m], [\bar43m], [4_z^{\star}/m_z^{\star}m_x^{\star}m_{xy}] 6 2 0 3
    [\bar4_zm_x2_{xy}] [(\bar4_xm_y2_{yz} ], [\bar4_ym_z2_{zx})] [T_{2u} ] [4_z/m_zm_xm_{xy}] [m\bar3m], [4_z^{\star}/m_z^{\star}m_xm_{xy}^{\star} ] 6 2 0 3
    [4_zm_xm_{xy}] [(4_xm_ym_{yz}], [4_ym_zm_{zx})] [T_{1u} ] [4_z/m_zm_xm_{xy}] [m\bar3m], [4_z/m_z^{\star}m_xm_{xy} ] 6 2 6 3
    [4_z2_x2_{xy}] [(4_x2_y2_{yz}], [4_y2_z2_{zx})] [E_u] [4_z/m_zm_xm_{xy} ] [m\bar3m], [432], [4_z/m_z^{\star}m_x^{\star}m_{xy}^{\star} ] 6 2 0 3
    [4_z/m_z] [(4_x/m_x], [4_y/m_y)] [T_{1g} ] [4_z/m_zm_xm_{xy}] [m\bar3m], [4_z/m_zm_x^{\star}m_{xy}^{\star} ] 6 2 0 3
    [\bar4_z] [(\bar4_x], [\bar4_y)] Reducible [4_z/m_zm_xm_{xy}] [m\bar3m], [\bar43m], [\bar4_z2_x^{\star}m_{xy}^{\star}], [\bar4_zm_x^{\star}2_{xy}^{\star} ], [4_z^{\star}/m_z^{\star}] 12 4 0 3
    [4_z] [(4_x ], [4_y)] Reducible [4_z/m_zm_xm_{xy}] [m\bar3m], [432], [4_zm_x^{\star}m_{xy}^{\star} ], [4_z2_x^{\star}2_{xy}^{\star}], [4_z/m_z^{\star}] 12 4 6 3
    [m_xm_ym_z] [E_g] [m\bar3m ] [m\bar3], [4_z^{\star}/m_zm_xm_{xy}^{\star} ] 6 6 0 6
    [m_{x\bar{y}}m_{xy}m_z] [(m_{y\bar{z}}m_{yz}m_x ], [m_{z\bar{x}}m_{zx}m_y)] [T_{2g} ] [4_z/m_zm_xm_{xy}] [m\bar3m ], [4_z^{\star}/m_zm_xm_{xy}^{\star} ] 6 2 0 6
    [m_xm_y2_z] [(2_xm_ym_z], [m_x2_ym_z)] Reducible [4_z/m_zm_xm_{xy}] [m\bar3], [4_y/m_ym_zm_{zx}], [\bar4_z^{\star}m_x2_{xy}^{\star}], [4_z^{\star}m_xm_{xy}^{\star}], [ m_xm_ym_z^{\star}] 12 4 6 6
    [m_{x\bar{y}}m_{xy}2_z] [(m_{y\bar{z}}m_{yz}2_x ], [m_{z\bar{x}}m_{zx}2_y)] Reducible [4_z/m_zm_xm_{xy}] [m\bar3m], [\bar43m], [\bar4_z^{\star}2_x^{\star}m_{xy}], [4_z^{\star}m_x^{\star}m_{xy}], [m_{x\bar{y}}m_{xy}m_z^{\star}] 12 4 6 6
    [m_{x\bar{y}}2_{xy}m_z] [(m_{y\bar{z}}2_{yz}m_x ], [m_{z\bar{x}}2_{zx}m_y], [ 2_{x\bar{y}}m_{xy}m_z], [2_{y\bar{z}}m_{yz}m_x], [2_{z\bar{x}}m_{zx}m_y)] [T_{1u}], [T_{2u}] [m_{x\bar{y}}m_{xy}m_z ] [m\bar3m(m_{zx})], [m\bar3m(2_{zx}) ], [4_z/m_zm_xm_{xy}], [ m_{x\bar{y}}m_{xy}^{\star}m_z] 12 2 12 6
    [2_x2_y2_z] [E_u] [m\bar3m ] [m\bar3], [23], [\bar4_z^{\star}2_xm_{xy}^{\star} ], [4_z^{\star}2_x2^{\star}_{xy}], [m_x^{\star}m_y^{\star}m_z^{\star} ] 12 12 0 6
    [2_{x\bar{y}}2_{xy}2_z] [(2_{y\bar{z}}2_{yz}2_x ], [2_{z\bar{x}}2_{zx}2_y)] [E_u] [4_z/m_zm_xm_{xy}] [m\bar3m], [432], [\bar4_zm_x^{\star}2_{xy} ], [4^{\star}_z2^{\star}_x2_{xy}], [m_{x\bar{y}}^{\star}m_{xy}^{\star}m_z^{\star} ] 12 4 0 6
    [2_z/m_z] [(2_x/m_x, 2_y/m_y)] Reducible [4_z/m_zm_xm_{x\bar{y}}] [m\bar3], [ 4_y/m_ym_zm_{zx}], [4_z^{\star}/m_z], [m_x^{\star}m_y^{\star}m_z], [m_{x\bar{y}}^{\star}m_{xy}^{\star}m_z ] 12 4 0 12
    [2_{xy}/m_{xy}] [(2_{yz}/m_{yz} ], [2_{zx}/m_{zx}], [2_{x\bar{y}}/m_{x\bar{y}}], [2_{y\bar{z}}/m_{y\bar{z}} ], [2_{z\bar{x}}/m_{z\bar{x}})] [T_{1g}], [T_{2g}] [m_{x\bar{y}}m_{xy}m_z] [m\bar3m], [\bar3_rm_{xy} (2)], [4_z/m_zm_xm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}m_z^{\star}] 12 2 0 12
    [m_z] [(m_x], [m_y)] [T_{1u}], [T_{2u} ] [4_z/m_zm_xm_{x\bar{y}} ] [m\bar3], [\bar4_xm_z2_{yz} ], [4_ym_zm_{zx}], [4_z/m_z], [2_x^{\star}m_y^{\star}m_z (2) ], [m_{x\bar{y}}^{\star}2_{xy}^{\star}m_z(2)], [2_z^{\star}/m_z ] 24 8 24 12
    [m_{xy}] [(m_{yz} ], [m_{zx}], [m_{x\bar{y}}], [m_{y\bar{z}}], [m_{z\bar{x}})] [T_{1u}] [m_{x\bar{y}}m_{xy}m_z ] [m\bar3m], [\bar43m ], [\bar4_z2_xm_{xy}], [4_zm_xm_{xy}], [\bar3_rm_{xy} ], [\bar3_sm_{xy}], [3_rm_{xy}], [3_sm_{xy}], [m_{x\bar{y}}^{\star}m_{xy}2_z^{\star} ], [2_{xy}^{\star}/m_{xy}] 24 4 24 12
    [2_z] [(2_x], [2_y)] Reducible [4_z/m_zm_xm_{x\bar{y}} ] [m\bar3], [23], [\bar4_y2_zm_{zx}], [4_y2_z2_{zx}], [\bar4_z^{\star}], [4_z^{\star}], [m_x^{\star}m_y^{\star}2_z], [m_{x\bar{y}}^{\star}m_{xy}^{\star}2_z ], [2_x^{\star}2_y^{\star}2_z ], [2_{x\bar{y}}^{\star}2_{xy}^{\star}2_z ], [2_z/m_z^{\star}] 24 8 6 12
    [2_{xy}] [(2_{yz} ], [ 2_{zx}], [2_{x\bar{y}}], [2_{y\bar{z}}], [2_{z\bar{x}})] [T_{2u}] [m_{x\bar{y}}m_{xy}m_z ] [m\bar3m], [432 ], [\bar3_rm_{xy}], [\bar3_sm_{xy}], [3_r2_{xy}], [3_s2_{xy}], [\bar4_zm_x2_{xy}], [4_z2_x2_{xy}], [m_{x\bar{y}}^{\star}2_{xy}m_z^{\star} ], [2_{x\bar{y}}^{\star}2_{xy}2_z^{\star}], [2_{xy}/m_{xy}^{\star} ] 24 4 12 12
    [\bar 1] [T_{1g}], [T_{2g} ] [m\bar3m ] [\bar3_p(4)], [4_z/m_z(3)], [2_z^{\star}/m_z^{\star}(3)], [2_{xy}^{\star}/m_{xy}^{\star}(6) ] 24 24 0 24
    1 [T_{1u}], [T_{2u} ] [m\bar3m ] [\bar3_p(4)], [\bar4_z(3)], [4_z(3)], [m_z^{\star}(3)], [m_{xy}^{\star}(6) ], [2_z^{\star}(3)], [2_{xy}^{\star}(6)], [ \bar1^{\star} ] 48 48 48 24
    [u =x], y, z, [xy], [yz], [zx], [x\bar{y}], [y\bar{z}], [z\bar{x}], [x{^\prime}], [x{^\prime}{^\prime}], [y{^\prime} ], [y{^\prime}{^\prime}].
  • [F_1]: this point group is a proper subgroup of G given in the first column and expresses the symmetry of the ferroic phase in the first single-domain state [{\bf S}_1]. In accordance with IT A (2005[link]), five groups are given in two orientations (bold and normal type). Subscripts of generators in the group symbol specify their orientation in the Cartesian (rectangular) crystallophysical coordinate system of the group G (see Tables 3.4.2.5[link] and 3.4.2.6[link], and Figs. 3.4.2.3 [link]and 3.4.2.4[link]). In the cubic groups, the direction of the body diagonal is denoted by abbreviated symbols: [p\equiv[111]] (all positive), [q\equiv[\bar 1\bar 11]], [r\equiv[1\bar 1\bar 1] ], [s\equiv[\bar 1 1 \bar 1]]. In the hexagonal and trigonal groups, axes x′, y′ and x′′, y′′ of a Cartesian coordinate system are rotated about the z axis through 120 and 240°, respectively, from the crystallophysical Cartesian coordinate axes x and [y ].

    Symmetry groups in parentheses are groups conjugate to [F_1 ] under G (see Section 3.2.3.2[link] ). These are symmetry groups (stabilizers) of some domain states [{\bf S}_k] different from [{\bf S}_1] (for more details see Section 3.4.2.2.3[link]).

  • [\Gamma_{\eta}]: physically irreducible representation of the group G. This specifies the transformation properties of the principal tensor parameter of the phase transition in a continuum description and transformation properties of the primary order parameter [\eta] of the equitranslational phase transitions in the microscopic description. The letters A, B signify one-dimensional representations, and letters E and T two- and three-dimensional irreducible representations, respectively. Two letters T indicate that the symmetry descent [G\subset F_1] can be accomplished by two non-equivalent three-dimensional irreducible representations (see Table 3.1.3.2[link] ). `Reducible' denotes a reducible representation of G. In this case, there are always several non-equivalent reducible representations inducing the same descent [G\subset F_1 ] [for more detailed information see the software GI[\star ]KoBo-1 and Kopský (2001[link])].

    Knowledge of [\Gamma_{\eta}] enables one to determine for all ferroic transitions property tensors and their components that are different in all principal domain states, and, for equitranslational transitions only, microscopic displacements and/or ordering of atoms and molecules that are different in different basic (microscopic) domain states (for details see Section 3.1.3[link] , especially Table 3.1.3.1[link] , and Section 3.1.2[link] ).

  • [{N_G} {(F}{_1)}]: the normalizer of [F_1] in G (defined in Section 3.2.3.2.4[link] ) determines subgroups conjugate to [F_1] in G and specifies which domain states have the same symmetry (stabilizer in G). The number [n_F] of subgroups conjugate to [F_1 ] in G is [n_F =] [[G:N_G(F_{1})] =] [|G|:|N_G(F_{1})|] [see equation (3.4.2.36)[link]] and the number [d_F] of principal domain states with the same symmetry is [d_F =] [[N_G(F_{1}):F_1] =] [|N_G(F_{1})|:|F_1|] [see equation (3.4.2.35)[link]]. There are three possible cases:

    (i) [N_G(F_1)=G ]. There are no subgroups conjugate to [F_1] and the symmetry group [F_i] (stabilizer of [{\bf S}_i] in G) of all principal domain states [{\bf S}_1,{\bf S}_2,\ldots,{\bf S}_n] is equal to G, [F_i=G], for all [i=1,2,\ldots,n]; hence domain states cannot be distinguished by their symmetry. The group [F_1] is a normal subgroup of [G, F_1\triangleleft G] (see Section 3.2.3.2[link] ). This is always the case if there are just two single-domain states [{\bf S}_1], [{\bf S}_2], i.e. if the index of [F_1] in G equals two, [[G:F_1]=|G|:|F_1|=2].

    (ii) [N_G(F_1)=F_1]. Then any two domain states [{\bf S}_i], [{\bf S}_k] have different symmetry groups (stabilizers), [{\bf S}_i\neq{\bf S}_k\Leftrightarrow F_i\neq F_k ], i.e. there is a one-to-one correspondence between single-domain states and their symmetries, [{\bf S}_i\Leftrightarrow F_i]. In this case, principal domain states [{\bf S}_i] can be specified by their symmetries [F_i, i=1,2,\ldots,n]. The number [n_F] of different groups conjugate to [F_1] is equal to the index [[G:F_1]=|G|:|F_1|=n ].

    (iii) [F_1\subset N_G(F_1)\subset G]. Some, but not all, domain states [{\bf S}_i], [{\bf S}_k] have identical symmetry groups (stabilizers) [F_i=F_k]. The number [d_F] of domain states with the same symmetry group is [d_F=] [[N_G(F_1):F_1]=] [|N_G(F_1)|:|F_1| ] [see equation (3.4.2.35[link])], [1 \,\lt\, d_F \,\lt\, n ]. The number [n_F] of different groups conjugate to [F_1] is equal to the index [n_F=] [[G:N_G(F_1)] =] [|G|:|N_G(F_1)|] [see equation (3.4.2.36[link])] and in this case [1 \,\lt\, n_F \,\lt\, n]. It always holds that [n_Fd_F=n] [see equation (3.4.2.37[link])].

  • [{K}_{1{j}}]: twinning group of a domain pair ([{\bf S}_1], [{\bf S}_j]). This group is defined in Section 3.4.3.2[link]. It can be considered a colour (polychromatic) group involving c colours, where [c=[K_{1j}:F_1]], and is, therefore, defined by two groups [K_{1j}] and [F_1], and its full symbol is [K_{1j}[F_1]]. In this column only [K_{1j}] is given, since [F_1] appears in the second column of the table.

    If the group symbol of [K_{1j}] contains generators with the star symbol, [^{\star}], which signifies transposing operations of the domain pair ([{\bf S}_1], [{\bf S}_j]), then the symbol [K_{1j}[F_1]] denotes a dichromatic (`black-and-white') group signifying a completely transposable domain pair. In this special case, just the symbol [K_{1j}] containing stars [^{\star}] specifies the group [F_1 ] unequivocally.

    The number in parentheses after the group symbol of [K_{1j}] is equal to the number of twinning groups [K_{1k}] equivalent with [K_{1j}].

    In the continuum description, a twinning group is significant in at least in two instances:

    • (1) A twinning group [K_{1j}[F_1]] specifies the distinction of two domain states [{\bf S}_1] and [{\bf S}_j=g_{1j}{\bf S}_1 ], where [g_{1j} \in G] (see Sections 3.4.3.2[link] and 3.4.3.4[link]).

    • (2) A twinning group [K_{1j}[F_1]] may assist in signifying classes of equivalent domain pairs (orbits of domain pairs). In most cases, to a twinning group [F_{1j}] there corresponds just one class of equivalent domain pairs (an orbit) G([{\bf S}_1,{\bf S}_j ]); then a twinning group can represent this class of equivalent domain pairs. Nevertheless, in some cases two or more classes of equivalent domain pairs have a common twinning group. Then one has to add a switching operation [g_{1j}] to the twinning group, [K_{1j}[F_1](g_{1j})] (see the end of Section 3.4.3.2[link]). In this way, classes of equivalent domain pairs G([{\bf S}_1], [{\bf S}_j]) are denoted in synoptic Tables 3.4.2.7[link] and 3.4.3.6[link].

    Twinning groups given in column [K_{1j}] thus specify all G-orbits of domain pairs. The number of G-orbits and representative domain pairs for each orbit are determined by double cosets of group [F_1] (see Section 3.4.3.2[link]). Representative domain pairs from each orbit of domain pairs are further analysed in synoptic Table 3.4.3.4[link] (non-ferroelastic domain pairs) and in synoptic Table 3.4.3.6[link] (ferroelastic domain pairs).

    The set of the twinning groups [K_{1j}] given in this column is analogous to the concept of a complete twin defined as `an edifice comprising in addition to an original crystal (domain state [{\bf S}_1]) as many twinned crystals (domain states [{\bf S}_j]) as there are possible twin laws' (see Curien & Le Corre, 1958[link]). If a traditional definition of a twin law [`a geometrical relationship between two crystal components of a twin', see Section 3.3.2[link] and Koch (2004[link]); Curien & Le Corre (1958[link])] is applied sensu stricto to domain twins then one gets the following correspondence:

    • (i) a twin law of a non-ferroelastic domain twin is specified by the twinning group [K_{1j}] (see Section 3.4.3.3[link] and Table 3.4.3.4[link]);

    • (ii) two twin laws of two compatible ferroelastic domain twins, resulting from one ferroelastic single-domain pair [\{{\bf S}_1,{\bf S}_j\} ], are specified by two layer groups [\overline{\sf {J}}_{1j}] associated with the twinning group [K_{1j}] of this ferroelastic single-domain pair [\{({\bf S}_1,{\bf S}_j)\}] (see Section 3.4.3.4[link] and Table 3.4.3.6[link]).

  • n : number of principal single-domain states, the finest subdivision of domain states in a continuum description, [n=] [[G:F_1]=|G|:|F_1|] [see equation (3.4.2.11[link])].

  • [{d_{F}}]: number of principal domain states with the same symmetry group (stabilizer), [d_F=[N_G(F_1):F_1]=|N_G(F_1)|:|F_1| ] [see equation (3.4.2.35[link])]. If [d_F>1], then the group [F_1] does not specify the first single-domain state [{\bf S}_1]. The number [n_F] of subgroups conjugate with [F_1 ] is [n_F=n:d_F].

  • [{ n_e}]: number of ferroelectric single-domain states, [n_e=[G:C_1]=] [|G|:|C_1|], where [C_1] is the stabilizer (in G) of the spontaneous polarization in the first domain state [{\bf S}_1 ] [see equation (3.4.2.32)[link]]. The number [d_e] of principal domain states compatible with one ferroelectric domain state (degeneracy of ferroelectric domain states) equals [d_e=[C_1:F_1]=|C_1|:|F_1|] [see equation (3.4.2.33[link])].

    Aizu's classification of ferroelectric phases (Aizu, 1969[link]; see Table 3.4.2.3[link]): [n_e=n], fully ferroelectric; [1 \,\lt\, n_e \,\lt\, n], partially ferroelectric; [n_e=1], non-ferroelectric, the parent phase is polar and the spontaneous polarization in the ferroic phase is the same as in the parent phase; [n_e=0], non-ferroelectric, parent phase is non-polar.

  • [{ n_a}]: number of ferroelastic single-domain states, [n_a=] [[G:A_1]] [=|G|:|A_1|], where [A_1] is the stabilizer (in G) of the spontaneous strain in the first domain state [{\bf S}_1 ] [see equation (3.4.2.28[link])]. The number [d_a] of principal domain states compatible with one ferroelastic domain state (degeneracy of ferroelastic domain states) is given by [d_a=[A_1:F_1]=|A_1|:|F_1|] [see equation (3.4.2.29[link])].

    Aizu's classification of ferroelastic phases (Aizu, 1969[link]; see Table 3.4.2.3[link]): [n_a=n], fully ferroelastic; [1 \,\lt\, n_a \,\lt\, n ], partially ferroelastic; [n_e=1], non-ferroelastic.

Example 3.4.2.5. Orthorhombic phase of perovskite crystals.  The parent phase has symmetry [G=m\bar3m] and the symmetry of the ferroic orthorhombic phase is [F_1=m_{x\bar y}2_{xy}m_z]. In Table 3.4.2.7[link], we find that [n=n_e], i.e. the phase is fully ferroelectric. Then we can associate with each principal domain state a spontaneous polarization. In column [K_{1j}] there are four twinning groups. As explained in Section 3.4.3[link], these groups represent four `twin laws' that can be characterized by the angle between the spontaneous polarization in single-domain state [{\bf S}_1] and [{\bf S}_j], [j=2,3,4,5]. If we choose [{\bf P}_{(s)}^{(1)}] along the direction [110] ([F_1] does not specify unambiguously this direction, since [d_F=2]!), then the angles between [{\bf P}_{(s)}^{(1)} ] and [{\bf P}_{(s)}^{(j)}], representing the `twin law' for these four twinning groups [m\bar3m(m_{zx})], [m\bar3m(2_{zx})], [4_z/m_zm_xm_{xy}], [m_{x\bar{y}}m_{xy}^{\star}m_z], are, respectively, 60, 120, 90 and 180°.

3.4.2.5. Basic (microscopic) domain states and their partition into translation subsets

| top | pdf |

The examination of principal domain states performed in the continuum approach can be easily generalized to a microscopic description. Let us denote the space-group symmetry of the parent (high-symmetry) phase by [\cal G] and the space group of the ferroic (low-symmetry) phase by [{\cal F}_1], which is a proper subgroup of [{\cal G}], [{\cal F}_1\subset {\cal G} ]. Further we denote by [{\sf S}_1] a basic (microscopic) low-symmetry structure described by positions of atoms in the unit cell. The stabilizer [{\cal I}_{\cal G}]([{\sf S}_1]) of the basic structure [{\sf S}_1] in a single-domain orientation is equal to the space group [{\cal F}_1] of the ferroic (low-symmetry) phase, [{\cal I}_{\cal G}({\sf S}_1)={\cal F}_1. \eqno(3.4.2.41) ]

By applying a lost symmetry operation [{\sf g}_j] on [{\sf S}_1 ], one gets a crystallographically equivalent low-symmetry basic structure [{\sf S}_j], [{\sf g}_j{\sf S}_1 = {\sf S}_j \not= {\sf S}_1, \quad {\sf g}_j\in {\cal G}, \quad {\sf g}_j \not\in {\cal F}_1. \eqno(3.4.2.42) ]We may recall that [{\sf g}_j] is a space-group symmetry operation consisting of a rotation (point-group operation) [g_j] and a non-primitive translation [{\bf u}(g_j)], [{\sf g}_j=\{g_j|{\bf u}(g_j)\}] (see Section 1.2.3[link] ). The symbol [\{g_j|{\bf u}(g_j)\}] is called a Seitz space-group symbol (Bradley & Cracknell, 1972[link]). The product (composition law) of two Seitz symbols is [\{g_1|{\bf u}(g_1)\}\{g_2|{\bf u}(g_2)\}=\{g_1g_2|g_1{\bf u}(g_2)+ {\bf u}(g_1)\}. \eqno(3.4.2.43) ]

All crystallographically equivalent low-symmetry basic structures form a [{\cal G}]-orbit and can be calculated from the first basic structure [{\sf S}_1] in the following way: [{\cal G}{\sf S}_1=\{{\sf S}_1, {\sf S}_2,\ldots, {\sf S}_j\ldots, {\sf S}_N\} = \{{\sf e}{\sf S}_1, {\sf g}_2{\sf S}_1,\ldots, {\sf g}_j{\sf S}_1\ldots, {\sf g}_N{\sf S}_1\}, \eqno(3.4.2.44) ]where [{\sf g}_1=] [{\sf e},] [{\sf g}_2,\ldots, {\sf g}_j,\ldots,{\sf g}_N ] are the representatives of the left cosets [{\sf g}_j{\cal F}_{1} ] of the decomposition of [{\cal G}], [{\cal G}={\cal F}_{1} \cup {\sf g}_2{\cal F}_{1} \cup\ldots\cup {\sf g}_j{\cal F}_{1} \cup\ldots\cup {\sf g}_N{\cal F}_{1}. \eqno(3.4.2.45) ]These crystallographically equivalent low-symmetry structures are called basic (elementary) domain states.

The number N of basic domain states is equal to the number of left cosets in the decomposition (3.4.2.45[link]). As we shall see in next section, this number is finite [see equation (3.4.2.60[link])], though the groups [{\cal G}] and [{\cal F}_{1}] consist of an infinite number of operations.

In a microscopic description, a basic (elementary) domain state is described by positions of atoms in the unit cell. Basic domain states that are related by translations suppressed at the phase transition are called translational or antiphase domain states. These domain states have the same macroscopic properties. The attribute `to have the same macroscopic properties' divides all basic domain states into classes of translational domain states.

In a microscopic description, a ferroic phase transition is accompanied by a lowering of space-group symmetry from a parent space group [{\cal G} ], with translation subgroup [{\cal T}] and point group G, to a low-symmetry space group [{\cal F}_1], with translation subgroup [{\cal U}_1] and point group [F_1]. There exists a unique intermediate group [{\cal M}_1], called the Hermann group, which has translation subgroup [{\cal T}] and point group [M_1=F_1] (see e.g. Hahn & Wondratschek, 1994[link]; Wadhawan, 2000[link]; Wondratschek & Aroyo, 2001[link]): [\eqalignno{{\cal F}_{1} \,&{\buildrel {c}\over {\subseteq}}\, {\cal M}_1 \,{\buildrel {t} \over {\subseteq}}\, \cal{G}, &(3.4.2.46)\cr F_1 &= M_1 \subseteq G, &(3.4.2.47)\cr {\cal U}_1&\subseteq {\cal T} = {\cal T},&(3.4.2.48)}%fd3.4.2.48 ]where [{\buildrel {c}\over {\subset}}] denotes an equiclass subgroup (a descent at which only the translational subgroup is reduced but the point group is preserved) and [{\buildrel {t}\over {\subset}}] signifies a equitranslational subgroup (only the point group descends but the translational subgroup does not change). Group [{\cal M}_1] is a maximal subgroup of [{\cal G}] that preserves all macroscopic properties of the basic domain state [{\sf S}_1] with symmetry [{\cal F}_1].

At this point we have to make an important note. Any space-group symmetry descent [{\cal G}\subset {\cal F}_1] requires that the lengths of the basis vectors of the translation group [{\cal U}_1] of the ferroic space group [{\cal F}_1] are commensurate with basic vectors of the translational group [{\cal T}] of the parent space group [{\cal G}]. It is usually tacitly assumed that this condition is fulfilled, although in real phase transitions this is never the case. Lattice parameters depend on temperature and are, therefore, different in parent and ferroic phases. At ferroelastic phase transitions the spontaneous strain changes the lengths of the basis vectors in different ways and at first-order phase transitions the lattice parameters change abruptly.

To assure the validity of translational symmetry descents, we have to suppress all distortions of the crystal lattice. This condition, called the high-symmetry approximation (Zikmund, 1984[link]) or parent clamping approximation (PCA) (Janovec et al., 1989[link]; Wadhawan, 2000[link]), requires that the lengths of the basis vectors [{\bf a}^{\kern1pt f}] [{\bf b}^{\kern1pt f}] [{\bf c}^{\kern1pt f} ] of the translation group [{\cal U}_1] of the ferroic space group [{\cal F}_1] are either exactly the same as, or are integer multiples of, the basic vectors [{\bf a}^{\kern1pt p}] [{\bf b}^{\kern1pt p}] [{\bf c}^{\kern1pt p}] of the translational group [{\cal T}] of the parent space group [{\cal G}]. Then the relation between the primitive basis vectors [{\bf a}^{\kern1pt f}] [{\bf b}^{\kern1pt f} ] [{\bf c}^{\kern1pt f}] of [{\cal U}_1] and the primitive basis vectors [{\bf a}^{\kern1pt p}] [{\bf b}^{\kern1pt p}] [{\bf c}^{\kern1pt p}] of [{\cal T}] can be expressed as [\pmatrix{{\bf a}^{\kern1pt f},\!\! & {\bf b}^{\kern1pt f},\!\! & {\bf c}^{\kern1pt f}}= \pmatrix{{\bf a}^{\kern1pt p},\!\! &{\bf b}^{\kern1pt p},\!\! & {\bf c}^{\kern1pt p}} \left(\matrix{m_{11} &m_{12} &m_{13} \cr m_{21} &m_{22} &m_{23} \cr m_{13} &m_{23} &m_{33}}\right), \eqno(3.4.2.49) ]where [m_{ij}], [i, j=1,2,3], are integers.

Throughout this part, the parent clamping approximation is assumed to be fulfilled.

Now we can return to the partition of the set of basic domain states into translational subsets. Let [\{{\sf S}_1,{\sf S}_2,\ldots,{\sf S}_{d_{t}}\} ] be the set of all basic translational domain states that can be generated from [{\sf S}_1] by lost translations. The stabilizer (in [{\cal G} ]) of this set is the Hermann group, [{\cal I}_{\cal G}\{{\sf S}_1,{\sf S}_2,\ldots,{\sf S}_{d_t}\} = {\cal M}_1, \eqno(3.4.2.50) ]which plays the role of the intermediate group. The number of translational subsets and the relation between these subsets is determined by the decomposition of [{\cal G}] into left cosets of [{\cal M}_1]: [\eqalignno{{\cal G}&=\{g_1|{\bf v}(g_1)\}{\cal M}_1 \cup \{g_2|{\bf v}(g_2)\}{\cal M}_1 \cup\ldots\cup \{g_j|{\bf v}(g_j)\}{\cal M}_1 &\cr &\quad\cup\ldots\cup \{g_{n}|{\bf v}(g_{n})\}{\cal M}_1. &(3.4.2.51)} ]Representatives [{\sf g}_j=\{g_j|{\bf u}(g_j)\}] are space-group operations, where [g_j] is a point-group operation and [{\bf u}(g_j) ] is a non-primitive translation (see Section 1.2.3[link] ).

We note that the Hermann group [{\cal M}_1] can be found in the software GI[\star]KoBo-1 as the equitranslational subgroup of [{\cal G}] with the point-group descent [G \subset F_1 ] for any space group [{\cal G}] and any point group [F_1] of the ferroic phase.

The decomposition of the point group G into left cosets of the point group [F_1] is given by equation (3.4.2.10[link]): [G=g_1F_{1} \cup g_2F_{1} \cup\ldots\cup g_jF_{1}\cup\ldots\cup g_{n}F_{1}. \eqno(3.4.2.52) ]Since the space groups [{\cal M}_1] and [{\cal F}_1] have identical point groups, [M_1=F_1], the decomposition (3.4.2.51[link]) is identical with a decomposition of G into left cosets of [M_1]; one can, therefore, choose for the representatives in (3.4.2.10[link]) the point-group parts of the representatives [\{g_j|{\bf u}(g_j)\}] in decomposition (3.4.2.51[link]). Both decompositions comprise the same number of left cosets, i.e. corresponding indices are equal; therefore, the number of subsets, comprising only translational basic domain states, is equal to the number n of principal domain states: [n =[{\cal G}:{\cal M}_1]=[G:F_1]=|G|:|F_1|, \eqno(3.4.2.53) ]where [|G|] and [|F_1|] are the number of operations of G and [F_1], respectively.

The first `representative' basic domain state [{\sf S}_j] of each subset can be obtained from the first basic domain state [{\sf S}_1]: [{\sf S}_j =\{g_j|{\bf v}(g_j)\}{\sf S}_1, \quad j=1,2,\ldots,n, \eqno(3.4.2.54) ]where [\{g_j|{\bf v}(g_j)\}] are representatives of left cosets of [{\cal M}_1] in the decomposition (3.4.2.51[link]).

Now we determine basic domain states belonging to the first subset (first principal domain state). Equiclass groups [{\cal M}_1] and [{\cal F}_1 ] have the same point-group operations and differ only in translations. The decomposition of [{\cal M}_1] into left cosets of [{\cal F}_1 ] can therefore be written in the form[{\cal M}_1=\{e|{\bf t}_1\}{\cal F}_1\cup \{e|{\bf t}_2\}{\cal F}_1 \cup\ldots\cup \{e|{\bf t}_k\}{\cal F}_1 \cup\ldots\cup \{e|{\bf t}_{d_{t}}\}{\cal F}_1, \eqno(3.4.2.55) ]where e is the identity point-group operation and [{\cal T}_k ], [k=1,2\ldots, d_t], are lost translations that can be identified with the representatives in the decomposition of [{\cal T}] into left cosets of [{\cal U}_1]: [{\cal T}={\bf t}_{1}{\cal U}_1+{\bf t}_{2}{\cal U}_1+\ldots+ {\bf t}_k{\cal U}_1 +\ldots+ {\bf t}_{d_t}{\cal U}_1. \eqno(3.4.2.56) ]The number [d_t] of basic domain states belonging to one principal domain state will be called a translational degeneracy. For the translations [{\bf t}_1,{\bf t}_2,\ldots,{\bf t}_k,\ldots,{\bf t}_{d_t}], one can choose vectors that lead from the origin of a `superlattice' primitive unit cell of [{\cal U}_1] to lattice points of [{\cal T}] located within or on the side faces of this `superlattice' primitive unit cell. The number [d_t] of such lattice points is equal to the ratio [v_{\cal F}:v_{\cal G} ], where [v_{\cal F}] and [v_{\cal G}] are the volumes of the primitive unit cells of the low-symmetry and parent phases, respectively.

The number [d_t] can be also expressed as the determinant det[(m_{ij})] of the [(3\times 3)] matrix of the coefficients [m_{ij}] that in equation (3.4.2.49[link]) relate the primitive basis vectors [{\bf a}^{\kern1pt f},{\bf b}^{\kern1pt f},{\bf c}^{\kern1pt f}] of [{\cal U}_1] to the primitive basis vectors [{\bf a}^{\kern1pt p},{\bf b}^{\kern1pt p},{\bf c}^{\kern1pt p} ] of [{\cal T}] (Van Tendeloo & Amelinckx, 1974[link]; see also Example 2.5[link] in Section 3.2.3.3[link] ). Finally, the number [d_t] equals the ratio [Z_{\cal F}:Z_{\cal G}], where [Z_{\cal F}] and [Z_{\cal G}] are the numbers of chemical formula units in the primitive unit cell of the ferroic and parent phases, respectively. Thus we get for the translational degeneracy [d_f] three expressions: [d_t=[{\cal M}_1: {\cal F}_1]=[{\cal T}:{\cal U}]=v_{\cal F}:v_{\cal G}= {\rm det}(m_{ij})=Z_{\cal F}:Z_{\cal G}. \eqno(3.4.2.57) ]The basic domain states belonging to the first subset of translational domain states are [{\sf S}_j=\{e|{\bf t}_k\}{\sf S}_1, \quad k=1,2,\ldots,d_t, \eqno(3.4.2.58) ]where [\{e|{\bf t}_k\}] is a representative from the decomposition (3.4.2.55[link]).

The partitioning we have just described provides a useful labelling of basic domain states: Any basic domain state can be given a label [ab], where the first integer [a=1,2,\ldots,n] specifies the principal domain state (translational subset) and the integer [b=1,2,\ldots,d_t] designates the the domain state within a subset. With this convention the kth basic domain state in the jth subset can be obtained from the first basic domain state [{\sf S}_1={\sf S}_{11}] (see Proposition 3.2.3.30[link] in Section 3.2.3.3[link] ): [{\sf S}_{jk}=\{g_j|{\bf v}(g_j)\}\{e|{\bf t}_k\}{\sf S}_{11}, \quad j=1,2,\ldots,n, \quad k=1,2,\ldots,d_t. \eqno(3.4.2.59) ]In a shorthand version, the letter [{\sf S}] can be omitted and the symbol can be written in the form [a_b], where the `large' number a signifies the principal domain state and the subscript b (translational index) specifies a basic domain state compatible with the principal domain state a.

The number n of translational subsets (which can be associated with principal domain states) times the translational degeneracy [d_t] (number of translational domain states within one translational subset) is equal to the total number N of all basic domain states: [\eqalignno{N&=nd_t=(|G|:|F_1|)(v_{\cal F}:v_{\cal G})= (|G|:|F_1|){\rm det}(m_{ij})&\cr&=(|G|:|F_1|)(Z_{\cal F}:Z_{\cal G}). &(3.4.2.60)\cr} ]

Example 3.4.2.6. Basic domain states in gadolinium molybdate (GMO).  Gadolinium molybdate [Gd2(MoO4)3] undergoes a non-equitranslational ferroic phase transition with parent space group [{\cal G}=P\bar42_1m] [(D_{2d}^3)] and with ferroic space group [{\cal F}_1=Pba2] [(C_{2v}^8)] (see Section 3.1.2[link] ). From equation (3.4.2.53[link]) we get n = [|\bar42m|:|mm2| =] [8:4 = 2], i.e. there are two subsets of translational domain states corresponding to two principal domain states. In the software GI[\star]KoBo-1 one finds for the space group [P\bar42_1m] and the point group [mm2] the corresponding equitranslational subgroup [{\cal M}_1=Cmm2 ] [(C_{2v}^{11})] with vectors of the conventional orthorhombic unit cell (in the parent clamping approximation) [{\bf a}^{o}={\bf a}^{t}-{\bf b}^{t} ], [{\bf b}^{o}={\bf a}^{t}+{\bf b}^{t}], [{\bf c}^{o}={\bf c}^{t} ], where [{\bf a}^{t},] [{\bf b}^{t},] [{\bf c}^{t}] is the basis of the tetragonal space group [P\bar42_1m]. Hence, according to equation (3.4.2.49[link]), [\left(\matrix{{\bf a}^o,\!\! &{\bf b}^o,\!\! &{\bf c}^o}\right) = \left(\matrix{{\bf a}^t,\!\! &{\bf b}^t,\!\! &{\bf c}^t}\right) \left(\matrix{1 &1 &0 \cr -1 &1 &0 \cr 0 &0 &1}\right). \eqno(3.4.2.61) ]The determinant of the transformation matrix equals two, therefore, according to equation (3.4.2.57[link]), each principal domain state can contain [d_t=2] translational domain states that are related by lost translation [{\bf a}^{t}] or [{\bf b}^{t}]. In all, there are four basic domain states (for more details see Barkley & Jeitschko, 1973[link]; Janovec, 1976[link]; Wondratschek & Jeitschko, 1976[link]).

Example 3.4.2.7. Basic domain states in calomel crystals.  Crystals of calomel, Hg2Cl2, consist of almost linear Cl—Hg—Hg—Cl molecules aligned parallel to the c axis. The centres of gravity of these molecules form in the parent phase a tetragonal body-centred parent phase with the conventional tetragonal basis at, bt, ct and with space group [{\cal G}=I4/mmm]. The structure of this phase projected onto the [z=0] plane is depicted in the middle of Fig. 3.4.2.5[link] as a solid square with four full circles and one empty circle representing the centres of gravity of the Hg2Cl2 molecules at the levels [z=0] and [z=c/2], respectively.

[Figure 3.4.2.5]

Figure 3.4.2.5 | top | pdf |

Four basic single-domain states [{{\sf S}_1}\equiv 1_1], [{{\sf S}_2}\equiv 1_2], [{{\sf S}_3}\equiv 2_1], [{{\sf S}_4}\equiv 2_2 ] of the ferroic phase of a calomel (Hg2Cl2) crystal. Full [\bullet] and empty [\circ] circles represent centres of gravity of Hg2Cl2 molecules at the levels [z=0] and [z=c/2], respectively, projected onto the [z=0] plane. The parent tetragonal phase is depicted in the centre of the figure with a full square representing the primitive unit cell. Arrows are exaggerated spontaneous shifts of molecules in the ferroic phase. Dotted squares depict conventional unit cells of the orthorhombic basic domain states in the parent clamping approximation. If the parent clamping approximation is lifted, these unit cells would be represented by rectangles elongated parallel to the arrows.

The ferroic phase has point-group symmetry [F_1=m_{xy}m_{x\bar y}2_z ], hence there are n = [|\bar42m|:|m_{xy}m_{x\bar y}2_z|] = 2 ferroelastic principal domain states. The conventional orthorhombic basis is [{\bf a}^{o}={\bf a}^{t}-{\bf b}^{t}, ] [{\bf b}^{o}={\bf a}^{t}+{\bf b}^{t},] [{\bf c}^{ o}={\bf c}^{t} ] (see upper left corner of Fig. 3.4.2.5[link]). This is the same situation as in the previous example, therefore, according to equations (3.4.2.57[link]) and (3.4.2.61[link]), the translational degeneracy [d_t=2], i.e. each ferroelastic domain state can contain two basic domain states.

The structure [{\sf S}_1] of the ferroic phase in the parent clamping approximation is depicted in the left-hand part of Fig. 3.4.2.5[link] with a dotted orthorhombic conventional unit cell. The arrows represent exaggerated spontaneous shifts of the molecules. These shifts are frozen-in displacements of a transverse acoustic soft mode with the k vector along the [110] direction in the first domain state [{\sf S}_1], hence all molecules in the (110) plane passing through the origin O are shifted along the [[1\bar10]] direction, whereas those in the neighbouring parallel planes are shifted along the antiparallel direction [[\bar110]] (the indices are related to the tetragonal coordinate system). The symmetry of [{\sf S}_1] is described by the space group [{\cal F}_1=Amam] [(D^{17}_{2h})]; this symbol is related to the conventional orthorhombic basis and the origin of this group is shifted by [{\bf a}^{t}/2] or [{\bf b}] with respect to the origin 0 of the group [{\cal G}=I4/mmm ].

Three more basic domain states [{\sf S}_2], [{\sf S}_3] and [{\sf S}_4] can be obtained, according to equation (3.4.2.44[link]), from [{\sf S}_1] by applying representatives of the left cosets in the resolution of [{\cal G}] [see equation (3.4.2.42[link])], for which one can find the expression[{\cal G}=\{1|000\}{\cal F}_1 \cup \{1|100\}{\cal F}_1 \cup \{4_z|000\}{\cal F}_1 \cup \{4_{z}{^3}|000\}{\cal F}_1.\eqno(3.4.2.62) ]

All basic domain states [{{\sf S}_1},] [{{\sf S}_2},] [{{\sf S}_3}] and [{{\sf S}_4}] are depicted in Fig. 3.4.2.5[link]. Domain states [{{\sf S}_1}] and [{{\sf S}_2}], and similarly [{{\sf S}_3}] and [{{\sf S}_4}], are related by lost translation [{\bf a}^{t}] or [{\bf b}^{t}]. Thus the four basic domain states [{{\sf S}_1},] [{{\sf S}_2},] [{{\sf S}_3} ] and [{{\sf S}_4}] can be partitioned into two translational subsets [\{{{\sf S}_1},{{\sf S}_2}\}] and [\{{{\sf S}_3},{{\sf S}_4}\}]. Basic domain states forming one subset have the same value of the secondary macroscopic order parameter [\lambda], which is in this case the difference [{\varepsilon}_{11}-{\varepsilon}_{22}] of the components of a symmetric second-rank tensor [\varepsilon], e.g. the permittivity or the spontaneous strain (which is zero in the parent clamping approximation).

This partition provides a useful labelling of basic domain states: [{{\sf S}_1}\equiv 1_1,] [{{\sf S}_2}\equiv 1_2,] [{{\sf S}_3}\equiv 2_1, ] [{{\sf S}_1}\equiv 2_2], where the first number signifies the ferroic (orientational) domain state and the subscript (translational index) specifies the basic domain state with the same ferroic domain state.

Symmetry groups (stabilizers in [{\cal G}]) of basic domain states can be calculated from a space-group version of equation (3.4.2.13[link]): [\eqalign{{\cal F}_2&=\{1|100\}{\cal F}_2\{1|100\}^{-1}={\cal F}_1\semi\cr \ {\cal F}_3&= \{4_z|000\}{\cal F}_2\{4_z|000\}^{-1}=Bbmm,} ]with the same conventional basis, and [{\cal F}_4=\{1|100\}{\cal F}_3\{1|100\}^{-1}] [={\cal F}_3], where the origin of these groups is shifted by [{\bf a}^{t}/2 ] or [{\bf b}] with respect to the origin [] of the group [{\cal G}=I4/mmm].

In general, a space-group-symmetry descent [{\cal G}\supset {\cal F}_1 ] can be performed in two steps:

  • (1) An equitranslational symmetry descent [{\cal G}\,{\buildrel {t}\over {\supseteq}}\, {\cal M}_1 ], where [{\cal M}_1] is the equitranslational subgroup of [{\cal G}] (Hermann group), which is unequivocally specified by space group [{\cal G}] and by the point group [F_1] of the space group [{\cal F}_1]. The Hermann group [{\cal M}_1] can be found in the software GI[\star]KoBo-1 or, in some cases, in IT A (2005[link]) under the entry `Maximal non-isomorphic subgroups, type I'.

  • (2) An equiclass symmetry descent [{\cal M}_1\,{\buildrel{c}\over{\supseteq} }\,{\cal F}_1 ], which can be of three kinds [for more details see IT A (2005[link]), Section 2.2.15[link] ]:

    • (i) Space groups [{\cal M}_1] and [{\cal F}_1 ] have the same conventional unit cell. These descents occur only in space groups [{\cal M}_1] with centred conventional unit cells and the lost translations are some or all centring translations of the unit cell of [{\cal M}_1]. In many cases, the descent [{\cal M}_1\,{\buildrel{c}\over{\supseteq} }\,{\cal F}_1 ] can be found in the main tables of IT A (2005[link]), under the entry `Maximal non-isomorphic subgroups, type IIa'. Gadolinium molybdate belongs to this category.

    • (ii) The conventional unit cell of [{\cal M}_1 ] is larger than that of [{\cal F}_1]. Some vectors of the conventional unit cell of [{\cal U}_1] are multiples of that of [{\cal T}]. In many cases, the descent [{\cal M}_1\,{\buildrel{c}\over{\supseteq} }\,{\cal F}_1 ] can be found in the main tables of IT A (2005[link]), under the entry `Maximal non-isomorphic subgroups, type IIb'.

    • (iii) Space group [{\cal F}_1] is an isomorphic subgroup of [{\cal M}_1], i.e. both groups are of the same space-group type (with the same Hermann–Mauguin symbol) or of the enantiomorphic space-group type. Each space group has an infinite number of isomorphic subgroups. Maximal isomorphic subgroups of lowest index are tabulated in IT A (2005[link]), under the entry `Maximal non-isomorphic subgroups, type IIc'.

3.4.3. Domain pairs: domain twin laws, distinction of domain states and switching

| top | pdf |

Different domains observed by a single apparatus can exhibit different properties even though their crystal structures are either the same or enantiomorphic and differ only in spatial orientation. Domains are usually distinguished by their bulk properties, i.e. according to their domain states. Then the problem of domain distinction is reduced to the distinction of domain states. To solve this task, we have to describe in a convenient way the distinction of any two of all possible domain states. For this purpose, we use the concept of domain pair.

Domain pairs allow one to express the geometrical relationship between two domain states (the `twin law'), determine the distinction of two domain states and define switching fields that may induce a change of one state into the other. Domain pairs also present the first step in examining domain twins and domain walls.

In this section, we define domain pairs, ascribe to them symmetry groups and so-called twinning groups, and give a classification of domain pairs. Then we divide domain pairs into equivalence classes (G-orbits of domain pairs) – which comprise domain pairs with the same inherent properties but with different orientations and/or locations in space – and examine the relation between G-orbits and twinning groups.

A qualitative difference between the coexistence of two domain states provides a basic division into non-ferroelastic and ferroelastic domain pairs. The synoptic Table 3.4.3.4[link] lists representatives of all G-orbits of non-ferroelastic domain pairs, contains information about the distinction of non-ferroelastic domain states by means of diffraction techniques and specifies whether or not important property tensors can distinguish between domain states of a non-ferroelastic domain pair. These data also determine the external fields needed to switch the first domain state into the second domain state of a domain pair. Synoptic Table 3.4.3.6[link] contains representative ferroelastic domain pairs of G-orbits of domain pairs for which there exist compatible (permissible) domain walls and gives for each representative pair the orientation of the two compatible domain walls, the expression for the disorientation angle (obliquity) and other data. Table 3.4.3.7[link] lists representatives of all classes of ferroelastic domain pairs for which no compatible domain walls exist. Since Table 3.4.2.7[link] contains for each symmetry descent [G\supset F] all twinning groups that specify different G-orbits of domain pairs which can appear in the ferroic phase, one can get from this table and from Tables 3.4.3.4[link], 3.4.3.6[link] and 3.4.3.7[link] the significant features of the domain structure of any ferroic phase.

3.4.3.1. Domain pairs and their symmetry, twin law

| top | pdf |

A pair of two domain states, in short a domain pair, consists of two domain states, say [{\bf S}_i] and [{\bf S}_k], that are considered irrespective of their possible coexistence (Janovec, 1972[link]). Geometrically, domain pairs can be visualized as two interpenetrating structures of [{\bf S}_i] and [{\bf S}_k]. Algebraically, two domain states [{\bf S}_i] and [{\bf S}_k] can be treated in two ways: as an ordered or an unordered pair (see Section 3.2.3.1.2[link] ).

An ordered domain pair, denoted ([{\bf S}_i,{\bf S}_k]), consists of the first domain state [{\bf S}_i] and the second domain state. Occasionally, it is convenient to consider a trivial ordered domain pair ([{\bf S}_i,{\bf S}_i]) composed of two identical domain states [{\bf S}_i].

An ordered domain pair is a construct that in bicrystallography is called a dichromatic complex (see Section 3.3.3[link] ; Pond & Vlachavas, 1983[link]; Sutton & Balluffi, 1995[link]; Wadhawan, 2000[link]).

An ordered domain pair ([{\bf S}_i,{\bf S}_k]) is defined by specifying [{\bf S}_i] and [{\bf S}_k] or by giving [{\bf S}_i] and a switching operation [g_{ik}] that transforms [{\bf S}_i ] into [{\bf S}_k], [{\bf S}_k = g_{ik}{\bf S}_i, \quad {\bf S}_i,{\bf S}_k \in G{\bf S}_1, \quad g_{ik}\in G.\eqno(3.4.3.1) ]For a given [{\bf S}_i] and [{\bf S}_k], the switching operation [g_{ik}] is not uniquely defined since each operation from the left coset [g_{ik}{F}_i] [where [F_i] is the stabilizer (symmetry group) of [{\bf S}_i]] transforms [{\bf S}_i] into [{\bf S}_k], [g_{ik}{\bf S}_i =] [ (g_{ik}{F}_{i}){\bf S}_i =] [{\bf S}_k].

An ordered domain pair [({\bf S}_k,{\bf S}_i)] with a reversed order of domain states is called a transposed domain pair and is denoted [({\bf S}_i,{\bf S}_k)^t\equiv({\bf S}_k,{\bf S}_i)]. A non-trivial ordered domain pair [({\bf S}_i,{\bf S}_k)] is different from the transposed ordered domain pair, [({\bf S}_k,{\bf S}_i) \neq ({\bf S}_i,{\bf S}_k) \ \ {\rm for} \ \ i\neq k. \eqno(3.4.3.2) ]

If [g_{ik}] is a switching operation of an ordered domain pair [({\bf S}_i,{\bf S}_k)], then the inverse operation [g_{ik}^{-1}] of [g_{ik}] is a switching operation of the transposed domain pair [({\bf S}_k,{\bf S}_i)]: [{\rm if} \,\, ({\bf S}_i,{\bf S}_k)=({\bf S}_i,g_{ik}{\bf S}_i) \,\, {\rm and} \,\, ({\bf S}_k,{\bf S}_i)=({\bf S}_k,g_{ki}{\bf S}_k), \,\, {\rm then} \,\, g_{ki}=g_{ik}^{-1}.\eqno(3.4.3.3) ]

An unordered domain pair, denoted by [\{{\bf S}_i,{\bf S}_k\} ], is defined as an unordered set consisting of two domain states [{\bf S}_i] and [{\bf S}_k]. In this case, the sequence of domains states in a domain pair is irrelevant, therefore[\{{\bf S}_i,{\bf S}_k\}= \{{\bf S}_k,{\bf S}_i\}. \eqno(3.4.3.4)]

In what follows, we shall omit the specification `ordered' or `unordered' if it is evident from the context, or if it is not significant.

A domain pair [({\bf S}_i,{\bf S}_k)] can be transformed by an operation [g\in G] into another domain pair, [g({\bf S}_i,{\bf S}_k)\equiv (g{\bf S}_i,g{\bf S}_k)=({\bf S}_l,{\bf S}_m), \quad {\bf S}_i, {\bf S}_k, {\bf S}_l, {\bf S}_m\in G{\bf S}_1, \quad g\in G. \eqno(3.4.3.5) ]These two domain pairs will be called crystallographically equivalent (in G) domain pairs and will be denoted [({\bf S}_i,{\bf S}_k)] [{\buildrel {G}\over {\sim}}] [({\bf S}_l,{\bf S}_m)].

If the transformed domain pair is a transposed domain pair [({\bf S}_k,{\bf S}_i) ], then the operation g will be called a transposing operation, [g^{\star}({\bf S}_i,{\bf S}_k)=(g^{\star}{\bf S}_i,g^{\star}{\bf S}_k)=({\bf S}_k,{\bf S}_i), \quad {\bf S}_i, {\bf S}_k \in G{\bf S}_1, \quad g^{\star}\in G.\eqno(3.4.3.6) ]We see that a transposing operation [g^{\star}\in G] exchanges domain states [{\bf S}_i] and [{\bf S}_k]: [g^{\star}{\bf S}_i = {\bf S}_k, \quad g^{\star}{\bf S}_k = {\bf S}_i, \quad {\bf S}_i, {\bf S}_k \in G{\bf S}_1, \quad g^{\star}\in G.\eqno(3.4.3.7) ]Thus, comparing equations (3.4.3.1[link]) and (3.4.3.7[link]), we see that a transposing operation [g^{\star}] is a switching operation that transforms [{\bf S}_i] into [{\bf S}_k ], and, in addition, switches [{\bf S}_k] into [{\bf S}_i]. Then a product of two transposing operations is an operation that changes neither [{\bf S}_i] nor [{\bf S}_k].

What we call in this chapter a transposing operation is usually denoted as a twin operation (see Section 3.3.5[link] and e.g. Holser, 1958[link]a; Curien & Donnay, 1959[link]; Koch, 2004[link]). We are reserving the term `twin operation' for operations that exchange domain states of a simple domain twin in which two ferroelastic domain states coexist along a domain wall. Then, as we shall see, the transposing operations are identical with the twin operations in non-ferro­elastic domains (see Section 3.4.3.5[link]) but may differ in ferroelastic domain twins, where only some transposing operations of a single-domain pair survive as twin operations of the corresponding ferroelastic twin with a nonzero disorientation angle (see Section 3.4.3.6.3[link]).

Transposing operations are marked in this chapter by a star, [^\star] (with five points), which should be distinguished from an asterisk, [^*] (with six points), used to denote operations or symmetry elements in reciprocal space. The same designation is used in the software GI[\star]KoBo-1 and in the tables in Kopský (2001[link]). A prime, ′, is often used to designate transposing (twin) operations (see Section 3.3.5[link] ; Curien & Le Corre, 1958[link]; Curien & Donnay, 1959[link]). We have reserved the prime for operations involving time inversion, as is customary in magnetism (see Chapter 1.5[link] ). This choice allows one to analyse domain structures in magnetic and magnetoelectric materials (see e.g. Přívratská & Janovec, 1997[link]).

In connection with this, we invoke the notion of a twin law. Since this term is not yet common in the context of domain structures, we briefly explain its meaning.

In crystallography, a twin is characterized by a twin law defined in the following way (see Section 3.3.2[link] ; Koch, 2004[link]; Cahn, 1954[link]):

  • (i) A twin law describes the geometrical relation between twin components of a twin. This relation is expressed by a twin operation that brings one of the twin components into parallel orientation with the other, and vice versa. A symmetry element corresponding to the twin operation is called the twin element. (Requirement `and vice versa' is included in the definition of Cahn but not in that of Koch; for the most common twin operations of the second order the `vice versa' condition is fulfilled automatically.)

  • (ii) The relation between twin components deserves the name `twin law' only if it occurs frequently, is reproducible and represents an inherent feature of the crystal.

An analogous definition of a domain twin law can be formulated for domain twins by replacing the term `twin components' by `domains', say [{\bf D}_i({\bf S}_j,B_k)] and [{\bf D}_m({\bf S}_n,B_p)], where [{\bf S}_j], [B_k] and [{\bf S}_n], [B_p] are, respectively, the domain state and the domain region of the domains [{\bf D}_i({\bf S}_j,B_k) ] and [{\bf D}_m({\bf S}_n,B_p)], respectively (see Section 3.4.2.1[link]). The term `transposing operation' corresponds to transposing operation [g_{12}^{\star}] of domain pair [({\bf S}_1,{\bf S}_2) =] [({\bf S}_j, g_{jn}^{\star}{\bf S}_n)] as we have defined it above if two domains with domain states [{\bf S}_1] and [{\bf S}_2 ] coexist along a domain wall of the domain twin.

Domain twin laws can be conveniently expressed by crystallographic groups. This specification is simpler for non-ferroelastic twins, where a twin law can be expressed by a dichromatic space group (see Section 3.4.3.5[link]), whereas for ferroelastic twins with a compatible domain wall dichromatic layer groups are adequate (see Section 3.4.3.6.3[link]).

Restriction (ii)[link], formulated by Georges Friedel (1926[link]) and explained in detail by Cahn (1954[link]), expresses a necessity to exclude from considerations crystal aggregates (intergrowths) with approximate or accidental `nearly exact' crystal components resembling twins (Friedel's macles d'imagination) and thus to restrict the definition to `true twins' that fulfil condition (i)[link] exactly and are characteristic for a given material. If we confine our considerations to domain structures that are formed from a homogeneous parent phase, this requirement is fulfilled for all aggregates consisting of two or more domains. Then the definition of a `domain twin law' is expressed only by condition (i)[link]. Condition (ii)[link] is important for growth twins.

We should note that the definition of a twin law given above involves only domain states and does not explicitly contain specification of the contact region between twin components or neighbouring domains. The concept of domain state is, therefore, relevant for discussing the twin laws. Moreover, there is no requirement on the coexistence of interpenetrating structures in a domain pair. One can even, therefore, consider cases where no real coexistence of both structures is possible. Nevertheless, we note that the characterization of twin laws used in mineralogy often includes specification of the contact region (e.g. twin plane or diffuse region in penetrating twins).

Ordered domain pairs [({\bf S}_1, {\bf S}_2)] and [({\bf S}_1, {\bf S}_3) ], formed from domain states of our illustrative example (see Fig. 3.4.2.2[link]), are displayed in Fig. 3.4.3.1[link](a) and (b), respectively, as two superposed rectangles with arrows representing spontaneous polarization. In ordered domain pairs, the first and the second domain state are distinguished by shading [the first domain state is grey (`black') and the second clear (`white')] and/or by using dashed and dotted lines for the first and second domain state, respectively.

[Figure 3.4.3.1]

Figure 3.4.3.1 | top | pdf |

Transposable domain pairs. Single-domain states are those from Fig. 3.4.2.2[link]. (a) Completely transposable non-ferroelastic domain pair. (b) Partially transposable ferroelastic domain pair.

In Fig. 3.4.3.2[link], the ordered domain pair [({\bf S}_1, {\bf S}_2) ] and the transposed domain pair [({\bf S}_2, {\bf S}_1)] are depicted in a similar way for another example with symmetry descent [G=] [6_z/m_z \supset 2_z/m_z] = [ F_1 ].

[Figure 3.4.3.2]

Figure 3.4.3.2 | top | pdf |

Non-transposable domain pairs. (a) The parent phase with symmetry [G=6_z/m_z] is represented by a dotted hexagon and the three ferroelastic single-domain states with symmetry [F_1=F_2=F_3=2_z/m_z] are depicted as drastically squeezed hexagons. (b) Domain pair [({\bf S}_1,{\bf S}_2) ] and transposed domain pair [({\bf S}_2,{\bf S}_1)]. There exists no operation from the group [6_z/m_z] that would exchange domain states [{\bf S}_1] and [{\bf S}_2], i.e. that would transform one domain pair into a transposed domain pair.

Let us now examine the symmetry of domain pairs. The symmetry group [{F}_{ik}] of an ordered domain pair [({\bf S}_i,{\bf S}_k) =] [({\bf S}_i,g_{ik}{\bf S}_i)] consists of all operations that leave invariant both [{\bf S}_i] and [{\bf S}_k ], i.e. [{F}_{ik}] comprises all operations that are common to stabilizers (symmetry groups) [F_i] and [F_k] of domain states [{\bf S}_i] and [{\bf S}_k], respectively, [{F}_{ik} \equiv {F}{_i} \cap {F}_k = {F}_i \cap g_{ik}{F}_{i}g_{ik}^{-1}, \eqno(3.4.3.8) ]where the symbol [\cap] denotes the intersection of groups [F_i] and [F_k]. The group [{F}_{ik}] is in Section 3.3.4[link] denoted by [{\cal H}^*] and is called an intersection group.

From equation (3.4.3.8[link]), it immediately follows that the symmetry [F_{ki}] of the transposed domain pair [({\bf S}_k,{\bf S}_i) ] is the same as the symmetry [F_{ik}] of the initial domain pair [({\bf S}_i,{\bf S}_k)]: [{F}_{ki} = {F}{_k} \cap {F}_i={F}{_i} \cap {F}_k=F_{ik}.\eqno(3.4.3.9) ]

Symmetry operations of an unordered domain pair [\{{\bf S}_i,{\bf S}_k\} ] include, besides operations of [F_{ik}] that do not change either [{\bf S}_i] or [{\bf S}_k], all transposing operations, since for an unordered domain pair a transposed domain pair is identical with the initial domain pair [see equation (3.4.3.4[link])]. If [g^{\star}_{ik} ] is a transposing operation of [({\bf S}_i,{\bf S}_k)], then all operations from the left coset [g^{\star}_{ik}F_{ik}] are transposing operations of that domain pair as well. Thus the symmetry group [J_{ik}] of an unordered domain pair [\{{\bf S}_i,{\bf S}_k\} ] can be, in a general case, expressed in the following way:[J_{ik} = F_{ik} \cup g^{\star}_{ik}F_{ik}, \quad g^{\star}_{ik}\in G.\eqno(3.4.3.10) ]

Since, for an unordered domain, the order of domain states in a domain pair is not significant, the transposition of indices [i, k] in [J_{ik}] does not change this group, [J_{ik}=F_{ik} \cup g^{\star}_{ik}F_{ik}=F_{ki} \cup g^{\star}_{ki}F_{ki}=J_{ki}, \eqno(3.4.3.11) ]which also follows from equations (3.4.3.3[link]) and (3.4.3.9[link]).

A basic classification of domain pairs follows from their symmetry. Domain pairs for which at least one transposing operation exists are called transposable (or ambivalent) domain pairs. The symmetry group of a transposable unordered domain pair [({\bf S}_i,{\bf S}_k)] is given by equation (3.4.3.10[link]).

The star in the symbol [J_{ik}^{\star}] indicates that this group contains transposing operations, i.e. that the corresponding domain pair [({\bf S}_i,{\bf S}_k)] is a transposable domain pair.

A transposable domain pair [({\bf S}_i,{\bf S}_k)] and transposed domain pair [({\bf S}_k,{\bf S}_i)] belong to the same G-orbit: [G({\bf S}_i,{\bf S}_k)=G({\bf S}_k,{\bf S}_i).\eqno(3.4.3.12) ]

If [\{{\bf S}_i,{\bf S}_k\}] is a transposable pair and, moreover, [F_i =F_k =F_{ik}], then all operations of the left coset [g_{ik}^{\star}F_i] simultaneously switch [{\bf S}_i] into [{\bf S}_k] and [{\bf S}_k] into [{\bf S}_i]. We call such a pair a completely transposable domain pair. The symmetry group [J_{ik}] of a completely transposable pair [\{{\bf S}_i,{\bf S}_k\} ] is [J_{ik}^{\star} = F_i \cup g^{\star}_{ik}F_i, \quad g^{\star}_{ik}\in G, \quad F_i=F_k. \eqno(3.4.3.13) ]We shall use for symmetry groups of completely transposable domain pairs the symbol [J_{ik}^{\star}].

If [F_i \neq F_k], then [F_{ik}\subset F_i] and the number of transposing operations is smaller than the number of operations switching [{\bf S}_i] into [{\bf S}_k]. We therefore call such pairs partially transposable domain pairs. The symmetry group [J_{ik}] of a partially transposable domain pair [\{{\bf S}_i,{\bf S}_k\}] is given by equation (3.4.3.10[link]).

The symmetry groups [J_{ik}] and [J_{ik}^{\star}], expressed by (3.4.3.10[link]) or by (3.4.3.13[link]), respectively, consists of two left cosets only. The first is equal to [F_{ik}] and the second one [g^{\star}_{ik}F_{ik}] comprises all the transposing operations marked by a star. An explicit symbol [J_{ik}[F_{ik}] ] of these groups contains both the group [J_{ik}] and [F_{ik} ], which is a subgroup of [J_{ik}] of index 2.

If one `colours' one domain state, e.g. [{\bf S}_i], `black' and the other, e.g. [{\bf S}_k], `white', then the operations without a star can be interpreted as `colour-preserving' operations and operations with a star as `colour-exchanging' operations. Then the group [J_{ik}[F_{ik}] ] can be treated as a `black-and-white' or dichromatic group (see Section 3.2.3.2.7[link] ). These groups are also called Shubnikov groups (Bradley & Cracknell, 1972[link]), two-colour or Heesch–Shubnikov groups (Opechowski, 1986[link]), or antisymmetry groups (Vainshtein, 1994[link]).

The advantage of this notation is that instead of an explicit symbol [J_{ik}[F_{ik}]], the symbol of a dichromatic group specifies both the group [J_{ik}] and the subgroup [F_{ij}] or [F_1], and thus also the transposing operations that define, according to equation (3.4.3.7[link]), the second domain state [{\bf S}_j] of the pair.

We have agreed to use a special symbol [J_{ik}^{\star}] only for completely transposable domain pairs. Then the star in this case indicates that the subgroup [F_{ik}] is equal to the symmetry group of the first domain state [{\bf S}_i] in the pair, [F_{ik}=F_i]. Since the group [F_i] is usually well known from the context (in our main tables it is given in the first column), we no longer need to add it to the symbol of [J_{ik}].

Domain pairs for which an exchanging operation [g^{\star}_{ik}] cannot be found are called non-transposable (or polar) domain pairs. The symmetry [J_{ij}] of a non-transposable domain pair is reduced to the usual `monochromatic' symmetry group [F_{ik}] of the corresponding ordered domain pair [({\bf S}_i,{\bf S}_k)]. The G-orbits of mutually transposed polar domain pairs are disjoint (Janovec, 1972[link]): [G({\bf S}_i,{\bf S}_k) \cap G({\bf S}_k,{\bf S}_i)= \emptyset.\eqno(3.4.3.14) ]Transposed polar domain pairs, which are always non-equivalent, are called complementary domain pairs.

If, in particular, [F_{ik}=F_i=F_k], then the symmetry group of the unordered domain pair is [J_{ik}=F_{i}=F_k.\eqno(3.4.3.15) ]In this case, the unordered domain pair [\{{\bf S}_i,{\bf S}_k\}] is called a non-transposable simple domain pair.

If [F_i\neq F_k], then the number of operations of [F_{ik}] is smaller than that of [F_i] and the symmetry group [J_{ik}] is equal to the symmetry group [F_{ik}] of the ordered domain pair [({\bf S}_i,{\bf S}_k)], [J_{ik}=F_{ik}, \quad F_{ik}\subset F_i. \eqno(3.4.3.16) ]Such an unordered domain pair [\{{\bf S}_i,{\bf S}_k\}] is called a non-transposable multiple domain pair. The reason for this designation will be given later in this section.

We stress that domain states forming a domain pair are not restricted to single-domain states. Any two domain states with a defined orientation in the coordinate system of the parent phase can form a domain pair for which all definitions given above are applicable.

Example 3.4.3.1.  Now we examine domain pairs in our illustrative example of a phase transition with symmetry descent G = [4_z/m_zm_xm_{xy}\supset 2_xm_ym_z =] [F_1] and with four single-domain states [{\bf S}_1,] [{\bf S}_2,] [{\bf S}_3] and [{\bf S}_3], which are displayed in Fig. 3.4.2.2[link]. The domain pair [\{{\bf S}_1,{\bf S}_2\}] depicted in Fig. 3.4.3.1[link](a) is a completely transposable domain pair since transposing operations exist, e.g. [g^{\star}_{12}=m^{\star}_x], and the symmetry group [F_{12}] of the ordered domain pair [({\bf S}_1,{\bf S}_2) ] is [F_{12}=F_1\cap F_2=F_1=F_2=2_xm_ym_z.\eqno(3.4.3.17) ]The symmetry group [J_{12}] of the unordered pair [\{{\bf S}_1,{\bf S}_2\} ] is a dichromatic group, [J_{12}^{\star}=2_xm_ym_z \cup m^{\star}_x\{2_xm_ym_z\} = m_x^{\star}m_ym_z.\eqno(3.4.3.18) ]

The domain pair [\{{\bf S}_1,{\bf S}_3\}] in Fig. 3.4.3.1[link](b) is a partially transposable domain pair, since there are operations exchanging domain states [{\bf S}_1] and [{\bf S}_3], e.g. [g^{\star}_{13}=m^{\star}_{x\bar{y}}], but the symmetry group [F_{13}] of the ordered domain pair [({\bf S}_1,{\bf S}_3)] is smaller than [F_1]: [F_{13}=F_1 \cap F_3 = 2_xm_ym_z \cap m_x2_ym_z = \{1,m_z\}\equiv \{m_z\},\eqno(3.4.3.19) ]where 1 is an identity operation and [\{1,m_z\}] denotes the group [m_z]. The symmetry group of the unordered domain pair [\{{\bf S}_1,{\bf S}_3\} ] is equal to a dichromatic group, [J_{13} = \{m_z\} \cup 2^{\star}_{xy}.\{m_z\} = 2_{xy}^{\star}m_{\bar{x}y}^{\star}m_z.\eqno(3.4.3.20) ]

The domain pair [({\bf S}_1,{\bf S}_2)] in Fig. 3.4.3.2[link](b) is a non-transposable simple domain pair, since there is no transposing operation of [G=6_z/m_z] that would exchange domain states [{\bf S}_1 ] and [{\bf S}_2], and [F_1=F_2=2_z/m_z]. The symmetry group [J_{12}] of the unordered domain pair [\{{\bf S}_1,{\bf S}_3\}] is a `monochromatic' group, [J_{12}=F_{12}=F_1=F_2=2_z/m_z. \eqno(3.4.3.21) ]The G-orbit [6_z/m_z({\bf S}_1,{\bf S}_2)] of the pair [({\bf S}_1,{\bf S}_2)] has no common domain pair with the G-orbit [6_z/m_z({\bf S}_2,{\bf S}_1)] of the transposed domain pair [({\bf S}_2,{\bf S}_1) ]. These two `complementary' orbits contain mutually transposed domain pairs.

Symmetry groups of domain pairs provide a basic classification of domain pairs into the four types introduced above. This classification applies to microscopic domain pairs as well.

3.4.3.2. Twinning group, distinction of two domain states

| top | pdf |

We have seen that for transposable domain pairs the symmetry group [J_{1j}] of a domain pair [({\bf S}_1,{\bf S}_j)] specifies transposing operations [g^{\star}_{1j}F_1] that transform [{\bf S}_1] into [{\bf S}_j]. This does not apply to non-transposable domain pairs, where the symmetry group [J_{1j}=F_{1j}] does not contain any switching operation. Another group exists, called the twinning group, which is associated with a domain pair and which does not have this drawback. The twinning group determines the distinction of two domain states, specifies the external fields needed to switch one domain state into another one and enables one to treat domain pairs independently of the transition [G\supset F_1]. This facilitates the tabulation of the properties of non-equivalent domain pairs that appear in all possible ferroic phases.

The twinning group [K_{1j}] of a domain pair [({\bf S}_1,{\bf S}_j)] is defined as the minimal subgroup of G that contains both [F_1] and a switching operation [g_{1j} ] of the domain pair [({\bf S}_1,{\bf S}_j)], [{\bf S}_j=g_{1j}{\bf S}_1 ] (Fuksa & Janovec, 1995[link]; Janovec et al., 1995[link]; Fuksa, 1997[link]), [F_1\subset K_{1j}\subseteq G, \quad g_{1j}\in K_{1j},\eqno(3.4.3.22)]where no group [K'_{1j}] exists such that [F_1\subset K'_{1j}\subset K_{1j}, \quad g_{1j}\in K'_{1j}.\eqno(3.4.3.23) ]

The twinning group [K_{1j}] is identical to the embracing (fundamental) group used in bicrystallography (see Section 3.2.2[link] ). In Section 3.3.4[link] it is called a composite symmetry of a twin.

Since [K_{1j}] is a group, it must contain all products of [g_{1j}] with operations of [F_1], i.e. the whole left coset [g_{1j}F_1]. For completely transposable domain pairs, the union of [F_1] and [g_{1j}^{\star}F_1] forms a group that is identical with the symmetry group [J_{1j}^{\star}] of the unordered domain pair [\{{\bf S}_1,{\bf S}_j\}]: [K_{1j}^{\star}=J_{1j}^{\star} = F_1 \cup g^{\star}_{1j}F_1, \quad g_{1j}^{\star}\in K_{1j}, \quad F_1=F_j.\eqno(3.4.3.24) ]

In a general case, the twinning group [K_{1j}], being a supergroup of [F_1], can always be expressed as a decomposition of the left cosets of [F_1], [K_{1j} = F_1 \cup g_{1j}F_1 \cup g_{1k}F_1 \cup\ldots\cup g_{1c}\in G. \eqno(3.4.3.25) ]

We can associate with the twinning group a set of c domain states, the [K_{1j}]-orbit of [{\bf S}_1], which can be generated by applying to [{\bf S}_1] the representatives of the left cosets in decomposition (3.4.3.25[link]), [K_{1j}{\bf S}_1=\{{\bf S}_1,{\bf S}_j,\ldots, {\bf S}_c\}.\eqno(3.4.3.26) ]This orbit is called the generic orbit of domain pair [({\bf S}_1,{\bf S}_j) ].

Since the generic orbit (3.4.3.26[link]) contains both domain states of the domain pair [({\bf S}_1,{\bf S}_j)], one can find different and equal nonzero tensor components in two domain states [{\bf S}_1 ] and [{\bf S}_j] by a similar procedure to that used in Section 3.4.2.3[link] for ascribing principal and secondary tensor parameters to principal and secondary domain states. All we have to do is just replace the group G of the parent phase by the twinning group [K_{1j}]. There are, therefore, three kinds of nonzero tensor components in [{\bf S}_1] and [{\bf S}_j]:

  • (1) Domain states [{\bf S}_1] and [{\bf S}_j ] differ in the principal tensor parameters [\kappa_a] of the `virtual' phase transition with symmetry descent [K_{1j}\supset F_1], [\kappa_a^{(1)}\neq \kappa_a^{(j)}, \quad a=1,2,\ldots,\eqno(3.4.3.27) ]where [\kappa_a^{(1)}] and [\kappa_a^{(j)}] are the principal tensor parameters in domain states [{\bf S}_1] and [{\bf S}_j]; in the symbol of the principal tensor parameter [\kappa_{a}] we explicitly write only the lower index a, which numbers different principal tensor parameters, but omit the upper index labelling the representation of [K_{1j}], according to which [\kappa_a] transforms, and the second lower index denoting the components of the principal tensor parameter (see Section 3.4.2.3[link] and the manual of the software GI[\star]KoBo-1, path: Subgroups\View\Domains and Kopský, 2001[link]).

    The principal tensor parameters [\kappa_a^{(1)}] of lower rank in domain state [{\bf S}_1] can be found for [G=K_{1j}] in Table 3.1.3.1[link] of Section 3.1.3.3[link] , where we replace G by [K_{1j}], and for all important property tensors in the software GI[\star]KoBo-1, path: Subgroups\View\Domains and in Kopský (2001[link]), where we again replace G by [K_{1j}]. Tensor parameters in domain state [{\bf S}_j ] can be obtained by applying to the principal tensor parameters in [{\bf S}_1] the operation [g_{1j}].

  • (2) If there exists an intermediate group [L_{1j} ] in between [F_1] and [K_{1j}] that does not – contrary to [K_{1j}] – contain the switching operation [g_{1j}] of the domain pair [({\bf S}_1,{\bf S}_j)], [F_1\subset L_{1j}\subseteq K_{1j}, \quad g_{1j}\in L_{1j}, \eqno(3.4.3.28) ][cf. relation (3.4.3.23[link])] then domain states [{\bf S}_1] and [{\bf S}_j] differ not only in the principal tensor parameters [\kappa_a], but also in the secondary tensor parameters [\lambda_{b}]: [\lambda_{b}^{(1)}\neq \lambda_{b}^{(j)}, \quad I_{K_{1j}}(\lambda_{b}^{(1)})=L_{1j}, \quad b=1,\ldots, \eqno(3.4.3.29) ]where [\lambda_{b}^{(1)}] and [\lambda_{b}^{(j)}] are the secondary tensor parameters in domain states [{\bf S}_1] and [{\bf S}_j ]; the last equation, in which [I_{K_{1j}}(\lambda_{b}^{(1)})] is the stabilizer of [\lambda_{b}^{(1)}] in [K_{1j}], expresses the condition that [\lambda_{b}] is the principal tensor parameter of the transition [K_{1j}\supset L_{1j}] [see equation (3.4.2.40[link])].

    The secondary tensor parameters [\lambda_{b}^{(1)} ] of lower rank in domain state [{\bf S}_1] can be found for [G=K_{1j}] in Table 3.1.3.1[link] of Section 3.1.3.3[link] , and for all important property tensors in the soft­ware GI[\star]KoBo-1, path: Subgroups\View\Domains and in Kopský (2001[link]). Tensor parameters [\lambda_{b}^{(j)}] in domain state [{\bf S}_j] can be obtained by applying to the secondary tensor parameters [\lambda_{b}^{(1)}] in [{\bf S}_1] the operation [g_{1j}].

  • (3) All nonzero tensor components that are the same in domain states [{\bf S}_1] and [{\bf S}_j] are identical with nonzero tensor components of the group [K_{1j}]. These components are readily available for all important material tensors in Section 1.1.4[link] , in the software GI[\star]KoBo-1, path: Subgroups\View\Domains and in Kopský (2001[link]).

Cartesian tensor components corresponding to the tensor parameters can be calculated by means of conversion equations [for details see the manual of the software GI[\star]KoBo-1, path: Subgroups\View\Domains and Kopský (2001[link])].

Let us now illustrate the above recipe for finding tensor distinctions by two simple examples.

Example 3.4.3.2.  The domain pair [({\bf S}_1,{\bf S}_2)] in Fig. 3.4.3.1[link](a) is a completely transposable pair, therefore, according to equations (3.4.3.24[link]) and (3.4.3.18[link]), [K_{12}^{\star}=J_{12}^{\star}=2_xm_ym_z \cup m^{\star}_x\{2_xm_ym_z\} = m_x^{\star}m_ym_z.\eqno(3.4.3.30) ]

In Table 3.1.3.1[link] , we find that the first principal tensor parameter [\kappa^{(1)}] of the transition G = [K_{1j}=] [m_xm_ym_z\supset 2_xm_ym_z =] [F_1] is the x-component [P_1] of the spontaneous polarization, [\kappa_1^{(1)}=P_1]. Since the switching operation [g_{12}^{\star}] is for example the inversion [\bar 1], the tensor parameter [\kappa_1^{(2)}] in the second domain state [{\bf S}_2] is [\kappa_1^{(2)}=-P_1].

Other principal tensor parameters can be found in the software GI[\star]KoBo-1 or in Kopský (2001[link]), p. 185. They are: [\kappa_2^{(1)}=d_{12},] [ \kappa_3^{(1)}=d_{13},] [\kappa_4^{(1)}=d_{26},] [\kappa_5^{(1)}=d_{35}] (the physical meaning of the components is explained in Table 3.4.3.5[link]). In the second domain state [{\bf S}_2], these components have the opposite sign. No other tensor components exist that would be different in [{\bf S}_1 ] and [{\bf S}_2], since there is no intermediate group [L_{1j} ] in between [F_1] and [K_{1j}].

Nonzero components that are the same in both domain states are nonzero components of property tensors in the group [mmm] and are listed in Section 1.1.4.7[link] or in the software GI[\star]KoBo-1 or in Kopský (2001[link]).

The numbers of independent tensor components that are different and those that are the same in two domain states are readily available for all non-ferroelastic domain pairs and important property tensors in Table 3.4.3.4[link].

Example 3.4.3.3.  The twinning group of the partially transposable domain pair [({\bf S}_1,{\bf S}_3) ] in Fig. 3.4.3.1[link](b) with [{\bf S}_3=2_{xy}{\bf S}_1 ] has the twinning group [\eqalignno{K_{13} &= 2_xm_ym_z \cup 2_{xy}\{2_xm_ym_z\} \cup 2_z\{2_xm_ym_z\} \cup 2_{x\bar y}\{2_xm_ym_z\} &\cr&= 4_z/m_zm_xm_{xy}. &(3.4.3.31)} ]Domain states [{\bf S}_1] and [{\bf S}_3] differ in the principal tensor parameter of the transition [4_z/m_zm_xm_{xy} \subset 2_xm_ym_z ], which is two-dimensional and which we found in Example 3.4.2.4[link]: [\kappa_1^{(1)}=(P,0)]. Then in the domain state [{\bf S}_3] it is [\kappa_1^{(3)}=D(2_{xy})(P,0)=(0,P)]. Other principal tensors are: [\kappa_2^{(1)}=(g_4,0),] [\kappa_3^{(1)}=(d_{11},0), ] [\kappa_4^{(1)}=(d_{12},0),] [\kappa_5^{(1)}=(d_{13},0),] [\kappa_6^{(1)}=(d_{26},0),] [\kappa_7^{(1)}=(d_{35},0)] (the physical meaning of the components is explained in Table 3.4.3.5[link]). In the domain state [{\bf S}_3] they keep their absolute value but appear as the second nonzero components, as with spontaneous polarization.

There is an intermediate group [L_{13}=m_xm_ym_z] between [F_1=2_xm_ym_z ] and [K_{13}=4_z/m_zm_xm_{xy}], since [L_{13}=m_xm_ym_z] does not contain [g_{13}=2_{xy}]. The one-dimensional secondary tensor parameters for the symmetry descent [K_{13}=] [4_z/m_zm_xm_{xy}\supset L_{13}=] [m_xm_ym_z] was also found in Example 3.4.2.4[link]: [\lambda_{1}^{(1)}=u_1-u_2; ] [\lambda_{2}^{(1)}=A_{14}+A_{25},A_{36};] [\lambda_{3}^{(1)}=s_{11}-s_{22}, ] [s_{13}-s_{23},] [s_{44}-s_{55};] [\lambda_{4}^{(1)}=Q_{11}-Q_{22}, ] [Q_{12}-Q_{21},] [Q_{13}-Q_{23},] [Q_{31}-Q_{32},] [Q_{44}-Q_{55}]. All these parameters have the opposite sign in [{\bf S}_3 ].

The tensor distinction of two domain states [{\bf S}_1] and [{\bf S}_j] in a domain pair [({\bf S}_1,{\bf S}_j)] provides a useful classification of domain pairs given in the second and the third columns of Table 3.4.3.1[link]. This classification can be extended to ferroic phases which are named according to domain pairs that exist in this phase. Thus, for example, if a ferroic phase contains ferroelectric (ferroelastic) domain pair(s), then this phase is a ferroelectric (ferroelastic) phase. Finer division into full and partial ferroelectric (ferro­elastic) phases specifies whether all or only some of the possible domain pairs in this phase are ferroelectric (ferroelastic) ones. Another approach to this classification uses the notions of principal and secondary tensor parameters, and was explained in Section 3.4.2.2[link].

Table 3.4.3.1 | top | pdf |
Classification of domain pairs, ferroic phases and of switching (state shifts)

[P_{0i}^{(k)}] and [u_{0\mu}^{(k)}] are components of the spontaneous polarization and spontaneous strain in the domain state [{\bf S}_k], where [k=1] or [k=j]; similarly, [d_{i\mu}^{(k)} ] are components of the piezoelectric tensor, [{\varepsilon}_0{\kappa}_{ij}^{(k)} ] are components of electric susceptibility, [s_{\mu\nu}^{(k)}] are compliance components and [Q_{ij\mu}^{(k)}] are components of electrostriction (components with Greek indices are expressed in matrix notation) [see Chapter 1.1[link] or e.g. Nye (1985[link]) and Sirotin & Shaskolskaya (1982[link])]. Text in italics concerns the classification of ferroic phases. [{\bf E}] is the electric field and [{\sigma} ] is the mechanical stress.

Ferroic class Domain pair – at least in one pair Domain pair – phase Switching (state shift) Switching field
Primary At least one [P_{0i}^{(j)}-P_{0i}^{(1)}\neq 0 ] Ferroelectric Electrically first order E
At least one [u_{0\mu}^{(j)}-u_{0\mu}^{(1)}\neq 0 ] Ferroelastic Mechanically first order [{\sigma} ]
Secondary At least one [P_{0i}^{(j)}-P_{0i}^{(1)}\neq 0 ] and at least one [u_{0\mu}^{(j)}-u_{0\mu}^{(1)}\neq 0] Ferroelastoelectric Electromechanically first order E [\sigma]
  All [P_{0i}^{(j)}-P_{0i}^{(1)}= 0] and at least one [{\varepsilon}_0({\kappa}_{ik}^{(j)}-{\kappa}_{ik}^{(1)})\neq 0 ] Ferrobielectric Electrically second order EE
  All [u_{0\mu}^{(j)}-u_{0\mu}^{(1)}= 0 ] and at least one [s_{\mu\nu}^{(j)}-s_{\mu\nu}^{(1)}\neq 0] Ferrobielastic Mechanically second order [{\sigma\sigma}]
[\ldots] [\ldots] [\ldots] [\ldots] [\ldots]

[i,j = 1,2,3]; [\mu, \nu = 1,2, \ldots,6].

A discussion of and many examples of secondary ferroic phases are available in papers by Newnham & Cross (1974a[link],b[link]) and Newnham & Skinner (1976[link]), and tertiary ferroic phases are discussed by Amin & Newnham (1980[link]).

We shall now show that the tensor distinction of domain states is closely related to the switching of domain states by external fields.

3.4.3.3. Switching of ferroic domain states

| top | pdf |

We saw in Section 3.4.2.1[link] that all domain states of the orbit [G{\bf S}_1] have the same chance of appearing. This implies that they have the same free energy, i.e. they are degenerate. The same conclusion follows from thermodynamic theory, where domain states appear as equivalent solutions of equilibrium values of the order parameter, i.e. all domain states exhibit the same free energy [\Psi] (see Section 3.1.2[link] ). These statements hold under a tacit assumption of absent external electric and mechanical fields. If these fields are nonzero, the degeneracy of domain states can be partially or completely lifted.

The free energy [\Psi^{(k)}] per unit volume of a ferroic domain state [{\bf S}_k], [k=1,2,\ldots,n], with spontaneous polarization [{\bf P}_0^{(k)}] with components [P_{0i}^{(k)}], [i=1,2,3], and with spontaneous strain components [u_{0\mu}^{(k)}], [\mu=1,2,\ldots, 6, ] is (Aizu, 1972[link])[\eqalignno{\Psi^{(k)}&={\Psi}_0-P_{0i}^{(k)}E_{i}-u_{0\mu}^{(k)}\sigma_{\mu}- d_{i\mu}^{(k)}E_i\sigma_{\mu}-\textstyle{{1}\over{2}}{\varepsilon}_0{\kappa}_{ik}^{(k)}E_iE_k&\cr&\quad- \textstyle{{1}\over{2}}s_{\mu\nu}^{(k)}\sigma_{\mu}\sigma_{\nu}-\textstyle{{1}\over{2}} Q_{ik\mu}E_iE_{k}\sigma_{\mu}-\ldots, &(3.4.3.32)} ]where the Einstein summation convention (summation with respect to suffixes that occur twice in the same term) is used with [i,j=1,2,3] and [\mu,\nu=1,2,\ldots,6]. The symbols in equation (3.4.3.32[link]) have the following meaning: [E_{i}] and [u_{\mu}] are components of the external electric field and of the mechanical stress, respectively, [d_{i\mu}^{(k)}] are components of the piezoelectric tensor, [{\varepsilon}_0{\kappa}_{ij}^{(k)} ] are components of the electric susceptibility, [s_{\mu\nu}^{(k)} ] are compliance components, and [Q_{ij\mu}^{(k)}] are components of electrostriction (components with Greek indices are expressed in matrix notation) [see Section 3.4.5[link] (Glossary), Chapter 1.1[link] or Nye (1985[link]); Sirotin & Shaskolskaya (1982[link])].

We shall examine two domain states [{\bf S}_1] and [{\bf S}_j ], i.e. a domain pair [({\bf S}_1,{\bf S}_j)], in electric and mechanical fields. The difference of their free energies is given by[\eqalignno{\Psi^{(j)}-\Psi^{(1)}&=-(P_{0i}^{(j)}-P_{0i}^{(1)})E_{i}- (u_{0\mu}^{(j)}-u_{0\mu}^{(1)}){\sigma}_{\mu}-(d_{i\mu}^{(j)}-d_{i\mu}^{(1)})E_i{\sigma}_{\mu}&\cr &\quad-\textstyle{{1}\over{2}}{\varepsilon}_0({\kappa}_{ik}^{(j)}-{\kappa}_{ik}^{(1)})E_iE_k-\textstyle{{1}\over{2}}(s_{\mu\nu}^{(j)}-s_{\mu\nu}^{(1)})\sigma_{\mu}\sigma_{\nu}&\cr &\quad-\textstyle{{1}\over{2}}(Q_{ik\mu}^{(j)}-Q_{ik\mu}^{(1)})E_iE_k\sigma_{\mu}-\ldots. &(3.4.3.33)} ]

For a domain pair [({\bf S}_1,{\bf S}_j)] and given external fields, there are three possibilities:

  • (1) [{\Psi}^{(j)}=\Psi^{(1)}]. Domain states [{\bf S}_1] and [{\bf S}_j] can coexist in equilibrium in given external fields.

  • (2) [{\Psi}^{(j)} \,\lt\, \Psi^{(1)}]. In given external fields, domain state [{\bf S}_j] is more stable than [{\bf S}_1]; for large enough fields (higher than the coercive ones), the state [{\bf S}_1] switches into the state [{\bf S}_j].

  • (3) [{\Psi}^{(j)}\,\gt\,\Psi^{(1)}]. In given external fields, domain state [{\bf S}_j] is less stable than [{\bf S}_1]; for large enough fields (higher than the coercive ones), the state [{\bf S}_j] switches into the state [{\bf S}_1].

A typical dependence of applied stress and corresponding strain in ferroelastic materials has a form of a elastic hysteresis loop (see Fig. 3.4.1.3[link]). Similar dielectric hysteresis loops are observed in ferroelectric materials; examples can be found in books on ferroelectric crystals (e.g. Jona & Shirane, 1962[link]).

A classification of switching (state shifts in Aizu's terminology) based on equation (3.4.3.33[link]) was put forward by Aizu (1972[link], 1973[link]) and is summarized in the second and fourth columns of Table 3.4.3.1[link]. The order of the state shifts specifies the switching fields that are necessary for switching one domain state of a domain pair into the second state of the pair.

Another distinction related to switching distinguishes between actual and potential ferroelectric (ferroelastic) phases, depending on whether or not it is possible to switch the spontaneous polarization (spontaneous strain) by applying an electric field (mechanical stress) lower than the electrical (mechanical) breakdown limit under reasonable experimental conditions (Wadhawan, 2000[link]). We consider in our classification always the potential ferroelectric (ferroelastic) phase.

A closer look at equation (3.4.3.33[link]) reveals a correspondence between the difference coefficients in front of products of field components and the tensor distinction of domain states [{\bf S}_1] and [{\bf S}_j ] in the domain pair [({\bf S}_1,{\bf S}_j)]: If a morphic Cartesian tensor component of a polar tensor is different in these two domain states, then the corresponding difference coefficient is nonzero and defines components of fields that can switch one of these domain states into the other. A similar statement holds for the symmetric tensors of rank two (e.g. the spontaneous strain tensor).

Tensor distinction for all representative non-ferroelastic domain pairs is available in the synoptic Table 3.4.3.4[link]. These data also carry information about the switching fields.

3.4.3.4. Classes of equivalent domain pairs and their classifications

| top | pdf |

Two domain pairs that are crystallographically equivalent, [({\bf S}_i,{\bf S}_k)\,{\buildrel{G}\over {\sim}}\,({\bf S}_l,{\bf S}_m) ] [see equation (3.4.3.5[link])], have different orientations in space but their inherent properties are the same. It is, therefore, useful to divide all domain pairs of a ferroic phase into classes of equivalent domain pairs. All domain pairs that are equivalent (in G) with a given domain pair, say [({\bf S}_i,{\bf S}_k)], can be obtained by applying to [({\bf S}_i,{\bf S}_k)] all operations of G, i.e. by forming a G-orbit [G({\bf S}_i,{\bf S}_k)].

One can always find in this orbit a domain pair [({\bf S}_1,{\bf S}_j) ] that has in the first place the first domain state [{\bf S}_1]. We shall call such a pair a representative domain pair of the orbit. The initial orbit [G({\bf S}_i,{\bf S}_k)] and the orbit [G({\bf S}_1,{\bf S}_j) ] are identical: [G({\bf S}_i,{\bf S}_k)=G({\bf S}_1,{\bf S}_j). ]

The set [{\sf P}] of [n^2] ordered pairs (including trivial ones) that can be formed from n domain states can be divided into G-orbits (classes of equivalent domain pairs): [{\sf P} = G({\bf S}_1,{\bf S}_1) \cup G({\bf S}_1,g_2{\bf S}_1) \cup\ldots\cup ({\bf S}_1,g_j{\bf S}_1) \cup\ldots\cup G({\bf S}_1,g_q{\bf S}_1).\eqno(3.4.3.34) ]

Similarly, as there is a one-to-one correspondence between domain states and left cosets of the stabilizer (symmetry group) [F_1] of the first domain state [see equation (3.4.2.9[link])], there is an analogous relation between G-orbits of domain pairs and so-called double cosets of [F_1].

A double coset [F_1g_jF_1] of [F_1] is a set of left cosets that can be expressed as [fg_jF_1], where [f\in F_1 ] runs over all operations of [F_1] (see Section 3.2.3.2.8[link] ). A group G can be decomposed into disjoint double cosets of [F_1\subset G]: [\displaylines{G=F_{1}eF_1 \cup F_{1}g_2F_{1} \cup\ldots\cup F_{1}g_jF_{1}\cup\ldots \cup F_{1}g_qF_{1},\cr\hfill j=1,2,\ldots, q,\hfill(3.4.3.35)} ]where [g_1=e,g_2,{\ldots}g_j,{\ldots}g_q] is the set of representatives of double cosets.

There is a one-to-one correspondence between double cosets of the decomposition (3.4.3.35[link]) and G-orbits of domain pairs (3.4.3.34[link]) (see Section 3.2.3.3.6[link] , Proposition 3.2.3.35[link] ): [G({\bf S}_1,{\bf S}_j)\leftrightarrow F_{1}g_jF_{1}, \ \ {\rm where} \ \ {\bf S}_j=g_j{\bf S}_1, \quad j=1,2,{\ldots},q.\eqno(3.4.3.36) ]

We see that the representatives [g_j] of the double cosets in decomposition (3.4.3.35[link]) define domain pairs [({\bf S}_1,g_j{\bf S}_1) ] which represent all different G-orbits of domain pairs. Just as different left cosets [g_iF_1] specify all domain states, different double cosets determine all classes of equivalent domain pairs (G-orbits of domain pairs).

The properties of double cosets are reflected in the properties of corresponding domain pairs and provide a natural classification of domain pairs. A specific property of a double coset is that it is either identical or disjoint with its inverse. A double coset that is identical with its inverse, [(F_1g_{j}F_1)^{-1}=F_1g_j^{-1}F_1=F_1g_jF_1,\eqno(3.4.3.37)]is called an invertible (ambivalent) double coset. The corresponding class of domain pairs consists of transposable (ambivalent) domain pairs.

A double coset that is disjoint with its inverse, [(F_1g_{j}F_1)^{-1}=F_1g_j^{-1}F_1\cap F_1g_jF_1=\emptyset, \eqno(3.4.3.38) ]is a non-invertible (polar) double coset ([\emptyset] denotes an empty set) and the corresponding class of domain pairs comprises non-transposable (polar) domain pairs. A double coset [F_1g_jF_1] and its inverse [(F_1g_{j}F_1)^{-1}] are called complementary double cosets. Corresponding classes called complementary classes of equivalent domain pairs consist of transposed domain pairs that are non-equivalent.

Another attribute of a double coset is the number of left cosets which it comprises. If an invertible double coset consists of one left coset, [F_1g_jF_1=g_jF_1=(g_jF_1)^{-1}, \eqno(3.4.3.39) ]then the domain pairs in the G-orbit [G({\bf S}_1,g_j{\bf S}_1) ] are completely transposable. An invertible double coset comprising several left cosets is associated with a G-orbit consisting of partially transposable domain pairs. Non-invertible double cosets can be divided into simple non-transposable double cosets (complementary double cosets consist of one left coset each) and multiple non-transposable double cosets (complementary double cosets comprise more than one left coset each).

Thus there are four types of double cosets (see Table 3.2.3.1[link] in Section 3.2.3.2[link] ) to which there correspond the four basic types of domain pairs presented in Table 3.4.3.2[link].

Table 3.4.3.2 | top | pdf |
Four types of domain pairs

[F_{1j}] [J_{1j}] [K_{1j}] Double coset Domain pair name symbol
[F_{1}=F_j ] [F_{1}\cup g_{1j}^{\star}F_1] [F_{1}\cup g_{1j}^{\star}F_1] [F_1g_{1j}F_1=g_{1j}F_1=(g_{1j}F_1)^{-1} ] [\underline{t}]ransposable [\underline{c} ]ompletely tc
[F_{1j}\subset F_1 ] [F_{1j}\cup g_{1j}^{\star}F_{1j} ] [F_{1}\cup g_{1j}^{\star}F_{1}\cup\,\ldots \, ] [F_1g_{1j}F_1=(F_1g_{1j}F_1)^{-1}] [\underline{t}]ransposable [\underline{p} ]artially tp
[F_{1}=F_j ] [F_{1} ] [F_{1}\cup g_{1j}F_{1}\cup g_{1j}^{-1}F_{1} ] [F_1g_{1j}F_1=g_{1j}F_1\cap (g_{1j}F_1)^{-1}=\emptyset ] [\underline{n}]on-transposable [\underline{s}]imple ns
[F_{1j}\subset F_1 ] [F_{1j} ] [F_{1}\cup g_{1j}F_{1}\cup (g_{1j}F_{1})^{-1}\cup\,\ldots ] [F_1g_{1j}F_1\cap (F_1g_{1j}F_1)^{-1}=\emptyset ] [\underline{n}]on-transposable [\underline{m}]ultiple nm

These results can be illustrated using the example of a phase transition with G = [4_z/m_zm_xm_{xy}\supset 2_xm_ym_z =] [F_1] with four domain states (see Fig. 3.4.2.2[link]). The corresponding four left cosets of [2_xm_ym_z] are given in Table 3.4.2.1[link]. Any operation from the first left coset (identical with [F_1]) transforms the second left coset into itself, i.e. this left coset is a double coset. Since it consists of an operation of order two, it is a simple invertible double coset. The corresponding representative domain pair is [({\bf S}_1,\bar 1{\bf S}_1)=({\bf S}_1,{\bf S}_2) ]. By applying operations of [G=4_z/m_zm_xm_{xy}] on [({\bf S}_1,{\bf S}_2) ], one gets the class of equivalent domain pairs (G-orbit): [({\bf S}_1,{\bf S}_2)] [{\buildrel {G}\over {\sim}}] [({\bf S}_2,{\bf S}_1) ] [{\buildrel {G} \over {\sim}}] [({\bf S}_3,{\bf S}_4)] [{\buildrel {G} \over {\sim}}] [({\bf S}_4,{\bf S}_3)]. These domain pairs can be labelled as `180° pairs' according to the angle between the spontaneous polarization in the two domain states.

When one applies operations from the first left coset on the third left coset, one gets the fourth left coset, therefore a double coset consists of these two left cosets. An inverse of any operation of this double coset belongs to this double coset, hence it is a multiple invertible double coset. Corresponding domain pairs are partially transposable ones. A representative pair is, for example, [({\bf S}_1,2_{xy}{\bf S}_1)=] [({\bf S}_1,{\bf S}_3) ] which is indeed a partially transposable domain pair [cf. (3.4.3.19) and (3.4.3.20)]. The class of equivalent ordered domain pairs is [({\bf S}_1,{\bf S}_3) ] [{\buildrel {G}\over{\sim}}] [({\bf S}_3,{\bf S}_1)] [{\buildrel{G}\over {\sim}}] [({\bf S}_1,{\bf S}_4)] [{\buildrel{G}\over {\sim}} ] [({\bf S}_4,{\bf S}_1)] [{\buildrel{G}\over {\sim}}] [({\bf S}_3,{\bf S}_2)] [{\buildrel{G}\over {\sim}}] [({\bf S}_2,{\bf S}_3) ] [{\buildrel{G}\over {\sim}}] [({\bf S}_2,{\bf S}_4)] [{\buildrel{G}\over {\sim}}] [({\bf S}_4,{\bf S}_2)]. These are `90° domain pairs'.

An example of non-invertible double cosets is provided by the decomposition of the group [G=6_z/m_z] into left and double cosets of [F_1=2_z/m_z ] displayed in Table 3.4.3.3[link]. The inverse of the second left coset (second line) is equal to the third left coset (third line) and vice versa. Each of these two left cosets thus corresponds to a double coset and these double cosets are complementary double cosets. Corresponding representative simple non-transposable domain pairs are [({\bf S}_1,{\bf S}_2)] and [({\bf S}_2,] [{\bf S}_1)], and are depicted in Fig. 3.4.3.2[link].

Table 3.4.3.3 | top | pdf |
Decomposition of [G=6_z/m_z] into left cosets of [F_1=2_z/m_z ]

Left coset Principal domain state
1 [2_z] [\bar 1] [m_z] [{\bf S}_1]
[3_z ] [6_z^5 ] [{\bar 3}_z ] [{\bar 6}_z^5 ] [{\bf S}_2]
[3_z^2 ] [6_z ] [{\bar 3}_z^5 ] [{\bar 6}_z ] [{\bf S}_3]

We conclude that double cosets determine classes of equivalent domain pairs that can appear in the ferroic phase resulting from a phase transition with a symmetry descent [G\supset F_1]. Left coset and double coset decompositions for all crystallographic point-group descents are available in the software GI[\star]KoBo-1, path: Subgroups\View\Twinning groups.

A double coset can be specified by any operation belonging to it. This representation is not very convenient, since it does not reflect the properties of corresponding domain pairs and there are many operations that can be chosen as representatives of a double coset. It turns out that in a continuum description the twinning group [K_{1j}] can represent classes of equivalent domain pairs [G({\bf S}_1,{\bf S}_j)] with two exceptions:

  • (i) Two complementary classes of non-transposable domain pairs have the same twinning group. This follows from the fact that if a twinning group contains the double coset, then it must comprise also the inverse double coset.

  • (ii) Different classes of transposable domain pairs have different twining groups except in the following case (which corresponds to the orthorhombic ferroelectric phase in perov­skites): the group [F_1 =m_{x\bar y}2_{xy}m_z] generates with switching operations [g=2_{yz} ] and [g_3=m_{yz}] two different double cosets with the same twinning group [K_{12}=K_{13}=m\bar 3m] (one can verify this in the software GI[\star]KoBo-1, path: Subgroups\View\Twinning groups). Domain states are characterized in this ferroelectric phase by the direction of the spontaneous polarization. The angles between the spontaneous polarizations of the domain states in domain pairs [({\bf S}_1,2_{yz}{\bf S}_1)] and [({\bf S}_1,m_{yz}{\bf S}_1) ] are 120 and 60°, respectively; this shows that these representative domain pairs are not equivalent and belong to two different G-orbits of domain pairs. To distinguish these two cases, we add to the twinning group [m\bar 3m[m_{x\bar y}2_{xy}m_z]] either the switching operation [2_{yz}] or [m_{yz}], i.e. the two distinct orbits are labelled by the symbols [m\bar 3m(2_{xy})] and [m\bar 3m(m_{xy}) ], respectively.

Bearing in mind these two exceptions, one can, in the continuum description, represent G-orbits of domain pairs [G({\bf S}_1,{\bf S}_j) ] by twinning groups [K_{1j}[F_1]].

We have used this correspondence in synoptic Table 3.4.2.7[link] of symmetry descents at ferroic phase transitions. For each symmetry descent [G\supset F_1], the twinning groups given in column [K_{1j}] specify possible G-orbits of domain pairs that can appear in the domain structure of the ferroic phase (Litvin & Janovec, 1999[link]). We divide all orbits of domain pairs (represented by corresponding twinning groups [K_{1j}]) that appear in Table 3.4.2.7[link] into classes of non-ferroelastic and ferroelastic domain pairs and present them with further details in the three synoptic Tables 3.4.3.4[link], 3.4.3.6[link] and 3.4.3.7[link] described in Sections 3.4.3.5[link] and 3.4.3.6[link].

As we have seen, a classification of domain pairs according to their internal symmetry (summarized in Table 3.4.3.2[link]) introduces a partition of all domain pairs that can be formed from domain states of the G-orbit [G{\bf S}_1] into equivalence classes of pairs with the same internal symmetry. Similarly, any inherent physical property of domain pairs induces a partition of all domain pairs into corresponding equivalence classes. Thus, for example, the classification of domain pairs, based on tensor distinction or switching of domain states (see Table 3.4.3.1[link], columns two and three), introduces a division of domain pairs into corresponding equivalence classes.

3.4.3.5. Non-ferroelastic domain pairs: twin laws, domain distinction and switching fields, synoptic table

| top | pdf |

Two domain states [{\bf S}_1] and [{\bf S}_j] form a non-ferroelastic domain pair [({\bf S}_1,{\bf S}_j)] if the spontaneous strain in both domain states is the same, [{\bf u}_0^{(1)}={\bf u}_0^{(j)} ]. This is so if the twinning group [K_{1j}] of the pair and the symmetry group [F_1] of domain state [{\bf S}_1] belong to the same crystal family (see Table 3.4.2.2[link]): [{\rm Fam}K_{1j}={\rm Fam}F_{1}.\eqno(3.4.3.40)]

It can be shown that all non-ferroelastic domain pairs are completely transposable domain pairs (Janovec et al., 1993[link]), i.e. [F_{1j}=F_1=F_j\eqno(3.4.3.41)]and the twinning group [K_{1j}] is equal to the symmetry group [J_{1j} ] of the unordered domain pair [see equation (3.4.3.24[link])]: [K_{1j}^{\star}=J_{1j}^{\star} = F_{1} \cup g^{\star}_{1j}F_{1}.\eqno(3.4.3.42) ](Complete transposability is only a necessary, but not a sufficient, condition of a non-ferroelastic domain pair, since there are also ferroelastic domain pairs that are completely transposable – see Table 3.4.3.6[link].)

The relation between domain states in a non-ferroelastic domain twin, in which two domain states coexist, is the same as that of a corresponding non-ferroelastic domain pair consisting of single-domain states. Transposing operations [g^{\star}_{1j}] are, therefore, also twinning operations.

Synoptic Table 3.4.3.4[link] lists representative domain pairs of all orbits of non-ferroelastic domain pairs. Each pair is specified by the first domain state [{\bf S}_1] with symmetry group [F_1] and by transposing operations [g^{\star}_{1j}] that transform [{\bf S}_1] into [{\bf S}_j], [{\bf S}_j =g^{\star}_{1j}{\bf S}_1 ]. Twin laws in dichromatic notation are presented and basic data for tensor distinction and switching of non-ferroelastic domains are given.

3.4.3.5.1. Explanation of Table 3.4.3.4[link]

| top | pdf |

The first three columns specify domain pairs.

Table 3.4.3.4 | top | pdf |
Non-ferroelastic domain pairs, domain twin laws and distinction of non-ferroelastic domains

[F_1]: symmetry of [{\bf S}_1]; [g_{1j}^{\star} ]: twinning operations of second order; [K_{1j}^{\star}]: twinning group signifying the twin law of domain pair [({\bf S}_1,g_{1j}^{\star}{\bf S}_1) ]; [J_{1j}^{\star}]: symmetry group of the pair; [\Gamma_{\alpha} ]: irreducible representation of [K_{1j}^{\star}]; [\epsilon ], [{\bi P}_i,\ldots], [{\bi Q}_{{ij}{\mu}} ]: components of property tensors (see Table 3.4.3.5[link]): [a|c]: number of distinct[|]equal nonzero tensor components of property tensors.

[F_1] [g_{1j}^{\star} ] [K_{1j}^{\star}=J_{1j}^{\star} ] [\Gamma_{\alpha}] Diffraction intensities [\epsilon] [{\bi P}_i ] [g_{\mu}] [{\bi d}_{i{\mu}} ] [A_{i\mu}] [{\bi s}_{\mu\nu} ] [{\bi Q}_{{ij}{\mu}} ]
1 [\bar1^{\star}] [\bar1^{\star}] [A_u] = [1|0] [3|0] [6|0] [18|0] [0|18] [0|21] [0|36]
[2_u] [\bar1^{\star}], [m^{\star}_u] [2_u/m_u^{\star}] [A_u] = [1|0 ] [1|0 ] [4|0 ] [8|0 ] [0|8 ] [0|13] [0|20]
[m_u] [\bar1^{\star}], [2^{\star}_u] [2_u^{\star}/m_u] [B_u] = [0|0] [2|0] [2|0] [10|0] [0|8] [0|13] [0|20]
[2_x2_y2_z] [\bar1^{\star}], [m^{\star}_x], [m^{\star}_y], [m^{\star}_z] [m_x^{\star}m_y^{\star}m_z^{\star}] [A_{u}] = [1|0 ] [0|0 ] [3|0 ] [3|0 ] [0|3 ] [0|9 ] [0|12]
[2_{x\bar{y}}2_{xy}2_z] [\bar1^{\star}], [m^{\star}_{xy} ], [m^{\star}_{x\bar{y}}], [m^{\star}_z] [m_{x\bar{y}}^{\star}m_{xy}^{\star}m_z^{\star} ] [A_{u}] [= ] [1|0 ] [0|0 ] [3|0 ] [3|0 ] [0|3 ] [0|9 ] [0|12]
[m_xm_y2_z] [\bar1^{\star}], [m^{\star}_z], [2^{\star}_x], [2^{\star}_y] [m_xm_ym_z^{\star}] [B_{1u}] [= ] [0|0 ] [1|0 ] [1|0 ] [5|0 ] [0|3 ] [0|9 ] [0|12]
[2_xm_ym_z] [\bar1^{\star}], [m^{\star}_x], [2^{\star}_y], [2^{\star}_z] [m_x^{\star}m_ym_z] [B_{1u}] = [0|0 ] [1|0 ] [1|0 ] [5|0 ] [0|3 ] [0|9 ] [0|12]
[m_x2_ym_z] [\bar1^{\star}], [m^{\star}_y], [2^{\star}_x], [2^{\star}_z] [m_xm_y^{\star}m_z] [B_{1u}] [= ] [0|0 ] [1|0 ] [1|0 ] [5|0 ] [0|3 ] [0|9 ] [0|12]
[m_{x\bar{y}}m_{xy}2_z] [\bar1^{\star}], [m^{\star}_z], [2^{\star}_{xy}], [2^{\star}_{x\bar{y}}] [m_{x\bar{y}}m_{xy}m_z^{\star}] [B_{1u}] [= ] [0|0 ] [1|0 ] [1|0 ] [5|0 ] [0|3 ] [0|9 ] [0|12]
[4_z] [\bar1^{\star}], [m^{\star}_z] [4_z/m_z^{\star}] [A_{u}] [= ] [1|0 ] [1|0 ] [2|0 ] [4|0 ] [0|4 ] [0|7 ] [0|10]
[4_z] [2_x^{\star}], [2^{\star}_y], [2^{\star}_{xy}], [2^{\star}_{x\bar y}] [4_z2_x^{\star}2_{xy}^{\star}] [A_{2}] [\not= ] [0|1 ] [1|0 ] [0|2 ] [3|1 ] [3|1 ] [1|6 ] [3|7]
[4_z] [m_x^{\star}], [m^{\star}_y], [m^{\star}_{xy}], [m^{\star}_{x\bar y}] [4_zm_x^{\star}m_{xy}^{\star}] [A_{2} ] [\not=] [1|0 ] [0|1 ] [2|0 ] [1|3 ] [3|1 ] [1|6 ] [3|7]
[\bar{4}_z ] [\bar1^{\star}], [m^{\star}_z] [4^{\star}_z/m_z^{\star}] [B_{u}] [= ] [0|0 ] [0|0 ] [2|0 ] [4|0 ] [0|4 ] [0|7 ] [0|10]
[\bar{4}_z] [m_{xy}^{\star}], [m^{\star}_{x\bar y} ], [2^{\star}_x], [2^{\star}_y] [\bar{4}_z2_x^{\star}m_{xy}^{\star}] [A_{2}] [\not= ] [0|0 ] [0|0 ] [1|1 ] [2|2 ] [3|1 ] [1|6 ] [3|7]
[\bar{4}_z] [m_x^{\star}], [m^{\star}_y], [2^{\star}_{xy}], [2^{\star}_{x\bar y}] [\bar{4}_zm_x^{\star}2_{xy}^{\star}] [A_{2}] [\not= ] [0|0 ] [0|0 ] [1|1 ] [2|2 ] [3|1 ] [1|6 ] [3|7]
[4_z/m_z] [m^{\star}_x], [m^{\star}_y], [m^{\star}_{xy}], [m^{\star}_{x\bar y}], [2_x^{\star}], [2^{\star}_y], [2^{\star}_{xy}], [ 2^{\star}_{x\bar y}] [4_z/m_zm_x^{\star}m_{xy}^{\star}] [A_{2g}] [\not=] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [3|1 ] [1|6 ] [3|7]
[4_z2_x2_{xy} ] [\bar1^{\star}], [m^{\star}_z], [m_x^{\star}], [m^{\star}_y], [m^{\star}_{xy}], [m^{\star}_{x\bar y} ] [4_z/m_z^{\star}m_x^{\star}m_{xy}^{\star} ] [A_{1u}] [= ] [1|0 ] [0|0 ] [2|0 ] [1|0 ] [0|1 ] [0|6 ] [0|7]
[4_zm_xm_{xy}] [\bar1^{\star}], [m^{\star}_z], [2_x^{\star}], [2^{\star}_y], [2^{\star}_{xy}], [2^{\star}_{x\bar y} ] [4_z/m_z^{\star}m_xm_{xy} ] [A_{2u}] [= ] [0|0 ] [1|0 ] [0|0 ] [3|0 ] [0|1 ] [0|6 ] [0|7]
[\bar{4}_z2_xm_{xy}] [\bar1^{\star}], [m^{\star}_z], [m_x^{\star}], [m^{\star}_y], [2^{\star}_{xy}], [2^{\star}_{x\bar y} ] [4_z^{\star}/m_z^{\star}m_x^{\star}m_{xy} ] [B_{1u}] [= ] [0|0 ] [0|0 ] [1|0 ] [2|0 ] [0|1 ] [0|6 ] [0|7]
[\bar{4}_zm_x2_{xy}] [\bar1^{\star}], [m^{\star}_z], [m^{\star}_{xy}], [m^{\star}_{x\bar y}], [2_x^{\star} ], [2^{\star}_y] [4_z^{\star}/m_z^{\star}m_xm_{xy}^{\star} ] [B_{1u}] [= ] [0|0 ] [0|0 ] [1|0 ] [2|0 ] [0|1 ] [0|6 ] [0|7]
[3_v] [\bar{1}^{\star} ] [\bar{3}_v^{\star} ] [A_{u}] [= ] [1|0 ] [1|0 ] [2|0 ] [6|0 ] [0|6 ] [0|7 ] [0|12]
[3_z ] [2_x^{\star}], [2^{\star}_{x{^\prime}} ], [2^{\star}_{x{^\prime}{^\prime}}] [3_z2_x^{\star} ] [A_{2}] [\not=] [0|1 ] [1|0 ] [0|2 ] [4|2 ] [4|2] [1|6 ] [4|8]
[3_z] [2_y^{\star}], [2^{\star}_{y{^\prime}} ], [2^{\star}_{y{^\prime}{^\prime}}] [3_z2_y^{\star} ] [A_{2}] [\not=] [0|1 ] [1|0 ] [0|2 ] [4|2 ] [4|2 ] [1|6 ] [4|8]
[3_p] [2_{x\bar{y}}^{\star}], [2^{\star}_{y\bar{z}} ], [2^{\star}_{z\bar{x}}] [3_p2_{x\bar{y}}^{\star} ] [A_{2}] [\not=] [0|1 ] [1|0 ] [0|2 ] [4|2 ] [4|2] [1|6 ] [4|8]
[3_z ] [m_x^{\star}], [m^{\star}_{x{^\prime}} ], [m^{\star}_{x{^\prime}{^\prime}}] [3_zm_x^{\star} ] [A_{2}] [\not=] [1|0 ] [0|1 ] [2|0 ] [2|4 ] [4|2 ] [1|6 ] [4|8]
[3_z] [m_y^{\star}], [m^{\star}_{y{^\prime}} ], [m^{\star}_{y{^\prime}{^\prime}}] [3_zm_y^{\star} ] [A_{2}] [\not= ] [1|0 ] [0|1 ] [2|0 ] [2|4 ] [4|2 ] [1|6 ] [4|8]
[3_p ] [m_{x\bar{y}}^{\star}], [m^{\star}_{y\bar{z}} ], [m^{\star}_{z\bar{x}}] [3_pm_x^{\star} ] [A_{2}] [\not=] [1|0 ] [0|1 ] [2|0 ] [2|4 ] [4|2 ] [1|6 ] [4|8]
[3_z] [2_z^{\star} ] [6_z^{\star} ] B [\not=] [0|1 ] [0|1 ] [0|2 ] [2|4 ] [2|4 ] [2|5 ] [4|8]
[3_z ] [m_z^{\star} ] [\bar{6}_z^{\star} ] [A^{{^\prime}{^\prime}}] [\not= ] [1|0 ] [1|0 ] [2|0 ] [4|2 ] [2|4 ] [2|5 ] [4|8]
[\bar{3}_z] [m_x^{\star}], [m^{\star}_{x{^\prime}} ], [m^{\star}_{x{^\prime}{^\prime}}], [2_x^{\star} ], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}}] [\bar{3}_zm_x^{\star}] [A_{2g}] [\not=] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [4|2 ] [1|6 ] [4|8]
[\bar{3}_z] [m_y^{\star}], [m^{\star}_{y{^\prime}} ], [m^{\star}_{y{^\prime}{^\prime}}], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}}] [\bar{3}_zm_y^{\star}] [A_{2g}] [\not= ] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [4|2 ] [1|6 ] [4|8]
[\bar{3}_p ] [m_{x\bar{y}}^{\star}], [m^{\star}_{y\bar{z}} ], [m^{\star}_{z\bar{x}}], [2_{x\bar{y}}^{\star}], [2^{\star}_{y\bar{z}}], [2^{\star}_{z\bar{x}}] [\bar{3}_pm_x^{\star}] [A_{2g}] [\not=] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [4|2 ] [1|6 ] [4|8]
[\bar{3}_z] [m^{\star}_z], [2_z^{\star}] [6_z^{\star}/m_z^{\star}] [B_{g} ] [\not= ] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [2|4 ] [2|5] [4|8]
[3_z2_x ] [\bar1^{\star}], [m_x^{\star}], [m^{\star}_{x{^\prime}}], [m^{\star}_{x{^\prime}{^\prime}}] [\bar{3}_z^{\star}m_x^{\star}] [A_{1u}] [= ] [1|0 ] [0|0] [2|0 ] [2|0 ] [0|2 ] [0|6 ] [0|8]
[3_z2_y ] [\bar1^{\star}], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}}] [\bar{3}_z^{\star}m_y^{\star}] [A_{1u}] = [1|0 ] [0|0] [2|0 ] [2|0 ] [0|2 ] [0|6 ] [0|8]
[3_z2_x ] [2_z^{\star}], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}}] [6_z^{\star}2_x2_y^{\star}] [B_{1}] [\not= ] [0|1 ] [0|0 ] [0|2 ] [1|1 ] [1|1 ] [1|5] [2|6]
[3_z2_y] [2_z^{\star}], [2_x^{\star}], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}}] [6_z^{\star}2_x^{\star}2_y ] [B_{1}] [\not=] [0|1 ] [0|0 ] [0|2 ] [1|1 ] [1|1 ] [1|5 ] [2|6]
[3_p2_{x\bar{y}}] [\bar1^{\star}], [m_{x\bar{y}}^{\star} ], [m^{\star}_{y\bar{z}}], [m^{\star}_{z\bar{x}}] [\bar{3}_p^{\star}m_x^{\star}] [A_{1u}] [= ] [1|0] [0|0 ] [2|0 ] [2|0 ] [0|2 ] [0|6 ] [0|8]
[3_z2_x ] [m_z^{\star}], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}}] [\bar{6}_z^{\star}2_xm_y^{\star}] [A^{{^\prime}{^\prime}}_{1}] [\not= ] [1|0 ] [0|0] [2|0 ] [1|1 ] [1|1 ] [1|5 ] [2|6]
[3_z2_y ] [m_z^{\star}], [m_x^{\star}], [m^{\star}_{x{^\prime}}], [m^{\star}_{x{^\prime}{^\prime}}] [\bar{6}_z^{\star}m_x2_y^{\star}] [A^{{^\prime}{^\prime}}_{1}] [\not= ] [1|0 ] [0|0] [2|0 ] [1|1 ] [1|1 ] [1|5 ] [2|6]
[3_pm_{x\bar{y}} ] [\bar1^{\star}], [2_{x\bar{y}}^{\star} ], [2^{\star}_{y\bar{z}}], [2^{\star}_{z\bar{x}}] [\bar{3}_p^{\star}m_x ] [A_{2u}] = [0|0 ] [1|0 ] [0|0 ] [4|0 ] [0|2 ] [0|6 ] [0|8]
[3_zm_x ] [\bar1^{\star}], [2_x^{\star}], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}}] [\bar{3}_z^{\star}m_x ] [A_{2u}] = [0|0 ] [1|0 ] [0|0 ] [4|0 ] [0|2 ] [0|6 ] [0|8]
[3_zm_y] [\bar1^{\star}], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}}] [\bar{3}_z^{\star}m_y ] [A_{2u}] = [0|0 ] [1|0 ] [0|0 ] [4|0 ] [0|2 ] [0|6 ] [0|8]
[3_zm_x ] [2_z^{\star}], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}}] [6_z^{\star}m_xm_y^{\star} ] [B_{2}] [\not=] [0|0 ] [0|1 ] [0|0 ] [1|3 ] [1|1 ] [1|5 ] [2|6]
[3_zm_y] [m_x^{\star}], [m^{\star}_{x{^\prime}} ], [m^{\star}_{x{^\prime}{^\prime}}] [6_z^{\star}m_x^{\star}m_y ] [B_{2}] [\not=] [0|0 ] [0|1 ] [0|0 ] [1|3 ] [1|1 ] [1|5 ] [2|6]
[3_zm_x ] [m_z^{\star}], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}}] [\bar{6}_z^{\star}m_x2_y^{\star}] [A_{2}^{{^\prime}{^\prime}}] [\not=] [0|0 ] [1|0 ] [0|0 ] [3|1 ] [1|1 ] [1|5 ] [2|6]
[3_zm_y ] [m_z^{\star}], [2_x^{\star}], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}}] [\bar{6}_z^{\star}2_x^{\star}m_y] [A_{2}^{{^\prime}{^\prime}}] [\not=] [0|0 ] [1|0 ] [0|0 ] [3|1 ] [1|1 ] [1|5 ] [2|6]
[\bar{3}_zm_x ] [m_z^{\star}], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}}] [6_z^{\star}/m_z^{\star}m_xm_y^{\star} ] [B_{1g}] [\not=] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [1|1 ] [1|5 ] [2|6]
[\bar{3}_zm_y ] [m_z^{\star}], [m_x^{\star}], [m^{\star}_{x{^\prime}}], [m^{\star}_{x{^\prime}{^\prime}}] [6_z^{\star}/m_z^{\star}m_x^{\star}m_y ] [B_{1g}] [\not= ] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [1|1 ] [1|5 ] [2|6]
[6_z ] [\bar1^{\star}], [m^{\star}_z] [6_z/m_z^{\star} ] [A_{u}] [= ] [1|0 ] [1|0 ] [2|0 ] [4|0 ] [0|4 ] [0|5 ] [0|8]
[6_z ] [2_x^{\star}], [2^{\star}_{x{^\prime}} ], [2^{\star}_{x{^\prime}{^\prime}}], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}}] [6_z2_x^{\star}2_y^{\star} ] [A_{2}] [\not=] [0|1 ] [1|0 ] [0|2 ] [3|1 ] [3|1 ] [0|5 ] [2|6]
[6_z] [m_x^{\star}], [m^{\star}_{x{^\prime}} ], [m^{\star}_{x{^\prime}{^\prime}}], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}} ] [6_zm_x^{\star}m_y^{\star} ] [A_{2}] [\not=] [1|0 ] [0|1 ] [2|0 ] [1|3 ] [3|1 ] [0|5 ] [2|6]
[\bar{6}_z ] [\bar1^{\star}], [2^{\star}_z] [6_z^{\star}/m_z ] [B_{u}] [= ] [0|0 ] [0|0 ] [0|0 ] [2|0 ] [0|4 ] [0|5 ] [0|8]
[\bar{6}_z ] [m_x^{\star}], [m^{\star}_{x{^\prime}} ], [m^{\star}_{x{^\prime}{^\prime}}], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}} ] [\bar{6}_zm_x^{\star}2_y^{\star}] [A_{2}^{{^\prime}}] [\not= ] [0|0 ] [0|0 ] [0|0 ] [1|1 ] [3|1 ] [0|5 ] [2|6]
[\bar{6}_z ] [m_y^{\star}], [m^{\star}_{y{^\prime}} ], [m^{\star}_{y{^\prime}{^\prime}}], [2_x^{\star}], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}} ] [\bar{6}_z2_x^{\star}m_y^{\star}] [A_{2}^{{^\prime}}] [\not= ] [0|0 ] [0|0 ] [0|0 ] [1|1 ] [3|1 ] [0|5 ] [2|6]
[6_z/m_z ] [m_x^{\star}], [m^{\star}_{x{^\prime}} ], [m^{\star}_{x{^\prime}{^\prime}}], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}}], [2_x^{\star}], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}} ], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}} ] [6_z/m_zm_x^{\star}m_y^{\star} ] [A_{2g}] [\not=] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [3|1 ] [0|5 ] [2|6]
[6_z2_x2_y ] [\bar1^{\star}], [m_z^{\star}], [m_x^{\star}], [m^{\star}_{x{^\prime}}], [m^{\star}_{x{^\prime}{^\prime}} ], [m_y^{\star}], [m^{\star}_{y{^\prime}}][, m^{\star}_{y{^\prime}{^\prime}} ] [6_z/m_z^{\star}m_x^{\star}m^{\star}_y ] [A_{1u}] [= ] [1|0 ] [0|0 ] [2|0 ] [1|0 ] [0|1 ] [0|5 ] [0|6]
[6_zm_xm_y ] [\bar1^{\star}], [m^{\star}_z], [2_x^{\star}], [2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}} ], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}} ] [6_z/m_z^{\star}m_xm_y ] [A_{2u}] = [0|0 ] [1|0 ] [0|0 ] [3|0 ] [0|1 ] [0|5 ] [0|6]
[\bar{6}_z2_xm_y ] [\bar1^{\star}], [2^{\star}_z], [m_x^{\star}], [m^{\star}_{x{^\prime}}], [m^{\star}_{x{^\prime}{^\prime}} ], [2_y^{\star}], [2^{\star}_{y{^\prime}}], [2^{\star}_{y{^\prime}{^\prime}} ] [6_z^{\star}/m_zm_x^{\star}m_y] [B_{2u}] [= ] [0|0 ] [0|0 ] [0|0 ] [1|0 ] [0|1 ] [0|5 ] [0|6]
[\bar{6}_zm_x2_y ] [\bar1^{\star}], [2^{\star}_z], [m_y^{\star}], [m^{\star}_{y{^\prime}}], [m^{\star}_{y{^\prime}{^\prime}} ], [2_x^{\star}, 2^{\star}_{x{^\prime}}], [2^{\star}_{x{^\prime}{^\prime}} ] [6_z^{\star}/m_zm_xm_y^{\star}] [B_{2u}] [= ] [0|0 ] [0|0 ] [0|0 ] [1|0 ] [0|1 ] [0|5 ] [0|6]
23 [\bar1^{\star}], [m_x^{\star}], [m^{\star}_y], [m^{\star}_z] [m^{\star}\bar{3} ] [A_{u}] = [1|0 ] [0|0 ] [1|0 ] [1|0 ] [0|1 ] [0|3 ] [0|4]
23 [2_{xy}^{\star}], [2_{yz}^{\star} ], [2_{zx}^{\star}], [2^{\star}_{x\bar y}], [2^{\star}_{y\bar z} ], [2^{\star}_{z\bar x}] [4^{\star}32^{\star}] [A_{2}] [\not= ] [0|1 ] [0|0 ] [0|1] [1|0 ] [1|0 ] [0|3 ] [1|3]
23 [m_{xy}^{\star}], [m_{yz}^{\star} ], [m_{zx}^{\star}], [m^{\star}_{x\bar y}], [m^{\star}_{y\bar z} ], [m^{\star}_{z\bar x}] [\bar{4}^{\star}3m^{\star} ] [A_{2}] [\not= ] [1|0 ] [0|0 ] [1|0 ] [0|1 ] [1|0 ] [0|3 ] [1|3]
[m\bar{3}] [m_{xy}^{\star}], [m_{yz}^{\star} ], [m_{zx}^{\star}], [m^{\star}_{x\bar y}], [m^{\star}_{y\bar z} ], [m^{\star}_{z\bar x}], [2^{\star}_{xy}], [2^{\star}_{yz} ], [2^{\star}_{zx}], [2^{\star}_{x\bar y}], [2^{\star}_{y\bar z} ], [2^{\star}_{z\bar x}] [m\bar{3}m^{\star}] [A_{2g}] [\not=] [0|0 ] [0|0 ] [0|0 ] [0|0 ] [1|0 ] [0|3 ] [1|3]
432 [\bar1^{\star}], [m_x^{\star}], [m^{\star}_y], [m^{\star}_z], [m_{xy}^{\star}], [m_{yz}^{\star} ], [m_{zx}^{\star}], [m^{\star}_{x\bar y}], [m^{\star}_{y\bar z} ], [m^{\star}_{z\bar x}] [m^{\star}\bar{3}m^{\star}] [A_{1u}] [= ] [1|0] [0|0] [1|0] [0|0] [0|0] [0|3] [0|3]
[\bar{4}3m ] [\bar1^{\star}], [m_x^{\star}], [m^{\star}_y], [m^{\star}_z], [2^{\star}_{xy}], [2^{\star}_{yz} ], [2^{\star}_{zx}], [2^{\star}_{x\bar y} ], [2^{\star}_{y\bar z} ], [2^{\star}_{z\bar x}] [m^{\star}\bar{3}m] [A_{2u}] [= ] [0|0] [0|0] [0|0] [1|0] [0|0] [0|3] [0|3]

  • [F_1]: point-group symmetry (stabilizer in [K_{1j}]) of the first domain state [{\bf S}_1)] in a single-domain orientation. There are two domain states with the same [F_1]; one has to be chosen as [{\bf S}_1]. Subscripts of generators in the group symbol specify their orientation in the Cartesian (rectangular) crystallophysical coordinate system of the group [K_{1j}] (see Tables 3.4.2.5[link], 3.4.2.6[link] and Figs. 3.4.2.3[link], 3.4.2.4[link]).

  • [g_{1j}^{\star}]: switching operations that specify domain pair [({\bf S}_1,g_{1j}^{\star}{\bf S}_1)=] [({\bf S}_1,{\bf S}_j)]. Subscripts of symmetry operations specify the orientation of the corresponding symmetry element in the Cartesian (rectangular) crystallophysical coordinate system of the group [K_{1j} ]. In hexagonal and trigonal systems, [x', y'] and [x'', y'' ] denote the Cartesian coordinate system rotated about the z axis through 120 and 240°, respectively, from the Cartesian coordinate axes x and y; diagonal directions are abbreviated: [p =[111]], [q=[\bar 1\bar 11]], [r=[1\bar 1\bar 1]], [s=[\bar 1 1 \bar 1]] (for further details see Tables 3.4.2.5[link] and 3.4.2.6[link], and Figs. 3.4.2.3[link] and 3.4.2.4[link]).

    All switching operations of the second order are given, switching operations of higher order are omitted. The star symbol signifies that the operation is both a transposing and a twinning operation.

  • [K_{1j}^{\star}=J_{1j}^{\star}]: twinning group of the domain pair [({\bf S}_1,{\bf S}_j)]. This group is equal to the symmetry group [J_{1j}^{\star}] of the completely transposable unordered domain pair [\{{\bf S}_1,{\bf S}_j\}] [see equation (3.4.3.24[link])]. The dichromatic symbol of the group [K^{\star}_{1j}=J_{1j}^{\star} ] designates the twin law of the non-ferroelastic domain pair [\{{\bf S}_1,{\bf S}_j \}] and the twin law of all non-ferroelastic twins with domains containing [{\bf S}_1] and [{\bf S}_j] (see Section 3.4.3.1[link]).

The second part of the table concerns the distinction and switching of domain states of the non-ferroelastic domain pair [({\bf S}_1,{\bf S}_j) =] [({\bf S}_1,g_{1j}^{\star}{\bf S}_1)].

  • [{\Gamma}_{\alpha} ]: irreducible representation of [K_{1j}] that defines the transformation properties of the principal tensor parameters of the symmetry descent [K_{1j}\supset F_1] and thus specifies the components of principal tensor parameters that are given explicitly in Table 3.1.3.1[link] , in the software GI[\star]KoBo-1 and in Kopský (2001[link]), where one replaces G by [K_{1j}].

  • Diffraction intensities : the entries in this column characterize the differences of diffraction intensities from two domain states of the domain pair:

    = signifies that the twinning operations belong to the Laue class of [F_1]. Then the reflection intensities per unit volume are the same for both domain states if anomalous scattering is zero, i.e. if Friedel's law is valid. For nonzero anomalous scattering, the intensities from the two domain states differ, but when the partial volumes of both states are equal the diffraction pattern is centrosymmetric;

    [\neq] signifies that the twinning operations do not belong to the Laue class of [F_1]. Then the reflection intesities per unit volume of the two domain states are different [for more details, see Chapter 3.3[link] ; Catti & Ferraris (1976[link]); Koch (2004[link])].

  • [\epsilon], [{\bi P}_i], [g_{\mu},\ldots, ] [{\bi Q}_{ij\mu}]: components (in matrix notation) of important property tensors that are specified in Table 3.4.3.5[link]. The same symbol may represent several property tensors (given in the same row of Table 3.4.3.5[link]) of the same rank and intrinsic symmetry. Bold-face symbols signify polar tensors. For each type of property tensor two numbers [a|c] are given; number a in front of the vertical bar | is the number of independent covariant components (in most cases identical with Cartesian components) that have the same absolute value but different sign in domain states [{\bf S}_1 ] and [{\bf S}_j]. The number c after the vertical bar | gives the number of independent nonzero tensor parameters that have equal values in both domain states of the domain pair [({\bf S}_1,{\bf S}_j)]. These tensor components are already nonzero in the parent phase.

    The principal tensor parameters are one-dimensional and have the same absolute value but opposite sign in [{\bf S}_1] and [{\bf S}_j=g^{\star}_{1j}{\bf S}_1]. Principal tensor parameters for symmetry descents [K_{1j}\supset F_1] and the associated [\Gamma_{\alpha} ] of all non-ferroelastic domain pairs can be found for property tensors of lower rank in Table 3.1.3.1[link] and for all tensors appearing in Table 3.4.3.4[link] in the software GI[\star]KoBo-1 and in Kopský (2001[link]), where one replaces G by [K_{1j}].

    When [a\neq 0] for a polar tensor (in bold-face components), then switching fields exist in the combination given in the last column of Table 3.4.3.5[link]. Components of these fields can be determined from the explicit form of corresponding principal tensor parameters expressed in Cartesian components.

Table 3.4.3.5[link] lists important property tensors up to fourth rank. Property tensor components that appear in the column headings of Table 3.4.3.4[link] are given in the first column, where bold face is used for the polar tensors significant for specifying the switching fields appearing in schematic form in the last column. In the third and fourth columns, those propery tensors appear for which hold all the results presented in Table 3.4.3.4[link] for the symbols given in the first column of Table 3.4.3.5[link].

Table 3.4.3.5 | top | pdf |
Property tensors and switching fields

[i =1,2,3]; [{\mu},{\nu} =1,2,\ldots,6].

Symbol Property tensor Symbol Property tensor Switching fields
[\epsilon] Enantiomorphism   Chirality  
[{\bi P}_i] Polarization [p_i] Pyroelectricity E
[\boldvarepsilon_{ij}] Permittivity     EE
[g_{\mu}] Optical activity      
[{\bi d}_{i{\mu}}] Piezoelectricity [r_{ijk}] Electro-optics Eu
[A_{i{\mu}}] Electrogyration      
[{\bi s}_{\mu\nu}] Elastic compliances     uu
[{\bi Q}_{{ij}{\mu}}] Electrostriction [\pi_{{ij}{\mu}}] Piezo-optics EEu

We turn attention to Section 3.4.5[link] (Glossary), which describes the difference between the notation of tensor components in matrix notation given in Chapter 1.1[link] and those used in the software GI[\star]KoBo-1 and in Kopský (2001[link]).

The numbers a in front of the vertical bar | in Table 3.4.3.4[link] provide global information about the tensor distinction of two domain states and enables one to classify domain pairs. Thus, for example, the first number a in column [P_i] gives the number of nonzero components of the spontaneous polarization that differ in sign in both domain states; if [a\neq 0], this domain pair can be classified as a ferroelectric domain pair.

Similarly, the first number a in column [g_{\mu}] determines the number of independent components of the tensor of optical activity that have opposite sign in domain states [{\bf S}_1] and [{\bf S}_j]; if [a\neq 0], the two domain states in the pair can be distinguished by optical activity. Such a domain pair can be called a gyrotropic domain pair. As in Table 3.4.3.1[link] for the ferroelectric (ferroelastic) domain pairs, we can define a gyrotropic phase as a ferroic phase with gyrotropic domain pairs. The corresponding phase transition to a gyrotropic phase is called a gyrotropic phase transition (Koňák et al., 1978[link]; Wadhawan, 2000[link]). If it is possible to switch gyrotropic domain states by an external field, the phase is called a ferrogyrotropic phase (Wadhawan, 2000[link]). Further division into full and partial subclasses is possible.

One can also define piezoelectric (electro-optic) domain pairs, electrostrictive (elasto-optic) domain pairs and corresponding phases and transitions.

As we have already stated, domain states in a domain pair [({\bf S}_1,{\bf S}_j) ] differ in principal tensor parameters of the transition [K_{1j}\supset F_1 ]. These principal tensor parameters are Cartesian tensor components or their linear combinations that transform according to an irreducible representation [\Gamma_{\alpha}] specifying the primary order parameter of the transition [K_{1j}\supset F_1] (see Section 3.1.3[link] ). Owing to a special form of [K_{1j}] expressed by equation (3.4.3.42[link]), this representation is a real one-dimensional irreducible representation of [K_{1j}]. Such a representation associates +1 with operations of [F_1 ] and −1 with operations from the left coset [g^{\star}_{1j} ]. This means that the principal tensor parameters are one-dimensional and have the same absolute value but opposite sign in [{\bf S}_1] and [{\bf S}_j=g^{\star}_{1j}{\bf S}_1]. Principal tensor parameters for symmetry descents [K_{1j}\supset F_1] and associated [\Gamma_{\alpha}]'s of all non-ferroelastic domain pairs can be found for property tensors of lower rank in Table 3.1.3.1[link] and for all tensors appearing in Table 3.4.3.5[link] in the software GI[\star]KoBo-1 and in Kopský (2001[link]).

These specific properties of non-ferroelastic domain pairs allow one to formulate simple rules for tensor distinction that do not use principal tensor parameters and that are applicable for property tensors of lower rank.

  • (i) Symmetry descents [K_{1j}\supset F_1 ] of non-ferroelastic domain pairs for lower-rank property tensors lead only to the appearance of independent Cartesian morphic tensor components and not to the breaking of relations between these components. These morphic Cartesian tensor components can be found by comparing matrices of property tensors in the twinning group [K_{1j}] and the low-symmetry group [F_1] as those components that appear in [F_1] but are zero in [K_{1j}].

  • (ii) As follows from Table 3.4.3.4[link], one can always find a twinning operation that is either inversion, or a twofold axis or a mirror plane with a prominent crystallographic orientation. By applying the method of direct inspection (see Section 1.1.4.6.3[link] ), one can in most cases easily find morphic Cartesian components in the second domain state of the domain pair considered and prove that they differ only in sign.

Example 3.4.3.4. Tensor distinction of domains and switching in lead germanate.  Lead germanate (Pb5Ge3O11) undergoes a phase transition with symmetry descent [G=\bar 6 \supset 3=F_1] for which we find in Table 3.4.2.7[link], column [K_{1j} ], just one twinning group [K_{1j}={\bar 6}^{\star}], i.e. [K_{1j}^{\star}=G]. This means that there is only one G-orbit of domain pairs. Since Fam3 = Fam[\bar 6] [see Table 3.4.2.2[link] and equation (3.4.3.40[link])] this orbit comprises non-ferroelastic domain pairs. In Table 3.4.3.4[link], we find for [F_1=3] and [F_{1j}^{\star}=\bar 6] that the two domain states differ in some components of all property tensors listed in this table. The first polar tensor is the spontaneous polarization (the pair is ferroelectric) with one component [(a=1)] that has opposite sign in the two domain states. In Table 3.1.3.1[link] , we find for [G (=K_{1j})=\bar6] and [F_1 = 3] that this component is [P_3=P_z ]. From Table 3.4.3.1[link], it follows that the state shift is electrically first order with switching field [{\bf E}=({0,0,E}_z)].

The first optical tensor, which could enable the visualization of the domain states, is the optical activity [g_\mu] with two independent components which have opposite sign in the two domain states. In the software GI[\star]KoBo-1, path: Subgroups\View\Domains or in Kopský (2001[link]) we find these components: [g_3, g_1+g_2]. Shur et al. (1989[link]) have visualized in this way the domain structure of lead germanate with excellent black and white contrast (see Fig. 3.4.3.3[link]). Other examples are given in Shuvalov & Ivanov (1964[link]) and especially in Koňák et al. (1978[link]).

[Figure 3.4.3.3]

Figure 3.4.3.3 | top | pdf |

Domain structure in lead germanate observed using a polarized-light microscope. Visualization based on the opposite sign of the optical activity coefficient in the two domain states. Courtesy of Vl. Shur, Ural State University, Ekaterinburg.

Table 3.4.3.4[link] can be used readily for twinning by merohedry [see Chapter 3.3[link] and e.g. Cahn (1954[link]); Koch (2004[link])], where it enables an easy determination of the tensor distinction of twin components and the specification of external fields for possible switching and detwinning.

Example 3.4.3.5. Tensor distinction and switching of Dauphiné twins in quartz.  Quartz undergoes a phase transition from [G=6_z2_x2_y] to [F_1=3_z2_x]. Using the same procedure as in the previous example, we come to following conclusions: There are only two domain states [{\bf S}_1], [{\bf S}_2] and the twinning group, expressing the twin law, is equal to the high-symmetry group [K_{12}^{\star}=6_z_x2_y]. In Table 3.4.3.4[link], we find that these two states differ in one independent component of the piezoelectric tensor and in one elastic compliance component. Comparison of the matrices for [6_z2_x2_y] and [3_z2_x] (see Sections 1.1.4.10.3[link] and 1.1.4.10.4[link] ) yields the following morphic tensor components in the first domain state [{\bf S}_1]: [d^{(1)}_{11} =] [-d^{(1)}_{12}=] [-2d^{(1)}_{26}] and [s^{(1)}_{14}=] [-s^{(1)}_{24} =] [2s^{(1)}_{56}]. According to the rule given above, the values of morphic components in the second domain state [{\bf S}_2] are [d^{(2)}_{11}=] [-d^{(1)}_{11} =] [-d^{(2)}_{12}=] [d^{(1)}_{12}=] [-2d^{(2)}_{26}=] [2d^{(1)}_{26}] and [s^{(2)}_{14}=] [-s^{(1)}_{14}=] [-s^{(2)}_{24}=] [s^{(1)}_{24}=] [2s^{(2)}_{56}=] [-2s^{(1)}_{56}] [see Section 3.4.5[link] (Glossary)]. These results show that there is an elastic state shift of second order and an electromechanical state shift of second order. Nonzero components [d_{14}=-d_{25}] in [6_z2_x2_y] are the same in both domain states. Similarly, one can find five independent components of the tensor [s_{\mu\nu} ] that are nonzero in [6_z2_x2_y] and equal in both domain states. For the piezo-optic tensor [\pi_{\mu\nu} ], one can proceed in a similar way. Aizu (1973[link]) has used the ferrobielastic character of the domain pairs for visualizing domains and realizing switching in quartz. Other methods for switching and visualizing domains in quartz are known (see e.g. Bertagnolli et al., 1978[link], 1979[link]).

3.4.3.6. Ferroelastic domain pairs

| top | pdf |

A ferroelastic domain pair consists of two domain states that have different spontaneous strain. A domain pair [({\bf S}_1,{\bf S}_j) ] is a ferroelastic domain pair if the crystal family of its twinning group [K_{1j}] differs from the crystal family of the symmetry group [F_1] of domain state [{\bf S}_1], [{\rm Fam}K_{1j}\neq {\rm Fam}F_1.\eqno(3.4.3.43)]

Before treating compatible domain walls and disorientations, we explain the basic concept of spontaneous strain.

3.4.3.6.1. Spontaneous strain

| top | pdf |

A strain describes a change of crystal shape (in a macroscopic description) or a change of the unit cell (in a microscopic description) under the influence of mechanical stress, temperature or electric field. If the relative changes are small, they can be described by a second-rank symmetric tensor [\bf u] called the Lagrangian strain. The values of the strain components [u_{ik},] [i,k=1,2,3] (or in matrix notation [u_{\mu},] [\mu=1, \ldots, 6]) can be calculated from the `undeformed' unit-cell parameters before deformation and `deformed' unit-cell parameters after deformation (see Schlenker et al., 1978[link]; Salje, 1990[link]; Carpenter et al., 1998[link]).

A spontaneous strain describes the change of an `undeformed' unit cell of the high-symmetry phase into a `deformed' unit cell of the low-symmetry phase. To exclude changes connected with thermal expansion, one demands that the parameters of the undeformed unit cell are those that the high-symmetry phase would have at the temperature at which parameters of the low-symmetry phase are measured. To determine these parameters directly is not possible, since the parameters of the high-symmetry phase can be measured only in the high-symmetry phase. One uses, therefore, different procedures in order to estimate values for the high-symmetry parameters under the external conditions to which the measured values of the low-symmetry phase refer (see e.g. Salje, 1990[link]; Carpenter et al., 1998[link]). Three main strategies are illustrated using the example of leucite (see Fig. 3.4.3.4[link]):

  • (i) The lattice parameters of the high-symmetry phase are extrapolated from values measured in the high-symmetry phase (a straight line [a_0] in Fig. 3.4.3.4[link]). This is a preferred approach.

    [Figure 3.4.3.4]

    Figure 3.4.3.4 | top | pdf |

    Temperature dependence of lattice parameters in leucite. Courtesy of E. K. H Salje, University of Cambridge.

  • (ii) For certain symmetry descents, it is possible to approximate the high-symmetry parameters in the low-symmetry phase by average values of the lattice parameters in the low-symmetry phase. Thus for example in cubic [\rightarrow] tetragonal transitions one can take for the cubic lattice parameter [a_0=(2a+c)/3] (the dotted curve in Fig. 3.4.3.4[link]), for cubic [\rightarrow] orthorhombic transitions one may assume [a_0=(abc)^{1/3}], where [a, b, c] are the lattice parameters of the low-symmetry phase. Errors are introduced if there is a significant volume strain, as in leucite.

  • (iii) Thermal expansion is neglected and for the high-symmetry parameters in the low-symmetry phase one takes the lattice parameters measured in the high-symmetry phase as close as possible to the transition. This simplest method gives better results than average values in leucite, but in general may lead to significant errors.

Spontaneous strain has been examined in detail in many ferroic crystals by Carpenter et al. (1998[link]).

Spontaneous strain can be divided into two parts: one that is different in all ferroelastic domain states and the other that is the same in all ferroelastic domain states. This division can be achieved by introducing a modified strain tensor (Aizu, 1970[link]b), also called a relative spontaneous strain (Wadhawan, 2000[link]): [{\bf{u}}_{(s)}^{(i)}={\bf{u}}^{(i)}-{\bf{u}}_{(s)}^{({\rm av})}, \eqno(3.4.3.44) ]where [{\bf{u}}_{(s)}^{(i)}] is the matrix of relative (modified) spontaneous strain in the ferroelastic domain state [{\bf R}_i], [{\bf{u}}^{(i)}] is the matrix of an `absolute' spontaneous strain in the same ferroelastic domain state [{\bf R}_i] and [{\bf{u}}_{(s)}^{({\rm av})} ] is the matrix of an average spontaneous strain that is equal to the sum of the matrices of absolute spontaneous strains over all [n_a ] ferroelastic domain states, [{\bf{u}} ^{({\rm av})}={{1}\over{n_a}}\sum_{j=1}^{n_a}{\bf{u}}^{(j)}. \eqno(3.4.3.45) ]

The relative spontaneous strain [{\bf b}_{(s)}^{(i)}] is a symmetry-breaking strain that transforms according to a non-identity representation of the parent group G, whereas the average spontaneous strain is a non-symmetry breaking strain that transforms as the identity representation of G.

Example 3.4.3.6.  We illustrate these concepts with the example of symmetry descent [4_z/m_zm_xm_{xy}\supset 2_xm_ym_z] with two ferroelastic domain states [{\bf R}_1] and [{\bf R}_2] (see Fig. 3.4.2.2[link]). The absolute spontaneous strain in the first ferroelastic domain state [{\bf R}_1] is [{\bf u}^{(1)}= \left(\matrix{{{a-a_0}\over{a_0}} &0 &0 \cr 0 &{{b-a_0}\over{a_0}} &0 \cr 0 &0 &{{c-c_0}\over{c_0}} }\right) = \left(\matrix{u_{11} &0 &0 \cr 0 &u_{22} &0 \cr 0 &0 &u_{33}}\right),\eqno(3.4.3.46) ]where [a,b,c] and [a_0,b_0,c_0] are the lattice parameters of the orthorhombic and tetragonal phases, respectively.

The spontaneous strain [{\bf u}^{(2)}] in domain state [{\bf R}_2 ] is obtained by applying to [{\bf{u}}^{(1)}] any switching operation that transforms [{\bf R}_1] into [{\bf R}_2] (see Table 3.4.2.1[link]), [{\bf u}^{(2)}= \left(\matrix{u_{22} &0 &0 \cr 0 &u_{11} &0 \cr 0 & 0 &u_{33}}\right).\eqno(3.4.3.47) ]

The average spontaneous strain is, according to equation (3.4.3.45[link]), [{\bf u}^{(\rm av)}=\textstyle{{1}\over{2}} \left(\matrix{u_{11}+u_{22} &0 &0 \cr 0 &u_{11}+u_{22} &0 \cr 0 &0 &u_{33}+u_{33} }\right). \eqno(3.4.3.48) ]This deformation is invariant under any operation of G.

The relative spontaneous strains in ferroelastic domain states [{\bf R}_1 ] and [{\bf R}_2] are, according to equation (3.4.3.44[link]), [\eqalignno{{\bf u}_{(s)}^{(1)}&={\bf u}^{(1)}-{\bf u}^{(\rm av)}= \pmatrix{{{1}\over{2}}(u_{11}-u_{22}) &0 &0 \cr 0 &-{{1}\over{2}}(u_{11}-u_{22}) &0 \cr 0 &0 &0},&\cr &&(3.4.3.49)\cr{\bf u}_{(s)}^{(2)}&={\bf u}^{(2)}-{\bf u}^{(\rm av)}= \pmatrix{-{{1}\over{2}}(u_{11}-u_{22}) &0 &0 \cr 0 &{{1}\over{2}}(u_{11}-u_{22}) &0 \cr 0 &0 &0}. &\cr &&(3.4.3.50)}%fd3.4.3.50 ]

Symmetry-breaking nonzero components of the relative spontaneous strain are identical, up to the factor [{{1}\over{2}}], with the secondary tensor parameters [\lambda_{b}^{(1)}] and [\lambda_{b}^{(2)}] of the transition [4_z/m_zm_xm_{xy}\supset2_xm_ym_z] with the stabilizer [I_{4_z/m_zm_xm_{xy}}({\bf R}_1)=] [I_{4_z/m_zm_xm_{xy}}({\bf R}_2) =] [m_xm_ym_z]. The non-symmetry-breaking component [u_{33}] does not appear in the relative spontaneous strain.

The form of relative spontaneous strains for all ferroelastic domain states of all full ferroelastic phases are listed in Aizu (1970[link]b).

3.4.3.6.2. Equally deformed planes of a ferroelastic domain pair

| top | pdf |

We start with the example of a phase transition with the symmetry descent [G=4_z/m_zm_xm_{xy}\supset 2_xm_ym_z], which generates two ferroelastic single-domain states [{\bf R}_1] and [{\bf R}_2] (see Fig. 3.4.2.2[link]). An `elementary cell' of the parent phase is represented in Fig. 3.4.3.5[link](a) by a square [B_0E_0C_0F_0 ] and the corresponding domain state is denoted by [{\bf R}_0].

[Figure 3.4.3.5]

Figure 3.4.3.5 | top | pdf |

Formation of a ferroelastic domain twin. (a) Formation of ferroelastic single-domain states [{\bf R}_1, {\bf R}_2] from the parent phase [{\bf R}_0]; p and [p^\prime] are two perpendicular planes of equal deformation. (b) Formation of a ferroelastic twin: (i) by rotating the single-domain states [{\bf S}_1, {\bf S}_2] in (a) through an angle [\pm{{1}\over{2}}\varphi] about the domain-pair axis A ([{\bf R}_1^+] and [{\bf R}_2^-] are the resulting disoriented ferroelastic domain states); (ii) by a simple shear deformation with a shear angle (obliquity) [\varphi].

In the ferroic phase, the square [B_0E_0C_0F_0] can change either under spontaneous strain [{\bf u}^{(1)}] into a spontaneously deformed rectangular cell [B_1E_1C_1F_1] representing a domain state [{\bf R}_1 ], or under a spontaneous strain [{\bf u}^{(2)}] into rectangular [B_2E_2C_2F_2] representing domain state [{\bf R}_2]. We shall use the letter [{\bf R}_0] as a symbol of the parent phase and [{\bf R}_1, ] [{\bf R}_2] as symbols of two ferroelastic single-domain states.

Let us now choose in the parent phase a vector [{\buildrel{\longrightarrow}\over{AB_0}} ]. This vector changes into [{\buildrel{\longrightarrow}\over{AB_1}} ] in ferroelastic domain state [{\bf R}_1] and into [{\buildrel{\longrightarrow}\over{AB_2}} ] in ferroelastic domain state [{\bf R}_1]. We see that the resulting vectors [{\buildrel{\longrightarrow}\over{AB_1}}] and [{\buildrel{\longrightarrow}\over{AB_2}} ] have different direction but equal length: [|{\buildrel{\longrightarrow}\over{AB_1}}|= |{\buildrel{\longrightarrow}\over{AB_2}}| ]. This consideration holds for any vector in the plane p, which can therefore be called an equally deformed plane (EDP). One can find that the perpendicular plane [p'] is also an equally deformed plane, but there is no other plane with this property.

The intersection of the two perpendicular equally deformed planes p and [p'] is a line called an axis of the ferroelastic domain pair [({\bf R}_1,{\bf R}_2)] (in Fig. 3.4.3.5[link] it is a line at A perpendicular to the paper). This axis is the only line in which any vector chosen in the parent phase exhibits equal deformation and has its direction unchanged in both single-domain states [{\bf R}_1 ] and [{\bf R}_2] of a ferro­elastic domain pair.

This consideration can be expressed analytically as follows (Fousek & Janovec, 1969[link]; Sapriel, 1975[link]). We choose in the parent phase a plane p and a unit vector [{\bf v}(x_1,x_2,x_3) ] in this plane. The changes of lengths of this vector in the two ferroelastic domain states [{\bf R}_1] and [{\bf R}_2] are [u^{(1)}_{ik}x_ix_k ] and [u^{(2)}_{ik}x_ix_k], respectively, where [u^{(1)}_{ik} ] and [u^{(2)}_{ik}] are spontaneous strains in [{\bf R}_1] and [{\bf R}_2], respectively (see e.g. Nye, 1985[link]). (We are using the Einstein summation convention: when a letter suffix occurs twice in the same term, summation with respect to that suffix is to be understood.) If these changes are equal, i.e. if [u^{(1)}_{ik}x_ix_k=u^{(2)}_{ik}x_ix_k,\eqno(3.4.3.51)]for any vector [{\bf v}(x_1,x_2,x_3)] in the plane p this plane will be an equally deformed plane. If we introduce a differential spontaneous strain[\Delta u_{ik}\equiv u^{(2)}_{ik}-u^{(1)}_{ik}, \quad i, k=1,2,3, \eqno(3.4.3.52) ]the condition (3.4.3.51[link]) can be rewritten as [\Delta u_{ik}x_ix_j=0. \eqno(3.4.3.53)]This equation describes a cone with the apex at the origin. The cone degenerates into two planes if the determinant of the differential spontaneous strain tensor equals zero, [{\rm det}\Delta u_{ik}=0. \eqno(3.4.3.54) ]If this condition is satisfied, two solutions of (3.4.3.53[link]) exist: [Ax_1+Bx_2+Cx_3=0, \quad A{^\prime}{x_1}+B{^\prime}{x_2}+C{^\prime}{x_3}=0. \eqno(3.4.3.55) ]These are equations of two planes p and [p^\prime] passing through the origin. Their normal vectors are [{\bf n}=[ABC]] and [{\bf n'}=[A'B'C']]. It can be shown that from the equation [\Delta u_{11}+\Delta u_{22}+\Delta u_{33}=0, \eqno(3.4.3.56)]which holds for the trace of the matrix [{\rm det}\Delta u_{ik}], it follows that these two planes are perpendicular: [AA^\prime+BB^\prime+CC^\prime=0. \eqno(3.4.3.57) ]

The intersection of these equally deformed planes (3.4.3.53[link]) is the axis of the ferroelastic domain pair [({\bf R}_1,{\bf R}_2) ].

Let us illustrate the application of these results to the domain pair [({\bf R}_1,{\bf R}_2)] depicted in Fig. 3.4.3.1[link](b) and discussed above. From equations (3.4.3.41[link]) and (3.4.3.47[link]), or (3.4.3.49[link]) and (3.4.3.50[link]) we find the only nonzero components of the difference strain tensor are[\Delta u_{11}=u_{22}-u_{11}, \quad \Delta u_{22}=u_{11}-u_{22}. \eqno(3.4.3.58) ]Condition (3.4.3.54[link]) is fulfilled and equation (3.4.3.53[link]) is [\Delta u_{11}x_1^2+\Delta u_{22}x_2^2=(u_{22}-u_{11})x_1^2+(u_{11}-u_{22})x_2^2=0. \eqno(3.4.3.59) ]There are two solutions of this equation: [x_1=x_2, \quad x_1=-x_2. \eqno(3.4.3.60)]These two equally deformed planes p and [p^\prime] have the normal vectors [{\bf n}=[\bar 110] ] and [{\bf n}=[110]]. The axis of this domain pair is directed along [001].

Equally deformed planes in our example have the same orientations as have the mirror planes [m_{\bar{x}y}] and [m_{xy}] lost at the transition [4_z/m_zm_xm_{xy} \supset m_xm_ym_z]. From Fig. 3.4.3.5[link](a) it is clear why: reflection [m_{\bar{x}y}], which is a transposing operation of the domain pair ([{\bf R}_1,{\bf R}_2]), ensures that the vectors [{\buildrel{\longrightarrow}\over{AB_1}}] and [{\buildrel{\longrightarrow}\over {AB_2}} ] arising from [{\buildrel{\longrightarrow}\over{AB_0}}] have equal length. A similar conclusion holds for a 180° rotation and a plane perpendicular to the corresponding twofold axis. Thus we come to two useful rules:

Any reflection through a plane that is a transposing operation of a ferroelastic domain pair ensures the existence of two planes of equal deformation: one is parallel to the corresponding mirror plane and the other one is perpendicular to this mirror plane.

Any 180° rotation that is a transposing operation of a ferroelastic domain pair ensures the existence of two equally deformed planes: one is perpendicular to the corresponding twofold axis and the other one is parallel to this axis .

A reflection in a plane or a 180° rotation generates at least one equally deformed plane with a fixed prominent crystallographic orientation independent of the magnitude of the spontaneous strain; the other perpendicular equally deformed plane may have a non-crystallographic orientation which depends on the spontaneous strain and changes with temperature. If between switching operations there are two reflections with corresponding perpendicular mirror planes, or two 180° rotations with corresponding perpendicular twofold axes, or a reflection and a 180° rotation with a corresponding twofold axis parallel to the mirror, then both perpendicular equally deformed planes have fixed crystallographic orientations. If there are no switching operations of the second order, then both perpendicular equally deformed planes may have non-crystallographic orientations, or equally deformed planes may not exist at all.

Equally deformed planes in ferroelastic–ferroelectric phases have been tabulated by Fousek (1971[link]). Sapriel (1975[link]) lists equations (3.4.3.55[link]) of equally deformed planes for all ferro­elastic phases. Table 3.4.3.6[link] contains the orientation of equally deformed planes (with further information about the walls) for representative domain pairs of all orbits of ferroelastic domain pairs. Table 3.4.3.7[link] lists representative domain pairs of all ferroelastic orbits for which no compatible walls exist.

3.4.3.6.3. Disoriented domain states, ferroelastic domain twins and their twin laws

| top | pdf |

To examine another possible way of forming a ferroelastic domain twin, we return once again to Fig. 3.4.3.5[link](a) and split the space along the plane p into a half-space [{\cal B}_1 ] on the negative side of the plane p (defined by a negative end of normal [{\bf n}]) and another half-space [{\cal B}_2] on the positive side of p. In the parent phase, the whole space is filled with domain state [{\bf R}_0] and we can, therefore, treat the crystal in region [{\cal B}_1] as a domain [{\bf D}_1({\bf R}_0,{\cal B}_{1}) ] and the crystal in region [{\cal B}_2] as a domain [{\bf D}_2({\bf R}_0,{\cal B}_2) ] (we remember that a domain is specified by its domain region, e.g. [{\cal B}_1], and by a domain state, e.g. [{\bf R}_1 ], in this region; see Section 3.4.2.1[link]).

Now we cool the crystal down and exert the spontaneous strain [{\bf u}^{(1)} ] on domain [{\bf D}_1({\bf R}_0,{\cal B}_1)]. The resulting domain [{\bf D}_1({\bf R}_1,{\cal B}_1^{-})] contains domain state [{\bf R}_1 ] in the domain region [{\cal B}_1^{-}] with the planar boundary along [(\overline{B_1C_1})] (the overbar `−' signifies a rotation of the boundary in the positive sense). Similarly, domain [{\bf D}_2({\bf R}_0,{\cal B}_2) ] changes after performing spontaneous strain [{\bf u}^{(2)}] into domain [{\bf D}_2({\bf R}_2,{\cal B}_2^{+})] with domain state [{\bf R}_2] and the planar boundary along [(\overline{B_2C_2})]. This results in a disruption in the sector [B_1AB_2] and in an overlap of [{\bf R}_1] and [{\bf R}_2] in the sector [C_1AC_2].

The overlap can be removed and the continuity recovered by rotating the domain [{\bf D}_1({\bf R}_1,{\cal B}_1^{-})] through angle [\varphi /2 ] and the domain [{\bf D}_2({\bf R}_2,{\cal B}_2^{+})] through [-\varphi /2] about the domain-pair axis A (see Fig. 3.4.3.5[link]a and b). This rotation changes the domain [{\bf D}_1({\bf R}_1,{\cal B}_1^{-})] into domain [{\bf D}_1({\bf R}_1^{+},{\cal B}_1) ] and domain [{\bf D}_2({\bf R}_2,{\cal B}_2^{-})] into domain [{\bf D}_1({\bf R}_2^{-},{\cal B}_2)], where [{\bf R}_1^{+}] and [{\bf R}_2^{-}] are domain states rotated away from the single-domain state orientation through [\varphi /2] and [-\varphi /2], respectively. Domains [{\bf D}_1({\bf R}_1,{\cal B}_1)] and [{\bf D}_1({\bf R}_2,{\cal B}_2) ] meet without additional strains or stresses along the plane p and form a simple ferroelastic twin with a compatible domain wall along p. This wall is stress-free and fulfils the conditions of mechanical compatibility.

Domain states [{\bf R}_1^{+}] and [{\bf R}_2^{-}] with new orientations are called disoriented (misoriented) domain states or suborientational states (Shuvalov et al., 1985[link]; Dudnik & Shuvalov, 1989[link]) and the angles [\varphi /2 ] and [-\varphi /2] are the disorientation angles of [{\bf R}_1^{+}] and [{\bf R}_2^{-}], respectively.

We have described the formation of a ferroelastic domain twin by rotating single-domain states into new orientations in which a stress-free compatible contact of two ferroelastic domains is achieved. The advantage of this theoretical construct is that it provides a visual interpretation of disorientations and that it works with ferroelastic single-domain states which can be easily derived and transformed.

There is an alternative approach in which a domain state in one domain is produced from the domain state in the other domain by a shear deformation. The same procedure is used in mechanical twinning [for mechanical twinning, see Section 3.3.8.4[link] and e.g. Cahn (1954[link]); Klassen-Neklyudova (1964[link]); Christian (1975[link])].

We illustrate this approach again using our example. From Fig. 3.4.3.5[link](b) it follows that domain state [{\bf R}_2^{-} ] in the second domain can be obtained by performing a simple shear on the domain state [{\bf R}_1^{+}] of the first domain. In this simple shear, a point is displaced in a direction parallel to the equally deformed plane p (in mechanical twinning called a twin plane) and to a plane perpendicular to the axis of the domain pair (plane of shear). The displacement [{\bf q}] is proportional to the distance d of the point from the domain wall. The amount of shear is measured either by the absolute value of this displacement at a unit distance, [s=q/d], or by an angle [\varphi] called a shear angle (sometimes [2\varphi] is defined as the shear angle). There is no change of volume connected with a simple shear.

The angle [\varphi] is also called an obliquity of a twin (Cahn, 1954[link]) and is used as a convenient measure of pseudosymmetry of the ferroelastic phase.

The high-resolution electron microscopy image in Fig. 3.4.3.6[link] reveals the relatively large shear angle (obliquity) [\varphi] of a ferroelastic twin in the monoclinic phase of tungsten trioxide (WO3). The plane (101) corresponds to the plane p of a ferro­elastic wall in Fig. 3.4.3.5[link](b). The planes [(\bar1 01) ] are crystallographic planes in the lower and upper ferroelastic domains, which correspond in Fig. 3.4.3.5[link](b) to domain [{\bf D}_1({\bf R}_1^{+},{\cal B}_1)] and domain [{\bf D}_2({\bf R}_2^{-},{\cal B}_2) ], respectively. The planes [(\bar1 01)] in these domains correspond to the diagonals of the elementary cells of [{\bf R}_1^{+}] and [{\bf R}_2^{-}] in Fig. 3.4.3.5[link](b) and are nearly perpendicular to the wall. The angle between these planes equals [2\varphi], where [\varphi] is the shear angle (obliquity) of the ferroelastic twin.

[Figure 3.4.3.6]

Figure 3.4.3.6 | top | pdf |

High-resolution electron microscopy image of a ferroelastic twin in the orthorhombic phase of WO3. Courtesy of H. Lemmens, EMAT, University of Antwerp.

Disorientations of domain states in a ferroelastic twin bring about a deviation of the optical indicatrix from a strictly perpendicular position. Owing to this effect, ferroelastic domains exhibit different colours in polarized light and can be easily visualized. This is illustrated for a domain structure of YBa2Cu3O7−δ in Fig. 3.4.3.7[link]. The symmetry descent G = [4_z/m_zm_xm_{xy}\supset m_xm_ym_z =] [F_1=] [F_2] gives rise to two ferroelastic domain states [{\bf R}_1] and [{\bf R}_2 ]. The twinning group [K_{12}] of the non-trivial domain pair [({\bf R}_1,{\bf R}_2)] is [K_{12}[m_xm_ym_z]= J_{12}^{\star}=m_xm_ym_z \cup 4_z^{\star}\{2_xm_ym_z\} = 4_z^{\star}/m_zm_xm_{xy}^{\star}. \eqno(3.4.3.61) ]The colour of a domain state observed in a polarized-light microscope depends on the orientation of the index ellipsoid (indicatrix) with respect to a fixed polarizer and analyser. This index ellipsoid transforms in the same way as the tensor of spontaneous strain, i.e. it has different orientations in ferroelastic domain states. Therefore, different ferroelastic domain states exhibit different colours: in Fig. 3.4.3.7[link], the blue and pink areas (with different orientations of the ellipse representing the spontaneous strain in the plane of of figure) correspond to two different ferroelastic domain states. A rotation of the crystal that does not change the orientation of ellipses (e.g. a 180° rotation about an axis parallel to the fourfold rotation axis) does not change the colours (ferroelastic domain states). If one neglects disorientations of ferroelastic domain states (see Section 3.4.3.6[link]) – which are too small to be detected by polarized-light microscopy – then none of the operations of the group [F_1=] [F_2=m_xm_ym_z] change the single-domain ferroelastic domain states [{\bf R}_1], [{\bf R}_2], hence there is no change in the colours of domain regions of the crystal. On the other hand, all operations with a star symbol (operations lost at the transition) exchange domain states [{\bf R}_1] and [{\bf R}_2 ], i.e. also exchange the two colours in the domain regions. The corresponding permutation is a transposition of two colours and this attribute is represented by a star attached to the symbol of the operation. This exchange of colours is nicely demonstrated in Fig. 3.4.3.7[link] where a −90° rotation is accompanied by an exchange of the pink and blue colours in the domain regions (Schmid, 1991[link], 1993[link]).

[Figure 3.4.3.7]

Figure 3.4.3.7 | top | pdf |

Ferroelastic twins in a very thin YBa2Cu3O7−δ crystal observed in a polarized-light microscope. Courtesy of H. Schmid, Université de Geneve.

It can be shown (Shuvalov et al., 1985[link]; Dudnik & Shuvalov, 1989[link]) that for small spontaneous strains the amount of shear s and the angle [\varphi] can be calculated from the second invariant [\Lambda_2] of the differential tensor [\Delta u_{ik}]: [\eqalignno{s&=2\sqrt{-\Lambda_2}, &(3.4.3.62)\cr \varphi&=\sqrt{-\Lambda_2 }, &(3.4.3.63)}%fd3.4.3.63 ]where [\Lambda_2 = \left|\matrix{\bigtriangleup u_{11} &\bigtriangleup u_{12} \cr \bigtriangleup u_{21} &\bigtriangleup u_{22}}\right| + \left|\matrix{\bigtriangleup u_{22} &\bigtriangleup u_{23} \cr \bigtriangleup u_{32} &\bigtriangleup u_{33}}\right| + \left|\matrix{\bigtriangleup u_{11} &\bigtriangleup u_{13} \cr \bigtriangleup u_{31} &\bigtriangleup u_{33}}\right|.\eqno(3.4.3.64) ]

In our example, where there are only two nonzero components of the differential spontaneous strain tensor [see equation (3.4.3.58[link])], the second invariant [\Lambda_2=] [-(\Delta u_{11}\Delta u_{22}) =] [-(u_{22}-u_{11})^2] and the angle [\varphi] is [\varphi=\pm|u_{22}-u_{11}|.\eqno(3.4.3.65) ]In this case, the angle [\varphi] can also be expressed as [\varphi=\pi /2-2\,{\rm arctan}\,a/b], where a and b are lattice parameters of the orthorhombic phase (Schmid et al., 1988[link]).

The shear angle [\varphi] ranges in ferroelastic crystals from minutes to degrees (see e.g. Schmid et al., 1988[link]; Dudnik & Shuvalov, 1989[link]).

Each equally deformed plane gives rise to two compatible domain walls of the same orientation but with opposite sequence of domain states on each side of the plane. We shall use for a simple domain twin with a planar wall a symbol [({\bf R}_1^{+}|\bf{n}|{\bf R}_2^{-})] in which n denotes the normal to the wall. The bra–ket symbol [(\,\,|] and [|\,\,)] represents the half-space domain regions on the negative and positive sides of [{\bf n}], respectively, for which we have used letters [{\cal B}_1] and [{\cal B}_2], respectively. Then [({\bf R}_1^{+}|] and [|{\bf R}_2^{-})] represent domains [{{\bf D}_1}({\bf R}_1^{+},{\cal B}_1) ] and [{{\bf D}_2}({\bf R}_2^{-},{\cal B}_2)], respectively. The symbol [({\bf R}_1^{+}|{\bf R}_2^{-})] properly specifies a domain twin with a zero-thickness domain wall.

A domain wall can be considered as a domain twin with domain regions restricted to non-homogeneous parts near the plane p. For a domain wall in domain twin [({\bf R}_1^{+}|{\bf R}_2^{-})] we shall use the symbol [[{\bf R}_1^{+}|{\bf R}_2^{-}] ], which expresses the fact that a domain wall of zero thickness needs the same specification as the domain twin.

If we exchange domain states in the twin [({\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-}) ], we get a reversed twin (wall) with the symbol [({\bf R}_2^{-}|{\bf n}|{\bf R}_1^{+}) ]. These two ferroelastic twins are depicted in the lower right and upper left parts of Fig. 3.4.3.8[link], where – for ferroelastic–non-ferroelectric twins – we neglect spontaneous polarization of ferroelastic domain states. The reversed twin [{\bf R}_2^{-}|{\bf n}'|{\bf R}_1^{+}] has the opposite shear direction.

[Figure 3.4.3.8]

Figure 3.4.3.8 | top | pdf |

Exploded view of four ferroelastic twins with disoriented ferroelastic domain states [{\bf R}_1^+, {\bf R}_2^-] and [{\bf R}_1^-, {\bf R}_2^+ ] formed from a single-domain pair [({\bf S}_1,{\bf S}_2)] (in the centre).

Twin and reversed twin can be, but may not be, crystallographically equivalent. Thus e.g. ferroelastic–non-ferroelectric twins [({\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-}) ] and [({\bf R}_2^{-}|{\bf n}|{\bf R}_1^{+})] in Fig. 3.4.3.8[link] are equivalent, e.g. via [2_z], whereas ferroelastic–ferroelectric twins [({\bf S}_1^{+}|{\bf n}|{\bf S}_3^{-}) ] and [({\bf S}_3^{-}|{\bf n}|{\bf S}_1^{+})] are not equivalent, since there is no operation in the group [K_{12}] that would transform [({\bf S}_1^{+}|{\bf n}|{\bf S}_3^{-})] into [({\bf S}_3^{-}|{\bf n}|{\bf S}_1^{+}) ].

As we shall show in the next section, the symmetry group [{\sf T}_{12}({\bf n}) ] of a twin and the symmetry group [{\sf T}_{21}({\bf n})] of a reverse twin are equal, [{\sf T}_{12}({\bf n}) = {\sf T}_{21}({\bf n}).\eqno(3.4.3.66) ]

A sequence of repeating twins and reversed twins [\ldots {\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-}|{\bf n}|{\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-} |{\bf n}|{\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-}|{\bf n}|{\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-}\ldots\eqno(3.4.3.67) ]forms a lamellar ferroelastic domain structure that is very common in ferroelastic phases (see e.g. Figs. 3.4.1.1[link] and 3.4.1.4[link]).

Similar considerations can be applied to the second equally deformed plane [p^\prime] that is perpendicular to p. The two twins and corresponding compatible domain walls for the equally deformed plane [p^\prime] have the symbols [({\bf R}_1^{-}|{\bf n}^\prime|{\bf R}_2^{+})] and [({\bf R}_2^{-}|{\bf n}^\prime|{\bf R}_1^{+})], and are also depicted in Fig. 3.4.3.8[link]. The corresponding lamellar domain structure is [\ldots{\bf R}_1^{-}|{\bf n}^\prime|{\bf R}_2^{+}|{\bf n}^\prime|{\bf R}_1^{-}|{\bf n}^\prime|{\bf R}_2^{+} |{\bf n}^\prime|{\bf R}_1^{-}|{\bf n}^\prime|{\bf R}_2^{+}|{\bf n}^\prime|{\bf R}_1^{-}|{\bf n}^\prime|{\bf R}_2^{+}\ldots\,.\eqno(3.4.3.68) ]

Thus from one ferroelastic single-domain pair [({\bf R}_1, {\bf R}_2) ] depicted in the centre of Fig. 3.4.3.8[link] four different ferroelastic domain twins can be formed. It can be shown that these four twins have the same shear angle [\varphi] and the same amount of shear s. They differ only in the direction of the shear.

Four disoriented domain states [{\bf R}_1^{-}, {\bf R}_1^{+}] and [{\bf R}_2^{-}, {\bf R}_2^{+}] that appear in the four domain twins considered above are related by lost operations (e.g. diagonal, vertical and horizontal reflections), i.e. they are crystallographically equivalent. This result can readily be obtained if we consider the stabilizer of a disoriented domain state [{\bf R}_1^{+}], which is [I_{4/mmm}({\bf R}_1^{+})=2_z/m_z ]. Then the number [n_a^{\rm dis}] of disoriented ferroelastic domain states is given by [n^{\rm dis}_a=[G:I_g({\bf R}_1^{+})]=|4_z/m_zm_xm_{xy}|:|2_z/m_z]=16:4=4.\eqno(3.4.3.69) ]All these domain states appear in ferroelastic polydomain structures that contain coexisting lamellar structures (3.4.3.67[link]) and (3.4.3.68[link]).

Disoriented domain states in ferroelastic domain structures can be recognized by diffraction techniques (e.g. using an X-ray precession camera). The presence of these four disoriented domain states results in splitting of the diffraction spots of the high-symmetry tetragonal phase into four or two spots in the orthorhombic ferroelastic phase. This splitting is schematically depicted in Fig. 3.4.3.9[link]. For more details see e.g. Shmyt'ko et al. (1987[link]), Rosová et al. (1993[link]), and Rosová (1999[link]).

[Figure 3.4.3.9]

Figure 3.4.3.9 | top | pdf |

Splitting of diffraction spots from the four domain twins in Fig. 3.4.3.8[link]. (a) Diffraction spots of the tetragonal parent phase of the domain state [{\bf R}_1]. (b) Diffraction pattern of the domain structure with four domain twins: white circles, [{\bf R}_1^+]; black circles, [{\bf R}_1^-]; white squares, [{\bf R}_2^+]; black squares, [{\bf R}_2^-].

Finally, we turn to twin laws of ferroelastic domain twins with compatible domain walls. In a ferroelastic twin, say [({\bf R}_1^{+}|{\bf n}|{\bf R}_2^{-}) ], there are just two possible twinning operations that interchange two ferroelastic domain states [{\bf R}_1^{+}] and [{\bf R}_2^{-} ] of the twin: reflection through the plane of the domain wall ([m^{\star}_{\bar{x}y} ] in our example) and 180° rotation with a rotation axis in the intersection of the domain wall and the plane of shear ([2^{\star}_{x y} ]). These are the only transposing operations of the domain pair [({\bf R}_1,{\bf R}_2)] that are preserved by the shear; all other transposing operations of the domain pair [({\bf R}_1,{\bf R}_2)] are lost. (This is a difference from non-ferroelastic twins, where all transposing operations of the pair become twinning operations of a non-ferroelastic twin.)

Consider the twin [({\bf S}_1^{+}|{\bf n}|{\bf S}_3^{-})] in Fig. 3.4.3.8[link]. By non-trivial twinning operations we understand transposing operations of the domain pair [({\bf S}_1^{+},{\bf S}_3^{-}) ], whereas trivial twinning operations leave invariant [{\bf S}_1^{+} ] and [{\bf S}_3^{-}]. As we shall see in the next section, the union of trivial and non-trivial twinning operations forms a group [{\sf T}_{1^{+}2^{-}}({\bf n}) ]. This group, called the symmetry group of the twin [({\bf S}_1^{+}|{\bf n}|{\bf S}_3^{-}) ], comprises all symmetry operations of this twin and we shall use it for designating the twin law of the ferroelastic twin, just as the group [J_{1j}^{\star}] of the domain pair [({\bf S}_1,{{\bf S}_j)} ] specifies the twin law of a non-ferroelastic twin. This group [{\sf T}_{1^{+}2^{-}}({\bf n}) ] is a layer group (see Section 3.4.4.2[link]) that keeps the plane p invariant, but for characterizing the twin law, which specifies the relation of domain states of two domains in the twin, one can treat [{\sf T}_{1^{+}2^{-}}({\bf n})] as an ordinary (dichromatic) point group [T_{1^{+}2^{-}}({\bf n})]. Thus the twin law of the domain twin [({\bf S}_1^{+}|{\bf n}|{\bf S}_3^{-})] is designated by the group [T_{1^{+}3^{-}}({\bf n})=2_{xy}^{\star}m_{x\bar y}^{\star}m_z=T_{3^{-}1^{+}}{({\bf n})}, \eqno(3.4.3.70) ]where (3.4.3.70[link]) expresses the fact that a twin and the reversed twin have the same symmetry, see equation (3.4.3.66[link]). We see that this group coincides with the symmetry group [J_{1^{+}2^{-}} ] of the single-domain pair [({\bf S}_1,{\bf S}_3)] (see Fig. 3.4.3.1[link]b).

The twin law of two twins [({\bf S}_1^{-}|{\bf n^\prime}|{\bf S}_3^{+}) ] and [({\bf S}_3^{+}|{\bf n^\prime}|{\bf S}_1^{-})] with the same equally deformed plane [p^\prime] is expressed by the group [T_{1^{-}3^{+}}({\bf n^\prime})=m_z= T_{3^{-}1^{+}}({\bf n}^\prime), \eqno(3.4.3.71) ]which is different from the [T_{1^{+}3^{-}}({\bf n})] of the twin [({\bf S}_1^{+}|{\bf n}|{\bf S}_3^{-})].

Representative domain pairs of all orbits of ferroelastic domain pairs (Litvin & Janovec, 1999[link]) are listed in two tables. Table 3.4.3.6[link] contains representative domain pairs for which compatible domain walls exist and Table 3.4.3.7[link] lists ferroelastic domain pairs where compatible coexistence of domain states is not possible. Table 3.4.3.6[link] contains, beside other data, for each ferroelastic domain pair the orientation of two equally deformed planes and the corresponding symmetries of the corresponding four twins which express two twin laws.

3.4.3.6.4. Ferroelastic domain pairs with compatible domain walls, synoptic table

| top | pdf |

As we have seen, for each ferroelastic domain pair for which condition (3.4.3.54[link]) for the existence of coherent domain walls is fulfilled, there exist two perpendicular equally deformed planes. On each of these planes two ferroelastic twins can be formed; these two twins are in a simple relation (one is a reversed twin of the other), have the same symmetry, and can therefore be represented by one of these twins. Then we can say that from one ferroelastic domain pair two different twins can be formed. Each of these twins represents a different `twin law' that has arisen from the initial domain pair. All four ferroelastic twins can be described in terms of mechanical twinning with the same value of the shear angle [\varphi].

3.4.3.6.4.1. Explanation of Table 3.4.3.6[link]

| top | pdf |

Table 3.4.3.6[link] presents representative domain pairs of all classes of ferroelastic domain pairs for which compatible domain walls exist. The first five columns concern the domain pair. In subsequent columns, each row splits into two rows describing the orientation of two associated perpendicular equally deformed planes and the symmetry properties of the four domain twins that can be formed from the given domain pair. We explain the meaning of each column in detail.

Table 3.4.3.6 | top | pdf |
Ferroelastic domain pairs and twins with compatible walls

[F_1:] symmetry of [{\bf S}_1]; [g_{1j}]: twinning operations; [K_{1j}]: twinning group; Axis: axis of domain pair; Equation: direction of the axis; [\varphi ]: disorientation angle; [{\sf \overline J}_{1j} ]: symmetry of the twin pair; [{\underline t}_{1j}^{\star}]: twinning operation; [{\sf \overline T}_{1j}]: symmetry of the twin and wall, twin law of the twin; Classification: see Table 3.4.4.3[link].

[F_1 ] [g_{1j} ] [K_{1j}] Axis Equation Wall normals [\varphi] [{\sf \overline J}_{1j}] [\underline t^{\star}_{1j} ] [{\sf\overline T}_{1j}] Classification
1 [2^{\star}_z] [2^{\star}_Z] [[B\bar{1}0]] (a)   [[001]] (1) [2^{\star}_z]   1 [{\rm AR}^{\star}]
  [[1B0]_e] [\underline2^{\star}_z] [\underline2^{\star}_z] [\underline2^{\star}_z] SI
1 [m^{\star}_z] [m^{\star}_z] [[B\bar{1}0] ] (a)   [[001]_e ] (1) [{\underline m}^{\star}_z ] [{\underline m}^{\star}_z ] [{\underline m}^{\star}_z ] SI
  [[1B0] ] [m^{\star}_z]   1 [{\rm AR}^{\star}]
[\bar{1}] [m^{\star}_z], [2^{\star}_z] [2^{\star}_z/m^{\star}_z ] [[B\bar{1}0]] (a)   [[001]] (1) [2_z^{\star}/\underline{m}^{\star}_z] [\underline{m}^{\star}_z ] [\underline{m}^{\star}_z ] SR
  [[1B0]] [\underline2^{\star}_z/m^{\star}_z] [\underline2^{\star}_z] [\underline2^{\star}_z] SR
[2_z] [2^{\star}_x], [2^{\star}_y] [2^{\star}_x2^{\star}_y2_z ] [[001]]     [[100] ] (2) [2^{\star}_x\underline2^{\star}_y\underline2_z ] [\underline2^{\star}_y] [\underline2^{\star}_y] SR
    [[010]] [\underline2^{\star}_x2^{\star}_y\underline2_z ] [\underline2^{\star}_x] [\underline2^{\star}_x] SR
[2_z] [m^{\star}_x], [m^{\star}_y] [m^{\star}_xm^{\star}_y2_z ] [[001]]     [[100] ] (2) [\underline{m}^{\star}_xm^{\star}_y\underline2_z ] [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x ] SR
  [[010]] [m^{\star}_x\underline{m}^{\star}_y\underline2_z ] [\underline{m}^{\star}_y ] [\underline{m}^{\star}_y ] SR
[2_z] [4^{\star}_z], [4_z^{3 \star}] [4^{\star}_z] [[001]] (b) [\Bigl[] [[1B0]] (3) [\underline2_z]   1 [{\rm A}\underline {\rm R}]
[[B\bar{1}0] ] [\underline2_z ] 1 [{\rm A}\underline{\rm R}]
[2_z] [\bar4^{\star}_z ], [\bar4_z^{*3}] [\bar{4}^{\star}_z ] [[001]] (b) [\Bigl[] [[1B0]] (3) [\underline2_z ]   1 [{\rm A}\underline{\rm R} ]
[[B\bar{1}0] ] [\underline2_z]   1 [{\rm A}\underline{\rm R} ]
[2_z] [3_z], [6_z^5] [6_{z}] [[001]] (c)   [[1B0]] (4) [\underline2_z]   1 [{\rm A}\underline{\rm R}]
  [[B\bar{1}0]] [\underline2_z]   1 [{\rm A}\underline{\rm R}]
[3_z^2], [6_z] [6_z] [[001]] (c)   [[1B0] ] (4) [\underline2_z]   1 [{\rm A}\underline{\rm R}]
  [[B\bar{1}0] ] [\underline2_z]   1 [{\rm A}\underline{\rm R}]
[2_z ] [\bar3_z^5], [\bar6_z ] [6_{z}/m_z] [[001]] (c)   [[1B0] ] (4) [\underline2_z ]   1 [{\rm A}\underline{\rm R} ]
  [[B\bar{1}0]] [\underline2_z]   1 [{\rm A}\underline{\rm R}]
[\bar3_z], [\bar6_z^5] [6_z/m_z] [[001]] (c)   [[1B0]] (4) [\underline2_z]   1 [{\rm A}\underline{\rm R} ]
  [[B\bar{1}0] ] [\underline2_z]   1 [{\rm A}\underline{\rm R}]
[2_x] [2^{\star}_{xy} ], [4_z] [4_z2_x2_{xy} ] [[\bar{C}C2] ] (d)   [[110] ] (5) [2^{\star}_{xy}]   1 [{\rm AR}^{\star}]
  [[1\bar{1}C]_e] [\underline2^{\star}_{xy} ] [\underline2^{\star}_{xy} ] [\underline2^{\star}_{xy}] SI
[2_x] [m^{\star}_{xy} ], [\bar4_z] [\bar4_z2_xm_{xy} ] [[\bar{C}C2]] (d)   [[110]_e ] (5) [\underline{m}^{\star}_{xy} ] [\underline{m}^{\star}_{xy} ] [\underline{m}^{\star}_{xy}] SI
  [[1\bar{1}C] ] [m^{\star}_{xy}]   [1] [{\rm AR}^{\star}]
[2_x] [2^{\star}_{x{^\prime}} ], [3_z^2] [3_z2_x] [[\sqrt{3}C], C, [\bar4]] (e)   [[\bar1\sqrt{3}0]] (6) [2^{\star}_{x{^\prime}}]   1 [{\rm AR}^{\star}]
  [[\sqrt{3}1C]_e ] [\underline2^{\star}_{x{^\prime}}] [\underline2^{\star}_{x{^\prime}} ] [\underline2^{\star}_{x{^\prime}} ] SI
[2_x] [m^{\star}_{x{^\prime}} ], [\bar3_z^5] [\bar{3}_zm_x ] [[\sqrt{3}C], C, [\bar4]] (e)   [[\bar1\sqrt{3}0]_e ] (6) [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}} ] SI
  [[\sqrt{3}1C] ] [m^{\star}_{x{^\prime}} ]   1 [{\rm AR}^{\star}]
[2_x] [2^{\star}_{y{^\prime}} ], [6_z] [6_z2_x2_y] [[\bar{C}], [\sqrt{3}C], [\bar4]] (f)   [[\sqrt{3}10]] (7) [2^{\star}_{y{^\prime}}]   1 [{\rm AR}^{\star}]
  [[\bar1\sqrt{3}C]_e] [\underline2^{\star}_{y{^\prime}} ] [\underline2^{\star}_{y{^\prime}} ] [\underline2^{\star}_{y{^\prime}} ] SI
[2_x] [m^{\star}_{y{^\prime}} ], [\bar6_z] [\bar{6}_z2_xm_y ] [[\bar{C}], [\sqrt{3}C], [\bar4]] (f)   [[\sqrt{3}10]_e ] (7) [\underline{m}^{\star}_{y{^\prime}}] [\underline{m}^{\star}_{y{^\prime}} ] [\underline{m}^{\star}_{y{^\prime}} ] SI
  [[\bar1\sqrt{3}C] ] [m^{\star}_{y{^\prime}}]   1 [{\rm AR}^{\star}]
[2_{xy}] [m^{\star}_{x} ], [\bar4_z^{3}] [\bar4_zm_x2_{xy} ] [[0C\bar{1}] ] (g)   [[100]_e ] (8) [\underline{m}^{\star}_{x} ] [\underline{m}^{\star}_{x} ] [\underline{m}^{\star}_{x}] SI
  [[01C]] [m^{\star}_{x}]   1 [{\rm AR}^{\star}]
[m_z] [m^{\star}_x], [2^{\star}_y] [m^{\star}_x2^{\star}_ym_z ] [[001]]     [[100]_e] (2) [\underline{m}^{\star}_x\underline2^{\star}_ym_z ] [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x\underline2^{\star}_ym_z ] SI
    [[010]] [m_x^{\star}2_y^{\star}m_z] [m_z] [{\rm AR}^{\star}]
[m_{z}] [4_z], [\bar4_z^3 ] [4_z/m_{z}] [[001]] (b)   [[1B0]_{e0}] (3) [m_z ]   [m_z ] AI
  [[B\bar{1}0]_{0e}] [m_z] [m_z] AI
[4_z^3], [\bar4_z] [4_z/m_z] [[001]] (b)   [[1B0]_{e0}] (3) [m_z]   [m_z] AI
  [[B\bar{1}0]_{0e}] [m_z] [m_z] AI
[m_z] [3_z], [\bar6^5_z ] [\bar6_{z}] [[001]] (c)   [[1B0]_{e0} ] (4) [m_z ]   [m_z] AI
  [[B\bar{1}0]_{0e} ] [m_z]   [m_z ] AI
[3^2_z], [\bar6_z] [\bar6_z] [[001]] (c)   [[1B0]_{e0} ] (4) [m_z]   [m_z] AI
  [[B\bar{1}0]_{0e} ] [m_z]   [m_z ] AI
[m_z] [\bar3_z], [6_z^5 ] [6_{z}/m_z] [[001]] (c)   [[1B0]_{e0}] (4) [m_z]   [m_z] AI
  [[B\bar{1}0]_{0e} ] [m_z]   [m_z] AI
[\bar3_z^5], [6_z] [6_z/m_z] [[001]] (c)   [[1B0]_{e0} ] (4) [m_z]   [m_z ] AI
  [[B\bar{1}0]_{0e} ] [m_z]   [m_z ] AI
[m_x] [m^{\star}_{xy} ], [4_z] [4_zm_xm_{xy} ] [[\bar{C}C2]] (d)   [[110]_e ] (5) [\underline{m}^{\star}_{xy}] [\underline{m}^{\star}_{xy}] [\underline{m}^{\star}_{xy}] SI
  [[1\bar{1}C]] [m^{\star}_{xy}]   1 [{\rm AR}^{\star}]
[m_x] [2^{\star}_{xy} ], [\bar4_z] [\bar{4}_zm_x2_{xy} ] [[\bar{C}C2]] (d)   [[110] ] (5) [2^{\star}_{xy}]   1 [{\rm AR}^{\star}]
  [[1\bar{1}C]_e] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] SI
[m_x] [m^{\star}_{x{^\prime}} ], [3_z^2] [3_zm_x] [[\sqrt{3}C], C, [\bar4]] (e)   [[\bar1\sqrt{3}0]_e ] (6) [\underline{m}^{\star}_{x{^\prime}}] [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}} ] SI
  [[\sqrt{3}1C] ] [m^{\star}_{x{^\prime}}]   [1] [{\rm AR}^{\star}]
[m_x] [2^{\star}_{x{^\prime}} ], [\bar3_z^5] [\bar3_zm_x] [[\sqrt{3}C], C, [\bar4]] (e)   [[\bar1\sqrt{3}0] ] (6) [2^{\star}_{x{^\prime}}]   [1] [{\rm AR}^{\star}]
  [[\sqrt{3}1C]_e] [\underline2^{\star}_{x{^\prime}}] [\underline2^{\star}_{x{^\prime}}] [\underline2^{\star}_{x{^\prime}}] SI
[m_x] [m^{\star}_{y{^\prime}} ], [6_z] [6_zm_xm_y] [[\bar{C}], [\sqrt{3}C], [\bar4]] (f)   [[\sqrt{3}10]_e] (7) [\underline{m}^{\star}_{y{^\prime}} ] [\underline{m}^{\star}_{y{^\prime}} ] [\underline{m}^{\star}_{y{^\prime}} ] SI
  [[\bar1\sqrt{3}C] ] [m^{\star}_{y{^\prime}}]   [1] [{\rm AR}^{\star}]
[m_x] [2^{\star}_{y{^\prime}} ], [\bar6_z] [\bar{6}_zm_x2_y ] [[\bar{C}], [\sqrt{3}C], [\bar4]] (f)   [[\sqrt{3}10] ] (7) [2^{\star}_{y{^\prime}}]   [1] [{\rm AR}^{\star}]
  [[\bar1\sqrt{3}C]_e ] [\underline2^{\star}_{y{^\prime}} ] [\underline2^{\star}_{y{^\prime}}] [\underline2^{\star}_{y{^\prime}} ] SI
[m_{xy}] [2^{\star}_{x} ], [\bar4^3_z] [\bar{4}_z2_xm_{xy} ] [[0C\bar{1}]] (g)   [[100]] (8) [2^{\star}_{x}]   [1] [{\rm AR}^{\star}]
  [[01C]_e] [\underline2^{\star}_{xy}] [\underline2^{\star}_{x}] [\underline2^{\star}_{x}] SI
[2_z/m_z] [m^{\star}_x], [m^{\star}_y] [m^{\star}_xm^{\star}_ym_z ] [[001] ]     [[100]] (2) [\underline{m}^{\star}_xm^{\star}_ym_z ] [\underline{m}^{\star}_x] [\underline{m}^{\star}_x\underline2^{\star}_ym_z ] SR
    [[010]] [m^{\star}_x\underline{m}^{\star}_ym_z ] [\underline{m}^{\star}_y] [\underline2^{\star}_x\underline{m}^{\star}_ym_z ] SR
[2_z/m_z] [4^{\star}_z], [4^{3 \star}_z] [4^{\star}_z/m_z ] [[001]] (b) [\Bigl[] [[1B0]] (3) [\underline2_z/m_z]   [m_z] [{\rm A}\underline {\rm R}]
[[B\bar{1}0]] [\underline2_z/m_z] [m_z] [{\rm A}\underline {\rm R}]
[2_z/m_z] [3_z], [6_z^5] [6_{z}/m_z] [[001]] (c)   [[1B0]] (4) [\underline2_z/m_z]   [m_z ] [{\rm A}\underline {\rm R}]
  [[B\bar{1}0] ] [\underline2_z/m_z]   [m_z] [{\rm A}\underline {\rm R}]
[3_z^2], [6_z] [6_z/m_z ] [[001]] (c)   [[1B0] ] (4) [\underline2_z/m_z]   [m_z] [{\rm A}\underline {\rm R}]
  [[B\bar{1}0] ] [\underline2_z/m_z]   [m_z] [{\rm A}\underline {\rm R}]
[2_x/m_x] [m^{\star}_{xy} ], [4_z] [4_z/m_zm_xm_{xy} ] [[\bar{C}C2]] (d)   [[110]] (5) [2^{\star}_{xy}/\underline{m}^{\star}_{xy} ] [\underline{m}^{\star}_{xy}] [\underline{m}^{\star}_{xy}] SR
  [[1\bar{1}C]] [\underline2^{\star}_{xy}/m^{\star}_{xy} ] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] SR
[2_x/m_x] [m^{\star}_{x{^\prime}} ], [3_z^2] [\bar{3}_zm_x ] [[\sqrt{3}CC\bar4] ] (e)   [[\bar1\sqrt{3}0]] (6) [2^{\star}_{x{^\prime}}/\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}}] [\underline{m}^{\star}_{x{^\prime}}] SR
  [[\sqrt{3}1C]] [\underline2^{\star}_{x{^\prime}}/m^{\star}_{x{^\prime}} ] [\underline2^{\star}_{x{^\prime}} ] [\underline2^{\star}_{x{^\prime}} ] SR
[2_x/m_x] [m^{\star}_{y{^\prime}} ], [6_z] [6_z/m_zm_xm_y ] [[\bar{C}], [\sqrt{3}C], [\bar4]] (f)   [[\sqrt{3}10] ] (7) [2^{\star}_{y{^\prime}}/\underline{m}^{\star}_{y{^\prime}} ] [\underline{m}^{\star}_{y{^\prime}}] [\underline{m}^{\star}_{y{^\prime}}] SR
  [[\bar1\sqrt{3}C]] [\underline2^{\star}_{y{^\prime}}/m^{\star}_{y{^\prime}} ] [\underline2^{\star}_{y{^\prime}} ] [\underline2^{\star}_{y{^\prime}} ] SR
[2_x2_y2_z] [2^{\star}_{x\bar{y}} ], [2^{\star}_{xy}] [4^{\star}_z2_x2^{\star}_{xy} ] [[001]]   [\Bigl[] [[110]] (10) [2^{\star}_{xy}\underline2^{\star}_{x\bar{y}}\underline2_z ] [\underline2^{\star}_{x\bar{y}}] [\underline2^{\star}_{x\bar{y}}] SR
  [[1\bar10]] [\underline2^{\star}_{xy}2^{\star}_{x\bar{y}}\underline2_z ] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] SR
[2_x2_y2_z] [m^{\star}_{x\bar{y}} ], [m^{\star}_{xy}] [\bar{4}^{\star}_z2_xm^{\star}_{xy} ] [[001]]   [\Bigl[] [[110]] (10) [\underline{m}^{\star}_{xy}m^{\star}_{x\bar{y}}\underline2_z ] [\underline{m}^{\star}_{xy} ] [\underline{m}^{\star}_{xy}] SR
  [[1\bar10]] [m^{\star}_{xy}\underline{m}^{\star}_{x\bar{y}}\underline2_z ] [\underline{m}^{\star}_{x\bar{y}} ] [\underline{m}^{\star}_{x\bar{y}} ] SR
[2_x2_y2_z] [2^{\star}_{x{^\prime}} ], [2^{\star}_{y{^\prime}}] [6_z2_x2_y] [[001]]     [[\bar1\sqrt{3}0] ] (9) [2^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}\underline2_z ] [\underline2^{\star}_{y{^\prime}}] [\underline2^{\star}_{y{^\prime}} ] SR
    [[\sqrt{3}10]] [\underline2^{\star}_{x{^\prime}}2^{\star}_{y{^\prime}}\underline2_z ] [\underline2^{\star}_{x{^\prime}}] [\underline2^{\star}_{x{^\prime}} ] SR
[2_x2_y2_z] [m^{\star}_{x{^\prime}} ], [m^{\star}_{y{^\prime}}] [6_z/m_zm_xm_y ] [[001]]     [[\bar1\sqrt{3}0]] (9) [\underline{m}^{\star}_{x{^\prime}}m^{\star}_{y{^\prime}}\underline2_z ] [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}}] SR
    [[\sqrt{3}10]] [{m}^{\star}_{x{^\prime}}\underline{m}^{\star}_{y{^\prime}}\underline2_z ] [\underline{m}^{\star}_{y{^\prime}} ] [\underline{m}^{\star}_{y{^\prime}}] SR
[2_{x\bar{y}}2_{xy}2_z ] [m^{\star}_{x} ], [m^{\star}_{y}] [{\bar 4}_z^\star m_x^\star 2_{xy} ] [[001]]   [\Bigl[] [[100]] (12) [\underline{m}^{\star}_{x}m^{\star}_{y}\underline2_z ] [\underline{m}^{\star}_{x} ] [\underline{m}^{\star}_{x}] SR
  [[010]] [m^{\star}_{x}\underline{m}^{\star}_{y}\underline2_z ] [\underline{m}^{\star}_{y} ] [\underline{m}^{\star}_{y}] SR
[2_{x\bar{y}}2_{xy}2_z ] [2^{\star}_{xz} ], [4_y] [4_z3_{p}2_{xy} ] [[B2\bar{B}]] (h)   [[101]] (11) [2^{\star}_{xz}]   1 [{\rm AR}^{\star}]
  [[\bar{1}B1]] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] SI
[2_{x\bar{y}}2_{xy}2_z ] [m_{xz}^{\star} ], [\bar4_y] [m_z\bar{3}_{p}m_{xy} ] [[B2\bar{B}]] (h)   [[101]] (11) [\underline{m}_{xz}^{\star}] [\underline{m}_{xz}^{\star}] [\underline{m}_{xz}^{\star}] SI
  [[\bar{1}B1] ] [m_{xz}^{\star}]   1 [{\rm AR}^{\star}]
[m_xm_y2_z] [m^{\star}_{x\bar{y}} ], [m^{\star}_{xy}] [4^{\star}_zm_xm^{\star}_{xy} ] [[001]]   [\Bigl[] [[110]] (10) [m^{\star}_{x\bar{y}}\underline{m}^{\star}_{xy}\underline2_z ] [\underline{m}^{\star}_{xy} ] [\underline{m}^{\star}_{xy}] SR
  [[1\bar10]] [\underline{m}^{\star}_{x\bar{y}}{m}^{\star}_{xy}\underline2_z ] [\underline{m}^{\star}_{x\bar{y}} ] [\underline{m}^{\star}_{x\bar{y}} ] SR
[m_xm_y2_z] [2^{\star}_{x\bar{y}} ], [2^{\star}_{xy}] [\bar{4}^{\star}_zm_x2^{\star}_{xy} ] [[001]]   [\Bigl[] [{[110]}] (10) [2^{\star}_{xy}\underline2^{\star}_{x\bar{y}}\underline2_z ] [\underline2^{\star}_{x\bar{y}} ] [\underline2^{\star}_{x\bar{y}} ] SR
  [[1\bar10]] [\underline2^{\star}_{xy}2^{\star}_{x\bar{y}}\underline2_z ] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy} ] SR
[m_xm_y2_z] [m^{\star}_{x{^\prime}} ], [m^{\star}_{y{^\prime}}] [6_zm_xm_y] [[001]]     [[\bar1\sqrt{3}0] ] (9) [\underline{m}^{\star}_{x{^\prime}}m^{\star}_{y{^\prime}}\underline2_z ] [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}} ] SR
    [[\sqrt{3}10] ] [{m}^{\star}_{x{^\prime}}\underline{m}^{\star}_{y{^\prime}}\underline2_z ] [\underline{m}^{\star}_{y{^\prime}} ] [\underline{m}^{\star}_{y{^\prime}} ] SR
[m_xm_y2_z] [2^{\star}_{x{^\prime}} ], [2^{\star}_{y{^\prime}}] [6_z/m_zm_xm_y ] [[001]]     [[\bar1\sqrt{3}0] ] (9) [2^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}\underline2_z ] [\underline2^{\star}_{y{^\prime}} ] [\underline2^{\star}_{y{^\prime}}] SR
    [[\sqrt{3}10] ] [\underline2^{\star}_{x{^\prime}}2^{\star}_{y{^\prime}}\underline2_z ] [\underline2^{\star}_{x{^\prime}} ] [\underline2^{\star}_{x{^\prime}} ] SR
[m_x2_ym_z] [m^{\star}_{x{^\prime}} ], [2^{\star}_{y{^\prime}}] [\bar{6}_zm_x2_y ] [[001]]     [[\bar1\sqrt{3}0]_e ] (9) [\underline{m}^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}m_z ] [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}m_z ] SI
    [[\sqrt{3}10] ] [{m}^{\star}_{x{^\prime}}2^{\star}_{y{^\prime}}m_z ]   [m_z] [{\rm AR}^{\star}]
[2_xm_ym_z] [m^{\star}_{x\bar{y}} ], [2^{\star}_{xy}] [4_z/m_zm_xm_{xy} ] [[001]]     [[110] ] (10) [2^{\star}_{xy}m^{\star}_{x\bar{y}}m_z ]   [m_z] [{\rm AR}^{\star}]
  [[1\bar10]_e] [\underline2^{\star}_{xy}\underline{m}^{\star}_{x\bar{y}}m_z ] [\underline{m}^{\star}_{x\bar{y}} ] [\underline2^{\star}_{xy}\underline{m}^{\star}_{x\bar{y}}m_z ] SI
[2_xm_ym_z] [m^{\star}_{y{^\prime}} ], [2^{\star}_{x{^\prime}}] [\bar{6}_z2_xm_y ] [[001]]     [[\bar1\sqrt{3}0] ] (9) [2^{\star}_{x{^\prime}}m^{\star}_{y{^\prime}}m_z ]   [m_z] [{\rm AR}^{\star}]
    [[\sqrt{3}10]_e ]   [\underline{2}^{\star}_{x{^\prime}}\underline{m}^{\star}_{y{^\prime}}m_z ] [\underline{m}_{y{^\prime}} ] [\underline{2}^{\star}_{x{^\prime}}\underline{m}^{\star}_{y{^\prime}}m_z ] SI
[2_xm_ym_z] [m^{\star}_{x{^\prime}} ], [2^{\star}_{y{^\prime}}] [6_z/m_zm_xm_y ] [[001]]     [[\bar1\sqrt{3}0]_e] (9) [\underline{m}^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}m_z ] [\underline{m}^{\star}_{x{^\prime}} ] [\underline{m}^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}m_z ] SI
    [[\sqrt{3}10] ] [m^{\star}_{x{^\prime}}2^{\star}_{y{^\prime}}m_z ]   [m_z] [{\rm AR}^{\star}]
[m_{x\bar{y}}m_{xy}2_z] [2^{\star}_{x}], [2^{\star}_{y} ] [\bar{4}^{\star}_z2^{\star}_xm_{xy}] [[001]]   [\Bigl[] [[100] ] (12) [2^{\star}_{x}\underline2^{\star}_{y}\underline2_z ] [\underline2^{\star}_{y}] [\underline2^{\star}_{y}] SR
          [[010]]   [\underline2^{\star}_{x}2^{\star}_{y}\underline2_z ] [\underline2^{\star}_{x} ] [\underline2^{\star}_{x}] SR
[m_{x\bar{y}}m_{xy}2_z ] [m^{\star}_{xz} ], [\bar4_y] [\bar{4}_z3_{p}m_{xy} ] [[B2\bar{B}]] (h)   [[101]_e] (11) [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SI
  [[\bar{1}B1] ]   [m^{\star}_{xz}]   1 [{\rm AR}^{\star}]
[m_{x\bar{y}}m_{xy}2_z ] [2^{\star}_{xz} ], [4_y] [m_z\bar{3}_{p}m_{xy} ] [[B2\bar{B}]] (h)   [[101] ] (11) [2^{\star}_{xz}]   1 [{\rm AR}^{\star}]
  [[\bar{1}B1]_e] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz} ] SI
[m_{x\bar{y}}2_{xy}m_z ] [m^{\star}_{xz} ], [4_y] [m_z\bar{3}_{p}m_{xy}(m^{\star}_{xz}) ] [[B2\bar{B}]] (h)   [[101]_e] (11) [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SI
  [[\bar{1}B1] ] [m^{\star}_{xz}]   [1] [{\rm AR}^{\star}]
[m_{x\bar{y}}2_{xy}m_z ] [2^{\star}_{xz} ], [\bar4_y] [m_z\bar{3}_{p}m_{xy}(2^{\star}_{xz}) ] [[B2\bar{B}]] (h)   [[101]] (11) [2^{\star}_{xz}]   1 [{\rm AR}^{\star}]
  [[\bar{1}B1]_e ] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] SI
[m_xm_ym_z] [m^{\star}_{xy} ], [m^{\star}_{x\bar{y}}] [4^{\star}_z/m_zm_xm^{\star}_{xy} ] [[001]]   [\Bigl[] [[110]] (10) [m^{\star}_{x\bar{y}}\underline{m}^{\star}_{xy}m_z ] [\underline{m}^{\star}_{xy} ] [\underline2^{\star}_{x\bar{y}}\underline{m}^{\star}_{xy}m_z ] SR
  [[1\bar10]] [\underline{m}^{\star}_{x\bar{y}}{m}^{\star}_{xy}m_z ] [\underline{m}^{\star}_{x\bar{y}} ] [\underline{m}^{\star}_{x\bar{y}}\underline{2}^{\star}_{xy}m_z ] SR
[m_xm_ym_z] [m^{\star}_{x{^\prime}} ], [m^{\star}_{y{^\prime}}] [6_z/m_zm_xm_y ] [[001]]     [[\bar1\sqrt{3}0] ] (9) [\underline{m}^{\star}_{x{^\prime}}m^{\star}_{y{^\prime}}m_z ] [\underline{m}^{\star}_{x{^\prime}}] [\underline{m}^{\star}_{x{^\prime}}\underline2^{\star}_{y{^\prime}}m_z ] SR
    [[\sqrt{3}10] ] [m^{\star}_{x'}\underline{m}^{\star}_{y'}m_z ] [\underline{m}^{\star}_{y'}] [\underline{2}^{\star}_{x'}\underline{m}^{\star}_{y'}m_z ] SR
[m_{xy}m_{\bar{x}y}m_z ] [m^{\star}_{xz} ], [4_y] [m_z\bar{3}_{p}m_{xy} ] [[B2\bar{B}]] (h)   [[101]] (11) [2^{\star}_{xz}/\underline{m}^{\star}_{xz} ] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SR
  [[\bar{1}B1]] [\underline2^{\star}_{xz}/m^{\star}_{xz} ] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz} ] SR
[4_z] [2^{\star}_{xz} ], [4_y] [4_z3_{p}2_{xy} ] [[010]]     [[101]] (13) [2^{\star}_{xz}]   1 [{\rm AR}^{\star}]
    [[\bar101]_e ] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] SI
[4_z] [m^{\star}_{xz} ], [\bar{4}_y] [m_z\bar{3}_{p}m_{xy} ] [[010]]     [[101]_e] (13) [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SI
    [[\bar101]] [m^{\star}_{xz}]   [1] [{\rm AR}^{\star}]
[\bar{4}_z] [m^{\star}_{xz} ], [\bar4_y] [\bar{4}_z3_{p}m_{xy} ] [[010]]     [[101]] (13) [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SI
    [[\bar101]] [m^{\star}_{xz}]   [1] [{\rm AR}^{\star}]
[\bar{4}_z] [2^{\star}_{xz} ], [4_y] [m_z\bar{3}_{p}m_{xy} ] [[010]]     [[101] ] (13) [2^{\star}_{xz}]   [1] [{\rm AR}^{\star}]
    [[\bar101]_e ] [\underline2_{xz}^{\star}] [\underline2_{xz}^{\star}] [\underline2_{xz}^{\star}] SI
[4_z/m_z] [m^{\star}_{xz} ], [4_y] [m_z\bar{3}_{p}m_{xy} ] [[010]]     [[101]] (13) [2^{\star}_{xz}/\underline{m}^{\star}_{xz} ] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SR
    [[\bar101]] [\underline2^{\star}_{xz}/m^{\star}_{xz} ] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}] SR
[4_z2_x2_{xy} ] [2^{\star}_{xz} ], [2^{\star}_{x\bar{z}}] [4_z3_{p}2_{xy} ] [[010]]   [\Bigl[] [[101]] (13) [2^{\star}_{xz}\underline2^{\star}_{x\bar{z}}\underline2_{y} ] [\underline2^{\star}_{\bar{x}z} ] [\underline2^{\star}_{\bar{x}z} ] SR
  [[\bar101]] [\underline2^{\star}_{xz}2^{\star}_{x\bar{z}}\underline2_y ] [\underline2^{\star}_{xz} ] [\underline2^{\star}_{xz} ] SR
[4_z2_x2_{xy} ] [m^{\star}_{xz} ], [m^{\star}_{x\bar{z}}] [m_z\bar{3}_{p}m_{xy} ] [[010]]   [\Bigl[] [[101] ] (13) [\underline{m}^{\star}_{xz}m^{\star}_{x\bar{z}}\underline2_y ] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SR
  [[\bar101]] [m^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}\underline2_y ] [\underline{m}^{\star}_{x\bar{z}}] [\underline{m}^{\star}_{x\bar{z}}] SR
[4_zm_xm_{xy} ] [m^{\star}_{x\bar{z}} ], [2^{\star}_{xz}] [m_z\bar{3}_{p}m_{xy} ] [[010]]     [[101]] (13) [2^{\star}_{xz}m^{\star}_{x\bar{z}}m_y ]   [m_y ] [{\rm AR}^{\star}]
    [[\bar101]_e ] [\underline2^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}m_y ] [\underline2^{\star}_{xz}] [\underline2^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}m_y ] SI
[\bar{4}_z2_xm_{xy} ] [m^{\star}_{xz} ], [m^{\star}_{x\bar{z}}] [\bar{4}_z3_{p}m_{xy} ] [[010]]   [\Bigl[] [[101]] (13) [\underline{m}^{\star}_{xz}m^{\star}_{x\bar{z}}\underline2_y ] [\underline{m}^{\star}_{xz}] [\underline{m}^{\star}_{xz}] SR
  [[\bar101]] [m^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}\underline2_y ] [\underline{m}^{\star}_{x\bar{z}} ] [\underline{m}^{\star}_{x\bar{z}} ] SR
[\bar{4}_zm_x2_{xy} ] [m_{x\bar{z}}^{\star} ], [2_{xz}^{\star}] [m_z\bar{3}_{p}m_{xy} ] [[010]]     [[101]] (13) [2^{\star}_{xz}m^{\star}_{x\bar{z}}m_y ]   [m_y] [{\rm AR}^{\star}]
    [[\bar101]] [\underline2^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}m_y ] [\underline{m}^{\star}_{x\bar{z}}] [\underline2^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}m_y ] SR
[\bar{4}_z2_xm_{xy} ] [2^{\star}_{xz} ], [2^{\star}_{x\bar{z}}] [m_z\bar{3}_{p}m_{xy} ] [[010]]     [[101]] (13) [2^{\star}_{xz}2^{\star}_{x\bar{z}}2_y ] [\underline2^{\star}_{x\bar{z}}] [\underline2^{\star}_{x\bar{z}}] SR
    [[\bar101] ] [\underline2^{\star}_{xz}\underline{2}^{\star}_{x\bar{z}}2_y ] [\underline{2}^{\star}_{x\bar{z}} ] [\underline2^{\star}_{xz}\underline{2}^{\star}_{x\bar{z}}2_y ] SI
[4_z/m_zm_xm_{xy} ] [m^{\star}_{xz} ], [m^{\star}_{\bar{x}z}] [m_z\bar{3}_{p}m_{xy} ] [[010]]   [\Bigl[] [[101]] (13) [\underline{m}^{\star}_{xz}m^{\star}_{\bar{x}z}m_y ] [\underline{m}^{\star}_{xz} ] [\underline{m}^{\star}_{xz}\underline{2}^{\star}_{\bar{x}z}m_y ] SR
  [[\bar101]] [m^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}m_y ] [\underline{m}^{\star}_{x\bar{z}} ] [\underline2^{\star}_{xz}\underline{m}^{\star}_{x\bar{z}}m_y ] SR
[3_{p}] [2^{\star}_{x} ], [3_r] [2_z 3_{p}] [[01\bar1]]     [[100]] (14) [2^{\star}_x]   [1] [{\rm AR}^{\star}]
    [[011]_e] [\underline2^{\star}_x] [\underline2^{\star}_x] [\underline2^{\star}_x] SI
[3_{p}] [m^{\star}_{x} ], [\bar3_r] [m_z \bar3_{p} ] [[01\bar1]]     [[100]_e] (14) [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x ] SI
    [[011] ] [m^{\star}_x ]   [1] [{\rm AR}^{\star}]
[3_{p}] [2^{\star}_{xy} ], [4_y] [4_z 3_{p} 2_{xy} ] [[1\bar10]]     [[001]_e] (14) [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] SI
    [[110]] [2^{\star}_{xy}]   [1] [{\rm AR}^{\star}]
[3_{p}] [m^{\star}_{xy} ], [\bar4_y] [\bar4_z3_{p}m_{xy} ] [[1\bar10]]     [[001]] (14) [m^{\star}_{xy}]   [1] [{\rm AR}^{\star}]
    [[110]_e] [\underline{m}^{\star}_{xy}] [\underline{m}^{\star}_{xy}] [\underline{m}^{\star}_{xy}] SI
[\bar3_{p}] [m^{\star}_{x} ], [3_r] [m_z\bar3_{p} ] [[01\bar1]]     [[100]] (14) [2^{\star}_x/\underline{m}^{\star}_x ] [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x ] SR
    [[011]] [\underline2^{\star}_x/m^{\star}_x] [\underline2^{\star}_x ] [\underline2^{\star}_x] SR
[\bar3_{p}] [m^{\star}_{xy} ], [4_y] [m_z\bar3_{p}m_{xy} ] [[1\bar10]]     [[001]] (14) [\underline2^{\star}_{xy}/m^{\star}_{xy} ] [\underline2^{\star}_{xy}] [\underline2^{\star}_{xy}] SR
    [[110]] [2^{\star}_{xy}/\underline{m}^{\star}_{xy} ] [\underline{m}^{\star}_{xy}] [\underline{m}^{\star}_{xy}] SR
[3_{p} 2_{x\bar{y}} ] [2^{\star}_{x} ], [2^{\star}_{yz}] [4_z3_{p} 2_{xy} ] [[01\bar1]]     [[100]] (14) [2^{\star}_x\underline2^{\star}_{yz}\underline2_{y\bar{z}} ] [\underline2^{\star}_{yz} ] [\underline2^{\star}_{yz} ] SR
    [[011]] [\underline2^{\star}_x2^{\star}_{yz}\underline2_{y\bar{z}} ] [\underline2^{\star}_x] [\underline2^{\star}_x] SR
[3_{p} 2_{x\bar{y}} ] [m^{\star}_{x} ], [m^{\star}_{yz}] [m_z\bar3_{p}m_{xy} ] [[01\bar1]]     [[100]] (14) [\underline{m}^{\star}_xm_{yz}^{\star}\underline2_{y\bar{z}} ] [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x ] SR
    [[011]] [m^{\star}_x\underline{m}^{\star}_{yz}\underline2_{y\bar{z}} ] [\underline{m}^{\star}_{yz}] [\underline{m}^{\star}_{yz}] SR
[3_{p} m_{x\bar{y}} ] [2^{\star}_{x} ], [m^{\star}_{yz}] [\bar4_z3_{p}m_{xy} ] [[01\bar1]]     [[100] ] (14) [m^{\star}_{yz}m_{y\bar{z}}2^{\star}_x ]   [m_{y\bar{z}}] [{\rm AR}^{\star}]
    [[011]_e] [\underline{m}_{yz}m_{y\bar{z}}\underline2^{\star}_x ] [\underline{m}_{yz} ] [\underline{m}_{yz}m_{y\bar{z}}\underline2^{\star}_x ] SI
[3_{p} m_{x\bar{y}} ] [m^{\star}_x], [2^{\star}_{yz}] [m_z\bar3_{p}m_{xy} ] [[01\bar1]]     [[100]_e] (14) [\underline{m}^{\star}_x\underline2^{\star}_{yz}m_{y\bar{z}} ] [\underline{m}^{\star}_x ] [\underline{m}^{\star}_x\underline2^{\star}_{yz}m_{y\bar{z}} ] SI
    [[011]] [m^{\star}_x2^{\star}_{yz}m_{y\bar{z}} ]   [m_{y\bar{z}}] [{\rm AR}^{\star}]
[\bar3_{p}m_{x\bar{y}}] [m^{\star}_x], [m^{\star}_{yz} ] [m_z \bar3_{p}m_{xy}] [[01\bar1]]     [[100]] (14) [\underline{m}^{\star}_xm^{\star}_{yz}m_{y\bar{z}} ] [\underline{m}^{\star}_x] [\underline{m}^{\star}_x\underline2^{\star}_{yz}m_{y\bar{z}} ] SR
    [[011]] [m^{\star}_x\underline{m}^{\star}_{yz}m_{y\bar{z}} ] [\underline{m}^{\star}_{yz}] [\underline2^{\star}_x\underline{m}^{\star}_{yz}m_{y\bar{z}} ] SR
Equations for directions of axes and shear angle [\varphi]:[\matrix{\pmatrix{a &f &e \cr f &b &d \cr e &d &c \cr}\quad\quad\quad\quad\quad\quad\quad\quad\quad\hfill& \pmatrix{a & d & 0 \cr d & b & 0 \cr 0 & 0 & c \cr}\hfill& \pmatrix{a & 0 & 0 \cr 0 & b & d \cr 0 & d & c \cr}\hfill\cr (a)\quad [001]\quad[1 {{\displaystyle d}\over{\displaystyle e}} 0]\hfill & (b)\quad [\bar1 \alpha 0]\quad [\alpha10]\hfill & (d)\quad[110] \quad [1\bar1 {{\displaystyle 2d}\over{\displaystyle a-b}}]\hfill\cr &\alpha={{\displaystyle 2d+\sqrt{(a-b)^2+4d^2}}\over{\displaystyle a-b}}\hfill& (e)\quad[\bar1\sqrt30]\quad [\sqrt31{{\displaystyle 4d}\over{\displaystyle b-a}}]\hfill\cr & (c)\quad[\bar1 \beta 0] \quad [\beta 10]\hfill& (f)\quad[\sqrt310]\quad [\bar1\sqrt3{{\displaystyle 4\sqrt3}\over{\displaystyle 3(a-b)}}]\hfill\cr & \beta={{\displaystyle(a-b)+2\sqrt{3}d+4\sqrt{(a-b)^2+4d^2}}\over{\displaystyle \sqrt3(a-b)-2d}}\hfill &\cr&&&\cr(1)\quad 2 \sqrt{d^2+e^2}\hfill & (2)\quad 2|d|\hfill & (5)\quad \sqrt{(a-b)^2+2d^2}\hfill\cr &(3)\quad {{1}\over{2}}\sqrt{(a-b)^2+4d^2}\hfill &(6)\quad {{\sqrt3}\over{2}}\sqrt{(a-b)^2+4d^2}\hfill\cr &(4)\quad {{\sqrt3}\over{2}}\sqrt{(a-b)^2+4d^2}\hfill & (7)\quad 2|d|\hfill\cr &&&\cr &&&\cr\pmatrix{a & b & -d \cr b & a & d \cr -d & d & c \cr}\hfill & \pmatrix{a & 0 & 0 \cr 0 & b & 0 \cr 0 & 0 & c \cr}\hfill & \pmatrix{a & d & 0 \cr d & a & 0 \cr 0 & 0 & c \cr}\hfill\cr (g)\quad[100]\quad[0\bar1 {{\displaystyle d}\over{\displaystyle b}}]\hfill & &(h)\quad[101]\quad[\bar1 {{\displaystyle 2d}\over{\displaystyle c-d}}1]\hfill\cr &&&\cr (8)\quad 2\sqrt{a^2+b^2}\hfill & (9)\quad {{\sqrt3}\over{2}} |a-b|\hfill & (11)\quad \sqrt{(a-c)^2+2d^2}\hfill\cr &(10)\quad|a-b|\hfill &(12)\quad 2|d|\hfill\cr &&&\cr &&&\cr\pmatrix{a & 0 & 0 \cr 0 & a & 0 \cr 0 & 0 & c \cr}\hfill&\pmatrix{a & d & d \cr d & a & d \cr d & d & a \cr}\hfill &\cr(13)\quad |a-c|\hfill & (14)\quad 2\sqrt2 |d|\hfill &\cr} ]

The first three columns specify domain pairs.

  • [F_1]: point-group symmetry (stabilizerin [K_{1j}]) of the first domain state [{\bf S}_1] in a single-domain orientation.

  • [g_{1j}]: switching operations (if available) that specify the domain pair [({\bf S}_1,{\bf S}_j =g_{1j}{\bf S}_1)]. Subscripts [x,y,z] specify the orientation of the symmetry operations in the Cartesian coordinate system of [K_{1j}]. Subscripts [x^\prime, y^\prime ] and [x^{\prime\prime}, y^{\prime\prime}] denote a Cartesian coordinate system rotated about the z axis through 120 and 240°, respectively, from the Cartesian coordinate axes x and y. Diagonal directions are abbreviated: [p=[111]], [q=[\bar 1\bar 11] ], [r=[1\bar 1\bar 1]], [s=[\bar 1 1 \bar 1]]. Where possible, reflections and 180° rotations are chosen such that the two perpendicular permissible walls have crystallographic orientations.

  • [K_{1j}]: twinning group of the domain pair [({\bf S}_1,{\bf S}_j)]. For the pair with [F_1=m_{x\bar y}2_{xy}m_{z} ] and [K=m\bar3 m], where the twinning group does not specify the domain pair unambiguously, we add after [K_{1j}] in parentheses a switching operation [2^{\star}_{xz}] or [m^{\star}_{xz}] that defines the domain pair.

  • Axis : axis of ferroelastic domain pair around which single-domain states must be rotated to establish a contact along a compatible domain wall. This axis is parallel to the intersection of the two compatible domain walls given in the column Wall normals and its direction [{\bf h}] is defined by a vector product [{\bf h}={\bf n}_1 \times {\bf n}_2 ] of normal vectors [{\bf n}_1] and [{\bf n}_2] of these walls. Letters B and C denote components of h which depend on spontaneous strain.

  • Equation : a reference to an expression, given at the end of the table, for the direction [{\bf h}] of the axis, where parameters B and C in the column Axis are expressed as functions of spontaneous strain components. The matrices above these expressions give the form of the `absolute' spontaneous strain.

  • Wall normals : orientation of equally deformed planes. As explained above, each plane represents two mutually reversed compatible domain walls. Numbers or parameters B, C given in parentheses can be interpreted either as components of normal vectors to compatible walls or as intercepts analogous to Miller indices: Planes of compatible domain walls [Ax_1+Bx_2+Cx_3=0] and [A{^\prime}x_1+B{^\prime}x_2+C{^\prime}x_3=0 ] [see equations (3.4.3.55[link])] pass through the origin of the Cartesian coordinate system of [K_{1j}] and have normal vectors [{\bf n}_1=[ABC]] and [{\bf n}_2=[A{^\prime}B{^\prime}C{^\prime}]]. It is possible to find a plane with the same normal vector [[ABC]] but not passing through the origin, e.g. [Ax_1+Bx_2+Cx_3=1]. Then parameters A, B and C can be interpreted as the reciprocal values of the oriented intercepts on the coordinate axes cut by this plane, [[x_1/(1/A)] + [x_2/(1/B)]+[x_3/(1/C)] =1]. In analogy with Miller indices, the symbol [(ABC)] is used for expressing the orientation of a wall. However, parameters A, B and C are not Miller indices, since they are expressed in an orthonormal and not a crystallographic coordinate system. A left square bracket [ in front of two equally deformed planes signifies that the two domain walls (domain twins) associated with one equally deformed plane are crystallographically equivalent (in [K_{1j}]) with two domain walls (twins) associated with the perpendicular equally deformed plane, i.e. all four compatible domain walls (domain twins) that can be formed from domain pair [({\bf S}_1,{\bf S}_j) ] are crystallographically equivalent in [K_{1j}].

    The subscript e indicates that the wall carries a nonzero polarization charge, [\hbox{Div }{\bf P}\neq0]. This can happen in ferroelectric domain pairs with spontaneous polarization not parallel to the axis of the pair. If one domain wall is charged then the perpendicular wall is not charged. In a few cases, polarization and/or orientation of the domain wall is not determined by symmetry; then it is not possible to specify which of the two walls is charged. In such cases, a subscript [e0] or [0e] indicates that one of the two walls is charged and the other is not.

  • [\varphi]: reference to an expression, given at the end of the table, in which the shear angle [\varphi] (in radians) is given as a function of the `absolute' spontaneous strain components, which are defined in a matrix given above the equations.

  • [{\sf \overline J}_{1j}]: symmetry of the `twin pair'. The meaning of this group and its symbol is explained in the next section. This group specifies the symmetry properties of a ferroelastic domain twin and the reversed twin with compatible walls of a given orientation and with domain states [{\bf S}_1^+], [{\bf S}_j^-] and [{\bf S}_1^+], [{\bf S}_j^-]. This group can be used for designating a twin law of the ferroelastic domain twin.

  • [{\underline t}_{1j}^{\star}]: one non-trivial twinning operation of the twin [{\bf S}_1[ABC]{\bf S}_j] and the wall. An underlined symbol with a star symbol signifies an operation that inverts the wall normal and exchanges the domain states (see the next section).

  • [{\sf \overline T}_{1j}]: layer-group symmetry of the ferroelastic domain twin and the reversed twin with compatible walls of a given orientation. Contains all trivial and non-trivial symmetry operations of the domain twin (see the next section).

  • Classification : symbol that specifies the type of domain twin and the wall. Five types of twins and domain walls are given in Table 3.4.4.3[link]. The letter S denotes a symmetric domain twin (wall) in which the structures in two half-spaces are related by a symmetry operation of the twin, A denotes an asymmetric twin where there is no such relation. The letters R (reversible) and I (irreversible) signify whether a twin and reversed twin are, or are not, crystallographically equivalent in [K_{1j}].

Example 3.4.3.7. The rhombohedral phase of perovskite crystals.  Examples include PZN-PT and PMN-PT solid solutions (see e.g. Erhart & Cao, 2001[link]) and BaTiO3 below 183 K. The phase transition has symmetry descent [m\bar 3m\supset 3m].

In Table 3.4.2.7[link] we find that there are eight domain states and eight ferroelectric domain states. In this fully ferroelectric phase, domain states can be specified by unit vectors representing the direction of spontaneous polarization. We choose [{\bf S}_1\equiv [111]] with corresponding symmetry group [F_1=3_pm_{z\bar y}].

From eight domain states one can form [7\times8=56] domain pairs. These pairs can be divided into classes of equivalent pairs which are specified by different twinning groups. In column [K_{1j}] of Table 3.4.2.7[link] we find three twinning groups:

  • (i) The first twin law [\bar3_p^{\star}m_{x\bar y} ] characterizes a non-ferroelastic pair (Fam[\bar3_p^{\star}m_{x\bar y} =] Fam[3_p^{\star}m_{x\bar y}]) with inversion [\bar 1] as a twinning operation of this pair. A representative domain pair is [({\bf S}_1,g_{12}{\bf S}_1=] [{\bf S}_2) =] [([111],[\bar1\bar1\bar1])], domain pairs consist of two domain states with antiparallel spontaneous polarization (`180° pairs'). Domain walls of low energy are not charged, i.e. they are parallel with the spontaneous polarization.

  • (ii) The second twinning group [K_{13}=\bar4 3m ] characterizes a ferro­elastic domain pair (Fam[\bar4 3m=] [m\bar 3m] [\neq] Fam[F_1=] [\bar3_pm_{z\bar y}]). In Table 3.4.3.6[link], we find [g_{13}^{\star}=2^{\star}_x], which defines the representative pair [([111],[1\bar 1\bar 1])] (`109° pairs'). Orientations of compatible domain walls of this domain pair are [(100)] and [(011)_e] (this wall is charged). All equivalent orientations of these compatible walls will appear if all crystallographically equivalent pairs are considered.

  • (iii) The third twinning group [K_{14}=m\bar 3m ] also represents ferroelastic domain pairs with representative pair [([111],m^{\star}_x[111])=] [([111],[\bar 111]) ] (`71° pairs') and compatible wall orientations [(100)_e ] and (011). We see that for a given crystallographic orientation both charged and non-charged domain walls exist; for a given orientation the charge specifies to which class the domain wall belongs.

These conclusions are useful in deciphering the `domain-engineered structures' of these crystals (Yin & Cao, 2000[link]).

3.4.3.6.5. Ferroelastic domain pairs with no compatible domain walls, synoptic table

| top | pdf |

Ferroelastic domain pairs for which condition (3.4.3.54[link]) for the existence of coherent domain walls is violated are listed in Table 3.4.3.7[link]. All these pairs are non-transposable pairs. It is expected that domain walls between ferroelastic domain states would be stressed and would contain dislocations. Dudnik & Shuvalov (1989[link]) have shown that in thin samples, where elastic stresses are reduced, `almost coherent' ferroelastic domain walls may exist.

Table 3.4.3.7 | top | pdf |
Ferroelastic domain pairs with no compatible domain walls

[F_1] is the symmetry of [{\bf S}_1], [g_{1j} ] is the switching operation, [K_{1j}] is the twinning group. Pair is the domain pair type, where ns is non-transposable simple and nm is non-transposable multiple (see Table 3.4.3.2[link]). [ v= z], [p=[111]], [q=[\bar 1\bar 11]], [r =[1\bar 1\bar 1]], [s=[\bar 1 1 \bar 1]] (see Table 3.4.2.5[link] and Fig. 3.4.2.3[link]).

[F_1] [g_{1j} ] [K_{1j} ] Pair
[1 ] [4_z ] [4_z ] ns
[1 ] [\bar4_z] [\bar4_z ] ns
[1 ] [3_v] [3_v] ns
[1 ] [\bar3_v] [\bar3_v] ns
[1 ] [6_z ] [6_z ] ns
[1 ] [\bar6_z] [\bar6_z ] ns
[\bar1 ] [4_z], [4_z^3] [4_z/m_z ] ns
[\bar1 ] [3_v], [3_v^2] [\bar3_v] ns
[\bar1 ] [6_z], [6_z^5] [6_z/m_z ] ns
[2_z ] [3_p], [3_p^2] [2_z3_p ] nm
[2_z ] [\bar3_p], [\bar3_p^5 ] [m_z\bar3_p] nm
[2_{xy} ] [3_p], [3_p^2] [4_z3_p2_{xy} ] nm
[2_{xy} ] [\bar3_p], [\bar3_p^5 ] [m_z\bar3_pm_{xy}] nm
[m_z ] [3_p], [3_p^2] [m_z3_p^2 ] nm
[m_{xy} ] [3_p], [3_p^2] [\bar4_z3_pm_{xy} ] nm
[m_{xy} ] [4_x], [4_x^3] [m_z\bar3_pm_{xy}] nm
[2_z/m_z ] [3_p], [3_p^2] [m_z\bar3_p] nm
[2_{xy}/m_{xy} ] [3_p], [3_p^2] [m_z\bar3_pm_{xy}] nm
[2_x2_y2_z ] [3_p], [3_p^2] [2_z3_p ] ns
[2_x2_y2_z ] [\bar3_p], [\bar3_p^5 ] [m_z\bar3_p ] ns
[m_xm_y2_z ] [3_p], [3_p^2 ] [m_z\bar3_p ] nm
[m_xm_ym_z ] [3_p], [3_p^2 ] [m_z\bar3_p ] ns

Example 3.4.3.8. Ferroelastic crystal of langbeinite.  Langbeinite K2Mg2(SO4)3 undergoes a phase transition with symmetry descent [23\supset 222] that appears in Table 3.4.3.7[link]. The ferroelastic phase has three ferroelastic domain states. Dudnik & Shuvalov (1989[link]) found, in accord with their theoretical predictions, nearly linear `almost coherent' domain walls accompanied by elastic stresses in crystals thinner than 0.5 mm. In thicker crystals, elastic stresses became so large that crystals were cracking and no domain walls were observed.

Similar effects were reported by the same authors for the partial ferroelastic phase of CH3NH3Al(SO4)2·12H2O (MASD) with symmetry descent [\bar 3m\supset mmm], where ferroelastic domain walls were detected only in thin samples.

3.4.3.7. Domain pairs in the microscopic description

| top | pdf |

In the microscopic description, two microscopic domain states [{\sf S}_{i}] and [{\sf S}_{k}] with space-group symmetries [{\cal F}_i] and [{\cal F}_k], respectively, can form an ordered domain pair ([{\sf S}_{i},{\sf S}_{k}]) and an unordered domain pair [\{{\sf S}_{i},{\sf S}_{k}\}] in a similar way to in the continuum description, but one additional aspect has to be considered. The definition of the symmetry group [{\cal F}_{ik}] of an ordered domain pair ([{\sf S}_{i},{\sf S}_{k} ]), [{\cal F}_{ik} = {\cal F}_i \cap {\cal F}_k, \eqno(3.4.3.72) ]is meaningful only if the group [{\cal F}_{ik}] is a space group with a three-dimensional translational subgroup (three-dimensional twin lattice in the classical description of twinning, see Section 3.3.8[link] ) [{\cal T}_{ik}={\cal T}_i \cap {\cal T}_k, \eqno(3.4.3.73) ]where [{\cal T}_i] and [{\cal T}_k] are translation subgroups of [{\cal F}_i] and [{\cal F}_k], respectively. This condition is fulfilled if both domain states [{\sf S}_{i}] and [{\sf S}_{k} ] have the same spontaneous strains, i.e. in non-ferroelastic domain pairs, but in ferroelastic domain pairs one has to suppress spontaneous deformations by applying the parent clamping approximation [see Section 3.4.2.2[link], equation (3.4.2.49[link])].

Example 3.4.3.9. Domain pairs in calomel.  Calomel undergoes a non-equitranslational phase transition from a tetragonal parent phase to an orthorhombic ferroelastic phase (see Example 3.4.2.7[link] in Section 3.4.2.5[link]). Four basic microscopic single-domain states are displayed in Fig. 3.4.2.5[link]. From these states, one can form 12 non-trivial ordered single-domain pairs that can be partitioned (by means of double coset decomposition) into two orbits of domain pairs.

Representative domain pairs of these orbits are depicted in Fig. 3.4.3.10[link], where the first microscopic domain state [{\sf S}_{i} ] participating in a domain pair is displayed in the upper cell (light grey) and the second domain state [{\sf S}_{j}], [j=2,3], in the lower white cell. The overlapping structure in the middle (dark grey) is a geometrical representation of the domain pair [\{{\sf S}_1,{\sf S}_j\} ].

[Figure 3.4.3.10]

Figure 3.4.3.10 | top | pdf |

Domain pairs in calomel. Single-domain states in the parent clamping approximation are those from Fig. 3.4.2.5. The first domain state of a domain pair is shown shaded in grey (`black'), the second domain state is colourless (`white'), and the domain pair of two interpenetrating domain states is shown shaded in dark grey. (a) Ferroelastic domain pair [({{\sf S}_1},{{\sf S}_3}) ] in the parent clamping approximation. This is a partially transposable domain pair. (b) Translational domain pair [({{\sf S}_1},{{\sf S}_2}) ]. This is a completely transposable domain pair.

The domain pair [\{{\sf S}_1,{\sf S}_3\}], depicted in Fig. 3.4.3.10[link](a), is a ferroelastic domain pair in the parent clamping approximation. Then two overlapping structures of the domain pair have a common three-dimensional lattice with a common unit cell (the dotted square), which is the same as the unit cells of domain states [{\sf S}_1 ] and [{\sf S}_3].

Domain pair [\{{\sf S}_1,{\sf S}_2\}], shown in Fig. 3.4.3.10[link](b), is a translational (antiphase) domain pair in which domain states [{\sf S}_1] and [{\sf S}_2] differ only in location but not in orientation. The unit cell (heavily outlined small square) of the domain pair [\{{\sf S}_1,{\sf S}_2\}] is identical with the unit cell of the tetragonal parent phase (cf. Fig. 3.4.2.5[link]).

The two arrows attached to the circles in the domain pairs represent exaggerated displacements within the wall.

Domain pairs represent an intermediate step in analyzing microscopic structures of domain walls, as we shall see in Section 3.4.4.[link]

3.4.4. Domain twins and domain walls

| top | pdf |

3.4.4.1. Formal description of simple domain twins and planar domain walls of zero thickness

| top | pdf |

In this section, we examine crystallographic properties of planar compatible domain walls and simple domain twins. The symmetry of these objects is described by layer groups. Since this concept is not yet common in crystallography, we briefly explain its meaning in Section 3.4.4.2[link]. The exposition is performed in the continuum description, but most of the results apply with slight generalizations to the microscopic treatment that is illustrated with an example in Section 3.4.4.7.[link]

We shall consider a simple domain twin [{\bf T}_{12}] that consists of two domains [{\bf D}_1] and [{\bf D}_2] which meet along a planar domain wall [{\bf W}_{12}] of zero thickness. Let us denote by p a plane of the domain wall, in brief wall plane of [{\bf W}_{12}]. This plane is specified by Miller indices [(hkl)], or by a normal [{\bf n}] to the plane which also defines the sidedness (plus and minus side) of the plane p. By orientation of the plane p we shall understand a specification which can, but may not, include the sidedness of p. If both the orientation and the sidedness are given, then the plane p divides the space into two half-spaces. Using the bra–ket symbols, mentioned in Section 3.4.3.6[link], we shall denote by [(\, |] the half-space on the negative side of p and by [| \,)] the half-space on the positive side of p.

A simple twin consists of two (theoretically semi-infinite) domains [{\bf D}_1] and [{\bf D}_2] with domain states [{\bf S}_1] and [{\bf S}_2], respectively, that join along a planar domain wall the orientation of which is specified by the wall plane p with normal n. A symbol [({\bf S}_1|{\bf n}|{\bf S}_2)] specifies the domain twin unequivocally: domain [({\bf S}_1|], with domain region [(\, | ] filled with domain state [{\bf S}_1], is on the negative side of p and domain [|\,{\bf S}_2)] is on the positive side of p (see Fig. 3.4.4.1[link]a).

[Figure 3.4.4.1]

Figure 3.4.4.1 | top | pdf |

Symbols of a simple twin. (a) Two different symbols with antiparallel normal [\bf n]. (b) Symbols of the reversed twin.

If we were to choose the normal of opposite direction, i.e. [-{\bf n}], the same twin would have the symbol [({\bf S}_2|-{\bf n}|{\bf S}_1) ] (see Fig. 3.4.4.1[link]a). Since these two symbols signify the same twin, we have the identity [({\bf S}_{1}|{\bf n}|{\bf S}_{2}) \equiv ({\bf S}_{2}|-{\bf n}|{\bf S}_{1}). \eqno(3.4.4.1) ]Thus, if we invert the normal n and simultaneously exchange domain states [{\bf S}_{1}] and [{\bf S}_{2}] in the twin symbol, we obtain an identical twin (see Fig. 3.4.4.1[link]a). This identity expresses the fact that the specification of the twin by the symbol introduced above does not depend on the chosen direction of the wall normal [{\bf n}].

The full symbol of the twin can be replaced by a shorter symbol [{\bf T}_{12}({\bf n}) ] if we accept a simple convention that the first lower index signifies the domain state that occupies the half space [(\, |] on the negative side of [{\bf n}]. Then the identity (3.4.4.1[link]) in short symbols is [{\bf T}_{12}({\bf n}) \equiv {\bf T}_{21}(-{\bf n}). \eqno(3.4.4.2) ]

If the orientation and sidedness of the plane p of a wall is known from the context or if it is not relevant, the specification of n in the symbol of the domain twin and domain wall can be omitted.

A twin [({\bf S}_{1}|{\bf n}|{\bf S}_2)], or [{\bf T}_{12}({\bf n}) ], can be formed by sectioning the ordered domain pair [({\bf S}_1,{\bf S}_2) ] by a plane p with normal [{\bf n}] and removing the domain state [{\bf S}_2] on the negative side and domain state [{\bf S}_2] on the positive side of the normal [{\bf n}]. This is the same procedure that is used in bicrystallography when an ideal bicrystal is derived from a dichromatic complex (see Section 3.2.2[link] ).

A twin with reversed order of domain states is called a reversed twin. The symbol of the twin reversed to the initial twin [({\bf S}_{1}|{\bf n}|{\bf S}_{2}) ] is [({\bf S}_{2}|{\bf n}|{\bf S}_{1}) \equiv ({\bf S}_{1}|-{\bf n}|{\bf S}_{2}) \, \eqno(3.4.4.3) ] or [{\bf T}_{21}({\bf n}) \equiv {\bf T}_{12}(-{\bf n}). \eqno(3.4.4.4) ]A reversed twin [({\bf S}_{2}|{\bf n}|{\bf S}_{1}) \equiv ({\bf S}_{1}|{\bf -n}|{\bf S}_{2}) ] is depicted in Fig. 3.4.4.1[link](b).

A planar domain wall is the interface between the domains [{\bf D}_1] and [{\bf D}_2] of the associated simple twin. Even a domain wall of zero thickness is specified not only by its orientation in space but also by the domain states that adhere to the minus and plus sides of the wall plane p. The symbol for the wall is, therefore, analogous to that of the twin, only in the explicit symbol the brackets ( ) are replaced by square brackets [ ] and T in the short symbol is replaced by W: [[{\bf S}_{1}|{\bf n}|{\bf S}_{2}] \equiv [{\bf S}_{2}|-{\bf n}|{\bf S}_{1}] \eqno(3.4.4.5) ]or by a shorter equivalent symbol [{\bf W}_{12}({\bf n}) \equiv {\bf W}_{21}(-{\bf n}). \eqno(3.4.4.6) ]

3.4.4.2. Layer groups

| top | pdf |

An adequate concept for characterizing symmetry properties of simple domain twins and planar domain walls is that of layer groups. A layer group describes the symmetry of objects that exist in a three-dimensional space and have two-dimensional translation symmetry. Typical examples are two-dimensional planes in three-dimensional space [two-sided planes and sectional layer groups (Holser, 1958[link]a,[link]b), domain walls and interfaces of zero thickness], layers of finite thickness (e.g. domain walls and interfaces of finite thickness) and two semi-infinite crystals joined along a planar and coherent (compatible) interface [e.g. simple domain twins with a compatible (coherent) domain wall, bicrystals].

A crystallographic layer group comprises symmetry operations (isometries) that leave invariant a chosen crystallographic plane p in a crystalline object. There are two types of such operations:

  • (i) side-preserving operations keep invariant the normal [{\bf n}] of the plane p, i.e. map each side of the plane p onto the same side. This type includes translations (discrete or continuous) in the plane p, rotations of 360°/n, n = 2, 3, 4, 6, around axes perpendicular to the plane p, reflections through planes perpendicular to p and glide reflections through planes perpendicular to p with glide vectors parallel to p. The corresponding symmetry elements are not related to the location of the plane p in space, i.e. they are the same for all planes parallel to p.

  • (ii) side-reversing operations invert the normal n of the plane i.e. exchange sides of the plane. Operations of this type are: an inversion through a point in the plane p, rotations of 360°/n, n = 3, 4, 6 around axes perpendicular to the plane followed by inversion through this point, 180° rotation and 180° screw rotation around an axis in the plane p, reflection and glide reflections through the plane p, and combinations of these operations with translations in the plane p. All corresponding symmetry elements are located in the plane p.

A layer group [{\cal L}] consists of two parts: [{\cal L}=\widehat{\cal L} \ \cup \ \underline{s}\widehat{\cal L},\eqno(3.4.4.7) ]where [\widehat{\cal L}] is a subgroup of [{\cal L}] that comprises all side-preserving operations of [{\cal L}]; this group is isomorphic to a plane group and is called a trivial layer group or a face group. An underlined character [\underline{s}] denotes a side-reversing operation and the left coset [\underline{s}\widehat{\cal L} ] contains all side-reversing operations of [{\cal L}]. Since [\widehat{\cal L}] is a halving subgroup, the layer group [{\cal L} ] can be treated as a dichromatic (black-and-white) group in which side-preserving operations are colour-preserving operations and side-reversing operations are colour-exchanging operations.

There are 80 layer groups with discrete two-dimensional translation subgroups [for a detailed treatment see IT E (2002)[link], or e.g. Vainshtein (1994[link]), Shubnikov & Kopcik (1974[link]), Holser (1958[link]a)]. Equivalent names for these layer groups are net groups (Opechowski, 1986[link]), plane groups in three dimensions (Grell et al., 1989[link]), groups in a two-sided plane (Holser, 1958[link]a,b[link]) and others.

To these layer groups there correspond 31 point groups that describe the symmetries of crystallographic objects with two-dimensional continuous translations. Holser (1958[link]b) calls these groups point groups in a two-sided plane, Kopský (1993[link]) coins the term point-like layer groups. We shall use the term `layer groups' both for layer groups with discrete translations, used in a microscopic description, and for crystallographic `point-like layer groups' with continuous translations in the continuum approach. The geometrical meaning of these groups is similar and most of the statements and formulae hold for both types of layer groups.

Crystallographic layer groups with a continuous translation group [point groups of two-sided plane (Holser, 1958[link]b)] are listed in Table 3.4.4.1[link]. The international notation corresponds to international symbols of layer groups with discrete translations; this notation is based on the Hermann–Mauguin (international) symbols of three-dimensional space groups, where the c direction is the direction of missing translations and the character `1' represents a symmetry direction in the plane with no associated symmetry element (see IT E , 2002[link]).

Table 3.4.4.1 | top | pdf |
Crystallographic layer groups with continuous translations

International Non-coordinate
[1 ] [1 ]
[\bar1 ] [\underline{\bar1} ]
[112 ] [2 ]
[11m ] [\underline{m} ]
[112/m ] [2/\underline{m} ]
[211 ] [\underline{2} ]
[m11 ] [\underline{m} ]
[2/m11 ] [\underline{2}/m ]
[222 ] [\underline{2}\underline{2}2 ]
[mm2 ] [mm2 ]
[m2m ] [m\underline{2}\underline{m} ]
[mmm ] [mm\underline{m}]
[4 ] [4 ]
[\bar4 ] [\underline{\bar4}]
[4/m ] [4/\underline{m}]
[422 ] [4\underline{2}\underline{2} ]
[4mm ] [4mm ]
[\bar42m ] [\underline{\bar4}\underline{2}m ]
[4/mmm ] [4/\underline{m}mm ]
[3 ] [3 ]
[\bar{3}] [\underline{\bar3}]
[32 ] [3\underline{2}]
[3m ] [3m ]
[\bar{3}m] [\underline{\bar3}m ]
[6 ] [6 ]
[\bar6] [\underline{\bar6} ]
[6/m] [6/\underline{m} ]
[622] [6\underline{2}\underline{2} ]
[6mm ] [6mm ]
[\bar{6}m2] [\underline{\bar6}m\underline{2} ]
[6/mmm ] [6/\underline{m}mm]

In the non-coordinate notation (Janovec, 1981[link]), side-reversing operations are underlined. Thus e.g. [\underline{2} ] denotes a [180^{\circ}] rotation around a twofold axis in the plane p and [{\underline m}] a reflection through this plane, whereas 2 is a side-preserving [180^{\circ}] rotation around an axis perpendicular to the plane and m is a side-preserving reflection through a plane perpendicular to the plane p. With exception of [\underline{\bar{1}}] and [{\underline 2}], the symbol of an operation specifies the orientation of the plane p. This notation allows one to signify layer groups with different orientations in one reference coordinate system. Another non-coordinate notation has been introduced by Shubnikov & Kopcik (1974[link]).

If a crystal with point-group symmetry G is bisected by a crystallographic plane p, then all operations of G that leave the plane p invariant form a sectional layer group = [\overline{{\sf G}(p)} ] of the plane p in G. Operations of the group [\overline{{\sf G}(p)}] can be divided into two sets [see equation (3.4.4.7[link])]: [\overline{{\sf G}(p)}=\widehat{{\sf G}(p)} \ \cup \ \underline{g}\widehat{{\sf G}(p)}, \eqno(3.4.4.8) ]where the trivial layer group [\widehat{{\sf G}(p)}] expresses the symmetry of the crystal face with normal [{\bf n}]. These face symmetries are listed in IT A (2005[link]), Part 10[link] , for all crystallographic point groups G and all orientations of the plane expressed by Miller indices [(hkl)]. The underlined operation [\underline{g}] is a side-reversing operation that inverts the normal [{\bf n}]. The left coset [\underline{g}\widehat{{\sf G}(p)}] contains all side-reversing operations of [\overline{{\sf G}(p)}].

The number [n_p] of planes symmetrically equivalent (in G) with the plane p is equal to the index of [\overline{{\sf G}(p)} ] in G: [n_p=[G:\overline{{\sf G}(p)}]=|G|:|\overline{{\sf G}(p)}|. \eqno(3.4.4.9) ]

Example 3.4.4.1.  As an example, we find the sectional layer group of the plane [(010) ] in the group [G=4_z/m_zm_xm_{xy}] (see Fig. 3.4.2.2[link]).[\eqalignno{{4_z/m_zm_xm_{xy}(010)}&=m_x2_ym_z \cup \ \underline{m}_y\{m_x2_ym_z\}\cr &= m_x2_ym_z \cup \ \{\underline{m}_y,\underline{2_z},\underline{\bar 1},\underline{2}_x\}\cr &=m_x\underline{m}_ym_z. &(3.4.4.10)} ]

In this example [n_p=|4_z/m_zm_xm_{xy}|:|m_x\underline{m}_ym_z|=16:8= 2 ] and the plane crystallographically equivalent with the plane [(010) ] is the plane (100) with sectional symmetry [\underline{m}_xm_ym_z ].

3.4.4.3. Symmetry of simple twins and planar domain walls of zero thickness

| top | pdf |

We shall examine the symmetry of a twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j}) ] with a planar zero-thickness domain wall with orientation and location defined by a plane p (Janovec, 1981[link]; Zikmund, 1984[link]; Zieliński, 1990[link]). The symmetry properties of a planar domain wall [{\bf W}_{1j}] are the same as those of the corresponding simple domain twin. Further, we shall consider twins but all statements also apply to the corresponding domain walls.

Operations that express symmetry properties of the twin must leave the orientation and location of the plane p invariant. We shall perform our considerations in the continuum description and shall assume that the plane p passes through the origin of the coordinate system. Then point-group symmetry operations leave the origin invariant and do not change the position of p.

If we apply an operation [g\in G] to the twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j}) ], we get a crystallographically equivalent twin [({\bf S}_{i}|{\bf n}_{m}|{\bf S}_{k}){\buildrel{G}\over {\sim}} ({\bf S}_{1}|{\bf n}|{\bf S}_{j}) ] with other domain states and another orientation of the domain wall, [g({\bf S}_{1}|{\bf n}|{\bf S}_{j}) = (g{\bf S}_{1}|g{\bf n}|g{\bf S}_{j}) = ({\bf S}_{i}|{\bf n}_{m}|{\bf S}_{k}), \quad g\in G. \eqno(3.4.4.11) ]

It can be shown that the transformation of a domain pair by an operation [g\in G] defined by this relation fulfils the conditions of an action of the group G on a set of all domain pairs formed from the orbit [G{\bf S}_1] (see Section 3.2.3.3[link] ). We can, therefore, use all concepts (stabilizer, orbit, class of equivalence etc.) introduced for domain states and also for domain pairs.

Operations g that describe symmetry properties of the twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j})] must not change the orientation of the wall plane p but can reverse the sides of p, and must either leave invariant both domain states [{\bf S}_{1}] and [{\bf S}_{j} ] or exchange these two states. There are four types of such operations and their action is summarized in Table 3.4.4.2[link]. It is instructive to follow this action in Fig. 3.4.4.2[link] using an example of the twin [({\bf S}_{1}[010]{\bf S}_{2})] with domain states [{\bf S}_1] and [{\bf S}_2] from our illustrative example (see Fig. 3.4.2.2[link]).

  • (1) An operation [f_{1j}] which leaves invariant the normal n and both domain states [{\bf S}_{1}, {\bf S}_{j} ] in the twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j})]; such an operation does not change the twin and is called the trivial symmetry operation of the twin. An example of such an operation of the twin [({\bf S}_{1}[010]{\bf S}_{2}) ] in Fig. 3.4.4.2[link] is the reflection [m_z].

    Table 3.4.4.2 | top | pdf |
    Action of four types of operations g on a twin [({\bf S}_1 |{\bf n}| {\bf S}_j) ]

    Operation g keeps the orientation of the plane p unchanged.

    g [g{\bf S}_1 ] [g{\bf S}_j ] [g{\bf n}] [g({\bf S}_1 | ] [g|{\bf S}_j) ] [g({\bf S}_1 |{\bf n}| {\bf S}_j)=(g{\bf S}_1|g{\bf n}|g{\bf S}_j) ] Resulting twin
    [f_{1j}] [{\bf S}_1 ] [{\bf S}_j ] [{\bf n}] [({\bf S}_1| ] [|{\bf S}_j) ] [({\bf S}_1 |{\bf n}|{\bf S}_j) ] Initial twin
    [\underline{s}_{1j} ] [{\bf S}_1 ] [{\bf S}_j ] [{\bf -n} ] [|{\bf S}_1)] [({\bf S}_j| ] [({\bf S}_1|-{\bf n}|{\bf S}_j)\equiv({\bf S}_j|{\bf n}|{\bf S}_1) ] Reversed twin
    [{r}_{1j}^{\star} ] [{\bf S}_j ] [{\bf S}_1 ] [{\bf n}] [({\bf S}_j| ] [|{\bf S}_1)] [({\bf S}_j |{\bf n}|{\bf S}_1) ] Reversed twin
    [\underline{t}_{1j}^{\star} ] [{\bf S}_j ] [{\bf S}_1] [-{\bf n}] [|{\bf S}_j) ] [({\bf S}_1| ] [({\bf S}_j |{\bf {-n}}| {\bf S}_1) \equiv ({\bf S}_1 |{\bf {n}}| {\bf S}_j) ] Initial twin
    [Figure 3.4.4.2]

    Figure 3.4.4.2 | top | pdf |

    A simple twin under the action of four types of operation that do not change the orientation of the wall plane p. Compare with Table 3.4.4.2[link].

  • (2) An operation [\underline{s}_{1j}] which inverts the normal n but leaves invariant both domain states [{\bf S}_{1} ] and [{\bf S}_{j}]. This side-reversing operation transforms the initial twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j})] into [({\bf S}_{1}|-{\bf n}|{\bf S}_{j}) ], which is, according to (3.4.4.1[link]), identical with the inverse twin [({\bf S}_{j}|{\bf n}|{\bf S}_{1})]. As in the non-coordinate notation of layer groups (see Table 3.4.4.1[link]) we shall underline the side-reversing operations. The reflection [\underline{m}_y ] in Fig. 3.4.4.2[link] is an example of a side-reversing operation.

  • (3) An operation [r_{1j}^{\star}] which exchanges domain states [{\bf S}_{1}] and [{\bf S}_{j}] but does not change the normal n. This state-exchanging operation, denoted by a star symbol, transforms the initial twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j}) ] into a reversed twin [({\bf S}_{j}|{\bf n}|{\bf S}_{1})]. A state-exchanging operation in our example is the reflection [m_x^{\star}].

  • (4) An operation [\underline{t}_{1j}^{\star} ] which inverts n and simultaneously exchanges [{\bf S}_{1} ] and [{\bf S}_{j}]. This operation, called the non-trivial symmetry operation of a twin, transforms the initial twin into [({\bf S}_{j}|-{\bf n}|{\bf S}_{1}) ], which is, according to (3.4.4.1[link]), identical with the initial twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j})]. An operation of this type can be expressed as a product of a side-exchanging operation (underlined) and a state-exchanging operation (with a star), and will, therefore, be underlined and marked by a star. In Fig. 3.4.4.2[link], a non-trivial symmetry operation is for example the 180° rotation [\underline{2}_{z}^{\star}].

We note that the star and the underlining do not represent any operation; they are just suitable auxiliary labels that can be omitted without changing the result of the operation.

To find all trivial symmetry operations of the twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j}) ], we recall that all symmetry operations that leave both [{\bf S}_1 ] and [{\bf S}_j] invariant constitute the symmetry group [F_{1j}] of the ordered domain pair [({\bf S}_1,{\bf S}_j)], [F_{1j}=F_1 \cap F_j], where [F_1] and [F_j] are the symmetry groups of [{\bf S}_1] and [{\bf S}_j], respectively. The sectional layer group of the plane p in group [F_{1j}] is (if we omit p) [\overline{\sf F}_{1j}=\widehat{\sf F}_{1j} \cup \underline{s}_{1j}\widehat{\sf F}_{1j}.\eqno(3.4.4.12) ]The trivial (side-preserving) subgroup [\widehat{\sf F}_{1j}] assembles all trivial symmetry operations of the twin [({\bf S}_1|{\bf n}|{\bf S}_j) ]. The left coset [\underline{s}_{1j}\widehat{\sf F}_{1j}], where [\underline{s}_{1j}] is a side-reversing operation, contains all side-reversing operations of this twin. In our example [\widehat{F}_{12}=\{1,m_z\}] and [\underline{s}_{1j}\widehat{\sf F}_{1j}=\underline{m}_y\{1,m_z\}=\{\underline{m}_y,\underline{2}_x\} ] (see Fig. 3.4.4.2[link]).

Similarly, the left coset [{r}_{1j}^{\star}\widehat{\sf F}_{1j}] contains all state-exchanging operations, and [\underline{t}_{1j}^{\star}\widehat{\sf F}_{1j} ] all non-trivial symmetry operations of the twin [({\bf S}_{1}|{\bf n}|{\bf S}_j) ]. In the illustrative example, [{r}_{1j}^{\star}\widehat{\sf F}_{1j} =] [m_{x}^{\star}\{1,m_z\}=] [\{m_x^{\star},2_{y}^{\star}\}] and [\underline{t}_{1j}^{\star}\widehat{\sf F}_{1j} =] [\underline{2}_z^{\star}\{1,m_z\}=] [\{\underline{2}_z^{\star},\underline{\bar 1}^{\star}\} ].

The trivial group [\widehat{F}_{1j}] and its three cosets constitute the sectional layer group [\overline{J}_{1j}] of the plane p in the symmetry group [J_{1j}=F_{1j} \cup g_{1j}^{\star}F_{1j}] of the unordered domain pair [\{{\bf S}_1,{\bf S}_j\}], [\overline{\sf J}_{1j} = \widehat{\sf J}_{1j} \cup \underline{s}_{1j}\widehat{\sf J}_{1j} =\widehat{\sf F}_{1j} \cup r^{\star}_{1j}\widehat{\sf F}_{1j} \cup \underline{s}_{1j}\widehat{\sf F}_{1j} \cup \underline{t}^{\star}_{1j}\widehat{\sf F}_{1j}, \eqno(3.4.4.13) ]where [r_{1j}^{\star}] is an operation of the left coset [g_{1j}^{\star}F_{1j}] that leaves the normal [\bf n] invariant and [\underline{t}^*_{1j}=\underline{s}_{1j}r^{\star}_{1j}].

Group [\overline{\sf J}_{1j}] can be interpreted as a symmetry group of a twin pair [({\bf S}_1,{\bf S}_{j}|{\bf n}|{\bf S}_{j},{\bf S}_1) ] consisting of a domain twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j})] and a superposed reversed twin [({\bf S}_{j}|{\bf n}|{\bf S}_{1})] with a common wall plane p. This construct is analogous to a domain pair (dichromatic complex in bicrystallography) in which two homogeneous domain states [{\bf S}_1] and [{\bf S}_j] are superposed (see Section 3.4.3.1[link]). In the same way as the group [J_{1j}] of domain pair [\{{\bf S}_1,{\bf S}_j\}] is divided into two cosets with different results of the action on this domain pair, the symmetry group [\overline{\sf J}_{1j} ] of the twin pair can be decomposed into four cosets (3.4.4.13[link]), each of which acts on a domain twin [({\bf S}_{j}|{\bf n}|{\bf S}_{1})] in a different way, as specified in Table 3.4.4.2[link].

We can associate with operations from each coset in (3.4.4.13[link]) a label. If we denote operations from [\widehat{\sf F}_{1j}] without a label by e, underlining by a and star by b, then the multiplication of labels is expressed by the relations [a^2=b^2=e, \quad ab=ba. \eqno(3.4.4.14)]The four different labels [e, a, b, ab] can be formally viewed as four colours, the permutation of which is defined by relations (3.4.3.14[link]); then the group [\overline{\sf J}_{1j}] can be treated as a four-colour layer group.

Since the symbol of a point group consists of generators from which any operation of the group can be derived by multiplication, one can derive from the international symbol of a sectional layer group, in which generators are supplied with adequate labels, the coset decomposition (3.4.4.13[link]).

Thus for the domain pair [\{{\bf S}_1,{\bf S}_2\}] in Fig. 3.4.4.2[link] with [J_{12}^{\star}=m_x^{\star}m_ym_z] [see equation (3.4.3.18[link])] and [p(010)] we get the sectional layer group [\overline{{\sf J}_{12}(010)}=m_x^{\star}\underline{m}_ym_z ]. Operations of this group (besides generators) are [m_x^{\star}\underline{m}_y =\underline{2}_z^{\star},] [\underline{m}_ym_z =\underline{2}_x,] [m_x^{\star}m_z=] [2_y^{\star},] [m_x^{\star}\underline{2}_x=\underline{\bar 1}^{\star}].

All operations [g\in G] that transform a twin into itself constitute the symmetry group [{\sf T}_{1j}({\bf n})] (or in short [{\sf T}_{1j})] of the twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j}) ]. This is a layer group consisting of two parts: [{\sf T}_{1j} = \widehat {\sf F}_{1j} \cup \underline t^{\star}_{1j}{\widehat {\sf F}}_{1j},\eqno(3.4.4.15) ]where [\widehat {\sf F}_{1j}] is a face group comprising all trivial symmetry operations of the twin and the left coset [\underline t^*_{1j}\widehat {\sf F}_{1j} ] contains all non-trivial operations of the twin that reverse the sides of the wall plane p and simultaneously exchange the states [({\bf S}_{1} ] and [{\bf S}_{j})].

One can easily verify that the symmetry [{\sf T}_{1j}({\bf n})] of the twin [({\bf S}_{1}|{\bf n}|{\bf S}_{j})] is equal to the symmetry [{\sf T}_{j1}({\bf n})] of the reversed twin [({\bf S}_{j}|{\bf n}|{\bf S}_{1}) ], [{\sf T}_{1j}({\bf n})={\sf T}_{j1}({\bf n}).\eqno(3.4.4.16) ]

Similarly, for sectional layer groups, [\overline{{\sf F}}_{1j}({\bf n})=\overline{{\sf F}}_{j1}({\bf n}) \ \ {\rm and} \ \ \overline{{\sf J}}_{1j}({\bf n})=\overline{{\sf J}}_{j1}({\bf n}). \eqno(3.4.4.17) ]

Therefore, the symmetry of a twin [{\sf T}_{1j}(p)] and of sectional layer groups [\overline{{\sf F}}_{1j}(p)], [\overline{{\sf J}}_{1j}(p)] is specified by the orientation of the plane p [expressed e.g. by Miller indices [(hkl)]] and not by the sidedness of p. However, the two layer groups [\overline{{\sf F}}_{1j}(p) ] and [\overline{{\sf F}}_{j1}(p)], and [{\sf T}_{1j}(p)] and [{\sf T}_{j1}(p)] express the symmetry of two different objects, which can in special cases (non-transposable pairs and irreversible twins) be symmetrically non-equivalent.

The symmetry [{\sf T}_{1j}({\bf n})] also expresses the symmetry of the wall [{\bf W}_{1j}({\bf n})]. This symmetry imposes constraints on the form of tensors describing the properties of walls. In this way, the appearance of spontaneous polarization in domain walls has been examined (Přívratská & Janovec, 1999[link]; Přívratská et al., 2000[link]).

According to their symmetry, twins and walls can be divided into two types: For a symmetric twin (domain wall), there exists a non-trivial symmetry operation [\underline t^{\star}_{1j}] and its symmetry is expressed by equation (3.4.4.15[link]). A symmetric twin can be formed only from transposable domain pairs.

For an asymmetric twin (domain wall), there is no non-trivial symmetry operation and its symmetry group is, therefore, confined to trivial group [\widehat{\sf F}_{1j}], [{\sf T}_{1j} = \widehat {\sf F}_{1j}. \eqno(3.4.4.18) ]The difference between symmetric and asymmetric walls can be visualized in domain walls of finite thickness treated in Section 3.4.4.6.[link]

The symmetry [{\sf T}_{1j}] of a symmetric twin (wall), expressed by relation (3.4.4.15[link]), is a layer group but not a sectional layer group of any point group. It can, however, be derived from the sectional layer group [\overline{\sf F}_{1j}] of the corresponding ordered domain pair [({\bf S}_1,{\bf S}_j)] [see equation (3.4.4.12[link])] and the sectional layer group [\overline{\sf J}_{1j}] of the unordered domain pair [\{{\bf S}_1,{\bf S}_j\}] [see equation (3.4.4.13[link])], [{\sf T}_{1j}=\overline{\sf J}_{1j} - \{\overline{\sf F}_{1j}-\widehat{\sf F}_{1j}\} - \{\widehat{\sf J}_{1j} - \widehat{\sf F}_{1j}\}. \eqno(3.4.4.19) ]

This is particularly useful in the microscopic description, since sectional layer groups of crystallographic planes in three-dimensional space groups are tabulated in IT E (2002[link]), where one also finds an example of the derivation of the twin symmetry in the microscopic description.

The treatment of twin (wall) symmetry based on the concept of domain pairs and sectional layer groups of these pairs (Janovec, 1981[link]; Zikmund, 1984[link]) is analogous to the procedure used in treating interfaces in bicrystals (see Section 3.2.2[link] ; Pond & Bollmann, 1979[link]; Pond & Vlachavas, 1983[link]; Kalonji, 1985[link]; Sutton & Balluffi, 1995[link]). There is the following correspondence between terms: domain pair [\rightarrow] dichromatic complex; domain wall [\rightarrow] interface; domain twin with zero-thickness domain wall [\rightarrow] ideal bicrystal; domain twin with finite-thickness domain wall [\rightarrow] real (relaxed) bicrystal. Terms used in bicrystallography cover more general situations than domain structures (e.g. grain boundaries of crystals with non-crystallographic relations, phase interfaces). On the other hand, the existence of a high-symmetry phase, which is missing in bicrystallography, enables a more detailed discussion of crystallographically equivalent variants (orbits) of various objects in domain structures.

The symmetry group [{\sf T}_{1j}] is the stabilizer of a domain twin (wall) in a certain group, and as such determines a class (orbit) of domain twins (walls) that are crystallographically equivalent with this twin (wall). The number of crystallographically equivalent twins is equal to the number of left cosets (index) of [{\sf T}_{1j}] in the corresponding group. Thus the number [n_{W(p)}] of equivalent domain twins (walls) with the same orientation defined by a plane p of the wall is [n_{W(p)}=[\overline{G(p)}:{\sf T}_{1j}]=|\overline{G(p)}|:|{\sf T}_{1j}|, \eqno(3.4.4.20) ]where [\overline{G(p)}] is a sectional layer group of the plane p in the parent group G, [[\overline{G(p)}:{\sf T}_{1j}] ] is the index of [{\sf T}_{1j}] in [\overline{G(p)}] and absolute value denotes the number of operations in a group.

The set of all domain walls (twins) crystallographically equivalent in G with a given wall [[{\bf S}_{1}|{\bf n}|{\bf S}_{j}]] forms a G-orbit of walls, [G{\sf W}_{1j}\equiv G[{\bf S}_{1}|{\bf n}|{\bf S}_{j}] ]. The number [n_W] of walls in this G-orbit is [\eqalignno{n_{W} &= [G:{\sf T}_{1j}] =|G|:|{\sf T}_{1j}| = (|G|:|\overline{G(p)}|)(|\overline{G(p)}|:|{\sf T}_{1j}|)\cr &=n_pn_{W(p)}, &(3.4.4.21)\cr} ]where [n_p] is the number of planes equivalent with plane p expressed by equation (3.4.4.9[link]) and [n_{W(p)} ] is the number of equivalent domain walls with the plane p [see equation (3.4.4.20[link])]. Walls in one orbit have the same scalar properties (e.g. energy) and their structure and tensor properties are related by operations that relate walls from the same orbit.

Another aspect that characterizes twins and domain walls is the relation between a twin and the reversed twin. A twin (wall) which is crystallographically equivalent with the reversed twin (wall) will be called a reversible twin (wall). If a twin and the reversed twin are not crystallographically equivalent, the twin will be called an irreversible twin (wall). If a domain wall is reversible, then the properties of the reversed wall are fully specified by the properties of the initial wall, for example, these two walls have the same energy and their structures and properties are mutually related by a crystallographic operation. For irreversible walls, no relation exists between a wall and the reversed wall. Common examples of irreversible walls are electrically charged ferroelectric walls (walls carrying a nonzero polarization charge) and domain walls or discommensurations in phases with incommensurate structures.

A necessary and sufficient condition for reversibility is the existence of side-reversing and/or state-exchanging operations in the sectional layer group [\overline{\sf J}_{1j}] of the unordered domain pair [\{{\bf S}_1,{\bf S}_j\} ] [see equation (3.4.4.13[link])]. This group also contains the symmetry group [{\sf T}_{1j}] of the twin [see equation (3.4.4.15[link])] and thus provides a full symmetry characteristic of twins and walls, [ \overline{\sf J}_{1j} ={\sf T}_{1j} \cup \underline{s}_{1j}\widehat{\sf F}_{1j} \cup \underline{t}^{\star}_{1j}\widehat{\sf F}_{1j}. \eqno(3.4.4.22) ]

Sequences of walls and reversed walls appear in simple lamellar domain structures which are formed by domains with two alternating domain states, say [{\bf S}_1] and [{\bf S}_2], and parallel walls [{\sf W}_{12} ] and reversed walls [{\sf W}_{21}] (see Fig. 3.4.2.1[link]).

The distinction `symmetric–asymmetric' and `reversible–irreversible' provides a natural classification of domain walls and simple twins. Five prototypes of domain twins and domain walls, listed in Table 3.4.4.3[link], correspond to five subgroups of the sectional layer group [\overline{\sf J}_{1j}]: the sectional layer group [\overline{\sf J}_{1j} ] itself, the layer group of the twin [{\sf T}_{1j} = \widehat{\sf F}_{1j} \cup \underline{t}^*_{1j}\widehat{\sf F}_{1j}, ] the sectional layer group [\overline{\sf F}_{1j} = \widehat{\sf F}_{1j} \cup \underline{s}_{1j}\widehat{\sf F}_{1j}, ] the trivial layer group [\widehat{\sf J}_{1j} = \widehat{\sf F}_{1j} \cup r^*_{1j}\widehat{\sf F}_{1j} ] and the trivial layer group [\widehat {\sf F}_{1j}].

Table 3.4.4.3 | top | pdf |
Classification of domain walls and simple twins

[{\sf T}_{1j} ] [\overline{\sf J}_{1j} ] Classification Symbol
[\widehat{\sf F}_{1j} \cup \underline{t}^*_{1j}\widehat{\sf F}_{1j} ] [\widehat{\sf F}_{1j} \cup \underline{t}^*_{1j}\widehat{\sf F}_{1j} \cup r^*_{1j}\widehat{\sf F}_{1j} \cup \underline{s}_{1j}\widehat{\sf F}_{1j} ] Symmetric reversible SR
[\widehat{\sf F}_{1j} \cup \underline{t}^*_{1j}\widehat{\sf F}_{1j} ] [\widehat{\sf F}_{1j} \cup \underline t^*_{1j}{\widehat {\sf F}}_{1j} ] Symmetric irreversible SI
[\widehat{\sf F}_{1j}] [\widehat{\sf F}_{1j} \cup \underline{s}_{1j}\widehat{\sf F}_{1j} ] Asymmetric side-reversible [{\rm A}\underline {\rm R}]
[\widehat {\sf F}_{1j} ] [\widehat{\sf F}_{1j} \cup r^*_{1j}\widehat{\sf F}_{1j} ] Asymmetric state-reversible [{\rm AR}^*]
[\widehat {\sf F}_{1j} ] [\widehat{\sf F}_{1j} ] Asymmetric irreversible AI

An example of a symmetric reversible (SR) twin (and wall) is the twin [({\bf S}_1[010]{\bf S}_2)] in Fig. 3.4.4.2[link] with a non-trivial twinning operation [\underline{2}_z^{\star}] and with reversing operations [\underline{m}_y] and [m_x^{\star}]. The twin [({\bf S}_1^{+}[\bar110]{\bf S}_3^{-})] and reversed twin [({\bf S}_3^{-}[\bar110]{\bf S}_1^{+})] in Fig. 3.4.3.8[link] are symmetric and irreversible (SI) twins with a twinning operation [\underline{m}_{{\bar x}y}^{\star} ]; no reversing operations exist (walls are charged and charged walls are always irreversible, since a charge is invariant with respect to any transformation of the space). The twin [({\bf S}_1^{-}[110]{\bf S}_3^{+})] and reversed twin [({\bf S}_3^{+}[110]{\bf S}_1^{-})] in the same figure are asymmetric state-reversible twins with state-reversing operation [m_{{\bar x}y}^{\star} ] and with no non-trivial twinning operation.

The same classification also applies to domain twins and walls in a microscopic description.

As in the preceding section, we shall now present separately the symmetries of non-ferroelastic simple domain twins [`twinning without a change of crystal shape (or form)'] and of ferroelastic simple domain twins [`twinning with a change of crystal shape (or form)'; Klassen-Neklyudova (1964[link]), Indenbom (1982[link])].

3.4.4.4. Non-ferroelastic domain twins and domain walls

| top | pdf |

Compatibility conditions impose no restriction on the orientation of non-ferroelastic domain walls. Any of the non-ferro­elastic domain pairs listed in Table 3.4.3.4[link] can be sectioned on any crystallographic plane p and the sectional group [\overline{\sf J}_{1j}] specifies the symmetry properties of the corresponding twin and domain wall. The analysis can be confined to one representative orientation of each class of equivalent planes, but a listing of all possible cases is too voluminous for the present article. We give, therefore, in Table 3.4.4.4[link] only possible symmetries [{\sf T}_{1j}] and [\overline{\sf J}_{1j}] of non-ferroelastic domain twins and walls, together with their classification, without specifying the orientation of the wall plane p.

Table 3.4.4.4 | top | pdf |
Symmetries of non-ferroelastic domain twins and walls

[{\sf T}_{1j}] [{\sf \overline{J}}_{1j} ] Classification
1 1 AI
1 [\underline{\bar1}] [{\rm A\underline{R}}]
[\underline2 ] [{\rm A\underline{R}}]
[2^{\star} ] [{\rm AR}^{\star}]
[m^{\star}] [{\rm AR}^{\star}]
[{\bar1}^\star] [{\bar1}^\star ] SI
[\underline2/m^{\star} ] SR
[2^{\star}/\underline{m} ] SR
[2 ] [2m^{\star}m^{\star} ] [{\rm AR}^{\star}]
[\underline2^{\star} ] [\underline2^{\star} ] SI
[\underline2^{\star}/m^{\star} ] SR
[2^{\star}\underline2^{\star}\underline2 ] SR
[\underline2^{\star}\underline{m}m^{\star} ] SR
m m AI
[\underline2/m ] [{\rm A\underline{R}}]
[2^{\star}mm^{\star} ] [{\rm AR}^{\star}]
[\underline{m}^{\star} ] [\underline{m}^{\star} ] SI
[2^{\star}/\underline{m}^{\star} ] SR
[\underline{m}^{\star}m^{\star}\underline2 ] SR
[\underline2^{\star}/m ] [\underline2^{\star}/m ] SI
[m\underline{m}m^{\star} ] SR
[2/\underline{m}^{\star} ] [2/\underline{m}^{\star} ] SI
[m^{\star}m^{\star}\underline{m}^{\star} ] SR
[4^{\star}/\underline{m}^{\star} ] SR
[2\underline2^{\star}\underline2^{\star} ] [2\underline2^{\star}\underline2^{\star} ] SI
[\underline{m}m^{\star}m^{\star} ] SR
[4^\star {\underline 2} {\underline 2}^\star ] SR
[\underline{\bar4}\underline2^{\star}m^{\star} ] SR
[\underline{m}^{\star}m\underline2^{\star} ] [\underline{m}^{\star}m\underline2^{\star} ] SI
[\underline{m}^{\star}mm^{\star}] SR
[mm\underline{m}^{\star} ] [mm\underline{m}^{\star} ] SI
[4^{\star}/\underline{m}^{\star}m^{\star}m ] SR
[4 ] [4m^{\star}m^{\star}] [{\rm AR}^{\star}]
[\underline{\bar4}^{\star} ] [\underline{\bar4}^{\star}\underline2m^{\star} ] SR
[4/\underline{m}^{\star} ] [4/\underline{m}^{\star} ] SI
[4/\underline{m}^{\star}m^{\star}m^{\star} ] SR
[4\underline2^{\star}\underline2^{\star} ] [4\underline2^{\star}\underline2^{\star} ] SI
[4/\underline{m}m^{\star}m^{\star} ] SR
[\underline{\bar4}^{\star}\underline2^{\star}m ] [4^{\star}/\underline{m}m^{\star}m] SR
[4/\underline{m}^{\star}mm ] [4/\underline{m}^{\star}mm ] SI
[3 ] [3m^{\star} ] [{\rm AR}^{\star}]
[6^{\star} ] [{\rm AR}^{\star}]
[\underline{\bar3}^{\star} ] [\underline{\bar3}^{\star} ] SI
[\underline{\bar3}^{\star} m^{\star} ] SR
[6^{\star}/\underline{m} ] SR
[3m ] [6^{\star}mm^{\star} ] [{\rm AR}^{\star}]
[3\underline2^{\star} ] [3\underline2^{\star} ] SI
[\underline{\bar3}m^{\star} ] SR
[6^{\star}\underline2\underline2^{\star} ] SR
[\underline{\bar6}\underline2^{\star}m^{\star} ] SR
[\underline{\bar3}^{\star}m ] [\underline{\bar3}^{\star}m ] SI
[6^{\star}/\underline{m}mm^{\star} ] SR
[6 ] [6m^{\star}m^{\star} ] [{\rm AR}^{\star}]
[\underline{\bar6}^{\star} ] [\underline{\bar6}^{\star} ] SI
[6^{\star}/\underline{m}^{\star} ] SR
[\underline{\bar6}^{\star}\underline2m^{\star} ] SR
[6/\underline{m}^{\star} ] [6/\underline{m}^{\star} ] SI
[6/\underline{m}^{\star}m^{\star}m^{\star} ] SR
[6\underline2^{\star}\underline2^{\star} ] [6\underline2^{\star}\underline2^{\star} ] SI
[6/\underline{m}m^{\star}m^{\star} ] SR
[\underline{\bar6}^{\star}m\underline2^{\star} ] [\underline{\bar6}^{\star}m\underline2^{\star} ] SI
[6^{\star}/\underline{m}^{\star}mm^{\star} ] SR
[6/\underline{m}^{\star}mm ] [6/\underline{m}^{\star}mm ] SI

Non-ferroelastic domain walls are usually curved with a slight preference for certain orientations (see Figs. 3.4.1.5[link] and 3.4.3.3[link]). Such shapes indicate a weak anisotropy of the wall energy [\sigma], i.e. small changes of [\sigma] with the orientation of the wall. The situation is different in ferroelectric domain structures, where charged domain walls have higher energies than uncharged ones.

A small energetic anisotropy of non-ferroelastic domain walls is utilized in producing tailored domain structures (Newnham et al., 1975[link]). A required domain pattern in a non-ferroelastic ferroelectric crystal can be obtained by evaporating electrodes of a desired shape (e.g. stripes) onto a single-domain plate cut perpendicular to the spontaneous polarization [{\bf P}_0]. Subsequent poling by an electric field switches only regions below the electrodes and thus produces the desired antiparallel domain structure.

Periodically poled ferroelectric domain structures fabricated by this technique are used for example in quasi-phase-matching optical multipliers (see e.g. Shur et al., 1999[link], 2001[link]; Rosenman et al., 1998[link]). An example of such an engineered domain structure is presented in Fig. 3.4.4.3[link].

[Figure 3.4.4.3]

Figure 3.4.4.3 | top | pdf |

Engineered periodic non-ferroelastic ferroelectric stripe domain structure within a lithium tantalate crystal with symmetry descent [\bar 6\supset 3 ]. The domain structure is revealed by etching and observed in an optical microscope (Shur et al., 2001[link]). Courtesy of Vl. Shur, Ural State University, Ekaterinburg.

Anisotropic domain walls can also appear if the Landau free energy contains a so-called Lifshitz invariant (see Section 3.1.3.3[link] ), which lowers the energy of walls with certain orientations and can be responsible for the appearance of an incommensurate phase (see e.g. Dolino, 1985[link]; Tolédano & Tolédano, 1987[link]; Tolédano & Dmitriev, 1996[link]; Strukov & Levanyuk, 1998[link]). The irreversible character of domain walls in a commensurate phase of crystals also containing (at least theoretically) an incommensurate phase has been confirmed in the frame of phenomenological theory by Ishibashi (1992[link]). The incommensurate structure in quartz that demonstrates such an anisotropy is discussed at the end of the next example.

Example 3.4.4.2. Domain walls in [\alpha]-phase of quartz.  Quartz (SiO2) undergoes a structural phase transition from the parent [\beta] phase (symmetry group [6_{z}2_x2_y]) to the ferroic [\alpha] phase (symmetry [3_{z}2_x]). The [\alpha ] phase can appear in two domain states [{\bf S}_1] and [{\bf S}_2 ], which have the same symmetry [F_1 =F_2 =3_z2_x]. The symmetry [J_{12}] of the unordered domain pair [\{{\bf S}_1,{\bf S}_2\}] is given by [J_{12}^{\star}=] [3_z2_x \cup 2_y^*\{3_z2_x\} =] [6_z^*2_x2_y^*].

Table 3.4.4.5[link] summarizes the results of the symmetry analysis of domain walls (twins). Each row of the table contains data for one representative domain wall [{\bf W}_{12}({\bf n}_{12})] from one orbit [G{\bf W}_{12}({\bf n}_{12})]. The first column of the table specifies the normal [{\bf n}] of the wall plane p, further columns list the layer groups [\widehat{\sf F}_{12} ], [{\sf T}_{12} ] and [{\overline{\sf J}}_{12}] that describe the symmetry properties and classification of the wall (defined in Table 3.4.4.3[link]), and [n_W] is the number of symmetrically equivalent domain walls [cf. equation (3.4.4.21[link])].

Table 3.4.4.5 | top | pdf |
Symmetry properties of domain walls in quartz

[|{\bf P}| \not=|{\bf P^\prime}|, \ P^\prime_{\alpha} \not= -P_{\alpha}, \ \alpha = x,y,z ].

[{\bf n}] [\widehat{\sf F}_{12} ] [{\sf T}_{12} ] [\overline{\sf J}_{12} ] Classification [n_W] [{\bf P}({\bf W}_{12}) ] [{\bf P}({\bf W}_{21}) ]
[[001] ] [3_z ] [3_z\underline{2}^*_{y} ] [6^*_z\underline{2}_{x}\underline{2}^*_{y} ] SR 2    
[[100] ] [2_{x} ] [2_{x}\underline{2}^*_{y}\underline{2}^*_z ] [2_{x}\underline{2}^*_{y}\underline{2}^*_z ] SI 3    
[[010] ] 1 [\underline{2}^*_z ] [\underline{2}_{x}2^*_{y}\underline{2}^*_z ] SR 6 [0,0,P_z ] [0,0,-P_z ]
[[0vw] ] 1 1 [\underline{2}_x ] [{\rm A}\underline{\rm R}] 12 [P_x,P_y,P_z ] [P_x,-P_y,-P_z ]
[[u0w] ] 1 [\underline{2}^*_{y} ] [\underline{2}^*_{y} ] SI 6 [0,P_y,0 ] [0,-P^\prime_y,0 ]
[[uv0] ] 1 [\underline{2}^*_z ] [\underline{2}^*_z ] SI 6 [0,0,P_z ] [0,0,P^\prime_z ]
[[uvw] ] 1 1 1 AI 12 [P_x,P_y,P_z ] [P^\prime_{x},P^\prime_{y},P^\prime_z ]

The last two columns give possible components of the spontaneous polarization P of the wall [{\bf W}_{12}({\bf n})] and the reversed wall [{\bf W}_{21}({\bf n})]. Except for walls with normals [001] and [100], all walls are polar, i.e. they can be spontaneously polarized. The reversal of the polarization in reversible domain walls requires the reversal of domain states. In irreversible domain walls, the reversal of [{\bf W}_{12} ] into [{\bf W}_{21}] is accompanied by a change of the polarization P into P′, which may have a different absolute value and direction different to that of P.

The structure of two domain states and two mutually reversed domain walls obtained by molecular dynamics calculations are depicted in Fig. 3.4.4.4[link] (Calleja et al., 2001[link]). This shows a projection on the ab plane of the structure represented by SiO4 tetrahedra, in which each tetrahedron shares four corners. The threefold symmetry axes in the centres of distorted hexagonal channels and three twofold symmetry axes (one with vertical orientation) perpendicular to the threefold axes can be easily seen. The two vertical dotted lines are the wall planes p of two mutually reversed walls [[{\bf S}_{1}[010]{\bf S}_{2}]=] [{\bf W}_{12}[010]] and [[{\bf S}_{2}[010]{\bf S}_{1}] =] [{\bf W}_{21}[010]]. In Table 3.4.4.5[link] we find that these walls have the symmetry [{\sf T}_{12}[010] =] [{\sf T}_{21}[010]=] [\underline{2}_x2_y^{\star}\underline{2}_z^{\star}], and in Fig. 3.4.4.4[link] we can verify that the operation [\underline{2}_x] is a `side-reversing' operation [\underline{s}_{12}] of the wall (and the whole twin as wall), operation [2_y^{\star}] is a `state-exchanging operation' [r_{12}^{\star}] and the operation [\underline{2}_z^{\star} ] is a non-trivial `side-and-state reversing' operation [\underline{t}_{12}^{\star} ] of the wall. The walls [{\bf W}_{12}[010]] and [{\bf W}_{21}[010] ] are, therefore, symmetric and reversible walls.

[Figure 3.4.4.4]

Figure 3.4.4.4 | top | pdf |

Microscopic structure of two domain states and two parallel mutually reversed domain walls in the [\alpha] phase of quartz. The left-hand vertical dotted line represents the domain wall [{\sf W}_{12}], the right-hand line is the reversed domain wall [{\sf W}_{12}]. To the left of the left-hand line and to the right of the right-hand line are domains with domain state [{\bf S}_1], the domain between the lines has domain state [{\bf S}_2]. For more details see text. Courtesy of M. Calleja, University of Cambridge.

During a small temperature interval above the appearance of the [\alpha ] phase at 846 K, there exists an incommensurate phase that can be treated as a regular domain structure, consisting of triangular columnar domains with domain walls (discommensurations) of negative wall energy [\sigma] (see e.g. Dolino, 1985[link]). Both theoretical considerations and electron microscopy observations (see e.g. Van Landuyt et al., 1985[link]) show that the wall normal has the [[uv0]] direction. From Table 3.4.4.5[link] it follows that there are six equivalent walls that are symmetric but irreversible, therefore any two equivalent walls differ in orientation.

This prediction is confirmed by electron microscopy in Fig. 3.4.4.5[link], where black and white triangles correspond to domains with domain states [{\bf S}_1] and [{\bf S}_2], and the transition regions between black and white areas to domain walls (discommensurations). To a domain wall of a certain orientation no reversible wall appears with the same orientation but with a reversed order of black and white. Domain walls in homogeneous triangular parts of the structure are related by 120 and 240° rotations and carry, therefore, parallel spontaneous polarizations; wall orientations in two differently oriented blocks (the middle of the right-hand part and the rest on the left-hand side) are related by 180° rotations about the axis [2_x ] in the plane of the photograph and are, therefore, polarized in antiparallel directions (for more details see Saint-Grégoire & Janovec, 1989[link]; Snoeck et al., 1994[link]). After cooling down to room temperature, the wall energy becomes positive and the regular domain texture changes into a coarse domain structure in which these six symmetry-related wall orientations still prevail (Van Landuyt et al., 1985[link]).

[Figure 3.4.4.5]

Figure 3.4.4.5 | top | pdf |

Transmission electron microscopy (TEM) image of the incommensurate triangular ([3-q] modulated) phase of quartz. The black and white triangles correspond to domains with domain states [{\bf S}_1] and [{\bf S}_2 ], and the transition regions between black and white areas to domain walls (discommensurations). For a domain wall of a certain orientation there are no reversed domain walls with the same orientation but reversed order of black and white; the walls are, therefore, non-reversible. Domain walls in regions with regular triangular structures are related by 120 and 240° rotations about the z direction and carry parallel spontaneous polarizations (see text). Triangular structures in two regions (blocks) with different orientations of the triangles are related e.g. by [2_x] and carry, therefore, antiparallel spontaneous polarizations and behave macroscopically as two ferroelectric domains with antiparallel spontaneous polarization. Courtesy of E. Snoeck, CEMES, Toulouse and P. Saint-Grégoire, Université de Toulon.

3.4.4.5. Ferroelastic domain twins and walls. Ferroelastic twin laws

| top | pdf |

As explained in Section 3.4.3.6[link], from a domain pair [({\bf S}_1,{\bf S}_j)] of ferroelastic single-domain states with two perpendicular equally deformed planes p and [p^\prime] one can form four different ferroelastic twins (see Fig. 3.4.3.8[link]). Two mutually reversed twins [({\bf S}_1|{\bf n}|{\bf S}_j)] and [({\bf S}_j|{\bf n}|{\bf S}_1) ] have the same twin symmetry [{\sf T}_{1j}(p)] and the same symmetry [\overline{\sf J}_{1j}(p)] of the twin pair [({\bf S}_1,{\bf S}_j|{\bf n}|{\bf S}_j,{\bf S}_1) ]. The ferroelastic twin laws can be expressed by the layer group [\overline{\sf J}_{1j}(p)] or, in a less complete way (without specification of reversibility), by the twin symmetry [{\sf T}_{1j}(p)]. The same holds for two mutually reversed twins [({\bf S}_1|{\bf n^\prime}|{\bf S}_j) ] and [({\bf S}_j|{\bf n^\prime}|{\bf S}_1)] with a twin plane [p^\prime] perpendicular to p.

Table 3.4.4.6[link] summarizes possible symmetries [{\sf T}_{1j}] of ferro­elastic domain twins and corresponding ferroelastic twin laws [\overline{\sf J}_{1j}]. Letters V and W signify strain-dependent and strain-independent (with a fixed orientation) domain walls, respectively. The classification of domain walls and twins is defined in Table 3.4.4.3[link]. The last column contains twinning groups [K_{1j}[F_1] ] of ordered domain pairs [({\bf S}_1,{\bf S}_j)] from which these twins can be formed. The symbol of [K_{1j}] is followed by a symbol of the group [F_1] given in square brackets. The twinning group [K_{1j}[F_1]] specifies, up to two cases, a class of equivalent domain pairs [orbit [G({\bf S}_1,{\bf S}_j)]] (see Section 3.4.3.4[link]). More details on particular cases (orientation of domain walls, disorientation angle, twin axis) can be found in synoptic Table 3.4.3.6[link]. From this table follow two general conclusions:

  • (1) All layer groups describing the symmetry of compatible ferroelastic domain walls are polar groups, therefore all compatible ferroelastic domain walls in dielectric crystals can be spontaneously polarized. The direction of the spontaneous polarization is parallel to the intersection of the wall plane p and the plane of shear (i.e. a plane perpendicular to the axis of the ferroelastic domain pair, see Fig. 3.4.3.5[link]b and Section 3.4.3.6.2[link]).

    Table 3.4.4.6 | top | pdf |
    Symmetry properties of ferroelastic domain twins and compatible domain walls

    [{\sf T}_{1j}] [{\sf \overline{J}}_{1j} ] Classification [K_{1j}[F_1] ]
    [1] [\underline2 ] V [{\rm A}\underline{\rm R}]   [4^{\star}[2]], [\bar4^{\star}[2]], [6[2]], [6/m[2]]
    [1] [\underline2] V [{\rm A}\underline{\rm R}]  
    [1] [2^{\star}] W [{\rm AR}^{\star}] [\Bigl\{] [2^{\star}[1]], [422[2]], [\bar42m[m]], [32[2]], [\bar3m[m]], [622[2]], [\bar6m2[m],] [432[222]], [m\bar3m[mm2]], [m\bar3m[2^{\star}_{xy}][mm2]]
    [\underline2^{\star}] [\underline2^{\star}] V SI
    [1] [2^{\star}] W [{\rm AR}^{\star}]   [23[3]], [432[4]], [432[3]], [m\bar3m[\bar4]]
    [\underline2^{\star}] [\underline2^{\star}] W SI  
    [1] [m^{\star}] V [{\rm AR}^{\star}] [\Bigl\{] [m^{\star}[1]], [4mm[m]], [\bar42m[2]], [3m[m]], [\bar3m[2]], [6mm[m]], [\bar6m2[2]], [\bar43m[mm2]], [m\bar3m[222]], [m\bar3m[m^{\star}_{xy}][m2m]]
    [{\underline m}^{\star} ] [\underline m^{\star}] W SI
    [1] [m^{\star}] W [{\rm AR}^{\star}]   [m\bar3[3]], [\bar43m[\bar4]], [\bar43m[3]], [m\bar3m[4] ]
    [{\underline m}^{\star} ] [\underline2^{\star}] W SI  
    [\underline2^{\star}] [2^{\star}\underline2^{\star}\underline2 ] W SR [\Bigl\{] [2^{\star}2^{\star}2[2]], [4^{\star}22^{\star}[222]], [\bar4^{\star}2^{\star}m[mm2] ], [622[222]], [6/mmm[mm2]], [432[422]], [432[32] ], [m\bar3m[\bar42m]]
    [\underline2^{\star}] [\underline2^{\star}2^{\star}\underline2 ] W SR
    [\underline2^{\star}] [\underline2^{\star}/m^{\star} ] V SR   [2^{\star}/m^{\star}[\bar1]], [4/mmm[2/m]], [\bar3m[2/m]], [6/mmm[2/m]], [m\bar3m[mmm]]
    [\underline{m}^{\star} ] [2^{\star}/\underline{m}^{\star} ] W SR  
    [\underline2^{\star}] [\underline2^{\star}/m^{\star} ] W SR   [m\bar3[\bar3]], [m\bar3m[4/m]], [m\bar3m[\bar3]]
    [\underline{m}^{\star} ] [2^{\star}/\underline{m}^{\star} ] W SR  
    [m ] m V AI   [4/m[m]], [\bar6[m]], [6/m[m]]
    [m ] m V AI  
    m [\underline{2}/m ] V [{\rm A}\underline{\rm R} ]   [4^{\star}/m[2/m]], [6/m[2/m] ]
    [m ] [\underline{2}/m ] V [{\rm A}\underline{\rm R} ]  
    [\underline{m}^{\star} ] [{m}^{\star}\underline{m}^{\star}\underline2 ] W SR [\Bigl\{] [m^{\star}m^{\star}2[2]], [4^{\star}mm^{\star}[mm2]], [\bar4^{\star}2m^{\star}[222] ], [6mm[mm2]], [6/mmm[222]], [\bar43m[\bar42m]], [m\bar3m[422]], [m\bar3m[32]]
    [\underline{m}^{\star} ] [\underline{m}^{\star}m^{\star}\underline2 ] W SR
    [m ] [m^{\star}2^{\star}m ] W [{\rm AR}^{\star}] [\Bigl\{] [m^{\star}2^{\star}m[m]], [4/mmm[2mm]], [\bar6m2[m2m]], [6/mmm [m2m]], [\bar43m[3m]], [m\bar3m[4mm]], [m\bar3m[\bar42m] ], [m\bar3m[3m]]
    [\underline{m}^{\star}\underline2^{\star}m ] [\underline{m}^{\star}\underline2^{\star}m ] W SI
    [\underline{m}^{\star}\underline2^{\star}m ] [\underline{m}^{\star}m^{\star}m ] W SR [\Bigl\{] [m^{\star}m^{\star}m[2/m] ], [4^{\star}/mmm^{\star}[mmm]], [6/mmm[mmm]], [m\bar3m[4/mmm] ], [m\bar3m[\bar3m]]
    [\underline{2}^{\star}\underline{m}^{\star}m ] [m^{\star}\underline{m}^{\star}m ] W SR
  • (2) Domain twin [({\bf S}_1|{\bf n}|{\bf S}_j) ] formed in the parent clamping approximation from a single-domain pair [({\bf S}_1,{\bf S}_j)] and the relaxed domain twin [({{\bf S}_1^+}|{\bf n}|{{\bf S}_j^-})] with disoriented domain states have the same symmetry groups [{\sf T}_{1j}] and [{\sf \overline{J}}_{1j} ].

This follows from simple reasoning: all twin symmetries [{\sf {T}}_{1j} ] in Table 3.4.4.6[link] have been derived in the parent clamping approximation and are expressed by the orthorhombic group [mm2] or by some of its subgroups. As shown in Section 3.4.3.6.2[link], the maximal symmetry of a mechanically twinned crystal is also [mm2]. An additional simple shear accompanying the lifting of the parent clamping approximation cannot, therefore, decrease the symmetry [{{\sf T}_{1j}}(p) ] derived in the parent clamping approximation. In a similar way, one can prove the statement for the group [{\sf\overline{J}}_{1j}(p)] of the twin pairs [({{\bf S}_1},{\bf S}_j|{\bf n}|{\bf S}_j,{{\bf S}_1})] and [({{\bf S}_1^+},] [{\bf S}_j^{-}|{\bf n}|{\bf S}_j^{-},] [{{\bf S}_1^+})].

3.4.4.6. Domain walls of finite thickness – continuous description

| top | pdf |

A domain wall of zero thickness is a geometrical construct that enabled us to form a twin from a domain pair and to find a layer group that specifies the maximal symmetry of that twin. However, real domain walls have a finite, though small, thickness. Spatial changes of the structure within a wall may, or may not, lower the wall symmetry and can be conveniently described by a phenomenological theory.

We shall consider the simplest case of a one nonzero component [\eta ] of the order parameter (see Section 3.1.2[link] ). Two nonzero equilibrium homogeneous values of [-\eta_{0}] and [+\eta_{0}] of this parameter correspond to two domain states [{\bf S}_1] and [{\bf S}_2]. Spatial changes of the order parameter in a domain twin [({\bf S}_1|{\bf n}|{\bf S}_2) ] with a zero-thickness domain wall are described by a step-like function [\eta(\xi)=-\eta_0] for [\xi \,\lt\, 0] and [\eta(\xi)=+\eta_0 ] for [\xi\,\gt\,0], where [\xi] is the distance from the wall of zero thickness placed at [\xi=0].

A domain wall of finite thickness is described by a function [\eta(\xi) ] with limiting values [-\eta_{0}] and [\eta_{0} ]: [\lim_{\xi \rightarrow -\infty} \eta(\xi) = -\eta_{0}, \quad \lim_{\xi \rightarrow +\infty} \eta(\xi) = \eta_{0}.\eqno(3.4.4.23) ]If the wall is symmetric, then the profile [\eta(\xi)] in one half-space, say [\xi \,\lt\, 0], determines the profile in the other half-space [\xi\,\gt\,0]. For continuous [\eta(\xi)] fulfilling conditions (3.4.4.23[link]) this leads to the condition [\eta(\xi)=-\eta(-\xi), \eqno(3.4.4.24)]i.e. [\eta(\xi)] must be an odd function. This requirement is fulfilled if there exists a non-trivial symmetry operation of a domain wall (twin): a side reversal [(\xi \rightarrow -\xi)] combined with an exchange of domain states [[\eta(\xi) \rightarrow -\eta(\xi)]] results in an identical wall profile.

A particular form of the wall profile [\eta(\xi)] can be deduced from Landau theory. In the simplest case, the dependence [\eta(\xi)] of the domain wall would minimize the free energy [\int^{\infty}_{-\infty} \left(\Phi_{0} + \textstyle{{1}\over{2}}\alpha(T-T_c)\eta^{2} + \textstyle{{1}\over{4}}\beta\eta^4 + \textstyle{{1}\over{2}}\delta\left({{d^{2}\eta}\over{d\xi^{2}}}\right)^{2}\right)\,{\rm d}\xi, \eqno(3.4.4.25) ]where [\alpha,] [\beta,] [\delta] are phenomenological coefficients and T and Tc are the temperature and the temperature of the phase transition, respectively. The first three terms correspond to the homogeneous part of the Landau free energy (see Section 3.2.1[link] ) and the last term expresses the energy of the spatially changing order parameter. This variational task with boundary conditions (3.4.4.23[link]) has the following solution (see e.g. Salje, 1990[link], 2000[link]b; Ishibashi, 1990[link]; Strukov & Levanyuk, 1998[link]) [\eta(\xi) = \eta_{0}\tanh({\xi}/{w}), \eqno(3.4.4.26)]where the value w specifies one half of the effective thickness [2w] of the domain wall and is given by [w = \sqrt{2\delta/\alpha(T_c-T)}. \eqno(3.4.4.27)]This dependence, expressed in relative dimensionless variables [\xi/w] and [\eta/\eta_{0}], is displayed in Fig. 3.4.4.6[link].

[Figure 3.4.4.6]

Figure 3.4.4.6 | top | pdf |

Profile of the one-component order parameter [\eta(\xi)] in a symmetric wall (S). The effective thickness of the wall is [2w].

The wall profile [\eta(\xi)] expressed by solution (3.4.4.26[link]) is an odd function of [\xi], [\eta(-\xi)=\eta_{0}\tanh({-\xi}/w)=-\eta_{0}\tanh({\xi}/w)=-\eta(\xi), \eqno(3.4.4.28) ]and fulfils thus the condition (3.4.4.24[link]) of a symmetric wall.

The wall thickness can be estimated from electron microscopy observations, or more precisely by a diffuse X-ray scattering technique (Locherer et al., 1998[link]). The effective thickness [2w] [see equation (3.4.4.26[link])] in units of crystallographic repetition length A normal to the wall ranges from [2w/A=2] to [2w/A=12], i.e. [2w] is about 10–100 nm (Salje, 2000[link]b). The temperature dependence of the domain wall thickness expressed by equation (3.4.4.27[link]) has been experimentally verified, e.g. on LaAlO3 (Chrosch & Salje, 1999[link]).

The energy [\sigma] of the domain wall per unit area equals the difference between the energy of the twin and the energy of the single-domain crystal. For a one nonzero component order parameter with the profile (3.4.4.26[link]), the wall energy [\sigma] is given by (Strukov & Levanyuk, 1998[link]) [\sigma = \int^{\infty}_{-\infty} \left[\Phi(\eta(\xi)) - \Phi(\eta_{0})\right]\,{\rm d}\xi = {{2\sqrt{2\delta}}\over{3\beta}}[\alpha(T_c-T)]^{3/2}, \eqno(3.4.4.29) ]where [2w] is the effective thickness of the wall [see equation (3.4.4.27[link])] and the coefficients are defined in equation (3.4.4.25[link]).

The order of magnitude of the wall energy [\sigma] of ferroelastic and non-ferroelastic domain walls is typically several millijoule per square metre (Salje, 2000[link]b).

Example 3.4.4.3.  In our example of a ferroelectric phase transition [4_z/m_zm_xm_{xy}\supset 2_xm_ym_z ], one can identify [\eta] with the [P_1] component of spontaneous polarization and [\xi] with the axis y. One can verify in Fig. 3.4.4.6[link] that the symmetry [{\sf T}_{12}[010]=\underline{2}_z^{\star}/m_z ] of the twin [({\bf S}_1[010]{\bf S}_2)] with a zero-thickness domain wall is retained in the domain wall with symmetric profile (3.4.4.26[link]): both non-trivial symmetry operations [\underline{2}_z^{\star} ] and [\underline{\bar1}^{\star}] transform the profile [\eta(y) ] into an identical function.

This example illustrates another feature of a symmetric wall: All non-trivial symmetry operations of the wall are located at the central plane [\xi=0 ] of the finite-thickness wall. The sectional group [{\sf T}_{12}] of this plane thus expresses the symmetry of the central layer and also the global symmetry of a symmetric wall (twin). The local symmetry of the off-centre planes [\xi\neq 0] is equal to the face group [\widehat{\sf F}_{12}] of the the layer group [{\sf T}_{12} ] (in our example [\widehat{\sf F}_{12}=\{1,m_z\}]).

The relation between a wall profile [\eta(\xi)] of a symmetric reversible (SR) wall and the profile [{\eta}^{\rm rev}(\xi) ] of the reversed wall is illustrated in Fig. 3.4.4.7[link], where the the dotted curve is the wall profile [{\eta}^{\rm rev}(\xi)] of the reversed wall. The profile [{\eta}^{\rm rev}(\xi)] of the reversed wall is completely determined by the the profile [{\eta}(\xi)] of the initial wall, since both profiles are related by equations [{\eta}^{\rm rev}(\xi)=-{\eta}(\xi)={\eta}(-\xi).\eqno(3.4.4.30) ]The first part of the equation corresponds to a state-exchanging operation [r_{12}^{\star}] (cf. point [r^{\star}{\rm A}] in Fig. 3.4.4.7[link]) and the second one to a side-reversing operation [\underline{s}_{12} ] (point [\underline{s}{\rm A}] in the same figure). In a symmetric reversible wall, both types of reversing operations exist (see Table 3.4.4.3[link]).

[Figure 3.4.4.7]

Figure 3.4.4.7 | top | pdf |

Profiles of the one-component order parameter [\eta(\xi)] in a symmetric wall (solid curve) and in the reversed wall (dotted curve). The wall is symmetric and reversible (SR).

In a symmetric irreversible (SI) wall both initial and reversed wall profiles fulfil symmetry condition (3.4.4.24[link]) but equations (3.4.4.30[link]) relating both profiles do not exist. The profiles [{\eta}(\xi)] and [{\eta}^{\rm rev}(\xi)] may differ in shape and surface wall energy. Charged domain walls are always irreversible.

A possible profile of an asymmetric domain wall is depicted in Fig. 3.4.4.8[link] (full curve). There is no relation between the negative part [{\eta}({\xi}) \,\lt\, 0] and positive part [{\eta}({\xi})\,\gt\,0] of the wall profile [{\eta}({\xi})]. Owing to the absence of non-trivial twin operations, there is no central plane with higher symmetry. The local symmetry (sectional layer group) at any location [\xi] within the wall is equal to the face group [\widehat{\sf F}_{12} ]. This is also the global symmetry [{\sf T}_{12}] of the entire wall, [{\sf T}_{12}=\widehat{\sf F}_{12}].

[Figure 3.4.4.8]

Figure 3.4.4.8 | top | pdf |

Profiles of the one-component order parameter [\eta(\xi)] in an asymmetric wall (solid curve) and in the reversed asymmetric wall (dotted curve). The wall is asymmetric and state-reversible ([{\rm AR}^{\star}]).

The dotted curve in Fig. 3.4.4.8[link] represents the reversed-wall profile of an asymmetric state-reversible (AR[^{\star}]) wall that is related to the initial wall by state-exchanging operations [r_{12}^{\star}\widehat{\sf F}_{12}] (see Table 3.4.4.5[link]), [{\eta}^{\rm rev}(\xi)=-{\eta}(\xi).\eqno(3.4.4.31) ]

An example of an asymmetric side-reversible (A[\underline {\rm R} ]) wall is shown in Fig. 3.4.4.9[link]. In this case, an asymmetric wall (full curve) and reversed wall (dotted curve) are related by side-reversing operations [\underline{s}_{12}\widehat{\sf F}_{12} ]: [{\eta}^{\rm rev}(\xi)={\eta}(-\xi). \eqno(3.4.4.32) ]

[Figure 3.4.4.9]

Figure 3.4.4.9 | top | pdf |

Profiles of the one-component order parameter [\eta(\xi)] in an asymmetric wall (solid curve) and in the reversed asymmetric wall (dotted curve). The wall is asymmetric and side-reversible (A[\underline{\rm R} ]).

In an asymmetric irreversible (AI) wall, both profiles [{\eta}(\xi)] and [{\eta}^{\rm rev}(\xi)] are asymmetric and there is no relation between these two profiles.

The symmetry [{\sf T}_{12}(\eta)] of a finite-thickness wall with a profile [\eta(\xi)] is equal to or lower than the symmetry [{\sf T}_{12}] of the corresponding zero-thickness domain wall, [{\sf T}_{12}\supseteq {\sf T}_{12}(\eta) ]. A symmetry descent [{\sf T}_{12}\supset {\sf T}_{12}(\eta)] can be treated as a phase transition in the domain wall (see e.g. Bul'bich & Gufan, 1989a[link],b[link]; Sonin & Tagancev, 1989[link]). There are [n_{W(\eta)}] equivalent structural variants of the finite-thickness domain wall with the same orientation and the same energy but with different structures of the wall, [n_{W(\eta)}=[{\sf T}_{12}:{\sf T}_{12}(\eta)]=|{\sf T}_{12}|:|{\sf T}_{12}(\eta)|. \eqno(3.4.4.33)]

Domain-wall variants – two-dimensional analogues of domain states – can coexist and meet along line defects – one-dimensional analogues of a domain wall (Tagancev & Sonin, 1989[link]).

Symmetry descent in domain walls of finite thickness may occur if the order parameter [\eta] has more than one nonzero component. We can demonstrate this on ferroic phases with an order parameter with two components [\eta_1 ] and [\eta_2]. The profiles [\eta_1(\xi)] and [\eta_2(\xi) ] can be found, as for a one-component order parameter, from the corresponding Landau free energy (see e.g. Cao & Barsch, 1990[link]; Houchmandzadeh et al., 1991[link]; Ishibashi, 1992[link], 1993[link]; Rychetský & Schranz, 1993[link], 1994[link]; Schranz, 1995[link]; Huang et al., 1997[link]; Strukov & Levanyuk, 1998[link]; Hatt & Hatch, 1999[link]; Hatch & Cao, 1999[link]).

Let us denote by [{\sf T}_{12}(\eta_1)] the symmetry of the profile [\eta_1(\xi)] and by [{\sf T}_{12}(\eta_2)] the symmetry of the profile [\eta_2(\xi)]. Then the symmetry of the entire wall [{\sf T}_{12}(\eta) ] is a common part of the symmetries [{\sf T}_{12}(\eta_1)] and [{\sf T}_{12}(\eta_2)], [{\sf T}_{12}(\eta)={\sf T}_{12}(\eta_1) \cap {\sf T}_{12}(\eta_2). \eqno(3.4.4.34) ]

Example 3.4.4.4.  In our illustrative phase transition [4_z/m_zm_xm_{xy}] [\supset] [2_xm_ym_z], the order parameter has two components [\eta_1, \eta_2] that can be associated with the x and y components [P_1] and [P_2] of the spontaneous polarization (see Table 3.1.3.1[link] and Fig. 3.4.2.2[link]). We have seen that the domain wall [[{\bf S}_1[010]{\bf S}_2]] of zero thickness has the symmetry [{\sf T}_{12}=\underline{2}_z^{\star}/m_z]. If one lets [\eta_1(y)] relax and keeps [\eta_2(y)=0] (a so-called linear structure), then [{\sf T}_{12}(\eta_1)=\underline{2}_z^{\star}/m_z ] (see Fig. 3.4.4.2[link] with [\xi=y]). If the last condition is lifted, a possible profile of a relaxed [\eta_2(y)] is depicted by the full curve in Fig. 3.4.4.10[link]. If both components [\eta_1(y)] and [\eta_2(y)] are nonzero within the wall, one speaks about a rotational structure of domain wall. In this relaxed domain wall the spontaneous polarization rotates in the plane (001), resembling thus a Néel wall in magnetic materials. The even profile [\eta_2(-y)=\eta_2(y)] has the symmetry [{\sf T}_{12}(\eta_2)=m_x^{\star}2_y^{\star}m_z ]. Hence, according to (3.4.4.34[link]), the symmetry of a relaxed wall with a rotational structure is [{\sf T}_{12}(\eta) =] [\underline{2}_z^{\star}/m_z ] [\cap] [m_x^{\star}2_y^{\star}m_z=] [\{1,m_z\}]. This is an asymmetric state-reversible (AR[^{\star}]) wall with two chiral variants [see equation (3.4.4.33[link])] that are related by [\underline{1}^{\star}] and [\underline{2}_z^{\star} ]; the profile [\eta_2(y)] of the second variant is depicted in Fig. 3.4.4.10[link] by a dashed curve.

[Figure 3.4.4.10]

Figure 3.4.4.10 | top | pdf |

A profile of the second order parameter component in a degenerate domain wall.

Similarly, one gets for a zero-thickness domain wall [[{\bf S}_1[001]{\bf S}_2] ] perpendicular to z the symmetry [{\sf T}_{12}=\underline{2}_y^{\star}/m_y ]. For a relaxed domain wall with profiles [\eta_1(z)] and [\eta_2(z)], displayed in Figs. 3.4.4.6[link] and 3.4.4.10[link] with [\xi=z], one gets [{\sf T}_{12}(\eta_1)=\underline{2}_y^{\star}/m_y ], [{\sf T}_{12}(\eta_2)] [=m_x^{\star}\underline{2}_y^{\star}m_z] and [{\sf T}_{12}(\eta)=\{1,\underline{2}_y^{\star}\}]. The relaxed domain wall with rotational structure has lower symmetry than the zero-thickness wall or the wall with linear structure, but remains a symmetric and reversible (SR) domain wall in which spontaneous polarization rotates in a plane (001), resembling thus a Bloch wall in magnetic materials. Two chiral right-handed and left-handed variants are related by operations [m_z] and [\underline{1}^{\star} ]. This example illustrates that the structure of domain walls may differ with the wall orientation.

We note that the stability of a domain wall with a rotational structure and with a linear structure depends on the values of the coefficients in the Landau free energy, on temperature and on external fields. In favourable cases, a phase transition from a symmetric linear structure to a less symmetric rotational structure can occur. Such phase transitions in domain walls have been studied theoretically by Bul'bich & Gufan (1989a[link],b[link]) and by Sonin & Tagancev (1989[link]).

3.4.4.7. Microscopic structure and symmetry of domain walls

| top | pdf |

The thermodynamic theory of domain walls outlined above is efficient in providing quantitative results (wall thickness, energy) in any specific material. However, since this is a continuum theory, it is not able to treat local structural changes on a microscopic level and, moreover, owing to the small thickness of domain walls (several lattice constants), the reliability of its conclusions is to some extent uncertain.

Discrete theories either use simplified models [e.g. pseudospin ANNNI (axial next nearest neighbour Ising) model] that yield quantitative results on profiles, energies and interaction energies of walls but do not consider real crystal structures, or calculate numerically for a certain structure the atomic positions within a wall from interatomic potentials.

Symmetry analysis of domain walls provides useful qualitative conclusions about the microscopic structure of walls. Layer groups with discrete two-dimensional translations impose, via the site symmetries, restrictions on possible displacements and/or ordering of atoms or molecules. From these conclusions, combined with a reasonable assumption that these shifts or ordering vary continuously within a wall, one gets topological constraints on the field of local displacements and/or ordering of atoms or molecules in the wall. The advantage of this treatment is its simplicity and general validity, since no approximations or simplified models are needed. The analysis can also be applied to domain walls of zero thickness, where thermodynamic theory fails. However, this method does not yield any quantitative results, such as values of displacements, wall thickness, energy etc.

The procedure is similar to that in the continuum description. The main relations equations (3.4.4.12)[link]–(3.4.4.17)[link] and the classification given in Table 3.4.4.3[link] hold for a microscopic description as well; one has only to replace point groups by space groups.

A significant difference is that the sectional layer groups and the wall symmetry depend on the location of the plane p in the crystal lattice. This position can by expressed by a vector [{s\bf d}], where d is the scanning vector (see IT E , 2002[link] and the example below) and s is a non-negative number smaller than 1, [0\leq s \,\lt\, 1]. An extended symbol of a twin in the microscopic description, corresponding to the symbol (3.4.4.1[link]) in the continuum description, is [({\sf S}_1|{\bf n};s{\bf d}|{\sf S}_2) \equiv ({\sf S}_2|{\bf -n};s{\bf d}|{\sf S}_1). \eqno(3.4.4.35) ]The main features of the analysis are demonstrated on the following example.

Example 3.4.4.5. Ferroelastic domain wall in calomel.  We examine a ferroelastic compatible domain wall in a calomel crystal (Janovec & Zikmund, 1993[link]; IT E, 2002[link], Chapter 5.2[link] ). In Section 3.4.2.5[link], Example 3.4.2.7[link], we found the microscopic domain states (see Fig. 3.4.2.5[link]) and, in Section 3.4.3.7[link], the corresponding ordered domain pair [({\sf S}_1,{\sf S}_3)] and unordered domain pair [\{{\sf S}_1,{\sf S}_3\}] (depicted in Fig. 3.4.3.10[link]). These pairs have symmetry groups [{\cal F}_{13}=Pn_{x\overline{y}}n_{xy}m_z ] and [{\cal J}_{13}=P4_{2z}^\star /m_zn_{xy}m_x^\star], respectively. Both groups have an orthorhombic basis [{\bf a}^{o}={\bf a}^{t}-{\bf b}^{t}, {\bf b}^{o}={\bf a}^{t}+{\bf b}^{t}, {\bf c}^{o}={\bf c}^{t} ], with a shift of origin [{\bf b}^t /2] for both groups.

Compatible domain walls in this ferroelastic domain pair have orientations (100) and (010) in the tetragonal coordinate system (see Table 3.4.3.6[link]). We shall examine the former case – the latter is crystallographically equivalent. Sectional layer groups of this plane in groups [{\cal F}_{13}] and [{\cal J}_{13}] have a two-dimensional translation group (net) with basic vectors [{\bf a}^{s}=2{\bf b}^{t}] and [{\bf b}^{s}={\bf c}^{t}], and the scanning vector [{\bf d}=2{\bf a}^{t} ] expresses the repetition period of the layer structure (cf. Fig. 3.4.3.10a[link]). From the diagram of symmetry elements of the group [{\cal F}_{13}] and [{\cal J}_{13}], available in IT A (2005[link]), one can deduce the sectional layer groups at any location [s{\bf d}, 0\leq s \,\lt\, 1]. These sectional layer groups are listed explicitly in IT E (2002[link]) in the scanning tables of the respective space groups.

The resulting sectional layer groups [\overline{{\cal F}}_{13}] and [\overline{{\cal J}}_{13}] are given in Table 3.4.4.7[link] in two notations, in which the letter p signifies a two-dimensional net with the basic translations [{\bf a}^{s}, ] [{\bf b}^{s}] introduced above. Standard symbols are related to the basis [{\bf a}^{s}, ] [{\bf b}^{s},] [{\bf c}^{s}=] [{\bf d}]. Subscripts in non-coordinate notation specify the orientation of symmetry elements in the reference Cartesian coordinate system of the tetragonal phase, the partial translation in the glide plane a and in the screw axis [2_1] is equal to [{{1}\over{2}}{\bf a}^{s}={\bf b}^{t}], i.e. the symbols a and [2_1] are also related to the basis [{\bf a}^{s}, ] [{\bf b}^{s},] [{\bf c}^{s}]. At special locations [s{\bf d}=0{\bf d},{{1}\over{2}}{\bf d}] and [s{\bf d}={{1}\over{4}}{\bf d},{{3}\over{4}}{\bf d} ], sectional groups contain both side-preserving and side-reversing operations, whereas for any other location [s{\bf d}] these layer groups are trivial (face) layer groups consisting of side-preserving operations only and are, therefore, also called floating groups in the direction d (IT E , 2002[link]).

Table 3.4.4.7 | top | pdf |
Sectional layer groups and twin (wall) symmetries of the twin [({\sf S}_1|[100];s{\bf d}|{\sf S}_3)] in a calomel crystal

Location [\overline{{\cal F}}_{13}] [\overline{{\cal J}}_{13}] [{\sf T}_{13}] Classification
[s{\bf d} ] Standard Non-coordinate Standard Non-coordinate Standard Non-coordinate
[{{1}\over{4}}{\bf d}, {{3}\over{4}}{\bf d} ] [p12/m1 ] [p\underline{2}_z/m_z ] [pmma ] [pm_{y}^{\star}m_z\underline{a}^{\star}_x ] [p2_1ma ] [p\underline{2}^{\star}_{1y}m_z\underline{a}^{\star}_x ] SR
[0{\bf d}, {{1}\over{4}}{\bf d}] [p12/m1] [p\underline{2}_z/m_z] [pmmm] [pm_{y}^{\star}m_z\underline{m}_{x}^{\star} ] [p2mm ] [p\underline{2}^{\star}_{y}m_z\underline{m}^{\star}_x ] SR
[s{\bf d} ] [p1m1 ] [pm_z ] [pmm2 ] [p2^{\star}_xm_{y}^{\star}m_z ] [p1m1 ] [pm_z ] [{\rm AR}^{\star}]
Shift of origin [{\bf b}_t/2].

The wall (twin) symmetry [{\sf T}_{13}] can be easily deduced from sectional layer groups [\overline{{\cal F}}_{13}] and [\overline{{\cal J}}_{13} ]: the floating group [\widehat{\cal F}_{13}] is just the sectional layer group [\overline{{\cal F}}_{13}] at a general location, [\widehat{\cal F}_{13}=\overline{{\cal F}}_{13}(s{\bf d})=pm_z]. Two other generators in the group symbol of [{\sf T}_{13}] are non-trivial twinning operations (underlined with a star) of [\overline{{\cal J}}_{13}]. The classification in the last column of Table 3.4.4.7[link] is defined in Table 3.4.4.3[link].

Local symmetry exerts constraints on possible displacements of the atoms within a wall. The site symmetry of atoms in a wall of zero thickness, or at the central plane of a finite-thickness domain wall, are defined by the layer group [{\sf T}_{13}]. The site symmetry of the off-centre atoms at [0 \,\lt\, |\xi| \,\lt\, \infty] are determined by floating group [{\widehat{\cal F}}_{13}] and the limiting structures at [\xi \rightarrow -\infty ] and [\xi \rightarrow \infty] by space groups [{\cal F}_1] and [{\cal F}_{3}], respectively. A reasonable condition that the displacements of atoms change continuously if one passes through the wall from [\xi \rightarrow -\infty ] to [\xi \rightarrow \infty] allows one to deduce a qualitative picture of the displacements within a wall.

Symmetry groups of domain pairs, sectional layer groups and the twin symmetry have been derived in the parent clamping approximation (PCA) (see Section 3.4.2.5[link]). As can be seen from Fig. 3.4.3.5[link], a relaxation process, accompanying a lifting of this approximation, consists of a simple shear (shear vector parallel to q) and an elongation (or contraction) in the domain wall along the shear direction (change of the vector [\buildrel {\longrightarrow} \over {AB_{0}}] into the vector [\buildrel {\longrightarrow} \over {AB_1^{+}} ]). These deformations influence neither the layer group [{\sf T}_{13} ] nor its floating group [\widehat{\cal F}_{13}]. Hence the wall (twin) symmetry [{\sf T}_{13}] derived in the parent clamping approximation expresses also the symmetry of a ferroelastic domain wall (twin) with nonzero spontaneous shear unless the simple shear is accompanied by a reshuffling of atoms or molecules in both domains. This useful statement holds for any ferroelastic domain wall (twin).

A microscopic structure of the ferroelastic domain wall in two symmetrically prominent positions is depicted in Fig. 3.4.4.11[link]. For better recognition, displacements of molecules are exaggerated and the changes of the displacement lengths are neglected. Since the symmetry of all groups involved contains a reflection [m_z], the atomic shifts are confined to planes (001). It can be seen in the figure that when one moves through the wall in the direction [110] or [[1\bar10]], the vector of the molecular shift experiences rotations through [{{1}\over{2}}\pi] about the [{\bf c}^{t}] direction in opposite senses for the `black' and `white' molecules.

[Figure 3.4.4.11]

Figure 3.4.4.11 | top | pdf |

Microscopic structure of a ferroelastic domain wall in calomel. (a) and (b) show a domain wall at two different locations with two different layer groups and two different structures of the central planes.

The `black' molecules in the central layer at location [{{1}\over{4}}{\bf d} ] or [{{3}\over{4}}{\bf d}] [wall (a) on the left-hand side of Fig. 3.4.4.11[link]] exhibit nearly antiparallel displacements perpendicular to the wall. Strictly perpendicular shifts would represent `averaged' displacements compatible with the layer symmetry [\overline{{\cal J}}_{13}=p\underline{2}^{\star}_{1y}m_z\underline{a}^{\star}_x ], which is, however, broken by a simple shear that decreases the symmetry to [{\sf T}_{13}=p\underline{2}^{\star}_{1y}m_z\underline{a}^{\star}_x ], which does not require perpendicular displacements of `black' molecules.

The wall with central plane location [0{\bf d}] or [{{1}\over{4}}{\bf d} ] (Fig. 3.4.4.11[link]b) has symmetry [{\sf T}_{13}=p\underline{2}^{\star}_{y}m_z\underline{m}^{\star}_x ], which restricts displacements of `white' molecules of the central layer to the y direction only; the `averaged' displacements compatible with [\overline{{\cal J}}_{13}=pm_{y}^\star m_z {\underline m}_x^\star ] (origin shift [{\bf b}^t /2]) would have equal lengths of shifts in the [+y] and [-y] directions, but the relaxed central layer with symmetry [{\sf T}_{13}=p\underline{2}^{\star}_{y}m_z\underline{m}^{\star}_x ] allows unequal shifts in the [-y] and [+y] directions.

Walls (a) and (b) with two different prominent locations have different layer symmetries and different structures of the central layer. These two walls have extremal energy, but symmetry cannot decide which one has the minimum energy. The two walls have the same polar point-group symmetry [\underline{m}^{\star}_x\underline{2}^{\star}_{y}m_z], which permits a spontaneous polarization along y.

Similar analysis of the displacement and ordering fields in domain walls has been performed for KSCN crystals (Janovec et al., 1989[link]), sodium superoxide NaO2 (Zieliński, 1990[link]) and for the simple cubic phase of fullerene C60 (Saint-Grégoire et al., 1997[link]).

3.4.5. Glossary

| top | pdf |

Note: the correspondence between contracted Greek indices and the Cartesian vector components used in Sections 3.1.3[link] , in the present chapter and in the software GI[\star]KoBo-1, is defined in the following way:[\matrix{\hbox{Cartesian components}\hfill & 11 & 22 & 33 & 23,32& 31,13 & 12,21\cr \hbox{Contracted notation}\hfill & 1 & 2 & 3 & 4 & 5 & 6} ]

In this designation, coefficients with contracted indices 4, 5, 6 appear two times, e.g. index 4 replaces yz in one coefficient and zy in the other coefficient. With this convention, the coefficients transform in tensor space as vector components, but some coefficients differ from the usual matrix notation (Voigt matrices) by numerical factors [see Section 1.1.4.10[link] ; Nye (1985[link]); Sirotin & Shaskolskaya, Appendix E (1982[link])].

(a) Objects

[B_m] domain region
d scanning vector (basis vector of a scanning group)
[{\bf D}_i({\bf S}_k], [B_m)] the ith domain, with domain state [{\bf S}_k] in the mth domain region [{\cal B}_m]
[G{\bf S}_1] G -orbit of principal single-domain states
[G({\bf S}_{1},{\bf S}_{j})] G -orbit of domain pairs
[G({\bf S}_{1}|{\bf n}|{\bf S}_{j})] G -orbit of simple domain twins
n normal to a plane p
p plane of a domain wall, domain wall plane
[{\bf R}_{1},] [{\bf R}_{2},] [\ldots,] [{\bf R}_{i}, ] [\ldots] secondary ferroic single-domain states
[{\bf R}_{1}^+,] [{\bf R}_{1}^-,] [{\bf R}_{2}^+,] [{\bf R}_{2}^-,] [\ldots] disoriented secondary ferroic domain states
[s{\bf d}] [(0\leq s \,\lt\, 1)] location of a plane in crystal lattice
[{\bf S}_{1},] [{\bf S}_{2},] [\ldots,] [{\bf S}_{i}, ] [\ldots] principal single-domain states (orientation states, variants)
[{\bf S}_{1}^+,] [{\bf S}_{1}^-,] [{\bf S}_{2}^+,] [{\bf S}_{2}^-, ] [\ldots] disoriented domain states
[{\sf S}_1,] [{\sf S}_2,] [\ldots,] [{\sf S}_i,] [\ldots] basic (microscopic) single-domain states (structural variants)
[({\bf S}_{i},{\bf S}_{k})] ordered domain pair = ordered pair of domain states [{\bf S}_{i}] and [{\bf S}_{k} ]
[\{{\bf S}_i,{\bf S}_k\}] unordered domain pair = unordered pair of domain states [{\bf S}_{i}] and [{\bf S}_{k}]
[({\bf S}_{i}|{\bf n}|{\bf S}_{k})] simple domain twin formed from single-domain states
[({\bf S}_{i}^{+}|{\bf n}|{\bf S}_{k}^{-})] simple ferroelastic domain twin with a compatible domain wall
[[{\bf S}_{i}|{\bf n}|{\bf S}_{k}]] domain wall in the simple twin [({\bf S}_{i}|{\bf n}|{\bf S}_{k})]
[{\bf T}_{ik}({\bf n})] or [{\bf T}_{ik}] simple domain twin – short symbol
[{\bf W}_{ik}({\bf n})] or [{\bf W}_{ik}] domain wall – short symbol
[\varphi] shear angle, obliquity
[\pm{{1}\over{2}}\varphi] disorientation angle of a domain state

(b) Symmetry groups – point groups in a continuum description and space groups in a microscopic description

F point-group symmetry of the ferroic phase (domain state not specified)
[{\cal F}] space-group symmetry of the ferroic phase (domain state not specified)
[ F_i ] point-group symmetry of a principal domain state [{\bf S}_i]
[{\cal F}_i] space-group symmetry of a basic (microscopic) domain state [{\sf S}_i]
[F_{ik}] point-group symmetry (stabilizer in G) of the ordered domain pair [({\bf S}_{i},{\bf S}_{k})]
[{\cal F}_{ik}] space-group symmetry (stabilizer in [{\cal G}]) of the ordered domain pair [({\sf S}_i,{\sf S}_k)]
[\overline {\sf F}_{ik}] sectional layer group of [F_{ik}]
[\widehat {\sf F}_{ik} ] face group, trivial layer group, scanning group of [F_{ik}]
Fam[\,G] crystal family of the group G
G point-group symmetry of the parent phase
[{\cal G}] space-group symmetry of the parent phase
g point-group symmetry operation of the group [G({\cal G})]
[{\sf g}] space-group symmetry operation of the group [{\cal G}]
[g_{ik}] switching operation in domain pair [({\bf S}_i,{\bf S}_k)], transforms [{\bf S}_i] into [{\bf S}_k]
[g_{ik}^{\star}] transposing operation in domain pair [({\bf S}_i,{\bf S}_k)], exchanges [{\bf S}_i] and [{\bf S}_k], twinning operation of a non-ferroelastic domain pair [({\bf S}_i, {\bf S}_k)]
[I_{G}({\bf S}_i)] stabilizer (isotropy group) of [{\bf S}_i] in G
[{\cal I}_{\cal G}({\sf S}_{i})] stabilizer (isotropy group) of [{\sf S}_i] in [{\cal G}]
[J_{ik}] point-group symmetry (stabilizer in G) of the unordered domain pair [\{{\bf S}_i, {\bf S}_k\}]
[J_{ik}^{\star}] point-group symmetry (stabilizer in G) of a completely transposable domain pair [\{{\bf S}_i,{\bf S}_k\} ]
[{\cal J}_{ik}] space-group symmetry (stabilizer in [{\cal G}]) of the unordered domain pair [\{{\sf S}_i,{\sf S}_k\}]
[K_{ik}] twinning group of the domain pair [({\bf S}_i,{\bf S}_k)]
[K_{ik}^{\star}] twinning group of a completely transposable domain pair [({\bf S}_i,{\bf S}_k)]
[L_i] intermediate group, [F_i \in L_i \in G]
[{\overline {\sf J}}_{ik}] sectional layer group of [J_{ik}]
[{\widehat {\sf J}}_{ik} ] face group, trivial subgroup, floating subgroup of sectional group of [J_{ik}]
[r^{\star}_{ik}] symmetry operation of [{\overline {\sf J}}_{ik}] that exchanges [{\bf S}_{i}] and [{\bf S}_{k} ]
[{\underline s}_{ik}] symmetry operation of [{\overline {\sf J}}_{ik}] that inverts n into −n
[\underline t^{\star}_{ik}] symmetry operation of [{\overline {\sf J}}_{ik}] that exchanges [{\bf S}_{i}] and [{\bf S}_{k} ] and inverts n into −n
[{\sf T}_{ik}({\bf n})] symmetry group of the twin [{\bf T}_{ik}({\bf n})]
[{\sf W}_{ik}({\bf n})] symmetry group of the domain wall [{\bf W}_{ik}({\bf n})]
[{\cal T}_i] translational subgroup of [{\cal F}_i]
[{\cal T}_{ik}] translational subgroup of [{\cal F}_{ik}]

(c) Components of property tensors

[\varepsilon] enantiomorphism
[P_i] polarization
[u_{\mu}] strain
[g_{\mu}] optical activity
[d_{i\mu}] piezoelectricity
[A_{i\nu}] electrogyration
[s_{\mu\nu}] linear elasticity
[Q_{\mu\nu}] electrostriction
   
[i = 1, 2, 3; \mu,\nu = 1, 2, ..., 6.]

Acknowledgements

In the final period this work was supported by Ministry of Education of the Czech Republic under project MCM242200002.

References

First citation Aizu, K. (1969). Possible species of `ferroelastic' crystals and of simultaneously ferroelectric and ferroelastic crystals. J. Phys. Soc. Jpn, 27, 387–396.Google Scholar
First citation Aizu, K. (1970a). Possible species of ferromagnetic, ferroelectric and ferroelastic crystals. Phys. Rev. B, 2, 754–772. Google Scholar
First citation Aizu, K. (1970b). Determination of the state parameters and formulation of spontaneous strain for ferroelastics. J. Phys. Soc. Jpn, 28, 706–716.Google Scholar
First citation Aizu, K. (1972). Electrical, mechanical and electromechanical orders of state shifts in nonmagnetic ferroic crystals. J. Phys. Soc. Jpn, 32, 1287–1301.Google Scholar
First citation Aizu, K. (1973). Second-order ferroic state shifts. J. Phys. Soc. Jpn, 34, 121–128.Google Scholar
First citation Altmann, S. L. & Herzig, P. (1994). Point-group theory tables. Oxford: Clarendon Press.Google Scholar
First citation Amin, A. & Newnham, R. E. (1980). Tertiary ferroics. Phys. Status Solidi A, 61, 215–219.Google Scholar
First citation Barkley, J. R. & Jeitschko, W. (1973). Antiphase boundaries and their interactions with domain walls in ferroelastic–ferroelectric Gd2(MoO4)3. J. Appl. Phys. 44, 938–944.Google Scholar
First citation Bertagnolli, E., Kittinger, E. & Tichý, J. (1978). Observation of reversible elastic Dauphiné twinning in alpha-quartz. J. Phys. (Paris) Lett. 39, 295–297.Google Scholar
First citation Bertagnolli, E., Kittinger, E. & Tichý, J. (1979). Ferrobielastic hysteresis in alpha-quartz. J. Appl. Phys. 50, 6267–6271.Google Scholar
First citation Boulesteix, C. (1984). A survey of domains and domain walls generated by crystallographic phase transitions causing a change of the lattice. Phys. Status Solidi A, 86, 11–42.Google Scholar
First citation Bradley, C. J. & Cracknell, A. P. (1972). The mathematical theory of symmetry in solids. Oxford: Clarendon Press.Google Scholar
First citation Bul'bich, A. A. & Gufan, Yu. M. (1989a). Inevitable symmetry lowering in a domain wall near a reordering phase transition. Sov. Phys. JETP, 67, 1153–1157.Google Scholar
First citation Bul'bich, A. A. & Gufan, Yu. M. (1989b). Phase transition in domain walls. Ferroelectrics, 172, 351–359. Google Scholar
First citation Cahn, R. W. (1954). Twinned crystals. Adv. Phys. 3, 363–445.Google Scholar
First citation Calleja, M., Dove, M. T. & Salje, E. K. H. (2001). Anisotropic ionic transport in quartz: the effect of twin boundaries. J. Phys. Condens. Matter, 13, 9445–9454.Google Scholar
First citation Cao, W. & Barsch, G. R. (1990). Landau–Ginzburg model of interphase boundaries in improper ferroelastic perovskites of [D{_{4h}^{18}}] symmetry. Phys. Rev. B, 41, 4334–4348. Google Scholar
First citation Carpenter, M. A., Salje, E. K. H. & Graeme-Barber, A. (1998). Spontaneous strain as a determinant of thermodynamic properties for phase transitions in minerals. Eur. J. Mineral. 10, 621–691.Google Scholar
First citation Catti, M. & Ferraris, G. (1976). Twinning by merohedry and X-ray structure determination. Acta Cryst. A32, 163–165. Google Scholar
First citation Chen, X. J., Liu, J. S., Zhu, J. S. & Wang, Y. N. (2000). Group theoretical analysis of the domain structure of SrBi2Ta2O9 ferroelectric ceramic. J Phys. Condens. Matter, 12, 3745–3749.Google Scholar
First citation Christian, J. W. (1975). The theory of transformations in metals and alloys. Oxford: Pergamon Press.Google Scholar
First citation Chrosch, J. & Salje, E. K. H. (1999). Temperature dependence of the domain wall width in LaAlO3. J. Appl. Phys. 85, 722–727.Google Scholar
First citation Curien, H. & Donnay, J. D. H. (1959). The symmetry of the complete twin. Am. Mineral. 44, 1067–1070.Google Scholar
First citation Curien, H. & Le Corre, Y. (1958). Notation des macles à l'aide du symbolisme des groupes de couleurs de Chubnikov. Bull. Soc. Fr. Mineral. Cristallogr. 81, 126–132.Google Scholar
First citation Dolino, G. (1985). In Modern problems in condensed matter sciences. Vol. 2, Incommensurate phases in dielectrics, edited by R. Blinc & A. P. Levanyuk, pp. 207–231. Amsterdam: North-Holland. Google Scholar
First citation Dudnik, E. F. & Shuvalov, L. A. (1989). Domain structure and phase boundaries in ferroelastics. Ferroelectrics, 98, 207–234. Google Scholar
First citation Erhart, J. & Cao, W. (2001). Effective symmetry and physical properties of twinned perovskite ferroelectric single crystals. J. Mater. Res. 16, 570–578.Google Scholar
First citation Fatuzzo, E. & Merz, W. J. (1967). Ferroelectricity. Amsterdam: North-Holland.Google Scholar
First citation Fesenko, E. G., Gavrilyachenko, B. G. & Semenchev, A. F. (1990). Domain structure in multiaxial ferroelectric crystals. Rostov on Don: Publishing House of the Rostov University (in Russian).Google Scholar
First citation Fousek, J. (1971). Permissible domain walls in ferroelectric species. Czech. J. Phys. B, 21, 955–968.Google Scholar
First citation Fousek, J. & Janovec, V. (1969). The orientation of domain walls in twinned ferroelectric crystals. J. Appl. Phys. 40, 135–142.Google Scholar
First citation Friedel, G. (1926). Leçons de cristallographie. Nancy, Paris, Strasbourg: Berger-Levrault. Reprint (1964). Paris: Blanchard. Google Scholar
First citation Fuksa, J. (1997). The role of the twinning group of a domain pair in tensor distinction of domain states. Ferroelectrics, 204, 135–155.Google Scholar
First citation Fuksa, J. & Janovec, V. (1995). Permutation classification of domain pairs. Ferroelectrics, 172, 343–350.Google Scholar
First citation Grell, H., Krause, C. & Grell, J. (1989). Tables of the 80 plane space groups in three dimensions. Berlin: Akademie der Wissenschaften der DDR.Google Scholar
First citation Hahn, Th. & Wondratschek, H. (1994). Symmetry of crystals. Introduction to International Tables for Crystallography Vol. A. Sofia: Heron Press.Google Scholar
First citation Hatch, D. M. & Cao, W. (1999). Determination of domain and domain wall formation at ferroic transitions. Ferroelectrics, 222, 1–10.Google Scholar
First citation Hatch, D. M., Ghose, S. & Stokes, H. (1990). Phase transitions in leucite, KAl2O6. I. Symmetry analysis with order parameter treatment and the resulting microscopic distortions. Phys. Chem. Mineral. 17, 220–227.Google Scholar
First citation Hatt, R. A. & Hatch, D. M. (1999). Order parameter profiles in ferroic phase transitions. Ferroelectrics, 226, 61–78. Google Scholar
First citation Holser, W. T. (1958a). Relation of symmetry to structure in twinning. Z. Kristallogr. 110, 249–265.Google Scholar
First citation Holser, W. T. (1958b). Point groups and plane groups in a two-sided plane and their subgroups. Z. Kristallogr. 110, 266–281.Google Scholar
First citation Houchmandzadeh, B., Lajzerowicz, J. & Salje, E. K. H. (1991). Order parameter coupling and chirality of domain walls. J. Phys. Condens. Matter, 3, 5163–5169.Google Scholar
First citation Huang, X. R., Jiang, S. S., Hu, X. B. & Liu, W. J. (1997). Theory of twinning structures in the orthorhombic phase of ferroelectric perovskites. J. Phys. Condens. Matter, 9, 4467–4482.Google Scholar
First citation IEEE Standards on Piezoelectricity (1987). IEEE Std 176–987. New York: The Institute of Electrical and Electronics Engineers, Inc.Google Scholar
First citation Indenbom, V. L. (1982). In Modern crystallography II, edited by B. K. Vainshtein, V. M. Fridkin & V. L. Indenbom, pp. 387–396. Berlin: Springer.Google Scholar
First citation International Tables for Crystallography (2002). Vol. E, Subperiodic groups, edited by V. Kopský & D. B. Litvin. Dordrecht: Kluwer Academic Publishers.Google Scholar
First citation International Tables for Crystallography (2005). Vol. A, Space-group symmetry, 5th edition, edited by Th. Hahn. Heidelberg: Springer.Google Scholar
First citation Ishibashi, Y. (1990). Structure and physical properties of domain walls. Ferroelectrics, 104, 299–310.Google Scholar
First citation Ishibashi, Y. (1992). Domain walls in crystals with incommensurate phases. II. J. Phys. Soc. Jpn, 61, 357–362.Google Scholar
First citation Ishibashi, Y. (1993). The 90°-wall in the tetragonal phase of BaTiO3-type ferroelectrics. J. Phys. Soc. Jpn, 62, 1044–1047.Google Scholar
First citation Janovec, V. (1972). Group analysis of domains and domain pairs. Czech. J. Phys. B, 22, 974–994.Google Scholar
First citation Janovec, V. (1976). Symmetry approach to domain structures. Ferroelectrics, 12, 43–53.Google Scholar
First citation Janovec, V. (1981). Symmetry and structure of domain walls. Ferroelectrics, 35, 105–110.Google Scholar
First citation Janovec, V., Litvin, D. B. & Fuksa, J. (1995). Transposable domain pairs and domain distinction. Ferroelectrics, 172, 351–359. Google Scholar
First citation Janovec, V., Richterová, L. & Litvin, D. B. (1993). Non-ferroelastic twin laws and distinction of domains in non-ferroelastic phases. Ferroelectrics, 140, 95–100.Google Scholar
First citation Janovec, V., Schranz, W., Warhanek, H. & Zikmund, Z. (1989). Symmetry analysis of domain structure in KSCN crystals. Ferroelectrics, 98, 171–189.Google Scholar
First citation Janovec, V. & Zikmund, Z. (1993). Microscopic structure of domain walls and antiphase boundaries in calomel crystals. Ferroelectrics, 140, 89–93.Google Scholar
First citation Jerphagnon, J., Chemla, D. & Bonneville, R. (1978). The description of the physical properties of condensed matter using irreducible tensors. Adv. Phys. 27, 609–650.Google Scholar
First citation Jona, F. & Shirane, G. (1962). Ferroelectric crystals. Oxford: Pergamon Press.Google Scholar
First citation Kalonji, G. (1985). A roadmap for the use of interfacial symmetry groups. J. Phys. (Paris) Colloq. 46, 49–556.Google Scholar
First citation Känzig, W. (1957). Ferroelectrics and antiferroelectrics. In Solid state physics IV, edited by F. Seitz & D. Turnbull, pp. 1–197. New York: Academic Press.Google Scholar
First citation Klassen-Neklyudova, M. V. (1964). Mechanical twinning of crystals. New York: Consultants Bureau.Google Scholar
First citation Knox, R. S. & Gold, A. (1967). Symmetry in the solid state. Introduction. New York: W. A. Benjamin.Google Scholar
First citation Koch, E. (2004). Twinning. In International tables for crystallography, Vol. C, Mathematical, physical and chemical tables, 3rd edition, edited by E. Prince, ch. 1.3. Dordrecht: Kluwer Academic Publishers.Google Scholar
First citation Koňák, Č., Kopský, V. & Smutný, F. (1978). Gyrotropic phase transitions. J. Phys. Solid State Phys. 11, 2493–2518.Google Scholar
First citation Kopský, V. (1979). Tensorial covariants for the 32 crystal point groups. Acta Cryst. A35, 83–95.Google Scholar
First citation Kopský, V. (1982). Group lattices, subduction of bases and fine domain structures for magnetic point groups. Prague: Academia.Google Scholar
First citation Kopský, V. (1993). Translation normalizers of Euclidean motion groups. J. Math. Phys. 34, 1548–1576.Google Scholar
First citation Kopský, V. (2001). Tensor parameters of ferroic phase transitions I. Theory and tables. Phase Transit. 73, 1–422.Google Scholar
First citation Lines, M. E. & Glass, A. M. (1977). Principles and applications of ferroelectric and related materials. Oxford: Clarendon Press.Google Scholar
First citation Litvin, D. B. & Janovec, V. (1999). Classification of domain pairs and tensor distinction. Ferroelectrics, 222, 87–93. Google Scholar
First citation Locherer, K. R., Chrosch, J. & Salje, E. K. H. (1998). Diffuse X-ray scattering in WO3. Phase Transit. 67, 51–63.Google Scholar
First citation Mitsui, T., Tatsuzaki, I. & Nakamura, E. (1976). An introduction to the physics of ferroelectrics. New York: Gordon & Breach.Google Scholar
First citation Newnham, R. E. (1974). Domains in minerals. Am. Mineral. 59, 906–918.Google Scholar
First citation Newnham, R. E. (1975). Structure–property relations. Berlin: Springer.Google Scholar
First citation Newnham, R. E. & Cross, L. E. (1974a). Symmetry of secondary ferroics I. Mater. Res. Bull. 9, 927–934.Google Scholar
First citation Newnham, R. E. & Cross, L. E. (1974b). Symmetry of secondary ferroics II. Mater. Res. Bull. 9, 1021–1032.Google Scholar
First citation Newnham, R. E., Miller, C. S., Cross, L. E. & Cline, T. W. (1975). Tailored domain patterns in piezoelectric crystals. Phys. Status Solidi A, 32, 69–78.Google Scholar
First citation Newnham, R. E. & Skinner, D. P. Jr (1976). Polycrystalline secondary ferroics. Mater. Res. Bull. 11, 1273–1284.Google Scholar
First citation Nye, J. F. (1985). Physical properties of crystals. Oxford: Clarendon Press.Google Scholar
First citation Opechowski, W. (1986). Crystallographic and metacrystallographic groups. Amsterdam: North-Holland.Google Scholar
First citation Palmer, D. C., Putnis, A. & Salje, E. K. H. (1988). Twinning in tetragonal leucite. Phys. Chem. Mineral. 16, 298–303. Google Scholar
First citation Pond, R. C. & Bollmann, W. (1979). The symmetry and interfacial structure of bicrystals. Philos. Trans. R. Soc. London Ser. A, 292, 449–472.Google Scholar
First citation Pond, R. C. & Vlachavas, D. S. (1983). Bicrystallography. Proc. R. Soc. London Ser. A, 386, 95–143.Google Scholar
First citation Přívratská, J. & Janovec, V. (1997). Pyromagnetic domain walls connecting antiferromagnetic non-ferroelastic domains. Ferroelectrics, 204, 321–331.Google Scholar
First citation Přívratská, J. & Janovec, V. (1999). Spontaneous polarization and/or magnetization in non-ferroelastic domain walls: symmetry predictions. Ferroelectrics, 222, 23–32.Google Scholar
First citation Přívratská, J., Janovec, V. & Machonský, L. (2000). Tensor properties discriminating domain walls from non-ferroelastic domains. Ferroelectrics, 240, 83–92.Google Scholar
First citation Putnis, A. (1992). Introduction to mineral sciences. Cambridge University Press.Google Scholar
First citation Rosenman, G., Skliar, A., Eger, D., Oron, M. & Katz, M. (1998). Low temperature periodic electrical poling of flux-grown KTiOPO4 and isomorphic crystals. Appl. Phys. Lett. 73, 3650–3652. Google Scholar
First citation Rosová, A. (1999). Real domain structure origination in (110) mechanical twinning in YBa2Cu3O7−y. In Studies of high temperature superconductors, Vol. 28, edited by A. Narlikar, pp. 125–148. New York: Nova Science Publishers. Google Scholar
First citation Rosová, A., Boulesteix, C. & Vávra, I. (1993). Role of microtwins in twin lamella intersections and interconnections in YBa2Cu3O7−y. Physica C, 214, 247–256. Google Scholar
First citation Rychetský, I. & Schranz, W. (1993). Antiphase boundaries in Hg2Br2 and KSCN. J. Phys. Condens. Matter, 5, 1455–1472. Google Scholar
First citation Rychetský, I. & Schranz, W. (1994). Ferroelastic domain walls in Hg2Br2 and KSCN. J. Phys. Condens. Matter, 6, 11159–11165.Google Scholar
First citation Saint-Grégoire, P. & Janovec, V. (1989). Modulated phases in crystals: Symmetry of walls and wall lattices. Example of quartz. In Nonlinear coherent structures, edited by M. Bartes & J. Léon. Lecture notes in physics, Vol. 353, pp. 117–126. Berlin: Springer.Google Scholar
First citation Saint-Grégoire, P., Janovec, V. & Kopský, V. (1997). A sample analysis of domain walls in simple cubic phase of C60. Ferroelectrics, 191, 73–78.Google Scholar
First citation Salje, E. K. H. (1990). Phase transitions in ferroelastic and co-elastic crystals, 1st edition. Cambridge University Press.Google Scholar
First citation Salje, E. K. H. (1991). Strain-related transformation twinning in minerals. Neues Jahrb. Mineral. Abh. 163, 43–86.Google Scholar
First citation Salje, E. K. H. (2000a). Mesoscopic twin patterns in ferroelastic and co-elastic minerals. Rev. Mineral. Geochem. 39, 65–84.Google Scholar
First citation Salje, E. K. H. (2000b). Ferroelasticity. Contemp. Phys. 41, 79–91.Google Scholar
First citation Sapriel, J. (1975). Domain-wall orientations in ferroelastics. Phys. Rev. B, 12, 5128–5140.Google Scholar
First citation Schlenker, J. L., Gibbs, G. V. & Boisen, M. B. (1978). Strain-tensor components expressed in terms of lattice parameters. Acta Cryst. A34, 52–54.Google Scholar
First citation Schmid, H. (1991). Polarized light microscopy of the ferroelastic domains of YBa2Cu3O7−x. Phase Transit. 30, 205–214.Google Scholar
First citation Schmid, H. (1993). Polarized light microscopy (PLM) of ferroelectric and ferroelastic domains in transmitted and reflected light. In Ferroelectric ceramics, edited by N. Setter & E. L. Colla, pp. 107–126. Basel: Birkhäuser.Google Scholar
First citation Schmid, H., Burkhardt, E., Walker, E., Brixel, W., Clin, M., Rivera, J.-P., Jorda, J.-L., François, M. & Yvon, K. (1988). Polarized light and X-ray precession study of the ferroelastic domains of YBa2Cu3O7−δ. Z. Phys. B Condens. Matter, 72, 305–322.Google Scholar
First citation Schranz, W. (1995). Domains and interfaces near ferroic phase transitions. Key Eng. Mater. 101102, 41–60. Google Scholar
First citation Scott, J. (1998). The physics of ferroelectric ceramic thin films for memory applications. Ferroelectrics Rev. 1, 1–129. Google Scholar
First citation Scott, J. (2000). Ferroelectric memories. Heidelberg: Springer.Google Scholar
First citation Shmyt'ko, I. M., Shekhtman, V. Sh., Ossipyan, Yu. A. & Afonikova, N. S. (1987). Twin structure and structure of twin boundaries in 1–2–3-O7−δ crystals. Ferroelectrics, 97, 151–170.Google Scholar
First citation Shubnikov, A. V. & Kopcik, V. A. (1974). Symmetry in science and art. New York: Plenum Press.Google Scholar
First citation Shur, V. Ya., Batchko, R. G., Rumyantsev, E. L., Miller, G. D., Fejer, M. M. & Byer, R. L. (1999). Domain engineering: periodic domain pattering in lithium niobate. Proc. 11th ISAF, pp. 399–406. Piscataway, NJ: IEEE.Google Scholar
First citation Shur, V. Ya., Gruverman, A. L., Letuchev, V. V., Rumyantsev, E. L. & Subbotin, A. L. (1989). Domain structure of lead germanate. Ferroelectrics, 98, 29–49.Google Scholar
First citation Shur, V. Ya., Rumyantsev, E. L., Nikolaeva, E. V., Shishkin, E. I., Batchko, R. G., Fejer, M. M. & Byer, R. L. (2001). Recent achievements in domain engineering in lithium niobate and lithium tantalate. Ferroelectrics, 257, 191–202.Google Scholar
First citation Shuvalov, L. A. (1988). Editor. Modern crystallography IV. Physical properties of crystals. Berlin: Springer.Google Scholar
First citation Shuvalov, L. A., Dudnik, E. F. & Wagin, S. V. (1985). Domain structure geometry of real ferroelastics. Ferroelectrics, 65, 143–152.Google Scholar
First citation Shuvalov, L. A. & Ivanov, N. R. (1964). Change in the optical activity of ferroelectric crystals on reversal of polarization. Sov. Phys. Crystallogr. 9, 290–299. (Kristallografiya, 9, 363–372.)Google Scholar
First citation Sidorkin, A. S. (2002). Domain structure in ferroelectrics and related materials. Moscow: Fizmatlit (in Russian). Google Scholar
First citation Sirotin, Yu. I. & Shaskolskaya, M. P. (1982). Fundamentals of crystal physics. Moscow: Mir.Google Scholar
First citation Smolenskii, G. A., Bokov, V. A., Isupov, V. A., Krainik, N. N., Pasynkov, R. E. & Shur, M. S. (1984). Physics of ferroelectric phenomena. New York: Gordon & Breach.Google Scholar
First citation Snoeck, E., Saint-Grégoire, P., Janovec, V. & Roucau, C. (1994). TEM study of 3-q modulated phase of quartz-type under electric field. Ferroelectrics, 155, 171–176.Google Scholar
First citation Sonin, E. B. & Tagancev, A. K. (1989). Structure and phase transitions in antiphase boundaries of improper ferroelectrics. Ferroelectrics, 98, 291–295.Google Scholar
First citation Strukov, B. A. & Levanyuk, A. P. (1998). Ferroelectric phenomena in crystals. Berlin: Springer.Google Scholar
First citation Sutton, A. P. & Balluffi, R. W. (1995). Interfaces in crystalline materials. Oxford: Clarendon Press.Google Scholar
First citation Tagancev, A. R. & Sonin, E. B. (1989). Linear singularities and their motion in improper ferroelectrics. Ferroelectrics, 98, 297–300.Google Scholar
First citation Tolédano, J.-C. & Tolédano, P. (1987). The Landau theory of phase transitions. Singapore: World Scientific. Google Scholar
First citation Tolédano, P. & Dmitriev, V. (1996). Reconstructive phase transitions. Singapore: World Scientific.Google Scholar
First citation Tomaszewski, P. E. (1992). Structural phase transitions in crystals. I. Database. II. Statistical analysis. Phase Transit. 38, 127–220, 221–228.Google Scholar
First citation Uchino, K. (2000). Ferroelectric devices. New York: Marcel Dekker.Google Scholar
First citation Vainshtein, B. K. (1994). Modern crystallography I. Symmetry of crystals. Berlin: Springer.Google Scholar
First citation Van Landuyt, J., Van Tendeloo, G., Amelinckx, S. & Walker, M. B. (1985). Interpretation of Dauphiné-twin-domain configurations resulting from the α–β phase transition in quartz and aluminium phosphate. Phys. Rev. B, 31, 2986–2992.Google Scholar
First citation Van Tendeloo, G. & Amelinckx, S. (1974). Group-theoretical considerations concerning domain formation in ordered alloys. Acta Cryst. A30, 431–440.Google Scholar
First citation Wadhawan, V. K. (1991). Ferroelasticity: introductory survey and present status. Phase Transit. 34, 3–18.Google Scholar
First citation Wadhawan, V. K. (2000). Introduction to ferroic materials. The Netherlands: Gordon and Breach.Google Scholar
First citation Wondratschek, H. & Aroyo, M. I. (2001). The application of Hermann's group [{\cal M}] in group–subgroup relations between space groups. Acta Cryst. A57, 311–320.Google Scholar
First citation Wondratschek, H. & Jeitschko, W. (1976). Twin domains and antiphase domains. Acta Cryst. A32, 664–666.Google Scholar
First citation Xu, Y. (1991). Ferroelectric materials and their applications. Amsterdam: North-Holland.Google Scholar
First citation Yin, J. & Cao, W. (2000). Domain configurations in domain engineered 0.995Pb(Zn1/3Nb2/3)O30.045PbTiO3 single crystals. J. Appl. Phys. 87, 7438–7441. Google Scholar
First citation Zheludev, I. S. (1988). Electrical properties of crystals. In Modern crystallography IV. Physical properties of crystals, edited by L. A. Shuvalov, pp. 178–266. Berlin: Springer-Verlag.Google Scholar
First citation Zieliński, P. (1990). Group-theoretical description of domains and phase boundaries in crystalline solids. Surf. Sci. Rep. 11, 179–223.Google Scholar
First citation Zikmund, Z. (1984). Symmetry of domain pairs and domain walls. Czech. J. Phys. B, 34, 932–949.Google Scholar








































to end of page
to top of page