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

International Tables for Crystallography (2006). Vol. D. ch. 3.2, pp. 388-390

Section 3.2.3.3.5. Intermediate subgroups and partitions of an orbit into suborbits

V. Janovec,a* Th. Hahnb and H. Klapperc

a Department of Physics, Technical University of Liberec, Hálkova 6, 461 17 Liberec 1, Czech Republic,bInstitut für Kristallographie, Rheinisch–Westfälische Technische Hochschule, D-52056 Aachen, Germany, and cMineralogisch-Petrologisches Institut, Universität Bonn, D-53113 Bonn, Germany
Correspondence e-mail:  janovec@fzu.cz

3.2.3.3.5. Intermediate subgroups and partitions of an orbit into suborbits

| top | pdf |

Proposition 3.2.3.30  . Let [G{\bf S}_1] be a G orbit from Proposition 3.2.3.23[link] and [L_1] an intermediate group, [F_1 \subset L_1 \subset G. \eqno(3.2.3.78)]A successive decomposition of G into left cosets of [L_1] and [L_1] into left cosets of [F_1] [see (3.2.3.25)[link]] introduces a two-indices relabelling of the objects of a G orbit defined by the one-to-one correspondence [h_jp_kF_1 \leftrightarrow {\bf S}_{jk}, \quad j=1,2,\ldots m, \quad k=1,2,\ldots, d, \eqno(3.2.3.79)]where [\{h_1,h_2,\ldots,h_m\}] are the representatives of the decompositions of G into left cosets of [L_1], [G=h_1L_{1} \cup h_2L_{1} \cup\ldots\cup h_jL_{1}\cup\ldots \cup h_mL_{1}, \quad m=[G:L_1], \eqno(3.2.3.80)]and [\{p_1,p_2,\ldots,p_d\}] are the representatives of the decompositions of [L_1] into left cosets of [F_1], [L_{1}=p_1F_{1} \cup p_2F_{1} \cup\ldots \cup p_kF_{1}\cup\ldots \cup p_dF_{1}, \quad d=[L_1:F_1]. \eqno(3.2.3.81)]

The index n of [F_{1}] in G can be expressed as a product of indices m and d [see (3.2.3.26)[link]], [n=[G:F_1]=[G:L_{1}][L_{1}:F_{1}]=md. \eqno(3.2.3.82)]If G is a finite group, then the index n can be expressed in terms of orders of groups G, [F_1] and [L_1]: [n=|G|:|F_1|=(|G|:|L_{1}|)(|L_{1}|:|F_{1}|)=md. \eqno(3.2.3.83)]

When one chooses [{\bf S}_{1} ={\bf S}_{11}], then the members of the orbit [G{\bf S}_{11}] can be arranged into an [m\times d] array, [\matrix{{\bf S}_{11} & {\bf S}_{12} &\ldots &{\bf S}_{1k} &\ldots&{\bf S}_{1d} \cr {\bf S}_{21} & {\bf S}_{22} &\ldots &{\bf S}_{2k} &\ldots &{\bf S}_{2d} \cr\vdots &\vdots&\ddots &\vdots &\ddots &\vdots\cr{ \bf S}_{j1} &{ \bf S}_{j2} &\ldots &{\bf S}_{jk} &\ldots &{\bf S}_{jd} \cr\vdots &\vdots &\ddots &\vdots &\ddots &\vdots\cr {\bf S}_{m1} & {\bf S}_{m2} &\ldots &{\bf S}_{mk} &\ldots &{\bf S}_{md}} \eqno(3.2.3.84)]

The set of objects of the jth row of this array forms an [L_j] orbit with the representative [{\bf S}_{j 1}], [\eqalignno{ &\{ {\bf S}_{j1}, {\bf S}_{j2},\ldots, {\bf S}_{jk},\ldots, {\bf S}_{jd}\}&\cr&\quad = \{h_jp_1{\bf S}_{11}, h_jp_2{\bf S}_{11},\ldots, h_jp_k{\bf S}_{11},\ldots, h_jp_d{\bf S}_{11}\}&\cr&\quad= L_j{\bf S}_{j 1},&(3.2.3.85)}] where [L_j=h_jL_1h_j^{-1}, \quad {\bf S}_{j 1}=h_j{\bf S}_{11}, \quad j=1,2,\ldots, m. \eqno(3.2.3.86)]The intermediate group [L_1] thus induces a splitting of the orbit [G{\bf S}_{11}] into m suborbits [L_j{\bf S}_{j 1}], [j=1,2,\ldots,m]: [G{\bf S}_{11}=L_1{\bf S}_{11} \cup L_2{\bf S}_{21} \cup\ldots\cup L_j{\bf S}_{j1}\cup\ldots\cup L_m{\bf S}_{m1}, \quad m=[G:L_1]. \eqno(3.2.3.87)]Aizu (1972[link]) denotes this partitioning factorization of species.

The relation (3.2.3.79)[link] is just the application of the correspondence (3.2.3.69)[link] of Proposition 3.2.3.23[link] on the successive decomposition (3.2.3.25)[link]. Derivation of the second part of Proposition 3.2.3.30[link] can be sketched in the following way: [\eqalignno{\{ {\bf S}_{j1}, {\bf S}_{j2},\ldots, {\bf S}_{jd}\}&= h_j\{p_1{\bf S}_{11},p_2{\bf S}_{11},\ldots, p_d{\bf S}_{11}\}&\cr&=h_j\{p_1, p_2,\ldots, p_d\}F_1{\bf S}_{11}&\cr &= h_jL_1{\bf S}_{11} = h_jL_1h_j^{-1}{\bf S}_{j1} =L_j{\bf S}_{j1}, &\cr j&=1,2,\ldots, m, &(3.2.3.88)}]where the relation (3.2.3.70)[link] is used.

We note that the described partitioning of an orbit into suborbits depends on the choice of representative of the first suborbit [{\bf S}_{11}] and that the number of conjugate subgroups [L_j] may be equal to or smaller than the number m of suborbits (see Example [oS] 3.2.3.34[link]).

Each intermediate group [L_1] in Proposition 3.2.3.30[link] can usually be associated with a certain attribute, e.g. a secondary order parameter, which specifies the suborbits.

Example [oP] 3.2.3.31  . Let G be a point group and [X_1] a point of general position [(I_G(X_1)=e)] in the point space. A symmetry descent to a subgroup [L_1\subset G] is accompanied by a splitting of the orbit [GX_1] of [|G|] equivalent points into [m=|G|:|L_1|] suborbits each consisting of [|L_1|] equivalent points. The first suborbit is [L_1X_1], the others are [L_jX_j], [L_j=h_jL_1h_j^{-1}], [X_j=h_jX_1], [j=1,2,\ldots,m], where [h_j] are representatives of left cosets of [L_1] in the decomposition of G [see (3.2.3.80)[link]].

Splitting of orbits of points of general position is a special case in which [I_L(X_1)=I_G(X_1)]. Splitting of orbits of points of special position is more complicated if [I_L(X_1)\subset I_G(X_1)] (see Wondratschek, 1995[link]).

Example [oC] 3.2.3.32  . Let us consider a phase transition accompanied by a lowering of space-group symmetry from a parent space group [\cal G] with translation subgroup T and point group G to a low-symmetry space group [\cal F] with translation subgroup U and point group F. There exists a unique intermediate group [\cal M], called the group of Hermann, which has translation subgroup T and point group [M=F] (see e.g. Hahn & Wondratschek, 1994[link]; Wadhawan, 2000[link]; Wondratschek & Aroyo, 2001[link]).

The decomposition of [{\cal G}] into left cosets of [{\cal M}], corresponding to (3.2.3.80)[link], is in a one-to-one correspondence with the decomposition of G into left cosets of F, since [{\cal G}] and [{\cal M}] have the same translation subgroup [\bf{T}] and [{\cal M}] and [{\cal F}] have the same point group. Therefore, the index [n\equiv [{\cal G}:{\cal M}]=[G:F]=|G|:|F|].

Since [{\cal M}] and [{\cal F}] have the same point group F, the decomposition of [{\cal M}] into left cosets of [{\cal F}], corresponding to (3.2.3.81)[link], is in a one-to-one correspondence with the decomposition of [\bf{T}] into left cosets of [\bf{U}], [{\bf T}={\bf t}_1{\bf U}+{\bf t}_2{\bf U}+\ldots+ {\bf t}_d{\bf U}. \eqno(3.2.3.89)]Representatives [{\bf t}_1, {\bf t}_2,\ldots {\bf t}_d] are translations. The corresponding vectors lead from the origin of a `superlattice' primitive unit cell of the low-symmetry phase to lattice points of [\bf{T}] located within or on the side faces of this `superlattice' primitive unit cell (Van Tendeloo & Amelinckx, 1974[link]). The number [d_t] of these vectors is equal to the ratio [v_{\cal F}:v_{\cal G}=Z_{\cal F}:Z_{\cal G}], where [v_{\cal F}] and [v_{\cal G}] are the volumes of the primitive unit cell of the low-symmetry phase and the parent phase, respectively, and [Z_{\cal F}] and [Z_{\cal G}] are the number of chemical formula units in the primitive unit cell of the low-symmetry phase and the parent phase, respectively.

There is another useful formula for expressing [d_t=[{\bf T}:{\bf U}]]. The primitive basis vectors [{\bf b}_1, {\bf b}_2, {\bf b}_3] of [\bf U] are related to the primitive basis vectors [{\bf a}_1, {\bf a}_2, {\bf a}_3] of [\bf{T}] by a linear relation, [{\bf b}_i=\textstyle\sum\limits_{j=1}^{3}{\bf a}_jm_{ji}, \quad i=1,2,3, \eqno(3.2.3.90)]where [m_{ji}] are integers. The volumes of primitive unit cells are [v_{\cal G}={\bf a}_1({\bf a}_2\times {\bf a}_3)] and [v_{\cal F}={\bf b}_1({\bf b}_2\times {\bf b}_3)]. Using (3.2.3.90)[link], one gets [v_{\cal F}={\rm det}(m_{ij})v_{\cal G}], where [ {\rm det}(m_{ij})] is the determinant of the [(3\times 3)] matrix of the coefficients [m_{ij}]. Hence the index [d_t=(v_{\cal F}:v_{\cal G})={\rm det}(m_{ij})].

Thus we get for the index N of [{\cal F}] in [{\cal G}] [\eqalignno{N&= [{\cal G}:{\cal F}]= [G:F][{\bf T}:{\bf U}]&\cr &= (|G|:|F|)(v_{\cal F}:v_{\cal G})=(|G|:|F|)(Z_{\cal F}:Z_{\cal G})&\cr&= (|G|:|F|) {\rm det}(m_{ij})=nd_t. &(3.2.3.91)}]

Each suborbit, represented by a row in the array (3.2.3.84)[link], contains all basic (microscopic) domain states that are related by pure translations. These domain states exhibit the same tensor properties, i.e. they belong to the same ferroic domain state.

Example [sT] 3.2.3.33  . Let us consider a phase transition with a symmetry descent [G \supset F_1] with an orbit [G{\bf S}_{11}] of domain states. Let [L_1] be an intermediate group, [F_1 \subset L_1 \subset G], and [{\lambda}^{(1)}] the principal order parameter associated with the symmetry descent [G \supset L_1] [cf. (3.2.3.58)[link]], [I_G({\lambda}^{(1)})=L_1]. Since [L_1] is an intermediate group, the quantity [{\lambda}^{(1)}] represents a secondary order parameter of the symmetry descent [G \supset F_1]. The G orbit of [{\lambda}^{(1)}] is [G{\lambda}^{(1)} =\{{\lambda}^{(1)}, {\lambda}^{(2)},\ldots, {\lambda}^{(m)}\}, \quad m=[G:L_1]. \eqno(3.2.3.92)]As in Example [oT] 3.2.3.28[link], there is a bijection between left cosets of the decomposition of G into left cosets of [L_1] [see (3.2.3.80)[link]] and the G orbit of secondary order parameters (3.2.3.92)[link]. One can, therefore, associate with the suborbit [L_j{\bf S}_{j1}] the value [{\lambda}^{(j)}] of the secondary order parameter [\lambda], [L_j{\bf S}_{j1} \leftrightarrow {\lambda}^{(j)}, \quad j=1,2,\ldots, m. \eqno(3.2.3.93)]A suborbit [L_j{\bf S}_{j1}] is thus comprised of objects of the orbit [G{\bf S}_{11}] with the same value of the secondary order parameter [\lambda^{(j)}].

Example [oS] 3.2.3.34  . Let us choose for the intermediate group [L_1] the normalizer [N_G(F_1)]. Then the suborbits equal [\displaylines{N_G(F_j){\bf S}_{j1}=\{h_j{\bf S}_{11},h_jp_{2}{\bf S}_{11},\ldots, h_jp_{d}{\bf S}_{11}\},\cr\hfill\hfill j=1,2,\ldots, m=[G:N_G(F_1)], \hfill(3.2.3.94)}]where [p_{1}=e,p_{2},\ldots, p_{d}] are representatives of left cosets [p_{k}F_1] in the decomposition of [N_G(F_1)], [N_G(F_{1})=p_1F_{1} \cup p_2F_{1} \cup\ldots \cup p_dF_{1}, \quad d=[N_G(F_1):F_1], \eqno(3.2.3.95)]and [h_j] are representatives of the decomposition (3.2.3.77)[link]. The suborbit [ F_j{\bf S}_{j1}] consists of all objects with the same stabilizer [F_j], [\displaylines{I_G({\bf S}_{j1})=I_G({\bf S}_{j2})=\ldots= I_G({\bf S}_{jd})=F_j, \cr j=1,2,\ldots, m=[G:N_G(F_1)]. \cr\hfill(3.2.3.96)}]

Propositions 3.2.3.23[link] and 3.2.3.30[link] are examples of structures that a group action induces from a group G on a G-set. Another important example is a permutation representation of the group G which associates operations of G with permutations of the objects of the orbit [G{\bf S}_i] [see e.g. Kerber (1991[link], 1999[link]); for application of the permutation representation in domain-structure analysis and domain engineering, see e.g. Fuksa & Janovec (1995[link], 2002[link])].

References

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 Fuksa, J. & Janovec, V. (1995). Permutation classification of domain pairs. Ferroelectrics, 172, 343–350.Google Scholar
First citation Fuksa, J. & Janovec, V. (2002). Macroscopic symmetries and domain configurations of engineered domain structures. J. Phys. Condens. Matter, 14, 3795–3812.Google Scholar
First citation Hahn, Th. & Wondratschek, H. (1994). Symmetry of crystals. Sofia: Heron Press Ltd.Google Scholar
First citation Kerber, A. (1991). Algebraic combinatorics via finite group action. Mannheim: B. I. Wissenschaftsverlag.Google Scholar
First citation Kerber, A. (1999). Applied finite group actions. Berlin: Springer.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. (2000). Introduction to ferroic materials. Amsterdam: Gordon and Breach.Google Scholar
First citation Wondratschek, H. (1995). Splitting of Wyckoff positions (orbits). Z. Kristallogr. 210, 567–573.Google Scholar
First citation Wondratschek, H. & Aroyo, M. (2001). The application of Hermann's group [\cal M] in group–subgroup relations between space groups. Acta Cryst. A57, 311–320.Google Scholar








































to end of page
to top of page