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

Journal logoFOUNDATIONS
ADVANCES
ISSN: 2053-2733
Volume 71| Part 2| March 2015| Pages 143-149

Partial order among the 14 Bravais types of lattices: basics and applications

CROSSMARK_Color_square_no_text.svg

aResearch with Neutrons and Muons, Paul Scherrer Institut, WHGA/342, Villigen PSI, CH-5232, Switzerland
*Correspondence e-mail: hans.grimmer@psi.ch

(Received 13 November 2014; accepted 15 December 2014; online 29 January 2015)

Neither International Tables for Crystallography (ITC) nor available crystallography textbooks state explicitly which of the 14 Bravais types of lattices are special cases of others, although ITC contains the information necessary to derive the result in two ways, considering either the symmetry or metric properties of the lattices. The first approach is presented here for the first time, the second has been given by Michael Klemm in 1982. Metric relations between conventional bases of special and general lattice types are tabulated and applied to continuous equi-translation phase transitions.

1. Introduction

The mathematician Michael Klemm (1982[Klemm, M. (1982). Symmetrien von Ornamenten und Kristallen. Berlin: Springer-Verlag.]) published a text `Symmetrien von Ornamenten und Kristallen' intended for students and lecturers of mathematics. Considering conventional cells for the 14 types of Bravais lattices, he determined what lattice types are special cases of others and illustrated the result in a figure. This result is of basic importance but it is mentioned neither in Volume A of International Tables for Crystallography (Hahn, 2002[Hahn, Th. (2002). Editor. International Tables for Crystallography, Vol. A, Space-Group Symmetry, 5th ed. Dordrecht: Kluwer Academic Publishers.]), which we shall refer to as ITC-A, nor in any of the standard crystallography textbooks. To improve general understanding, Grimmer & Nespolo (2006[Grimmer, H. & Nespolo, M. (2006). Z. Kristallogr. 221, 28-50.]) gave a figure where the result was shown using standard crystallographic notations for the lattice types. Klemm (1982[Klemm, M. (1982). Symmetrien von Ornamenten und Kristallen. Berlin: Springer-Verlag.]) and Grimmer & Nespolo (2006[Grimmer, H. & Nespolo, M. (2006). Z. Kristallogr. 221, 28-50.]) arranged the lattice types on five levels, which correspond to different numbers of independent lattice parameters, as shown in Fig. 1[link].

[Figure 1]
Figure 1
The lattice type of the three-dimensional lattice at the upper end of a line is a special case of the type at its lower end.

Looking at graphical representations of the various lattice types as given e.g. in Fig. 9.1.7.1 of ITC-A one finds that certain relations are obvious, e.g. the relations between the primitive lattices aPmPoPtPcP, where the arrow points from general to special case. However, there are also pitfalls: (i) hP is not a special case of oP, although the holohedry of hP contains the one of oP. (ii) The centred monoclinic lattice type mS is called mC in Fig. 9.1.7.1. However, C centring a lattice of type mP gives a new type only if the unique monoclinic axis is not c, otherwise it remains of type mP.

Unfortunately, misunderstandings of the relations between lattice types seem to be frequent. Grimmer & Nespolo's (2006[Grimmer, H. & Nespolo, M. (2006). Z. Kristallogr. 221, 28-50.]) figure has not had the required effect. This led the author to investigate other ways of deriving the result. Considering Niggli-reduced primitive cells instead of the conventional centred ones did not appear advantageous. Finally, the author found that ITC-A contains results that allow one to derive the relations in two different ways, one based on the metric properties of lattices, the other on their symmetries.

In §2[link] it will be shown how Fig. 1[link] can be obtained from metric results given in ITC-A. §3[link] presents for the first time the approach based on the space-group symmetry of the lattice types. In retrospect, it may be surprising that this approach has not been published earlier, considering that it is based on results available already in Internationale Tabellen zur Bestimmung von Kristallstrukturen (Hermann, 1935[Hermann, C. (1935). Editor. Internationale Tabellen zur Bestimmung von Kristallstrukturen, 1. Bd (in German, English and French). Berlin: Borntraeger.]). These tables list the translationengleiche (i.e. equi-translation) subgroups of the space groups in three dimensions, based on the work of Hermann (1929[Hermann, C. (1929). Z. Kristallogr. 69, 533-555.]). In §4[link] metric relations between conventional bases of special and minimally more general lattice types are tabulated. They are applied to continuous equi-translation phase transitions in §5[link].

2. Metric-based derivation of the partial order among the 14 lattice types

Similar to Klemm (1982[Klemm, M. (1982). Symmetrien von Ornamenten und Kristallen. Berlin: Springer-Verlag.]), Gruber (2002[Gruber, B. (2002). International Tables for Crystallography, Vol. A, ch. 9.3. Dordrecht: Kluwer Academic Publishers.]) gives in Table 9.3.4.1 of ITC-A necessary and sufficient conditions for conventional cells of lattices belonging to one of the 14 lattice types. Except for hP and the three cubic types, the conditions contain at least one `<' or `≠'. Replacing one of them in turn by `=', the Bravais types that are minimal special cases of the given type can be determined. In the more complicated cases, Gruber (2002[Gruber, B. (2002). International Tables for Crystallography, Vol. A, ch. 9.3. Dordrecht: Kluwer Academic Publishers.]) does this already in footnotes to his table. Let us do it in all cases.

There are no special cases of cP, cI, cF (a = b = c, α = β = γ = 90°) and of hP (a = b, α = β = 90°, γ = 120°).

For the primitive conventional tetragonal cell the conditions for tP are according to Table 9.3.4.1: a = bc, α = β = γ = 90°. If b = c the type is cP.

For the body-centred conventional tetragonal cell the conditions for tI are: [c/\sqrt 2]a = bc, α = β = γ = 90°. If [c/\sqrt 2] = a the type is cF, if b = c the type is cI.

For the primitive conventional orthorhombic cell the conditions for oP are: a < b < c, α = β = γ = 90°. If a = b or b = c the type is tP.

For the body-centred conventional orthorhombic cell the conditions for oI are: a < b < c, α = β = γ = 90°. If a = b or b = c the type is tI.

For the all-face-centred conventional orthorhombic cell the conditions for oF are: a < b < c, α = β = γ = 90°. If a = b or b = c the type is tI.

For the C-face-centred conventional orthorhombic cell the conditions for oC are: a < b[a\sqrt3], α = β = γ = 90°. If a = b the type is tP, if b = [a\sqrt3] the type is hP.

For the primitive conventional rhombohedral cell the conditions for hR are: a = b = c, α = β = γ, α ≠ 60°, α ≠ 90°, α ≠ arccos(−1/3) = 109°28′16″. If α = 60° the type is cF, if α = 90° the type is cP, if α = arccos(−1/3) the type is cI.

For the primitive conventional monoclinic cell (unique axis b) the conditions for mP are: −2c cosβ < a < c, α = γ = 90° < β. If β = 90° the type is oP, if −2c cosβ = a or a = c the type is oS.

For the body-centred conventional monoclinic cell (unique axis b) the conditions for mI are: −c cosβ < a < c, α = γ = 90° < β and additional conditions given by Gruber that exclude hR. If β = 90° the type is oI, if −c cosβ = a the type is oC, if a = c the type is oF.

These considerations show that all types with less than four independent lattice parameters are special cases of at least one of the two monoclinic types. Fig. 1[link] follows because, obviously, the two monoclinic types are special cases of the anorthic (= triclinic) one.

3. Symmetry-based derivation of a partial order among the 14 lattice types

Each lattice type can be characterized by the space-group type to which its lattices belong. These types are symmorphic and possess the point group of the holohedry. Table 1[link] gives the corresponding information, which can also be found in Vainshtein (1981[Vainshtein, B. K. (1981). Modern Crystallography I. Berlin: Springer-Verlag.]) and in Borchardt-Ott (1997[Borchardt-Ott, W. (1997). Kristallographie, 5th ed. Berlin: Springer.]).

Table 1
The 14 lattice types and the corresponding space-group types (SGTs)

Lattice type International SGT symbol Schoenflies SGT symbol Number of SGT Order of the holohedry
aP P[{\overline 1}] Ci1 2 2
mP P2/m C2h1 10 4
mS (mA, mB, mC, mI) C2/m C2h3 12 4
oP Pmmm D2h1 47 8
oS (oA, oB, oC) Cmmm D2h19 65 8
oF Fmmm D2h23 69 8
oI Immm D2h25 71 8
tP P4/mmm D4h1 123 16
tI I4/mmm D4h17 139 16
hR R[{\overline 3}]m D3d5 166 12
hP P6/mmm D6h1 191 24
cP Pm[{\overline 3}]m Oh1 221 48
cF Fm[{\overline 3}]m Oh5 225 48
cI Im[{\overline 3}]m Oh9 229 48

The notations mS and oS, where S stands for `side-face-centred' (i.e. seitenflächenzentriert), have been proposed as standard ones by de Wolff et al. (1985[Wolff, P. M. de, Belov, N. V., Bertaut, E. F., Buerger, M. J., Donnay, J. D. H., Fischer, W., Hahn, Th., Koptsik, V. A., Mackay, A. L., Wondratschek, H., Wilson, A. J. C. & Abrahams, S. C. (1985). Acta Cryst. A41, 278-280.]). Note that Gruber (2002[Gruber, B. (2002). International Tables for Crystallography, Vol. A, ch. 9.3. Dordrecht: Kluwer Academic Publishers.]) uses a body-centred (innenzentriert) cell mI.

To answer the question `Which lattice types are special cases of others?' we shall make use of information given in ITC-A or, in more detail, in Volume A1 of International Tables for Crystallography (Wondratschek & Müller, 2004[Wondratschek, H. & Müller, U. (2004). Editors. International Tables for Crystallography, Vol. A1, Symmetry Relations Between Space Groups. Dordrecht: Kluwer Academic Publishers.]).

Consider a lattice type g and let G be the corresponding space-group type. Find for G those maximal translation­engleiche (i.e. type I) subgroup types that occur in Table 1[link]. Examples:

(i) g = oC. P2/m appears once and C2/m twice in the list of maximal subgroups of type I of G = Cmmm.

(ii) g = cP. A set of three conjugate P4/mmm and a set of four conjugate R[\overline 3]m appear in the list of maximal subgroups of type I of G = Pm[\overline 3]m.

This procedure leads to the result shown in Fig. 2[link].

[Figure 2]
Figure 2
The Bravais type of the three-dimensional lattice at the upper end of a line is a special case of the type at its lower end. Solid lines indicate normal subgroups, dashed lines sets of conjugate subgroups. The number of conjugate groups in a set is equal to the subgroup index, i.e. the quotient of the orders of the corresponding point groups (4 for the transition cubic to rhombohedral and 3 in the other cases).

For discussing the numbers of subgroups shown in Fig. 2[link], the definition of conventional cells, as given in Part 2 of ITC-A, is needed. This is shown in Table 2[link] for lattices in three dimensions.

Table 2
Conventional bases for the seven lattice systems

Note that Gruber (2002[Gruber, B. (2002). International Tables for Crystallography, Vol. A, ch. 9.3. Dordrecht: Kluwer Academic Publishers.]) used stricter conventions, which generally depend also on the lattice type, e.g. a < b < c for oP, oF and oI.

Crystal family Lattice system Holohedry Restrictions Free parameters Lattice types
Cubic Cubic m[{\overline 3}]m |a| = |b| = |c| = a, α = β = γ = 90° a cP, cF, cI
Tetragonal Tetragonal 4/mmm |a| = |b| = a, α = β = γ = 90° a, c tP, tI
Hexagonal Hexagonal 6/mmm |a| = |b| = a, α = β = 90°, γ = 120° a, c hP
Rhombohedral [{\overline 3}]m |a| = |b| = |c| = a, α = β = γ a, α hR
Orthorhombic Orthorhombic mmm α = β = γ = 90° a, b, c oP, oS, oF, oI
Monoclinic Monoclinic 2/m α = γ = 90° (b unique) a, b, c, β mP, mS
      α = β = 90° (c unique) a, b, c, γ  
Anorthic Anorthic [{\overline 1}]   a, b, c, α, β, γ aP

Alternatively, two lattice systems, hexagonal and rhombohedral, are combined in the hexagonal crystal family, where hR is considered as a rhombohedrally centred hexagonal lattice instead of a primitive rhombohedral lattice.

Fig. 2[link] shows the changes of lattice types that are possible in phase transitions where the lattice changes continuously: the corresponding pairs of lattice types are connected by lines. Note that the two types of these pairs always belong to different crystal families, whence they have different conventional bases.

Let us now discuss the numbers of subgroups shown in Fig. 2[link].

A rhombohedral deformation of the conventional cubic cell lets only one of the four threefold axes survive, transforming the lattice types cP, cF and cI into hR. A tetragonal deformation of the conventional cubic cell lets only one of the three fourfold axes survive, transforming cP into tP, cF and cI into tI.

The plane perpendicular to the sixfold axis of a lattice of type hP contains three pairs of mutually orthogonal twofold axes. An orthorhombic deformation lets only one of these three pairs survive, transforming hP into oC.

A plane perpendicular to the fourfold axis of a tetragonal lattice cuts the conventional tetragonal cell into a square, which contains two pairs of mutually orthogonal twofold axes, parallel either to the edges or the diagonals of the square. An orthorhombic deformation lets only one of these two pairs survive, transforming tP either into oP or oC, and tI either into oI or oF, depending on whether the square is deformed into a rectangle or a rhombus.

Perpendicular to the threefold axis in a primitive rhombohedral cell, there are three twofold axes at 120° one to another. A monoclinic deformation of this cell can be done in three ways, preserving one of the twofold axes and transforming hR into mC.

A monoclinic deformation of the conventional ortho­rhombic cell preserves only one of the three mutually perpendicular twofold orthorhombic axes. In all three cases oP is transformed into mP whereas oF and oI are transformed into mC. The type oC is transformed into mP if the twofold axis perpendicular to the C-face survives, into mC if one of the two twofold axes in the C-face survives.

Finally, an anorthic deformation removes the twofold monoclinic axis and transforms mP and mC into aP.

In two dimensions one obtains Fig. 3[link].

[Figure 3]
Figure 3
The type of the two-dimensional lattice at the upper end of a line is a special case of the type at its lower end. Solid lines indicate normal subgroups, dashed lines sets of conjugate subgroups.

Note that the partially ordered set formed by the two-dimensional lattice types mp, op, oc, tp and hp has the same structure as the partially ordered set formed by mP, oP, oC, tP and hP, as indicated by the notation.

4. Metric relations between conventional bases of `neighbouring' lattice types

In this section we express the conventional basis of each lattice type in terms of the conventional basis of each lattice type that is minimally more general. In the case of rhombohedral lattices both conventions are considered in Table 3[link], the one where a primitive cell is used for hR and the one where a rhombohedrally centred hexagonal cell is used in the usual obverse setting with lattice points at 0, 0, 0, [{2\over 3},{1\over 3},{1\over 3}] and [{1\over 3},{2\over 3},{2\over 3}].

Table 3
Metric relations between the conventional unit cells of lattice types related by conjugate subgroups, i.e. joined by dashed lines in Fig. 2[link]

`hR, hex' considers hR as a rhombohedrally centred hexagonal lattice, `hR, rho' considers hR as a primitive rhombohedral lattice; d′ = −a′ − b′.

Lattice types L1 → L2 Corresponding space-group types Basis a′, b′, c′ of conventional cell for L2 expressed in terms of basis a, b, c of conventional cell for L1 Det Limiting case of L1 for which it becomes L2 Basis a, b, c of conventional cell for L1 expressed in terms of basis a′, b′, c′ of conventional cell for L2
    a b d c     a b c
hR, hex → cP R[{\overline 3}]mPm[{\overline 3}]m [{1\over 3}](2a+b+c) [{1\over 3}](−a+b+c)   [{1\over 3}](a+2bc) [{1\over 3}] c = ½(61/2) a a′−b b′−c a′+b′+c
[{1\over 3}](2a+b+c) [{1\over 3}](−a+b+c)   [{1\over 3}](a+2bc)     a′+b −b′+c a′−b′−c
[{1\over 3}](2a+b+c) [{1\over 3}](−a+b+c)   [{1\over 3}](a+2bc)     −a′−b b′+c −a′+b′−c
[{1\over 3}](2a+b+c) [{1\over 3}](−a+b+c)   [{1\over 3}](a+2bc)     a′+b b′−c a′−b′+c
hR, hex → cF R[{\overline 3}]mFm[\bar 3]m [{1\over 3}](4a+2bc) [{1\over 3}](2a−2b+c)   [{1\over 3}](2a+4b+c) [{4\over 3}] c = 61/2 a ½(−a′+b′) ½(−b′+c′) a′+b′+c
[{1\over 3}](4a+2bc) [{1\over 3}](2a−2b+c)   [{1\over 3}](2a+4b+c)     ½(−a′−b′) ½(b′−c′) a′−b′−c
[{1\over 3}](4a+2bc) [{1\over 3}](2a−2b+c)   [{1\over 3}](2a+4b+c)     ½(a′+b′) ½(−b′−c′) a′+b′−c
[{1\over 3}](4a+2bc) [{1\over 3}](2a−2b+c)   [{1\over 3}](2a+4b+c)     ½(a′−b′) ½(b′+c′) a′−b′+c
hR, hex → cI R[{\overline 3}]mIm[{\overline 3}]m [{1\over 3}](2a+b−2c) [{1\over 3}](ab+2c)   [{1\over 3}](a+2b+2c) [{2\over 3}] c = ¼(61/2) a a′+b b′+c ½(a′+b′+c′)
[{1\over 3}](2a+b−2c) [{1\over 3}](ab+2c)   [{1\over 3}](a+2b+2c)     a′−b b′−c ½(a′−b′−c′)
[{1\over 3}](2a+b−2c) [{1\over 3}](ab+2c)   [{1\over 3}](a+2b+2c)     a′+b b′−c ½(−a′+b′−c′)
[{1\over 3}](2a+b−2c) [{1\over 3}](ab+2c)   [{1\over 3}](a+2b+2c)     a′−b b′+c ½(−a′−b′+c′)
hR, rho → cP R[{\overline 3}]mPm[{\overline 3}]m a b   c 1 cos α = 0 → α = 90° a b c
a b   c     a b c
a b   c     a b c
a b   c     a b c
hR, rho → cF R[{\overline 3}]mFm[{\overline 3}]m a+b+c ab+c   a+bc 4 cos α = ½ → α = 60° ½(b′+c′) ½(c′+a′) ½(a′+b′)
a+b+c a+bc   ab+c     ½(−b′−c′) ½(−c′+a′) ½(a′−b′)
abc ab+c   ab+c     ½(b′−c′) ½(−c′−a′) ½(−a′+b′)
abc a+bc   a+bc     ½(−b′+c′) ½(c′−a′) ½(−a′−b′)
hR, rho → cI R[{\overline 3}]mIm[{\overline 3}]m b+c a+c   a+b 2 cos α = −[{1\over 3}]α = 109.47° ½(−a′+b′+c′) ½(a′−b′+c′) ½(a′+b′−c′)
b+c ac   ab     ½(−a′−b′−c′) ½(a′+b′−c′) ½(a′−b′+c′)
bc a+c   ab     ½(a′+b′−c′) ½(−a′−b′−c′) ½(−a′+b′+c′)
bc ac   a+b     ½(a′−b′+c′) ½(−a′+b′+c′) ½(−a′−b′−c′)
tPcP P4/mmmPm[{\overline 3}]m a b   c 1 c = a a b c
c a   b     b c a
b c   a     c a b
tIcF I4/mmmFm[{\overline 3}]m a+b −a+b   c 2 c = 21/2 a ½(a′−b′) ½(a′+b′) c
c a+b   a+b     ½(b′−c′) ½(b′+c′) a
a+b c   a+b     ½(c′−a′) ½(c′+a′) b
tIcI I4/mmmIm[{\overline 3}]m a b   c 1 c = a a b c
c a   b     b c a
b c   a     c a b
oShP CmmmP6/mmm −½(a+b) ½(ab) b c ½ a = 31/2 b a′+b d c
b −½(a+b) ½(ab) c     b′+d a c
½(ab) b −½(a+b) c     d′+a b c
mShR, hex C12/m1 → R[{\overline 3}]m −½(3a+b+2c) ½(3ab+2c) b c [{3\over 2}] cos β = −2c/(3a) and 9a2 = 3b2 + 4c2 [{1\over 3}](−a′+b′−2c′) d c
b −½(3a+b+2c) ½(3ab+2c) c   [{1\over 3}](−b′+d′−2c′) a c
½(3ab+2c) b −½(3a+b+2c) c   [{1\over 3}](−d′+a′−2c′) b c
A112/mR[{\overline 3}]m −½(3b+c+2a) ½(3bc+2a) c a cos γ = −2a/(3b) and 9b2 = 3c2 + 4a2 c [{1\over 3}](−a′+b′−2c′) d
c −½(3b+c+2a) ½(3bc+2a) a     c [{1\over 3}](−b′+d′−2c′) a
½(3bc+2a) c −½(3b+c+2a) a     c [{1\over 3}](−d′+a′−2c′) b
mShR, rho C12/m1 → R[{\overline 3}]m −½(a+b) a+c   −½(ab) ½ cos β = −2c/(3a) and 9a2 = 3b2 + 4c2 a′−c a′+c a′+b′+c
−½(ab) −½(a+b)   a+c   b′−a b′+a a′+b′+c
a+c −½(ab)   −½(a+b)   c′−b c′+b a′+b′+c
A112/mR[{\overline 3}]m −½(b+c) a+b   −½(bc) cos γ = −2a/(3b) and 9b2 = 3c2 + 4a2 a′+b′+c a′−c a′+c
−½(bc) −½(b+c)   a+b     a′+b′+c b′−a b′+a
a+b −½(bc)   −½(b+c)     a′+b′+c c′−b c′+b

In the case of conjugate subgroups the various expressions for a′, b′, c′ are equivalent by symmetry. For the transitions hR to cubic, the three other possibilities are obtained from the first one keeping one of the vectors a′, b′, c′ and changing the signs of the other two. For the transitions tetragonal to cubic, the three possibilities are related by cyclic permutations a′ → b′ → c′ → a′; for the transition oS to hP they are related by cyclic permutations a′ → b′ → d′ → a′. Note that for the transition mS to hR the three possibilities are related by cyclic permutations a′ → b′ → c′ → a′ if hR is considered as a primitive rhombohedral lattice, and by a′ → b′ → d′ → a′ if hR is considered as a rhombohedrally centred hexagonal lattice.

The transition mS → hR is given (for hR, hex and hR, rho) first with unique monoclinic axis b, then with axis c.

The column `Det' in Tables 3[link] and 4[link] gives the determinant of the matrix M expressing a′, b′, c′ in terms of a, b, c. It equals the number of lattice points in a conventional cell of L2 divided by the number of lattice points in a conventional cell of L1.

Table 4
Metric relations between the conventional unit cells of lattice types related by normal subgroups, i.e. joined by full lines in Fig. 2[link]

Lattice types L1 → L2 Corresponding space-group types Basis a′, b′, c′ of conventional cell for L2 expressed in terms of basis a, b, c of conventional cell for L1 Det Limiting case of L1 for which it becomes L2 Basis a, b, c of conventional cell for L1 expressed in terms of basis a′, b′, c′ of conventional cell for L2
    a b c     a b c
oPtP PmmmP4/mmm a b c 1 b = a a b c
oCtP CmmmP4/mmm ½(a+b) ½(−a+b) c ½ b = a a′−b a′+b c
oFtI FmmmI4/mmm ½(a+b) ½(−a+b) c ½ b = a a′−b a′+b c
oItI ImmmI4/mmm a b c 1 b = a a b c
mPoP P112/mPmmm a b c 1 γ = 90° a b c
  P12/m1 → Pmmm a b c   β = 90° a b c
  P12/m1 → Pmmm b c a   α = 90° c a b
mPoS P112/mCmmm a+b −a+b c 2 b = a ½(a′−b′) ½(a′+b′) c
mSoS C12/m1 → Cmmm a b c 1 β = 90° a b c
mSoS C12/m1 → Cmmm b a c 1 α = 90° −b a c
mSoF A112/mFmmm 2a+b b c 2 cos γ = −b/(2a) ½(a′−b′) b c
  C12/m1 → Fmmm a b 2c+a   cos β = −a/(2c) a b ½(c′−a′)
  C12/m1 → Fmmm b a 2c+a   cos α = −c/(2b) −b a ½(b′+c′)
mSoI A112/mImmm −ba a c 1 cos γ = −a/b b a′−b c
  C12/m1 → Immm c b ac   cos β = −c/a −c−a b a
  C12/m1 → Immm b c a+c   cos α = −b/c b′+c a b
aPmP P[{\overline 1}]P12/m1 a b c 1 α = γ = 90° a b c
P[{\overline 1}]P112/m a b c α = β = 90° a b c
aPmS P[{\overline 1}]C12/m1 a+b −a+b c 2 b = a, β = α ½(a′−b′) ½(a′+b′) c
P[{\overline 1}]A112/m a b+c −b+c b = c, β = γ a ½(b′−c′) ½(b′+c′)

The restrictions given in Table 2[link] do not determine a unique conventional basis. In accordance with Volume A1 of International Tables for Crystallography (Wondratschek & Müller, 2004[Wondratschek, H. & Müller, U. (2004). Editors. International Tables for Crystallography, Vol. A1, Symmetry Relations Between Space Groups. Dordrecht: Kluwer Academic Publishers.]), conventional bases for lattice-type pairs L1 and L2 have been chosen in Tables 3[link] and 4[link] such that the matrix M becomes as simple as possible (e.g. M is the identity matrix for each of the transitions aPmPoPtPcP). As a consequence, the conditions given in the column `Limiting case of L1 for which it becomes L2' of Tables 3[link] and 4[link] take a particularly simple form. This column contains two conditions that must be satisfied for the transitions mShR and anorthic → monoclinic, in accordance with Fig. 1[link], which shows that the number of independent lattice parameters is reduced by 2 in these cases.

Remark to Table 4[link]: for the transitions monoclinic to orthorhombic, the unique monoclinic axis is the orthorhombic axis c′ in the first, b′ in the second and a′ in the third line; for the transitions anorthic to monoclinic, the unique monoclinic axis is b′ in the first line, c′ in the second.

5. Applications

Let us consider two applications to continuous equi-translation phase transitions. Both concern transitions mShR, hR being considered as a rhombohedrally centred hexagonal lattice in the first example and as a primitive rhombohedral lattice in the second.

(i) Przeniosło et al. (2014[Przeniosło, R., Sosnowska, I., Stekięl, M., Wardecki, D., Fitch, A. & Jasiński, J. B. (2014). Physica B, 449, 72-76.]) measured the monoclinic deformation of the crystal lattice of hematite (α-Fe2O3) at room temperature. Hematite is paramagnetic with space group R[\overline 3]c above its Néel temperature TN = 955 K. Below TN it is weakly ferromagnetic (canted antiferromagnet) with space group C2/c down to the Morin temperature TM = 260 K. The lattice type is therefore mC at room temperature and changes to hR at TN. Equations A1–A3 in Appendix A of Przeniosło et al. (2014[Przeniosło, R., Sosnowska, I., Stekięl, M., Wardecki, D., Fitch, A. & Jasiński, J. B. (2014). Physica B, 449, 72-76.]) show that they chose a conventional C-centred monoclinic cell with basis a = [{1\over 3}](−a′ + b′ − 2c′), b = −a′ − b′, c = c′, where a′, b′, c′ is the conventional basis of the rhombo­hedrally centred hexagonal cell, as suggested in our Table 3[link]. The authors found for their sample I at room temperature: a = 961.935 (12), b = 503.575 (7), c = 1375.277 (17) pm and β = 162.4049 (2)°. Neglecting experimental uncertainties, we obtain from the last column of Table 3[link] that the lattice becomes rhombohedral if β = arccos[−2c/(3a)] = 162.3889° and a = [{1\over 3}](3b2 + 4c2)1/2 = 961.845 pm.

It follows that although the measurement was performed approximately 660 K below the phase-transition temperature, a deviates by only 0.1 pm and β by only 1 minute of arc from the values for a rhombohedral lattice. The calculations given in Table 2[link] of Przeniosło et al. (2014[Przeniosło, R., Sosnowska, I., Stekięl, M., Wardecki, D., Fitch, A. & Jasiński, J. B. (2014). Physica B, 449, 72-76.]), which take account of experimental uncertainties, lead to a similar deviation for a and an even smaller deviation for β of the order of 1 second of arc. For their sample VI, the deviation for β even has opposite sign. We conclude that the magnetic ordering, which destroys the trigonal symmetry, affects the lattice parameters so little that high-resolution synchrotron radiation diffraction is necessary to measure the effect.

(ii) Pyridinium tetrafluoroborate [C5H6N]+BF4 has been investigated by Czarnecki et al. (1998[Czarnecki, P., Katrusiak, A., Szafraniak, I. & Wąsicki, J. (1998). Phys. Rev. B, 57, 3326-3332.]). It is paraelectric at room temperature with space group R[\overline 3]m and undergoes at T = 238.7 K a continuous transition to a ferroelectric phase with space group C2. It follows that the lattice type is hR at room temperature and changes to mC at T. Using high-resolution neutron powder diffraction, Czarnecki et al. (1998[Czarnecki, P., Katrusiak, A., Szafraniak, I. & Wąsicki, J. (1998). Phys. Rev. B, 57, 3326-3332.]) found for the conventional primitive rhombohedral cell a′, b′, c′ at 293 K: |a′| = |b′| = |c′| = a′ = 567.074 (7) pm and α′ = β′ = γ′ = 97.305 (1)°. The entry at the bottom of our Table 3[link] tells us that C-centring the conventional monoclinic cell with basis a = −a′ − c′, b = −a′ + c′, c = a′ + b′ + c′ produces a primitive cell that coincides with the rhombohedral one in the limiting case that mC becomes hR. According to their Fig. 3[link], Czarnecki et al. (1998[Czarnecki, P., Katrusiak, A., Szafraniak, I. & Wąsicki, J. (1998). Phys. Rev. B, 57, 3326-3332.]) chose a different conventional monoclinic cell [{\tilde{\bf a}}] = a′ + c′, [{\tilde{\bf b}}] = a′ − c′, [{\tilde{\bf c}}] = b′. They found at 230 K: [\tilde a] = 734.68 (2), [\tilde b] = 839.95 (2), [\tilde c] = 571.14 (2) pm and [{\tilde \beta}] = 101.952 (2)°. With our cell choice a = −[{\tilde{\bf a}}], b = −[{\tilde{\bf b}}], c = [{\tilde{\bf a}}] + [{\tilde{\bf c}}], we obtain at 230 K: a = [\tilde a], b = [\tilde b], c = 831.96 pm and β = 137.808°. It follows that arccos[−2c/(3a)] − β = 1.212° and [{1\over 3}][3b2 + 4c2]1/2a = 2.07 pm, showing that 8.7 K below the transition temperature T the lattice already differs considerably from a rhombohedral one, in contrast to the first example.

6. Discussion and conclusions

Figs. 1[link][link]–3[link] show basic relations between the lattice types. These relations are of importance also for applications: they tell us the changes of lattice type possible in continuous equi-translation phase transitions, as discussed e.g. by Landau & Lifshitz (1980[Landau, L. D. & Lifshitz, E. M. (1980). Statistical Physics, Part 1, 3rd ed. Translated from the Russian by J. B. Sykes & M. J. Kearsley. Oxford: Pergamon Press.]) or Burns & Glazer (1990[Burns, G. & Glazer, A. M. (1990). Space Groups for Solid State Scientists. San Diego: Academic Press.]). They can be useful also in cases of twinning by `metric merohedry' as defined by Nespolo & Ferraris (2000[Nespolo, M. & Ferraris, G. (2000). Z. Kristallogr. 215, 77-81.]). The use of Fig. 2[link] and of Tables 3[link] and 4[link], which give metric relations between the conventional bases of lattice types joined by lines in Fig. 2[link], has been illustrated with two examples. Tables 3[link] and 4[link] also illustrate how the distinction between conjugate and normal subgroups made in Fig. 2[link] affects the relation between the conventional bases.

For all these reasons, it is suggested that figures like Figs. 2[link] and 3[link] and tables like Tables 3[link] and 4[link] be introduced in future editions of ITC-A. Symmetry aspects are central to ITC-A; its Fig. 10.1.3.2 and Fig. 10.1.3.1 show relations between the types of crystallographic point groups in three and two dimensions in a similar way as our Figs. 2[link] and 3[link] give relations between the lattice types. To balance the information given in ITC-A on the point group and lattice aspects of the space groups, also the Bravais lattice type should be given in the banner line of each space-group type.

Acknowledgements

The author wishes to thank Dr Howard Flack for suggestions and corrections that led to significant improvement of the paper.

References

First citationBorchardt-Ott, W. (1997). Kristallographie, 5th ed. Berlin: Springer.  Google Scholar
First citationBurns, G. & Glazer, A. M. (1990). Space Groups for Solid State Scientists. San Diego: Academic Press.  Google Scholar
First citationCzarnecki, P., Katrusiak, A., Szafraniak, I. & Wąsicki, J. (1998). Phys. Rev. B, 57, 3326–3332.  Web of Science CrossRef CAS Google Scholar
First citationGrimmer, H. & Nespolo, M. (2006). Z. Kristallogr. 221, 28–50.  Web of Science CrossRef CAS Google Scholar
First citationGruber, B. (2002). International Tables for Crystallography, Vol. A, ch. 9.3. Dordrecht: Kluwer Academic Publishers.  Google Scholar
First citationHahn, Th. (2002). Editor. International Tables for Crystallography, Vol. A, Space-Group Symmetry, 5th ed. Dordrecht: Kluwer Academic Publishers.  Google Scholar
First citationHermann, C. (1929). Z. Kristallogr. 69, 533–555.  Google Scholar
First citationHermann, C. (1935). Editor. Internationale Tabellen zur Bestimmung von Kristallstrukturen, 1. Bd (in German, English and French). Berlin: Borntraeger.  Google Scholar
First citationKlemm, M. (1982). Symmetrien von Ornamenten und Kristallen. Berlin: Springer-Verlag.  Google Scholar
First citationLandau, L. D. & Lifshitz, E. M. (1980). Statistical Physics, Part 1, 3rd ed. Translated from the Russian by J. B. Sykes & M. J. Kearsley. Oxford: Pergamon Press.  Google Scholar
First citationNespolo, M. & Ferraris, G. (2000). Z. Kristallogr. 215, 77–81.  Web of Science CrossRef CAS Google Scholar
First citationPrzeniosło, R., Sosnowska, I., Stekięl, M., Wardecki, D., Fitch, A. & Jasiński, J. B. (2014). Physica B, 449, 72–76.  Google Scholar
First citationVainshtein, B. K. (1981). Modern Crystallography I. Berlin: Springer-Verlag.  Google Scholar
First citationWolff, P. M. de, Belov, N. V., Bertaut, E. F., Buerger, M. J., Donnay, J. D. H., Fischer, W., Hahn, Th., Koptsik, V. A., Mackay, A. L., Wondratschek, H., Wilson, A. J. C. & Abrahams, S. C. (1985). Acta Cryst. A41, 278–280.  CrossRef Web of Science IUCr Journals Google Scholar
First citationWondratschek, H. & Müller, U. (2004). Editors. International Tables for Crystallography, Vol. A1, Symmetry Relations Between Space Groups. Dordrecht: Kluwer Academic Publishers.  Google Scholar

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

Journal logoFOUNDATIONS
ADVANCES
ISSN: 2053-2733
Volume 71| Part 2| March 2015| Pages 143-149
Follow Acta Cryst. A
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds