International
Tables for
Crystallography
Volume A
Space-group symmetry
Edited by Th. Hahn

International Tables for Crystallography (2006). Vol. A. ch. 9.2, pp. 750-755
https://doi.org/10.1107/97809553602060000518

Chapter 9.2. Reduced bases

P. M. de Wolffa

a Laboratorium voor Technische Natuurkunde, Technische Hogeschool, Delft, The Netherlands

In this chapter, reduced bases (reduced cells) are described. The treatment starts with the definition of the reduced basis and the reduced form in terms of the metric tensor and lists the conditions (both main and special) for reduced bases of type-I cells and type-II cells. This is followed by a detailed and systematic geometric explanation of these conditions. The resulting 44 lattice characters are defined and tabulated, and the relations between the lattice characters and the conventional cell parameters of the 14 Bravais lattices are listed and discussed.

Keywords: reduced bases; lattice characters; unit cell; lattices; Bravais types; transformations.

9.2.1. Introduction

| top | pdf |

Unit cells are usually chosen according to the conventions mentioned in Chapter 9.1[link] so one might think that there is no need for another standard choice. This is not true, however; conventions based on symmetry do not always permit unambiguous choice of the unit cell, and unconventional descriptions of a lattice do occur. They are often chosen for good reasons or they may arise from experimental limitations such as may occur, for example, in high-pressure work. So there is a need for normalized descriptions of crystal lattices.

Accordingly, the reduced basis1 (Eisenstein, 1851[link]; Niggli, 1928[link]), which is a primitive basis unique (apart from orientation) for any given lattice, is at present widely used as a means of classifying and identifying crystalline materials. A comprehensive survey of the principles, the techniques and the scope of such applications is given by Mighell (1976)[link]. The present contribution merely aims at an exposition of the basic concepts and a brief account of some applications.

The main criterion for the reduced basis is a metric one: choice of the shortest three non-coplanar lattice vectors as basis vectors. Therefore, the resulting bases are, in general, widely different from any symmetry-controlled basis, cf. Chapter 9.1[link] .

9.2.2. Definition

| top | pdf |

A primitive basis a, b, c is called a `reduced basis' if it is right-handed and if the components of the metric tensor G (cf. Chapter 9.1[link] ) [\matrix{{\bf a \cdot a} &{\bf b \cdot b} &{\bf c \cdot c}\cr {\bf b \cdot c} &{\bf c \cdot a} &{\bf a \cdot b}\cr} \eqno(9.2.2.1)] satisfy the conditions shown below. The matrix (9.2.2.1)[link] for the reduced basis is called the reduced form.

Because of lattice symmetry there can be two or more possible orientations of the reduced basis in a given lattice but, apart from orientation, the reduced basis is unique.

Any basis, reduced or not, determines a unit cell – that is, the parallelepiped of which the basis vectors are edges. In order to test whether a given basis is the reduced one, it is convenient first to find the `type' of the corresponding unit cell. The type of a cell depends on the sign of [T = ({\bf a \cdot b}) ({\bf b \cdot c}) ({\bf c \cdot a}).] If [T \gt 0], the cell is of type I, if [T \leq 0] it is of type II. `Type' is a property of the cell since T keeps its value when a, b or c is inverted. Geometrically speaking, such an inversion corresponds to moving the origin of the basis towards another corner of the cell. Corners with all three angles acute occur for cells of type I (one opposite pair, the remaining six corners having one acute and two obtuse angles). The other alternative, specified by main condition (ii)[link] of Section 9.2.3[link], viz all three angles non-acute, occurs for cells of type II (one or more opposite pairs, the remaining corners having either one or two acute angles).

The conditions can all be stated analytically in terms of the components (9.2.2.1)[link], as follows:

  • (a) Type-I cell

    Main conditions: [\eqalignno{{\bf a \cdot a} \leq {\bf b \cdot b} \leq {\bf c \cdot c}\semi \quad& |{\bf b \cdot c}| \leq {\textstyle{1 \over 2}} {\bf b \cdot b}\semi \quad |{\bf a \cdot c}| \leq {\textstyle{1 \over 2}} {\bf a \cdot a}\semi & \cr & |{\bf a \cdot b}| \leq {\textstyle{1 \over 2}} {\bf a \cdot a} &(9.2.2.2a)\cr {\bf b \cdot c} \;\gt\; 0\semi \quad& \phantom{|}{\bf a \cdot c} \;\gt\; 0\semi \quad{\bf a \cdot b} \;\gt\; 0. & (9.2.2.2b)}] Special conditions: [\eqalignno{&\hbox{if } {\bf a \cdot a} = {\bf b \cdot b}\quad\hbox{then}\quad {\bf b \cdot c} \leq {\bf a \cdot c} & (9.2.2.3a) \cr &\hbox{if } {\bf b \cdot b} = {\bf c \cdot c}\quad \hbox{then}\quad {\bf a \cdot c} \leq {\bf a \cdot b} & (9.2.2.3b) \cr &\hbox{if } {\bf b \cdot c} = \textstyle{{1 \over 2}} {\bf b \cdot b}\quad\hbox{then}\quad{\bf a \cdot b} \leq 2{\bf a \cdot c}& (9.2.2.3c)\cr &\hbox{if } {\bf a \cdot c} = \textstyle{{1 \over 2}} {\bf a \cdot a}\quad\hbox{then}\quad{\bf a \cdot b} \leq 2{\bf b \cdot c}& (9.2.2.3d)\cr &\hbox{if } {\bf a \cdot b} = \textstyle{{1 \over 2}} {\bf a \cdot a}\quad\hbox{then}\quad {\bf a \cdot c} \leq 2{\bf b \cdot c.} & (9.2.2.3e)}]

  • (b) Type-II cell

    Main conditions: [\displaylines{\hfill \hbox{as } (9.2.2.2a) \hfill (9.2.2.4a)\cr \hfill (|{\bf b \cdot c}| + |{\bf a \cdot c}| + |{\bf a \cdot b}|) \leq {\textstyle{1 \over 2}} ({\bf a \cdot a} + {\bf b \cdot b}) \hfill (9.2.2.4b)\cr \hfill {\bf b \cdot c} \leq 0; \ \qquad {\bf a \cdot c} \leq 0; \ \qquad {\bf a \cdot b} \leq 0. \hfill (9.2.2.4c)}] [link]Special conditions: [\eqalignno{&\hbox{if }{\bf a \cdot a} = {\bf b \cdot b}\quad\hbox{then}\quad |{\bf b \cdot c}| \leq |{\bf a \cdot c}| &(9.2.2.5a)\cr & \hbox{if } {\bf b \cdot b} = {\bf c \cdot c} \quad \hbox{then}\quad |{\bf a \cdot c}| \leq |{\bf a \cdot b}| & (9.2.2.5b)\cr & \hbox{if } |{\bf b \cdot c}| = {\textstyle{1 \over 2}}{\bf b \cdot b}\quad \hbox{then}\quad {\bf a \cdot b} = 0 & (9.2.2.5c)\cr & \hbox{if } |{\bf a \cdot c}| = {\textstyle{1 \over 2}}{\bf a \cdot a}\quad \hbox{then}\quad{\bf a \cdot b} = 0 & (9.2.2.5d)\cr& \hbox{if } |{\bf a \cdot b}| = {\textstyle{1 \over 2}}{\bf a \cdot a}\quad \hbox{then}\quad{\bf a \cdot c}  = 0& (9.2.2.5e)\cr & \hbox{if } (|{\bf b \cdot c}| + |{\bf a \cdot c}| + |{\bf a \cdot b}|) = {\textstyle{1 \over 2}}({\bf a \cdot a} + {\bf b \cdot b})&\cr &\quad\hbox{then}\quad{\bf a \cdot a} \leq 2|{\bf a \cdot c}| + |{\bf a \cdot b}|.&(9.2.2.5f)\cr}]

The geometrical interpretation in the following sections is given in order to make the above conditions more explicit rather than to replace them, since the analytical form is obviously the most suitable one for actual verification.

9.2.3. Main conditions

| top | pdf |

The main conditions2 express the following two requirements:

  • (i) Of all lattice vectors, none is shorter than a; of those not directed along a, none is shorter than b; of those not lying in the ab plane, none is shorter than c. This requirement is expressed analytically by (9.2.2.2a)[link], and for type-II cells by (9.2.2.4b)[link], which for type-I cells is redundant.

  • (ii) The three angles between basis vectors are either all acute or all non-acute, conditions (9.2.2.2b)[link] and (9.2.2.4c)[link]. As shown in Section 9.2.2[link] for a given unit cell, the origin corner can always be chosen so as to satisfy either the first alternative of this condition (if the cell is of type I) or the second (if the cell is of type II).

Condition (i)[link] is by far the most essential one. It uniquely defines the lengths a, b and c, and limits the angles to the range [60 \leq \alpha, \beta, \gamma \leq 120^{\circ}]. However, there are often different unit cells satisfying (i),[link] cf. Gruber (1973)[link]. In order to find the reduced basis, starting from an arbitrary one given by its matrix (9.2.2.1)[link], one can: (a) find some basis satisfying (i)[link] and (ii)[link] and if necessary modify it so as to fulfil the special conditions as well; (b) find all bases satisfying (i)[link] and (ii)[link] and test them one by one with regard to the special conditions until the reduced form is found. Method (a) relies mainly on an algorithm by Buerger (1957,[link] 1960[link]), cf. also Mighell (1976)[link]. Method (b) stems from a theorem and an algorithm, both derived by Delaunay (1933)[link]; the theorem states that the desired basis vectors a, b and c are among seven (or fewer) vectors – the distance vectors between parallel faces of the Voronoi domain – which follow directly from the algorithm. The method has been established and an example is given by Delaunay et al. (1973)[link], cf. Section 9.1.3[link] where this method is described.

9.2.4. Special conditions

| top | pdf |

For a given lattice, the main condition (i)[link] defines not only the lengths a, b, c of the reduced basis vectors but also the plane containing a and b, in the sense that departures from special conditions can be repaired by transformations which do not change this plane. An exception can occur when [b = c]; then such transformations must be supplemented by interchange(s) of b and c whenever either (9.2.2.3b)[link] or (9.2.2.5b)[link] is not fulfilled. All the other conditions can be conveniently illustrated by projections of part of the lattice onto the ab plane as shown in Figs. 9.2.4.1[link][link][link][link] to 9.2.4.5[link]. Let us represent the vector lattice by a point lattice. In Fig. 9.2.4.1[link], the net in the ab plane (of which OBAD is a primitive mesh; [OA = a], [OB = b]) is shown as well as the projection (normal to that plane) of the adjoining layer which is assumed to lie above the paper. In general, just one lattice node [P_{0}] of that layer, projected in Fig. 9.2.4.1[link] as P, will be closer to the origin than all others. Then the vector [\overrightarrow{OP_{0}}] is [\pm {\bf c}] according to condition (i)[link]. It should be stressed that, though the ab plane is most often (see above) correctly established by (i)[link], the vectors a, b and c still have to be chosen so as to comply with (ii)[link], with the special conditions, and with right-handedness. The result will depend on the position of P with respect to the net. This dependence will now be investigated.

[Figure 9.2.4.1]

Figure 9.2.4.1| top | pdf |

The net of lattice points in the plane of the reduced basis vectors a and b; OBAD is a primitive mesh. The actual choice of a and b depends on the position of the point P, which is the projection of the point [P_{0}] in the next layer (supposed to lie above the paper, thin dashed lines) closest to O. Hence, P is confined to the Voronoi domain (dashed hexagon) around O. For a given interlayer distance, P defines the complete lattice. In that sense, P and [P'] represent identical lattices; so do Q, [Q'] and [Q''], and also R and [R']. When P lies in a region marked [-c^{\rm II}] (hatched), the reduced type-II basis is formed by [{\bf a}^{\rm II}], [{\bf b}^{\rm II}] and [{\bf c} = -\overrightarrow{OP}_{0}]. Regions marked [c^{\rm I}] (cross-hatched) have the reduced type-I basis [{\bf a}^{\rm I},{\bf b}^{\rm I}] and [{\bf c} = + \overrightarrow{OP}_{0}]. Small circles in O, M etc. indicate twofold rotation points lying on the region borders (see text).

[Figure 9.2.4.2]

Figure 9.2.4.2| top | pdf |

The effect of the special conditions. Border lines of type-I and type-II regions are drawn as heavy lines if included. The type-I and type-II regions are marked as in Fig. 9.2.4.1[link]. A heavy border line of a region stops short of an end point if the latter is not included in the region to which the border belongs. a, b net primitive orthogonal; special conditions (9.2.2.5c)[link], (9.2.2.5d)[link].

[Figure 9.2.4.3]

Figure 9.2.4.3| top | pdf |

The effect of the special conditions. Border lines of type-I and type-II regions are drawn as heavy lines if included. Type-I and type-II regions are marked as in Fig. 9.2.4.1[link]. [n_{b}] belongs to the type-II region. A heavy border line of a region stops short of an end point if the latter is not included in the region to which the border belongs. a, b net oblique; special conditions (9.2.2.3c)[link], (9.2.2.3d)[link], (9.2.2.5f)[link].

[Figure 9.2.4.4]

Figure 9.2.4.4| top | pdf |

The effect of the special conditions. Border lines of type-I and type-II regions are drawn as heavy lines if included. The type-I region is cross-hatched; the type-II region is a mere line. A heavy border line of a region stops short of an end point if the latter is not included in the region to which the border belongs. a, b net centred orthogonal (elongated); special conditions (9.2.2.3e)[link], (9.2.2.5e)[link].

[Figure 9.2.4.5]

Figure 9.2.4.5| top | pdf |

The effect of the special conditions. Border lines of type-I and type-II regions are drawn as heavy lines if included. Type-I and type-II regions are marked as in Fig. 9.2.4.1[link]. [n_{b}] belongs to the type-II region. A heavy border line of a region stops short of an end point if the latter is not included in the region to which the border belongs. a, b net centred orthogonal (compressed); special conditions (9.2.2.3a)[link], (9.2.2.5a)[link].

The inner hexagon shown, which is the two-dimensional Voronoi domain around O, limits the possible projected positions P of [P_{0}]. Its short edges, normal to OD, result from (9.2.2.4b)[link]; the other edges from (9.2.2.2a)[link]. If the spacing d between ab net planes is smaller than b, the region allowed for P is moreover limited inwardly by the circle around O with radius [(b^{2} - d^{2})^{1/2}], corresponding to the projection of points [P_{0}] for which [OP_{0} = c = b]. The case [c = b] has been dealt with, so in order to simplify the drawings we shall assume [d \;\gt\; b]. Then, for a given value of d, each point within the above-mentioned hexagonal domain, regarded as the projection of a lattice node [P_{0}] in the next layer, completely defines a lattice based on [\overrightarrow{OA}], [\overrightarrow{OB}] and [\overrightarrow{OP}_{0}]. Diametrically opposite points like P and [P'] represent the same lattice in two orientations differing by a rotation of 180° in the plane of the figure. Therefore, the systematics of reduced bases can be shown completely in just half the domain. As a halving line, the [n_{a}] normal to OA is chosen. This is an important boundary in view of condition (ii)[link], since it separates points P for which the angle between [OP_{0}] and OA is acute from those for which it is obtuse.

Similarly, [n_{b}], normal to OB, separates the sharp and obtuse values of the angles [P_{0} OB]. It follows that if P lies in the obtuse sector (cross-hatched area) between [n_{a}] and [n_{b}], the reduced cell is of type I, with basis vectors [{\bf a}^{\rm I}], [{\bf b}^{\rm I}], and [OP_{0} = + {\bf c}]. Otherwise (hatched area), we have a type-II reduced cell, with [OP_{0} = - {\bf c}] and [+ {\bf a}] and [+ {\bf b}] as shown by [{\bf a}^{\rm II}] and [{\bf b}^{\rm II}].

Since type II includes the case of right angles, the borders of this region on [n_{a}] and [n_{b}] are inclusive. Other borderline cases are points like R and [R'], separated by b and thus describing the same lattice. By condition (9.2.2.5c)[link] the reduced cell for such cases is excluded from type II (except for rectangular a, b nets, cf. Fig. 9.2.4.2[link]); so the projection of c points to R, not [R']. Accordingly, this part of the border is inclusive for the type-I region and exclusive (at [R']) for the type-II region as indicated in Fig. 9.2.4.3.[link] Similarly, (9.2.2.5d)[link] defines which part of the border normal to OA is inclusive.

The inclusive border is seen to end where it crosses OA, OB or OD. This is prescribed by the conditions (9.2.2.3d)[link], (9.2.2.3c)[link] and (9.2.2.5f)[link], respectively. The explanation is given in Fig. 9.2.4.1[link] for (9.2.2.3c)[link]: The points Q and [Q''] represent the same lattice because [Q'] (diametrically equivalent to Q as shown before) is separated from [Q''] by the vector b. Hence, the point M halfway between O and B is a twofold rotation point just like O. For a primitive orthogonal a, b net, only type II occurs according to (9.2.2.5c)[link] and (9.2.2.5d)[link], cf. Fig. 9.2.4.2.[link] A centred orthogonal a, b net of elongated character (shortest net vector in a symmetry direction, cf. Section 9.2.5[link]) is depicted in Fig. 9.2.4.4.[link] It yields type-I cells except when [\beta = 90^{\circ}] [condition (9.2.2.5c)[link]]. Moreover, (9.2.2.3c)[link] eliminates part of the type-I region as compared to Fig. 9.2.4.3.[link] Finally, a centred net with compressed character (shortest two net vectors equal in length) requires criteria allowing unambiguous designation of a and b. These are conditions (9.2.2.3a)[link] and (9.2.2.5a)[link], cf. Fig. 9.2.4.5.[link] The simplicity of these bisecting conditions, similar to those for the case [b = c] mentioned initially, is apparent from that figure when compared with Fig. 9.2.4.3.[link] This compressed type of centred orthogonal a, b net is limited by the case of a hexagonal net (where it merges with the elongated type, Fig. 9.2.4.4)[link] and by the centred quadratic net (where it merges with the primitive orthogonal net, Fig. 9.2.4.2[link]). In the limit of the hexagonal net, the triangle Ohh in Figs. 9.2.4.4[link] and 9.2.4.5[link] is all that remains, it is of type I except for the point O. For the quadratic net, only the type-II region in Fig. 9.2.4.5,[link] then a triangle with all edges inclusive, is left. It corresponds to the triangle Oqq in Fig. 9.2.4.2[link].

9.2.5. Lattice characters

| top | pdf |

Apart from being unique, the reduced cell has the further advantage of allowing a much finer differentiation between types of lattices than is given by the Bravais types. For two-dimensional lattices, this is apparent already in the last section where the centred orthogonal class is subdivided into nets with elongated character and those with compressed character, depending on whether the shortest net vector is, or is not, a symmetry direction. It is impossible to perform a continuous deformation – within the centred orthogonal type – of an elongated net into a compressed one, since one has to pass through either a hexagonal or a quadratic net.

In three dimensions, lattices are of the same character if, first, a continuous deformation of one into the other is possible without leaving the Bravais type. Secondly, it is required that all matrix elements of the reduced form (9.2.2.1)[link] change continuously during such a deformation. These criteria lead to 44 different lattice characters (Niggli, 1928[link]; Buerger, 1957[link]). Each of them can be recognized easily from the relations between the elements of the reduced form given in Table 9.2.5.1[link] [adapted from Table 5.1.3.1 in IT (1969)[link], which was recently improved by Mighell & Rodgers (1980)[link]]. The numbers in column 1 of this table are at the same time used as a general notation of the lattice characters themselves. We speak, for example, about the lattice character No. 7 (which is part of the Bravais type tI) etc.

Table 9.2.5.1| top | pdf |
The parameters [D = {\bf b}\cdot {\bf c}], [E = {\bf a}\cdot {\bf c}] and [F = {\bf a}\cdot {\bf b}] of the 44 lattice characters ([A = {\bf a}\cdot {\bf a},\ B = {\bf b}\cdot {\bf b},\ C = {\bf c}\cdot {\bf c}])

The character of a lattice given by its reduced form (9.2.2.1)[link] is the first one that agrees when the 44 entries are compared with that reduced form in the sequence given below (suggested by Gruber). Such a logical order is not always obeyed by the widely used character numbers (first column), which therefore show some reversals, e.g. 4 and 5.

No.TypeDEFLattice symmetryBravais typeTransformation to a conventional basis (cf. the last footnote to Table 9.2.5.2)
[A = B = C]
1I[A/2][A/2][A/2]CubiccF[1\bar{1}1/11\bar{1}/\bar{1}11]
2IDDDRhombohedralhR[1\bar{1}0/\bar{1}01/\bar{1}\bar{1}\bar{1}]
3II000CubiccP[100/010/001]
5II[- A/3][- A/3][- A/3]CubiccI[101/110/011]
4IIDDDRhombohedralhR[1\bar{1}0/\bar{1}01/\bar{1}\bar{1}\bar{1}]
6IID§DFTetragonaltI[011/101/110]
7IID§EETetragonaltI[101/110/011]
8IID§EFOrthorhombicoI[\bar{1}\bar{1}0/\bar{1}0\bar{1}/0\bar{1}\bar{1}]
[A = B], no conditions on C
9I[A/2][A/2][A/2]RhombohedralhR[100/\bar{1}10/\bar{1}\bar{1}3]
10IDDFMonoclinicmC[110/1\bar{1}0/00\bar{1}]
11II000TetragonaltP[100/010/001]
12II00[- A/2]HexagonalhP[100/010/001]
13II00FOrthorhombicoC[110/\bar{1}10/001]
15II[- A/2][- A/2]0TetragonaltI[100/010/112]
16IID§DFOrthorhombicoF[\bar{1}\bar{1}0/1\bar{1}0/112]
14IIDDFMonoclinicmC[110/\bar{1}10/001]
17IID§EFMonoclinicmC[1\bar{1}0/110/\bar{1}0\bar{1}]
[B = C], no conditions on A
18I[A/4][A/2][A/2]TetragonaltI[0\bar{1}1/1\bar{1}\bar{1}/100]
19IDA/2A/2OrthorhombicoI[\bar{1}00/0\bar{1}1/\bar{1}11]
20IDEEMonoclinicmC[011/01\bar{1}/\bar{1}00]
21II000TetragonaltP[010/001/100]
22II[- B/2]00HexagonalhP[010/001/100]
23IID00OrthorhombicoC[011/0\bar{1}1/100]
24IID§[- A/3][- A/3]RhombohedralhR[121/0\bar{1}1/100]
25IIDEEMonoclinicmC[011/0\bar{1}1/100]
No conditions on A, B, C
26I[A/4][A/2][A/2]OrthorhombicoF[100/\bar{1}20/\bar{1}02]
27ID[A/2][A/2]MonoclinicmC[\bar{1}20/\bar{1}00/0\bar{1}1]
28ID[A/2]2DMonoclinicmC[\bar{1}00/\bar{1}02/010]
29ID2D[A/2]MonoclinicmC[100/1\bar{2}0/00\bar{1}]
30I[B/2]E2EMonoclinicmC[010/01\bar{2}/\bar{1}00]
31IDEFTriclinicaP[100/010/001]
32II000OrthorhombicoP[100/010/001]
40II[- B/2]00OrthorhombicoC[0\bar{1}0/012/\bar{1}00]
35IID00MonoclinicmP[0\bar{1}0/\bar{1}00/00\bar{1}]
36II0[- A/2]0OrthorhombicoC[100/\bar{1}0\bar{2}/010]
33II0E0MonoclinicmP[100/010/001]
38II00[- A/2]OrthorhombicoC[\bar{1}00/120/00\bar{1}]
34II00FMonoclinicmP[\bar{1}00/00\bar{1}/0\bar{1}0]
42II[- B/2][- A/2]0OrthorhombicoI[\bar{1}00/0\bar{1}0/112]
41II[- B/2]E0MonoclinicmC[0\bar{1}\bar{2}/0\bar{1}0/\bar{1}00]
37IID[- A/2]0MonoclinicmC[102/100/010]
39IID0[- A/2]MonoclinicmC[\bar{1}\bar{2}0/\bar{1}00/00\bar{1}]
43IIDEFMonoclinicmI[\bar{1}00/\bar{1}\bar{1}\bar{2}/0\bar{1}0]
44IIDEFTriclinicaP[100/010/001]
[2|D + E + F| = A + B].
[2|D + E + F| = A + B] plus [|2D + F| = B].
§The symbols for Bravais lattices were adopted by the International Union of Crystallography in 1985; cf. de Wolff et al. (1985)[link]. The capital letter of the symbols in this column indicates the centring type of the cell as obtained by the transformation in the last column. For this reason, the standard symbols mS and oS are not used here.

In Table 9.2.5.2,[link] another description of lattice characters is given by grouping together all characters of a given Bravais type and by indicating for each character the corresponding interval of values of a suitable parameter p, expressed in the usual parameters of a conventional cell. In systems where no generally accepted convention exists, the choice of this cell has been made for convenience in the last column of this table.

Table 9.2.5.2| top | pdf |
Lattice characters described by relations between conventional cell parameters

Under each of the roman numerals below `Lattice characters in', numbers of characters (cf. Table 9.2.5.1,[link] first column) are listed for which the key parameter p lies in the interval defined by the same roman numeral below `Intervals of p'. For instance, a lattice with character No. 15 under IV has [p = c/a]; so it falls in the interval IV with [\surd 2 \lt c/a\ (\lt \infty)]; No. 33 under II has [p = b]; therefore the interval [a - c] for II yields the relation [a \lt b \lt c].

Lattice symmetryBravais typep = key parameterLattice characters inIntervals of pConventions‡‡
IIIIIIIV I II III IV 
TetragonaltP[c/a]21110 1    
TetragonaltI[c/a]1867150 [\surd (2/3)] 1 [\surd 2]  
HexagonalhP[c/a]22120 1    
RhombohedralhR[c/a]244290 [\surd (3/8)] [\surd (3/2)] [\surd 6] Hexagonal axes
OrthorhombicoP32 no relations    [a \lt b \lt c]
OrthorhombicoS               
[\quad b \lt a/\surd 3] c23130 d   [a \lt b]
[\quad b \gt a/\surd 3] c4036380 a d   
OrthorhombicoIr§819420 a b  [a \lt b \lt c]
OrthorhombicoF[b/a]16261 [\surd 3]   [a \lt b \lt c]
MonoclinicmPb3533340 a c  [a \lt c]††
MonoclinicmS               
Centred net [\Bigg\{][=1] [b/a][\cases{28\cr 29\cr}][\Bigg\}]0 [\surd(1/3)] 1 [\surd 3] C centred††
[=2] [b/a]30
[=2,3] [b/a][\cases{37\cr 41\cr}]2025
[=1,2,3] [b/a][\cases{27\cr 39\cr}][\cases{10\cr 17\cr}]14
[=3] [b/a]431    I centred††
TriclinicaP[\alpha, \beta, \gamma]314460° 90° 120°  [a \lt b \lt c]
CubiccP[\left.\matrix{3\cr 5\cr 1\cr}\right\}\hbox{ no relations}]     
cI     
cF     
The symbols for Bravais lattices were adopted by the International Union of Crystallography in 1985; cf. de Wolff et al. (1985)[link].
[d = {1 \over 2} \surd (a^{2} + b^{2})].
§[r = {1 \over 2} \surd (a^{2} + b^{2} + c^{2})].
This number specifies the centred net among the three orthogonal nets parallel to the twofold axis and passing through (1) the shortest, (2) the second shortest, and (3) the third shortest lattice vector perpendicular to the axis. For example, `2, 3' means that either net (2) or net (3) is the centred one.
††Setting with unique axis b; [\beta \gt 90^{\circ}]; [a \lt c] for both P and I cells, [a \lt c] or [a \gt c] for C cells.
‡‡These conventions refer to the cells obtained by the transformations of Table 9.2.5.1[link]. They have been chosen for convenience in this table.

The subdistinctions `centred [\hbox{net} = 1, 2] or 3' for the monoclinic centred type are closely related to the description in other conventions. For instance, they correspond to C-, A- or I-centred cells, respectively, if b is the unique axis and a and c are the shortest vectors [(a \;\lt\; c)] perpendicular to b; note that in Table 9.2.5.2 only C and I, not A, cells are listed. From the multiple entries in Table 9.2.5.2[link] for this type, it follows that the description in terms of b/a is not exhaustive; the distinctions depend upon rather intricate relations (cf. Mighell et al., 1975[link]; Mighell & Rodgers, 1980[link]).

No attempt has been made in Table 9.2.5.2[link] to specify whether the end points of p intervals are inclusive or not. For practical purposes, they can always be taken to be non-inclusive. Indeed, the end points correspond either to a different Bravais type or to a purely geometric singularity without physical significance. If p is very close to an interval limit of the latter kind, one should be aware of the fact that different measurements of such a lattice may yield different characters, with totally differing aspects of the reduced form.

9.2.6. Applications

| top | pdf |

9.2.6.1. Classification

| top | pdf |

The reduced basis can be used to derive the Bravais-lattice type and the conventional cell parameters, starting from an arbitrary description of the lattice. For this purpose, the reduced form is first derived from the given description, e.g. by means of the algorithm of Křivý & Gruber (1976)[link]. Subsequently it is compared with the reduced forms (Table 9.2.5.1)[link] for the 44 lattice characters and transformed to the appropriate conventional cell. Thus the reduced cell is helpful as an accessory in classifications based on conventional cells.

Alternatively, the parameters of the reduced form itself (either of the direct lattice or of the reciprocal lattice) can be used as a basis for determinative classification.

9.2.6.2. Comparison of lattices

| top | pdf |

Two lattices, defined by their reduced cells, can be compared on a rigorous basis to find out whether they are identical lattices or are related by one cell being a subcell of the other (Santoro et al., 1980[link]).

Further properties of lattices are discussed in Chapter 9.3[link] .

Acknowledgements

The author wishes to thank Dr B. Gruber (Prague) and Dr A. Santoro (Washington) for reading the manuscript and for suggesting several improvements as well as pointing out errors, especially in Tables 9.2.5.1[link] and 9.2.5.2[link].

References

First citation Buerger, M. J. (1957). Reduced cells. Z. Kristallogr. 109, 42–60.Google Scholar
First citation Buerger, M. J. (1960). Note on reduced cells. Z. Kristallogr. 113, 52–56.Google Scholar
First citation Delaunay, B. N. (1933). Neuere Darstellung der geometrischen Kristallographie. Z. Kristallogr. 84, 109–149.Google Scholar
First citation Delaunay, B. N., Galiulin, R. V., Dolbilin, N. P., Zalgaller, V. A. & Stogrin, K. I. (1973). On three successive minima of a three-dimensional lattice (in Russian). Dokl. Akad. Nauk SSSR, 209, 309–313.Google Scholar
First citation Eisenstein, G. (1851). Tabelle der reducierten positiven quadratischen Formen nebst den Resultaten neuer Forschungen über diese Formen, insbesondere Berücksichtigung auf ihre tabellarische Berechung. J. Math. (Crelle), 41, 141–190.Google Scholar
First citation Gruber, B. (1973). The relationship between reduced cells in a general Bravais lattice. Acta Cryst. A29, 433–440.Google Scholar
First citation Gruber, B. (1978). Comments on Bravais lattices. Research report, Charles University, Prague, Czech Republic. Details can be obtained from the author on request.Google Scholar
First citation Gruber, B. (1989). Reduced cells based on extremal principles. Acta Cryst. A45, 123–131.Google Scholar
First citation International Tables for X-ray Crystallography (1969). Vol. I, 3rd ed., edited by N. F. M. Henry & K. Lonsdale. Birmingham: Kynoch Press. [Abbreviated as IT (1969).]Google Scholar
First citation Křivý, I. & Gruber, B. (1976). A unified algorithm for determining the reduced (Niggli) cell. Acta Cryst. A32, 297–298.Google Scholar
First citation Mighell, A. D. (1976). The reduced cell: its use in the identification of crystalline materials. J. Appl. Cryst. 9, 491–498.Google Scholar
First citation Mighell, A. D. & Rodgers, J. R. (1980). Lattice symmetry determination. Acta Cryst. A36, 321–326.Google Scholar
First citation Mighell, A. D., Santoro, A. & Donnay, J. D. H. (1975). Addenda to International Tables for X-ray Crystallography. Acta Cryst. B31, 2942.Google Scholar
First citation Niggli, P. (1928). Kristallographische und strukturtheoretische Grundbegriffe. Handbuch der Experimentalphysik, Vol. 7, Part 1. Leipzig: Akademische Verlagsgesellschaft.Google Scholar
First citation Santoro, A., Mighell, A. D. & Rodgers, J. R. (1980). The determination of the relationship between derivative lattices. Acta Cryst. A36, 796–800.Google Scholar
First citation 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). Nomenclature for crystal families, Bravais-lattice types and arithmetic classes. Report of the International Union of Crystallography Ad hoc Committee on the Nomenclature of Symmetry. Acta Cryst. A41, 278–280.Google Scholar








































to end of page
to top of page