International
Tables for
Crystallography
Volume B
Reciprocal space
Edited by U. Shmueli

International Tables for Crystallography (2006). Vol. B. ch. 2.1, pp. 203-209   | 1 | 2 |

Section 2.1.8. Non-ideal distributions: the Fourier method

U. Shmuelia* and A. J. C. Wilsonb

a School of Chemistry, Tel Aviv University, Tel Aviv 69 978, Israel, and bSt John's College, Cambridge, England
Correspondence e-mail:  ushmueli@post.tau.ac.il

2.1.8. Non-ideal distributions: the Fourier method

| top | pdf |

The starting point of the method described in the previous section is the central-limit theorem approximation, and the method consists of finding correction factors which result in better approximations to the actual p.d.f. Conceptually, this is equivalent to improving the approximation of the characteristic function [cf. equation (2.1.4.10)[link]] over that which led to the central-limit theorem result.

The method to be described in this section does not depend on any initial approximation and will be shown to utilize the dependence of the exact value of the characteristic function on the space-group symmetry, atomic composition and other factors. This approach has its origin in a simple but ingenious observation by Barakat (1974[link]), who noted that if a random variable has lower and upper bounds then the corresponding p.d.f. can be non-zero only within these bounds and can therefore be expanded in an ordinary Fourier series and set to zero (identically) outside the bounded interval. Barakat's (1974[link]) work dealt with intensity statistics of laser speckle, where sinusoidal waves are involved, as in the present problem. This method was applied by Weiss & Kiefer (1983[link]) to testing the accuracy of a steepest-descents approximation to the exact solution of the problem of random walk, and its first application to crystallographic intensity statistics soon followed (Shmueli et al., 1984[link]). Crystallographic (e.g. Shmueli & Weiss, 1987[link]; Rabinovich et al., 1991a[link],b[link]) and noncrystallographic (Shmueli et al., 1985[link]; Shmueli & Weiss, 1985a[link]; Shmueli, Weiss & Wilson, 1989[link]; Shmueli et al., 1990[link]) symmetry was found to be tractable by this approach, as well as joint conditional p.d.f.'s of several structure factors (Shmueli & Weiss, 1985b[link], 1986[link]; Shmueli, Rabinovich & Weiss, 1989[link]). The Fourier method is illustrated below by deriving the exact counterparts of equations (2.1.7.5)[link] and (2.1.7.6)[link] and specifying them for some simple symmetries. We shall then indicate a method of treating higher symmetries and present results which will suffice for evaluation of Fourier p.d.f.'s of [|E|] for a wide range of space groups.

2.1.8.1. General representations of p.d.f.'s of [|E|] by Fourier series

| top | pdf |

We assume, as before, that (i) the atomic phase factors [\vartheta_{j} = 2\pi {\bf h}^{T}{\bf r}_{j}] [cf. equation (2.1.1.2)[link]] are uniformly distributed on (0–2[\pi]) and (ii) the atomic contributions to the structure factor are independent. For a centrosymmetric space group, with the origin chosen at a centre of symmetry, the random variable is the (real) normalized structure factor E and its bounds are [-E_{M}] and [E_{M}], where [E_{M} = {\textstyle\sum\limits_{j = 1}^{N}}n_{j}, \hbox{ with } n_{j} = {{f_{j}}\over{\left({\textstyle\sum_{k = 1}^{N}}\;f_{k}^{2}\right)^{1/2}}}. \eqno(2.1.8.1)] Here, [E_{M}] is the maximum possible value of E and [f_{j}] is the conventional scattering factor of the jth atom, including its temperature factor. The p.d.f., [p(E)], can be non-zero in the range ([-E_{M},E_{M}]) only and can thus be expanded in the Fourier series [p(E) = (\alpha/2)\textstyle\sum\limits_{k = -\infty}^{\infty}C_{k}\exp(-\pi ik\alpha E), \eqno(2.1.8.2)] where [\alpha = 1/E_{M}]. Only the real part of [p(E)] is relevant. The Fourier coefficients can be obtained in the conventional manner by integrating over the range ([-E_{M},E_{M}]),[C_{k} = \textstyle\int\limits_{-E_{M}}^{E_{M}}p(E)\exp(\pi ik\alpha E)\;{\rm d}E. \eqno(2.1.8.3)] Since, however, [p(E) = 0] for [E \;\lt\; -E_{M}] and [E \;\gt\; E_{M}], it is possible and convenient to replace the limits of integration in equation (2.1.8.3)[link] by infinity. Thus [C_{k} = \textstyle\int\limits_{-\infty}^{\infty}p(E)\exp(\pi ik\alpha E)\;{\rm d}E = \langle \exp(\pi ik\alpha E) \rangle. \eqno(2.1.8.4)] Equation (2.1.8.4)[link] shows that [C_{k}] is a Fourier transform of the p.d.f. [p(E)] and, as such, it is the value of the corresponding characteristic function at the point [t_{k} = \pi\alpha k] [i.e., [C_{k} = C(\pi\alpha k)], where the characteristic function [C(t)] is defined by equation (2.1.4.1)[link]]. It is also seen that [C_{k}] is the expected value of the exponential [\exp(\pi ik\alpha E)]. It follows that the feasibility of the present approach depends on one's ability to evaluate the characteristic function in closed form without the knowledge of the p.d.f.; this is analogous to the problem of evaluating absolute moments of the structure factor for the correction-factor approach, discussed in Section 2.1.7[link]. Fortunately, in crystallographic applications these calculations are feasible, provided individual isotropic motion is assumed. The formal expression for the p.d.f. of [|E|], for any centrosymmetric space group, is therefore [p(|E|) = \alpha\left[1+2\textstyle\sum\limits_{k = 1}^{\infty}C_{k}\cos(\pi k\alpha |E|) \right], \eqno(2.1.8.5)] where use is made of the assumption that [p(E) = p(-E)], and the Fourier coefficients are evaluated from equation (2.1.8.4)[link].

The p.d.f. of [|E|] for a noncentrosymmetric space group is obtained by first deriving the joint p.d.f. of the real and imaginary parts of E and then integrating out its phase. The general expression for E is [E = A + iB = |E|\cos\varphi + i|E|\sin\varphi, \eqno(2.1.8.6)] where [\varphi] is the phase of E. The required joint p.d.f. is [p(A,B) = (\alpha^{2}/4)\textstyle\sum\limits_{m}\textstyle\sum\limits_{n}C_{mn}\exp[-\pi i\alpha(mA+nB)], \eqno(2.1.8.7)] and introducing polar coordinates [m = r\sin\Delta] and [n = r\cos\Delta], where [r = \sqrt{m^{2}+n^{2}}] and [\Delta = \tan^{-1}(m/n)], we have [\eqalignno{p(|E|,\varphi) &= (\alpha^{2}/4)|E|\textstyle\sum\limits_{m}\textstyle\sum\limits_{n}C_{mn}\exp[-\pi i \alpha |E| & \cr &\quad\times{}\sqrt{m^{2}+n^{2}}\sin(\varphi + \Delta)]. &(2.1.8.8) \cr}] Integrating out the phase [\varphi], we obtain [p(|E|) = (\pi\alpha^{2} |E|/2)\textstyle\sum\limits_{m}\textstyle\sum\limits_{n}C_{mn}J_{0}(\pi\alpha |E|\sqrt{m^{2}+n^{2}}), \eqno(2.1.8.9)] where [J_{0}(x)] is the Bessel function of the first kind (e.g. Abramowitz & Stegun, 1972[link]). This is a general form of the p.d.f. of [|E|] for a noncentrosymmetric space group. The Fourier coefficients are obtained, similarly to the above, as [C_{mn} = \langle \exp[\pi i\alpha(mA+nB)] \rangle \eqno(2.1.8.10)] and the average in equation (2.1.8.10)[link], just as that in equation (2.1.8.4)[link], is evaluated in terms of integrals over the appropriate trigonometric structure factors. In terms of the characteristic function for a joint p.d.f. of A and B, the Fourier coefficient in equation (2.1.8.10)[link] is given by [C_{mn} = C(\pi\alpha m,\pi\alpha n)].

We shall denote the characteristic function by [C(t_{1})] if it corresponds to a Fourier coefficient of a Fourier series for a centrosymmetric space group and by [C(t_{1},t_{2})] or by [C(t,\Delta)], where [t = (t_{1}^{2}+t_{2}^{2})^ {1/2}] and [\Delta = \tan^{-1}(t_{1}/t_{2})], if it corresponds to a Fourier series for a noncentrosymmetric space group.

2.1.8.2. Fourier–Bessel series

| top | pdf |

Equations (2.1.8.5)[link] and (2.1.8.9)[link] are the exact counterparts of equations (2.1.7.5)[link] and (2.1.7.6)[link], respectively. The computational effort required to evaluate equation (2.1.8.9)[link] is somewhat greater than that for (2.1.8.5)[link], because a double Fourier series has to be summed. The p.d.f. for any noncentrosymmetric space group can be expressed by a double Fourier series, but this can be simplified if the characteristic function depends on [t = (t_{1}^{2}+t_{2}^{2})^{1/2}] alone, rather than on [t_{1}] and [t_{2}] separately. In such cases the p.d.f. of [|E|] for a noncentrosymmetric space group can be expanded in a single Fourier–Bessel series (Barakat, 1974[link]; Weiss & Kiefer, 1983[link]; Shmueli et al., 1984[link]). The general form of this expansion is [p(|E|) = 2\alpha^{2}|E|\textstyle\sum\limits_{u = 1}^{\infty}D_{u}J_{0}(\alpha\lambda_{u}|E|), \eqno(2.1.8.11)] where [D_{u} = {{C(\alpha\lambda_{u})}\over{J_{1}^{2}(\lambda_{u})}} \eqno(2.1.8.12)] and [C(\alpha\lambda_{u}) = \textstyle\prod\limits_{j = 1}^{N/g}C_{ju}, \eqno(2.1.8.13)] where [J_{1}(x)] is the Bessel function of the first kind, and [\lambda_{u}] is the uth root of the equation [J_{0}(x) = 0]; the atomic contribution [C_{ju}] to equation (2.1.8.13)[link] is computed as [C_{ju} = C(\alpha n_{j}\lambda_{u}). \eqno(2.1.8.14)] The roots [\lambda_{u}] are tabulated in the literature (e.g. Abramowitz & Stegun, 1972[link]), but can be most conveniently computed as follows. The first five roots are given by [\eqalign{ \lambda_1& = 2.4048255577\cr \lambda_2& = 5.5200781103 \cr \lambda_3& = 8.6537279129\cr \lambda_4& = 11.7915344390\cr \lambda_5& = 14.9309177085 }] and the higher ones can be obtained from McMahon's approximation (cf. Abramowitz & Stegun, 1972[link]) [{\lambda_{u} = \beta+{{1}\over{8\beta}}-{{124}\over{3(8\beta)^{3}}}+{{120928}\over{15(8\lambda)^{5}}}-{{401743168}\over{105(8\lambda)^{7}}}+\ldots,} \eqno(2.1.8.15)] where [\beta = (u-{\textstyle {1\over4}})\pi]. For [u \;\gt\; 5] the values given by equation (2.1.8.15)[link] have a relative error less than 10−11 so that no refinement of roots of higher orders is needed (Shmueli et al., 1984[link]). Numerical computations of single Fourier–Bessel series are of course faster than those of the double Fourier series, but both representations converge fairly rapidly.

2.1.8.3. Simple examples

| top | pdf |

Consider the Fourier coefficient of the p.d.f. of [|E|] for the centrosymmetric space group [P\bar{1}]. The normalized structure factor is given by [E = 2\textstyle\sum\limits_{j = 1}^{N/2}n_{j}\cos\vartheta_{j}, \quad{\rm with}\quad \vartheta_{j} = 2\pi {\bf h}^{T}\cdot{\bf r}_{j}, \eqno(2.1.8.16)] and the Fourier coefficient is [\eqalignno{C_{k} & = \langle \exp(\pi ik\alpha E) \rangle &(2.1.8.17) \cr& = \left\langle \exp\left[2\pi ik\alpha\textstyle\sum\limits_{j = 1}^{N/2}n_{j}\cos\vartheta_{j}\right] \right\rangle &(2.1.8.18) \cr & = \left\langle \textstyle\prod\limits_{j = 1}^{N/2} \exp(2\pi ik\alpha n_{j}\cos\vartheta_{j}) \right\rangle &(2.1.8.19) \cr & = \textstyle \prod\limits_{j = 1}^{N/2} \langle \exp(2\pi ik\alpha n_{j}\cos\vartheta_{j}) \rangle &(2.1.8.20) \cr & = \textstyle\prod\limits_{j = 1}^{N/2}\left\{{({1}/{2\pi})}\int\limits_{-\pi}^{\pi}\exp(2\pi ik\alpha n_{j}\cos\vartheta)\;{\rm d}\vartheta \right\}\cr &&(2.1.8.21) \cr & = \textstyle\prod\limits_{j = 1}^{N/2}J_{0}(2\pi k\alpha n_{j}). &(2.1.8.22)}%2.1.8.22] Equation (2.1.8.20)[link] is obtained from equation (2.1.8.19)[link] if we make use of the assumption of independence, the assumption of uniformity allows us to rewrite equation (2.1.8.20)[link] as (2.1.8.21)[link], and the expression in the braces in the latter equation is just a definition of the Bessel function [J_{0}(2\pi k\alpha n_{j})] (e.g. Abramowitz & Stegun, 1972[link]).

Let us now consider the Fourier coefficient of the p.d.f. of [|E|] for the noncentrosymmetric space group [P1]. We have [A = \textstyle\sum\limits_{j = 1}^{N}n_{j}\cos\vartheta_{j} \quad {\rm and}\quad B = \textstyle\sum\limits_{j = 1}^{N}n_{j}\sin\vartheta_{j}. \eqno(2.1.8.23)] These expressions for A and B are substituted in equation (2.1.8.10)[link], resulting in [\eqalignno{C_{mn} & = \left\langle \textstyle\prod\limits_{j = 1}^{N} \exp[\pi i\alpha n_{j}(m\cos\vartheta_{j} + n\sin\vartheta_{j})] \right\rangle & \cr & & (2.1.8.24) \cr & = \left\langle \textstyle\prod\limits_{j = 1}^{N} \exp[\pi i\alpha n_{j}\sqrt{m^{2}+n^{2}} \sin(\vartheta_{j}+\Delta)] \right\rangle & \cr & & (2.1.8.25)\cr &= \textstyle \prod\limits_{j = 1}^{N} J_{0}(\pi\alpha n_{j}\sqrt{m^{2}+n^{2}}). &(2.1.8.26)}] Equation (2.1.8.24)[link] leads to (2.1.8.25)[link] by introducing polar coordinates analogous to those leading to equation (2.1.8.8)[link], and equation (2.1.8.26)[link] is then obtained by making use of the assumptions of independence and uniformity in an analogous manner to that detailed in equations (2.1.8.12)[link]–(2.1.8.22)[link] above.

The right-hand side of equation (2.1.8.26)[link] is to be used as a Fourier coefficient of the double Fourier series given by (2.1.8.9)[link]. Since, however, this coefficient depends on [(m^{2}+n^{2})^{1/2}] alone rather than on m and n separately, the p.d.f. of [|E|] for P1 can also be represented by a Fourier–Bessel series [cf. equation (2.1.8.11)[link]] with coefficient [D_{u} = {{1}\over{J_{1}^{2}(\lambda_{u})}}\prod_{j = 1}^{N}J_{0}(\alpha n_{j}\lambda_ {u}), \eqno(2.1.8.27)] where [\lambda_{u}] is the uth root of the equation [J_{0}(x) = 0].

2.1.8.4. A more complicated example

| top | pdf |

We now illustrate the methodology of deriving characteristic functions for space groups of higher symmetries, following the method of Rabinovich et al. (1991a[link],b[link]). The derivation is performed for the space group P [\bar{6}] [No. 174]. According to Table A1.4.3.6[link] , the real and imaginary parts of the normalized structure factor are given by [\eqalignno{A & = 2\textstyle\sum\limits_{j = 1}^{N/6}n_{j}[C(hki)c(lz)]_{j} \cr & = 2\textstyle\sum\limits_{j = 1}^{N/6}n_{j}\cos\tau_{j}\textstyle\sum\limits_{k = 1}^{3}\cos\alpha_{jk} &(2.1.8.28)}] and [\eqalignno{B & = 2\textstyle\sum\limits_{j = 1}^{N/6}n_{j}[S(hki)c(lz)]_{j} \cr & = 2\textstyle\sum\limits_{j = 1}^{N/6}n_{j}\cos\tau_{j}\textstyle\sum\limits_{k = 1}^{3}\sin\alpha_{jk}, &(2.1.8.29)}] where[\eqalign{ \alpha_{j1}& = 2\pi (hx_j+ky_j),\cr \alpha_{j2}& = 2\pi (kx_j+iy_j),\cr \alpha_{j3}& = 2\pi (ix_j+hy_j),\cr \tau_j& = 2\pi lz_j. }] Note that [\alpha_{j1}+\alpha_{j2}+\alpha_{j3} = 0], i.e., one of these contributions depends on the other two; this is a recurring problem in calculations pertaining to trigonal and hexagonal systems. For brevity, we write directly the general form of the characteristic function from which the functional form of the Fourier coefficient can be readily obtained. The characteristic function is given by [\eqalignno{C(t_{1},t_{2})& = \langle \exp[i(t_{1}A+t_{2}B)] \rangle &(2.1.8.30) \cr & = \textstyle\prod\limits_{j = 1}^{N/6}\left\langle \exp\left[2in_{j}\cos\tau_{j}\textstyle\sum\limits_{k = 1}^{3}(t_{1} \cos\alpha_{jk}+t_{2}\sin\alpha_{jk})\right]\right\rangle & \cr & & (2.1.8.31) \cr & = \textstyle \prod\limits_{j = 1}^{N/6}\biggl\langle \exp\biggl[2in_{j}t\cos\tau_{j}\textstyle\sum\limits_{k = 1}^{3}(\sin \Delta\cos\alpha_{jk} & \cr & \quad+ \cos\Delta\sin\alpha_{jk})\biggr]\biggr\rangle & (2.1.8.32) \cr & = \textstyle\prod\limits_{j = 1}^{N/6}\left\langle \exp\left[2in_{j}t\cos\tau_{j}\textstyle\sum\limits_{k = 1}^{3} \sin(\alpha_{jk}+\Delta)\right]\right\rangle , & \cr & &(2.1.8.33)}%2.1.8.33] where [\Delta = \tan^{-1}(t_{1}/t_{2})], [t = ({t_{1}^{2}+t_{2}^{2}}){}^{1/2}] and the assumption of independence was used. If we further employ the assumption of uniformity, while remembering that the angular variables [\alpha_{jk}] are not independent, the characteristic function can be written as [\eqalignno{C(t_{1},t_{2}) & = \textstyle\prod\limits_{j = 1}^{N/6}\left\{(1/2\pi)\textstyle\int\limits_{-\pi}^{\pi}\;{\rm d}\tau \biggl([1/(2\pi)^{2}] \right. \cr &\quad \times\textstyle\int\limits_{-\pi}^{\pi}\textstyle\int\limits_{-\pi}^{\pi}\textstyle\int\limits_{-\pi}^{\pi} {\rm d}\alpha_{1}\;{\rm d}\alpha_{2}\;{\rm d}\alpha_{3}\delta_{2\pi}(\alpha_{1}+\alpha_{2}+ \alpha_{3}) \cr &\quad\times \left.\left. \exp\left[2in_{j}t\cos\tau\textstyle\sum\limits_{k = 1}^{3}\sin(\alpha_{k} +\Delta)\right] \right) \right\}, & \cr & &(2.1.8.34)}] where [\delta_{2\pi}(\alpha) = {{1}\over{2\pi}}\sum_{k = -\infty}^{\infty}\exp(-ik\alpha) \eqno(2.1.8.35)] is the Fourier representation of the periodic delta function. Equation (2.1.8.34)[link] then becomes [\eqalignno{C(t_{1},t_{2}) &= \textstyle\prod\limits_{j = 1}^{N/6}\biggl\{(1/2\pi)\textstyle\int\limits_{-\pi}^{\pi}\;{\rm d}\tau \textstyle\sum\limits_{k = -\infty}^{\infty}\biggl[(1/2\pi) & \cr &\quad\times \textstyle\int\limits_{-\pi}^{\pi}\exp\left(-ik \alpha+2in_{j}t\cos\tau\sin(\alpha+\Delta)\right)\;{\rm d}\alpha\biggr]^{3}\biggr\}. & \cr & &(2.1.8.36) \cr}] If we change the variable [\alpha] to [\alpha'-\Delta], [\sin(\alpha+\Delta)] becomes [\sin\alpha'] and [-ik\alpha = -ik\alpha'+ik\Delta]. Hence [C(t_{1},t_{2}) = \textstyle\prod\limits_{j = 1}^{N/6}\biggl\{(1/2\pi)\textstyle\int\limits_{-\pi}^{\pi}\;{\rm d}\tau\textstyle\sum\limits_{k = -\infty}^{\infty}\exp(3ik\Delta) J_{k}^{3}(2n_{j}t\cos\tau)\biggr\}. \eqno(2.1.8.37)] The imaginary part of the summation, involving Bessel functions of odd orders, vanishes upon integration and the latter is restricted to the positive quadrant in [\tau]. Thus, upon replacing cosines by sines (this is permissible at this stage) the atomic contribution to the characteristic function becomes [\eqalignno{C_{j}(t,\Delta) &= (2/\pi)\textstyle\int\limits_{0}^{\pi/2}\biggl[J_{0}^{3}(2n_{j}t \sin\tau) & \cr &\quad + 2\textstyle\sum\limits_{k = 1}^{\infty}\cos(6k\Delta)J_{k}^{3}(2n_{j}t\sin\tau) \biggr]\;{\rm d}\tau & \cr & & (2.1.8.38) \cr}] and a double Fourier series must be used for the p.d.f.

2.1.8.5. Atomic characteristic functions

| top | pdf |

Expressions for the atomic contributions to the characteristic functions were obtained by Rabinovich et al. (1991a[link]) for a wide range of space groups, by methods similar to those described above. These expressions are collected in Table 2.1.8.1[link] in terms of symbols which are defined below. The following abbreviations are used in the subsequent definitions of the symbols: [\eqalign{s_{\pm} &= 2an_{j}\sin(\tau\pm\rho), \cr c_{\pm} &= 2an_{j}\cos(\tau\pm \rho)\quad {\rm and}\cr \sigma_{\pm} &= 2an_{j}\sin(\tau\pm 2\pi/3 + \rho),} ] and the symbols appearing in Table 2.1.8.1[link] are given below: [\eqalign{ {}^{(a)}L_j(a,\rho) & = \langle J_0(s_+)J_0(s_-) \rangle_\tau \cr & = \textstyle\sum\limits_{k = -\infty}^\infty\cos(4k\rho)J_k^4 \cr & = J_0^4(an_j) + 2\textstyle\sum\limits_{k = 1}^\infty\cos(4k\rho)J_k^4(an_j),\cr {}^{(b)}Q_j^{(1)} (a,\rho) & = \langle J_0^2(s_+)J_0^2(s_-) \rangle_\tau,\cr {}^{(c)}Q_j^{(2)} (a,\rho) & = \langle J_0(s_+)J_0(s_-)J_0(c_+)J_0 (c_-) \rangle_\tau,\cr {}^{(d)}T_j(a,\rho) & = \textstyle\sum\limits_{k = -\infty}^\infty\exp(6ik\rho)J_k^6(an_j) \cr & = J_0^6(an_j)+2\textstyle\sum\limits_{k = 1}^\infty\cos(6k\rho)J_k^6(an_j),\cr {}^{(e)}H_j^{(1)} (a,\mu) & = \left\langle {\cal R} \left [S_j^{(1)} (\tau\hbox{; } a,\mu,0) \right] \right\rangle _\tau, \cr {}^{(f)}H_j^{(2)} (a,\mu) & = \left\langle {\cal R} \left [S_j^{(2)} (\tau\hbox{; } a,\mu,0) \right] \right\rangle _\tau,\cr {}^{(g)}\tilde{H}_j^{(1)} (a,\mu_1,\mu_2,\rho) & = \biggl\langle {{\cal R}} \biggl[S_j^{(1)} (\tau\hbox{; }a,\mu_1,\rho) \cr &\quad\times{}S_j^{(1)} (\tau\hbox{; }a,\mu_2,-\rho) \biggr] \biggr\rangle _\tau, \cr {}^{(h)}\tilde{H}_j^{(2)} (a,\mu_1,\mu_2,\rho) & = \biggl\langle {\cal R} \biggl[S_j^{(2)} (\tau\hbox{; }a,\mu_1,\rho) \cr &\quad\times{}S_j^{(2)} (\tau\hbox{; }a,\mu_2,-\rho) \biggr] \biggr\rangle _\tau, } ] where [S_{j}^{(1)}(\tau\hbox{; }a,\mu,\rho) = \textstyle\sum\limits_{k = -\infty}^{\infty}e^{3ik\mu}J_{k}^{3} (s_{+})] and [S_{j}^{(2)}(\tau\hbox{; }a,\mu,\rho) = \textstyle\sum\limits_{k = -\infty}^{\infty}e^{3ik\mu}J_{k} (s_{+})J_{k}(\sigma_{+})J_{k}(\sigma_{-}).]

Table 2.1.8.1 | top | pdf |
Atomic contributions to characteristic functions for [p(|E|)]

The table lists symbolic expressions for the atomic contributions to exact characteristic functions (abbreviated as c.f.) for [p(|E|)], to be computed as single Fourier series (centric), double Fourier series (acentric) and single Fourier–Bessel series (acentric), as defined in Sections 2.1.8.1[link] and 2.1.8.2[link]. The symbolic expressions are defined in Section 2.1.8.5[link]. The table is arranged by point groups, space groups and parities of the reflection indices analogously to the table of moments, Table 2.1.7.1[link], and covers all the space groups and statistically different parities of hkl up to and including space group Fd [\bar{3}]. The expressions are valid for atoms in general positions, for general reflections and presume the absence of noncrystallographic symmetry and of dispersive scatterers.

Space group(s) g Atomic c.f. Remarks
Point group: 1      
P 1 1 [ J_{0}(t n_{j})]  
Point group: [\bar{\bf 1}]      
P [\bar{1}] 2 [J_{0}(2t_{1}n_{j})]  
Point groups: 2, m      
All P 2 [J_{0}^{2}(t n_{j})]  
All C 4 [J_{0}^{2}(2t n_{j})]  
Point group: [{\bf 2/}{\bi m}]      
All P 4 [J_{0}^{2}(2t_{1} n_{j})]  
All C 8 [J_{0}^{2}(4t_{1} n_{j})]  
Point group: 222      
All P 4 [ L_{j}(t,\Delta)^{(a)}]  
All C and I 8 [ L_{j}(2t,\Delta)]  
F 222 16 [ L_{j}(4t,\Delta)]  
Point group: mm2      
All P 4 [ L_{j}(t,0)]  
All C and I 8 [ L_{j}(2t,0)]  
Fmm 2 16 [ L_{j}(4t,0)]  
Fdd 2 16 [L_{j}(4t,0)] [ h+k+l=2n]
  16 [ L_{j}(4t,\pi/4)] [ h+k+l=2n+1]
Point group: mmm      
All P 8 [ L_{j}(2t_{1},0)]  
All C and I 16 [ L_{j}(4t_{1},0)]  
Fmmm 32 [ L_{j}(8t_{1},0)]  
Fddd 32 [ L_{j}(8t_{1},0)] [ h+k+l=2n]
  32 [ L_{j}(8t_{1},\pi/4)] [ h+k+l=2n+1]
Point group: 4      
P 4, P42 4 [ L_{j}(t,0)]  
[P4_1] 4 [ L_{j}(t,0)] [ l=2n]
  4 [ L_{j}(t,\pi/4)] [ l=2n+1]
I 4 8 [ L_{j}(2t,0)]  
[ I4_{1}] 8 [ L_{j}(2t,0)] [2h+l=2n]
  8 [ L_{j}(2t,\pi/4)] [ 2h+l=2n+1]
Point group: [\bar{\bf 4}]      
P [\bar{4}] 4 [L_{j}(t,\Delta)]  
I [\bar{4}] 8 [ L_{j}(2t,\Delta)]  
Point group: [{\bf 4/}{\bi m}]      
All P 8 [ L_{j}(2t_{1},0)]  
[ I4/m] 16 [ L_{j}(4t_{1},0)]  
[ I4_1/a] 16 [ L_{j}(4t_{1},0)] [ l=2n]
  16 [L_{j}(4t_{1},\pi/4)] [ l=2n+1]
Point group: 422      
P 422, P4212, P4222, P42212 8 [ Q_{j}^{(1)}(t,\Delta)^{(b)}]  
P 4122, P41212 8 [ Q_{j}^{(1)}(t,\Delta)] [ l=2n]
  8 [ Q_{j}^{(2)}(t,\Delta)^{(c)}] [ l=2n+1]
I 422 16 [ Q_{j}^{(1)}(2t,\Delta)]  
I 4122 16 [ Q_{j}^{(1)}(2t,\Delta)] [ 2k+l=2n]
  16 [ Q_{j}^{(2)}(2t,\Delta)] [ 2k+l=2n+1]
Point group: 4mm      
All P 8 [ Q_{j}^{(1)}(t,0)]  
I 4mm, I4cm 16 [ Q_{j}^{(1)}(2t,0)]  
I 41 md, I41 cd 16 [Q_{j}^{(1)}(2t,0)] [2k+l=2n]
  16 [ Q_{j}^{(1)}(2t,\pi/4)] [2k+l=2n+1]
Point groups: [\bar{\bf 4}]2m, [\bar{\bf 4}]m2      
All P 8 [ Q_{j}^{(1)}(t,\Delta)]  
I [\bar{4}]2m, I [\bar{4}] m2, I [\bar{4}] c2 16 [Q_{j}^{(1)}(2t,\Delta)]  
I [\bar{4}]2d 16 [Q_{j}^{(1)}(2t,\Delta)] [2h+l=2n]
  16 [Q_{j}^{(2)}(2t,\Delta)] [2h+l=2n+1]
Point group: [{\bf 4/}{\bi mmm}]      
All P 16 [ Q_{j}^{(1)}(2t_{1},0)]  
[I4/mmm], [I4/mcm] 32 [Q_{j}^{(1)}(4t_{1},0)]  
[ I4_{1}/amd], [I4_{1}/acd] 32 [Q_{j}^{(1)}(4t_{1},0)] [ l=2n]
  32 [ Q_{j}^{(1)}(4t_{1},\pi/4)] [ l=2n+1]
Point group: 3      
All P and R 3 [ J_{0}^{3}(t n_{j})]  
Point group: [\bar{\bf 3}]      
All P and R 6 [ J_{0}^{3}(2t_{1} n_{j})]  
Point group: 32      
All P and R 6 [T_{j}(t,\Delta)^{(d)}]  
Point group: 3m      
P 3m1, P31m, R3m 6 [ T_{j}(t,\pi/2)]  
P 3c1, P31c, R3c 6 [ T_{j}(t,\pi/2)] [ l=2n\;(P)], [h+k+l=2n\;(R)]
  6 [ T_{j}(t,0)] [ l=2n+1\;(P)], [h+k+l =2n+1\;(R)]
Point group: [\bar{\bf 3}] m      
P [\bar{3}] m1, P [\bar{3}]1m, R [\bar{3}] m 12 [ T_{j}(2t_{1},\pi/2)]  
P [\bar{3}] c1, P [\bar{3}]1c, R [\bar{3}] c 12 [ T_{j}(2t_{1},\pi/2)] [ l=2n\;(P)], [h+k+l=2n\;(R)]
  12 [ T_{j}(2t_{1},0)] [ l=2n+1\;(P)], [h+k+l=2n+1\;(R)]
Point group: 6      
P 6 6 [H_{j}^{(1)}(t, \pi/2)^{(e)}]  
[ P6_{1}] 6 [H_{j}^{(1)}(t, \pi/2)] [ l=6n]
  6 [ H_{j}^{(2)}(t, 0)^{(f)}] [ l=6n+1], [6n+5]
  6 [H_{j}^{(2)}(t, \pi/2)] [l=6n+2], [6n+4]
  6 [ H_{j}^{(1)}(t, 0)] [ l=6n+3]
[ P6_{2}] 6 [H_{j}^{(1)}(t, \pi/2)] [ l=3n]
  6 [H_{j}^{(2)}(t, \pi/2)] [l=3n \pm 1]
[ P6_{3}] 6 [ H_{j}^{(1)}(t, \pi/2)] [ l=2n]
  6 [ H_{j}^{(1)}(t, 0)] [ l=2n+1]
Point group: [\bar{\bf 6}]      
[P\bar{6}] 6 [H_{j}^{(1)}(t, \Delta)]  
Point group: [{\bf 6/}{\bi m}]      
[ P6/m] 12 [H_{j}^{(1)}(2t_{1}, \pi/2)]  
[ P6_{3}/m] 12 [ H_{j}^{(1)}(2t_{1}, \pi/2)] [ l=2n]
  12 [ H_{j}^{(1)}(2t_{1}, 0)] [l=2n+1]
Point group: 622      
P 622 12 [\tilde{H}_{j}^{(1)} (t,\pi/2], [-\pi/2,\Delta)^{(g)}]  
[ P6_{1}22] 12 [ \tilde{H}_{j}^{(1)}(t,\pi/2], [-\pi/2,\Delta)] [l=6n]
  12 [ \tilde{H}_{j}^{(2)}(t,0,0,\Delta)^{(h)}] [ l=6n+1], [6n+5]
  12 [ \tilde{H}_{j}^{(2)}(t,\pi/2,\pi/2,\Delta)] [ l=6n+2], [6n+4]
  12 [ \tilde{H}_{j}^{(1)}(t,0,0,\Delta)] [ l=6n+3]
[ P6_{2}22] 12 [ \tilde{H}_{j}^{(1)}(t,\pi/2], [-\pi/2,\Delta)] [ l=3n]
  12 [ \tilde{H}_{j}^{(2)}(t,\pi/2,\pi/2,\Delta)] [ l=3n \pm 1]
[ P6_{3}22] 12 [ \tilde{H}_{j}^{(1)}(t,\pi/2], [-\pi/2,\Delta)] [ l=2n]
  12 [ \tilde{H}_{j}^{(1)}(t,0,0,\Delta)] [ l=2n+1]
Point group: 6mm      
P 6mm 12 [ \tilde{H}_{j}^{(1)}(t,\pi/2,\pi/2,0)]  
P 6cc 12 [ \tilde{H}_{j}^{(1)}(t,\pi/2,\pi/2,0)] [ l=2n]
  12 [ \tilde{H}_{j}^{(1)}(t,\pi/2], [-\pi/2,0)] [ l=2n+1]
[P6_{3}cm], [P6_{3}mc] 12 [ \tilde{H}_{j}^{(1)}(t,\pi/2,\pi/2,0)] [l=2n]
  12 [ \tilde{H}_{j}^{(1)}(t,0,0,0)] [l=2n+1]
Point groups: [\bar{\bf 6}]2m, [\bar{\bf 6}]m2      
P [\bar{6}]2m, P[\bar{6}]m2 12 [ \tilde{H}_{j}^{(1)}(t,\Delta,\Delta,0)]  
P [\bar{6}]2c, P[\bar{6}]c2 12 [ \tilde{H}_{j}^{(1)}(t,\Delta,\Delta,0)] [ l=2n]
  12 [ \tilde{H}_{j}^{(1)}(t,\Delta+\pi/2], [-\Delta-\pi/2,0)] [ l=2n+1]
Point group: [{\bf 6/}{\bi mmm}]      
[P6/mmm] 24 [\tilde{H}_{j}^{(1)}(2t_{1},\pi/2,\pi/2,0)]  
[ P6/mcc] 24 [\tilde{H}_{j}^{(1)}(2t_{1},\pi/2,\pi/2,0)] [ l=2n]
  24 [ \tilde{H}_{j}^{(1)}(2t_{1}], [\pi/2], [-\pi/2,0)] [ l=2n+1]
[P6_{3}/mcm], [P6_{3}/mmc] 24 [ \tilde{H}_{j}^{(1)}(2t_{1},\pi/2,\pi/2,0)] [l=2n]
  24 [ \tilde{H}_{j}^{(1)}(2t_{1},0,0,0)] [ l=2n+1]
Point group: 23      
P 23, P213 12 [ L_{j}^{3}(t,\Delta)]  
I 23, [I2_{1}3] 24 [ L_{j}^{3}(2t,\Delta)]  
F 23 48 [ L_{j}^{3}(4t,\Delta)]  
Point group: m [\bar{\bf 3}]      
Pm [\bar{3}], Pn[\bar{3}], Pa[\bar{3}] 24 [ L_{j}^{3}(2t_{1},0)]  
Im [\bar{3}], Ia[\bar{3}] 48 [ L_{j}^{3}(4t_{1},0)]  
Fm [\bar{3}] 96 [ L_{j}^{3}(8t_{1},0)]  
Fd [\bar{3}] 96 [ L_{j}^{3}(8t_{1},0)] [h+k+l=2n]
  96 [ L_{j}^{3}(8t_{1},\pi/4)] [h+k+l=2n+1]
And the enantiomorphous space group.

The averages appearing in the above summary are, in general, computed as[\left\langle f(\tau)\right\rangle = ({{2}/{\pi}})\textstyle\int\limits_{0}^{\pi/2}f(\tau)\;{\rm d}\tau, \eqno(2.1.8.39)] except [H_{j}^{(2)}] and [\tilde{H}_{j}^{(2)}] which are computed as [\left\langle f(\tau)\right\rangle = ({{3}/{\pi}})\textstyle\int\limits_{0}^{\pi/3}f(\tau)\;{\rm d}\tau, \eqno(2.1.8.40)] where [f(\tau)] is any of the atomic characteristic functions indicated above. The superscripts preceding the symbols in the above summary are appended to the corresponding symbols in Table 2.1.8.1[link] on their first occurrence.

2.1.8.6. Other non-ideal Fourier p.d.f.'s

| top | pdf |

As pointed out above, the representation of the p.d.f.'s by Fourier series is also applicable to effects of noncrystallographic symmetry. Thus, Shmueli et al. (1985[link]) obtained the following Fourier coefficient for the bicentric distribution in the space group [P\bar{1}] [C_{k} = (2/\pi)\textstyle\int\limits_{0}^{\pi/2}\left[\textstyle\prod\limits_{j = 1}^{N/4}J_{0}(4\pi k\alpha n_{j}\cos\vartheta)\right]\;{\rm d}\vartheta \eqno(2.1.8.41)] to be used with equation (2.1.8.5)[link]. Furthermore, if we use the important property of the characteristic function as outlined in Section 2.1.4.1[link], it is easy to write down the Fourier coefficient for a [P\bar{1}] asymmetric unit containing a centrosymmetric fragment centred at a noncrystallographic centre and a number of atoms not related by symmetry. This Fourier for the above partially bicentric arrangement is a product of expressions (2.1.8.17)[link] and (2.1.8.41)[link], with the appropriate number of atoms in each factor (Shmueli & Weiss, 1985a[link]). While the purely bicentric p.d.f. obtained by using (2.1.8.41)[link] with (2.1.8.5)[link] is significantly different from the ideal bicentric p.d.f. given by equation (2.1.5.13)[link] only when the atomic composition is sufficiently heterogeneous, the above partially bicentric p.d.f. appears to be a useful development even for an equal-atom structure.

The problem of the coexistence of several noncrystallographic centres of symmetry within the asymmetric unit of P [\bar{1}], and its effect on the p.d.f. of [|E|], was examined by Shmueli, Weiss & Wilson (1989[link]) by the Fourier method. The latter study indicates that the strongest effect is produced by the presence of a single noncrystallographic centre.

Another kind of noncrystallographic symmetry is that arising from the presence of centrosymmetric fragments in a noncentrosymmetric structure – the subcentric arrangement already discussed in Section 2.1.5.4[link]. A Fourier-series representation of a non-ideal p.d.f. corresponding to this case was developed by Shmueli, Rabinovich & Weiss (1989[link]), and was also applied to the mathematically equivalent effects of dispersion and presence of heavy scatterers in centrosymmetric special positions in a noncentrosymmetric space group.

A variety of other non-ideal p.d.f.'s occur when heavy atoms are present in special positions (Shmueli & Weiss, 1988[link]). Without going into the details of this development, it can be noted that if the atoms are distributed among k types of Wyckoff positions, the characteristic function corresponding to the p.d.f. of [|E|] is a product of the k characteristic functions, each of which is related to one of these special positions; the same property of the characteristic function as that in Section 2.1.4.1[link] is here utilized.

2.1.8.7. Comparison of the correction-factor and Fourier approaches

| top | pdf |

The need for theoretical non-ideal distributions was exemplified by Fig. 2.1.7.1[link](a), referred to above, and the performance of the two approaches described above, for this particular example, is shown in Fig. 2.1.7.1[link](b). Briefly, the Fourier p.d.f. shows an excellent agreement with the histogram of recalculated [|E|] values, while the agreement attained by the Hermite correction factor is much less satisfactory, even for the (longest available to us) five-term expansion. It must be pointed out that (i) the inadequacy of `short' correction factors, in the example shown, is due to the large deviation from the ideal behaviour and (ii) the number of terms used there in the Fourier summation is twenty, whereafter the summation is terminated. Obviously, the computation of twenty (or more) Fourier coefficients is easier than that of five terms in the correction factor. The convergence of the Fourier series is very satisfactory. It appears that the (analytically) exact Fourier approach is the preferred one in cases of large or intermediate deviations, while the correction-factor approach may cope well with small ones. As far as the availability of symmetry-dependent centric and acentric p.d.f.'s is concerned, correction factors are available for all the space groups (see Table 2.1.7.1[link]), while Fourier coefficients of p.d.f.'s are available for the first 206 space groups (see Table 2.1.8.1[link]). It should be pointed out that p.d.f.'s based on the correction-factor method cope very well with cubic symmetries higher than [Fd\bar{3}], even if the asymmetric unit of the space group is strongly heterogeneous (Rabinovich et al., 1991b[link]).

Both approaches described in this section are related to the characteristic function of the required p.d.f. The correction-factor p.d.f.'s (2.1.7.5)[link] and (2.1.7.6)[link] can be obtained by expanding the logarithm of the appropriate characteristic function in a series of cumulants [e.g. equation (2.1.4.13)[link]; see also Shmueli & Wilson (1982[link])], truncating the series and performing its term-by-term Fourier inversion. The Fourier p.d.f., on the other hand, is computed by forming a Fourier series whose coefficients are exact analytical forms of the characteristic function at points related to the summation indices [e.g. equations (2.1.8.5)[link], (2.1.8.9)[link] and (2.1.8.11)[link], and Table 2.1.8.1[link]] and truncating the series when the terms become small enough.

References

First citation Abramowitz, M. & Stegun, I. A. (1972). Handbook of mathematical functions. New York: Dover.Google Scholar
First citation Barakat, R. (1974). First-order statistics of combined random sinusoidal waves with application to laser speckle patterns. Opt. Acta, 21, 903–921.Google Scholar
First citation Faggiani, R., Lippert, B. & Lock, C. J. L. (1980). Heavy transition metal complexes of biologically important molecules. 4. Crystal and molecular structure of pentahydroxonium chloro(uracilato-N(1))(ethylenediamine)platinum(II)chloride (H5O2)[PtCl(NH2CH2CH2NH2)(C4H5N2O2)]Cl, and chloro(thyminato-N(1))(ethylenediamine)platinum(II), PtCl(NH2CH2CH2NH2)(C5H5N2O2). Inorg. Chem. 19, 295–300.Google Scholar
First citation Rabinovich, S., Shmueli, U., Stein, Z., Shashua, R. & Weiss, G. H. (1991a). Exact random-walk models in crystallographic statistics. VI. P.d.f.'s of [|E|] for all plane groups and most space groups. Acta Cryst. A47, 328–335.Google Scholar
First citation Rabinovich, S., Shmueli, U., Stein, Z., Shashua, R. & Weiss, G. H. (1991b). Exact random-walk models in crystallographic statistics. VII. An all-space-group study of the effects of atomic heterogeneity on the p.d.f.'s of [|E|]. Acta Cryst. A47, 336–340.Google Scholar
First citation Shmueli, U., Rabinovich, S. & Weiss, G. H. (1989). Exact conditional distribution of a three-phase invariant in the space group P1. I. Derivation and simplification of the Fourier series. Acta Cryst. A45, 361–367.Google Scholar
First citation Shmueli, U., Rabinovich, S. & Weiss, G. H. (1990). Exact random-walk models in crystallographic statistics. V. Non-symmetrically bounded distributions of structure-factor magnitudes. Acta Cryst. A46, 241–246.Google Scholar
First citation Shmueli, U. & Weiss, G. H. (1985a). Centric, bicentric and partially bicentric intensity statistics. Structure and statistics in crystallography, edited by A. J. C. Wilson, pp. 53–66. Guilderland: Adenine Press.Google Scholar
First citation Shmueli, U. & Weiss, G. H. (1985b). Exact joint probability distributions for centrosymmetric structure factors. Derivation and application to the Σ1 relationship in the space group [P\bar{1}]. Acta Cryst. A41, 401–408.Google Scholar
First citation Shmueli, U. & Weiss, G. H. (1986). Exact joint distribution of Eh , Ek and Eh+k, and the probability for the positive sign of the triple product in the space group [P\bar{1}]. Acta Cryst. A42, 240–246.Google Scholar
First citation Shmueli, U. & Weiss, G. H. (1987). Exact random-walk models in crystallographic statistics. III. Distributions of [|E|] for space groups of low symmetry. Acta Cryst. A43, 93–98.Google Scholar
First citation Shmueli, U. & Weiss, G. H. (1988). Exact random-walk models in crystallographic statistics. IV. P.d.f.'s of [|E|] allowing for atoms in special positions. Acta Cryst. A44, 413–417.Google Scholar
First citation Shmueli, U., Weiss, G. H. & Kiefer, J. E. (1985). Exact random-walk models in crystallographic statistics. II. The bicentric distribution in the space group [P\bar{1}]. Acta Cryst. A41, 55–59.Google Scholar
First citation Shmueli, U., Weiss, G. H., Kiefer, J. E. & Wilson, A. J. C. (1984). Exact random-walk models in crystallographic statistics. I. Space groups [P\bar{1}] and [P1]. Acta Cryst. A40, 651–660.Google Scholar
First citation Shmueli, U., Weiss, G. H. & Wilson, A. J. C. (1989). Explicit Fourier representations of non-ideal hypercentric p.d.f.'s of [|E|]. Acta Cryst. A45, 213–217.Google Scholar
First citation Shmueli, U. & Wilson, A. J. C. (1982). Intensity statistics: non-ideal distributions in theory and practice. In Crystallographic statistics: progress and problems, edited by S. Ramaseshan, M. F. Richardson & A. J. C. Wilson, pp. 83–97. Bangalore: Indian Academy of Sciences.Google Scholar
First citation Weiss, G. H. & Kiefer, J. E. (1983). The Pearson random walk with unequal step sizes. J. Phys. A, 16, 489–495.Google Scholar








































to end of page
to top of page