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

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767

Solving molecular crystal structures from laboratory X-ray powder diffraction data with DASH: the state of the art and challenges

aDepartment of Pharmaceutical Sciences, University of Strathclyde, 27 Taylor Street, Glasgow G4 0NR, UK, bCrystallografX Ltd, 38 Queen Street, Glasgow G1 3DX, UK, cISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 0QX, UK, dCambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK, eDepartment of Pure and Applied Chemistry, University of Strathclyde, Glasgow G1 1XL, UK, fSchool of Pharmacy, The Robert Gordon University, Schoolhill, Aberdeen AB10 1FR, UK, gUniversity of Durham, Department of Chemistry, University Science laboratories, South Road, Durham DH1 3LE, UK, hAstraZeneca R&D Charnwood, Loughborough, Leicestershire LE11 5RH, UK, and iPharmorphix Ltd, 250 Cambridge Science Park, Milton Road, Cambridge CB4 0WE, UK
*Correspondence e-mail: alastair.florence@strath.ac.uk

(Received 10 August 2004; accepted 9 December 2004)

The crystal structures of 35 molecular compounds have been redetermined from laboratory monochromatic capillary transmission X-ray powder diffraction data using the simulated-annealing approach embodied within the DASH structure solution package. The compounds represent industrially relevant areas (pharmaceuticals; metal coordination compounds; nonlinear optical materials; dyes) in which the research groups in this multi-centre study are active. The molecules were specifically selected to form a series within which the degree of structural complexity (i.e. degrees of freedom in the global optimization) increased systematically, the degrees of freedom increasing with increasing number of optimizable torsion angles in the structural model and with the inclusion of positional disorder or multiple fragments (counterions; crystallization solvent; Z′ > 1). At the lower end of the complexity scale, the structure was solved with excellent reproducibility and high accuracy. At the opposite end of the scale, the more complex search space offered a significant challenge to the global optimization procedure and it was demonstrated that the inclusion of modal torsional constraints, derived from the Cambridge Structural Database, offered significant benefits in terms of increasing the frequency of successful structure solution by restricting the magnitude of the search space in the global optimization.

1. Introduction

Global optimization methods for crystal structure determination from powder diffraction data (SDPD) have become widely available in recent years and have successfully been applied to solve the structures of organic (Harris & Cheung, 2003[Harris, K. D. M. & Cheung, E. Y. (2003). Org. Process Res. Dev. 7, 970-976.]; Johnston et al., 2004[Johnston, A., Florence, A. J., Shankland, K., Markvardsen, A., Shankland, N., Steele, G. & Cosgrove, S. D. (2004). Acta Cryst. E60, o1751-o1753.]; Rukiah et al., 2004[Rukiah, M., Lefebvre, J., Descamps, M., Hemon, S. & Dzyabchenko, A. (2004). J. Appl. Cryst. 37, 464-471.]; Zaske et al., 2004[Zaske, L., Perrin, M.-A., Daiguebonne, C. & Guillou, O. (2004). Mater. Sci. Forum, 443-444, 411-421.]), inorganic (Deem & Newsam, 1989[Deem, M. W. & Newsam, J. M. (1989). Nature (London), 342, 260-262.]; Edgar et al., 2002[Edgar, M., Carter, V. J., Tunstall, D. P., Grewal, P., Favre-Nicolin, V., Cox, P. A., Lightfoot, P. & Wright, P. A. (2002). Chem. Commun. 8, 808-809.]; Reinaudi et al., 2000[Reinaudi, L., Leiva, E. P. M. & Carbonio, R. E. (2000). J. Chem. Soc. Dalton Trans. 23, 4258-4262.]) and organometallic (Ivashevskaja et al., 2002[Ivashevskaja, S. N., Aleshina, L. A., Andreev, V. P., Nizhnik, Y. P., Chernyshev V. V. & Schenk, H. (2002). Acta Cryst. C58, m300-m301.]; Dinnebier et al., 2000[Dinnebier, R. E., Wagner, M., Peters, F., Shankland, K. & David, W. I. F. (2000). Z. Anorg. Allg. Chem. 626, 1400-1405.]) materials, to cite but a few examples. The basis of global optimization strategies has been fully described elsewhere (Shankland & David, 2002[Shankland, K. & David, W. I. F. (2002). Structure Determination from Powder Diffraction Data, edited by W. I. F. David, K. Shankland, L. B. McCusker & Ch. Baerlocher, pp. 252-285. IUCr/Oxford University Press.]) and software implementing global optimization methods is now widely available [e.g. DASH (David et al., 2001[David, W. I. F., Shankland, K., Cole, J., Maginn, S., Motherwell, W. D. S. & Taylor, R. (2001). DASH user manual. Cambridge Crystallographic Data Centre, Cambridge, UK.]), ESPOIR (Le Bail, 2001[Le Bail, A. (2001). Mater. Sci. Forum, 378-381, 65-70.]), FOX (Favre-Nicolin & Cerny, 2002[Favre-Nicolin, V. & Cerny, R. (2002). J. Appl. Cryst. 35, 734-743.]), PowderSolve (Engel et al., 1999[Engel, G. E., Wilke, S., König, O., Harris, K. D. M. & Leusen, F. J. J. (1999). J. Appl. Cryst. 32, 1169-1179.]), TOPAS (Coelho, 2003[Coelho, A. A. (2003). J. Appl. Cryst. 36, 86-95.])].

It is the application of global optimization methods to structure determination from data collected on standard, widely available, laboratory diffractometers that concerns us here. Specifically, the aim is to quantify the accuracy of a series of crystal structures solved from laboratory X-ray powder diffraction (XRPD) data using the simulated-annealing (SA) approach (David et al., 1998[David, W. I. F., Shankland, K. & Shankland, N. (1998). Chem. Commun. pp. 931-932.]) implemented in the DASH structure solution package and to investigate factors influencing the chances of successful structure solution.

1.1. Data quality

To maximize the chances of successfully and accurately solving crystal structures from laboratory XRPD data, the following data requirements should be addressed: accurate measurement of reflection positions and intensities; high angular resolution (i.e. small FWHM) and spatial resolution (ca 2 Å or better); good signal-to-background ratios across the full pattern and minimal preferred orientation (PO) effects. These requirements are best achieved in the laboratory with the sample mounted in a rotating capillary and the data collected in transmission geometry using monochromatic Cu Kα1 radiation. Linear one-dimensional position-sensitive detectors (PSDs) combine excellent angular resolution with favourable count rates; recent developments in solid-state PSDs offer the prospect of even greater improvements in performance with respect to background, sensitivity and data acquisition rates.

1.2. Maximizing the chances of success

There are a range of strategies which can be generally applied to maximize the chances of successfully solving a crystal structure from laboratory XRPD data, including those summarized in Table 1[link]. Of particular interest in the study of complex structures is the incorporation of prior chemical information in the form of torsion-angle constraints. These constraints do not reduce the number of degrees of freedom (DOF) to be optimized during the search, but do reduce the extent of the search space explored during the SA process, a strategy that has been shown to be highly effective in SDPD (Middleton et al., 2002[Middleton, D. A., Peng, X., Saunders, D., Shankland, K., David, W. I. F. & Markvardsen, A. J. (2002). Chem. Commun. pp. 1976-1977.]; torsion-angle constraints derived by solid-state NMR conformational analysis). For the constraints approach to become amenable to routine application, the derivation of the constraints for any given problem has to be as straightforward as possible. Fortunately, this is readily tractable with the Cambridge Structural Database (CSD; Allen, 2002[Allen, F. H. (2002). Acta Cryst. B58, 380-388.]) and single-range torsion-angle constraints (e.g. 40 to 160°) derived from the CSD have previously been used to increase the frequency of success in global optimization structure solution (Shankland et al., 1998[Shankland, K., David, W. I. F., Csoka, T. & McBride, L. (1998). Int. J. Pharm. 165, 117-126.]).

Table 1
Approaches for optimizing data quality and maximizing the chance of successfully solving crystal structures from laboratory XRPD data

Approach Aim/advantage Comment/reference
Recrystallization Minimize intrinsic sample line width; improve angular resolution Risk of phase transformation or texture effects
Low-T data collection Improve signal-to-noise, particularly at high 2θ angles; improve accuracy of reflection intensities Differential thermal expansion (Zachariasen & Ellinger, 1963[Zachariasen, W. H. & Ellinger, F. H. (1963). Acta Cryst. 16, 369-375.]; Shankland et al., 1997[Shankland, K., David, W. I. F. & Sivia, D. (1997). J. Mater. Chem. 7, 569-572.]); risk of phase transformation
Variable count-time data collection As `low-T data collection' Madsen & Hill (1994[Madsen, I. C. & Hill, R. J. (1994). J. Appl. Cryst. 27, 385-392.])
Optimize SA control parameters Increase probability of locating global minimum For example, reduce the cooling rate to avoid quenching (Shankland et al., 2002[Shankland, K., McBride, L., David, W. I. F., Shankland, N. & Steele, G. (2002). J. Appl. Cryst. 35, 443-454.])
Crystallographic constraints Reduce number of degrees of freedom to be optimized during search; increase probability of locating global minimum For example, in space groups such as P1, with floating origins, fixing the x, y and z coordinates of an atom in the formula unit removes three degrees of freedom
Chemical constraints As `crystallographic constraints' For example, fixing amide torsion angle (H—N—C=O) to an exact value of 180°, eliminating it from the optimization

Although it has always been possible to input single-range constraints into the DASH program, modal torsion constraints, whereby multiple ranges are defined for each individual torsion angle, offer a more selective means of constraining complex optimization problems. It has been found that, in general, the values of specific torsion angles within crystal structures in the CSD will form distributions that can be classified as uni-, bi- or trimodal. For example, the C—C—C—C torsion angle defined by CH2—CH2—CH2—C(=O) adopts values that fall into three ranges; 40 to 80°, 160 to −160° and −80 to −40° (Fig. 1[link]; see also §3[link]). Accordingly, this torsion angle is classified as `trimodal', with the middle of each discrete cluster separated by 120°. Thus, a lower and upper bound for a single mode of the torsion angle may be input via the DASH interface (e.g. 40 to 80°) along with the modal type (i.e. trimodal), whereupon the program automatically generates the bounds for the other two modes. The SA algorithm then samples torsion-angle values from these three ranges during the search. The ability to sample only relevant regions of torsion-angle space is potentially advantageous in solving crystal structures with a large number of internal DOF. One important caveat to the general applicability of this approach is the finite possibility that the conformation adopted at a specific torsion angle within the molecule of interest may lie outside the ranges measured from known structures within the CSD.

[Figure 1]
Figure 1
A polar plot showing the torsion-angle values obtained from a search of the CSD for the fragment CH2–CH2–CH2–C(=O). All C atoms were defined as acyclic. The distribution shows three clear modes (i.e. a trimodal distribution) centred on ca 60, 180, −60°, with the largest number of structures within the distribution adopting a trans conformation at this torsion angle, i.e. in the mode centred around 180°.

2. Data collection

The 35 compounds used in the study were selected to cover a wide range of structural complexity, including significant conformational (torsional) flexibility, salts, solvates, positional disorder (27 and 33) and Z′ > 1 (34) (Table 2[link], Fig. 2[link]). A prerequisite for inclusion in the study was the availability of reference crystal structures for the purpose of evaluating the accuracy of the structures solved using the SDPD approach (see §3[link]). All polycrystalline samples (except compound 8) were lightly ground in an agate mortar and pestle and filled into 0.7 mm borosilicate glass capillaries prior to being mounted and aligned on a Bruker-AXS D8 Advance powder diffractometer (Table 3[link]). Compound 8, the orthorhombic form of paracetamol (form II), was prepared in situ by cooling a molten sample of paracetamol to room temperature inside a 0.7 mm borosilicate glass capillary.

Table 2
Compound name and molecular formula with the reference code used throughout the text

Code Name Molecular formula
1 Hydrochlorothiazide C7H8ClN3O4S2
2 2-Mercaptobenzoic acid C7H6O2S
3 N,N′-Bis[1-pyridin-4-yl-meth-(E)-ylidene]­hydrazine C12H10N4
4 Carbamazepine (β polymorph) C15H12N2O
5 Dapsone C12H12N2O2S
6 Hydroflumethiazide C8H8F3N3O4S2
7 Paracetamol (form I polymorph) C8H9NO2
8 Paracetamol (form II polymorph) C8H9NO2
9 Phenylacetic acid C8H8O2
10 2-(Phenylsulfonyl)acetamide C8H9NO3S
11 Captopril C9H15NO3S
12 Methyl 4-[(4-aminophenyl)ethynyl]benzoate C16H13NO2
13 trans-Dichlorobis(triphenylphos­phine)nickel(II) C36H30Cl2NiP2
14 2-(4-Hydroxy-2-oxo-2,3-dihydro-1,3-benzothiazol-7-yl)ethylammonium chloride C9H11N2O2S·Cl
15 Salbutamol C13H21NO3
16 trans-Diisothiocyanatobis(triphenylphosphine)nickel(II) C38H30N2NiP2S2
17 Dopamine hydrobromide C8H12NO2·Br
18 Methyl 4-{[4-(dimethylamino)phenyl]­ethynyl}benzoate C18H17NO2
19 cis-Thiothixene C23H29N3O2S2
20 Chlorpropamide C10H13ClN2O3S
21 Creatine monohydrate C4H9N3O2·H2O
22 1,4-Bis(2-phenethyloxyethanesulfonyl)piperazine C24H34N2O6S2
23 Clomipramine hydrochloride C19H24ClN2·Cl
24 α-Lactose monohydrate C12H22O11·H2O
25 Promazine hydrochloride C17H21N2S·Cl
26 Tolbutamide C12H18N2O3S
27 Carbamazepine dihydrate C15H12N2O·2H2O
28 Famotidine C8H15N7O2S3
29 Diltiazem hydrochloride C22H27N2O4S·Cl
30 Zopiclone dihydrate C17H17ClN6O3·2H2O
31 Capsaicin C18H27NO3
32 Sodium 4-[(E)-(4-hydroxyphenyl)diazenyl]benzene sulfonate dihydrate C12H9N2O4S·Na·2H2O
33 2-{[3-(2-Phenylethoxy)propyl]sul­fonyl}­ethyl benzoate C20H24O5S
34 S-Ibuprofen C13H18O2
35 Verapamil hydrochloride C27H39N2O4·Cl

Table 3
Instrumental and data collection parameters

Typical instrument settings  
System D8 Advance θ/2θ
Generator 50 kV, 40 mA
Measuring diameter (mm) 435
Radiation (Å) Cu Kα1, λ = 1.54056 Å
Monochromator Primary, focusing curved Ge 111
Geometry Transmission capillary configuration
Sample holder 0.7 mm borosilicate glass capillary
Detector PSD system MBraun OED-50M
   
Typical measuring conditions  
Range (° 2θ) 5–65
Step size (° 2θ) 0.0145
Step time (s) 10.0
Total data collection time (h) ca 10
[Figure 2]
Figure 2
Molecular structures of compounds 135.

All data were collected at room temperature and can be accessed at http://www.powderdata.info.

3. Data analysis and simulated annealing

Diffraction patterns were indexed using DICVOL91 (Boultif & Louër, 1991[Boultif, A. & Louër, D. (1991). J. Appl. Cryst. 24, 987-993.]) to obtain lattice parameters that were subsequently refined (Table 4[link]) along with background, zero point, peak shape parameters and reflection intensities in a Pawley fit (Pawley, 1981[Pawley, G. S. (1981). J. Appl. Cryst. 14, 357-361.]) using DASH. All samples gave sharp diffraction, with good to moderate angular resolution and a mean FWHM = 0.099 ± 0.015 Å (Table 5[link]). Data were truncated as necessary to allow up to 350 reflections to be extracted from each pattern, with the spatial resolution across all of the data sets ranging from 1.44 to 2.18 Å (Table 5[link]).

Table 4
Space group and refined unit-cell parameters (this work) for compounds 135

The last column identifies the reference crystal structures (typically CSD refcode/CCDC deposition number) used to calculate the RMSD values in Table 5[link]. The structures of 6, 14, 22, 27, 28 and 33, in CIF format, can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033; e-mail: deposit@ccdc.cam.ac.uk). For the other structures, the individual citations corresponding to each refcode are given in Table 8[link].

Pawley χ2 is the profile χ2 for the Pawley fit as described in the DASH manual: χ2 = [\big\{\textstyle\sum_{i}^{N} w_{i} [y_{i} ({\rm obs}) - y_{i} ({\rm calc})]^{2}\big\}/](NP + C), where yi(obs) is the observed intensity at the ith step in the powder diffraction pattern, yi(calc) is the associated calculated intensity; wi = 1/[\sigma _{i}^{2}], where σi is the standard deviation of the observed intensity at that point. The summation is performed over all N data points; (NP + C) = (number of data points) − (number of parameters) + (number of parameter constraints).

Code Space group a (Å) b (Å) c (Å) α (o) β (o) γ (o) Pawley χ2 Reference
1 P21 7.400 8.506 10.006 90 111.72 90 2.18 HCSBTZ
2 P21/c 7.885 5.976 14.949 90 100.48 90 2.84 ZZZLWW01
3 P21/c 3.848 11.005 12.727 90 92.38 90 4.25 LIZCUS
4 P21/n 7.537 11.157 13.918 90 92.87 90 2.98 CBMZPN10
5 P212121 25.538 8.061 5.762 90 90 90 3.10 DAPSUO10
6 P21 7.636 8.662 9.743 90 110.30 90 3.92 CCDC 198487
7 P21/n 7.100 9.380 11.708 90 97.42 90 5.42 HXACAN07
8 Pbca 17.142 11.822 7.404 90 90 90 2.82 HXACAN08
9 P21/a 10.226 4.967 14.467 90 99.25 90 7.72 ZZZMLY01
10 P21/c 8.884 5.408 19.469 90 101.66 90 4.77 Frampton (2004[Frampton, C. S. (2004). Unpublished single-crystal data.])
11 P212121 8.810 17.948 6.834 90 90 90 2.65 MCPRPL
12 P21 7.572 5.908 14.141 90 95.34 90 2.80 Marder (2004[Marder, T. B. (2004). Unpublished single-crystal data.])
13 P2/c 11.638 8.197 17.388 90 107.03 90 3.93 CLTPNI03
14 P21/a 7.555 14.640 10.246 90 109.30 90 1.86 CCDC 247129
15 Pbca 21.657 8.783 14.555 90 90 90 2.94 BHHPHE
16 P[\bar{1}] 7.958 10.488 11.500 111.10 74.56 92.29 4.36 GEBZUI
17 Pbc21 10.671 11.459 7.950 90 90 90 7.61 QQQAEJ01
18 Pna21 6.121 7.472 33.002 90 90 90 2.97 Marder (2004[Marder, T. B. (2004). Unpublished single-crystal data.])
19 P21 10.141 8.695 13.683 90 110.60 90 3.13 THTHXN01
20 P212121 9.078 5.220 26.658 90 90 90 6.87 BEDMIG
21 P21/c 12.506 5.046 12.169 90 108.88 90 3.06 CREATH03
22 P21/a 13.442 5.182 19.796 90 108.74 90 3.10 CCDC 247131
23 P21/c 15.514 8.610 14.035 90 96.93 90 3.81 CIMPRA
24 P21 7.937 21.573 4.814 90 109.75 90 2.59 LACTOS10
25 P21/c 11.806 11.497 13.429 90 111.71 90 2.41 PROMZC01
26 Pna21 20.218 7.820 9.072 90 90 90 7.67 ZZZPUS02
27 Cmca 19.775 4.937 28.719 90 90 90 7.87 CCDC 247132
28 P21/c 17.767 5.334 18.311 90 123.64 90 2.70 CCDC 198488
29 P212121 42.190 9.075 6.037 90 90 90 8.29 CEYHUJ01
30 P21/c 16.479 7.145 17.398 90 108.80 90 3.72 UCUVET
31 P21/c 12.672 14.980 9.426 90 93.69 90 6.90 Frampton (2004[Frampton, C. S. (2004). Unpublished single-crystal data.])
32 Pbcn 14.591 5.831 32.952 90 90 90 4.39 YAYWUQ
33 P21/n 5.137 37.934 9.844 90 98.50 90 4.22 CCDC 247130
34 P21 12.463 8.029 13.538 90 112.93 90 2.09 JEKNOC10
35 P[\bar{1}] 7.089 10.593 19.207 100.11 93.75 101.56 4.70 CURHOM

Table 5
Summary of results of the SA runs for compounds 135

FWHM = average full width at half-maximum of 8 reflections measured in the XRPD data sets out to ca 30° 2θ; DOF (ext) = number of optimized external degrees of freedom; DOF (int) = number of optimized internal degrees of freedom; Data range = data range used in the Pawley fit; Nrefs = number of reflections in the fitted data range; Res. = spatial resolution of data used in the SA runs; Profile χ2 = range of profile χ2 values observed at the end of 20 SA runs (χ2 calculated as per footnote to Table 4[link]); Nsol = number of correct structure solutions obtained from a batch of 20 SA runs; χ2 ratio = [\chi ^{2}_{\rm profile} / \chi ^{2}_{\rm Pawley}] for the best solution, i.e. that with the lowest profile χ2; RMSD = root mean square displacement (as defined in §3[link]) for the best solution.

Code FWHM (° 2θ) DOF (ext) DOF (int) Data range (° 2θ) Nrefs Res. (Å) Profile χ2 Nsol χ2 ratio RMSD (Å)
1 0.097 6 1 8.0–55.8 157 1.65 6.75–18.97 20 3.1 0.102
2 0.095 6 1 8.5–64.6 244 1.44 5.41–5.42 20 1.9 0.026
3 0.102 6 1 5.0–54.7 121 1.68 9.69–9.88 20 2.3 0.026
4 0.093 6 1 7.5–56.4 284 1.63 9.04–12.28 20 3.0 0.017
5 0.108 6 2 5.0–56.3 196 1.63 11.45–13.77 20 3.7 0.027
6 0.117 6 2 9.0–45.7 94 1.99 13.62–13.98 20 3.5 0.131
7 0.080 6 2 10.0–64.7 268 1.44 13.56–20.05 20 2.5 0.092
8 0.138 6 2 9.0–64.9 265 1.44 19.26–23.87 20 6.8 0.140
9 0.112 6 2 5.0–56.9 171 1.63 22.92–24.26 20 3.0 0.077
10 0.085 6 3 8.0–49.9 153 1.83 9.96–10.24 20 2.1 0.079
11 0.101 6 4 9.0–55.2 169 1.66 12.63–12.73 20 4.8 0.077
12 0.138 6 4 4.0–44.6 97 2.03 4.16–4.82 20 1.5 0.109
13 0.084 6 4 5.0–52.2 288 1.75 8.04–19.45 20 2.0 0.070
14 0.092 6+3 2 5.0–52.1 315 1.75 4.63–7.07 20 2.5 0.058
15 0.095 6 5 6.0–51.8 263 1.76 27.51–29.05 20 9.4 0.088
16 0.108 6 5 6.5–47.9 264 1.90 15.63–18.15 20 3.6 0.058
17 0.104 6 + 3 2 7.0–49.6 101 1.84 34.25–43.23 20 4.5 0.075
18 0.125 6 5 4.0–44.3 98 2.04 5.14–6.05 20 1.7 0.066
19 0.097 6 5 5.0–48.9 214 1.88 22.95–55.94 19 7.3 0.129
20 0.097 6 6 5.0–49.2 149 1.85 23.43–24.95 20 3.4 0.070
21 0.088 6 + 3 3 5.0–62.4 232 1.49 15.56–161.00 § 19 5.1 0.059
22 0.115 6 7 4.0–51.3 242 1.77 7.02–9.50 20 2.3 0.138
23 0.084 6 + 3 4 10.0–49.0 306 1.86 15.47–184.54 § 19 4.1 0.109
24 0.085 6 + 3 4 5.0–49.9 145 1.82 20.21–63.52 19 7.8 0.073
25 0.079 6 + 3 4 6.7–52.0 330 1.76 10.33–12.16 20 4.3 0.107
26 0.111 6 7 7.5–59.9 228 1.54 17.39–72.45 20 2.3 0.127
27 0.092 6 + 3 + 3 2 5.0–60.4 218 1.53 38.05–39.27 20 4.8 0.116
28 0.105 6 9 5.0–48.9 229 1.86 10.00–204.18 8 3.7 0.063
29 0.098 6 + 3 7 7.0–41.5 164 2.18 23.97–229.63 8 2.9 0.118
30 0.086 6 + 3 + 3 4 4.7–50.5 337 1.80 13.07–223.82 17 3.5 0.096
31 0.092 6 11 5.0–51.8 338 1.76 29.13–192.38 5 4.2 0.168
32 0.093 6 + 3 + 3 + 3 3 4.5–56.4 341 1.63 21.90–122.20 § 19 5.0 0.136
33 0.086 6 13 4.0–46.1 302 1.96 17.24–221.32 3 4.1 0.165
34 0.090 6 + 6 4 + 4 6.0–54.7 321 1.68 8.87–96.95 7 4.2 0.073
35 0.082 6 + 3 13 4.4–42.6 308 2.12 45.02–232.50 1 9.6 0.204
†PO correction included in SA runs for 8 (r = 1.50, [001]), 19 (r = 0.88, [010]), 21 (r = 1.06, [100]).
‡Six DOF for the cation and three for the anion in 14. In other structures comprising >1 fragment, the number of DOF for each fragment that was optimized is identified in the table.
§For 21, 23 and 32, the largest profile χ2 among the 19 correct solutions in each case was 22.79, 27.17 and 36.27, respectively.
¶Disordered (half-occupancy) phenyl ring model.

Z matrices describing the molecular topology of the fragments in each compound were generated automatically from the appropriate reference crystal structure1 using DASH, and all optimizable torsion angles were automatically assigned to vary in the range −180 to 180°. A single O atom was used to approximate each water molecule of crystallization in hydrates 21, 24, 27, 30 and 32 and the Z matrices of 27 and 33 were manually altered to accommodate positional disorder.

Global optimization of all external (rotational and translational) and internal (torsion angles) DOF (Table 5[link]) against the extracted intensities was carried out with all DASH SA control parameters set to default values. 20 runs with 1 × 107 SA moves per run were implemented for each structure determination, with a simplex refinement being executed upon completion. The structure of the best solution (i.e. that with the lowest profile χ2) was overlaid upon the corresponding reference crystal structure and the root mean square dis­place­ment (RMSD, Å) calculated for all non-H atoms (Table 5[link]). The majority of the reference data comprised single-crystal structures retrieved from the CSD (Table 4[link]). In instances where the data collection temperatures for the XRPD and reference structures are not matched, the magnitude of the RMSD value necessarily contains a contribution which is attributable to this temperature difference.

SA runs for 28, 29, 31, 34 and 35 were repeated using modal torsional constraints (§1.2[link]). The CSD (November 2002, v5.24) was searched for fragments of molecules corresponding to the torsion angle of interest using Conquest (Bruno et al., 2002[Bruno, I. J., Cole, J. C., Edgington, P. R., Kessler, M., Macrae, C. F., McCabe, P., Pearson, J. & Taylor, R. (2002). Acta Cryst. B58, 389-397.]). The torsion angle was specified as a geometric parameter so that the appropriate torsion values from hits were recorded. Hits of the search were viewed in Vista (CCDC software) and the torsion-angle ranges to be used in DASH were chosen by inspection (Table 6[link]). A modal torsion-angle range was not specified if there were less than 30 observations or if there was no clear distribution.

Table 6
CSD searches and modal ranges utilized in the modal constraints within DASH

CSD search. Atom numbers correspond with the schemes used by Golič et al. (1989[Golič, L., Djinović, K. & Florjanič, M. (1989). Acta Cryst. C45, 1381-1384.]) (28); Kojic Prodic et al. (1984[Kojic Prodic, B., Ruzic Toros, Z., Sunjic, V., Decorte, E. & Moimas, F. (1984). Helv. Chim. Acta, 67, 916-925.]) (29); CCDC 171602 (David et al., 1998[David, W. I. F., Shankland, K. & Shankland, N. (1998). Chem. Commun. pp. 931-932.]) (31); Freer et al. (1993[Freer, A. A., Bunyan, J. M., Shankland, N. & Sheen, D. B. (1993). Acta Cryst. C49, 1378-1380.]) (34) and Carpy et al. (1985[Carpy, A., Leger, J.-M. & Melchiorre, C. (1985). Acta Cryst. C41, 624-627.]) (35). The four atoms of the torsion angle were specified as `cyclic' or `acyclic' (subscripts c and a, respectively) and the appropriate bond types between the atoms of the torsion angle were also defined in the search (`∼' is an unspecified bond type). In addition, where appropriate, the environment of the torsion-angle atoms was specified to narrow the Conquest search to include only very closely related fragments, e.g. number of H atoms attached or total number of coordinated atoms (T2 = only two connections to atom allowed; T3 = only three connections to atom allowed; T4 = only four connections to atom allowed).

Mode B = bimodal. Planar torsion-angle ranges (centred around 0° and 180°) are searched by inputting into DASH the bounds for a single mode of the torsion angle (e.g. −160° to 160°), along with the modal type (bimodal). The program automatically generates the complementary bounds for the other mode (in this case, −20° to 20°). For non-planar torsion angles, inputting, say, 30° to 50°, and specifying `bimodal' will automatically generate −30° to −50° for the bounds of the other mode.

Mode T = trimodal. The modal type and the bounds for a single mode of a torsion angle (e.g. −160° to 160°) are input into DASH and the program automatically generates the bounds for the other two modes.

Code Torsion angle CSD search Nobs Range (°) Mode
28 N1:C1:N3:C2 NaH2—Ca=Na—Cc 51 −160 to 160 B
  S3:N5:C8:C7 SaO2—Na∼Ca—CaH2 124 60 to 180 B
  C8:C7:C6:S2 Ca—CaH2—CaH2—Sa 613 −160 to 160 T
  C3:C5:S2:C6 Cc—CaH2—Sa—CaH2 173 −150 to 150 T
  C7:C6:S2:C5 CaH2—CaH2—Sa—CaH2 233 −150 to 150 T
  N4:C3:C5:S2 Nc—Cc—CaH2—SaT2 49 −150 to 150 T
  O1:S3:N5:C8 Oa=Sa—Na∼Ca 750 −150 to 150 T
  C1:N3:C2:N4 Ca=Na—Cc=Nc 38 20 to −20 T
  N5:C8:C7:C6 Na=Ca—CaH2—CaH2 49 No clear distribution  
           
29 C17:O2:C2:C1 Ca—Oa—CaH—CaH 1566 60 to 180 B
  C12:C10:C1:C2 CcH—Cc—CaH—CaH 986 30 to 150 B
  O3:C17:O2:C2 O=Ca—Oa—CaH 1958 −20 to 0 B
  C20:C19:N1:C3 CaH2—CaH2—NaT3—Ca=O 892 60 to 180 B
  C16:O1:C13:C11 CaH3—Oa—Cc—CcH 4126 −160 to 160 B
  N2:C20:C19:N1 NaH+—CaH2—CaH2—NaT3 63 −150 to 150 T
  C21:N2:C20:C19 Ca—NaH+—CaH2—CaH2 1504 −150 to 150 T
           
31 C8:N1:C9:C10 CaH2—NaH—Ca(=O)—CaH2 326 −160 to −180 B
  C1:C8:N1:C9 Cc—CaH2—NaH—CaH2 82 60 to 125 B
  C15:C14:C13:C12 Ca—CaH=CaH—CaH2—CaH2 159 90 to 160 B
  H20:C15:C14:C13 H—Ca=CaH—CaH2 3877 −160 to 160 B
  C7:O1:C3:C2 CaH3—Oa—Cc—CcH 4126 −160 to 160 B
  N1:C9:C10:C11 C—NaH—Ca(=O)—CaH2—CaH2 280 60 to 180 B
  C13:C12:C11:C10 CaH2—CaH2—CaH2—CaH2 >2000 −150 to 150 T
  C14:C13:C12:C11 CaH2—CaH2—CaH2—CaH2 >2000 −150 to 150 T
  C12:C11:C10:C9 CaH2—CaH2—CaH2—Ca(=O) 1075 −160 to 160 T
  C17:C16:C15:C14 Ca—CaT4—CaH=Ca 2501 −30 to 30 T
  C2:C1:C8:N1 Cc—Cc—CaH2—NaHT3 516 No clear distribution  
           
34 C1:C2:C4:C5 Cc—Cc—CaHT4—CaT4 >10000 20 to 120 B
  C11:C10:C7:C6 Cc—Cc—CaH2—CaT4 4087 50 to 130 B
  C12:C11:C10:C7 Cc—CaH2—CaT4—CaT4 1255 −150 to 150 T
  O1:C1:C2:C4 OH—C(=O)—CaHT4—Cc 190 −150 to 150 T
           
35 C14:C13:C12:C11 Cc—Cc—CaH2—CaT4 4087 50 to 130 B
  C2:C1:C7:C8 Cc—Cc—CaH2—CaT4 4087 50 to 130 B
  C27:O4:C17:C18 CaH3—Oa—Cc—CcH 4126 −160 to 160 B
  C26:O3:C16:C15 CaH3—Oa—Cc—CcH 4126 −160 to 160 B
  C20:O2:C5:C6 CaH3—Oa—Cc—CcH 4126 −160 to 160 B
  C19:O1:C4:C3 CaH3—Oa—Cc—CcH 4126 −160 to 160 B
  C12:C11:C10:C9 CaH2—CaH2—CaH2—CaH2 >2000 −150 to 150 T
  C23:C22:C12:C11 CaH2—CaH2—CaH2—CaH2 >2000 −150 to 150 T
  C8:N1:C9:C10 CaH2—Na [C(H)]—CaH2—CaH2 >1000 −150 to 150 T
  C9:N1:C8:C7 CaH2—Na [C(H)]—CaH2—CaH2 >1000 −150 to 150 T
  C13:C12:C11:C10 Cc—Ca(C2)—CaH2—CaH2 204 −150 to 150 T
  N1:C8:C7:C1 NHT4—CaH2—CaH2—C 2090 −150 to 150 T
  N1:C9:C10:C11 NHT4—CaH2—CaH2—C 2090 −150 to 150 T
†See Fig. 1[link].

4. Results and discussion

The crystal structures of all compounds were solved successfully2 and the correct solution obtained with excellent reproducibility in the majority of cases (Table 5[link]). For compounds 127, with DOF < 15, correct solutions were generated in ∼100% of SA runs, with a relatively narrow spread in the [\chi ^{2}_{\rm profile}] range observed for any one compound.

For the more complex structures 2835, with DOF ≥ 15, the adverse effect of local minima in the agreement hypersurface is reflected in the reduced frequency of success and the accompanying increased spread in [\chi ^{2}_{\rm profile}] for a particular compound. The exceptions to this are 30 (DOF = 16; Nsol = 17) and 32 (DOF = 18; Nsol = 19), which have in common a high degree of planar aromatic structure and a small number of internal DOF (4 and 3 DOF, respectively). This combination of molecular features clearly favoured success in reaching the global minimum in the SA runs.

For the level of success achieved with 2834, a batch size of 20 SA runs proved sufficient to solve the structure reproducibly and, therefore, convincingly. The same does not hold true for the most complex example, 35, which returned only one solution in 20 runs; this aspect of verapamil hydrochloride is discussed further in §4.2.3[link].

The excellent accuracy of the best SA solution, across the full range of 35 structures, is reflected in each case by the favourable [\chi ^{2}_{\rm profile} / \chi ^{2}_{\rm Pawley}] ratio and the small RMSD. The latter ranged in value from 0.017 Å (4) to 0.204 Å (35), with a mean across all 35 structures equal to 0.093 ± 0.043 Å. This close agreement arises from the ability of the SA algorithm to `fine tune' both the internal and external DOF. In each case, the atomic displacements of the best SA solution are within the radius of convergence of a typical Rietveld refinement (see §4.2.3[link] for the example of the restrained Rietveld refinement of the SA solution for structure 35).

Critical examination of the fit to the diffraction data returned by the SA process is a key step in the structure determination process. The observation of a high [\chi ^{2}_{\rm profile} / \chi ^{2}_{\rm Pawley}] ratio or significant misfit in any region of the diffraction pattern, or unfavourably short atom–atom contacts, is diagnostic of problems that are best addressed at the structure solution stage, prior to refinement. In practice, this means checking for the possibility of PO and consulting other available experimental data for any evidence of disorder in the structure.

Three representative examples from each of the two populations identified above (127 and 2835) are now considered in §§4.1[link] and 4.2[link].

4.1. Compounds 127 with <15 DOF

4.1.1. Paracetamol form II

SDPD is a powerful means of solving the structures of metastable phases crystallized in situ in a glass capillary (Shankland et al., 2001[Shankland, N., David, W. I. F., Shankland, K., Kennedy, A. R., Frampton, C. S. & Florence, A. J. (2001). Chem. Commun. pp. 2204-2205.]). In the case of metastable orthorhombic form II paracetamol (8), a polycrystalline sample was readily obtained in a capillary by cooling molten paracetamol to room temperature. Unsurprisingly, oriented growth of crystallites within the capillary necessitated a significant March–Dollase correction of intensities for PO in the data (r = 1.50, [001]; r determined as an optimizable parameter in the SA runs) (Dollase, 1986[Dollase W. A. (1986). J. Appl. Cryst. 19, 267-272.]). With this PO correction, an accurate structure solution was obtained, with [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 6.8 and RMSD = 0.140 Å (cf. [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 22.2 and RMSD = 0.428 Å for the best SA solution with no PO correction included; Fig. 3[link]).

[Figure 3]
Figure 3
Overlay of the best SA solutions for 8, with (black) and without (grey) a March–Dollase PO correction of intensities included in the SA searches. Without the PO correction, the aromatic ring is tilted ca 10° out of its correct position and the C–O–N plane suffers a rotation of some 37° (H atoms in this and subsequent figures have been omitted for clarity).
4.1.2. Promazine hydrochloride

Promazine hydrochloride (25) falls in the mid-range of structural complexity represented by compounds 135. The diffraction data are particularly high quality (Fig. 4[link]) and yielded the lowest FWHM value in Table 4[link] (0.079°; cf. 0.059° for the 100 reflection of LaB6 collected under the same conditions). The structure was solved with 100% success, the best SA solution returning [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 4.3 and RMSD = 0.107 Å.

[Figure 4]
Figure 4
Observed profile (circles), calculated profile (line) and difference plot [(yobsycalc)/σ(yobs)] of the Pawley fit for 25 (Pawley χ2 = 2.41) in the range 6.7–51.5° 2θ. Inset: high-angle data in the range 31.5–51.5° 2θ showing the excellent fit to the data out to 1.76 Å resolution.
4.1.3. cis-Thiothixene

The initial runs on cis-Thiothixene (19) returned a narrow profile χ2 range (29.25–29.75; cf. [\chi ^{2}_{\rm Pawley}] = 3.13), but the high [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] ratio raised the strong suspicion that the global minimum had not been reached. Inclusion of a PO correction in the SA re-runs (r = 0.88, [010]) confirmed this to be the case, with the [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] ratio improving significantly to 7.3 (RMSD = 0.129 Å). A subsequent comparison of the initial structure with the global minimum structure showed that the orientation of the —N(CH3)2 group (i.e. eclipsing —SO2) in the former was incorrect.

4.2. Compounds 2835 with ≥15 DOF

4.2.1. 2-{[3-(2-Phenylethoxy)propyl]sulfonyl}ethyl benzo­ate

The single-crystal structure of compound 33 has a rotationally disordered phenyl ring, resulting in four of the C atoms being disordered over two sites. Accordingly, a Z matrix allowing for phenyl rotational disorder (19 DOF) was constructed by incorporating two independent half-occupancy phenyl rings.

The structure was solved successfully, the best SA solution yielding [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 4.1, RMSD = 0.165 Å and accurate orientations for the half-occupancy phenyl rings (Fig. 5[link]). As expected, a simplified model (18 DOF) with a fully ordered phenyl ring gave a solution that was largely correct, except for a compromise in the ring orientation (Fig. 5[link]). Thus, as has been found elsewhere with disordered fragments (Graham et al., 2004[Graham, D., Kennedy, A. R., McHugh, C. J., Smith, W. E., David, W. I. F., Shankland, K. & Shankland, N. (2004). New J. Chem. 28, 161-165.]; Johnston et al., 2004[Johnston, A., Florence, A. J., Shankland, K., Markvardsen, A., Shankland, N., Steele, G. & Cosgrove, S. D. (2004). Acta Cryst. E60, o1751-o1753.]), a significant improvement in the fit to the diffraction data can be derived by including fractional-occupancy atoms in the global optimization, albeit at the cost of reducing the frequency of success (Table 5[link]; eliminating one DOF in the fully ordered model increased the number of SA solutions with [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] < 10 from three to nine).

[Figure 5]
Figure 5
The best SA solution for 33 (black; disordered model) overlaid on the single-crystal structure (grey), showing the excellent agreement between corresponding half-occupancy phenyl ring positions. Inset: with the phenyl ring in the SA model set to full occupancy, the best solution (black) returned [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 4.4, with the ring positioned approximately midway between the disordered phenyl positions in the single-crystal structure (grey).

There is a good correspondence among the top six SA solutions ([\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 4.1–14.2) with regard to the positions of the phenyl ring centroids and the S atoms (Fig. 6[link]) and Fig. 7[link] confirms that the higher [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] ratios reflect a decrease in the accuracy with which the positions of the acyclic backbone atoms (other than the S atom) are located.

[Figure 6]
Figure 6
Comparison of the positions of the phenyl ring centroids and S atoms in the top six SA solutions for 33 and in the equivalent single-crystal structure (centroid 1 corresponds to the disordered phenyl ring).
[Figure 7]
Figure 7
Comparison of the best SA solution for 33 (top; [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 4.1) with the fourth-ranked structure in the batch of 20 SA runs (bottom: [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 10.9). The corresponding phenyl ring centroids in each structure are separated by 0.10 Å (disordered ring) and 0.20 Å, with the S atoms 0.17 Å apart. The numerals in the bottom structure indicate the separation (Å) from the corresponding atom in the top structure.

In summary, in those instances where an SA search fails to reach an acceptably low [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] ratio for any particular compound, it is probable that the regions of relatively high X-ray scattering power, at least, are located reliably, giving a partial structure around which further constrained SA runs may be instigated.

4.2.2. Capsaicin

Capsaicin (31; Nsol = 5) has two fewer internal DOF than disordered 33 (Nsol = 3). As was observed with 33, the aromatic rings are located with good reproducibility and the higher [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] ratios reflect a lower accuracy in locating the acyclic chain atoms (Fig. 8[link]).

[Figure 8]
Figure 8
Overlay of the top five SA solutions for 31, spanning the range [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 4.2–6.3.
4.2.3. Verapamil hydrochloride

The best solution for verapamil hydrochloride (35, 22 DOF) yielded [\chi ^{2}_{\rm profile}/\chi ^{2}_{\rm Pawley}] = 9.6 and RMSD = 0.204 Å. A restrained Rietveld refinement (TOPAS; Coelho, 2003[Coelho, A. A. (2003). J. Appl. Cryst. 36, 86-95.]) of the SA structure (Rwp = 6.3) against the raw data re-fitted to 65° 2θ (Pawley Rwp = 2.2) resulted in a significant improvement in accuracy, with a final Rwp of 3.1 and a reduced RMSD value of 0.134 Å, only slightly greater than the mean RMSD across all structures. A PO correction in the direction [100] was included in the refinement, the magnitude of the correction (r = 0.91) being consistent with mild PO in the sample.

Whilst the single answer required to `solve the structure' was obtained, the frequency with which low-lying areas of χ2 space were visited was low (Nsol = 1). In such instances, a better sampling of low-lying space can be achieved by increasing the number of SA runs, increasing the number of moves per run and decreasing the cooling rate (Table 1[link]). It is also reasonable to expect that the inclusion of modal torsional constraints will also increase the number of runs successfully locating low-lying regions of the search space (§1.2[link]) and this aspect is reported in §4.2.4[link].

4.2.4. Results of constrained SA structure determinations

The application of modal torsional constraints increased the frequency with which the structures of 28, 29, 31 and 34 were solved (Table 7[link]), yielding solutions with comparable RMSD values to those reported in Table 5[link]. In the case of verapamil hydrochloride (35), the first attempt to apply the modal constraints failed to yield any correct structures (lowest profile χ2 among 20 constrained SA runs equalled 110.51). Repeat batches of unconstrained runs indicated that the frequency of success for 35 was somewhat less than the 5% observed for the initial batch of 20 runs reported in Table 5[link]. Following this realization, the structure was solved reproducibly (in replicate batches) by combining a bigger batch size (minimum of 50 runs) and an increased number of SA moves per run (2 × 107) with a reduced SA cooling rate (0.01). Thereafter, the application of modal torsion constraints doubled the frequency with which the structure of 35 was solved, from Nsol = 1 to 2.

Table 7
Summary of results for SA runs constrained using the modal torsion constraints

Column headings are as defined in Table 5.[link]

Code Profile χ2 Nsol χ2 ratio RMSD (Å)
28 9.97–216.23 13 3.7 0.084
29 24.03–194.05 12 2.9 0.093
31 36.83–195.19 8 5.3 0.226
34 11.26–70.22 9 5.4 0.056
35 23.20–248.91 2 4.9 0.092
†Batch size = 50 runs for compound 35 (20 runs for the others). The corresponding batch of 50 unconstrained SA runs for 35 (§4.2.4[link]; cooling rate = 0.01, moves per run = 2 × 107) returned one correct solution in a profile χ2 range of 28.08–230.10.
‡Profile χ2 = 23.20 and 28.56.

5. Conclusions

At a time when the development of experimental methods for increasing the efficiency and throughput of drug development is a priority within the pharmaceutical industry, valuable savings in time, materials and analysis can be achieved by wider reliance on high-quality laboratory XRPD data and SDPD.

Given XRPD data collected to 2 Å resolution or better, the results of this investigation substantiate the following conclusions: (a) structures with <15 DOF present little challenge to the SA process, reproducibly yielding accurate structure solutions; (b) for structures with greater complexity (DOF = 15–20), where the preponderance of local minima in the agreement hypersurface reduces the frequency of success, the SA algorithm is still able to locate the global minimum with a reasonable frequency.

It is at higher levels of complexity (DOF > 20) where the greatest challenges remain, with structures such as 35 (DOF = 22) and AR-C69457CC (DOF = 26; Johnston et al., 2004[Johnston, A., Florence, A. J., Shankland, K., Markvardsen, A., Shankland, N., Steele, G. & Cosgrove, S. D. (2004). Acta Cryst. E60, o1751-o1753.]) representing the current state of the art of SDPD from laboratory XRPD data. Modal torsion-angle constraints can significantly increase the frequency of success and offer a convenient means by which to introduce prior chemical knowledge to reduce the size of the search space. Automation of the process of determining the constraints from a knowledge base of molecular geometry (MOGUL; Bruno et al., 2004[Bruno, I. J., Cole, J. C., Kessler, M., Luo, J., Motherwell, W. D. S., Purkis, L. H., Smith, B. R., Taylor, R., Cooper, R. I., Harris, S. E. & Orpen, A. G. (2004). J. Chem. Inf. Comput. Sci. In preparation.]) in the latest version of DASH (v3.0) should enable this knowledge to be used more routinely.

One might also consider the application of modified or alternative search algorithms such as parallel tempering (Hansmann, 1997[Hansmann, U. H. E. (1997). Chem. Phys. Lett. 281, 140-150.]) or hybrid Monte Carlo (Johnston et al., 2002[Johnston, J. C., Markvardsen, A. J., David, W. I. F. & Shankland, K. (2002). Acta Cryst. A58, 441-447.]), and/or different evaluation functions (e.g. maximum likelihood; Markvardsen et al., 2002[Markvardsen, A. J., David, W. I. F. & Shankland, K. (2002). Acta Cryst. A58, 316-326.]). Not all of these approaches have been implemented beyond the proof-of-concept stage in SDPD, and the challenge therefore remains to implement such approaches for routine structure solution.

Powder diffractometers at synchrotron sources continue to offer considerable advantages over laboratory-based instrumentation (e.g. increased incident flux and higher instrumental resolution) and these advantages often translate to an increased diffraction information content that allows more complex structures to be determined and refined.

Rietveld refinement of a solved structure can always be recommended, with the caveat that an improved fit to the diffraction data should not be pursued at the expense of chemical sense. This typically means the careful application of restraints, or perhaps rigid-body refinements. The latter option is similar to the simplex refinement at the end of a DASH run and thus it is not uncommon to find little improvement on refining an SA solution where the structure concerned is either relatively simple or substantially rigid.

Table 8
CSD refcodes and references

CSD refcode Reference
HCSBTZ Dupont & Dideberg (1972[Dupont, L. & Dideberg, O. (1972). Acta Cryst. B28, 2340-2347.]).
ZZZLWW01 Steiner (2000[Steiner, T. (2000). Acta Cryst. C56, 876-877.])
LIZCUS Raj et al. (2000[Raj, S. S. S., Fun, H. -K., Zhang, J., Xiong, R. -G. & You, X.-Z. (2000). Acta Cryst. C56, e274-e275.])
CBMZPN10 Himes et al. (1981[Himes, V. L., Mighell, A. D. & DeCamp, W. H. (1981). Acta Cryst. B37, 2242-2245.])
DAPSUO10 Alleaume (1967[Alleaume, M. (1967). Thesis, Bordeaux, France.])
HXACAN07 Nichols & Frampton (1998[Nichols, C. & Frampton, C. S. (1998). J. Pharm. Sci. 87, 684-693.])
HXACAN08 Nichols & Frampton (1998[Nichols, C. & Frampton, C. S. (1998). J. Pharm. Sci. 87, 684-693.])
ZZZMLY01 Hodgson & Asplund (1991[Hodgson, D. J. & Asplund, R. O. (1991). Acta Cryst. C47, 1986-1987.])
MCPRPL Fujinaga & James (1980[Fujinaga, M. & James, M. N. G. (1980). Acta Cryst. B36, 3196-3199.])
CLTPNI03 Brammer & Stevens (1989[Brammer, L. & Stevens, E. D. (1989). Acta Cryst. C45, 400-403.])
BHHPHE Beale & Grainger (1972[Beale, J. P. & Grainger, C. T. (1972). Cryst. Struct. Commun. 1, 71-74.])
GEBZUI Bamgboye & Sowerby (1986[Bamgboye, T. T. & Sowerby, D. B. (1986). Polyhedron, 5, 1487-1488.])
QQQAEJ01 Shankland et al. (1996[Shankland, N., Love, S. W., Watson, D. G., Knight, K. S., Shankland, K. & David, W. I. F. (1996). J. Chem. Soc. Faraday Trans. 92, 4555-4559.])
THTHXN01 David et al. (1998[David, W. I. F., Shankland, K. & Shankland, N. (1998). Chem. Commun. pp. 931-932.])
BEDMIG Koo et al. (1980[Koo, C. H., Cho, S. I. & Yeon, Y. H. (1980). Arch. Pharm. Res. 3, 37-49.])
CREATH03 Kato et al. (1979[Kato, Y., Haimoto, Y. & Sakurai, K. (1979). Bull. Chem. Soc. Jpn, 52, 233-234.])
CIMPRA Post & Horn (1977[Post, M. L. & Horn, A. S. (1977). Acta Cryst. B33, 2590-2595.])
LACTOS10 Fries et al. (1971[Fries, D. C., Rao, S. T. & Sundaralingam, M. (1971). Acta Cryst. B27, 994-1005.])
PROMZC01 David et al. (1998[David, W. I. F., Shankland, K. & Shankland, N. (1998). Chem. Commun. pp. 931-932.])
ZZZPUS02 Donaldson et al. (1981[Donaldson, J. D., Leary, J. R., Ross, S. D., Thomas, M. J. K. & Smith, C. H. (1981). Acta Cryst. B37, 2245-2248.])
CEYHUJ01 Kojic-Prodic et al. (1984[Kojic Prodic, B., Ruzic Toros, Z., Sunjic, V., Decorte, E. & Moimas, F. (1984). Helv. Chim. Acta, 67, 916-925.])
UCUVET Shankland et al. (2001[Shankland, N., David, W. I. F., Shankland, K., Kennedy, A. R., Frampton, C. S. & Florence, A. J. (2001). Chem. Commun. pp. 2204-2205.])
YAYWUQ Kennedy et al. (2001[Kennedy, A. R., Hughes, M. P., Monaghan, M. L., Staunton, E., Teat, S. J. & Smith, W. E. (2001). J. Chem. Soc. Dalton Trans. pp. 2199-2205.])
JEKNOC10 Freer et al. (1993[Freer, A. A., Bunyan, J. M., Shankland, N. & Sheen, D. B. (1993). Acta Cryst. C49, 1378-1380.])
CURHOM Carpy et al. (1985[Carpy, A., Leger, J.-M. & Melchiorre, C. (1985). Acta Cryst. C41, 624-627.])

Footnotes

1The recommended, and indeed simplest, way to construct an accurate Z matrix is from a related crystal structure, such as a polymorphic, salt or solvated form of the molecule of interest. In the absence of such a structure, mean values for bond lengths, covalent bond angles and non-optimizable torsion angles are preferably extracted from the CSD (Allen, 2002[Allen, F. H. (2002). Acta Cryst. B58, 380-388.]) and input into the Z matrix. This requirement to input the chemical formula and connectivity of fragments is, in fact, one of the drawbacks of global optimization methods for solving unknown crystal structures.

2A reliable indicator of how close the structure is to the global minimum is obtained by taking the ratio [\chi ^{2}_{\rm profile} / \chi ^{2}_{\rm Pawley}]; the smaller the ratio, the more likely it is that the correct solution has been obtained. Favourable values for this ratio typically range from 2 to 10, the higher ratios often indicating that additional details (such as PO or positional disorder) need to be factored into the model. In determining Nsol in Table 5[link], all structures with [\chi ^{2}_{\rm profile} / \chi ^{2}_{\rm Pawley}] < 10 were considered to be solved and the solutions were confirmed by subsequent comparison with the known crystal structure.

Acknowledgements

The authors gratefully acknowledge EPSRC for funding this research through grant GR/N07462/01 and AstraZeneca R&D Charnwood for financial support for AJ, and for compounds 14, 22 and 33.

References

First citationAlleaume, M. (1967). Thesis, Bordeaux, France.  Google Scholar
First citationAllen, F. H. (2002). Acta Cryst. B58, 380–388.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationBamgboye, T. T. & Sowerby, D. B. (1986). Polyhedron, 5, 1487–1488.  CSD CrossRef CAS Web of Science Google Scholar
First citationBeale, J. P. & Grainger, C. T. (1972). Cryst. Struct. Commun. 1, 71–74.  CAS Google Scholar
First citationBoultif, A. & Louër, D. (1991). J. Appl. Cryst. 24, 987–993.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBrammer, L. & Stevens, E. D. (1989). Acta Cryst. C45, 400–403.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBruno, I. J., Cole, J. C., Edgington, P. R., Kessler, M., Macrae, C. F., McCabe, P., Pearson, J. & Taylor, R. (2002). Acta Cryst. B58, 389–397.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationBruno, I. J., Cole, J. C., Kessler, M., Luo, J., Motherwell, W. D. S., Purkis, L. H., Smith, B. R., Taylor, R., Cooper, R. I., Harris, S. E. & Orpen, A. G. (2004). J. Chem. Inf. Comput. Sci. In preparation.  Google Scholar
First citationCarpy, A., Leger, J.-M. & Melchiorre, C. (1985). Acta Cryst. C41, 624–627.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationCoelho, A. A. (2003). J. Appl. Cryst. 36, 86–95.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDavid, W. I. F., Shankland, K., Cole, J., Maginn, S., Motherwell, W. D. S. & Taylor, R. (2001). DASH user manual. Cambridge Crystallographic Data Centre, Cambridge, UK.  Google Scholar
First citationDavid, W. I. F., Shankland, K. & Shankland, N. (1998). Chem. Commun. pp. 931–932.  Web of Science CSD CrossRef Google Scholar
First citationDeem, M. W. & Newsam, J. M. (1989). Nature (London), 342, 260–262.  CrossRef CAS Web of Science Google Scholar
First citationDinnebier, R. E., Wagner, M., Peters, F., Shankland, K. & David, W. I. F. (2000). Z. Anorg. Allg. Chem. 626, 1400–1405.  CrossRef CAS Google Scholar
First citationDollase W. A. (1986). J. Appl. Cryst. 19, 267–272.  CrossRef IUCr Journals Google Scholar
First citationDonaldson, J. D., Leary, J. R., Ross, S. D., Thomas, M. J. K. & Smith, C. H. (1981). Acta Cryst. B37, 2245–2248.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationDupont, L. & Dideberg, O. (1972). Acta Cryst. B28, 2340–2347.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationEdgar, M., Carter, V. J., Tunstall, D. P., Grewal, P., Favre-Nicolin, V., Cox, P. A., Lightfoot, P. & Wright, P. A. (2002). Chem. Commun. 8, 808–809.  Web of Science CSD CrossRef Google Scholar
First citationEngel, G. E., Wilke, S., König, O., Harris, K. D. M. & Leusen, F. J. J. (1999). J. Appl. Cryst. 32, 1169–1179.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFavre-Nicolin, V. & Cerny, R. (2002). J. Appl. Cryst. 35, 734–743.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFrampton, C. S. (2004). Unpublished single-crystal data.  Google Scholar
First citationFreer, A. A., Bunyan, J. M., Shankland, N. & Sheen, D. B. (1993). Acta Cryst. C49, 1378–1380.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationFries, D. C., Rao, S. T. & Sundaralingam, M. (1971). Acta Cryst. B27, 994–1005.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationFujinaga, M. & James, M. N. G. (1980). Acta Cryst. B36, 3196–3199.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationGolič, L., Djinović, K. & Florjanič, M. (1989). Acta Cryst. C45, 1381–1384.  CSD CrossRef Web of Science IUCr Journals Google Scholar
First citationGraham, D., Kennedy, A. R., McHugh, C. J., Smith, W. E., David, W. I. F., Shankland, K. & Shankland, N. (2004). New J. Chem. 28, 161–165.  Web of Science CSD CrossRef CAS Google Scholar
First citationHansmann, U. H. E. (1997). Chem. Phys. Lett. 281, 140–150.  Web of Science CrossRef CAS Google Scholar
First citationHarris, K. D. M. & Cheung, E. Y. (2003). Org. Process Res. Dev. 7, 970–976.  Web of Science CrossRef CAS Google Scholar
First citationHimes, V. L., Mighell, A. D. & DeCamp, W. H. (1981). Acta Cryst. B37, 2242–2245.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationHodgson, D. J. & Asplund, R. O. (1991). Acta Cryst. C47, 1986–1987.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationIvashevskaja, S. N., Aleshina, L. A., Andreev, V. P., Nizhnik, Y. P., Chernyshev V. V. & Schenk, H. (2002). Acta Cryst. C58, m300–m301.  CrossRef IUCr Journals Google Scholar
First citationJohnston, J. C., Markvardsen, A. J., David, W. I. F. & Shankland, K. (2002). Acta Cryst. A58, 441–447.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationJohnston, A., Florence, A. J., Shankland, K., Markvardsen, A., Shankland, N., Steele, G. & Cosgrove, S. D. (2004). Acta Cryst. E60, o1751–o1753.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationKato, Y., Haimoto, Y. & Sakurai, K. (1979). Bull. Chem. Soc. Jpn, 52, 233–234.  CrossRef CAS Web of Science Google Scholar
First citationKennedy, A. R., Hughes, M. P., Monaghan, M. L., Staunton, E., Teat, S. J. & Smith, W. E. (2001). J. Chem. Soc. Dalton Trans. pp. 2199–2205.  Web of Science CSD CrossRef Google Scholar
First citationKojic Prodic, B., Ruzic Toros, Z., Sunjic, V., Decorte, E. & Moimas, F. (1984). Helv. Chim. Acta, 67, 916–925.  CAS Google Scholar
First citationKoo, C. H., Cho, S. I. & Yeon, Y. H. (1980). Arch. Pharm. Res. 3, 37–49.  CrossRef CAS Google Scholar
First citationLe Bail, A. (2001). Mater. Sci. Forum, 378–381, 65–70.  CrossRef CAS Google Scholar
First citationMadsen, I. C. & Hill, R. J. (1994). J. Appl. Cryst. 27, 385–392.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationMarder, T. B. (2004). Unpublished single-crystal data.  Google Scholar
First citationMarkvardsen, A. J., David, W. I. F. & Shankland, K. (2002). Acta Cryst. A58, 316–326.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMiddleton, D. A., Peng, X., Saunders, D., Shankland, K., David, W. I. F. & Markvardsen, A. J. (2002). Chem. Commun. pp. 1976–1977.  Web of Science CrossRef Google Scholar
First citationNichols, C. & Frampton, C. S. (1998). J. Pharm. Sci. 87, 684–693.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationPawley, G. S. (1981). J. Appl. Cryst. 14, 357–361.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationPost, M. L. & Horn, A. S. (1977). Acta Cryst. B33, 2590–2595.  CSD CrossRef CAS IUCr Journals Google Scholar
First citationRaj, S. S. S., Fun, H. -K., Zhang, J., Xiong, R. -G. & You, X.-Z. (2000). Acta Cryst. C56, e274–e275.  CSD CrossRef CAS IUCr Journals Google Scholar
First citationReinaudi, L., Leiva, E. P. M. & Carbonio, R. E. (2000). J. Chem. Soc. Dalton Trans. 23, 4258–4262.  Web of Science CrossRef Google Scholar
First citationRukiah, M., Lefebvre, J., Descamps, M., Hemon, S. & Dzyabchenko, A. (2004). J. Appl. Cryst. 37, 464–471.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationShankland, K. & David, W. I. F. (2002). Structure Determination from Powder Diffraction Data, edited by W. I. F. David, K. Shankland, L. B. McCusker & Ch. Baerlocher, pp. 252–285. IUCr/Oxford University Press.  Google Scholar
First citationShankland, K., David, W. I. F., Csoka, T. & McBride, L. (1998). Int. J. Pharm. 165, 117–126.  Web of Science CrossRef CAS Google Scholar
First citationShankland, N., David, W. I. F., Shankland, K., Kennedy, A. R., Frampton, C. S. & Florence, A. J. (2001). Chem. Commun. pp. 2204–2205.  Web of Science CSD CrossRef Google Scholar
First citationShankland, K., David, W. I. F. & Sivia, D. (1997). J. Mater. Chem. 7, 569–572.  CSD CrossRef CAS Google Scholar
First citationShankland, N., Love, S. W., Watson, D. G., Knight, K. S., Shankland, K. & David, W. I. F. (1996). J. Chem. Soc. Faraday Trans. 92, 4555–4559.  CSD CrossRef CAS Web of Science Google Scholar
First citationShankland, K., McBride, L., David, W. I. F., Shankland, N. & Steele, G. (2002). J. Appl. Cryst. 35, 443–454.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationSteiner, T. (2000). Acta Cryst. C56, 876–877.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationZachariasen, W. H. & Ellinger, F. H. (1963). Acta Cryst. 16, 369–375.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationZaske, L., Perrin, M.-A., Daiguebonne, C. & Guillou, O. (2004). Mater. Sci. Forum, 443–444, 411–421.  CrossRef CAS Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Follow J. Appl. Cryst.
Sign up for e-alerts
Follow J. Appl. Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds