International
Tables for
Crystallography
Volume F
Crystallography of biological macromolecules
Edited by M. G. Rossmann and E. Arnold

International Tables for Crystallography (2006). Vol. F. ch. 22.4, pp. 558-574   | 1 | 2 |
https://doi.org/10.1107/97809553602060000713

Chapter 22.4. The relevance of the Cambridge Structural Database in protein crystallography

F. H. Allen,a* J. C. Colea and M. L. Verdonka

aCambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, England
Correspondence e-mail:  allen@ccdc.cam.ac.uk

References

First citation Abola, E. E., Sussman, J. L., Prilusky, J. & Manning, N. O. (1997). Protein Data Bank archives of three-dimensional macromolecular structures. Methods Enzymol. 277, 556–571.Google Scholar
First citation Allen, F. H., Baalham, C. A., Lommerse, J. P. M. & Raithby, P. R. (1998). Carbonyl–carbonyl interactions can be competitive with hydrogen bonds. Acta Cryst. B54, 320–329.Google Scholar
First citation Allen, F. H., Bird, C. M., Rowland, R. S. & Raithby, P. R. (1997a). Resonance-induced hydrogen bonding at sulfur acceptors in R1R2C=S and [R_{1}CS_{2}^{-}] systems. Acta Cryst. B53, 680–695.Google Scholar
First citation Allen, F. H., Bird, C. M., Rowland, R. S. & Raithby, P. R. (1997b). Hydrogen-bond acceptor and donor properties of divalent sulfur. Acta Cryst. B53, 696–701.Google Scholar
First citation Allen, F. H., Davies, J. E., Galloy, J. J., Johnson, O., Kennard, O., Macrae, C. F., Mitchell, E. M., Mitchell, G. F., Smith, J. M. & Watson, D. G. (1991). The development of versions 3 and 4 of the Cambridge Structural Database system. J. Chem. Inf. Comput. Sci. 31, 187–204.Google Scholar
First citation Allen, F. H., Doyle, M. J. & Auf der Heyde, T. P. E. (1991). Automated conformational analysis from crystallographic data. 6. Principal-component analysis for n-membered carbocyclic rings (n = 4, 5, 6): symmetry considerations and correlations with ring-puckering parameters. Acta Cryst. B47, 412–424.Google Scholar
First citation Allen, F. H., Doyle, M. J. & Taylor, R. (1991). Automated conformational analysis from crystallographic data. 3. Three-dimensional pattern recognition within the Cambridge Structural Database system: implementation and practical examples. Acta Cryst. B47, 50–61.Google Scholar
First citation Allen, F. H., Harris, S. E. & Taylor, R. (1996). Comparison of conformer distributions in the crystalline state with conformational energies calculated by ab initio techniques. J. Comput.-Aided Mol. Des. 10, 247–254.Google Scholar
First citation Allen, F. H., Howard, J. A. K. & Pitchford, N. A. (1996). Symmetry-modified conformational mapping and classification of the medium rings from crystallographic data. IV. Cyclooctane and related eight-membered rings. Acta Cryst. B52, 882–891.Google Scholar
First citation Allen, F. H., Kennard, O., Watson, D. G., Brammer, L., Orpen, A. G. & Taylor, R. (1987). Tables of bond lengths determined by X-ray and neutron diffraction. Part 1. Bond lengths in organic compounds. J. Chem. Soc. Perkin Trans. 2, pp. S1–S19.Google Scholar
First citation Allen, F. H., Lommerse, J. P. M., Hoy, V. J., Howard, J. A. K. & Desiraju, G. R. (1997). Halogen···O(nitro) supramolecular synthon in crystal engineering: a combined crystallographic database and ab initio molecular orbital study. Acta Cryst. B53, 1006–1016.Google Scholar
First citation Allen, F. H., Motherwell, W. D. S., Raithby, P. R., Shields, G. P. & Taylor, R. (1999). Systematic analysis of the probabilities of formation of bimolecular hydrogen bonded ring motifs in organic crystal structures. New J. Chem. 23, 25–34.Google Scholar
First citation Allen, F. H., Raithby, P. R., Shields, G. P. & Taylor, R. (1998). Probabilities of formation of bimolecular cyclic hydrogen bonded motifs in organic crystal structures: a systematic database study. Chem. Commun. pp. 1043–1044.Google Scholar
First citation Allen, F. H., Rowland, R. S., Fortier, S. & Glasgow, J. I. (1990). Knowledge acquisition from crystallographic databases: towards a knowledge-based approach to molecular scene analysis. Tetrahedron Comput. Methodol. 3, 757–774.Google Scholar
First citation Ashida, T., Tsunogae, Y., Tanaka, I. & Yamane, T. (1987). Peptide chain structure parameters, bond angles and conformation angles from the Cambridge Structural Database. Acta Cryst. B43, 212–218.Google Scholar
First citation Balasubramanian, R., Chidambaram, R. & Ramachandran, G. N. (1970). Potential functions for hydrogen-bond interactions. II. Formulation of an empirical potential function. Biochim. Biophys. Acta, 221, 196–206.Google Scholar
First citation Berkovitch-Yellin, Z. & Leiserowitz, L. (1984). The role played by C—H···O and C—H···N interactions in determining molecular packing and conformation. Acta Cryst. B40, 159–165.Google Scholar
First citation Berman, H. M., Olson, W. K., Beveridge, D. L., Westbrook, J., Gelbin, A., Demeny, T., Hsieh, S.-H., Srinivasan, A. R. & Schneider, B. (1992). The Nucleic Acid Database. A comprehensive relational database of three-dimensional structures of nucleic acids. Biophys. J. 63, 751–759.Google Scholar
First citation Berman, H. M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T. N., Weissig, H., Shindyalov, I. N. & Bourne, P. E. (2000). The Protein Data Bank. Nucleic Acids Res. 28, 235–242.Google Scholar
First citation Bertolasi, V., Gilli, P., Ferretti, V. & Gilli, G. (1996). Resonance-assisted O—H···O hydrogen bonding. Its role in the crystalline self-recognition of beta-diketone enols and its structural and IR characterisation. Chem. Eur. J. 2, 925–934.Google Scholar
First citation Blundell, T. L., Sibanda, B. L., Sternberg, M. J. E. & Thornton, J. M. (1987). Knowledge-based prediction of protein structures and the design of novel molecules. Nature (London), 326, 347–352.Google Scholar
First citation Böhm, H.-J. (1992a). The computer program LUDI: a new method for the de novo design of enzyme inhibitors. J. Comput.-Aided Mol. Des. 6, 61–78.Google Scholar
First citation Böhm, H.-J. (1992b). LUDI: rule-based automatic design of new substituents for enzyme inhibitors. J. Comput.-Aided Mol. Des. 6, 593–606.Google Scholar
First citation Bondi, A. (1964). van der Waals volumes and radii. J. Phys. Chem. 68, 441–452.Google Scholar
First citation Boström, J., Norrby, P.-O. & Liljefors, T. (1998). Conformational energy penalties of protein-bound ligands. J. Comput.-Aided Mol. Des. 12, 383–396.Google Scholar
First citation Bower, M., Cohen, F. E. & Dunbrack, R. L. Jr (1997). Prediction of protein side-chain rotamers from a backbone-dependent rotamer library. J. Mol. Biol. 267, 1268–1282.Google Scholar
First citation Brammer, L., Zhao, D., Ladipo, F. T. & Braddock-Wilking, J. (1995). Hydrogen bonds involving transition metal centres – a brief review. Acta Cryst. B51, 632–640.Google Scholar
First citation Bruno, I. J., Cole, J. C., Lommerse, J. P. M., Rowland, R. S., Taylor, R. & Verdonk, M. L. (1997). IsoStar: a library of information about nonbonded interactions. J. Comput.-Aided Mol. Des. 11, 525–537.Google Scholar
First citation Bürgi, H.-B. & Dunitz, J. D. (1983). From crystal statics to chemical dynamics. Acc. Chem. Res. 16, 153–161.Google Scholar
First citation Bürgi, H.-B. & Dunitz, J. D. (1988). Can statistical analysis of structural parameters from different crystal environments lead to quantitative energy relationships? Acta Cryst. B44, 445–451.Google Scholar
First citation Bürgi, H.-B. & Dunitz, J. D. (1994). Structure correlation. Weinheim: VCH Publishers.Google Scholar
First citation Carbonell, J. (1989). Editor. Machine learning – paradigms and methods. Amsterdam: Elsevier.Google Scholar
First citation Carrell, A. B., Shimoni, L., Carrell, C. J., Bock, C. W., Murray-Rust, P. & Glusker, J. P. (1993). The stereochemistry of the recognition of nitrogen-containing heterocycles by hydrogen bonding and by metal ions. Receptor, 3, 57–76.Google Scholar
First citation Carrell, C. J., Carrell, H. L., Erlebacher, J. & Glusker, J. P. (1988). Structural aspects of metal ion–carboxylate interactions. J. Am. Chem. Soc. 110, 8651–8656.Google Scholar
First citation Ceccarelli, C., Jeffrey, G. A. & Taylor, R. (1981). A survey of O—H···O hydrogen bond geometries determined by neutron diffraction. J. Mol. Struct. 70, 255–271.Google Scholar
First citation Chakrabarti, P. (1990a). Interaction of metal ions with carboxylic and carboxamide groups in protein structures. Protein. Eng. 4, 49–56.Google Scholar
First citation Chakrabarti, P. (1990b). Geometry of interaction of metal ions with histidine residues in protein structures. Protein Eng. 4, 57–63.Google Scholar
First citation Chakrabarti, P. & Dunitz, J. D. (1982). Directional preferences of ether O-atoms towards alkali and alkaline earth cations. Helv. Chim. Acta, 65, 1482–1488.Google Scholar
First citation Chatfield, C. & Collins, A. J. (1980). Introduction to multivariate analysis. London: Chapman & Hall.Google Scholar
First citation Conklin, D., Fortier, S., Glasgow, J. I. & Allen, F. H. (1996). Conformational analysis from crystallographic data using conceptual clustering. Acta Cryst. B52, 535–549.Google Scholar
First citation Cremer, D. & Pople, J. A. (1975). A general definition of ring puckering coordinates. J. Am. Chem. Soc. 97, 1354–1358.Google Scholar
First citation Deane, C. M., Allen, F. H., Taylor, R. & Blundell, T. L. (1999). Carbonyl–carbonyl interactions stabilise the partially allowed Ramachandran conformations of asparagine and aspartic acid. Protein Eng. 12, 1025–1028.Google Scholar
First citation Derewenda, Z. S., Lee, L. & Derewenda, U. (1995). The occurrence of C—H···O hydrogen bonds in proteins. J. Mol. Biol. 252, 248–262.Google Scholar
First citation Desiraju, G. R. (1989). Crystal engineering: the design of organic solids. New York: Academic Press.Google Scholar
First citation Desiraju, G. R. (1991). The C—H···O hydrogen bond in crystals. What is it? Acc. Chem. Res. 24, 270–276.Google Scholar
First citation Desiraju, G. R. (1995). Supramolecular synthons in crystal engineering – a new organic synthesis. Angew. Chem. Int. Ed. Engl. 34, 2311–2327.Google Scholar
First citation Desiraju, G. R., Kashino, S., Coombs, M. M. & Glusker, J. P. (1993). C—H···O packing motifs in some cyclopenta[a]phenanthrenes. Acta Cryst. B49, 880–892.Google Scholar
First citation Desiraju, G. R. & Murty, B. N. (1987). Correlation between crystallographic and spectroscopic properties for C—H···O bonds in terminal acetylenes. Chem. Phys. Lett. 139, 360–361.Google Scholar
First citation Donohue, J. (1952). The hydrogen bond in organic crystals. J. Phys. Chem. 56, 502–510.Google Scholar
First citation Donohue, J. (1968). Selected topics in hydrogen bonding. In Structural chemistry and molecular biology, edited by W. Davidson & E. Rich, pp. 443–465. San Francisco: W. H. Freeman.Google Scholar
First citation Dunbrack, R. L. Jr & Karplus, M. (1993). Backbone-dependent rotamer library for proteins: applications to side-chain prediction. J. Mol. Biol. 230, 534–571.Google Scholar
First citation Dunitz, J. D. & Taylor, R. (1997). Organic fluorine hardly ever accepts hydrogen bonds. Chem. Eur. J. 3, 83–90.Google Scholar
First citation Einspahr, H. & Bugg, C. E. (1981). The geometry of calcium–carboxylate interactions in crystalline complexes. Acta Cryst. B37, 1044–1052.Google Scholar
First citation Engh, R. A. & Huber, R. (1991). Accurate bond and angle parameters for X-ray protein structure refinement. Acta Cryst. A47, 392–400.Google Scholar
First citation Everitt, B. (1980). Cluster analysis. New York: Wiley.Google Scholar
First citation Fortier, S., Castleden, I., Glasgow, J., Conklin, D., Walmsley, C., Leherte, L. & Allen, F. H. (1993). Molecular scene analysis: the integration of direct-methods and artificial-intelligence strategies for solving protein crystal structures. Acta Cryst. D49, 168–178.Google Scholar
First citation Glusker, J. P. (1980). Citrate conformation and chelation: enzymatic implications. Acc. Chem. Res. 13, 345–352.Google Scholar
First citation Glusker, J. P., Lewis, M. & Rossi, M. (1994). Crystal structure analysis for chemists and biologists. Weinheim: VCH Publishers.Google Scholar
First citation Goodford, P. J. (1985). A computational procedure for determining energetically favourable binding sites on biologically important molecules. J. Med. Chem. 28, 849–857.Google Scholar
First citation Gould, R. O., Gray, A. M., Taylor, P. & Walkinshaw, M. D. (1985). Crystal environments and geometries of leucine, isoleucine, valine and phenylalanine provide estimates of minimum nonbonded contact and preferred van der Waals interaction distances. J. Am. Chem. Soc. 107, 5921–5927.Google Scholar
First citation Hall, S. R., Allen, F. H. & Brown, I. D. (1991). The crystallographic information file (CIF): a new standard archive file for crystallography. Acta Cryst. A47, 655–685.Google Scholar
First citation Hayes, I. C. & Stone, A. J. (1984). An intermolecular perturbation theory for the region of moderate overlap. J. Mol. Phys. 53, 83–105.Google Scholar
First citation Hooft, R. W. W., Vriend, G., Sander, C. & Abola, E. E. (1996). Errors in protein structures. Nature (London), 381, 272.Google Scholar
First citation Jeffrey, G. A. & Maluszynska, H. (1982). A survey of hydrogen bond geometries in the crystal structures of amino acids. Int. J. Biol. Macromol. 4, 173–185.Google Scholar
First citation Jeffrey, G. A. & Maluszynska, H. (1986). A survey of the geometry of hydrogen bonds in the crystal structures of barbiturates, purines and pyrimidines. J. Mol. Struct. 147, 127–142.Google Scholar
First citation Jeffrey, G. A. & Mitra, J. (1984). Three-centre (bifurcated) hydrogen bonding in the crystal structures of amino acids. J. Am. Chem. Soc. 106, 5546–5553.Google Scholar
First citation Jeffrey, G. A. & Saenger, W. (1991). Hydrogen bonding in biological structures. Berlin: Springer Verlag.Google Scholar
First citation Jones, G., Willett, P. & Glen, R. C. (1995). Molecular recognition of receptor sites using a genetic algorithm with a description of solvation. J. Mol. Biol. 245, 43–53.Google Scholar
First citation Jones, G., Willett, P., Glen, R. C., Leach, A. R. & Taylor, R. (1997). Development and validation of a genetic algorithm for flexible docking. J. Mol. Biol. 267, 727–748.Google Scholar
First citation Joris, L., Schleyer, P. & Gleiter, R. (1968). Cyclopropane rings as proton acceptor groups in hydrogen bonding. J. Am. Chem. Soc. 90, 327–336.Google Scholar
First citation Kennard, O. (1962). In International tables for X-ray crystallography, Vol. II. Birmingham: Kynoch Press.Google Scholar
First citation Kennard, O. & Allen, F. H. (1993). The Cambridge Crystallographic Data Centre. Chem. Des. Autom. News, 8, 1, 31–37.Google Scholar
First citation Klebe, G. (1994). The use of composite crystal-field environments in molecular recognition and the de novo design of protein ligands. J. Mol. Biol. 237, 212–235.Google Scholar
First citation Klebe, G. & Mietzner, T. (1994). A fast and efficient method to generate biologically relevant conformations. J. Comput.-Aided Mol. Des. 8, 583–594.Google Scholar
First citation Kroon, J. & Kanters, J. A. (1974). Non-linearity of hydrogen bonds in molecular crystals. Nature (London), 248, 667–669.Google Scholar
First citation Kroon, J., Kanters, J. A., van Duijneveldt-van de Rijdt, J. G. C. M., van Duijneveldt, F. B. & Vliegenthart, J. A. (1975). O—H···O hydrogen bonds in molecular crystals: a statistical and quantum chemical analysis. J. Mol. Struct. 24, 109–129.Google Scholar
First citation Kuntz, I. D., Meng, E. C. & Stoichet, B. K. (1994). Structure-based molecular design. Acc. Chem. Res. 27, 117–122.Google Scholar
First citation Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993). PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Cryst. 26, 283–291.Google Scholar
First citation Laskowski, R. A., Thornton, J. M., Humblet, C. & Singh, J. (1996). X-SITE: use of empirically derived atomic packing preferences to identify favourable interaction regions in the binding sites of proteins. J. Mol. Biol. 259, 175–201.Google Scholar
First citation Lehn, J.-M. (1988). Perspectives in supramolecular chemistry – from molecular recognition towards molecular information processing and self-organization. Angew. Chem. Int. Ed. Engl. 27, 90–112.Google Scholar
First citation Lemmen, C. & Lengauer, T. (1997). Time-efficient flexible superposition of medium sized molecules. J. Comput.-Aided Mol. Des. 11, 357–368.Google Scholar
First citation Levitt, M. & Perutz, M. (1988). Aromatic rings act as hydrogen bond acceptors. J. Mol. Biol. 201, 751–754.Google Scholar
First citation Lommerse, J. P. M., Price, S. L. & Taylor, R. (1997). Hydrogen bonding of carbonyl, ether and ester oxygen atoms with alkanol hydroxyl groups. J. Comput. Chem. 18, 757–780.Google Scholar
First citation Lommerse, J. P. M., Stone, A. J., Taylor, R. & Allen, F. H. (1996). The nature and geometry of intermolecular interactions between halogens and oxygen or nitrogen. J. Am. Chem. Soc. 118, 3108–3116.Google Scholar
First citation Maccallum, P. H., Poet, R. & Milner-White, E. J. (1995a). Coulombic interactions between partially charged main-chain atoms not hydrogen bonded to each other influence the conformations of α-helices and antiparallel β-sheets. J. Mol. Biol. 248, 361–373.Google Scholar
First citation Maccallum, P. H., Poet, R. & Milner-White, E. J. (1995b). Coulombic interactions between partially charged main-chain atoms stabilise the right-handed twist found in most β-strands. J. Mol. Biol. 248, 374–384.Google Scholar
First citation Miller, J., McLachlan, A. D. & Klug, A. (1985). Repetitive zinc-binding domains in the protein transcription factor IIIA from Xenopus oocytes. EMBO J. 4, 1609–1614.Google Scholar
First citation Murray-Rust, P. & Bland, R. (1978). Computer retrieval and analysis of molecular geometry. II. Variance and its interpretation. Acta Cryst. B34, 2527–2533.Google Scholar
First citation Murray-Rust, P. & Glusker, J. P. (1984). Directional hydrogen bonding to sp2 and sp3 hybridized oxygen atoms and its relevance to ligand–macromolecule interactions. J. Am. Chem. Soc. 105, 1018–1025.Google Scholar
First citation Murray-Rust, P. & Motherwell, S. (1978). Computer retrieval and analysis of molecular geometry. III. Geometry of the β-1′-aminofuranoside fragment. Acta Cryst. B34, 2534–2546.Google Scholar
First citation Nicklaus, M. C., Wang, S., Driscoll, J. S. & Milne, G. W. A. (1995). Conformational changes of small molecules binding to proteins. Bioorg. Med. Chem. 3, 411–428.Google Scholar
First citation Nobeli, I., Price, S. L., Lommerse, J. P. M. & Taylor, R. (1997). Hydrogen bonding properties of oxygen and nitrogen acceptors in aromatic heterocycles. J. Comput. Chem. 18, 2060–2074.Google Scholar
First citation Nyburg, S. C. & Faerman, C. H. (1985). A revision of van der Waals atomic radii for molecular crystals: N, O, F, S, Cl, Se, Br and I bonded to carbon. Acta Cryst. B41, 274–279.Google Scholar
First citation Orpen, A. G., Brammer, L., Allen, F. H., Kennard, O., Watson, D. G. & Taylor, R. (1989). Tables of bond lengths determined by X-ray and neutron diffraction. Part II: organometallic compounds and coordination complexes of the d- and f-block metals. J. Chem. Soc. Dalton Trans. pp. S1–S83.Google Scholar
First citation Pauling, L. (1939). The nature of the chemical bond. Ithaca: Cornell University Press.Google Scholar
First citation Pedireddi, V. R. & Desiraju, G. R. (1992). A crystallographic scale of carbon acidity. Chem. Commun. pp. 988–990.Google Scholar
First citation Pimentel, G. C. & McClellan, A. L. (1960). The hydrogen bond. San Francisco: W. H. Freeman.Google Scholar
First citation Price, S. L., Stone, A. J., Lucas, J., Rowland, R. S. & Thornley, A. E. (1994). The nature of —Cl···Cl— intermolecular interactions. J. Am. Chem. Soc. 116, 4910–4918.Google Scholar
First citation Rappoport, Z., Biali, S. E. & Kaftory, M. (1990). Application of the structure correlation method to ring-flip processes in benzophenone. J. Am. Chem. Soc. 112, 7742–7750.Google Scholar
First citation Rarey, M., Kramer, B., Lengauer, T. & Klebe, G. (1996). Predicting receptor-ligand interactions by an incremental construction algorithm. J. Mol. Biol. 261, 470–481.Google Scholar
First citation Robertson, J. M. (1953). Organic crystals and molecules. Ithaca: Cornell University Press.Google Scholar
First citation Rosenfield, R. E., Parthasarathy, R. & Dunitz, J. D. (1977). Directional preferences of non-bonded atomic contacts with divalent sulphur. 1. Electrophiles and nucleophiles. J. Am. Chem. Soc. 99, 4860–4862.Google Scholar
First citation Rosenfield, R. E. Jr, Swanson, S. M., Meyer, E. F. Jr, Carrell, H. L. & Murray-Rust, P. (1984). Mapping the atomic environment of functional groups: turning 3D scatterplots into pseudo-density contours. J. Mol. Graphics, 2, 43–46.Google Scholar
First citation Rowland, R. S. & Taylor, R. (1996). Intermolecular nonbonded contact distances in organic crystal structures: comparison with distances expected from van der Waals radii. J. Phys. Chem. 100, 7384–7391.Google Scholar
First citation Schweizer, W. B. & Dunitz, J. D. (1982). Structural characteristics of the carboxylic acid ester group. Helv. Chim. Acta, 65, 1547–1552.Google Scholar
First citation Singh, J., Saldanha, J. & Thornton, J. M. (1991). A novel method for the modelling of peptide ligands to their receptors. Protein Eng. 4, 251–261.Google Scholar
First citation Steiner, T., Kanters, J. A. & Kroon, J. (1996). Acceptor directionality of sterically unhindered C—H···O=C hydrogen bonds donated by acidic C—H groups. J. Chem. Soc. Chem. Commun. 11, 1277–1278.Google Scholar
First citation Steiner, T. & Saenger, W. (1992). Geometry of C—H···O hydrogen bonds in carbohydrate crystal structures. Analysis of neutron diffraction data. J. Am. Chem. Soc. 114, 10146–10154.Google Scholar
First citation Steiner, T., Starikov, E. B., Amado, A. M. & Teixeira-Dias, J. J. C. (1995). Weak hydrogen bonding. Part 2. The hydrogen-bonding nature of short C—H··· · contacts. Crystallographic, spectroscopic and quantum mechanistic studies of some terminal alkynes. J. Chem. Soc. Perkin Trans. 2, pp. 1312–1326.Google Scholar
First citation Strynadka, N. C. J. & James, M. N. G. (1989). Crystal structures of the helix-loop-helix calcium-binding proteins. Annu. Rev. Biochem. 58, 951–998.Google Scholar
First citation Sutton, L. E. (1956). Tables of interatomic distances and configuration in molecules and ions. Special Publication No. 11. London: The Chemical Society.Google Scholar
First citation Sutton, L. E. (1959). Tables of interatomic distances and configuration in molecules and ions (supplement). Special Publication No. 18. London: The Chemical Society.Google Scholar
First citation Taylor, R. (1986). The Cambridge Structural Database in molecular graphics: techniques for the rapid identification of conformational minima. J. Mol. Graphics, 4, 123–131.Google Scholar
First citation Taylor, R. & Allen, F. H. (1994). Statistical and numerical methods of data analysis. In Structure correlation, edited by H.-B. Bürgi & J. D. Dunitz. Weinheim: VCH Publishers.Google Scholar
First citation Taylor, R. & Kennard, O. (1982). Crystallographic evidence for the existence of C—H···O, C—H···N and C—H···Cl hydrogen bonds. J. Am. Chem. Soc. 104, 5063–5070.Google Scholar
First citation Taylor, R. & Kennard, O. (1983). Comparison of X-ray and neutron diffraction results for the N—H···O=C hydrogen bond. Acta Cryst. B39, 133–138.Google Scholar
First citation Taylor, R., Kennard, O. & Versichel, W. (1983). Geometry of the N—H···O=C hydrogen bond. 1. Lone-pair directionality. J. Am. Chem. Soc. 105, 5761–5766.Google Scholar
First citation Taylor, R., Kennard, O. & Versichel, W. (1984a). Geometry of the N—H···O=C hydrogen bond. 2. Three-centre (bifurcated) and four-centre (trifurcated) bonds. J. Am. Chem. Soc. 106, 244–248.Google Scholar
First citation Taylor, R., Kennard, O. & Versichel, W. (1984b). Geometry of the N—H···O=C hydrogen bond. 3. Hydrogen-bond distances and angles. Acta Cryst. B40, 280–288.Google Scholar
First citation Taylor, R., Mullaley, A. & Mullier, G. W. (1990). Use of crystallographic data in searching for isosteric replacements: composite field environments of nitro and carbonyl groups. Pestic. Sci. 29, 197–213.Google Scholar
First citation Thornton, J. M. & Gardner, S. P. (1989). Protein motifs and database searching. Trends Biochem. Sci. 14, 300–304.Google Scholar
First citation Tintelnot, M. & Andrews, P. (1989). Geometries of functional group interactions in enzyme–ligand complexes: guides for receptor modelling. J. Comput.-Aided Mol. Des. 3, 67–84.Google Scholar
First citation Verdonk, M. L. (1998). Unpublished results.Google Scholar
First citation Verdonk, M. L., Cole, J. C. & Taylor, R. (1999). SuperStar: a knowledge-based approach for identifying interaction sites in proteins. J. Mol. Biol. 289, 1093–1108.Google Scholar
First citation Viswamitra, M. A., Radhakrishnan, R., Bandekar, J. & Desiraju, G. R. (1993). Evidence for O—H···C and N—H···C hydrogen bonding. J. Am. Chem. Soc. 115, 4868–4869.Google Scholar
First citation Whitesides, G. M., Simanek, E. E., Mathias, J. P., Seto, C. T., Chin, D. N., Mammen, M. & Gordon, D. M. (1995). Non-covalent synthesis: using physical-organic chemistry to make aggregates. Acc. Chem. Res. 28, 37–43.Google Scholar