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. 1.3, pp. 10-25   | 1 | 2 |
https://doi.org/10.1107/97809553602060000656

Chapter 1.3. Macromolecular crystallography and medicine

W. G. J. Hola* and C. L. M. J. Verlindea

aBiomolecular Structure Center, Department of Biological Structure, Howard Hughes Medical Institute, University of Washington, Seattle, WA 98195-7742, USA
Correspondence e-mail:  hol@gouda.bmsc.washington.edu

References

First citation Achari, A., Somers, D. O., Champness, J. N., Bryant, P. K., Rosemond, J. & Stammers, D. K. (1997). Crystal structure of the anti-bacterial sulfonamide drug target dihydropteroate synthase. Nature Struct. Biol. 4, 490–497.Google Scholar
First citation Adman, E. T., Stenkamp, R. E., Sieker, L. C. & Jensen, L. H. (1978). A crystallographic model for azurin at 3 Å resolution. J. Mol. Biol. 123, 35–47.Google Scholar
First citation Aertgeerts, K., De Bondt, H. L., De Ranter, C. J. & Declerck, P. J. (1995). Mechanisms contributing to the conformational and functional flexibility of plasminogen activator inhibitor-1. Nature Struct. Biol. 2, 891–897.Google Scholar
First citation Ævarsson, A., Chuang, J. L., Wynn, R. M., Turley, S., Chuang, D. T. & Hol, W. G. J. (2000). Crystal structure of human branched-chain α-ketoacid dehydrogenase and the molecular basis of multienzyme complex deficiency in maple syrup urine disease. Struct. Fold. Des. 8, 277–291.Google Scholar
First citation Allaire, M., Chernaia, M. M., Malcolm, B. A. & James, M. N. (1994). Picornaviral 3C cysteine proteinases have a fold similar to chymotrypsin-like serine proteinases. Nature (London), 369, 72–76.Google Scholar
First citation Allured, V. S., Collier, R. J., Carroll, S. F. & McKay, D. B. (1986). Structure of exotoxin A of Pseudomonas aeruginosa at 2.0-Å resolution. Proc. Natl Acad. Sci. USA, 83, 1320–1324.Google Scholar
First citation Almassy, R. J. & Dickerson, R. E. (1978). Pseudomonas cytochrome c551 at 2.0 Å resolution: enlargement of the cytochrome c family. Proc. Natl Acad. Sci. USA, 75, 2674–2678.Google Scholar
First citation Amos, L. A. & Lowe, J. (1999). How Taxol stabilises microtubule structure. Chem. Biol. 6, R65–R69.Google Scholar
First citation Arnold, E., Das, K., Ding, J., Yadav, P. N., Hsiou, Y., Boyer, P. L. & Hughes, S. H. (1996). Targeting HIV reverse transcriptase for anti-AIDS drug design: structural and biological considerations for chemotherapeutic strategies. Drug Des. Discov. 13, 29–47.Google Scholar
First citation Arnold, E. & Rossmann, M. G. (1988). The use of molecular-replacement phases for the refinement of the human rhinovirus 14 structure. Acta Cryst. A44, 270–283.Google Scholar
First citation Arnold, E. & Rossmann, M. G. (1990). Analysis of the structure of a common cold virus, human rhinovirus 14, refined at a resolution of 3.0 Å. J. Mol. Biol. 211, 763–801.Google Scholar
First citation Arnold, G. F. & Arnold, E. (1999). Using combinatorial libraries to develop vaccines. ASM News, 65, 603–610.Google Scholar
First citation Arnold, G. F., Resnick, D. A., Li, Y., Zhang, A., Smith, A. D., Geisler, S. C., Jacobo-Molina, A., Lee, W., Webster, R. G. & Arnold, E. (1994). Design and construction of rhinovirus chimeras incorporating immunogens from polio, influenza, and human immunodeficiency viruses. Virology, 198, 703–708.Google Scholar
First citation Arnoux, P., Haser, R., Izadi, N., Lecroisey, A., Delepierre, M., Wandersman, C. & Czjzek, M. (1999). The crystal structure of HasA, a hemophore secreted by Serratia marcescens. Nature Struct. Biol. 6, 516–520.Google Scholar
First citation Athanasiadis, A., Vlassi, M., Kotsifaki, D., Tucker, P. A., Wilson, K. S. & Kokkinidis, M. (1994). Crystal structure of PvuII endonuclease reveals extensive structural homologies to EcoRV. Nature Struct. Biol. 1, 469–475.Google Scholar
First citation Baca, A. M., Sirawaraporn, R., Turley, S., Athappilly, F., Sirawaraporn, W. & Hol, W. G. J. (2000). Crystal structure of Mycobacterium tuberculosis 6-hydroxymethyl-1,8-dihydropteroate synthase in complex with pterin monophosphate: new insight into the enzymatic mechanism and sulfa-drug action. J. Mol. Biol. 302, 1193–1212.Google Scholar
First citation Baldwin, E. T., Bhat, T. N., Gulnik, S., Hosur, M. V., Sowder, R. C. I., Cachau, R. E., Collins, J., Silva, A. M. & Erickson, J. W. (1993). Crystal structures of native and inhibited forms of human cathepsin D: implications for lysosomal targeting and drug design. Proc. Natl Acad. Sci. USA, 90, 6796–6800.Google Scholar
First citation Baldwin, E. T., Weber, I. T., St Charles, R., Xuan, J. C., Appella, E., Yamada, M., Matsushima, K., Edwards, B. F., Clore, G. M. & Gronenborn, A. M. (1991). Crystal structure of interleukin 8: symbiosis of NMR and crystallography. Proc. Natl Acad. Sci. USA, 88, 502–506.Google Scholar
First citation Baldwin, J. J., Ponticello, G. S., Anderson, P. S., Christy, M. E., Murcko, M. A., Randall, W. C., Schwam, H., Sugrue, M. F., Springer, J. P., Gautheron, P., Grove, J., Mallorga, P., Viader, M. P., McKeever, B. M. & Navia, M. A. (1989). Thienothiopyran-2-sulfonamides: novel topically active carbonic anhydrase inhibitors for the treatment of glaucoma. J. Med. Chem. 32, 2510–2513.Google Scholar
First citation Ban, N., Nissen, P., Hansen, J., Capel, M., Moore, P. B. & Steitz, T. A. (1999). Placement of protein and RNA structures into a 5 Å-resolution map of the 50S ribosomal subunit. Nature (London), 400, 841–847.Google Scholar
First citation Banbula, A., Potempa, J., Travis, J., Fernandez-Catalan, C., Mann, K., Huber, R., Bode, W. & Medrano, F. (1998). Amino-acid sequence and three-dimensional structure of the Staphylococcus aureus metalloproteinase at 1.72 Å resolution. Structure, 6, 1185–1193.Google Scholar
First citation Banner, D. W., D'Arcy, A., Chene, C., Winkler, F. K., Guha, A., Konigsberg, W. H., Nemreson, Y. & Kirchhofer, D. (1996). The crystal structure of the complex of blood coagulation factor VIIa with soluble tissue factor. Nature, 380, 41–46.Google Scholar
First citation Banner, D. W., D'Arcy, A., Janes, W., Gentz, R., Schoenfeld, H. J., Broger, C., Loetscher, H. & Lesslauer, W. (1993). Crystal structure of the soluble human 55 kd TNF receptor–human TNF beta complex: implications for TNF receptor activation. Cell, 7, 431–445.Google Scholar
First citation Baumann, U. (1994). Crystal structure of the 50 kDa metallo protease from Serratia marcescens. J. Mol. Biol. 242, 244–251.Google Scholar
First citation Beaman, T. W., Binder, D. A., Blanchard, J. S. & Roderick, S. L. (1997). Three-dimensional structure of tetrahydrodipicolinate N-succinyltransferase. Biochemistry, 36, 489–494.Google Scholar
First citation Beaman, T. W., Sugantino, M. & Roderick, S. L. (1998). Structure of the hexapeptide xenobiotic acetyltransferase from Pseudomonas aeruginosa. Biochemistry, 37, 6689–6696.Google Scholar
First citation Becker, J. W., Marcy, A. I., Rokosz, L. L., Axel, M. G., Burbaum, J. J., Fitzgerald, P. M., Cameron, P. M., Esser, C. K., Hagmann, W. K., Hermes, J. D. & Springer, J. P. (1995). Stromelysin-1: three dimensional structure of the inhibited catalytic domain and of the C-truncated proenzyme. Protein Sci. 4, 1966–1976.Google Scholar
First citation Bentley, G., Dodson, E., Dodson, G., Hodgkin, D. & Mercola, D. (1976). Structure of insulin in 4-zinc insulin. Nature (London), 261, 166–168.Google Scholar
First citation Bernstein, B. E., Williams, D. M., Bressi, J. C., Kuhn, P., Gelb, M. H., Blackburn, G. M. & Hol, W. G. J. (1998). A bisubstrate analog induces unexpected conformational changes in phosphoglycerate kinase from Trypanosoma brucei. J. Mol. Biol. 279, 1137–1148.Google Scholar
First citation Bernstein, F. C., Koetzle, T. F., Williams, G. J., Meyer, E. F. Jr, Brice, M. D., Rodgers, J. R., Kennard, O., Shimanouchi, T. & Tasumi, M. (1977). The Protein Data Bank. A computer-based archival file for macromolecular structures. Eur. J. Biochem. 80, 319–324.Google Scholar
First citation Betz, M., Huxley, P., Davies, S. J., Mushtaq, Y., Pieper, M., Tschesche, H., Bode, W. & Gomis-Ruth, F. X. (1997). 1.8-Å crystal structure of the catalytic domain of human neutrophil collagenase (matrix metalloproteinase-8) complexed with a peptidomimetic hydroxamate primed-side inhibitor with a distinct selectivity profile. Eur. J. Biochem. 247, 356–363.Google Scholar
First citation Bienkowska, J., Cruz, M., Atiemo, A., Handin, R. & Liddington, R. (1997). The von Willebrand factor A3 domain does not contain a metal ion-dependent adhesion site motif. J. Biol. Chem. 272, 25162–25167.Google Scholar
First citation Birrer, P. (1995). Proteases and antiproteases in cystic fibrosis: pathogenetic considerations and therapeutic strategies. Respiration, 62, S25–S28.Google Scholar
First citation Bjorkman, P. J. & Burmeister, W. P. (1994). Structures of two classes of MHC molecules elucidated: crucial differences and similarities. Curr. Opin. Struct. Biol. 4, 852–856.Google Scholar
First citation Bjorkman, P. J., Saper, M. A., Samraoui, B., Bennett, W. S., Strominger, J. L. & Wiley, D. C. (1987). Structure of human class I histocompatibility antigen, HLA-A2. Nature (London), 329, 506–512.Google Scholar
First citation Blaber, M., DiSalvo, J. & Thomas, K. A. (1996). X-ray crystal structure of human acidic fibroblast growth factor. Biochemistry, 35, 2086–2094.Google Scholar
First citation Blake, C. C., Geisow, M. J., Oatley, S. J., Rerat, B. & Rerat, C. (1978). Structure of prealbumin: secondary, tertiary and quaternary interactions determined by Fourier refinement at 1.8 Å. J. Mol. Biol. 121, 339–356.Google Scholar
First citation Blankenfeldt, W., Nowicki, C., Montemartini-Kalisz, M., Kalisz, H. M. & Hecht, H. J. (1999). Crystal structure of Trypanosoma cruzi tyrosine aminotransferase: substrate specificity is influenced by cofactor binding mode. Protein Sci. 8, 2406–2417.Google Scholar
First citation Bochkarev, A., Barwell, J. A., Pfuetzner, R. A., Furey, W. Jr, Edwards, A. M. & Frappier, L. (1995). Crystal structure of the DNA-binding domain of the Epstein–Barr virus origin-binding protein EBNA 1. Cell, 83, 39–46.Google Scholar
First citation Bode, W., Mayr, I., Baumann, U., Huber, R., Stone, S. R. & Hofsteenge, J. (1989). The refined 1.9 Å crystal structure of human alpha-thrombin: interaction with D-Phe-Pro-Arg chloromethylketone and significance of the Tyr-Pro-Pro-Trp insertion segment. EMBO J. 8, 3467–3475.Google Scholar
First citation Bode, W., Reinemer, P., Huber, R., Kleine, T., Schnierer, S. & Tschesche, H. (1994). The X-ray crystal structure of the catalytic domain of human neutrophil collagenase inhibited by a substrate analogue reveals the essentials for catalysis and specificity. EMBO J. 13, 1263–1269.Google Scholar
First citation Bode, W., Wei, A. Z., Huber, R., Meyer, E., Travis, J. & Neumann, S. (1986). X-ray crystal structure of the complex of human leukocyte elastase (PMN elastase) and the third domain of the turkey ovomucoid inhibitor. EMBO J. 5, 2453–2458.Google Scholar
First citation Borkakoti, N., Winkler, F. K., Williams, D. H., D'Arcy, A., Broadhurst, M. J., Brown, P. A., Johnson, W. H. & Murray, E. J. (1994). Structure of the catalytic domain of human fibroblast collagenase complexed with an inhibitor. Nature Struct. Biol. 1, 106–110.Google Scholar
First citation Borst, P. (1999). Multidrug resistance: a solvable problem? Ann. Oncol. 10, S162–S164.Google Scholar
First citation Brandhuber, B. J., Boone, T., Kenney, W. C. & McKay, D. B. (1987). Three-dimensional structure of interleukin-2. Science, 238, 1707–1709.Google Scholar
First citation Brange, J. (1997). The new era of biotech insulin analogues. Diabetologia, 40, S48–S53.Google Scholar
First citation Breton, R., Housset, D., Mazza, C. & Fontecilla-Camps, J. C. (1996). The structure of a complex of human 17-beta-hydroxysteroid dehydrogenase with estradiol and NADP+ identifies two principal targets for the design of inhibitors. Structure, 4, 905–915.Google Scholar
First citation Brown, J. H., Jardetzky, T. S., Gorga, J. C., Stern, L. J., Urban, R. G., Strominger, J. L. & Wiley, D. C. (1993). Three-dimensional structure of the human class II histocompatibility antigen HLA-DR1. Nature (London), 364, 33–39.Google Scholar
First citation Browner, M. F., Smith, W. W. & Castelhano, A. L. (1995). Matrilysin-inhibitor complexes: common themes among metalloproteases. Biochemistry, 34, 6602–6610.Google Scholar
First citation Bruns, C. M., Hubatsch, I., Ridderstrom, M., Mannervik, B. & Tianer, J. A. (1999). Human glutathione transferase A4-4 crystal structures and mutagenesis reveal the basis of high catalytic efficiency with toxic lipid peroxidation products. J. Mol. Biol. 288, 427–439.Google Scholar
First citation Bruns, C. M., Nowalk, A. J., Arvai, A. S., McTigue, M. A., Vaughan, K. G., Mietzner, T. A. & McRee, D. E. (1997). Structure of Hemophilus influenzae Fe(+3)-binding protein reveals convergent evolution within a superfamily. Nature Struct. Biol. 4, 919–924.Google Scholar
First citation Brzozowski, A. M., Pike, A. C., Dauter, Z., Hubbard, R. E., Bonn, T., Engstrom, O., Ohman, L., Greene, G. L., Gustafsson, J. A. & Carlquist, M. (1997). Molecular basis of agonism and antagonism in the oestrogen receptor. Nature (London), 389, 753–758.Google Scholar
First citation Bullock, T. L., Roberts, T. M. & Stewart, M. (1996). 2.5 Å resolution crystal structure of the motile major sperm protein (MSP) of Ascaris suum. J. Mol. Biol. 263, 284–296.Google Scholar
First citation Burke, K. L., Dunn, G., Ferguson, M., Minor, P. D. & Almond, J. W. (1988). Antigen chimaeras of poliovirus as potential new vaccines. Nature (London), 332, 81–82.Google Scholar
First citation Bussiere, D. E., Pratt, S. D., Katz, L., Severin, J. M., Holzman, T. & Park, C. H. (1998). The structure of VanX reveals a novel amino-dipeptidase involved in mediating transposon-based vancomycin resistance. Mol. Cell, 2, 75–84.Google Scholar
First citation Cameron, A. D., Sinning, I., L'Hermite, G., Olin, B., Board, P. G., Mannervik, B. & Jones, T. A. (1995). Structural analysis of human alpha-class glutathione transferase A1-1 in the apo-form and in complexes with ethacrynic acid and its glutathione conjugate. Structure, 3, 717–727.Google Scholar
First citation Carfi, A., Pares, S., Duee, E., Galleni, M., Duez, C., Frere, J. M. & Dideberg, O. (1995). The 3-D structure of a zinc metallo-beta-lactamase from Bacillus cereus reveals a new type of protein fold. EMBO J. 14, 4914–4921.Google Scholar
First citation Carrell, R. W. & Gooptu, B. (1998). Conformational changes and diseases – serpins, prions and Alzheimer's. Curr. Opin. Struct. Biol. 8, 799–809.Google Scholar
First citation Carrell, R. W., Stein, P. E., Fermi, G. & Wardell, M. R. (1994). Biological implications of a 3 Å structure of dimeric antithrombin. Structure, 2, 257–270.Google Scholar
First citation Carter, D. C. & Ho, J. X. (1994). Structure of serum albumin. Adv. Protein Chem. 45, 153–203.Google Scholar
First citation Cate, J. H., Yusupov, M. M., Zusupova, G. Z., Earnest, T. N. & Noller, H. F. (1999). X-ray crystal structures of 70S ribosome functional complexes. Science, 285, 2095–2104.Google Scholar
First citation Champness, J. N., Achari, A., Ballantine, S. P., Bryant, P. K., Delves, C. J. & Stammers, D. K. (1994). The structure of Pneumocystis carinii dihydrofolate reductase to 1.9 Å resolution. Structure, 2, 915–924.Google Scholar
First citation Chan, D. C., Fass, D., Berger, J. M. & Kim, P. S. (1997). Core structure of gp41 from the HIV envelope glycoprotein. Cell, 89, 263–273.Google Scholar
First citation Chang, G., Spencer, R. H., Lee, A. T., Barclay, M. T. & Rees, D. C. (1998). Structure of the MscL homolog from Mycobacterium tuberculosis: a gated mechanosensitive ion channel. Science, 282, 2220–2226.Google Scholar
First citation Chang, Y., Mochalkin, I., McCance, S. G., Cheng, B., Tulinsky, A. & Castellino, F. J. (1998). Structure and ligand binding determinants of the recombinant kringle 5 domain of human plasminogen. Biochemistry, 37, 3258–3271.Google Scholar
First citation Charifson, P. S. (1997). Practical application of computer-aided drug design. New York: Marcel Dekker Inc.Google Scholar
First citation Chitarra, V., Holm, I., Bentley, G. A., Petres, S. & Longacre, S. (1999). The crystal structure of C-terminal merozoite surface protein 1 at 1.8 Å resolution, a highly protective malaria vaccine candidate. Mol. Cell, 3, 457–464.Google Scholar
First citation Cho, Y., Gorina, S., Jeffrey, P. D. & Pavletich, N. P. (1994). Crystal structure of a p53 tumor suppressor–DNA complex: understanding tumorigenic mutations. Science, 265, 346–355.Google Scholar
First citation Choe, S., Bennett, M. J., Fujii, G., Curmi, P. M., Kantardjieff, K. A., Collier, R. J. & Eisenberg, D. (1992). The crystal structure of diphtheria toxin. Nature, 357, 216–222.Google Scholar
First citation Choi, H. J., Kang, S. W., Yang, C. H., Rhee, S. G. & Ryu, S. E. (1998). Crystal structure of a novel human peroxidase enzyme at 2.0 Å resolution. Nature Struct. Biol. 5, 400–406.Google Scholar
First citation Choudhury, D., Thompson, A., Stojanoff, V., Langermann, S., Pinkner, J., Hultgren, S. J. & Knight, S. D. (1999). X-ray structure of the Fim-C-FimH chaperone–adhesin complex from uropathogenic Escherichia coli. Science, 285, 1061–1066.Google Scholar
First citation Chudzik, D. M., Michels, P. A., de Walque, S. & Hol, W. G. J. (2000). Structures of type 2 peroxisomal targeting signals in two trypanosomatid aldolases. J. Mol. Biol. 300, 697–707.Google Scholar
First citation Cirilli, M., Zheng, R., Scapin, G. & Blanchard, J. S. (1993). Structural symmetry: the three-dimensional structure of Haemophilus influenzae diaminopimelate epimerase. Biochemistry, 37, 16452–16458.Google Scholar
First citation Ciszak, E. & Smith, G. D. (1994). Crystallographic evidence for dual coordination around zinc in the T3R3 human insulin hexamer. Biochemistry, 33, 1512–1517.Google Scholar
First citation Clackson, T., Yang, W., Rozamus, L. W., Hatada, M., Amara, J. F., Rollins, C. T., Stevenson, L. F., Magari, S. R., Wood, S. A., Courage, N. L., Lu, X., Cerasoli, F. J., Gilman, M. & Holt, D. A. (1998). Redesigning an FKBP–ligand interface to generate chemical dimerizers with novel specificity. Proc. Natl Acad. Sci. USA, 95, 10437–10442.Google Scholar
First citation Cleasby, A., Wonacott, A., Skarzynski, T., Hubbard, R. E., Davies, G. J., Proudfoot, A. E., Bernard, A. R., Payton, M. A. & Wells, T. N. (1996). The X-ray crystal structure of phosphomannose isomerase from Candida albicans at 1.7 Å resolution. Nature Struct. Biol. 3, 470–479.Google Scholar
First citation Clemons, W. M. J., May, J. L., Wimberly, B. T., McCutcheon, J. P., Capel, M. S. & Ramakrishnan, V. (1999). Structure of a bacterial 30S ribosomal subunit at 5.5 Å resolution. Nature (London), 400, 833–840.Google Scholar
First citation Cobessi, D., Tete-Favier, F., Marchal, S., Azza, S., Branlant, G. & Aubry, A. (1999). Apo and holo crystal structures of an NADP-dependent aldehyde dehydrogenase from Streptococcus mutans. J. Mol. Biol. 290, 161–173.Google Scholar
First citation Colby, T. D., Vanderveen, K., Strickler, M. D., Markham, G. D. & Goldstein, B. M. (1999). Crystal structure of human type II inosine monophosphate dehydrogenase: implications for ligand binding and drug design. Proc. Natl Acad. Sci. USA, 96, 3531–3536.Google Scholar
First citation Cole, S. T., Brosch, R., Parkhill, J., Garnier, T., Churcher, C., Harris, D., Gordon, S. V., Eiglmeier, K., Gas, S., Barry, C. E. III, Tekaia, F., Badcock, K., Basham, D., Brown, D., Chillingworth, T., Connor, R., Davies, R., Devlin, K., Feltwell, T., Gentles, S., Hamlin, N., Holroyd, S., Hornby, T., Jagels, K., Krogh, A., McLean, J., Moule, S., Murphy, L., Oliver, K., Osborne, J., Quail, M. A., Rajandream, M. A., Rogers, J., Rutter, S., Seeger, K., Skelton, J., Squares, R., Squares, S., Sulston, J. E., Taylor, K., Whitehead, S. & Barrell, B. G. (1998). Deciphering the biology of Mycobacterium tuberculosis from the complete genome sequence. Nature (London), 393, 537–544.Google Scholar
First citation Collins, F. S. (1992). Cystic fibrosis: molecular biology and therapeutic implications. Science, 256, 774–779.Google Scholar
First citation Concha, N. O., Rasmussen, B. A., Bush, K. & Herzberg, O. (1996). Crystal structure of the wide-spectrum binuclear zinc beta-lactamase from Bacteroides fragilis. Structure, 4, 823–836.Google Scholar
First citation Conlon, H. D., Baqai, J., Baker, K., Shen, Y. Q., Wong, B. L., Noiles, R. & Rausch, C. W. (1995). 2-step immobilized enzyme conversion of cephalosporin-c to 7-aminocephalosporanic acid. Biotechnol. Bioeng. 46, 510–513.Google Scholar
First citation Cooper, J. B., McIntyre, K., Badasso, M. O., Wood, S. P., Zhang, Y., Garbe, T. R. & Young, D. (1995). X-ray structure analysis of the iron-dependent superoxide dismutase from Mycobacterium tuberculosis at 2.0 Å resolution reveals novel dimer–dimer interactions. J. Mol. Biol. 246, 531–544.Google Scholar
First citation Correll, C. C., Batie, C. J., Ballou, D. P. & Ludwig, M. L. (1992). Phthalate dioxygenase reductase: a modular structure for electron transfer from pyridine nucleotides to [2Fe–2S]. Science, 258, 1604–1610.Google Scholar
First citation Crennell, S., Garman, E., Laver, G., Vimr, E. & Taylor, G. (1994). Crystal structure of Vibrio cholerae neuraminidase reveals dual lectin-like domains in addition to the catalytic domain. Structure, 2, 535–544.Google Scholar
First citation Curry, S., Mandelkow, H., Brick, P. & Franks, N. (1998). Crystal structure of human serum albumin complexed with fatty acid reveals an asymmetric distribution of binding sites. Nature Struct. Biol. 5, 827–835.Google Scholar
First citation Cushman, D. W. & Ondetti, M. A. (1991). History of the design of captopril and related inhibitors of angiotensin converting enzyme. Hypertension, 17, 589–592.Google Scholar
First citation Cutfield, S. M., Dodson, E. J., Anderson, B. F., Moody, P. C., Marshall, C. J., Sullivan, P. A. & Cutfield, J. F. (1995). The crystal structure of a major secreted aspartic proteinase from Candida albicans in complexes with two inhibitors. Structure, 3, 1261–1271.Google Scholar
First citation Das, K., Ding, J., Hsiou, Y., Clark, A. D. Jr, Moereels, H., Koymans, L., Andries, K., Pauwels, R., Janssen, P. A. J., Boyer, P. L., Clark, P., Smith, R. H. Jr, Kroeger Smith, M. B., Michejda, C. J., Hughes, S. H. & Arnold, E. (1996). Crystal structures of 8-Cl and 9-Cl TIBO complexed with wild-type HIV-1 RT and 8-Cl TIBO complexed with the Tyr181Cys HIV-1 RT drug-resistant mutant. J. Mol. Biol. 264, 1085–1100.Google Scholar
First citation Davies, J. F., Delcamp, T. J., Prendergast, N. J., Ashford, V. A., Freisheim, J. H. & Kraut, J. (1990). Crystal structures of recombinant human dihydrofolate reductase complexed with folate and 5-deazafolate. Biochemistry, 29, 9467–9479.Google Scholar
First citation De Bondt, H. L., Rosenblatt, J., Jancarik, J., Jones, H. D., Morgan, D. O. & Kim, S. H. (1993). Crystal structure of cyclin-dependent kinase 2. Nature (London), 363, 595–602.Google Scholar
First citation Deller, M. C. & Jones, E. Y. (2000). Cell surface receptors. Curr. Opin. Struct. Biol. 10, 213–219.Google Scholar
First citation Derewenda, U., Derewenda, Z., Dodson, E. J., Dodson, G. G., Reynolds, C. D., Smith, G. D., Sparks, C. & Swenson, D. (1989). Phenol stabilizes more helix in a new symmetrical zinc insulin hexamer. Nature (London), 338, 594–596.Google Scholar
First citation Dessen, A., Quemard, A., Blanchard, J. S., Jacobs, W. R. J. & Sacchettini, J. C. (1995). Crystal structure and function of the isoniazid target of Mycobacterium tuberculosis. Science, 267, 1638–1641.Google Scholar
First citation DeVos, A. M., Tong, L., Milburn, M. V., Matias, P. M., Jancarik, J., Noguchi, S., Nishimura, S., Miura, K., Ohtsuka, E. & Kim, S. H. (1988). Three-dimensional structure of an oncogene protein: catalytic domain of human c-H-ras p21. Science, 239, 888–893.Google Scholar
First citation DeVos, A. M., Ultsch, M. & Kossiakoff, A. A. (1992). Human growth hormone and extracellular domain of its receptor: crystal structure of the complex. Science, 255, 306–312.Google Scholar
First citation Dhanaraj, V., Ye, Q.-Z., Johnson, L. L., Hupe, D. J., Otwine, D. F., Dunbar, J. B. J., Rubin, J. R., Pavlovsky, A., Humblet, C. & Blundell, T. L. (1996). X-ray structure of a hydroxamate inhibitor complex of stromelysin catalytic domain and its comparison with members of the zinc metalloproteinase superfamily. Structure, 4, 375–386.Google Scholar
First citation Dickerson, R. E. & Geis, I. (1983). Hemoglobin. Menlo Park: Benjamin Cummings Publishing Co.Google Scholar
First citation Ding, J., Das, K., Tantillo, C., Zhang, W., Clark, A. D. Jr, Jessen, S., Lu, X., Hsiou, Y., Jacobo-Molina, A., Andries, K., Pauwels, R., Moereels, H., Koymans, L., Janssen, P. A. J., Smith, R. H. Jr, Kroeger Koepke, M., Michejda, C. J., Hughes, S. H. & Arnold, E. (1995). Structure of HIV-1 reverse transcriptase in a complex with the non-nucleoside inhibitor alpha-APA R 95845 at 2.8 Å resolution. Structure, 3, 365–379.Google Scholar
First citation Ding, J., McGrath, W. J., Sweet, R. M. & Mangel, W. F. (1996). Crystal structure of the human adenovirus proteinase with its 11 amino acid cofactor. EMBO J. 15, 1778–1783.Google Scholar
First citation Duggleby, H. J., Tolley, S. P., Hill, C. P., Dodson, E. J., Dodson, G. & Moody, P. C. (1995). Penicillin acylase has a single-amino-acid catalytic centre. Nature (London), 373, 264–268.Google Scholar
First citation Dyda, F., Hickman, A. B., Jenkins, T. M., Engelman, A., Craigie, R. & Davies, D. R. (1994). Crystal structure of the catalytic domain of HIV-1 integrase: similarity to other polynucleotidyl transferases. Science, 266, 1981–1986.Google Scholar
First citation Eads, J. C., Scapin, G., Xu, Y., Grubmeyer, C. & Sacchettini, J. C. (1994). The crystal structure of human hypoxanthine-guanine phosphoribosyltransferase with bound GMP. Cell, 78, 325–334.Google Scholar
First citation Ealick, S. E., Cook, W. J., Vijay-Kumar, S., Carson, M., Nagabhushan, T. L., Trotta, P. P. & Bugg, C. E. (1991). Three-dimensional structure of recombinant human interferon-gamma. Science, 252, 698–702.Google Scholar
First citation Ealick, S. E., Rule, S. A., Carter, D. C., Greenhough, T. J., Babu, Y. S., Cook, W. J., Habash, J., Helliwell, J. R., Stoeckler, J. D., Parks, R. E. J., Chen, S. F. & Bugg, C. E. (1990). Three-dimensional structure of human erythrocytic purine nucleoside phosphorylase at 3.2-Å resolution. J. Biol. Chem. 265, 1812–1820.Google Scholar
First citation Eckert, D. M., Malashkevish, V. N., Hong, L. H., Carr, P. A. & Kim, P. S. (1999). Inhibiting HIV-1 entry: discovery of D-peptide inhibitors that target the gp41 coiled-coil pocket. Cell, 99, 103–115.Google Scholar
First citation Ekstrom, J. L., Mathews, I. I., Stanley, B. A., Pegg, A. E. & Ealick, S. E. (1999). The crystal structure of human S-adenosylmethionine decarboxylase at 2.25 Å resolution reveals a novel fold. Struct. Fold. Des. 7, 583–595.Google Scholar
First citation Emsley, J., Cruz, M., Handin, R. & Liddington, R. (1998). Crystal structure of the von Willebrand factor A1 domain and implications for the binding of platelet glycoprotein Ib. J. Biol. Chem. 273, 10396–10401.Google Scholar
First citation Emsley, P., Charles, I. G., Fairweather, N. F. & Isaacs, N. W. (1996). Structure of Bordetella pertussis virulence factor P.69 pertactin. Nature (London), 381, 90–92.Google Scholar
First citation Erickson, J., Neidhart, D. J., VanDrie, J., Kempf, D. J., Wang, X. C., Norbeck, D. W., Plattner, J. J., Rittenhouse, J. W., Turon, M., Wideburg, N., Kohlbrenner, W. E., Simmer, R., Helfrich, R., Paul, D. A. & Knigge, M. (1990). Design, activity, and 2.8 Å crystal structure of a C2 symmetric inhibitor complexed to HIV-1 protease. Science, 249, 527–533.Google Scholar
First citation Erickson, J. W. & Burt, S. K. (1996). Structural mechanisms of HIV drug resistance. Annu. Rev. Pharmacol. Toxicol. 36, 545–571.Google Scholar
First citation Erlandsen, H., Fusetti, F., Martinez, A., Hough, E., Flatmark, T. & Stevens, R. C. (1997). Crystal structure of the catalytic domain of human phenylalanine hydroxylase reveals the structural basis for phenylketonuria. Nature Struct. Biol. 4, 995–1000.Google Scholar
First citation Esser, C. K., Bugianesi, R. L., Caldwell, C. G., Chapman, K. T., Durette, P. L., Girotra, N. N., Kopka, I. E., Lanza, T. J., Levorse, D. A., Maccoss, M., Owens, K. A., Ponpipom, M. M., Simeone, J. P., Harrison, R. K., Niedzwiecki, L., Becker, J. W., Marcy, A. I., Axel, M. G., Christen, A. J., McDonnell, J., Moore, V. L., Olsqewski, J. M., Saphos, C., Visco, D. M., Shen, F., Colletti, A., Kriter, P. A. & Hagmann, W. K. (1997). Inhibition of stromelysin-1 (MMP-3) by P1′-biphenylylethyl carboxyalkyl dipeptides. J. Med. Chem. 40, 1026–1040.Google Scholar
First citation Fan, C., Moews, P. C., Walsh, C. T. & Knox, J. R. (1994). Vancomycin resistance: structure of D-alanine:D-alanine ligase at 2.3 Å resolution. Science, 266, 439–443.Google Scholar
First citation Fan, E., Zhang, Z., Minke, W. E., Hou, Z., Verlinde, C. L. M. J. & Hol, W. G. J. (2000). A 105 gain in affinity for pentavalent ligands of E. coli heat-labile enterotoxin by modular structure-based design. J. Am. Chem. Soc. 122, 2663–2664.Google Scholar
First citation Ferrer, M., Kapoor, T. M., Strassmaier, T., Weissenhorn, W., Skehel, J. J., Oprian, D., Schreiber, S. L., Wiley, D. C. & Harrison, S. C. (1999). Selection of gp41-mediated HIV-1 cell entry inhibitors from biased combinatorial libraries of non-natural binding elements. Nature Struct. Biol. 6, 953–960.Google Scholar
First citation Fields, B. A., Malchiodi, E. L., Li, H., Ysern, X., Stauffacher, C. V., Schlievert, P. M., Karjalainen, K. & Mariuzza, R. A. (1996). Crystal structure of a T-cell receptor beta-chain complexed with a superantigen. Nature (London), 384, 188–192.Google Scholar
First citation Filman, D. J., Wien, M. W., Cunningham, J. A., Bergelson, J. M. & Hogle, J. M. (1998). Structure determination of echovirus 1. Acta Cryst. D54, 1261–1272.Google Scholar
First citation Finzel, B. C., Baldwin, E. T., Bryant, G. L. J., Hess, G. F., Wilks, J. W., Trepod, C. M., Mott, J. E., Marshall, V. P., Petzold, G. L., Poorman, R. A., O'Sullivan, T. J., Schostarez, H. J. & Mitchell, M. A. (1998). Structural characterizations of nonpeptidic thiadiazole inhibitors of matrix metalloproteinases reveal the basis for stromelysin selectivity. Protein Sci. 7, 2118–2126.Google Scholar
First citation Focia, P. J., Craig, S. P. III, Nieves-Alicea, R., Fletterick, R. J. & Eakin, A. E. (1998). A 1.4 Å crystal structure for the hypoxanthine phosphoribosyltransferase of Trypanosoma cruzi. Biochemistry, 37, 15066–15075.Google Scholar
First citation Foster, B. A., Coffey, H. A., Morin, M. J. & Rastinejad, F. (1999). Pharmacological rescue of mutant p53 conformation and function. Science, 286, 2507–2510.Google Scholar
First citation Fox, M. P., Otto, M. J. & McKinlay, M. A. (1986). Prevention of rhinovirus and poliovirus uncoating by WIN 51711, a new antiviral drug. Antimicrob. Agents Chemother. 30, 110–116. Google Scholar
First citation Frankenberg, N., Erskine, P. T., Cooper, J. B., Shoolingin-Jordan, P. M., Jahn, D. & Heinz, D. W. (1999). High resolution crystal structure of a Mg2+-dependent porphobilinogen synthase. J. Mol. Biol. 289, 591–602.Google Scholar
First citation Fremont, D. H., Matsumura, M., Stura, E. A., Peterson, P. A. & Wilson, I. A. (1992). Crystal structures of two viral peptides in complex with murine MHC class I H-2Kb. Science, 257, 919–927.Google Scholar
First citation Freymann, D., Down, J., Carrington, M., Roditi, I., Turner, M. & Wiley, D. (1990). 2.9 Å resolution structure of the N-terminal domain of a variant surface glycoprotein from Trypanosoma brucei. J. Mol. Biol. 216, 141–160.Google Scholar
First citation Fulop, V., Ridout, C. J., Greenwood, C. & Hajdu, J. (1995). Crystal structure of the di-haem cytochrome c peroxidase from Pseudomonas aeruginosa. Structure, 3, 1225–1233.Google Scholar
First citation Futterer, K., Wong, J., Grucza, R. A., Chan, A. C. & Waksman, G. (1998). Structural basis for Syk tyrosine kinase ubiquity in signal transduction pathways revealed by the crystal structure of its regulatory SH2 domains bound to a dually phosphorylated ITAM peptide. J. Mol. Biol. 281, 523–537.Google Scholar
First citation Gaboriaud, C., Serre, L., Guy-Crotte, O., Forest, E. & Fontecilla-Camps, J. C. (1996). Crystal structure of human trypsin 1: unexpected phosphorylation of Tyr151. J. Mol. Biol. 259, 995–1010.Google Scholar
First citation Gamblin, S. J., Cooper, B., Millar, J. R., Davies, G. J., Littlechild, J. A. & Watson, H. C. (1990). The crystal structure of human muscle aldolase at 3.0 Å resolution. FEBS Lett. 264, 282–286.Google Scholar
First citation Garboczi, D. N., Ghosh, P., Utz, U., Fan, Q. R., Biddison, W. E. & Wiley, D. C. (1996). Structure of the complex between human T-cell receptor, viral peptide and HLA-A2. Nature (London), 384, 134–141.Google Scholar
First citation Garcia, K. C., Degano, M., Stanfield, R. L., Brunmark, A., Jackson, M. R., Petereson, P. A., Teyton, L. & Wilson, I. A. (1996). An αβ T cell receptor structure at 2.5 Å and its orientation in the TCR–MHC complex. Science, 274, 209–219.Google Scholar
First citation Gardner, M. J., Tettelin, H., Carucci, D. J., Cummings, L. M., Aravind, L., Koonin, E. V., Shallom, S., Mason, T., Yu, K., Fujii, C., Pederson, J., Shen, K., Jing, J., Aston, C., Lai, Z., Schwartz, D. C., Pertea, M., Salzburg, S., Zhou, L., Sutton, G. G., Clayton, R., White, O., Smith, H. O., Fraser, C. M., Adams, M. D., Venter, J. C. & Hoffman, S. L. (1998). Chromosome 2 sequence of the human malaria parasite Plasmodium falciparum. Science, 282, 1126–1132.Google Scholar
First citation Gatti, D. L., Palfey, B. A., Lah, M. S., Entsch, B., Massey, V., Ballou, D. P. & Ludwig, M. L. (1994). The mobile flavin of 4-OH benzoate hydroxylase. Science, 266, 110–114.Google Scholar
First citation Ghosh, D., Pletnev, V. Z., Zhy, D. W., Wawrzak, Z., Duax, W. L., Pangborn, W., Labrie, F. & Lin, S. W. (1995). Structure of human estrogenic 17-beta-hydroxysteroid dehydrogenase at 2.20-Å resolution. Structure, 3, 503–513.Google Scholar
First citation Giulian, D., Corpuz, M., Richmond, B., Wendt, E. & Hall, E. R. (1996). Activated microglia are the principal glial source of thromboxane in the central nervous system. Neurochem. Int. 29, 65–76.Google Scholar
First citation Gohlke, U., Gomis-Ruth, F. X., Crabbe, T., Murphy, G., Docherty, A. J. & Bode, W. (1996). The C-terminal (haemopexin-like) domain structure of human gelatinase A (MMP2): structural implications for its function. FEBS Lett. 378, 126–120.Google Scholar
First citation Gomis-Ruth, F. X., Gohlke, U., Betz, M., Knauper, V., Murphy, G., Lopez-Otin, C. & Bode, W. (1996). The helping hand of collagenase-3 (MMP-13): 2.7 Å crystal structure of its C-terminal haemopexin-like domain. J. Mol. Biol. 264, 556–566.Google Scholar
First citation Gomis-Ruth, F. X., Maskos, K., Betz, M., Bergner, A., Huber, R., Suzuki, K., Yoshida, N., Nagase, H., Brew, K., Bourenkov, G. P., Bartunik, H. & Bode, W. (1997). Mechanism of inhibition of the human matrix metalloproteinase stromelysin-1 by TIMP-1. Nature (London), 389, 77–81.Google Scholar
First citation Gong, W., O'Gara, M., Blumenthal, R. M. & Cheng, X. (1997). Structure of pvu II DNA-(cytosine N4) methyltransferase, an example of domain permutation and protein fold assignment. Nucleic Acids Res. 25, 2702–2715.Google Scholar
First citation Goodwill, K. E., Sabatier, C., Marks, C., Raag, R., Fitzpatrick, P. F. & Stevens, R. C. (1997). Crystal structure of tyrosine hydroxylase at 2.3 Å and its implications for inherited neurodegenerative diseases. Nature Struct. Biol. 4, 578–585.Google Scholar
First citation Gordon, D. B., Marshall, S. A. & Mayo, S. L. (1999). Energy functions for protein design. Curr. Opin. Struct. Biol. 9, 509–513.Google Scholar
First citation Gorina, S. & Pavletich, N. P. (1996). Structure of the p53 tumor suppressor bound to the ankyrin and SH3 domains of 53BP2. Science, 274, 1001–1005.Google Scholar
First citation Gouet, P., Jouve, H. M. & Dideberg, O. (1995). Crystal structure of Proteus mirabilis PT catalase with and without bound NADPH. J. Mol. Biol. 249, 933–954.Google Scholar
First citation Gourley, D. G., Shrive, A. K., Polikarpov, I., Krell, T., Coggins, J. R., Hawkins, A. R., Isaacs, N. W. & Sawyer, L. (1999). The two types of 3-dehydrogenase have distinct structures but catalyze the same overall reaction. Nature Struct. Biol. 6, 521–525.Google Scholar
First citation Grant, R. A., Hiemath, C. N., Filman, D. J., Syed, R., Andries, K. & Hogle, J. M. (1994). Structures of poliovirus complexes with anti-viral drugs: implications for viral stability and drug design. Curr. Biol. 4, 784–797.Google Scholar
First citation Graves, B. J., Hatada, M. H., Hendrickson, W. A., Miller, J. K., Madison, V. S. & Satow, Y. (1990). Structure of interleukin 1 alpha at 2.7-Å resolution. Biochemistry, 29, 2679–2684.Google Scholar
First citation Grimes, J., Basak, A. K., Roy, P. & Stuart, D. (1995). The crystal structure of bluetongue virus VP7. Nature (London), 373, 167–170.Google Scholar
First citation Hampele, I. C., D'Arcy, A., Dale, G. E., Kostrewa, D., Nielsen, J., Oefner, C., Page, M. G., Schonfeld, H. J., Stuber, D. & Then, R. L. (1997). Structure and function of the dihydropteroate synthase from Staphylococcus aureus. J. Mol. Biol. 268, 21–30.Google Scholar
First citation Han, S., Eltis, L. D., Timmis, K. N., Muchmore, S. W. & Bolin, J. T. (1995). Crystal structure of the biphenyl-cleaving extradiol dioxygenase from a PCB-degrading pseudomonad. Science, 270, 976–980.Google Scholar
First citation Hansen, J. L., Long, A. M. & Schultz, S. C. (1997). Structure of the RNA-dependent RNA polymerase of poliovirus. Structure, 5, 1109–1122.Google Scholar
First citation Harrington, D. J., Adachi, K. & Royer, W. E. Jr (1997). The high resolution crystal structure of deoxyhemoglobin S. J. Mol. Biol. 272, 398–407.Google Scholar
First citation Harris, S. F. & Botchan, M. R. (1999). Crystal structure of the human papillomavirus type 18 E2 activation domain. Science, 284, 1673–1677.Google Scholar
First citation He, X. M. & Carter, D. C. (1992). Atomic structure and chemistry of human serum albumin. Nature (London), 358, 209–215.Google Scholar
First citation Hegde, R. S. & Androphy, E. J. (1998). Crystal structure of the E2 DNA-binding domain from human papillomavirus type 16: implications for its DNA binding-site selection mechanism. J. Mol. Biol. 284, 1479–1489.Google Scholar
First citation Hennig, M., Dale, G. E., D'Arcy, A., Danel, F., Fischer, S., Gray, C. P., Jolidon, S., Muller, F., Page, M. G., Pattison, P. & Oefner, C. (1999). The structure and function of the 6-hydroxymethyl-7,8-dihydropterin pyrophosphokinase from Haemophilus influenzae. J. Mol. Biol. 287, 211–219.Google Scholar
First citation Hennig, M., D'Arcy, A., Hampele, I. C., Page, M. G., Oefner, C. & Dale, G. E. (1998). Crystal structure and reaction mechanism of 7,8-dihydroneopterin aldolase from Staphylococcus aureus. Nature Struct. Biol. 5, 357–362.Google Scholar
First citation Herzberg, O. & Moult, J. (1987). Bacterial resistance to beta-lactam antibiotics: crystal structure of beta-lactamase from Staphylococcus aureus PC1 at 2.5 Å resolution. Science, 236, 694–701.Google Scholar
First citation Hill, C. P., Worthylake, D., Bancroft, D. P., Christensen, A. M. & Sundquist, W. I. (1996). Crystal structures of the trimeric human immunodeficiency virus type 1 matrix protein: implications for membrane association and assembly. Proc. Natl Acad. Sci. USA, 93, 3099–3104.Google Scholar
First citation Hodel, A. E., Gershon, P. D., Shi, X. & Quiocho, F. A. (1996). The 1.85 Å structure of vaccinia protein VP39: a bifunctional enzyme that participates in the modification of both mRNA ends. Cell, 85, 247–256.Google Scholar
First citation Hodgkin, D. C. (1971). Insulin molecules: the extent of our knowledge. Pure Appl. Chem. 26, 375–384.Google Scholar
First citation Hofmann, B., Schomburg, D. & Hecht, H. J. (1993). Crystal structure of a thiol proteinase from Staphylococcus aureus V-8 in the E-64 inhibitor complex. Acta Cryst. A49, C-102.Google Scholar
First citation Hogle, J. M., Chow, M. & Filman, D. J. (1985). Three-dimensional structure of poliovirus at 2.9 Å resolution. Science, 229, 1358–1365.Google Scholar
First citation Hol, W. G. J. (1986). Protein crystallography and computer graphics – toward rational drug design. Angew. Chem. Int. Ed. Engl. 25, 767–778.Google Scholar
First citation Hoog, S. S., Smith, W. W., Qiu, X., Janson, C. A., Hellmig, B., McQueney, M. S., O'Donnell, K., O'Shannessy, D., DiLella, A. G., Debouck, C. & Abdel-Meguid, S. S. (1997). Active site cavity of herpesvirus proteases revealed by the crystal structure of herpes simplex virus protease/inhibitor complex. Biochemistry, 36, 14023–14029.Google Scholar
First citation Hough, E., Hansen, L. K., Birknes, B., Jynge, K., Hansen, S., Hordvik, A., Little, C., Dodson, E. & Derewenda, Z. (1989). High-resolution (1.5 Å) crystal structure of phospholipase C from Bacillus cereus. Nature (London), 338, 357–360.Google Scholar
First citation Hsiou, Y., Das, K., Ding, J., Clark, A. D. Jr, Kleim, J. P., Rosner, M., Winkler, I., Riess, G., Hughes, S. H. & Arnold, E. (1998). Structures of Tyr188Leu mutant and wild-type HIV-1 reverse transcriptase complexed with the non-nucleoside inhibitor HBY 097: inhibitor flexibility is a useful design feature for reducing drug resistance. J. Mol. Biol. 284, 313–323.Google Scholar
First citation Hu, S. H., Peek, J. A., Rattigan, E., Taylor, R. K. & Martin, J. L. (1997). Structure of TcpG, the DsbA protein folding catalyst from Vibrio cholerae. J. Mol. Biol. 268, 137–146.Google Scholar
First citation Huang, H., Chopra, R., Verdine, G. L. & Harrison, S. C. (1998). Structure of a covalently trapped catalytic complex of HIV-1 reverse transcriptase: implications for drug resistance. Science, 282, 1669–1675.Google Scholar
First citation Huang, K., Strynadka, N. C., Bernard, V. D., Peanasky, R. J. & James, M. N. (1994). The molecular structure of the complex of Ascaris chymotrypsin/elastase inhibitor with porcine elastase. Structure, 2, 679–689.Google Scholar
First citation Huang, S., Xue, Y., Sauer-Eriksson, E., Chirica, L., Lindskog, S. & Jonsson, B. H. (1998). Crystal structure of carbonic anhydrase from Neisseria gonorrhoeae and its complex with the inhibitor acetazolamide. J. Mol. Biol. 283, 301–310.Google Scholar
First citation Hubbard, S. R. (1997). Crystal structure of the activated insulin receptor tyrosine kinase in complex with peptide substrate and ATP analog. EMBO J. 16, 5572–5581.Google Scholar
First citation Hubbard, S. R., Wei, L., Ellis, L. & Hendrickson, W. A. (1994). Crystal structure of the tyrosine kinase domain of the human insulin receptor. Nature (London), 372, 746–754.Google Scholar
First citation Huizinga, E. G., Martijn van der Plas, R., Kroon, J., Sixma, J. J. & Gros, P. (1997). Crystal structure of the A3 domain of human von Willebrand factor: implications for collagen binding. Structure, 5, 1147–1156.Google Scholar
First citation Hulsmeyer, M., Hecht, H. J., Niefind, K., Hofer, B., Eltis, L. D., Timmis, K. N. & Schomburg, D. (1998). Crystal structure of cis-biphenyl-2,3-dihydrodiol-2,3-dehydrogenase from a PCB degrader at 2.0 Å resolution. Protein Sci. 7, 1286–1293.Google Scholar
First citation Hurley, T. D., Bosron, W. F., Hamilton, J. A. & Amzel, L. M. (1991). Structure of human beta 1 beta 1 alcohol dehydrogenase: catalytic effects of non-active-site substitutions. Proc. Natl Acad. Sci. USA, 88, 8149–8153.Google Scholar
First citation Isupov, M. N., Antson, A. A., Dodson, E. J., Dodson, G. G., Dementieva, I. S., Zakomirdina, L. N., Wilson, K. S., Dauter, Z., Lebedev, A. A. & Harutyunyan, E. H. (1998). Crystal structure of tryptophanase. J. Mol. Biol. 276, 603–623.Google Scholar
First citation Itzstein, M. von, Wu, W. Y., Kok, G. B., Pegg, M. S., Dyason, J. C., Jin, B., Van Phan, T., Smythe, M. L., White, H. F., Oliver, S. W., Colman, P. M., Varghese, J. N., Ryan, D. M., Woods, J. M., Bethell, R. C., Hotham, V. J., Cameron, J. M. & Penn, C. R. (1993). Rational design of potent sialidase-based inhibitors of influenza virus replication. Nature (London), 363, 418–423.Google Scholar
First citation Jackson, R. C. (1997). Contributions of protein structure-based drug design to cancer chemotherapy. Semin. Oncol. 24, 164–172.Google Scholar
First citation Jacobo-Molina, A., Ding, J., Nanni, R. G., Clark, A. D. Jr, Lu, X., Tantillo, C., Williams, R. L., Kamer, G., Ferris, A. L., Clark, P., Hizi, A., Hughes, S. H. & Arnold, E. (1993). Crystal structure of human immunodeficiency virus type 1 reverse transcriptase complexed with double-stranded DNA at 3.0 Å resolution shows bent DNA. Proc. Natl Acad. Sci. USA, 90, 6320–6324.Google Scholar
First citation Jain, S., Drendel, W. B., Chen, Z. W., Mathews, F. S., Sly, W. S. & Grubb, J. H. (1996). Structure of human beta-glucuronidase reveals candidate lysosomal targeting and active-site motifs. Nature Struct. Biol. 3, 375–381.Google Scholar
First citation Jia, Z., Vandonselaar, M., Quail, J. W. & Delbaere, L. T. (1993). Active-centre torsion-angle strain revealed in 1.6 Å-resolution structure of histidine-containing phosphocarrier protein. Nature (London), 361, 94–97.Google Scholar
First citation Kallarakal, A. T., Mitra, B., Kozarich, J. W., Gerlt, J. A., Clifton, J. G., Petsko, G. A. & Kenyon, G. L. (1995). Mechanism of the reaction catalyzed by mandelate racemase: structure and mechanistic properties of the K166R mutant. Biochemistry, 34, 2788–2797.Google Scholar
First citation Kallen, J., Spitzfaden, C., Zurini, M. G. M., Wider, G., Widmer, H., Wuethrich, K. & Walkinshaw, M. D. (1991). Structure of human cyclophilin and its binding site for cyclosporin A determined by X-ray crystallography and NMR spectroscopy. Nature (London), 353, 276–279.Google Scholar
First citation Kannan, K. K., Notstrand, B., Fridborg, K., Lovgren, S., Ohlsson, A. & Petef, M. (1975). Crystal structure of human erythrocyte carbonic anhydrase B. Three-dimensional structure at a nominal 2.2-Å resolution. Proc. Natl Acad. Sci. USA, 72, 51–55.Google Scholar
First citation Karpusas, M., Nolte, M., Benton, C. B., Meier, W., Lipscomb, W. N. & Goelz, S. (1997). The crystal structure of human interferon beta at 2.2-Å resolution. Proc. Natl Acad. Sci. USA, 94, 11813–11818.Google Scholar
First citation Ke, H., Zydowsky, L. D., Liu, J. & Walsh, C. T. (1991). Crystal structure of recombinant human T-cell cyclophilin A at 2.5-Å resolution. Proc. Natl Acad. Sci. USA, 88, 9483–9487.Google Scholar
First citation Kim, H., Certa, U., Döbelli, H., Jakob, P. & Hol, W. G. J. (1998). Crystal structure of fructose-1,6-bisphosphate aldolase from the human malaria parasite Plasmodium falciparum. Biochemistry, 37, 4388–4396.Google Scholar
First citation Kim, H., Feil, I. K., Verlinde, C. L. M. J., Petra, P. H. & Hol, W. G. J. (1995). Crystal structure of glycosomal glyceraldehyde-3-phosphate dehydrogenase from Leishmania mexicana: implication for structure-based drug design and a new position for the inorganic phosphate binding site. Biochemistry, 34, 14975–14986.Google Scholar
First citation Kim, K. K., Song, H. K., Shin, D. H., Hwang, K. Y. & Suh, S. W. (1997). The crystal structure of a triacylglycerol lipase from Pseudomonas cepacia reveals a highly open conformation in the absence of a bound inhibitor. Structure, 5, 173–185.Google Scholar
First citation Kim, Y., Yoon, K.-H., Khang, Y., Turley, S. & Hol, W. G. J. (2000). The 2.0 Å crystal structure of cephalosporin acylase. Struct. Fold. Des. 8, 1059–1068.Google Scholar
First citation Kissinger, C. R., Parge, H. E., Knighton, D. R., Lewis, C. T., Pelletier, L. A., Tempczyk, A., Kalish, V. J., Tucker, K. D., Showalter, R. E., Moomaw, E. W., Gastinel, L. N., Habuka, N., Chen, X., Maldonado, F., Barker, J. E., Bacquet, R. & Villafranca, J. E. (1995). Crystal structures of human calcineurin and the human FKBP12-FK506-calcineurin complex. Nature (London), 378, 641–644.Google Scholar
First citation Kitadokoro, K., Hagishita, S., Sato, T., Ohtan, M. & Miki, K. (1998). Crystal structure of human secretory phospholipase A2-IIA complex with the potent indolizine inhibitor 120–1032. J. Biochem. 123, 619–623.Google Scholar
First citation Kitov, P. I., Sadowska, J. M., Mulvey, G., Armstrong, G. D., Ling, H., Pannu, N. S., Read, R. J. & Bundle, D. R. (2000). Shiga-like toxins are neutralized by tailored multivalent carbohydrate ligands. Nature (London), 403, 669–672.Google Scholar
First citation Knighton, D. R., Kan, C. C., Howland, E., Janson, C. A., Hostomska, Z., Welsh, K. M. & Matthews, D. A. (1994). Structure of and kinetic channelling in bifunctional dihydrofolate reductase-thymidylate synthase. Nature Struct. Biol. 1, 186–194.Google Scholar
First citation Ko, T.-P., Safo, M. K., Musayev, F. N., Di Salvo, M. L., Wang, C., Wu, S.-H. & Abraham, D. J. (2000). Structure of human erythrocyte catalase. Acta Cryst. D56, 241–245.Google Scholar
First citation Kohara, M., Abe, S., Komatsu, T., Tago, K., Arita, M. & Nomoto, A. (1988). A recombinant virus between the Sabin 1 and Sabin 3 vaccine strains of poliovirus as a possible candidate for a new type 3 poliovirus live vaccine strain. J. Virol. 62, 2828–2835.Google Scholar
First citation Kohlstaedt, L. A., Wang, J., Friedman, J. M., Rice, P. A. & Steitz, T. A. (1992). Crystal structure at 3.5 Å resolution of HIV-1 reverse transcriptase complexed with an inhibitor. Science, 256, 1783–1790.Google Scholar
First citation Kohno, M., Funatsu, J., Mikami, B., Kugimiya, W., Matsuo, T. & Morita, Y. (1996). The crystal structure of lipase II from Rhizopus niveus at 2.2 Å resolution. J. Biochem. 120, 505–510.Google Scholar
First citation Kopka, M. L., Yoon, C., Goodsell, D., Pjura, P. & Dickerson, R. E. (1985). The molecular origin of DNA-drug specificity in netropsin and distamycin. Proc. Natl Acad. Sci. USA, 82, 1376–1380.Google Scholar
First citation Krengel, U., Petsko, G. A., Goody, R. S., Kabsch, W. & Wittinghofer, A. (1990). Refined crystal structure of the triphosphate conformation of H-ras p21 at 1.35-Å resolution: implications for the mechanism of GTP hydrolysis. EMBO J. 9, 2351–2359.Google Scholar
First citation Kuntz, I. D. (1992). Structure-based strategies for drug design and discovery. Science, 257, 1078–1082.Google Scholar
First citation Kurihara, H., Mitsui, Y., Ohgi, K., Irie, M., Mizuno, H. & Nakamura, K. T. (1992). Crystal and molecular structure of RNase Rh, a new class of microbial ribonuclease from Rhizopus niveus. FEBS Lett. 306, 189–192.Google Scholar
First citation Kuzin, A. P., Nukaga, M., Nukaga, Y., Hujer, A. M., Bonomo, R. A. & Knox, J. R. (1999). Structure of the SHV-1 beta-lactamase. Biochemistry, 38, 5720–5727.Google Scholar
First citation Kwong, P. D., Wyatt, R., Robinson, J., Sweet, R. W., Sodroski, J. & Hendrickson, W. A. (1998). Structure of an HIV gp120 envelope glycoprotein incomplex with the CD4 receptor and a neutralizing human antibody. Nature (London), 393, 648–659.Google Scholar
First citation Laba, D., Bauer, M., Huber, R., Fischer, S., Rudolph, R., Kohnert, U. & Bode, W. (1996). The 2.3 Å crystal structure of the catalytic domain of recombinant two-chain human tissue-type plasminogen activator. J. Mol. Biol. 258, 117–135.Google Scholar
First citation Lacy, D. B. & Stevens, R. C. (1998). Unraveling the structure and modes of action of bacterial toxins. Curr. Opin. Struct. Biol. 8, 778–784.Google Scholar
First citation Lacy, D. B., Tepp, W., Cohen, A. C., DasGupta, B. R. & Stevens, R. C. (1998). Crystal structure of botulinum neurotoxin type A and implications for toxicity. Nature Struct. Biol. 5, 898–902.Google Scholar
First citation Lantwin, C. B., Schlichting, I., Kabsch, W., Pai, E. F. & Krauth-Siegel, R. L. (1994). The structure of Trypanosoma cruzi trypanothione reductase in the oxidized and NADPH reduced state. Proteins, 18, 161–173.Google Scholar
First citation Le, H. V., Yao, N. & Weber, P. C. (1998). Emerging targets in the treatment of hepatitis C infection. Emerg. Ther. Targets, 2, 125–136.Google Scholar
First citation Lebron, J. A., Bennett, M. J., Vaughn, D. E., Chirino, A. J., Snow, P. M., Mintier, G. A., Feder, J. N. & Bjorkman, P. J. (1998). Crystal structure of the hemochromatosis protein HFE and characterization of its interaction with transferrin receptor. Cell, 93, 111–123.Google Scholar
First citation Lee, C. H., Saksela, K., Mirza, U. A., Chait, B. T. & Kuriyan, J. (1996). Crystal structure of the conserved core of HIV-1 Nef complexed with a Src family SH3 domain. Cell, 85, 931–942.Google Scholar
First citation Leonard, S. A., Gittis, A. G., Petrella, E. D., Pollard, T. D. & Lattman, E. E. (1997). Crystal structure of the actin-binding protein actophorin from Acanthamoeba. Nature Struct. Biol. 4, 369–373.Google Scholar
First citation Levine, M. M. & Noriega, F. (1995). A review of the current status of enteric vaccines. Papua New Guinea Med. J. 38, 325–331.Google Scholar
First citation Li, H., Dunn, J. J., Luft, B. J. & Lawson, C. L. (1997). Crystal structure of Lyme disease antigen outer surface protein A complexed with an Fab. Proc. Natl Acad. Sci. USA, 94, 3584–3589.Google Scholar
First citation Li, R., Sirawaraporn, P., Chitnumsub, P., Sirawaraporn, W. & Hol, W. G. J. (2000). Three-dimensional structure of M. tuberculosis dihydrofolate reductase reveals opportunities for the design of novel tuberculosis drugs. J. Mol. Biol. 295, 307–323.Google Scholar
First citation Li de la Sierra, I., Pernot, L., Prange, T., Saludjian, P., Schiltz, M., Fourme, R. & Padron, G. (1997). Molecular structure of the lipoamide dehydrogenase domain of a surface antigen from Neisseria meningitidis. J. Mol. Biol. 269, 129–141.Google Scholar
First citation Libson, A. M., Gittis, A. G., Collier, I. E., Marmer, B. L., Goldberg, G. I. & Lattman, E. E. (1995). Crystal structure of the haemopexin-like C-terminal domain of gelatinase A. Nature Struct. Biol. 2, 938–942.Google Scholar
First citation Liljas, A., Kannan, K. K., Bergsten, P. C., Waara, I., Fridborg, K., Strandberg, B., Carlbom, U., Jarup, L., Lovgren, S. & Petef, M. (1972). Crystal structure of human carbonic anhydrase C. Nature (London), New Biol. 235, 131–137.Google Scholar
First citation Lin, J. H., Ostovic, D. & Vacca, J. P. (1998). The integration of medicinal chemistry, drug metabolism, and pharmaceutical research and development in drug discovery and development. The story of Crixivan, an HIV protease inhibitor. Pharm. Biotechnol. 99, 233–255.Google Scholar
First citation Ling, H., Boodhoo, A., Hazes, B., Cummings, M. D., Armstrong, G. D., Brunton, J. L. & Read, R. J. (1998). Structure of the shiga-like toxin I B-pentamer complexed with an analogue of its receptor Gb3. Biochemistry, 37, 1777–1788.Google Scholar
First citation Liu, S., Fedorov, A. A., Pollard, T. D., Lattman, E. E., Almo, S. C. & Magnus, K. A. (1998). Crystal packing induces a conformational change in profilin-I from Acanthamoeba castellanii. J. Struct. Biol. 123, 22–29.Google Scholar
First citation Liu, Y., Gong, W., Huang, C. C., Herr, W. & Cheng, X. (1999). Crystal structure of the conserved core of the herpes simplex virus transcriptional regulatory protein VP16. Genes Dev. 13, 1692–1703.Google Scholar
First citation Livnah, O., Johnson, D. L., Stura, E. A., Farrell, F. X., Barbone, F. P., You, Y., Liu, K. D., Goldsmith, M. A., He, W., Krause, C. D., Petska, S., Jolliffe, L. K. & Wilson, I. A. (1998). An antagonist peptide-EPO receptor complex suggests that receptor dimerization is not sufficient for activation. Nature Struct. Biol. 5, 993–1004.Google Scholar
First citation Livnah, O., Stura, E. A., Johnson, D. L., Middleton, S. A., Mulcahy, L. S., Wrighton, N. C., Dower, W. J., Jolliffe, L. K. & Wilson, I. A. (1996). Functional mimicry of a protein hormone by a peptide agonist: the EPO receptor complex at 2.8 Å. Science, 273, 464–471.Google Scholar
First citation Lobkovsky, E., Moews, P. C., Liu, H., Zhao, H., Frere, J. M. & Knox, J. R. (1993). Evolution of an enzyme activity: crystallographic structure at 2-Å resolution of cephalosporinase from the ampC gene of Enterobacter cloacae P99 and comparison with a class A penicillinase. Proc. Natl Acad. Sci. USA, 90, 11257–11261.Google Scholar
First citation Loebermann, H., Tokuoka, R., Deisenhofer, J. & Huber, R. (1984). Human alpha 1-proteinase inhibitor. Crystal structure analysis of two crystal modifications, molecular model and preliminary analysis of the implications for function. J. Mol. Biol. 177, 531–557.Google Scholar
First citation Loll, P. J. & Lattman, E. E. (1989). The crystal structure of the ternary complex of staphylococcal nuclease, Ca2+, and the inhibitor pdTp, refined at 1.65 Å. Proteins, 5, 183–201.Google Scholar
First citation Love, R. A., Parge, H. E., Wickersham, J. A., Hostomsky, Z., Habuka, N., Moomaw, E. W., Adachi, T. & Hostomska, Z. (1996). The crystal structure of hepatitis C virus NS3 proteinase reveals a trypsin-like fold and a structural zinc binding site. Cell, 87, 331–342.Google Scholar
First citation Lovejoy, B., Cleasby, A., Hassell, A. M., Longley, K., Luther, M. A. W. D., McGeehan, G., McElroy, A. B., Drewry, D., Lambert, M. H. & Jordan, S. R. (1994). Structure of the catalytic domain of fibroblast collagenase complexed with an inhibitor. Science, 263, 375–377.Google Scholar
First citation Lovejoy, B., Hassell, A. M., Luther, M. A., Weigl, D. & Jordan, S. R. (1994). Crystal structures of recombinant 19-kDa human fibroblast collagenase complexed to itself. Biochemistry, 33, 8207–8217.Google Scholar
First citation Lukatela, G., Krauss, N., Theis, K., Selmer, T., Gieselmann, V., von Figura, K. & Saenger, W. (1998). Crystal structure of human arylsulfatase A: the aldehyde function and the metal ion at the active site suggest a novel mechanism for sulfate ester hydrolysis. Biochemistry, 37, 3654–3664.Google Scholar
First citation McGrath, M. E., Eakin, A. E., Engel, J. C., McKerrow, J. H., Craik, C. S. & Fletterick, R. J. (1995). The crystal structure of cruzain: a therapeutic target for Chagas' disease. J. Mol. Biol. 247, 251–259.Google Scholar
First citation McGrath, M. E., Palmer, J. T., Bromme, D. & Somoza, J. R. (1998). Crystal structure of human cathepsin S. Protein Sci. 7, 1294–1302.Google Scholar
First citation McLaughlin, P. J., Gooch, J. T., Mannherz, H. G. & Weeds, A. G. (1993). Structure of gelsolin segment 1-actin complex and the mechanism of filament severing. Nature (London), 364, 685–692.Google Scholar
First citation McTigue, M. A., Wickersham, J. A., Pinko, C., Showalter, R. E., Parast, C. V., Tempczyk-Russell, A., Gehring, M. R., Mroczkowski, B., Kan, C. C., Villafranca, J. E. & Appelt, K. (1999). Crystal structure of the kinase domain of human vascular endothelial growth factor receptor 2: a key enzyme in angiogenesis. Struct. Fold. Des. 7, 319–330.Google Scholar
First citation McTigue, M. A., Williams, D. R. & Tainer, J. A. (1995). Crystal structure of a schistosomal drug and vaccine target: glutathione S-transferase from Schistosoma japonica and its complex with the leading antischistosomal drug praziquantel. J. Mol. Biol. 246, 21–27.Google Scholar
First citation Maldonado, E., Soriano-Garcia, M., Moreno, A., Cabrera, N., Garza-Ramos, G., de Gomez-Puyou, M., Gomez-Puyou, A. & Perez-Montfort, R. (1998). Differences in the intersubunit contacts in triosephosphate isomerase from two closely related pathogenic trypanosomes. J. Mol. Biol. 283, 193–203.Google Scholar
First citation Malik, P. & Perham, R. N. (1997). Simultaneous display of different peptides on the surface of filamentous bacteriophage. Nucleic Acids Res. 25, 915–916.Google Scholar
First citation Mande, S. C., Mainfroid, V., Kalk, K. H., Goraj, K., Marital, J. A. & Hol, W. G. J. (1994). Crystal structure of recombinant human triosephosphate isomerase at 2.8 Å resolution. Protein Sci. 3, 810–821.Google Scholar
First citation Mande, S. C., Mehra, F., Bloom, B. R. & Hol, W. G. J. (1996). Structure of the heat shock protein chaperonin-10 of Mycobacterium leprae. Science, 271, 203–207.Google Scholar
First citation Martin, A., Wychowski, C., Couderc, T., Crainic, R., Hogle, J. & Girard, M. (1988). Engineering a poliovirus type 2 antigenic site on a type 1 capsid results in a chimaeric virus which is neurovirulent for mice. EMBO J. 7, 2839–2847.Google Scholar
First citation Mather, T., Oganesseyan, V., Hof, P., Huber, R., Foundling, S., Esmon, S. & Bode, W. (1996). The 2.8 Å crystal structure of Gla-domainless activated protein C. EMBO J. 15, 6822–6831.Google Scholar
First citation Mathews, I. I., Vanderhoff-Hanaver, P., Castellino, F. J. & Tulinsky, A. (1996). Crystal structures of the recombinant kringle 1 domain of human plasminogen in complexes with the ligands epsilon-aminocaproic acid and trans-4-(aminomethyl)cyclohexane-1-carboxylic acid. Biochemistry, 35, 2567–2576.Google Scholar
First citation Matthews, D. A., Alden, R. A., Bolin, J. T., Freer, S. T., Hamlin, R., Xuong, N., Kraut, J., Poe, M., Williams, M. & Hoogsteen, K. (1977). Dihydrofolate reductase: X-ray structure of the binary complex with methotrexate. Science, 197, 452–455.Google Scholar
First citation Matthews, D. A., Smith, W. W., Ferre, R. A., Condon, B., Budahazi, G., Sisson, W., Villafranca, J. E., Janson, C. A., McElroy, H. E., Gribskov, C. L. & Worland, S. (1994). Structure of human rhinovirus 3C protease reveals a trypsin-like polypeptide fold, RNA-binding site, and means for cleaving precursor polyprotein. Cell, 77, 761–771.Google Scholar
First citation Meng, W., Sawasdikosol, S., Burakoff, S. J. & Eck, M. J. (1999). Structure of the amino-terminal domain of Cbl complexed to its binding site on ZAP-70 kinase. Nature (London), 398, 84–90.Google Scholar
First citation Merritt, E. A., Sarfaty, S., van den Akker, F., L'hoir, C., Martial, J. A. & Hol, W. G. J. (1994). Crystal structure of cholera toxin B-pentamer bound to receptor GM1 pentasaccharide. Protein Sci. 3, 166–175.Google Scholar
First citation Merritt, E. A., Sarfaty, S., Feil, I. K. & Hol, W. G. J. (1997). Structural foundation for the design of receptor antagonists targeting E. coli heat-labile enterotoxin. Structure, 5, 1485–1499.Google Scholar
First citation Metcalf, P. & Fusek, M. (1993). Two crystal structures for cathepsin D: the lysosomal targeting signal and active site. EMBO J. 12, 1293–1302.Google Scholar
First citation Mikami, B., Adachi, M., Kage, T., Sarikaya, E., Nanmori, T., Shinke, R. & Utsumi, S. (1999). Structure of raw starch-digesting Bacillus cereus beta-amylase complexed with maltose. Biochemistry, 38, 7050–7061.Google Scholar
First citation Mikol, V., Ma, D. & Carlow, C. K. (1998). Crystal structure of the cyclophilin-like domain from the parasitic nematode Brugia malayi. Protein Sci. 7, 1310–1316.Google Scholar
First citation Milburn, M. V., Hassell, A. M., Lambert, M. H., Jordan, S. R., Proudfoot, A. E. I., Graber, P. & Wells, T. N. C. (1993). A novel dimer configuration revealed by the crystal structure at 2.4-Å resolution of human interleukin-5. Nature (London), 363, 172–176.Google Scholar
First citation Miller, M. D., Tanner, J., Alpaugh, M., Benedik, M. J. & Krause, K. L. (1994). 2.1 Å structure of Serratia endonuclease suggests a mechanism for binding to double-stranded DNA. Nature Struct. Biol. 1, 461–468.Google Scholar
First citation Minke, W. E., Hong, F., Verlinde, C. L. M. J., Hol, W. G. J. & Fan, E. (1999). Using a galactose library for exploration of a novel hydrophobic pocket in the receptor binding site of the E. coli heat-labile enterotoxin. J. Biol. Chem. 274, 33469–33473.Google Scholar
First citation Miyatake, H., Hata, Y., Fujii, T., Hamada, K., Morihara, K. & Katsube, Y. (1995). Crystal structure of the unliganded alkaline protease from Pseudomonas aeruginosa IFO3080 and its conformational changes on ligand binding. J. Biochem. (Tokyo), 118, 474–479.Google Scholar
First citation Moser, J., Gerstel, B., Meyer, J. E., Chakraborty, T., Wehland, J. & Heinz, D. W. (1997). Crystal structure of the phosphatidylinositol-specific phospholipase C from the human pathogen Listeria monocytogenes. J. Mol. Biol. 273, 269–282.Google Scholar
First citation Mottonen, J., Strand, A., Symersky, J., Sweet, R. M., Danley, D. E., Gioghegan, K. F., Gerard, R. D. & Goldsmith, E. J. (1992). Structural basis of latency in plasminogen activator inhibitor-1. Nature (London), 355, 270–273.Google Scholar
First citation Muchmore, S. W., Sattler, M., Liang, H., Meadows, R. P., Harlan, J. E., Yoon, H. S., Nettesheim, D., Chang, B. S., Thompson, C. B., Wong, S. L., Ng, S. L. & Fesik, S. W. (1996). X-ray and NMR structure of human Bcl-xL, an inhibitor of programmed cell death. Nature (London), 381, 335–341.Google Scholar
First citation Mulichak, A. M., Tulinsky, A. & Ravichandran, K. G. (1991). Crystal and molecular structure of human plasminogen kringle 4 refined at 1.9-Å resolution. Biochemistry, 30, 10576–10588.Google Scholar
First citation Mulichak, A. M., Wilson, J. E., Padmanabhan, K. & Garavito, R. M. (1998). The structure of mammalian hexokinase-1. Nature Struct. Biol. 5, 555–560.Google Scholar
First citation Muller, Y. A., Ultsch, M. H. & DeVos, A. M. (1996). The crystal structure of the extracellular domain of human tissue factor refined to 1.7-Å resolution. J. Mol. Biol. 256, 144–159.Google Scholar
First citation Muller, Y. A., Ultsch, M. H., Kelley, R. F. & DeVos, A. M. (1994). Structure of the extracellular domain of human tissue factor: location of the factor VIIa binding site. Biochemistry, 33, 10864–10870.Google Scholar
First citation Murray, C. J. & Salomon, J. A. (1998). Modeling the impact of global tuberculosis control strategies. Proc. Natl Acad. Sci. USA, 95, 13881–13886.Google Scholar
First citation Murray, I. A., Cann, P. A., Day, P. J., Derrick, J. P., Sutcliffe, M. J., Shaw, W. V. & Leslie, A. G. (1995). Steroid recognition by chloramphenicol acetyltransferase: engineering and structural analysis of a high affinity fusidic acid binding site. J. Mol. Biol. 254, 993–1005.Google Scholar
First citation Murray, M. G., Kuhn, R. J., Arita, M., Kawamura, N., Nomoto, A. & Wimmer, E. (1988). Poliovirus type 1/type 3 antigenic hybrid virus constructed in vitro elicits type 1 and type3 neutralizing antibodies in rabbits and monkeys. Proc. Natl Acad. Sci. USA, 85, 3203–3207.Google Scholar
First citation Murthy, H. M., Clum, S. & Padmanabhan, R. (1999). Dengue virus NS3 serine protease. Crystal structure and insights into interaction of the active site with substrates by molecular modeling and structural analysis of mutational effects. J. Biol. Chem. 274, 5573–5580.Google Scholar
First citation Musil, D., Zucic, D., Turk, D., Engh, R. A., Mayr, I., Huber, R., Popovic, T., Turk, V., Towatari, T., Katunuma, N. & Bode, W. (1991). The refined 2.15 Å X-ray crystal structure of human liver cathepsin B: the structural basis for its specificity. EMBO J. 10, 2321–2330.Google Scholar
First citation Nagar, B., Jones, R. G., Diefenbach, R. J., Isenman, D. E. & Rini, J. M. (1998). X-ray crystal structure of C3d: a C3 fragment and ligand for complement receptor 2. Science, 280, 1277–1281.Google Scholar
First citation Nam, H. J., Haser, W. G., Roberts, T. M. & Frederick, C. A. (1996). Intramolecular interactions of the regulatory domains of the Bcr-Abl kinase reveal a novel control mechanism. Structure, 4, 1105–1114.Google Scholar
First citation Narayana, N., Matthews, D. A., Howell, E. E. & Nguyen-huu, X. (1995). A plasmid-encoded dihydrofolate reductase from trimethoprim-resistant bacteria has a novel D2-symmetric active site. Nature Struct. Biol. 2, 1018–1025.Google Scholar
First citation National Institute of Allergy and Infectious Diseases (1998). The Jordan report: accelerated development of vaccines. NIAID, Bethesda, MD.Google Scholar
First citation Navia, M. A., Fitzgerald, P. M., McKeever, B. M., Leu, C. T., Heimbach, J. C., Herber, W. K., Sigal, I. S., Darke, P. L. & Springer, J. P. (1989). Three-dimensional structure of aspartyl protease from human immunodeficiency virus HIV-1. Nature (London), 337, 615–620.Google Scholar
First citation Navia, M. A., McKeever, B. M., Springer, J. P., Lin, T. Y., Williams, H. R., Fluder, E. M., Dorn, C. P. & Hoogsteen, K. (1989). Structure of human neutrophil elastase in complex with a peptide chloromethyl ketone inhibitor at 1.84-Å resolution. Proc. Natl Acad. Sci. USA, 86, 7–11.Google Scholar
First citation Naylor, C. E., Eaton, J. T., Howells, A., Justin, N., Moss, D. S., Titball, R. W. & Basak, A. K. (1998). Structure of the key toxin in gas gangrene. Nature Struct. Biol. 5, 738–746.Google Scholar
First citation Nurizzo, D., Silvestrini, M. C., Mathieu, M., Cutruzzola, F., Bourgeois, D., Fulop, V., Hajdu, J., Brunori, M., Tegoni, M. & Cambillau, C. (1997). N-terminal arm exchange is observed in the 2.15 Å crystal structure of oxidized nitrite reductase from Pseudomonas aeruginosa. Structure, 5, 1157–1171.Google Scholar
First citation Oefner, C., D'Arcy, A. & Winkler, F. K. (1988). Crystal structure of human dihydrofolate reductase complexed with folate. Eur. J. Biochem. 174, 377–385.Google Scholar
First citation Oinonen, C., Tikkanen, R., Rouvinen, J. & Peltonen, L. (1995). Three-dimensional structure of human lysosomal aspartylglucosaminidase. Nature Struct. Biol. 2, 1102–1108.Google Scholar
First citation Ozaki, H., Sato, T., Kubota, H., Hata, Y., Katsube, Y. & Shimonishi, Y. (1991). Molecular structure of the toxin domain of heat-stable enterotoxin produced by a pathogenic strain of Escherichia coli. A putative binding site for a binding protein on rat intestinal epithelial cell membranes. J. Biol. Chem. 266, 5934–5941.Google Scholar
First citation Padmanabhan, K., Padmanabhan, K. P., Tulinsky, A., Park, C. H., Bode, W., Huber, R., Blankenship, D. T., Cardin, A. D. & Kisiel, W. (1993). Structure of human des(1–45) factor Xa at 2.2-Å resolution. J. Mol. Biol. 232, 947–966.Google Scholar
First citation Pai, E. F., Kabsch, W., Krengel, U., Holmes, K. C., John, J. & Wittinghofer, A. (1989). Structure of the guanine-nucleotide-binding domain of the Ha-ras oncogene product p21 in the triphosphate conformation. Nature (London), 341, 209–214.Google Scholar
First citation Papageorgiou, A. C., Acharya, K. R., Shapiro, R., Passalacqua, E. F., Brehm, R. D. & Tranter, H. S. (1995). Crystal structure of the superantigen enterotoxin C2 from Staphylococcus aureus reveals a zinc-binding site. Structure, 3, 769–779.Google Scholar
First citation Papageorgiou, A. C., Tranter, H. S. & Acharya, K. R. (1998). Crystal structure of microbial superantigen staphylococcal enterotoxin B at 1.5 Å resolution: implications for superantigen recognition by MHC class II molecules and T-cell receptors. J. Mol. Biol. 277, 61–79.Google Scholar
First citation Pares, S., Mouz, N., Petillot, Y., Hakenbeck, R. & Dideberg, O. (1996). X-ray structure of Streptococcus pneumoniae PBP2x, a primary penicillin target enzyme. Nature Struct. Biol. 3, 284–289.Google Scholar
First citation Parge, H. E., Forest, K. T., Hickey, M. J., Christensen, D. A., Getzoff, E. D. & Tainer, J. A. (1995). Structure of the fibre-forming protein pilin at 2.6 Å resolution. Nature (London), 378, 32–38.Google Scholar
First citation Parge, H. E., Hallewell, R. A. & Tainer, J. A. (1992). Atomic structures of wild-type and thermostable mutant recombinant human Cu,Zn superoxide dismutase. Proc. Natl Acad. Sci. USA, 89, 6109–6113.Google Scholar
First citation Patskovsky, Y. V., Patskovska, L. N. & Listowsky, I. (1999). Functions of His107 in the catalytic mechanism of human glutathione s-transferase hGSTM1a-1a. Biochemistry, 38, 1193–1202.Google Scholar
First citation Pauptit, R. A., Karlsson, R., Picot, D., Jenkins, J. A., Niklaus-Reimer, A. S. & Jansonius, J. N. (1988). Crystal structure of neutral protease from Bacillus cereus refined at 3.0 Å resolution and comparison with the homologous but more thermostable enzyme thermolysin. J. Mol. Biol. 199, 525–537.Google Scholar
First citation Pearl, L., O'Hara, B., Drew, R. & Wilson, S. (1994). Crystal structure of AmiC: the controller of transcription antitermination in the amidase operon of Pseudomonas aeruginosa. EMBO J. 13, 5810–5817.Google Scholar
First citation Pedelacq, J. D., Maveyraud, L., Prevost, G., Bab-Moussa, L., Gonzalez, A., Coucelle, E., Shepard, W., Monteil, H., Samama, J. P. & Mourey, L. (1999). The structure of a Staphylococcus aureus leucocidin component (LukF-PV) reveals the fold of the water-soluble species of a family of transmembrane pore-forming toxins. Struct. Fold. Des. 7, 277–287.Google Scholar
First citation Pedersen, L. C., Benning, M. M. & Holden, H. M. (1995). Structural investigation of the antibiotic and ATP-binding sites in kanamycin nucleotidyltransferase. Biochemistry, 34, 13305–13311.Google Scholar
First citation Perrakis, A., Tews, I., Dauter, Z., Oppenheim, A. B., Chet, I., Wilson, K. S. & Vorgias, C. E. (1994). Crystal structure of a bacterial chitinase at 2.3 Å resolution. Structure, 2, 1169–1180.Google Scholar
First citation Perutz, M. (1992). Protein structure. New approaches to disease and therapy. New York: W. H. Freeman & Co.Google Scholar
First citation Petosa, C., Collier, R. J., Klimpel, K. R., Leppla, S. H. & Liddington, R. C. (1997). Crystal structure of the anthrax toxin protective antigen. Nature (London), 385, 833–838.Google Scholar
First citation Pfuegl, G., Kallen, J., Schirmer, T., Jansonius, J. N., Zurini, M. G. M. & Walkinshaw, M. D. (1993). X-ray structure of a decameric cyclophilin–cyclosporin crystal complex. Nature (London), 361, 91–94.Google Scholar
First citation Phillips, C., Dohnalek, J., Gover, S., Barrett, M. P. & Adams, M. J. (1998). A 2.8 Å resolution structure of 6-phosphogluconate dehydrogenase from the protozoan parasite Trypanosoma brucei: comparison with the sheep enzyme accounts for differences in activity with coenzyme and substrate analogues. J. Mol. Biol. 282, 667–681.Google Scholar
First citation Phillips, C. L., Ullman, B., Brennan, R. G. & Hill, C. P. (1999). Crystal structures of adenine phosphoribosyltransferase from Leishmania donovani. EMBO J. 18, 3533–3545.Google Scholar
First citation Pineo, G. F. & Hull, R. D. (1999). Thrombin inhibitors as anticoagulant agents. Curr. Opin. Hematol. 6, 298–303.Google Scholar
First citation Poljak, R. J., Amzel, L. M., Avey, H. P., Chen, B. L., Phizackerley, R. P. & Saul, F. (1973). Three-dimensional structure of the Fab′ fragment of a human immunoglobulin at 2.8-Å resolution. Proc. Natl Acad. Sci. USA, 70, 3305–3310.Google Scholar
First citation Porter, R. (1999). The greatest benefit to mankind: a medical history of humanity. New York: W. W. Norton and Co., Inc.Google Scholar
First citation Prasad, G. S., Earhart, C. A., Murray, D. L., Novick, R. P., Schlievert, P. M. & Ohlendorf, D. H. (1993). Structure of toxic shock syndrome toxin 1. Biochemistry, 32, 13761–13766.Google Scholar
First citation Pratt, K. P., Cote, H. C., Chung, D. W., Stenkamp, R. E. & Davie, E. W. (1997). The primary fibrin polymerization pocket: three-dimensional structure of a 30-kDa C-terminal gamma chain fragment complexed with the peptide Gly-Pro-Arg-Pro. Proc. Natl Acad. Sci. USA, 94, 7176–7181.Google Scholar
First citation Pratt, K. P., Shen, B. W., Takeshima, K., Davie, E. W., Fujikawa, K. & Stoddard, B. L. (1999). Structure of the C2 domain of human factor VIII at 1.5 Å resolution. Nature (London), 402, 439–442.Google Scholar
First citation Priestle, J. P., Schar, H. P. & Grutter, M. G. (1988). Crystal structure of the cytokine interleukin-1 beta. EMBO J. 7, 339–343.Google Scholar
First citation Qiu, X., Culp, J. S., DiLella, A. G., Hellmig, B., Hoog, S. S., Janson, C. A., Smith, W. W. & Abdel-Meguid, S. S. (1996). Unique fold and active site in cytomegalovirus protease. Nature (London), 383, 275–279.Google Scholar
First citation Qiu, X., Janson, C. A., Culp, J. S., Richardson, S. B., Debouck, C., Smith, W. W. & Abdel-Meguid, S. S. (1997). Crystal structure of varicella-zoster virus protease. Proc. Natl Acad. Sci. USA, 94, 2874–2879.Google Scholar
First citation Qiu, X., Verlinde, C. L. M. J., Zhang, S., Schmitt, M. P., Holmes, R. K. & Hol, W. G. J. (1995). Three-dimensional structure of the diphtheria toxin repressor in complex with divalent cation co-repressors. Structure, 3, 87–100.Google Scholar
First citation Rabijns, A., De Bondt, H. L. & De Ranter, C. (1997). Three-dimensional structure of staphylokinase, a plasminogen activator with therapeutic potential. Nature Struct. Biol. 4, 357–360.Google Scholar
First citation Radhakrishnan, R., Walter, L. J., Hruza, A., Reichert, P., Trotta, P. P., Nagabhushan, T. L. & Walter, M. R. (1996). Zinc mediated dimer of human interferon-alpha 2b revealed by X-ray crystallography. Structure, 4, 1453–1463.Google Scholar
First citation Raghunathan, S., Chandross, R. J., Kretsinger, R. H., Allison, T. J., Penington, C. J. & Rule, G. S. (1994). Crystal structure of human class mu glutathione transferase GSTM2–2. Effects of lattice packing on conformational heterogeneity. J. Mol. Biol. 238, 815–832.Google Scholar
First citation Rano, T. A., Timkey, T., Peterson, E. P., Rotonda, J., Nicholson, D. W., Becker, J. W., Chapman, K. T. & Thornberry, N. A. (1997). A combinatorial approach for determining protease specificities: application to interleukin-1 beta converting enzyme (ICE). Chem. Biol. 4, 149–155.Google Scholar
First citation Rao, Z., Handford, P., Mayhew, M., Knott, V., Brownlee, G. G. & Stuart, D. (1995). The structure of a Ca(2+)-binding epidermal growth factor-like domain: its role in protein–protein interactions. Cell, 82, 131–141.Google Scholar
First citation Read, J. A., Wilkinson, K. W., Tranter, R., Sessions, R. B. & Brady, R. L. (1999). Chloroquine binds in the cofactor binding site of Plasmodium falciparum lactate dehydrogenase. J. Biol. Chem. 274, 10213–10218.Google Scholar
First citation Redinbo, M. R., Stewart, L., Kuhn, P., Champoux, J. J. & Hol, W. G. J. (1998). Crystal structures of human topoisomerase I in covalent and noncovalent complexes with DNA. Science, 279, 1504–1513.Google Scholar
First citation Reichmann, L., Clark, M., Waldmann, H. & Winter, G. (1988). Reshaping human antibodies for therapy. Nature (London), 332, 323–327.Google Scholar
First citation Reinemer, P., Dirr, H. W., Ladenstein, R., Huber, R., Lo Bello, M., Federici, G. & Parker, M. W. (1992). Three-dimensional structure of class pi glutathione S-transferase from human placenta in complex with S-hexylglutathione at 2.8 Å resolution. J. Mol. Biol. 227, 214–226.Google Scholar
First citation Reinemer, P., Grams, F., Huber, R., Kleine, T., Schnierer, S., Piper, M., Tschesche, H. & Bode, W. (1994). Structural implications for the role of the N terminus in the `superactivation' of collagenases. A crystallographic study. FEBS Lett. 338, 227–233.Google Scholar
First citation Ren, J., Esnouf, R., Garman, E., Somers, D., Ross, C., Kirby, I., Keeling, J., Darby, G., Jones, Y., Stuart, D. & Stammers, D. (1995). High resolution structures of HIV-1 RT from four RT-inhibitor complexes. Nature Struct. Biol. 2, 293–302.Google Scholar
First citation Ren, J., Esnouf, R. M., Hopkins, A. L., Jones, E. Y., Kirby, I., Keeling, J., Ross, C. K., Larder, B. A., Stuart, D. I. & Stammers, D. K. (1998). 3-Azido-3′-deoxythymidine drug resistance mutations in HIV-1 reverse transcriptase can induce long range conformational changes. Proc. Natl Acad. Sci. USA, 95, 9518–9523.Google Scholar
First citation Renwick, S. B., Snell, K. & Baumann, U. (1998). The crystal structure of human cytosolic serine hydroxymethyltransferase: a target for cancer chemotherapy. Structure, 15, 1105–1116.Google Scholar
First citation Resnick, D. A., Smith, A. D., Geisler, S. C., Zhang, A., Arnold, E. & Arnold, G. F. (1995). Chimeras from a human rhinovirus 14–human immunodeficiency virus type 1 (HIV-1) V3 loop seroprevalence library induce neutralizing responses against HIV-1. J. Virol. 69, 2406–2411.Google Scholar
First citation Rey, F. A., Heinz, F. X., Mandl, C., Kunz, C. & Harrison, S. C. (1995). The envelope glycoprotein from tick-borne encephalitis virus at 2 Å resolution. Nature (London), 375, 291–298.Google Scholar
First citation Rigden, D. J., Phillips, S. E., Michels, P. A. & Fothergill-Gilmore, L. A. (1999). The structure of pyruvate kinase from Leishmania mexicana reveals details of the allosteric transition and unusual effector specificity. J. Mol. Biol. 291, 615–635.Google Scholar
First citation Ripka, W. C. (1997). Design of antithrombotic agents directed at factor Xa. In Structure-based drug design, edited by P. Veerapandian, pp. 265–294. New York: Marcel Dekker. Google Scholar
First citation Roberts, M. M., White, J. L., Grutter, M. G. & Burnett, R. M. (1986). Three-dimensional structure of the adenovirus major coat protein hexon. Science, 232, 1148–1151.Google Scholar
First citation Rodgers, D. W., Bamblin, S. J., Harris, B. A., Ray, S., Culp, J. S., Hellmig, B., Woolf, D. J., Debouck, C. & Harrison, S. C. (1995). The structure of unliganded reverse transcriptase from the human immunodeficiency virus type 1. Proc. Natl Acad. Sci. USA, 92, 1222–1226.Google Scholar
First citation Roe, S. M., Barlow, T., Brown, T., Oram, M., Keeley, A., Tsaneva, I. R. & Pearl, L. H. (1998). Crystal structure of an octameric RuvA–Holliday junction complex. Mol. Cell, 2, 361–372.Google Scholar
First citation Rolan, P. E., Parker, J. E., Gray, S. J., Weatherley, B. C., Ingram, J., Leavens, W., Wootton, R. & Posner, J. (1993). The pharmacokinetics, tolerability and pharmacodynamics of tucaresol (589C80; 4[2-formyl-3-hydroxyphenoxymethyl] benzoic acid), a potential anti-sickling agent, following oral administration to healthy subjects. Br. J. Clin. Pharmacol. 35, 419–425.Google Scholar
First citation Rossjohn, J., Feil, S. C., McKinstry, W. J., Tweten, R. K. & Parker, M. W. (1997). Structure of a cholesterol-binding, thiol-activated cytolysin and a model of its membrane form. Cell, 89, 685–692.Google Scholar
First citation Rossjohn, J., Feil, S. C., Wilce, M. C. J., Sexton, J. L., Spithill, T. W. & Parker, M. W. (1997). Crystallization, structural determination and analysis of a novel parasite vaccine candidate: Fasciola hepatica glutathione S-transferase. J. Mol. Biol. 273, 857–872.Google Scholar
First citation Rossjohn, J., McKinstry, W. J., Oakley, A. J., Verger, D., Flanagan, J., Chelvanayagam, G., Tan, K. L., Board, P. G. & Parker, M. W. (1998). Human theta class glutathione transferase: the crystal structure reveals a sulfate-binding pocket within a buried active site. Structure, 6, 309–322.Google Scholar
First citation Rossjohn, J., Polekhina, G., Feil, S. C., Allocati, N., Masulli, M., De Illio, C. & Parker, M. W. (1998). A mixed disulfide bond in bacterial glutathione transferase: functional and evolutionary implications. Structure, 6, 721–734.Google Scholar
First citation Rossmann, M. G., Arnold, E., Erickson, J. W., Frankenberger, E. A., Griffith, J. P., Hecht, H. J., Johnson, J. E., Kamer, G., Luo, M., Mosser, A. G., Rueckert, R. R., Sherry, B. & Vriend, G. (1985). Structure of a human common cold virus and functional relationship to other picornaviruses. Nature (London), 317, 145–153.Google Scholar
First citation Roussel, A., Anderson, B. F., Baker, H. M., Fraser, J. D. & Baker, E. N. (1997). Crystal structure of the streptococcal superantigen SPE-C: dimerization and zinc binding suggest a novel mode of interaction with MHC class II molecules. Nature Struct. Biol. 4, 635–643.Google Scholar
First citation Rudenko, G., Bonten, E., d'Azzo, A. & Hol, W. G. J. (1995). Three-dimensional structure of the human `protective protein': structure of the precursor form suggests a complex activation mechanism. Structure, 3, 1249–1259.Google Scholar
First citation Rudenko, G., Bonten, E., Hol, W. G. J. & d'Azzo, A. (1998). The atomic model of the human protective protein/cathepsin A suggests a structural basis for galactosialidosis. Proc. Natl Acad. Sci. USA, 95, 621–625.Google Scholar
First citation Russo, A. A., Tong, L., Lee, J. O., Jeffrey, P. D. & Pavletich, N. P. (1998). Structural basis for inhibition of the cyclin-dependent kinase Cdk6 by the tumour suppressor p16INK4a. Nature (London), 395, 237–243.Google Scholar
First citation Rydel, T. J., Ravichandran, K. G., Tulinsky, A., Bode, W., Huber, R., Roitsch, C. & Fenton, J. W. I. (1990). The structure of a complex of recombinant hirudin and human alpha-thrombin. Science, 249, 277–280.Google Scholar
First citation Rydel, T. J., Yin, M., Padmanabhan, K. P., Blankenship, D. T., Cardin, A. D., Correa, P. E., Fenton, J. W. I. & Tulinsky, A. (1994). Crystallographic structure of human gamma-thrombin. J. Biol. Chem. 269, 22000–22006.Google Scholar
First citation Sarafianos, S. G., Das, K., Ding, J., Boyer, P. L., Hughes, S. H. & Arnold, E. (1999). Touching the heart of HIV-1 drug resistance: the fingers close down on the dNTP at the polymerase active site. Chem. Biol. 6, R137–R146.Google Scholar
First citation Sauer, F. G., Futterer, K., Pinkner, J. S., Dodson, K. W., Hultgren, S. J. & Waksman, G. (1999). Structural basis of chaperone function and pilus biogenesis. Science, 285, 1058–1061.Google Scholar
First citation Savva, R., McAuley-Hecht, K., Brown, T. & Pearl, L. (1995). The structural basis of specific base-excision repair by uracil-DNA glycosylase. Nature (London), 373, 487–493.Google Scholar
First citation Scheffzek, K., Ahmadian, M. R., Kabsch, W., Wiesmuller, L., Lautwein, A., Schmitz, F. & Wittinghofer, A. (1997). The Ras–RasGAP complex: structural basis for GTPase activation and its loss in oncogenic Ras mutants. Science, 277, 333–338.Google Scholar
First citation Schiffer, C. A., Clifton, I. J., Davisson, V. J., Santi, D. V. & Stroud, R. M. (1995). Crystal structure of human thymidylate synthase: a structural mechanism for guiding substrates into the active site. Biochemistry, 34, 16279–16287.Google Scholar
First citation Schlagenhauf, E., Etges, R. & Metcalf, P. (1998). The crystal structure of the Leishmania major surface proteinase leishmanolysin (gp63). Structure, 6, 1035–1046.Google Scholar
First citation Schonbrunn, E., Sack, S., Eschenberg, S., Perrakis, A., Krekel, F., Amrhein, N. & Mandelkow, E. (1996). Crystal structure of UDP-N-acetylglucosamine enolpyruvyltransferase, the target of the antibiotic fosfomycin. Structure, 4, 1065–1075.Google Scholar
First citation Schreuder, H., Tardif, C., Trump-Kallmeyer, S., Soffientini, A., Sarubbi, E., Akeson, A., Bowlin, T., Yanofsky, S. & Barrett, R. W. (1997). A new cytokine-receptor binding mode revealed by the crystal structure of the IL-1 receptor with an antagonist. Nature (London), 386, 194–200.Google Scholar
First citation Schreuder, H. A., de Boer, B., Dijkema, R., Mulders, J., Theunissen, H. J. M., Grootenhuis, P. D. J. & Hol, W. G. J. (1994). The intact and cleaved human antithrombin III complex as a model for serpin-proteinase interactions. Nature Struct. Biol. 1, 48–54.Google Scholar
First citation Schumacher, M. A., Carter, D., Ross, D. S., Ullman, B. & Brennan, R. G. (1996). Crystal structures of Toxoplasma gondii HGXPRTase reveal the catalytic role of a long flexible loop. Nature Struct. Biol. 3, 881–887.Google Scholar
First citation Schumacher, M. A., Carter, D., Scott, D. M., Roos, D. S., Ullman, B. & Brennan, R. G. (1998). Crystal structures of Toxoplasma gondii uracil phosphoribosyltransferase reveal the atomic basis of pyrimidine discrimination and prodrug binding. EMBO J. 17, 3219–3232.Google Scholar
First citation Schwabe, J. W. E., Chapman, L., Finch, J. T. & Rhodes, D. (1993). The crystal structure of the estrogen receptor DNA-binding domain bound to DNA: how receptors discriminate between their response elements. Cell, 75, 567–578.Google Scholar
First citation Scott, D. L., White, S. P., Browning, J. L., Rosa, J. J., Gelb, M. H. & Sigler, P. B. (1991). Structures of free and inhibited human secretory phospholipase A2 from inflammatory exudate. Science, 254, 1007–1010.Google Scholar
First citation Sha, B. & Luo, M. (1997). Structure of a bifunctional membrane-RNA binding protein, influenza virus matrix protein M1. Nature Struct. Biol. 4, 239–244.Google Scholar
First citation Shah, S. A., Shen, B. W. & Brunger, A. T. (1997). Human ornithine aminotransferase complexed with L-canaline and gabaculine: structural basis for substrate recognition. Structure, 5, 1067–1075.Google Scholar
First citation Sharma, A., Hanai, R. & Mondragon, A. (1994). Crystal structure of the amino-terminal fragment of vaccinia virus DNA topoisomerase I at 1.6 Å resolution. Structure, 2, 767–777.Google Scholar
First citation Sharma, V., Grubmeyer, C. & Sacchettini, J. C. (1998). Crystal structure of quinolinic acid phosphoribosyltransferase from Mycobacterium tuberculosis: a potential TB drug target. Structure, 6, 1587–1599.Google Scholar
First citation Sherry, B., Mosser, A. G., Colonno, R. J. & Rueckert, R. R. (1986). Use of monoclonal antibodies to identify four neutralization immunogens on a common cold picornavirus, human rhinovirus 14. J. Virol. 57, 246–257.Google Scholar
First citation Sherry, B. & Rueckert, R. (1985). Evidence for at least two dominant neutralization antigens on human rhinovirus 14. J. Virol. 53, 137–143.Google Scholar
First citation Shi, D., Morizono, H., Ha, Y., Aoyagi, M., Tuchman, M. & Allewell, N. M. (1998). 1.85-Å resolution crystal structure of human ornithine transcarbamoylase complexed with N-phosphonacetyl-L-ornithine. Catalytic mechanism and correlation with inherited deficiency. J. Biol. Chem. 273, 34247–34254.Google Scholar
First citation Shi, W., Li, C. M., Tyler, P. C., Furneaux, R. H., Cahill, S. M., Girvin, M. E., Grubmeyer, C., Schramm, V. L. & Almo, S. C. (1999). The 2.0 Å structure of malarial purine phosphoribosyltransferase in complex with a transition-state analogue inhibitor. Biochemistry, 38, 9872–9880.Google Scholar
First citation Shi, W., Schramm, V. L. & Almo, S. C. (1999). Nucleoside hydrolase from Leishmania major. Cloning, expression, catalytic properties, transition state inhibitors, and the 2.5-Å crystal structure. J. Biol. Chem. 274, 21114–21120.Google Scholar
First citation Shieh, H. S., Kurumbail, R. G., Stevens, A. M., Stegeman, R. A., Sturman, E. J., Pak, J. Y., Wittwer, A. J., Palmier, M. O., Wiegand, R. C., Holwerda, B. C. & Stallings, W. C. (1996). Three-dimensional structure of human cytomegalovirus protease. Nature (London), 383, 279–282.Google Scholar
First citation Sielecki, A. R., Hayakawa, K., Fujinaga, M., Murphy, M. E. P., Fraser, M., Muir, A. K., Carilli, C. T., Lewicki, J. A., Baxter, J. D. & James, M. N. G. (1989). Structure of recombinant human renin, a target for cardiovascular-active drugs, at 2.5-Å resolution. Science, 243, 1346–1351.Google Scholar
First citation Silva, A. M., Lee, A. Y., Gulnik, S. V., Maier, P., Collins, J., Bhat, T. N., Collins, P. J., Cachau, R. E., Luker, K. E., Gluzman, I. Y., Francis, S. E., Oksman, A., Goldberg, D. E. & Erickson, J. W. (1996). Structure and inhibition of plasmepsin II, a hemoglobin-degrading enzyme from Plasmodium falciparum. Proc. Natl Acad. Sci. USA, 93, 10034–10039.Google Scholar
First citation Silvian, L. F., Wang, J. & Steitz, T. A. (1999). Insights into editing from an ile-tRNA synthetase structure with tRNAile and mupirocin. Science, 285, 1074–1077.Google Scholar
First citation Sinning, I., Kleywegt, G. J., Cowan, S. W., Reinemer, P., Dirr, H. W., Huber, R., Gilliland, G. L., Armstrong, R. N., Ji, X., Board, P. G., Olin, B., Mannervik, B. & Jones, T. A. (1993). Structure determination and refinement of human alpha class glutathione transferase A1-1, and a comparison with the mu and pi class enzymes. J. Mol. Biol. 232, 192–212.Google Scholar
First citation Sixma, T. K., Pronk, S. E., Kalk, K. H., Wartna, E. S., van Zanten, B. A. M., Witholt, B. & Hol, W. G. J. (1991). Crystal structure of a cholera toxin-related heat-labile enterotoxin from E. coli. Nature (London), 351, 371–377.Google Scholar
First citation Skinner, R., Abrahams, J. P., Whisstock, J. C., Lesk, A. M., Carrell, R. W. & Wardell, M. R. (1997). The 2.6 Å structure of antithrombin indicates a conformational change at the heparin binding site. J. Mol. Biol. 266, 601–609.Google Scholar
First citation Skinner, R., Chang, W. S. W., Jin, L., Pei, X. Y., Huntington, J. A., Abrahams, J. P., Carrell, R. W. & Lomas, D. A. (1998). Implications for function and therapy of a 2.9 Å structure of binary-complexed antithrombin. J. Mol. Biol. 283, 9–14.Google Scholar
First citation Smith, A. D., Geisler, S. C., Chen, A. A., Resnick, D. A., Roy, B. M., Lewi, P. J., Arnold, E. & Arnold, G. F. (1998). Human rhinovirus type 14:human immunodeficiency virus type 1 (HIV-1) V3 loop chimeras from a combinatorial library induce potent neutralizing antibody responses against HIV-1. J. Virol. 72, 651–659.Google Scholar
First citation Somers, W., Ultsch, M., DeVos, A. M. & Kossiakoff, A. A. (1994). The X-ray structure of a growth hormone–prolactin receptor complex. Nature (London), 372, 478–481.Google Scholar
First citation Song, L., Hobaugh, M. R., Shustak, C., Cheley, S., Bayley, H. & Gouaux, J. E. (1996). Structure of staphylococcal alpha-hemolysin, a heptameric transmembrane pore. Science, 274, 1859–1866.Google Scholar
First citation Souza, D. H., Garratt, R. C., Araujo, A. P., Guimaraes, B. G., Jesus, W. D., Michels, P. A., Hannaert, V. & Oliva, G. (1998). Trypanosoma cruzi glycosomal glyceraldehyde-3-phosphate dehydrogenase: structure, catalytic mechanism and targeted inhibitor design. FEBS Lett. 424, 131–135.Google Scholar
First citation Spraggon, G., Everse, S. J. & Doolittle, R. F. (1997). Crystal structures of fragment D from human fibrinogen and its crosslinked counterpart from fibrin. Nature (London), 389, 455–462.Google Scholar
First citation Spraggon, G., Phillips, C., Nowak, U. K., Ponting, C. P., Saunders, D., Dobson, C. M., Stuart, D. I. & Jones, E. Y. (1995). The crystal structure of the catalytic domain of human urokinase-type plasminogen activator. Structure, 3, 681–691.Google Scholar
First citation Spurlino, J. C., Smallwood, A. M., Carlton, D. D., Banks, T. M., Vavra, K. J., Johnson, J. S., Cook, E. R., Falvo, J., Wahl, R. D., Pulvino, T. A., Wendoloski, J. J. & Smith, D. L. (1994). 1.56-Å structure of mature truncated human fibroblast collagenase. Proteins, 19, 98–109.Google Scholar
First citation Stams, T., Spurlino, J. C., Smith, D. L., Wahl, R. C., Ho, T. F., Qoronfleh, M. H., Banks, T. M. & Rubin, B. (1994). Structure of human neutrophil collagenase reveals large S1′ specificity pocket. Nature Struct. Biol. 1, 119–123.Google Scholar
First citation Stein, P. E., Boodhoo, A., Armstrong, G. D., Cockle, S. A., Klein, M. H. & Read, R. J. (1994). The crystal structure of pertussis toxin. Structure, 2, 45–57.Google Scholar
First citation Stein, P. E., Boodhoo, A., Tyrrell, G. J., Brunton, J. L. & Read, R. J. (1992). Crystal structure of the cell-binding B oligomer of verotoxin-1 from E. coli. Nature (London), 355, 748–750.Google Scholar
First citation Stewart, L., Redinbo, M. R., Qiu, X., Hol, W. G. J. & Champoux, J. J. (1998). A model for the mechanism of human topoisomerase I. Science, 279, 1534–1541.Google Scholar
First citation Stubbs, M. T., Laber, B., Bode, W., Huber, R., Jerala, R., Lenarcic, B. & Turk, V. (1990). The refined 2.4 Å X-ray crystal structure of recombinant human stefin B in complex with the cysteine proteinase papain: a novel type of proteinase inhibitor interaction. EMBO J. 9, 1939–1947.Google Scholar
First citation Stuckey, J. A., Schubert, H. L., Fauman, E. B., Zhang, Z. Y., Dixon, J. E. & Saper, M. A. (1994). Crystal structure of Yersinia protein tyrosine phosphatase at 2.5 Å and the complex with tungstate. Nature (London), 370, 571–575.Google Scholar
First citation Sugio, S., Kashima, A., Mochizuki, S., Noda, M. & Kobayashi, K. (1999). Crystal structure of human serum albumin at 2.5 Å resolution. Protein Eng. 12, 439–446.Google Scholar
First citation Suguna, K., Bott, R. R., Padlan, E. A., Subramanian, E., Sheriff, S., Cohen, G. H. & Davies, D. R. (1987). Structure and refinement at 1.8 Å resolution of the aspartic proteinase from Rhizopus chinensis. J. Mol. Biol. 196, 877–900.Google Scholar
First citation Sundstrom, M., Hallen, D., Svensson, A., Schad, E., Dohlsten, M. & Abrahmsen, L. (1996). The co-crystal structure of staphylococcal enterotoxin type A with Zn2+ at 2.7 Å resolution. Implications for major histocompatibility complex class II binding. J. Biol. Chem. 271, 32212–32216.Google Scholar
First citation Symersky, J., Patti, J. M., Carson, M., House-Pompeo, K., Teale, M., Moore, D., Jin, L., Schneider, A., DeLucas, L. J., Hook, M. & Narayana, S. V. (1997). Structure of the collagen-binding domain from a Staphylococcus aureus adhesin. Nature Struct. Biol. 4, 833–838.Google Scholar
First citation Taylor, P., Page, A. P., Kontopidis, G., Husi, H. & Walkinshaw, M. D. (1998). The X-ray structure of a divergent cyclophilin from the nematode parasite Brugia malayi. FEBS Lett. 425, 361–366.Google Scholar
First citation Testa, B. (1994). Drug metabolism. In Burger's medicinal chemistry and drug discovery, 5th ed., edited by M. E. Wolf, Vol. 1. New York: John Wiley & Sons.Google Scholar
First citation Tews, I., Perrakis, A., Oppenheim, A., Dauter, Z., Wilson, K. S. & Vorgias, C. E. (1996). Bacterial chitobiase structure provides insight into catalytic mechanism and the basis of Tay–Sachs disease. Nature Struct. Biol. 3, 638–648.Google Scholar
First citation Thayer, M. M., Flaherty, K. M. & McKay, D. B. (1991). Three-dimensional structure of the elastase of Pseudomonas aeruginosa at 1.5-Å resolution. J. Biol. Chem. 266, 2864–2871.Google Scholar
First citation Thieme, R., Pai, E. F., Schirmer, R. H. & Schulz, G. E. (1981). Three-dimensional structure of glutathione reductase at 2 Å resolution. J. Mol. Biol. 152, 763–782.Google Scholar
First citation Thompson, S. K., Halbert, S. M., Bossard, M. J., Tomaszek, T. A., Levy, M. A., Zhao, B., Smith, W. W., Abdel-Meguid, S. S., Janson, C. A., D'Alessio, K. J., McQueney, M. S., Amegadzie, B. Y., Hanning, C. R., DesJarlais, R. L., Briand, J., Sarkar, S. K., Huddleston, M. J., James, C. F., Carr, S. A., Garges, K. T., Shu, A., Heys, J. R., Bradbeer, J., Zembryki, D. & Veber, D. F. (1997). Design of potent and selective human cathepsin K inhibitors that span the active site. Proc. Natl Acad. Sci. USA, 94, 14249–14254.Google Scholar
First citation Thylefors, B., Negrel, A. D., Pararajasegaram, R. & Dadzie, K. Y. (1995). Global data on blindness. Bull. WHO, 73, 115–121.Google Scholar
First citation Tiffany, K. A., Roberts, D. L., Wang, M., Paschke, R., Mohsen, A. W., Vockley, J. & Kim, J. J. (1997). Structure of human isovaleryl-CoA dehydrogenase at 2.6 Å resolution: structural basis for substrate specificity. Biochemistry, 36, 8455–8464.Google Scholar
First citation Toney, M. D., Hohenester, E., Cowan, S. W. & Jansonius, J. N. (1993). Dialkylglycine decarboxylase structure: bifunctional active site and alkali metal sites. Science, 261, 756–759.Google Scholar
First citation Tong, L., Pav, S., White, D. M., Rogers, S., Crane, K. M., Cywin, C. L., Brown, M. L. & Pargellis, C. A. (1997). A highly specific inhibitor of human p38 MAP kinase binds in the ATP pocket. Nature Struct. Biol. 4, 311–316.Google Scholar
First citation Tong, L., Qian, C., Massariol, M. J., Bonneau, P. R., Cordingley, M. G. & Lagace, L. (1996). A new serine-protease fold revealed by the crystal structure of human cytomegalovirus protease. Nature (London), 383, 272–275.Google Scholar
First citation Tran, P. H., Korszun, Z. R., Cerritelli, S., Springhorn, S. S. & Lacks, S. A. (1998). Crystal structure of the DpnM DNA adenine methyltransferase from the Dpnii restriction system of Streptococcus pneumoniae bound to S-adenosylmethionine. Structure, 6, 1563–1575.Google Scholar
First citation Tskovsky, Y. V., Patskovska, L. N. & Listowsky, I. (1999). Functions of His107 in the catalytic mechanism of human glutathione s-transferase hGSTM1a-1a. Biochemistry, 38, 1193–1202.Google Scholar
First citation Turner, B. G. & Summers, M. F. (1999). Structural biology of HIV. J. Mol. Biol. 285, 1–32.Google Scholar
First citation Umland, T. C., Wingert, L. M., Swaminathan, S., Furey, W. F., Schmidt, J. J. & Sax, M. (1997). Structure of the receptor binding fragment HC of tetanus neurotoxin. Nature Struct. Biol. 4, 788–792.Google Scholar
First citation Van Duyne, G. D., Standaert, R. F., Karplus, P. A., Schreiber, S. L. & Clardy, J. (1991). Atomic structure of FKBP-FK506, an immunophilin–immunosuppressant complex. Science, 252, 839–842.Google Scholar
First citation Van Duyne, G. D., Standaert, R. F., Karplus, P. A., Schreiber, S. L. & Clardy, J. (1993). Atomic structures of the human immunophilin FKBP-12 complexes with FK506 and rapamycin. J. Mol. Biol. 229, 105–124.Google Scholar
First citation Van Duyne, G. D., Standaert, R. F., Schreiber, S. L. & Clardy, J. (1991). Atomic structure of the rapamycin human immunophilin FKBP-12 complex. J. Am. Chem. Soc. 113, 7433–7434.Google Scholar
First citation Varghese, J. N., Laver, W. G. & Colman, P. M. (1983). Structure of the influenza virus glycoprotein antigen neuraminidase at 2.9 Å resolution. Nature (London), 303, 35–40.Google Scholar
First citation Varney, M. D., Palmer, C. L., Romines, W. H. R., Boritzki, T., Margosiak, S. A., Almassy, R., Janson, C. A., Bartlett, C., Howland, E. J. & Ferre, R. (1997). Protein structure-based design, synthesis, and biological evaluation of 5-thia-2,6-diamino-4(3H)-oxopyrimidines: potent inhibitors of glycinamide ribonucleotide transformylase with potent cell growth inhibition. J. Med. Chem. 40, 2502–2524.Google Scholar
First citation Vath, G. M., Earhart, C. A., Rago, J. V., Kim, M. H., Bohach, G. A., Schlievert, P. M. & Ohlendorf, D. H. (1997). The structure of the superantigen exfoliative toxin A suggests a novel regulation as a serine protease. Biochemistry, 36, 1559–1566.Google Scholar
First citation Veerapandian, P. (1997). Editor. Structure-based drug design. New York: Marcel-Dekker.Google Scholar
First citation Velanker, S. S., Ray, S. S., Gokhale, R. S., Suma, S., Balaram, P. & Murthy, M. R. (1997). Triosephosphate isomerase from Plasmodium falciparum: the crystal structure provides insights into antimalarial drug design. Structure, 5, 751–761.Google Scholar
First citation Vellieux, F. M., Hajdu, J., Verlinde, C. L., Groendijk, H., Read, R. J., Greenhough, T. J., Campbell, J. W., Kalk, K. H., Littlechild, J. A., Watson, H. C. & Hol, W. G. J. (1993). Structure of glycosomal glyceraldehyde-3-phosphate dehydrogenase from Trypanosoma brucei determined from Laue data. Proc. Natl Acad. Sci. USA, 90, 2355–2359.Google Scholar
First citation Verlinde, C. L. M. J. & Hol, W. G. J. (1994). Structure-based drug design: progress, results and challenges. Structure, 2, 577–587.Google Scholar
First citation Vigers, G. P., Anderson, L. J., Caffes, P. & Brandhuber, B. J. (1997). Crystal structure of the type-I interleukin-1 receptor complexed with interleukin-1beta. Nature (London), 386, 190–194.Google Scholar
First citation Villeret, V., Tricot, C., Stalon, V. & Dideberg, O. (1995). Crystal structure of Pseudomonas aeruginosa catabolic ornithine transcarbamoylase at 3.0-Å resolution: a different oligomeric organization in the transcarbamoylase family. Proc. Natl Acad. Sci. USA, 92, 10762–10766.Google Scholar
First citation Vondrasek, J., van Buskirk, C. P. & Wlodawer, A. (1997). Database of three-dimensional structure of HIV proteinases. Nature Struct. Biol. 4, 8.Google Scholar
First citation Waksman, G., Shoelson, S. E., Pant, N., Cowburn, D. & Kuriyan, J. (1993). Binding of a high affinity phosphotyrosyl peptide to the Src SH2 domain: crystal structures of the complexed and peptide-free forms. Cell, 72, 779–790.Google Scholar
First citation Walker, N. P. C., Talanian, R. V., Brady, K. D., Dang, L. C., Bump, N. J., Ferenz, C. R., Franklin, S., Ghayur, T., Hackett, M. C., Hammill, L. D., Herzog, L., Hugunin, M., Houy, W., Mankovich, J. A., McGuiness, L., Orlewicz, E., Paskind, M., Pratt, C. A., Reis, P., Summani, A., Terranova, M. & Welch, J. P. (1994). Crystal structure of the cysteine protease interleukin-1 beta-converting enzyme: a (p20/p10)2 homodimer. Cell, 78, 343–352.Google Scholar
First citation Walsh, C. T., Fisher, S. L., Park, I. S., Prahalad, M. & Wu, Z. (1996). Bacterial resistance to vancomycin: five genes and one missing hydrogen bond tell the story. Chem. Biol. 3, 21–28.Google Scholar
First citation Walter, M. R., Windsor, W. T., Nagabhushan, T. L., Lundell, D. J., Lunn, C. A., Zauodny, P. J. & Narula, S. K. (1995). Crystal structure of a complex between interferon-gamma and its soluble high-affinity receptor. Nature (London), 376, 230–235.Google Scholar
First citation Wang, A. H., Ughetto, G., Quigley, G. J. & Rich, A. (1987). Interactions between an anthracycline antibiotic and DNA: molecular structure of daunomycin complexed to d(CpGpTpApCpG) at 1.2 Å resolution. Biochemistry, 26, 1152–1163.Google Scholar
First citation Warner, P., Green, R. C., Gomes, B. & Strimpler, A. M. (1994). Non-peptidic inhibitors of human leukocyte elastase. 1. The design and synthesis of pyridone-containing inhibitors. J. Med. Chem. 37, 3090–3099.Google Scholar
First citation Watanabe, K., Hata, Y., Kizaki, H., Katsube, Y. & Suzuki, Y. (1997). The refined crystal structure of Bacillus cereus oligo-1,6-glucosidase at 2.0 Å resolution: structural characterization of proline-substitution sites for protein thermostabilization. J. Mol. Biol. 269, 142–153.Google Scholar
First citation Weber, P. C. & Czarniecki, M. (1997). Structure-based design of thrombin inhibitors. In Structure-based drug design, edited by P. Veerapandian, pp. 247–264. New York: Marcel Dekker.Google Scholar
First citation Wei, A. Z., Mayr, I. & Bode, W. (1988). The refined 2.3-Å crystal structure of human leukocyte elastase in a complex with a valine chloromethyl ketone inhibitor. FEBS Lett. 234, 367–373.Google Scholar
First citation Weissenhorn, W., Carfi, A., Lee, K. H., Skehel, J. J. & Wiley, D. C. (1998). Crystal structure of the Ebola virus membrane fusion subunit, GP2, from the envelope glycoprotein ectodomain. Mol. Cell, 2, 605–616.Google Scholar
First citation Wery, J. P., Schevitz, R. W., Clawson, D. K., Bobbitt, J. L., Dow, E. R., Gamboa, G., Goodson, T. J., Hermann, R. B., Kramer, R. M., McClure, D. B., Mihelich, E. D., Putnam, J. E., Sharp, J. D., Stark, D. H., Teater, C., Warrick, M. W. & Jones, N. D. (1991). Structure of recombinant human rheumatoid arthritic synovial fluid phospholipase A2 at 2.2-Å resolution. Nature (London), 352, 79–82.Google Scholar
First citation Weston, S. A., Camble, R., Colls, J., Rosenbrock, G., Taylor, I., Egerton, M., Tucker, A. D., Tunnicliffe, A., Mistry, A., Macia, F., de La Fortelle, E., Irwin, J., Bricogne, G. & Pauptit, R. A. (1998). Crystal structure of the anti-fungal target N-myristoyl transferase. Nature Struct. Biol. 5, 213–221.Google Scholar
First citation Whitlow, M., Howard, A. J., Stewart, D., Hardman, K. D., Kuyper, L. F., Baccanari, D. P., Fling, M. E. & Tansik, R. L. (1997). X-ray crystallographic studies of Candida albicans dihydrofolate reductase. High resolution structure of the holoenzyme and an inhibited ternary complex. J. Biol. Chem. 272, 30289–30298.Google Scholar
First citation Whittingham, J. L., Edwards, D. J., Antson, A. A., Clarkson, J. M. & Dodson, G. G. (1998). Interactions of phenol and m-cresol in the insulin hexamer, and their effect on the association properties of B28 pro Asp insulin analogues. Biochemistry, 37, 11516–11523.Google Scholar
First citation Whittle, P. J. & Blundell, T. L. (1994). Protein structure-based drug design. Annu. Rev. Biophys. Biomol. Struct. 23, 349–375.Google Scholar
First citation Wierenga, R. K., Kalk, K. H. & Hol, W. G. J. (1987). Structure determination of the glycosomal triosephosphate isomerase from Trypanosoma brucei brucei at 2.4 Å resolution. J. Mol. Biol. 198, 109–121.Google Scholar
First citation Wild, K., Bohner, T., Aubry, A., Folkers, G. & Schulz, G. E. (1995). The three-dimensional structure of thymidine kinase from herpes simplex virus type 1. FEBS Lett. 368, 289–292.Google Scholar
First citation Williams, J. C., Zeelen, J. P., Neubauer, G., Vriend, G., Backmann, J., Michels, P. A., Lambeir, A. M. & Wierenga, R. K. (1999). Structural and mutagenesis studies of leishmania triosephosphate isomerase: a point mutation can convert a mesophilic enzyme into a superstable enzyme without losing catalytic power. Protein Eng. 12, 243–250.Google Scholar
First citation Williams, P. A., Cosme, J., Sridhar, V., Johnson, E. F. & McRee, D. E. (2000). Mammalian microsomal cytochrome P450 monooxygenase: structural adaptations for membrane binding and functional diversity. Mol. Cell, 5, 121–131.Google Scholar
First citation Williams, S. P. & Sigler, P. B. (1998). Atomic structure of progesterone complexed with its receptor. Nature (London), 393, 392–396.Google Scholar
First citation Wilson, D. K., Bohren, K. M., Gabbay, K. H. & Quiocho, F. A. (1992). An unlikely sugar substrate site in the 1.65 Å structure of the human aldose reductase holoenzyme implicated in diabetic complications. Science, 257, 81–84.Google Scholar
First citation Wilson, I. A., Skehel, J. J. & Wiley, D. C. (1981). Structure of the haemagglutinin membrane glycoprotein of influenza virus at 3 Å resolution. Nature (London), 289, 366–373.Google Scholar
First citation Wilson, K. P., Fitzgibbon, M. J., Caron, P. R., Griffith, J. P., Chen, W., McCaffrey, P. G., Chambers, S. P. & Su, M. S. (1996). Crystal structure of p38 mitogen-activated protein kinase. J. Biol. Chem. 271, 27696–27700.Google Scholar
First citation Wireko, F. C. & Abraham, D. J. (1991). X-ray diffraction study of the binding of the antisickling agent 12C79 to human hemoglobin. Proc. Natl Acad. Sci. USA, 88, 2209–2211.Google Scholar
First citation Wlodawer, A., Miller, M., Jaskolski, M., Sathyanarayana, B. K., Baldwin, E., Weber, I. T., Selk, L. M., Clawson, L., Schneider, J. & Kent, S. B. (1989). Conserved folding in retroviral proteases: crystal structure of a synthetic HIV-1 protease. Science, 245, 616–621.Google Scholar
First citation Wlodawer, A. & Vondrasek, J. (1998). Inhibitors of HIV-1 protease: a major success of structure-assisted drug design. Annu. Rev. Biophys. Biomol. Struct. 27, 249–284.Google Scholar
First citation Wolf, E., Vassilev, A., Makino, Y., Sali, A., Nakatani, Y. & Burley, S. K. (1998). Crystal structure of a GCN5-related N-acetyltransferase: Serratia marcescens aminoglycoside 3-N-acetyltransferase. Cell, 94, 439–449.Google Scholar
First citation Worthylake, D. K., Wang, H., Yoo, S., Sundquist, W. I. & Hill, C. P. (1999). Structures of the HIV-1 capsid protein dimerization domain at 2.6 Å resolution. Acta Cryst. D55, 85–92.Google Scholar
First citation Wynne, S. A., Crowther, R. A. & Leslie, A. G. (1999). The crystal structure of the human hepatitis B virus capsid. Mol. Cell, 3, 771–780.Google Scholar
First citation Xia, D., Henry, L. J., Gerard, R. D. & Deisenhofer, J. (1994). Crystal structure of the receptor-binding domain of adenovirus type 5 fiber protein at 1.7 Å resolution. Structure, 2, 1259–1270.Google Scholar
First citation Xie, X., Gu, Y., Fox, T., Coll, J. T., Fleming, M. A., Markland, W., Caron, P. R., Wilson, K. P. & Su, M. S. (1998). Crystal structure of JNK3: a kinase implicated in neuronal apoptosis. Structure, 6, 983–991.Google Scholar
First citation Xu, W., Harrison, S. C. & Eck, M. J. (1997). Three-dimensional structure of the tyrosine kinase c-Src. Nature (London), 385, 595–602.Google Scholar
First citation Xue, Y., Bjorquist, P., Inghardt, T., Linschoten, M., Musil, D., Sjolin, L. & Deinum, J. (1998). Interfering with the inhibitory mechanism of serpins: crystal structure of a complex formed between cleaved plasminogen activator inhibitor type 1 and a reactive-centre loop peptide. Structure, 6, 627–636.Google Scholar
First citation Yan, Y., Li, Y., Munshi, S., Sardana, V., Cole, J. L., Sardana, M., Steinkuehler, C., Tomei, L., De Francesco, R., Kuo, L. C. & Chen, Z. (1998). Complex of NS3 protease and NS4A peptide of BK strain hepatitis C virus: a 2.2 Å resolution structure in a hexagonal crystal form. Protein Sci. 7, 837–847.Google Scholar
First citation Yang, J., Kloek, A. P., Goldberg, D. E. & Mathews, F. S. (1995). The structure of Ascaris hemoglobin domain I at 2.2 Å resolution: molecular features of oxygen avidity. Proc. Natl Acad. Sci. USA, 92, 4224–4228.Google Scholar
First citation Yang, X. & Moffat, K. (1995). Insights into specificity of cleavage and mechanism of cell entry from the crystal structure of the highly specific Aspergillus ribotoxin, restrictocin. Structure, 4, 837–852.Google Scholar
First citation Yao, N., Hesson, T., Cable, M., Hong, Z., Kwong, A. D., Le, H. V. & Weber, P. C. (1997). Structure of the hepatitis C virus RNA helicase domain. Nature Struct. Biol. 4, 463–467.Google Scholar
First citation Yee, V. C., Pedersen, L. C., LeTrong, I., Bishop, P. D., Stenkamp, R. E. & Teller, D. C. (1994). Three-dimensional structure of a transglutaminase: human blood coagulation factor XIII. Proc. Natl Acad. Sci. USA, 91, 7296–7300.Google Scholar
First citation Yeh, J. I., Claiborne, A. & Hol, W. G. J. (1996). Structure of the native cystein-sulfenic acid redox center of enterococcal NADH peroxidase refined at 2.8 Å resolution. Biochemistry, 35, 9951–9957.Google Scholar
First citation Yoshimoto, T., Kabashima, T., Uchikawa, K., Inoue, T., Tanaka, N., Nakamura, K. T., Tsuru, M. & Ito, K. (1999). Crystal structure of prolyl aminopeptidase from Serratia marcescens. J. Biochem. (Tokyo), 126, 559–565.Google Scholar
First citation Zaitseva, I., Zaitsev, V., Card, G., Moshkov, K., Bax, B., Ralph, A. & Lindley, P. (1996). The X-ray structure of human serum ceruloplasmin at 3.1 Å: nature of the copper centres. J. Biol. Inorg. Chem. 1, 15–23.Google Scholar
First citation Zdanov, A., Schalk-Hihi, C., Menon, S., Moore, K. W. & Wlodawer, A. (1997). Crystal structure of Epstein–Barr virus protein BCRF1, a homolog of cellular interleukin-10. J. Mol. Biol. 268, 460–467.Google Scholar
First citation Zhang, A., Geisler, S. C., Smith, A. D., Resnick, D. A., Li, M. L., Wang, C. Y., Looney, D. J., Wong-Staal, F., Arnold, E. & Arnold, G. F. (1999). A disulfide-bound HIV-1 V3 loop sequence on the surface of human rhinovirus 14 induces neutralizing responses against HIV-1. J. Biol. Chem. 380, 365–374.Google Scholar
First citation Zhang, H., Gao, Y. G., van der Marel, G. A., van Boom, J. H. & Wang, A. H. (1993). Simultaneous incorporations of two anticancer drugs into DNA. The structures of formaldehyde-cross-linked adducts of daunorubicin-d(CG(araC)GCG) and doxorubicin-d(CA(araC)GTG) complexes at high resolution. J. Biol. Chem. 268, 10095–10101.Google Scholar
First citation Zhang, R., Evans, G., Rotella, F. J., Westbrook, E. M., Beno, D., Huberman, E., Joachimiak, A. & Collart, F. R. (1999). Characteristics and crystal structure of bacterial inosine-5′-monophosphate dehydrogenase. Biochemistry, 38, 4691–4700.Google Scholar
First citation Zhang, R. G., Scott, D. L., Westbrook, M. L., Nance, S., Spangler, B. D., Shipley, G. G. & Westbrook, E. M. (1995). The three-dimensional crystal structure of cholera toxin. J. Mol. Biol. 251, 563–573.Google Scholar
First citation Zhang, X., Morera, S., Bates, P. A., Whitehead, P. C., Coffer, A. I., Hainbucher, K., Nash, R. A., Sternberg, M. J., Lindahl, T. & Freemont, P. S. (1998). Structure of an XRCC1 BRCT domain: a new protein–protein interaction module. EMBO J. 17, 6404–6411.Google Scholar
First citation Zhu, X., Kim, J. L., Newcomb, J. R., Rose, P. E., Stover, D. R., Toledo, L. M., Zhao, H. & Morgenstern, K. A. (1999). Structural analysis of the lymphocyte-specific kinase Lck in complex with non-selective and Src family selective kinase inhibitors. Struct. Fold. Des. 7, 651–661.Google Scholar
First citation Zuccola, H. J., Rozzelle, J. E., Lemon, S. M., Erickson, B. W. & Hogle, J. M. (1998). Structural basis of the oligomerization of hepatitis delta antigen. Structure, 6, 821–830.Google Scholar