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. 4.2, p. 99   | 1 | 2 |

Section 4.2.7. General recommendations

H. Michela*

aMax-Planck-Institut für Biophysik, Heinrich-Hoffmann-Strasse 7, D-60528 Frankfurt/Main, Germany
Correspondence e-mail: michel@mpibp-frankfurt.mpg.de

4.2.7. General recommendations

| top | pdf |

When trying to crystallize a membrane protein the first, and most important, task is to obtain a pure, stable and homogeneous preparation of the membrane protein under investigation. Unfortunately, general methods for producing substantial amounts of membrane proteins using heterologous expression systems in a native membrane environment do not exist, nor is refolding of membrane proteins from inclusion bodies well established. Once a pure and homogeneous preparation of a membrane protein has been obtained, its stability in various detergents has to be investigated at different pH values. Frequently, a sharp optimum pH for stability is observed with detergents of shorter alkyl-chain length. In crystallization attempts, the usual parameters (nature of the precipitant, buffer, pH, addition of inhibitors and/or substrates, see Chapter 4.1[link] ) should be varied. The most important variable, however, is the detergent. Detergents can be exchanged most conveniently by binding the membrane protein to column materials (e.g. ion-exchange resins, hydroxyapatite, affinity matrices), washing with a buffer containing the new detergent at concentrations above the CMC under non-eluting conditions, and then eluting the bound membrane protein with a buffer containing the new detergent. The completeness of the detergent exchange should be checked. For this purpose, the use of 14C- or 3H-labelled detergents is recommended. Washing with about 20 column volumes is frequently required for a complete detergent exchange. It is more difficult to exchange a detergent with a low CMC (this means long alkyl chains) against a detergent with a high CMC (shorter alkyl chains). The detergent can often be exchanged during a final step in the purification. Less satisfying methods for detergent exchange are molecular-sieve chromatography, ultracentrifugation or repeated ultrafiltration and dilution. If the amount of membrane protein is limited, it is advisable to restrict the usual parameters to a set of, e.g., 48 standard combinations with respect to precipitating agent, pH, buffer etc., but to try all available detergents. If crystals of insufficient quality or size are obtained, trying the antibody Fv fragment approach is highly recommended. Alternatives are to use a different source (species) for the membrane protein under investigation, or to remove flexible parts of the membrane protein by proteolytic digestion. I am convinced that 50% of all membrane proteins will yield well ordered, three-dimensional crystals within five years, once the problem of obtaining a pure, stable and homogeneous preparation has been solved.

References

First citation Allen, J. P., Feher, G., Yeates, T. O., Komiya, H. & Rees, D. C. (1987). Structure of the reaction center from Rhodobacter sphaeroides R-26: the protein subunits. Proc. Natl Acad. Sci. USA, 84, 6162–6166.Google Scholar
First citation Buchanan, S. K., Smith, B. S., Venkatrami, L., Xia, D., Esser, L., Palnitkar, M., Chakraborty, R., van der Helm, D. & Deisenhofer, J. (1999). Crystal structure of the outer membrane active transporter FepA from Escherichia coli. Nature Struct. Biol. 6, 56–63.Google Scholar
First citation Chang, C. H., El-Kabbani, D., Tiede, D., Norris, J. & Schiffer, M. (1991). Structure of the membrane-bound protein photosynthetic reaction center from Rhodobacter sphaeroides. Biochemistry, 30, 5352–5360.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 Cowan, S. W., Garavito, R. M., Jansonius, J. N., Jenkins, J. A., Karlsson, R., König, N., Pai, E. F., Pauptit, R. A., Rizkallah, P. J., Rosenbusch, J. P., Rummel, G. & Schirmer, T. (1995). The structure of OmpF porin in a tetragonal crystal form. Structure, 3, 1041–1050.Google Scholar
First citation Cowan, S. W., Schirmer, T., Rummel, G., Steiert, M., Gosh, R., Pauptit, R. A., Jansonius, J. N. & Rosenbusch, J. P. (1992). Crystal structures explain functional properties of two E. coli porins. Nature (London), 358, 727–733.Google Scholar
First citation Deisenhofer, J., Epp, O., Miki, K., Huber, R. & Michel, H. (1985). Structure of the protein subunits in the photosynthetic reaction center of Rhodopseudomonas viridis at 3 Å. Nature (London), 318, 618–642.Google Scholar
First citation Deisenhofer, J., Epp, O., Sinning, I. & Michel, H. (1995). Crystallographic refinement at 2.3 Å resolution and refined model of the photosynthetic reaction centre from Rhodopseudomonas viridis. J. Mol. Biol. 246, 429–457.Google Scholar
First citation Doyle, D. A., Cabral, J. M., Pfuetzner, R. A., Kuo, A. L., Gulbis, J. M., Cohen, S. L., Chait, B. T. & MacKinnon, R. (1998). The structure of the potassium channel: molecular basis of [K^{+}] conduction and selectivity. Science, 280, 69–77.Google Scholar
First citation Ermler, U., Fritzsch, G., Buchanan, S. K. & Michel, H. (1994). Structure of the photosynthetic reaction centre from Rhodobacter sphaeroides at 2.65 Å resolution: cofactors and protein–cofactor interactions. Structure, 2, 925–936.Google Scholar
First citation Essen, L. O., Siegert, R., Lehmann, W. D. & Oesterhelt, D. (1998). Lipid patches in membrane protein oligomers – crystal structure of the bacteriorhodpsin–lipid complex. Proc. Natl Acad. Sci. USA, 95, 11673–11678.Google Scholar
First citation Ferguson, A. D., Hofmann, E., Coulton, J. W., Diederichs, K. & Welte, W. (1998). Siderophore mediated iron transport: crystal structure of FhuA with bound lipopolysaccharide. Science, 282, 2215–2220.Google Scholar
First citation Forst, D., Welte, W., Wacker, T. & Diederichs, K. (1998). Structure of the sucrose-specific porin ScrY from Salmonella typhimurium and its complex with sucrose. Nature Struct. Biol. 5, 37–46.Google Scholar
First citation Grigorieff, N., Ceska, T. A., Downing, K. H., Baldwin, J. M. & Henderson, R. (1996). Electron crystallographic refinement of the structure of bacteriorhodopsin. J. Mol. Biol. 259, 393–421.Google Scholar
First citation Henderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, E. & Downing, K. H. (1990). Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopy. J. Mol. Biol. 213, 899–929.Google Scholar
First citation Hirsch, A., Breed, J., Saxena, K., Richter, O. M. H., Ludwig, B., Diederichs, K. & Welte, W. (1997). The structure of porin from Paracoccus denitrificans at 3.1 Å resolution. FEBS Lett. 404, 208–210.Google Scholar
First citation Iwata, S., Lee, J. W., Okada, K., Lee, J. K., Iwata, M., Rasmussen, B., Link, T. A., Ramaswamy, S. & Jap, B. K. (1998). Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex. Science, 281, 64–71.Google Scholar
First citation Iwata, S., Ostermeier, C., Ludwig, B. & Michel, H. (1995). Structure at 2.8 Å resolution of cytochrome c oxidase from Paracoccus denitrificans. Nature (London), 376, 660–669.Google Scholar
First citation Kim, H., Xia, D., Yu, C. A., Xia, J. Z., Kachurin, A. M., Li, Z., Yu, L. & Deisenhofer, J. (1998). Inhibitor binding changes domain mobility in the iron–sulfur protein of the mitochondrial bc1 complex from bovine heart. Proc. Natl Acad. Sci. USA, 95, 8026–8033.Google Scholar
First citation Kimura, Y., Vassylyev, D. G., Miyazawa, A., Kidera, A., Matsushima, M., Mitsuoka, K., Murata, K., Hirai, T. & Fujiyoshi, Y. (1997). Surface of bacteriorhodopsin revealed by high-resolution electron crystallography. Nature (London), 389, 206–211.Google Scholar
First citation Koepke, J., Hu, X., Muenke, C., Schulten, K. & Michel, H. (1996). The crystal structure of the light-harvesting complex II (B800–850) from Rhodospirilum molischianum. Structure, 4, 581–597.Google Scholar
First citation Kreusch, A., Neubüser, A., Schiltz, E., Weckesser, J. & Schulz, G. E. (1994). Structure of the membrane channel porin from Rhodopseudomonas blastica at 2.0 Å resolution. Protein Sci. 3, 58–63.Google Scholar
First citation Kurumbail, R. G., Stevens, A. M., Gierse, J. K., McDonald, J. J., Stegeman, R. A., Pak, J. Y., Gildehaus, D., Miyashiro, J. M., Penning, T. D., Seibert, K., Isakson, P. C. & Stallings, W. C. (1996). Structural basis for selective inhibition of cyclooxygenase-2 by anti-inflammatory agents. Nature (London), 384, 644–648.Google Scholar
First citation Kühlbrandt, W., Wang, D. N. & Fujiyoshi, Y. (1994). Atomic model of plant light-harvesting complex by electron crystallography. Nature (London), 367, 614–621.Google Scholar
First citation Lancaster, C. R. D. & Michel, H. (1997). The coupling of light-induced electron transfer and proton uptake as derived from crystal structures of reaction centres from Rhodopseudomonas viridis modified at the binding site of the secondary quinone, QB. Structure, 5, 1339–1359.Google Scholar
First citation Lancaster, C. R. D. & Michel, H. (1999). Refined crystal structures of reaction centres from Rhodopseudomonas viridis in complexes with the herbicide atrazine and two chiral atrazine derivatives also lead to a new model of the bound carotenoid. J. Mol. Biol. 286, 883–898.Google Scholar
First citation Locher, K. P., Rees, B., Koebnik, R., Mitschler, A., Moulinier, L., Rosenbusch, J. P. & Moras, D. (1998). Transmembrane signaling across the ligand-gated FhuA receptor: crystal structures of free and ferrichrome-bound states reveal allosteric changes. Cell, 98, 771–778.Google Scholar
First citation Luecke, H., Richter, H. T. & Lamyi, J. K. (1998). Proton transfer pathways in bacteriorhodopsin at 2.3 Å resolution. Science, 280, 1934–1937.Google Scholar
First citation Luong, C., Miller, A., Barnett, J., Chow, J., Ramesha, C. & Browner, M. F. (1996). Flexibility of the NSAID binding site in the structure of human cyclooxygenase-2. Nature Struct. Biol. 3, 927–933.Google Scholar
First citation McDermott, G., Prince, S. M., Freer, A. A., Hawthornthwaite-Lawless, A. M., Papiz, M. Z., Cogdell, R. J. & Isaacs, N. W. (1995). Crystal-structure of an integral membrane light-harvesting complex from photosynthetic bacteria. Nature (London), 374, 517–521.Google Scholar
First citation Meyer, J. E. W., Hofnung, M. & Schulz, G. E. (1997). Structure of maltoporin from Salmonella typhimurium ligated with a nitrophenyl-maltotrioside. J. Mol. Biol. 266, 761–775.Google Scholar
First citation Ostermeier, C., Harrenga, A., Ermler, U. & Michel, H. (1997). Structure at 2.7 Å resolution of the Paracoccus denitrificans two-subunit cytochrome c oxidase complexed with an antibody Fv fragment. Proc. Natl Acad. Sci. USA, 94, 10547–10553.Google Scholar
First citation Pautsch, A. & Schulz, G. E. (1998). Structure of the outer membrane protein: a transmembrane domain. Nature Struct. Biol. 5, 1013–1017.Google Scholar
First citation Pebay-Peyroula, E., Rummel, G., Rosenbusch, J. P. & Landau, E. M. (1997). X-ray structure of bacteriorhodopsin at 2.5 Å from microcrystals grown in lipidic cubic phases. Science, 277, 1676–1681.Google Scholar
First citation Picot, D., Loll, P. J. & Garavito, M. (1994). The X-ray crystal structure of the membrane protein prostaglandin H2 synthase-1. Nature (London), 367, 243–249.Google Scholar
First citation Schirmer, T., Keller, T. A., Wang, Y. F. & Rosenbusch, J. P. (1995). Structural basis for sugar translocation through maltoporin channels at 3.1 Å resolution. Science, 267, 512–514.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 Stowell, M. H. B., McPhillips, T. M., Rees, D. C., Soltis, S. M., Abresch, E. & Feher, G. (1997). Light-induced structural changes in photosynthetic reaction center: implications for mechanism of electron–proton transfer. Science, 276, 812–816.Google Scholar
First citation Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T., Yamaguchi, H., Shinzawa-Itoh, K., Nakashima, R., Yaono, R. & Yoshikawa, S. (1995). Structures of metal sites of oxidized bovine heart cytochrome c oxidase at 2.8 Å. Science, 269, 1069–1074.Google Scholar
First citation Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T., Yamaguchi, H., Shinzawa-Itoh, K., Nakashima, R., Yaono, R. & Yoshikawa, S. (1996). The whole structure of the 13-subunit oxidized cytochrome c oxidase at 2.8 Å. Science, 272, 1136–1144.Google Scholar
First citation Weiss, M. S., Abele, U., Weckesser, J., Welte, W., Schiltz, E. & Schulz, G. E. (1991). Molecular architecture and electrostatic properties of a bacterial porin. Science, 254, 1627–1630.Google Scholar
First citation Wendt, K. U., Poralla, K. & Schulz, G. E. (1997). Structure and function of a cyclase. Science, 277, 1811–1815.Google Scholar
First citation Xia, D., Yu, C. A., Kim, H., Xia, J. Z., Kachurin, A. M., Zhang, L., Yu, L. & Deisenhofer, J. (1997). Crystal structure of the cytochrome bc1 complex from bovine heart mitochondria. Science, 277, 60–66.Google Scholar
First citation Yoshikawa, S., Shinzawa-Itoh, K., Nakashima, R., Yaono, R., Yamashita, E., Inoue, N., Yao, M., Fei, M. J., Libeu, C. P., Mizushima, T., Yamaguchi, H., Tomizaki, T. & Tsukihara, T. (1998). Redox-coupled crystal structural changes in bovine heart cytochrome c oxidase. Science, 280, 1723–1729.Google Scholar
First citation Zhang, Z. L., Huang, L. S., Shulmeister, V. M., Chi, Y. I., Kim, K. K., Hung, L. W., Crofts, A. R., Berry, E. A. & Kim, S. H. (1998). Electron transfer by domain movement in cytochrome bc1. Nature (London), 392, 677–684.Google Scholar








































to end of page
to top of page