Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-28T02:43:46.263Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 November 2012

Igor N. Serdyuk
Affiliation:
Institute of Protein Research, Moscow
Nathan R. Zaccai
Affiliation:
University of Bristol
Joseph Zaccai
Affiliation:
Institut de Biologie Structurale, Grenoble
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Methods in Molecular Biophysics
Structure, Dynamics, Function
, pp. 1076 - 1102
Publisher: Cambridge University Press
Print publication year: 2007

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Agalarov, S. C., and Williamson, J. R. (2000). A hierarchy of RNA subdomains in assembly of the central domain of the 30S ribosomal subunit. RNA, 6, 402–408.CrossRefGoogle Scholar
Agalarov, S. C., Selivanova, O. M., et al. (1999). Independent in vitro assembly of all three major morphological parts of the 30S ribosomal subunit of Thermus thermophilus. Eur. J. Biochem., 266, 533–537.CrossRefGoogle ScholarPubMed
Agalarov, S. C., Sheleznyakova, E. N., et al. (1998). In vitro assembly of a ribonucleoprotein particle corresponding to the platform domain of the 30S ribosomal subunit. Proc. Natl. Acad. Sci. USA, 95, 999–1003.CrossRefGoogle ScholarPubMed
Allemand, J.-F., Bensimon, D., Lavery, R., and Croquette, V. (1998). Stretched and overwound DNA forms a Pouling-like structure with exposed base. Proc. Natl. Acad. Sci. USA, 95, 14152–14157.CrossRefGoogle Scholar
Allison, S. A. (1999). Low Reynolds number transport properties of axisymmetric particles employing stick and slip boundary conditions. Macromolecules, 32, 5304–5312.CrossRefGoogle Scholar
Allison, S. A. (2001). Boundary element modelling of biomolecular transport. Biophys. Chem., 93, 197–213.CrossRefGoogle Scholar
Altieri, A. S., Hinton, D., and Byrd, R. A. (1995). Association of biomolecular system via pulse field gradient NMR self-diffusion measurements. JACS, 117, 7566–7567.CrossRefGoogle Scholar
Altose, M. D., Zheng, Y., Dong, J., Palfey, B. A., and Carey, P. R. (2001). Comparing protein–ligand interactions in solution and single crystals by Raman spectroscopy. PNAS, 98, 3006–3011.CrossRefGoogle ScholarPubMed
Bacia, K., and Schwille, P. (2003). A dynamic view of cellular processes by in vivo fluorescence auto- and cross-correlation spectroscopy. Methods, 29, 74–85.CrossRefGoogle ScholarPubMed
Bailey, B., Farkas, D. L., Taylor, D. L., and Lanni, F. (1993). Enhancement of axial resolution in fluorescence microscopy by standing-wave excitation. Nature, 366, 44–48.CrossRefGoogle ScholarPubMed
Ban, N., Nissen, P., et al. (1999). Placement of protein and RNA structures into a 5 å-resolution map of the 50S ribosomal subunit. Nature, 400(6747), 841–847.CrossRefGoogle ScholarPubMed
Banachowicz, E., Gapinski, J., and Patkowski, A. (2000). Solution structure of biopolymers: A new method of constructing a bead model. Biophys. J., 78, 70–78.CrossRefGoogle ScholarPubMed
Bandecar, J. (1992). Amide modes and protein conformation. Biochim. Biophys. Acta, 1120, 123–143.CrossRefGoogle Scholar
Barnes, W. L., Dereux, A., and Ebbesen, T. W. (2003). Surface plasmon subwavelength optics. Nature, 424(6950), 824–830.CrossRefGoogle ScholarPubMed
Barron, L. D., Hecht, L., Blanch, E. W., and Bell, A. F. (2000). Solution structure and dynamics of biomolecules from Raman optical activity. Prog. Biophys. and Mol. Biol., 73, 1–49.CrossRefGoogle ScholarPubMed
Basavappa, R., and Sigler, P. B. (1991). EMBO J., 10, 3105–3111.
Bastiaens, P. I. H., and Pepperkok, R. (2000). Observing proteins in their natural habitat: the living cell. TIBS, 25, 631–636.Google ScholarPubMed
Baumann, C. G., Bloomfield, V. A., Smith, S. B., Bustamante, C., Wang, M. D., and Block, S. M. (2000). Stretching of single collapsed DNA molecules. Biophys. J., 78, 1965–1978.CrossRefGoogle ScholarPubMed
Bax, A. (2003). Weak alignment offers new NMR opportunities to study protein structure and dynamics. Protein Sci., 12, 1–16.CrossRefGoogle ScholarPubMed
Belke, J., and Ristau, O. (1997). Analysis of interacting biopolymer systems by analytical centrifugation. Eur. Biophys. J., 25, 325–332.CrossRefGoogle Scholar
Bellissent-Funel, M. C., Zanotti, J. M., et al. (1996). Slow dynamics of water molecules on the surface of globular proteins. Faraday Discuss., 103, 281–294.CrossRefGoogle Scholar
Belov, M. E., Gorshkov, M. V., Udeseth, H. R., Anderson, G. A., and Smith, R. D. (2000). Zeptomole-sensititivity electrospray ionization – Fourier transform ion cyclotron resonance mass spectrometry proteins. Anal. Chem., 72, 2271–2279.CrossRefGoogle ScholarPubMed
Benner, W. H. (1997). A gated electrostatic ion trap to repetitiously measure the charge and m/z of large electrospray ions. Anal. Chem., 69, 4162–4168.CrossRefGoogle Scholar
Bennett, M. J., and Eisenberg, D. (1994). Refined structure of monomeric diphtheria toxin at 2.3 å resolution. Protein Sci., 3(9), 1464–1475.CrossRefGoogle ScholarPubMed
Bennink, M. L., Leuba, S. H., Leno, G. H., Zlatanova, J., Grooth, B. G., and Greve, J. (2001). Unfolding individual nucleosomes by stretching single chromatin fibers with optical tweezers. Nature Str. Biol, 8, 606–610.CrossRefGoogle ScholarPubMed
Berg, H. (1983). Random Walks in Biology. Princeton: Princeton University Press.Google Scholar
Bergethon, P. R. (1995). The Physical Basis of Biochemistry. The Foundation of Molecular Biophysics. New York: Springer.Google Scholar
Bernado, P., Garcia de la Torre, J., and Pons, M. (2002). Interpretation of 15N NMR relaxation data for globular proteins using hydrodynamic calculations with HYDRONMR. J. Biomol. NMR, 23, 139–150.CrossRefGoogle ScholarPubMed
Bernal, J. D., and Crowfoot, D. (1934). X-ray photographs of crystalline pepsin. Nature, 134, 794–795.CrossRefGoogle Scholar
Biemann, K. (1992). Mass spectrometry of peptides and proteins. Annu. Rev. Biochem., 61, 977–1010.CrossRefGoogle ScholarPubMed
Bischler, N., Brino, L., et al. (2002). Localization of the yeast RNA polymerase I-specific subunits. Embo. J., 21(15), 4136–4144.CrossRefGoogle ScholarPubMed
Bjorkman, P. J., Saper, M. A., et al. (1987). Structure of the human class I histocompatibility antigen, HLA-A2. Nature, 329(6139), 506–512.CrossRefGoogle ScholarPubMed
Blattner, F. R. (1997). The complete genome sequence of Eshcherichia coli K-12. Science, 277, 1453–1474.CrossRefGoogle Scholar
Block, S. M., Blair, D. F., and Berg, H. C. (1989). Compliance of bacterial flagella measured wih optical tweezers. Nature, 338, 514–518.CrossRefGoogle Scholar
Bon, C., Dianoux, A. J., et al. (2002). A model for water motion in crystals of lysozyme based on an incoherent quasielastic neutron-scattering study. Biophys. J., 83(3), 1578–1588.CrossRefGoogle ScholarPubMed
Bon, C., Lehmann, M. S., et al. (1990). Quasi Laue neutron-diffraction study of the water arrangement in crystals of triclinic hen egg-white lysozyme. Acta Crystallogr. D, 55, 978–987.CrossRefGoogle Scholar
Booth, D. R., Sunde, M., et al. (1997). Instability, unfolding and aggregation of human lysozyme variants underlying amyloid fibrillogenesis. Nature, 385, 787–793.CrossRefGoogle ScholarPubMed
Bottcher, B., Tsuji, N., et al. (1998). Peptides that block hepatitis B virus assembly: analysis by cryomicroscopy, mutagenesis and transfection. Embo. J., 17(23), 6839–6845.CrossRefGoogle ScholarPubMed
Bowie, J. U., Luthy, R., and Eisenberg, D. (1991). A method to identify protein sequences that fold into a known three-dimensional structure. Science, 253(5016), 164–170.CrossRefGoogle ScholarPubMed
Braiman, M. S., and Rothschild, K. J. (1988). Fourier transform infrared techniques for probing membrane protein structure. Ann. Rev. Biophys. Biophysical Chem., 17, 541–570.CrossRefGoogle ScholarPubMed
Brändén, C.-I., and Tooze, J. (1999). Introduction to Protein Structure. New York: Garland Pub.Google Scholar
Brant, D. A. (1999). Novel approaches to the analysis of polysaccharide structure. Curr. Opin. Struct. Biol., 9, 556–562.CrossRefGoogle Scholar
Brey, W. S. (ed.) (1988). Pulse Methods in 1D and 2D Liquid-Phase NMR. San-Diego: Academic.Google Scholar
Brooks, B. R., Bruccoleri, R. E., et al. (1983). CHARMM: A program for macromolecular empirical energy modelling. J. Comp. Chem., 4, 187–230.CrossRefGoogle Scholar
Brower-Toland, B. R., Smith, C. L., Yeh, R. S., Lis, J. T., Peterson, C. L., and Wang, M. D. (2002). Mechanical disruption of individual nucleosomes reveals a reversible multistage release of DNA. Proc. Natl. Acad. Sci. USA, 99, 1960–1966.CrossRefGoogle ScholarPubMed
Brudler, R., Rammelsberg, R., et al. (2001). Structure of the I1 early intermediate of photoactive yellow protein by FTIR spectroscopy. Nat. Struct. Biol., 8(3), 265–270.CrossRefGoogle ScholarPubMed
Brune, D., and Kim, S. (1993). Predicting protein diffusion coefficients. J. Am. Chem. Soc., 90, 3835–3839.Google ScholarPubMed
Brunger, A. T., and Adams, P. D. (2002). Molecular dynamics applied to X-ray structure refinement. Acc. Chem. Res., 35(6), 404–412.CrossRefGoogle ScholarPubMed
Brunger, A. T., Adams, P. D., et al. (1998). Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallogr. D Biol. Crystallogr., 54 (Pt 5), 905–921.CrossRefGoogle ScholarPubMed
Buchanan, M. V., and Hettich, R. L. (1993). Fourier transform mass spectrometry of high mass molecules. Anal. Chem., 65, 245A–259A.CrossRefGoogle Scholar
Burlingame, A. L., and Carr, S. A. (1996). Mass Spectrometry in the Biological Sciences. Totowa: Humana Press.CrossRefGoogle Scholar
Burlingame, A. L., Carr, S. A., et al. (2000). Mass Spectrometry in Biology and Medicine. Totowa: Humana Press.CrossRefGoogle Scholar
Burton, A., and Sinsheimer, R. L. (1965). The process of infection with bacteriophage X174. VII. Ultracentrifugal analysis of the replicative form. J. Mol. Biol., 14, 327–347.CrossRefGoogle ScholarPubMed
Bustamante, C., Erie, D. A., and Keller, D. (1994). Biochemical and structural applications of scanning force microscopy. Curr. Opin. Struct. Biol., 4, 750–760.CrossRefGoogle Scholar
Byron, O. (1997). Construction of hydrodynamic bead models from high resolution x-ray crystallographic or nuclear magnetic resonance data. Bioph. J., 72, 406–415.CrossRefGoogle ScholarPubMed
Caffrey, M. (2000). A lipid's eye view of membrane protein crystallization in mesophases. Curr. Opin. Struct. Biol., 10(4), 486–497.CrossRefGoogle ScholarPubMed
Cai, S., and Singh, B. R. (1999). Identification of beta-turn and random coil amide III infrared bands for secondary structure estimation of proteins. Biophys. Chem., 80(1), 7–20.CrossRefGoogle ScholarPubMed
Callender, R., and Deng, H. (1994). Nonresonance Raman difference spectroscopy: a general probe of protein structure, ligand binding, enzymatic catalysis, and the structures of other biomacromolecules. Annu. Rev. Biophys. Biomol. Struct., 23, 215–245.CrossRefGoogle ScholarPubMed
Callis, P. R., and Davidson, N. (1969). Hydrodynamic relaxation times of DNA from decay of flow dichroism measurements. Biopolymers, 8, 379–390.CrossRefGoogle Scholar
Campos-Olivas, R., Horr, I., Bormann, C., Jung, G., and Gronenborn, A. M. (2001). Solution structure, backbone dynamics and chitin binding of the anti-fungal protein from Streptomyces tendae TU901. J. Mol. Biol., 308, 765–782.CrossRefGoogle ScholarPubMed
Canet, D., Doering, K., Dobson, C. M., and Dupont, Y. (2001). High-sensitivity fluorescence anisotropy detection of protein-folding events: application to α-lactalbumin. Biophys. J., 80, 1996–2003.CrossRefGoogle ScholarPubMed
Cantor, C., and Schimmel, P. (1980). Biophysical Chemistry. Part II. Technique for the Study of Biological Structure and Function. San Francisco: W. H. Freeman and Company.Google Scholar
Caprioli, R. M., and Suter, M. J. F. (1995). Mass spectrometry. In Introduction to Biophysical Methods for Protein and Nucleic Acid Research. San Diego: Academic Press.Google Scholar
Carr, S. A., and Burlingame, A. L. (1996). The meaning and usage of the terms monoisotopic mass, average mass, mass resolution, and mass accuracy for measurements of biomolecules. In Mass Spectrometry in the Biological Sciences, eds. Burlingame, A. L. and Carr, S. A., pp. 546–552. Totowa: Humana Press.Google Scholar
Carra, J. H., Murphy, E. C., et al. (1996). Thermodynamic effects of mutations on the denaturation of T4 lysozyme. Biophys. J., 71(4), 1994–2001.CrossRefGoogle ScholarPubMed
Carrasco, B., and Garcia de la Torre, J. (1999). Hydrodynamic properties of rigid particles: comparison of different modelling and computational procedure. Biophys. J., 75, 3044–30572.CrossRefGoogle Scholar
Carrion-Vazquez, M., Overhauser, A. F., et al. (2000). Mechanical design of proteins studied by single-molecule force spectroscopy and protein engineering. Prog. Biophys. Mol. Biol., 74, 63–91.CrossRefGoogle ScholarPubMed
Castro, A., Fairfield, F. R., and Shera, E. B. (1993). Fluorescence detection and size measurement of single DNA molecules. Anal. Chem., 65, 849–852.CrossRefGoogle Scholar
Cate, J. H., Yusupov, M. M., et al. (1999). X-ray crystal structures of 70S ribosome functional complexes. Science, 285(5436), 2095–2104.CrossRefGoogle ScholarPubMed
Chacon, P., Moran, F., et al. (1998). Low-resolution structures of proteins in solution retrieved from X-ray scattering with a genetic algorithm. Biophys. J., 74(6), 2760–75.CrossRefGoogle ScholarPubMed
Chacon, P., Diaz, J. F., et al. (2000). Reconstruction of protein form with X-ray solution scattering and a genetic algorithm. J. Mol. Biol., 299(5), 1289–1302.CrossRefGoogle Scholar
Charney, E., Chen, H.-H., and Rau, D. (1991). The flexibility of A-form DNA. J. Biolmol. Struct. Dyn., 9, 353–362.CrossRefGoogle ScholarPubMed
Che, Z., N. Olson, H., et al. (1998). Antibody-mediated neutralization of human rhinovirus 14 explored by means of cryoelectron microscopy and X-ray crystallography of virus–Fab complexes. J. Virol., 72(6), 4610–4622.Google ScholarPubMed
Checovich, W. J., Bolger, R. E., and Burke, T. (1995). Fluorescence polarization – a new tool for cell and molecular biology. Nature, 375, 254–256.CrossRefGoogle ScholarPubMed
Chen, X., Wu, H., Mao, C., and Whitesides, G. M. (2002). A prototype two-dimensional capillary electrophoresis system fabricated in poly(dimethylsiloxane). Anal. Chem., 74, 1772–1778.CrossRefGoogle Scholar
Cherry, R. J., and Schneider, G. (1976). A spectroscopic technique for measuring slow rotational diffusion of macromolecules. 2: Determination of rotational correlation times of protein in solution. Biochemistry, 15, 3657–3661.CrossRefGoogle Scholar
Chervenka, , 1969. A Manual of Methods for the Analytical Ultracentrifuge. Palo Alto: Spinco Division, Beckman Instruments.Google Scholar
Chong, B. E., Lubman, D. M., et al. (1999). Rapid screening of protein profiles of human breast cancer cell lines using non-porous reversed-phase high performance liquid chromatography separation with matrix-assisted laser desorption/ionization time-of-flight mass spectral analysis. Rapid. Commun. Mass Spectr., 13(18), 1808–1812.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Chong, B. E., Lubman, D. M., Rosenspire, A., and Miller, F. (1998). Protein profiles and identification of high performance liquid chromatography isolated proteins of cancer cell lines using matrix-asisted laser desorption/ionization time-of-flight mass spectrometry. Rapid Commun. Mass Spectr., 12, 1986–1993.3.0.CO;2-H>CrossRefGoogle Scholar
Clore, G. M., and Gronenborn, A. M. (1991). Two-, three-, and four-dimensional NMR methods for obtaining more precise three-dimensional structure of proteins in solution. Ann. Rev. Biophys. Chem., 20, 29–63.CrossRefGoogle Scholar
Cluzel, P., Lebrun, A., Heller, C., Lavery, R., Viovy, J.-L., Chatenay, D., Caron, F. (1996). DNA: an extensible molecule. Science, 271, 792–794.CrossRefGoogle Scholar
Cohn, E. J., and Edsall, J. T. (1965). In Proteins, Amino Acids and Peptides as Ions and Dipolar Ions, eds. Cohn, E. J. and Edsall, J. T., pp. 370–381. New York: Hafner Publ. Co.Google Scholar
Collins, K. D. (1997). Charge density-dependent strength of hydration and biological structure. Biophys. J., 72(1), 65–76.CrossRefGoogle ScholarPubMed
Colon, L. A., Guo, Y., and Fermier, A. (1997). Capillary electrochromatography. Anal. Chem. News, 69, 461A–467A.CrossRefGoogle Scholar
Cordone, L., Ferrand, M., et al. (1999). Harmonic behavior of trehalose-coated carbon-monoxy-myoglobin at high temperature. Biophys. J., 76, 1043–1047.CrossRefGoogle ScholarPubMed
Crichton, R. R., Engelman, D. M., et al. (1977). Contrast variation study of specifically deuterated Escherichia coli ribosomal subunits. Proc. Natl. Acad. Sci. USA, 74, 5547–5550.CrossRefGoogle ScholarPubMed
Crick, F. H. (1968). The origin of the genetic code. J. Mol. Biol., 38, 367–379.CrossRefGoogle ScholarPubMed
Crothers, D. M., and Zimm, B. H. (1965) Viscosity and sedimentation of the DNA from bacteriophages T2 and T7 and the relation to molecular weight. J. Mol. Biol., 12, 527–536.CrossRefGoogle ScholarPubMed
Dasgupta, S., and Spiro, T. G. (1980). Resonance Raman characterization of the 7-ns photoproduct of (carbonmonoxy) hemoglobin: implications for hemoglobin dynamics. Biochemistry, 25, 5941–5948.CrossRefGoogle Scholar
Davidson, I. W., and Secrest, W. L. (1972). Determination of chromium in biological materials by atomic absorption spectrometry using a graphite furnace atomizer. Anal. Chem., 44(13), 1808–1813.CrossRefGoogle ScholarPubMed
Torre, G. (2001) Hydration from hydrodynamics. General consideration and applications to bead modelling to globular proteins. Biophys. Chem., 93, 159–170.CrossRefGoogle Scholar
Dekker, N. H., Rybenkov, V. V., et al. (2002). The mechanism of type IA topoisomerases. Proc. Natl. Acad. Sci. USA, 99, 12126–12131.CrossRefGoogle ScholarPubMed
Delano, W. L., and Brunger, A. T. (1995). The direct rotation function: rotational Patterson correlation search applied to molecular replacement. Acta Cryst. D, 51, 740–748.CrossRefGoogle Scholar
Derome, A. E. (1987). Modern NMR Techniques for Chemistry Research. New York: Pergamon.Google Scholar
Dessen, P., Blanquet, S., Zaccai, G., and Jacrot, B. (1978). Antico-operative binding of initiator transfer RNAMet to methionyl-transfer RNA synthetase from Escherichia coli: neutron scattering studies. J. Mol. Biol., 126, 293–313.CrossRefGoogle ScholarPubMed
Dickerson, R., and Geiss, I. (1969). The Structure and Action of Proteins. Menlo Park: Benjamin Cummings.Google Scholar
Diehl, M., Doster, W., et al. (1997). Water-coupled low-frequency modes of myoglobin and lysozyme observed by inelastic neutron scattering. Biophys. J., 73, 2726–32.CrossRefGoogle ScholarPubMed
Dobo, A., and Kaltashov, I. A. (2001). Detection of multiple protein conformational ensembles in solution via deconvolution of charge-state distribution in ESI MS. Anal. Chem., 73, 4763–4773.CrossRefGoogle Scholar
Dolgikh, D. A., Gilmanshin, R. I., et al. (1981). Alpha-lactalbumin: compact state with fluctuating tertiary structure? FEBS Lett., 136, 311–315.CrossRefGoogle ScholarPubMed
Dong, J., Wan, Z., Popov, M., Carey, P. R., and Weiss, M. A. (2003). Insulin assembly damps conformational fluctuations: Raman analysis of amide I linewidths in native states and fibrils. JMB, 330, 431–442.CrossRefGoogle ScholarPubMed
Doster, W., Cusack, S., et al. (1989). Dynamical transition of myoglobin revealed by inelastic neutron scattering. Nature, 337, 754–756.CrossRefGoogle ScholarPubMed
Doty, P., Bradbury, J. H., and Holtzer, A. M. (1956). Polypeptides. IV. The molecular weight, configuration and association of poly-γ-glutamate in various solvents. J. Am. Chem. Soc., 78, 947–954.CrossRefGoogle Scholar
Dubin, S. B., Clark, N. A., and Benedek, G. B. (1971). Measurement of the rotational diffusion coefficient of lysozyme by depolarised light scattering: configuration of lysozyme in solution. J. Chem. Phys., 54, 5158–5164.CrossRefGoogle Scholar
Dunkerk, A. K., and Williams, R. W. (1979). Ultraviolet and LASER Raman investigation of the buried tyrosines in fd phage. J. Biol. Chem., 254, 6446.Google Scholar
Dutta, R. K., Hammons, K., Willibey, B., and Haney, M. A. (1991). Analysis of protein denaturation by high-performance continuous differential viscometry. J. Cromatog., 536, 113–121.CrossRefGoogle ScholarPubMed
Dykxhoorn, D. M., Novina, C. D., et al. (2003). Killing the messenger: short RNAs that silence gene expression. Nat. Rev. Mol. Cell Biol., 4, 457–467.CrossRefGoogle ScholarPubMed
Eastman, J. E., Taguchi, A. K., et al. (2000). Characterization of a Rhodobacter capsulatus reaction center mutant that enhances the distinction between spectral forms of the initial electron donor. Biochemistry, 39, 14787–14798.CrossRefGoogle ScholarPubMed
Eden, D., Luu, B. Q., Zapata, D. J., Sablin, E. P., and Kuul, F. J. (1995). Solution structure of two molecular motor domains: nonclaret disjunctional and kinesin. Biophys. J., 68, 59s–65s.Google ScholarPubMed
Eigen, M., and Rigler, R. (1994). Sorting single molecules: applications to diagnostic and evolutionary biotechnology. Proc. Natl. Acad. Sci. USA, 91, 5740–5747.CrossRefGoogle Scholar
Eimer, W., and Pecora, R. (1991). Rotational and translational diffusion of short rodlike molecules in solution: oligonucleotides. J. Chem. Phys., 94, 2324–2329.CrossRefGoogle Scholar
Eimer, W., Williamson, J. R., Boxer, S. G., and Pecora, R. (1990). Characterization of the overall and internal dynamics of short oligonucleotides by depolarised dynamic light scattering and NMR relaxation measurements. Biochemistry, 29, 799–811.CrossRefGoogle Scholar
Eisenberg, H. (1981). Forward scattering of light, X-rays and neutrons. Q. Rev. Biophys., 14, 141–172.CrossRefGoogle ScholarPubMed
Elias, J. G., and Eden, D. (1981). Transient electric birefringence study of the persistence length and electrical polarizability of restriction fragments of DNA. Macromolecules, 14, 410–419.CrossRefGoogle Scholar
Elöve, G. A., Chaffotte, A. F., et al. (1992). Early steps in cytochrome c folding probed by time-resolved circular dichroism and fluorescence spectroscopy. Biochemistry, 31, 6876–6883.CrossRefGoogle ScholarPubMed
Engel, A., Lyubchenko, Y., and Muller, D. (1999). Atomic force microscopy: a powerful tool to observe biomolecules at work. Trends Cell Biol., 9, 77–80.CrossRefGoogle ScholarPubMed
Ernst, R. R., Bodenhausen, G., and Wokaun, A. (1987). Principles of Nuclear Magnetic Resonance in One and Two Dimensions. Oxford: Oxford University Press.Google Scholar
Essavaz-Roulet, B., Bockelman, U., and Heslot, F. (1997). Mechanical separation of the complementary strands of DNA. PNAS, 94, 11935–11940.CrossRefGoogle Scholar
Fan, E., Merritt, E. A., et al. (2000). AB(5) toxins: structures and inhibitor design. Curr. Opin. Struct. Biol., 10, 680–686.CrossRefGoogle ScholarPubMed
Fasman, G. D. (1996). Circular Dichroism and the Conformational Analysis of Biomolecules. New York: Plenum Press.CrossRefGoogle Scholar
Ferrer, M. L., Duchowicz, R., Carrasco, B., Garcia de la Torre, J., and Acuna, A. U. (2001). The conformation of serum albumin in solution: a combined phosphorescence depolarization–hydrodynamic modelling study. Biophys. J., 80, 2422–2430.CrossRefGoogle Scholar
Feynman, R. P., Leighton, R. B., et al. (1963). The Feynman Lectures on Physics. Reading, MA: Addison-Wesley Pub. Co.Google Scholar
Fisher, T. E., Marszalek, P. E., and Fernandez, J. M. (2000). Stretching single molecules into novel conformations using the atomic force microscopy. Nature Stuctr. Biol., 7, 719–724.Google Scholar
Flaux, J., Bertelsen, E. B., Horwich, A. L., and Wuthrich, K. (2002). NMR analysis of a 900K GroEL–GroES complex. Nature, 418, 207–211.Google Scholar
Flynn, P. F., and Wand, A. J. (2001). High-resolution nuclear magnetic resonance of Encapsulated proteins dissolved in low viscosity fluids, Methods Enzymol., 339, 54–70.CrossRefGoogle ScholarPubMed
Franklin, S. E., and Gosling, R. G. (1953). Molecular configuration in sodium thymonucleate. Nature, 171, 740–741.CrossRefGoogle ScholarPubMed
Franks, F., Gent, M., et al. (1963). Solubility of benzene in water. J. Chem. Soc., 8, 2716–2723.CrossRefGoogle Scholar
Franzen, S., and Boxer, S. G. (1997). On the origin of heme absorption band shifts and associated protein structural relaxation in myoglobin following flash photolysis. J. Biol. Chem., 272, 9655–9660.CrossRefGoogle ScholarPubMed
Frauenfelder, H., Parak, F., et al. (1988). Conformational substates in proteins. Ann. Rev. Biophys. Chem., 17, 451–479.CrossRefGoogle ScholarPubMed
Freifelder, D. (1970). Molecular weights of coliphages and coliphage DNA. IV, Molecular weights of DNA from bacteriophages T4, T5, and T7 and general problem of determination of molecular weight. J. Mol. Biol., 54, 567.CrossRefGoogle Scholar
Freire, E. (1995). Thermal denaturation methods in the study of protein folding. Methods Enzymol., 259, 144–168.CrossRefGoogle Scholar
Frey, W., Schief, W. R. Jr., et al. (1996). Two-dimensional protein crystallization via metal-ion coordination by naturally occurring surface histidines. Proc. Natl. Acad. Sci. USA, 93, 4937–4941.CrossRefGoogle ScholarPubMed
Frohn, J. T., Knapp, H. F., and Stemmer, A. (2000). True optical resolution beyond the Rayleigh limit achieved by standing wave illumination. Proc. Natl. Acad. Sci. USA, 97, 7232–7236.CrossRefGoogle ScholarPubMed
Fuller, S. D., and Argos, P. (1987). Is Sindbis a simple picornavirus with an envelope? Embo. J., 6, 1099–1105.Google ScholarPubMed
Fuller, W., Forsyth, T., et al. (2004). Water-DNA interactions as studied by X-ray and neutron fibre diffraction. Phil. Trans. R. Soc. Lond B Biol. Sci., 359, 1237–1247; discussion 1247–1248.CrossRefGoogle ScholarPubMed
Gabel, F., Bicout, D., et al. (2002). Protein dynamics studied by neutron scattering. Q. Rev. Biophys., 35, 327–367.CrossRefGoogle ScholarPubMed
Gabel, F. (2005). Protein dynamics in solution and powder measured by incoherent elastic neutron scattering: the influence of Q−range and energy resolution. Eur. Biophys. J., 34, 1–12.CrossRefGoogle ScholarPubMed
Gadola, S. D., Zaccai, N. R., et al. (2002). Structure of human CD1b with bound ligands at 2.3 å, a maze for alkyl chains. Nat. Immunol., 3, 721–726.CrossRefGoogle ScholarPubMed
Ganem, B. (1993). Detecting noncovalent interactions: new frontiers for mass spectrometry. Am. Biotechnol. Lab., 11, 32–34.Google ScholarPubMed
Ganem, B., Li, Y.-T., and Henion, J. D. (1991). Observation of noncovalent enzyme. Substrate and enzyme product complexes by ion spray mass spectrometry. J. Am. Chem. Soc., 113, 7818–7819.CrossRefGoogle Scholar
Garces-Chavez, V., McGloin, D., Melville, H., Sibbett, W., and Dholakia, K. (2002). Simultaneous micromanipulation in multiple planes using a self-reconstruction light beam. Nature, 419, 145–147.CrossRefGoogle Scholar
Garcia de la Torre, J. (2001). Hydration from hydrodynamics. General consideration and applications of bead modelling to globular proteins. Biophys. Chem., 93, 159–170.CrossRefGoogle Scholar
Garcia de la Torre, J., Martinez, M. C. L., and Tirado, M. M. (1984). Dimensions of short, rodlike macromolecules from translational and rotational diffusion coefficients. Study of the gramicidin dimer. Biopolymers, 23, 611–615.CrossRefGoogle Scholar
Garcia de la Torre, J., Navarro, S., and Lopez Martinez, M. C. (1994). Hydrodynamic properties of a double-helical model for DNA. Biophys. J., 66, 1573–1579.CrossRefGoogle ScholarPubMed
Garfin, D. E. (1995). Electrophoretic methods. In: Introduction to Biophysical Methods for Protein and Nucleic Acid Research, eds. Glaser, J. A. and Deutscher, M. P.. San Diego: Academic Press.Google Scholar
Garret, D. S., Seok, Y.-J., Liao, D.-I., Peterkofsky, A., Gronenborn, A. M., and Clore, G. M. (1997). Solution structure of the 30 kDa N-terminal domain of enzyme I of the Escherichia coli phosphoenolpuruvate: sugar phosphotransferase system by multidimensional NMR. Biochemistry, 36, 2517–2530.CrossRefGoogle Scholar
Gelles, J., and Landick, R. (1998). RNA polymerase as a molecular motor, Cell, 93, 13–16.CrossRefGoogle ScholarPubMed
Giege, R., Lorber, B., et al. (1982). Formation of a catalytically active complex between tRNAAsp and aspartyl-tRNA synthetase from yeast in high concentrations of ammonium sulphate. Biochimie, 64, 357–362.CrossRefGoogle ScholarPubMed
Gilbet, W. (1986). The RNA world. Nature, 319, 618.CrossRefGoogle Scholar
Gimzewski, J. K., and Joachum, C. (1999). Nanoscale science of single molecules using molecular probes. Science, 283, 1683–1688.CrossRefGoogle Scholar
Gluehmann, M., Zarivach, R., et al. (2001). Ribosomal crystallography: from poorly diffracting microcrystals to high-resolution structures. Methods, 25, 292–302.CrossRefGoogle ScholarPubMed
Go, N., Noguti, T., et al. (1983). Dynamics of a small globular protein in terms of low-frequency vibrational modes. Proc. Natl. Acad. Sci. USA, 80, 3696–3700.CrossRefGoogle ScholarPubMed
Godovach-Zimmermann, J., and Brown, L. R. (2001). Perspectives for mass spectrometry and functional proteomics. Mass Spectr. Rev., 20, 1–57.3.0.CO;2-J>CrossRefGoogle Scholar
Goldberg, D. E. (1989). Genetic Algorithms in Search, Optimization, and Machine Learning. Reading, MA: Addison-Wesley Pub. Co.Google Scholar
Gomez, J., Hilser, V. J., et al. (1995). The heat capacity of proteins. Proteins, 22, 404–412.CrossRefGoogle ScholarPubMed
Goodsell, D. S., and Olson, A. J. (1993). Soluble proteins: size, shape and function. TIBS, 18, 65–68.Google ScholarPubMed
Gordon, D. B. (2000). Mass spectrometric technique. In Principles and Techniques of Practical Biochemistry, 5th edn., eds. Wilson, K. and Walker, J.. Ch. 11. Cambridge: Cambridge University Press.Google Scholar
Greis, K. D., Hayes, B. K., et al. (1996). Selective detection and site-analysis of O-GlcNAc-modified glycopeptides by beta-elimination and tandem electrospray mass spectrometry. Anal. Biochem., 234, 38–49.CrossRefGoogle ScholarPubMed
Grier, D. (2003). A revolution in optical manipulation. Nature, 424, 810–816.CrossRefGoogle ScholarPubMed
Griko, Y. V., Freire, E., et al. (1995). The unfolding thermodynamics of c-type lysozymes: a calorimetric study of the heat denaturation of equine lysozyme. J. Mol. Biol., 252, 447–459.CrossRefGoogle ScholarPubMed
Griko, Y. V., Makhatadze, G. I., et al. (1994). Thermodynamics of barnase unfolding. Protein Sci., 3, 669–676.CrossRefGoogle ScholarPubMed
Gross, S. (2003). Application of optical traps in vivo. Methods in Enzymol.,V 361, 162–174.CrossRefGoogle ScholarPubMed
Grotjahn, L., Frank, R., and Blocker, H. (1982). Ultrafast sequencing of oligodeoxyribonucleotides by FAB-mass spectrometry. Nucl. Acids Res., 10, 4671–4677.CrossRefGoogle ScholarPubMed
Gutsche, I., Holzinger, J., et al. (2001). ATP-induced structural change of the thermosome is temperature-dependent. J. Struct. Biol., 135, 139–146.CrossRefGoogle ScholarPubMed
Haag, L., Garoff, H., Xing, L., Hammar, L., Kan, S. T., Cheng, R. H. (2002). Acid-induced movements in the glycoprotein shell of an alphavirus turn the spikes into membrane fusion mode. EMBO J., 21, 4402–4410.CrossRefGoogle ScholarPubMed
Hafner, J. H., Cheung, C.-L., Wooley, A. T., and Lieber, C. M. (2001). Structural and functional imaging with carbon nanotube AFM probes. Progr. Biophys. Mol. Biol., 77, 73–110.CrossRefGoogle ScholarPubMed
Hagerman, P. J. (1981). Investigation of the flexibility of DNA using transient electric birefringence. Biopolymers, 20, 1503–1535.CrossRefGoogle ScholarPubMed
Hagerman, P. J. (1985). Application of transient electric birefringence to the study of biopolymer structure. Methods Enzymol., 117, 199–215.Google Scholar
Hagerman, P. J. (2000). Transient electric birefringence for determining global conformations of non-helix elements and protein-induced bends in RNA. Methods Enzymol., 317, 440–453.CrossRefGoogle Scholar
Hahn, T. and International Union of Crystallography (2002). International Tables for Crystallography. Brief teaching edition of volume A, Space-group symmetry. Dordrecht; Boston, Published for the International Union of Crystallography by Kluwer Academic Publishers.Google Scholar
Hamm, P., Lim, M., and Hochstrasser, R. M. (1999). Structure of the amide I band of peptides measured by femtosecond non-linear-infrared spectroscopy. PNAS, 96, 6123–6128.Google Scholar
Han, W., Lindsay, S. M., Dlakic, M., and Harrington, R. E. (1997). Kinked DNA. Nature, 386, 563.CrossRefGoogle ScholarPubMed
Hansen, J. C., Lebowitz, J., and Demeler, B. (1994). Analytical ultracentrifugation of complex macromolecular systems. Biochemistry, 33, 13155–13163.CrossRefGoogle ScholarPubMed
Hansen, M. R., Mueller, L., and Pardi, A. (1998). Tunable alignment of macromolecules by filamentous phage yields dipolar coupling interaction. Nat. Struct. Biol., 5, 1065–1074.CrossRefGoogle Scholar
Harding, S. E. (1980). The combination of the viscosity increment with the harmonic mean rotational relaxation time for determining the conformation of biological macromolecules in solution. Biochem. J., 189, 359–361.CrossRefGoogle ScholarPubMed
Harding, S. E. (1981). A compound hydrodynamics shape function derived from viscosity and molecular covolume measurements. Int. J. Biol. Macromol., 3, 398–399.CrossRefGoogle Scholar
Harding, S. E. (1995). On the hydrodynamic analysis of macromolecular conformation. Biophys. Chem., 55, 69–93.CrossRefGoogle ScholarPubMed
Harding, S. E., and Rowe, A. (1982). Modelling biological macromolecules in solution: 1. The ellipsoid of revolution. Int. J. Biol. Macromol., 4, 160–164.CrossRefGoogle Scholar
Harding, S. E., Horton, J. C., and Colfen, H. (1997). The ELLIPS suite of macromolecular conformation algorithms. Eur. Biophys, J., 25, 347–359.CrossRefGoogle ScholarPubMed
Harpaz, Y., Gerstein, M., Chothia, C. (1994). Volume changes on protein folding. Structure, 2, 641–649.CrossRefGoogle ScholarPubMed
Harris, R. (1983). Nuclear Magnetic Resonance Spectroscopy. London: Pitman.Google Scholar
Haupts, U., Tittor, J., et al. (1997). General concept for ion translocation by halobacterial retinal proteins: the isomerization/switch/transfer (IST) model. Biochemistry, 36, 2–7.CrossRefGoogle ScholarPubMed
Haupts, U., Tittor, J., and Oesterhelt, D. (1999). Closing in on bacteriorhodopsin: progress in understanding the molecule. Ann. Rev. Biophys. Biomol. Struct., 28, 367–399.CrossRefGoogle ScholarPubMed
Hausten, E., and Schwille, P. (2003). Ultrasensitive investigations of biological systems by fluorescence correlation spectroscopy. Methods, 29, 153–166.CrossRefGoogle Scholar
Hazlett, T. L., Moore, K. J. M., Lowe, P. N., Jameson, D. M., and Eccleston, J. F. (1993). Solution of p21ras proteins bound with fluorescent nucleotides: a time-resolved fluorescence study. Biochemistry, 32, 13575–13583.CrossRefGoogle ScholarPubMed
Heberle, J., and Gensch, T. (2001). When FT-IR spectroscopy meets X-ray crystallography. Nat. Struct. Biol., 8, 195–197.CrossRefGoogle ScholarPubMed
Hellweg, T., Eimer, W., Krahn, E., Schneider, K., and Muller, A. (1997). Hydrodynamic properties of nitrogenase – the MoFe protein from Azotobacter vinelandii studied by dynamic light-scattering and hydrodynamic modelling. Biochim. Biophys. Acta, 1337, 311–318.CrossRefGoogle ScholarPubMed
Hensley, P. (1996). Defining the structure and stability of macromolecular assemblies in solution: the re-emergence of analytical ultracentrifugation as a practical tool. Structure, 4, 367–373.CrossRefGoogle ScholarPubMed
Hillisch, A., Lorenz, M., and Diekmann, S. (2001). Recent advances in FRET: distance determination in protein–DNA complexes. Curr. Opin. Struct. Biol., 11, 201–207.CrossRefGoogle ScholarPubMed
Hirao, I., and Ellington, A. D. (1995). Re-creating the RNA world. Curr. Biol., 5, 1017–1022.CrossRefGoogle ScholarPubMed
Homans, S.W., Edge, C. J., Ferguson, M. A., and Dwek, R. A. (1989). Solution structure of the glycosylphosphatidylinositol membrane anchor glycan of Trypanosoma bruccei variant surface glycoprotein. Biochemistry, 28, 2881–2887.CrossRefGoogle ScholarPubMed
Hore, P. J. (1995). Nuclear Magnetic Resonance. Oxford: Oxford University Press.Google Scholar
Horwitz, J., Strickland, E. H., and Billups, C. (1970). Analysis of the vibrational structure in the near-ultraviolet circular dichroism and absorption spectra of tyrosine derivatives and ribonuclease-A at 77 K. J. Am. Chem. Soc., 92, 2119–2129.CrossRefGoogle Scholar
Hu, C.-M., and Zwanzig, R. (1974). Rotational friction coefficients for spheroids with the slipping boundary conditions. J. Chem. Phys., 60, 4354–4357.CrossRefGoogle Scholar
Hunt, J. F., McCrea, P. D., et al. (1997). Assessment of the aggregation state of integral membrane proteins in reconstituted phospholipid vesicles using small angle neutron scattering. J. Mol. Biol., 273, 1004–1019.CrossRefGoogle ScholarPubMed
Hutchens, J. O. (1970). Handbook of Chemistry and Selected Data for Molecular Biology, ed. Sober, H. A.Cleveland, OH: Chemical Rubber Co; International Tables for Crystallography: Spzae Group Symmetry (2002). Dordrecht: Kluwer.Google Scholar
Huygens, C. (1690). Treatise on Light, New York: Dover (1962) of the English translation first published by Macmillan and Co. in 1912.Google Scholar
Ishijima, A., Kojima, H., et al. (1998). Simultaneous observation of individual ATPase and mechanical events by a single myosin molecule during interaction with actin. Cell, 92, 161–171.CrossRefGoogle ScholarPubMed
Ishima, R., and Torchia, D. (2000). Protein dynamics from NMR. Nat. Struct. Biol., 7, 740–743.CrossRefGoogle ScholarPubMed
Jacobs, R. E., Ahrens, E. T., Meade, T. J., and Fraser, S. E. (1999). Looking deeper into vertebrate development. TIBS, 9, 73–76.Google ScholarPubMed
Jacrot, B. (1976). The study of biological structures by neutron scattering from solution. Rep. Prog. Phys., 39, 911–953.CrossRefGoogle Scholar
Jacrot, B., and Zaccai, G. (1981). Determination of molecular weight by neutron scattering. Biopolymers, 20, 2414–2426.CrossRefGoogle Scholar
Jacrot, B., Chauvin, C., and Witz, J. (1977). Comparative neutron small-angle scattering study of small spherical RNA viruses. Nature, 266(5601), 417–421.CrossRefGoogle ScholarPubMed
Jaenicke, R. (2000). Do ultrastable proteins from hyperthermophiles have high or low conformational rigidity? Proc. Natl. Acad. Sci. USA, 97, 2962–2962.CrossRefGoogle ScholarPubMed
Jancarik, J., and Kim, S.-H. (1991). Sparse matrix sampling: a screening method for crystallization of proteins. J. Appl. Cryst., 24, 409–411.CrossRefGoogle Scholar
Jancarik, J., Scott, W. G., et al. (1991). Crystallization and preliminary X-ray diffraction study of the ligand-binding domain of the bacterial chemotaxis-mediating aspartate receptor of Salmonella typhimurium. J. Mol. Biol., 221, 31–34.CrossRefGoogle ScholarPubMed
Jeener, J. (1996). In Encyclopedia of Nuclear Magnetic Resonance, eds. Grant, D. M., Harris, R. K., Vol. 1, p. 40, Chichester: John Wiley and Sons.Google Scholar
Jeffrey, P. D., Nichol, L. W., Turner, D. R., and Winzor, D. J. (1977). The combination of molecular covolume and frictional coefficient to determine the shape and axial ratio of a rigid macromolecule. Studies on Ovalbumin. J. Phys. Chem., 81, 776–781.CrossRefGoogle Scholar
Jeruzalmi, D. and Steitz, T. A. (1997). Use of organic cosmotropic solutes to crystallize flexible proteins: application to T7 RNA polymerase and its complex with the inhibitor T7 lysozyme. J. Mol. Biol., 274, 748–756.CrossRefGoogle ScholarPubMed
Jia, Y., Sytnic, A., Li, L., Vladimirov, S., Cooperman, B. S., and Hochstrasser, R. M. (1997). Nonexponencial kinetics of a single tRNA Phe molecule under physiological conditions. Proc. Natl. Acad. Sci. USA, 94, 7932–7936.CrossRefGoogle ScholarPubMed
Jiang, Y., Ruta, V., et al. (2003). The principle of gating charge movement in a voltage-dependent K+ channel. Nature, 423, 42–48.CrossRefGoogle Scholar
Johnson, W. C. (1985). Circular Dichroism and Its Empirical Application to Biopolymers. Methods of Biochemical Analysis, Vol. 31, ed. Glick, D.. Chichester: John Wiley and Sons Inc.CrossRefGoogle ScholarPubMed
Johnson, K. H., and Gray, D. M. (1992). Analysis of an RNA pseudoknot structure by CD spectroscopy. J. Biomol. Struct. Dyn., 9, 733–745.CrossRefGoogle ScholarPubMed
Jolly, D., and Eisenberg, H. (1976). Photon correlation spectroscopy, total intensity light with laser radiation, and hydrodynamic studies of a well fractionated DNA sample. Biopolymers, 15, 61–95.CrossRefGoogle ScholarPubMed
Jones, J. A., Wilkins, D. K., Smith, L. J., and Dobson, C. M. (1997). Characterization of protein unfolding by NMR diffusion measurements. J. Biomolecular NMR, 10, 199–203.CrossRefGoogle Scholar
Jones, T. A., Zou, J. Y., et al. (1991). Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallogr. A, 47 (Pt 2), 110–119.CrossRefGoogle ScholarPubMed
Kabsch, W. (1988). Evaluation of single-crystal X-ray diffraction data from a position-sensitive detector. J. Appl. Cryst., 21, 916–924.CrossRefGoogle Scholar
Karger, B. L., Chu, Y.-H., and Foret, F. (1995). Capillary electrophoresis of proteins and nucleic acids. Annu. Rev. Biophys. Biomol. Struct., 24, 579–610.CrossRefGoogle ScholarPubMed
Kassas, S., Thomson, N. H., et al. (1997). Esherichia coli RNA polymerase activity observed using atomic force microscopy. Biochemistry, 36, 461–468.CrossRefGoogle Scholar
Kebbekus, P., Draper, D. E., and Hagerman, P. (1995). Persistence length of RNA. Biochemistry, 34, 4354–4357.CrossRefGoogle ScholarPubMed
Keiderling, T. A. (2002). Protein and peptide secondary structure and conformational determination with vibrational circular dichroism. Curr. Opin. Chem. Biol., 6, 682–688.CrossRefGoogle ScholarPubMed
Kincaid, J. R. (1995). Structure and dynamics of transient species using time-resolved resonance Raman spectroscopy. Methods Enzymol., 246, 460–501.CrossRefGoogle ScholarPubMed
King, R. W., and Williams, K. R. (1989a). The Fourier transform in chemistry. Part 1, Nuclear magnetic resonance: introduction. J. Chem. Education, 66, A213–A219.CrossRefGoogle Scholar
King, R. W., and Williams, K. R. (1989b). The Fourier transform in chemistry, Part 2, Nuclear magnetic resonance: the single pulse experiment. J. Chem. Education, 66, A243–A248.CrossRefGoogle Scholar
Kinosita, K. Jr., Yasuda, R., Noji, H., Ishiwata, S., and Yoshida, M. (1998). F1-ATPase: a rotary motor made of a single molecule. Cell, 93, 21–24.CrossRefGoogle ScholarPubMed
Klar, T. A., Jacobs, S., Dyba, M., Egner, A., and Hell, S. W. (2000). Fluorescence microscopy with diffraction resolution barrier broken by stimulated emission. Proc. Natl. Acad. Sci. USA, 97, 8206–8210.CrossRefGoogle ScholarPubMed
Klein, G., Satre, M., et al. (1982). Spontaneous aggregation of the mitochondrial natural ATPase inhibitor in salt solutions as demonstrated by gel filtration and neutron scattering. Application to the concomitant purification of the ATPase inhibitor and F1-ATPase. Biochim. Biophys. Acta, 681, 226–232.CrossRefGoogle ScholarPubMed
Kleywegt, G. J., and Jones, T. A. (2002). Homo crystallographicus – quo vadis? Structure (Camb), 10, 465–472.CrossRefGoogle ScholarPubMed
Klotz, L. C., and Zimm, B. H. (1972). Size of DNA determined by viscoelastic measurements: results on bacteriophage, Bacillus subtilus and Escherichia coli, 72, 779–800.Google Scholar
Koch, M. H., and Stuhrmann, H. B. (1979). Neutron-scattering studies of ribosomes. Methods Enzymol., 59, 670–706.CrossRefGoogle ScholarPubMed
Konermann, L., and Douglas, D. J. (1998a). Equilibrum unfolding of proteins monitored by electrospray ionization mass spectrometry: distinguishing two-state from multi-state transition. Rapid Commin. Mass Spectr., 12, 435–442.3.0.CO;2-F>CrossRefGoogle Scholar
Konermann, L., and Douglas, D. J. (1998b). Unfolding of proteins monitored by electrospray ionization mass spectrometry: a comparison of positive and negative ion modes. J. Am. Soc. Mass Spectrom., 9, 1248–1254.CrossRefGoogle Scholar
Koppel, D. E. (1979). Fluorescence redistribution after photobleaching. Biophys. J., 28, 281–291.CrossRefGoogle ScholarPubMed
Korgel, B. A., Zanten, J. H., and Monbouquette, H. G. (1998). Vesicle size distributions measured by flow field-flow fractionation coupled with multiangle light scattering. Biophys. J., 74, 3264–3272.CrossRefGoogle ScholarPubMed
Kossiakoff, A. A. (1983). Neutron protein crystallography: advances in methods and applications. Ann. Rev. Biophys. Bioeng., 12, 159–182.CrossRefGoogle ScholarPubMed
Kovacic, R. T., and Holde, K. E. (1977). Sedimentation of homogeneous double-strand DNA molecules. Biochemistry, 16, 1490–1498.CrossRefGoogle ScholarPubMed
Kroes, S. J., Canters, G. W., Giardi, G., Hoek, A., and Visser, A. J. W. G. (1998). Time-resolved fluorescence study of azurin variants: conformational heterogeneity and tryptophan mobility. Biophys. J., 75, 2441–2450.CrossRefGoogle ScholarPubMed
Kumar, A., Ernst, R. R., and Wuthrich, K. (1980). A two-dimensional nuclear Overhauser enhancement (2D NOE) experiment for the elucidation of complete proton–proton cross-relaxation networks in biological macromolecules. Biochem. Biophys. Res. Commun., 95, 1–6.CrossRefGoogle ScholarPubMed
Kunst, F., Oyasawara, N., et al. (1997). The complete genome sequence of the Gram-positive bacterium Bacillus subtilis. Nature, 390, 249–256.CrossRefGoogle ScholarPubMed
Kuntz, I. D. Jr., and Kauzmann, W. (1974). Hydration of Proteins and Polypeptides. In Advances in Protein Chemistry, eds. Anfinsen, C. B., Edsall, J. T. and Richards, F. M., Vol. 28. New York: Academic Press.Google Scholar
Lakowicz, J. R. (ed.) (1999). Principles of Fluorescence Spectroscopy, second edn. New York: Kluwer Academic/Plenum Publ.CrossRefGoogle Scholar
Lakowicz, J. R., Gryczynski, I., et al. (2000). Microsecond dynamics of biological macromolecules. Methods Enzymol., 323, 473–509.CrossRefGoogle ScholarPubMed
Langan, P., Nishiyama, Y., et al. (1999). A revised structure and hydrogen bonding system in cellulose II from a neutron fibre diffraction analysis. J. Am. Chem. Soc., 121, 9940–9946.CrossRefGoogle Scholar
Langley, K. H. (1992). Developments in electrophoretic laser light scattering and some biochemical applications. In Laser Scattering in Biochemistry, eds. Harding, S. E., Sattelle, D. B., and Bloomfield, V. A.. Cambridge: Royal Society of Chemistry.Google Scholar
Langowski, J., Kremer, W., and Kapp, U. (1992). Dynamic light scattering for study of solution conformation and dynamics of superhelical DNA. Methods Enzymol., 211, 431–448.Google ScholarPubMed
Laue, T. M., Ridgeway, T. M., Wool, J. O., and Shepard, H. K. (1996). Insights into a new analytical electrophoresis apparatus. J. Pharm. Sci., 85, 1331–1335.CrossRefGoogle ScholarPubMed
Lay, J. O. (2001). MALDI-TOF mass spectrometry of bacteria. Mass Spectr. Rev., 20, 172–194.CrossRefGoogle Scholar
Leavitt, S., and Freire, E. (2001). Direct measurement of protein binding energetics by isothermal titration calorimetry. Curr. Opin. Struct. Biol., 11, 560–566.CrossRefGoogle ScholarPubMed
Lehnert, U. (2002). Hydration dependence of local thermal motions in the Purple Membrane explored by neutron scattering and isotopic labeling. Ph D Thesis. Université Joseph Fourier, Grenoble.Google Scholar
Lemasters, J. J., Chacon, E., Zahrebelski, G., Reece, J. M., and Nieminen, A.-L. (1993). Laser scanning confocal microscopy of living cells. In: Optical Microscopy. Emerging Methods and Application, eds. Herman, B. and Lemasters, J. J.. San Diego: Academic Press.Google Scholar
Leone, M., Cupane, A., et al. (1994). Thermal broadening of the Soret band in heme complexes and in heme-proteins: role of iron dynamics. Eur. Biophys. J., 23, 349–352.CrossRefGoogle ScholarPubMed
Lescrinier, E., Froeyen, M., and Herdewijn, P. (2003). Difference in conformational diversity between nucleic acids with six-membered ‘sugar’ unit and natural ‘furanose’ nucleic acids. Nucleic Acids. Res., 31, 2975–2989.CrossRefGoogle ScholarPubMed
Leslie, A. G. (1999). Integration of macromolecular diffraction data. Acta Crystallogr. D Biol. Crystallogr., 55 (Pt 10), 1696–1702.CrossRefGoogle ScholarPubMed
Levitt, M., Sander, C., et al. (1985). Protein normal-mode dynamics: trypsin inhibitor, crambin, ribonuclease and lysozyme. J. Mol. Biol., 181, 423–447.CrossRefGoogle ScholarPubMed
Lewis, A., Lieberman, K., et al. (1995). New design and imaging concepts in NSOM, Ultramicroscopy, 61, 215–220.CrossRefGoogle Scholar
Lewis, A., Radko, A., Ami, N. B., Palanker, D., and Lieberman, K. (1999). Near-field scanning optical microscopy in cell biology. TIBS, 9, 70–73.Google ScholarPubMed
Li, H., Cocco, M. J., Steitz, T., and Engelman, D. M. (2001). Conversion of phospholamban into a soluble pentameric helical bundle. Biochemistry, 40, 6636–6645.CrossRefGoogle ScholarPubMed
Li, Y., Hunter, R. L., and Mciver, R. T. Jr. (1994). High-resolution mass spectrometer for protein chemistry. Nature, 370, 393–395.CrossRefGoogle ScholarPubMed
Liphardt, J., Onoa, B., Smith, S. B., Tinoco, I. Jr., and Bustamante, C. (2001). Reversible unfolding of single RNA moleculees by mechanical forces. Science, 292, 733–737.CrossRefGoogle Scholar
Longsworth, L. G. (1953). Diffusion measurements, at 25ο, of aqueous solution of amino acids, peptides and sugars. J. Am. Chem. Soc., 75, 5705–5709.CrossRefGoogle Scholar
Lu, H. P., Xun, L., and Xie, X. S. (1998). Single-molecule enzymatic dynamics. Science, 282, 1877–1882.CrossRefGoogle ScholarPubMed
Luthy, R., Bowie, J. U., and Eisenberg, D. (1992). Assessment of protein models with three-dimensional profiles. Nature, 356, 83–85.CrossRefGoogle ScholarPubMed
Luz, Z., and Meiboom, S. (1963). J. Chem. Phys., 39, 366–370.CrossRef
Luzzati, V., and Tarclieu, A. (1980). Recent developments in solution X-ray scattering. Annu. Rev. Biophys. Bioeng., 9, 1–29.CrossRefGoogle ScholarPubMed
Ma, J., Flynn, T. C., et al. (2002). A dynamic analysis of the rotation mechanism for conformational change in F(1)-ATPase. Structure (Camb), 10, 921–931.CrossRefGoogle ScholarPubMed
MacDonald, D., and Lu, P. (2002). Residual dipolar couplings in nucleic acid structure determination. Curr. Opinion Str. Biol., 12, 337–343.CrossRefGoogle ScholarPubMed
MacGregor, I. K., Anderson, A. L., and Laue, T. M. (2004). Fluorescence detection for the XLI analytical ultracentrifuge. Biophys. Chem., 108, 165–185.CrossRefGoogle ScholarPubMed
Machtle, W. (1999). High-resolution, submicron particle size distribution analysis using gravitational-sweep sedimentation. Biophys. J., 76, 1080–1091.CrossRefGoogle ScholarPubMed
Madern, D., Ebel, C., et al. (2000). Halophilic adaptation of enzymes. Extremophiles, 4, 91–98.CrossRefGoogle ScholarPubMed
Makhatadze, G. I., and Privalov, P. L. (1990). Heat capacity of proteins. I. Partial molar heat capacity of individual amino acid residues in aqueous solution: hydration effect. J. Mol. Biol., 213, 375–384.CrossRefGoogle ScholarPubMed
Mancini, E. J., Clarke, M., et al. (2000). Cryo-electron microscopy reveals the functional organization of an enveloped virus, Semliki Forest virus. Mol. Cell. 5, 255–266.CrossRefGoogle Scholar
Mandelkow, E., and Mandelkow, E. M. (2002). Kinesin motor and disease. TICB, 12, 585–591.CrossRefGoogle ScholarPubMed
Manor, D., Weng, G., Deng, H., Cosloy, S., Chen, C. X. (1991). An isotope edited classical Raman difference spectroscopic study of the interactions of guanine nucleotides with elongation factor Tu and H-ras p21. Biochemistry, 30, 10914–10920.CrossRefGoogle ScholarPubMed
Mark, J. E. (1996). Physical Properties of Polymers Handbook. Woodbury, NY: AIP Press.Google Scholar
Marassi, F. M., Ma, C., et al. (1999). Correlation of the structural and functional domains in the membrane protein Vpu from HIV-1. PNAS, 96, 14336–14341.CrossRefGoogle ScholarPubMed
Mattei, B., Borch, J., and Roepstorff, P. (2004). Biomolecular interaction analysis and MS. Anal. Chem., 76, 18A–26A.CrossRefGoogle Scholar
McLafferty, F. W. (1993). Interpretation of Mass Spectra, fourth edn. Mill Valley, CA: University Science Books.Google Scholar
Medalia, O., Weber, I., et al. (2002). Macromolecular architecture in eukaryotic cells visualized by cryoelectron tomography. Science, 298, 1209–1213.CrossRefGoogle ScholarPubMed
Medek, A., Olejniczak, E. T., Meadows, R. P., and Fesik, S. W. (2000). An approach for high-throughput structure determination of proteins by NMR spectroscopy. J. Biomol. NMR, 18, 229–238.CrossRefGoogle ScholarPubMed
Mehta, A. D., Rief, M., Spudlich, J. A., Smith, D. A., and Simmons, R. M. (1999). Single-molecule biomechanics with optical methods. Science, 283, 1689–1695.CrossRefGoogle ScholarPubMed
Merritt, E. A., Sarfaty, S., et al. (1994). Crystal structure of cholera toxin B-pentamer bound to receptor GM1 pentasaccharide. Protein Sci, 3, 166–175.CrossRefGoogle ScholarPubMed
Merzel, F., and Smith, J. C. (2002). Is the first hydration shell of lysozyme of higher density than bulk water? Proc. Natl. Acad. Sci. USA, 99, 5378–5383.CrossRefGoogle ScholarPubMed
Meselson, M., and Stahl, F. W. (1958). The replication of DNA in Escherichia coli. Proc. Natl. Acad. Sci. USA, 44, 671–682.CrossRefGoogle ScholarPubMed
Meselson, M., Stahl, F. W., and Vinograd, J. (1957). Equilibrium sedimentation of macromolecules in density gradient. Proc. Natl. Acad. Sci. USA, 43, 581–588.CrossRefGoogle Scholar
Michielsen, S., and Pecora, R. (1981). Solution dimensions of the gramicidin dimer by dynamic light scattering. Biochemistry, 20, 6994–6997.CrossRefGoogle ScholarPubMed
Miele, A. E., Federici, L., et al. (2003). Analysis of the effect of microgravity on protein crystal quality: the case of a myoglobin triple mutant. Acta Crystallogr. D Biol. Crystallogr., 59 (Pt 6), 982–988.CrossRefGoogle ScholarPubMed
Mitellbach, P., and Porod, G. (1962). Zur Röntgenkleinwinkelstreuung die Berechning der Steukurven von Dreiachsigen Ellipsoiden. Acta Physica Austriaca, 15, 122–147.Google Scholar
Moerner, W. E., and Orrit, M. (1999). Illuminating single molecules in condensed matter. Science, 283, 1670–1676.CrossRefGoogle ScholarPubMed
Mollova, E. T., Hansen, M. R., and Pardi, A. (2000). Global structure of RNA determined with residual dipolar coupling, JACS, 122, 11561–11562.CrossRefGoogle Scholar
Monkos, K., and Turczynski, B. (1991). Determination of the axial ratio of globular proteins in the aqueous solution using viscometric measurements. Int. J. Biol. Macromol. 13, 341–344.CrossRefGoogle ScholarPubMed
Montelione, G. T., Zheng, D., Huang, Y. J., Gundalus, K. S., and Szyperski, T. (2000). Protein NMR spectroscopy in structural genomics. Nat. Struct. Biol., 7, 982–985.CrossRefGoogle ScholarPubMed
Moscowitz, A. (1962). Theoretical aspects of optical activity. Part one: Small molecules. Adv. Chem. Phys., 4, 67–112.Google Scholar
Mou, J., Csajkovsky, D. M., Zhang, Y., and Shao, Z. (1995). High-resolution atomic force microscopy of DNA: the pitch of the double helix. FEBS Lett., 371, 279–282.Google ScholarPubMed
Muller, D. J., Janovjak, H., Lehto, T., Kuerschner, L., and Anderson, K. (2002). Observing structure, function and assembly of single proteins by AFM. Progress Biophys. Mol. Biol., 79, 1–43.CrossRefGoogle ScholarPubMed
Mullett, W. M., Lai, E. P., et al. (2000). Surface plasmon resonance-based immunoassays. Methods, 22, 77–91.CrossRefGoogle ScholarPubMed
Munro, I., Pecht, I., and Stryer, L. (1979). Subnanosecond motions of tryptophan residues in proteins. Proc. Natl. Acad. Sci. USA, 76, 56–60.CrossRefGoogle ScholarPubMed
Myers, E. W., Sutton, G. G., et al. (2000). A whole-genome assembly of Drosophila. Science, 287, 2196–2204.CrossRefGoogle ScholarPubMed
Myszka, D. G., Sweet, R. W., et al. (2000). Energetics of the HIV gp120-CD4 binding reaction. Proc. Natl. Acad. Sci. USA, 97, 9026–9031.CrossRefGoogle ScholarPubMed
Nagorni, M., and Hell, S. W. (1998). 4Pi-Confocal microscopy provides three-dimensional images of the microtubule network with 100- to 150-nm resolution. J. Struct. Biol., 123, 236–247.CrossRefGoogle ScholarPubMed
Navarro, S., Lopez Martinez, M. C., Garcia de la Torre, J. (1995). Relaxation times in electric birefringence of flexible polymer chains. J. Chem. Phys., 103, 7631–7639.CrossRefGoogle Scholar
Navaza, J., and Saludjian, P. (1997). AMoRe: an automated molecular replacement program package. Methods Enzymol., 276, 581–594.CrossRefGoogle ScholarPubMed
Newcomb, W. W., Juhas, R. M., et al. (2001). The UL6 gene product forms the portal for entry of DNA into the herpes simplex virus capsid. J. Virol., 75, 10923–10932.CrossRefGoogle ScholarPubMed
Nie, S., and Zare, R. N. (1997). Optical detection of single molecule. Annu. Rev. Biophys. Biomol. Struct., 26, 567–596.CrossRefGoogle Scholar
Nie, S., Chlu, D. T., and Zare, R. N. (1995). Real-time detection of single molecules in solution by confocal fluorescence microscopy. Anal. Chem., 67, 2849–2857.CrossRefGoogle Scholar
Nielsen, M. L., Bennet, K. L., Larsen, B., Monniate, M., and Mann, M. (2002). Peptide end sequencing by orthogonal MALDI tandem mass spectroscopy. J. Proteome Res., 1, 63–71.CrossRefGoogle Scholar
Nir, S., and Stein, W. D. (1971). Two modes of diffusion in liquids. J. Chem. Phys., 55, 1598–1603.CrossRefGoogle Scholar
Nishiyama, Y., Langan, P., et al. (2002). Crystal structure and hydrogen-bonding system in cellulose Ibeta from synchrotron X-ray and neutron fiber diffraction. J. Am. Chem. Soc., 124, 9074–9082.CrossRefGoogle ScholarPubMed
Nishiyama, Y., Okano, T., et al. (1999). High resolution neutron fibre diffraction data on hydrogenated and deuterated cellulose. Int. J. Biol. Macromol., 26, 279–283.CrossRefGoogle ScholarPubMed
Noji, H., Yasuda, R.et al. (1997). Direct observation of the rotation of F1-ATPase. Nature, 386, 299–302.CrossRefGoogle ScholarPubMed
Nollert, P., Navarro, J., and Landau, E. M. (2002). Crystallization of membrane proteins in cubo. Methods in Enzym., 343, 183–99.CrossRefGoogle ScholarPubMed
Oberg, K. A., Ruysschaert, J. M., et al. (2004). The optimization of protein secondary structure determination with infrared and circular dichroism spectra. Eur. J. Biochem., 271, 2937–2948.CrossRefGoogle ScholarPubMed
Ogorzalek Loo, R. R., Cavalcoli, J. D., et al. (2001). Virtual 2-D gel electrophoresis: visualization and analysis of the E. coli proteome by mass spectrometry. Anal. Chem., 73, 4063–4070.CrossRefGoogle Scholar
Opella, S. J., and Stewart, P. L. (1989). Solid-state nuclear magnetic resonance structural studies of proteins. Methods Enzymol., 176, 242–275.CrossRefGoogle ScholarPubMed
Orgel, L. E. (1968). Evolution of the genetic apparatus. J. Mol. Biol., 38(3), 381–93.CrossRefGoogle ScholarPubMed
Ormo, M., Cubbit, A. B., Kallio, K., Gross, L. A., and Tsien, R. Y. (1996). Crystal structure of the Aequorea victoria green fluorescent protein. Science, 273, 1392–1395.CrossRefGoogle ScholarPubMed
Oster, G., and Wang, H. (2003). Rotary protein motor. Trends Cell Biol., 13, 114–121.CrossRefGoogle Scholar
Otting, G., and Wuthrich, K. (1990) Heteronuclear filters in two-dimensional [1H,1H]-NMR spectroscopy: combined use with isotope labelling for studies of macromolecular conformation and intermolecular interactions. Q. Rev. Biophys., 23, 39–56.CrossRefGoogle ScholarPubMed
Otwinowski, Z., and Minor, W. (1997). Processing of X-ray diffraction data in oscillation mode. Methods Enzymol., 276, 307–326.CrossRefGoogle ScholarPubMed
Pauling, L., and Corey, R. (1953). A proposed structure for the nucleic acids. PNAS, 39, 84–97.CrossRefGoogle ScholarPubMed
Pecora, R. (1968). Spectral distribution of light scattered by monodisperse rigid rods. J. Chem. Phys., 48, 4126–4128.CrossRefGoogle Scholar
Perrakis, A., Sixma, T. K., et al. (1997). wARP: improvement and extension of crystallographic phases by weighted averaging of multiple-refined dummy atomic models. Acta Cryst. D., 53, 448–455.CrossRefGoogle ScholarPubMed
Peticolas, W. L. (1995). Raman spectroscopy of DNA and proteins. Methods Enzymol., 246, 389–416.CrossRefGoogle ScholarPubMed
Peticolas, W. L., and Evertsz, E. (1992). Conformation of DNA in vitro and in vivo from laser Raman scattering. Methods in Enzymol., 211, 335–352.CrossRefGoogle ScholarPubMed
Pfuhl, M., Chen, H. A., Kristinsen, S., Driscool, P. C. (1999). NMR exchange broadening arising from specific low affinity protein self-association: Analysis of nitrogen-15 relaxation for rat CD2 domain 1. J. Biomolecular NMR, 14, 307–320.CrossRefGoogle ScholarPubMed
Piston, D. W. (1999). Imaging living cells and tissues by two-photon excitation microscopy. Trends Cell Biol., 9, 66–69.CrossRefGoogle ScholarPubMed
Pitner, T. P., Walter, R., and Glickson, J. D. (1976). Mechanism of the intramolecular 1H nuclear Overhauser effect in peptides and depsipeptides. Biochem. Biophys. Res. Commun., 70, 746–751.CrossRefGoogle ScholarPubMed
Plenert, M. L., and Shear, J. B. (2003). Microsecond electrophoresis. Proc. Natl. Acad. Sci. USA, 100, 3853–3857.CrossRefGoogle ScholarPubMed
Pohl, F. M., and Jovin, T. M. (1972). Salt-induced co-operative conformational change of a synthetic DNA: equilibrium and kinetic studies with poly (dG-dC). J. Mol. Biol., 67, 375–396.CrossRefGoogle Scholar
Popot, J. L., Berry, E. A., et al. (2003). Amphipols: polymeric surfactants for membrane biology research. Cell Mol. Life Sci., 60, 1559–1574.CrossRefGoogle ScholarPubMed
Prestegard, J. H. (1998). New techniques in structural NMR – anisotropic interactions. Nature Struct. Biol., 5, 517–522.CrossRefGoogle ScholarPubMed
Prestegard, J. H., Valafar, H., Glushka, J., and Tian, F. (2001). Nuclear magnetic resonance in the era of structural Genomics. Biochemistry, 40, 8677–8685.CrossRefGoogle ScholarPubMed
Price, N. C., Dwek, R. A., et al. (2001). Principles and Problems in Physical Chemistry for Biochemists. Oxford: Oxford University Press.Google Scholar
Price, P. B. (2000). A habitat for psychrophiles in deep Antarctic ice. Proc. Natl. Acad. Sci. USA, 97, 1247–1251.CrossRefGoogle ScholarPubMed
Privalov, G. P., and Privalov, P. L. (2000). Problems and prospects in microcalorimetry of biological macromolecules. Methods Enzymol., 323, 31–62.CrossRefGoogle ScholarPubMed
Privalov, G., Kavina, V., et al. (1995). Precise scanning calorimeter for studying thermal properties of biological macromolecules in dilute solution. Anal. Biochem., 232, 79–85.CrossRefGoogle ScholarPubMed
Privalov, P. L. (1980). Scanning microcalorimeters for studying macromolecules. Pure & Appl. Chem., 52, 479–497.CrossRefGoogle Scholar
Privalov, P. L. (1982). Stability of proteins. Proteins which do not present a single cooperative system. Adv. Protein Chem., 35, 1–104.CrossRefGoogle ScholarPubMed
Privalov, P. L., and Khechinashvili, N. N. (1974). A thermodynamic approach to the problem of stabilization of globular protein structure: a calorimetric study. J. Mol. Biol., 86, 665–684.CrossRefGoogle ScholarPubMed
Privalov, P. L., and Makhatadze, G. I. (1990). Heat capacity of proteins. II. Partial molar heat capacity of the unfolded polypeptide chain of proteins: protein unfolding effects. J. Mol. Biol., 213, 385–391.CrossRefGoogle ScholarPubMed
Privalov, P. L., and Makhatadze, G. I. (1992). Contribution of hydration and non-covalent interactions to the heat capacity effect on protein unfolding. J. Mol. Biol., 224, 715–723.CrossRefGoogle ScholarPubMed
Purcell, E. M. (1977). Life at low Reynolds number. Am. J. Phys., 45, 311.CrossRefGoogle Scholar
Ramakrishnan, V., and Moore, P. B. (2001). Atomic structures at last: the ribosome in 2000. Curr. Opinion in Str. Biol., 11, 144–154.CrossRefGoogle ScholarPubMed
Rau, D. C., and Bloomfield, V. A. (1979) Transient electric birefringence of T7 virial DNA. Biopolymers, 18, 2783–2805.CrossRefGoogle Scholar
Rayment, I. (1996). Kinesin and myosin: molecular motor with similar engine. Structure, 4, 501–504.CrossRefGoogle Scholar
Records, M. T. Jr., Woodbury, C. P., and Inman, R. B. (1975). Characterization of rodlike DNA fragments. Biopolymers, 14, 393–408.Google Scholar
Rhee, K. H., Scarborough, G. A., et al. (2002). Domain movements of plasma membrane H(+)-ATPase: 3D structures of two states by electron cryo-microscopy. Embo. J., 21, 3582–3589.CrossRefGoogle ScholarPubMed
Richard, S. B., Madern, D., et al. (2000). Halophilic adaptation: novel solvent protein interactions observed in the 2.9 and 2.6 A resolution structures of the wild type and a mutant of malate dehydrogenase from Haloarcula marismortui. Biochemistry, 39, 992–1000.CrossRefGoogle Scholar
Rief, M., Fernandez, J. M., and Gaub, H. E. (1998). Elastically coupled two-level system as a model for biopolymer extensibility, Phys. Rev. Letters, 81, 4764–4767.CrossRefGoogle Scholar
Rief, M., Oesterhelt, T. F., Heymann, B., and Gaub, H. E. (1997). Single molecule force spectroscopy on polysaccharides by atomic force microscopy. Science, 275, 1295–1297.CrossRefGoogle ScholarPubMed
Rief, M., Pascual, J.Saraste, M., and Gaub, H. E. (1999). Single molecule force spectroscopy of spectrin repeats: low unfolding forces in helix bundles. J. Mol. Biol., 286, 553–561.CrossRefGoogle ScholarPubMed
Riek, R., Hornemann, S., Wider, G., Glockshuber, R., and Wuthrich, K. (1997). NMR characterization of the full-length recombinant murine prion protein, mPrP(23–231). FEBS Lett., 413, 282–288.CrossRefGoogle Scholar
Roberts, M. M., Coker, A. R., et al. (1999). Crystallization, X-ray diffraction and preliminary structure analysis of Mycobacterium tuberculosis chaperonin 10. Acta Crystallogr. D Biol. Crystallogr., 55 (Pt 4), 910–914.CrossRefGoogle ScholarPubMed
Rosenheck, K., and Doty, P. (1961). The far ultraviolet absorption spectra of polypeptide and protein solutions and their dependence on conformation. Proc. Natl. Acad. Sci. USA, 47, 1775–1785.CrossRefGoogle ScholarPubMed
Rosenthal, P. B., and Henderson, R. (2003). Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol., 333, 721–745.CrossRefGoogle ScholarPubMed
Rostom, A. A., and Robinson, C. V. (1999). Disassembly of intact multiprotein complexes in the gas phase. Curr. Opin. Struct. Biol., 9, 135–141.CrossRefGoogle ScholarPubMed
Rostom, A., Fucini, P., et al. (2000). Detection and selective dissociation of intact ribosomes in a mass spectrometer. Proc. Natl. Acad. Sci. USA, 97, 5185–5190.CrossRefGoogle Scholar
Rould, M. A. (1997). Screening for heavy-atom derivatives and obtaining accurate isomorphous differences. Methods in Enzym, 276 461–472.CrossRefGoogle ScholarPubMed
Rould, M. A., and Carter, C. W. Jr. (2003). Isomorphous difference methods. Methods in Enzym., 374, 145–63.CrossRefGoogle ScholarPubMed
Rowe, A. (1977). The concentration dependence of transport processes: a general description applicable to the sedimentation, translational diffusion, and viscosity coefficients of macromolecular solutes. Biopolymers, 16, 2595–2611.CrossRefGoogle Scholar
Rubtsov, I. V., Wang, J., and Hochstrasser, R. M. (2003). Dual-frequency 2D-IR spectroscopy heterodyned photon echo of the peptide bond, PNAS, 5601–5606.CrossRefGoogle ScholarPubMed
Salmeen, I., Rimai, L., Liebes, L., Rich, M. A., and McCormick, J. J. (1975). Hydrodynamic diameters of RNA tumor viruses. Studies by laser beat frequency light scattering spectroscopy of avian myeblastosos and Rauscher murine leukemia viruses. Biochemistry, 14, 134–141.CrossRefGoogle ScholarPubMed
Sanders, C. R., Hare, B. J., Howard, K. P., Prestegard, J. H. (1994). Magnetically oriented phospholipid micelles as a tool for the study of membrane-associated molecules. Prog. NMR Spectros., 26, 5.CrossRefGoogle Scholar
Schachman, H. K. (1989). Analytical ultracentrifugation reborn. Nature, 341, 259–260.CrossRefGoogle Scholar
Schmidt, B., and Reisner, D. (1992). A fluorescence detection system for analytical ultracentrifuge and its application to proteins, nucleic acids, viroid and viruses. In Analytical Ultracentrifugation in Biochemistry and Polymer Science, eds. Harding, S. E., Rowe, A. J. and Horton, J. C.. Cambridge: Royal Society of Chemistry.Google Scholar
Schmidt, T., Schultz, G. J., Baumgartner, W., Gruber, H. J., Schindler, H. (1996). Imaging of single molecule diffusion. PNAS, 93, 2926–2929.CrossRefGoogle ScholarPubMed
Schmitz, K. S. (1993). An Introduction to Dynamic Light Scattering of Molecules. Boston: Academic Press.Google Scholar
Schmitz, K. S., and Schurr, J. M. (1973). Rotational relaxation of macromolecules determined by dynamic light scattering. II. Temperature dependence for DNA. Biopolymers, 12, 1543–1564.CrossRefGoogle ScholarPubMed
Scholtan, W., and Lange, H. (1972). Bestimmung der Teilchegrobenverteilung von Latices mit der Ultracentrifuge. Kolloid-Z., u. Z. Polimere., 250, 782–796.CrossRefGoogle Scholar
Schouten, S., Hopmans, E. C., et al. (2000). Widespread occurrence of structurally diverse tetraether membrane lipids: evidence for the ubiquitous presence of low-temperature relatives of hyperthermophiles. Proc. Natl Acad. Sci. USA, 97, 14421–14426.CrossRefGoogle ScholarPubMed
Schrader, M., Bahlmann, K., Giese, G., and Hell, S. W. (1998). 4Pi-confocal imaging in fixed biological specimens. Biophys. J., 75, 1659–1668.CrossRefGoogle ScholarPubMed
Schuck, P. (2004). A model for sedimentation in inhomogeneous media. I. Dynamic density gradients from sedimenting co-solutes. Biophys. Chem., 108, 187–200.CrossRefGoogle ScholarPubMed
Schultz, D. A. (2003). Plasmon resonant particles for biological detection. Curr. Opin. Biotechnol., 14, 13–22.CrossRefGoogle ScholarPubMed
Schwabe, J. W. R., Chapman, L., Finch, J. T, Rhodes, D., and Neuhaus, D. (1993). DNA recognition by the oestrogen receptor: from solution to the crystal. Structure, 1, 187–204.CrossRefGoogle ScholarPubMed
Schwille, P., Meyer-Almes, F. J., et al. (1997). Dual-colour fluorescence cross-correlation spectroscopy for multicomponent diffusional analysis in solution. Biophys. J., 72, 1878–1886.CrossRefGoogle Scholar
Seelig, J. (2004). Thermodynamics of lipid–peptide interactions. Biochim. Biophys. Acta., 1666, 40–50.CrossRefGoogle ScholarPubMed
Seils, J., and Dorfmuller, Th. (1991). Internal dynamics of linear and superhelical DNA as studied by photon correlation spectroscopy. Biopolymers, 31, 813–825.CrossRefGoogle ScholarPubMed
Serdyuk, I. N., Grenader, A. K., et al. (1979). Study of the internal structure of Escherichia coli ribosomes by neutron and x-ray scattering. J. Mol. Biol., 135, 691–707.CrossRefGoogle ScholarPubMed
Serdyuk, I. N., Pavlov, M., et al. (1994). The triple isotopic substitution method in small angle neutron scattering. Application to the study of the ternary complex EF-Tu.GTP.aminoacyl-tRNA. Biophys. Chem., 53, 123–130.CrossRefGoogle Scholar
Serdyuk, I., Ulitin, A., et al. (1999). Structure of a beheaded 30S ribosomal subunit from Thermus thermophilus. J. Mol. Biol., 292, 633–639.CrossRefGoogle Scholar
Sheetz, M. P., Turne, Y. S., Qian, H., and Elson, E. L. (1989). Nanometre level analysis demonstrates that lipid flow does not drive membrane glycoprotein movements. Nature, 340, 284–285.CrossRefGoogle Scholar
Sheraga, H. A., and Mandelkern, L. (1953). Consideration of the hydrodynamic properties of proteins. J. Am. Chem. Soc., 75, 179–184.CrossRefGoogle Scholar
Sheraga, H. A., Edsall, J. T., and Gadd, J. O. (1951). Double refraction of flow: numerical evaluation of extinction angle and birefringence as a function of velocity gradient. J. Chem. Phys., 19, 1101–1108.CrossRefGoogle Scholar
Shiku, H., and Dunn, R. C. (1999). Near field scanning optical microscopy. Anal. Chem., 71, 23A–29A.CrossRefGoogle ScholarPubMed
Shingyoji, C.Higuchi, H., Yoshimura, M., Katayama, E., and Yanagida, T. (1998). Dynein arms are oscillating force generators. Nature, 393, 711–714.CrossRefGoogle ScholarPubMed
Shiryaev, V. M., Selivanova, O. M., Hartsch, T., Nazimov, I. V., and Spirin, A. S. (2002). Ribosomal protein S1 from Thermus thermophilus: its detection, identification and overproduction. FEBS Letters, 525, 88–92.CrossRefGoogle ScholarPubMed
Shivashankar, G. V., and Livchaber, A. (1997). Single DNA molecule grafting and manipulation using a combined atomic force microscope and an optical tweezer. Appl. Phys. Lett., 71, 3727–3729.CrossRefGoogle Scholar
Sigler, P. B., Xu, Z., et al. (1998). Structure and function in GroEL-mediated protein folding. Ann. Rev. Biochem., 67, 581–608.CrossRefGoogle ScholarPubMed
Simpson, A. A., Tao, Y., et al. (2000). Structure of the bacteriophage Φ29 DNA packaging motor. Nature, 408, 745–750.CrossRefGoogle ScholarPubMed
Skoog, D. A., Holler, F. J., and Nieman, T. A. (1995). Principle of Instrumental Analysis. Philadelphia: Saunders College Publishing.Google Scholar
Sliz, P., Harrison, S. C., and Rosenbaum, G. (2003). How does radiation damage in protein crystals depend on X-ray dose? Structure (Camb), 11, 13–19.CrossRefGoogle ScholarPubMed
Smith, D. E., Babcock, H. P., and Chu, S. (1999). Single-polymer dynamics in steady shear flow. Science, 283, 1724–1727.CrossRefGoogle ScholarPubMed
Smith, D. E., Tans, S. J., et al. (2001). The bacteriophage Φ29 portal motor can package DNA against a large internal force. Nature, 413, 748–752.CrossRefGoogle ScholarPubMed
Smith, L. J., Redfield, C., et al. (1994). Comparison of four independently determined structure of human recombinant interleikin-4. Struct. Biol., 1, 301–310.CrossRefGoogle Scholar
Smith, M. H. (1970). Molecular weight of proteins and some other materials including sedimentation diffusion and frictional coefficients and partial specific volumes. In Handbook of Biochemistry. Selected Data for Molecular Biology, ed. Sober, H. A., pp. C3–C47. Cleveland, OH: The Chemical Rubber Company.Google Scholar
Smith, R. D., Bruce, J. E., et al. (1996). The role of Fourier transform ion cyclotron resonance mass spectrometry in biological research – new development and applications. In Mass Spectrometry in the Biological Science, eds. Burlingame, A. L. and Carr, S. A., pp. 25–68. Totowa, NJ: Humana Press.Google Scholar
Smith, S. B., Cui, Y., and Bustamante, C. (1996). Overstretching B-DNA: the elastic response of individual double-stranded and single-stranded DNA molecules. Science, 271, 795–799.CrossRefGoogle ScholarPubMed
Smith, S. O., and Peersen, O. B. (1992). Solid-state NMR approaches for studying membrane protein structure. Ann. Rev. Biophys. Biomol. Str., 21, 25–47.CrossRefGoogle ScholarPubMed
Snatzke, G. (1994). Circular dichroism: an introduction. In Circular Dichroism: Principles and Applications., eds, Nakanishi, K., Berova, N. and Woody, R. W.. New York, VCH Publishers.Google Scholar
Sodano, P., Chary, K. V., et al. (1991). Nuclear magnetic resonance studies of recombinant Escherichia coli glutaredoxin. Sequence-specific assignments and secondary structure determination of the oxidised form. Eur. J. Biochem, 200, 369–377.CrossRefGoogle Scholar
Sober, H. A. (ed.) (1970). Handbook of Biochemistry, 2nd edn. Cleveland: CRC Press.Google Scholar
Sorlie, S. S., and Pecora, R. (1988). A dynamic light scattering study of a 2311 base pair DNA restriction fragment. Macromolecules, 21, 1437–1441.CrossRefGoogle Scholar
Sosa, H., Dias, D. P., et al. (1997). A model for the microtubule-Ncd motor protein complex obtained by cryo-electron microscopy and image analysis. Cell, 90, 217–224.CrossRefGoogle ScholarPubMed
Spirin, A. S. (1963). Some Problems of Macromolecular Structure of Ribonucleic Acids (in Russian). Moscow: Academy of Science USSR.Google Scholar
Spirin, A. S. (2000). Ribosomes. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Spiro, T. G., and Chernuszevich, R. S. (1995). Resonance Raman spectroscopy of metalloprotein. Methods Enzymol., 246, 416–459.CrossRefGoogle Scholar
Spolar, R. S., and Record, M. T. Jr. (1994). Coupling of local folding to site-specific binding of proteins to DNA. Science, 263, 777–784.CrossRefGoogle Scholar
Sportsman, J. R. (2003). Fluorescence anisotropy in pharmacologic screening. Methods Enzymol., 361, 505–529.CrossRefGoogle ScholarPubMed
Spronk, C. A. E. M., Linge, J. P., Hilbers, C. W., and Vuister, G. W. (2002). Improving the quality of protein structures derived by NMR spectroscopy, J. Biomolecular NMR, 22, 281–289.CrossRefGoogle ScholarPubMed
Spudlish, J. A. (1994). How molecular motors works. Nature, 372, 515–518.Google Scholar
Squire, P. G. (1970). An equation of consistency relating the harmonic mean relaxation time to sedimentation data. Biochim. Biophys. Acta, 221, 425–429.CrossRefGoogle ScholarPubMed
Srajer, V., Ren, Z., et al. (2001). Protein conformational relaxation and ligand migration in myoglobin: a nanosecond to millisecond molecular movie from time-resolved Laue X-ray diffraction. Biochemistry, 40, 13802–13815.CrossRefGoogle ScholarPubMed
Sreerama, N., and Woody, R. W. (2003). Structural composition of betaI- and betaII-proteins. Protein Sci, 12, 384–388.CrossRefGoogle ScholarPubMed
Stafford, W. F., and Braswell, E. H. (2004). Sedimentation velocity, multi-speed method for analyzing polydisperse solutions. Biophys. Chem., 108, 273–279.CrossRefGoogle ScholarPubMed
Steely, H. T. Jr., Gray, D. M., and Lang, D. (1986a). Study of the circular dichroism of bacteriophage φ6 and φ6 nucleocapsid. Biopolymers 25, 171–188.CrossRefGoogle Scholar
Steely, H. T. Jr., Gray, D. M., Lang, D., and Maestre, M. F. (1986b). Circular dichroism of double-stranded RNA in the presence of salt and ethanol. Biopolymers 25, 91–117.CrossRefGoogle Scholar
Stejskal, E. O., and Tanner, J. E. (1965). Spin diffusion measurements: Spin echoes in the presence of a time dependent field gradient. J. Chem. Phys, 42, 288–292.CrossRefGoogle Scholar
Stellwagen, N. C. (1981). Electric birefringence of restriction enzyme fragments of DNA: Optical factor and electric polarizability as a function of molecular weight. Biopolymers, 20, 399–434.CrossRefGoogle ScholarPubMed
Stellwagen, N. C. (1996). Electric birefringence of kilobase-sized DNA molecules. Biophys. Chem., 58, 117–124.CrossRefGoogle ScholarPubMed
Stellwagen, N. C., Gelfi, C., and Righetti, P. G. (1997). The free solution mobility of DNA. Biopolymers, 42, 687–703.3.0.CO;2-Q>CrossRefGoogle ScholarPubMed
Stoeckli, M., Chaurand, P., Hallahan, D. E., and Caprioli, R. (2001). Imaging mass spectrometry: A new technology for the analysis of protein expression in mamalian tissues. Nat. Medicine, 7, 493–496.CrossRefGoogle Scholar
Stoeckli, M., Chaurand, P., et al. (2001). Imaging mass spectrometry: a new technology for the analysis of protein expression in mammalian tissues. Nat. Med., 7, 493–96.CrossRefGoogle ScholarPubMed
Stone, M. J., Fairbrother, W. J., et al. (1992). Backbone dynamics of the Bacillus subtilis glucose permease IIA domain determined from 15N NMR relaxation measurements. Biochemistry, 31, 4394–4406.CrossRefGoogle ScholarPubMed
Strick, T. R., Allemand, J. -F., Bensimon, D., Bensimon, A., Croquettet, V. (1996). The elasticity of a single supercoiled DNA molecule. Science, 271, 1835–1837.CrossRefGoogle ScholarPubMed
Strick, T., Allemand, J. -F., Croquete, V., and Bensimon, D. (2000). Twisting and stretching single DNA molecules. Progr. Biophys. Mol. Biol., 74, 115–140.CrossRefGoogle ScholarPubMed
Stroebel, D., Choquet, Y., Popot, J. L., and Picot, D. (2003). An atypical haem in the cytochrome b(6)f complex. Nature, 426, 413–418.CrossRefGoogle ScholarPubMed
Stryer, L. (1968). Fluorescence spectroscopy of proteins. Science, 162, 526–533.CrossRefGoogle ScholarPubMed
Stryer, L., and Hougland, R. P. (1967). Energy transfer: A spectroscopic ruler. PNAS, 58, 719–726.CrossRefGoogle ScholarPubMed
Stuhrmann, H. B., and Miller, A. (1978). Small-angle scattering of biological structures. J. Appl. Cryst., 11, 325–345.CrossRefGoogle Scholar
Subramanian, S., and Henderson, R. (1999). Electron crystallography of bacteriorhodopsin with millisecond time resolution. J. Struct. Biol., 128, 19–25.CrossRefGoogle Scholar
Subramanian, S., Hirai, T., and Henderson, R. (2002). From structure to mechanism: electron crystallographic studies of bacteriorhodopsin. Phil. Trans. A Math. Phys. Eng. Sci., 360, 859–874.CrossRefGoogle Scholar
Subramaniam, S., and Henderson, R. (2000). Molecular mechanism of vectorial proton translocation by bacteriorhodopsin. Nature, 10, 653–657.CrossRefGoogle Scholar
Subramaniam, S., Lindahl, M., et al. (1999). Protein conformational changes in the bacteriorhodopsin photocycle. J. Mol. Biol., 287, 145–1461.CrossRefGoogle ScholarPubMed
Surewicz, W. K., Mantsch, H. H., et al. (1993). Determination of protein secondary structure by Fourier transform infrared spectroscopy: a critical assessment. Biochemistry, 32, 389–394.CrossRefGoogle ScholarPubMed
Susi, H., and Byler, D. M. (1986). Resolution-enhanced Fourier transform infrared spectroscopy of enzymes. Meth. Enzymol., 130, 290–311.CrossRefGoogle ScholarPubMed
Suzuki, Y., Yasunaga, T., Ohkura, R., Wakabayashi, T., and Sutoh, K. (1998). Swing of the lever arm of a myosin motor at the isomerization and phosphate-release steps. Nature, 396, 380–383.CrossRefGoogle ScholarPubMed
Svergun, D. I. (1999). Restoring low resolution structure of biological macromolecules from solution scattering using simulated annealing. Biophys. J., 76, 2879–2886.CrossRefGoogle ScholarPubMed
Svergun, D. I. (2000). Advanced solution scattering data analysis methods and their applications. J. Appl. Crystallog., 33, 530–534.CrossRefGoogle Scholar
Svergun, D. I., Barberato, C., et al. (1995). CRYSOL – a program to evalate X-ray solution scattering of biological macromolecules from atomic coordinates. J. Appl. Crystallogr., 28, 768–773.CrossRefGoogle Scholar
Svergun, D. I., Malfois, M., et al. (2000). Low resolution structure of the sigma54 transcription factor revealed by X-ray solution scattering. J. Biol. Chem., 275, 4210–4214.CrossRefGoogle ScholarPubMed
Svergun, D. I., Richard, S., et al. (1998). Protein hydration in solution: experimental observation by x-ray and neutron scattering. Proc. Natl. Acad. Sci. USA, 95, 2267–2272.CrossRefGoogle ScholarPubMed
Svergun, D. I., Volkov, V. V., et al. (1996). New developments in direct shape determination from small-angle scattering 2. Uniqueness. Acta. Crystallog., A 52, 419–426.CrossRefGoogle Scholar
Taillandier, E., and Liquier, J. (1992). Infrared spectroscopy of DNA. Meth. Enzymol., 211, 307–335.CrossRefGoogle ScholarPubMed
Takahashi, H., Nakanishi, T., Kami, K., Arata, Y., and Shimada, I. (2000). A novel NMR method for determining the interfaces of large protein–protein complexes. Natur. Struct. Biol., 7, 220–223.Google ScholarPubMed
Tamm, L. K. (1993). Total internal reflectance florescence microscopy in optical microscopy. In Emerging Methods and Applications, eds. Herman, and Lemasters, . Academic Press.Google Scholar
Tanford, C. (1968). Protein denaturation. Adv. Protein Chem., 23, 121–282.CrossRefGoogle ScholarPubMed
Tanford, C., Kawahara, K., and Lapanje, S. (1967). Proteins as random coils. I. Intrinsic viscosities and sedimentation coefficients in concentrated guanidine hydrochloride. J. Am. Chem. Soc., 89, 729–736.CrossRefGoogle Scholar
Tardieu, A., Vachette, P., et al. (1981). Biological macromolecules in solvents of variable density: characterization by sedimentation equilibrium, densimetry, and X-ray forward scattering and an application to the 50S ribosomal subunit from Escherichia coli. Biochemistry, 20, 4399–4406.CrossRefGoogle ScholarPubMed
Tarek, M., and Tobias, D. J. (2002). Single-particle and collective dynamics of protein hydration water: a molecular dynamics study. Phys. Rev. Lett., 89, 275501.CrossRefGoogle ScholarPubMed
Tcien, R. Y., and Miyawaki, A. (1998). Seeing the machinary of live cells. Science, 280, 1954–1955.Google Scholar
Tehei, M., Franzetti, B., et al. (2004). Adaptation to extreme environments: macromolecular dynamics in bacteria compared in vivo by neutron scattering. EMBO Rep., 5, 66–70.CrossRefGoogle ScholarPubMed
Tehei, M., Madern, D., et al. (2001). Fast dynamics of halophilic malate dehydrogenase and BSA measured by neutron scattering under various solvent conditions influencing protein stability. Proc. Natl. Acad. Sci. USA, 98, 14356–1461.CrossRefGoogle ScholarPubMed
Tenford, C. (1965). Physical Chemistry of Macromolecules. New York: John Wiley and Sons.Google Scholar
Thalhammer, S., Stark, R. W., Müller, S., Weinberg, J., and Heck, W. M. (1997). J. Struct. Biol., 119, 232–237.CrossRef
Thomas, G. J., and Tsuboi, M. (1993). Raman spectroscopy of nucleic acids and their complexes. Adv. Biophys. Chem., 3, 1–69.Google Scholar
Thompson, D. S., and Gill, S. J. (1967). Polymer relaxation times from birefringence relaxation measurements. J. Chem. Phys., 47, 5008–5017.CrossRefGoogle ScholarPubMed
Tian, F., Al-Hashimi, H. M., Craighead, J. L., and Prestegard, J. H. (2001). Conformational analysis of a flexible oligosaccharide using residual dipolar coupling, JACS, 123, 485–492.CrossRefGoogle Scholar
Tinoco, I. Jr., Sauer, K., and Wang, J. C., Physical Chemistry. Principles and Applications in Biological Science. Prentice Hall. New Jersey (1998).Google Scholar
Tirado, M. M., Martinez, C. L., and Garcia de la Torre, J. (1984). Comparison of theories for the translational and rotational diffusion coefficients of rod-like macromolecules. Application to short DNA fragments. J. Chem. Phys. 81, 2047–2051.CrossRefGoogle Scholar
Tjandra, N., Tate, S., Ono, A., Kainosho, M., and Bax, A. (2000). The NMR structure of a DNA dodecamer in an aqueous dilute liquid crystalline phase. JACS, 122, 6190–6200.CrossRefGoogle Scholar
Tolbert, T. J. and Williamson, J. R. (1996). Preparation of specifically deuterated RNA for NMR studies using a combination of chemical and enzymatic synthesis. JACS, 118, 7929–7940.CrossRefGoogle Scholar
Tong, L., and Rossmann, M. G. (1997). Rotation function calculations with GLRF program. Methods in Enzymology, 276, 594–611.CrossRefGoogle ScholarPubMed
Toyoshima, C., Sasabe, H., and Stokes, D. L. (1993). Three-dimensional cryo-electron microscopy of the calcium ion pump in the sarcoplasmic reticulum membrane. Nature, 362, 469–471.CrossRefGoogle ScholarPubMed
Tristram-Nagle, S., and Nagle, J. F. (2004). Lipid bilayers: thermodynamics, structure, fluctuations, and interactions. Chem. Phys. Lipids, 127(1), 3–14.CrossRefGoogle ScholarPubMed
Tsvetkov, V. N. (1989). Rigid-chain Polymers. Hydrodynamic and Optical Properties in Solution. Consultants Bureau, New York and London.Google Scholar
Tsvetkov, V. N., Eskin, V. E., and Frenkel, S. Y. (1971). Structure of Macromolecules in Solution (translated from Russian), V. 1. Chapter 7. National Lending Library for Science and Technology, Boston, UK.Google Scholar
Ulitin, A. B., Agalarov, S. C., and Serdyuk, I. N. (1997). Preparation of a “beheaded” derivative of the 30S ribosomal subunit. Biochimie, 79, 523–526.CrossRefGoogle ScholarPubMed
Unger, K. K., Huber, M., Walhagen, K., Hennessy, T. P., and Hearn, M. T. W. (2002). A critical appraisal of capillary electrochromatography. Anal. Chem., 74, 200A–207A.CrossRefGoogle ScholarPubMed
Vagin, A., and Teplyakov, A. (2000). An approach to multi-copy search in molecular replacement. Acta Crystallogr. D Biol. Crystallogr., 56, 1622–1624.CrossRefGoogle ScholarPubMed
Vale, R. D. (1996). Switches, latches, and amplifier: common themes of proteins and molecular motors. J. Cell Biol., 135, 291–302.CrossRefGoogle Scholar
Valle, M., Zavialov, A., et al. (2003). Locking and unlocking of ribosomal motions. Cell, 114, 123–34.CrossRefGoogle ScholarPubMed
Groot, F. G., Gonzàlez-Mañas, J. M., Lakey, J. H. and Pattus, F. (1991). A ‘molten globule’ membrane-insertion intermediate of the pore-forming domain of colicin A. Nature, 354, 408–410.Google Scholar
Heel, M., Gowen, B., et al. (2000). Single-particle electron cryo-microscopy: towards atomic resolution. Q. Rev. Biophys., 33, 307–369.CrossRefGoogle ScholarPubMed
Holde, (1985). Physical Biochemistry. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Varki, A. (1999). Essentials of Glycobiology. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Váró, G., and Lanyi, J. K. (1991). Effects of the crystalline structure of purple membrane on the kinetics and energetics of the bacteriorhodopsin photocycle. Biochemistry, 30, 7165–7171.CrossRefGoogle ScholarPubMed
Velazquez-Campoy, A., Kiso, Y., et al. (2001). The binding energetics of first- and second-generation HIV-1 protease inhibitors: implications for drug design. Arch. Biochem. Biophys., 390, 169–175.CrossRefGoogle ScholarPubMed
Velazquez-Campoy, A., Leavitt, S. A., et al. (2004). Characterization of protein-protein interactions by isothermal titration calorimetry. Methods Mol. Biol., 261, 35–54.Google ScholarPubMed
Venyaminov, S. Y. and Vassilenko, K. S. (1994). Determination of protein tertiary structure class from circular dichroism spectra. Anal. Biochem., 222, 176–184.CrossRefGoogle ScholarPubMed
Vénien-Bryan, C., and Fuller, S. D. (1994). The organization of the spike complex of Semliki Forest virus. J. Mol. Biol., 236, 572–583.CrossRefGoogle ScholarPubMed
Vysotski, E. S., Liu, Z. J., Rose, J., Wang, B. C., and Lee, J. (1999). Preparation and preliminary study of crystals of the recombinant calcium regulated photoprotein obelin from the bioluminescent hydroid Obelia longissima. Acta Crystallogr., D55, 1965–1966.Google Scholar
Wakia, S. (1971). Slow motion in shear flow of a doublet of two spheres in contact. J. Phys. Soc. Jpn, 31, 1581–1587.CrossRefGoogle Scholar
Walker, J. M. (2000). Electrophoretic techniques. In Principles and Techniques of Practical Biochemistry, eds. Wilson, K. and Walker, J.. Cambridge: Cambridge University Press.Google Scholar
Walther, D., Cohen, F. E., and Doniak, S. (2000). Reconstruction of low-resolution three dimensional density maps from one dimensional small-angle X-ray solution scattering data for biomolecules. J. Appl. Crystallognr., 33, 350–363.CrossRefGoogle Scholar
Wand, A. J., Ehrhardt, M. R., and Flynn, P. F. (1998). High-resolution NMR of encapsulated proteins dissolved in low-viscosity fluids. PNAS, 95, 15299–15302.CrossRefGoogle ScholarPubMed
Wang, K., Forbes, J. G., and Jin, A. J. (2001). Single molecule measurements of titin elasticity. Progr. Biophys. Mol. Biol., 77, 1–44.CrossRefGoogle ScholarPubMed
Wang, M. (1999). Manipulation of single molecules in biology. Curr. Opin. Biotech., 10, 81–86.CrossRefGoogle ScholarPubMed
Wang, M. D., Schnitzer, M. J., et al. (1998). Force and velosity measured for single molecules of RNA polymerase. Science, 282, 902–907.CrossRefGoogle Scholar
Ward, T. J. (1994). Chiral media for capillary electrophoresis. Anal. Chem., 66, 633A–640A.Google ScholarPubMed
Watson, J. D., and Crick, F. H. (1953). Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature, 171, 737–738.CrossRefGoogle ScholarPubMed
Webb, R. H. (1996). Confocal optical microscopy. Rep. Progr. Phys., 59, 427–471.CrossRefGoogle Scholar
Weber, G. (1953). Rotational Brownian motion and polarization of the fluorescence of solutions. Adv. Protein Chem., 8, 415–459.CrossRefGoogle ScholarPubMed
Weber, K., and Osborn, M. (1969). The reliability of molecular weight determinations by dodecyl sulfate-polyacrylamide gel electrophoresis. J. Biol. Chem., 244, 4406–4412.Google ScholarPubMed
Weber, P. C., and Salemme, F. R. (2003). Applications of calorimetric methods to drug discovery and the study of protein interactions. Curr. Opin. Struct. Biol., 13, 115–121.CrossRefGoogle Scholar
Weik, M. (2003). Low-temperature behavior of water confined by biological macromolecules and its relation to protein dynamics. Eur. Phys. J. E. Soft Matter, 12, 153–158.CrossRefGoogle ScholarPubMed
Weik, M., Lehnert, U., and Zaccai, G. (2005). Liquid-like water confined in stacks of biological membranes at 200 K and its relation to protein dynamics. Biophys. J., 89, 3639–3646.CrossRefGoogle ScholarPubMed
Weik, M., Ravelli, R. B., et al. (2000). Specific chemical and structural damage to proteins produced by synchrotron radiation. Proc. Natl. Acad. Sci. USA, 97, 623–628.CrossRefGoogle ScholarPubMed
Weiskopf, A. S., Vouros, P., and Harvey, D. (1997). Characterization of oligosaccaride composition and structure by quadropole ion trap mass spectrometry. Rapid Commun. Mass Spectr., 11, 1493–1504.3.0.CO;2-1>CrossRefGoogle Scholar
Weiss, S. (1999). Fluorescence spectroscopy of single biomolecules. Science, 283, 1676–1683.CrossRefGoogle ScholarPubMed
Weiss, S. (2000). Shattering the diffraction limit of light: a revolution in fluorescence microscopy? Proc. Nat. Acad. Sci. USA, 97, 8747–8749.CrossRefGoogle ScholarPubMed
Weissman, M., Schindler, H., and Feher, G. (1976). Determination of molecular weights by fluctuation spectroscopy: application to DNA. Proc. Nat. Acad. Sci. USA, 73, 2776–2780.CrossRefGoogle Scholar
Westhof, E., Dumas, P., and Moras, D. (1988). Acta Cryst., A44, 122–123.
Wetlaufer, D. B. (1962). Ultraviolet spectra of proteins and nucleic acids. Adv. Prot. Chem., 17, 303–390.CrossRefGoogle Scholar
Wider, G., and Wuthrich, K. (1999). NMR spectroscopy of large molecules and multimolecular assemblies in solution. Curr. Opin. Struct. Biol., 9, 594–601.CrossRefGoogle ScholarPubMed
Wilkinson, S. R., and Thurston, G. B. (1976). The optical birefringence of DNA solutions induced by oscillatory electric and hydrodynamic fields. Biopolymers, 15, 1555–1572.CrossRefGoogle ScholarPubMed
Willcox, B. E., Gao, G. F., et al. (1999). TCR binding to peptide-MHC stabilizes a flexible recognition interface. Immunity, 10, 357–65.CrossRefGoogle ScholarPubMed
Williams, K. R., and King, R. W. (1990a). The Fourier transform in chemistry – NMR. Part 3. Multiple-pulse experiments. J. Chem. Educ., 67, A93–A99.CrossRefGoogle Scholar
Williams, K. R., and King, R. W. (1990b). The Fourier transform in chemistry – NMR. Part 4. Two-dimensional methods. J. Chem. Educ., 67, A125–A137.CrossRefGoogle Scholar
Wilson, K., and Walker, J. M. (2000). Principles and Techniques of Practical Biochemistry. Cambridge: Cambridge University Press.Google Scholar
Wimberly, B. T., Brodersen, D. E., et al. (2000). Structure of the 30S ribosomal subunit. Nature, 407, 327–339.Google ScholarPubMed
Winston, R. L., and Fitzgerald, M. C. (1997). Mass spectrometry as a readout of protein structure and function. Mass Spectr. Rev., 16, 165–179.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Woese, C. R. (1967). The Genetic Code; The Molecular Basis for Genetic Expression. New York: Harper & Row.Google Scholar
Wuthrich, K. (1986). NMR of Proteins and Nucleic Acids, New York: Wiley-Interscience.Google Scholar
Wuthrich, K. (1995). NMR – this other method for protein and nucleic acid structure determination. Acta Cryst., D51, 249–270.Google Scholar
Wuthrich, K. (2000). Protein recognition by NMR, Nat. Struct. Biol., 7, 188–189.CrossRefGoogle Scholar
Wuthrich, K. (2001). The way to NMR structures of protein. Nat. Struct. Biol., 8, 923–925.CrossRefGoogle Scholar
Wuthrich, K., Wider, G., Wagner, G., and Braun, W. (1982). Sequential resonance assignments as a basis for determination of spatial protein structures by high resolution proton nuclear magnetic resonance. J. Mol. Biol., 155, 311–319.CrossRefGoogle ScholarPubMed
Wyer, J. R., Willcox, B. E., et al. (1999). T cell receptor and coreceptor CD8 alphaalpha bind peptide-MHC independently and with distinct kinetics. Immunity, 10, 219–225.CrossRefGoogle ScholarPubMed
Xie, X. S., and Dunn, R. C. (1994). Probing single molecule dynamics. Science, 265, 361–364.CrossRefGoogle ScholarPubMed
Xu, H.-X., and Yeung, E. S. (1997). Direct measurement of single-molecule diffusion and photodecomposition in free solution. Science, 275, 1106–1109.CrossRefGoogle ScholarPubMed
Xu, Z., Horwich, A. L., and Sigler, P. B. (1997). Nature, 388, 741–750.CrossRef
Xue, Q., and Yeung, E. S. (1995). Difference in the chemical reactivity of individual molecules of an enzyme. Nature, 373, 681–683.CrossRefGoogle Scholar
Yan, Y., Winograd, E., et al. (1993). Crystal structure of the repetitive segments of spectrin. Science, 262, 2027–2030.CrossRefGoogle ScholarPubMed
Yanagida, T., Kitamura, K., Tanaka, H., Iwane, A. H., and Esaki, S. (2000). Single molecule analysis of the actomyosin motor. Curr. Opin. Cell. Biol., 12, 20–25.CrossRefGoogle ScholarPubMed
Yguerabide, J., Epstein, H. F., and Stryer, L. (1970). Segmental flexibility in an antibody molecule. J. Mol. Biol., 51, 573–590.CrossRefGoogle Scholar
Yonath, A., Miissig, J., et al. (1980). Crystallization of the large ribosomal subunit from B. stearothermophilus. Biochem. Int., 1, 428.Google Scholar
Yoshida, K., Yoshimoto, M., et al. (1998). Fabrication of new substrate for atomic force microscopic observation of DNA molecules from an ultrasmooth sapphire plate. Biophys. J., 74, 1654–1657.CrossRefGoogle ScholarPubMed
Yoshizaki, T., and Yamakawa, H. (1980). Dynamics of spheroid–cylindrical molecules in dilute solution. J. Chem. Phys., 72, 57–69.CrossRefGoogle Scholar
Zaccai, G. (2000). How soft is a protein? A protein dynamics force constant measured by neutron scattering. Science, 288, 1604–1607.CrossRefGoogle ScholarPubMed
Zaccai, G., Morin, P., et al. (1979). Interactions of yeast valyl-tRNA synthetase with RNAs and conformational changes of the enzyme. J. Mol. Biol., 129, 483–500.CrossRefGoogle ScholarPubMed
Zanni, M., and Hochstrasser, R. (2001). Two-dimensional infrared spectroscopy: a promising new method for the time resolution of structure. Curr. Opin. Struct. Biol., 11, 516–562.CrossRefGoogle Scholar
Zheng, R., Zheng, X., Dong, J., and Carey, P. R. (2004). Proteins can convert to β-sheet in single crystals. Protein Science, 13, 1288–1294.CrossRefGoogle ScholarPubMed
Zhou, H.-X. (1995). Calculation of translational friction and intrinsic viscosity. I. General formulation for arbitrarily shaped particles. Biophys. J., 69, 2286–2297.CrossRefGoogle ScholarPubMed
Zhou, H.-X. (2001). A unified picture of protein hydration: prediction of hydrodynamic properties from known structures. Biophys. Chem. 93, 171–179.CrossRefGoogle ScholarPubMed
Zhuang, X., Bartley, L. E., et al. (2000). A single-molecule study of RNA catalysis and folding. Science, 288, 2048–2051.CrossRefGoogle ScholarPubMed
Zidek, L., Stefl, R., and Sklenar, V. (2001). NMR methodology for the stydy of nucleic acids. Curr. Opin. Struct. Biol., 11, 275–281.CrossRefGoogle Scholar
Zipper, P., and Durchshlag, H. (2000). Prediction of hydrodynamic and small angle scattering parameters from crystal and electron microscopic structure. J. Appl. Cryst., 33, 788–792.CrossRefGoogle Scholar
Zlatanova, J., Lindsay, S. M., and Leuba, A. H. (2000). Single molecule force spectroscopy in biology using the atomic force microscope. Progr. Biophys. Mol. Biol., 74, 37–61.CrossRefGoogle ScholarPubMed
Agalarov, S. C., and Williamson, J. R. (2000). A hierarchy of RNA subdomains in assembly of the central domain of the 30S ribosomal subunit. RNA, 6, 402–408.CrossRefGoogle Scholar
Agalarov, S. C., Selivanova, O. M., et al. (1999). Independent in vitro assembly of all three major morphological parts of the 30S ribosomal subunit of Thermus thermophilus. Eur. J. Biochem., 266, 533–537.CrossRefGoogle ScholarPubMed
Agalarov, S. C., Sheleznyakova, E. N., et al. (1998). In vitro assembly of a ribonucleoprotein particle corresponding to the platform domain of the 30S ribosomal subunit. Proc. Natl. Acad. Sci. USA, 95, 999–1003.CrossRefGoogle ScholarPubMed
Allemand, J.-F., Bensimon, D., Lavery, R., and Croquette, V. (1998). Stretched and overwound DNA forms a Pouling-like structure with exposed base. Proc. Natl. Acad. Sci. USA, 95, 14152–14157.CrossRefGoogle Scholar
Allison, S. A. (1999). Low Reynolds number transport properties of axisymmetric particles employing stick and slip boundary conditions. Macromolecules, 32, 5304–5312.CrossRefGoogle Scholar
Allison, S. A. (2001). Boundary element modelling of biomolecular transport. Biophys. Chem., 93, 197–213.CrossRefGoogle Scholar
Altieri, A. S., Hinton, D., and Byrd, R. A. (1995). Association of biomolecular system via pulse field gradient NMR self-diffusion measurements. JACS, 117, 7566–7567.CrossRefGoogle Scholar
Altose, M. D., Zheng, Y., Dong, J., Palfey, B. A., and Carey, P. R. (2001). Comparing protein–ligand interactions in solution and single crystals by Raman spectroscopy. PNAS, 98, 3006–3011.CrossRefGoogle ScholarPubMed
Bacia, K., and Schwille, P. (2003). A dynamic view of cellular processes by in vivo fluorescence auto- and cross-correlation spectroscopy. Methods, 29, 74–85.CrossRefGoogle ScholarPubMed
Bailey, B., Farkas, D. L., Taylor, D. L., and Lanni, F. (1993). Enhancement of axial resolution in fluorescence microscopy by standing-wave excitation. Nature, 366, 44–48.CrossRefGoogle ScholarPubMed
Ban, N., Nissen, P., et al. (1999). Placement of protein and RNA structures into a 5 å-resolution map of the 50S ribosomal subunit. Nature, 400(6747), 841–847.CrossRefGoogle ScholarPubMed
Banachowicz, E., Gapinski, J., and Patkowski, A. (2000). Solution structure of biopolymers: A new method of constructing a bead model. Biophys. J., 78, 70–78.CrossRefGoogle ScholarPubMed
Bandecar, J. (1992). Amide modes and protein conformation. Biochim. Biophys. Acta, 1120, 123–143.CrossRefGoogle Scholar
Barnes, W. L., Dereux, A., and Ebbesen, T. W. (2003). Surface plasmon subwavelength optics. Nature, 424(6950), 824–830.CrossRefGoogle ScholarPubMed
Barron, L. D., Hecht, L., Blanch, E. W., and Bell, A. F. (2000). Solution structure and dynamics of biomolecules from Raman optical activity. Prog. Biophys. and Mol. Biol., 73, 1–49.CrossRefGoogle ScholarPubMed
Basavappa, R., and Sigler, P. B. (1991). EMBO J., 10, 3105–3111.
Bastiaens, P. I. H., and Pepperkok, R. (2000). Observing proteins in their natural habitat: the living cell. TIBS, 25, 631–636.Google ScholarPubMed
Baumann, C. G., Bloomfield, V. A., Smith, S. B., Bustamante, C., Wang, M. D., and Block, S. M. (2000). Stretching of single collapsed DNA molecules. Biophys. J., 78, 1965–1978.CrossRefGoogle ScholarPubMed
Bax, A. (2003). Weak alignment offers new NMR opportunities to study protein structure and dynamics. Protein Sci., 12, 1–16.CrossRefGoogle ScholarPubMed
Belke, J., and Ristau, O. (1997). Analysis of interacting biopolymer systems by analytical centrifugation. Eur. Biophys. J., 25, 325–332.CrossRefGoogle Scholar
Bellissent-Funel, M. C., Zanotti, J. M., et al. (1996). Slow dynamics of water molecules on the surface of globular proteins. Faraday Discuss., 103, 281–294.CrossRefGoogle Scholar
Belov, M. E., Gorshkov, M. V., Udeseth, H. R., Anderson, G. A., and Smith, R. D. (2000). Zeptomole-sensititivity electrospray ionization – Fourier transform ion cyclotron resonance mass spectrometry proteins. Anal. Chem., 72, 2271–2279.CrossRefGoogle ScholarPubMed
Benner, W. H. (1997). A gated electrostatic ion trap to repetitiously measure the charge and m/z of large electrospray ions. Anal. Chem., 69, 4162–4168.CrossRefGoogle Scholar
Bennett, M. J., and Eisenberg, D. (1994). Refined structure of monomeric diphtheria toxin at 2.3 å resolution. Protein Sci., 3(9), 1464–1475.CrossRefGoogle ScholarPubMed
Bennink, M. L., Leuba, S. H., Leno, G. H., Zlatanova, J., Grooth, B. G., and Greve, J. (2001). Unfolding individual nucleosomes by stretching single chromatin fibers with optical tweezers. Nature Str. Biol, 8, 606–610.CrossRefGoogle ScholarPubMed
Berg, H. (1983). Random Walks in Biology. Princeton: Princeton University Press.Google Scholar
Bergethon, P. R. (1995). The Physical Basis of Biochemistry. The Foundation of Molecular Biophysics. New York: Springer.Google Scholar
Bernado, P., Garcia de la Torre, J., and Pons, M. (2002). Interpretation of 15N NMR relaxation data for globular proteins using hydrodynamic calculations with HYDRONMR. J. Biomol. NMR, 23, 139–150.CrossRefGoogle ScholarPubMed
Bernal, J. D., and Crowfoot, D. (1934). X-ray photographs of crystalline pepsin. Nature, 134, 794–795.CrossRefGoogle Scholar
Biemann, K. (1992). Mass spectrometry of peptides and proteins. Annu. Rev. Biochem., 61, 977–1010.CrossRefGoogle ScholarPubMed
Bischler, N., Brino, L., et al. (2002). Localization of the yeast RNA polymerase I-specific subunits. Embo. J., 21(15), 4136–4144.CrossRefGoogle ScholarPubMed
Bjorkman, P. J., Saper, M. A., et al. (1987). Structure of the human class I histocompatibility antigen, HLA-A2. Nature, 329(6139), 506–512.CrossRefGoogle ScholarPubMed
Blattner, F. R. (1997). The complete genome sequence of Eshcherichia coli K-12. Science, 277, 1453–1474.CrossRefGoogle Scholar
Block, S. M., Blair, D. F., and Berg, H. C. (1989). Compliance of bacterial flagella measured wih optical tweezers. Nature, 338, 514–518.CrossRefGoogle Scholar
Bon, C., Dianoux, A. J., et al. (2002). A model for water motion in crystals of lysozyme based on an incoherent quasielastic neutron-scattering study. Biophys. J., 83(3), 1578–1588.CrossRefGoogle ScholarPubMed
Bon, C., Lehmann, M. S., et al. (1990). Quasi Laue neutron-diffraction study of the water arrangement in crystals of triclinic hen egg-white lysozyme. Acta Crystallogr. D, 55, 978–987.CrossRefGoogle Scholar
Booth, D. R., Sunde, M., et al. (1997). Instability, unfolding and aggregation of human lysozyme variants underlying amyloid fibrillogenesis. Nature, 385, 787–793.CrossRefGoogle ScholarPubMed
Bottcher, B., Tsuji, N., et al. (1998). Peptides that block hepatitis B virus assembly: analysis by cryomicroscopy, mutagenesis and transfection. Embo. J., 17(23), 6839–6845.CrossRefGoogle ScholarPubMed
Bowie, J. U., Luthy, R., and Eisenberg, D. (1991). A method to identify protein sequences that fold into a known three-dimensional structure. Science, 253(5016), 164–170.CrossRefGoogle ScholarPubMed
Braiman, M. S., and Rothschild, K. J. (1988). Fourier transform infrared techniques for probing membrane protein structure. Ann. Rev. Biophys. Biophysical Chem., 17, 541–570.CrossRefGoogle ScholarPubMed
Brändén, C.-I., and Tooze, J. (1999). Introduction to Protein Structure. New York: Garland Pub.Google Scholar
Brant, D. A. (1999). Novel approaches to the analysis of polysaccharide structure. Curr. Opin. Struct. Biol., 9, 556–562.CrossRefGoogle Scholar
Brey, W. S. (ed.) (1988). Pulse Methods in 1D and 2D Liquid-Phase NMR. San-Diego: Academic.Google Scholar
Brooks, B. R., Bruccoleri, R. E., et al. (1983). CHARMM: A program for macromolecular empirical energy modelling. J. Comp. Chem., 4, 187–230.CrossRefGoogle Scholar
Brower-Toland, B. R., Smith, C. L., Yeh, R. S., Lis, J. T., Peterson, C. L., and Wang, M. D. (2002). Mechanical disruption of individual nucleosomes reveals a reversible multistage release of DNA. Proc. Natl. Acad. Sci. USA, 99, 1960–1966.CrossRefGoogle ScholarPubMed
Brudler, R., Rammelsberg, R., et al. (2001). Structure of the I1 early intermediate of photoactive yellow protein by FTIR spectroscopy. Nat. Struct. Biol., 8(3), 265–270.CrossRefGoogle ScholarPubMed
Brune, D., and Kim, S. (1993). Predicting protein diffusion coefficients. J. Am. Chem. Soc., 90, 3835–3839.Google ScholarPubMed
Brunger, A. T., and Adams, P. D. (2002). Molecular dynamics applied to X-ray structure refinement. Acc. Chem. Res., 35(6), 404–412.CrossRefGoogle ScholarPubMed
Brunger, A. T., Adams, P. D., et al. (1998). Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallogr. D Biol. Crystallogr., 54 (Pt 5), 905–921.CrossRefGoogle ScholarPubMed
Buchanan, M. V., and Hettich, R. L. (1993). Fourier transform mass spectrometry of high mass molecules. Anal. Chem., 65, 245A–259A.CrossRefGoogle Scholar
Burlingame, A. L., and Carr, S. A. (1996). Mass Spectrometry in the Biological Sciences. Totowa: Humana Press.CrossRefGoogle Scholar
Burlingame, A. L., Carr, S. A., et al. (2000). Mass Spectrometry in Biology and Medicine. Totowa: Humana Press.CrossRefGoogle Scholar
Burton, A., and Sinsheimer, R. L. (1965). The process of infection with bacteriophage X174. VII. Ultracentrifugal analysis of the replicative form. J. Mol. Biol., 14, 327–347.CrossRefGoogle ScholarPubMed
Bustamante, C., Erie, D. A., and Keller, D. (1994). Biochemical and structural applications of scanning force microscopy. Curr. Opin. Struct. Biol., 4, 750–760.CrossRefGoogle Scholar
Byron, O. (1997). Construction of hydrodynamic bead models from high resolution x-ray crystallographic or nuclear magnetic resonance data. Bioph. J., 72, 406–415.CrossRefGoogle ScholarPubMed
Caffrey, M. (2000). A lipid's eye view of membrane protein crystallization in mesophases. Curr. Opin. Struct. Biol., 10(4), 486–497.CrossRefGoogle ScholarPubMed
Cai, S., and Singh, B. R. (1999). Identification of beta-turn and random coil amide III infrared bands for secondary structure estimation of proteins. Biophys. Chem., 80(1), 7–20.CrossRefGoogle ScholarPubMed
Callender, R., and Deng, H. (1994). Nonresonance Raman difference spectroscopy: a general probe of protein structure, ligand binding, enzymatic catalysis, and the structures of other biomacromolecules. Annu. Rev. Biophys. Biomol. Struct., 23, 215–245.CrossRefGoogle ScholarPubMed
Callis, P. R., and Davidson, N. (1969). Hydrodynamic relaxation times of DNA from decay of flow dichroism measurements. Biopolymers, 8, 379–390.CrossRefGoogle Scholar
Campos-Olivas, R., Horr, I., Bormann, C., Jung, G., and Gronenborn, A. M. (2001). Solution structure, backbone dynamics and chitin binding of the anti-fungal protein from Streptomyces tendae TU901. J. Mol. Biol., 308, 765–782.CrossRefGoogle ScholarPubMed
Canet, D., Doering, K., Dobson, C. M., and Dupont, Y. (2001). High-sensitivity fluorescence anisotropy detection of protein-folding events: application to α-lactalbumin. Biophys. J., 80, 1996–2003.CrossRefGoogle ScholarPubMed
Cantor, C., and Schimmel, P. (1980). Biophysical Chemistry. Part II. Technique for the Study of Biological Structure and Function. San Francisco: W. H. Freeman and Company.Google Scholar
Caprioli, R. M., and Suter, M. J. F. (1995). Mass spectrometry. In Introduction to Biophysical Methods for Protein and Nucleic Acid Research. San Diego: Academic Press.Google Scholar
Carr, S. A., and Burlingame, A. L. (1996). The meaning and usage of the terms monoisotopic mass, average mass, mass resolution, and mass accuracy for measurements of biomolecules. In Mass Spectrometry in the Biological Sciences, eds. Burlingame, A. L. and Carr, S. A., pp. 546–552. Totowa: Humana Press.Google Scholar
Carra, J. H., Murphy, E. C., et al. (1996). Thermodynamic effects of mutations on the denaturation of T4 lysozyme. Biophys. J., 71(4), 1994–2001.CrossRefGoogle ScholarPubMed
Carrasco, B., and Garcia de la Torre, J. (1999). Hydrodynamic properties of rigid particles: comparison of different modelling and computational procedure. Biophys. J., 75, 3044–30572.CrossRefGoogle Scholar
Carrion-Vazquez, M., Overhauser, A. F., et al. (2000). Mechanical design of proteins studied by single-molecule force spectroscopy and protein engineering. Prog. Biophys. Mol. Biol., 74, 63–91.CrossRefGoogle ScholarPubMed
Castro, A., Fairfield, F. R., and Shera, E. B. (1993). Fluorescence detection and size measurement of single DNA molecules. Anal. Chem., 65, 849–852.CrossRefGoogle Scholar
Cate, J. H., Yusupov, M. M., et al. (1999). X-ray crystal structures of 70S ribosome functional complexes. Science, 285(5436), 2095–2104.CrossRefGoogle ScholarPubMed
Chacon, P., Moran, F., et al. (1998). Low-resolution structures of proteins in solution retrieved from X-ray scattering with a genetic algorithm. Biophys. J., 74(6), 2760–75.CrossRefGoogle ScholarPubMed
Chacon, P., Diaz, J. F., et al. (2000). Reconstruction of protein form with X-ray solution scattering and a genetic algorithm. J. Mol. Biol., 299(5), 1289–1302.CrossRefGoogle Scholar
Charney, E., Chen, H.-H., and Rau, D. (1991). The flexibility of A-form DNA. J. Biolmol. Struct. Dyn., 9, 353–362.CrossRefGoogle ScholarPubMed
Che, Z., N. Olson, H., et al. (1998). Antibody-mediated neutralization of human rhinovirus 14 explored by means of cryoelectron microscopy and X-ray crystallography of virus–Fab complexes. J. Virol., 72(6), 4610–4622.Google ScholarPubMed
Checovich, W. J., Bolger, R. E., and Burke, T. (1995). Fluorescence polarization – a new tool for cell and molecular biology. Nature, 375, 254–256.CrossRefGoogle ScholarPubMed
Chen, X., Wu, H., Mao, C., and Whitesides, G. M. (2002). A prototype two-dimensional capillary electrophoresis system fabricated in poly(dimethylsiloxane). Anal. Chem., 74, 1772–1778.CrossRefGoogle Scholar
Cherry, R. J., and Schneider, G. (1976). A spectroscopic technique for measuring slow rotational diffusion of macromolecules. 2: Determination of rotational correlation times of protein in solution. Biochemistry, 15, 3657–3661.CrossRefGoogle Scholar
Chervenka, , 1969. A Manual of Methods for the Analytical Ultracentrifuge. Palo Alto: Spinco Division, Beckman Instruments.Google Scholar
Chong, B. E., Lubman, D. M., et al. (1999). Rapid screening of protein profiles of human breast cancer cell lines using non-porous reversed-phase high performance liquid chromatography separation with matrix-assisted laser desorption/ionization time-of-flight mass spectral analysis. Rapid. Commun. Mass Spectr., 13(18), 1808–1812.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Chong, B. E., Lubman, D. M., Rosenspire, A., and Miller, F. (1998). Protein profiles and identification of high performance liquid chromatography isolated proteins of cancer cell lines using matrix-asisted laser desorption/ionization time-of-flight mass spectrometry. Rapid Commun. Mass Spectr., 12, 1986–1993.3.0.CO;2-H>CrossRefGoogle Scholar
Clore, G. M., and Gronenborn, A. M. (1991). Two-, three-, and four-dimensional NMR methods for obtaining more precise three-dimensional structure of proteins in solution. Ann. Rev. Biophys. Chem., 20, 29–63.CrossRefGoogle Scholar
Cluzel, P., Lebrun, A., Heller, C., Lavery, R., Viovy, J.-L., Chatenay, D., Caron, F. (1996). DNA: an extensible molecule. Science, 271, 792–794.CrossRefGoogle Scholar
Cohn, E. J., and Edsall, J. T. (1965). In Proteins, Amino Acids and Peptides as Ions and Dipolar Ions, eds. Cohn, E. J. and Edsall, J. T., pp. 370–381. New York: Hafner Publ. Co.Google Scholar
Collins, K. D. (1997). Charge density-dependent strength of hydration and biological structure. Biophys. J., 72(1), 65–76.CrossRefGoogle ScholarPubMed
Colon, L. A., Guo, Y., and Fermier, A. (1997). Capillary electrochromatography. Anal. Chem. News, 69, 461A–467A.CrossRefGoogle Scholar
Cordone, L., Ferrand, M., et al. (1999). Harmonic behavior of trehalose-coated carbon-monoxy-myoglobin at high temperature. Biophys. J., 76, 1043–1047.CrossRefGoogle ScholarPubMed
Crichton, R. R., Engelman, D. M., et al. (1977). Contrast variation study of specifically deuterated Escherichia coli ribosomal subunits. Proc. Natl. Acad. Sci. USA, 74, 5547–5550.CrossRefGoogle ScholarPubMed
Crick, F. H. (1968). The origin of the genetic code. J. Mol. Biol., 38, 367–379.CrossRefGoogle ScholarPubMed
Crothers, D. M., and Zimm, B. H. (1965) Viscosity and sedimentation of the DNA from bacteriophages T2 and T7 and the relation to molecular weight. J. Mol. Biol., 12, 527–536.CrossRefGoogle ScholarPubMed
Dasgupta, S., and Spiro, T. G. (1980). Resonance Raman characterization of the 7-ns photoproduct of (carbonmonoxy) hemoglobin: implications for hemoglobin dynamics. Biochemistry, 25, 5941–5948.CrossRefGoogle Scholar
Davidson, I. W., and Secrest, W. L. (1972). Determination of chromium in biological materials by atomic absorption spectrometry using a graphite furnace atomizer. Anal. Chem., 44(13), 1808–1813.CrossRefGoogle ScholarPubMed
Torre, G. (2001) Hydration from hydrodynamics. General consideration and applications to bead modelling to globular proteins. Biophys. Chem., 93, 159–170.CrossRefGoogle Scholar
Dekker, N. H., Rybenkov, V. V., et al. (2002). The mechanism of type IA topoisomerases. Proc. Natl. Acad. Sci. USA, 99, 12126–12131.CrossRefGoogle ScholarPubMed
Delano, W. L., and Brunger, A. T. (1995). The direct rotation function: rotational Patterson correlation search applied to molecular replacement. Acta Cryst. D, 51, 740–748.CrossRefGoogle Scholar
Derome, A. E. (1987). Modern NMR Techniques for Chemistry Research. New York: Pergamon.Google Scholar
Dessen, P., Blanquet, S., Zaccai, G., and Jacrot, B. (1978). Antico-operative binding of initiator transfer RNAMet to methionyl-transfer RNA synthetase from Escherichia coli: neutron scattering studies. J. Mol. Biol., 126, 293–313.CrossRefGoogle ScholarPubMed
Dickerson, R., and Geiss, I. (1969). The Structure and Action of Proteins. Menlo Park: Benjamin Cummings.Google Scholar
Diehl, M., Doster, W., et al. (1997). Water-coupled low-frequency modes of myoglobin and lysozyme observed by inelastic neutron scattering. Biophys. J., 73, 2726–32.CrossRefGoogle ScholarPubMed
Dobo, A., and Kaltashov, I. A. (2001). Detection of multiple protein conformational ensembles in solution via deconvolution of charge-state distribution in ESI MS. Anal. Chem., 73, 4763–4773.CrossRefGoogle Scholar
Dolgikh, D. A., Gilmanshin, R. I., et al. (1981). Alpha-lactalbumin: compact state with fluctuating tertiary structure? FEBS Lett., 136, 311–315.CrossRefGoogle ScholarPubMed
Dong, J., Wan, Z., Popov, M., Carey, P. R., and Weiss, M. A. (2003). Insulin assembly damps conformational fluctuations: Raman analysis of amide I linewidths in native states and fibrils. JMB, 330, 431–442.CrossRefGoogle ScholarPubMed
Doster, W., Cusack, S., et al. (1989). Dynamical transition of myoglobin revealed by inelastic neutron scattering. Nature, 337, 754–756.CrossRefGoogle ScholarPubMed
Doty, P., Bradbury, J. H., and Holtzer, A. M. (1956). Polypeptides. IV. The molecular weight, configuration and association of poly-γ-glutamate in various solvents. J. Am. Chem. Soc., 78, 947–954.CrossRefGoogle Scholar
Dubin, S. B., Clark, N. A., and Benedek, G. B. (1971). Measurement of the rotational diffusion coefficient of lysozyme by depolarised light scattering: configuration of lysozyme in solution. J. Chem. Phys., 54, 5158–5164.CrossRefGoogle Scholar
Dunkerk, A. K., and Williams, R. W. (1979). Ultraviolet and LASER Raman investigation of the buried tyrosines in fd phage. J. Biol. Chem., 254, 6446.Google Scholar
Dutta, R. K., Hammons, K., Willibey, B., and Haney, M. A. (1991). Analysis of protein denaturation by high-performance continuous differential viscometry. J. Cromatog., 536, 113–121.CrossRefGoogle ScholarPubMed
Dykxhoorn, D. M., Novina, C. D., et al. (2003). Killing the messenger: short RNAs that silence gene expression. Nat. Rev. Mol. Cell Biol., 4, 457–467.CrossRefGoogle ScholarPubMed
Eastman, J. E., Taguchi, A. K., et al. (2000). Characterization of a Rhodobacter capsulatus reaction center mutant that enhances the distinction between spectral forms of the initial electron donor. Biochemistry, 39, 14787–14798.CrossRefGoogle ScholarPubMed
Eden, D., Luu, B. Q., Zapata, D. J., Sablin, E. P., and Kuul, F. J. (1995). Solution structure of two molecular motor domains: nonclaret disjunctional and kinesin. Biophys. J., 68, 59s–65s.Google ScholarPubMed
Eigen, M., and Rigler, R. (1994). Sorting single molecules: applications to diagnostic and evolutionary biotechnology. Proc. Natl. Acad. Sci. USA, 91, 5740–5747.CrossRefGoogle Scholar
Eimer, W., and Pecora, R. (1991). Rotational and translational diffusion of short rodlike molecules in solution: oligonucleotides. J. Chem. Phys., 94, 2324–2329.CrossRefGoogle Scholar
Eimer, W., Williamson, J. R., Boxer, S. G., and Pecora, R. (1990). Characterization of the overall and internal dynamics of short oligonucleotides by depolarised dynamic light scattering and NMR relaxation measurements. Biochemistry, 29, 799–811.CrossRefGoogle Scholar
Eisenberg, H. (1981). Forward scattering of light, X-rays and neutrons. Q. Rev. Biophys., 14, 141–172.CrossRefGoogle ScholarPubMed
Elias, J. G., and Eden, D. (1981). Transient electric birefringence study of the persistence length and electrical polarizability of restriction fragments of DNA. Macromolecules, 14, 410–419.CrossRefGoogle Scholar
Elöve, G. A., Chaffotte, A. F., et al. (1992). Early steps in cytochrome c folding probed by time-resolved circular dichroism and fluorescence spectroscopy. Biochemistry, 31, 6876–6883.CrossRefGoogle ScholarPubMed
Engel, A., Lyubchenko, Y., and Muller, D. (1999). Atomic force microscopy: a powerful tool to observe biomolecules at work. Trends Cell Biol., 9, 77–80.CrossRefGoogle ScholarPubMed
Ernst, R. R., Bodenhausen, G., and Wokaun, A. (1987). Principles of Nuclear Magnetic Resonance in One and Two Dimensions. Oxford: Oxford University Press.Google Scholar
Essavaz-Roulet, B., Bockelman, U., and Heslot, F. (1997). Mechanical separation of the complementary strands of DNA. PNAS, 94, 11935–11940.CrossRefGoogle Scholar
Fan, E., Merritt, E. A., et al. (2000). AB(5) toxins: structures and inhibitor design. Curr. Opin. Struct. Biol., 10, 680–686.CrossRefGoogle ScholarPubMed
Fasman, G. D. (1996). Circular Dichroism and the Conformational Analysis of Biomolecules. New York: Plenum Press.CrossRefGoogle Scholar
Ferrer, M. L., Duchowicz, R., Carrasco, B., Garcia de la Torre, J., and Acuna, A. U. (2001). The conformation of serum albumin in solution: a combined phosphorescence depolarization–hydrodynamic modelling study. Biophys. J., 80, 2422–2430.CrossRefGoogle Scholar
Feynman, R. P., Leighton, R. B., et al. (1963). The Feynman Lectures on Physics. Reading, MA: Addison-Wesley Pub. Co.Google Scholar
Fisher, T. E., Marszalek, P. E., and Fernandez, J. M. (2000). Stretching single molecules into novel conformations using the atomic force microscopy. Nature Stuctr. Biol., 7, 719–724.Google Scholar
Flaux, J., Bertelsen, E. B., Horwich, A. L., and Wuthrich, K. (2002). NMR analysis of a 900K GroEL–GroES complex. Nature, 418, 207–211.Google Scholar
Flynn, P. F., and Wand, A. J. (2001). High-resolution nuclear magnetic resonance of Encapsulated proteins dissolved in low viscosity fluids, Methods Enzymol., 339, 54–70.CrossRefGoogle ScholarPubMed
Franklin, S. E., and Gosling, R. G. (1953). Molecular configuration in sodium thymonucleate. Nature, 171, 740–741.CrossRefGoogle ScholarPubMed
Franks, F., Gent, M., et al. (1963). Solubility of benzene in water. J. Chem. Soc., 8, 2716–2723.CrossRefGoogle Scholar
Franzen, S., and Boxer, S. G. (1997). On the origin of heme absorption band shifts and associated protein structural relaxation in myoglobin following flash photolysis. J. Biol. Chem., 272, 9655–9660.CrossRefGoogle ScholarPubMed
Frauenfelder, H., Parak, F., et al. (1988). Conformational substates in proteins. Ann. Rev. Biophys. Chem., 17, 451–479.CrossRefGoogle ScholarPubMed
Freifelder, D. (1970). Molecular weights of coliphages and coliphage DNA. IV, Molecular weights of DNA from bacteriophages T4, T5, and T7 and general problem of determination of molecular weight. J. Mol. Biol., 54, 567.CrossRefGoogle Scholar
Freire, E. (1995). Thermal denaturation methods in the study of protein folding. Methods Enzymol., 259, 144–168.CrossRefGoogle Scholar
Frey, W., Schief, W. R. Jr., et al. (1996). Two-dimensional protein crystallization via metal-ion coordination by naturally occurring surface histidines. Proc. Natl. Acad. Sci. USA, 93, 4937–4941.CrossRefGoogle ScholarPubMed
Frohn, J. T., Knapp, H. F., and Stemmer, A. (2000). True optical resolution beyond the Rayleigh limit achieved by standing wave illumination. Proc. Natl. Acad. Sci. USA, 97, 7232–7236.CrossRefGoogle ScholarPubMed
Fuller, S. D., and Argos, P. (1987). Is Sindbis a simple picornavirus with an envelope? Embo. J., 6, 1099–1105.Google ScholarPubMed
Fuller, W., Forsyth, T., et al. (2004). Water-DNA interactions as studied by X-ray and neutron fibre diffraction. Phil. Trans. R. Soc. Lond B Biol. Sci., 359, 1237–1247; discussion 1247–1248.CrossRefGoogle ScholarPubMed
Gabel, F., Bicout, D., et al. (2002). Protein dynamics studied by neutron scattering. Q. Rev. Biophys., 35, 327–367.CrossRefGoogle ScholarPubMed
Gabel, F. (2005). Protein dynamics in solution and powder measured by incoherent elastic neutron scattering: the influence of Q−range and energy resolution. Eur. Biophys. J., 34, 1–12.CrossRefGoogle ScholarPubMed
Gadola, S. D., Zaccai, N. R., et al. (2002). Structure of human CD1b with bound ligands at 2.3 å, a maze for alkyl chains. Nat. Immunol., 3, 721–726.CrossRefGoogle ScholarPubMed
Ganem, B. (1993). Detecting noncovalent interactions: new frontiers for mass spectrometry. Am. Biotechnol. Lab., 11, 32–34.Google ScholarPubMed
Ganem, B., Li, Y.-T., and Henion, J. D. (1991). Observation of noncovalent enzyme. Substrate and enzyme product complexes by ion spray mass spectrometry. J. Am. Chem. Soc., 113, 7818–7819.CrossRefGoogle Scholar
Garces-Chavez, V., McGloin, D., Melville, H., Sibbett, W., and Dholakia, K. (2002). Simultaneous micromanipulation in multiple planes using a self-reconstruction light beam. Nature, 419, 145–147.CrossRefGoogle Scholar
Garcia de la Torre, J. (2001). Hydration from hydrodynamics. General consideration and applications of bead modelling to globular proteins. Biophys. Chem., 93, 159–170.CrossRefGoogle Scholar
Garcia de la Torre, J., Martinez, M. C. L., and Tirado, M. M. (1984). Dimensions of short, rodlike macromolecules from translational and rotational diffusion coefficients. Study of the gramicidin dimer. Biopolymers, 23, 611–615.CrossRefGoogle Scholar
Garcia de la Torre, J., Navarro, S., and Lopez Martinez, M. C. (1994). Hydrodynamic properties of a double-helical model for DNA. Biophys. J., 66, 1573–1579.CrossRefGoogle ScholarPubMed
Garfin, D. E. (1995). Electrophoretic methods. In: Introduction to Biophysical Methods for Protein and Nucleic Acid Research, eds. Glaser, J. A. and Deutscher, M. P.. San Diego: Academic Press.Google Scholar
Garret, D. S., Seok, Y.-J., Liao, D.-I., Peterkofsky, A., Gronenborn, A. M., and Clore, G. M. (1997). Solution structure of the 30 kDa N-terminal domain of enzyme I of the Escherichia coli phosphoenolpuruvate: sugar phosphotransferase system by multidimensional NMR. Biochemistry, 36, 2517–2530.CrossRefGoogle Scholar
Gelles, J., and Landick, R. (1998). RNA polymerase as a molecular motor, Cell, 93, 13–16.CrossRefGoogle ScholarPubMed
Giege, R., Lorber, B., et al. (1982). Formation of a catalytically active complex between tRNAAsp and aspartyl-tRNA synthetase from yeast in high concentrations of ammonium sulphate. Biochimie, 64, 357–362.CrossRefGoogle ScholarPubMed
Gilbet, W. (1986). The RNA world. Nature, 319, 618.CrossRefGoogle Scholar
Gimzewski, J. K., and Joachum, C. (1999). Nanoscale science of single molecules using molecular probes. Science, 283, 1683–1688.CrossRefGoogle Scholar
Gluehmann, M., Zarivach, R., et al. (2001). Ribosomal crystallography: from poorly diffracting microcrystals to high-resolution structures. Methods, 25, 292–302.CrossRefGoogle ScholarPubMed
Go, N., Noguti, T., et al. (1983). Dynamics of a small globular protein in terms of low-frequency vibrational modes. Proc. Natl. Acad. Sci. USA, 80, 3696–3700.CrossRefGoogle ScholarPubMed
Godovach-Zimmermann, J., and Brown, L. R. (2001). Perspectives for mass spectrometry and functional proteomics. Mass Spectr. Rev., 20, 1–57.3.0.CO;2-J>CrossRefGoogle Scholar
Goldberg, D. E. (1989). Genetic Algorithms in Search, Optimization, and Machine Learning. Reading, MA: Addison-Wesley Pub. Co.Google Scholar
Gomez, J., Hilser, V. J., et al. (1995). The heat capacity of proteins. Proteins, 22, 404–412.CrossRefGoogle ScholarPubMed
Goodsell, D. S., and Olson, A. J. (1993). Soluble proteins: size, shape and function. TIBS, 18, 65–68.Google ScholarPubMed
Gordon, D. B. (2000). Mass spectrometric technique. In Principles and Techniques of Practical Biochemistry, 5th edn., eds. Wilson, K. and Walker, J.. Ch. 11. Cambridge: Cambridge University Press.Google Scholar
Greis, K. D., Hayes, B. K., et al. (1996). Selective detection and site-analysis of O-GlcNAc-modified glycopeptides by beta-elimination and tandem electrospray mass spectrometry. Anal. Biochem., 234, 38–49.CrossRefGoogle ScholarPubMed
Grier, D. (2003). A revolution in optical manipulation. Nature, 424, 810–816.CrossRefGoogle ScholarPubMed
Griko, Y. V., Freire, E., et al. (1995). The unfolding thermodynamics of c-type lysozymes: a calorimetric study of the heat denaturation of equine lysozyme. J. Mol. Biol., 252, 447–459.CrossRefGoogle ScholarPubMed
Griko, Y. V., Makhatadze, G. I., et al. (1994). Thermodynamics of barnase unfolding. Protein Sci., 3, 669–676.CrossRefGoogle ScholarPubMed
Gross, S. (2003). Application of optical traps in vivo. Methods in Enzymol.,V 361, 162–174.CrossRefGoogle ScholarPubMed
Grotjahn, L., Frank, R., and Blocker, H. (1982). Ultrafast sequencing of oligodeoxyribonucleotides by FAB-mass spectrometry. Nucl. Acids Res., 10, 4671–4677.CrossRefGoogle ScholarPubMed
Gutsche, I., Holzinger, J., et al. (2001). ATP-induced structural change of the thermosome is temperature-dependent. J. Struct. Biol., 135, 139–146.CrossRefGoogle ScholarPubMed
Haag, L., Garoff, H., Xing, L., Hammar, L., Kan, S. T., Cheng, R. H. (2002). Acid-induced movements in the glycoprotein shell of an alphavirus turn the spikes into membrane fusion mode. EMBO J., 21, 4402–4410.CrossRefGoogle ScholarPubMed
Hafner, J. H., Cheung, C.-L., Wooley, A. T., and Lieber, C. M. (2001). Structural and functional imaging with carbon nanotube AFM probes. Progr. Biophys. Mol. Biol., 77, 73–110.CrossRefGoogle ScholarPubMed
Hagerman, P. J. (1981). Investigation of the flexibility of DNA using transient electric birefringence. Biopolymers, 20, 1503–1535.CrossRefGoogle ScholarPubMed
Hagerman, P. J. (1985). Application of transient electric birefringence to the study of biopolymer structure. Methods Enzymol., 117, 199–215.Google Scholar
Hagerman, P. J. (2000). Transient electric birefringence for determining global conformations of non-helix elements and protein-induced bends in RNA. Methods Enzymol., 317, 440–453.CrossRefGoogle Scholar
Hahn, T. and International Union of Crystallography (2002). International Tables for Crystallography. Brief teaching edition of volume A, Space-group symmetry. Dordrecht; Boston, Published for the International Union of Crystallography by Kluwer Academic Publishers.Google Scholar
Hamm, P., Lim, M., and Hochstrasser, R. M. (1999). Structure of the amide I band of peptides measured by femtosecond non-linear-infrared spectroscopy. PNAS, 96, 6123–6128.Google Scholar
Han, W., Lindsay, S. M., Dlakic, M., and Harrington, R. E. (1997). Kinked DNA. Nature, 386, 563.CrossRefGoogle ScholarPubMed
Hansen, J. C., Lebowitz, J., and Demeler, B. (1994). Analytical ultracentrifugation of complex macromolecular systems. Biochemistry, 33, 13155–13163.CrossRefGoogle ScholarPubMed
Hansen, M. R., Mueller, L., and Pardi, A. (1998). Tunable alignment of macromolecules by filamentous phage yields dipolar coupling interaction. Nat. Struct. Biol., 5, 1065–1074.CrossRefGoogle Scholar
Harding, S. E. (1980). The combination of the viscosity increment with the harmonic mean rotational relaxation time for determining the conformation of biological macromolecules in solution. Biochem. J., 189, 359–361.CrossRefGoogle ScholarPubMed
Harding, S. E. (1981). A compound hydrodynamics shape function derived from viscosity and molecular covolume measurements. Int. J. Biol. Macromol., 3, 398–399.CrossRefGoogle Scholar
Harding, S. E. (1995). On the hydrodynamic analysis of macromolecular conformation. Biophys. Chem., 55, 69–93.CrossRefGoogle ScholarPubMed
Harding, S. E., and Rowe, A. (1982). Modelling biological macromolecules in solution: 1. The ellipsoid of revolution. Int. J. Biol. Macromol., 4, 160–164.CrossRefGoogle Scholar
Harding, S. E., Horton, J. C., and Colfen, H. (1997). The ELLIPS suite of macromolecular conformation algorithms. Eur. Biophys, J., 25, 347–359.CrossRefGoogle ScholarPubMed
Harpaz, Y., Gerstein, M., Chothia, C. (1994). Volume changes on protein folding. Structure, 2, 641–649.CrossRefGoogle ScholarPubMed
Harris, R. (1983). Nuclear Magnetic Resonance Spectroscopy. London: Pitman.Google Scholar
Haupts, U., Tittor, J., et al. (1997). General concept for ion translocation by halobacterial retinal proteins: the isomerization/switch/transfer (IST) model. Biochemistry, 36, 2–7.CrossRefGoogle ScholarPubMed
Haupts, U., Tittor, J., and Oesterhelt, D. (1999). Closing in on bacteriorhodopsin: progress in understanding the molecule. Ann. Rev. Biophys. Biomol. Struct., 28, 367–399.CrossRefGoogle ScholarPubMed
Hausten, E., and Schwille, P. (2003). Ultrasensitive investigations of biological systems by fluorescence correlation spectroscopy. Methods, 29, 153–166.CrossRefGoogle Scholar
Hazlett, T. L., Moore, K. J. M., Lowe, P. N., Jameson, D. M., and Eccleston, J. F. (1993). Solution of p21ras proteins bound with fluorescent nucleotides: a time-resolved fluorescence study. Biochemistry, 32, 13575–13583.CrossRefGoogle ScholarPubMed
Heberle, J., and Gensch, T. (2001). When FT-IR spectroscopy meets X-ray crystallography. Nat. Struct. Biol., 8, 195–197.CrossRefGoogle ScholarPubMed
Hellweg, T., Eimer, W., Krahn, E., Schneider, K., and Muller, A. (1997). Hydrodynamic properties of nitrogenase – the MoFe protein from Azotobacter vinelandii studied by dynamic light-scattering and hydrodynamic modelling. Biochim. Biophys. Acta, 1337, 311–318.CrossRefGoogle ScholarPubMed
Hensley, P. (1996). Defining the structure and stability of macromolecular assemblies in solution: the re-emergence of analytical ultracentrifugation as a practical tool. Structure, 4, 367–373.CrossRefGoogle ScholarPubMed
Hillisch, A., Lorenz, M., and Diekmann, S. (2001). Recent advances in FRET: distance determination in protein–DNA complexes. Curr. Opin. Struct. Biol., 11, 201–207.CrossRefGoogle ScholarPubMed
Hirao, I., and Ellington, A. D. (1995). Re-creating the RNA world. Curr. Biol., 5, 1017–1022.CrossRefGoogle ScholarPubMed
Homans, S.W., Edge, C. J., Ferguson, M. A., and Dwek, R. A. (1989). Solution structure of the glycosylphosphatidylinositol membrane anchor glycan of Trypanosoma bruccei variant surface glycoprotein. Biochemistry, 28, 2881–2887.CrossRefGoogle ScholarPubMed
Hore, P. J. (1995). Nuclear Magnetic Resonance. Oxford: Oxford University Press.Google Scholar
Horwitz, J., Strickland, E. H., and Billups, C. (1970). Analysis of the vibrational structure in the near-ultraviolet circular dichroism and absorption spectra of tyrosine derivatives and ribonuclease-A at 77 K. J. Am. Chem. Soc., 92, 2119–2129.CrossRefGoogle Scholar
Hu, C.-M., and Zwanzig, R. (1974). Rotational friction coefficients for spheroids with the slipping boundary conditions. J. Chem. Phys., 60, 4354–4357.CrossRefGoogle Scholar
Hunt, J. F., McCrea, P. D., et al. (1997). Assessment of the aggregation state of integral membrane proteins in reconstituted phospholipid vesicles using small angle neutron scattering. J. Mol. Biol., 273, 1004–1019.CrossRefGoogle ScholarPubMed
Hutchens, J. O. (1970). Handbook of Chemistry and Selected Data for Molecular Biology, ed. Sober, H. A.Cleveland, OH: Chemical Rubber Co; International Tables for Crystallography: Spzae Group Symmetry (2002). Dordrecht: Kluwer.Google Scholar
Huygens, C. (1690). Treatise on Light, New York: Dover (1962) of the English translation first published by Macmillan and Co. in 1912.Google Scholar
Ishijima, A., Kojima, H., et al. (1998). Simultaneous observation of individual ATPase and mechanical events by a single myosin molecule during interaction with actin. Cell, 92, 161–171.CrossRefGoogle ScholarPubMed
Ishima, R., and Torchia, D. (2000). Protein dynamics from NMR. Nat. Struct. Biol., 7, 740–743.CrossRefGoogle ScholarPubMed
Jacobs, R. E., Ahrens, E. T., Meade, T. J., and Fraser, S. E. (1999). Looking deeper into vertebrate development. TIBS, 9, 73–76.Google ScholarPubMed
Jacrot, B. (1976). The study of biological structures by neutron scattering from solution. Rep. Prog. Phys., 39, 911–953.CrossRefGoogle Scholar
Jacrot, B., and Zaccai, G. (1981). Determination of molecular weight by neutron scattering. Biopolymers, 20, 2414–2426.CrossRefGoogle Scholar
Jacrot, B., Chauvin, C., and Witz, J. (1977). Comparative neutron small-angle scattering study of small spherical RNA viruses. Nature, 266(5601), 417–421.CrossRefGoogle ScholarPubMed
Jaenicke, R. (2000). Do ultrastable proteins from hyperthermophiles have high or low conformational rigidity? Proc. Natl. Acad. Sci. USA, 97, 2962–2962.CrossRefGoogle ScholarPubMed
Jancarik, J., and Kim, S.-H. (1991). Sparse matrix sampling: a screening method for crystallization of proteins. J. Appl. Cryst., 24, 409–411.CrossRefGoogle Scholar
Jancarik, J., Scott, W. G., et al. (1991). Crystallization and preliminary X-ray diffraction study of the ligand-binding domain of the bacterial chemotaxis-mediating aspartate receptor of Salmonella typhimurium. J. Mol. Biol., 221, 31–34.CrossRefGoogle ScholarPubMed
Jeener, J. (1996). In Encyclopedia of Nuclear Magnetic Resonance, eds. Grant, D. M., Harris, R. K., Vol. 1, p. 40, Chichester: John Wiley and Sons.Google Scholar
Jeffrey, P. D., Nichol, L. W., Turner, D. R., and Winzor, D. J. (1977). The combination of molecular covolume and frictional coefficient to determine the shape and axial ratio of a rigid macromolecule. Studies on Ovalbumin. J. Phys. Chem., 81, 776–781.CrossRefGoogle Scholar
Jeruzalmi, D. and Steitz, T. A. (1997). Use of organic cosmotropic solutes to crystallize flexible proteins: application to T7 RNA polymerase and its complex with the inhibitor T7 lysozyme. J. Mol. Biol., 274, 748–756.CrossRefGoogle ScholarPubMed
Jia, Y., Sytnic, A., Li, L., Vladimirov, S., Cooperman, B. S., and Hochstrasser, R. M. (1997). Nonexponencial kinetics of a single tRNA Phe molecule under physiological conditions. Proc. Natl. Acad. Sci. USA, 94, 7932–7936.CrossRefGoogle ScholarPubMed
Jiang, Y., Ruta, V., et al. (2003). The principle of gating charge movement in a voltage-dependent K+ channel. Nature, 423, 42–48.CrossRefGoogle Scholar
Johnson, W. C. (1985). Circular Dichroism and Its Empirical Application to Biopolymers. Methods of Biochemical Analysis, Vol. 31, ed. Glick, D.. Chichester: John Wiley and Sons Inc.CrossRefGoogle ScholarPubMed
Johnson, K. H., and Gray, D. M. (1992). Analysis of an RNA pseudoknot structure by CD spectroscopy. J. Biomol. Struct. Dyn., 9, 733–745.CrossRefGoogle ScholarPubMed
Jolly, D., and Eisenberg, H. (1976). Photon correlation spectroscopy, total intensity light with laser radiation, and hydrodynamic studies of a well fractionated DNA sample. Biopolymers, 15, 61–95.CrossRefGoogle ScholarPubMed
Jones, J. A., Wilkins, D. K., Smith, L. J., and Dobson, C. M. (1997). Characterization of protein unfolding by NMR diffusion measurements. J. Biomolecular NMR, 10, 199–203.CrossRefGoogle Scholar
Jones, T. A., Zou, J. Y., et al. (1991). Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallogr. A, 47 (Pt 2), 110–119.CrossRefGoogle ScholarPubMed
Kabsch, W. (1988). Evaluation of single-crystal X-ray diffraction data from a position-sensitive detector. J. Appl. Cryst., 21, 916–924.CrossRefGoogle Scholar
Karger, B. L., Chu, Y.-H., and Foret, F. (1995). Capillary electrophoresis of proteins and nucleic acids. Annu. Rev. Biophys. Biomol. Struct., 24, 579–610.CrossRefGoogle ScholarPubMed
Kassas, S., Thomson, N. H., et al. (1997). Esherichia coli RNA polymerase activity observed using atomic force microscopy. Biochemistry, 36, 461–468.CrossRefGoogle Scholar
Kebbekus, P., Draper, D. E., and Hagerman, P. (1995). Persistence length of RNA. Biochemistry, 34, 4354–4357.CrossRefGoogle ScholarPubMed
Keiderling, T. A. (2002). Protein and peptide secondary structure and conformational determination with vibrational circular dichroism. Curr. Opin. Chem. Biol., 6, 682–688.CrossRefGoogle ScholarPubMed
Kincaid, J. R. (1995). Structure and dynamics of transient species using time-resolved resonance Raman spectroscopy. Methods Enzymol., 246, 460–501.CrossRefGoogle ScholarPubMed
King, R. W., and Williams, K. R. (1989a). The Fourier transform in chemistry. Part 1, Nuclear magnetic resonance: introduction. J. Chem. Education, 66, A213–A219.CrossRefGoogle Scholar
King, R. W., and Williams, K. R. (1989b). The Fourier transform in chemistry, Part 2, Nuclear magnetic resonance: the single pulse experiment. J. Chem. Education, 66, A243–A248.CrossRefGoogle Scholar
Kinosita, K. Jr., Yasuda, R., Noji, H., Ishiwata, S., and Yoshida, M. (1998). F1-ATPase: a rotary motor made of a single molecule. Cell, 93, 21–24.CrossRefGoogle ScholarPubMed
Klar, T. A., Jacobs, S., Dyba, M., Egner, A., and Hell, S. W. (2000). Fluorescence microscopy with diffraction resolution barrier broken by stimulated emission. Proc. Natl. Acad. Sci. USA, 97, 8206–8210.CrossRefGoogle ScholarPubMed
Klein, G., Satre, M., et al. (1982). Spontaneous aggregation of the mitochondrial natural ATPase inhibitor in salt solutions as demonstrated by gel filtration and neutron scattering. Application to the concomitant purification of the ATPase inhibitor and F1-ATPase. Biochim. Biophys. Acta, 681, 226–232.CrossRefGoogle ScholarPubMed
Kleywegt, G. J., and Jones, T. A. (2002). Homo crystallographicus – quo vadis? Structure (Camb), 10, 465–472.CrossRefGoogle ScholarPubMed
Klotz, L. C., and Zimm, B. H. (1972). Size of DNA determined by viscoelastic measurements: results on bacteriophage, Bacillus subtilus and Escherichia coli, 72, 779–800.Google Scholar
Koch, M. H., and Stuhrmann, H. B. (1979). Neutron-scattering studies of ribosomes. Methods Enzymol., 59, 670–706.CrossRefGoogle ScholarPubMed
Konermann, L., and Douglas, D. J. (1998a). Equilibrum unfolding of proteins monitored by electrospray ionization mass spectrometry: distinguishing two-state from multi-state transition. Rapid Commin. Mass Spectr., 12, 435–442.3.0.CO;2-F>CrossRefGoogle Scholar
Konermann, L., and Douglas, D. J. (1998b). Unfolding of proteins monitored by electrospray ionization mass spectrometry: a comparison of positive and negative ion modes. J. Am. Soc. Mass Spectrom., 9, 1248–1254.CrossRefGoogle Scholar
Koppel, D. E. (1979). Fluorescence redistribution after photobleaching. Biophys. J., 28, 281–291.CrossRefGoogle ScholarPubMed
Korgel, B. A., Zanten, J. H., and Monbouquette, H. G. (1998). Vesicle size distributions measured by flow field-flow fractionation coupled with multiangle light scattering. Biophys. J., 74, 3264–3272.CrossRefGoogle ScholarPubMed
Kossiakoff, A. A. (1983). Neutron protein crystallography: advances in methods and applications. Ann. Rev. Biophys. Bioeng., 12, 159–182.CrossRefGoogle ScholarPubMed
Kovacic, R. T., and Holde, K. E. (1977). Sedimentation of homogeneous double-strand DNA molecules. Biochemistry, 16, 1490–1498.CrossRefGoogle ScholarPubMed
Kroes, S. J., Canters, G. W., Giardi, G., Hoek, A., and Visser, A. J. W. G. (1998). Time-resolved fluorescence study of azurin variants: conformational heterogeneity and tryptophan mobility. Biophys. J., 75, 2441–2450.CrossRefGoogle ScholarPubMed
Kumar, A., Ernst, R. R., and Wuthrich, K. (1980). A two-dimensional nuclear Overhauser enhancement (2D NOE) experiment for the elucidation of complete proton–proton cross-relaxation networks in biological macromolecules. Biochem. Biophys. Res. Commun., 95, 1–6.CrossRefGoogle ScholarPubMed
Kunst, F., Oyasawara, N., et al. (1997). The complete genome sequence of the Gram-positive bacterium Bacillus subtilis. Nature, 390, 249–256.CrossRefGoogle ScholarPubMed
Kuntz, I. D. Jr., and Kauzmann, W. (1974). Hydration of Proteins and Polypeptides. In Advances in Protein Chemistry, eds. Anfinsen, C. B., Edsall, J. T. and Richards, F. M., Vol. 28. New York: Academic Press.Google Scholar
Lakowicz, J. R. (ed.) (1999). Principles of Fluorescence Spectroscopy, second edn. New York: Kluwer Academic/Plenum Publ.CrossRefGoogle Scholar
Lakowicz, J. R., Gryczynski, I., et al. (2000). Microsecond dynamics of biological macromolecules. Methods Enzymol., 323, 473–509.CrossRefGoogle ScholarPubMed
Langan, P., Nishiyama, Y., et al. (1999). A revised structure and hydrogen bonding system in cellulose II from a neutron fibre diffraction analysis. J. Am. Chem. Soc., 121, 9940–9946.CrossRefGoogle Scholar
Langley, K. H. (1992). Developments in electrophoretic laser light scattering and some biochemical applications. In Laser Scattering in Biochemistry, eds. Harding, S. E., Sattelle, D. B., and Bloomfield, V. A.. Cambridge: Royal Society of Chemistry.Google Scholar
Langowski, J., Kremer, W., and Kapp, U. (1992). Dynamic light scattering for study of solution conformation and dynamics of superhelical DNA. Methods Enzymol., 211, 431–448.Google ScholarPubMed
Laue, T. M., Ridgeway, T. M., Wool, J. O., and Shepard, H. K. (1996). Insights into a new analytical electrophoresis apparatus. J. Pharm. Sci., 85, 1331–1335.CrossRefGoogle ScholarPubMed
Lay, J. O. (2001). MALDI-TOF mass spectrometry of bacteria. Mass Spectr. Rev., 20, 172–194.CrossRefGoogle Scholar
Leavitt, S., and Freire, E. (2001). Direct measurement of protein binding energetics by isothermal titration calorimetry. Curr. Opin. Struct. Biol., 11, 560–566.CrossRefGoogle ScholarPubMed
Lehnert, U. (2002). Hydration dependence of local thermal motions in the Purple Membrane explored by neutron scattering and isotopic labeling. Ph D Thesis. Université Joseph Fourier, Grenoble.Google Scholar
Lemasters, J. J., Chacon, E., Zahrebelski, G., Reece, J. M., and Nieminen, A.-L. (1993). Laser scanning confocal microscopy of living cells. In: Optical Microscopy. Emerging Methods and Application, eds. Herman, B. and Lemasters, J. J.. San Diego: Academic Press.Google Scholar
Leone, M., Cupane, A., et al. (1994). Thermal broadening of the Soret band in heme complexes and in heme-proteins: role of iron dynamics. Eur. Biophys. J., 23, 349–352.CrossRefGoogle ScholarPubMed
Lescrinier, E., Froeyen, M., and Herdewijn, P. (2003). Difference in conformational diversity between nucleic acids with six-membered ‘sugar’ unit and natural ‘furanose’ nucleic acids. Nucleic Acids. Res., 31, 2975–2989.CrossRefGoogle ScholarPubMed
Leslie, A. G. (1999). Integration of macromolecular diffraction data. Acta Crystallogr. D Biol. Crystallogr., 55 (Pt 10), 1696–1702.CrossRefGoogle ScholarPubMed
Levitt, M., Sander, C., et al. (1985). Protein normal-mode dynamics: trypsin inhibitor, crambin, ribonuclease and lysozyme. J. Mol. Biol., 181, 423–447.CrossRefGoogle ScholarPubMed
Lewis, A., Lieberman, K., et al. (1995). New design and imaging concepts in NSOM, Ultramicroscopy, 61, 215–220.CrossRefGoogle Scholar
Lewis, A., Radko, A., Ami, N. B., Palanker, D., and Lieberman, K. (1999). Near-field scanning optical microscopy in cell biology. TIBS, 9, 70–73.Google ScholarPubMed
Li, H., Cocco, M. J., Steitz, T., and Engelman, D. M. (2001). Conversion of phospholamban into a soluble pentameric helical bundle. Biochemistry, 40, 6636–6645.CrossRefGoogle ScholarPubMed
Li, Y., Hunter, R. L., and Mciver, R. T. Jr. (1994). High-resolution mass spectrometer for protein chemistry. Nature, 370, 393–395.CrossRefGoogle ScholarPubMed
Liphardt, J., Onoa, B., Smith, S. B., Tinoco, I. Jr., and Bustamante, C. (2001). Reversible unfolding of single RNA moleculees by mechanical forces. Science, 292, 733–737.CrossRefGoogle Scholar
Longsworth, L. G. (1953). Diffusion measurements, at 25ο, of aqueous solution of amino acids, peptides and sugars. J. Am. Chem. Soc., 75, 5705–5709.CrossRefGoogle Scholar
Lu, H. P., Xun, L., and Xie, X. S. (1998). Single-molecule enzymatic dynamics. Science, 282, 1877–1882.CrossRefGoogle ScholarPubMed
Luthy, R., Bowie, J. U., and Eisenberg, D. (1992). Assessment of protein models with three-dimensional profiles. Nature, 356, 83–85.CrossRefGoogle ScholarPubMed
Luz, Z., and Meiboom, S. (1963). J. Chem. Phys., 39, 366–370.CrossRef
Luzzati, V., and Tarclieu, A. (1980). Recent developments in solution X-ray scattering. Annu. Rev. Biophys. Bioeng., 9, 1–29.CrossRefGoogle ScholarPubMed
Ma, J., Flynn, T. C., et al. (2002). A dynamic analysis of the rotation mechanism for conformational change in F(1)-ATPase. Structure (Camb), 10, 921–931.CrossRefGoogle ScholarPubMed
MacDonald, D., and Lu, P. (2002). Residual dipolar couplings in nucleic acid structure determination. Curr. Opinion Str. Biol., 12, 337–343.CrossRefGoogle ScholarPubMed
MacGregor, I. K., Anderson, A. L., and Laue, T. M. (2004). Fluorescence detection for the XLI analytical ultracentrifuge. Biophys. Chem., 108, 165–185.CrossRefGoogle ScholarPubMed
Machtle, W. (1999). High-resolution, submicron particle size distribution analysis using gravitational-sweep sedimentation. Biophys. J., 76, 1080–1091.CrossRefGoogle ScholarPubMed
Madern, D., Ebel, C., et al. (2000). Halophilic adaptation of enzymes. Extremophiles, 4, 91–98.CrossRefGoogle ScholarPubMed
Makhatadze, G. I., and Privalov, P. L. (1990). Heat capacity of proteins. I. Partial molar heat capacity of individual amino acid residues in aqueous solution: hydration effect. J. Mol. Biol., 213, 375–384.CrossRefGoogle ScholarPubMed
Mancini, E. J., Clarke, M., et al. (2000). Cryo-electron microscopy reveals the functional organization of an enveloped virus, Semliki Forest virus. Mol. Cell. 5, 255–266.CrossRefGoogle Scholar
Mandelkow, E., and Mandelkow, E. M. (2002). Kinesin motor and disease. TICB, 12, 585–591.CrossRefGoogle ScholarPubMed
Manor, D., Weng, G., Deng, H., Cosloy, S., Chen, C. X. (1991). An isotope edited classical Raman difference spectroscopic study of the interactions of guanine nucleotides with elongation factor Tu and H-ras p21. Biochemistry, 30, 10914–10920.CrossRefGoogle ScholarPubMed
Mark, J. E. (1996). Physical Properties of Polymers Handbook. Woodbury, NY: AIP Press.Google Scholar
Marassi, F. M., Ma, C., et al. (1999). Correlation of the structural and functional domains in the membrane protein Vpu from HIV-1. PNAS, 96, 14336–14341.CrossRefGoogle ScholarPubMed
Mattei, B., Borch, J., and Roepstorff, P. (2004). Biomolecular interaction analysis and MS. Anal. Chem., 76, 18A–26A.CrossRefGoogle Scholar
McLafferty, F. W. (1993). Interpretation of Mass Spectra, fourth edn. Mill Valley, CA: University Science Books.Google Scholar
Medalia, O., Weber, I., et al. (2002). Macromolecular architecture in eukaryotic cells visualized by cryoelectron tomography. Science, 298, 1209–1213.CrossRefGoogle ScholarPubMed
Medek, A., Olejniczak, E. T., Meadows, R. P., and Fesik, S. W. (2000). An approach for high-throughput structure determination of proteins by NMR spectroscopy. J. Biomol. NMR, 18, 229–238.CrossRefGoogle ScholarPubMed
Mehta, A. D., Rief, M., Spudlich, J. A., Smith, D. A., and Simmons, R. M. (1999). Single-molecule biomechanics with optical methods. Science, 283, 1689–1695.CrossRefGoogle ScholarPubMed
Merritt, E. A., Sarfaty, S., et al. (1994). Crystal structure of cholera toxin B-pentamer bound to receptor GM1 pentasaccharide. Protein Sci, 3, 166–175.CrossRefGoogle ScholarPubMed
Merzel, F., and Smith, J. C. (2002). Is the first hydration shell of lysozyme of higher density than bulk water? Proc. Natl. Acad. Sci. USA, 99, 5378–5383.CrossRefGoogle ScholarPubMed
Meselson, M., and Stahl, F. W. (1958). The replication of DNA in Escherichia coli. Proc. Natl. Acad. Sci. USA, 44, 671–682.CrossRefGoogle ScholarPubMed
Meselson, M., Stahl, F. W., and Vinograd, J. (1957). Equilibrium sedimentation of macromolecules in density gradient. Proc. Natl. Acad. Sci. USA, 43, 581–588.CrossRefGoogle Scholar
Michielsen, S., and Pecora, R. (1981). Solution dimensions of the gramicidin dimer by dynamic light scattering. Biochemistry, 20, 6994–6997.CrossRefGoogle ScholarPubMed
Miele, A. E., Federici, L., et al. (2003). Analysis of the effect of microgravity on protein crystal quality: the case of a myoglobin triple mutant. Acta Crystallogr. D Biol. Crystallogr., 59 (Pt 6), 982–988.CrossRefGoogle ScholarPubMed
Mitellbach, P., and Porod, G. (1962). Zur Röntgenkleinwinkelstreuung die Berechning der Steukurven von Dreiachsigen Ellipsoiden. Acta Physica Austriaca, 15, 122–147.Google Scholar
Moerner, W. E., and Orrit, M. (1999). Illuminating single molecules in condensed matter. Science, 283, 1670–1676.CrossRefGoogle ScholarPubMed
Mollova, E. T., Hansen, M. R., and Pardi, A. (2000). Global structure of RNA determined with residual dipolar coupling, JACS, 122, 11561–11562.CrossRefGoogle Scholar
Monkos, K., and Turczynski, B. (1991). Determination of the axial ratio of globular proteins in the aqueous solution using viscometric measurements. Int. J. Biol. Macromol. 13, 341–344.CrossRefGoogle ScholarPubMed
Montelione, G. T., Zheng, D., Huang, Y. J., Gundalus, K. S., and Szyperski, T. (2000). Protein NMR spectroscopy in structural genomics. Nat. Struct. Biol., 7, 982–985.CrossRefGoogle ScholarPubMed
Moscowitz, A. (1962). Theoretical aspects of optical activity. Part one: Small molecules. Adv. Chem. Phys., 4, 67–112.Google Scholar
Mou, J., Csajkovsky, D. M., Zhang, Y., and Shao, Z. (1995). High-resolution atomic force microscopy of DNA: the pitch of the double helix. FEBS Lett., 371, 279–282.Google ScholarPubMed
Muller, D. J., Janovjak, H., Lehto, T., Kuerschner, L., and Anderson, K. (2002). Observing structure, function and assembly of single proteins by AFM. Progress Biophys. Mol. Biol., 79, 1–43.CrossRefGoogle ScholarPubMed
Mullett, W. M., Lai, E. P., et al. (2000). Surface plasmon resonance-based immunoassays. Methods, 22, 77–91.CrossRefGoogle ScholarPubMed
Munro, I., Pecht, I., and Stryer, L. (1979). Subnanosecond motions of tryptophan residues in proteins. Proc. Natl. Acad. Sci. USA, 76, 56–60.CrossRefGoogle ScholarPubMed
Myers, E. W., Sutton, G. G., et al. (2000). A whole-genome assembly of Drosophila. Science, 287, 2196–2204.CrossRefGoogle ScholarPubMed
Myszka, D. G., Sweet, R. W., et al. (2000). Energetics of the HIV gp120-CD4 binding reaction. Proc. Natl. Acad. Sci. USA, 97, 9026–9031.CrossRefGoogle ScholarPubMed
Nagorni, M., and Hell, S. W. (1998). 4Pi-Confocal microscopy provides three-dimensional images of the microtubule network with 100- to 150-nm resolution. J. Struct. Biol., 123, 236–247.CrossRefGoogle ScholarPubMed
Navarro, S., Lopez Martinez, M. C., Garcia de la Torre, J. (1995). Relaxation times in electric birefringence of flexible polymer chains. J. Chem. Phys., 103, 7631–7639.CrossRefGoogle Scholar
Navaza, J., and Saludjian, P. (1997). AMoRe: an automated molecular replacement program package. Methods Enzymol., 276, 581–594.CrossRefGoogle ScholarPubMed
Newcomb, W. W., Juhas, R. M., et al. (2001). The UL6 gene product forms the portal for entry of DNA into the herpes simplex virus capsid. J. Virol., 75, 10923–10932.CrossRefGoogle ScholarPubMed
Nie, S., and Zare, R. N. (1997). Optical detection of single molecule. Annu. Rev. Biophys. Biomol. Struct., 26, 567–596.CrossRefGoogle Scholar
Nie, S., Chlu, D. T., and Zare, R. N. (1995). Real-time detection of single molecules in solution by confocal fluorescence microscopy. Anal. Chem., 67, 2849–2857.CrossRefGoogle Scholar
Nielsen, M. L., Bennet, K. L., Larsen, B., Monniate, M., and Mann, M. (2002). Peptide end sequencing by orthogonal MALDI tandem mass spectroscopy. J. Proteome Res., 1, 63–71.CrossRefGoogle Scholar
Nir, S., and Stein, W. D. (1971). Two modes of diffusion in liquids. J. Chem. Phys., 55, 1598–1603.CrossRefGoogle Scholar
Nishiyama, Y., Langan, P., et al. (2002). Crystal structure and hydrogen-bonding system in cellulose Ibeta from synchrotron X-ray and neutron fiber diffraction. J. Am. Chem. Soc., 124, 9074–9082.CrossRefGoogle ScholarPubMed
Nishiyama, Y., Okano, T., et al. (1999). High resolution neutron fibre diffraction data on hydrogenated and deuterated cellulose. Int. J. Biol. Macromol., 26, 279–283.CrossRefGoogle ScholarPubMed
Noji, H., Yasuda, R.et al. (1997). Direct observation of the rotation of F1-ATPase. Nature, 386, 299–302.CrossRefGoogle ScholarPubMed
Nollert, P., Navarro, J., and Landau, E. M. (2002). Crystallization of membrane proteins in cubo. Methods in Enzym., 343, 183–99.CrossRefGoogle ScholarPubMed
Oberg, K. A., Ruysschaert, J. M., et al. (2004). The optimization of protein secondary structure determination with infrared and circular dichroism spectra. Eur. J. Biochem., 271, 2937–2948.CrossRefGoogle ScholarPubMed
Ogorzalek Loo, R. R., Cavalcoli, J. D., et al. (2001). Virtual 2-D gel electrophoresis: visualization and analysis of the E. coli proteome by mass spectrometry. Anal. Chem., 73, 4063–4070.CrossRefGoogle Scholar
Opella, S. J., and Stewart, P. L. (1989). Solid-state nuclear magnetic resonance structural studies of proteins. Methods Enzymol., 176, 242–275.CrossRefGoogle ScholarPubMed
Orgel, L. E. (1968). Evolution of the genetic apparatus. J. Mol. Biol., 38(3), 381–93.CrossRefGoogle ScholarPubMed
Ormo, M., Cubbit, A. B., Kallio, K., Gross, L. A., and Tsien, R. Y. (1996). Crystal structure of the Aequorea victoria green fluorescent protein. Science, 273, 1392–1395.CrossRefGoogle ScholarPubMed
Oster, G., and Wang, H. (2003). Rotary protein motor. Trends Cell Biol., 13, 114–121.CrossRefGoogle Scholar
Otting, G., and Wuthrich, K. (1990) Heteronuclear filters in two-dimensional [1H,1H]-NMR spectroscopy: combined use with isotope labelling for studies of macromolecular conformation and intermolecular interactions. Q. Rev. Biophys., 23, 39–56.CrossRefGoogle ScholarPubMed
Otwinowski, Z., and Minor, W. (1997). Processing of X-ray diffraction data in oscillation mode. Methods Enzymol., 276, 307–326.CrossRefGoogle ScholarPubMed
Pauling, L., and Corey, R. (1953). A proposed structure for the nucleic acids. PNAS, 39, 84–97.CrossRefGoogle ScholarPubMed
Pecora, R. (1968). Spectral distribution of light scattered by monodisperse rigid rods. J. Chem. Phys., 48, 4126–4128.CrossRefGoogle Scholar
Perrakis, A., Sixma, T. K., et al. (1997). wARP: improvement and extension of crystallographic phases by weighted averaging of multiple-refined dummy atomic models. Acta Cryst. D., 53, 448–455.CrossRefGoogle ScholarPubMed
Peticolas, W. L. (1995). Raman spectroscopy of DNA and proteins. Methods Enzymol., 246, 389–416.CrossRefGoogle ScholarPubMed
Peticolas, W. L., and Evertsz, E. (1992). Conformation of DNA in vitro and in vivo from laser Raman scattering. Methods in Enzymol., 211, 335–352.CrossRefGoogle ScholarPubMed
Pfuhl, M., Chen, H. A., Kristinsen, S., Driscool, P. C. (1999). NMR exchange broadening arising from specific low affinity protein self-association: Analysis of nitrogen-15 relaxation for rat CD2 domain 1. J. Biomolecular NMR, 14, 307–320.CrossRefGoogle ScholarPubMed
Piston, D. W. (1999). Imaging living cells and tissues by two-photon excitation microscopy. Trends Cell Biol., 9, 66–69.CrossRefGoogle ScholarPubMed
Pitner, T. P., Walter, R., and Glickson, J. D. (1976). Mechanism of the intramolecular 1H nuclear Overhauser effect in peptides and depsipeptides. Biochem. Biophys. Res. Commun., 70, 746–751.CrossRefGoogle ScholarPubMed
Plenert, M. L., and Shear, J. B. (2003). Microsecond electrophoresis. Proc. Natl. Acad. Sci. USA, 100, 3853–3857.CrossRefGoogle ScholarPubMed
Pohl, F. M., and Jovin, T. M. (1972). Salt-induced co-operative conformational change of a synthetic DNA: equilibrium and kinetic studies with poly (dG-dC). J. Mol. Biol., 67, 375–396.CrossRefGoogle Scholar
Popot, J. L., Berry, E. A., et al. (2003). Amphipols: polymeric surfactants for membrane biology research. Cell Mol. Life Sci., 60, 1559–1574.CrossRefGoogle ScholarPubMed
Prestegard, J. H. (1998). New techniques in structural NMR – anisotropic interactions. Nature Struct. Biol., 5, 517–522.CrossRefGoogle ScholarPubMed
Prestegard, J. H., Valafar, H., Glushka, J., and Tian, F. (2001). Nuclear magnetic resonance in the era of structural Genomics. Biochemistry, 40, 8677–8685.CrossRefGoogle ScholarPubMed
Price, N. C., Dwek, R. A., et al. (2001). Principles and Problems in Physical Chemistry for Biochemists. Oxford: Oxford University Press.Google Scholar
Price, P. B. (2000). A habitat for psychrophiles in deep Antarctic ice. Proc. Natl. Acad. Sci. USA, 97, 1247–1251.CrossRefGoogle ScholarPubMed
Privalov, G. P., and Privalov, P. L. (2000). Problems and prospects in microcalorimetry of biological macromolecules. Methods Enzymol., 323, 31–62.CrossRefGoogle ScholarPubMed
Privalov, G., Kavina, V., et al. (1995). Precise scanning calorimeter for studying thermal properties of biological macromolecules in dilute solution. Anal. Biochem., 232, 79–85.CrossRefGoogle ScholarPubMed
Privalov, P. L. (1980). Scanning microcalorimeters for studying macromolecules. Pure & Appl. Chem., 52, 479–497.CrossRefGoogle Scholar
Privalov, P. L. (1982). Stability of proteins. Proteins which do not present a single cooperative system. Adv. Protein Chem., 35, 1–104.CrossRefGoogle ScholarPubMed
Privalov, P. L., and Khechinashvili, N. N. (1974). A thermodynamic approach to the problem of stabilization of globular protein structure: a calorimetric study. J. Mol. Biol., 86, 665–684.CrossRefGoogle ScholarPubMed
Privalov, P. L., and Makhatadze, G. I. (1990). Heat capacity of proteins. II. Partial molar heat capacity of the unfolded polypeptide chain of proteins: protein unfolding effects. J. Mol. Biol., 213, 385–391.CrossRefGoogle ScholarPubMed
Privalov, P. L., and Makhatadze, G. I. (1992). Contribution of hydration and non-covalent interactions to the heat capacity effect on protein unfolding. J. Mol. Biol., 224, 715–723.CrossRefGoogle ScholarPubMed
Purcell, E. M. (1977). Life at low Reynolds number. Am. J. Phys., 45, 311.CrossRefGoogle Scholar
Ramakrishnan, V., and Moore, P. B. (2001). Atomic structures at last: the ribosome in 2000. Curr. Opinion in Str. Biol., 11, 144–154.CrossRefGoogle ScholarPubMed
Rau, D. C., and Bloomfield, V. A. (1979) Transient electric birefringence of T7 virial DNA. Biopolymers, 18, 2783–2805.CrossRefGoogle Scholar
Rayment, I. (1996). Kinesin and myosin: molecular motor with similar engine. Structure, 4, 501–504.CrossRefGoogle Scholar
Records, M. T. Jr., Woodbury, C. P., and Inman, R. B. (1975). Characterization of rodlike DNA fragments. Biopolymers, 14, 393–408.Google Scholar
Rhee, K. H., Scarborough, G. A., et al. (2002). Domain movements of plasma membrane H(+)-ATPase: 3D structures of two states by electron cryo-microscopy. Embo. J., 21, 3582–3589.CrossRefGoogle ScholarPubMed
Richard, S. B., Madern, D., et al. (2000). Halophilic adaptation: novel solvent protein interactions observed in the 2.9 and 2.6 A resolution structures of the wild type and a mutant of malate dehydrogenase from Haloarcula marismortui. Biochemistry, 39, 992–1000.CrossRefGoogle Scholar
Rief, M., Fernandez, J. M., and Gaub, H. E. (1998). Elastically coupled two-level system as a model for biopolymer extensibility, Phys. Rev. Letters, 81, 4764–4767.CrossRefGoogle Scholar
Rief, M., Oesterhelt, T. F., Heymann, B., and Gaub, H. E. (1997). Single molecule force spectroscopy on polysaccharides by atomic force microscopy. Science, 275, 1295–1297.CrossRefGoogle ScholarPubMed
Rief, M., Pascual, J.Saraste, M., and Gaub, H. E. (1999). Single molecule force spectroscopy of spectrin repeats: low unfolding forces in helix bundles. J. Mol. Biol., 286, 553–561.CrossRefGoogle ScholarPubMed
Riek, R., Hornemann, S., Wider, G., Glockshuber, R., and Wuthrich, K. (1997). NMR characterization of the full-length recombinant murine prion protein, mPrP(23–231). FEBS Lett., 413, 282–288.CrossRefGoogle Scholar
Roberts, M. M., Coker, A. R., et al. (1999). Crystallization, X-ray diffraction and preliminary structure analysis of Mycobacterium tuberculosis chaperonin 10. Acta Crystallogr. D Biol. Crystallogr., 55 (Pt 4), 910–914.CrossRefGoogle ScholarPubMed
Rosenheck, K., and Doty, P. (1961). The far ultraviolet absorption spectra of polypeptide and protein solutions and their dependence on conformation. Proc. Natl. Acad. Sci. USA, 47, 1775–1785.CrossRefGoogle ScholarPubMed
Rosenthal, P. B., and Henderson, R. (2003). Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol., 333, 721–745.CrossRefGoogle ScholarPubMed
Rostom, A. A., and Robinson, C. V. (1999). Disassembly of intact multiprotein complexes in the gas phase. Curr. Opin. Struct. Biol., 9, 135–141.CrossRefGoogle ScholarPubMed
Rostom, A., Fucini, P., et al. (2000). Detection and selective dissociation of intact ribosomes in a mass spectrometer. Proc. Natl. Acad. Sci. USA, 97, 5185–5190.CrossRefGoogle Scholar
Rould, M. A. (1997). Screening for heavy-atom derivatives and obtaining accurate isomorphous differences. Methods in Enzym, 276 461–472.CrossRefGoogle ScholarPubMed
Rould, M. A., and Carter, C. W. Jr. (2003). Isomorphous difference methods. Methods in Enzym., 374, 145–63.CrossRefGoogle ScholarPubMed
Rowe, A. (1977). The concentration dependence of transport processes: a general description applicable to the sedimentation, translational diffusion, and viscosity coefficients of macromolecular solutes. Biopolymers, 16, 2595–2611.CrossRefGoogle Scholar
Rubtsov, I. V., Wang, J., and Hochstrasser, R. M. (2003). Dual-frequency 2D-IR spectroscopy heterodyned photon echo of the peptide bond, PNAS, 5601–5606.CrossRefGoogle ScholarPubMed
Salmeen, I., Rimai, L., Liebes, L., Rich, M. A., and McCormick, J. J. (1975). Hydrodynamic diameters of RNA tumor viruses. Studies by laser beat frequency light scattering spectroscopy of avian myeblastosos and Rauscher murine leukemia viruses. Biochemistry, 14, 134–141.CrossRefGoogle ScholarPubMed
Sanders, C. R., Hare, B. J., Howard, K. P., Prestegard, J. H. (1994). Magnetically oriented phospholipid micelles as a tool for the study of membrane-associated molecules. Prog. NMR Spectros., 26, 5.CrossRefGoogle Scholar
Schachman, H. K. (1989). Analytical ultracentrifugation reborn. Nature, 341, 259–260.CrossRefGoogle Scholar
Schmidt, B., and Reisner, D. (1992). A fluorescence detection system for analytical ultracentrifuge and its application to proteins, nucleic acids, viroid and viruses. In Analytical Ultracentrifugation in Biochemistry and Polymer Science, eds. Harding, S. E., Rowe, A. J. and Horton, J. C.. Cambridge: Royal Society of Chemistry.Google Scholar
Schmidt, T., Schultz, G. J., Baumgartner, W., Gruber, H. J., Schindler, H. (1996). Imaging of single molecule diffusion. PNAS, 93, 2926–2929.CrossRefGoogle ScholarPubMed
Schmitz, K. S. (1993). An Introduction to Dynamic Light Scattering of Molecules. Boston: Academic Press.Google Scholar
Schmitz, K. S., and Schurr, J. M. (1973). Rotational relaxation of macromolecules determined by dynamic light scattering. II. Temperature dependence for DNA. Biopolymers, 12, 1543–1564.CrossRefGoogle ScholarPubMed
Scholtan, W., and Lange, H. (1972). Bestimmung der Teilchegrobenverteilung von Latices mit der Ultracentrifuge. Kolloid-Z., u. Z. Polimere., 250, 782–796.CrossRefGoogle Scholar
Schouten, S., Hopmans, E. C., et al. (2000). Widespread occurrence of structurally diverse tetraether membrane lipids: evidence for the ubiquitous presence of low-temperature relatives of hyperthermophiles. Proc. Natl Acad. Sci. USA, 97, 14421–14426.CrossRefGoogle ScholarPubMed
Schrader, M., Bahlmann, K., Giese, G., and Hell, S. W. (1998). 4Pi-confocal imaging in fixed biological specimens. Biophys. J., 75, 1659–1668.CrossRefGoogle ScholarPubMed
Schuck, P. (2004). A model for sedimentation in inhomogeneous media. I. Dynamic density gradients from sedimenting co-solutes. Biophys. Chem., 108, 187–200.CrossRefGoogle ScholarPubMed
Schultz, D. A. (2003). Plasmon resonant particles for biological detection. Curr. Opin. Biotechnol., 14, 13–22.CrossRefGoogle ScholarPubMed
Schwabe, J. W. R., Chapman, L., Finch, J. T, Rhodes, D., and Neuhaus, D. (1993). DNA recognition by the oestrogen receptor: from solution to the crystal. Structure, 1, 187–204.CrossRefGoogle ScholarPubMed
Schwille, P., Meyer-Almes, F. J., et al. (1997). Dual-colour fluorescence cross-correlation spectroscopy for multicomponent diffusional analysis in solution. Biophys. J., 72, 1878–1886.CrossRefGoogle Scholar
Seelig, J. (2004). Thermodynamics of lipid–peptide interactions. Biochim. Biophys. Acta., 1666, 40–50.CrossRefGoogle ScholarPubMed
Seils, J., and Dorfmuller, Th. (1991). Internal dynamics of linear and superhelical DNA as studied by photon correlation spectroscopy. Biopolymers, 31, 813–825.CrossRefGoogle ScholarPubMed
Serdyuk, I. N., Grenader, A. K., et al. (1979). Study of the internal structure of Escherichia coli ribosomes by neutron and x-ray scattering. J. Mol. Biol., 135, 691–707.CrossRefGoogle ScholarPubMed
Serdyuk, I. N., Pavlov, M., et al. (1994). The triple isotopic substitution method in small angle neutron scattering. Application to the study of the ternary complex EF-Tu.GTP.aminoacyl-tRNA. Biophys. Chem., 53, 123–130.CrossRefGoogle Scholar
Serdyuk, I., Ulitin, A., et al. (1999). Structure of a beheaded 30S ribosomal subunit from Thermus thermophilus. J. Mol. Biol., 292, 633–639.CrossRefGoogle Scholar
Sheetz, M. P., Turne, Y. S., Qian, H., and Elson, E. L. (1989). Nanometre level analysis demonstrates that lipid flow does not drive membrane glycoprotein movements. Nature, 340, 284–285.CrossRefGoogle Scholar
Sheraga, H. A., and Mandelkern, L. (1953). Consideration of the hydrodynamic properties of proteins. J. Am. Chem. Soc., 75, 179–184.CrossRefGoogle Scholar
Sheraga, H. A., Edsall, J. T., and Gadd, J. O. (1951). Double refraction of flow: numerical evaluation of extinction angle and birefringence as a function of velocity gradient. J. Chem. Phys., 19, 1101–1108.CrossRefGoogle Scholar
Shiku, H., and Dunn, R. C. (1999). Near field scanning optical microscopy. Anal. Chem., 71, 23A–29A.CrossRefGoogle ScholarPubMed
Shingyoji, C.Higuchi, H., Yoshimura, M., Katayama, E., and Yanagida, T. (1998). Dynein arms are oscillating force generators. Nature, 393, 711–714.CrossRefGoogle ScholarPubMed
Shiryaev, V. M., Selivanova, O. M., Hartsch, T., Nazimov, I. V., and Spirin, A. S. (2002). Ribosomal protein S1 from Thermus thermophilus: its detection, identification and overproduction. FEBS Letters, 525, 88–92.CrossRefGoogle ScholarPubMed
Shivashankar, G. V., and Livchaber, A. (1997). Single DNA molecule grafting and manipulation using a combined atomic force microscope and an optical tweezer. Appl. Phys. Lett., 71, 3727–3729.CrossRefGoogle Scholar
Sigler, P. B., Xu, Z., et al. (1998). Structure and function in GroEL-mediated protein folding. Ann. Rev. Biochem., 67, 581–608.CrossRefGoogle ScholarPubMed
Simpson, A. A., Tao, Y., et al. (2000). Structure of the bacteriophage Φ29 DNA packaging motor. Nature, 408, 745–750.CrossRefGoogle ScholarPubMed
Skoog, D. A., Holler, F. J., and Nieman, T. A. (1995). Principle of Instrumental Analysis. Philadelphia: Saunders College Publishing.Google Scholar
Sliz, P., Harrison, S. C., and Rosenbaum, G. (2003). How does radiation damage in protein crystals depend on X-ray dose? Structure (Camb), 11, 13–19.CrossRefGoogle ScholarPubMed
Smith, D. E., Babcock, H. P., and Chu, S. (1999). Single-polymer dynamics in steady shear flow. Science, 283, 1724–1727.CrossRefGoogle ScholarPubMed
Smith, D. E., Tans, S. J., et al. (2001). The bacteriophage Φ29 portal motor can package DNA against a large internal force. Nature, 413, 748–752.CrossRefGoogle ScholarPubMed
Smith, L. J., Redfield, C., et al. (1994). Comparison of four independently determined structure of human recombinant interleikin-4. Struct. Biol., 1, 301–310.CrossRefGoogle Scholar
Smith, M. H. (1970). Molecular weight of proteins and some other materials including sedimentation diffusion and frictional coefficients and partial specific volumes. In Handbook of Biochemistry. Selected Data for Molecular Biology, ed. Sober, H. A., pp. C3–C47. Cleveland, OH: The Chemical Rubber Company.Google Scholar
Smith, R. D., Bruce, J. E., et al. (1996). The role of Fourier transform ion cyclotron resonance mass spectrometry in biological research – new development and applications. In Mass Spectrometry in the Biological Science, eds. Burlingame, A. L. and Carr, S. A., pp. 25–68. Totowa, NJ: Humana Press.Google Scholar
Smith, S. B., Cui, Y., and Bustamante, C. (1996). Overstretching B-DNA: the elastic response of individual double-stranded and single-stranded DNA molecules. Science, 271, 795–799.CrossRefGoogle ScholarPubMed
Smith, S. O., and Peersen, O. B. (1992). Solid-state NMR approaches for studying membrane protein structure. Ann. Rev. Biophys. Biomol. Str., 21, 25–47.CrossRefGoogle ScholarPubMed
Snatzke, G. (1994). Circular dichroism: an introduction. In Circular Dichroism: Principles and Applications., eds, Nakanishi, K., Berova, N. and Woody, R. W.. New York, VCH Publishers.Google Scholar
Sodano, P., Chary, K. V., et al. (1991). Nuclear magnetic resonance studies of recombinant Escherichia coli glutaredoxin. Sequence-specific assignments and secondary structure determination of the oxidised form. Eur. J. Biochem, 200, 369–377.CrossRefGoogle Scholar
Sober, H. A. (ed.) (1970). Handbook of Biochemistry, 2nd edn. Cleveland: CRC Press.Google Scholar
Sorlie, S. S., and Pecora, R. (1988). A dynamic light scattering study of a 2311 base pair DNA restriction fragment. Macromolecules, 21, 1437–1441.CrossRefGoogle Scholar
Sosa, H., Dias, D. P., et al. (1997). A model for the microtubule-Ncd motor protein complex obtained by cryo-electron microscopy and image analysis. Cell, 90, 217–224.CrossRefGoogle ScholarPubMed
Spirin, A. S. (1963). Some Problems of Macromolecular Structure of Ribonucleic Acids (in Russian). Moscow: Academy of Science USSR.Google Scholar
Spirin, A. S. (2000). Ribosomes. New York: Kluwer Academic/Plenum Publishers.Google Scholar
Spiro, T. G., and Chernuszevich, R. S. (1995). Resonance Raman spectroscopy of metalloprotein. Methods Enzymol., 246, 416–459.CrossRefGoogle Scholar
Spolar, R. S., and Record, M. T. Jr. (1994). Coupling of local folding to site-specific binding of proteins to DNA. Science, 263, 777–784.CrossRefGoogle Scholar
Sportsman, J. R. (2003). Fluorescence anisotropy in pharmacologic screening. Methods Enzymol., 361, 505–529.CrossRefGoogle ScholarPubMed
Spronk, C. A. E. M., Linge, J. P., Hilbers, C. W., and Vuister, G. W. (2002). Improving the quality of protein structures derived by NMR spectroscopy, J. Biomolecular NMR, 22, 281–289.CrossRefGoogle ScholarPubMed
Spudlish, J. A. (1994). How molecular motors works. Nature, 372, 515–518.Google Scholar
Squire, P. G. (1970). An equation of consistency relating the harmonic mean relaxation time to sedimentation data. Biochim. Biophys. Acta, 221, 425–429.CrossRefGoogle ScholarPubMed
Srajer, V., Ren, Z., et al. (2001). Protein conformational relaxation and ligand migration in myoglobin: a nanosecond to millisecond molecular movie from time-resolved Laue X-ray diffraction. Biochemistry, 40, 13802–13815.CrossRefGoogle ScholarPubMed
Sreerama, N., and Woody, R. W. (2003). Structural composition of betaI- and betaII-proteins. Protein Sci, 12, 384–388.CrossRefGoogle ScholarPubMed
Stafford, W. F., and Braswell, E. H. (2004). Sedimentation velocity, multi-speed method for analyzing polydisperse solutions. Biophys. Chem., 108, 273–279.CrossRefGoogle ScholarPubMed
Steely, H. T. Jr., Gray, D. M., and Lang, D. (1986a). Study of the circular dichroism of bacteriophage φ6 and φ6 nucleocapsid. Biopolymers 25, 171–188.CrossRefGoogle Scholar
Steely, H. T. Jr., Gray, D. M., Lang, D., and Maestre, M. F. (1986b). Circular dichroism of double-stranded RNA in the presence of salt and ethanol. Biopolymers 25, 91–117.CrossRefGoogle Scholar
Stejskal, E. O., and Tanner, J. E. (1965). Spin diffusion measurements: Spin echoes in the presence of a time dependent field gradient. J. Chem. Phys, 42, 288–292.CrossRefGoogle Scholar
Stellwagen, N. C. (1981). Electric birefringence of restriction enzyme fragments of DNA: Optical factor and electric polarizability as a function of molecular weight. Biopolymers, 20, 399–434.CrossRefGoogle ScholarPubMed
Stellwagen, N. C. (1996). Electric birefringence of kilobase-sized DNA molecules. Biophys. Chem., 58, 117–124.CrossRefGoogle ScholarPubMed
Stellwagen, N. C., Gelfi, C., and Righetti, P. G. (1997). The free solution mobility of DNA. Biopolymers, 42, 687–703.3.0.CO;2-Q>CrossRefGoogle ScholarPubMed
Stoeckli, M., Chaurand, P., Hallahan, D. E., and Caprioli, R. (2001). Imaging mass spectrometry: A new technology for the analysis of protein expression in mamalian tissues. Nat. Medicine, 7, 493–496.CrossRefGoogle Scholar
Stoeckli, M., Chaurand, P., et al. (2001). Imaging mass spectrometry: a new technology for the analysis of protein expression in mammalian tissues. Nat. Med., 7, 493–96.CrossRefGoogle ScholarPubMed
Stone, M. J., Fairbrother, W. J., et al. (1992). Backbone dynamics of the Bacillus subtilis glucose permease IIA domain determined from 15N NMR relaxation measurements. Biochemistry, 31, 4394–4406.CrossRefGoogle ScholarPubMed
Strick, T. R., Allemand, J. -F., Bensimon, D., Bensimon, A., Croquettet, V. (1996). The elasticity of a single supercoiled DNA molecule. Science, 271, 1835–1837.CrossRefGoogle ScholarPubMed
Strick, T., Allemand, J. -F., Croquete, V., and Bensimon, D. (2000). Twisting and stretching single DNA molecules. Progr. Biophys. Mol. Biol., 74, 115–140.CrossRefGoogle ScholarPubMed
Stroebel, D., Choquet, Y., Popot, J. L., and Picot, D. (2003). An atypical haem in the cytochrome b(6)f complex. Nature, 426, 413–418.CrossRefGoogle ScholarPubMed
Stryer, L. (1968). Fluorescence spectroscopy of proteins. Science, 162, 526–533.CrossRefGoogle ScholarPubMed
Stryer, L., and Hougland, R. P. (1967). Energy transfer: A spectroscopic ruler. PNAS, 58, 719–726.CrossRefGoogle ScholarPubMed
Stuhrmann, H. B., and Miller, A. (1978). Small-angle scattering of biological structures. J. Appl. Cryst., 11, 325–345.CrossRefGoogle Scholar
Subramanian, S., and Henderson, R. (1999). Electron crystallography of bacteriorhodopsin with millisecond time resolution. J. Struct. Biol., 128, 19–25.CrossRefGoogle Scholar
Subramanian, S., Hirai, T., and Henderson, R. (2002). From structure to mechanism: electron crystallographic studies of bacteriorhodopsin. Phil. Trans. A Math. Phys. Eng. Sci., 360, 859–874.CrossRefGoogle Scholar
Subramaniam, S., and Henderson, R. (2000). Molecular mechanism of vectorial proton translocation by bacteriorhodopsin. Nature, 10, 653–657.CrossRefGoogle Scholar
Subramaniam, S., Lindahl, M., et al. (1999). Protein conformational changes in the bacteriorhodopsin photocycle. J. Mol. Biol., 287, 145–1461.CrossRefGoogle ScholarPubMed
Surewicz, W. K., Mantsch, H. H., et al. (1993). Determination of protein secondary structure by Fourier transform infrared spectroscopy: a critical assessment. Biochemistry, 32, 389–394.CrossRefGoogle ScholarPubMed
Susi, H., and Byler, D. M. (1986). Resolution-enhanced Fourier transform infrared spectroscopy of enzymes. Meth. Enzymol., 130, 290–311.CrossRefGoogle ScholarPubMed
Suzuki, Y., Yasunaga, T., Ohkura, R., Wakabayashi, T., and Sutoh, K. (1998). Swing of the lever arm of a myosin motor at the isomerization and phosphate-release steps. Nature, 396, 380–383.CrossRefGoogle ScholarPubMed
Svergun, D. I. (1999). Restoring low resolution structure of biological macromolecules from solution scattering using simulated annealing. Biophys. J., 76, 2879–2886.CrossRefGoogle ScholarPubMed
Svergun, D. I. (2000). Advanced solution scattering data analysis methods and their applications. J. Appl. Crystallog., 33, 530–534.CrossRefGoogle Scholar
Svergun, D. I., Barberato, C., et al. (1995). CRYSOL – a program to evalate X-ray solution scattering of biological macromolecules from atomic coordinates. J. Appl. Crystallogr., 28, 768–773.CrossRefGoogle Scholar
Svergun, D. I., Malfois, M., et al. (2000). Low resolution structure of the sigma54 transcription factor revealed by X-ray solution scattering. J. Biol. Chem., 275, 4210–4214.CrossRefGoogle ScholarPubMed
Svergun, D. I., Richard, S., et al. (1998). Protein hydration in solution: experimental observation by x-ray and neutron scattering. Proc. Natl. Acad. Sci. USA, 95, 2267–2272.CrossRefGoogle ScholarPubMed
Svergun, D. I., Volkov, V. V., et al. (1996). New developments in direct shape determination from small-angle scattering 2. Uniqueness. Acta. Crystallog., A 52, 419–426.CrossRefGoogle Scholar
Taillandier, E., and Liquier, J. (1992). Infrared spectroscopy of DNA. Meth. Enzymol., 211, 307–335.CrossRefGoogle ScholarPubMed
Takahashi, H., Nakanishi, T., Kami, K., Arata, Y., and Shimada, I. (2000). A novel NMR method for determining the interfaces of large protein–protein complexes. Natur. Struct. Biol., 7, 220–223.Google ScholarPubMed
Tamm, L. K. (1993). Total internal reflectance florescence microscopy in optical microscopy. In Emerging Methods and Applications, eds. Herman, and Lemasters, . Academic Press.Google Scholar
Tanford, C. (1968). Protein denaturation. Adv. Protein Chem., 23, 121–282.CrossRefGoogle ScholarPubMed
Tanford, C., Kawahara, K., and Lapanje, S. (1967). Proteins as random coils. I. Intrinsic viscosities and sedimentation coefficients in concentrated guanidine hydrochloride. J. Am. Chem. Soc., 89, 729–736.CrossRefGoogle Scholar
Tardieu, A., Vachette, P., et al. (1981). Biological macromolecules in solvents of variable density: characterization by sedimentation equilibrium, densimetry, and X-ray forward scattering and an application to the 50S ribosomal subunit from Escherichia coli. Biochemistry, 20, 4399–4406.CrossRefGoogle ScholarPubMed
Tarek, M., and Tobias, D. J. (2002). Single-particle and collective dynamics of protein hydration water: a molecular dynamics study. Phys. Rev. Lett., 89, 275501.CrossRefGoogle ScholarPubMed
Tcien, R. Y., and Miyawaki, A. (1998). Seeing the machinary of live cells. Science, 280, 1954–1955.Google Scholar
Tehei, M., Franzetti, B., et al. (2004). Adaptation to extreme environments: macromolecular dynamics in bacteria compared in vivo by neutron scattering. EMBO Rep., 5, 66–70.CrossRefGoogle ScholarPubMed
Tehei, M., Madern, D., et al. (2001). Fast dynamics of halophilic malate dehydrogenase and BSA measured by neutron scattering under various solvent conditions influencing protein stability. Proc. Natl. Acad. Sci. USA, 98, 14356–1461.CrossRefGoogle ScholarPubMed
Tenford, C. (1965). Physical Chemistry of Macromolecules. New York: John Wiley and Sons.Google Scholar
Thalhammer, S., Stark, R. W., Müller, S., Weinberg, J., and Heck, W. M. (1997). J. Struct. Biol., 119, 232–237.CrossRef
Thomas, G. J., and Tsuboi, M. (1993). Raman spectroscopy of nucleic acids and their complexes. Adv. Biophys. Chem., 3, 1–69.Google Scholar
Thompson, D. S., and Gill, S. J. (1967). Polymer relaxation times from birefringence relaxation measurements. J. Chem. Phys., 47, 5008–5017.CrossRefGoogle ScholarPubMed
Tian, F., Al-Hashimi, H. M., Craighead, J. L., and Prestegard, J. H. (2001). Conformational analysis of a flexible oligosaccharide using residual dipolar coupling, JACS, 123, 485–492.CrossRefGoogle Scholar
Tinoco, I. Jr., Sauer, K., and Wang, J. C., Physical Chemistry. Principles and Applications in Biological Science. Prentice Hall. New Jersey (1998).Google Scholar
Tirado, M. M., Martinez, C. L., and Garcia de la Torre, J. (1984). Comparison of theories for the translational and rotational diffusion coefficients of rod-like macromolecules. Application to short DNA fragments. J. Chem. Phys. 81, 2047–2051.CrossRefGoogle Scholar
Tjandra, N., Tate, S., Ono, A., Kainosho, M., and Bax, A. (2000). The NMR structure of a DNA dodecamer in an aqueous dilute liquid crystalline phase. JACS, 122, 6190–6200.CrossRefGoogle Scholar
Tolbert, T. J. and Williamson, J. R. (1996). Preparation of specifically deuterated RNA for NMR studies using a combination of chemical and enzymatic synthesis. JACS, 118, 7929–7940.CrossRefGoogle Scholar
Tong, L., and Rossmann, M. G. (1997). Rotation function calculations with GLRF program. Methods in Enzymology, 276, 594–611.CrossRefGoogle ScholarPubMed
Toyoshima, C., Sasabe, H., and Stokes, D. L. (1993). Three-dimensional cryo-electron microscopy of the calcium ion pump in the sarcoplasmic reticulum membrane. Nature, 362, 469–471.CrossRefGoogle ScholarPubMed
Tristram-Nagle, S., and Nagle, J. F. (2004). Lipid bilayers: thermodynamics, structure, fluctuations, and interactions. Chem. Phys. Lipids, 127(1), 3–14.CrossRefGoogle ScholarPubMed
Tsvetkov, V. N. (1989). Rigid-chain Polymers. Hydrodynamic and Optical Properties in Solution. Consultants Bureau, New York and London.Google Scholar
Tsvetkov, V. N., Eskin, V. E., and Frenkel, S. Y. (1971). Structure of Macromolecules in Solution (translated from Russian), V. 1. Chapter 7. National Lending Library for Science and Technology, Boston, UK.Google Scholar
Ulitin, A. B., Agalarov, S. C., and Serdyuk, I. N. (1997). Preparation of a “beheaded” derivative of the 30S ribosomal subunit. Biochimie, 79, 523–526.CrossRefGoogle ScholarPubMed
Unger, K. K., Huber, M., Walhagen, K., Hennessy, T. P., and Hearn, M. T. W. (2002). A critical appraisal of capillary electrochromatography. Anal. Chem., 74, 200A–207A.CrossRefGoogle ScholarPubMed
Vagin, A., and Teplyakov, A. (2000). An approach to multi-copy search in molecular replacement. Acta Crystallogr. D Biol. Crystallogr., 56, 1622–1624.CrossRefGoogle ScholarPubMed
Vale, R. D. (1996). Switches, latches, and amplifier: common themes of proteins and molecular motors. J. Cell Biol., 135, 291–302.CrossRefGoogle Scholar
Valle, M., Zavialov, A., et al. (2003). Locking and unlocking of ribosomal motions. Cell, 114, 123–34.CrossRefGoogle ScholarPubMed
Groot, F. G., Gonzàlez-Mañas, J. M., Lakey, J. H. and Pattus, F. (1991). A ‘molten globule’ membrane-insertion intermediate of the pore-forming domain of colicin A. Nature, 354, 408–410.Google Scholar
Heel, M., Gowen, B., et al. (2000). Single-particle electron cryo-microscopy: towards atomic resolution. Q. Rev. Biophys., 33, 307–369.CrossRefGoogle ScholarPubMed
Holde, (1985). Physical Biochemistry. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Varki, A. (1999). Essentials of Glycobiology. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Váró, G., and Lanyi, J. K. (1991). Effects of the crystalline structure of purple membrane on the kinetics and energetics of the bacteriorhodopsin photocycle. Biochemistry, 30, 7165–7171.CrossRefGoogle ScholarPubMed
Velazquez-Campoy, A., Kiso, Y., et al. (2001). The binding energetics of first- and second-generation HIV-1 protease inhibitors: implications for drug design. Arch. Biochem. Biophys., 390, 169–175.CrossRefGoogle ScholarPubMed
Velazquez-Campoy, A., Leavitt, S. A., et al. (2004). Characterization of protein-protein interactions by isothermal titration calorimetry. Methods Mol. Biol., 261, 35–54.Google ScholarPubMed
Venyaminov, S. Y. and Vassilenko, K. S. (1994). Determination of protein tertiary structure class from circular dichroism spectra. Anal. Biochem., 222, 176–184.CrossRefGoogle ScholarPubMed
Vénien-Bryan, C., and Fuller, S. D. (1994). The organization of the spike complex of Semliki Forest virus. J. Mol. Biol., 236, 572–583.CrossRefGoogle ScholarPubMed
Vysotski, E. S., Liu, Z. J., Rose, J., Wang, B. C., and Lee, J. (1999). Preparation and preliminary study of crystals of the recombinant calcium regulated photoprotein obelin from the bioluminescent hydroid Obelia longissima. Acta Crystallogr., D55, 1965–1966.Google Scholar
Wakia, S. (1971). Slow motion in shear flow of a doublet of two spheres in contact. J. Phys. Soc. Jpn, 31, 1581–1587.CrossRefGoogle Scholar
Walker, J. M. (2000). Electrophoretic techniques. In Principles and Techniques of Practical Biochemistry, eds. Wilson, K. and Walker, J.. Cambridge: Cambridge University Press.Google Scholar
Walther, D., Cohen, F. E., and Doniak, S. (2000). Reconstruction of low-resolution three dimensional density maps from one dimensional small-angle X-ray solution scattering data for biomolecules. J. Appl. Crystallognr., 33, 350–363.CrossRefGoogle Scholar
Wand, A. J., Ehrhardt, M. R., and Flynn, P. F. (1998). High-resolution NMR of encapsulated proteins dissolved in low-viscosity fluids. PNAS, 95, 15299–15302.CrossRefGoogle ScholarPubMed
Wang, K., Forbes, J. G., and Jin, A. J. (2001). Single molecule measurements of titin elasticity. Progr. Biophys. Mol. Biol., 77, 1–44.CrossRefGoogle ScholarPubMed
Wang, M. (1999). Manipulation of single molecules in biology. Curr. Opin. Biotech., 10, 81–86.CrossRefGoogle ScholarPubMed
Wang, M. D., Schnitzer, M. J., et al. (1998). Force and velosity measured for single molecules of RNA polymerase. Science, 282, 902–907.CrossRefGoogle Scholar
Ward, T. J. (1994). Chiral media for capillary electrophoresis. Anal. Chem., 66, 633A–640A.Google ScholarPubMed
Watson, J. D., and Crick, F. H. (1953). Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature, 171, 737–738.CrossRefGoogle ScholarPubMed
Webb, R. H. (1996). Confocal optical microscopy. Rep. Progr. Phys., 59, 427–471.CrossRefGoogle Scholar
Weber, G. (1953). Rotational Brownian motion and polarization of the fluorescence of solutions. Adv. Protein Chem., 8, 415–459.CrossRefGoogle ScholarPubMed
Weber, K., and Osborn, M. (1969). The reliability of molecular weight determinations by dodecyl sulfate-polyacrylamide gel electrophoresis. J. Biol. Chem., 244, 4406–4412.Google ScholarPubMed
Weber, P. C., and Salemme, F. R. (2003). Applications of calorimetric methods to drug discovery and the study of protein interactions. Curr. Opin. Struct. Biol., 13, 115–121.CrossRefGoogle Scholar
Weik, M. (2003). Low-temperature behavior of water confined by biological macromolecules and its relation to protein dynamics. Eur. Phys. J. E. Soft Matter, 12, 153–158.CrossRefGoogle ScholarPubMed
Weik, M., Lehnert, U., and Zaccai, G. (2005). Liquid-like water confined in stacks of biological membranes at 200 K and its relation to protein dynamics. Biophys. J., 89, 3639–3646.CrossRefGoogle ScholarPubMed
Weik, M., Ravelli, R. B., et al. (2000). Specific chemical and structural damage to proteins produced by synchrotron radiation. Proc. Natl. Acad. Sci. USA, 97, 623–628.CrossRefGoogle ScholarPubMed
Weiskopf, A. S., Vouros, P., and Harvey, D. (1997). Characterization of oligosaccaride composition and structure by quadropole ion trap mass spectrometry. Rapid Commun. Mass Spectr., 11, 1493–1504.3.0.CO;2-1>CrossRefGoogle Scholar
Weiss, S. (1999). Fluorescence spectroscopy of single biomolecules. Science, 283, 1676–1683.CrossRefGoogle ScholarPubMed
Weiss, S. (2000). Shattering the diffraction limit of light: a revolution in fluorescence microscopy? Proc. Nat. Acad. Sci. USA, 97, 8747–8749.CrossRefGoogle ScholarPubMed
Weissman, M., Schindler, H., and Feher, G. (1976). Determination of molecular weights by fluctuation spectroscopy: application to DNA. Proc. Nat. Acad. Sci. USA, 73, 2776–2780.CrossRefGoogle Scholar
Westhof, E., Dumas, P., and Moras, D. (1988). Acta Cryst., A44, 122–123.
Wetlaufer, D. B. (1962). Ultraviolet spectra of proteins and nucleic acids. Adv. Prot. Chem., 17, 303–390.CrossRefGoogle Scholar
Wider, G., and Wuthrich, K. (1999). NMR spectroscopy of large molecules and multimolecular assemblies in solution. Curr. Opin. Struct. Biol., 9, 594–601.CrossRefGoogle ScholarPubMed
Wilkinson, S. R., and Thurston, G. B. (1976). The optical birefringence of DNA solutions induced by oscillatory electric and hydrodynamic fields. Biopolymers, 15, 1555–1572.CrossRefGoogle ScholarPubMed
Willcox, B. E., Gao, G. F., et al. (1999). TCR binding to peptide-MHC stabilizes a flexible recognition interface. Immunity, 10, 357–65.CrossRefGoogle ScholarPubMed
Williams, K. R., and King, R. W. (1990a). The Fourier transform in chemistry – NMR. Part 3. Multiple-pulse experiments. J. Chem. Educ., 67, A93–A99.CrossRefGoogle Scholar
Williams, K. R., and King, R. W. (1990b). The Fourier transform in chemistry – NMR. Part 4. Two-dimensional methods. J. Chem. Educ., 67, A125–A137.CrossRefGoogle Scholar
Wilson, K., and Walker, J. M. (2000). Principles and Techniques of Practical Biochemistry. Cambridge: Cambridge University Press.Google Scholar
Wimberly, B. T., Brodersen, D. E., et al. (2000). Structure of the 30S ribosomal subunit. Nature, 407, 327–339.Google ScholarPubMed
Winston, R. L., and Fitzgerald, M. C. (1997). Mass spectrometry as a readout of protein structure and function. Mass Spectr. Rev., 16, 165–179.3.0.CO;2-F>CrossRefGoogle ScholarPubMed
Woese, C. R. (1967). The Genetic Code; The Molecular Basis for Genetic Expression. New York: Harper & Row.Google Scholar
Wuthrich, K. (1986). NMR of Proteins and Nucleic Acids, New York: Wiley-Interscience.Google Scholar
Wuthrich, K. (1995). NMR – this other method for protein and nucleic acid structure determination. Acta Cryst., D51, 249–270.Google Scholar
Wuthrich, K. (2000). Protein recognition by NMR, Nat. Struct. Biol., 7, 188–189.CrossRefGoogle Scholar
Wuthrich, K. (2001). The way to NMR structures of protein. Nat. Struct. Biol., 8, 923–925.CrossRefGoogle Scholar
Wuthrich, K., Wider, G., Wagner, G., and Braun, W. (1982). Sequential resonance assignments as a basis for determination of spatial protein structures by high resolution proton nuclear magnetic resonance. J. Mol. Biol., 155, 311–319.CrossRefGoogle ScholarPubMed
Wyer, J. R., Willcox, B. E., et al. (1999). T cell receptor and coreceptor CD8 alphaalpha bind peptide-MHC independently and with distinct kinetics. Immunity, 10, 219–225.CrossRefGoogle ScholarPubMed
Xie, X. S., and Dunn, R. C. (1994). Probing single molecule dynamics. Science, 265, 361–364.CrossRefGoogle ScholarPubMed
Xu, H.-X., and Yeung, E. S. (1997). Direct measurement of single-molecule diffusion and photodecomposition in free solution. Science, 275, 1106–1109.CrossRefGoogle ScholarPubMed
Xu, Z., Horwich, A. L., and Sigler, P. B. (1997). Nature, 388, 741–750.CrossRef
Xue, Q., and Yeung, E. S. (1995). Difference in the chemical reactivity of individual molecules of an enzyme. Nature, 373, 681–683.CrossRefGoogle Scholar
Yan, Y., Winograd, E., et al. (1993). Crystal structure of the repetitive segments of spectrin. Science, 262, 2027–2030.CrossRefGoogle ScholarPubMed
Yanagida, T., Kitamura, K., Tanaka, H., Iwane, A. H., and Esaki, S. (2000). Single molecule analysis of the actomyosin motor. Curr. Opin. Cell. Biol., 12, 20–25.CrossRefGoogle ScholarPubMed
Yguerabide, J., Epstein, H. F., and Stryer, L. (1970). Segmental flexibility in an antibody molecule. J. Mol. Biol., 51, 573–590.CrossRefGoogle Scholar
Yonath, A., Miissig, J., et al. (1980). Crystallization of the large ribosomal subunit from B. stearothermophilus. Biochem. Int., 1, 428.Google Scholar
Yoshida, K., Yoshimoto, M., et al. (1998). Fabrication of new substrate for atomic force microscopic observation of DNA molecules from an ultrasmooth sapphire plate. Biophys. J., 74, 1654–1657.CrossRefGoogle ScholarPubMed
Yoshizaki, T., and Yamakawa, H. (1980). Dynamics of spheroid–cylindrical molecules in dilute solution. J. Chem. Phys., 72, 57–69.CrossRefGoogle Scholar
Zaccai, G. (2000). How soft is a protein? A protein dynamics force constant measured by neutron scattering. Science, 288, 1604–1607.CrossRefGoogle ScholarPubMed
Zaccai, G., Morin, P., et al. (1979). Interactions of yeast valyl-tRNA synthetase with RNAs and conformational changes of the enzyme. J. Mol. Biol., 129, 483–500.CrossRefGoogle ScholarPubMed
Zanni, M., and Hochstrasser, R. (2001). Two-dimensional infrared spectroscopy: a promising new method for the time resolution of structure. Curr. Opin. Struct. Biol., 11, 516–562.CrossRefGoogle Scholar
Zheng, R., Zheng, X., Dong, J., and Carey, P. R. (2004). Proteins can convert to β-sheet in single crystals. Protein Science, 13, 1288–1294.CrossRefGoogle ScholarPubMed
Zhou, H.-X. (1995). Calculation of translational friction and intrinsic viscosity. I. General formulation for arbitrarily shaped particles. Biophys. J., 69, 2286–2297.CrossRefGoogle ScholarPubMed
Zhou, H.-X. (2001). A unified picture of protein hydration: prediction of hydrodynamic properties from known structures. Biophys. Chem. 93, 171–179.CrossRefGoogle ScholarPubMed
Zhuang, X., Bartley, L. E., et al. (2000). A single-molecule study of RNA catalysis and folding. Science, 288, 2048–2051.CrossRefGoogle ScholarPubMed
Zidek, L., Stefl, R., and Sklenar, V. (2001). NMR methodology for the stydy of nucleic acids. Curr. Opin. Struct. Biol., 11, 275–281.CrossRefGoogle Scholar
Zipper, P., and Durchshlag, H. (2000). Prediction of hydrodynamic and small angle scattering parameters from crystal and electron microscopic structure. J. Appl. Cryst., 33, 788–792.CrossRefGoogle Scholar
Zlatanova, J., Lindsay, S. M., and Leuba, A. H. (2000). Single molecule force spectroscopy in biology using the atomic force microscope. Progr. Biophys. Mol. Biol., 74, 37–61.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Igor N. Serdyuk, Institute of Protein Research, Moscow, Nathan R. Zaccai, University of Bristol, Joseph Zaccai, Institut de Biologie Structurale, Grenoble
  • Book: Methods in Molecular Biophysics
  • Online publication: 05 November 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511811166.044
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Igor N. Serdyuk, Institute of Protein Research, Moscow, Nathan R. Zaccai, University of Bristol, Joseph Zaccai, Institut de Biologie Structurale, Grenoble
  • Book: Methods in Molecular Biophysics
  • Online publication: 05 November 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511811166.044
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Igor N. Serdyuk, Institute of Protein Research, Moscow, Nathan R. Zaccai, University of Bristol, Joseph Zaccai, Institut de Biologie Structurale, Grenoble
  • Book: Methods in Molecular Biophysics
  • Online publication: 05 November 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511811166.044
Available formats
×