Hostname: page-component-5c6d5d7d68-sv6ng Total loading time: 0 Render date: 2024-08-22T05:21:25.275Z Has data issue: false hasContentIssue false

Sequence-specific assignments and their use in NMR studies of DNA structure

Published online by Cambridge University Press:  17 March 2009

B. R. Reid
Affiliation:
Chemistry Department and Biochemistry Department, University of Washington, Seattle, WA 98195, U.S.A.

Extract

There has been a surge of recent interest, reflected by a sharp increase in the number of publications, in the area of high-resolution nuclear magnetic resonance (NMR) studies of DNA. The goal of many of these studies is to monitor the structure of biologically important DNA sequences directly in solution; the impetus for such studies was the realization, from early single-crystal X-ray structures, that nearest-neighbor context effects are a major determinant of local structure in short double-helical DNAs (Dickerson & Drew, 1981; Dickerson, 1983). Thus, instead of the previously assumed regular averaged structure of the double helix derived from fibre diffraction analysis, the more interesting concept emerged that specific sequence-dependent distortions from ‘classical’ DNA structure might be responsible for the recognition of such sequences by a variety of ligands such as repressors, polymerases, drugs, etc.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1987

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

Abragam, A. (1961). Principles of Nuclear Magnetism, Oxford University Press.Google Scholar
Angerman, N. S., Victor, T. A., Bell, C. L. & Danyluk, S. (1972). Proton magnetic resonance study of the aggregation of actinomycin D in deuterated water. Biochemistry 11, 24022411.Google Scholar
Arnott, S. & Chandrasekaran, R. (1983). personal communication.Google Scholar
Arnott, S., Chandrasekaran, R., Puigjaner, L. C., Walker, J. K., Hall, I. H., Birdsall, D. L. & Ratcliff, R. L. (1983). Wrinkled DNA. Nucl. Acids Res. 11, 14571474.Google Scholar
Arnott, S. & Hukins, D. W. (1972). Optimized parameters for A-DNA and B-DNA. Biochem. biophys. Res. Commun. 47, 15041510.CrossRefGoogle Scholar
Aue, W. P., Bartholdi, E. & Ernst, R. R. (1976). Two-dimensional spectroscopy. Applications to nuclear magnetic resonance J. chem. Phys. 64, 22292246.Google Scholar
Boelens, R., Scheek, R. M., Dijkstra, K. & Kaptein, R. (1985). Sequential assignment of imino- and amino-proton resonances in proton NMR spectra of oligonucleotides by two-dimensional NMR spectroscopy. Application to a lac operator fragment. J. magn. Reson. 62, 378386.Google Scholar
Boelens, R., Scheek, R. M., van Boom, J. H. & Kaptein, R. (1986). The complex of lac repressor headpiece with a 14 base pair lac operator fragment studied by two-dimensional NMR. J. molec. Biol. (in the press.)Google Scholar
Broido, M. S., James, T. L., Zon, G. & Keepers, J. W. (1985). Investigation of the solution structure of a DNA octamer [d(GGAATTCC)]2, using two-dimensional nuclear Overhauser enhancement spectroscopy. Eur. J. Biochem. 150, 117125.CrossRefGoogle ScholarPubMed
Brown, S. C., Mullis, K., Levenson, C. & Shafer, R. H. (1984). Aqueous solution structure on an intercalated actinomycin D-dATGCAT complex by two-dimensional and one-dimensional proton NMR. Biochemistry 23, 403498.CrossRefGoogle Scholar
Buck, F., Hahn, K. D., Brill, W., Ruterjans, H., Chernov, B. K., Skryabin, K. G., Kirpichnikov, M. P. & Bayev, A. A. (1986). NMR studies of DNA recognition sequences and their interaction with proteins. The phage λ OR1 operator, a symmetric lac operator and their specific complexes with cro protein and lac repressor ‘head-piece’. J. Biomol. Struct. Dyn 3, 899911.CrossRefGoogle Scholar
Buck, F., Hahn, K. D., Zemann, W., Ruterjans, H., Sadler, J. R., Beyreuther, K., Kaptein, R., Scheek, R. M. & Hull, W. E. (1983). NMR study of the interaction between the lac repressor and the lac operator. Eur. J. Biochem. 132, 321328.CrossRefGoogle ScholarPubMed
Chazin, W. J., Wüthrich, K., Hyberts, S., Rance, M., Denny, W. A. & Leupin, W. (1986). 1H nuclear magnetic resonance assignments for d(GCATTAATGC)2 using experimental refinements of established procedures. J. molec. Biol. 190, 439453.CrossRefGoogle ScholarPubMed
Chou, S. H., Hare, D. R., Wemmer, D. E. & Reid, B. R. (1983). Sequence-specific recognition of deoxyribonucleic acid. Chemical synthesis and nuclear magnetic resonance assignment of the imino protons of λ OR3 operator deoxyribonucleic acid. Biochemistry 22, 30373041.CrossRefGoogle ScholarPubMed
Chou, S. H., Wemmer, D. E., Hare, D. R. & Reid, B. R. (1984). Sequence-specific recognition of DNA: NMR studies of the imino protons of a synthetic RNA polymerase promoter. Biochemistry 22, 22572262.CrossRefGoogle Scholar
Clore, G. M. & Gronenborn, A. M. (1983). Sequence-dependent structural variations in two right-handed alternating pyrimidine-purine DNA oligomers in solution determined by nuclear Overhauser enhancement measurements. EMBO J. 2, 21092115.Google Scholar
Clore, G. M. & Gronenborn, A. M. (1984 a). A nuclear-Overhauser-enhancement study of the solution structure of a double-stranded DNA undecamer comprising a portion of the specific target site for the cyclic-AMP-receptor protein in the gal operon. Sequential resonance assignment. Eur. J. Biochem. 141, 119136.CrossRefGoogle ScholarPubMed
Clore, G. M. & Gronenborn, A. M. (1984 b). Internal mobility in a double-stranded B DNA hexamer and undecamer -A time dependent proton-proton nuclear Over-hauser enhancement study. FEBS Lett. 172, 219226.CrossRefGoogle Scholar
Clore, G. M. & Gronenborn, A. M. (1984 c). Interproton distance measurements in solution for a double-stranded DNA undecamer comprising a porton of the specific target site for the cyclic AMP receptor protein in the gal-operon. A nuclear Overhauser enhancement study. FEBS Lett. 175, 117139.Google Scholar
Clore, G. M. & Gronenborn, A. M. (1985). The solution structure of a B-DNA undecamer comprising a portion of the specific target site for the cAMP receptor protein in the gal operon. Refinement on the basis of interproton distance data. EMBO J. 4, 820835.Google Scholar
Clore, G. M., Gronenborn, A. M., Moss, D. S. & Tickel, I. J. (1985). Refinement of the solution structure of the B DNA hexamer 5′d(CGTACG)2 on the basis of inter-proton distance data. J. molec. Biol. 185, 219226.Google Scholar
Crippen, G. M. (1981). Distance Geometry and Conformational Calculations, Research Studies Press/Wiley, Chichester.Google Scholar
Crippen, G. M. & Havel, T. F. (1978). Stable circulation of coordinates from distance information. Acta Cryst. A34, 282284.CrossRefGoogle Scholar
Dickerson, R. E. (1983). Base sequence and helix structure variations in B and A DNA J. molec. Biol. 166, 419441.Google Scholar
Dickerson, R. E. & Drew, H. R. (1981). Structure of a B-DNA Dodecamer II. Influence of base sequence on helix structure. J. molec. Biol. 149, 761786.CrossRefGoogle ScholarPubMed
Dickerson, R. E. & Kopka, M. L. (1985). Nuclear Overhauser data and sterochemical considerations suggest that netropsin binds symmetrically within the minor groove of poly (dA). poly (dT) forming hydrogen bonds with both strands of the double helix. J. Biomol. Struct. Dyn. 3, 423431.CrossRefGoogle Scholar
Eigen, M. (1964). Proton transfer, acid-base catalysis and enzymatic hydrolysis. Angew. Chem. Int. Ed. 3, 172.CrossRefGoogle Scholar
Englander, S. W. & Kallenbach, N. R. (1984). Hydrogen exchange and structural dynamics of proteins and nucleic acids. Quart. Rev. Biophys. 16, 521569.CrossRefGoogle Scholar
Feigon, J., Leupin, W., Denny, W. A. & Kearns, D. R. (1983). Two-dimensional proton nuclear magnetic resonance investigation of the synthetic deoxyribonucleic acid decamer d(ATATCGATAT)2 Biochemistry 22, 59435951.Google Scholar
Feigon, J., Wright, J. M., Leupin, W., Denny, W. A. & Kearns, D. R. (1982). Use of two-dimensional nmr in the study of a double stranded DNA decamer. J. Am. chem. Soc. 104, 55405541.Google Scholar
Freeman, R. & Morris, G. A. (1979). Two-dimensional fourier transforms in nmr. Bull. magn. Reson. 1, 526.Google Scholar
Frey, M. H., Leupin, W., Sorensen, O. W., Denny, W. A., Ernst, R. R. & Wüthrich, K. (1985). Sequence-specific assignment of the backbone proton-and phosphorus-31 NMR lines in a short DNA duplex with homo-and heteronuclear correlated spectroscopy. Biopolymers 24, 23712380.Google Scholar
Gorenstein, D. G. (1983). Non-biological aspects of phosphorus-31 nmr spectroscopy. Prog. nucl. mag. Reson. Spectrosc. 16, 198.CrossRefGoogle Scholar
Gorenstein, D. G., Lai, K. & Shah, D. O. (1984). Phosphorus-31 and two-dimensional phosphorus-31 proton correlated NMR spectra of duplex d(AP[17O]Gp[18O]Cp [16O]T) and assignment of phosphorus-31 signals in d(ApGpCpT)2-actinomycin D complex. Biochemistry 23, 6717A6723A.Google Scholar
Graves, D. E., Stone, M. P. & Krugh, T. R. (1985) Structure of the Anthramycind(ATGCAT)2 adduct from one- and two- dimensional proton NMR experiments in solution. Biochemistry 24, 7573–5785.CrossRefGoogle Scholar
Gronenborn, A. M. & Clore, G. M. (1985) Investigations of the solution structures of short nucleic acid fragments by means of nuclear Overhauser enhancement measurements. Prog. nucl. magn. Reson. Spectrosc. 17, 132.CrossRefGoogle Scholar
Gronenborn, A. M., Clore, G. M. & Kimber, B. J. (1984) An investigation into the solution structures of two self-complementary DNA oligomers, 5′-d(C-G-T-A-C-G) and 5′-d(A-C-G-C-G-C-G-T) by means of nuclear-Overhauser-enhancement measurements. Biochem. J. 221, 723736.Google Scholar
Gueron, M., Kochoyan, M. & Leroy, J. L. (1986). Nucleic acid kinetics: NMR studies of proton exchange. Abst. L17.XII Int. Conf. Mag. Reson. Biol. Systems,Todtmoos, Germany,Sept. 1986.Google Scholar
Haasnoot, C. A. G., de Bruin, S. H., Berendsen, R. G., Janssen, H. G. J. M., Binnendijk, T. J. J., Hilbers, C. W., van der Marel, G. A. & van Boom, J. H. (1983 b). Structure, kinetics and thermodynamics of DNA hairpin fragments in solution. J. Biomol. Struct. Dyn. 1, 115149.Google Scholar
Haasnoot, C. A. G. & Hilbers, C. W. (1983). Effective water resonance suppression in 1D and 2D-FT 1H -NMR spectroscopy of biopolymers in aqueous solution. Biopolymers 22, 12591266.Google Scholar
Haasnoot, C. A. G., Hilbers, C. W., van der Marel, G. A., van Boom, J. H., Singh, U. C., Pattabiraman, N. & Kollman, P. A. (1986). On loopfolding in nucleic acid hairpin structures. J. Biomol. Struct. Dyn. 3, 843857.Google Scholar
Haasnoot, C. A. G., Westerink, H. P., van der Marel, G. A. & van Boom, J. H. (1983 a). Conformational analysis of a hybrid DNA-RNA double helical oligonucleotide in aqueous solution: d (CG)r(CG)d(CG) studied by 1-D and 2D-1H NMR spectroscopy. J. Biomol. Struct. Dyn. 1, 31149.CrossRefGoogle Scholar
Haneef, I., Moss, D. S., Stanford, M. J. & Borkakoti, N. (1985). Restrained structure-factor least-squares refinement of protein structures using a vector processing computer. Acta Cryst. A41, 426433.Google Scholar
Hare, D. R. & Reid, B. R. (1982). Nuclear Overhauser assignment of the imino protons of the acceptor helix and the ribothymidine helix in the nuclear magnetic resonance spectrum of Escherichia coli isoleucine transfer ribonucleic acid: evidence for costacked helixes in solution. Biochemistry 21, 51295135.CrossRefGoogle ScholarPubMed
Hare, D. R. & Reid, B. R. (1986). Three-dimensional structure of a DNA hairpin in solution: Two-dimensional nmr studies and distance geometry calculations on d(CGCGTTTTCGCG). Biochemistry 25, 53415350.Google Scholar
Hare, D. R., Ribeiro, N. S., Wemmer, D. E. & Reid, B. R. (1985). Complete assignment of the imino protons of Escherichia coli valine transfer RNA: two dimensional NMR studies in water. Biochemistry 24, 43004306.CrossRefGoogle ScholarPubMed
Hare, D. R., Shapiro, L. & Patel, D. J. (1986) – personal communication.Google Scholar
Hare, D. R., Wemmer, D. E., Chou, S. H., Drobny, G. & Reid, B. R. (1983). Assignment of the non-exchangeable proton resonances of d(CGCGAATTCGCG) using two-dimensional nuclear magnetic resonance methods. J. molec. Biol. 171, 319336.Google Scholar
Havel, T. F., Kuntz, I. D. & Crippen, G. M. (1983). The combinatorial distance geometry method for the calculation of molecular conformation. II. Sample problems and computational statistics. J. theor. Biol. 104, 383400.Google Scholar
Havel, T. F. & Wüthrich, K. (1984). A distance geometry program for determining the structure of small proteins and other macromolecules from nuclear magnetic resonance measurements of intramolecular 1H-1H proximities in solution. Bull Math. Biol. 46, 673704.Google Scholar
Havel, T. F. & Wüthrich, K. (1985). An evaluation of the combined use of nuclear magnetic resonance and distance geometry for the determination of protein conformations in solution. J. molec. Biol. 182, 281294.Google Scholar
Jamin, N., James, T. L. & Zon, G. (1985). Two-dimensional nuclear Overhauser enhancement investigation of the solution structure and dynamics of the DNA octamer [d(GGTATACC)]2. Eur. J. Biochem. 152, 157.CrossRefGoogle ScholarPubMed
Jeener, J. (1971). Informal lecture presented at Ampere Summer School, Basko Poljie, Yugoslavia.Google Scholar
Kalk, A. & Berendsen, H. J. C. (1976). Proton magnetic relaxation and spin diffusion in proteins. J. magn. Reson. 24, 343366.Google Scholar
Kaptein, R., Zuiderwég, E. R. P., Scheek, R. M., Boelens, R. & van Gunsteren, W. F. (1985). A protein structure from nuclear magnetic resonance data. The lac repressor headpiece. J. molec. Biol. 182, 179182.Google Scholar
Keepers, J. W. & James, T. L. (1984). A theoretical study of distance determination from NMR. Two-dimensional nuclear Overhauser effect spectra. J. magn. Reson. 57, 404426.Google Scholar
Kirpichnikov, M. P., Hahn, K. D., Buck, F., Ruterjans, H., Chernov, B. K., Kurochkin, A. V., Skryabin, K. G. & Bayev, A. A. (1984). 1H-nmr study of interaction of bacteriophage λ cro protein with OR3 operator. Evidence for a change in the conformation of the OR3 operator on binding. Nucl. Acids Res. 12, 35513561.Google Scholar
Klevit, R. E., Wemmer, D. E. & Reid, B. R. (1986). 1H-NMR studies on the interaction between distamycin A and a symmetrical DNA dodecamer. Biochemistry 25, 32963303.CrossRefGoogle Scholar
Klug, A., Jack, A., Viswamitra, M. A., Kennard, O., Shakkad, Z. & Steitz, T. A. (1979). A hypothesis on a specific sequence-dependent conformation of DNA and its relation to the binding of the lac-repressor protein. J. molec. Biol. 131, 669680.Google Scholar
Kopka, M. L., Yoon, C., Goodsell, D., Pjura, P. & Dickerson, R. E. (1985). Binding of an antitumor drug to DNA. Netropsin and C-G-C-G-A-A-T-T-BrC-G-C-G. J. molec. Biol. 183, 553563.Google Scholar
Lefevre, J. F., Lane, A. N. & Jardetzky, O. (1985). Nuclear magnetic resonance study of the proton exchange rate in the operator- promoter DNA sequence of the trp operon of Escherichia coli. J. molec. Biol. 185, 689699.Google Scholar
Leroy, J. L., Bolo, N., Figueroa, N., Plateau, P. & Gueron, M. (1985 b). Internal motions of transfer RNA: a study of exchanging protons by magnetic resonance. J. Biomol. Struct. Dyn. 2, 915939.Google Scholar
Leroy, J. L., Broseta, D. & Gueron, M. (1985 a). Proton exchange and base-pair kinetics of poly(rA)ׂpoly(rU) and poly(rI)ׂpoly(rC). J. molec. Biol. 184, 165168.Google Scholar
Lu, P., Cheung, S. & Arndt, K. (1983). Possible molecular detent in the DNA structure at regulatory sequences. J. Biomol. Struct. Dynam. 1, 509521.CrossRefGoogle ScholarPubMed
Macura, S. & Ernst, R. R. (1980). Elucidation of cross relaxation in liquids by two-dimensional N.M.R. spectroscopy. Molec. Phys. 41, 95117.Google Scholar
Matthews, B. W., Ohlendorf, D. H., Anderson, W. F. & Takeda, Y. (1982). Structure of the DNA-binding region of lac repressor inferred from its homology with cro repressor. Proc. Natn. Acad. Sci USA 79, 14281432.Google Scholar
Mirau, P. A. & Shafer, R. H. (1982). Role of actinomycin pentapeptides in actinomycin-DNA binding and kinetics. Biochemistry 21, 26262631.CrossRefGoogle Scholar
Muller, W. & Crothers, D. M. (1968). Studies of the binding of actinomycin and related compounds to DNA. J. molec. Biol. 35, 251290.Google Scholar
Nagayama, K., Kumar, A., Wüthrich, K. & Ernst, R. R. (1980) Experimental techniques of two-dimensional correlated spectroscopy. J. magn. Reson. 40, 321334.Google Scholar
Nerdal, W., Hare, D. R. & Reid, B. R. (1986). The three-dimensional structure of the wild type lac Pribnow promoter DNA in solution: Two-dimensional NMR studies and distance geometry calculations. (Submitted.)Google Scholar
Nilsson, L., Clore, G. M., Gronenborn, A. M., Brunger, A. T. & Karplus, M. (1986). Structure refinement of oligonucleotides by molecular dynamics with nuclear Overhauser effect interproton distance restraints: application to 5′d(C-G-T-A-C-G)2. J. molec. Biol. 188, 455475.Google Scholar
Noggle, J. H. & Schirmer, R. E. (1971). The Nuclear Overhauser Effect: Chemical Applications. Academic Press, New York.Google Scholar
Ohlendorf, D. H. & Matthews, B. W. (1983). Structural studies of protein-nucleic acid interactions. Ann. Rev. biophys. Bioeng. 12, 259273.Google Scholar
Pabo, C. O. & Sauer, R. T. (1984). Protein-DNA Recognition. Ann. Rev. Biochem. 53, 293321.Google Scholar
Patel, D. J. (1982). Antibiotic-DNA interactions: Intermolecular nuclear Overhauser effects in the netropsin-d(CGCGAATTCGCG) complex in solution. Proc. Natn. Acad. Sci. USA 79, 64246428.Google Scholar
Patel, D. J., Ikuta, S., Kozlowski, S. & Itakura, K. (1983). Sequence dependence of hydrogen exchange kinetics in DNA duplexes at the individual base pair level in solution. Proc. Natn. Acad. Sci. USA 80, 21842188.Google Scholar
Patel, D. J., Kozlowski, S. A., Rice, J. A., Broka, C. & Itakura, K. (1981). Mutual interaction between adjacent d G·dC actinomycin binding sites and dA·dT netropsin binding sites on the self-complementary d(CGCGAATTCGCG) duplex in solution. Proc. Natn. Acad. Sci. USA 78 72817284.Google Scholar
Reid, D. G., Salisbury, S. A., Bellard, S, Shakked, Z. & Williams, D. H. (1983 a). Proton nuclear Overhauser effect study of the structure of a deoxyoligonucleotide duplex in aqueous solution. Biochemistry 22, 20192025.Google Scholar
Reid, D. G., Salisbury, S. A. & Williams, D. H. (1983 b). Proton nuclear Overhauser effect study of the structure of an actinomycin D complex with a self-complementary tetranucleoside triphosphate. Biochemistry 22, 13771385.Google Scholar
Roy, S. & Redfield, A. G. (1981). Nuclear Overhauser effect study and assignment of D stem and reverse-Hoogsteen base pair proton resonances in yeast tRNAAsp. Nucl. Acids Res. 9, 70737083.Google Scholar
Saenger, W., Hunter, W. N. & Kennard, O. (1986). DNA conformation is determined by economics in the hydration of phosphate groups. Nature 324, 385388.Google Scholar
Sarma, M. H., Gupta, G. & Sarma, R. H. (1985 a). Netropsin specifically recognizes one of the two conformationally equivalent strands of poly(dA)·poly(dT). One dimensional NMR study at 500 MHz involving NOE transfer between netropsin and DNA protons. J. Biomol. Struct. Dyn. 2, 10851095.Google Scholar
Sarma, M. H., Gupta, G. & Sarma, R. H. (1985B) Structure of poly(dA)·poly(dT) is not identical to the AT rich regions of the single crystal structure of CGCGAATTBrCGCG. The consequence of this to netropsin binding to poly(dA)·poly(dT). J. Biomol. Struct. Dyn. 3, 433436.Google Scholar
Scheek, R. M., Boelens, R., Russo, N., van Boom, J. H. & Kaptein, R. (1984). Sequential resonance assignments in 1H spectra or oligonucleotides by two dimensional nmr spectroscopy. Biochemistry 23, 13711376.CrossRefGoogle ScholarPubMed
Scheek, R. M., Boelens, R., Russo, N. & Kaptein, R. (1985). Sequential resonance assignments in the 1H-NMR spectrum of a lac operator fragment by two-dimensional NOE spectroscopy at 500 MHz, in Structure and Motion: Membranes, Nucleic Acids and Proteins (Ed. Clementi, E.Corongiu, G.Sarma, M. H. and Sarma, R. H., p. 485, Adenine Press, Guilderland N.Y.Google Scholar
Scheek, R. M., Russo, N., Boelens, R. & Kaptein, R. (1983 a). Sequential resonance assignments in DNA 1H NMR spectra by two-dimensional NOE spectroscopy. J. Am. chem. Soc. 105, 29142916.Google Scholar
Scheek, R. M., Zuiderweg, E. R. P., Klappe, K. J. M., van Boom, J. H., Kaptein, R., Ruterjans, H. & Beyreuther, K. (1983 b). Lac repressor headpiece binds specifically to half of the lac operator: a proton nuclear magnetic resonance study. Biochemistry 22, 228235.Google Scholar
Shakked, Z., Rabinovich, D., Cruse, W. B. T., Egert, E., Kennard, O., Sala, G., Salisbury, S. A. & Viswamitra, M. A. (1981). Crystalline A-DNA: the X-ray analysis of the fragment d(G-G-T-A-T-A-C-C) Proc. R. Soc. London Ser. B213, 479487.Google Scholar
Sobell, H. M. & Jain, S. C. (1972). Stereochemistry of actinomycin binding to DNA. I. Refinement and further structural details of the actinomycin-deoxy guanosine crystalline complex. J. molec. Biol. 68, 2134.Google Scholar
Suzuki, E., Pattabiraman, N., Zon, G. & James, T. L. (1986). Solution structure of [d(AT)5]2 via complete relaxation matrix analysis of 2D NOE spectra and molecular mechanics calculations: Evidence for an hydration tunnel. Biochemistry (In the Press.)Google Scholar
Takusagawa, F., Dabrow, M., Neidle, S. & Berman, H. (1982). The structure of a pseudointercalated complex between actinomycin and the DNA binding sequence d(GpC). Nature (London) 296, 466469.Google Scholar
Teitelbaum, H. & Englander, S. W. (1975). Open states in native polynucleotides. II. Hydrogen-exchange study of cytosine-containing double helixes. J. molec Biol. 92, 7992.CrossRefGoogle Scholar
Ulrich, E. L., John, E. M. M., Gough, G. R., Brunden, M. J., Gilham, P. T., Westler, W. M. & Markley, J. L. (1983). Imino proton assignments in the proton nuclear magnetic resonance spectrum of the λ phage OR3 deoxyribonucleic acid fragment. Biochemistry 22, 43624365.Google Scholar
Weiss, M. A., Patel, D. J., Sauer, R. T. & Karplus, M. (1984 a). Two-dimensional 1H nmr study of the λ operator site OL1: A sequential assignment and its applications. Proc. Natn. Acad. Sci. USA 81, 130134.Google Scholar
Weiss, M. A., Patel, D. J., Sauer, R. T. & Karplus, M. (1984 b). Proton NMR study of the λ operator site OL1: assignment of the imino and adenine H2 resonances. Nucl. Acids Res. 12, 40354047.Google Scholar
Wemmer, D. E., Chou, S. H., Hare, D. R. & Reid, B. R. (1985). Duplex-hairpin transitions in DNA: NMR studies on CGCGTATACGCG. Nucl. Acids Res. 13, 37553772.Google Scholar
Wemmer, D. E. & Reid, B. R. (1983) unpublished observations.Google Scholar
Wemmer, D. E. & Reid, B. R. (1985). High resolution NMR studies of nucleic acids and proteins. Ann. Rev. phys. Chem. 36, 105137.Google Scholar
Williamson, M. P., Havel, T. F. & Wüthrich, K. (1985). Solution conformation of proteinase inhibitor IIA from bull seminal plasma by proton nuclear magnetic resonance and distance geometry. J. molec. Biol. 182, 295315.Google Scholar
Wüthrich, K., Wider, G., Wagner, G. & Braun, W. (1982). Sequential resonance assignments as a basis for determination of spatial protein structures by high resolution proton nuclear magnetic resonance. J. molec. Biol. 155, 311319.Google Scholar
Zuiderweg, E. R. P., Kaptein, R. & Wüthrich, K. (1983). Sequence specific resonance assignments in the 1H nuclear magnetic resonance spectrum of the lac repressor DNA-binding domain 1–51 from E. coli by two-dimension spectroscopy. Eur. J. Biochem. 137, 279292.CrossRefGoogle Scholar
Zuiderweg, E. R. P., Scheek, R. M. & Kaptein, R. (1985). Two-dimensional proton-NMR studies on the lac repressor DNA binding domain: further resonance assignments and identification of nuclear Overhauser enhancements. Biopolymers 24, 22572277.Google Scholar