Show simple item record

Chemical Reagents for Investigating the Major Groove of DNA

dc.contributor.authorRokita, Steven E.
dc.date.accessioned2020-01-13T15:02:39Z
dc.date.available2020-01-13T15:02:39Z
dc.date.issued2001-07
dc.identifier.citationRokita, Steven E. (2001). "Chemical Reagents for Investigating the Major Groove of DNA." Current Protocols in Nucleic Acid Chemistry 5(1): 6.6.1-6.6.16.
dc.identifier.issn1934-9270
dc.identifier.issn1934-9289
dc.identifier.urihttps://hdl.handle.net/2027.42/152488
dc.description.abstractChemical modification provides an inexpensive and rapid method for characterizing the structure of DNA and its association with drugs and proteins. Numerous conformation‐specific probes are available, but most investigations rely on only the most common and readily available of these. The major groove of DNA is typically characterized by reaction with dimethyl sulfate, diethyl pyrocarbonate, potassium permanganate, osmium tetroxide, and, quite recently, bromide with monoperoxysulfate. This commentary discusses the specificity of these reagents and their applications in protection, interference, and missing contact experiments.
dc.publisherWiley Periodicals, Inc.
dc.publisherOxford University Press
dc.titleChemical Reagents for Investigating the Major Groove of DNA
dc.typeArticle
dc.rights.robotsIndexNoFollow
dc.subject.hlbsecondlevelBiological Chemistry
dc.subject.hlbsecondlevelChemical Engineering
dc.subject.hlbsecondlevelChemistry
dc.subject.hlbsecondlevelPublic Health
dc.subject.hlbtoplevelEngineering
dc.subject.hlbtoplevelHealth Sciences
dc.subject.hlbtoplevelScience
dc.description.peerreviewedPeer Reviewed
dc.description.bitstreamurlhttps://deepblue.lib.umich.edu/bitstream/2027.42/152488/1/cpnc0606.pdf
dc.identifier.doi10.1002/0471142700.nc0606s05
dc.identifier.sourceCurrent Protocols in Nucleic Acid Chemistry
dc.identifier.citedreferencePu, W.T. and Struhl, K. 1992. Uracil interference, a rapid and general method for defining protein‐DNA interactions involving the 5‐methyl group of thymines: The GCN4‐DNA complex. Nucl. Acids Res. 20: 771 ‐ 775.
dc.identifier.citedreferenceSen, D. and Gilbert, W. 1988. Formation of parallel four‐stranded complexes by guanine‐rich motifs in DNA and its implications for meiosis. Nature 334: 364 ‐ 366.
dc.identifier.citedreferenceServa, S., Weinhold, E., Roberts, R.J., and Klimasauskas, S. 1998. Chemical display of thymine residues flipped out by DNA methyltransferease. Nucl. Acids Res. 26: 3473 ‐ 3479.
dc.identifier.citedreferenceSinger, B. and Grunberger, D. 1983. Reactions of directly acting agents with nucleic acids. In Molecular Biology of Mutagens and Carcinogens ( B. Singer and D. Grunberger, eds.)Chapter 4 pp. 45 ‐ 96. Plenum Press, New York.
dc.identifier.citedreferenceSip, M., Schwartz, A., Vovelle, F., Ptak, M., and Leng, M. 1992. Distortions induced in DNA by cis ‐platinum interstrand adducts. Biochemistry 31: 2508 ‐ 2513.
dc.identifier.citedreferenceSriram, M. and Wang, A.H.‐J. 1996. Structure of DNA and RNA. Bioorganic Chemistry: Nucleic Acids ( S.M. Hecht ed.)pp. 105 ‐ 143. Oxford University Press, New York. In
dc.identifier.citedreferenceStevens, S.Y. and Glick, G.D. 1999. Evidence for sequence‐specific recognition of DNA by anti‐single‐stranded DNA autoantibodies. Biochemistry 38: 560 ‐ 568.
dc.identifier.citedreferenceSzabó, P., Pfeifer, G.P., Miao, F., O’Connor, T.R., and Mann, J.R. 2000. Improved in vivo dimethyl sulfate footprinting using AlkA protein: DNA‐protein interactions at the mouse H19 gene promoter in primary embryo fibroblasts. Anal. Biochem. 283: 112 ‐ 116.
dc.identifier.citedreferenceTimsit, Y. and Moras, D. 1992. Crystallization of DNA. Methods Enzymol. 211: 409 ‐ 429.
dc.identifier.citedreferenceTruss, M., Chalepakis, G., Slater, E.P., Mader, S., and Beato, M. 1991. Functional interaction of hybrid response elements with wild‐type and mutant steroid hormone receptors. Mol. Cell. Biol. 11: 3247 ‐ 3258.
dc.identifier.citedreferenceTsodikov, O.V., Craig, M.L., Saecker, R.M., and Record, M.T. 1998. Quantitative analysis of multiple‐hit footprinting studies to characterize DNA conformational changes in protein‐DNA complexes. J. Mol. Biol. 283: 757 ‐ 769.
dc.identifier.citedreferenceTullius, T.D. 1991. The use of chemical probes to analyze DNA and RNA structures. Curr. Opin. Struct. Biol. 1: 428 ‐ 434.
dc.identifier.citedreferenceVenczel, E.A. and Sen, D. 1993. Parallel and antiparallel G‐DNA structures from a complex telomeric sequence. Biochemistry 32: 6220 ‐ 6228.
dc.identifier.citedreferenceVincze, A., Henderson, R.E.L., McDonald, J.J., and Leonard, N.J. 1973. Reaction of diethyl pyrocarbonate with nucleic acid components. Bases and nucleosides derived from guanine cytosine and uracil. J. Am. Chem. Soc. 95: 2677 ‐ 2682.
dc.identifier.citedreferenceWebb, M. and Thomas, J.O. 1999. Structure‐specific binding of the two tandem HMG boxes of HMG1 to four‐way junction DNA is mediated by the A domain. J. Mol. Biol. 294: 373 ‐ 387.
dc.identifier.citedreferenceWeeks, K.M. and Crothers, D.M. 1993. Major groove accessibility of RNA. Science 261: 1574 ‐ 1577.
dc.identifier.citedreferenceWilliamson, J.R. and Celander, D.W. 1990. Rapid procedure for chemical sequencing of small oligonucleotides without ethanol precipatation. Nucl. Acids Res. 18: 379.
dc.identifier.citedreferenceWilliamson, J.R., Raghuraman, M.K., and Cech, T.R. 1989. Monovalent cation‐induced structure of telomeric DNA: The G‐quartet model. Cell 59: 871 ‐ 880.
dc.identifier.citedreferenceWissmann, A. and Hillen, W. 1991. DNA contacts probed by modification protection and interference studies. Methods Enzymol. 208: 365 ‐ 379.
dc.identifier.citedreferenceWong, E. and Giandomenico, C.M. 1999. Current status of platinum‐based antitumor drugs. Chem. Rev. 99: 2451 ‐ 2466.
dc.identifier.citedreferenceWurdeman, R.L. and Gold, B. 1988. The effect of DNA sequenceionic strength, and cationic DNA affinity binder on the methylation of DNA by N‐methyl‐N‐nitrosourea. Chem. Res. Toxicol. 1: 146 ‐ 147.
dc.identifier.citedreferenceYeung, A.T., Dinehart, W.J., and Jones, B.K. 1988. Modification of guanine bases during oligonucleotide synthesis. Nucl. Acids Res. 16: 4539 ‐ 4554.
dc.identifier.citedreferenceAddess, K.J. and Feigon, J. 1996. Introduction to 1 H NMR spectroscopy of DNA. In Bioorganic Chemistry: Nucleic Acids ( S.M. Hecht ed.)pp. 163 ‐ 185. Oxford University Press, New York.
dc.identifier.citedreferenceAkman, S.A., Doroshow, J.H.., and Dizdaroglu, M. 1990. Base modification in plasmid DNA caused by potassium permanganate. Arch. Biochem. Biophys. 282: 202 ‐ 205.
dc.identifier.citedreferenceAnsari, A.Z., Chael, M.L., and O’Halloran, T.V. 1992. Allosteric underwinding of DNA is a critical step in positive control of transcription by Hg‐MerR. Nature 355: 87 ‐ 89.
dc.identifier.citedreferenceArmitage, B., Koch, T., Frydenlund, H., Ørum, H., Batz, H.‐G., and Schuster, G.B. 1997. Peptide nucleic acid‐anthraquinone conjugates: Strand invasion and photoinduced cleavage of duplex DNA. Nucl. Acids Res. 25: 4674 ‐ 4678.
dc.identifier.citedreferenceBailly, C., Gentle, D., Hamy, F., Purcell, M., and Waring, M.J. 1994. Localized chemical reactivity in DNA associated with the sequence‐specific bisintercalation of echinomycin. Biochem. J. 300: 165 ‐ 173.
dc.identifier.citedreferenceBalagurumoorthy, P. and Brahmachari, S.K. 1994. Structure and stability of human telomeric sequence. J. Biol. Chem. 269: 21858 ‐ 21869.
dc.identifier.citedreferenceBarrios, H., Grande, R., Olvera, L., and Morett, E. 1998. In vivo genomic footprinting analysis reveals that the complex Bradyrhizobium japonicum fixRnifA promoter region is differently occupied by two distinct RNA polymerase holoenzymes. Proc. Natl. Acad. Sci. U.S.A. 95: 1014 ‐ 1019.
dc.identifier.citedreferenceBavykin, S.G. and Pruss, D. 1997. DNA‐protein cross‐linking studies: Beyond the seemingly invariable nucleosome core structure. Chemtracts; Biochem. Mol. Biol. 10: 723 ‐ 736.
dc.identifier.citedreferenceBeal, P.A. and Dervan, P.B. 1992. Recognition of double helical DNA by alternative strand triple helix formation. J. Am. Chem. Soc. 114: 4976 ‐ 4982.
dc.identifier.citedreferenceBochkarev, A., Bochkareva, E., Frappier, L., and Edwards, A.M. 1998. The 2.2 Å structure of a permanganate‐sensitive DNA site bound by the Epstein‐Barr virus origin binding protein, EBNA1. J. Mol. Biol. 284: 1273 ‐ 1278.
dc.identifier.citedreferenceBorowiec, J.A. and Gralla, J.D. 1986. High‐resolution analysis of lac transcription complexes inside cells. Biochemistry 25: 5051 ‐ 5057.
dc.identifier.citedreferenceBorowiec, J.A., Zhang, L., Sasse‐Dwight, S., and Gralla, J.D. 1987. DNA supercoiling promotes formation of a bent repression loop in lac DNA. J. Mol. Biol. 196: 101 ‐ 111.
dc.identifier.citedreferenceBreuer, B., Steuer, B., and Alonso, A. 1993. Basal level transcription of the histone H1 0 gene is mediated by a 80 bp promoter fragment. Nucl. Acids Res. 21: 927 ‐ 934.
dc.identifier.citedreferenceBrewer, A.C., Marsh, P.J., and Patient, R.K. 1990. A simplified method for in vivo footprinting using DMS. Nucl. Acids Res. 18: 5574.
dc.identifier.citedreferenceBrunelle, A. and Schleif, R.F. 1987. Missing contact probing of DNA‐protein interactions. Proc. Natl. Acad. Sci. U.S.A. 84: 6673 ‐ 6676.
dc.identifier.citedreferenceChang, C.‐F., Tada, H., and Khalili, K. 1994. The role of a pentanucleotide repeat sequence, AGGGAAGGGA, in the regulation of JC virus DNA replication. Gene 148: 309 ‐ 314.
dc.identifier.citedreferenceChepanoske, C.L., Porello, S.L., Fujiwara, T., Sugiyama, H., and David, S.S. 1999. Substrate recognition by Escherichia coli MutY using substrate analogs. Nucl. Acids Res. 27: 3197 ‐ 3204.
dc.identifier.citedreferenceChow, C.S. and Barton, J.K. 1992. Transition‐metal complexes as probes of nucleic acids. Methods Enzymol. 212: 219 ‐ 242.
dc.identifier.citedreferenceClark, J.M. and Beardsley, G.P. 1987. Functional effects of cis ‐thymine glycol lesions on DNA synthesis in vitro. Biochemistry 26: 5398 ‐ 5403.
dc.identifier.citedreferenceCollier, D.A., Mergny, J.‐L., Thuong, N.T., and Hélène, C. 1991. Site‐specific intercalation at the triplex‐duplex junction induces a conformational change which is detectable by hypersensitivity to DEPC. Nucl. Acids Res. 19: 4219 ‐ 4224.
dc.identifier.citedreferenceCotton, R.G.H. 1989. Detection of single base changes in nucleic acids. Biochem. J. 263: 1 ‐ 10.
dc.identifier.citedreferenceCotton, R.G.H. and Campbell, R.D. 1989. Chemical reactivity of matched cytosine and thymine bases near mismatched and unmatched bases. Nucl. Acids Res. 17: 4223 ‐ 4233.
dc.identifier.citedreferenceCotton, R.G.H., Rodrigues, N.R., and Campbell, R.D. 1988. Reactivity of cytosine and thymine in single‐base‐pair mismatches. Proc. Natl. Acad. Sci. U.S.A. 85: 4397 ‐ 4401.
dc.identifier.citedreferenceDobi, A. and Agoston, D.V. 1998. Submillimolar levels of calcium regulates DNA structure at the dinucleotide repeat (TG/AC) n. Proc. Natl. Acad. Sci. U.S.A. 95: 5981 ‐ 5986.
dc.identifier.citedreferenceEgholm, M., Christensen, L., Dueholm, K.L., Buchardt, O., Coull, J., and Nielsen, P.E. 1995. Efficient pH‐independent sequence‐specific DNA binding by pseudoisocytosine‐containing bis‐PNA. Nucl. Acids Res. 23: 217 ‐ 222.
dc.identifier.citedreferenceEhrsemann, C., Baudin, F., Mougel, M., Romby, P., Ebel, J.‐P., and Ehresmann, B. 1987. Probing the structure of RNAs in solution. Nucl. Acids Res. 15: 9109 ‐ 9128.
dc.identifier.citedreferenceFalazka, G., Paleček, E., Wells, R.D., and Klysik, J. 1986. Site‐specific OsO 4 modification of the B‐Z junctions formed at the (dA‐dC) 32 region in supercoiled DNA. J. Biol. Chem. 261: 7093 ‐ 7098.
dc.identifier.citedreferenceFeng, B., Xiao, X., and Marzluf, G.A. 1993. Recognition of specific nucleotide bases and cooperative DNA binding by the trans ‐acting nitrogen regulatory protein NIT2. Nucl. Acids Res. 21: 3989 ‐ 3996.
dc.identifier.citedreferenceFenton, M.S., Lee, S.J., and Gralla, J.D. 2000. Escherichia coli promoter opening and ‐10 recognition: Mutational analysis of σ70. EMBO J. 19: 1130 ‐ 1137.
dc.identifier.citedreferenceFitzgerald, D.J. and Anderson, J.N. 1999. DNA distortion as a factor in nucleosome positioning. J. Mol. Biol. 293: 477 ‐ 491.
dc.identifier.citedreferenceFord, H., Chang, C.‐H., and Behrman, E.J. 1981. Sequence‐specific osmium reagents for polynucleotides. J. Am. Chem. Soc. 103: 7773 ‐ 7779.
dc.identifier.citedreferenceFox, K.R. 2000. Targeting DNA with triplexes. Curr. Med. Chem. 7: 17 ‐ 37.
dc.identifier.citedreferenceFox, K.R. and Grigg, G.W. 1988. Diethyl pyrocarbonate and permanganate provide evidence for an unusual DNA conformation induced by binding of the antitumor antibiotics bleomycin and phleomycin. Nucl. Acids Res. 16: 2063 ‐ 2075.
dc.identifier.citedreferenceFriedman, T. and Brown, D.M. 1978. Base‐specific reactions useful for DNA sequencing: Methylene blue‐sensitized photooxidation of guanine and osmium tetroxide modification of thymine. Nucl. Acids Res. 5: 615 ‐ 622.
dc.identifier.citedreferenceFurlong, J.C. and Lilley, D.M.J. 1986. Highly selective chemical modification of cruciform loops by diethyl pyrocarbonate. Nucl. Acids Res. 14: 3995 ‐ 4007.
dc.identifier.citedreferenceGiovannangeli, C. and Hélène, C. 1997. Progress in development of triplex‐based strategies. Antisense & Nucleic Acid Drug Development 7: 413 ‐ 421.
dc.identifier.citedreferenceGogos, J.A., Karayiorgou, M., Aburatani, H., and Kafatos, F.C. 1990. Detection of single base mismatches of thymine and cytosine. Nucl. Acids Res. 18: 6807 ‐ 6814.
dc.identifier.citedreferenceGough, G.W., Sullivan, K.M., and Lilley, D.M.J. 1986. The structure of cruciforms in supercoiled DNA: Probing the single‐stranded character of nucleotide bases with bisulfite. EMBO J. 1986: 191 ‐ 196.
dc.identifier.citedreferenceGruffat, N., Manet, E., Rigolet, A., and Sergeant, A. 1990. The enhancer factor R of Epstein‐Barr virus is a sequence‐specific DNA binding protein. Nucl. Acids Res. 18: 6835 ‐ 6843.
dc.identifier.citedreferenceGuo, Q., Lu, M., and Kallenbach, N.R. 1995. Effect of hemimethylation and methylation of adenine on the structure and stability of model DNA duplexes. Biochemistry 34: 16359 ‐ 16364.
dc.identifier.citedreferenceHänsler, U. and Rokita, S.E. 1993. Electrostatics rather than conformation control the oxidation of DNA by the anionic reagent permanganate. J. Am. Chem. Soc. 115: 8554 ‐ 8557.
dc.identifier.citedreferenceHartley, J.A. 1993. Selectivity in alkylating agent‐DNA interactions. In Molecular Aspects of Anticancer Drug‐DNA Interactions ( S. Neidle and M. Waring eds.) pp. 1 ‐ 31. CRC Press, Boca Raton, Fla.
dc.identifier.citedreferenceHartley, J.A., Forrow, S.M., and Souhami, R.L. 1990. Effect of ionic strength and cationic DNA affinity binders on the DNA sequence selective alkylation of guanine N7‐positions by nitrogen mustards. Biochemistry 29: 2985 ‐ 2991.
dc.identifier.citedreferenceHayatsu, H. 1976. Bisulfite modification of nucleic acids and their constituents. Prog. Nucleic Acid Res. Mol. Biol. 16: 75 ‐ 124.
dc.identifier.citedreferenceHayatsu, H. and Ukita, T. 1967. The selective degradation of pyrimidines in nucleic acids by permanganate oxidation. Biochem. Biophys. Res. Commun. 29: 556 ‐ 561.
dc.identifier.citedreferenceHerr, W. 1985. Diethyl pyrocarbonate: A chemical probe for secondary structure in negatively supercoiled DNA. Proc. Natl. Acad. Sci. U.S.A. 82: 8009 ‐ 8013.
dc.identifier.citedreferenceHerr, W., Corbin, V., and Gilbert, W. 1982. Nucleotide sequence of the 3′ half of AKV. Nucl. Acids Res. 10: 6931 ‐ 6943.
dc.identifier.citedreferenceHowgate, P., Jones, A.S., and Tittensor, J.R. 1968. The permanganate oxidation of thymidine. J. Chem. Soc. C 2 75 ‐ 279.
dc.identifier.citedreferenceHtun, H. and Johnston, B.H. 1992. Mapping adducts of DNA structural probes using transcription and primer extension approaches. Methods Enzymol. 212: 272 ‐ 294.
dc.identifier.citedreferenceIde, M., Kow, Y.W., and Wallace, S.S. 1985. Thymine glycols and urea residues in M13 DNA constitute replicative blocks in vitro. Nucl. Acids Res. 13: 8035 ‐ 8052.
dc.identifier.citedreferenceJames, T.L. 1995. Nuclear Magnetic Resonance and Nucleic Acids. Academic Press, New York.
dc.identifier.citedreferenceJamieson, E.R. and Lippard, S.J. 1999. Structure recognition and processing of cisplatin‐DNA adducts. Chem. Rev. 99: 2467 ‐ 2498.
dc.identifier.citedreferenceJeppesen, C. and Nielsen, P.E. 1988. Detection of intercalation‐induced changes in DNA structure by reaction with diethyl pyrocarbonate or potassium permanganate. FEBS Lett. 231: 172 ‐ 176.
dc.identifier.citedreferenceJeppesen, C. and Nielsen, P.E. 1989. Uranyl mediated photofootprinting reveals strong RNA polymerase‐DNA backbone contacts in the +10 region of the Deo P1 promoter open complex. Nucl. Acids Res. 17: 4947 ‐ 4956.
dc.identifier.citedreferenceJiang, H., Zacharias, W., and Amirhaeri, S. 1991. Potassium permanganate as an in situ probe for B‐Z and Z‐Z junctions. Nucl. Acids Res. 19: 6943 ‐ 6948.
dc.identifier.citedreferenceJohnston, B.H. 1992. Hydroxylamine and methoxyl‐amine as probes of DNA structure. Methods Enzymol. 212: 180 ‐ 194.
dc.identifier.citedreferenceJohnston, B.H. and Rich, A. 1985. Chemical probes of DNA conformation: Detection of Z‐DNA at nucleotide resolution. Cell 42: 713 ‐ 724.
dc.identifier.citedreferenceKasparkova, J., Farrell, N., and Brabec, V. 2000. Sequence specificity, conformation, and recognition by HMG1 protein of major DNA interstrand cross‐links of antitumor dinuclear platinum complexes. J. Biol. Chem. 275: 15789 ‐ 15798.
dc.identifier.citedreferenceKinoshita, H., Shi, Y., Sandefur, C., and Jarrard, D.F. 2000. Screening hypermethylated regions by methylation‐sensitive single‐strand conformational polymorphism. Anal. Biochem. 278: 165 ‐ 169.
dc.identifier.citedreferenceKladde, M.P. and Simpson, R.T. 1996. Chromatin structure mapping in vivo using methyltransferases. Methods Enzymol. 274: 214 ‐ 233.
dc.identifier.citedreferenceKlysik, J., Rippe, K., and Jovin, T.M. 1990. Reactivity of parallel‐stranded DNA to chemical modification reagents. Biochemistry 29: 9831 ‐ 9839.
dc.identifier.citedreferenceKoizume, S., Inoue, H., Kamiya, H., and Ohtsuka, E. 1998. Neighboring base damage induced by permanganate oxidation of 8‐oxoguanine in DNA. Nucl. Acids Res. 26: 3599 ‐ 3607.
dc.identifier.citedreferenceKönig, P., Giraldo, R., Chapman, L., and Rhodes, D. 1996. The crystal structure of the DNA‐binding domain of yeast RAP1 in complex with telomeric DNA. Cell 85: 125 ‐ 136.
dc.identifier.citedreferenceKostrhunova, H. and Brabec, V. 2000. Conformational analysis of site‐specific DNA cross‐links of cisplatin‐distamycin conjugates. Biochemistry 39: 12639 ‐ 12649.
dc.identifier.citedreferenceLambrinakos, A., Humphrey, K.E., Babon, J.J., Ellis, T.P., and Cotton, R.G.H. 1999. Reactivity of KMnO 4 and tetraethylammonium chloride with mismatched bases and a simple mutation detection protocol. Nucl. Acids Res. 27: 1866 ‐ 1874.
dc.identifier.citedreferenceLeonard, N.J., McDonald, J.J., Henderson, R.E.L., and Reichmann, M.E. 1971. Reaction of diethyl pyrocarbonate with nucleic acid components. Adenosine. Biochemistry 10: 3335 ‐ 3342.
dc.identifier.citedreferenceLilley, D.M. 1992. Probes of DNA structure. Methods Enzymol. 212: 133 ‐ 139.
dc.identifier.citedreferenceLundblad, R.L. 1995. Techniques in Protein Modification. CRC Press, Boca Raton, Fla.
dc.identifier.citedreferenceMarrot, L. and Leng, M. 1989. Chemical probes of the conformation of DNA modified by cis ‐diamminedichloroplatinum(II). Biochemistry 28: 1454 ‐ 1461.
dc.identifier.citedreferenceMäueler, W., Kyas, A., Keyl, H.‐G., and Epplen, J.T. 1998. A genome‐derived (gaa.ttc)24 trinucleotide block binds nuclear protein(s) specifically and forms triple helices. Gene 215: 389 ‐ 403.
dc.identifier.citedreferenceMaxam, A.M. and Gillbert, W. 1980. Sequencing end‐labeled DNA with base‐specific chemical cleavages. Methods Enzymol. 65: 499 ‐ 560.
dc.identifier.citedreferenceMcBoom, L.D.B. and Sadowski, P.D. 1994. Contacts of the ABF1 protein of Saccharomyces cerevisiae with a DNA‐binding site at MAT a. J. Biol. Chem. 269: 16455 ‐ 16460.
dc.identifier.citedreferenceMcCarthy, J.G. 1989. An improvement in thymine specific chemical DNA sequencing. Nucl. Acids Res. 17: 7541.
dc.identifier.citedreferenceMcCarthy, J.G. and Rich, A. 1991. Detection of an unusual distortion in A‐tract DNA using KMnO 4: Effect of temperature and distamycin on the altered conformation. Nucl. Acids Res. 19: 3421 ‐ 3429.
dc.identifier.citedreferenceMcCarthy, J.G., Williams, L.D., and Rich, A. 1990. Chemical reactivity of KMnO 4 and diethyl pyrocarbonate with B‐DNA: Specific reactivity with short A‐tracts. Biochemistry 29: 6071 ‐ 6081.
dc.identifier.citedreferenceMcCarthy, J.G., Frederick, C.A., and Nicolas, A. 1993. A structural analysis of the bent kinetoplast DNA from Crithidia fasciculata by high resolution chemical probing. Nucl. Acids Res. 21: 3309 ‐ 3317.
dc.identifier.citedreferenceMiller, P.S. 1996. Development of antisense and antigene oligonucleotide analogs. Prog. Nucleic Acid Res. Mol. Biol. 52: 261 ‐ 291.
dc.identifier.citedreferenceMirzabekov, A.D., Bavykin, S.G., Belyavsky, A.V., Karpov, V.L., Preobrazhenskaya, O.V., Shick, V.V., and Ebralidse, K.K. 1989. Mapping DNA‐protein interactions by cross‐linking. Methods Enzymol. 170: 386 ‐ 408.
dc.identifier.citedreferenceMorett, E. and Buck, M. 1989. In vivo studies on the interaction of RNA polymerase‐σ 54 with the Klebsiella pneumoniae and Rhizobium meliloti nif H promoters. J. Mol. Biol. 210: 65 ‐ 77.
dc.identifier.citedreferenceNakatani, K., Hagihara, S., Sando, S., Miyazaki, H., Tanabe, K., and Saito, I. 2000. Site selective formation of thymine glycol‐containing oligodeoxynucleotides by oxidation with osmium tetroxide and a bipyridine‐tethered oligonucleotide. J. Am. Chem. Soc. 122: 6309 ‐ 6310.
dc.identifier.citedreferenceNawamura, T., Negishi, K., and Hayatsu, H. 1994. 8‐Hydroxyguanine is not produced by permanganate oxidation of DNA. Arch. Biochem. Biophys. 311: 523 ‐ 524.
dc.identifier.citedreferenceNeidle, S. and Stuart, D.I. 1976. The crystal and molecular structure of an osmium bispyridine adduct of thymine. Biochim. Biophys. Acta 418: 226 ‐ 231.
dc.identifier.citedreferenceNejedlý, K., Kwinkowski, M., Galazka, G., Klysik, J., and Paleček, E. 1985. Recognition of the structural distortions at the junctions between B and Z segments in negatively supercoiled DNA by osmium tetroxide. J. Biomol. Struct. Dyn. 3: 467 ‐ 478.
dc.identifier.citedreferenceNejedlý, K., Skorová, E., Diekmann, S., and Paleček, E. 1998. Analysis of a curved DNA constructed from alternating dA n:dT n ‐tracts in linear and supercoiled form by high resolution chemical probing. Biophys. Chem. 73: 205 ‐ 216.
dc.identifier.citedreferenceNick, H. and Gilbert, W. 1985. Detection in vivo of protein‐DNA interactions within the lac operon of Escherichia coli. Nature 313: 795 ‐ 798.
dc.identifier.citedreferenceNickel, J., Short, M.L., Schmitz, A., Eggert, M., and Renkawitz, R. 1995. Methylation of the mouse M‐lysozyme downstream enhancer inhibits heterotetrameric GABP binding. Nucl. Acids Res. 23: 4785 ‐ 4792.
dc.identifier.citedreferenceNielsen, P.E. 1990. Chemical and photochemical probing of DNA complexes. J. Mol. Recogn. 3: 1 ‐ 25.
dc.identifier.citedreferenceO’Halloran, T.V., Frantz, B., Shin, M.K., Ralston, D.M., and Wright, J.G. 1989. The MerR heavy metal receptor mediates positive activation in a topologically novel transcription complex. Cell 56: 119 ‐ 129.
dc.identifier.citedreferencePaleček, E. 1992a. Probing DNA structure with osmium tetroxide complexes in vitro. Methods Enzymol. 212: 139 ‐ 155.
dc.identifier.citedreferencePaleček, E. 1992b. Probing of DNA structure in cells with osmium tetroxide‐2,2′‐bipyridine. Methods Enzymol. 212: 305 ‐ 318.
dc.identifier.citedreferencePeattie, D.A. 1979. Direct chemical method for sequencing RNA. Proc. Natl. Acad. Sci. U.S.A. 76: 1760 ‐ 1764.
dc.identifier.citedreferencePeattie, D.A. and Gilbert, W. 1980. Chemical probes for higher‐order structure in RNA. Proc. Natl. Acad. Sci. U.S.A. 77: 4679 ‐ 4682.
dc.identifier.citedreferencePullman, A. and Pullman, B. 1981. Molecular electrostatic potential of the nucleic acids. Quart. Rev. Biophys. 14: 289 ‐ 380.
dc.identifier.citedreferenceRahmouni, A.R. and Wells, R.D. 1989. Stabilization of Z‐DNA in vivo by localized supercoiling. Science 246: 358 ‐ 363.
dc.identifier.citedreferenceReddy, Y.V.R. and Rao, D.N. 2000. Binding of EcoP151 DNA methyltransferase to DNA reveals a large structural distortion within the recognition sequence. J. Mol. Biol. 298: 597 ‐ 610.
dc.identifier.citedreferenceRoberts, E., Deeble, V.J., Woods, C.G., and Taylor, G.R. 1997. KMnO 4 and tetraethylammonium chloride are a safe and effective substitute for OsO 4 in solid‐phase fluorescent chemical cleavage of mismatch. Nucl. Acids Res. 25: 3377 ‐ 378.
dc.identifier.citedreferenceRoss, S.A. and Burrows, C.J. 1996. Cytosine‐specific chemical probing of DNA using bromide and monoperoxysulfate. Nucl. Acids Res. 24: 5062 ‐ 5063.
dc.identifier.citedreferenceRoss, S.A. and Burrows, C.J. 1997. Bromination of pyrimidines using bromide and monoperoxysulfate: A competition study between cytidine, uridine and thymidine. Tetrahedron Lett. 38: 2805 ‐ 2808.
dc.identifier.citedreferenceRother, K.I., Silke, J., Georgiev, O., Schaffner, W., and Matsuo, K. 1995. Influence of DNA sequence and methylation status on bisulfite conversion of cytosine residues. Anal. Biochem. 231: 263 ‐ 265.
dc.identifier.citedreferenceRubin, C.M. and Schmid, C.W. 1980. Pyrimidine‐specific chemical reactions useful for DNA sequencing. Nucl. Acids Res. 8: 4613 ‐ 4619.
dc.identifier.citedreferenceRunkel, L. and Nordheim, A. 1986. Chemical footprinting of the interaction between left‐handed Z‐DNA and anti‐Z‐DNA antibodies by diethyl pyrocarbonate carbethoxylation. J. Mol. Biol. 189: 487 ‐ 501.
dc.identifier.citedreferenceSaluz, H. and Jost, J.‐P. 1989. A simple high‐resolution procedure to study DNA methylation and in vivo DNA‐protein interactions on a single‐copy gene level in higher eukaryotes. Proc. Natl. Acad. Sci. U.S.A. 86: 2602 ‐ 2606.
dc.identifier.citedreferenceSasse‐Dwight, S. and Gralla, J.D. 1989. KMnO 4 as a probe for lac promoter DNA melting and mechanism in vivo. J. Biol. Chem. 264: 8074 ‐ 8081.
dc.identifier.citedreferenceScholten, P.M. and Nordheim, A. 1986. Diethyl pyrocarbonate: A chemical probe for DNA cruciforms. Nucl. Acids Res. 14: 3981 ‐ 3993.
dc.owningcollnameInterdisciplinary and Peer-Reviewed


Files in this item

Show simple item record

Remediation of Harmful Language

The University of Michigan Library aims to describe library materials in a way that respects the people and communities who create, use, and are represented in our collections. Report harmful or offensive language in catalog records, finding aids, or elsewhere in our collections anonymously through our metadata feedback form. More information at Remediation of Harmful Language.

Accessibility

If you are unable to use this file in its current format, please select the Contact Us link and we can modify it to make it more accessible to you.