Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-sh8wx Total loading time: 0 Render date: 2024-07-18T19:11:02.204Z Has data issue: false hasContentIssue false

6 - Circular DNA

Published online by Cambridge University Press:  05 March 2015

Alexander Vologodskii
Affiliation:
New York University
Get access

Summary

In 1963 Dulbecco and Vogt, and Weil and Vinograd, discovered that dsDNA of the polyoma virus exists in a closed circular form (Dulbecco & Vogt 1963, Weil & Vinograd 1963). It turned out that this form is typical of bacterial DNA and of cytoplasmic DNA in animals. The distinctive feature of closed circular molecules is that its topological state cannot be altered by any conformational rearrangement that does not involve breaking DNA strands. This topological constraint is the basis for the fascinating properties of circular DNA molecules. The physical properties of circular DNA molecules is a subject of the current chapter.

Linking number of complementary strands and DNA supercoiling

Two forms of circular DNA molecules are extracted from the cell; they were designated as form I and form II (Weil & Vinograd 1963). The more compact form I was found to turn into form II after a single-stranded break was introduced into one chain of the double helix. Subsequent studies performed by Vinograd and co-workers linked the compactness of form I, in which both DNA strands are intact, to supercoiling. Form I is called the closed circular form. In this form each of the two strands that make up the DNA molecule are closed in on themselves. A diagram of closed circular DNA is presented in Fig. 6.1. The two strands of the double helix in closed circular DNA are topologically linked. In topological terms, the links between two strands of the double helix belong to a particular class, called the torus class (see Section 6.7). The quantitative description of such links is called the linking number, Lk, which may be determined in the following way (Fig. 6.2). One of the strands defines the edge of an imaginary surface (any such surface gives the same result). Lk is the algebraic (i.e. sign-dependent) number of intersections of the surface by the other strand.

Type
Chapter
Information
Biophysics of DNA , pp. 184 - 250
Publisher: Cambridge University Press
Print publication year: 2015

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

Aboul-ela, F., Bowater, R. P. & Lilley, D. M. J. (1992). Competing B–Z and helix–coil conformational transitions in supercoiled plasmid DNA. J. Biol. Chem. 267, 1776–85.Google ScholarPubMed
Adams, C. C. (1994). The Knot Book. New York: Freeman.
Adrian, M., ten Heggeler-Bordier, B., Wahli, W., Stasiak, A. Z., Stasiak, A. & Dubochet, J. (1990). Direct visualization of supercoiled DNA molecules in solution. EMBO J. 9, 4551–4.Google ScholarPubMed
Alexander, J. W. (1928). Topological invariants of knots and links. Trans. Am. Math. Soc. 30, 275–306.CrossRefGoogle Scholar
Bauer, W. & Vinograd, J. (1968). The interaction of closed circular DNA with intercalative dyes. I. The superhelix density of SV40 DNA in the presence and absence of dye. J. Mol. Biol. 33, 141–71.CrossRefGoogle ScholarPubMed
Bauer, W. & Vinograd, J. (1970). The interaction of closed circular DNA with intercalative dyes. II. The free energy of superhelix formation in SV40 DNA. J. Mol. Biol. 47, 419–35.CrossRefGoogle ScholarPubMed
Bauer, W. R. (1978). Structure and reactions of closed duplex DNA. Annu. Rev. Biophys. Bioeng. 7, 7–287.CrossRefGoogle ScholarPubMed
Bauer, W. R. & Benham, C. J. (1993). The free energy, enthalpy and entropy of native and of partially denatured closed circular DNA. J. Mol. Biol. 234, 1184–96.CrossRefGoogle ScholarPubMed
Bauer, W. R., Crick, F. H. C. & White, J. H. (1980). Supercoiled DNA. Sci. Am. 243, 100–13.Google ScholarPubMed
Beard, P., Morrow, J. F. & Berg, P. (1973). Cleavage of circular, superhelical simian virus 40 DNA to a linear duplex by S1 nuclease. J. Virol. 12, 1303–13.Google ScholarPubMed
Bednar, J., Furrer, P., Stasiak, A., Dubochet, J., Egelman, E. H. & Bates, A. D. (1994). The twist, writhe and overall shape of supercoiled DNA change during counterion-induced transition from a loosely to a tightly interwound superhelix. J. Mol. Biol. 235, 825–47.CrossRefGoogle ScholarPubMed
Benham, C. J. (1978). The statistics of superhelicity. J. Mol. Biol. 123, 361–70.CrossRefGoogle ScholarPubMed
Benham, C. J. (1979). Torsional stress and local denaturation in supercoiled DNA. Proc. Natl. Acad. Sci. U. S. A. 76, 3870–4.CrossRefGoogle ScholarPubMed
Benham, C. J. (1983). Statistical mechanical analysis of competing conformational transitions in superhelical DNA. Cold Spring Harbor Symp. Quant. Biol. 47, 219–27.CrossRefGoogle ScholarPubMed
Bliska, J. B. & Cozzarelli, N. R. (1987). Use of site-specific recombination as a probe of DNA structure and metabolism in vivo. J. Mol. Biol. 194, 205–18.CrossRefGoogle ScholarPubMed
Boles, T. C., White, J. H. & Cozzarelli, N. R. (1990). Structure of plectonemically supercoiled DNA. J. Mol. Biol. 213, 931–51.CrossRefGoogle ScholarPubMed
Borst, P., Overdulve, J. P., Weijers, P. J., Fase-Fowler, F. & Van den Berg, M. (1984). DNA circles with cruciforms from Isospora (Toxoplasma) gondii. Biochim. Biophys. Acta 781, 100–11.Google ScholarPubMed
Brahms, S., Nakasu, S., Kikuchi, A. & Brahms, J. G. (1989). Structural changes in positively and negatively supercoiled DNA. Eur. J. Biochem. 184, 297–303.CrossRefGoogle ScholarPubMed
Brahms, S., Vergne, J., Brahms, J. G., Di Capua, E., Bucher, P. & Koller, T. (1982). Natural DNA sequences can form left-handed helices in low salt solution under conditions of topological constraint. J. Mol. Biol. 162, 473–93.CrossRefGoogle ScholarPubMed
Brown, P. O. & Cozzarelli, N. R. (1979). A sign inversion mechanism for enzymatic supercoiling of DNA. Science 206, 1081–3.CrossRefGoogle ScholarPubMed
Brown, P. O. & Cozzarelli, N. R. (1981). Catenation and knotting of duplex DNA by type 1 topoisomerases: a mechanistic parallel with type 2 topoisomerases. Proc. Natl. Acad. Sci. U. S. A. 78, 843–7.CrossRefGoogle ScholarPubMed
Browning, D. F., Grainger, D. C. & Busby, S. J. W. (2010). Effects of nucleoid-associated proteins on bacterial chromosome structure and gene expression. Curr. Opin. Microbiol. 13, 773–80.CrossRefGoogle ScholarPubMed
Buck, G. R. & Zechiedrich, E. L. (2004). DNA disentangling by type-2 topoisomerases. J. Mol. Biol. 340, 933–9.CrossRefGoogle ScholarPubMed
Calugareanu, G. (1961). Surlas classes d'isotopie des noeuds tridimensionnels et leurs invariants. Czech. Math. J. 11, 588–625.Google Scholar
Champoux, J. J. (2001). DNA topoisomerases: structure, function, and mechanism. Annu. Rev. Biochem. 70, 369–413.CrossRefGoogle ScholarPubMed
Champoux, J. J. & Dulbecco, R. (1972). An activity from mammalian cells that untwists superhelical DNA – a possible swivel for DNA replication. Proc. Natl. Acad. Sci. U. S. A. 69, 143–6.CrossRefGoogle ScholarPubMed
Charvin, G., Bensimon, D. & Croquette, V. (2003). Single-molecule study of DNA unlinking by eukaryotic and prokaryotic type-II topoisomerases. Proc. Natl. Acad. Sci. U. S. A. 100, 9820–5.CrossRefGoogle ScholarPubMed
Courey, A. J. & Wang, J. C. (1983). Cruciform formation in a negatively supercoiled DNA may be kinetically forbidden under physiological conditions. Cell 33, 817–29.CrossRefGoogle Scholar
Crisona, N. J., Kanaar, R., Gonzalez, T. N., Zechiedrich, E. L., Klippel, A. & Cozzarelli, N. R. (1994). Processive recombination by wild-type Gin and an enhancer-independent mutant: Insight into the mechanisms of recombination selectivity and strand exchange. J. Mol. Biol. 243, 243–437.CrossRefGoogle ScholarPubMed
Crisona, N. J., Strick, T. R., Bensimon, D., Croquette, V. & Cozzarelli, N. R. (2000). Preferential relaxation of positively supercoiled DNA by E. coli topoisomerase IV in single-molecule and ensemble measurements. Genes Dev. 14, 2881–92.CrossRefGoogle ScholarPubMed
Dayn, A., Malkhosyan, S., Duzhy, D., Lyamichev, V., Panchenko, Y. & Mirkin, S. (1991). Formation of (dA-dT)n cruciform in E. coli cells under different environmental conditions. J. Bacteriol. 173, 2658–64.CrossRefGoogle Scholar
Dayn, A., Malkhosyan, S. & Mirkin, S. M. (1992). Transcriptionally driven cruciform formation in vivo. Nucleic Acids Res. 20, 5991–7.CrossRefGoogle ScholarPubMed
Dedon, P. C., Dederich, D. A. & Barth, M. C. (2009). An improved method for large-scale preparation of negatively and positively supercoiled plasmid DNA. Bio Techniques 47, 633–5.Google Scholar
Dekker, N. H., Rybenkov, V. V., Duguet, M., Crisona, N. J., Cozzarelli, N. R., Bensimon, D. & Croquette, V. (2002). The mechanism of type IA topoisomerases. Proc. Natl. Acad. Sci. U. S.A. 99, 12126–31.CrossRefGoogle ScholarPubMed
Depew, R. E. & Wang, J. C. (1975). Conformational fluctuations of DNA helix. Proc. Natl. Acad. Sci. U. S. A. 72, 4275–9.CrossRefGoogle ScholarPubMed
DiGate, R. J. & Marians, K. J. (1988). Identification of a potent decatenating enzyme from Escherichia coli. J. Biol. Chem. 263, 13366–73.Google Scholar
Dong, K. C. & Berger, J. M. (2007). Structural basis for gate-DNA recognition and bending by type IIA topoisomerases. Nature 450, 1201–5.CrossRefGoogle ScholarPubMed
Drlica, K. (1992). Control of bacterial DNA supercoiling. Mol. Microbiol. 6, 425–33.CrossRefGoogle ScholarPubMed
Du, Q., Livshits, A., Kwiatek, A., Jayaram, M. & Vologodskii, A. (2007). Protein-induced local DNA bends regulate global topology of recombination products. J. Mol. Biol. 368, 170–82.CrossRefGoogle ScholarPubMed
Dulbecco, R. & Vogt, M. (1963). Evidence for a ring structure of polyoma virus DNA. Proc. Natl. Acad. Sci. U. S. A. 50, 236–43.CrossRefGoogle ScholarPubMed
Ellison, M. J., Fenton, M. J., Ho, P. S. & Rich, A. (1987). Long-range interactions of multiple DNA structural transitions within a common topological domain. EMBO J. 6, 1513–22.Google ScholarPubMed
Ellison, M. J., Kelleher, R. J., 3rd, Wang, A. H., Habener, J. F. & Rich, A. (1985). Sequence-dependent energetics of the B-Z transition in supercoiled DNA containing nonalternating purine-pyrimidine sequences. Proc. Natl. Acad. Sci. U. S. A. 82, 8320–4.CrossRefGoogle ScholarPubMed
Frank-Kamenetskii, M. D., Lukashin, A. V., Anshelevich, V. V. & Vologodskii, A. V. (1985). Torsional and bending rigidity of the double helix from data on small DNA rings. J. Biomol. Struct. Dyn. 2, 1005–12.CrossRefGoogle ScholarPubMed
Frank-Kamenetskii, M. D. & Vologodskii, A. V. (1981). Topological aspects of the physics of polymers: the theory and its biophysical applications. Sov. Phys. Usp. 24, 679–96.CrossRefGoogle Scholar
Frank-Kamenetskii, M. D. & Vologodskii, A. V. (1984). Thermodynamics of the B–Z transition in superhelical DNA. Nature 307, 481–2.CrossRefGoogle ScholarPubMed
Fuller, F. B. (1971). The writhing number of a space curve. Proc. Natl. Acad. Sci. U. S. A. 68, 815–19.CrossRefGoogle ScholarPubMed
Gagua, A. V., Belintsev, B. N. & Lyubchenko, Y. L. (1981). Effect of base-pair stability on the melting of superhelical DNA. Nature 294, 662–3.CrossRefGoogle ScholarPubMed
Gebe, J. A., Delrow, J. J., Heath, P. J., Fujimoto, B. S., Stewart, D. W. & Schurr, J. M. (1996). Effects of Na+ and Mg2+ on the structures of supercoiled DNAs: comparison of simulations with experiments. J. Mol. Biol. 262, 105–28.CrossRefGoogle ScholarPubMed
Gebe, J. A. & Schurr, J. M. (1996). Thermodynamics of the first transition in writhe of a small circular DNA by Monte Carlo simulation. Biopolymers 38, 493–503.3.0.CO;2-O>CrossRefGoogle ScholarPubMed
Geggier, S., Kotlyar, A. & Vologodskii, A. (2011). Temperature dependence of DNA persistence length. Nucleic Acids Res. 39, 1419–26.CrossRefGoogle ScholarPubMed
Geggier, S. & Vologodskii, A. (2010). Sequence dependence of DNA bending rigidity. Proc. Natl. Acad. Sci. U. S. A. 107, 15421–6.CrossRefGoogle ScholarPubMed
Gellert, M., Mizuuchi, K., O'Dea, M. H. & Nash, H. A. (1976). DNA gyrase: an enzyme that introduces superhelical turns into DNA. Proc. Natl. Acad. Sci. U. S. A. 73, 3872–6.CrossRefGoogle ScholarPubMed
Gellert, M., Mizuuchi, K., O'Dea, M. H., Ohmori, H. & Tomizawa, J. (1979). DNA gyrase and DNA supercoiling. Cold Spring Harbor Symp. Quant. Biol. 43, 35–40.CrossRefGoogle ScholarPubMed
Gellert, M., O'Dea, M. H. & Mizuuchi, K. (1983). Slow cruciform transitions in palindromic DNA. Proc. Natl. Acad. Sci. U. S. A. 80, 5545–9.CrossRefGoogle ScholarPubMed
Gray, H. B., Jr., Ostrander, D. A., Hodnett, J. L., Legerski, R. J. & Robberson, D. L. (1975). Extracellular nucleases of Pseudomonas BAL 31. I. Characterization of single strand-specific deoxyriboendonuclease and double-strand deoxyriboexonuclease activities. Nucleic Acids Res. 2, 1459–92.CrossRefGoogle ScholarPubMed
Greaves, D. R., Patient, R. K. & Lilley, D. M. (1985). Facile cruciform formation by an (A-T)34 sequence from a Xenopus globin gene. J. Mol. Biol. 185, 461–78.CrossRefGoogle ScholarPubMed
Halford, S. E., Welsh, A. J. & Szczelkun, M. D. (2004). Enzyme-mediated DNA looping. Annu. Rev. Biophys. Biomol. Struct. 33, 1–24.CrossRefGoogle ScholarPubMed
Hammermann, M., Brun, N., Klenin, K. V., May, R., Toth, K. & Langowski, J. (1998). Salt-dependent DNA superhelix diameter studied by small angle neutron scattering measurements and Monte Carlo simulations. Biophys. J. 75, 3057–63.CrossRefGoogle ScholarPubMed
Haniford, D. B. & Pulleyblank, D. E. (1983a). Facile transition of poly[d(TG) • d(CA)] into a left-handed helix in physiological conditions. Nature 302, 632–4.CrossRefGoogle ScholarPubMed
Haniford, D. B. & Pulleyblank, D. E. (1983b). The in-vivo occurrence of Z DNA. J. Biomol. Struct. Dyn. 1, 593–609.CrossRefGoogle ScholarPubMed
Haniford, D. B. & Pulleyblank, D. E. (1985). Transition of a cloned d(AT)n × d(AT)n tract to a cruciform in vivo. Nucleic Acids Res. 13, 4343–63.CrossRefGoogle Scholar
Hentschel, C. C. (1982). Homocopolymer sequences in the spacer of a sea urchin histone gene repeat are sensitive to S1 nuclease. Nature 295, 714–6.CrossRefGoogle ScholarPubMed
Hiasa, H. & Marians, K. J. (1994). Topoisomerase III, but not topoisomerase I, can support nascent chain elongation during theta-type DNA replication. J. Biol. Chem. 269, 32655–9.Google ScholarPubMed
Higgins, N. P. & Vologodskii, A. V. (2004). Topological behavior of plasmid DNA. In Plasmid Biology, eds. G., Phillips & B., Funnell, 181–201. Washington, DC: ASM Press.Google Scholar
Ho, P. S., Ellison, M. J., Quigley, G. J. & Rich, A. (1986). A computer aided thermodynamic approach for predicting the formation of Z-DNA in naturally occurring sequences. EMBO. J. 5, 2737–44.Google ScholarPubMed
Hochschild, A. (1990). Protein–protein interactions and DNA loop formation. In DNA Topology and its Biological Effects, eds. N. R., Cozzarelli & J. C., Wang, 107–38. Cold Spring Harbor, NY: Cold Spring Habor Laboratory Press.Google Scholar
Hochschild, A. & Ptashne, M. (1986). Cooperative binding of lambda repressors to sites separated by integral turns of the DNA helix. Cell 44, 681–7.CrossRefGoogle ScholarPubMed
Horowitz, D. S. & Wang, J. C. (1984). Torsional rigidity of DNA and length dependence of the free energy of DNA supercoiling. J. Mol. Biol. 173, 75–91.CrossRefGoogle ScholarPubMed
Hsieh, T. S. & Wang, J. C. (1975). Thermodynamic properties of superhelical DNAs. Biochemistry 14, 14–527.CrossRefGoogle ScholarPubMed
Htun, H., Lund, E. & Dahlberg, J. E. (1984). Human U1 RNA genes contain an unusually sensitive nuclease S1 cleavage site within the conserved 3′ flanking region. Proc. Natl. Acad. Sci. U. S. A. 81, 7288–92.CrossRefGoogle ScholarPubMed
Kampranis, S. C., Bates, A. D. & Maxwell, A. (1999). A model for the mechanism of strand passage by DNA gyrase. Proc. Natl. Acad. Sci. U. S. A. 96, 8414–9.CrossRefGoogle ScholarPubMed
Kelleher, R. J., 3rd, Ellison, M. J., Ho, P. S. & Rich, A. (1986). Competitive behavior of multiple, discrete B-Z transitions in supercoiled DNA. Proc. Natl. Acad. Sci. U. S. A. 83, 6342–6.CrossRefGoogle ScholarPubMed
Keller, W. (1975). Determination of the number of superhelical turns in simian virus 40 DNA by gel electrophoresis. Proc. Natl. Acad. Sci. U. S. A. 72, 4876–80.CrossRefGoogle ScholarPubMed
Kikuchi, A. & Asai, K. (1984). Reverse gyrase – a topoisomerase which introduces positive superhelical turns into DNA. Nature 309, 677–81.CrossRefGoogle ScholarPubMed
King, I. F., Yandava, C. N., Mabb, A. M., Hsiao, J. S., Huang, H. S., Pearson, B. L., Calabrese, J. M., Starmer, J., Parker, J. S., Magnuson, T., Chamberlain, S. J., Philpot, B. D. & Zylka, M. J. (2013). Topoisomerases facilitate transcription of long genes linked to autism. Nature 501, 58–62.CrossRefGoogle ScholarPubMed
Kirkegaard, K. & Wang, J. C. (1985). Bacterial DNA topoisomerase I can relax positively supercoiled DNA containing a single-stranded loop. J. Mol. Biol. 185, 625–37.CrossRefGoogle ScholarPubMed
Klenin, K. V., Vologodskii, A. V., Anshelevich, V. V., Dykhne, A. M. & Frank-Kamenetskii, M. D. (1988). Effect of excluded volume on topological properties of circular DNA. J. Biomol. Struct. Dyn. 5, 1173–85.CrossRefGoogle ScholarPubMed
Klenin, K. V., Vologodskii, A. V., Anshelevich, V. V., Klisko, V. Y., Dykhne, A. M. & Frank-Kamenetskii, M. D. (1989). Variance of writhe for wormlike DNA rings with excluded volume. J. Biomol. Struct. Dyn. 6, 707–14.CrossRefGoogle ScholarPubMed
Klug, A. & Lutter, L. C. (1981). The helical periodicity of DNA on the nucleosome. Nucleic Acids Res. 17, 4267–83.Google Scholar
Koster, D. A., Croquette, V., Dekker, C., Shuman, S. & Dekker, N. H. (2005). Friction and torque govern the relaxation of DNA supercoils by eukaryotic topoisomerase IB. Nature 434, 671–4.CrossRefGoogle ScholarPubMed
Kouzine, F., Gupta, A., Baranello, L., Wojtowicz, D., Ben-Aissa, K., Liu, J., Przytycka, T. M. & Levens, D. (2013). Transcription-dependent dynamic supercoiling is a short-range genomic force. Nat. Struct. Mol. Biol. 20, 396–403.CrossRefGoogle ScholarPubMed
Kramer, H., Amouyal, M., Nordheim, A. & Muller-Hill, B. (1988). DNA supercoiling changes the spacing requirement of two lac operators for DNA loop formation with lac repressor. EMBO J. 7, 547–56.Google ScholarPubMed
Kramer, P. R., Bat, O. & Sinden, R. R. (1999). Measurement of localized DNA supercoiling and topological domain size in eukaryotic cells. Methods Enzymol. 304, 639–50.Google ScholarPubMed
Kramer, P. R. & Sinden, R. R. (1997). Measurement of unrestrained negative supercoiling and topological domain size in living human cells. Biochemistry 36, 3151–8.CrossRefGoogle ScholarPubMed
Krasilnikov, A. S., Podtelezhnikov, A., Vologodskii, A. & Mirkin, S. M. (1999). Large-scale effects of transcriptional DNA supercoiling in vivo. J. Mol. Biol. 292, 1149–60.CrossRefGoogle ScholarPubMed
Krasnow, M. A., Stasiak, A., Spengler, S. J., Dean, F., Koller, T. & Cozzarelli, N. R. (1983). Determination of the absolute handedness of knots and catenanes of DNA. Nature 304, 559–60.CrossRefGoogle ScholarPubMed
Kreuzer, K. N. & Cozzarelli, N. R. (1980). Formation and resolution of DNA catenanes by DNA gyrase. Cell 20, 245–54.CrossRefGoogle ScholarPubMed
Krueger, A., Protozanova, E. & Frank-Kamenetskii, M. D. (2006). Sequence-dependent base pair opening in DNA double helix. Biophys. J. 90, 3091–9.CrossRefGoogle ScholarPubMed
Krylov, D. Y., Makarov, V. L. & Ivanov, V. I. (1990). The B-A transition in superhelical DNA. Nucleic Acids Res. 18, 759–61.CrossRefGoogle Scholar
LaMarr, W. A., Sandman, K. M., Reeve, J. N. & Dedon, P. C. (1997). Large scale preparation of positively supercoiled DNA using the archaeal histone HMf. Nucleic Acids Res. 25, 1660–1.CrossRefGoogle ScholarPubMed
Larsen, A. & Weintraub, H. (1982). An altered DNA conformation detected by S1 nuclease occurs at specific regions in active chick globin chromatin. Cell 29, 609–22.CrossRefGoogle ScholarPubMed
Laundon, C. H. & Griffith, J. D. (1988). Curved helix segments can uniquely orient the topology of supertwisted DNA. Cell 52, 545–9.CrossRefGoogle ScholarPubMed
Le Bret, M. (1979). Catostrophic variation of twist and writhing of circular DNAs with constraint?Biopolymers 18, 1709–25.CrossRefGoogle Scholar
Le Bret, M. (1980). Monte Carlo computation of supercoiling energy, the sedimentation constant, and the radius of gyration of unknotted and knotted circular DNA. Biopolymers 19, 619–37.CrossRefGoogle ScholarPubMed
Lee, C. H., Mizusawa, H. & Kakefuda, T. (1981). Unwinding of double-stranded DNA helix by dehydration. Proc. Natl. Acad. Sci. U. S. A. 78, 2838–42.CrossRefGoogle ScholarPubMed
Lee, D. W. & Schleif, R. F. (1989). In vivo DNA loops in araCBAD: Size limits and helical repeat. Proc. Natl. Acad. Sci. U. S. A. 86, 476–80.CrossRefGoogle ScholarPubMed
Lilley, D. M. (1980). The inverted repeat as a recognizable structural feature in supercoiled DNA molecules. Proc. Natl. Acad. Sci. U. S. A. 77, 6468–72.CrossRefGoogle ScholarPubMed
Liu, L. F., Liu, C.-C. & Alberts, B. M. (1980). Type II DNA topoisomerases: enzymes that can unknot a topologically knotted DNA molecule via a reversible double-strand break. Cell 19, 697–707.CrossRefGoogle Scholar
Liu, L. F., Liu, C. C. & Alberts, B. M. (1979). T4 DNA topoisomerase: a new ATP-dependent enzyme essential for initiation of T4 bacteriophage DNA replication. Nature 281, 456–61.CrossRefGoogle ScholarPubMed
Liu, L. F. & Wang, J. C. (1987). Supercoiling of the DNA template during transcription. Proc. Natl. Acad. Sci. U. S. A. 84, 7024–7.CrossRefGoogle ScholarPubMed
Ljungman, M. & Hanawalt, P. C. (1992). Localized torsional tension in the DNA of human cells. Proc. Natl. Acad. Sci. U. S. A. 89, 6055–9.Google ScholarPubMed
Lyamichev, V., Panyutin, I. & Mirkin, S. (1984). The absence of cruciform structures from pAO3 plasmid DNA in vivo. J. Biomol. Struct. Dyn. 2, 291–301.CrossRefGoogle ScholarPubMed
Lyamichev, V. I., Mirkin, S. M. & Frank-Kamenetskii, M. D. (1985). A pH-dependent structural transition in the homopurine–homopyrimidine tract in superhelical DNA. J. Biomol. Struct. Dyn. 3, 327–38.CrossRefGoogle ScholarPubMed
Lyamichev, V. I., Mirkin, S. M. & Frank-Kamenetskii, M. D. (1986). Structures of homopurine–homopyrimidine tract in superhelical DNA. J. Biomol. Struct. Dyn. 3, 667–9.CrossRefGoogle ScholarPubMed
Lyamichev, V. I., Mirkin, S. M. & Frank-Kamenetskii, M. D. (1987). Structure of (dG)n·(dC)n under superhelical stress and acid pH. J. Biomol. Struct. Dyn. 5, 275–82.CrossRefGoogle ScholarPubMed
Lyamichev, V. I., Mirkin, S. M., Kumarev, V. P., Baranova, L. V., Vologodskii, A. V. & Frank-Kamenetskii, M. D. (1989). Energetics of the B-H transition in supercoiled DNA carrying d(CT)x · d(AG)x and d(C)n × d(G)n inserts. Nucleic Acids Res. 17, 9417–23.CrossRefGoogle Scholar
Lyamichev, V. I., Panyutin, I. G. & Frank-Kamenetskii, M. D. (1983). Evidence of cruciform structures in superhelical DNA provided by two-dimensional gel electrophoresis. FEBS Lett. 153, 298–302.CrossRefGoogle ScholarPubMed
Lyubchenko, Y. L. & Shlyakhtenko, L. S. (1997). Visualization of supercoiled DNA with atomic force microscopy in situ. Proc. Natl. Acad. Sci. U. S. A. 94, 496–501.CrossRefGoogle ScholarPubMed
McClellan, J., Boublikova, P., Palecek, E. & Lilley, D. (1990). Superhelical torsion in cellular DNA responds directly to enironmental and genetic factors. Proc. Natl. Acad. Sci. U. S. A. 87, 87–8373.CrossRefGoogle Scholar
Menzel, R. & Gellert, M. (1987). Regulation of the genes for E. coli DNA gyrase: homeostatic control of DNA supercoiling. Cell 34, 105–13.Google Scholar
Mirkin, S. M. & Frank-Kamenetskii, M. D. (1994). H-DNA and related structures. Annu. Rev. Biophys. Biomol. Struct. 23, 541–76.CrossRefGoogle ScholarPubMed
Mirkin, S. M., Lyamichev, V. I., Drushlyak, K. N., Dobrynin, V. N., Filippov, S. A. & Frank-Kamenetskii, M. D. (1987a). DNA H form requires a homopurine–homopyrimidine mirror repeat. Nature 330, 495–7.CrossRefGoogle ScholarPubMed
Mirkin, S. M., Lyamichev, V. I., Kumarev, V. P., Kobzev, V. F., Nosikov, V. V. & Vologodskii, A. V. (1987b). The energetics of the B-Z transition in DNA. J. Biomol. Struct. Dyn. 5, 79–88.CrossRefGoogle ScholarPubMed
Mizuuchi, K., Fisher, L. M., O'Dea, M. H. & Gellert, M. (1980). DNA gyrase action involves the introduction of transient double-strand breaks into DNA. Proc. Natl. Acad. Sci. U. S. A. 77, 1847–51.CrossRefGoogle ScholarPubMed
Naughton, C., Avlonitis, N., Corless, S., Prendergast, J. G., Mati, I. K., Eijk, P. P., Cockroft, S. L., Bradley, M., Ylstra, B. & Gilbert, N. (2013). Transcription forms and remodels supercoiling domains unfolding large-scale chromatin structures. Nat. Struct. Mol. Biol. 20, 387–95.CrossRefGoogle ScholarPubMed
Naylor, L. H., Lilley, D. M. & van de Sande, J. H. (1986). Stress-induced cruciform formation in a cloned d(CATG)10 sequence. EMBO J. 5, 2407–13.Google Scholar
Neale, B. M., Kou, Y., Liu, L., Ma'ayan, A., Samocha, K. E., Sabo, A., Lin, C. F., Stevens, C., Wang, L. S., Makarov, V., Polak, P., Yoon, S., Maguire, J., Crawford, E. L., Campbell, N. G., Geller, E. T., Valladares, O., Schafer, C., Liu, H., Zhao, T., Cai, G., Lihm, J., Dannenfelser, R., Jabado, O., Peralta, Z., Nagaswamy, U., Muzny, D., Reid, J. G., Newsham, I., Wu, Y., Lewis, L., Han, Y., Voight, B. F., Lim, E., Rossin, E., Kirby, A., Flannick, J., Fromer, M., Shakir, K., Fennell, T., Garimella, K., Banks, E., Poplin, R., Gabriel, S., DePristo, M., Wimbish, J. R., Boone, B. E., Levy, S. E., Betancur, C., Sunyaev, S., Boerwinkle, E., Buxbaum, J. D., Cook, E. H., Jr., Devlin, B., Gibbs, R. A., Roeder, K., Schellenberg, G. D., Sutcliffe, J. S. & Daly, M. J. (2012). Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 485, 242–5.CrossRefGoogle ScholarPubMed
Panayotatos, N. & Wells, R. D. (1981). Cruciform structures in supercoiled DNA. Nature 289, 466–70.CrossRefGoogle ScholarPubMed
Panyutin, I., Klishko, V. & Lyamichev, V. (1984). Kinetics of cruciform formation and stability of cruciform structure in superhelical DNA. J. Biomol. Struct. Dyn. 1, 1311–24.CrossRefGoogle ScholarPubMed
Panyutin, I., Lyamichev, V. & Mirkin, S. (1985). A structural transition in d(AT)n × d(AT)n inserts within superhelical DNA. J. Biomol. Struct. Dyn. 2, 1221–34.CrossRefGoogle Scholar
Peck, L. J., Wang, J. C., Nordheim, A. & Rich, A. (1986). Rate of B to Z structural transition of supercoiled DNA. J. Mol. Biol. 190, 125–7.CrossRefGoogle Scholar
Pettijohn, D. & Phenninger, O. (1980). Supercoils in prokaryotic DNA restrained in vivo. Proc. Natl. Acad. Sci. 77, 1331–5.CrossRefGoogle ScholarPubMed
Pohl, F. M. (1986). Dynamics of the B-to-Z transition in supercoiled DNA. Proc. Natl. Acad. Sci. U. S. A. 83, 4783–7.CrossRefGoogle ScholarPubMed
Pohl, F. M., Thomae, R. & Di Capua, E. (1982). Antibodies to Z-DNA interact with form V DNA. Nature 300, 545–6.CrossRefGoogle ScholarPubMed
Pulleyblank, D. E., Shure, M., Tang, D., Vinograd, J. & Vosberg, H. P. (1975). Action of nicking-closing enzyme on supercoiled and nonsupercoiled closed circular DNA: formation of a Boltzmann distribution of topological isomers. Proc. Natl. Acad. Sci. U. S. A. 72, 4280–4.CrossRefGoogle ScholarPubMed
Rahmouni, A. R. & Wells, R. D. (1989). Stabilization of Z DNA in vivo by localized supercoiling. Science 246, 358–63.CrossRefGoogle ScholarPubMed
Roca, J., Berger, J. M., Harrison, S. C. & Wang, J. C. (1996). DNA transport by a type II topoisomerase – direct evidence for a two-gate mechanism. Proc. Natl. Acad. Sci. U. S. A. 93, 93–4057.CrossRefGoogle ScholarPubMed
Roca, J. & Wang, J. (1994). DNA transport by a type II DNA topoisomerase – evidence in favor of a two-gate mechanism. Cell 77, 609–16.CrossRefGoogle ScholarPubMed
Rodriguez, A. C. & Stock, D. (2002). Crystal structure of reverse gyrase: insights into the positive supercoiling of DNA. EMBO J. 21, 418–26.CrossRefGoogle ScholarPubMed
Rolfsen, D. (1976). Knots and Links. Berkeley, CA: Publish or Perish.
Rudolph, M. G., del Toro Duany, Y., Jungblut, S. P., Ganguly, A. & Klostermeier, D. (2012). Crystal structures of Thermotoga maritima reverse gyrase: inferences for the mechanism of positive DNA supercoiling. Nucleic Acids Res. DOI: 10.1093/nar/gks1073Google ScholarPubMed
Rybenkov, V. V., Cozzarelli, N. R. & Vologodskii, A.V (1993). Probability of DNA knotting and the effective diameter of the DNA double helix. Proc. Natl. Acad. Sci. U. S. A. 90, 5307–11.CrossRefGoogle ScholarPubMed
Rybenkov, V. V., Ullsperger, C., Vologodskii, A. V. & Cozzarelli, N. R. (1997a). Simplification of DNA topology below equilibrium values by type II topoisomerases. Science 277, 690–3.CrossRefGoogle ScholarPubMed
Rybenkov, V. V., Vologodskii, A. V. & Cozzarelli, N. R. (1997b). The effect of ionic conditions on DNA helical repeat, effective diameter, and free energy of supercoiling. Nucleic Acids Res. 25, 25–1412.CrossRefGoogle ScholarPubMed
Rybenkov, V. V, Vologodskii, A. V & Cozzarelli, N. R. (1997c). The effect of ionic conditions on the conformations of supercoiled DNA. I. Sedimentation analysis. J. Mol. Biol. 267, 299–311.Google ScholarPubMed
Rybenkov, V. V., Vologodskii, A. V. & Cozzarelli, N. R. (1997d). The effect of ionic conditions on the conformations of supercoiled DNA. II. Equilibrium catenation. J. Mol. Biol. 267, 312–23.Google ScholarPubMed
Schoeffler, A. J. & Berger, J. M. (2008). DNA topoisomerases: harnessing and constraining energy to govern chromosome topology. Q. Rev. Biophys. 41, 41–101.CrossRefGoogle ScholarPubMed
Shaw, S. Y. & Wang, J. C. (1993). Knotting of a DNA chain during ring closure. Science 260, 533–6.CrossRefGoogle ScholarPubMed
Shimada, J. & Yamakawa, H. (1988). Moments for DNA topoisomers: the helical wormlike chain. Biopolymers 27, 657–73.Google ScholarPubMed
Sinden, R. R., Broyles, S. S. & Pettijohn, D. E. (1983). Perfect palindromic lac operator DNA sequence exists as a stable cruciform structure in supercoiled DNA in vitro but not in vivo. Proc. Natl. Acad. Sci. U. S. A. 80, 1797–801.CrossRefGoogle Scholar
Sinden, R. R., Carlson, J. O. & Pettijohn, D. E. (1980). Torsional tension in the DNA double helix measured with trimethylpsoralen in living E. coli cells: analogous measurements in insect and human cells. Cell 21, 773–83.CrossRefGoogle ScholarPubMed
Singleton, C. K., Klysik, J., Stirdivant, S. M. & Wells, R. D. (1982). Left-handed Z-DNA is induced by supercoiling in physiological ionic conditions. Nature 299, 312–6.CrossRefGoogle ScholarPubMed
Slesarev, A. I. (1988). Positive supercoiling catalysed in vitro by ATP-dependent topoisomerase from Desulfurococcus amylolyticus. Eur. J. Biochem. 173, 395–9.CrossRefGoogle ScholarPubMed
Stark, W. M. & Boocock, M. R. (1995). Topological selectivity in site-specific recombination. In Mobile Genetic Elements, ed. D. J., Sherratt, 101–29. Oxford: Oxford University Press.Google Scholar
Stewart, L., Redinbo, M. R., Qiu, X., Hol, W. G. & Champoux, J. J. (1998). A model for the mechanism of human topoisomerase I. Science 279, 1534–41.CrossRefGoogle ScholarPubMed
Stone, M. D., Bryant, Z., Crisona, N. J., Smith, S. B., Vologodskii, A., Bustamante, C. & Cozzarelli, N. R. (2003). Chirality sensing by Escherichia coli topoisomerase IV and the mechanism of type II topoisomerases. Proc. Natl. Acad. Sci. U. S. A. 100, 8654–9.CrossRefGoogle ScholarPubMed
Tse, Y. & Wang, J. C. (1980). E. coli and M. luteus DNA topoisomerase I can catalyze catenation of decatenation of double-stranded DNA rings. Cell 22, 269–76.CrossRefGoogle Scholar
Ullsperger, C. J., Vologodskii, A. V. & Cozzarelli, A. V. (1995). Unlinking of DNA by topoisomerases during DNA replication. Nucl. Acids Mol. Biol. 9, 115–42.Google Scholar
Vinograd, J., Lebowitz, J., Radloff, R., Watson, R. & Laipis, P. (1965). The twisted circular form of polyoma viral DNA. Proc. Natl. Acad. Sci. U. S. A. 53, 1104–11.CrossRefGoogle ScholarPubMed
Vinograd, J., Lebowitz, J. & Watson, R. (1968). Early and late helix-coil transitions in closed circular DNA. The number of superhelical turns in polyoma DNA. J. Mol. Biol. 33, 173–97.CrossRefGoogle ScholarPubMed
Vologodskaia, M. Y. & Vologodskii, A. V. (1999). Effect of magnesium on cruciform extrusion in supercoiled DNA. J. Mol. Biol. 289, 851–9.CrossRefGoogle ScholarPubMed
Vologodskii, A. 2007. Monte Carlo simulation of DNA topological properties. In Topology in Molecular Biology, ed. M., Monastryrsky, 23–41. Berlin: Springer.Google Scholar
Vologodskii, A. (2009). Theoretical models of DNA topology simplification by type IIA DNA topoisomerases. Nucleic Acids Res. 37, 3125–33.CrossRefGoogle ScholarPubMed
Vologodskii, A. V, Anshelevich, V. V., Lukashin, A. V. & Frank-Kamenetskii, M. D. (1979a). Statistical mechanics of supercoils and the torsional stiffness of the DNA. Nature 280, 294–298.CrossRefGoogle ScholarPubMed
Vologodskii, A. V. & Cozzarelli, N. R. (1994). Conformational and thermodynamic properties of supercoiled DNA. Annu. Rev. Biophys. Biomol. Struct. 23, 609–43.CrossRefGoogle ScholarPubMed
Vologodskii, A. V. & Cozzarelli, N. R. (1996). Effect of supercoiling on the juxtaposition and relative orientation of DNA sites. Biophys. J. 70, 2548–56.CrossRefGoogle ScholarPubMed
Vologodskii, A. V, Crisona, N. J., Laurie, B., Pieranski, P., Katritch, V., Dubochet, J. & Stasiak, A. (1998). Sedimentation and electrophoretic migration of DNA knots and catenanes. J. Mol. Biol. 278, 1–3.CrossRefGoogle ScholarPubMed
Vologodskii, A. V. & Frank-Kamenetskii, M. D. (1983). The relaxation time for a cruciform structure in superhelical DNA. FEBS Lett. 160, 173–6.CrossRefGoogle ScholarPubMed
Vologodskii, A. V., Levene, S. D., Klenin, K. V., Frank-Kamenetskii, M. & Cozzarelli, N. R. (1992). Conformational and thermodynamic properties of supercoiled DNA. J. Mol. Biol. 227, 227–1224.CrossRefGoogle ScholarPubMed
Vologodskii, A. V., Lukashin, A. V., Anshelevich, V. V. & Frank-Kamenetskii, M. D. (1979b). Fluctuations in superhelical DNA. Nucleic Acids Res. 6, 967–82.CrossRefGoogle ScholarPubMed
Vologodskii, A. V, Zhang, W., Rybenkov, V. V, Podtelezhnikov, A. A., Subramanian, D., Griffith, J. D. & Cozzarelli, N. R. (2001). Mechanism of topology simplification by type II DNA topoisomerases. Proc. Natl. Acad. Sci. U. S. A. 98, 3045–9.CrossRefGoogle ScholarPubMed
Wang, A. H., Gessner, R. V., van der Marel, G. A., van Boom, J. H. & Rich, A. (1985). Crystal structure of Z-DNA without an alternating purine-pyrimidine sequence. Proc. Natl. Acad. Sci. U.S.A. 82, 3611–5.CrossRefGoogle ScholarPubMed
Wang, J. C. (1971). Interaction between DNA and an Escherichia coli protein ω. J. Mol. Biol. 55, 523–33.CrossRefGoogle ScholarPubMed
Wang, J. C. (1974a). The degree of unwinding of the DNA helix by ethidium. I. Titration of twisted PM2 DNA molecules in alkaline cesium chloride density gradients. J. Mol. Biol. 89, 783–801.Google ScholarPubMed
Wang, J. C. (1974b). Interactions between twisted DNAs and enzymes: the effects of superhelical turns. J. Mol. Biol. 87, 797–816.CrossRefGoogle ScholarPubMed
Wang, J. C. (1996). DNA topoisomerases. Annu. Rev. Biochem. 65, 635–95.CrossRefGoogle ScholarPubMed
Wang, J. C. (1998). Moving one DNA double helix through another by a type II DNA topoisomerase: the story of a simple molecular machine. Q. Rev. Biophys. 31, 107–44.CrossRefGoogle ScholarPubMed
Wang, J. C. (2009). Untangling the Double Helix: DNA Entanglement and the Action of the DNA Topoisomerases. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Wang, J. C., Peck, L. J. & Becherer, K. (1983). DNA supercoiling and its effects on DNA structure and function. Cold SpringHarbor Symp. Quant. Biol. 47, 85–91.Google ScholarPubMed
Wang, J. C. & Schwartz, H. (1967). Noncomplementarity in base sequences between the cohesive ends of coliphages 186 and lambda and the formation of interlocked rings between the two DNAs. Biopolymers 5, 953–66.CrossRefGoogle Scholar
Wasserman, S. A. & Cozzarelli, N. R. (1986). Biochemical topology: applications to DNA recombination and replication. Science 232, 951–60.CrossRefGoogle Scholar
Wasserman, S. A., White, J. H. & Cozzarelli, N. R. (1988). The helical repeat of double-stranded DNA varies as a function of catenation and supercoiling. Nature 334, 448–50.CrossRefGoogle ScholarPubMed
Weil, R. & Vinograd, J. (1963). The cyclic helix and cyclic coil forms of polyoma viral DNA. Proc. Natl. Acad. Sci. U. S. A. 50, 730–8.CrossRefGoogle ScholarPubMed
White, J. H. (1969). Self-linking and the Gauss integral in higher dimensions. Am. J. Math. 91, 693–728.CrossRefGoogle Scholar
White, J. H., Cozzarelli, N. R. & Bauer, W. R. (1988). Helical repeat and linking number of surface-wrapped DNA. Science 241, 323–7.CrossRefGoogle ScholarPubMed
Wittig, B., Wolfl, S., Dorbic, T., Vahrson, W. & Rich, A. (1992). Transcription of human c-myc in permeabilized nuclei is associated with formation of Z-DNA in three discrete regions of the gene. EMBO J. 11, 4653–63.Google ScholarPubMed
Wolfl, S., Martinez, C., Rich, A. & Majzoub, J. A. (1996). Transcription of the human corticotropin-releasing hormone gene in NPLC cells is correlated with Z-DNA formation. Proc. Natl. Acad. Sci. U. S. A. 93, 3664–8.CrossRefGoogle ScholarPubMed
Wu, H.-Y., Shyy, S., Wang, J. C. & Liu, L. F. (1988). Transcription generates positively and negatively supercoiled domains in the template. Cell 53, 433–40.CrossRefGoogle ScholarPubMed
Yan, J., Magnasco, M. O. & Marko, J. F. (1999). A kinetic proofreading mechanism for disentanglement of DNA by topoisomerases. Nature 401, 932–5.CrossRefGoogle ScholarPubMed
Zacharias, W., Jaworski, A., Larson, J. E. & Wells, R. D. (1988). The B- to Z-DNA equilibrium in vivo is perturbed by biological processes. Proc. Natl. Acad. Sci. 85, 7069.CrossRefGoogle Scholar

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.

  • Circular DNA
  • Alexander Vologodskii, New York University
  • Book: Biophysics of DNA
  • Online publication: 05 March 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781139542371.007
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.

  • Circular DNA
  • Alexander Vologodskii, New York University
  • Book: Biophysics of DNA
  • Online publication: 05 March 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781139542371.007
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.

  • Circular DNA
  • Alexander Vologodskii, New York University
  • Book: Biophysics of DNA
  • Online publication: 05 March 2015
  • Chapter DOI: https://doi.org/10.1017/CBO9781139542371.007
Available formats
×