physics0609139/ctn.tex
1: \documentclass[aps,rmp,twocolumn]{revtex4}
2: 
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{graphicx}
6: \usepackage{psfrag}
7: \usepackage{mathrsfs}
8: \usepackage[usenames]{color}
9: 
10: \newcommand{\ecoli}{\emph{E.coli\/} }
11: \newcommand{\tr}{\mathsf{t}}
12: 
13: 
14: \newcommand{\AC}{{\it Acta Crystallogr.} }
15: \newcommand{\AM}{{\it Acta Metall.} }
16: \newcommand{\AP}{{\it Ann. Phys., Lpz.} }
17: \newcommand{\APNY}{{\it Ann. Phys., NY\/} }
18: \newcommand{\APP}{{\it Ann. Phys., Paris\/} }
19: \newcommand{\CJP}{{\it Can. J. Phys.} }
20: \newcommand{\JAP}{{\it J. Appl. Phys.} }
21: \newcommand{\JCP}{{\it J. Chem. Phys.} }
22: \newcommand{\JJAP}{{\it Japan. J. Appl. Phys.} }
23: \newcommand{\JP}{{\it J. Physique\/} }
24: \newcommand{\JPhCh}{{\it J. Phys. Chem.} }
25: \newcommand{\JMMM}{{\it J. Magn. Magn. Mater.} }
26: \newcommand{\JMP}{{\it J. Math. Phys.} }
27: \newcommand{\JOSA}{{\it J. Opt. Soc. Am.} }
28: \newcommand{\JPSJ}{{\it J. Phys. Soc. Japan\/} }
29: \newcommand{\JQSRT}{{\it J. Quant. Spectrosc. Radiat. Transfer\/} }
30: \newcommand{\NC}{{\it Nuovo Cimento\/} }
31: \newcommand{\NIM}{{\it Nucl. Instrum. Methods\/} }
32: \newcommand{\NP}{{\it Nucl. Phys.} }
33: \newcommand{\PL}{{\it Phys. Lett.} }
34: \newcommand{\PR}{{\it Phys. Rev.} }
35: \newcommand{\PRE}{{\it Phys. Rev. E.} }
36: \newcommand{\PRL}{{\it Phys. Rev. Lett.} }
37: \newcommand{\PRS}{{\it Proc. R. Soc.} }
38: \newcommand{\PS}{{\it Phys. Scr.} }
39: \newcommand{\PSS}{{\it Phys. Status Solidi\/} }
40: \newcommand{\PTRS}{{\it Phil. Trans. R. Soc.} }
41: \newcommand{\RMP}{{\it Rev. Mod. Phys.} }
42: \newcommand{\PNAS}{{\it Proc. Natl. Acad. Sci. U.S.A.\/} }
43: %
44: \newcommand{\etal}{{\it et al\/}\ }
45: \newcommand{\dash}{------}
46: \newcommand{\nonum}{\par\item[]}                %\par added 1/9/93
47: \newcommand{\mat}[1]{\underline{\underline{#1}}}
48: %
49: % abbreviations for IOPP journals
50: %
51: \newcommand{\CQG}{{\it Class. Quantum Grav.} }
52: \newcommand{\CTM}{{\it Combust. Theory Modelling\/} }
53: \newcommand{\DSE}{{\it Distrib. Syst. Engng\/} }
54: \newcommand{\EJP}{{\it Eur. J. Phys.} }
55: \newcommand{\HPP}{{\it High Perform. Polym.} }              % added 4/5/93
56: \newcommand{\IP}{{\it Inverse Problems\/} }
57: \newcommand{\JHM}{{\it J. Hard Mater.} }                    % added 4/5/93
58: \newcommand{\JO}{{\it J. Opt.} }
59: \newcommand{\JOA}{{\it J. Opt. A: Pure Appl. Opt.} }
60: \newcommand{\JOB}{{\it J. Opt. B: Quantum Semiclass. Opt.} }
61: \newcommand{\JPA}{{\it J. Phys. A: Math. Gen.} }
62: \newcommand{\JPB}{{\it J. Phys. B: At. Mol. Phys.} }      %1968-87
63: \newcommand{\jpb}{{\it J. Phys. B: At. Mol. Opt. Phys.} } %1988 and onwards
64: \newcommand{\JPC}{{\it J. Phys. C: Solid State Phys.} }   %1968--1988
65: \newcommand{\JPCM}{{\it J. Phys.: Condens. Matter\/} }    %1989 and onwards
66: \newcommand{\JPD}{{\it J. Phys. D: Appl. Phys.} }
67: \newcommand{\JPE}{{\it J. Phys. E: Sci. Instrum.} }
68: \newcommand{\JPF}{{\it J. Phys. F: Met. Phys.} }
69: \newcommand{\JPG}{{\it J. Phys. G: Nucl. Phys.} }         %1975--1988
70: \newcommand{\jpg}{{\it J. Phys. G: Nucl. Part. Phys.} }   %1989 and onwards
71: \newcommand{\MSMSE}{{\it Modelling Simulation Mater. Sci. Eng.} }
72: \newcommand{\MST}{{\it Meas. Sci. Technol.} }                 %1990 and onwards
73: \newcommand{\NET}{{\it Network: Comput. Neural Syst.} }
74: \newcommand{\NJP}{{\it New J. Phys.} }
75: \newcommand{\NL}{{\it Nonlinearity\/} }
76: \newcommand{\NT}{{\it Nanotechnology} }
77: \newcommand{\PAO}{{\it Pure Appl. Optics\/} }
78: \newcommand{\PM}{{\it Physiol. Meas.} }                        % added 4/5/93
79: \newcommand{\PMB}{{\it Phys. Med. Biol.} }
80: \newcommand{\PPCF}{{\it Plasma Phys. Control. Fusion\/} }      % added 4/5/93
81: \newcommand{\PSST}{{\it Plasma Sources Sci. Technol.} }
82: \newcommand{\PUS}{{\it Public Understand. Sci.} }
83: \newcommand{\QO}{{\it Quantum Opt.} }
84: \newcommand{\QSO}{{\em Quantum Semiclass. Opt.} }
85: \newcommand{\RPP}{{\it Rep. Prog. Phys.} }
86: \newcommand{\SLC}{{\it Sov. Lightwave Commun.} }               % added 4/5/93
87: \newcommand{\SST}{{\it Semicond. Sci. Technol.} }
88: \newcommand{\SUST}{{\it Supercond. Sci. Technol.} }
89: \newcommand{\WRM}{{\it Waves Random Media\/} }
90: \newcommand{\JMM}{{\it J. Micromech. Microeng.\/} }
91: \newcommand{\NAT}{{\it Nature\/} }
92: 
93: 
94: \begin{document}
95: 
96: \title{Single DNA conformations and biological function}
97: 
98: \author{Ralf Metzler}
99: \email{metz@nordita.dk}
100: \affiliation{NORDITA---Nordic Institute for Theoretical Physics,
101: Blegdamsvej 17, 2100 Copenhagen \O, Denmark}
102: \affiliation{Department of Physics, University of Ottawa, 150 Louis
103: Pasteur, Ottawa, Ontario, K1N 6N5, Canada (New address)}
104: \author{Tobias Ambj{\"o}rnsson}
105: \affiliation{NORDITA---Nordic Institute for Theoretical Physics,
106: Blegdamsvej 17, 2100 Copenhagen \O, Denmark}
107: \affiliation{Department of Chemistry, Massachusetts Institute of Technology,
108: 77 Massachusetts Avenue, Cambridge, Massachusetts 02139 (New address)}
109: \author{Andreas Hanke}
110: \affiliation{Department of Physics and Astronomy, University of Texas at
111: Brownsville, 80 Fort Brown, Brownsville, TX 78520}
112: \affiliation{Institute of Biomedical Sciences and Technology, University of
113: Texas at Dallas, Richardson, TX 75083}
114: \author{Yongli Zhang}
115: \affiliation{Department of Physiology and Biophysics, Albert Einstein
116: College
117: of Medicine, Bronx, NY 10461}
118: \author{Stephen Levene}
119: \affiliation{Institute of Biomedical Sciences and Technology and
120: Department of Molecular and Cell Biology, University of Texas at Dallas,
121: Richardson, TX 75083}
122: 
123: \begin{abstract}
124: Abstract. From a nanoscience perspective, cellular processes and their
125: reduced in vitro imitations provide extraordinary examples for highly
126: robust few or single molecule reaction pathways. A prime example are
127: biochemical reactions involving DNA molecules, and the coupling of these
128: reactions to the physical conformations of DNA. In this review,
129: we summarise recent results on the following phenomena: We investigate the
130: biophysical properties of DNA-looping and the
131: equilibrium configurations of DNA-knots, whose relevance to biological
132: processes are increasingly appreciated. We discuss how random DNA-looping
133: may be related to the efficiency of the target search process of proteins for
134: their specific binding site on the DNA molecule. And we dwell on the
135: spontaneous formation of intermittent DNA nanobubbles and their importance
136: for biological processes, such as transcription initiation. The physical
137: properties of DNA may indeed turn out to be particularly suitable for the
138: use of DNA in nanosensing applications.\\[0.2cm]
139: Key words: DNA, single molecules, DNA looping, DNA denaturation, knots,
140: gene regulation
141: \end{abstract}
142: 
143: \maketitle
144: 
145: \vspace*{8cm}
146: 
147: \newpage
148: \clearpage
149: 
150: \tableofcontents
151: 
152: \clearpage
153: 
154: \section{Introduction}
155: 
156: Deoxyribonucleic acid (DNA) is the molecule of life as we know
157: it.\footnote{Our DNA world during
158: biotic and prebiotic evolution was supposedly preceded by an RNA world and,
159: quite likely, by sugarless nucleic acids.} It contains all information of
160: an entire organism.\footnote{A small fraction of genetic information is stored
161: on DNA that is kept at other regions of the cell and not replicated on cell
162: division, such as mitochondrial or ribosomal DNA.}
163: This information is copied during cell division with an extremely
164: high fidelity by the replication mechanism. Despite the rather high
165: chemical and physical stability of DNA, due to constant action of enzymes
166: and other binding proteins (mismatches, rupture) as well as potential
167: environmentally induced damage (radiation, chemicals), this low error rate,
168: i.e., the suppression of the liability to mutations, is only possible with the
169: constant action of repair mechanisms \cite{alberts,snustad,kornberg,kornberg1}.
170: Although DNA's structural and mechanical properties are rather well established
171: for isolated DNA molecules (starting with Rosalind Franklin's X-ray diffraction
172: images \cite{franklin}), the characterisation
173: of DNA in its cellular environment, and even in vitro during interaction with
174: binding proteins, is subject of ongoing investigations.
175: 
176: %\begin{figure}
177: %\includegraphics[width=8cm]{franklin.eps}
178: %\caption{The diffraction photograph of the B form of DNA taken by Rosalind
179: %Franklin in May 1952 \cite{franklin}. Data derived from this photograph were
180: %instrumental in allowing James Watson and Francis Crick to construct their
181: %Nobel Prize­winning model for DNA \cite{watsoncrick}. $\copyright$ Nature.
182: %\label{franklin}}
183: %\end{figure}
184: 
185: Recent advances in experimental techniques such as fluorescence methods,
186: atomic force microscopy, or optical tweezers have leveraged the potential
187: to both probe and manipulate the equilibrium and out of equilibrium
188: behaviour of \emph{single\/} DNA molecules, making it possible to explore DNA's
189: physical and mechanical properties as well as its interaction with other
190: biopolymers, such as the DNA-protein interplay during gene regulation or repair
191: processes. An important ingredient is the coupling to thermal activation due
192: to the highly
193: Brownian environment. Although mostly performed in vitro, these experiments
194: provide access to increasingly refined information on the nature of DNA and
195: its environment-controlled behaviour.
196: 
197: In addition to chromosomal packaging inside the nucleus of eukaryotic cells
198: and the concentration of DNA in the membraneless nucleoid region of
199: prokaryotes, the global
200: structure of the DNA molecule can be affected by topological entanglements.
201: Thus, by error or design a DNA molecule can attain a knotted or concatenated
202: state, reducing or inhibiting biologically relevant functions, for instance,
203: replication or transcription. Such entangled states
204: can be actively reduced by enzymes of the topoisomerase family. Their precise
205: action, in particular, how they determine the presence of an entangled state,
206: is not fully known. Current studies therefore aim at shedding light on
207: possible mechanisms, in particular, in view of the importance of
208: topoisomerase action (or better, its inhibition) in tumour proliferation.
209: Other applications may be directed towards the treatment of viral deceases
210: by modifying the packaging of viral DNA to create knots in the virus capsid
211: and prevent ejection of the DNA into a host, and thereby infection.
212: DNA knots are also being recognised as a potential complication in the use
213: of nanochannels for DNA separation and sequencing. In such confined geometries
214: DNA knots are created with appreciable probability, affecting the reliability
215: of these techniques. Similarly to DNA knots, DNA looping is intimately
216: connected to the function of DNA. Current results on DNA looping and DNA
217: knot behaviour are summarised in the first parts of this review.
218: 
219: The Watson-Crick double-helix represents the thermodynamically stable state
220: of DNA at moderate salt concentrations and below the melting temperature.
221: This stability is effected by Watson-Crick hydrogen bonding and the
222: stronger base stacking of neighbouring base-pairs (bps). However, even at room
223: temperature DNA locally opens up intermittent flexible single-stranded domains,
224: so-called DNA-bubbles. Their size typically ranges from a few broken
225: bps, increasing to some 200 broken bps closer to the
226: melting temperature. The thermal melting of DNA has traditionally been used
227: to obtain the sequence-dependent stability parameters of DNA. More recently,
228: the role of intermittent bubble domains has been investigated with respect
229: to the liability of DNA-denaturation induced by proteins that selectively
230: bind to single-stranded DNA. It has been speculated that due to the liability
231: to denaturation of the TATA motif bubble formation may add in
232: transcription initiation. The dynamics of single bubbles can be monitored
233: by fluorescence methods, opening a window to both study the breathing of
234: DNA experimentally, but also to obtain high precision DNA stability data.
235: Finally, bubble dynamics has been suggested as a useful tool in optical
236: nanosensing. DNA breathing is the topic of the second part of this work.
237: 
238: Essentially all the biological functions of DNA rely on site-specific
239: DNA-binding proteins locating their targets (cognate sites) on the DNA
240: molecule, and therefore require searching through megabases of non-target
241: DNA in a highly efficient manner. For instance, gene regulation is performed
242: by specific regulatory proteins. On binding to a promoter area on the DNA,
243: they recruit or inhibit binding of RNA polymerase and subsequent
244: transcription of the associated gene. The search for the cognate site is
245: in fact facilitated by the DNA molecule: in addition to three-dimensional
246: search it enables the proteins to also move one-dimensionally along the
247: DNA while being non-specifically bound. Moreover, at points where the DNA
248: loops back on itself, this polymeric conformation provides shortcuts for
249: the proteins in the chemical coordinate along the DNA, approximately giving
250: rise to search-efficient L{\'e}vy flights. Target search is currently a very
251: active field of research, and single molecule methods have been shown to
252: provide essential new information. Moreover, the architecture of more complex
253: promoters relying on the simultaneous presence of several regulatory
254: proteins is being investigated to create in silico circuits for highly
255: sensitive chemical probes in small volumes. Such nanosensing applications
256: are expected to be of great importance in microarrays or other nano- and
257: microapplications. The third part of this review deals with diffusional
258: aspects of gene regulation.
259: 
260: At the same time DNA's role in classical polymer physics is increasingly
261: appreciated. With the possibility to reproduce DNA with extremely low
262: error rate by the PCR\footnote{Polymerase Chain Reaction: thermal
263: denaturation of a DNA molecule into two single strands and subsequent
264: cooling in a solution of single nucleotides and invariable primers,
265: produces two new complete double-stranded DNA molecules. Cycling of this
266: process produces large, monodisperse quantities of DNA.}, monodisperse
267: samples can be prepared. While shorter single-stranded DNA can be
268: used as a model for flexible polymers, the double strand exhibits a
269: semiflexible behaviour with a persistence length, that can be easily
270: probed experimentally. Moreover, DNA
271: is orders of magnitude longer than conventional polymers. Combined with
272: the potential of single molecule probing, DNA is advancing as a model
273: polymer.
274: 
275: After an introduction to the properties of DNA we address these functional
276: properties of DNA from the perspective of biological relevance, physical
277: behaviour and nanotechnological potential. Most emphasis will be put on
278: the single molecular aspects of DNA. We note that this is not intended to
279: be an exhaustive review on the physical properties of DNA. Rather, we
280: present some important features and their consequences from a personal
281: perspective.
282: 
283: 
284: \section{Physical properties and biological function of DNA}
285: 
286: Biomolecules, that occur naturally in biological systems, can be grouped
287: into unspecific oligo- and macromolecules and biopolymers in the stricter sense
288: \cite{alberts}. Unspecific biomolecules are
289: produced by biological organisms in a large range of molecular weight and
290: structure, such as polysaccharides (cellulose, chitin, starch, etc.),
291: higher fatty acids, actin filaments or microtubules. Also
292: the natural `india-rubber' from the {\em Hevea Brasiliensis\/} tree,
293: historically important for both industrial purposes and the development of
294: polymer physics \cite{treloar} belongs to this group.
295: 
296: \begin{figure}
297: \includegraphics[width=6.4cm]{mycrick.eps}
298: \caption{Central dogma of molecular biology after F. Crick:
299: Potentially, information flow is completely symmetric
300: between the three levels of cellular biopolymers (DNA, RNA, proteins).
301: However, the recognised pathways are only those represented here,
302: where solid lines represent probable transfers, and dotted lines for
303: (in principle) possible transfers \protect\cite{crick}.
304: \label{crick}}
305: \end{figure}
306: 
307: Biopolymers in the stricter sense we are going to assume here comprise
308: the polynucleotides DNA and RNA consisting of the four-letter nucleotide
309: alphabet with A-T and G-C (A-U and G-C for RNA) bps, and the
310: polypeptidic proteins consisting of 20 different amino acids, each coded for
311: by 3 bases (codons) in the RNA \cite{alberts,snustad,kornberg,kornberg1}.
312: We will come back to proteins later when reviewing binding protein-DNA
313: interactions. Biopolymers are copied and/or created according
314: to the information flow sketched in figure \ref{crick}, the so-called central
315: dogma of molecular biology, a term originally coined by Frances Crick
316: \cite{crick1}. Accordingly, starting from the genetic code stored in the DNA
317: (in some cases in RNA) DNA is copied by DNA polymerase (replication), and
318: the proteins as the actually task-performing biopolymers are created via
319: messenger RNA (created by DNA transcription through RNA polymerase) and
320: further by translation in ribosomes to proteins.\footnote{Alternatively, the
321: genetic code can be transcribed into transfer and ribosomal RNA that is not
322: translated into proteins.}
323: 
324: DNA is made up of the four bases
325: \cite{alberts,snustad,kornberg,kornberg1,bloomfield,rnaworld}:
326: A(denine), G(uanine), C(ytosine), and
327: T(hymine) that form the DNA ladder structure shown in figure \ref{dh}.
328: \begin{figure}
329: \includegraphics[height=7.68cm]{dna_bb.eps}
330: \caption{Ladder structure of the DNA formed by its four building-blocks
331: A, G, C, and T, giving rise to the typical double-helical structure
332: of DNA. A-T bps establish 2 H-bonds, G-C bps 3 H-bonds.
333: \label{dh}}
334: \end{figure}
335: These building-blocks A, G, C, T bp according to the key-lock principle
336: as A-T and G-C, where the AT bond is weaker than the GC bond in terms of
337: stability. Apart from the Watson-Crick base-pairing energy, the stability
338: of dsDNA is effected by the stacking interactions, the specific matching
339: of subsequent bps along the double-strand, i.e., bp-bp
340: interactions. In standard literature, the stacking interactions are listed
341: for pairs of bps (e.g., for AT-GC, AT-AT, AT-TA, etc.), see
342: below.\footnote{Longer ranging bp-bp interactions are most likely small in
343: comparison.}
344: 
345: \begin{figure*}
346: \includegraphics[width=18cm]{Ecoli_structureatlas.ps}
347: \caption{Structure atlas of the \ecoli genome. Figure courtesy David Ussery,
348: Technical University of Denmark. The structure atlas is available under
349: the URL {\tt www.cbs.dtu.dk/services/GenomeAtlas/}.
350: \label{structure}}
351: \end{figure*}
352: 
353: Based on this AGCT alphabet, the
354: primary structure of DNA can be specified. DNA's six local structural elements
355: twist, tilt, roll, shift, slide, and rise are effected by the stacking
356: interactions between vicinal bps. In figure \ref{structure}, we show a map
357: with the structure elements of the entire \ecoli genome, demonstrating the
358: degree of structural information currently available. These structural
359: elements define the local geometrical structure of DNA within a typical
360: correlation (persistence) length\footnote{The persistence length of a polymer
361: chain defines the characteristic length scale above which the polymer is
362: susceptible to bending induced by thermal fluctuations, i.e., it is the
363: length scale above which the tangent-tangent correlation decays along the
364: chain, see the Appendix.} of about 150 bps corresponding to 50 nm
365: (the bp-bp distance measures 3.4 {\AA}, reflecting the rather complex chemical
366: structure of a nucleotide in comparison to the monomer size of man-made
367: polymers such as polyethylene) \cite{frank,frank1,marko,marko1}.
368: On a larger scale, much longer than the persistence length,
369: DNA becomes flexible. On this level, tertiary structural elements come
370: into play. One example is DNA looping, that is the formation of polymeric
371: lasso loops induced by chemical bonds between binding proteins attached to
372: the DNA at specific bps which are remote along the DNA backbone
373: \cite{alberts,snustad,ptashne1,revet,bell,bell1,hame_looping}.
374: An extreme limit of tertiary structure is the
375: packaging of DNA onto histones and further wrapping into the chromosomes
376: of eukaryotic cells \cite{schiessel,kreth,alberts}. At the same time, dsDNA may
377: locally open into floppy ssDNA bubbles, with a persistence length of a few
378: bases.\footnote{In fact, it has been questioned whether there is a meaningful
379: value of the persistence length of ssDNA at all, due to its significant
380: apparent sequence dependence \cite{goddard}.} These fluctuation-induced bubbles
381: increase their statistical weight at higher temperatures, until the dsDNA fully
382: denatures (melts). We will come back to DNA denaturation
383: bubbles below. Depending on the
384: external conditions, DNA occurs in several configurations. Under physiological
385: conditions,
386: one is concerned with B-DNA, but there are other states such as A, B', Z, ps,
387: triplex DNA, quadruplex DNA, cruciform, and H, reviewed, for instance, in
388: \cite{frank,frank1}. DNA occurs naturally in a large
389: range of length scales. In viruses, DNA is of the order of a few $\mu$m
390: long. In bacteria, it already reaches lengths of several mm, and in mammalian
391: cells it can reach the order of a few m, roughly 2 m in a human cell and 35 m
392: in a cell of the South American lungfish, albeit split up into the
393: individual chromosomes \cite{kornberg}. DNA in bacteria in
394: vivo, or extracted from bacteria and higher cells for our purposes can
395: therefore be viewed a fully flexible polymer with a persistence length of
396: roughly 50 nm, being governed by generic effects independent of the detailed
397: sequence. On short scales DNA becomes semiflexible and governed by the
398: worm-like chain model (Kratky-Porod model) \cite{grosberg}; on even shorter
399: scales, local structural elements become important (in particular, for
400: recognition by binding proteins \cite{alberts}), and eventually molecular
401: resolution is reached.
402: 
403: Stacking interactions govern the local structure of dsDNA. Globally,
404: an additional constraint arises due to the circular nature of the DNA,
405: since it has to satisfy the conservation law \cite{calu,white1,fuller}
406: \begin{equation} \label{ltw}
407: {\rm Lk}={\rm Tw}+{\rm Wr},
408: \end{equation}
409: where Lk stands for the linking number, Tw for the twist, and Wr for the
410: writhe of the double helix. The linking number $Lk$ is an integer and
411: formally
412: given by one-half the number of signed crossings of one DNA strand with the
413: other in any regular projection of the molecule. $Lk$ is a topological
414: property, and no deformation of a closed DNA, without breaking and rejoining
415: the DNA strands, will alter it. $Tw$ is equal to the number
416: of times that the two strands of DNA wind about the central
417: axis of the molecule, and $Wr$ is a number whose absolute value equals
418: approximately the number of times that the DNA axis winds about
419: itself.\footnote{For details about the calculation of Tw and Wr for
420: representative models of DNA, see \cite{white}.}
421: Whereas Tw is a property of the double-helical structure of DNA,
422: Wr is a property of the DNA axis alone. Tw and Wr do not need to
423: be integers and are not conserved, but coupled through
424: Lk by equation (\ref{ltw}).
425: A nicked circular DNA, i.e., when the twist can fully relax, carries
426: ${\rm Lk}_0=N/h$ links, where $N$ is the number of bp and $h$ ($h\simeq 10.5$
427: in B-DNA) the number of bps per turn.
428: 
429: \begin{figure}
430: \includegraphics[width=7.6cm]{wind1.ps}
431: \caption{Right-handed (negative), normal, and left-handed (positive)
432: superhelix. The DNA of virtually all terrestrial organisms is
433: negatively supercoiled.}
434: \label{wind1}
435: \end{figure}
436: 
437: \begin{figure}
438: \includegraphics[height=6.0cm]{plasmid1.eps}
439: \includegraphics[height=6.0cm]{plasmid2.eps}
440: \caption{Electron micrographs of nicked (left) and supercoiled (right)
441: 6996-bp plasmid DNAs. The supercoiled example is from a population of DNA
442: molecules with an average superhelix density, $\bar {\sigma }$= -0.027,
443: close to the value expected \textit{in vivo}.}
444: \label{wind2}
445: \end{figure}
446: 
447: The degree of \emph{supercoiling\/} of DNA can be expressed in terms of
448: the linking number difference, $\Delta Lk=Lk-Lk_0$.
449: The DNA of virtually all terrestrial organisms is underwound or
450: {\em negatively supercoiled}, i.e., $\Delta Lk<0$
451: (figures \ref{wind1} and \ref{wind2}).\footnote{An exception are
452: thermophilic organisms living near
453: undersea geothermal vents that have positively supercoiled DNA in order to
454: stabilise the double helix at extreme temperatures.}
455: Often, the superhelical density $\sigma=\Delta Lk/Lk_0$ is used; most
456: supercoiled DNA molecules isolated from either prokaryotes or eukaryotes
457: have $\sigma$ values between $-0.05$ and $-0.07$ \cite{bauer}.
458: Negative supercoiling is regulated in prokaryotes by DNA gyrase; eukaryotes
459: lack gyrase but maintain negative supercoiling through winding of DNA around
460: nucleosomes and interactions with DNA-unwinding proteins.
461: There are two forms of intracellular supercoiling, the {\it plectonemic}
462: form, characteristic of plasmid DNA and accessible, nucleosome-free
463: regions of chromatin, and the {\it toroidal} or {\it solenoidal} form,
464: where supercoiling is
465: attained by DNA wrapped around histone octamers or prokaryotic non-histone
466: DNA-binding proteins (figure \ref{wind3}).
467: The former is the active form of supercoiled DNA and
468: is freely accessible to proteins involved in transcription, replication,
469: recombination and DNA repair. The latter is the stored form of supercoiled
470: DNA and is largely responsible for the extraordinary degree of compaction
471: required to condense typical genomes into the cell's nucleus.\footnote{The
472: nucleus of a human cell has a radius of circa 5 $\mu$m and stores the 2 m
473: of the human genome \cite{sun}.}
474: Negative supercoiling facilitates the local unwinding of DNA by providing
475: a ubiquitous source of free energy that augments the unwinding free energy
476: accompanying the interactions of many proteins with their cognate DNA
477: sequences. The local unwinding of DNA, in turn, is an integral part of many
478: biological processes such as gene regulation and DNA replication
479: (see section \ref{generegulation}). Therefore, understanding the interplay of
480: supercoiling and local helical structure is essential to the
481: understanding of biological mechanisms
482: \cite{marko1,benham,goetze,levene,levene1}.
483: 
484: \begin{figure}
485: \includegraphics[width=7.2cm]{wind3.ps}
486: \caption{Toroidal (left) and plectonemic (right) forms of supercoiled DNA.}
487: \label{wind3}
488: \end{figure}
489: 
490: Ribonucleic acid (RNA)
491: consists of the same building blocks as DNA, with the exception
492: that T(hymine) is replaced by U(racile) \cite{rnaworld}. RNA typically occurs
493: in single-stranded form. Therefore, its secondary structure is richer, being
494: characterised by sequences of hairpins: Smaller regions in which
495: chemically remote sequences
496: of bases match, pair and form hairpins which are stiff and energy-dominated,
497: similar to dsDNA. The remaining regions form entropy-dominated floppy loops,
498: analogous to the ssDNA bubbles. Additional
499: tertiary structure in RNA comes about by the formation of so-called
500: pseudoknots, chemical bonds established between bases sitting on chemically
501: distant segments of the secondary structure. In RNA-modelling
502: the incorporation of pseudoknots is a non-trivial problem, which currently
503: receives considerable interest; see, for instance,
504: references \cite{rnaworld,orland,orland1,baiesips}.
505: 
506: 
507: \section{DNA-looping}
508: 
509: The formation of DNA loops mediated by proteins bound at distant sites 
510: along a single molecule is an essential mechanistic aspect of many biological 
511: processes including gene regulation, DNA replication, and recombination 
512: (for reviews, see \cite{loop1,loop2}).
513: In {\em E. coli}, DNA looping represses gene expression 
514: at the {\em ara}, {\em gal}, {\em lac}, and {\em deo} 
515: operons \cite{loop3,loop4,loop5,loop6}
516: and activates transcription from 
517: the {\em gln}ALG operon \cite{loop7}.  
518: The size of DNA loops formed in these systems varies 
519: between approximately 100 and 600 bps.  In eukaryotes, a variety of
520: transcription factors bind to enhancers that are hundreds to several 
521: thousand bps away from their promoters and interact with RNA 
522: polymerases directly or through mediators in order to achieve combinatorial 
523: gene regulation \cite{loop8}.
524: DNA looping is required to juxtapose two recombination 
525: sites in intramolecular site-specific recombination 
526: \cite{loop9,loop10,loop11} and is also
527: employed by a number of restriction endonucleases such 
528: as {\em Sfi}I and {\em Ngo}MIV, 
529: which recognise and cut two copies of well-separated cognate sites 
530: simultaneously \cite{loop12,loop13,loop14}.
531: Here we describe a recent statistical-mechanical theory of 
532: loop formation that 
533: connects global mechanical and geometric properties of both DNA and 
534: protein and demonstrates the importance of protein flexibility in 
535: loop-mediated protein-DNA interactions \cite{loop26,loop51}.
536: 
537: 
538: \subsection{Biological significance of DNA looping}
539: \label{loopbio}
540: 
541: The biological importance of DNA loop formation 
542: is underscored by the abundance of architectural proteins in the cell such 
543: as HU, IHF, and HMG, which facilitate looping by bending the intervening
544: DNA between protein-recognition sites \cite{loop15}.
545: Moreover, DNA looping has been 
546: shown to be subject to regulation through the binding of effector 
547: molecules that alter protein conformation or protein-DNA 
548: interactions \cite{loop16}.
549: 
550: Two characteristics of DNA looping have been demonstrated by {\em in vitro}
551: and {\em in vivo} experiments. One is cooperative binding of a protein 
552: to its two cognate sites, which can be demonstrated by footprinting 
553: methods \cite{loop17}.  
554: DNA looping can increase the occupancies of both binding sites; in particular,
555: it can significantly enhance protein association to the lower-affinity site
556: because of the tethering effect of DNA looping.  This is a general mechanism 
557: by which many transcription factors recruit RNA polymerases in gene
558: regulation. Another hallmark is the helical dependence of loop formation
559: \cite{loop1,loop3}, which arises because of DNA's limited torsional 
560: flexibility and the requirement for correct 
561: torsional alignment of the two protein-binding sites.
562: Although many methods have been developed to directly observe DNA looping 
563: {\em in vitro}, such as scanning-probe \cite{loop7}
564: and electron microscopy \cite{bloomfield}, and single-molecule techniques
565: \cite{loop19}, assays based on helical dependence have 
566: been the only way to identify DNA looping {\em in vivo}. In these experiments, 
567: the DNA length between two protein binding sites is varied and the yield 
568: of DNA loop formation is monitored, for example by the repression or 
569: activation of a reporter gene \cite{loop20}.
570: Using this helical-twist assay, DNA 
571: looping in the {\em ara} operon was first discovered \cite{loop3}.
572: 
573: Our knowledge about the roles of DNA bending, twist, and their respective 
574: energetics in DNA looping has come largely from analyses of DNA 
575: cyclisation \cite{loop1,loop21,loop22}.
576: Circularisation efficiencies of DNA fragments, which are 
577: quantitatively described by $J$-{\em factors}, 
578: oscillate with DNA length and therefore torsional 
579: phase \cite{loop23,loop24}. The $J$-factor is defined as the
580: ratio of the partition function of a circularised polymer chain
581: to that of an open chain. Since there is a dimension reduction due to
582: circularisation constraints (two polymer ends have to meet), the ratio
583: has a unit of concentration, or $1/ L^3$ with $L$ representing length;
584: see \cite{loop26} for details.
585: In the present context, the $J$-factor is equal to the free DNA-end
586: concentration whose bimolecular ligation efficiency
587: equals that of the two ends of a cyclising DNA molecule \cite{loop25}.
588: For short DNA fragments $J$-factors are 
589: limited by the significant bending and twisting energies required to form 
590: closed circles, whereas for long DNA, the chain entropy loss during
591: circularisation exceeds the elastic-energy decrease and reduces the 
592: $J$-factor.  Because of this competition between bending and twisting 
593: energetics and entropy, there is an optimal DNA length for 
594: cyclisation \cite{loop26}.
595: Analogous behaviour has been expected for DNA looping, especially with 
596: respect to the helical dependence discussed above.
597: 
598: Quantitative analyses of DNA looping and cyclisation are challenging 
599: problems in statistical mechanics and have been largely limited to 
600: Monte Carlo or Brownian dynamics simulations 
601: \cite{loop27,loop28,loop29,loop30,loop31}. Analytical 
602: solutions are available only for some ideal and special cases.
603: An important contribution in this area is the theory of Shimada and 
604: Yamakawa \cite{loop32}, which is based on a 
605: homogeneous and continuous elastic 
606: rod model of DNA.  This theory has been applied extensively to DNA 
607: cyclisation \cite{loop23,loop33} and also DNA looping 
608: \cite{loop21,loop22,loop34}. The Shimada-Yamakawa
609: theory makes use of a perturbation approach, in which small configurational 
610: fluctuations of a DNA chain around the most probable configuration are 
611: accounted for in the evaluation of the partition function.
612: 
613: The elastic-equilibrium conformation is obvious for the homogeneous DNA 
614: circle studied by Shimada and Yamakawa \cite{loop32}.
615: However, the search for the 
616: elastic-energy minimum of homogeneous DNA molecules with complex geometry, 
617: such as in DNA looping, supercoiling, and the case of inhomogeneous DNA 
618: sequences containing curvature and nonuniform DNA flexibility, is not 
619: trivial \cite{loop4,loop35,loop36}. 
620: Recently, a statistical-mechanical theory for sequence-dependent 
621: DNA circles has been developed \cite{loop26} and applied to the
622: problem of DNA cyclisation \cite{loop26} and DNA looping \cite{loop51}. 
623: In this model, the DNA configuration is described by 
624: parameters defined at dinucleotide steps, i.e., tilt, roll, and twist, 
625: which allows straightforward incorporation of intrinsic or protein-induced 
626: DNA curvature at the bp level. Following Shimada and Yamakawa's
627: method, the theory first determines the mechanical equilibrium configuration
628: in small DNA circles (i.e., less than $\sim 1000$ bp) under certain
629: constraints; fluctuations around the equilibrium configuration are then
630: taken into account using an harmonic approximation.
631: The new method is much more computationally 
632: efficient than Monte Carlo simulation, has comparable accuracy, 
633: and has been applied successfully to analyse experimental results from 
634: DNA cyclisation \cite{loop26}. 
635: 
636: The basis of the extension of the model
637: to DNA looping \cite{loop51} is 
638: to treat the protein subunits as connected rigid bodies and to allow for 
639: a limited number of degrees of freedom between the subunits.  Motions of 
640: the subunits are assumed to be governed by harmonic potentials and an 
641: associated set of force constants, neglecting the anharmonic terms often 
642: required for proteins undergoing large conformational fluctuations among 
643: their modular domains.  Indeed, the use of a harmonic approximation is 
644: supported by the success of continuum elastic models that are based only 
645: on shape and mass-distribution information in descriptions of protein 
646: motion \cite{loop37}.
647: Similar to the description used for individual DNA bps 
648: in the model, protein geometry and dynamics are described by three 
649: rigid-body
650: rotation angles (tilt, roll, and twist).  Therefore, DNA looping can be 
651: viewed as a generalisation of DNA cyclisation in which the protein component 
652: is characterised by a particular set of local geometric constraints and 
653: elastic constants. This treatment not only unifies the theoretical 
654: descriptions of DNA cyclisation and looping, but also allows consideration 
655: of flexibilities at protein-DNA and protein-protein interfaces and 
656: application of the concepts of linking number and writhe. In previous 
657: work, proteins were considered rigid and their effects on DNA configuration 
658: were represented by a set of constraints applied to DNA ends 
659: \cite{loop1,loop38,loop39}.
660: With the present approach, programs developed for analysing DNA cyclisation 
661: can be used to analyse DNA looping with only minor modifications.
662: 
663: The new method \cite{loop26,loop51} is most applicable to the problem 
664: of short DNA loops, in which 
665: the free energy of a wormlike chain is dominated by bending and torsional 
666: elasticity \cite{loop26,loop51}.
667: Possible modes of 
668: DNA self contact and contacts between protein and DNA at positions other 
669: than the binding sites are not considered.
670: For large loops 
671: contributions to the free energy from chain entropy and DNA-DNA contacts 
672: can become highly significant. Several alternative treatments of DNA 
673: looping have appeared recently. One of these addresses the excluded-volume 
674: contribution to DNA looping within large open-circular molecules 
675: \cite{hame_looping},
676: whereas two others consider the effect on looping of traction at the ends 
677: of a DNA chain \cite{loop41,loop42}.
678: None of these treatments includes helical phasing
679: effects on DNA looping.  In contrast, a method based on the Kirchhoff 
680: elastic-rod model, which includes the helical-phase dependence, has been 
681: presented \cite{loop39,loop43}.
682: However, this approach does not include thermal 
683: fluctuations {\em per se} and therefore is not directly applicable to 
684: calculations of the $J$-factor.  The comprehensive treatment of small DNA
685: loops described in \cite{loop26,loop51}
686: is thus far unique to the extent that it accounts for 
687: sequence- and protein-dependent conformational and flexibility parameters, 
688: thermal fluctuations, and helical phasing effects.
689: 
690: 
691: \subsection{DNA loop model}
692: 
693: The protein subunits that mediate loop 
694: formation are modelled as two identical and connected rigid 
695: bodies, as shown in figure \ref{loopfig1} \cite{loop51}.
696: There are three additional sets of rigid-body 
697: rotation angles that are defined in addition to those for dinucleotide 
698: steps: two sets for the interfaces between protein and the last (DP) and 
699: first bps (PD) of the DNA and one set for the interface between the 
700: two protein domains (PP), where the symbols in parentheses are used to 
701: indicate the corresponding angles through subscripts. The local 
702: Cartesian-coordinate frames for protein subunits are defined such that 
703: their origins coincide with vertices of a circular chain and their 
704: $z$-axes point toward the next vertex in succession.  Thus protein 
705: dimensions can be modelled in terms of a non-canonical value for the helix 
706: rise corresponding to particular segments within a circular polymer chain.
707: 
708: \begin{figure}
709: \begin{center}
710: \includegraphics[width=7.6cm]{bk_loopfig1.ps}
711: \end{center}
712: \caption{Rigid-body models for studies of protein-mediated DNA looping.
713: (a) A prototype 137-bp DNA loop generated by interactions with a pair of 
714: rigid, DNA-binding protein subunits is shown.  DNA bps are 
715: represented by rectangular slabs (red) with axes (blue) that indicate 
716: the orientation of the local Cartesian coordinate frame whose origin 
717: lies at the centre of each bp.  Two sets of coordinate axes 
718: (green) represent the local coordinate frames embedded in the protein 
719: subunits (gold ellipsoids) that mediate DNA looping.  The coupling of 
720: protein and DNA geometry is characterised by tilt, roll, and twist 
721: values for the DNA-protein, protein-protein, and protein-DNA interfaces.  
722: Three of these variables are shown here: the DNA-protein roll angle, 
723: $\phi_{DP}$; the protein-protein twist angle, $\tau_{PP}$; 
724: and the protein-DNA roll angle, $\phi_{PD}$.  
725: (b) Prototype 179-bp loop with protein-protein twist angle, $\tau_{PP}$,
726: equal to $-60$ degrees.  The view is from the base of the loop toward 
727: the DNA apex.  (c) Loop conformation shown in (b) viewed from the side, 
728: perpendicular to the loop dyad axis.
729: \label{loopfig1}}
730: \end{figure}
731: 
732: Angles are expressed in degrees, and length in units of the DNA helical 
733: rise, $\ell_{bp} = 3.4$ {\AA}. All 
734: calculations used canonical mechanical parameters for duplex DNA: helical 
735: twist $\tau_{0} = 34.45^{\circ}$, a sequence-independent twist-angle 
736: standard deviation, or twisting flexibility, $\sigma_{\tau} = 4.388^{\circ}$,
737: and standard deviations, or bending flexibilities, for all tilt and roll 
738: angles,
739: $\sigma_{\theta}$ and $\sigma_{\phi}$, respectively, of $4.678^{\circ}$
740: (equivalent to a persistence length of 150 bp). Except for specific cases 
741: where intrinsic DNA bending is considered, the average values of tilt and 
742: roll are taken to be zero.
743: 
744: 
745: \subsection{Simplified protein geometries and flexibility parameters}
746: \label{loopsection}
747: 
748: For DNA loops with either zero or nonzero end-to-end distances, constraints 
749: are directly applied to the DNA ends, as in the case of DNA cyclisation.  
750: We modelled DNA loops formed during site synapsis using protein-dependent 
751: parameters
752: ${\rm{roll }} = \phi_{DP}  = \phi_{PD} = 90^{\circ}$ and
753: ${\rm{twist}} = \tau_{DP}  = \tau_{PD} = 34.45^{\circ}$.
754: The angle was considered an adjustable parameter that we denote the 
755: {\em axial angle} and, unless specified, all other protein-related angular 
756: parameters were set equal to $0^{\circ}$. In these cases the DNA ends 
757: (the centres of two protein-binding sites on DNA) are separated by twice 
758: the protein-arm length $\ell_p$ and displaced from one another along 
759: the $+x$ direction, or toward the major groove of DNA. Projected along 
760: the $x$-axis, the axial angle is the included angle between the tangents 
761: to the DNA at the two protein binding sites and is altered by varying the 
762: twist between protein subunits (figure \ref{loopfig1} b, c). 
763: An axial angle equal to $0^{\circ}$
764: corresponds to antiparallel axes at the ends as shown in 
765: figure \ref{loopfig1}a.  
766: The case of a rigid protein assembly is modelled by setting the standard 
767: deviations of the DP, PP, and PD sets of rigid-body rotation angles to 
768: $1 \cdot 10^{-8}$ deg.
769: 
770: 
771: \subsection{DNA loops having zero end-to-end distance and antiparallel 
772: helical axes}
773: 
774: DNA loops containing $N$ bps in which the two ends meet in an 
775: antiparallel orientation can be empirically described by the following 
776: formula:
777: %
778: \begin{eqnarray} \label{1}
779: & & {\rm{Tilt:}}  \quad \, \, \, \, 
780: \theta_i = - A_i \, \cos(180 + \delta) \\ \nonumber
781: & & {\rm{Roll:}}  \quad \, \, \, \phi_i = A_i \, 
782: \sin (180 + \delta ) \\ \nonumber
783: & & {\rm{Twist:}} \quad \tau_i  = \tau^0 \nonumber
784: \end{eqnarray}
785: %
786: where $\tau^{0}$ is the intrinsic DNA twist and $\delta$ an arbitrary angle 
787: related to the unconstrained torsional degree of freedom of DNA. The 
788: coefficients $A_i$ are given by
789: %
790: \begin{equation} \label{2}
791: A_i = \frac{1}{N} \, f\left( {\frac{i}{{N - 1}}} \right) \, ,
792: \quad i = 0, \ldots, N - 1
793: \end{equation}
794: %
795: with
796: %
797: \begin{equation} \label{3}
798: f(x) = \left\{ \begin{array}{l}
799: g(x) \, \, , \qquad \quad \, \, \, 0 \le x \le 0.5 \\
800: g(0.5 - x) \, \, , \quad 0.5 < x \le 1 \\
801: \end{array} \right.
802: \end{equation}
803: %
804: where
805: %
806: \begin{equation} \label{4}
807: g(x) = \sum\limits_{i = 1}^5 {a_{i\,} x^i } ,\quad 0 \le x \le 0.5 \, \, .
808: \end{equation}
809: %
810: \begin{figure}
811: \begin{center}
812: \includegraphics[width=7.6cm]{bk_loopfig2.ps}
813: \end{center}
814: \caption{Conformation of an antiparallel, 150-bp DNA loop with zero 
815: end-to-end distance.  (a) Computed space-filling model of the loop 
816: generated with 3DNA \cite{loop49}.  
817: The ends of the DNA juxtapose exactly with 
818: antiparallel helical axes and exact torsional phasing. (b) Equilibrium 
819: roll and magnitude of the loop shown in (a).  The bending magnitude of 
820: each dinucleotide step is defined as $\sqrt{\theta_i^2 + \phi_i^2}$
821: where $\theta_i$ and $\phi_i$ are the tilt and roll of $i$-th 
822: dinucleotide step, respectively.
823: \label{loopfig2}}
824: \end{figure}
825: %
826: The coefficients in equation (\ref{4}) were obtained by fitting the space 
827: curve corresponding to the DNA helical axis that gives the minimum elastic 
828: energy conformation of DNA loops of different sizes and are as follows:
829: $a_0 = -335.0142$, $a_1 = 2318.881$, $a_2 = -1299.164$, 
830: $a_3 = -4483.366$, $a_4 = 38169.74$, $a_5 = -54753.5$. 
831: The error for end-to-end distances computed using 
832: equation (\ref{1}) is less
833: than $2\%$ of DNA length from 50 bp to 100 bp, and less than $0.5\%$ 
834: from 100 bp to 500 bp. The torsional phase angle between two ends is
835: $\xi  =  - \left( {N - 2} \right)\tau  - 2\delta$. The entire loop lies 
836: in a plane, and the angle between the normal vector of the plane and
837: the $x$-axis of the external coordinate can be shown to be 
838: $\psi  = 180 + \tau  - \delta$. The expressions for $\xi$ and $\psi$ 
839: suggest that $\delta$ is related to DNA bending isotropy. Loop 
840: configurations with different $\delta$ values are related to each other 
841: by globally twisting DNA molecules. Since the orientation of the first 
842: bp is fixed, this global twist is equivalent to rotation of the 
843: loop plane, which corresponds to the rotational symmetry met in DNA 
844: cyclisation of homogeneous DNA with bending isotropy \cite{loop26}. 
845: Therefore, $J$-factors for configurations with different $\delta$ values 
846: are identical.
847: 
848: If DNA looping needs to be torsionally in-phase, only two degenerate 
849: loop configurations are available, breaking the rotational symmetry. 
850: These loop geometries can be expressed
851: by equation (\ref{1}) with two different $\delta$ values:
852: $\delta_1 = - (N - 2) \tau / 2$ and
853: $\delta_2  = 180 - (N - 2) \tau / 2$,
854: which satisfy the torsional phase requirement 
855: $\xi  = 360 \cdot n,\;n = 0, \pm 1, \pm 2, \ldots$
856: In contrast to DNA cyclisation, no twist change is involved in forming these 
857: ideal DNA loops for any DNA length and thus the helical dependence vanishes 
858: in this case.  From the expression given above for $\psi$ it is clear that 
859: the helical axes of the two loops are coincident and their directions are 
860: reversed. Figure \ref{loopfig2}
861: shows the bending profile of the loop configuration 
862: corresponding to $\delta_1$ for a 150 bp DNA. Surprisingly, the maximal 
863: $J$-factor occurs at approximately the same DNA length, or 460 bp (data 
864: not shown), as in DNA cyclisation \cite{loop26}.
865: This can be partly explained by 
866: the fact that the total bending magnitude of the loop is $290$ degrees, 
867: close to a full circle, instead of $180$ degrees.
868: 
869: 
870: 
871: \subsection{DNA looping with finite end-to-end distance, antiparallel 
872: helical axes, and in-phase torsional constraint}
873: 
874: \begin{figure}
875: \begin{center}
876: \includegraphics[width=6.8cm]{fig9.eps}
877: \end{center}
878: \caption{The DNA-length-dependent $J$-factor and loop configuration as a 
879: function of end-to-end separation (the $J$-factor is defined
880: in section \ref{loopbio}).
881: (a) The helical dependence of DNA looping 
882: is shown for values of the end-to-end separation equal to 10 bp and 30 bp. 
883: The two configurations for the 10-bp separation are obtained from 
884: corresponding configurations with zero end-to-end separation by using an 
885: iterative algorithm. Therefore the two configurations are designated by the 
886: initial configurations with phase angles 
887: $\delta = -(N-2) \tau / 2 + 0$ ($0^{\circ}$, dashed line) and
888: $\delta = -(N-2) \tau / 2 + 180$ ($180^{\circ}$, solid line)
889: as described in the text. (b) and (c) show
890: stereo models of the two equilibrium 
891: configurations for 210-bp (b) and 215-bp (c) antiparallel DNA loops with 
892: end-to-end separation equal to 10 bp. The 210- and 215-bp DNA correspond 
893: to an adjacent peak and valley of the curve in (a), respectively. 
894: Conformations shown in blue correspond to $\delta = 0$; 
895: those shown in red are for $\delta = 180^{\circ}$. Note that for $N$-bp DNA, 
896: the chain contour length is equal to $(N-1)\ell_{bp}$.
897: \label{loopfig3}}
898: \end{figure}
899: 
900: 
901: Separation of the DNA ends breaks the rotational symmetry, restoring the 
902: dependence on helical twist. Figure \ref{loopfig3}a 
903: shows the $J$-factor as a function 
904: of DNA length for end-to-end distances of $10$ bp and $30$ bp. The helical 
905: dependence increases with end-to-end separation.  Starting from the two 
906: loop configurations (corresponding to $\delta_1$ and $\delta_2$) with zero
907: end-to-end distance and in-phase torsional alignment as initial
908: configurations, two mechanical equilibrium configurations are obtained by 
909: using the iterative algorithm described in \cite{loop26}.
910: The $J$-factor 
911: in figure \ref{loopfig3}a is the sum of separate $J$-factors 
912: calculated for the two 
913: configurations. Note that in all cases involving configurations that differ 
914: in linking number, equilibration between the two forms requires breakage of
915: at least one of the protein-DNA interfaces. The contributions from each of 
916: these configurations are shown in detail for the case where the ends are 
917: separated by $10$ bp. Interestingly, the length dependence of $J$ computed 
918: from the individual configurations are out of phase and have a periodicity 
919: of 2 helical turns, which results from the half-twist dependence of the 
920: phase angles $\delta_1$ and $\delta_2$.  However, their sum displays a 
921: periodicity of one helical turn. Figures \ref{loopfig3} b and c
922: show two such configurations for DNA molecules that are torsionally 
923: in-phase ($N = 210$ bp) or out-of-phase ($N = 215$ bp).
924: 
925: In the case of cyclisation, the helical-phase dependence of the $J$-factor 
926: persists at DNA lengths well beyond that corresponding to the maximum 
927: value of $J$, which lies near $500$ bp. This is clearly not the case for 
928: DNA looping.  In figure \ref{loopfig3}a, 
929: the periodic dependence of $J$ on DNA length 
930: for $10$-bp end-to-end separation decays nearly to zero well before the 
931: maximum $J$ value is reached.  Although the periodicity of $J$ is not 
932: attenuated quite as strongly for $30$-bp end separation, there is less 
933: than four-fold variation in the value of $J$ near $300$ bp, as opposed 
934: to the more than ten-fold variation in cyclisation $J$-factors expected
935: in this length range.  The differences between looping and cyclisation 
936: are largely due to substantial differences in the relative contributions 
937: of DNA writhe in the two processes, as discussed below.
938: 
939: 
940: \subsection{DNA looping in synapsis}
941: \label{loopsyn}
942: 
943: Intramolecular reactions of most site-specific recombination systems 
944: \cite{loop9,loop10,loop11}
945: and a number of DNA restriction endonucleases such as {\em Sfi}I and 
946: {\em Ngo}MIV \cite{loop12},
947: proceed through protein-mediated intermediate structures 
948: in which a pair of DNA sites are brought together in space and the
949: intervening DNA is looped out.  The intermediate nucleoprotein complex 
950: involved in site pairing and strand cleavage (and also exchange, in the 
951: case of recombinases) is termed the {\em synaptic complex}. In these systems, 
952: two characteristic geometric parameters are of interest: the average 
953: through-space distance between the sites and the average crossing angle 
954: between the two ends of the loop, which we denote the axial angle
955: (see section \ref{loopsection}).
956: The latter quantity can be described in terms of the twist angle between 
957: the protein domains, $\tau_{PP}$ (figure \ref{loopfig1}b), 
958: and we use these terms interchangeably.
959: 
960: Figure \ref{loopfig4} shows the helical dependence of looping 
961: (figure \ref{loopfig4}a) and the 
962: elastic-minimum configuration of DNA loops 
963: (figure \ref{loopfig4}b) for different 
964: values of the axial angle. The most prominent feature of these results 
965: is that the phase of the helical dependence is shifted as a function of the
966: axial angle, characterised by a relative global shift of the curve along 
967: the $x$-axis. This implies that DNA looping does not always occur most 
968: efficiently when two sites are separated by an integral number of helical 
969: turns, as has been suggested for some simple DNA looping systems studied 
970: previously. The axial angle also globally modulates $J$-factors, which 
971: is apparent from the vertical shift in the $J$ versus length curve and 
972: effects on the amplitude of the helical dependence. 
973: The torsion-angle-independent value of $J$, averaged over a full helical 
974: turn, decreases with increasing axial angle, whereas the amplitude of 
975: the helical dependence increases.  The above observations can be 
976: qualitatively explained by analogous results from DNA cyclisation.  
977: As in cyclisation, DNA forms loops most efficiently when the number of 
978: helical turns in the loop is close to an integer value. It is therefore 
979: appropriate to consider this issue in terms of the linking number for the
980: looped conformation, $Lk$, which involves contributions from the geometries 
981: of both the protein and DNA.
982: 
983: \begin{figure}
984: \begin{center}
985: \includegraphics[width=7.6cm]{bk_loopfig4.ps}
986: \end{center}
987: \caption{Dependence of the $J$-factor on axial angle 
988: (the $J$-factor is defined in section \ref{loopbio},
989: and the axial angle is defined as 
990: the average crossing angle between the two 
991: ends of the loop, see section \ref{loopsection}).
992: (a) DNA-length dependence of $J$ for axial angles of 
993: $0^{\circ}$, $60^{\circ}$, and $120^{\circ}$ 
994: with the end-to-end separation set equal to 40 bp.
995: Note that the positions 
996: of the extrema shift to the left with increasing values of the axial angle.  
997: (b) Stereo models of minimum elastic-energy conformations of 179-bp loops 
998: colour coded in accord with the corresponding axial-angle values in (a).
999: \label{loopfig4}}
1000: \end{figure}
1001: 
1002: We define the {\em loop helical turn} $H_{t,loop}$ as the sum of the 
1003: DNA twist and the twist introduced by the protein subunits, divided by 
1004: $360$.  Therefore, changing the twist angle, the axial angle will 
1005: shift the phase of the helical dependence relative to that of
1006: the DNA alone.  For a loop with $N = 179$ bp and $\tau_{PP} = 0$, the 
1007: total twist is simply equal to that for the DNA loop.  Because this loop 
1008: has $17.0$ helical turns, only one loop topoisomer contributes to the 
1009: $J$-factor.  The value of $J$ is a local maximum at $\tau_{PP} = 0$ and, 
1010: as shown in figure \ref{loopfig5}a,
1011: decreases monotonically for both $\tau_{PP} > 0$
1012: and $\tau_{PP} < 0$.  Contributions to $J$ from other topoisomers of 
1013: the $179$-bp loop are less than $5$ percent over the range 
1014: $-135^{\circ} < \tau_{PP} < +120^{\circ}$. The twist for the planar 
1015: equilibrium conformation of a $173$-bp loop is $16.5$ helical turns;
1016: thus there are two alternative loops that can be efficiently formed 
1017: (figure \ref{loopfig5}a): 
1018: either a loop with $H_{t,loop} = 17.0$ and $\tau_{PP} > 0$, 
1019: or a loop with $H_{t,loop} = 16.0$ and $\tau_{PP} < 0$. The $J$ value 
1020: at $\tau_{PP} = 0$ is a local minimum and there is a bimodal dependence 
1021: on axial angle for loops in which the DNA twist is half-integral. 
1022: We investigated the phase shift of the $J$-factor and found that 
1023: this quantity is a non-linear function of the axial angle.  From 
1024: figure \ref{loopfig4}a, 
1025: the calculated phase shifts for $60^{\circ}$ and $120^{\circ}$ 
1026: axial angles relative to $0^{\circ}$ are approximately $52^{\circ}$ 
1027: and $103^{\circ}$, respectively.  Moreover, the local maxima for the total
1028: $J$ curve for $N = 173$ shown in figure \ref{loopfig5}a
1029: are located at $-58.5^{\circ}$ 
1030: and $63^{\circ}$, positions that are not in agreement with predicted 
1031: angle values based solely on $H_{t,loop}$ ($-166^{\circ}$ and 
1032: $194^{\circ}$, respectively).
1033: 
1034: \begin{figure}
1035: \begin{center}
1036: \includegraphics[width=7.6cm]{bk_loopfig5.ps}
1037: \end{center}
1038: \caption{$J$-factor, loop-geometry parameters, and elastic-free energies as 
1039: functions of axial angle; compare figure \ref{loopfig4}.
1040: (a) $J$-factor values for loop topoisomers 
1041: corresponding to 179-bp and 173-bp loops in figure \ref{loopfig4}.
1042: The principal contribution to $J$ for $N = 179$ bp comes from a single 
1043: loop topoisomer with $Lk = 17$.  For $N = 173$ bp, the overall $J$-factor 
1044: is the sum of contributions from two loop topoisomers with $Lk$ values of 
1045: 16 and 17, generating a bimodal dependence of $J$ on axial angle as 
1046: described in the text. (b) Excess helical twist, $\Delta H_t$, 
1047: and writhe of the loop formed by the $Lk = 16$ topoisomer for $N = 173$ bp
1048:  as a function of axial angle.  Excess twist is computed from the expression 
1049: $H_{t, loop} - 16$, where $H_{t, loop}$ is the loop helical turn value 
1050: described in the text, and depends linearly on the axial angle.  
1051: The writhing number of the loop was calculated using the method of 
1052: Vologodskii \cite{loop28,loop50}. (c) Elastic-free energies of the 
1053: $Lk = 16$ loop topoisomer for $N = 173$ bp calculated according to 
1054: equation 38 of Zhang and Crothers \cite{loop26}. The individual 
1055: contributions of bending and twisting energies are shown along 
1056: with their sum.
1057: \label{loopfig5}}
1058: \end{figure}
1059: 
1060: These deviations can be explained by the fact that writhe makes an 
1061: important contribution to the overall $Lk$ for the loop.  This aspect 
1062: of DNA looping is dramatically different from that in the cyclisation 
1063: of small DNA molecules.  The conformations of small DNA circles are
1064: close to planar and the writhe contribution is small relative to DNA 
1065: twist \cite{loop26,loop30,loop44,loop45}.
1066: In the case of protein-mediated looping, nonzero 
1067: values of the axial angle impose an intrinsically nonplanar conformation 
1068: on the DNA.  The relative contributions of loop writhe and twist for 
1069: the $Lk = 16$ topoisomer of a $173$-bp loop are shown as a function of 
1070: axial angle in figure \ref{loopfig5}b.
1071: 
1072: In figure \ref{loopfig5}c, 
1073: we plot the axial-angle-dependent values of the bending and 
1074: twisting free energies for the $Lk = 16$ topoisomer and their sum, which 
1075: is the total elastic-free energy of the loop. The minimum value of the 
1076: total elastic energy occurs at $\tau_{PP} = -58.5^{\circ}$, coincident 
1077: with the position of the $J$-factor maximum for this topoisomer 
1078: (figure \ref{loopfig5}a).
1079: This mechanical state can be achieved with very little twist deformation 
1080: of the loop, but at the expense of significant bending energy.  Further 
1081: reduction of the axial angle requires even less twisting energy; however, 
1082: the bending energy increases monotonically. In contrast, for 
1083: $\tau_{PP} > -58.5^{\circ}$, somewhat less bending energy is required,
1084: but the twisting energy begins to increase significantly with increasing 
1085: axial angle. Since the sense of the bending deformation for 
1086: $\tau_{PP} > 0$ opposes the needed reduction in loop linking number, 
1087: the elastic energy cannot be decreased by increasing the axial angle.
1088: The only way that the loop geometry can compensate for this is through 
1089: twist deformation. This asymmetry arises because we are considering the 
1090: contribution of only one loop topoisomer to the elastic free energy.
1091: 
1092: \subsection{Conclusion}
1093: 
1094: The statistical-mechanical theory for DNA looping discussed above
1095: \cite{loop26,loop51} 
1096: suggests that the helical dependence of DNA looping is 
1097: affected by many factors and leads to the conclusion that whereas a 
1098: positive helical-twist assay can often confirm DNA looping, a negative 
1099: result cannot exclude DNA looping.  Since it is difficult to explore 
1100: the architecture of DNA loops with current experimental techniques, 
1101: this theory will be useful for more reliably analysing DNA looping
1102: with limited experimental data.  The model has advantages over 
1103: previous approaches based exclusively on DNA mechanics, particularly 
1104: when protein flexibility is taken into account. In these cases, entropy 
1105: effects become important and are responsible for the observed decay of 
1106: looping efficiency with DNA length.  
1107: 
1108: 
1109: \section{DNA knots and their consequences: entropy and targeted knot
1110: removal}
1111: 
1112: 
1113: Bacterial DNA occurs largely in circular form. Notably, instead of a simply
1114: connected ring shape (the unknot), the DNA often exhibits permanently
1115: entangled states, such as catenated and knotted DNA. An example for a DNA
1116: trefoil knot is shown in figure \ref{dna_trefoil}. Such configurations have
1117: potentially devastating effects on the cell development. Conversely, however,
1118: knots might have designed purposes in gene regulation, separating different
1119: regions of the genome, or, alternatively, locking chemically
1120: remote parts of the genome proximate in geometrical space. In eukaryotic
1121: cells additional topological effects occur in the likely entanglement of
1122: individual chromosomes. Here, we concentrate on the prokaryotic case.
1123: 
1124: \begin{figure}
1125: \includegraphics[height=5.4cm]{dna_trefoil.ps}
1126: \caption{Electron microscope image of a DNA trefoil knot, from
1127: \cite{wassermann2}. $\copyright$ Science, with permission.
1128: \label{dna_trefoil}}
1129: \end{figure}
1130: 
1131: 
1132: \subsection{Physiological background of knots}
1133: 
1134: The discovery how one can use molecular biological tools to create knotted
1135: DNA resolved a long-standing argument against the
1136: Watson-Crick double helix picture of DNA \cite{frank}, namely that the
1137: replication of DNA could not work as the opening up of the double helix
1138: would produce a superstructure such that the two daughter strands could
1139: not be separated. In fact, the topology of both ssDNA and dsDNA is continuously
1140: changed in vivo, and this can readily be mimicked in vitro, although the
1141: activity of enzymes in vivo is much more restricted than in vitro
1142: \cite{deibler,dna_topo}: Different concentrations of enzymes versus knotted
1143: DNA molecules accessible in vitro,
1144: that is, makes it possible to probe topology-altering effects by enzymes
1145: which in vivo do not contribute to such effects.
1146: 
1147: Although it would be likely
1148: with a probability of roughly $\frac{1}{2}$ that the linear DNA injected by
1149: bacteriophage $\lambda$ into its host \ecoli would create a knot before
1150: cyclisation, it turned out to be difficult to detect \cite{frank}. First studies
1151: therefore concentrated on the fact that under physiological conditions knots
1152: are introduced by enzymes, DNA replication and recombination, DNA repair, and
1153: topoisomerisation, using these enzymes to prove both knotting and unknotting
1154: \cite{liu,mizuuchi,pollock,spengler,wassermann,wassermann1,wassermann2}.
1155: DNA-knotting is also prone to occur behind a stalled replication fork
1156: \cite{viguera,sogo}. Some of the typical topology-altering reactions
1157: undergoing in \ecoli are summarised in figure \ref{topogyro}.
1158: Knots can efficiently be created from nicked\footnote{One of the two strands
1159: is cut.} dsDNA under action of
1160: topoisomerase I at non-physiological concentrations \cite{topoknot}.
1161: Another possibility is by active packaging of a DNA mutant into phage
1162: capsids \cite{arsuaga}, and then denaturing the capsid proteins. Both
1163: methods produce a distribution of different knot types. They can be
1164: separated by electrophoresis \cite{electro}.
1165: 
1166: \begin{figure}
1167: \includegraphics[width=8.6cm]{topo_gyro.eps}
1168: \caption{Enzymes changing the topology of dsDNA by cutting and pasting of one
1169: or both strands (example for \ecoli): (A) Torsional stress resulting from
1170: the Lk deficit causes the DNA double helix to writhe about itself (negative
1171: supercoiling). In \ecoli, gyrase introduces negative supercoils into DNA
1172: and is countered by topoisomerase I (topo I) and topo IV, which relax negative
1173: supercoils. (B) Topo IV unlinks catenanes generated by replication or
1174: recombination in vivo. (C) Topo IV unknots DNA in vivo. After \cite{deibler}.
1175: \label{topogyro}}
1176: \end{figure}
1177: 
1178: The existence of DNA-knots has far-reaching effects on physiological processes,
1179: and knottedness of DNA has therefore to be eliminated in order to
1180: maintain proper functioning of the cell. Among other possible effects, it
1181: is immediately clear that the presence of a knot in a circular DNA impedes
1182: replication of the DNA, i.e., the full separation of the two daughter
1183: strands \cite{alberts,frank}. Moreover, even transcription is impaired
1184: \cite{portugal}. The presence of knots inhibits the assembly of chromatin
1185: \cite{rodriguez}, knotted chromosomes cannot be separated during mitosis
1186: \cite{alberts}, and knots in a chromosome may serve as topological barriers
1187: between different sections of chromosomes, such that the genomic structural
1188: organisation is altered, and certain sections of the chromosomal DNA may no
1189: longer interact \cite{staczek}. Conversely, it is conceivable that knots,
1190: analogously to protein induced DNA looping, lock remote segments of the genome
1191: close together in geometric space. Finally, knots may lead to double-strand
1192: breaks, as they weaken biopolymers considerably due to creation
1193: of localised sharp bends \cite{arai,pieranski,saitta,stasiak1}
1194: as well as macroscopic lines and ropes \cite{mcnally}.\footnote{The weakness
1195: of strings at the site of the knot can be experienced easily by pulling apart
1196: a linear nylon string in comparison to a knotted one \protect\cite{pieranski}.}
1197: 
1198: Above we said that knots can be introduced, inter alia, by the different
1199: enzymes of the topoisomerase family. To remove a knot
1200: from a dsDNA, it is necessary to cut both strands, and then pass one
1201: segment through the created gap, before resealing the two open ends. In
1202: vivo, this is usually achieved by topoisomerases II and IV. A reconstruction
1203: of topo II is shown in figure \ref{topoII}, indicating the upper clamp
1204: holding a segment of the DNA, while the bulge-clamp introduces the cut
1205: through which the upper segment is passed. In the figure, the segment
1206: visible in the pocket of the lower clamp has already been passed through
1207: the gap. After resealing, topo II detaches. This process requires energy,
1208: provided by ATP. Notably, topo II is extremely efficient, for circular dsDNA
1209: of length $\simeq 10$ kbp it was found that topo reduced the knotted state in
1210: between 50 and 100-fold, in comparison to a `dumb' enzyme, which would simply
1211: pass segments through at
1212: random \cite{rybenkov}. We note that the step-wise action of topoisomerase
1213: II was recorded in a single molecule setup using magnetic tweezers
1214: \cite{stasiak,strick}. Topoisomerases are surveyed in the review of
1215: \cite{wang}.
1216: 
1217: \begin{figure}
1218: \includegraphics[width=6.8cm]{topo2_bollen.ps}
1219: \caption{Topoisomerase II. This enzyme can
1220: actively change the topology of DNA by cutting the double-strand and
1221: passing another segment of double-stranded DNA through the gap before
1222: resealing it. The image depicts a short stretch of DNA (horizontally
1223: at the bulge of the enzyme, as well as another segment in the lower
1224: clamp (perpendicular to the image) after passage through the gap from
1225: the upper clamp. This mechanism makes sure that no additional strand
1226: passage through the open gap can take place \cite{berger,wang_to}.
1227: Figure courtesy James M Berger, UC Berkeley.
1228: \label{topoII}}
1229: \end{figure}
1230: 
1231: 
1232: \subsection{Classification of knots}
1233: 
1234: Knottedness can only be defined on a closed (circular) chain. This is
1235: intuitively clear as in an open linear chain a knot can always be tied, or
1236: an existing knot released. Mathematically, this means that knot invariants
1237: are only well-defined for a closed space-curve. However, a linear chain
1238: whose ends are permanently attached to one, or two walls, or whose ends are
1239: extended towards infinity, can be considered as (un)knotted in the proper
1240: mathematical sense, i.e., their knottedness cannot change. In a loser
1241: sense, we will also speak of knots on an open piece of DNA, appealing to
1242: intuition.
1243: 
1244: The classification of knots, or graphs in general, in terms of invariants
1245: can essentially be traced back to Euler, recalling his graph theoretical
1246: elaboration in connection with the Bridges of K{\"o}nigsberg problem
1247: \cite{euler}, determining a closed path by crossing each K{\"o}nigsberg bridge
1248: exactly once. However, the first investigations of topological problems in
1249: modern science is most probably due to Kepler, who studied surface tiling to
1250: great detail (therefore the notion of Kepler tiling in mathematical literature)
1251: \cite{kepler}. Further initial steps
1252: were due to Leibniz, Vandermonde and Gauss, in whose collection of papers
1253: drawings of various knots were found\footnote{Probably copies from an English
1254: original.} whose linking (`Umschlingungen'=windings) number is indeed a knot
1255: invariant \cite{adams,kauffman,reidemeister}. Gauss' student, Listing, in fact
1256: introduced the term `topology', and his work on knots may be viewed as the
1257: real starting point of knot theory \cite{listing}, although his complexions
1258: number was proved by Tait not to be an invariant.
1259: 
1260: Inspired by Helmholtz' theory of an ideal fluid and building on Listing's
1261: early contributions to knot theory, Scotsmen and chums Maxwell, Tait and
1262: Thomson (Lord Kelvin) started to discuss the possible implications of
1263: knottedness in physics and chemistry, ultimately distilled into Thomson's
1264: theory of vortex atoms \cite{thomson,thomson1}. Out of this endeavour emerged
1265: Tait's interest in knots, and he devoted most of his career on the
1266: classification of knots. Numerous charts and still unresolved conjectures
1267: on knots document his pioneering work \cite{tait,tait1,tait2,tait3}. The
1268: studies were carried on by Kirkman and Little
1269: \cite{kirkman,kirkman1,little,little1}.
1270: A more detailed historical account of knot theory may be found in the review
1271: article by van de Griend \cite{degriend}, and on the St. Andrews history of
1272: mathematics webpages\footnote{The MacTutor History of Mathematics
1273: archive, URL: {\tt http://turnbull.mcs.st-and.ac.uk/~history/}}.
1274: 
1275: Planar projections of knots were rendered unique by Listing's introduction
1276: of the handedness of a crossing, i.e., the orientational information assigned
1277: to a point where in the projection two lines intersect. With this information,
1278: projections are the standard representation for knot studies. On their basis,
1279: the minimum
1280: number of crossings (`essential crossings') can be immediately read off as
1281: one of the simplest knot invariants. To arrive at the minimum number, one
1282: makes use of the Reidemeister moves, three fundamental permitted moves of the
1283: \begin{figure}
1284: \includegraphics[width=8cm]{reidemeist.eps}
1285: \caption{The three Reidemeister moves. All topology-preserving moves of a
1286: knot projection can be decomposed into these three fundamental moves.
1287: \label{reidemeister}}
1288: \end{figure}
1289: lines in a knot projection, as shown in figure \ref{reidemeister}. More complex
1290: knot invariants include polynomials of the Alexander, Kauffman and HOMFLY types
1291: \cite{adams,kauffman,reidemeister}.\footnote{These polynomials all start to
1292: be degenerate for higher order knots, i.e., above a certain knot complexity
1293: several knots may correspond to one given polynomial \cite{adams,kauffman}.
1294: In the case of the simpler knots attained in most DNA configurations and in
1295: knot simulations, the Alexander polynomials are unique, in contrast to the
1296: Gauss or Edwards invariant, compare, e.g., reference \protect\cite{volo}.}
1297: Here, we will only employ the number of essential crossings as
1298: classification of knots, in particular, we do not concern ourselves with the
1299: question of degeneracy for a given knot invariant. However, the bookkeeping
1300: of knot types is vital in knot simulations.
1301: 
1302: 
1303: \subsection{Long chains are almost always entangled.}
1304: 
1305: During the polymerisation and final cyclisation of a polymer grown
1306: in a solvent under freely floating conditions, a knot is created
1307: with probability 1. This Frisch-Wassermann-Delbr{\"u}ck conjecture
1308: \cite{frisch,delbrueck} could be mathematically proved for a self-avoiding
1309: chain \cite{sumners,pippenger}, compare also \cite{vanderzande}. This is
1310: consistent with numerical findings
1311: that the probability of unknot formation decreases dramatically with
1312: chain length \cite{frank3,volo}. Indeed, recent simulations results indicate
1313: that the probability of finding the unknot in such a cyclisised polymer
1314: decays exponentially with chain length \cite{koniaris,michels,michels1,janse}:
1315: \begin{equation}
1316: P_\emptyset(N)\propto \exp\left(-\frac{N}{N_c}\right).
1317: \end{equation}
1318: However, there exist theoretical arguments and simulations results indicating
1319: that the characteristic number of monomers $N_c$ occurring in this relation
1320: may become
1321: surprisingly large \cite{frank2,grosberg4,klenin,shimamura1}. The probability
1322: to find a given knot type ${\cal K}$ on random circular polymer formation has
1323: been fitted with the functional form \cite{katritch,dobay,shimamura1}
1324: \begin{equation}
1325: P_{\cal K}(N)=a\big(N-N_0\big)^b\exp\left(-\frac{N^c}{d}\right),
1326: \end{equation}
1327: where $a$, $b$, and $d$ are free parameters depending on ${\cal K}$, and $c
1328: \approx 0.18$. $N_0$ is the minimal number of segments required to form a
1329: knot ${\cal K}$, without the closing segment \cite{dobay}. The tendency
1330: towards knotting during polymer cyclisation creates problems in industrial
1331: and laboratory processes.
1332: 
1333: 
1334: \subsection{Entropic localisation in the figure-eight slip-link structure.}
1335: 
1336: To obtain a feeling for how and when entropy leads to the localisation of
1337: a permanently entangled structure, we consider the simplest polymer object
1338: with non-trivial (non-unknot) geometry, the 
1339: figure-eight structure (F8) displayed in figure \ref{fig8}. In this compound,
1340: a pair contact is enforced by a slip-link, separating off two loops in the
1341: circular polymer, such that none of the loops can fully retract, and both
1342: loops can freely exchange length among each other. We denote the loop sizes
1343: by $n$ and $N-n$, where $N$ is the (conserved) total length of the polymer
1344: chain. For such an object, we can actually perform a closed statistical
1345: mechanical analysis based on results from scaling theory of polymers,
1346: and compare the result with Monte Carlo simulations of the F8.
1347: \begin{figure}
1348: \includegraphics[height=4cm]{figu-8.eps}
1349: \caption{Figure-eight structure, in which a slip-link separates two loops
1350: of size $n$ and $N-n$, such that they can freely exchange length among
1351: each other, but none of the loops can
1352: completely retract from the slip-link. On the right, a schematic drawing
1353: of a slip-link, which may be thought of as a small belt buckle.
1354: \label{fig8}}
1355: \end{figure}
1356: 
1357: The statistical quantities that are of particular interest are the gyration
1358: radius, $R_g$, and the number of degrees of freedom, $\omega$
1359: \cite{degennes}. $R_g$, as defined in equation (\ref{appgyr}),
1360: measures the root mean squared distance of the monomers along the chain to the
1361: gyration centre,
1362: and is therefore a good measure of its extension. It can, for instance, be
1363: measured by light scattering experiment. The degrees of freedom $\omega$ count all possible
1364: different configurations of the chain. For a circular polymer (i.e., a polymer
1365: with $\mathbf{r} (0)=\mathbf{r}(N)$), the gyration radius becomes
1366: \begin{equation}
1367: \label{gyrra}
1368: R_g\simeq AN^{\nu}
1369: \end{equation}
1370: with exponent $\nu=1/2$ for a Gaussian chain, and $\nu=0.588$ in the 3D
1371: excluded volume case ($\nu=3/5$ in the Flory model, and $\nu=3/4$
1372: in 2D). Whereas in 2D this scaling contains truly a ring polymer, in 3D the
1373: exponent $\nu$ emerges from averaging over all possible topologies, and
1374: necessarily includes knots of all types \cite{duplantier1,deutsch,degennes}.
1375: For a circular chain, the number of degrees of freedom contains the number of all possible
1376: ways to place an $N$-step walk on the lattice with connectivity $\mu$
1377: (e.g., $\mu=2d$ on a cubic lattice in $d$ dimensions),
1378: $\mu^N$, and the entropy loss for requiring a closed loop, $N^{-d\nu}$,
1379: involving the same Flory exponent $\nu$. For the Gaussian case, we recognise
1380: in this entropy loss factor the returning probability of the random walk. In
1381: the excluded volume case, $N^{-d\nu}$ is an analogous measure
1382: \cite{degennes,grosberg,hughes}. Thus, for a circular chain embedded in
1383: $d$-dimensional space, the number of degrees of freedom is
1384: \begin{equation}
1385: \label{dof}
1386: \omega\simeq\mu^NN^{-d\nu}.
1387: \end{equation}
1388: Let us evaluate these measures for the F8 from figure \ref{fig8}.
1389: 
1390: As a first approximation, consider the F8 as a Gaussian (phantom) random walk,
1391: demonstrating that, like in the charged knot case \cite{paul}, entropic effects
1392: give rise to long-range interactions. The two loops correspond to returning
1393: random walks, i.e., the number of degrees of freedom for the F8 in the phantom chain case
1394: becomes \cite{degennes,flory1,grosberg}\footnote{Here and in the following we
1395: consider two configurations of a polymer chain different if they cannot be
1396: matched by translation. In addition, the origin of a given structure is fixed
1397: by a vertex point (see below), i.e., a point where several legs of the polymer
1398: chain are joint. In the F8-structure, this vertex naturally coincides with the
1399: slip-link. For a simply connected ring polymer, such a vertex is a two-vertex
1400: anywhere along the chain. \label{counting}}
1401: \begin{equation}
1402: \label{fig8id}
1403: \omega_{\rm F8,PC}\simeq\mu^Nn^{-d/2}(N-n)^{-d/2},
1404: \end{equation}
1405: where $d$ is the embedding dimension. We note that normalisation of this
1406: expression produces the probability density function for finding
1407: the F8 with a given loop size $\ell=na$ ($L=Na$),
1408: \begin{equation}
1409: \label{f8_pdf}
1410: p_{\rm F8,PC}(\ell)\simeq\mathscr{N}\ell^{-d/2}(L-\ell)^{-d/2},
1411: \end{equation}
1412: where $\mathscr{N}$ denotes a normalisation factor.
1413: The conversion from expressing the chain size in
1414: terms of the number of monomers to its actual length is of advantage in what
1415: follows, as it allows to more easily keep track of dimensions. Here, we use
1416: the length unit $a$, which may be interpreted as the monomer size (lattice
1417: constant), or as the size of a Kuhn statistical segment.
1418: 
1419: \begin{figure}
1420: \includegraphics[width=6.8cm]{sl2D_equilibrium.eps}
1421: \caption{Bead-and-tether chain used in Monte Carlo simulation, showing a
1422: typical equilibrium configuration for a self-avoiding chain: the localisation
1423: of the smaller loop is distinct. Note that in this 2D simulation the slip-link
1424: is represented by the three tethered black beads.
1425: \label{sl}}
1426: \end{figure}
1427: 
1428: To classify different grades of localisation, we follow the convention from
1429: references \cite{slili2d,paraknot}.
1430: The average loop size $\langle \ell\rangle$ determined through
1431: $\langle \ell \rangle = \int_{a}^{L-a} \ell p(\ell) d\ell$
1432: is trivially $\langle\ell\rangle=L/2$ by symmetry of
1433: the structure. Here, we introduce a short-distance cutoff set by the lattice
1434: constant $a$. However, as the probability density function
1435: is strongly peaked at $\ell=0$ and $\ell=L$,
1436: the two poles caused by the returning probabilities, and therefore a {\em
1437: typical\/} shape consists of one small ({\em tight\/}) and one large ({\em
1438: loose\/}) loop, compare figure \ref{sl}. This can be quantified in terms of
1439: the average size of the smaller loop,
1440: \begin{equation}
1441: \label{sls_f8}
1442: \langle \ell \rangle_<\equiv 2\int_a^{L/2} \ell p(\ell) d\ell \, .
1443: \end{equation}
1444: In $d=2$, we obtain
1445: \begin{equation}
1446: \langle \ell \rangle_<\sim \frac{L}{|\log(a/L)|},
1447: \end{equation}
1448: such that with the logarithmic correction the smaller loop is only marginally
1449: smaller than the big one. In contrast, one observes {\em weak localisation}
1450: \begin{equation}
1451: \label{fig8d3}
1452: \langle \ell \rangle_<\sim a^{1/2}L^{1/2}
1453: \end{equation}
1454: in $d=3$, in the sense that the relative size $\langle \ell \rangle_</L$
1455: tends to zero for large chains. By comparison, for $d>4$ one encounters
1456: $\langle \ell \rangle_<\sim a$, corresponding to {\em strong localisation},
1457: as the size of the smaller loop does not depend on $L$ but is set by the
1458: short-distance cutoff $a$. Above four dimensions, excluded volume effects
1459: become negligible, and therefore both Gaussian and self-avoiding chains are
1460: strongly localised in $d\ge 4$.\footnote{Consideration of higher than the
1461: physical 3 space dimensions is often useful in polymer physics.}
1462: 
1463: To include self-avoiding interactions, we make use of results for general
1464: polymer networks obtained by Duplantier \cite{duplantier,duplantier1}, which
1465: are summarised in the appendix at the end of this review. In terms of such
1466: networks, our F8-structure corresponds to the following parameters: the
1467: number ${\cal N}=2$ of polymer segments with lengths $s_1=\ell=na$ and $s_2=
1468: L-\ell=(N-n)a$, forming ${\cal L}=2$ physical loops, connected by $n_4=1$
1469: vertex of order four.
1470: By virtue of equation (\ref{network}), the number of configurations
1471: of the F8 with fixed $\ell$ follows the scaling form
1472: \begin{equation}
1473: \label{t_scaling}
1474: \omega_{\rm F8}(\ell)\simeq \mu^L (L-\ell)^{\gamma_{\rm F8}-1}
1475: {\cal X}_{\rm F8}\left(\frac{\ell}{L-\ell}\right),
1476: \end{equation}
1477: with the configuration exponent $\gamma_{\rm F8}=1-2d\nu+\sigma_4$. In the
1478: limit $\ell\ll L$, the contribution of the large loop
1479: in equation (\ref{t_scaling})
1480: should not be affected by a small appendix, and therefore should exhibit the
1481: regular Flory scaling $\sim (L-\ell)^{-d\nu}$ \cite{kafri,kafri1,duplantier3}.
1482: This fixes the scaling behaviour of the scaling function ${\cal X}_{\rm F8}
1483: (x)\sim x^{\gamma_{\rm F8}-1+d\nu}$ in this limit ($x \to 0$ in dimensionless
1484: variable $x$), such that
1485: \begin{equation}
1486: \label{sl_limit}
1487: \omega_{\rm F8}(\ell)\simeq\mu^L(L-\ell)^{-d\nu}\ell^{-c},
1488: \quad \ell\ll L,
1489: \end{equation}
1490: where $c=-(\gamma_{\rm F8}-1+d\nu)=d\nu-\sigma_4$. Using $\sigma_4=-19/16$
1491: and $\nu = 3/4$ in $d=2$ \cite{duplantier,duplantier1}, we obtain
1492: \begin{equation}
1493: \label{expon}
1494: c=43/16=2.6875, \quad d=2.
1495: \end{equation}
1496: In $d = 3$, $\sigma_4\approx -0.48$ \cite{schaefer,kafri,kafri1} and
1497: $\nu\approx 0.588$, so that
1498: \begin{equation}
1499: c\approx 2.24.
1500: \end{equation}
1501: In both cases the result $c>2$ enforces that the loop of length
1502: $\ell$ is strongly localised in the sense defined above.
1503: This result is self-consistent with the a priori assumption $\ell\ll L$.
1504: Note that for self-avoiding chains, in $d=2$ the
1505: localisation is even {\em stronger\/} than in $d=3$, in contrast
1506: to the corresponding trend for ideal chains.
1507: 
1508: We performed Monte Carlo (MC) simulations of the 2D figure-eight structure,
1509: in which the slip-link was represented by three tethered
1510: beads enforcing a sliding pair contact such that the loops cannot fully
1511: retract (see figure \ref{fig8_sim}). We used a 2D hard core bead-and-tether
1512: chain with 512 monomers, starting off from a symmetric initial condition
1513: with $\ell=L/2$.
1514: Self-crossings were prevented by keeping a maximum bead-to-bead
1515: distance of 1.38 times the bead diameter, and a maximum step length of 0.15
1516: times the bead diameter.
1517: As shown in figure \ref{sl1}, the size distribution
1518: for the small loop can be fitted to a power law with exponent $c=2.68$
1519: in good agreement with equation (\ref{expon}).
1520: 
1521: \begin{figure}
1522: \includegraphics[width=6.8cm]{loops.eps}
1523: \caption{Monte Carlo simulation of an F8-structure in 2D: loop
1524: sizes $\ell$ and $L-\ell$ as a function of Monte Carlo steps for a chain
1525: with 512 monomers. The symmetry breaking after the symmetric initial
1526: condition is distinct.
1527: \label{fig8_sim}}
1528: \end{figure}
1529: 
1530: \begin{figure}
1531: \includegraphics[width=6.8cm]{pdf_slN512.eps}
1532: \caption{Power-law fit to the probability density function of the smaller
1533: loop. The fit
1534: produces a slope of -2.68, in excellent agreement with the calculated value.
1535: \label{sl1}}
1536: \end{figure}
1537: 
1538: An experimental study of entropic tightening of a macroscopic F8-structure
1539: was reported in reference \cite{bennaim}. There, a granular chain consisting
1540: of hollow steel spheres connected by steel rods was once twisted and then
1541: put on a vibrating table. From digital imaging, the distribution of loop
1542: sizes could be determined and compared to a power-law with index 43/16
1543: as calculated for the 2D excluded volume chain. The agreement was found to
1544: be consistent \cite{bennaim}.
1545: 
1546: 
1547: \subsection{Simulations of entropic knots in 2D and 3D.}
1548: 
1549: Much of our knowledge about the interaction of knots with thermal fluctuations
1550: is based on simulations of knotted chains. Before going further into the
1551: theoretical modelling of knotted chains, we report some of the results based
1552: on simulations studies of both Gaussian and self-avoiding
1553: walks.\footnote{Although per se a Gaussian chain cannot have a fixed topology
1554: due to its phantom character, such simulations introduce a fixed topology
1555: by rejecting moves that result in a different knot type.}
1556: Such simulations
1557: either start with a given knot configuration and then perform moves of
1558: specific segments, each time making sure that the topology is preserved;
1559: or, each new configuration emanates from a new
1560: random walk, whose correct topology may be checked by calculating the
1561: corresponding knot invariant, usually the Alexander polynomial, and
1562: created configurations that do not match the desired topology are discarded.
1563: We note that it is of lesser significance that knot invariants such as the
1564: Alexander polynomials in fact are no longer unique for more complex knots,
1565: because for typical chain lengths with the highest probability simpler knots
1566: are created, for which
1567: the invariants are unique. For more details we refer to the works
1568: quoted below.
1569: 
1570: In fact, the fixed topology turns out to have a highly non-trivial effect
1571: on chains without self-excluded volume. As conjectured in \cite{descloizeaux},
1572: a Gaussian circular chain, whose permitted set of configurations is restricted
1573: to a fixed topology, will exhibit self-avoiding behaviour. This was proved in a
1574: numerical analysis in \cite{deutsch}. The required number of monomers to
1575: reach this self-avoiding exponent was estimated to be of the order of 500.
1576: Keeping this non-trivial scaling of a Gaussian chain at fixed topology in mind,
1577: knot simulations on the basis of phantom Gaussian chains were performed in
1578: \cite{dobay1}, always making sure that the configurations taken into the
1579: statistics fulfil the desired knot topology.
1580: 
1581: The dependence of the gyration radius $R_g$ on the knot type was investigated
1582: for simpler knots in 3D in reference \cite{janse}. On the basis of the
1583: expansion
1584: \begin{equation}
1585: R_g^2\simeq A_{\cal K}\Big(1+B_{\cal K}N^{-\Delta}+C_{\cal K}N^{-1}+o(1/N)\Big)
1586: N^{2\nu_{\cal K}},
1587: \end{equation}
1588: including a confluent correction \cite{janse,leguillou,zinn} in comparison to
1589: the standard expression (\ref{gyrra}), it was found that the Flory exponent
1590: $\nu_{\cal K}$ is independent of the knot type ${\cal K}$ and has the
1591: 3D value $0.588$. This was interpreted via a {\em localisation\/} of knots
1592: such that the influence of tight knots on $R_g$ is vanishingly small. In fact,
1593: $\Delta$ is
1594: of the order of $0.5$ according to the investigations in references
1595: \cite{leguillou,zinn,janse1,li}.
1596: Based on longer chains in comparison to reference
1597: \cite{janse}, the study of \cite{orlandini} thus corroborates the
1598: independence of $\nu_{\cal K}\approx 0.588$ of the knot type ${\cal K}$.
1599: In recent AFM experiments analysing single DNA knots, the Flory scaling
1600: $R_g\simeq N^{\nu}$ was confirmed for both simple and complex knots
1601: \cite{valle}.
1602: 
1603: For the number of degrees of freedom $\omega_{\cal K}$, it was found for the
1604: form\footnote{Note that
1605: we changed the exponent by 1 in comparison to the original work, making the
1606: counting of non-translatable configurations consistent with the counting
1607: convention specified in footnote \protect\ref{counting}.}
1608: \begin{equation}
1609: \label{orland}
1610: \omega_{\cal K}\simeq A_{\cal K}N^{\alpha_{\cal K}-2}\mu_{\cal K}^N
1611: \left(1+\frac{B_{\cal K}}{N^{\Delta_{\cal K}}}+\ldots\right)
1612: \end{equation}
1613: with confluent corrections, that while for the unknot with $\alpha_{\emptyset}
1614: \approx 0.27$ expression (\ref{orland}) is consistent with the standard
1615: result (\ref{dof}) ($[0.27-2]/3\approx -0.58\approx \nu$), for prime knots
1616: $\alpha_{\cal K}=\alpha_{\emptyset}+1$, and for composite knots with $N_f$
1617: prime components,
1618: \begin{equation}
1619: \alpha_{\cal K}=\alpha_{\emptyset}+N_f.
1620: \end{equation}
1621: This finding is in agreement
1622: with the view that each prime component of a knot ${\cal K}$
1623: is tightly localised and statistically able to move around one central loop,
1624: each prime component counting an additional factor $N$ of degrees of freedom. The fact that
1625: for a chain of finite thickness the size of the big central loop is in fact
1626: diminished by the size of the tight knot is a confluent effect, such that the
1627: confluent exponent $\Delta$ should be related to the size distribution of
1628: the knot region.
1629: Not surprisingly, the connectivity factor $\mu_{\cal K}\approx 4.68$ was found
1630: to be independent of ${\cal K}$, assuming the standard value for a cubic
1631: lattice \cite{guttmann}. Also the amplitude $A_{\cal K}$ and the exponent
1632: $\Delta_{\cal K}$ of the confluent correction turned out to be ${\cal
1633: K}$-independent. We note that a similar analysis in (pseudo) 2D\footnote{The
1634: simulated polymer chain moves in 2D, however, crossings are permitted at
1635: which one chain passes underneath another. In that, the simulated polymers
1636: are in fact equivalent to knot projections with a certain orientations of
1637: individual crossings.} also strongly points towards tight localisation of
1638: the knot \cite{guitter}.
1639: 
1640: In contrast to the above results, 3D simulations undertaken in \cite{quake}
1641: (also compare \cite{quake1}) show the dependence
1642: \begin{equation}
1643: \label{quake}
1644: R_g\simeq N^{3/5}C^{-4/15}
1645: \end{equation}
1646: of the gyration radius on the knot type, characterised by the number $C$ of
1647: essential crossings. $R_g$, that is, decreases as a power-law with $C$,
1648: where the exponent $-4/15=1/3-\nu$ \cite{quake}. The functional form
1649: (\ref{quake})) was derived from a Flory-type argument for a polymer construct
1650: of $C$ interlocked loops of equal length $N/C$ by arguing that each loop
1651: occupies a volume $\simeq (N/C)^{3\nu}$, and the volume of the knot is given
1652: by $V\simeq C(N/C)^{3\nu}$ (i.e., assuming that due to self-avoiding repulsion
1653: the volume of individual loops adds up to the total volume). Equation
1654: (\ref{quake}) then follows immediately. This model of equal loop sizes is
1655: equivalent to a completely delocalised knot. It may therefore be speculated,
1656: albeit rather long chain sizes of up to 400 were used, whether the numerical algorithm
1657: employed for the simulations in \cite{quake} causes finite-size effects that,
1658: in turn, prevent a knot localisation. We note that the Flory-type
1659: scaling assumed to derive expression (\ref{quake}) is consistent with a
1660: modelling brought forward in reference \cite{grosberg2}, in which the knot is
1661: quantified by the aspect ratio in a configuration corresponding to a maximally
1662: inflated tube with the given topology (i.e., a state corresponding to
1663: complete delocalisation). In reference \cite{quake}, the temporal relaxation
1664: behaviour of a given knot was also studied. While regular
1665: Rouse behaviour was found for the case of the unknot, the knotted chains
1666: displayed somewhat surprising long time contributions to the relaxation
1667: time spectrum \cite{treloar,ward,ferry,erman}, a phenomenon
1668: already pointed out by de Gennes within an
1669: activation argument to create free volume in a tight knot in order to move
1670: along the chain \cite{degennesma}. Note that relatively lose knots in shorter
1671: chains do not appear to exhibit such extremely long relaxation time behaviour
1672: \cite{lai1}.
1673: 
1674: Simulation of a 3D knot with varying excluded volume showed, if only the
1675: excluded volume becomes large enough, the gyration radius of the knot is
1676: independent of the knot type \cite{shimamura}. The picture of tight knots
1677: is further corroborated in the study by Katritch {\it et al.} using a Gaussian
1678: chain model with fixed topology to demonstrate that the size distribution
1679: of the knot is distinctly peaked at rather small sizes \cite{katritch}.
1680: 
1681: Apart from determining the statistical quantities $R_g$ and $\omega_{\cal K}$
1682: from simulations, there also exist indirect methods for quantifying the size
1683: of the knot region in a knotted polymer. One such method is to confine an
1684: open chain containing a knot between two walls, and measuring the finite
1685: size corrections of the force-extension curve due to the knot size. This is
1686: based on the idea that the gyration radius for a system depending on more
1687: than one length scale (i.e., apart from the chain length $N$) shows above
1688: mentioned confluent corrections, such that \cite{farago}
1689: \begin{equation}
1690: R_g=AN^{\nu}\Phi\left(\frac{N_0}{N},\frac{N_1}{N},\ldots\right)\simeq
1691: AN^{\nu}\Big(1-BN^{-\Delta}\Big)
1692: \end{equation}
1693: when only the largest correction is considered, and in 3D $\Delta\approx 0.5$
1694: is supposed to be universal \cite{leguillou,zinn,janse1,li}.
1695: If this leading correction is due to the argument $N_0/N$ in the scaling
1696: function $\Phi$, the length scale $N_0$ depends on $N$ through the scaling
1697: $N_0\sim N^t$ with $\Delta=1-t$. From Monte Carlo simulations of a bead and
1698: tether chain model, it could then
1699: be inferred that the size of the knot scales like \cite{farago}
1700: \begin{equation}
1701: N_k\sim N^t, \quad t=0.4\pm0.1
1702: \end{equation}
1703: This, in turn, enters the force-extension curve $f'=G(R')$ with the
1704: dimensionless force $f'=fAN^{\nu}/(k_BT)$ and distance $R'=R/(AN^{\nu})$ of
1705: the walls, in the form with confluent correction
1706: \begin{equation}
1707: f'\simeq G(R')\Big(1+g(R')N^{-\Delta}\Big).
1708: \end{equation}
1709: From the simulation, $t=0.4$ corresponds to the best data collapsing, assuming
1710: the validity of
1711: the scaling arguments. An argument in favour of this approach is the
1712: consistency of the exponent $t=0.4$ with the inferred $\Delta=0.6$, which is
1713: close to the known value. Note that the force-extension of a chain with a
1714: slip-link was discussed in reference \cite{pull} and shown that a loop
1715: separated off by a slip-link is confined within a Pincus-de Gennes blob.
1716: We also note that results
1717: corresponding to delocalisation in force-size relations were reported in
1718: \cite{sheng,sheng1}. An entropic scale was conceived in \cite{roya}: Separating
1719: two chains with fixed topology but allowing them to exchange length (e.g.,
1720: through a small hole in a wall) would enable one to infer the localisation
1721: behaviour of a knot by comparing the equilibrium balance of this knot with
1722: a slip-link construct of known degrees of freedom until the
1723: average length on both sides coincides. The preliminary results in \cite{roya}
1724: are shadowed by finite-size effects of the accessible system size,
1725: as limited by computation power. The analysis in reference \cite{marcone} of a
1726: self-avoiding polygon model uses the
1727: method of closure of a short fragment of the knot and subsequent determination
1728: of its Alexander polynomial to obtain the scaling exponent $t=0.75$; in a
1729: second variant, the authors find a consistent result by a variant of the
1730: knot scale method. Another recent study uses a more realistic model for a
1731: polymer chain, namely, a simplified model of polyethylene; with up to 1000
1732: monomers in the simulation, the exponent $t\approx0.65$ is found (and
1733: delocalisation is obtained in the dense phase) \cite{virnau}.
1734: 
1735: Thus, there exist simulations results pointing in both directions, knot
1736: localisation and delocalisation. As the latter may be explained by finite
1737: size effects, it seems likely that (at least simple) knots in 2D and 3D
1738: localise in the sense that the knot region occupies a portion of the chain
1739: that is significantly smaller in comparison to the entire chain. In particular,
1740: this would imply that the average size of the knot region $\langle\ell\rangle$
1741: scales with the chain length $Na$ with an exponent less than one, such that
1742: \begin{equation}
1743: \lim_{N\to\infty}\frac{\langle\ell\rangle}{Na}=0.
1744: \end{equation}
1745: Below, we show from analytical grounds that such a localisation is a natural
1746: consequence of interactions of a chain of fixed topology with fluctuations.
1747: We note, however, that conclusive results for knot localisation may in fact
1748: come from experiments: Manipulation of single chains such as DNA can be
1749: performed for rather long chains, making it possible to reach beyond the
1750: finite-size corrections inherent in, e.g., the force-extension simulations
1751: mentioned above. The aforementioned AFM studies on single DNA knots indeed
1752: reveal knot localisation of flattened knots \cite{valle}; due to experimental
1753: limitations, presently only one DNA length was investigated, such that the
1754: scaling exponent $t$ currently cannot be obtained.
1755: 
1756: Before proceeding to these analytical approaches, we note that there have
1757: also been performed simulations of knotted chains under non-dilute conditions
1758: \cite{stella,orlanddense}.
1759: In (pseudo) 2D, these have found delocalisation of the knot, i.e., $\lim_{
1760: N\to\infty}\langle\ell\rangle/N=const$. We come back to these
1761: simulations below in connection with the modelling of dense and $\Theta$-knots.
1762: 
1763: 
1764: \subsection{Flattened knots in dilute and dense phases.}
1765: 
1766: Analytically, knots are a hard problem to tackle. Statistical mechanical
1767: treatments of permanently entangled polymers are so difficult to treat
1768: since topological restrictions cannot be formulated as a Hamiltonian problem
1769: but appear as hard constraints partitioning the phase space
1770: \cite{degennes,grosberg,vilgis,kholodenko}.\footnote{For comparison,
1771: self-avoidance in 3D is usually treated as a perturbation, i.e., as a ``soft
1772: constraint'', in analytical studies \protect\cite{degennes}.} A segment of a 3D
1773: knot, in other words, can move without feeling the constraints due to the
1774: non-trivial topology of a knotted state, until it actually collides with
1775: another segment. The accessible phase space of degrees of freedom is therefore characterised
1776: by inequalities.\footnote{Although a similar statement is true for polymer
1777: networks in 3D, the field theoretical results for their critical exponents are
1778: in fact obtained as {\em averages\/} over {\em all\/} topologies. For
1779: instance, the exponent $\nu$ entering the gyration radius of a a 3D polymer
1780: ring counts all knotted states \protect\cite{duplantier1}.}
1781: 
1782: Consequently, only a relatively small range of problems have
1783: been treated analytically, starting with the seminal papers by Edwards
1784: \cite{edwards1,edwards2}, in which he considers the classification of
1785: topological constraints in polymer physics. De Gennes addressed the problem
1786: of tight knot motion along a polymer chain using scaling arguments for the
1787: activation of free length inside the knot region, producing a double-exponential
1788: expression for the corresponding time scale \cite{degennesma}, which might
1789: explain the extreme long-time contributions in the relaxation time
1790: spectrum of permanently entangled polymers \cite{doi,treloar,ward,erman,ferry}.
1791: Some analytical results were obtained for a pair, or an `Olympic' gel of
1792: entangled polymer rings, see for instance,
1793: \cite{otto,otto1,vilgis1,ferrari,ferrari1}. In a mean field approach based on
1794: the Kauffman invariant the entropy of knots was investigated in references
1795: \cite{grone,grone1,grone2}. Similarly, some statistical properties of random
1796: knot diagrams were investigated in \cite{nechaev,nechaev1}. However, some
1797: insight
1798: can be gained on the basis of phenomenological models, which we will come back
1799: to below. Here, we continue with an analytical study of flat knots.
1800: 
1801: One possibility to treat knotted polymer chains analytically is to confine
1802: the degrees of freedom of the knot to motion in 2D, only. The knot, that is, is
1803: preserved, as at the crossings the chain is allowed to form an
1804: over-/underpassing, while the rest of the knot is confined to 2D. Such a
1805: confinement can in fact be experimentally realised in various ways. Thus,
1806: the chain can be confined between two close-by glass slabs, as demonstrated
1807: in \cite{craighead}; it can be pressed flat on a surface by gravitation or
1808: similar forces, for instance in macroscopic systems \cite{bennaim,bennaim1};
1809: the chain can be adhesively bound to a membrane and still reach
1810: configurational equilibrium, as experimentally shown for DNA in references
1811: \cite{valle,maier}. Or it can be adsorbed to a mica surface either by APTES
1812: coating or by providing bivalent Mg ions in solution, as shown in
1813: figure \ref{valle}. From such flat knots as discussed in
1814: the remainder of this section, we will
1815: be able to infer certain generic features also for 3D knots.
1816: 
1817: \begin{figure}
1818: \begin{center}
1819: \includegraphics{knot42.ps}
1820: \end{center}
1821: \caption{AFM tapping-mode images of flattened, complex DNA-knots with
1822: approximately 30-40 essential crossings, see \protect\cite{valle}.
1823: The substrate surface used is AP-mica (freshly cleaved mica
1824: reacted with an amino terminal silane to obtain a positively charged surface).
1825: The DNA knots used are extracted from bacteriophage P4; the DNA is a 11.4
1826: kbp molecule (with a 1.4 kbp deletion resulting in a final length of 10 kbp)
1827: which has two cohesive ends. They are not covalently closed, thus no
1828: supercoiling is present. The knot adsorbed out of the 3D bulk
1829: on to the surface is strongly trapped, i.e., the knot is `projected' onto the
1830: surface without any equilibration. The knot appears rather delocalised.
1831: Courtesy F. Valle and G. Dietler.
1832: \label{valle}}
1833: \end{figure}
1834: 
1835: A flat knot therefore corresponds to a polymer network in 2D, but the
1836: orientation of the crossings is preserved, such that the network graph
1837: actually coincides with a typical knot projection
1838: \cite{kauffman,reidemeister,adams}, as shown in figure \ref{tref_proj}
1839: on the left. This projection of the trefoil, and similar projections for
1840: all knots, displays the knot with the essential crossings. A flat knot can,
1841: in principle acquire an arbitrary number of crossings by Reidemeister moves;
1842: for instance, the bottom left segment of the flat trefoil can slide under
1843: the vicinal segment, creating a new pair of vertices, and so on. However,
1844: we suppose that such transient additional loops are sufficiently short-lived
1845: so that we can neglect them in our analysis. Then, we can apply results from
1846: scaling analysis of polymer networks of the most general type shown in
1847: figure \ref{network}, see the primer in the appendix. We note that from the
1848: Monte Carlo simulations we performed it may be concluded that such additional
1849: vertices can in fact be neglected.
1850: 
1851: 
1852: \subsubsection{Flat knots in dilute phase.}
1853: 
1854: We had previously found that for the F8-structure the probability density
1855: function for the size of
1856: each loop is peaked at $\ell\to 0$ and $\ell\to L$.
1857: From the scaling analysis for
1858: self-avoiding polymer networks, we concluded strong localisation of one
1859: subloop. For more complicated structures, the joint probability to find
1860: the individual segments with given lengths $s_i$ is expected to peak at
1861: the edges of the higher-dimensional configuration hyperspace. Some analysis
1862: is necessary to find the characteristic shapes. Let us consider here the
1863: simplest non-trivial knot, the (flat) trefoil knot $3_1$ shown in figure
1864: \ref{tref_proj}.
1865: \begin{figure}
1866: \begin{center}
1867: \includegraphics[width=7.2cm]{tref_proj.eps}
1868: \end{center}
1869: \caption{Flat trefoil knot with segment labels. On the right, a schematic
1870: representation of a localised flat trefoil with one large segment is
1871: shown.
1872: \label{tref_proj}}
1873: \end{figure}
1874: Each of the three crossings is replaced with a vertex with four
1875: outgoing legs, and the resulting network is assumed to separate into
1876: a large loop and a multiply connected region which includes the vertices.
1877: Let $\ell=\sum_{i=1}^5 s_i$ be the total length of all segments contained
1878: in the multiply connected knot region. Accordingly, the length of the large
1879: loop is $s_6=L-\ell$.
1880: 
1881: In the limit $\ell\ll L$, the number of configurations of this network can be
1882: derived in a similar way as in the scaling approach followed for the F8. This
1883: procedure determines the concrete behaviour of the scaling form
1884: \begin{equation}
1885: \omega_{\rm III}\simeq\mu^L{\cal W}_{\rm III}\left(L-\ell,\ell,
1886: \frac{s_1}{\ell},\frac{s_2}{\ell},\frac{s_3}{\ell},\frac{s_4}{\ell}\right)
1887: \end{equation}
1888: including the scaling function ${\cal W}$ that depends on altogether six
1889: arguments. The index III is chosen according to figure \ref{knot_zoo},
1890: where the flat trefoil configuration in the dense phase appears at
1891: position III of the scheme (explained below). After
1892: some manipulations, the number of degrees of freedom yields in the form \cite{slili2d}
1893: \begin{equation}
1894: \label{t_limit}
1895: \omega_{\text{III}}(\ell,L)\sim\mu^L (L-\ell)^{-d \nu}\ell^{-c},
1896: \end{equation}
1897: with the scaling exponent
1898: \begin{equation}
1899: c=-(\gamma_{\text{III}}-1+d\nu)-m, \qquad m=4.
1900: \end{equation}
1901: Here, $m=4$ corresponds to the number of independent integrations over the
1902: segments $s_i$ ($i=1,\ldots,4$) of the knot region, as we only retain the
1903: cumulative size $\ell=\sum_{i=1}^5s_i$ of the knot region. Putting numerical
1904: values, we find $c=65/16$, i.e., {\em strong localisation}.
1905: 
1906: \begin{figure*}
1907: \begin{center}
1908: \includegraphics{comschem.eps}
1909: \end{center}
1910: \caption{Hierarchy of the flat trefoil knot $3_1$. Upper row: dilute phase.
1911: Middle row: $\Theta$-phase. Bottom row: dense phase. To the left of each
1912: row, the trefoil projection is shown. It splits up into the hierarchies of
1913: configurations, with exponents $c$ below each contraction. The small protruding
1914: legs represent the big central loop, compare, for instance, figure
1915: \protect\ref{tref_proj} on the right with position III of the top row. See
1916: text for details.
1917: \label{knot_zoo}}
1918: \end{figure*}
1919: 
1920: However, some care is necessary in performing these integrations,
1921: since the scaling function ${\cal W}_{\rm III}$ may
1922: exhibit non-integrable singularities if one or more of the arguments
1923: $s_i/\ell$ tend to $0$. The geometries corresponding to these
1924: limits (edges of the configuration hyperspace)
1925: represent {\em contractions\/} of the original
1926: trefoil network ${\cal G}_{\text{III}}$ in the sense that the length of
1927: one or more of the segments $s_i$ is of the order of the
1928: short-distance cutoff $a$. If such a short segment connects
1929: different vertices, they cannot be resolved on larger length
1930: scales, but appear as a single, new vertex. Thus, each
1931: contraction corresponds to a different network ${\cal G}$,
1932: which may contain a vertex with up to eight outgoing legs. For
1933: the flat trefoil knot, there exist six different contractions,
1934: as grouped in figure \ref{knot_zoo} around the original flat
1935: trefoil at position III. As an example, in the top row of figure \ref{knot_zoo}
1936: contraction VI follows from the original trefoil III if the uppermost segment
1937: becomes very small, and similarly the network VII emanates from contraction VI
1938: if one of the four symmetric segments becomes very small.
1939: For each of these networks, one can calculate the corresponding
1940: exponent $c$ in a similar way as above, leading to the general
1941: expression
1942: \begin{equation}
1943: \label{euler}
1944: c=2+\sum_{N\ge 4}n_N\left\{\frac{N}{2}\Big(d\nu-1\Big)+
1945: \Big(|\sigma_N|-d\nu\Big)\right\}.
1946: \end{equation}
1947: The $\sigma_N$ are given in equation (\ref{top_coeff}). In figure
1948: \ref{knot_zoo}, the various contractions are arranged in increasing
1949: exponent $c$.
1950: 
1951: Our scaling analysis relies on an expansion in
1952: $a/\ell\ll 1$, and the values of $c$ determine a sequence of
1953: contractions according to higher orders in $a/\ell$: The {\em smallest\/}
1954: value of $c$ corresponds to the most likely contraction, while the
1955: others represent corrections to this leading scaling behaviour, and
1956: are thus less and less probable (see figure \ref{knot_zoo}).
1957: To lowest order, the trefoil behaves like a large ring polymer at whose fringe
1958: the point-like knot region is located. At the next level of resolution,
1959: it appears contracted to the figure-eight shape ${\cal G}_\mathrm{I}$.
1960: For more accurate data, the higher order
1961: shapes II to VII may be found with decreasing
1962: probability. Interestingly, the original uncontracted trefoil
1963: configuration ranks third in the hierarchy of shapes.
1964: 
1965: These predictions were checked by MC simulations with the same conditions as
1966: described above, to prevent intersection. The flat trefoil knot was prepared
1967: from a symmetric, harmonic 3D representation with 512 monomers, which was
1968: collapsed and then kept on a hard wall by the
1969: ``gravitational'' field $V=-k_BTh/h^*$ perpendicular to the wall, where $h$ is
1970: the height of a monomer, and $h^*$ was set to 0.3 times the bead diameter.
1971: Configurations corresponding to contraction I
1972: are then selected by requiring that besides a large loop, they contain
1973: only one segment larger than a preset cutoff length (taken to be 5 monomers),
1974: and similarly for contraction II.
1975: The size distributions for such contractions, as well as for all
1976: possible knot shapes are shown in figure \ref{figsl1}.
1977: \begin{figure}
1978: \begin{center}
1979: \includegraphics[width=8cm]{trefoil_pdf.eps}
1980: \end{center}
1981: \caption{Power law tails in probability density functions
1982: for the size $\ell$ of tight segments:
1983: As defined in the figure, we show results for the smaller loop in a
1984: figure-eight structure, the overall size of the trefoil knot,
1985: as well as the two leading contractions of the latter.
1986: \label{figsl1}}
1987: \end{figure}
1988: The tails of the distributions are indeed consistent with the predicted
1989: power laws, although the data (especially for contraction II) is too
1990: noisy for a definitive statement.
1991: 
1992: Our scaling results pertain to all flat prime knots. In particular, the
1993: dominating contribution for {\em any} prime knot corresponds to the
1994: figure-eight contraction ${\cal G}_{\text{I}}$, as equation (\ref{euler}) predicts
1995: a larger value of the scaling exponent $c$ for any network ${\cal G}$ other
1996: than ${\cal G}_{\text{I}}$. Accordingly, figure \ref{fig_mcknots} demonstrates
1997: \begin{figure}
1998: \begin{center}
1999: \includegraphics[height=4.4cm]{k3.eps}
2000: \includegraphics[height=4.4cm]{k8.eps}
2001: \includegraphics[height=4.4cm]{k33.eps}
2002: \end{center}
2003: \caption{Typical configurations of 256-mer chains for the trefoil $3_1$,
2004: the prime knot $8_{19}$, and the composite knot $3_1$\#$3_1$
2005: consisting of two trefoils, in $d=2$. The initial conditions were symmetric
2006: in all cases.
2007: \label{fig_mcknots}}
2008: \end{figure}
2009: the tightness of the prime knot $8_{19}$. Composite knots, however, can
2010: maximise the number of configurations by
2011: splitting into their prime factors as indicated
2012: in figure \ref{fig_mcknots} for $3_1$\#$3_1$. Each prime factor is
2013: tight and located at the fringe of one large loop, and accounts for
2014: an additional factor of $L$ for the number of configurations
2015: as compared to a ring of length $L$ without a knot.
2016: Indeed, this gain in entropy leads to the tightness of knots.
2017: Flat knots can experimentally be produced by `projecting' a dilute 3D knot
2018: from the bulk onto a mica surface, on which the knot is adsorbed. Variation
2019: of the ionic strength in the solution determines whether the knot is going to
2020: be strongly trapped on the surface such that, once captured on the surface, it
2021: is completely immobilised (small ionic strength); or whether the adsorption
2022: is weaker such that the knot can (partially) equilibrate while being confined
2023: to 2D, i.e., equilibrate as a flat knot.
2024: Figure \ref{valle} shows a strongly trapped complex knot, whereas figure
2025: \ref{erika} depicts a weakly adsorbed simple knot, compare \cite{valle}.
2026: 
2027: \begin{figure}
2028: \includegraphics[width=6.8cm,angle=90]{erika.eps}
2029: \caption{Flat knot imaged by AFM, similar to the one shown in figure
2030: \ref{valle}. However, this knot is rather simple (likely a trefoil)
2031: and was allowed to relax while attaching to freshly cleaved mica in
2032: the presence of bivalent Mg counterions. Courtesy E. Ercolini, J. Adamcik,
2033: and G. Dietler. Note the close resemblance to
2034: the trefoil configuration shown in figure \ref{fig_mcknots}.
2035: \label{erika}}
2036: \end{figure}
2037: 
2038: 
2039: \subsubsection{Flat knots under $\Theta$ and dense conditions.}
2040: 
2041: In many situations, polymer chains are not dilute.
2042: Polymer melts, gels, or rubbers exhibit fairly high densities of
2043: chains, and the behaviour of an individual chain in such systems
2044: is significantly different compared to the dilute phase
2045: \cite{degennes,doi,erman}. Similar considerations apply to biomolecules:
2046: in bacteria, the gyration radius of the almost freely floating
2047: ring DNA may sometimes be larger than the cell radius itself.
2048: Moreover, under certain conditions, there is a non-negligible
2049: osmotic pressure due to vicinal layers of protein molecules,
2050: which tends to confine the DNA \cite{walter,vasilevskaya,lerman}.
2051: In protein folding studies, globular proteins in their native
2052: state are often modelled as compact polymers on a lattice
2053: (see \cite{garel} for a recent review).
2054: 
2055: A polymer is considered dense if, on a lattice, the fraction $f$ of occupied
2056: sites has a finite value $f > 0$. This can be obtained by considering a single
2057: polymer of total length $L$ inside a box of volume $V$ and taking the limit
2058: $L \to \infty$, $V \to \infty$ in such a way that $f = L/V$ remains
2059: finite \cite{duplantier2,duplantier3,jacobsen}.
2060: Alternatively, dense polymers can be obtained in an infinite volume
2061: through the action of an attractive force between monomers.
2062: Then, for temperatures $T$ below the collapse
2063: (Theta) temperature $\Theta$, the polymers collapse to a dense phase,
2064: with a density  $f > 0$, which is a function of $T$
2065: \cite{duplantier3,duplantier4,owczarek,owczarek1}.
2066: For a dense polymer in $d$ dimensions, the exponent
2067: $\nu$, defined by the radius of gyration $R_g \sim L^{\nu}$, becomes
2068: $\nu = 1 / d$.
2069: The limit $f = 1$ is realised in Hamiltonian paths,
2070: where a random walk visits every site of a given
2071: lattice exactly once \cite{duplantier7,kondev}. Dense polymers may be related
2072: to 2D vesicles and lattice animals (branched polymers)
2073: \cite{fisher,seno,dekeyser,cardy}.
2074: 
2075: As studied in reference \cite{dense},
2076: the value of the exponent $c$ for the 2D dense F8 is (compare to the appendix)
2077: \begin{equation} \label{c8}
2078: c=-\gamma_{\rm F8}=11/8=1.375,
2079: \end{equation}
2080: implying that the smaller loop is {\em weakly localised\/}. This means
2081: that the probability for the size of each loop is peaked
2082: at $\ell=0$ and, by symmetry, at $\ell=L$.
2083: An analogous reasoning for the 2D F8 at the $\Theta$ point gives
2084: \begin{equation} \label{c8_theta}
2085: c=11/7=1.571.
2086: \end{equation}
2087: In both cases the smaller loop is weakly localised in the sense that $\langle
2088: l\rangle_</L\to 0$. Figure \ref{fig-8} shows
2089: \begin{figure}
2090: \begin{center}
2091: \includegraphics[width=4.2cm]{dense_f8_1.eps}
2092: \includegraphics[width=4.2cm]{dense_f8_3.eps}
2093: \end{center}
2094: \caption{Symmetric ($\ell=L/2=128$) initial configuration of a 2D dense
2095: F8 (left) and equilibrium configuration (right) with
2096: periodic boundary conditions. The two different grey values correspond to the
2097: two subloops created by the slip-link. The slip-link itself is
2098: represented by the three (tethered) black dots.\label{fig-8}}
2099: \end{figure}
2100: the symmetric initial and a typical equilibrium configuration for periodic
2101: boundary conditions obtained from Monte Carlo (MC) simulations, see reference
2102: \cite{dense} for details.
2103: In figure \ref{fig-8}, the lines represent the bonds (tethers) between the
2104: monomers (beads, not shown here). The three black dots mark the locations
2105: of the tethered beads forming the slip-link in 2D.
2106: The initial symmetric configuration soon gives way to a
2107: configuration with $\ell\ll L$ on approaching equilibrium.
2108: Figure \ref{fig-8_1} shows the development of this symmetry breaking as a
2109: function of the number of MC steps.
2110: We note, however,
2111: that the fluctuations of the loop sizes in the ``stationary'' regime
2112: appear to be larger in comparison to the dilute case studied in reference
2113: \cite{paraknot}, compare figure \ref{fig8_sim}.
2114: We checked that for densities (area coverage) above
2115: 40\% the scaling behaviour becomes independent of the density.
2116: (The above simulation results correspond to a density of 55\%.)
2117: The size distribution data is well fitted to a power law (for over 1.5 decades
2118: with 1024 monomers),
2119: and the corresponding exponent  with 512 and 1024
2120: monomers in figure \ref{fig-8_1}
2121: is in good agreement with the predicted value (\ref{c8}).
2122: 
2123: For our MC analysis, we again used a hard core
2124: bead-and-tether chain, in which self-crossings were prevented by keeping a
2125: maximum bead-to-bead distance of 1.38 times the bead diameter, and a maximum
2126: step length of 0.15 times the bead diameter. To create the dense F8 initial
2127: condition, a free F8 is squeezed into a quadratic box with hard
2128: walls. This is achieved by starting off from the free F8, surrounding it by a
2129: box, and turning on a force directed towards one of the edges. Then,
2130: the opposite edge is moved towards the centre of the box, and so on.
2131: During these steps, the slip-link is locked,
2132: i.e., the chain cannot slide through it, and the two loops are of equal length
2133: during the entire preparation. Finally, when the envisaged density is
2134: reached, the hard walls are replaced by periodic boundary conditions,
2135: and the slip-link is unlocked.
2136: After each step, the system is allowed to relax for times larger than the
2137: localisation times occurring at the main stage of the run.
2138: 
2139: \begin{figure}
2140: \begin{center}
2141: \includegraphics[height=6cm]{pdf_f8_dense.eps}
2142: \end{center}
2143: \caption{The loop size probability distribution  $p(\ell)$ at $\rho=55\%$ area
2144: coverage, for the F8 with 512 (top) and 1024 (bottom) monomers.
2145: The power law with with the predicted exponent $c=1.375$ in equation (\ref{c8})
2146: is indicated by the dotted line.
2147: \label{fig-8_1}}
2148: \end{figure}
2149: 
2150: A similar analysis as for the dense/$\Theta$-F8 structure and the dilute flat
2151: trefoil above, reveals the number of degrees of freedom for the flat dense trefoil
2152: in the form \cite{dense}
2153: \begin{equation}
2154: \label{3_limit}
2155: \omega_{3}(\ell,L)\sim\omega_{0}(L)\ell^{-c}
2156: \end{equation}
2157: with $c=-\gamma_3-m$, where $\gamma_3=-33/8$ from equation
2158: (\ref{nexp}) in the appendix (${\cal L} = 4$, $n_4 = 3$)
2159: and $m = 4$ is the number of independent integrations over chain
2160: segments. Thus, $c = 1/8 < 1$ which implies that the dense 2D
2161: trefoil is {\em delocalised}.
2162: As above, we have to consider the various possible contractions of the flat
2163: knot. For dense polymers, the present scaling results show
2164: that both the original trefoil shape ($c = 1/8 < 1$, see above)
2165: and position II ($c = 3/4 < 1$) are in fact {\em delocalised\/} and
2166: represent equally the leading scaling order (cf.~top part of figure
2167: \ref{knot_zoo}). The F8 is only found at the third position and is weakly
2168: localised ($c = 11/8 > 1$).
2169: In an MC simulation of the dense 2D trefoil, we predict that
2170: one mainly observes delocalised shapes corresponding to the original
2171: trefoil and position II in figure \ref{knot_zoo}, and further,
2172: with decreasing probability, the weakly localised F8 and the other
2173: shapes of the hierarchy (top part) in figure \ref{knot_zoo}.
2174: 
2175: These predictions are consistent with the numerical simulations of
2176: reference \cite{stella}, who observe that the mean
2177: value of the second largest segment of the simulated 2D dense trefoil
2178: configurations grows linearly with $L$, and conjecture the same
2179: behaviour also for the other segments, corresponding to the
2180: delocalisation of the trefoil obtained above.
2181: 
2182: An analogous reasoning can be applied to the 2D trefoil in the
2183: $\Theta$ phase. We find that
2184: in this case that the leading shape is again the
2185: original (uncontracted) trefoil, with $c = 5/7 < 1$. This
2186: implies that the 2D trefoil is {\em delocalised\/} also at the
2187: $\Theta$ point. All other shapes are at least weakly localised,
2188: and subdominant to the leading scaling order represented by the
2189: original trefoil. The resulting hierarchy of shapes is shown
2190: in figure \ref{knot_zoo} (bottom part).
2191: 
2192: 
2193: \subsection{3D knots defy complete analytical treatment.}
2194: 
2195: As already mentioned, 3D knots correspond to a problem involving hard
2196: constraints that defy a closed analytical treatment. It may be possible,
2197: however, that by a suitable mapping to, for instance, a field theory,
2198: an analytical description may be found. This may in fact be connected
2199: to the study of knots in diagrammatic solutions in
2200: high energy physics \cite{gambini}. There exists a fundamental relation
2201: between knots and gauge theory as knot projections and Feynman graphs
2202: share the same basic ingredients corresponding to a Hopf algebra
2203: \cite{kauffman}. However, up to now no such mapping has been found, and
2204: a theoretical description of 3D knots based on first principles is
2205: presently beyond hope. To obtain some insight into the statistical
2206: mechanical behaviour of knotted chains, one therefore has to resort
2207: to simulations studies or experiments. In addition, a few phenomenological
2208: models for both the equilibrium and dynamical behaviour of knots have been
2209: suggested such as in references
2210: \cite{quake,slili3d,lai1,sheng1,sheng,grosberg1,grosberg2,grosberg4}.
2211: 
2212: When discussing numerical knot studies, we already mentioned the Flory-type
2213: model brought leading to equation \ref{quake} \cite{quake,quake1}.
2214: One may argue that the
2215: differences in the knot size for the different knot types corresponding to
2216: the same $C$ may be included in the prefactor, that is independent of the
2217: chain length $N$. Obviously, this model of equal loop sizes is equivalent to
2218: a completely delocalised knot.
2219: This statement is in fact equivalent to another Flory-type approach to
2220: knotted polymers reported in \cite{grosberg2}. In this model, the knot is
2221: thought of as an inflatable tube: for a very thin tube diameter, the
2222: tube is equivalent to the original knot conformation; inflating the tube more
2223: and more will increasingly smoothen out the shape until a maximally inflated
2224: state is reached. The knot is then characterised by the aspect ratio
2225: \begin{equation}
2226: p=\frac{L}{D}\,\,, \, \mbox{therefore} \,\,
2227: 1\le p\le N,
2228: \end{equation}
2229: between length $L$ and maximum tube diameter $D$. It appears that $p$ is a
2230: (weak) knot invariant, and can be used to characterise the gyration radius of
2231: the knot. It is clear that, by construction, the aspect ratio described a
2232: totally delocalised knot, and indeed it turns out that in good solvent, the
2233: gyration radius shows the dependence $R_g\simeq AN^{3/5}\tau^{1/5}p^{-4/15}$,
2234: where $\tau$ is the (dimensionless) deviation from the $\Theta$ temperature
2235: \cite{grosberg2}. Obviously, the aspect ratio appears to be proportional to
2236: the number of essential crossings in comparison to expression (\ref{quake}).
2237: We note that similar considerations are employed in reference \cite{grosberg1},
2238: including a comparison to the entropy of a tight knot, finding comparable
2239: entropic likelihood. The modelling based on the aspect ratio $p$ is further
2240: refined in \cite{grosberg4}.
2241: 
2242: Knot localisation is a subtle interplay between the degrees of freedom of one big loop,
2243: and the internal degrees of freedom of the various segments in the knot region. Under
2244: localisation, the number of degrees of freedom
2245: \begin{equation}
2246: \omega\simeq\mu^NN^{1-d\nu}
2247: \end{equation}
2248: includes an additional factor $N$ from the knot region encircling the
2249: big loop. For flat knots, the competition between the single big loop and
2250: the knot region is indeed won by the big loop.
2251: In the case of 3D knots, this balance is presently not resolved for
2252: knots of all complexity. Probably only detailed simulations studies of higher
2253: order knots will make it possible to decide for the various models of 3D knots.
2254: Major contributions are also expected from single molecule experiments, for
2255: instance, from force-extension measurements along the lines of the simulations
2256: study in \cite{farago}, the advantage of experiments being the fact that it
2257: should be possible to go towards rather high chain lengths that are inaccessible
2258: in simulations.
2259: To overcome similar difficulties in the context of the entropic
2260: elasticity for rubber networks, Ball, Doi, Edwards and coworkers
2261: replaced permanent entanglements by slip-links
2262: \cite{edwards2a,edwards2b,edwards2c,edwards2d}.
2263: Gaussian networks containing slip-links have been successful in the prediction
2264: of important physical quantities of rubber networks \cite{erman}, and they
2265: have been used to study a small number of entangled chains \cite{sommer}. In a
2266: similar fashion, one may investigate the statistical behaviour of single polymer
2267: chains in which a fixed topology is created by a number of slip-links. Such
2268: `paraknots' can be studied analytically using the Duplantier scaling results
2269: \cite{paraknot}. As mentioned previously, knowledge of the statistical
2270: behaviour of paraknots can be used to create a knot scale for calibrating
2271: the degrees of freedom of real knots, and therefore also important to understand or design
2272: indirect experiments on knot entropy, such as by force-extension measurements
2273: \cite{pull}. Paraknots may also be useful in the design of entropy-based
2274: functional molecules \cite{hame_cpl,meamb_ctn}.
2275: 
2276: 
2277: \section{DNA breathing: local denaturation zones and biological implications}
2278: 
2279: "A most remarkable physical feature of the DNA helix, and one that is
2280: crucial to its functions in replication and transcription, is the ease
2281: with which its component chains can come apart and rejoin. Many techniques
2282: have been used to measure this melting and reannealing behaviour. Nevertheless,
2283: important questions remain about the kinetics and thermodynamics of denaturation
2284: and renaturation and how these processes are influenced by other molecules in
2285: the test tube and cell'' \cite{kornberg}. This remarkable quotation, despite
2286: 30 years old, still summarises the challenge of understanding local and global
2287: denaturation of DNA, in particular, its dynamics. In this section, we report
2288: recent findings on the spontaneous formation of intermittent denaturation
2289: zones within an intact DNA double helix. Such denaturation \emph{bubbles\/}
2290: fluctuate in size by (random) motion of the zipper forks relative to each
2291: other. The opening and subsequent closing of DNA bubbles is often called
2292: \emph{DNA breathing}.
2293: 
2294: \begin{figure}
2295: \includegraphics[width=8.6cm]{melt.eps}
2296: \caption{Fraction $\theta_h$ of double-helical domains within the DNA as a
2297: function of temperature. Schematic representation of $\theta_h(T)$,
2298: showing the increased formation of bubbles and unzipping from the ends,
2299: until full denaturation has been reached.
2300: \label{melting_curve}}
2301: \end{figure}
2302: 
2303: 
2304: \subsection{Physiological background of DNA denaturation}
2305: \label{breathintro}
2306: 
2307: The Watson-Crick double-helix is the
2308: thermodynamically stable configuration of a DNA molecule under physiological
2309: conditions (normal salt and room/body temperature). This stability is effected
2310: (a) by Watson-Crick H-bonding, that is essential for the specificity of
2311: base-pairing, i.e., for the key-lock principle according to which the
2312: nucleotide Adenine exclusively binds to Thymine, and Guanine only to
2313: Cytosine. Base-pairing therefore guarantees the high level of fidelity
2314: during replication and transcription. (b) The second contribution to
2315: DNA-helix stability comes from base-stacking between neighbouring bps:
2316: through hydrophobic interactions between the planar aromatic bases, that
2317: overlap geometrically and electronically, the bp stacking stabilises
2318: the helical structure against the repulsive electrostatic force between the
2319: negatively charged phosphate groups located at the outside of the DNA
2320: double-strand. While hydrogen bonds contribute only little
2321: to the helix stability, the major support comes from base-stacking
2322: \cite{kornberg,delcourt}.
2323: 
2324: The quoted ease with which its component chains can come apart and rejoin,
2325: without damaging the chemical structure of the two single-strands, is
2326: crucial to many physiological
2327: processes such as replication via the proteins DNA helicase and polymerase,
2328: and transcription through RNA polymerase. During these processes, the proteins
2329: unzip a certain region of the double-strand, to obtain access to the genetic
2330: information stored in the bases in the core of the double-helix
2331: \cite{kornberg,revzin,watsoncrick}. This unzipping corresponds to breaking
2332: the hydrogen bonds between the bps. Classically, the so-called melting
2333: and reannealing behaviour of DNA has been studied in solution in vitro by
2334: increasing the temperature, or by titration with acid or alkali. During thermal
2335: melting, the stability of the DNA duplex is related to the content of
2336: triple-hydrogen-bonded G-C bps: the larger the fraction of G-C pairs,
2337: the higher the required melting temperature or pH value. Thus, under thermal
2338: melting, dsDNA starts to unwind in regions rich in A-T bps, and then
2339: proceeds to regions of progressively higher G-C content
2340: \cite{kornberg,delcourt}. Conversely, molten, complementary chains of
2341: single-stranded DNA (ssDNA) begin to reassociate
2342: and eventually reform the original double-helix under incubation at roughly
2343: 25$^\circ$ below the melting temperature $T_m$ \cite{kornberg}. The relative
2344: amount of molten DNA in a solution can be measured by UV spectroscopy,
2345: revealing large changes in absorption in the presence of perturbed
2346: base-stacking \cite{wartell}. Careful melting studies allow one to obtain
2347: accurate values for the stacking energies of the various combinations
2348: of neighbouring bps, a basis for detailed thermodynamic modelling
2349: of DNA-melting and DNA-structure per se \cite{blake,blossey}. In fact,
2350: thermal melting data have been successfully
2351: used to identify coding sequences of the genome due to the different
2352: G-C content \cite{yeramian,yeramian1,carlon}.
2353: 
2354: Complementary to thermal or pH induced denaturation, dsDNA can be driven
2355: toward denaturation mechanically, by applying a tensional stress along
2356: the DNA in an optical tweezer trap \cite{williams}. As shown in figure
2357: \ref{mark_stretch}, the force per extension increases in worm-like chain
2358: fashion, until a plateau at approximately 65 pN is reached. This plateau
2359: is sometimes interpreted as new DNA configuration, the S form \cite{busta}.
2360: By a series of experiments, it appears more likely that the plateau corresponds
2361: to the mechanical denaturation transition \cite{williams1}. To first order,
2362: the effect of
2363: the longitudinal pulling translates into an external torque $\mathfrak{T}$,
2364: whose effect is a decrease in the free energy for melting a bp:
2365: \begin{equation}
2366: \Delta G_F=\Delta G_{F=0}-\mathfrak{T}\theta_0,
2367: \end{equation}
2368: where $\theta_0=2\pi/10.35$ is the twist angle per bp of the double helix
2369: \cite{hwa}.
2370: 
2371: \begin{figure}
2372: \includegraphics[width=8cm]{williams.eps}
2373: \caption{Overstretching of double-stranded DNA. The black curve shows the
2374: typical force-extension behaviour of DNA following the rapid worm-like chain
2375: increase until at around 65 pN a plateau is reached. Crossing of the plateau
2376: corresponds to progressive mechanical denaturation. See text for details.
2377: Figure courtesy Mark C. Williams.}
2378: \label{mark_stretch}
2379: \end{figure}
2380: 
2381: An important application of thermal DNA melting is the Polymerase Chain
2382: Reaction (PCR). In PCR, dsDNA is melted at elevated temperatures into two
2383: strands of ssDNA. By lowering the temperature in a
2384: solution of invariable primers and single nucleotides, each ssDNA is completed
2385: to dsDNA by the key-lock principle of base-pairing \cite{pcr,pcrrev}.
2386: By many such cycles,
2387: of the order of $10^9$ copies of the original DNA can be produced within the
2388: range of hours.\footnote{Most proteins denature
2389: at temperatures between 40 to 60$^\circ$C, including polymerases. In early
2390: PCR protocols, after each heating step new polymerase had to be washed into
2391: the reaction chamber. Modern protocols make use of heat-resistant polymerases
2392: that survive the temperatures necessary in melting. Such heat-resistant proteins
2393: occur, for instance, in bacteria dwelling near undersea thermal vents.}
2394: Again, the error rate due to the underlying biochemistry can be considered
2395: negligible for most purposes. In particular, from the viewpoint
2396: of polymer physics/chemistry, the obtained sample is monodisperse and free
2397: of parasitic reactions, creating (almost) ideal samples for physical studies,
2398: in particular, as any designed sequence of bases can be custom-made in modern
2399: molecular biology labs \cite{snustad}.
2400: 
2401: While the double-helix is the thermodynamically stable configuration
2402: of the DNA molecule below $T_m$ (at non-denaturing pH), even at
2403: physiological conditions there exist
2404: local denaturation zones, so-called DNA-bubbles, predominantly in A-T-rich
2405: regions of the genome \cite{wartell,poland}. Driven by ambient thermal
2406: fluctuations, a DNA-bubble is a dynamical entity whose size varies by
2407: thermally activated zipping and unzipping of successive bps at the two
2408: forks where the ssDNA-bubble is bordered by the dsDNA-helix. This incessant
2409: zipping and unzipping leads to a random walk in the bubble-size coordinate,
2410: and to a finite lifetime of DNA-bubbles under non-melting conditions, as
2411: eventually the bubble closes due to the energetic preference for the closed
2412: state \cite{wartell,poland}. DNA-breathing typically
2413: opens up a few bps \cite{gueron,krueger}. It has been demonstrated
2414: recently that by fluorescence correlation methods the fluctuations of
2415: DNA-bubbles can be explored on the single molecule level, revealing a
2416: multistate kinetics that corresponds to the picture of successive
2417: zipping and unzipping of single bps.\footnote{Essentially, the zipper
2418: model advocated by Kittel \protect\cite{kittel}.} At room temperature, the
2419: characteristic closing time of an unbounded bp was found to be in
2420: the range 10 to 100 $\mu$sec corresponding to an overall bubble lifetime in
2421: the range of a few msec \cite{altan}. The multistate nature of the
2422: DNA-breathing was confirmed by a UV-light absorption study \cite{zocchi}.
2423: The zipping dynamics of DNA is also investigated by NMR methods
2424: \cite{nmr,nmr1,nmr2}, revealing considerably shorter time scales than the
2425: fluorescence experiments. An interesting finding from NMR studies is the
2426: dramatically different denaturation dynamics in B' DNA, where more than
2427: three AT bps occur in a row \cite{russu}. It is conceivable that fluorescence
2428: correlation and NMR probe
2429: different levels of the denaturation dynamics. Our analysis of the single
2430: DNA fluorescence data reported below demonstrates that, albeit the much
2431: longer time scale, the dependence of the measured autocorrelation function
2432: on the stacking along the sequence is very sensitive, and agrees well with
2433: the quantitative behaviour predicted from the stability data.
2434: 
2435: The presence of fluctuating DNA-bubbles is essential to the
2436: understanding of the binding of single-stranded DNA binding proteins
2437: (SSBs) that selectively bind to ssDNA, and that play important roles
2438: in replication, recombination and repair of DNA \cite{kornberg1}. One
2439: of the key tasks of SSBs is to prevent the formation of secondary
2440: structure in ssDNA \cite{alberts,snustad}. From the thermodynamical
2441: point of view one would therefore expect SSBs to be of an effectively
2442: helix-destabilising nature, and thus to lower $T_m$
2443: \cite{vonhippel}. However, it was found that neither the gp32 protein
2444: from the T4 phage nor E.coli SSBs do
2445: \cite{vonhippel,vonhippel1,rich}. An explanation to this apparent
2446: paradox was suggested to consist in a kinetic block, i.e., a kinetic regulation
2447: such that the rate constant for the binding of SSBs is smaller than
2448: the one for bubble closing \cite{vonhippel1,rich1}. This hypothesis
2449: could recently be verified in extensive single molecule setups using
2450: mechanical overstretching of dsDNA by optical tweezers in the presence
2451: of T4 gene 32 protein \cite{pant,pant1,pant2}, as detailed below.
2452: 
2453: 
2454: \subsection{The Poland-Scheraga model of DNA melting}
2455: 
2456: The most widely used approach to DNA melting in bioinformatics is the
2457: statistical, Ising model-like Poland-Scheraga model (sometimes also referred
2458: to as Bragg-Zimm model) and its variations \cite{poland,cantor,wartell};
2459: see also \cite{kafri,kafri1,richard,hanke}. It
2460: defines
2461: the partition function $\mathscr{Z}$ of a DNA molecule in a grand canonical
2462: picture with arbitrary many bubbles. For simplicity, we will restrict the
2463: following discussion to a single bubble. Below the melting temperature
2464: $T_m$, the one bubble picture is a good approximation: due to the high
2465: energy cost of bubble initiation, the distance between bubbles on a DNA
2466: molecule is large, and bubbles behave statistically independently. In
2467: typical experimental setups for measuring the bubble dynamics (see below),
2468: the used DNA construct is actually designed to host an individual bubble.
2469: For a homopolymer stretch of double-stranded DNA with 400 bps, figure \ref{400}
2470: shows the probabilities to find zero, one or, or two bubbles as a function
2471: of the Boltzmann factor $u=\exp(\Delta G/RT)$ for denaturation of a
2472: single bp.\footnote{In biochemistry, energies are usually measured in calories
2473: per mol. Instead of the Boltzmann factor $\beta=1/k_BT$ commonly used in physics
2474: and engineering, it is therefore convenient to replace the Boltzmann constant
2475: $k_B$ by the gas constant $R=k_BN_A$, where $N_A$ is the Avogadro-Loschmidt
2476: number.} Even at the denaturation transition $\Delta G=0$, it is
2477: quite unlikely to find two bubbles simultaneously.
2478: 
2479: \begin{figure}
2480: \includegraphics[width=8cm]{bubble_M400.eps}
2481: \caption{Probability of having 0, 1, or 2 bubbles as a function of $u$
2482: for a DNA region of chain length 400 bps. The cooperativity parameter
2483: was $\sigma_0=10^{-3}$ and the loop correction exponent c=1.76 (see text).}
2484: \label{400}
2485: \end{figure}
2486: 
2487: The free energy $\Delta G$ to break an individual bp are constructed as
2488: follows. We mention two different approaches. Common for both is the
2489: Poland-Scheraga construction of the partition function. We start with
2490: the case that a linear DNA molecule denatures from one of its ends. The
2491: corresponding partition function is \cite{poland,wartell}
2492: \begin{equation}
2493: \label{endpart}
2494: \mathscr{Z}_{\mathrm{end}}(m)=\prod_{x=1}^me^{\Delta G_{x,x+1}/RT}
2495: \end{equation}
2496: where $m$ is the number of broken bps, and $\Delta G_{x,x+1}$ is the stacking
2497: free energy for disruption of the bp at position $x$ measured from the end
2498: of the DNA. The notation explicitly refers to the stacking between the bp at
2499: $x$ and $x+1$.
2500: The first closed bp is located at $x=m+1$. For a homonucleotide, $\Delta
2501: G_{x,x+1}=\Delta G$, while for a given sequence of bps, there come into play
2502: the different stacking energies for the possible combinations of pairs of bps
2503: in sequence\footnote{I.e., an AT bp followed by another AT as different from
2504: an AT followed by a TA, etc.} The stacking energies $\Delta G_{x,x+1}$ have
2505: the following contributions.
2506: 
2507: The more traditional way to determine the stacking interactions is by fit of
2508: bulk melting curves of DNA constructs containing exclusively pairs of the
2509: specific bp-bp combination such as $(\mathrm{AT/TA})_n$
2510: (see, e.g., \cite{santalucia,delcourt} and references therein). The free energy
2511: used in this automated fit procedure using the MELTSIM algorithm \cite{blake},
2512: \begin{equation}
2513: \label{eblake}
2514: \Delta G^{\mathrm{Mix}}_{x,x+1}=\Delta H^{\mathrm{ST}}_{x,x+1}-T\Delta
2515: S_{x,x+1},
2516: \end{equation}
2517: combines the stacking enthalpy difference $\Delta H^{\mathrm{ST}}_{x,x+1}$
2518: for both hydrogen bond and actual stacking energies, and the entropy difference
2519: $\Delta S_{x,x+1}$ chosen to explicitly depend on the nature of the broken bp.
2520: A recent alternative to determine the stability parameters of DNA was
2521: developed in the group of Frank-Kamenetskii, leading to the free energy
2522: \cite{krueger} 
2523: \begin{equation}
2524: \label{ekrueger}
2525: \Delta G^{\mathrm{Sep}}_{x,x+1}=\Delta G^{\mathrm{ST}}_{x,x+1}+\Delta G
2526: ^{\mathrm{HB}}_x
2527: \end{equation}
2528: where the Gibbs free energies $\Delta G^{\mathrm{ST}}_{x,x+1}$ and
2529: $\Delta G^{\mathrm{HB}}_x$ measure the stacking of bps $x$ and $x+1$ and
2530: the hydrogen bonding of bp $x$ \emph{including} the entropy release on
2531: disruption. Note that $\Delta G^{\mathrm{HB}}_x$ is chosen such that it only
2532: depends on the broken bp and has two values for AT and
2533: GC bps, irrespective of the orientation (3' or 5'). The stacking free energies
2534: $\Delta G^{\mathrm{ST}}$ were determined from denaturation at a DNA nick and
2535: show a pronounced asymmetry between AT/TA and TA/AT bonds \cite{krueger}. For an
2536: end-denaturing DNA both descriptions are equivalent (though somewhat different
2537: when one puts numbers), as the breaking of each bp involves the disruption of
2538: one hydrogen bonds of bp $x$ and one stacking with its neighbour.
2539: 
2540: The difference between the two approaches becomes apparent when we consider
2541: the initiation of a bubble, i.e., a denatured coil enclosed by intact
2542: double-helix. Now, the partition function for a bubble with left fork
2543: position at $x_L$ and consisting of $m$ broken bps,
2544: \begin{equation}
2545: \label{midpart}
2546: \mathscr{Z}_{\mathrm{mid}}(x_L,m)=\Lambda\Sigma(m)\prod_{x=x_L}^{x_L+m}
2547: e^{\Delta G_{x,x+1}/RT},
2548: \end{equation}
2549: differs from (\ref{endpart}) in three respects: (i) While the bubble consists
2550: of $m$ molten bps, $m+1$ stacking interactions need to be broken to create
2551: two boundaries between intact double-strand and the single-strand in the
2552: bubble; the extra stacking interaction is effectively incorporated into
2553: $\Lambda$. (ii) The polymeric nature of the flexible single-stranded
2554: bubbles involves the entropy loss factor $\Sigma(m)=1/(m+D)^c$ with
2555: critical exponent $c$ and the parameter $D$ to take care of finite
2556: size effects\footnote{Usually, $D=1$ is chosen.} \cite{richard,blake,fixman};
2557: (iii) the factor $\Lambda$: In the standard notation, $\Lambda
2558: \equiv\sigma_0=\exp(-F_s/RT)\simeq10^{-4}-10^{-5}$ \cite{lazurkin,blake,%
2559: wartell,poland}, while according to \cite{krueger}, $\Lambda=\xi\simeq10^{-3}$
2560: is called the ring factor.
2561: Interestingly, the cooperativity parameter $\sigma_0$ is of the order of
2562: what corresponds to the singular unbalanced stacking enthalpy for breaking
2563: the first bp to initiate the bubble.
2564: The new stability data lead to a more pronounced asymmetry in opening
2565: probabilities between different bp-bp combinations. The analysis in
2566: references \cite{tobias_prl,tobias_long} demonstrates that the parameters
2567: from \cite{krueger} appear to support better the biological relevance
2568: of the TATA motif\footnote{The four bp {\small $\begin{array}{l}\cdot\mathtt{
2569: TATA}\cdot\\[-0.16cm]\cdot\mathtt{ATAT}\cdot\end{array}$} sequence is one of
2570: the typical codes marking where RNA polymerase starts the transcription process
2571: \cite{alberts,snustad}.} in natural sequences, that is, show a more pronounced
2572: simultaneous opening probability for the TATA motif.
2573: 
2574: As demonstrated for the autocorrelation function measuring the breathing
2575: dynamics in figure \ref{autocorr}, the description in terms of the
2576: partition function $\mathscr{Z}$ based on the stability parameters from
2577: \cite{krueger} reproduces the experimental data well. The analysis in
2578: \cite{tobias_prl,tobias_long} also indicates that the accuracy of the
2579: model predictions for the bubble dynamics is rather sensitive to the
2580: parameters. It is therefore conceivable that improved fluorescence
2581: measurement of the bubble dynamics may be employed to obtain accurate DNA,
2582: stability parameters, complementing the more traditional melting, NMR, and
2583: gel electrophoresis
2584: bulk methods. It should be noted that the dynamics is strongly influenced
2585: by local deviations from the B configuration of the DNA double helix. Thus,
2586: in local stretches of more than three AT bps in sequence, the B' structure
2587: is assumed, leading to pronouncedly different zipping dynamics \cite{russu}.
2588: 
2589: Two major questions remain in the thermodynamic formulation of DNA denaturation
2590: and its dynamics. Namely, the exact origin of the bubble initiation factor
2591: $\sigma_0$ (or, alternatively, the ring factor $\xi$ from \cite{krueger}), and
2592: a method to properly calibrate the zipping rate $k$. The
2593: factor $\sigma_0$ is related to the entropic imbalance on opening the
2594: first bp of a bubble: While this requires the breaking of two stacking
2595: interactions, only one bp has access to an increased amount of degrees
2596: of freedom. Still, these degrees of freedom must be influenced by the
2597: fact that the single open bp is coupled to two zipper forks. Currently,
2598: $\sigma_0$ remains a fit parameter. The exact value of the zipping rate
2599: $k$ remains open. While NMR experiments indicate much faster rates in
2600: the nanosecond range ($\simeq 10^8\mbox{ sec}^{-1}$), the fluorescence
2601: correlation measurements produce values in the microsecond range ($\simeq
2602: 10^4-10^5\mbox{ sec}^{-1}$). This large discrepancy may be based on the
2603: fact that both methods have different sensitivity to the amplitude of
2604: intra-bp separation. Currently, $k$ is taken as a fit parameter. In the
2605: analysis in \cite{tobias_prl,tobias_long}, we use the stacking parameters
2606: from \cite{krueger} including the value of the ring factor $\xi$, so that
2607: $k$ (a shift along the logarithmic abscissa) is the only adjustable
2608: parameter.
2609: 
2610: It has been under debate what exact value should be taken for the critical
2611: exponent $c$ entering in the entropy loss factor for a denaturation bubble.
2612: This is connected to the fact that $c>2$ would imply a first order denaturation
2613: transition on melting, while $1<c<2$ would stand for a second order
2614: transition \cite{fisher,richard}. Speculations about a possible first order
2615: transition are connected to the rather distinct spikes in the differential
2616: melting curves \cite{wartell}.\footnote{Due to the rather small DNA samples
2617: used in melting experiments (5000 bp and less \cite{wartell}), claims about
2618: the order of the underlying thermodynamical phase transition should be
2619: considered with some care.} Theoretical polymer physics approaches to explain
2620: a $c>2$ are either based on the inclusion of polymeric self-avoidance
2621: interactions of the bubble with the remainder of the chain \cite{kafri}; or
2622: built on a directed polymer model \cite{garel1}. Despite the elegance of both
2623: approaches, it is an open question how truly they represent the detailed
2624: denaturation behaviour of real DNA \cite{hame_comment,somendra_comment}.
2625: Applying the MELTSIM algorithm to typical sequences, it was found that there
2626: is a connection between the fit result for the cooperativity parameter $\sigma
2627: _0$, whose value is reduced from $\approx 10^{-5}$ to $\approx 10^{-3}$
2628: by assuming $c=2.12$ instead of $1.76$ \cite{blossey}. Below the melting
2629: transition, the typical bubble size is only a few bps, and in that regime the
2630: polymeric treatment of the loop entropy loss is of approximative nature.
2631: Indeed, in the analysis of \cite{krueger} no entropy loss due to polymer
2632: ring formation was included. For the breathing dynamics, we include $c$,
2633: to cover higher temperatures with somewhat larger bubbles,
2634: but find no significant change in the behaviour between $c<2$ and $c>2$,
2635: as long as the exponent is sufficiently close to 2.
2636: We therefore use the value $c=1.76$, that is consistent with the traditional
2637: data fits employed in the determination of the stacking parameters.
2638: 
2639: 
2640: \subsection{Fluctuation dynamics of DNA bubbles: DNA breathing}
2641: 
2642: Below the melting temperature $T_m$, DNA bubbles are intermittent, i.e.,
2643: they form spontaneously due to thermal fluctuations, and after some time close
2644: again. DNA-breathing can be thought of as a biased random walk in the phase
2645: space spanned by the bubble size $m$ and its position denoted, e.g., by the
2646: left zipper fork position $x_L$ \cite{tobias_prl,tobias_long}. The bubble
2647: creation can be viewed
2648: as a nucleation process \cite{landau}, whereas the bubble lifetime corresponds
2649: to the survival time of the first passage problem of relaxing to the $m=0$
2650: state after a random walk in the $m>0$ halfspace
2651: \cite{hame_jpa,tobias_jpc,tobias_prl,tobias_long,tobias_pre}. Bubble
2652: breathing on the
2653: single DNA-bubble level was measured by fluorescence correlation spectroscopy
2654: in \cite{altan}. This technique employs a designed stretch of DNA, in which
2655: weaker AT bps form the bubble domain, that is clamped by stronger GC bonds.
2656: In the bubble domain, a fluorophore-quencher pair is attached. Once the bubble
2657: is created, fluorophore and quencher are separated, and fluorescence occurs.
2658: A schematic of this setup is shown in figure \ref{fig:bubbles}.
2659: 
2660: \begin{figure}
2661: \begin{center}
2662: \includegraphics[width=6.8cm]{DNA_cartoon_xT.eps}
2663: \end{center}
2664: \caption{Clamped DNA domain with internal bps $x=1$ to $M$,
2665: statistical weights $u_{\rm hb}(x)$, $u_{\rm st}(x)$, and tag position
2666: $x_T$. The DNA sequence enters through the statistical weights $u_{\rm
2667: st}(x)$ and $u_{\rm hb}(x)$ for disrupting stacking and hydrogen bonds
2668: respectively. The bubble breathing process consists of the initiation
2669: of a bubble and the subsequent motion of the forks at positions $x_L$
2670: and $x_R$. See \cite{tobias_long} for details.}
2671: \label{fig:bubbles}
2672: \end{figure}
2673: 
2674: The zipper forks move stepwise $x_{L/R}\rightarrow x_{L/R}\pm 1$ with rates
2675: $\mathsf{t}^{\pm}_{L/R}(x_{L/R},m)$. We define for bubble size decrease
2676: \begin{equation}
2677: \label{eq:t_L_plus}
2678: \mathsf{t}^+_L(x_L,m)=\mathsf{t}^-_R(x_L,m)=k/2 \qquad (m\ge 2)
2679: \end{equation}
2680: for the two forks.\footnote{Due to intrachain coupling (e.g., Rouse), larger
2681: bubbles may involve an additional `hook factor' $m^{-\mu}$ \cite{tobias_jpc}.}
2682: The rate $k$ characterises a single bp zipping.
2683: Its independence of $x$ corresponds to the view that bp closure requires the
2684: diffusional encounter of the two bases and bond formation; as sterically AT
2685: and GC bps are very similar, $k$ should not significantly vary with bp stacking.
2686: The rate
2687: $k$ is the only adjustable parameter of our model, and has to be determined
2688: from experiment or future MD simulations. The factor $1/2$ is introduced for
2689: consistency \cite{tobias_jpc,tobias_pre,tobias_prl,tobias_long,tobias_leif}.
2690: Bubble size increase is controlled by
2691: \begin{eqnarray}
2692: \nonumber
2693: \mathsf{t}_{L}^{-}(x_L,m)&=&ku_{\rm st}(x_L) u_{\rm hb}(x_{L})
2694: s(m)/2,
2695: \label{eq:t_L_minus}\\
2696: \mathsf{t}_{R}^{+}(x_L,m)&=&ku_{\rm st}(x_R+1) u_{\rm hb}(x_{R})
2697: s(m)/2,
2698: \label{eq:t_R_plus}
2699: \end{eqnarray}
2700: for $m\ge 1$, where $s(m)=\{(1+m)/(2+m)\}^c$. Finally, bubble initiation and
2701: annihilation from and to the zero-bubble ground state, $m=0 \leftrightarrow
2702: 1$ occur with rates
2703: \begin{eqnarray}
2704: \nonumber
2705: \mathsf{t}_G^+(x_L)&=&k\xi's(0)u_{\rm st}(x_L+1)u_{\rm hb}(x_L+1)u_{\rm
2706: st}(x_L+2)\\
2707: \mathsf{t}_G^-(x_L)&=&k.
2708: \label{eq:t_G_plus}
2709: \end{eqnarray}
2710: The rates $\mathsf{t}$ fulfil detailed balance conditions. The annihilation
2711: rate $\mathsf{t}_G^-(x_L)$ is twice the zipping rate of a single fork, since
2712: the last open bp can close either from the left or right. Due to the clamping,
2713: $x_L\ge 0$ and $x_R\le M+1$, ensured by reflecting conditions $\mathsf{t}
2714: _L^-(0,m)=\mathsf{t}_R^+(x_L,M-x_L)=0$. The rates $\mathsf{t}$ together
2715: with the boundary conditions fully determine the bubble dynamics.
2716: 
2717: In the FCS experiment fluorescence occurs if the bps in a
2718: $\Delta$-neighbourhood of the fluorophore position $x_T$ are open
2719: \cite{altan}. Measured fluorescence time series thus correspond to
2720: the stochastic variable $I(t)$, that takes the value 1 if at least all bps
2721: in $[x_T-\Delta,x_T+\Delta]$ are open, else it is 0. The time averaged
2722: ($\overline{\,\,\cdot\,\,}$) fluorescence autocorrelation
2723: \begin{equation}
2724: \label{autocf}
2725: A_t(x_T,t)=\overline{I(t)I(0)}-\overline{I(t)}^2
2726: \end{equation}
2727: for the sequence AT9 from \cite{altan} are rescaled in figure \ref{autocorr}.
2728: 
2729: \begin{figure}
2730: \begin{center}
2731: \includegraphics[width=9.1cm]{autocorr_at9.eps}
2732: \end{center}
2733: \caption{Time-dependence of the autocorrelation function $A_t(x_T,t)$ for
2734: the sequence AT9 measured in the FCS setup of reference \cite{altan} at 100mM
2735: NaCl. The full
2736: lines show the result from the master equation, based on the DNA stability
2737: parameters from Krueger et al.~\cite{krueger}. The inset shows the broadening
2738: of the relaxation time spectrum with increasing temperature.}
2739: \label{autocorr}
2740: \end{figure}
2741: 
2742: We note that an alternative method to obtain precise DNA stability data
2743: may be provided by a DNA construct with two AT-rich zones
2744: between which a shorter GC-rich bridge is located. The first passage
2745: problem corresponding to bubble merging at temperatures between the
2746: melting temperatures of the AT and GC zones was recently calculated
2747: \cite{hansen}, and provides the framework for modified fluorescence
2748: correlation setups similar to the one from reference \cite{altan}.
2749: 
2750: 
2751: \subsection{Probabilistic modelling---the master equation (ME)}
2752: 
2753: DNA breathing is described by the probability
2754: distribution $P(x_L,m,t)$ to find a bubble of size $m$ located at $x_L$
2755: whose time evolution follows the ME
2756: \cite{tobias_jpc,tobias_pre,tobias_prl,tobias_long,tobias_leif}
2757: \begin{equation}
2758: \label{master}
2759: \frac{\partial}{\partial t}P(x_L,m,t)=\mathbb{W} P(x_L,m,t).
2760: \end{equation}
2761: The transfer matrix $\mathbb{W}$ incorporates
2762: the rates $\mathsf{t}$. Detailed balance guarantees equilibration toward
2763: $\lim_{t\to\infty}P(x_L,m,t)=\mathscr{Z}(x_L,m)/\mathscr{Z}$, with
2764: $\mathscr{Z}=\sum_{x_L,m}\mathscr{Z}(x_L,m)$
2765: \cite{vankampen}. The
2766: ME and the explicit construction of $\mathbb{W}$ are discussed at
2767: length in references \cite{tobias_jpc,tobias_long}. Eigenmode analysis and matrix
2768: diagonalisation produces all quantities of interest such as the ensemble
2769: averaged autocorrelation function
2770: \begin{equation}
2771: \label{auto}
2772: A(x_T,t)=\langle I(t)I(0)\rangle-(\langle I\rangle)^2.
2773: \end{equation}
2774: $\langle I(t)I(0)\rangle$ is proportional to the survival density that the
2775: bp is open at $t$ and that it was open initially \cite{tobias_prl,tobias_long}.
2776: 
2777: In figure \ref{autocorr} the blue curve shows the predicted behaviour of $A(
2778: x_T,t)$, calculated for $T=49^\circ$C with the parameters from
2779: \cite{krueger}. As in the experiment we assumed that fluorophore and quencher
2780: attach to bps $x_T$ and $x_T+1$, that both are required open to produce
2781: a fluorescence signal. From the scaling plot, we calibrate the zipping
2782: rate as $k=7.1\times 10^4/$s, in good agreement with the findings from
2783: reference \cite{altan}. The calculated behaviour reproduces the data within
2784: the error bars, while the model prediction at $T=35^\circ$C shows more
2785: pronounced deviation. Potential causes are destabilising effects of the
2786: fluorophore and quencher, and additional modes that broaden the decay
2787: of the autocorrelation. The latter is underlined by the fact that for
2788: lower temperatures the relaxation time distribution $f(\tau)$, defined
2789: by $A(x_T,t)=\int\exp(-t/\tau)f(\tau)d\tau$, becomes narrower
2790: (figure \ref{autocorr} inset).
2791: Deviations may also be associated with the correction for diffusional motion
2792: of the DNA construct, measured without quencher and neglecting
2793: contributions from internal dynamics \cite{oleg1}. Indeed, the black
2794: curve shown in figure \ref{autocorr} was obtained by a 3\% reduction of the
2795: diffusion time;\footnote{For diffusion time $\tau_D=150\mu$s measured for an RNA
2796: construct of comparable length in \cite{oleg1}.} see details in
2797: \cite{tobias_long}.
2798: 
2799: A remark on a prominent alternative approach to DNA breathing appears in order. 
2800: This is the Peyrard Bishop Dauxois (PBD) model \cite{peyrard,dauxois} based
2801: on the set of Langevin equations \cite{peyrard_cm}
2802: \begin{eqnarray}
2803: \nonumber
2804: m\frac{d^2y_n}{dt^2}&=&-\frac{dV(y_n)}{dy_n}-\frac{dW(y_{n+1},y_n)}{dy_n}-
2805: \frac{dW(y_n,y_{n-1})}{dy_n}\\
2806: &&-m\gamma\frac{dy_n}{dt}=\xi_n(t).
2807: \label{peyrard}
2808: \end{eqnarray}
2809: Here, $V(y_n)=D_n\left[\exp(-a_ny_n)-1\right]^2$ is a Morse potential
2810: for the hydrogen bonding, $D_n$ and $a_n$ assuming two different values
2811: for AT and GC bps; $W(y,y')=\frac{k}{2}\left[1+\rho\exp\{-\beta(y+y')\}
2812: \right](y-y')^2$ is a nonlinear potential to include bp-bp stacking
2813: interactions between adjacent bps $y$ and $y'$. The parameters $k$, $\rho$,
2814: $\beta$, $\gamma$, and $m$ are invariant of the sequence. Usually, the 
2815: stochastic equations (\ref{peyrard}) is integrated numerically
2816: \cite{peyrard_cm}.
2817: Due to its formulation in terms of a set of Langevin equations, the DPB model
2818: is very appealing, and it is a useful model to study some generic features of
2819: DNA denaturation. The disadvantage of the current formulation of the DBP model
2820: is the fact that it does not include enough parameters to account for the
2821: known independent stability constants of double stranded DNA (in fact, only
2822: two parameters are allowed to vary with the sequence, in contrast to the 12
2823: independent parameters needed to fully describe the bp stacking and hydrogen
2824: bonding \cite{krueger}).
2825: Moreover, there appear to be certain ambiguities in the proper formulation
2826: of boundary conditions in the stochastic integration \cite{the_prl_footnote},
2827: and also with respect to the interpretation of the biological relevance and
2828: computational limitations of the PBD model \cite{peyrard_comment}.
2829: The master equation and Gillespie approach brought forth in references
2830: \cite{tobias_jpc,tobias_pre,tobias_prl,tobias_long,tobias_leif} bridges the
2831: gap between the thermodynamic data for
2832: the bp stacking and hydrogen bonding obtained by various experimental methods,
2833: and the dynamical nature of DNA breathing. It is hoped that both dynamic
2834: models will synergetically be developed further and eventually lead to a
2835: better understanding of DNA denaturation fluctuations.
2836: 
2837: 
2838: \subsection{Stochastic modelling---the Gillespie algorithm}
2839: 
2840: \begin{figure*}
2841: \includegraphics[width=14cm]{ts_bubble.eps}
2842: \caption{Time series of single bubble-breathing dynamics for
2843: $\sigma_0=10^{-3}$, $M=20$, and (a) $u=0.6$ and (b) $u=0.9$. The lower
2844: panel shows a zoom-in of how single bubbles of size $m(t)$ open up and close.}
2845: \label{suman}
2846: \end{figure*}
2847: 
2848: Despite its mathematically simple form, the master equation (\ref{master})
2849: needs to be solved numerically by inverting the transfer matrix
2850: \cite{tobias_jpc,tobias_long}. Moreover, it produces
2851: ensemble-averaged information. Given the access to single molecule data,
2852: it is of relevance to obtain a model for the fully stochastic time evolution
2853: of a single DNA-bubble, providing a description for pre-noise-averaged
2854: quantities such as the step-wise (un)zipping. With this scope, we
2855: introduced a stochastic simulation scheme for the (un)zipping dynamics,
2856: using the Gillespie algorithm to update the state of the system by determining
2857: (i) the random time between individual (un)zipping events, and (ii) which
2858: reaction direction (zipping, $\leftarrow$, or unzipping, $\rightarrow$) will
2859: occur \cite{suman}. This scheme is efficient computationally, easy to
2860: implement, and amenable to generalisation.
2861: 
2862: To define the model, we denote a bubble state of $m$ broken bps by the
2863: occupation numbers $b_m=1$ and $b_{m'}=0$ ($m'\neq m$). The stochastic
2864: simulation then corresponds to the nearest-neighbour jump process
2865: \begin{equation}
2866: \label{eq1}
2867: b_0\rightleftarrows b_1\rightleftarrows\ldots
2868: \rightleftarrows b_m\rightleftarrows\ldots
2869: \rightleftarrows b_{M-1}\rightleftarrows b_M,
2870: \end{equation}
2871: with reflecting boundary conditions at $b_0$ and $b_M$. Each jump away
2872: from state $b_m$ occurs after a random time $\tau$, and in random direction
2873: to either $b_{m-1}$ or $b_{m+1}$, governed by the reaction probability
2874: density function\footnote{The original expression for the reaction probability
2875: density function,
2876: $P(\tau,\mu)=b_m\tr^{\mu}(m)$ $\exp\left(-\tau\sum_{m,\mu}b_m\tr^{\mu}(m)
2877: \right)$, that is relevant for consideration of multi-bubble states,
2878: simplifies here due to the particular choice of the
2879: $b_m$.} \cite{gillespie,gillespie1}
2880: \begin{equation}
2881: \label{rpdf}
2882: P(\tau,\mu)=\tr^{\mu}(m)e^{-\left(\tr^+(m)+\tr^-(m)\right)\tau},
2883: \end{equation}
2884: where $\mu\in\{+,-\}$ denotes the unzipping ($+$) or zipping ($-$) of a
2885: bp, and the
2886: jump rates $\tr^{\pm}(m)$ are defined below. From the joint probability
2887: density function (\ref{rpdf}),
2888: the waiting time probability
2889: density function that a jump away from $b_m$ occurs is given by $\psi(\tau)
2890: =\sum_{\mu}P(\tau,\mu)$, i.e., it is Poissonian. The probability that the
2891: bubble size does not change in the time interval $[0,t]$ is given by
2892: $\phi(t)=1-\int_0^t\psi(\tau)d\tau$. The fork position $x_L$ (and thereby
2893: the sequence of bps) is straightforwardly incorporated
2894: \cite{tobias_prl,tobias_long}.
2895: 
2896: We start the simulations from the completely zipped state, $b_0=1$ at
2897: $t=0$, and measure the bubble size at time $t$ in terms of
2898: $m(t)=\sum_{m=0}^Mmb_m(t)$. The time series of $m(t)$ for a single stochastic
2899: realization is shown in figure \ref{suman}. It is distinct that the bubble
2900: events are very sharp (note the time windows of the zoom-ins), and
2901: most of the time the zero-bubble state $b_0$ prevails due to $\sigma_0\ll 1$.
2902: Moreover, raising the temperature increases the bubble size and lifetime, as
2903: it should. By construction of the simulation procedure, it is guaranteed that an
2904: occupation number $b_m=1$ ($m\neq 0$) corresponds to exactly one bubble.
2905: 
2906: In a careful analysis, it was shown that the stochastic simulation method
2907: provides accurate information of the statistical quantities of the bubble,
2908: such as opening probability and autocorrelation function \cite{suman}. It
2909: can therefore be used to obtain the same information as the master equation
2910: with the advantage of also giving access to the noise in the system.
2911: With the Gillespie technique, we also obtained the data points in the
2912: graphs in this section.
2913: 
2914: 
2915: \subsection{Bacteriophage T7 promoter sequence analysis.}
2916: 
2917: An example from the analysis of the promoter sequence
2918: \begin{equation}
2919: \begin{array}{l}
2920: \mbox{\small\texttt{\textcolor{white}{AAAA}1\textcolor{white}{%
2921: AAAAAAAAAAAAAAAAAA}20}}\\[-0.1cm]
2922: \mbox{\small\texttt{\textcolor{white}{AAAA}|\textcolor{white}{%
2923: AAAAAAAAAAAAAAAAAA}|\textcolor{white}{AAAAAAAAA}\textcolor{white}{AAA}}}\\[-0.1cm]
2924: \mbox{\small\texttt{5'-aTGACCAGTTGAAGGACTGGAAGTAATACGACTC}}\\
2925: \mbox{\small\texttt{\textcolor{white}{AAA}AG}\textcolor{red}{
2926: \texttt{TATA}}\texttt{GGGACAATGCTTAAGGTCGCTCTCTAGGAg-3'}}\\[-0.1cm]
2927: \mbox{\small\texttt{\textcolor{white}{AAAAA}|\textcolor{white}{AAA}|
2928: \textcolor{white}{AAAAAAAAAAAAAAAAAAAAAAAAAA}|\textcolor{white}{AAA}}}\\[-0.1cm]
2929: \mbox{\small\texttt{\textcolor{white}{AAAAA}\textcolor{red}{38}%
2930: \textcolor{white}{AA}\textcolor{blue}{41}\textcolor{white}{%
2931: AAAAAAAAAAAAAAAAAAAAAAAAAA}68\textcolor{white}{AAA}}}
2932: \end{array}
2933: \label{promoter}
2934: \end{equation}
2935: of bacteriophage T7 is shown in figure \ref{signal} \cite{tobias_prl}. It
2936: depicts the time series of $I(t)$ for the tag positions
2937: $x_T=38$ at the beginning of TATA, and $x_T=41$ at the first GC bp after
2938: TATA. It is distinct how frequent bubble events are in TATA in comparison
2939: to the vicinal GC-rich domain (note that AT/TA bps are particularly weak
2940: \cite{krueger}). This is quantified by the waiting time density $\psi(\tau)$,
2941: whose characteristic time scale is more than an order of magnitude longer
2942: for the $x_T=41$ position. In contrast, we observe almost identical
2943: behaviour for the bubble survival density $\phi(\tau)$. Due to the proximity
2944: of $x_T=41$ to TATA, the typical bubble sizes for both tag positions are
2945: similar, and therefore the relaxation time. However, as shown in
2946: figure \ref{signal} bottom, the variation of the mean lifetime obtained from
2947: the master equation is quite small (within a factor 2) for the entire sequence.
2948: The latter graph also shows the insignificant variation according to the
2949: earlier stability parameters by Blake et al.~\cite{blake}.
2950: 
2951: The results summarised in figure \ref{signal} and further studies in
2952: \cite{tobias_prl,tobias_long} may indicate that it is not solely the
2953: increased opening probability at the TATA motif, as studied in \cite{choi}.
2954: Given the rather short bubble opening times of order of a few $k^{-1}$,
2955: it might be sufficient to induce binding of transcription enzymes (or
2956: other single stranded DNA binding proteins) if only bubble events are
2957: repeated often enough. In the present example, the waiting time between
2958: individual bubble events is increased by a factor of 25 inside the TATA
2959: motif. Guided by such results, detailed future studies combining optical
2960: tweezers overstretching and monitoring transcription initiation may be a
2961: step toward better understanding of this important biochemical process.
2962: 
2963: We note that the influence of noise (e.g., due to repetition of single
2964: molecule experiments) on the bubble dynamics can also be studied in the
2965: weak noise limit by a WKB method \cite{hans}. This model provides 
2966: information, for instance, about the time it takes a DNA to denature under
2967: temperatures above $T_m$ (mathematically corresponding to a finite time
2968: singularity). Bubble breathing can be mapped on the Coulomb problem of
2969: the Schr{\"o}dinger equation, and the corresponding phase transition
2970: studied \cite{hans_short}.
2971: 
2972: \begin{figure}
2973: \begin{center}
2974: \includegraphics[width=8.8cm]{signal.eps}
2975: \end{center}
2976: \caption{Time series $I(t)$ for the T7 promoter, with
2977: $x_T=38$, 41. Middle: Waiting time ($\psi(\tau)$) and bubble
2978: survival time ($\varphi(\tau)$) densities. Bottom: Mean bubble
2979: survival time, $\Delta=2$.}
2980: \label{signal}
2981: \end{figure}
2982: 
2983: 
2984: \subsection{Interaction of DNA bubbles with selectively single-strand
2985: binding proteins}
2986: 
2987: Let us now come back to the destabilising effect of single-stranded DNA
2988: binding proteins (SSBs) mentioned in section \ref{breathintro}. In a
2989: homopolymer approach, this was studied in a master equation approach
2990: in references \cite{tobias_pre,tobias_jpc}. The quantity of interest is
2991: the joint probability $P(m,n,t)$ to have a bubble consisting of $m$ broken
2992: bps, and $n$ SSBs bound to the two arches of the bubble. In addition to the
2993: rates $\mathsf{t}^{\pm}$ for bubble increase and decrease, the rates
2994: $\mathsf{r}^{\pm}$ for SSB binding and unbinding are necessary to define
2995: the breathing dynamics in the presence of SSBs. On the statistical level,
2996: the effect of the SSBs becomes coupled to the motion of the zipper forks.
2997: Thus, the rate for bubble size decrease is proportional to the probability
2998: that no SSB is located right next to the corresponding zipper fork; and
2999: the rate for SSB binding is proportional to the probability that there is
3000: sufficient unoccupied space on the bubble. Binding is allowed to be
3001: asymmetric, and is related to a parking lot problem in the following sense.
3002: The number $\lambda$ of bps occupied by a bound SSB is usually (considerably)
3003: larger than one. In order to be able to bind in between two already bound SSBs,
3004: the distance between these two SSBs must be larger than $\lambda$. The larger
3005: $\lambda$ the less efficient the SSB-binding becomes, similar to parking
3006: large cars on a parking lot desgined for small vehicles. Apart from the
3007: binding size $\lambda$ of the SSBs, two additional physical parameters
3008: come into play: the unbinding rate $q$ of the SSBs, and their binding
3009: strength $\kappa=c_0K^{\mathrm{eq}}$ consisting of the volume concentration
3010: $c_0$ of SSBs and the equilibrium binding constant $K^{\mathrm{eq}}=v_0\exp
3011: \left(\beta|E_{\mathrm{SSB}}|\right)$, with the typical SSB volume and binding
3012: energy $E_{\mathrm{SSB}}$.
3013: 
3014: \begin{figure}
3015: \includegraphics[width=8cm]{F_m_n_kappa_0p5.eps}
3016: \includegraphics[width=8cm]{F_m_n_kappa_1p5.eps}
3017: \caption{Effective free energy in the limit $\gamma\gg 1$ (---),
3018: and `free energy' for various
3019: fixed $n$ ($u=0.6$, $M=40$, $c=1.76$, $\lambda=5$). Top: $\kappa=0.5$; bottom:
3020: stronger binding, $\kappa=1.5$.
3021: \label{ssb_breath}}
3022: \end{figure}
3023: 
3024: The coupled dynamics of SSB-binding and bubble breathing is discussed in
3025: references \cite{tobias_pre,tobias_jpc}; similar effects in end-denaturing
3026: DNA was studied in \cite{tobias_jpca} in detail. Here, we report
3027: the behaviour of the effective free energy landscape in the limit of fast
3028: SSB-binding in the sense that the dimensionless parameter $\gamma\equiv q/
3029: k$ of SSB-unbinding and bubble zipping rates is large, $\gamma\gg 1$. This
3030: limit allows one to average out the SSB-dynamics and to calculate an effective
3031: free energy, in which the bubble dynamics with the slow variable $m$ runs
3032: off. The result for two different binding strengths $\kappa$ is shown in
3033: figure \ref{ssb_breath}, along with the free energies corresponding to keeping
3034: $n$ fixed. It is distinct that while for lower $\kappa$ the presence of SSBs
3035: diminishes the slope of the effective free energy, for larger $\kappa$ the
3036: slope actually becomes negative. In the first case, that is, the bubble
3037: opening is more likely, but still globally unfavourable. In the latter
3038: case, the presence of SSBs indeed leads to full denaturation. One observes
3039: distinct finite size effects due to $\lambda>1$: only when the bubble reaches
3040: a minimal size $m\ge\lambda$, SSB-binding may occur, a second SSB is allowed
3041: to bind to the same arch only once $m\ge2\lambda$, etc. This effect also
3042: produces the nucleation barrier for full denaturation in the lower plot of
3043: figure \ref{ssb_breath}.
3044: Similar finite size effects were investigated for biopolymer translocation
3045: in references \cite{tobias_pb,amlomme}.
3046: We note that the transition to denaturation could also be achieved by
3047: reaching a smaller positive slope of the effective free energy in the
3048: presence of SSBs, and additional titration or change of the effective
3049: temperature through actual temperature change or mechanical stretching
3050: as performed in the experiments reported in references \cite{pant,pant1,pant2}.
3051: 
3052: 
3053: \section{Role of DNA conformations in gene regulation}
3054: \label{generegulation}
3055: 
3056: Our current understanding of gene regulation to large extent is based on the
3057: experiments by Andr{\'e} Lwoff at Institut Pasteur more than 50 years ago
3058: \cite{lwoff}. Lwoff and his collaborators discovered that while a strain of
3059: \emph{E.coli}, a common intestinal bacterium, divided regularly when
3060: undisturbed,
3061: an unexpected phenomenon occurred when the strain was exposed to UV light:
3062: the bacteria stop growing and after some 90 minutes they burst (lyse),
3063: releasing a load of viruses. These viruses then invade new
3064: \emph{E.coli}.\footnote{Often these viruses are called phages or
3065: bacteriophages---bacteria eaters.} Some
3066: of the newly infected bacteria immediately lyse again, while the rest
3067: divides normally---while carrying the virus in them. This dormant state
3068: (lysogeny) of the bacterium can then be driven toward lysis by renewed UV
3069: exposure.
3070: 
3071: \begin{figure}
3072: \includegraphics[width=8.2cm]{ptashne.eps}
3073: \caption{Gene regulation, here the example of the (divergent) bacteriophage
3074: $\lambda$ switch after infection of \ecoli. Figure from \cite{ptashne1},
3075: with permission from M. Ptashne. This figure was modified by the author
3076: from the corresponding figure in M. Ptashne, A Genetic Switch: Phage Lambda
3077: and Higher Organisms, 2nd edition  \copyright Blackwell Science, Malden, MA
3078: and Cell Press, Cambridge, MA, with permission.
3079: \label{ptashne}}
3080: \end{figure}
3081: 
3082: The UV exposure-induced transition from lysogeny to lysis occurs as sketched
3083: in figure \ref{ptashne}. On infection, the bacteriophage $\lambda$ injects its
3084: DNA into \ecoli. In the lysogenic pathway, the viral DNA is integrated into
3085: the host DNA. During lysogeny, repressor dimers bind to certain operator
3086: sites on the $\lambda$ part of the DNA, recruiting RNA polymerase to bind
3087: to the overlapping promoter region(s) and blocking of the vicinal promoter
3088: for the divergent \emph{cro\/} gene. RNA polymerase then transcribes the
3089: \emph{cI} gene to the left of the operator, leading to the expression of
3090: new repressor molecules. UV light, however, leads to cleavage of the
3091: repressor dimers.\footnote{By activation of RecA proteins.} Now the basal
3092: transcription of the gene \emph{cro}, opposite to the \emph{cI\/} gene with
3093: respect to the operator region, leads to the expression of the Cro protein.
3094: Cro bound to the operator then recruits RNA polymerase to the operator,
3095: stabilising the Cro production and blocking \emph{cI}. Simultaneously, a whole
3096: sequence of genes
3097: is being expressed, and the virus reproduces itself inside \ecoli until
3098: lysis occurs. UV light flips the switch from transcription of the
3099: gene \emph{cI\/} maintaining the dormant lysogenic pathway, inducing lysis
3100: that is fostered by transcription of the \emph{cro\/} gene
3101: \cite{ptashne,ptashne1}.
3102: 
3103: The activity of a gene can be monitored even on the single genome level,
3104: by combining the targeted gene \emph{gI} with the gene leading to synthesis
3105: of GFP, the green fluorescent protein, i.e., when \emph{gI} is transcribed,
3106: then so is the gene for GFP. Occurrence of fluorescence then reports
3107: transcription of \emph{gI}. Connected to questions such as the stability of
3108: a genetic pathway is the search process of a specific gene by regulatory
3109: proteins, that is, how dynamically the binding protein actually locates
3110: the operator on the genome. We address these points in what follows.
3111: 
3112: 
3113: \subsection{Physiological background of gene regulation and expression}
3114: 
3115: The $\lambda$ switch from figure \ref{ptashne} is an example of a relatively
3116: simple mechanism. Even simpler is the well-studied Lac repressor. There,
3117: the \emph{lacZ} gene is expressed by recruitment through the CAP protein
3118: when \ecoli is starved of glucose and exposed to lactose. This enables
3119: \ecoli to digest lactose. In absence of lactose, \emph{lacZ} is blocked
3120: by the rep protein. In general, the expression of a certain gene is just
3121: one element in a cascade of simultaneous and/or hierarchical control units,
3122: such as in the developmental regulatory network of the sea urchin embryo
3123: \cite{davidson}. The basic physiological background is common to all of them:
3124: 
3125: Genes are the blueprints of proteins. They control physiological processes
3126: but also developmental pathways: from a fertilised egg cell, eventually
3127: all cell types (skin, hair, liver, brain, etc cells) of a human body emerge,
3128: or a skin cell changes colour on sunlight exposure. A gene is but a stretch
3129: of a DNA molecule, typically comprising some 200-1000 bps. Roughly
3130: speaking, a gene is on when it is being transcribed by RNA polymerase,
3131: otherwise it is off. RNA polymerase binds at the promoter region consisting
3132: of some 60 bps close to the beginning of the gene. It then converts
3133: the A, T, G, C code of the gene into a complementary messenger RNA (from
3134: which, in turn, the protein is produced during translation). The stop of
3135: transcription is triggered by a certain sequence at the end of the gene.
3136: Depending on specific conditions of the recruitment by regulatory proteins,
3137: RNA polymerase binding to the promoter of a certain gene is either blocked
3138: (the gene is off), facilitated (high binding affinity of RNA polymerase due
3139: to the (simultaneous) presence of certain protein(s)), or basal (in absence
3140: of any bound regulatory protein, RNA polymerase can still have a minor
3141: affinity to the promoter and then autonomously start transcription).
3142: The transcription mechanism is part and parcel of the central dogma of
3143: molecular biology summarised in figure \ref{crick}.
3144: 
3145: Molecular switches such as the $\lambda$ switch are surprisingly stable
3146: against noise, despite the fact that there are only about 100 repressor
3147: dimers in the entire bacteria cell \cite{reichardt}. Thus, apart from
3148: external induction, lysis occurs by spontaneous induction due to absence of
3149: CI from the operators \cite{metzi,mewo}.
3150: Such noise-induced errors are estimated to occur once
3151: in $10^7$ cell generations \cite{rozanov,little2}. The stability of the
3152: $\lambda$ switch against noise was analysed in terms of a Wentzell-Freidlin
3153: approach \cite{aurell} and by a simulation analysis \cite{aurell1}. The
3154: latter confirmed that the currently known molecular mechanisms used in
3155: modelling the $\lambda$ switch appear sufficient. While the classical
3156: Shea-Ackers model based on a statistical mechanical approach \cite{shea}
3157: is well established and studied numerically \cite{arkin},
3158: it relies on the knowledge of 13 fundamental Gibbs
3159: free binding energies composed to 40 different binding states of regulatory
3160: proteins and RNA polymerase at the two promoters of $\lambda$. Simulation
3161: of the complete $\lambda$ regulatory system proved the understanding of the
3162: mechanisms of the switch \cite{arkin}. Two more recent studies show that the
3163: $\lambda$ switch remains stable even when each of the fundamental Gibbs free
3164: energies is varied within its (appreciable) experimental error. Moreover,
3165: effects of potential mutations resulting in more significant changes of
3166: the binding energies were studied, and it was shown that certain mutations
3167: can even be compensated by parallel mutations influencing other binding
3168: energies (suppressors) \cite{bakk2,bakk3}. A typical result is shown in
3169: figure \ref{bakk}.
3170: 
3171: \begin{figure}
3172: \includegraphics[width=8.6cm]{Bakk_BJ.ps}
3173: \caption{Activity of the two $\lambda$ promoters as function of repressor
3174: concentration for vanishing Cro concentration. The full line corresponds
3175: to wild type data, whereas the dashed lines correspond to "mutations". The
3176: thin vertical line corresponds to lysogenic CI concentration. See
3177: reference \cite{bakk2} for details.
3178: \label{bakk}}
3179: \end{figure}
3180: 
3181: 
3182: \subsection{Binding proteins: Specific and nonspecific binding modes}
3183: 
3184: Given their very specific function, DNA-binding proteins must recognise
3185: a specific (cognate) sequence of nucleotides along the genome. In fact,
3186: without opening the double helix, the outside of the DNA can be read by
3187: proteins, as the edge of each bp is exposed at the surface. These
3188: patterns are unique only in the major groove of the DNA, this being the
3189: reason why gene regulatory proteins generally bind to the major groove.
3190: Apart from single bp pattern recognition, the protein binding is
3191: sensitive to the special surface features of a certain DNA region. This
3192: local structure of DNA needs to be complementary to the protein structure.
3193: Typical structure patterns (motifs) include helix-turn-helix, zinc 
3194: fingers, leucine zipper, and helix-loop-helix motifs \cite{alberts}.
3195: In bacteria, typical DNA-binding proteins cover some 20 bps or
3196: less. For instance, the lac repressor has a cognate sequence of 21 base
3197: pairs, the CAP protein 16, and the $\lambda$ repressor cI 17 bps
3198: \cite{alberts}. Although the interaction with a single nucleotide within
3199: such a DNA-protein bond is relatively weak, the sum of all matching
3200: nucleotides reaches appreciable values for the overall binding enthalpy,
3201: see below. Moreover, regulatory proteins bound simultaneously can 
3202: significantly enhance the stability of their individual bonds.
3203: 
3204: A simple model for the binding interaction goes back to the work of Berg
3205: and von Hippel \cite{bvh,berg}. Accordingly, the binding free energy is comprised
3206: of two contributions: (i) the (average) non-specific binding free energy
3207: due to electrostatic interaction with the DNA; and (ii) additional binding
3208: free energy if the sequence of the binding site is sufficiently close to the
3209: best (perfectly matching) sequence. The transition to the non-specific
3210: binding is supposed to occur via a conformational change of the regulatory
3211: protein from one that allows more hydrogen bond-formation to another that
3212: permits closer contact between the positive charges of the binding protein
3213: to the negatively charged DNA backbone \cite{wbvh}. This is supported by
3214: more recent structural studies: While in the non-specific binding mode
3215: the Lac repressor is bound to DNA in a rather loose and fuzzy way
3216: \cite{kaladimos}, it appears much more ordered in the specific mode. In
3217: fact, in the latter the protein induces a bend in the DNA \cite{lewis}.
3218: In case (ii), the additional binding free energy can, to good approximation,
3219: be considered independent and additive. reference \cite{gerland} provides a
3220: review of these issues, and derives the following result.
3221: Accordingly, to satisfy both the thermodynamic and kinetic constraints of
3222: the DNA-binding protein interaction, each additional base mismatch in
3223: comparison to the best sequence amounts to the loss of roughly 2 $k_BT$,
3224: and the optimal value for the transition between best specific binding to
3225: the cognate site and non-specific binding is shown to be some 16 $k_BT$
3226: below the energy of the best binding. This value is quite close to the
3227: $\approx 14$ $k_BT$ found for the difference between specific and nonspecific
3228: binding in \cite{bakk,bakk1}.
3229: 
3230: The fact that regulatory proteins bind with varying affinity is an important
3231: ingredient in gene regulation: Not all promoters should have the same activity,
3232: because some proteins are required by the cell at much higher levels than
3233: others. Thus, one given regulatory protein, that controls the recruitment to
3234: the promoters of several genes, can act with different strength depending
3235: on the degree of matching with the local sequence.
3236: 
3237: Non-specific binding can become quite appreciable. It was discovered in
3238: reference \cite{kao} that in the case of the Lac repressor less than 10\% of
3239: the proteins were unbound. In a more recent study using in vivo data of
3240: the $\lambda$ switch, it was found that in a lysogen nearly 90\% of the
3241: repressor protein cI is non-specifically bound. This implies that only
3242: 10-20 free cI dimers exist in the \ecoli cell at any time, pointing at
3243: the important role of non-specific binding in the search process of the
3244: cognate site addressed in the following subsection. Under different 
3245: conditions, both cI and Cro are always non-specifically bound by more than
3246: 50\%. The corresponding non-specific binding energies were estimated as
3247: 7 $k_BT$ \cite{bakk,bakk1}.
3248: 
3249: We note that in contrast to regulatory proteins, restriction enzymes
3250: have an approximate all-or-nothing matching
3251: condition: If a defined sequence matches the restriction enzyme, it will cut,
3252: otherwise not. Even a single mismatch reduces the action of the restriction
3253: enzyme by orders of magnitude. This distinction from regulatory
3254: protein makes sense as restriction enzymes are survival mechanisms and 
3255: should not just cut the cell's own DNA \cite{stormo}. This does not mean
3256: that restriction enzymes do not bind non-specifically---in fact, this is
3257: an important ingredient of their search process in total analogy to 
3258: regulatory proteins. However, their sole active role occurs on complete
3259: matching.
3260: 
3261: 
3262: \subsection{The search process for the specific target sequence}
3263: 
3264: \begin{figure}
3265: \includegraphics[width=8cm]{ownmechanisms.eps}
3266: \caption{Schematic of the search mechanisms in equation (\ref{eq:themodel}).}
3267: \label{fig:ownmechanisms}
3268: \end{figure}
3269: 
3270: To find their specific (cognate) binding site along the genome, DNA-binding
3271: proteins such as restriction enzymes or transcription factors have to search
3272: megabases along the DNA molecule. The high accuracy of gene expression control
3273: by binding proteins such as in the $\lambda$-switch requires
3274: a fast search and recognition of the target sequence by the proteins.
3275: A simple 3-dimensional (3D) search of the target sequence by the proteins
3276: is not sufficient to explain experimentally measured
3277: target search rates. It has been suggested relatively early \cite{adam,eigen}
3278: that additional search mechanisms such as 1D sliding along the genome are
3279: needed to account for the actual efficiency of the search process. In their
3280: pioneering work, Berg, von Hippel and coworkers established a statistical
3281: model for target search comprising the four fundamental steps, as shown in
3282: figure \ref{fig:ownmechanisms}: (i) 3D macrohops during which the 
3283: protein fully detaches from the genome until after a volume excursion it
3284: rebinds to the DNA (as a good approximation, the landing site on the DNA
3285: after a macrohop can be assumed to be equidistributed and uncorrelated);
3286: (ii) microhops
3287: during which the protein detaches from the DNA but always stays very close
3288: to it (i.e., the microhop takes place within a cylinder whose radius
3289: corresponds to the escape distance of the protein from
3290: the DNA, see \cite{berg}); (iii) 1D sliding along the genome (while
3291: preserving a certain bonding to the DNA due to nonspecific binding);
3292: and (iv) intersegmental jumps.
3293: The latter are mediated by DNA-loops bringing two chemically remote segments
3294: of the DNA close in Euclidean space, see, for instance, \cite{hame_looping} and
3295: references therein. A protein like Lac repressor, which can establish bonds to
3296: two different stretches of dsDNA simultaneously, can then jump from one to
3297: the neighbouring segment.\footnote{Possibly, also other binding proteins are
3298: able to perform intersegmental jumps.} This process might lead to a
3299: paradoxical diffusion behaviour \cite{paradox,dirk}. However, if the
3300: conformational changes in the DNA
3301: are not too slow, both the bulk mediated macrohops and the intersegmental
3302: transfer lead to fast mixing of the enzymes' positions along the chain (as
3303: it was shown for the related problem in \cite{JLum}), and on the mean-field
3304: level can be described by a desorption followed by the absorption at a
3305: random place.
3306: 
3307: Recently, there has been renewed interest in the targeting problem, both
3308: theoretically (see, for instance, \cite{gerland,slutsky,marko2,coppey}) and
3309: experimentally (e.g., \cite{grillo,record}), including single molecule studies
3310: \cite{shima,pant1,pant2}. Despite the extensive knowledge of specific binding
3311: rates and both specific and non-specific binding free energies, the precise
3312: relative
3313: contributions of the different search mechanisms (and, to some extent, also
3314: the stringent criteria to define these four elementary interactions) are not
3315: fully resolved. Moreover, it has been suggested that under tight(er) binding
3316: conditions, the sliding of the protein becomes subdiffusive due to the local
3317: structure landscape of a heteropolymer DNA \cite{slutsky1}. This complication,
3318: however, is expected to be relaxed in a more loosely bound search mode of the
3319: searching protein \cite{slutsky}. We here adopt the latter view of normal
3320: diffusion, which is corroborated by the experimental study in the next
3321: subsection.
3322: 
3323: 
3324: \subsection{A unique situation: Pure one-dimensional search of SSB mutants}
3325: 
3326: In previous studies, the 1D sliding problem had always been considered
3327: as a problem of 3D diffusion which is enhanced by 1D diffusion. Thus,
3328: workers such as Berg, Winter, and von Hippel \cite{berg} assumed that proteins
3329: nonspecifically bound would on average unbind before finding their
3330: specific binding sites. This results in an enhancement of specific binding
3331: rates that is proportional to the 1D sliding rate, but the overall specific
3332: binding rate depends linearly on protein concentration. These studies neglect
3333: the possibility that the protein finds its specific site before unbinding.
3334: Given the experimental conditions under which transcription factor binding has
3335: been previously studied, this approximation is appropriate. However, as
3336: demonstrated in \cite{igor_prot}, this mechanism, in which the unbinding rate is
3337: much lower than the specific binding rate, occurs for the 1D search of DNA by
3338: the single-stranded DNA binding protein T4 gene 32 protein (gp32). This fast 1D
3339: search rate is essential for gp32 to be able to quickly find specific locations
3340: on DNA molecules that are undergoing replication, and which have large sections
3341: of single-stranded DNA exposed for gp32 binding. The resulting
3342: nonlinear concentration dependence of gp32 binding will likely
3343: have significant effects on gp32's ability to find its replication sites as
3344: well as its ability to recruit other proteins during replication.
3345: If these nonlinear effects also occur for TFs, this
3346: characteristic will strongly affect regulatory processes governed by protein
3347: binding.
3348: 
3349: \begin{figure}
3350: \begin{center}
3351: \includegraphics[width=6.2cm,angle=270]{figure5.eps}
3352: \end{center}
3353: \caption{Dimensional binding rate $k_a$ in 1/s as function of protein
3354: concentration $C$ in M, for parameters corresponding to 100 mM salt. The
3355: fitted 1D diffusion constant
3356: for sliding along the dsDNA is $D_{\mathrm{1d}}=3.3\cdot 10^{-9}\mathrm{cm}^2/
3357: \mathrm{sec}$, located nicely within the experimental value $10^{-8}\ldots
3358: 10^{-9}\mathrm{cm}^2/\mathrm{sec}$ \protect\cite{pant}.}
3359: \label{rates}
3360: \end{figure}
3361: 
3362: Results from the single DNA overstretching experiment are shown in
3363: figure \ref{rates} along with the results from the theoretical and simulations
3364: analysis from references \cite{igor_prot}. The scaling of search
3365: rate as function of concentration is described by the relation
3366: \begin{equation}
3367: k_a=D_{\mathrm{1d}}n_0^2
3368: \end{equation}
3369: obtained for the pure 1D search of random walkers of line density $n_0^2$
3370: searching along the DNA. For low concentrations, the McGhee and von Hippel
3371: isotherm \cite{mcghee} predicts a linear relation between $n_0$ and the
3372: volume concentration $C$; thus, $k_a\propto C^2$. The experimental evidence
3373: for the purely linear search process, as shown in figure \ref{rates} for 100
3374: mM salt, was found for a large range of salt concentrations, see
3375: references \cite{igor_prot,pant2} for details. The case of high line density
3376: of proteins was discussed in \cite{igor_finite}.
3377: 
3378: 
3379: \subsection{L{\'e}vy flights and target search}
3380: 
3381: We now address the general search process with interchange of 1D and 3D
3382: diffusion, and intersegmental jumps. To this end, we first quickly review
3383: the definition of L{\'e}vy flights \cite{klafternature,physicstoday,report,%
3384: report1,coffey}.
3385: 
3386: L{\'e}vy flights (LFs) are random walks whose
3387: jump lengths $x$ are distributed like $\lambda(x)\simeq|x|^{-1-\alpha}$ with
3388: exponent $0<\alpha<2$ \cite{hughes}. Their probability density
3389: to be at position $x$ at time $t$ has the characteristic function
3390: $P(q,t)\equiv\int_{-\infty}^{\infty}e^{iqx}P(x,t)dx=\exp\left(-D_{\mathrm{L}}
3391: |q|^{\alpha}t\right)$, a consequence of the generalised central limit theorem
3392: \cite{levy,gnedenko}; in that sense, LFs are a natural extension of normal
3393: Gaussian diffusion ($\alpha=2$). LFs occur in a wide range of systems
3394: \cite{report}; in particular, they represent an optimal search
3395: mechanism in contrast to locally oversampling Gaussian search
3396: \cite{stanley}. Dynamically, LFs can be described by a space-fractional
3397: diffusion equation $\partial P/\partial t=D_{\mathrm{L}}\partial^{\alpha}
3398: P(x,t)/\partial|x|^{ \alpha}$, a convenient basis to introduce additional
3399: terms, as shown below. $D_{\mathrm{L}}$ is a diffusion constant of
3400: dimension $\mathrm{cm}^{\alpha}/\mathrm{sec}$, and the fractional derivative is
3401: defined via its Fourier transform, $\mathscr{F}\{\partial^\alpha P(x,t)
3402: /\partial |x|^\alpha\}=-|q|^\alpha P(q,t)$ \cite{report,report1,coffey}. LFs
3403: exhibit
3404: superdiffusion in the sense that $\langle|x|^{\zeta}\rangle^{2/\zeta}\simeq
3405: D_{\mathrm{L}}t^{2/\alpha}$ ($0<\zeta<\alpha$) \cite{report}, spreading
3406: faster than the linearly growing mean squared displacement of standard
3407: diffusion ($\alpha=2$).
3408: A prime example of an LF is linear particle diffusion to next neighbour
3409: sites on a fast folding (`annealed') polymer that permits intersegmental
3410: jumps at chain contact points (see figure \ref{fig:ownmechanisms}) caused
3411: by polymer looping \cite{paradox,dirk}. In fact, the contour length $|x|$
3412: stored in a loop between such contact points is distributed in 3D like
3413: $\lambda(x)\simeq|x|^{-1-\alpha}$, where $\alpha=1/2$ for Gaussian chains
3414: ($\theta$ solvent), and $\alpha\approx 1.2$ for self-avoiding walk chains
3415: (good solvent) \cite{duplantier}.
3416: 
3417: In our description of the target
3418: search process, we use the density per length $n(x,t)$ of proteins on
3419: the DNA as the relevant dynamical quantity ($x$ is the distance along
3420: the DNA contour). Apart from intersegmental transfer, we include 1D
3421: sliding along
3422: the DNA with diffusion constant $D_{\mathrm{B}}$, protein dissociation
3423: with rate $k_{\mathrm{off}}$ and (re)adsorbtion with rate $k_{\mathrm{on}}$
3424: from a bath of proteins of concentration $n_{\mathrm{bulk}}$.
3425: The dynamics of $n(x,t)$ is thus governed by the equation
3426: \cite{michael}
3427: \begin{eqnarray}
3428: \nonumber
3429: \frac{\partial}{\partial t}n(x,t)=&&\left(D_{\rm B} \frac{\partial^2}{\partial
3430: x^2}+D_{\rm L} \frac{\partial^\alpha}{\partial |x|^\alpha}-k_{\rm
3431: off}\right)n(x,t)\nonumber\\
3432: &&+k_{\rm on}n_{\rm bulk}-j(t)\delta(x).
3433: \label{eq:themodel}
3434: \end{eqnarray}
3435: Here, $j(t)$ is the flux into the target located at $x=0$. We determine
3436: the flux $j(t)$ by assuming that the target is perfectly absorbing: $n(
3437: 0,t)=0$ \cite{chechkin_bvp}. Be initially the system at equilibrium, except
3438: that the target is
3439: unoccupied; then, the initial protein density is $n_0=n(x,0)=k_{\rm on}
3440: n_{\rm bulk}/k_{\rm off}$.\footnote{Note that the dimension of the on and
3441: off rates differ; while
3442: $[k_{\mathrm{off}}]=\mathrm{sec}^{-1}$, we chose $[k_{\mathrm{on}}]=
3443: \mathrm{cm}^2/\mathrm{sec}$.} The total number
3444: of particles that have arrived at the target up to time $t$ is
3445: $J(t)=\int_0^t d t'\;j(t')$. We derive explicit analytic
3446: expressions for $J(t)$ in different limiting regimes, and study
3447: the general case numerically. We use $J(t)$ to obtain
3448: the mean first arrival time $T$ to the target; in
3449: particular, to find the value of $k_{\rm off}$ that minimises $T$.
3450: 
3451: \begin{figure}
3452: \includegraphics[width=6.8cm]{k_opt.epsi}
3453: \caption{Optimal choice of off rate $k_{\rm off}$ as function of the
3454: LF diffusion constant, from numerical evaluation of the model in reference
3455: \cite{michael}. The circle on the abscissa marks where $k_{\rm off}^{\rm opt}$
3456: becomes 0 in the case $\alpha<1/2$.}
3457: \label{searcheff}
3458: \end{figure}
3459: 
3460: The various regimes of target search embodied in equation (\ref{eq:themodel})
3461: are discussed in detail in references \cite{michael,michael_long}. The main
3462: result for the efficiency of the related search process is summarised in
3463: figure \ref{searcheff}, i.e., which protein unbinding rate $k_{\mathrm{off}}$
3464: optimises the mean search time $T$. Three regimes can be distinguished:
3465: 
3466: (i) Without L{\'e}vy flights, we obtain $k_{\rm off}^{\rm opt}=k_{\rm on}'$:
3467: the proteins should spend equal amounts of time in bulk and on the DNA.
3468: This corresponds to the result obtained for single protein searching on
3469: a long DNA \cite{slutsky,coppey}. 
3470: 
3471: (ii) For $\alpha>1$, i.e., when DNA is in the self-avoiding regime,
3472: we find
3473: \begin{equation}
3474: k_{\rm off}^{\rm opt}\sim (\alpha-1)k_{\rm on}':
3475: \end{equation}
3476: The optimal off rate shrinks linearly with decreasing $\alpha$.
3477: 
3478: (iii) For $\alpha<1$, i.e., when DNA leaves the self-avoiding phase
3479: (e.g., by lowering the temperature or introducing attractive interactions)
3480: the value of $k_{\rm off}^{\rm opt}$ approaches zero as the frequency of
3481: intersegmental jumps ($\propto D_L$) increases: The L{\'e}vy flight
3482: mechanism becomes so efficient that bulk excursions become irrelevant.
3483: At $\alpha=1/2$, the case of the ideal Gaussian chain, we observe a
3484: qualitative change:
3485: When $\alpha<1/2$, the rate $k_{\rm off}^{\rm opt}$ reaches
3486: zero for \emph{finite} values of the rate for intersegmental jumps.
3487: Note that when $\alpha<1$, the spread of the L{\'e}vy flight ($\simeq
3488: t^{1/\alpha}$) grows
3489: faster than the number of sites visited ($\simeq t$), rendering the mixing
3490: effect of bulk excursions insignificant. A scaling argument to understand
3491: the crossover at $\alpha=1/2$ relates the probability density of first
3492: arrival with the width ($\simeq t^{1/\alpha}$) of the Green's function of a
3493: L{\'e}vy flight
3494: $p_{\mathrm{fa}}\simeq t^{-1/\alpha}$. We see that the associated mean
3495: arrival time becomes finite for $0<\alpha<1/2$, even for the infinite chain
3496: limit considered here.
3497: 
3498: We remark that this model is valid for an annealed DNA only. This means
3499: that the chain can equilibrate (at least, locally) on the typical time
3500: scale between intersegmental jumps. Even though real DNA in solution
3501: might not be fully annealed, features of this analysis will reflect on
3502: the target search. A more detailed study of different regimes of DNA
3503: is under way.
3504: 
3505: 
3506: \subsection{Viruses---extreme nanomechanics}
3507: 
3508: Viruses have played an important role in the discovery of the mechanisms
3509: underlying gene regulation, see, for instance, reference \cite{lwoff}. From
3510: a nanoscience perspective, viruses are of interest on their own part.
3511: During the assembly of many viruses, the viral DNA of several $\mu$m length
3512: is packaged into the capsid, the protein container making up most of the virus, 
3513: by a motor protein. This motor packages the DNA by exerting forces of up to
3514: 60 pN or more, causing pressures building up in the capsid of the order
3515: of 6 MPa \cite{smith,zandi}. The size of the capsid spans few tens
3516: of $\mu$m, and is therefore comparable to the persistence length of DNA
3517: \cite{feiss,morita,smith,hud1,hud}. Therefore, fluctuation-based undulations are
3518: suppressed, and the chain can be approximately thought of as being wound
3519: up helically like thread on a bobbin, or like a ball of
3520: yarn. Ultimately, a relatively highly ordered 3D
3521: configuration of the DNA inside the capsid is achieved, which under certain
3522: conditions may even lead to local crystallisation of the DNA
3523: \cite{feiss,morita,pack,hud,kindt,kondev1,ali}.
3524: It is generally argued that this
3525: ordered
3526: arrangement helps to avoid the creation of entanglements or even knots of the
3527: wound-up DNA, thus enabling easy ejection, i.e., release of the DNA once
3528: the phage docks to a new host cell; this ejection is not assisted by the
3529: packaging motor, but it can be facilitated by host cellular DNA polymerase,
3530: which starts to transcribe the DNA and thereby pulls it out of the capsid
3531: \cite{alberts,snustad,kindt,arsuaga}.
3532: Details on the packaging energetics can be found in reference \cite{pack}, and
3533: the works cited therein. Model calculations for the
3534: entropy loss, binding and twist energy, and electrostatic forces that need
3535: to be overcome on packaging reveal, that at higher packaging ratios the
3536: packaging force almost exclusively comes from the electrostatic repulsion.
3537: 
3538: 
3539: \section{Functional molecules and nanosensing}
3540: 
3541: Complex molecules can be endowed with the distinct feature
3542: that they contain subunits which are linked to each other
3543: mechanically rather than chemically \cite{schill}. The investigation of
3544: the structure and properties of such interlocked {\em topological
3545: molecules\/} is
3546: subject of the growing field of chemical topology \cite{frisch};
3547: while speculations about the possibility of catenanes
3548: \footnote{{\em catena\/} (lat.), the chain.} (Olympic rings) date
3549: back to the early 20th century lectures of Willst{\"a}tter, the
3550: actual synthesis of catenanes and rotaxanes
3551: \footnote{{\em rota\/} (lat.), the wheel; {\em axis\/} (lat.),
3552: the axle.} succeeded in 1958 \cite{schill}. Modern organic chemistry
3553: has seen the development of refined synthesis methods to generate topological
3554: molecules.
3555: 
3556: 
3557: \subsection{Functional molecules}
3558: 
3559: In parallel to the miniaturisation in electronics \cite{bishop} and the
3560: possibility of manipulating single (bio)molecules \cite{strick1},
3561: supramolecular chemistry which makes use of chemical topology properties is
3562: coming of age \cite{lehn,dietrich}. Thus, rotaxane-type molecules are believed
3563: to be the building blocks for certain nanoscale machines and motors
3564: \cite{blanco}, so-called hermaphrodite molecules have been shown to perform
3565: linear relative motion (``contraction and stretching'') \cite{jimenez},
3566: and pirouetting molecules have been synthesised \cite{raehm}.
3567: Moreover, topological molecules are thought to become
3568: components for molecular electronics switching devices in memory and
3569: computing applications \cite{pease,lehn1}.
3570: These molecular machines are usually of lower molecular weight, and their
3571: behaviour is essentially energy-dominated in the sense that their
3572: conformations and dynamical properties are governed by external and thermal
3573: activation in an energy landscape. The understanding of the physical
3574: properties and the theoretical modelling of such designer molecules and their
3575: natural biological counterparts has increasingly gained momentum, and the
3576: stage is already set for the next generation of applications
3577: \cite{bishop,strick1,lehn,dietrich,blanco,jimenez,raehm,pease,lehn1,%
3578: montemagno,soong,yurke,metz,motors,frey}.
3579: 
3580: In reference \cite{hame_cpl} we introduced some basic
3581: concepts for functional molecules whose driving force is entropic rather
3582: than energetic, see also the more recent publications in chemistry journals
3583: \cite{functional,functional1}.
3584: Entropy-functional molecules will be of higher molecular weight
3585: (hundred monomers or above)
3586: in order to provide sufficient degrees of freedom such that entropic
3587: effects can determine the behaviour of the molecule.
3588: The potential for such {\em entropy-driven functional molecules\/}
3589: can be anticipated from the classical Gibbs Free energy
3590: \begin{equation} \label{free}
3591: \mathscr{F} = U - T S \, ;
3592: \end{equation}
3593: in functional molecules, $\mathscr{F}$ is minimised mainly by variation of the
3594: internal energy $U$ representing the shape of the energy landscape of
3595: the functional unit. New types of molecules were proposed for which
3596: $\mathscr{F}$ is minimised by variations of the entropy $S$, while the energies
3597: and chemical bondings are left unchanged \cite{hame_cpl}. The entropy-functional
3598: units of such molecules can be specifically controlled by external parameters
3599: like temperature, light flashes, or other electromagnetic fields
3600: \cite{lehn,dietrich}. We note that DNA is already being studied as
3601: a macromolecular prototype
3602: building block for molecular machines \cite{yurke}.
3603: 
3604: \begin{figure}
3605: \includegraphics[width=7.8cm]{fig4.eps}
3606: \caption{Molecular muscle consisting of two interlocked rings
3607: 1 and 2 with attached rod-like molecules. Within this structure, sliding
3608: rings, 3, can be placed, which, if activated, tend to contract the muscle
3609: by entropic forces.}
3610: \label{muscle}
3611: \end{figure}
3612: 
3613: A typical example is the molecule shown in figure \ref{muscle}.
3614: According to the arrangement of the sliding rings 1 and 2, this compound
3615: exhibits the unique feature of a molecule that it can slide {\em laterally\/}.
3616: Suggested as precursors of molecular muscles \cite{jimenez}, this compound
3617: could be propelled with internal entropy-motors, which entropically
3618: adjust the elongation of the muscle. In the configuration shown in
3619: figure \ref{muscle}, the sliding ring 3 creates, if activated, an
3620: entropic force which tends to contract the ``muscle'';
3621: at $T = 300 \, {\mbox K}$ and on a typical scale $x = 10 \, \mbox{nm}$,
3622: the entropic force $k_B T / x$
3623: is of the order of pN, and thus comparable to the force created
3624: in biological muscle cells \cite{gittes}. Molecular muscles of such a make
3625: can be viewed as the nano-counterpart of macroscopic muscle models proposed
3626: by de Gennes \cite{degennesm}, in which the contraction is based on
3627: the entropy difference between the isotropic and nematic phases in liquid
3628: crystalline elastomer films \cite{thomsen}.
3629: 
3630: Similarly, one might speculate
3631: whether the DNA helix-coil transition \cite{poland}
3632: in multiplication setups could be
3633: facilitated in the presence of pre-ring molecules which in vitro attach
3634: to an opened loop of the double strand and close, creating an entropy
3635: pressure which tends to open up the vicinal parts of the DNA which are
3636: still in the helix state. Finally, considering molecular motors,
3637: it would be interesting to design an externally controllable,
3638: purely entropy-driven rotating nanomotor.
3639: 
3640: Numerous additional nanoapplications of biopolymers appear in current
3641: literature. An interesting example is the nanomotor created by a DNA
3642: ring in a periodically driven external field, for instance, a focused
3643: light beam inducing localised temperature variations \cite{kulic_schiessel,%
3644: kulic_schiessel_jpc}. The speeds possibly attained by such a device are
3645: of the order of those reached by biological organisms. Such a nanorotor
3646: could be used to stir smallest volumes in higher viscous environments.
3647: 
3648: 
3649: \subsection{Nanosensing}
3650: 
3651: The advances in minituarization of reactors and devices also brings along
3652: the need of probes, by which smallest volumes can be tested. For instance,
3653: microarrays used in genomics require sensors to detect the presence of
3654: certain proteins (often at small concentrations) in a microdish, without
3655: disturbing the environment in the small volume too much. Similarly, single
3656: molecule experiments require specific local detection possibilities.
3657: 
3658: \begin{figure}
3659: \includegraphics[width=6.8cm]{colourbeacon.eps}
3660: \caption{Molecular beacon based on local DNA denaturation. The green blobs
3661: may represent single-stranded DNA binding proteins, or more specifically
3662: binding proteins binding or other molecule to a custom designed DNA sequence
3663: along the denaturation fork. Bound proteins stabilize the denatured fork
3664: and change the spectrum of the beacon.}
3665: \label{beacon}
3666: \end{figure}
3667: 
3668: A fine example for a potential nanosensore is
3669: the blinking behaviour of a fluorophore-quencher pair mounted on
3670: the denaturation wedge as shown in figure \ref{beacon}. This setup, similar
3671: to the ones described in references \cite{altan,oleg1} works as follows.
3672: As long as the dsDNA is intact, fluorophore and quencher are in close
3673: proximity. Once they come apart from one another when the denaturation
3674: wedge opens up, the incident laser light causes fluorescence of the dye.
3675: The on/off blinking of this "molecular beacon" can be monitored in the
3676: focus of a confocal microscope, or, depending on the intensity of the
3677: emitted light, by a digital camera.
3678: The blinking renders immediate information about the state of the bp,
3679: that is tagged by the dye-quencher pair. Fluorescence, that is, indicates that
3680: the bp is currently broken. It is therefore advantageous to define
3681: the random variable $I(t)$ with the property
3682: \begin{equation}
3683: I(t)=\left\{\begin{array}{ll}
3684: 0 & \mbox{if base-pair at $x=x_T$ is closed}\\
3685: 1 & \mbox{if base-pair at $x=x_T$ is open}\end{array}\right.,
3686: \end{equation}
3687: and in experiments one typically measures the corresponding blinking
3688: autocorrelation function
3689: \begin{equation}
3690: A(t)=\langle I(t)I(0)\rangle-\langle I\rangle_{\mathrm{eq}^2},
3691: \label{A_t}
3692: \end{equation}
3693: where $\langle I\rangle_{\mathrm{eq}}$ is the (ensemble) equilibrium value,
3694: or its spectral decomposition
3695: \begin{equation}
3696: A(t)=\int_0^{\infty}f(\tau)\exp\left(-\frac{t}{\tau}\right)d\tau,
3697: \end{equation}
3698: where
3699: \begin{equation}
3700: f(\tau)=\sum_{p\neq 0}T_p^2\delta\left(\tau-\tau_p\right).
3701: \end{equation}
3702: is called the relaxation time spectrum.
3703: 
3704: Figure \ref{blinkbeac} shows an example for the achievable sensitivity of
3705: such nanobeacons, in an example where the denaturation wedge is in
3706: solution together with a certain concentration (proportional to $\kappa$,
3707: compare Sec. IVG) of selectively single-stranded
3708: DNA binding proteins, as discussed previously. It is distinct how both
3709: measurable signals, $A(t)$ and $f(\tau)$ change with varying SSB-concentration.
3710: \begin{figure}
3711: \includegraphics[width=7.8cm]{f_tau_gamma2.eps}
3712: \includegraphics[width=7.8cm]{A_t_gamma2.eps}
3713: \caption{Spectral response of the denaturation beacon in the presence of
3714: single-stranded DNA binding proteins. Top: Relaxation time spectrum, bottom:
3715: blinking autocorrelation function.}
3716: \label{blinkbeac}
3717: \end{figure}
3718: 
3719: 
3720: \section{Summary}
3721: 
3722: Biopolymers such as DNA, RNA, and proteins are indispensable for their
3723: specificity and robustness in all forms of life. Given their detailed
3724: physical properties such as DNA's persistence length of some 50nm or
3725: its local denaturation in nano-bubbles already at room temperature,
3726: and biochemically relevant interfaces such as 10-20 bps, they 
3727: deeply stretch into the nanoscience domain. This statement is twofold
3728: in the following sense. Firstly, nanotechniques such as atomic force
3729: microscopes become important tools to manipulate and probe biomolecules
3730: and their interaction even on the single molecule level. Secondly,
3731: biomolecules are entering the stage as nanotools such as nanosensors,
3732: functional molecules, or highly sensitive force transducers.
3733: 
3734: The possibility to perform controlled experiments on biomolecules, for
3735: instance, to measure the force-extension curves of single biopolymers,
3736: also opens up novel possibilities to test new physical theories. The
3737: foremost examples may be the exploration of persistence lengths and
3738: other polymer physics properties, and the statistical mechanical
3739: concepts relevant for small system sizes. The latter are known under
3740: the keyword of the Jarzynski relation connecting the non-equilibrium
3741: work performed on a physical system with the difference in the thermodynamic
3742: (i.e., equilibrium) potential between initial and final states
3743: \cite{jarzinsky,udo}.
3744: However, there exist by now several similar theories addressing different
3745: physical quantities, such as the concept of entropy production along a single
3746: particle trajectory \cite{udo1}.
3747: 
3748: This review summaries fundamental physical properties of DNA, and their
3749: relevance for both biological processes and technological applications.
3750: The extensive list of references will be useful for further studies on
3751: specific topics covered herein. We are confident that the role of 
3752: biomolecules in technology, not at least for biomedical applications,
3753: will experience a dramatic increase during the coming years and will
3754: enable us to extend current physical understanding of fundamental
3755: processes.
3756: 
3757: 
3758: \acknowledgments
3759: 
3760: RM and TA acknowledge many helpful and enjoyable discussions with
3761: Jozef Adamcik, Audun Bakk,
3762: Suman Banik, Erika Ercolini, Giovanni Dietler, Hans Fogedby, Yacov Kantor,
3763: Mehran Kardar, Joseph Klafter, Oleg Krichevsky, Michael Lomholt, Maxim
3764: Frank-Kamenetskii, Igor Sokolov, Andrzej Stasiak, Francesco Valle, and
3765: Mark Williams.
3766: RM acknowledges partial funding from the Natural
3767: Sciences and Engineering Research Council (NSERC) of Canada and the Canada
3768: Research Chairs programme. TA acknowledges partial funding from the Wallenberg
3769: foundation.
3770: AH acknowledges funding by the AFOSR (FA9550-05-1-0472) and by the NIH
3771: (SCORE program GM068855-03S1).
3772: SDL acknowledges funding by the Joint DMS/NIGMS
3773: Mathematical Biology Initiative (NIH GM 67242)
3774: and by the National Foundation for Cancer Research
3775: through the Yale-NFCR Center for Protein and Nucleic Acid Chemistry.
3776: 
3777: \clearpage
3778: 
3779: \section*{\large Biographical notes}
3780: 
3781: \begin{minipage}[t]{3.6cm}
3782: \unitlength=1cm
3783: \begin{picture}(4,0)
3784: \put(-5.48,-11.72){\special{psfile=metz.eps vscale=65 hscale=65 voffset=0
3785: hoffset=0}}
3786: \end{picture}
3787: \end{minipage}
3788: \begin{minipage}[t]{4.6cm} \noindent
3789: \textbf{Ralf Metzler} received his doctoral degree in physics from the
3790: University of Ulm, FRG. He then went to Tel Aviv University as Humboldt
3791: Feodor Lynen fellow and later as Minerva Amos de Shalit fellow, to work
3792: with Joseph Klafter. As DFG Emmy Noether fellow, after a period as visiting
3793: scientist at the University of Illinois at Urbana-Champaign with Peter Wolynes,
3794: Ralf moved to the Massachusetts Institute of\\\vspace*{-0.26cm}
3795: \end{minipage}
3796: \noindent Technology (MIT) in Cambridge,
3797: MA, where he worked with Mehran Kardar. In 2002 Ralf was appointed Assistant
3798: Professor at
3799: the Nordic Institute for Theoretical Physics (NORDITA) in Copenhagen,
3800: Denmark. In summer 2006, Ralf assumed his post as
3801: Associate Professor and Canada Research Chair in Biological Physics at
3802: the University of Ottawa, Canada.
3803: Ralf works extensively on biological physics problems and anomalous
3804: stochastic processes. These include DNA physics such as DNA denaturation,
3805: DNA topology and the role of DNA conformations in gene regulation, as well
3806: as anomalous diffusion in biological systems. The latter questions are connected
3807: with L{\'e}vy statistics leading to, respectively, subdiffusion or L{\'e}vy
3808: flights. In his spare time, Ralf enjoys the company of his daughter and wife,
3809: he listens to classical music and is a keen reader of murder mysteries. In
3810: Canada, Ralf is looking forward to nature walks, cross-country skiing and
3811: ice-skating.
3812: 
3813: \hspace{0.2cm}
3814: 
3815: \begin{minipage}[t]{3.6cm}
3816: \unitlength=1cm
3817: \begin{picture}(4,0)
3818: \put(-10.82,-19.78){\special{psfile=tobias.eps vscale=118 hscale=118 voffset=0
3819: hoffset=0}}
3820: \end{picture}
3821: \end{minipage}
3822: \begin{minipage}[t]{4.6cm} \noindent
3823: \textbf{Tobias Ambj{\"o}rnsson} received his PhD from Chalmers and
3824: Gothenburg University, where he worked on the electromagnetic response of
3825: matter. In 2003 he became a NORDITA postdoctoral fellow, working with
3826: Ralf Metzler on the modelling of single biomolecule problems such as
3827: DNA breathing and biopolymer translocation through nanopores. Tobias
3828: re-\\\vspace*{-0.26cm}
3829: \end{minipage}
3830: \noindent
3831: recently received a prestigious fellowship from the Wallenberg foundation,
3832: allowing him to join the group of Robert Silbey at MIT in autumn 2006,
3833: to pursue studies on nanosensors. Tobias enjoys listening to pop music,
3834: travelling, and spending time with friends.
3835: 
3836: \hspace{0.2cm}
3837: 
3838: \begin{minipage}[t]{3.6cm}
3839: \unitlength=1cm
3840: \begin{picture}(4,0)
3841: \put(-5.32,-11.12){\special{psfile=andreas1.eps vscale=64 hscale=64 voffset=0
3842: hoffset=0}}
3843: \end{picture}
3844: \end{minipage}
3845: \begin{minipage}[t]{4.52cm} \noindent
3846: \textbf{Andreas Hanke}
3847: received his doctoral degree in physics from the University of
3848: Wuppertal in Germany, FRG. He then went to the Massachusetts Institute of
3849: Technology (MIT) for his postdoctoral studies with Mehran Kardar.
3850: After a period as visiting scientist with Michael Schick at the University of
3851: Washington, Se\-attle, he moved to a postdoc-\\\vspace*{-0.26cm}
3852: \end{minipage}
3853: \noindent
3854: toral position in John Cardy's
3855: group at the University of Oxford, UK, followed by postdocs with Udo Seifert
3856: in Stuttgart, FRG, and again with John Cardy in Oxford. In 2004 Andreas
3857: became Assistant Professor of Physics at the University of Texas at
3858: Brownsville. He is also Adjunct Assistant Professor at the Department of
3859: Physics at The University of Texas at Dallas and a member of the Institute
3860: of Biomedical Sciences and Technology at UT Dallas. In addition, his
3861: research includes summer appointments at the UT Dallas NanoTech Institute.
3862: Andreas has worked in the fields of mesoscopic quantum systems, soft
3863: condensed matter physics, and biological physics. Currently he is building
3864: up a theory division in Molecular Biophysics and Nanoscience at the
3865: University of Texas at Brownsville. In his spare time, Andreas enjoys
3866: latin and salsa music and dancing, all
3867: kinds of outdoor activities, and excursions to Mexico.
3868: 
3869: \hspace{0.2cm}
3870: 
3871: \begin{minipage}[t]{4.26cm}
3872: \unitlength=1cm
3873: \begin{picture}(4,0)
3874: \put(-7.72,-15.28){\special{psfile=Yongli_passport_photo.ps vscale=92
3875: hscale=92 voffset=0
3876: hoffset=0}}
3877: \end{picture}
3878: \end{minipage}
3879: \begin{minipage}[t]{3.82cm} \noindent
3880: \textbf{Yongli Zhang} is a postdoctoral fellow of the Jane Coffin Childs
3881: Memorial Fund for Medical Research in Carlos Bustamante’s group at University
3882: of California at Berkeley. He received his Ph.D. in Molecular Biophysics
3883: and Biochem-\\[-0.26cm]
3884: \end{minipage}
3885: \noindent
3886: Biochemistry from Yale University in 2003, under supervision of Donald
3887: Crothers. In December 2006, he will move to the Department of Physiology
3888: and Biophysics at Albert Einstein College of Medicine as an assistant
3889: professor. Yongli has been doing both experimental and theoretical
3890: research in DNA mechanics, chromatin dynamics, and mechanism of chromatin
3891: remodeling. His lab will mainly use single-molecule techniques, such as
3892: optical tweezers and atomic force microscopy, to study mechanism of
3893: molecular motors and dynamics of macromolecular assemblies.
3894: 
3895: \hspace{0.2cm}
3896: 
3897: \begin{minipage}[t]{4.26cm}
3898: \unitlength=1cm
3899: \begin{picture}(4,0)
3900: \put(-6.72,-14.48){\special{psfile=stephen.eps vscale=83 hscale=83 voffset=0
3901: hoffset=0}}
3902: \end{picture}
3903: \end{minipage}
3904: \begin{minipage}[t]{3.74cm} \noindent
3905: \textbf{Stephen Levene}
3906: is Associate Professor of Molecular and Cell Biology at
3907: The University of Texas at Dallas. Dr. Levene received his Ph.D. in
3908: Chemistry from Yale University in 1985 and was an American Cancer Society
3909: postdoctoral fellow with Bruno Zimm at University of
3910: California,\\\vspace*{-0.26cm}
3911: \end{minipage}
3912: \noindent
3913: San Diego
3914: until 1989. He then spent one year as a staff scientist at the Human Genome
3915: Center at Lawrence Berkeley Laboratory and was a Program in Mathematics and
3916: Molecular Biology Fellow at University of California, Berkeley in Nicholas
3917: Cozzarelli’s laboratory. Dr. Levene's research interests are in the area of
3918: nucleic-acid structure and flexibility, mechanisms of DNA recombination, the
3919: structural organization of human telomeres, and applications of these areas
3920: to biotechnology. He leads the focus group in Molecular Diagnostics and
3921: Bioimaging in the UT-Dallas Institute of Biomedical Sciences and Technology,
3922: is recipient of an Obermann Interdisciplinary Research Fellowship, a member
3923: of the Biophysical Society, and has served on several NIH study sections.
3924: In addition to scientific pursuits, Dr. Levene is an avid snow skier and
3925: cyclist, having previously competed in both disciplines.
3926: 
3927: \clearpage
3928: 
3929: \begin{appendix}
3930: 
3931: \section*{A polymer primer.}
3932: \label{appA}
3933: 
3934: In this section, we introduce some basic concepts from polymer physics.
3935: Starting from the random walk model, we define the fundamental measures
3936: of a polymer chain, before introducing excluded volume. For more details,
3937: we refer to the monographs \cite{degennes,doi,grosberg,flory}.
3938: 
3939: The simplest polymer model is due to Orr \cite{orr}. It models the polymer
3940: chain a a random walk on a periodic lattice with lattice spacing $a$. Then,
3941: each monomer of index $i$ is characterised by a position vector $\mathbf{R}_i$
3942: with $i=0,1,\ldots,N$. The distance between monomers $i$ and $i+1$ is called
3943: $\mathbf{a}_{i+1}=\mathbf{R}_{i+1}-\mathbf{R}_i$. Consequently, the end-to-end
3944: vector of the polymer is
3945: \begin{equation}
3946: \mathbf{r}=\sum_i\mathbf{a}_i.
3947: \end{equation}
3948: Different $\mathbf{a}_i$ have completely independent orientations, such that
3949: we immediately obtain the average ($\langle\cdot\rangle$ over different
3950: configurations) squared end-to-end distance
3951: \begin{equation}
3952: \mathbf{R}_0^2=\langle\mathbf{r}^2\rangle=\sum_{i,j}\langle\mathbf{a}_i\cdot
3953: \mathbf{a}_j\rangle=\sum_i\langle\mathbf{a}_i^2\rangle=Na^2.
3954: \end{equation}
3955: $R_0\simeq N^{1/2}a$ is a measure for the size of the random walk.
3956: An alternative measure of the size of a polymer chain is provided by its radius
3957: of gyration $R_g$, which may be measured by light scattering experiments. It is
3958: defined by
3959: \begin{equation}
3960: \label{appgyr}
3961: R_g^2=\frac{1}{1+N}\sum_{i=0}^N\langle\left(\mathbf{R}_i-\mathbf{R}_G\right)^2
3962: \rangle,
3963: \end{equation}
3964: and measures the average squared distance to the centre of gravity,
3965: \begin{equation}
3966: \mathbf{R}_G=\frac{1}{1+N}\sum_{i=0}^N\mathbf{R}_i.
3967: \end{equation}
3968: Expression (\ref{appgyr}) can be rewritten as
3969: \begin{equation}
3970: R_g^2=(1+N)^{-2}\sum_{i=0}^{N-1}\sum_{j=i+1}
3971: ^N\langle\left(\mathbf{R}_i-\mathbf{R}_j\right)^2\rangle.
3972: \end{equation}
3973: With $\mathbf{R}_j-\mathbf{R}_i=\sum_{n=i+1}^j\mathbf{a}_n$, one can easily
3974: show that $R_g^2=a^2N(N+2)/[6(N+1)]$. For large $N$, that is, $R_g\simeq
3975: \frac{a^2}{6}N$, and therefore:
3976: \begin{equation}
3977: R_g\sim R_0\sim aN^{1/2}.
3978: \end{equation}
3979: 
3980: On a cubic lattice in $d$ dimensions, each step can go in $2d$ directions,
3981: and for a general lattice, each vector $\mathbf{a}_i$ will have $\mu$ possible
3982: directions. The number of distinct walks with $N$ steps is therefore $\mu^N$.
3983: Denote $\mathfrak{N}_N(\mathbf{r})$ the number of distinct walks with
3984: end-to-end vector $\mathbf{r}$, the probability density function for a
3985: given $\mathbf{r}$ is
3986: \begin{equation}
3987: p(\mathbf{r})=\frac{\mathfrak{N}_N(\mathbf{r})}{\sum_{\mathbf{r}}\mathfrak{N}_N(
3988: \mathbf{r})}.
3989: \end{equation}
3990: For large $N$, due to the independence of individual $\mathbf{a}_i$, this
3991: probability density function will acquire a Gaussian shape,
3992: \begin{equation}
3993: p(\mathbf{r})=\left(\frac{d}{2\pi Na^2}\right)^{d/2}\exp\left(-\frac{dr^2}{
3994: 2Na^2}\right),
3995: \end{equation}
3996: where the normalisation is such that $\langle\mathbf{r}^2\rangle=Na^2$.
3997: From this expression, we can deduce that the number of degrees of freedom
3998: of a closed random walk chain is proportional to $N^{-d/2}$, the entropy
3999: loss suffered by a chain subject to the constraint $\mathbf{r}=0$. On a
4000: general lattice,
4001: \begin{equation}
4002: \omega\simeq \mu^NN^{-d/2}
4003: \end{equation}
4004: with the connectivity constant $\mu$, a measure for in how many different
4005: directions the next bond vector can point ($\mu=2d$ in a cubic lattice).
4006: At fixed end-to-end distance, the entropy of the random walk becomes
4007: $S(\mathbf{r})=S_0-dr^2/(2Na^2)$ where $S_0$ absorbs all constants.
4008: For the free energy $\mathscr{F}(\mathbf{r})=E-k_BTS(\mathbf{r})$ we therefore
4009: obtain
4010: \begin{equation}
4011: \mathscr{F}(\mathbf{r})=\mathscr{F}_0+\frac{dk_BTr^2}{2R_0^2},
4012: \end{equation}
4013: i.e., the random walk likes to coil, the restoring force $-\nabla
4014: \mathscr{F}(\mathbf{r})$ being linear in $\mathbf{r}$. This is often called
4015: the entropic spring character of a Gaussian polymer. Note that the `spring
4016: constant' increases with temperature (`entropy elasticity').
4017: 
4018: In this random walk model of a polymer chain, it is straightforward to define
4019: the persistence length of the chain. By this we mean that successive vectors
4020: $\mathbf{a}_i$ are not independent, but tend to be parallel. Over long
4021: distance, this correlation is lost, and the chain behaves like a random
4022: walk. Due to the quantum chemistry of the monomers, an adjacent pair of
4023: vectors $\mathbf{a}_i,\mathbf{a}_{i+1}$ includes preferred angles, for
4024: carbon chains leading to the trans/gauche configurations. This feature is
4025: captured schematically in the freely rotating chain as depicted in figure
4026: \ref{fjc}.
4027: \begin{figure}
4028: \includegraphics[height=6cm]{f_j_chain.eps}
4029: \caption{Freely jointed chain, in which successive bond vectors include
4030: an angle $\theta$.
4031: \label{fjc}}
4032: \end{figure}
4033: Following \cite{doi}, we can obtain the correlation $\langle\mathbf{a}_n
4034: \cdot\mathbf{a}_m\rangle$ as follows. If we fix all vectors $\mathbf{a}_m,
4035: \ldots,\mathbf{a}_{n-1}$, then the average $\langle\mathbf{a}_n\rangle_{
4036: \mathbf{a}_m,\mathbf{a}_{m+1},\ldots,\mathbf{a}_{n-1}\mbox{ fixed}}=
4037: \mathbf{a}_{n-1}\cos\theta$. Multiplication by $\mathbf{a}_m$ produces
4038: \begin{equation}
4039: \langle\mathbf{a}_m\cdot\mathbf{a}_n\rangle_{\mathbf{a}_m,\ldots,\mathbf{a}
4040: _{n-1}\mbox{ fixed}}=\mathbf{a}_m\cdot\mathbf{a}_{n-1}\cos\theta.
4041: \end{equation}
4042: Averaging
4043: over the $\mathbf{a}_m,\ldots,\mathbf{a}_{n-1}$ leads to the recursion
4044: relation $\langle\mathbf{a}_m\cdot\mathbf{a}_n\rangle=\langle\mathbf{a}_m
4045: \cdot\mathbf{a}_{n-1}\rangle\cos\theta$. With the initial condition
4046: $\langle\mathbf{a}^2\rangle=a^2$, we find
4047: \begin{equation}
4048: \langle\mathbf{a}_m\cdot\mathbf{a}_n\rangle=a^2\cos^{|n-m|}\theta.
4049: \end{equation}
4050: Thus, if $\theta=0$, we obtain a rigid rod behaviour, while for $\theta\neq 0$,
4051: there occurs an exponential decay of the correlation between any two bond
4052: vectors $\mathbf{a}_n$ and $\mathbf{a}_m$. This defines a length scale
4053: \begin{equation}
4054: \ell_p\equiv\frac{a}{\log\cos\theta},
4055: \end{equation}
4056: the `persistence length' of the chain. It diverges for $\theta\to 0$, while
4057: for $\theta=90^\circ$, it vanishes, corresponding to the random walk model
4058: discusses above (`freely jointed chain'). As
4059: \begin{equation}
4060: \sum_{k=-\infty}^{\infty}\langle\mathbf{a}_{n+k}\cdot\mathbf{a}_n\rangle=
4061: a^2\left(1+2\sum_{k=1}^{\infty}\cos^k\theta\right)=a^2\frac{1+\cos\theta}{
4062: 1-\cos\theta},
4063: \end{equation}
4064: we find $R_0^2=a^2N(1+\cos\theta)/(1-\cos\theta)$, i.e., statistically, the
4065: freely jointed chain behaves the same as the random walk chain, but with a
4066: rescaled monomer length. The statistical unit in a polymer chain is often
4067: taken to be the Kuhn length $\ell_K=2\ell_p$.
4068: 
4069: Above chain models are often referred to as being phantom, i.e., the chain
4070: can freely cross itself. A physical polymer possesses an excluded volume
4071: and behaves like a so-called self-avoiding chain. Mathematically, this can
4072: be modelled by self-avoiding walks. To include the major effects, it is
4073: sufficient to follow a simple argument due to Flory. Consider a chain with
4074: unknown radius $R$ and internal monomer concentration $c_{\rm int}\simeq N/
4075: R^d$. Assuming that the self-avoiding character is due to monomer-monomer
4076: interactions, the repulsive energy is proportional to the squared concentration,
4077: i.e.,
4078: \begin{equation}
4079: \mathscr{F}_{\rm rep}=\frac{1}{2}Tv(T)c^2,
4080: \end{equation}
4081: with the excluded volume parameter $v(T)$ ($v(T)\equiv(1-2\chi)a^d$ in
4082: Flory's notation, where the $\theta$ condition $\chi=1/2$ corresponds to
4083: ideal chain behaviour). To obtain the total averaged repulsive energy
4084: $\mathscr{F}_{\rm rep|tot}$, we need to average over $c^2$. In a mean field
4085: approach, we take $\langle c^2\rangle\longrightarrow\langle c\rangle^2\sim
4086: c_{\rm int}^2$. We therefore obtain
4087: \begin{equation}
4088: \mathscr{F}_{\rm rep|tot}\simeq Tv(T)c_{\rm int}^2R^d=Tv(T)\frac{N^2}{R^d},
4089: \end{equation}
4090: favouring large values of $R$. This `swelling' competes with the entropic
4091: elasticity contribution $\mathscr{F}_{\rm el}\simeq TR^2/(Na^2)$. The total
4092: free energy becomes
4093: \begin{equation}
4094: \frac{\mathscr{F}}{T}\simeq v(T)\frac{N^2}{R^d}+\frac{R^2}{Na^2},
4095: \end{equation}
4096: with a minimum at $R_F^{d+2}=v(T)a^2N^3$, so that the Flory radius scales
4097: like
4098: \begin{equation}
4099: R_F\sim AN^{\nu} \,\,, \, \mbox{therefore} \,\, \nu=\frac{3}{2+d} \,.
4100: \end{equation}
4101: The values of the exponent $\nu(d=2)=3/4$ and $\nu(d=3)=3/5$ are extremely
4102: close to the best known values $0.75$ and $0.588$.\footnote{An interesting
4103: discussion about the flaws underlying this reasoning can be
4104: found in reference \protect\cite{degennes}.}
4105: 
4106: 
4107: \subsection*{Polymer networks.}
4108: \label{duplantier}
4109: 
4110: A linear excluded volume polymer chain has the size
4111: \begin{equation}
4112: R_g^2\simeq AN^{2\nu}
4113: \end{equation}
4114: with $\nu=0.75$ in $d=2$, and $\nu=0.588$ in $d=3$. Its number of degrees of freedom is
4115: given in terms of the configuration exponent $\gamma$ such that
4116: \begin{equation}
4117: \omega\simeq\mu^NN^{\gamma-1},
4118: \end{equation}
4119: where $\gamma=1.33$ in $d=2$ and $\gamma=1.16$ in $d=3$.
4120: 
4121: Remarkably, similar critical exponents can be obtained for a general polymer
4122: network of the type shown in figure \ref{network}, as originally
4123: by Duplantier \cite{duplantier,duplantier1},
4124: compare also and in references \cite{ohno,schaefer}: In a network ${\cal G}$
4125: consisting of ${\cal N}$ chain segments of lengths $s_1,\ldots,s_{\cal N}$ and
4126: total length $L=\sum_{i=1}^{\cal N}s_i$, the number of configurations
4127: $\omega_{\cal G}$ scales as
4128: \begin{equation}
4129: \label{network}
4130: \omega_{\cal G}(s_1,\ldots,s_{\cal N})=\mu^{L}s_{\cal N}^{\gamma_{\cal G}-1}
4131: {\cal Y}_{\cal G}\left( \frac{s_1}{s_{\cal N}},\ldots,\frac{s_{{\cal N}-1}}{s_
4132: {\cal N}}\right),
4133: \end{equation}
4134: where ${\cal Y}_{\cal G}$ is a scaling function, and $\mu$ is the effective
4135: connectivity constant for self-avoiding walks. The exponent
4136: $\gamma_{\cal G}$ is given by $\gamma_{\cal G}=1-d\nu{\cal L}+\sum_{N
4137: \ge 1}n_N\sigma_N$, where $\nu$ is the swelling exponent, ${\cal L}$
4138: is the number
4139: of independent loops, $n_N$ is the number of vertices with $N$ outgoing legs,
4140: and $\sigma_N$ is an exponent associated with such a vertex.
4141: In $d=2$, this exponent is given by \cite{duplantier,duplantier1}
4142: \begin{equation}
4143: \label{top_coeff}
4144: \sigma_N=\frac{(2-N)(9N+2)}{64}.
4145: \end{equation}
4146: 
4147: In the dense phase in 2D
4148: \cite{duplantier2,duplantier3,owczarek,owczarek1,duplantier4,duplantier5},
4149: and at the $\Theta$ transition \cite{duplantier6}, analogous results can be
4150: obtained.
4151: 
4152: First, consider the dense phase in 2D. If all segments have equal
4153: length $s$ and $L = {\cal N} s$, the configuration number
4154: $\omega_{\cal G}$ of such a network scales as
4155: \cite{duplantier2,duplantier3}
4156: \footnote{Note that due to the factor $\omega_0(L)$ the exponent of
4157: $s$ is $\gamma_{\cal G}$, and not $\gamma_{\cal G}-1$ like in the
4158: expressions used in the dilute phase \cite{duplantier} or at the
4159: $\Theta$ point, for which  $\omega_0(L) \sim L^{- d \nu}$. However,
4160: for 2D dense polymers one has $d \nu = 1$, so that both definitions of
4161: $\gamma_{\cal G}$ are equivalent, cf.~section 3 in reference
4162: \cite{duplantier3}.}
4163: \begin{equation} \label{noc_md}
4164: \omega_{\cal G}(s) \sim \omega_0(L) \, s^{\gamma_{\cal G}} \, \, ,
4165: \end{equation}
4166: %
4167: where $\omega_0(L)$ is the configuration number of a simple ring of
4168: length $L$. For dense polymers, and in contrast to the dilute phase
4169: or at the $\Theta$ point, $\omega_0(L)$ (and thus $\omega_{\cal G}$)
4170: depends on the boundary conditions and even on the shape of the system
4171: \cite{duplantier3,owczarek,owczarek1,duplantier4,duplantier5}. For example,
4172: for periodic boundary conditions (which
4173: we focus on in this study) corresponding to a 2D torus, one finds
4174: $\omega_0(L) \sim \mu^L \, L^{\Psi - 1}$ with a connectivity
4175: constant $\mu$ and $\Psi = 1$ \cite{duplantier3}. However, the network
4176: exponent
4177: %
4178: \begin{equation} \label{nexp}
4179: \gamma_{\cal G} = 1 - {\cal L} + \sum_{N\ge 1}n_N\sigma_N
4180: \end{equation}
4181: %
4182: is {\em universal\/} and depends only on the topology of the
4183: network by the number ${\cal L}$ of independent loops, and by
4184: the number $n_N$ of vertices of order $N$ with vertex exponents
4185: $\sigma_N = (4 - N^2)/32$ \cite{duplantier2,duplantier3}. For a linear chain,
4186: the corresponding exponent $\gamma_{\rm lin} = 19/16$ has been
4187: verified by numerical simulations \cite{duplantier3,grassberger}. For a network
4188: made up of different segment lengths $\left\{s_i\right\}$ of total
4189: length $L = \sum_{i=1}^{\cal N} s_i$, equation (\ref{noc_md})
4190: generalises to (cf.~section 4 in reference \cite{duplantier3})
4191: %
4192: \begin{equation} \label{noc}
4193: \omega_{\cal G}(s_1, \ldots, s_{\cal N})
4194: \, \sim \, \omega_0(L) \, s_{\cal N}^{\gamma_{\cal G}} \,
4195: {\cal Y}_{\cal G}\left(\frac{s_1}{
4196: s_{\cal N}},\ldots,\frac{s_{{\cal N}-1}}{s_{\cal N}}
4197: \right) \,  ,
4198: \end{equation}
4199: %
4200: which involves the scaling function ${\cal Y}_{\cal G}$.
4201: 
4202: \begin{figure}
4203: \includegraphics[width=6.8cm]{network.eps}
4204: \caption{Polymer network ${\cal G}$ with vertices ($\bullet$) of different
4205: order $N$, where $N$ self-avoiding walks are joined ($n_1=5$, $n_3=4$, $n_4=3$,
4206: $n_5=1$).
4207: \label{netw}}
4208: \end{figure}
4209: For polymers in an infinite volume and endowed with an attractive
4210: interaction between neighbouring monomers, a different scaling behaviour
4211: emerges if the system is not below but right at the $\Theta$ point
4212: \cite{duplantier3}. In this case the number of configurations of a general
4213: network ${\cal G}$ is given by
4214: %
4215: \begin{equation} \label{noc_theta}
4216: \overline{\omega}_{\cal G}(s_1, \ldots, s_{\cal N})
4217: \sim \mu^L \, s_{\cal N}^{\overline{\gamma}_{\cal G} - 1} \,
4218: \overline{{\cal Y}}_{\cal G}\left(\frac{s_1}{s_{\cal N}},
4219: \ldots,\frac{s_{{\cal N}-1}}{s_{\cal N}}
4220: \right),
4221: \end{equation}
4222: %
4223: with the network exponent
4224: \begin{equation} \label{nexp_theta}
4225: \overline{\gamma}_{\cal G} =
4226: 1 - d \nu {\cal L} + \sum_{N\ge 1}n_N\overline{\sigma}_N \, \, .
4227: \end{equation}
4228: %
4229: Overlined symbols refer to polymers at the $\Theta$ point.
4230: In $d = 2$, $\nu = 4/7$ and $\overline{\sigma}_N = (2-N)(2N+1)/42$
4231: \cite{duplantier3}.
4232: 
4233: 
4234: \end{appendix}
4235: 
4236: 
4237: \clearpage
4238: 
4239: \begin{thebibliography}{999}
4240: 
4241: \bibitem{alberts} B.~Alberts, A.~Johnson, J.~Lewis, M.~Raff,
4242: K.~Roberts, and P.~Walter, \emph{Molecular biology of the cell\/}
4243: (Garland, New York, 2002).
4244: 
4245: \bibitem{snustad} D.~P.~Snustad and M.~J.~Simmons,
4246: \emph{Principles of Genetics} (John Wiley \& Sons, New York, 2003).
4247: 
4248: \bibitem{kornberg} A.~Kornberg, \emph{DNA synthesis} (W.~H.~Freeman,
4249: San Francisco, CA, 1974).
4250: 
4251: \bibitem{kornberg1} A.~Kornberg and T.~A.~Baker, \emph{DNA Replication}
4252: (W.~H.~Freeman, New York, 1992).
4253: 
4254: \bibitem{franklin} R.~E.~Franklin and R.~G.~Gosling, \NAT \textbf{171},
4255: 740 (1953).
4256: 
4257: \bibitem{watsoncrick} J.~D.~Watson and F.~H.~C.~Crick, \NAT \textbf{171},
4258: 737 (1953).
4259: 
4260: \bibitem{treloar} L.~R.~G.~Treloar, \emph{The physics of rubber elasticity\/}
4261: (Clarendon Press, Oxford, 1975).
4262: 
4263: \bibitem{crick} F.~Crick, \NAT \textbf{227}, 561 (1970).
4264: 
4265: \bibitem{crick1} F.~C.~H.~Crick, in {\em Symp.~Soc.~Exp.~Biol.,
4266: The Biological Replication of Macromolecules}, XII, 138 (1958).
4267: 
4268: \bibitem{bloomfield} V.~A.~Bloomfield, D.~M.~Crothers, and I.~Tinoco,
4269: {\em Physical Chemistry of Nucleic Acids} (Harper \& Row, New York, 1974).
4270: 
4271: \bibitem{rnaworld} R.~F.~Gesteland and J.~F.~Atkins, editors, {\em The
4272: RNA world} (Cold Spring Harbor Laboratory Press, Cold Spring Harbor,
4273: New York, 1993).
4274: 
4275: \bibitem{frank} M.~D.~Frank-Kamenetskii, {\em
4276: Unraveling DNA: The Most Important Molecule of Life\/} (Perseus,
4277: Cambridge, MA, 1997).
4278: 
4279: \bibitem{frank1} M.~D.~Frank-Kamenetskii, Phys.~Rep.~{\bf 288}, 13 (1997).
4280: 
4281: \bibitem{marko} J.~F.~Marko and E.~D.~Siggia, Macromolecules {\bf 28}, 8759
4282: (1995).
4283: 
4284: \bibitem{marko1} J.~F.~Marko and E.~D.~Siggia, Phys.~Rev.~E {\bf 52}, 2912
4285: (1995).
4286: 
4287: \bibitem{ptashne1} M. Ptashne and A. Gunn, {\em Genes and signals}
4288: (Cold Spring Harbor Laboratory Press, Cold Spring Harbor,
4289: New York, 2002).
4290: 
4291: \bibitem{revet} B. R{\'e}vet, B. v. Wilcken-Bergmann, H. Besset, A. Barker,
4292: and B. M{\"u}ller-Hill, Curr. Biol. {\bf 9}, 151 (1999).
4293: 
4294: \bibitem{bell} C. E. Bell and M. Lewis, J. Mol. Biol. {\bf 314}, 1127 (2001).
4295: 
4296: \bibitem{bell1} C. E. Bell, P. Frescura, A. Hochschild, and M. Lewis, Cell
4297: {\bf 101}, 801 (2000).
4298: 
4299: \bibitem{hame_looping} A. Hanke and R. Metzler, Biophys. J. \textbf{85}, 167
4300: (2003).
4301: 
4302: \bibitem{schiessel} H. Schiessel, J. Phys. Cond Mat. \textbf{15}, R699 (2003).
4303: 
4304: \bibitem{kreth} G. Kreth, J. Finsterle, J. von Hase, M. Cremer, and C. Cremer,
4305: Biophys. J. \textbf{86}, 2803 (2004).
4306: 
4307: \bibitem{goddard} N. L. Goddard, G. Bonnet, O. Krichevsky, and A. Libchaber,
4308: Phys. Rev. Lett. \textbf{85}, 2400 (2000).
4309: 
4310: \bibitem{grosberg} A.~Yu. Grosberg and A.~R.  Khokhlov, {\em Statistical
4311: Mechanics of Macromolecules\/} (AIP Press, New York, 1994).
4312: 
4313: \bibitem{calu} G.~C\u{a}lug\u{a}reanu, Czech.
4314: Math.~J.~{\bf 11}, 588 (1961).
4315: 
4316: \bibitem{white1} J.~H.~White, Am.~J.~Math.~{\bf 91}, 693 (1969).
4317: 
4318: \bibitem{fuller} F.~B.~Fuller, Proc.~Natl.~Acad.~Sci.~USA {\bf
4319: 75}, 3557 (1971).
4320: 
4321: \bibitem{white} J. H. White and W. R. Bauer, J. Mol. Biol. \textbf{189},
4322: 329 (1986).
4323: 
4324: \bibitem{bauer} W. R. Bauer, Ann. Rev. Biophys. Bioeng. \textbf{7},
4325: 287 (1978).
4326: 
4327: \bibitem{sun} H. B. Sun, J. Shen, and H. Yokata, Biophys. J. \textbf{79}, 184
4328: (2000).
4329: 
4330: \bibitem{benham} C.~J.~Benham, Comput.~Appl.~Biosci.~{\bf 12}, 375 (1996).
4331: 
4332: \bibitem{goetze} S.~Goetze, A.~Gluch, C.~Benham, and J.~Bode,
4333: Biochemistry {\bf 42}, 154 (2003).
4334: 
4335: \bibitem{levene} A.~V.~Vologodskii, S.~D.~Levene, K.~V.~Klenin,
4336: M.~Frank-Kamenetskii, and N.~R.~Cozzarelli, J.~Mol.~Biol.~{\bf 227},
4337: 1224 (1992).
4338: 
4339: \bibitem{levene1} H.~Tsen, A. Hanke, and S.~D.~Levene, in preparation
4340: (2006).
4341: 
4342: \bibitem{orland} M.~Pillsbury, H.~Orland, and A.~Zee, \PRE
4343: \textbf{72}, 011911 (2005).
4344: 
4345: \bibitem{orland1} H.~Orland and A.~Zee, Nucl.~Phys.~B {\bf 620}, 456 (2002).
4346: 
4347: \bibitem{baiesips} M.~Baiesi, E.~Orlandini, and A.~L.~Stella, \PRL
4348: \textbf{91}, 198102 (2003).
4349: 
4350: \bibitem{loop1}
4351: R. Schleif, Annu. Rev. Biochem. {\bf 61}, 199 (1992).
4352: 
4353: \bibitem{loop2}
4354: K.~S. Matthews, Microbiol. Rev. {\bf 56} 123 (1992).
4355: 
4356: \bibitem{loop3}
4357: T.~M. Dunn, S. Hahn, S. Ogden, and R.~F. Schleif,
4358: Proc. Nat. Acad. Sci. USA {\bf 81}, 5017 (1984).
4359: 
4360: \bibitem{loop4}
4361: M. Geanacopoulos, G. Vasmatzis, V.~B. Zhurkin, and S. Adhya,
4362: Nat. Struct. Biol. {\bf 8}, 432 (2001).
4363: 
4364: \bibitem{loop5}
4365: M.~C. Mossing and M.~T. Record, Science {\bf 233} 889 (1986).
4366: 
4367: \bibitem{loop6}
4368: P. Valentin-Hansen, B. Albrechtsen, and J.~E. Love Larsen,
4369: Embo J. {\bf 5} 2015 (1986).
4370: 
4371: \bibitem{loop7}
4372: K. Rippe, M. Guthold, P.~H. von Hippel, and C. Bustamante,
4373: J. Mol. Biol. {\bf 270}, 125 (1997).
4374: 
4375: \bibitem{loop8}
4376: K. Ogata, K. Sato, and T. Tahirov,
4377: Curr. Opin. Struct. Biol. {\bf 13}, 40 (2003).
4378: 
4379: \bibitem{loop9}
4380: G.~D. van Duyne, Annu. Rev. Biophys. Biomolec. Struct. {\bf 30}, 87 (2001).
4381: 
4382: \bibitem{loop10}
4383: H. Benjamin and N. Cozzarelli, in
4384: {The Robert A. Welch Foundation Conferences on Chemical Research},
4385: vol. 26, p. 107 (1986).
4386: 
4387: \bibitem{loop11}
4388: H. Tsen and S.~D. Levene, Proc. Natl. Acad. Sci. USA. {\bf 94}, 2817 (1997).
4389: 
4390: \bibitem{loop12}
4391: M.~A. Watson, D.~M. Gowers, and S.~E. Halford, J. Mol. Biol. {\bf 298},
4392: 461 (2000).
4393: 
4394: \bibitem{loop13}
4395: M.~L. Embleton, S.~A. Williams, M.~A. Watson, and
4396: S.~E. Halford, J. Mol. Biol. {\bf 289}, 785 (1999).
4397: 
4398: \bibitem{loop14}
4399: M. Deibert, S. Grazulis, G. Sasnauskas, V. Siksnys, and R. Huber,
4400: Nat. Struct. Biol. {\bf 7}, 792 (2000).
4401: 
4402: \bibitem{loop26}
4403: Y.~L. Zhang and D.~M. Crothers, Biophys. J. {\bf 84}, 136 (2003).
4404: 
4405: \bibitem{loop51}
4406: Y. Zhang, A.~E. McEwen, D.~M. Crothers, and S.~D. Levene,
4407: Biophys. J. {\bf 90}, 1903 (2006).
4408: 
4409: \bibitem{loop15}
4410: E.~D. Ross, P.~R. Hardwidge, and L.~J. Maher,
4411: Mol. Cell. Biol. {\bf 21}, (2001).
4412: 
4413: \bibitem{loop16}
4414: R. Schleif, Trends Genet. {\bf 16}, 559 (2000).
4415: 
4416: \bibitem{loop17}
4417: A. Hochschild, Method Enzymol. {\bf 208}, 343 (1991).
4418: 
4419: \bibitem{loop19}
4420: L. Finzi and J. Gelles, Science {\bf 267}, 378 (1995).
4421: 
4422: \bibitem{loop20}
4423: J. M\"uller, S. Oehler, and B. M\"uller-Hill,
4424: J. Mol. Biol. {\bf 257}, 21 (1996).
4425: 
4426: \bibitem{loop21}
4427: K. Rippe, P.~H. von Hippel, and J. Langowski,
4428: Trends Biochem. Sci. {\bf 20}, 500 (1995).
4429: 
4430: \bibitem{loop22}
4431: K. Rippe, Trends Biochem. Sci. {\bf 26}, 733 (2001).
4432: 
4433: \bibitem{loop23}
4434: D. Shore, J. Langowski, and R.~L. Baldwin,
4435: Proc. Nat. Acad. Sci. USA. {\bf 78}, 4833 (1981).
4436: 
4437: \bibitem{loop24}
4438: D. Shore and R.~L. Baldwin, J. Mol. Biol. {\bf 170}, 957 (1983).
4439: 
4440: \bibitem{loop25}
4441: D.~M. Crothers, J. Drak, J.~D. Kahn, and S.~D. Levene,
4442: Method Enzymol. {\bf 212}, 3 (1992).
4443: 
4444: \bibitem{loop27}
4445: K. Klenin, H. Merlitz, and J. Langowski, Biophys. J. {\bf 74}, 780 (1998).
4446: 
4447: \bibitem{loop28}
4448: A.~V. Vologodskii, V.~V. Anshelevich, A.~V. Lukashin,
4449: and M.~D. Frank-Kamenetskii, Nature {\bf 280}, 294 (1979).
4450: 
4451: \bibitem{loop29}
4452: P.~J. Hagerman, Biopolymers {\bf 24}, 1881 (1985).
4453: 
4454: \bibitem{loop30}
4455: S.~D. Levene and D.~M. Crothers, J. Mol. Biol. {\bf 189}, 61 (1986).
4456: 
4457: \bibitem{loop31}
4458: L.~M. Edelman, R. Cheong, and J.~D. Kahn, Biophys. J. {\bf 84}, 1131 (2003).
4459: 
4460: \bibitem{loop32}
4461: J. Shimada and H. Yamakawa, Macromolecules {\bf 17} 689 (1984).
4462: 
4463: \bibitem{loop33}
4464: A. Bacolla, R. Gellibolian, M. Shimizu, S. Amirhaeri, S. Kang, K. Ohshima,
4465: J.~E. Larson, S.~C. Harvey, B.~D. Stollar, and R.~D. Wells,
4466: J. Biol. Chem. {\bf 272}, 16783 (1997).
4467: 
4468: \bibitem{loop34}
4469: S.~M. Law, G.~R. Bellomy, P.~J. Schlax, and M.~T. Record,
4470: J. Mol. Biol. {\bf 230}, 161 (1993).
4471: 
4472: \bibitem{loop35}
4473: M.~H. Hao and W.~K. Olson, Macromolecules {\bf 22}, 3292 (1989).
4474: 
4475: \bibitem{loop36}
4476: B.~D. Coleman, W.~K. Olson, and D. Swigon,
4477: J. Chem. Phys. {\bf 118}, (2003).
4478: 
4479: \bibitem{loop37}
4480: D. Ming, Y. Kong, M.~A. Lambert, Z. Huang, and J. Ma,
4481: Proc. Natl. Acad. Sci. USA {\bf 99}, 8620 (2002).
4482: 
4483: \bibitem{loop38}
4484: I. Tobias, B.~D. Coleman, and W.~K. Olson,
4485: J. Chem. Phys. {\bf 101}, 10990 (1994).
4486: 
4487: \bibitem{loop39}
4488: A. Balaeff, L. Mahadevan, and K. Schulten,
4489: Phys. Rev. Lett. {\bf 83}, 4900 (1999).
4490: 
4491: %\bibitem{loop40}
4492: %A. Hanke and R. Metzler, Biophys. J. {\bf 85}, 167 (2003).
4493: 
4494: \bibitem{loop41}
4495: J. Yan, R. Kawamura, and J.~F. Marko, Phys. Rev. E {\bf 71}, 061905 (2005).
4496: 
4497: \bibitem{loop42}
4498: S. Blumberg, A.~V. Tkachenko, and J.~C. Meiners,
4499: Biophys. J. {\bf 88}, 1692 (2005).
4500: 
4501: \bibitem{loop43}
4502: A. Balaeff, C.~R. Koudella, L. Mahadevan, and K. Schulten,
4503: Philos. Transact. A {\bf 362}, 1355 (2004).
4504: 
4505: \bibitem{loop44}
4506: D. Shore and R.~L. Baldwin, J. Mol. Biol. {\bf 170}, 983 (1983).
4507: 
4508: \bibitem{loop45}
4509: D.~S. Horowitz and J.~C. Wang, J. Mol. Biol. {\bf 173}, 75 (1984).
4510: 
4511: \bibitem{loop49}
4512: X.~J. Lu and W.~K. Olson, Nucleic Acids Res. {\bf 31}, 5108 (2003).
4513: 
4514: \bibitem{loop50}
4515: K. Klenin and J. Langowski, Biopolymers {\bf 54}, 307 (2000).
4516: 
4517: \bibitem{wassermann2} S.~A. Wassermann,
4518: J.~M. Dungan, and N.~R. Cozzarelli, Science {\bf 229}, 171 (1985).
4519: 
4520: \bibitem{deibler} R.~W. Deibler, S. Rahmati, and
4521: E.~L. Zechiedrich, Genes \& Development {\bf 15}, 748 (2001).
4522: 
4523: \bibitem{dna_topo} N.~R. Cozzarelli and J.~C. Wang, {\em DNA topology and
4524: its biological effects} (Cold Spring
4525: Harbor Laboratory Press, Cold Spring Harbor, NY, 1990).
4526: 
4527: \bibitem{liu} L.~F. Liu, C.-C. Liu, and B.~M. Alberts,
4528: Cell {\bf 19}, 697 (1980).
4529: 
4530: \bibitem{mizuuchi} K. Mizuuchi, L.~M. Fisher, M.~H.~O. Dea,
4531: and M. Gellert, Proc. Natl. Acad. Sci. {\bf 77}, 1847 (1980).
4532: 
4533: \bibitem{pollock} T.~J. Pollock and H.~A. Nash,
4534: J. Mol. Biol. {\bf 170}, 1 (1983).
4535: 
4536: \bibitem{spengler} S.~J. Spengler, A. Stasiak, and
4537: N.~R. Cozzarelli, Cell {\bf 42}, 325 (1985).
4538: 
4539: \bibitem{wassermann} S.~A. Wassermann and
4540: N.~R. Cozzarelli, Science {\bf 232}, 951 (1986).
4541: 
4542: \bibitem{wassermann1} S.~A. Wassermann and
4543: N.~R. Cozzarelli, Proc. Natl. Acad. Sci. USA {\bf 82}, 1079 (1984).
4544: 
4545: \bibitem{viguera} E. Viguera, P. Hernandez, D.~B.
4546: Krimer, A.~S. Boistov, R. Lurz, J.~C. Alonso, and J.~B. Schvartzman,
4547: J. Biol. Chem. {\bf 271}, 22414 (1996).
4548: 
4549: \bibitem{sogo} J. Sogo, A. Stasiak, M.~L. Mart\'{\i}nez-Robles, D.~B. Krimer,
4550: P. H{\'e}rnandez, and J.~B. Schvartzman, J. Mol. Biol. {\bf 286}, 637 (1999).
4551: 
4552: \bibitem{topoknot} F. B. Dean, A. Stasiak, T. Koller, and N. R. Cozzarelli,
4553: J. Biol. Chem. {\bf 260}, 4975 (1985).
4554: 
4555: \bibitem{arsuaga} J. Arsuaga, M. Vazquez, S. Trigueros, D. Sumners,
4556: and J. Roca, Proc. Natl. Acad. Sci. USA {\bf 99}, 5373 (2002).
4557: 
4558: \bibitem{electro} A. V. Vologodskii, N. J. Crisona, B. Laurie, P. Pieranski,
4559: V. Katritch, J. Dubochet, and A. Stasiak, J. Mol. Biol. {\bf 278}, 1 (1998).
4560: 
4561: \bibitem{portugal} J. Portugal and A.
4562: Rodriguez-Campos, Nucleic Acids Res. {\bf 24}, 4890 (1996).
4563: 
4564: \bibitem{rodriguez} A. Rodriguez-Campos, J. Biol.
4565: Chem. {\bf 271}, 14150 (1996).
4566: 
4567: \bibitem{staczek} P. Staczek and N.~P.
4568: Higgins, Mol. Microbiol. {\bf 29}, 1435 (1998).
4569: 
4570: \bibitem{arai} Y. Arai, R. Yasuda, K.-I. Akashi, Y. Harada,
4571: H. Miyata, K. Kinosita Jr., and H. Itoh, Nature {\bf 399}, 446 (1999).
4572: 
4573: \bibitem{pieranski} P. Pieranski, S. Kasas, G.
4574: Dietler, J. Dubochet, and A. Stasiak, New J. Phys. {\bf 3}, 101 (2001).
4575: 
4576: \bibitem{saitta} A.~M. Saitta, P.~D. Soper, E. Wasserman,
4577: and M.~L. Klein, Nature {\bf 399}, 46 (1999).
4578: 
4579: \bibitem{stasiak1} A. Stasiak, A. Dobay, J. Dubochet,
4580: and G. Dietler, Science {\bf 286}, 11a (1999).
4581: 
4582: \bibitem{mcnally} T. McNally, {\em The complete book of fly
4583: fishing\/} (Mountain Press, Camden, ME).
4584: 
4585: \bibitem{rybenkov} V.~V. Rybenkov, C. Ullsperger,
4586: A.~V. Vologodskii, and N.~R. Cozzarelli, Science {\bf 277}, 690 (1997).
4587: 
4588: \bibitem{stasiak} A. Stasiak, Curr. Biol. {\bf 10}, R526
4589: (2000).
4590: 
4591: \bibitem{strick} T.~R. Strick, V.~Croquette, and D.
4592: Bensimon, Nature {\bf 404}, 901 (2000).
4593: 
4594: \bibitem{wang} J.~C. Wang, Annu. Rev. Biochem. {\bf 65}, 635
4595: (1996).
4596: 
4597: \bibitem{berger} J.~M. Berger, S.~J. Gamblin, S.~C. Harrrison, and J.~C. Wang,
4598: Nature {\bf 379}, 225 (1996).
4599: 
4600: \bibitem{wang_to} J.~C. Wang, Quart. Rev. Biophys. {\bf 31}, 107 (1998).
4601: 
4602: \bibitem{euler} L. Euler, {\em Solutio problematis ad geometriam
4603: situs pertinentis} Comment. Academiae Sci. Imp. Petropolitanae {\bf 8}, 128
4604: (1736); English translation: Sci. Amer. {\bf 189}, 66 (1953).
4605: 
4606: \bibitem{kepler} J. Kepler, in W. v. Dyck and M. Caspar
4607: (Editors), {\em Johannes Kepler, Gesammelte Werke\/} (Beck, M{\"u}nchen, 1937).
4608: 
4609: \bibitem{adams} A.~C.~C. Adams, {\em The knot book: an elementary
4610: introduction to the mathematical theory of knots\/} (Freeman, New York, 1994).
4611: 
4612: \bibitem{kauffman} L.~H. Kauffman, {\em Knots and physics\/},
4613: Series on Knots and Everything, Vol. I (World Scientific, Singapore, 1993).
4614: 
4615: \bibitem{reidemeister} K. Reidemeister, {\em
4616: Knotentheorie\/} (Springer, Berlin, 1931) [{\em Knot theory} (BSC Assocs.,
4617: Moscow, Idaho, 1983)].
4618: 
4619: \bibitem{listing} J.~B. Listing, {\em Vorstudien zur Topologie},
4620: G{\"o}ttinger Studien (Vandenhoeck und Ruprecht, G{\"o}ttingen, 1848).
4621: 
4622: \bibitem{thomson} W. Thomson, Lord Kelvin, Phil. Mag. {\bf 34},
4623: 15 (1867).
4624: 
4625: \bibitem{thomson1} W. Thomson, Proc. Roy. Soc. Edinb.
4626: {\bf 27}, 59 (1875-76).
4627: 
4628: \bibitem{tait} P.~G. Tait, Trans. Roy. Soc. Edinb. {\bf 28},
4629: 145 (1876-77).
4630: 
4631: \bibitem{tait1} P.~G. Tait, Trans. Roy. Soc. Edinb.
4632: {\bf 32}, 327 (1883-4).
4633: 
4634: \bibitem{tait2} P.~G. Tait, Trans. Roy. Soc. Edinb. {\bf 33},
4635: 493 (1884-85).
4636: 
4637: \bibitem{tait3} P.~G. Tait, {\em Scientific papers\/}
4638: (Cambridge University Press, London, 1898).
4639: 
4640: \bibitem{kirkman} T.~P. Kirkman, Proc. Royal Soc. Edinb.
4641: {\bf 13, 120}, 363 (1884-5).
4642: 
4643: \bibitem{kirkman1} T.~P. Kirkman, Proc. Royal Soc. Edinb.
4644: {\bf 32}, 483 (1884-85).
4645: 
4646: \bibitem{little}  C.~N. Little, Trans. Connecticut Acad. Sci.
4647: 18, {\bf 7}, 27 (1885).
4648: 
4649: \bibitem{little1} C.~N. Little, Trans. Roy. Soc. Edinb.
4650: {\bf 35}. 771 (1890).
4651: 
4652: \bibitem{degriend} P. van de Griend, in {\em History
4653: and science of knots,} edited by J. C. Turner and P. van de Griend
4654: (World Scientific, Singapore, 1996).
4655: 
4656: \bibitem{volo} A.~V. Vologodskii, A.~V. Lukashin,
4657: M.~D. Frank-Kamenetskii, and V.~V. Anshelvich, Sov. Phys. JETP {\bf 39},
4658: 1059 (1974).
4659: 
4660: \bibitem{frisch} H.~L. Frisch and E. Wassermann,
4661: J. Am. Chem. Soc. {\bf 83}, 3789 (1961).
4662: 
4663: \bibitem{delbrueck} M. Delbr{\"u}ck, in {\em Mathematical
4664: problems in biological sciences} (Proc. Symp. Appl. Math. {\bf 14}, 55 (1962)),
4665: edited by R. E. Bellman.
4666: 
4667: \bibitem{sumners} D.~W. Sumners and S.~G.
4668: Whittington, J. Phys. A {\bf 21}, 1689 (1988).
4669: 
4670: \bibitem{pippenger} N. Pippenger, Discrete Appl. Math.
4671: {\bf 25}, 273 (1989).
4672: 
4673: \bibitem{vanderzande} C. Vanderzande, J. Phys. A {\bf 28},
4674: 3681 (1995).
4675: 
4676: \bibitem{frank3} M.~D. Frank-Kamenetskii,
4677: A.~V. Lukashin, and A.~V. Vologodskii, Nature {\bf 258}, 398 (1975).
4678: 
4679: \bibitem{koniaris} K. Koniaris and
4680: M. Muthukumar, Phys. Rev. Lett. {\bf 66}, 2211 (1991).
4681: 
4682: \bibitem{michels} J.~P.~J. Michels and F.~W.
4683: Wiegel, Proc. Roy. Soc. (London) A {\bf 403}, 269 (1986).
4684: 
4685: \bibitem{michels1} J.~P.~J. Michels and F.~W.
4686: Wiegel, Phys. Lett. {\bf 90A}, 381 (1982).
4687: 
4688: \bibitem{janse} E.~J. Janse van Rensburg
4689: and S.~G. Whittington, J. Phys. A {\bf 24}, 3935 (1991).
4690: 
4691: \bibitem{frank2} M.~D. Frank-Kamenetskii,
4692: A.~V. Lukashin, V.~V. Anshelevich, and A.~V. Vologodskii, J. Biomol. Struct.
4693: Dyn. {\bf 2}, 1005 (1985).
4694: 
4695: \bibitem{grosberg4} A. Yu. Grosberg, E-print cond-mat/0207427 (2002).
4696: 
4697: \bibitem{klenin} K.~V. Klenin, A.~V. Vologodskii,
4698: V.~V. Anshelevich, A.~M. Dykhne, and M.~D. Frank-Kamenetskii, J. Biomol.
4699: Struct. Dyn. {\bf 5}, 1173 (1988).
4700: 
4701: \bibitem{shimamura1}  M.~K. Shimamura and
4702: T. Deguchi, Phys. Lett. A {\bf 274}, 184 (2000).
4703: 
4704: \bibitem{katritch} V. Katritch, W.~K. Olson, A.
4705: Vologodskii, J. Dubochet, and A. Stasiak, Phys. Rev. E {\bf 61}, 5545 (2000).
4706: 
4707: \bibitem{dobay} A. Dobay, P.~E. Sottas, J. Dubochet,
4708: and A. Stasiak, Lett. Math. Phys. {\bf 55}, 239 (2001).
4709: 
4710: \bibitem{degennes} P.~G. de Gennes, {\em Scaling concepts in
4711: polymer physics\/} (Cornell University Press, Ithaca, New York, 1979).
4712: 
4713: \bibitem{duplantier1} B. Duplantier, J. Stat. Phys.
4714: {\bf 54}, 581 (1989).
4715: 
4716: \bibitem{deutsch} J.~M. Deutsch, Phys. Rev. E {\bf 59}, R2539
4717: (1999).
4718: 
4719: \bibitem{paul} P. G. Dommersnes, Y. Kantor, and M. Kardar, Phys. Rev. E
4720: \textbf{66}, 031802 (2002).
4721: 
4722: \bibitem{hughes} B.~D. Hughes, {\em Random Walks and Random
4723: Environments, Volume 1: Random Walks\/} (Oxford University Press, Oxford,
4724: 1995).
4725: 
4726: \bibitem{flory1} P.~J. Flory, {\em Statistical Mechanics of
4727: Chain Molecules} (Oxford University Press, Oxford, 1989).
4728: 
4729: \bibitem{slili2d} R. Metzler, A. Hanke, P.~G.  Dommersnes, Y. Kantor,
4730: and M. Kardar, Phys. Rev. Lett. {\bf 88}, 188101 (2002).
4731: 
4732: \bibitem{paraknot} R. Metzler, A. Hanke, P.~G.
4733: Dommersnes, Y. Kantor, and M. Kardar, Phys. Rev. E {\bf 65}, 061103 (2002).
4734: 
4735: \bibitem{duplantier} B. Duplantier, Phys. Rev. Lett.
4736: {\bf 57}, 941 (1986).
4737: 
4738: \bibitem{kafri} Y. Kafri, D. Mukamel, and L. Peliti, Phys.
4739: Rev. Lett. {\bf 85}, 4988 (2000).
4740: 
4741: \bibitem{kafri1} Y. Kafri, D. Mukamel, and L. Peliti,
4742: Eur. Phys. J. B {\bf 27}, 135 (2002).
4743: 
4744: \bibitem{schaefer} L. Sch{\"a}fer, C. v.
4745: Ferber, U. Lehr, and B. Duplantier, Nucl. Phys. B {\bf 374}, 473 (1992).
4746: 
4747: \bibitem{bennaim} M.~B. Hastings, Z.~A. Daya, E.
4748: Ben-Naim, and R.~E. Ecke, Phys. Rev. E {\bf 66}, 025102(R) (2002).
4749: 
4750: \bibitem{descloizeaux} J. des Cloizeaux, J. Phys.
4751: (France) Lett. {\bf 42}, L433 (1981).
4752: 
4753: \bibitem{dobay1} A. Dobay, J. Dubochet, K. Millett,
4754: P.-E. Sottas, and A. Stasiak, Proc. Natl. Acad. Sci. USA {\bf 100},
4755: 5611 (2003).
4756: 
4757: \bibitem{leguillou} J.~C. LeGuillou and J.
4758: Zinn-Justin, Phys. Rev. B {\bf 21}, 3976 (1989).
4759: 
4760: \bibitem{zinn} J.~C. LeGuillou and J.
4761: Zinn-Justin, J. Physique (France) {\bf 50}, 1365 (1989).
4762: 
4763: \bibitem{janse1} E.~J. Janse van Rensburg,
4764: S.~G. Whittington, and N. Madras, J. Phys. A {\bf 23}, 1589 (1990).
4765: 
4766: \bibitem{li} B. Li, S.~N. Madras, and A.~D. Sokal, J. Stat.
4767: Phys. {\bf 80}, 661 (1995).
4768: 
4769: \bibitem{orlandini} E. Orlandini, M.~C. Tesi, E.~J.
4770: Janse van Rensburg, and S.~G. Whittington, J. Phys. A {\bf 31}, 5953 (1998).
4771: 
4772: \bibitem{valle} E. Ercolini, F. Valle, J. Adamcik, G. Witz, R. Metzler, P.
4773: de los Rios, J. Roca, and G. Dietler, E-print cond-mat/0609084.
4774: 
4775: \bibitem{guttmann} A.~J. Guttmann, J. Phys. A {\bf 22},
4776: 2807 (1989).
4777: 
4778: \bibitem{guitter} E. Guitter and E. Orlandini,
4779: J. Phys. A {\bf 32}, 1359 (1999).
4780: 
4781: \bibitem{quake} S.~R. Quake, Phys. Rev. Lett. {\bf 73}, 3317
4782: (1994).
4783: 
4784: \bibitem{quake1} S.~R. Quake, Phys. Rev. E {\bf 52}, 1176 (1986).
4785: 
4786: \bibitem{grosberg2} A.~Yu. Grosberg, A. Feigel, and
4787: Y. Rabin, Phys. Rev. E {\bf 54}, 6618 (1996).
4788: 
4789: \bibitem{erman} B. Erman and J.~E. Mark, {\em
4790: Structures and properties of rubberlike networks\/} (Oxford, New York, 1997).
4791: 
4792: \bibitem{ferry} J.~D. Ferry, {\em Viscoelastic properties of
4793: polymers\/} (Wiley, New York, 1970).
4794: 
4795: \bibitem{ward} I.~M. Ward and D. Hadley, {\em
4796: An Introduction to the Mechanical Properties of Solid Polymers\/}
4797: (John Wiley and Sons Ltd., New York).
4798: 
4799: \bibitem{degennesma} P.-G. de Gennes, Macromolecules {\bf 17},
4800: 703 (1984).
4801: 
4802: \bibitem{lai1} P.~K. Lai, Y.~J. Sheng, and H.~K. Tsao,
4803: Phys. Rev. Lett. {\bf 87}, 175503 (2001).
4804: 
4805: \bibitem{shimamura} M.~K. Shimamura and
4806: T. Deguchi, Phys. Rev. E {\bf 64}, 020801 (2001).
4807: 
4808: \bibitem{farago} O. Farago, Y. Kantor, and M. Kardar,
4809: Europhys. Lett. {\bf 60}, 53 (2002).
4810: 
4811: \bibitem{pull} R. Metzler, Y. Kantor, and M. Kardar,
4812: Phys. Rev. E {\bf 66}, 022102 (2002).
4813: 
4814: \bibitem{sheng} Y.~J. Sheng, P.~K. Lai, and H.~K. Tsao,
4815: Phys. Rev. E {\bf 61}, 2895 (2000).
4816: 
4817: \bibitem{sheng1} P.~K. Lai, Y.~J. Sheng, and H.~K. Tsao,
4818: Macromol. Theor. Simul. {\bf 9}, 578 (2000).
4819: 
4820: \bibitem{roya} R. Zandi, Y. Kantor, and M. Kardar,
4821: E-print cond-mat/0306587 (2003).
4822: 
4823: \bibitem{marcone} B. Marcone, E. Orlandini, A. L. Stella, and F. Zonta,
4824: J. Phys. A \textbf{38}, L15 (2005).
4825: 
4826: \bibitem{virnau} P. Virnau, Y. Kantor, and M. Kardar, J. Amer. Chem. Soc.
4827: \textbf{127}, 15102 (2005).
4828: 
4829: \bibitem{stella} E. Orlandini, A.~L. Stella, and C.
4830: Vanderzande, Phys. Rev. E {\bf 68}, 031804 (2003).
4831: 
4832: \bibitem{orlanddense} E. Orlandini, A.~L. Stella, and C. Vanderzande,
4833: J. Stat. Phys. {\bf 115}, 681 (2004).
4834: 
4835: \bibitem{vilgis} T.~A. Vilgis, Phys. Rep. {\bf 336}, 167 (2000).
4836: 
4837: \bibitem{kholodenko} A.~L. Kholodenko and T.~A.
4838: Vilgis, Phys. Rep. {\bf 298}, 251 (1998).
4839: 
4840: \bibitem{edwards1} S.~F. Edwards, Proc. Phys. Soc. (London)
4841: {\bf 91}, 513 (1967).
4842: 
4843: \bibitem{edwards2} S.~F. Edwards, J. Phys. A {\bf 1}, 15
4844: (1968).
4845: 
4846: \bibitem{doi} M. Doi and S.~F. Edwards {\em The Theory
4847: of Polymer Dynamics\/} (Clarendon Press, Oxford, 1986).
4848: 
4849: \bibitem{otto} M. Otto and T.~A. Vilgis,
4850: Phys. Rev. Lett. {\bf 80}, 881 (1998).
4851: 
4852: \bibitem{otto1} M. Otto and T.~A. Vilgis, J. Phys. A
4853: {\bf 29}, 3893 (1996).
4854: 
4855: \bibitem{vilgis1} T.~A. Vilgis and M. Otto, Phys.
4856: Rev. E {\bf 56}, R1314 (1997).
4857: 
4858: \bibitem{ferrari} F. Ferrari, H. Kleinert, and I.
4859: Lazzizzera, Int. J. Mod. Phys. B {\bf 14}, 3881 (2000).
4860: 
4861: \bibitem{ferrari1} F. Ferrari, H. Kleinert, and I.
4862: Lazzizzera, Phys. Lett. A {\bf 276}, 31 (2000).
4863: 
4864: \bibitem{grone2} A.~Yu. Grosberg and S. Nechaev,
4865: Adv. Polym. Sci. {\bf 106}, 1 (1993).
4866: 
4867: \bibitem{grone1} A.~Yu. Grosberg and S. Nechaev,
4868: Europhys. Lett. {\bf 20}, 613 (1992).
4869: 
4870: \bibitem{grone} A.~Yu. Grosberg and S. Nechaev,
4871: J. Phys. A {\bf 25}, 4659 (1992a).
4872: 
4873: \bibitem{nechaev} S. Nechaev, in {\em Topological aspects of
4874: low dimensional systems\/}, Lecture notes of Les Houches 1998 summer school;
4875: extended version available as E-print cond-mat/9812205.
4876: 
4877: \bibitem{nechaev1} O. Vasilyev and S. Nechaev,
4878: JETP {\bf 93}, 1119 (2001).
4879: 
4880: \bibitem{craighead} S.~W.~P. Turner, M. Cabodi, and H.~G.
4881: Craighead, Phys. Rev. Lett. {\bf 88}, 128103 (2002).
4882: 
4883: \bibitem{bennaim1} E. Ben-Naim, Z.~A. Daya, P.
4884: Vorobieff, and R.~E. Ecke, Phys. Rev. Lett. {\bf 86}, 1414 (2001).
4885: 
4886: \bibitem{maier} B. Maier and J.~O. R{\"a}dler,
4887: Phys. Rev. Lett. {\bf 82}, 1911 (1999).
4888: 
4889: \bibitem{walter} H. Walter and D.~E. Brooks, FEBS
4890: Lett. {\bf 361}, 135 (1995).
4891: 
4892: \bibitem{vasilevskaya} V.~V. Vasilevskaya, A.~R.
4893: Khokhlov, Y. Matsuzawa, and K. Yoshikawa, J. Chem. Phys. {\bf 102},
4894: 6595 (1995).
4895: 
4896: \bibitem{lerman} L. Lerman, Proc. Natl. Acad. Sci. USA {\bf 68},
4897: 1886 (1971).
4898: 
4899: \bibitem{garel} T. Garel, {\em Remarks on homo- and hetero-polymeric
4900: aspects of protein folding}, E-print cond-mat/0305053 (2003).
4901: 
4902: \bibitem{duplantier2} B. Duplantier, J. Phys. A {\bf 19},
4903: L1009 (1986).
4904: 
4905: \bibitem{duplantier3} B. Duplantier and H. Saleur,
4906: Nucl. Phys. B {\bf 290}, 291 (1987).
4907: 
4908: \bibitem{jacobsen} J.~L. Jacobsen, N. Read, and H.
4909: Saleur, Phys. Rev. Lett. {\bf 90}, 090601 (2003).
4910: 
4911: \bibitem{duplantier4} B. Duplantier, Phys. Rev. Lett.
4912: {\bf 71}, 4274 (1993).
4913: 
4914: \bibitem{owczarek} A.~L. Owczarek, T. Prellberg, and
4915: R. Brak, Phys. Rev. Lett. {\bf 70}, 951 (1993).
4916: 
4917: \bibitem{owczarek1} A.~L. Owczarek, T. Prellberg,
4918: and R. Brak, Phys. Rev. Lett. {\bf 71}, 4275 (1993).
4919: 
4920: \bibitem{duplantier7} B. Duplantier and F.
4921: David, J. Stat. Phys. {\bf 51}, 327 (1988).
4922: 
4923: \bibitem{kondev} J. Kondev and J.~L. Jacobsen,
4924: Phys. Rev. Lett. {\bf 81}, 2922 (1998).
4925: 
4926: \bibitem{fisher} M.~E. Fisher, Physica D {\bf 38}, 112 (1989).
4927: 
4928: \bibitem{seno} F. Orlandini, F. Seno, A.~L. Stella,
4929: and M.~C. Tesi, Phys. Rev. Lett. {\bf 68}, 488 (1992).
4930: 
4931: \bibitem{dekeyser} R. Dekeyser, E. Orlandini, A.~L.
4932: Stella, and M.~C. Tesi, Phys. Rev. E {\bf 52}, 5214 (1995).
4933: 
4934: \bibitem{cardy} J. Cardy, J. Phys. A {\bf 34}, L665 (2001).
4935: 
4936: \bibitem{dense} A. Hanke, R. Metzler, P.~G. Dommersnes,
4937: Y. Kantor, and M. Kardar, Eur. Phys. J. E {\bf 12}, 347 (2003).
4938: 
4939: \bibitem{gambini} R. Gambini and J. Pullin,
4940: {\em Loops, knots, gauge theories and quantum gravity} (Cambridge University
4941: Press, Cambridge, UK, 1996).
4942: 
4943: \bibitem{slili3d} R. Metzler, New J. Phys. {\bf 4}, 91.1 (2002).
4944: 
4945: \bibitem{grosberg1} A.~Yu. Grosberg, Phys. Rev. Lett.
4946: {\bf 85}, 3858 (2000).
4947: 
4948: \bibitem{edwards2a} R.~C. Ball, M.~Doi, S.~F. Edwards,
4949: and M. Warner, Polymer {\bf 22}, 1010 (1981).
4950: 
4951: \bibitem{edwards2d} M. Doi and S.~F. Edwards, J. Chem.
4952: Soc. Farad. Trans. {\bf 274}, 1802 (1978).
4953: 
4954: \bibitem{edwards2c} S.~F. Edwards and T.~A. Vilgis,
4955: Polymer {\bf 27}, 483 (1986).
4956: 
4957: \bibitem{edwards2b} P.~G. Higgs and R.~C. Ball,
4958: Europhys. Lett. {\bf 8}, 357 (1989).
4959: 
4960: \bibitem{sommer} J.~U. Sommer, J. Chem. Phys. {\bf 97}, 5777
4961: (1992).
4962: 
4963: \bibitem{hame_cpl} A. Hanke and R. Metzler, Chem. Phys. Lett. \textbf{359},
4964: 22 (2002).
4965: 
4966: \bibitem{meamb_ctn} R. Metzler and T. Ambj{\"o}rnsson, J. Comp. Theoret.
4967: Nanosc. \textbf{2}, 389 (2005).
4968: 
4969: \bibitem{revzin} A. Revzin, Editor, {\em The biology of
4970: non-specific DNA-protein interactions} (CRC Press, Boca Raton, FL, 1990).
4971: 
4972: \bibitem{delcourt}
4973: S.~G.~Delcourt and  R.D.~Blake, J.~Biol.~Chem.~{\bf 266}, 15160 (1991).
4974: 
4975: \bibitem{wartell} R. M. Wartell and A. S. Benight, Phys. Rep. \textbf{126},
4976: 67 (1985).
4977: 
4978: \bibitem{blake} R.~D.~Blake, R.~D., Bizzaro, J.~W., Blake, J.~D., Day, G.~R.,
4979: Delcourt, S.~G., Knowles, J., Marx, K.~A., \& SantaLucia, J., Jr. (1999)
4980: Bioinf.~{\bf 15}, 370 (1999).
4981: 
4982: \bibitem{blossey} R. Blossey and E. Carlon, Phys. Rev. E \textbf{68}, 061911
4983: (2003).
4984: 
4985: \bibitem{yeramian} E. Yeramian, Gene {\bf 255}, 139 (2000).
4986: 
4987: \bibitem{yeramian1} E. Yeramian, Gene {\bf 255}, 151-168 (2000).
4988: 
4989: \bibitem{carlon} E. Carlon, M. L. Malki, and R. Blossey,
4990: Phys. Rev. Lett. {\bf 94}, 178101 (2005).
4991: 
4992: \bibitem{williams} M. C. Williams, Biophys. Textbook online
4993: {\em Optical tweezers: measuring piconewton forces},
4994: {\tt www.biophysics.org/btol/}.
4995: 
4996: \bibitem{busta} S. B. Smith, Y. J. Cui, and C. Bustamante, Science \textbf{271},
4997: 795 (1996).
4998: 
4999: \bibitem{williams1} M. C. Williams, I. Rouzina, and V. A. Bloomfield,
5000: Accts. Chem. Res. \textbf{35}, 159 (2002).
5001: 
5002: \bibitem{hwa}  T. Hwa, E. Marinari, K. Sneppen, and L.-H. Tang,
5003: Proc. Natl. Acad. Sci. USA {\bf 100}, 4411 (2003).
5004: 
5005: \bibitem{pcrrev} C. W. Dieffenbach and G. S. Dycksler, {\em PCR primer: a
5006: laboratory manual} (Cold Spring Harbor Laboratory Press, Cold Spring Harbor,
5007: NY, 1995).
5008: 
5009: \bibitem{pcr} M. J. McPherson and S. G. M{\o}ller, {\em PCR Basics}
5010: (Springer-Verlag Telos, New York, 2000.
5011: 
5012: \bibitem{gueron} M. Gu{\'e}ron, M. Kochoyan, and
5013: J.~L. Leroy, Nature {\bf 328} 89 (1987).
5014: 
5015: \bibitem{krueger} A.~Krueger, E.~Protozanova, and M.~D.~Frank-Kamenetskii,
5016: Biophys. J., at press. (2006).
5017: 
5018: \bibitem{kittel} C. Kittel, Am. J. Phys. {\bf 37}, 917 (1969).
5019: 
5020: \bibitem{altan} G. Altan-Bonnet, A. Libchaber,
5021: and O. Krichevsky, \PRL {\bf 90}, 138101 (2003).
5022: 
5023: \bibitem{zocchi} Y.~Zeng, A.~Montrichok and G.~Zocchi, J. Mol. Biol.
5024: {\bf 339}, 67 (2004).
5025: 
5026: \bibitem{nmr} M. Gu{\'e}ron, M. Kochoyan, and J. L. Leroy, Nature \textbf{328},
5027: 89 (1987).
5028: 
5029: \bibitem{nmr1} M. Gu{\'e}ron and J. L. Leroy, Methods Enzymol. \textbf{261},
5030: 383 (1995).
5031: 
5032: \bibitem{nmr2} I. M. Russu, Methods Enzymol. \textbf{261}, 383 (1995).
5033: 
5034: \bibitem{russu} C. Chen and I. M. Russu, Biophys. J. \textbf{87}, 2545 (2004).
5035: 
5036: \bibitem{vonhippel} D.~E.~Jensen and P.~H.~von Hippel, J. Biol. Chem.
5037: {\bf 251}, 7198 (1976).
5038: 
5039: \bibitem{rich} R.~L.~Karpel, in reference \protect\cite{revzin}.
5040: 
5041: \bibitem{vonhippel1} D.~E.~Jensen, R.~C.~Kelly and P.~H.~von Hippel, J. Biol.
5042: Chem. {\bf 251}, 7215 (1976).
5043: 
5044: \bibitem{rich1} R.~L.~Karpel, IUBMB Life {\bf 53}, 161 (2002).
5045: 
5046: \bibitem{pant} K.~Pant, R.~L.~Karpel and M.~C.~Williams, J. Mol. Biol. {\bf
5047: 327}, 571 (2003).
5048: 
5049: \bibitem{pant1} K.~Pant, R.~L.~Karpel, I.~Rouzina and M.~C.~Williams, J. Mol.
5050: Biol. {\bf 336}, 851 (2004).
5051: 
5052: \bibitem{pant2} K. Pant, R. L. Karpel, I. Rouzina, and M. C. Williams, J. Mol.
5053: Biol. \textbf{349}, 317 (2005).
5054: 
5055: \bibitem{poland} D. Poland and H.~A. Scheraga,
5056: {\em Theory of Helix-Coil Transitions in Biopolymers\/} (Academic Press,
5057: New York, 1970).
5058: 
5059: \bibitem{cantor} C. R. Cantor and P. R. Schimmel, Biophysical Chemistry
5060: (W. H. Freeman and Company, New York, NY, 1980).
5061: 
5062: \bibitem{richard} C. Richard and A.~J. Guttmann, J. Stat. Phys. {\bf 115},
5063: 925 (2004).
5064: 
5065: \bibitem{hanke} A. Hanke, J. Phys.: Condens. Matter {\bf 17}, S1731 (2005).
5066: 
5067: \bibitem{fixman} M. Fixman and J.~J. Freiere, Biopol. {\bf 16}, 2693 (1977).
5068: 
5069: \bibitem{santalucia} J. SantaLucia Jr., Proc.~Natl.~Acad.~Sci.~{\bf 95},
5070: 1460 (1998).
5071: 
5072: \bibitem{lazurkin}  Yu. S. Lazurkin, M. Frank-Kamenetskii, and
5073: E. N. Trifonov, Biopol. \textbf{9}, 1253 (1970).
5074: 
5075: \bibitem{tobias_prl} T.~Ambj{\"o}rnsson, S.~K.~Banik, O.~Krichevsky, and
5076: R.~Metzler, Phys.~Rev.~Lett. \textbf{97}, 128105 (2006).
5077: 
5078: \bibitem{tobias_long} T.~Ambj{\"o}rnsson, S.~K.~Banik, O.~Krichevsky, and
5079: R.~Metzler, in preparation.
5080: 
5081: \bibitem{garel1} T. Garel, C. Monthus, and H. Orland, Europhys. Lett. {\bf 55},
5082: 132 (2001).
5083: 
5084: \bibitem{hame_comment} A. Hanke and R. Metzler,
5085: Phys. Rev. Lett. {\bf 90}, 159801 (2003).
5086: 
5087: \bibitem{somendra_comment} S.~M. Bhattacharjee, Europhys. Lett. {\bf 57}, 772.
5088: 
5089: \bibitem{landau} L.~D. Landau and E.~M. Lifshitz, {\em Physical Kinetics}
5090: (Butterworth-Heinemann, Oxford, UK, 1981).
5091: 
5092: \bibitem{hame_jpa} A. Hanke and R. Metzler, J. Phys. A \textbf{36}, L473
5093: (2003).
5094: 
5095: \bibitem{tobias_jpc} T. Ambj{\"o}rnsson and R. Metzler, J. Phys. Cond. Mat.
5096: \textbf{17}, S1841 (2005).
5097: 
5098: \bibitem{tobias_pre} T. Ambj{\"o}rnsson and R. Metzler, Phys. Rev. E
5099: \textbf{72}, 030901(R) (2005).
5100: 
5101: \bibitem{tobias_leif} R. Metzler and T. Ambj{\"o}rnsson, J. Biol. Phys.
5102: \textbf{31}, 339 (2005).
5103: 
5104: \bibitem{hansen} T. Novotn{\'y}, J. N. Pedersen, M. S. Hansen, T.
5105: Ambj{\"o}rnsson, and R. Metzler, E-print cond-mat/0610752.
5106: 
5107: \bibitem{vankampen} N.~G.~van Kampen, Stochastic Processes
5108: in Physics and Chemistry (North-Holland, Amsterdam, 1992).
5109: 
5110: \bibitem{oleg1} O.~Krichevsky and G. Bonnet, Rep.~Prog.~Phys.~\textbf{65},
5111: 251 (2002).
5112: 
5113: \bibitem{peyrard} M. Peyrard and A. R. Bishop, Phys. Rev. Lett. \textbf{62},
5114: 2755 (1989).
5115: 
5116: \bibitem{dauxois} T. Dauxois, M. Peyrard, and A. R. Bishop, Phys. Rev. E
5117: \textbf{44}, R44 (1993).
5118: 
5119: \bibitem{peyrard_cm} B. S. Alexandrov, L. T. Wille, K. {\O}. Rasmussen, A. R.
5120: Bishop, and K. B. Blagoev, E-print cont-mat/0601555.
5121: 
5122: \bibitem{campa} A. Campa and A. Giansanti, Phys. Rev. E \textbf{58}, 3585
5123: (1998).
5124: 
5125: \bibitem{the_prl_footnote} S. Ares, N. K. Voulgarakis, {\O}. Rasmussen,
5126: and A. R. Bishop, Phys. Rev. Lett. \textbf{94}, 035504 (2005); cf. reference
5127: [19].
5128: 
5129: \bibitem{peyrard_comment} T. S. van Erp, S. Cuesta-Lopez, J.-G. Hagmann,
5130: and M. Peyrard, Phys. Rev. Lett. \textbf{95}, 218104 (2005).
5131: 
5132: \bibitem{suman} S. K. Banik, T. Ambj{\"o}rnsson, and R. Metzler, Europhys.
5133: Lett. \textbf{71}, 852 (2005).
5134: 
5135: \bibitem{gillespie} D. T. Gillespie, J. Comp. Phys. \textbf{22}, 403 (1976).
5136: 
5137: \bibitem{gillespie1} D. T. Gillespie, J. Phys. Chem. \textbf{81}, 2340 (1977).
5138: 
5139: \bibitem{choi} Ch. H. Choi, G. Kalosakas, K. {\O}. Ramsussen, M. Hiromura, A. R.
5140: Bishop, and A. Usheva, Nucl. Acids Res. \textbf{32}, 1584 (2004).
5141: 
5142: \bibitem{hans} H.~C.~Fogedby and R.~Metzler, in preparation.
5143: 
5144: \bibitem{hans_short} H.~C.~Fogedby and R.~Metzler, E-print cond-mat/0608458.
5145: 
5146: \bibitem{tobias_jpca} T.~Ambj{\"o}rnsson and R. Metzler, J. Phys. Cond. Mat.
5147: \textbf{17}, S4305 (2005).
5148: 
5149: \bibitem{amlomme} T.~Ambj{\"o}rnsson, M. A. Lomholt, and R. Metzler, J. Phys.
5150: Cond. Mat. \textbf{17}, S3945 (2005).
5151: 
5152: \bibitem{tobias_pb} T.~Ambj{\"o}rnsson and R. Metzler, Phys. Biol. \textbf{1},
5153: 77 (2004).
5154: 
5155: \bibitem{lwoff} A.~Lwoff, \emph{Bacteriol.~Rev.\/} \textbf{17}, 269 (1953).
5156: 
5157: \bibitem{ptashne} M.~Ptashne, \emph{A Genetic Switch\/}, 3rd edition
5158: (Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York, 2004)
5159: 
5160: \bibitem{davidson} E.~H.~Davidson et al., \emph{Science\/} \textbf{295},
5161: 1669 (2002).
5162: 
5163: \bibitem{reichardt} L.~Reichardt and A.~D.~Kaiser, \PNAS \textbf{68},
5164: 2185 (1971).
5165: 
5166: \bibitem{metzi} R.~Metzler, \PRL \textbf{87}, 068103 (2001).
5167: 
5168: \bibitem{mewo} R.~Metzler and P.~G.~Wolynes, \emph{Chem. Phys.} \textbf{284},
5169: 469 (2002).
5170: 
5171: \bibitem{little2} J.~W.~Little, D.~P.~Shipley, and D.~W.~Wert, \emph{EMBO J.}
5172: \textbf{18}, 4299 (1999).
5173: 
5174: \bibitem{rozanov} D.~V.~Rozanov, R.~D'Ari, and S.~P.~Sineoky, \emph{J.
5175: Bacteriol.} \textbf{180}, 6306 (1998).
5176: 
5177: \bibitem{aurell} E.~Aurell and K.~Sneppen, \PRL \textbf{88}, 048101 (2002).
5178: 
5179: \bibitem{aurell1} E.~Aurell, S.~Brown, J.~Johanson, and K.~Sneppen, \PRE
5180: \textbf{65}, 051914 (2002).
5181: 
5182: \bibitem{shea} M.~A.~Shea and G.~K.~Ackers, \emph{J. Mol. Biol.} \textbf{181},
5183: 211 (1985).
5184: 
5185: \bibitem{arkin} A.~Arkin, J.~Ross, and H.~H.~McAdams, \emph{Genetics}
5186: \textbf{149}, 1633 (1998).
5187: 
5188: \bibitem{bakk2} A.~Bakk, R.~Metzler, and K.~Sneppen, \emph{Biophys. J.}
5189: \textbf{86}, 58 (2004).
5190: 
5191: \bibitem{bakk3} A.~Bakk, R.~Metzler, and K.~Sneppen, \emph{Isr. J. Chem.}
5192: \textbf{44}, 309 (2004).
5193: 
5194: \bibitem{berg} O.~G.~Berg, R.~B.~Winter, and P.~H.~von Hippel, Biochem.
5195: \textbf{20}, 6929 (1981).
5196: 
5197: \bibitem{wbvh} R.~B.~Winter, O.~G.~Berg, and P.~H.~von Hippel, \emph{Biochem.}
5198: \textbf{20}, 6961 (1981).
5199: 
5200: \bibitem{bvh} P.~H.~von Hippel and O.~G.~Berg, \PNAS \textbf{83}, 1608 (1986).
5201: 
5202: \bibitem{kaladimos} C.~G.~Kalodimos, N.~Biris, A.~M.~Bonvin, M.~M.~Levandoski,
5203: M.~Guennuegues, R.~Boelens, and R.~Kaptein, \emph{Science\/} \textbf{305},
5204: 350 (2004).
5205: 
5206: \bibitem{lewis} M.~Lewis, G.~Chang, N.~C.~Horton, M.~A.~Kercher, H.~C.~Pace,
5207: M.~A.~Schumacher, R.~G.~Brennan, and P.~Lu, \emph{Science\/} \textbf{271},
5208: 1247 (1996).
5209: 
5210: \bibitem{gerland} U.~Gerland, J.~D.~Moroz, and T.~Hwa, \PNAS \textbf{99}, 12015
5211: (2002).
5212: 
5213: \bibitem{bakk} A.~Bakk and R.~Metzler, \emph{FEBS Lett.} \textbf{563}, 66
5214: (2004).
5215: 
5216: \bibitem{bakk1} A.~Bakk and R.~Metzler, \emph{J. Theor. Biol.} \textbf{231},
5217: 525 (2004).
5218: 
5219: \bibitem{kao} Y.~Kao-Huang, A.~Revzin, A.~P.~Butler, P.~O'Conner, D.~W.~Noble,
5220: and P.~H.~von Hippel, \PNAS \textbf{74}, 4228 (1977).
5221: 
5222: \bibitem{stormo} G.~D.~Stormo, \emph{Bioinformatics\/} \textbf{16}, 16
5223: (2000).
5224: 
5225: \bibitem{adam} G. Adam and M. Delbr{\"{u}}ck, in A. Rich, and N. Davidson,
5226: Eds., Structural Chemistry and Molecular Biology  (W. H. Freeman, San
5227: Francisco, CA, 1968).
5228: 
5229: \bibitem{eigen} P. H. Richter and M. Eigen, Biophys. Chem. \textbf{2}, 255
5230: (1974).
5231: 
5232: \bibitem{paradox} I.~M. Sokolov, J. Mai and A. Blumen,
5233: Phys. Rev. Lett. {\bf 79}, 857 (1997).
5234: 
5235: \bibitem{dirk} D.~Brockmann and T.~Geisel, Phys.~Rev.~Lett.~\textbf{91},
5236: 048303.
5237: 
5238: \bibitem{JLum} I.M. Sokolov, J. Mai, and A. Blumen, J. Luminesc. \textbf{76-77},
5239: 377 (1998).
5240: 
5241: \bibitem{slutsky} M. Slutsky and L. A. Mirny, Biophys. J. {\bf 87}, 4021
5242: (2004).
5243: 
5244: \bibitem{coppey} M. Coppey, O. B{\'e}nichou, R. Voituriez,
5245: and M. Moreau, Biophys. J. {\bf 87}, 1640 (2004).
5246: 
5247: \bibitem{marko2} J. Yan and J.~F. Marko, Phys.~Rev.~Lett.~{\bf 93}, 108108
5248: (2004).
5249: 
5250: \bibitem{grillo} A. O. Grillo, M. P. Brown, and C. A. Royer, J. Mol. Biol.
5251: \textbf{287}, 539 (1999).
5252: 
5253: \bibitem{record} R. S. Spolar and M. T. Record, Science \textbf{263}, 777
5254: (1994).
5255: 
5256: \bibitem{shima} N. Shimamoto, J. Biol. Chem. \textbf{274}, 15293 (1999).
5257: 
5258: \bibitem{slutsky1} M. Slutsky, M. Kardar, and L. A. Mirny, Phys. Rev. E
5259: \textbf{69}, 061903 (2004).
5260: 
5261: \bibitem{igor_prot} I. M. Sokolov, R. Metzler, K. Pant, and M. C. Williams,
5262: Biophys. J. \textbf{89}, 895 (2005).
5263: 
5264: \bibitem{mcghee} J. D. McGhee and P. H. von Hippel, J. Mol. Biol. \textbf{86},
5265: 469 (1974).
5266: 
5267: \bibitem{igor_finite} I. M. Sokolov, R. Metzler, K. Pant, and M. C. Williams,
5268: Phys. Rev. E \textbf{72}, 041102 (2005).
5269: 
5270: \bibitem{klafternature} M. F. Shlesinger, G. M. Zaslavsky, and J. Klafter,
5271: Nature {\bf 363}, 31 (1993).
5272: 
5273: \bibitem{physicstoday} J. Klafter, M. F. Shlesinger and G. Zumofen, Phys.
5274: Today \textbf{49} (2), 33 (1996).
5275: 
5276: \bibitem{report} R. Metzler and J. Klafter, Phys. Rep. \textbf{339}, 1
5277: (2000).
5278: 
5279: \bibitem{report1} R. Metzler and J. Klafter, J. Phys. A \textbf{37}, R161
5280: (2004).
5281: 
5282: \bibitem{coffey} A. V. Chechkin, V. Yu. Gonchar, R. Metzler, and J. Klafter,
5283: Adv. Chem. Phys., at press (2006).
5284: 
5285: \bibitem{stanley} G.~M.~Viswanathan, S.~V.~Buldyrev, S.~Havlin, M.~G.~.E~.~da
5286: Luz, E.~P.~.~Raposo, and H.~E.~Stanley, Nature {\bf 401}, 911 (1999).
5287: 
5288: \bibitem{levy}  P. L{\'{e}}vy, \emph{Th{\'{e}}orie de l'addition des
5289: variables al{\'{e}}atoires\/} (Gauthier-Villars, Paris, 1954).
5290: 
5291: \bibitem{gnedenko}  B. V. Gnedenko and A. N. Kolmogorov \emph{Limit
5292: Distributions for Sums of Random Variables\/} (Addison--Wesley, Reading,
5293: 1954).
5294: 
5295: \bibitem{michael} M.~A.~Lomholt, T.~Ambj{\"o}rnsson, and R.~Metzler,
5296: \PRL \textbf{95}, 260603 (2005).
5297: 
5298: \bibitem{michael_long} M.~A.~Lomholt, T.~Ambj{\"o}rnsson, and R.~Metzler,
5299: in preparation
5300: 
5301: \bibitem{chechkin_bvp} A. V. Chechkin, R. Metzler, J. Klafter, V.~Y. Gonchar,
5302: and L.~V. Tanatarov, J. Phys. A {\bf 36}, L537 (2003).
5303: 
5304: \bibitem{smith}  D. E. Smith, S. J. Tans, S. B. Smith, S. Grimes,
5305: D. L. Anderson, and C. Bustamante, Nature \textbf{413}, 748 (2001).
5306: 
5307: \bibitem{zandi} R. Zandi and D. Reguera, Phys. Rev. E \textbf{}, (2005).
5308: 
5309: \bibitem{feiss} C. E. Catalano, D. Cue, and M. Feiss, Mol. Microbiol.
5310: \textbf{16}, 1075 (1995).
5311: 
5312: \bibitem{morita} H. Fujisawa and M. Morita, Genes to Cells \textbf{2},
5313: 537 (1997).
5314: 
5315: \bibitem{hud1} N. V. Hud and K. H. Downing, Proc. Natl. Acad. Sci. USA
5316: \textbf{98}, 14925 (2001).
5317: 
5318: \bibitem{hud} N. V. Hud, Biophys. J. \textbf{69}, 1355 (1995).
5319: 
5320: \bibitem{pack} M. E. Cerritelli, N. Cheng, A. H. Rosenberg, C. E.
5321: McPherson, F. P. Booy, and A. C. Steven, Cell \textbf{17}, 271 (1997).
5322: 
5323: \bibitem{kindt} J. Kindt, S. Tzlil, A. Ben-Shaul, and W. M. Gelbert,
5324: Proc. Natl. Acad. Sci. USA \textbf{98}, 13671 (2001).
5325: 
5326: \bibitem{kondev1} P. K. Purohit, M. M. Inamdar, P. D. Grayson, T. M. Squires,
5327: J. Kondev, and R. Phillips, Biophys. J. \textbf{88}, 851 (2005).
5328: 
5329: \bibitem{ali} I. Ali, D. Marenduzzo, and J. M. Yeomans, J. Chem. Phys.
5330: \textbf{121}, 8635 (2004).
5331: 
5332: \bibitem{schill} G. Schill, Catenanes, rotaxanes and knots,
5333: Academic Press, New York, 1971.
5334: 
5335: \bibitem{bishop} D. Bishop, P. Gammel, C. R. Giles, Phys. Today 54(10) (2001).
5336: 
5337: \bibitem{strick1} T. Strick, J.-F. Allemand, V. Croquette, D. Bensimon,
5338: Phys. Today 54(10) (2001) 46.
5339: 
5340: \bibitem{lehn} J.-M. Lehn, Supramolecular Chemistry, VCH, Weinheim, 1995.
5341: 
5342: \bibitem{dietrich} J.-P. Sauvage and C. Dietrich-Buchecker (editors),
5343: Molecular catenanes, rotaxanes, and knots: a journey
5344: through the world of molecular topology, Wiley-VCH, Weinheim, 1999.
5345: 
5346: \bibitem{blanco} M.-J. Blanco, M. C. Jim{\'e}nez, J.-C. Chambron, V. Heitz,
5347: M. Linke, J.-P. Sauvage, Chem. Soc. Rev. 28 (1999) 293.
5348: 
5349: \bibitem{jimenez} M. C. Jim{\'e}nez, C. Dietrich-Buchecker, J.-P. Sauvage,
5350: Angew. Chem. Int. Ed. 39 (2000) 3284.
5351: 
5352: \bibitem{raehm} L. Raehm, J.-M. Kern, J.-P. Sauvage, Chem. Eur. J. 5 (1999)
5353: 3310.
5354: 
5355: \bibitem{pease} A. R. Pease, J. O. Jeppesen, J. F. Stoddart, Y. Luo,
5356: C. P. Collier, J. R. Heath, Acc. Chem. Res. 34 (2001) 433.
5357: 
5358: \bibitem{lehn1} J.-M. Lehn, Chem. Eur. J. 6 (2000) 2097.
5359: 
5360: \bibitem{montemagno} C. Montemagno, G. Bachand, Nanotechnology 10 (1999) 225.
5361: 
5362: \bibitem{soong}
5363: R. K. Soong, G. D. Bachand, H. P. Neves, A. G. Olkhovets,
5364: H. G. Craighead, C. D. Montemagno, Science 290 (2000) 1555.
5365: 
5366: \bibitem{yurke} B. Yurke, A. J. Turberfield, A. P. Mills jr., F. C. Simmel,
5367: J. L. Neumann, Nature 406 (2000) 605.
5368: 
5369: \bibitem{metz} R. Metzler, Phys. Rev. E 63 (2001) 12103.
5370: 
5371: \bibitem{motors} A. Ajdari, F. J{\"u}licher, J. Prost, Rev. Mod. Phys.
5372: 69 (1998) 1269.
5373: 
5374: \bibitem{frey} A. Vilfan, E. Frey, F. Schwabl, M. Thormahlen, Y. H. Song,
5375: E. Mandelkow, J. Mol. Biol. 312 (2001) 1011.
5376: 
5377: \bibitem{functional} G. Bottari, F. Dehez, D. A. Lehigh, P. J. Nash, E. M.
5378: Perez, J. K. Y. Wong, and F. Zerbetto, Angew. Chemie Intl. Ed. \textbf{42},
5379: 5886 (2003).
5380: 
5381: \bibitem{functional1} S. Garaudee, S. Silvi, M. Venturi, A. Credi, A. H.
5382: Flodd, and J. F. Stoddart, Chem. Phys. Chem. \textbf{6}, 2145 (2005).
5383: 
5384: \bibitem{gittes} F. Gittes, E. Meyhofer, S. Baek, J. Howard, Biophys. J.
5385: 70 (1996) 418.
5386: 
5387: \bibitem{degennesm} P.-G. de Gennes, Comptes Rendus Acad. Sci. Paris S{\'e}r.
5388: II 324 (1997) 343.
5389: 
5390: \bibitem{thomsen} D. L. Thomsen, P. Keller, J. Naciri, R. Pink, H. Jeon,
5391: D. Shenoy, B. R. Ratna, Macromol. 34 (2001) 5868.
5392: 
5393: \bibitem{kulic_schiessel} I. M. Kulic, R. Thaokar, and H. Schiessel, Europhys.
5394: Lett. \textbf{72}, 527 (2005).
5395: 
5396: \bibitem{kulic_schiessel_jpc} I. M. Kulic, R. Thaokar, and H. Schiessel,
5397: J. Phys. Cond. Mat. \textbf{17}, S3965 (2005).
5398: 
5399: \bibitem{jarzinsky} C. Jarzinsky, Phys. Rev. Lett. \textbf{78}, 2690 (1997).
5400: 
5401: \bibitem{udo} O. Braun, A. Hanke, and U. Seifert, Phys. Rev. Lett. \textbf{93},
5402: 158105 (2004).
5403: 
5404: \bibitem{udo1} U. Seifert, Phys. Rev. Lett. \textbf{95}, 040602 (2005).
5405: 
5406: \bibitem{flory}  P.~J. Flory, {\em Principles of Polymer
5407: Chemistry\/} (Cornell University Press, Ithaca, New York, 1953).
5408: 
5409: \bibitem{orr} W.~J. Orr, Trans. Faraday Soc. {\bf 43}, 12 (1947).
5410: 
5411: \bibitem{ohno} K. Ohno and K. Binder, J. Phys.
5412: (Paris) {\bf 49}, 1329 (1988).
5413: 
5414: \bibitem{duplantier5} B. Duplantier and F. David, J. Stat.
5415: Phys. {\bf 51}, 327 (1988).
5416: 
5417: \bibitem{duplantier6} B. Duplantier and H. Saleur, Phys.
5418: Rev. Lett. {\bf 59}, 539 (1987).
5419: 
5420: \bibitem{grassberger} P. Grassberger and R. Hegger, Ann. Phys. {\bf 4}, 230
5421: (1995).
5422: 
5423: \end{thebibliography}
5424: 
5425: 
5426: \end{document}
5427: