1: \documentstyle[prl,floats,epsfig,aps]{revtex}
2: \begin{document}
3: \title{%
4: Theoretical models for single-molecule DNA and RNA experiments: \\
5: from elasticity to unzipping
6: }
7: %
8: \author{%
9: S. Cocco $^{\text{a}}$, J.F. Marko $^{\text{b}}$, R. Monasson $^{\text{c}}$
10: }
11: %
12: \address{%
13: \begin{itemize}\labelsep=2mm\leftskip=-5mm
14: \item[$^{\text{a}}$]
15: CNRS--Laboratoire de Dynamique
16: des Fluides Complexes, 3 rue de l'Universit{\'e}, 67000 Strasbourg, France;
17: \item[$^{\text{b}}$]
18: Department of Physics, The University of Illinois at Chicago,
19: 845 W. Taylor St., Chicago, IL 60607;
20: \item[$^{\text{c}}$]
21: CNRS--Laboratoire de Physique Th{\'e}orique de l'ENS, 24 rue
22: Lhomond, 75005 Paris, France.
23: \end{itemize}
24: }
25: %
26: \bibliographystyle{unsrt} % ,apalike
27: \maketitle
28: \thispagestyle{empty}
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: %%% Abstract %%%
31: %%%%%%%%%%%%%%%%%%
32: \begin{abstract}
33: We review statistical-mechanical theories
34: of single-molecule micromanipulation experiments on nucleic acids.
35: First, models for describing polymer elasticity are
36: introduced. We then review how these models are used to interpret
37: single-molecule force-extension experiments on single-stranded
38: and double-stranded DNA. Depending on the force and the molecules
39: used, both smooth elastic behaviors and abrupt structural transitions
40: are observed. Third, we show how combining the elasticity
41: of two single nucleic acid strands with a description of the base-pairing
42: interactions between them explains much of the phenomenology and kinetics of
43: RNA and DNA `unzipping' experiments.
44: \end{abstract}
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: %%% French title %%%
47: %%%%%%%%%%%%%%%%%%%%%%
48:
49:
50: \par\medskip\centerline{\rule{2cm}{0.2mm}}\medskip
51: \setcounter{section}{0}
52:
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: %%% Main text (in English) %%%
55: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56: %\tableofcontents
57:
58: \section{Introduction}
59:
60:
61: Single-molecule studies that provide information on properties of one
62: or a few interacting biomolecules are becoming increasingly
63: important in biophysics. The precision of control and quantitative
64: measurement, and simple interpretation of these experiments,
65: make detailed theoretical analyses appropriate.
66: The theory of single molecule micromanipulation
67: experiments is a new development of polymer physics, emphasizing the
68: structural richness of biopolymers (inhomogeneity of sequence,
69: sequence-specific monomer interactions, transformations of
70: secondary structure...). Both equilibrium and non-equilibrium
71: aspects of single-molecule experiments reveal new basic physical problems.
72:
73: This review presents some of the theoretical ideas that have been
74: useful for description of single-molecule micromanipulation studies of
75: nucleic acids.
76: First, models useful for describing biopolymer elasticity will be
77: presented. We will then review how these models are used to interpret
78: single-molecule DNA force-extension experiments, which show
79: both smooth elastic behaviors and abrupt structural transitions.
80: Third, we will show how combining the elasticity
81: of two single nucleic acid strands with a description of the base-pairing
82: interactions between them explains much of the phenomenology of
83: RNA and DNA `unzipping' experiments.
84:
85: The theoretical studies that we review use
86: a wide range of tools and concepts from statistical mechanics and
87: quantum mechanics. Single molecules are
88: composed of a large number of elementary units (monomers).
89: The nearest-neighbour character of the interactions between monomers
90: often leads to partition functions with the form of path integrals
91: (the curvilinear coordinate plays the role of time) which can be
92: analyzed using the tools of quantum mechanics.
93: Ideas from the theory of phase transitions are also extensively employed,
94: for example to describe the abrupt, first-order-like structural
95: changes frequently observed in stretching experiments. The kinetics
96: of such transitions are thus related to problems
97: from non-equilibrium statistical mechanics.
98:
99: \section{Theoretical models of flexible polymers}
100:
101: A number of polymer models have been used to model single-molecule
102: experiments. Here we focus on applications relevant to
103: double-stranded DNA, which is important biologically and also
104: nearly ideal as an object for theoretical study.
105:
106: \subsection{Gaussian Model}
107:
108: The simplest description of a polymer is the
109: Gaussian polymer (GP) model\cite{Flo89,Doi86}, which essentially considers a
110: polymer to be a series of particles joined by Hookean springs (Fig.~1).
111: Let us call ${\bf r}_n=(x_n,y_n,z_n)$
112: the location of monomer $n$.
113: The vector leading from monomer $n-1$ to monomer $n$,
114: ${\bf r}_n-{\bf r}_{n-1}$,
115: obeys a Gaussian distribution of average zero and variance
116: $ < ({\bf r}_ n- {\bf r}_{n-1})^2>=b^2,$
117: \begin{equation}
118: D({\bf r}_n-{\bf r}_{n-1})= \left(\frac{3}{2\pi b^2}
119: \right)^{3/2}\; \exp \left ( -
120: \frac{3\, {(\bf r}_n-{\bf r}_{n-1}) ^2}{2 b^2} \right)
121: \end{equation}
122: where the index $n$ runs from 0 to $N$.
123:
124: \begin{figure}[t]
125: \begin{center}
126: \psfig{figure=gc.eps,height=1.125cm,angle=0}\break
127: \hskip .25cm
128: \psfig{figure=fjc.eps,height=1.5cm,angle=0}\break
129: \hskip .25cm
130: \psfig{figure=wlc.eps,height=1.5cm,angle=0} \break
131: \hskip .05cm
132: \psfig{figure=wlrc.eps,height=1.05cm,angle=0} \break
133: \end{center}
134: \caption[]{Mathematical representations of polymeric chains.
135: GP: the Gaussian polymer is made of $N$ monomers represented
136: by harmonic springs. FJC: the Freely Jointed Chain is
137: composed of $N$ bonds of fixed length $b$, with no correlation
138: between the orientation of adjacent segments.
139: WLC: the Worm Like Chain is a continuous model, characterized by a
140: persistence length $A$; the orientation of the chain tangent ${\bf t}(s)$ is
141: changing appreciably over contour lengths greater than $A$.
142: WLRC: the Worm Like Rod Chain is described by a rotating three-dimensional
143: coordinate system, with local triedron (${\bf t}, {\bf n}_1, {\bf
144: n}_2$), along the curvilinear coordinate $s$.
145: \label{fjcfig}}
146: \end{figure}
147:
148: When submitted to a force $f$ along
149: the $z$ direction, the Hamiltonian thus has a Gaussian elastic term,
150: and a force-distance term,
151: \begin{equation}
152: H _{GP}
153: =\sum_{n=0}^{N} \frac{3\, k_B T}{2 b^2}\, ({\bf r}_n-{\bf r}_{n-1}) ^2 -
154: f \,(z_N -z_0)
155: \end{equation}
156: where the force is taken to act in the z-direction.
157: The statistics of the z-component of the end-to-end vector are
158: easily computed; for example
159: the average end-to-end distance as a function of the force is
160: \begin{equation}
161: \label{spring}
162: <z> _{GP}= \frac{N b^ 2}{3 k_B T}\; f
163: \end{equation}
164: The Gaussian polymer as a whole behaves as a Hookean spring
165: of zero rest length and stiffness $C= 3 k_B T/N b ^2$ proportional to the
166: temperature and inversely proportional to the length $N$.
167: This effective elasticity is a model for the entropic elasticity
168: resulting from the decrease in the number of a polymer's configurations
169: as is it extended. This basic picture of flexible polymer
170: elasticity is the basis of rubber elasticity and a starting point for
171: polymer physics\cite{Doi86}.
172:
173: \subsection{Freely-Jointed Chain model}
174: The GP has the unphysical feature that it can be indefinitely extended,
175: and is therefore useful only for weakly stretched polymers, and even
176: then only when physio-chemical details of the monomers are not of interest.
177: A model which corrects the indefinite extensibility problem but which
178: is still elementary and in wide use is
179: the Freely Jointed Chain model\cite{Flo89,Doi86}, which
180: %\cite{Smi92,Aus97,All97},
181: consists of $N$ bonds of fixed length $b$ (the Kuhn length),
182: allowed to point in any direction independently of each other
183: (see Fig~\ref{fjcfig}). When under zero force, the mean-squared
184: end-to-end distance is $R^2 = N b^2$, the familiar result for
185: a random walk of $N$ steps.
186:
187: When this chain is subjected to a force $f$, the bonds tends to align
188: along the force, as dipoles in an electric field, with an
189: energy
190: \begin{equation}
191: \label{fjc1}
192: H_{FJC}=-f\,b\, \sum _{n=1}^N \cos \theta_n
193: \end{equation}
194: where $\theta_n$ is the angle between the force and the $n^{th}$ bond
195: directions. Since the segment orientations are decoupled, the partition function
196: is easily calculated. The mean average end to end distance when a
197: force $f$ is applied is
198: \begin{equation}
199: \label{fjc2}
200: <z> _{FJC}= N\, b <\cos{\theta}>=
201: N\, b \left[ \coth
202: \left(\frac{f\,b}{k_B\,T}\right)-\frac{k_B\,T}{f\,b}\right]
203: \end{equation}
204: The small force behavior coincides with that of the GP expression (\ref{spring}).
205: However, (\ref{fjc2}) departs from the GP at large forces,
206: since the FJC model properly takes into account that
207: the extension of the molecule cannot exceed the total contour length $Nb$.
208: For large $f$, $<z> _{FJC}/[Nb] \approx 1 - k_B T/[b f]$.
209:
210: \subsection{Worm-Like Chain}
211: The Worm-Like Chain (WLC) \cite{Flo89} is a continuous
212: model without the sharp bends of the FJC. The chain is described
213: by its unit tangent vector ${\bf{t}(s)}$, as a function of contour
214: length $s$ along the chain. If no forces are applied, the tangent vector
215: is presumed to undergo Gaussian fluctuations with zero mean and
216: variance $<(d{\bf t}/ds)^2> = 1/A$ (Fig.~1). The energy
217: for this model in presence of a force $f$, is given by
218: \begin{equation}
219: \label{wlc}
220: H_{WLC}=k_BT \, \frac{A}{2}\int_0^{L} ds
221: \left ( \frac{\partial {\bf t}}{\partial s} \right)^2
222: -f \int_0^{L} \left ( {\bf z} . {\bf t}(s) \right) ds
223: \end{equation}
224: The first term is curvature energy that accounts for the resistance
225: of the chain to bending, and the second term is the stretching energy
226: due to application of the external force $f$.
227: The partition function of a chain of length $L$, with tangent vectors
228: at extremities ${\bf t}(s=0)= { \bf t}_0$,
229: ${\bf t}(s=L)= {\bf t}_1$, can be written as a path integral,
230: \begin{equation}
231: \label{wlcz}
232: Z(L,f,{\bf t}_0,{\bf t}_1)=\int D {\bf t}\; e^{\;-{H_{WLC}}/{k_BT}}
233: \end{equation}
234: The significance of the parameter $A$ is made clear
235: by the expression of average scalar product between
236: tangent vectors at coordinates
237: $s$ and $s'$ at zero force,
238: \begin{equation}
239: <{\bf t}(s) \cdot {\bf t}(s') >=\exp\,(\,-|s-s'|/A\,)\,
240: \end{equation}
241: Therefore $A$ represents the characteristic distance above which
242: tangent vectors decorrelate; $A$ is called the persistence length.
243:
244: For a long chain under zero tension the WLC mean-squared end-to-end distance
245: is $R^2 = 2 A L$ for $L>> A$ (the formula for general $L$ is often useful,
246: see \cite{Doi86}).
247: Therefore the unperturbed random coil properties of the WLC are equivalent
248: to those of the FJC and GP if we make the identification $2A=b$ and $L=Nb$.
249: The unstretched WLC on large scales becomes a random walk
250: of $N=L/(2A)$ steps each of length $b=2A$.
251:
252: From a physical point of view, the FJC represents
253: $N$ uncorrelated dipoles in an electrical
254: field, and the average orientation of one dipole at equilibrium is obtained
255: by classical statistical mechanics (eqns \ref{fjc1}, \ref{fjc2}).
256: By contrast, the WLC describes, the ``time'' evolution of a
257: dipole with moment of inertia $A$ in an electric field, with the role of
258: time played by the contour length coordinate $s$. The introduction
259: of the time dimension makes WLC equivalent to a quantum mechanical problem.
260: The Schr\"odinger equation for the associated
261: wave function can be analytically solved for small and large force limits, and
262: can be numerically solved for general force\cite{Bus94,Mar95}.
263:
264: At small forces $<< k_B T/A$ the Hookean behavior (\ref{spring}) is recovered
265: (i.e. with $b \rightarrow 2A$ and $N \rightarrow L/b$) while
266: for large forces $<z>_{WLC}/L\approx 1-\sqrt{k_B T/(4\,A\, f)}$.
267: \begin{figure}[t]
268: \begin{center}
269: \psfig{figure=forcext.eps,height=10cm,angle=-90}
270: %\psfig{introdurre la figura 1 di Bus[00] }
271: \caption[]{Force-extension curve for a $\lambda$-dsDNA with
272: equilibrium contour length $L=16.4 \mu$m in 10mM $Na^+$ buffer.
273: Experimental data (x) are from [24] and (+) from [14].
274: The plateau at $f\simeq 70$ pN indicates the cooperative
275: transition to SDNA. The extensible WLC model with persistence
276: length $A=53$ nm, Young modulus $\gamma \approx 1000$ pN, and
277: two states dsDNA-SDNA reproduces accurately the experimental
278: behavior [23]. Marko and Siggia's interpolation formula [3]
279: (\ref{mark2}) is very accurate up to forces of 10 pN.
280: Predictions from the GP and FJC ($b=2A=106$ nm) models
281: are plotted in the Inset.
282: \label{forzaesfig}}
283: \end{center}
284: \end{figure}
285: \section{Elasticity of Double- and Single-stranded DNAs:
286: experiments and theory}
287:
288: \subsection{Double-stranded DNA (dsDNA) under tension }
289: \label{tensionds}
290:
291: Reviews of experiments and theory on dsDNA elasticity can be
292: found in \cite{Aus97,Bus00,Wan97,Str00}.
293: In Fig.~2 we report the force extension curve of a single dsDNA.
294: Experimental data obtained by \cite{Clu96,Str98}
295: for $\lambda$-DNA of length 48502 bp, or $L=16.4$ microns are shown,
296: with fits of the GP, FJC, and
297: WLC models with $A=b/2= 53$~nm in 10 mM $Na^{+}$ buffer.
298: For dsDNA, the Kuhn length $b=106$~nm is much larger than the
299: natural base pair spacing of 3.4 $\AA$.
300: dsDNA is not naturally described by the FJC model because
301: consecutive bases, stacked onto each other, are not free to reorient
302: independently of each other. The success of the WLC shows that
303: dsDNA behaves as a semiflexible polymer, with a bending modulus $A\; k_BT$.
304:
305: Several analytical interpolation formulae for the WLC, and modifications
306: of the FJC introduced to fit accurately the data are discussed and compared
307: in \cite{Wan97,Bou99}.
308: Marko and Siggia proposed a simple interpolation formula that
309: is close to the exact numerical solution of the force-extension response
310: of the WLC\cite{Mar95,Bus94},
311: \begin{equation}\label{mark2}
312: f _{WLC} (z)\simeq \frac{k_B T}{ A}\left[\frac{1}{4\,(1-z/L)^2}-\frac
313: 14+\frac{z}{L}\right]
314: \end{equation}
315: This expression reduces to the exact solution as either $z \to 0$
316: or $z \to L$, but differs from the exact solution by up to $\approx 10 \%$
317: near $f=0.1$ pN (Fig~2). Bouchiat {\em et al.} \cite{Bou99} have introduced
318: correction terms to eqn (\ref{mark2}), in the form of
319: a seventh order polynomial in $z/L$. The resulting approximation for
320: $f _{WLC}(z)$ is accurate to 0.1\%.
321: According to formula (\ref{spring}), a force $f_0 = 3 k_B T/b$
322: is required to extend dsDNA by a fraction of its contour length;
323: from $b\simeq 100$ nm we see that the characteristic force associated
324: with the entropic elasticity of dsDNA is $f_0 \approx 0.1$ pN.
325:
326: Fig~2B shows that experimental data are well fitted by
327: the FJC model for forces $f<0.1$ pN, and by the WLC
328: up to $f<5$ pN.
329: Various experiments analyzed in terms of the WLC
330: give $A= 50 \pm 5 $ nm in 10 mM
331: $Na^{+}$ \cite{Smi96,Wan97}. The persistence length
332: of DNA is reduced in high salt concentrations by electrostatic
333: screening of the repulsive charge
334: along the backbone; electrostatic effects have been
335: taken into account in the WLC model by Barrat and Joanny
336: through Debye-Huckel interactions \cite{Bar93,Mar95,Bau97,Wan97}.
337:
338: Fitting larger-force experimental data demands the introduction
339: of the stretching elastic modulus of the molecule,
340: $\gamma \simeq 1000 \pm 200$ pN, quantitatively
341: consistent with the relation between the bending modulus, $A\, k_B T$
342: and the Young modulus $Y=\gamma/( \pi R^2)\simeq 300$ MPa
343: ($R=10$ $\AA$ is the double helix radius) for an elastic rod,
344: \begin{equation}
345: A =\frac{\pi}{4\, k_BT} Y \;R^4\equiv \frac{\gamma \, R^2} {4 k_B T}
346: \end{equation}
347: The value of the elastic modulus of DNA indicates that thermal
348: fluctuations of the axial base pair distance $h$ of the order of
349: $ <(h -<h>)^2>\simeq k_B T/(\gamma <h>) \simeq 0.14 \AA$.
350: This order of magnitude is in agreement with molecular
351: dynamics simulations, providing a consistent picture of the elasticity
352: at the atomic and mesoscopic scale\cite{Mar03}.
353: Axial vibrational modes have been
354: studied in \cite{Coc99b} and compared to Raman spectroscopy measurements.
355:
356: When dsDNA molecule is subjected to a force
357: of $f=65$ pN it undergoes a structural transition to another conformation,
358: S-DNA, with a contour length $1.7$ times larger than its B-DNA
359: counterpart\cite{Smi92,Smi96,Clu96}.
360: Numerical investigations of the structure of S-DNA
361: have been performed by Lavery and collaborators\cite{Clu96,Leb96}.
362: The force plateau around $f=65$ pN corresponds to a highly cooperative
363: transition, reminiscent of a first order phase transition.
364: A two state model proposed by Cluzel
365: {\em et al.}\cite{Clu96,Ciz97}, inspired from models introduced in the
366: context of thermally-induced denaturation~\cite{Hil59,Zim59,Kafri},
367: is able to reproduce the B-to-S transition. A recent study has
368: suggested that the extended S state is actually strand-separated
369: with the S phase described as stretched ssDNAs
370: \cite{Rouzina}.
371:
372: In the simplest two-state model of the B to S transition,
373: the molecule is described as a chain of $N$ elements (base pairs), which
374: can be in states B (energy $E_B$, length $l_B$) or S (energy
375: $E_S$, length $l_S>l_B$). Each element is associated a
376: spin variable, $s=1$ (respectively $-1$) for the
377: B (resp. S) state. The energy of the chain reads,
378: \begin{equation}
379: \label{is}
380: H=\sum_{i=1}^{N}(E_{s _i}-f\,l_{s _i})+
381: \frac{\omega}{2} \sum_{i=1}^{N-1} (1-s_i\,s_{i+1})
382: \end{equation}
383: $\omega$ represents a 'domain wall' energetic cost of a B-S
384: frontier. Up to an additive constant,
385: % $N=n_S+n_B$ and $\sum_i s_i=n_S-n_B$ we obtain
386: \begin{equation}
387: \label{ising}
388: H=-\frac{\omega}{2} \sum_{i=1}^{N-1} s_i s_{i+1}
389: -\frac 1 2 (\Delta E -f \,\Delta l) \sum_{i=1}^{N} s_i
390: \end{equation}
391: where $\Delta E=E_S-E_B$ and $\Delta l= l_S-l_B.$
392: Notice that eqn (\ref{ising}) is simply the Hamiltonian of a
393: one-dimensional Ising model with magnetic field
394: $h=\Delta E -f \, \Delta l $.
395: The extension is obtained from the derivative of the free energy with
396: respect to the force $f$. Comparison with experiments allows
397: to determine quantitatively the domain wall energy, $\omega\simeq 4 k_BT$
398: \cite{Ciz97}.
399: The extensible WLC including nonlinearities which define two states
400: of extension provides a way to fit the force-extension curve
401: over a wide range of forces $0.01<f<100$ pN \cite{Mar98}.
402:
403: \subsection{Supercoiled DNA under tension}
404: \begin{figure}[t]
405: \begin{center}
406: \psfig{figure=plec.eps,height=6cm,angle=0}
407: \end{center}
408: \caption[]{Phase coexistence in a supercoiled DNA under tension.
409: A fraction $x$ of the length is plectonemic supercoil
410: with radius $R$ and pitch $P$, while the remaining fraction $1-x$ is in
411: an extended conformation.
412: \label{plecfig}}
413: \end{figure}
414:
415:
416: dsDNA differs from simpler polymers because it exhibits torsional
417: and bending stiffness. Try to impose a twist to an elastic rod while keeping
418: it extended and fixed at one end. Then, if you relieve the tension,
419: an interwound structure called a plectoneme will appear, Fig~\ref{plecfig}
420: (twisted telephone cords often form plectonemic supercoils).
421: Similarly, dsDNAs under sufficient torsional stress interwinds to
422: form plectonemic supercoils. Formally, the over- or underwinding of
423: DNA is measured by the change in double-helical linking number.
424: This is often expressed as $\sigma$, the fractional change
425: in linking number relative to that of relaxed dsDNA
426: (one right-handed, or positive link per 10.5 base pairs).
427: Supercoiling of DNA is of tremendous importance to eubacteria.
428: For example, in E. coli all the DNA is held under torsional stress
429: and is topologically constrained with $\sigma \approx -0.06$.
430:
431: The elasticity of a single supercoiled DNA molecule has been
432: experimentally measured by Strick {\em et al.}
433: \cite{Str96,Str98} and by
434: L\'eger {\em et al.} \cite{Leg99}. A rich behavior was observed.
435: At small forces the molecule responds to increasing
436: positive or negative supercoiling by first having its conformations
437: slightly chirally perturbed, and then
438: by forming plectonemes with appreciable shortening of its length. At
439: forces $f > 0.5$ pN, negative supercoiling is released
440: through the opening of the double helix into denaturation bubbles.
441: At forces $f>3$ pN, positive supercoiling induces the formation
442: of regions exhibiting a new conformation called P-DNA.
443: The structure of P-DNA has been deduced by molecular modeling\cite{All98};
444: it is essentially characterized by its exposed bases. P-DNA can
445: be thought of as two tightly interwound ssDNAs.
446:
447: The theory of stretched supercoiled DNA was initiated by
448: Marko and Siggia \cite{Mar95b,Mar97}, who
449: considered phase coexistence
450: of linear, plectonemic, and denatured DNA in different regions
451: of a supercoiled molecule. The relative extensions of these portions are
452: determined by the degree of supercoiling $\sigma$ and the stretching
453: force $f$.
454:
455: At small forces, a fraction $x$ of the molecule is in the
456: plectonemic ($p$) regime, whereas the remaining $1-x$ fraction is
457: extended ($s$) (Fig~\ref{plecfig}). The free energy per unit of length
458: ${\cal F}=F/L$ reads,
459: \begin{equation}
460: {\cal F} (\sigma, z/L) =x\; {\cal F}_p (\sigma_p)+ (1-x)
461: \, {\cal F}_s(f,\sigma_s)
462: \end{equation}
463: $\sigma= \Delta L_k/L^0_k$ is the excess of density of supercoiling with
464: respect to the rest configuration
465: (dsDNA making a double helix turn in 10.4 bases,
466: $L^0_k =L/10.4$); $\sigma$ is
467: partitioned into extended and plectonemic regions:
468: $\sigma = x\, \sigma_s+ (1-x) \sigma_p$.
469: The free energy of the extended phase equals
470: the WLC free energy plus the twisting energy,
471: ${\cal F}_s(f,\sigma_s)=
472: {\cal F}_{WLC}(f)+k_BT\, C\, L_k^2 /(2 L^2)\, (2\pi\sigma_s)^ 2$.
473: $C$ is the twist persistence length known from experiment to be
474: $C=75 \pm 30$ nm.
475:
476: The free energy of the plectonemes is thus considered to be the sum of elastic
477: (bending and twisting), electrostatic and entropic contributions,
478: minimized over plectonemic parameters e.g. the pitch $P$ and radius $R$
479: (Fig~\ref{plecfig}).
480: The entropic term represents the
481: entropy lost by confining the DNA in the superhelix formation.
482: The total free energy is obtained by minimization with respect to the
483: plectonemic portion $x.$
484: At large forces the structural transition to denatured DNA
485: is also included in the model by allowing the plectonemic phase
486: to be a mixture of denatured and normal plectonemic DNA.
487: The theoretical force-extension curve at
488: fixed supercoiling reproduces the experimental behavior \cite{Mar97,Str96}.
489:
490: A semi-microscopic model has been used to describe thermal-
491: and torque-induced denaturation in one phase diagram\cite{Coc99a}.
492: This work described the formation of denaturation bubbles when DNA
493: is stretched at $f >0.5$ pN and negatively supercoiled.
494: The critical torque at room temperature $\Gamma \approx -2k_BT $ is in
495: good agreement with the value inferred from the experiments by Strick
496: {\em et al.} \cite{Str98}.
497:
498: A generalization of the two-state Ising description (\ref{ising})
499: of the overstretching transition
500: has been introduced by Sarkar {\em et al.} \cite{Sar01} to model the
501: structural transition of a twisted and stretched DNA molecule
502: observed in \cite{Leg99}.
503: For each site five possible state are introduced:
504: dsDNA, S-DNA, P-DNA, sc-PDNA (a supercoiled P state), and a left
505: handed double helix Z-DNA. This last state, with a supercoiling
506: degree $\sigma_Z = -1.3$, is proposed as an alternative to denatured ssDNA
507: ($\sigma =-1$).
508: A force-torque diagram is derived that agrees with the experiments on the
509: critical unwinding torque at zero force, $\simeq -2 k_B T$, and the
510: torque to drive DNA into the P structure, $\simeq 7k_B T.$
511:
512: The elasticity of supercoiled DNA has also been studied at a more
513: microscopic level. Twisting and bending deformations can be
514: described by extending the WLC to include description of base-pair
515: orientation using a triad of unit vectors (WLRC) (Fig~1) \cite{Fai97}.
516: %parametrized by three Euler angles
517: Moroz and Nelson \cite{Mor97}, and Bouchiat and
518: M\'ezard\cite{Bou98} have written the partition function of this model
519: as a path integral in the space of the Euler angles
520: parametrizing the orientations, limiting the
521: integration measure to the paths with a fixed linking number $L_k$.
522: The free energy is obtained from the ground state energy of a Schr\"odinger
523: equation describing a particle moving on a unit sphere in the presence of
524: electric and magnetic fields.
525:
526: The WLRC model does not take into account self avoidance: the WLRC chain
527: is a phantom chain that can pass through itself. The result is that
528: linking number fluctuations are not well-defined in the continuum limit.
529: This problem is reflected in a divergence of the ground state energy.
530: This is strictly a technical problem since real DNA has self-avoidance
531: interactions.
532: To avoid this problem, Moroz and Nelson considered the
533: unambiguous infinite force situation (fully stretched molecule),
534: and obtained finite force results by means of perturbation theory.
535: Bouchiat and Mezard have introduced a short distance
536: cutoff (discretization of the chain) to suppress the singularity.
537: Fitting the theory to experimental data \cite{Str96},
538: the twist persistence length $C$ can be determined but is largely dependent
539: on the theoretical scheme followed: $C=120$ nm is obtained by Moroz and Nelson,
540: while $C= 82$ nm is obtained by
541: Bouchiat and Mezard for a cutoff length of $7$ nm.
542:
543: \subsection{Single-stranded DNA under tension}
544:
545: Single-stranded DNA (ssDNA) is more flexible and can reach a larger extension
546: per base pair than dsDNA.
547: A sensible simple model of ssDNA is, at first sight, a FJC with
548: a Kuhn length equal to the sugar-phosphate monomer backbone length $b=7 \AA$.
549: However the ssDNA elasticity is complicated by
550: nucleotide interactions, and as a result simple polymer models
551: do not describe ssDNA elasticity over a wide range of forces.
552:
553: \begin{figure}
554: \begin{center}
555: \psfig{figure=ssforcext.eps,height=8cm,angle=-90}
556: \end{center}
557: \vskip .5cm
558: \caption{Top: Force extension curve for a $\lambda$-ssDNA with
559: equilibrium contour length $L_{ss}=l_{ss} N \equiv 0.56\, nm \times
560: 48502\,bp=27 \mu m$ in 150 mM Na$^ +$. Experimental data (+),
561: dashed line: FJCL with $b=1.5 nm$ , dotted line: extensible
562: FJCL with $\gamma=800\, pN$, from[11].
563: Bottom: Force extension curve for a Charomid ssDNA in 10 mM
564: phosphate buffer, 5 mM Mg$^{++}$ buffer,
565: data are from [38], the fits are with the FJCL with $b=1.9\,
566: nm$ and $\gamma=800\, pN$ (dotted line) and with the hairpin model
567: (full line) of [41]. }
568: \label{ssfext}
569: \end{figure}
570:
571:
572:
573: For forces $f< 20$ pN the experimental force extension curve for a
574: 48502 base $\lambda$ ss-DNA in 150 mM $Na^{+}$ has been fitted
575: with a FJC--like
576: (FJCL) model by Smith {\em et al.} \cite{Smi96} with two effective
577: parameters: a Kuhn length $d=15 \AA$~ and a contour length
578: per base pair $l_{ss}=5.6 \AA$~ that differs from the backbone
579: distance (see Fig.~4).
580: Note that due to the higher flexibility, the characteristic entropic force
581: of the single strand, $f_0= 3 k_B T/ \sqrt{b l_{ss}} \approx 10$ pN
582: (\ref{spring}), is much higher than for dsDNA.
583: At forces $f>15$ pN the fit requires the introduction of a stretching modulus
584: $\gamma=800$ pN (Section~\ref{tensionds}).
585:
586: ssDNA elasticity depends strongly on salt concentration.
587: At low salt (1 mM Na$^+$, self avoidance interactions
588: due to electrostatic self-repulsion along the charged sugar-phosphate
589: backbone occurs.
590: The experimentally observed logarithmic-like dependence of the extension
591: upon force is well reproduced by Monte Carlo simulations \cite{Des03,Zha01}.
592: Electrostatic selfavoiding effects can be taken analytically into account
593: using Barrat and Joanny formalism \cite{Coc03}, or with
594: a Hartree-Fock calculation from the WLC models \cite{Bou03}.
595:
596: At higher salt concentration ($>$ 100 mM Na$^+$, or in presence of Mg$^{++}$ )
597: formation of secondary structure (`hairpins') by hydrogen
598: bonding between complementary bases on the same strand strongly influences
599: elastic properties. Experiments show that the force--extension behavior
600: curve depends on the strand GC vs. AT content, and can be modulated
601: using denaturing chemical agents that suppress hydrogen bonding
602: \cite{Mai00,Des03}.
603: A theoretical analysis of the elasticity of a polymer with hairpin secondary
604: structure has been developed by Montanari and Mezard \cite{Mon01}.
605: Conformations of hairpins taken into account are such that any two
606: pairs of paired bases ($i<j$, $k<l$) are independent
607: ($i<j<k<l$), or nested ($i<k<l<j$) \cite{Bun99,Pag00}.
608: This representation do not include pseudoknots \cite{Isa00} but
609: leads to a solvable
610: recursion relation relating the partition functions of successively larger sequences
611: \cite{Bun99}, under the simplifying hypothesis that any two bases e.g.
612: AT, GC, AG, ... have the same pairing free-energy. For an infinite molecule,
613: a phase transition takes
614: place between a folded (zero extension, $f<f_s$) and an extended
615: phase ($f>f_s$) with $f_s$ of the order of 1 pN, given reasonable choices
616: of parameters (see Fig.4).
617:
618: \subsection{Elasticity of DNA in presence of DNA-folding proteins.}
619: In eukaryote cells, DNA is wrapped around octamers of histone
620: proteins to form a more compact structure called a nucleosome.
621: The long chromosomal DNAs of eukaryote cells
622: are thus organized into long strings of nucleosomes, or
623: `chromatin fibers'. A single eukaryote chromosome may contain
624: more than $10^8$ base pairs of DNA and roughly $10^6$ nucleosomes.
625:
626: The elasticity of chromatin fiber has been experimentally studied by Cui
627: and Bustamante \cite{Wan98}. The experimental curves can be fitted
628: with polymer models composed of units which,
629: independently of each other, can be in a
630: folded (short) or unfolded (long) state \cite{Mar97a,Coc03}.
631: These states are taken to correspond to stacked and unstacked nucleosomes.
632: The elastic response of whole mitotic chromosomes can be related back
633: to this fiber elastic response\cite{Coc03}.
634:
635: Very roughly, models of DNA folding by proteins will generally show
636: a characteristic force at which the proteins will dissociate in
637: equilibrium\cite{Mar97a}.
638: Given a free energy difference between the folded and unfolded
639: states of $g$ per fold, and given an end-to-end length reduction
640: of $d$, this characteristic force will be about $g/d$. Note that
641: for large values of $d$ (e.g. by formation of large DNA loops,
642: a common feature of gene-regulatory proteins) this implies
643: low on-off equilibrium forces. It must be kept in
644: mind that if the enthalpic component of the binding free energy
645: is large, there may be large barriers for such loops to open
646: and close, making equilibrium difficult to reach.
647: Such situations should show theoretically interesting
648: many-body thermal barrier-crossing kinetic phenomena.
649:
650: \section{DNA and RNA unzipping}
651:
652: Essevaz-Roulet, Bockelman and Heslot
653: have shown that the two strands of a dsDNA
654: can be pulled apart by a force $\approx 12$ pN
655: \cite{Ess97} (Fig.~\ref{unzfig}).
656: Variations of the `unzipping' force
657: correspond to the DNA sequence, through the known relationship between
658: DNA sequence and base-pair interaction strengths\cite{Bre86}.
659: Higher forces were shown to correspond to DNA regions with
660: higher GC densities, which in general have
661: stronger base-pairing interactions than AT-rich sequences.
662: Experimental unzipping force traces show a series of sawtooth signals
663: attributed to stick-slip motion, with the sticking generated by DNA regions
664: with higher GC content.
665: This kind of experiment amounts to 'feeling' DNA sequence.
666:
667: Current techniques are able to observe GC-content over long
668: stretches of DNA ($\approx 10$ kb) with about 10 base pair resolution.
669: It should be noted that it has been demonstrated that unzipping
670: is sensitive to at least some single-base substitutions\cite{Boc02}.
671: We now discuss some equilibrium and dynamical aspects
672: of DNA and RNA unzipping.
673:
674: %figure:
675: \begin{figure}[t]
676: \begin{center}
677: \psfig{figure=unz.eps,height=8cm,angle=0}
678: \caption[]{Sketch of a DNA molecule with $n$ base pairs
679: unzipped, as a result of a mechanical stress (applied force $f$).
680: The distance between the two ssDNA ends is defined to be $2 r$.
681: \label{unzfig}}
682: \end{center}
683: \end{figure}
684:
685: \subsection{Thermodynamics of DNA-RNA unzipping}
686:
687: Control parameters for unzipping vary from experiment to experiment.
688: Roughly speaking, either the force or the distance between strand
689: extremities may be kept fixed (Fig.~\ref{unzfig}).
690:
691: \subsubsection{Fixed pulling force}
692:
693: %figura del paesaggio energetico per P5ab.
694: %figura del paesaggio energetico dell' unzipping fig. 20 pag 1550 boc02
695: With a fixed force $f$ on the
696: molecule ends, the free energy $G$ of the molecule
697: with $n$ base pairs opened is the difference between the free energy of
698: the two extended single strands of $n$ bases each,
699: and the free energy lost in unpairing the $n$ first base pairs ($i=1, \ldots ,n$):
700: \begin{equation}
701: \label{unz1}
702: G (f,n)=2\; n\, {\cal F}_{ss} (f) -\sum_{i=1}^{n}\; g_{ds} (i)
703: \end{equation}
704: As discussed above, the free energy per base pair of stretched ssDNA,
705: ${\cal F}_{ss} (f)$, can be expressed using the FJCL model for forces
706: $f \simeq 10\,$ pN \cite{Smi92}. At this
707: high tension, nucleotide hairpin-formation effects are absent.
708: Quadratic expansion of ${\cal F}_{ss} (f)$ around $f=10\,$ pN
709: gives the free energy of a Gaussian polymer,
710: ${\cal F}^{GP}_{ss} (f) =- f^2 b^2 /(6 k_B T) $
711: with an effective Kuhn length $b=7 \AA$.
712:
713: We start by considering an homogeneous sequence, where all base pairs
714: have pairing free energy $g_{ds}= - g_0$.
715: The unzipping critical force $f_u$ is simply given by the condition
716: \begin{equation}
717: \label{fc}
718: G(f,n)\equiv n\; g(f)= [2\, {\cal F}_{ss}(f)+ g_0] n = 0
719: \end{equation}
720: For $f < f_u$, dsDNA is stable, and if $f > f_u$, the double helix unzips
721: as in a first order phase transition.
722: Using the Gaussian model for the ssDNA, we obtain
723: $f_u^{GP}=\sqrt{3 k_B T g_0}/b$.
724: The unzipping force of a homogeneous sequence is therefore directly
725: related to the pairing free energy.
726:
727: Rief {\it et al} measured the unzipping forces $f_u$ for DNAss of
728: various repeated sequences\cite{Rie99}. It was found that
729: $f_u$(poly~dA- dT)=9$\pm 3$~pN and $f_u$({poly~dG-dC})=20$\pm 3$~pN, giving
730: $g_0^{GC}({A-T})=0.8\; k_B T$,$\ g_0^{FJCL}({A-T})=1.1\; k_B T$
731: $g_0^{GC}({G-C})=4.2\; k_B T$, and
732: $g_0^{FJCL} ({G-C})=3.5\; k_B T$ respectively.
733: These values of the free energies of denaturation are compatible
734: with thermodynamic data based on DNA melting\cite{Bre86}.
735: It is to be noted that unzipping experiments give the only direct
736: measurement of the relative free energies of ss and dsDNA at equal
737: temperatures.
738:
739: Unzipping has been
740: discussed in the language of continuous phase transitions
741: by Lubensky and Nelson \cite{Lub00,Lub02}.
742: The unzipping free energy per base pair, $g(f)$, vanishes as $f_u- f$
743: with a discontinuous slope, as in a first order phase
744: transition. However (as in interfacial wetting)
745: as the unzipping transition is approached from below {\em i.e.} $f \to f_u ^-$,
746: the molecule the average number $<n>$ of open base pairs
747: undergoes a continuous power-law
748: divergence. From the probability to have $n$ open base pairs,
749: $P(n)= {g(f)}/({k_B T})\; \exp[- (n\, g(f))/({k_BT})]$,
750: one obtains
751: \begin{equation}
752: <n>=\frac{ k_B T}{g(f)} \simeq \frac 1 {f_u -f}
753: \end{equation}
754:
755: \begin{figure}[t]
756: \begin{center}
757: \psfig{figure=df1.eps,height=7cm,angle=-90} \break
758: %\psfig{figure=p5ab.eps,height=4cm,angle=90}
759: \end{center}
760: \caption[]{DNA unzipping phase diagram as a function of torque
761: $\Gamma$ in units of $k_B T$ and force $f$ in pN, from [54].
762: The solid line shows
763: the results for the FJCL model of ssDNA elasticity, while the dashed
764: line shows the result within the Gaussian approximation. At zero torque,
765: the unzipping force is $f_u \simeq 12$ pN; positive torque increase $f_u$,
766: negative torques reduce $f_u$ until it vanishes at $\Gamma= -2.4 k_B T$. }
767: \label{phasediagfig}
768: \end{figure}
769:
770: When the DNA molecule is subjected to a torque $\Gamma$,
771: a torque-angle work contribution occurs in the unzipping free energy,
772: \begin{equation}
773: \label{fct}
774: g (f,\Gamma)= 2 {\cal F}_{ss}(f) +g_0 - \Theta_0\Gamma \qquad .
775: \end{equation}
776: $\Theta_0=2\pi/10.4$ is the change in twist angle during conversion of
777: dsDNA to separated strands\cite{Coc01}.
778: The phase diagram for the unzipping transition in the force, torque
779: plane is shown in Fig.~\ref{phasediagfig}.
780:
781: \begin{figure}[t]
782: \begin{center}
783: \psfig{figure=gfeqP5ab.eps,height=6cm,angle=-90} \break
784: \psfig{figure=p5ab.eps,height=5.7cm,angle=90} \break
785: \psfig{figure=bock.eps,height=5cm,angle=0}
786: %\psfig{figure=lambda2.eps,height=4cm,angle=90}
787: \caption[]{Top: Free energy (in $k_B T$) and probability
788: distribution of the opening fork at the critical force
789: $f=15$ pN for the P5ab molecule (middle) from [59].
790: Bottom: Schematic representation of the free energy landscape
791: for a displacement of 4970 nm (left, corresponding to slip phase in the
792: force) and of 5050 nm (right, corresponding to the stick phase in the
793: force) during the unzipping of a $\lambda$-DNA (from [58]).
794: \label{p5abfig}}
795: \end{center}
796: \end{figure}
797:
798: Along heterogeneous sequences, the free energy to open the first $n$ base
799: pairs, $G(f,n)$ (\ref{unz1}), can be calculated using the sequence dependent
800: pairing free energy $g_{ds}(i)$ (e.g. from the Mfold server \cite{Zuk00}).
801: In Fig.~\ref{p5abfig}, we show $G(f,n)$ for the RNA molecule called
802: P5ab, mechanically unzipped by Liphardt {\em et al.} \cite{Lip01} with
803: a force maintained fixed at the extremity through a feedback mechanism.
804: The critical force is defined by the condition
805: that the closed and open state have equal minimal free energies.
806: Contrary to the homogeneous case,
807: the free energy landscape at the critical force is not flat.
808: It is characterized by high energy barriers $G^*\approx 10 k_B T$.
809: The probability to have $n$ open base pairs is essentially zero
810: if $n$ differs from the open and the closed configuration
811: (Fig~\ref{p5abfig}). Indeed, experiments show \cite{Lip01} that
812: at the critical force the molecule essentially hops between open and
813: closed configurations.
814:
815: Lubensky and Nelson \cite{Lub00,Lub02} have shown that
816: the critical behavior around $f_u$ changes for large random sequence
817: with respect to homogeneous sequences. Instead of the divergence
818: $1/(f_u-f)$ for the averaged number $<n>$ of base pairs, a stronger
819: singularity $<n> \approx 1/(f_u-f)^2$ appears.
820:
821: \subsubsection{Fixed distance between extremities}
822: \label{refutile}
823: If the ssDNA ends are held apart at some
824: distance $2r$ (Fig. \ref{unzfig}), some average number of bases $n$ will open.
825: In the ideal case of
826: a rigid opening device, the free energy cost to open $n$ base pairs
827: is a sum of chain stretching and denaturation contributions,
828: \begin{equation}
829: \label{fixdis}
830: H_{d} (r,n)=
831: W_{ss}(2r,2n) -\sum_{i=1}^{n} g_{ds}(i)
832: \end{equation}
833: $W_{ss}(2r,2n)$ is the work done by the force to stretch $2n$ base pairs
834: of ssDNA at a distance $2r$. For simplicity we consider
835: the GP free energy (\ref{spring}),
836: considering as in the previous section the effective Kuhn length $b=7 \AA$~,
837: from the interpolation formula (\ref{mark2})
838: \cite{Tho95}, or from numerical inversion of the FJC extension
839: vs. force (\ref{fjc2})~\cite{Boc98}.
840: %The partition function and the total free energy
841: %are obtained as usual by the sum on n of the Gibbs weights
842: %$P(2n,2r) =\exp [- H_{d}(r,n)/k_B T]$.
843: Thus, unzipping at fixed extension can bedescribed
844: using
845: \begin{equation}\label{gcea}
846: W^{GP}_{ss}(2r,2n)= 3 k_B T \; \frac{r^2} {n\; b^2}
847: \end{equation}
848:
849: The most probable value of the number of opened base pairs $\bar n$ is
850: obtained by minimization of the free energy (\ref{fixdis},\ref{gcea})
851: with respect to $n$.
852: The number of unzipped based pairs is found to scale linearly with
853: the distance, $\bar n (r)= r/d_u $
854: where $d_u=\sqrt{g_0/3 k_B T } d =5 \AA$~
855: is the projection of the monomer length along the force direction.
856: The resulting free energy reads
857: \begin{equation}
858: {\cal F}(r)=2\,\sqrt {3 k_B T g_0}\;\, r/b= 2\, \bar n (r)\; g_0
859: \end{equation}
860: The tension $\bar f$ in the chain is simply the derivative of ${\cal F}$
861: with respect to $2r$,
862: $\bar{f}= \sqrt { 3 k_B T g_0}/b = 12$ pN.
863: This simple calculation shows that as unzipping proceeds quasi-statically,
864: the ssDNA tension is a constant, just $f_u$.
865: Note that the excess free energy per unpaired base for fixed extension
866: is double the free energy of denaturation because the
867: work done extending the ssDNAs adds to the work done when opening the
868: molecule. This indicates a strategy to determine $g_{ds}$
869: unambiguously from unzipping force data.
870:
871: The analysis of the fluctuations around the minimum free energy gives
872: $\left< f -f_u \right> = O(1/r ^2) $. Note that in the
873: constant--force ensemble result, $f -f_u \approx 1/<r>$ \cite{Lub02}.
874: The fixed-distance and fixed-force ensembles are equivalent only in the
875: thermodynamic limit $r \to \infty$.
876:
877: The unzipping force at small distance between extremities is
878: sensitive to the detailed structure of the pairing potential
879: as a function of the interbase distance. The semimicroscopic model
880: introduced in \cite{Coc01,Coc02} predicts a force barrier of $\simeq
881: 300$~pN at a distance $2r \simeq 21 \AA$~,
882: at which the hydrogen bond are broken but the bases are still stacked in
883: the double helix configuration. This force barrier might be directly
884: observable in an AFM experiment.
885:
886: To take into account the experimental apparatus,
887: Bockelmann {\em et al.} have included in their theory
888: the effects of the two dsDNA linker arms
889: of $N_{ds}$ base pairs and extension $r_{ds}$),
890: and the cantilever stiffness \cite{Boc98,Boc02}.
891: The free energy (\ref{fixdis}) is then
892: \begin{equation}
893: H_d^c (r,r_{ss},r_{ds},n)=
894: W_{ss}(2 r_{ss}, 2n)+W_{ds}(r_{ds},N_{ds}) -\sum_{i=1}^{n} g_{ds}(i) -
895: k_{lever} (r-2r_{ss}-r_{ds}) \ .
896: \end{equation}
897: The partition function is obtained by summing over all possibles value of
898: $n,r_{ss},r_{ds}$. The free energy $g_{ds}(i)$ can be computed
899: using e.g. the Mfold program for base-pair interactions \cite{Zuk00}.
900: The ssDNA, dsDNA and the lever can be considered as three springs
901: in series, with an effective stiffness is
902: $k_{tot}^{-1}=k_{ss}^{-1} + k_{ds}^{-1} +k_{lever}^{-1}$.
903: The spring constant of dsDNA,
904: $k_{ds} \simeq 0.03$~ pN/nm for a
905: dsDNA total length of $15000$ bases in the experiment
906: of Bockelmann {\em et al.}, is obtained from the
907: derivative of $f^{WLC}$ (including the Young modulus)
908: calculated at a force of 12 pN. The ssDNA stiffness is
909: $k_{ss} \simeq 6 k_{B T} /(b^{2} n) \simeq 50/n$~pN/nm, and
910: the cantilever stiffness equals $k_{lever}=0.25$~ pN/nm.
911: For less than $\approx 1500$ unzipped base pairs, $k_{tot}$ essentially
912: reduces to the dsDNA linker stiffness. When more bases are unzipped,
913: the dominant contribution comes from the ssDNA stiffness.
914:
915: The DNA sequence dependence results in a complicated free energy landscape
916: that generates a 'stick-slip' variation of the force during unzipping
917: \cite{Boc02}.
918: The stick phase corresponds to the presence of one deep
919: minimum, and the slip phase to a flat free energy landscape
920: see Fig~6. The analytical description of Bockelmann
921: {\em et al.} predicts that the
922: height $h$ of the potential barrier increases as $h \approx \delta^2$
923: with the fluctuations $\delta$ of the pairing free energy,
924: and decreases $h \approx k_{tot}^{-1}$ with the effective stiffness.
925:
926: \subsection{Kinetics of unzipping}
927:
928: The kinetics of unzipping at short length scales
929: is affected by the presence of barriers
930: with various physical origins, which makes it an activated process.
931:
932: \subsubsection{Homogeneous sequences and unzipping initiation barriers}
933:
934: Unzipping of homogeneous DNA requires the crossing of a barrier
935: whose physical origin is the following. We imagine $r$, the half
936: distance between the two bases of one pair to be a reaction coordinate
937: indicating whether the pair is bonded ($r\simeq 10 \AA$) or open
938: ($r>11-12 \AA$~ due to the very short range of H-bond).
939: The effective free energy $V(r)$
940: of the base pair as a function of $r$ is shown in Fig.~\ref{rub}.
941: It is low for both small (pairing energy) and large (entropy gain)
942: values of $r$, and exhibits a maximum around $r\simeq 10.5 \AA$~ where
943: the H-bond is broken but bases are still stacked in the double helix
944: conformation and are not free to move \cite{Coc01}.
945: The set of half distances $r(n)$ between the two strands defines an
946: abstract polymer (Fig.~\ref{rub}). At low enough forces this polymer
947: is confined to the potential well (closed state). When a force $f$ larger
948: than $f_u$ is applied at one extremity, the polymer escapes from
949: the well (unzipping). As in a first order phase transition, nucleation
950: theory can be employed to understand the opening kinetics\cite{Lan67}.
951:
952: The kinetics are slowed because of the activated
953: crossing of the free energy barrier.
954: A saddle-point calculation provides the optimal configuration of the
955: polymer for crossing the barrier\cite{Coc01}. This configuration is made of a
956: transition `bubble' of a few $\simeq 4$ bases long,
957: and free energy cost $G^*(f)$.
958: The time of unzipping initiation grows exponentially with
959: the activation free energy $G^*(f)$, $t(f)=t_0 \exp (G^*/k_BT)$.
960: The elementary time $t_0$ corresponds to the time necessary for the
961: polymer to escape from the saddle-point configuration along the unstable
962: direction in the free energy landscape\cite{Lan67}.
963: When the applied force is smaller than $f_u$, the molecule may
964: still unzip. The opening time, which still depends on the barrier
965: free energy, is now exponentially large in the equilibrium
966: free energy $N g(f)$ where $N$ is the number of base pairs.
967:
968: The determination of the opening time $t(f)$ at fixed force provides
969: in turn the distribution of unzipping forces when the molecule is
970: loaded with a fixed rate\cite{Eva92}. The most probable unzipping
971: force exhibits a rich pattern
972: depending on the loading rate and on the length of the
973: sequence\cite{Coc01,Coc02}.
974:
975:
976: \begin{figure}
977: \begin{center}
978: \psfig{figure=rub.eps,height=7cm,angle=-90}
979: \end{center}
980: \vskip .5cm
981: \caption{Schematic representation of the saddle-point polymer
982: conformation $r^*(n)$ which crosses the barrier of $V(r)$;
983: $r_i$ and $r_f$ are the radii of the edges of the nucleation bubble.}
984: \label{rub}
985: \end{figure}
986:
987:
988: \subsubsection{Sequence-dependent barriers}
989:
990: The above mechanism is mostly relevant in the kinetics of initiation
991: of opening of the double helix. During unzipping of a long
992: double helix, the sequence dependent free energy
993: landscape of Fig~\ref{p5abfig} with barriers of $\approx 10 k_BT$ is
994: responsible for a slow hopping dynamics between the open and closed
995: states, i.e. the slip-stick behavior observed by Bockelmann et al
996: \cite{Ess97,Boc02}.
997:
998: In experiments by Liphardt et al\cite{Lip01}, small helix-loop RNA structures
999: (essentially short regions of double-helical RNA terminated with a short
1000: loop) were held in such a way that equilibrium fluctuation between
1001: open and closed states occurred. The timescale observed was
1002: close to 1 sec \cite{Lip01}, remarkably long for a few-nm-long molecule.
1003: In \cite{Coc03b} a dynamical model was introduced for the motion of
1004: the `fork' separating the base paired and opened regions of the
1005: molecule, allowing computation of the opening and closing rate as a
1006: function of the force.
1007: The model, which describes the experimental data well,
1008: is based on the following elementary
1009: rates of opening and closing base pair $n$
1010: \begin{equation}
1011: r_0(n)= r\; e^{-g_0(n)/k_B T}, \;\; r_c(f,n)= r\; e^{-2 {\cal F}_{ss,}
1012: (f,n)}
1013: \end{equation}
1014: The opening rate is taken to depend only on the pairing free energy since
1015: the short range hydrogen bond is broken before the force-length work
1016: over the longer $\approx 0.7$ nm distance can be done. Conversely,
1017: to close the base pair, work must be first done against the
1018: applied force, and so the closing rate is taken to be depend only on
1019: the force.
1020: The separation of length scales of the range of base
1021: pair interaction and base pair extension after unzipping is thus
1022: used to justify placing most of the force-dependence in the zipping rate,
1023: with most of the interaction dependence in the unzipping rate.
1024: More sophisticated rate models will require further experiments to
1025: determine their form.
1026:
1027: \section{Conclusion}
1028:
1029: We have presented a very brief overview of the theory used to think
1030: about single-molecule nucleic acid micromanipulation experiments.
1031: The field of single-molecule experiments is evolving so rapidly at present
1032: that we have been forced to omit many exciting topics.
1033: Here we present a few general comments about
1034: what has been learned and suggest some directions that might be particularly
1035: interesting for study in the near future.
1036:
1037: A feature common to all the studies described above, and to the theory
1038: of other types of single-biomolecule
1039: experiments, is the central role of statistical mechanics.
1040: The interaction of this field with statistical mechanics is fundamental:
1041: the understanding (in some cases, even the primary data analysis)
1042: of single-molecule DNA experiments requires statistical
1043: mechanics. Additionally, previously unimagined
1044: statistical-mechanical problems are resulting from
1045: the huge range of experimental possibilities for DNA and DNA-protein
1046: micromechanical experiments.
1047:
1048: The first phase of single-DNA experiments involved basic characterization
1049: of dsDNA and ssDNA, and from the theoretical side involved development
1050: and solution of statistical-mechanical theories for the molecules subjected
1051: to stresses. The studies reviewed above essentially fall into this
1052: first class, and are characterized by a degree of
1053: quantitative success, thanks both to the efforts of experimentalists
1054: and theorists, which is unprecedented in soft condensed matter physics.
1055:
1056: The second phase, which we are in at present, involves
1057: the study of modifications of the basic molecules, e.g. by unzipping,
1058: or by action of proteins acting on DNAs under
1059: mechanical control. The statistical mechanics of this second class
1060: of problems is less well developed, and includes problems far from
1061: thermal equilibrium such as DNAs acted on by processive, ATP-powered
1062: motor-like enzymes. The diversity of challenging and new statistical
1063: physics problems in this second class is mind-boggling.
1064: The second phase is also forcing theorists to confront the
1065: information content of nucleic acids since sequence plays an
1066: essential role in targeted nucleic acid-protein interactions.
1067:
1068: The third class of problems involves applications of the lessons
1069: learned, to the description of cell machinery in vivo, or at least
1070: under in-vivo-like conditions. Such experiments are in their
1071: infancy, and extension of theoretical physics into this domain
1072: is still in a dark age. However, we can look forward
1073: to a time when we understand processes such as cell division and
1074: growth, gene regulation, and other cell biological functions in
1075: statistical-mechanistic terms. The lessons being learned now
1076: about the importance of statistical mechanical ideas to
1077: biochemical-micromechanical experiments on nucleic acids will thus become
1078: an important component of future quantitative understanding
1079: of cell biology.
1080:
1081:
1082: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1083: %%% Acknowledgements %%%
1084: %%%%%%%%%%%%%%%%%%%%%%%%%%
1085: {\bf Acknowledgements}
1086: We thank A. Sarkar, V. Croquette and D. Bensimon for their helpful comments
1087: on this manuscript. This work was in part supported by
1088: the NSF(USA; JM and SC: DMR-9734178, RM: DMR-9808595),
1089: the Research Corporation (JM), the A. della Riccia Foundation (SC),
1090: and by the Focused Giving Program of Johnson and Johnson (JM).
1091:
1092: \begin{thebibliography}{999999}
1093: \bibitem{Flo89}
1094: P.J. Flory,
1095: {\em Statistical Mechanics of Chain Molecules}. Hanser, Munich (1989).
1096:
1097: \bibitem{Doi86}
1098: M. Doy, S.F. Edwards, {\em The Theory of Polymer Dynamics}.
1099: Oxford University Press, Oxford (1986).
1100:
1101: \bibitem{Bus94}
1102: C. Bustamante, J.F. Marko, E.D. Siggia, S. Smith,
1103: Entropic elasticity of lambda-phage DNA.
1104: {\em Science} {\bf 265}, 1599 (1994).
1105:
1106: \bibitem{Mar95}
1107: J.F. Marko, E.D. Siggia,
1108: Stretching DNA.
1109: {\em Macromolecules} {\bf 28}, 209 (1995).
1110:
1111: \bibitem{Aus97}
1112: R.H. Austin, J.P. Brody, E.C. Cox, T. Duke, W. Volkmuth,
1113: Stretch genes.
1114: {\em Physics Today} {\bf 50}, 32 (Feb. 1997).
1115:
1116: \bibitem{Bus00}
1117: C. Bustamante, S.B. Smith, J. Liphardt, D. Smith,
1118: Single-molecule studies of DNA mechanics,
1119: {\em Curr. Op. Struct. Biol} {\bf 10}, 279 (2000).
1120:
1121: \bibitem{Wan97}
1122: M.D. Wang, H. Yin, R. Landick, J. Gelles, S. Block,
1123: Stretching DNA with optical tweezers,
1124: {\em Biophys. J.} {\bf 72}, 1335 (1997).
1125:
1126: \bibitem{Str00}
1127: T. Strick, J.F. Allemand, V. Croquette, D. Bensimon,
1128: Twisting and stretching single DNA molecules,
1129: {\em Progress in Biophysics and Molecular Biology} {\bf 74}, 115 (2000).
1130:
1131: \bibitem{Smi92}
1132: S.B. Smith, L. Finzi, C. Bustamante,
1133: Direct mechanical measurements of the elasticity of single DNA
1134: molecules by using magnetic beads.
1135: {\em Science} {\bf 258}, 1122 (1992).
1136:
1137: \bibitem{Bou99}
1138: C. Bouchiat, M.D. Wang, J F. Allemand, T. Strick, S.M. Block,
1139: V. Croquette,
1140: Estimating the persistence length of a worm-like chain molecule from
1141: force extension measurements,
1142: {\em Biophys. J.} {\bf 76}, 409 (1999).
1143:
1144: \bibitem{Smi96}
1145: S. Smith, Y. Cui, C. Bustamante,
1146: Overstretching B-DNA: The elastic response of individual
1147: double-stranded and single-stranded DNA molecules.
1148: {\em Science} {\bf 271}, 795 (1996).
1149:
1150: \bibitem{Bar93}
1151: J.L. Barrat, J.F. Joanny,
1152: Persistence length of polyelectrolytes chains,
1153: {\em Europhys. Lett.} {\bf 24}, 333 (1993).
1154:
1155: \bibitem{Bau97}
1156: C.G. Baumann, S.B. Smith, V.A. Bloomfield, C. Bustamante,
1157: Ionic effects on the elasticity of single DNA molecules.
1158: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 94}, 6185 (1997).
1159:
1160:
1161: \bibitem{Clu96}
1162: P. Cluzel, A. Lebrun, C. Heller, R. Lavery, J.L. Viovy, D. Chatenay,
1163: F. Caron,
1164: DNA: An extensible molecule.
1165: {\em Science} {\bf 271}, 792 (1996).
1166:
1167: \bibitem{Leb96}
1168: A. Lebrun, R. Lavery,
1169: Modelling extreme deformations of DNA,
1170: {\em Nucl. Acids Res.} {\bf 24},2260 (1996).
1171:
1172: \bibitem{Ciz97}
1173: P. Cizeau, J.L. Viovy ,
1174: Modelling extreme extension of DNA,
1175: {\em Biopolymers} {\bf 42}, 383-385 (1997).
1176:
1177: \bibitem{Mar03}
1178: J. Marko, M. Feig, B.M. Pettitt,
1179: Unification of the microscopic atomic fluctuations with mesoscopic
1180: elasticity of the DNA double helix,
1181: preprint (2001).
1182:
1183: \bibitem{Coc99b}
1184: S. Cocco, R. Monasson,
1185: Theoretical study of collective modes in DNA at ambient temperature,
1186: {\em J. Chem. Phys. } {\bf 112}, 10017 (2000).
1187:
1188: \bibitem{Hil59}
1189: T.L. Hill,
1190: {\em J. Chem. Phys.} {\bf 30}, 383 (1959).
1191:
1192: \bibitem{Zim59}
1193: B. Zimm, J. Bragg,
1194: {\em J. Chem. Phys.} {\em 31}, 526 (1959).
1195:
1196: \bibitem{Kafri}
1197: Y. Kafri, D. Mukamel, L. Peliti,
1198: Why is the DNA denaturation transition first order?
1199: {\em Phys. Rev. Lett.} {\bf 85}, 4988 (2000).
1200:
1201: \bibitem{Mar98}
1202: J.F. Marko, DNA under high tension: overstretching, undertwisting,
1203: and relaxation dynamics,
1204: {\em Phys. Rev. E} {\bf 57}, 2134 (1998).
1205:
1206: \bibitem{Str96}
1207: T.R. Strick, J-F. Allemand, D. Bensimon, A. Bensimon, V. Croquette,
1208: The elasticity of a single supercoiled DNA Molecule,
1209: {\em Science} {\bf 271}, 1835 (1996).
1210:
1211: \bibitem{Str98}
1212: T.R. Strick, J.F. Allemand, D. Bensimon, V. Croquette,
1213: Behavior of supercoiled DNA,
1214: {\em Biophys. J.} {\bf 74}, 2016 (1998).
1215:
1216:
1217: \bibitem{Leg99}
1218: J-F. L{\'e}ger, G. Romano, A. Sarkar, J. Robert, L. Bourdieu, D. Chatenay,
1219: J.F. Marko,
1220: Structural Transitions of a Twisted and Stretched DNA Molecule,
1221: {\em Phys. Rev. Lett.} {\bf 83}, 1066 (1999).
1222:
1223: \bibitem{All98}
1224: J.F. Allemand, D. Bensimon, R. Lavery, V. Croquette,
1225: Stretched and overwound DNA forms a Pauling-like structure with exposed bases,
1226: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 74}, 2016 (1998).
1227:
1228: \bibitem{Mar95b}
1229: J.F. Marko , E.D. Siggia,
1230: Statistical mechanics of supercoiled DNA.
1231: {\em Phys. Rev. E} {\bf 52}, 2912 (1995).
1232:
1233: \bibitem{Mar97}
1234: J.F. Marko,
1235: Supercoiled and braided DNA under tension,
1236: {\em Phys. Rev. E} {\bf 55}, 1758 (1997).
1237:
1238: \bibitem{Coc99a}
1239: S. Cocco, R Monasson,
1240: Statistical mechanics of torque induced denaturation of DNA,
1241: {\em Phys. Rev. Lett.} {\bf 83}, 5178 (1999).
1242:
1243: \bibitem{Sar01}
1244: A. Sarkar, J-F. L\'eger, D. Chatenay, J.F. Marko,
1245: Structural transitions in DNA driven by external force and torque,
1246: {\em Phys. Rev. E} {\bf 63}, 051903 (2001).
1247:
1248: \bibitem{Fai97}
1249: B. Fain, J. Rudnick, S. Ostlund,
1250: Conformations of linear DNA,
1251: {\em Phys. Rev. E} {\bf 55}, 7364 (1997).
1252:
1253: \bibitem{Mor97}
1254: J.D. Moroz, P. Nelson,
1255: Torsional directed walks, entropic elasticity, and DNA twist stiffness,
1256: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 94}, 1441 (1997).
1257:
1258: \bibitem{Bou98}
1259: C. Bouchiat, M. M{\'e}zard,
1260: Elasticity model of a supercoiled DNA molecule,
1261: {\em Phys. Rev. Lett.} {\bf 80}, 1556 (1998).
1262:
1263: \bibitem{Des03}
1264: M.N. Dessinges, B. Maier, M. Peliti, D. Bensimon, V. Croquette,
1265: Stretching single stranded DNA,
1266: a real self avoiding and interacting heteropolymer.
1267: preprint (2001).
1268:
1269: \bibitem{Zha01}
1270: Y. Zhang, H. Zhou, Z.C. Ou-Yang,
1271: Stretching single-stranded DNA: interplay of electrostatic, base-pairing, and
1272: base-pair stacking interactions,
1273: {\em Biophys. J.} {\bf 81}, 1133 (2001).
1274:
1275: \bibitem{Coc03}
1276: S. Cocco, R. Monasson, J. Yan, A. Sarkar, J.F. Marko,
1277: Elastic response of folding polymers,
1278: preprint (2002).
1279:
1280: \bibitem{Bou03}
1281: C. Bouchiat,
1282: Hartree-Fock computation of self avoiding flexible polymer elasticity,
1283: preprint (2001).
1284:
1285: \bibitem{Mai00}
1286: B. Maier, D. Bensimon, V. Croquette,
1287: Replication by a single DNA-polymerase of a stretched single strand DNA.
1288: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 97}, 12002 (2000).
1289:
1290: \bibitem{Bun99}
1291: R. Bundschuh, T. Hwa,
1292: Statistical mechanics of secondary structures fromed by random RNA
1293: sequences.
1294: {\em Phys. Rev. E} {\bf 65}, 031903 (2002).
1295:
1296: \bibitem{Pag00}
1297: A. Pagnani, G. Parisi, F. Ricci-Tersenghi,
1298: Glassy transition in a disordered model for the RNA secondary structure.
1299: {\em Phys. Rev. Lett.} {\bf 84}, 2026 (2000).
1300:
1301: \bibitem{Isa00}
1302: H. Isambert, E.D. Siggia,
1303: Modeling RNA folding paths with pseudoknots: Application to hepatitis
1304: delta virus ribozyme
1305: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 97}, 6515 (2000)
1306: \bibitem{Mon01}
1307: A. Montanari, M. Mezard,
1308: Hairpin formation and elongation of biomolecules,
1309: {\em Phys. Rev. Lett.} {\bf 86}, 2178 (2001).
1310:
1311: \bibitem{Mar97a}
1312: J.F. Marko, E.D. Siggia,
1313: Driving proteins off DNA using applied tension,
1314: { \em Biophys. J.} {\bf 73}, 2173 (1997).
1315:
1316: \bibitem{Wan98}
1317: Y. Cui, C. Bustamante,
1318: Pulling a single chromatin fiber reveals the forces that maintain its
1319: higher-order structure.
1320: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 97}, 127-132 (2000).
1321:
1322: \bibitem{Ess97}
1323: B. Essevaz-Roulet, U. Bockelmann, F. Heslot,
1324: Mechanical separation of the complementary strands of DNA,
1325: {\em Proc. Natl. Acad. Sci. (USA) } {\bf 94}, 11935 (1997).
1326:
1327: \bibitem{Rie99}
1328: M. Rief, H. Clausen-Schaumann, H.E. Gaub,
1329: Sequence-dependent mechanics of single DNA molecules,
1330: { \em Nat. Struct. Biol.} {\bf 6}, 346 (1999).
1331:
1332: \bibitem{Bre86}
1333: K. Breslauer, R. Frank, H. Blocker, L.A. Marky,
1334: Predicting DNA duplex stability from the base sequence,
1335: {\em Proc. Natl. Acad. Sci. (USA)} {\bf 83}, 3746 (1986).
1336:
1337: \bibitem{Lub00}
1338: D.K. Lubensky, D.R. Nelson,
1339: Pulling pinned polymers and unzipping DNA,
1340: {\em Phys. Rev. Lett.} {\bf 85}, 1572 (2000).
1341:
1342: \bibitem{Lub02}
1343: D.K. Lubensky, D.R. Nelson,
1344: Single molecule statistics and the polynucleotide unzipping transition,
1345: {\em Phys. Rev. E} {\bf 65}, 031917 (2002).
1346:
1347: \bibitem{Zuk00}
1348: M. Zuker,
1349: Calculating nucleic acid secondary structure,
1350: {\em Curr. Opin. Struct. Biol.} {\bf 10}, 303-310 (2000).
1351:
1352: \bibitem{Lip01}
1353: J. Liphardt, B. Onoa, S.B. Smith, I. Tinoco JR, C. Bustamante,
1354: Reversible unfolding of single RNA molecules by mechanical force.
1355: {\em Science} {\bf 292}, 733-737 (2001).
1356:
1357: \bibitem{Tho95}
1358: E.R. Thompson, E.D. Siggia,
1359: Physical limits on the mechanical measurement of the secondary
1360: structure of Bio-molecules.
1361: {\em Europhys. Lett.} {\bf 31}, 335-340 (1995).
1362:
1363: \bibitem{Boc98}
1364: U. Bockelmann, B. Essevaz-Roulet, F. Heslot,
1365: DNA strand separation studied by single molecule force measurements.
1366: {\em Phys. Rev. E} {\bf 58}, 2386 (1998).
1367:
1368: \bibitem{Coc01}
1369: S. Cocco, R. Monasson, J.F. Marko,
1370: Force and kinetic barriers to unzipping of the DNA double helix
1371: {\em Proc. Natl. Aca. Sci. (USA)} {\bf 98}, 8608-8613 (2001).
1372:
1373: \bibitem{Coc02}
1374: S. Cocco, R. Monasson, J.F. Marko,
1375: Force and kinetic barriers to initiation of DNA unzipping.
1376: {\em Phys. Rev. E} {\bf 65}, 041907 (2002).
1377:
1378: \bibitem{Lan67}
1379: J.S. Langer, Statistical theory of the decay of metastable states.
1380: {\em Ann. Phys. (N.Y.)} {\bf 54}, 258-275 (1967).
1381:
1382: \bibitem{Eva92}
1383: E. Evans, K. Ritchie, Dynamic strength of molecular adhesion bonds.
1384: {\em Biophys. J.} {\bf 72}, 1541-1555 (1997).
1385:
1386: \bibitem{Rouzina}
1387: I. Rouzina, V.A. Bloomfield, Force-induced melting of the DNA double helix.
1388: {\em Biophys. J.} {\bf 80}, 882--893 (2001).
1389:
1390: \bibitem{Boc02}
1391: U. Bockelmann, P. Thomen, B. Essevaz-Roulet, V. Viasnoff, F. Heslot,
1392: Unzipping DNA with optical tweezers: high sequence sensitivity and
1393: force flips.
1394: {\em Biophys J.} {\bf 82}, 1537-1553 (2002).
1395:
1396: \bibitem{Coc03b}
1397: S. Cocco, J.F. Marko J.F., R. Monasson,
1398: Slow nucleic acid unzipping kinetics from sequence-defined barriers.
1399: preprint (2002).
1400:
1401: \end{thebibliography}
1402:
1403: \end{document}
1404:
1405:
1406:
1407:
1408:
1409:
1410:
1411:
1412:
1413:
1414:
1415:
1416:
1417:
1418:
1419:
1420: