cond-mat0412504/ens4.tex
1: \documentstyle[preprint,aps,prl,epsf]{revtex}
2: %\documentstyle[aps,prl,epsf,twocolumn]{revtex}
3: \begin{document}
4: \def\bea{\begin{eqnarray}}
5: \def\eea{\end{eqnarray}}
6: \def\a{\alpha}
7: \def\d{\delta}
8: \def\p{\partial} 
9: \def\nn{\nonumber}
10: \def\r{\rho}
11: \def\rv{\bar{r}}
12: \def\la{\langle}
13: \def\ra{\rangle}
14: \def\e{\epsilon}
15: \def\o{\omega}
16: \def\n{\eta}
17: \def\g{\gamma}
18: \def\break#1{\pagebreak \vspace*{#1}}
19: \def\f{\frac}
20: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname
21: %@twocolumnfalse\endcsname
22: \draft
23: \title{Inequivalence of Statistical Ensembles \\in\\ 
24: Single Molecule Measurements}
25: \author{Supurna Sinha$^{}$\cite{SUP} and Joseph Samuel$^{}$\cite{SAM}}
26: \address{Harish Chandra Research Institute,\\
27: Chhatnag Road, Jhunsi, Allahabad 211 019, India\\
28: and\\
29: Raman Research Institute, Bangalore 560 080, India\\}
30: %\date{\today}
31: \maketitle
32: \widetext
33: \begin{abstract}
34: We study the role of fluctuations in 
35: single molecule experimental measurements of force-extension ($f-\zeta$)
36: curves. We use the Worm Like Chain (WLC) model to bring out the 
37: connection between the Helmholtz ensemble characterized by the 
38: Free Energy ($F(\zeta)$) and the Gibbs ensemble characterized by 
39: the Free Energy ($G(f)$). 
40: We consider the rigid rod limit of the 
41: WLC model as an instructive special case to bring out the issue of 
42: ensemble inequivalence. 
43: We point out the need for taking into account the free energy of
44: transition when one goes from one ensemble to another. 
45: We also comment on the ``phase transition'' noticed in an 
46: isometric setup for semiflexible 
47: polymers and propose a realization of its thermodynamic limit.  
48: We present general arguments which rule out non-monotonic force-extension
49: curves in some ensembles
50: and note that these do {\it not} apply to the isometric ensemble. 
51: 
52: \end{abstract}
53: \pacs{PACS numbers: 87.15.-v,05.40.-a,36.20.-r}
54: \narrowtext
55: \textwidth=12cm
56: {\bf Introduction:}
57: 
58: In the past, experiments on polymers were confined to studying their
59: bulk properties, which involved probing 
60: large numbers of molecules\cite{bulk}. 
61: The results of these experiments could be analyzed by using 
62: the traditional tools of thermodynamics. 
63: In recent years, however, researchers have been successful in 
64: micromanipulating {\it single}
65: biological molecules such as DNA, proteins and RNA to probe 
66: their elastic
67: properties \cite{bust}. 
68: Such studies serve a twofold role. 
69: On one hand, they shed light on 
70: mechanical properties of semiflexible polymers,
71: which are of clear biological importance in processes such as 
72: gene regulation and transcription \cite{Strick,twist,writhe}.  
73: On the other hand, they provide physicists with a concrete testing
74: ground for understanding some of the fundamental ideas of 
75: statistical mechanics.
76: In statistical mechanics, an isometric setup would be described by  
77: the 
78: Helmholtz free energy, whereas an isotensional setup would be
79: described by the Gibbs free energy\cite{landau}. In the thermodynamic 
80: limit 
81: these 
82: two descriptions agree, but semiflexible polymers (those with contour 
83: lengths comparable to their persistence lengths) are {\it not} at the 
84: thermodynamic limit. Experimentally, both isometric and isotensional 
85: ensembles are realizable. 
86: Typically the polymer molecule is suspended (in a suitable medium) between 
87: a translation stage
88: and a force sensor.
89: The force sensor could be realized by using an Atomic 
90: Force
91: Microscope (AFM) cantilever or by optical or magnetic forces. 
92: As noted by Kreuzer and Payne \cite{kreuz},
93: an isometric setup can be realized using a stiff trap and an isotensional 
94: setup by using a soft trap. 
95: In a more sophisticated
96: version, an electronic feedback circuit is used to control the 
97: force (or the extension) and one measures the fluctuations in the 
98: extension
99: (or the force) \cite{keller}.
100: Here we will focus on the role of 
101: fluctuations in single molecule experiments.
102: In order to correctly interpret such experiments one needs to 
103: understand the effect of fluctuations on the measured quantities. 
104: For instance, it turns out, that an experiment in which the 
105: ends of a polymer molecule are fixed (isometric) and the 
106: force fluctuates yields a different
107: result from one in which the force between the ends is held fixed 
108: (isotensional) and 
109: the end-to-end distance fluctuates\cite{keller,neumann}. 
110: This difference can be traced to 
111: large fluctuations about the mean value of the force or the extension,
112: depending on the experimental setup. These fluctuations
113: vanish only in the thermodynamic limit of very long polymers. 
114: 
115: Here we use the Worm Like Chain (WLC) model \cite{kratky,bend,Kleinert} to 
116: study the inequivalence of ensembles due to 
117: finite size effects.
118: The WLC model 
119: has been very successful in achieving quantitative agreement with  
120: experimentally measured force-extension curves\cite{bust,Marko}. 
121: The paper is organized as follows. In Sec. $II$ we discuss the Helmholtz
122: and Gibbs ensembles. In Sec. $III$ we consider the rigid rod limit which
123: forcefully brings out the main issues dealt with in this paper. 
124: In Sec. $IV$ we draw attention to the importance of taking into 
125: account the free energy of transition in going from one ensemble to 
126: another. In Sec. $V$ we discuss the thermodynamic
127: limit of a ``phase transition'' recently seen in semiflexible polymers. 
128: Finally, we end the paper with a discussion in Sec. $VI$. 
129:   
130: {\bf II Helmholtz and Gibbs Ensembles:}
131: 
132: Consider an idealized experiment in which one end of a molecule is 
133: held fixed at $(x_0,y_0,0)$ and the other end is attached to a dielectric
134: bead confined to a harmonic optical trap described by the potential
135: \begin{equation}
136: V(x,y,z) = A\frac{[(x-x_0)^2+(y-y_0)^2]}{2} + C\frac{(z-z_0)^2}{2}.
137: \label{trap}
138: \end{equation}
139: with $(x_0,y_0,z_0)$ defining the center of the trap. 
140: Consider $A$ to be small 
141: so that the bead is free to move in the plane $z=z_0$.
142: For a polymer of contour length $L$ and persistence length $L_P$
143: it is convenient to introduce the following dimensionless variables:
144: $\zeta =\frac{z}{L}$, $\zeta_0 =\frac{z_0}{L}$, $\beta =\frac{L}{L_P}$ and 
145: $f=\frac{FL_P}{k_BT}$ where $F$ is an applied stretching force and
146: $k_BT$ is the thermal energy at a temperature $T$.
147: Consider $P(\zeta)d{\zeta}$, the number of 
148: configurations (counted with 
149: Boltzmann weight)\cite{bend} for a polymer of length $L$ 
150: starting from the origin and ending anywhere on the $x-y$ plane
151: in an interval $d\zeta$ of $\zeta$. 
152: The free energy defined by ${\cal F}(\zeta)= -\frac{1}{\beta}ln P(\zeta)$ 
153: is the 
154: Helmholtz
155: free energy.  
156: The partition function\cite{partfoot} for the combined system consisting 
157: of the 
158: polymer molecule
159: {\it and} the trap is given by:
160: \begin{equation}
161: Z(\zeta_0,\beta)=\sqrt{\frac{\tilde C}{2 
162: \pi}}\int_{-\infty}^{+\infty}{d\zeta e^{-\beta 
163: {\cal F}(\zeta)}e^{-\tilde{C}\frac{(\zeta 
164: -\zeta_0)^2}{2}}}
165: \label{part}
166: \end{equation}
167: where $\tilde{C} =C L^2$.
168: By tuning the longitudinal stiffness $\tilde{C}$ we can realize the two 
169: limiting
170: cases. 
171: ${\it Helmholtz:}$
172: In the limit of a stiff trap ($\tilde{C} \rightarrow \infty$), the 
173: Gaussian factor 
174: pertaining to the trap approaches a delta function and one gets
175: \begin{equation}
176: Z(\zeta,\beta)= e^{-\beta {\cal F}(\zeta)}
177: \label{stifftrap}
178: \end{equation} 
179: Here we have switched notation to write $\zeta$ in place of $\zeta_0$.
180: Thus a stiff trap realizes the Helmholtz 
181: ensemble by constraining fluctuations in the $\zeta$ coordinate.
182: To extend the molecule from $\zeta$ to $\zeta+d\zeta$ one needs
183: to apply a force $<f>=\frac{\partial F}{\partial \zeta}$
184: in order to compensate for the change of entropy. Plotting
185: $<f>$ versus $\zeta$ we find the $(<f>,\zeta)$ force-extension relation.
186: 
187: ${\it Gibbs:}$ 
188: In the opposite limit of a soft trap ($\tilde{C} \rightarrow 0$ and 
189: $\zeta_0 
190: \rightarrow \infty$ such that $\tilde{C}\zeta_0 =\beta f$ 
191: remains finite),
192: one gets\cite{partfoot}
193: \begin{equation}
194: {\tilde Z}(f,\beta)=\int_{-\infty}^{+\infty}{d\zeta e^{-\beta 
195: F(\zeta)}e^{\beta 
196: f\zeta}}
197: \label{softrap}
198: \end{equation}
199: Thus a soft trap permits fluctuations in the $\zeta$ coordinate but 
200: constrains
201: the force fluctuations\cite{kreuz} and thus realizes the 
202: Gibbs ensemble.
203: ${\tilde Z}(f)$ is the generating function for the $\zeta$ distribution.
204: Defining the Gibbs free energy $G(f)={-\beta}ln{\tilde{Z}}(f)$ we
205: can work out the mean extension $<\zeta>=-\frac{\partial G}{\partial f}$
206: and the $(<\zeta>,f)$ force-extension relation.
207: 
208: Notice that ${\tilde Z}(f)$ is the Laplace transform 
209: of 
210: $Z(\zeta)$.    
211: In the thermodynamic limit of long polymers ($\beta 
212: \rightarrow \infty$) the Laplace transform integral Eq.(\ref{softrap}) 
213: is dominated by 
214: the saddle point value and therefore  
215: ${\cal F}(\zeta)$ and $G(f)$ are related by a Legendre transform:
216: \begin{equation}
217: {\cal F}(\zeta) = G(f) +f\zeta.
218: \label{legendre}
219: \end{equation}
220: For finite $\beta$ i.e. for a polymer of finite extent, 
221: the saddle point approximation 
222: no longer holds true and fluctuations about the saddle point value
223: of the free energy become important.  
224: Thus one finds that ${\cal F}(\zeta)$ and $G(f)$ are {\it not} 
225: Legendre transforms of each other.
226: We notice that this difference between the Legendre transform [Eq. 
227: (\ref{legendre})] and 
228: the Laplace transform [Eq. (\ref{softrap})] is the mathematical origin 
229: of the finite size fluctuation effects described in this paper. 
230: These fluctuations are of thermal origin and 
231: can ultimately be traced to 
232: collisions of the polymer molecule
233: with the molecules of the suspending medium. 
234: In this section we have recovered the results of \cite{kreuz}.
235: We have also gone beyond \cite{kreuz} and 
236: traced the origin of the 
237: difference between the two 
238: ensembles to the difference between the 
239: Legendre and Laplace transforms.
240: 
241: {\bf III Rigid Rod Limit: An Instructive Extreme Case}
242: 
243: We noticed in the last section that because of fluctuation effects
244: the Helmholtz and Gibbs free energies are not related by a 
245: Legendre Transform. 
246: An important consequence of this is that the $(F,<z>)$ relation 
247: is different from the $(<F>, z)$ relation. In other words, the 
248: force-extension curves plotted in the two ensembles are {\it distinct}
249: due to finite size fluctuation effects.  
250: Fluctuations dominate at finite $\beta$ 
251: and disappear in the thermodynamic limit ($\beta \rightarrow \infty$) of 
252: flexible polymers. 
253: We bring out the ensemble dependence 
254: of the force-extension 
255: relations explicitly and most dramatically 
256: in the limiting case of a very stiff polymer ($\beta \rightarrow 0$).
257: 
258: In this extremely rigid limit\cite{frey}, the end of the polymer is uniformly 
259: distributed over the sphere of directions. In the Helmholtz ensemble we  
260: thus have $P(z)d{z}= \frac{d z}{2L}$ for $-L<z<L$ and 
261: $P(z) =0$ otherwise. 
262: Since the free energy is constant in the range $-L<z<L$ and diverges 
263: otherwise, we find that the average force $<F>$ vanishes for 
264: $|z|<L$ and diverges for 
265: $|z| =L$ (See Fig. $1 a$). In the Gibbs ensemble we find by standard 
266: manipulations that 
267: \begin{equation}
268: <z>=[LCoth{FL}-\frac{1}{F}]
269: \label{zfrel}
270: \end{equation}
271: which differs from the $<F>-z$ relation determined above in 
272: the Helmholtz ensemble (See Fig. $1 b$). 
273: The theoretical analysis of the ensemble dependence of the force-extension 
274: relation based on the rigid rod limit is a new result of this paper.
275: 
276: Thus an experimenter making force-extension measurements on, 
277: for instance, Actin filaments\cite{loic}, would find that a measurement
278: in which the force is controlled and the end-to-end distance is measured
279: leads to a different force-extension curve 
280: from a measurement in which the end-to-end separation is controlled and 
281: the force is measured. A theorist interpreting the curves also needs 
282: to keep in mind whether the curves are obtained in a constant-force
283: setup or a constant-extension setup since a proper interpretation of
284: the curves requires a knowledge of the ensemble used in the 
285: measurement.   
286: 
287: \vspace{.6 cm}
288: \vbox{
289: \epsfxsize=13.0cm
290: \epsfysize=6.0cm
291: \epsffile{figure1.eps}
292: \begin{figure}
293: \caption{Force-Extension Curve in the Helmholtz (Fig. 1a) and Gibbs 
294: (Fig. 1b)
295: ensembles for $\beta=0.$ We have set $L=1$. }
296: \label{contrast}
297: \end{figure}} 
298: 
299: 
300: {\bf IV Inequivalence Of Ensembles and the Free Energy Of Transition}
301: 
302: In an isotensional setup one controls the force and one 
303: measures the mean extension and plots
304: it as a function of force.
305: In an isometric setup the roles of extension and force are interchanged.
306: In both setups zero force corresponds to zero extension ($\zeta=0$) 
307: and a large force corresponds to maximal extension (i.e. $\zeta = 1$).
308: Imagine going from zero force to a large force via the 
309: isotensional setup and returning from maximal extension to 
310: zero extension via the isometric setup.
311: Since the ``equation of state'' depends on the
312: chosen ensemble, in general there will be two distinct curves describing
313: the extension in one ensemble followed by contraction in the other 
314: \cite{foot}.
315: In such a situation there could be a net area enclosed in the
316: Force-Extension plane. 
317: This poses a puzzle because it appears that a cyclic process can 
318: extract work from the system. This puzzle is easily resolved. 
319: In completing the cycle and returning to the initial
320: state one is in fact changing ensembles twice at the two end points.
321: These correspond to {\it finite} free energy changes which need to be
322: taken into account. 
323:  
324: Let $(\zeta_1,f_1)$ and $(\zeta_2,f_2)$ be two points which lie on 
325: both isometric and isotensional force-extension curves. In our example
326: $(\zeta_1,f_1)=(0,0)$ and $(\zeta_2,f_2)=(1,\infty)$.
327: Let us suppose that we go from $(\zeta_1,f_1)$ to $(\zeta_2,f_2)$
328: in the isotensional ensemble and return via the isometric ensemble.
329: We find that 
330: \begin{equation}
331: G(f_2)-G(f_1) =\int_{f_1}^{f_2}{\frac{\partial G}{\partial 
332: f}df}
333: =-\int_{f_1}^{f_2}{<\zeta> df}. 
334: \end{equation}
335: Similarly,
336: \begin{equation}
337: F(\zeta_2)-F(\zeta_1) =\int_{\zeta_1}^{\zeta_2}{\frac{\partial 
338: F}{\partial 
339: \zeta}d\zeta}
340: =\int_{\zeta_1}^{\zeta_2}{<f> d\zeta}.
341: \end{equation}
342: The area enclosed between the two curves is given by 
343: $$W=\int_{\zeta_1}^{\zeta_2}{f d<\zeta>} 
344: -\int_{\zeta_1}^{\zeta_2}{<f> d\zeta}$$
345: which can be rewritten as 
346: \begin{eqnarray}
347: W=\int_{\zeta_1}^{\zeta_2}{d(f <\zeta>)}
348: -\int_{f_1}^{f_2}{<\zeta> df} - 
349: \int_{\zeta_1}^{\zeta_2}{<f> d{\zeta}}\\{\nonumber}
350: =f_2\zeta_2 - f_1\zeta_1 +G(f_2) - G(f_1)-F(\zeta_2) + 
351: F(\zeta_1)\\{\nonumber}
352: =[f\zeta+G(f)-F(\zeta)]_1^{2} 
353: =[\tilde{F}(\zeta)-F(\zeta)]_1^{2}\\{\nonumber}
354: =\Delta F^{tr}(\zeta)|_1^{2} 
355: \label{freetr}
356: \end{eqnarray}
357: 
358: where we define $\Delta F^{tr}(\zeta)$ as the free energy of transition. 
359: $\Delta F^{tr}(\zeta)$ is the difference between 
360: $\tilde{F}(\zeta)=f\zeta+G(f)$,
361: the Legendre transform of $G(f)$ and the Helmholtz 
362: free energy $F(\zeta)$. Since these are not equal (except in the limit
363: of long polymers) the free energy of transition between ensembles must
364: be considered in order that the total free energy change in a cyclic
365: process vanishes. 
366:    
367: In order to understand this issue more explicitly we consider 
368: corrections to the saddle point approximation which is valid 
369: in the long polymer limit.
370: Let us expand $\phi({\zeta}) = {\cal F}(\zeta)-f\zeta$, the argument of 
371: the exponential appearing 
372: on the 
373: right hand side of Eq. (\ref{softrap})
374: around the saddle point value $\zeta = \zeta_{*}$ (which dominates the 
375: integral in 
376: the long polymer limit) and retain terms upto second order in 
377: the fluctuations about the saddle point value: 
378: $$\phi({\zeta}) = \phi({\zeta_{*}}) 
379: +\frac{1}{2}{\phi}^{''}(\zeta)|_{\zeta=\zeta_{*}}(\zeta 
380: -\zeta_{*})^2.$$
381: If we plug in this expansion in Eq. (\ref{softrap}) and identify ${\tilde 
382: Z}(f,\beta)$ with 
383: $e^{-\beta 
384: G(f)}$
385: we arrive at the following equation:
386: \begin{equation}
387: G(f)=[{\cal F}(\zeta_{*})-f\zeta_{*}]+\frac{1}{2\beta}
388: ln{{\cal F}^{''}(\zeta_{*})}
389: \label{saddle}
390: \end{equation}
391: The free energy due to fluctuations around the saddle point value is
392: $\frac{1}{2\beta}ln{{\cal F}^{''}(\zeta_{*})}$. 
393: Notice that in the long polymer limit of $\beta \rightarrow \infty$,
394: this term vanishes. For finite $\beta$, this nonzero contribution 
395: to the free energy accounts for the transition between the constant
396: extension ensemble and the constant force ensemble.
397: In going from a soft trap to a stiff trap work is done 
398: on the bead by the trap. Similarly in going from a stiff trap to a soft 
399: trap work is extracted from the bead by the trap. The net work done is the 
400: difference between the work done at the two ends
401: of the force-extension curves in switching ensembles.
402: This net work exactly cancels out the nonzero 
403: area enclosed in the force-extension plane. 
404: 
405: {\bf V ``First Order Phase Transition'' and the Thermodynamic Limit in 
406: Semiflexible Polymers}
407: 
408: In Sec. $II$ we considered the bead to be in a potential which was 
409: soft in the $x$ and $y$ directions. Let us now consider what happens
410: when the trap is stiff in all three directions
411: ($A$ as well as $C$ in Eq. ({\ref{trap}}) are large) and the vector 
412: position 
413: of the bead is constrained to be at $(x_0,y_0,z_0)$.
414: Let $Q({\vec r})$ be the number of polymer configurations 
415: which start at the origin and end in the volume element $d{\vec r}$ 
416: centered at ${\vec r}$ \cite{bend,dabhi}. 
417: $Q({\vec r})$ is related to $P(z)$ via the equation 
418: \begin{equation}
419: P(z)=\int{d\vec{r} Q(\vec{r})\delta(r_3 -z)} 
420: \label{pofz}
421: \end{equation}
422: which in words, means that $P(z)$ is obtained by 
423: integrating $Q({\vec r})$ over a plane of constant $z$\cite{bend}.
424: The distribution $Q({\vec r})$ was studied in \cite{dabhi} 
425: where it was noticed that 
426: in an intermediate range (around $3.8$) of $\beta$
427: the free energy 
428: $A({\vec r})= {\frac{-1}{\beta}}log{Q(\vec r)}$ had multiple minima.
429: For a fixed contour length as one varies $\beta$ by tuning the persistence
430: length $L_P$ one finds a competition between flexible and rigid phases of 
431: the polymer for intermediate values of $\beta$. Thus the polymer 
432: undergoes a flexible to rigid ``first order phase transition''
433: via a two peaked profile of $Q({\vec r})$. 
434: This leads to a curious 
435: force-extension relation. 
436: As one pulls the bead, the restoring force at first increases, 
437: then decreases to zero and then goes 
438: negative and becomes a destabilizing force. 
439: The 
440: molecule is unstable and goes to a new extended state. 
441: 
442: The words ``first order phase transition'' above were in quotes as a
443: finite system does not show phase transitions. If one takes the 
444: thermodynamic limit by taking the length of the polymer to 
445: infinity ($\beta \rightarrow \infty$) one loses the multiple minima
446: structure which is present only in a small range of $\beta$ around
447: $3.8$. Is it possible for this phase transition to survive the 
448: thermodynamic limit? As we will see below, this is indeed possible 
449: provided one takes the replica thermodynamic limit. 
450: We take $N$ replicas of the molecule with fixed $\beta$ and let $N$ 
451: tend to $\infty$. Consider $N$ identical polymers with $\beta= 
452: \frac{L}{L_P}$ fixed, their two ends anchored to flat surfaces 
453: $S_1$ and $S_2$  (See Fig. $2$). One could realize the above
454: arrangement by using 
455: $(i)$ two planar arrays of optical traps or
456: $(ii)$ by introducing suitably synthesized supramolecular lamellar
457: structures. 
458: The anchoring is such that the tangent vectors to the molecule 
459: at the fixed ends are free to swivel. If one applies a 
460: force $F$ to pull $S_1$ and $S_2$ 
461: apart the $N$ molecules are also stretched. We consider the molecules
462: to be well separated so that they can be regarded as independent.
463: One could now look
464: at the mean force $F$ needed to maintain the separation $r$. It is easily
465: seen
466: that the force is proportional to $N$ and also the mean square
467: fluctuation
468: $<(\Delta F)^2>$ in the force is proportional to $N$. 
469: This is because the mean force and its variance are 
470: respectively the first and second derivatives of the free energy,
471: which being an extensive quantity is proportional to $N$. 
472: %Heuristically, this
473: %is because of cancellations of the force fluctuations from different
474: %polymers.
475: It follows that as
476: $N$ goes to infinity, the fluctuations ($\Delta F/F$) in $F$ die out as
477: $1/\sqrt{N}$. We can now regard the mean force $F$ as a control parameter
478: (i.e. consider a constant $F$ ensemble)
479: and observe that if we tune the applied force $F$, there is a
480: discontinuous change in the
481: separation $r$ between the two sets of optical traps 
482: signalling a first order phase transition with the inter-trap separation
483: $r$ as the order parameter. Thus, 
484: in the replica way of taking the thermodynamic limit 
485: the double humped
486: form of the distribution function  $Q({\vec r})$ results in a true first 
487: order phase transition. 
488: 
489: \vbox{
490: %\vspace{-0.5cm}
491: \epsfxsize=6.0cm
492: \epsfysize=6.0cm
493: \epsffile{figure2.eps}
494: \begin{figure}
495: \caption{Schematic Experimental Design for Replica Thermodynamic Limit,
496: shown above for 
497: $N=4.$}
498: \label{replica}
499: \end{figure}}
500: 
501: {\bf VI Discussion:}
502: 
503: In this paper we point out the importance of considering the free energy 
504: of transition in going between the Helmholtz and the Gibbs ensembles
505: in the context of single molecule force-extension measurements. 
506: We also study two distinct ways of taking the thermodynamic limit
507: $(i)$ by letting the length of the polymer tend to infinity
508: (i. e. $\beta \rightarrow \infty$) and 
509: $(ii)$ by considering replicas. In particular we notice that 
510: the flexible to rigid transition mentioned in Sec. $V$ for a 
511: single semiflexible polymer survives the replica thermodynamic limit. 
512: In contrast, this feature disappears in the 
513: usual thermodynamic limit of $\beta \rightarrow \infty$. 
514: 
515: The non-monotonic behavior of $Q(\vec r)$ is an intriguing feature of 
516: molecular elasticity. 
517: It was noticed in recent simulations\cite{dabhi} and subsequently in 
518: a semi-analytical treatment\cite{bend}. It was also commented on in a 
519: recent paper\cite{Kleinert2}.
520: This double humped form of $Q({\vec r})$
521: has gained renewed interest in the context of cyclization 
522: of polymers and its significance to gene regulation \cite{gautam,CELL}.
523:  
524: 
525: We note that this remarkable qualitatively
526: distinctive feature is special to the isometric ensemble (fixed vector
527: end-to-end separation) and is not shared by other ensembles. 
528: In other ensembles one can argue generally that
529: such non-monotonic behavior cannot occur. For example, the conjugate
530: distribution 
531: ${\tilde P}({\vec F})=\int Q({\vec r})e^{{\vec F}.{\vec 
532: r}}d{\vec r}$ 
533: is monotonic. This follows from noticing that the $3\times 3$ matrix  
534: \begin{equation}
535: \frac{\partial{ln {\tilde P}}}{\partial{F_i}\partial{F_j}}
536: =<(r_i -<r_i>)(r_j -<r_j>)> 
537: \label{nonmon}
538: \end{equation}
539: is positive definite. Such 
540: arguments do not apply to $Q({\vec r})$ since there is no analogous
541: formula to Eq. (\ref{softrap}) in the conjugate distribution. Indeed,
542: if there were, a double humped form could not appear in $Q({\vec r})$,
543: for one could express the second derivative of $Q({\vec r})$ as the 
544: variance of the force.
545: Can $P(z)$ 
546: show non-monotonic behavior? The answer is no,
547: for it has been shown in \cite{bend} that  
548: $${\frac{-2}{z}}{\frac{d P}{d z}} = Q(z)$$
549: Since $Q(\vec r)$ is a probability density and therefore non-negative,
550: it follows that $\frac{d P}{d z} \leq 0$ for $z>0$, thus 
551: ruling out multiple peaks in $P(z)$. 
552: This argument which rules out multiple peaks in $P(z)$ is a new 
553: observation. 
554: Note that a $P(z)$ measurement differs from $Q(\vec r)$ only in the 
555: transverse stiffness $A$ of the trap. By tuning $A$ we can permit 
556: fluctuations in the transverse direction and therefore 
557: destroy the 
558: phase transition present in the stiff $A$ limit.
559: One would expect to see a critical stiffness $A=A_c$ below which the 
560: phase transition is destroyed. Alternately, one could tune the mean 
561: force and expect to see the phase transition vanishing below a critical 
562: mean force $F=F_c$ for a fixed value of $\beta$ in the intermediate
563: range of $\beta$.  
564: We emphasize that the non-monotonic 
565: features of $Q({\vec r})$ in the semiflexible range $\beta \approx 3.8$
566: are predictions of the WLC model which can be tested against experiments.
567: A single molecule with its ends confined in optical traps is expected
568: to show this flexible to rigid transition. The effect can be dramatic
569: however, if a large number of molecules co-operatively show such a 
570: transition. One could attach the ends of 
571: a collection of semiflexible
572: polymers to supramolecular 
573: layers\cite{drozdov} and detect the 
574: flexible to rigid
575: transition signalled by a change in the inter-layer 
576: separation via a suitable probe. 
577: It may be possible to exploit this dramatic transition from flexible to 
578: rigid behavior in technological applications.   
579: 
580: If one considers the replica thermodynamic limit of a semiflexible polymer
581: one sees that force-extension curves continue to remain distinct in 
582: the Helmholtz and the Gibbs ensembles. 
583: So while interpreting a force-extension curve obtained for a collection 
584: of semiflexible polymers suspended between two arrays of traps,
585: one needs to know if the curve is obtained using
586: the soft trap setup or a stiff trap setup.
587: However, for $\beta \rightarrow 
588: \infty$, which is the usual thermodynamic limit, the two ensembles 
589: give rise to the {\it same} force-extension curve.
590: This observation is consistent with 
591: the fact that inequivalence of ensembles can 
592: survive 
593: at the thermodynamic limit for systems with long range 
594: interactions\cite{kreuz,keller,Lynd}. 
595: In the context of semiflexible polymers,
596: the persistence length $L_P$ plays the role of 
597: the range of interactions.
598: There has been some work\cite{baro} on the thermodynamics
599: of particle systems in the presence of external macroscopic 
600: fields in classical and quantal contexts. In these papers
601: the authors have dealt with the macroscopic limit of the 
602: definition of pressure which is analogous to the thermodynamic 
603: limit of the definition of force in our work. 
604: In particular the authors of Ref. \cite{baro} discuss the 
605: connection between the values of the
606: pressure defined by two different thermodynamic limit procedures:
607: in the first, the system is confined successively in a
608: sequence of boxes which grows to fill up the whole space.  In the 
609: second,
610: the system is in an external potential 
611: similar to the present context.
612: In the case of a
613: semiflexible polymer the force corresponding to a given extension is 
614: the same in the Gibbs and Helmholtz ensembles only in the thermodynamic
615: limit of $\beta \rightarrow \infty$. This is analogous to the second
616: procedure of taking the thermodynamic limit 
617: given in \cite{baro}, where they recover the thermodynamic notion 
618: of pressure from an underlying microscopic definition when they 
619: let the scale factor go to infinity in the macroscopic limit. 
620: The results of this paper are therefore, consistent with the general 
621: treatment in \cite{baro}. 
622: 
623: The fluctuation effects mentioned here also 
624: have some biological significance.
625: In particular, the process of gene regulation involves interaction
626: between parts of a DNA molecule which are about 
627: less than one persistence length
628: apart ($\approx 34 nm$)\cite{CELL}. Over such short segments of the 
629: DNA fluctuation effects would be significant.
630: The replica thermodynamic
631: limit would also play a role in the concrete biological 
632: context of a network of  
633: actin filaments forming the cytoskeletal structure. 
634: 
635: 
636: {\it Acknowledgements:}
637: We thank Dipanjan Bhattacharya for drawing attention to Ref. 
638: \cite{keller}
639: and Abhishek Dhar, Deepak Dhar, Erwin Frey and Richard Neumann for 
640: discussions.
641: We thank one of the referees for drawing attention to Ref.\cite{baro}.
642: We thank Harish-Chandra Research Institute for their hospitality and 
643: the wonderful working atmosphere in which this paper was written.
644: \begin{references}
645: %\vspace{-1.5cm}
646: \bibitem[\dagger]{SUP} e-mail\,:supurna@mri.ernet.in
647: \bibitem[\star]{SAM} e-mail\,:sam@mri.ernet.in
648: \bibitem{bulk} M. Doi, {\it Introduction to Polymer Physics}, Oxford,
649: Clarendon Press (1996). 
650: \bibitem{bust} C. Bustamante et al,
651: Current Opinion in Structural Biology {\bf 10}, 279 (2000).
652: \bibitem{Strick}T. R. Strick et al, {\it Science} {\bf 271}, 1835 (1996).
653: \bibitem{twist} J. Samuel and S. Sinha, {\it Phys. Rev. Lett.} {\bf 90},
654: 098305 (2003).
655: \bibitem{writhe} Supurna Sinha, {\it Physical Review E} {\bf 70 },
656: 011801 (2004). 
657: %\bibitem{Jarz} C. Jarzynski, {\it Phys. Rev. Lett.} {\bf 78}, 
658: %2690 (1997); G. Gallavotti and E. G. D. Cohen, {\it Phys. Rev. Lett.}
659: %{\bf 74}, 2694 (1995); G. E. Crooks, {\it J. Stat. Phys.} {\bf 90}, 1481
660: %(1998).
661: \bibitem{landau} L. D. Landau and E. M. Lifshitz, {\it Statistical 
662: Physics, Part I }, Pergamon Press, {\it 3rd ed.} (1980).  
663: \bibitem{kreuz}
664: H. J. Kreuzer and S. H. Payne, {\it Phys. Rev.} {\bf E 63}, 021906 (2001).
665: \bibitem{keller} D. Keller, D. Swigon and C. Bustamante, 
666: {\it Biophysical Journal} {\bf 84}, 733 (2003).
667: \bibitem{neumann} Richard M. Neumann, 
668: {\it Biophysical Journal} {\bf 85}, 3418 (2003) and references therein.
669: \bibitem{kratky}
670: O. Kratky and G. Porod, {\it Recl. Tral. Chim,} {\bf 68}, 1106 (1949).
671: \bibitem{bend} J. Samuel and S. Sinha, 
672: {\it Physical Review E}, {\bf 66} 050801(R) (2002). 
673: \bibitem{Kleinert}
674: H. Kleinert, {\it Path Integrals in Quantum Mechanics, Statistics and 
675: Polymer Physics},
676: {\it World Scientific, Singapore}, Ist ed.,1990.
677: \bibitem{Marko}
678: J.Marko and E.D.Siggia, Macromolecules {\bf 28}, 8759 (1995);
679: H.E.Daniels, Proc.R.Soc.Edinb. {\bf 63A}, 29 (1952).
680: \bibitem{partfoot} We drop additive constants in the free energies or
681: equivalently, multiplicative constants in the partition functions.
682: \bibitem{frey}
683: J. Wilhelm and E. Frey, {\it Phys. Rev. Lett.} {\bf 77}, 2581 (1996). 
684: \bibitem{loic}
685: L. Le Goff, O. Hallatschek, E. Frey and F. Amblard,
686: {\it Phys. Rev. Lett.} {\bf 89}, 258101 (2002).    
687: \bibitem{foot}In fact we have illustrated this point explicitly in the 
688: context of the rigid rod limit in Sec. $III$.
689: \bibitem{dabhi} A. Dhar and D. Chaudhuri, {\it Phys. Rev. Lett.} {\bf 89}, 
690: 065502-1 (2002).
691: \bibitem{Kleinert2}
692: B. Hamprecht, W. Janke and H. Kleinert, cond-mat/0307530.
693: \bibitem{gautam} P. Ranjith, P. B. Sunil Kumar and G. I. Menon
694: cond-mat/0411723.
695: \bibitem{CELL} B. Alberts et al, {\it Molecular Biology Of The Cell}
696: (Garland Publishing, New York, 1994), 3rd ed; T. E. 
697: Cloutier and J. Widom, {\it Mol. Cell}, {\bf 14}, 355 (2004).
698: \bibitem{drozdov}
699: A. D. Drozdov, cond-mat/0408474.
700: \bibitem{Lynd}
701: R. M. Lynden-Bell, {\it Mol. Phys.} {\bf 86}, 1353 (1995);
702: J. Barre, D. Mukamel and S. Ruffo, {\it Phys. Rev. Lett} {\bf 87}, 030601
703: (2001);
704: F. Gulminelli and Ph. Chomaz, {\it Physical Review E}, {\bf 66}, 046108 
705: (2002).
706: \bibitem{baro} C. Marchioro and E. Presutti,{\it Communications
707: in Mathematical Physics}, {\bf 27}, 146 (1972);
708: C. Marchioro and E. Presutti,{\it Communications
709: in Mathematical Physics}, {\bf 29}, 265 (1973).
710: \end{references}
711: \end{document}
712: 
713: