1: \documentclass[pre,twocolumn,showpacs,preprintnumbers,superscriptaddress,floatfix]{revtex4}
2: \usepackage{graphicx,citesort,psfrag,amsmath,amssymb}
3: \usepackage[dvips]{color}
4:
5: \renewcommand{\vec}[1]{\pmb{#1}}
6: \newcommand{\avg}[1]{\left\langle #1\right\rangle}
7: \newcommand{\spd}[2]{\left( #1|#2\right) }
8: \newcommand{\abs}[1]{\left\lvert #1\right\rvert}
9: \newcommand{\mint}[4]{\int_{#2}^{#3}\!\!#1\,#4}
10: \newcommand{\rem}[1]{\textbf{#1}}
11: \newcommand{\kil}[1]{{\scriptsize #1}}
12: \newcommand{\Ord}[1]{{\cal O}(#1)}
13: \newcommand{\lOrd}[1]{{\cal o}(#1)}
14: \newcommand{\ord}[1]{{\text{ord}}(#1)}
15: \newcommand{\pe}{\perp}
16: \newcommand{\pa}{\parallel}
17: \newcommand{\rpe}{r_\perp}
18: \newcommand{\rpa}{r_\parallel}
19: \newcommand{\dpe}{\avg{\delta\rpe^2}}
20: \newcommand{\dpa}{\avg{\delta R_\pa^2}}
21: \newcommand{\epe}{\hat{\vec e}_\pe}
22: \newcommand{\epa}{\hat{\vec e}_\pa}
23: \newcommand{\cpe}{\chi_\pe}
24: \newcommand{\tv}{\vec {\mathfrak t}}
25: \newcommand{\nv}{\vec {\mathfrak n}}
26: \newcommand{\bv}{\vec {\mathfrak b}}
27:
28: \newcommand{\kr}{\chi}
29: \newcommand{\ts}{\tau}
30:
31: \newcommand{\fopt}{f^{\text{OPT}}}
32:
33: \newcommand{\fe}{\mathfrak f \,}
34: \newcommand{\efe}{\hat{\vec e}_\fe}
35: \newcommand{\fel}{\vec{f}_{\text{el}}\,}
36: \newcommand{\gex}{\mathfrak g \,}
37: \newcommand{\gext}{\mathfrak g_0}
38: \newcommand{\gfl}{\vec{ g}_{\text{fr}}}
39: \newcommand{\gel}{\vec{ g}_{\text{el}}}
40:
41: \newcommand{\Hext}{{\cal H}_{\text{ext}}}
42: \newcommand{\Hb}{{\cal H}_{\text{WLC}}}
43: \newcommand{\Ht}{{\cal H}_{\text{tors}}}
44: \newcommand{\Hc}{{\cal H}_{\text{comp}}}
45: \newcommand{\Htot}{{\cal H}}
46:
47: \newcommand{\msd}{\text{MSD}}
48: \newcommand{\msdl}{\text{msd}}
49:
50: \newcommand{\fc}{c}
51: \newcommand{\ludo}{\ell_\text{SWN}}
52: \newcommand{\lralf}{\ell_\text{EJAM}}
53: \newcommand{\lpierre}{\ell_\text{BBG}}
54: \newcommand{\lflower}{\ell_\text{Flower}}
55: \newcommand{\lp}{\ell_p}
56: \newcommand{\lpe}{\ell_\perp}
57: \newcommand{\lpa}{\ell_\|}
58: \newcommand{\tfin}{t_L^\|}
59: \newcommand{\tf}{t_\fe}
60: \renewcommand{\sf}{s_\fe}
61: \newcommand{\tot}{\tau_{\text{tot}}}
62: \newcommand{\totpe}{t_L^\pe}
63: \newcommand{\totpa}{t_L^\pa}
64:
65: \newcommand{\eom}{equations of motion}
66: \newcommand{\leom}{longitudinal equation of motion}
67: \newcommand{\teom}{transverse equation of motion}
68: \newcommand{\rhs}{right hand side}
69:
70: \newcommand{\crr}[1]{\color{red}{#1 }\color{black}}
71: \newcommand{\crb}[1]{\color{blue}{#1 }\color{black}}
72:
73: \newcounter{fnt}
74:
75: \newcommand{\pull}{\text{\emph{Pulling}}}
76: \newcommand{\epull}{\text{\emph{Towing}}}
77: \newcommand{\lelo}{\text{\emph{Release}}}
78: \newcommand{\tjump}{\text{$\lp$--\emph{Quench}}}
79: \newcommand{\push}{\text{\emph{Pushing}}}
80:
81:
82: \DeclareMathOperator{\sgn}{sgn}
83: \DeclareMathOperator{\erf}{erf}
84:
85:
86: \sloppy
87:
88: \preprint{LMU-ASC 60/06}
89:
90: \begin{document}
91: \bibliographystyle{apsrev}
92:
93:
94: \title{Tension dynamics in semiflexible polymers. \\Part~II: Scaling
95: solutions and applications }
96:
97:
98: \author{Oskar Hallatschek} \email{ohallats@fas.harvard.edu}
99: \affiliation{Lyman Laboratory of Physics, %
100: Harvard University, Cambridge, Massachusetts 02138, USA}
101:
102: \author{Erwin Frey} \affiliation{Arnold Sommerfeld Center for Theoretical Physics
103: and Center for NanoScience, %
104: LMU M\"unchen, Theresienstr.~37, 800333 M\"unchen,
105: Germany}
106:
107: \author{Klaus Kroy}
108: \affiliation{Institut f\"ur Theoretische Physik,%
109: Universit\"at Leipzig, Augustusplatz 10/11, 04109 Leipzig, Germany}
110:
111:
112:
113: \date{\today}
114:
115: \begin{abstract}
116: In Part~I of this contribution, a systematic coarse-grained
117: description of the dynamics of a weakly-bending semiflexible polymer
118: was developed. Here, we discuss analytical solutions of the
119: established deterministic partial integro-differential equation for
120: the spatio-temporal relaxation of the backbone tension. For
121: prototypal experimental situations, such as the sudden application
122: or release of a strong external pulling force, it is demonstrated
123: that the tensile dynamics reflects the self-affine conformational
124: fluctuation spectrum in a variety of intermediate asymptotic power
125: laws. Detailed and explicit analytical predictions for the tension
126: propagation and relaxation and corresponding results for common
127: observables, such as the end-to-end distance, are obtained.
128: \end{abstract}
129:
130:
131:
132: \pacs{ 87.15.He, 87.15.Aa, 87.16.Ka, 83.10.-y}
133:
134:
135:
136: \maketitle
137:
138:
139:
140: \section{Introduction}
141: \label{sec:intro}
142: Polymer physics has traditionally focused on very flexible polymers that admit a
143: highly coarse-grained description in terms of Gaussian chains and exhibit universal
144: physical behavior that can be explained using methods from statistical mechanics such
145: as the renormalization group and scaling arguments~\cite{degennes:79}. Further
146: research has explored new terrain that lies beyond the realm of applicability of this
147: highly successful approach either because the polymers of interest are too stiff or
148: because they are subject to extreme forces. Many of these instances have recently
149: appeared in applications involving biopolymers. Notorious examples are the nonlinear
150: mechanical response of DNA~\cite{BustamanteBS03}, which has turned out to be pivotal
151: to protein-DNA interactions, and the problem of force transduction through the
152: cytoskeleton~\cite{JanmeyW04,YasuhiroSawada02182002,GaborForgacs06012004}, which is a
153: major mechanism by which cells explore their environment and react to external
154: mechanical stimuli. Clearly, in neither of these situations can theorists contend
155: themselves with the convenient Gaussian chain representation, but have to resort to
156: more realistic, yet still schematic descriptions, such as the freely-jointed chain
157: (e.g.\ for single-stranded DNA) or the wormlike chain model (for double-stranded DNA,
158: F--actin, microtubules etc.)~\cite{doi-edwards:86,degennes:79}.
159:
160: Suspicions that this might entail a substantial loss of universality and render
161: systematic analytical approaches forbiddingly complex have turned out to be
162: unfounded. The wormlike chain model provides an analytically tractable standard model
163: for many of the above mentioned new applications, in particular for calculating the
164: non-equilibrium dynamical response of stiff and semiflexible but weakly bending
165: polymers to strong external fields. As established in Part~I of this contribution,
166: the weakly bending wormlike chain lends itself to a multiple-scale perturbation
167: theory (MSPT) based on a length scale separation between longitudinal and transverse
168: dynamic correlation lengths. In the present Part~II, we demonstrate that the
169: self-affine roughness, acquired by the weakly-bending contour in thermal equilibrium,
170: plays an analogous role as the more familiar fractal conformational correlations in
171: the case of flexible polymers~\cite{degennes:79}. The self-similarity of the static
172: conformational fluctuations entails self-similar dynamics. It manifests itself in a
173: variety of \emph{intermediate asymptotic dynamic power laws}. Apart from the
174: restriction to polymers with a (locally) rodlike structure, these predictions are as
175: universal as those of classical polymer physics. They are moreover derived in a
176: direct way, usually including exact amplitudes, from a controlled perturbation
177: expansion.
178:
179:
180:
181: As a major result of the multiple scale theory developed in Part~I, we
182: obtained a coarse-grained reformulation of the free-draining Langevin
183: equations of motion of a weakly bending rod in the form of the
184: deterministic equation
185: \begin{equation}
186: \label{eq:cg-eom}
187: \partial_s^2 F(s,t)=-\zeta_\pa \avg{\Delta
188: \overline{\varrho}}\left[F(s,\tilde t \leq t),t\right] \;.
189: \end{equation}
190: It describes the long wave-length (all time) dynamics of the
191: time-integrated tension
192: \begin{equation}
193: \label{eq:integrated-tension}
194: F(s,t)\equiv\mint{dt'}{0}{t}f(s,t')
195: \end{equation}
196: with $\zeta_{\parallel}$ being the friction coefficient for longitudinal motion and
197: $\avg{\Delta \overline{\varrho}}\left[F(s,\tilde t \leq t),t\right]$ the average
198: release of contour length stored in the transverse undulations up to time $t$.
199: Written out in terms of the transverse normal mode contributions, the latter reads
200: \begin{eqnarray}
201: \avg{\Delta\overline{\varrho}}(t)&\equiv& \mint{\frac{dq}{\pi\lp}}{0}{\infty}
202: \left\{\frac{1}{q^2+f_<}\left(e^{-2q^2[\kappa q^2 t+ F(t)]/\zeta_\perp}-1\right) \right.\nonumber\\
203: &&\left.+2 q^2\mint{d\tilde t}{0}{t}e^{-2q^2\left[\kappa q^2(t-\tilde
204: t)+F(t)-F(\tilde t)\right]/\zeta_\perp}\right\},
205: \label{eq:change-stored-length}
206: \end{eqnarray}
207: where $\kappa$, $\lp=\kappa/(k_B T)$ and $\zeta_\perp$ are the bending stiffness,
208: persistence length and friction coefficient for transverse motion, respectively. The
209: parameter $f_<\equiv f(t<0)=$const.~allows to take into account a constant
210: prestress~\footnote{The parameter $\theta$ occurring in the dynamic force extension
211: relation of Part~I will be set to one, $\theta=1$, throughout this paper. It
212: describes the effect of sudden changes in persistence length, which will be
213: discussed elsewhere~\cite{obermayer-kroy-frey-hallatschek3:tbp}.}. Upon inserting
214: this dynamical force-extension relation into Eq.~(\ref{eq:cg-eom}), we arrived at our
215: central result, the closed partial integro-differential equation (PIDE) for the
216: time-integrated tension $F(s,t)$,
217: \begin{eqnarray}
218: \label{eq:pide}
219: &\partial_s^2 F(s,t)= \hat \zeta
220: \mint{\frac{dq}{\pi\lp}}{0}{\infty}
221: \left\{\frac{1}{q^2+f_<}\left(1-e^{-2q^2[ q^2 t+ F(s,t)]}\right)
222: \right.& \nonumber \\
223: &\left. -2 q^2\mint{d\tilde
224: t}{0}{t}e^{-2q^2\left[ q^2(t-\tilde t)+F(s,t)-F(s,\tilde
225: t)\right]}\right\}\;.&
226: \end{eqnarray}
227: For convenience, we have made the following choice of units: Time and
228: tension, respectively, are rescaled according to
229: \begin{eqnarray}
230: \label{eq:rescaling-time-tension}
231: t&\to&\zeta_\perp t/\kappa \;, \\
232: f&\to& \kappa f\;.
233: \end{eqnarray}
234: This corresponds to setting $\kappa\equiv\zeta_\pe\equiv 1$ and $\hat
235: \zeta\equiv 1/2=\zeta_\pa$. As a consequence all variables represent
236: powers of length, e.g., $t$ and $f$ are a length$^4$ and a
237: length$^{-2}$, respectively.
238:
239: To leading order in the small contour undulations, the deterministic coarse-grained
240: tension dynamics, described by Eq.~(\ref{eq:pide}), together with the microscopic
241: transverse equation of motion, represents a valid reformulation of the constrained
242: Langevin dynamics of a weakly bending rod subject to a putative pre-stress. The PIDE
243: Eq.~(\ref{eq:pide}) is the basis not only for discussing the tension dynamics itself,
244: but also the starting point for analytical and numerical calculations of the
245: longitudinal and transverse nonlinear response of a weakly bending polymer. It is the
246: purpose of the present Part~II to treat the former case in detail, while the latter,
247: somewhat more complex case is reserved for a future
248: communication~\cite{obermayer-hallatschek2:tbp}. In Sec.~\ref{sec:driving}, we derive
249: detailed solutions to Eq.~(\ref{eq:pide}) for idealized experimental protocols
250: involving a representative selection of external fields. The analytical scaling
251: solutions obtained for a semi-infinite polymer suddenly pulled (or released) at its
252: end, reveal the non-trivial short-time phenomenon of tension propagation
253: (Sec.~\ref{sec:tp}). At long times, the finite contour length comes in as an
254: additional characteristic length scale, which gives rise to additional scaling
255: regimes, discussed in Sec.~\ref{term-rel-tensprop}. To make contact with experiments,
256: we finally identify the repercussions of the tension dynamics on pertinent
257: observables like the (projected) end-to-end distance (Sec.~\ref{sec:MSDII}) and
258: comment on novel experimental perspectives brought up by our analysis
259: (Sec.~\ref{sec:possible-exps}).
260:
261:
262:
263: \section{Generic longitudinal driving forces}
264: \label{sec:driving}
265: In general, the tension dynamics depends on how the filament is driven externally,
266: i.e., on the boundary and initial conditions imposed on Eq.~(\ref{eq:pide}). With
267: the definition of generic experimental force protocols, this section shall provide a
268: framework for the analysis of Eq.~(\ref{eq:pide}). We introduce the scenarios
269: {\pull}, {\epull}, and {\lelo} and report on existing investigations. Apart from
270: being directly relevant for experiments, we have chosen to consider these scenarios
271: because of two properties that render them attractive from a theoretical perspective.
272: Firstly, they correspond to \emph{sudden} changes of the environment and thus do not
273: introduce an additional delay time scale. Secondly, since external forces are
274: assumed to act at the ends, these scenarios only change boundary conditions and leave
275: the equations of motion unchanged. In more complicated scenarios, that involve
276: forces applied not only at the ends, these problems show up as subproblems. For
277: instance, if a single point force is applied somewhere within the bulk of a polymer,
278: the filament can be partitioned into two sections that perceive the external force
279: only at their ends.
280:
281:
282:
283: \subsubsection{{\pull}}
284: \label{sec:pull-def}
285: The polymer is supposed to be free for negative times, such that it is equilibrated
286: under zero tension at time zero, i.e., we require
287: \begin{equation}
288: \label{eq:pulling-ic}
289: f_<=f(s,t<0)=0 \;.
290: \end{equation}
291: Then, for positive times the polymer is pulled in longitudinal direction at both ends
292: with a constant force $\fe$~\footnote{Throughout, we denote \emph{external} forces or
293: force fields by fraktur letters.}. The corresponding external force density
294: $\fe\left[\delta(s-L)-\delta(s)\right]$ provides the boundary conditions for the
295: tension (see, e.g., the longitudinal equation of motion in Part~I)
296: \begin{equation}
297: \label{eq:pulling-bc}
298: f(s=0,t>0)=\fe \;,\; f(L,t>0)=\fe \;.
299: \end{equation}
300:
301:
302:
303: {\pull} was first considered by Seifert et al.~\cite{seifert-wintz-nelson:96} (SWN).
304: They predict that a ``large'' tension spreads within a time $t$ a characteristic
305: length $\lpa(t)=\ludo(t)\equiv \lp^{1/2} (\fe t)^{1/4}$ from the ends into the bulk
306: of the filament. Their analysis neglects bending forces and thermal forces for the
307: dynamics, albeit the self-affine thermal initial conditions are used
308: (\emph{taut-string approximation}). The contribution of Everaers et
309: al.~\cite{everaers-Maggs:99} (EJAM) shed light on the linear response to longitudinal
310: forces. Their simulations established a typical propagation length of
311: $\lpa(t)=\lralf(t)\equiv \lp^{1/2} t^{1/8}$ for weak forces, which was previously
312: predicted by Morse~\cite{morse:98II}, and made it plausible by scaling arguments.
313: Brochard--Wyart et al.~\cite{brochard-buguin-de_gennes:99} (BBG) proposed a theory
314: for tension propagation claimed to be valid on scales much larger than $\lp$
315: supposing, however, the weakly bending approximation. A \emph{quasi-static
316: approximation} underlies their analysis, in which the polymer is at any instant of
317: time considered to be equilibrated with the local tension. Applying their results to
318: the situation considered here, tension should propagate a distance
319: $\lpa(t)\propto\lpierre(t)\equiv \lp^{1/2} \fe^{3/4} t^{1/2}$.
320:
321: Naturally, the scaling arguments used to predict the three different scaling regimes
322: did not address the crossover and the range of validity. Below, we show that in fact
323: only two scaling regimes exist.
324:
325: \subsubsection{{\epull}}
326: \label{sec:epull-def}
327:
328: Pulling of a filament can also be studied for time dependent external forces. A
329: dynamic force protocol of particular experimental relevance is given by the constant
330: velocity ensemble: The polymer is pulled by a time-dependent external force $\fe(t)$
331: at the left end such that this end moves with a \emph{constant velocity} $v$
332: ({\epull}). Besides the growth law of the boundary layer, we wish to understand the
333: time dependence of the external force. We will see that both quantities are
334: proportional to each other because the external force essentially has to drag a
335: polymer section of length $\lpa(t)$ through the viscous solvent with the constant
336: velocity $v$, hence $\fe(t)\simeq \hat \zeta v \lpa(t)$. By measuring the
337: time-dependent force in a constant velocity experiment one can thus directly monitor
338: $\lpa(t)$. Possible experimental realizations are outlined in
339: Sec.~\ref{sec:possible-exps}.
340:
341: The external force density field corresponding to {\epull} is given by $-\fe(t)
342: \delta(s)$ with an external force $\fe(t)$ determined to fulfill the requirement that
343: $\partial_t r_\pa(0,t)=v$. Recall from the equations of motion derived in Part~I,
344: that, up to terms of order $\Ord{\epsilon}$, the gradient of the tension is given by
345: the longitudinal friction (as in a rigid rod). This implies the boundary condition
346: \begin{eqnarray}
347: \label{eq:efpulling-bc}
348: \partial_s f(s=0,t>0)&=&-\hat\zeta v+\Ord{\epsilon} \\
349: f(L,t>0)&=&0\;. \nonumber
350: \end{eqnarray}
351:
352:
353: \subsubsection{{\lelo}}
354: \label{sec:lelo-def}
355:
356: {\lelo} refers to the process ``inverse'' to {\pull}: the filament is supposed to be
357: equilibrated at $t=0$ under a constant pulling force,
358: \begin{equation}
359: \label{eq:let-loose-ic}
360: f_<=f(s,t<0)=\fe>0 \;.
361: \end{equation}
362: Then, at $t=0$, the external force is suddenly switched off and the filament begins to
363: relax. The ends are considered to be free for $t>0$,
364: \begin{equation}
365: \label{eq:let-loose-bc}
366: f(s=0,t>0)=0 \;,\; f(L,t>0)=0 \;.
367: \end{equation}
368: {\lelo} has been discussed by Brochard et al.~\cite{brochard-buguin-de_gennes:99}. According to that work
369: the characteristic size of the boundary layer, where the tension is appreciably
370: decreased from $\fe$, should be given by $\lpa(t)\propto\lpierre(t)$ (the same as for
371: {\pull}).
372:
373: Furthermore, Brochard et al.~predict that the tension is relaxed as soon as the
374: tension has spread over the whole filament yielding a relaxation time $\totpa$ for
375: the tension that satisfies $\lpa(\totpa)=L$. This is in conflict with what we will
376: find Sec.~\ref{term-rel-tensprop}, where we identify a novel scaling regime of
377: homogeneous tension relaxation.
378:
379: \section{Tension propagation}
380: \label{sec:tp}
381: To unravel the physical implications of Eq.~(\ref{eq:pide}) for the scenarios
382: introduced above, we begin with the tension propagation regime $\lpa\ll L$, where the
383: total length $L$ of the polymer is irrelevant. In this regime, it is legitimate to
384: discuss the dynamics on a (formally) semi-infinite arc length interval $[ 0,
385: \infty[$. Problems like {\pull} and {\lelo} still depend on four independent length
386: scales ($\lp, \fe^{-1/2},s , t^{1/4}$). Yet, it is shown in
387: Sec.~\ref{sec:scaling-forms} that, Eq.~(\ref{eq:pide}) is \emph{solved exactly} by a
388: tension profile that obeys a crossover scaling form depending on only two arguments,
389: which can be identified as a reduced time and arc length variable, respectively. In
390: Sec.~\ref{sec:asymptotic-growth}, we then argue that for asymptotically short ($<$)
391: and long ($>$) times the scaling function reduces to a function of only one scaling
392: variable $\xi^{\gtrless}=s/\lpa^{\gtrless}(t)$. Our major results concerning tension
393: propagation, particularly our classification of tension propagation laws
394: $\lpa^{\gtrless}(t)$, are summarized in Sec.~\ref{sec:tensprop-summary}.
395:
396:
397:
398: \subsection{Scaling forms}
399: \label{sec:scaling-forms}
400: For each of the generic problems introduced in Sec.~\ref{sec:driving}, we shall see
401: that the tension profile obeys certain crossover scaling forms, that cannot be
402: inferred from dimensional analysis. These scaling forms greatly simplify the further
403: analysis of the tension dynamics by reducing the number of independent parameters.
404:
405: To solve the equation of motion, Eq.~(\ref{eq:pide}), for a given force protocol, we
406: proceed in the following way. A scaling ansatz is postulated and shown to eliminate
407: the parameter dependence in Eq.~(\ref{eq:pide}) and the boundary conditions after a
408: suitable choice of length, time and force scales. These crossover scales turn out to
409: separate two different regimes, a short- and long-time regime,
410: respectively~\footnote{Such a crossover was also found in Part~I for the stored
411: length under a spatially constant tension from ordinary perturbation theory.}.
412:
413: Although being ultimately interested in the tension profile $f(s,t)$, it is
414: convenient to first discuss the \emph{time integrated} tension $F(s,t)$, defined in
415: Eq.~(\ref{eq:integrated-tension}), because the equation of motion for the tension,
416: Eq.~(\ref{eq:pide}), is naturally formulated in terms of $F(s,t)$. The physically
417: more intuitive quantity $f(s,t) =\partial_t F(s,t)$ is extracted afterwards by a
418: differentiation with respect to time.
419:
420: We make the following ansatz for the time integral $F(s,t)$ of the tension
421: \begin{equation}
422: \label{eq:sansatz}
423: F(s,t)=\fe \tf \; \phi\left(\frac s {\sf } , \frac t {\tf} \right) \;,
424: \end{equation}
425: in terms of as yet unknown crossover time and length scales $\tf$ and $\sf$ to be
426: determined below. While a force scale $\fe$ is given explicitely in the case of
427: {\pull} and {\lelo} by the pulling/pre-stretching force, a natural force scale
428: \begin{equation}
429: \label{eq:ext-force-towing}
430: \fe\equiv \hat \zeta v
431: \sf \qquad \text{(\epull)}\:
432: \end{equation}
433: for {\epull} is provided not unless $\sf$ is fixed. The combination of variables in
434: Eq.~(\ref{eq:ext-force-towing}) represents the force necessary to drag a polymer
435: section of length $\sf$ (longitudinally) through the fluid with the imposed towing
436: velocity $v$.
437:
438: As long as the polymer is in equilibrium, $t<0$, the dimensionless scaling function
439: $\phi(\sigma,\tau)$ for the integrated tension is zero for both pulling scenarios,
440: but linearly increasing with time for {\lelo} due to the constant pre-stretching
441: force,
442: \begin{equation}
443: \label{eq:ic-sansatz}
444: \phi(\sigma,\tau<0)\equiv \fc \tau=\left\{{0 \;, \atop \tau \;,}
445: \qquad {{\pull}/ {\epull} \atop \lelo }
446: \right.
447: \end{equation}
448: The constant $c$ entering the initial condition, Eq.~(\ref{eq:ic-sansatz}), is
449: given by $c= 0$ for {\pull}/{\epull} and $c= 1$ for {\lelo}, respectively.
450:
451: Since we expect that the signal of a sudden change at the end of the polymer,
452: i.e.~at $\sigma=0$, takes time to propagate into the bulk of the polymer, which
453: corresponds to $\sigma\to\infty$, we look for solutions that have a time-independent
454: stored length at $\sigma\to\infty$. According to Eq.~(\ref{eq:cg-eom}), this
455: corresponds to the boundary condition of a vanishing curvature of the tension profile
456: at infinity,
457: \begin{equation}
458: \label{eq:bc-infty}
459: \partial_\sigma^2\phi(\sigma\to\infty,\tau>0) =0
460: \;.
461: \end{equation}
462: At the origin, the force, respectively, the gradient of the force are prescribed by
463: the considered experimental setup,
464: \begin{subequations}
465: \label{eq:bc-origin}
466: \begin{eqnarray}
467: \label{eq:bc-origin-1}
468: \phi(\sigma=0,\tau>0)&=&\left\{{\tau \;, \atop 0 \;,}
469: \qquad {{\pull} \atop \lelo } \right. \\
470: \partial_\sigma\phi(\sigma=0,\tau>0) &=&-\tau \;, \qquad {\epull}
471: \label{eq:bc-origin-2}
472: \end{eqnarray}
473: \end{subequations}
474: Inserting the scaling ansatz, Eq.~(\ref{eq:sansatz}), into Eq.~(\ref{eq:pide}) yields
475: after the variable substitutions $q\to q\sqrt{\fe}$, $t\to \tau \tf$ and $\tilde t
476: \to \tilde \tau \tf$
477: \begin{eqnarray}
478: \label{eq:pide-inserted-1}
479: &&\lp \frac{\fe^{3/2} \tf}{\hat \zeta \sf ^2} \partial_\sigma^2
480: \phi(\sigma,\tau) =\\
481: &&\mint{\frac{dq}{2\pi}}{-\infty}{\infty}
482: \left\{\frac{1}{q^2+\fc}\left[1-e^{-2q^2\left[q^2\tau+
483: \phi(\sigma,\tau)\right] \tf \fe^2}\right] \right. \nonumber\\
484: &&\left. -2 q^2 \tf \fe^2\mint{d\tilde
485: \tau}{0}{\tau}e^{-2 q^2 \left[ q^2 (\tau-\tilde
486: \tau)+\phi(\sigma,\tau)-\tilde \tau \phi(\sigma, \tilde
487: \tau)\right] \tf \fe^2} \right\}\;.\qquad \nonumber
488: \end{eqnarray}
489: By fixing the scales $\tf$ and $\sf $ appropriately,
490: \begin{subequations} \label{eq:scales1}
491: \begin{eqnarray}
492: \tf&=&\fe^{-2} \label{eq:scales1-time} \\
493: \sf &=&\hat \zeta^{-1/2} \lp^{1/2} \fe^{-1/4}
494: \label{eq:scales1-length} \;,
495: \end{eqnarray}
496: \end{subequations}
497: we can eliminate the parameter dependence of
498: Eq.~(\ref{eq:pide-inserted-1}),
499: \begin{eqnarray}
500: \label{eq:pide-nondimensional-1}
501: &&\partial_\sigma^2 \phi(\sigma,\tau)
502: =\mint{\frac{dq}{2\pi}}{-\infty}{\infty} \left\{\frac{1}{q^2+\fc
503: }\left[1-e^{-2q^2\left[q^2 \tau+
504: \phi(\sigma,\tau)\right]}\right] \right.\nonumber\\
505: &&\qquad \left. -2 q^2 \mint{d\hat \tau}{0}{\tau}e^{-2
506: q^2 \left[ q^2 (\tau-\hat \tau)+\phi(\sigma,\tau)-
507: \phi(\sigma, \hat \tau)\right] } \right\}\;.
508: \end{eqnarray}
509: Note that for {\epull}, the conditions in Eqs.~(\ref{eq:scales1-time},
510: \ref{eq:scales1-length}) imply the scales
511: \begin{subequations} \label{eq:scales2}
512: \begin{eqnarray}
513: \tf&=&\left(v^2 \hat \zeta \lp \right)^{-4/5}
514: \label{eq:scales2-time} \\
515: \sf&=&\left(v{\hat \zeta}^{3}\lp^{-2}\right)^{-1/5}
516: \label{eq:scales2-length} \qquad ({\epull})\;,
517: \end{eqnarray}
518: \end{subequations}
519: since $\fe$ depends on $\sf$ via Eq.~(\ref{eq:ext-force-towing}).
520:
521:
522:
523:
524: \subsection{Asymptotic scaling}
525: \label{sec:asymptotic-growth}
526: We have thus removed the parameter dependence of the differential equation as well as
527: the boundary and initial conditions. The remaining task is to solve
528: Eq.~(\ref{eq:pide-nondimensional-1}) for $\phi$ under the initial/boundary conditions
529: given by Eqs.~(\ref{eq:ic-sansatz},~\ref{eq:bc-infty},~\ref{eq:bc-origin}), and to
530: extract the tension
531: \begin{subequations}
532: \label{eq:closure-1}
533: \begin{eqnarray}
534: f(s,t)&=&\partial_t F(s,t) \label{eq:closure-a-1} \\
535: &=&\fe \; \varphi\left(\frac s {\sf },\frac t {\tf}\right)
536: \end{eqnarray}
537: \end{subequations}
538: in terms of the scaling function
539: \begin{equation}
540: \label{eq:close-scaling-form-1}
541: \varphi(\sigma,\tau)\equiv\partial_\tau \phi(\sigma,\tau) \;.
542: \end{equation}
543:
544: Although this remaining task of finding a solution is analytically not possible in
545: generality, we now know, at least, that it should be a function of only two
546: variables, an effective space and time variable $\sigma=s/\sf$ and $\tf$,
547: respectively. From the conditions in Eqs.~(\ref{eq:scales1-time},
548: \ref{eq:scales1-length}), it is seen that the scales $\sf $ and $\tf$ are given by a
549: combination of the characteristic force scale $\fe$ (a length$^{-2}$ in our units)
550: and the persistence length, and hence not simply a consequence of dimensional
551: analysis. The significance of these nontrivial scales is that they mark a crossover
552: in the behavior of the tension. For it turns out that, the two-parameter scaling form
553: $\varphi(\sigma,\tau)$ collapses onto one-parameter scaling form in the limit of
554: large and small arguments, i.e.,
555: \begin{equation}
556: \label{eq:scaling}
557: \varphi(\sigma,\tau)\to \tau^\alpha \hat \varphi
558: \left(\frac{\sigma}{\tau^z}\right)
559: \;,\qquad\text{for } \tau{\ll \atop \gg}1
560: \end{equation}
561: with a positive and monotonous scaling function $\hat \varphi(\xi)$
562: that is bounded as $\xi\to \{0,\;\infty\}$ and exponents $\alpha$ and
563: $z$ depending on the problem and the limit, i.e., short or long time
564: limit. Equation (\ref{eq:scaling}) expresses the asymptotic
565: self-similarity of the tension profile: by stretching the tension
566: profile at a given time in the arclength coordinate $\sigma$ one
567: obtains the tension profile at a later time, a property inherited
568: from the self-affine conformational fluctuation spectrum of the weakly
569: bending wormlike chain.
570:
571: Rewriting the scaling variable in Eq.~(\ref{eq:scaling}) as
572: $\sigma/\tau^z\equiv s/\lpa(t)$ identifies the tension propagation length
573: \begin{equation}
574: \label{eq:def-lf}
575: \lpa(t)=\sf \left( \frac{t}{\tf} \right)^z \;.
576: \end{equation}
577: In Tab.~\ref{tab:lf-growth-laws} the actual growth laws $\lpa(t)$ are
578: tabulated depending on the problem and the asymptotic limit.
579:
580: Before deriving these growth laws from an asymptotic analysis of
581: Eq.~(\ref{eq:pide-nondimensional-1}), let us give a simple argument as to what the
582: exponent $z$ should be on short times. As usual, we assume the crossover should occur
583: when the scaling variable $\sigma/\tau^z$ is of order one, hence
584: \begin{equation}
585: \label{eq:crossover-requirement}
586: \lpa^\gtrless(\tf)\simeq\sf \;.
587: \end{equation}
588: If we further assume, that on very short times the propagation length $\lpa^<(\tau)$
589: of the tension should actually be independent of the external force, we can
590: immediately infer
591: \begin{equation}
592: \label{eq:short-time-guess}
593: \lpa^<(t)\propto \lp^{1/2}t^{1/8} \;,
594: \end{equation}
595: which is the correct short time growth law, as will be shown in
596: Sec.~\ref{sec:short-time-asymptotics}.
597:
598:
599:
600: The derivations we present in the following are consistent in the sense that we use
601: assumptions that are a posteriori legitimized by the solutions. In particular, we
602: exploit Eq.~(\ref{eq:scaling}) as a scaling ansatz in order to derive asymptotic
603: differential equations for the tension. Those equations are then shown to be indeed
604: solved by similarity solutions of the postulated type. In addition, let us assume
605: that the exponent $\alpha$ in Eq.~(\ref{eq:scaling}) is larger than $-1/2$,
606: \begin{equation}
607: \label{eq:alpha-assumption}
608: \alpha>-1/2 \;,
609: \end{equation}
610: i.e., the tension should increase (decrease) less rapidly than $\tau^{-1/2}$ for
611: $\tau\to 0$ ($\tau\to\infty$). The assumption is reasonable for {\pull}, because at
612: the ends, we have $\varphi(0,\tau)=1$ and therefore $\alpha=0$ in this case. It turns
613: out that Eq.~(\ref{eq:alpha-assumption}) is correct for all considered problems of
614: tension propagation except for sudden temperature changes, discussed in
615: Ref.~\cite{obermayer-kroy-frey-hallatschek3:tbp}. As a consequence, the
616: approximations that are made in the following do not apply to sudden changes in
617: persistence length, which is an indication that it is an exceptional problem. In
618: Sec.~\ref{term-rel-tensprop}, where we consider the scaling regime \emph{succeeding}
619: tension propagation, we encounter an asymptotic regime of {\lelo} characterized by an
620: exponent $\alpha=-2/3$ as another important exception of
621: Eq.~(\ref{eq:alpha-assumption}).
622:
623: Since the central PIDE Eq.~(\ref{eq:pide-nondimensional-1}) is
624: expressed in terms of the time integrated tension $\phi(\sigma,\tau)$,
625: it is useful for the following discussion to reformulate the scaling
626: assumption, Eq.~(\ref{eq:scaling}), in terms of $\phi$,
627: \begin{equation}
628: \label{eq:scaling-phi}
629: \phi(\sigma,\tau)\to \tau^{\alpha+1} \hat \phi
630: \left(\frac{\sigma}{\tau^z}\right)
631: \;,\qquad\text{for } \tau{\ll \atop \gg}1 \;.
632: \end{equation}
633:
634:
635:
636: \subsubsection{Short times ($t \ll \tf$)}
637: \label{sec:short-time-asymptotics}
638: In case Eq.~(\ref{eq:alpha-assumption}) holds we can linearize
639: Eq.~(\ref{eq:pide-nondimensional-1}) for short times in $\phi$. This is at first
640: sight only correct for small wave numbers that satisfy $q^2\phi=\Ord{ q^2
641: \tau^{\alpha+1}}\ll 1$. However, upon a closer inspection of the region of large
642: wave vectors $q>\tau^{-(\alpha+1)/2}$ that do not allow for a linearization, it is
643: seen that there is another term in the exponent, $q^4 \tau>\tau^{-2\alpha-1}\gg1$
644: (for $\tau\ll1$ and $\alpha>-1/2$), which renders the considered exponential
645: essentially zero. Therefore, we can approximate Eq.~(\ref{eq:pide-nondimensional-1})
646: by
647: \begin{equation}
648: \label{eq:pide-linearized}
649: \begin{split}
650: & \partial_\sigma^2 \phi(\sigma,\tau) \approx \\
651: & \mint{\frac{dq}{2\pi}}{-\infty}{\infty} \left\{\frac{1}{q^2+\fc}
652: \left[1-\left(1-2q^2\phi(\sigma,\tau)\right)e^{-2q^4
653: \tau}\right] \right.\\
654: & \left. -2 q^2 \mint{d\hat
655: \tau}{0}{\tau}\left[1-2q^2\left(\phi(\sigma,\tau)-\phi(\sigma,\hat
656: \tau)\right)\right]e^{-2 q^4 (\tau-\hat \tau)} \right\}\\
657: &\qquad\qquad=\mint{\frac{dq}{2\pi}}{-\infty}{\infty}
658: \left\{-\frac{\fc}{q^2(q^2+\fc)}\left(1-e^{-2q^4\tau}\right)
659: \right. \\
660: &\qquad\qquad \left. +2\phi(\sigma,\tau)\left[1-\left(\frac{\fc}
661: {q^2+\fc}\right)e^{-2q^4\tau}\right]\right.\\
662: &\qquad\qquad\left.- 4q^4 \mint{d\hat \tau}{0}{\tau}\phi(\sigma,\hat \tau) e^{-2
663: q^4 (\tau-\hat \tau)} \right\}\;.
664: \end{split}
665: \end{equation}
666: Since $\tau\ll1$ we can neglect the parameter $\fc$ in the denominator of the first
667: term and set the exponential function in the second term equal to one. Furthermore,
668: we observe that the lower bound of the time-integral can be set to $-\infty$ by
669: defining $\phi(\sigma,\tau<0)\equiv 0$. After the variable substitution $\hat \tau
670: \to \hat \tau+\tau$ we obtain
671: \begin{equation}
672: \label{eq:pide-linearized-11}
673: \begin{split}
674: &\partial_\sigma^2 \phi(\sigma,\tau)
675: =\mint{\frac{dq}{\pi}}{-\infty}{\infty} \left[\fc \frac{e^{-2q^4
676: \tau}-1}{2q^4}+
677: \right. \\
678: &\left.\frac{q^2 \phi(\sigma,\tau)}{q^2+\fc} - 2q^4\mint{d\hat
679: \tau}{-\infty}{0}\phi(\sigma,\hat \tau+\tau) e^{2 q^4 \hat
680: \tau} \right] \;.
681: \end{split}
682: \end{equation}
683: Now we introduce the Laplace transform of $\phi$ according to
684: \begin{subequations}
685: \label{eq:laplace-trafo}
686: \begin{align}
687: \phi(\sigma,z)& \equiv \mint{d\tau}{0}{\infty}e^{-z \tau}
688: \phi(\sigma,\tau) \\
689: \phi(\sigma,\tau)&=\mint{\frac{dz}{2\pi i}}{c-i \infty}{c+i
690: \infty}e^{z \tau} \phi(\sigma,z)\;,
691: \end{align}
692: \end{subequations}
693: such that Eq.~(\ref{eq:pide-linearized-11}) in Laplace space reads
694: \begin{eqnarray}
695: \label{eq:pide-linearized-lt-1}
696: &&\partial_\sigma^2 \phi(\sigma,z) =
697: \nonumber\\
698: &&\mint{\frac{dq}{\pi}}{-\infty}{\infty}\left[
699: -\frac{\fc}{2q^4}\left(\frac1 z-\frac1{z+2q^4}\right)+
700: \frac{q^2 \phi(\sigma,z)}{q^2+\fc}\right.\nonumber\\
701: &&\left.-\mint{d\tau}{0}{\infty}e^{-
702: z \tau}\mint{d \hat\tau}{-\infty}{0}\phi(\sigma,\hat \tau+
703: \tau)(2
704: q^4)e^{2q^4 \hat \tau}\right] \nonumber\\
705: &&= \mint{\frac{
706: dq}{\pi}}{-\infty}{\infty}\left[-\frac{\fc}{z(z+2q^4)}+
707: \frac{z \phi(\sigma,z)}{2
708: q^4+z}-\frac{\fc \phi(\sigma,z)}{q^2+\fc}\right] \nonumber\\
709: &&=2^{-3/4}\left(z^{1/4}\phi-\fc z^{-7/4}\right)-\fc^{1/2}\phi
710: \;.
711: \end{eqnarray}
712: In the short time limit $z\gg 1$, this reduces to
713: \begin{equation}
714: \label{eq:linear-pide-laplace}
715: \partial_\sigma^2 \phi(\sigma,z)=2^{-3/4}z^{1/4}\phi\;.
716: \end{equation}
717: Choosing the decaying solution we find for the boundary conditions in
718: Eqs.~(\ref{eq:bc-infty},~\ref{eq:bc-origin-1}) of {\pull}
719: \begin{equation}
720: \label{eq:phi-small-t}
721: \phi=\phi_P(\sigma,z)\equiv z^{-2} e^{-\sigma (z/8)^{1/8}} \qquad
722: (\pull) \;.
723: \end{equation}
724: Since Eq.~(\ref{eq:pide-linearized-lt-1}) is a linear differential equation, we can
725: express the solutions for the boundary conditions of {\epull} and {\lelo} in terms of
726: $\phi_P$,
727: \begin{subequations}
728: \label{eq:letloose-from-pulling}
729: \begin{align}
730: \phi&=z^{-2}-\phi_P(\sigma,\tau) \qquad (\lelo) \\
731: \phi&=\mint{d\hat \sigma}{\sigma}{\infty}\phi_P(\hat
732: \sigma,\tau)\qquad (\epull) \;.
733: \end{align}
734: \end{subequations}
735: Ultimately, we are interested in the tension
736: \begin{equation}
737: \label{eq:tension-from-phi}
738: \varphi_P(\sigma,\tau)=\partial_\tau
739: \phi_P(\sigma,\tau) \;,
740: \end{equation}
741: which is given by the inverse Laplace transform of $z
742: \phi_P(\sigma,z)$,
743: \begin{eqnarray}
744: \varphi_P(\sigma,\tau)&=&\hat\varphi_P\left(\frac{\sigma}{\tau^{1/8}}\right) \label{eq:lt-inverse-a}
745: \\
746: \hat\varphi_P(\xi)&=&\mint{\frac{dz}{2\pi i z}}{-i \infty+\epsilon}{i\infty+\epsilon}e^{-\xi
747: (z/8)^{1/8}+z }\;. \label{eq:lt-inverse-b}
748: \end{eqnarray}
749: After deforming the contour of integration such that it encloses the branch cut at the
750: negative real axis, the integral in Eq.~(\ref{eq:lt-inverse-b}) becomes
751: \begin{eqnarray}
752: \label{eq:small-time-result}
753: \hat \varphi_P(\xi)&=&\mint{\frac{dx}{\pi}}{0}{\infty}\frac8x \sin\left[\xi x\sin{\frac \pi 8}
754: \right] e^{-\xi x \cos{\frac \pi 8}}\nonumber\\
755: & &\times\left( 1-e^{-8x^8} \right) \;,
756: \end{eqnarray}
757: which is to our knowledge not tabulated, but can be easily evaluated numerically, see
758: Fig.~\ref{fig:small-t-profile}. Upon using the known Laplace transform,
759: \begin{equation}
760: \label{eq:power-law-back}
761: \frac{\Gamma(\nu)}{z^\nu}=\mint{d\tau}{0}{\infty}e^{-z \tau} \tau^{\nu-1}\;,\quad
762: \Re \nu>0\;,
763: \end{equation}
764: and Taylor expanding the integrand of Eq.~(\ref{eq:lt-inverse-b}) one obtains an
765: expansion of $\varphi_P(\xi)$ that is particular useful for small $\xi$,
766: \begin{equation}
767: \label{eq:small-sigma-exp}
768: \hat \varphi_P(\xi)=\sum_{{n=0\atop n/8\notin\mathbb{N}}}^{\infty}
769: \frac{(-2^{-3/8}\xi)^n}{n!\Gamma(1-n/8)}\;.
770: \end{equation}
771: With an absolute error less than one percent the scaling function is approximated by
772: an exponential,
773: \begin{equation}
774: \label{eq:approx-lr}
775: \hat \varphi_P(\xi)\approx \exp\left(-\frac{\xi}{2^{3/8}\Gamma(7/8)}
776: \right) \;,
777: \end{equation}
778: where the pre-factor of $\xi$ in the exponent is the initial slope
779: $\partial_\xi \hat\varphi\vert_{\xi=0}$ of the scaling function.
780: \begin{figure}
781: \centerline{\includegraphics[width=\columnwidth]
782: {small-time-profile-psfraged.eps}}
783: \caption{The scaling function $\hat \varphi_P(\xi)$ describing the shape
784: of the tension profile of {\pull} on short times.}\label{fig:small-t-profile}
785: \end{figure}
786:
787: The tension profiles of the problems under consideration are all
788: related to the scaling function $\hat \varphi_P(\xi)$,
789: \begin{subequations}
790: \label{eq:letloose-from-pulling-2}
791: \begin{align}
792: f(s,t)&=\fe \hat\varphi_P\left(\frac s {\lpa(t)}\right)\qquad (\pull) \\
793: f(s,t)&=\fe \left[1-\hat\varphi_P\left(\frac s
794: {\lpa(t)}\right)\right]
795: \qquad (\lelo) \\
796: f(s,t)&=\hat \zeta v \lpa(t)\mint{d\xi}{s/\lpa(t)}{\infty}
797: \hat\varphi_P(\xi)\qquad (\epull) \label{eq:efp-t<1} \;,
798: \end{align}
799: \end{subequations}
800: where the boundary layer at time $t$ has the typical size
801: \begin{equation}
802: \label{eq:lambda-short-time}
803: \lpa(t)=\hat \zeta^{-1/2} \lp^{1/2} t^{1/8}\propto\lralf(t) \;.
804: \end{equation}
805: The scaling was anticipated in Eq.~(\ref{eq:short-time-guess}) and in an ad hoc
806: scaling argument presented in Part~I. The result for {\epull},
807: Eq.~(\ref{eq:efp-t<1}), shows that the force at the grafted end, at $\xi=0$, scales
808: like $\hat \zeta v \lpa(t)$. This scaling will be shown to hold also in the long-time
809: limit and can be understood in the sense that the graft has to balance only the drag
810: arising within the boundary layer, since the bulk of the filament is not moving
811: longitudinally. Thus, measuring the force at the grafted end, we can monitor the
812: spreading of the tension. This gives special experimental relevance to the {\epull}
813: scenario. In particular, we predict
814: \begin{equation}
815: \label{eq:force-at-graft}
816: f(0,t)=\hat \zeta v \lpa(t)2^{3/8}/\Gamma(9/8)
817: \end{equation}
818: for the force at the grafted end on short times, $\tau\ll1$. The
819: pre-factor in Eq.~(\ref{eq:force-at-graft}) has been found by
820: evaluating the remaining integral in Eq.~(\ref{eq:efp-t<1}), which
821: yields a Taylor series very similar to Eq.~(\ref{eq:small-sigma-exp}),
822: \begin{equation}
823: \label{eq:efp-t<1-full}
824: \mint{d\hat\xi}{\xi}{\infty}
825: \hat\varphi_P(\hat \xi)=2^{3/8}\sum_{{n=-1\atop
826: n/8-1\notin\mathbb{N}}}^{\infty}
827: \frac{(-2^{-3/8}\xi)^{n+1}}{(n+1)!\Gamma\left(1-n/8\right)}\;.
828: \end{equation}
829:
830:
831: \subsubsection{Long times ($t \gg \tf$)}
832: \label{sec:long-time-asymptotics}
833: The present subsection deals with the dynamics of the tension on a semi-infinite
834: filament on asymptotically long times. We identify and interprete the terms
835: dominating the stored length release in this limit. Neglecting subdominant terms in
836: the continuity equation, Eq.~(\ref{eq:cg-eom}), results in differential equations
837: that can be solved by similarity solutions.
838:
839: The right hand side of the nondimensionalized PIDE,
840: Eq.~(\ref{eq:pide-nondimensional-1}), represents the negative change of stored length
841: in adapted units. It can be written as the sum of two terms, $A$ and $B$, where
842: \begin{subequations}
843: \begin{align}
844: \label{eq:A-def}
845: A&\equiv\mint{\frac{dq}{2\pi}}{-\infty}{+\infty}\frac 1 {q^2+\fc
846: }\left[1-e^{-2q^2\left[q^2 \tau+
847: \phi(\sigma,\tau)\right]}\right]\;,\\
848: \label{eq:B}
849: B&\equiv-\mint{\frac{dq}{2\pi}}{-\infty}{\infty}
850: 2 q^2 \mint{d\hat \tau}{0}{\tau}e^{-2 q^2 \left[ q^2
851: (\tau-\hat \tau)+\phi(\sigma,\tau)- \phi(\sigma, \hat
852: \tau)\right] }\;.
853: \end{align}
854: \end{subequations}
855: We already pointed out in Part~I that the term $A$ can be interpreted as the
856: ``deterministic relaxation'' of stored length (for the fictitious situation ``T=0'',
857: i.e.~no thermal noise, for $t>0$). The term $B$ describes the increase in stored
858: length due to the thermal kicks and is strictly positive. We analyze both terms
859: separately.
860:
861: For pre-stretched initial conditions ($\fc=1$) the long-time limit
862: of $A$ follows from setting the exponential to zero,
863: \begin{equation}
864: \label{eq:A-let-loose}
865: A\stackrel{\tau\gg1}{\to}\mint{\frac{dq}{2\pi}}{-\infty}{+\infty}
866: \frac1{q^2+1}=\frac 1 2\;, \quad \text{for }\fc=1 \;.
867: \end{equation}
868: The term $A$ for $\tau\to \infty$ is nothing but the initially stored length, which
869: is completely relaxed after a purely deterministic relaxation.
870:
871: The same reasoning cannot be applied in cases of a tension-free
872: initial state (both pulling problems): setting the exponential to zero
873: in Eq.~(\ref{eq:A-def}) for $\fc=0$ yields an infrared divergence. The
874: exponential has to be retained to render the integrand finite at small
875: wave numbers. For the dominant small wave numbers one can, however,
876: neglect the term $2 q^4 \tau$ in the exponent because it is small
877: compared to the term $2q^2\Phi$, so that we arrive at the asymptotic
878: expression
879: \begin{equation}
880: \label{eq:a-pulling}
881: \begin{split}
882: A\stackrel{\tau\gg1}{\to}&\mint{\frac{dq}{2\pi}}{-\infty}{+\infty}
883: \frac1{q^2}
884: \left[1-e^{-2q^2\phi(\sigma,\tau)}\right]\\
885: &=\sqrt{\frac2\pi} \sqrt{\phi(\sigma,\tau)}\;,\quad \text{for } \fc=0 \;.
886: \end{split}
887: \end{equation}
888: I.e. the ``deterministic relaxation'' is dominated by the tension term
889: and bending can be neglected (as heuristically assumed by Seifert et
890: al.~\cite{seifert-wintz-nelson:96}). A more formal justification for
891: Eqs.~(\ref{eq:A-let-loose},~\ref{eq:a-pulling}) is given in
892: App.~\ref{sec:A}.
893:
894:
895: The term $B$, describing the stored length generated by the thermal
896: kicks, takes for asymptotically large $\tau\gg1$ the form
897: \begin{equation}
898: \label{eq:B-large-times}
899: B\stackrel{\tau\gg1}{\to}-\mint{\frac{dq}{2\pi}}{-\infty}{\infty}
900: \frac 1 {q^2+\partial_\tau \phi(\sigma,\tau)}
901: =-\frac1{2\sqrt{\partial_\tau\phi(\sigma,\tau)}}
902: \end{equation}
903: independent of the initial conditions. This is shown in App.~\ref{sec:B} upon using
904: the scaling assumptions in Eqs.~(\ref{eq:scaling-phi},~\ref{eq:alpha-assumption}).
905: The result, Eq.~(\ref{eq:B-large-times}), should not come as a surprise. It simply
906: represents the (negative) stored length of a stiff polymer equilibrated at the
907: (rescaled) tension $\partial_\tau \phi$. For a constant force, i.e., $\partial_\tau
908: \phi=$const., it is obvious that the stored length should saturate for long times at
909: the corresponding equilibrium value. But also if the tension is varying slowly
910: enough in time ($\alpha>-1/2$) the ``noise-generated'' stored length can be
911: considered as quasi-statically equilibrated with the tension.
912:
913: Finally, we combine the ``relaxed stored length'' expressed in $A$ and
914: the ``noise-generated stored length'' $B$.
915:
916:
917:
918: \paragraph{Pulling and Towing:} For $\fc=0$ and $\alpha>-1/2$ the term $A\propto
919: \sqrt{\phi}\propto\tau^{\alpha/2+1/2}$ is much larger than $B\propto
920: (\partial_\tau\phi)^{-1/2}=\Ord{\tau^{-\alpha/2}}$. I.e. the effects of noise can be
921: neglected on long times (as presumed by Seifert et
922: al.~\cite{seifert-wintz-nelson:96}). In this limit, the thermal noise is merely
923: relevant in preparing the initial state. The relaxation after force application for
924: $\tau\gg 1$ is purely mechanical, like for a pulled string that is initially prepared
925: with some contour roughness (\emph{taut-string approximation}). If we replace the
926: right-hand-side of Eq.~(\ref{eq:pide-nondimensional-1}) by the asymptotic form of
927: $A$, Eq.~(\ref{eq:a-pulling}), we obtain the partial differential equation
928: \begin{equation}
929: \label{eq:pulling-large-time-pde}
930: \partial_\sigma^2\phi(\sigma,\tau)=\sqrt{\frac2\pi}
931: \sqrt{\phi(\sigma,\tau)}
932: \end{equation}
933: for the dynamics of the integrated tension $\phi$.
934:
935: Eq.~(\ref{eq:pulling-large-time-pde}) represents a Newtonian equation of motion for a
936: particle moving in a conservative force field $\propto\sqrt{\phi}$ and can be
937: integrated straightforwardly. In the case of {\pull} the solution, which satisfies
938: the boundary conditions in Eqs.~(\ref{eq:bc-infty},~\ref{eq:bc-origin-1}) of {\pull} (in
939: particular $\phi(0,\tau)=\tau$), is given by
940: \begin{equation}
941: \label{eq:pulling-large-time-phi}
942: \phi(\sigma,\tau)=\tau\left[\sigma/(72\pi\tau)^{1/4}
943: -1\right]^4
944: \end{equation}
945: for $\sigma<(72\pi \tau)^{1/4}$ and $\phi=0$ otherwise. The dimensionless tension
946: $\varphi$ is derived from $\phi$ via differentiation, see
947: Eq.~(\ref{eq:close-scaling-form-1}), and obeys a scaling form,
948: \begin{equation}
949: \label{eq:pulling-large-time-varphi}
950: f(s,t)=\fe \varphi(s,t)=\fe \hat\varphi\left(\frac s {\lpa(t)}\right)\;,
951: \end{equation}
952: with a typical boundary layer size proportional to $\ludo(t)$,
953: \begin{equation}
954: \label{eq:lambda-large-time}
955: \lpa(t)=t^{1/4}\hat
956: \zeta^{-1/2}\lp^{1/2}\fe^{1/4}
957: \end{equation}
958: and a scaling function $\hat \varphi$ given by
959: \begin{equation}
960: \label{eq:scalingfct-large-times}
961: \hat\varphi(\xi)=\left[1-\xi/(72\pi)^{1/4}\right]^3
962: \end{equation}
963: for $ \xi<(72 \pi)^{1/4}$ and $\hat \varphi=0$ otherwise. {\epull} starts with the same initial conditions ($\fc=0$) but with
964: different boundary conditions, Eqs.~(\ref{eq:bc-infty},~\ref{eq:bc-origin-2}).
965: Again, we have to solve Eq.~(\ref{eq:pulling-large-time-pde}), but now
966: under the boundary condition $\partial_\sigma \phi(0,\tau)=-\tau$. The
967: solution is
968: \begin{equation}
969: \label{eq:phi-electroforesis}
970: \phi(\sigma,\tau)=\tau^{4/3}\left( \frac{9\pi}{32} \right)^{1/3}\left[
971: \frac{\sigma}{(18\pi\tau)^{1/3}}-1 \right]^4
972: \end{equation}
973: for $\sigma<(18\pi\tau)^{1/3}$ and $\phi=0$ otherwise. This implies a tension profile
974: of
975: \begin{equation}
976: \label{eq:tension-orig-electroforesis}
977: f(s,t)=\hat \zeta v \lpa(t)\; \hat\varphi\left(\frac s
978: {\lpa(t)}\right)\;,
979: \end{equation}
980: with the typical boundary layer size
981: \begin{equation}
982: \label{eq:lambda-ef-large-time}
983: \lpa(t)= t^{1/3}\lp^{2/3}(v/\hat \zeta)^{1/3} \;,
984: \end{equation}
985: and the scaling function $\hat \varphi$ given by
986: \begin{equation}
987: \label{eq:scalingfct-towing-large-times}
988: \hat\varphi(\xi)=\left(\frac{2\pi}{3}
989: \right)^{1/3}\left[1-\xi/(18\pi)^{1/3}\right]^3
990: \end{equation}
991: for $ \xi<(18 \pi)^{1/3}$ and $\hat\varphi=0$ otherwise. As on short times, the
992: absolute value of the reduced tension at the left end is proportional to the size of
993: the boundary layer. Its precise value is predicted to be
994: \begin{equation}
995: \label{eq:left-end-tension-electroforese}
996: f(0,t)=\left(2\pi/3 \right)^{1/3}\hat \zeta v \lpa(t)\propto t^{1/3} \;,
997: \end{equation}
998: and should be directly accessible to single molecule experiments.
999:
1000:
1001: \paragraph{Release:}
1002: The stored length release is asymptotically given by
1003: \begin{equation}
1004: \label{eq:letting-loose-rho-release-large-t}
1005: -\avg{\Delta\overline{\varrho}}\propto A+B\stackrel{\tau\gg1}{\sim} \frac 1 2-
1006: \frac1{2\sqrt{\partial_\tau \phi(\sigma,\tau)}}\;.
1007: \end{equation}
1008: This expression for the change in stored length can also be directly obtained if one
1009: assumes that the filament was at any time equilibrated with the current tension
1010: $\varphi=\partial_\tau \phi$ (\emph{quasi-static approximation}). Our derivations
1011: show, that this assumption, used by Brochard et
1012: al.~\cite{brochard-buguin-de_gennes:99}, is only valid in the long time limit for the
1013: scenario of {\lelo}. The dynamics of the integrated tension is then described by
1014: \begin{equation}
1015: \label{eq:pde-letting-loose-long-t}
1016: \partial_\sigma^2\phi=\frac 1 2-\frac1{2\sqrt{\partial_\tau
1017: \phi(\sigma,\tau)}}\;,
1018: \end{equation}
1019: or, in terms of the tension $\varphi=\partial_\tau \phi(\sigma,\tau)$,
1020: \begin{equation}
1021: \label{eq:pde-letting-loose-long-t-2}
1022: \partial_\sigma^2\varphi=\frac1 4 \varphi^{-3/2}\partial_\tau\varphi\;.
1023: \end{equation}
1024: The solution satisfying the correct boundary conditions,
1025: Eqs.~(\ref{eq:bc-infty},~\ref{eq:bc-origin-1}), is given by a scaling form
1026: \begin{equation}
1027: \label{eq:scform-letting-loose-long-time}
1028: \varphi(s,t)=\hat \varphi\left(\frac s {\lpa(t)}\right) \;,
1029: \end{equation}
1030: with the typical boundary layer size now growing like
1031: \begin{equation}
1032: \label{eq:lambda-letloose-large-t}
1033: \lpa(t)=t^{1/2}\hat\zeta^{-1/2}\lp^{1/2}\fe^{3/4} \propto \lpierre(t)
1034: \end{equation}
1035: and a scaling function $\hat\varphi(\xi)$ satisfying the ordinary differential
1036: equation
1037: \begin{equation}
1038: \label{eq:scfct-ode}
1039: \partial_\xi^2\hat\varphi=-\frac1 8 \xi
1040: {\hat\varphi}^{-3/2}\partial_\xi\hat\varphi \;.
1041: \end{equation}
1042: The scaling function depicted in Fig.~\ref{fig:degennes-scfct} was already obtained
1043: numerically in Ref.~\cite{brochard-buguin-de_gennes:99}. The slope at the origin is
1044: $\partial_\xi \hat \varphi\vert_{\xi=0}\approx0.6193$.
1045:
1046: \begin{figure}
1047: \includegraphics[width=\columnwidth]{degennes-corrected-psfraged.eps}
1048: \caption{The scaling function $\hat \varphi(\xi)$ for
1049: {\lelo} obtained by a numerical solution of
1050: Eq.~(\ref{eq:scfct-ode}).}\label{fig:degennes-scfct}
1051: \end{figure}
1052:
1053:
1054:
1055: \subsection{Tension propagation (summary)}
1056: \label{sec:tensprop-summary}
1057:
1058: This section summarizes the picture of tension propagation that emerges from the
1059: above solutions of the dynamical equation for the tension, Eq.~(\ref{eq:pide}), for
1060: sudden changes in boundary conditions.
1061:
1062: For the considered problems {\pull}, {\epull} and {\lelo}, we have shown that the
1063: tension profile in units of $\fe$ has to obey a crossover scaling form
1064: $\varphi(\sigma,\tau)$ depending on a reduced time variable $\tau\equiv t/t_\fe$ and
1065: a reduced arc length variable $\sigma\equiv s/s_\fe$. The scaling function $\varphi$
1066: describes how sudden changes of the tension at the ends spread into the bulk of the
1067: polymer. Its crossover structure and the expressions for $t_\fe$ and
1068: $s_\fe\approx\lpa(t_\fe)$ are consistent with our heuristic discussion of {\pull} in
1069: Part~I. In the limits $\tau \ll1$ and $\tau\gg1$ the function $\varphi(\sigma,\tau)$
1070: assumes a simple (one-variable) scaling form
1071: \begin{equation}
1072: \label{eq:tension-scform}
1073: \varphi\sim \tau^{\alpha_{\gtrless}}
1074: \hat\varphi^{\gtrless} [\sigma/\tau^{z_{\gtrless}}] \;.
1075: \end{equation}
1076: The notation $\gtrless$
1077: indicates that the asymptotic form of $\hat \varphi$, $\alpha$, and $z$ will
1078: generally not only depend on the kind of external perturbation applied, but will also
1079: differ for times $t\gtrless t_\fe$. Rewriting $\sigma/\tau^z\equiv s/\lpa$
1080: identifies the tension propagation length $\lpa\equiv s_\fe\,\tau^z$.
1081:
1082: For $t\ll t_\fe$ Eq.~(\ref{eq:pide}) could be linearized in $f$ and the scaling
1083: function $\hat \varphi^<$ was obtained analytically. Whereas $\hat \varphi^<$ depends
1084: on the considered force protocol, the corresponding exponent $z_<=1/8$ is independent
1085: of the boundary conditions. As established by our heuristic discussion of {\pull},
1086: this is due to the relaxation of modes with Euler forces $\lpe^{-2}\gg \fe$ much
1087: larger than the external force, for which the equilibrium mode spectrum is hardly
1088: perturbed by the external force. The self-affinity of the equilibrium mode spectrum
1089: translates into a self-similar relaxation dynamics. The dynamic exponent $z$ for the
1090: growth of the boundary layer could already be anticipated from requiring
1091: $\hat\varphi^<$ to become $\fe-$independent as in linear response, see
1092: Eq.~(\ref{eq:short-time-guess}). The short-time dynamics for strong external force is
1093: thus closely related to the linear response. Note, however, that the limit $\fe\to0$
1094: is problematic, as it does not interchange with $\epsilon\to0$. Our identification
1095: of arc length averages with (local) ensemble averages in Part~I breaks down for
1096: $\fe<(\zeta/\lp)^{1/4}t^{-7/16}$, where fluctuations in the tension become comparable
1097: to its average value. In fact, extending Eq.~(\ref{eq:pide}) to linear response
1098: amounts to an uncontrolled factorization approximation $\avg{f r_\pe^2}\to
1099: \avg{f}\avg{r_\pe^2}$. Even in the stiff limit the linear longitudinal dynamic
1100: response remains an open problem. The limit where fluctuations in the tension become
1101: important and its consequences will be detailed in Sec.~\ref{sec:tens-fluct}.
1102:
1103: For $t\gg t_\fe$ the dynamics becomes nonlinear in the external force and starts to
1104: depend on the kind of external perturbation and on how precisely it is applied to the
1105: polymer. Previously predicted power laws were recovered from Eq.~(\ref{eq:pide}) by
1106: employing different approximations to its right hand side. In the \emph{taut-string
1107: approximation} of Ref.~\cite{seifert-wintz-nelson:96}, one neglects for $t>0$
1108: bending and thermal forces against the tension, i.e., one drops the $q^4-$term in the
1109: relaxation time $\tau_q=q^4+f q^2$ of a mode with wave number $q$ and sets
1110: $\lp\to\infty$ for positive times (i.e.~$\theta=0$). The complementary
1111: \emph{quasi-static approximation} of Ref.~\cite{brochard-buguin-de_gennes:99} amounts
1112: to the omission of memory effects, i.e., to the assumption of instantaneous
1113: equilibration of tension and stored length (as it would be the case for vanishing
1114: \emph{transverse} friction, $\zeta_\pe\to 0$). In cases, where either of these
1115: approximations applies, a power-law dispersion relation combines with a self-affine
1116: mode spectrum to produce self-similar tension dynamics. Our analysis of
1117: Eq.~(\ref{eq:pide}) showed that either of these approximations becomes rigorous in
1118: the intermediate asymptotic regime defined by $t\gg t_\fe$, $\lpa\ll L$. The
1119: quasi-equilibrium approximation applies to {\lelo} and the taut-string approximation
1120: to {\pull}. We could rule out the applicability of the \emph{taut-string
1121: approximation} for {\lelo} and of the \emph{quasi-static approximation} for
1122: {\pull}~\cite{brochard-buguin-de_gennes:99} and {\epull}. The ``pure'' scenarios of
1123: self-similar dynamics are summarized in Tab.~\ref{tab:lf-growth-laws} and
1124: Fig.~\ref{fig:lf}.
1125:
1126:
1127:
1128: \begin{table}
1129: \caption{Asymptotic growth laws for the dynamic
1130: size $\lpa(t)$ of the tension boundary
1131: layer.}
1132: \label{tab:lf-growth-laws}
1133: \begin{center}
1134: \begin{tabular}{c|c|c}
1135: Problem&$t\ll \tf$&$ \tf\ll t \ll t_L^\pa$\\ \hline
1136: {\pull}&${\displaystyle \sqrt{ \lp /\hat \zeta }\;
1137: t^{1/8}}$&${\displaystyle
1138: \sqrt{ \lp /\hat \zeta }\; \fe^{1/4}t^{1/4}}$\\
1139: {\epull}&${\displaystyle \sqrt{\lp/ \hat \zeta }\;
1140: t^{1/8}}$&${\displaystyle
1141: (\lp^2 v/\hat \zeta)^{1/3}t^{1/3}}$\\
1142: {\lelo}&${\displaystyle \sqrt{ \lp/ \hat \zeta }\;t^{1/8}}$
1143: &${\displaystyle
1144: \sqrt{ \lp / \hat \zeta }\;\fe^{3/4}t^{1/2}}$\\
1145: \end{tabular}
1146: \end{center}
1147: \end{table}
1148:
1149: \begin{figure}[t]
1150: \includegraphics[width=\columnwidth,height=.5\columnwidth
1151: ]{growth-laws-pre-psfraged.eps}
1152: \caption{Schematic of the tension propagation laws $\lpa(t)\propto t^z$ on a
1153: double-logarithmic scale. At $t_\fe=\fe^{-2}$ they crossover from a universal
1154: short-time regime to (problem-specific) tension-dominated intermediate
1155: asymptotics, except for weak forces, $\fe < \lp^2/L^4$. The propagation ends when
1156: $\lpa(t)\approx L$.}
1157: \label{fig:lf}
1158: \end{figure}
1159:
1160:
1161:
1162: \section{Terminal stress relaxation}
1163: \label{term-rel-tensprop}
1164: Up to now, we have considered the growth of the boundary layer in a
1165: stiff polymer that has a (formally) semi-infinite arc length
1166: parameter space, $s\in [0,\infty [$, which is an idealization. However,
1167: the foregoing discussion obviously applies to a polymer of
1168: \emph{finite} length $L$ for sufficiently short times: As long as the
1169: size of the boundary layer is much smaller than the total length $L$
1170: the presence of a second end is irrelevant to the boundary layer
1171: at the first end. The time where the boundary layers span the whole
1172: polymer marks the crossover to a new behavior. For definiteness, we
1173: define the crossover time $\totpa$ by
1174: \begin{equation}
1175: \label{tension-rel-time-tp}
1176: \lpa(\totpa)\equiv L \;.
1177: \end{equation}
1178: What happens for $t>\totpa$? The straightforward way to answer this question is to
1179: solve for the intermediate asymptotics of the PIDE, Eq.~(\ref{eq:pide}), for a polymer
1180: of finite length. The finiteness of $L$ amounts to replacing the boundary condition
1181: $\partial_s^2 f(s\to\infty)=0$ by the correct problem-specific one, i.e., by
1182: \begin{eqnarray}
1183: \label{eq:bc-tension-rend}
1184: f(L,t>0)&=&\fe\qquad {\pull} \nonumber\\
1185: f(L,t>0)&=&0 \qquad {\lelo}, {\epull}.\nonumber
1186: \end{eqnarray}
1187: One could now proceed as in Sec.~\ref{sec:asymptotic-growth} by identifying proper
1188: scaling forms and extracting their asymptotic behavior. As compared to the
1189: semi-infinite polymer limit, this procedure is more complicated for a finite polymer
1190: because of the additional scaling variable $\lpe(t)/L$. Therefore, we prefer to take
1191: the following short-cut, which consists in two steps.
1192:
1193: \subsubsection{Trivial tension profiles for $t\gg t_\star$.} In the heuristic
1194: analysis of Part~I, we found that ordinary perturbation theory (OPT) should become
1195: valid for times larger than some crossover time $t_\star$. With the exception of
1196: {\epull}, which we discuss separately below, the introduced scenarios treat both ends
1197: equally, such that the tension profile in OPT has a trivial time and arc length
1198: dependence, namely
1199: \begin{equation}
1200: \label{eq:OPT-tension}
1201: \fopt(s,t)=f(0,t)=\text{const.}\;,
1202: \end{equation}
1203: up to small terms of order $\Ord{\epsilon}$. For constant tension we
1204: can easily extract the longitudinal dynamics from
1205: Eq.~(\ref{eq:change-stored-length}). The corresponding
1206: predictions for the evolution of the end-to-end distance will be
1207: discussed in Sec.~\ref{sec:tension-relaxed}.
1208:
1209:
1210: \subsubsection{Possibility of homogeneous tension relaxation for $\totpa \ll t \ll
1211: t_\star$.} Up to now, we have argued that the tension propagates for $t\ll \totpa$
1212: and is constant in time and space for $t\gg t_\star$. There remains the question
1213: whether there is a non-trivial regime of homogeneous tension relaxation in the
1214: time- interval $[\totpa;t_\star]$. Using the systematic approach outlined in
1215: App.~\ref{sec:tstar} to determine $t_\star$, we actually find for most of the
1216: problems that $t_\star\simeq \totpa$, i.e., there is no scaling regime between
1217: tension propagation and the stationary tension profiles dictated by OPT. The
1218: {\lelo}-scenario, however, provides an important exception, as it allows for a time
1219: scale separation $\totpa\ll t_{\star}$, as we demonstrate explicitely below. For
1220: intermediate times the tension relaxation is shown to exhibit a novel behavior with
1221: an almost parabolic tension profile and an amplitude that slowly decays in time
1222: according to a power law.
1223:
1224: \subsection{{\lelo} for large pre-stretching force}
1225: \label{sec:homogeneous-relaxation}
1226: Let us first determine the time $t_\star$, at which OPT becomes valid, for {\lelo}.
1227: In the OPT regime, the tension should be so small, that we can calculate the change in
1228: stored length accurately by means of Eq.~(\ref{eq:change-stored-length}) (with the
1229: pre-stretched initial conditions of {\lelo}, i.e.~$f_<=\fe$) under the assumption of
1230: a vanishing tension, $\fopt=0$,
1231: \begin{subequations}
1232: \label{eq:stole-lelo-OPT}
1233: \begin{eqnarray}
1234: &&\avg{\Delta \overline{\varrho}}(t)=\mint{\frac{dq}{\pi\lp}}{0}{\infty}
1235: \left\{\frac{1}{q^2+\fe}\left(e^{-2q^4 t}-1\right)\right.\nonumber\\
1236: &&\qquad\left.+2
1237: q^2\mint{d\tilde t}{0}{t}e^{-2q^4 (t-\tilde
1238: t)}\right\} \label{eq:stole-lelo-OPT-1}\\
1239: &&=\mint{\frac{dq}{\pi\lp}}{0}{\infty} \left\{\left(
1240: \frac{1}{q^2}-\frac{1}{q^2+\fe} \right)\left(1-e^{-2q^4
1241: t}\right) \right\} \label{eq:stole-lelo-OPT-2}\\
1242: &&=\frac{t^{1/4}}{\lp}\mint{\frac{dq}{\pi}}{0}{\infty}\\
1243: &&\qquad \times\left\{\left( \frac{1}{q^2}-\frac{1}{q^2+\sqrt{t/\tf}}
1244: \right)\left(1-e^{-2q^4} \right)
1245: \right\}\qquad\label{eq:stole-lelo-OPT-3} \\
1246: &&\stackrel{t\gg\tf}{\sim}\frac{2^{3/4}} {\Gamma(1/4)}
1247: \frac{t^{1/4}}{\lp}\;.\label{eq:stole-lelo-OPT-4}
1248: \end{eqnarray}
1249: \end{subequations}
1250: Here, we have substituted $q\to q t^{-1/4}$ and replaced $\fe$ by $\tf^{-1/2}$ in
1251: order to obtain Eq.~(\ref{eq:stole-lelo-OPT-3}). In the final step, we took the
1252: long-time limit $t\gg\tf$. Let us now determine the order of magnitude of the
1253: variation $\delta f$ of the tension due to the longitudinal friction. Estimating
1254: $\delta f\simeq \hat \zeta \avg{\Delta\overline\varrho}/(t L^2)$ from Eq.~(\ref{eq:cg-eom}), we
1255: obtain
1256: \begin{equation}
1257: \label{eq:delta-f-lelo}
1258: \delta f\approx \hat \zeta \frac{L^2}{\lp}t^{-3/4}\;.
1259: \end{equation}
1260: As in the heuristic discussion of Part~I, we find a diverging tension in the limit
1261: $t\to 0$ for the OPT result, so that the OPT result can only be valid after the
1262: effect of $\delta f$ on the evolution of $\avg{\Delta\overline{\varrho}}(t)$ can be
1263: neglected. For the particular case of {\lelo}, this time can be determined as
1264: follows. As discussed in Part~I, the stress-free dynamics is at the time $t$
1265: characterized by relaxation of modes with wave length $\lpe(t)\simeq t^{1/4}$. When
1266: $\delta f$ is larger than the critical Euler Buckling force $\lpe(t)^{-2}$
1267: corresponding to the length $\lpe(t)$ we expect that the tension cannot be neglected
1268: for the evolution of $\avg{\Delta\overline{\varrho}}$. Hence, the above result
1269: $\avg{\Delta\overline{\varrho}}\propto L t^{1/4}/\lp$, obtained from OPT, can only be
1270: valid if
1271: \begin{equation}
1272: \label{eq:val-cond-lelo}
1273: \lpe(t)^2\delta f\simeq \frac{L^2}{\lp}t^{-1/4}\ll1 \;,
1274: \end{equation}
1275: i.e., for long enough times
1276: \begin{equation}
1277: \label{eq:val-cond-lelo-2}
1278: t\gg t_\star=L^8/\lp^4 \;.
1279: \end{equation}
1280: The time $t_\star(L)$ is obviously not identical with the time
1281: \begin{equation}
1282: \label{eq:tf-2}
1283: \totpa=L^2 \lp^{-1} \hat\zeta \fe^{-3/2} \qquad (t\gg\tf)
1284: \end{equation}
1285: it takes for the boundary layer to spread over the filament. To compare them, we
1286: first notice that $\totpa\gg\tf=\fe^{-2}$ implies that the polymer must have been
1287: pre-stretched by a large enough force,
1288: \begin{equation}
1289: \label{eq:large-force}
1290: \fe\gg \lp^2/L^4=\epsilon^{-2} L^{-2} \;,
1291: \end{equation}
1292: larger than the pre-stretching force at least necessary to enter
1293: the propagation regime $\tf\ll t\ll t_L^\pa$. By using the estimate
1294: in Eq.~(\ref{eq:large-force}) we can compare $t_\star$ and $\totpa$,
1295: \begin{equation}
1296: \label{eq:tstar-vs-tf}
1297: t_\star(L)= \frac{L^2}{\lp}\left(\frac{L^2}{\lp}\right)^3
1298: \gg\frac{L^2}{\lp} \fe^{-3/2} \simeq \totpa \;.
1299: \end{equation}
1300: It is seen that the time window $\totpa\ll t\ll t_\star$ grows with the
1301: pre-stretching force, which means, in particular, that it describes the limit of an
1302: initially straight polymer.
1303:
1304: To determine the physics of this novel regime, we have to solve the equation of
1305: motion for the tension on the finite arc length interval $[0;L]$ using the
1306: approximations developed in Sec.~\ref{sec:asymptotic-growth}. There, we found that
1307: in the limit $t\gg\tf$ the tension profile of {\lelo} is described by
1308: Eq.~(\ref{eq:pde-letting-loose-long-t-2})
1309: \begin{equation}
1310: \label{eq:degennes-eq}
1311: \partial_\sigma^2\varphi=\frac1 4 \varphi^{-3/2}\partial_\tau\varphi\;.
1312: \end{equation}
1313: The right-hand-side represents the time-derivative of the stored
1314: length in the \emph{quasi-static approximation}. Going back to
1315: variables $f$, $s$ and $t$, Eq.~(\ref{eq:degennes-eq}) takes the form
1316: \begin{equation}
1317: \label{eq:degennes-eq-2}
1318: \partial_s^2 f=\frac1 4 \frac{\hat \zeta}{\lp}f^{-3/2}\partial_t f \;.
1319: \end{equation}
1320: For the following, we assume that this \emph{quasi-static
1321: approximation} is not only valid in the tension propagation regime for
1322: $\tf\ll t \ll\totpa$, but also for longer times until the OPT
1323: regime begins, $\tf\ll t \ll t_\star$. This is justified a posteriori
1324: in App.~\ref{sec:lelo-app}, where it is shown that the change in
1325: stored length for the solution $f(s,t)$ of
1326: Eq.~(\ref{eq:degennes-eq-2}) can indeed be calculated quasi-statically
1327: for times $t \ll t_\star$. We solve Eq.~(\ref{eq:degennes-eq-2}) for
1328: the boundary conditions
1329: \begin{equation}
1330: \label{eq:bc}
1331: f(s=0,t)=0 \;,\;f(s=L,t)=0
1332: \end{equation}
1333: and the initial conditions
1334: \begin{equation}
1335: \label{eq:ini-cond}
1336: f(s,0)=\fe \;,\qquad \text{ for }0<s<L \;.
1337: \end{equation}
1338: In Sec.~\ref{sec:long-time-asymptotics}, we solved the differential equation
1339: Eq.~(\ref{eq:degennes-eq}) with the scaling ansatz $f=\fe \hat \varphi[s/\lpa(t)]$
1340: numerically for the correct initial condition, but ignored the boundary conditions of
1341: at one end by sending $L\to\infty$. In contrast, we now look for a simple solution
1342: that obeys the tension boundary conditions exactly, at the expense of a possible
1343: mismatch with the initial conditions. To this end, we make the product ansatz
1344: \begin{equation}
1345: \label{product-ansatz-2}
1346: f(s,t)=g(t) h(s) \;.
1347: \end{equation}
1348: Separation of variables yields
1349: \begin{equation}
1350: \label{eq:sep-var}
1351: \frac{\hat \zeta}{4\lp}g^{-5/2}\partial_t g=C=\sqrt{h}h'' \;.
1352: \end{equation}
1353: Choosing $C=-1/(6L^2)$, for convenience, we find the long time asymptotics
1354: \begin{equation}
1355: \label{eq:g-eq}
1356: g(t)\sim\left(\frac{\hat \zeta L^2}{\lp t}\right)^{2/3}\;.
1357: \end{equation}
1358: The spatial part obeys a scaling form $h(s)=\hat h(s/L)$, where $\hat
1359: h$ satisfies the equation
1360: \begin{equation}
1361: \label{eq:h-eq}
1362: \partial_\xi^2 \hat h(\xi)=-\frac 1 6 \hat h^{-1/2}
1363: \end{equation}
1364: with the boundary condition
1365: \begin{equation}
1366: \label{eq:bc-hat-h}
1367: \hat h(0)=\hat h(1)=0 \;.
1368: \end{equation}
1369: The analytical solution can be found in a standard way. The main characteristics are
1370: the slope at $\xi=0$,
1371: \begin{equation}
1372: \label{eq:h-ini-slope}
1373: \partial_\xi h(\xi)\vert_{\xi=0}=12^{-1/3}\approx 0.4368 \;.
1374: \end{equation}
1375: and the maximum value of $h$,
1376: \begin{equation}
1377: \label{eq:hmax}
1378: h(1/2)=\frac 1 {16} \left( \frac 3 2 \right)^{2/3}\approx
1379: 0.0819 \;.
1380: \end{equation}
1381:
1382:
1383: \begin{figure}
1384: \includegraphics[width=\columnwidth]{degennes3d-psfraged.eps}
1385: \label{fig:lelo-3d}
1386: \caption{The initial ($t\ll t_\star$) time evolution of {\lelo}. A
1387: regime of (slow) tension relaxation begins after the sudden change
1388: in boundary condition has propagated through the filament.}
1389: \end{figure}
1390:
1391:
1392: Up to now we have investigated the quasi-static approximation in the tension
1393: propagation regime ($t\ll \tfin$) and in the regime of tension relaxation ($t\gg
1394: \tfin$) separately. In order to illustrate the crossover, we have also solved the
1395: corresponding PIDE, Eq.~(\ref{eq:degennes-eq-2}), numerically. The result shown in
1396: Fig.~\ref{fig:lelo-3d} unveils the transient nature of the tension propagation
1397: regime.
1398:
1399:
1400:
1401: \section{Nonlinear response of the projected length}
1402: \label{sec:MSDII}
1403: After having discussed the rich tension dynamics of stiff polymers in detail, we wish
1404: to derive its consequences for pertinent observables in order to make contact with
1405: experiments. Although the tension may in some situations be monitored directly, see
1406: Sec.~\ref{sec:possible-exps}, a more conventional observable is the longitudinal
1407: extension $R_\pa(t)$ of the polymer, which is defined to be the end-to-end distance
1408: projected onto the longitudinal axis. Tension dynamics strongly affects the nonlinear
1409: response of the projected length, which shall be detailed in the following for the
1410: force protocols {\pull}, {\epull} and {\lelo}.
1411:
1412: The average temporal change in the projected length $R_\pa(t)$ is directly related to
1413: the stored length release,
1414: \begin{equation}
1415: \label{eq:end-end-length}
1416: \avg{\Delta R_\pa} (t)= -\mint{ds}{0}{L}\avg{\Delta\varrho}(s,t)+o(\epsilon)\;,
1417: \end{equation}
1418: which was already noted in Part~I. As long as modes with wave length on the order of
1419: the total length are irrelevant (i.e.~both ends of the polymer are not correlated),
1420: \begin{equation}
1421: \label{eq:ends-uncorrelated}
1422: \lpe(t)\ll L\;,
1423: \end{equation}
1424: half-space solutions for $\avg{\Delta\varrho}$ may be used to evaluate
1425: Eq.~(\ref{eq:end-end-length}).
1426:
1427: Recall from Part~I, that $\avg{\Delta\varrho}$ can be decomposed into a bulk
1428: contribution $\avg{\Delta \overline{\varrho}}$ and a term influenced by the boundary
1429: conditions, which vanishes under a spatial average. We show in
1430: Sec.~\ref{sec:microscale} that in the cases of hinged and clamped ends, as opposed to
1431: free ends, the contribution of the boundary term to the integral in
1432: Eq.~(\ref{eq:end-end-length}) is subdominant in the limit $t\to 0$.
1433: Nevertheless, these boundary effects represent important corrections that should be
1434: taken into account in any experimental situation (i.e.~with finite $t$) of {\pull}
1435: and {\epull}.
1436:
1437: At first, however, let us consider the \emph{bulk} contribution to the end-to-end
1438: distance,
1439: \begin{equation}
1440: \label{eq:end-end-bulk}
1441: \avg{\Delta \overline{R}_\pa}(t)\equiv\mint{ds}{0}{L}
1442: \avg{\Delta\overline{\varrho}}\left[F(s,\tilde t
1443: \leq t),t\right] \;,
1444: \end{equation}
1445: which is universal in the sense that it does not depend on the boundary conditions of
1446: $\vec r_\pe$. In Eq.~(\ref{eq:end-end-bulk}), and in the following, we neglect contributions
1447: of higher order than $\Delta\varrho=\Ord{\epsilon}$ to the projected end-to-end
1448: distance. According to Eqs.~(\ref{eq:end-end-bulk},~\ref{eq:cg-eom}),
1449: the bulk change of the end-to-end distance is given by the arclength integral of
1450: the curvature of the time integrated tension $F(s,t)$,
1451: \begin{subequations}
1452: \begin{align}
1453: \label{eq:EED-from-tension}
1454: \avg{\Delta \overline{R}_\pa}(t)&=\hat \zeta^{-1} \mint{ds}{0}{L}F''(s,t)\\
1455: \label{eq:EED-from-tension-slopes}
1456: &=\hat \zeta^{-1} \left[ F'(L,t)-F'(0,t) \right]\;.
1457: \end{align}
1458: \end{subequations}
1459:
1460: The predictions are presented separately for the regime of tension propagation ($t\ll
1461: \totpa$) and the regime of OPT ($t\gg t_\star$), where the tension profiles are flat
1462: and time-independent up to sub-leading terms. For weak forces, an explicit expression
1463: for times $t<t_\star$ is given in Sec.~\ref{sec:small-forces-crossover}, which
1464: captures the crossover from tension propagation to the tension saturated OPT regime.
1465: As discussed in Sec.~\ref{sec:homogeneous-relaxation}, {\lelo} turns out to also have
1466: an additional time-window $\totpa< t < t_\star$, whose consequences for the
1467: end-to-end distance is described in
1468: Sec.~\ref{sec:homogeneous-relaxation-e-e-distance}. Since {\epull} is the only
1469: problem, in which both ends do not behave in the same way, we discuss this
1470: ``asymmetrical'' problem separately in Sec.~\ref{sec:e-e-epull}.
1471:
1472: \subsection{Tension propagation regime ($t\ll \totpa$)}
1473: \label{sec:unrelaxed-tension}
1474: In the time domain of tension propagation, where the boundary layers are growing in
1475: from both ends ($t\ll \totpa$), we may use the tension profiles for the
1476: semi-infinite (pseudo-) polymer from Sec.~\ref{sec:scaling-forms} (labeled by
1477: $F_\infty$ here) to approximate Eq.~(\ref{eq:EED-from-tension-slopes}) by
1478: \begin{equation}
1479: \label{eq:EED-from-tension-2}
1480: \avg{\Delta \overline{R}_\pa}(t)\stackrel{t\ll \totpa}{\sim}-2\hat
1481: \zeta^{-1} F_\infty'(0,t)
1482: \end{equation}
1483: for scenarios with \emph{two} equally treated ends. Now, if $t$ falls into a regime
1484: where the tension exhibits scaling,
1485: \begin{equation}
1486: \label{eq:F-scaling}
1487: F_\infty(s,t)\propto t^{\alpha+1} \hat F_\infty\left(\frac s {t^z} \right) \;,
1488: \end{equation}
1489: we immediately obtain from Eq.~(\ref{eq:EED-from-tension-2}) the
1490: power-law
1491: \begin{equation}
1492: \label{eq:ee-growth-law-principle}
1493: \avg{\Delta \overline{R}_\pa}(t)\propto - t^{\alpha+1-z} 2 \partial_{\xi} \hat
1494: F_\infty(\xi=0)
1495: \end{equation}
1496: for the growth of the end-to-end distance. The pre-factors can be calculated for all
1497: cases, because the scaling functions are known. In this way we obtain the list
1498: Tab.~\ref{tab:ww-growth-laws} of growth laws.
1499:
1500: \begin{table}
1501: \caption{Universal bulk contribution $\avg{\Delta\overline{R}_\pa}(t)$ to the dynamic change
1502: of the end-to-end distance in the limit
1503: $t\ll \totpa\ll t_L^\pe$.}
1504: \label{tab:ww-growth-laws}
1505: \begin{center}
1506: \begin{tabular}{c|c|c}
1507: Problem&$t\ll \tf$&$t\gg \tf$\\ \hline
1508: {\pull}&${\displaystyle \frac{2^{5/8}}{\Gamma(15/8)}
1509: \frac{\fe}{\sqrt{ \hat \zeta \lp }}t^{7/8}}$&${\displaystyle
1510: \frac{4}{\sqrt{3}}
1511: \left(\frac{2}{\pi}\right)^{1/4}\frac{\fe^{3/4}}{\sqrt{ \hat
1512: \zeta \lp }}t^{3/4}}$\\
1513: {\lelo}&$-{\displaystyle \frac{2^{5/8}}{\Gamma(15/8)}
1514: \frac{\fe}{\sqrt{ \hat \zeta \lp }}t^{7/8}}$&${\displaystyle
1515: - 2.477\frac{\fe^{1/4}}{\sqrt{ \hat
1516: \zeta \lp }}t^{1/2}}$\\
1517: \end{tabular}
1518: \end{center}
1519: \end{table}
1520:
1521:
1522: \subsection{OPT regime ($t\gg t_{\star}$)}
1523: \label{sec:tension-relaxed}
1524: When calculating the release of stored length in OPT the tension profile
1525: $\fopt=f(s=0)=\text{const.}$ is assumed to be stationary and flat, so that
1526: \begin{eqnarray}
1527: \label{eq:OPT-Rpa}
1528: &\avg{\Delta \overline{R}_\pa}(t)=-L \avg{\Delta \overline{\varrho}}
1529: \left[ \fopt,t \right]\;,\nonumber &\\
1530: &\qquad\qquad\qquad \text{for }t\gg t_\star \;.&
1531: \end{eqnarray}
1532: In the case of {\lelo}, the quantity $\avg{\Delta \overline{\varrho}}(\fopt=0,t)$ has
1533: been explicitely calculated in Eq.~(\ref{eq:stole-lelo-OPT}). For {\pull},
1534: $\avg{\Delta \overline{\varrho}}(\fopt=\fe,t)$ can be evaluated from
1535: Eq.~(\ref{eq:change-stored-length}) in a similar straightforward manner, because the
1536: tension is spatially constant. The corresponding growth laws are summarized in
1537: Tab.~\ref{tab:ww-growth-laws-2}.
1538:
1539: Compared to the tension propagation regime, the growth laws in the OPT regime are
1540: slowed down, see Tabs.~\ref{tab:ww-growth-laws},~\ref{tab:ww-growth-laws-2}.
1541: Actually, for all cases except {\lelo} the corresponding growth laws obey
1542: \[ \avg{\Delta \overline{R}_\pa}^{\text{OPT}}\propto \avg{ \Delta
1543: \overline{R}_\pa}^{\text{MSPT}} t^{-z}\;.\] This can be understood in terms of the
1544: scaling arguments used Part~I. There, we took tension propagation heuristically into
1545: account by assuming that the stored length release, as given by OPT, is restricted to
1546: the boundary layer of size $\lpa(t)$. Then, one has
1547: \begin{equation}
1548: \label{eq:dr-estimate1}
1549: \avg{\Delta \overline{R}_\pa}^{\text{OPT}}\simeq -\lpa(t) \avg{\Delta
1550: \overline{\varrho}}(\fopt,t) \qquad \text{for }t\ll\totpa
1551: \end{equation}
1552: as compared to
1553: \begin{equation}
1554: \label{eq:dr-estimate1}
1555: \avg{\Delta \overline{R}_\pa}^{\text{MSPT}}\simeq - L
1556: \avg{\Delta\overline{\varrho}}(\fopt,t) \qquad \text{for }t\gg t_\star \;.
1557: \end{equation}
1558: This conforms with the heuristic rule noticed above,
1559: \begin{equation}
1560: \label{eq:OPT-MSPT}
1561: \avg{\Delta \overline{R}_\pa}^{\text{OPT}}\approx
1562: \frac{L}{\lpa(t)}
1563: \avg{\Delta \overline{R}_\pa}^{\text{MSPT}}
1564: \sim t^{\alpha+1-2z}\;.
1565: \end{equation}
1566:
1567:
1568:
1569: \begin{table}
1570: \caption{Universal bulk contribution $\avg{\Delta\overline{R}_\pa}(t)$ to the dynamic change
1571: of the end-to-end distance in the limit
1572: $t_\star\ll t \ll t_L^\pe$.}
1573: \label{tab:ww-growth-laws-2}
1574: \begin{center}
1575: \begin{tabular}{c|c|c}
1576: Problem&$t\ll \tf$&$t\gg \tf$\\ \hline
1577: {\pull}&${\displaystyle
1578: \frac{L
1579: \fe}{\lp\Gamma(7/4)}\left(\frac{t}{2}\right)^{3/4}}$~\cite{granek:97,gittes-mackintosh:98_pub}
1580: &${\displaystyle
1581: \frac{L}{\lp}\sqrt{\frac{2\fe t}{\pi}}}$\\
1582: {\lelo}&$-{\displaystyle
1583: \frac{L \fe}{\lp\Gamma(7/4)}\left(\frac{t}{2}\right)^{3/4}}
1584: $ &${\displaystyle \frac{-2^{3/4}}
1585: {\Gamma(1/4)} \frac{L}{\lp}t^{1/4}}$\\
1586: \end{tabular}
1587: \end{center}
1588: \end{table}
1589:
1590:
1591: \subsection{{\lelo} in the limit $\totpa\ll t\ll t_{\star}$}
1592: \label{sec:homogeneous-relaxation-e-e-distance}
1593: The intuitive rule in Eq.~(\ref{eq:OPT-MSPT}) fails for {\lelo} in the limit $t\gg\tf$,
1594: which indicates that this is an exceptional scenario. As discussed in
1595: Sec.~\ref{sec:homogeneous-relaxation}, $t_\star=L^8/\lp^4$ cannot be identified with
1596: $\totpa=L^2 \lp^{-1} \hat\zeta \fe^{-3/2}$ in this case. There exists a time window
1597: $\totpa\ll t \ll t_\star$ that expands in the limit of large forces, $\fe\gg
1598: \lp^2/L^4$. We have shown that the tension exhibits homogeneous relaxation in this
1599: novel regime. With the slope of the tension at the ends
1600: \begin{equation}
1601: \label{eq:lelo-tension-slopes}
1602: \partial_s f\vert_{s=\left\{ {0 \atop L}\right\}}=\pm \frac 1 {16} \left( \frac 3 2
1603: \right)^{2/3} \frac{1}{L}\left(\frac{\hat \zeta L^2}{\lp t}\right)^{2/3}
1604: \end{equation}
1605: the growth law
1606: \begin{equation}
1607: \label{eq:homogenous-lelo-3}
1608: \avg{\Delta \overline{R}_\pa}(t)\simeq - 18^{1/3} \left(\frac{L t}{\hat \zeta
1609: \lp^2}\right)^{1/3}
1610: \qquad (\lelo)
1611: \end{equation}
1612: follows from Eq.~(\ref{eq:EED-from-tension-slopes}). We expect the growth law
1613: $\avg{\Delta \overline{R}_\pa}\propto t^{1/3}$ during homogeneous tension relaxation to hold even
1614: for chains with $L\gg\ell_p$. The example of retracting DNA will be discussed as an
1615: experimental outlook in Sec.~\ref{sec:possible-exps}. The exponent $1/3$ coincides
1616: with that obtained by an adiabatic application of the stationary force-extension
1617: relation~\cite{bustamante-marko-siggia-smith:94} to a
1618: ``frictionless''~\cite{bhobot-wiggins-granek:04} polymer with attached beads at its
1619: ends~\cite{Hallatschek:F:K::94:p077804:2005}.
1620:
1621: For times $t\gg t_\star=L^8/\lp^4$ the growth law in Eq.~(\ref{eq:homogenous-lelo-3})
1622: crosses over to the one noted in Tab.~\ref{tab:ww-growth-laws-2}. Interestingly, both
1623: growth laws appearing for $t\gg \totpa$ are independent of the initial tension $\fe$.
1624: In both cases, the initial conditions are completely ``forgotten'' once the tension
1625: has propagated through the whole polymer. An overview over the time scales separating
1626: the diverse regimes for {\lelo} (as compared to {\pull}) will be given in
1627: Sec.~\ref{sec:lelo-vs-pull}.
1628:
1629:
1630: \subsection{{\pull} and {\lelo} for small forces}
1631: \label{sec:small-forces-crossover}
1632: Provided the external force is smaller than the critical Euler buckling force of the
1633: polymer, $\fe< L^{-2}$, the crossover time $\tf$ exceeds the terminal relaxation time
1634: $\totpe$, hence the linearized PIDE, Eq.~(\ref{eq:pide-linearized-lt-1}), applies
1635: throughout the contour relaxation. The linearity allows to solve the PIDE for a
1636: polymer of finite length, and we obtain an analytic description of the crossover
1637: between the asymptotic power laws $\Delta\overline{R}_\pa(t)\propto t^{7/8}$ for
1638: $t\ll t_L^\pa$ and $\Delta\overline{R}_\pa(t)\propto t^{3/4}$ for $t_\star\ll t \ll
1639: t_L^\pe$ (cf.~Tabs.~\ref{tab:ww-growth-laws},~\ref{tab:ww-growth-laws-2}). To this
1640: end, let us first reinstall original units (which are better adapted for the
1641: present purpose) into the linearized PIDE, Eq.~(\ref{eq:pide-linearized-lt-1}),
1642: \begin{equation}
1643: \label{eq:PIDE-linear-units}
1644: \partial_s^2 F(s,z)=\frac{z^{1/4} \hat \zeta}{2^{3/4} \lp} F(s,z)\;,
1645: \end{equation}
1646: where $F(s,z)$ is the Laplace transformation of the integrated tension,
1647: \begin{equation}
1648: \label{eq:Laplace trafo}
1649: F(s,z)\equiv \mint{dt}{0}{\infty}e^{-z t}F(s,t)\;.
1650: \end{equation}
1651: For the boundary conditions of {\pull}, Eqs.~(\ref{eq:bc-infty},
1652: \ref{eq:bc-origin-1}), the solution to Eq.~(\ref{eq:PIDE-linear-units}) is given by
1653: \begin{equation}
1654: \label{eq:weak-pulling-finite-polymer}
1655: F(s,z)=\fe z^{-2}\frac{\cosh\left[(\hat \zeta/\lp)^{1/2} (z/8)^{1/8} (s-L/2)
1656: \right]} {\cosh\left[(\hat \zeta/\lp)^{1/2} (z/8)^{1/8} L/2 \right]}
1657: \end{equation}
1658: This implies a growth of the end-to-end distance of
1659: \begin{equation}
1660: \label{eq:eed-weak-forces-finitep-laplace}
1661: \begin{split}
1662: \avg{\Delta \overline R_\pa}(z)&=-2\hat \zeta^{-1} \partial_s F|_{s=0} \\
1663: &=\frac{2^{5/8}\fe}{z^{15/8} (\hat \zeta\lp)^{1/2}}\tanh\left[ \left( \frac{\hat
1664: \zeta}{\lp} \right)^{1/2}\left( \frac{z}{8} \right)^{1/8} \frac{L}{2}
1665: \right]\;.
1666: \end{split}
1667: \end{equation}
1668: The Laplace back-transform of Eq.~(\ref{eq:eed-weak-forces-finite-polymer}) takes
1669: the form
1670: \begin{equation}
1671: \label{eq:eed-weak-forces-finite-polymer}
1672: \begin{split}
1673: \avg{\Delta \overline R_\pa}(t)&=\mint{\frac{dz}{2\pi
1674: i}}{-i\infty+\epsilon}{i\infty+\epsilon} e^{z t} \avg{\Delta \overline
1675: R_\pa}(t)\\
1676: &=\frac{\fe \hat \zeta^3 L^7}{\lp^4} \Upsilon\left[ \left( \frac{t}{L^4} \right)
1677: \left( \frac{\lp}{\hat \zeta L} \right)^4 \right]
1678: \end{split}
1679: \end{equation}
1680: where $\Upsilon(\tau)$ is a scaling function, given by
1681: \begin{equation}
1682: \label{eq:Upsilon}
1683: \begin{split}
1684: \Upsilon(\tau)&= \frac{2^{5/8}}{\pi}\mint{ds}{0}{\infty}\frac{e^{-x
1685: \tau}-1}{x^{15/8}}\\
1686: &\times \Im\left[ e^{-i\frac 78 \pi} \tanh\left( 2^{-11/8}x^{1/8}e^{i
1687: \pi/8} \right) \right]
1688: \end{split}
1689: \end{equation}
1690: We have depicted $\Upsilon(\tau)$ in a double logarithmic plot in
1691: Fig.~\ref{fig:pulling-finite-polymer-weak-forces} together with the corresponding
1692: asymptotics from Tabs.~\ref{tab:ww-growth-laws},~\ref{tab:ww-growth-laws-2}. Note
1693: that the time $\tau_c\approx 6.55 \times 10^{-4}$ where the asymptotic lines cross is
1694: a good indication for the crossover occurring in the exact solution. Assuming that
1695: this also holds quite generally, it is possible to obtain estimations of when the
1696: crossover between linear and nonlinear regimes should occur in the natural time units
1697: $\tf$. To this end, one simple equates the corresponding asymptotic power laws
1698: (including the exact pre-factors) for the growth of the end-to-end distance. As for
1699: the present case with $\tau_c\approx 6.55 \times 10^{-4}$, these crossover times are
1700: typically not of order one in natural units because of the numerical proximity of the
1701: exponents of the asymptotic power laws, which are $7/8$ and $3/4$ in the present
1702: case. In a given experimental situation, one should therefore check carefully which
1703: regime is expected by comparing experimental time scales with these unusual crossover
1704: times.
1705: \begin{figure}
1706: \includegraphics[width=\columnwidth]{pulling-finite-polymer-psfraged.eps}
1707: \caption{The scaling function $\Upsilon(\tau)$ in a double logarithmic plot. The
1708: crossover of the scaling behavior is close to the time $\tau_c =6.55\times
1709: 10^{-4}$ where the asymptotic short-(blue) and longtime (red) asymptotics
1710: cross.}
1711: \label{fig:pulling-finite-polymer-weak-forces}
1712: \end{figure}
1713:
1714:
1715:
1716: \subsection{FDT and linear response}
1717: \label{sec:tens-fluct}
1718: According to the fluctuation-dissipation theorem
1719: (FDT~\cite{kubo83bookI,kubo91bookII}), the fluctuations in the end-to-end distance
1720: should be related to the linear response of the polymer. It is tempting to interpret
1721: the linear short-time regime, where $\avg{\Delta \overline{R}}\simeq \fe t^{7/8}(\hat
1722: \zeta \lp)^{-1/2}$, as linear response in the usual sense, in which case the
1723: fluctuations of the end-to-end distance should, according to the FDT, scale as
1724: $\avg{\Delta \overline{R}^2}\simeq t^{7/8}{\hat \zeta}^{-1/2} \lp^{-3/2}$ in
1725: equilibrium. However, fluctuations of this strength are inconsistent with a
1726: deterministic tension dynamics, because they generate friction forces over a scale
1727: $\lpa(t)\simeq t^{7/8}(\lp/\hat \zeta)^{1/2}$ and thus imply tension fluctuations of
1728: magnitude
1729: \begin{equation}
1730: \label{eq:tens-fluct}
1731: \delta f\simeq \hat \zeta \lpa(t) \avg{\Delta \overline{R}}/t\simeq
1732: (\zeta/\lp)^{1/4} t^{-7/16}
1733: \end{equation}
1734: Tension fluctuations exceed the applied force in magnitude at any given time for
1735: small enough external forces. Hence, in the limit $\fe\to 0$ while $\epsilon\ll1$ is
1736: fixed, the tension cannot be considered as a deterministic quantity. Recall however,
1737: that our MSPT analysis in Part~I was based on the limit $\epsilon\to 0$ while $\fe$
1738: is fixed, and only in this limit the self-averaging argument of Part~I applies.
1739: Extending our results to the usual linear response limit corresponds to the
1740: uncontrolled approximation $f \rpe'^2\to f\avg{\rpe'^2}$. To analyze this limit more
1741: carefully, one has to solve the stochastic PIDE that obtained in Part~I before the
1742: self-averaging argument was employed. However, since the $7/8-$scaling of the
1743: fluctuations has already been confirmed in simulations~\cite{everaers-Maggs:99}, we
1744: expect that such a more rigorous analysis would yield the same scaling but a
1745: pre-factor different from the one of the deterministic short-time law
1746: (Tab.~\ref{tab:ww-growth-laws}).
1747:
1748:
1749: \subsection{{\lelo} versus {\pull} (overview)}
1750: \label{sec:lelo-vs-pull}
1751: The diverse regimes and their range of validity are summarized in
1752: Fig.~\ref{fig:lelo-vs-pull} for the {\pull} and the complementary
1753: {\lelo} problem. It depicts the crossover time scales as a function of
1754: the externally applied tension $\fe$. The line $\tf\simeq \fe^{-2}$
1755: separates ``linear'' from ``nonlinear'' behavior. The symmetry of the
1756: graph for $t<\tf$ with respect to the $\fe=0$-axis indicates that the
1757: scenario of {\pull} can indeed be considered as the inverse scenario
1758: of {\lelo} for weak forces $f\ll f_c$, or, more generally, on short
1759: times. This symmetry is lost in the nonlinear regime. The growing
1760: importance of uniform tension relaxation for {\lelo} with increasing
1761: initial tension $\fe$ becomes particularly apparent on the log-scale
1762: of the figure. Due to the particular choice of units ($t/L^4$ and
1763: $f/L^{-2}$) the regimes of nonlinear growth of the boundary layers
1764: appear relatively narrow in Fig.~\ref{fig:lelo-vs-pull}. Which of the
1765: various regimes will pre-dominantly be observed in measurements
1766: actually depends strongly on the ratio $L/\lp$ and on the
1767: experimentally accessible time scales.
1768: \begin{figure}
1769: \includegraphics[width=\columnwidth]
1770: {lelo-vs-pulling-sregion-psfraged.eps}
1771: \caption{Characteristic times (logarithmic scale) for {\pull} and {\lelo} against
1772: the applied external force $\fe$ (linear scale). The time $t_\star$ (stars)
1773: separates regions where ordinary perturbation theory (OPT) applies (dark shaded)
1774: from regions (light shaded) of linear (hatched) and nonlinear tension propagation
1775: and from homogeneous tension relaxation (white). Whereas longitudinal friction is
1776: negligible for $t>t_\star$, it limits the dynamics for $t<t_\star$. The innermost
1777: funnel-shaped region indicates the regime where the tension fluctuations are
1778: important (defined by $f<\delta f$ with $\delta f$ given by
1779: Eq.~(\ref{eq:tens-fluct})).}
1780: \label{fig:lelo-vs-pull}
1781: \end{figure}
1782:
1783:
1784: \subsection{\epull}
1785: \label{sec:e-e-epull} {\epull} was excluded from the preceeding discussion because it
1786: is the only ``asymmetric'' problem, as both ends of the polymer behave differently.
1787: In the tension propagation regime ($t\ll\totpa$), the left end is constrained to move
1788: with constant velocity, while the right end experiences no driving force. Hence, we
1789: have
1790: \begin{equation}
1791: \label{eq:epull-growth-law}
1792: \avg{\Delta \overline{R}_\pa}(t)=v t \;,\qquad \text{for } t\ll\totpa\;.
1793: \end{equation}
1794: After the boundary layer of non-zero tension has reached the free end, $t\gg \totpa$,
1795: the right end starts to move. Then, the tension profile becomes linear as for a
1796: straight rod dragged through a viscous solvent. The further contour relaxation is up
1797: to pre-factors identical to the {\pull} problem for $t\gg \totpa\approx t_\star$.
1798:
1799: \subsection{\push}
1800: \label{sec:pushing}
1801:
1802: For completeness, we mention the scenario of a filament being compressed by external
1803: longitudinal forces. This scenario has some subtleties. {\push} increases the stored
1804: length exponentially for $t\gg \tf$ by virtue of the Euler Buckling instability and
1805: generates a situation where the weakly bending approximation is not valid anymore.
1806: Then, hairpins are generated~\cite{ranjith:02} and for $t\gg t_f$ the rigidly
1807: oriented driving forces \emph{pull} on those hair pins. Our theory is only applicable
1808: at short times. For $t\ll t_f$ the response of the system is linear in the driving
1809: force, irrespective of the sign.
1810:
1811:
1812: \subsection{Boundary effects}
1813: \label{sec:microscale}
1814: Up to now, we only discussed the change $\avg{\Delta \overline{R}_\pa}(t)$ in
1815: projected end-to-end distance corresponding to the change
1816: $\avg{\Delta\overline{\varrho}}$ of the stored length in the bulk. For a hinged,
1817: respectively, clamped semi-infinite polymer, the as yet missing boundary
1818: contribution $X^{h/c}(t)\equiv\avg{\Delta R^{h/c}_\pa}-\avg{\Delta\overline{R}_\pa}$
1819: is given by
1820: \begin{eqnarray}
1821: \label{eq:bc-contri-1}
1822: && X^{h/c}(t)=\pm2\mint{ds}{0}{\infty}\mint{\frac{dq}{\pi\lp}}{0}{\infty}
1823: \left\{\frac{e^{-2q^2[q^2 t+ F(s,t)]}-1}{q^2+f_<}\right. \nonumber\\
1824: &&+\left. 2
1825: q^2\mint{d\tilde t}{0}{t}e^{-2q^2\left[q^2(t-\tilde
1826: t)+F(s,t)-F(s,\tilde t)\right]}\right\}\cos(2 q s) \;.\nonumber\\
1827: \end{eqnarray}
1828: As discussed in Part~I, the boundary dependent term
1829: of the stored length decays on a length scale of $\Ord{1}$ due to the cosine factor.
1830: Since the tension decays on a much larger length scale of order
1831: $\Ord{\epsilon^{-1/2}}$, it is permissible to use the (integrated) tension at $s=0$
1832: to evaluate the arclength integral in Eq.~(\ref{eq:bc-contri-1}).
1833:
1834: Upon replacing $F(s,t)\to F(0,t)$ and using
1835: \[\mint{ds}{0}{\infty}\cos(2 q s)=(\pi/2)\delta(q)\;,\] the integral in Eq.~(\ref{eq:bc-contri-1}) can be evaluated,
1836: \begin{equation}
1837: \label{eq:Rpa-boundary}
1838: X^{h/c}(t)=\pm\lim_{q\to 0}\frac 2 {\lp} \frac{q^2(q^2 t +F(0,t))}{q^2+f_<}\;,
1839: \end{equation}
1840: which vanishes, unless
1841: \begin{equation}
1842: \label{eq:Rpa-boundary2}
1843: f_<=0 \quad \Rightarrow \quad X^{h/c}(t)=\pm\frac {2F(0,t)} {\lp}\;.
1844: \end{equation}
1845: On the semi-infinite arclength interval, these boundary contributions are therefore
1846: nonzero only for {\pull} and {\epull}, and are summarized in
1847: Tab.~\ref{tab:tab:bc-effects}. Note that, the boundary effects of clamped (hinged)
1848: ends tend to reduce (increase) the longitudinal response of the polymer in comparison
1849: to the bulk response. This may be explained as follows. Close to a clamped end, a
1850: polymer is more stretched out than in the bulk because $\vec r_\pe'^2$ is constraint
1851: to approach zero at the end. As a consequence, the end portion of the polymer is less
1852: able to store or release excess length. For hinged ends, the boundary conditions act
1853: just in the reverse direction.
1854:
1855: \begin{table}
1856: \caption{Boundary contribution to the change $\avg{\Delta R^{h/c}_\pa}(t)$ of the end-to-end distance for
1857: hinged, respectively, clamped ends in the limit $t_L^\pe \ll t$.}
1858: \label{tab:tab:bc-effects}
1859: \begin{tabular}{c|c}
1860: Problem&$\avg{\Delta R^{h/c}_\pa}(t)-\avg{\Delta \overline{R}_\pa}(t)$\\ \hline
1861: {\pull}&$\pm 2t \fe/\lp$\\
1862: {\epull}&$\pm 2t \fe/\lp$\\
1863: {\lelo}&$\left\{ {\pm 2t \fe/\lp\text{, for }f_<\ll L^{-2} \atop
1864: 0 \text{, for }f_<\gg L^{-2}}
1865: \right.$ \\
1866: \end{tabular}
1867: \end{table}
1868:
1869:
1870: The strict vanishing of the boundary term for any finite $f_<$ and the discontinuity
1871: at $f_<=0$ is a consequence of the assumed infinite half-space. For a polymer of
1872: finite length, it turns out that the boundary term approaches zero for
1873: pre-stretching forces larger than the critical Euler-Buckling force, $f_<\gg
1874: f_c\equiv L^{-2}$. This can be seen by studying the finite integral
1875: $\mint{ds}{0}{L}(\avg{\Delta \varrho^{h/c}}(s,t)-\avg{\Delta \overline{\varrho}}(s,t))$. For
1876: forces $f_<\ll f_c$, the integrand saturates at a plateau of magnitude
1877: $\Ord{f_<^{3/2}t/\lp }$ for $\lpe(t)<s<f_<^{-1/2}$ before it finally decays to zero.
1878: Within the semi-infinite integral, the integral over this long plateau cancels the
1879: contribution stemming from $s<\lpe(t)$, as required by Eq.~(\ref{eq:Rpa-boundary}).
1880: However, for $L\ll f_<^{-1/2}$ the contribution from the plateau may be neglected. As
1881: a consequence, the value of the integral is for $f_<\ll f_c$ given by $\pm 2f_<
1882: t/\lp$. This asymptotic behavior is important for {\lelo}, as noted in
1883: Tab.~\ref{tab:tab:bc-effects}.
1884:
1885:
1886:
1887:
1888:
1889:
1890: Upon comparing Tab.~\ref{tab:tab:bc-effects} with Tabs.~\ref{tab:ww-growth-laws},
1891: \ref{tab:ww-growth-laws-2}, one may think that boundary effects are always
1892: subdominant in the short time limit. However, our calculations were specialized to
1893: hinged or clamped boundary conditions. In many experimental situations, one has to
1894: deal with \emph{free} boundary conditions. Somewhat tedious but rigorously, these
1895: boundary conditions can be taken into account by means of the correct susceptibility,
1896: which can only be given in terms of an integral. Here, we discuss effects related to
1897: free boundary conditions on a heuristic basis and show that they generate a
1898: \emph{dominant} contribution to the change in the end-to-end distance for {\pull},
1899: which is proportional to $t^{3/4}$.
1900:
1901:
1902:
1903:
1904:
1905: The argument is based on the observation, that external forces that act in the
1906: longitudinal direction ($\pa$) while the polymer is free, automatically introduce
1907: small (of order $\Ord{\epsilon^{1/2}}$) transverse forces at the ends. This follows
1908: from the force balance equation of a semiflexible polymer (cf.~Part~I)
1909: \begin{equation}
1910: \label{eq:force-balance}
1911: \kappa\, \vec{r}''' +\fel = f\, \vec{r}' \;,
1912: \end{equation}
1913: where $\fel(s)$ is the elastic force acting at arc length $s$. At the ends, where
1914: $\fel$points in longitudinal direction to cancel the external pulling force, one obtains
1915: from the projection of Eq.~(\ref{eq:force-balance}) onto the transverse axis
1916: \begin{equation}
1917: \label{eq:perp-endforces}
1918: \vec r_\pe'''=f \vec r_\pe'=\fe \vec r_\pe'+\Ord{\epsilon} \qquad \text{(at the ends)} \;.
1919: \end{equation}
1920: Thus, $\vec r_\pe'''$ is typically non-zero at the ends because the slope $\vec
1921: r'=\Ord{\epsilon^{1/2}}$ fluctuates. This, however, corresponds to a transverse
1922: force at the end, which results in a transverse deformation. The corresponding bulge
1923: of contour is only visible on the micro-scale because it spreads with the transverse
1924: correlation length $\lpe(t)\simeq t^{1/4}$. Nevertheless this deformation may
1925: dominate the growth of the end-to-end distance, as is demonstrated for {\pull}: With
1926: the transverse bulk susceptibility scaling as $\overline{\chi}(t)\simeq t^{-1/4}$ we
1927: estimate the magnitude of transverse deformation $\Delta r_\pe(s=0,t)\equiv r_\pe
1928: (0,t)-r_\pe(0,0)$ induced by a transverse force of magnitude
1929: $\ord{\fe\epsilon^{1/2}}$ by
1930: \begin{equation}
1931: \label{eq:tdispl-estimate}
1932: \ord{\avg{\abs{\Delta r_\pe(0,t)}}}\simeq \fe
1933: \epsilon^{1/2} t \, \overline{\chi}(t)\simeq \epsilon^{1/2} t^{3/4}\fe \;.
1934: \end{equation}
1935: The displacement of the end couples to the projected length because of the mismatch
1936: of the end-tangent with the $\pa$-axis by a small angle of typical magnitude
1937: $\avg{\abs{r_\pe'(s=0)}}\simeq \ord{\epsilon^{1/2}}$. The expected growth of
1938: end-to-end distance due to this effect is therefore estimated by
1939: \begin{equation}
1940: \label{eq:micro-fluctuations}
1941: \begin{split}
1942: &\avg{\Delta R^{h/c}_\pa}(t)-\avg{\Delta
1943: \overline{R}_\pa}(t)\simeq\\&
1944: \avg{\abs{r_\pe'(s=0)}} \times\avg{\abs{\Delta
1945: r_\pe(0,t)}}\simeq\epsilon t^{3/4}
1946: \fe=\frac{L}{\lp}t^{3/4}\fe \;.
1947: \end{split}
1948: \end{equation}
1949: This dominates over the growth law of the bulk, which scales like
1950: $t^{7/8}/\sqrt{\lp}$ on short times. The only way to avoid the outlined effect is to
1951: apply the external force strictly tangentially to the end-tangents, which is however
1952: somewhat unrealistic. The same problem will experimentally arise in the short time
1953: limit of {\lelo} if the pre-stretching force was not applied strictly tangentially.
1954: However, the long-time limits are unaffected by this subtlety because the bulk
1955: contributions dominate over contributions from the end.
1956:
1957: To our knowledge, these end-effects have so far masked the subdominant
1958: $t^{7/8}$-contribution in experiments that monitored the time-dependent end-to-end
1959: distance (we note that in Ref.~\cite{legoff-amblard-furst:02} the $7/8$-scaling is
1960: inferred from a corresponding scaling of the measured shear modulus of an active
1961: gel). As we outline in Sec.~\ref{sec:smfs}, force spectroscopy, on the contrary, may
1962: allow to measure the tension dynamics, which is itself truly \emph{independent} of
1963: the boundary conditions imposed on the transverse displacements.
1964:
1965:
1966:
1967:
1968:
1969:
1970:
1971: \section{Suggestions for experiments}
1972: \label{sec:possible-exps}
1973: While many experiments have been done concerning flexible polymers in external force
1974: fields~\cite{perkins-smith-chu:97,quake-babcock-chu:97}, the available measurements
1975: on driven stiff or pre-stretched polymers is not sufficient to verify our
1976: predictions. Most of these experiments have monitored the transverse and
1977: longitudinal response on intermediate times where OPT is valid
1978: (e.g.~\cite{legoff-hallatschek-frey:02}). Investigations concerning the longitudinal
1979: short-time dynamics are
1980: scarce~\cite{bhobot-wiggins-granek:04,legoff-amblard-furst:02,maier-seifert-raedler:02}.
1981: In the following, we propose several assays that might be able to fill this gap.
1982:
1983: To facilitate the application of our predictions, we reintroduce the parameters
1984: $\kappa$ and $\zeta$ for the following.
1985:
1986: \subsection{Strongly stretched DNA}
1987: \label{sec:dna-exp}
1988:
1989: The experimental verification of most of our results requires stiff
1990: polymers with a total length much smaller than their persistence
1991: length. A remarkable exception is {\lelo} in the regime of
1992: homogeneous tension relaxation, which was discussed in
1993: Sec.~\ref{sec:homogeneous-relaxation}. For polymers with $L\gg\lp$,
1994: like a typical DNA molecule, this novel regime appears if the
1995: pre-stretching force obeys
1996: \begin{equation}
1997: \label{eq:lelo-exp-large-forces}
1998: \fe\gg \fe_{\lp}\equiv \kappa \lp^{-2}=\frac{k_B T}{\lp} \;,
1999: \end{equation}
2000: where $\fe_{\lp}$ is the Euler buckling force corresponding to the
2001: buckling length $\lp$. For forces much larger than the cross-over
2002: force $\fe_{\lp}$, the polymer may be considered as weakly bending.
2003: For those long, but strongly stretched polymers our analysis of
2004: {\lelo} predicts the following.
2005:
2006: After the stretching force has been released the tension first
2007: propagates through the filament. This
2008: regime~\cite{brochard-buguin-de_gennes:99,Hallatschek:F:K::94:p077804:2005}
2009: ends at the time
2010: \begin{equation}
2011: \label{eq:lelo-exp-crossover}
2012: \totpa=\frac{\zeta_\pa L^2 \lp}{k_b T}\left( \frac{\fe}{\fe_{\lp}}
2013: \right)^{-3/2}\equiv \tau_{R} \left( \frac{\fe}{\fe_{\lp}} \right)^{-3/2}
2014: \end{equation}
2015: when the tension has to propagated through the filament. Here, $\tau_{R}\equiv
2016: \zeta_\pa L^2 \lp /(k_B T)$ scales like the longest relaxation time of a Rouse chain
2017: with segment length $\lp$. The characteristic time $\totpa\ll \tau_{R}$ marks the
2018: cross-over to a regime where the tension profile is roughly parabolic and slowly
2019: decays according to the power law $f(t)\propto t^{-2/3}$. This is associated with the
2020: projected length $R_\pa$ growing like
2021: \begin{equation}
2022: \label{eq:lelo-exp-growth-law}
2023: \avg{\Delta R_\pa(t)}\equiv \avg{R_\pa(t)-R_\pa(0)}=
2024: 18^{1/3} L \left( \frac{t}{\tau_R}
2025: \right)^{1/3} \;.
2026: \end{equation}
2027:
2028: The above analysis strictly holds for the portion of the polymer that
2029: stays weakly bending. This is not the case at the boundaries, for
2030: which we refer to existing theories. According to the stem-flower
2031: model of Refs.~\cite{brochard:93,brochard:95,manneville:96} the
2032: boundaries will develop in time $t$ a ``flower'' of arc length
2033: \begin{equation}
2034: \label{eq:flower-length}
2035: \lflower(t)\simeq L \left( \frac{t}{\tau_{R}} \right)^{1/2} \;,
2036: \end{equation}
2037: thereby reducing the end-to-end distance by an amount of the same order of
2038: magnitude as $\lflower$ itself. It is seen, that for $t\ll \tau_R$ the shrinkage of
2039: the end-to-end distance due to the flower is much smaller than that due to the
2040: weakly bending part of the polymer (the stem), Eq.~(\ref{eq:lelo-exp-growth-law}).
2041:
2042: Thus, the evolution of the end-to-end distance should be described
2043: by Eq.~(\ref{eq:lelo-exp-growth-law}) even for flexible polymers if
2044: the pre-stretching force is large enough. Since DNA can be stretched
2045: by very large forces without un-zipping or destroying the covalent
2046: bonds, we think that the scaling $\avg{\Delta R(t)}\propto t^{1/3}$
2047: should be visible in a {\lelo}-experiment with DNA. The relevant
2048: quantities in an experiment with $\lambda$-phage DNA (as in
2049: Ref.~\cite{maier-seifert-raedler:02}) in aqueous solution would be
2050: \begin{subequations}
2051: \label{eq:lelo-exp-numbers}
2052: \begin{align}
2053: \lp&\approx 50\text{nm} \nonumber \\
2054: L&\approx 20 \mu\text{m} \nonumber\\
2055: a&\approx 2 \mu\text{m} \qquad \text{(thickness)}\nonumber\\
2056: \zeta_\pe&\approx 4 \pi \eta/\log(L/a)\approx 1.3 \times 10^{-3}
2057: \text{Pa s} \nonumber\\
2058: \fe_{\lp}&\approx 0.08 \text{pN} \nonumber\\
2059: \tau_R&\approx 7 \text{s} \;. \nonumber
2060: \end{align}
2061: \end{subequations}
2062: Though it might require extreme conditions to reach the asymptotic limit in
2063: Eq.~(\ref{eq:lelo-exp-growth-law}), any experiment with finite pre-stretching forces
2064: larger than $f_c$ should be suitable for deciding the question whether or not the
2065: stem dominates the retraction. One would then compare the data with a numerical
2066: solution~\cite{obermayer-kroy-frey-hallatschek:tbp} of the governing equation,
2067: Eq.~(\ref{eq:cg-eom}), which describes the stress relaxation in stiff worm-like
2068: chains.
2069:
2070:
2071:
2072: \subsection{Single molecule force spectroscopy}
2073: \label{sec:smfs}
2074:
2075: \subsubsection{{\epull}}
2076:
2077: As we already mentioned Sec.~\ref{sec:tensprop-summary}, {\epull} offers the possibility
2078: to measure the tension propagation by a force measurement. The experimental idea is
2079: illustrated in Fig.~\ref{fig:tweezer} (a). The polymer's left end is suddenly pulled,
2080: e.g., by an optical tweezer whose focus moves with constant velocity. As a
2081: consequence the pulled end follows the laser beam with (almost~\footnote{Due to the
2082: finite stiffness of the harmonic Laser potential, the trapped end will not be moved
2083: with \emph{exactly} the same velocity as the trap during a transient, in which the
2084: end approaches its steady state location within the Laser potential.}) constant
2085: velocity. By measuring the deflection of the trapped end from the center of the beam
2086: one can, in principle, extract the pulling force $\fe(t)$ and thus the size
2087: $\lpa(t)\simeq \fe(t)/(\hat \zeta v)$ of the boundary layer. However, since the
2088: laser beam is moving, it might represent some problems to dynamically extract the
2089: deflection.
2090:
2091: \begin{figure}
2092: \centerline{\includegraphics[width=.7\columnwidth]
2093: {tweezer1-psfraged.eps}}
2094: \caption{Two possible realizations of {\epull}. In (a) the polymer's left end is
2095: being dragged through the solvent with constant velocity $\vec v$, whereas in (b)
2096: the optical tweezer is immobile while the solvent flows with constant velocity
2097: $\vec v$. In both experiments, the length $\lpa(t)$ of the boundary layer is
2098: derived from the pulling force of the tweezer, which can be inferred from the
2099: displacement of the end within the focus of the tweezer.}\label{fig:tweezer}
2100: \end{figure}
2101:
2102:
2103: A solution to the latter problem is suggested by the following Gedanken experiment.
2104: Consider the above realization of {\epull} in the coordinate frame co-moving with
2105: the left tip, as illustrated in Fig.~\ref{fig:tweezer}(b). Then, it seems as if the
2106: bulk of the polymer was dragged by a homogeneous force field to the right while the
2107: left end is held fixed by the optical tweezer, just as if the optical tweezer was
2108: spatially fixed while the solvent was homogeneously flowing to the right. In fact,
2109: from the polymer's perspective there is no difference between both experiments. Thus,
2110: we propose to graft one end of a polymer by a tweezer or by the cantilever of an
2111: atomic force microscope. Then, a homogeneous force field (electric field or fluid
2112: flow) is suddenly turned on that pulls the bulk of the polymer to the right. The
2113: deflection of the tip gives the pulling force and hence the length of the boundary
2114: layer.
2115:
2116:
2117: \subsubsection{Onset of the nonlinear regime}
2118:
2119: One would like to estimate typical time sales for the diverse regimes
2120: of tension propagation and relaxation. As we have seen above, most of
2121: the time-scales crucially depend on the total length of the polymer,
2122: e.g., $\lpa(t)\propto L^8$ in the linear regime. This high tunability
2123: is of experimental advantage because one can adjust the setup to the
2124: observable time scales. On the other hand, it forbids to give
2125: \emph{typical} time-scales for those quantities.
2126:
2127:
2128: \begin{table}
2129: \caption{Threshold values $I_c=\sqrt{\kappa \zeta_\pe}$ for diverse biopolymers.}
2130: \label{tab:threshold-ic}
2131: \begin{center}
2132: \begin{tabular}{c|c}
2133: Polymer&$I_c$\\ \hline
2134: microtubuli & $1.6 \times 10^2$ pN\,ms$^{1/2}$ \\
2135: F--actin & $10$ pN\,ms$^{1/2}$\\
2136: Intermediate fil.s & $4$ pN\,ms$^{1/2}$ \\
2137: DNA & $0.9$ pN\, ms$^{1/2}$ \\
2138: \end{tabular}
2139: \end{center}
2140: \end{table}
2141:
2142:
2143: However, there are characteristic quantities that do not depend on the
2144: length of the polymer. The probably most interesting one is the
2145: threshold to the nonlinear regime: when the product of driving force
2146: and the square root of the applied time exceeds a certain value $I_c$
2147: the polymer response becomes nonlinear,
2148: \begin{equation}
2149: \label{eq:nonlinear regime}
2150: \fe \sqrt{t} \gg I_c\equiv\sqrt{\kappa \zeta_\pe}=\sqrt{k_B T \lp \zeta_\pe}
2151: \;.
2152: \end{equation}
2153: With persistence lengths of $\lp=7$mm, $17\mu$m, $2\mu$m, $50$nm for
2154: microtubuli\cite{mickey-howard:95,FrancescoPampaloni07052006},
2155: F-actin~\cite{legoff-hallatschek-frey:02}, intermediate filaments and
2156: DNA~\cite{bustamante-marko-siggia-smith:94}, respectively, and corresponding friction
2157: coefficients $\zeta_\pe\approx 4 \pi \eta /\ln(\lp/a)$ (transverse friction
2158: coefficient per length of a rod of length $\lp$), we have evaluated $I_c$ for some
2159: common biopolymers, see Tab.~\ref{tab:threshold-ic}. Those values can be used to
2160: decide whether the response of a given biopolymer under a ``typical'' time-dependent
2161: external longitudinal force is predominantly nonlinear or linear. During a power
2162: stroke, for instance, the molecular motor myosin exerts a force of about $5$pN on an
2163: actin filament during a time of roughly $1$ms (cycle time of the power stroke).
2164: Hence, the impulse of about $0.5 I_c$ is somewhat smaller than $I_c$ for actin, so
2165: that the actin response should be linear.
2166:
2167:
2168:
2169:
2170:
2171: \section{Summary}
2172: \label{cha:Summary}
2173:
2174: In this paper, we have studied the tension dynamics of a weakly bending semiflexible
2175: polymer in a viscous fluid theoretically. Starting from the coarse-grained equation
2176: of motion for the tension, Eq.~(\ref{eq:pide}), we elaborated the non-linear
2177: longitudinal dynamic response to various external perturbations (mechanical
2178: excitations, hydrodynamic flows, electrical fields \dots) that can be represented as
2179: sudden changes of boundary conditions. For the various scenarios we identified
2180: two-parameter scaling forms, that capture the crossover from linear to nonlinear
2181: tension dynamics. In the limit of large and small arguments, where the equilibrium
2182: structure of the polymer is self-affine, they were shown to reduce to one-parameter
2183: scaling forms, which could be calculated analytically in most cases. The growth law
2184: $\lpa(t)\sim t^z$ of the tension profiles could be inferred from the scaling
2185: variables of the respective scenarios. This enabled us to develop a unified theory of
2186: tension propagation. Not only does it contain all cases (correctly) studied in the
2187: literature so far. It also identifies their ranges of validity and provides new
2188: predictions. The recovered known results and our new predictions are summarized in
2189: Figs.~\ref{fig:lf},~\ref{fig:lelo-vs-pull} and
2190: Tabs.~\ref{tab:lf-growth-laws},~\ref{tab:ww-growth-laws} and
2191: \ref{tab:ww-growth-laws-2}.
2192:
2193: Various dynamic regimes should be well realizable for certain biopolymers. A novel
2194: regime of homogeneous tension relaxation is a particularly remarkable result from the
2195: experimental point of view (Sec.~\ref{sec:dna-exp}). In contrast to previous
2196: expectations, this new regime is predicted to dominate the relaxation of strongly
2197: stretched DNA. Moreover, it is an intriguing question, whether the tension
2198: propagation laws $\ell_\|(t)$ govern mechanical signal transduction through the
2199: cytoskeleton~\cite{shankar-pasquali-morse:2002,GardelVCBW03}. We expect that the
2200: force spectroscopical methods, proposed in Sec.~\ref{sec:smfs}, might be helpful to
2201: answer these questions.
2202:
2203: Inclusion of hydrodynamic interactions merely produce logarithmic corrections but
2204: would give rise to more interesting effects for polymerized membranes to which our
2205: discussion could be generalized with otherwise little change. Other natural
2206: generalizations including the transverse nonlinear response of
2207: polymers~\cite{obermayer-hallatschek2:tbp}, quenches in the persistence
2208: length~\cite{obermayer-kroy-frey-hallatschek3:tbp} and more complex force
2209: protocols~\cite{obermayer-hallatschek-frey-kroy:tbp} are currently also under
2210: investigation.
2211:
2212:
2213:
2214: \section{Acknowledgments}
2215: \label{sec:ack}
2216: It is a pleasure to acknowledge helpful conversations with Benedikt Obermayer, who,
2217: in addition, corrected several pre-factors. This research was supported by the
2218: German Academic Exchange Service (DAAD) through a fellowship within the
2219: Postdoc-Program (OH) and by the Deutsche Forschungsgemeinschaft through grant no.~Ha
2220: 5163/1 (OH) and SFB 486 (EF).
2221:
2222:
2223: \appendix
2224:
2225:
2226: \begin{table}[b]
2227: \caption{Some important notations}
2228: \label{tab:common-notation}
2229: \begin{ruledtabular}
2230: \begin{tabular}{c|l}
2231: Symbol(s) & General Meaning \\ \hline
2232: $L$ & total length of the worm-like chain \\
2233: $\kappa$ & bending stiffness \\
2234: $\epsilon$ & small parameter, defined such that $\vec r_\pe'^2=\Ord{\epsilon}$ \\
2235: $\vec r_\pe(s,t)$ & transverse displacement; $\vec r_\pe=\Ord{\epsilon^{1/2}}$ \\
2236: $r_\pa(s,t)$ & longitudinal displacement; $\vec r_\pa=\Ord{\epsilon}$ \\
2237: $\varrho(s,t)$ & stored-length density; $\varrho=\vec
2238: r_\pe'^2/2+\Ord{\epsilon^2}=\Ord{\epsilon}$ \\
2239: $R$ & end-to-end distance \\
2240: $R_\pa$ & end-to-end vector projected onto the long.~axis
2241: \\
2242: $\simeq$ & equal up to numerical factors of
2243: order $1$ \\
2244: $\propto$ & proportional to \\
2245: $\sim$ & asymptotically equal \\
2246: $\lp$ & persistence length \\
2247: $k_B T$ & thermal energy \\
2248: $f(s,t)$ & line tension \\
2249: $\lpe(t)$ & equilibration scale for transverse bending modes
2250: \\
2251: $\lpa(t)$ & scale of tension variations \\
2252: $\vec \fe$ & external force \\
2253: $\vec \xi(s,t)$ & thermal force per arc length
2254: \end{tabular}
2255: \end{ruledtabular}
2256: \end{table}
2257:
2258:
2259: \section{Crossover scaling (details)}
2260: \label{sec:cscaling-det}
2261:
2262:
2263: \subsection{Deterministic relaxation on long times ($t \gg \tf$)}
2264: \label{sec:A}
2265: We have for $\fc=0$
2266: \begin{equation}
2267: \label{eq:A-step1}
2268: \begin{split}
2269: &A-\mint{\frac{dq}{2\pi}}{-\infty}{\infty} \frac 1
2270: {q^2}\left(1-e^{-2q^2 \phi}\right) = \\
2271: &\sqrt{\phi}\mint{\frac{dq}{2\pi}}{-\infty}{\infty} \frac 1
2272: {q^2}\left(e^{-2q^2}-e^{-2q^2(q^2\tau\phi^{-2}+1)}\right) \\
2273: &\stackrel{\tau\gg1}{\to}0 \;,\qquad \text{if }\phi^{-2}=
2274: o(\tau^{-1}) \;,
2275: \end{split}
2276: \end{equation}
2277: and for $\fc=1$
2278: \begin{equation}
2279: \label{eq:A-step2}
2280: \begin{split}
2281: A-\mint{\frac{dq}{2\pi}}{-\infty}{\infty}
2282: \frac 1 {q^2+1}&=-\mint{\frac{dq}{2\pi}}{-\infty}{\infty}
2283: \frac 1 {q^2+1} e^{-2q^2(q^2\tau+\phi)}\\
2284: &\stackrel{\tau\gg1}{\to}0
2285: \;,\qquad \text{if }\phi^{-2}= o(\tau^{-1}) \;.
2286: \end{split}
2287: \end{equation}
2288: As indicated, both expressions, Eqs.~(\ref{eq:A-step1},~\ref{eq:A-step2}), go to zero
2289: for large $\tau$ if $\phi^{-2}= o(\tau^{-1})$. The latter, however, follows from the
2290: assumptions stated in the main text, namely that the tension satisfies the scaling
2291: form Eq.~(\ref{eq:scaling-phi}) with the requirement in Eq.~(\ref{eq:alpha-assumption}).
2292: N.b.~it turns out that $\phi=\Ord{\tau}$ for {\pull} and {\lelo} and
2293: $\phi=\Ord{\tau^{4/3}}$ for \epull.
2294:
2295:
2296:
2297: \subsection{Thermal excitation on long times ($t \gg \tf$)}
2298: \label{sec:B}
2299: We want to show that it is justified to calculate (the negative of) the thermally
2300: generated stored length represented by the term $B$, Eq.~(\ref{eq:B}), on long times
2301: \emph{quasi-statically} for the semi-infinite polymer. To this end, we first
2302: insert the scaling ansatz, Eq.~(\ref{eq:scaling-phi}), for $\phi(\sigma,\tau)$,
2303: \begin{equation}
2304: \label{eq:B-app-1}
2305: \begin{split}
2306: B&=-4 \mint{\frac{dq}{2\pi}}{\Lambda^{-1}}{\infty} q^2 \mint{d\hat
2307: \tau}{0}{\tau}\\
2308: &\times e^{-2 q^2 \left[ q^2 (\tau-\hat \tau)+\tau^{\alpha+1}\left(\hat
2309: \phi(\xi)- (\hat \tau/ \tau)^{\alpha+1}\hat\phi[\xi (\tau/\hat \tau)^{z}
2310: ]\right)\right] } \;,
2311: \end{split}
2312: \end{equation}
2313: where we introduced the scaling variable $\xi\equiv
2314: \sigma/\tau^{z}$. Then we substitute $\hat \tau\to x \tau$ and
2315: $q\to q \tau^{-1/4}$,
2316: \begin{equation}
2317: \label{eq:B-app-2}
2318: \begin{split}
2319: B&=-4 \tau^{1/4} \mint{\frac{dq}{2\pi}}{\Lambda^{-1}\tau^{1/4}}{\infty} q^2
2320: \mint{dx}{0}{1} \\
2321: &\times e^{-2 q^2 \left[ q^2
2322: (1-x)+\tau^{\alpha+1/2}\left(\hat \phi(\xi)- x^{\alpha+1} \hat\phi(\xi
2323: x^{-z})\right)\right] } \;.
2324: \end{split}
2325: \end{equation}
2326: Note that for $\alpha>-1/2$ (as assumed in Eq.~(\ref{eq:alpha-assumption})) the
2327: factor $\tau^{\alpha+1/2}$ in the exponent diverges in the long-time limit. Hence,
2328: for any given wave number the $x$-integral will be dominated by $x$ close to $1$ for
2329: large enough $\tau\gg1$. This allows us to linearize the exponent in $1-x$ when
2330: performing the $x$-integral for this given wave number. In contrast, for a given time
2331: $\tau$ and $\hat \phi(\xi)=\Ord{1}$, the exponent can be linearized only for large
2332: enough wave numbers, $q\gg q_\star(\tau)\equiv \tau^{-\alpha/2-1/4}$, for which the
2333: factor $q^2 \tau^{\alpha+1/2}\gg1$ in the exponent of Eq.~(\ref{eq:B-app-2}) is much
2334: larger than $1$.
2335:
2336: Since the integral over $q$ runs over all $q$-vector, we have also to
2337: care about the small wave numbers, for which the exponent cannot be
2338: linearized in $1-x$. To this end, we split the $q$-integral at a wave
2339: vector $K$ satisfying
2340: \begin{equation}
2341: \label{eq:K}
2342: q_\star^{1/3}\gg K\gg q_\star\equiv\tau^{-\alpha/2-1/4} \;,
2343: \end{equation}
2344: which can be found in the limit $\tau\gg1$ under the premise of
2345: Eq.~(\ref{eq:alpha-assumption}), $\alpha>-1/2$. The first inequality
2346: in Eq.~(\ref{eq:K}) is required for reasons that become clear later
2347: on. For the upper part $B_>$ of the integral we can linearize the
2348: exponent in $1-x$,
2349: \begin{equation}
2350: \label{eq:B-app-upper}
2351: \begin{split}
2352: B_>&\equiv-4 \tau^{1/4} \mint{\frac{dq}{2\pi}}{K}{\infty} q^2
2353: \mint{dx}{0}{1}\\& \times e^{-2 q^2 \left[ q^2
2354: (1-x)+\tau^{\alpha+1/2} \left(\hat \phi(\xi)-x^{\alpha+1} \hat\phi(\xi
2355: x^{-z})\right)\right] }\\
2356: &\sim-4 \tau^{1/4} \mint{\frac{dq}{2\pi}}{K}{\infty} q^2
2357: \mint{dx}{0}{1} \\& \times e^{-2 q^2 (1-x)\left[ q^2
2358: +\tau^{1/2}\partial_\tau \left(\tau^{\alpha+1} \hat
2359: \phi(\sigma/\tau^z)\right)\right] } \\
2360: &=-2 \tau^{1/4} \mint{\frac{dq}{2\pi}}{K}{\infty} \frac{1-e^{-2 q^2 \left[ q^2
2361: +\tau^{1/2}\partial_\tau \left(\tau^{\alpha+1} \hat
2362: \phi(\sigma/\tau^z)\right)\right] }}{q^2
2363: +\tau^{1/2}\partial_\tau \left(\tau^{\alpha+1} \hat
2364: \phi(\sigma/\tau^z)\right)}
2365: \end{split}
2366: \end{equation}
2367: where we eliminated the scaling variable $\xi=\sigma/\tau^z$, again.
2368: Using the second inequality in Eq.~(\ref{eq:K}) it is seen that we can
2369: drop the exponential for the ``interesting'' regime
2370: $\hat\phi(\sigma/\tau^z)=\Ord{1}$, where the tension has an
2371: appreciable value. Inserting back
2372: $\phi=\tau^{\alpha+1}\hat\phi(\sigma,\tau)$ we obtain
2373: \begin{equation}
2374: \label{eq:B-app-upper-2}
2375: \begin{split}
2376: B_> &\sim -2 \tau^{1/4} \mint{\frac{dq}{2\pi}}{K}{\infty}
2377: \frac{1}{q^2
2378: +\tau^{1/2}\partial_\tau \phi(\sigma,\tau)} \\
2379: & = \frac{2}{\sqrt{\partial_\tau \phi(\sigma,\tau)}}
2380: \mint{\frac{dq}{2\pi}}{K/\left(\tau^{1/4}\sqrt{\partial_\tau
2381: \phi(\sigma,\tau)}\right)}{\infty} \frac{1}{q^2 +1 } \\
2382: &\sim - \frac{1}{2\sqrt{\partial_\tau \phi(\sigma,\tau)}}\;.
2383: \end{split}
2384: \end{equation}
2385: To obtain the last asymptotics, we have approximated the lower bound
2386: of the integral by zero. This can be justified by the first inequality
2387: in Eq.~(\ref{eq:K}),
2388: \[\frac{K}{\tau^{1/4}\sqrt{\partial_\tau \phi(\sigma,\tau)}}= K\,
2389: \Ord{\tau^{-\frac 1 4-\frac \alpha 2}=q_\star} \ll q_\star^{4/3}\ll 1 \;.\]
2390:
2391: The remaining lower part $B_<$ of the $q$-integral in
2392: Eq.~(\ref{eq:B-app-2}) is estimated to be small as compared to $B_>$,
2393: \begin{equation}
2394: \label{eq:B-app-lower}
2395: \begin{split}
2396: B_<&\equiv-4 \tau^{1/4} \mint{\frac{dq}{2\pi}}{\Lambda^{-1}t^{1/4}}{K} q^2
2397: \mint{dx}{0}{1} \\& \times e^{-2 q^2 \left[ q^2
2398: (1-x)+\tau^{\alpha+1/2} \left(\hat \phi(\xi)-x^{\alpha+1} \hat\phi(\xi
2399: x^{-z})\right)\right] }\\
2400: &< -4 \tau^{1/4} \mint{\frac{dq}{2\pi}}{0}{K} q^2
2401: \mint{dx}{0}{1} \\
2402: &= -\frac{4}{3} \tau^{1/4} K^3 /(2 \pi)\\
2403: &\ll -\frac{4}{3} \tau^{-\alpha/2} /(2 \pi) \\
2404: &\propto B_> \;,
2405: \end{split}
2406: \end{equation}
2407: where the first inequality in Eq.~(\ref{eq:K}) has been applied. Therefore, $B$ is
2408: asymptotically given by Eq.~(\ref{eq:B-app-upper-2}).
2409:
2410:
2411: \section{Relaxation of a completely stretched polymer}
2412: \label{sec:lelo-app}
2413:
2414: In this section, we consider more closely the intermediate asymptotics
2415: \begin{equation}
2416: \label{eq:homogeneous-lelo}
2417: f\propto \left(\frac{\hat \zeta L^2}{l_p t}\right)^{2/3} \;,
2418: \end{equation}
2419: that is approached in the limit
2420: \begin{equation}
2421: \label{eq:homo-lelo-time-frame}
2422: \totpa=\frac{\hat \zeta L^2}{l_p {f_<}^{3/2}}\ll t \ll
2423: t_\star=\frac{\hat \zeta^4 L^8}{l_p^4}\;
2424: \end{equation}
2425: of {\lelo}, which we analyzed in terms of a quasi-static approximation in
2426: Sec.~\ref{sec:tension-relaxed}. The purpose of this section is to justify the
2427: quasi-static assumption in the limit of $f_<\to\infty$ (i.e.~$\tf\to0$) where we
2428: start with a completely stretched polymer and all stored length is generated by the
2429: action of stochastic forces.
2430:
2431: To this end, we show that, in the limit $t\to 0$, the change in stored length
2432: $\avg{\Delta\overline \varrho}$ given by Eq.~(\ref{eq:change-stored-length}) for the
2433: force history given by Eq.~(\ref{eq:homogeneous-lelo}) asymptotically approaches the
2434: value one obtains from the quasi-static calculation,
2435: \begin{eqnarray}
2436: &\avg{\Delta\overline
2437: \varrho}(t)=\mint{\frac{dq}{\pi\lp}}{0}{\infty}
2438: \left\{\frac{1}{q^2+f_<}\left(e^{-2q^2[q^2 t+ F(t)]}-1\right)
2439: \right.& \nonumber\\
2440: &\left.+2 q^2\mint{d\tilde t}{0}{t}e^{-2q^2\left[q^2(t-\tilde
2441: t)+F(t)-F(\tilde t)\right]}\right\}&\nonumber\\
2442: &\stackrel{f_<\to\infty}{=}\mint{\frac{dq}{\pi\lp}}{0}{\infty} 2
2443: q^2\mint{d\tilde t}{0}{t}e^{-2q^2\left[q^2(t-\tilde
2444: t)+F(t)-F(\tilde t)\right]} & \label{eq:to-show-1} \\
2445: &\stackrel{!}{\sim} \mint{\frac{dq}{\pi\lp}}{0}{\infty}
2446: \frac{1}{q^2+f(t)}=\frac{1}{2l_p}\left[f(t)\right]^{-1/2}\;,\qquad\text{for } t
2447: \to 0\;.&\nonumber
2448: \end{eqnarray}
2449: This will comprise an a posteriori justification of the quasi-static
2450: assumption that entered in the derivation of the right-hand-side of
2451: Eq.~(\ref{eq:degennes-eq-2}).
2452:
2453: The argument closely follows App.~\ref{sec:B}. Inserting the force history
2454: \begin{equation}
2455: \label{eq:F-1/3}
2456: F(t)=\mint{d\hat t}{0}{t}f(\hat t)= C t^{\alpha+1}
2457: \end{equation}
2458: with
2459: \begin{equation}
2460: \label{eq:prefactor}
2461: \alpha=-2/3\stackrel{!}{<}-1/2 \qquad \text{and}\qquad C\approx \left(\frac{\hat \zeta L^2}{l_p}\right)^{2/3}
2462: \end{equation}
2463: into Eq.~(\ref{eq:to-show-1}) and changing variables $\hat t \to x t$
2464: and $q\to q t^{-1/4}$ yields
2465: \begin{eqnarray}
2466: \label{eq:1/3-noise-rescaled}
2467: &&\avg{\Delta\overline
2468: \varrho}(t)=2t^{1/4}\mint{\frac{dq}{\pi\lp}}{\Lambda^{-1}t^{1/4}}{\infty}
2469: q^2 \nonumber \\
2470: &&\qquad\times \mint{dx}{0}{1}e^{-2q^2\left[q^2(1-x)+C t^{\alpha+1/2}\left(1-x^{\alpha+1}\right)\right]}
2471: \end{eqnarray}
2472: As in Sec.~\ref{sec:B} an approximation to the integral can be found in
2473: the limit
2474: \begin{equation}
2475: \label{eq:central-limit}
2476: C t^{\alpha+1/2}\gg1 \;.
2477: \end{equation}
2478: We split the $q$-integral at $K$ satisfying
2479: \begin{equation}
2480: \label{eq:K2}
2481: q_\star^{1/3}\gg K\gg q_\star\equiv (C t^{\alpha+1/2})^{-1/2} \;,
2482: \end{equation}
2483: which can be found in the limit $t\ll 1$ ($\Rightarrow
2484: q_\star\ll1$) because $\alpha<-1/2$. The upper part of
2485: the integral
2486: \begin{equation}
2487: \label{eq:rho>}
2488: \avg{\Delta\overline
2489: \varrho^>}(t)\equiv \left(\dots\right)\mint{dq}{K}{\infty}\left(\dots\right)
2490: \end{equation}
2491: is dominated by values of $x$ close to $1$ and we can linearize the
2492: exponent in $1-x$,
2493: \begin{equation}
2494: \label{eq:rho>-2}
2495: \begin{split}
2496: \avg{\Delta\overline{
2497: \varrho}^>}(t)&=2t^{1/4}\mint{\frac{dq}{\pi}}{K}{\infty}q^2\mint{dx}{0}{1}
2498: \\&\times e^{-2q^2 (1-x) \left[q^2+(\alpha+1)C t^{\alpha+1/2} \right]} \\
2499: &= \tau^{1/4}\mint{\frac{dq}{\pi}}{K}{\infty} \frac{1-e^{-2q^2 (1-x)
2500: \left[q^2+(\alpha+1)C t^{\alpha+1/2}
2501: \right]}} {q^2+(\alpha+1)C t^{\alpha+1/2}} \\
2502: &\sim \tau^{1/4}\mint{\frac{dq}{\pi}}{K}{\infty}
2503: \frac{1}{q^2+(\alpha+1)C t^{\alpha+1/2}} \\
2504: &=\frac{1}{l_p\sqrt{(\alpha+1)C t^\alpha}}\mint{\frac{dq}{1
2505: \pi}}{K/\sqrt{(\alpha+1)C t^{\alpha+1/2}}}{\infty}\frac{1}{q^2+1} \\
2506: &\sim \frac{1}{2 l_p \sqrt{f(t)}}
2507: \end{split}
2508: \end{equation}
2509: where the asymptotics follow from both inequalities in
2510: Eqs.~(\ref{eq:central-limit},~\ref{eq:K2}). The lower part $\avg{\Delta\overline{
2511: \varrho}^<}$ of the integral is estimated to be subdominant as compared to
2512: $\avg{\Delta\overline{ \varrho}^>}$,
2513: \begin{equation}
2514: \label{eq:rho<}
2515: \begin{split}
2516: \avg{\Delta\overline{
2517: \varrho}^<}(t)&=(\dots)\mint{dq}{0}{K}(\dots) \\
2518: &<\frac{2 t^{1/4}}{l_p}\mint{\frac{dq}{\pi}}{0}{K} q^2
2519: \mint{dx}{0}{1} \\
2520: &=\frac{t^{1/4}K^3}{3\pi l_p}\\
2521: &\ll\frac{C^{-1/2}t^{-\alpha/2}}{3\pi l_p} \\
2522: &\propto \avg{\Delta\overline{
2523: \varrho}^>}(t) \;,
2524: \end{split}
2525: \end{equation}
2526: where the first inequality in Eq.~(\ref{eq:K2}) has been applied. Therefore
2527: $\avg{\Delta\overline{ \varrho}}(t)$ is asymptotically given by
2528: Eq.~(\ref{eq:rho>-2}).
2529:
2530: Finally, we want to emphasize the central condition for the validity
2531: of the quasi-static approximation,
2532: \begin{equation}
2533: \label{eq:central-limit-2}
2534: C t^{\alpha+1/2}\approx\left(\frac{\hat \zeta L^2}{l_p}\right)^{2/3}t^{-1/6}\gg 1\;,
2535: \end{equation}
2536: or $t\ll t_\star$ with $t_\star=(L^2/\lp)^4$ as given by
2537: Eq.~(\ref{eq:val-cond-lelo-2}).
2538:
2539:
2540: \section{Defining $t_\star$}
2541: \label{sec:tstar}
2542:
2543: As anticipated Part~I, there is a problem-specific time $t_\star$ limiting the
2544: short-time validity of OPT. Physically, the crossover at $t_\star$ can be understood
2545: as follows. For $t\gg t_\star$ the ``speed'' of the structural relaxation is
2546: determined solely by the relaxation times of the bending modes, which are related to
2547: the transverse friction while the longitudinal friction is irrelevant. In contrast,
2548: for $t\ll t_\star$ the longitudinal friction substantially limits the speed of the
2549: relaxation. This suggests to estimate the time $t_\star$ as follows. From the
2550: continuity equation, Eq.~(\ref{eq:cg-eom}), derived via the multiple-scale
2551: perturbation theory (MSPT), we can estimate the order of magnitude of the correction
2552: $\delta f(s,t)=f(s,t)-f^{\text{OPT}}$ to the flat tension profile,
2553: Eq.~(\ref{eq:OPT-tension}), by
2554: \begin{equation}
2555: \label{eq:f-correction}
2556: \delta f \approx \hat \zeta \partial_t \avg{\Delta \overline{\varrho}}\left(
2557: f^{\text{OPT}},t \right) L^2 \;.
2558: \end{equation}
2559: OPT can only be applicable if the correction $\delta f$ has negligible
2560: effect on the evolution of the stored length,
2561: \begin{equation}
2562: \label{eq:OPT-validity}
2563: \abs{1-\frac{\avg{\Delta\overline{\varrho}}\left(
2564: f^{\text{OPT}}+\delta f,t \right)}{\avg{\Delta\overline{\varrho}}\left(
2565: f^{\text{OPT}},t \right)}}\ll 1 \;,\qquad \text{for } t\gg t_\star\;.
2566: \end{equation}
2567: The time for which the left hand side of Eq.~(\ref{eq:OPT-validity}) becomes of order
2568: may thus be identified with the time $t_\star$ before which OPT is not valid.
2569:
2570:
2571:
2572:
2573:
2574:
2575:
2576: \bibliographystyle{apsrev}
2577:
2578: \bibliography{bibis/tdpre06,bibis/elastic-rod-dynamics,bibis/journals,bibis/sfpdynamics-tcited,bibis/unpub,bibis/actin-viscoelastic,bibis/semiflexibleA04,bibis/sf,bibis/klaussf,bibis/mysf,bibis/sfnet,bibis/unpub,bibis/mysfnet}
2579:
2580: \end{document}