1: %\documentstyle[aps,epsf]{revtex}
2: \documentstyle[aps,epsf,multicol]{revtex}
3:
4: \def\be{\begin{equation}}
5: \def\ee{\end{equation}}
6: \def\bea{\begin{eqnarray}}
7: \def\eea{\end{eqnarray}}
8:
9: \parskip=5pt
10: %\renewcommand{\baselinestretch}{1.7}
11: \newcommand{\e}{{\rm e}}
12: \newcommand{\half}{\mbox{$\frac{1}{2}$}}
13: \newcommand{\tmu}{{\tilde \mu}}
14: \newcommand{\Bt}{{\tilde B}}
15: \newcommand{\Bhat}{{\hat B}}
16: \newcommand{\lb}{\label}
17: \renewcommand{\d}{{\rm d}}
18:
19: \begin{document}
20:
21: \draft \title{Embedding a Native State into a Random \\
22: Heteropolymer Model: The Dynamic Approach}
23:
24: \author{Z. Konkoli$^{1,2}$ and J. Hertz$^{2}$}
25: \address{
26: $^1${Department of Applied Physics, \\
27: Chalmers University of Technology and G\"oteborg University, \\
28: SE 412 96 G\"oteborg, Sweden} \\
29: $^2${NORDITA, Blegdamsvej 17, DK 2100 K\o benhavn, Denmark}
30: }
31: \date{\today}
32: \maketitle
33:
34: \begin{abstract}
35: We study a random heteropolymer model with Langevin dynamics, in the
36: supersymmetric formulation. Employing a procedure similar to one that
37: has been used in static calculations, we construct an ensemble in
38: which the affinity of the system for a native state is controlled by a
39: ``selection temperature" $T_0$. In the limit of high $T_0$, the model
40: reduces to a random heteropolymer, while for $T_0 \rightarrow 0$ the
41: system is forced into the native state. Within the Gaussian
42: variational approach that we employed previously for the random
43: heteropolymer, we explore the phases of the system for high and low
44: $T_0$. For high $T_0$, the system exhibits a (dynamical) spin glass
45: phase, like that found for the random heteropolymer, below a
46: temperature $T_g$. For low $T_0$, we find an ordered phase,
47: characterized by a nonzero overlap with the native state, below a
48: temperature $T_n \propto 1/T_0 > T_g$. However, the random-globule
49: phase remains locally stable below $T_n$, down to the dynamical glass
50: transition at $T_g$. Thus, in this model, folding is rapid for
51: temperatures between $T_g$ and $T_n$, but below $T_g$ the system can
52: get trapped in conformations uncorrelated with the native state. At a
53: lower temperature, the ordered phase can also undergo a dynamical
54: glass transition, splitting into substates separated by large
55: barriers.
56: \end{abstract}
57:
58: \pacs{05.70.Ln, 87.14.Ee}
59:
60: \maketitle
61:
62: \begin{multicols}{2}
63: \narrowtext
64:
65: \section{Introduction}
66:
67: The protein folding process is relevant for all aspects of life:
68: once read off from the RNA chain, proteins perform a variety of
69: functions, from mechanical work to attacking viruses.~\cite{Boy}
70: The key factor which determines the function of a protein molecule
71: is its 3D structure, which, in turn, is determined by the sequence
72: of amino acids forming the protein
73: chain.~\cite{PGT1,Wol1,WE,Creig} Furthermore, a protein that has
74: been denatured (by stretching it for example) finds its native
75: state relatively quickly. Protein folding has attracted an
76: enormous amount of scientific attention, but still there is no
77: generic understanding of this process. Nevertheless, one thing is
78: clear: a proteins generally has a potential energy surface which
79: results in a stable free energy minimum, corresponding to the
80: native state~\cite{Wol1}.
81:
82: Random heteropolymer models (RHP) have been used extensively as
83: candidate systems which might help us to understand the generic
84: features of the potential energy surfaces of proteins and their
85: connection with thermodynamic
86: ~\cite{SG1,SG2,GHLO,GOP,GLO,TW,SGS,SW} and dynamical
87: ~\cite{LT,Pit,TPW,TAB,PS,Olem1,Olem2} properties. The RHP model
88: is characterized by quenched random monomer-monomer interactions,
89: meant to mimic the variety of interactions between amino-acids in
90: random sequences. It turns out that the potential energy surface
91: of the RHP is quite similar to that of a particular class of spin
92: glasses~\cite{SpGl}: Its complex form, with exponentially large
93: numbers of local minima and saddle points, constrains the motion
94: of the system drastically, and it cannot explore its full
95: configuration space and reach Gibbs equilibrium. In a previous
96: paper (\cite{KHS}, henceforth referred to as paper I), we
97: demonstrated, in mean field theory, the existence of a sharp
98: transition to a ``dynamical glassy state'' in which the
99: equilibration time diverges and the dynamics exhibit aging. (The
100: potential importance of spin glass physics to proteins was first
101: discussed in Ref.~\cite{Wol2}). Obviously, the random
102: heteropolymer model does not describe a protein with a native
103: state, but it alerts us to the need to examine possible glassiness
104: in models for protein dynamics.
105:
106: Why are real proteins not glassy? Evidently, nature has tuned amino
107: acid sequences to avoid glassy behavior. To understand how such
108: tuning might be done, it is worthwhile to study models which contain
109: competition between glassiness and a tendency to form a native state,
110: by choosing interactions which are not completely random. Several
111: studies along the lines of this suggestion have been made in {\em
112: statics} (using the replica treatment, see, e.g., Ref.~\cite{Wol2}).
113: The tendency toward a particular state can be built in by choosing
114: sequences from a distribution correlated with the native sequence
115: \cite{PGT1,RS,PGT2,WS}. A dynamical treatment of similar models is
116: highly desirable, not only to help gain insight into results obtained
117: in replica approaches, but also because knowledge of the correct
118: thermodynamics alone may not be sufficient: it is known that in
119: related (mean field models) static and dynamic phase diagrams can be
120: different. Thus (at least on sufficiently short time scales) only a
121: dynamical approach can describe the measurable properties of the
122: system. In this paper we undertake such a study.
123:
124: We extend the RHP model studied in \cite{SG1,SG2} to include the
125: existence of a native state: the original random monomer-monomer
126: interactions are biased so as to favor the native state
127: conformation. The problem is formulated as a Langevin model. To
128: the best of our knowledge, there is so far neither a static nor a
129: dynamic treatment available for a model of this sort: Static
130: studies have been based on random monomer sequences, i.e., using
131: only $N$ random parameters, see Refs.~\cite{PGT1,RS,PGT2,WS},
132: rather than the $N(N-1)/2$ in the RHP model.
133:
134: Admittedly, the model does not describe a realistic protein (e.g.,
135: it does not give rise to secondary structure such as
136: $\alpha$-helices or $\beta$-sheets). However, it does contain
137: important generic features: the polymeric structure and the
138: mixture of attractive and repulsive interactions. Together, these
139: features lead to frustration in the structural dynamics. In our
140: view, ours is the simplest such model that includes competition
141: between glassy and native states. As we will see, it teaches us
142: that one can not get rid of glassiness so easily.
143:
144: As in paper I, we simplify the model further by omitting
145: three-body interactions in the polymer. (A review describing how
146: to include three-body terms is given in~\cite{GOP}.) The price we
147: have to pay for this simplification is that we have to introduce a
148: somewhat arbitrary confining potential, which we take to have a
149: quadratic form. We adjust its strength so that the radius of
150: gyration $R_g$ of a polymer of size $N$ scales like $N^{-1/d}$,
151: where $d$ is the dimensionality of the system. In this way we
152: attempt to describe a globular state. Of course, we can not
153: describe the $\theta$-point transition in such a model, but here
154: we are only interested in transitions between different globular
155: states.
156:
157: Our formal starting point is the Martin-Siggia-Rose generating
158: functional for the Langevin dynamics of the model
159: \cite{MSR,Dom,Jans1,Jans2}, written, for convenience and
160: compactness, in its supersymmetric form \cite{Kur}. To derive
161: equations of motion for correlation and response functions we use
162: a variational ansatz with a quadratic action. This approach has
163: been used to study the problem of a manifold in a random
164: potential, in both statics \cite{MP1,MP2} and dynamics
165: \cite{CKD,CD}.
166:
167: In paper I we showed that the RHP model exhibited broken
168: ergodicity (formally, a spontaneous supersymmetry breaking) in a
169: low-temperature dynamical glassy phase. In the present study, with
170: interactions biased in favor of a native state to a controlled
171: degree, we find, in addition, a well-folded phase, if the bias is
172: strong enough. It can coexist with either the disordered
173: (random-globule) state or the frozen-globule glass phase,
174: depending on the temperature. Furthermore, we find that at low
175: temperature the native phase can itself undergo a dynamical
176: freezing into a different glassy phase. In this phase the
177: conformation of the protein is always highly correlated with the
178: native state, but cooperative kinetic constraints still lead to a
179: divergent equilibration time, as for the frozen-globule state.
180:
181:
182: \section{The Model}
183:
184: The model is defined as follows. The Langevin dynamics is assumed to
185: be governed by a Hamiltonian $H[x]$,
186: \begin{equation}
187: \partial x(s,t)/\partial t = - \delta H[x] / \delta x(s,t) + \eta(s,t).
188: \label{eq:dxdt}
189: \end{equation}
190: Here $x(s,t)$ is the position of monomer $s$ at time $t$. The monomers
191: are numbered continuously from $s=0$ to $s=N$. $\eta(s,t)$ is Gaussian
192: noise
193: \begin{equation}
194: \langle \eta(s,t)\eta(s',t') \rangle_T = 2T\delta(s-s')\delta(t-t'),
195: \label{eq:etas}
196: \end{equation}
197: resulting from coupling to a heat bath at temperature $T$.
198:
199: The Hamiltonian $H[x]$ contains a deterministic part $H_0[x,\mu]$
200: and a random part $H[x,\{B\}]$. $H_0[x,\mu]$ is defined as
201: \begin{equation}
202: H_0[x,\mu]= \frac{T}{2} \int_{0}^{N} ds
203: \{[\partial x(s,t)/\partial s]^2+\mu x(s,t)^2\}.
204: \label{eq:H0}
205: \end{equation}
206: It describes the elastic properties of the chain and a confinement
207: potential which fixes the density of the protein. The radius of
208: gyration $R_g \sim \mu^{-1/4}$, so, in order that the protein is
209: compact, i.e., $R_g \sim N^{1/d}$, we require $\mu \sim
210: N^{-4/d}$. Thus, since we are interested in very long proteins (to
211: obtain the thermodynamic limit) we need to solve the model for
212: $\mu$ close to zero.
213:
214: The random part $H[x,\{B\}]$ describes the quenched random
215: interactions between monomers,
216: \begin{equation}
217: H[x,\{B\}]= \frac{1}{2} \int_{0}^{N} ds ds'
218: B_{ss'} V(x(s,t)-x(s',t)).
219: \label{eq:Hrand}
220: \end{equation}
221: We take $B_{ss'}$ Gaussian, with variance $B^2$. The
222: quenched average over $B_{ss'}$ is performed as $\langle (.) \rangle_B
223: = \int \prod_{s>s'} dB_{ss'} (.) P(\{B\})$.
224: $V(\Delta x)$ is a short-range potential, and, for simplicity, we
225: take it to have a Gaussian form, as in Ref.~\cite{TPW},
226: \begin{equation}
227: V(\Delta x)=\left(\frac{1}{2\pi\sigma}\right)^{d/2}
228: \e^{-(\Delta x)^2/2\sigma}.
229: \label{eq:V}
230: \end{equation}
231: $d$ is the dimensionality of the system and $\sqrt{\sigma}$ the range of
232: the potential. Large (small) $\sigma$ corresponds to a long (short)
233: range potential. In particular, for $\sigma\rightarrow 0$, $V(\Delta
234: x)\rightarrow\delta(\Delta x)$, and we recover the potential used in
235: \cite{SG1,SG2,PS}. Here and in the following $\Delta x$ refers to a
236: monomer-monomer distance: $\Delta x=x(s,t)-x(s',t)$ for a pair of monomers
237: $s$, $s'$.
238:
239: We use reasoning similar to that employed in statics to define $P(\{B\})$
240: (see Refs.~\cite{PGT1,RS,PGT2,WS}), adapting it to the random-bond model:
241: \begin{equation}
242: P(\{B\}) \propto \e^{ - \frac{1}{T_0} H[x_0,{B}]
243: - \frac{1}{2}\int ds ds' B_{ss'}^2 / 2 B^2 }
244: \label{eq:PB1}
245: \end{equation}
246: $T_0$ is called the selection temperature, and $x_0(s)$ is some arbitrary
247: native state conformation. Thus the symmetric bond distribution of the
248: RHP model is distorted so as to give bigger weight to $B_{ss'}$'s which
249: are attractive between monomers which lie close to each other in the
250: configuration $x_0(s)$. Explicitly, the properly normalized
251: $P(\{B\})$ is given by
252: \begin{eqnarray}
253: && P(\{B\}) = (2\pi B^2)^{-N(N-1)/4}
254: \e^{-\beta_0^2 B^2 /4 \int ds ds' V(x_0(s)-x_0(s'))^2}
255: \nonumber \\
256: && \times \e^{-\beta_0/2\int dsds'B_{ss'} V(x_0(s)-x_0(s'))
257: -1/2\int dsds'B_{ss'}^2/2B^2},
258: \label{eq:PB2}
259: \end{eqnarray}
260: from which we see that the distribution of $B_{ss'}$ is peaked around
261: $B^{max}_{ss'} = - \beta_0 B^2 V(x_0(s)-x_0(s'))$. Thus, if monomers
262: $s$ and $s'$ are close in the native state ($V(x_0(s)-x_0(s'))\ne 0$),
263: their coupling constant $B_{ss'}$ is pulled down, as in a Go model
264: \cite{Go1,Go2}. For $T_0\rightarrow\infty$ we recover the RHP model.
265: For $T_0\rightarrow 0$, $P(\{B\})$ picks a specific set of
266: $B_{ss'}$. For this set, by construction, $x_0(s)$ is the deepest
267: minimum of $H[x,\{B\}]$ given in Eq.~(\ref{eq:Hrand}). This is the
268: mechanism that embeds the native state $x_0(s)$.
269:
270: This mechanism is somewhat arbitrary. However, the fact that the
271: strength of embedding of the native state is controlled by the single
272: parameter $T_0$ facilitates the study of transitions between random
273: and native-like states (and, as we will show, of possible coexistence
274: of such phases).
275:
276: So far, the configuration $x_0(s)$ is arbitrary. Thus $x_0(s)$ has to
277: be considered a quenched random function, to be averaged over just
278: like $B_{ss'}$ in order to obtain generic results. We will carry this
279: average out later.
280:
281: All our results are obtained in the thermodynamic limit, where the
282: length $N$ of the heteropolymer chain goes to infinity. Also, for
283: simplicity, we join the polymer ends to form a ring. This neglect
284: of end effects is valid for a long chain.
285:
286:
287:
288: \section{Mapping to the Field Theory}
289:
290: To solve the model we map the Langevin dynamics onto a
291: supersymmetric (SUSY) field theory. Using the standard
292: Martin-Siggia-Rose formalism \cite{MSR,Dom,Jans1,Jans2} and
293: supersymmetric (SUSY) notation \cite{Olem1,Olem2,Kur,Olem3}, the
294: dynamical average of any observable, for fixed $\{B\}$, can be
295: calculated as (see, e.g., Paper I for details),
296: %
297: \begin{eqnarray}
298: && \langle {\cal O}[\Phi] \rangle_T =\int D\Phi
299: {\cal O}[\Phi] e^{-S[\Phi] }, \label{eq:avSUSY} \\
300: && S[\Phi] = S_1[\Phi]+S[\Phi, x_0, \{B\}], \label{eq:SSUSY}
301: \end{eqnarray}
302: %
303: where
304: %
305: \begin{eqnarray}
306: & & S_1[\Phi] = 1/2 \int ds d1 ds' d2 \Phi(s,1) K_{12}^{ss'} \Phi(s'2),
307: \label{eq:S0} \\
308: & & S[\Phi,x_0,\{B\}] = 1/2 \int d1 ds ds' \times \nonumber \\
309: & & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
310: \ \ \ \ \ \ \times B_{s,s'} V(\Phi(s,1)-\Phi(s',1)),
311: \label{eq:Srand}
312: \end{eqnarray}
313: %
314: and
315: %
316: \begin{eqnarray}
317: K_{12}^{ss'} && \equiv \delta_{12} \delta_{ss'} K_1^s \ , \ \
318: K_1^s = T \left[ \mu-(\partial/\partial s)^2 \right] - D_1^{(2)}, \\
319: D_1^{(2)} && =2 T \frac{\partial^2}{\partial\theta_1\partial\bar\theta_1}
320: + 2 \theta_1 \frac{\partial^2}{\partial\theta_1\partial t_1} -
321: \frac{\partial}{\partial t_1}.
322: \end{eqnarray}
323: %
324: The $\Phi(s,1)$ denotes a superfield
325: %
326: \begin{eqnarray}
327: & \Phi(s,1) = & x(s,t_1) + \bar\theta_1 \eta(s,t_1) + \nonumber \\
328: & & + \bar\eta(s,t_1) \theta_1
329: + \bar\theta_1\theta_1\tilde x(s,t_1)
330: \label{Phis1}
331: \end{eqnarray}
332: %
333: containing the physical coordinate $x(s,t)$, the MSR auxiliary field
334: $\tilde x(s,t)$, ghost fields $\eta(s,t)$ and $\bar\eta(s,t)$ that
335: enforce the normalization of the distribution, and Grassmann variables
336: $\theta$ and $\bar\theta$. We use the notation $1\equiv
337: (\theta_1,\bar\theta_1,t_1)$, likewise $\int d1 \equiv \int d\bar\theta_1
338: d\theta_1 dt_1$.
339:
340: Of course, the solution can be obtained without the aid of the
341: supersymmetric formalism, but we find it conveniently compact.
342:
343:
344: As noticed by De Dominicis \cite{Dom} the expression in
345: Eq.(\ref{eq:avSUSY}) is already normalized, so the average over the
346: quenched disorder $B_{s,s'}$ can be done directly on
347: (\ref{eq:avSUSY}):
348: %
349: \begin{equation}
350: \langle\langle {\cal O}[\Phi] \rangle_T\rangle_B = \int D\Phi
351: {\cal O}[\Phi] \e^{-(S_1[\Phi]+S_2[\Phi, x_0])}
352: \label{avO},
353: \end{equation}
354: %
355: where $\exp(-S_2[\Phi,
356: x_0])\equiv\langle\exp(-S[\Phi,x_0,\{B\}])\rangle_B$, and
357: %
358: \begin{eqnarray}
359: & S_2[\Phi, x_0] = & -\frac{B^2}{4} \int ds ds'
360: \left[
361: \int d1 V(\Phi(s,1)-\Phi(s',1))
362: \right]^2 \nonumber \\
363: && - \frac{\beta_0 B^2}{2} \int ds ds' d1 V(\Phi(s,1)-\Phi(s',1)) \times
364: \nonumber \\
365: && \times V(x_0(s)-x_0(s')). \label{eq:S2}
366: \end{eqnarray}
367: %
368: Thus, the native state $x_0(s)$ enters the action in the second term
369: of Eq.~(\ref{eq:S2}). Note that there is no term
370: $\beta_0^2V(x_0(s)-x_0(s'))^2$, since it gets cancelled by a similar
371: normalization factor for P(\{B\}) in Eq.~(\ref{eq:PB2}). It is useful
372: to rewrite Eq.~(\ref{eq:S2}) as
373: %
374: \begin{eqnarray}
375: & S_2 = & -\frac{B^2}{4} \int d^dx\,d^dy\,d1\,d2\,
376: A_{12}^{(V)}(x,y) A_{12}^{(\delta)}(x,y) \nonumber \\
377: & & -\frac{\beta_0B^2}{2} \int d^dx\,d^dy\,d1\,
378: A_{10}^{(V)}(x,y) A_{10}^{(\delta)}(x,y)
379: \end{eqnarray}
380: %
381: with the notation $A_{12}^{(f)}(x,y) = \int ds f(\Phi(s,1)-x)
382: f(\Phi(s,2)-y)$, $A_{10}^{(f)}(x,y) = \int ds f(\Phi(s,1)-x)
383: f(x_0(s)-y)$; $f\in\{V,\delta\}$. In the long-chain limit, as
384: discussed in Paper I (and references therein), one obtains a
385: self-consistent field theoretic formulation, with $S_2$ simplified to,
386: %
387: \begin{eqnarray}
388: && S_2[\Phi,x_0] = \frac{B^2}{4} \int d^dx d^dy d1 d2
389: \left[
390: \langle A_{12}^{(V)}(x,y) \rangle
391: \langle A_{12}^{(\delta)}(x,y) \rangle -
392: \right.
393: \nonumber \\
394: && \left.
395: - A_{12}^{(V)}(x,y) \langle A_{12}^{(\delta)}(x,y) \rangle
396: - \langle A_{12}^{(V)}(x,y) \rangle A_{12}^{(\delta)}(x,y)
397: \right] \nonumber \\
398: && + \frac{\beta_0B^2}{2} \int d^dx d^dy d1
399: \left[
400: \langle A_{10}^{(V)}(x,y) \rangle
401: \langle A_{10}^{(\delta)}(x,y) \rangle
402: \right.
403: \nonumber \\
404: && \left.
405: - A_{10}^{(V)}(x,y) \langle A_{10}^{(\delta)}(x,y) \rangle
406: - \langle A_{10}^{(V)}(x,y) \rangle A_{10}^{(\delta)}(x,y)
407: \right].
408: \label{eq:S'}
409: \end{eqnarray}
410: %
411: All averages of the type $\langle A^{(V,\delta)} \rangle$ have to be
412: calculated self-consistently with $S[\Phi]=S_1[\Phi]+S_2[\Phi]$. (We
413: have abbreviated the double average $\langle \langle.\rangle_T
414: \rangle_B$ simply by $\langle.\rangle$.) In the limit
415: $N\rightarrow\infty$ Eqs.~(\ref{avO}) and (\ref{eq:S'}) provide an
416: exact description of the dynamics for an arbitrary native state
417: $x_0(s)$.
418:
419:
420:
421: \section{Average over native state conformations}
422:
423:
424: It is impossible to solve the model for a general native state
425: configuration $x_0(s)$. We therefore consider a distribution of
426: native states and perform the average
427: \begin{equation}
428: \overline{<O[\Phi,x_0]>}=\int Dx_0 <O[\Phi,x_0]> \e^{-S_0[x_0]},
429: \end{equation}
430: %
431: where $S_0[x_0]$ weights each native state conformation in the
432: ensemble as
433: %
434: \begin{equation}
435: S_0[x_0]=1/2\int ds x_0(s) K^{ss'}_{00} x_0(s'),
436: \end{equation}
437: %
438: with
439: %
440: \begin{equation}
441: K^{ss'}_{00} \equiv \delta_{ss'} ( \mu_0-\partial^2/\partial s'^2).
442: \end{equation}
443: %
444: The parameter $\mu_0$ fixes a size of the globule in this
445: ensemble,
446: %
447: \begin{equation}
448: \langle x_0(s)^2 \rangle = \frac{1}{2\sqrt{\mu_0}}
449: \label{x0^2}
450: \end{equation}
451: %
452: Since the polymer ends are joined, there is translational
453: invariance along the coordinate $s$ and $\langle x_0(s)^2\rangle$
454: does not depend on $s$. Thus, with this procedure, the dynamical
455: generating functional for the problem is calculated as
456: %
457: \begin{equation}
458: e^{-F_{dyn}}=\int Dx_0 D\Phi e^{ -(S_0[x_0]+S_1[\Phi]+S_2[\Phi,x_0])
459: }.
460: \label{Fdyn}
461: \end{equation}
462: %
463: There is some formal similarity between the dynamical functional
464: $F_{dyn}$ and the static replica partition function. The integration
465: over $Dx_0$ enters in the same way as the extra replica in the static
466: formalism.
467:
468:
469: \section{Correlation functions}
470:
471: The SUSY correlation functions
472: %
473: \begin{eqnarray}
474: & & G_{12}^{ss'} \equiv \langle \Phi(s,1) \Phi(s',2) \rangle \label{G12} \\
475: & & G_{10}^{ss'} \equiv \langle \Phi(s,1) x_0(s') \rangle \label{G10} \\
476: & & G_{00}^{ss'} \equiv \langle x_0(s)x_0(s') \rangle \label{G00}
477: \end{eqnarray}
478: %
479: contain all the information we are interested in.
480:
481: $G_{12}^{ss'}$ encodes 16 correlation functions, out of which only
482: two, correlation and response function, are independent and nonzero:
483: %
484: \begin{eqnarray}
485: & G_{12}^{ss'} = & C(s,t_1;s',t_2)
486: + (\bar\theta_2-\bar\theta_1) \times \nonumber \\
487: & & \times [ \theta_2 R(s,t_1;s',t_2) - \theta_1 R(s',t_2;s,t_1) ],
488: \end{eqnarray}
489: %
490: with
491: %
492: \begin{eqnarray}
493: & & C(s,t;s',t') \equiv \langle x(s,t)x(s',t') \rangle, \\
494: & & R(s,t;s',t') \equiv \langle x(s,t) \tilde x(s',t') \rangle
495: = \frac{\delta\langle x(s,t)\rangle}{\delta h(s',t')}.
496: \end{eqnarray}
497: %
498: The field $h(s',t')$ entering the description of response function is
499: an arbitrary external field that couples to $x(s',t')$. The fact that
500: only two correlation functions survive is related to Ward identities
501: originating from SUSY invariance of the original action $S$.
502:
503: The supersymmetry of the theory is associated with equilibrium.
504: One of the Ward identities resulting from SUSY is the
505: fluctuation-dissipation theorem (FDT) which relates correlation
506: and response functions. In the present case, the glassy state
507: manifests itself as a spontaneous breaking of supersymmetry,
508: leading to a modified FDT, as in previous treatments of other
509: models \cite{Kur,CKD}.
510:
511: $G_{10}^{ss'}$ describes the overlap with the native state. Due
512: to Ward identities, only a single correlation function survives
513: (see Appendix A for details):
514: %
515: \begin{equation}
516: G_{10}^{ss'} = \langle x(s,t)x_0(s') \rangle \equiv \phi(s,t_1;s').
517: \label{G10a}
518: \end{equation}
519: %
520: Similarly, the native state ensemble is described by
521: %
522: \begin{equation}
523: G_{00}^{ss'} = \langle x_0(s)x_0(s') \rangle \equiv \Gamma(s;s').
524: \end{equation}
525: %
526: $G_{12}^{ss'}$ alone is sufficient to describe the RHP model.
527: Here we need the two extra functions $G_{10}^{ss'}$ and $G_{00}^{ss'}$.
528:
529: Also, in what follows, we exploit the translational invariance
530: along the $s$ coordinate and define Fourier transforms of all
531: correlation functions: $X(s,s')= \int \frac{dk}{2\pi} e^{ik(s-s')}
532: X_k$ where $X=C,R,\phi,\Gamma$.
533:
534:
535: \section{Equations of Motion}
536:
537:
538: To solve the model we proceed by making a Gaussian variational ansatz
539: (GVA), assuming that the fields $\Phi$ are described by the
540: approximate action
541: %
542: \begin{eqnarray}
543: & S_{var} = &
544: \frac{1}{2} \int d1 ds d2 ds'
545: \Phi(s,1) (G^{-1})_{12}^{ss'} \Phi(s',2) + \nonumber \\
546: && + \int d1 ds ds'
547: \Phi(s,1) (G^{-1})_{10}^{ss'} x_0(s') + \nonumber \\
548: && + \frac{1}{2} \int ds ds'
549: x_0(s) (G^{-1})_{00}^{ss'} x_0(s').
550: \label{Svar}
551: \end{eqnarray}
552: %
553: Technically, this implies the following approximation for $F_{dyn}$:
554: %
555: \begin{equation}
556: F_{dyn}\approx \langle S \rangle_{var} + F_{var}.
557: \label{Fdyn1}
558: \end{equation}
559: %
560: where
561: %
562: \begin{eqnarray}
563: & & \e^{-F_{var}}\equiv \int Dx_0 D\Phi \e^{-S_{var}} =\e^{(d/2)Tr\ln G}, \\
564: & & \langle . \rangle_{var}=\e^{F_{var}}\int Dx_0 D\Phi( . )\e^{-S_{var}}.
565: \end{eqnarray}
566: %
567: The stationarity condition
568: %
569: \begin{equation}
570: \frac{\delta F_{dyn}}{\delta G_{12}^{ss'}} = 0
571: \label{dFdG}
572: \end{equation}
573: %
574: translates into the equation of motion for Green's function
575: $G_{12}^{ss'}$ (see Eqs.~\ref{G12k}-\ref{G00k}). We have derived
576: identical equations of motion by using the approach of Ref. \cite{CD},
577: where standard field theoretic identities (e.g. $\langle \Phi \delta
578: S/\delta\Phi \rangle=0$ ) are used. It can be shown that for
579: quadratic $S_{var}$ the two procedures give the same result. We omit
580: this analysis here to save space.
581:
582: In a corresponding equilibrium problem, the stationarity condition
583: is also an extremum condition and provides a bound on the free
584: energy. Here, since $F_{dyn}$ contains integrations over complex
585: fields and Grassmann variables, the GVA does not give a bound on
586: $F_{dyn}$. Nevertheless, it is the first step in a systematic
587: approximation procedure, as outlined in Appendix B.
588:
589: The GVA has been applied to the problem of a manifold in a random
590: potential, in both statics \cite{MP1,MP2} and dynamics
591: \cite{CKD,CD}. The method is exact when the dimensionality of the
592: manifold is infinite but is only approximate for finite
593: dimensionality. Nevertheless, even for rather low dimensionality
594: it has been shown to be a very good approximation in the
595: random-manifold problem, where it has been checked numerically
596: \cite{CD}. We have shown in Paper I that the present model is
597: closely related to the random-manifold problem. Thus, we hope that
598: the GVA will also be reasonable here, although we have not
599: strictly checked its validity.
600:
601: Using (\ref{Fdyn1}), (\ref{Svar}) and (\ref{eq:SSUSY}) gives the following
602: expression for $F_{dyn}$:
603: %
604: \begin{eqnarray}
605: & & F_{dyn} = \frac{d}{2} \int dsds' K^{ss'}_{00} G^{ss'}_{00}
606: + \nonumber \\
607: & & + \frac{d}{2} \int dsds'd1d2K^{ss'}_{12}G^{ss'}_{12}
608: - \frac{d}{2} Tr\ln G \nonumber \\
609: & & - \frac{B^2}{4}
610: \int d^dxd^dy d1d2 \langle A_{12}^{(V)}(x,y) \rangle
611: \langle A_{12}^{(\delta)}(x,y) \rangle
612: \nonumber \\
613: & & - \frac{\beta_0 B^2}{2}
614: \int d^dxd^dy d1d2 \langle A_{10}^{(V)}(x,y) \rangle
615: \langle A_{10}^{(\delta)}(x,y) \rangle ,
616: \label{Fdyn2}
617: \end{eqnarray}
618: %
619: where all averages are to be calculated using $S_{var}$ (see
620: Eq.~\ref{Svar}). Performing averages, the fourth and fifth term on the
621: right hand side of (\ref{Fdyn2}) become
622: %
623: \begin{eqnarray}
624: & & F_{dyn}^{(4)} =
625: - \frac{d}{2N}
626: \int d1d2dsds'{\cal V}\left[ (B_{12}^s+B_{12}^{s'})/2 \right], \\
627: & & F_{dyn}^{(5)} =
628: - \frac{\beta_0 d}{N}
629: \int d1d2dsds'{\cal V}\left[ (B_{10}^s+B_{10}^{s'})/2 \right],
630: \end{eqnarray}
631: %
632: where
633: %
634: \begin{eqnarray}
635: & & B_{12}^s = \langle [ \Phi(s,1)-\Phi(s,2) ]^2 \rangle =
636: G_{11}^{ss} + G_{22}^{ss} - 2 G_{12}^{ss}, \\
637: & & B_{10}^s = \langle [ \Phi(s,1)-x_0(s) ]^2 \rangle =
638: G_{11}^{ss} + G_{00}^{ss} - 2 G_{10}^{ss},
639: \end{eqnarray}
640: %
641: and
642: %
643: \begin{equation}
644: {\cal V}(z) = - \frac{\tilde B^2}{d}(z+\sigma)^{-d/2}\ , \ \
645: \tilde B^2 = \frac{B^2}{2} \frac{N}{v} (4\pi)^{-d/2}.
646: \end{equation}
647: %
648:
649: Performing the variational ansatz (i.e., evaluating Eq.~\ref{dFdG})
650: results in the following equations of motion:
651: \begin{eqnarray}
652: & & \left[ T(\mu+k^2)-D^{(2)}_1 \right] G_{12}^k = \delta_{12}
653: + 2 \int d3 {\cal V}'(B_{13}) \times \nonumber \\
654: & & \times (G_{32}^k-G_{12}^k)
655: + 2 \beta_0 {\cal V}'(B_{10}) (G_{02}^k-G_{12}^k),
656: \label{G12k}
657: \end{eqnarray}
658: \begin{eqnarray}
659: & & \left[ T(\mu+k^2)-D^{(2)}_1 \right] G_{10}^k =
660: 2 \int d2 {\cal V}'(B_{12}) \times \nonumber \\
661: & & (G_{20}^k-G_{10}^k) + 2 \beta_0 {\cal V}'(B_{10})(G_{00}^k-G_{10}^k),
662: \end{eqnarray}
663: \begin{eqnarray}
664: & ( \mu_0 + k^2) G_{01}^k = & 2 \beta_0 \int d2 {\cal V}'(B_{20})
665: (G_{21}^k-G_{01}^k),
666: \end{eqnarray}
667: \begin{equation}
668: ( \mu_0 + k^2) G_{00}^k = 1 + 2 \beta_0 \int d1 {\cal V}'(B_{10})
669: (G_{10}^k-G_{00}^k),
670: \label{G00k}
671: \end{equation}
672: %
673: and after disentangling the SUSY notation one gets (see Paper I for
674: related details)
675: %
676: \begin{eqnarray}
677: & [T & (\mu+k^2)+\partial/\partial t] C_k(t,t') = \nonumber \\
678: & & 2 T R_k(t',t)
679: + 2 \int_{0}^{t} dt'' {\cal V}'\left[B(t,t'')\right] R_k(t',t'')
680: \nonumber \\
681: & & + 4 \int_{0}^{t} dt'' {\cal V}''\left[B(t,t'')\right]
682: r(t,t'') \left[ C_k(t,t')-C_k(t'',t') \right] \nonumber \\
683: & & - 2 \beta_0 {\cal V}'[A(t)] [ C_k(t,t')-\phi_k(t') ] ,
684: \label{Ck}
685: \end{eqnarray}
686: \begin{eqnarray}
687: & [T & (\mu+k^2)+\partial/\partial t] R_k(t,t') = \delta(t-t') + \nonumber \\
688: & & + 4 \int_{0}^{t} dt'' {\cal V}''\left[B(t,t'')\right] r(t,t'') \left[
689: R_k(t,t')-R_k(t'',t') \right] \nonumber \\
690: & & - 2 \beta_0 {\cal V}'[A(t)]R_k(t,t'),
691: \label{Rk}
692: \end{eqnarray}
693: \begin{eqnarray}
694: & [ T & (\mu+ k^2)+\partial/\partial t] \phi_k(t) = \nonumber \\
695: & & 4 \int_{0}^{t} dt'' {\cal V}''\left[B(t,t'')\right]
696: r(t,t'') \left[ \phi_k(t)-\phi_k(t'') \right] \nonumber \\
697: & & + 2 \beta_0 {\cal V}'[A(t)] (\Gamma_k-\phi_k(t)),
698: \label{Phik1}
699: \end{eqnarray}
700: \begin{eqnarray}
701: & & (\mu_0+k^2) \phi_k(t) =
702: 2 \beta_0 \int_{0}^{t}dt''{\cal V}'[A(t'')] R_k(t,t''),
703: \label{Phik2}
704: \end{eqnarray}
705: \begin{eqnarray}
706: & & (\mu_0+k^2) \Gamma_k = 1,
707: \end{eqnarray}
708: %
709: where $B(t,t')$ and $A(t)$ are defined as $B(t,t') = \langle (x(s,t) -
710: x(s,t'))^2 \rangle = C(s,t;s,t) + C(s,t';s,t') - 2 C(s,t;s,t')$ and
711: $A(t) = \langle (x(s,t) - x_0(s))^2 \rangle = C(s,t;s,t) -
712: 2\phi(s,t;s) + \Gamma(s,s)$. Note that due to translational
713: invariance with respect to $s$ both $B(t,t')$ and $A(t)$ are
714: $s$-independent. The equations of motion for $C_k(t,t')$ and
715: $R_k(t,t')$ are almost identical to those for the pure RHP
716: model. Coupling to the native state enters through the terms
717: proportional to $\beta_0$. Again, for large selection temperature
718: $\beta_0\rightarrow 0$ and one recovers the RHP model.
719:
720:
721:
722:
723: \section{Extracting order parameters} \label{sec:order_parameters}
724:
725:
726: The equations of motion are coupled integro-differential equations
727: with initial conditions given by $C_k(0,0)$, $\phi(0)$, and (we use
728: Ito's convention) $R(t+\epsilon,t)\rightarrow 1$ as
729: $\epsilon\rightarrow 0$. To solve the equations analytically we have
730: to consider several assumptions (which can be checked by numerical
731: solution).
732:
733: First, we make the (rather strong) standard assumptions from aging
734: theory for spin glasses about the asymptotic behavior of the
735: solutions: In the regime of time translational invariance (TTI),
736: %
737: \begin{eqnarray}
738: & & \lim_{t\rightarrow\infty} C_k(t+\tau,t) = C_k(\tau), \\
739: & & \lim_{t\rightarrow\infty} R_k(t+\tau,t) = R_k(\tau),
740: \end{eqnarray}
741: %
742: and, in the aging regime,
743: %
744: \begin{eqnarray}
745: & & \lim_{t\rightarrow\infty} C_k(t,\lambda t) =
746: q_k \hat C_k(\lambda), \\
747: & & \lim_{t\rightarrow\infty} R_k(t,\lambda t) =
748: \frac{1}{t} \hat R_k(\lambda).
749: \end{eqnarray}
750: %
751: The validity of these assumptions could be checked numerically. Since
752: this has been done for equations of similar type elsewhere~\cite{CD},
753: we omit it in the present analysis.
754:
755: Second, it is well known that asymptotic solutions of such equations
756: can be characterized by a few order parameters
757: \cite{CKD,CD,CK1,BCKP,CK2}. They are defined as
758: %
759: \begin{eqnarray}
760: & & \tilde q_k = \lim_{t\rightarrow\infty} C_k(t,t), \\
761: & & q_k = \lim_{\tau\rightarrow\infty} C_k(\tau), \\
762: & & q_{0,k} = \lim_{\lambda\rightarrow 0} q_k \hat C_k(\lambda), \\
763: & & \varphi_k=\lim_{t\rightarrow\infty}\phi_k(t).
764: \end{eqnarray}
765: %
766: The following $k$-integrated quantities will also be useful:
767: %
768: \begin{eqnarray}
769: & & \tilde q \equiv \int \frac{dk}{2\pi} \tilde q_k =
770: \lim_{t\rightarrow\infty} \langle x(s,t)x(s,t) \rangle , \\
771: & & q \equiv \int \frac{dk}{2\pi} q_k =
772: \lim_{\tau\rightarrow\infty}
773: \lim_{t\rightarrow\infty} \langle x(s,t)x(s,t+\tau) \rangle ,\\
774: & & q_0 \equiv \int \frac{dk}{2\pi} q_{0,k} =
775: \lim_{\lambda\rightarrow 0}
776: \lim_{t\rightarrow\infty} \langle x(s,t)x(s,\lambda t) \rangle ,\\
777: & & \varphi \equiv \int \frac{dk}{2\pi} \varphi_k =
778: \lim_{t\rightarrow\infty} \langle x(s,t) x_0(s) \rangle.
779: \end{eqnarray}
780: %
781: $\tilde q$ measures the size of the globule, $q$ measures the
782: persistent correlation in the TTI regime, $q_0$ the asymptotic
783: correlation in the aging regime, and $\varphi$ the overlap with native
784: state. Also, it is useful to define
785: %
786: \begin{eqnarray}
787: & & b = 2 ( \tilde q - q ) \, , \,\,\, b_0 = 2 ( \tilde q - q_0 ),
788: \label{bqq0} \\
789: & & a \equiv \lim_{t\rightarrow\infty}
790: \langle [x(s,t)-x_0(s)]^2 \rangle
791: = \tilde q - 2 \varphi + \frac{1}{2\sqrt{\mu_0}}.
792: \end{eqnarray}
793: %
794:
795:
796: Third, we assume that the generalized fluctuation dissipation theorem
797: is valid in the form
798: %
799: \begin{equation}
800: \hat R_k(\lambda) = \frac{x}{T}
801: \, q_k \frac{ d\hat C_k(\lambda) }{ d\lambda },
802: \label{GFDT}
803: \end{equation}
804: %
805: $x$ could in principle depend on $k$ and $C_k$. However, related
806: models have been studied in detail and they exhibit one step replica
807: symmetry breaking with a $k$-independent $x$. This one step replica
808: symmetry breaking ansatz in our dynamical study translates exactly to
809: Eq.~(\ref{GFDT}).
810:
811:
812: \section{Relating order parameters}
813:
814: For $t=t'$ and $t\rightarrow\infty$ Eq.~(\ref{Ck}) gives
815: %
816: \begin{eqnarray}
817: & & T(\mu+k^2) \tilde q_k = T
818: + \frac{2}{T} {\cal V}'(b) (1-x) ( \tilde q_k - q_k ) \nonumber \\
819: & & \ \ \ \ \ \
820: + \frac{2}{T} {\cal V}'(b_0) x ( \tilde q_k - q_{0,k})
821: -2\beta_0 {\cal V}'(a)(\tilde q_k - \phi_k).
822: \label{eq:qtk2}
823: \end{eqnarray}
824: %
825: With $t=t'+\tau$ and $t'\rightarrow\infty$ and then
826: $\tau\rightarrow\infty$ Eq.~(\ref{Ck}) becomes
827: %
828: \begin{eqnarray}
829: & & T(\mu+k^2) q_k =
830: \frac{2}{T} ( {\cal V}'(b)
831: - x {\cal V}'(b_0) ) ( \tilde q_k - q_k ) \nonumber \\
832: & & \ \ \ \ \ \
833: + \frac{2}{T} {\cal V}'(b_0) x ( \tilde q_k - q_{0,k} )
834: -2\beta_0 {\cal V}'(a)(q_k - \phi_k).
835: \label{eq:qk2}
836: \end{eqnarray}
837: %
838: Eq.~(\ref{Ck}) in the aging regime $t'=\lambda t$, first for
839: $t\rightarrow\infty$ and then $\lambda\rightarrow 0$, gives
840: %
841: \begin{eqnarray}
842: & & T(\mu+k^2) q_{0,k} =
843: \frac{2}{T} {\cal V}'(b_0) (1-x) ( \tilde q_k - q_k ) \nonumber \\
844: & & \ \ \ \ \ \
845: + \frac{2}{T} {\cal V}'(b_0) x ( \tilde q_k - q_{0,k} )
846: -2\beta_0 {\cal V}'(a)(q_{0,k} - \phi_k).
847: \label{eq:q0k2}
848: \end{eqnarray}
849: %
850: Eqs.~(\ref{Phik1}) and (\ref{Phik2}) result in two equations for
851: $\varphi_k$,
852: %
853: \begin{eqnarray}
854: & T ( \mu + k^2 ) \varphi_k = &
855: \frac{2}{T_0} {\cal V}'(a) ( \Gamma_k - \varphi_k )
856: \label{varphik1} \\
857: & (\mu_0 + k^2 ) \varphi_k = &
858: \frac{2}{TT_0} {\cal V}'(a) x ( q_k - q_{o,k} ) \nonumber \\
859: & & + \frac{2}{TT_0} {\cal V}'(a) ( \tilde q_k - q_k )
860: \label{varphik2}
861: \end{eqnarray}
862: %
863: They are equivalent; one can chose to solve for the order parameters
864: working with either (\ref{varphik1}) or (\ref{varphik2}). This seems a
865: rather remarkable coincidence. We believe that it originates from the
866: SUSY invariance of the original action $S$. For example, a similar
867: comment holds for equations (\ref{Ck}) and (\ref{Rk}); they are
868: equivalent in the TTI regime and one can derive one from the other.
869: The `conspiracy' of (\ref{Phik1}) and (\ref{Phik2}) not contradicting
870: each other is very likely a similar phenomenon. Eq.(\ref{Rk}) for
871: $\lambda=1$ reduces to
872: %
873: \begin{equation}
874: \hat R_k(1) (\tilde\mu+k^2+\Sigma) = - (\tilde q_k - q_k )
875: \frac{4{\cal V}''(b)}{T^2} \hat r(1),
876: \label{MSCk}
877: \end{equation}
878: %
879: where
880: \begin{equation}
881: \hat r(\lambda) \equiv \int \frac{dk}{2\pi} \hat R_k(\lambda)
882: \end{equation}
883: and $\Sigma$ is defined by
884: %
885: \begin{equation}
886: \Sigma = x \frac{2}{T^2} \left( {\cal V}'(b) - {\cal V}'(b_0) \right).
887: \label{Sigma}
888: \end{equation}
889: %
890:
891: Solving these equations for the order parameters gives
892: %
893: \begin{eqnarray}
894: & & b = \frac{1}{\sqrt{\tilde\mu+\Sigma}}, \label{qtq}
895: \end{eqnarray}
896: \begin{eqnarray}
897: & & b_0 =
898: \frac{1}{x} \frac{1}{\sqrt{\tilde\mu}}
899: + \frac{x-1}{x} \frac{1}{\sqrt{\tilde\mu+\Sigma}}
900: \label{qq0},
901: \end{eqnarray}
902: \begin{eqnarray}
903: & \tilde q = & \frac{b_0}{2}
904: + \frac{{\cal V}'(b_0)}{4T^2\tilde\mu^{3/2}}
905: + \frac{1}{4\sqrt{\mu_0}}
906: \left(
907: \frac{1-\frac{\mu}{\tilde\mu}}{1-\frac{\mu_0}{\tilde\mu}}
908: \right)^2 \times \nonumber \\
909: & & \times \left(2+\sqrt{\frac{\mu_0}{\tilde\mu}} \right)
910: \left( 1 -\sqrt{\frac{\mu_0}{\tilde\mu}} \right)^2,
911: \label{qt}
912: \end{eqnarray}
913: \begin{eqnarray}
914: & a = & \frac{b_0}{2}
915: + \frac{{\cal V}'(b_0)}{4T^2\tilde\mu^{3/2}}
916: + \frac{1}{4\sqrt{\mu_0}}
917: \frac{1}{
918: \left( 1+\sqrt{\frac{\mu_0}{\tilde\mu}} \right)^2
919: } \times \nonumber \\
920: & & \times
921: \left[
922: \sqrt{\frac{\mu_0}{\tilde\mu}}
923: \left( 1+2\sqrt{\frac{\mu_0}{\tilde\mu}} \right)
924: + 2 \frac{\mu}{\tilde\mu}\sqrt{\frac{\mu_0}{\tilde\mu}}
925: \right.
926: \nonumber \\
927: & & \left.
928: \,\,\,\,\,\,\,\,\,\, + \left( \frac{\mu}{\tilde\mu} \right)^2
929: \left( 2+\sqrt{\frac{\mu_0}{\tilde\mu}} \right)
930: \right],
931: \label{a}
932: \end{eqnarray}
933: \begin{eqnarray}
934: & & \tilde\mu = \mu + \frac{2}{TT_0}{\cal V}'(a), \label{mut}
935: \end{eqnarray}
936: %
937: and the combination of Eq.~(\ref{qtq}) and (\ref{MSCk}) gives
938: %
939: \begin{eqnarray}
940: && 0 = \hat r(1) \left[ T^2 + b^3 {\cal V}''(b) \right]. \label{MSC}
941: \end{eqnarray}
942: Furthermore, the overlap $\varphi$ with the native state
943: is given by
944: %
945: \begin{equation}
946: \varphi = \frac{1}{2\sqrt{\mu_0}}
947: \frac{
948: 1-\frac{\mu}{\tilde\mu}
949: }{
950: 1+\sqrt{
951: \frac{\mu_0}{\tilde\mu}
952: }
953: }.
954: \label{varphi}
955: \end{equation}
956: %
957: All overlap order parameters are positive. However, this result is
958: not obvious and has to be obtained after some algebra.
959:
960: These equations have two kinds of solutions. In one kind, $b=b_0$, so
961: there is no glassiness (aging). For this kind of solution, the
962: parameter $x$ is irrelevant. We call such solutions ``ergodic''.
963: (While it will turn out that some of them are not truly ergodic, in
964: the sense of describing states where the entire configuration space is
965: visited with Boltzmann probabilities, they violate ergodicity in a
966: rather trivial way, like a ferromagnet below the Curie temperature.
967: We could call them ``non-glassy'', but we prefer not to use a negative
968: term.)
969:
970: For an ergodic solution, with $b=b_0$, $\Sigma=0$. Furthermore, $\hat
971: r(\lambda)=0$, so Eqn.~(\ref{MSC}) is trivially satisfied. One then
972: has to solve the four equations (\ref{qtq}) and (\ref{qt}-\ref{mut})
973: for $b$, $\tilde q$, $a$ and $\tilde \mu$.
974:
975: The stability of such a phase against glassiness can be determined
976: using the analysis we presented in Paper I (see Fig.~1). There, we
977: studied a model with no native-state bias in its interactions ($T_0 =
978: \infty$) for finite $\mu$. The boundary of the glassy state as a
979: function of $\mu$ has a form qualitatively like that in the p-spin
980: glass as a function of field \cite{CS,CHS}. In the present model, the
981: presence of the native state enters the calculation solely through the
982: replacement of $\mu$ by $\tmu$. Therefore, if a particular $T$ and
983: $\tmu$ fall in the glassy regime (the region below the full and dashed
984: lines) in Fig.~1, the ergodic ansatz has to be given up.
985:
986: The instability can occur in two ways, according to whether $\tmu$ is
987: bigger or smaller than the critical value $\tmu_c$. Above $\tmu_c$,
988: the line separating glassy from ergodic regions is an Almeida-Thouless (AT)
989: line; below it the stability condition
990: %
991: \begin{equation}
992: T^2 + b^3 {\cal V}''(b) >0, \label{eq:ATinequality}
993: \end{equation}
994: %
995: is violated. For $\tmu < \tmu_c$, there is no AT instability. The
996: transition is like that for the completely random heteropolymer.
997: To find such a transition, we have to solve for a glassy phase,
998: characterized in part by a value of the FDT-violation parameter $x <
999: 1$ and then find where in the parameter space $x \rightarrow 1$. In
1000: the region where the $x<1$ solution exists, the associated ergodic
1001: phase is unstable and is replaced by the glassy one.
1002:
1003: In a glassy phase, aging is present: $\hat r(1)\ne 0$, so the quantity
1004: in brackets in Eq.~(\ref{MSC}) has to vanish, i.e., the AT condition
1005: has to be satisfied as an equality, rather than an inequality. This
1006: so-called marginal stability condition determines $b$ as a function of
1007: temperature. In this case we have three more unknowns, $\Sigma$,
1008: $b_0$ and $x$, making a total of seven, and seven equations,
1009: (\ref{qtq}-\ref{MSC}), to solve for them.
1010:
1011: We look for ergodic solutions first in the next section, and we
1012: examine their stability. Then, in the following section, we study
1013: glassy solutions (within the 1-step aging ansatz of section VII) and
1014: identify the regions in the parameter space where they hold.
1015:
1016:
1017: \section{Ergodic phases} \label{sec:erg_phase}
1018:
1019: For ergodic phases, Eqns.~(\ref{qtq}-\ref{mut}) reduce to
1020: %
1021: \begin{eqnarray}
1022: && b=b_0=\frac{1}{\sqrt{\tilde\mu}} \label{berg} \\
1023: & \tilde q = & \frac{1}{2\sqrt{\tilde\mu}}
1024: + \frac{{\cal V}'(1/\sqrt{\tilde\mu})}{4T^2\tilde\mu^{3/2}}
1025: + \frac{1}{4\sqrt{\mu_0}}
1026: \left(
1027: \frac{1-\frac{\mu}{\tilde\mu}}{1-\frac{\mu_0}{\tilde\mu}}
1028: \right)^2 \times \nonumber \\
1029: & & \times \left(2+\sqrt{\frac{\mu_0}{\tilde\mu}} \right)
1030: \left( 1- \sqrt{\frac{\mu_0}{\tilde\mu}} \right)^2
1031: \label{qterg} \\
1032: & a = & \frac{1}{2\sqrt{\tilde\mu}}
1033: + \frac{{\cal V}'(\frac{1}{\sqrt{\tilde\mu}})}{4T^2\tilde\mu^{3/2}}
1034: + \frac{1}{4\sqrt{\mu_0}}
1035: \frac{1}{
1036: \left( 1+\sqrt{\frac{\mu_0}{\tilde\mu}} \right)^2
1037: } \times \nonumber \\
1038: & & \times
1039: \left[
1040: \sqrt{\frac{\mu_0}{\tilde\mu}}
1041: \left( 1+2\sqrt{\frac{\mu_0}{\tilde\mu}} \right)
1042: + 2 \frac{\mu}{\tilde\mu}\sqrt{\frac{\mu_0}{\tilde\mu}}
1043: \right.
1044: \nonumber \\
1045: & & \left.
1046: \,\,\,\,\,\,\,\,\,\, + \left( \frac{\mu}{\tilde\mu} \right)^2
1047: \left( 2+\sqrt{\frac{\mu_0}{\tilde\mu}} \right)
1048: \right]
1049: \label{aerg} \\
1050: & & \tilde\mu = \mu + \frac{2}{TT_0} {\cal V}'(a) \label{muterg}
1051: \end{eqnarray}
1052: %
1053: They can be solved numerically: given $\mu$, $\mu_0$, $T$ and $T_0$
1054: one can find $\tilde\mu$, which in turn determines $\tilde q$, $b=b_0$
1055: (equivalently $q=q_0$), and $\varphi$. However, it is possible to gain
1056: some analytic understanding in a few soluble limits.
1057:
1058: In this discussion we will concentrate on the limit of small $\mu$.
1059: As we noted in paper I, if we want to confine $N$ monomers within a
1060: gyration radius $\sqrt{\tilde q} \propto \mu^{-1/4}$, we need $\mu
1061: \propto N^{-4/d}$. Thus, for a long polymer $\mu \rightarrow 0$. We
1062: will also take $\mu=\mu_0$ to simplify the algebra a bit.
1063:
1064: The pair of equations (\ref{aerg}) and (\ref{muterg}) fully determine
1065: $\tmu$ as function of $T$ and $T_0$. For $\mu_0=\mu$ they take the
1066: form
1067: %
1068: \bea
1069: a(\tmu) &=& \frac{1}{2\sqrt{\tmu}} +\frac{\Bt^2}{8T^2 \tmu^{3/2}
1070: (\sigma + \tmu^{-1/2})^{\frac{d}{2}+1}} \nonumber \\
1071: &+& \frac{1}{4\sqrt{\tmu}}\left( 1 + \frac{\mu}{\tmu}\right)
1072: \lb{eq:a} \\
1073: \tmu(a) &=& \mu + \frac{\Bt^2}{TT_0 (\sigma + a)^{\frac{d}{2}+1}}
1074: \lb{eq:tmu}
1075: \eea
1076: %
1077: Given $\tmu$, $T$, $T_0$ one can find the overlap with the native
1078: state $\varphi$ and the size of the polymer from (\ref{varphi}).
1079:
1080:
1081: \subsection{Random-globule state}
1082:
1083: It is immediately evident that when both the temperature $T$ and the
1084: selection temperature $T_0$ are large, $\tilde \mu \approx \mu$ in
1085: (\ref{muterg}), leading to a random-globule solution $a = b =
1086: \mu^{-1/2}$, $\tilde q = \mu^{-1/2}/2$, $\varphi = 0$. What is not so
1087: obvious is that in the $\mu \rightarrow 0$ limit a solution very close
1088: to this exists all the way down to very low temperatures, even for
1089: small $T_0$. In this subsection we examine this state in detail.
1090:
1091: We look first for solutions of Eqs.~(\ref{eq:a}) and (\ref{eq:tmu})
1092: with the ansatz $\alpha \equiv \tmu/\mu$ fixed and $\mu\rightarrow
1093: 0$. We call this the random globule ansatz, since, as will be shown,
1094: the polymer does not have any fixed conformation (it is melted), and
1095: on the average the conformations it adopts have zero overlap with the
1096: native state. (Strictly speaking, this is the only truly ergodic
1097: phase we find.) For $a$ we get,
1098: %
1099: \begin{equation}
1100: a = \frac{1}{\sqrt{\mu}}
1101: \left[
1102: \frac{
1103: 3+\frac{1}{\alpha}
1104: }{
1105: 4\sqrt{\alpha}
1106: }
1107: +
1108: {\cal O}(\mu^{(d-2)/2})
1109: \right],
1110: \end{equation}
1111: %
1112: which, after inserting into (\ref{eq:tmu}), gives
1113: %
1114: \begin{equation}
1115: \alpha \approx 1 + \frac{\Bt^2}{TT_0} \mu^{(d-2)/4}
1116: \left(
1117: \frac{
1118: 4 \sqrt{\alpha}
1119: }{
1120: 3 + \frac{1}{\alpha}
1121: }
1122: \right)^{d/2+1}.
1123: \label{alpha}
1124: \end{equation}
1125: %
1126: Eqn.~(\ref{alpha}) can be used to calculate $\alpha$ as a function
1127: of $\mu$. One can see easily that $\alpha\rightarrow 1$ when
1128: $\mu\rightarrow 0$. This shows that our ansatz is self-consistent
1129: in the limit of small $\mu$. Also, (\ref{qterg}) and
1130: (\ref{varphi}) become
1131: %
1132: \begin{eqnarray}
1133: & & \varphi=\frac{1}{2\sqrt{\mu}}
1134: \left[\frac{\alpha-1}{2}+{\cal O}(\alpha-1)^2\right]
1135: \label{varphi2} \\
1136: & & \tilde q=\frac{1}{2\sqrt{\mu}}\left[1+{\cal O}(\alpha-1)\right].
1137: \label{qterg2}
1138: \end{eqnarray}
1139:
1140: The normalized overlap between the polymer conformation and the native
1141: state is:
1142: %
1143: \begin{equation}
1144: \cos \theta =
1145: \frac{
1146: \lim_{t\rightarrow\infty}\langle x(s,t)x_0(s)\rangle
1147: }{
1148: \sqrt{
1149: \lim_{t\rightarrow\infty} \langle x(s,t)^2 \rangle
1150: \langle x_0(s)^2 \rangle
1151: }
1152: } =
1153: \frac{\varphi
1154: }{
1155: \sqrt{\tilde q (\frac{1}{2\sqrt{\mu}}})
1156: }.
1157: \label{costh}
1158: \end{equation}
1159: %
1160: >From (\ref{varphi2}) and (\ref{qterg2}) we get
1161: $\cos(\theta)\sim(\alpha-1)\sim\mu^{(d-2)/4}$. Thus there is no
1162: overlap with native state as $\mu\rightarrow 0$.
1163:
1164: Furthermore, to check that polymer does not freeze into some other
1165: conformation, we calculate the normalized overlap between two
1166: configurations taken at very different times,
1167: %
1168: \begin{equation}
1169: \cos \theta' =
1170: \frac{
1171: \lim_{\tau,t\rightarrow\infty} \langle x(s,t)x(s,t+\tau)\rangle
1172: }{
1173: \sqrt{ [\lim_{t\rightarrow\infty} \langle x(s,t)^2 \rangle ]^2 }
1174: } =
1175: \frac{q}{\tilde q}.
1176: \label{costh'}
1177: \end{equation}
1178: %
1179: After rewriting
1180: %
1181: \begin{equation}
1182: q/\tilde q=1-\frac{b}{2\tilde q}=
1183: 1-\frac{1}{2\sqrt{\tmu\tilde q}},
1184: \label{q/qt}
1185: \end{equation}
1186: %
1187: and, using (\ref{qterg2}), we get $\cos \theta'={\cal
1188: O}(\alpha-1)$. Again, as $\mu\rightarrow 0$, $\cos \theta' \rightarrow
1189: 0$. This confirms that the ansatz $\alpha ={\cal O}(1)$ and $\mu
1190: \rightarrow 0$ leads to a melted random-globule-like phase. This
1191: phase is identical to that found at high temperatures for the
1192: completely random heteropolymer in paper I.
1193:
1194: The validity of the present ansatz rests upon the fact that we can
1195: solve Eq.~(\ref{alpha}). Clearly, for $\mu\rightarrow 0$ a solution
1196: can always be found, namely $\alpha=1$. Since the physically relevant
1197: $\mu$ is $\propto N^{-4/d}$, we can always satisfy this equation, for
1198: any $T_0$, in the limit $N \rightarrow\infty$.
1199:
1200: We now address briefly the question of what happens for finite $N$
1201: (and $\mu$). One can easily see that Eq.~(\ref{alpha}) has two
1202: solutions when $\mu^{(d-2)/4}/(TT_0)$ is not too large (e.g. by
1203: plotting the left- and right-hand side as functions of $\alpha$). The
1204: solution close to 1 is lost when the slopes of the left- and
1205: right-hand sides become roughly equal. Evaluating these slopes leads
1206: to the condition
1207: %
1208: \be
1209: \frac{3\Bt^2}{4TT_0}\left(\frac{d}{2}+1\right)\mu^{\frac{d-2}{4}} <1
1210: \label{eq:coilcond}
1211: \ee
1212: %
1213: for the existence of a random-globule-like state.
1214:
1215: Some caution is in order. Working this out for finite $N$, $d=3$, and
1216: an average density of $1$, we find that the inequality
1217: (\ref{eq:coilcond}) is violated below a temperature
1218: %
1219: \be
1220: T_x = \left( \frac{\pi}{6}\right)^{1/2}\frac{15\Bt^2}{8T_0} N^{-1/3}.
1221: \ee
1222: %
1223: With the small power of $N^{-1}$, one has to go to quite large $N$ to
1224: make this temperature very low. Thus our statement that the
1225: random-globule-like state exists for all temperatures in the $\mu
1226: \rightarrow 0$ limit may be of limited relevance for real
1227: 3-dimensional heteropolymers of the length of typical proteins.
1228: Nevertheless, here we are just considering this simple limit.
1229:
1230: We now discuss the stability of this solution. In the large-$N$ limit,
1231: it is locally stable against spontaneous formation of a native-like
1232: state at any $T$ and $T_0$. However, it is unstable against glass
1233: formation at low temperatures: Since it is identical with the
1234: random-globule solution of the completely random heteropolymer
1235: problem, we can take over the result from paper I that it is unstable
1236: below a temperature $T_g \propto \Bt$, with the constant of
1237: proportionality of order 1. This glass temperature is independent of
1238: $T_0$. (In Fig.~1 this is the transition at $\tmu \rightarrow 0$.)
1239: Thus, wherever the system is in a random-globule-like state at $T >
1240: T_g$, it will no longer equilibrate if the temperature is lowered
1241: below $T_g$. Instead, it will become glassy and its dynamics will
1242: show aging.
1243:
1244:
1245: \subsection{Ergodic native state}
1246:
1247: At low $T_0$ and $T$, one expects that the polymer should be very
1248: close to its native state, i.e., small $a$. Therefore we also look
1249: for such solutions of Eqs.~(\ref{eq:a}) and (\ref{eq:tmu}). We will
1250: try to solve equations (\ref{eq:a}) and (\ref{eq:tmu}) in the limit
1251: where $\mu \rightarrow 0$ and $\tmu$ stays finite. The limit
1252: $\mu\rightarrow 0$ turns out not to involve any subtleties when $\tmu$
1253: is kept constant, so we will just set $\mu = 0$ from the outset.
1254: Eqs.~(\ref{eq:a}) and (\ref{eq:tmu}) become
1255: %
1256: \begin{eqnarray}
1257: a(\tmu) &=& \frac{3}{4\sqrt{\tmu}} +\frac{\Bt^2}{8T^2 \tmu^{3/2}
1258: (\sigma + \tmu^{-1/2})^{\frac{d}{2}+1}} \lb{eq:a1} \\
1259: \tmu(a) &=& \frac{\Bt^2}{TT_0 (\sigma + a)^{\frac{d}{2}+1}}
1260: \lb{eq:tmu1}
1261: \end{eqnarray}
1262: %
1263: These equations can be solved for $\tmu$ as function of $T$ and
1264: $T_0$. However, one has to keep in mind that $\mu\rightarrow 0$ has
1265: been taken. This implies that (\ref{varphi}) and (\ref{qterg}) become
1266: %
1267: \begin{equation}
1268: \varphi \approx \frac{1}{2\sqrt{\mu}} , \ \ \
1269: \tilde q \approx \frac{1}{2\sqrt{\mu}}
1270: \label{varphiqt},
1271: \end{equation}
1272: %
1273: and, inserting (\ref{varphiqt}) into (\ref{costh}), the normalized
1274: overlap between native state and polymer conformations, becomes $\cos
1275: \theta \approx 1$. Furthermore, because of its large overlap with the
1276: native state, the polymer is essentially frozen. This can be seen by
1277: calculating the normalized overlap between two polymer conformations
1278: after a very long time interval, as in previous section. Inserting
1279: (\ref{varphiqt}) into (\ref{costh'}) and (\ref{q/qt}) gives $\cos
1280: \theta' \approx 1-\sqrt{\mu/\tmu}\rightarrow 1$.
1281:
1282: There is interesting behavior associated with the limit
1283: $\mu\rightarrow 0$ for very long polymers. When the polymer gets
1284: longer and longer ($N\rightarrow\infty$) a finite part of the chain is
1285: not in the native state conformation, since $a$ stays constant. The
1286: rest of the chain is in the native state, which can be seen from the
1287: fact that overlap with native state approaches 1. Thus, in the limit
1288: of a very long polymer, the fraction of chain not in the native
1289: state conformation becomes negligible: the recipe for biasing the
1290: coupling constants $B_{ss'}$ described in chapter II works best for
1291: long polymers.
1292:
1293: In the following we will proceed with the solution of equations
1294: (\ref{eq:a1}) and (\ref{eq:tmu1}). Before continuing, it will be
1295: useful to compactify notation a bit. Making the change of variables
1296: $\hat X = X / \sigma$ for $X=b,\,b_0,\,a,\,q,\,\tilde q$; $\hat Y = Y
1297: \sigma^2$ for $Y=\mu,\,\tmu$; and $\hat Z=Z\sigma^{(d-2)/4}/\Bt$ for
1298: $Z=T,\,T_0$, we get equations of the same form, with $X \rightarrow
1299: \hat X$, $Y \rightarrow \hat Y$ and $Z \rightarrow \hat Z$, but with
1300: $\sigma =1$ and $\Bt=1$. Thus, without loss of generality, we can
1301: choose units with $\sigma=1$ and $\Bt=1$ (and remove the hats). From
1302: now on we do this.
1303:
1304: The working strategy for solving the equations is as follows. For
1305: fixed $T$, one can consider $T_0$ as a function of $\tmu$. This can be
1306: easily done by inserting the expression for $a$ from Eq.~(\ref{eq:a1})
1307: into (\ref{eq:tmu1}), thus writing $\beta_0 = 1/T_0$ as
1308: %
1309: \begin{equation}
1310: \beta_0(\tmu,T) = T\tmu[1+a(\tmu,T)]^{d/2+1}
1311: \label{invt0tmu}
1312: \end{equation}
1313: %
1314: The four panels of Fig.~2 shows the shape of $\beta_0(\tmu,T)$ as a
1315: function of $\tmu$ for four different temperatures. We want ultimately
1316: to construct a phase diagram in the $(\beta_0,T)$ plane. Therefore we
1317: have to specify $T$ (one panel of the figure) and $\beta_0$ and ask
1318: whether one or more solutions, i.e., particular values of $\tmu$ which
1319: solve Eq.~(\ref{invt0tmu}), exist. For example, in panel (a) in
1320: Fig.~2, a horizontal line at $\beta_0>\beta_0^{\rm min}$ intersects
1321: $\beta_0(\tmu,T)$ curve at two places, indicating two solutions
1322: $\tmu=\tmu_{1}, \tmu_{2}$. To make the figures more readable we have
1323: shown such a horizontal line, at a particular values of $\beta_0$,
1324: only in panel (a). If this horizontal line is moved below
1325: $\beta_0^{\rm min}$, it will never intersect the $\beta_0(\tmu,T)$
1326: curve. Thus, we can see that for every $T$, there is a value
1327: $\beta_0^{\rm min}(T)$ below which no solutions exist.
1328:
1329: We proceed with the analysis of Fig.~2. For sufficiently high
1330: temperatures (panels (a)-(c)) there are exactly two solutions for all
1331: $\beta_0 > \beta_0^{\rm min}$. Of these, the one with the larger
1332: value of $\tmu$ is a stable solution (local free energy minimum)
1333: describing the ergodic native phase. For example, the solution
1334: labeled $\tmu_2$ in panel (a) is of this sort. The one with the
1335: smaller value of $\tmu$ (e.g., the one labeled by $\tmu_1$ in panel
1336: (a)) is unstable. It describes a free energy maximum between the
1337: minima at the random-globule and ergodic native states. We will call
1338: such states ``unstable stationary'' (abbreviated US). (We have not
1339: done a static calculation to show this, but the situation here is
1340: analogous to that in an ordinary ferromagnet below the Curie
1341: temperature. There, one has three solutions of the mean field
1342: equations, one with positive, one with negative, and one with zero
1343: magnetization. The middle one, with zero magnetization, is
1344: unstable). The US state has a lower overlap with the native
1345: conformation than the ergodic-native solution does, because it has a
1346: smaller value of $\tmu$. As $\beta_0$ is increased from below through
1347: $\beta_0^{\rm min}$, the native-state and US-state solutions appear
1348: together and separate. For the temperatures of panels (a)-(c), they
1349: both exist for all $\beta_0 >\beta_0^{\rm min}$.
1350:
1351: Panel (d) (at the lowest of the temperatures) shows a more complex
1352: behavior where double-minimum structure appears. We have found
1353: numerically that this happens below $T\approx 0.20$. Here the
1354: behavior around $\beta_0^{\rm min}$ is just as in the other cases, but
1355: we note that at this temperature $\beta_0(\tmu,T)$ has a second local
1356: minima at a smaller value of $\tmu$. Thus there is a range
1357: $\beta_1^{\rm min}<\beta_0<\beta_1^{\rm max}$ for which there are four
1358: solutions. The rightmost one is stable and describes the
1359: ergodic-native phase, as before. Moving from right to left, the
1360: solutions alternate between stability and instability. Thus the
1361: second solution from the left represents a locally stable
1362: conformation. It is also correlated with the native state, since
1363: $\tmu$ is finite (though we always find $\tmu \ll 1$ in 3 dimensions).
1364: The remaining two solutions (with $\partial \beta_0/\partial \tmu <0$)
1365: represent US states (local free energy maxima) between it and the
1366: random-globule phase in one direction and the ergodic-native phase in
1367: the other.
1368:
1369: Plotting $\beta_0^{\rm min}$ against $T$, we obtain the stability
1370: boundary indicated by the thick solid curve in Fig.~3. Within our
1371: present assumption of ergodicity, everywhere to the right of this line
1372: the ergodic-native phase is dynamically stable. One can invert the
1373: relation $\beta_0^{\rm min}(T)$, obtaining a transition temperature
1374: $T_n(\beta_0)$, the maximum temperature for which the ergodic native
1375: phase is dynamically stable. It is separated from the (also stable)
1376: random-globule phase by a barrier, the top of which is described by
1377: the unstable solution.
1378:
1379: In Fig.~3 we also indicate the region in the $(\beta_0,T)$ plane where
1380: the second locally-stable solution is found. This region has the form
1381: of a kind of sliver extending out toward large $\beta_0$ at low
1382: temperatures.
1383:
1384: So far we have not examined the stability of these solutions against
1385: glassiness. As indicated above, we do this with the help of Fig.~1:
1386: Stable solutions can not lie in the range $\tmu_{min} < \tmu <
1387: \tmu_{AT}$. In Fig.~2, these limits are marked on the $\tmu$ axes.
1388: We thus see, for example, in Panel (c), that the native-state
1389: solutions found for the range of $\beta_0$ corresponding to values of
1390: $\tmu$ between $\tmu_*$ and $\tmu_{AT}$ are not acceptable: they
1391: violate the AT stability condition (\ref{eq:ATinequality}).
1392:
1393: Similarly, in panel (b) the US solutions found for a range of
1394: $\beta_0$ values can also be seen to lie in the forbidden region. And
1395: the intermediate locally-stable states that we identified in panel (d)
1396: always lie in a glassy region.
1397:
1398: In Fig.~3 we also plot the AT line (\ref{MSC}) in the $(\beta_0,T)$
1399: plane, indicating the regions where the various kinds of ergodic
1400: solutions are forbidden. For the native-phase solutions, the
1401: forbidden region is a strip mostly at low values of $T$ (diagonally
1402: cross-hatched region between thick and AT line). However, it ``wraps
1403: around'' at the leftmost part of the region where those solutions are
1404: found.
1405:
1406: The forbidden region for the US solutions occupies most of the region
1407: where these solutions occur below $T_{max}$, the maximum temperature
1408: for a glass transition shown in Fig.~1, including the entire portion
1409: of it below $T_g$, the glass instability temperature of the
1410: random-globule state.
1411:
1412: The structure in a tiny region near the minimum value of $\beta_0$ for
1413: which ergodic-native solution are found is a bit complicated and
1414: cannot be seen in the top panel of Fig.~3. Therefore, the lower panel
1415: shows an enlargement of this region.
1416:
1417: In summary, we have found four kinds of ergodic solutions. One
1418: essentially describes a random globule state. It is locally stable
1419: (in the limit of a large globule) at all $T$ and $\beta_0$ against
1420: condensation into a native-like state, but unstable against glass
1421: formation everywhere below a transition temperature $T_g$. The second
1422: kind of solution describes a phase which is highly correlated with the
1423: native state conformation, and it is stable in most of the region
1424: where the solution exists. The third kind of solution describes a
1425: locally stable state, correlated with the native state but more weakly
1426: so than the ergodic native phase just described. It is never stable
1427: against glass formation. Finally, there are unstable solutions, found
1428: whenever the ergodic-native solutions exist. They describe US states,
1429: free energy maxima between pairs of the previously-described
1430: solutions. However, in a large part of the region where these
1431: solutions are found (roughly, everywhere below $T_{max} \approx T_g$)
1432: they violate the AT stability condition and so are not physically
1433: relevant.
1434:
1435: Outside the regions where these ergodic solutions are allowed, we have
1436: to look for glassy solutions. We do this in the next section.
1437:
1438:
1439: \section{Glassy phases}
1440:
1441: In a glassy phase, $\hat r(1)\ne 0$ and Eq.~(\ref{MSC}) has to be
1442: kept, which gives,
1443: %
1444: \begin{equation}
1445: T^2 = - b^3 {\cal V}''(b)
1446: \label{b(T)}
1447: \end{equation}
1448: %
1449: Also, equations (\ref{Sigma}), (\ref{qtq}) and (\ref{qq0}) can be
1450: rewritten in the form
1451: %
1452: \begin{eqnarray}
1453: && \frac{ {\cal V}'(b) - {\cal V}'(b_0) }{ b_0 - b } = \frac{T^2}{2}
1454: \frac{\sqrt{\tilde\mu}}{b}
1455: \left( \frac{1}{b} + \sqrt{\tilde\mu} \right),
1456: \label{b0b} \\
1457: && b_0 - b = \frac{1}{x} \left( \frac{1}{\sqrt{\tilde\mu}} - b \right).
1458: \label{xbb0}
1459: \end{eqnarray}
1460: %
1461: and, with $\mu_0=\mu$, (\ref{a}) and (\ref{mut}) become
1462: %
1463: \begin{eqnarray}
1464: a &=& \frac{b_0}{2} +\frac{1}{8T^2 \tmu^{3/2}
1465: (1 + b_0)^{\frac{d}{2}+1}} + \frac{1}{4\sqrt{\tmu}}
1466: \left( 1 + \frac{\mu}{\tmu}\right)
1467: \lb{eq:a2} \\
1468: \tmu &=& \mu + \frac{1}{TT_0 (1 + a)^{\frac{d}{2}+1}}
1469: \lb{eq:tmu2}
1470: \end{eqnarray}
1471: %
1472: The above equations can be solved as follows. Eq.~(\ref{b(T)}) gives
1473: $b$ as a function of $T$, and then (\ref{b0b}), (\ref{eq:a2}) and
1474: (\ref{eq:tmu2}) can be used to find $b_0$ and $\tmu$ as functions of
1475: $T$ and $T_0$. Once $b_0$ and $\tmu$ are found one can calculate
1476: $\tilde q$ as
1477: %
1478: \begin{eqnarray}
1479: & \tilde q = & \frac{b_0}{2} +
1480: \frac{1}{8T^2 \tilde\mu^{3/2}
1481: (1 + b_0)^{\frac{d}{2}+1}} + \nonumber \\
1482: & & +\frac{1}{4\sqrt{\mu}} ( 2 + \sqrt{\frac{\mu}{\tilde\mu}})
1483: (1-\sqrt{\frac{\mu}{\tilde\mu}})
1484: \label{qtglas1}
1485: \end{eqnarray}
1486:
1487: As in our analysis of ergodic solutions in the preceding section, we
1488: will try two types of ansatz: one with $\tmu/\mu=const$ as
1489: $\mu\rightarrow 0$ and one with $\tmu=const$ as $\mu\rightarrow 0$
1490: leading to what we call frozen-globule and glassy native phases,
1491: respectively.
1492:
1493:
1494:
1495: \subsection{Frozen-globule phase}
1496:
1497: The limit where $\alpha=\tmu/\mu$ is kept constant and $\mu\rightarrow
1498: 0$ is easily treated. Eq.~(\ref{b(T)}) stay the same, while
1499: Eq.~(\ref{b0b}) gives
1500: %
1501: \begin{equation}
1502: \frac{{\cal V}(b)-{\cal V}(b_0)}{b_0-b}
1503: \approx\frac{T^2\sqrt{\alpha}}{2b^2}\sqrt{\mu}.
1504: \end{equation}
1505: %
1506: Since $b$ is kept fixed the only solution of equation above is
1507: $b_0\rightarrow\infty$ as
1508: %
1509: \begin{equation}
1510: b_0 \approx \frac{\psi(b)}{\sqrt{\alpha\mu}},
1511: \label{b0glas}
1512: \end{equation}
1513: %
1514: where $\psi(b)$ is a function which depends only on $b$, as $\mu$ is
1515: sent to $0$. Inserting (\ref{b0glas}) into (\ref{eq:tmu2}) gives
1516: %
1517: \begin{equation}
1518: \alpha = 1 + {\cal O}(\mu^{(d-2)/4})
1519: \end{equation}
1520: %
1521: and $\alpha$ stays very close to $1$, as in the ergodic random globule
1522: case. Also, $\varphi$ is given by (\ref{varphi2}), while
1523: (\ref{qtglas1}) gives
1524: %
1525: \begin{equation}
1526: \tilde q = \frac{1}{2\sqrt{\mu}} \left[ \psi(b)
1527: + {\cal O}(\alpha-1) \right]
1528: \label{qtglas2}
1529: \end{equation}
1530: %
1531: which can be compared with ergodic globule result, Eq.~(\ref{qterg2}).
1532: Eq.~(\ref{costh}) stays the same, and one gets $\cos \theta\sim
1533: \alpha-1$ which goes to zero as $\mu\rightarrow 0$. There is no
1534: overlap with native state. Does the system freeze into some other
1535: configuration? To find out, we calculate overlap angles between
1536: configurations at time $t$ and a much later time $t'$. As discussed
1537: in section \ref{sec:order_parameters} there are two ways in which the
1538: limit $t,t\rightarrow\infty$ can be taken, leading to $q_0\ne q$.
1539:
1540: In the first limit, the equivalent of Eq.~(\ref{costh'}) for the
1541: ansatz used here reads
1542: %
1543: \begin{equation}
1544: \cos \theta'_g =
1545: \frac{
1546: \lim_{\lambda,t\rightarrow\infty} \langle x(s,t)x(s,\lambda t)\rangle
1547: }{
1548: \sqrt{ [\lim_{t\rightarrow\infty} \langle x(s,t)^2 \rangle ]^2 }
1549: } =
1550: \frac{q_0}{\tilde q}
1551: \label{costh'g}
1552: \end{equation}
1553: %
1554: Using Eq.~(\ref{bqq0}), we can write $\cos \theta'_g =1-b_0/2\tilde
1555: q$, and Eqs~(\ref{qtglas2}) and (\ref{b0glas}) give $\cos \theta'_g
1556: \sim \alpha-1$ which goes to $0$ as $\mu\rightarrow 0$. (This behavior
1557: is analogous to that found in p-spin glasses.) However, at
1558: not-too-long time scales (shorter than the waiting time), as in
1559: equation (\ref{costh''g}), the polymer is frozen:
1560: %
1561: \begin{equation}
1562: \cos \theta''_g =
1563: \frac{
1564: \lim_{\tau,t\rightarrow\infty} \langle x(s,t)x(s,t+\tau)\rangle
1565: }{
1566: \sqrt{ [\lim_{t\rightarrow\infty} \langle x(s,t)^2 \rangle ]^2 }
1567: } =
1568: \frac{q}{\tilde q}
1569: \label{costh''g}
1570: \end{equation}
1571: %
1572: Using Eq.~(\ref{bqq0}), we can write $\cos \theta''_g=1-b/2\tilde q$,
1573: and Eq.~(\ref{qtglas2}) gives $\cos \theta''_g\sim
1574: 1-\sqrt{\mu}b/\psi(b)$, which goes to $1$ as $\mu\rightarrow 0$.
1575: Thus, this glassy phase has no overlap with the native state.
1576:
1577: As discussed above, there is an upper temperature limit $T_g$
1578: (independent of $\beta_0$) above which this phase melts, leaving the
1579: system in the random-globule state. $T_g$ can be found from
1580: Eqs.~(\ref{b(T)}) and (\ref{b0b}), using $b_0 \rightarrow \infty$ and
1581: (\ref{xbb0}) with $x \rightarrow1$. This leads to a value $b =
1582: 2/(\half d - 1) = {\cal O}(1)$ at the transition and $T_g = 2 (\half d
1583: -1)^{\half (\half d - 1)}/(\half d +1)^{\half( \half d +1)}$. For
1584: $d=3$, $T_g \approx 0.535$.
1585:
1586:
1587:
1588: \subsection{Glassy native states}
1589:
1590: We also have to study the possible glassy phases with overlap with the
1591: native state, i.e., with finite $\tmu$ (and, accordingly, finite $a$)
1592: when $\mu\rightarrow 0$. In such a phase, as in the ergodic
1593: native-like states described above, the system moves only in the
1594: neighborhood of the native state configuration. However, in a
1595: ``glassy native'' state even this restricted motion is strongly
1596: suppressed by the complexity of the local potential energy surface,
1597: and a glassy phase results.
1598:
1599: As in the ergodic ansatz, the limit $\mu\rightarrow 0$ introduces no
1600: problems. Eqns.~(\ref{b(T)}) and (\ref{b0b}) remain the same as in
1601: the frozen-globule case, while the equations for $a$ and $\tmu$,
1602: (\ref{eq:a2}) and (\ref{eq:tmu2}) become
1603: %
1604: \begin{eqnarray}
1605: a(\tmu) &=& \frac{b_0}{2} +\frac{1}{8T^2 \tmu^{3/2}
1606: (1 + b_0)^{\frac{d}{2}+1}} + \frac{1}{4\sqrt{\tmu}}
1607: \lb{eq:a3} \\
1608: \tmu(a) &=& \frac{1}{TT_0 (1 + a)^{\frac{d}{2}+1}}
1609: \lb{eq:tmu3}
1610: \end{eqnarray}
1611: %
1612: Again, Eq.~(\ref{b(T)}) specifies $b$ as a function of $T$, and
1613: (\ref{b0b}), (\ref{eq:a3}) and (\ref{eq:tmu3}) determine $b_0$ and
1614: $\tmu$ as functions of $T$ and $T_0$. $\tilde q$ and $\varphi$ are
1615: given by $\varphi,\tilde q\approx 1/(2\sqrt{\mu})$.
1616:
1617: The overlap with the native state is the largest possible: $\cos
1618: \theta_g=1$, as can be easily seen from Eq.~(\ref{costh}) and the
1619: values for $\varphi$ and $\tilde q$ we have just given. The overlap
1620: between two conformations at very different times also takes its
1621: largest possible value. From Eqs.~(\ref{costh'g}) and
1622: (\ref{costh''g}), knowing that $b_0$ and $b$ do not depend on $\mu$ we
1623: have $\cos \theta'_g =1-b_0/2\tilde q\approx 1 - b_0
1624: \sqrt{\mu}\rightarrow 1$ and $\cos \theta''_g =1-b/2\tilde q\approx
1625: 1-b\sqrt{\mu}\rightarrow 1$. Thus, the polymer is frozen almost
1626: everywhere into the native conformation. However, the freezing is not
1627: total, since $a$ in (\ref{eq:a3}) is not zero. Furthermore, there is
1628: aging in the system, since $x$ in Eq.~(\ref{xbb0}) is not equal to 1.
1629:
1630: We turn now to the solution of the equations (\ref{b(T)}),
1631: (\ref{b0b}), (\ref{eq:a3}) and (\ref{eq:tmu3}). As for the
1632: corresponding ergodic phases we have to resort to numerical solution;
1633: here we describe the analysis. The working strategy is similar to the
1634: one presented in subsection IX.B; the goal is to find $\beta_0$ as
1635: function of $\tmu$ for fixed $T$ since, as in the ergodic native case,
1636: extrema of the function $\beta_0(\tmu,T)$ govern the phase boundaries.
1637:
1638: The procedure for finding value of the function $T_0(\tmu,T)$ is as
1639: follows. Eq~(\ref{b(T)}) determines $b$ as a function of $T$, to be
1640: referred to as $b(T)$. Once $b(T)$ is found from (\ref{b(T)}) it is
1641: inserted into Eq.~(\ref{b0b}), which determines $b_0(T,\tmu)$. The
1642: value found for $b_0$ is inserted into Eq.~(\ref{eq:a3}) to find $a$,
1643: and finally $\beta_0 =1/T_0$ is calculated from
1644: Eq.~(\ref{eq:tmu3}). Thus, at each temperature for which glassy
1645: solutions are possible, we can construct a graph of $\beta_0(\tmu)$,
1646: as we did for ergodic solutions in Fig.~2. We have used Mathematica
1647: to do these calculations. We can use these curves, together with the
1648: ergodic ones previously analyzed, to identify the possible states of
1649: the system at a given temperature and $\beta_0$ (Fig.~4). The
1650: procedure is fairly simple. At any given $\tmu$, only one of the
1651: solutions is physical: In the region $\tmu_{min}<\tmu<\tmu_{AT}$ one
1652: has to follow the glassy $\beta_0(\tmu)$ curve, while outside it one
1653: follows the ergodic one. In Fig.~4 the physical solution is indicated
1654: as the thick dashed curve. One then looks for solutions as
1655: intersections of this curve with a horizontal line at a given value of
1656: $\beta_0$, as done previously (e.g., as in Fig.~2, panel (a)) within
1657: the ergodic ansatz.
1658:
1659:
1660: In Fig.~4 this procedure is shown for several different values of $T$.
1661: In the first panel, $T$ lies just a little below $T_{max}$ (as in
1662: Panel (b) of Fig.~2). Suppose we start in the ergodic native phase at
1663: large $\beta_0$ and then lower $\beta_0$. (In Fig.~5, this would
1664: correspond to moving along a horizontal line (constant $T$) slightly
1665: above point B in panel (b) or (c)). We can lower $\beta_0$ all the
1666: way down to $\beta_0^{\rm min}$ without encountering an AT
1667: instability. So, just as in the ergodic analysis of section X.B,
1668: beyond $\beta_0^{\rm min}$ the ergodic native phase melts into the
1669: random-globule phase.
1670:
1671: In the same panel we can also analyze the what happens to the unstable
1672: stationary state in the same range of $\beta_0$ for this temperature.
1673: At very large $\beta_0$ we have an ergodic solution, but as we lower
1674: $\beta_0$ we pass through a range of $\tmu$, between $\tmu_{min}$ and
1675: $\tmu_{AT}$, where the ergodic solution is unstable against
1676: glassiness. In this region we must follow the glassy curve instead of
1677: the ergodic one. We interpret this glassy solution in the following
1678: way: The free energy landscape near the US maximum becomes rough in
1679: this range of values of $\beta_0$ (at this temperature), the same way
1680: the free energy landscape near the minimum corresponding to a
1681: thermodynamic phase becomes rough in a glassy state. We call it a
1682: ``glassy US state''.
1683:
1684: The next panel is for a slightly lower temperature (but still above
1685: $T_g$). Here, as we lower $\beta_0$ in the ergodic-native phase, we
1686: reach an AT instability before we get all the way down to the minimum
1687: on the ergodic curve. (In Fig.~5, this would correspond to moving on a
1688: line of constant $T$, meeting the AT line somewhere between points A
1689: and B in panel (b) or (c)). Furthermore, the only available glassy solution
1690: for $\tmu < \tmu_{AT}$ is one with negative $\partial \beta_0/\partial \tmu$,
1691: that is, it corresponds to the kind of glassy US state discussed
1692: above. As this is not a stable phase, we conclude that for this $T$,
1693: the minimum value of $\beta_0$ lies at this AT line, and beyond it
1694: there is no stable native-like state. We can follow the glassy US
1695: state back up to larger $\beta_0$, seeing that we eventually cross
1696: over to a normal (non-glassy) transition state.
1697:
1698: In the last panel, the temperature is lowered a bit more (below
1699: $T_g$). Again, starting in the ergodic native phase at large
1700: $\beta_0$ and lowering $\beta_0$, we encounter an AT instability and a
1701: glassy solution appears. (Equivalently, in Fig.~5 one moves on a
1702: horizontal line somewhere below $T_g$ until meeting the AT line for
1703: the first time.) For smaller $\beta_0$, we switch to the glassy
1704: curve, which has positive $\partial \beta_0/\partial \tmu$, describing
1705: a glassy native phase. We can follow this curve down to its minimum
1706: $\beta_0$, beyond which no phases correlated with the native phase
1707: exist. But, of course, following it back up toward large $\beta_0$ on
1708: the unstable branch, we can identify the glassy US state between
1709: the phase correlated with the native state and the one uncorrelated
1710: with it. (Above $T_g$, the latter is the random-globule phase; below
1711: it, it is the frozen-globule phase.)
1712:
1713:
1714:
1715: \section{Features of the phase diagram}
1716:
1717: The phase structure implied by this simple model is not so simple.
1718: Fig.~5 shows the phase diagram constructed from the above analysis.
1719: For clarity, we show in the top panel only the solutions that
1720: correspond to stable phases. The second panel shows the details in
1721: the region where the ergodic-native, glassy-native, random-globule and
1722: frozen globule states come together (or nearly so). The third panel
1723: shows the regions where ergodic and glassy US states are found.
1724:
1725: There are six distinct regions in the phase diagram. In region I
1726: (high $T$, small $\beta_0$), the only stable phase is the random
1727: globule. In region II (small $\beta_0$, $T <T_g$) it undergoes a
1728: glass transition to the frozen-globule phase. The properties of the
1729: system in this part of the phase diagram are the same as in the
1730: completely random heteropolymer model of paper I; the bias of the
1731: interactions toward a native state does not have any effect until a
1732: (temperature-dependent) threshold $\beta_0^c(T)$ is reached. This
1733: threshold is marked on the diagram by the lines separating region I
1734: from regions III and V and region II from region VI.
1735:
1736: To help thinking about these phases, we offer the schematic free
1737: energy-surface pictures of Fig.~6. They show how we imagine the free
1738: energy varies as a function of the native-state overlap coordinate
1739: $\varphi$. Fig.~6A depicts this cross-section through the free-energy
1740: surface in region I, where there is a single smooth minimum around
1741: $\varphi = 0$, representing the random-globule phase. Fig.~6B shows
1742: what happens in region II, where this phase is replaced by the
1743: frozen-globule phase. We represent this by drawing the free energy
1744: surface with many local minima. Fig.~6C shows what happens in the
1745: middle of region III, where there are two (smooth) minima, the new one
1746: corresponding to the ergodic native phase. (It will lie above or
1747: below that at $\varphi = 0$ according to whether we are above or below
1748: a first-order transition that we expect to occur at a temperature $T_1
1749: <T_n$, see below.) In Fig.~6D we depict the situation in region IV,
1750: where both the $\varphi \approx 0$ region and the region around the
1751: maximum become rough. Fig.~6E represents region V, with the native
1752: valley and the maximum rough, but the region near $\varphi = 0$ still
1753: smooth. In Fig.~6F (region VI), that, too, becomes rough.
1754:
1755: An important feature is the fact that the random-globule and
1756: frozen-globule phases remain (dynamically) stable in their respective
1757: temperature ranges for all $\beta_0$. Thus the horizontal line
1758: separating region I from region II continues across the diagram,
1759: separating region V from region VI and region II from region IV.
1760:
1761: In each of regions II, IV, V and VI (Fig.~6, panels (b), (d), (e),
1762: (f)) there are two stable states. Thus, regions I and III are
1763: separated by the line $T_n(\beta_0)$, below or to the right of which,
1764: in addition to the random-globule state, an ergodic native
1765: state exists and is stable. In region IV, the frozen-globule and this
1766: ergodic native state are both stable. In region V, the random globule
1767: state and a glassy native state are stable, while in region VI the
1768: stable states are the frozen globule and the glassy native phase.
1769:
1770: The diagram shows some interesting fine structure in the neighborhood
1771: of the region where regions I, III, and V meet (second panel, point
1772: A). In particular, below point A, the boundary between regions III
1773: and V (i.e., between the ergodic-native and glassy-native state) is an
1774: AT line. It comes about as can be seen in the last panel of Fig.~4,
1775: where, following the ergodic phase down from high $\beta_0$ and
1776: $\tmu$, we reach $\tmu_{AT}$ and thereafter have to switch to the
1777: glassy solution.
1778:
1779: On the other hand, the boundary $T_n(\beta_0)$ above point B is
1780: reached as in the first panel of Fig.~4: one can come all the way down
1781: to the minimum value of $\beta_0$ for the ergodic solution before
1782: reaching $\tmu_{AT}$. The full line continuing upwards and to the
1783: right of point B is an AT line which goes over into an $x=1$ line
1784: at its maximum, $T_{max}$. Below this line US solutions become
1785: glassy.
1786:
1787: Between points A and B, the boundary is marked by reaching $\tmu_{AT}$
1788: in the way shown in the second panel of Fig.~4: There is no stable
1789: ergodic native solution to the left of this line, since, upon lowering
1790: $\beta_0$ below $\beta_0^{\rm min}$ in Panel (b), the solution with
1791: $\partial \beta_0/\partial \tmu>0$ is lost. Thus in this region the
1792: line AB is an AT line for both the native phase and the US solutions.
1793: The dashed line marks the boundary found if one ignores the AT line
1794: (i.e., it is a portion of the boundary found using the ergodic ansatz
1795: and shown in Fig.~3).
1796:
1797: Everywhere below the AT line (both the portion AB and its extension
1798: upward to the right) is the region where the US solutions are glassy,
1799: as shown in the last panel of Fig.~5. At a given $\beta_0$, these
1800: features all have an onset at a temperature between $T_g$ and
1801: $T_{max}$. As can be seen in Fig.~1, this is a very small temperature
1802: range. This is also the reason why there is fine structure, as shown
1803: in both the second and the third panels of Fig.~5, in such a small
1804: temperature range in the phase diagrams.
1805:
1806: As remarked above, we have not done an equilibrium calculation, but we
1807: expect that the first-order transition temperature $T_1(\beta_0)$
1808: where the free energies of the random-globule and ergodic-native
1809: phases are equal will also rise with $\beta_0$. For large $\beta_0$,
1810: we expect $T_1$, like $T_n$, to be proportional to $\beta_0$ (but $T_1
1811: < T_n$, of course).
1812:
1813: Thus, at fairly large $\beta_0$, we expect the following sequence of
1814: stable states as we lower $T$ from a high value. Initially, only the
1815: random-globule state is stable. Then, below $T_n$, the ergodic native
1816: state is also stable, and below $T_1$ it becomes the lowest-free
1817: energy state. Going further down in $T$, we cross the boundary (last
1818: panel in Fig.~5) where the US state between the ergodic native and
1819: random-globule states becomes glassy (i.e., acquires a rough local
1820: free energy landscape). Very soon thereafter, we cross $T_g$, where
1821: the random-globule state undergoes glassy freezing. Continuing, we
1822: reach a temperature where the ergodic native state undergoes glassy
1823: freezing. Finally, we reach the stability limit of this glassy-native
1824: phase, leaving the system with nowhere to go but the frozen-globule
1825: state.
1826:
1827: What lessons are there in these findings for protein folding? We
1828: start from the assumption that the initial state in the folding
1829: process is uncorrelated with the native state (i.e., in Fig.~6 we
1830: start in a local minimum at $\varphi\approx 0$). Folding requires the
1831: system to find its way to the (ergodic) native state.
1832:
1833: One feature that is evident is that such a path in configuration space
1834: always requires an uphill free-energy step. This is because either
1835: the random-globule or frozen-globule phases is always locally stable.
1836:
1837: If we stick to our mean-field dynamical picture, where barriers are
1838: infinite, folding is, strictly speaking, impossible. In dynamical
1839: terms the ``infinite'' barriers translate into the fact that the
1840: equations which govern the motion of order parameters, $\varphi$
1841: included, have basins of attraction corresponding to the plots shown
1842: in Fig.~6. For example, starting from $\varphi$ somewhere close to $0$
1843: and given a free energy profile like that in Fig.~6C, dynamical
1844: equations will never carry $\varphi$ to the large value describing the
1845: native state. On the contrary, $\varphi$ will approach $0$ as time
1846: goes on.
1847:
1848: But, if we relax this assumption and imagine finite barriers
1849: (associated with local nucleation of a native
1850: phase\cite{BW90,TakWol97,PTW01}), we may ask (informally) when
1851: activated motion over the barrier to nucleation is least hindered. We
1852: argue that the free energy landscape features present globally in our
1853: mean-field picture will also be relevant locally: when our calculation
1854: here finds a glassy US state, we expect that free energy surface near
1855: the true transition state will also be rough. Thus, from the
1856: preceding description of the phases and the transition states between
1857: them, we can see that folding should be easiest for large $\beta_0$ in
1858: a window between $T_1(\beta_0)$ and the upper boundary of the region
1859: where the US states become glassy (and passage across the transition
1860: region is kinetically impeded by the tortuous nature of the local free
1861: energy landscape). The latter boundary lies, in turn, just barely
1862: above $T_g$, where the landscape in the (large) portion of the
1863: configuration space uncorrelated with the native state also becomes
1864: rough, further impeding escape from it. At still lower temperatures,
1865: things become even worse, first with the onset of glassiness in the
1866: native-like region of configuration space itself and finally with the
1867: disappearance of native-like solutions. But these features probably
1868: have minor consequences, since folding will already have been so
1869: strongly impeded by the effects (with onset near $T_g$) that tend to
1870: confine it in a region of configuration space uncorrelated with the
1871: native state.
1872:
1873:
1874:
1875: \section{Discussion}
1876:
1877: We have introduced what we might call a generic model for a protein,
1878: based on what seems to us to be the simplest way to incorporate a
1879: tendency to form a native state in an otherwise random heteropolymer
1880: model. To make it possible to calculate typical properties, we follow
1881: previous authors \cite{PGT1,RS,PGT2,WS} and do not specify a
1882: particular native state, but rather an ensemble of them, constrained
1883: only by chain-entropic constraints and confinement to the appropriate
1884: volume. This ensemble is characterized by the selection temperature
1885: $T_0$. Our model differs from previous ones in that they are based on
1886: random-sequence heteropolymers, while we start from a model
1887: \cite{SG1,SG2} in which each monomer-monomer interaction is an
1888: independent random variable.
1889:
1890: While it might be argued that random-sequence models are more relevant
1891: to proteins, they approach the model we consider here in the limit
1892: where the number of monomer types becomes large. Thus, what we find
1893: out about our model may be relevant to proteins (with 20 different
1894: amino acids). Of course, it is also important to study what happens
1895: away from the large-monomer-type limit; our analysis here can help in
1896: solving that more difficult problem.
1897:
1898: Furthermore, naively, one might assume that by adjusting $N(N-1)/2$
1899: parameters one could imprint a native state more strongly than for
1900: models with only $N$ parameters. Our model shows that this is not
1901: necessarily true. Parts of the phase diagram are glassy, even for
1902: very low selection temperature $T_0$, when the native state should be
1903: strongly imprinted into the model.
1904:
1905: Instead of the quadratic confinement term $\mu x(s,t)^2$ one could add
1906: three-body terms, which are commonly used to fix the globule
1907: density. It would be interesting to extend the analysis presented here
1908: to such models. Also, in our treatment, translational invariance
1909: within the globule is put in by hand. Keeping three-body terms would
1910: lead to automatic translational invariance. We have seen that if
1911: translational averaging is omitted (see paper I) then the equations
1912: become coupled in the $k$ variable and are thus a lot harder to solve.
1913:
1914: Within out model, we have made just two approximations: the Gaussian
1915: variational ansatz of section VI, and the assumption of 1-step
1916: ergodicity breaking (analogous to 1-step replica symmetry breaking in
1917: the replica approach). Otherwise the solution is complete and exact
1918: to the accuracy we were able to achieve numerically.
1919:
1920: Our most important result is the existence of the various different
1921: phases at large $\Bt/T_0$, where the interactions that favor a native
1922: state are strong. While it is natural to anticipate that the
1923: native-like configurations will be thermally disrupted above a
1924: temperature of order $\Bt^2/T_0$, it is not so obvious that at low
1925: temperatures there will be other impediments to efficient folding. We
1926: identify two of these:
1927:
1928: (1) The frozen-globule state, which is uncorrelated with the native
1929: state, always exists below $T_g$, no matter how big $\beta_0$ is.
1930: This means that in a large part of configuration space, the system may
1931: be trapped in a rough energy landscape and never (in MFT) get to the
1932: native-state region where it can fold rapidly. Furthermore, in almost
1933: the same temperature, range, we expect that the energy landscape is
1934: also rough around the transition region on the way to the
1935: correctly-folded state, further impeding the folding process. Thus,
1936: while lower temperature favors well-folded over random-globule-like
1937: configurations energetically, the rough energy landscape of the glassy
1938: phase will hinder correct folding. Our conclusion here is consistent
1939: with that of Goldstein {\em et al.}\cite{GLW92}, who found (albeit in
1940: a different kind of model) that a large $T_n/T_g$ (or $T_1/T_g$) ratio
1941: favors folding.
1942:
1943: (2) At even lower temperatures, the native state itself is
1944: unstable against a glass transition where it splits into a large
1945: number of substates. Transitions between these substates are
1946: blocked by high barriers (infinite, in MFT). A phase of this kind
1947: was found earlier by Bryngelson and Wolynes in a phenomenological
1948: model\cite{BW87}. It is tempting to associate the substates with
1949: the glassy conformations observed at temperatures below ~200 K in
1950: myoglobin \cite{Ibenetal}.
1951:
1952: Of course, MFT is an approximation. The escape from the tortuous part
1953: of the energy landscape to the smooth region will not take forever,
1954: nor will transitions between low-$T$ substates. Nevertheless, MFT
1955: does indicate when we can expect relaxational dynamics, including
1956: folding, to be slow or fast, as well as give us some insight into the
1957: physics behind these differences.
1958:
1959: Our analysis here is a purely dynamical one. We do not compute
1960: equilibrium partition functions. A complete analysis would include
1961: such calculations, but we defer them to future work. Nevertheless,
1962: the purely dynamical analysis can reveal important properties of the
1963: system that cannot be seen in an equilibrium analysis. For example,
1964: it has been know for a long time that for a large class of models ---
1965: namely, those which have a glass transition where $x \rightarrow 1$
1966: --- the dynamic and static glass transition temperatures are different
1967: \cite{CK1,CS,CHS}. This is expected to be the case for the transition
1968: at $T_g$ in our model: The equilibrium glass transition temperature is
1969: lower than the dynamical one. Thus, in a temperature range just below
1970: the dynamical $T_g$, the equilibrium analysis does not reveal the slow
1971: dynamics (accompanied by aging) that we are able to identify and
1972: analyze here.
1973:
1974: In other glassy models for which it has been possible to do a more
1975: complete analysis \cite{CKD,CD,CK1,CS,CHS}, the static and dynamic
1976: transitions coincide when they occur as a result of an
1977: Almeida-Thouless instability (the marginal stability condition,
1978: Eqn~\ref{MSC}) \cite{AT}. This is the case here at the phase boundary
1979: where the native-like state becomes glassy.
1980:
1981: Gillin and Sherrington \cite{GS} and Gillin {\em et al.} \cite{GNS}
1982: have been able to analyze both the statics and the dynamics of several
1983: classes of mean-field spin glass models with a competition between
1984: glassy and ferromagnetic states (see also ref.~\cite{HSN} for a
1985: special case). Some features of the phase diagram of our model that
1986: we have been able to discover so far are also seen in their models.
1987:
1988: Gillin {\em et al.} also studied full (as well as 1-step) replica
1989: symmetry breaking solutions, which we have not. In some of their
1990: models, the counterpart of our glassy native phase undergoes full RSB
1991: at low temperatures, and the counterpart of our II-VI boundary becomes
1992: vertical. It is possible in our model as well that, in particular
1993: regions of the phase diagram, our 1-step solutions are not stable and
1994: full RSB is necessary. More generally, it will be an interesting
1995: problem to try to explore what kinds and degrees of universality there
1996: are in the phase structures of various systems where the glassiness
1997: induced by disorder competes with some kind of order analogous to the
1998: native state in our problem or the ferromagnetic state in theirs.
1999:
2000:
2001:
2002:
2003: \begin{acknowledgements}
2004: It is a pleasure to thank Silvio Franz for discussions leading to our
2005: formulation of this problem.
2006: \end{acknowledgements}
2007:
2008: \appendix
2009:
2010: \section{correlation function $G_{10}^{ss'}$ }
2011:
2012: Here we derive Eq.~(\ref{G10a}). Inserting Eq.~(\ref{Phis1}) for
2013: the superfield into (\ref{G10}) gives
2014: %
2015: \begin{eqnarray}
2016: G_{10}^{ss'}
2017: & = & \langle x(s,t_1) x_0(s') \rangle
2018: + \langle \bar\eta(s,t_1) x_0(s') \rangle \theta_1 + \nonumber \\
2019: & & + \bar\theta_1 \langle \eta(s,t_1) x_0(s') \rangle
2020: + \bar\theta_1\theta_1 \langle \tilde x(s,t_1) x_0(s')
2021: \rangle .
2022: \label{G10b}
2023: \end{eqnarray}
2024: %
2025: One can show that the action of the dynamical generating
2026: functional $F_{dyn}$ (see Eq.~\ref{Fdyn}) is invariant under the
2027: infinitesimal transformation (BRS symmetry)
2028: %
2029: \begin{equation}
2030: \delta \Phi(s,t,\theta,\bar\theta) = \epsilon
2031: \frac{\partial}{\partial\bar\theta} \Phi(s,t,\theta,\bar\theta)
2032: \label{deltaPhi}.
2033: \end{equation}
2034: %
2035: This follows in two steps. First one notices that for any function
2036: $f$
2037: %
2038: \begin{equation}
2039: \delta\int d\bar\theta f(\Phi(\bar\theta)) =
2040: \epsilon \int d\bar\theta \frac{\partial}{\partial\bar\theta}
2041: f(\Phi(\bar\theta)) = 0,
2042: \end{equation}
2043: %
2044: due to the identity $\int d\bar\theta
2045: \frac{\partial}{\partial\bar\theta} = 0$. This means that any term
2046: involving a local function of $\Phi$ (i.e., not containing
2047: derivatives over $\theta$ and $\bar\theta$), e.g., $S_2[\Phi,x_0]$
2048: (see Eq.~\ref{eq:S2}), is invariant under the transformation
2049: (\ref{deltaPhi}). $S_0[x_0]$ is trivially invariant since it does
2050: not contain the superfield $\Phi$. (The same reasoning holds for
2051: a transformation like (\ref{deltaPhi}) with a derivative with
2052: respect to $\theta$ instead of with respect to $\bar\theta$.) The
2053: only term left is the $S_1[\Phi]$ (i.e., the part of the action
2054: quadratic in the superfield) and it is straightforward to see that
2055: this term is also invariant under (\ref{deltaPhi}) (though not
2056: under the transformation involving the derivative with respect to
2057: $\theta$).
2058:
2059: The fact that the action is invariant under (\ref{deltaPhi})
2060: implies the Ward identity
2061: %
2062: \begin{equation}
2063: \frac{\partial}{\partial\bar\theta_1} G_{12}^{ss'} = 0
2064: \label{ward1}
2065: \end{equation}
2066: %
2067: which gives $0=\langle \eta x_0 \rangle + \theta_1 \langle \tilde
2068: x x_0 \rangle$ (we have suppressed the arguments of the fields to
2069: simplify the notation). This implies that separately one has
2070: %
2071: \begin{equation}
2072: \langle \eta x_0 \rangle=0, \ \ \langle \tilde x x_0 \rangle=0
2073: \label{ward2}
2074: \end{equation}
2075: %
2076: Inserting (\ref{ward2}) into (\ref{G10b}) gives the desired
2077: result, Eq.~(\ref{G10a}).
2078:
2079: \section{Details on the GVA and how to improve it}
2080:
2081: Here we give more background on the use of Eq.~(\ref{dFdG}). In
2082: the dynamical calculation, the fields are complex and contain
2083: Grassmann variables; thus, $F_{dyn}$ is not a real number. This
2084: means that any interpretation of Eq.~(\ref{dFdG}) as an extremum
2085: condition for $F_{dyn}$ has to be given up. Nevertheless, we can
2086: still make some sense of the GVA as the first step in a systematic
2087: approximation scheme.
2088:
2089: Formally, one starts from Eq.~(\ref{Fdyn}), which we rewrite in the
2090: shorter form
2091: %
2092: \begin{equation}
2093: e^{-F_{dyn}}=\int D\Psi e^{ -S[\Psi] }.
2094: \label{Fdyn3}
2095: \end{equation}
2096: %
2097: where $\Psi$ stands for the pair $(x_0,\Phi)$ and, likewise,
2098: $D\Psi$ for $Dx_0D\Phi$. One can express $F_{dyn}$ in a slightly
2099: different form
2100: %
2101: \begin{equation}
2102: e^{-F_{dyn}} = \langle e^{-(S-S_{var})} \rangle_{var} e^{-F_{var}}
2103: \label{Fdyn4}
2104: \end{equation}
2105: %
2106: where
2107: %
2108: \begin{equation}
2109: e^{-F_{var}}=\int D\Psi e^{-S_{var}}
2110: \end{equation}
2111: %
2112: and
2113: %
2114: \begin{equation}
2115: \langle (...) \rangle_{var}=
2116: \frac{
2117: \int D\Psi ( ... ) e^{-S_{var}}
2118: }{
2119: \int D\Psi e^{-S_{var}}
2120: }
2121: \end{equation}
2122: %
2123: In a static calculation one would proceed with the inequality
2124: %
2125: \begin{equation}
2126: e^{-F} \ge e^{ - \langle (S-S_{var}) \rangle_{var} }
2127: e^{ -F_{var} }
2128: \label{Fdyn5}
2129: \end{equation}
2130: %
2131: to conclude that
2132: %
2133: \begin{equation}
2134: F \le \langle S-S_{var} \rangle_{var} + F_{var}
2135: \label{Fdyn6}.
2136: \end{equation}
2137: %
2138: Thus, in a static calculation, the variationally-obtained $F$
2139: gives an upper bound on the true $F$. What is allowed to vary is
2140: the form of $S_{var}$, most often, the parameters describing it.
2141: (In the GVA, $S_{var}$ is specified by $G_{12}^{ss'}$ and
2142: $G_{10}^{ss'}$.)
2143:
2144: In the dynamical problem we follow another route, starting exactly
2145: at the problematic step, Eq.~(\ref{Fdyn5}), along the lines of
2146: ref.~\cite{Kleinert}. Instead of the inequality (\ref{Fdyn6}) we
2147: use Eq.~(\ref{Fdyn4}) in a slightly modified form
2148: %
2149: \begin{equation}
2150: F_{dyn} = F_{var}
2151: - \ln \langle e^{-\Delta S} \rangle_{var}
2152: \label{Fdyn7},
2153: \end{equation}
2154: %
2155: where $\Delta S=S-S_{var}$. Applying a cumulant expansion
2156: %
2157: \begin{eqnarray}
2158: \langle {\rm exp}(-\Delta S)\rangle_{var} & & =
2159: {\rm exp}\left[ -\langle \Delta S \rangle_{var} + \right. \nonumber \\
2160: & & \left. + \frac{1}{2} \left(
2161: \langle \Delta S^2 \rangle_{var}
2162: - \langle \Delta S \rangle_{var}^2
2163: \right)
2164: + \cdots \right],
2165: \label{Fdyn8}
2166: \end{eqnarray}
2167: %
2168: one gets
2169: %
2170: \begin{equation}
2171: F_{dyn} = F_{var} + \langle S-S_{var} \rangle_{var} + \Delta F
2172: \end{equation}
2173: %
2174: where $\Delta F$ contains second and higher order corrections in
2175: $\Delta S$. In any approximation made by keeping a finite number
2176: of terms in (\ref{Fdyn8}) (the simplest being to set $\Delta
2177: F=0$), $F_{dyn}$ depends on $G_{12}^{ss'}$. To minimize this
2178: dependence, we chose $G_{12}^{ss'}$ so that the derivative of the
2179: approximate form for $F_{dyn}$ with respect to $G_{12}^{ss'}$
2180: vanishes. This gives Eq.~(\ref{dFdG}). Furthermore, if all terms
2181: in $\Delta F$ are kept, this procedure, by construction, formally
2182: gives back the exact $F_{dyn}$.
2183:
2184: The meaning of minimizing the dependence with respect to
2185: quantities involving Grassmann numbers may seem obscure, but we
2186: note that we are using the SUSY representation only for
2187: compactness. The entire GVA calculation could have been presented
2188: equivalently without any Grassmann variables, with no change in
2189: meaning or result. Thus, we are really only minimizing the
2190: dependence on parameters of physically well-defined correlation
2191: and response functions.
2192:
2193: %\bibliographystyle{unsrt}
2194: %\bibliography{refs}
2195:
2196: \begin{thebibliography}{10}
2197:
2198: \bibitem{Boy}
2199: B. Rensberger, {\em Life itself: exploring the realm of the living cell},
2200: (Oxford University Press, 1996).
2201:
2202: \bibitem{PGT1}
2203: V.S. Pande, A.Y. Grosberg and T. Tanaka, Rev. Mod. Phys. {\bf 72}, 259-314
2204: (2000).
2205:
2206: \bibitem{Wol1}
2207: P.G. Wolynes, H. Frauenfelder and R.H. Austin, {\em More things in heaven and
2208: earth. A celebration of physics at the millennium}, (Springer, 1999), page
2209: 706-725.
2210:
2211: \bibitem{WE}
2212: P.G. Wolynes and W.A. Eaton, Physics World {\bf 9}, 39 (1999).
2213:
2214: \bibitem{Creig}
2215: M. Karplus and E. Shakhnovich, in {\em Protein Folding}, edited by T.E.
2216: Creighton, ( Freeman, New York, 1992), page 127.
2217:
2218: \bibitem{SG1}
2219: E. I. Shakhnovich and A. M. Gutin, Europhys. Lett. {\bf 8}, 327 (1989).
2220:
2221: \bibitem{SG2}
2222: E. I. Shakhnovich and A. M. Gutin, J. Phys. A {\bf 22}, 1647 (1989).
2223:
2224: \bibitem{GHLO}
2225: T. Garel, D.A. Huse, L. Leibler, H. Orland, Europhys. Lett. {\bf 6}, 307
2226: (1988).
2227:
2228: \bibitem{GOP}
2229: T. Garel, H. Orland and E. Pitard, cond-mat/9706125.
2230:
2231: \bibitem{GLO}
2232: T. Garel, L. Leibler and H. Orland, J. Phys. II (France) {\bf 4}, 2139 (1994).
2233:
2234: \bibitem{TW}
2235: S. Takada and P.G. Wolynes, Phys. Rev. E {\bf 55}, 4562 (1997).
2236:
2237: \bibitem{SGS}
2238: C.D. Sfatos, A.M. Gutin and E.I. Shakhnovich, Phys. Rev. E {\bf 48}, 465-475
2239: (1993).
2240:
2241: \bibitem{SW}
2242: M. Sasai and P.G. Wolynes, Phys. Rev. Lett. {\bf 65}, 2740 (1990).
2243:
2244: \bibitem{LT}
2245: N. Lee and D. Thirumalai, J. Chem. Phys. {\bf 113}, 5126-5129 (2000).
2246:
2247: \bibitem{Pit}
2248: E. Pitard, Eur. Phys. J B {\bf 7}, 665-673 (1999).
2249:
2250: \bibitem{TPW}
2251: S. Takada, J.J. Portman and P.G. Wolynes, Proc. Natl. Acad. Sci. USA {\bf 94},
2252: 2318 (1997).
2253:
2254: \bibitem{TAB}
2255: D. Thirumalai, V. Ashwin and J.K. Bhattacharjee, Phys. Rev. Lett. {\bf 77},
2256: 5385-5388 (1996).
2257:
2258: \bibitem{PS}
2259: E. Pitard and E.I. Shakhnovich, Phys. Rev. E {\bf 63}, 041501 (2001).
2260:
2261: \bibitem{Olem1}
2262: A.I.Olemskoi, Physica A {\bf 270}, 444-452 (1999).
2263:
2264: \bibitem{Olem2}
2265: A.I.Olemskoi, V.A.Brazhnyi, Physics of the Solid State {\bf 43}, 386-396
2266: (2001).
2267:
2268: \bibitem{SpGl}
2269: M. Mezard, G. Parisi and M.A. Virasoro, {\em Spin Glass Theory and Beyond}
2270: (World Scientific, 1987).
2271:
2272: \bibitem{KHS}
2273: Z. Konkoli, J. Hertz and S. Franz, Phys. Rev. E {\bf 64}, 051910
2274: (2001).
2275:
2276: \bibitem{Wol2}
2277: J.D. Bryngelson and P.G. Wolynes, Proc. Natl. Acad. Sci. USA {\bf 84}, 7524
2278: (1987).
2279:
2280: \bibitem{RS}
2281: S. Ramanathan and E. Shakhnovich, Phys Rev E {\bf 50A}, 1303-1312 (1994).
2282:
2283: \bibitem{PGT2}
2284: V.S. Pande, A.Y. Grosberg and T. Tanaka, J. Phys. II {\bf 4}, 1771-1784 (1994).
2285:
2286: \bibitem{WS}
2287: J. Wilder and E.I. Shakhnovich, Phys. Rev. E {\bf 62B}, 7100-7110 (2000).
2288:
2289: \bibitem{MSR}
2290: P.C. Martin, E.D. Siggia and H.A. Rose, Phys. Rev. A {\bf 8}, 423
2291: (1973).
2292:
2293: \bibitem{Dom}
2294: C. De Dominicis, Phys. Rev. B {\bf 18}, 4913 (1978).
2295:
2296: \bibitem{Jans1} H. K. Janssen, {\em On the Renormalized Field Theory
2297: of Nonlinear Critical Relaxation}, in {\em From Phase Transitions to
2298: Chaos}, page 68--91, year 1992, eds. Gy{\"o}rgyi and I. Kondor and
2299: L. Sasv{\'a}ri and T. Tel, World Scientific, Singapore
2300:
2301: \bibitem{Jans2} H. K. Janssen, {\em Field-theoretic method applied to
2302: critical dynamics}, in {\em Dynamical Critical Phenomena and Related
2303: Topics}, page 25--47, year 1979, ed. C. P. Enz, Springer-Verlag,
2304: Berlin
2305:
2306: \bibitem{Kur}
2307: J. Kurchan, J. Phys. I (France) {\bf 2}, 1333 (1992).
2308:
2309: \bibitem{MP1}
2310: M. M\'{e}zard and G. Parisi, J. Phys. I {\bf 1}, 809 (1991).
2311:
2312: \bibitem{MP2}
2313: M. M\'{e}zard and G. Parisi, J. Phys. I {\bf 2}, 2231 (1992).
2314:
2315: \bibitem{CKD}
2316: L.F. Cugliandolo, J. Kurchan and P. Le Doussal, Phys. Rev. Lett. {\bf 76}, 2390
2317: (1996).
2318:
2319: \bibitem{CD}
2320: L.F. Cugliandolo and P. Le Doussal, Phys. Rev. E {\bf 53}, 1525 (1996).
2321:
2322: \bibitem{Go1}
2323: N. Go, Biopolymers {\bf 20}, 991-1011 (1981).
2324:
2325: \bibitem{Go2}
2326: Y. Ueda, H. Taketomi and N. Go, Int. J. Peptide. Res. {7}, 445 (1975).
2327:
2328: \bibitem{Olem3}
2329: A.I.Olemskoi, V.A.Brazhnyi, Physica A {\bf 273}, 368-400 (1999).
2330:
2331: \bibitem{CK1}
2332: L.F. Cugliandolo and J. Kurchan, Phys. Rev. Lett. {\bf 71}, 173 (1993), J.
2333: Phys. A {\bf 27}, 5749 (1994).
2334:
2335: \bibitem{BCKP}
2336: A. Baldassarri, L.F. Cugliandolo, J. Kurchan and G. Parisi, J. Phys. A {\bf
2337: 27}, 5749 (1994).
2338:
2339: \bibitem{CK2}
2340: L.F. Cugliandolo and J. Kurchan, Phil. Mag. B {\bf 71}, 501 (1995).
2341:
2342: \bibitem{CS}
2343: A. Crisanti and H.-J. Sommers, Z. Phys. B {\bf 87}, 341 (1992).
2344:
2345: \bibitem{CHS}
2346: A. Crisanti, H. Horner and H.-J. Sommers, Z. Phys. B {\bf 92}, 257 (1993).
2347:
2348: \bibitem{BW90}
2349: J.D. Bryngelson and P.G. Wolynes, Biopolymers {\bf 30}, 177 (1990).
2350:
2351: \bibitem{TakWol97}
2352: S. Takada and P.G. Wolynes, J. Chem. Phys. {\bf 107}, 9585 (1997).
2353:
2354: \bibitem{PTW01}
2355: J.J. Portman, S. Takada and P.G. Wolynes, J. Chem. Phys. {\bf 114}, 5069, 5082
2356: (2001).
2357:
2358: \bibitem{GLW92}
2359: R.A. Goldstein, Z.A. Luthey-Schulten and P.G. Wolynes, Proc. Nat. Acad. Sci.
2360: USA {\bf 89}, 4918 (1992).
2361:
2362: \bibitem{BW87}
2363: J.D. Bryngelson and P.G. Wolynes, Proc. Nat Acad. Sci. USA {\bf 84}, 7524
2364: (1987).
2365:
2366: \bibitem{Ibenetal}
2367: I.E.T. Iben, D. Braunstein, W. Doster, H. Frauenfelder, M.K. Hong, J.B.
2368: Johnson, S. Luck, P. Ormos, A. Schulte, P.J. Steinbach, A.H. Xie, and R.D.
2369: Young, Phys. Rev. Lett. {\bf 62}, 1916 (1989).
2370:
2371: \bibitem{AT}
2372: J.R.L. de Almeida and D.J. Thouless, J. Phys. A {\bf 11}, 983 (1978).
2373:
2374: \bibitem{GS}
2375: P. Gillin and D. Sherrington, J. Phys A {\bf 33}, 3081 (2000).
2376:
2377: \bibitem{GNS}
2378: P. Gillin, H. Nishimori and D. Sherrington, J. Phys. A {\bf 34}, 2949 (2001).
2379:
2380: \bibitem{HSN}
2381: J.A. Hertz, D. Sherrington and Th.M. Nieuwenhuizen, Phys. Rev. E {\bf 60}, 2460
2382: (1998).
2383:
2384: \bibitem{Kleinert} H. Kleinert, {\em Path integrals in quantum
2385: mechanics, statistics and polymer physics}, (World Scientific, 1995).
2386:
2387: \end{thebibliography}
2388:
2389: \newpage
2390:
2391: \begin{figure}
2392: %%%%%%%%%%%%%%%%%%%% will be taken away later BEGIN %%%%%%%%%%%%%%%%%%
2393: \epsfxsize=8cm
2394: \epsfbox{fig1.eps}
2395: %%%%%%%%%%%%%%%%%%%% will be taken away later END %%%%%%%%%%%%%%%%%%
2396: \caption{Boundary of the glassy phase in the $(\tmu, T)$ plane.
2397: The boundary is same as in the case of the random-heteropolymer
2398: model from paper I, except that native-state correlations lead to
2399: the replacement of $\mu$ by $\tilde\mu$. We have used parameters
2400: $d=3$ and $\sigma=1$. $T_{max}$ is the maximum $T$ for which
2401: Eq.~(\ref{MSC}) has a solution (see Paper I for further details).
2402: $\mu_c$ is the value of $\tmu$ where $T(\tilde\mu)$ attains this
2403: maximum. The solid part of the boundary is an AT line, and the
2404: dash-dotted part marks a transition where $x \rightarrow 1$.
2405: Approaching the AT line from below, $b-b_0 \rightarrow 0$, while
2406: $x$ remains strictly less than $1$. Approaching the $x=1$ line
2407: from below, $x \rightarrow 1$ smoothly, while $b-b_0$ is
2408: discontinuous there. Above both lines, $b=b_0$ and $x$ is
2409: undetermined (any $x \neq 0$ solves (74)-(80), and no physical
2410: quantity depends on it). The same holds for all figures where
2411: these lines appear.}
2412: \end{figure}
2413:
2414:
2415:
2416: \begin{figure}
2417: %%%%%%%%%%%%%%%%%%%% will be taken away later BEGIN %%%%%%%%%%%%%%%%%%
2418: \epsfxsize=9cm
2419: \epsfbox{fig2.eps}
2420: %%%%%%%%%%%%%%%%%%%% will be taken away later END %%%%%%%%%%%%%%%%%%
2421: \caption{Analysis of ergodic-native solutions: $\beta_0(\tmu,T)$ as
2422: function of $\tmu$ for four values of $T$ (see text for explanation).
2423: Panel (d) shows the existence of the two extra solutions (one stable,
2424: the other unstable) in the range $[\beta_1^{\rm min}, \beta_1^{\rm
2425: max}]$. }
2426: \end{figure}
2427:
2428:
2429:
2430: \begin{figure}
2431: %%%%%%%%%%%%%%%%%%%% will be taken away later BEGIN %%%%%%%%%%%%%%%%%%
2432: \epsfxsize=8cm
2433: \epsfbox{fig3.eps}
2434: %%%%%%%%%%%%%%%%%%%% will be taken away later END %%%%%%%%%%%%%%%%%%
2435: \caption{Regions of existence and stability of ergodic-native solutions.
2436: Panel (a): Ergodic-native-phase and US solutions exist everywhere to
2437: the right of the thick solid curve. The ergodic native phase is
2438: stable against glassiness everywhere there except in the diagonally
2439: cross-hatched region. The US states are also unstable against
2440: glassiness there, and additionally in the horizontally cross-hatched
2441: region. The vertical cross-hatching marks the region where the extra
2442: phase seen in Panel (d) of Fig.~2 is found. (This phase is never
2443: stable against glassiness.) Panel (b): Enlargement of the circled
2444: region in Panel (a). The AT line is tangent to the ergodic phase
2445: boundary (thick line). At its maximum, at $T_{max}$, it becomes an
2446: $x=1$ line (dashed-dotted line, see also Fig.~1). Lowering $T$ from
2447: the white region into the horizontally cross-hatched region results in
2448: two different types of transitions depending on whether one crosses
2449: the AT or the $x=1$ line. In both cases the US state becomes glassy.
2450: }
2451:
2452: \end{figure}
2453:
2454:
2455:
2456: \begin{figure}
2457: %%%%%%%%%%%%%%%%%%%% will be taken away later BEGIN %%%%%%%%%%%%%%%%%%
2458: \epsfxsize=8cm
2459: \epsfbox{fig4.eps}
2460: %%%%%%%%%%%%%%%%%%%% will be taken away later END %%%%%%%%%%%%%%%%%%
2461: \caption{Analysis of glassy-native solutions: $\beta_0(\tmu,T)$ as a
2462: function of $\tmu$ for three fixed values of $T$ (see text for
2463: explanation). Full lines: $\beta_0(\tmu,T)$ calculated within the
2464: ergodic ansatz (as in Fig.~2). Dashed lines: $\beta_0(\tmu,T)$
2465: calculated with the glassy ansatz. The actual curves vary with $\tmu$
2466: in a way that is difficult to plot in a useful way, so here we have
2467: distorted them in such a way as to make their qualitative form (number
2468: and ordering of maxima and minima) evident. When the two curves cross
2469: at $\tmu=\tmu_{AT}$, one has to change from the ergodic to the glassy
2470: solution (when approaching from $\tmu=\infty$). Similarly, when
2471: $\tmu=\tmu_{min}$ one has to go back to the ergodic solution. The
2472: thick dashed line indicates the physically relevant states (both
2473: stable phases and US states). }
2474: \end{figure}
2475:
2476:
2477:
2478: \begin{figure}
2479: %%%%%%%%%%%%%%%%%%%% will be taken away later BEGIN %%%%%%%%%%%%%%%%%%
2480: \epsfxsize=9cm
2481: \epsfbox{fig5.eps}
2482: %%%%%%%%%%%%%%%%%%%% will be taken away later END %%%%%%%%%%%%%%%%%%
2483: \caption{The final phase diagram.
2484: Panel (a): Stable phases.
2485: Region I: Random globule is the only stable phase.
2486: Region II: Frozen globule is the only stable phase.
2487: Region III: Ergodic native and random globule phases stable.
2488: Region IV: Ergodic native and frozen globule phases stable.
2489: Region V (only visible in panel (b)): Glassy
2490: native and random globule stable.
2491: Region VI: Glassy native and frozen globule stable. The dashed line
2492: marks the boundary of the (unphysical) ergodic native state from
2493: Fig.~3, to emphasize that the phase boundary of the glassy native
2494: state (solid) does not coincide with it.
2495: Panel (b): Enlargement showing structure in the region near $T =
2496: T_{max} \approx T_g$ and $\beta_0 = 1.45$ (including region V). Below
2497: point B the boundary of region III is given by the AT line. Above
2498: point B the boundary is the ergodic-native stability limit (the
2499: uppermost line in Panel (a)). The continuation of the AT lie is shown
2500: as a dotted line (which turns into dash-dot $x=1$ line).
2501: Panel (c): US states are ergodic in the vertically hatched region,
2502: glassy in the horizontally hatched region. The boundary above and to
2503: the right of point A is an AT line.
2504: Beyond the region shown the $x=1$ line falls
2505: off monotonically, and for $\beta_0 \rightarrow \infty$ it approaches
2506: $T_g$. Below point A, the small-$\beta_0$ boundary coincides with the
2507: line between regions II and VI in panels (a) and (b). }
2508: \end{figure}
2509:
2510:
2511:
2512: \begin{figure}
2513: %%%%%%%%%%%%%%%%%%%% will be taken away later BEGIN %%%%%%%%%%%%%%%%%%
2514: \epsfxsize=8cm
2515: \epsfbox{fig6.eps}
2516: %%%%%%%%%%%%%%%%%%%% will be taken away later END %%%%%%%%%%%%%%%%%%
2517: \caption{Schematic free energy surfaces in different regions of the
2518: phase diagram. (See text for explanation.) (a): region I. (b):
2519: Region II. (c): Region III. (d): Region IV. (e): Region V. (f):
2520: Region VI. }
2521: \end{figure}
2522:
2523:
2524: \end{multicols}
2525:
2526: \end{document}
2527: