q-bio0409010/paper.tex
1: \documentclass[aps,pre,twocolumn,final,floatfix,showpacs,amsmath,amssymb,10pt]{revtex4}
2: 
3: 
4: % RNA sequence stuff definitions
5: \newcommand*{\Base}[1][i]{\ensuremath{r_{#1}}}
6: \newcommand*{\Sequence}{\ensuremath{\mathcal{R}}}
7: \newcommand*{\Structure}{\ensuremath{\mathcal{S}}}
8: \newcommand*{\epsStructure}{\ensuremath{\Structure_\varepsilon}}
9: \newcommand*{\SM}[1][i,j]{\ensuremath{S_{#1}}}
10: \newcommand*{\Ensemble}{\ensuremath{\mathcal{E}}}
11: \newcommand*{\PairE}{\ensuremath{E_p}} % pair energy
12: \newcommand*{\StackE}{\ensuremath{E_s}}% stacking energy
13: \newcommand*{\randE}{\ensuremath{\eta}}% random field energy
14: %
15: \newcommand*{\N}{\ensuremath{N}}
16: \newcommand*{\Nh}{\ensuremath{\hat N}}
17: \newcommand*{\Z}{\ensuremath{Z}}
18: \newcommand*{\Zh}{\ensuremath{\hat Z}}
19: \newcommand*{\kb}{\ensuremath{k_\mathrm{B}}}
20: \newcommand*{\kbT}{\ensuremath{k_\mathrm{B}T}}
21: \newcommand*{\overlap}{\ensuremath{q}}
22: \newcommand*{\distance}{\ensuremath{d}}
23: \newcommand*{\Ssize}{\ensuremath{t}} % average stacking size
24: \newcommand*{\BC}{\ensuremath{B}} % Binder Cumulant
25: \newcommand*{\PC}{\ensuremath{A}} % Parisi Cumulant
26:                                 
27: \newcommand*{\average}[1]{\ensuremath{\left\langle #1\right\rangle}}
28: \newcommand*{\qaverage}[1]{\ensuremath{\left[ #1\right]}}
29: \newcommand*{\D}{\,\mathrm{d}}
30: \newcommand*{\define}{:=}
31: \newcommand*{\E}{\mathrm{e}}
32: \newcommand*{\vek}[1]{\boldsymbol{#1}}
33: \newcommand*{\order}[1]{\ensuremath{\mathcal{O}(#1)}}
34: 
35: \newcommand*{\qvariance}{\chi_{\Sequence}}
36: 
37: \usepackage{xspace}
38: \newcommand*{\eg}{e.g.,\xspace} % z.B.
39: \newcommand*{\ie}{i.e.,\xspace} % d.h.
40: \newcommand*{\refeq}[1]{eq.\@ \eqref{#1}\xspace}
41: \newcommand*{\reffig}[1]{Fig.\@ \ref{#1}\xspace}
42: \newcommand*{\reftab}[1]{Tab.\@ \ref{#1}\xspace}
43: \newcommand*{\refsec}[1]{section\@ \ref{#1}\xspace}
44: \newcommand*\epsCoup{$\varepsilon$-coupling\xspace}
45: 
46: \renewcommand\atop[2]{\genfrac{}{}{0pt}{}{#1}{#2}}
47: 
48: 
49: 
50: \usepackage{graphicx}
51: \newcommand*\picturewidth{1.0\columnwidth}
52: 
53: \newcommand*\guidelines{Calculated data points are indicated by symbols. Lines
54:   are  drawn to guide the eye.} 
55: 
56: \newcommand*\errorbars{Missing error bars are of the size of the symbols or smaller,  and omitted for legibility.}
57: 
58: 
59: \newcommand*\legendsymbols{%
60: \label{legendsymbols}% use only once
61: For system size $L$ the following symbols are used: 
62: $\circ$~128, $\triangle$~192, $\square$~256, $\triangledown$~384, 
63: $\Diamond$~512, $\triangleleft$~768, $\blacksquare$~1024.
64: \guidelines\ \errorbars
65: }
66: \newcommand*\legendref{For explanation of symbols see caption of
67:   \reffig{legendsymbols}.\xspace}
68: 
69: 
70: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71: \begin{document}
72: 
73: \author{Bernd Burghardt}
74: \email{burghard@theorie.physik.uni-goettingen.de}
75: \author{Alexander K. Hartmann}
76: \email{hartmann@theorie.physik.uni-goettingen.de}
77: \affiliation{Institut f{\"u}r Theoretische Physik, Universit{\"a}t G{\"o}ttingen,
78:   Friedrich-Hund-Platz 1, D--37077 G{\"o}ttingen, Germany}
79: 
80: \title{Dependence of RNA secondary structure on the energy model}
81: \date{\today}
82: \begin{abstract}
83: We analyze a microscopic RNA model, which includes two widely used models as
84: limiting cases,   namely it contains terms for bond as
85:   well as for stacking energies.
86: We numerically investigate possible changes in the qualitative and quantitative
87: behaviour while going from one model to the other; 
88: in particular we test, whether a transition occurs, when continuously
89: moving from one model to the other. 
90: For this we calculate various thermodynamic quantities, both at zero
91: temperature as well as at finite temperatures.
92:    All calculations can be done efficiently in polynomial time by a dynamic
93: programming algorithm.  
94: We do not find a sign for  transition between the models, 
95: but the critical
96: exponent $\nu$ of the correlation length, 
97: describing the phase transition in all models to an
98: ordered low-temperature phase, seems to depend continuously 
99: on the model. Finally,
100: we apply the \epsCoup method, to study low excitations. The
101: exponent $\theta$ describing the energy-scaling of the excitations
102: seems to depend not much on the energy model. 
103:   \end{abstract}
104: 
105: \pacs{64.60.Fr, 87.15.Aa}
106: 
107: \maketitle
108: %%%%%%%%%%%%%%%
109: 
110: \section{Introduction}
111: %
112: RNA plays an important rule in the biochemistry of all living
113: systems~\cite{GCA99,Hig00}.  Similar to the DNA, it is a linear chain-molecule
114: build from four types of bases, \ie adenine (A), cytosine (C), guanine (G),
115: and uracil (U).  
116: It does not only transmit pure genetic information, but, \eg works as a
117: catalyst. While for the former the primary structure, \ie the sequence of
118: the bases, is relevant, for the later the kind of higher order structures, \ie
119: secondary and tertiary structures, are relevant. 
120: 
121: %%%% hydrogen bonds
122: Like in the double helix of the DNA, in RNA complementary bases can build
123: hydrogen bonds between each other. Compared to DNA, where the bonds are built
124: between two different strands, RNA builds bonds between bases of the same RNA
125: strand. The information, which bases of the strand are paired, gives the
126: secondary structure, and the spatial structure is called the tertiary
127: structure. 
128: The tertiary structure is stabilized by a much weaker interaction than the
129: secondary structure. This leads to a separation of energy scales between
130: secondary and tertiary structure, and gives the justification  to neglect the
131: later~\cite{BH99}. 
132: Therefore we deal here with the secondary structure only. 
133: 
134: One crucial point in calculating the secondary structure is the used energy
135: model: On the one hand, if one aims to get minimum structures close to the
136: experimentally observed one, one uses energy models that take into account
137: structure elements~\cite{Zuk89,McC90,HFS*94,LZP99}, \eg hair pin loops. 
138: On the other hand, if one is interested in the qualitative behaviour, one uses
139: models as simple as possible that keep the general behaviour, \eg only one
140: kind of base~\cite{LB04} or using energies depending only on the number and on
141: the type of paired bases~\cite{Hig96,BH02,MPR02,PPR00}. 
142: %
143: Here we will consider only models with the later kind of interaction energy. 
144: In recent years several authors examined this kind of models with regard to
145: the thermodynamic behaviour, \ie searching for phase transitions and
146: describing the type of phase involved~\cite{BH02,PPR00,Har01,MPR02}; Liu and
147: Bundschuh~\cite{LB03,LB04} recently discussed, whether  native RNA is already
148: in the regime of the thermodynamic limit or finite size effect have to be
149: taken into account. 
150: In this paper we numerically investigate a hybrid model of two well known
151: energy models \cite{HB03,MET*03},
152:  \ie a \emph{pair energy model}, where only base pairs are
153: considered regardless of their neighbourhood, and a \emph{stacking energy
154:   model}, where only consecutive paired bases, \ie forming a stack, gives an
155: energy contribution. 
156:  It has been claimed 
157: that the stacking energy is more relevant than just
158: the pair energy in real RNA \cite{TB99}.
159: Our model contains terms for {\em both} types of
160:   interactions and allows to move continuously from one model to the
161:   other. We are interested, whether the two limiting models are
162:   qualitatively different, in particular, whether a phase transition
163:   occurs, when moving from one model to the other.
164: 
165: 
166: The paper is organized as follows. In section \refsec{sec:model}, we define
167: our model, \ie we formally define secondary structures and introduce our energy
168: model. 
169: In \refsec{sec:algorithm}, we explain how to calculate the partition function
170: with a dynamic programming algorithm. 
171: In \refsec{sec:observables}, we introduce the observables which we
172: investigate in the following \refsec{sec:results}. 
173: While in \refsec{sec:binder_cumulant} and \refsec{sec:temperature_dependence}
174: we do  finite temperature calculations, in \refsec{sec:epsilon_coupling}
175: we use the \epsCoup method at zero temperature.
176: 
177: 
178: 
179: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
180: \section{The Model}
181: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
182: \label{sec:model}
183: %
184: Because RNA molecules are linear chains of bases, they  can be described as a
185: (quenched) sequence 
186: $\Sequence=(\Base[i])_{i=1,\dots,L}$ of bases
187: $\Base[i]\in\{\mathrm{A,C,G,U}\}$, where $L$ is the length of the sequence. 
188: Within this single stranded molecule some bases can pair and build a secondary
189: structure. 
190: Typically Watson-Crick base pairs, \ie A-U and C-G have the strongest
191: affinity to each other, they are also called
192:   complementary base pairs. 
193: Each base can be paired at most once.
194: For a given sequence $\Sequence$ of bases the secondary structure can be
195: described by a set $\Structure$ of pairs $(i,j)$ (with the convention $1\leq
196: i<j\leq L$), meaning that bases \Base\ and $\Base[j]$ are paired. 
197: For convenience of notation we further define a Matrix $(\SM)_{i,j=1,\dots,L}$
198: with $\SM=1$ if $(i,j)\in\Structure$, and $\SM=0$ otherwise.  
199: Two restriction are used: 
200: (i)
201: Here we exclude so called \emph{pseudo-knots}, that means, for any
202: $(i,j),(i',j')\in\Structure$, either $i<j<i'<j'$ or $i<i'<j'<j$ must hold, 
203: \ie we follow the notion of pseudo knots being more an element of the
204: tertiary structure \cite{TB99}.
205: 
206: 
207: (ii) Between two paired bases a minimum distance is required: $|j-i|\geq s$ is
208: required, granting some flexibility of the molecule (here $s=2$).
209: 
210: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
211: \subsection{Energy models}
212: \label{sec:energy_models}
213: %
214: Every secondary structure \Structure\ is assigned a certain energy
215: $E(\Structure)$; note that this energy in general depends on the \Sequence\ as
216: well, so it is more precisely to write $E(\Structure,\Sequence)$, but we assume
217: that the structure also includes the information about the sequence. 
218: With such an energy model it is possible to calculate the canonical partition function $Z$
219: of a given sequence \Sequence\ by summing over all possible structures
220: $Z=\sum_\Structure \E^{-\beta E(\Structure)}$, but it is  computationally
221: more efficient to compute it by using the partition functions of the subsequences, \ie by a dynamic programming approach.
222: 
223: Motivated by the observation that the secondary structure is due to building
224: of numerous base pairs where every pair of bases is bound by hydrogen
225: bonds, one assigns each pair $(i,j)$ a
226: certain energy $e(\Base,\Base[j])$ depending only on the kind of bases.
227: The total energy is the sum over all pairs
228: \begin{equation}
229:   \label{eq:paired_energy_model}
230: E_p(\Structure)=\sum_{(i,j)\in\Structure}e(\Base,\Base[j])\,,
231: \end{equation}
232: %
233: \eg by choosing $e(\Base[], \Base[]')=+\infty$ for non-complementary bases
234: \Base[], $\Base[]'$ pairings of this kind are suppressed.
235: 
236: Another possible model is to assign an energy \StackE\ to a pair
237: $(i,j)\in\Structure$  iff also $(i+1,j-1)\in\Structure$. This \emph{stacking
238:   energy} can be motivated by the fact that a single pairing gives some gain
239: in the binding energy, but also reduces the entropy of the molecule, because
240: through this additional binding it looses some flexibility.
241:  Formally the total
242: energy of a structure can be written as
243: %
244: \begin{equation}
245:   \label{eq:stacked_energy_model}
246:   E_s(\Structure)=\sum_{(i,j)\in\Structure}\StackE\SM[i+1,j-1]\,,
247: \end{equation}
248: %
249: assuming that for every  pair $(i,\,j)\in\Structure$ the bases \Base\ and
250: \Base[j] are complementary bases.
251: The total number \Ssize\ of consecutive base pairs is called the
252: \emph{stacking size}. 
253: Single base pairs are not considered as stacks,
254: therefore $\Ssize\geq2$ for any stack. 
255: 
256: Both types of energy models are discussed in the literature~% 
257: \cite{HB03,MET*03}, but, to our knowledge, so far no one has discussed a hybrid
258: model. 
259: We examine the sum of both models at once.
260: \begin{equation}
261:   \label{eq:energy_model}
262:   E(\Structure)\define E_p(\Structure)+ E_s(\Structure)
263: \end{equation}
264: where the parameters \StackE\ and $e(\Base[],\Base[]')$ can be adjusted freely,
265: including both models discussed above. 
266: Here we use 
267: \begin{equation}
268: \label{eq:pair_energy}
269:  e(\Base[],\Base[]')=
270:  \begin{cases}
271:    \PairE& \text{if $r$ and $r'$ are compl.\ bases}\\
272:    +\infty& \text{otherwise}
273:  \end{cases}
274: \end{equation}
275: with a pair energy $\PairE\leq0$ independent of the kind of bases. 
276: 
277: Due to the simple structure of the energy model, \eg the energies depend 
278: not on the position of the bases within the sequence or whether the paired
279: bases include some structure elements like hairpins, the ground state is highly
280: degenerated~\cite{PPR00,Har01}.
281: 
282: 
283: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
284: \subsection{Calculating the partition function}
285: \label{sec:algorithm}
286: %
287: Due to the fact that pseudo knots are excluded from our model (see
288: \refsec{sec:energy_models}), the calculation of the partition function can be 
289: done recursively. The algorithm is similar to that of Nussinov~%
290: \cite{NPG*78,DEK*98}.  
291: The algorithms calculates the elements of two upper-triangular matrices
292: $(\Z_{i,j})_{1\leq i\leq j\leq L}$ and $(\Zh_{i,j})_{1\leq i\leq j\leq L}$,
293: where $Z_{i,j}$ is the partition function of the
294: subsequence from base \Base\ to \Base[j] under the boundary condition that
295: bases \Base[i-1]  and \Base[j+1] are not paired, and $\Zh_{i,j}$ the partition
296: function under the boundary condition that bases \Base[i-1]  and \Base[j+1] are
297: complementary; $\Zh_{i,j}$ is only used as an auxiliary matrix. 
298: Then $\Z_{i,j}$ can be computed from the partition functions \Z\ and \Zh\ of
299: smaller subsequences in the following way 
300: (remember that $s$ denotes the minimum
301: distance between two bases of a pair)
302: %
303:   \begin{align}
304:   \label{eq:recursion}
305:  &\Z_{i,j}=\Z_{i,j-1}+\sum_{k=i}^{j-s-1}  
306:      \Z_{i,k-1}\,\E^{-e(\Base[k], \Base[j])/\kbT}\,\Zh_{k+1,j-1}\nonumber\\
307:     & \begin{aligned}
308:        \Zh_{i,j}=\Z_{i,j-1}
309:      +\E^{-(e(\Base[i], \Base[j])+\StackE)/\kbT}\,\Zh_{i+1,j-1}  \\
310:      +\sum_{k=i+1}^{j-s-1}  
311:      \Z_{i,k-1}\,\E^{-e(\Base[k], \Base[j])/\kbT}\,\Zh_{k+1,j-1}   
312:      \end{aligned}\\
313:       &\Z_{i,j}=\Zh_{i,j}=1\quad\text{for } i\ge j \nonumber\\
314:       &\Zh_{i,j}=0 \quad\text{for }0<j-i<s-2\nonumber
315:    \end{align}
316: %
317: which is schematically explained in \reffig{fig:partition_function}.
318: \begin{figure}[htbp]
319:   \includegraphics*[width=\picturewidth]{fig1}%{algorithm}
320:   \caption{Schematical explanation of \refeq{eq:recursion} and
321:     \refeq{eq:nuss_recursion}, \eg white boxes represent \Z, gray boxes \Zh.}  
322:   \label{fig:partition_function}
323: \end{figure}
324: %
325: %
326: Because both matrices depend on each other, they must be calculated
327: simultaneously, starting along the diagonal and continuing along the
328: off-diagonals. 
329: The calculation of the partition function can be done in $\order{L^3}$ steps,
330: where $L$ is the length of the  sequence.
331: The partition function $Z$ of the entire sequence is $\Z_{1,L}$, but also the
332: other matrix elements are useful for generating ensembles of structures
333: according to the Boltzmann distribution (see
334: \refsec{sec:observables}). 
335: 
336: A similar algorithm can be used to calculate the ground state energy: 
337:   \begin{align}
338:   \label{eq:nuss_recursion}
339:  &\N_{i,j}=\min(\N_{i,j-1}, 
340:  \min_{k=i}^{j-s-1}(\N_{i,k-1}+e(\Base[k], \Base[j]))+\Nh_{k+1,j-1})\nonumber\\
341:     & \begin{aligned}
342:        \Nh_{i,j}=\min(\N_{i,j-1}, e(\Base[i], \Base[j])+\StackE+\Nh_{i+1,j-1},
343:        \\
344:      \min_{k=i+1}^{j-s-1}(  
345:      \N_{i,k-1}+e(\Base[k], \Base[j])+\Nh_{k+1,j-1}   ))
346:      \end{aligned}\\
347:      & \N_{i,j}=\Nh_{i,j}=0\quad\text{for } i\ge j \nonumber\\
348:      & \Nh_{i,j}=+\infty \quad\text{for }0<j-i<s-2\nonumber
349:    \end{align}
350: In comparison to \refeq{eq:recursion} additions are replace by
351: $\min$-operations, multiplications by additions and the exponentials of the
352: energies by the energies them self. 
353: 
354: 
355: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
356: \section{Observables}
357: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
358: \label{sec:observables}
359: %
360: After calculating the partition function \Z\ for a given random sequence, we
361:  want to measure some quantities to compare the members of the ensemble. 
362: In principle most quantities could be calculated by a similar dynamic
363: programming algorithm introduced above, but in general the running time
364: behaviour would be of higher order (than three) in the sequence
365: length. For this reason we use a different method, where an ensemble of
366: structures is generated due to its Boltzmann weight~\cite{Hig96}. 
367: The procedure to build a sequence is essentially a backtracking algorithm:
368: Starting with the entire sequence a partner for an outermost base, \eg base
369: $L$, is chosen with the appropriate probability,  \eg base $k$, after this the
370: procedure is applied to the subsequences 1 to $k-1$ and $k+1$ to $L$. 
371: If base $L$ is chosen not to be paired at all, one uses the sequence shortened
372: by base $L$.
373:  Considering
374: a subsequence from base $k$ to $l$, the 
375: probability $p_{i,j;k,l}$ that bases $i$ and $j$ ($k\leq i<j\leq l$) are paired is given by
376: %
377: %\begin{widetext}
378:   \begin{equation}
379:     \label{eq:probability}
380:     p_{i,j;k,l}=
381:     \begin{cases}
382:       \begin{aligned}
383:   \Z_{k,l}^{-1}\,&\Z_{k,i-1}\E^{-(e(\Base[k], \Base[j])+E_0)/\kbT}\Zh_{j+1,l} \\
384:   &\; \text{bases $i-1$ and $j+1$ paired}\\ 
385:       \end{aligned}\\ \\
386:       \begin{aligned}
387:       \Z_{k,l}^{-1}\,&\Z_{k,i-1}\E^{
388: -e(\Base[k], \Base[j])/\kbT}\Zh_{j+1,l} \\
389:       &\;\text{bases $i-1$ and $j+1$ unpaired}\\
390:       \end{aligned}
391:     \end{cases}
392:   \end{equation}
393: %\end{widetext}
394: 
395:  For each
396: member of this ensemble \Ensemble\ the quantity  of interest $X$ is calculated
397: and the average
398: $\average{X}=\frac1{|\Ensemble|}\sum_{\Structure}X(\Structure)$ is used as
399: an approximation to the expectation value of the Gibbs-Boltzmann-ensemble; 
400: for large enough ensembles \Ensemble\ this average approaches the true expectation value. 
401: 
402: A simple observable is the energy $E$ and its fluctuations $(\Delta E)^2$, the
403: later is directly connected with the specific heat $c_V=(\Delta E)^2/L\,\kb
404: T^2$. 
405: 
406: Of particular importance is the overlap \overlap\ between two structures
407: $\Structure$ and $\Structure'$
408: \begin{equation}
409:   \label{eq:overlap}
410:   \overlap(\Structure,\,\Structure')\define
411:   \frac2L\sum_{1\leq i<j\leq L}\SM\,\SM'
412: \end{equation}
413: %
414: that is the number of bases paired to the same base in both structures
415: normalized such that \overlap\ lies always between zero and one. This is a
416: measure of how similar two structures are. 
417: Note however, that with this definition the overlap of one structure with
418: itself is $\overlap(\Structure,\,\Structure)\lneqq1$ unless all bases are
419: paired, where it is equal to one. 
420: A definition of $\overlap$ where also bases unpaired in both structures are
421: counted is used in~\cite{Hig96} resulting in an overlap definition that is
422: normalized, however this similarity measurement has the drawback that the less
423: bases are paired the more two structures get similar.
424: We further remark that for any two structures \Structure, $\Structure'$ the Cauchy-Schwarz inequality $(\overlap(\Structure, \Structure'))^2\leq
425: \overlap(\Structure, \Structure)\,\overlap(\Structure', \Structure')$ holds.
426: With this quantity \overlap\ two ensembles $\Ensemble$ and $\Ensemble'$ can be
427: compared, \eg looking at the distribution of
428: $\overlap(\Structure,\,\Structure')$ of all
429: $\Structure\in\Ensemble,\Structure'\in\Ensemble'$. 
430: 
431: 
432: The ensemble averages $\average{\cdot}_{\Ensemble}$ in general depend on the
433: chosen sequence, therefore a further averaging over several (random) sequences
434: is required. This sequence average is denoted by $\qaverage{\cdot}$.
435: We again approximate this average by summing over a finite set of sequences.
436: 
437: Because of this two stage averaging, it is probably preferable instead of
438: looking  at $\qaverage{\average\overlap}$ directly to use
439: functions of the first and higher moments of \overlap, \eg the
440: Binder cumulant \cite{Bin81,BY85,BY88}, which  is defined by
441: %
442: \begin{equation}
443:   \label{eq:binder_cumulant}
444:   \BC\define\frac12\left(3-\frac{\qaverage{\average{q^4}}}{\qaverage{\average{q^2}^2}}\right)
445:  \end{equation} 
446: %
447: where $\average{\overlap^n}$ is either the average over all pairs of one
448: ensemble 
449: \begin{equation}
450:   \label{eq:ensemble_overlap}
451:   \overlap_{\Ensemble}\define \frac1{|\Ensemble|\,(|\Ensemble|-1)} 
452:             \sum_{\atop{\Structure,\,\Structure'\in\Ensemble}{\Structure\neq\Structure'}}  
453:             \overlap(\Structure,\,\Structure')\,
454: \end{equation}
455:  or the average over all pairs of two ensembles
456: \begin{equation}
457:   \label{eq:reference_overlap}
458:   \overlap_{\Ensemble,\,\Ensemble'}\define \frac1{|\Ensemble|\,|\Ensemble'|} 
459:             \sum_{\Structure\in\Ensemble,\,\Structure'\in\Ensemble'}  
460:             \overlap(\Structure,\,\Structure')\,.
461: \end{equation}
462: %
463: The later choice is appropriate when one is looking for a change in the
464: behaviour of the models when one varies the parameters in comparison 
465: to a reference model, 
466: while the former  is the better choice, if an external parameter, \eg the
467: temperature, is varied.
468: 
469: The Binder cumulant \BC\ vanishes at high temperatures, while for low
470: temperatures it approaches a finite value in the thermodynamic limit.
471: 
472: A similar quantity has been used in~\cite{PPR00}:
473: \begin{align}
474:   \label{eq:parisi_cumulant}
475:   \PC&\define
476:   \frac{\qaverage{\qvariance^2}-\qaverage{\qvariance}^2}%
477:        {\qaverage{\qvariance}^2}\\
478: \intertext{where}
479: \label{eq:qvariance}
480: \qvariance&\define L\,\left(\average{\overlap^2}-\average{\overlap}^2\right)
481: \end{align}
482: %
483: is  the variance of the $\overlap$ distribution.
484: This parameter \PC\ measures how the probability distribution of $\overlap$
485: varies from sequence to sequence. A value close to zero indicates a
486: self-averaging behaviour. 
487: 
488: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
489: % Results                    %
490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
491: \section{Numerical Results}
492: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
493: \label{sec:results}
494: 
495: In order to find some possible differences in the behaviour of the energy
496: model \refeq{eq:energy_model} for different parameters \PairE, \StackE,
497:  as introduced in \refeq{eq:pair_energy} and \refeq{eq:stacked_energy_model},
498: respectively, 
499: we
500: numerically calculated the quantities mentioned in \refsec{sec:observables}
501: above. 
502: In all our examples we averaged over randomly generated sequences
503: $\Sequence=(\Base)$,  where the probability for a specific base \Base\ at
504: position $i$ is $\frac14$ for all base types independent of the other bases
505: \Base[j\neq i].
506: The size of the sequence varied from $L=128$ up to $L=1024$, for the disorder
507: average 2000 up to 6000 random sequences were used. 
508: Pairing of bases are only allowed for complementary bases and the minimum
509: distance between to bases was chosen as $s=2$.
510: 
511: In \refsec{sec:basic_observables} and \ref{sec:binder_cumulant}, we vary
512: continuously the energy parameters between the two extreme cases
513: $(\PairE=0,\StackE=-1)$ and $(\PairE=-1,\StackE=0)$.
514: In \refsec{sec:basic_observables} we shortly discuss the averages of the
515: stacking size \Ssize\ and of the \overlap\ for different energy parameters.
516: In \refsec{sec:binder_cumulant} we examine the cross overlap
517: $\overlap_{\Ensemble,\,\Ensemble^\text{ref}}$, 
518: see \refeq{eq:reference_overlap},
519: between a reference ensemble $\Ensemble^\text{ref}$ generated for fixed
520: energy parameters, and  ensembles generated for different energy parameters.
521: In the following \refsec{sec:temperature_dependence} we look at the
522: temperature variation of various quantities without using any reference
523: ensemble to estimate some critical parameters. 
524: In the last \refsec{sec:epsilon_coupling} we apply the \epsCoup
525: method at $T=0$ to determine 
526: the critical exponent $\theta$ describing the behavior of
527: low-lying excitations. 
528: 
529: 
530: 
531: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
532: \subsection{Basic observables}
533: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
534: \label{sec:basic_observables}
535: %
536: %
537: In \refsec{sec:energy_models}, where we introduced the energy model, we
538: opposed the pair energy to the stacking energy model. 
539: In \reffig{fig:average_stacking_size} we show the average size \Ssize\ of a
540: stacking as a function of the energy parameter $\PairE$ at temperature
541: $T=0$. 
542: We keep  $\PairE+\StackE=-1$ constant, to fix the overall
543: energy scale. 
544: %
545: \begin{figure}[ht]
546:   \includegraphics*[width=\picturewidth]{fig2}%{StackingSize}  
547:   \caption{ \label{fig:average_stacking_size}
548:      Average stacking size as a function of energy parameter $\PairE$ for
549:      different system sizes. At $\PairE=-1$, \ie $\StackE=0$, and $\PairE=0$
550:      the curves are discontinuous.
551:     \legendref
552:     \errorbars
553:   }
554: \end{figure}
555: %
556: For all fixed energy parameters the average stacking
557: size \Ssize\ increases with increasing system size,
558: which is expected as with increasing length the probability for longer stacks
559: increases. 
560: Also as expected is  the overall increase in the average stacking size with
561: the energy parameter $\StackE$, because the stacking energy prefers to build
562: stacks. 
563: The large increase of \Ssize\ while changing $\StackE$ from zero  to a nonzero
564: value can be explained as following:
565: The ground state for $\PairE=-1.0, \StackE=0.0$ is highly degenerated, while
566: changing to a nonzero \StackE\ only those states stay  ground states which
567: have a high stacking contribution to the energy. This selection increases the
568: average of \Ssize.
569: A similar argument applies at the other end, where \PairE\ changes from a nonzero value to zero.
570: 
571: Similarly the overlap \overlap\ jumps to a larger value when changing from
572: $\PairE=0$ to a nonzero value (\reffig{fig:average_overlap}).
573: In addition at positions, where $\PairE/\StackE$ are fractions with small
574: numerator and denominator, \eg $\frac11$ or $\frac13$,  
575: %
576: \begin{figure}[ht]
577:   \includegraphics*[width=\picturewidth]{fig3}%{Overlaps}  
578:   \caption{  \label{fig:average_overlap}
579: Average overlap $\overlap_\Ensemble$ as a function of the  energy
580:     parameter \PairE\ for different system sizes. The local minima at
581:     $\PairE=-0.75,-0.5,-0.25$ are due to the commensurability of \PairE\ and
582:     \StackE\, and indicate a broad distribution of the ground states in
583:     configuration space.
584:     \legendref
585:     The left inset is an enlargement to show the discontinuity.
586:     The right inset is the $\overlap$-distribution for $\StackE=\PairE=-0.5$
587:     (solid lines) and $\StackE=-0.49, \PairE=-0.51$ (broken lines) for
588:     sequence lengths $L=512$ (thin lines) and $L=1024$ (thick lines).
589:   }
590: \end{figure}
591: %
592: for this energy parameters  the ground states in configuration space are more
593: broadly distributed than for slightly different parameters. This can be seen
594: in the right inset of \reffig{fig:average_overlap}, where $q$-Distribution in
595: the symmetric case ($\StackE=\PairE=-0.5$) is broader than in the slightly
596: asymmetric case ($\StackE=-0.49, \PairE=-0.51$).
597: The width of this minima in the main plot, as well as that in
598: \reffig{fig:binder_cumulant} and \reffig{fig:binder_cross_cumulant}, is below
599: $\Delta E=0.001$, as one can estimate from the left inset of \reffig{fig:average_overlap}. 
600: 
601: For both the average stacking size and the average overlap, the
602: behavior changes smoothly when moving from one model to the other,
603: with the exception of the highly degenerate points, where we can
604: observe the jumps in the overlap $\overlap$. Hence, there is no sign
605: of a transition in between the two extremal models. To confirm this, we
606: next study the Binder cumulant.
607: 
608: 
609: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610: \subsection{Binder cumulant}
611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
612: \label{sec:binder_cumulant}
613: %
614: Since we introduced in \refeq{eq:energy_model} a whole class of energy
615: models depending on the pair energy $E_p$ and the stacking energy $E_s$, we
616: examined the behaviour of the Binder cumulant depending on this two energy
617: parameters and the sequence length $L$.
618: %We generated the reference Ensemble for fixed chosen $\PairE$ and
619: %$\StackE^\text{ref}$, and compare it with a ensembles for varying $E_p$ (the 
620: %stacking energy was kept constant at $\StackE=\StackE^\text{ref}$).
621: Second order phase transitions are characterized by crossing of the
622: Binder cumulant for different system sizes at the transition point.
623: 
624: 
625: In \reffig{fig:binder_cumulant}the Binder cumulant \refeq{eq:binder_cumulant}
626: is shown at $T=0$ using the ``self-overlap''
627: \refeq{eq:ensemble_overlap}. 
628: %
629: \begin{figure}[ht]
630:   \includegraphics*[width=\picturewidth]{fig4}%{Binder}  
631:   \caption{  \label{fig:binder_cumulant}
632:     The Binder Cumulant \BC\ of \refeq{eq:binder_cumulant} with
633:     $q=\overlap_{\Ensemble}$ (see \refeq{eq:ensemble_overlap})  at temperature
634:     $T=0$. 
635:     The curves for different system sizes do not cross, and therefore \BC\
636:     gives no hint for a phase transition.
637:   \legendref
638: }
639: \end{figure}
640: %
641: Again, the energy parameter \PairE\ and \StackE\ are varied such that always
642: $\PairE+\StackE=-1$ holds. 
643: The value of \BC\ increases with increasing system size for \PairE, \ie the
644: curves do not cross each other and therefore give no evidence of a phase
645: transition.
646:   The local minima are -- as the minima in \reffig{fig:average_overlap} -- due
647: to the commensurability of the energy parameter $\PairE$ and $\StackE$.
648: 
649: 
650: 
651: Further we used a reference ensemble generated for energy parameters
652: $\PairE=\StackE=-0.5$ and used the ``cross overlap'' of
653: \refeq{eq:reference_overlap}.
654: % 
655: \begin{figure}[ht]
656:   \includegraphics*[width=\picturewidth]{fig5}%{BinderCross}
657:   \caption{  \label{fig:binder_cross_cumulant}
658:     The Binder Cumulant \BC\ of \refeq{eq:binder_cumulant} with
659:     $q=\overlap_{\Ensemble,\Ensemble'}$ (see \refeq{eq:ensemble_overlap})  at
660:     temperature  $T=0$.  
661:   \legendref}
662: \end{figure}
663: %
664: First, the values of \BC\ at $\PairE=\StackE=-0.5$ coincide with the values
665: shown in \reffig{fig:binder_cumulant}.
666: Similar to the observation above, \BC\ roughly increases with the
667:  system size, although the separation of the curves is not as clear
668: as in \reffig{fig:binder_cumulant}, especially in the range from $\PairE=-0.4$
669: to $\PairE=0.0$, where the curves coincide within the error-bars. 
670: To summarize, the dependence of the Binder cumulant on the energy parameters
671: does not indicate a phase transition.
672: 
673: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
674: % end: \subsection{Binder cumulant}
675: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
676: 
677: \subsection{Temperature dependence of the energy models}
678: \label{sec:temperature_dependence}
679: %
680: Another possible method to discriminate between the different energy
681: parameters is to examine the temperature dependence of some quantities,
682: especially the behaviour at a critical temperature. 
683:   In Ref.\ \onlinecite{PPR00} it was shown for a similar model that below a
684: critical
685:   temperature, 
686: almost all sequences fold to a compact
687:   structure, but for most sequences not into a single structure. 
688:   They point out that  this kind of
689:  low temperature behaviour is well known from spin
690:   glass and other disordered systems.
691: %
692: \begin{figure}[ht]
693:   \includegraphics*[width=\picturewidth]{fig6}%{SpecHeat}
694:   \caption{  \label{fig:C_V__T}
695:     Specific heat $c_V$ as a function of temperature for parameters
696:     $\StackE=-1.0$ and $\PairE=0.0$. The curves for different system sizes
697:     crosses at $T\approx 0.25$ and $T\approx 0.11$. The inset is an
698:   enlargement of the main plot.
699:   \legendref}
700: \end{figure}
701: %
702: In \reffig{fig:C_V__T} we plotted the specific heat for different system sizes
703: as a function of temperature for $\StackE=-1.0$ and $\PairE=0.0$. 
704: All curves cross each other close to $T=0.11$ and $T=0.25$, which is an
705: evidence for a phase transition at this temperature region. 
706: The data for other energy parameters $(E_s, E_p)$ look similar, but the
707: curves cross at different temperatures.
708: 
709: To determine the critical behaviour quantitatively we investigated the width
710: $\qvariance$ of the overlap distribution.
711: As can be seen from \reffig{fig:qvariance_T} all curves have a maximum and the
712: position of this maximum is decreasing with increasing sequence length. 
713: %
714: \begin{figure}[ht]
715:   \includegraphics*[width=\picturewidth]{fig7}%{Xi_Es1Ep0}
716:   \caption{  \label{fig:qvariance_T}
717:     $\qvariance$ as a function of temperature for energy parameter 
718:     $\StackE=-1.0$, $\PairE=0.0$. 
719:     \legendsymbols%explanation of symbols only here
720:   }
721: \end{figure}
722: %
723: \begin{figure}[ht]
724:   \centering
725:   \includegraphics*[width=\picturewidth]{fig8}%{Fit}
726:   \caption{  \label{fig:scaling_fit}
727:     The position of the maxima of $\qvariance$ for different energy parameter. 
728:     The curves are fitted to the form $T_c(L)=T_c+a\,L^{-1/\nu}$, and gives
729:     the critical parameters of \reftab{tab:critical_parameter_chi}.
730:     \errorbars
731:   }
732: \end{figure}
733: %
734: We assume that the maximum position has the form
735: $T_c(L)=T_c+a\,L^{-1/\nu}$ and  fit the data to this form to get the
736: critical parameters (see \reffig{fig:scaling_fit}).
737:  The results for three different pairs of energy parameters are shown in
738: \reftab{tab:critical_parameter_chi}. 
739: %
740: \begin{table}[ht]
741:   \begin{tabular}{|l|l|l|l||l|}
742:     \hline
743:     \multicolumn{2}{|c|}{Energy Model} & $1/\nu$ & $T_c$& $\theta$\\
744:     \hline                                              % for \epsilon<0.5 
745:     $\StackE=0$   & $\PairE=-1$  & 0.93(15)& 0.086(3)&  0.229(38)\\
746:     $\StackE=-0.5$& $\PairE=-0.5$& 0.70(36)& 0.109(7)&  0.237(50)\\
747:     $\StackE=-1$  & $\PairE=0$   & 0.36(17)& 0.125(21)& 0.194(67) \\
748:     \hline
749:   \end{tabular}
750:   \caption{  \label{tab:critical_parameter_chi}
751:     Critical parameter of a $\qvariance$-maximum fit.
752:     Comparing the two limiting cases $(\StackE,\PairE)=(-1,0)$ and
753:     $(\StackE,\PairE)=(0,-1)$ the parameters $\nu$ and $T_c$ are different and
754:     indicate a quantitative change. 
755:     The last column belongs to the \epsCoup method in
756:     \refsec{sec:epsilon_coupling}.}
757: \end{table}
758: %
759: The critical exponent $\nu$ for the parameter pair $(E_s=0, E_p=-1)$ is
760: clearly different from that of the parameter pair $(E_s=-1, E_p=0)$, showing
761: that the quantitative behaviour changes when varying from one limiting case to
762: the other. On the other hand for the intermediate parameter set the critical
763: exponent is consistent with both within the error range.   
764: 
765: Finally, we also found (not shown) that 
766:  the behaviour of \PC\ of \refeq{eq:parisi_cumulant} is in agreement with
767: this observation and with the results~\cite{PPR00} for a two-letter RNA
768: model. 
769: For all three cases studied here,
770: the width of the $\overlap$--distribution varies only slightly from
771: realization to realization at high temperatures, while for low
772: temperatures the self averaging behaviour disappears.  
773: 
774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
775: % end: \subsection{Temperature dependence of the energy models}
776: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
777: 
778: \subsection{\epsCoup}
779: \label{sec:epsilon_coupling}
780: %
781: Previously the \emph{\epsCoup method} has been used\cite{MPR02,KMM02}
782: to investigate low-energy excitations of RNA secondary structures.
783: The basic idea is to add another term to the energy model in
784: \refeq{eq:energy_model}, which depends on the ground state of the original
785: problem:
786: Assume $\Structure_0$ is the unique ground state of $E(\Structure)$, then a new
787: energy function is defined as following
788: \begin{equation}
789:   \label{eq:epsilon_energy}
790:   E_\varepsilon(\Structure)=E(\Structure)
791:              +\varepsilon\,\overlap(\Structure,\,\Structure_0)
792: \end{equation}
793: with %--- of course --- 
794: $\varepsilon>0$.
795: The additional term penalizes structures similar to $\Structure_0$, where
796: $\varepsilon\,\overlap(\Structure,\Structure_0)$ is largest for
797: $\Structure=\Structure_0$.
798: In general the ground state $\epsStructure$ of the new energy model
799: $E_\varepsilon$ will be different from $\Structure_0$. 
800: The difference $\Delta E(\varepsilon)\define E(\epsStructure)-E(\Structure_0)$
801: is an increasing function of $\varepsilon$ and $\Delta E(\varepsilon)\leq
802: \varepsilon$ holds.  
803: The later implies that $\epsStructure$ is for small enough $\varepsilon$ a low
804: lying excitation of the original energy model, and has the smallest overlap
805: with $\Structure_0$.  
806: 
807: The average distance $\distance(\varepsilon, L)\define
808: 1-\overlap(\epsStructure, \Structure_0)$ between the new and the old ground
809: state scales as $\distance(\varepsilon, L)\propto \varepsilon\,L^{-\theta}$,
810: $\varepsilon$ constant,
811: while the average energy difference scales as $\Delta E(\distance, L)\propto
812: L^\theta$, $d$ constant, with the critical exponents $\theta$. For details see
813: Refs. \onlinecite{MPR02} or \onlinecite{KMM02}.
814: 
815: The assumption of a unique ground state used above does not hold for our model
816: used so far: in general the ground state is highly degenerated because only two
817: energy parameters (\StackE\ and \PairE) are used, many structures will have the
818: same energy. 
819: The degeneracy of the ground state renders the \epsCoup method as
820: described above almost useless. 
821: Since in natural RNA the contributions to the energy are more complex
822: different structure  will never be degenerated. 
823: This justifies to change the energy model slightly by adding a random
824: energy $\randE_{i,j}$ to the pair energies introduced in
825: \refeq{eq:paired_energy_model}: $e(\Base,\Base[j])\to
826: e(\Base,\Base[j])+\randE_{i,j}$.
827: There are several possibilities to choose the distributions of the $\randE$s 
828: (see \cite{MPR02}), here we use identical independently distributed Gaussian 
829: random number with zero mean $\average{\randE}=0$ and variance
830: $\average{\randE^2}=\randE_0^2/L$ with $\randE_0=0.1$ (the QD model in
831: Ref. \onlinecite{MPR02}). 
832: 
833: 
834: \begin{figure}[ht]
835: \includegraphics*[width=\picturewidth]{fig9}%{EpsCoupling}
836:   \caption{  \label{fig:epsilon_coupling}
837:     \epsCoup results for $\StackE=-0.5$, $\PairE=-0.5$. The energy difference
838:   $\Delta E(\Structure)$ between the original ground state structure
839:   $\Structure_0$ and the ground state structure $\epsStructure$ of the
840:   disturbed model is plotted as a function of the distance between this
841:   structures.
842:   The inset is a scaled plot of the data with $\varepsilon<1$ of the main
843:   plot. ($\theta=0.24$, see  \reftab{tab:critical_parameter_chi})
844:   \legendref
845: } 
846: \end{figure}
847: 
848: %
849: The raw result for different values of $\varepsilon \in [0.01,100]$ is
850: shown in  \reffig{fig:epsilon_coupling}.
851: A scaling plot of the
852:  data for $\varepsilon<1$ according the scaling form $\Delta E
853:  L^{-\theta}=f(\distance)$ is shown 
854: in the inset of  \reffig{fig:epsilon_coupling}. 
855: The scaling parameters $\theta$ leading to the best data collapse
856: for different energy parameters  are shown
857: in the right most column of \reftab{tab:critical_parameter_chi}. They
858:  are equal
859: within the error margins and thus does not give us a further hint of a
860: quantitative different behaviour for different energy parameters.
861: This difficulties in doing a good scaling of the data in the QD model were
862: already pointed out \cite{MPR02}.
863: However, for a different model using a scaling function
864: with finite-size corrections a critical exponent $\theta=0.23\pm0.05$
865: was obtained \cite{KMM02}, which
866: is close to our findings.     
867: 
868: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
869: % end: \subsection{$\varepsilon$-Coupling}
870: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
871: 
872: 
873: \section{Summary}
874: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
875: \label{sec:summary}
876: %
877: We have introduced a RNA model which continuously interpolates between two
878: well known models.
879: We sought for the answer to the question, whether there is any phase
880: transition of the thermodynamic behaviour.
881: We used both zero as well as finite temperature data.
882: 
883: Zero temperature results give no evidence for a phase transition apart from
884: trivial transitions, \eg discontinuity of \Ssize\ at points with $\StackE=0$
885: or $\PairE=0$. 
886: The curves of the Binder cumulant do not cross at a certain point, which would
887: be an indication of a phase transition.
888: Similar, the critical exponent $\theta$ derived from the \epsCoup method seems
889: to be independent of the energy parameters \StackE\ and \PairE, and therefore
890: gives no evidence for a quantitative difference in the thermodynamic
891: properties.
892: But as stated in \cite{MPR02}, the determination of the critical parameter is
893: rather difficult in the quasi-degenerated case studied here. 
894: 
895: The finite temperature results show -- in contrast to the zero temperature
896: data -- a quantitative dependence on the energy parameters.
897: The critical exponent $\nu$ for the correlation length seems to depend
898: on the energy model and may vary continuously while going from one
899: limiting model to the other.
900: 
901: 
902: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
903: % end: \section{Summary}
904: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
905: 
906: \begin{acknowledgments}
907: The authors have obtained financial support from the
908: \emph{Volkswagenstiftung} (Germany) within the program
909: ``Nachwuchsgruppen an Universit\"aten''.
910: The simulations were performed at the Paderborn Center
911: for Parallel Computing in Germany and on a workstation
912: cluster at the Institut f\"ur Theoretische Physik, Universit\"at
913: G\"ottingen, Germany.
914: We thank E.~Yewande for helpful remarks.
915: \end{acknowledgments}
916: 
917: %
918: %%\bibliography{bbas_def,bbase}
919: \providecommand{\noopsort}[1]{}
920: \begin{thebibliography}{23}
921: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
922: \expandafter\ifx\csname bibnamefont\endcsname\relax
923:   \def\bibnamefont#1{#1}\fi
924: \expandafter\ifx\csname bibfnamefont\endcsname\relax
925:   \def\bibfnamefont#1{#1}\fi
926: \expandafter\ifx\csname citenamefont\endcsname\relax
927:   \def\citenamefont#1{#1}\fi
928: \expandafter\ifx\csname url\endcsname\relax
929:   \def\url#1{\texttt{#1}}\fi
930: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
931: \providecommand{\bibinfo}[2]{#2}
932: \providecommand{\eprint}[2][]{\url{#2}}
933: 
934: \bibitem[{\citenamefont{Gesteland et~al.}(1999)\citenamefont{Gesteland, Cech,
935:   and Atkins}}]{GCA99}
936: \bibinfo{editor}{\bibfnamefont{R.~F.} \bibnamefont{Gesteland}},
937:   \bibinfo{editor}{\bibfnamefont{T.~R.} \bibnamefont{Cech}}, \bibnamefont{and}
938:   \bibinfo{editor}{\bibfnamefont{J.~F.} \bibnamefont{Atkins}}, eds.,
939:   \emph{\bibinfo{title}{The RNA World}} (\bibinfo{publisher}{Cold Spring Harbor
940:   Laboratory Press}, \bibinfo{address}{New York}, \bibinfo{year}{1999}),
941:   \bibinfo{edition}{2nd} ed.
942: 
943: \bibitem[{\citenamefont{Higgs}(2000)}]{Hig00}
944: \bibinfo{author}{\bibfnamefont{P.~G.} \bibnamefont{Higgs}},
945:   \bibinfo{journal}{Quarterly Review of Biophysics}
946:   \textbf{\bibinfo{volume}{33}}, \bibinfo{pages}{199} (\bibinfo{year}{2000}).
947: 
948: \bibitem[{\citenamefont{Bundschuh and Hwa}(1999)}]{BH99}
949: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bundschuh}} \bibnamefont{and}
950:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Hwa}},
951:   \bibinfo{journal}{Phys.\ Rev.\ Lett.} \textbf{\bibinfo{volume}{83}},
952:   \bibinfo{pages}{1479} (\bibinfo{year}{1999}).
953: 
954: \bibitem[{\citenamefont{Zuker}(1989)}]{Zuk89}
955: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zuker}},
956:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{244}},
957:   \bibinfo{pages}{48} (\bibinfo{year}{1989}).
958: 
959: \bibitem[{\citenamefont{McCaskill}(1990)}]{McC90}
960: \bibinfo{author}{\bibfnamefont{J.~S.} \bibnamefont{McCaskill}},
961:   \bibinfo{journal}{Biopolymers} \textbf{\bibinfo{volume}{29}},
962:   \bibinfo{pages}{1105} (\bibinfo{year}{1990}).
963: 
964: \bibitem[{\citenamefont{Hofacker et~al.}(1994)\citenamefont{Hofacker, Fontana,
965:   Stadler, Bonhoeffer, Tacker, and Schuster}}]{HFS*94}
966: \bibinfo{author}{\bibfnamefont{I.~L.} \bibnamefont{Hofacker}},
967:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Fontana}},
968:   \bibinfo{author}{\bibfnamefont{P.~F.} \bibnamefont{Stadler}},
969:   \bibinfo{author}{\bibfnamefont{L.~S.} \bibnamefont{Bonhoeffer}},
970:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tacker}}, \bibnamefont{and}
971:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Schuster}},
972:   \bibinfo{journal}{Monatsh.\ Chemie} \textbf{\bibinfo{volume}{125}},
973:   \bibinfo{pages}{167} (\bibinfo{year}{1994}).
974: 
975: \bibitem[{\citenamefont{Lyngs{\o} et~al.}(1999)\citenamefont{Lyngs{\o}, Zuker,
976:   and Pedersen}}]{LZP99}
977: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Lyngs{\o}}},
978:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Zuker}}, \bibnamefont{and}
979:   \bibinfo{author}{\bibfnamefont{C.~N.~S.} \bibnamefont{Pedersen}},
980:   \bibinfo{journal}{Bioinformatics} \textbf{\bibinfo{volume}{15}},
981:   \bibinfo{pages}{440} (\bibinfo{year}{1999}).
982: 
983: \bibitem[{\citenamefont{Liu and Bundschuh}(2004)}]{LB04}
984: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Liu}} \bibnamefont{and}
985:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bundschuh}},
986:   \bibinfo{journal}{Phys.\ Rev.\ E} \textbf{\bibinfo{volume}{69}},
987:   \bibinfo{eid}{061912} (pages~\bibinfo{numpages}{10}) (\bibinfo{year}{2004}),
988:   \urlprefix\url{http://link.aps.org/abstract/PRE/v69/e061912}.
989: 
990: \bibitem[{\citenamefont{Pagnani et~al.}(2000)\citenamefont{Pagnani, Parisi, and
991:   Ricci-Tersenghi}}]{PPR00}
992: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Pagnani}},
993:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Parisi}}, \bibnamefont{and}
994:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ricci-Tersenghi}},
995:   \bibinfo{journal}{Phys.\ Rev.\ Lett.} \textbf{\bibinfo{volume}{84}},
996:   \bibinfo{pages}{2026} (\bibinfo{year}{2000}).
997: 
998: \bibitem[{\citenamefont{Higgs}(1996)}]{Hig96}
999: \bibinfo{author}{\bibfnamefont{P.~G.} \bibnamefont{Higgs}},
1000:   \bibinfo{journal}{Phys.\ Rev.\ Lett.} \textbf{\bibinfo{volume}{76}},
1001:   \bibinfo{pages}{704} (\bibinfo{year}{1996}).
1002: 
1003: \bibitem[{\citenamefont{Bundschuh and Hwa}(2002)}]{BH02}
1004: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bundschuh}} \bibnamefont{and}
1005:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Hwa}},
1006:   \bibinfo{journal}{Phys.\ Rev.\ E} \textbf{\bibinfo{volume}{65}},
1007:   \bibinfo{pages}{031903} (\bibinfo{year}{2002}).
1008: 
1009: \bibitem[{\citenamefont{Marinari et~al.}(2002)\citenamefont{Marinari, Pagnani,
1010:   and Ricci-Tersenghi}}]{MPR02}
1011: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Marinari}},
1012:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Pagnani}}, \bibnamefont{and}
1013:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ricci-Tersenghi}},
1014:   \bibinfo{journal}{Phys.\ Rev.\ E} \textbf{\bibinfo{volume}{65}},
1015:   \bibinfo{eid}{041919} (pages~\bibinfo{numpages}{7}) (\bibinfo{year}{2002}),
1016:   \urlprefix\url{http://link.aps.org/abstract/PRE/v65/e041919}.
1017: 
1018: \bibitem[{\citenamefont{Hartmann}(2001)}]{Har01}
1019: \bibinfo{author}{\bibfnamefont{A.~K.} \bibnamefont{Hartmann}},
1020:   \bibinfo{journal}{Phys.\ Rev.\ Lett.} \textbf{\bibinfo{volume}{86}},
1021:   \bibinfo{pages}{1382} (\bibinfo{year}{2001}).
1022: 
1023: \bibitem[{\citenamefont{Liu and Bundschuh}(2003)}]{LB03}
1024: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Liu}} \bibnamefont{and}
1025:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bundschuh}},
1026:   \emph{\bibinfo{title}{Large finite size effects in {RNA} secondary
1027:   structures}}, \bibinfo{howpublished}{physics/0304108} (\bibinfo{year}{2003}).
1028: 
1029: \bibitem[{\citenamefont{Han and Byun}(2003)}]{HB03}
1030: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Han}} \bibnamefont{and}
1031:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Byun}},
1032:   \bibinfo{journal}{Nucleic Acids Research} \textbf{\bibinfo{volume}{31}},
1033:   \bibinfo{pages}{3432} (\bibinfo{year}{2003}),
1034:   \eprint{http://nar.oupjournals.org/cgi/reprint/31/13/3432.pdf},
1035:   \urlprefix\url{http://nar.oupjournals.org/cgi/content/abstract/31/13/3432}.
1036: 
1037: \bibitem[{\citenamefont{Mukhopadhyay et~al.}(2003)\citenamefont{Mukhopadhyay,
1038:   Emberly, Tang, and Wingreen}}]{MET*03}
1039: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mukhopadhyay}},
1040:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Emberly}},
1041:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Tang}}, \bibnamefont{and}
1042:   \bibinfo{author}{\bibfnamefont{N.~S.} \bibnamefont{Wingreen}},
1043:   \bibinfo{journal}{Phys.\ Rev.\ E} \textbf{\bibinfo{volume}{68}},
1044:   \bibinfo{pages}{041904} (\bibinfo{year}{2003}).
1045: 
1046: \bibitem[{\citenamefont{Tinoco and Bustamante}(1999)}]{TB99}
1047: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Tinoco}, \bibfnamefont{Jr}}
1048:   \bibnamefont{and}
1049:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Bustamante}},
1050:   \bibinfo{journal}{J.\ Mol.\ Biol.} \textbf{\bibinfo{volume}{293}},
1051:   \bibinfo{pages}{271} (\bibinfo{year}{1999}).
1052: 
1053: \bibitem[{\citenamefont{Nussinov et~al.}(1978)\citenamefont{Nussinov,
1054:   Pieczenik, Griggs, and Kleitman}}]{NPG*78}
1055: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Nussinov}},
1056:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Pieczenik}},
1057:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Griggs}},
1058:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{D.~J.}
1059:   \bibnamefont{Kleitman}}, \bibinfo{journal}{SIAM Journal of Applied
1060:   Mathematics} \textbf{\bibinfo{volume}{35}}, \bibinfo{pages}{68}
1061:   (\bibinfo{year}{1978}).
1062: 
1063: \bibitem[{\citenamefont{Durbin et~al.}(1998)\citenamefont{Durbin, Eddy, Krogh,
1064:   and Mitchison}}]{DEK*98}
1065: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Durbin}},
1066:   \bibinfo{author}{\bibfnamefont{S.~R.} \bibnamefont{Eddy}},
1067:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Krogh}}, \bibnamefont{and}
1068:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Mitchison}},
1069:   \emph{\bibinfo{title}{Biological sequence analysis}}
1070:   (\bibinfo{publisher}{Cambridge University Press}, \bibinfo{year}{1998}).
1071: 
1072: \bibitem[{\citenamefont{Binder}(1981)}]{Bin81}
1073: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Binder}}, \bibinfo{journal}{Z.\
1074:   Phys.\ B - Condensed Matter} \textbf{\bibinfo{volume}{43}},
1075:   \bibinfo{pages}{119} (\bibinfo{year}{1981}).
1076: 
1077: \bibitem[{\citenamefont{Bhatt and Young}(1985)}]{BY85}
1078: \bibinfo{author}{\bibfnamefont{R.~N.} \bibnamefont{Bhatt}} \bibnamefont{and}
1079:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Young}},
1080:   \bibinfo{journal}{Phys.\ Rev.\ Lett.} \textbf{\bibinfo{volume}{54}},
1081:   \bibinfo{pages}{924} (\bibinfo{year}{1985}).
1082: 
1083: \bibitem[{\citenamefont{Bhatt and Young}(1988)}]{BY88}
1084: \bibinfo{author}{\bibfnamefont{R.~N.} \bibnamefont{Bhatt}} \bibnamefont{and}
1085:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Young}},
1086:   \bibinfo{journal}{Phys.\ Rev.\ B} \textbf{\bibinfo{volume}{37}},
1087:   \bibinfo{pages}{5606} (\bibinfo{year}{1988}).
1088: 
1089: \bibitem[{\citenamefont{Krzakala et~al.}(2002)\citenamefont{Krzakala,
1090:   M{\'e}zard, and M{\"u}ller}}]{KMM02}
1091: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Krzakala}},
1092:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{M{\'e}zard}},
1093:   \bibnamefont{and}
1094:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{M{\"u}ller}},
1095:   \bibinfo{journal}{Europhys.\ Lett.} \textbf{\bibinfo{volume}{57}},
1096:   \bibinfo{pages}{752} (\bibinfo{year}{2002}).
1097: 
1098: \end{thebibliography}
1099: 
1100: \end{document}
1101: 
1102: