q-bio0508042/VROZ1.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
2: %\\  
3: %Title:         The Topology of Pseudoknotted Homopolymers 
4: %Authors:       Graziano Vernizzi, Paolo Ribeca, Henri Orland, A. Zee 
5: %Comments:      RevTeX 4 pages, 6 figures
6: %Report-no:     SPhT-T05/127, HU-EP-05/37, SFB/CPP-05-38
7: %\\
8: %  
9: %Abstract: 
10: % We consider the folding of a self-avoiding homopolymer on a lattice,
11: % with saturating hydrogen bond interactions. Our goal is to numerically
12: % evaluate the statistical distribution of the topological genus of
13: % pseudoknotted configurations. The genus has been recently proposed for
14: % classifying pseudoknots (and their topological complexity) in the
15: % context of RNA folding.  We compare our results on the distribution of
16: % the genus of pseudoknots, with the theoretical predictions of an
17: % existing combinatorial model for an infinitely flexible and
18: % stretchable homopolymer. We thus obtain that steric and geometric
19: % constraints considerably limit the topological complexity of
20: % pseudoknotted configurations, as it occurs for instance in real RNA
21: % molecules. We also analyze the scaling properties at large homopolymer
22: % length, and the genus distributions above and below the critical
23: % temperature between the swollen phase and the compact-globule phase,
24: % both in two and three dimensions. 
25: %
26: %  
27: % PACS  87.14.Gg DNA, RNA  
28: %       87.15.Cc Folding and sequence analysis                
29: %       02.70.Uu Applications of Monte Carlo methods
30: %       05.10.Ln Monte Carlo methods
31: %       87.53.Wz Monte Carlo applications
32: %       87.15.By Structure and bonding   
33: %       82.39.Pj Nucleic acids, DNA and RNA bases
34: %       02.10.Ox Combinatorics; graph theory
35: %
36: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
37: 
38: \documentclass[prl,twocolumn,superscriptaddress]{revtex4}
39: \usepackage{amssymb} 
40: \usepackage{graphicx}
41: \usepackage{epsfig}  
42: \usepackage{subfigure} 
43: \usepackage{psfrag} 
44:  
45:  
46: \begin{document} 
47: \begin{flushleft}
48: %SPhT-T05/127,   
49: %HU-EP-05/37,
50: %SFB/CPP-05-38
51: \end{flushleft}
52:  
53: \title{The Topology of Pseudoknotted Homopolymers} 
54:  
55: \author{Graziano Vernizzi} 
56: %\email{vernizzi@cea.fr} 
57: \affiliation{Service de Physique Th\'eorique, CEA Saclay, 91191 
58: Gif-sur-Yvette Cedex, France}  
59:  
60: \author{Paolo Ribeca} 
61: %\email{ribeca@physik.hu-berlin.de} 
62: \affiliation{Humboldt Universit\"at, Newtonstr.~15, 12489 Berlin, Germany}  
63: 
64: \author{Henri Orland} 
65: %\email{orland@cea.fr} 
66: \affiliation{Service de Physique Th\'eorique, CEA Saclay, 91191 
67: Gif-sur-Yvette Cedex, France}  
68: 
69:  
70: \author{A. Zee} 
71: %\email{zee@itp.ucsb.edu} 
72: \affiliation{Department of Physics, University of California, Santa 
73: Barbara, CA 93106, USA} 
74: \affiliation{Institute for Theoretical Physics, University of 
75: California, Santa Barbara, CA 93106, USA} 
76:  
77: \begin{abstract} 
78: We consider the folding of a self-avoiding homopolymer on a lattice,
79: with saturating hydrogen bond interactions. Our goal is to numerically
80: evaluate the statistical distribution of the topological genus of
81: pseudoknotted configurations. The genus has been recently proposed for
82: classifying pseudoknots (and their topological complexity) in the
83: context of RNA folding.  We compare our results on the distribution of
84: the genus of pseudoknots, with the theoretical predictions of an
85: existing combinatorial model for an infinitely flexible and
86: stretchable homopolymer. We thus obtain that steric and geometric
87: constraints considerably limit the topological complexity of
88: pseudoknotted configurations, as it occurs for instance in real RNA
89: molecules. We also analyze the scaling properties at large homopolymer
90: length, and the genus distributions above and below the critical
91: temperature between the swollen phase and the compact-globule phase,
92: both in two and three dimensions.
93: \end{abstract} 
94: \maketitle 
95: 
96: %\section{Introduction}
97: One of the most exciting fields in modern computational molecular
98: biology is the search for tools predicting the complex foldings of
99: bio-polymers such as RNA \cite{SBS,tinoco,Higgs}, when homologous
100: sequences are not available.
101: \begin{figure}[t]
102: \centering 
103: \includegraphics[width=0.48\textwidth]{fig1-complete.eps}
104: \caption{Top, from left to right: 
105: %\cite{rasmol}: 
106: a hairpin loop (PDB number 1NA2), its squiggle-plot and disk diagram
107: representation which is of genus zero since it is planar.  Bottom:
108: $H$-type pseudoknot (PDB number 1RNK). In this case the disk diagram
109: is not planar and has genus one.}
110: \label{fig:1}
111: \end{figure}
112: The prediction of the full tertiary structure of a RNA
113: molecule is still an open issue \cite{mfold}, mainly because of its
114: intrinsic high computational complexity \cite{lyngso}. It is known 
115: that the tertiary structure involves an important set of
116: structural motifs, the so-called {\it pseudoknots}
117: \cite{Bosch}. These are conformations such that the
118: associated disk diagram  
119: (which represents all nucleotides along the RNA backbone as points on an
120: oriented circle from the 5' end to the 3' end, and where each base-pair is
121: represented by an arc joining the two interacting
122: nucleotides, inside the circle; see Figure \ref{fig:1}) is not planar, 
123: i.e.~it contains intersecting
124: arcs. RNA pseudoknots have been identified in nearly every organism,
125: and they proved to play important regulatory and functional roles 
126: \cite{c1,c2}. Their ubiquity manifests in a large variety
127: of possible shapes and structures \cite{plos}, and their existence
128: should not be neglected in structure prediction algorithms, as they
129: account for the 10\%-30\% on average of the total number of base
130: pairs. Actually, several computer programs have been proposed for
131: predicting RNA secondary structures including pseudoknots
132: \cite{eddyrivas, POZ, PTOZ, HI, Gul, abrah, Tab} (the list is not 
133: exhaustive), but the complexity of the problem and the approximations
134: involved are usually such that the issue is far from being solved
135: \cite{eddyBIO}.\\
136: An analytic mathematical tool which can fully describe any RNA contact
137: structure including all possible pseudoknots, appeared first in
138: \cite{OZ}. There, all  RNA disk diagrams are  considered as Feynman diagrams
139: of a suitable field theory of $N \times N$ Hermitian matrices (a
140: combinatorial tool borrowed from quantum field theory). The latter is
141: known to organize all the diagrams according to an asymptotic
142: $1/N^2$ topological expansion at large-$N$ \cite{thooft}. This
143: provides in fact a rigorous way to classify non-planar diagrams, and
144: therefore it induces a natural topological classification of
145: pseudoknots \cite{PTOZ}.  Namely, to any given pseudoknotted
146: configuration (and more generally, to any contact structure of an
147: heteropolymer with binary saturating interactions), one can associate
148: an integer number $g$, the {\it genus}. It is defined as the
149: topological genus of the associated disk diagram, i.e.~by $\chi=1-2g$
150: where $\chi$ is the Euler characteristic number of the diagram. As
151: reviewed in \cite{VOZ}, the genus is the minimum number of handles the
152: disk should have in order that all the cords are not intersecting (see
153: Figure \ref{fig:1}). Other characterizations of pseudoknots have been
154: proposed (e.g.~\cite{stella, Aeddy, Lucas}). The classification 
155: \cite{OZ} is truly topological, meaning that it is independent from the 
156: way the diagram is drawn, and dependent only
157: on the intrinsic complexity of the contact structure.
158: 
159: The large-$N$ asymptotics of the analytical model in \cite{OZ} is hard
160: to obtain exactly.  However, in \cite{PRL} a special case of the
161: general model \cite{OZ} has been considered and solved. It was the 
162: simple case of an infinitely flexible and stretchable homopolymer, where
163: there is no dependence on the primary sequence, and any saturating
164: base pair between all the ``nucleotides'' is allowed. An analytical
165: asymptotic expansion was evaluated and the distribution of the genus of
166: pseudoknotted contact structures was obtained. One of the
167: results is that an homopolymer with $L$ nucleotides has an average 
168: genus close to the maximal one, that is $L/4$. Of course, real RNA 
169: molecules are not infinitely flexible and stretchable homopolymers. 
170: It is customary to assume that the bases $i$ and $j$ can interact only 
171: if they are sufficiently far apart along the chain 
172: (e.g.~$\left|i-j\right|\geq4$,
173: \cite{mfold}) because of bending rigidity.  Moreover helices have a long
174: persistence length ($\sim 200$ base pairs) and this 
175: ne\-ces\-sa\-ri\-ly con\-strains the al\-lowed pairings even more.  
176: We expect that including 
177: all steric and geometrical constraints should considerably decrease
178: the genus of allowed pseudoknots, compared to the purely combinatorial
179: case \cite{PRL} where the actual three-dimensional conformation was
180: neglected. The purpose of this Letter is to numerically analyze the 
181: effects of
182: steric and geometric constraints on the genus distribution of
183: pseudoknots topologies in homopolymers, in the same spirit of
184: \cite{PRL}.
185: \smallskip
186: 
187: %\section{The Model}
188: \underline{The Model:} 
189: we model the system by considering a polymer on a cubic
190: lattice, i.e.~a self-avoiding random walk with short-range attractive
191: interaction \cite{degennes}.  A self-avoiding walk (SAW) is a sequence
192: of neighboring lattice sites $i=0,1,\ldots,n$ with coordinates $\{
193: \mathbf{r}_i \}$, such that the same lattice-site cannot be visited
194: more than once. This is a standard approach
195: in polymer physics (and RNA: see 
196: e.g.~\cite{stella, Lucas}). The attractive interaction is usually 
197: used to describe bad solvent quality, but in our case we insist
198: more on the saturating nature of the hydrogen bond interactions. Such a 
199: requirement is crucial here, since the concept of the topological genus
200: for a contact structure can be defined unambiguously only when the
201: interactions are saturating. One of the most natural ways to model 
202: the interaction is by considering a ``spin'' model (see 
203: e.g.~\cite{GOO,FS, Tiana}).  Strictly speaking, our
204: model is a variation of the standard $\theta$-polymer model, and
205: similar interaction models for RNA on the lattice have been already
206: proposed (e.g.~\cite{orl}).  To each vertex $i$ we
207: associate a unit spin $\mathbf{s}_i$ which represents the
208: nucleotide direction with respect to the backbone. The only
209: allowed directions for $\mathbf{s}_i$ are the lattice ones. Moreover,
210: the spins cannot overlap with the backbone because of the excluded
211: volume between the nucleotides and the backbone.
212: %, i.e.~$\mathbf{s}_i
213: %\cdot (\mathbf{r}_{i \pm 1}-\mathbf{r}_{i})\neq a$ where $a$ is the 
214: %lattice spacing. 
215: The saturating nucleotide-nucleotide interaction
216: occurs when two spins $\mathbf{s}_i,\mathbf{s}_j$ on neighboring
217: sites, $\left| \mathbf{r}_i-\mathbf{r}_j \right|=1$, are pointing to
218: each other. The energy of a configuration $\{\mathbf{r}_i,\mathbf{s}_i \}$ is thus defined by the
219: Hamiltonian:
220: \begin{equation}
221: {\mathcal H}=-\epsilon \sum_{i<j} \, 
222: \delta(\mathbf{r}_i+ \mathbf{s}_i - \mathbf{r}_j) 
223: \delta(\mathbf{s}_i+ \mathbf{s}_j) 
224: \delta(\left| \mathbf{r}_i - \mathbf{r}_{j}\right|-1) \,  ,
225: \label{Hamiltonian}
226: \end{equation}
227: where $\epsilon\geq 0$ is an effective hydrogen-binding energy, the
228: same for all monomers of the chain. Let us note that since we are not
229: aimidgng to set up a realistic lattice model for RNA-folding, but rather
230: to understand steric effects on the genus distributions of a
231: homopolymer, we do not take into account stacking energies.
232: 
233: The basic features of our model are clear:
234: At high temperatures, we expect the system to be in
235: a swollen SAW state (entropy dominated coil state), 
236: whereas at lower temperatures 
237: we expect a kind of ``compact globule''-like phase
238: \cite{degennes}. The transition temperature $T_\theta$
239: defines the so-called $\theta$-point. However, details on the
240: thermodynamics, kinetics, phase diagram, etc. can be rather complex
241: \cite{orl,Lucas}. We limit ourselves here only to the analysis of the
242: genus distribution of pseudoknotted structures for comparing the
243: effects of stericity constraints versus the purely combinatorial model
244: of \cite{PRL}.  All other considerations are postponed elsewhere.
245: \smallskip
246: 
247: %\section{The Method}
248: \underline{The Method:}
249: The numerical sampling of the statistical distribution ${\mathcal Z}=
250: \sum_{\mathrm{SAW},\{\mathbf{s}_i\}} \exp({-\mathcal H/k_B T})$, where
251: $k_B$ is Boltzmann's constant, $T$ is the absolute temperature, and
252: the sum is restricted to SAWs and configurations of spins $\{
253: \mathbf{s}_i\}$ satisfying the aforementioned constraints, is
254: im\-ple\-men\-ted by using the Monte Carlo Growth Method. It was
255: originally proposed by T. Garel and H. Orland in \cite{GO} and has
256: been applied to several statistical systems since then (see references
257: in \cite{HO}).  It
258: consists in starting with an ensemble of chains at equilibrium and
259: then growing each chain by adding one monomer at a time with a
260: probability proportional to the Boltzmann factor for the energy of the
261: chain. At each step the ensemble remains at equilibrium (a detailed
262: description of the algorithm with applications can be found in
263: \cite{HO}). It belongs to the family of so-called
264: ``population Monte Carlo algorithms'' \cite{Iba}, where, contrary to
265: the ``dynamical'' Markov Chain Monte Carlo methods, the population is
266: fully grown and evolved, non-dynamically. 
267: %The main advantage is that,
268: %in principle, there are no slowing-down effects close to critical
269: %points. 
270: At high temperatures we considered populations with a 
271: variable number of chains in the range 10000-40000, and with a 
272: typical  length of $L=500$ monomers (up to $L=1200$ in some cases). 
273: Accuracy and statistical averages were computed by
274: taking several independent populations (of the order of 40). 
275: At low temperatures we considered populations of up to 100000 chains. 
276: All the simulations have also been performed on a square 
277: lattice in two dimensions.
278: \smallskip
279: 
280: %\section{Results and discussion}
281: \underline{Results and discussion:}
282: We expect different genus distributions above and below $T_\theta$.
283: We therefore first determine $T_\theta$ , which can be done efficiently by
284: computing the end-to-end distance $R_e^2=(
285:  \mathbf{r}_L-\mathbf{r}_0)^2$, and the  radius of
286: gyration $R_g^2=\sum_{i<j} (\mathbf{r}_i-\mathbf{r}_j)^2/L^2$. 
287: It is known that the ratio $\rho^2=\langle R_e^2\rangle 
288: /\langle R_g^2 \rangle$ is universal in 
289: the limit $L \to \infty$ and converges to a step function as a
290: function of $T$, with a universal critical value at $T_\theta$
291: \cite{degennes, PHA, Nickel}. In Figure \ref{fig:2} we plot $\rho$ 
292: which shows a transition temperature $T_\theta=0.39 \pm 0.01$ ($
293: T^{2D}_\theta=0.48 \pm 0.02$ in two dimensions), in units 
294: where $k_B=1$ and $\epsilon=1$.
295: \begin{figure}[hbt]  
296: \centering \includegraphics[width=0.48\textwidth]{rho-3D-COL.final.eps}
297: \caption{The ratio $\rho$ in $3d$
298: as a function of the temperature $T$, for  several values
299: of $L$ (with error bars plotted only for $L=60$). At low temperatures
300: the error bars are larger than the variations of the curves.  
301: The inset shows that as $L$ increases,  the curves approach 
302: a  universal step function about $T_\theta\sim 0.39\pm 0.01$.}  
303: \label{fig:2}
304: \end{figure}
305: We also verified that $\rho_\infty=2.5\pm 0.05$ for $T\gg T_\theta$ 
306: asymptotically, 
307: and we find an intermediate value $\rho_\theta=2.35\pm 0.05$ at $T_\theta$ 
308: (and $\rho_\infty=2.67 \pm 0.01 $ and $\rho_\theta=2.39 \pm 0.01$ in 2d, 
309: respectively).  At 
310: large-$L$ we find the following scalings: for $T>T_\theta$, $\langle R_g 
311: \rangle  \sim L^\nu$ with $\nu=0.59 \pm 0.01$ ($\nu=0.75\pm 0.02$ in 
312: two dimensions), which is consistent with the critical exponent of a
313: swollen SAW; for $T<T_\theta$, $\nu=0.32 \pm 0.02$ ($\nu=0.50\pm 0.01$
314: in two dimensions) which is consistent with a compact phase.  All
315: these results are in agreement with high-accuracy simulations of
316: similar models \cite{grassberger,madras}.  We then proceed with
317: extracting the genus distributions in the two phases. The results are
318: in Figure \ref{fig:3} and \ref{fig:4}.
319: \begin{figure}[t]  
320: \centering 
321: \includegraphics[width=0.48\textwidth]{genus-2D-below-COL.final.eps}
322: \includegraphics[width=0.48\textwidth]{genus-3D-below-COL.final.eps}
323: \caption{The genus distributions  of pseudoknots at fixed 
324: $L$, in $2d$ (top) and $3d$ (bottom), 
325: in the compact phase at $T=0.225$ and $T=0.2$, respectively. 
326: The insets represent the 
327: behavior  of $\langle g\rangle$ at large-$L$.} 
328: \label{fig:3}
329: \end{figure}
330: \begin{figure}[hbt]
331: \centering
332: \includegraphics[width=0.48\textwidth]{genus-2D-above-COL.final.eps}
333: \caption{The genus distribution of a two dimensional homopolymer, 
334: in the swollen phase $T=10>T^{2D}_\theta$, at various values of $L$.}
335: \label{fig:4}
336: \end{figure}
337: When comparing them with the combinatorial results of \cite{PRL}, we
338: see that the genus at a fixed $L$ is on the average much smaller.
339: More precisely, below the $\theta$-point the average genus scales like
340: by $\langle g / L \rangle \sim 0.141 \pm 0.003$ and $\langle g / L
341: \rangle \sim 0.1318\pm 0.0025$, in $3d$ (at $T=0.2$) and $2d$ (at
342: $T=0.225$), respectively. In both cases the scaling is at a lower rate
343: (about 50\% less) than the value $L/4$ computed in \cite{PRL}.  In the
344: swollen-phase (e.g.~$T=10>T_\theta$), the average genus is given by
345: $\langle g / L \rangle \sim (585 \pm 8)10^{-6}$ in $3d$, and $\langle
346: g / L \rangle \sim (410 \pm 1) 10^{-5}$ in $2d$.  Such a low rate
347: comes from the tendency of a homopolymer to develop long rectilinear
348: sub-chains in the swollen phase.  In two dimensions the entropic
349: factor is smaller than in three dimensions and the genus growth rate
350: is therefore larger (see Figure \ref{fig:4}).  Moreover, the genus
351: distributions for $T>T_\theta$ are numerically consistent with
352: Poissonian distributions (see Figure \ref{fig:4}), whereas at smaller
353: temperatures they are closer to Gaussian ones.
354: 
355: It turns out that the average genus is an extensive quantity, like the
356: energy, and their ratio is shown in Figure \ref{fig:5}.
357: \begin{figure}[t]  
358: \centering
359: \includegraphics[width=0.48\textwidth]{genus-3D-over-energy-COL.final.eps}
360:  \caption{The ratio genus/energy of an homopolymer on a cubic lattice, as a 
361: function of $T$, at different values of $L$.}  
362: \label{fig:5}
363: \end{figure}
364: All these results confirm that the genus distribution behaves
365: differently in the two phases, as expected.  They also quantify how
366: much the restrictions induced by the actual three-dimensional
367: arrangement of the chain can limit the number and complexity of
368: pseudoknots (compared to \cite{PRL}). We find values closer to what
369: seems to happen in pseudoknots of real RNA molecules. In fact, real
370: RNA molecules typically have small genus. For instance, a simple
371: $H$-type pseudoknot ($\sim$20 bases or more) or the classical
372: kissing-hairpins pseudoknot ($\sim$30 bases or more) both have genus
373: 1, much less than the toy-model prediction in \cite{PRL}.  Even tRNAs
374: ($\sim$80 bases) mostly contain 4 helices, two of them linked together
375: by a kissing-hairpin pseudoknot, has still genus 1. Typical tmRNAs
376: ($\sim$350 bases long) contain four $H$-type pseudoknots, and its
377: total genus is 4, far below the theoretical upper bound $L/4$.  Our
378: numerical results would instead indicate, for instance for a $80$
379: bases long homopolymer, a genus of about $11.2$ in three dimensions
380: ($\sim$10.5 in two dimensions). Even if it is smaller than the value
381: suggested in \cite{PRL} (because of the steric constraints), it is
382: still too high when compared to real RNA molecules.  The obvious
383: reasons are that we neither included the primary sequence nor
384: realistic stacking energies.  We have nevertheless been able to
385: quantify the general effect of steric constraints on the genus
386: distribution of a pseudoknotted homopolymer on a lattice, as a first
387: step towards a model which includes a more realistic energy
388: function.\\
389: \underline{Acknowledgements}: 
390: We wish to thank T. Garel and R. Guida  for  discussions.  
391: This work was supported in part by the National Science 
392: Foundation under grant number PHY 99-07949, and by 
393: Sonderforschungsbereich-Transregio ``Computational Particle Physics''
394: (SFB-TR9). GV acknowledges  the support of the European Fellowship 
395: MEIF-CT-2003-501547. 
396: 
397: \begin{thebibliography}{99} 
398: \bibitem{SBS} R. Schroeder, A. Barta and  K. Semrad, 
399: Nature Rev. Mol. Cell Biol. {\bf 5}, 908 (2004). 
400: %908-919.
401: %STRATEGIES FOR RNA FOLDING AND ASSEMBLY
402: 
403: \bibitem{tinoco} I. Tinoco Jr. and C. Bustamante, 
404: J. Mol. Biol. {\bf 293}, 271 (1999).  
405: %How RNA folds
406: 
407: \bibitem{Higgs}  P.G. Higgs, 
408: Quart. Rev. Biophys. {\bf 33}, 199 (2000). 
409: %199-253. 
410: %RNA Secondary Structure: Physical and Computational Aspects. 
411: 
412: %\bibitem{rasmol} The 3d plots are generated with RasMol: R. 
413: %Sayle and E.J. Milner-White, Trends in Biochem. Sci. {\bf 20},  374 
414: %(1995).
415: %"RasMol: Biomolecular graphics for all", 
416: 
417: \bibitem{mfold} M. Zuker, Nucleic Acids Res. {\bf 31}, 3406 (2003);
418: %3406-3415
419: % Mfold web server for nucleic acid folding and hybridization prediction
420: See also  I.L. Hofacker, Nucleic Acids Research, {\bf 31}, 3429(2003). 
421: %3429-3431.
422: %Vienna RNA secondary structure server 
423: 
424: \bibitem{lyngso}  R.B. Lyngs{\o}  and  C.N. Pedersen, 
425: J. Comput. Biol.  {\bf 7}, 409 (2000).
426: %409-427.
427: %RNA Pseudoknot Prediction in Energy-Based Models
428: 
429: \bibitem{Bosch} C.W. Pleij, K. Rietveld and  L. Bosch, 
430: Nucleic Acids Res. {\bf 11}, 1717 (1985). 
431: %1717-1731.
432: % A new principle of RNA folding based on pseudoknotting
433: 
434: \bibitem{c1} L.X. Shen and I. Tinoco Jr, 
435: J. Mol. Biol. {\bf 247}, 963 (1995).
436: % 963-978.
437: %The structure of an RNA pseudoknot that causes efficient frameshifting in mouse mammary tumor virus. 
438: 
439: \bibitem{c2} P.L. Adams, M.R. Stahley, A.B. Kosek, J. Wang and S.A. Strobel, 
440: Nature {\bf 430}, 45 (2004).
441: %45-50.
442: %Crystal structure of a self-splicing group I intron with both exons. 
443: 
444: \bibitem{plos} D.W. Staple and  S.E. Butcher, 
445: PLoS Biol {\bf 3}, 213 (2005).
446: %Pseudoknots: RNA Structures with Diverse Functions.
447: 
448: \bibitem{eddyrivas} E. Rivas and S.R. Eddy, 
449: J. Mol. Biol. {\bf 285}, 2053 (1999).  
450: 
451: \bibitem{POZ} M.~Pillsbury, H.~Orland and A.~Zee, 
452: Phys. Rev. E {\bf 72}, 011911 (2005).
453: 
454: \bibitem{PTOZ} M.~Pillsbury, J.A.~Taylor, H.~Orland and A. Zee, 
455: http:// arXiv.org/\-cond-mat/\-0310505.  
456: %An Algorithm for RNA Pseudoknots 
457: 
458: \bibitem{HI} A. Xayaphoummine, T. Bucher and H. Isambert, 
459: Nucleic Acids Res. {\bf 33}, 605 (2005). 
460: %605-610. 
461: %Kinefold web server for RNA/DNA folding path and structure prediction including  pseudoknots and knots
462: 
463: \bibitem{Gul} A.P. Gultyaev, 
464: Nucleic Acids Res. {\bf 19}, 2489 (1991).
465: % 2489-2494.
466: %The computer simulation of RNA folding involving pseudoknot formation. 
467:  
468: \bibitem{abrah} J.P. Abrahams, M. van den Berg, E. van Batenburg and C.W.A.  Pleij, 
469: Nucleic Acids Res. {\bf 18}, 3035 (1990).
470: %3035-3044.   
471: %Prediction of RNA secondary structure, including pseudoknotting, by computer simulation. Nucleic Acids Res. 18, 3035-3044.   
472: 
473: \bibitem{Tab}  J.E. Tabaska, R. B. Cary, H.N. Gabow and G.D. Stormo,  
474: Bioinformatics  {\bf 14}, 691 (1998).
475: % 691-699.
476: %An RNA folding method capable of identifying pseudoknots and base triples.
477:  
478: \bibitem{eddyBIO}  S.R. Eddy,  
479: Nature Biotechnology  {\bf 22}, 1457 (2004).
480: % 1457-1458.
481: %How do RNA folding algorithms work?
482: 
483: \bibitem{OZ} H. Orland and A. Zee,  
484: Nucl. Phys. B  {\bf 620}, 456  (2002).
485: % 456-476. 
486: 
487: \bibitem{thooft}  G. 't Hooft, 
488: Nucl. Phys. B {\bf 72}, 461 (1974). 
489: 
490: %\bibitem{VOZ} G.~Vernizzi, H.~Orland and A.~Zee, 
491: %http://arxiv.org/q-bio.BM/0405014; see also G.~Vernizzi, H.~Orland
492: %Acta Phys. Pol. B {\bf 36}, 2821 (2005); A.~Zee, {\it ibid.} {\bf 36},
493: %2829 (2005).
494: 
495: \bibitem{VOZ} G.~Vernizzi, H.~Orland and A.~Zee, 
496: http://arxiv.org/q- bio.BM/0405014; see also Acta Phys. Pol. B {\bf
497: 36}, 2821 and 2829 (2005).
498: 
499: 
500: \bibitem{stella} A. Kabak\c{c}io\v{g}lu and A.L. Stella, 
501: Phys. Rev. E {\bf 70}, 011802 (2004).
502: %Pseudoknots in a homopolymer
503: 
504: \bibitem{Aeddy} E. Rivas and S.R. Eddy, 
505: Bioinformatics {\bf 16}, 334 (2000).
506: % 334-340.
507: %The language of RNA: a formal grammar that includes pseudoknots 
508: 
509: \bibitem{Lucas} A. Lucas and K.A. Dill, 
510: J. Chem. Phys. {\bf 119}, 2414 (2003).
511: %% 2414-2421.
512: %Statistical mechanics of pseudoknot polymers
513: 
514: \bibitem{PRL} G. Vernizzi, H. Orland and A. Zee, 
515: Phys. Rev. Lett. {\bf 94} 168103 (2005); Virt. J. Biol. Phys. Res. 9,
516: May 1, 2005.
517: %Enumeration of RNA structures by Matrix Models.
518: 
519: \bibitem{degennes} P.G. De Gennes, {\it Scaling Concepts in Polymer Physics}, 
520: Cornell University Press (1979); C. Vanderzande,{\it Lattice Models of
521: Polymers}, Cambridge University Press (1998).
522: 
523: \bibitem{GOO} T. Garel, H. Orland and E. Orlandini, 
524: Eur. Phys. J. B 12, 261-268 (1999).
525: 
526: \bibitem{FS} D.P. Foster and F. Seno, 
527: J. Phys. A {\bf 34}, 9939 (2001).
528: % 9939-9957
529: %Two-dimensional self-avoiding walk with hydrogen-like bonding: phase diagram and critical behaviour
530: 
531: \bibitem{Tiana} J. Borg, M.H. Jensen, K. Sneppen and G. Tiana, 
532: Phys. Rev. Lett. {\bf 86}, 1031 (2001).
533: % 1031-1033.
534: %Hydrogen bonds in polymer folding, Phys. Rev. Lett., 86(2001) 1031 [article]
535: 
536: 
537: \bibitem{orl} M. Baiesi, E. Orlandini and A.L. Stella,
538: Phys. Rev Lett. {\bf 91}, 198102 (2003);
539: %RNA denaturation: excluded volume, pseudoknots and transition scenarios
540: P. Leoni and C. Vanderzande, Phys. Rev. E. {\bf 68}, 051904 (2003).
541: %Statistical mechanics of RNA folding: a lattice approach
542: 
543: \bibitem{GO} T. Garel and H. Orland, 
544: J. Phys. A {\bf 23}, L621 (1990).
545: % Guided replication of random chains: a new Monte Carlo method
546: 
547: \bibitem{HO} P.G. Higgs and H. Orland, J. Chem. Phys. {\bf 95}, 4506 
548: (1991).
549: 
550: \bibitem{Iba} Y. Iba, Trans. Jap. Soc. for Artif. Intell. {\bf 
551: 16}, 279 (2001).
552: %279-286.
553: %Population Monte Carlo algorithms
554: 
555: \bibitem{PHA} V. Provman, P.C. Hohenberg and A. Aharony, 
556: {\it Phase Transitions and Critical Phenomena}, 
557: edited by C. Domb and J.L Lebowitz, Academic Press, New York (1991). 
558: 
559: \bibitem{Nickel} B.G. Nickel, Macromolecules {\bf 24}, 1358 (1991).
560: 
561: \bibitem{grassberger} P. Grassberger, Phys. Rev. E {\bf 56}, 3682 (1997).
562: 
563: \bibitem{madras} B.Li, N. Madras and A.D. Sokal, J. Stat. Phys. {\bf 
564: 80}, 661 (1995).
565: %661-754.
566: 
567: \end{thebibliography} 
568: \end{document}
569: 
570: