q-bio0604014/CDG2.tex
1: \documentclass[aps,pre,groupedaddress]{revtex4}
2: %\documentclass[epsf]{article}
3: \usepackage{graphicx,amsmath}
4: \input{ggmacro}
5: \newcommand{\eq}{\begin{equation}}
6: \newcommand{\feq}{\end{equation}}
7: \newcommand{\eqn}{\begin{eqnarray}}
8: \newcommand{\feqn}{\end{eqnarray}}
9: \newcommand{\arr}{\begin{eqnarray*}}
10: \newcommand{\farr}{\end{eqnarray*}}
11: \newcommand{\beq}{\begin{equation}}
12: \newcommand{\eeq}{\end{equation}}
13: \newcommand{\bea}{\begin{eqnarray}}
14: \newcommand{\eea}{\end{eqnarray}}
15: \newcommand{\lb}{\label}
16: 
17: \def\Y{Yakushevich }
18: 
19: \def\Dpsi{\theta^{(a)}_{i+1}-\theta^{(a)}_{i}}
20: \def\psip{{\psi}}
21: \def\psim{{\chi}}
22: %\def\dpsip{{\psip_{n+1}-\psip_n}}
23: %\def\dpsim{{\psim_{n+1}-\psim_n}}
24: \def\dpsip{{\Delta_n\psip}}
25: \def\dpsim{{\Delta_n\psim}}
26: \def\spsip{{{\rm S}_n\psip}}
27: \def\spsim{{{\rm S}_n\psim}}
28: \def\Omegap{{\xi}}
29: \def\Omegam{{\eta}}
30: %\def\dOmegap{{\Omegap_{n+1}-\Omegap_n}}
31: %\def\dOmegam{{\Omegam_{n+1}-\Omegam_n}}
32: \def\dOmegap{{\Delta_n\Omegap}}
33: \def\dOmegam{{\Delta_n\Omegam}}
34: \def\sOmegap{{{\rm S}_n\Omegap}}
35: \def\sOmegam{{{\rm S}_n\Omegam}}
36: \def\a{\alpha}
37: \def\energy#1#2#3#4{$E^{^{#1}}_{_{(#2,#3)}}=#4$}
38: 
39: \numberwithin{equation}{section}
40: 
41: \begin{document}
42: 
43: \title{A composite model for DNA torsion dynamics\footnote{Work
44: supported in part by the Italian MIUR under the program COFIN2004,
45: as part of the PRIN project {\it ``Mathematical Models for DNA
46: Dynamics ($M^2 \times D^2$)''}.}}
47: 
48: \author{Mariano Cadoni}
49: \email{mariano.cadoni@ca.infn.it} \affiliation {Dipartimento di
50: Fisica, Universit\`a di Cagliari and INFN, Sezione di Cagliari,
51: Cittadella Universitaria 09042 Monserrato, Italy}
52: \author{Roberto De Leo}
53: \email{roberto.deleo@ca.infn.it} \affiliation {Dipartimento di
54: Fisica, Universit\`a di Cagliari and INFN, sezione di Cagliari,
55: Cittadella Universitaria 09042 Monserrato, Italy}
56: \author{Giuseppe Gaeta}
57: \email{gaeta@mat.unimi.it} \affiliation {Dipartimento di
58: Matematica, Universit\`a di Milano, via Saldini 50, I--20133
59: Milano, Italy}
60: 
61: \begin{abstract}\noindent
62: DNA torsion dynamics is essential in the transcription process; a
63: simple model for it, in reasonable agreement with experimental
64: observations, has been proposed by Yakushevich (Y) and developed
65: by several authors; in this, the DNA subunits made of a nucleoside
66: and the attached nitrogen bases are described by a single degree
67: of freedom. In this paper we propose and investigate, both
68: analytically and numerically, a ``composite'' version of the Y
69: model, in which the nucleoside and the base are described by
70: separate degrees of freedom. The model proposed here contains as a
71: particular case the Y model and shares with it many features and
72: results, but represents an improvement from both the conceptual
73: and the phenomenological point of view. It provides a more
74: realistic description of DNA and possibly a justification for the
75: use of models which consider the DNA chain as uniform. It shows
76: that the existence of solitons is a generic feature of the
77: underlying nonlinear dynamics and is to a large extent independent
78: of the detailed modelling of DNA. The model we consider supports
79: solitonic solutions, qualitatively and quantitatively very similar
80: to the Y solitons, in a fully realistic range of all the physical
81: parameters characterizing the DNA.
82: \end{abstract}
83: \maketitle
84: 
85: \section{Introduction}
86: 
87: The possibility that nonlinear excitations -- in particular, kink
88: solitons or breath\-ers -- in DNA chains play a functional role
89: has attracted the attention of biophysicists as well as nonlinear
90: scientists since the pioneering paper of Englander et al.
91: \cite{Eng}, and the works by Davydov on solitons in biological
92: systems \cite{Dav}.
93: 
94: A number of mechanical models of the DNA double chain have been
95: proposed over the years, focusing on different aspects of the DNA
96: molecule and on different biological, physical and chemical
97: processes in which DNA is involved.
98: 
99: Here we will not discuss these, but just refer the reader to the
100: discussions of such attempts given in the book by Yakushevich
101: \cite{YakuBook} and in the review paper by Peyrard \cite{PeyNLN}
102: (see also the conference \cite{PeyHouches}), also for what
103: concerns earlier attempts which constituted the basis on which the
104: models considered below were first formulated.\footnote{It should
105: be stressed that when we speak of ``mechanical models of DNA'' we
106: exclude consideration of the all-important interactions between
107: DNA and its environment. The latter includes at least the fluid in
108: which DNA is immersed, and interaction with this leads to energy
109: exchanges; one should thus include in the equations describing DNA
110: dynamics both dissipation terms and random terms due to
111: interaction with molecules in the fluid. We will work here at a
112: purely mechanical level, i.e. do not consider at present these
113: effects. Moreover, it should be mentioned that even forgetting
114: dissipative and brownian motion effects, one could consider
115: interaction with the solvent by including effective terms in the
116: intrapair potential $V_p$ (see below), as done e.g. in \cite{ZC};
117: it has been recently shown that in the context of the
118: Peyrard-Bishop model this leads to a sharpening of certain
119: transitions \cite{Web}.} Similarly, we will not describe the
120: structure and functioning of DNA, but just refer e.g. to
121: \cite{CD,FK2,Sae}. See also \cite{PeyHouches,Pushchino} for the
122: role of Nonlinear Dynamics modelling in the understanding of DNA,
123: and \cite{Lavery,Strick} for DNA single-molecule experiments
124: (these were initiated about fifteen years ago \cite{SFB}, but
125: their range and precision has dramatically increased in recent
126: years; the formation of bubbles in a double-stranded DNA has been
127: observed in \cite{Lib}).
128: 
129: In recent years, two models have been extensively studied in the
130: Nonlinear Physics literature; these are the model by Peyrard and
131: Bishop \cite{PB} (and the extensions of this formulated by Dauxois
132: \cite{DauPLA} and later on by Barbi, Cocco, Peyrard and Ruffo
133: \cite{BCP,BCPR}; see also Cocco and Monasson \cite{CM}. More
134: recent advances are discussed in \cite{PeyNLN} and
135: \cite{BLPT,CuS,TPM}) and the one by Yakushevich \cite{YakPLA}; we
136: will refer to these as the PB and the Y models respectively.
137: 
138: Original versions of these models are discussed in \cite{GRPD};
139: they are put in perspective within a ``hierarchy'' of DNA models
140: in \cite{YakPRE}. An attempt to blend together the two is given in
141: \cite{YakSBP}; see also \cite{Joy}. Interplay between radial and
142: torsional degrees of freedom is considered more organically in
143: \cite{BCP,BCPR}.
144: 
145: The PB model is primarily concerned with DNA denaturation, and
146: describes degrees of freedom related to ``straight'' (or
147: ``radial'') separation of the two helices which are wound together
148: in the DNA double helical molecule. On the other hand, the Y model
149: -- on which we focus in this note -- is primarily concerned with
150: rotational and torsional degrees of freedom of the DNA molecule,
151: which play a central role in the process of DNA transcription
152: \footnote{It is appropriate, in this context, to mention earlier
153: models proposed by Fedyanin, Gochev and Lisy \cite{Fedyanin}, Muto
154: {\it et al.} \cite{Muto}, Prohofsky \cite{Proho}, Takeno and Homma
155: \cite{Takeno}, van Zandt \cite{VanZ}, Yomosa \cite{Yomosa}, and
156: Zhang \cite{Zhang}.}.
157: 
158: In this model, one studies a system of nonlinear equations which
159: in the continuum limit reduce to a double sine-Gordon type
160: equation; the relevant nonlinear oscillations are kink solitons --
161: which are solitons in both dynamical and topological sense --
162: which describe the unwinding of the double helix in the
163: transcription region. The latter is a ``bubble'' of about 20
164: bases, to which RNA Polymerase (RNAP) binds in order to read the
165: base sequence and produce the RNA Messenger; the RNAP travels
166: along the DNA double chain, and so does the unwound region. The
167: proposal of Englander et al. \cite{Eng} was that if the nonlinear
168: excitations are not created or forced by the RNAP but are anyway
169: present due to the nonlinear dynamics of the DNA double helix
170: itself, a number of questions -- in particular, concerning energy
171: flows -- receive a simple explanation.
172: 
173: The Y model has been studied in a number of paper, in particular
174: for what concerns its solitonic solutions; here we will quote in
175: particular \cite{GaePLA1,GaePLA2,Gonz,YakSBP,YakPhD,YakPRE}. It
176: has been shown that it gives a correct prediction of quantities
177: related to small amplitude dynamics, such as the frequency of
178: small torsional oscillations; as well as of quantities related to
179: fully nonlinear dynamics, such as the size of solitonic
180: excitations describing transcription bubbles \cite{GRPD,YakuBook}.
181: Moreover, in its ``helicoidal'' version, it provides a scenario
182: for the formation of nonlinear excitation out of linear normal
183: modes lying at the bottom of the dispersion relation branches
184: \cite{GRPD}. On the other hand, if we try to fit the observed
185: speed of waves along the chain \cite{YakuBook}, this is possible
186: only upon assuming unphysical values for the coupling constants
187: \cite{YakPRE}.
188: 
189: The Y model is also a very simple one, and adopts the same kind of
190: simplification as in the PB model. In particular, two quite strong
191: features of the models are:
192: \begin{itemize}
193: \item {\it (a)} there is a single (angular) degree of freedom for
194: each nucleotide; \item $(b)$ all bases are considered as
195: identical.
196: \end{itemize}
197: 
198: These were in a sense at the basis of the success of the model, in
199: that thanks to these features the model can be solved exactly and
200: one can check that predictions allowed by the model correspond to
201: the real world situations for certain specific quantities. But the
202: features mentioned above are of course not in agreement with the
203: real situation.
204: 
205: Indeed, it is well known that bases are quite different from each
206: other, and in particular purines are much bigger than pirimidines;
207: hence feature (b) -- albeit necessary for an analytical treatment
208: of the model -- is definitely unrealistic.
209: Moreover, it is quite justified to consider several groups of
210: atoms within a single nucleotide (the phosphodiester chain, the
211: sugar ring, and the nitrogen base) as substantially rigid
212: subunits; but these -- in particular the sugar ring -- have some
213: degree of flexibility, and what's more they have a considerable
214: freedom of displacement -- in particular for what concerns
215: torsional and rotational movements -- with respect to each other.
216: Thus, even in a simple modelling, feature (a) is not justified
217: {\it per se}, and it seems quite appropriate to consider several
218: subunits within each nucleotide. In this sense, we will speak of a
219: {\it composite Yakushevich (Y) model}.
220: Needless to say, if such a more detailed modelling would produce
221: results very near to those of the simple Y model, this should be
222: seen as a confirmation that the latter correctly captures the
223: relevant features of DNA torsional dynamics -- hence justify {\it
224: a posteriori} feature (a) of the Y model.
225: 
226: In this work we propose and study a composite Y model (in the
227: sense mentioned above), in which we describe  with two independent
228:  angular degrees of freedom, the nucleoside (i.e. the segment of the
229: sugar-phosphate backbone pertaining to the nucleotide) and the
230: nitrogen base in each nucleotide.
231: 
232: The purpose of our study of such a model is to shed light on the
233: following points.
234: 
235: \begin{itemize}
236: \item {\it (A)} We want to take care (to some
237: extent) of correcting feature (a) above and thus investigating --
238: by comparing results -- how justified is the original Y modelling
239: in terms of one degree of freedom per nucleotide.
240: 
241: \item {\it (B)} We  aim at opening the way
242: to correct -- or justify -- feature (b) of the Y model. Indeed, in
243: our model we will consider separately the part of the nucleotide
244: which precisely replicates identical in each nucleotide (the unit
245: of the sugar-phosphate backbone), and the part which varies from
246: one nucleotide to the other (the bases). We will then be able to
247: study the different role of the two.
248: 
249: \item{\it (C)} We want to check the dependence of the solitonic
250: solution of the  model both from the geometry and from the value
251: of the physical parameters chosen. In particular, we would like to
252: understand how far the existence of solitons is a generic feature
253: of DNA and if a more realistic choice of the model geometry is
254: consistent with phenomenologically acceptable values of the
255: physical parameters.
256: \end{itemize}
257: 
258: It will turn out that the Y model, which can be considered as a
259: particular case of our model, captures the essential features of
260: DNA nonlinear dynamics. The more realistic geometry of the model
261: we use in this paper enables a drastic improvement of the
262: descriptive power of our model at both the conceptual and the
263: phenomenological level: on the one hand the composite Y model
264: keeps almost all the relevant features of the Y model, but on the
265: other hand it allows for more realistic choice of the physical
266: parameters.
267: 
268: It will turn out that the different degrees of freedom we use play
269: a fundamentally different role in the description of DNA nonlinear
270: dynamics. The backbone degrees of freedom are ``topological'' and
271: play to some extent a ``master'' role, while those associated to
272: the base are ``nontopological'' and are (in fluid dynamics
273: language) ``slaved'' to former ones. This opens an interesting
274: possibility, i.e. to consider a more realistic model, in which
275: differences among bases are properly considered, as a perturbation
276: of our idealized uniform model. As the essential features of the
277: fully nonlinear dynamics are related only to backbone degrees of
278: freedom, we expect that such a perturbation -- albeit with
279: relevant difference in the quantitative values of some parameters
280: entering in the model (the base dynamical and geometrical
281: parameters) -- will show the same kind of nonlinear dynamics as
282: our uniform model studied here.
283: 
284: The paper is organized as follows. In Sect. \ref{s2}  we will
285: briefly review some basic know facts about the DNA  structure and
286: modelling. In Sects.  \ref{s3} and \ref{s4} we will set up our
287: model,  describe the interaction and write down the equations of
288: motions that govern its dynamics. The physical parameters
289: characterizing our model are discussed in Sect. \ref{s5}. In Sect.
290: \ref{s6} we discuss the linear approximation  of our dynamical
291: system, in particular its dispersion relations.  In Sect. \ref{s7}
292: we will set up the framework for the investigation of the
293: nonlinear dynamics and  the topological excitations of our model.
294: In sect. \ref{s8}, we will show how the Y model and Y solitons
295: emerge as a particular case of our composite Y model and its
296: solitons. Sect. \ref{s9}  we investigate and derive numerically
297: the solitonic solutions of our model. Finally in Sect. \ref{s10}
298: we summarize our work and present our conclusions.
299: 
300: 
301: 
302: \section{DNA structure and modelling}
303: \lb{s2}
304: 
305: DNA is a gigantic polymer, made of two helices wound together; the
306: helices have a directionality and the two helices making a DNA
307: molecule run in opposite direction. We will refer for definiteness
308: to the standard conformation of the molecule (B-DNA); in this the
309: pitch of the helix corresponds to ten base pairs, and the distance
310: along the axis of the helix between successive base pairs is $\de
311: = 3.4~\AA$.
312: 
313: The general structure of each helix can be described as follows.
314: The helix is made of a sugar-phosphate backbone, to which bases
315: are attached. The backbone has a regular structure consisting of
316: repeated identical units ({\it nucleosides}); bases are attached
317: to a specific site on each nucleoside and are of four possible
318: types. These are either purines, which are adenine (A) or guanine
319: (G), or pyrimidines, which are cytosine (C) or thymin (T) in DNA.
320: It should be noted that the bases are rather rigid structures, and
321: have an essentially planar configuration. A unit of each helix is
322: called a {\it nucleotide}; this is the complex of a nucleoside and
323: the attached base.
324: 
325: The winding together of the two helices makes that to each base
326: site on the one helix corresponds a base site on the other helix.
327: Bases at corresponding sites form a base pair; each base has only
328: a possible partner in a base pair; bases
329: in a pair are linked together via hydrogen bonds (two for A-T
330: pairs, three for G-C pairs). The base pairs can be ``opened''
331: quite easily, the dissociation energy for each H-bond being of the
332: order of 0.04 eV, hence $\Delta E \simeq 0.1$ eV per base pair.
333: Opening is instrumental to a number of processes undergone by DNA,
334: among which notably transcription, denaturation and replication.
335: 
336: The backbone structure (see Fig.1)  is made of a phosphodiester
337: chain and a sugar ring. To one of the $C$ atoms of the sugar ring
338: is attached a base; this is one of the four possible bases
339: $A,C,G,T$, whose sequence represents the information content of
340: DNA and is different for different species, and to some extent for
341: each individual.
342: Thus, each helix is made of a succession of identical nucleosides,
343: and attached bases which can be different at each site. Bases at
344: corresponding sites on the two helices form a base pair, and these
345: can be only of two types, G-C and A-T.
346: 
347: \begin{figure}
348: \includegraphics[width=300pt]{DNA.ps}\\
349:   \caption{The structure of a DNA helix. A nucleotide is shown; nitrogen bases
350:   are attached to the $C_1$ atom in the sugar ring.}\label{figDNA}
351: \end{figure}
352: 
353: %\subsection{Interactions}
354: 
355: The atoms on each helix are of course held together by covalent
356: bonds; apart from these, other interactions should be taken into
357: account when attempting a description of the DNA molecule.
358: The backbone structure has some rigidity; in particular, it would
359: resist movements which represent a torsion of one nucleoside with
360: respect to neighboring ones. We will refer to the interaction
361: responsible for the forces resisting these torsion as {\it
362: torsional interactions}.
363: 
364: As already mentioned, the two bases composing a base pair are
365: linked together by hydrogen bonds; we will refer to the
366: interaction mediated by these as {\it pairing interaction}.
367: Each base interacts with bases at neighboring sites on the same
368: chain via electrostatic forces (bases are strongly polar); these
369: make energetically favorable the conformation in which bases are
370: stacked on top of each other, and therefore are referred to as
371: {\it stacking interaction}.
372: 
373: Finally, water filaments -- thus, essentially, bridges of hydrogen
374: bonds -- link units at different sites; these are also known as
375: Bernal-Fowler filaments \cite{Dav}. In particular, they have a
376: good probability to form between nucleosides or bases which are
377: half-turn of the helix apart on different chains, i.e. which are
378: near to each other in space due to the double helix geometry;
379: these water filaments-mediated interactions are therefore also
380: called {\it helicoidal interactions} \footnote{It is usual, for
381: ease of language, to refer to models in which helicoidal
382: interactions are taken into account as ``helicoidal'', and to
383: models in which they are overlooked as ``planar''. Needless to
384: say, the geometry of the model is the same in both cases.}.
385: We stress that these are quite weaker than other interactions, and
386: can be safely overlooked when we consider the fully nonlinear
387: regime. They are instead of special interest when discussing small
388: amplitude (low energy) dynamics, as -- just because of their
389: weakness -- they are easily excited and introduce a length scale
390: in the dispersion relations (see below).
391: 
392: If we consider large amplitude deviations from the equilibrium
393: configurations, then motions will not be completely free: the
394: molecule is densely packed, and the presence of the
395: sugar-phosphate backbone -- and of neighboring bases -- will cause
396: steric hindrances to the base movements. In particular, for the
397: rotations in a plane perpendicular to the double helix axis, the
398: bases will not be able to rotate around the $C_1$ atom for more
399: than a maximum angle $\phi_0$ without colliding with the
400: nucleoside.
401: 
402: This will lead of course to complex behaviors as the DNA helix
403: gets unwound; in particular, as $\phi$ gets near to its limit
404: value $\phi_0$ we expect some kind of essentially (if not
405: mathematically) discontinuous behavior. This should not be seen as
406: shortcoming of the model: it is indeed well known that bases
407: rotate in a complex way while flipping about the DNA axis (see
408: e.g. \cite{Banavali}).
409: 
410: Finally, we mention that here we consider a DNA molecule without
411: taking into account its macro-conformational features; that is, we
412: consider an ``ideal'' molecule, disregarding supercoiling,
413: organization in istones, and all that \cite{CD}.
414: 
415: 
416: \section{Composite Y model}
417: \lb{s3}
418: 
419: As mentioned above, we will model the molecule as made of
420: different parts (units), each of them behaving as a single
421: element, i.e. as a rigid body. We consider each nucleoside N as a
422: unit, to which a base B (considered again as a single unit) is
423: attached.
424: 
425: \subsection{General features}
426: 
427: We will hence model each of the helices in the DNA double chain as
428: an array of elements (nucleotides) made of two subunits; one of
429: these subunits model the nucleoside, the other the nitrogen base.
430: We will consider the bases as all equal, thus disregarding the
431: substantial difference between them \footnote{This assumption (see
432: also section 9) is common to all the mathematical -- as opposed to
433: physico-chemical -- models of DNA, and as already remarked is
434: necessary to be able to perform an analytical study of the model.
435: We refer e.g. to \cite{PeyNLN,YakuBook} for discussion about this
436: point. Study of real sequences, i.e. with different
437: characteristics for different bases, is possible numerically; see
438: e.g. \cite{Sal,ZC}.}. The chains -- and thus the arrays -- will be
439: considered as infinite.
440: 
441: We will use a superscript $a=1,2$ to distinguish elements on the
442: two chains, and a subscript $i \in {\bf Z}$ to identify the site
443: on the chains. Thus the base pairing will be between bases
444: $B_i^{(1)}$ and $B_i^{(2)}$, while stacking interaction will be
445: between base $B_i^{(a)}$ and bases $B_{i+1}^{(a)}$ and
446: $B_{i-1}^{(a)}$.
447: 
448: We will consider each nucleoside $N^{(a)}_i$ as a disk; bases will
449: be seen as disks themselves, with a point on the border of
450: $B^{(a)}_i$ attached via an inextensible rod to a point $p_c$ on
451: the border of $N^{(a)}_i$; these points on B and N represent the
452: locations of the $N$ atom on B and of the $C_1$ atom on N involved
453: in the chemical bond attaching the base to the nucleoside. The rod
454: can rotate by an angle $\pm \phi_0$ before B collides with N; on
455: the other hand, the disk N can rotate completely around its axis.
456: 
457: \begin{figure}
458: \includegraphics[width=300pt,bb = 110 610 502 762]{modello1.ps}\\
459: % \includegraphics[width=300pt]{modello1.ps}\\
460:   \caption{A base pair in our model. The origin of the coordinate system is in O.
461:   The angles $\theta_1$ between the lines AO and AB and $\theta_2$
462:   between A'O and A'B' correspond
463:   to torsion of sugar-phosphate backbone with respect to the equilibrium
464:   B-DNA conformation; the angles $\phi_1$ between the line AB and the line BC, and
465:   $\phi_2$ between A'B' and B'C' correspond to rotation of bases around
466:   the $C_1 - N$ bond linking them to the nucleotide. All angles are in counterclockwise
467:   direction; thus the angles $\theta_2$ and $\phi_2$ in the figure are negative.}
468:   \label{modello1}
469: \end{figure}
470: 
471: 
472: 
473: 
474: We also single out a point $p_h$ on the border of the disk
475: modelling the base; this represents the atom(s) which form the H
476: bond with the corresponding base on the other DNA chain.
477: 
478: The disks (i.e. the elements of our model) are subject to
479: different kinds of forces, corresponding to those described above:
480: torsional forces resisting the rotation of one disk $N^{(a)}_i$
481: with respect to neighboring disks $N^{(a)}_{i+1}$ and
482: $N^{(a)}_{i-1}$ on the same chain; stacking forces between a base
483: $B_i^{(a)}$ and neighboring bases $B_{i+1}^{(a)}$ and
484: $B_{i-1}^{(a)}$ on the same chain; pairing forces between bases
485: $B_i^{(1)}$ and $B_i^{(2)}$ in the same base pair; and finally,
486: helicoidal forces correspond to hydrogen bonded Bernal-Fowler
487: filaments linking bases $B_i^{(1)}$ and $B_{i\pm 5}^{(2)}$ (and
488: $B_i^{(2)}$ and $B_{i\pm 5}^{(1)}$).
489: 
490: \subsection{The Lagrangian}
491: 
492: We should now translate the above discussion into a Lagrangian
493: defining our model. This will be written as
494: %
495: \beq\lb{lag} L \ = \ T
496: \ - \ \( U_t \ + \ U_s \ + \ U_p \ + \ U_h \) \feq
497: %
498: where $T$ is
499: the kinetic energy, and $U_a$ are the potential energies for the
500: different interactions listed above, i.e.
501: \begin{itemize}
502: \item $U_t$ is the backbone torsional potential, \item $U_s$ is
503: the stacking potential, \item $U_p$ is the pairing potential,
504: \item $U_h$ is the helicoidal potential. \end{itemize}
505: These will be modelled by two-body potentials, for which we use
506: the notation $V_a$, to be summed over all interacting pairs in
507: order to produce the $U_a$.
508: 
509: We denote by $I$ the moment of inertia (around center of mass) of
510: disks modelling the nucleosides, and by $I_B$ the moment of
511: inertia of bases around the $C_1$ atom in the sugar ring; as the
512: bases can not rotate around their center of mass, this is equal to
513: $m r^2$, where $m$ is the base mass and $r$ is the distance
514: between the $C_1$ atom in the sugar and the center of mass of the
515: base.
516: 
517: \subsection{The degrees of freedom}
518: 
519: We are primarily interested in the torsional dynamics. Thus, for
520: each element we will consider torsional movements, hence a
521: rotation angle (with respect to the equilibrium conformation,
522: which we take for definiteness to be B-DNA); these will be denoted
523: as $\theta^{(a)}_i$ for the nucleoside $N^{(a)}_i$, and
524: $\phi^{(a)}_i$ for the base $B^{(a)}_i$. Only these rotations will
525: be allowed in our model. All angles will be positive in
526: counterclockwise sense.
527: 
528: The angles $\theta$ represent a torsion of the sugar-phosphate
529: backbone with respect to the equilibrium configuration; thus they
530: are related to unwinding of the double helix. On the other hand,
531: the angles $\phi$ represent a rotation of the base with respect to
532: the corresponding nucleoside; the motion described by $\phi$ can
533: be thought of as a rotation around the $C_1$ atom in the sugar
534: ring. Note that the hindrances due to the presence of backbone
535: atoms constrain rotation of the base around the $C_1$ atom.
536: 
537: Thus, as mentioned above, the angles will have a different range
538: of values: \beq\lb{bc} \theta^{(a)}_i \in \R \ , \ \ \phi^{(a)}_i
539: \in [- \phi_0,\phi_0] \ \ \phi_0 < \pi \ . \feq
540: %
541: The actual value of $\phi_0$ is not essential. The important
542: feature is that the base can not rotate freely around the $C_1$
543: atom, but only pivot between certain limits. At the level of the
544: numerical analysis the simplest way to implement the previous
545: boundary condition is to use a confining potential, which
546: reproduces  approximately the form of a box. For this reason in
547: Sect. \ref{s9} we will add to the Hamiltonian of the system a
548: confining potential $V_{w}= K \tan^{4}\phi^{(a)}$.
549: 
550: It should be stressed that -- just on the basis of these different
551: ranges of variations -- there will be a substantial difference
552: between the degrees of freedom described by the angles: those
553: described by $\theta$ angles will be {\it topological} degrees of
554: freedom, while those described by $\phi$ angles will only describe
555: local and (relatively) small motions -- hence $\phi$ describe {\it
556: non topological} degrees of freedom.
557: 
558: \subsection{Cartesian coordinates}
559: 
560: In computing the kinetic energy, it will be convenient to consider
561: cartesian coordinates. With reference to Fig. \ref{modello1}, the cartesian
562: coordinates in the $(x,y)$ plane orthogonal to the double helix
563: axis of relevant points will be as follow.
564: 
565: The center of disks, representing the position of the
566: phosphodiester chain, will be $(x^{(a)}_o , y^{(a)}_o)$; the point
567: on the border of the disks representing the $C_1$ atom to which
568: the base disks are attached will be $(x^{(a)}_c , y^{(a)}_c)$. The
569: center of mass of the bases will be $(x^{(a)}_b , y^{(a)}_b)$, and
570: the point on the border of the disks modelling bases representing
571: the atom(s) forming the H bonds will be $(x^{(a)}_h , y^{(a)}_h)$.
572: 
573: In terms of the $\{ \theta , \phi \}$ angles, these are given by
574: (we omit the site index $i$ for ease of writing, and give
575: condensed formulas for the two chains, with first sign referring
576: to chain 1): \beq\lb{cartesian} \begin{array}{ll}
577: x^{(1,2)}_o = \mp a , & y^{(1,2)}_o = 0 ; \\
578: x^{(1,2)}_c = x^{(1,2)}_o \pm R \cos ( \theta^{(1,2)} ) , &
579: y^{(1,2)}_c = \pm R \sin ( \theta^{(1,2)} ) ; \\
580: x^{(1,2)}_b = x^{(1,2)}_c \pm r \cos ( \theta^{(1,2)} +
581: \phi^{(1,2)} ) , & y^{(1,2)}_b = y^{(1,2)}_c \pm r \sin (
582: \theta^{(1,2)} + \phi^{(1,2)} ) ; \\
583: x^{(1,2)}_h = x^{(1,2)}_c \pm d_h \cos ( \theta^{(1,2)} +
584: \phi^{(1,2)} ) , & y^{(1,2)}_h = y^{(1,2)}_c \pm d_h \sin (
585: \theta^{(1,2)} + \phi^{(1,2)} ). \\
586: \end{array} \feq
587: Here and in the following we denote by $R$ the radius of disks
588: describing nucleosides, i.e. the length of the segments AB and
589: A'B' in Fig. \ref{modello1} (this is the distance from the
590: phosphodiester chain to the $C_1$ atom); by $r$ the distance
591: between the center of mass of bases and the border of the disk
592: modelling the nucleoside (i.e. the $C_1$ atom). We also denote by
593: $d_h$ the lengths (supposed equal) of the segments BC and B'C'
594: joining the $C_1$ atom on the nucleoside and the atoms of the
595: bases forming the hydrogen bond linking this to the complementary
596: base. The parameter $a$ corresponds to the distance between the
597: double helix axis and the phosphodiester chain, whereas
598: $\rho_{0}$ is the distance between points $C$ and $C'$ in the
599: equilibrium configuration. The previous parameters are obviously
600: related by the equation $2a=2R+2d_{h}+\rho_{0}$.
601: 
602: \subsection{Kinetic energy}
603: 
604: With this notation, and standard computations, the kinetic energy
605: of each nucleotide is written as
606: $$ T_i^{(a)} = \frac {1}{2}
607: \[ m  r^2 {\dot\phi}^2 + 2  m  r  ( r + R \cos \phi )
608: {\dot\theta}  {\dot\phi} + \( I + m_B (R^2 + r^2 ) + 2 m R r
609:  \cos \phi \) {\dot\theta}^2 \] , $$
610:  where we have suppressed super- and sub-scripts for ease of
611:  reading. Thus, the total kinetic energy for the double chain is
612: \beq\lb{kin} \begin{array}{rl}
613: T \ =& \ \sum_a \ \sum_i \ T^{(a)}_i \ = \\
614:  =& \ \frac {1} {2} \, \sum_a \sum_i
615:  \ \[ m \, r^2 \, \( {\dot\phi}^{(a)}_i \)^2 \ + \ 2 \, m \, r \,
616:  ( r + R \cos (\phi^{(a)}_i) ) \,
617: {\dot\theta}^{(a)}_i \, {\dot\phi}^{(a)}_i \ + \right. \\
618:  & \left. + \ \( I + m_B (R^2 + r^2 ) + 2 m R r
619:  \cos (\phi^{(a)}_i) \) \( {\dot\theta}^{(a)}_i \)^2 \] \  \ .
620: \end{array} \feq
621: 
622: We have thus considered a general class of composite Y models; so
623: far we have not specified the interaction potentials, which are
624: needed to have a definite model.
625: 
626: 
627: \subsection{Modelling the interactions}
628: 
629: We have now to specify our model by fixing analytical expressions
630: for terms modelling potential interactions in the lagrangian
631: (\ref{lag}).
632: As one of our aims is to compare the results obtained by a
633: composite Y model with those obtained with the simple Y model, we
634: will make choices with the same physical content as those made by
635: Yakushevich.
636: 
637: \subsubsection{Torsional interactions}
638: 
639: Torsional forces will depend only on difference of angles
640: (measured with respect to the equilibrium B-DNA configuration) of
641: neighboring units on the same phosphodiester chain; thus we
642: introduce a torsion potential $V_t$ and have
643: %
644: \beq\lb{tp}
645: U_t \ = \ \sum_a
646: \sum_i V_t \( \theta^{(a)}_{i+1} - \theta^{(a)}_i \) \ . \feq
647: %
648: The potential $V_t$ must have a minimum in zero, and be $2
649: \pi$-periodic in order to take into account the fundamentally
650: discrete and quantum nature of the phosphodiester chain. Here we
651: will take the simplest such function \footnote{A more realistic
652: choice could have important consequences of qualitative -- and not
653: just quantitative -- behavior of nonlinear excitations. This point
654: is discussed in \cite{SacSgu}.}, i.e. (adding an inessential
655: constant so that the minimum corresponds to zero energy) \beq V_t
656: (x) \ = \ K_t \( 1 \ - \ \cos (x) \) \feq where $K_t$ is a
657: dimensional constant.
658: Thus, our choice for torsional interactions will be
659: \beq\lb{Ut}
660: U_t \ = \ K_t \ \sum_a \sum_i \, \[ 1 \, - \, \cos \(
661: \theta^{(a)}_{i+1} - \theta^{(a)}_i \) \] \ . \feq
662: %
663: The harmonic
664: approximation for this is of course $$ U_t^q \ = \frac{1}{2} K_t \
665: \sum_a \sum_i \, \( \theta^{(a)}_{i+1} - \theta^{(a)}_i \)^2  \ .
666: $$
667: 
668: \subsubsection{Stacking interactions}
669: 
670: Stacking between bases will only depend on the relative
671: displacement of neighboring bases on the same helix in the plane
672: orthogonal to the double helix axis \footnote{The stacking
673: interactions are of essentially electrostatic nature; thus it is
674: reasonable in this context to see the bases as dipoles. If we have
675: two identical dipoles made of charges $\pm \alpha$ a distance $d$
676: apart, their separation vector being along the $z$ direction, and
677: force them to move in parallel planes orthogonal to the $z$ axis
678: and a distance $L$ apart, then denoting with $\s$ the distance of
679: their projections in the $(x,y)$ plane we have
680: $$ V (\rho ) = (1/2) \a ((d+L)^{-3} - 2 L^{-3} + (d-L)^{-3})\s^2 - (3/8)
681: \a ((d+L)^{-5} - 2 L^{-5} + (d-L)^{-5}) \s^4 + O (\s^6) \ .
682: $$}. That is, introducing a stacking potential $V_s$ we have
683: %
684: \beq\lb{Us1} U_s \ = \ \sum_a
685: \sum_i V_s \( \s^{(a)}_{i} \) \ , \feq where \beq \s^{(a)}_i \ :=
686: \ \sqrt{(x^{(a)}_{i+1} - x^{(a)}_i)^2 \ + \ (y^{(a)}_{i+1} -
687: y^{(a)}_i)^2} \ , \feq
688: %
689: where $x^{(a)}_i,\,y^{(a)}_i$ are the coordinates of the center of
690: mass of the bases.
691: The simplest choice corresponds to a harmonic potential
692: \footnote{The interaction does more properly depend on the degree
693: of superposition of projections to bases on the plane orthogonal
694: to the double helix axis (and moreover depends on the details of
695: charge distribution on each base), and in particular quickly goes
696: to zero once the bases assume different positions. Moreover, once
697: the bases extrude from the double helix there are ionic
698: interactions between the bases and the solvent which should be
699: taken into account \cite{ZC}. In this note, however, we will just
700: consider harmonic stacking.}, $V_s = (1/2) K_s \s^2$. This will be
701: our choice -- which again corresponds to the one made in the PB
702: and in the Y models, so that
703: %
704: \beq U_s \ = \ \sum_a \sum_i
705: \frac{K_s}{ 2} \, (\s^{(a)}_{i} )^2 \ . \feq
706: %
707: We should however express this in terms of the $\theta$ and $\phi$
708: angles. With standard algebra, using Eqs. (\ref{cartesian}),  we obtain \beq\lb{Us}
709: \begin{array}{rl} U_s \ =& \ \frac{1}{2} \, K_s \ \sum_a \sum_i \
710: 2 \, \[ R^2 +
711: r^2 \ + \right. \\
712: & \left. - \ R^2 \cos (\theta^{(a)}_{i+1} - \theta^{(a)}_{i} ) \,
713: - \, r^2 \cos [ (\theta^{(a)}_{i+1} - \theta^{(a)}_{i} ) +
714: (\phi^{(a)}_{i+1} - \phi^{(a)}_{i}) ] \ + \right. \\
715:  & \left. - \ R r \( \cos [(\theta^{(a)}_{i+1} - \theta^{(a)}_{i})
716:  + \phi^{(a)}_{i+1} ] + \cos [(\theta^{(a)}_{i+1} - \theta^{(a)}_{i})
717:  - \phi^{(a)}_{i} ] \) \ + \right. \\
718:   & \left. + \ R r \( \cos (\phi^{(a)}_{i+1}) + \cos(\phi^{(a)}_{i})
719:  \) \] \ . \end{array} \feq
720: 
721: \subsubsection{Pairing interactions}
722: 
723: Pairing interactions are due to stretching of the hydrogen bonds
724: linking bases in a pair. Introducing a paring potential $V_p$
725: which models the H bonds, we have
726: %
727: \beq U_p \ = \ \sum_i \, V_p
728: (\theta^{(1)}_{i} , \theta^{(2)}_{i}, \phi^{(1)}_{i},
729: \phi^{(2)}_{i}) \ . \feq
730: %
731: We note that H bonds are strongly directional, so that they are
732: quickly disrupted once the alignment between pairing bases is
733: disrupted. This feature is traditionally disregarded in the Y
734: model, where it is assumed that $V_p$ only depends on the distance
735: %
736: \beq \rho_i \ := \sqrt{ ( x^{(1)}_{i} - x^{(2)}_{i} )^2  + (
737: y^{(1)}_{i} - y^{(2)}_{i} )^2} \feq between the interacting bases;
738: that is, \beq\lb{Up1} U_p \ = \ \sum_i \, V_p (\rho_{i} ) \ . \feq
739: %
740: As noted by Gonzalez and Martin-Landrove
741: \cite{Gonz} in the context of the Yakushevich model, one should be
742: careful in expanding a potential $V_p (\rho)$ in terms of the
743: rotation angles $\phi$ and $\theta$: indeed, unless $\rho_{0}=0$, i.e.
744: $a = R + d_h$,
745: one would get zero quadratic term in such an expansion (see
746: however \cite{GaeY1} for what concerns solitons in this context).
747: 
748: As for the potential $V_p$, there are two simple choices for this
749: appearing in the literature. On the one hand, Yakushevich
750: \cite{YakPLA} suggests to consider a potential harmonic in the
751: intrapair distance $\rho$ (this would appear nonlinear when
752: expressed through rotation angles) and this has been kept in
753: subsequent discussions and extensions of her model
754: \cite{YakuBook}; on the other hand, Peyrard and Bishop \cite{PB}
755: consider a Morse potential; again this has been kept in subsequent
756: discussions and extensions of their model \cite{PeyNLN}.
757: 
758: There is no doubt that the Morse potential is more justified in
759: physical terms; however, as we wish to compare our results with
760: those of the original Y model, we will at first
761: consider a harmonic
762: potential
763: %
764: \beq\lb{Vp} V_p^{(Y)} (\rho) \ = \ \frac{1}{2} \, K_p \,
765: (\rho - \rho_0 )^2 \ , \feq
766: %
767: where $\rho_0$ is the intrapair distance in the
768: equilibrium configuration. Moreover, again in order to compare our
769: results with those of the original Y model, we will later on set
770: $\rho_0 = 0$. This corresponds to setting $a = R + d_h$. These
771: approximations can appear very crude, but experience gained (as
772: preliminary work for the present investigation) with the standard
773: \Y model \cite{GaeY1,GaeY2} suggests they do not have a great
774: impact at the level of fully nonlinear dynamics.
775: 
776: We should express $V_p$ in terms of the rotations angles. Using
777: once again the expressions (\ref{cartesian}), we have with
778: standard computations that
779: \beq\lb{rho} \begin{array}{rl} \rho_i^2
780: \ :=& \( x^{(1)}_{i} - x^{(2)}_{i} \)^2 \ + \ \(
781: y^{(1)}_{i} - y^{(2)}_{i} \)^2 \ = \\
782:  =&
783: 2 \[ 2 a^2 + R^2 + d_{h}^2  +
784:     R^2 \cos (\theta^{(1)}_i  - \theta^{(2)}_i )  +
785:     d_{h}^2 \cos [(\theta^{(1)}_i  - \theta^{(2)}_i ) +
786:     (\phi^{(1)}_i  - \phi^{(2)}_i ) + \right. \\
787:  & \left. + \ R d_{h} \( \cos \phi^{(1)}_i  + \cos \phi^{(2)}_i +
788:     \cos [(\theta^{(1)}_i  - \theta^{(2)}_i ) + \phi^{(1)}_i  ] +
789:     \cos [(\theta^{(1)}_i  - \theta^{(2)}_i ) - \phi^{(2)}_i  ]  \) + \right. \\
790:  & \left. - 2 a R \( \cos (\theta^{(1)}_i ) + \cos (\theta^{(2)}_i
791:  )\) - 2 a d_{h}\(
792:     \cos (\phi^{(1)}_i  + \theta^{(1)}_i )  +
793:     \cos (\phi^{(2)}_i  + \theta^{(2)}_i ) \) \] \ . \end{array} \feq
794: With this, our choice for the pairing part of the hamiltonian will
795: be \beq\lb{Up} U_p \ = \ \sum_i \, V_p (\rho_i ) \ . \feq
796: 
797: \subsubsection{Helicoidal interactions}
798: 
799: Helicoidal interaction are mediated by water filaments
800: (Bernal-Fowler filaments \cite{Dav}) connecting different
801: nucleotides; in particular we will consider those being on
802: opposite helices at half-pitch distance, as they are near enough
803: in three-dimensional space due to the double helical geometry. As
804: the nucleotide move, the hydrogen bonds in these filaments -- and
805: those connecting the filaments to the nucleotides -- are stretched
806: and thus resist differential motions of the two connected
807: nucleotides.
808: 
809: We will, for the sake of simplicity and also in view of the small
810: energies involved, only consider filaments forming between
811: nucleosides; thus only the $\theta$ angles will be involved in
812: these interactions.
813: We have therefore, introducing a helicoidal potential $V_h$ and
814: recalling that the pitch of the helix corresponds to 10 bases in
815: the B-DNA equilibrium configuration, \beq\lb{Uh1} U_h \ = \ \sum_i
816: \, V_h (\theta^{(1)}_{i+5} - \theta^{(2)}_{i}) \ + \ V_h
817: (\theta^{(2)}_{i+5} - \theta^{(1)}_{i}) \ . \feq As the angles
818: $\theta$ are involved, the potential $V_h$ should be $2
819: \pi$-periodic \footnote{In physical terms, this is not obvious by
820: itself, as the filaments could have to wind around the double
821: helix if the two connected nucleosides are twisted by $2 \pi$ with
822: respect to each other; however, when this happens the filaments
823: are actually broken and then built again thanks to quantum
824: fluctuations.}.
825: 
826: Such water filament connections involve a large number (around 10)
827: of hydrogen bonds; hence each of them is only slightly stretched,
828: and it makes sense to consider the angular-harmonic approximation
829: \beq V_h (\tau ) \ = \ K_h [1 - \cos (\tau) ] \ \simeq \ \frac{1}{2} K_h
830: \tau^2  \ . \feq
831: 
832: Our choice will therefore be \beq\lb{Uh} U_h \ = \ K_h \, \sum_i
833: \, \[ 2 - \cos(\theta^{(1)}_{i+5} - \theta^{(2)}_{i}) -
834: \cos(\theta^{(2)}_{i+5} - \theta^{(1)}_{i}) \] \ . \feq
835: 
836: 
837: \section{Equations of motion}
838: \lb{s4}
839: 
840: In  the previous sections we have set up the  model and the
841: interactions. Let us now study its dynamics. We denote
842: collectively the variables as $\psi^a$, e.g. with $\psi =
843: (\phi^{(1)},\phi^{(2)},\theta^{(1)},\theta^{(2)} )$. The dynamics
844: of the model will be described by the Euler-Lagrange equations
845: corresponding to the Lagrangian (\ref{lag}) with the terms in the
846: interaction potential given respectively by Eqs. (\ref{Ut}),
847: (\ref{Us}), (\ref{Up}), (\ref{Uh}) \beq \frac{\pa L}{\pa \psi^a_i}
848: \ - \ \frac{\d } {\d t} \frac{\pa L} { \pa {\dot\psi}^a_i} \ = \ 0
849: \ . \feq With our choices for the different terms of $L$, and
850: writing $\hat a$ for the complementary chain of the chain $a$ (that
851: is, $\hat 1 = 2$, $\hat 2=1$), these read
852: \beq\lb{EuLag}
853: \begin{array}{l}
854: m r^2 {\ddot \phi}^{(a)}_i + m r [R \cos (\phi^{(a)}_i) + r]
855: {\ddot \theta}^{(a)}_i  + m r R \sin ( \phi^{(a)}_i ) ({\dot
856: \theta}^{(a)}_i )^2 \ = \\
857: \ = \ K_s r^2 \sin[\phi^{(a)}_{i-1} - \phi^{(a)}_{i} +
858: \theta^{(a)}_{i-1} - \theta^{(a)}_{i}] -
859:   2 a d_h K_p \sin (\phi^{(a)}_{i} + \theta^{(a)}_{i} )\\ -
860:   K_s r R \sin (\phi^{(a)}_{i} - \theta^{(a)}_{i-1} + \theta^{(a)}_{i} ) -
861:   K_s r R \sin ( \phi^{(a)}_{i} + \theta^{(a)}_{i} - \theta^{(a)}_{i+1} ) \\
862:   \ \ \ -
863:   K_s r^2 \sin ( \phi^{(a)}_{i} - \phi^{(a)}_{i+1} + \theta^{(a)}_{i}
864:   - \theta^{(a)}_{i+1} ) +
865:   d_h K_p R \sin ( \phi^{(a)}_{i} + \theta^{(a)}_{i} -
866:   \theta^{(\hat a)}_{i} ) +\\
867:   d_h^2 K_p \sin ( \phi^{(a)}_{i} - \phi^{(\hat a)}_{i} +
868:   \theta^{(a)}_{i} - \theta^{(\hat a)}_{i} ) +
869:   R (d_h K_p + 2 K_s r ) \sin[ \phi^{(a)}_{i} ] \ ; \\
870:   {} \\
871: m r R \cos ( \phi^{(a)}_{i} ) ({\ddot \phi}^{(a)}_{i} + 2 {\ddot
872: \theta}^{(a)}_{i} ) + m r^2 {\ddot \phi}^{(a)}_{i} + I {\ddot
873: \theta}^{(a)}_{i} +
874:   m r^2 {\ddot \theta}^{(a)}_{i} + m R^2 {\ddot \theta}^{(a)}_{i}\\
875:   -m r R \sin (\phi^{(a)}_{i}) {\dot \phi}^{(a)}_{i} ({\dot \phi}^{(a)}_{i} +
876:   2 {\dot \theta}^{(a)}_{i}) \ = \\
877: \ = \ (K_t + K_s R^2 ) \sin ( \theta^{(a)}_{i-1} -
878: \theta^{(a)}_{i} )
879:   + K_s r R \sin ( \phi^{(a)}_{i-1} - (\theta^{(a)}_{i} -
880:   \theta^{(a)}_{i-1}))\\
881:   -  K_s r^{2} \sin ( (\phi^{(a)}_{i} - \phi^{(a)}_{i-1})
882:   \ \ \ + (\theta^{(a)}_{i} - \theta^{(a)}_{i-1}) ) -
883:   2 a K_p R \sin (\theta^{(a)}_{i}) -\\
884:   2 a d_h K_p \sin ( \phi^{(a)}_{i} + \theta^{(a)}_{i} )
885:   -K_s r R \sin [ \phi^{(a)}_{i} + (\theta^{(a)}_{i} - \theta^{(a)}_{i-1} )
886:   ]  \\- K_{s} r R \sin (\phi^{(a)}_{i} - (\theta^{(a)}_{i+1} -
887:   \theta^{(a)}_{i}) )
888:   + (K_t + K_s R^2) \sin (\theta^{(a)}_{i+1} - \theta^{(a)}_{i} )\\
889:   + K_s r^2 \sin [(\phi^{(a)}_{i+1} - \phi^{(a)}_{i}) +
890:         (\theta^{(a)}_{i+1} - \theta^{(a)}_{i}) ]
891:    + K_s r R \sin [ \phi^{(a)}_{i+1} + (\theta^{(a)}_{i+1} -
892:   \theta^{(a)}_{i}) ] +\\
893:   K_p R^2 \sin ( \theta^{(a)}_{i} - \theta^{(\hat a)}_{i})  +
894:   d_h K_p R \sin (\phi^{(a)}_{i} + (\theta^{(a)}_{i} - \theta^{(\hat a)}_{i}) )
895:   \ \ \ +\\
896:   d_h^2 K_p \sin ((\phi^{(a)}_{i} - \phi^{(\hat a)}_{i}) +
897:   (\theta^{(a)}_{i} - \theta^{(\hat a)}_i ) ) -
898:   d_h K_p R \sin ( \phi^{(\hat a)}_{i} - (\theta^{(a)}_{i} -
899:   \theta^{(\hat a)}_{i}) )\\
900:     + K_h \( \theta^{(\hat a)}_{i+5} -
901:   2  \theta^{(a)}_{i} + \theta^{(\hat a)}_{i-5} \)
902: \end{array} \feq
903: Note that here $a,R,r,d_h$ are considered as independent
904: parameters, i.e. we have not enforced the Yakushevich condition $R
905: + d_h = a$ (i.e. $\rho_0 = 0$).
906: Needless to say, these are far too complex to be analyzed
907: directly, and we will need to introduce various kinds of
908: approximation.
909: 
910: We have thus completely specified the model we are going to study and
911:  derived the equations that govern its dynamics,
912: i.e. its lagrangian and the equations of motion. The choice of
913: torsion angles as variables to describe our dynamics led to
914: involved expressions, but our choices are very simple physically.
915: 
916: We have considered ``angular harmonic'' approximations (expansion up to
917: first Fourier mode) i.e. potentials of the form $V (x) = [1 - \cos
918: (x)]$ for the torsion and helicoidal interactions, harmonic
919: approximation for the base stacking interaction, and a harmonic
920: potential depending on the intrapair distance for the pairing
921: interaction.
922: Our approximations are coherent with those considered in the
923: literature when dealing with uniform models of the DNA chain, and
924: in particular when dealing with (extensions of) the Yakushevich
925: model.
926: Thus, when comparing the characteristic of our model with those of
927: these other models, we are really focusing on the differences
928: arising from considering separately the nucleoside and the base
929: within each nucleotide.
930: 
931: It would of course be possible to consider more realistic
932: expressions for the potentials; but we believe that at the present
933: stage this would rather obscure the relevant point here, i.e. the
934: discussion of how such ``composite'' models can retain the
935: remarkable good features of the Y model and at the same time
936: overcome some of the difficulties they encounter.
937: Finally, we note that it is quite obvious that the dynamical
938: equations describing the model are -- despite the simplifying
939: assumptions we made at various stage -- too hard to have any hope
940: to obtain a general solution, either in the discrete or in the
941: continuum version (see below) of the model.
942: 
943: In the next section we will   focus our attention on
944: the choice of the physical parameters appearing in our model.
945: Later,  we
946: will investigate the dynamics beginning with  the linear
947: approximation and then in the fully nonlinear regime.
948: 
949: \section{Physical values of parameters}
950: \lb{s5}
951: 
952: In order to have a well defined model
953: we should still assign concrete values to the
954: parameters -- both geometrical ones and coupling constants --
955: appearing in our Lagrangian (\ref{lag})  and in the  equation of
956: motion (\ref{EuLag}).
957: 
958: \subsection{Kinematical parameters}
959: 
960: Let us start by discussing kinematical parameters; in these we
961: include the geometrical parameters as well as the mass $m$ and the
962: moment of inertia $I$.
963: 
964: The masses can be readily evaluated by considering the chemical
965: structure of the bases. They can be calculated just by knowing
966: masses of the atoms and their multiplicity in the different bases.
967: As for the geometrical parameters like $R$, $a$, $r$ and $d_h$ (and the
968: moment of inertia $I$), quite surprisingly different authors seem
969: to provide different values for these. Rather than assuming the values
970: given by one or another author, we have preferred to estimate the
971: parameters using the available information about the DNA structure.
972: Position of atoms within the bases (which of course determine
973: $R$, $a$, $r$ and $d_h$, and hence $I$) and  geometrical descriptions of
974: DNA are widely available to the scientific community in form of
975: {\tt PDB} files \cite{PDBRep}. We will use this information
976: (which we accessed at \cite{1BNA} and \cite{PDBFiles}) to estimate
977: directly all static parameters in play on the basis of the atomic
978: positions.
979: 
980: The geometrical parameters which are relevant for our discussion
981: are the longitudinal width of bases  $l_{b}$ and of the sugar
982: $l_{s}$, the  distances of the bases from the relative sugars
983: $d_s$ and the distance of a base from  the relative dual base
984: $d_b$. We give our estimates for the masses, moments of inertia
985: and the parameters $l,d_{s}, d_{b}$ for the different bases and
986: their mean values in Table~\ref{tab:cynParms}.
987: >From those data and using the equations $ R = l_{s},\,
988: r = {\hat d}_{s}+ {\hat l}_{b}/2,\,
989: d_h =  {\hat l}_{b} + {\hat d}_{s},\,
990: a = l_{s} +
991: {\hat l}_{b} + {\hat d}_{s} + {\hat d}_{b} / 2$ (hats denote  mean
992: values),
993: one obtains  the average values for
994: the geometrical parameters appearing in our Lagrangian.
995: given in table \ref{tab:para}.
996: 
997: \begin{table}[ht]
998: \centering
999:  \begin{tabular}{|l|c|c|c|c|c|c|}
1000:   \hline
1001:     & A & T & G & C & mean & Sugar\\
1002:     \hline
1003:   $m$ & 134  & 125  & 150  & 110  & 130 & 85 \\
1004:   $I$ & $3.6\times10^3$ & $3.0\times10^3$ & $4.4\times10^3$ & $2.3\times10^3$ & $3.3\times10^3$ & $1.2\times10^2$\\
1005:   $l$ & $3.2$ & $4.0$ & $5.0$ & $2.4$ & $4.7$ & $3.3$\\
1006:   $d_s$ & $1.5$ & $1.5$ & $1.5$ & $1.5$ & $1.5$ & -\\
1007:   $d_b$ & $2.0$ & $2.0$ & $2.0$ & $2.0$ & $2.0$ & -\\
1008:  \hline
1009: \end{tabular}
1010: \caption{
1011:   Order of magnitude for the basic geometrical parameters of the DNA.
1012:   Units of measure are: atomic unit for masses $m$, $1.67 \times 10^{-47}
1013:   {\rm Kg}\cdot {\rm m}^2$
1014:   for the inertia momenta $I$, Angstrom for $l$, $d_s$ and $d_b$, respectively the
1015:   longitudinal width of bases and their distances from the relative sugars and
1016:   from the relative dual base. These values have
1017:   been extracted from the sample ``generic'' B-DNA PDB data~\cite{PDBFiles}, kindly
1018:   provided by the Glactone Project~\cite{Glac}, and double checked with the data from
1019:   \cite{1BNA}, that agree within $5\%$.
1020:   Inertia momenta of bases has been evaluated with respect to
1021:   rotations about the DNA's symmetry axis passing through the sugar's $C_1$ atom
1022:   the base is attached to; the inertia momentum of the sugar itself has been evaluated
1023:   with respect to rotations about its $C_3-C_4$ axis (see Fig.~\ref{figDNA})}
1024:  \label{tab:cynParms}
1025: \end{table}
1026: 
1027: \begin{table}[ht]
1028: \centering
1029:    \begin{tabular}{|c|c|c|c|}
1030:     \hline
1031:     $R$ & $r$ & $d_{h}Ą$ & $a$ \\
1032:     \hline
1033:     3.3~\AA  &3.8~\AA  & 6.2~\AA & 10.5~\AA  \\
1034:     \hline
1035:    \end{tabular}
1036:    \caption{Numerical values of the geometrical parameters
1037:    chracterizing our model}
1038:    \label{tab:para}
1039: \end{table}
1040: 
1041: \subsection{Coupling constants}
1042: 
1043: The determination of the four coupling constants appearing in our
1044: model is more
1045: problematic, due partly to the difficulties in making experiments
1046: to test single coupling constants and partly to the complexity of
1047: the system itself.
1048: 
1049: 
1050: 
1051: \subsubsection{Pairing}
1052: 
1053: The  coupling constant $K_p$, which appears in the pairing potential
1054: (\ref{Up}) can be easily determine by considering the typical energy
1055: of hydrogen bonds. The pairing interaction involves
1056: two (in the $A-T$ case) or three (in the $G-C$ case) electrostatic
1057: hydrogen bond. The pairing potential can be modelled
1058: with a Morse function
1059: \beq\lb{morse}
1060: V_p(x)=D(e^{-b d(x,x_0)}-1)^2=\frac{1}{2}(2D b^2)(\rho-\rho_0)^2+O(\rho^3)\,,
1061: \feq
1062: where $D$ is the potential depth, $\rho$ the distance from the
1063: equilibrium position $\rho_0$ and $b$ a parameter that defines
1064: the width of the well. Although throughout this paper we  use
1065: the harmonic potential (\ref{Up}) to model the pairing interaction,
1066: the use of the Morse function seems more appropriate for evaluating
1067: the parameter  $K_p$. The point is that the pairing coupling constant
1068: is physically determined by the behavior of the pairing potential
1069: away from its minimum. Using the harmonic approximation (\ref{Up})
1070: for estimate $K_p$ would result in a completely unphysical value for
1071: the parameter.
1072: 
1073: Different estimates  of the parameters appearing in the potential
1074: (\ref{morse}) are present in the literature.
1075: The estimates
1076: $$ D_{AT} = 0.030 {\rm eV} , \ D_{GC} = 0.045 {\rm eV}, \
1077: b_{AT} = 1.9 {\rm \AA}^{-1}, \ b_{GC} = 2.5 {\rm \AA}^{-1}
1078: $$ are given in \cite{CP} and used in \cite{ZC}.
1079: The values
1080: $$ D = 0.040 {\rm eV}, \ b = 4.45 {\rm \AA}^{-1} $$ are given in
1081: \cite{PBD} and used in \cite{PBD,BCP,DeLuca}.
1082: Finally,  the estimates
1083: $$ D_{AT} = 0.050 {\rm eV}, \ D_{GC} = 0.075 {\rm eV}, \
1084: b_{AT} = b_{GC} = 4 {\rm \AA}^{-1} $$ are given in \cite{Campa}
1085: and used in \cite{Campa,KS}. The values of coupling constants
1086: corresponding to these different values for the parameters
1087: appearing in the Morse potential range across a whole order of
1088: magnitude: \beq\lb{kd} 3.5 \, {\rm N/m} \ \leq \ K_p \, := \, 2
1089: b^2 D \ \leq \ 38 {\rm N/m} \ . \feq
1090: 
1091: In our numerical investigations we will use a value of $K_{p}$
1092: near to the lower bound given in \ref{kd}; that is, we adopt the
1093: value $K_p = 4 {\rm N/m}$, leading to an optical frequency of
1094: $\omega_0=\sqrt{2K_p/m}=36 {\rm cm}^{-1}$, so to be in agreement with
1095: \cite{Pow}.
1096: \bigskip
1097: 
1098: \subsubsection{Stacking}
1099: 
1100: The determination of the torsion and stacking coupling constants
1101: is more involved  and rests on a smaller amount of experimental
1102: data. The main information is the total torsional rigidity of the
1103: DNA chain  $C=S \de$, where $\delta=3.4 {\rm \AA}$
1104: is the base-pair spacing and $S$ is the torsional rigidity. It is
1105: known \cite{BZ,BFLG} that \beq 10^{-28}\, {\rm J  m} \ \leq \ C \
1106: \leq \ 4 \cdot 10^{-28} \, {\rm J  m} \ . \feq This information
1107: is used e.g. in \cite{Eng,Zhang}, whose estimate is based on the
1108: evaluation of the free energy of superhelical winding; this fixes
1109: the range for the total torsional energy to be \beq 180 \, {\rm K
1110: J/mol} \ \leq \ S \ \leq \ 720 \, {\rm K J/mol} \ . \feq
1111: 
1112: In our composite model the total torsional energy of the DNA chain
1113: has to be considered as the sum of two parts, the base stacking
1114: energy and the torsional energy of the sugar-phosphate backbone.
1115: In order to extract the stacking coupling constant we use the
1116: further information that $\pi-\pi$ stacking bonds amount at the
1117: most to $50 {\rm KJ/mol}$ \cite{SC}. Assuming a quadratic stacking
1118: potential, as we do, and a width of the potential well of about
1119: $2\AA$  we obtain the estimate $K_s = 68 {\rm N/m}$. The phonon
1120: speed induced by this is $c_1 = \de \sqrt{K_s/m} \simeq 6
1121: {\rm Km/s}$, see eq. (\ref{speed}); this is rather close to the
1122: the estimate of $1.8 {\rm Km/s} \leq c_1 \leq 3.5 {\rm Km/s}$
1123: given in \cite{YakPRE}.
1124: 
1125: As we shall see in detail in Sect. \ref{s9}, choosing smaller
1126: values for $K_s$ would have non-trivial consequences since
1127: solitons with small topological numbers become unstable in the
1128: discrete setting when the ratios $K_s/K_p$ and $K_t/K_p$ get small
1129: enough (see sect.~\ref{s9}). In particular, this value for $K_s$
1130: -- together with the $K_t$ below -- is barely enough to allow the
1131: existence of solitons, as discussed later on this paper.
1132: 
1133: \subsubsection{Torsion and helicoidal couplings}
1134: \label{sec:hel}
1135: 
1136: After extracting the stacking component, our estimate for the
1137: torsional coupling constant $K_t$ is in the range \beq 130 \, {\rm
1138: K J/mol} \ \leq \ K_t \ \leq \ 670 \, {\rm KJ/mol} \ . \feq Assuming
1139: (see below) that $K_h \simeq K_t/25$, so that
1140: $c_4=\sqrt{2K_t/I_s}$ (see eqs. (\ref{speed1}) and (\ref{speed})),
1141: all of these values for $K_t$ induce phonon speeds slightly higher
1142: with respect to the estimates cited above, between 5 Km/s and 11
1143: Km/s. For our numerical investigations, to keep the phonon speed as
1144: low as possible,  we will  set
1145: $K_t=130 {\rm KJ/mol}$.
1146: 
1147: Finally, for the helicoidal coupling constant, following
1148: \cite{GaeJBP}, we assume that $K_t$ and $K_h$ differ by about a
1149: factor $25$, so that $K_h=5 {\rm KJ/mol}$.
1150: 
1151: \subsubsection{Discussion}
1152: 
1153: It is interesting to point out how the geometry of the model
1154: nicely fits with the estimates of the binding energies so to
1155: induce optical frequencies and phonon speeds of the right order of
1156: magnitude (see also the discussion in Sect. \ref{s6}). This is not
1157: the case in simpler models, where in order to get the right phonon
1158: speed within a simple Y model one is obliged to assume for $K_t$
1159: the unphysical value $K_t = 6000 {\rm KJ/mol}$ \cite{YakPRE}.
1160: 
1161: Our estimates, and hence our choices for the values of the
1162: coupling constants appearing in our model, are summarized in
1163: table~\ref{tab:dynParms}. We will use these values of the physical
1164: parameters of DNA in the next sections, when discussing both the
1165: linear approximation and the dispersion relations as well as the
1166: full nonlinear regime and the solitonic solutions.
1167: 
1168: 
1169: \begin{table}[ht]
1170: \centering
1171:    \begin{tabular}{|c|c|c|c|}
1172:     \hline
1173:     $K_t$ & $K_s$ & $K_p$ & $K_h$ \\
1174:     \hline
1175:     130 KJ/mol & 68 N/m & 3.5 N/m & 5 KJ/mol \\
1176:     \hline
1177:    \end{tabular}
1178:    \caption{Values of the coupling constants for our DNA model}
1179:    \label{tab:dynParms}
1180: \end{table}
1181: 
1182: 
1183: 
1184: \section{Small amplitude excitations and dispersion relations}
1185: \lb{s6} In this section we will investigate the dynamical behavior
1186: of our model  for  small excitations in the linear regime. We will
1187: enforce the Yakushevich condition $R + d_h = a$ in order to keep
1188: the calculations and their results as simple as possible (see also
1189: \cite{GaeY1}).
1190: 
1191: Linearizing the   equation  of motion (\ref{EuLag}) around the
1192: equilibrium configuration
1193: %
1194: \beq \phi^{(a)}_i = \theta^{(a)}_i =
1195: {\dot\phi}^{(a)}_i = {\dot\theta}^{(a)}_i =  0,  \feq
1196: %
1197: we get using standard algebra,
1198: \beq\lb{linEL} \begin{array}{l}
1199: m r^2 {\ddot \phi}^{(a)}_i + m (r R + r^2) {\ddot \theta}^{(a)}_i \ = \\
1200: \ = \  K_s \[ (r R + r^2 ) (\theta^{(a)}_{i+1} - 2 \theta^{(a)}_i
1201: + \theta^{(a)}_{i-1}) +
1202:         r^2 (\phi^{(a)}_{i+1} - 2 \phi^{(a)}_i + \phi^{(a)}_{i-1}) \] \ + \\
1203:     \ \ \  - K_p \[
1204:   (a-R)^2 (\phi^{(a)}_i + \phi^{(\hat  a)}_i) +
1205:   a (a - R) (\theta^{(a)}_i + \theta^{(\hat  a)}_i ) \] \ ; \\
1206:  {} \\
1207:  m (r R + r^2 ) {\ddot \phi}^{(a)}_i + ( I + m (R+r)^2 ) {\ddot \theta}^{(a)}_i \ = \\
1208: \ = \ K_t \[ \theta^{(a)}_{i+1} - 2 \theta^{(a)}_i +
1209: \theta^{(a)}_{i-1} \] + \\
1210: \ \ \ + K_s (r + R) \[ (R+r) (\theta^{(a)}_{i+1} - 2
1211: \theta^{(a)}_i + \theta^{(a)}_{i-1}) +
1212:         r (\phi^{(a)}_{i+1} - 2 \phi^{(a)}_i + \phi^{(a)}_{i-1}) \] + \\
1213:         \ \ \ -  K_p \[ (a^2 -a R) (\phi^{(a)}_i + \phi^{(\hat  a)}_i) +
1214:         a^2 (\theta^{(a)}_i + \theta^{(\hat  a)}_i ) \] + \\
1215:         \ \ \ + K_h (\theta^{(\hat a)}_{i+5} - 2 \theta^{(a)}_i +
1216: \theta^{(\hat a)}_{i-5}) \ .
1217: \end{array} \feq
1218: %
1219: We are mainly interested in the  dispersion relations for the
1220: propagating waves, which are solution of the system (\ref{linEL}).
1221: To derive them it is convenient to introduce variables $\phi^{(\pm)}$ and
1222: $\theta^{(\pm)}$ defined as
1223: %
1224: \beq\lb{tpm}
1225:  \phi^\pm_i = \phi^{(1)}_i \pm \phi^{(2)}_i \ , \ \
1226: \theta^\pm_i = \theta^{(1)}_i \pm \theta^{(2)}_i \ . \feq
1227: %
1228: Let us now  Fourier transform our variables, i.e. set
1229: \beq\lb{fourier} \phi^\pm_n (t) \ = \ F^\pm_{k \om} \, \exp[i (k
1230: \delta n + \om t) ] \ ; \ \theta^\pm_n (t) \ = \ G^\pm_{k \om} \,
1231: \exp[i (k \delta n + \om t)] \ . \feq Here $k$ is the spatial wave
1232: number, $\om$ is the wave frequency, and $\delta$ is a parameter
1233: with dimension of length and set equal to the interpair distance
1234: ($\delta = 3.4~\AA$), introduced so that $k$ has dimension
1235: $[L]^{-1}$ and the physical wavelength is $\la = 2 \pi / k$. In
1236: this way, we should only consider $k \in [ - \pi / \delta , \pi /
1237: \delta ]$.
1238: 
1239: Using (\ref{tpm}) and (\ref{fourier}) into (\ref{linEL}), we get a set of
1240: linear equations for $(F^\pm_{k \om} , G^\pm_{k \om} )$; each set
1241: of coefficients with indices $(k, \om )$ decouples from other wave
1242: number and frequency coefficients, i.e. we have a set of four
1243: dimensional systems depending on the two continuous parameters $k$
1244: and $\om$. This is better rewritten in vector notation as
1245: \beq\lb{4.6} M \, \zeta_{k \om} \ = \ 0, \feq where $\zeta_{k \om}$
1246: is the vector of components \beq \zeta_{k \om} \ = \ \( F^+_{k
1247: \om} \, , \, F^-_{k \om} \, , \,  G^+_{k\om} \, , \,  G^-_{k \om}
1248: \) \feq and $M$ is a four by four matrix which we omit to write
1249: explicitly.
1250: 
1251: In order to simplify the calculations we will set to zero the
1252: radius of the disk modelling the base, i.e $d_{h}=r$. As in our
1253: model the disk describing the base cannot rotate around its axis,
1254: this assumption does not modify the physical outcome of the
1255: calculations.
1256: The condition for the existence of a solution to (\ref{4.6}) is
1257: the vanishing of the determinant of $M$. By explicit computation
1258: the latter is written as the product of three terms,
1259: apart from a constant factor $r^2$,
1260: \beq ||M|| \ = \ r^2
1261: \, \pi_1 \, \pi_2 \, \pi_3 \ . \feq
1262: The three factors being,
1263: \bea\lb{p1}
1264: \pi_1 \ &=& \ - 2 K_s + m \om^2 + 2 K_s
1265: \cos ( \delta k) \ ,\nonumber\\
1266: \pi_2 \ &=& \ I \om^2 - 2  ((K_h + K_t) -
1267: K_t \cos[k \delta] +
1268:        K_h \cos[5 k \delta]) \ ,\nonumber\\
1269: \pi_3&=&  \left(I m r^2\right) \om^4 - 2  \ r^2 \,
1270: \left( K_p I + 2 (K_s I + K_t m ) \sin^2 (k \de/2) +\right.\nonumber\\
1271: &+&2 \left.m K_h \sin^2 (5 k \de /2) \right) \om^2 + \ 8 r^2 \, \(
1272: K_p + 2 K_s \sin^2 (k \de / 2 ) \) \( K_h\sin^2 (5 k \de / 2) +
1273: K_t \sin^2 (k \de /2 ) \). \eea The equation $||M||=0$ has four
1274: solutions, given by \beq\label{drel}
1275: \begin{array}{l}
1276: \om^2_1 \ = \ 4 (K_s / m) \, \sin^2 (k \de/2 ) \ , \\
1277: \om^2_2 \ = \ 4 (K_t / I) \, \sin^2 (k \de/2) \ + \ 2 (K_h /I )
1278: \, [1+ \cos (5 k \de) ] \ , \\
1279: \om^2_3 \ = \ 2 (K_p / m) \ + \ 4 (K_s/m) \, \sin^2 (k \de/2) \ ,
1280: \\
1281: \om^2_4 \ = \ 4 (K_t/I) \, \sin^2 (k \de/2) \ + \ 4 (K_h/I) \,
1282: \sin^2 (5k \de /2) \ . \end{array} \feq
1283: 
1284: 
1285: Eqs. (\ref{drel}) provide the dispersion relations for our model.
1286: Physically, the four dispersion relations correspond to the four
1287: oscillation modes of the system in the linear regime. The relation
1288: involving $\om_{1}$ describes relative oscillations of the two
1289: bases in the chain with respect to the neighboring bases. As
1290: $\om_{1} (k)\to 0$ for $k\to 0$ there is no threshold for the
1291: generation of these phonon mode excitations.
1292: 
1293: The relations involving $\om_{2}$ and $\om_{4}$ are associated with
1294: torsional oscillations of the backbone. In case of $\om_{2}$
1295: there is a threshold for the generation of the excitation
1296: originating in the helicoidal interaction, whereas the second
1297: torsional mode $\om_{4}$ has no threshold and is thus also of
1298: acoustical type.
1299: The dispersion relation involving $\om_{3}$ describes relative
1300: oscillations of two bases in a pair. The threshold for the
1301: generation of the excitation is now determined by the pairing
1302: interaction.
1303: The dispersion relations (\ref{drel}) are plotted
1304: as
1305: $\om/(2\pi c)$, where $c$ is the speed of light  (we use the, in the
1306: literature widespread, convention of measuring frequencies in $2\pi c$
1307: units) versus $k\Delta/2$ in
1308: Fig.~\ref{fig:dispRel} for values of the physical parameters
1309: given in the tables  \ref{tab:cynParms}  and
1310: \ref{tab:dynParms}.
1311: The four dispersion relations take a simple  form if we consider
1312: excitations with wavelength $\la$ much bigger then the intrapair
1313: distance, i.e $\la >> \de$; this corresponds to the $\delta\to 0$
1314: limit. We have then \beq \om^2_\a \, - \, c^2_\a \, k^2 \ = \
1315: q^2_\a \ , \feq where $c_\a$ and $q_\a$ ($\a=1\ldots 4$) are, respectively, the
1316: velocity of propagation (in the limit $k >> q_{\alpha}$) and the
1317: excitation threshold. They are given by
1318: \beq\lb{speed1} \begin{array}{ll}
1319: c_1 = \delta \sqrt{K_s/m},\quad & q_1 = 0; \\
1320: c_2 = \delta \sqrt{(K_t - 25 K_h )/I},\quad & q_2 = 2 \sqrt{K_h/I}; \\
1321: c_3 = \delta \sqrt{K_s/m}, \quad & q_3 = \sqrt{2 K_p/m}; \\
1322: c_4 = \delta \sqrt{(K_t + 25 K_h )/I},\quad & q_4 = 0. \end{array}
1323: \feq
1324: Using the values of the parameters given
1325: in the tables \ref{tab:cynParms} and
1326: \ref{tab:dynParms}
1327: we have
1328: \beq
1329: \label{speed}
1330: \begin{array}{ll}
1331:   c_1 = 6.1 \, {\rm{ Km/s}},\quad & q_1 = 0 \ ; \\
1332:   c_2 = 0 \, {\rm{ Km/s}},\quad & q_2 = 22 \, {\rm{ cm}^{-1}} \ ; \\
1333:   c_3 = 6.1 \, {\rm{ Km/s}},\quad & q_3 = 36 \, {\rm {cm}}^{-1} \ ; \\
1334:   c_4 = 5.1 \, {\rm {Km/s}},\quad & q_4 = 0 \ ,
1335: \end{array}
1336: \feq
1337:  where $c_2=0$ comes from the fact that we are taking $K_t\simeq25K_h$
1338: (see  table \ref{tab:dynParms}). This of course  just means that
1339: $c_2$ is at least an order of magnitude smaller than the other
1340: $c_i$ -- and therefore negligible. Speeds can be converted to base
1341: per seconds by dividing each $c_i$ by $\delta=3.4\AA$; excitation
1342: thresholds can be converted in inverse of seconds by multiplying
1343: each $q_i$ by $2\pi c$, where $c$ is the speed of light.
1344: 
1345: 
1346: \begin{figure}
1347:  \includegraphics[width=130pt]{dispRel-bw.eps}\\
1348:  \caption{Graph of the dispersion relations (\ref{drel})
1349:  in the first Brillouin zone.  We plot $\omega_{\alpha}/(2\pi
1350:  c)$ ($c$  is the speed of light) as a function of $k\Delta/2$.
1351: $\omega_1,\omega_2,\omega_3,\omega_4$ are represented respectively
1352: by the thin continuous, thick continuous, thick dashed  and  thin
1353: dashed line. Units are $cm^{-1}$ in the vertical axis and radiants
1354: in the horizontal axis.}
1355:  \label{fig:dispRel}
1356: \end{figure}
1357: 
1358: \section{Nonlinear dynamics and travelling waves}
1359: \lb{s7}
1360: After studying the small amplitude dynamics of our model, we
1361: should now investigate the fully nonlinear dynamics. We are in
1362: particular interested in soliton solutions, and on physical basis
1363: they should have -- if the model has any relation with real DNA --
1364: a size of about twenty base pairs. This also means that such
1365: solutions vary smoothly on the length scale of the discrete chain,
1366: and we can pass to the continuum approximation. On the other hand,
1367: such a smooth variance assumption is not justified on the length
1368: scales (five base pairs) involved in the helicoidal interaction,
1369: and one should introduce nonlocal operators in order to take into
1370: account helicoidal interactions in the continuum approximation
1371: \cite{GRPD}.
1372: 
1373: Luckily, numerical experiments show that -- at least in the case
1374: of the original Yakushevich model -- soliton solutions are very
1375: little affected by the presence or  of the helicoidal
1376: terms (as could also be expected by their intrinsical smallness,
1377: in a context where they cannot play a qualitative role as for
1378: small amplitude dynamics) \cite{GRPD}. Thus, we will from now on
1379: simply drop the helicoidal terms, i.e. set $K_h = 0$.
1380: 
1381: \subsection{Continuum approximation and field equations}
1382: 
1383: The continuum description of the discrete model we are considering
1384: requires to introduce fields $\Phi^{(a)} (z,t) , \Theta^{(a)}
1385: (z,t) $ such that \beq\lb{6.1} \Phi^{(a)} (n \de ,t) \approx
1386: \phi^{(a)}_n \ , \ \Theta^{(a)} (n \de ,t) \approx \theta^{(a)}_n
1387: \ . \feq
1388: 
1389: The continuum approximation we wish to consider is the one where
1390: we take \beq\lb{6.2} \begin{array}{l} \Phi (x \pm \de , t) \
1391: \approx \ \Phi (x , t) \ \pm \ \de \,
1392: \Phi_x (x,t) \ + \ (1/2) \, \de^2 \, \Phi_{xx} (x,t) \ , \\
1393: \Theta (x \pm \de , t) \ \approx \ \Theta (x , t) \ \pm \ \de \,
1394: \Theta_x (x,t) \ + \ (1/2) \, \de^2 \, \Theta_{xx} (x,t) \ .
1395: \end{array} \feq
1396: 
1397: Inserting (\ref{6.2}), and taking $K_h=0$, into the Euler-Lagrange
1398: equations (\ref{EuLag}), we obtain a set of nonlinear coupled PDEs for
1399: $\Phi^{(a)} (x,t)$ and $\Theta^{(a)} (x,t)$, depending on the
1400: parameter $\de$. Coherently with (\ref{6.2}), we expand these equations to
1401: second order in $\de$ and drop higher order terms. The equations
1402: we obtain in this way  are symmetric in the chain exchange.
1403: We will be able to decompose their solutions into a symmetric and
1404: an antisymmetric part under the same exchange. In view of the
1405: considerable complication of the system, it is convenient to deal
1406: directly with the equations for these symmetric and antisymmetric
1407: part and to enforce the Y contact approximation $R+d_{h}=a$
1408: 
1409: That is  we will consider fields \beq\lb{6.4} \Phi^\pm =
1410: \Phi^{(1)} \pm \Phi^{(2)} \ , \ \Theta^\pm = \Theta^{(1)} \pm
1411: \Theta^{(2)} \ , \feq and discuss equations which result by
1412: setting two of these to zero. In the symmetric case, i.e. for
1413: \beq\lb{6.5} \Theta^{(1)} (x,t) = \Theta^{(2)} (x,t) = \Theta
1414: (x,t) \ , \ \Phi^{(1)} (x,t) = \Phi^{(2)} (x,t) = \Phi (x,t) \ ,
1415: \feq the resulting equations are
1416: %
1417: \beq\lb{6.6} \begin{array}{l} m
1418: r^2 \Phi_{tt} + (mr^2 + m R r \cos \Phi ) \Theta_{tt} \ = \ - 2
1419: a (a-R) K_p \sin (\Phi + \Theta ) + \\
1420: \ \ \ - R (2 K_p (R-a) + m r
1421: \Theta_t^2 ) \sin \Phi + \\
1422: \ \ \ + \de^2 \ K_s r \[ r (\Phi_{xx} + \Theta_{xx} ) + R
1423: \Theta_{xx} \cos \Phi + R \Theta_x^2 \sin \Phi \] \ ; \\
1424: ( m r^2 + m R r \cos \Phi ) \Phi_{tt} + (I + m (R^2 + r^2 + 2 R r
1425: \cos \Phi ) ) \Theta_{tt} \ = \\
1426: \ \ \ = - 2 a K_p (R \sin \Theta + (a - R) \sin (\Phi + \Theta) )
1427: + m \Phi_t R r (\Phi_t + 2 \Theta_t ) \sin \Phi + \\
1428: \ \ \ + \de^2 \[ K_s r^2 \Phi_{xx} + (K_t + K_s (R^2 + r^2 ))
1429: \Theta_{xx} + \right. \\
1430: \ \ \ \left. + K_s R r \( (\Phi_{xx} + 2 \Theta_{xx}) \cos \Phi -
1431: \Phi_x (\Phi_x + 2 \Theta_x ) \sin \Phi \) \] \ .
1432: \end{array} \feq
1433: %
1434: In the antisymmetric case
1435: \beq\lb{6.7} \Theta^{(1)} (x,t) = -
1436: \Theta^{(2)} (x,t) = \Theta (x,t) \ , \ \Phi^{(1)} (x,t) = -
1437: \Phi^{(2)} (x,t) = \Phi (x,t) \ , \feq the resulting equations are
1438: %
1439: \beq\lb{6.8} \begin{array}{l}
1440: m r^2 \Phi_{tt} + (m r^2 + m R r \cos \Phi ) \Theta_{tt} \ = \\
1441: \ \ \ = \ K_p (a-R) \( R \sin (\Phi + 2 \Theta ) + (a - R) \sin (2
1442: (\Phi + \Theta ) ) - 2 a \sin (\Phi + \Theta ) \) \ + \\
1443: \ \ \ + R (K_p (a-R) - m r \Theta_t^2 ) \sin \Phi + \\
1444: \ \ \ + \de^2 K-s r \[ r (\Phi_{xx} + \Theta_{xx} ) + R \cos
1445: (\Phi) \Theta_{xx} + R \sin (\Phi ) \Theta_x^2 \] \ ; \\
1446: (m r^2 + m R r \cos \Phi ) \Phi_{tt} + (I + m (R^2 + r^2 + 2 R r
1447: \cos \Phi)) \Theta_{tt} \ = \\
1448: \ \ \ = \ K_p \( -2 a R \sin \Theta + R^2  \sin (2 \Theta) + \right. \\
1449: \ \ \ \left. + (a - R) (-2 a \sin (\Phi + \Theta) + (a - R) \sin
1450: (2 (\Phi + \Theta)) +
1451: 2 R \sin (\Phi + 2 \Theta) ) \) + \\
1452: \ \ \ + m \Phi_t r R (\Phi_t + 2 \Theta_t) \sin \Phi + \\
1453: \ \ \ + \de^2 \[ K_s r^2 \Phi_{xx} + K_t \Theta_{xx} + K_s (R^2 +
1454: r^2 ) \Theta_{xx} + \right. \\
1455: \ \ \ \left. + K_s r R ((\Phi_{xx} + 2 \Theta_{xx}) \cos \Phi -
1456: \Phi_x (\Phi_x + 2 \Theta_x) \sin \Phi)\] \ . \end{array} \feq We
1457: will not write the equations in the cases of mixed symmetry, i.e.
1458: for $\Theta^{(1)} (x,t) = \Theta^{(2)} (x,t) = \Theta (x,t)$,
1459: $\Phi^{(1)} (x,t) = - \Phi^{(2)} (x,t) = \Phi (x,t)$ and for
1460: $\Theta^{(1)} (x,t) = - \Theta^{(2)} (x,t) = \Theta (x,t)$ and
1461: $\Phi^{(1)} (x,t) = \Phi^{(2)} (x,t) = \Phi (x,t)$.
1462: 
1463: \subsection{Soliton equations}
1464: 
1465: When studying DNA models, one is specially interested in
1466: travelling wave solutions, i.e. solutions depending only on $z :=
1467: x - v t$ with fixed speed $v$: \beq\lb{6.9} \Phi^{(a)} (x,t) =
1468: \vphi^{(a)} (x - v t) := \vphi^{(a)} (z) \ , \ \Theta^{(a)} (x,t)
1469: = \vth^{(a)} (x-vt) := \vth^{(a)} (z) \ . \feq
1470: 
1471: If we insert the ansatz (\ref{6.9}) into the equations (\ref{6.6}) and
1472: (\ref{6.8}), we get a set
1473: of four coupled second order ODEs; defining
1474: \beq\lb{6.10} \mu \ :=
1475: \ (m v^2 - K_s \de^2 ) \ , \ \ J \ := (I v^2 - K_t \de^2 ) \ ,
1476: \feq
1477: in the completely symmetric case (\ref{6.6}) we obtain
1478: 
1479: \beq\lb{6.12}
1480: \begin{array}{l}
1481: \mu r^2 \, \vphi'' \ + \ \mu r (r + R \cos \vphi ) \, \vth'' \ = \\
1482: \ \ = \ - 2 a K_p (a - R) \, \sin \(\vphi + \vth \) + K_s \de^2 R
1483: r \, \sin \( \vphi \) \, (\vth' )^2 \ + \\
1484: \ \ \ - R \, \sin (\vphi) \, (-2 K_p (a - R) + m r v^2 (\vth' )^2
1485: ) \ ; \\
1486: \mu r (r + R \cos \vphi ) \, \vphi'' \ + \ [ J + \mu  (R^2 + r^2 +
1487: 2 R r \cos \vphi ) \, \vth'' \ = \\
1488: \ \ = \ - 2 a K_p \( R \sin \vth + (a - R) \sin (\vphi + \vth) \)
1489: + \mu R r \, \sin (\vphi) [ (\vphi')^2 + 2 \vphi' \vth' ] \ .
1490: \end{array} \feq
1491: 
1492: In the completely antisymmetric case (\ref{6.8}), instead, we get
1493: \beq\lb{6.13}
1494: \begin{array}{l}
1495: \mu r^2 \, \vphi'' \ + \ \mu r (r + R \cos \vphi ) \, \vth'' \ = \\
1496: \ \ = \ - 2 K_p (a - R) (a - R \cos \vth - (a - R) \cos (\vphi +
1497: \vth )) \sin (\vphi + \vth ) + \\
1498: \ \ \ - \mu  R r (\sin
1499: \phi) (\vth' )^2 \ ; \\
1500: \mu r (r + R \cos \vphi ) \, \vphi'' \ + \ [ J + \mu (R^2 + r^2 +
1501: 2 R r \cos \vphi )
1502: \, \vth'' \ = \\
1503: \ \ = \ - K_p \[ 2 a R \sin \vth - R^2 \sin (2 \vth ) + \right.
1504: \\
1505: \ \ \ \left. + (a -  R) \(2 a \sin (\vphi + \vth) - (a - R) \sin
1506: (2 (\vphi + \vth)) - 2 R
1507: \sin (\vphi + 2 \vth) \) \] + \\
1508: \ \ \ + \mu r R (\sin \vphi) [(\vphi' )^2  + 2 \vphi' \vth' ] \ .
1509: \end{array} \feq
1510: 
1511: The previous equations appear too involved to be studied analytically
1512: at least in the general case.  Numerical
1513: results are discussed in Sect. \ref{s9}  below.
1514: Some understanding can be gained  at the analytical level by considering
1515: a particular case of the full equations (\ref{6.12}),(\ref{6.13}), when
1516: the system reduces essentially to the Y case.
1517: The next section is devoted to this.
1518: 
1519: \subsection{Boundary conditions}
1520: 
1521: We have so far just discussed the field equations (\ref{6.6}) and
1522: (\ref{6.8}) and their reductions; however these PDEs make sense
1523: only once we specify the function space to which their solutions
1524: are required to belong.
1525: The natural physical condition is that of {\it finite energy}; we
1526: now briefly discuss what it means in terms of our equations and
1527: the boundary conditions it imposes on their solutions.
1528: 
1529: The field equations (\ref{6.6}), (\ref{6.8}) are Euler-Lagrange
1530: equations for the Lagrangian obtained as continuum limit of
1531: (\ref{lag}). In the present case, the finite energy condition
1532: corresponds to requiring that for large $|z|$ the kinetic energy
1533: vanishes and the configuration correspond to points of minimum for
1534: the potential energy.
1535: The condition on kinetic energy yields \beq\lb{finkin} \Phi_t (\pm
1536: \infty,t) = 0 \ , \ \Theta_t (\pm \infty,t) = 0 \ , \feq where of
1537: course $\Phi_t (\pm \infty,t)$ stands for $\lim_{z \to \pm \infty}
1538: \Phi_t (z,t)$, and so for $\Theta_t$.
1539: 
1540: As for the condition involving potentials, by the explicit
1541: expression of our potentials, see above, this means (with the same
1542: shorthand notation as above) \beq\label{finpot}
1543: \begin{array}{l} \Phi (\pm \infty,t) = 0 \ , \ \Theta (\pm \infty,t) =
1544: 2 n_\pm \pi \ , \\
1545: \Phi_z (\pm \infty,t) = 0 \ , \ \Theta_z (\pm \infty,t) = 0 \ ;
1546: \end{array} \eeq
1547: 
1548: Let us now consider the reduction to travelling waves, i.e. eqs.
1549: (\ref{6.12}) and (\ref{6.13}). In this framework, conditions
1550: (\ref{finkin}) and (\ref{finpot}) imply we have to require the
1551: limit behavior described by \beq\label{fes}
1552: \begin{array}{l} \vphi
1553: (\pm \infty) = 0 \ , \ \vth (\pm \infty) = 2 n_\pm \pi \ , \\
1554: \vphi' (\pm \infty) = 0 \ , \ \vth' (\pm \infty) = 0 \ ,
1555: \end{array} \eeq for the functions $\vphi (\xi)$, $\vth (\xi )$.
1556: 
1557: We would like to stress that eqs. (\ref{6.12}) and (\ref{6.13})
1558: can also be seen as describing the motion (in the fictitious time
1559: $\xi$) of point masses of coordinates $\vth (\xi), \vphi (\xi)$ in
1560: an effective potential; such a motion can satisfy the boundary
1561: conditions (\ref{fes}) only if $(\vphi,\vth) = (0,2 \pi k )$ is a
1562: point of maximum for the effective potential. This would provide
1563: the condition $\mu < 0$ and hence a maximal speed for soliton
1564: propagation (as also happens for the standard Y model); we will
1565: not discuss this point here, as it is no variation with the
1566: standard Y case, and the condition $\mu < 0$ is satisfied with the
1567: values of parameters obtained and discussed in Sect. \ref{s5}.
1568: 
1569: The solutions satisfying (\ref{fes}) can be classified by the
1570: winding number $n := n_+ - n_-$.
1571: Needless to say, here we considered the equations describing
1572: symmetric or antisymmetric solutions, but a similar discussion
1573: also applies to the full equations, i.e. those in which we have
1574: not selected any symmetry of the solutions; in this case we would
1575: have two winding numbers, which we can associated either to
1576: $\vth^{(1)}$, $\vth^{(2)}$ or directly to their symmetric and
1577: antisymmetric combinations $\vth^{(1)} \pm \vth^{(2)}$.
1578: 
1579: 
1580: 
1581: 
1582: \section{Comparison with the Yakushevich model}
1583: \lb{s8}
1584: 
1585: The standard Y model \cite{YakPLA} can be seen as a particular
1586: limiting case of our model. Thus, a  check of the validity of our
1587: model -- and in particular of the fact we considered here physical
1588: assumptions which correspond to those by Y in her geometry -- can
1589: be obtained by going to a limit in which our composite Y model
1590: reduces to the standard Y model.
1591: The latter can be obtained as a limiting case of our composite
1592: model in two conceptually different ways.
1593: \begin{itemize}
1594: \item A first possibility, which we call {\it parametric}, is to
1595: choose the geometrical parameters of the model so that its
1596: geometry actually reduces to that of the standard Y model. \item A
1597: second possibility, which we call {\it dynamical} is to force the
1598: dynamics of our model by setting $\varphi^{a}=0$, i.e. by freezing
1599: the non-topological angles and constraining them to be zero.
1600: \end{itemize}
1601: 
1602: Let us briefly discuss these in some more detail. The parametric
1603: way to recover the standard Y model from our model consists in
1604: setting to zero the radius of the disks modelling the bases, and
1605: at the same time pushing it on the disk representing the
1606: nucleoside on the DNA chain. In this way the base corresponds to a
1607: point on the circle bounding the disk representing the nucleoside.
1608: Note that this would cause a change in the interbase equilibrium
1609: distance, unless we at the same time also change the radius of the
1610: disk representing the nucleoside.
1611: 
1612: This limiting procedure corresponds -- recalling we also want to
1613: recover the Y approximation of zero interbase distance -- to the
1614: following choice of the parameters:
1615: \beq\lb{8.1}
1616: \ m \ = \ K_{s}=d_{h} \ = \ r \ = \ 0 \ , \ \ a \ = R \ .
1617: \feq
1618: 
1619: After the base has been pushed on the disk, its mass enters to be
1620: part of the disk's mass -- hence contributes to its moment of
1621: inertia -- and we can thus just take $m=0$. Similarly, as the
1622: bases have lost their identity and are enclosed in the disk
1623: modelling the whole nucleotide, the effective stacking interaction
1624: has to be physically identified with the torsional interaction of
1625: the disk now modelling the entire nucleotide. For this reason in
1626: our equations we will take $K_s=0$ and $K_t \to K_s$.
1627: 
1628: Use of equations (\ref{8.1}) into the equations of motion (\ref{EuLag})  yields
1629: \beq\lb{8.2}
1630: \begin{array}{rl}
1631: I {\ddot \theta}^{(a)}_{i} \ =& \ K_s  \sin ( \theta^{(a)}_{i-1} -
1632: \theta^{(a)}_{i} ) \, + \, K_s \sin (\theta^{(a)}_{i+1} - \theta^{(a)}_{i} ) \,
1633: + \\
1634: & \ + \, K_p R^2 \sin ( \theta^{(a)}_{i} - \theta^{(\hat a)}_{i})
1635:    \, + \, K_h \( \theta^{(\hat a)}_{i+5} - 2  \theta^{(a)}_{i} +
1636:    \theta^{(\hat a)}_{i-5} \).
1637: \end{array}
1638: \feq The previous equations represents the equations of motions
1639: for the Y model. Some care has to be used when the values of the
1640: parameters given by Eq. (\ref{8.1}) correspond to singular points
1641: of the equations. This is for instance the case of the dispersion
1642: relations $\omega_{1}$,$\omega_{3}$  in Eqs. (\ref{drel}),  which
1643: are singular for $m=0$. The dispersion relations for the Y model
1644: can be easily found by linearizing the system (\ref{8.2}). One
1645: finds two dispersion relations; one is given by  $\om_2$ of the
1646: composite model, see Eqs. (\ref{drel}); the other is \beq\lb{8.3}
1647: \om^{2} \ = \ 2 R^{2}\frac{K_{p}}{I} \ + \ \frac{4
1648: K_{s}}{I}\sin^{2} (\frac{k \de}{2}) \ + \ \frac{4
1649: K_{h}}{I}\sin^{2} (\frac{5k \de}{2}) \ . \feq
1650: 
1651: To obtain the Y model dynamically from our model, we set $\Phi=0$
1652: into the continuum equations (\ref{6.6}) and (\ref{6.8}). We also
1653: enforce the Y condition $R+d_{h}=a$ and work in the zero
1654: radius approximation for the disk modelling the base, i.e we set
1655: $r=d_{h}$. In the fully symmetric case we get from Eqs.
1656: (\ref{6.6})
1657: \beq\lb{8.4}
1658: \begin{array}{l}
1659: m \Theta_{tt} \ = \ - 2 K_p \sin ( \Theta ) \, + \, \de^2 K_s \Theta_{xx} \ \ ;
1660: \quad \(\frac{I}{(R + r)^2} + m \) \, \Theta_{tt} \
1661:  = \ - 2K_{p} \sin \Theta \, + \, \de^2 \[ \frac{K_t}{(r+R)^2 } + K_s  \]
1662: \Theta_{xx}. \  \end{array}
1663: \feq
1664: Compatibility of the previous
1665: two equations requires that
1666: \beq\lb{8.5}
1667: I \, \Theta_{tt} \ = \ \de^{2} \, K_{t} \ \Theta_{xx} \ .
1668: \feq
1669: In the case of travelling wave solutions $\Theta(x,t)=\vth(x-vt)$
1670: the constraint (\ref{8.5}) reads
1671: \beq\lb{8.6}
1672:  v^{2} \ = \ \frac{\de^2}{I} \, K_{t}.
1673: \feq
1674: %
1675: We take from now on the
1676: positive determination of velocity for ease of discussion.
1677: Using the  Eqs. (\ref{6.9}),  (\ref{6.10}) and  (\ref{8.6}),
1678: Eqs. (\ref{8.4})  yields
1679: the travelling wave equation,
1680: %
1681: \beq\lb{6.20} \vth'' \ = - 2 (K_p / \mu_0 ) \ \sin \vth \,
1682: \feq
1683: where
1684: \beq\lb{6.19}  \mu_0= \ (m K_t - I K_s ) \ \frac{\de^2} { I} \ . \feq
1685: 
1686: With the usual boundary conditions $\vth'(\pm\infty)=0$,
1687: $\vth(\infty)= 2\pi$, $\vth (- \infty ) = 0$, Eq. (\ref{6.20}) has a
1688: solution for $\mu_0 <0$, given precisely by the  $(1,0)$ \Y
1689: soliton \beq\lb{6.21} \vth \ = \ 4 \ {\rm arctan} \[ e^{4\kappa z}
1690: \] \ , \ \ \ \kappa = \sqrt{2K_{p}|\mu_{0}|} \ . \feq
1691: We have thus recovered for the topological angles -- imposing the
1692: vanishing of non-topological angles as an external constraint --
1693: the Y solitons.
1694: The condition for the existence of the soliton, $\mu_0<0$,
1695:  implies that the physical parameters of
1696: our model must satisfy the condition
1697: \beq\lb{6.21a} \frac{K_{t}}{I}
1698: \ < \ \frac{K_{s}}{m} \ . \feq
1699: With the parameter values given in
1700: the tables  \ref{tab:cynParms} and \ref{tab:dynParms}, we have
1701: \beq\lb{6.22} \frac{m K_t} {I K_s } \ \simeq \ 0.3
1702: \ < \ 1 \ ; \feq hence (\ref{6.21a}) is satisfied and we are in the
1703: region of existence of the soliton.
1704: 
1705: Let us now consider  the antisymmetric equations.
1706: Using $\phi=0$ into the system (\ref{6.13}) and the compatibility
1707: equation (\ref{8.6}) we get
1708: \beq\lb{6.23} \vth'' \ = - 2 (K_p
1709: /\mu_0) \ (1 - \cos \vth) \ \sin \vth \ . \feq
1710: As expected this
1711: equation, with the usual boundary conditions (see above), admits a
1712: solution for $\mu_0<0$, and the solution is in this case given by
1713: $(0,1)$ \Y soliton \beq\lb{6.24} \vth \ = \ - 2 {\rm arccot} \[ -
1714: \kappa z \] \feq (with $\kappa$ as above). The allowed values of
1715: the physical parameters are determined by the same arguments used
1716: for the $(1,0)$ soliton.
1717: 
1718: 
1719: 
1720: It should be stressed that in the standard Y model the travelling
1721: waves speed is essentially a free parameter, provided the speed is
1722: lower than a limiting speed \cite{GaeSpeed,GaeJBP}. Here,
1723: recovering the standard Y model as a limiting case of the
1724: composite model produces a selection of the soliton speed given by
1725: Eq. (\ref{8.6}); this makes quite sense physically, as it
1726: coincides with the speed of long waves determined by the
1727: dispersion relations (\ref{speed1}).
1728: 
1729: 
1730: 
1731: 
1732: \section{Numerical analysis of soliton equations and soliton solutions}
1733: \label{s9}
1734: 
1735: Even after the several simplifying assumptions we made
1736: for our DNA model, the complete equations of motions
1737: given by Eqs. (\ref{6.6}) and (\ref{6.8}) respectively for
1738: symmetric and antisymmetric configurations,  are too complex to be
1739: solved analytically; the same applies to the reduced equations
1740: (\ref{6.12}), (\ref{6.13}) describing soliton
1741: solutions. We will thus look for solutions, and in particular for
1742: the soliton solutions we are interested in, numerically.
1743: 
1744: In order to determine the profile of the soliton solutions we will
1745: analyze the stationary case, with zero speed and kinetic energy,
1746: and apply the ``conjugate gradients'' algorithm (see e.g.
1747: \cite{NR,gsl}) to evaluate numerically the minima of our
1748: Hamiltonian
1749: 
1750: 
1751: This approach also allows a direct comparison with the results
1752: obtained for the standard \Y model, and shown in \cite{YakPRE},
1753: where authors proceed in the same way and by means of the same
1754: numerical algorithm; this again with the aim, as remarked several
1755: times above, to emphasize the differences which are due purely to
1756: the different geometry of our ``composite'' model. With the same
1757: motivation, we have also checked our numerical routines by
1758: applying them to the standard Y model; in doing this we have also
1759: considered with some care -- and fully confirmed -- certain
1760: nontrivial effects mentioned in \cite{YakPRE}.
1761: 
1762: \subsection{Solitons in the standard \Y model}
1763: 
1764: As mentioned above, we will first present   the  numerical
1765: investigation of  the stationary solitons of \Y model. Although
1766: the soliton solutions of the \Y model have already been the
1767: subject of previous numerical investigations \cite{YakPRE}, it is
1768: useful to repeat the analysis here in order to check our algorithm
1769: (and confirm the results reported by \cite{YakPRE}).
1770: 
1771: Moreover, in the next section we will compare the soliton
1772: solutions of our composite model with those of the \Y model. We
1773: will therefore need an explicit numerical results for the \Y model
1774: obtained using our code.
1775: 
1776: The homogeneous \Y Hamiltonian for static solutions, i.e. setting
1777: the kinetic term $T$ to zero, is \beq\label{eq:YakHam}
1778:   \begin{array}{rl}
1779:     H_Y \ =& \ \frac{I}{2} \, \sum_{i,a} \, (\Dpsi)^2 +
1780:     g \, \sum_{i,a} \, \sin^2 ((\Dpsi)/2) + K \sum_i \[3 + \cos (\theta^1_i-\theta^2_i) -
1781:     2 \cos\theta^1_i - 2 \cos\theta^2_i \]\\
1782:     =& \ I \, \sum_{n} \, [\dpsip^2 + \dpsim^2] +
1783:     g \, \sum_n \, [1-\cos\dpsip\cos\dpsim]
1784:     + 2 K \sum_n \[1 - 2\cos\psip_n\cos\psim_n + \cos^2\psim_n \] \ .
1785:   \end{array}
1786: \feq
1787: Here $I$ is the inertia momentum; $K$ and $g$ respectively
1788: the pairing and torsional coupling constants of the \Y model;
1789: $\theta^{(1,2)}$ are the angles describing the sugar rotation with
1790: respect to the backbone. Moreover, we write $\psip=\theta^+$,
1791: $\psim=\theta^-$ (the $\theta^{\pm}$ are defined as in
1792: Eq. (\ref{tpm})), and $\Delta_n\psip :=\psip_{n+1}-\psip_n$ and
1793: similarly for $\psim$.
1794: 
1795: If we select the physical value for $I$ (given in table
1796: \ref{tab:cynParms}) and factor out the $K$
1797: (equivalently, we measure energies in units of $K$; this takes the
1798: value $K=150$KJ/mol), then the only independent parameter left in
1799: $H$ is the coupling constant $g$. This can and will be used,
1800: as in \cite{YakPRE}, for a parametric study of the solutions.
1801: In our numerical investigations we used (in order to avoid any
1802: accidental spurious effect) two independent implementations of the
1803: conjugate-gradient
1804: algorithms, i.e. the one
1805: developed by Numerical Recipes \cite{NR} and the one provided by
1806: the GNU \cite{gsl}. The results obtained with the two
1807: implementations turned out to be in very good agreement.
1808: 
1809: The conjugate-gradient algorithms requires a ``starting point'',
1810:  i.e. an initial approximation of the minimizing
1811: configuration; in the case of multiple minima, the algorithm will
1812: actually determine a local minimum or the other depending on this
1813: initial approximation. Following \cite{YakPRE}, we have used as
1814: starting points hyperbolic tangent profiles
1815: %
1816: \beq\lb{start} \theta^{1,2}_n \ =
1817: \ q^{1,2} \ \pi (1 + \tanh (\beta (2n-N))) \ , \feq
1818: where
1819: $(q^1,q^2)$ is the topological type of the soliton, $N$ is the
1820: number of sites in the chain, and $\beta$ a parameter, whose
1821: reasonable range is about $0 \leq \beta \leq N/10$, used to adjust
1822: the profile.
1823: The parameter $\beta$ is crucial to determine the structure of the
1824: minima (hence of the solitonic solutions) of the \Y Hamiltonian.
1825: Choosing  different ranges of variation for $\beta$ we are probing
1826: different dynamical regions of the system, where different local
1827: minima of the Hamiltonian  may be present.
1828: 
1829: In the case of the ``elementary'' solitons -- the $(1,0)$ and the
1830: $(0,1)$ ones -- for which only one degree of freedom matters, we
1831: obtain the same local minimum, up to $10^{-5}$ in the energy and
1832: $10^{-2}$ in the angles, independently of $\beta$ (provided of
1833: course that $\beta$ is not too close to zero; in our case it
1834: suffices to keep $\beta \geq 4$) in agreement with the fact that
1835: in this case the minimum is known to be global.
1836: 
1837: \subsubsection {Quasi-degeneration of the energy minimizing configurations
1838: for higher topological numbers}
1839: 
1840: The situation appears to be different in the case of the $(1,1)$
1841: soliton. Here in facts numerical investigations
1842: show a strong dependence on $\beta$ of the local minimum
1843: determined by the algorithm for most of its range, in particular
1844: for $\beta > 6$,  while energies vary very little -- within
1845: $0.2\%$ -- about 49.3K. This  suggests we are in the presence of a rather
1846: complex structure of the phase space for the (1,1) limiting
1847: conditions. There still is however a small interval, ($4 \leq
1848: \beta \leq 6$), where the algorithm behaves exactly as in the
1849: $(1,0)$ and $(0,1)$ cases, i.e. yields almost the same energy minimizing
1850: configuration as $\beta$ is varied, and allows us to find what we
1851: believe is the global minimum of the Hamiltonian.
1852: 
1853: We have detected this same behavior also in solitons of higher
1854: topological types, e.g. $(2,0)$ and $(2,1)$. This suggests we are
1855: in the presence of a generic pattern \footnote{There are indeed
1856: qualitative arguments suggesting a degeneration of minimizing
1857: configuration whenever the topological numbers are not (1,0) or
1858: (0,1); we will not discuss these arguments nor the matter here.};
1859: we point out that the reason why this same behavior is not shared
1860: by the solitons of types $(1,0)$ and $(0,1)$ is related to the
1861: fact that in these two cases (and only in them) the problem
1862: reduces actually to a one-dimensional dynamics, while all others
1863: cases are intrinsically two-dimensional.
1864: 
1865: \begin{figure}
1866: $$\begin{array}{cc}
1867:     \includegraphics[width=200pt]{yak-1.0.eps} &
1868:     \includegraphics[width=200pt]{yak-1.1.eps}\\
1869:     a & c \\
1870:     \includegraphics[width=200pt]{yak-0.1.eps} &
1871:     \includegraphics[width=200pt]{yak-1.1-bad.eps} \\
1872:     b & d \end{array}$$
1873:   \caption{Static solitons profiles for the Yakushevich homogeneous Hamiltonian.
1874:   Thicker lines correspond to $g=150$, thinner ones to $g=23.4$;
1875:   dashed ones to the angle
1876:   $\theta^1$ and continuous ones to the angle $\theta^2$. Topological
1877:   numbers are as follows. Upper left (a): (1,0); lower left (b):
1878:   (0,1); upper right (c): (1,1); lower right (d): (1,1). Picture
1879:   (d) is obtained by a computation where $2\pi$ has been approximated with 6.28,
1880:   and shows a sensitive dependence of the solution on the boundary conditions.
1881:   The energies $E^g_{(p,q)}$ of the solitons are:
1882:   \energy{150}{1}{1}{153.4\, K}, \energy{150}{0}{1}{62.93\, K},
1883:   \energy{150}{1}{0}{62.93\, K}, \energy{23.4}{1}{1}{59.32\, K},
1884:   \energy{23.4}{0}{1}{24.63\, K}, \energy{23.4}{1}{0}{24.63\, K}.
1885:   Energies are measured in units of $K=150 {\rm KJ/mol}$.} \label{Yak}
1886: 
1887: \end{figure}
1888: 
1889: \begin{figure}
1890: $$\begin{array}{ccc}
1891:     \includegraphics[width=150pt]{yak2-1.0.eps}&
1892:     \includegraphics[width=150pt]{yak2-0.1.eps}&
1893:     \includegraphics[width=150pt]{yak2-1.1.eps}\\
1894:     a&b&c\end{array}$$
1895:     \caption{Static solitons profiles for the Yakushevich homogeneous
1896:       hamiltonian expressed in the coordinates $(\psip,\psim)$.
1897:       Thicker lines correspond to $g=150$, thinner ones to $g=23.4$;
1898:       dashed ones to the angle $\psip$ and continuous ones to
1899:       the angle $\psim$. Topological numbers are: (1,0) in
1900:       (a), (0,1) in (b), (1,1) in (c).
1901:       The energies $E^g_{(p,q)}$ of the solitons are:
1902:       \energy{150}{1}{1}{62.93\,K}, \energy{150}{0}{1}{153.4\,K},
1903:       \energy{150}{1}{0}{195.3\,K}, \energy{23.4}{1}{1}{24.64\,K},
1904:       \energy{23.4}{0}{1}{59.34\,K}, \energy{23.4}{1}{0}{75.56\,K}.
1905:       Energies are measured in units of $K=150 {\rm KJ/mol}$.}
1906:      \label{Yak2}
1907: \end{figure}
1908: 
1909: \subsubsection{Numerical instability of the solitons}
1910: 
1911: A noteworthy fact pointed out in \cite{YakPRE} is that the
1912: discrete version of the soliton solutions lose stability,
1913: namely switch from minima to saddle points,
1914: as $g$ gets close to $0$, a phenomenon that is not shared
1915: by its continuos counterpart.
1916: We have looked for this effect and confirmed the findings of
1917: \cite{YakPRE}. In our numerical computations the $g$ values at
1918: which the transition  takes place turned out to be different
1919: if computations are performed in the $\theta^{1,2}$ coordinates or
1920: in the $(\psi,\chi)$ ones. Transition values corresponding to the
1921: onset of the numerical instability are given in Table \ref{tab:instab}
1922: 
1923: \begin{table}
1924:   \centering
1925:   \begin{tabular}{|c||c|c|c|}
1926:   \hline
1927:     & (1,0) & (0,1) & (1,1) \\
1928:   \hline
1929:   ($\theta^1,\theta^2$) & 7.05 & 7.05 & 14.7 \\
1930:   ($\psi,\chi$) & 14.7 & 16.2 & 7.0 \\
1931:   \hline
1932: \end{tabular}
1933:   \caption{The transition values of $g_0$ for instability (arising
1934:   for $g < g_0$)
1935:   of the $(p,q)$ solitons.}\label{tab:instab}
1936: \end{table}
1937: 
1938: 
1939: The transition values are, in the $(\theta^{1,2})$ coordinates,
1940: $g=14.7$ for the $(1,1)$ soliton and $g=7.05$ for both the $(1,0)$
1941: and $(0,1)$ ones; in the $(\psi,\chi)$ coordinates they are $g=7$
1942: for the $(1,1)$ soliton, $g=16.2$ for the $(0,1)$ one and $g=14.7$
1943: for the $(1,0)$ one.
1944: When the instability sets in, what we observe is that the soliton --
1945: i.e. the discrete configuration smoothly interpolating between the
1946: boundary values -- breaks down and we have instead a configuration
1947: in which all angles are equal to the left boundary value for $n
1948: \le n_0$, and to the right boundary value for $n>n_0$.
1949: 
1950: In other words, the transition between the two boundary value does
1951: not take place over a (more and less extended) range of sites, but
1952: abruptly at a given site -- which we denoted above as $n_0$, but
1953: of course can be any site. This also shows that in this case we
1954: have a strong degeneration of the Hamiltonian also in the finite
1955: length case (for infinite chains, this is always the case as the
1956: Hamiltonian is invariant under translations), which will show up
1957: in a computational instability for the numerical energy
1958: minimization.
1959: 
1960: In Fig. \ref{Yak} we show the profiles of the \Y solitons we have
1961: obtained for $g=23.4$ (this is the value corresponding to the
1962: coupling constants in use in our model, see Eq. (\ref{eq:g}) below)
1963: and $g=150$ (this corresponds to the coupling constant value
1964: chosen in \cite{YakPRE}). The profiles we obtain are qualitatively
1965: identical to the inhomogeneous ones presented in
1966: \cite{YakPRE}. Incidentally, we noticed a peculiarly
1967: strong dependence on the initial conditions for the $(1,1)$ mode,
1968: so that those profiles change considerably depending on the
1969: approximation chosen for $2\pi$: e.g., in Fig.~\ref{Yak}c we used
1970: an approximation extremely precise while in Fig.~\ref{Yak}d we
1971: used $2\pi=6.28$. We believe that the first one represents the
1972: correct solution, e.g. also because it is the only one of the two
1973: that respects the symmetry of the equations.
1974: We also produced profiles for the solitons of the \Y Hamiltonian
1975: with respect to the angles $\psi,\chi$ (which
1976: are the coordinates we use in our Hamiltonian). The results are
1977: show in Fig. \ref{Yak2}; in these new coordinates no strong
1978: dependence on the initial conditions was spotted.
1979: 
1980: \subsection{Solitons in the composite  model}
1981: 
1982: Let us now turn to the numerical investigation of our  model. We
1983: will consider the case when the intrapair distance at the
1984: equilibrium is zero, i.e we will set $a=r+d_{h}$ (contact
1985: approximation).
1986: Notice that we are not considering the zero-radius approximation
1987: for the bases, so that in general $r\neq d_{h}$.
1988: The Hamiltonian of the system  can be easily derived from the
1989: Lagrangian (\ref{lag}) and is given by \beq H \ = \ T_B + T_b +
1990: V_t + V_s + V_p + V_h + V_w \ . \feq
1991: We use the shorthand notation
1992: $$ \begin{array}{l}
1993: \psip_n = \theta^+_n \ , \ \psim=\theta^-_n \ , \ \Omegap_n =
1994: \phi^+_n \ , \ \Omegam_n=\phi^-_n \ ; \\
1995: \dpsip = \psip_{n+1} - \psi_n \ , \ \spsip = \psip_{n+1}+\psi_n \
1996: ; \end{array} $$ and similarly for the other variables; we also
1997: write $$ \alpha = R/r \ , \ \beta = R / d_h \ ; $$ with the
1998: values given in table \ref{tab:cynParms}, it results $\a = 0.92$,
1999: $\b =0.53$. With these notations, we have
2000: \begin{equation}
2001: \label{eq:GaetaHam}
2002:   \begin{array}{rl}
2003:     T_B=&\sum_n I_B\left(\dpsip^2+\dpsim^2\right)\\
2004:     T_b=&\sum_n I_b    [
2005:       \dOmegap^2+\dOmegam^2+\dpsip^2+\dpsim^2+2\dpsip\dOmegap+2\dpsim\dOmegam +\alpha^2(\dpsip^2+\dpsim^2)
2006:       \\
2007:       &+2\alpha(\dpsip^2+\dpsim^2+\dpsip\dOmegap+\dpsim\dOmegam)\cos\Omegap_n\cos\Omegam_n\\
2008:       &+2\alpha(2\dpsip\dpsim+\dpsim\dOmegap+\dpsip\dOmegam)\sin\Omegap_n\sin\Omegam_n
2009:     ]\\
2010:     V_t=&2K_t \sum_n [\cos\Delta_n\psip\cos\Delta_n\psim-1]\\
2011:     V_s=&\frac{1}{2} K_s r^2\sum_n
2012:     4[ 1 + \alpha^2 - \alpha^2\cos(\dpsip)\cos(\dpsim) -
2013:     \cos(\dpsim + \dOmegam)\cos(\dpsip+\dOmegap)\\
2014:       &+2\alpha\cos((\sOmegap-\sOmegam)/2)\sin((\dpsip-\dpsim)/2)
2015:       \sin((\dpsip-\dpsim+\dOmegap-\dOmegam)/2)\\
2016:       &+2\alpha\cos((\sOmegap+\sOmegam)/2)\sin((\dpsip+\dpsim)/2)
2017:       \sin((\dpsip+\dpsim+\dOmegap+\dOmegam)/2)
2018:     ]\\
2019:     V_p=&\frac{1}{2} K_p d_h^2 \sum_n 4[
2020:       (1+\beta)^2+\cos^2(\psim_n+\Omegam_n)+
2021:       2\beta\cos\psim_n\cos\Omegap_n\cos(\psim_n+\Omegam_n)+\beta^2\cos^2\psim_n\\
2022:       &-2\beta(1+\beta)\cos\psip_n\cos\psim_n -2(1+\beta)\cos(\psip_n+\Omegap_n)
2023:       \cos(\psim_n+\Omegam_n)]\\
2024:     V_h=&K_h\sum_n [2 - \cos(\psip_{n+5}-\psim_n) -
2025:     \cos(\psim_{n+5}-\psip_n)]\\
2026:     V_w=&K_w\sum_n [\tanh(\Omegap_n+\Omegam_n)+\tanh(\Omegap_n-\Omegam_n)]
2027:   \end{array}
2028: \end{equation}
2029: 
2030: 
2031: Note that we have inserted in the Hamiltonian the confining
2032: potential $V_{w}$ in order to implement dynamically the constraint
2033: (\ref{bc}) for the non-topological angles $\phi^{(a)}$. Adding
2034: this term in the potential is also instrumental in stabilizing the
2035: numerical minimizations.
2036: 
2037: 
2038: The Hamiltonian (\ref{eq:GaetaHam}) reduces to that of the \Y
2039: model setting $\Omegap=\Omegam=0$ (and disregarding the helicoidal
2040: term); with this we get
2041: \begin{equation}
2042:   \label{eq:GaetaYakHam}
2043:   \begin{array}{rl}
2044:     T_B=&I_B\sum_n \left(\dpsip^2+\dpsim^2\right)\cr
2045:     T_b=&I_b(1+\alpha^2)\sum_n [\dpsip^2+\dpsim^2]\cr
2046:     V_t=&2K_t\sum_n [\cos\dpsip\cos\dpsim-1]\cr
2047:     V_s=&\frac{1}{2} K_s r^2\sum_n 4(1+\alpha)^2 [1-\cos\dpsip\cos\dpsim]\cr
2048:     V_p=&\frac{1}{2} K_p d_h^2\sum_n  4(1+\beta)^2
2049:     [1-2\cos\psip_n\cos\psim_n +\cos^2\psim_n]\cr
2050:     V_w=&0;
2051:   \end{array}
2052: \end{equation}
2053: Also note that in this case $V_t$ and $V_s$ differ just by a
2054: multiplicative function.
2055: 
2056: \begin{figure}
2057:   \includegraphics[width=150pt]{map.eps}
2058:   \caption{Region of instability (black region) for the discrete solitons
2059:   of topological type
2060:   $(0,1)$.
2061:     in the $(g_t,g_s)$ plane.}
2062:   \label{fig:unstReg}
2063: \end{figure}
2064: 
2065: The typical energies involved in the different interactions are
2066: given by the coefficients  in Eq. (\ref{eq:GaetaHam}),
2067: $E_{t}=2K_{t},\, E_{s}= \frac{1}{2} K_{s}r^{2},\, E_{p}=
2068: \frac{1}{2} K_{p}d_{h} ^{2},\,E_{h}=K_{h},\, E_{w}=K_{w}$,  which
2069: represent, respectively the typical torsional, stacking, pairing,
2070: helicoidal and confining energies. Using the values of the
2071: physical parameters given in the tables \ref{tab:cynParms},
2072: \ref{tab:dynParms}, and choosing $E_{w}$ in order to keep the
2073: confining energy at least a full order of magnitude smaller than
2074: any other one, we get for the typical interaction energies the
2075: values given in table \ref{tab:energies}.
2076: 
2077: In order to work with dimensionless quantities, throughout this
2078: section we will measure energies in terms of $E_p=(1/2) K_p
2079: d_h^2=4.0\cdot10^2 {\rm KJ/mol}$. Using the values of the
2080: kinematical parameters given in \ref{tab:para} and those of the
2081: dynamical parameters given by table \ref{tab:dynParms}, the
2082: dimensionless coupling constants turn out to be \beq\lb{nc}
2083: \begin{array}{l}
2084: g_t = E_t/E_p = 0.65 \ , \
2085: g_s = E_s/E_p = 7.2 \ , \ g_p = 1 \ , \\
2086: g_h = E_h/E_p = 0.026 \ , \ g_w = E_w /E_p = 0.001 \ .
2087: \end{array} \feq
2088: 
2089: \begin{table}
2090:   \centering
2091:   \begin{tabular}{|c|c|c|c|c|}
2092:   \hline
2093:     $E_t$ & $E_s$ & $E_p$ & $E_h$&$E_{w}$ \\
2094:   \hline
2095:     $2.6\cdot10^2 KJ/mol$ & $2.9\cdot10^3 KJ/mol$ & $4\cdot10^2KJ/mol$ & $10
2096:     KJ/mol$ &$4\cdot10^{-1}KJ/mol$ \\
2097:   \hline
2098: \end{tabular}
2099:   \caption{Values of the typical energies characterizing the different
2100:   interactions in the Hamiltonian of Eq.~\ref{eq:GaetaHam}}
2101:   \label{tab:energies}
2102: \end{table}
2103: %
2104: 
2105: Note that Eq. (\ref{eq:GaetaYakHam}) implies that, in the limit
2106: $\Omegap=\Omegam=0$, the \Y couplings ($K,g$) and our coupling
2107: constants are related by \beq K \ = \ 2 (1+\beta)^2 \, g_p \ , \ \
2108: g \ = \ 2 \, g_t \ + \ 4 (1+\alpha)^2 \, g_s \ ; \feq this also
2109: yields
2110: \begin{equation}
2111: \label{eq:g}
2112: \frac{g}{K} \ = \ \frac{g_t+2(1+\alpha)^2g_s}{(1+\beta)^2 g_p}
2113: \ \simeq \ \frac{g_t+7.4g_s}{2.3g_p} \ \simeq \ 23 \ .
2114: \end{equation}
2115: 
2116: Most of the statements made in the previous section for the \Y
2117: Hamiltonian apply almost verbatim  to our case. We obtain an
2118: approximate profile of a soliton, subject to the  boundary
2119: conditions (\ref{fes}), by
2120: minimizing numerically the Hamiltonian through the
2121: ``conjugate-gradient'' algorithm, in particular through its
2122: implementations in the GSL \cite{gsl} and in the Numerical Recipes
2123: \cite{NR}.
2124: 
2125: \begin{figure}
2126: $$\begin{array}{cc}
2127:     \includegraphics[width=150pt]{gaeta-0.1a.eps}&
2128:     \includegraphics[width=150pt]{gaeta-0.1b.eps}\\
2129:     a&b\\
2130:     \includegraphics[width=150pt]{gaeta-0.1c.eps}&
2131:     \includegraphics[width=150pt]{gaeta-0.1d.eps}\\
2132:     c&d\end{array} $$
2133:   \caption{Stationary solitonic solutions of our model with energy
2134:   $E=80.06 \,E_{p} $ of topological numbers $(1,0)$ (thick line) compared
2135:   with the solitonic solutions of the \Y model  of energy  $E=75.56 \, E_{p}$
2136:   (thin line).
2137:   Upper left (a): the angle $\psip$; upper right (b): the angle $\psim$;
2138:   lower left (c): the angle $\Omegap$; lower right (d): the angle $\Omegam$.
2139:   The thin line segments visible in (b) show the small difference between
2140:   the profile of the  solitons of our and  of the \Y model.
2141:   }
2142:   \label{fig:g1.0}
2143: \end{figure}
2144: 
2145: To enforce a particular topological type $(p,q)$ for the soliton
2146: under study we fix the angles at the extremes of the chain so that
2147: $\psip_{-\infty}=\psim_{-\infty}=0$ and $\psip_{+\infty}=2\pi p$,
2148: $\psim_{+\infty}=2\pi q$, while the non-topological angles are
2149: requested to be identically zero at the extremes.
2150: As ``starting point'' for the the algorithm (see above) we use the
2151: natural choice \cite{YakPRE} \beq \begin{array}{l} \psip_n = \pi
2152:  p(1+\tanh(\beta (2n-N))) \ , \\
2153:  \psim_n= \pi q(1+\tanh(\beta (2n-N))) \ , \\
2154: \Omegap_n=\Omegam_n=0 \ , \end{array} \feq where $\beta$ is a
2155: parameter used to adjust the profile of the initial configuration
2156: (the starting point) and $N$ is the number of sites on the chain.
2157: The number $N$ is of course much smaller than in real DNA (usually
2158: $N=4000$ in our simulations) but big enough to ensure that $\psip$
2159: and $\psim$ are constant at the beginning and the end of the chain
2160: within the numerical precision of our computations.
2161: 
2162: \begin{figure}
2163:   {} \hfill \includegraphics[width=150pt]{gaeta-1.0a.eps}
2164:   \hfill \includegraphics[width=150pt]{gaeta-1.0b.eps} \hfill {}
2165:   \caption{Stationary solitonic solutions of our model with
2166:   topological numbers $(0,1)$ for different values of the normalized
2167:   couplings.  The  angle $\psip$ is depicted
2168:   on the left whereas  $\Omegap$ is depicted
2169:     on the right.
2170:     The soliton relative to the physical coupling constants with energy
2171:     $E=189.9E_{p} $ (thick line)
2172:     is shown together
2173:     with those relative to the coupling constants $g_t=0$, $g_s=46$
2174:    with energy $E=492.4 E_{p} $
2175:     (thin dashed line) and
2176:     $g_t=345$, $g_s=0$ (thin line, $E=388.5 E_{p} $)
2177:     to show how the profile would change at the increasing of the
2178:     coupling constants in the two extreme cases of negligible torsional or stacking interactions.}
2179:   \label{fig:g0.1}
2180: \end{figure}
2181: 
2182: 
2183: Like in previous case, there is a threshold for the coupling
2184: constants that must be surpassed for the solitons to be stable; in
2185: Fig. \ref{fig:unstReg} we show the region of instability for
2186: solitons in the $(g_t,g_s)$ plane when we fix the values of the
2187: other coupling constants to be $g_h=0.026$, $g_p=1$, $g_w=0.001$.
2188: As above, in the case of solitons of topological type $(1,0)$ and
2189: $(0,1)$ we always reach the same minimum -- within $10^{-5}$ in
2190: the energy and $10^{-2}$ in the angle -- while $\beta$ varies
2191: across almost two orders of magnitude, provided that $\beta \geq
2192: 4$ to avoid falling on the step solution.
2193: 
2194: The $(1,1)$ soliton also shows the very same behavior as for the
2195: \Y Hamiltonian, namely a different local extrema for every value
2196: of $\beta$ except for a short interval $4 \leq \beta \leq 5.5$
2197:  and for a range of energies wider -- within 2\% --
2198: about 160K; in
2199: the latter we get the same stable behavior observed for the
2200: $(1,0)$ and $(0,1)$ solitons.
2201: In this case again, as in the \Y case, the sensitivity of the
2202: numerical solitonic solutions to the values of the parameter
2203: $\beta$ could be an indication of the existence of many, almost
2204: degenerate, solitonic solution for topological numbers different
2205: than (1,0) or (0,1).
2206: We would like to stress that that the existence of different,
2207: almost degenerate, local minima is a typical feature of most
2208: bio-physical systems. However, our results concerning these point
2209: have to  be regarded just has an indication. Further
2210: investigations, in particular at the analytical level, are needed
2211: in order to draw a definite conclusion.
2212: 
2213: In Fig. \ref{fig:g1.0} we plot the $(1,0)$ soliton of our model with
2214: the  physical values of the normalized coupling constants
2215: given by Eq. (\ref{nc}) and compare them with those obtained for
2216: the \Y model. The profiles of the topological angles change very
2217: little from the corresponding profiles of the \Y solitons.
2218: 
2219: %\subsubsection*{Dependence on the parameters}
2220: 
2221: We have also investigated the deformation of soliton profiles when
2222: adjusting selected parameters of our Hamiltonian.
2223: First, in order to see how the shape of the soliton changes upon
2224: increasing the strength of the torsional/stacking interactions, in
2225: Fig.~\ref{fig:g0.1} we compare the profiles of the $(0,1)$
2226: solitons with profiles corresponding to $g/K\simeq150$ (see
2227: Eq. (\ref{eq:g}), i.e. the coupling constant used in \cite{YakPRE}.
2228: 
2229: As it is not completely clear how to separate the interaction
2230: strength between torsional and stacking interactions (for our
2231: choice of physical constants in Sect. \ref{s5}) we have  used about
2232: the smallest reasonable value for $K_t$. We present
2233: the profiles corresponding to the two extreme possibilities: the
2234: one in which we put all the strength in the backbone torsional
2235: interaction ($g_t=345$, $g_s=0$), and the one in which we put all
2236: of it in the bases stacking ($g_t=0$, $g_s=46$). The effect is the
2237:  widening of both the soliton and the non-topological
2238: profiles by roughly a factor 4 in the first case and of a factor 2
2239: in the second case.
2240: 
2241: In Fig.~\ref{fig:g1.1} we compare the profiles of the soliton
2242: $(1,1)$ with those obtained by using the correct distance function
2243: for $V_p$, namely by replacing $g_p\rho^2$ with
2244: $g_p(\rho-d_0/d_h)^2$ (where $d_0 \simeq 2~\AA$ is the equilibrium
2245: distance between two bases in a pair \footnote{In order to avoid
2246: any confusion, we stress that here the ``distance'' between bases
2247: refers e.g. to the N--H distance in a N--H--O hydrogen bond; the
2248: total length of the bond is about $3 \AA$, and often one refers to
2249: this as the interbase distance. Here instead we consider the $H$
2250: atom -- which lies at about $1 \AA$ from the nearer atom in the H
2251: bond -- as part of one of the bases and hence consider the
2252: distance of it from the other atom as the interbase distance.}),
2253: and by varying the helicoidal interaction term. No relevant
2254: changes are detected in the first case: relative differences in
2255: energies and angles are of the order of $10^{-2}$ in energy and
2256: $10^{-1}$ in the angles; even increasing the base-pairs distances
2257: by two orders of magnitude these results do not  modify the
2258: situation.
2259: 
2260: \begin{figure}
2261: $$\begin{array}{cc}
2262:  \includegraphics[width=150pt]{gaeta-1.1a.eps}&
2263:  \includegraphics[width=150pt]{gaeta-1.1b.eps}\\
2264:     a&b\\
2265:     \includegraphics[width=150pt]{gaeta-1.1c.eps}&
2266:     \includegraphics[width=150pt]{gaeta-1.1d.eps}\\
2267:     c&d\end{array} $$
2268:   \caption{Comparison of stationary solitonic solutions  of our
2269:   model  with those obtained using a modified pairing potential
2270:   $V_{p}$. Upper left (a): the angle $\psip$; upper right (b): the angle $\psim$;
2271:   lower left (c): the angle $\Omegap$; lower right (d): the angle $\Omegam$.
2272:   The thick line represent a soliton with $E=80.06$ and topological numbers $(1,1)$).
2273:   The thin dashed line gives the profile for the same soliton with
2274:   energy $E=74.41 E_{p} $ and
2275:   with the ``correct''
2276:   pairing potential $V_p=g_p(\rho-d_0/d_h)^2$. The latter solitonic
2277:   solution has been derived taking $d_0=3.2d_h$,
2278:   namely a order of magnitude bigger than its physical value, to enhance the
2279:   profile differences (an almost identical profile is obtained if we suppress
2280:   the helicoidal term from the Hamiltonian). The thin continuos line
2281:   ($E=102.9 E_{p} $) is the
2282:   profile we get by increasing the helicoidal term to $g_h=1$.}
2283:   \label{fig:g1.1}
2284: \end{figure}
2285: 
2286: As for the helicoidal term, we get variations of the same order of
2287: magnitude as above if we simply turn it off. If we instead
2288: increase the coupling constant by one order of magnitude
2289: ($g_h=0.26$), then we get energy and angles changes of the order
2290: of $10^{-1}$ and by increasing it to $g_h=1$ we arrive to changes
2291: of the order of $10^0$ in both the angles and the energy.
2292: 
2293: Raising $g_h$ up to $g_h=2$ leads to the disappearance of the
2294: soliton; it seems reasonable to argue that this is due to such an
2295: interaction favoring a sharper transition between limit behaviors,
2296: so that the discreteness effect discussed in the previous
2297: subsection arises.
2298: 
2299: \subsection{Discussion}
2300: 
2301: The numerical analysis we have performed shows  the
2302: existence of solitonic solutions of our composite DNA model.
2303: The profiles of the topological solitons -- in particular, the
2304: part relating to the topological degree of freedom -- of our model
2305: are both qualitatively and quantitatively very similar to those of
2306: the Y model. This means that the most relevant (for DNA
2307: transcription) and characterizing feature of the nonlinear DNA
2308: dynamics present in the Y model is preserved by considering
2309: geometrically more complex and hence more realistic DNA models.
2310: 
2311: Moreover, the topological soliton profiles of our model seem to
2312: change very little when either the physical parameters change in a
2313: reasonable range or also the form of the potential modelling the
2314: pairing interaction is modified to a more realistic form.
2315: In particular, the form of the topological solitons are very
2316: little sensitive to the interchange of torsional and stacking
2317: coupling constant.
2318: 
2319: This feature add other reasons why the Y model, although based on
2320: a strong simplification of the DNA geometry, works quite well in
2321: describing solitonic excitations. The Y model, indeed, does not
2322: distinguish between torsional and stacking interaction; but, as we
2323: have shown, this distinction is not relevant -- at least as long
2324: as one is only interested in the existence and form of the soliton
2325: solutions.
2326: The ``compositeness'' of our model becomes relevant -- and
2327: rather crucial -- when it comes on the one hand to allowing the
2328: existence of solitons {\it together} with requiring a physically
2329: realistic choice of the physical parameters characterizing the
2330: DNA, and on the other hand to have also predictions fitting
2331: experimental observations for what concerns quantities related to
2332: small amplitude dynamics, such as transverse phonons speed. In
2333: other words, the somewhat more detailed description of DNA
2334: dynamics provided by our model allows it to be effective -- with
2335: the same parameters -- across regimes, and provide meaningful
2336: quantities in both the linear and the fully nonlinear regime.
2337: 
2338: The solitonic solutions of the composite model share also two
2339: other features with those of the Y model, namely the presence of a
2340: numerical instability and the existence of quasi-degenerate
2341: solutions for solitonic configurations with higher topological
2342: numbers.
2343: 
2344: 
2345: We expect that the model considered here is the simplest DNA model
2346: describing rotational degrees of freedom which, with physically
2347: realistic values of the coupling constants and other parameters,
2348: allows for the existence of topological solitons and at the same
2349: time is also compatible with observed values of bound energies and
2350: phonon speeds in DNA.
2351: 
2352: 
2353: 
2354: \section{Summary and conclusions}
2355: \lb{s10}
2356: 
2357: Let us, in the end, summarize our discussion and the results of
2358: our work, and state the conclusions which can be drawn from it.
2359: 
2360: \subsection{Summary and results}
2361: 
2362: Following the work by Englander {\it et al.} \cite{Eng}, different
2363: authors have considered simple models of the DNA double chain --
2364: focusing on rotational degrees of freedom -- able to support
2365: dynamical and topological solitons \cite{YakuBook}, supposedly
2366: related to the transcription bubbles present in real DNA and
2367: playing a key role in the transcription process.
2368: 
2369: These models usually consider a single (rotational) degree of
2370: freedom per nucleotide \cite{YakuBook}, albeit models with one
2371: rotational and one radial degree of freedom per nucleotide have
2372: also been considered \cite{BCP,BCPR,CM,Joy,YakSBP} (as an
2373: extension of ``purely radial'' models \cite{PB,PeyNLN}, considered
2374: in the study of DNA denaturation). A simple model which has been
2375: studied in depth is the so called Y model \cite{YakPLA}. This
2376: supports topological solitons (of sine-Gordon type) and provides
2377: correct orders of magnitude for several physically relevant
2378: quantities \cite{GRPD}; on the other hand, the soliton speed
2379: remains essentially a free parameter \cite{GaeSpeed,GaeJBP}, and
2380: the speed of transverse phonons can be made to have a physical
2381: value only by assigning unphysical values to the coupling
2382: constants of the model \cite{YakPRE}.
2383: 
2384: Here we have considered an extension of the Y model, with two
2385: degrees of freedom -- both rotational -- per nucleotide; one of
2386: these is associated to rotations of the nucleoside (unit of the
2387: backbone) around the phosphodiester chain and is topological --
2388: i.e. can go round the $S^1$ circle -- while the other is
2389: associated to rotations of the attached nitrogen base around the
2390: $C_1$ atom in the sugar ring, and due to sterical hindrances is
2391: non-topological -- i.e. rotations are limited to a relatively
2392: small range around the equilibrium position. We denoted this as a
2393: ``composite Y model''.
2394: 
2395: Several parameters appear in the model; some of these are related
2396: to the geometry and the kinematics of the DNA molecule, while
2397: other are coupling constants entering in the potential used to
2398: model intramolecular interactions. We have assigned values to the
2399: first kind of parameters following from available direct
2400: experimental observations, and for the second kind of parameters
2401: we used experimental data on the ionization energies of the
2402: concerned couplings and the form of the potentials appearing in
2403: the model. That is, these parameters were {\it not}  chosen by
2404: fitting dynamical predictions of the model; see sect. V for
2405: detail.
2406: We have first considered small amplitude dynamics (sect. VI); this
2407: yields the dispersion relations and produced some prediction on
2408: the phonon speed and the optical frequency for the different branches.
2409: These prediction are
2410: a first success of the present model, in that it was shown in
2411: sect. VI that taking the physical order of magnitude for the pairing,
2412: stacking, torsional and helicoidal interaction, one can obtain the order of
2413: magnitude of the
2414: experimentally observed speed for transverse phonon excitations and
2415: the frequency threshold for the optical branch. In particular, using
2416: a physical value for the stacking and torsional interaction energy we
2417: get  a value for the transverse phonon speed which is about three times the
2418: ``correct'' one.
2419: It should be considered, in looking at this
2420: value, that we modelled the intrapair interaction by a very simple
2421: and non realistic potential (with the aim of both keeping
2422: computations simple and allowing direct comparison with the
2423: standard Y model by making the same simplifying assumptions as
2424: there). As for comparison with standard Y model, hence for an
2425: evaluation of the advantages brought by considering a more
2426: articulated geometry of the nucleotide, it should be recalled that
2427: the numerical computations of Yakushevich, Savin and Manevitch
2428: \cite{YakPRE} (which we repeated, and fully confirmed) show in
2429: order to obtain the experimentally observed speed for transverse
2430: phonon excitations in the framework of the standard Y model, one
2431: should take a coupling constant for the transverse intrapair
2432: interaction which is about 6000 times the ``correct'' one.
2433: 
2434: We passed then to consider the fully nonlinear regime, and in
2435: particular to look for solitonic-like travelling excitations.
2436: These should have smooth variations on the space scale of
2437: nucleotides, hence we passed to a continuum description and field
2438: equations; by using the chain exchange symmetry, we considered
2439: fully symmetric and antisymmetric reductions, see (\ref{6.6}) and
2440: (\ref{6.8}). By a travelling wave ansatz we reduced these to a
2441: system of two coupled second order ODEs for $\vth (z)$ and $\vphi
2442: (z)$, see (\ref{6.12}) and (\ref{6.13}). Here $\vth$ is the
2443: topological angle, i.e. the variable associated to the topological
2444: field, and $\vphi$ is the non topological angle, i.e. the variable
2445: associated to the non topological field.
2446: 
2447: The finite energy condition (\ref{finkin}), (\ref{finpot})
2448: requires that the solutions to this system of ODEs satisfy certain
2449: limit conditions (see (\ref{fes} ). These in turn imply that solutions
2450: satisfying them can be classified according to two topological
2451: indices (winding numbers for the topological fields; in the
2452: symmetric or antisymmetric case, one index is enough to determine
2453: the other as well).
2454: 
2455: We have also shown that the standard Y model can be obtained from
2456: the composite Y model by a limiting procedure (sect. VII); this
2457: also reduces the solitons of the composite model to solitons of
2458: the Y model. However, the limiting procedure requires that a
2459: certain condition is satisfied, see eq. (\ref{8.5}), and this in
2460: turn constraints the speed of solitonic excitations; see
2461: (\ref{8.6}). Thus, the requirement to obtain the standard Y
2462: solitons in a certain limit fixes the speed of solitons; the
2463: resulting speed is just the speed of long waves as determined by
2464: the dispersion relations.
2465: 
2466: Finally, in sect. IX we conducted a careful numerical
2467: investigation of the simpler soliton solutions for the composite Y
2468: model. We preliminarily checked our numerical routines on the
2469: standard Y model and fully confirmed the results of Yakushevich,
2470: Savin and Manevitch \cite{YakPRE}, also confirming certain
2471: instability phenomena they reported. We then considered the
2472: solitons for the composite Y model with the value of parameters
2473: descending from their physical meaning (i.e. with no parameter
2474: fitting), confirming their existence, properties and stability. We
2475: also showed how the profile of the soliton component corresponding
2476: to the topological degree of freedom is extremely similar to the
2477: standard Y soliton with same topological numbers. We considered
2478: next the stability of these soliton solutions upon varying the
2479: parameters of the model, and observed that as in the standard Y
2480: case there is a stability threshold. Thus, the existence and
2481: stability of soliton solutions for physical values of the
2482: parameters is a nontrivial prediction.
2483: 
2484: \subsection{Discussion and conclusions}
2485: 
2486: The composite Y model considered here retains all the favorable
2487: features of the standard Y model. At the same time, its more
2488: articulated geometry allows at the same time -- and with physical
2489: values of the coupling constants and other parameters entering in
2490: the model -- to reproduce relevant value of physical quantities
2491: related to the linear regime (such as speed of transverse phonon,
2492: which was a critical test for the standard Y model) and support
2493: stable soliton solutions.
2494: 
2495: Further, and at difference with the standard Y model, it provides
2496: a precise prediction for the soliton speed; this is quite
2497: reasonable physically, as it corresponds to the speed of long
2498: waves as obtained from the dispersion relations for the model.
2499: Thus, our model passed some -- in our opinion, significant --
2500: quantitative test and provides precise predictions.
2501: 
2502: It should also be stressed that we used -- both to simplify the
2503: mathematics and to have a direct comparison with the standard Y
2504: model -- a very simple form for the intrapair coupling potential
2505: (and at some stage also resorted to the ''contact approximation''
2506: to get simpler formulas, again as in the standard Y model
2507: treatment). It is quite conceivable that adopting a more realistic
2508: potential will provide better estimates of relevant physical
2509: quantities, in particular for quantities related to the linear
2510: regime. However, experience recently gained with the standard Y
2511: model \cite{GaeY1,GaeY2} suggests that the predictions related to
2512: the fully nonlinear regime are rather little sensitive to the
2513: detailed form of the potential and to adopting or otherwise the
2514: contact approximation; we are thus rather confident that future
2515: work with more realistic potentials will confirm the results
2516: obtained in the simple setting considered here.
2517: 
2518: Finally, we would like to remark a very relevant feature of our
2519: model. All the DNA models amenable to analytic treatment look at
2520: homogeneous DNA, albeit the genetic information lies precisely in
2521: the non homogeneous part of the DNA (i.e. the base sequence; bases
2522: have rather different physical and geometrical characteristics).
2523: Our discussion was no exception, and we considered identical bases
2524: with ``average'' geometrical and physical characteristics. But,
2525: the degrees of freedom we considered for each nucleotide are one
2526: concerned with the uniform part of the DNA molecule (the
2527: nucleosides), the other with the non homogeneous part (the base
2528: sequence). Moreover, it turned out that -- for what concerns
2529: soliton excitations -- the most relevant role is played by the
2530: (topological) variables associated to the uniform part, which are
2531: directly at play in the topological solitons, while the (non
2532: topological) variables associated to the non uniform part are in a
2533: way just accompanying the soliton.
2534: 
2535: This suggests that, within the framework of composite models, the
2536: non homogeneous case can be studied as a (non-singular)
2537: perturbation of the homogeneous case; needless to say, by this we
2538: mean an analytical -- albeit approximated -- study, and not just a
2539: numerical one. This represents a significant advance with respect
2540: to what is possible with simple models considered so far.
2541: 
2542: 
2543: \section*{Acknowledgements}
2544: 
2545: This work received support by the Italian MIUR (Ministero
2546: dell'Istruzione, Universit\`a e Ricerca) under the program
2547: COFIN2004, as part of the PRIN project {\it ``Mathematical Models
2548: for DNA Dynamics ($M^2 \times D^2$)''}.
2549: 
2550: %\vfill\eject
2551: 
2552: 
2553: 
2554: \begin{thebibliography}{99}
2555: 
2556: 
2557: \bibitem{Lib} G. Altan-Bonnet, A. Libchaber and O. Krichevsky, ``Bubble dynamics
2558: in double-stranded DNA'', {\it Phys. Rev. Lett.} {\bf 90} (2003),
2559: 138101
2560: 
2561: \bibitem{Banavali} N.K. Banavali and A.D. MacKerell Jr., ``Free energy
2562: and structural pathways of base flipping in a DNA GCGC containing
2563: sequence'', {\it J. Mol. Bio.} {\bf 319} (2002), 141-160
2564: 
2565: \bibitem{BCP} M. Barbi, S. Cocco and M. Peyrard, ``Helicoidal model for DNA
2566: opening'', {\it Phys. Lett. A} {\bf 253} (1999), 358-369; ``Vector
2567: nonlinear Klein–Gordon lattices: general derivation of small
2568: amplitude envelope soliton solution'', {\it Phys. Lett. A} {\bf
2569: 253} (1999), 161-167
2570: 
2571: \bibitem{BCPR} M. Barbi, S. Cocco, M. Peyrard and S. Ruffo, ``A twist-opening
2572: model of DNA'', {it J. Biol. Phys.} {\bf 24} (1999), 97-114
2573: 
2574: \bibitem{BLPT} M. Barbi, S. Lepri, M. Peyrard, and N. Theodorakopoulos,
2575: ``Thermal denaturation of a helicoidal DNA model'', {\it Phys.
2576: Rev. E} {\bf 68} (2003), 061909
2577: 
2578: \bibitem{BZ} M.D. Barkley and B.H. Zimm, ``Theory of twisting and bending of
2579: chain macromolecules; analysis of the fluorescence depolarization
2580: of DNA'', {\it J. Chem. Phys.} {\bf 70} (1979), 2991-3007
2581: 
2582: \bibitem{BFLG} N. Bruant, D. Flatters, R. Lavery and D. Genest,
2583: ``From atomic to mesoscopic descriptions of the internel dynamics
2584: of DNA'', Biophysical Journal {\bf 77} (1999), 2366-2376
2585: 
2586: \bibitem{Campa} A. Campa, ``Bubble propagation in a helicoidal molecular chain'',
2587: {\it Phys. Rev. E} {\bf 63} (2000), 021901
2588: 
2589: \bibitem{CD} C. Calladine and H. Drew, {\it Understanding DNA},
2590: Academic Press (London) 1992; C. Calladine, H. Drew, B. Luisi and
2591: A. Travers, {\it Understanding DNA} ($3^{rd}$ edition), Academic
2592: Press (London) 2004
2593: 
2594: \bibitem{CP} Y.Z. Chen and E.W. Prohofsky, {\it Theory of presure-dependent
2595: melting of the DNA double helix: role of straining hydrogen
2596: bonds}, {\it Phys. Rev. E} {\bf 47} (1992)
2597: 
2598: \bibitem{CM} S. Cocco and R. Monasson, ``Statistical mechanics of torque
2599: induced denaturation of DNA'', {\it Phys. Rev. Lett.} {\bf 83}
2600: (1999), 5178-5181
2601: 
2602: \bibitem{CuS} S. Cuenda and A. Sanchez, ``On the discrete
2603: Peyrard-Bishop model of DNA: stationary solutions and stability'',
2604: preprint {\it q-bio.OT/0511036} (2006), to appear in {\it Chaos}
2605: 
2606: \bibitem{DauPLA} Th. Dauxois, ``Dynamics of breather modes in a
2607: nonlinear helicoidal model of DNA'', {\it Phys. Lett. A} {\bf 159}
2608: (1991), 390-395
2609: 
2610: \bibitem{Dav} A.S. Davydov, {\it Solitons in Molecular Systems},
2611: Kluwer (Dordrecht) 1981
2612: 
2613: \bibitem{DeLuca} J. De Luca, E.D. Filho, A. Ponno and J.P. Ruggiero,
2614: ``Energy localization in the Peyrard-Bishop DNA model'', {\it
2615: Phys. Rev. E} {\bf 70} (2004), 026213
2616: 
2617: \bibitem{1BNA} H.R. Drew, R.M. Wing, T. Takano, C. Broka,
2618: S. Tanaka, K. Itakura and R.E. Dickerson, ``Structure of a B-DNA
2619: dodecamer: conformation and dynamics'', {\it Proc. Natl. Acad.
2620: Sci. USA} {\bf 78} (1981) 2179-2183;
2621: 
2622: \bibitem{Eng} S.W. Englander, N.R. Kallenbach, A.J. Heeger, J.A.
2623: Krumhansl and A. Litwin, ``Nature of the open state in long
2624: polynucleotide double helices: possibility of soliton
2625: excitations'', {\it PNAS USA} {\bf 77} (1980), 7222-7226
2626: 
2627: 
2628: \bibitem{Fedyanin} V.K. Fedyanin, I. Gochev and V. Lisy,
2629: ``Nonlinear dynamics of bases in continual model of DNA helices'',
2630: {\it Stud. Biophys.} {\bf 116} (1984), 59-64; V.K. Fedyanin and V.
2631: Lisy, ``Soliton conformational excitations in DNA'', {\it Stud.
2632: Biophys.} {\bf 116} (1984), 65-71
2633: 
2634: \bibitem{FK2} M.D. Frank-Kamenetskii, ``Biophysics of the DNA
2635: molecule'', {\it Phys. Rep.} {\bf 288} (1997), 13-60
2636: 
2637: \bibitem{GaePLA1} G. Gaeta, ``On a model of DNA torsion dynamics'',
2638: {\it Phys. Lett. A} {\bf 143} (1990), 227-232
2639: 
2640: \bibitem{GaePLA2} G. Gaeta, ``Solitons in planar and helicoidal Yakushevich model of
2641: DNA dynamics'', {\it Phys. Lett. A} {\bf 168} (1992), 383-389
2642: 
2643: \bibitem{GaeSpeed} G. Gaeta: ``A realistic version of the Y model for
2644: DNA dynamics; and selection  of soliton speed''; {\it Phys. Lett.
2645: A} {\bf 190} (1994), 301-308
2646: 
2647: \bibitem{GaeJBP} G. Gaeta, ``Results and limitations of the soliton
2648: theory of DNA'', {\it Journal of Biological Physics} {\bf 24}
2649: (1999), 81--96
2650: 
2651: \bibitem{GaeY1} G. Gaeta, ``Solitons in the Yakushevich model of
2652: DNA beyond the contact approximation'', preprint {\it
2653: q-bio/0604003} (2006)
2654: 
2655: \bibitem{GaeY2} G. Gaeta, ``Solitons in Yakushevich-like models of DNA
2656: dynamics with improved intrapair potential'', preprint {\it
2657: q-bio/0604004} (2006)
2658: 
2659: \bibitem{GRPD} G. Gaeta, C. Reiss, M. Peyrard and Th. Dauxois,
2660: ``Simple models of non-linear DNA dynamics'', {\it Rivista del
2661: Nuovo Cimento} {\bf 17} (1994) n.4, 1--48
2662: 
2663: \bibitem{Glac} http://chemistry.gsu.edu/glactone/
2664: 
2665: \bibitem{gsl} http://www.gnu.org/software/gsl/manual/gsl-ref\_35.html\#SEC474
2666: 
2667: \bibitem{Gonz} J.A. Gonzalez and M. Martin-Landrove, ``Solitons in
2668: a nonlinear DNA model'', {\it Phys. Lett. A} {\bf 191} (1994),
2669: 409-415
2670: 
2671: \bibitem{SC} C.A. Hunter and J.K.M. Sanders,
2672: ``The Nature of n-n Interactions'', {\it J. Am. Chem. Soc.} {\bf 112}
2673: (1990), 5525-5534; R. Khairoutdinov, http://www.uaf.edu/chem/467Sp05/lecture4.pdf
2674: 
2675: \bibitem{Joy} M. Joyeux and S. Buyukdagli, ``Dynamical model based
2676: on finite stacking enthalpies for homogeneous and inhomogeneous
2677: DNA thermal denaturation'', {\it Phys. Rev. E} {\bf 72} (2005),
2678: 051902; S. Buyukdagli, M. Sanrey and M. Joyeux, ``Towards more
2679: realistic dynamical models for DNA secondary structure'', {\it
2680: Chem. Phys. Lett.} {\bf 419} (2006), 434-438
2681: 
2682: \bibitem{KS} N. Komarova and A. Soffer, ``Nonlinear waves in double stranded DNA'',
2683: {\it Bull. Math. Biol.} {\bf 67} (2005), 701-718
2684: 
2685: \bibitem{Lavery} R. Lavery, A. Lebrun, J.F. Allemand, D. Bensimon and V. Croquette,
2686: ``Structure and mechanics of single biomolecules: experiments and
2687: simulation'', {\it J. Phys.: Condens. Matter} {\bf 14} (2002),
2688: R383-R414
2689: 
2690: \bibitem{Muto} V. Muto, J. Holding, P.L. Christiansen and A.C. Scott, ``Solitons in DNA'',
2691: {\it J. Biomol. Str. Dyn.} {\bf 5} (1988), 873-894; V. Muto, P.S.
2692: Lomdahl and P.L. Christiansen, ``Two-dimensional discrete model
2693: for DNA dynamics: longitudinal wave propagation and
2694: denaturation'', {\it Phys. Rev. A} {\bf 42} (19890), 7452-7458
2695: 
2696: \bibitem{NR} Numerical Recipes, see
2697: {\it http://www.library.cornell.edu/nr/bookfpdf/f10-6.pdf}
2698: 
2699: \bibitem{PDBRep} PDB repository, {\tt http://www.rcsb.org/pdb/}
2700: 
2701: \bibitem{PDBFiles} PDB files at
2702: {http://chemistry.gsu.edu/glactone/PDB/pdb.html}
2703: 
2704: \bibitem{PeyHouches} M. Peyrard (editor), {\it Nonlinear
2705: excitations in biomolecules}, (Proceedings of a workshop held in
2706: Les Houches, 1994), Springer (Berlin) and Les Editions de Physique
2707: (Paris) 1995
2708: 
2709: \bibitem{PeyNLN} M. Peyrard, ``Nonlinear dynamics and statistical
2710: physics of DNA'', {\it Nonlinearity} {\bf 17} (2004) R1-R40
2711: 
2712: \bibitem{Pushchino} M. Peyrard ed., {\it Nonlinear phenomena in biology}
2713: (proceedings of the Pushchino conference, June 23--27, 1998),
2714: published as issues 2--4 of {\it J. Biol. Phys.} {\bf 24} (1999)
2715: 
2716: \bibitem{PB} M. Peyrard and A.R. Bishop, ``Statistical mechanics
2717: of a nonlinear model for DNA denaturation'', {\it Phys. Rev.
2718: Lett.} {\bf 62} (1989), 2755-2758
2719: 
2720: \bibitem{PBD} M. Peyrard, A.R. Bishop and Th. Dauxois, ``Dynamics and
2721: thermodynamics of a nonlinear model for DNA denaturation'', {\it
2722: PRE} {\bf 47} (1992)
2723: 
2724: \bibitem{Pow} J. W. Powell, G. S. Edwards, L. Genzel, F. Kremer,
2725: A. Wittlin, W. Kubasek and W. Peticolas, ``Investigation of
2726: far-infrared vibrational modes in polynucleotides'', {\it Phys.
2727: Rev. A} {\bf 35} (1987), 3929-3939
2728: 
2729: \bibitem{Proho} E.W. Prohofsky, ``Solitons hiding in DNA and their
2730: possible significance in DNA transcription'', {\it Phys. Rev. A}
2731: {\bf 38} (1988), 1538-1541
2732: 
2733: \bibitem{SacSgu} G. Saccomandi and I. Sgura, ``The relevance of
2734: nonlinear stacking interactions in simple models of
2735: double-stranded DNA'', preprint 2006, to appeare in {\it J. Royal
2736: Soc. Interfaces}
2737: 
2738: \bibitem{Sae} W. Saenger, {\it Principles of nucleic acid
2739: structure}, Springer (Berlin) 1984
2740: 
2741: \bibitem{Sal} M. Salerno, ``Discrete model for DNA-promoter
2742: dynamics'', {\it Phys. Rev. A} {\bf 44} (1991), 5292–5297
2743: 
2744: \bibitem{SFB} S.B. Smith, L. Finzi and C. Bustamante, ``Direct mechanical
2745: measurements of the elasticity of single DNA molecules by using
2746: magnetic beads'', {\it Science} {\bf 258} (1992), 1122-1126
2747: 
2748: \bibitem{Strick} T.R. Strick, M.N. Dessinges, G. Charvin, N.H.
2749: Dekker, J.F. Allemand, D. Bensimon and V. Croquette, ``Stretching
2750: of macromolecules and proteins'', {\it Rep. Prog. Phys.} {\bf 66}
2751: (2003), 1-45
2752: 
2753: \bibitem{Takeno} S. Takeno and S. Homma, ``Topological solitons
2754: and modulated structure of bases in DNA double helices'', {\it
2755: Progr. Theor. Phys.} {\bf 70} (1983), 308-311; S. Homma and S.
2756: Takeno, ``A coupled base-rotator model for structure and dynamics
2757: of DNA'', {\it Progr. Theor. Phys.} {\bf 72} (1984), 679-693
2758: 
2759: \bibitem{TPM} N. Theodorakopoulos, M. Peyrard and R.S. MacKay,
2760: ``Nonlinear structures and thermodynamic instabilities in a
2761: one-dimensional lattice system'', {\it Phys. Rev. Lett.} {\bf 93}
2762: (2004), 258101
2763: 
2764: \bibitem{YakPLA} L.V. Yakushevich, ``Nonlinear DNA dynamics: a new
2765: model'', {\it Phys. Lett. A} {\bf 136} (1989), 413-417
2766: 
2767: \bibitem{YakSBP} L.V. Yakushevich, ``Investigation of a system of nonlinear
2768: equations simulating DNA torsional dynamics'', {\it Stud.
2769: Biophys.} {\bf 140} (1991), 163-170
2770: 
2771: \bibitem{YakPhD} L.V. Yakushevich, ``Nonlinear DNA dynamics:
2772: hyerarchy of the models'', {\it Physica D} {\bf 79} (1994), 77-86
2773: 
2774: \bibitem{YakMMS} L.V. Yakushevich, ``DNA as a nonlinear dynamical system'',
2775: {\it Macromol. Symp.} {\bf 160} (2000), 61-68
2776: 
2777: \bibitem{YakuBook} L.V. Yakushevich, {\it Nonlinear Physics of
2778: DNA}, Wiley (Chichester) 1998; second edition 2004
2779: 
2780: \bibitem{YakPRE} L.V. Yakushevich, A.V. Savin and L.I. Manevitch,
2781: ``Nonlinear dynamics of topological solitons in DNA'', {\it Phys.
2782: Rev. E} {\bf 66} (2002), 016614
2783: 
2784: \bibitem{Yomosa} S. Yomosa, ``Soliton excitations in
2785: deoxyribonucleic acid (DNA) double helices'', {\it Phys. Rev. A}
2786: {\bf 27} (1983), 2120-2125; ``Solitary excitations in
2787: deoxyribonucleic acid (DNA) double helices'', {\it Phys. Rev. A}
2788: {\bf 30} (1984), 474-480
2789: 
2790: \bibitem{VanZ} L.L. van Zandt, ``DNA soliton realistic
2791: parameters'', {\it Phys. Rev. A} {\bf 40} (1989), 6134-6137
2792: 
2793: \bibitem{Web} G. Weber, ``Sharp DNA denaturation due to solvent
2794: interaction'', {\it Europhys. Lett.} {\bf 73} (2006), 806-811
2795: 
2796: \bibitem{Zhang} Ch.T. Zhang, ``Soliton excitations in
2797: deoxyribonucleic acid (DNA) double helices'', {\it Phys. Rev. A}
2798: {\bf 35} (1987), 886-891; ``Harmonic and subharmonic resonances of
2799: microwave absorption in DNA'', {\it Phys. Rev. A} {\bf 40} (1989),
2800: 40-45
2801: 
2802: \bibitem{ZC} F. Zhang and M.A. Collins, ``Model simulations of DNA
2803: dynamics'', {\it Phys. Rev. E} {\bf 52} (1995), 4217-4224
2804: 
2805: 
2806: \end{thebibliography}
2807: 
2808: 
2809: \end{document}
2810: