nlin0612017/hex7.tex
1: \documentclass[12pt]{iopart}
2: %\usepackage{iopams}
3: \usepackage{indentfirst}
4: \usepackage{graphicx,color}
5: \usepackage{subfigure}
6: \usepackage{psfrag}
7: \usepackage[hang, small, bf, margin=20pt, tableposition=top]{caption}
8: \setlength{\abovecaptionskip}{0pt}
9: \usepackage[T1]{fontenc}
10: \usepackage[square, comma, numbers, sort&compress]{natbib}
11: \def\newblock{\hskip .11em plus .33em minus .07em}
12: \newfont{\bbold}{msbm10 scaled\magstep1}
13: \newcommand{\bbf}[1]{\mbox{{\bbold #1}}}
14: 
15: % ABBREVIATIONS
16: \newcommand{\spc}{\quad}
17: \newcommand{\dt}{\dot}
18: \newcommand{\tbar}{\overline}
19: \newcommand{\bbar}{\underline}
20: \newcommand{\lft}{\left}
21: \newcommand{\rit}{\right}
22: \newcommand{\clyd}{\partial}
23: \newcommand{\mcal}{\mathcal}
24: \newcommand{\npg}{\newpage}
25: \newcommand{\eee}{\varepsilon}
26: \newcommand{\kkk}{\kappa}
27: \newcommand{\w}{\omega}
28: \newcommand{\om}{\Omega}
29: \newcommand{\ld}{\lambda}
30: \newcommand{\tht}{\theta}
31: \newcommand{\degrees}{\mbox{$^\circ$}}
32: \newcommand{\sech}{\mathrm{sech}}
33: \newcommand{\dd}{{\mathrm d}}
34: \newcommand{\ee}{{\mathrm e}}
35: \newcommand{\ii}{{\mathrm i}}
36: \newcommand{\sfr}[2]{\mbox{$\frac{#1}{#2}$}}
37: \newcommand{\shalf}{\sfr{1}{2}}
38: \newcommand{\elo}{E_0}
39: \newcommand{\wbr}{\mcal{W}_{\mathrm{br}}}
40: \newcommand{\ptwo}{\lft(\frac{p}{2}\rit)}
41: \newenvironment{eqn}{\begin{equation}}{\end{equation}}
42: \newenvironment{ary}{\begin{eqnarray}}{\end{eqnarray}}
43: \newenvironment{aln}{\begin{align}}{\end{align}}
44: \newenvironment{enum}{\begin{enumerate}}{\end{enumerate}}
45: \newenvironment{lst}{\begin{itemize}}{\end{itemize}}
46: \newcommand{\itc}{\emph}
47: \newcommand{\str}{\textrm}
48: \newcommand{\mb}{\mathbf}
49: \newcommand{\tb}{\textbf}
50: \newcommand{\fns}{\footnotesize}
51: \newcommand{\ssz}{\scriptsize}
52: 
53: %COMMANDS TO DO WITH NUMBERING
54: \newcommand{\lbl}{\label}
55: %\newcommand{\lbl}[1]{\framebox{\scriptsize #1}\qquad \label{#1}}
56: \makeatletter
57: \@addtoreset{equation}{section}
58: \makeatother
59: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
60: 
61: %----------------------------------------------------------------------------
62: \begin{document}
63: 
64: \submitto{\JPA}
65: 
66: \title{Discrete breathers in a two-dimensional 
67: hexagonal Fermi-Pasta-Ulam lattice}
68: \author{Imran A Butt and Jonathan A D Wattis}
69: \address{Theoretical Mechanics,  School of Mathematical Sciences,
70: University of Nottingham, University Park, Nottingham, NG7 2RD,
71: UK} \eads{\mailto{imran.butt@maths.nott.ac.uk},
72: \mailto{jonathan.wattis@nottingham.ac.uk}}
73: 
74: \begin{abstract} \lbl{abs}
75: We consider a two-dimensional Fermi-Pasta-Ulam (FPU) lattice 
76: with hexagonal symmetry.  Using asymptotic methods based on 
77: small amplitude ansatz, at third order we obtain a reduction to a 
78: cubic nonlinear Schr{\"o}dinger equation (NLS) for the breather 
79: envelope.  However, this does not support stable soliton solutions, 
80: so we pursue a higher-order analysis yielding a generalised NLS, 
81: which includes known stabilising terms.   We present numerical 
82: results which suggest that long-lived stationary and moving breathers 
83: are supported by the lattice. We find breather solutions which move 
84: in an arbitrary direction, an ellipticity criterion for the wavenumbers 
85: of the carrier wave, asymptotic estimates for the breather energy, 
86: and a minimum threshold energy below which breathers cannot be 
87: found.  This energy threshold is maximised for stationary breathers, 
88: and becomes vanishingly small near the boundary of the elliptic domain 
89: where breathers attain a maximum speed.  Several of the results obtained 
90: are similar to those obtained for the square FPU lattice (Butt \& Wattis, 
91: {\em J Phys A}, {\bf 39}, 4955, (2006)),  though we find that the square 
92: and hexagonal lattices exhibit different properties in regard to the 
93: generation of harmonics, and the isotropy of the generalised NLS equation.  
94: \end{abstract}
95: 
96: \pacs{05.45.-a, 05.45.Yv \\[2ex]This version: \today}
97: 
98: 
99: %--------------------------------------------------------------
100: \section{Introduction} \lbl{hexint}
101: 
102: 
103: Discrete breathers are time-periodic and spatially localised exact
104: solutions of translationally invariant nonlinear lattices.  For a
105: brief review of some general properties of breathers in
106: higher-dimensional systems, see our earlier work \citep{buts06}.  
107: In particular, it is known that while some fundamental properties
108: such as the existence of breathers are not affected by lattice
109: dimension (see Flach \etal \citep{flat94} and Mackay and Aubry
110: \citep{mac94}), other properties are affected profoundly, for
111: instance, the energy properties of breathers (see Flach \etal
112: \citep{fla97}, Kastner \citep{kas04} and Weinstein \citep{wei99}).
113: In this paper we investigate how the symmetry of the lattice 
114: influences the properties of discrete breathers found therein.  
115: 
116: 
117: In Hamiltonian systems, stationary breathers occur in one-parameter 
118: families.  For a certain class of Hamiltonian systems, a critical spatial 
119: dimension $d_c$ exists such that for systems with $d \geq d_c$, 
120: there exists a positive lower bound on the energy of a breather 
121: family, and the breather energies do {\em not} approach zero even 
122: as the amplitude tends to zero.   For lattices with dimension 
123: $d < d_c$, there is no positive lower bound on the energy of 
124: breathers.  In other words, the energy of a family of breathers goes 
125: to zero with amplitude, and breathers of arbitrarily small energy can 
126: be found.    The critical dimension $d_c$ is typically two.  A small 
127: amplitude expansion yields the NLS reduction to $\ii F_T+\Delta F 
128: + \kappa |F|^{2\sigma}F=0$ which has a critical dimension of  
129: $d_c=2/\sigma$, with blow-up in NLS occurring when $\kappa>0$ 
130: and $d>d_c$.   For typical lattice potentials, $\sigma=1$ again 
131: confirming the critical dimension of $d_c=2$. 
132: 
133: 
134: Marin, Eilbeck and Russell have performed extensive numerical 
135: investigations of breather dynamics in two-dimensional lattices 
136: \citep{mar98,mar00,mar01}.  Their results suggest that moving 
137: breather modes exist (or are at least extremely long-lived), and 
138: that the lattice exhibits a strong directional preference whereby 
139: breathers can only move along symmetries of the lattice, and in 
140: no other direction.  Such quasi-one-dimensional behaviour is also 
141: observed in higher-dimensional lattices.
142: 
143: 
144: The work in this paper follows on from our earlier study 
145: \citep{buts06} of breathers in a two-dimensional lattice with square 
146: rotational symmetry;  hereupon referred to as the ``square'' lattice.  
147: This lattice has $C_4$ symmetry, by which we mean that a rotation 
148: through any multiple of $2\pi/4 = \pi/2$ about an axis perpendicular to 
149: the lattice plane through a lattice site maps the lattice onto itself.  
150: %
151: In \cite{buts06}, using the semi-discrete multiple-scale method 
152: (see Remoissenet \citep{rem85}), we determined an approximate 
153: form for (as well as the properties of) small 
154: amplitude breathers in a two-dimensional square Fermi-Pasta-Ulam 
155: (FPU) lattice \citep{fer40}.  We found a third-order analysis to be 
156: inadequate, since the partial differential equation obtained at this order 
157: describing the breather envelope exhibits \itc{blow-up}.  To
158: overcome this, we incorporated higher-order effects in the model,
159: thereby obtaining a modified partial differential equation which
160: includes known stabilising terms. From this, we determined regions
161: of parameter space where breather solutions are expected.
162: Numerical simulations supported the results of our analysis, and
163: suggested that, in contrast to the two-component lattices studied
164: by Marin \etal\ \citep{mar98,mar00,mar01} (that is, with two
165: degrees of freedom at each lattice site), there is no restriction upon 
166: the permitted directions of travel within the one-component square 
167: FPU lattice. We also found asymptotic estimates for the breather 
168: energy which confirmed the existence of a minimum threshold 
169: energy, in agreement with the work of Flach \etal\ \citep{fla97}.
170: 
171: 
172: \begin{figure}[htbp]
173: \begin{center}
174: \resizebox{5in}{!}{\includegraphics{hexlatsmen.eps}}
175: \caption{The 2D hexagonal electrical transmission lattice (HETL).}
176: \lbl{hexetl}
177: \end{center}
178: \end{figure}
179:  
180: 
181: In this paper, we consider a hexagonal electrical transmission lattice 
182: (HETL).  This two-dimensional network possesses $C_6$ (or hexagonal) 
183: rotational symmetry.  That is, rotation through any multiple of the angle 
184: $2\pi/6 = \pi/3$ about a lattice site maps the lattice onto itself.  The HETL 
185: is shown in \Fref{hexetl}, pictured from a point vertically above the plane 
186: of the lattice.  We note from \Fref{hexetl} that geometrically, the HETL is 
187: an arrangement of tessellating triangles (not hexagons).   Nevertheless, the 
188: arrangement in \Fref{hexetl} is referred to as ``hexagonal'' rather than 
189: ``triangular'' since these descriptions refer to its symmetry properties and 
190: not to the geometrical shapes which comprise the array (not all authors 
191: follow this convention).  One might expect the analysis for the hexagonal 
192: lattice to be more involved, since it is geometrically more complicated in 
193: having more links emanating from each node.  However, the hexagonal 
194: symmetry results in greater isotropy and hence simpler equations than 
195: those obtained for the square lattice.
196: 
197: 
198: We derive the equations of motion and demonstrate a Hamiltonian
199: formalism in \Sref{hexder}.  In Sections \ref{hexsym} and
200: \ref{hexasymm}, we present two cases for which the hexagonal FPU
201: lattice equations can be reduced to a two-dimensional nonlinear
202: Schr{\"o}dinger (NLS) equation with cubic nonlinearity.  We
203: consider lattices with a symmetric interaction potential, in which
204: case reduction to a cubic NLS equation can be performed for moving
205: breathers.  We find an ellipticity criterion for the wavenumbers
206: of the carrier wave in \Sref{hexdell}.  A reduction to the cubic NLS
207: equation can also be carried out for lattices with an asymmetric
208: potential, provided we consider only stationary breathers.  
209: 
210: 
211: As expected we find a minimum energy below which breathers cannot 
212: exist in the hexagonal FPU lattice.  We find that the
213: energy threshold is dependent upon the wavevector of the carrier
214: wave. It is maximised for stationary breathers, and becomes
215: arbitrarily small near the boundary of the elliptic domain.   
216: 
217: 
218: The cubic NLS equation admits only unstable Townes solitons.  Hence, 
219: in \Sref{hex5}, we extend our asymptotic analysis to higher order and 
220: find an isotropic generalised NLS equation which incorporates known 
221: stabilising terms.  Our analytic work is supplemented by numerical simulations 
222: presented in \Sref{hexnumerics}, which suggest that long-lived stationary
223: and moving breather modes are supported by the system.
224: In \Sref{hdisc}, we discuss the results obtained in this paper. 
225: 
226: 
227: %==================================================
228: \section{A two-dimensional hexagonal Fermi-Pasta-Ulam lattice}
229: \lbl{hexfpulat}
230: 
231: 
232: \subsection{Preliminaries} \lbl{hexpre}
233: 
234: 
235: Before we derive the equations governing the HETL, we describe our 
236: scheme for indexing the lattice nodes and our choice of basis vectors.  We 
237: introduce a rectangular lattice with basis vectors $\mcal{B}~=~\{ \mb{i'} \, , 
238: \, \mb{j'} \}$, where $\mb{i'} = \mb{i} = [1,0]^T$ and $\mb{j'} = [0,h]^T$ 
239: ($\mb{j}$ being $[0,1]^T$), illustrated in \Fref{sqbase}.  We use only half of 
240: the $(m,n)$ indices, namely, those for which the sum $m+n$ is even.  We 
241: choose an origin with coordinates $(0,0)$;  the position of the site $(m,n)$ 
242: is $m\mb{i'} + n\mb{j'}$.  In order for the hexagonal lattice to be regular, 
243: we specify $h=\sqrt{3}$. 
244: 
245:  
246: \begin{figure}[htbp]
247: \begin{center}
248: \vspace{-1.0cm}
249: \psfrag{0}{{\scriptsize $(m,n)$}}
250: \psfrag{1}{{\scriptsize $(m+2,n)$}}
251: \psfrag{2}[B][]{{\scriptsize $(m+1,n+1)$}}
252: \psfrag{3}[B][]{{\scriptsize $(m-1,n+1)$}}
253: \psfrag{4}[B][r]{{\scriptsize $(m-2,n)$}}
254: \psfrag{5}[B][b]{{\scriptsize $(m-1,n-1)$}}
255: \psfrag{6}[B][b]{{\scriptsize $(m+1,n-1)$}}
256: \psfrag{i}{{\footnotesize $\mb{i'}$}}
257: \psfrag{j}{{\footnotesize $\mb{j'}$}}
258: \resizebox{4in}{!}{\includegraphics{sqbasecorr2.eps}}
259: \caption{Labelling of nodes in the HETL with basis 
260: $ \mcal{B} = \{ \mb{i'} \, , \, \mb{j'} \}$.}
261: \lbl{sqbase}
262: \end{center}
263: \end{figure}
264: 
265: 
266: At every node of the HETL lies a nonlinear capacitor (not shown in 
267: \Fref{hexetl}), and between every node and each of its six nearest 
268: neighbours is a linear inductor.  An enlarged view of the area surrounding 
269: the capacitor at $(m,n)$ is shown in \Fref{hexunitdir}, where the capacitor 
270: is visible.  The variable $V_{m,n}$ denotes the voltage across the 
271: capacitor $(m,n)$ and $Q_{m,n}$ denotes the total charge stored on the 
272: capacitor at $(m,n)$.  Also, $I_{m,n}$, $J_{m,n}$ and $K_{m,n}$ are the 
273: currents through the inductors immediately to the right of site $(m,n)$ in 
274: the directions $\mb{e_i} = [2,0]^T$, $\mb{e_j} = [1,-\sqrt{3}]^T$ and 
275: $\mb{e_k}=[1,\sqrt{3}]^T$ respectively, as illustrated in \Fref{hexunitdir}. 
276: 
277: 
278: \begin{figure}[htbp]
279: \begin{center}
280: \vspace{1.3cm}
281: \psfrag{a}{{\large $V_{m,n}$}}
282: \psfrag{b}[B][]{{\large $V_{m+1,n-1}$}}
283: \psfrag{c}[B][]{{\large $V_{m+2,n}$}}
284: \psfrag{d}{{\large $V_{m+1,n+1}$}}
285: \psfrag{e}{{\large $I_{m,n}$}}
286: \psfrag{f}{{\large $J_{m,n}$}}
287: \psfrag{g}{{\large $K_{m,n}$}}
288: \psfrag{i}{{\normalsize $\mb{e_i}$}}
289: \psfrag{j}{{\normalsize $\mb{e_j}$}}
290: \psfrag{k}{{\normalsize $\mb{e_k}$}}
291: \resizebox{5.5in}{!}{\includegraphics{hexundir.eps}}
292: \caption{Enlarged view of the HETL at site $(m,n)$.}
293: \lbl{hexunitdir}
294: \end{center}
295: \end{figure}
296: 
297: 
298: %----------------------------------------------------------------
299: \subsection{Derivation of model equations} \lbl{hexder}
300: 
301: 
302: To derive the equations relating current, charge and voltage in the lattice
303: we apply Kirchoff's law 
304: \begin{eqnarray}
305: V_{m+2,n} - V_{m,n} & = -L \frac{\dd I_{m,n}}{\dd t}, \lbl{hki} \\[1ex]
306: V_{m+1,n-1} - V_{m,n} & = -L \frac{\dd J_{m,n}}{\dd t}, \lbl{hkj} \\[1ex]
307: V_{m+1,n+1} - V_{m,n} & = -L \frac{\dd K_{m,n}}{\dd t}, \lbl{hkk}
308: \end{eqnarray}
309: where the inductance $L$ is constant.  Conservation of charge gives
310: \begin{eqn} \lbl{hcc}
311: I_{m-2,n}-I_{m,n} + J_{m-1,n+1}-J_{m,n} + K_{m-1,n-1}-K_{m,n} =
312: \frac{\dd Q_{m,n}}{\dd t}.
313: \end{eqn}
314: Differentiating \eref{hcc} with respect to time, and then using 
315: \eref{hki}--\eref{hkk} to find $\dot{I}_{m-2,n}$, $\dot{J}_{m-1,n+1}$ 
316: and $\dot{K}_{m-1,n-1}$, we have  
317: \begin{eqnarray}
318: L \frac{\dd^2 Q_{m,n}}{\dd t^2} & = & \spc (V_{m+2,n} - 2 V_{m,n} + 
319: V_{m-2,n}) + (V_{m+1,n-1} - 2 V_{m,n} + V_{m-1,n+1}) \nonumber \\ & & 
320: + (V_{m+1,n+1} - 2 V_{m,n} + V_{m-1,n-1}). 
321: \lbl{heqvqfull} \end{eqnarray}
322: Equation \eref{heqvqfull} may be written in the abbreviated form
323: $L \ddot{Q}_{m,n} = (\delta^2_I + \delta^2_J + \delta^2_K)V_{m,n}$,
324: where the centred second-difference operators are defined by
325: \begin{eqnarray}
326: \lbl{hdi} \delta^2_I A_{m,n} & = A_{m+2,n}   - 2 A_{m,n} + A_{m-2,n},\\[1ex]
327: \lbl{hdj} \delta^2_J A_{m,n} & = A_{m+1,n-1} - 2 A_{m,n} + A_{m-1,n+1},\\[1ex]
328: \lbl{hdk} \delta^2_K A_{m,n} & = A_{m+1,n+1} - 2 A_{m,n} + A_{m-1,n-1} . 
329: \end{eqnarray}
330: Here, $A_{m,n}$ is an arbitrary quantity referenced by two indices; $\delta^2_I$, 
331: $\delta^2_J$ and $\delta^2_K$ are centred second-difference operators in 
332: the directions of $\mb{e_i}$, $\mb{e_j}$ and $\mb{e_k}$ respectively.   
333: Since the voltage $V_{m,n}$ is known in terms of the charge, we reformulate 
334: \eref{heqvqfull} in terms of the single quantity $Q_{m,n}$.  We invert the 
335: capacitor's nonlinear charge-voltage relationship $Q=VC(V)$ (see 
336: equations (2.9)--(2.14) of \cite{buts06} for details), to obtain 
337: \begin{equation} 
338: V(Q) = ( Q + a Q^2 + b Q^3 + c Q^4 + d Q^5 ) / C_0 , 
339: \lbl{hqvexp} \end{equation}
340: where $C_0=C(0)$.  Hence, the HETL equations \eref{heqvqfull} can 
341: be written as 
342: \begin{eqn} \lbl{heqq}
343: \frac{\dd^2 Q_{m,n}}{\dd t^2} = (\delta^2_I + \delta^2_J + \delta^2_K) 
344: \lft[Q_{m,n}+aQ_{m,n}^2+bQ_{m,n}^3+cQ_{m,n}^4+dQ_{m,n}^5\rit], 
345: \end{eqn}
346: where $m,n \in {\bbf{Z}}$ and, by rescaling the time variable, we set 
347: $LC_0=1$ without loss of generality.   Thus we have shown that the 
348: equation governing charge in the HETL \eref{heqq} is a two-dimensional 
349: analogue of the Fermi-Pasta-Ulam equation 
350: \begin{equation}
351: \frac{\dd^2 Q_j}{\dd t^2} = W'(Q_{j+1}) - 2 W'(Q_j) + W'(Q_{j-1}) , 
352: \end{equation}
353: which is a Hamiltonian system that can be derived from both 
354: \begin{equation}
355: H = \sum_j \shalf \pi_j^2 + W(\phi_{j+1}-\phi_j) , \;\;\; {\rm and} \;\;\; 
356: \widetilde H = \sum_j \shalf (P_{j+1}-P_j)^2 + W(Q_j) , 
357: \end{equation}
358: where $Q_j=\phi_{j+1}-\phi_j$.  
359: 
360: 
361: The lattice equations \eref{heqq} can be derived from the Hamiltonian 
362: \begin{eqnarray}
363: \widetilde{H} & =\sum_{m,n} & 
364: \mbox{$\frac{1}{2}$} (P_{m+2,n}-P_{m,n})^2 + 
365: \mbox{$\frac{1}{2}$} (P_{m+1,n-1} - P_{m,n})^2 \nonumber \\ & & + 
366: \mbox{$\frac{1}{2}$} (P_{m+1,n+1} - P_{m,n})^2 + \Upsilon (Q_{m,n}), 
367: \lbl{hham} \end{eqnarray}
368: where $\Upsilon (Q_{m,n})$ satisfies $\Upsilon\!\ '(Q_{m,n}) =
369: V(Q_{m,n})$ given in \eref{hqvexp}, hence   
370: \begin{eqn}
371: \Upsilon(Q) = \mbox{$\frac{1}{2}$} Q^2 + 
372: \mbox{$\frac{1}{3}$} a Q^3 + \mbox{$\frac{1}{4}$} b Q^4 + 
373: \mbox{$\frac{1}{5}$} c Q^5 + \mbox{$\frac{1}{6}$} d Q^6 .
374: \end{eqn} 
375: We describe potentials which satisfy $\Upsilon(-Q)=\Upsilon(Q)$ (that is, 
376: $a=c=0$) as `symmetric'.  The variables $P_{m,n}$ and $Q_{m,n}$ are 
377: conjugate momenta and displacement variables of the system; and 
378: eliminating $P_{m,n}$ from the equations 
379: \begin{eqn} \lbl{hhameq}
380: \frac{\dd Q_{m,n}}{\dd t} = -(\delta^2_I + \delta^2_J  +
381: \delta^2_K)P_{m,n}, \qquad
382: \frac{\dd P_{m,n}}{\dd t} = -\Upsilon\!\ '(Q_{m,n}) , 
383: \end{eqn}
384: yields (\ref{heqq}). In \Sref{heeh}, we derive expressions for 
385: the energy of breathers given the small amplitude solutions 
386: which are obtained in the next section. 
387: 
388: 
389: %--------------------------------------------------------------------
390: \subsection{Asymptotic analysis} \lbl{haa}
391: 
392: 
393: We apply the method of multiple-scales to determine an approximate 
394: analytic form for small amplitude breather solutions of \eref{heqq}, with 
395: slowly varying envelope.  We introduce new variables defined by
396:  \begin{eqn} \lbl{hexvars}
397:  X=\eee m,\spc Y=\eee h n,\spc\tau=\eee t\spc\str{and}\spc T=\eee^2 t;
398:  \end{eqn}
399: note the presence of the scaling factor $h$ in the definition of $Y$.  
400: We seek solutions of \eref{heqq} of the form
401: \begin{eqnarray} \fl 
402: Q_{m,n}(t) = \hspace*{-3mm} & \!\!\!\! \eee \ee^{\ii \psi} F(X,Y,\tau,T) + 
403: \eee^2 G_0(X,Y,\tau,T) +
404: \eee^2 \ee^{\ii \psi} G_1(X,Y,\tau,T) + \nonumber \\ & \!\!\!\! 
405: \eee^2 \ee^{2\ii \psi} G_2(X,Y,\tau,T) + \eee^3 H_0 (X,Y,\tau,T) +
406: \eee^3 \ee^{\ii \psi} H_1(X,Y,\tau,T) + \nonumber \\
407: & \!\!\!\!  \eee^3 \ee^{2\ii \psi} H_2(X,Y,\tau,T)
408: + \eee^3\ee^{3\ii \psi}H_3(X,Y,\tau,T) + \eee^4\ee^{\ii\psi}I_1(X,Y,\tau,T)
409: + \nonumber\\ & \!\!\!\!  
410: \eee^4 \ee^{2\ii \psi} I_2(X,Y,\tau,T) + \eee^4 \ee^{3\ii\psi} I_3(X,Y,\tau,T)
411: + \eee^4\ee^{4\ii \psi} I_4(X,Y,\tau,T) + \nonumber \\
412: & \!\!\!\! \eee^5\ee^{\ii \psi} J_1(X,Y,\tau,T)+ \cdots + \mathrm{c.c.},
413: \lbl{hexanz} \end{eqnarray}
414: where the phase $\psi$ of the carrier wave is given by $km+lhn+\w t$ (once 
415: again noting the extra factor $h$), and $\mb{k}=[k,l]^T$ and $\w(\mb{k})$ 
416: are its wavevector and temporal frequency respectively.  We substitute the 
417: ansatz \eref{hexanz} into the lattice equations \eref{heqq} and equate 
418: coefficients of each harmonic frequency at each order of $\eee$.  After 
419: much simplification, this yields the following system of equations: \\
420: \\
421: $\mathcal{O}(\eee \ee^{\ii\psi})$:
422: \begin{eqn} \lbl{hexdisp}
423: \w^2 F = 4\sin^2(k) F + 4\sin^2 \lft( \frac{k+lh}{2} \rit) F + 
424: 4\sin^2 \lft( \frac{k-lh}{2} \rit) F,
425: \end{eqn}
426: $\mathcal{O}(\eee^2 \ee^{\ii\psi})$:
427: \begin{eqn} \lbl{hexvel}
428: \w F_{\tau} = 2\sin k [2\cos k + \cos(lh)]F_X + 2h\cos k \sin(lh) F_Y,
429: \end{eqn}
430: $\mcal{O}(\eee^2 \ee^{2\ii\psi})$:
431: \begin{eqn} \lbl{hexg2}
432: \w^2 G_2 = [\sin^2(2k) + \sin^2(k+lh) + \sin^2(k-lh)](G_2 + a F^2),
433: \end{eqn}
434: $\mcal{O}(\eee^3 \ee^{\ii\psi})$:
435: \begin{eqnarray}
436: \fl 2\ii \w F_T + F_{\tau\tau}  & 
437: = & [4\cos(2k) + 2\cos k \cos(lh)] F_{XX} + 2h^2\cos k \cos(lh)
438: F_{YY} \nonumber \\
439: & & - 4h\sin k \sin(lh)F_{XY} \nonumber \\
440: & & - \ 8a \lft[ \sin^2(k) + \sin^2 \lft( \frac{k+lh}{2} \rit)
441: + \sin^2 \lft( \frac{k-lh}{2} \rit) \rit] [F(G_0+\tbar{G}_0)+\tbar{F}G_2] 
442: \nonumber\\
443: & & - \ 12 b \lft[ \sin^2(k) + \sin^2 \lft( \frac{k+lh}{2} \rit)
444: + \sin^2 \lft( \frac{k-lh}{2} \rit) \rit] |F|^2 F,  \lbl{hexnls}
445: \end{eqnarray}
446: $\mcal{O}(\eee^3 \ee^{3\ii\psi})$:
447: \begin{eqn} \fl 
448: 9\w^2H_3 =  4 \lft[ \sin^2(3k) + \sin^2 \lft( \frac{3k+3lh}{2} \rit) + 
449: \sin^2 \lft( \frac{3k-3lh}{2} \rit) \rit] \left( H_3 + 2aFG_2+bF^3 \right), 
450: \lbl{hexh3} \end{eqn}
451: $\mcal{O}(\eee^4 \ee^0)$:
452: \begin{eqn} \lbl{hexfo}
453: G_{0\tau\tau} = 6\, G_{0XX} + 2h^2 G_{0YY} + 
454: a \! \lft[ 6 \lft( |F|^2 \rit)_{XX} + 2h^2 \lft(|F|^2 \rit)_{YY} \rit] .
455: \end{eqn}
456: 
457: 
458: Though each equation plays a similar role to its counterpart in the square 
459: lattice, equations \eref{hexdisp}--\eref{hexfo} are more complicated. 
460: Equation \eref{hexdisp} is the dispersion relation for the system \eref{heqq}.  
461: Since we are interested only in solutions for which $F \neq 0$, \eref{hexdisp} yields 
462: \begin{eqn} \lbl{hexdisp2} 
463: \w^2  = 4\sin^2(k)  + 4\sin^2 \lft( \frac{k+lh}{2} \rit) + 
464: 4\sin^2 \lft( \frac{k-lh}{2} \rit),
465: \end{eqn}
466: which does not simplify significantly.  From equation \eref{hexvel}, 
467: we determine the velocity of the travelling wave $F$, finding 
468: \begin{eqn}  \lbl{hexvels}
469: F(X,Y,\tau,T) \equiv F(Z,W,T),
470: \end{eqn}
471: where $Z=X-u\tau$ and $W=Y-v\tau$, and the horizontal and vertical
472: velocity components ($u$ and $v$) are found to be
473: \begin{eqn}  \lbl{velsuv}
474: u = \frac{-2\sin (k)[2\cos(k)+\cos(lh)]}{\w} \spc\str{and}\spc
475: v = \frac{-2h\cos(k)\sin(lh)}{\w}.
476: \end{eqn}
477: Equation \eref{velsuv}, along with \eref{hexdisp2} enables the elimination 
478: of terms involving $G_1$ from \eref{hexnls} which are not shown.  We 
479: denote the angle at which the envelope $F$ propagates through the 
480: lattice by $\Psi$, which is measured from the direction of the basis vector 
481: $\mb{e_i}$ to the line of travel and hence is given by $\tan^{-1}(v/u)$, 
482: which in turn depends upon the wavevector  $\mb{k}$.   For both the cases 
483: that we consider (namely, symmetric and asymmetric interaction potentials), 
484: we find constraints upon the wavenumbers $k$ and $l$ which affect the 
485: velocity components $u$ and $v$.   By taking $k=\pi/3$ and  $l=\pi/h$ we 
486: find $u=v=0$, which corresponds to a static breather; and by choosing 
487: $k,l$ values which circle this point, we find breathers which can propagate 
488: in any direction ($0\leq\Psi<2\pi$); in other words, our analysis suggests 
489: that there is no restriction upon the direction of travel through the lattice.
490: 
491: 
492: Our aim is to reduce \eref{hexnls} to a nonlinear
493: Schr{\"o}dinger (NLS) equation for $F$.  Before this
494: can be done, the quantities $G_0$ and $G_2$ in \eref{hexnls} must
495: be found in terms of $F$.  As for the square lattice, it is straightforward
496: to determine $G_2$ from the algebraic equation \eref{hexg2}.
497: However, the partial differential equation \eref{hexfo} for $G_0$
498: can be solved for two special cases only; namely, 
499: symmetric potentials, in which case the reduction of \eref{hexnls}
500: to an NLS equation can be completed even for moving breathers, and
501: also asymmetric potentials, provided we confine our attention to
502: stationary breathers.  These two cases are considered in Sections 
503: \ref{hexsym} and \ref{hexasymm} respectively, where we also use 
504: our breather formulae to generate estimates for the breather energy.
505: 
506: 
507: %-------------------------------------------------------------------
508: \subsection{Asymptotic estimates for breather energy} \lbl{heeh}
509: 
510: 
511: The HETL is a lossless network, meaning that the total electrical energy 
512: is conserved.  This quantity is related to the Hamiltonian 
513: (\ref{hham}) by $E=\widetilde{H}/C_0$ and so is given by 
514: \begin{eqn} \lbl{hexentot}
515: E = \sum_{m,n} e_{m,n} =  \sum_{m,n} \frac{\Upsilon(Q_{m,n})}{C_0} 
516: + \mbox{$\frac{1}{2}$}L \lft( I_{m,n}^2 + J_{m,n}^2 + K_{m,n}^2 \rit) . 
517: \end{eqn}
518: The electrical energy $e_{m,n}$ associated with a unit of the lattice is 
519: (see \Fref{hexunitdir})
520: \begin{eqn} \lbl{hexenun} 
521: e_{m,n} = \frac{\Upsilon(Q_{m,n})}{C_0} + 
522: \mbox{$\frac{1}{2}$}L ( I_{m,n}^2 + J_{m,n}^2 + K_{m,n}^2 ) . 
523: \end{eqn}
524: To derive a leading-order estimate for the electrical energy, defined by 
525: \begin{eqn}  \lbl{hentotlo}
526: E_0 = \sum_{m,n} e_{m,n}^{(0)}=\sum_{m,n}\frac{Q_{m,n}^2}{2C_0}
527: + \frac{L}{2} \lft( I_{m,n}^2 + J_{m,n}^2 + K_{m,n}^2 \rit) , 
528: \end{eqn}
529: we use leading-order expressions for each of the terms in the terms in 
530: the summand of \eref{hexentot}.  The first term is $Q_{m,n}^2/(2C_0)$;  
531: from \eref{hqvexp}, it follows that $V_{m,n} \sim Q_{m,n}/C_0$, and 
532: so to leading order, (\ref{hexentot}) agrees with the linear approximation 
533: to the energy in the capacitor being $QV/2$.   To find leading-order 
534: expressions for the currents  $I_{m,n}$, $J_{m,n}$ and $K_{m,n}$, we 
535: use equations \eref{hki}--\eref{hkk}, which imply 
536: \begin{eqnarray}
537: Q_{m+2,n} - Q_{m,n} & = - \frac{\dd I_{m,n}}{\dd t}, \lbl{hkiq} \\[1ex]
538: Q_{m+1,n-1} - Q_{m,n} & = - \frac{\dd J_{m,n}}{\dd t}, \lbl{hkjq} \\[1ex]
539: Q_{m+1,n+1} - Q_{m,n} & = - \frac{\dd K_{m,n}}{\dd t}, \lbl{hkkq}
540: \end{eqnarray}
541: where $LC_0=1$.  The currents are determined by substituting the
542: expression for the breather $Q_{m,n}$ into \eref{hkiq}--\eref{hkkq} 
543: and then integrating with respect to time.  We obtain leading-order 
544: estimates for the energy of moving breathers in systems with symmetric 
545: potentials Section \ref{hemv}, and in Section \ref{hest}, the energies 
546: of stationary breathers with asymmetric potentials.
547: 
548: 
549: %----------------------------------------------------------------
550: \subsection{The dispersion relation for the HETL} \lbl{wplots}
551: 
552: 
553: In this section, we analyse the dispersion relation \eref{hexdisp2} 
554: for the system \eref{heqq}.  A contour plot of $\w$ against $k$ and 
555: $l$ is shown in \Fref{wcont}.  Since $w$ is periodic in both $k$ and 
556: $l$, with period $2\pi$ along the $k$-direction, and $2\pi/h$ in the 
557: $l$-direction, we consider only $k$ and $l$ such that 
558: $(k,l) \in \mcal{T}^2 = [0,2\pi] \times [0,2\pi/h]$.
559: 
560: 
561: The function $\w$ is minimised, and assumes the value zero, at the
562: centre of the circular patterns in \Fref{wcont}.  The minima are located 
563: at $(0,0)$, $(2\pi,0)$, $(2\pi,2\pi/h)$, $(0,2\pi/h)$ and $(\pi,\pi/h)$.  
564: Plus signs (`$+$') mark the points in $(k,l)$-space at which $\w$ is 
565: maximised and takes the value $\w=3$.  The maxima lie at the centres 
566: of the equilateral triangles, the wavevectors corresponding to these 
567: points are denoted $\mb{k_1}, \ldots, \mb{k_6}$, where
568: \begin{equation} \begin{array}{ lclcl}
569: \mb{k_1}=[\pi/3,\pi/h]^T,  && 
570: \mb{k_2}=[2\pi/3,0]^T,  && 
571: \mb{k_3}=[4\pi/3,0]^T,    \\ 
572: \mb{k_4}=[5\pi/3,\pi/h]^T, && 
573: \mb{k_5}=[4\pi/3,2\pi/h]^T, &&
574: \mb{k_6}=[2\pi/3,2\pi/h]^T.
575: \end{array} \end{equation} 
576: The arrangement of the points corresponding to these wavevectors
577: in $(k,l)$-space reflects the hexagonal symmetry properties of the
578: lattice \eref{heqq}.  It may be verified using \eref{velsuv} that
579: the velocity components $u$ and $v$ are both zero for each of the
580: wavevectors $\{ \mb{k_1},\ldots, \mb{k_6} \}$.
581: 
582: 
583: \begin{figure}[htbp]
584: \begin{center}
585: \resizebox{4.5in}{!}{\includegraphics{wcont.ps}}
586: \put(-250,103){\footnotesize $\mb{k_1}$}
587: \put(-200,44){\footnotesize $\mb{k_2}$}
588: \put(-100,44){\footnotesize $\mb{k_3}$}
589: \put(-48,103){\footnotesize $\mb{k_4}$}
590: \put(-100,183){\footnotesize $\mb{k_5}$}
591: \put(-200,183){\footnotesize $\mb{k_6}$}
592: \caption{Contour plot of $\w(\mb{k})$; $\w$ attains its maximum 
593: value of 3 at the points $\mb{k_1},\ldots,\mb{k_6}$, and 
594: is minimised at the points marked `o', where $\w=0$. }
595: \lbl{wcont}
596: \end{center}
597: \end{figure}
598: 
599: 
600: %------------------------------------------------------------------
601: \subsection{Lattices with a symmetric potential} \lbl{hexsym}
602: 
603: 
604: In this section, we consider lattices with a symmetric interaction potential, 
605: where $\Upsilon (Q)$ is even and $\Upsilon\!\ '(Q)$  has odd symmetry. 
606: This corresponds to $a=c=0$ in \eref{heqq}.  Since there are no even 
607: harmonics for vibrations controlled by symmetric potentials, it follows 
608: that $G_0$ and $G_2$ are both zero.  In \eref{hexnls}, the term 
609: $F_{\tau\tau}$ is eliminated using $F_{\tau\tau} = u^2F_{ZZ} + 2uvF_{ZW} 
610: + v^2F_{WW}$ which is derived from \eref{hexvels}.  This 
611: leads to (\ref{hexnls}) being reduced to an NLS equation 
612: \begin{eqnarray}
613: \fl 2\ii\w F_T + \lft[ u^2 - 4\cos(2k) - 2\cos k\cos(lh) \rit] F_{ZZ}
614: + \lft[ v^2 - 2h^2\cos k \cos(lh) \rit] F_{WW} \nonumber \\
615: + \lft[ 2uv + 4h \sin k \sin(lh) \rit] F_{ZW} + 3b\w^2|F|^2F = 0, 
616: \lbl{hexrnls} \end{eqnarray}
617: where the velocities $u$ and $v$ are given by \eref{velsuv}.  By
618: applying an appropriate change of variables, we remove 
619: the mixed derivative term from  \eref{hexrnls}, and reduce the
620: equation to a standard form.  To simplify the appearance of
621: subsequent expressions, we denote the coefficients of $F_{ZZ}$,
622: $F_{WW}$ and $F_{ZW}$ by $D_1 = u^2 - 4\cos(2k) - 2\cos
623: k\cos(lh)$, $D_2 =  v^2 - 2h^2\cos k \cos(lh)$ and $D_3 = 2uv + 4h
624: \sin k \sin(lh)$ respectively.  A suitable transformation is thus
625: \begin{eqn} \lbl{hextr}
626: \xi = \frac{hZ}{\sqrt{D_1}} \spc \str{and} \spc
627: \eta = \frac{h(2D_1 W - D_3 Z)}{\sqrt{D_1 (4D_1 D_2 - D_3^2)}},
628: \end{eqn}
629: which implies \eref{hexrnls} becomes
630: \begin{eqn} \lbl{hexcan}
631: 2\ii\w F_T + 3\nabla^2 F + 3b\w^2|F|^2F = 0,
632: \end{eqn}
633: where the differential operator $\nabla^2 F \equiv F_{\xi\xi} +
634: F_{\eta\eta}$ is isotropic in the $(\xi,\eta)$ variables.   An approximation 
635: to the Townes soliton solution of (\ref{hexcan}) is given by equation 
636: (\ref{vartown}) in the appendix.  Substituting the resulting expression for $F$ 
637: into \eref{hexanz} yields a leading-order expression for the breather
638: \begin{eqn} \lbl{hexlo}
639: Q_{m,n}(t) = 2 \eee \alpha \cos[km+lhn+(\w+\eee^2\lambda)t] 
640: \ \sech(\beta r) + \mcal{O}(\eee^3),
641: \end{eqn}
642: where $\alpha$ and $\beta$ are determined as described by equation 
643: (\ref{alpbet}) in the appendix, with $D=3/2\w$, $B=3b\w/2$.  Further,
644: $r=\sqrt{\xi^2+\eta^2}$ is found in terms of the physical discrete
645: variables $m$ and $n$ by inverting the transformations
646: \eref{hextr} and reverting back to the variables $Z$ and $W$, using
647: \begin{eqnarray} 
648: r^2 &=& \xi^2 + \eta^2 = \frac{ 4 h^2 \eee^2 
649: ( D_2(m\!-\!ut)^2 + D_1(hn\!-\!vt)^2 - D_3(m\!-\!ut)(hn\!-\!vt) )}
650: {4D_1D_2 -D_3^2}.  \nonumber \\ && \lbl{hexr}
651: \end{eqnarray}
652: The terms $D_1$, $D_2$ and $D_3$ are known from \eref{hexrnls}, the 
653: velocities $u$ and $v$ are given by \eref{velsuv} and $\w$ is given in 
654: \eref{hexdisp2}.
655: 
656: 
657: %------------------
658: \subsubsection{Determining the domain of ellipticity. } \lbl{hexdell}
659: 
660: 
661: We confine our attention to elliptic NLS equations.  We seek to determine 
662: the region $\mcal{D}$ of $(k,l)$-parameter space (that is, the two-torus 
663: $\mcal{T}^2 = [(0,2\pi)]\times [0,2\pi/h]$) where the NLS equation 
664: \eref{hexrnls} is elliptic. By definition, this equation is elliptic when 
665: $D_3^2 < 4 D_1 D_2$, where $D_1$, $D_2$ and $D_3$ are given in 
666: \eref{hexrnls}.  Whilst the region $\mcal{D}$ cannot be specified 
667: explicitly,  it is simple to find numerically, and is illustrated in  \Fref{econt}.  
668: Defining the function $e(k,l) = 4D_1(k,l) \! \cdot \! D_2(k,l) - D_3(k,l)^2$, 
669: we are concerned with the region where $e(k,l)>0$.  The subdomains 
670: have been labelled $\{ \mcal{D}_1 ,\ldots, \mcal{D}_6\}$ in \Fref{econt}, 
671: where $\mcal{D} = \mcal{D}_1 \cup \ldots \cup \mcal{D}_6$.  
672: 
673: 
674: \addtolength{\captionmargin}{-15pt}
675: \begin{figure}[ht]
676: \begin{center}
677: \resizebox{4.5in}{!}{\includegraphics{econt.ps}}
678: \put(-250,130){\footnotesize $\mcal{D}_1$}
679: \put(-200,44){\footnotesize $\mcal{D}_2$}
680: \put(-100,44){\footnotesize $\mcal{D}_3$}
681: \put(-48,130){\footnotesize $\mcal{D}_4$}
682: \put(-100,220){\footnotesize $\mcal{D}_5$}
683: \put(-200,220){\footnotesize $\mcal{D}_6$}
684: \caption{The domain $\mcal{D} = \mcal{D}_1 \cup \ldots \cup 
685: \mcal{D}_6$ in which the NLS equation \eref{hexrnls} is elliptic.}
686: \lbl{econt}
687: \end{center}
688: \end{figure}
689: \addtolength{\captionmargin}{+15pt}
690: 
691: 
692: Again, the hexagonal symmetry properties of the HETL are reflected clearly 
693: in the function $e(k,l)$, which has six maxima (at which $e(k,l)=36$), each
694: lying at the centre of one of the closed curves in \Fref{econt}.  The maxima 
695: of $e(k,l)$ coincide with the six maxima of $\w(k,l)$ shown in \Fref{wcont}, 
696: namely, at the wavevectors $\{\mb{k_1},\ldots,\mb{k_6}\}$ in $\mcal{T}^2$; 
697: $e(k,l)$ is minimised ($e(k,l)=-48$) at the six midpoints of the line 
698: segments which connect adjacent maxima. 
699: 
700: 
701: \begin{figure}[ht]
702: \begin{center}
703: \resizebox{2.5in}{!}{\includegraphics{WD1.eps}} \quad 
704: \resizebox{2.5in}{!}{\includegraphics{SPD1.eps}}
705: \end{center}
706: \caption{Contour plots of $\w(k,l)$ (left) and speed 
707: ($\sqrt{u^2+v^2}$, right) for wavevectors $\mb{k}\in\mcal{D}_1$. 
708: Contours are for $2.82\leq\w\leq3$ in steps of $0.02$ and speeds from 
709: zero to 0.6 in steps of 0.1.  The central point corresponds to $\w=3$ and 
710: zero speed, as one considers wave vectors nearer to the edge of 
711: $\mcal{D}_1$, the frequency $\w$ decreases and the speed increases. }
712: \lbl{wspcont}
713: \end{figure}
714: 
715: 
716: Figure \ref{wspcont} shows a contour plot for the frequency $\w$ in 
717: $\mcal{D}_1$: the ellipticity constraint only permits breathers with a 
718: relatively high frequency, that is, with $\w>2.82$.  This constraint in turn 
719: implies that not all breather envelope velocities are attainable, only breathers 
720: with speeds upto about $0.7$ lattice sites per second are permitted; the plot 
721: on the right of Figure \ref{wspcont} shows that only breathers with speeds 
722: upto about $0.3$ sites per second can move in any direction. There is then an 
723: intermediate range of speeds, between $0.3$ and $0.7$ sites per second where 
724: breathers can only move in certain directions, these correspond to the lattice 
725: directions. The larger the speed, the more restricted is the direction of motion. 
726: 
727: 
728: %-------------------
729: \subsubsection{Breather energy. } \lbl{hemv}
730: 
731: 
732: To calculate the leading-order energy $\elo$ of moving breathers in 
733: lattices with a symmetric potential we use \eref{hexlo} which we write as
734: \begin{eqn} \lbl{hxs2}
735: Q_{m,n}(t) \sim 2 \eee \alpha \cos \Phi \ \sech(\beta r),
736: \end{eqn}
737: where $\Phi = km + lhn + \om t$ is the phase of the carrier wave,
738: $(k,l) \in \mcal{D}$, $\om = \w + \eee^2\lambda$ is the breather
739: frequency including the first correction term,  ($\psi=km+lhn+\w t$ 
740: is only the leading-order expression).  $\ld$ parameterises the 
741: breather amplitude, $\alpha=\alpha(\lambda)$, $\beta=\beta(\lambda)$ 
742: and $r^2$ are as described in equations \eref{hexlo} and \eref{hexr}.  
743: 
744: 
745: We now find expressions for the currents $I_{m,n}$, $J_{m,n}$ and
746: $K_{m,n}$: the current $I_{m,n}$ is obtained by substituting the 
747: expression for $Q_{m,n}$ \eref{hxs2} into \eref{hkiq} and integrating 
748: with respect to time.   Owing to the complexity of the expression for $r$ 
749: given by \eref{hexr}, the left-hand side of \eref{hkiq} can not be integrated 
750: with respect to time. However, the variable $r$ varies more slowly in time 
751: than $\Phi$. Hence,  integration by parts (using $\int f'(t) g(\eee t) \dd t = 
752: [f(t) g(\eee t)] - \eee \int f(t) g'(\eee t)\dd t$) gives, to leading-order, 
753: \begin{eqn} \lbl{himov} \fl
754: I_{m,n}\sim\frac{2\eee\alpha}{\w}\lft[1-\cos(2k)\rit]\sin\Phi\,\sech(\beta r)
755: - \frac{2\eee\alpha}{\w} \sin(2k) \cos \Phi \,\sech(\beta r),
756: \end{eqn}
757: where we have taken the constant of integration to be zero, and
758: $\om \sim \w$ to leading order.  Similarly, substituting for $Q_{m,n}$ in 
759: equations \eref{hkjq} and \eref{hkkq} and integrating, we find 
760: \begin{eqnarray} \fl
761: J_{m,n} & \sim & \frac{2\eee\alpha}{\w} \lft[ 1-\cos(k-lh) \rit] 
762: \sin \Phi\, \sech(\beta r) - 
763: \frac{2\eee\alpha}{\w}\sin(k-lh)\cos\Phi\,\sech(\beta r), \lbl{hjmov}\\[1ex] 
764: \fl  K_{m,n} & \sim & \frac{2\eee\alpha}{\w} \lft[ 1-\cos(k+lh) \rit] 
765: \sin \Phi\, \sech(\beta r)
766: - \frac{2\eee\alpha}{\w} \sin(k+lh) \cos \Phi \,\sech(\beta r). \lbl{hkmov}
767: \end{eqnarray}
768: Substituting these expressions into \eref{hentotlo}, we obtain 
769: \begin{eqnarray}
770: E_0 & \sim \sum_{m,n} & \frac{2\eee^2\alpha^2}{C_0} 
771: \cos^2 \Phi \ \sech^2(\beta r) \nonumber \\ & & + 
772: \frac{2L\eee^2\alpha^2}{\w^2}\sech^2(\beta r) \lft\{ \lft[ 
773: (1-\cos(2k))\sin\Phi - \sin(2k)\cos\Phi  \rit]^2 \right. \nonumber \\
774: & & + \lft[ (1-\cos(k-lh))\sin\Phi - \sin(k-lh)\cos\Phi \rit]^2 \nonumber \\ 
775: & & + \left. \lft[ (1-\cos(k+lh))\sin\Phi - \sin(k+lh)\cos\Phi \rit]^2 \rit\} .  
776: \lbl{hmvan} \end{eqnarray}
777: We replace the sum by an integral since the variables $Z=\eee (m-ut)$ 
778: and $W=\eee (hn-v t)$  vary slowly with $m$ and $n$.  To simplify the 
779: resulting integral, we approximate the terms $\cos^2\Phi$, $\sin^2\Phi$ 
780: and $\sin\Phi\cos\Phi$ by their average values of $\sfr{1}{2}$, $\sfr{1}{2}$ 
781: and $0$ respectively.  Hence \eref{hmvan} becomes 
782: \begin{eqn} \lbl{henmvint}
783: E_0 \sim \sum_{m,n} \frac{2\eee^2\alpha^2}{C_0} \sech^2(\beta r).
784: \end{eqn}
785: 
786: 
787: {}From the definition of $r$ (\ref{hexr}) we note the function 
788: $\sech(\beta r)$ is not in general radially symmetric in $m$, $n$.  Hence 
789: we work in $(\xi,\eta)$-space to facilitate evaluation of the double integral 
790: which approximates the double sum in (\ref{henmvint}).   Evaluating the 
791: Jacobian associated with the transformation from $(m,n)$ to $(\xi,\eta)$ 
792: coordinates, we find 
793: \begin{eqn}
794: E_0 \sim \frac{\alpha^2}{h^3C_0}\sqrt{4D_1D_2 - D_3^2} \int \!\!\! \int
795: \sech(\beta\sqrt{\xi^2+\eta^2}) \ \dd \xi \dd \eta. 
796: \lbl{hexemtri} \end{eqn}
797: Evaluating this and substituting for $\alpha$ and $\beta$ in terms of 
798: $D=3/2\w(k,l)$ and $B=3b\w(k,l)/2$ (see the appendix) we find 
799: \begin{eqn} \lbl{henar}
800: E_0 \sim \frac{4\pi\log 2(2\log 2 + 1)}
801: {3\, h^3 \, C_0 \, b \, \w^2(4\log 2 - 1)} \sqrt{4D_1D_2 - D_3^2}.
802: \end{eqn}
803: It is evident from \eref{henar} that the leading-order energy $E_0$ is 
804: independent of the breather amplitude, again confirming the existence 
805: of a minimum energy of moving breathers in the two-dimensional HETL 
806: with symmetric potential.  However, the threshold energy does depend 
807: upon the wavenumbers $k$ and $l$ hence moving breathers have a 
808: different threshold energy.   A plot of the expression \eref{henar} is 
809: shown in \Fref{henmov}.  We see that $\elo$, given by \eref{henar}, is 
810: strictly positive in the region of ellipticity $\mcal{D}$, and is maximised 
811: (attaining the same value) at each of the points corresponding to 
812: wavevectors $\{ \mb{k_1},\ldots, \mb{k_6} \}$, that is, at the points which
813: correspond to stationary breathers.  The threshold energy (\ref{henar}) 
814: decays to zero towards the boundary of the elliptic domain $\mcal{D}$ (see 
815: \Fref{econt}).   Hence, as for breathers in the square lattice, the energy 
816: threshold for moving breathers is {\em lower} than that for stationary 
817: breathers.  The threshold becomes arbitrarily small at the boundary of the 
818: domain of ellipticity.
819: 
820: 
821: \begin{figure}[htbp]
822: \begin{center}
823: \resizebox{3.5in}{!}{\includegraphics{henmov.ps}}
824: \caption{Plot of $E_0(k,l)$ for lattices with a symmetric potential.}
825: \lbl{henmov}
826: \end{center}
827: \end{figure}
828: 
829: 
830: %-------------------------------------------------------------------
831: \subsection{Lattices with an asymmetric potential} \lbl{hexasymm}
832: 
833: 
834: In this section, we consider the more general scenario where the potential 
835: $\Upsilon\!\ '(Q)$ is asymmetric.  In this case, the terms $a$ and $c$ in 
836: \eref{hqvexp} and \eref{heqq} are not both zero.   If \eref{hexnls} is to 
837: be reduced to an NLS equation in  $F$, both $G_0$ and $G_2$ must be 
838: found in terms of $F$.  Whilst $G_2$ is given by the simple algebraic 
839: equation \eref{hexg2}; in order to find $G_0$, the partial differential 
840: equation \eref{hexfo} must be solved.  We  assume that $G_0$ travels 
841: at the same velocity as $F$, that is, $G_0(X,Y,\tau,T) \equiv G_0(Z,W,T)$.  
842: Eliminating $G_{0\tau\tau}$,  \eref{hexfo} becomes
843: \begin{eqn} \lbl{hexg0}
844: (u^2 - 6)G_{0ZZ} + (v^2 - 2h^2)G_{0WW} + 2uvG_{0ZW} =
845: 6a|F|_{ZZ}^2 + 2ah^2|F|_{WW}^2.
846: \end{eqn}
847: For general $k$ and $l$, it is difficult to solve for $G_0$ explicitly.  
848: However, for any of the wavevectors $\{\mb{k_1},\ldots,\mb{k_6}\}$, 
849: the velocities $u$ and $v$ become zero, and so \eref{hexg0} becomes 
850: $\nabla^2 G_0 = -a\nabla^2|F|^2$, where the operator $\nabla^2$, 
851: defined by $\nabla_{(Z,W)}^2 \equiv \clyd_Z^2 + \clyd_W^2$, is 
852: equivalent to $\nabla_{(X,Y)}^2 \equiv \clyd_X^2 + \clyd_Y^2$. It 
853: follows that $G_0 = -a|F|^2$ for each of these wavevectors.
854: 
855: 
856: Equations \eref{hexdisp}--\eref{hexfo} are the same whichever 
857: of the wavevectors $\{\mb{k_1},\ldots,\mb{k_6}\}$ is used.  Thus, 
858: \eref{hexdisp} gives $\w=3$, and  \eref{hexg2} implies $G_2 = aF^2/3$.  
859: Substituting these expressions for $G_0$ and $G_2$ in  \eref{hexnls} 
860: gives the following NLS equation for asymmetric potentials for any 
861: of the wavevectors $\{\mb{k_1},\ldots,\mb{k_6}\}$
862: \begin{eqn} \lbl{hnlsasy}
863: 2\ii\w F_T + 3 \nabla^2 F + \w^2(3b-\mbox{$\frac{10}{3}$}a^2)|F|^2F=0,
864: \end{eqn}
865: The anomalous dispersion regime thus corresponds to $b>10a^2/9$.  
866: Soliton solutions are derived in the appendix;   \eref{hnlsasy} 
867: corresponds to $D = 1/2$ and $B =(9b-10a^2)/2$, and also 
868: $r^2 = X^2 + Y^2$.   Substituting the solution for $F$ into the lattice 
869: ansatz \eref{hexanz}, along with the known expressions for $G_0$ and 
870: $G_2$ gives the following second-order formula for stationary 
871: breathers in lattices with an asymmetric potential 
872: \begin{eqnarray} 
873: \fl Q_{m,n}(t) = 2\ \! \eee  \ \! \alpha\cos[km+lhn+(\w+\eee^2\lambda)t]\ 
874: \sech(\beta r) \nonumber \\ 
875: + \mbox{$\frac{2}{3}$}\ \! a \ \! \eee^2 \ \! \alpha^2 \sech^2(\beta r)\!\lft\{ 
876: \cos[2km+2lhn+(2\w+2\eee^2\lambda)t] - 3 
877: \rit\} + \mcal{O}(\eee^3), \lbl{hexqasym} 
878: \end{eqnarray} 
879: where $\alpha$ and $\beta$ are determined in the appendix, 
880: $r=\sqrt{X^2+Y^2}$, and $X=\eee m$ and $Y=\eee h n$ in terms of the 
881: original discrete variables $m$ and $n$.
882: 
883: 
884: %------------------
885: \subsubsection{Breather energy. } \lbl{hest}
886: 
887: 
888: We calculate the leading-order energy $\elo$ of stationary breathers in 
889: lattices with an asymmetric potential.  In this case, stationary breathers 
890: are given by \eref{hexqasym}, from which we take only the leading order term 
891: \begin{eqn} \lbl{hxa2}
892: Q_{m,n}(t) \sim 2 \eee \alpha \cos \Theta \ \sech(\beta r),
893: \end{eqn}
894: where $\Theta = km + lhn + \om t$ is the phase of the carrier wave 
895: with $(k,l)$ corresponding to one of $\{ \mb{k_1},\ldots, \mb{k_6}\}$, 
896: $\om = \w+\eee^2\lambda$ is the breather frequency, and $w=3$. 
897: 
898: 
899: The currents $I_{m,n}$, $J_{m,n}$ and $K_{m,n}$ are obtained 
900: by substituting \eref{hxa2} into equations \eref{hkiq}--\eref{hkkq} 
901: and integrating with respect to time, taking the constant of 
902: integration to be zero.   Thus, to leading order, we find
903: \begin{eqnarray}
904: I_{m,n} & \sim & \frac{3\eee\alpha}{\w} \sin \Theta\, \sech(\beta r)
905: - \frac{\sqrt{3}\eee\alpha}{\w} \cos \Theta \,\sech(\beta r),\lbl{heni} \\[1ex]
906: J_{m,n} & \sim & \frac{3\eee\alpha}{\w} \sin \Theta\, \sech(\beta r)
907: + \frac{\sqrt{3}\eee\alpha}{\w} \cos \Theta \,\sech(\beta r), \lbl{henj} \\[1ex]
908: K_{m,n} & \sim & \frac{3\eee\alpha}{\w} \sin \Theta\, \sech(\beta r)
909: + \frac{\sqrt{3}\eee\alpha}{\w} \cos \Theta \,\sech(\beta r).  \lbl{henk}
910: \end{eqnarray}
911: Substituting these expressions into  \eref{hentotlo} 
912: gives the leading-order energy 
913: \begin{eqnarray} \fl
914: E_0 \sim \sum_{m,n} \frac{2\eee^2\alpha^2}{C_0} 
915: \cos^2 \Theta \ \sech^2(\beta r) \nonumber \\
916: +  \frac{Lh^2}{2\w^2}\,\eee^2\alpha^2  \lft[  3h^2\sin^2\Theta + 
917: 2h\sin\Theta\cos\Theta + 3\cos^2\Theta  \rit]  \sech^2(\beta r)  . 
918: \lbl{henas}
919: \end{eqnarray}
920: We approximate the term in square brackets by taking the average
921: values of $\cos^2 \Theta$, $\sin^2\Theta$, $\sin\Theta\cos\Theta$ 
922: as \mbox{$\frac{1}{2}$}, \mbox{$\frac{1}{2}$} and $0$ respectively. 
923: 
924: 
925: We replace the double sum by an integral giving
926: \begin{eqn} \lbl{henint}
927: E_0 \sim  \frac{2\alpha^2}{hC_0} \int \!\!\! \int
928: \sech^2(\beta \sqrt{X^2+Y^2}) \ \dd X \dd Y.
929: \end{eqn}
930: This can be evaluated to 
931: \begin{eqn} \lbl{henex}
932: E_0 \sim \frac{4\pi\log 2}{h C_0} \frac{\alpha^2}{\beta^2}
933: = \frac{8 \pi \log 2(2\log 2+1)}{C_0h(9b-10a^2)(4\log 2-1)} . 
934: \end{eqn}
935: Again, this estimate for energy is independent of the breather
936: amplitude $\lambda$, demonstrating the energy threshold properties
937: of the two-dimensional HETL; namely, the activation energy
938: required to create a breather in the HETL is an $\mcal{O}(1)$
939: quantity, irrespective of its amplitude ($2\eee\alpha\ll 1$).
940: 
941: 
942: %==============================================
943: \section{Higher-order asymptotic analysis}  \lbl{hex5}
944: 
945: 
946: The cubic NLS equation exhibits blow-up, which cannot occur in a discrete 
947: system since energy is conserved, and even if all the energy were localised 
948: at a single site, the amplitude would still be finite.   The two-dimensional 
949: cubic NLS equation description of the breather envelope is lacking.   
950: Higher-order dispersive and nonlinear effects play an important role in the 
951: dynamics of such discrete systems and therefore must be incorporated.  In 
952: this section we extend our analysis of the lattice equations \eref{heqq} to 
953: fifth-order, and derive a generalised NLS equation which includes higher-order 
954: dispersive and nonlinear terms.  It then remains to determine whether this 
955: generalised NLS equation supports stable soliton solutions.  
956: 
957: 
958: Due to the complexity of a fifth-order analysis of  a general asymmetric 
959: potential, we consider only lattices with a symmetric potential, that is, 
960: those for which $a=c=0$ in \eref{heqq}. Since no second or fourth harmonic
961: terms are generated by the nonlinearity, we use a much simpler ansatz, namely 
962: \begin{eqn} \lbl{h5anzgen}
963: Q_{m,n}(t) = \eee \ee^{\ii \psi} F(X,Y,\tau,T)
964: + \eee^3 \ee^{3\ii \psi}H_3(X,Y,\tau,T) + \cdots + \mathrm{c.c.,}
965: \end{eqn}
966: where the phase $\psi = km+lhn+\w t$.  In this case, in addition to the 
967: equations \eref{hexdisp2}--\eref{velsuv}, \eref{hexrnls} and \eref{hexh3}, 
968: we also have \\ 
969: \\
970: $\mcal{O}(\eee^5 \ee^{\ii\psi})$:
971: \begin{eqnarray} \fl 
972: F_{TT} =&& \hspace*{-12mm}
973: \mbox{$\frac{1}{6}$} \lft[ 8\cos(2k) + \cos k \cos(lh) \rit] F_{XXXX} 
974: + h^2\cos k \cos(lh)F_{XXYY}  \nonumber \\ &&\hspace*{-16mm}
975: + \mbox{$\frac{1}{6}$} h^4 \cos k \cos(lh)F_{YYYY} 
976: - \mbox{$\frac{2}{3}$} h \sin k \sin(lh) F_{XXXY} 
977: - \mbox{$\frac{2}{3}$} h^3 \sin k \sin(lh)F_{XYYY} 
978: \nonumber\\ &&\hspace*{-16mm}
979: + 6b \lft[ 2\cos(2k) + \cos k \cos(lh) \rit](|F|^2F)_{XX}
980: + 6bh^2\cos k\cos(lh)(|F|^2F)_{YY} \nonumber \\ && \hspace*{-16mm}
981: -12b\sin k \sin(lh) \lft[  2hFF_X\tbar{F}_Y  + 2hFF_Y \tbar{F}_X
982: + 2h\tbar{F}F_XF_Y + hF^2\tbar{F}_{XY} + 2hF\tbar{F}F_{XY} \rit] 
983: \nonumber \\  && \hspace*{-16mm} 
984: - 3b\w^2\tbar{F}^2H_3 - 10\w^2 d |F|^4F . 
985: \lbl{hfogen} \end{eqnarray}
986: We simplify this by restricting attention to stationary breathers.
987: Accordingly, only one extra timescale, $T=\eee^2 t$, is required, and we 
988: fix the wavenumbers $k$ and $l$ to correspond to one of the wavevectors 
989: $\{ \mb{k_1},\ldots, \mb{k_6} \}$.  Hence we obtain the equations: \\
990: \\
991: $\mcal{O}(\eee^3 \ee^{\ii\psi})$: \\
992: \begin{eqn} \lbl{h5nls}
993: 2\ii \w F_T + 3\nabla^2 F + 3b\w^2 |F|^2F = 0,
994: \end{eqn}
995: $\mcal{O}(\eee^5 \ee^{\ii\psi})$: \\
996: \begin{eqn} \lbl{h550}
997: F_{TT} = -\mbox{$\frac{3}{4}$} \nabla^4F - 9b\nabla^2(|F|^2F) -
998: 10\w^2d|F|^4F , 
999: \end{eqn} 
1000: in addition to $\w=3$ (from the $\mcal{O}(\eee \ee^{\ii\psi})$
1001: equation) and $H_3=0$ (from $\mcal{O}(\eee^3 \ee^{3\ii\psi})$). 
1002: A consequence of the hexagonal symmetry of the HETL is that all 
1003: differentials on the right-hand side of \eref{h550} are isotropic.  
1004: 
1005: 
1006: In order to obtain a generalised NLS equation, we combine the
1007: higher-order equation \eref{h550} with the cubic two-dimensional
1008: NLS equation \eref{h5nls}.  The $F_{TT}$ term on the
1009: left-hand side of \eref{h550} is eliminated  by
1010: differentiating \eref{h5nls} with respect to $T$ and substituting
1011: for $F_{TT}$ into \eref{h550}.  The resulting expression is 
1012: \begin{eqnarray}
1013: 6 \ii F_T + 3\nabla^2 F + 27b|F|^2F + \frac{\eee^2}{2} \nabla^4 F
1014: + \frac{9\eee^2}{4}(40d-27b^2)|F|^4F & \nonumber \\ \qquad \qquad 
1015: + \frac{27b\eee^2}{4} \nabla^2(|F|^2F) - \frac{9b\eee^2}{2}|F|^2\nabla^2F
1016: - \frac{9b\eee^2}{4} F^2\nabla^2\tbar{F} = 0 , & \lbl{hexgennls}
1017: \end{eqnarray}
1018: which is isotropic. 
1019: 
1020: 
1021: To the best of our knowledge, this perturbed form of the NLS equation 
1022: has not been studied in the literature before; although it is similar to 
1023: and slightly simpler than the corresponding equation derived for the 
1024: square lattice \cite{buts06}.  It is also similar to the perturbed NLS 
1025: equation considered by Davydova \etal \citep{dav03}, namely 
1026: \begin{equation}
1027: \ii F_T + D \nabla^2 F + B |F|^2 F + P \nabla^4 F + K |F|^4 F =0 , 
1028: \lbl{davydova} 
1029: \end{equation} 
1030: which is known to have stable soliton solutions.  The anomalous 
1031: dispersion case is $BD>0$, which in our equation (\ref{hexgennls}) 
1032: corresponds to $b>0$;   Davydova's criteria for soliton existence is 
1033: $PK>0$ which implies $40d > 27b^2$.  Hence, it is in this parameter 
1034: regime that we seek breathers solutions of the HETL.  The numerical 
1035: results presented in  \Sref{hexnumerics} show that long-lived breather 
1036: solutions are supported by the two-dimensional hexagonal FPU lattice, 
1037: suggesting that the additional perturbing terms in \eref{hexgennls}
1038: do not destabilise the Townes soliton.
1039: In fact, numerical simulations (presented in \cite{thesis} but not here) of the 
1040: case $40d<27b^2$ show the breather mode to be long-lived, suggesting 
1041: that the terms on the second line of (\ref{hexgennls}) are stabilising. 
1042: 
1043: 
1044: We have been unable to find a variational formulation of
1045: \eref{hexgennls}, and are therefore unable to use the methods of 
1046: Davydova {\em et al.}\ \cite{dav03} and Kuznetsov {\em et al.}\ 
1047: \citep{kuz}.  Alternative possible methods include the modulation 
1048: theory of Fibich and Papanicolaou \citep{fib98,fib99}.
1049: 
1050: 
1051: %=============================================
1052: \section{Numerical results}  \lbl{hexnumerics}
1053: 
1054: \subsection{Preliminaries} \lbl{hnumpre}
1055: 
1056: 
1057: Using a fourth-order Runge-Kutta scheme, we solve the equations 
1058: \begin{eqnarray}
1059: \frac{\dd Q_{m,n}}{\dd t} & = & R_{m,n}, \nonumber \\[1ex]
1060: \frac{\dd R_{m,n}}{\dd t} & = & (\delta^2_I + \delta^2_J +
1061: \delta^2_K)
1062: \lft[ Q_{m,n} + a Q_{m,n}^2 + b Q_{m,n}^3 + c Q_{m,n}^4 + d
1063: Q_{m,n}^5 \rit], \lbl{heqqfo}
1064: \end{eqnarray}
1065: numerically.  Introducing the variable $R_{m,n}$ converts the 
1066: system of second-order ordinary differential equations \eref{heqq} 
1067: to an equivalent system of first-order differential equations.  
1068: 
1069: 
1070: We present the results of simulations for a range of parameter
1071: values.  From \eref{velsuv}, the velocity of the envelope $(u,v)$ 
1072: depends upon the wavevector $\mb{k}=[k,l]^T$, and hence we 
1073: obtain moving breathers by choosing $(k,l) \in \mcal{D}$.  
1074: In \Fref{ellcontnum}, we show the points in  $\mcal{D}_1 \subset  
1075: \mcal{D}$ for which we solve the lattice equations \eref{heqqfo} 
1076: numerically.  These points correspond to the wavevectors $\mb{k_1} 
1077: = [\pi/3,\pi/h]^T$, $\mb{k_a} = [1.4,\pi/h]^T$, $\mb{k_b} = 
1078: [0.79,1.7324]^T$ and $\mb{k_c} = [0.8,1.9987]^T$.  
1079: Selecting $(k,l)$ too near to the boundary of $\mcal{D}$ results 
1080: in a sharply elongated breather, which are difficult to simulate as 
1081: they require a large domain.  Wavevectors close to any of 
1082: $\{ \mb{k_1},\ldots, \mb{k_6} \}$, lead to breather modes with small 
1083: speeds, a check of breather velocity would then require a long-time 
1084: simulation.  In practice we have chosen wavenumbers which do not 
1085: lead to a severely elongated breather envelope, and yet have velocities 
1086: which result in observable displacements over reasonable times.   We 
1087: present simulations of stationary breathers in systems with asymmetric 
1088: potentials ($a,c \neq 0$) in \Sref{hnra} and, in Sections 
1089: \ref{hnstsy}--\ref{hpsi130}, simulations of stationary and moving breathers 
1090: in lattices with symmetric potentials (that is, $a=c=0$ in \eref{heqqfo}). 
1091: 
1092: 
1093: \begin{figure}[htbp]
1094: \begin{center}
1095: \resizebox{4in}{!}{\includegraphics{ellcontnum.eps}}
1096: \put(-136,114){\scriptsize $\mb{k_1}$}
1097: \put(-89,114){\scriptsize $\mb{k_a}$}
1098: \put(-172,91){\scriptsize $\mb{k_b}$}
1099: \put(-170,140){\scriptsize $\mb{k_c}$}
1100: \put(-87,164){\fns $\mcal{D}_1$}
1101: \caption{Wavevectors in $\mcal{D}_1 \subset \mcal{D}$ for 
1102: which breathers are simulated.}
1103: \lbl{ellcontnum}
1104: \end{center}
1105: \end{figure}
1106: 
1107: 
1108: %-----------------------------------------------------------------
1109: \subsection{Initial data and boundary conditions} \lbl{hbcics}
1110: 
1111: 
1112: We generate initial data by using the analytic expressions for
1113: breather solutions derived in \Sref{haa}. The formulae for $Q_{m,n}$ 
1114: and $R_{m,n}$ are found in terms of the original discrete variables 
1115: $m$ and $n$, and then shifted horizontally and vertically so that initially 
1116: the breather lies at the centre of the lattice.   We impose periodic 
1117: boundary conditions for the lattice in both horizontal and vertical 
1118: directions, converting the two-dimensional arrangement illustrated 
1119: in \Fref{hexetl} into a two-torus.  In long-time simulations, moving 
1120: breathers which approach an edge of the lattice reappear from the 
1121: opposite edge.  We select the site $(1,1)$ to lie at the bottom left-hand 
1122: corner of the arrangement as illustrated in \Fref{hexbcs}.  Sites along 
1123: the boundaries and corners are missing between one and four neighbours.  
1124: We introduce fictitious sites along the boundaries and corners 
1125: where necessary to effect periodic boundary conditions.  
1126: 
1127: 
1128: An illustration of a small finite lattice is shown in \Fref{hexbcs}.  
1129: The dots represent capacitors located at lattice sites, and the 
1130: lines represent some of the inter-connecting inductors.  From 
1131: equations \eref{heqqfo},  the charge $Q_{m,n}$ stored on 
1132: each capacitor depends upon the charge stored on the capacitors 
1133: located at its six neighbouring sites,  two in each of the directions 
1134: $\mb{e_i}$, $\mb{e_j}$ and $\mb{e_k}$.  
1135: For the sake of clarity, the inductors connecting the capacitors at
1136: the center of each hexagon to its six nearest neighbours are not
1137: shown (see \Fref{hexetl}).  It is not necessary that the lattice
1138: should be ``square,'' meaning that the lattice could comprise $M
1139: \times N$ lattice sites, with $M \neq N$.  However, typically, we
1140: consider $M=N$ as in \Fref{hexbcs}, since the numerical routines
1141: are simpler to encode when the lattice \itc{is} square, and in  
1142: the examples below, we consider lattices with $N \leq 50$.  
1143: 
1144: 
1145: \addtolength{\abovecaptionskip}{10pt}
1146: \begin{figure}[htbp]
1147: \begin{center}
1148: \vspace{1.0cm}
1149: \resizebox{2.5in}{!}{\includegraphics{hexbcs.eps}}
1150: \put(-204,-10){${Q_{1,1}}$}
1151: \put(-10,-10){${Q_{2N-1,1}}$}
1152: \put(-173,153){${Q_{2,N}}$}
1153: \put(-18,153){${Q_{2N,N}}$}
1154: \caption{Periodic boundary conditions for the two-dimensional HETL.}
1155: \lbl{hexbcs}
1156: \end{center}
1157: \end{figure}
1158: \addtolength{\abovecaptionskip}{-10pt}
1159: 
1160: 
1161: %--------------------------------------------------------------------
1162: \subsection{Numerical computation of breather energy} \lbl{hnumen}
1163: 
1164: 
1165: Since the HETL is lossless the total energy is conserved and 
1166: can be used as a check of the accuracy of our numerical scheme.   
1167: We compute the leading-order energy by expressing the summand 
1168: $e_{m,n}^{(0)}$ \eref{hentotlo} in terms of the output variables of 
1169: the numerical routine, namely $Q_{m,n}$ and $R_{m,n}$.  The first 
1170: term of $e_{m,n}^{(0)}$ is simply dependent on $Q_{m,n}$.  It
1171: remains to find the currents $I_{m,n}$, $J_{m,n}$ and $K_{m,n}$ in
1172: terms of $Q_{m,n}$ and $R_{m,n}$.  The details of this calculation 
1173: depend upon whether the interaction potential is symmetric or asymmetric.
1174: As a check of the numerical scheme we verify that the sum is 
1175: conserved and that it agrees with the asymptotic estimate \eref{henar}.
1176: 
1177: 
1178: %-------------------------------------------------------------
1179: \subsubsection{Lattices with a symmetric potential. }
1180: 
1181: 
1182: First, we find the current $I_{m,n}$ in terms of $Q_{m,n}$ and
1183: $R_{m,n}$.  Differentiating the leading-order expression for the breather 
1184: given by \eref{hxs2}, we have (retaining leading-order terms only)
1185: \begin{eqn} \lbl{hnem}
1186: \dot{Q}_{m,n} = R_{m,n} \sim -2\eee\alpha \w \sin \Phi \ \sech (\beta r),
1187: \end{eqn}
1188: where $\alpha$, $\beta$ and $r^2$ are as defined in \Sref{hexsym}.
1189: Comparing the analytic expression \eref{himov} for the current
1190: $I_{m,n}$ with the expressions for $Q_{m,n}$ \eref{hxs2} and
1191: $R_{m,n}$ \eref{hnem}, we find
1192: \begin{eqn} \lbl{hniqr}
1193: I_{m,n} = \frac{\cos(2k)-1}{\w^2} R_{m,n} - \frac{\sin(2k)}{\w}Q_{m,n}.
1194: \end{eqn}
1195: Similarly, comparing the expressions \eref{hjmov} and \eref{hkmov}
1196: for the currents $J_{m,n}$ and $K_{m,n}$ respectively with
1197: equations \eref{hxs2} and \eref{hnem}, we find 
1198: \begin{eqnarray}
1199: J_{m,n} & = & \frac{\cos(k-lh)-1}{\w^2} R_{m,n} -
1200: \frac{\sin(k-lh)}{\w}Q_{m,n},
1201: \lbl{hnjqr} \\[1ex]
1202: K_{m,n} & = & \frac{\cos(k+lh)-1}{\w^2} R_{m,n} -
1203: \frac{\sin(k+lh)}{\w}Q_{m,n}. \lbl{hnkqr}
1204: \end{eqnarray}
1205: Substituting these expressions into \eref{hentotlo}, gives an expression 
1206: for $E_0$ which does not simplify, so we do not reproduce it here.  
1207: 
1208: 
1209: %------------------- 
1210: \subsubsection{Lattices with an asymmetric potential. }
1211: 
1212: 
1213: Differentiating the leading-order expression \eref{hxa2} for the
1214: charge $Q_{m,n}$ in lattices with an asymmetric potential gives
1215: \begin{eqn} \lbl{hxa2dot}
1216: \dot{Q}_{m,n} = R_{m,n} \sim 
1217: -2\eee\alpha \w \sin \Theta \ \sech (\beta r) , 
1218: \end{eqn}
1219: where $\Theta=\pi m/3 +\pi n + \om t$ with $\om=3$.  Using \eref{hxa2} 
1220: and \eref{hxa2dot}, the current $I_{m,n}$  given by \eref{heni} can be 
1221: expressed in terms of $Q_{m,n}$ and $R_{m,n}$ as
1222: \begin{eqn} \lbl{hiqqd}
1223: I_{m,n} \sim -\frac{\sqrt{3}}{2\w} \lft[ \frac{\sqrt{3}R_{m,n}}{\w} + 
1224: Q_{m,n} \rit]\!.
1225: \end{eqn}
1226: Similarly, from  \eref{henj}--\eref{henk}, we have
1227: \begin{eqn} \lbl{hjkqqd}
1228: J_{m,n}  = K_{m,n} \sim -\frac{\sqrt{3}}{2\w} 
1229: \lft[ \frac{\sqrt{3}R_{m,n}}{\w} - Q_{m,n} \rit]\!.
1230: \end{eqn}
1231: These are equivalent to substituting $k=\pi/3$ and $l=\pi/h$ into 
1232: (\ref{hniqr})--(\ref{hnkqr}).  Inserting the expressions for $I_{m,n}$, 
1233: $J_{m,n}$ and $K_{m,n}$ (\eref{hiqqd}--\eref{hjkqqd}) into \eref{hentotlo} yields 
1234: \begin{eqn} \lbl{hentotqqd}
1235: E_0 = \sum_{m,n} \frac{1}{72 C_0} \lft[ 
1236: 45Q_{m,n}^2 + 3R_{m,n}^2 - 2\sqrt{3}Q_{m,n}R_{m,n} \rit]\! . 
1237: \end{eqn}
1238: 
1239: 
1240: %------------------------
1241: \subsubsection{Effective breather width. }
1242: 
1243: 
1244: Numerical computation of the energy as described above allows us to 
1245: check that the total lattice energy is conserved, but does not indicate 
1246: whether a breather changes shape over time.  To remedy this, we define 
1247: breather widths in the $m$ and $n$ directions, the sum of which we 
1248: denote $\mcal{W}_{\mathrm{br}}$, where
1249: \begin{eqn} \lbl{hbrwdth}
1250: \mcal{W}_{\mathrm{br}}^2 = \frac{r_{20}}{\elo} + \frac{r_{02}}{\elo}
1251: - \lft( \frac{r_{10}}{\elo} \rit)^2 -  \lft( \frac{r_{01}}{\elo} \rit)^2,
1252: \end{eqn}
1253: and $r_{10}$, $r_{01}$, $r_{20}$ and $r_{02}$ are defined by
1254: \addtolength{\arraycolsep}{0.5cm}
1255: \begin{displaymath}
1256: \begin{array}{c c}
1257: r_{10}=\sum_{m,n} m e_{m,n},\quad&r_{20}=\sum_{m,n}m^2 e_{m,n},\\[8pt]
1258: r_{01}=\sum_{m,n} hn e_{m,n},&r_{02} = \sum_{m,n} h^2n^2 e_{m,n}. 
1259: \end{array}
1260: \end{displaymath}
1261: \addtolength{\arraycolsep}{-0.5cm}
1262: The variation in $\wbr$ over time gives a measure of the distortion 
1263: suffered by a breather.
1264: 
1265: 
1266: %------------------------------------------------------------------
1267:  \subsection{Stationary breather in a lattice with symmetric potential}
1268: \lbl{hnstsy}
1269: 
1270: 
1271: We investigate stationary breather solutions in the anomalous dispersion
1272: regime which corresponds to $b>0$ (see \eref{hexgennls}).  We set 
1273: $k=\pi/3$ and $l=\pi/h$, corresponding to $\mb{k_1}$ in \Fref{ellcontnum}, 
1274: so that the velocities $u$ and $v$ are zero.   Davydova's result 
1275: implies that we expect to find stable soliton solutions when $PK>0$.  
1276: {}From \eref{davydova}, $P=\eee^2/2$, hence this inequality implies 
1277: $40d > 27b^2$.  Hence we choose $b=d=1$ as well as $N=30$, $\eee=0.2$ 
1278: and $\lambda=1$.  Using the technique outlined in the appendix we 
1279: calculate $\alpha=1.0212$, $\beta=1.8670$.  The breather frequency is 
1280: $\w+\eee^2\lambda=3.040$, and therefore the period of oscillation is $T=2.0668$. 
1281: 
1282: 
1283: As is the case for all our simulations, the initial profile of the breather 
1284: is located at the centre of the lattice, as illustrated in  \Fref{simh1}\,(a). 
1285: At $t=0$, the breather energy is $\elo=0.7606$, and $\wbr=3.74$.
1286: The asymptotic estimate for $E_0$ given by \eref{henar} is 0.7523, 
1287: which is only 1\% different from the numerically obtained value.
1288: In \Fref{simh1}\,(c) we show the breather after 30 oscillations.  
1289: Plots of the local energy $e_{m,n}$ at $t=0$
1290: and $t=30T$ are presented in Figures \ref{simh1}\,(b) and
1291: \ref{simh1}\,(d).  After thirty oscillations the breather has shed 
1292: a small amount of energy, which is manifested as small amplitude 
1293: radiation throughout the lattice.  Accordingly, the breather 
1294: appears a little distorted in shape compared to its initial profile.  
1295: In particular, at $t=30T$, we find $\wbr=4.21$. The energy of the 
1296: breather at $t=30T$ is 0.7087, giving $\Delta E_0/E_0=-0.0682$, 
1297: 
1298: 
1299: \begin{figure}[htbp]
1300: \centering % PICTURE 1
1301: \subfigure[Profile at $t=0$.]{
1302: \includegraphics[width=.45\textwidth]{hs1a.eps}}
1303: \hspace{.3in} % PICTURE 2
1304: \subfigure[Plot of $e_{m,n}$, $E_0=0.7606$.]{
1305: \includegraphics[width=.45\textwidth]{hs1aen.eps}}\\
1306: \vspace{.3in} % PICTURE 3
1307: \subfigure[Profile at $t=30T=62.82$.]{
1308: \includegraphics[width=.45\textwidth]{hs1b.eps}}
1309: \hspace{.3in} % PICTURE 4
1310: \subfigure[Plot of $e_{m,n}$, $E_0=0.7087$.]{
1311: \includegraphics[width=.45\textwidth]{hs1ben.eps}}
1312: \caption{Stationary breather in a lattice with symmetric potential, 
1313: see Section \ref{hnstsy} for details.}
1314: \label{simh1}
1315: \end{figure}
1316: 
1317: 
1318: %---------------------------------------------------------------------
1319: \subsection{Breather moving along a lattice direction ($\Psi=0\degrees$)}
1320: \lbl{hpsi0}
1321: 
1322: 
1323: In \Fref{simh2}, we show a simulation of a breather moving along a
1324: lattice direction parallel to the $m$-axis, that is, $\Psi=0\degrees$.  We 
1325: have chosen $\mb{k} = \mb{k_a} = [1.4,\pi/h]^T$ so that $u=0.4445$ 
1326: and $v=0$ and thus $\Psi=\tan^{-1}(v/u) = 0\degrees$.  The breather 
1327: frequency is $\w=2.9265$, and hence the period $T=2.1470$. Following 
1328: the calculation outlined in the appendix, the amplitude and width 
1329: parameters are $\alpha=1.0339$ and $\beta=1.8440$; the remaining 
1330: parameters being $b=1$, $d=1$, $N=30$, $\eee=0.1$ and $\lambda=1$.
1331: 
1332: 
1333: The initial profile of the breather is shown in \Fref{simh2}\,(a), and at 
1334: this time, the calculated energy is $E_0=0.5537$, whilst the asymptotic 
1335: estimate \eref{henar} is $E_0=0.5522$.  It may be observed that the 
1336: breather is not radially symmetric. In fact, it is slightly elongated in the 
1337: direction parallel to the $m$-axis, that is, parallel to the direction of 
1338: motion. This is because the point corresponding to the wavevector 
1339: $\mb{k_a}$ is near the boundary of the region of ellipticity $\mcal{D}_1$ 
1340: as illustrated in \Fref{econt}. 
1341: 
1342: 
1343: \Fref{simh2}\,(b) shows the breather at the later time of  $t=21.58T=46.3513$, 
1344: at which time we find $\elo=0.5562$.  By this time, the breather has reached 
1345: the right-hand edge of the lattice and, owing to periodic boundary conditions, 
1346: it reemerges from the left-hand side as shown in \Fref{simh2}\,(c). The 
1347: breather remains localised, without spreading, though it leaves behind a 
1348: small amount of energy in its path (visible in Figures \ref{simh2}\,(b) and 
1349: \ref{simh2}\,(c)).   The computed energy changes much less than for the 
1350: previous simulation (here, $\Delta E_0/E_0=0.0287$). 
1351: 
1352: 
1353: The velocity of the breather can be measured from a plot of the energy 
1354: $e_{m,n}$, which is shown in \Fref{simh2}\,(d), viewed from directly 
1355: above the plane of the lattice.  From this plot, we note that  the 
1356: breather has travelled 42 units at an average speed of 0.42 units per 
1357: second, which is 5.5\% lower than our predicted speed of 0.4445 units 
1358: per second.  This is consistent with a small amount of energy being 
1359: shed as the system transforms from our approximated initial conditions 
1360: into the precise shape of the breather.
1361: 
1362: 
1363: \begin{figure}[t]
1364: \centering % PICTURE 1
1365: \subfigure[Profile at $t=0$, $E_0=0.5537$.]{
1366: \includegraphics[width=.45\textwidth]{hs2a.eps}}
1367: \hspace{.3in} % PICTURE 2
1368: \subfigure[Profile at $21.58T$, $E_0=0.5562$.]{
1369: \includegraphics[width=.45\textwidth]{hs2b.eps}}\\
1370: \vspace{.3in} % PICTURE 3
1371: \subfigure[Profile at $t=46.59T$, $E_0=0.5696$.]{
1372: \includegraphics[width=.45\textwidth]{hs2c.eps}}
1373: \hspace{.3in} % PICTURE 4
1374: \subfigure[Plot of $e_{m,n}$.]{
1375: \includegraphics[width=.45\textwidth]{hs2cen.eps}}
1376: \caption{Breather moving along a lattice direction, $\Psi=0\degrees$, 
1377: see Section \ref{hpsi0} for details.}
1378: \label{simh2}
1379: \end{figure}
1380: 
1381: 
1382: %--------------------------------------------------------------------
1383: \subsection{Breather moving at $\Psi=210\degrees$} \lbl{hpsi210}
1384: 
1385: 
1386: We now show that it is possible to simulate breathers moving in
1387: directions other than a lattice direction (that is, $\Psi \neq 0\degrees$).  
1388: We set $k=0.79$ and $l=1.7324$, which corresponds to $\mb{k_b}$ in
1389: \Fref{ellcontnum}.  It may be verified that $u=-0.1999$ and
1390: $v=-0.1154$ units per second, leading to $\Psi = 210\degrees$ as 
1391: required.  Although not a lattice vector, this direction is an axis of 
1392: symmetry of the lattice. For the wavevector $\mb{k_b}$, we have 
1393: $\w=2.9675$ and
1394: hence $T=2.1174$,  the remaining parameters being $b=1$, $d=1$,
1395: $N=30$, $\eee=0.1$ and $\lambda=1$.  The variational parameters
1396: $\alpha$ and $\beta$ are 1.0267 and 1.8568 respectively.  The
1397: breather is shown at times $t=25$, 50, 75 and 100 seconds in
1398: \Fref{simh6}.  Clearly, the breather does not deform significantly
1399: as it travels, nor does it radiate much energy.  The initial energy is 
1400: computed to be $E_0=0.6306$, and at $t=100$, the energy is 0.6275, 
1401: a loss of 0.5\%.   The asymptotic estimate for the energy is 
1402: $E_0=0.6184$ --- 2\% different from the numerically computed value.
1403: The motion of the breather is charted in \Tref{tsimh6}, the final 
1404: measured values for the velocities $u$ and $v$ give an average 
1405: speed of 0.2152 units per second, 6.8\% below the predicted 
1406: speed of 0.2308 units per second. The angle of travel is 
1407: almost identical to the expected value of $\Psi=210\degrees$.
1408: 
1409: 
1410: \begin{figure}[t]
1411: \centering % PICTURE 1
1412: \subfigure[Profile at $t=25$, $\elo = 0.6295$.]{
1413: \includegraphics[width=.45\textwidth]{hs6a.eps}}
1414: \hspace{.3in} % PICTURE 2
1415: \subfigure[Profile at $t=50$, $\elo = 0.6292$.]{
1416: \includegraphics[width=.45\textwidth]{hs6b.eps}}\\
1417: \vspace{.3in} % PICTURE 3
1418: \subfigure[Profile at $t=75$, $\elo = 0.6235$.]{
1419: \includegraphics[width=.45\textwidth]{hs6c.eps}}
1420: \hspace{.3in} % PICTURE 4
1421: \subfigure[Profile at $t=100$, $\elo = 0.6275$.]{
1422: \includegraphics[width=.45\textwidth]{hs6d.eps}}
1423: \caption{Breather moving at $\Psi=210\degrees$, 
1424: see Section \ref{hpsi210} for details.}
1425: \label{simh6}
1426: \end{figure}
1427: 
1428: 
1429: \setlength{\arrayrulewidth}{1pt}
1430: \begin{table}
1431: \begin{center}
1432: \begin{tabular}[htbp]{| c | c | c | c | c | c | c |} \hline 
1433: {\ssz{Time}} & {\ssz{Horizontal}} & {\ssz{Vertical}} & 
1434: {\ssz{Average horizontal}} & {\ssz{Average vertical}} & {} & {} \\[-8pt]  
1435: {\ssz{(s)}} & {\ssz{displacement}} & {\ssz{displacement}} & 
1436: {\ssz{velocity (units s$^{-1}$)}} & {\ssz{velocity (units s$^{-1}$)}} & 
1437: {\ssz{$\tan\Psi$}} & {\ssz{$\Psi$}} \\  \hline 
1438: {\ssz{$25$}} & {\ssz{$-3.5$}}  & {\ssz{$-2$}}  & {\ssz{$-0.14$}}  & {\ssz{$-0.08$}}  & {\ssz{$0.5714$}} & {\ssz{$209.74\degrees$}} \\ \hline 
1439: {\ssz{$50$}} & {\ssz{$-8$}}  & {\ssz{$-4.5$}}  & {\ssz{$-0.16$}}  & {\ssz{$-0.09$}}  & {\ssz{$0.5625$}} & {\ssz{$209.36\degrees$}} \\ \hline 
1440: {\ssz{$75$}} & {\ssz{$-13$}}  & {\ssz{$-7.5$}}  & {\ssz{$-0.1733$}}  & {\ssz{$-0.1$}}  & {\ssz{$0.5770$}} & {\ssz{$209.99\degrees$}} \\ \hline 
1441: {\ssz{$100$}} & {\ssz{$-18.5$}}  & {\ssz{$-11$}}  & {\ssz{$-0.185$}}  & {\ssz{$-0.11$}}  & {\ssz{$0.5946$}} & {\ssz{$210.74\degrees$}} \\ \hline
1442: \end{tabular}
1443: \end{center}
1444: \caption{Summary of breather motion ($\Psi=210\degrees$).}
1445: \label{tsimh6}
1446: \end{table}
1447: 
1448: 
1449: %-------------------------------------------------------------------
1450: \subsection{Breather moving at $\Psi=130\degrees$} \lbl{hpsi130}
1451: 
1452: 
1453: We have presented simulations of breathers which move along axes of 
1454: symmetry of the lattice.  In \Sref{hexdell}, from the results of our asymptotic 
1455: analysis, we found that breather solutions could be constructed for any 
1456: direction of travel.  In this section, we test the mobility of breathers along 
1457: directions which do not correspond to axes of symmetry of the lattice.
1458: We have successfully propagated breathers in a range of directions, 
1459: and here we show only one such breather, moving at an angle
1460: $\Psi=130\degrees$.  We set $k=0.8$ and $l=1.9987$, from which we 
1461: find  $u=-0.2160$, $v=0.2575$, $\Psi = 130\degrees$, $\w=2.9502$ 
1462: and the period is $T=2.1298$.  The remaining parameters are $b=1$, 
1463: $d=1$, $N=30$, $\eee=0.1$, $\lambda=1$, $\alpha=1.0297$ and 
1464: $\beta=1.8514$.  From Figure \ref{simh7}, 
1465: Even at 120 seconds, the breather remains localised without suffering
1466: appreciable degradation nor is much radiation left behind in its wake.   
1467: Initially the energy $E_0$ is computed as 0.5977, only 0.2\% different 
1468: from the asymptotic estimate of $E_0=0.5965$.   
1469: After 120 seconds, $E_0$ is 0.5897, a change of 1.3\%.  
1470: To find the velocity of the breather, we have recorded its motion
1471: and presented relevant data in \Tref{tsimh7}.  From (\ref{velsuv}), 
1472: the average speed is predicted to be 0.3254 units per second; 
1473: whilst our numerical simulation shows an average speed of 0.3361 
1474: units per second,  3.2\% lower than the expected speed.  The 
1475: direction of motion of the breather is predicted accurately.
1476: 
1477: 
1478: \begin{figure}[t]
1479: \centering % PICTURE 1
1480: \subfigure[Profile at $t=30$, $\elo = 0.5961$.]{
1481: \includegraphics[width=.45\textwidth]{hs7a.eps}}
1482: \hspace{.3in} % PICTURE 2
1483: \subfigure[Profile at $t=60$, $\elo = 0.5966$.]{
1484: \includegraphics[width=.45\textwidth]{hs7b.eps}}\\
1485: \vspace{.3in} % PICTURE 3
1486: \subfigure[Profile at $t=90$, $\elo = 0.5939$.]{
1487: \includegraphics[width=.45\textwidth]{hs7c.eps}}
1488: \hspace{.3in} % PICTURE 4
1489: \subfigure[Profile at $t=120$, $\elo = 0.5897$.]{
1490: \includegraphics[width=.45\textwidth]{hs7d.eps}}
1491: \caption{Breather moving at $\Psi=130\degrees$, see Section 
1492: \ref{hpsi130} for details.}
1493: \label{simh7}
1494: \end{figure}
1495: 
1496: 
1497: \begin{table}
1498: \begin{center}
1499: \begin{tabular}[htbp]{| c | c | c | c | c | c | c |} \hline 
1500: {\ssz{Time}} & {\ssz{Horizontal}} & {\ssz{Vertical}} & 
1501: {\ssz{Average horizontal}} & {\ssz{Average vertical}} & {} & {} \\[-8pt] 
1502: {\ssz{(s)}} & {\ssz{displacement}} & {\ssz{displacement}} & 
1503: {\ssz{velocity (units s$^{-1}$)}} & {\ssz{velocity (units s$^{-1}$)}} & 
1504: {\ssz{$\tan\Psi$}} & {\ssz{$\Psi$}} \\ \hline 
1505: {\ssz{$30$}} & {\ssz{$-6$}}  & {\ssz{$7$}}  & {\ssz{$-0.2$}}  & {\ssz{$0.2333$}}  & {\ssz{$-1.1667$}} & {\ssz{$130.60\degrees$}} \\ \hline 
1506: {\ssz{$60$}} & {\ssz{$-11$}}  & {\ssz{$13.5$}}  & {\ssz{$-0.1833$}}  & {\ssz{$0.225$}}  & {\ssz{$-1.2273$}} & {\ssz{$129.17\degrees$}} \\ \hline 
1507: {\ssz{$90$}} & {\ssz{$-17.5$}}  & {\ssz{$21.5$}}  & {\ssz{$-0.1944$}}  & {\ssz{$0.2389$}}  & {\ssz{$-1.2286$}} & {\ssz{$129.14\degrees$}} \\ \hline  
1508: {\ssz{$120$}} & {\ssz{$-25$}}  & {\ssz{$30$}}  & {\ssz{$-0.2083$}}  & {\ssz{$0.25$}}  & {\ssz{$-1.2$}} & {\ssz{$129.81\degrees$}} \\ \hline
1509: \end{tabular}
1510: \end{center}
1511: \caption{Summary of breather motion ($\Psi=130\degrees$) 
1512: see Section \ref{hpsi130} for details.}
1513: \label{tsimh7}
1514: \end{table}
1515: 
1516: 
1517: %----------------------------------------------------------------------
1518: \subsection{Stationary breather in a lattice with asymmetric potential} 
1519: \lbl{hnra}
1520: 
1521: 
1522: In Sections \ref{hnstsy}--\ref{hpsi130}, we have shown simulations
1523: of lattices with a symmetric potential, that is, with $a$ and $c$ not 
1524: necessarily zero in \eref{heqqfo}.   We now consider the more general 
1525: case for which the interaction potential is asymmetric ($a\neq0 \neq c$).  
1526: To illustrate a stationary breather, we generate initial data using 
1527: \eref{hexqasym},  with the wavevector $\mb{k} = \mb{k_1} = 
1528: [\pi/3,\pi/h]^T$.  As discussed in \Sref{hexasymm}, anomalous dispersion 
1529: corresponds to $b>10a^2/9$, hence we choose $a=1$, $b=2.5$, $c=0$ 
1530: and $d=1$.  The remaining parameter values are $N=30$, $\eee=0.1$, 
1531: $\lambda=1$, hence $\w=3$, $T=2.0668$, $\alpha=0.8665$ and $\beta=1.8670$. 
1532: 
1533: 
1534: The breather is shown in \Fref{simh5} after $10$, $30$ and $40$ full 
1535: oscillations.  Initially, we find the breather's width and energy are  
1536: $\wbr=7.46$ and $\elo=0.5429$, the latter being only 0.26\% different 
1537: from the asymptotic estimate \eref{henex} of 0.5416.   The accompanying 
1538: plots of $e_{m,n}$ demonstrate clearly that the breather preserves its 
1539: form and remains localised, even after 80 seconds.  At $t=40T$, we find 
1540: that $\wbr=6.83$, a narrowing of the breather by 8\%.  Also, at $t=40T$, 
1541: the numerically computed value of the energy $E_0=0.5371$, showing 
1542: that the energy does not fluctuate significantly.  
1543: 
1544: 
1545: \begin{figure}[htbp]
1546: \centering % PICTURE 1
1547: \subfigure[$t=10T=20.69$.]{
1548: \includegraphics[width=.45\textwidth]{hs5b.eps}}
1549: \hspace{.3in} % PICTURE 2
1550: \subfigure[Plot of $e_{m,n}$, $E_0=0.5453$.]{
1551: \includegraphics[width=.45\textwidth]{hs5ben.eps}} \\
1552: \vspace{.3in} % PICTURE 3
1553: \subfigure[$t=30T=41.34$.]{
1554: \includegraphics[width=.45\textwidth]{hs5c.eps}}
1555: \hspace{.3in} % PICTURE 4
1556: \subfigure[Plot of $e_{m,n}$, $E_0=0.5425$.]{
1557: \includegraphics[width=.45\textwidth]{hs5cen.eps}} \\
1558: \vspace{.3in} % PICTURE 5
1559: \subfigure[$t=40T=82.67$.]{
1560: \includegraphics[width=.45\textwidth]{hs5d.eps}}
1561: \hspace{.3in} % PICTURE 4
1562: \subfigure[Plot of $e_{m,n}$, $E_0=0.5371$.]{
1563: \includegraphics[width=.45\textwidth]{hs5den.eps}}
1564: \caption{Stationary breather in a lattice with an asymmetric potential, 
1565: see Section \ref{hnra} for details.}
1566: \label{simh5}
1567: \end{figure}
1568: 
1569: 
1570: %=================================================
1571: \section{Discussion} \lbl{hdisc}
1572: 
1573: 
1574: In this paper, we have found approximations to discrete breathers
1575: in a two-dimensional hexagonal FPU lattice with a scalar-valued 
1576: function at each node.  We have shown that
1577: the lattice equations can be reduced to a cubic NLS equation for
1578: two special cases.  In \Sref{hexsym}, we obtained moving
1579: breathers when the interaction potential is symmetric; in
1580: this case an ellipticity criterion for the wavevector is found.
1581: In the appendix we summarise our technique for approximating 
1582: soliton solutions of the two-dimensional NLS equation. 
1583: 
1584: 
1585: In \Sref{hexasymm}, we considered lattices with an asymmetric 
1586: potential, in which case a reduction to NLS could only be performed
1587: for stationary breathers.   
1588: The theoretical methods employed here are similar to those used on 
1589: the square lattice of \citep{buts06}; however,  several important 
1590: differences emerge in the course of the analysis.  
1591: We found that the anomalous dispersion regime corresponds to 
1592: $b>4a^2/3$ in the square lattice and $b>10a^2/9$ in the hexagonal.  
1593: Furthermore we find that stationary breathers involve the 
1594: generation of second harmonics in the hexagonal lattice 
1595: ($G_2=aF^2/3$), whereas they are suppressed in the square 
1596: lattice ($G_2=0$).  In both lattices, the quadratic nonlinearity 
1597: generates a small amplitude slowly-varying mode ($G_0=-a|F|^2$). 
1598: 
1599: 
1600: For symmetric interactions we found an associated ellipticity constraint: 
1601: moving breathers only occur for certain wavevectors, this means that 
1602: all breathers have (i) a relatively high frequency, and (ii) a maximum 
1603: speed, which depends on the wave vector and hence on the 
1604: direction of travel, (iii) a threshold energy which also depends on 
1605: wavevector and hence is related to speed and direction of travel.  
1606: We find that the threshold energy is lower for breathers at the 
1607: edge of the domain of ellipticity in wavevector space. These breathers 
1608: have lower frequencies and faster speeds.  The fastest-moving 
1609: breathers are restricted to moving along lattice directions. 
1610: 
1611: 
1612: We also presented asymptotic estimates for the breather energy in
1613: Sections \ref{hexsym} and \ref{hexasymm}.  As expected, we found a
1614: minimum energy required to create breathers in the hexagonal
1615: lattice.  The threshold energy for moving breathers is smaller
1616: than that required for stationary breathers and becomes vanishingly
1617: small at the boundary of the domain of ellipticity.
1618: 
1619: 
1620: In \Sref{hex5}, we extended the small amplitude expansion to fifth-order 
1621: and derived a higher-order equation which more correctly describes 
1622: the shape and stability properties of the breather envelope.  We 
1623: obtained a generalised NLS equation \eref{hexgennls} with a variety 
1624: of perturbation terms, some of which are known to be stabilising.  This 
1625: equation is slightly simpler than the corresponding equation obtained 
1626: for the square lattice, namely (3.11) of \citep{buts06}.  In particular, 
1627: the higher-order dispersive terms in \eref{hexgennls} are isotropic, 
1628: reflecting the hexagonal rotational symmetry of the lattice.  For stationary 
1629: breathers in the case of a symmetric potential we find that the cubic 
1630: nonlinearity does not give rise to third harmonics, that is, $H_3=0$ in 
1631: the hexagonal lattice, in contrast to $H_3=bF^3/8$ in the square lattice.
1632: 
1633: 
1634: In \Sref{hexnumerics}, we illustrated  these breather modes, showing 
1635: that both stationary and moving breathers are long-lived.  The 
1636: breather profiles change little over time, a small amount of 
1637: energy is shed due to the initial conditions being only approximate.  
1638: We have successfully propagated long-lived breathers moving in 
1639: directions which are not axes of symmetry of the lattice (for 
1640: instance, $\Psi=130\degrees$),  suggesting that there is no absolute 
1641: restriction upon the direction of travel; this is in contrast to the 
1642: observations reported by Marin \etal \citep{mar00, mar98} for 
1643: mechanical (two-component) two-dimensional lattices, who could 
1644: only find breathers which travelled along axes of symmetry of the lattice. 
1645: 
1646: 
1647: %--------------------------------
1648: \ack
1649: 
1650: IAB would like to thank both the UK Engineering and Physical Sciences 
1651: Research Council for financial assistance, and also Qamran Yaqoob for his 
1652: assistance in the preparation of several diagrams that appear in this paper. 
1653: 
1654: 
1655: %------------------------------------------------------------
1656: \appendix 
1657: \renewcommand{\theequation}{\Alph{section}\arabic{equation}}
1658: \section{Approximation to Townes soliton} \lbl{app}
1659: 
1660: 
1661: Since analytic formulae for Townes solitons are unavailable, we use the 
1662: Rayleigh-Ritz method to find time-harmonic radially symmetric solutions of
1663: \begin{equation}
1664: \ii F_T + D \nabla^2 F + B |F|^2 F= 0 ,  
1665: \lbl{app-pde} 
1666: \end{equation}
1667: of the form $F({\bf x},T)=e^{\ii\lambda T}\phi(r)$ where 
1668: $r=|{\bf x}|=\sqrt{\xi^2 + \eta^2}$.   The function $\phi$ satisfies 
1669: \begin{equation} \lbl{nlsrl}
1670: -\lambda \phi + D\nabla^2 \phi + B\phi^3=0 ,
1671: \end{equation}
1672: where $\nabla^2=\partial_\xi^2 + \partial_\eta^2$.
1673: Equation (\ref{nlsrl}) arises from a variational derivative of  
1674: \begin{equation} \lbl{hvar}
1675: \mcal{E}(\phi) = \int \int \shalf \lambda |\phi|^2 +
1676: \shalf D |\nabla \phi|^2 - \sfr{1}{4} B |\phi|^4 d^2{\bf r}.
1677: \end{equation}
1678: Using a trial solution of the form $\phi=\alpha\,\sech(\beta r)$, 
1679: where $\alpha$ and $\beta$ are undetermined parameters, we find 
1680: \begin{equation} \lbl{hab}
1681: \mcal{E}(\alpha,\beta) = \frac{D(1+2\str{ln}2)}{12} \alpha^2 -
1682: \frac{B(4\str{ln}2-1)}{24}\frac{\alpha^4}{\beta^2} + \frac{\lambda
1683: \str{ln}2}{2}\frac{\alpha^2}{\beta^2} ; 
1684: \end{equation}
1685: $\alpha$ and $\beta$ are determined by seeking stationary 
1686: points of the action $\mcal{E}$, namely  ${\partial \mcal{E}}/
1687: { \partial \alpha} = {\partial \mcal{E}}/{ \partial \beta} = 0$.
1688: Hence we find 
1689: \begin{equation} \lbl{alpbet}
1690: \alpha = \sqrt{\frac{12\lambda\str{ln}2}{B(4\str{ln}2 -1)}} , \qquad 
1691: \beta = \sqrt{\frac{6\lambda\str{ln}2}{D(2\str{ln}2 + 1)}} , 
1692: \end{equation}
1693: and so our approximation to the Townes soliton solution of (\ref{app-pde}) is
1694: \begin{equation} \lbl{vartown}
1695: F = \sqrt{\frac{12\lambda\log2}{B(4\log2 -1)}} \ \exp (\ii \lambda T) \
1696: \sech \left( \sqrt{\frac{6\lambda\log 2}{D(2\log2 + 1)}}
1697: \sqrt{\xi^2 + \eta^2}\right) .
1698: \end{equation}
1699: 
1700: 
1701: \footnotesize
1702: %--------------------------------------------------------------
1703: \begin{thebibliography}{99}
1704: 
1705: \renewcommand{\itemsep}{0pt}
1706: \renewcommand{\parsep}{0pt}
1707: 
1708: \bibitem{thesis} IA~Butt. PhD Thesis. University of Nottingham, 
1709: to be published at {\tt http://etheses.nottingham.ac.uk/} (2006). 
1710: 
1711: \bibitem{buts06} IA~Butt and JAD Wattis.  Discrete breathers 
1712: in a two-dimensional {F}ermi-{P}asta-{U}lam lattice.
1713: \emph{J. Phys. A: Math. Gen.}, {\bf 39}, 4955, (2006).
1714: 
1715: \bibitem{but1d} IA~Butt and JAD Wattis.   Asymptotic analysis of 
1716: combined breather-kink modes in a Fermi-Pasta-Ulam chain. 
1717: \emph{submitted}, (2006). 
1718: 
1719: \bibitem{dav03}  TA~Davydova, AI~Yakimenko, and YA~Zaliznyak.
1720: Two-dimensional solitons and vortices in normal and anomalous 
1721: dispersive media.  \emph{Phys. Rev. E}, {\bf 67}, 026402, (2003).
1722: 
1723: \bibitem{fer40} E~Fermi, J~Pasta, and S~Ulam.  Studies of nonlinear 
1724: problems.   Los Alamos Scientific Report, 1940, originally unpublished. 
1725: Later published in {\em Lectures in Applied Mathematics}, {\bf 15}, 143, (1974).
1726: 
1727: \bibitem{fib98} G~Fibich and G~Papanicolaou.   A modulation 
1728: method for self-focusing in the perturbed critical nonlinear 
1729: {S}chr{\"o}dinger equation. \emph{Phys. Lett. A}, {\bf 239}, 167, (1998).
1730: 
1731: \bibitem{fib99}  G~Fibich and G~Papanicolaou.   Self-focusing in the 
1732: perturbed and unperturbed nonlinear {S}chr{\"o}dinger equation in 
1733: critical dimension.   \emph{SIAM J. Appl. Math.}, {\bf 60}, 183, (1999).
1734: 
1735: \bibitem{fla97}  S~Flach, K~Kladko, and RS~MacKay.  Energy thresholds 
1736: for discrete breathers in one-, two-, and  three-dimensional 
1737: lattices.  \emph{Phys. Rev. Lett.}, {\bf 78}, 1207, (1997).
1738: 
1739: \bibitem{flat94} S~Flach, K~Kladko, and CR~Willis.  Localized excitations 
1740: in two-dimensional lattices. \emph{Phys. Rev. E}, {\bf 50}, 2293, (1994).
1741: 
1742: \bibitem{gia04} J~Giannoulis and A~Mielke.   The nonlinear 
1743: {S}chr{\"o}dinger equation as a macroscopic limit for an oscillator chain 
1744: with cubic nonlinearities.  \emph{Nonlinearity}, {\bf 17}, 551, (2004).
1745: 
1746: \bibitem{gia06} J~Giannoulis and A~Mielke.  Dispersive evolution 
1747: of pulses in oscillator chains with general interaction potentials.
1748: \emph{Disc. Cont. Dyn. Sys. B}, {\bf 6}, 493, (2006).
1749: 
1750: \bibitem{kas04}  M~Kastner.  Energy thresholds for discrete 
1751: breathers.  \emph{Phys. Rev. Lett.}, {\bf 92}, 104301, (2004).
1752: 
1753: \bibitem{kuz} EA Kuznetsov, AM Rubenchik and VE Zakharov. Soliton 
1754: stability in plasmas and hydrodynamics. \emph{Phys Rep}, {\bf 142}, 103, (1986).
1755: 
1756: \bibitem{mac94} RS~MacKay and S~Aubry.  Proof of existence of 
1757: breathers for time-reversible or {H}amiltonian networks of weakly 
1758: coupled oscillators.  \emph{Nonlinearity}, {\bf 7}, 1623, (1984).
1759: 
1760: \bibitem{mar98} JL~Marin, JC~Eilbeck, and FM~Russell.  
1761: Localised moving breathers in a 2{D} hexagonal lattice. 
1762: \emph{Phys. Lett. A}, {\bf 248}, 225, (1998).
1763: 
1764: \bibitem{mar00} JL~Marin, JC~Eilbeck, and FM~Russell.
1765: \emph{2D breathers and applications (Nonlinear Science 
1766: at the Dawn of the 21st Century)}.  Eds: PL Christiansen 
1767: and MP Soerensen.   Springer, Berlin, (2000).
1768: 
1769: \bibitem{mar01}  JL~Marin, FM~Russell, , and JC~Eilbeck.  Breathers 
1770: in cuprate-like lattices.  \emph{Phys. Lett. A}, {\bf 281}, 21, (2001).
1771: 
1772: \bibitem{rem85} M~Remoissenet.  Low-amplitude breather 
1773: and envelope solitons in quasi-one-dimensional physical models.
1774: \emph{Phys. Rev. B}, {\bf 33}, 2386, (1985).
1775: 
1776: \bibitem{wei99} MI~Weinstein.  Excitation thresholds for nonlinear 
1777: localized modes on lattices. \emph{Nonlinearity}, {\bf 12}, 673, (1999).
1778: 
1779: \end{thebibliography}
1780: 
1781: 
1782: \end{document}
1783: