nlin0001049/bh.tex
1: \documentstyle[epsfig]{elsart}
2: \begin{document}
3: \begin{frontmatter}
4: \title{Period Stabilization in the Busse-Heikes Model of the 
5: K\"uppers-Lortz Instability\thanksref{joel}}
6: \author{R. Toral\thanksref{email}}, \author{M. San Miguel and R. Gallego} 
7: \address{
8: Instituto Mediterr\'aneo de Estudios Avanzados\thanksref{www} (UIB-CSIC),
9: E-07071 Palma de Mallorca, Spain
10: }
11: %\date{\today}
12: \thanks[joel] 
13: {This paper is dedicated to Joel Lebowitz on the occasion of his 70th birthday.}
14: %\begin{twocolumn}
15: \thanks[email]{email: {\tt raul@imedea.uib.es}}
16: \thanks[www]{Web site: http://www.imedea.uib.es/PhysDept/}
17: 
18: \begin{abstract}
19: The Busse-Heikes dynamical model is described in terms of relaxational and
20: nonrelaxational dynamics. Within this dynamical picture a diverging alternating
21: period is calculated in a reduced dynamics given by a time-dependent Hamiltonian
22: with decreasing energy. A mean period is calculated which results from noise
23: stabilization of a mean energy. The consideration of spatial-dependent
24: amplitudes leads to vertex formation. The competition of front motion around
25: the vertices and the K\"uppers-Lortz instability in determining an alternating
26: period is discussed.
27: \end{abstract}
28: 
29: \end{frontmatter}
30: 
31: \section{Introduction}
32: 
33: One of the most extensively studied systems, in the field of pattern formation
34: in nonequilibrium systems, is Rayleigh-B\'enard thermal convection. In many
35: geophysical and astrophysical systems,  thermally induced convection is
36: combined with Coriolis forces induced by rotation. Therefore, Rayleigh-B\'enard
37: convection in fluid layers rotating around a vertical axis is a hydrodynamical
38: system of significant importance. Specially interesting is a spatio-temporal
39: regime that takes place above a critical rotation angular velocity. The system
40: breaks up into a persistent dynamical state such that set of parallel
41: convection rolls are seen to change orientation with a characteristic period.
42: This phenomenon is known as the K\"uppers-Lortz instability
43: \cite{KuppersLortz}.  This instability can be described as follows: for  an
44: angular rotation speed $\Omega$ greater than some critical  value $\Omega_c$,
45: convective rolls lose their stability with respect to rolls inclined at an
46: angle of about $60^0$ in the sense of rotation. The new rolls undergo the same
47: instability, so that there is no stable steady-state pattern. As a result
48: spatially disordered patterns arise already arbitrarily close to the onset of
49: convection. Experimental characterization of this regime of spatio-temporal
50: chaos has been reported in~\cite{KLexp}. Most experiments have been performed
51: for small Prandtl numbers and have been theoretically described in the realm of
52: Swift-Hohenberg models with \cite{SHE_mf} and without \cite{SHE_nmf} the
53: inclusion of mean flow effects. On the other hand, large Prandtl numbers lead to
54: more rigid convection rolls. In this situation mean flow coupling can be
55: neglected and, in the limit of infinite Prandtl numbers, three-mode models
56: have been shown \cite{CrossTu,CrossSHE} to exhibit the same qualitative
57: features as more sophisticate Swift-Hohenberg models that take into account the
58: full range of possible roll orientations.
59: 
60: We consider in this paper a three-mode model proposed by Busse and Heikes
61: \cite{BusseHeikes} to study the K\"uppers-Lortz instability. Each mode
62: represents the amplitude of a set of parallel rolls with an orientation of
63: $60^\circ$ to each other. This model contains an attracting heteroclinic cycle
64: connecting three fixed points corresponding to the three different roll
65: solutions. The model predicts successfully the existence of a region in
66: parameter space  in which the roll solution is unstable, but fails to reproduce
67: the experimental observation of an approximately constant period between roll
68: alternation. Whereas Busse and Heikes speculated that  such a constant period
69: would be obtained by the addition of noise, a conclusion confirmed by Stone and
70: Holmes \cite{stone},  no systematic study of the relation of the period to the
71: system parameters has been performed so far. Another explanation for period
72: stabilization has been given by Cross and Tu \cite{CrossTu} who have performed
73: numerical investigations of an extension of the Busse-Heikes equations, where a
74: spatial variation of the amplitudes has been introduced. In this paper, we
75: study in detail these two proposed mechanisms for period stabilization  in the
76: Busse-Heikes model: (i) addition of noise and (ii) the consideration of
77: spatial-dependent terms. 
78: 
79: The paper is organized as follows: in section \ref{II} we present a description
80: of the Busse-Heikes model and give a clear physical explanation of the
81: period divergence. We describe the dynamics in terms of a relaxational and a
82: nonrelaxational part. The alternating period is calculated for the latter part
83: which is associated with a slowly varying time-dependent Hamiltonian. In
84: section \ref{BHnoise} we consider the same model with the inclusion of additive
85: white noise terms and we calculate the mean period stabilized by noise in terms
86: of the previous dynamical picture. Sections \ref{II} and \ref{BHnoise} discuss
87: ordinary differential equations for the amplitudes of the three modes. In
88: section \ref{IV} we consider the more physically appropriate situation of
89: spatial-dependent amplitudes in a $d=2$ model and study the influence on the
90: dynamics of
91: isotropic and anisotropic spatial-dependent terms. We describe the formation of
92: vertices and how the period of roll alternation is determined by the competition
93: of front motion around the vertices and the K\"uppers-Lortz
94: instability.
95: 
96: 
97: \section{Busse-Heikes Model}
98: \label{II}
99: 
100: Based on the fact that, in a first approximation, only three directions are
101: relevant to this problem, Busse and Heikes \cite{BusseHeikes} proposed a
102: dynamical model to study the K\"uppers-Lortz instability. The vertical
103: component of the velocity field $\Psi({\mathbf r},t)$ is written as:
104: \begin{equation}
105: \Psi({\mathbf r},t) = \sum_{j=1}^3 A_j({\mathbf r},t){\rm e}^{iq_0\hat{\mathbf
106: e}_j\cdot{\mathbf r}} + {\mathrm c.c.}
107: \end{equation}
108: (``c.c." denotes complex conjugate).  The vectors $\hat{\mathbf e}_j$ are unit
109: vectors in directions $j=1,2,3$ which form an angle of $60^\circ$ between them, and
110: $q_0$ is the selected wavenumber of the convection pattern. In this model the
111: (complex) amplitudes of the three rotating modes, $A_1$, $A_2$, $A_3$, are
112: independent of space and follow the evolution equations \cite{BusseHeikes}:
113: \begin{eqnarray}
114: \dot A_1 & = & A_1[\nu-|A_1|^2-(1+\mu+\delta)|A_2|^2-
115: (1+\mu-\delta)|A_3|^2], \nonumber\\
116: \dot A_2 & = & A_2[\nu-|A_2|^2-(1+\mu+\delta)|A_3|^2-
117: (1+\mu-\delta)|A_1|^2], \label{eq:bh}\\
118: \dot A_3 & = & A_3[\nu-|A_3|^2-(1+\mu+\delta)|A_1|^2-
119: (1+\mu-\delta)|A_2|^2]. \nonumber
120: \end{eqnarray}
121: The parameter $\nu$ is proportional to the difference between the Rayleigh number
122: and the the critical Rayleigh number for convection. We will consider
123: exclusively in this paper the case of well-developed convection for which the
124: parameter $\nu$ can be rescaled to $1$, i.e. $\nu=1$ henceforth. The exact relation
125: of $\mu$ and $\delta$ to the fluid properties has been given in
126: \cite{KuppersLortz}. We mention here that  $\mu$ is a parameter related to the
127: temperature gradient and the Taylor number (proportional to the rotation speed
128: $\Omega$) in such a way that it takes a nonzero value in the case of no
129: rotation, $\Omega=0$, whereas  $\delta$ is related to the Taylor number in such
130: a way  that $\Omega=0$ implies $\delta=0$. We will consider only $\Omega > 0$,
131: or $\delta > 0$; the case $\Omega < 0$ ($\delta < 0$) follows by a simple
132: change of the coordinate system. Although the dynamical equations are defined
133: for all values of the parameters, only the case $\mu \ge 0$ is physically
134: relevant. Writing $A_j=\sqrt{R_j}{\rm e}^{i\theta_j}$ we obtain equations for
135: the modulus square of the amplitudes $R_j$:
136: \begin{eqnarray}
137: \dot R_1 & = & 2R_1[1-R_1-(1+\mu+\delta)R_2-
138: (1+\mu-\delta)R_3], \nonumber \\
139: \dot R_2 & = & 2R_2[1-R_2-(1+\mu+\delta)R_3-
140: (1+\mu-\delta)R_1], \label{bhr}\\
141: \dot R_3 & = & 2R_3[1-R_3-(1+\mu+\delta)R_1-
142: (1+\mu-\delta)R_2], \nonumber 
143: \end{eqnarray}
144: and for the phases $\theta_j$:
145: \begin{eqnarray}
146: \dot \theta_1 & = & 0, \nonumber\\
147: \dot \theta_2 & = & 0, \label{bha}\\
148: \dot \theta_3 & = & 0. \nonumber
149: \end{eqnarray}
150: It follows that the phases are simply arbitrary constants fixing the location
151: of the rolls. A solution of the  form $\Psi({\mathbf r})=\sqrt{R_j}{\rm
152: e}^{i(q_0\hat{\mathbf e}_j\cdot{\mathbf r}+\theta_j)}+{\mathrm c.c.}$ 
153: represents a set of rolls of wavelength $2\pi/q_0$, oriented in a direction
154: perpendicular to the vector $\hat{\mathbf e}_j$. Hence, in this model one can
155: simply consider the equations for the real variables $R_j$ instead of the
156: equations for the complex variables $A_j$.  A similar set of equations has been
157: proposed to study population  competition dynamics. For a single biological
158: species,  the Verhulst or logistic model assumes that  its population $N(t)$
159: satisfies the evolution equation:
160: \begin{equation}
161: \frac{dN}{dt} = r N (1-\lambda N),
162: \end{equation}
163: where $r$ is the reproductive growth rate and $\lambda$ is a coefficient
164: denoting competition amongst the members of the species. If three species are
165: competing together, it is adequate in some occasions to model this competition
166: by introducing a Gause--Lotka--Volterra type of equations \cite{ML,bennaim}:
167: \begin{eqnarray}
168: \dot N_1 & = & r N_1\left(1-\lambda N_1-\alpha N_2-\beta N_3\right), \nonumber \\
169: \dot N_2 & = & r N_2\left(1-\lambda N_2-\alpha N_3-\beta N_1\right), \\
170: \dot N_3 & = & r N_3\left(1-\lambda N_3-\alpha N_1-\beta N_2\right), \nonumber 
171: \end{eqnarray}
172: which are the same that the Busse-Heikes equations (\ref{bhr}) for the modulus
173: square of the amplitudes $R_j$ with the identifications: $r=2$, $\lambda=1$,
174: $\alpha=1+\mu+\delta$, $\beta=1+\mu-\delta$. These  equations are the basis of
175: May and Leonard analysis \cite{ML}. We also mention the work of
176: Soward \cite{Soward} which is concerned with the study of the nature of the
177: bifurcations mainly, but not limited to, close to the convective instability
178: for small $\nu$, in a slightly more general model that includes also quadratic
179: nonlinearities in the equations. In the remaining of the section we will
180: analyze some of the properties of the solutions of the Busse-Heikes equations
181: (\ref{eq:bh}). Although our analysis essentially reobtains the results of May
182: and Leonard, we find it convenient to give it in some detail because, besides
183: obtaining some further analytical expressions for the time variation of the
184: amplitudes, we are able in some cases of rewriting the dynamics in terms of a
185: Lyapunov potential. The existence of this Lyapunov potential allows us to
186: interpret the asymptotic dynamics for $\mu=0$ as a residual (conservative)
187: Hamiltonian dynamics. For $\mu >0$ we will use an adiabatic approximation with
188: a time-dependent Hamiltonian. This interpretation will turn out to be very
189: useful in the case that noise terms are added to the dynamical equations,
190: because the found Lyapunov potential governs approximately the stationary
191: probability distribution. 
192: 
193: We first look for stationary solutions of the Busse-Heikes equations
194: (\ref{eq:bh}). The fixed point solutions are the following:
195: 
196: (a) The {\sl null} solution: $R_1=R_2=R_3=0$.
197: 
198: (b) {\sl Roll} solutions. There are three families of these solutions, each
199: characterized by a unique nonvanishing amplitude, for instance:
200: $(R_1,R_2,R_3)=(1,0,0)$ is a roll solution with rolls perpendicular to the
201: $\hat{\mathbf e}_1$ direction,  and so on.
202: 
203: (c) {\sl Hexagon} solutions. The three amplitudes are equal and different from 0, namely 
204: $R_1=R_2=R_3= \frac{1}{3+2\mu}$. They only exist for $\mu > -3/2$.
205: 
206: (d) {\sl Rhombus} solutions. There are three families of these solutions, in which 
207: two amplitudes are different from $0$ and the
208: third amplitude vanishes. For instance:
209: $(R_1,R_2,R_3)=(\frac{\mu+\delta}{\mu(\mu+2)-\delta^2},
210: \frac{\mu-\delta}{\mu(\mu+2)-\delta^2},0)$.
211: They only exist for $\mu > \delta$, or  $-1-\sqrt{1+\delta^2} < \mu < 
212: - \delta$.
213: 
214: The stability of the previous solutions can be studied by  means
215: of  a linear stability analysis.  The result is summarized in  Fig.
216: \ref{fig1}.  For $\mu < -3/2$ there are no stable solutions and the amplitudes
217: grow without limit. The rhombus and null solutions are never stable.  The
218: hexagon solutions are stable for $-3/2 < \mu < 0$. The roll solutions are stable
219: for $\mu > \delta$. For $0 < \mu < \delta$ there are no stable solutions, but 
220: the amplitudes remain bounded. This instability can be described as follows:
221: consider the unstable roll solution $(R_1,R_2,R_3) = (1,0,0)$. The amplitude of the $A_2$ mode starts growing and that of $A_1$
222: decreasing in order to reach the roll solution $(0,1,0)$. However, this new
223: roll solution is also unstable, and before it can be reached, the dynamical
224: system starts evolving towards the roll solution $(0,0,1)$, which is unstable
225: and evolves towards the solution $(1,0,0)$ which is unstable, and so on.
226: Schematically, we can represent the situation as follows:
227: \begin{equation}
228: (1,0,0) \to (0,1,0) \to (0,0,1) \to (1,0,0) \to (0,1,0) \dots
229: \end{equation}
230: This is the K\"uppers-Lortz instability that 
231: shows up in the rotation of the convective rolls.
232: The K\"uppers-Lortz unstable region is characterized by the presence
233: of three unstable fixed points, and a
234: heteroclinic cycle connecting them.
235: 
236: The novelty of our treatment
237: consists in writing the Busse-Heikes equations of motion in the form:
238: \begin{equation}
239: \label{dotap}
240: \dot A_j = -\frac{\partial V}{\partial A_j^\ast} + \delta \, v_j, ~~~~~~j=1,2,3,
241: \end{equation}
242: with the {\sl potential function}:
243: \begin{eqnarray} 
244: V(A_1,A_2,A_3) & = & -\left(|A_1|^2+|A_2|^2+|A_3|^2\right) + \frac{1}{2}
245: \left(|A_1|^4+|A_2|^4+|A_3|^4\right)+\nonumber\\
246: & & (1+\mu)\left(|A_1|^2|A_2|^2+
247: |A_2|^2|A_3|^2+|A_3|^2|A_1|^2\right) \label{potential} \\
248: & = & -\left(R_1+R_2+R_3\right) + \frac{1}{2}
249: \left(R_1^2+R_2^2+R_3^2\right)+(1+\mu)\left(R_1R_2+
250: R_2R_3+R_3R_1\right), \nonumber
251: \end{eqnarray}
252: and 
253: \begin{eqnarray}
254: \label{vi}
255: v_1& =& A_1(-|A_2|^2+|A_3|^2) = A_1(-R_2+R_3), \nonumber \\
256: v_2& =& A_2(-|A_3|^2+|A_1|^2) = A_2(-R_3+R_1), \\
257: v_3& =& A_3(-|A_1|^2+|A_2|^2) = A_3(-R_1+R_2). \nonumber
258: \end{eqnarray}
259: The first term in the right-hand side of (\ref{dotap}) describes relaxation in
260: the potential $V(A_1,A_2,A_3)$. In the case $\delta=0$, hence, the dynamics is
261: described simply as the relaxation, along the gradient lines of the potential
262: $V$, in order to reach a minimum of $V$. In the case $\delta >0$ there is
263: another contribution to the dynamics. Its effect can be analyzed partly by
264: looking at the time evolution of the potential:
265: \begin{equation}\label{dvdta}
266: \frac{dV}{dt}=\sum_{j=1}^3\frac{\partial V}{\partial
267: A_j}\frac{dA_j}{dt}+{\mathrm c.c.} = -2\sum_{j=1}^3\left|\frac{\partial
268: V}{\partial A_j}\right|^2 + \delta \left[\sum_{j=1}^3\frac{\partial V}{\partial
269: A_j}v_j+{\mathrm c.c.}\right].
270: \end{equation}
271: 
272: \noindent Therefore, when the so-called {\sl orthogonality condition} is
273: satisfied:
274: %
275: \begin{equation}\label{ortho}
276:   \delta\left[\sum_{j=1}^3\frac{\partial V}{\partial A_j}v_j+{\mathrm c.c.}\right]=0,
277: \end{equation}
278: 
279: \noindent the function $V$ decreases along the dynamical trajectories and it
280: becomes a Lyapunov potential \cite{lyapunov} if it is bounded from below (which
281: is the case for $\mu>-3/2$). Using equations (\ref{potential})
282: and (\ref{vi}), (\ref{dvdta}) can be rewritten as:
283: %
284: \begin{equation}
285: \label{dVdt}
286: \frac{dV}{dt}=-2\sum_{j=1}^3\left|\frac{\partial V}{\partial A_j}\right|^2 -2
287: \mu \delta (|A_1|^2-|A_2|^2)(|A_2|^2-|A_3|^2)(|A_3|^2-|A_1|^2),
288: \end{equation}
289: 
290: \noindent so that the orthogonality condition is seen to be satisfied for
291: $\mu\delta=0$. In the case $\delta=0$, the system is purely relaxational in the
292: potential $V$ and the corresponding stability diagram can be obtained also by
293: looking at the minima of $V$. For the null solution, the potential takes the
294: value $V=0$; for the rhombus, $V=-1/(2+ \mu)$; for the roll solution, $V=
295: -1/2$; and, finally, for the hexagon solution,  $V=-3/(6+4\mu)$. The study of
296: the potential (for $\delta=0$) shows that the null and rhombus solutions
297: correspond always to maxima of the potential and are, therefore, unstable
298: everywhere. It turns out that the rolls (hexagons) are maxima (minima) of the
299: potential for $\mu <0$ and minima (maxima) for $\mu>0$. Also, the potential for
300: the roll solution is smaller that the potential for the other solutions
301: whenever $\mu>0$, indicating that the rolls are the most stable (and indeed the
302: only stable ones) solutions in this case. Unfortunately, this  simple criterion
303: does not have an equivalent in the nonrelaxational case,  $\delta> 0$, for
304: which one has to perform the full linear stability analysis.
305: 
306: \subsection{The case $\mu=0$}
307: 
308: According to the result (\ref{dVdt}), the function $V(A_1,A_2,A_3)$ is a
309: Lyapunov potential whenever $\mu \delta=0$. As discussed in the previous
310: section, the case $\delta=0$ implies a relaxational gradient dynamics in which
311: all variables tend to fixed values.  In the case $\mu=0$, $\delta >0$, the
312: dynamics is nonrelaxational potential \cite{msmmahg,mhgmsm,SMT} and, whereas
313: the dynamics still leads to the surface of minima of the Lyapunov function,
314: there is a residual motion in this surface for which $dV/dt=0$. In other words:
315: the relaxational terms in the dynamics make the system evolve towards the
316: degenerate minimum of the potential (which for $\mu=0$ occurs at
317: $R_1+R_2+R_3=1$).  The residual motion is governed by the nonrelaxational part
318: which is proportional to $\delta$ and this residual motion disappears for
319: $\delta=0$, the relaxational gradient case.
320: 
321: According to this reduction of the dynamics as a residual motion in the surface
322: of minima of the potential $V$, strictly valid only for $\mu=0$, it turns out
323: that it is possible to solve essentially the equations of motion. By
324: ``essentially" we mean that after a transient time in which the system is
325: driven to the minima of $V$, the residual motion is a conservative one in which
326: it is possible to define a Hamiltonian-like function that allows one to find
327: explicit expressions for the time variation of the dynamical variables.  Let us
328: define the variable 
329: \begin{equation}
330: X(t)=R_1+R_2+R_3.
331: \end{equation}
332: It is straightforward to show
333: that, for arbitrary $\mu$ and $\delta$, $X$ satisfies the evolution equation:
334: \begin{equation}
335: \dot X = 2X(1-X)-4\mu Y,
336: \end{equation}
337: where:
338: \begin{equation}
339: Y(t)=R_1R_2+R_2R_3+R_3R_1.
340: \end{equation}
341: In the case $\mu=0$ the equation for $X(t)$ is a closed equation 
342: whose solution is 
343: \begin{equation}
344: X(t) = \frac{1}{\left( \frac{1}{X_0}-1 \right) {\rm e}^{-2t} +1}.
345: \end{equation}
346: Here $X_0=X(t=0)$. From this expression it turns out that  $\lim_{t \to \infty}
347: X(t) = 1$ independently of the initial condition. In practice, and due to the
348: exponential decay towards $1$ of the above expression, after a transient time
349: of order $1$, $X(t)$ already takes its asymptotic value $X(t)=1$. Therefore, we
350: can substitute $R_1(t)$, say, by $1-R_2(t)-R_3(t)$ to obtain evolution
351: equations for $R_2(t)$ and  $R_3(t)$. In this way, the original 3-variable
352: problem, Eqs. (\ref{bhr}), is reduced to a residual dynamics in a 2-variable subspace:
353: \begin{eqnarray}
354: \dot R_2 & = &  2\delta R_2(1-R_2-2R_3), \\
355: \dot R_3 & = & - 2\delta R_3(1-2R_2-R_3). 
356: \end{eqnarray}
357: These are Hamilton's equations:
358: \begin{eqnarray}
359: \dot R_2 & = & 2\delta \frac{\partial {\cal H}}{\partial R_3}, \\
360: \dot R_3 & = & -2\delta \frac{\partial {\cal H}}{\partial R_2},
361: \end{eqnarray}
362: corresponding to the Hamiltonian:
363: \begin{equation} 
364: \label{hamiltonian}
365: {\cal H}(R_2,R_3) = R_2R_3(1-R_2-R_3).
366: \end{equation}
367: As a consequence, in the asymptotic dynamics for which the Hamiltonian
368: description is valid, ${\cal H}(t)$ is a constant of motion, ${\cal H}=E$,
369: which will be called the ``energy". The Hamiltonian dynamics is valid
370: only after a transient time, but the value of $E$ depends only on initial
371: conditions at $t=0$. The dependence of $E$ on the initial conditions can be
372: found by introducing the variable $\hat {\cal H}$:
373: \begin{equation}
374: {\hat {\cal H}} = R_1 R_2 R_3
375: \end{equation}
376: which, in the asymptotic limit ($t\to\infty$) is equivalent to $\cal H$. 
377: It is easy to show that, for arbitrary values of $\mu$ and $\delta$,
378: $\hat {\cal H}$ satisfies the following evolution equation:
379: \begin{equation}
380: {\hat {\cal H}}^{-1}\frac{d \hat {\cal H}}{dt} = 6-(6+4\mu)X
381: \end{equation}
382: (one can reduce the original dynamical problem to variables
383: $\{X,~Y,~\hat{\cal H}\}$  but the equation for $\dot Y$ turns out to be too
384: complicated, see \cite{Soward}). If we substitute the solution for $X(t)$ valid
385: in the case $\mu=0$ we obtain:
386: \begin{equation}
387: \hat{\cal H}(t)= \hat {\cal H}_0\left[(1-X_0)e^{-2t}+X_0\right]^{-3},
388: \end{equation}
389: with $\hat {\cal H}_0=\hat {\cal H}(t=0)$.
390: The asymptotic value for ${\cal H}$ is 
391: \begin{equation}
392: E = \lim_{t \to \infty} {\cal H}(t) = \lim_{t \to \infty} \hat{\cal H}(t)= 
393: \frac{\hat{\cal H}_0}{X_0^3} = 
394: \frac{R_1(0)R_2(0)R_3(0)}{(R_1(0)+R_2(0)+R_3(0))^{3}}.
395: \end{equation}
396: Again, this asymptotic value is reached after a transient time of 
397: order 1. This expression suggests to define the time-dependent variable:
398: \begin{equation} 
399: \label{defet}
400: E(t) = \frac{\hat {\cal H}}{X^3}=\frac{ R_1 R_2 R_3}{(R_1+R_2+R_3)^3},
401: \end{equation}
402: whose evolution equation (again, for arbitrary $\mu$, $\delta$) is:
403: \begin{equation}
404: \label{dqdt}
405: \frac{dE}{dt} = -4\mu \left(X-3\frac{Y}{X}\right) E \equiv -4\mu f(t) E.
406: \end{equation}
407: Therefore, in the case $\mu=0$, $E(t)=E$ is a constant of motion that
408: coincides, in the asymptotic limit when $X=1$, with the numerical value of the
409: Hamiltonian $\cal H$. According to their definition, $E(t)$ is a bounded
410: function $0 \le E(t) \le 1/27$ and $f(t)\ge 0$ for $R_j\geq 0,~j=1,2,3$.
411: 
412: The problem in the case $\mu=0$ can now be given an explicit solution. 
413: After a transient time (or order 1), the motion occurs on
414: the plane $R_1+R_2+R_3=1$, see Fig. \ref{fig2}. The motion
415: is periodic because it corresponds to 
416: a Hamiltonian orbit with a fixed energy. The exact shape of the trajectory 
417: depends on the value of
418: the energy $E$ which, in turn, depends on initial conditions. More
419: interestingly, the period of the orbit can also be computed. For this,
420: we solve the evolution equation (again asymptotically) for, say, $R_3$.
421: By elimination of $R_2$ by setting ${\cal H}=E$ in Eq. (\ref{hamiltonian}):
422: \begin{equation} 
423: R_2= \frac{1}{2}\left( 1-R_3 \pm \sqrt{(1-R_3)^2-4E/R_3}\right),
424: \end{equation}
425: we obtain a closed equation for $R_3$:
426: \begin{equation}
427: \label{dotr3}
428: \dot R_3 = \pm 2\delta \sqrt{R_3^2(1-R_3)^2-4ER_3}.
429: \end{equation}
430: Let $b$ and $c$ be the return points, i.e. the solutions of 
431: \begin{equation}
432: R_3(1-R_3)^2-4E =0,
433: \end{equation} 
434: lying in the interval $(0,1)$ \cite{r3ne0}. The three roots, $a,b,c$, of the
435: above third-degree  equation are real and two of them (the return points
436: $b,c$)  lie in the interval $(0,1)$. The explicit expression for the roots is:
437: \begin{eqnarray}
438: a  & = & \frac{2}{3}\left[1+\cos{\frac{\theta}{3}}\right], \\
439: b  & = & \frac{2}{3}\left[1+\cos{\frac{\theta-2\pi}{3}}\right], \\
440: c  & = & \frac{2}{3}\left[1+\cos{\frac{\theta+2\pi}{3}}\right],
441: \end{eqnarray}
442: where 
443: \begin{equation}
444: \theta=\arccos{(54E-1)}.
445: \end{equation}
446: Integration of (\ref{dotr3}) yields the equation of motion for $R_3(t)$:
447: \begin{equation}
448: \int_c^{R_3(t)} \frac{dx}{\sqrt{x(x-a)(x-b)(x-c)}} = 2 \delta \int_{t_0}^t dt',
449: \end{equation}
450: where we have chosen the initial time $t_0$ to correspond to the minimum value 
451: when $R_3(t)=c$. The integral in the left hand side can be expressed in terms
452: of the Jacobi elliptic function \cite{abra} ${\rm sn}[x|q]$, to yield:
453: \begin{equation}
454: R_3(t)=\frac{bc}{b+(c-b){\rm sn}^2[\delta\sqrt{b(a-c)} (t-t_0)|q]},
455: \end{equation}
456: where
457: \begin{equation}
458: q  = \frac{a(b-c)}{b(a-c)}.
459: \end{equation}
460: The period of the orbit $T$ can be expressed in terms of the complete elliptic
461: function of the first kind $K(q)$:
462: \begin{equation}
463: \label{te2}
464: T= \frac{2}{\delta\sqrt{b(a-c)}} K(q)
465: \end{equation}
466: and $R_3(t)$ can be written as:
467: \begin{equation}
468: \label{r3t}
469: R_3(t)=\frac{bc}{b+(c-b){\rm sn}^2\left[
470: \frac{2K(q)}{T}(t-t_0)|q\right]},
471: \end{equation}
472: Finally, the evolution equations for the other variables are:
473: \begin{eqnarray}
474: R_1(t)& = & R_3(t-T/3), \label{r1t}\\
475: R_2(t)& = & R_3(t-2T/3). \label{r2t}
476: \end{eqnarray}
477: Summarizing, the behavior of the dynamical system in the case $\mu=0$  can be
478: described as follows: after a transient time (or order $1$) the three variables
479: $R_1$, $R_2$, $R_3$ vary periodically in time on the plane $R_1+R_2+R_3=1$.
480: When $R_1$ decreases, $R_2$ increases, etc. The period of the orbit depends
481: only on the initial conditions through a constant of motion $E$.  The explicit
482: expression for the period, Eq. (\ref{te2}), shows that  the period
483: diverges logarithmically when $E$ tends to zero, namely
484: \begin{equation}
485: \label{te}
486: T(E) = -\frac{3}{2\delta}\ln E \times (1+O(E)),
487: \end{equation} 
488: and the amplitude of the oscillations $\Delta\equiv b-c$ depends
489: also on the constant $E$. When $E$ tends to $0$ the amplitude approaches
490: $1$:
491: \begin{equation}
492: \label{bce}
493: \Delta = (1-2E^{1/2}) \times (1+O(E)).
494: \end{equation}
495: All these relations have been confirmed by a numerical integration of the
496: Busse-Heikes equations. In Fig. \ref{fig3} we plot the time evolution of
497: the amplitudes in the case $\mu=0$, $\delta=1.3$. In this figure we can observe
498: that, after an initial transient time, there is a periodic motion
499: (characteristic of the K\"uppers-Lortz instability) well described by the
500: previous analytical expressions. 
501: 
502: \subsection{The case $\mu > 0$}
503:  
504: Once we have understood the case $\mu=0$, we now turn to $\mu > 0$. In this
505: case, the function $V$ is no longer a Lyapunov potential and we can not
506: reduce the motion to a Hamiltonian one on the surface of minima of $V$. However, since the main
507: features of the K\"uppers-Lortz dynamics are already present in the case
508: $\mu=0$ one would like to perform some kind of perturbative analysis valid for
509: small $\mu$ in order to characterize the K\"uppers-Lortz instability. We
510: exploit these ideas in order to develop some heuristic arguments that will
511: allow us to make some quantitative predictions about the evolution of the
512: system.
513: 
514: According to Eq. (\ref{dqdt}), one can infer that $E(t)$ decreases with time in
515: a characteristic time scale of order $\mu^{-1}$. If $\mu$ is small, $E(t)$
516: decreases very slowly and we can extend the picture of the previous section by
517: using an adiabatic approximation. We assume, then, that the evolution for
518: $\mu>0$ can be described by a Hamiltonian dynamics with an energy that slowly
519: decreases with time. Hence, in reducing the energy, the system evolves by
520: spiraling from a periodic orbit to another (similarly to a damped harmonic
521: oscillator). Assuming this picture of a time-dependent energy $E(t)$, the main
522: features of the case $\mu=0$ can now be extended. This model has several
523: predictions:\\
524: 
525: $\bullet$ After a transient time of order $1$, the motion occurs near the plane
526: $R_1+R_2+R_3 =1$. This is checked in the simulations as we can see in Fig.
527: \ref{fig4} where we plot the time evolution of the three amplitudes as well as
528: their sum, in the case $\delta=1.3,~\mu=0.1$.\\
529: 
530: $\bullet$ The period of the orbits is now a function of time. Since the energy
531: decreases towards zero, it follows from Eq. (\ref{te}) that the period
532: diverges with time. Moreover, it is possible to give an approximate expression
533: for the time dependence of the period. By integration of equation (\ref{dqdt}),
534: we obtain:
535: \begin{equation}
536: \label{et}
537: E(t) = E(t_0){\rm e}^{-4\mu\int_{t_0}^{t}f(t')dt'} \approx E(t_0){\rm e}^{-4
538: \mu (t-t_0)},
539: \end{equation} 
540: where we have approximated $f(t)$ by its asymptotic value $f(t)=1$. Once we
541: have the time evolution of the energy, we can compute the time dependence of
542: the period by using $T(t)=T(E(t))$ as given by (\ref{te2}). For late times, the
543: energy is small and the asymptotic result (\ref{te}) leads to:
544: \begin{equation}
545: \label{tmd}
546: T(t) = T_0+\frac{6 \mu}{\delta} t.
547: \end{equation}
548: This shows that the period increases linearly with time, in agreement with the
549: results of \cite{ML} in which the residence period was shown to behave also
550: linearly with time (although with a different prefactor). In order to check
551: this relation, we have performed a numerical integration of Eqs. (\ref{eq:bh})
552: and computed the period $T$, defined as the time it takes for a given amplitude
553: to cross a reference level (taken arbitrarily as $R_j=0.5$), as a function of
554: time. The results for $\delta=\{1.3,~3\}$ and $\mu=\{0.1,~0.01\}$, plotted in
555: Fig. \ref{fig5}, show that there is a perfect agreement between the
556: theoretical expression and the numerical results.\\
557: 
558: $\bullet$ The amplitude of the oscillations, as given by the return points
559: $\Delta(t)=b(t)-c(t)$ is now a function of time. Using expression (\ref{bce})
560: with an energy that decreases with time as in Eq. (\ref{et}) we obtain that the
561: amplitude of the oscillations increases  with time, see Fig. \ref{fig4},
562: and that it approaches $1$ in a time or order $t \sim \mu^{-1}$. More
563: specifically, we have:
564: \begin{equation}
565: 1-\Delta(t)= (1-\Delta_0){\rm e}^{-2 \mu t}.
566: \end{equation}
567: 
568: In summary, for the case $\mu>0$, the period of the orbits,  which is a
569: function of the energy, increases linearly with time and the amplitude of the
570: oscillations approaches $1$.  We characterize in this way the increase of the
571: period between successive alternation of the dominating modes, see Fig.
572: \ref{fig4}, as an effect of the Hamiltonian dynamics with a slowly decreasing
573: energy. This prediction of the Busse--Heikes model for the K\"uppers-Lortz
574: instability is unphysical, since the experimental results do not show such an
575: increase of the period. Busse and Heikes  were fully aware of this problem and
576: suggested that noise terms (``small amplitude disturbances"), that are present
577: at all times, prevent the amplitudes from decaying to arbitrary small levels and
578: a motion which is essentially periodic but with a fluctuating period is 
579: established. In the next section we study the effect of noise in the dynamical
580: equations.
581: 
582: \section{Busse-Heikes model in the presence of noise}
583: \label{BHnoise}
584: 
585: In order to account for the effect of the fluctuations, we modify the
586: Busse-Heikes equations by the inclusion of noise terms:
587: \begin{eqnarray}
588: \dot A_1 & = & A_1[1-|A_1|^2-(1+\mu+\delta)|A_2|^2-
589: (1+\mu-\delta)|A_3|^2] + \xi_1(t), \nonumber\\
590: \dot A_2 & = & A_2[1-|A_2|^2-(1+\mu+\delta)|A_3|^2-
591: (1+\mu-\delta)|A_1|^2] + \xi_2(t), \label{eq:bhn}\\
592: \dot A_3 & = & A_3[1-|A_3|^2-(1+\mu+\delta)|A_1|^2-
593: (1+\mu-\delta)|A_2|^2] + \xi_3(t). \nonumber
594: \end{eqnarray}
595: We take the simplest case in which the $\xi_i(t)$ are, complex, white--noise
596: processes \cite{vanKampen} with correlations:
597: \begin{equation}
598: \langle \xi_i(t) \xi_j^\ast(t') \rangle = 2 \epsilon \delta(t-t') \delta_{ij}.
599: \end{equation}
600: As mentioned before, and in the case of parameter values lying inside the
601: K\"uppers-Lortz instability region, noise prevents the system from spending an
602: increasing amount of time near any of the (unstable) fixed points.  The
603: mechanism for this is that fluctuations are amplified when the trajectory comes
604: close to one of the (unstable) fixed points of the dynamics and the trajectory
605: is then repelled towards another fixed point \cite{stone}. Hence, a
606: fluctuating, but periodic on average, trajectory is sustained by noise.
607: Within the general picture developed in the previous section, the main role of
608: noise for $\mu>0$ is that of preventing $E(t)$ from decaying to zero. This can
609: be understood in the following qualitative terms: when noise is absent, the
610: dynamics brings the system to the surface of minima of $V$, where the
611: dissipative terms act by decreasing the energy in a time scale of order
612: $\mu^{-1}$, see Eq. (\ref{et}). The inclusion of noise has the effect of
613: counteracting this energy decrease that occurs in the surface of minima of $V$. As a consequence, $E(t)$ no longer
614: decays to zero but it stabilizes  around a mean value $\langle E \rangle$. By
615: stabilizing the orbit around that one corresponding to the mean value $\langle
616: E \rangle$, fluctuations in the residual motion stabilize the mean period to a
617: finite value.  In order to check this picture, we have performed numerical
618: simulations of Eqs. (\ref{eq:bhn}) for small noise amplitude $\epsilon$, using
619: a stochastic Runge-Kutta algorithm \cite{SMT}. The numerical simulations, see
620: Fig. \ref{fig6}, show indeed that the trajectories have a well defined
621: average period $\langle T \rangle$. 
622: 
623: From a more quantitative point of view, and according to the previous
624: picture, we can compute the mean period $\langle T \rangle$, which in the purely
625: Hamiltonian case was a function of $E$, see Eq. (\ref{te2}), by using  the same function applied to
626: the mean value of $E$, i.e. $\langle T \rangle =T(\langle E  \rangle )$. This
627: relation has been checked in the numerical simulations. In Fig. \ref{fig7}
628: we plot the mean period $\langle T \rangle$ versus the period calculated from
629: the mean energy, $\langle E  \rangle$, which has also been evaluated
630: numerically. From this figure it appears that our qualitative argument of a
631: trajectory stabilized around the Hamiltonian orbit, corresponding to the
632: average energy, is well supported by the numerical simulations. 
633: 
634: In order to proceed further, we consider the probability distribution for the
635: amplitude variables, $P(A_1,A_2,A_3;t)$ which obeys a Fokker-Planck equation
636: \cite{FP}. For a general dynamics of the type given by Eq. (\ref{dotap}), it is
637: possible to show \cite{SMT} that the stationary probability distribution for
638: the $A_j$ variables is given by
639: %
640: \begin{eqnarray}
641: P_{\rm st}(A_1,A_2,A_3)& =& Z^{-1}\exp[-V(A_1,A_2,A_3)/\epsilon], \label{pdf} \\
642: Z & =&\int dA_1dA_1^\ast dA_2dA_2^\ast dA_3dA_3^\ast\, {\mathrm e}^{-V/\epsilon} 
643: \nonumber 
644: \end{eqnarray}
645: 
646: \noindent whenever two conditions are satisfied:
647: \begin{enumerate}
648:   \item[a)] Orthogonality condition (\ref{ortho}).
649:   \item[b)] The residual dynamics [nonrelaxational part of (\ref{dotap})] is
650:   divergence free:
651:   %
652:   \begin{equation}\label{diver}
653:     \sum_{j=1}^3 \frac{\partial v_j}{\partial A_j}=0.
654:   \end{equation}
655: \end{enumerate}
656: 
657: \noindent In our case of the Busse-Heikes equations the orthogonality condition
658: is satisfied for $\mu=0$, $\delta>0$, and (\ref{diver}) is satisfied
659: independently of $\mu$ and $\delta$. For $\mu >0 $ this is no longer true but
660: we expect that for small $\mu$ a relation similar to (\ref{pdf}) would be valid
661: if we replace $V$ by a function $\Phi$ that differs from $V$ in terms that vanish for
662: vanishing
663: $\mu$.  Using this probability distribution, one can compute the average value
664: of the variable $E$ as:
665: \begin{equation}
666: \langle E \rangle = Z^{-1}\int dA_1dA_1^\ast dA_2dA_2^\ast dA_3dA_3^\ast\, E
667: \exp[-\Phi/\epsilon].
668: \end{equation}
669: We take the crude approximation $\Phi=V$ and, after a change of variables to
670: amplitude and phase, the mean value of the energy can then be computed as:
671: \begin{equation}
672: \langle E \rangle = \frac{\int_0^{\infty} dR_1 \int_0^{\infty}dR_2 
673: \int_0^{\infty}dR_3 E \exp[-V/\epsilon]}
674: {\int_0^{\infty}dR_1 \int_0^{\infty}dR_2 \int_0^{\infty}dR_3 
675: \exp[-V/\epsilon]},
676: \end{equation}
677: where $V$ and $E$ are given in terms of the variables $R_1,R_2,R_3$ in
678: Eqs. (\ref{potential}) and (\ref{defet}), respectively. In the case $\mu=0$ (for
679: which the above expression is exact) we obtain the value $\langle E \rangle =
680: 1/60$, independent of $\epsilon$, and  $T=T(\langle E \rangle = 1/60) \approx
681: 6.4467/\delta$. 
682: 
683: In the case $\mu>0$, the above integral can be performed by means of
684: a steepest descent calculation, valid in the limit $\epsilon \to 0$, where it yields the
685: asymptotic behavior $\langle E \rangle \to (\epsilon/\mu)^2$. The mean period can now be computed, in this limit of small $\epsilon$, using (\ref{te}), with the result that the period, as a function of the system parameters $\delta,~\mu,~\epsilon$, behaves as:
686: \begin{equation}
687: \label{tdme}
688: T(\epsilon,\mu,\delta) \approx \frac{3}{\delta}\ln(\mu/\epsilon),
689: \end{equation}
690: 
691: \noindent a relation that is expected to hold in the limit of small $\epsilon$
692: and for small values of $\mu$. The dependence with $\epsilon$ is
693: the same than the one holding for the mean first passage time in the decay from
694: an unstable state \cite{SMT} and also follows from the general arguments of 
695: \cite{stone}. In Fig. \ref{fig8} we show that there is indeed a
696: linear relation between the period computed in the numerical simulations and 
697: $\delta^{-1}\ln(\mu/\epsilon)$, as predicted by the above formula, although the
698: exact prefactor $3$ is not reproduced. We find it remarkable that, in view of
699: the simplifications involved in our treatment, this linear relation holds for a
700: large range of values for the parameters $\mu$, $\delta$ and $\epsilon$.
701: 
702: 
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
704: \section{Spatial-dependent terms}
705: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
706: \label{IV}
707: 
708: Tu and  Cross \cite{CrossTu} have proposed an alternative explanation for the
709: stabilization of the period without the necessity of the inclusion of the noise
710: terms: they modify the  Busse-Heikes equations by considering two-dimensional
711: amplitude fields, $A_j({\mathbf r},t)$, and including terms accounting for the
712: spatial variation of those fields:
713: \begin{eqnarray}
714: \partial_t A_1 & = & {\cal L}_1A_1+A_1 [1-|A_1|^2-(1+\mu+\delta)|A_2|^2-
715: (1+\mu-\delta)|A_3|^2],  \nonumber\\
716: \partial_t A_2 & = & {\cal L}_2A_2+A_2 [1-|A_2|^2-(1+\mu+\delta)|A_3|^2-
717: (1+\mu-\delta)|A_1|^2], \label{eq:spatial}\\
718: \partial_t A_3 & = & {\cal L}_3A_3+A_3 [1-|A_3|^2-(1+\mu+\delta)|A_1|^2-
719: (1+\mu-\delta)|A_2|^2].\nonumber 
720: \end{eqnarray}
721: 
722: \noindent Here ${\cal L}_j$ ($j=1,2,3$) are linear differential operators. Two
723: main classes of operators can be considered: isotropic and anisotropic. Whereas
724: a multiple scale analysis of the convective instability usually leads to
725: anisotropic terms, the isotropic terms are often justified for the sake of
726: mathematical and numerical simplicity. There are also the natural choice in
727: problems of population dynamics\cite{bennaim}. The simplest isotropic terms are the Laplacian
728: operators:
729: %
730: \begin{equation}
731: \label{ID}
732: {\cal L}_j^{\rm I}=\nabla^2, \quad j=1,2,3.
733: \end{equation}
734: 
735: Two types of anisotropic terms have been proposed for similar fluid problems in
736: the literature: (i) the Newell-Whitehead-Segel (NWS) terms \cite{NWS} and (ii)
737: the Gunaratne-Ouyang-Swinney (GOS) terms \cite{GOS,Graham}. Without altering
738: the essentials of the problem, both NWS and GOS terms can be further simplified
739: leading to second-order directional derivatives along three directions with a
740: relative orientation of $60^\circ$~\cite{CrossTu,echebarria}: 
741: \begin{equation} 
742: \label{AD} {\cal L}_j^{\rm A}=(\hat{\mathbf
743: e}_j\cdot\nabla)^2, \quad j=1,2,3, 
744: \end{equation} 
745: which are the only anisotropic terms considered henceforth. These are more
746: tractable numerically and will be used to compute the alternating period as
747: explained below.  
748: 
749: In this section we will compare the dynamical evolution corresponding to each
750: one of the isotropic and anisotropic spatial dependent terms presented before,
751: Eqs. (\ref{ID}) and (\ref{AD}), respectively.
752: 
753: Common to all of them is that,
754: as in section \ref{II}, we can recast system (\ref{eq:spatial}) into the form:
755: %
756: \begin{equation}\label{eq:BH_model2}
757: \partial_tA_j({\mathbf r},t)=-\frac{\delta{\cal F}_{\mathrm BH}}{\delta 
758: A_j^\ast}+\delta v_j,\quad j=1,2,3,
759: \end{equation}
760: where ${\cal F}_{\mathrm BH}$ is a real functional of the fields
761: given by:
762: \begin{eqnarray}\label{eq:BH_FBH}
763:   {\cal F}_{\mathrm BH}[A_1,A_2,A_3]& = &\int d{\mathbf r}\biggl[
764:   \sum_{j=1}^{3} \left(\frac{1}{2}|{\cal L}_j^{1/2}A_j|^2-
765:   |A_j|^2+\frac{1}{2}|A_j|^4 \right) 
766:   + \nonumber \\
767:   & &  (1+\mu)(|A_1|^2|A_2|^2+|A_2|^2|A_3|^2+|A_3|^2|A_1|^2\biggr]
768: \end{eqnarray}
769: and the functions $v_j$ are given by (\ref{vi}).
770: 
771: As in the zero-dimensional case of sections \ref{II} and \ref{BHnoise},
772: $\delta=0$ entails a relaxational gradient type dynamics and ${\cal F}_{\mathrm
773: BH}$ acts as a Lyapunov functional that decreases monotonically with time.
774: Since this potential is minimized by homogeneous solutions (because the
775: spatial-dependent term gives always a positive contribution) the stationary
776: solutions (and their stability) in the case $\delta=0$ are the same as in the
777: zero-dimensional case. Unfortunately, the orthogonality condition
778: %
779: \begin{equation}
780:   \delta\sum_{j=1}^3 \int d{\mathbf r}\,\frac{\delta{\cal F}_{\mathrm BH}}
781:   {\delta A_j}v_j+{\mathrm c.c.}=0
782: \end{equation}
783: is not trivially satisfied in the case $\mu=0$ for any of the spatial dependent
784: terms mentioned before, and the dynamical equations can not be reduced for
785: $\mu=0$ as in the zero-dimensional case. 
786: 
787: In general, for $0<\delta<\mu$, when the amplitudes grow from random initial
788: conditions around $A_j=0$, $j=1,2,3$, we expect the formation of interfaces
789: between the roll homogeneous states.  Those interfaces move due to curvature
790: and non-potential ($\delta > 0$) effects. Moreover, the fact of dealing with
791: three fields allows the formation of vertices, or points at which the three
792: amplitudes take the same value. In the potential case, $\delta=0$, the
793: interface motion is such that a final state in which a unique roll solution
794: fills the whole space is obtained (a process defined as ``coarsening"). On
795: the other hand, the nonpotential dynamics induces the rotation of front lines
796: around vertices giving rise to the formation of rotating spiral structures
797: \cite{GallegoCPC}. Similar structures have been observed in other three
798: competing species systems, such as lattice voter models \cite{porto}. For small
799: values of $\mu$, the interfaces are wide (it can be shown that an interface
800: varies over a length scale of order $1/\sqrt{\mu}$) and the density of vertices
801: is low. For large $\mu$ the interfaces are sharp and the density of vertices
802: increases. The exact shape of the spirals depends upon the spatial-derivative
803: terms used. With the isotropic terms, Eq. (\ref{ID}), interface propagation
804: follows the normal direction at each point so that closed domains have
805: spherical shape and spiral structures are close to Archimedes' spirals.  On the
806: other hand, for anisotropic spatial derivatives, Eq. (\ref{AD}), interface
807: propagation no longer follows the normal direction and closed domains stretch
808: or collapse along preferential directions so that they adopt an elliptic shape
809: rather than a spherical one. 
810: 
811: An important effect is that the rotation of interfaces around vertices, driven
812: by nonpotential effects, prevents the system from reaching a single roll
813: solution filling the whole space, even outside the K\"uppers-Lortz instability
814: region, i.e. for $\delta < \mu$ \cite{GallegoBH2d}. While this is true both for
815: isotropic and anisotropic derivatives, the dynamical mechanism that prevents
816: this coarsening is different for isotropic and anisotropic terms. For the
817: isotropic terms, vertices of opposite sense of rotation annihilate initially
818: with each other if located closer than a critical distance
819: $d_c\sim\delta^{-1}$. After a transient time in which vertices are formed, they
820: place each other outside the range of effective attraction of other vertices so
821: that their number is essentially constant, thus preventing coarsening. For the
822: anisotropic terms, two interfaces associated with the same vertex (and thus
823: rotating in the same sense) may collide and generate continuously new vertices
824: which, in turn, annihilate against each other
825: again preventing coarsening outside the instability region. A consequence of
826: interface motion is that a fixed point in space sees a change of the dominating
827: amplitude. This alternation change is essentially periodic in time and presents
828: a characteristic period which has nothing in common with the K\"uppers-Lortz
829: instability mechanism in the bulk. Therefore, the period associated to this
830: rotation is continuous at $\delta=\mu$, the instability point.
831: 
832: Before discussing what happens when this interface motion appears together with
833: the instability in the bulk, we mention that for the isotropic terms it is
834: possible to establish an analytical result concerning the front and spiral
835: motion. In this case, using the fact that interfaces move in the normal
836: direction to each point, it is possible to show  that the rotation angular
837: velocity of the interfaces around an isolated vertex scales, for small
838: $\delta$, as $\omega\sim\delta^2$ \cite{GallegoBH2d}. This predicts that, for
839: the isotropic derivatives, the average period in a fixed point of space coming
840: from the rotating spirals scales as $\langle T \rangle \sim \delta^{-2}$. 
841: 
842: As mentioned above, the mechanism of front motion due to the nonpotential
843: effects coexists with the K\"uppers-Lortz bulk instability. We will show in the
844: remaining of the section some results that follow, mainly, from a numerical
845: integration of Eqs. (\ref{eq:spatial}) in two spatial dimensions. It appears
846: from the numerical simulations that  the behavior beyond the instability point
847: (for $\delta>\mu$) depends strongly on the type of spatial derivatives used as
848: well on the magnitude of the parameter $\mu$. We discuss first each type of
849: derivatives separately.
850: 
851: {\bf Isotropic derivatives}: For $\mu$ small, the bulk instability is such that
852: the intrinsic K\"uppers-Lortz period stabilizes to a statistically constant
853: value. In a given point of space, we can see that the dominant
854: amplitude changes due both to invasion from a rotating interface and a new
855: amplitude growing inside the bulk. We give evidence of this combined mechanism
856: in Fig. \ref{fig:smallmu} where we have used the value $\mu=0.1$ and we present
857: representative configurations inside and outside the instability region.
858: 
859: For higher values of $\mu$, the K\"uppers-Lortz intrinsic period in the bulk is
860: observed to increase with time. This is the same phenomenon that occurs in the
861: zero-dimensional model without noise, see section II. Therefore at long times
862: the K\"uppers-Lortz period is so large that we only see rotating interfaces
863: around vertices, just like below the instability point. The two images of the
864: upper row in Fig. \ref{fig:largemu} show domain configurations at long times
865: for $\mu=2.5$, below ($\delta=2$) and beyond ($\delta=3.5$) the K\"uppers-Lortz
866: instability point in the case of the isotropic terms. Apart from the typical
867: size of the domains, it appears that there is no qualitative difference between
868: them.  The period of alternating amplitudes is entirely dominated by front
869: motion.
870: 
871: {\bf Anisotropic derivatives}: Both for small and large $\mu$, in the
872: K\"uppers-Lortz regime, $\delta>\mu$, we observe, in addition to the front
873: motion, domains of one phase emerging in the bulk of other domains; this is
874: seen at all times, indicating that, at variance with the isotropic derivative
875: case, the period associated with the K\"uppers-Lortz instability does not diverge
876: with time. Evidence is given in Fig. \ref{fig:smallmu} for $\mu=0.1$ and
877: Fig. \ref{fig:largemu} for $\mu=2.5$, both figures showing results inside
878: and outside the instability region.
879: 
880: For small $\mu$, in summary, the morphology of domains inside and outside the
881: instability region turns out to be similar with both kinds of spatial dependent
882: terms, Fig. \ref{fig:smallmu}. The alternating period for $\delta>\mu$ is
883: dominated by the K\"uppers-Lortz instability and is similar with isotropic and
884: anisotropic spatial derivatives. This shows up in the fact that the period
885: computed in a single point of space does not depend essentially of the type of
886: derivatives used, as shown in Fig. \ref{fig:period}a. 
887: 
888: For large $\mu$, on the
889: other hand, the morphology is different for isotropic and anisotropic terms.
890: For the isotropic ones, spiral rotation dominates the dynamics because of the
891: very large period associated with the bulk instability. For the anisotropic
892: terms, both front motion and bulk instability are present.
893: Finally in Fig. \ref{fig:period}b (large $\mu$) we show how the alternating
894: period changes when going through the K\"uppers-Lortz instability. We first note
895: that the period does not vanish in the stable regime ($\delta<\mu$). In this
896: regime it is entirely due to front and spiral motion. For isotropic derivatives
897: the period changes smoothly through the point $\delta=\mu$. This supports the
898: fact that the period is still given by front motion for $\delta>\mu$. On the
899: contrary, for anisotropic derivatives a jump in $T$ is observed at $\delta=\mu$.
900: In the K\"uppers-Lortz unstable regime and for anisotropic derivatives, $T$ is
901: determined by a combination of bulk instability and front motion.
902: 
903: %%%%%%%%%%%%%%%%%%%%%%
904: \section{Conclusions}
905: %%%%%%%%%%%%%%%%%%%%%%
906: 
907: We have analyzed the Busse-Heikes equations for Rayleigh-B\'enard convection in
908: a rotating fluid. For the situation of spatial-independent amplitudes, a case
909: previously analyzed by May and Leonard, we find a Lyapunov potential that
910: allows us, for $\mu=0$, to split the dynamics into a relaxational plus a residual
911: part. Since the residual dynamics is Hamiltonian, we are able to give explicit
912: relations for the time variation of the amplitudes and to compute the period of
913: the orbits as a function of the energy, which, in turn, is a function of
914: initial conditions. For $\mu>0$ we extend the previous picture by using an
915: adiabatic approximation in which the energy slowly decreases with time. This
916: allows us to compute the variation of the alternation period between the three
917: modes in the K\"uppers-Lortz instability regime.  We next consider the effect
918: of fluctuations and show how noise can stabilize the mean period to a finite
919: value. By using the Lyapunov potential employed in the deterministic case, we
920: can deduce an approximate expression that yields the period as a function of
921: the system parameters, $\mu,~\delta$ as well as a function of the noise
922: intensity $\epsilon$. The conclusion is that the period increases
923: logarithmically with decreasing noise intensity, a result that is well
924: confirmed by numerical simulations
925: 
926: The two-dimensional version of this problem exhibits rather different dynamical
927: behavior grossly dominated by vertices where three domain walls meet and which
928: have no parallel in lower dimensional systems. The rotation of interfaces
929: around vertices is driven by nonpotential effects and this inhibits coarsening
930: for sufficiently large systems. We investigated the influence on the dynamics
931: of the type of spatial dependent terms. For small values of the parameter
932: $\mu$, the morphology of domains inside the K\"uppers-Lortz region turns out to
933: be similar for both isotropic and anisotropic spatial derivatives. The
934: alternating period is dominated by the K\"uppers-Lortz instability and is
935: similar for both kinds of spatial-dependent terms. For large $\mu$, on the
936: contrary, the morphology of patterns as well as the alternating mean period are
937: different for isotropic and anisotropic terms. While the intrinsic period of
938: the instability diverges with time with isotropic derivatives, it saturates to
939: a finite value in the anisotropic case.
940: 
941: 
942: We acknowledge financial support from DGESIC (Spain) projects numbers
943: PB94-1167 and PB97-0141-C02-01.
944: 
945: 
946: \newpage
947: \begin{thebibliography}{10}
948: 
949: \bibitem{KuppersLortz} G. K\"uppers and D. Lortz, J. Fluid Mech. {\bf 35}, 609
950: (1969).
951: \bibitem{KLexp} Y. Hu, R.E. Ecke and G. Ahlers, Phys. Rev. E {\bf 55}, 6928
952: (1997); Phys. Rev. Lett. {\bf 74}, 5040 (1995). 
953: \bibitem{SHE_mf} Y. Ponty, T. Passot, and P.L. Sulem, Phys. Rev. Lett. {\bf
954: 79}, 71 (1997); H. Xi, J. D. Gunton, and A. Markish, Physica A {\bf 204}, 741
955: (1994).
956: \bibitem{SHE_nmf} M. Neufeld and R. Friedrich, Phys. Rev. E {\bf 51}, 2033
957: (1995); J. Mill\'an-Rodr\'{\i}guez {\it et.~al.}, Chaos {\bf 4}, 369 (1994); J.
958: Mill\'an-Rodr\'{\i}guez {\it et.~al.}, Phys. Rev. A {\bf 46}, 4729 (1992).
959: \bibitem{CrossTu} Y. Tu and M. C. Cross, Phys. Rev. Lett. {\bf 69}, 2515 (1992).
960: \bibitem{CrossSHE} M.C. Cross, D. Meiron, and Y. Tu, Chaos {\bf 4}, 607 (1994).
961: \bibitem{BusseHeikes} F.H. Busse and K.E. Heikes, Science {\bf 208}, 173 (1980).
962: \bibitem{stone} E. Stone and P. Holmes, SIAM J. Appl. Math. {\bf 50}, 726 (1990).
963: \bibitem{ML} R. May and W.J. Leonard, SIAM J. App. Math. {\bf 29}, 243 (1975).
964: \bibitem{bennaim} L. Fracheborug, P.L. Krapivsky and E. Ben-Naim, Phys. Rev.
965: E {\bf 54}, 6186 (1996).
966: \bibitem{Soward} A.M. Soward, Physica D {\bf 14}, 227 (1985).
967: \bibitem{lyapunov} See, for example, J. Guckenheimer and P. Holmes, {\sl Nonlinear Oscillations, Dynamical Systems and Bifurcations of Vector Fields}, Springer-Verlag (1983).
968: 
969: \bibitem{msmmahg} M. San Miguel, R. Montagne, A. Amengual and E.
970: Hern\'andez-Garc\'{\i}a, in {\sl Instabilities and Nonequilibrium Structures V},
971: E. Tirapegui and W. Zeller, eds. Kluwer Academic Pub. (1996).
972: 
973: \bibitem{mhgmsm} R. Montagne, E. Hern\'andez-Garc\'{\i}a, and M. San Miguel,
974: Physica D {\bf 96}, 47 (1996).
975: 
976: \bibitem{SMT} M. San Miguel and R. Toral, in {\sl Instabilities and Nonequilibrium Structures, VI}, E. Tirapegui and W. Zeller, eds. Kluwer Academic Pub. (1999).
977: \bibitem{r3ne0} The case $R_3=0$ necessarily leads to $E=0$ and the dynamics
978: stops.
979: \bibitem{abra} M. Abramowitz and I. Stegun, eds. {\sl Handbook of Mathematical Functions}, Dover Pub. New York (1970).
980: \bibitem{vanKampen} N.G. van Kampen, {\sl Stochastic Processes in Physics and
981: Chemistry}, North-Holland (1981).
982: \bibitem{FP} H. Risken, {\sl The Fokker-Planck equation}, Springer-Verlag,
983: Berlin (1984).
984: \bibitem{NWS} A.C. Newell and J.A. Whitehead, J. Fluid Mech. {\bf 38}, 279
985: (1969); L.A. Segel, J. Fluid Mech. {\bf 38}, 279 (1969).
986: \bibitem{GOS} G. H. Gunaratne, Q. Ouyang, and H. L. Swinney, Phys. Rev. E {\bf
987: 50}, 2802 (1994).
988: \bibitem{Graham} R. Graham, Phys. Rev. Lett. {\bf 76}, 2185 (1996).
989: \bibitem{echebarria} B. Echebarria and H. Riecke, preprint patt-sol/9912002.
990: \bibitem{GallegoCPC} R. Gallego, M. San Miguel, and R. Toral, Comp.
991: Phys. Comm. {\bf 121-122}, 324 (1999). 
992: \bibitem{porto} G. Szab\'o, M.A. Santos and J.F.F. Mendes, preprint,
993: cond-mat/9907333.
994: \bibitem{GallegoBH2d} R. Gallego, R. Toral, and M. San Miguel, in preparation.
995: \end{thebibliography}
996: 
997: 
998: \newpage
999: \begin{figure}
1000: \begin{center}
1001: \psfig{figure=fig1.eps,width=0.6\textwidth}
1002: \end{center}
1003: %\figBH/sta_dia3.eps
1004: \caption{\label{fig1}
1005: Stability regions for the Busse-Heikes dynamical system 
1006: (\ref{eq:bh}). The region `H' is where the hexagon solution (three equal
1007: amplitudes) is stable. In the `R' region, the three roll
1008: solutions are stable, and in region `KL' there are 
1009: no stable fixed points.}
1010: \end{figure}
1011: 
1012: 
1013: \begin{figure}[h]
1014: \begin{center}
1015: \psfig{figure=fig2.eps,width=0.6\textwidth}
1016: \end{center}
1017: \caption{\label{fig2} Dynamics for $\mu=0$ in the variables $R_1,~R_2,~R_3$ for
1018: two different initial conditions.
1019: After a transient time of order $1$ the motion is on the plane 
1020: $R_1+R_2+R_3=1$}
1021: \end{figure}
1022: 
1023: \begin{figure}[h]
1024: \begin{center}
1025: \psfig{figure=fig3.eps,width=0.6\textwidth}
1026: \end{center}
1027: \caption{\label{fig3} 
1028: Time evolution of amplitudes in the case $\delta =1.3$, $\mu=0$. After a
1029: transient time of order $1$, the three variables $R_1,~R_2,~R_3$ vary
1030: periodically in time. The lines are the theoretical predictions that come from
1031: Eqs. (\ref{r3t},\ref{r1t},\ref{r2t}).}
1032: \end{figure}
1033: 
1034: \begin{figure}[h]
1035: \begin{center}
1036: \psfig{figure=fig4.eps,width=0.6\textwidth}
1037: \end{center}
1038: \caption{\label{fig4}  Time evolution of amplitudes in the case $\delta
1039: =1.3$, $\mu=0.1$. The characteristic alternation time of the three variables
1040: $R_1,~R_2,~R_3$ increases with time. Notice that the envelope of the amplitudes
1041: approaches one asymptotically, and that their sum, $R_1+R_2+R_3$ is
1042: approximately equal to $1$.}
1043: \end{figure}
1044: 
1045: \begin{figure}[h]
1046: \begin{center}
1047: \psfig{figure=fig5.eps,width=0.6\textwidth}
1048: \end{center}
1049: \caption{\label{fig5} 
1050: Time evolution of the period, defined as the time it takes a given amplitude
1051: to cross the reference level $R_j=0.5$ plotted versus time for several values
1052: of  $\delta$ and $\mu$. We also plot straight lines with slopes $\frac{6
1053: \mu}{\delta}$ as predicted by Eq. (\ref{tmd}).}
1054: \end{figure}
1055: 
1056: 
1057: \begin{figure}[h]
1058: \begin{center}
1059: \psfig{figure=fig6.eps,width=0.6\textwidth}
1060: \end{center}
1061: \caption{\label{fig6} 
1062: Time evolution of amplitudes in the presence of noise for $\delta =1.30$,
1063: $\mu=0.1$,  $\epsilon = 10^{-7}$. In this case, the motion is such that the
1064: time interval between dominations of a single mode fluctuates around a mean
1065: value (compare with the equivalent deterministic case shown in Fig.
1066: \ref{fig4}).}
1067: \end{figure}
1068: 
1069: \begin{figure}
1070: \begin{center}
1071: \psfig{figure=fig7.eps,width=0.6\textwidth}
1072: \end{center}
1073: \caption{\label{fig7} Plot of the average period $\langle T \rangle$ plotted
1074: versus the theoretical value $T(\langle E \rangle)$ computed using the value of
1075: $\langle E \rangle$ obtained numerically. For each value of $\mu$ and $\delta$
1076: (same symbols meaning than in Fig. \ref{fig5}) we use values of $\epsilon$
1077: ranging from $\epsilon=10^{-2}$ to $\epsilon=10^{-7}$.}
1078: \end{figure}
1079: 
1080: \begin{figure}[h]
1081: \begin{center}
1082: \psfig{figure=fig8.eps,width=0.6\textwidth}
1083: \end{center}
1084: \caption{\label{fig8}  Average period, $\langle T \rangle$, plotted as a
1085: function of $\delta^{-1}\log(\mu/\epsilon)$ in order to check the predicted
1086: linear relation (\ref{tdme}). The straight line is the best fit and has a
1087: slope of $1.73$. Same symbols meanings than in Fig. \ref{fig5} and values
1088: of $\epsilon$ ranging from $\epsilon=10^{-2}$ to $\epsilon=10^{-7}$.}
1089: \end{figure}
1090: 
1091: 
1092: \begin{figure}
1093: \centering
1094: \psfig{file=fig9.eps,width=0.8\textwidth}
1095: \caption{Four snapshots at long times corresponding to a numerical simulation
1096: of the Busse-Heikes model [eq. (\ref{eq:spatial})] with isotropic (${\cal
1097: L}^I_j=\nabla^2$) and anisotropic (${\cal L}^A_j=(\hat{\mathbf
1098: e}_j\cdot\nabla)^2$) spatial
1099: derivatives. Parameter values are: $\mu=0.1$ and $\delta=0.05$ ($1.3$) outside
1100: (inside) the K\"uppers-Lortz instability region.
1101: \label{fig:smallmu}}
1102: \end{figure}
1103: 
1104: \begin{figure}
1105: \centering
1106: \psfig{file=fig10.eps,width=0.8\textwidth}
1107: \caption{Four snapshots at long times corresponding to a numerical simulation
1108: of the Busse-Heikes model [eq. (\ref{eq:spatial})] with isotropic (${\cal
1109: L}^I_j=\nabla^2$) and anisotropic (${\cal L}^A_j=(\hat{\mathbf
1110: e}_j\cdot\nabla)^2$) spatial
1111: derivatives. Parameter values are: $\mu=2.5$ and $\delta=2$ ($3.5$) outside
1112: (inside) the K\"uppers-Lortz instability region.
1113: \label{fig:largemu}}
1114: \end{figure}
1115: 
1116: \begin{figure}
1117: \centering
1118: \psfig{figure=fig11.eps,width=\textwidth}
1119: \caption{Inverse of the alternating mean period as a function of $\delta^2$ for
1120: the two-dimensional Busse-Heikes model with isotropic and anisotropic
1121: spatial-dependent terms. We have chosen the coordinates in order to emphasize
1122: the linear relation between the inverse of the period and $\delta^2$ valid for
1123: small $\delta$\cite{GallegoBH2d}. Each plot
1124: corresponds to a different value of the parameter $\mu$. The K\"uppers-Lortz
1125: instability takes place at the right of the vertical dotted lines.
1126: \label{fig:period}}
1127: \end{figure}
1128: 
1129: \end{document}
1130: