hep-th0003252/qb.tex
1: \documentstyle[12pt,a4wide]{article}
2: \newcommand{\be}{\begin{equation}}
3: \newcommand{\ee}{\end{equation}}
4: \newcommand{\bea}{\begin{eqnarray}}
5: \newcommand{\eea}{\end{eqnarray}}
6: \newcommand{\bean}{\begin{eqnarray*}}
7: \newcommand{\rhobf}{\mbox{\boldmath $\rho$}}
8: \newcommand{\eean}{\end{eqnarray*}}
9: \font\upright=cmu10 scaled\magstep1
10: \font\sans=cmss10
11: \newcommand{\ssf}{\sans}
12: \newcommand{\stroke}{\vrule height8pt width0.4pt depth-0.1pt}
13: \newcommand{\Z}{\hbox{\upright\rlap{\ssf Z}\kern 2.7pt {\ssf Z}}}
14: \newcommand{\ZZ}{\Z\hskip -10pt \Z_2}
15: \newcommand{\C}{{\rlap{\rlap{C}\kern 3.8pt\stroke}\phantom{C}}}
16: \newcommand{\R}{\hbox{\upright\rlap{I}\kern 1.7pt R}}
17: \newcommand{\CP}{\C{\upright\rlap{I}\kern 1.5pt P}}
18: \newcommand{\PP}{\hbox{\upright\rlap{I}\kern 1.5pt P}}
19: \newcommand{\half}{\frac{1}{2}}
20: \newcommand{\identity}{{\upright\rlap{1}\kern 2.0pt 1}}
21: \newcommand{\bm}{\boldmath}
22: \newcommand{\pibf}{\mbox{\boldmath $\pi$}}
23: \newcommand{\taubf}{\mbox{\boldmath $\tau$}}
24: \newcommand{\bn}{{\bf n}}
25: \newcommand{\hf}{\frac{1}{2}}
26: \newcommand{\pp}{\Delta}
27: \newcommand{\HH}{\mbox{\hbox{\upright\rlap{I}\kern 1.7pt H}}}
28: \newcommand{\kk}{\kappa}
29: \newcommand{\zb}{{\bar z}}
30: \newcommand{\vp}{\vert\phi\vert}
31: 
32: %%%%%%%If you do not have the msbm fonts, delete the following 10 lines
33: \font\mybb=msbm10 at 11pt
34: \font\mybbb=msbm10 at 17pt
35: \def\bb#1{\hbox{\mybb#1}}
36: \def\bbb#1{\hbox{\mybbb#1}}
37: \def\bZ {\bb{Z}}
38: \def\bR {\bb{R}}
39: \def\bE {\bb{E}}
40: \def\bT {\bb{T}}
41: \def\bM {\bb{M}}
42: \def\bC {\bb{C}}
43: \def\bA {\bb{A}}
44: \def\e  {\epsilon}
45: \def\bbC {\bbb{C}}
46: %%%%%%%%%%%%%%%%%
47: \renewcommand{\CP}{\bC {\rm P}}
48: \newcommand{\CPP}{\bbC {\rm P}}
49: 
50: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
51: \newcommand{\news}{\setcounter{equation}{0}}
52: \input{epsf}
53: \begin{document}
54: \title{\vskip -70pt
55: \begin{flushright}
56: %{\normalsize UKC/IMS/99/22} \\
57: \end{flushright}\vskip 50pt
58: {\bf \LARGE \bf Q-ball Dynamics }\\[30pt]
59: \author{Richard A. Battye$^{\ \dagger}$ and Paul M. Sutcliffe$^{\ \ddagger}$
60: \\[10pt]
61: \\{\normalsize $\dagger$ {\sl Department of Applied Mathematics and Theoretical
62: Physics,}}
63: \\{\normalsize {\sl Centre for Mathematical Sciences, University of Cambridge,}}
64: \\{\normalsize {\sl Wilberforce Road, Cambridge CB3 0WA, U.K.}}
65: \\{\normalsize {\sl Email : R.A.Battye@damtp.cam.ac.uk}}\\
66: \\{\normalsize $\ddagger$  {\sl Institute of Mathematics, University of Kent at Canterbury,}}\\
67: {\normalsize {\sl Canterbury, CT2 7NZ, U.K.}}\\
68: {\normalsize{\sl Email : P.M.Sutcliffe@ukc.ac.uk}}\\}}
69: \date{March 2000}
70: \maketitle
71: 
72: 
73: \begin{abstract}
74: We investigate the dynamics of Q-balls in one, two and three space dimensions,
75: using numerical simulations of the full nonlinear equations of motion.
76: We find that the dynamics of Q-balls is extremely complex, involving
77: processes such as charge transfer and Q-ball fission.
78:  We present results of simulations which illustrate the salient features of 2-Q-ball interactions
79:  and give qualitative arguments to explain them in terms of the evolution of the time-dependent phases.
80: \end{abstract}
81: 
82: 
83: \newpage
84: \section{Introduction}
85: \news\ \ \ \ \ \
86: One of the most fascinating areas of inter-disciplinary
87: research at the interface between mathematics and physics is the study
88: of {\it solitons}. This word has as many definitions as there are
89: people who study them, but in general terms they are stable, localized energy distributions. From a
90: purely mathematical perspective, solitons are described as
91: extended solutions to a set of hyperbolic or parabolic partial differential
92: equations, which can travel without dissipation at a uniform velocity
93: and maintain, at least asymptotically, their shape during collisions;
94: often these properties of solitons are attributable to the existence of 
95: an infinite number of conserved quantities, connected with the notions integrability,
96: and radiation-free soliton collisions can be constructed.
97: 
98: In the context of particle physics, which is our main interest here,
99:  the usage of the word soliton is less rigorous and any kind of
100:  localized energy distribution falls under this broad umbrella. Only
101:  in very specialised circumstances will soliton collisions not generate
102: radiation and solutions which radiate substantially during, for
103:  example, a highly relativistic 2-soliton collision are included. Of course 
104:  very little exact analytic progress is possible in these more general
105: settings since radiative processes are
106:  notoriously difficult to model analytically, thus numerical simulations
107:  are necessary to probe more complicated situations. The usual
108:  development of  understanding in this subject follows an intricate
109:  interplay between the two, with analytic work putting on solid ground
110:  more qualitative observations from simulations. The domain of
111:  validity of analytic approximations can then be checked and extended
112:  by further simulations. Examples of the classes of solitons which
113:  have been investigated in this way are vortices~\cite{vort}, monopoles~\cite{mon} and skyrmions~\cite{sky}, 
114: whose existence and stability is essentially due to conserved topological currents and charges, 
115: along with energy bounds related to the charge and stable scaling laws.
116: In these examples the topological features constrain the amount of radiation produced
117: in low energy collisions and allows approximations to be applied which ignore the
118: radiative effects.
119: 
120: The subject of this paper is a particular class of solitons known as
121: Q-balls~\cite{lee,coleman}. These are different in many ways to the
122: topological solitons mentioned above. Firstly, they are time-dependent
123: with a rotating internal phase. Secondly, the conserved charge
124: associated with their stability, Noether charge (Q), is
125: not topological and therefore their stability is also a dynamical issue.
126:  As we will see these two features lead to a much more complicated variety of interaction properties
127:  than seen in the study of topological solitons. The main difference is that the charge is quantized in
128:  topological models, it usually being scaled to be an exact integer, whereas we shall see that the  charge Q 
129: can take any value (in a specified range) allowing for the possibility of charge transfer between solitons 
130:  and/or fission during the interaction process.
131: 
132: Although the concepts associated with Q-balls are extremely general and
133: they are likely to exist in a wide variety of physical contexts (for
134: example, see ref.~\cite{lowtemp}), the main motivation for the current
135: study is the recent realization that they are a generic consequence of
136: the Minimal Supersymmetric Standard Model (MSSM)~\cite{k2} due to
137: the existence of D-flat directions in the effective potential created
138: by tri-linear couplings. In
139: this context the conserved charge is that associated with the U(1)
140: symmetries of Baryon and Lepton number conservation and the relevant U(1)
141: fields correspond to either squark or slepton particles. Therefore, the
142: Q-balls can be thought of as  condensates of either a large number of squarks or
143: sleptons. It has been suggested that such condensates can be involved
144: in baryogenesis via the Affleck-Dine mechanism~\cite{AD} after an
145: epoch of inflation in the early universe.
146: If this is the case then there are two interesting
147: possibilities. If the Q-balls can avoid evaporation into lighter,
148: stable  particles such as  protons, then it might be possible for them
149: to be important cosmologically as cold dark matter~\cite{KS}. Whereas
150: if they are unstable, they would decay in a non-trivial way
151: into baryons and could  lead
152: to observable isocurvature baryon fluctuations~\cite{EM}.
153: 
154: Underpinning these interesting suggestions are assumptions as to how
155: Q-balls actually interact and it is our intention here to make an
156: exhaustive study of this issue. Our approach will be to identify
157: numerically the important dynamical processes that can occur in
158: general situations of two interacting Q-balls, which we will then
159: explain qualitatively, leaving a more detailed analytic exposition of
160: the dynamics~\cite{BMS} and the cosmological implications to
161: subsequent papers. In the next section we will discuss the basic
162: properties of static U(1) Q-balls. Then we will present a detailed and
163: extensive study of Q-balls on the line. As we will see, many of the
164: properties of interest such as fission and charge transfer are
165: observed in one-dimension and given the simplicity of simulations, it
166: seems sensible to make the most exhaustive study there. In the
167: subsequent sections on planar Q-balls and fully three-dimensional
168: Q-balls we will show to what extent the one-dimensional simulations can
169: be used to understand the dynamics in higher dimensions and what
170: effects are clearly of higher dimensional origin. In a penultimate section
171: we will discuss the interactions of Q-balls with anti-Q-balls which have
172: an equal and opposite rotation, before making a concluding summary
173: in the final section.
174: 
175: We should note that there is a disparate literature~\cite{lit} on
176: Q-balls in which some (but by no means all) of the processes we will discuss  have already
177: been noted, but not necessarily completely understood. In particular
178: we should mention recent work~\cite{greek} which presented results for
179: Q-balls in one and two dimensions. There it was suggested that the
180: right-angle scattering of solitons seen in two-dimensional
181: topological soliton models is also prevalent in these non-topological
182: models. At the relevant points we will point out in what ways we disagree
183: with their explanation of this phenomena, and demonstrate that it is by no means general.
184: 
185: \section{U(1) Q-balls}
186: \news\ \ \ \ \ \
187: 
188: Given our motivation for studying Q-balls it seems sensible
189: to work with the U(1) Goldstone model, although Q-balls can exist in a
190: variety of field theoretic models. To be precise, the model we consider is that of a single complex scalar
191: field $\phi$ in $D=1,2,3$ spatial dimensions with a potential $U(\vp).$
192: Explicitly, the Lagrangian is
193: \be{\cal L}=\frac{1}{2}\partial_\mu\phi \partial^\mu\bar\phi-
194: U(\vp)\,,\label{lag}\ee
195: with the key feature being the fact that the potential is only a
196: function of $\vp$.
197: The model has a global $U(1)$ symmetry
198: and the associated conserved Noether current $J_{\mu}$ is given by 
199: \be
200: J_{\mu}={1\over
201: 2i}\left(\bar\phi\partial_{\mu}\phi-\phi\partial_\mu\bar\phi\right)\,,
202: \ee
203: whose covariant conservation $\partial^{\mu}J_{\mu}=0$ leads to the
204: existence of the conserved Noether charge Q, given by
205: \be 
206: Q={1\over 2i}\int\left(\bar\phi\partial_t\phi-\phi\partial_t\bar\phi\right)d^Dx=\int \mbox{Im}(\bar\phi \partial_t\phi) \ d^Dx\,.
207: \label{q}\ee
208: A stationary Q-ball solution has the form
209: \be
210: \phi=e^{i\omega t}f(r)\,,
211: \label{qbform}\ee
212: where  $f(r)$ is a real
213: radial profile function which satisfies the ordinary differential equation
214: \be
215: \frac{d^2f}{dr^2}=\frac{(1-D)}{r}\frac{df}{dr}-\omega^2 f+U'(f)\,,
216: \label{profile}\ee
217: with the boundary conditions that $f(\infty)=0$ and $\frac{df}{dr}(0)=0.$
218: 
219: This equation can either be interpreted as describing the motion of a
220: point particle moving in a potential with friction~\cite{coleman}, or
221: in terms of Euclidean bounce solutions~\cite{k1}; in
222: each case the effective potential being $U_{\rm
223: eff}(f)=\omega^2f^2/2-U(f)$. This  leads to constraints
224: on the potential $U(f)$ and the frequency $\omega$ in order for a Q-ball solution to
225: exist. Firstly, the effective mass of $f$ must be negative. If we
226: consider a potential $U(f)$ which is non-negative and satisfies
227: $U(0)=U'(0)=0$, $U''(0)=\omega_+^2>0$, then one can deduce that
228: $\omega<\omega_+$. Furthermore, the minimum of $U(f)/f^2$ must be
229: attained at some positive value of $f$, say $0<f_0<\infty,$ and existence of
230: the solution requires that $\omega>\omega_-$ where 
231: \be \omega_-^2=2U(f_0)/f_0^2\,.
232: \label{range}\ee
233: Hence,  Q-ball solutions exist for all $\omega$ in the range
234: $\omega_-<|\omega|<\omega_+$. Note (i) that solutions exist for
235: positive and negative values of $\omega$, the negative ones being
236: termed anti-Q-balls, (ii) it is often interesting to think of the
237: Q-balls as being akin to charged bubbles; their profiles being very similar. 
238: 
239: The classical stability of the
240: solutions is a more sensitive issue. For sufficiently large $Q$ these
241: solutions are guaranteed to be stable, as can be seen using the \lq thin wall
242: limit\rq\ \cite{coleman}, where the profile function can be
243: approximated by a smoothed-out step function. For small Q it is
244: necessary to perform a full stability analysis using the second
245: variation of the action. In general, the results depend on the details of the
246: potential, but it can be shown that arbitrarily small Q-balls are
247: stable for certain potentials~\cite{k1}. For a rigorous approach to the
248: classical stability of Q-balls see ref.~\cite{St}
249: and references therein. From the quantum mechanical
250: point of view, the solutions are always stable for large enough $Q$ since the energy per
251: unit charge approaches $\omega_-$, which is always less than that for the $\phi$ particle itself, which
252: is $\omega_{+}$.
253: 
254: In choosing a simple potential which admits Q-balls there are
255:  three natural classes  which have been considered, although there are
256:  obviously many other possibilities,
257: \bea
258: &I&: \ \ U(f)=\alpha_1f^2+\alpha_2f^4+\alpha_3f^6\,,\\
259: &II&: \ \ U(f)=\beta_1f^2+\beta_2f^3+\beta_3f^4\,,\\
260: &III&: \ \ U(f)=\gamma_1f^2(1-\gamma_2\log(\gamma_3 f^2))+\gamma_4f^{2p}\,.
261: \eea
262: In each case it is possible to remove two of the parameters by
263:  rescaling the units of energy and time. Note therefore that the
264:  potentials of type I and II have one free parameter, while potentials
265:  of type III have two free parameters for a fixed value of $p$.
266: 
267: The type I potential is the simplest allowed potential which is a polynomial
268: in $f^2$, while type II is the simplest  which is polynomial in
269: $f$. Finally, those of type III  mimic the D-flat
270: direction in the MSSM. Here $p\ge 6$ is some integer that ensures the 
271: growth of the potential for large $f$, but does not destroy the
272: flatness property for intermediate values of $f.$ We should note that
273: none of these types of potential are the kind which might be
274: associated with a renormalizable quantum field theory, but are typical of
275: effective theories incorporating radiative or finite temperature
276: corrections to a bare potential.
277: 
278: In this paper we will mainly be concerned with the type I potential,
279: although we have also studied the type II case. We should note
280: that although the existence and stability properties of Q-balls with
281: these potentials are somewhat different, the qualitative features of the 
282: dynamics appears to be almost independent of the potential. The reason for this is that the
283: main interaction processes that we will describe, charge transfer  and
284: fission, are due mainly to the time-dependent nature of the solution,
285: rather than the precise profile function.
286: 
287: Explicitly, we shall choose our potential so that
288: \be
289: U(f)=f^2(1+(1-f^2)^2)\,,
290: \label{pot}
291: \ee
292: and therefore in terms of the earlier notation we have that
293: $w_+=2$ and $w_-=\sqrt{2}$, so that stable Q-balls exist
294: for $\sqrt{2}<\omega<2$. To illustrate the important features of Q-ball
295: solutions we shall focus on the case of one dimension where the profile function equation (\ref{profile})
296: can be solved exactly~\cite{lee} to give
297: \be
298: f_\omega(r)=\sqrt{\frac{4-\omega^2}{2+\sqrt{2\omega^2-4}
299: \cosh(2r\sqrt{4-\omega^2})}}\,.
300: \label{pro1d}
301: \ee
302: The associated energy $E_\omega$ and charge $Q_\omega$ can then be
303: computed to be 
304: \be
305: Q_\omega=\sqrt{2}\omega\tanh^{-1}\left(\frac{2-\sqrt{2\omega^2-4}}
306: {\sqrt{2}\sqrt{4-\omega^2}}\right)
307: \,,\quad E_\omega=\frac{\sqrt{4-\omega^2}}{2}
308: +\frac{(\omega^2+2)}{2\omega}Q_\omega.
309: %\tanh^{-1}\bigl(\frac{2-\sqrt{2\omega^2-4}}
310: %{\sqrt{2}\sqrt{4-\omega^2}}\bigr)\,.
311: \label{qe1d}\ee
312: 
313: \begin{figure}[ht]
314: \begin{center}
315: \leavevmode
316: \epsfxsize=12cm\epsffile{a01.ps}
317: \caption{The charge $Q$ and energy $E$ as a function of the frequency
318: $\omega.$ } 
319: \label{qvw}
320: \end{center}
321: \end{figure}
322: 
323: In figure~\ref{qvw} we plot the charge $Q_\omega$ and energy
324: $E_\omega$ for $\omega$ in the allowed range  $\sqrt{2}<\omega<2.$
325: and we see that both $Q_\omega$ and  $E_\omega$  are monotonically
326: decreasing functions of $\omega.$ From (\ref{pro1d}) we can deduce that
327: \be
328: f_\omega(0)=\sqrt{{(4-\omega^2)}/{(2+\sqrt{2\omega^2-4)}}}\,,
329: \ee
330: and therefore 
331: $f_\omega(0)$ increases with the charge $Q_\omega$, since it is a decreasing
332: function of $\omega.$ In figure~\ref{evq} we display the energy per unit
333: charge $E_\omega/Q_\omega$ as a function of the charge $Q_\omega$, 
334: from which it can be seen that $E_\omega/Q_\omega$ is a decreasing
335:  function of the charge.
336: Recall that the asymptotic limit as $Q_\omega\rightarrow\infty$ is
337: $E_\omega/Q_\omega=\omega_-=\sqrt{2}.$  
338: Thus, these Q-balls are stable against decay into a number of smaller Q-balls
339: preserving the total charge.
340: \begin{figure}[ht]
341: \begin{center}
342: \leavevmode
343: \epsfxsize=12cm\epsffile{a02.ps}
344: \caption{The energy per unit charge $E/Q$ as a 
345: function of the charge $Q.$ }
346: \label{evq}
347: \end{center}
348: \end{figure}
349: 
350: Most of the general properties of Q-balls in any dimension
351: and for differing choices of the potential are captured by this
352:  one-dimensional example, where explicit formulae are available.
353: However, there are some slight differences, for example
354: if $D=3$ then there is a lower bound on the charge of a Q-ball,
355: whereas in the $D=1$ case considered above Q-balls can have an arbitrarily
356: small charge. This is not a generic feature of every potential and,
357: for example, it has been shown that for $D=3$ arbitrarily small Q-balls can be found
358:  with a potential of type II using a `thick
359: wall limit' in ref.~\cite{k1}. These kind of  
360: details are easily determined by solving the profile function
361: equation (\ref{profile}) numerically and a complete treatment of these
362: issues can be found in ref.~\cite{MV}. But, as we have already noted, we
363: don't believe that they are important for the dynamical processes 
364: which we focus on in the subsequent sections.
365: 
366: \vfill\eject
367: 
368: \section{Q-ball dynamics on the line}\news\ \ \ \ \ \
369: 
370: \noindent In this section we shall investigate the dynamics of 
371: Q-balls in one-dimension;  even for $D=1$ we shall see that
372: multi-Q-ball dynamics is a complicated issue, there being a rich variety of
373: phenomena associated with the non-quantization of the charge and the
374: time-dependent phase.
375: 
376: To investigate the dynamics of Q-balls we numerically solve the field
377: equations which follow from the Lagrangian (\ref{lag}) with
378: the potential (\ref{pot}), namely
379: \be
380: \ddot\phi-\nabla^2\phi+2\phi(2-4\vp^2+3\vp^4)=0\,,
381: \label{eom}
382: \ee
383: which is valid for any value of D.
384: The numerical methods we use are simple finite difference
385: schemes involving either second or fourth order accurate
386: spatial derivatives and a second order leapfrog algorithm
387: for the time evolution with 1000 points, the spatial step size $\Delta x=0.1$ 
388: and the time step size $\Delta t=0.02$. We apply absorbing boundary conditions,
389: which allows radiation to leave the grid and therefore simulates an
390: infinite domain (see refs.~\cite{BS,Bat} for details on
391: how to apply these boundary conditions).
392: 
393: As initial conditions to describe two well-separated Q-balls
394: we use the ansatz
395: \be
396: \phi=e^{i\omega_1t+i\alpha}f_{\omega_1}(\vert x+a\vert)
397: +e^{i\omega_2t}f_{\omega_2}(\vert x-a\vert )\,,
398: \label{ansatz1}
399: \ee
400: in one dimension, which can be trivially generalized to higher dimensions.
401: This ansatz describes a Q-ball with frequency $\omega_1$
402: at the position $x=-a$ and a second Q-ball with frequency
403: $\omega_2$ at the position $x=a.$ The $U(1)$ symmetry
404: of the theory means that for a single Q-ball the phase
405: of $\phi$ can be set to zero at $t=0$ without loss
406: of generality. However, for multi-Q-ball configurations only the
407: initial overall phase can be removed and so for a 2-Q-ball
408: configuration the relative phase, $\alpha$, remains as an important
409: parameter. 
410: 
411: The total charge of this configuration is 
412: \be
413: Q=Q_{\omega_1}+Q_{\omega_2}
414: +(\omega_1+\omega_2)\cos\alpha
415: \int_{-\infty}^\infty f_{\omega_1}(\vert x+a\vert)
416: f_{\omega_2}(\vert x-a\vert)\ dx.
417: \label{totalq}
418: \ee
419: The final term in the above expression is exponentially small
420: in the separation parameter $a$, since the profile functions have an 
421: exponential fall-off. However, we see that the relative phase $\alpha$ 
422: does affect the value of the total charge. Thus, strictly speaking, it
423: is not valid to substitute this ansatz into the energy functional to
424: determine how the energy depends on the relative phase, since this would
425: involve comparing configurations with differing values of $Q.$
426: The same remark also applies to any attempt to determine how the
427: potential energy depends upon the relative separation $a,$ which would
428: help to determine the nature of the interaction force between
429: Q-balls. This issue and its resolution using the methods of
430: ref.~\cite{ams} will be discussed in ref.~\cite{BMS}.
431: \begin{figure}[ht]
432: \begin{center}
433: \leavevmode
434: \epsfxsize=12cm\epsffile{a03.ps}
435: \caption{The charge density at times $0\le t\le 300$,
436: for the parameters $\omega_1=\omega_2=1.5$, $a=4$, $\alpha=0.$ The two
437: Q-balls form a much larger Q-ball, almost stationary at the origin, and
438: the excess is taken away by the fission products: two small Q-balls.} 
439: \label{inphase}
440: \end{center}
441: \end{figure}
442: 
443: To begin with we consider configurations in which the two Q-balls have
444: the same charge, that is $\omega_1=\omega_2=\omega.$  In figure
445: \ref{inphase}, and in all subsequent figures illustrating the dynamics
446: of Q-balls,  we plot the charge density (the integrand in equation
447: (\ref{q})) for the initial conditions and at later times\footnote{We
448: could have made similar plots of the energy density which are not equivalent.
449: We believe the charge density gives a  better
450: representation of the dynamics.}. The
451: parameter values used for this simulation are $\omega=1.5$ and $a=4$,
452: with the Q-balls initially in phase so that $\alpha=0.$ The two Q-balls
453: slowly attract and coalesce to  form one larger Q-ball which has a charge
454: which  is slightly less than the sum of the charges of the two
455: original Q-balls; the charge deficit being  carried away by the fission
456: of two additional Q-balls which, in figure~\ref{inphase}, can just be
457: seen moving away from the almost stationary large Q-ball at the origin. The attraction of the two Q-balls is simple to explain;
458: it being a consequence of the ratio $E/Q$ decreasing as $Q$
459: increases. However, the process of fission is less intuitive and is a
460: novel  concept to those who might have studied the dynamics of
461: topological solitons in an attractive potential. In the
462: topological  case the charge is an integer and so the fission of
463: higher charge solitons can only be achieved when the solitons are
464: moving sufficiently fast for the kinetic energy to overcome the
465: attraction and release a soliton. But here the charge of an isolated Q-ball can have any
466: value, arbitrarily close to zero, and so one might imagine that the
467: energy barrier to  fission at low interaction speeds is small, particularly when the charge is high.
468: 
469: \begin{figure}[ht]
470: \begin{center}
471: \leavevmode \epsfxsize=12cm\epsffile{a04.ps}
472: \caption{The charge density at times $0\le t\le 100$, for a Q-ball with
473: $Q=8.4$ and a scale distortion $\lambda=1.6.$ Notice that the Q-ball
474: splits up into two equal parts. }
475: \label{largesquash}
476: \end{center}
477: \end{figure}
478: \begin{figure}[ht]
479: \begin{center}
480: \leavevmode \epsfxsize=12cm\epsffile{a05.ps}
481: \caption{The charge density at times $0\le t\le 3000$, for a Q-ball
482: with $Q=5.6$ and a scale distortion $\lambda=1.6.$ For this lower
483: charge the Q-ball does not fission and the oscillations damp with time.} 
484: \label{smallsquash}
485: \end{center}
486: \end{figure}
487: 
488: The fission of Q-balls is a process which can occur when a Q-ball suffers a
489: large distortion, for example, during a collision. This can be
490: demonstrated by taking a single Q-ball and squashing it by applying the
491: scale transformation 
492: \be x\mapsto \lambda x, \ \ \ \ \phi\mapsto\sqrt{\lambda}\phi\,,
493: \label{squash}
494: \ee 
495: where $\lambda>1$ is a scale factor. The scaling of the $\phi$
496: field is required in order that the charge of the Q-ball is not changed
497: by the scaling. For small enough values of $\lambda$ the Q-ball
498: oscillates, but then eventually returns to the original un-squashed Q-ball
499: corresponding to $\lambda=1$, which is to be expected since  these
500: Q-balls are stable. However, for a sufficiently large distortion the
501: Q-ball will break up into smaller Q-balls. This is illustrated in figure
502: \ref{largesquash}, where we take a Q-ball with charge $Q=8.4$ and
503: perturb it with a scale $\lambda=1.6.$
504: 
505: To consider the efficiency of the fission process as a function of the
506: charge, we define the quantity \be \Delta(Q)=(2E(Q/2)-E(Q))/E(Q) \,,\ee
507: where $E(Q)$ denotes the energy of a Q-ball with charge $Q.$
508: $\Delta(Q)$ is the fractional increase in energy required to allow a
509: charge $Q$ Q-ball to fission into two charge $Q/2$ Q-balls.  This is a
510: monotonically decreasing function of $Q$ with $\Delta(\infty)=0$,
511: with the limit $Q\rightarrow\infty$ being a Bogomolny-like limit in
512: which the energy is proportional to the charge. Thus we expect that
513: the fission of Q-balls  is more easily stimulated when the charge is
514: large.  To verify this we apply the same distortion factor
515: $\lambda=1.6$ as displayed in figure~\ref{largesquash}, but this time
516: we take a smaller Q-ball with charge $Q=5.6$,  and the results are displayed
517: in figure~\ref{smallsquash}. The Q-ball performs a breather-like motion
518: in which two structures initially begin to form but then recombine.
519: This motion persists for many cycles with a slowly decreasing
520: amplitude until it eventually settles down to a configuration which is
521: very close to the original Q-ball without a distortion (the
522: configuration at $t=3000$ is almost identical to the  stationary
523: $Q=5.6$ Q-ball). 
524: 
525: The above expectations of the fission of Q-balls are confirmed by
526: performing simulations, as in figure~\ref{inphase}, with two
527: stationary Q-balls and varying the charge (increasing or decreasing the
528: value of $\omega$). The charge of the additional Q-balls produced decreases
529: as the charge of the initial Q-balls is reduced  and for small enough
530: Q-balls no fission takes place; rather the sole Q-ball formed
531: oscillates for some time, with a decreasing amplitude.
532: 
533: If the two Q-balls are initially Lorentz boosted toward one another, each with a
534: velocity $v$ say, then if $v$ is large enough the two Q-balls can be
535: made to pass through each other. In figure~\ref{boost} we display the
536: charge density  for a simulation with $\omega_1=\omega_2=1.5$, $a=10$ and $v=0.3.$
537: \begin{figure}[ht]
538: \begin{center}
539: \leavevmode
540: \epsfxsize=12cm\epsffile{a06.ps}
541: \caption{The charge density at times $0\le t\le 150$,
542: for the parameters $\omega_1=\omega_2=1.5$, $a=10$, $\alpha=0$,
543: $v=0.3.$ The Q-balls pass through each other at this interaction speed and
544: coalesce at lower speeds.}
545: \label{boost}
546: \end{center}
547: \end{figure}
548: In this case the two Q-balls pass through each other, although they do
549: lose some charge via radiation during the interaction process and
550: their velocities are reduced. For a slightly lower value
551: of the velocity the two Q-balls pass through each other,
552: but do not escape to infinite separation. Rather they subsequently 
553: recombine, forming a stationary Q-ball at the origin and producing two
554: additional Q-balls in the same manner as described above. In figure
555: \ref{position} we plot the positions of the two main Q-balls
556: (determined as the location of the maximum of the charge density) as a function of time for the velocity $v=0.28.$
557: \begin{figure}[ht]
558: \begin{center}
559: \leavevmode
560: \epsfxsize=12cm\epsffile{a07.ps}
561: \caption{Q-ball positions as a function of time for an
562: initial velocity of $v=0.28$ and all other parameters as in figure
563: \ref{boost}. We see that the Q-balls do not initially coalesce but have
564: insufficient energy to escape as in the case of $v=0.3$ and eventually
565: coalesce at the second attempt.} 
566: \label{position}
567: \end{center}
568: \end{figure}
569: 
570: To study how the relative phase affects the interaction we consider the
571: same initial configuration used to produce figure~\ref{inphase} except
572: that we set $\alpha=\pi$, so that the two Q-balls are exactly out of phase.
573: In this case the resulting evolution is very different and the two Q-balls
574: simply drift apart with no change in their shape or charge, even
575: though the crude arguments based on $E/Q$ suggest that they should
576: attract. The effects of changing the overall phase are similar in many
577: ways to the overall isospin rotations possible in 2-skyrmion
578: interactions. There it is possible to make the skyrmions attract or
579: repel by an internal SU(2) rotation about the line joining the
580: two soliton centres. Another way of understanding the relative phase
581: is as a current between two charged bubbles, if the Q-balls are thought of as
582: bubbles. The results of this repulsive interaction channel are 
583: displayed in figure~\ref{outphase}.
584: 
585: \begin{figure}[ht]
586: \begin{center}
587: \leavevmode
588: \epsfxsize=12cm\epsffile{a08.ps}
589: \caption{As figure~\ref{inphase} except $\alpha=\pi.$ This is the
590: repulsive channel of Q-ball interactions.}
591: \label{outphase}
592: \end{center}
593: \end{figure}
594: \begin{figure}[ht]
595: \begin{center}
596: \leavevmode \epsfxsize=12cm\epsffile{a09.ps}
597: \caption{As figure~\ref{inphase} except $\alpha=\pi/9.$ Notice the
598: novel process of charge transfer where the soliton in the left hand
599: half plane loses charge to the one in the right hand half
600: plane. Clearly the one which has lost charge is moving faster than the
601: other.} 
602: \label{piby4}
603: \end{center}
604: \end{figure}
605: \begin{figure}[ht]
606: \begin{center}
607: \leavevmode
608: \epsfxsize=12cm\epsffile{a10.ps}
609: \caption{The charge $Q_{\rm half}$ on the half-line $x>0$ as a function
610: of time for initial relative phases $\alpha=\pi/9,\pi/4,\pi/2.$}
611: \label{qhalf}
612: \end{center}
613: \end{figure}
614: 
615: As we have seen, for $\alpha=0$ the Q-balls attract and for
616: $\alpha=\pi$ they repel. However, for intermediate values of $\alpha$
617: the dynamics is much more complicated. We display the evolution for
618: $\alpha=\pi/9$ in figure~\ref{piby4}, illustrating a novel process
619: which we shall call charge transfer. The Q-balls initially move very
620: slightly towards each other, but they eventually repel and separate off
621: to infinity. The most interesting aspect is that the charge of the
622: first Q-ball has clearly decreased and that of the second Q-ball has
623: increased, despite the fact that the Q-balls remain two distinct
624: objects with only a very small overlap throughout the time evolution.
625: As they separate the smaller Q-ball moves at a greater speed than the
626: larger one.
627: 
628: In figure~\ref{qhalf} we plot the total charge $Q$ in the
629:  right half of the line, $x>0,$
630: as a function of time for the simulation displayed in figure~\ref{piby4},
631:  where $\alpha=\pi/9$,
632: and also for the cases $\alpha=\pi/4$ and $\alpha=\pi/2.$ We see that as the
633: relative phase is decreased the rate at which the charge transfer initially
634: takes place is reduced but the total charge transfered is increased. The
635: in-phase limit $\alpha=0,$ where the two Q-balls form a single larger Q-ball,
636: is a smooth limit if we interpret it as a total charge transfer. 
637: In the out-of-phase limit $\alpha=\pi$, as we have already seen,
638: there is no charge transfer.
639: If $\alpha<0$ then the same amount of charge is transferred as in the case
640: of a phase $-\alpha$, but this time it is the Q-ball in the left half of the
641: interval which increases in charge.
642: 
643: It may seem surprising that there is charge transfer, but this
644: result can be understood, at least qualitatively, by considering a
645: simplified mechanical analogue of the field dynamics associated
646: with two well-separated Q-balls.
647: Consider two equal charge Q-balls, fixed at the positions $x=\pm a$, 
648: then the ansatz (\ref{ansatz1}) may
649: be written in the form
650: \be
651: \phi=e^{i\theta_1}f(\vert x+a\vert)+e^{i\theta_2}f(\vert x-a\vert)\,,
652: \label{ansatz2}
653: \ee
654: where $\theta_1\equiv\theta_1(t),\theta_2\equiv\theta_2(t)$ are the time-dependent phases of the two Q-balls
655: and $f$ is a profile function.
656: To leading order in the separation $a$, corresponding to the
657: limit of large separation, the  contribution to the
658: Lagrangian is given by 
659: \be
660: {\cal L}=\frac{1}{2}M(\dot\theta_1^2+\dot\theta_2^2)
661: -\epsilon^2\cos(\theta_1-\theta_2)-4M\,,
662: \label{mech}
663: \ee
664: where 
665: $M=\int_{-\infty}^\infty f(\vert x\vert)^2\ dx$
666:  is treated as a constant moment of inertia and
667: $\epsilon^2= 4\int_{-\infty}^\infty f(\vert x+a\vert)f(\vert x-a\vert)\ dx$
668: is a small interaction coefficient. To derive this Lagrangian we have made
669: the assumption that the profile function is time independent, which obviously
670: has a very limited range of validity as we shall discuss further
671: below; it is nonetheless instructive.
672: The equations of motion which follow from (\ref{mech}) are
673: \be
674: \ddot\theta_1+\ddot\theta_2=0\,, \ \ \
675: \ddot\theta_1-\ddot\theta_2=\frac{2\epsilon^2}{M}\sin(\theta_1-\theta_2)\,.
676: \label{phasedynamics}
677: \ee
678: The first of these equations simply represents the fact that
679: the sum of the rotation frequencies $\dot\theta_1+\dot\theta_2$ is conserved, and
680: the second equation determines the dynamics of the relative phase.
681: There are symmetric solutions, $\theta_1=\theta_2$, and 
682: $\theta_1=\theta_2+ \pi$, where the phase difference remains constant,
683: corresponding to the two Q-balls being exactly
684: in-phase or exactly out-of-phase for all time, but for general values
685: of the initial relative phase, $\alpha=\theta_1(0)-\theta_2(0)$,
686:  there will be a non-trivial time dependence.
687: For all $\alpha\in(0,\pi)$ there will be an initial positive acceleration
688: in the relative phase, so that $\dot\theta_1>\dot\theta_2$ for 
689: small $t>0.$ Thus the first Q-ball will have a higher frequency than
690: the second Q-ball and, since we know that the charge of a Q-ball decreases with increasing frequency,
691: then this corresponds to the charge of the first Q-ball decreasing and the
692: charge of the second Q-ball increasing. This simple analysis also predicts that
693: the initial rate of charge transfer will be greatest for a relative phase
694: $\alpha=\pi/2$ and will decrease as $\alpha$ decreases.
695: This agrees with the observation we made earlier by an examination
696: of the plots in figure~\ref{qhalf} for small times. 
697: 
698: However, what we are clearly not able to study with our simple
699: restricted mechanical model is the  whole charge transfer process for
700: later times. One reason for this is that  we assumed that the profile
701: function $f$ was fixed when of course we know that it is highly
702: dependent on the rotation frequency (see equation (\ref{pro1d})). In
703: particular this dependence constrains the rotation frequencies to
704: satisfy $\omega_-<\dot\theta_1,\dot\theta_2<\omega_+$ and as either of
705: these limits are approached our simple model breaks down. One might be
706: tempted to improve our simple model to deal with this issue by
707: including the known frequency dependence of the profile function, but
708: it is not obvious how to do this since the profile function depends
709: upon the frequency $\dot\theta$ so using such an ansatz in the
710: Lagrangian would lead to a Lagrangian for a mechanical system with
711: second order derivatives $\ddot\theta$ and hence a fourth order
712: equation of motion. Furthermore, we have assumed that the positions of
713: the Q-balls are fixed when in fact the results of the full simulations
714: show that they eventually drift apart. This effect will also serve to
715: cut-off the relative phase dynamics since it will correspond to
716: reducing the $\epsilon^2$ coefficient in our simple mechanical model. 
717: 
718: In summary, we have shown that a simple  mechanical model is useful in
719: understanding the qualitative features of the charge transfer process,
720: but a more sophisticated analysis is required to explain the
721: quantitative behaviour found. The analysis of relative phase dynamics
722: in mechanical systems, such as discrete breathers,  has been studied in
723: some detail and the phase space trajectories are well understood
724: \cite{ams}. These methods can be extended to study
725: the more complicated relative phase dynamics, and hence charge transfer, 
726: of Q-balls~\cite{BMS}.
727: 
728: \begin{figure}[ht]
729: \begin{center}
730: \leavevmode
731: \epsfxsize=12cm\epsffile{a11.ps}
732: \caption{The charge density at times $0\le t\le 450$ for the parameters
733: $\omega_1=1.8, \omega_2=1.5, a=3, \alpha=0.$ The two Q-balls repel with
734: virtually no charge transfer since they never get close enough.} 
735: \label{unequal}
736: \end{center}
737: \end{figure}
738:    \begin{figure}[ht]
739: \begin{center}
740: \leavevmode
741: \epsfxsize=12cm\epsffile{a12.ps}
742: \caption{The charge density at times $0\le t\le 95$ for the 
743: parameter values $\omega_1=1.8, \omega_2=1.5, a=6, \alpha=0, v=0.2.$
744: The non-zero relative velocity allows the interaction and charge transfer
745: takes place.}
746: \label{unequalboost}
747: \end{center}
748: \end{figure}
749: 
750: So far we have only considered initial conditions in which the two
751: Q-balls have the same charge. For two Q-balls which have different charges
752:  the initial relative phase does not have the same importance as for
753: Q-balls of the same charge, due to the fact that the Q-balls have
754: different frequencies and so the initial relative phase will not be
755: preserved, even with no interaction. This can easily been seen using
756: the simple mechanical system above. 
757: 
758: In figure~\ref{unequal} we plot the charge density for the initial
759: conditions  $\omega_1=1.8, \omega_2=1.5, a=3, \alpha=0.$ It can be
760: seen that the two Q-balls repel and there is virtually  no charge
761: transfer since the solitons never get close enough for the charge
762: transfer process to become important. Similarly, if a non-zero
763: relative phase is introduced then virtually no charge transfer takes
764: place, although the rate of separation does vary very
765: slightly. However, if the Q-balls are Lorentz boosted towards each
766: other with a sufficiently large velocity $v$ so that they collide and
767: pass through each other, it is possible to induce charge transfer as
768: illustrated in figure~\ref{unequalboost} for the parameters
769: $\omega_1=1.8, \omega_2=1.5, a=6, \alpha=0, v=0.2.$  The amount of
770: charge transferred depends on the value of the relative phase as the
771: Q-balls collide, as can be verified by changing the initial phase.  It
772: can be checked that this is equivalent to varying the initial
773: separation, since the time to collision is then altered and hence the
774: relative phase is different by an amount equal to the change in
775: collision time multiplied by the frequency difference. Just using the
776: simple mechanical analogy, one might have naively expected that an
777: initial difference in the rotation speeds  would be on
778:  a similar footing to an
779: initial phase difference, but this is clearly not the case. It is
780: evident that there is a non-trivial interaction between the relative
781: dynamics of the Q-balls and the charge transfer process.
782: 
783: \section{Planar Q-balls}\news\ \ \ \ \ \
784: 
785: The main features of one-dimensional Q-balls which we have described in
786: the previous section, such as charge transfer and the dependence of
787: the interaction force on the relative phase, carry through to the
788: two-dimensional case. We demonstrate this by again performing
789: numerical simulations of the field equations via an equivalent finite
790: difference scheme to the one dimensional case. We find that a grid
791: containing $200^2$ points with $\Delta x=0.2$ and $\Delta t=0.05$  gives
792: an accurate representation of the dynamics in this case.  In
793: contrast to the one-dimensional case  an exact solution is not known
794: for the profile function in two-dimensions, but it is a simple matter
795: to numerically obtain the profile function using a standard shooting
796: method.  
797: 
798: One might assume that head-on collisions of Q-balls with a small charge
799: (for example, $\omega=1.6$) are equivalent to those in  one-dimension
800: with attraction, repulsion and charge transfer taking place  as before
801: and indeed this is the case as we will discuss later.  The
802: phase-dependent force is, however, a more general concept which we  will
803: illustrate in fully two dimensional interactions with a non-zero impact
804: parameter.  In particular, we consider the collision of two equal
805: charge Q-balls ($\omega_1=\omega_2=1.6$) with a non-zero impact
806: parameter, where the initial positions of the Q-balls are
807: $(x_1,x_2)=\pm(6,3)$, each is boosted along the $x_1$-axis with a
808: velocity $v=0.05$, and the relative phase $\alpha$  is set to zero. In
809: figure~\ref{d10} we plot the charge density for this interaction  at
810: $t=0,104,112,3200.$ 
811: 
812: \begin{figure}
813: \begin{center}
814: \leavevmode
815: %\ \vskip -3cm
816: \epsfxsize=7cm\epsffile{a13.ps}
817: %\ \vskip -2cm
818:  \caption{The charge density at $t=0,104,112,3200$ for two Q-balls with
819: $\omega_1=\omega_2=1.6$, positions $\pm(6,3)$, velocities $v=0.05$, and
820: relative phase $\alpha=0.$ This interaction with non-zero impact parameter 
821: shows the attractive potential of the two Q-balls, which coalesce into a
822: larger Q-ball.}
823: \label{d10}
824: \end{center}
825: \end{figure}
826: 
827: \begin{figure}
828: \begin{center}
829: \leavevmode
830: %\ \vskip -2cm
831: \epsfxsize=7cm
832: \epsffile{a14.ps}
833: \ \vskip 0cm \caption{The charge density at $t=0,104,140,200$ for all
834: parameters as 
835: in figure~\ref{d10} except $\alpha=\pi.$ We now see the repulsive interaction
836: of the two Q-balls as in the one-dimensional case, where it could only be observed through head-on collisions.}
837: \label{d11}
838: \end{center}
839: \end{figure}
840: 
841: \begin{figure}
842: \begin{center}
843: \leavevmode
844: %\ \vskip -1cm
845: \epsfxsize=7cm\epsffile{a15.ps}
846: %\ \vskip -2cm
847:  \caption{As figure~\ref{d11} except $\alpha=\pi/4.$ Charge 
848: transfer takes place in an almost analogous way to the one-dimensional 
849: interactions.}
850: \label{d09}
851: \end{center}
852: \end{figure}
853: 
854: \begin{figure}
855: \begin{center}
856: \leavevmode
857: \ \vskip -3cm
858: \epsfxsize=12cm
859: \centerline{\epsffile{a16.ps}}
860: \caption{The charge density at $t=0,20,24,28,32,36,40,44,52$ for two Q-balls
861: with $\omega_1=\omega_2=1.5$, positions $\pm(10,0)$, velocities $v=0.4$ and
862: relative phase $\alpha=0.$ The two Q-ball under go a complicated interaction 
863: process which eventually leads to them being scattered at right angles to the 
864: incident direction.}
865: \label{d05}
866: \end{center}
867: \end{figure}
868: 
869: As in the one-dimensional case (where collisions are head-on) the two Q-balls
870: attract and form a single larger Q-ball, although the large Q-ball has some 
871: angular momentum due to the fact that the collision was not head-on.
872: Figure~\ref{d11} displays the results of the same simulation except
873: that the two Q-balls are exactly out-of-phase, that is $\alpha=\pi,$
874: where the Q-balls clearly repel. Finally, in figure~\ref{d09}
875: we display the simulation with a relative phase $\alpha=\pi/4$, where
876: there is an initial attraction, followed charge transfer and 
877: finished off by a repulsion which forces the two Q-balls apart. We conclude,
878: therefore, that while the head-on collisions of small Q-balls in two-dimensions
879: can be thought of as being effectively one-dimensional, the same 
880: dynamical processes are active in the case of a non-zero impact parameter.
881: 
882: Next, we turn our attention to head-on collisions of Q-balls with
883: higher charge,  where --- based on the intuition of the
884: one-dimensional interactions --- one would expect things to be
885: slightly different.  We take two Q-balls with $\omega_1=\omega_2=1.5$
886: at positions $(x_1,x_2)=\pm(10,0)$ with each Lorentz boosted towards
887: the other with a velocity $v=0.4.$ Figure~\ref{d05} displays the
888: charge density at $t=0,20,24,28,32,36,40,44,52$ for the in-phase case,
889: $\alpha=0.$ As can be seen from the figure, there is a very
890: complicated interaction process involving the charge being strongly
891: deformed and the emission of some radiation. Eventually, two Q-balls
892: emerge from the  interaction region at right angles to the initial
893: direction of approach.  Naively one may think that this is a simple
894: $90^\circ$ scattering process as seen in a number of topological
895: soliton  models~\cite{vort,mon,sky} and suggested for Q-balls  in
896: ref.~\cite{greek}\footnote{We should note that the work presented in
897: ref.~\cite{greek} uses a potential of type II, not type I as used in
898: our work, but that the qualitative nature of
899: this process is independent of
900: the choice of potential.}. However, the scattering of Q-balls is a  complicated
901: dynamical issue rather than being topological, and the underlying
902: mechanisms are very different. In particular, there is no associated
903: geometry of a moduli space which forces the Q-balls to scatter at right
904: angles.  Rather, during collisions the Q-matter becomes highly
905: deformed with huge charge densities and it is this deformation, and
906: its associated pressures, that lies at the heart of the interaction
907: process and the fission of Q-balls in the plane perpendicular to the
908: incident direction.  As we demonstrated in the one-dimensional case, a
909: sufficient distortion of a Q-ball will induce fission, and it this same
910: process which is responsible for this more complicated phenomena in
911: two-dimensions.
912: 
913: This point can be illustrated immediately by considering the same
914: scattering process, but this time we set the Q-balls to be exactly
915: out-of-phase, that is $\alpha=\pi$ with the results displayed in 
916: figure~\ref{d13}. Although the initial conditions look identical
917: in figures \ref{d05} and \ref{d13}, the evolution is clearly very
918: different. As the two Q-balls are now in a repulsive phase
919: the Q-matter gets distorted in a very different way. Rather
920: than forming a single structure, as in figure~\ref{d05}c, the two
921: individual Q-balls never actually coalesce because of the repulsive 
922: interaction, getting squashed separately and this distortion
923: induces the fission of each. Thus, each Q-ball splits into two and the
924: two pairs repel each other, producing four Q-balls in all, which are
925: clearly visible in figure~\ref{d13}g. For this particular set of 
926: initial parameters the Q-ball pairs do not have enough energy to escape
927: each others attraction and eventually recombine leaving
928: two Q-balls which move off to infinity. By, for example, increasing the
929: initial velocity it is found that the four Q-balls can be produced in such
930: a way that they all separately move off to infinity without
931: any subsequent recombination. If the initial velocity is small enough
932: then the distortion is not large enough to induce fission and the
933: two Q-balls eventually repel keeping their individual structure intact.
934: 
935: \begin{figure}
936: \begin{center}
937: \leavevmode
938: \ \vskip -3cm
939: \epsfxsize=12cm
940: \centerline{\epsffile{a17.ps}}
941: \caption{As figure~\ref{d05} except $\alpha=\pi.$ The two Q-balls come together,
942: but never coalesce. Due to their large charge,
943:  distortion of the Q-matter takes place
944: inducing the fission of four Q-balls in the plane perpendicular to the incident
945: direction.}
946: \label{d13}
947: \end{center}
948: \end{figure}
949: 
950: \begin{figure}
951: \begin{center}
952: \leavevmode
953: \ \vskip -3cm
954: \epsfxsize=12cm
955: \centerline{\epsffile{a18.ps}}
956: \caption{As figure~\ref{d05} except $\alpha=\pi/2.$ Now charge transfer takes
957: place, as well as fission into the plane perpendicular to the incident
958: direction.}
959: \label{d14}
960: \end{center}
961: \end{figure}
962: 
963: \begin{figure}
964: \begin{center}
965: \leavevmode
966: \ \vskip -3cm
967: \epsfxsize=12cm
968: \centerline{\epsffile{a19.ps}}
969: \caption{The charge density at $t=0,20,24,28,36,40,44,52,80$ for two Q-balls
970: with $\omega_1=\omega_2=1.6$, positions $\pm(10,0)$, velocities $v=0.4$ and
971: relative phase $\alpha=0.$ After some oscillations around the centre, the 
972: final configuration settles down to a single Q-ball at the centre.}
973: \label{d06}
974: \end{center}
975: \end{figure}
976: 
977: \begin{figure}
978: \begin{center}
979: \leavevmode
980: %\ \vskip -3cm
981: \epsfxsize=7cm
982: \centerline{\epsffile{a20.ps}}
983: %\ \vskip -2cm
984:  \caption{The charge density at $t=0,16,20,32$ for two Q-balls
985: with $\omega_1=\omega_2=1.6$, positions $\pm(10,0)$, velocities $v=0.6$ and
986: relative phase $\alpha=0.$ The extra kinetic energy results in a highly 
987: inelastic collision, with four small Q-balls left at the end, plus a large
988: amount of radiation.}
989: \label{d07}
990: \end{center}
991: \end{figure}
992: 
993: \begin{figure}
994: \begin{center}
995: \leavevmode
996: %\ \vskip -3cm
997: \epsfxsize=7cm
998: \centerline{\epsffile{a21.ps}}
999: %\ \vskip -2cm 
1000: \caption{The charge density at $t=0,16,24,40$ for two Q-balls
1001: with $\omega_1=\omega_2=1.6$, positions $\pm(10,0)$, velocities $v=0.8$ and
1002: relative phase $\alpha=0.$ The Q-balls are moving so fast now that their 
1003: momentum carries them through each other before they have time to interact.}
1004: \label{d08}
1005: \end{center}
1006: \end{figure}
1007: 
1008: In figure~\ref{d14} we investigate the same simulation as in figure~\ref{d05} 
1009: but with an
1010: initial relative phase $\alpha=\pi/2.$ In this case we expect that the
1011: distortion will also be accompanied by a charge transfer, and 
1012: indeed this is what we find, as the first Q-ball loses charge to the
1013: second one and then each Q-ball undergoes fission to produce a total of two
1014: small Q-balls and two large Q-balls. The two small Q-balls move away from
1015: the interaction region at a faster rate than the large Q-balls, and they
1016: do not recombine. The two large Q-balls move away from the interaction
1017: region at only a very slow speed and in fact they eventually recombine after
1018: a time of around $t=200$ (the final plot in figure~\ref{d14} is only
1019: at $t=52$).
1020: 
1021: As we discussed in the one-dimensional case, large Q-balls are more
1022: susceptible to fission than smaller Q-balls, so we expect that the
1023: scattering processes we have described above will vary depending on
1024: the charge of the initial Q-balls. For example, our reasoning predicts
1025: that the fission process in figure~\ref{d05}, which produced two
1026: Q-balls moving at  right-angles to the initial line of approach, will
1027: be more difficult to reproduce for smaller charges. To test this we
1028: perform the same simulation, with $v=0.4$ again, but decrease the
1029: charge of the initial Q-balls by taking $\omega=1.6$ rather than
1030: $\omega=1.5.$ The resulting evolution is shown in figure~\ref{d06} and
1031: confirms that now the distortion is not sufficient to liberate two
1032: Q-balls. The configuration oscillates for some time before settling
1033: down to a single larger Q-ball, after a small amount of charge has been
1034: dissipated through radiation. Fission can be produced for these
1035: smaller Q-balls by increasing the impact velocity, but this also has
1036: the result that some of the charge passes straight through the
1037: interaction region producing small Q-balls which continue to travel
1038: along the direction of approach. A collision at increased velocities
1039: is also a more violent process and more charge is lost to radiation in
1040: these circumstances. In figure~\ref{d07} we display the evolution for
1041: the case where the velocity is increased to $v=0.6$, with all other
1042: parameters kept the same as in figure \ref{d06}. Just visible in
1043: figure~\ref{d07}d  are the four very small Q-balls which are produced
1044: by this collision together with a ring of charge carried away by the
1045: radiation generated. If the collision velocity is increased further
1046: then the two Q-balls have less time to interact and their momentum
1047: carries them through the collision process with no deflection. This is
1048: demonstrated in figure~\ref{d08} where $v=0.8$ and no additional
1049: Q-balls are produced. In summary, as the charge is reduced an increased
1050: velocity is required in order for a sufficient deformation to be
1051: generated to produce fission, but this also results in  more of the
1052: charge  being carried straight through the collision. Thus for small
1053: charge there is a very limited window of velocities for which
1054: collisions of the form displayed in figure \ref{d05} may occur.
1055: 
1056: %\newpage
1057: 
1058: \section{3D Q-balls}\news\ \ \ \ \ \
1059: 
1060: In the previous two sections we have built up a picture of the dynamics of
1061: Q-balls in one and two dimensions. In going from the extensive study
1062: in one dimension to two dimensions we have noted a number of subtle
1063: effects associated with the extra dimension. However, the basic processes
1064: involved are the same: attraction, repulsion and charge transfer. In
1065: this section we will apply the same numerical techniques to the case
1066: of three dimensions.  To begin with we conducted an extensive study of
1067: the dynamics on grids containing $100^3$ points  and have once again
1068: found that in many cases the dynamics are very similar to those in
1069: one-dimension. At the risk of labouring  the point we found that for
1070: small Q-balls, if they were initially in-phase they coalesced, while if
1071: they were out-of-phase they repelled, and if they had any other phase
1072: they engaged in  charge transfer. However, we did find some extremely
1073: complicated interactions which are related  to those seen in the case
1074: of two dimensions.  As was pointed out in the previous section when
1075: the Q-balls have a large charge, their interactions can have some
1076: interesting variants in two dimensions and it is these particular
1077: cases in three dimensions on which we will focus in this section.
1078: 
1079: \begin{figure}
1080: \begin{center}
1081: \leavevmode
1082: \ \vskip -3cm \hskip 0.5cm
1083: \epsfxsize=12cm\epsffile{one.ps}
1084: \caption{The charge density at t=0,27,33,39,45,60,75,90,105,120,135,150
1085: for two Q-balls with $\omega_1=\omega_2=1.5$, initial positions $\pm(15,0,0)$,
1086: velocities $v=0.4$ and relative phase $\alpha=0$. Shown is the
1087: three-dimensional isosurface and a two dimensional slice through the
1088: centre of the interaction region which should be compared to the
1089: corresponding two dimensional interaction in figure~\ref{d05}. Note
1090: the production of a loop in the plane perpendicular to the line of
1091: incidence, which expands before recollapsing into Q-balls along the
1092: incident direction.}
1093: \label{3d1}
1094: \end{center}
1095: \end{figure}
1096: 
1097: In particular we have focussed on the analogues of figures~\ref{d05},
1098: \ref{d13} and \ref{d14}  in which complicated dynamical phenomena were
1099: identified. In each of these cases we have  performed the analogous
1100: simulations on a grid of $300^3$ points with $\Delta x=0.3$ and
1101: $\Delta t=0.03$ in order to accurately simulate the complicated
1102: dynamical processes. In each of the three cases we start with 2 Q-balls
1103: each with $\omega=1.5$ at $\pm(15,0,0)$, Lorentz boosted toward each
1104: other with a velocity $v=0.4$. The results of the simulations are
1105: displayed in figures~\ref{3d1}, \ref{3d2} and \ref{3d3}.
1106: 
1107: The in-phase case (figure~\ref{3d1}) has some marked similarities
1108: to the equivalent case in two dimensions (figure~\ref{d05}) if one
1109: looks at the two-dimensional slice through the centre of the
1110: Q-balls\footnote{It should be noted that the two interactions are
1111: not equivalent  even when the parameters are almost identical since
1112: the relationship between the charge $Q$ and the frequency $\omega$ is
1113: not the same in two and three dimensions.}. However, the extra dimension has
1114: one remarkable effect: it allows for the production of a loop in the plane
1115: perpendicular to line joining the two incident Q-balls. This phenomena
1116: is the three dimensional analogue of the right-angled fission process
1117: described in the two dimensional case. But in three dimensions the
1118: fission takes place symmetrically in all directions in the plane while
1119: respecting the cylindrical symmetry of the initial configuration. The
1120: loop expands leaving some  charge in the centre which later expands to
1121: create a second, much smaller loop. Later both the loops collapse and
1122: Q-balls emerge back along the incident direction.
1123: 
1124: \begin{figure}
1125: \begin{center}
1126: \leavevmode
1127: \ \vskip -3cm \hskip 0.5cm
1128: \epsfxsize=12cm\epsffile{two.ps}
1129: \caption{As figure~\ref{3d1} except that $\alpha=\pi$, with the
1130: equivalent two dimensional interaction being figure~\ref{d13}. Fission
1131: of the Q-balls produces two loops in the plane perpendicular to the
1132: incident direction, which are subsequently repelled while recollapsing
1133: and emitting Q-balls along the incident direction.}
1134: \label{3d2}
1135: \end{center}
1136: \end{figure}
1137: 
1138: The formation of the loop in this three dimensional simulation adds
1139: further weight  to our earlier discussion of the two dimensional case
1140: where we pointed out that right-angled fission of two Q-balls was not
1141: of topological origin, nor was it even related to the topological
1142: interactions of, for example, vortices~\cite{vort},
1143: monopoles~\cite{mon} and skyrmions~\cite{sky}. The reason being that
1144: in a topological interaction in three dimensions one would have
1145: expected there to have been a preferred direction (this is because in
1146: the case of topological solitons, for example skyrmions, the field
1147: configuration of a single soliton is not spherically symmetric,
1148: although the change in the  field due to a spatial rotation can be
1149: undone by acting with a symmetry of the theory, which means quantities
1150: such as the energy density are spherically symmetric). But as we have
1151: already pointed out the formation of a loop is reliant on all
1152: directions being on an equal footing as far as the fission process is
1153: concerned, which of course is due to the fact that the field itself is
1154: spherically symmetric for a single Q-ball.
1155: 
1156: This explanation of the formation of a loop during the interaction of
1157: two large Q-balls in three dimensions is compatible with the results
1158: of the out-of-phase case (figure~\ref{3d2}) and that of a general
1159: relative phase (figure~\ref{3d3}). In both cases, the two dimensional
1160: slice through the interaction region is very similar to that of the
1161: equivalent two dimensional interactions (figures~\ref{d13} and
1162: \ref{d14} respectively). In the out-of-phase interaction, two
1163: identical loops form which are then repelled back along the direction
1164: from which they came. As they move away they begin to collapse, the
1165: final outcome being a series of symmetrically placed Q-balls along the
1166: incident direction. The interaction for $\alpha=\pi/2$ is similar,
1167: except that, as expected, charge transfer takes place during the
1168: interaction and the two loops created have very different charge.
1169: 
1170: \begin{figure}
1171: \begin{center}
1172: \leavevmode \ \vskip -3cm \hskip 0.5cm
1173: \epsfxsize=12cm\epsffile{three.ps}
1174: \caption{As figure~\ref{3d1} except that $\alpha=\pi/2$, with the
1175: equivalent two dimensional interaction being figure~\ref{d14}. Charge
1176: transfer takes place during the interaction leading to the formation
1177: of two loops which are not of equal charge. These two loops  are then
1178: repelled, the smaller one having a higher speed.}
1179: \label{3d3}
1180: \end{center}
1181: \end{figure}
1182: 
1183: Just to finish this section off and by way of illustrating that this
1184: process is reasonably generic, we have performed an equivalent
1185: simulation to figure~\ref{3d1} with a much higher speed of incidence
1186: $(v=0.8)$. Due to the Lorentz contraction of the initial conditions,
1187: this requires a smaller value of $\Delta x=0.15$ and consequently
1188: $\Delta t=0.015$, and the results of this simulation are displayed in
1189: figure~\ref{3dfast}. We see the production of a big loop at the point
1190: of interaction plus two others which are repelled from the centre
1191: along the line of interaction. At the end of the simulation the loops
1192: are still expanding in size and are also getting close to the size of
1193: the discrete grid. It is an interesting question as to whether loops
1194: can be stable, and this question will be addressed in a separate
1195: publication~\cite{BSut}.
1196: 
1197: \begin{figure}
1198: \begin{center}
1199: \leavevmode \ \vskip -3cm \hskip 0.5cm
1200: \epsfxsize=12cm\epsffile{fast1.ps}
1201: \caption{The charge density at $t=0,27,33,39,45,60,75,90$ for two
1202: Q-balls with $\omega_1=\omega_2=1.5$, initial positions $\pm(15,0,0)$,
1203: velocities $v=0.8$ and relative phase $\alpha=0$. The high speed
1204: collision leads to the formation of three loops, one at the point of
1205: interaction, and two others which propagate back in opposite
1206: directions along the line of interaction.}
1207: \label{3dfast}
1208: \end{center}
1209: \end{figure}
1210: 
1211: \section{Q-ball Anti-Q-ball Dynamics}\news\ \ \ \ \ \
1212: 
1213: In the preceding sections we have studied in detail  2-Q-ball
1214: interactions. The charge $Q$, however,  can also be negative; this
1215: being achieved in the Q-ball solution by replacing $\omega$ with
1216: $-\omega$ and these solutions are known as anti-Q-balls.  In this
1217: section we will study the interactions of Q-ball/anti-Q-ball pairs  in
1218: two and three dimensions.
1219: 
1220: Intuitively, one would expect slow soliton/anti-soliton interactions
1221: with equal and opposite charges to result in annihilation into
1222: radiation. However, only for a very small range of parameters does
1223: this annihilation take place in the case of Q-balls due to the
1224: complicated nature of the time-dependent interaction potential which
1225: we have highlighted in the case of 2-Q-ball interactions.
1226: 
1227: In general a Q-ball/anti-Q-ball interaction will result in either the
1228: two solitons bouncing back, or them passing through each other. In
1229: both cases, the charge is partially annihilated, but only for a very
1230: limited region of the interaction parameter space can it be thought of
1231: as being complete. This absence of the annihilation can be attributed
1232: to the concept of charge transfer which we have discussed  in 2-Q-ball
1233: interactions\footnote{Here, as in the case of two Q-ball interactions
1234: with different charges, the initial relative phase is a less important
1235: concept. But charge transfer can still take place since the difference
1236: between the two rotation speeds of the Q-balls is maximal.}  The main
1237: difference between the charge transfer process for 2-Q-ball
1238: interactions and the Q-ball/anti-Q-ball interactions under
1239: consideration here is the two solitons now have opposite charge and
1240: hence when charge is transfered it results in annihilation. However,
1241: we have showed that the charge transfer never takes place fully in
1242: 2-Q-ball interactions and hence annihilation also never takes place
1243: fully in a single interaction. When there is a lower bound on the
1244: charge of a Q-ball it is more likely that sufficient charge transfer
1245: can take place for complete annihilation, whereas in models where
1246: arbitrarily small Q-balls can exist complete annihilation is likely to
1247: be much more difficult.
1248: 
1249: The process of partial annihilation, via charge transfer, is
1250: illustrated in figure~\ref{e04} where we have displayed the charge
1251: density at $t=0,15,20,50$ for the collision of a Q-ball and an
1252: anti-Q-ball in two dimensions. The solitons were initially at
1253: $\pm(6,0)$, with $\omega_1=-\omega_2=1.8$ and were Lorentz boosted
1254: together with a velocity $v=0.3$. It is clear that the momentum of the
1255: Q-ball carries it through the interaction region and that there is
1256: some annihilation of the charge; the maximum charge density of the
1257: outgoing Q-ball being lower than that for the incoming one. As we have
1258: already discussed this is generically what takes place during a
1259: Q-ball/anti-Q-ball collision. A variant on this kind of interaction is
1260: that the Q-balls bounce back, again partially annihilating charge,
1261: which takes place at low incident velocities for this particular
1262: charge.
1263: 
1264: This picture of partial annihilation with bounce back at low
1265: velocities and the solitons passing through each other at high,
1266: suggests that there exists some critical incident velocity $v_{\rm
1267: c}$, a function of $\omega$ at which annihilation takes place, and
1268: indeed this is what we find\footnote{In fact, there will exist a small
1269: range of velocities around $v_{\rm c}$ for which complete annihilation
1270: takes place.}. Figure~\ref{e03} illustrates this by displaying the
1271: charge density at $t=0,25,50,100,150,250,300,350,400$ for a
1272: Q-ball/anti-Q-ball collision, with the solitons initially positioned
1273: at $\pm(6,0)$, charges $\omega_1=-\omega_2=1.5$, Lorentz boosted
1274: together with velocity $v=0.3$. These are the same parameters as in
1275: figure~\ref{e04}, except that the charge is much larger. It can be
1276: clearly seen that annihilation eventually takes place, but even in
1277: this case the mechanism is complicated, involving a number of
1278: oscillations of the system before it is achieved. The charge here is
1279: much higher than for the example above which had $\omega=1.8$, and in
1280: the Q-ball interactions we saw complicated fission processes for
1281: interactions involving solitons with this charge. This is also the
1282: case in Q-ball/anti-Q-ball interactions as is illustrated in
1283: figure~\ref{e02}, which uses the same parameters as in
1284: figure~\ref{e03} except that the incident velocity is now much higher,
1285: $v=0.6$.  The figures shown are at times
1286: $t=0,10,15,25,30,35,40,45,50.$  One can see, after some partial
1287: annihilation of charge, that the fission of the incident Q-balls takes
1288: place in the direction perpendicular to the line of incidence and that
1289: now there is an effective bounce back of the incident solitons with a
1290: reduced charge.
1291: 
1292: These processes are also prevalent in both one and three
1293: dimensions. By way of illustration we have also included two examples
1294: in three dimensions. In figure~\ref{3danti3}  the solitons initially
1295: at $\pm(15,0,0)$, are Lorentz boosted together with a velocity $v=0.3$ and
1296: have $\omega_1=-\omega_2=1.5$. In this particular interaction there
1297: appears to be very little annihilation and the solitons effectively
1298: pass through each other. Figure~\ref{3danti2} has the same initial
1299: configuration, except that the solitons are boosted together with an
1300: initial velocity of $v=0.6$. The subsequent interaction is complicated
1301: involving first the formation of two loops, which is the equivalent of
1302: the right-angled fission observed in two dimensions, and then what
1303: appears to be almost total annihilation once the loops collapse. Two
1304: small Q-balls are emitted along the line of incidence, along with much
1305: radiation. 
1306: 
1307: \begin{figure}
1308: \begin{center}
1309: \leavevmode \ \vskip -3cm \hskip 0.5cm
1310:  \epsfxsize=12cm\epsffile{e04.ps}
1311: %\vskip -6cm
1312: \caption{Q-ball/anti-Q-ball interactions in two dimensions. The two
1313: solitons are placed initially at $\pm(6,0)$, are Lorentz boosted
1314: together with a velocity $v=0.3$ and have
1315: $\omega_1=-\omega_2=1.8$. Partial annihilation of the Q-balls takes
1316: place during the interaction and the momentum of the incident Q-ball is
1317: sufficient to take it through to the other side.}
1318: \label{e04}
1319: \end{center}
1320: \end{figure}
1321: 
1322: \begin{figure}
1323: \begin{center}
1324: \leavevmode \ \vskip -3cm \hskip 0.5cm
1325: \epsfxsize=12cm\epsffile{e03.ps}
1326: \caption{As figure~\ref{e04} but the charge has been increased so that
1327: $\omega_1=-\omega_2=1.5$. Complete annihilation takes place during a
1328: complicated oscillatory interaction.}
1329: \label{e03}
1330: \end{center}
1331: \end{figure}
1332: 
1333: \begin{figure}
1334: \begin{center}
1335: \leavevmode \ \vskip -3cm \hskip 0.5cm
1336: \epsfxsize=12cm\epsffile{e02.ps}
1337: \caption{As figure~\ref{e03} but with an incident velocity of
1338: $v=0.6$. Notice that fission takes place in the plane perpendicular to
1339: the line of incidence.}
1340: \label{e02}
1341: \end{center}
1342: \end{figure}
1343: 
1344: \begin{figure}
1345: \begin{center}
1346: \leavevmode \ \vskip -3cm \hskip 0.5cm
1347: \epsfxsize=12cm\epsffile{anti3.ps}
1348: \caption{Q-ball/anti-Q-ball interaction in three dimensions. Included
1349: are isosurfaces of the modulus of the charge density and a slice of
1350: the charge density itself through the centre of the interaction
1351: region. The two solitons are initially at $\pm(10,0,0)$, Lorentz
1352: boosted together with a velocity $v=0.3$ and have
1353: $\omega_1=-\omega_2=1.5$. The two solitons pass through each other
1354: with very little annihilation.}
1355: \label{3danti3}
1356: \end{center}
1357: \end{figure}
1358: 
1359: \begin{figure}
1360: \begin{center}
1361: \leavevmode
1362: \ \vskip -3cm \hskip 0.5cm
1363: \epsfxsize=12cm\epsffile{anti2.ps}
1364: \caption{As figure~\ref{3danti3} but with an incident velocity of
1365: $v=0.6$. Two loops are formed in the initial interaction, which
1366: subsequently collapse, leading to almost total annihilation.}
1367: \label{3danti2}
1368: \end{center}
1369: \end{figure}
1370: 
1371: \section{Summary and conclusions}\news\ \ \ \ \ \
1372: 
1373: We have identified the key parameters in the interactions of Q-balls to
1374: be the relative phase, the incident velocity and the charge of Q-balls,
1375: with the resulting interactions being strongly dependent on these parameters.
1376: The generic interaction involves attraction if the relative phase
1377: $\alpha=0$ and repulsion if $\alpha=\pi$ when the two Q-balls have the
1378: same charge. If they have an initial relative phase other than
1379: $\alpha=0$ or $\alpha=\pi$, or if the charges of the Q-balls are
1380: different, the dynamics of the Q-balls results in 
1381: charge transfer, a phenomena
1382: which is analogous to that observed in discrete breather
1383: systems (see, for example, ref.~\cite{ams} and references therein),
1384: although this behaviour often has to be induced by making the solitons
1385: come together via a Lorentz boost. With no Lorentz boost such breather
1386: systems naturally repel and this is also seen after the charge
1387: transfer process is complete. If the incident velocity is extremely
1388: high Q-balls can be made to pass through each other with very little
1389: interaction taking place.
1390: 
1391: This picture is almost independent of
1392: the number of dimensions if the charge of the Q-balls is small (for
1393: example, $\omega\approx 1.6$, but if the charge is much larger
1394: fission can take place during the interaction process due to the
1395: compression of the charge. In one dimension this process results in
1396: the emission of Q-balls during the slow interaction of two large
1397: Q-balls. In higher dimensions complicated, but analogous,
1398: phenomena are observed in which fission takes place in the direction
1399: perpendicular to the line of incidence. This leads, under specialized
1400: circumstances to the right angled fission of Q-balls in two dimensions
1401: and the production of loops in three. These fission effects can also
1402: be coupled with those of attraction, repulsion and charge transfer to
1403: produce complicated compound phenomena.
1404: 
1405: Interestingly, the naive expectation that an Q-ball/anti-Q-ball should
1406: annhilate into radiation is modified by the complicated
1407: breather-type interactions which take place. Since the difference
1408: between the two charges is large this can be thought of a special case
1409: of the charge transfer process, leading to the phenomenon of partial
1410: charge annihilation, the case of complete annihilation being very
1411: special. Fission of the Q-ball and anti-Q-ball can also take
1412: place if the charge is high.
1413: 
1414: Our original motivation was to understand the
1415: microphysical interactions of Q-balls formed by the Affleck-Dine
1416: mechanism for baryogenesis within the framework of the MSSM. The
1417: potential that we have concentrated on in this paper is different to
1418: that expected in the MSSM, but as we have pointed out the main
1419: interaction processes that we have identified are related to the
1420: phase. Therefore, we expect our result to be qualitatively independent
1421: of potential, and hence our results have some bearing on this
1422: case. We have repeated several of the simulations described in this
1423: paper for a potential of type II and found the same qualitative results.
1424: The next step in this research is an analytic description of the
1425: dynamics in terms of a slow manifold approach~\cite{BMS}, before their application to the
1426: problem of Q-ball formation in the Early Universe.
1427: 
1428: 
1429: 
1430: \def\jnl#1#2#3#4#5#6{\hang{#1 [#2], {\it #4\/} {\bf #5}, #6.} }
1431: \def\jnldot#1#2#3#4#5#6{\hang{#1 [#2], {\it #4\/} {\bf #5}, #6} }
1432: \def\jnltwo#1#2#3#4#5#6#7#8{\hang{#1 [#2], {\it #4\/} {\bf #5},
1433: #6;{\bf #7} #8.} }
1434: \def\prep#1#2#3#4{\hang{#1 [#2],`#3', #4.} } 
1435: \def\proc#1#2#3#4#5#6{{#1 [#2], in {\it #4\/}, #5, eds.\ (#6).} }
1436: \def\book#1#2#3#4{\hang{#1 [#2], {\it #3\/} (#4).} }
1437: \def\jnlerr#1#2#3#4#5#6#7#8{\hang{#1 [#2], {\it #4\/} {\bf #5}, #6.
1438: {Erratum:} {\it #4\/} {\bf #7}, #8.} }
1439: \def\prl{Phys.\ Rev.\ Lett.}
1440: \def\pr{Phys.\ Rev.}
1441: \def\pl{Phys.\ Lett.}
1442: \def\np{Nucl.\ Phys.}
1443: \def\prp{Phys.\ Rep.}
1444: \def\rmp{Rev.\ Mod.\ Phys.}
1445: \def\cmp{Comm.\ Math.\ Phys.}
1446: \def\mpl{Mod.\ Phys.\ Lett.}
1447: \def\apj{Ap.\ J.}
1448: \def\apjl{Ap.\ J.\ Lett.}
1449: \def\aap{Astron.\ Ap.}
1450: \def\cqg{Class.\ Quant.\ Grav.} 
1451: \def\grg{Gen.\ Rel.\ Grav.}
1452: \def\mn{M.$\,$N.$\,$R.$\,$A.$\,$S.}
1453: \def\ptp{Prog.\ Theor.\ Phys.}
1454: \def\jetp{Sov.\ Phys.\ JETP}
1455: \def\jetpl{Sov.\ Phys.\ JETP Lett.}
1456: \def\jmp{J.\ Math.\ Phys.}
1457: \def\zpc{Z.\ Phys.\ C}
1458: \def\ijmp{Int.\ J.\ Mod.\ Phys.}
1459: \def\cupress{Cambridge University Press}
1460: \def\oup{Oxford University Press}
1461: \def\pup{Princeton University Press}
1462: \def\wss{World Scientific, Singapore}
1463: 
1464: \begin{thebibliography}{99}
1465: 
1466: \bibitem{vort}
1467: See for example
1468: \jnl{E.P.S. Shellard}{1987}{}{\np}{B283}{624}\jnl{E.P.S. Shellard and
1469: P.J. Ruback}{1988}{}{\pl}{209B}{262}
1470: 
1471: \bibitem{mon}
1472: See for example
1473: \jnldot{P.M. Sutcliffe}{1997}{}{Int. J. Mod. Phys}{A12}{4663} and
1474: references therein.
1475: 
1476: \bibitem{sky}
1477: See for example \jnl{M.F. Atiyah and
1478: N.S. Manton}{1993}{}{\cmp}{153}{391}\jnl{R.A. Battye and
1479: P.M. Sutcliffe}{1997}{}{\pl}{391B}{366}\jnl{R.A. Battye and
1480: P.M. Sutcliffe}{1997}{}{\prl}{79}{363}\jnl{C.J. Houghton, N.S. Manton
1481: and P.M. Sutcliffe}{1998}{}{\np}{510}{507}
1482: 
1483: \bibitem{lee}
1484: \book{T.D. Lee}{1981}{Particle physics and introduction to field
1485: theory}{Harwood, London}
1486: 
1487: \bibitem{coleman}
1488: \jnl{S. Coleman}{1985}{}{\np}{B262}{263}
1489: 
1490: \bibitem{lowtemp}
1491: \jnl{D.K. Hong}{1998}{}{J. Low Temp. Phys.}{71}{483}
1492: 
1493: \bibitem{k2}
1494: \jnl{A. Kusenko}{1997}{}{\pl}{405B}{108}
1495: 
1496: \bibitem{AD}
1497: \jnl{I. Affleck and M. Dine}{1985}{}{\pl}{B249}{361}
1498: 
1499: \bibitem{KS}
1500: \jnl{A. Kusenko and
1501: M. Shaposhnikov}{1998}{}{\pl}{418B}{46}\jnl{A. Kusenko, V. Kuzmin and
1502: M. Shaposhnikov}{1998}{}{\prl}{80}{3185}
1503: 
1504: \bibitem{EM}
1505: \jnl{K. Enqvist and
1506: J. MacDonald}{1998}{}{\pl}{425B}{309}\jnl{K. Enqvist and
1507: J. MacDonald}{1999}{}{\np}{B538}{321}\jnl{K. Enqvist and
1508: J. MacDonald}{1998}{}{\prl}{81}{3071}
1509: 
1510: \bibitem{BMS}
1511: \prep{R.A. Battye, R.S. Mackay and P.M. Sutcliffe}{2000}{An effective
1512: Hamiltonian for Q-ball dynamics}{In preparation}
1513: 
1514: \bibitem{lit}
1515: \jnl{J.K. Drohm, L.P. Yok, Y.A. Simonov, J.A. Tyon and
1516: V.I. Veselov}{1981}{}{\pl}{101B}{204}
1517: \jnl{T.I. Belova and A.E. Kudryavtsev}{1989}{}{Sov. Phys. JETP}{68}{7}
1518: \jnl{G. Jaroszkiewcz}{1995}{}{J. Phys.}{G21}{501}
1519: 
1520: \bibitem{greek}
1521: \prep{M. Axenides, S. Komineas, L. Perivolopoulos and
1522: M. Floratos}{1999}{Dynamics of non-topological solitons}{hep-ph/9910388}
1523: 
1524: \bibitem{k1}
1525: \jnl{A. Kusenko}{1997}{}{\pl}{404B}{285}
1526: 
1527: \bibitem{St}
1528: \prep{D.M.A Stuart}{1999}{Modulational approach to the stability of
1529: non-topological solitons in semi-linear wave equations}{preprint
1530: DAMTP}
1531: 
1532: \bibitem{BS}
1533: \jnl{R.A. Battye and E.P.S. Shellard}{1994}{}{\np}{B423}{260}
1534: 
1535: \bibitem{Bat}
1536: \prep{R.A. Battye}{1995}{String radiation, interactions and
1537: cosmological constraints}{PhD Thesis (University of Cambridge)}
1538: 
1539: \bibitem{ams}
1540: \prep{T. Ahn, R.S. MacKay, J.A. Sepulchre}{1999}{Dynamics of relative
1541: phases in generalised multi-breathers}{Preprint DAMTP}
1542: 
1543: \bibitem{MV}
1544: \prep{T. Multam\"aki and I. Vilja}{1999}{Analytic and numerical
1545: properties of Q-balls}{hep-ph/9908446}
1546: 
1547: \bibitem{BSut}
1548: \prep{R.A. Battye and P.M. Sutcliffe}{2000}{Q-rings and loops}{In preparation}
1549: 
1550: \end{thebibliography}
1551: 
1552: \end{document}
1553: 
1554: