0811.3966/u3.tex
1: \documentclass[a4paper]{iopart}
2: \usepackage{iopams}
3: \usepackage{graphicx}% Include figure files
4: \usepackage{psfrag}
5: \usepackage{dcolumn}% Align table columns on decimal point
6: \usepackage{slashbox,multirow}
7: \usepackage{amsthm,amssymb}
8: \usepackage{cases,color}
9: \usepackage{mathrsfs}
10: \usepackage{times}
11: \setlength{\unitlength}{1mm}
12: 
13: \DeclareSymbolFontAlphabet{\mathrsfs}{rsfs} \DeclareMathAlphabet{\mathcal}{OMS}{cmsy}{m}{n}
14: \newcommand{\scri}{\mathrsfs{I}}
15: \newcommand{\izero}{\imath^0}
16: \newcommand{\const}{\text{const}}
17: \newcommand{\be}{\begin{equation}}
18: \newcommand{\ee}{\end{equation}}
19: \providecommand{\todo}[1]{\textsl{\textbf{TODO:}#1}} \marginparsep1mm \marginparwidth2.1cm
20: \providecommand{\margin}[1]{\marginpar{\scriptsize{\textsl{#1}}}}
21: \renewcommand{\figurename}{Figure}
22: 
23: \def\tet{{\tilde{t}}}
24: \def\ter{{\tilde{r}}}
25: \def\teg{{\tilde{g}}}
26: 
27: \newcommand{\old}[1]{{\color{green}$\blacksquare$~\textsf{[old: #1]}}}
28: \newcommand{\new}[1]{{\color{blue}$\blacksquare$~\textsf{[new: #1]}}}
29: 
30: \begin{document}
31: 
32: \title{Universality of global dynamics for the cubic wave equation}
33: 
34: \author{Piotr Bizo\'n$^{1}$ and An{\i}l Zengino\u{g}lu$^{2,3}$}
35: 
36: 
37: \address{$^{1}$ M. Smoluchowski Institute of
38:  Physics, Jagiellonian University, Krak\'ow, Poland}
39: 
40: \address{$^{2}$ Max-Planck-Institut f\"ur Gravitationsphysik
41:  (Albert-Einstein-Institut), Golm, Germany}
42: 
43: \address{$^{3}$ Department of Physics, and Center for Scientific
44:  Computation and Mathematical Modeling, University of Maryland,
45:  College Park, MD 20742, USA}
46: 
47: %
48: \date{\today}
49: %
50: \begin{abstract}
51: We consider the initial value problem for the spherically symmetric, focusing cubic wave equation
52: in three spatial dimensions. We give numerical and analytical evidence for the existence of a
53: universal attractor which encompasses both global and blowup solutions. As a byproduct we get an
54: explicit description of the critical behavior at the threshold of blowup.
55: \end{abstract}
56: 
57: %\pacs{Valid PACS appear here}% PACS, the Physics and Astronomy
58:                             % Classification Scheme.
59: %\keywords{Suggested keywords}%Use showkeys class option if keyword
60:                              %display desired \maketitle
61: 
62: \section{Introduction} We consider a  semilinear wave equation in three spatial dimensions with
63: the focusing cubic nonlinearity
64: \begin{equation}\label{eqo}
65: \partial_{tt} \phi - \Delta\phi- \phi^3=0\,.
66: \end{equation}
67: Heuristically, the dynamics of solutions of this equation can be viewed as a competition between the
68: Laplacian  which tends to disperse the waves and the nonlinearity which tends to concentrate the
69: waves. For small initial data the dispersion "wins" leading to global solutions which decay to
70: zero as $t\rightarrow \infty$. For large initial data the dispersive spreading is too weak to
71: counterbalance the focusing nonlinearity and solutions blow up in  finite time. For each of these
72: generic evolutions the leading asymptotic behavior  is known: small solutions decay as $1/t^2$ at
73: timelike infinity \cite{ch1,nik}, while large solutions diverge as $\sqrt{2}/(T-t)$ for $t$
74: approaching a blowup time $T$ \cite{mz1,bct}. The dichotomy of dispersion and blowup brings up
75: the question of what determines the boundary between these two behaviors and, in particular, what
76: is the evolution of critical initial data which lie on the boundary.
77: During numerical investigations of this question we observed  that for a
78: large set of initial data the solutions rapidly converge to a universal attractor which is given
79: by a two-parameter family of explicit solutions of equation (\ref{eqo}). The aim of this paper is
80: to give analytic and numerical evidence for this behavior (which is rather surprising for a
81: conservative wave equation). The description of the critical dynamics, which motivated our
82: investigations, emerges as a special case.
83: 
84: %
85: The paper is organized as follows. In section~2 we recall some basic properties of solutions of
86: equation (\ref{eqo}) and we formulate three conjectures about the existence of the attractor.
87: Analytic evidence for these conjectures is given in section~3. After a short description of our
88: numerical methods and tests in section~4, we present numerical evidence for the conjectures in
89: section~5. Finally, in section~6 we make some general remarks.
90: 
91: \section{Preliminaries and conjectures}
92: In this paper we restrict our attention to spherically symmetric solutions $\phi=\phi(t,r)$, so
93: equation (\ref{eqo}) reduces to
94: \begin{equation}\label{eq}
95: \partial_{tt} \phi - \partial_{rr}\phi-\frac{2}{r}\partial_r \phi - \phi^3=0\,.
96: \end{equation}
97: This equation  is invariant under the following transformations:
98: \begin{itemize}
99: \item translation in time $T_a$ by a constant $a$
100: \begin{equation}\label{trans}
101: T_a:    \phi(t,r)\, \rightarrow \, \phi(t+a,r)\,,
102: \end{equation}
103: \item scaling $S_{\lambda}$ by a positive constant $\lambda$
104: \begin{equation}\label{scal}
105: S_\lambda:    \phi(t,r)\, \rightarrow \, \frac{1}{\lambda}
106:    \phi\left(\frac{t}{\lambda},\frac{r}{\lambda}\right)\,,
107: \end{equation}
108: \item conformal inversion $I$ (which is an involution)
109: \begin{equation}\label{inv}
110: I:    \phi(t,r)\, \rightarrow \, \frac{1}{t^2-r^2} \, \phi\left(\frac{t}{r^2-t^2},
111:    \frac{r}{t^2-r^2}\right)\,,
112: \end{equation}
113: \item reflection $\phi \rightarrow -\phi$.
114: \end{itemize}
115: %
116: \noindent Neglecting the Laplacian in (\ref{eqo}) and solving the ordinary differential equation
117: $\partial_{tt}\phi-\phi^3=0$, one gets the one-parameter family of spatially homogeneous solutions
118: \begin{equation}\label{u0}
119:    \phi_b=\frac{\sqrt{2}}{b-t}\,.
120: \end{equation}
121: %
122: This family is a special case of the two-parameter family of solutions \cite{anco}
123: \begin{equation}\label{orbit}
124:    \phi_{(a,b)}(t,r)=\frac{\sqrt{2}}{t+a+b\left((t+a)^2-r^2\right)}\,,
125: \end{equation}
126: which can be obtained from (\ref{u0}) by the action of conformal inversion $I$ followed by the
127: time translation $T_a$.
128: The scaling transformation $S_{\lambda}$ acting on
129: $\phi_{(a,b)}$ only rescales the parameters of the solution without changing its form.
130: Note
131: that the solution $\phi_{(a,b)}(t,r)$ is singular on the two-sheeted hyperboloid $t= -a -1/(2b)\pm
132: \sqrt{1/(4b^2)+r^2}$.
133: 
134: The main result of this paper is the observation that the family (\ref{orbit}) is an attractor
135: for a large set of initial data. More precisely, we have the following conjectures about the
136: forward-in-time behavior of solutions of equation (\ref{eq}) starting from smooth, compactly
137: supported (or suitably localized) initial data (by time reflection symmetry analogous conjectures
138: can be formulated for the backward-in-time behavior):
139: %
140: \vskip 0.2cm \noindent{\emph{Conjecture~1.}} For any generic globally regular solution
141: $\phi(t,r)$ there exist parameters $(a,b)\in \mathbb{R}\times \mathbb{R}^+$ and $\kappa=\pm 1$
142: such that $t^4 \left(\phi(t,r)-\kappa\, \phi_{(a,b)}(t,r)\right)$ is bounded for all finite $r$
143: and $t\rightarrow \infty$.
144: %
145: \vskip 0.2cm \noindent{\emph{Conjecture~2.}}
146: For any solution $\phi(t,r)$ which blows up at the origin in finite time $T$,
147:  there exist parameters $(a,b)\in \mathbb{R}\times \mathbb{R}^-$ and $\kappa=\pm 1$ such that
148:    $(T-t)^{-2} \left(\phi(t,r)-\kappa\, \phi_{(a,b)}(t,r)\right)$
149: is bounded for $t\rightarrow T^-$ inside the past light cone of the blowup point.
150: %
151: \vskip 0.2cm \noindent{\emph{Conjecture~3.}} The borderline between dispersive and blowup
152: solutions, described respectively in Conjectures~1 and~2, consists of codimension-one globally
153: regular solutions $\phi(t,r)$ for which there exist a parameter $a\in \mathbb{R}$ and $\kappa=\pm
154: 1$ such that
155:    $t^4 \left(\phi(t,r)-\kappa\, \phi_{(a,0)}(t,r)\right)$
156: is bounded for all finite $r$ and $t\rightarrow \infty$. \vskip 0.2cm \noindent The parameters
157: $(a,b)$ for which the above assertions hold will be referred to as optimal. \vskip 0.2cm
158: 
159: \noindent \emph{Remark~1:} Conjectures 1 and 2 are refinements of the well-known asymptotic
160: behavior of solutions of equation (\ref{eq}): $t^{-2}$ decay near timelike infinity in the case
161: of global regularity (for small initial data) and $\sqrt{2}/(T-t)$ growth in the case of blowup
162: (for large initial data), respectively.
163: 
164: \noindent \emph{Remark 2:} Although Conjecture~3 is a special limiting case
165: of Conjecture~1, we state it as a separate conjecture to emphasize the critical character
166: of the attractor solution $\phi_{(a,0)}$. Note that the slowly decaying global solutions
167: described in Conjecture~3 are not asymptotically free, i.e., they do not scatter.
168: 
169: \noindent \emph{Remark~3:} The genericity condition in Conjecture~1 is essential
170: because, as we shall see below, there do exist non-generic very rapidly
171: decaying globally regular solutions which do not converge to the attractor $\phi_{(a,b)}$.
172: 
173: \noindent \emph{Remark~4:} For very large amplitudes the blowup first occurs on a sphere
174: \cite{bct} and then Conjecture~2 does not hold.
175: 
176: In the remainder of the paper we give evidence for the above conjectures. The evidence is based
177: mainly on numerical simulations, however, before presenting numerics, we will give two analytic
178: arguments: one based on linearized stability analysis and one based on an explicit solution.
179: 
180: \section{Analytic evidence}\label{sec:3}
181: %
182: \subsection{Linearized stability}
183: In this section we discuss the linearized stability of the attractor solution
184: $\phi_{(a,b)}(t,r)$. Since linearization commutes with symmetries, we may set $a=b=0$ without
185: loss of generality. We denote the resulting solution by $\phi_0$, thus $\phi_0=\sqrt{2}/t$. In
186: order to determine the spectrum of small perturbations around this forward self-similar solution
187: we will use the symmetry under conformal inversions. To this end, consider the region $I^+(O)$,
188: the interior of the future light cone of the origin $(0,0)$ of the Minkowski spacetime, and its
189: foliation by spacelike hyperboloids (in this paragraph we
190:  follow Christodoulou \cite{ch2})
191: \begin{equation}\label{hyp}
192:    t=c+\sqrt{c^2+r^2}\,,
193: \end{equation}
194: where $c$ is a positive constant. The hyperboloid (\ref{hyp}) is asymptotic to $\partial I^+(c)$,
195: the future light cone of the point $(c,0)$. Let us make the conformal inversion
196: \begin{equation}\label{ci}
197:    I: (t,r) \mapsto (\bar t,\bar
198:    r)=\left(\frac{t}{r^2-t^2},\frac{r}{t^2-r^2}\right)\,.
199: \end{equation}
200: This transformation maps $I^+(O)$ (in the original coordinate system) to $I^-(O)$ (the interior
201: of the past light cone of the origin in the barred coordinate system). In particular, the
202: hyperboloids (\ref{hyp}) are mapped to spacelike hyperplanes
203: \begin{equation}\label{hplanes}
204:    \bar t = - \bar c\,, \qquad \mbox{where} \qquad \bar c=\frac{1}{2c}\,.
205: \end{equation}
206: %
207: 
208: Now, consider a global-in-time solution $\phi(t,r)$ inside $I^+(O)$. In the barred coordinates we
209: have from (\ref{inv})
210: \begin{equation}\label{phib}
211:    \bar \phi(\bar t,\bar r)=(t^2-r^2) \phi(t,r)\,.
212: \end{equation}
213: It follows from the above paragraph that the study of the asymptotics of $\phi(t,r)$ for
214: $t\rightarrow \infty$ is equivalent to the study of the asymptotics of $\bar\phi(\bar t,\bar r)$
215: for $\bar t \rightarrow 0^-$. In particular, there is equivalence between linearized
216: perturbations of $\phi_0=\sqrt{2}/t$ for $t\rightarrow \infty$ and linearized perturbations of
217: $\bar \phi_0=\sqrt{2}/(-\bar t)$ for $\bar t\rightarrow 0^-$. But the latter have already been
218: determined in \cite{bct,gp}. Namely, it has been shown there that the spectrum of smooth linearized
219: perturbations about the backward self-similar solution $\sqrt{2}/(-\bar t)$ consists of a
220: discrete set of eigenmodes of the form $(-\bar t)^{\lambda_n} \xi_n(y)$, where
221: $y=\bar r/(-\bar t)$ and $\lambda_0=-2,\lambda_1=0,\lambda_n=n \,(n\geq 2)$. More precisely, we have the
222: following eigenmode expansion around $\bar \phi_0$
223: \begin{equation}\label{bct}
224:  \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!  \delta\bar \phi(\bar t,\bar r)=c_0 (-\bar t)^{-2}
225:    +c_1 (1-y^2) +c_2 (-\bar t)^{2}
226:    (1-\frac{2}{3}y^2+\frac{1}{5}y^4)+
227:    \mathcal{O}((-\bar t)^3)\,.
228: \end{equation}
229: By (\ref{phib}), this translates into the following eigenmode expansion around $\phi_0$ (where
230: now $y=r/t$)
231: \begin{equation}\label{exp2}
232: \!\!\!\!\!\! \delta\phi(t,r)= c_0 (1-y^2)+c_1 \frac{1}{t^2} +
233:    c_2 \frac{1}{t^4} \frac{1-2y^2/3+y^4/5}{(1-y^2)^3}+
234:    \mathcal{O}(1/t^5)\,.
235: \end{equation}
236: The first two eigenmodes in the expansions (\ref{bct}) and (\ref{exp2}) correspond to the perturbations
237: along the symmetry
238: orbits $\bar \phi_{(a,b)}(\bar t,\bar r)$ and $\phi_{(a,b)}(t,r)$ respectively. For example, we have
239: \begin{equation}\label{gena}
240: \!\!\!\!\!\!\!\!\! \frac{\partial}{\partial b}
241:    \phi_{(a,b)}(t,r) \vert_{(a=0,b=0)}
242:    \sim 1-y^2\,,\quad
243:    \frac{\partial}{\partial a} \phi_{(a,b)}(t,r)
244:    \vert_{(a=0,b=0)} \sim \frac{1}{t^2}\,.
245: \end{equation}
246: The choice of the optimal parameters $(a,b)$ for the attractor amounts to tuning away the
247: coefficients of the symmetry modes $c_0$ and $c_1$, hence the rate of convergence to the
248: attractor is expected to be governed by the third eigenmode in the expansions (\ref{bct}) and
249: (\ref{exp2}):
250: \begin{equation}\label{conver0}
251:    \bar \phi(\bar t,\bar r)-\bar \phi_{(a,b)}(\bar t,\bar r) \sim (-\bar t)^2 \qquad \mbox{for}
252:    \,\,\bar t\rightarrow 0^-\,,
253: \end{equation}
254: and
255: \begin{equation}\label{conver}
256:    \phi(t,r)-\phi_{(a,b)}(t,r) \sim 1/t^4 \qquad \mbox{for}
257:    \,\,t\rightarrow \infty\,,
258: \end{equation}
259: which is consistent with our conjectures. Below we will verify this expectation numerically, but
260: first we want to give a simple example which corroborates  (\ref{conver}).
261: %
262: \subsection{The conformal solution}
263: %
264: Equation (\ref{eq}) has the following globally regular explicit solution
265: \begin{equation}\label{A2}
266:    \phi_{conf}(t,r)= \frac{2}{\sqrt{(1+(t-r)^2)(1+(t+r)^2)}}\,,
267: \end{equation}
268: which we will refer to  as the conformal solution. Note that this solution corresponds to
269: time symmetric initial data
270: \begin{equation}\label{ic}
271:    \phi(0,r)=\frac{2}{1+r^2}\,,\qquad \partial_t \phi(0,r)=0\,.
272: \end{equation}
273: The conformal solution can be easily found in the framework of conformal compactification which
274: maps Minkowski space with the flat metric $\eta$ into the Einstein universe with the conformal
275: metric $g=\Omega^{2} \eta$, where the conformal factor $\Omega=2[(1+(t-r)^2)(1+(t+r)^2)]^{-1/2}$
276: \cite{Penrose65}.
277: Under this mapping the cubic wave equation $\Box \phi+\phi^3=0$ transforms to $\Box_{g} \Phi
278: -\Phi+\Phi^3=0$, where $\Phi=\Omega^{-1}\phi$. The trivial constant solution $\Phi=1$ gives
279: $\phi_{conf}=\Omega$.
280: 
281: By elementary calculation we find that for the conformal solution  the optimal parameters of the
282: attractor $\phi_{(a,b)}$ are $(a,b)=(-1/\sqrt{2},1/\sqrt{2})$. More precisely, we have for
283: $t\rightarrow\infty$
284: \begin{equation}\label{a2v}
285:    \phi_{conf}(t,r)-\phi_{(-\frac{1}{\sqrt{2}},\frac{1}{\sqrt{2}})}(t,r)
286:    = -\frac{1}{t^4}\, \frac{3+y^2}{(1-y^2)^3} +
287:    \mathcal{O}(t^{-6})\,,
288: \end{equation}
289: %
290: in agreement with Conjecture~1.
291: 
292: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
293: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
294: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
295: 
296: \section{The numerical method and tests}
297: \subsection{The hyperboloidal initial value problem for the cubic wave equation}
298: In order to study the asymptotic behavior of solutions and verify the conjectured convergence to
299: the attractor, it is convenient to foliate Minkowski spacetime by hyperboloids. A time coordinate
300: $\tau$ adapted to such a foliation can be written as \be\label{eq:tau} \tau = t -
301: \sqrt{\frac{9}{K^2}+r^2}.\ee The level surfaces of $\tau$ are standard hyperboloids shifted in
302: the time direction. They have constant mean curvature $K$ which is a free parameter. In the
303: following we set $K=3$ for simplicity. A foliation of Minkowski spacetime by level sets of $\tau$
304: is depicted in Figure \ref{fig:1}.
305: 
306: \begin{figure}[t]
307:  \begin{centering}
308:    {\psfrag{t}{$t$} \psfrag{r}{$r$}
309:      \includegraphics[width=0.31\textwidth,height=0.17\textheight]{figures/figure1}
310:      \hspace{23mm}} {\psfrag{ip}{$i^+$} \psfrag{i0}{$i^0$}
311:      \psfrag{scrp}{$\scri^+$}
312:      \includegraphics[width=0.27\textwidth]{figures/figure2a}} \caption{
313:      The future domain of a $t=0$ surface is partially depicted on
314:      the left panel in standard coordinates and entirely represented
315:      on the right panel in a Penrose diagram
316:      \cite{Penrose65,Hawking73,Zeng07}.  Dashed lines are level sets of $t$,
317:      solid lines are hyperboloids shifted in time as given by
318:      (\ref{eq:tau}), the thick straight lines depict outgoing
319:      characteristics indicating the location of the approximate
320:      wavefront. \label{fig:1}}
321:  \end{centering}
322: \end{figure}
323: 
324: In order to be able to analyze the propagation of the outgoing waves to infinity we
325: introduce a compactifying radial coordinate $\rho$ along the surfaces of our foliation. For the
326: regularity of our equations in this compactified setting, we perform a suitable conformal
327: rescaling of the metric. The rescaling factor denoted by $\Omega$ must be a function of $\rho^2$
328: to ensure regularity at the origin in the conformal manifold. We choose $\Omega=(1-\rho^2)/2$
329: following \cite{Moncrief00, Husa02, Fodor03}. The compactifying coordinate is then chosen
330: according to $r=\rho/\Omega$.
331: 
332: We express the standard Minkowski metric $\eta$ in the new coordinates $(\tau,\rho)$ and rescale
333: it with the conformal factor $\Omega^2$ to obtain \be\label{eq:metric} g = \Omega^2\eta =
334: -\Omega^2\,d\tau^2 - 2\rho\,d\tau d\rho + d\rho^2 + \rho^2\,d\sigma^2, \ee where $d\sigma^2$ is
335: the standard metric on the unit two-sphere. In these coordinates the zero set of the conformal factor corresponds to future null infinity denoted by $\scri^+$. We rewrite the cubic wave equation on Minkowski
336: spacetime in conformally covariant form \be \label{eq:conf} \Box_g \Phi -\frac{1}{6} R[g]\, \Phi
337: + \Phi^3 = 0, \quad \mathrm{where} \quad \Phi = \frac{\phi}{\Omega}\,, \quad g=\Omega^2 \eta\,.
338: \ee
339: The Ricci scalar of the metric $g$ from (\ref{eq:metric}) appearing in the above equation is given by
340: \[R[g]=\frac{12\, (1-\rho^2)\, (3+\rho^2)}{(1+\rho^2)^3}\,.\]
341: With the auxiliary variables,
342: \[
343: \psi:=
344: \partial_\rho \Phi\quad \mbox{and}
345: \quad \pi := \frac{2}{1+\rho^2}(\partial_\tau\Phi+\rho\,\partial_\rho \Phi)\,,\]
346: we can rewrite the
347: system (\ref{eq:conf}) in first order symmetric hyperbolic form as
348: \begin{eqnarray}\label{eq:wave}
349: \partial_\tau \Phi &=&\frac{1+\rho^2}{2}\,\pi - \rho\,\psi, \nonumber
350: \\ \partial_\tau \psi &=&\partial_\rho\left(\frac{1+\rho^2}{2}\pi -
351: \rho\,\psi\right), \\ \partial_\tau \pi &=&
352: \frac{1}{\rho^2}\,\partial_\rho\left(\rho^2\left(\frac{1+\rho^2}{2}
353: \psi-\rho\,\pi\right)\right) +
354: \frac{1+\rho^2}{2}\left(\Phi^3-\frac{1}{6}\,R[g] \Phi \right)
355: \nonumber.
356: \end{eqnarray}
357: 
358: The initial data are arbitrary in our conjectures. In the numerical studies presented below we
359: choose a Gaussian pulse \be\label{eq:id} \Phi(0,\rho) = A\,e^{-(\rho-\rho_c)^2/\sigma^2}, \qquad
360: \partial_\tau \Phi(0,\rho) = 0, \ee  with fixed  parameters $\rho_c=0.3,~\sigma=0.07$, and
361: varying amplitude $A$. Tests for different initial data give the same qualitative results which
362: makes us feel confident that the phenomena described below are universal.
363: \subsection{The code}
364: We solve the hyperboloidal initial value problem (\ref{eq:wave},\ref{eq:id}) numerically
365: using 4th order Runge-Kutta integration in time
366: and 6th order finite differencing in space. At the origin we apply the regularity condition
367: $\psi(\tau,0) = 0$. No boundary conditions are needed at the outer boundary because there are no
368: incoming characteristics due to the compactification. One-sided finite differencing is applied on
369: the numerical boundaries.
370: 
371: \begin{figure}[t]
372:  \begin{centering}
373:    \psfrag{t}{$\tau$} \psfrag{Q}{$Q$}
374:    \includegraphics[width=0.42\textwidth]{figures/conv}\hspace{15mm}
375:    \includegraphics[width=0.42\textwidth]{figures/conv_long}
376:    \caption{Convergence factors in time measured in the
377:      $L_2$-norm. The convergence factor $Q$ is defined by
378:      \mbox{$Q=\log_2\frac{ \| \Phi^{low}-\Phi^{med}\|}{\|
379:          \Phi^{med}-\Phi^{high}\|}$}. On the left panel we see that
380:      after a short transient phase the code converges with 6th
381:      order. The dashed curve corresponds to a simulation with a 10
382:      times smaller Courant factor and shows 6th order convergence
383:      from the beginning. The right panel shows results of a long time
384:      convergence test and indicates loss of convergence after about
385:      $\tau=1000$. \label{fig:conv}}
386:  \end{centering}
387: \end{figure}
388: 
389: To test the code we performed a three level convergence study with
390: 200, 400 and 800 grid cells on the coordinate domain $\rho\in [0,1]$
391: that has infinite physical extent. For this study, we used an
392: amplitude of $A=2$ which is below the critical amplitude, and a
393: Courant factor of $\triangle \tau/\triangle \rho=0.8$. The convergence
394: factors shown in Figure \ref{fig:conv} indicate that, as expected, our
395: code is 6th order convergent after a short transient phase. The dashed
396: curve has been calculated using a much smaller Courant factor of
397: $0.08$ to show that the initial transient phase is due to numerical
398: errors in the time integration that converge with 4th order. The long
399: time convergence plot on the right panel in Figure \ref{fig:conv}
400: shows that convergence is lost at about $\tau=1000$ for the number of
401: cells used in this study. This is due to accumulation of numerical
402: errors and occurs at later times in tests with higher resolution. We
403: give numerical evidence for universal dynamics only in the convergent
404: regime. Two-level convergence tests using the explicit conformal
405: solution (\ref{A2}) give the same qualitative results.
406: 
407: Further tests can be performed using known properties of solutions to the cubic wave equation as
408: studied in \cite{bct}. First, it is known that generic small solutions decay as $t^{-2}$ near
409: timelike infinity and $t^{-1}$ along null infinity \cite{nik}. It is clear from (\ref{orbit})
410: that the attractor solution has this behavior for $b>0$. In order to accurately measure the
411: decay rate of the solution along null infinity, near timelike infinity and in the transition
412: domain between these two asymptotic regimes, we calculate the local power index, $p_\rho(\tau)$,
413: defined by
414: \[ p_\rho(\tau) = \frac{d\ln |\Phi(\tau,\rho)|}{d\ln \tau}. \]
415: Figure \ref{fig:tests} shows that our code reproduces the predicted
416: decay rates very accurately.
417: 
418: \begin{figure}[t]
419:  \begin{centering}
420:    {\psfrag{t}{$\tau$} \psfrag{p}{$p_\rho$}
421:      \includegraphics[width=0.41\textwidth]{figures/test_decay}\hspace{13mm}}
422:    {\psfrag{t}{$\frac{1}{(T-\tau)}$}\psfrag{phi}{$\Phi$}
423:      \includegraphics[width=0.41\textwidth]{figures/test_blowup}}
424:    \caption{Left panel: The local power index calculated along
425:      $\scri^+$, \mbox{$r=\{100,20\}$} and at the origin from top
426:      to bottom. Right panel: A numerical solution that blows up at
427:      the origin is depicted by small squares and the theoretical
428:      prediction of the blowup is depicted by the solid line on a
429:      log-log scale. The small squares are not distributed uniformly
430:      in time because we reduce our time steps while the solution
431:      grows. The figures indicate that the code reproduces the known
432:      decay rates for small solutions and the blowup rate for large
433:      solutions very accurately. \label{fig:tests}}
434:  \end{centering}
435: \end{figure}
436: 
437: Second, it is known that large initial data lead to blowup in finite
438: time with the blowup rate at the origin $\sqrt{2}/(T-t)$
439: \cite{mz1,bct}. This is in accordance with the attractor solution
440: (\ref{orbit}) for $b<0$. The right panel in Figure \ref{fig:tests}
441: shows the numerical solution at the origin for large data on a log-log
442: scale against the theoretical prediction of the blowup rate. The two
443: curves match over many orders of magnitude indicating that our code
444: can reliably handle the blowup.
445: \section{Numerical evidence for universal dynamics}
446: The space of solutions to the cubic wave equation can be divided into three parts: decay, blowup
447: and criticality. These parts correspond respectively to $b>0$, $b<0$, and $b=0$ for the attractor
448: solution (\ref{orbit}). We will follow this natural classification in our presentation of the
449: numerical evidence for the universality of dynamics.
450: 
451: In many cases, the numerical evidence will be presented in terms of the conformally rescaled
452: solution in the coordinates presented in the previous section. In these variables the
453: two-parameter family of solutions (\ref{orbit}) takes the form (using the abbreviation
454: $\tilde{\tau}=\tau+a$)
455: \begin{equation} \label{phiab_hypal} \!\!\!\!\!\!\!\Phi_{(a,b)}(\tau,\rho) =
456: \frac{2 \sqrt{2}}{(\tilde{\tau}+1) (b\, (\tilde{\tau}+1)+1)-\rho^2
457:  (\tilde{\tau}-1) (b\, (\tilde{\tau}-1)+1)}\,. \end{equation}
458: 
459: 
460: \subsection{Decay}
461: \subsubsection{Convergence to the attractor}
462: According to Conjecture~1, the difference between the attractor solution (with optimal
463: parameters) and a generic solution should decay as $t^{-4}$ for small data (\ref{conver}). This
464: behavior is confirmed numerically in Figure \ref{fig:decay}.
465: 
466: \begin{figure}[t]
467:  \begin{centering}
468:    \psfrag{t}{$\tau$} {\psfrag{phi}{$\|\Phi-\Phi_{(a,b)}\|_{L_2}$}
469:      \includegraphics[width=0.41\textwidth]{figures/decay}}\hspace{13mm}
470:    \psfrag{delta}{$\delta$}
471:    \includegraphics[width=0.41\textwidth]{figures/modulation}
472:    \caption{Left panel: The $L_2$-norm of the difference between a
473:      generic decaying numerical solution and the fitted attractor
474:      solution is depicted by the small squares on a log-log
475:      scale. The solid line, shown for comparison, has the slope $-4$, in accordance with Conjecture~1.
476:      Right panel: Modulation in the relative errors
477:      $\delta$ of the parameters $a$ (solid line) and $b$ (dashed
478:      line) on a log-log scale. The modulation for $a$ seems to decay
479:      as $t^{-1}$ and for $b$ as $t^{-2}$. The relative error, say
480:      $\delta_a$ for $a$, is defined as $\delta_a=|a(t)-a|/a$ where
481:      $a$ is the value of the parameter at the last time
482:      step of the numerical evolution. \label{fig:decay}}
483:  \end{centering}
484: \end{figure}
485: %
486: For this plot, we first fit the numerical solution to the attractor (\ref{phiab_hypal}) to
487: determine the optimal parameters $(a,b)$. We applied two methods for the fit. Fit in time, and
488: fit in space. For the fit in time we fix a grid point $\rho=\rho_f$, fit the solution
489: $\Phi(\tau,\rho_f)$ in $\tau$ and repeat the process for each grid point $\rho_f$. This gives us
490: a set of values $(a,b)_{\rho_f}$ depending on $\rho_f$. The average of these values over all grid
491: points gives the optimal parameters. For this method, the starting time for the fit needs to be
492: taken well behind the approximate wavefront because the convergence of the numerical solution to
493: the attractor family, and therefore the variation of $(a,b)_{\rho_f}$ over the grid, will depend
494: strongly on the foliation in early times. Note that our grid has infinite physical extent due to
495: compactification. The variation is expected to decrease in time which we have checked
496: numerically.  Choosing a large value of $\tau$ as the starting time of the time fit ensures
497: higher accuracy in the determination of the optimal parameters.
498: 
499: The second method that we applied is the fit in space in which we fix
500: a time $\tau=\tau_f$, fit the solution $\Phi(\tau_f,\rho)$ in $\rho$
501: and repeat the process on each time surface $\tau_f$. This gives us a
502: set of values $(a,b)_{\tau_f}$ depending on $\tau_f$. The resulting
503: time variation of the parameters is commonly referred to as
504: modulation. If our conjecture is correct, the modulation should be
505: small.
506: 
507: At the end, both methods should deliver the same values up to a small
508: error. This is used as a cross-check for the quality of fitting. The
509: modulation of the parameters $a$ and $b$ in time is shown on the right
510: panel in Figure \ref{fig:decay}. We observe that the accuracy of the
511: fitting in $b$ is better than in $a$. This is due to the fact that the
512: perturbation generated by changing $a$ decays in time while the
513: perturbation generated by changing $b$ does not (\ref{gena}), hence an
514: error in determining $b$ is easier to spot.
515: 
516: Once the optimal parameters have been determined, we compute the difference between the attractor
517: solution and the numerical solution at each time step. The difference $\Phi-\Phi_{(a,b)}$ at late
518: times has a
519:  strict sign for all values of $\rho$ on our numerical
520:  grid. Therefore the $L_2$-norm of this difference over the grid
521:  as a function of time gives a strong uniform measure of how close the
522:  solutions are to each other. Figure \ref{fig:decay} indicates that this
523: difference falls off as $t^{-4}$, as claimed in Conjecture~1.
524: 
525: \subsubsection{Exceptional solutions}
526: As mentioned above in Remark~3, not all globally regular solutions converge to the attractor
527: $\phi_{(a,b)}$.  The existence of exceptional solutions with different asymptotic behavior can be
528: seen as follows. Consider a one-parameter family of initial data, such as (\ref{eq:id}). It turns
529: out that along such a family typically there occurs a flip of sign of the attractor, that is,
530: there is a (subcritical) amplitude $A_f$, such that for $A<A_f$ the solutions converge to, say,
531: $\phi_{(a,b)}$ and for $A>A_f$ they converge to $-\phi_{(a,b)}$. Performing a bisection we may
532: easily fine-tune the amplitude to $A_f$ and generate a special solution (below referred to as the
533: flip solution) that is not an element of the attractor family. One can expect that the flip
534: solution will have a faster decay rate than the generic global solutions. This expectation is
535: confirmed in Figure \ref{fig:flip} which shows that the flip solution decays as $t^{-3}$ at
536: timelike infinity and as $t^{-2}$ along null infinity. By modifying the initial amplitude
537: slightly away from the flip solution, we can see that the decay rates of the generic solutions
538: are attained at late times after a transient phase. Note that by construction the flip solutions
539: correspond to codimension-one initial data. It is likely that there exist initial data of higher
540: codimensions which lead to globally regular solutions with even faster decay rates, however the
541: numerical construction of such solutions would be very difficult.
542: 
543: \begin{figure}[t]
544:  \begin{centering}
545:    \psfrag{t}{$\tau$} \psfrag{p}{$p_\rho$}
546:      \includegraphics[width=0.37\textwidth]{figures/flip}\hspace{13mm}
547:      \includegraphics[width=0.37\textwidth]{figures/flipma}
548:      \caption{On the left panel we plot decay rates for the flip
549:        solution with $A_f=2.4913$ along the surfaces $\scri^+$,
550:        \mbox{$r=\{100,33,14,0\}$} from top to bottom. On the right
551:        panel we plot the decay rates for a solution with initial
552:        amplitude $A_f-0.01$ along the same surfaces. Here, the
553:        generic decay rates are obtained after a much longer time than
554:        that in Figure \ref{fig:tests}. \label{fig:flip}}
555:  \end{centering}
556: \end{figure}
557: 
558: \subsection{Blowup}
559: \subsubsection{Convergence to the attractor}
560: Merle and Zaag proved in \cite{mz1} that the ODE solution $\sqrt{2}/(T-t)$ determines the
561: universal rate of blowup for  equation (\ref{eqo}), however the problem of profile of blowup
562: remains open.  Numerical simulations in spherical symmetry \cite{bct} showed that for a solution
563: which blows up at the origin, its deviation from $\sqrt{2}/(T-t)$ near the tip of the past light
564: cone of the blowup point is very well approximated by the second eigenmode in the expansion
565: (\ref{bct}) (see Figure 4 in \cite{bct})
566: \begin{equation}\label{bct2}
567:    \phi(t,r)-\frac{\sqrt{2}}{T-t} \approx c_1 \left(1-\frac{r^2}{(T-t)^2}\right)\,,
568: \end{equation}
569: however an error on the right hand side was not quantified in \cite{bct}. According to
570: Conjecture~2
571: the approximation (\ref{bct2}) may be improved to
572: \begin{equation}\label{conj22}
573:     \phi(t,r)-\phi_{(a,b)}(t,r) \approx  \mathcal{O}((T-t)^2)\,,
574: \end{equation}
575: provided that the parameters $(a,b)$ are optimal.
576: 
577: The numerical verification of the formula (\ref{conj22}) is shown in Figure~\ref{fig:st_blup}.
578: The data for this plot were produced using a code based on standard Minkowski coordinates.
579: 
580: \begin{figure}[t]
581:  \begin{centering}
582:    \psfrag{del}{$T-t$} \psfrag{dif}{$\phi-\phi_{(a,b)}$}
583:    \includegraphics[width=0.46\textwidth]{figures/standard_blowup}
584:      \caption{Blowup at the origin in standard coordinates on a
585:        log-log scale. The difference between the attractor solution
586:        and the numerical solution is depicted by small squares. The linear fit (solid line)
587:        gives the slope $2.02$, which confirms Conjecture~2.
588:        The deviation from the straight line seen at times very close
589:        to the blowup time is due to the fact that the domain of
590:        analysis extends beyond the past light cone of the blowup
591:        point.
592: \label{fig:st_blup}}
593:  \end{centering}
594: \end{figure}
595: 
596: We point out that there is no genericity condition in Conjecture~2. This might appear surprising
597: in view of existence of a countable family of regular self-similar solutions of equation
598: (\ref{eq}) \cite{bbmw}. It seems that these self-similar solutions do not play any role in the
599: Cauchy evolution, which is probably due to the fact that they contain singularities outside the
600: past light cone of the blowup point (see section 6 in \cite{bbmw}).
601: 
602: \subsubsection{Blowup surface}
603: %
604: For $b<0$ the attractor solutions blow up along a hyperboloid
605: \begin {equation}\label{blowatt}
606: t= -a -\frac{1}{2b} + \sqrt{\frac{1}{4b^2}+r^2}\,.
607: \end{equation}
608: This surface has the form (\ref{eq:tau}) with the mean extrinsic curvature $K=-6b$. Hence, for
609: our choice of the hyperboloidal foliation with $K=3$ the blowup surface of the attractor with
610: $b=-1/2$ coincides with one leaf of the foliation. Therefore, if Conjecture~2 is correct, in the
611: case of the blowup solution converging to the attractor with the optimal parameter $b=-1/2$ we
612: should observe an approximately simultaneous blowup along the whole grid. This expectation is
613: verified in Figure \ref{fig:globup}.
614: 
615: \begin{figure}[t]
616:  \begin{centering}
617:    \psfrag{r}{$\rho$} \psfrag{phi}{$\Phi$}
618:      \includegraphics[width=0.46\textwidth]{figures/globup}
619:      \caption{Simultaneous blowup along the numerical grid for fine
620:          tuned initial data. The solution is plotted on a log
621:        scale at various time steps close to the blowup time $T$. The
622:        solid lines depict an attractor solution with $b=-1/2$ at the
623:        corresponding times. The times are from bottom to top
624:        $\triangle T = T-\tau =
625:        \{0.34,0.1,0.02,0.007,0.003,0.0008\}$. \label{fig:globup}}
626:  \end{centering}
627: \end{figure}
628: 
629: Note that, for a given constant mean curvature foliation, the simultaneous blowup is not generic.
630: To produce data for Figure \ref{fig:globup} we used the dependence of $b$ on the amplitude $A$ of
631: the gaussian. For relatively small (but supercritical) values of $A$ we have $-1/2<b<0$ and the
632: blowup first occurs at null infinity, while for larger amplitudes we have $b<-1/2$ and the blowup
633: first occurs at the origin. Performing bisection between these two states
634: we fine-tuned to the blowup solution with $b=-1/2$.
635: 
636: We would like to emphasize that a rather counter-intuitive phenomenon of blowup at null infinity
637: is a mere coordinate effect. This only occurs if the mean extrinsic curvature of the blowup
638:  surface, given by $-6b$, is smaller then the mean extrinsic curvature
639:  of the hyperboloidal foliation, $K$, used in the numerical
640:  simulation.  Nevertheless, as discussed above, we can use this
641: effect to our advantage to probe the shape of the blowup surface. Note
642: also that one can predict the blowup at null infinity by fitting the
643: attractor to the numerical solution at the origin. If this fit gives a
644: value of $b \in (-1/2,0)$, then we know that the solution will blow up
645: at null infinity even though it is decaying at the origin (see Figure
646: \ref{fig:nullup}).
647: 
648: \begin{figure}[t]
649:  \begin{centering}
650:    \psfrag{rho}{$\rho$} \psfrag{phi}{$\Phi$}
651:      \includegraphics[width=0.46\textwidth]{figures/nullup}
652:      \caption{Blowup at null infinity. The numerical
653:        solution (small squares) at the times (counting, near the origin, from top to bottom)
654:        $\tau = \{2.3,3.2,5,7,25\}$ is compared to
655:        the attractor solution (solid line) with
656:        $b=-0.02$.  We see that the solution grows at
657:        null infinity while it decays near the
658:        origin. \label{fig:nullup}}
659:  \end{centering}
660: \end{figure}
661: 
662: 
663: \subsection{Critical behavior}
664: Now we consider the behavior of solutions for initial data lying at the boundary between
665: dispersion and blowup. Let us recall that this problem was addressed before in \cite{bct} for the
666: focusing wave equation $\partial_{tt} \phi - \Delta\phi- \phi^p=0$ with three values of the
667: exponent $p$ (corresponding to three different criticality classes with respect to scaling of
668: energy): $p=3$ (subcritical), $p=5$ (critical), and $p=7$ (supercritical). It was shown there
669: that the nature of the critical solution, whose codimension-one stable manifold separates blowup
670: from dispersion, depends on $p$: for $p=7$ the critical solution is self-similar, while for $p=5$
671: it is static. For $p=3$ the critical solution was not determined because of numerical
672: difficulties (although with hindsight it could have been inferred from Figure~11 in \cite{bct}).
673: Now, in view of Conjectures~1 and 2 which assert that the solution $\phi_{(a,b)}$ is an attractor
674: for generic dispersive solutions if $b>0$ and for blowup solutions if $b<0$, it is easy to guess
675: that the critical solution corresponds to $b=0$, hence it has the form $\phi_c=\sqrt{2}/(t+a)$.
676: 
677: \begin{figure}[t]
678:  \begin{centering}
679:    {\psfrag{rho}{$\rho$}\psfrag{phi}{$\Phi$}
680:      \includegraphics[width=0.44\textwidth]{figures/crit_fit}\hspace{13mm}}
681:    {\psfrag{t}{$\tau$} \psfrag{phi}{$\Phi|_{\scri^+}$}
682:      \includegraphics[width=0.44\textwidth]{figures/crit_dev}}
683:    \caption{Left panel: The critical solution on the grid at times
684:      $\tau=\{2,3,5,8,13,34,89\}$ (counting from top to bottom). We see
685:      that at $\tau=2$ the solution is not yet described well by the
686:      attractor solution, but already after $\tau=3$ the agreement is
687:      good. Right panel: Numerical solutions close to the critical
688:      solution evaluated at null infinity. Denoting the initial
689:      amplitude for the solution that corresponds to the almost constant solid line at $\sqrt{2}$ by
690:      $A_c$, the deviating dashed curves have $A=A_c \pm 10^{-9}$, the
691:      deviating solid ones have $A=A_c\pm 10^{-8}$. \label{fig:critical}}
692:  \end{centering}
693: \end{figure}
694: 
695: The critical solution is difficult to study with standard numerical methods by bisection because
696: it is a globally decaying solution, but it can be studied very accurately with the conformal
697: method. The reason is that the conformal scaling factors out the leading asymptotic
698: behavior implying that the $1/t$ decay of the critical solution is factored out at the conformal
699: boundary.  Specifically, the rescaled critical solution $\Phi_c=\phi_c/\Omega$ in the new
700: coordinates as given in (\ref{phiab_hypal}) evaluated at null infinity becomes
701: $\Phi_c|_{\scri^+}=\Phi_{(a,0)}(\tau,1) = \sqrt{2}$, hence the deviation of $\Phi|_{\scri^+}$
702: from $\sqrt{2}$ can be used as a bisection criterion. On the left panel of Figure
703: \ref{fig:critical} we plot the critical solution on the grid at various time steps and compare it
704: to the attractor solution with $b=0$ and fitted $a$. We see that the solution is decaying while
705: its value at $\scri^+=\{\rho=1\}$ is constant. The instability of the critical solution can be seen
706: by evolving initial data that differ slightly from the critical amplitude $A_c$ -- for such data
707: the deviation from $\sqrt{2}$ at $\scri^+$ grows linearly with time. This is depicted on the
708: right panel of Figure \ref{fig:critical} for four different values of marginally critical
709: amplitude.
710: 
711: 
712: \section{Final remarks}
713: 
714: We wish to emphasize that the use of the hyperboloidal foliation (\ref{eq:tau}) in combination with
715: the conformal method was instrumental in unraveling the dynamics of global solutions of
716: equation (\ref{eq}). First and foremost, this method eliminates the need of introducing an
717: artificial boundary, which is a notorious problem in computing wave propagation on unbounded
718: domains. Second, the intersection of $\tau=const$ hyperboloids with $\scri^+$ increases
719: monotonically with $\tau$ leading to the dispersive dissipation of energy along the leaves of the
720: foliation which is a mechanism responsible for convergence to the attractor. Third,
721: the conformal rescaling
722: allows a very accurate computation of the critical solution by factoring out its leading
723: asymptotic behavior. Finally, with this approach we can probe efficiently the shape of the blowup
724: surface and observe the simultaneous blowup on the whole grid.
725: 
726: 
727: Our main observation that a simple family of exact solutions can act as a universal attractor for
728: solutions of the nonlinear wave equation was unexpected to us. It is clear that this surprising
729: phenomenon is intimately related to the conformal invariance of the cubic wave equation, and
730: therefore it is more a curiosity than a stable property, in particular it is absent for
731: semilinear focusing wave equations
732: \begin{equation}\label{focusp}
733: \Box \phi +|\phi|^{p-1}\phi=0
734: \end{equation}
735: with $p\neq 3$. However, we conjecture that the threshold behavior
736: mediated by the slowly decaying global solution is structurally stable in the sense that the ODE
737: solution of equation~(\ref{focusp})
738: \begin{equation}\label{diffp}
739:    \phi=\frac{c}{(t+a)^{\frac{2}{p-1}}}\,,\quad \mbox{where} \qquad
740:    c=\left[\frac{2(p+1)}{(p-1)^2}\right]^{\frac{1}{p-1}}\,,
741: \end{equation}
742: is a critical solution for all exponents satisfying $1+\sqrt{2}<p\leq 3$. Note that the decay
743: rate exponent of the critical solution, $2/(1-p)$, and the decay rate exponent of generic
744: solutions, $1-p$, merge for $p\rightarrow 1+\sqrt{2}$, which is consistent with the fact that for
745: $p \leq 1+\sqrt{2}$ all solutions of equation (\ref{focusp}) with compactly supported initial
746: data blow up in finite time \cite{fjohn}.
747: 
748: \ack We thank Helmut Friedrich, Sascha Husa, Vince Moncrief, and Nikodem Szpak for discussions.
749: PB thanks the Albert Einstein Institute in Golm for hospitality and support during the initial
750: phase of this project. PB acknowledges support by the MNII grants: NN202 079235 and
751: 189/6.PRUE/2007/7. AZ acknowledges support by the NSF grant PHY0801213 to the University of
752: Maryland.
753: 
754: \section*{References}
755: \bibliography{references}\bibliographystyle{utphys}%plain
756: 
757: \end{document}
758: 
759: