math0701337/jcp.tex
1: \documentclass[12pt]{article}
2: \usepackage{amsmath,amssymb}
3: \usepackage{graphicx}
4: \setlength{\textwidth}{6.5in}
5: \setlength{\textheight}{8.75in}
6: \setlength{\headheight}{0cm}
7: \setlength{\topmargin}{-1cm}
8: \setlength{\oddsidemargin}{0.in}
9: \setlength{\evensidemargin}{0.in}
10: \setlength{\parskip}{1.0mm}
11: \newcommand{\triangleup}[0]{\nabla}
12: 
13: %\renewcommand{\baselinestretch}{2}
14: %\newcommand{\mybf}[2]{{\bf{#2}}}
15: \newcommand{\mybf}[2]{#2}
16: \renewcommand{\mod}[1]{\mathrm{#1}}
17: 
18: \begin{document}
19: 
20: \title{Computing Nearly Singular Solutions Using Pseudo-Spectral Methods}
21: \author{Thomas Y. Hou\thanks{Applied and Comput. Math, 217-50, Caltech, Pasadena, 
22: CA 91125. Email: hou@acm.caltech.edu, and LSEC, 
23: Academy of Mathematics and Systems Sciences, Chinese Academy of Sciences,
24: Beijing 100080, China.}
25: \and Ruo Li\thanks{LMAM\&School of Mathematical Sciences, Peking
26:   University, Beijing 100871, China.
27: Email: rli@acm.caltech.edu.}}
28: 
29: \maketitle
30: 
31: \begin{abstract}
32: In this paper, we investigate the performance of pseudo-spectral
33: methods in computing nearly singular solutions of fluid dynamics
34: equations. We consider two different ways of removing the
35: aliasing errors in a pseudo-spectral method. The first one
36: is the traditional 2/3 dealiasing rule. The second one 
37: is a high (36th) order Fourier smoothing which keeps a significant
38: portion of the Fourier modes beyond the 2/3 cut-off point in the
39: Fourier spectrum for the 2/3 dealiasing method. Both the 1D Burgers 
40: equation and the 3D incompressible Euler equations are considered. We 
41: demonstrate that the pseudo-spectral method with the high order 
42: Fourier smoothing gives a much better performance than the
43: pseudo-spectral method with the 2/3 dealiasing rule. Moreover, 
44: we show that the high 
45: order Fourier smoothing method captures about $12 \sim 15\%$ 
46: more effective Fourier modes in each dimension than the 2/3 
47: dealiasing method. For the 3D Euler equations, the gain in the
48: effective Fourier codes for the high order Fourier smoothing
49: method can be as large as 20\% over the 2/3 dealiasing method. 
50: Another interesting observation is that the error produced by 
51: the high order Fourier smoothing method is highly localized near 
52: the region where the solution is most singular, while the 2/3 
53: dealiasing method tends to produce oscillations in the entire 
54: domain. The high order Fourier
55: smoothing method is also found be very stable dynamically.
56: No high frequency instability has been observed.
57: \end{abstract}
58: 
59: \section{Introduction}
60: 
61: Pseudo-spectral methods have been one of the most commonly used
62: numerical methods in solving nonlinear partial differential
63: equations with periodic boundary conditions. Pseudo-spectral
64: methods have the advantage of computing a nonlinear convection
65: term very efficiently using the Fast Fourier Transform. On the
66: other hand, the discrete Fourier transform of a periodic
67: function introduces the so-called aliasing error
68: \cite{GO77,CHQZ88,CHQZ06, Boyd00},
69: which is partially due to the artificial periodicity of the 
70: discrete Fourier coefficient as a function of the wave number. 
71: The aliasing error pollutes
72: the accuracy of the high frequency modes, especially those
73: last 1/3 of the high frequency modes.  Without using any 
74: dealiasing or Fourier smoothing, the pseudo-spectral method may 
75: suffer from some mild numerical instability \cite{GHT94}.  
76: One of the most commonly used dealiasing methods is the 
77: so-called 2/3 dealiasing rule, in which one sets to zero
78: the last 1/3 of the high frequency modes and keeps 
79: the first 2/3 of the Fourier modes unchanged.
80: Another way to control the aliasing errors is to apply
81: a smooth cut-off function or Fourier smoothing to the 
82: Fourier coefficients. However, many existing Fourier 
83: smoothing methods damp the last 1/3 of the high frequency 
84: modes just like the 2/3 dealiasing method.
85: 
86: In this paper, we investigate the performance of pseudo-spectral
87: methods using the 2/3 dealiasing rule and a high order Fourier 
88: smoothing. In the Fourier smoothing method, we use a 36th order 
89: Fourier smoothing function which keeps a significant portion of 
90: the Fourier modes beyond the 2/3 cut-off point in the Fourier
91: spectrum for the 2/3 dealiasing 
92: rule. We apply these two methods to compute nearly singular 
93: solutions in fluid flows. Both the 1D Burgers equation
94: and the 3D incompressible Euler equations will be considered.
95: The advantage of using the Burgers equation is that it shares some
96: essential difficulties as other fluid dynamics equations, and yet we 
97: have a semi-analytic formulation for its solution. By using the 
98: Newton iterative method, we can obtain an approximate solution 
99: to the exact solution up to 13 digits of accuracy. Moreover, 
100: we know exactly when a shock singularity will form in time. This 
101: enables us to perform a careful convergence study in both the 
102: physical space and the spectral space very close to the singularity 
103: time. 
104: 
105: We first perform a careful convergence study of the two pseudo-spectral 
106: methods in both physical and spectral spaces for the 1D Burgers 
107: equation. Our extensive numerical results demonstrate that the 
108: pseudo-spectral method with the high order Fourier smoothing
109: (the Fourier smoothing method for short) gives a much better performance 
110: than the pseudo-spectral method with the 2/3 dealiasing rule
111: (the 2/3 dealiasing method for short). In particular, we show that 
112: the unfiltered high frequency coefficients in the Fourier smoothing method 
113: approximate accurately the corresponding exact Fourier coefficients. 
114: More precisely, we demonstrate that the Fourier 
115: smoothing method captures about $12 \sim 15\%$ more effective Fourier 
116: modes than the 2/3 dealiasing method in each dimension. The gain is 
117: even higher for the 3D Euler equations since the number of effective 
118: modes in the Fourier smoothing method is higher in three 
119: dimensions. Thus the Fourier smoothing method gives a 
120: more accurate approximation than the 2/3 dealiasing method. We 
121: will illustrate this improved accuracy by studying the errors in
122: $L^\infty$-norm and $L^1$-norm as a function of time, and by studying
123: the spatial distribution of the pointwise error and the convergence of 
124: the Fourier spectrum at a sequence of times very close to the singularity
125: time. Another interesting observation is that the error produced by 
126: the Fourier smoothing method is highly localized near the 
127: region where the solution is most singular and decays exponentially fast 
128: with respect to the distance from the singularity point. The error in 
129: the smooth region is several orders of magnitude smaller than that in 
130: the singular region. On the other hand, the 
131: 2/3 dealiasing method produces noticeable oscillations in the entire 
132: domain as we approach the singularity time. This is to some extent
133: due to the Gibbs phenomenon and the loss of the $L^2$ energy associated
134: with the solution.
135: Moreover, our computational results show that in the smooth region
136: the error produced by the Fourier smoothing method is several orders
137: of magnitude smaller than that produced by the 2/3 dealiasing method.
138: This is an important advantage of the Fourier smoothing 
139: method over the 2/3 dealiasing method.   
140: 
141: Next, we apply the two pseudo-spectral methods to solve the 
142: nearly singular solution of the 3D incompressible Euler equations.
143: We would like to see if the comparison we make regarding the convergence
144: properties of the two pseudo-spectral methods for the 1D Burgers equation
145: is still valid for the more challenging 3D incompressible Euler equations. 
146: In order to make our comparison meaningful, we choose a smooth initial 
147: condition which could potentially develop a finite time singularity.
148: There have been many computational efforts in searching for finite time
149: singularities of the 3D Euler equations, see e.g.
150: \cite{Chorin82,PS90,KH89,GS91,SMO93,Kerr93,Caf93,BP94,FZG95,Pelz98,GMG98,Kerr05}.
151: One of the frequently cited numerical evidences for a finite time blowup
152: of the Euler equations is the two slightly perturbed anti-parallel vortex tubes
153: initial data studied by Kerr \cite{Kerr93,Kerr05}. In Kerr's computations, 
154: a pseudo-spectral discretization with the 2/3 dealiasing rule was used 
155: in the $x$ and $y$ directions while a Chebyshev polynomial discretization 
156: was used along the $z$ direction. His best space resolution was of the 
157: order $512\times 256\times 192$. In \cite{Kerr93,Kerr05}, Kerr reported 
158: that the maximum vorticity blows up like $O((T-t)^{-1})$ and the velocity 
159: field blows up like $O((T-t)^{-1/2})$. The alleged singularity time $T$ 
160: is equal to 18.7 while his computations beyond $t=17$ were not considered 
161: as the primary evidence for a singularity since they were polluted by 
162: noises \cite{Kerr93}.
163: 
164: We perform a careful convergence study of the two pseudo-spectral methods
165: using Kerr's initial condition with a sequence of resolutions up to $T=19$, 
166: beyond the singularity time alleged in \cite{Kerr93,Kerr05}. The largest 
167: space resolution we use is $1536\times 1024 \times 3072$. Convergence in 
168: both physical and spectral space has been observed for the two
169: pseudo-spectral methods. Both numerical methods converge to the same 
170: solution under mesh refinement. Our computational study also demonstrates
171: that the Fourier smoothing method offers better computational accuracy 
172: than the 2/3 dealiasing method. For a given resolution, the Fourier 
173: smoothing method captures about 20\% more effective Fourier modes 
174: than the 2/3 dealiasing method does. We also find that the 2/3 
175: dealiasing method produces some oscillations near the 2/3 
176: cut-off point of the spectrum. This abrupt cut-off of the Fourier 
177: spectrum generates noticeable oscillations in the vorticity contours 
178: at later times. Even using a relative high resolution 
179: $1024\times 786 \times 2048$, we find that the vorticity contours 
180: obtained by the 2/3 dealiasing method still suffer relatively large 
181: oscillations in the late stage of the computations, which are to
182: some extent caused by the Gibbs phenomenon and the loss of enstrophy 
183: due to the abrupt cut-off of the high frequency modes.
184: On the other hand, the vorticity contours 
185: obtained by the Fourier smoothing method remains smooth throughout the 
186: computations. Our spectral computations using both the 2/3 dealiasing 
187: rule and the high order Fourier smoothing confirm the finding 
188: reported in \cite{HL06}, i.e. the maximum vorticity does not grow 
189: faster than double exponential in time and the velocity field remains 
190: bounded up to $T=19$.
191: 
192: We would like to emphasize that the Fourier smoothing method is very 
193: stable and robust in all our computational experiments. We do not observe 
194: any high frequency instability in our computations for both the 1D Burgers 
195: equation and the 3D incompressible Euler equations. The resolution study 
196: that we conduct is completely based on the consideration of accuracy, not 
197: by the consideration of stability. 
198: 
199: We would like to mention that Fourier smoothing has been also used 
200: effectively to approximate discontinuous solutions of linear hyperbolic
201: equations, see e.g. \cite{MMO77,ML78,AGT86}. To compute discontinuous
202: solutions for nonlinear conservation laws, the spectral viscosity method 
203: has been introduced and analyzed, see \cite{Tad89,MT89} and the review 
204: article \cite{Tad93}. Like Fourier smoothing, the purpose of introducing
205: the spectral viscosity is to localize the Gibbs oscillations, maintaining
206: stability without loss of spectral accuracy.
207: 
208: The remaining of the paper is organized as follows. In Section 2, we 
209: present a careful convergence study of the two pseudo-spectral methods 
210: for the 1D Burgers equation. In Section 3, we present a similar convergence
211: study for the incompressible 3D Euler equations using Kerr's initial 
212: data. Some concluding remarks are made in Section 4.
213: 
214: 
215: \section{Convergence study of the two pseudo-spectral methods for the 1D Burgers equation}
216: 
217: In this section, we perform a careful convergence study of the 
218: two pseudo-spectral methods for the 1D Burgers equation. The 1D Burgers
219: equation shares some of the essential difficulties in many fluid
220: dynamic equations. In particular, it has the same type of quadratic 
221: nonlinear convection term as other fluid dynamics equations.
222: It is well known that the 1D Burgers equation can form a shock 
223: discontinuity in a finite time \cite{LeVeque92}. The advantage 
224: of using the 1D Burgers equation as a prototype is that we have 
225: a semi-analytical solution formulation for the 1D Burgers equation. 
226: This allows us to use the Newton iterative method to obtain a very accurate 
227: approximation (up to 13 digits of accuracy) to the exact solution of 
228: the 1D Burgers equation arbitrarily close to the singularity time. This 
229: provides a solid foundation in our convergence study of the two
230: spectral methods.
231: 
232: We consider the inviscid 1D Burgers equation 
233: \begin{equation}\label{eq.burgers}
234: u_t + \left(\frac{u^2}{2}\right)_x = 0, \quad -\pi \leq x \leq \pi,
235: \end{equation}
236: with an initial condition given by
237: \[ 
238: u|_{t=0} = u_0(x). 
239: \]
240: We impose a periodic boundary condition over $[-\pi, \pi]$. 
241: By the method of characteristics,
242: it is easy to show that the solution of the 1D Burgers equation is given by
243: \begin{equation}
244: \label{burgers-sol}
245: u(x,t)=u_0(x-t u(x,t)) .
246: \end{equation}
247: The above implicit formulation defines a unique solution for 
248: $u(x,t)$ up to the time when the first shock singularity 
249: develops. After the shock singularity develops, equation (\ref{burgers-sol})
250: gives a multi-valued solution. An entropy condition is required to
251: select a unique physical solution beyond the shock singularity \cite{LeVeque92}. 
252: 
253: We now use a standard pseudo-spectral method to
254: approximate the solution. Let $N$ be an integer, and let 
255: $h=\pi/N$. We denote by $x_j = j h$ ($j=-N,...,N$) the discrete 
256: mesh over the interval $[-\pi,\pi]$. To describe the pseudo-spectral
257: methods, we recall that the discrete Fourier transform of a periodic
258: function $u(x)$ with period $2 \pi$ is defined by
259: \[ 
260: \hat{u}_k = \frac{1}{2N}\sum_{j=-N+1}^{N} u(x_j) e^{-i k x_j} \;.
261: \]
262: The inversion formula reads
263: \[
264: u(x_j) = \sum_{k=-N+1}^{N} \hat{u}_k e^{i k x_j} \; .
265: \]
266: We note that $\hat{u}_k $ is periodic in $k$ with period $2N$. This
267: is an artifact of the discrete Fourier transform, and the source of
268: the aliasing error. To remove the aliasing error, one usually applies
269: some kind of dealiasing filtering when we compute the discrete 
270: derivative. Let $\rho(k/N)$ be a cut-off function in the spectrum
271: space. A discrete derivative operator may be expressed in the
272: Fourier transform as 
273: \begin{equation}
274: \label{spec_deriv}
275: \widehat{(D_h u)}_k = ik \rho (k/N) \widehat{u}_k , \quad k=-N+1, ...,N.
276: \end{equation}
277: Both the 2/3 dealiasing rule and the Fourier
278: smoothing method can be described by a specific choice of the
279: high frequency cut-off function, $\rho$ (also known as Fourier
280: filter). For the 2/3 dealiasing rule, the cut-off function is 
281: chosen to be 
282: \begin{equation}
283: \label{cutoff-23rd}
284: \rho (k/N) = \left \{ \begin{array}{ll} 1, & \mbox{if $ |k/N| \leq 2/3$},\\
285:                                       0, & \mbox{if $ |k/N| > 2/3$.} 
286:                     \end{array}
287:             \right .
288: \end{equation}
289: In our computations, in order to obtain an alias-free computation on
290: a grid of $M$ points for a quadratic nonlinear equation, we apply the 
291: above filter to the high wavenumbers so as to retain only $(2/3)M$ 
292: unfiltered wavenumbers before making the coefficient-to-grid Fast 
293: Fourier Transform. This dealiasing procedure is alternatively known
294: as the the 3/2 dealiasing rule because to obtain $M$ unfiltered 
295: wavenumbers, one must compute nonlinear products in physical space on 
296: a grid of $(3/2)M$ points, see page 229 of \cite{Boyd00} for more 
297: discussions.
298: 
299: For the Fourier smoothing method, we choose $\rho$ as follows:
300: \begin{equation}
301: \label{cutoff-sm}
302: \rho (k/N) =  e^{-\alpha (|k|/N)^m},
303: \end{equation}
304: with $\alpha = 36$ and $m=36$. In our implementation, both filters are
305: applied on the numerical solution at every time step. For the $2/3$ 
306: dealiasing rule, the Fourier modes with wavenumbers $|k| \ge 2/3N$ 
307: are always set to zero. Thus there is no aliasing error being 
308: introduced in our approximation of the nonlinear convection term.
309: 
310: The Fourier smoothing method we choose is based on three
311: considerations.  The first one is that the aliasing instability is 
312: introduced by the highest frequency Fourier modes. As demonstrated 
313: in \cite{GHT94}, as long as one can damp out a small portion of 
314: the highest frequency Fourier modes, the mild instability
315: caused by the aliasing error can be under control. The second 
316: observation is that the magnitude of the Fourier coefficient is 
317: decreasing with respect to the wave number $|k|$ for a function 
318: that has certain degree of regularity. Typically, we have
319: $|\widehat{u}_k| \leq C/(1+|k|^m)$ if the $m$th derivative of a 
320: function $u$ is bounded in $L^1$.
321: Thus the high frequency Fourier modes have a relatively smaller
322: contribution to the overall solution than the low to intermediate
323: frequency modes. The third observation is that one should not 
324: cut off high frequency Fourier modes abruptly to avoid the 
325: Gibbs phenomenon and the loss of the $L^2$ energy associated with the 
326: solution. This is especially important when we compute a nearly 
327: singular solution whose high frequency Fourier coefficient 
328: has a very slow decay. 
329: 
330: Based on the above considerations, we choose a smooth cut-off 
331: function which 
332: decays exponentially fast with respect to the high wave number. In 
333: our cut-off function, we choose the parameters $\alpha=36$ and $m=36$. 
334: These two parameters are chosen to achieve two objectives: (i) When 
335: $|k|$ is close to $N$, the cut-off function reaches the machine 
336: precision, i.e. $10^{-16}$; (ii) The cut-off function remains very 
337: close to 1 for $|k| < 4N/5$, and decays rapidly and smoothly to zero 
338: beyond $|k| = 4N/5$. In Figure \ref{fig.fourier_smoother}, we plot
339: the cut-off function $\rho (x)$ as a function of $x$. The cut-off 
340: function used by the 2/3rd dealiasing rule is plotted on top of
341: the cut-off function used by the Fourier smoothing method.
342: We can see that the Fourier smoothing method keeps about 
343: $12 \sim 15\%$ more modes than the 2/3 dealiasing method. In
344: this paper, we will demonstrate by our numerical experiments 
345: that the extra modes we keep by the Fourier smoothing method give 
346: an accurate approximation of the correct high frequency Fourier 
347: modes. 
348: 
349: \begin{figure}
350: \begin{center}
351: \includegraphics[width=8cm]{pics/fourier_smoother.eps}
352: \end{center}
353: \caption{The profile of the Fourier smoothing, $\exp(-36 (x)^{36})$,
354: as a function of $x$. The vertical line corresponds to the cut-off point 
355: in the Fourier spectrum in the $2/3$ 
356: dealiasing rule. We can see that using this Fourier smoothing we keep 
357: about $12 \sim 15\%$ more modes than those using the 2/3 dealiasing rule.
358: \label{fig.fourier_smoother}}
359: \end{figure}
360: 
361: Next, we will present a convergence study of the two pseudo-spectral methods using
362: a generic initial condition, $u_0(x) = \sin(x)$. For this initial condition, 
363: the solution will develop a shock singularity at $t=1$ at $x=\pi$ and $x=-\pi$.
364: For $t < 1$ but sufficiently close to the singularity time, a sharp layer will 
365: develop at $x=\pi$ and $x=-\pi$. We will perform a sequence of well-resolved 
366: computations sufficiently close to the singularity time using the two 
367: pseudo-spectral methods and study their convergence properties. In our 
368: computations for both methods, we will use a standard compact three 
369: step Runge-Kutta scheme for the time integration. In order to compute the 
370: errors of the two pseudo-spectral methods accurately, we need to solve for the 
371: ``exact solution'' on a computational grid. We do this by solving the 
372: implicit solution formula (\ref{burgers-sol}) using the Newton iterative 
373: method up to 13 digits of accuracy.
374: 
375: Our computational studies demonstrate convincingly that the 
376: Fourier smoothing method gives a much better performance than the
377: 2/3 dealiasing method. In Figure \ref{fig.burgers_l0_error}, we plot the 
378: $L^\infty$ error of the two pseudo-spectral methods as a function of time 
379: using three different resolutions. The errors are plotted in a log-log scale. 
380: The left figure is the result obtained by the 2/3 dealiasing method and the 
381: right figure is the result obtained by the Fourier smoothing method. We can 
382: see clearly that the $L^\infty$ error obtained by the Fourier smoothing 
383: method is smaller than that obtained by the 2/3rd dealiasing method. With 
384: increasing resolution, the errors obtained by both methods decay rapidly, 
385: confirming the spectral convergence of both methods. In Figure 
386: \ref{fig.burgers_l1_error}, we plot the $L^1$ errors
387: of the two methods as a function of time using three different
388: resolutions. We can see that the convergence of the Fourier smoothing
389: method is much faster than the 2/3 dealiasing method. 
390: 
391: It is interesting to study the spatial distribution of the pointwise 
392: errors obtained by the two methods. In Figure 
393: \ref{fig.burgers_pointwise_error}, we plot the pointwise errors of the 
394: two methods over one period $[-\pi,\pi]$ at $t=0.985$. In the
395: computations presented on the left picture, we use resolution $N=1024$, 
396: while the computations in the right picture corresponds 
397: to $N=2048$. The errors are plotted in a log scale. We can see that the
398: error of the 2/3rd dealiasing method, which is colored in red, is highly 
399: oscillatory and spreads out over the entire domain. This is caused by the
400: Gibbs phenomenon and the loss of the $L^2$ energy associated with the solution.
401: On the other hand, the error of the Fourier smoothing method
402: is highly localized near the location of the shock singularity at $x=\pi$
403: and $x=-\pi$, and decays exponentially fast with respect to the distance 
404: from the singularity point. The error in the smooth region is several 
405: orders of magnitude smaller than that in the singular region. This is 
406: a very interesting phenomenon. This property makes the Fourier smoothing 
407: method a better method for computing nearly singular solutions.
408: 
409: \begin{figure}
410: \begin{center}
411: \includegraphics[width=8cm]{pics/burgers_l0_error_da.eps}
412: \includegraphics[width=8cm]{pics/burgers_l0_error_fs.eps}
413: \end{center}
414: \caption{The $L^\infty$ errors of the two pseudo-spectral methods as a function
415: of time using three different resolutions. The plot is in a log-log scale. 
416: The initial condition is $u_0(x) = \sin(x)$.
417: The left figure is the result obtained by the 2/3 dealiasing method and 
418: the right figure is the result obtained by the Fourier smoothing method. 
419: It can be seen clearly that the $L^\infty$
420: error obtained by the Fourier smoothing method is smaller than that
421: obtained by the 2/3 dealiasing method. 
422: %With the increasing of the
423: %  resolution, the error obtained by the Fourier smoothing method is even
424: %  smaller which indicate a high order convergence of the Fourier
425: %  smoothing method. It is known that though the spectral method has
426: %  spectral accuracy theoretically, while practically, the convergence
427: %  order is often less than 6. The numerical result above illustrate
428: %  that the convergence order can be improved if the Fourier smoothing
429: %  method is adopted instead of the 2/3rd dealiasing method.
430:   \label{fig.burgers_l0_error}}
431: \end{figure}
432: 
433: \begin{figure}
434: \begin{center}
435: \includegraphics[width=8cm]{pics/burgers_l1_error_da.eps}
436: \includegraphics[width=8cm]{pics/burgers_l1_error_fs.eps}
437: \end{center}
438: \caption{The $L^1$ errors of the two pseudo- spectral methods as a function
439: of time using three different resolutions. The plot is in a log-log scale. 
440: The initial condition is given by $u_0(x) = \sin(x)$.
441:   The left figure is the result obtained by the 2/3
442:   dealiasing method and the right figure is the result obtained by the
443:   Fourier smoothing method. One can see that the $L^1$ error obtained by 
444:   the Fourier smoothing method is much smaller than corresponding $L^\infty$ error. 
445: \label{fig.burgers_l1_error}}
446: \end{figure}
447: 
448: \begin{figure}
449: \begin{center}
450: \includegraphics[width=8cm]{pics/burgers_pointwise_error_1024_9875.eps}
451: \includegraphics[width=8cm]{pics/burgers_pointwise_error_2048_9875.eps}
452: \end{center}
453: \caption{The pointwise errors of the two pseudo-spectral methods as a function
454: of time using three different resolutions. The plot is in a log scale. 
455: The initial condition is given by $u_0(x) = \sin(x)$.
456:   The error of the 2/3rd dealiasing method is highly oscillatory and 
457:   spreads out over the entire domain, while the error of the Fourier 
458:   smoothing method is highly localized near the location of the shock 
459:   singularity. 
460: \label{fig.burgers_pointwise_error}} 
461: \end{figure}
462: 
463: To gain further insight of the two pseudo-spectral methods, we study 
464: the convergence of the two methods in the spectral space. In Figure 
465: \ref{fig.burgers_spec}, we plot the Fourier spectra of the two
466: spectral methods at a sequence of times with two different
467: resolutions $N=4096$ and $N=8192$ respectively. We observe 
468: that when the solution is relatively smooth and can be resolved
469: by the computational grid, the spectra of the two methods are
470: almost indistinguishable. However, when the solution becomes
471: more singular and cannot be completely resolved by the computational
472: grid, the two methods give a very different performance. 
473: For a given resolution, the Fourier smoothing method
474: keeps about 20\% more Fourier modes than the 2/3 dealiasing method.
475: When we compare with the ``exact'' spectrum obtained by using
476: the Newton iterative method, we can see clearly that the extra
477: Fourier modes that are kept by the Fourier smoothing method give 
478: indeed an accurate approximation to the correct Fourier modes.
479: This explains why the Fourier smoothing method offers better
480: accuracy than the 2/3 dealiasing method. On the other hand,
481: we observe that the Fourier spectrum of the 2/3 dealiasing method
482: develops noticeable oscillations near the 2/3 cut-off point of
483: the Fourier spectrum. This abrupt cut-off in the high frequency
484: spectrum gives rise to the well-known Gibbs phenomenon and the
485: loss of the $L^2$ energy, which is 
486: the main cause for the highly oscillatory and widespread pointwise 
487: error that we observe in Figure \ref{fig.burgers_pointwise_error}. 
488: 
489: 
490: \begin{figure}
491: \begin{center}
492: \includegraphics[width=8cm]{pics/burgers_u_spec_comp_4096.eps}
493: \includegraphics[width=8cm]{pics/burgers_u_spec_comp_8192.eps}
494: \end{center}
495: \caption{Comparison of Fourier spectra of the two methods on different 
496: resolutions at a sequence of times. The initial condition is given
497: by $u_0(x) = \sin(x)$.
498: The left picture corresponds to $N=4096$ and the right picture 
499: corresponds to $N=8192$.
500: \label{fig.burgers_spec}}
501: \end{figure}
502: 
503: We have also performed similar numerical experiments for several
504: other initial data. They all give the same qualitative behavior as the
505: one we have demonstrated above.  In Figure \ref{fig.ex1_spec}, we plot
506: the Fourier spectra of the two methods at a sequence of times using
507: a different initial condition: 
508: $u_0(x) = (0.1 + \sin^2 x )^{-1/2}$. The picture on the left
509: corresponds to resolution $N=1024$, while the picture on the right
510: corresponds to resolution $N=2048$. One can see that the convergence
511: properties of the two methods are essentially the same as those 
512: presented for the initial condition $u_0(x)=\sin(x)$.
513: 
514: Finally, we would like to point out that the numerical computations
515: using the Fourier smoothing method have been very stable and robust.
516: No high frequency instability has been observed throughout our computations.
517: This indicates that the high order Fourier smoothing we use has effectively
518: eliminated the mild numerical instability introduced by the 
519: aliasing error \cite{GHT94}.
520: 
521: %\begin{figure}
522: %\begin{center}
523: %\includegraphics[width=8cm]{pics/ex1_l0_error_da.eps}
524: %\includegraphics[width=8cm]{pics/ex1_l0_error_fs.eps}
525: %\end{center}
526: %\caption{The $L^\infty$ error with time in log-log scale.\label{fig.ex1_l0_error}}
527: %\end{figure}
528: %
529: %\begin{figure}
530: %\begin{center}
531: %\includegraphics[width=8cm]{pics/ex1_l1_error_da.eps}
532: %\includegraphics[width=8cm]{pics/ex1_l1_error_fs.eps}
533: %\end{center}
534: %\caption{The $L^1$ error with time in log-log scale.\label{fig.ex1_l1_error}}
535: %\end{figure}
536: %
537: \begin{figure}
538: \begin{center}
539: \includegraphics[width=8cm]{pics/ex1_spec_comp_1024.eps}
540: \includegraphics[width=8cm]{pics/ex1_spec_comp_2048.eps}
541: \end{center}
542: \caption{Comparison of Fourier spectra of the two methods on different 
543: resolutions at a sequence of times. The initial condition is given by 
544: $u_0(x) = (0.1 + \sin^2 x)^{-1/2}$.
545: The left picture corresponds to $N=1024$ and the right picture 
546: corresponds to $N=2048$. 
547: Notice that only even modes are plotted in these figures since the odd
548: modes are vanished in this example.
549: \label{fig.ex1_spec}}
550: \end{figure}
551: 
552: \section{Computing nearly singular solutions of the 3D Euler equations using
553: pseudo-spectral methods}
554: 
555: In this section, we will apply the two pseudo-spectral methods to solve
556: the nearly singular solution of the 3D incompressible Euler equations.
557: The spectral computation of the 3D incompressible Euler equations is
558: much more challenging due to the nonlocal and nonlinear nature of the
559: problem and the possible formation of a finite time singularity.
560: It would be interesting
561: to find out if the comparison we have made regarding the convergence 
562: property of the two pseudo-spectral methods for the 1D Burgers equation
563: is still valid for the 3D Euler equations. To make our comparison useful,
564: we choose a smooth initial condition which could potentially develop a 
565: finite time singularity. There have been many computational efforts in 
566: searching for finite time singularities of the 3D Euler 
567: equations, see e.g.
568: \cite{Chorin82,PS90,KH89,GS91,SMO93,Kerr93,Caf93,BP94,FZG95,Pelz98,GMG98,Kerr05}.
569: Of particular interest is the numerical study of the interaction of two 
570: perturbed antiparallel vortex tubes by Kerr \cite{Kerr93,Kerr05}, in which 
571: a finite time blowup of the 3D Euler equations was reported. In this
572: section, we will perform the comparison of the two pseudo-spectral
573: methods using Kerr's initial condition.
574: 
575: The 3D incompressible Euler equations in the vorticity stream function
576: formulation are given as follows (see, e.g., \cite{CM93,MB02}):
577: \begin{eqnarray}\label{3deuler}
578: \vec{\omega}_t+(\vec{u}\cdot\nabla) \vec{\omega} & = & \nabla 
579: \vec{u} \cdot \vec{\omega}, \\
580: - \bigtriangleup \vec{ \psi} &= & \vec{\omega}, \quad
581: \vec{u} = \nabla \times \vec{\psi},
582: \end{eqnarray} 
583: with initial condition $\vec{\omega}\mid_{t=0} =  \vec{\omega}_{0}$,
584: where $\vec{u}$ is velocity, $\vec{\omega}$ is vorticity, and 
585: $\vec{\psi}$ is stream function. Vorticity is related to velocity 
586: by $\vec{\omega} = \nabla \times \vec{u}$. The incompressibility
587: implies that 
588: \[
589: \nabla \cdot \vec{u} = \nabla \cdot \vec{\omega} = \nabla \cdot \vec{\psi} = 0.
590: \]
591: We consider periodic boundary conditions with period $4 \pi$ in all three 
592: directions.  The initial condition is the same as the one used by Kerr 
593: (see Section III of \cite{Kerr93}, and also \cite{HL06} for corrections of 
594: some typos in the description of the initial condition in \cite{Kerr93}). 
595: Following \cite{Kerr93}, we call the $x$-$y$ plane 
596: as the ``dividing plane'' and the $x$-$z$ plane as the ``symmetry plane''. 
597: There is one vortex tube above and below the dividing plane respectively. 
598: The term ``antiparallel'' refers to the anti-symmetry of the vorticity 
599: with respect to the dividing plane in the following sense: 
600: $\vec{\omega}(x,y,z) = -\vec{ \omega}(x,y,-z)$. Moreover, with respect to 
601: the symmetry plane, the vorticity is symmetric in its $y$ component and 
602: anti-symmetric in its $x$ and $z$ components. Thus we have 
603: $\omega_x(x,y,z)=-\omega_x(x,-y,z)$, $\omega_y(x,y,z)=\omega_y(x,-y,z)$ and
604: $\omega_z(x,y,z)=-\omega_z(x,-y,z)$. Here $\omega_x, \; \omega_y , \; \omega_z$
605: are the $x$, $y$, and $z$ components of vorticity respectively. These symmetries 
606: allow us to compute only one quarter of the whole periodic cell.
607: 
608: To compare the performance of the two pseudo-spectral methods, we will 
609: perform a careful convergence study for the two methods. To get a better 
610: idea how the solution evolves dynamically, we present the 3D plot of the 
611: vortex tubes at $t=0$ 
612: and $t=6$ respectively in Figure \ref{fig.vorttube}. As we can see, the 
613: two initial vortex tubes are very smooth and essentially symmetric. Due 
614: to the mutual attraction of the two antiparallel vortex tubes, the 
615: two vortex tubes approach to each one and experience severe deformation 
616: dynamically. By time $t=6$, there is already a significant flattening 
617: near the center of the tubes. In Figure \ref{fig.local_struc_3d_17}, we 
618: plot the local 3D vortex structure of the upper vortex tube at $t=17$.
619: By this time, the 3D vortex tube has essentially turned into a
620: thin vortex sheet with rapidly decreasing thickness. The vortex sheet
621: rolls up near the left edge of the sheet. It is interesting to note 
622: that the maximum vorticity is actually located near the rolled-up 
623: region of the vortex sheet.  
624: 
625: \subsection{Convergence study of the two pseudo-spectral methods in the
626: spectral space}
627: 
628: In this subsection, we perform a convergence study for the
629: two numerical methods using a sequence of resolutions.
630: For the Fourier smoothing method, we use the resolutions 
631: $768\times 512\times 1536$,
632: $1024\times 768\times 2048$, and $1536\times 1024\times 3072$
633: respectively. Except for the computation on the largest resolution
634: $1536\times 1024\times 3072$, all computations are carried out from 
635: $t=0$ to $t=19$. The computation on the final resolution
636: $1536\times 1024\times 3072$ is started from $t=10$ with the
637: initial condition given by the computation with the resolution
638: $1024\times 768\times 2048$. For the 2/3 dealiasing method, we use
639: the resolutions $512 \times 384 \times 1024$,
640: $768\times 512\times 1536$ and $1024\times 768\times 2048$
641: respectively. The computations using these three resolutions
642: are all carried out from $t=0$ to $t=19$. The time integration is 
643: performed using the classical fourth order Runge-Kutta method. 
644: Adaptive time stepping is used to satisfy 
645: the CFL stability condition with CFL number equal to $\pi/4$. 
646: 
647: 
648: \begin{figure}
649: \begin{center} 
650: \includegraphics[width=8cm]{pics/vorttubet=0.eps}
651: \includegraphics[width=8cm]{pics/vorttubet=6.eps}
652: \end{center}
653: \caption{The 3D view of the vortex tube for $t=0$ and $t=6$. The 
654: tube is the isosurface at $60\%$ of the maximum vorticity.
655: The ribbons on the symmetry plane are the contours
656: at other different values.
657: \label{fig.vorttube}}
658: \end{figure}                                                                                 
659:                                                                                 
660: \begin{figure}
661: \begin{center}
662: \includegraphics[width=10cm]{pics/l17_label.eps}
663: %\includegraphics[width=10cm]{pics/l17.epsf}
664: \end{center}
665: \caption{The local 3D vortex structure and vortex lines around the maximum
666: vorticity at $t=17$. The size of the box on the left is
667: $0.075^3$ to demonstrate the scale of the picture.
668: \label{fig.local_struc_3d_17}}
669: \end{figure}
670:                                                                                 
671: In Figure \ref{fig.energy-spec-comp}, we compare the Fourier 
672: spectra of the energy obtained by using the $2/3$ dealiasing method
673: with those obtained by the Fourier smoothing method. 
674: For a fixed resolution $1024\times 768\times 2048$, we can see
675: that the Fourier spectra obtained by the Fourier smoothing
676: method retains more effective Fourier modes than those obtained 
677: by the $2/3$ dealiasing method. This can be seen by comparing the 
678: results with the corresponding computations using a higher 
679: resolution $1536\times 1024 \times 3072$. Moreover, the 
680: Fourier smoothing method does not give the spurious oscillations 
681: in the Fourier spectra which are present in the computations using 
682: the $2/3$ dealiasing method near the $2/3$ cut-off point. Similar 
683: convergence study has been made in the enstrophy spectra computed 
684: by the two methods. The results are given in Figures 
685: \ref{fig.enstrophy-spec-comp} and \ref{fig.enstrophy_spec_2}. 
686: They give essentially the same results.
687: 
688: \begin{figure}
689: \begin{center}
690: \includegraphics[width=12cm,height=6cm]{pics/energy_spec_comp_1024.eps}
691: \end{center}
692: \caption{The energy spectra versus wave numbers. We compare the energy
693: spectra obtained using the 
694: Fourier smoothing method with those using the 2/3 dealiasing method. The dashed
695: lines and the dashed-dotted lines are the energy spectra with the
696: resolution $1024\times 768\times 2048$ using the 2/3 dealiasing method and the
697: Fourier smoothing method, respectively. The solid lines are the energy spectra
698: obtained by the Fourier smoothing method with the highest resolution 
699: $1536\times 1024 \times 3072$. The times for the spectra lines
700: are at $t=15, 16, 17, 18, 19$ respectively.
701: \label{fig.energy-spec-comp}}
702: \end{figure}
703: 
704: \begin{figure}
705: \begin{center}
706: \includegraphics[width=12cm,height=6cm]{pics/energy_spec_comp_768.eps}
707: \end{center}
708: \caption{The energy spectra versus wave numbers. We compare the energy
709: spectra obtained using the 
710: Fourier smoothing method with those using the 2/3 dealiasing method. The dashed
711: lines and solid lines are the energy spectra with the
712: resolution $768\times 512\times 1536$ using the 2/3 dealiasing method and the
713:  Fourier smoothing, respectively. The times for the spectra lines
714: are at $t=8, 10, 12, 14, 16, 18$ respectively.
715: \label{fig.energy-spec-comp-early}}
716: \end{figure}
717: 
718: \begin{figure}
719: \begin{center}
720: \includegraphics[width=12cm,height=6cm]{pics/enstrophy_spec_comp_1024.eps}
721: \end{center}
722: \caption{The enstrophy spectra versus wave numbers. We compare the enstrophy 
723: spectra obtained using the 
724: Fourier smoothing method with those using the 2/3 dealiasing method. The dashed 
725: lines and dashed-dotted lines are the enstrophy spectra with the
726: resolution $1024\times 768\times 2048$ using the 2/3 dealiasing method and the
727:  Fourier smoothing, respectively. The solid lines are the enstrophy spectra
728: with resolution $1536\times 1024\times 3072$ obtained using the 
729: Fourier smoothing. The times for the spectra lines 
730: are at $t= 15, 16, 17, 18, 19$ respectively. 
731: \label{fig.enstrophy-spec-comp}}
732: \end{figure}
733: 
734: \begin{figure} 
735: \begin{center}
736: \includegraphics[width=12cm,height=6cm]{pics/enstrophy_spec_comp_da.eps}
737: \end{center}
738: \caption{Convergence study for enstrophy spectra obtained by
739: the 2/3 dealiasing method using different resolutions. The solid line 
740: is computed with
741: resolution $512\times 384\times 1024$, the dashed line is 
742: computed with resolution $786\times 512\times 1536$, and
743: the dashed-dotted line is computed with resolution 
744: $1024\times 768\times 2048$. The times for the lines 
745: from bottom to top are $t=8, 10, 12, 14, 16, 18, 19$.
746: \label{fig.enstrophy_spec_2}}
747: \end{figure}
748:                                                                                          
749: We perform further comparison of the two methods using the
750: same resolution. In  Figure \ref{fig.energy-spec-comp-early},
751: we plot the energy spectra computed by the two methods using 
752: resolution $768\times 512\times 1536$. We can see that 
753: there is almost no difference in the Fourier spectra 
754: generated by the two methods in early times, $t=8, 10$,
755: when the solution is still relatively smooth. The difference  
756: begins to show near the cut-off point when the Fourier spectra 
757: rise above the round-off error level starting from $t=12$. 
758: We can see that the spectra computed by the 2/3 dealiasing 
759: method introduces noticeable oscillations near the 2/3 
760: cut-off point. The spectra computed by the Fourier
761: smoothing method, on the other hand, extend smoothly
762: beyond the 2/3 cut-off point. As we see from 
763: Figures \ref{fig.energy-spec-comp-early} and
764: \ref{fig.energy-spec-comp}, a significant portion 
765: of those Fourier modes beyond the 2/3 cut-off position 
766: are still accurate. This portion of the Fourier modes that
767: go beyond the 2/3 cut-off point is about $12\sim 15\%$ of total 
768: number of modes in each dimension. For 3D problems, the total 
769: number of effective modes in the Fourier smoothing method is 
770: about 20\% more than that in the 2/3 dealiasing method.
771: This is a very significant increase in the resolution for
772: a large scale computation. In our largest resolution, 
773: the effective Fourier modes in our Fourier smoothing method
774: are more than 320 millions, which has 140 millions more
775: effective modes than the corresponding 2/3 dealiasing method. 
776: 
777: \subsection{Comparison of the two methods in the physical space}
778: 
779: Next, we compare the solutions obtained by the two methods 
780: in the physical space for the velocity field and the vorticity. 
781: In Figure \ref{fig.max-u-comp-1024}, 
782: we compare the maximum velocity as a function of time computed by 
783: the two methods using resolution $1024\times 768\times 2048$. The 
784: two solutions are almost indistinguishable. In Figure 
785: \ref{fig.max-vort-comp-1024}, we plot the maximum vorticity
786: as a function of time. The two solutions also agree reasonably 
787: well. However, the comparison of the solutions obtained by
788: the two methods at resolutions lower than $1024\times 768\times 2048$ 
789: shows more significant differences of the two methods, see
790: Figures \ref{fig.max-vort-comp-768}, \ref{fig.omega} and
791: \ref{fig.omega_2}.
792: 
793: To understand better how the two methods differ in their 
794: performance, we examine the contour plots of the axial vorticity 
795: in Figures \ref{fig.vort-cont-comp-1024-17},
796: \ref{fig.vort-cont-comp-1024-18} and \ref{fig.vort-cont-comp-1024-19}. 
797: As we can see, the vorticity computed by the 2/3 dealiasing method
798: already develops small oscillations at $t=17$. The oscillations grow 
799: bigger by $t=18$ (see Figure \ref{fig.vort-cont-comp-1024-18}), 
800: and bigger still at $t=19$ (see Figure \ref{fig.vort-cont-comp-1024-19}). 
801: We note that the oscillations in the axial vorticity contours concentrate
802: near the region where the magnitude of vorticity is close to zero. Thus
803: they have less an effect on the maximum vorticity. On the other hand, 
804: the solution computed by the Fourier smoothing method is still relatively 
805: smooth.
806: 
807: \begin{figure}
808: \begin{center}
809: \includegraphics[width=10cm,height=6cm]{pics/max_velo_comp_1024.eps}
810: \end{center}
811: \caption{Comparison of maximum velocity as a function of time
812: computed by two methods. The solid line represents the solution
813: obtained by the Fourier smoothing method,
814: and the dashed line represents the solution obtained by
815: the 2/3 dealiasing method.
816: The resolution is $1024\times 768\times 2048$ for both methods.
817: \label{fig.max-u-comp-1024}}
818: \end{figure}
819: 
820: \begin{figure}
821: \begin{center}
822: \includegraphics[width=10cm]{pics/max_vort_comp_1024.eps}
823: \end{center}
824: \caption{Comparison of maximum vorticity as a function of time
825: computed by two methods. The solid line represents the solution
826: obtained by the Fourier smoothing method,
827: and the dashed line represents the solution obtained by
828: the 2/3 dealiasing method.
829: The resolution is $1024\times 768\times 2048$ for both methods.
830: \label{fig.max-vort-comp-1024}}
831: \end{figure}
832: 
833: \begin{figure}
834: \begin{center}
835: \includegraphics[width=10cm]{pics/max_vort_comp_768.eps}
836: \end{center}
837: \caption{Comparison of maximum vorticity as a function of time
838: computed by two methods. The solid line represents the solution
839: obtained by the Fourier smoothing method,
840: and the dashed line represents the solution obtained by
841: the 2/3 dealiasing method.
842: The resolution is $768\times 512\times 1024$ for both methods.
843: \label{fig.max-vort-comp-768}}
844: \end{figure}
845: 
846: \begin{figure}
847: \begin{center}
848: \includegraphics[width=12cm,height=4cm]{pics/vort_cont_da_1024_17.epsf}
849: \includegraphics[width=12cm,height=4cm]{pics/vort_cont_fs36_1024_17.epsf}
850: \end{center}
851: \caption{Comparison of axial vorticity contours at $t=17$ 
852: computed by two methods. The upper picture is the solution
853: obtained by the 2/3 dealiasing method,
854: and the picture on the bottom is the solution obtained by
855: the Fourier smoothing method.
856: The resolution is $1024\times 768\times 2048$ for both methods.
857: \label{fig.vort-cont-comp-1024-17}}
858: \end{figure}
859:                                                                                           
860: \begin{figure}
861: \begin{center}
862: \includegraphics[width=12cm,height=4cm]{pics/vort_cont_da_1024_18.epsf}
863: \includegraphics[width=12cm,height=4cm]{pics/vort_cont_fs36_1024_18.epsf}
864: \end{center}
865: \caption{Comparison of axial vorticity contours at $t=18$ 
866: computed by two methods. The upper picture is the solution
867: obtained by the 2/3 dealiasing method,
868: and the picture on the bottom is the solution obtained by
869: the Fourier smoothing method.
870: The resolution is $1024\times 768\times 2048$ for both methods.
871: \label{fig.vort-cont-comp-1024-18}}
872: \end{figure}
873:                                                                                           
874: \begin{figure}
875: \begin{center}
876: \includegraphics[width=12cm,height=4cm]{pics/vort_cont_da_1024_19.epsf}
877: \includegraphics[width=12cm,height=4cm]{pics/vort_cont_fs36_1024_19.epsf}
878: \end{center}
879: \caption{Comparison of axial vorticity contours at $t=19$ 
880: computed by two methods. The upper picture is the solution
881: obtained by the 2/3 dealiasing method,
882: and the picture on the bottom is the solution obtained by
883: the Fourier smoothing method.
884: The resolution is $1024\times 768\times 2048$ for both methods.
885: \label{fig.vort-cont-comp-1024-19}}
886: \end{figure}
887:                                                                                           
888: %\begin{figure}
889: %\begin{center}
890: %\includegraphics[width=12cm,height=6cm]{pics/energy_spec_comp_1024_1536.eps}
891: %%\includegraphics[width=12cm,height=6cm]{pics/energy_spec_1.eps}
892: %\end{center}
893: %\caption{Convergence study for energy spectra obtained by
894: %the pseudo-spectral method with high order smoothing using different 
895: %resolutions.
896: %The dashed lines and the solid lines are the energy spectra on resolution
897: %$1536\times 1024\times 3072$ and $1024\times 768\times 2048$,
898: %respectively. The times for the lines from bottom to top are $t=15, 16, 17,
899: %18, 19$.
900: %\label{fig.energy_spec_1}}
901: %\end{figure}
902: %
903: To further demonstrate the accuracy of our computations we compare the 
904: maximum vorticity obtained by the Fourier smoothing method for three 
905: different resolutions: $768\times 512\times 1536$, 
906: $1024\times 768\times 2048$, and $1536\times 1024\times 3072$ respectively.
907: The result is plotted in Figure \ref{fig.omega}.
908: We have performed a similar convergence study for the
909: 2/3 dealiasing method for the maximum vorticity. The result is
910: given in Figure \ref{fig.omega_2}. Two conclusions
911: can be made from this resolution study. First, by comparing
912: Figure \ref{fig.omega} with Figure \ref{fig.omega_2},
913: we can see that the Fourier smoothing method is indeed more accurate 
914: than the 2/3 dealiasing method for a given resolution. The 2/3
915: dealiasing method gives a slower growth rate in the maximum vorticity
916: with resolution $768\times 512\times 1536$. Secondly, the 
917: resolution $768\times 512\times 1536$ is not good enough to resolve 
918: the nearly singular solution at later times. On the other hand, we 
919: observe that the difference between the numerical solution obtained by
920: the resolution $1024\times 768\times 2048$ and that obtained by
921: the resolution $1536\times 1024\times 3072$ is relatively small.
922: This indicates that the vorticity is reasonably well-resolved
923: by our largest resolution $1536\times 1024\times 3072$. 
924: 
925: We have also performed a similar resolution study for the maximum
926: velocity in Figure \ref{fig.velocity}. The solutions obtained by
927: the two largest resolutions are almost indistinguishable,
928: which suggests that the velocity is well-resolved by
929: our largest resolution $1536\times 1024\times 3072$.
930: 
931: \begin{figure}
932: \begin{center}
933: %\includegraphics[width=9cm]{pics/max_vort_comp.eps}
934: \includegraphics[width=9cm]{pics/max_vort_comp_fs36.eps}
935: \end{center} 
936: \caption{The maximum vorticity $\|\vec{\omega}\|_\infty$ in time 
937: computed by the Fourier smoothing method using
938: different resolutions.
939: \label{fig.omega}}
940: \end{figure} 
941: 
942: \begin{figure}
943: \begin{center}
944: \includegraphics[width=9cm]{pics/max_vort_comp_da.eps}
945: %\includegraphics[width=8cm]{pics/max_vort_comp_1024.eps}
946: \end{center}
947: \caption{The maximum vorticity $\|\vec{\omega}\|_\infty$ in time
948: computed by the 2/3 dealiasing method using different resolutions.
949: %(b) Comparison of the maximum vorticity in time computed by the two methods
950: %with resolutions $1024\times 768\times 2048$ (using high order smoothing)
951: %and $1024\times 768\times 2048$ (using 2/3 dealiasing) respectively.
952: \label{fig.omega_2}} 
953: \end{figure}
954: 
955: The resolution study given by Figures 
956: \ref{fig.max-vort-comp-768} and \ref{fig.omega_2} 
957: also suggests that the the computation obtained by the 
958: pseudo-spectral method with the 2/3 dealiasing rule using 
959: resolution $768\times 512\times 1536$ is significantly
960: under-resolved after $t=18$. It is interesting 
961: to note from Figure \ref{fig.max-vort-comp-768} that 
962: the computational results obtained by the two methods 
963: with resolution $768\times 512\times 1536$
964: begin to deviate from each other precisely around $t=18$. By 
965: comparing the result from Figure \ref{fig.max-vort-comp-768} 
966: with that from Figure \ref{fig.omega}, we confirm again 
967: that for a given resolution, the Fourier smoothing method
968: gives a more accurate approximation
969: than the 2/3 dealiasing method.
970: 
971: We remark that our numerical computations for the 3D incompressible
972: Euler equations using the Fourier smoothing method are very 
973: stable and robust. No high frequency instability has been observed 
974: throughout the computations. The resolution study we perform here is
975: completely based on the consideration of accuracy, not on stability.
976: This again confirms that the Fourier smoothing method offers a very 
977: stable and accurate computational method for the 3D incompressible flow.
978: 
979: \subsection{Does a finite time singularity develop?}
980: 
981: Before we conclude this section, we would like to have 
982: further discussions how to interpret the numerical results
983: we have obtained. Specifically, given the fast growth of
984: maximum vorticity, does a finite time singularity develop
985: for this initial condition? 
986: 
987: In \cite{Kerr93}, Kerr presented numerical evidence which suggested
988: a finite time singularity of the 3D Euler equations for the same
989: initial condition that we use in this paper. Kerr used a 
990: pseudo-spectral discretization with the 2/3 dealiasing rule
991: in the $x$ and $y$ directions, and a Chebyshev method in the $z$ 
992: direction with resolution of order $512\times 256 \times 192$. 
993: His computations showed that the growth of the peak vorticity, 
994: the peak axial strain, and the enstrophy production obey 
995: $(T-t)^{-1}$ with $T = 18.9$. In his recent paper \cite{Kerr05},
996: Kerr applied a high wave number filter to the data obtained 
997: in his original computations to ``remove the noise that 
998: masked the structures in earlier graphics'' presented in 
999: \cite{Kerr93}. With this filtered solution, he presented 
1000: some scaling analysis of the numerical solutions up to 
1001: $t=17.5$. Two new properties were presented in this recent
1002: paper \cite{Kerr05}. First, the velocity field was shown to blow 
1003: up like $O(T-t)^{-1/2}$ with $T$ being revised to $T=18.7$.
1004: Secondly, he showed that the blowup is characterized by two 
1005: anisotropic length scales, $\rho \approx (T-t) $ and 
1006: $R \approx (T-t)^{1/2}$.
1007: 
1008: From the resolution study we present in Figure \ref{fig.omega},
1009: we find that the maximum vorticity increases rapidly from the
1010: initial value of $0.669$ to $23.46$ at the final time $t=19$,
1011: a factor of 35 increase from its initial value. Kerr's
1012: computations predicted a finite time singularity at $T=18.7$.
1013: Our computations show no sign of finite time blowup of the 3D Euler
1014: equations up to $T=19$, beyond the singularity time predicted by
1015: Kerr. From Figures \ref{fig.vort-cont-comp-1024-17},
1016: \ref{fig.vort-cont-comp-1024-18} and \ref{fig.vort-cont-comp-1024-19},
1017: we can see that a thin layer (or a vortex sheet) is formed dynamically.
1018: Beyond $t=17$, the vortex sheet has rolled up and traveled backward 
1019: for some distance. With only 192 grid points along the $z$-direction,
1020: Kerr's computations did not have enough grid points to resolve the
1021: nearly singular vortex sheet that travels backward and away from
1022: the $z$-axis. In comparison, we have 3072 grid points along the
1023: $z$-direction. This gives about 16 grid points across the nearly 
1024: singular layered structure at $t=18$ and about $8$ grid points at 
1025: $t=19$.
1026: 
1027: \begin{figure}
1028: \begin{center}
1029: \includegraphics[width=8cm]{pics/growth_rate.eps}
1030: \end{center}
1031: \caption{Study of the vortex stretching term in time. This computation is 
1032: performed by the Fourier smoothing method with resolution $1536\times
1033: 1024\times 3072$. We take $c_1 = 1/8.128$, $c_2 = 1/23.24$ to match the
1034: same starting value for all three plots.
1035: \label{fig.growth_rate}}
1036: \end{figure}
1037:                                                                                 
1038: \begin{figure}
1039: \begin{center}
1040: \includegraphics[width=8cm]{pics/loglog_omega.eps}
1041: \end{center}
1042: \caption{The plot of $ \log \log \|\omega\|_\infty$ vs time. 
1043: This computation is performed by the Fourier smoothing 
1044: method with resolution $1536\times 1024 \times 3072$.
1045: \label{fig.omega_loglog}}
1046: \end{figure}
1047:                                                                                 
1048: In order to understand the nature of the dynamic growth in
1049: vorticity, we examine the degree of nonlinearity in the vortex 
1050: stretching term. In Figure \ref{fig.growth_rate}, we plot
1051: the quantity,
1052: $\|\xi \cdot \nabla \vec{u} \cdot \vec{\omega}\|_\infty$,
1053: as a function of time, where $\xi$ is the unit vorticity vector.
1054: If the maximum vorticity indeed blew up like $O((T-t)^{-1})$,
1055: as alleged in \cite{Kerr93}, this quantity should have been
1056: quadratic as a function of maximum vorticity. We find that
1057: there is tremendous cancellation in this vortex stretching
1058: term. It actually grows slower than
1059: $C\|\vec{\omega} \|_\infty \log (\|\vec{\omega} \|_\infty )$,
1060: see Figure \ref{fig.growth_rate}. It is easy to show that 
1061: such weak nonlinearity in vortex stretching would imply 
1062: only doubly exponential growth in the maximum vorticity.
1063: Indeed, as demonstrated by Figure \ref{fig.omega_loglog}, the
1064: maximum vorticity does not grow faster than doubly exponential
1065: in time. In fact, a closer inspection reveals that the location 
1066: of the maximum vorticity has moved away from the dividing plane
1067: for $t \geq 17.5$. This implies that the compression mechanism
1068: between the two vortex tubes becomes weaker toward the end of the
1069: computation, leading to a slower growth rate in maximum vorticity
1070: \cite{HL06}.
1071:                                                                                 
1072: \begin{figure}
1073: \begin{center}
1074: %\includegraphics[width=9cm]{pics/max_velo_comp.eps}
1075: \includegraphics[width=9cm]{pics/max_velo_comp_fs36.eps}
1076: \end{center}
1077: \caption{Maximum velocity $\|\vec{u}\|_\infty$ in time
1078: computed by the Fourier smoothing method using different resolutions.
1079: \label{fig.velocity}} 
1080: \end{figure} 
1081: 
1082: Another important evidence which supports the non-blowup of
1083: the solution up to $t=19$ is that the maximum velocity remains
1084: bounded, see Figure \ref{fig.velocity}. This is in contrast
1085: with the claim in \cite{Kerr05} that the maximum velocity 
1086: blows up like $O(T-t)^{-1/2}$ with $T=18.7$. With the velocity
1087: field being bounded, the local non-blowup criteria of Deng-Hou-Yu
1088: \cite{DHY05a,DHY05b} can be applied, which implies that the solution
1089: of the 3D Euler equations remains smooth at least up to $T=19$,
1090: see also \cite{HL06}.
1091: 
1092: 
1093: \section{Conclusion Remarks}
1094: 
1095: In this paper, we have performed a systematic convergence study of
1096: the two pseudo-spectral methods. The first pseudo-spectral method
1097: uses the traditional 2/3 dealiasing rule, while the second
1098: pseudo-spectral method uses a high order Fourier smoothing.
1099: The Fourier smoothing method is designed to cut
1100: off the high frequency modes smoothly while retaining a significant
1101: portion of the Fourier modes beyond the 2/3 cut-off point in
1102: the spectral space. We apply both methods to compute nearly
1103: singular solutions of the 1D Burgers equation and the 3D 
1104: incompressible Euler equations. In the case of the 1D Burgers
1105: equation, we can obtain a very accurate approximation of the
1106: exact solution sufficiently close to the singularity time
1107: with 13 digits of accuracy. This allows us to estimate
1108: the numerical errors of the two methods accurately and provides
1109: a solid ground in our convergence study. In our study of
1110: the 3D incompressible Euler equations, we use the highest
1111: resolution that we can afford to perform our convergence study.
1112: In both cases, we demonstrate convincingly that the Fourier smoothing
1113: method gives a more accurate approximation than the 2/3 dealiasing method.
1114: 
1115: Our extensive convergence studies in both physical and spectral
1116: spaces show that the Fourier smoothing method offers
1117: several advantages over the 2/3 dealiasing method when computing
1118: a nearly singular solution. First of all, the error in the
1119: Fourier smoothing method is highly localized near the region
1120: where the solution is most singular. The error in the smooth region 
1121: is several orders of magnitude smaller than that near the 
1122: ``singular'' region. The 2/3 dealiasing method, on the other hand, 
1123: has a wide spread pointwise error distribution, and produces relatively 
1124: large oscillations even in the smooth region. Secondly, for the same 
1125: resolution, the Fourier smoothing method offers a 
1126: more accurate approximation to the physical solution than the 
1127: 2/3 dealiasing method. Our numerical study shows that for a 
1128: given resolution, the Fourier smoothing method retains 
1129: about $12\sim 15\%$ more effective Fourier modes than the 2/3 
1130: dealiasing method in each dimension. For a 3D problem, the gain 
1131: is as large as 20\%. This gain is quite significant in a 
1132: large scale computation. Thirdly, the Fourier smoothing 
1133: method is very stable and robust when computing nearly singular 
1134: solutions of fluid dynamics equations. Spectral convergence is 
1135: clearly observed in all our computational experiments without 
1136: suffering from the Gibbs phenomenon. Moreover, there is no additional 
1137: computational cost in implementing the Fourier 
1138: smoothing method. We have also implemented the Fourier
1139: smoothing method for the incompressible 3D Navier-Stokes equations
1140: and observed a similar performance.
1141: 
1142: We have applied both spectral methods to study the 
1143: potentially singular solution of the 3D Euler equation using 
1144: the same initial condition as Kerr \cite{Kerr93}. Both the
1145: Fourier smoothing method and the 2/3 dealiasing
1146: method give qualitatively the same result except that the
1147: 2/3 dealiasing method suffers from the Gibbs phenomenon
1148: and produces relative large oscillations at late times.  
1149: Our convergence study in both the physical and spectral spaces 
1150: shows that the maximum vorticity does not grow faster
1151: than double exponential in time and the maximum velocity field
1152: remains bounded up to $T=19$, beyond the singularity time
1153: $T = 18.7$ alleged in \cite{Kerr93,Kerr05}. Tremendous
1154: cancellation seems to take place in the vortex stretching
1155: term. The local geometric regularity of the vortex lines
1156: near the region of the maximum vorticity seems to be 
1157: responsible for this dynamic depletion of vortex stretching
1158: \cite{CFM96,DHY05a,DHY05b,HL06}.
1159: 
1160: \vspace{0.2in}
1161: \noindent
1162: {\bf Acknowledgments.}
1163: We would like to thank Prof. Lin-Bo Zhang from the Institute of
1164: Computational Mathematics in Chinese Academy of Sciences (CAS) for 
1165: providing us with the computing resource to perform this large
1166: scale computational project. Additional computing resource was
1167: provided by the Center of High Performance Computing in CAS. We
1168: also thank Prof. Robert Kerr for providing us with his Fortran
1169: subroutine that generates his initial data. This work was in part
1170: supported by NSF under the NSF FRG grant DMS-0353838 and ITR
1171: Grant ACI-0204932. Part of this work was done while Hou visited
1172: the Academy of Systems and Mathematical Sciences of CAS in the
1173: summer of 2005 as a member of the Oversea Outstanding Research
1174: Team for Complex Systems. Li was supported by the National Basic 
1175: Research Program of China under the grant 2005CB321701. Finally,
1176: we would like to thank Professors Alfio Quarteroni, Jie Shen, and
1177: Eitan Tadmor for their valuable comments on our draft manuscript.  
1178: 
1179: \bibliographystyle{amsplain}
1180: \bibliography{bib}
1181: 
1182: \end{document}
1183: 
1184: