nlin0409025/turb.tex
1: \documentclass[12pt]{article}
2: \usepackage{amsmath}
3: \usepackage{psfig}
4: \setlength{\oddsidemargin}{-.05in} 
5: \setlength{\evensidemargin}{00in}
6: \setlength{\textwidth}{6.50in}
7: \setlength{\topmargin}{-.5in}
8: \setlength{\textheight}{8.50in}
9: \renewcommand{\theequation}{\thesection .\arabic{equation}}
10: \renewcommand{\baselinestretch}{1.4}
11: \newcommand{\bfchi}{\mbox{\boldmath $\chi$}}
12: \newcommand{\bfxi}{\mbox{\boldmath $\xi$}}
13: %\newtheorem{Theorem}{Theorem}[section]
14: %\newenvironment{Proof}{P{\sc roof. }}{$\Box$}
15: %\newtheorem{Corollary}{Corollary}[section]
16: %\newtheorem{Lemma}{Lemma}[section]
17: %\def\mybibliography#1{\list{}{\setlength{\leftmargin}{2em}\setlength{\labelsep}{0%pt}\itemindent=-\leftmargin}
18: %\def\newblock{\hskip .11em plus .33 em minus -.07em}\sloppy\clubpenalty4000\widow%penalty4000\sfcode`\.=1000\relax}
19: \newcommand{\bfphi}{\mbox{\boldmath $\phi$}}
20: \newcommand{\bfomega}{\mbox{\boldmath $\omega$}}
21: %\setlength{\parskip}{2mm}
22: %\pagestyle{empty}
23: 
24: \begin{document}
25: 
26: \title{Assessing Turbulence Strength via Lyaponuv Exponents}
27:                    
28: \author{Mayer Humi\\
29: Department of Mathematical Sciences\\
30: Worcester Polytechnic Institute\\
31: 100 Institute Road\\
32: Worcester, MA  01609}
33: 
34: \maketitle
35: \thispagestyle{empty}
36: 
37: \begin{abstract}
38: In this paper we study the link between `turbulence strength' in a flow
39: and the leading Lyaponuv exponent that characterize it. To this end we
40: use two approaches. The first, analytical, considers the truncated
41: convection equations in 2-dimensions with three (Lorenz model) and six
42: components and study their leading Lyaponuv exponent as a function of the
43: Rayleigh number. For the second approach we analyze fifteen time series 
44: of measurements taken by a plane flying at constant height in the upper 
45: troposphere. For each of these time series we estimate the leading Lyaponuv 
46: exponent which we then correlate with the structure constant for the 
47: temperature.
48: \end{abstract}
49: 
50: \newpage
51: 
52: 
53: {\bf Turbulence is still considered as one of the major unsolved problems of
54: modern science. Usually one attempts to characterize the turbulent state 
55: of the system in terms of various numbers such as the Reynolds number,
56: Grashof number, Rayleigh number, etc. How a combination of these numbers 
57: actually characterize the strength of turbulence in the system is
58: not answered easily. The situation is even more complex when the flow is
59: represented only by a time series. Furthermore is there any relation 
60: between this `strength' and chaos theory?  Although chaos owes it modern 
61: origins to the work of Lorenz who investigated the onset of convection, 
62: a characterization of turbulence in terms of dynamical invariants such
63: as Lyapunov exponents, fractal dimension and so on is still an open
64: question which needs further investigation. This paper attempts to 
65: present a modest contribution in this direction.}
66: 
67: \newpage
68: 
69: \section{Introduction}
70: 
71: Atmospheric convection in two dimensions is described [1,2] by the following 
72: system of partial differential equation:
73: \begin{equation}
74: \label{1}
75: \frac{\partial}{\partial t} \nabla^2\psi = -
76: \frac{\partial(\psi,\nabla^2\psi)}{\partial(x,z)} + \nu \nabla^4\psi +
77: \alpha g \frac{\partial\theta}{\partial x}
78: \end{equation}
79: \begin{equation}
80: \label{2}
81: \frac{\partial\theta}{\partial t} =
82: -\frac{\partial(\psi,\theta)}{\partial(x,z)} + \frac{\Delta T}{H}
83:  \;\frac{\partial\psi}{\partial x} + \kappa\nabla^2\theta
84: \end{equation}
85: 
86: Subject to the stress free boundary conditions:
87: 
88: \begin{equation}
89: \psi=0,\,\,\nabla^2\psi=0,\,\,\theta=0 ,\,\,\,\, z=0,1
90: \end{equation}
91: In this system $\psi$ is the stream function and $\theta$ is the potential
92: temperature. The constants $\alpha, \nu, \kappa, g$ denote respectively
93: the coefficient of thermal expansion,the kinematic viscosity,the thermal
94: conductivity and the acceleration of gravity. $H$ is the fluid layer 
95: thickness
96: and $\Delta T$ is the temperature difference between the upper and lower
97: surface of the fluid (which is assumed to be held constant). We also have
98: \begin{equation}
99: \label{3}
100: \frac{\partial(f,g)}{\partial(x,z)} = \left| \begin{array}{cc}
101: \frac{\partial f}{\partial x}    &\frac{\partial f}{\partial z}\\
102: \frac{\partial g}{\partial x}   &\frac{\partial g}{\partial z}
103: \end{array}
104: \right|
105: \end{equation}
106: 
107: 
108: In his famous 1963 paper [2] Lorenz introduced and studied a model in which
109: the solution to eq.(\ref{1})-(\ref{2}) is approximated by a three Fourier modes
110: \begin{equation}
111: \label{4}
112: \psi = \frac{\kappa(1+a^2)\sqrt{2}}{a}X(t)\sin\left(\frac{\pi
113: ax)}{H}\right) \sin\left(\frac{\pi z}{H}\right)
114: \end{equation}
115: \begin{equation}
116: \label{5}
117: \theta = \frac{R_c}{\pi R_a}\left\{\sqrt{2} Y(t) \cos\left(\frac{\pi a
118: x}{H}\right)\sin\left(\frac{\pi z}{H}\right) -
119: Z(t)\sin\left(\frac{2\pi z}{H}\right)\right\}
120: \end{equation}
121: where $a$ is a parameter and
122: \begin{equation}
123: R_a = \frac{g \alpha H^3 \Delta T}{\kappa \nu}
124: \end{equation}
125: \begin{equation}
126: R_c = \frac{\pi^4 (1+a^2)^3}{a^2}
127: \end{equation}
128: are the Rayleigh number and the critical Rayleigh numbers for the flow.
129: This led to the following three coupled equations for $X,Y,Z$.
130: \begin{eqnarray}
131: \label{6}
132: \dot{X} &=& -\sigma X + \sigma Y\\ \nonumber
133: \dot{Y} &=& XZ + rX - Y\\ \nonumber
134: \dot{Z} &=& XY - bZ
135: \end{eqnarray}
136: where
137: \begin{equation}
138: \label{7}
139: \sigma = \frac{\nu}{\kappa},\,\, b = \frac{4}{(1 + a^2)}, \,\, 
140: r = \frac{R_a}{R_c}.
141: \end{equation}
142: 
143:    
144: These equations are usually referred to as the `Lorenz model'. Since its
145: appearance 
146: this model and its implications were studied in great detail in hundreds of 
147: publications with special attention to its bifurcations as a function 
148: of the parameters $\sigma, r$ and $b$. It has been recognized [3] 
149: however that (as expected) the approximation of the solution to eqs 
150: (\ref{1}) and (\ref{2}) which is provided by (\ref{3})-(\ref{7}) becomes 
151: `poor' as $r$ increases.  This has 
152: led several authors to develop and study models with a larger number of 
153: modes [4,21]. 
154: 
155: The basic conjecture that we want to validate in this paper, through some
156: case study, is that stronger turbulence [22] in a system is linked to larger 
157: value of the leading Lyaponuv exponents [5,6] for the the flow. 
158: To test this conjecture we
159: study the Lyaponuv exponents of two truncated systems that were derived form 
160: eqs (\ref{1})-(\ref{2}). We then use Rosenstein algorithm [7] to
161: estimate the 
162: Lyaponuv exponents for fifteen atmospheric time series that were obtained
163: from measurements in the upper troposphere (10Km approximately).
164: The Lyaponuv exponent is then correlated with the structure constant 
165: for the temperature [8,9] which is representative of the 
166: density fluctuations
167: in the atmosphere and hence with the strength of turbulence in the flow.
168: 
169: To begin with we study the dependence of the leading
170: Lyaponuv exponent on the parameter $r$ in the Lorenz model. This parameter
171: is representative of the Rayleigh number in eqs. (1.1)-(1.2).
172: We show that for small values of $r$ the leading Lyaponuv exponent in 
173: the Lorenz model is either negative or small. Increasing
174: $r$ leads at the `onset of chaos' to a steep increase in the value of
175: this exponent. Further increase in $r$ leads to further modest increase 
176: in the value of the leading Lyaponuv exponent (which is indicative of the
177: fact that these modes are 'saturated').
178: 
179: One expects to obtain a better approximation to the solution of eqs.
180: (\ref{1})-(\ref{2}) as the number of modes allowed in the model increases. 
181: Hence as a second step in our study we consider a model with six modes.
182: The leading Lyaponuv exponent for this model displays a similar behavior
183: as for the Lorenz model. However the highest value of the leading Lyaponuv 
184: exponent in this model is larger than the one attained by the Lorenz model. 
185: This is consistent with our conjecture since a model with a larger number 
186: of modes can represent stronger turbulent state of the system due 
187: to the larger number of interacting modes. 
188: 
189: From a practical point of view it is important to assess turbulence
190: strength in a flow from a representative time series of measurements.
191: In this case although the equations governing the flow are well known
192: one has no initial or boundary conditions to simulate them. Furthermore
193: it is impossible to estimate correctly the parameters that appear in these 
194: equations 
195: from the data at hand. To gauge the turbulence strength under these
196: circumstances one has to proceed indirectly. One well known manifestation
197: of turbulence strength in the atmosphere is through density fluctuations
198: (this leads to the well known `twinkling of the stars' phenomena).
199: Hence turbulence strength is related to the value $C_N^2$ (the refraction
200: structure constant [8,9]). In the upper troposphere (at
201: heights of about 10km)
202: the main contributor to $C_N^2$ is $C_T^2$ -the structure constant for the 
203: temperature. It follows then that if our conjecture is correct then
204: the leading Lyaponuv exponent for the temperature time series should 
205: increase as $C_T^2$ increases.
206: 
207: The plan of the paper is as follows:\,\, In section 2 we introduce the 
208: six mode truncated model for eqs (\ref{1}) -(\ref{2}) and study the 
209: behaviors of 
210: the leading Lyaponuv exponent for the Lorenz model and this system as a 
211: function of $r$. In section 3 we give a short overview about the atmospheric
212: structure constants and their computation. In section 4 we consider the time 
213: series of measurements for the temperature and show that there exist a 
214: strong correlation between $C_T^2$ and the leading Lyaponuv exponent. 
215: We end up in section 5 with summary and conclusions.
216: 
217: \setcounter{equation}{0}
218: \section{Analytical Approach}
219: 
220: In an attempt to relate the leading Lyaponuv exponent  to
221: the turbulence strength in the flow we consider a truncated expansion 
222: of eqs.  (\ref{1})-(\ref{2}) with 3 and 6 Fourier components. For the six 
223: components model we let 
224: \begin{eqnarray}
225: \label{21}
226: \psi &=&
227: \frac{\sqrt{2}(1+a^2)\kappa}{a}\left\{\left[X_1(t)\sin\left(\frac{\pi
228: ax}{H}\right) + X_2(t)\sin\left(\frac{2\pi
229: ax}{H}\right)\right]\right.\\ \nonumber
230: &&\sin\left(\frac{\pi z}{H}\right)\left.  +
231: X_3(t)\sin\left(\frac{\pi ax}{H}\right)\sin\left(\frac{2\pi
232: z}{H}\right)\right\}
233: \end{eqnarray}
234: \begin{equation}
235: \label{2.2}
236: \theta = \frac{\sqrt{2}\Delta T R_c}{\pi R_a}
237: \left[Y_1(t)\cos\left(\frac{\pi ax}{H}\right) + Y_2(t)\cos\left(\frac{2\pi
238: ax}{H}\right)\right]\sin\left(\frac{\pi z}{H}\right) - \frac{\Delta
239: T R_a}{\pi R_c} Z(t)\sin\left(\frac{2\pi z}{H}\right)
240: \end{equation}
241: which leads after some lengthy algebra to the following six equations
242: for ${X_1,X_2,X_3,Y_1,Y_2,Z}$
243: \begin{equation}
244: \label{23}
245: \dot{X}_1 = \sigma Y_1 - \sigma
246: X_1+\frac{9}{4}\frac{(a^2-1)\sqrt{2}}{1 + a^2} X_2X_3
247: \end{equation}
248: \begin{equation}
249: \label{24}
250: \dot{X}_2 = \frac{2\sigma(1+a^2)}{1+4a^2}Y_2 - \frac{\sigma(1+4a^2)}{1+a^2}
251: X_2 + \frac{9\sqrt{2}}{4(1+4a^2)}\;X_1X_3
252: \end{equation}
253: \begin{equation}
254: \label{25}
255: \dot{X}_3 = -\frac{\sigma(4+a^2)}{1+a^2}\;X_3 -
256: \frac{9a^2\sqrt{2}}{4(4+a^2)} X_1X_2
257: \end{equation}
258: \begin{equation}
259: \label{26}
260: \dot{Y}_1 = \frac{3}{4}\sqrt{2} X_3Y_2 - X_1 Z - Y_1 + rX_1
261: \end{equation}
262: \begin{equation}
263: \label{27}
264: \dot{Y}_2 = -\frac{3}{4}\sqrt{2} X_3Y_1 - \frac{(1+4a^2)}{1+a^2}Y_2 -
265: 2X_2Z + 2rX_2
266: \end{equation}
267: \begin{equation}
268: \label{28}
269: \dot{Z} = X_1Y_1 + 2X_2Y_2 - bZ
270: \end{equation}
271: 
272: 
273: For the systems (\ref{5})-(\ref{7}) and (\ref{21})-(\ref{28}) we applied
274: (the analytical) Wolf 
275: algorithm [10] to compute the leading Lyaponuv exponent as a function
276: of $r$. The results of these computations are shown in Fig. 1 for
277: values of $r$ in the range of [10,115]. We observe that both of these systems 
278: represent a truncation of the original equations. As a result the flow 
279: which is represented by eqs.(\ref{1})-(\ref{2}) can not be approximated well by 
280: the solution of the truncated equations when $r$ is large as additional modes 
281: become active. As a result the relationship between the solutions of eqs.
282: (\ref{5})-(\ref{7}) and (\ref{23})-(\ref{28}) and actual convective turbulence 
283: phenomena is lost as $r$ becomes large. 
284: 
285: From Fig. 1 we see that for both truncated systems under consideration
286: the leading Lyaponuv exponent remain small or negative for small values
287: of $r$. It then ``bifurcates'' and goes up sharply at the onset of convective 
288: state and increases modestly as $r$ is increased further. For the
289: three mode model as $r$ increases over 70 the modes become saturated 
290: and the model ceases to approximate the true solution of the convective 
291: problem. It then follows it own characteristics. 
292: 
293: For the six mode solution, the leading Lyaponuv exponent 
294: follows the same trajectory as the three mode model up
295: to $r ~ 70$. However for higher values of $r$ the additional modes in this
296: model become active and as a result the leading Lyaponuv exponent continue 
297: to increase up to $r ~ 110$ (where the modes of this model get saturated).
298: 
299: The obvious explanation of this behavior of the leading Lyaponuv exponent
300: is that the  
301: the six mode system provides a better approximation to the solution of eqs.
302: (\ref{1})-(\ref{2}) than the three mode system. As a result the six mode system
303: captures more of the turbulence phenomena which are related to the solution
304: of the original system. Accordingly Fig. 1 confirms our conjecture and shows
305: that the flow in a system with stronger turbulence has also a larger
306: leading Lyaponuv exponent. 
307: 
308: \setcounter{equation}{0}
309: \section{Structure Constants}
310: 
311: The structure function of a geophysical variable e.g. the temperature is
312: defined as
313: 
314: \begin{equation}
315: D({\bf r})= \langle T^\prime[({\bf r}_1 + {\bf r}),T^\prime ({\bf r}_1)]^{2}\rangle
316: \end{equation}
317: 
318: where $T^\prime$ are the turbulent fluctuations in the temperature and 
319: ${\bf r}$ is the vector from one point to another.
320: 
321: Kolmogorov showed that for isotropic turbulence in the inertial range this
322: function depends only on $d$ = $|{\bf r}|$ and scales as
323: 
324: \begin{equation}
325: D(d) = C_T^{2} d^{2/3} .
326: \end{equation}
327: 
328: $C_T^2$ which appears as the proportionality constant in this equation
329: is referred to as the ``temperature structure constant''.
330: 
331: The determination of the atmospheric structure constants [8,9,11,12]
332: and in particular the temperature structure constant $C_T^2$ is important
333: in many applications e.g the propagation of electromagnetic signals
334: [8,9]. Local peaks in the values of these constants, which are indicative 
335: of strong turbulence and reflect on the structure of the atmospheric flow, 
336: can have a negative effect on the operation of various optical instruments.
337: 
338: To estimate these structure constants in the upper troposphere or the
339: stratosphere it is a common practice to send high flying airplanes
340: that collect data about the basic meteorological variables (such as wind,
341: temperature and pressure) along its flight path which may extend up
342: to 200 kms. To estimate the averaged value of the structure constants
343: along this path one must decompose first the meteorological data into
344: mean flow, waves and turbulent residuals [13,14]. From the spectrum of the
345: turbulent residuals one can estimate an averaged value of the structure
346: constants using Kolmogorov inertial range scaling and Taylor's frozen
347: turbulence hypothesis [15]. For $C_T^2$ in particular we have
348: \begin{equation}
349:             C_T^2(k) = 4 F(k) k^{5/3}
350: \end{equation}
351: where $F(k)$ is the temperature spectral density in the inertial range
352: and $k$ is the wave number. An averaged value for $C_T^2$ (over all wave
353: numbers in the inertial range) is obtained by averaging these
354: values over $k$.
355: 
356: In the upper troposphere where humidity is low $C_T^2$ is the main
357: contributor to the refraction structure constant $C_N^2$ [8,9] which has 
358: an important impact on the operation of ground telescopes and the 
359: twinkling of the stars.
360: 
361: \section{Analysis of Atmospheric Time Series}
362: 
363: During 1999 a series of measurements were made over Australia and Japan
364: by a specially equipped aeroplane [14]. These measurements were taken with 
365: sampling frequency of 55.1 Hz along a flight path of 200 kms (approximately). 
366: For each data set the plane flew at almost constant height along a 
367: straight line at an approximate speed of 103m/sec [14,15].
368: 
369: To use this data to estimate $C_T^2$ we have to split the original
370: measurements into a sum of mean flow, waves and turbulent residuals.
371: To accomplish this task we used Karahunen-Loeve(K-L) decomposition
372: algorithm which has been used by many researchers [15,16].
373: Actually the last paper applies the K-L decomposition to
374: the same data considered here and the details of the decomposition
375: which were given there will not be repeated here.
376: 
377: Based on instrument specifications the data noise should be at a relative
378: error level of $10^{-3}$. This is confirmed by the eigenvalues obtained
379: in the K-L decomposition where the last few eigenvalues (which reflect
380: the noise level in the data) are of order $10^{-3}$ of the leading 
381: eigenvalue.
382: 
383: For fifteen data sets that were obtained during these flights averaged 
384: values for $C_T^2$ were obtained using the methodology described above 
385: [8,9,10].
386: 
387: To compute the leading Lyaponuv exponents for these time series we used
388: Rosenstein algorithm and its implementation in the TISEAN package [7,17]. 
389: (However one should note that other algorithms are available for this 
390: purpose [18]).
391: To apply this algorithm one has to determine first the `optimal'
392: delay coordinates and embedding dimension. These were determined using the
393: mutual information and false neighborhood algorithms [5,19,20] which are 
394: also implemented in the above mentioned package. This analysis led us to 
395: choose a four dimensional embedding space with delay coordinates of 600
396: data points. Fig 2. presents a log-log plot of the results 
397: for this exponent versus $C_T^2$. We also show on this plot the least
398: square line for this data. From this plot we see that a change in $C_T^2$
399: over three orders of magnitudes correlates well with the leading Lyaponuv 
400: exponents. The fluctuations around the least squares line can be attributed
401: to wave activity and possible measurements errors. This demonstrates that
402: the Lyaponuv exponent can be used as a second global measure of turbulence
403: strength in the data. In cases of discrepancy between $C_T^2$ and 
404: the leading Lyaponuv exponent one must trace out the reasons for this 
405: mismatch and correct them.
406: 
407: \section{Summary and Conclusions}
408: 
409: Our main objective in this paper was to link turbulence strength to
410: the leading Lyaponuv exponent that is related to the flow. This program
411: was carried out in two contexts. In the first we used a truncated expansion 
412: of the flow and tested for leading Lyaponuv exponent of the resulting dynamical 
413: system. In the second part we tested this hypothesis for time series
414: which represent the flow. In both instances we found a strong supporting 
415: evidence for our conjecture. Although this does not constitutes a proof
416: of this conjecture in general we still feel that our results will
417: useful in many applied contexts especially as a check for the validity of
418: other global invariants that characterize the flow. In particular the 
419: determination of the leading Lyaponuv exponent can help verify the value
420: of $C_T^2$ (or other structure constants) that have important practical 
421: applications. 
422: 
423: 
424: \section*{References}
425: 
426: \begin{itemize}
427: 
428: \item[1] Saltzman, B., 1962:\, Finite amplitude free convection as an 
429: initial value problem, {\it J. Atmos. Sci}, {\bf 19}, 329-341.
430: 
431: \item[2] Lorenz, E.N., 1963:\,  Deterministic nonperiodic flow,
432: {\it J. Atmos. Sci}, {\bf 20}, 130-141.
433: 
434: \item[3] Marcus, P. S., 1981:\, Effects of truncation in modal representation
435: of thermal convection, {\it J. Fluid Mech.}, {\bf 103}, 241-255. 
436: 
437: \item[4] Curry, J. H., 1978:\, A generalized Lorenz system, 
438: {\it Commun. Math. Phys.}, {\bf 60}, 193-204.
439: 
440: \item[5] Abarbanel, H. D. I., Brown, R., Sidorowich, J.J., and Tsimring L.S.,
441: 1993:\, The analysis of observed chaotic data in physical systems,
442: {\it Rev. Mod. Phys.}, {\bf 65}, 1331.
443: 
444: \item[6] Ott, E., Sauer, T., Yorke, J. A., 1994:\, Coping with Chaos, 
445: Wiley, NY. 
446: 
447: \item[7] Rosenstein, M. T., Collins, J. J., and De Luca, C. J., 1993:\,
448: {\it Physica D}, {\bf 65}, 117.
449: 
450: \item[8] Panofsky, H.A., and Dutton, J.A., 1989:\,  Atmospheric
451: Turbulence,  Wiley, NY.
452: 
453: \item[9] Dewan, E., 1980: \, Optical Turbulence Forecasting,  {\it Air Force
454: Geophysical Lab.}, AFGL-TR-80-0030, Lab. Report.
455: 
456: \item[10] Fanta, R.L., 1985:\,  Wave Propagation in Random Media,  {\it
457: Progress in Optics}, {\bf 22}, 342-399.
458: 
459: \item[11] Wolf, A, Swift, J. B., Swinney, H.L., and Vastano, J.A., 1985:\;
460: Determining Lyapunov exponents from a time series, {\it Physica D},
461: {\bf 16}, 285-317.
462: 
463: \item[12] Humi, M., 2004:\; Estimation of Atmospheric Structure Constants
464: from Airplane Data, {\it J. of Atmospheric and Oceanic Technology},
465: {\bf 21}, 495-500. 
466: 
467: \item[13] Jumper, G.Y. and Beland, R.R., Progress in the understandings and
468: modeling of atmospheric optical turbulence, {\it AIAA}, 2000-2355,
469: Proc. 31st AIAA Plasmadynamics and Lasers Conference, 19-22 June 2000,
470: Denver, CO.
471: 
472: \item[14] Cote, O.R., Hacker J.M., Crawford, T.L., and Dobosy, R.J., 2000:\, 
473: Clear air turbulence and refractive turbulence in the upper troposphere
474: and lower stratosphere, Ninth conference on Aviation, Range and Aerospace
475: Meteorology, Paper No. 16386, Sep 2000.
476: 
477: \item[15] Humi, M., 2003:\, Approximate Simulation of Stratospheric Flow 
478: from Aircraft Data,
479: {\it Inter. J. Numerical Methods in Fluids}, {\bf 41}, 209-223.
480: 
481: \item[16] Penland, C., Ghil, M., and Weickmann, K.M., 1991:\, Adaptive filtering
482: and maximum entropy spectra with applications to changes in atmospheric
483: angular momentum, {\it J. Geo. Res.}, {\bf 96}, 22659-22671.
484: 
485: \item[17] Hegger, R,  Kantz, H., and Schreiber, T., 1991:\, Practical 
486: implementation of nonlinear time series methods:\, The TISEAN package, 
487: Chaos {\bf 9}, 413.
488: 
489: \item[18] Liu, H., Yang, Y., Dai, Z., Yu, Z., 2003:\, The largest Lyapunov
490: exponent of a chaotic dynamical system in scale space and its
491: application, {\it Chaos}, {\bf 13}, 839-844.
492: 
493: \item[19] Fraser, A.M. and Swinney, H.L., 1986:\, Independent coordinates for 
494: strange attractors from mutual information, Phys. Rev. A {\bf 33}, 1134.
495: 
496: \item[20] Kennel, M.B., Brown, R. and Abarbanel, H.D.I., 1992:\, Determining 
497: embedding dimension for phase-space reconstruction using a geometrical 
498: construction, Phys. Rev. A, {\bf 45}, 3403.
499: 
500: \item[21] Reiterer P.,Lainscsek C., Schurrer F., Letellier C. and Maquet J.
501: 1998: A nine dimensional Lorenz system to study high dimensional chaos
502: J. Phys. A {\bf 31} 7121-7139.
503: 
504: \item[22] Frisch U. and Orszag S. -Turbulence: Challenges for theory and 
505: experiment, Physics today {\bf 43} ,24-29 
506: 
507: \end{itemize}
508: 
509: \newpage
510: 
511: \centerline{\Large{\bf List of Captions}}
512: 
513: Fig. 1: Plot of the leading Lyapunov exponent versus $r=\frac{R_a}{R_c}$
514: for the Lorenz model(dashed line) and the 6-mode model (solid line)
515: which was developed in Sec.2.
516: 
517: Fig. 2: Log-log plot of E - the leading Lyapunov exponent- versus $C_T^2$. 
518: The first order least squares fit for this data yields the relation
519: $E=10.85*(C_T^2)^(0.2760)$ .
520: 
521: \newpage
522: 
523: \begin{figure}[htb]
524: \centering
525: \centerline{\psfig{file=fig01.ps,height=4in,width=5.5in}}
526: \caption{}
527: \end{figure}
528: 
529: \newpage
530: 
531: \begin{figure}[htb]
532: \centering
533: \centerline{\psfig{file=fig02.ps,height=4in,width=5.5in}}
534: \caption{}
535: \end{figure}
536: 
537: \end{document} 
538: