nlin0512007/chm.tex
1: \documentclass{jfm}
2: %\documentclass[referee]{jfm}
3: \usepackage{graphicx,bm}
4: \def\dt{{\rm d}t}
5: \def\d{{\rm d}}
6: \def\dk{{\rm d}k}
7: \def\dx{{\rm d}{\bm x}}
8: \def\x{{\bm x}}
9: \def\k{{\bm k}}
10: \def\etal{{\it et al. }}
11: \def\norm#1{\left|\mkern-2mu\left|#1\right|\mkern-2mu\right|}
12: 
13: \title[Impeded inverse transfer in the CHM model of quasi-geostrophic 
14: flows]{Impeded inverse energy transfer in the Charney--Hasegawa--Mima 
15: model of quasi-geostrophic flows}
16: \author[C.V. Tran and D. G. Dritschel]
17: {C\ls H\ls U\ls O\ls N\ls G\ns V.\ns T\ls R\ls A\ls N \ns \and
18: D\ls A\ls V\ls I\ls D\ns G.\ns D\ls R\ls I\ls T\ls S\ls C\ls H\ls E\ls L}
19: \affiliation{School of Mathematics and Statistics, University of St Andrews, 
20: St Andrews KY16 9SS, UK}
21: 
22: %\pubyear{2005}
23: %\volume{538}
24: %\pagerange{119--126}
25: %\date{12 September 2005}
26: %\setcounter{page}{119}
27: 
28: \begin{document}
29: 
30: \maketitle
31: 
32: \begin{abstract}
33: 
34: The behaviour of turbulent flows within the single-layer quasi-geostrophic
35: (Charney--Hasegawa--Mima) model is shown to be strongly dependent on the
36: Rossby deformation wavenumber $\lambda$ (or free-surface elasticity). 
37: Herein, we derive a bound on the inverse energy transfer, specifically on 
38: the growth rate $\d\ell/\dt$ of the characteristic length scale $\ell$ 
39: representing the energy centroid. It is found that 
40: $\d\ell/\dt\le2\norm q_\infty/(\ell_s\lambda^2)$, where $\norm q_\infty$ 
41: is the supremum of the potential vorticity and $\ell_s$ represents the 
42: potential enstrophy centroid of the reservoir, both invariant. This result 
43: implies that in the potential energy dominated regime 
44: ($\ell\ge\ell_s\gg\lambda^{-1}$), the inverse energy transfer is strongly 
45: impeded, in the sense that under the usual time scale no significant 
46: transfer of energy to larger scales occurs. The physical implication is 
47: that the elasticity of the free surface impedes turbulent energy transfer 
48: in wavenumber space, effectively rendering large-scale vortices long-lived 
49: and inactive. Results from numerical simulations of forced-dissipative 
50: turbulence confirm this prediction.
51: 
52: \end{abstract}
53: 
54: \section{Introduction}
55: 
56: The atmosphere and oceans are hugely influenced by the background planetary 
57: rotation and stable density stratification. These features virtually suppress 
58: motion in the vertical direction, rendering geophysical fluid systems, 
59: intrinsically vast in their horizontal extent compared with their limited 
60: depth, approximately two-dimensional (2D) at large scales. Many models 
61: derived from the theory of quasi-geostrophy exploit this simplification.
62: One such model, governed by the Charney--Hasegawa--Mima (CHM) equation, 
63: is the subject of the present study.
64: 
65: The CHM equation, which describes the evolution of potential vorticity in
66: a shallow fluid layer with a free surface, takes the form (Pedlosky 1987)
67: \begin{eqnarray}
68: \label{CHM}
69: \frac{\partial q}{\partial t}+J(\psi,q) &=& 0.
70: \end{eqnarray}
71: Here $\psi(x,y,t)$, the streamfunction, is assumed to be periodic in both
72: $x$ and $y$. Note that $\psi$ is proportional to the free surface height
73: anomaly. The quantity $q=(\Delta-\lambda^2)\psi$ is the potential 
74: vorticity,\footnote{This form of potential vorticity ignores the 
75: so-called $\beta$-term. For some studies concerning the effects of 
76: this term, see Kukhakin \& Orszag (1996), Okunu \& Masuda (2003) and 
77: Smith (2004).} where $\lambda$ is the Rossby deformation wavenumber 
78: ($\lambda=f/c$, where $f$ is the Coriolis frequency and $c$ is the 
79: small-scale gravity waves speed, both assumed constant). The differential 
80: operators $J(\cdot,\cdot)$ and $\Delta$ are, respectively, the Jacobian 
81: and two-dimensional Laplacian. Eq.~(\ref{CHM}) expresses material 
82: conservation of potential vorticity. It also governs the evolution of 
83: quasi-2D fluctuations of the electrostatic potential in a plane 
84: perpendicular to a strong magnetic field applied uniformly to a plasma, 
85: in which case $\psi$ is the electrostatic potential and $\lambda^{-1}$ 
86: is the ion Larmor radius (Hasegawa \& Mima 1978). 
87: 
88: The CHM equation has been a subject of active research for decades (see 
89: for example, Hasegawa \& Mima 1978; Fyfe \& Montgomery 1979; Yanase \& 
90: Yamada 1984; Larichev \& McWilliams 1991; Ottaviani \& Krommes 1992; 
91: Kukharkin, Orszag \& Yakhot 1995; Iwayama, Shepherd \& Watanabe 2002; 
92: Arbic \& Flierl 2003; Tran \& Bowman 2003). A majority of these studies 
93: have been concerned with the dual transfer of the total energy 
94: $\langle-\psi q\rangle/2=\langle|\nabla\psi|^2\rangle/2+\lambda^2\langle
95: \psi^2\rangle/2$ and potential enstrophy $\langle\Delta\psi q\rangle/2=
96: \langle|\Delta\psi|^2\rangle/2+\lambda^2\langle|\nabla\psi|^2\rangle/2$, 
97: which are conserved by the CHM nonlinearities. Here $\langle\cdot\rangle$ 
98: denotes a spatial average. The kinetic energy and the usual enstrophy are 
99: $\langle|\nabla\psi|^2\rangle/2$ and $\langle|\Delta\psi|^2\rangle/2$, 
100: respectively (these are not conserved). The quantity 
101: $\lambda^2\langle\psi^2\rangle/2$ is the potential energy. In addition
102: to the total energy and potential enstrophy, the CHM dynamics also 
103: conserve an infinite class of potential vorticity norms, including
104: the $L^\infty$ norm $\norm q_\infty$. Similar to 2D Navier--Stokes (NS) 
105: turbulence, the dual conservation law implies that the total energy 
106: (potential enstrophy) is preferentially transferred to lower (higher) 
107: wavenumbers. Unlike 2D NS turbulence, each of the invariants in CHM 
108: turbulence consists of two components. The presence of the second 
109: component is due to the elasticity of the free surface, which is known 
110: to have a fundamental influence on the dynamics. A remarkable effect is 
111: that the inverse energy transfer for wavenumbers $k\ll\lambda$ (the 
112: so-called potential energy regime)\footnote{For scales smaller than 
113: $\lambda^{-1}=c/f$, the motion is inherently three-dimensional (cf. 
114: Dritschel \& de la Torre Ju\'arez 1996; Dritschel, de la Torre Ju\'arez 
115: \& Ambaum 1999) and therefore poorly described by (\ref{CHM}).} is 
116: predominantly a transfer of the potential energy (Tran \& Bowman 2003), 
117: i.e. of the quadratic quantity $\langle\psi^2\rangle$. 
118: 
119: Numerical studies of CHM turbulence appear to suggest that in the 
120: potential energy regime, the inverse cascade is strongly impeded. For 
121: turbulence decaying from an initial energy reservoir, \cite{Larichev91} 
122: observe a ``slowness'' in the turbulent evolution. Iwayama \etal (2002) 
123: notice the formation of a sharp spectral peak, a manifestation of an 
124: inverse energy transfer with limited extent in wavenumber space. 
125: \cite{Arbic03} find that the inverse cascade slows down and that vortex 
126: merging takes a very long time. For forced turbulence, Kukharkin \etal 
127: (1995) observe the formation of vortical ``quasicrystals'', states of 
128: long-lived inactive vortices. While these observations are consistent 
129: with the finding of Tran \& Bowman (2003) that the inverse energy transfer 
130: involves virtually no kinetic energy when $k\ll\lambda$, they are not 
131: fully explained by this fact. A more satisfactory explanation would be 
132: that the inverse transfer of potential energy is also impeded in this regime. 
133: 
134: In this Note we derive an upper bound for the growth rate of a 
135: characteristic length scale associated with the inverse energy transfer.
136: We infer from this upper bound that in the potential energy regime,
137: the inverse energy transfer is strongly impeded by the elasticity of
138: the free surface. An interesting physical interpretation of this effect 
139: is that this elasticity prohibits the merger of large-scale vortices to 
140: form larger energy-carrying scales, effectively rendering such vortices 
141: long-lived and inactive. Although this effect is deduced in the framework 
142: of unforced and inviscid dynamics, it also manifests itself in a 
143: forced-dissipative version of (\ref{CHM}), as is illustrated by some 
144: numerical results. 
145: 
146: \section{Inverse energy transfer}
147: 
148: The nonlinear transfer of energy to lower wavenumbers of an energy reservoir
149: (when it occurs) necessarily entails an increase of a suitably defined length 
150: scale representing the energy centroid of the reservoir. Here we consider the 
151: length scale $\ell$ defined by  
152: \begin{eqnarray}
153: \label{ell}
154: \ell=\frac{\langle(-\Delta)^{-1/2}\psi q\rangle}{\langle\psi q\rangle}
155: =\frac{\langle|(-\Delta)^{1/4}\psi|^2\rangle+\lambda^2\langle|(-\Delta)^
156: {-1/4}\psi|^2\rangle}{\langle|\nabla\psi|^2\rangle+\lambda^2\langle
157: \psi^2\rangle}. 
158: \end{eqnarray}
159: In (\ref{ell}) the operator $(-\Delta)^\theta$ is defined by
160: $(-\Delta)^\theta\widehat\psi(\k)=k^{2\theta}\widehat\psi(\k)$, where
161: $k=|\k|$ is the wavenumber and $\widehat\psi(\k)$ the Fourier transform
162: of $\psi(\x)$. The growth rate $\d\ell/\dt$ provides a quantitative 
163: measure of the inverse transfer. In particular, a smaller rate 
164: corresponds to an inverse transfer with a more limited extent in 
165: wavenumber space than one with a greater rate. In this respect, 
166: $\d\ell/\dt$ may be considered as a measure of locality of the 
167: inverse transfer, in the sense that for a given energy distribution 
168: a smaller rate implies a more local transfer than a greater one.
169: 
170: In the case where most of the kinetic energy resides in wavenumbers 
171: $k\ll\lambda$ (the potential energy regime),  
172: $\langle|(-\Delta)^{1/4}\psi|^2\rangle$ is negligible compared with 
173: $\lambda^2\langle|(-\Delta)^{-1/4}\psi|^2\rangle$, and similarly 
174: $\langle|\nabla\psi|^2\rangle$ is negligible compared with
175: $\lambda^2\langle\psi^2\rangle$. More precisely, in this regime we 
176: have $$\frac{\langle|(-\Delta)^{1/4}\psi|^2\rangle}{\lambda^2\langle
177: |(-\Delta)^{-1/4}\psi|^2\rangle}\le\frac{\langle|\nabla\psi|^2\rangle}
178: {\lambda^2\langle\psi^2\rangle}\ll1,$$ where the first inequality is due
179: to H\"older's inequality (see also Tran \& Shepherd 2002 and Tran 2004), 
180: implying 
181: $\ell\approx\langle|(-\Delta)^{-1/4}\psi|^2\rangle/\langle\psi^2\rangle$. 
182: For an energy reservoir in the potential energy regime, the length scale 
183: $\ell$ specifies where, in wavenumber space, most of $\langle\psi^2\rangle$ 
184: (or equivalently most of the potential energy) is distributed. Its growth 
185: rate is thus the rate at which the spectral profile of the potential 
186: energy proceeds toward lower wavenumbers.
187: 
188: The evolution equation for the characteristic length scale $\ell$, defined 
189: in the preceding paragraphs, is obtained by multiplying (\ref{CHM}) by 
190: $(-\Delta)^{-1/2}\psi/\langle\psi q\rangle$ and taking the spatial average 
191: of the resulting equation:
192: \begin{eqnarray}
193: \label{evol1}
194: \frac{1}{2}\frac{\d\ell}{\dt} &=&
195: \frac{\langle(-\Delta)^{-1/2}\psi J(\psi,q)\rangle}
196: {\langle|\nabla\psi|^2\rangle+\lambda^2\langle\psi^2\rangle}
197: =-\frac{\langle q J(\psi,(-\Delta)^{-1/2}\psi)\rangle}
198: {\langle|\nabla\psi|^2\rangle+\lambda^2\langle\psi^2\rangle}\nonumber\\
199: &\le&
200: \frac{\langle|q||\nabla\psi||\nabla(-\Delta)^{-1/2}\psi|\rangle}
201: {\langle|\nabla\psi|^2\rangle+\lambda^2\langle\psi^2\rangle}
202: \le \frac{\langle|\nabla\psi|^2\rangle^{1/2}\langle\psi^2\rangle^{1/2}}
203: {\langle|\nabla\psi|^2\rangle+\lambda^2\langle\psi^2\rangle}
204: \norm q_\infty,
205: \end{eqnarray} 
206: where the various manipulations of the nonlinear term are straightforward.
207: In particular, the last step can be readily seen by expressing the 
208: streamfunction in terms of a Fourier series.
209: 
210: The ratio in front of $\norm q_\infty$ on the right-hand side of 
211: (\ref{evol1}) can be closely estimated in terms of conserved quantities. 
212: For that purpose, let us denote by $\ell_s$ the conserved length scale 
213: representing the enstrophy centroid of the reservoir:
214: \begin{eqnarray}
215: \label{enscentroid}
216: \ell_s &=& 
217: \frac{\langle-\psi q\rangle^{1/2}}{\langle\Delta\psi q\rangle^{1/2}}=\frac
218: {\left(\langle|\nabla\psi|^2\rangle+\lambda^2\langle\psi^2\rangle\right)^{1/2}}
219: {\left(\langle|\Delta\psi|^2\rangle+\lambda^2\langle|\nabla\psi|^2\rangle
220: \right)^{1/2}}.
221: \end{eqnarray} 
222: Solving (\ref{enscentroid}) for $\langle\psi^2\rangle$ and 
223: substituting the result into the said ratio we obtain
224: \begin{eqnarray}
225: \label{ratio}
226: \frac{\langle|\nabla\psi|^2\rangle^{1/2}\langle\psi^2\rangle^{1/2}}
227: {\langle|\nabla\psi|^2\rangle+\lambda^2\langle\psi^2\rangle} 
228: &=& \frac{\langle|\nabla\psi|^2\rangle^{1/2}
229: \left[\ell_s^2\langle|\Delta\psi|^2\rangle+(\ell_s^2\lambda^2-1)\langle
230: |\nabla\psi|^2\rangle\right]^{1/2}}{\lambda\left[\langle|\nabla\psi|^2
231: \rangle+\ell_s^2\langle|\Delta\psi|^2\rangle+(\ell_s^2\lambda^2-1)\langle
232: |\nabla\psi|^2\rangle\right]}\nonumber\\ 
233: &\le& \frac{\langle|\nabla\psi|^2\rangle^{1/2}
234: \left(\langle|\Delta\psi|^2\rangle+\lambda^2\langle
235: |\nabla\psi|^2\rangle\right)^{1/2}}{\ell_s\lambda\left(\langle|
236: \Delta\psi|^2\rangle+\lambda^2\langle
237: |\nabla\psi|^2\rangle\right)}\nonumber\\ 
238: &=& \frac{\langle|\nabla\psi|^2\rangle^{1/2}}{\ell_s\lambda
239: \left(\langle|\Delta\psi|^2\rangle+\lambda^2\langle
240: |\nabla\psi|^2\rangle\right)^{1/2}}
241: \le\frac{1}{\ell_s\lambda^2}.
242: \end{eqnarray} 
243: This is a very sharp estimate, especially for a reservoir in the 
244: potential energy regime. We note that in general the inequality 
245: $\ell\ge\ell_s$ holds. This is a consequence of the H\"older inequality
246: and can be shown as follows. Let us denote by $U(k)$ the power spectrum 
247: of $\langle-(-\Delta)^{-1/2}\psi q\rangle$. The length sales $\ell$
248: and $\ell_s$ can then be expressed in terms of $U(k)$ as 
249: $$\ell=\frac{\int U(k)\,\dk}{\int kU(k)\,\dk}$$ and 
250: $$\ell_s=\left(\frac{\int kU(k)\,\dk}{\int k^3U(k)\,\dk}\right)^{1/2}.$$ 
251: Now by H\"older's inequality we have 
252: $$\int kU(k)\,\dk\le\left(\int U(k)\,\dk\right)^{2/3}
253: \left(\int k^3U(k)\,\dk\right)^{1/3},$$ from which the inequality
254: $\ell\ge\ell_s$ indeed follows immediately.
255: 
256: Putting these results together we have
257: \begin{eqnarray}
258: \label{evol2}
259: \frac{\d\ell}{\dt} &\le& 2\frac{\norm q_\infty}{\ell_s\lambda^2}
260: \le 2\frac{\norm q_\infty}{\ell_s^2\lambda^2}\ell
261: \end{eqnarray}
262: using $\ell\ge\ell_s$. Hence the exponential growth rate of $\ell$ is 
263: bounded from above by $2\norm q_\infty/(\ell_s^2\lambda^2)$. Given a 
264: finite $\norm q_\infty$, the upper bound $2\norm q_\infty/(\ell_s^2\lambda^2)$ 
265: can be made arbitrarily small provided $\ell_s^2\lambda^2$ is sufficiently 
266: large. This condition is satisfied for an initial energy reservoir in a 
267: sufficiently remote wavenumber region of the potential energy regime,
268: where both $\ell_s\lambda\gg1$ and $\ell\lambda\gg1$. Under these 
269: circumstances, no significant growth of $\ell$ occurs when the initial 
270: energy reservoir freely spreads out, and hence no significant inverse 
271: energy transfer is possible. In passing it is worth mentioning that the 
272: upper bound for the exponential growth rate of $\ell$ in (\ref{evol2}) 
273: consists of the constant $\lambda$ and the two invariants $\norm q_\infty$ 
274: and $\ell_s$, and therefore can be completely determined by initial data.
275: 
276: The result in the preceding paragraph is an analytic version of the 
277: qualitative statement in the literature that the turbulence evolves 
278: on a rescaled (slower) time (cf. Larichev \& McWilliams 1991; Kukharkin, 
279: Orszag \& Yakhot 1995). To clarify this point let us rewrite 
280: (\ref{evol2}) as
281: \begin{eqnarray}
282: \label{evol3}
283: \ell_s^2\lambda^2\frac{\d\ell}{\dt} &\le& 2\norm q_\infty\ell.
284: \end{eqnarray}
285: The left-hand side of (\ref{evol3}) is the (instantaneous) growth rate 
286: of $\ell$ on the slow time $\tau=t/(\ell_s^2\lambda^2)$. For an initial
287: reservoir with a finite $\norm q_\infty$, the inverse transfer of 
288: potential energy can be considerable only on this slow time scale. In 
289: the limit $\ell_s\lambda\rightarrow\infty$, the invariants become 
290: $\langle\psi^2\rangle/2$ and $\langle|\nabla\psi|^2\rangle/2$ and the 
291: length scales $\ell$ and $\ell_s$ become
292: $\langle|(-\Delta)^{-1/4}\psi|^2\rangle/\langle\psi^2\rangle$ and 
293: $\langle\psi^2\rangle^{1/2}/\langle|\nabla\psi|^2\rangle^{1/2}$, 
294: respectively. The rescaled time $\tau$ becomes infinitely long. 
295: 
296: An estimate which may be significantly sharper than (\ref{evol2}) can be 
297: derived by replacing the potential vorticity $q$ in (\ref{evol1}) by the 
298: relative vorticity $\Delta\psi$. By doing that, instead of (\ref{evol2}), 
299: we obtain
300: \begin{eqnarray}
301: \label{evol4}
302: \frac{\d\ell}{\dt} &\le& 2\frac{\norm{\Delta\psi}_\infty}
303: {\ell_s^2\lambda^2}\ell.
304: \end{eqnarray}
305: In remote regions of the potential energy regime, where 
306: $\ell\lambda\ge\ell_s\lambda\gg1$, we expect $\lambda^2\psi$ to 
307: dominate $\Delta\psi$. Therefore in these regions it is plausible that
308: $\norm q_\infty\gg\norm{\Delta\psi}_\infty$, allowing (\ref{evol4}) to 
309: become significantly sharper than (\ref{evol2}). The price for this 
310: improvement is that $\norm{\Delta\psi}_\infty$ is not conserved. Note
311: that in such regions $\ell$ and $\ell_s$ are the length scales of 
312: potential and kinetic energy, respectively.
313: 
314: A straightforward physical interpretation of the above result is that 
315: the elasticity of the free surface suppresses energy transfer for scales 
316: larger than the Rossby deformation radius. This effect has a profound 
317: influence on the vortex dynamics. Let us consider a hypothetical flow 
318: consisting of vortices (having smooth potential vorticity---no small 
319: length scales) of comparable scales, which are much larger than the 
320: deformation radius $\lambda^{-1}$. For these scales the inverse energy 
321: transfer is strongly impeded, allowing for virtually no formation of 
322: larger energy-carrying scales. This keeps the vortices from merging, 
323: effectively establishing a state of long-lived inactive vortices. This 
324: may explain the crystallisation of vortices observed by Kukharkin \etal 
325: (1995). The impeded inverse energy transfer implies a similar effect on 
326: the direct transfer of potential enstrophy, a consequence of the dual 
327: conservation law. One can see from the ``symmetry'' of the dual 
328: conservation law that the available potential enstrophy for direct 
329: transfer depends on the inversely-transferred energy. Similarly, the 
330: available energy for inverse transfer depends on the directly-transferred 
331: potential enstrophy. Hence impeded inverse energy transfer means impeded 
332: direct potential enstrophy transfer (and vice versa). This implies that 
333: smooth potential vorticity with scales $\gg\lambda^{-1}$ remains smooth.  
334: 
335: \section{Numerical results}
336: 
337: We now consider results obtained from numerical simulations of a 
338: forced-dissipative version of (\ref{CHM}). These convincingly illustrate 
339: the strong reduction of the inverse transfer in the potential energy 
340: regime, and may help explain an observation by Iwayama \etal (2002) in 
341: their numerical simulations of freely decaying CHM turbulence wherein an 
342: initial energy reservoir evolves into a single sharp peak. A plausible 
343: explanation for this observation is that the sharp peak is due to both 
344: the viscous dissipation (which suppresses the formation of small scales) 
345: and the impeded inverse transfer deduced in the preceding section: their 
346: collective effect ``squeezes'' energy into a narrow band of wavenumbers, 
347: resulting in the observed peak. In our forced-dissipative case we observe 
348: that in the potential energy regime the inverse energy flux is arrested 
349: and deposits most of the potential energy at the lower-wavenumber end of 
350: the quasi-steady energy range. We also observe a relaxation of the kinetic 
351: energy spectrum of the inertial range (from $k^{-5/3}$ toward $k^{-1}$) 
352: as the inverse cascade proceeds from $k>\lambda$ to $k<\lambda$. 
353:  
354: We simulate Eq. (\ref{CHM}), with the addition of molecular viscosity 
355: and a body force, in a doubly periodic square $2\pi\times2\pi$ with 
356: a dealiased $1024^2$ pseudospectral method. No large-scale dissipation
357: mechanisms are employed and no steady dynamics are sought since we are
358: primarily interested in the near arrest of the inverse energy cascade in
359: a quasi-steady state. (For an analysis of steady CHM turbulence in the
360: presence of large-scale damping, see Smith \etal 2002.) Three different 
361: values of $\lambda$ are used: $\lambda=20,40$ and $60$. The forcing is 
362: spectrally localised in the wavenumber interval $K=[49.5,50.5]$, in the 
363: sense that its Fourier components $\widehat f(\k)$ are nonzero only for 
364: those modes $\k$ having magnitudes $k$ lying in the interval $K$: 
365: \begin{eqnarray}
366: \label{forcing}
367: \widehat{f}(\bm k)&=&\frac{\epsilon}{N}\frac{\widehat{\psi}(\bm{k})}
368: {\displaystyle\sum_{|\bm{p}|=k}|\widehat{\psi}(\bm{p})|^2}.
369: \end{eqnarray}
370: Here $\epsilon=1$ is the constant energy injection rate and~$N$ is the 
371: number of distinct wavenumbers in~$K$. The viscosity coefficient is 
372: $\nu=5\times10^{-4}$. The simulations were initialised with the kinetic 
373: energy spectrum $E(k)=10^{-5}\pi k/(50^2+k^2)$.
374: 
375: \begin{figure}
376: \centerline{\includegraphics{atmosp234}}
377: \caption{The kinetic energy spectrum $E(k)$ {\it vs.}\ $k$ averaged over 
378: periods $[139,140]$, $[155,156]$ and $[160,161]$ for $\lambda=20$, $40$ 
379: and $60$, respectively.} 
380: \label{atmosp234}
381: \end{figure} 
382: 
383: Figure \ref{atmosp234} shows the quasi-steady kinetic energy spectrum 
384: $E(k)$ vs. $k$ 
385: averaged over periods $[139,140]$, $[155,156]$ and $[160,161]$ for 
386: $\lambda=20$, $40$ and $60$, respectively. In all three cases, the 
387: accumulation of energy at low wavenumbers is clearly visible. This is 
388: attributed to the impeding effect discussed above, and not finite-size 
389: effects since the energy levels at the lowest wavenumbers are still 
390: small compared with the energy peaks. This is true for both kinetic 
391: and potential components. For $\lambda=20$ we have a limited kinetic 
392: energy regime in the wavenumber interval $[20,50]$. Before the energy 
393: peak reaches $\lambda=20$, the dynamics of this kinetic energy regime 
394: are expected to resemble those of 2D NS turbulence: the energy cascades 
395: to low wavenumbers via a $k^{-5/3}$ kinetic energy spectrum. We indeed 
396: observe a discernible $k^{-5/3}$ spectrum (not shown) in the wavenumber 
397: range $[20,50]$ before the energy peak reaches $\lambda=20$. The dynamics 
398: after the energy peak passes $\lambda$ undergo profound changes. Most 
399: notably, the growth of potential energy occurs mainly around the energy 
400: peak, consistent with the reduction of the inverse energy transfer in 
401: the conservative case. Another remarkable change is in the spectral 
402: slope of the energy spectrum from $k=\lambda=20$ up to the forcing 
403: wavenumber $k=50$. The slope of the kinetic energy spectrum in this 
404: region appears to relax from $-5/3$ to $-1$. The quasi-steady\footnote
405: {Note that after the energy peak passes $\lambda$, growth of energy 
406: is due primarily to that of the potential energy and the kinetic 
407: energy becomes quasi-steady.} spectrum becomes $k^{-1}$ with an 
408: excessive accumulation of energy at the lower-wavenumber end. This 
409: unexpected feature is common to the other two cases, for which hardly 
410: any inverse transfer region exists between the forcing wavenumber and 
411: $\lambda$ and for which no clear $k^{-5/3}$ spectrum is formed initially.
412: 
413: \begin{figure}
414:     \begin{center}
415:     \includegraphics[scale=0.625]{ell-ori.eps} \,\,
416:     \includegraphics[scale=0.625]{ell-sca.eps} 
417:     \end{center}
418: \caption{The energy centroid $\ell$ versus time $t$ (left) and scaled
419: time $t(20/\lambda)^2$ (right) for $\lambda=20$, $40$ 
420: and $60$, as labelled.} 
421: \label{ell_vs_time}
422: \end{figure} 
423: 
424: The evolution of the energy centroid $\ell$ for these three cases is
425: shown in Figure \ref{ell_vs_time}, on the left versus unscaled time $t$
426: and on the right versus scaled time $t(20/\lambda)^2$.  Both the 
427: reduction in ${\d\ell}/{\dt}$ with increasing $\lambda$ {\it and} 
428: the collapse of the curves when plotted versus scaled time are 
429: consistent with the growth rate bound (\ref{evol2}).\footnote
430: {We cannot directly compare the numerical growth rates with 
431: the theoretical bound because the imposed forcing violates 
432: conservation of $\norm{q}_\infty$ and $\ell_s$ used 
433: in (\ref{evol2}).}
434: Only the $\lambda = 20$ case stands out slightly, due to 
435: an early kinetic energy transfer which is virtually absent in 
436: the other two cases.  
437: 
438: \section{Summary and discussion}
439: 
440: The main result of this work is an upper bound for the growth rate of
441: the characteristic length scale associated with the inverse energy 
442: transfer in CHM turbulence. This upper bound is expressible in terms of 
443: the supremum of the potential vorticity, the characteristic length scale
444: representing the potential enstrophy centroid of the reservoir and the 
445: Rossby deformation radius. These are constants of motion, allowing the 
446: bound to be completely described in terms of initial data. It is found 
447: that in remote regions of the potential energy regime, the turbulent 
448: transfer from an initial energy reservoir becomes strongly impeded, in 
449: the sense that hardly any potential energy is transferred to lower 
450: wavenumbers. This effect has been illustrated by numerical simulations 
451: of a forced-dissipative version of the CHM equation. The physical 
452: implication of the present finding is that the elasticity of the free
453: surface hinders the nonlinear transfer of potential energy to larger scales. 
454: 
455: An interesting implication of the present result is that vortices having
456: smooth potential vorticity and scales much larger than the Rossby 
457: deformation radius are kept from merging to form larger energy-carrying 
458: scales and remain smooth. Such vortices are inactive and long-lived 
459: compared to their counterparts in 2D NS turbulence.
460: 
461: The present theory is fully consistent with other results for both forced
462: (Kukharkin \etal 1995) and freely decaying (Larichev \& McWilliams 1991; 
463: Iwayama \etal 2002; Arbic \& Flierl 2003) CHM turbulence, as discussed above.
464: These, we argue, are explained by the growth rate bound (\ref{evol3}) and 
465: its physical interpretations given at the end of \S\,2. 
466: 
467: Analyses of atmospheric and oceanic data yield results that could plausibly 
468: be explained by the present work. \cite{Boer83} find that the spectral flux 
469: in atmospheric wind data is arrested with essentially no inertial range. 
470: \cite{Scott05} find the inverse energy transfer in the oceans to be 
471: qualitatively similar to the result of \cite{Boer83}. These are in 
472: remarkable agreement with the idealised numerical results noted above and 
473: with the present findings.
474: 
475: Iwayama \etal (2002) observe ``well-behaved'' tails of the potential 
476: vorticity probability density function (PDF)---much steeper than observed
477: in NS turbulence, see their figure 5---and raise the question why it 
478: is that CHM turbulence in the potential energy regime exhibits no 
479: intermittency. A speculative answer provided by the authors is based 
480: on an asymptotic version of the CHM equation wherein the relative 
481: vorticity is dropped from the time-derivative term: 
482: $$\lambda^2\,\frac{\partial\psi}{\partial t}+J(\Delta\psi,\psi) = 0.$$ 
483: This equation, as noted by Iwayama \etal (2002), describes advection of
484: $\psi$ by $\Delta\psi$ (i.e. by the smaller scales of motion). These 
485: authors argue that this may lead to more ``random walk'' behaviour, 
486: effectively suppressing intermittency. The present study provides an 
487: alternative answer: the reduction of turbulent transfer both renders 
488: large-scale vortices inactive and discourages erratic small-scale 
489: motion. This also can account for the observed steep PDF. In some sense 
490: CHM turbulence in the potential energy regime is far less turbulent 
491: than its 2D NS counterpart. 
492: 
493: \begin{acknowledgments}
494: The numerical simulations in \S\,3 were carried out, using a computer 
495: code developed by John Bowman, when CVT was a Pacific Institute for the 
496: Mathematical Sciences postdoctoral fellow at the University of Alberta. 
497: Constructive comments from two anonymous referees were very much appreciated.
498: \end{acknowledgments} 
499: 
500: \begin{thebibliography}{}
501: 
502: \bibitem[Arbic \& Flierl (2003)]{Arbic03} 
503:      \textsc{Arbic, B. K. \& Flierl, G. R.} 2003
504:      {Coherent vortices and kinetic energy ribbons in asymptotic, quasi 
505:       two-dimensional f-plane turbulence.} 
506:      \textit{Phys. Fluids} {\bf 15}, 2177--2189.
507: 
508: \bibitem[Boer \& Shepherd (1983)]{Boer83} 
509:      \textsc{Boer, G. J. \& Shepherd, T. G.} 1983
510:      {Large-scale two-dimensional turbulence in the atmosphere.} 
511:      \textit{J. Atmos. Sci.} {\bf 40}, 164--184.
512: 
513: \bibitem[Dritschel, de la Torre Ju\'arez \& Ambaum (1999)]{Dritschel99}
514:      \textsc{Dritschel, D. G., de la Torre Ju\'arez, M. \& Ambaum, M. H. P.} 
515:       1999
516:      {The three dimensional vortical nature of atmospheric and oceanic
517:       turbulent flows.} 
518:      \textit{Phys. Fluids} \textbf{11}, 1512--1520.
519: 
520: \bibitem[Dritschel \& de la Torre Ju\'arez (1996)]{Dritschel96}
521:      \textsc{Dritschel, D. G. \& de la Torre Ju\'arez, M.} 1996
522:      {The instability and breakdown of tall columnar vortices in a
523:       quasi-geostrophic fluid.} 
524:      \textit{J. Fluid Mech.} \textbf{328}, 129--160.
525: 
526: \bibitem[Fyfe \& Montgomery (1979)]{Fyfe79}
527:      \textsc{Fyfe, D. \& Montgomery, D.} 1979
528:      {Possible inverse cascade behaviour for drift-wave turbulence.} 
529:      \textit{Phys. Fluids} \textbf{22}, 246--248.
530: 
531: \bibitem[Hasegawa (1978)]{Hasegawa78} 
532:      \textsc{Hasegawa, A. \& Mima, K.} 1978
533:      {Pseudo-three-dimensional turbulence in magnetized nonuniform plasma.} 
534:      \textit{Phys. Fluids} {\bf 21}, 87--92.
535: 
536: \bibitem[Iwayama, Shepherd \& Watanabe (2002)]{Iwayama02}
537:      \textsc{Iwayama, T., Shepherd, T. G. \& Watanabe, T.} 2002
538:      {An `ideal' form of decaying two-dimensional turbulence.}
539:      \textit{J. Fluid Mech.} \textbf{456}, 183--198.
540: 
541: \bibitem[Kukharkin \& Orszag (1996)]{Kukharkin96}
542:      \textsc{Kukharkin, N. \& Orszag, S. A.} 1996
543:      {Generation and structure of Rossby vortices in rotating fluids.}
544:      \textit{Phys. Rev.} E \textbf{54}, 4524--4527.
545: 
546: \bibitem[Kukharkin, Orszag \& Yakhot (1995)]{Kukharkin95}
547:      \textsc{Kukharkin, N., Orszag, S. A. \& Yakhot, V.} 1995
548:      {Quasicrystallization of vortices in drift-wave turbulence.}
549:      \textit{Phys. Rev. Lett.} \textbf{75}, 2486--2489.
550: 
551: \bibitem[Larichev \& McWilliams (1991)]{Larichev91}
552:      \textsc{Larichev, V. D. \& McWilliams, J. C.} 1991
553:      {Weakly decaying turbulence in an equivalent-barotropic fluid.}
554:      \textit{Phys. Fluids} A \textbf{3}, 938--950. 
555: 
556: \bibitem[Okuno \& Masuda (2003)]{Okuno03}
557:      \textsc{Okuno, A. \& Masuda, A.} 2003
558:      {Effects of horizontal divergence on the geostrophic turbulence on a
559:       beta-plane: suppression of the Rhines effect.}
560:      \textit{Phys. Fluids} \textbf{15}, 56--65. 
561: 
562: \bibitem[Ottaviani \& Krommes (1992)]{Ottaviani92}
563:      \textsc{Ottaviani, M. \& Krommes, J. A.} 1992
564:      {Weak- and strong-turbulence regimes of the forced Hasegawa--Mima
565:       equation.}
566:      \textit{Phys. Rev. Lett.} \textbf{69}, 2923--2026.
567: 
568: \bibitem[Pedlosky (1987)]{Pedlosky87}
569:      \textsc{Pedlosky, J.} 1987
570:      \textit{Geophysical Fluid Dynamics,} 2nd Edn. Springer. 
571: 
572: \bibitem[Scott \& Wang (2005)]{Scott05} 
573:      \textsc{Scott, R. B. \& Wang, F.} 2005
574:      {Direct evidence of an oceanic inverse kinetic energy cascade from 
575:       satellite altimetry.}
576:      \textit {J. Phys. Ocean.} \textbf{35}, 1650--1666.
577: 
578: \bibitem[Smith (2004)]{Smith04}
579:      \textsc{Smith, K. S.} 2004
580:      {A local model for planetary atmospheres forced by small-scale 
581:       convection.}
582:      \textit{J. Atmos. Sci.} \textbf{61}, 1420--1433.
583: 
584: \bibitem[Smith \etal (2002)]{Smith02}
585:      \textsc{Smith, K. S., Boccaletti, G., Henning, C. C., Marinov, I. N.,
586:       Tam, C. Y., Held, I. M. \& Vallis, G. K.} 2002
587:      {Turbulence diffusion in the geostrophic inverse cascade.}
588:      \textit{J. Fluid Mech.} \textbf{469}, 13--48.
589: 
590: \bibitem[Tran (2004)]{T04} 
591:      \textsc{Tran, C. V.} 2004
592:      {Nonlinear transfer and spectral distribution of energy in 
593:      $\alpha$ turbulence.}
594:      \textit {Physica} D \textbf{191}, 137--155.
595: 
596: \bibitem[Tran \& Bowman (2003)]{TB03} 
597:      \textsc{Tran, C. V. \& Bowman, J. C.} 2003
598:      {Energy budgets in Charney--Hasegawa--Mima and surface 
599:      quasigeostrophic turbulence.}
600:      \textit{Phys. Rev.} E \textbf{68}, 036304.
601: 
602: \bibitem[Tran \& Shepherd (2002)]{TS02}
603:      \textsc{Tran, C. V. \& Shepherd, T. G.} 2002 
604:      {Constraints on the spectral distribution of energy and enstrophy 
605:      dissipation in forced two-dimensional turbulence.} 
606:      \textit{Physica} D \textbf{165}, 199--212. 
607: 
608: \bibitem[Yanase \& Yamada (1984)]{Yanase84}
609:      \textsc{Yanase, S. \& Yamada, M.} 1984
610:      {The effect of the finite Rossby radius on two-dimensional isotropic
611:      turbulence.}
612:      \textit{J. Phys. Soc. Japan} \textbf{53}, 2513--2520.
613: 
614: \end{thebibliography}
615: 
616: \end{document}
617: