nlin0410050/Ic.tex
1: \documentclass
2: [aps,pre,twocolumn,groupedaddress,showpacs,amsmath,floatfix]{revtex4}
3: \usepackage{bm,graphicx}
4: 
5: \begin{document}
6: 
7: \preprint{PUPT-2140}
8: 
9: \title{Field theory of the inverse cascade in two-dimensional turbulence}
10: 
11: \author{Jackson R. Mayo}
12: \email{jmayo@princeton.edu}
13: \affiliation{Department of Physics, Princeton University, Princeton, New
14: Jersey 08544-0708}
15: 
16: \begin{abstract}
17: A two-dimensional fluid, stirred at high wavenumbers and damped by both
18: viscosity and linear friction, is modeled by a statistical field theory. The
19: fluid's long-distance behavior is studied using renormalization-group (RG)
20: methods, as begun by Forster, Nelson, and Stephen [Phys.\ Rev.\ A
21: \textbf{16}, 732 (1977)]. With friction, which dissipates energy at low
22: wavenumbers, one expects a stationary inverse energy cascade for strong
23: enough stirring. While such developed turbulence is beyond the quantitative
24: reach of perturbation theory, a combination of exact and perturbative results
25: suggests a coherent picture of the inverse cascade. The zero-friction
26: fluctuation-dissipation theorem (FDT) is derived from a generalized
27: time-reversal symmetry and implies zero anomalous dimension for the velocity
28: even when friction is present. Thus the Kolmogorov scaling of the inverse
29: cascade cannot be explained by any RG fixed point. The $\beta$ function for
30: the dimensionless coupling $\hat g$ is computed through two loops; the $\hat
31: g^3$ term is positive, as already known, but the $\hat g^5$ term is negative.
32: An ideal cascade requires a linear $\beta$ function for large $\hat g$,
33: consistent with a Pad\'e approximant to the Borel transform. The conjecture
34: that the Kolmogorov spectrum arises from an RG flow through large $\hat g$ is
35: compatible with other results, but the accurate $k^{-5/3}$ scaling is not
36: explained and the Kolmogorov constant is not estimated. The lack of scale
37: invariance should produce intermittency in high-order structure functions, as
38: observed in some but not all numerical simulations of the inverse cascade.
39: When analogous RG methods are applied to the one-dimensional Burgers equation
40: using an FDT-preserving dimensional continuation, equipartition is obtained
41: instead of a cascade---in agreement with simulations.
42: \end{abstract}
43: 
44: \pacs{47.27.Gs}
45: 
46: \maketitle
47: 
48: \section{\label{Intro}Introduction}
49: 
50: The cascade of energy to low wavenumbers in two-dimensional turbulence
51: \cite{K67}, more than other turbulence problems, is suited to the standard
52: methods of statistical field theory and the renormalization group (RG). These
53: methods \cite{Z96}, originating in quantum field theory, show that arbitrary
54: short-distance interactions lead to long-distance behavior described by a
55: local, renormalizable action---an effective field theory. Correlation
56: functions computed from this action contain ultraviolet (UV) divergences that
57: can be eliminated by redefining the parameters and fields. The divergences
58: leave their mark, however, in the dependence on the renormalization scale and
59: the resulting anomalous scaling laws.
60: 
61: The classic application of statistical field theory is to critical phenomena
62: (second-order phase transitions) in condensed matter \cite{Z96}. The infrared
63: (IR) scale invariance of correlation functions at the transition temperature
64: is explained by a fixed point of the RG flow. The inverse energy cascade of
65: two-dimensional turbulence is likewise believed to be nearly scale invariant,
66: and one might suspect that a similar fixed point is responsible. We will see,
67: however, that no fixed point can reproduce the observed $k^{-5/3}$ energy
68: spectrum. Rather, we will argue that the inverse cascade arises from a
69: nontrivial RG flow and thus is not expected to be completely scale invariant.
70: 
71: In the study of turbulence, a deviation from scale invariance (a dependence
72: of dimensionless physical quantities on scale) is referred to as
73: intermittency \cite{F95}. While intermittency is recognized as a property of
74: the three-dimensional direct cascade of energy, its existence in the
75: two-dimensional inverse cascade is unsettled. For a nonstationary inverse
76: cascade, in which energy is not dissipated but progresses to ever-lower
77: wavenumbers, intermittency is not observed in numerical simulations
78: \cite{SY93,SY94}; a theoretical explanation has been given \cite{Y99}. In
79: this paper we deal solely with the stationary regime, where an inverse
80: cascade requires a low-wavenumber energy sink. With few exceptions,
81: simulations of such a cascade confirm the $k^{-5/3}$ energy spectrum
82: initially predicted \cite{K67} on the basis of scale invariance. But one set
83: of simulations \cite{BDF95} finds strong intermittency in fourth- and
84: higher-order velocity correlations. Other simulations \cite{BCV00} and
85: experiments \cite{PT98}, though, find no significant intermittency. The
86: various studies differ mainly in the precise form of the dissipation terms.
87: The evidence suggests that intermittency in the stationary inverse cascade,
88: permitted by our theory, is at least possibly realized.
89: 
90: It is our restricted focus on the inverse cascade that allows us to work with
91: a purely local field theory. The random force that stirs the fluid is
92: correlated over a limited range and is effectively local in a long-distance
93: description. As with the RG treatment of quantum fields and condensed matter,
94: we expect all short-distance details to become irrelevant except as they are
95: manifested in local, renormalizable couplings. The two-dimensional direct
96: cascade of enstrophy to high wavenumbers \cite{K67} thus falls outside our
97: scope. A previous RG analysis of two-dimensional turbulence \cite{H98} is
98: formally similar to ours, but it follows three-dimensional studies by
99: adopting the long-range force correlations necessary for a direct cascade;
100: even its derivation of the inverse cascade relies on nonlocal forcing. Here
101: we apply RG methods in the familiar domain of local field theory, which
102: should allow a physical treatment of the inverse cascade. The explanation of
103: the direct enstrophy cascade may rest on entirely different foundations, such
104: as conformal invariance \cite{P93}.
105: 
106: The work most closely aligned with our theoretical approach is due to
107: Forster, Nelson, and Stephen (FNS) \cite{FNS77}. At the technical level, our
108: contribution is to add linear friction to FNS model A in $d = 2$ and compute
109: the RG flow to the next order of perturbation theory, two loops. The
110: inclusion of friction, which dissipates energy at low wavenumbers, makes our
111: theory capable in principle of describing the inverse cascade and its
112: $k^{-5/3}$ spectrum---unlike FNS model A, which gives a $k^1$ spectrum
113: corresponding to energy equipartition in $d = 2$. We also note a difference
114: in our viewpoint from that of FNS and others \cite{DM79,H98,AAV99} who apply
115: RG methods to turbulence by analogy with critical phenomena. These authors
116: seek a controlled IR-stable fixed point by starting with a logarithmically
117: divergent field theory and then decreasing the dimension of space or the
118: exponent of the stirring-force correlation by $\epsilon$. This adds to the
119: $\beta$ function a negative linear term proportional to $\epsilon$. With the
120: usual positive one-loop term, there exists an IR-stable fixed point at a
121: coupling that goes to zero with $\epsilon$; the fixed-point theory can then
122: be expanded in $\epsilon$ instead of the original coupling. Like FNS, we work
123: in $d = 2 - \epsilon$ to regulate UV divergences, but we ultimately take
124: $\epsilon = 0$, so that the fixed point is trivial. Our inverse-cascade model
125: lies not at a fixed point but in the region of large dimensionless coupling.
126: 
127: Naturally the use of perturbation theory is questionable for strong coupling.
128: Our perturbative results will have direct quantitative application only to
129: the extreme IR limit controlled by the trivial fixed point, which is of some
130: interest in itself. Nevertheless, we will make reasonable conjectures about
131: the theory's strong-coupling behavior that are consistent with the inverse
132: cascade, bearing in mind the dangers of the nonperturbative regime. Besides
133: the concern with the numerical accuracy of extrapolations, there are
134: fundamental difficulties at strong coupling. The anomalous dimensions of
135: operators may be large, and the relevance of terms in the action may differ
136: from the weak-coupling case. Furthermore, at strong coupling, there is no
137: simple relation between the couplings in different renormalization schemes,
138: such as the Wilsonian cutoff (useful for physical interpretation) and minimal
139: subtraction (convenient for systematic calculation). We may hope that these
140: subtleties do not affect the main conclusions even at very strong coupling.
141: At least we know that the theory of critical phenomena in $d = 4 - \epsilon$
142: is extrapolated to $\epsilon = 1$ (moderate coupling) with acceptable
143: results.
144: 
145: In Sec.~\ref{Framework} we describe the basis of our theory and our method of
146: calculation, confirming the one-loop RG flow of FNS \cite{FNS77}. In
147: Sec.~\ref{Properties} we present symmetries and other properties of the
148: theory that do not involve a dubious extrapolation to strong coupling. In
149: Sec.~\ref{TwoLoop} we compute the two-loop term of the $\beta$ function. In
150: Sec.~\ref{Cascade} we relate the plausible strong-coupling behavior of the
151: theory to the phenomenology of the inverse cascade. In Sec.~\ref{Burgers}, as
152: a test of our methods, we consider a rather different model, the UV-stirred
153: one-dimensional Burgers equation. A summary and discussion are presented in
154: Sec.~\ref{Discussion}.
155: 
156: \section{\label{Framework}Framework}
157: 
158: \subsection{\label{Path}Path integral for the Navier--Stokes equation}
159: 
160: The Navier--Stokes equation for the velocity field $v_j$ of an incompressible
161: two-dimensional fluid is
162: \begin{equation}
163: \label{NSv}
164: \dot v_j + v_k \nabla_k v_j + \nabla_j P - \nu\nabla^2 v_j + \alpha v_j =
165: f_j,
166: \end{equation}
167: where $P$ is the pressure divided by the density, $\nu$ is the kinematic
168: viscosity, $\alpha$ is the friction coefficient, and $f_j$ is the force per
169: unit mass. The incompressibility condition $\nabla_i v_i = 0$ allows the
170: velocity to be expressed as
171: \begin{equation}
172: v_i = \epsilon_{ij} \nabla_j \psi,
173: \end{equation}
174: where $\psi$ is a pseudoscalar field called the stream function and
175: $\epsilon_{ij}$ is the alternating tensor, which in two dimensions satisfies
176: \begin{equation}
177: \epsilon_{ij} \epsilon_{kl} = \delta_{ik} \delta_{jl} - \delta_{il}
178: \delta_{jk}.
179: \end{equation}
180: Upon writing Eq.~(\ref{NSv}) in terms of $\psi$ and applying the operator
181: $\epsilon_{ij} \nabla_i$, we obtain \cite{H98}
182: \begin{equation}
183: \label{NSpsi}
184: -\nabla^2 \dot\psi - \epsilon_{ij} \nabla_i \nabla^2 \psi \nabla_j \psi +
185: \nu\nabla^4 \psi - \alpha\nabla^2 \psi = \eta.
186: \end{equation}
187: Here
188: \begin{equation}
189: \eta = \bm{\nabla} \times \mathbf{f},
190: \end{equation}
191: with the notation
192: \begin{equation}
193: \mathbf{a} \times \mathbf{b} \equiv \epsilon_{ij} a_i b_j.
194: \end{equation}
195: 
196: In the real three-dimensional world, Eq.~(\ref{NSv}) is a good approximation
197: for a thin fluid film provided either (a) the boundary surface(s) and the
198: coordinate system are rotating rapidly about a perpendicular axis or (b) the
199: fluid is conducting and subject to a strong perpendicular magnetic field
200: \cite{L97}. In either case, boundary-layer effects produce a linear friction
201: parametrized by $\alpha$, which has units of frequency.
202: 
203: For convenience, our field-theory calculations will use the method of
204: dimensional regularization \cite{TV72}, based on continuation to a noninteger
205: spatial dimension $d = 2 - \epsilon$. Because in the end we are concerned
206: only with $d = 2$, we adopt a formal continuation of the Navier--Stokes
207: equation that preserves its two-dimensional features. For general $d$, we
208: retain the stream-function representation by choosing a ``physical''
209: two-dimensional subspace that contains the tensor $\epsilon_{ij}$ and the
210: external wavevectors of correlation functions. Denoting by $\Theta_{ij}$ the
211: projector onto the physical subspace, we now have
212: \begin{equation}
213: \label{epseps}
214: \epsilon_{ij} \epsilon_{kl} = \Theta_{ik} \Theta_{jl} - \Theta_{il}
215: \Theta_{jk}.
216: \end{equation}
217: We take Eq.~(\ref{NSpsi}) as the equation of motion for $\psi$, with
218: $\nabla^2 \equiv \nabla_k \nabla_k$ interpreted as the $d$-dimensional
219: Laplacian. In this way we preserve (for $\nu=\alpha=\eta=0$) the formal
220: conservation of the energy and enstrophy,
221: \begin{align}
222: E &= \tfrac{1}{2} \int d^d x\, \nabla_i \psi \nabla_i \psi,\\
223: \mathnormal{\Omega} &= \tfrac{1}{2} \int d^d x\, \nabla^2 \psi \nabla^2 \psi.
224: \end{align}
225: 
226: We assume that the fluid is stirred by a Gaussian random force that is
227: uncorrelated in time \cite{FNS77}, with
228: \begin{align}
229: \langle f_i(\omega,\mathbf{k})\, f_j(\omega',\mathbf{k}')\rangle =
230: {}&(2\pi)^{d+1} \delta(\omega+\omega')\,
231: \delta(\mathbf{k}+\mathbf{k}')\nonumber\\
232: \label{forcing}
233: &\times (\delta_{ij} - k_i k_j/k^2)\, D(k^2),
234: \end{align}
235: so that $\eta$ is also Gaussian with
236: \begin{equation}
237: \langle\eta(\omega,\mathbf{k})\, \eta(\omega',\mathbf{k}')\rangle =
238: (2\pi)^{d+1} \delta(\omega+\omega')\, \delta(\mathbf{k}+\mathbf{k}')\, k^2
239: D(k^2).
240: \end{equation}
241: A classical system with random forcing can be treated in the formalism of
242: quantum field theory \cite{MSR73}, including the path-integral representation
243: \cite{BJW76,P77,AVP83}. Upon introduction of a pseudoscalar field $p$
244: conjugate to $\eta$, correlation functions of $\psi$ are given by the path
245: integral
246: \begin{equation}
247: \langle F[\psi]\rangle \propto \int \mathcal{D}\psi\, \mathcal{D}p\,
248: F[\psi]\, e^{-S},
249: \end{equation}
250: with the action
251: \begin{align}
252: S = {}&{\int dt}\, d^d x\, \bigl[\tfrac{1}{2} (-\nabla^2 p)\,
253: D(-\nabla^2)p\nonumber\\
254: &+ ip(-\nabla^2 \dot\psi - \epsilon_{ij} \nabla_i \nabla^2 \psi \nabla_j \psi
255: + \nu\nabla^4 \psi - \alpha\nabla^2 \psi)\bigr]\nonumber\\
256: = {}&{\int dt}\, d^d x\, \bigl[\tfrac{1}{2} (-\nabla^2 p)\, D(-\nabla^2)p +
257: i\nu p\nabla^4 \psi\nonumber\\
258: \label{Sppsi}
259: &- i\alpha p\nabla^2 \psi - ip\nabla^2 \dot\psi - i\epsilon_{ij} \nabla_i
260: \nabla_k p \nabla_j \psi \nabla_k \psi\bigr],
261: \end{align}
262: where we have integrated the $p\psi\psi$ term by parts. The Jacobian
263: determinant from changing variables from $\eta$ to $\psi$ is an unimportant
264: constant by virtue of causality \cite{AVP83}.
265: 
266: \subsection{\label{Relevance}Relevance of couplings}
267: 
268: In the field theory based on the action (\ref{Sppsi}), the long-distance
269: behavior is governed by just the renormalizable terms---those with
270: coefficients whose scaling dimensions with respect to wavenumber in $d = 2$
271: are zero (marginal) or positive (relevant) \cite{Z96}. We assign scaling
272: dimensions $d_\psi$, $d_p$, $d_t$ to the fields $\psi$ and $p$ and to the
273: time $t$ by requiring that the highest-derivative quadratic terms in the
274: action have dimensionless coefficients in $d = 2$, since these terms control
275: the asymptotic behavior of propagators and thus the UV convergence or
276: divergence of diagrams. Because of the additional derivatives, the viscous
277: term $i\nu p\nabla^4 \psi$ is clearly less relevant than the friction term
278: $-i\alpha p\nabla^2 \psi$. In fact, it is commonly said that viscosity is
279: irrelevant to the inverse cascade, but we now show that this cannot be taken
280: in the technical sense. If the viscous term is ignored, then the $p\nabla^2
281: \psi$ and $p\nabla^2 \dot\psi$ terms have dimensionless coefficients only if
282: $d_t = 0$ and $d_\psi = -d_p$. For a nontrivial theory, the $p\psi\psi$ term
283: must be renormalizable, giving $2 \ge 4 + d_p + 2d_\psi = 4 - d_p$. With $d_p
284: \ge 2$, there exists no local renormalizable forcing term quadratic in $p$.
285: 
286: Let us therefore retain the viscous term and recompute the scaling
287: dimensions. The $p\nabla^4 \psi$ and $p\nabla^2 \dot\psi$ terms possess
288: dimensionless coefficients only if $d_t = -2$ and $d_\psi = -d_p$.
289: Renormalizability of the $p\psi\psi$ term now gives $4 \ge 4 + d_p + 2d_\psi
290: = 4 - d_p$, so that $d_p \ge 0$. It remains to specify the forcing term. We
291: assume that the external forcing is confined to a band of high wavenumbers,
292: as in model C of FNS \cite{FNS77}. The effective forcing at low wavenumbers
293: is generated by renormalization; because the interaction in Eq.~(\ref{Sppsi})
294: has two spatial derivatives acting on $p$, at least two derivatives must
295: accompany each factor of $p$ in any term so generated. The only
296: renormalizable forcing term is then $\frac{1}{2} D_0 \nabla^2 p \nabla^2 p$,
297: whose coefficient is dimensionless for $d_p = d_\psi = 0$. This effective
298: forcing has $D(k^2) = D_0 k^2$, as in model A of FNS. All terms in the action
299: are now marginal, except for the friction term, which is relevant
300: (coefficient of dimension $2$). The only other renormalizable terms that
301: could be generated are ones containing only $\psi$, but these are not
302: generated (see Sec.~\ref{Feynman}).
303: 
304: Next we label the fields and the time in Eq.~(\ref{Sppsi}) with the subscript
305: ``phys'' and introduce rescaled variables to simplify the action. Tentatively
306: seeking to set all coefficients other than forcing and friction equal to $1$,
307: we take
308: \begin{equation}
309: \begin{aligned}
310: \label{altpsipt}
311: \psi_{\text{phys}} &= \nu\psi, & p_{\text{phys}} &= \nu^{-1} p, &
312: t_{\text{phys}} &= \nu^{-1} t.
313: \end{aligned}
314: \end{equation}
315: The result is
316: \begin{align}
317: S = {}&{\int dt}\, d^d x\, \bigl(\tfrac{1}{2} g^2 \nabla^2 p \nabla^2 p +
318: ip\nabla^4 \psi\nonumber\\
319: \label{Sippsi}
320: &- i\nu^{-1} \alpha p\nabla^2 \psi - ip\nabla^2 \dot\psi - i\epsilon_{ij}
321: \nabla_i \nabla_k p \nabla_j \psi \nabla_k \psi\bigr),
322: \end{align}
323: where
324: \begin{equation}
325: g = D_0^{1/2} \nu^{-3/2}.
326: \end{equation}
327: We adopt, however, a different rescaling that will be particularly convenient
328: in Sec.~\ref{FDT}:
329: \begin{equation}
330: \begin{aligned}
331: \label{ppsit}
332: \psi_{\text{phys}} &= g\nu\psi, & p_{\text{phys}} &= (g\nu)^{-1} p, &
333: t_{\text{phys}} &= (g\nu)^{-1} t.
334: \end{aligned}
335: \end{equation}
336: The final form of the action is then
337: \begin{align}
338: S = {}&{\int dt}\, d^d x\, \bigl(\tfrac{1}{2} g^{-1} \nabla^2 p \nabla^2 p +
339: ig^{-1} p\nabla^4 \psi\nonumber\\
340: \label{Sfppsi}
341: &- i\rho p\nabla^2 \psi - ip\nabla^2 \dot\psi - i\epsilon_{ij} \nabla_i
342: \nabla_k p \nabla_j \psi \nabla_k \psi\bigr),
343: \end{align}
344: where
345: \begin{equation}
346: \rho = (g\nu)^{-1} \alpha = D_0^{-1/2} \nu^{1/2} \alpha.
347: \end{equation}
348: For general $d = 2 - \epsilon$, the scaling dimensions in Eq.~(\ref{Sfppsi})
349: are
350: \begin{equation}
351: \label{deps}
352: \begin{aligned}
353: d_\psi = d_p &= -\tfrac{1}{2} \epsilon, & d_t &= -2 + \tfrac{1}{2}
354: \epsilon,\\
355: d_g &= +\tfrac{1}{2} \epsilon, & d_\rho &= +2 - \tfrac{1}{2} \epsilon.
356: \end{aligned}
357: \end{equation}
358: In two dimensions, $g$ is a dimensionless coupling and $\rho$ is analogous to
359: a mass parameter in quantum field theory.
360: 
361: \subsection{\label{Feynman}Feynman rules}
362: 
363: \begin{figure}
364: \includegraphics{icfig1.eps}%
365: \caption{\label{FeynRules}Feynman rules for the action
366: (\protect\ref{Sfppsi}).}
367: \end{figure}
368: 
369: Correlation functions can be calculated for the action (\ref{Sfppsi}) using
370: Feynman diagrams whose lines carry both frequencies and wavevectors. We
371: represent the fields $\psi$ and $p$ by wiggly and plain lines respectively.
372: The ingredients of the diagrams, shown in Fig.~\ref{FeynRules}, are the
373: propagators $\langle p\psi\rangle_0$, $\langle\psi\psi\rangle_0$, obtained
374: from the quadratic terms in the action, and the vertex factor
375: $-\Gamma^{\psi\psi p}_0$, obtained from the cubic term. We label these
376: quantities with the subscript $0$ because they are the tree-level
377: contributions to the exact two-point correlation functions $\langle
378: p\psi\rangle$, $\langle\psi\psi\rangle$ and the exact three-point
379: one-particle-irreducible (1PI) function $-\Gamma^{\psi\psi p}$.
380: 
381: The remaining contributions arise from diagrams containing loops. For these
382: diagrams, we integrate over each loop frequency and wavevector according to
383: \begin{equation}
384: \int \frac{d\omega}{2\pi} \frac{d^d k}{(2\pi)^d}.
385: \end{equation}
386: Because the integrand is a rational function of the frequencies, the $\omega$
387: integrations can easily be performed by the contour method before integrating
388: over wavevectors. This contour integration shows that a 1PI diagram (or
389: subdiagram) vanishes if all its external lines are $\psi$ \cite{AVP83}, since
390: there is a closed loop of $\langle p\psi\rangle_0$ propagators and the
391: integrand is an analytic function of the loop frequency in the upper
392: half-plane.
393: 
394: Unlike many field theories in a low number of spatial dimensions, ours does
395: not contain IR divergences even for $\rho = 0$. This is because, after
396: integration over frequencies, internal-line propagators scale as $k^{-2}$,
397: but at least one further factor of the wavevector arises from each of the two
398: vertices that a line connects. Hence the integrand does not diverge as the
399: wavevector of any internal line goes to zero. The frequencies and wavevectors
400: of the external lines act as an IR cutoff. For simplicity, our calculations
401: will adopt another IR cutoff by assuming that $\rho > 0$; then 1PI diagrams
402: are analytic at zero external frequencies and wavevectors.
403: 
404: \subsection{\label{OneLoop}One-loop renormalization}
405: 
406: \begin{figure}
407: \includegraphics{icfig2.eps}%
408: \caption{\label{1L1PI}One-loop diagrams for the two-point 1PI functions
409: $-\Gamma^{\psi p}$ and $-\Gamma^{pp}$ at zero frequency.}
410: \end{figure}
411: 
412: The coefficients of the quadratic terms in the action (with frequency $\chi$
413: and wavevector $\mathbf{q}$) are
414: \begin{align}
415: \Gamma^{\psi p}_0 &= q^2 (\chi + i\rho + ig^{-1} q^2),\\
416: \Gamma^{pp}_0 &= g^{-1} q^4.
417: \end{align}
418: These are corrected at one loop by the two-point 1PI diagrams in
419: Fig.~\ref{1L1PI}. Because of the external wavevectors in the vertex factors,
420: the diagrams are $O(q^4)$, and so we can set $\chi = 0$. We now demonstrate
421: the computation of the $-\Gamma^{\psi p}_1$ diagram, to show the basic
422: methods to be used for two-loop diagrams in Sec.~\ref{TwoLoop}.
423: 
424: The frequency integral of the propagators is
425: \begin{align}
426: I &\equiv \int_{-\infty}^\infty \frac{d\omega}{2\pi} \bigl[g^{-1} k^{-2}
427: (\omega + i\rho + ig^{-1} k^2)^{-1}\nonumber\\
428: &\times (\omega + i\rho + ig^{-1} |\mathbf{k}-\mathbf{q}|^2)^{-1} (\omega -
429: i\rho - ig^{-1} |\mathbf{k}-\mathbf{q}|^2)^{-1}\bigr]\nonumber\\
430: &= \frac{-ig/4k^2}{(g\rho + k^2 - \mathbf{k}\cdot\mathbf{q} + \frac{1}{2}q^2)
431: (g\rho + k^2 - 2\mathbf{k}\cdot\mathbf{q} + q^2)},
432: \end{align}
433: obtained conveniently by closing the contour in the upper half-plane and
434: picking up one pole. We next multiply by the vertex factors and expand to
435: $O(q^4)$:
436: \begin{align}
437: i(k^2 - 2\mathbf{k}&\cdot\mathbf{q}) (\mathbf{k}\times\mathbf{q})\,
438: i(2\mathbf{k}\cdot\mathbf{q} - q^2) (\mathbf{k}\times\mathbf{q})\,
439: I\nonumber\\
440: = {}&(k^2 - 2\mathbf{k}\cdot\mathbf{q}) (q^2 - 2\mathbf{k}\cdot\mathbf{q})
441: [k^2_\parallel q^2 - (\mathbf{k}\cdot\mathbf{q})^2] I\nonumber\\
442: \to {}&\frac{6ig (\mathbf{k}\cdot\mathbf{q})^2 [k^2_\parallel q^2 -
443: (\mathbf{k}\cdot\mathbf{q})^2]} {4(g\rho + k^2)^3}\nonumber\\
444: \label{q4}
445: &- \frac{ig [k^2 q^2 + 4(\mathbf{k}\cdot\mathbf{q})^2] [k^2_\parallel q^2 -
446: (\mathbf{k}\cdot\mathbf{q})^2]}{4k^2 (g\rho + k^2)^2},
447: \end{align}
448: where $k^2_\parallel \equiv \Theta_{ij} k_i k_j$ is the squared projection of
449: $\mathbf{k}$ onto the physical subspace and we infer from Eq.~(\ref{epseps})
450: that
451: \begin{equation}
452: (\mathbf{k} \times \mathbf{q})^2 = k^2_\parallel q^2 - (\mathbf{k} \cdot
453: \mathbf{q})^2.
454: \end{equation}
455: We omit $O(q^3)$ terms in Eq.~(\ref{q4}) because they will now disappear when
456: we average over directions of $\mathbf{k}$.
457: 
458: With the $d$-dimensional isotropization formulas
459: \begin{equation}
460: \begin{aligned}
461: k_i k_j &\to k^2 \frac{\delta_{ij}}{d}, & k_i k_j k_k k_l &\to k^4
462: \frac{\delta_{ij}\delta_{kl} + \delta_{ik}\delta_{jl} +
463: \delta_{il}\delta_{jk}}{d(d+2)},
464: \end{aligned}
465: \end{equation}
466: which imply
467: \begin{equation}
468: \begin{aligned}
469: \label{isotrop}
470: k^2_\parallel &\to \frac{2k^2}{d}, & (\mathbf{k}\cdot\mathbf{q})^2 &\to
471: \frac{k^2 q^2}{d},\\
472: (\mathbf{k}\cdot\mathbf{q})^2 k^2_\parallel &\to \frac{4k^4 q^2}{d(d + 2)}, &
473: (\mathbf{k}\cdot\mathbf{q})^4 &\to \frac{3k^4 q^4}{d(d + 2)},
474: \end{aligned}
475: \end{equation}
476: the value of the diagram to $O(q^4)$ becomes
477: \begin{align}
478: -\Gamma^{\psi p}_1 = {}&{-}igq^4\nonumber\\
479: &\times \int_0^\infty \frac{dk}{(2\pi)^d}
480: \frac{2\pi^{d/2}k^{d-1}}{\Gamma(\frac{1}{2}d)} k^2 \frac{(6 + d) g\rho +
481: dk^2} {4d(d + 2) (g\rho + k^2)^3}\nonumber\\
482: = {}&{-}igq^4\frac{6 - d}{(2 + d) 2^{4+d}\, \Gamma(\frac{1}{2}d)
483: \sin(\frac{1}{2}\pi d)} \left(\frac{\pi}{g\rho}\right)^{1-d/2}.
484: \end{align}
485: For $d = 2 - \epsilon$ with $\epsilon \to 0$, we find
486: \begin{equation}
487: \label{1Lpsip}
488: +\Gamma^{\psi p}_1 = igq^4 \left(\frac{1}{32\pi\epsilon} -
489: \frac{\ln(g\rho/4\pi) + \gamma_{\text{E}} - 1}{64\pi} + O(\epsilon)\right),
490: \end{equation}
491: where $\gamma_{\text{E}}$ is the Euler constant. The result for the other
492: one-loop diagram (taking into account the symmetry factor) is very similar:
493: \begin{equation}
494: \label{1Lpp}
495: +\Gamma^{pp}_1 = gq^4 \left(\frac{1}{32\pi\epsilon} - \frac{\ln(g\rho/4\pi) +
496: \gamma_{\text{E}}}{64\pi} + O(\epsilon)\right).
497: \end{equation}
498: We have carefully obtained the $O(\epsilon^0)$ terms, which will be important
499: in Sec.~\ref{TwoLoop}.
500: 
501: The method of minimal subtraction \cite{T73} expresses the bare couplings $g$
502: and $\rho$, of respective dimensions $\frac{1}{2}\epsilon$ and $2 -
503: \frac{1}{2}\epsilon$, in terms of a renormalization scale $\mu$ (dimension
504: $1$) and dimensionless renormalized couplings $\bar g$ and $\bar\rho$ as
505: \begin{align}
506: g &= \mu^{\epsilon/2} \bigl[\bar g + \bar g^3 a_1 \epsilon^{-1} + O(\bar
507: g^5)\bigr],\\
508: \rho &= \mu^{2-\epsilon/2} \bar\rho,
509: \end{align}
510: where the corrections (called counterterms) involve only negative powers of
511: $\epsilon$. The coefficient of the $O(\bar g^3)$ counterterm, and the absence
512: of any counterterms for $\bar\rho$, are obtained by requiring that
513: $\Gamma^{\psi p}$ and $\Gamma^{pp}$ to $O(q^4)$ be finite at $\epsilon = 0$
514: when expressed in terms of $\bar g$ and $\bar\rho$:
515: \begin{alignat}{2}
516: \Gamma^{\psi p} = {}&q^2 (\chi + i\rho)\span\span\nonumber\\
517: &+ iq^4 \left[g^{-1} + g \left(\frac{1}{32\pi\epsilon} + O(\epsilon^0)\right)
518: + O(g^3)\right]\span\span\nonumber\\
519: = {}&q^2 (\chi + i\mu^{2-\epsilon/2} \bar\rho)\span\span\nonumber\\
520: &+ {}& &iq^4 \mu^{-\epsilon/2}\nonumber\\
521: &&&\times \left[\bar g^{-1} + \bar g \left(\frac{1}{32\pi\epsilon} -
522: \frac{a_1}{\epsilon} + O(\epsilon^0)\right) + O(\bar g^3)\right],
523: \displaybreak[0]\\
524: \Gamma^{pp} = {}&q^4 \mu^{-\epsilon/2}\span\span\nonumber\\
525: &\times \left[\bar g^{-1} + \bar g \left(\frac{1}{32\pi\epsilon} -
526: \frac{a_1}{\epsilon} + O(\epsilon^0)\right) + O(\bar g^3)\right].\span\span
527: \end{alignat}
528: Hence
529: \begin{equation}
530: a_1 = \frac{1}{32\pi},
531: \end{equation}
532: and no counterterms of any order are needed for $\bar\rho$ because the $q^2$
533: term of $\Gamma^{\psi p}$ is not renormalized.
534: 
535: As for the three-point 1PI function $-\Gamma^{\psi\psi p}$, we will see in
536: Sec.~\ref{Galilean} that it is related by Galilean invariance to the
537: $q^2\chi$ term of $\Gamma^{\psi p}$. Because the latter term is not
538: renormalized, we have
539: \begin{equation}
540: -\Gamma^{\psi\psi p} = -\Gamma^{\psi\psi p}_0 = i(k_1^2 - k_2^2) \mathbf{k}_1
541: \times \mathbf{k}_2,
542: \end{equation}
543: up to irrelevant terms with more factors of wavevector. We have thus rendered
544: the theory finite at one loop by renormalizing only the coupling $\bar g$,
545: without the need for counterterms to rescale the fields $\psi$ and $p$ or the
546: time $t$. Crucial for this was the equality of the $1/32\pi\epsilon$ terms in
547: Eqs.\ (\ref{1Lpsip}) and (\ref{1Lpp}). We conclude that the anomalous
548: dimensions are zero at one loop, and in Sec.~\ref{FDT} we will show that in
549: minimal subtraction they are exactly zero,
550: \begin{equation}
551: \gamma_\psi = \gamma_p = \gamma_t = 0.
552: \end{equation}
553: In a different renormalization scheme, or with a different definition of the
554: fields and the time such as Eq.~(\ref{altpsipt}), the anomalous dimensions
555: would not vanish identically, but at any RG fixed point their values are
556: universal \cite{Z96} and so they would be zero there.
557: 
558: The $\beta$ functions for the dimensionless couplings are determined by the
559: RG invariance of the bare couplings,
560: \begin{equation}
561: \left(\mu \frac{\partial}{\partial\mu} + \beta(\bar g)
562: \frac{\partial}{\partial\bar g} + \beta(\bar\rho)
563: \frac{\partial}{\partial\bar\rho}\right) \begin{Bmatrix} g\\ \rho
564: \end{Bmatrix} = 0.
565: \end{equation}
566: We obtain
567: \begin{align}
568: \beta(\bar g) &= -\tfrac{1}{2} \epsilon \bar g + \frac{\bar g^3}{32\pi} +
569: O(\bar g^5),\\
570: \beta(\bar\rho) &= \bigl(-2 + \tfrac{1}{2} \epsilon\bigr) \bar\rho.
571: \end{align}
572: The force correlation adopted by FNS \cite{FNS77} differs from
573: Eq.~(\ref{forcing}) by a factor of $2$, and consequently the dimensionless
574: coupling of FNS is $\bar\lambda = 2^{-1/2} \bar g$. Thus we have confirmed
575: the FNS result
576: \begin{equation}
577: \beta(\bar\lambda) = -\tfrac{1}{2} \epsilon \bar\lambda +
578: \frac{\bar\lambda^3}{16\pi} + O(\bar\lambda^5).
579: \end{equation}
580: 
581: \section{\label{Properties}General properties}
582: 
583: \subsection{\label{Galilean}Galilean invariance}
584: 
585: In two dimensions, the equation of motion (\ref{NSpsi}) and thus the action
586: (\ref{Sfppsi}) have the important physical property of invariance under a
587: Galilean transformation to a reference frame moving with constant velocity
588: $\mathbf{u}$ \cite{FNS77,DM79,AVP83}. This property depends on the assumption
589: that the stirring force is uncorrelated in time, since otherwise there would
590: exist a link between a point in space at one time and a ``corresponding''
591: point in space at a different time. Specifically, the action (\ref{Sfppsi})
592: is invariant under the transformation
593: \begin{equation}
594: \begin{aligned}
595: \label{GalTrans}
596: \psi(t, \mathbf{x}) &\to \psi(t, \mathbf{x} + \mathbf{u}t) + \epsilon_{ij}
597: x_i u_j,\\
598: p(t, \mathbf{x}) &\to p(t,\mathbf{x} + \mathbf{u}t),
599: \end{aligned}
600: \end{equation}
601: which induces the familiar transformation of the velocity,
602: \begin{equation}
603: \mathbf{v}(t, \mathbf{x}) \to \mathbf{v}(t, \mathbf{x} + \mathbf{u}t) -
604: \mathbf{u}.
605: \end{equation}
606: The original Navier--Stokes equation (\ref{NSv}) is not Galilean invariant
607: because of the friction term, which introduces a preferred state of rest; but
608: the differentiation in deriving the stream-function equation (\ref{NSpsi})
609: eliminates the constant shift in $\mathbf{v}$ \cite{H98}.
610: 
611: We would not expect Galilean invariance in $d = 2$ to be preserved by our
612: renormalization method unless the theory remains Galilean invariant when
613: dimensionally regulated. We now show that our formal stream-function
614: representation in arbitrary $d$ is invariant under the transformation
615: (\ref{GalTrans}), provided that $\mathbf{u}$ lies in the physical subspace.
616: For convenience we use the infinitesimal form
617: \begin{equation}
618: \begin{aligned}
619: \label{InfGalTrans}
620: \delta\psi &= t u_i \nabla_i \psi + \epsilon_{ij} x_i u_j, & \delta p &= t
621: u_i \nabla_i p.
622: \end{aligned}
623: \end{equation}
624: In the corresponding variation of the action (\ref{Sfppsi}), terms
625: proportional to $t$ vanish automatically because they correspond to a simple
626: spatial translation. The interesting terms are those where the $t$ is
627: differentiated ($-ip \nabla^2 \dot\psi$) and where $\psi$ is varied by
628: $\epsilon_{ij} x_i u_j$. Neither of these affects the forcing, viscous, or
629: friction terms, which contain $\psi$ only as $\nabla^2 \psi$ and are
630: trivially invariant. We are left with
631: \begin{align}
632: \delta S = -\int dt\, d^d x\, (&ipu_i \nabla_i \nabla^2 \psi + i\epsilon_{ij}
633: \nabla_i \nabla_k p\, \epsilon_{jl} u_l \nabla_k \psi\nonumber\\
634: &+ i\epsilon_{ij} \nabla_i \nabla_k p \nabla_j \psi\, \epsilon_{kl} u_l).
635: \end{align}
636: Upon integration by parts and use of $\Theta_{il} u_l = u_i$, the first two
637: terms cancel and the third vanishes.
638: 
639: In Fourier space, the transformation (\ref{InfGalTrans}) becomes
640: \begin{align}
641: \delta\psi(\omega,\mathbf{k}) &= \mathbf{u} \cdot \mathbf{k}\,
642: \frac{\partial}{\partial\omega} \psi(\omega,\mathbf{k}) - i(2\pi)^{d+1}
643: \delta(\omega)\, \mathbf{u} \times \bm{\nabla} \delta(\mathbf{k}),\nonumber\\
644: \delta p(\omega,\mathbf{k}) &= \mathbf{u} \cdot \mathbf{k}\,
645: \frac{\partial}{\partial\omega} p(\omega,\mathbf{k}).
646: \end{align}
647: The action is Galilean invariant by virtue of the relation
648: \begin{equation}
649: \mathbf{u} \cdot \mathbf{k}\, \frac{\partial}{\partial\omega} \Gamma^{\psi
650: p}_0(\omega,\mathbf{k}) = i \mathbf{u} \times \bm{\nabla}' \Gamma^{\psi\psi
651: p}_0(\omega,\mathbf{k}; 0,\mathbf{k}') |_{\mathbf{k}'=\mathbf{0}};
652: \end{equation}
653: both sides equal $\mathbf{u} \cdot \mathbf{k}\, k^2$. The same relation must
654: then hold between the exact 1PI functions: The coefficient of $k^2 \omega$ in
655: $\Gamma^{\psi p}$ equals the coefficient of $-i(k_1^2 - k_2^2) \mathbf{k}_1
656: \times \mathbf{k}_2$ in $\Gamma^{\psi\psi p}$. Since the former is not
657: renormalized, neither is the latter.
658: 
659: \subsection{\label{FDT}Fluctuation-dissipation theorem}
660: 
661: For zero friction ($\rho = 0$), the action (\ref{Sfppsi}) is equivalent to
662: model A of FNS \cite{FNS77}, who note that it obeys detailed balance and thus
663: is subject to a classical fluctuation-dissipation theorem (FDT) \cite{DH75}.
664: A complicated diagrammatic argument demonstrates that the FDT is preserved to
665: all orders of renormalization \cite{DH75}. Here we reach this conclusion by
666: obtaining the FDT from an exact symmetry of the action in arbitrary $d$.
667: Under the formal discrete transformation
668: \begin{equation}
669: p \to p - 2i\psi,
670: \end{equation}
671: the action with $\rho = 0$ changes only by reversing the sign of the viscous
672: term. The change in the interaction term vanishes upon integration by parts,
673: just as in deriving conservation of energy. To restore the sign of the
674: viscous term, we further perform a complete time reversal,
675: \begin{equation}
676: \begin{aligned}
677: t &\to -t, & \psi &\to -\psi,
678: \end{aligned}
679: \end{equation}
680: which naturally reverses the sign of the dissipation. The net effect is the
681: transformation
682: \begin{equation}
683: \begin{aligned}
684: p(t) &\to p(-t) + 2i\psi(-t), & \psi(t) &\to -\psi(-t),
685: \end{aligned}
686: \end{equation}
687: a generalized time reversal that is its own inverse and leaves the action
688: invariant.
689: 
690: The FDT is derived from this symmetry by expressing the invariance of
691: $\langle p\psi\rangle$:
692: \begin{equation}
693: \label{FDTeq}
694: \langle p\psi\rangle = -\langle\psi p\rangle - 2i \langle\psi\psi\rangle.
695: \end{equation}
696: By invariance under time translations and spatial rotations, negating the
697: times in a two-point correlation function is equivalent to interchanging the
698: points. As a first application of the FDT, we make use of the theorem that
699: the equal-time correlation function $\langle p\psi\rangle_= = \langle\psi
700: p\rangle_=$ is exact at tree level \cite{AVP83}. Thus, for $\rho = 0$, the
701: exact equal-time stream-function correlation is
702: \begin{align}
703: \langle\psi\psi\rangle_= &= \tfrac{1}{2}i \bigl(\langle p\psi\rangle_= +
704: \langle\psi p\rangle_=\bigr)\nonumber\\
705: &= \tfrac{1}{2}i \int_{-\infty}^\infty \frac{d\omega}{2\pi} \bigl(\langle
706: p\psi\rangle_0 + \langle\psi p\rangle_0\bigr)\nonumber\\
707: &= \tfrac{1}{2}i \int_{-\infty}^\infty \frac{d\omega}{2\pi}
708: \frac{-2ig^{-1}}{\omega^2 + g^{-2} k^4} = \frac{1}{2k^2}.
709: \end{align}
710: The energy spectrum, in the units implied by Eq.~(\ref{ppsit}), is then
711: \begin{equation}
712: E(k) = \frac{\pi k}{(2\pi)^2} k^2 \langle\psi\psi\rangle_= = \frac{k}{8\pi}.
713: \end{equation}
714: This is an equipartition spectrum, in agreement with FNS model A
715: \cite{FNS77}.
716: 
717: For $\rho = 0$, the FDT (\ref{FDTeq}) also implies that, if $\langle
718: p\psi\rangle$ and thus $\langle\psi p\rangle$ are made finite by
719: renormalizing $\bar g$, then $\langle\psi\psi\rangle$ is likewise finite,
720: without the need for field or time rescalings. As we have seen, the
721: three-point 1PI function is automatically finite. Hence the anomalous
722: dimensions vanish exactly for $\rho = 0$. But in minimal subtraction, the
723: counterterms for rescalings and for dimensionless couplings are independent
724: of the mass parameter \cite{T73}. We conclude that in minimal subtraction,
725: even with friction,
726: \begin{equation}
727: \gamma_\psi = \gamma_p = \gamma_t = 0,
728: \end{equation}
729: and $\beta(\bar g)$ depends only on $\bar g$. Indeed, we have seen these
730: statements verified to one loop in Sec.~\ref{OneLoop}.
731: 
732: \subsection{\label{RG}Renormalization-group flows}
733: 
734: We have shown that the anomalous dimensions vanish, and that the renormalized
735: couplings in exactly two dimensions obey
736: \begin{align}
737: \label{RGgbar}
738: \mu \frac{d\bar g}{d\mu} &\equiv \beta(\bar g) = \frac{\bar g^3}{32\pi} +
739: O(\bar g^5),\\
740: \label{RGrhobar}
741: \bar\rho &= \mu^{-2} \rho.
742: \end{align}
743: Thus $\bar\rho$ is very simply related to $\mu$ and can be used to
744: parametrize it. In the RG flow to low wavenumbers, $\bar\rho$ steadily
745: increases as friction becomes more important. Meanwhile, $\bar g$ flows in
746: its own characteristic way regardless of the value of $\bar\rho$; the most we
747: can say is that once $\bar g$ becomes small, it decreases further and
748: further, approaching zero in the IR limit. The solution of Eq.~(\ref{RGgbar})
749: in this limit is
750: \begin{equation}
751: \begin{aligned}
752: \bar g(\mu) &= \sqrt{\frac{16\pi}{\ln(k_g/\mu)}} & (\mu &\ll k_g),
753: \end{aligned}
754: \end{equation}
755: where $k_g$ is the scale at which $\bar g$ becomes large.
756: 
757: In the IR limit, we expect good accuracy from perturbative results such as
758: the tree-level expression for the energy spectrum,
759: \begin{equation}
760: \label{Ekp}
761: E(k) = \frac{\pi k}{(2\pi)^2} k^2 \langle\psi\psi\rangle_= =
762: \frac{k^3}{8\pi(g\rho + k^2)}.
763: \end{equation}
764: We might suppose that the true asymptotic behavior is given by replacing $g$
765: with the renormalized coupling $\bar g(k)$. This would follow from RG theory
766: if the loop corrections to Eq.~(\ref{Ekp}) contained $\ln(k^2)$. But with
767: $g\rho$ providing an IR cutoff, the diagrams are regular as $k \to 0$ and
768: instead contain $\ln(g\rho)$. Hence, for the purposes of correlation
769: functions, the RG flow effectively halts for wavenumbers below the ``mass''
770: $\sqrt{g\rho}$. If $m$ is the suitably renormalized value of this mass, then
771: as $k \to 0$ we expect that
772: \begin{equation}
773: \label{Ekp1}
774: E(k) = \frac{k^3}{8\pi\rho\, \bar g(m)}.
775: \end{equation}
776: On the other hand, in the bare tree-level result (\ref{Ekp}), $g$ can be
777: interpreted as a coupling renormalized at a very high wavenumber (the forcing
778: scale or UV cutoff). With $\beta(\bar g) > 0$, we have $\bar g(m) < g$, and
779: so the RG result (\ref{Ekp1}) gives a greater $E(k)$ at low $k$. Whereas the
780: bare tree-level calculation ignores all interactions between scales, the
781: effect of renormalization is to place more energy and dissipation at low $k$
782: (and therefore less at high $k$), consistent with the inverse cascade.
783: 
784: From Eq.~(\ref{RGrhobar}), any RG fixed point $(\bar g_*,\bar\rho_*)$ must
785: have $\bar\rho_* = 0$. Not only do the anomalous dimensions vanish at any
786: fixed point, but for $\bar\rho = 0$ we have the exact equipartition spectrum
787: $E(k) \propto k^1$, whether $\bar g$ is at a fixed point or not. It is clear
788: that, despite the suggestive evidence of scale invariance of the inverse
789: cascade, an RG fixed point in our framework cannot be the explanation for the
790: observed $k^{-5/3}$ spectrum. Nevertheless our theory contains all the
791: essential ingredients that have produced the stationary inverse cascade
792: experimentally and numerically. A natural explanation is that the $k^{-5/3}$
793: spectrum arises from the nonperturbative behavior of correlation functions at
794: $\bar\rho > 0$ and at large values of $\bar g$ that flow rapidly with scale.
795: Although it is far from obvious how such an RG flow can produce approximate
796: scale invariance with an effective anomalous dimension, we are motivated to
797: seek hints about the theory's strong-coupling behavior. The first step, which
798: can be useful for more mundane purposes as well, is to extend the
799: renormalization to the next order of perturbation theory.
800: 
801: \section{\label{TwoLoop}Two-loop renormalization}
802: 
803: \subsection{\label{Prescrip}Renormalization prescription and diagrams}
804: 
805: Calculating $\beta(\bar g)$ consistently to two loops requires a precise
806: specification of the renormalization scheme. Remarkably, though, as long as
807: $\beta(\bar g)$ depends only on $\bar g$, the $\beta$ function to two loops
808: (but no further) is independent of the particular scheme chosen \cite{Z96}.
809: As in Sec.~\ref{OneLoop}, we take $\rho > 0$, compute two-point 1PI diagrams
810: to $O(q^4)$ expanded about $\epsilon = 0$, and renormalize by minimal
811: subtraction. The two-loop expression for the bare coupling in terms of the
812: renormalized coupling is
813: \begin{equation}
814: \label{2Lg}
815: g = \mu^{\epsilon/2} \bigl[\bar g + \bar g^3 a_1 \epsilon^{-1} + \bar g^5
816: (a_2 \epsilon^{-2} + a_2' \epsilon^{-1}) + O(\bar g^7)\bigr].
817: \end{equation}
818: With zero anomalous dimensions, $\beta(\bar g)$ is the only RG function to be
819: determined and we need only compute a single 1PI function to two loops. Below
820: we will choose $-\Gamma^{pp}$ because its two-loop diagrams are technically
821: simpler than those of $-\Gamma^{\psi p}$.
822: 
823: We have seen that $g\rho$, which has dimension $2$ independent of $\epsilon$,
824: acts as the IR cutoff for wavevector integrals. By dimensional analysis,
825: $\Gamma^{pp}$ to $O(q^4)$ has the form
826: \begin{alignat}{2}
827: \Gamma^{pp} &= q^4 &&\biggl[g^{-1} + g (g\rho)^{-\epsilon/2}
828: \left(\frac{b_1}{\epsilon} + b_1' + O(\epsilon)\right)\nonumber\\
829: &&&+ g^3 (g\rho)^{-\epsilon} \left(\frac{b_2}{\epsilon^2} +
830: \frac{b_2'}{\epsilon} + O(\epsilon^0)\right) + O(g^5)\biggr]\nonumber\\
831: &= q^4 &&\biggl[g^{-1} + g \left(\frac{b_1}{\epsilon} + b_1' - \tfrac{1}{2}
832: b_1 \ln(g\rho) + O(\epsilon)\right)\nonumber\\
833: \label{2Lpp}
834: &&&+ g^3 \left(\frac{b_2}{\epsilon^2} + \frac{b_2' - b_2
835: \ln(g\rho)}{\epsilon} + O(\epsilon^0)\right) + O(g^5)\biggr].
836: \end{alignat}
837: To simplify the two-loop calculations we formally set $\rho = g^{-1}$ and
838: restore the $\ln(g\rho)$ terms at the end.
839: 
840: \begin{figure}
841: \includegraphics{icfig3.eps}%
842: \caption{\label{2L1PI}Nonvanishing two-loop diagrams for the two-point 1PI
843: function $-\Gamma^{pp}$. Along with the overall divergence, each diagram has
844: at most one divergent subdiagram (bold lines).}
845: \end{figure}
846: 
847: Multiloop diagrams can be characterized by their overall UV divergence (as
848: all loop frequencies and wavenumbers go to infinity together) and their
849: subdivergences (as a subset of loop frequencies and wavenumbers go to
850: infinity while the rest remain finite). Overlapping divergences occur when
851: two or more divergent subdiagrams share a propagator; this is a definite
852: complication in evaluating a diagram, though not insurmountable \cite{Z96}.
853: Some two-loop diagrams for the 1PI function $-\Gamma^{\psi p}$ contain
854: overlapping divergences, so we choose to calculate the diagrams for
855: $-\Gamma^{pp}$, which fortunately do not (Fig.~\ref{2L1PI}). A three-point
856: 1PI subdiagram is divergent only if its external lines are $\psi\psi p$, two
857: wiggly and one plain; Galilean invariance does not eliminate this
858: subdivergence when we treat each two-loop diagram separately, because the
859: finiteness of $-\Gamma^{\psi\psi p}_1$ results from a sum over distinct
860: one-loop diagrams. In Fig.~\ref{2L1PI} we have omitted one conceivable
861: diagram containing the 1PI subdiagram $-\Gamma^{\psi\psi}_1$, which vanishes
862: as noted in Sec.~\ref{Feynman}.
863: 
864: Previous two-loop calculations of similar complexity have been made for
865: different problems using diagrams of the same topology: the
866: Burgers--Kardar-Parisi-Zhang equation for interface growth \cite{FT94} and
867: the Navier--Stokes equation in more than two physical dimensions, where there
868: are fewer divergences \cite{AAKV03}.
869: 
870: \subsection{\label{Analytic}Analytic calculations}
871: 
872: We have programmed \textsc{mathematica} \cite{W99} to automate the steps in
873: evaluating each two-loop diagram. The loop frequencies are integrated one
874: after the other by adding the residues of all poles in the upper half-plane.
875: The resulting integrand, containing the external wavevector $\mathbf{q}$ and
876: the loop wavevectors $\mathbf{k}_{1,2}$, is expanded to $O(q^4)$ and averaged
877: over directions of $\mathbf{q}$ in the physical subspace. The numerator of
878: the integrand is now rife with the projector $\Theta_{ij}$, both from
879: averaging $q_i q_j$ and from applying Eq.~(\ref{epseps}) to cross products.
880: To eliminate $\Theta_{ij}$, we average the integrand over orientations of the
881: physical subspace within $d$-dimensional space: We introduce an orthonormal
882: physical basis, write $\Theta_{ij} = \hat a_i \hat a_j + \hat b_i \hat b_j$,
883: average over directions of $\hat{\mathbf{a}}$ in the $d - 1$ dimensions
884: perpendicular to $\hat{\mathbf{b}}$, and then average over directions of
885: $\hat{\mathbf{b}}$ in $d$-dimensional space. The result is a function of
886: $k_1^2$, $k_2^2$, and $\mathbf{k}_1 \cdot \mathbf{k}_2$ to be integrated over
887: $d^d k_1\, d^d k_2$.
888: 
889: Because we need only the divergent parts of the two-loop diagrams as
890: indicated in Eq.~(\ref{2Lpp}), we eliminate numerator terms that produce
891: neither an overall divergence nor a subdivergence by power counting. We seek
892: to integrate first over the loop wavevector (say $\mathbf{k}_1$) associated
893: with the subdivergence (if any). To simplify the denominator, Feynman
894: parameters \cite{Z96} are introduced, and $\mathbf{k}_1$ is translated by a
895: multiple of $\mathbf{k}_2$. After the numerator is isotropized, the
896: $\mathbf{k}_1$ integration is done analytically; there remains an integral
897: over Feynman parameters and over the magnitude $k_2$. From Eq.~(\ref{2Lpp}),
898: the integrand (excluding $g^3 q^4$) has dimension $-1 - 2\epsilon$. The part
899: that behaves like $k_2^{-1-2\epsilon}$ as $k_2 \to \infty$ is subtracted and
900: separately integrated over $k_2$, producing another $\epsilon^{-1}$ factor;
901: all the $\epsilon^{-2}$ terms arise here and are found analytically, but the
902: $\epsilon^{-1}$ terms involve intractable Feynman-parameter integrals.
903: Meanwhile, the remaining subtracted integral converges at $k_2 = \infty$ but
904: contains analytically integrable $\epsilon^{-1}$ poles from the
905: $\mathbf{k}_1$ integration.
906: 
907: Adding the divergent parts of the eight diagrams gives
908: \begin{align}
909: \Gamma^{pp}_2 = g^3 q^4 \biggl(&{-}\frac{1}{2048\pi^2 \epsilon^2}\nonumber\\
910: \label{an2Lpp}
911: &+ \frac{-2\ln(4\pi) + 2\gamma_{\text{E}} + 5 + X}{4096\pi^2 \epsilon} +
912: O(\epsilon^0)\biggr).
913: \end{align}
914: Here $X$ is a sum of Feynman-parameter integrals of complicated rational
915: functions with integer coefficients; thus we expect that $X$ may be a
916: rational number. Though we are unable to calculate $X$ analytically,
917: Eq.~(\ref{an2Lpp}) already displays some important features: If the
918: $\epsilon^{-2}$ term were different, the renormalization performed below
919: would become inconsistent \cite{T73}; and if the coefficients of
920: $\gamma_{\text{E}}$ and $\ln(4\pi)$ were different, these constants would
921: (contrary to expectation) appear in the two-loop $\beta$ function.
922: 
923: \subsection{\label{Numerical}Numerical calculations}
924: 
925: We have evaluated $X$ numerically by multidimensional Monte Carlo
926: integration, using the technique of importance sampling \cite{PTVF92} to
927: select more points in the ``corners'' of Feynman-parameter space (with one
928: parameter near $1$ and the others near $0$). This technique improves the
929: statistics because the integrands tend to diverge in these corners (but
930: slowly enough that the integrals converge). We treat the integrals that make
931: up $X$ separately, since they vary in complexity and in number of Feynman
932: parameters. To optimize the precision of the result for $X$ in a given
933: computation time $T$, we reason as follows. The contribution to $X$ from
934: diagram $i$, evaluated with $n_i$ sample points, is obtained with a precision
935: $s_i = \sigma_i n_i^{-1/2}$ in a time $t_i = \tau_i n_i$, where $\sigma_i$
936: and $\tau_i$ are characteristics of the integrand and the computer.
937: Constrained optimization shows that we achieve the best overall precision
938: \begin{equation}
939: S^2 = \sum_i s_i^2
940: \end{equation}
941: in the time
942: \begin{equation}
943: T = \sum_i t_i
944: \end{equation}
945: by choosing
946: \begin{equation}
947: n_i = \frac{T\sigma_i}{\tau_i^{1/2} \sum_j \sigma_j \tau_j^{1/2}}.
948: \end{equation}
949: Short runs are made to estimate $\sigma_i$ and $\tau_i$, and then the
950: high-precision integrations are performed with this plan. Our computations
951: for a total of $3.6 \times 10^8$ sample points yield
952: \begin{equation}
953: X = -3.995 \pm 0.005,
954: \end{equation}
955: strongly suggesting that the exact value is
956: \begin{equation}
957: X = -4.
958: \end{equation}
959: 
960: With the one-loop result (\ref{1Lpp}) and the two-loop result (\ref{an2Lpp}),
961: we have
962: \begin{multline}
963: \Gamma^{pp} = q^4 \biggl[g^{-1} + g \left(\frac{1}{32\pi\epsilon} -
964: \frac{\ln(g\rho/4\pi) + \gamma_{\text{E}}}{64\pi} + O(\epsilon)\right)\\
965: + g^3 \left(-\frac{1}{2048\pi^2 \epsilon^2} + \frac{2\ln(g\rho/4\pi) +
966: 2\gamma_{\text{E}} + 1}{4096\pi^2 \epsilon} + O(\epsilon^0)\right)\\
967: + O(g^5)\biggr].
968: \end{multline}
969: By substituting Eq.~(\ref{2Lg}) and requiring a finite expression at
970: $\epsilon = 0$, we determine
971: \begin{equation}
972: \begin{aligned}
973: a_1 &= \frac{1}{32\pi}, & a_2 &= \frac{3}{2048\pi^2}, & a_2' &=
974: -\frac{1}{4096\pi^2}.
975: \end{aligned}
976: \end{equation}
977: The RG invariance of the bare coupling $g$ finally gives
978: \begin{equation}
979: \beta(\bar g) = -\tfrac{1}{2} \epsilon \bar g + \frac{\bar g^3}{32\pi} -
980: \frac{\bar g^5}{2048\pi^2} + O(\bar g^7).
981: \end{equation}
982: This is reminiscent of the $\beta$ function in four-dimensional $\phi^4$
983: theory, where similarly the one-loop term is positive and the two-loop term
984: is negative \cite{Z96}. A positive $\beta$ function that grows too quickly
985: (faster than linearly) at large coupling raises the question of whether the
986: coupling becomes infinite at a large but finite renormalization scale. In our
987: theory this would suggest an absolute limit on the extent of the scaling
988: range for any inverse cascade. To avoid such a fate, the expansion of the
989: $\beta$ function must contain many negative terms to slow its initial
990: superlinear growth. It is pleasing to find such a term already at two loops.
991: 
992: \section{\label{Cascade}Inverse-cascade range}
993: 
994: \subsection{\label{Phenom}Inverse-cascade phenomenology}
995: 
996: The initial prediction of the two-dimensional inverse energy cascade
997: \cite{K67} assumed zero friction and small but nonzero viscosity. Kinetic
998: energy, continually injected at high wavenumbers, is expected to cascade down
999: through a quasisteady inertial range extending to lower and lower wavenumbers
1000: as time passes. Within this range, the assumption of scale invariance implies
1001: the energy spectrum
1002: \begin{equation}
1003: \label{Ek}
1004: E(k) = C\, \mathcal{E}^{2/3} k^{-5/3},
1005: \end{equation}
1006: where $C$ is the two-dimensional Kolmogorov constant and $\mathcal{E}$ is the
1007: rate of energy injection per unit mass. Numerical simulations with zero
1008: friction \cite{FS84,SY93,SY94} confirm this spectrum until the cascade
1009: approaches the minimum wavenumber associated with a finite system. Both
1010: nonstationary behavior and finite-size effects, however, lie outside our
1011: theoretical framework, which treats a fluid of infinite size that has been
1012: stirred for an infinite time. In the absence of friction, such a fluid has an
1013: equipartition spectrum as shown in Sec.~\ref{FDT}.
1014: 
1015: When friction is introduced, it is natural to expect a mirror image of the
1016: viscous energy dissipation at high wavenumbers in the three-dimensional
1017: direct cascade: A stationary inverse cascade should develop, with dissipation
1018: by friction at low wavenumbers $k \sim k_{\text{fr}}$ and with the spectrum
1019: (\ref{Ek}) at wavenumbers $k \gg k_{\text{fr}}$ where dissipation is
1020: unimportant. Dimensional analysis gives
1021: \begin{equation}
1022: k_{\text{fr}} = \mathcal{E}^{-1/2} \alpha^{3/2},
1023: \end{equation}
1024: where $\alpha$ is the friction coefficient \cite{BCV00,DG01}. Several
1025: numerical simulations \cite{SY94,BCV00} and laboratory experiments
1026: \cite{S86,PT98} confirm this picture. Other numerical studies obtain similar
1027: results but are more difficult to relate to our framework, since they modify
1028: the friction term by removing derivatives \cite{BDF95} or by applying
1029: friction only below a cutoff wavenumber \cite{MV91}. We are less concerned
1030: about the common practice of adding derivatives to the viscous term
1031: (hyperviscosity), because such a modified term is irrelevant and the RG flow
1032: should introduce a normal viscous term to replace it. But friction modified
1033: with inverse derivatives (hypofriction) is certainly relevant and is believed
1034: to alter the dynamics of the inverse cascade \cite{SY94,BCV00}. In an extreme
1035: case, hypofriction with eight inverse Laplacians destroys the $k^{-5/3}$
1036: spectrum \cite{B94}. Hypofriction is intended to confine dissipation
1037: explicitly to low wavenumbers; ordinary linear friction accomplishes the same
1038: thing more gently, but makes it difficult in practice to achieve an inertial
1039: range \cite{DG01}. We have used linear friction because it is physically
1040: realistic and leads to a local field theory.
1041: 
1042: For convenience in relating the observed inverse cascade to our field theory
1043: with the action (\ref{Sfppsi}), we adopt units of time in which
1044: \begin{equation}
1045: g\nu \equiv D_0^{1/2} \nu^{-1/2} = 1,
1046: \end{equation}
1047: so that the rescalings in Eq.~(\ref{ppsit}) are trivial. Then the kinematic
1048: viscosity is
1049: \begin{equation}
1050: \label{nug}
1051: \nu = g^{-1},
1052: \end{equation}
1053: the friction coefficient is
1054: \begin{equation}
1055: \label{alpharho}
1056: \alpha = \rho,
1057: \end{equation}
1058: and the force correlation is
1059: \begin{equation}
1060: \label{fcg}
1061: D(k^2) = g^{-1} k^2.
1062: \end{equation}
1063: 
1064: In contrast to the formal methods of dimensional regularization and minimal
1065: subtraction, a simple UV cutoff renders our theory finite in a physically
1066: meaningful way while still preserving the symmetries noted in
1067: Sec.~\ref{Properties}. A cutoff would have been inconvenient for the two-loop
1068: calculations of Sec.~\ref{TwoLoop}, but it is appropriate for understanding
1069: experimental and numerical results on the inverse cascade. The local,
1070: renormalizable action (\ref{Sfppsi}) applies only with a UV cutoff $\Lambda$
1071: below the wavenumbers of the external forcing \cite{FNS77}. A corresponding
1072: renormalization prescription is obtained by staying in exactly two spatial
1073: dimensions and writing the coupling $g$ in Eq.~(\ref{Sfppsi}) as a
1074: cutoff-dependent quantity $\hat g(\Lambda)$, such that the long-distance
1075: behavior is independent of $\Lambda$. No cutoff dependence is needed for
1076: $\rho$, because the friction term is not renormalized, as we saw in
1077: Sec.~\ref{OneLoop}.
1078: 
1079: Special properties of minimal subtraction allowed us to conclude
1080: (Sec.~\ref{FDT}) that in that scheme the anomalous dimensions vanish and
1081: $\beta(\bar g)$ depends only on $\bar g$. We may expect that in the cutoff
1082: scheme these statements remain approximately true, at least for small values
1083: of the dimensionless friction parameter $\Lambda^{-2} \rho$. Because the
1084: $\beta$ function is scheme independent through two loops \cite{Z96}, we have
1085: \begin{equation}
1086: \label{Lbeta}
1087: \Lambda \frac{d\hat g}{d\Lambda} \equiv \beta(\hat g) = \frac{\hat
1088: g^3}{32\pi} - \frac{\hat g^5}{2048\pi^2} + O(\hat g^7).
1089: \end{equation}
1090: This result is directly useful for weak coupling, but can only hint at the
1091: possible strong-coupling behavior. We will now show that a specific
1092: strong-coupling form of the $\beta$ function is required for a fully
1093: developed inverse cascade, and that it can be naturally interpolated with the
1094: two-loop result (\ref{Lbeta}).
1095: 
1096: \subsection{\label{Strong}Strong-coupling behavior}
1097: 
1098: The key condition for the ideal inverse cascade is that the dissipation of
1099: energy is dominated by friction and is almost totally confined to low
1100: wavenumbers. This means that the dissipation rate
1101: \begin{equation}
1102: \label{diss}
1103: \mathcal{E} = 2\rho \int_0^\Lambda dk\, E(k) = 2\rho \int_0^\Lambda dk\,
1104: \frac{\pi k}{(2\pi)^2} k^2 \langle\psi\psi\rangle_=
1105: \end{equation}
1106: should be independent of the UV cutoff $\Lambda$, because a change in
1107: $\Lambda$ produces neither a renormalization of $\rho$ nor a rescaling of
1108: $\psi$. Under stationary conditions, $\mathcal{E}$ equals the rate of energy
1109: injection, given by \cite{AAV99}
1110: \begin{equation}
1111: \label{EgL}
1112: \mathcal{E} = \int_0^\Lambda dk\, \frac{\pi k}{(2\pi)^2}\, D(k^2) =
1113: \frac{\Lambda^4}{16\pi\, \hat g(\Lambda)},
1114: \end{equation}
1115: in terms of the force correlation (\ref{fcg}). For $\mathcal{E}$ to be
1116: independent of $\Lambda$, we must have the strong-coupling behavior
1117: \begin{equation}
1118: \hat g(\Lambda) \simeq \text{const} \times \Lambda^4,
1119: \end{equation}
1120: corresponding to the asymptotically linear $\beta$ function
1121: \begin{equation}
1122: \begin{aligned}
1123: \beta(\hat g) &\simeq 4\hat g & (\hat g &\to \infty).
1124: \end{aligned}
1125: \end{equation}
1126: We now attempt to connect this form with the two-loop $\beta$ function
1127: (\ref{Lbeta}).
1128: 
1129: Perturbative expansions such as Eq.~(\ref{Lbeta}) are usually divergent, but
1130: a Borel transformation is expected to produce a finite radius of convergence
1131: about zero coupling \cite{Z96}. We see that the true expansion parameter is
1132: $\hat g^2$, and the alternatively normalized action (\ref{Sippsi}) makes it
1133: clear that the theory is unstable for $\hat g^2 < 0$. Thus we write
1134: \begin{equation}
1135: \label{BT}
1136: \frac{\beta(\hat g)}{\hat g} \equiv A(\hat g^2) = \int_0^\infty
1137: \frac{dz}{\hat g^2}\, B(z) \exp\frac{-z}{\hat g^2}.
1138: \end{equation}
1139: For the perturbation series
1140: \begin{equation}
1141: A(\hat g^2) = \sum_{n=1}^\infty A_n \hat g^{2n},
1142: \end{equation}
1143: the Borel transform is
1144: \begin{equation}
1145: B(z) = \sum_{n=1}^\infty \frac{A_n}{n!} z^n.
1146: \end{equation}
1147: In our case,
1148: \begin{equation}
1149: \label{Bser}
1150: B(z) = \frac{z}{32\pi} - \frac{z^2}{4096\pi^2} + O(z^3).
1151: \end{equation}
1152: 
1153: The Borel transform often has poles on the negative real axis associated with
1154: instantons \cite{Z96}, but $\beta(\hat g)$ is well-defined from
1155: Eq.~(\ref{BT}) as long as $B(z)$ is regular on the positive real axis. A
1156: simple and suitable rational (Pad\'e) approximant to $B(z)$ is
1157: \begin{equation}
1158: \begin{aligned}
1159: B(z) &\simeq \frac{z}{32\pi + yz} & (y &\ge 0);
1160: \end{aligned}
1161: \end{equation}
1162: fortunately, the choice $y = \frac{1}{4}$ agrees with Eq.~(\ref{Bser}). We
1163: thus take
1164: \begin{equation}
1165: B(z) \simeq \frac{4z}{128\pi + z},
1166: \end{equation}
1167: but we do not here attempt to study the possible instanton solutions
1168: corresponding to the pole at $z = -128\pi$. Eq.~(\ref{BT}) then gives
1169: precisely the desired asymptotic behavior
1170: \begin{equation}
1171: \begin{aligned}
1172: \beta(\hat g) &\simeq 4\hat g & (\hat g &\to \infty).
1173: \end{aligned}
1174: \end{equation}
1175: 
1176: We now ask whether the observed $k^{-5/3}$ energy spectrum (\ref{Ek}) is
1177: consistent with our theory, although we are unable to derive it
1178: systematically. We conjecture that strong-coupling effects produce the
1179: spectrum
1180: \begin{equation}
1181: \label{Ek1}
1182: E(k) \sim \mathcal{E}^{2/3} k^{-5/3}
1183: \end{equation}
1184: for
1185: \begin{equation}
1186: k \gtrsim k_{\text{fr}} = \mathcal{E}^{-1/2} \rho^{3/2}.
1187: \end{equation}
1188: From Eq.~(\ref{EgL}), the running coupling is
1189: \begin{equation}
1190: \hat g(\Lambda) \sim \mathcal{E}^{-1} \Lambda^4.
1191: \end{equation}
1192: For nonperturbative effects to be operative down to the wavenumber
1193: $k_{\text{fr}}$, we must have
1194: \begin{equation}
1195: \hat g(k_{\text{fr}}) \sim (\mathcal{E}^{-1} \rho^2)^3 \gtrsim 1.
1196: \end{equation}
1197: In the borderline case where $\hat g(k_{\text{fr}}) \sim 1$, we have
1198: $k_{\text{fr}} \sim \mathcal{E}^{1/4} \sim \rho^{1/2}$, and we can match
1199: Eq.~(\ref{Ek1}) in order of magnitude with the perturbative energy spectrum
1200: (\ref{Ekp}):
1201: \begin{equation}
1202: E(k_{\text{fr}}) \sim \mathcal{E}^{2/3} k_{\text{fr}}^{-5/3} \sim
1203: \mathcal{E}^{1/4} \sim \frac{k_{\text{fr}}^3}{\rho + k_{\text{fr}}^2}.
1204: \end{equation}
1205: Even when $\hat g(k_{\text{fr}}) \gg 1$, we may guess from Eq.~(\ref{Ekp})
1206: that in order of magnitude
1207: \begin{equation}
1208: E(k_{\text{fr}}) \sim \frac{k_{\text{fr}}^3}{\rho\, \hat g(k_{\text{fr}}) +
1209: k_{\text{fr}}^2} \sim \mathcal{E}^{3/2} \rho^{-5/2},
1210: \end{equation}
1211: and this again matches Eq.~(\ref{Ek1}) at $k_{\text{fr}}$.
1212: 
1213: In the case $\hat g(k_{\text{fr}}) \gg 1$, the running coupling eventually
1214: becomes $\sim 1$ at a lower wavenumber
1215: \begin{equation}
1216: k_g \sim \mathcal{E}^{1/4},
1217: \end{equation}
1218: below which perturbation theory is applicable and the energy spectrum is
1219: given roughly by Eq.~(\ref{Ekp1}). Thus we envision a varied but continuous
1220: behavior of the energy spectrum in the different regions:
1221: \begin{equation}
1222: E(k) \sim
1223: \begin{cases}
1224: \rho^{-1} k^3 & (k \lesssim \mathcal{E}^{1/4}),\\
1225: \mathcal{E} \rho^{-1} k^{-1} & (\mathcal{E}^{1/4} \lesssim k \lesssim
1226: \mathcal{E}^{-1/2} \rho^{3/2}),\\
1227: \mathcal{E}^{2/3} k^{-5/3} & (k \gtrsim \mathcal{E}^{-1/2} \rho^{3/2}).
1228: \end{cases}
1229: \end{equation}
1230: As a further check, we note that the contribution to the energy dissipation
1231: rate (\ref{diss}) for each of the three regions is $\sim \mathcal{E}$ (up to
1232: a logarithmic factor for the middle region). This suggests that a fully
1233: developed stationary inverse cascade has several distinct dissipation ranges.
1234: Testing by experiments and simulations is not straightforward, because
1235: finite-size effects may become important before the lower-wavenumber ranges
1236: are reached.
1237: 
1238: \subsection{\label{Flux}Energy flux and third moment}
1239: 
1240: We have described the ideal inverse cascade in terms of external forcing
1241: confined to high wavenumbers (assumed in Sec.~\ref{Relevance}) and energy
1242: dissipation practically confined to low wavenumbers (made plausible in Secs.\
1243: \ref{RG} and \ref{Strong}). This separation is equivalent to the conventional
1244: criterion of an inertial range with a constant energy flux. However, unlike
1245: the energy spectrum (a simple correlation function), the energy flux in
1246: wavenumber space is not invariant under the RG flow and is not given
1247: straightforwardly by our local field theory. Renormalization introduces the
1248: forcing (\ref{fcg}), which is nonzero even for wavenumbers in the inertial
1249: range and so makes the flux appear nonconstant. This effective forcing simply
1250: represents the energy transfer from wavenumbers $k > \Lambda$ that have been
1251: integrated out.
1252: 
1253: We can resolve the flux problem by working instead in physical space, where
1254: the effective forcing is a differentiated delta function that is zero at
1255: finite distances, including the inertial range of lengths. Thus we inquire
1256: whether the physical-space energy flux \cite{F95}
1257: \begin{equation}
1258: \label{er}
1259: \varepsilon(r) = -\tfrac{1}{4} \bm{\nabla}_{\mathbf{r}} \cdot \langle
1260: |\delta\mathbf{v}(\mathbf{r})|^2 \delta\mathbf{v}(\mathbf{r}) \rangle
1261: \end{equation}
1262: is preserved under renormalization. The flux is given in terms of the third
1263: moment of the velocity increment
1264: \begin{equation}
1265: \delta\mathbf{v}(\mathbf{r}) = \mathbf{v}(\mathbf{x}+\mathbf{r}) -
1266: \mathbf{v}(\mathbf{x})
1267: \end{equation}
1268: and is independent of $\mathbf{x}$ by homogeneity. Under stationary
1269: conditions, the Navier--Stokes equation (\ref{NSv}) yields \cite{B99}
1270: \begin{align}
1271: \varepsilon(r) &= (\nu\nabla_{\mathbf{r}}^2 - \alpha) \langle
1272: \mathbf{v}(\mathbf{x}) \cdot \mathbf{v}(\mathbf{x}+\mathbf{r}) \rangle +
1273: \tfrac{1}{2} \hat C(r)\nonumber\\
1274: \label{er1}
1275: &= [\hat g(\Lambda)^{-1} \nabla_{\mathbf{r}}^2 - \rho] \langle
1276: \mathbf{v}(\mathbf{x}) \cdot \mathbf{v}(\mathbf{x}+\mathbf{r}) \rangle,
1277: \end{align}
1278: where $\hat C(r)$ is the vanishing physical-space force correlation, and we
1279: have used Eqs.\ (\ref{nug}) and (\ref{alpharho}).
1280: 
1281: The right-hand side of Eq.~(\ref{er1}) can be interpreted as minus the rate
1282: of energy dissipation at length scales larger than $r$. Since $\langle
1283: \mathbf{v}(\mathbf{x}) \cdot \mathbf{v}(\mathbf{x}+\mathbf{r}) \rangle$ is an
1284: RG-invariant correlation function, only the viscosity $\hat g(\Lambda)^{-1}$
1285: introduces cutoff dependence. In accordance with our argument in
1286: Sec.~\ref{Strong}, as long as the energy dissipation is dominated by
1287: friction, the dissipation rate is independent of $\Lambda$, giving a
1288: well-defined energy flux
1289: \begin{equation}
1290: \varepsilon(r) = -\rho \langle \mathbf{v}(\mathbf{x}) \cdot
1291: \mathbf{v}(\mathbf{x}+\mathbf{r}) \rangle
1292: \end{equation}
1293: proportional to the Fourier transform of the energy spectrum. And if this
1294: spectrum is almost entirely concentrated at low wavenumbers $k \lesssim
1295: k_{\text{fr}}$, it follows that in the inertial range ($r \ll
1296: k_{\text{fr}}^{-1}$) the flux is constant,
1297: \begin{equation}
1298: \varepsilon(r) \simeq \varepsilon(0) = -\rho \langle v^2 \rangle =
1299: -\mathcal{E}.
1300: \end{equation}
1301: 
1302: \subsection{\label{Intermitt}Intermittency}
1303: 
1304: A striking feature observed in the inverse cascade is the approximate scale
1305: invariance of inertial-range velocity correlations (structure functions). The
1306: $k^{-5/3}$ energy spectrum (\ref{Ek}) gives for the second moment \cite{F95}
1307: \begin{equation}
1308: \label{vv}
1309: \langle |\delta\mathbf{v}(\mathbf{r})|^2 \rangle \propto \mathcal{E}^{2/3}
1310: r^{2/3},
1311: \end{equation}
1312: and the constant energy flux (\ref{er}) gives for the third moment \cite{B99}
1313: \begin{equation}
1314: \label{vvv}
1315: \langle |\delta\mathbf{v}(\mathbf{r})|^2 \delta\mathbf{v}(\mathbf{r}) \rangle
1316: \propto \mathcal{E} \mathbf{r}.
1317: \end{equation}
1318: In our field theory, since the anomalous dimensions vanish, such moments can
1319: be written in the form
1320: \begin{equation}
1321: \label{vn}
1322: \langle (\delta v)^n \rangle = r^{-n} f_n\bm{(}\hat g(r^{-1}), \rho
1323: r^2\bm{)},
1324: \end{equation}
1325: where $r^{-n}$ is the kinematic scaling and $f_n$ is a function of the
1326: dimensionless running couplings. Eq.~(\ref{vn}) should reduce to the observed
1327: forms in the inertial range
1328: \begin{equation}
1329: \label{inr}
1330: r \ll k_{\text{fr}}^{-1} = \mathcal{E}^{1/2} \rho^{-3/2},
1331: \end{equation}
1332: subject to the condition for a fully developed cascade,
1333: \begin{equation}
1334: \label{gkfr}
1335: \hat g(k_{\text{fr}}) \sim (\mathcal{E}^{-1} \rho^2)^3 \gg 1.
1336: \end{equation}
1337: Conditions (\ref{inr}) and (\ref{gkfr}) are equivalent to
1338: \begin{equation}
1339: \label{cond}
1340: (\rho r^2)^{-2} \ll \hat g(r^{-1}) \ll (\rho r^2)^{-3}.
1341: \end{equation}
1342: 
1343: To achieve complete scale invariance of inertial-range structure functions
1344: (absence of intermittency), we must have
1345: \begin{equation}
1346: \label{fn}
1347: f_n\bm{(}\hat g(r^{-1}), \rho r^2\bm{)} \propto \hat g(r^{-1})^{-n/3}
1348: \end{equation}
1349: for the range of arguments (\ref{cond}); then $f_n \propto \mathcal{E}^{n/3}
1350: r^{4n/3}$, and
1351: \begin{equation}
1352: \langle (\delta v)^n \rangle \propto \mathcal{E}^{n/3} r^{n/3}.
1353: \end{equation}
1354: Eq.~(\ref{fn}) requires a specific nonperturbative behavior of velocity
1355: correlations at strong coupling, mimicking the effect of an anomalous
1356: dimension. RG theory alone does not constrain the functions $f_n$. The
1357: well-established second moment (\ref{vv}) and third moment (\ref{vvv})
1358: strongly suggest that an exact calculation in our theory would yield the
1359: required behavior of $f_2$ and $f_3$. But in the absence of a field-theoretic
1360: reason why Eq.~(\ref{fn}) should persist for $n \ge 4$, high-order structure
1361: functions may generically be expected to violate scale invariance and produce
1362: intermittency. In sum, we would not be at all surprised by observations of
1363: intermittency in the inverse cascade, but a seeming total absence of
1364: intermittency would raise questions about unknown properties of our theory
1365: that enforce effective scale invariance.
1366: 
1367: Some numerical studies of higher moments \cite{SY93,SY94} deal with a
1368: nonstationary inverse cascade and find no signs of intermittency. Such
1369: results are not directly relevant to this paper but have been explained
1370: theoretically \cite{Y99} based on the growth of the inertial range with time.
1371: For the stationary inverse cascade, laboratory experiments \cite{PT98,DBPT01}
1372: reveal no evidence of intermittency. The results of stationary numerical
1373: simulations, however, are mixed: A study using linear friction \cite{BCV00}
1374: confirms the absence of intermittency, but a study using hypofriction with
1375: one inverse Laplacian \cite{BDF95} obtains intermittency that is described as
1376: strong and as similar to that observed in the three-dimensional direct
1377: cascade. Surprisingly, results from the latter simulation are also presented
1378: in a subsequent paper \cite{DBPT01} where the intermittency is described as
1379: insignificant. The evidence on intermittency in the stationary inverse
1380: cascade is unclear, and further numerical studies would be useful to resolve
1381: the question. Results exhibiting strong intermittency are most natural, from
1382: the viewpoint of our field theory.
1383: 
1384: The mild hypofriction used in the numerical study obtaining strong
1385: intermittency \cite{BDF95} is unlikely to alter the qualitative structure of
1386: our theory. While the inverse Laplacian renders the initial Navier--Stokes
1387: equation (\ref{NSv}) nonlocal, the differentiated form (\ref{NSpsi}) and the
1388: field-theory action (\ref{Sfppsi}) are still local in terms of the stream
1389: function $\psi$. The hypofriction term certainly violates Galilean
1390: invariance, but this symmetry holds for zero hypofriction and so its
1391: implications for the RG functions persist in minimal subtraction even when
1392: hypofriction is added. The same arguments used for linear friction in this
1393: paper suggest that intermittency should be expected with hypofriction as
1394: well. It would be interesting to perform a direct numerical study of the
1395: effect of modified friction on intermittency in the stationary inverse
1396: cascade.
1397: 
1398: \section{\label{Burgers}Generalized Burgers equation}
1399: 
1400: \subsection{\label{Continuation}Dimensional continuation}
1401: 
1402: As an example of an alternative statistical fluid model to which our RG
1403: methods can be applied but which exhibits different behavior, let us briefly
1404: consider the one-dimensional Burgers equation \cite{YS88}
1405: \begin{equation}
1406: \label{B1}
1407: \dot v + v\nabla v - \nu\nabla^2 v + \alpha v = f.
1408: \end{equation}
1409: Here $v$ is the velocity field of a fluid without pressure, $f$ is the force
1410: per unit mass, $\nu$ is the kinematic viscosity, and $\alpha$ is the friction
1411: coefficient (not normally included in the Burgers equation but useful in
1412: controlling the long-distance behavior). As with the two-dimensional
1413: incompressible fluid, we assume that the forcing is confined to a band of
1414: high wavenumbers, and we study the response at lower wavenumbers.
1415: 
1416: FNS \cite{FNS77} found that the UV-stirred Burgers equation, like the
1417: Navier--Stokes equation, has a dimensionless coupling in two spatial
1418: dimensions and can be analyzed by an $\epsilon$-expansion in $d = 2 -
1419: \epsilon$. But their continuation of the Burgers equation to $d > 1$ does not
1420: preserve important one-dimensional properties, and their $\epsilon$-expansion
1421: is not fully consistent \cite{FNS77,YS88}. With $\nu = \alpha = f = 0$,
1422: Eq.~(\ref{B1}) yields conservation of the ``energy''
1423: \begin{equation}
1424: E = \tfrac{1}{2} \int dx\, v^2.
1425: \end{equation}
1426: This $E$ is not proportional to the physical energy of a pressure-free fluid,
1427: because it does not account for the varying density; nevertheless, $E$ is
1428: conserved in $d = 1$, and leads to a fluctuation-dissipation theorem (FDT).
1429: It is not easy, however, to generalize Eq.~(\ref{B1}) to $d > 1$ so that a
1430: similar ``energy'' is conserved, while maintaining other key properties such
1431: as Galilean invariance. This is the task we now address.
1432: 
1433: If we neglect dissipation and forcing, the conservation of $E$ in $d = 1$
1434: follows from the relation
1435: \begin{equation}
1436: 0 = v\dot v + \tfrac{1}{3} \nabla(v^3) \equiv v(\dot v + v\nabla v).
1437: \end{equation}
1438: The analogue in $d > 1$ that would yield conservation of
1439: \begin{equation}
1440: E = \tfrac{1}{2} \int d^d x\, |\mathbf{v}|^2
1441: \end{equation}
1442: is
1443: \begin{align}
1444: 0 &= \mathbf{v} \cdot \dot{\mathbf{v}} + A \bm{\nabla} \cdot (v^2
1445: \mathbf{v})\nonumber\\
1446: &\equiv \mathbf{v} \cdot [\dot{\mathbf{v}} + A (\bm{\nabla} \cdot \mathbf{v})
1447: \mathbf{v} + (A - B) \bm{\nabla}(v^2) + 2B (\mathbf{v} \cdot \bm{\nabla})
1448: \mathbf{v}]\nonumber\\
1449: &\equiv \mathbf{v} \cdot \mathbf{w}.
1450: \end{align}
1451: Unfortunately, whatever the choice of the constants $A$ and $B$, the quantity
1452: $\mathbf{w}$ is not Galilean covariant and is not a suitable generalization
1453: of $(\dot v + v\nabla v)$.
1454: 
1455: To remedy this problem, we impose the potential-flow condition
1456: \begin{equation}
1457: \mathbf{v} = \bm{\nabla}\psi,
1458: \end{equation}
1459: as is commonplace when considering a multidimensional Burgers equation. Then,
1460: because
1461: \begin{equation}
1462: \dot E = \int d^d x\, \mathbf{v} \cdot \mathbf{w} = \int d^d x\, \psi
1463: (-\bm{\nabla} \cdot \mathbf{w}),
1464: \end{equation}
1465: a scalar equation of motion can conserve
1466: \begin{equation}
1467: E = \tfrac{1}{2} \int d^d x\, \nabla_i \psi \nabla_i \psi.
1468: \end{equation}
1469: We take
1470: \begin{align}
1471: 0 = -\bm{\nabla} \cdot \mathbf{w} \equiv {}&{-}\nabla^2 \dot\psi - A\nabla^2
1472: \psi \nabla^2 \psi - 2A\nabla_i \nabla_j \psi \nabla_i \nabla_j
1473: \psi\nonumber\\
1474: &- 3A\nabla_i \psi \nabla_i \nabla^2 \psi.
1475: \end{align}
1476: If $A = \tfrac{1}{3}$, this expression is covariant under the Galilean
1477: transformation
1478: \begin{equation}
1479: \psi(t, \mathbf{x}) \to \psi(t, \mathbf{x} + \mathbf{u}t) - \mathbf{u} \cdot
1480: \mathbf{x},
1481: \end{equation}
1482: which induces
1483: \begin{equation}
1484: \mathbf{v}(t, \mathbf{x}) \to \mathbf{v}(t, \mathbf{x} + \mathbf{u}t) -
1485: \mathbf{u}.
1486: \end{equation}
1487: 
1488: Upon restoring dissipation and forcing, we therefore propose
1489: \begin{align}
1490: &{-}\nabla^2 \dot\psi - \nabla_i \psi \nabla_i \nabla^2 \psi - \tfrac{1}{3}
1491: \nabla^2 \psi \nabla^2 \psi\nonumber\\
1492: \label{Bd}
1493: &- \tfrac{2}{3} \nabla_i \nabla_j \psi \nabla_i \nabla_j \psi + \nu\nabla^4
1494: \psi - \alpha\nabla^2 \psi = \eta
1495: \end{align}
1496: as a fully satisfactory multidimensional Burgers equation that reduces to
1497: Eq.~(\ref{B1}) in $d = 1$. Eq.~(\ref{Bd}) is formally very similar to the
1498: incompressible stream-function equation (\ref{NSpsi}), and we will follow our
1499: previous analysis closely.
1500: 
1501: \subsection{\label{OneLoopB}One-loop renormalization}
1502: 
1503: As in Sec.~\ref{Framework}, we assume Gaussian forcing and introduce a path
1504: integral with the action
1505: \begin{alignat}{2}
1506: S = {}&{\int dt}\, d^d x\, \bigl[\tfrac{1}{2} (-\nabla^2 p)\,
1507: D(-\nabla^2)p\span\span\nonumber\\
1508: &+ ip(&&{-}\nabla^2 \dot\psi - \nabla_i \psi \nabla_i \nabla^2 \psi -
1509: \tfrac{1}{3} \nabla^2 \psi \nabla^2 \psi\nonumber\\
1510: &&&- \tfrac{2}{3} \nabla_i \nabla_j \psi \nabla_i \nabla_j \psi + \nu\nabla^4
1511: \psi - \alpha\nabla^2 \psi)\bigr]\nonumber\\
1512: = {}&{\int dt}\, d^d x\, \bigl[\tfrac{1}{2} (-\nabla^2 p)\, D(-\nabla^2)p +
1513: i\nu p\nabla^4 \psi - i\alpha p\nabla^2 \psi\span\span\nonumber\\
1514: &- ip\nabla^2 \dot\psi - i\nabla_i \nabla_j p\, \bigl(\tfrac{1}{6}
1515: \delta_{ij} \nabla_k \psi \nabla_k \psi + \tfrac{1}{3} \nabla_i \psi \nabla_j
1516: \psi\bigr)\bigr],\span\span
1517: \end{alignat}
1518: where we have again integrated the $p\psi\psi$ term by parts. Analysis of
1519: dimensions in $d = 2 - \epsilon$ proceeds as before, since all terms contain
1520: the same numbers of derivatives as for the incompressible fluid. The final
1521: Burgers action, to be compared with Eq.~(\ref{Sfppsi}), is
1522: \begin{align}
1523: S = {}&{\int dt}\, d^d x\, \bigl[\tfrac{1}{2} g^{-1} \nabla^2 p \nabla^2 p +
1524: ig^{-1} p\nabla^4 \psi - i\rho p\nabla^2 \psi\nonumber\\
1525: &- ip\nabla^2 \dot\psi - i\nabla_i \nabla_j p\, \bigl(\tfrac{1}{6}
1526: \delta_{ij} \nabla_k \psi \nabla_k \psi + \tfrac{1}{3} \nabla_i \psi \nabla_j
1527: \psi\bigr)\bigr].
1528: \end{align}
1529: The scaling dimensions are again given by Eq.~(\ref{deps}).
1530: 
1531: The only change to the Feynman rules in Fig.~\ref{FeynRules} is the vertex
1532: factor, which now becomes
1533: \begin{equation}
1534: -\Gamma^{\psi\psi p}_0 = \tfrac{1}{3} i |\mathbf{k}_1 + \mathbf{k}_2|^2
1535: \mathbf{k}_1 \cdot \mathbf{k}_2 + \tfrac{2}{3} i (k_1^2 + \mathbf{k}_1 \cdot
1536: \mathbf{k}_2) (k_2^2 + \mathbf{k}_1 \cdot \mathbf{k}_2).
1537: \end{equation}
1538: Thanks to conservation of the ``energy'' $E$, the FDT is obtained just as in
1539: Sec.~\ref{FDT}, and together with Galilean invariance it implies that the
1540: anomalous dimensions vanish. Thus, to find the new one-loop $\beta$ function,
1541: we need only recompute one of the diagrams in Fig.~\ref{1L1PI}. We calculate
1542: \begin{equation}
1543: \label{1LpsipB}
1544: \Gamma^{\psi p}_1 = igq^4 \left(\frac{1}{32\pi\epsilon} -
1545: \frac{\ln(g\rho/4\pi) + \gamma_{\text{E}} + \frac{8}{9}}{64\pi} +
1546: O(\epsilon)\right),
1547: \end{equation}
1548: to be compared with Eq.~(\ref{1Lpsip}). Curiously, the $O(\epsilon^{-1})$
1549: term is exactly the same and again leads to the one-loop $\beta$ function
1550: \begin{equation}
1551: \label{betaB}
1552: \beta(\bar g) = -\tfrac{1}{2} \epsilon \bar g + \frac{\bar g^3}{32\pi} +
1553: O(\bar g^5).
1554: \end{equation}
1555: But the $O(\epsilon^0)$ term in Eq.~(\ref{1LpsipB}), which would enter a
1556: calculation of the two-loop $\beta$ function, differs from
1557: Eq.~(\ref{1Lpsip}), and we have no reason to expect an identity between the
1558: $\beta$ functions beyond one loop.
1559: 
1560: Our multidimensional Burgers equation (\ref{Bd}) appears not to have been
1561: investigated previously, and it is possible that numerical simulations in two
1562: or three dimensions could reveal interesting and unexpected behavior. The
1563: focus here, however, is on the extrapolation to one dimension ($\epsilon =
1564: 1$), where we seek to explain the known behavior \cite{YS88} of the ordinary
1565: Burgers equation (\ref{B1}). If $\epsilon$ is positive and small, then the
1566: $\beta$ function (\ref{betaB}) has a nontrivial IR-stable fixed point at
1567: \begin{equation}
1568: \bar g_*^2 = 16\pi\epsilon + O(\epsilon^2).
1569: \end{equation}
1570: By contrast, when a simpler continuation of the Burgers equation was used,
1571: the fixed point appeared to be IR-unstable near two dimensions
1572: \cite{FNS77,YS88}. From our results it is natural to expect an IR-stable
1573: strong-coupling fixed point in one dimension ($\epsilon = 1$).
1574: 
1575: \subsection{\label{Comparison}Comparison with inverse-cascade model}
1576: 
1577: Due to the nontrivial fixed point, the response of the one-dimensional
1578: Burgers equation to UV forcing is different from that of the two-dimensional
1579: Navier--Stokes equation---even in the absence of friction. At wavenumbers low
1580: enough that the fixed point is reached, Burgers correlation functions (at
1581: equal or unequal times) exhibit purely kinematic scaling, since all anomalous
1582: dimensions vanish. For example, from Eq.~(\ref{deps}), the scaling of time
1583: and frequency is given by
1584: \begin{equation}
1585: \begin{aligned}
1586: d_t &= -2 + \tfrac{1}{2} \epsilon = -\tfrac{3}{2}, & \omega &\propto k^{3/2}.
1587: \end{aligned}
1588: \end{equation}
1589: Meanwhile, in both models, the FDT implies that the energy spectrum (an
1590: equal-time correlation function) obeys equipartition. These conclusions agree
1591: with previous analytic \cite{FNS77} and numerical \cite{YS88} studies of the
1592: frictionless one-dimensional Burgers equation.
1593: 
1594: We have argued that the inclusion of friction dramatically alters the
1595: behavior of the two-dimensional Navier--Stokes model, from equipartition to a
1596: stationary inverse cascade. This is plausible only in the presence of a
1597: large, rapidly flowing coupling $\hat g$. We noted in Sec.~\ref{Cascade} how
1598: this RG flow could give rise to a constant energy flux and a $k^{-5/3}$
1599: spectrum. No such profound effect is expected for friction in the
1600: one-dimensional Burgers equation. Correlation functions will be modified at
1601: very low wavenumbers $\lesssim \rho^{2/3}$, but as long as $\hat g$ remains
1602: close to the fixed point, higher wavenumbers will retain the equipartition
1603: spectrum.
1604: 
1605: An open question concerns the strong-coupling behavior of our generalized
1606: Burgers equation (\ref{Bd}) in two dimensions with friction. If the $\beta$
1607: function happens to grow asymptotically in the same way as the Navier--Stokes
1608: one, then the arguments of Sec.~\ref{Cascade} can be repeated and it is at
1609: least possible that this Burgers model could exhibit an inverse energy
1610: cascade. If true, this should also be evident if the equation is studied
1611: numerically. It is unclear, however, whether this multidimensional
1612: Galilean-invariant scalar field theory has a direct physical application.
1613: 
1614: \section{\label{Discussion}Discussion}
1615: 
1616: We have presented a statistical field theory capable of describing the
1617: stationary inverse energy cascade in two-dimensional incompressible
1618: turbulence, and computed the RG functions through two loops. By contrast with
1619: previous RG studies of turbulence, we have taken advantage of the nature of
1620: the inverse cascade, with external forcing confined to high wavenumbers, to
1621: work with a conventional local field theory. The consequences of Galilean
1622: invariance and the fluctuation-dissipation theorem have been systematically
1623: derived based on the underlying symmetries of the action. After taking these
1624: symmetries into account, we have evaluated the necessary two-loop diagrams in
1625: dimensional regularization to obtain the two-loop $\beta$ function
1626: \begin{equation}
1627: \beta(\bar g) = \frac{\bar g^3}{32\pi} - \frac{\bar g^5}{2048\pi^2} + O(\bar
1628: g^7),
1629: \end{equation}
1630: which is independent of the renormalization scheme to precisely this order.
1631: 
1632: Because the anomalous dimensions vanish identically in minimal subtraction,
1633: no RG fixed point can yield the observed $k^{-5/3}$ energy spectrum of the
1634: inverse cascade. Instead, we have found that the inverse cascade could
1635: plausibly arise from nonperturbative strong-coupling effects in the presence
1636: of friction. The apparent anomalous dimension of the velocity must arise from
1637: the rapid RG flow of the coupling. Cutoff independence of the energy
1638: dissipation rate (or constancy of the energy flux) requires the
1639: strong-coupling behavior
1640: \begin{equation}
1641: \begin{aligned}
1642: \beta(\hat g) &\simeq 4\hat g & (\hat g &\to \infty),
1643: \end{aligned}
1644: \end{equation}
1645: and we have also obtained this in a heuristic way from a Borel transformation
1646: of the two-loop $\beta$ function. The observed energy spectrum in the
1647: inertial range has been matched with perturbative results at the wavenumbers
1648: where dissipation becomes important. Inertial-range intermittency (violation
1649: of scale invariance) is generically expected because the coupling flows
1650: rapidly with scale, but the evidence on intermittency from numerical
1651: simulations is mixed. On the other hand, a similar RG analysis of the
1652: one-dimensional Burgers equation confirms the simpler behavior of that model,
1653: controlled by a strong-coupling fixed point.
1654: 
1655: Our greatest difficulty is that the inverse cascade appears to be
1656: intrinsically a nonperturbative phenomenon. To make quantitative predictions,
1657: such as the exponent of the energy spectrum or the value of the Kolmogorov
1658: constant, it may be useful to combine a high-order perturbative calculation
1659: with an appropriate resummation method, as in our analysis of the
1660: strong-coupling $\beta$ function. Of course evaluating additional diagrams
1661: with two or more loops will be very challenging, especially if the finite
1662: parts are needed. At the present stage, the most intriguing prediction of our
1663: theory is that substantial intermittency should be expected in the stationary
1664: inverse cascade. It would be helpful to have more robust and consistent
1665: results from experiments and simulations to determine whether this
1666: expectation is realized, and if it is not, to identify the field-theoretic
1667: explanation.
1668: 
1669: \begin{acknowledgments}
1670: 
1671: I thank A. M. Polyakov, S. L. Sondhi, H. L. Verlinde, and V. Yakhot for
1672: helpful discussions.
1673: 
1674: This material is based upon work supported by the National Science Foundation
1675: under Grant No.\ 0243680. Any opinions, findings, and conclusions or
1676: recommendations expressed in this material are those of the author and do not
1677: necessarily reflect the views of the National Science Foundation.
1678: 
1679: \end{acknowledgments}
1680: 
1681: \begin{thebibliography}{34}
1682: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1683: \expandafter\ifx\csname bibnamefont\endcsname\relax
1684:   \def\bibnamefont#1{#1}\fi
1685: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1686:   \def\bibfnamefont#1{#1}\fi
1687: \expandafter\ifx\csname citenamefont\endcsname\relax
1688:   \def\citenamefont#1{#1}\fi
1689: \expandafter\ifx\csname url\endcsname\relax
1690:   \def\url#1{\texttt{#1}}\fi
1691: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1692: \providecommand{\bibinfo}[2]{#2}
1693: \providecommand{\eprint}[2][]{\url{#2}}
1694: 
1695: \bibitem[{\citenamefont{Kraichnan}(1967)}]{K67}
1696: \bibinfo{author}{\bibfnamefont{R.~H.} \bibnamefont{Kraichnan}},
1697:   \bibinfo{journal}{Phys. Fluids} \textbf{\bibinfo{volume}{10}},
1698:   \bibinfo{pages}{1417} (\bibinfo{year}{1967}).
1699: 
1700: \bibitem[{\citenamefont{Zinn-Justin}(1996)}]{Z96}
1701: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Zinn-Justin}},
1702:   \emph{\bibinfo{title}{Quantum Field Theory and Critical Phenomena}}
1703:   (\bibinfo{publisher}{Oxford University Press}, \bibinfo{address}{Oxford},
1704:   \bibinfo{year}{1996}), \bibinfo{edition}{3rd} ed.
1705: 
1706: \bibitem[{\citenamefont{Frisch}(1995)}]{F95}
1707: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Frisch}},
1708:   \emph{\bibinfo{title}{Turbulence: The Legacy of A. N. Kolmogorov}}
1709:   (\bibinfo{publisher}{Cambridge University Press},
1710:   \bibinfo{address}{Cambridge}, \bibinfo{year}{1995}).
1711: 
1712: \bibitem[{\citenamefont{Smith and Yakhot}(1993)}]{SY93}
1713: \bibinfo{author}{\bibfnamefont{L.~M.} \bibnamefont{Smith}} \bibnamefont{and}
1714:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Yakhot}},
1715:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{71}},
1716:   \bibinfo{pages}{352} (\bibinfo{year}{1993}).
1717: 
1718: \bibitem[{\citenamefont{Smith and Yakhot}(1994)}]{SY94}
1719: \bibinfo{author}{\bibfnamefont{L.~M.} \bibnamefont{Smith}} \bibnamefont{and}
1720:   \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Yakhot}},
1721:   \bibinfo{journal}{J. Fluid Mech.} \textbf{\bibinfo{volume}{274}},
1722:   \bibinfo{pages}{115} (\bibinfo{year}{1994}).
1723: 
1724: \bibitem[{\citenamefont{Yakhot}(1999)}]{Y99}
1725: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Yakhot}},
1726:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{60}},
1727:   \bibinfo{pages}{5544} (\bibinfo{year}{1999}).
1728: 
1729: \bibitem[{\citenamefont{Babiano et~al.}(1995)\citenamefont{Babiano, Dubrulle,
1730:   and Frick}}]{BDF95}
1731: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Babiano}},
1732:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Dubrulle}},
1733: \bibnamefont{and}
1734:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Frick}},
1735:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{52}},
1736:   \bibinfo{pages}{3719} (\bibinfo{year}{1995}).
1737: 
1738: \bibitem[{\citenamefont{Boffetta et~al.}(2000)\citenamefont{Boffetta, Celani,
1739:   and Vergassola}}]{BCV00}
1740: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Boffetta}},
1741:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Celani}}, \bibnamefont{and}
1742:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Vergassola}},
1743:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{61}},
1744:   \bibinfo{pages}{R29} (\bibinfo{year}{2000}).
1745: 
1746: \bibitem[{\citenamefont{Paret and Tabeling}(1998)}]{PT98}
1747: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Paret}} \bibnamefont{and}
1748:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tabeling}},
1749:   \bibinfo{journal}{Phys. Fluids} \textbf{\bibinfo{volume}{10}},
1750:   \bibinfo{pages}{3126} (\bibinfo{year}{1998}).
1751: 
1752: \bibitem[{\citenamefont{Honkonen}(1998)}]{H98}
1753: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Honkonen}},
1754:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{58}},
1755:   \bibinfo{pages}{4532} (\bibinfo{year}{1998}).
1756: 
1757: \bibitem[{\citenamefont{Polyakov}(1993)}]{P93}
1758: \bibinfo{author}{\bibfnamefont{A.~M.} \bibnamefont{Polyakov}},
1759:   \bibinfo{journal}{Nucl. Phys. B} \textbf{\bibinfo{volume}{396}},
1760:   \bibinfo{pages}{367} (\bibinfo{year}{1993}).
1761: 
1762: \bibitem[{\citenamefont{Forster et~al.}(1977)\citenamefont{Forster, Nelson,
1763: and
1764:   Stephen}}]{FNS77}
1765: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Forster}},
1766:   \bibinfo{author}{\bibfnamefont{D.~R.} \bibnamefont{Nelson}},
1767:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~J.}
1768:   \bibnamefont{Stephen}}, \bibinfo{journal}{Phys. Rev. A}
1769:   \textbf{\bibinfo{volume}{16}}, \bibinfo{pages}{732} (\bibinfo{year}{1977}).
1770: 
1771: \bibitem[{\citenamefont{DeDominicis and Martin}(1979)}]{DM79}
1772: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{DeDominicis}}
1773: \bibnamefont{and}
1774:   \bibinfo{author}{\bibfnamefont{P.~C.} \bibnamefont{Martin}},
1775:   \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{19}},
1776:   \bibinfo{pages}{419} (\bibinfo{year}{1979}).
1777: 
1778: \bibitem[{\citenamefont{Adzhemyan et~al.}(1999)\citenamefont{Adzhemyan,
1779:   Antonov, and Vasiliev}}]{AAV99}
1780: \bibinfo{author}{\bibfnamefont{L.~Ts.} \bibnamefont{Adzhemyan}},
1781:   \bibinfo{author}{\bibfnamefont{N.~V.} \bibnamefont{Antonov}},
1782:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.~N.}
1783:   \bibnamefont{Vasiliev}}, \emph{\bibinfo{title}{The Field Theoretic
1784:   Renormalization Group in Fully Developed Turbulence}}
1785:   (\bibinfo{publisher}{Gordon and Breach}, \bibinfo{address}{Amsterdam},
1786:   \bibinfo{year}{1999}).
1787: 
1788: \bibitem[{\citenamefont{Lesieur}(1997)}]{L97}
1789: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Lesieur}},
1790:   \emph{\bibinfo{title}{Turbulence in Fluids}} (\bibinfo{publisher}{Kluwer},
1791:   \bibinfo{address}{Boston}, \bibinfo{year}{1997}), \bibinfo{edition}{3rd}
1792: ed.
1793: 
1794: \bibitem[{\citenamefont{'t~Hooft and Veltman}(1972)}]{TV72}
1795: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{'t~Hooft}} \bibnamefont{and}
1796:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Veltman}},
1797:   \bibinfo{journal}{Nucl. Phys. B} \textbf{\bibinfo{volume}{44}},
1798:   \bibinfo{pages}{189} (\bibinfo{year}{1972}).
1799: 
1800: \bibitem[{\citenamefont{Martin et~al.}(1973)\citenamefont{Martin, Siggia, and
1801:   Rose}}]{MSR73}
1802: \bibinfo{author}{\bibfnamefont{P.~C.} \bibnamefont{Martin}},
1803:   \bibinfo{author}{\bibfnamefont{E.~D.} \bibnamefont{Siggia}},
1804:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.~A.}
1805: \bibnamefont{Rose}},
1806:   \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{8}},
1807:   \bibinfo{pages}{423} (\bibinfo{year}{1973}).
1808: 
1809: \bibitem[{\citenamefont{Bausch et~al.}(1976)\citenamefont{Bausch, Janssen,
1810: and
1811:   Wagner}}]{BJW76}
1812: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bausch}},
1813:   \bibinfo{author}{\bibfnamefont{H.~K.} \bibnamefont{Janssen}},
1814:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Wagner}},
1815:   \bibinfo{journal}{Z. Phys. B} \textbf{\bibinfo{volume}{24}},
1816:   \bibinfo{pages}{113} (\bibinfo{year}{1976}).
1817: 
1818: \bibitem[{\citenamefont{Phythian}(1977)}]{P77}
1819: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Phythian}},
1820:   \bibinfo{journal}{J. Phys. A} \textbf{\bibinfo{volume}{10}},
1821:   \bibinfo{pages}{777} (\bibinfo{year}{1977}).
1822: 
1823: \bibitem[{\citenamefont{Adzhemyan et~al.}(1983)\citenamefont{Adzhemyan,
1824:   Vasil'ev, and Pis'mak}}]{AVP83}
1825: \bibinfo{author}{\bibfnamefont{L.~D.} \bibnamefont{Adzhemyan}},
1826:   \bibinfo{author}{\bibfnamefont{A.~N.} \bibnamefont{Vasil'ev}},
1827:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{Yu.~M.}
1828:   \bibnamefont{Pis'mak}}, \bibinfo{journal}{Theor. Math. Phys. (USSR)}
1829:   \textbf{\bibinfo{volume}{57}}, \bibinfo{pages}{1131}
1830: (\bibinfo{year}{1983}).
1831: 
1832: \bibitem[{\citenamefont{'t~Hooft}(1973)}]{T73}
1833: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{'t~Hooft}},
1834:   \bibinfo{journal}{Nucl. Phys. B} \textbf{\bibinfo{volume}{61}},
1835:   \bibinfo{pages}{455} (\bibinfo{year}{1973}).
1836: 
1837: \bibitem[{\citenamefont{Deker and Haake}(1975)}]{DH75}
1838: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Deker}} \bibnamefont{and}
1839:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Haake}},
1840:   \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{11}},
1841:   \bibinfo{pages}{2043} (\bibinfo{year}{1975}).
1842: 
1843: \bibitem[{\citenamefont{Frey and T{\"a}uber}(1994)}]{FT94}
1844: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Frey}} \bibnamefont{and}
1845:   \bibinfo{author}{\bibfnamefont{U.~C.} \bibnamefont{T{\"a}uber}},
1846:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{50}},
1847:   \bibinfo{pages}{1024} (\bibinfo{year}{1994}).
1848: 
1849: \bibitem[{\citenamefont{Adzhemyan et~al.}(2003)\citenamefont{Adzhemyan,
1850:   Antonov, Kompaniets, and Vasil'ev}}]{AAKV03}
1851: \bibinfo{author}{\bibfnamefont{L.~Ts.} \bibnamefont{Adzhemyan}},
1852:   \bibinfo{author}{\bibfnamefont{N.~V.} \bibnamefont{Antonov}},
1853:   \bibinfo{author}{\bibfnamefont{M.~V.} \bibnamefont{Kompaniets}},
1854:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.~N.}
1855:   \bibnamefont{Vasil'ev}}, \bibinfo{journal}{Int. J. Mod. Phys. B}
1856:   \textbf{\bibinfo{volume}{17}}, \bibinfo{pages}{2137}
1857: (\bibinfo{year}{2003}).
1858: 
1859: \bibitem[{\citenamefont{Wolfram}(1999)}]{W99}
1860: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Wolfram}},
1861:   \emph{\bibinfo{title}{The Mathematica Book}} (\bibinfo{publisher}{Cambridge
1862:   University Press}, \bibinfo{address}{Cambridge}, \bibinfo{year}{1999}),
1863:   \bibinfo{edition}{4th} ed.
1864: 
1865: \bibitem[{\citenamefont{Press et~al.}(1992)\citenamefont{Press, Teukolsky,
1866:   Vetterling, and Flannery}}]{PTVF92}
1867: \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Press}},
1868:   \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Teukolsky}},
1869:   \bibinfo{author}{\bibfnamefont{W.~T.} \bibnamefont{Vetterling}},
1870:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~P.}
1871:   \bibnamefont{Flannery}}, \emph{\bibinfo{title}{Numerical Recipes in C: The
1872:   Art of Scientific Computing}} (\bibinfo{publisher}{Cambridge University
1873:   Press}, \bibinfo{address}{Cambridge}, \bibinfo{year}{1992}),
1874:   \bibinfo{type}{section} \bibinfo{chapter}{7.8}, \bibinfo{edition}{2nd} ed.
1875: 
1876: \bibitem[{\citenamefont{Frisch and Sulem}(1984)}]{FS84}
1877: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Frisch}} \bibnamefont{and}
1878:   \bibinfo{author}{\bibfnamefont{P.~L.} \bibnamefont{Sulem}},
1879:   \bibinfo{journal}{Phys. Fluids} \textbf{\bibinfo{volume}{27}},
1880:   \bibinfo{pages}{1921} (\bibinfo{year}{1984}).
1881: 
1882: \bibitem[{\citenamefont{Danilov and Gurarie}(2001)}]{DG01}
1883: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Danilov}} \bibnamefont{and}
1884:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Gurarie}},
1885:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{63}},
1886:   \bibinfo{pages}{020203(R)} (\bibinfo{year}{2001}).
1887: 
1888: \bibitem[{\citenamefont{Sommeria}(1986)}]{S86}
1889: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Sommeria}},
1890:   \bibinfo{journal}{J. Fluid Mech.} \textbf{\bibinfo{volume}{170}},
1891:   \bibinfo{pages}{139} (\bibinfo{year}{1986}).
1892: 
1893: \bibitem[{\citenamefont{Maltrud and Vallis}(1991)}]{MV91}
1894: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Maltrud}}
1895: \bibnamefont{and}
1896:   \bibinfo{author}{\bibfnamefont{G.~K.} \bibnamefont{Vallis}},
1897:   \bibinfo{journal}{J. Fluid Mech.} \textbf{\bibinfo{volume}{228}},
1898:   \bibinfo{pages}{321} (\bibinfo{year}{1991}).
1899: 
1900: \bibitem[{\citenamefont{Borue}(1994)}]{B94}
1901: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Borue}},
1902:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{72}},
1903:   \bibinfo{pages}{1475} (\bibinfo{year}{1994}).
1904: 
1905: \bibitem[{\citenamefont{Bernard}(1999)}]{B99}
1906: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bernard}},
1907:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{60}},
1908:   \bibinfo{pages}{6184} (\bibinfo{year}{1999}).
1909: 
1910: \bibitem[{\citenamefont{Dubos et~al.}(2001)\citenamefont{Dubos, Babiano,
1911: Paret,
1912:   and Tabeling}}]{DBPT01}
1913: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Dubos}},
1914:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Babiano}},
1915:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Paret}}, \bibnamefont{and}
1916:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Tabeling}},
1917:   \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{64}},
1918:   \bibinfo{pages}{036302} (\bibinfo{year}{2001}).
1919: 
1920: \bibitem[{\citenamefont{Yakhot and She}(1988)}]{YS88}
1921: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Yakhot}} \bibnamefont{and}
1922:   \bibinfo{author}{\bibfnamefont{Z.-S.} \bibnamefont{She}},
1923:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{60}},
1924:   \bibinfo{pages}{1840} (\bibinfo{year}{1988}).
1925: 
1926: \end{thebibliography}
1927: 
1928: \end{document}
1929: