Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Journal of the Physical Society of Japan Vol. 70, No. 2, February, 2001, pp. 376-386 Infrared Dynamics of Decaying Two-Dimensional Turbulence Governed by the Charney-Hasegawa-Mima Equation Takahiro Iwayama1,∗ , Takeshi Watanabe2,∗∗ and Theodore G. Shepherd1,∗∗∗ 1 Department 2 Division of Physics, University of Toronto, Toronto, Ontario M5S 1A7, Canada of Global Development Science, Graduate School of Science and Technology, Kobe University, Kobe 657-8501 (Received July 27, 2000) The low wave number range of decaying turbulence governed by the Charney-Hasegawa-Mima (CHM) equation is examined theoretically and by direct numerical simulation. Here, the low wave number range is defined as values of the wave number k below the wave number kE corresponding to the peak of the energy spectrum, or alternatively the centroid wave number of the energy spectrum. The energy spectrum in the low wave number range in the infrared regime (k → 0) is theoretically derived to be E(k) ∼ k5 , using a quasinormal Markovianized model of the CHM equation. This result is verified by direct numerical simulation of the CHM equation. The wave number triads (k, p, q) responsible for the formation of the low wave number spectrum are also examined. It is found that the energy flux Π(k) for k < kE can be entirely expressed by Π(−) (k), which is the total net input of energy to wave numbers < k arising from interactions with wave numbers p, q > k. Furthermore, the contribution of nonlocal triad interactions to the energy flux is found to be predominant in the range log(k/kE ) < −0.5, where the nonlocal interactions are defined to be those triad interactions for which the ratio of the largest leg of the triad to the smallest leg is larger than four. KEYWORDS: energy spectrum, infrared range, low wave number range, nonlocal interaction, two-dimensional turbulence, Charney-Hasegawa-Mima equation §1. case. Equation (1.1) contains two characteristic regimes. The first is the 2D Navier-Stokes (NS) regime that is obtained when λ → 0. The governing equation in this regime is the well-known 2D vorticity equation, Introduction The Charney-Hasegawa-Mima (CHM) equation, a two-dimensional (2D) turbulent system, has been actively studied both theoretically and numerically over the past decade.1-5) The equation describes the temporal evolution of quasi-2D fluctuations of the electrostatic field on the plane perpendicular to a strong magnetic field uniformly applied to a plasma.6) It also describes the temporal evolution of geostrophic motion in geophysical fluids and is called the quasi-geostrophic potential vorticity equation.7) The CHM equation in the strong turbulent state neglecting the effects of waves can be written as follows: ¢ ∂ ¡ 2 ∇ φ − λ2 φ + J(φ, ∇2 φ) = ν∇4 φ, (1.1) ∂t where all quantities are made nondimensional and J(•, •) is the Jacobian operator, ∇2 the 2D Laplacian, and φ(x, y) the electrostatic potential for the plasma case or the variable part of the free surface of the fluid for the geophysical case. The damping coefficient ν is the reciprocal of the Reynolds number. The constant λ is either the ratio of the horizontal length scale of interest L to the ion Larmor radius in the plasma case, or the ratio of L to the Rossby deformation radius in the geophysical ∂∇2 φ + J(φ, ∇2 φ) = ν∇4 φ. (1.2) ∂t The other regime is obtained asymptotically for λ → ∞. The governing equation is ∂φ + J(∇2 φ, φ) = −ν∇4 φ, ∂T (1.3) where T = t/λ2 is a rescaled time. Since eq. (1.3) has been called the asymptotic model (AM),1) we shall call this regime the AM regime. Although the AM regime is characteristic of the CHM equation, there are resemblances between turbulent solutions of (1.2) and those of (1.3). It is well known that the 2D vorticity equation has two quadratic inviscid invariants, the kinetic R −1 2 = (2A) (∇φ) dxdy = energy per unit area, E K A R∞ E (k)dk, and the enstrophy per unit area, Zr = K 0 R∞ R (2A)−1 A (∇2 φ)2 dxdy = 0 k2 EK (k)dk, where A is the area in which the field φ is determined and EK (k) is the kinetic energy spectrum. The existence of two quadratic inviscid invariants causes the dual cascade; i.e., the kinetic energy is transported to the small wave number side (inverse cascade) and the enstrophy is transported to the large wave number side (direct cascade). This fact yields two types of energy spectra, EK (k) ∼ k−5/3 in the energy inertial range and EK (k) ∼ k−3 in the ∗ On leave from Division of Global Development Science, Graduate School of Science and Technology, Kobe University, Kobe 6578501. E-mail: [email protected] ∗∗ E-mail: [email protected] ∗∗∗ E-mail: [email protected] 376 2001) Infrared Dynamics of Decaying Two-Dimensional Turbulence Governed. . . . enstrophy inertial range.8) Similar to the 2D vorticity equation, eq. (1.1) also has two quadratic inviscid invariants: the R ∞E = EK + Eλ = R total energy per unit area, (2A)−1 A [(∇φ)2 + λ2 φ2 ] dx dy = 0 E(k)dk and the potential = Zr + Zλ = R enstrophy per unit area, ZR ∞ (2A)−1 A [(∇2 φ)2 + λ2 (∇φ)2 ] dx dy = 0 k2 E(k)dk. In the AM regime, one thus obtains the invariants Eλ and Zλ . Moreover, the energy spectra E(k) ∼ k−11/3 in the energy inertial range and E(k) ∼ k−5 in the enstrophy inertial range are also derived by dimensional arguments similar to those for 2D NS turbulence.1, 2, 9) Recently, Watanabe et al.5) derived scaling laws for the temporal evolution of a characteristic wave number R k̄ ≡ E −1 k E(k) dk and of the energy spectral density at the scale k̄ for decaying CHM turbulence in the AM regime. Their discussion is similar to Batchelor’s analysis for decaying 2D NS turbulence.10) That is, provided that the energy spectrum evolves in a self-similar way and the energy E is an invariant in the limit of high Reynolds number (ν → 0), the energy spectrum can be expressed as ³ ´ E(k) = E 9/8 t1/4 h k E 1/8 t1/4 , (1.4) where h(x) is a function of universal form.11) This argument yields a temporal evolution of the characteristic wave number, k̄ ∼ t−1/4 , and of the energy spectral density at the scale k̄, ¡ ¢ E k̄ ∼ t1/4 . (1.5) (1.6) The same type of similarity spectrum (1.4) was also derived by Yanase and Yamada,12) but it was not confirmed by numerical simulations of the modified zero-fourth cumulant approximated equation of the CHM equation due to the insufficient calculation time. On the other hand, direct numerical simulation of decaying turbulence in (1.1) confirmed the scaling laws (1.5) and (1.6).5) Moreover, the self-similarity assumption was well satisfied. From the numerical simulation, the asymptotic form of h(x) was obtained as follows: h(x) ∼ x−6 for the high wave number range x > 1 and h(x) ∼ x4 for the low wave number range x < 1, where x ≡ k/k̄. However, theoretical explanation of the asymptotic form of h(x) was left as an unsolved problem. In this study we thus consider the asymptotic form of h(x). The asymptotic form of h(x) for x > 1 is similar to the spectrum in the enstrophy inertial range, h(x) ∼ x−5 , but slightly steeper. Although the reason why the spectrum in this range deviates from the prediction by the dimensional arguments remains unknown, we focus our attention on the asymptotic form of h(x) in the low wave number range, i.e., h(x) for x < 1. For theoretical analysis of h(x) in the low wave number range, we follow the study of Basdevant et al. (henceforth BLS78).13) They derived the energy spectrum EK (k) ∼ k3 in the limit k → 0 (infrared range) of 2D NS turbulence by using a quasinormal Markovianized closure theory. Existence of the spectrum EK (k) ∼ k3 was noted in direct numerical simulations by Frisch and Sulem14) and Chasnov.15) On the other hand, Yanase 377 and Yamada12) derived the energy spectrum in the infrared range of CHM turbulence, EK (k) ∼ k7 , but their result was not numerically confirmed in their paper as mentioned above. Indeed, there are no numerical studies focusing on the low wave number spectrum of CHM turbulence. In this paper, we derive the power-law exponent of the energy spectrum in the infrared range of CHM turbulence theoretically by using the method proposed by BLS78, and confirm our results by performing direct numerical simulations of (1.1). Moreover, the energy flux and the triad interactions in wave number space responsible for the formation of the energy spectrum in the low wave number range are also examined. The derivation of the infrared energy spectrum proposed by BSL78 relies on nonlocality of the triad interactions.13) Nonlocality of the triad interactions in the enstrophy inertial range of 2D NS turbulence is well known16, 17) and has been studied numerically by many researchers.18-21) However, there are no studies on the triad interactions in the low wave number range even for 2D NS turbulence. In this article, we focus only on the triad interactions responsible for the formation of the energy spectrum in the low wave number range of CHM turbulence. However, this study may contribute to the study of the low wave number range of 2D NS turbulence because the two systems have similar triad interaction terms. This paper is organized as follows. In §2, we introduce a quasinormal Markovianized model of the CHM equation. In §3, we derive the power-law exponent of the energy spectrum in the infrared range of decaying CHM turbulence using the method proposed by BSL78. In §4, we present results from direct numerical simulations of the CHM equation, including analysis of the energy transfer function and the triad interactions responsible for the formation of the energy spectrum in the low wave number range. Finally, we summarize the results in §5. §2. Formulation 2.1 Spectral form and energy equation of the CHM equation We consider a system which is confined within the square domain [0, L]2 and adopt doubly periodic boundary conditions. Then the field φ is expanded as X φ̂(k) exp(ik · r), (2.1) φ(r) = k and eq. (1.1) is rewritten as ¶ µ νk4 ∂ + φ̂(k) ∂t k2 + λ2 = ∆ X 1 q 2 − p2 (p × q)z 2 φ̂(p) φ̂(q), (2.2a) 2 k + λ2 k =p + q ∆ X k =p + q ≡ X δ k , p +q . (2.2b) p, q Here, k = 2πn/L is the wave vector and summation is taken over the integer vector n = (nx , ny ), k = |k|, p = |p|, q = |q|. The time argument is omitted for brevity. 378 Takahiro Iwayama, Takeshi Watanabe and Theodore G. Shepherd The temporal evolution of the energy spectrum, which is defined as X E(k), (2.3a) E= k E(k) = X0 1 2 k0 is governed by µ D E (k02 + λ2 ) |φ̂(k0 )|2 , 2νk4 ∂ + 2 ∂t k + λ2 ¶ E(k) = T (k), (2.4) P0 0 where k0 is the shell summation in k − ∆k 2 ≤ |k | < ∆k k + 2 , ∆k = 2π/L, T (k) the energy transfer function, and the angle bracket denotes the ensemble average. The energy transfer function T (k) can be expressed in terms of the triad energy transfer function as X1 T (k, p, q), (2.5a) T (k) = 2 p,q X0 T (k, p, q) = 1 0 (p × q 0 )z (q 02 − p02 ) δk0 , p0 +q 0 2 k 0 , p0 , q 0 0 0 0 Π(k) = Π (k) − Π (−) (k), (2.7a) X T (k0 , p, q). (2.7c) p>k k<q<p 2.2 A quasinormal Markovianized model of the CHM equation In the limit L → ∞, eqs. (2.3a), (2.5a), (2.6), (2.7b), and (2.7c) are represented in terms of integrals, i.e., Z ∞ E(k)dk, (2.8) E= 0 1 2 Z Z ∞Z ∞ T (k, p, q) dp dq, 0 ∞ Π(k) = k Z Π(+) (k) = ∞ k Z 0 can be divided into two parts,16) (2.7b) Π(+) (k) is the net energy input into all wave numbers > k from interactions with p and q both < k, while Π(−) (k) is the net energy input into all wave numbers < k from interactions with p and q both > k. If we set λ → 0, eq. (2.4) reduces to the evolution equation for the energy spectrum of the 2D NS equation. However, the energy transfer function (2.5) is always equivalent to that for the 2D NS equation, because (2.5) is independent of λ. Π(−) (k) = k0 >k T (k0 , p, q), p<k q<p X X Π(−) (k) = (2.5b) Since T (k, p, q) satisfies a detailed balance for wave numbers which form the triangle k = p+q, and is symmetric with respect to p and q, the energy flux X T (k0 ) (2.6) Π(k) = (+) k0 >k T (k) = ×[hφ̂(−k )φ̂(p )φ̂(q )i +hφ̂(k0 )φ̂(−p0 )φ̂(−q 0 )i]. X X X Π(+) (k) = k0 <k (2.3b) (Vol. 70, k (2.9) 0 T (k0 ) dk0 , Z dk0 Z dk0 (2.10) Z k p dp 0 dq T (k0 , p, q), (2.11) 0 Z ∞ p dp k dq T (k0 , p, q). (2.12) k Using a quasinormal Markovianized approximation,22) eq. (2.5b) is reduced to T (k, p, q) = k 2 p2 q 2 2k2 θkpq 2 [2a2 (k, p, q) k E(p) E(q)−b2 (k, p, q) p E(q) E(k)−b2 (k, q, p) q E(k) E(p)] , πpq (k + λ2 )(p2 + λ2 )(q 2 + λ2 ) (2.13) where T (k, p, q) = 0 outside of the domain in the (p, q) plane such that k, p, q can be the sides of the triangle k = p + q (Fig. 1), b2 (k, p, q) + b2 (k, q, p) , 2 (z 2 − x2 )(z 2 − y 2 ) , b2 (k, p, q) = 2 (1 − x2 )3/2 a2 (k, p, q) = b2 (k, q, p) = 2 (y 2 − x2 )(y 2 − z 2 ) , (1 − x2 )3/2 (2.14a) (2.14b) (2.14c) are the geometrical coefficients, and x, y, z refer to the cosines of the interior angles of the triangle facing respectively the sides k, p, q. The function θkpq is the relaxation time of the third order moments associated with the triad (k, p, q), the functional form of which is different for different theories. If θkpq is chosen as θkpq = 1 − exp(−νkpq t) , νkpq (2.15a) νk4 , (2.15b) + λ2 that is equivalent to the fourth-order cumulant being zero, and eq. (2.4) is reduced to the modified zero-fourth cumulant approximated equation derived by Yanase and Yamada.12) It is well known that this approximation includes only viscous damping and cannot generally lead to the inertial energy spectra that are derived by dimensional arguments, although the enstrophy inertial range spectrum of 2D NS turbulence, EK (k) ∝ k−3 , was derived by numerical simulations of the modified zero-fourth cumulant approximated equation.23) On the other hand, there is an approximation that introduces eddy damping effects µkpq in θkpq , referred to as the Eddy-Damped Quasi-Normal Markovianized approximation (E.D.Q.N.M.).24) In this approximation, θkpq can be expressed as follows: νkpq = νk + νp + νq , νk = k2 2001) Infrared Dynamics of Decaying Two-Dimensional Turbulence Governed. . . . q= p- k q= p+ k q k 0 k p Fig. 1. Domain in the (p, q) plane such that k, p, q can be the sides of the triangle k = p + q. The triad transfer function T (k, p, q) has a non zero value within the hatched region. θkpq = 1 − exp[−(νkpq + µkpq )t] , νkpq + µkpq µkpq = µk + µp + µq , £ ¤1/2 . µk ∼ λ−6 k9 E(k)g(k/λ)−3 (2.16a) (2.16b) (2.16c) Equation (2.16c) is derived by eliminating the energy transfer rate ² or the enstrophy transfer rate η from eqs. (9) and (10) in Watanabe et al.4) Here, g(x) is a dimensionless function, which is finite for x ¿ 1 and approaches g(x) = x2 for x À 1. Therefore µk ∼ ¤1/2 £ 9 in the AM regime. The proportionality conk E(k) stant in eq. (2.16c) has not yet been determined. However, as will be seen in the next section, the derivation of the energy spectrum in the infrared range proposed by BSL78 is independent of the functional form of θkpq . Thus we do not mention the functional form of θkpq hereafter. In the AM regime, eq. (2.13) reduces to T (k, p, q) = 2k4 pq θkpq [2a2 (k, p, q) k E(p) E(q) πλ6 −b2 (k, p, q) p E(q) E(k) −b2 (k, q, p) q E(k) E(p)] . (2.17) Equations (2.3b), (2.5)–(2.7) are used in the analysis of the direct numerical simulations of (1.1), while eqs. (2.9)–(2.12), (2.14), and (2.17) are used in the theoretical derivation of the energy spectrum in the infrared range of CHM turbulence. §3. Derivation of the Energy Spectrum in the Infrared Range of Decaying CHM Turbulence In this section, we derive the energy spectrum in the infrared range of decaying CHM turbulence using the method proposed by BSL78.13) In their method, the wave number dependence of the energy transfer function T (k) reflects directly on the functional form of the energy spectrum in the infrared range. This feature is in 379 contrast to the energy spectrum in the energy and enstrophy inertial ranges, where Kolmogorov-type dimensional analysis assumes the energy and enstrophy fluxes to be constant with respect to wave number.8) The method proposed by BSL7813) is summarized as follows:22) 1. Interactions among triad wave numbers k, p, q can be divided into two categories: either all of k, p, q have comparable magnitude, or one is very small in comparison with the others. The former refers to the local interactions and the latter to the nonlocal interactions. One may then divide the energy transfer function T (k) into two parts: TL (k) constructed from the local triad interactions, and TN L (k) from the nonlocal triad interactions. 2. Assume that the nonlocal triad interactions with elongated triads k ¿ p ' q are dominant in the infrared range. Then, one introduces the small parameter ² which is the ratio of two interacting wave numbers, ² ≡ k/p, and expresses TN L (k) in terms of power series in ². 3. Practical calculation is performed through the energy flux Π(k). Instead of calculating the energy transfer function TN L (k) directly, one can evaluate (−) the nonlocal part of the energy flux −ΠN L (k), which is equivalent to ΠN L (k) because the elongated triads k ¿ p ' q cannot contribute to the flux Π(+) (k). (+) To say this another way, ΠN L (k) is geometrically forbidden. Differentiating the resulting energy flux with respect to wave number, one recovers the nonlocal part of the energy transfer function. Then, the wave number dependence of the energy transfer function gives the functional form of the energy spectrum in the infrared range. Note also that generally the characteristic time θkpq is not expanded. Following the above method, we derive the energy spectrum in the infrared range of decaying turbulence of the CHM equation. The nonlocal part of Π(−) (k) is given by Z k Z p Z ∞ (−) 0 dk dp dq T (k0 , p, q), (3.1) ΠN L (k) ≡ 0 sup(k,k0 /²) p−k0 where ² = k0 /p and the lower limit on the q integral is a geometrical constraint. By using the cosine law q 2 = k02 + p2 − 2k0 pz, we transform the variable q in (3.1) to z. Then, eq. (3.1) is rewritten as Z k Z ∞ Z 1 k0 p (−) T (k0 , p, q), dp dz ΠN L (k) = dk0 q 0 0 0 sup(k,k /²) k /(2p) (3.2) Next, we expand the integrand in (3.2) with respect to the smallness parameter. Since the expansion of the geometrical coefficients and of the energy spectrum are the same as those for the 2D NS equation, we use the results of Appendix B of BSL78. The approximate expressions of the geometrical coefficients and the energy spectrum are given by b2 (k0 , p, q) 380 Takahiro Iwayama, Takeshi Watanabe and Theodore G. Shepherd =2 µ 02 ¶¸ · k p2 k0 2 1/2 2 (1 + 2z (1 − z ) ) + O −2z + , k02 p p2 (Vol. 70, p 2 z (1 − z 2 )1/2 + O(1), k0 µ 02 ¶ k ∂E k0 +O . E(q) = E(p) − zp p ∂p p2 a2 (k0 , p, q) = 4 (3.3) 0 b2 (k , q, p) µ 02 ¶¸ · k k0 p2 , = 2 02 (1 − z 2 )1/2 2z + (2z 2 − 1) + O k p p2 (3.5) (3.6) As a result of substituting these expressions into (2.17) with k0 in place of k, and keeping the leading order terms, the integrand of (3.2) is reduced to (3.4) T (k0 , p, q) ¾ ½ 8 2 k0 p 2 1/2 05 3 2 04 4 0 ∂ [p E(p)] 0 pp ' z (1 − z ) θ p E(p) − k p E(k ) . 2k k q πλ6 ∂p Since, to the leading order in ², Z Z 1 k0 /(2p) we obtain Z k dk0 k05 − Z ∞ dp sup(k,k0 /²) 0 Z k dk0 k04 E(k0 ) Z Moreover, we obtain the energy transfer function associated with (3.8), (−) ∂ΠN L (k) ∂k Z ∞ 3 p 2k4 2 θ E(p) dp − νT (k) E(k), ' k5 kpp 6 λ2 k/² λ TN L (k) = (3.9) where Z ∞ νT (k) = k/² p4 ∂ [p E(p)] dp. θkpp 4λ4 ∂p (3.10) The quantity νT (k) is the AM version of the turbulent viscosity derived by Kraichnan for the 2D NS equation.25) The second term on the right-hand side of (3.9) is negligible in the infrared range, because E(k) → 0 in the limit k → 0 for decaying turbulence. Moreover, the viscous damping term, the second term on the lefthand side of (2.4), is also negligible in the infrared range. Furthermore, the integral in (3.9) is nearly independent of its lower limit since it is dominated by the energy ∂ E(k) ' TN L (k) ∼ k5 containing scales. Therefore, ∂t and we obtain the energy spectrum E(k) ∼ k5 in the infrared range. This is consistent with the energy spectrum EK (k) ∼ k7 derived by Yanase and Yamada using the modified zero-fourth cummulant (quasi-normal Markovianized) approximation.12) The analytical result E(k) ∼ k5 is steeper than the low wave number range spectrum of numerical simulations in Larichev and McWilliams1) and Watanabe et al.5) In the next section, we perform direct numerical simulations of decaying turbulence of the CHM equation, and the resulting energy spectra and energy transfers are examined. π , 16 p3 θk0 pp E(p)2 λ6 ∞ dp sup(k,k0 /²) 0 (−) z 2 (1 − z 2 )1/2 dz = 0 (−) ΠN L (k) ' 1 z 2 (1 − z 2 )1/2 dz ' (3.7) §4. p4 ∂ [p E(p)] . θk0 pp 6 2λ ∂p (3.8) Numerical Simulations and Discussion In this section, we report direct numerical simulations of decaying turbulence governed by (1.1) with the hyperviscosity term −ν2 ∇6 φ instead of the normal viscosity term ν∇4 φ. The pseudospectral method is used in double precision arithmetic and at a resolution N 2 , which is the number of grid points in the computational domain, and the truncation wave number is taken as kT = [(N − 1)/3]∆k to suppress aliasing errors, where [ ] denotes the Gaussian symbol. Three simulations (run 1, run 2 and run 3) are conducted for different resolutions N = 256, N = 512, N = 1024, with the minimum grid size held fixed. Initial conditions are made by generating Gaussian random numbers with a mean value of 0 and a variance of 2π for the phase of each Fourier component of φ. We normalize the initial value of the kinetic energy per unit area to be 0.5, and the initial form of the energy spectrum is specified by E(k) ∼ k30 . (k + k0 )60 (4.1) These initial condition are the same as those used in Larichev and McWilliams1) and Watanabe et al.5, 26) Details of conditions in the simulations are listed in Table I. Time integration is done by the AdamsBashforth scheme. In the previous work,5) the fourth order Runge-Kutta scheme for time integration was adopted. We have checked that the results are independent of the time integration scheme in the case N = 256. Thus, we adopt the Adams-Bashforth scheme in this study for computational efficiency. All numerical simulations are done up to t = 200. This time corresponds to about 90 times the characteristic advection time τ , which is defined by Larichev and McWilliams1) as follows: 2001) Infrared Dynamics of Decaying Two-Dimensional Turbulence Governed. . . . Table I. computational domain boundary condition space difference time integration resolution in physical space minimum wave number truncation wave number ∆t dissipation term viscosity coefficient ν2 λ initial kinetic energy initial energy spectrum 381 Conditions of direct numerical simulations. run 1 run 2 run 3 2π × 2π 4π × 4π doubly periodic pseudospectral method Adams-Bashforth scheme 5122 0.5 85 2.5 × 10−3 −ν2 ∇6 φ 3.0 × 10−8 50 0.5 k30 E(k) ∼ (k+k )60 8π × 8π 2562 1 10242 0.25 0 k0 15 Z τ (t) = 0 t {2Zr (t0 )} 0 ¡ ¢2 dt , 1 + λ/k̄K R 1/2 k̄K ≡ kEK (k) dk . (4.2) EK When λ = 0, this definition is equivalent to the eddy turnover time commonly used in studies of 2D NS turbulence.15) Figure 2 shows the temporal evolution of the characteristic advection time τ (t). As shown in Fig. 2, τ (t) is almost the same for each simulation. The reason for the equivalence of τ for each simulation is that the definition of τ is based on the enstrophy which depends mainly on information in the large wave numbers. The evolution of the characteristic wave number k̄ in the three runs also considerably resemble each other (see Fig. 3). We note that the wave number kE corresponding to the peak of the energy spectrum changes discretely due to the finite resolution of wave number space in the direct numerical simulations. However, k̄ and kE have nearly the same values, i.e., they obey the same decay law in time (figure not shown). Thus we use kE for normalizing the energy spectrum in this study, although the charac- Fig. 2. Evolution of the characteristic time τ from (4.2). Broken line, run 1; dashed and dotted line, run 2; solid line, run 3. Since the results for run 2 and run 3 are nearly equivalent, the differences between them are indistinguishable. teristic wave number k̄ was used for the normalization of the energy spectrum in the previous study.5) Figure 4 shows normalized energy spectra (the abscissa is normalized by kE and the ordinate by E(kE )) at times t = 0, 4, 10, 20, 50, 100, 150 and 200. It shows that the energy spectra develop temporally in a self-similar way. Thus, Fig. 4 manifests the universal function h(x) in eq. (1.4). The energy spectrum adjusts to h(x) from the initial condition within τ ' 10. After that, h(x) exists for about 80 times the characteristic advection time. Therefore we stress that h(x) is not evanescent in our simulation. In order to examine the energy spectrum in the low wave number range in detail, we show compensated energy spectra, (k/kE )−δ [E(k)/E(kE )], in Fig. 5. If E(k) ∼ kδ in a range, then the compensated spectrum has a plateau in the corresponding range. The exponent δ of the energy spectrum in the low wave number is δ ' 4.1 for run 1, δ ' 4.45 for run 2 and δ ' 4.6 for run 3. Therefore, as one extends the system size progressively, h(x) in the range x < 1 becomes steeper and approaches the theoretically predicted infrared en- Fig. 3. Evolution of the characteristic wave number k̄. Broken line, run 1; dashed and dotted line, run 2; solid line, run 3. 382 Takahiro Iwayama, Takeshi Watanabe and Theodore G. Shepherd Fig. 4. Normalized energy spectra for (a) run 1, (b) run 2 and (c) run 3. The abscissa is normalized by kE , where the energy spectrum peaks, and the ordinate by E(kE ). Broken line, t = 0; solid lines, t = 4, 10, 20, 50, 100, 150, 200. ergy spectrum h(x) ∼ x5 , or E(k) ∼ k5 . This suggests that the previous studies were under-resolved. It may be noted in passing, for possible future reference, that the high wave number regime k > kE exhibits Fig. 5. Compensated times t = 4, 10, 20, run 2 and (c) run 3. solid lines, t = 100, reference. (Vol. 70, energy spectra (k/kE )−δ E(k)/E(kE ) at 50, 100, 150, and 200 for (a) run 1, (b) Thin solid lines, t = 4, 10, 20, 50; thick 150, 200. The dashed line is a plateau a power-law scaling of approximately E(k) ∼ k−5.5 in both run 2 and run 3 (Figs. 5(b) and 5(c)). All wave numbers shown have k < λ so even the high wave number dynamics are in the AM regime. The energy spectra in 2001) Infrared Dynamics of Decaying Two-Dimensional Turbulence Governed. . . . Fig. 6. Energy fluxes, Π(k), Π(+) (k), and −Π(−) (k) for run 2. Solid thin line, Π(k) at t = 20; broken thin line, −Π(−) (k) at t = 20; dashed and dotted thin line, Π(+) (k) at t = 20; solid thick line, Π(k) at t = 100; broken thick line, −Π(−) (k) at t = 100; dashed and dotted thick line, Π(+) (k) at t = 100. The values of the energy fluxes at t = 100 are multiplied by a factor of 2. the high wave number range in this study are somewhat shallower than those in the previous studies.1, 5) In the derivation of the infrared energy spectrum E(k) ∼ k5 , we (or rather BLS78) made the assumption that the nonlocal interactions with k ¿ p ' q dominated in the triad interactions in wave number space. We can check the validity of the above assumption by examining the results of our direct numerical simulations. (Due to lack of computational resource, the following analyses are done only for run 1 and run 2.) Figure 6 shows the energy fluxes Π(k), Π(+) (k) and −Π(−) (k) for run 2 at t = 20, 100. When Π(+) (k) has a positive value, this means a downward cascade of energy due to the triad interactions (k0 , p, q) with p, q < k and k0 > k. On the other hand, when Π(−) (k) has a positive value, this means an upward cascade of energy due to the triad interactions (k0 , p, q) with p, q > k and k0 < k. In order to redistribute energy within a triad under the constraint of conserving both the total energy and the potential enstrophy, only energy flow to or from the middle wave number among the triad can be permitted.27, 28) Both Π(+) (k) > 0 and Π(−) (k) > 0 as shown in Fig. 6, suggesting that energy flow away from the middle wave number of the triads is statistically dominant. Thus Batchelor’s hypothesis,29) that an initially narrow spectrum spreads out in time about its centroid, is supported in this system. As shown in Fig. 6, Π(k) is entirely expressed by −Π(−) (k) in the low wave number range k/kE < 1. In contrast, in the range k/kE > 1, Π(k) is determined by a competition between Π(+) (k) and −Π(−) (k). That the triad interactions contributing to Π(+) (k), which consist of triads (k0 , p, q) with p, q < k and k0 > k, are insignificant in the low wave number range is a geometrical constraint: for sufficiently small p, q, such triads do not exist. The result that Π(k) can be approximated by −Π(−) (k) in the low wave num- 383 Fig. 7. The ratio of© the nonlocal part of the energy flux £ ¤ª to the total energy flux, Π(k) − Π(+) (k, 2k) − Π(−) (k, 4k) /Π(k), at some instant, for both run 1 and run 2. Broken thin line, run 1 at t = 20; broken thick line, run 1 at t = 100; dashed and dotted thin line, run 2 at t = 20; dashed and dotted thick line, run 2 at t = 100. ber range was also verified for run 1 at t = 20 and 100. To determine which triads of wave numbers contribute to the energy fluxes Π(+) (k) and Π(−) (k) in detail, we calculate filtered fluxes Π(+) (k, kf ) and Π(−) (k, kf ) as introduced by Smith and Yakhot,30) X X 1 X Π(+) (k, kf ) = T (k0 , p, q), (4.3a) 2 0 k<k <kf 0<p<k 0<q<k 1 X Π(−) (k, kf ) = 2 0 X X T (k0 , p, q). (4.3b) 0<k <k k<p<kf k<q<kf Moreover, following Smith and Yakhot,30) we also define nonlocal interactions as those triad interactions for which the ratio of the largest leg to the smallest leg is larger than four. Since the length of one side of a triangle is smaller than the sum of the lengths of the other two sides, one can see that Π(+) (k) = Π(+) (k, 2k). Thus only local triad interactions contribute to Π(+) (k). Figure 7 shows the ratio of the nonlocal part£ of the energy flux to the ¤ª total energy flux, © Π(k) − Π(+) (k, 2k) − Π(−) (k, 4k) /Π(k), at some instant for both run 1 and run 2. The relative contribution of the local and nonlocal interactions to the energy flux changes dramatically at log(k/kE ) ' −0.5 (k/kE ' 0.32). The location of this transition of course depends on our choice kf = 4k. It is approximately determined by the relation k/kE ' k/kf , due to the dominance of the triad interactions between the low wave number range and the energy containing scale kE , as will be seen later. For log(k/kE ) < −0.5, the nonlocal interactions are dominant. On the other hand, in the vicinity of the energy containing scale log(k/kE ) = 0, the local interactions are entirely dominant. The reason why there is such a large deviation of the energy spectrum from the theoretical infrared energy spectrum in the low wave number range for run 1 is the extreme narrowness of the 384 Takahiro Iwayama, Takeshi Watanabe and Theodore G. Shepherd (Vol. 70, wave number range for which nonlocality of the triad interactions is satisfied. In particular, at t ≥ 100 there are only a few wave numbers which satisfy the condition k/kE < 0.32. Note that the energy spectra in the low wave number range in the late stage (t ≥ 100) for run 1 become shallower than those in the early stage (t < 100), which is further support for this argument. Next, we directly examine the triad interactions in wave number space for the direct numerical simulations. T (k, p, q) does not vanish inside of the domain in the (p, q) plane such that k, p, q can be the sides of the triangle k = p +√q, and the width of the nonzero region of T (k, p, q) is 2k (Fig. 1). When we consider the low wave number range k < kE , it is difficult to represent T (k, p, q) by a contour plot since the nonzero region of T (k, p, q) is narrow and data points in the region are not sufficiently dense. Therefore, we consider the azimuthal summation of T (k, p, q) in the (p, q) plane, X T (k, p, q). (4.4) T (k, ρ) = √ 2 2 ∆k ∆k ρ− 2 ≤ p +q <ρ+ 2 Note that the azimuthally summed triad transfer func√ tion T (k, ρ) vanishes for ρ < k/ 2. Moreover, the energy √ containing scale p, q ' kE corresponds to ρ ' 2kE because if we set q = p p(1 + ²), where √ ² is a small pap2 + q 2 = 2p [1 + O(²)]. Figrameter, then ρ = ure 8 shows T (k, ρ) for various k < kE for run 2 at t = 100. For wave numbers k ≤ 3 (for which the conthe T (k, ρ) have signifdition k/kE ≤ 0.32 is satisfied), √ icant values near log[ρ/( 2kE )] ' 0. Thus this result can be interpreted as demonstrating that interactions between wave numbers k ≤ 3 and the energy containing scale p, q ' kE are dominant. The above property of T (k, ρ) was also verified for run 2 at t = 20 and for run 1 at t = 20 and 100. Therefore, we conclude that if the system size is expanded in order to realize k → 0 approximately, the nonlocality of the triad interactions k ¿ p ' q becomes increasingly well satisfied. We try to interpret these properties of the triad interactions by using the evolution equation (2.2a). We divide the wave number range into three categories: the low wave number range denoted by L, the energy containing range E, and the high wave number range H. Then the evolution equation for the low wave number modes can be written formally as ∂ φ̂(L) = CLL φ̂(L)φ̂(L) ∂t +CEE φ̂(E)φ̂(E) + CHH φ̂(H)φ̂(H). (4.5) Here, φ̂ are the Fourier coefficients of φ, and C the interaction coefficients; the viscosity term is omitted for simplicity. Note that only three types of nonlinear terms are possible because the other types of combination of modes, e.g. φ̂(H)φ̂(L) and φ̂(E)φ̂(L), cannot form a triad with φ̂(L), i.e., they are geometrically prohibited. Multiplying (4.5) by φ̂(L), we obtain the evolution equation for the energy spectrum and the energy transfer function as follows: Fig. 8. Azimuthally summed triad transfer functions T (k, ρ) from (4.4) for (a) k = 1, (b) k = 2 and (c) k = 3 for run 2 at t = 100. ∂ φ̂(L)2 = DLL φ̂(L)φ̂(L)φ̂(L) ∂t +DEE φ̂(L)φ̂(E)φ̂(E) + DHH φ̂(L)φ̂(H)φ̂(H). (4.6) We now estimate the magnitude of each term on the right-hand side of (4.6). The interaction coefficients 2001) Infrared Dynamics of Decaying Two-Dimensional Turbulence Governed. . . . £ ¤ D are explicitly written as (p × q)z (q 2 − p2 ) /2 from (2.5b). Because k is the wave number of interest, it belongs to the low wave number range L and is considered as a fixed value in this calculation. On the other hand, both p and q belong to the same wave number range and are treated as variables. Then, we can set q = p(1 + ²), where ² is a small parameter. It follows that q 2 − p2 is proportional to p. Moreover, by using the cosine law k2 = p2 + q 2 − 2pqx, we obtain p 2 |p × q| = [k − (p − q)2 ] [(p + q)2 − k2 ]/2 ∝ p. As a result, the interaction coefficient D is proportional to p2 . Furthermore, in the AM regime, φ̂(p) ∝ [pE(p)]1/2 . Let the magnitudes of the characteristic wave number in the range L, E, and H be kL , kE , and kH , respectively. Using the facts that E(p) ∼ p5 if p is in L and E(p) ∼ p−5.5 if p in H, the ratios can be estimated as follows: µ ¶8 kL DLL φ̂(L)φ̂(L)φ̂(L) ∝ , kE DEE φ̂(L)φ̂(E)φ̂(E) µ ¶−2.5 kH DHH φ̂(L)φ̂(H)φ̂(H) ∝ . (4.7) kE DEE φ̂(L)φ̂(E)φ̂(E) Both ratios are considerably smaller than unity provided kL ¿ kE ¿ kH . Therefore we conclude that the second term on the right-hand side of (4.6), representing the triad interactions between the energy containing range and the low wave number range, is strongly dominant. This discussion is consistent with Fig. 8. Moreover, the abrupt change of the ratio of the nonlocal part of the energy flux to the total energy flux at k/kE ' 0.32 in Fig. 7 originates from the large exponent in the first relation of (4.7). The triads that contribute to the energy flux Π(+) (k) in the low wave number range originate entirely from the first term on the right-hand side of (4.6). Thus the smallness of the first term on the right-hand side of (4.6) is also consistent with Fig. 6. Note that the dominance of the interactions between the low wave number range and the energy containing range is not trivial. If the energy spectrum in the high wave number range were shallower than E(p) ∼ p−3 , then the ratio of the third term to the second term on the right-hand side of (4.6) would become larger than unity. In that case, the interaction of the low wave number range and the high wave number range would be significant. Therefore, the dominance of the interaction between the low wave number range and the energy containing scale is associated with the sharpness of the energy spectrum in CHM turbulence. On the other hand, the local triad interactions in the low wave number range are insignificant as long as the energy spectrum in the low wave number range has a positive exponent. The above discussion can also be applied to turbulence in the NS regime because the energy transfer function in the NS regime is equivalent to that in the AM regime (see eq. (2.5b)). In the NS regime, from the relation φ̂(p) ∝ [p−1 E(p)]1/2 and the fact that E(p) ∼ p3 if p is in L13) and E(p) ∼ p−4 if p in H,31) the ratios can be estimated as follows: µ ¶4 kL DLL φ̂(L)φ̂(L)φ̂(L) , ∝ kE DEE φ̂(L)φ̂(E)φ̂(E) DHH φ̂(L)φ̂(H)φ̂(H) DEE φ̂(L)φ̂(E)φ̂(E) 385 µ ∝ kH kE ¶−3 . (4.8) Therefore, the interaction between the low wave number range and the energy containing scale is dominant in the NS regime as well. §5. Summary In this paper, we have examined the energy spectrum in the low wave number range (k < kE ) of decaying turbulence governed by the Charney-HasegawaMima equation (1.1) in the AM regime where k ¿ λ. The energy spectrum of decaying CHM turbulence develops in a self-similar way, i.e., the universal function h(x) = E(k)/E(kE ), where x = k/kE , exists. This paper is the first to discuss the functional form of h(x) in the low wave number range both theoretically and numerically. By using a quasinormal Markovianized model, we derived the energy spectrum in the infrared range k → 0: E(k) ∼ k5 as k → 0 (or h(x) ∼ x5 as x → 0). We also performed direct numerical simulations of eq. (1.1). When we increase the system size from [0, 2π]2 to [0, 8π]2 , the low wave number spectrum comes closer to the theoretical infrared spectrum. Moreover, we examined the triad interactions responsible for the formation of the low wave number spectrum and found that the energy flux Π(k) can be entirely expressed as −Π(−) (k) for k < kE and that nonlocal triad interactions between the low wave number range and the energy containing scale are dominant. In decaying 2D NS turbulence, predictions from closure theory or the classical theory of 2D turbulence fail due to the existence of coherent vortices and the associated intermittency.32) However, in the system considered here, the prediction from closure theory is in fairly good agreement with direct numerical simulations. This suggests that intermittency is weak in the AM regime of CHM turbulence. This is confirmed by the fact that the flatness of the potential vorticity field q = (∇2 −λ2 )φ calculated from the direct numerical simulations is close to Fig. 9. Evolution of flatness of the potential vorticity field. Broken line, run 1; dashed and dotted line, run 2; solid line, run 3. 386 Takahiro Iwayama, Takeshi Watanabe and Theodore G. Shepherd the Gaussian value of 3 (Fig. 9), consistent with Polvani et al.33) Therefore we can anticipate that the system considered here is one for which the classical theory of 2D turbulence is applicable. However, in order to verify this conjecture, we would have to examine higher order moments of the potential vorticity field. This point should be the subject of a future study. Acknowledgements T. I. expresses his deep thanks to Prof. H. Fujisaka of Kyoto University and Prof. M. Takahashi of the University of Tokyo for helpful comments and encouragement. He also wishes to thank Prof. T. Nakano of Chuo University, Profs. A. Yoshizawa and F. Hamba, Dr. N. Yokoi and the other members of the Turbulence Seminar at the University of Tokyo for many stimulating discussions. T. W. was supported by a JSPS Research Fellowship for Young Scientists. This work was partially supported by the Grant-in-Aid for Scientific Research No. 11740267 from the Ministry of Education, Science, Sports and Culture of Japan and by the Center for Climate System Research, University of Tokyo. T. G. S. is supported by the Natural Sciences and Engineering Research Council and by the Meteorological Service of Canada. The GFD-DENNOU Library was used for drawing figures. 1) V. D. Larichev and J. C. McWilliams: Phys. Fluids A 3 (1991) 938. 2) M. Ottaviani and J. A. Krommes: Phys. Rev. Lett. 69 (1992) 2923. 3) N. Kukharkin, S. A. Orszag and V. Yakhot: Phys. Rev. Lett. 75 (1995) 2486. 4) T. Watanabe, H. Fujisaka and T. Iwayama: Phys. Rev. E 55 (1997) 5575. 5) T. Watanabe, T. Iwayama and H. Fujisaka: Phys. Rev. E 57 (1998) 1636. (Vol. 70, 6) A. Hasegawa and K. Mima: Phys. Fluids 21 (1978) 87. 7) J. Pedlosky: Geophysical Fluid Dynamics, 2nd ed. (SpringerVerlag, New York, 1987). 8) R. H. Kraichnan: Phys. Fluids 10 (1967) 1417. 9) D. Fyfe and D. Montgomery: Phys. Fluids 22 (1979) 246. 10) G. K. Batchelor: Phys. Fluids Suppl. II 12 (1969) II-233. 11) T. Watanabe: Ph. D. Thesis, Department of Physics, Kyushu University, Fukuoka, 2000. 12) S. Yanase and M. Yamada: J. Phys. Soc. Jpn. 53 (1984) 2513. 13) C. Basdevant, M. Lesieur and R. Sadourny: J. Atmos. Sci. 35 (1978) 1028. 14) U. Frisch and P. L. Sulem: Phys. Fluids 27 (1984) 1921. 15) J. R. Chasnov: Phys. Fluids 9 (1997) 171. 16) R. H. Kraichnan: J. Fluid Mech. 47 (1971) 525. 17) A. Pouquet, M. Lesieur, J. C. André and C. Basdevant: J. Fluid Mech. 72 (1975) 305. 18) K. Ohkitani: Phys. Fluids A 2 (1990) 1529. 19) M. E. Maltrud and G. K. Vallis: Phys. Fluids A 5 (1993) 1760. 20) V. Borue: Phys. Rev. Lett. 71 (1993) 3967. 21) S. Nazarenko and J.-P. Laval: J. Fluid Mech. 408 (2000) 301. 22) M. Lesieur: Turbulence in Fluids, Third Revised and Enlarged Edition (Kluwer, Dordrecht, 1997). 23) T. Tatsumi and S. Yanase: J. Fluid Mech. 110 (1981) 475. 24) S. A. Orszag: J. Fluid Mech. 41 (1970) 363. 25) R. H. Kraichnan: J. Atmos. Sci. 33 (1976) 1521. 26) Watanabe et al.5) normalized the initial condition by taking the total energy to be 0.5. However, the difference in normalization conditions between Larichev and McWilliams1) and Watanabe et al.5) is unimportant. 27) R. Fjørtoft: Tellus 5 (1953) 225. 28) P. E. Merilees and H. Warn: J. Fluid Mech. 69 (1975) 625. 29) G. K. Batchelor: The Theory of Homogeneous Turbulence (Cambridge University Press, 1953) p.186. 30) L. M. Smith and V. Yakhot: J. Fluid Mech. 274 (1994) 115. 31) P. Santangelo, R. Benzi and B. Legras: Phys. Fluids A 1 (1989) 1027. 32) J. R. Herring and J. C. McWilliams: J. Fluid Mech. 153 (1985) 229. 33) L. M. Polvani, J. C. McWilliams, M. A. Spall and R. Ford: Chaos 4 (1994) 177.