Vortex statistics for turbulence in a container with rigid boundaries Clercx, H.J.H.; Nielsen, A.H.

Similar documents
Geometry explains the large difference in the elastic properties of fcc and hcp crystals of hard spheres Sushko, N.; van der Schoot, P.P.A.M.

A note on non-periodic tilings of the plane

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Minimum analysis time in capillary gas chromatography. Vacuum- versus atmospheric-outlet column operation Leclercq, P.A.; Cramers, C.A.M.G.

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Spatial decay rate of speech in open plan offices: the use of D2,S and Lp,A,S,4m as building requirements Wenmaekers, R.H.C.; Hak, C.C.J.M.

Two-dimensional turbulence in square and circular domains with no-slip walls

The M/G/1 FIFO queue with several customer classes

On Decaying Two-Dimensional Turbulence in a Circular Container

Influence of the shape of surgical lights on the disturbance of the airflow Zoon, W.A.C.; van der Heijden, M.G.M.; Hensen, J.L.M.; Loomans, M.G.L.C.

Multi (building)physics modeling

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

On the decay of two-dimensional homogeneous turbulence

Thermodynamics of monomer partitioning in polymer latices: effect of molar volume of the monomers

Max Planck Institut für Plasmaphysik

On a set of diophantine equations

Massively parallel molecular-continuum simulations with the macro-micro-coupling tool Neumann, P.; Harting, J.D.R.

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

The dipole moment of a wall-charged void in a bulk dielectric

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Online algorithms for parallel job scheduling and strip packing Hurink, J.L.; Paulus, J.J.

Kommerciel arbejde med WAsP

Polydiagnostic study on a surfatron plasma at atmospheric pressure

The IEA Annex 20 Two-Dimensional Benchmark Test for CFD Predictions

Notes on cooperative research in the measurement of Gottwein-temperature Veenstra, P.C.

Prediction of airfoil performance at high Reynolds numbers.

The calculation of stable cutting conditions in turning

Validation of Boundary Layer Winds from WRF Mesoscale Forecasts over Denmark

Sound absorption properties of a perforated plate and membrane ceiling element Nilsson, Anders C.; Rasmussen, Birgit

Aalborg Universitet. Lecture 14 - Introduction to experimental work Kramer, Morten Mejlhede. Publication date: 2015

Aalborg Universitet. Specification of a Two-Dimensional Test Case Nielsen, Peter Vilhelm. Publication date: 1990

Reducing Uncertainty of Near-shore wind resource Estimates (RUNE) using wind lidars and mesoscale models

Influence of turbulence model on thermal plume in indoor air flow simulation Zelensky, P.; Bartak, M.; Hensen, J.L.M.; Vavricka, R.

Decaying 2D Turbulence in Bounded Domains: Influence of the Geometry

Steady State Comparisons HAWC2 v12.2 vs HAWCStab2 v2.12

A Note on Powers in Finite Fields

Lidar calibration What s the problem?

Radioactivity in the Risø District January-June 2014

Aalborg Universitet. Comparison between Different Air Distribution Systems Nielsen, Peter Vilhelm. Publication date: 2006

A comparison study of the two-bladed partial pitch turbine during normal operation and an extreme gust conditions

Published in: Proceeding of 9th SENVAR / 2nd ISESEE, Shah Alam, Malaysia, December 2008

Lagrangian acceleration in confined 2d turbulent flow

Graphs with given diameter maximizing the spectral radius van Dam, Edwin

Wake decay constant for the infinite wind turbine array Application of asymptotic speed deficit concept to existing engineering wake model

Simulations with a dynamic reaction-diffusion model of the polymer grating preparation by patterned ultraviolet illumination

Generating the optimal magnetic field for magnetic refrigeration

A Procedure for Classification of Cup-Anemometers

Measurements of electric-field strengths in ionization fronts during breakdown Wagenaars, E.; Bowden, M.D.; Kroesen, G.M.W.

Radioactivity in the Risø District July-December 2013

Design of Robust AMB Controllers for Rotors Subjected to Varying and Uncertain Seal Forces

Simulating wind energy resources with mesoscale models: Intercomparison of stateof-the-art

Crack Tip Flipping: A New Phenomenon yet to be Resolved in Ductile Plate Tearing

Benchmark of Femlab, Fluent and Ansys

Numerical modeling of the conduction and radiation heating in precision glass moulding

Multiple Scenario Inversion of Reflection Seismic Prestack Data

Published in: Proceedings of the 5th European Conference on Computational Fluid Dynamics, ECCOMAS CFD 2010

An Equivalent Source Method for Modelling the Lithospheric Magnetic Field Using Satellite and Airborne Magnetic Data

Modifications of the V2 Model for Computing the Flow in a 3D Wall Jet Davidson, L.; Nielsen, Peter Vilhelm; Sveningsson, A.

An iterative algorithm for nonlinear wavelet thresholding: Applications to signal and image processing

Determination of the minimum size of a statistical representative volume element from a fibre-reinforced composite based on point pattern statistics

A numerical study of vortex-induced vibrations (viv) in an elastic cantilever

Validity of PEC Approximation for On-Body Propagation

Solution to problem 87-6* : The entropy of a Poisson distribution Boersma, J.

Strongly 2-connected orientations of graphs

An optimization problem related to the storage of solar energy van der Wal, J.

Vortex Generator Induced Flow in a High Re Boundary Layer

Published in: 11th International Radiance Workshop, 12 September 2012, Copenhagen, Denmark

Surface composition of ordered alloys: An off-stoichiometric effect

Tilburg University. Strongly Regular Graphs with Maximal Energy Haemers, W. H. Publication date: Link to publication

Reflection of sound from finite-size plane and curved surfaces

A test reaction for titanium silicalite catalysts

Surfactant-free emulsion polymerization of styrene using crosslinked seed particles Eshuis, A.; Leendertse, H.J.; Thoenes, D.

Visualizing multiple quantile plots

On numerical evaluation of two-dimensional phase integrals

Analytical Studies of the Influence of Regional Groundwater Flow by on the Performance of Borehole Heat Exchangers

Measurements of the angular distribution of diffuse irradiance

Lattice Boltzmann Modeling From the Macro- to the Microscale - An Approximation to the Porous Media in Fuel Cells -

Accelerating interior point methods with GPUs for smart grid systems

Optimizing Reliability using BECAS - an Open-Source Cross Section Analysis Tool

Exposure Buildup Factors for Heavy Metal Oxide Glass: A Radiation Shield

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Application of mesoscale models with wind farm parametrisations in EERA-DTOC

Modelling Salmonella transfer during grinding of meat

The effect of microbubbles on developed turbulence

Aalborg Universitet. Selection of Air Terminal Device. Nielsen, Peter V. Publication date: 1988

Validation and comparison of numerical wind atlas methods: the South African example

Aalborg Universitet. Seawall Overtopping Tests sea erosion study : Sohar Cornice Frigaard, Peter. Publication date: 1990

Realtime 3D stress measurement in curing epoxy packaging

An Introduction to Theories of Turbulence. James Glimm Stony Brook University

Impact of the interfaces for wind and wave modeling - interpretation using COAWST, SAR and point measurements

Comparisons of winds from satellite SAR, WRF and SCADA to characterize coastal gradients and wind farm wake effects at Anholt wind farm

Design possibilities for impact noise insulation in lightweight floors A parameter study

Optimal free parameters in orthonormal approximations

Alignment of stress, mean wind, and vertical gradient of the velocity vector

Aalborg Universitet. Windows Heiselberg, Per Kvols; Svidt, Kjeld; Nielsen, Peter Vilhelm. Publication date: 2000

Determination of packaging induced 3D stress utilizing a piezocoefficient mapping device

Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Diffusion of sulfuric acid in protective organic coatings

Transcription:

Vortex statistics for turbulence in a container with rigid boundaries Clercx, H.J.H.; Nielsen, A.H. Published in: Physical Review Letters DOI: 0.03/PhysRevLett.85.752 Published: 0/0/2000 Document Version Publisher s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: A submitted manuscript is the author's version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. The final author version and the galley proof are versions of the publication after peer review. The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication Citation for published version (APA): Clercx, H. J. H., & Nielsen, A. H. (2000). Vortex statistics for turbulence in a container with rigid boundaries. Physical Review Letters, 85(4), 752-755. DOI: 0.03/PhysRevLett.85.752 General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. Users may download and print one copy of any publication from the public portal for the purpose of private study or research. You may not further distribute the material or use it for any profit-making activity or commercial gain You may freely distribute the URL identifying the publication in the public portal? Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Download date: 07. Dec. 208

Vortex Statistics for Turbulence in a Container with Rigid Boundaries H. J. H. Clercx and A. H. Nielsen 2 Department of Physics, Eindhoven University of Technology, P.O. Box 53, NL-5600 MB Eindhoven, The Netherlands 2 Optics and Fluid Dynamics Department, Forskningscenter Risø, DK-4000 Roskilde, Denmark (Received 29 November 999; revised manuscript received 4 April 2000) The evolution of vortex statistics for decaying two-dimensional turbulence in a square container with rigid no-slip walls is compared with a few available experimental results and with the scaling theory of two-dimensional turbulent decay as proposed by Carnevale et al. Power-law exponents, computed from an ensemble average of several numerical runs, coincide with some experimentally obtained values, but not with data obtained from numerical simulations of decaying two-dimensional turbulence with periodic boundary conditions. PACS numbers: 47.27.Eq, 47.32.Cc The emergence and the temporal evolution of a hierarchy of coherent vortices in decaying two-dimensional (2D) turbulence has been subject to many analytical, numerical, and experimental studies. One of the most remarkable theoretical results of the last decade is the scaling theory as proposed by Carnevale et al. [,2]. In this theory it is supposed that the average number density of vortices decreases algebraically, r t ~ t 2z, with z so far undetermined. Numerical simulations with a simple, punctuated- Hamiltonian, dynamical model for the evolution of a system of coherent vortices [] and also numerical simulations of the Navier-Stokes equations, although hyperviscosity has been used, show that z is approximately 0.72 6 0.03 [3]. This value of the decay exponent suggests that the straightforward scaling theory by Batchelor [4] is incorrect. His theory is based on the assumption that for nearly inviscid 2D turbulent flows the energy might still be considered as conserved, but the enstrophy not, because viscosity is preferentially dissipative on the smallest scales of the flow. Dimensional analysis then results in r t ~ t 22, which shows a large discrepancy with the Carnevale approach. Beside energy conservation Carnevale et al. also assumed that the extremum of vorticity, v ext, is conserved. Dimensional analysis then yields for the average number density r t, the average enstrophy Z t, the average vortex radius a t, and the average mean vortex separation r t r t ~ L 22 t T 2z, Z t ~ T 22 t T 2z 2, () a t ~ L t T z 4, r t ~ L t T z 2. (2) The characteristicplength scale L and time scale T are defined by L vext 2 E and T v 2 ext, respectively. As mentioned before, the power-law exponent is a free parameter which has to be predicted on the basis of numerical simulations. Another remarkable consequence of this scaling theory is the rather small decay exponent for the enstrophy, Z t ~ t 20.36 (z~0.72), compared to Z t ~ t 22 derived by Batchelor [4]. Measurements of the decay exponent z have been performed in experiments of decaying 2D turbulence in thin electrolyte solutions in a rectangular container [5,6]. A 2D array of magnets is located just below the bottom of the container, and in combination with an electric current the fluid is forced electromagnetically. In the experiments arrays of 6 3 6, 8 3 8, and 0 3 0 magnets, where nearest neighbors have oppositely signed magnetic polarity, have been used. The interaction of the magnetic field with the electric current results in a regular array of counterrotating vortices. After the current is switched off, the flow will relax and the process of decaying 2D turbulence can be investigated. The initial value of the Reynolds number, based on the scale of the container, is typically Re 2000. The experiments yielded power laws for the average number density of vortices and associated quantities which could be correlated with the scaling theory introduced by Carnevale. These experiments [5,6] are considered as a confirmation of the validity of this scaling theory despite the differences between the theoretical and experimental setup such as the presence of no-slip walls and finite Reynolds number effects in the latter. Recent numerical simulations of 2D turbulence [7,8] have elucidated the role of no-slip boundaries in general, and vortex-wall interactions in particular. It is therefore necessary to carry out numerical simulations of decaying 2D turbulence in containers with no-slip boundaries with analogous initial conditions as those employed in the experiments discussed above (i.e., an array of N 3 N vortices). Numerical experiments with the same initial flow field have also been repeated with periodic boundary conditions. The ensemble averaged power-law exponents of the no-slip and the Fourier runs are compared with each other and with the experimental results. It should be emphasized that in the experiments as well as in the numerical simulations isotropy and homogeneity are absent when noslip walls confine the flow. A direct comparison with the Carnevale theory is therefore hampered because this theory assumes isotropy and homogeneity. The numerical simulations of the 2D Navier-Stokes equations on a bounded domain were performed with a 2D dealiased Chebyshev pseudospectral method [9], with a maximum of 53 Chebyshev modes in each direction 752 003-9007 00 85(4) 752(4)$5.00 2000 The American Physical Society

for Re 20 000 (36 modes for Re 0 000 and 257 modes for Re 5000). The numerical computations of decaying 2D Navier-Stokes turbulence with periodic boundary conditions were carried out with a standard 2D Fourier pseudospectral method with a maximum of 34 active Fourier modes in each direction. In both cases neither hyperviscosity nor any other artificial dissipation has been used. The integral-scale Reynolds number of the flow is defined as Re UW n with U the rms velocity of the initial flow field, W the half width of the container, and n the kinematic viscosity of the fluid. Time has been made dimensionless by W U and vorticity by U W. The initial microscale Reynolds number is defined as Re micr 2Re v 0, with v 0 the (dimensionless) initial rms vorticity. In our numerical experiments v 0 38.0 6 0.5, thus corresponding with Re micr 263, 526, and 052, respectively. The time t is defined as t v 0t N, with t the dimensionless time and N 2 the number of vortices, and t corresponds approximately with the (initial) eddy turnover time. The initial condition for the velocity field consists of 00 nearly equal-sized Gaussian vortices (thus N 0). The vortices have a dimensionless radius of 0.05 and a dimensionless absolute vortex amplitude of jv max j 00. Half of the vortices have positive circulation, and the other vortices have negative circulation. The vortices are placed on a regular lattice, well away from the boundaries, with a random displacement of the vortex centers equal to approximately 6% of the dimensionless lattice parameter l, with l 0.7. A smoothing function, similar to the one employed in Refs. [7,8], has been used in order to ensure the no-slip condition exactly. The initial conditions for the simulations with periodic boundary conditions are the same. Figure shows two snapshots of the vorticity distribution of decaying turbulence in a container with no-slip walls (a) and for the case with periodic boundary conditions (b). The power laws discussed in this Letter are computed from ensemble averages of 2 runs for Re 5000 and 8 runs for Re 0 000 [2 runs have been performed for Re 20 000 and only power-law exponents for Z t could be computed]. The ensemble averaged decay exponents are also representative for the power laws 0.5 0-0.5 - - -0.5 0 0.5 (a) no-slip 0.5 0-0.5 - - -0.5 0 0.5 (b) periodic FIG.. Vorticity contour plots of the simulations with no-slip (a) and periodic (b) boundary conditions (Re 20 000, t 40). The contour level increment in units of v 0 is 0.2. obtained for each individual run. The slight variation observed in the set of individual decay exponents will be indicated by the error margin. The error margins are typically somewhat smaller for the runs with periodic boundary conditions compared to those for the runs with no-slip walls. Because of the finiteness of the initial Reynolds number in our simulations, the domain averaged kinetic energy E of the flow decreases during the decay process. The initial value of the kinetic energy is E 2, while at t 80 and Re 5000 the ensemble averaged kinetic energy of the flows with periodic boundary conditions has decreased to E periodic 0.93 and for flows with no-slip boundaries to E no-slip 0.55, with denoting the ensemble average. For Re 0 000 we find E periodic.25 and E no-slip 0.83 (note that all quantities are in dimensionless units). The interaction time scale of the vortices (e.g., the time scale associated with merging processes which is of the order of a few initial eddy turnover times) is so short that the energy might be considered as constant during interaction. From the simulations with no-slip and with periodic boundary conditions it is also clear that the conservation of v ext is strongly violated even when its value is normalized by p E. Because of the production of very strong vortices in the boundary layers, the runs with noslip boundaries show a slower algebraic decay compared to those with periodic boundary conditions; see Table I. It TABLE I. numbers. Re Power-law exponents for runs with no-slip and with periodic boundary conditions for different integral-scale Reynolds vext p E no-slip Z no-slip Z E no-slip r no-slip r no-slip 5000 20.35 6 0.05 2.00 6 0. 20.80 6 0. 20.85 6 0. 0.45 6 0.07 0 000 20.30 6 0.05 20.85 6 0. 20.70 6 0. 20.75 6 0. 0.40 6 0.07 20 000 20.80 6 0. 20.65 6 0. Re vext p E periodic Z periodic Z E periodic r periodic r periodic 5000 20.47 6 0.04 2.30 6 0.05 2.3 6 0.05 2.3 6 0. 0.65 6 0.05 0 000 20.38 6 0.04 2.5 6 0.05 20.98 6 0.05 2.03 6 0. 0.60 6 0.05 20 000 20.97 6 0.05 20.90 6 0.05 753

is interesting to note that few numerical runs with periodic boundary conditions have been carried out with hyperviscosity instead of Newtonian viscosity. A run with initial conditions as employed in the present study and a few runs with initial conditions as used by Weiss and McWilliams [3] show that v ext ~t 2a, with a 0.05 0.0. Similar behavior has been observed in a few experiments [6], and in the computed final decay regime of evolving 2D Navier- Stokes turbulence [7]. However, the flows in these particular experiments and simulations are characterized by relatively small Reynolds numbers (Re ø 000) in this regime and any agreement with the theory proposed by Carnevale et al. is therefore rather accidental. In Fig. 2 we have plotted Z t for Re 5000 (Fig. 2a) and Re 0 000 (Fig. 2b). From these results it is clear that the decay of the enstrophy for the no-slip runs is distinctly different from the results computed with periodic boundary conditions. It should be noted that our data for the runs with periodic boundary conditions are consistent with the numerical study by Chasnov [0] for decaying 2D turbulence with similar initial microscale Reynolds numbers. It is more difficult to find power-law behavior for the runs with no-slip boundaries. Because of strong vortex-wall interactions the enstrophy is very spiky (note that the plotted data already represent an average over many runs). The final stage (t.40) indicates, at least for the runs with Re 0 000 and 20 000, that a regime with a significantly lower decay exponent is reached [ Z t ~t 20.5 for Re 0 000]. The computation of the average number density of vortices r t, the average vortex radius a t, and the average mean vortex separation r t is based on a discrete wavelet packet transform (WPT) technique [ 3]. In order to consider a structure as a vortex the following conditions should be satisfied: The aspect ratio of the long and short axis of the (nearly) ellipsoidal patch should be smaller than two. The (absolute) value of the vorticity extremum should always be larger than 20% of the absolute value of the vorticity extremum of the strongest vortex. The computation of the vortex strength and the vortex radius is performed by taking into account the 20% of the strongest vortices detected by the WPT algorithm only [except in the final stage when the number of vortices becomes too small, i.e., r t, 0]. An important condition for the present comparison is that the data from the simulations with no-slip and with periodic boundary conditions are treated in exactly the same way. In Fig. 3 we have plotted r t and r t (based on the distance of the vortex centers), respectively. It is obvious from these data that the no-slip boundaries result in a denser system of vortices: The number of vortices decreases more slowly and the average mean vortex separation increases more slowly in the no-slip case compared to the results obtained for the periodic runs (see Table I). The final stage (t.40) indicates again a regime with a somewhat smaller decay exponent for the number of vortices (see Fig. 3): r t ~t 20.5 (Re 5000) and r t ~t 20.45 (Re 0 000). The NUMBER DENSITY OF VORTICES 00 0 n=-.3 n=-0.85 n=-0.5 (a) 0 00 000 NUMBER DENSITY OF VORTICES 00 0 n=-.03 Re=0,000 n=-0.75 n=-0.45 (b) 0 00 000 ENSTROPHY 000 00 0 n=-.30 n=-.00 ENSTROPHY 000 00 0 n=-0.85 n=-0.5 n=-.5 Re=0,000 MEAN VORTEX SEPARATION n=0.65 n=0.45 MEAN VORTEX SEPARATION Re=0,000 n=0.65 n=0.40 (a) 0 00 000 (b) 0 00 000 FIG. 2. Ensemble averaged enstrophy plotted as a function of t. Drawn lines represent the no-slip results and the dashed lines the results for periodic boundary conditions. 0. (c) 0 00 000 0. (d) 0 00 000 FIG. 3. Ensemble averaged number density (a),(b) and mean vortex separation (c),(d) plotted as a function of t. Drawn lines (with ) represent the no-slip results and the dashed lines (with 3) the results for periodic boundary conditions. 754

average vortex radius grows like a t ~t 0.2560.03 for the no-slip runs as well as the periodic case (both for Re 5000 and 0 000). The data displayed in Fig. 3 suggest parallel evolution for the periodic and no-slip runs for t#20. Inspection of the power-law compensated ensemble averages [e.g., for the vortex density with Re 5000: r t no-slip t 0.85 and r t periodic t.3 ] and particularly the associated horizontal plateau-value regime, however, clearly supports the values of the exponents summarized in Table I. For t#0 it is difficult to discriminate between the two cases. This is mainly due to the finite time necessary for boundary layers to develop and to become dynamically active. From our data it can be conjectured that this process takes some 4 to 8 initial eddy turnover times. It is clear that a Reynolds number dependence is still present, and it is not yet known how the vortex statistics will be modified by increasing the Reynolds number substantially. However, the different power laws obtained for the no-slip and the periodic runs give strong evidence that the vortex statistics are modified by the presence of boundaries. Comparison of our data from simulations with no-slip walls with some experimental power-law exponents obtained by Hansen et al. [6], r t ~t 20.7060., Z t E t ~t20.4760.06, (3) a t ~t 0.260.06, r t ~t 0.3860.08, (4) requires simulations with a smaller large-scale Reynolds number (Re 000 and 2000). The rather steep decay of r t and Z t E t, observed for the simulations with Re 5000 and 0 000 for t,40, are virtually absent for these relatively low Reynolds number simulations. The computed power-law exponents are r t ~ t 20.7, Z t E t ~t 20.5, a t ~t 0.2, and r t ~ t 0.5. Analogous power-law behavior is found during the final decay stage (t.40) of the runs with substantially higher Reynolds number [ r t ~t 20.5 and Z t E t ~t 20.4 ]. In the laboratory experiments with electromagnetically forced electrolytes the energy decay is enhanced by bottom drag. We have repeated several of the present simulations for Re 5000 and 0 000 in bounded domains with an additional friction term representing the bottom drag. The numerical values for the bottom friction has been chosen in the same range as for the experiments. It is found that up to t 00 the computed power-law exponents for v ext p E, Z E, r, a, and r are approximately the same for runs with and without bottom friction. For larger eddy turnover times the nonlinearity of the flow starts to become depleted due to bottom friction and a comparison between simulations with and without bottom friction is not opportune. An alternative approach to analyze the numerical data is by normalizing r t, Z t, a t, and r t with T and L [see Eqs. () and (2)]. Note that L and T depend on time, instead of being constants, due to finite Reynolds number effects. The power-law exponent z is then determined by computing L 2 r t, T 2 Z t, a t L, and r t L. The power-law exponents calculated in this way show some agreement with the Carnevale approach for the runs with periodic boundary conditions. We were not able to observe power laws for the average number density and other quantities from the runs with no-slip boundaries, and computing a single power-law exponent describing their decay properties is impossible. To summarize, we can conclude that vortex statistics for decaying 2D turbulence in containers with no-slip walls differ considerably from the case with periodic boundary conditions. We believe that any agreement between experimentally obtained power-law exponents and the Carnevale approach is rather accidental: The experimental data also agree with our numerical simulations with no-slip walls. Moreover, the power-law exponents obtained from our simulations with periodic boundary conditions, which is the closest numerical test for the scaling theory, disagree with the experimental data. One of us (H. J. H. C.) is grateful for support by the European Science Foundation (ESF-TAO/998/3). This work was sponsored by the Stichting Nationale Computerfaciliteiten (National Computing Facilities Foundation, NCF) for the use of supercomputer facilities, with financial support from the Netherlands Organization for Scientific Research (NWO). [] G. F. Carnevale, J. C. McWilliams, Y. Pomeau, J. B. Weiss, and W. R. Young, Phys. Rev. Lett. 66, 2735 (99). [2] G. F. Carnevale, J. C. McWilliams, Y. Pomeau, J. B. Weiss, and W. R. Young, Phys. Fluids A 4, 34 (992). [3] J. B. Weiss and J. C. McWilliams, Phys. Fluids A 5, 608 (993). [4] G. K. Batchelor, Phys. Fluids Suppl. II 2, 233 (969). [5] O. Cardoso, D. Marteau, and P. Tabeling, Phys. Rev. E 49, 454 (994). [6] A. E. Hansen, D. Marteau, and P. Tabeling, Phys. Rev. E 58, 726 (998). [7] H. J. H. Clercx, S. R. Maassen, and G. J. F. van Heijst, Phys. Fluids, 6 (999). [8] H. J. H. Clercx and G. J. F. van Heijst, Phys. Rev. Lett. (to be published). [9] H. J. H. Clercx, J. Comput. Phys. 37, 86 (997). [0] J. R. Chasnov, Phys. Fluids 9, 7 (997). [] A. Siegel and J. B. Weiss, Phys. Fluids 9, 988 (997). [2] M. V. Wickerhauser, Adapted Wavelet Analysis from Theory to Software Algorithms (A.K. Peters, Ltd., Wellesley, MA, 994). [3] M. Farge, E. Goirand, Y. Meyer, F. Pascal, and M. V. Wickerhauser, Fluid Dyn. Res. 0, 229 (992). 755