Annual Journal of Hydraulic Engineering, JSCE, Vol.52, 2008, February Annual Journal of Hydraulic Engineering, JSCE, Vol.52, 2008, February NUMERICAL MODELLING OF BEDFORM DEVELOPMENT; FORMATION OF SAND-WAVELETS IN HYDRODYNAMICALLY SMOOTH FLOW OVER AN ERODIBLE BED Y Q. NGUYEN1, John C. WELLS2 1 Graduate student, 2Member of JSCE, Dr. of Eng., Associate Professor Department of Civil & Environmental Engineering, Ritsumeikan University (1-1-1 Nojihigashi, Kusatsu, Shiga, Japan) Formation of the initial bedforms, known as sand-wavelets, developing from a flat erodible bed in hydrodynamically smooth flow has been studied numerically. The computational model consists of hydrodynamic and sediment transport equations. The unsteady hydrodynamic equations are solved by Large-Eddy-Simulation (LES) and Immersed-Boundary-Method (IBM) on a non-bodyfitted grid. Dynamics of individual grains are neglected and only bedload is considered, as calculated from Van Rijn’s equation (1984, Journal of Hydraulic Engineering, Vol. 110). Two-dimensional (2D) bed evolution is coupled with the above 3-dimensional (3D) hydrodynamic equations. The computed wavelength agrees with linear stability analysis and with experimental data in the literature. The length of the sand-wavelets is observed to correspond to the disturbance region of bed shear stress caused by discontinuities on the bed surface, as postulated by Yalin (1977, Mechanics of Sediment Transport, Pergamon press). Keywords: numerical simulation, bedform, sand-wavelets, bedload transport. 1. INTRODUCTION When an erodible riverbed or seabed is exposed to a current, the bed usually becomes unstable and bedforms appear1),2). Coleman & Melville3) and Coleman et al4) reported that the first sand-waves observed on a flat bed are instigated by random pileups of sediment. When the height of these pileups is about that of the bed roughness, small sand-waves of very low height but finite wavelength are formed. They defined these nascent sand-waves as sand-wavelets; such structures grow further in time to mature ripples or dunes depending on the flow characteristics. Various trends for the length of sand-wavelets have been reported. Coleman et al4) showed from their experimental results and data in the literature, the sand-wavelet’s wavelength is scaled only with the grainsize, nearly independent of the flow conditions for laminar, transitional and turbulent flows with particle Reynolds number up to 50. By fitting all the data, they proposed a relation between wavelet length and grainsize as 175 . , where λ and d are the wavelength and mean grainsize, respectively, in millimeters. In contrast, Langlois & Valance5) reported that the wavelength is roughly independent of grainsize for particle Reynolds number in their experimental range of 1-14 in turbulent flows. Kuru et al6) concluded that the wavelength increases with an increase in the flow rate of the carrier fluid in the transitional flows. To re-examine these different conclusions for turbulent flows, the test cases of Coleman & Melville3), Coleman et al4), Langlois & Valance5), and Kuru et al6) have been re-plotted (see Fig. 3). With this arrangement, the wavelength when non-dimensionalized by the grainsize is found to be a function of grain Reynolds number. For particle Reynolds number above about 4, all data sets show a wavelength around 200. By contrast, for the lower range of the Reynolds number, there is a strong discrepancy between the results of Coleman et al4) and Langlois and Valance5) and a strong scatter of Kuru et al’s6) results, and little trend is apparent. Aiming at obtaining a clear trend of variation of sand-wavelet’s wavelength for grain Reynolds number smaller than 3, i.e. the regime of hydrodynamically smooth flows, the formation of initial bedforms is numerically studied in this paper. With numerical simulation, we also hope to get more understanding of the interaction between the bedform and the flow during the initial stage of the bedform as described by Yalin1),7) (see Section 3). The computational method consists of a model for fluid flow and a model for the sediment transport. The fluid flow is solved with unsteady 3D Large-Eddy-Simulation (LES) method while the deformed boundary is treated with the Immersed-Boundary-Method (IBM) on a fixed - 163 - Cartesian non-bodyfitted grid. The model for sediment transport includes equations of sediment flux and equations of bed profile evolution in 2D which is coupled with the 3D LES equations. By employing the IBM, remeshing to fit the bed surface when it evolves in time is not required. Moreover, Cartesian solvers are used instead of body-fitted solvers, yielding better cost-efficiency8). As observed in the experiments3),4), the sand-wavelets are instigated by very small disturbances on the bed surface which grow in time, interact with the flow, and cause the local shear stress distribution to deviate from the standard one1),7). Therefore, to model correct behavior of the bedform at this initial stage, the flow solver must accurately reproduce local turbulence, particularly bed shear stress fluctuations around the disturbances. Among available methods which explicitly represent the turbulence, hence can satisfy the above requirements, LES and RANS (Reynolds-Averaged Navier-Stokes equations) can be considered as practical candidates9). Between the two, Keylock et al9) showed that LES is preferable for this kind of problem since most RANS models are intended for accurate representations of the mean flow field only. For example, Chang & Scotti10) , comparing LES and with the ω RANS model for separating flows over ripples, reported that RANS substantially underpredicted Reynolds stress and overpredicted vertical velocity, while LES agreed very well with DNS and experiment, even in the nonseparating regions. The present flow is non-separating, and with suitable tuning, it is possible that RANS could yield performance that is similar to LES but at much lower cost. If the present model were intended as a design tool, the RANS model would then be clearly preferable. However, our purpose here is to compute a highly reliable turbulent flow field for the physical (mechanistic) study of wavelet formation, and we believe LES/IBM to be clearly more credible for such a purpose. Furthermore, it is planned to proceed to study later stages of bedform development in which flow separation, and 3D beform formation, occurs. Clearly, it will be simpler and more credible to use a common CFD method, namely LES/IBM, that allows comparative analysis of early versus late stages of bedform formation. In the LES method, the flow equations have to be solved in 3D. A 2D model for the bed evolution facilitates the computation considerably, and should be sufficient because it is observed in experiments3)5) that the sand-wavelets at low particle Reynolds number are essentially two-dimensional. To the authors’ knowledge, this is the first numerical study on the initiation of sandwave from a flat bed with a highly reliable solution of the turbulent flow field. 2. NUMERICAL METHOD (1) LES and IBM method for unsteady hydrodynamic equations The governing equations for LES are non-dimensionalized by the total flow depth H and mean friction velocity : 0 (1a) % !"#$ &' (1b) , where (̃ ' is the subgrid stress; *+ .- is the flow Reynolds number based on the mean friction velocity and fluid viscosity ν; &' is the artificial body-force accounting for the presence of solid bodies. /0 and /̃ indicate filtered and non-dimensionalized quantities, respectively. The subgrid stress model is the Shear-Improved Smagorinsky model proposed by Leveque et al11): 0 (̃ ' 2' (̃ 33 256 7' (2) 1 :, <̃= 8>? ∆=A 8B7089 :, <̃=B BC7089 :, <̃=DB=, With 56 89 0 89 :, <̃= E 89 :, <̃= 89 :, <̃=F, and 7' A 0 7' 0 = :, <̃=B 827' B7089 .A . where Cs=0.16 and ∆ G∆ ∆H ∆I J 1 . Here, K, KH, KI are local grid spacings in the stream x, vertical y, span z directions, respectively. C. D denotes ensemble average which is evaluated along the homogenous, z, direction. This model was selected because of its simplicity, gentle restriction on time step, moderate computational cost, with performance that is comparable to the dynamic Smagorinsky model11). The artificial body-force is evaluated by the direct forcing method12),13). Eq. (1b) is time-discretized as: MN M ' ' K<G*,7 &' J (3) where K< is the computational time step; RHS contains the advection, pressure, and viscous term. To impose the desired value OP ' within the solid body, the artificial body-force should be: QR S &' *,7 (4) T inside the flow region occupied by the solid body and zero elsewhere. Usually, the solid surface does not coincide with the grid points. In this case, &' at the grid point closest to the solid surface but inside the solid body is linearly interpolated between the value that would yield OP ' inside the solid body and the value of zero inside the fluid portion. &' is evaluated at every time step. This method is shown to have 2nd-order accuracy, and the condition for pressure is satisfied automatically on the solid surface. Eq.(1-2) are solved by a finite difference method on a fixed Cartesian, staggered non-bodyfitted grid with 4th order central discretization in space and 2nd order - 164 - is approximately four times denser in x direction than that of the DNS case (64×124×64). However, Balaras8) showed that, even though the IBM method requires higher grid resolution than the body-fitted method, using the IBM with a Cartesian solver is still more cost-efficient in computational time. The LES and DNS have the same periodic boundary conditions in the stream x and span z directions, and no-slip condition on the lower surface while they have different boundary condition on the upper surface: free slip in the LES+IBM case, to approximate the free surface, and rigid in the DNS case. The bulk Reynolds number, Re=Ub*(H-α)/ν, where Ub is the bulk velocity; H is the total flow depth; α is the wave amplitude, in both cases are matched to about 4300. All the compared quantities have good agreements on the lower boundary, and because of different boundary conditions on the upper surface, they differ slightly far away from the lower wall. Grid refinement tests of the LES code have been done with two different grid resolutions and excellent agreements have been observed. The agreements between the LES+IBM and the DNS results show the accuracy of the present hydrodynamic model. Fig.1 Comparison between the present LES results with DNS data by Ohta et al15). From top to bottom: mean streamwise velocity, mean vertical velocity, Reynolds stress, turbulent kinetic energy scaled by the bulk velocity Ub; (○) DNS, (---) LES (128×72×64), (━)LES (256×128×64). Flow from left to right. Wave amplitude α=0.05*(H-α). Wave length=3.84*(H-α). Adams-Bashforth method for time marching14). The grid is uniform in the stream x and span z directions. On the vertical y direction, grid is equal-spaced until the highest point of the solid surface and then a hypertangential distribution is used up to the upper boundary. The boundary conditions are periodic in the x and z directions, and non-slip and free-slip on the lower and upper surface, respectively. To validate the LES+IBM code, a test case of flow over a sinusoidal bed has been conducted and the results compared to DNS data by Ohta et al15) as shown in Fig. 1. The computational domain in the present simulation is one wavelength in the stream direction, while it is four wavelengths in the DNS case, and half a wavelength in the span direction, which is the same as the DNS case. The grid resolution, over one wavelength, in the present simulation (256×128×64) (2) Model for bedload transport For the present range of particle Reynolds number, Rep<3, the viscosity influence is large, dynamics of individual particles can be neglected and the granular medium can be treated as a continuum1). Therefore, change of bed profile is corresponding to shear stress distribution on the bed surface, which determines the sediment flux. Moreover, Coleman & Melville3), Coleman et al4), and Langlois & Valance5) reported that formation of sand-wavelets is mainly due to bedload transport. Therefore, only bedload transport is included in the model. The bedload equation by Van Rijn16) has been selected because it takes particle Reynolds number into account and covers the range of Rep≤5. For consistency with the hydrodynamic equations, equations in this section are also non-dimensionalized with total flow depth H and mean friction velocity . Van Rijn16) employed two non-dimensional parameters. The particle parameter is U V *+XA Y W$ .1 \ A W$ ] A Z [ ^ *+ _ .1 (5a) with *+X ⁄- the particle Reynolds number, and the transport stage parameter is b a 1.0 (5b) bc with de the critical Shields parameters, and # A ^ d #$ (5c) .A \ A 6 %.j Z [ ^ _ 0.053 $.k \U W$ ] (5d) % d 8?S #=f\ [ is the local Shields parameter obtained from local friction velocity uτ. Then his equilibrium bedload flux equation becomes: - 165 - g"h Gravity effects are accounted as2),19): ye ye E1 25 λ 15 10 5 0 *+ 1.5 0.5 -0.5 -1.5 -2.5 Fig.2 Sand-wavelet developing from plane bed with an initial perturbation of one sinusoidal wave for the test case of Rep=1.0; Flow from left to right. From bottom: t̃=0.0, 4.5, 8.5, 12.5… the sand-wavelet is marked with the dash rectangular at t̃=16.5 with length, indicated by the arrows, of λ/d≃420. The dashed line highlights the development path of the sandwave. Vertical dimensions are stretched for legibility. At left, time development of y+max and y+min around the sandwave is shown. % #$ 8?S =f\ is the mean Shields number, s is the ratio of grain density to fluid density, g is the gravitational acceleration. It is assumed that the local bedload flux does not adapt instantaneously to the local bed shear stress17), hence there is a lag time, or equivalently, a lag distance between the two. This lag distance is shown to play a central role on formations of sandwaves on the bed17). The local, non-equilibrium, or lagged, bedload flux,g is obtained via the lag distance, or the non-equilibrium adaption-length m"h , through a relaxation law18),19): h n hShop qop (6a) Following Philips & Sutherland18) and Bui et al19), the non-equilibrium adaption-length m"h is taken as the saltation length proposed by Van Rijn16): m"h 3[.,^U.r a .s (6b) Variation of bed profile is calculated by Exner-Polya equation1): 81 t= u vh vn (8) u g"h g"h [1.0 ^ (9) ye : critical Shields parameter for zero-slope bed; determined from White’s experiment20), and }? : friction angle, average value of tan}? 0.63 as quoted in Richards21). Eq. (6a) and (7) are also solved by the finite difference method on the same grid as Eq.(1-2) with 2nd order accuracy in space and time. It is assumed that the time scale of the flow development is much smaller than that of the bedform development21). Moreover, for Rep≤2.5, the bed can be treated as smooth, fixed one when the hydrodynamic equations, Eq.(1a,1b) are solved1),7)22). The local friction velocity is then calculated from the shear stress, as averaged in the span z direction during an interval ∆a of 50 or 100 LES time steps. 20 where d . u F zM{| (7) Where w is the bed elevation; n is the bed porosity, taken to be 0 in this model, since the granular medium is treated as a continuum one; and x is the bed surface tangential direction. # ] *+ : (10) where n is the velocity component along the bed surface; is the surface-normal direction. Eq. (10) is : averaged from N #$ 2 4 After the Eq.(1-2) are solved and the bed shear stress is obtained by Eq.(10), the bedload flux is computed by Eq. (5,6,8,9) and bed profile is updated by Eq.(7). At the next step, the Eq.(1-2) are solved over the new bed profile, and the process is repeated until the desired bedform (sand-wavelet) has been observed. The interval ∆a is necessary to assure that the flow have enough time to adapt to the new bed profile. 3. RESULTS AND DISCUSSIONS Fig. 2 shows an example of the evolution of the initial bed profile with one sinusoidal wave occupying half of the domain, and the maximum and minimum bed elevation. From the crest of the initial wave, a small bump appears and grows in height, while its two ends are not clear until the height saturates at a certain value. The wave crest becomes longer in time until the downstream end of the wave starts being eroded strongly, with the scoured sediment deposited further downstream to form a new wave. At this instant, two troughs fore and aft of the wave can be specified, and the wave is identified as sand-wavelet, consistent with the wave-detection method by Coleman et al4). The trough-to-trough length, i.e. wavelength λ indicated by arrows in Fig.2, of this sand-wavelet is found to be independent of the wavelength and height of the initial sinusoidal wave on the initial bed profile provided that the disturbance produced by the initial bed profile is strong enough to help the bedform evolve. Due to the high computational cost of the 3D hydrodynamic equations, the computational domain is limited. Accordingly, further development of the - 166 - Fig.3 Computed wavelets’ wavelengths vs. reported experimental data for turbulent flow. bedform has not been computed. The top of the wave is seen to be quite symmetric fore-and-aft, which is similar to the bed profile in Figure 4 of Coleman et al4) for runs with Rep=1.5→ 2.1. Tests with different values of the mean Shields number d show that the wavelength at a specific value of Rep is independent of d ; Higher d only shortens the time for the sand-wavelet to appear. In addition, as in the case of ripples1), formation of sand-wavelets should not be dependent on the flow depth which, however, appears in the bedload equations through the non-dimensionalizing process. However, changing H does not change the results significantly. For example, at Rep=1.0, tests with H/d=300 and 500 yield λ/d=419 and 428, respectively. The simulations were run with values of Rep from 0.5 to 2.5 and the wavelengths of the sand-wavelets obtained at each Rep are plotted in Fig. 3, together with the reported experimental ones. That the computed points are distributing on the lower limit of the experimental ones shows the agreement between them. In the simulation results, the sand-wavelets are caught when they just appear, hence their wavelength should be corresponding to the smallest observed in the experiments. In experiments, the wavelength is averaged in the whole observed field. Therefore, it is possible that second-generation waves (induced by the first waves downstream) are also taken into account, and the averaged wavelength may be greater than that of the first-generation sand-wavelets that are obtained by simulation. Sumer et al22) used linear stability analysis, consisting of a linearized model for hydrodynamic equations and Bagnold’s bedload formula for sediment transport, in the hydrodynamically smooth regime (Rep<4.0) and found that the fastest growing wavelength is dependent on the friction angle }? : 220 ⁄- 610 for tan}? 0.32 0.75 and for the average value of tan}? 0.63, #$ A1 273 (11) \ !" As shown in Fig.3, the wavelength can be fitted as: \ !" present computed (12) Obviously, from Eq. (11) and (12), the results from Fig.4 Distribution of friction velocity (above) around small bumps (below). Bulk Reynolds number Re=5000. the two methods differ only by the prefactor. Since we are at present short of factors that determine }? , such as the particle properties, the flow properties, tests with other values of }? have not been done. Yalin7) gives an empirical formula for developed ripple for Rep ≤2.5: AA (13) \ !" Since Eq.(13)= constant×Eq.(12), it seems likely that the formation of the sand-wavelets (initial bedform) and the ripples (mature bedform) should be related to similar origins in this range of Rep. According to Yalin1),7), one possible origin is the instability of the bed surface. Incidental discontinuities on the bed surface can cause the derivative of bed shear stress (⁄ , hence g ⁄ to deviate from zero and the bed starts growing in time following Eq. (7). This hypothesis is consistent with the observations by Coleman & Melville3) and Coleman et al4) that the sand-wavelets are instigated by random pileups of sediment (see Section 1). In the present simulations, the height of the first waves when its top stops growing is wN w ⁄- 1.5 2.0 . This wave can be considered as the ‘discontinuity’ mentioned by Yalin1),7). To understand effects of such a discontinuity on the bed shear stress distribution, flows over bumps on a flat bed with heights wN 1.0 2.0 and with different bump lengths were solved with the LES+IBM method, Eq.(1-2), and are shown in Fig. 4. As expected, the peak of the friction velocity distributions are always before the peak of the bumps, i.e. there is a phase lag between the two. Although the bump does not cause flow separation, i.e. the friction velocity is always positive, there is a region, or ‘shadow’ with length Lx+, where the bed shear stress deviates from the standard value. Interestingly, Lx+ is almost independent of the bump size, and is about 300 which is similar to the sand-wavelet’s wavelength, Eq. (12). Therefore, the sand-wavelets appear to be formed adapting to the length of this ‘shadow’, and hence, - 167 - Yalin’s1),7) postulation is confirmed. Note that the ‘shadow’ length is from the top of the bump while the wavelength is measured on the whole wave. In Fig. 4, the longest bump has the gentlest gradient of the friction velocity on the top but the highest gradient at the downstream-end point. This may explain for observations in the simulations as presented in Fig. 2. Initially, disturbances caused by the initial bed profile are responsible for the appearance of the first small wave. This wave continues growing in both height and length because of sharp shear stress gradient on its top and downstream side, similar to the smallest bump in Fig. 4. When its length becomes larger, the shear stress gradient on its top becomes gentler until is not strong enough to help the wave gain in height while it can still gain in length because the downstream shear stress gradient is becoming stronger. When the wavelength is equivalent to that of the ‘shadow’, the shear stress gradient at the downstream end of the wave is strongest, and causes the strong erosion which help to identify the sand-wavelet. 4. SUMMARY Formation of sand-wavelets in hydrodynamically smooth flow over an erodible bed has been studied numerically. The non-dimensional wavelengths vary as a function of particle Reynolds number as ⁄ 400/*+X up to Rep=2.5. Agreement with experimental and analytical results in the literature has been obtained. Similar to the case of ripples, the sand-wavelets are seen to develop adapting to the disturbance of the bed shear stress caused by the discontinuities on the bed surface, which is consistent with the postulation by Yalin1),7) and the observations by Coleman & Melville3) and Coleman et al4). ACKNOWLEDGMENT: We are grateful for support from the Monbukagakusho Ministry for a scholarship for the first author. Dr. Dinh Xuan Thien at CFD laboratory, Ritsumeikan University is appreciated for giving the initial LES code and for helpful discussions. Professor Y. Shimizu and Dr. S. Giri at Hokkaido University are also thanked for useful discussions. The anonymous reviewers are grateful for helping to improve the quality of this paper. REFERENCES 1) Yalin. M. S.: Mechanics of Sediment Transport, 2nd edition, Pergamon Press, 1977. 2) Fredsoe J., Deigaard, R.: Mechanics of coastal sediment transport, Advanced series on ocean engineering, Vol.3, World Scientific, 1994. 3) Coleman, S.E., Melville, B.W.: Initiation of bed forms on a flat sand bed, Journal of Hydraulic Engineering, Vol. 122, No.6, 301-310, 1996. 4) Coleman, S.E., Fedele, J.J., Garcia, M.H.: Closed-conduit bed-form initiation and development, Journal of Hydraulic Engineering, ASCE, Vol. 129, No.12, 956-965, 2003 5) Langlois, V., Valance, A.: Initiation and evolution of current ripples on a flat sand bed under turbulent water flow, The European Physical Journal E, Vol. 22, 201-208, 2007. 6) Kuru, W.C., Leighton, D.T., McCready, M.J.: Formation of waves on a horizontal erodible bed of particles, International Journal of Multiphase Flow, Vol. 21, No. 6, 1123-1140, 1995. 7) Yalin, M.S.: On the determination of ripple length, Journal of Hydraulics Division, ASCE, Vol. 103 (HY4),439-442, 1977. 8) Balaras E.: Modeling complex boundaries using an external force field on fixed Cartesian grids in large-eddy simulations, Computers & Fluids, Vol. 33, 375-404, 2004. 9) Keylock C.J., Hardy R.J., Parsons D.R., Ferguson R.I., Lane S.N., Richards, K.S.: The theoretical foundations and potential for large-eddy simulation (LES) in fluvial geomorphic and sedimentological research, Earth-Science Reviews, Vol. 71, 27-304, 2005. 10) Chang Y. S., Scotti A.: Modelling unsteady flows over ripples: Reynolds-averaged Navier-Stokes equations (RANS) versus large-eddy simulation (LES), Journal of Geophysical Research, Vol. 109, C09012, 2004. 11) Leveque, E., Toschi, F., Shao, L., Bertoglio, J.P.: Shear-improved Smagorinsky Model for Large-eddy Simulation of Wall-bounded Turbulent Flows, Journal of Fluid Mechanics, Vol. 570, 491-502, 2007. 12) Fadlun, E.A., Verzicco, R., Orlandi, P., Mohd-Yusof, J.: Combined Immersed-Boundary Finite-Difference Methods for Three-Dimensional Complex Flow Simulation, Journal of Computational Physics, Vol. 161, 35-60, 2000. 13) Verzicco, R., Mohd-Yusof, J., Orlandi, P., Haworth, D.: Large eddy simulation in complex geometric configurations using boundary body forces, AIAA Journal, Vol. 38, Issue 3, 427-433, 2000. 14) Geurts B. J.: Elements of Direct and Large-Eddy Simulation, Edwards, 2004. 15) Ohta, T., Miyake, Y., Kajishima, T.: Direct numerical simulation of turbulent flow in a wavy channel, JSME International Journal, Ser. B, Vol. 41, No. 2, 447-453, 1998. 16) Van Rijn L. C.: Sediment transport, Part I: Bed load transport, Journal of Hydraulic Engineering, Vol. 110, No. 10, 1431-1456, 1984 17) Kennedy J.F.: The mechanics of dunes and antidunes in erodible-bed channels, Journal of Fluid Mechanics, Vol. 16, 521-544, 1963. 18) Phillips B.C. and Sutherland, A.J.: Spatial lag effects on a bed load sediment transport, Journal of Hydraulic Research, Vol. 27, No.1, 115-133, 1989. 19) Bui M. D., Wenda, T., Rodi W.: Numerical modeling of bed deformation in laboratory channels, Journal of Hydraulic Engineering, Vol. 130, No.9, 894-904,2004. 20) Raudkivi, A.J.: Loose Boundary Hydraulics, 3rd edition, Pergamon Press, 1990. 21) Richards, K.J.: The formation of ripples and dunes on an erodible bed, Journal of Fluid Mechanics, Vol. 99, part 3, 597-618 22) Sumer B.M., Bakioglu, M., Bulutoglu, A.: Ripple formation on a bed of fine, cohensionless, granular sediment, Euromech 156: Mechanics of Sediment Transport, 99-10, 1982. 23) Giri, S., Shimizu, Y.: Numerical computation of sand dune migration with free surface flow, Water Resources Research, Vol. 42, W10422, 2006. 24) http://www-fluid.mech.eng.osaka-u.ac.jp/~ota/ - 168 - (Received September 30, 2007)