The Role of Splatting Effect in High Schmidt Number Turbulent Mass Transfer Across an Air-Water Interface

Similar documents
LARGE EDDY SIMULATION OF MASS TRANSFER ACROSS AN AIR-WATER INTERFACE AT HIGH SCHMIDT NUMBERS

The Effect of Schmidt Number on Air-Water Interface Mass Transfer

SECONDARY MOTION IN TURBULENT FLOWS OVER SUPERHYDROPHOBIC SURFACES

INFLUENCE OF THE BOUNDARY CONDITIONS ON A TEMPERATURE FIELD IN THE TURBULENT FLOW NEAR THE HEATED WALL

Available online at ScienceDirect. Procedia Engineering 90 (2014 )

Transport of a passive scalar at a shear-free boundary in fully developed turbulent open channel flow

RECONSTRUCTION OF TURBULENT FLUCTUATIONS FOR HYBRID RANS/LES SIMULATIONS USING A SYNTHETIC-EDDY METHOD

Very large-scale structures observed in DNS of turbulent channel flow with passive scalar transport

DNS STUDY OF TURBULENT HEAT TRANSFER IN A SPANWISE ROTATING SQUARE DUCT

Turbulent boundary layer

Contribution of Reynolds stress distribution to the skin friction in wall-bounded flows

Principles of Convection

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Numerical simulations of heat transfer in plane channel flow

An evaluation of a conservative fourth order DNS code in turbulent channel flow

Tutorial School on Fluid Dynamics: Aspects of Turbulence Session I: Refresher Material Instructor: James Wallace

LES of turbulent shear flow and pressure driven flow on shallow continental shelves.

DIRECT NUMERICAL SIMULATION OF SPATIALLY DEVELOPING TURBULENT BOUNDARY LAYER FOR SKIN FRICTION DRAG REDUCTION BY WALL SURFACE-HEATING OR COOLING

ROLE OF THE VERTICAL PRESSURE GRADIENT IN WAVE BOUNDARY LAYERS

Fluid Dynamics Exercises and questions for the course

Turbulence Modeling I!

Turbulent Heat Transfer in A Channel Flow with Arbitrary. Directional System Rotation

INVESTIGATION OF THE FLOW OVER AN OSCILLATING CYLINDER WITH THE VERY LARGE EDDY SIMULATION MODEL

Direct numerical simulation of turbulence in a sheared air water flow with a deformable interface

On the transient modelling of impinging jets heat transfer. A practical approach

LIQUID FILM THICKNESS OF OSCILLATING FLOW IN A MICRO TUBE

elements remain in high frequency region and sometimes very large spike-shaped peaks appear. So we corrected the PIV time histories by peak cutting an

Numerical Simulation of Rocket Engine Internal Flows

2.3 The Turbulent Flat Plate Boundary Layer

The Johns Hopkins Turbulence Databases (JHTDB)

Chapter 2 Mass Transfer Coefficient

Applied Mathematics and Mechanics (English Edition)

+ = + t x x x x u. The standard Smagorinsky model has been used in the work to provide the closure for the subgridscale eddy viscosity in (2):

Preliminary Study of the Turbulence Structure in Supersonic Boundary Layers using DNS Data

The JHU Turbulence Databases (JHTDB)

Particle dispersion in stably stratified open channel flow

Large-Eddy Simulation for Turbulent Nature of Flow and Pressure Fields over Urban Building Arrays C. Hirose*, A. Hagishima, N. Ikegaya, and J. Tanimot

WALL PRESSURE FLUCTUATIONS IN A TURBULENT BOUNDARY LAYER AFTER BLOWING OR SUCTION

Numerical Methods in Aerodynamics. Turbulence Modeling. Lecture 5: Turbulence modeling

The JHU Turbulence Databases (JHTDB)

Engineering. Spring Department of Fluid Mechanics, Budapest University of Technology and Economics. Large-Eddy Simulation in Mechanical

Anisotropic grid-based formulas. for subgrid-scale models. By G.-H. Cottet 1 AND A. A. Wray

Observations of Giant Bursts Associated with Microscale Breaking Waves

Numerical Simulation of Rotating Turbulent Channel Flow by the Vortex in Cell Method

Numerical analysis of a fully developed non-isothermal particle-laden turbulent channel flow

Lagrangian statistics in turbulent channel flow

Soft Bodies. Good approximation for hard ones. approximation breaks when objects break, or deform. Generalization: soft (deformable) bodies

Divergence free synthetic eddy method for embedded LES inflow boundary condition

Large eddy simulation of turbulent flow over a backward-facing step: effect of inflow conditions

ROLE OF LARGE SCALE MOTIONS IN TURBULENT POISEUILLE AND COUETTE FLOWS

Basic Features of the Fluid Dynamics Simulation Software FrontFlow/Blue

Open boundary conditions in numerical simulations of unsteady incompressible flow

Convective Mass Transfer

Direct numerical simulation of interfacial instability in gas-liquid flows

DNS, LES, and wall-modeled LES of separating flow over periodic hills

Some remarks on grad-div stabilization of incompressible flow simulations

There are no simple turbulent flows

Large Eddy Simulation as a Powerful Engineering Tool for Predicting Complex Turbulent Flows and Related Phenomena

Analysis and modelling of subgrid-scale motions in near-wall turbulence

Hybrid LES RANS Method Based on an Explicit Algebraic Reynolds Stress Model

Direct and Large Eddy Simulation of stably stratified turbulent Ekman layers

CHAPTER 7 NUMERICAL MODELLING OF A SPIRAL HEAT EXCHANGER USING CFD TECHNIQUE

Turbulence: Basic Physics and Engineering Modeling

ME 144: Heat Transfer Introduction to Convection. J. M. Meyers

LES modeling of heat and mass transfer in turbulent recirculated flows E. Baake 1, B. Nacke 1, A. Umbrashko 2, A. Jakovics 2

1. Introduction, tensors, kinematics

BOUNDARY LAYER ANALYSIS WITH NAVIER-STOKES EQUATION IN 2D CHANNEL FLOW

A NOVEL VLES MODEL FOR TURBULENT FLOW SIMULATIONS

On the feasibility of merging LES with RANS for the near-wall region of attached turbulent flows

arxiv: v1 [physics.flu-dyn] 11 Oct 2012

FLOW-NORDITA Spring School on Turbulent Boundary Layers1

Fluid Dynamics: Theory, Computation, and Numerical Simulation Second Edition

Modeling Complex Flows! Direct Numerical Simulations! Computational Fluid Dynamics!

Direct Numerical Simulation of the Neutrally Stratified Turbulent Ekman Boundary Layer

J OURNAL OF TURBULENCE. JoT 5 (2004) 020. Large-eddy simulation of heat transfer downstream of a backward-facing step

Turbulent eddies in the RANS/LES transition region

Publication 97/2. An Introduction to Turbulence Models. Lars Davidson, lada

Numerical Heat and Mass Transfer

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders

Numerical simulation of dispersion around an isolated cubic building: Model evaluation of RANS and LES. Yoshihide Tominaga a and Ted Stathopoulos b

IHTC DRAFT MEASUREMENT OF LIQUID FILM THICKNESS IN MICRO TUBE ANNULAR FLOW

A Low Reynolds Number Variant of Partially-Averaged Navier-Stokes Model for Turbulence

Numerical simulation of wave breaking in turbulent two-phase Couette flow

Estimation of Turbulent Dissipation Rate Using 2D Data in Channel Flows

Detailed 3D modelling of mass transfer processes in two phase flows with dynamic interfaces

Turbulence Laboratory

Boundary-Layer Theory

Probability density function (PDF) methods 1,2 belong to the broader family of statistical approaches

An investigation of turbulent oscillatory heat transfer in channel flows by large eddy simulation

FEDSM TURBULENCE MODULATION BY AN ARRAY OF LARGE-SCALE STREAMWISE STRUCTURES OF EHD ORIGIN

PIV STUDY OF LONGITUDINAL VORTICES IN A TURBULENT BOUNDARY LAYER FLOW

Accommodating LES to high Re numbers: RANS-based, or a new strategy?

A Discussion on The Effect of Mesh Resolution on Convective Boundary Layer Statistics and Structures Generated by Large-Eddy Simulation by Sullivan

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

Chapter 7 The Time-Dependent Navier-Stokes Equations Turbulent Flows

model and its application to channel ow By K. B. Shah AND J. H. Ferziger

MEASUREMENTS OF TIME-SPACE DISTRIBUTION OF CONVECTIVE HEAT TRANSFER TO AIR USING A THIN CONDUCTIVE-FILM

Turbulent momentum and heat transfer in ducts of rhombic cross section

On Decaying Two-Dimensional Turbulence in a Circular Container

Pressure-velocity correction method Finite Volume solution of Navier-Stokes equations Exercise: Finish solving the Navier Stokes equations

Transcription:

Turbulence, Heat and Mass Transfer 4 K. Hanjalic, Y. Nagano and M. Tummers (Editors) 3 Begell House, Inc. The Role of Splatting Effect in High Schmidt Number Turbulent Mass Transfer Across an Air-Water Interface Y. Hasegawa and N. Kasagi Department of Mechanical Engineering, The University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-8656, Japan Abstract- High Schmidt number gas absorption into liquid across an air-water interface is simulated by means of a hybrid DNS/ES scheme. The results are well compared with those previously obtained by a agrangian method. Due to the wind shear, quasi-streamwise vortices are generated beneath the interface. They induce upwelling flows which impinge on the interface, i.e., splatting. At high Schmidt numbers, the concentration field near the interface is characterized by low concentration spots, which occur intermittently associated with the splatting events, and fine-scale high concentration streaks produced by the interfacial shear. With an assumption that the surface renewal period should be equal to the typical time scale of the quasi-streamwise vortex, the gas transfer rate is well predicted. Therefore, it is the quasi-streamwise vortex scaled by the shear units that governs the mass transfer across a shear-driven air-water interface. 1. Introduction Gas exchange across the sea surface plays a crucial role in the global carbon cycle on the earth, and better understanding of the mechanism controlling the interfacial mass transfer is indispensable to improve the model prediction of the global air-sea CO fluxes. For slightly soluble gases such as carbon dioxide, most mass transfer resistance occurs in liquid. Additionally, with the Schmidt number of about 1, the transport process is governed by near-surface turbulence within a thin concentration boundary layer ( ~ mm). Since the simultaneous measurement of the velocity and scalar fields is extremely difficult in experiments, the mechanism of the mass transfer is not understood well. When predicting the gas transfer rate K, the time scale of vortices which govern the mass transfer has been a subject of discussion. For instance, if one selects the time scale of the large eddies scaled by l / u *, one obtains K µ Sc -1/ Re* -1/, while K µ Sc -1/ Re* -1/4 by assuming the transfer is governed by the small eddies responsible for the dissipation. Here, Re* =u * l / n and l is the integral length scale. In the last decade, direct numerical simulation (DNS) has been applied to free-surface flows by several groups [1-3], and the complex interaction between a free surface and turbulent structures, and the associated transport process have been studied. However, since the number of grid points for resolving the scalar field increases in proportion to Sc 3/, the Schmidt number of scalar has been limited to Sc ~ O(1) in most of the previous calculations. In this study, DNS of a coupled air-water turbulent flow is carried out in order to investigate the structures of the scalar field and microscopic mechanism of the mass transfer near the interface. A hybrid DNS/ES scheme, in which DNS is applied in a near-surface region by concentrating grids, while ES in the outer region with relatively coarse grids, is proposed to calculate the scalar field at the high Schmidt number of Sc = 1. The obtained gas transfer rate is well compared with the agrangian result and the detailed mass transfer mechanism is discussed. 113

4 th Int. Symp. on Turbulence, Heat and Mass Transfer free slip c = 1. δ air δ water πδ.5πδ free slip c = zg xg yg y x z Fig. 1: Computational domain and coordinate system DNS (y < 16.5) Buffer layer (16.5 < y < 1.5) ES ( > 1.5) air-water interface Fig. : Grid system for solving scalar field with hybrid DNS / ES scheme. Numerical Methodology. 1 Computational Condition A fully developed air-water counter flow driven by a constant pressure gradient and the associated interfacial mass transfer are considered as shown in Fig. 1. The governing equations in the two phases are the incompressible Navier-Stokes, continuity, and scalar transport equations, i.e., = - ui ui p 1 ui ui uj t x j xi Re t xj x, = j x, (1) i c c 1 c uj = t xj Sc Re xj x, () t j where u i is i-th velocity component, x i the coordinates, c the scalar concentration. Subscripts 1,, and 3 correspond to streamwise, vertical and spanwise directions, respectively. All variables are normalized by the friction velocity u t and the depth d in each phase. The Reynolds number and the Schmidt number are defined as Re t = u t d /n and Sc = n / D, where n and D are the kinematic viscosity and the molecular diffusivity of scalar, respectively. In this study, the Reynolds numbers in the two phases are set to be Re t = Re t G = 15, which corresponds to the air-water flow at the wind velocity of m/s and the phase depth of 4 cm under the standard condition. The Schmidt number is changed from 1. to 1 in liquid, while kept constant Sc G = 1. in gas. Since it is known that the interfacial deformation is negligibly small in the low Reynolds number flow considered here, the interface is assumed to be flat. Therefore, the resultant boundary conditions at the interface are the continuity of the horizontal velocities and shear stress for the velocity field, and the Henry s law and the continuity of the scalar flux for the scalar field. They are given as: u = R u i r ig, u = ug =, (3) u x i = uig Rn x, (i = 1, 3) (4) G c c = a cg, = R R Sc cg r n x ScG x. (5) G Here, a is dimensionless Ostwald s solubility. The two non-dimensional parameters of R r and 114

Y. Hasegawa and N. Kasagi Table. 1: Number of modes k x, k z and grid points N y, and their spacing Dx, Dy and Dz for three regions in hybrid DNS/ ES scheme. Region k x, N, k Dx y z Dy Dz D NS < y < 16. 5 51, 95, 51. 3. ~.34. 9 B uffer 16.5 < y < 1. 6 51, 14, 51. 3.35 ~.38. 9 ES 1.6 < y < 15 64, 34, 64 18. 4. ~.38 7. R n in Eqs. (4) and (5) represent a degree of dynamical coupling between the two fluids, and are defined as: - Rr = r / rg, Rn =( ng / n) Rr 1. (6) In this study, R r and R n are set to be 9 and 1., respectively. At the outer boundaries, a freeslip condition is imposed for the velocity field in order to minimize the effect of the assumed outer boundary condition, and a constant concentration condition for the scalar fields, i.e., c = 1. and at the top and bottom boundaries, respectively.. Numerical Scheme A pseudo-spectral method with Fourier expansions in the x and z directions and Chebyshev polynomials in the y direction, is employed for spatial discretization of the velocity and scalar fields at Sc = 1., 3. and 5.. The numbers of modes employed for the velocity field in the three directions are (k x, k y, k z ) = (64, 88, 64) and their spacings are (Dx, Dy, Dz ) = (18.4,.45 ~.8, 7.). Time advancement is made by the second-order Adams-Bashforth method for the convective terms and the Crank-Nicolson method for the diffusion terms. The details of the numerical algorithm employed are essentially the same as those of ombardi et al. [1]... 1. Hybrid DNS / ES scheme In order to calculate the concentration filed at the Schmidt number of Sc = 1, we propose a hybrid DNS / ES scheme in which DNS with high-resolution grids is applied within the nearinterface region, while ES with coarser grids for the outer layer (see, Fig. ). For spatial discretization, Fourier expansions are used in the x and z directions, while the second-order finite difference scheme is employed in the y direction in order to avoid complexity of numerical procedures. The computational domain in liquid is divided into three regions, i.e., DNS region ( < < 16.8), buffer region (16.8 < y < 1.5), and ES region (y > 1.5). The numbers of modes and grid points employed in each region are listed in Table. 1. Since the finer grid system is employed for the scalar field in the DNS region, the fluid velocities should be appropriately interpolated in order to evaluate the convective terms accurately. In this study, the Fourier-Chebyshev interpolation scheme, in which fast Fourier and Chebyshev transforms are used by setting the components of higher wave numbers to be zero, is employed. In the outer layer, the same grid system as that of the velocity field is employed. Since the grids are not fine enough to resolve the scalar dissipation, the constant Smagorinsky model is introduced in order to dissipate the scalar fluctuation appropriately. This results in: 115

4 th Int. Symp. on Turbulence, Heat and Mass Transfer 3. 3..5 u i rms / u t. 1. G u i rms / u t. 1.5 1..5. 1 3 4 5 6 7. 1 3 4 5 6 7 c c 1 c = x j uj t xj Sc Re t xj xj x, (7) j c x j = -D s D S ij x, D = ( Dx Dz) 1 /. (8) j The model coefficient D s is determined by the previous a priori test data for a low Schmidt number, since it was found that the model coefficient is almost independent of the Schmidt number near the interface [5]. It is possible to use more complex models such as dynamic Smagorinsky model or mixed model to improve the model prediction. It should be noted, however, that most change of the mean concentration occurs in the DNS region (see, Fig. 4) and therefore, the effect of the SGS model on the total scalar transfer is insignificant. As is well known, employment of a hybrid scheme such as detached eddy simulation, frequently leads to a mismatch of the mean profile at the connection point between two schemes. In order to avoid this, we establish a buffer region between the two regions, where the SGS model is gradually imposed as going away from the interface by increasing the Smagorinsky constant.... agrangian method We also applied a agrangian method to the mass transfer at the high Schmidt numbers from 1 to 1. In this method, the concentration field is represented by the probability density function of a large number of scalar markers, which move due to the convective and molecular effects. The details of the numerical scheme can be found in Papavassiliou and Hanratty [4]. 3. Results Fig. 3: Velocity fluctuations in a) liquid and b) gas. Present study; : u rms, : v rms, : w rms, ombardi et al.; : u rms, : v rms, : w rms, Channel; : u rms, : v rms, : w rms. 3. 1 Velocity field The velocity fluctuations in gas and liquid are shown in Fig. 3 as a function of the distance from the interface. In liquid, the horizontal velocity fluctuations u rms and w rms have peaks at their interface, while those in gas are decreased with approaching the interface. These differences result from a high density ratio R r of the two fluids in Eq. (4). Hence, the interface is similar to a solid wall for gas, while a free surface for liquid. However, coherent structures scaled by the shear units, such as low and high speed streaks and quasi-streamwise vortices, are observed in both phases despite the difference in the effects of the boundary. Since the present paper focuses on the scalar field, the statistics of the velocity fields are referred to the previous work [1]. y G 116

Y. Hasegawa and N. Kasagi y G 1-1 - -3.4 Gas.5.6.7 c mean.8 iquid DNS Buffer ES.9 1. Fig. 4: Mean concentration profiles. : Sc = 1., : Sc = 3., : Sc = 5., : Sc = 1. c rms / DC.5..15.1.5. a) 1. b) 1. 5 1 Fig. 5: Concentration fluctuation profiles in liquid. For symbols see Fig. 4. 15.8.8 R cu.6.4 R cv.6.4... 4 6. 8 1 1 14 4 6 8 1 1 14 Fig. 6: Correlation coefficient a) R cu and b) R cv. For symbols see Fig. 4. 3. Scalar field 3.. 1. Eulerian analysis The mean concentration profiles for different Schmidt numbers near the interface are shown in Fig. 4. In all cases, the change of the mean concentration mainly occurs in the liquid phase, and the thickness of the concentration boundary layer rapidly decreases with increasing the Schmidt number. At Sc = 1, more than 95% of the mean concentration change occurs in the DNS region ( < 16.8), and a smooth profile without a mismatch at the connection between the DNS and ES regions is obtained by the hybrid scheme. In Fig. 5, the concentration fluctuations in liquid are shown. Although the location of the peak approaches the interface in inverse proportion to Sc 1/, the concentration fluctuation outside the DNS region is not negligible even at Sc = 1, which indicates the importance of coupling between the concentration fields in the DNS and ES regions. In Fig. 6 a) and b), the correlation coefficients R cu, R cv between the concentration c and the streamwise and normal velocities u, v are shown. Although the concentration field at higher Schmidt numbers is less analogous to the velocity field far from the interface, the correlation coefficient between the concentration and normal velocity R cv is kept high around.6 near the interface. This suggests that the scalar fluctuation responds to the normal velocity fluctuation quickly even at high Schmidt numbers. This tendency presents a striking contrast to that of R cv near a solid wall [6], in which the scalar fluctuations at higher wave numbers are more damped and the analogy between momentum and mass transfer breaks down. In Fig. 7 a) and b), instan- 117

4 th Int. Symp. on Turbulence, Heat and Mass Transfer a) 35 1. b) q 3 1 β -1 z 5 q 15 a) 5 6 7 8 x 5 6 7 8 x Fig. 7: Distribution of a) interfacial scalar flux q at Sc = 1, and b) surface divergence b and the velocity vectors. 1. 1.8.4 b) -1. - -15-1 -5 5 1 15 z DNS - -4-6 Fig. 8: a) Profiles of interfacial scalar flux q at Sc = 1:, and surface divergence b:, b) Instantaneous scalar field at Sc = 1, and velocity vectors in y-z plane. taneous distributions of the interfacial scalar flux q at Sc = 1 and the surface divergence b are shown. The surface divergence can be interpreted as a degree of replacement of the fluids near the interface with those beneath the interface, i.e., surface renewal, which is defined as: b = u w x z int erface = - v y 118 int erface. (9) In Fig. 7 a), fine-scale streaky structures and large spotty structures with low and high interfacial scalar flux q, are observed. The spotty high scalar flux regions coincide with high surface divergence regions in Fig. 7 b), where the low concentration fluid impinges to the interface. These spotty structures are observed in both low and high Schmidt numbers, and result in high correlation R cv near the interface (see Fig. 6 b)). In Fig. 8 a) and b), distributions of the interfacial scalar flux q and the surface divergence b along the z direction, and the simultaneous velocity and scalar fields in the y-z plane are shown. The spatial distribution of the interfacial scalar flux exhibits sharp changes, which correspond to the fine-scale streaky structures with low interfacial scalar flux observed in Fig. 7 a). On the other hand, the striking peaks coincide with the high surface divergence regions caused by upwelling flows associated with quasi-streamwise vortices, i.e., splatting. This indicates that the β ES 1. c

Y. Hasegawa and N. Kasagi z z z a) 1 3 4 b) 1 3 4 c) 1 3 4 x x 1 8 6 4 1 8 6 4 K = K / u t 5. u' -5. y G.1 8 7 6 5 4 3 45 3 15-15 -3-45 -45-3 -15 15 3 45 z G or z Fig. 1: SE velocity field around a marker (Sc = 1) which passes downward the plane of = 3. u' Fig. 9: x 1 8 6 4 Distributions of scalar markers near the interface in liquid. a) Sc =.1, b) Sc = 1., c) Sc = 1. splatting penetrates the concentration boundary layer and enhances the interfacial mass transfer. 3... agrangian analysis In the agrangian method, 18 x 18 scalar makers are uniformly disposed at the interface as an initial condition and the time development of the probability density functions of the makers was calculated to investigate the mechanism of scalar transfer from the interface to the bulk in liquid. In Fig. 9, the distributions of the scalar markers at different Schmidt numbers near the interface ( < 3) at t = 1 after release are shown. Since the convective effect governs the movement of the scalar makers at high Schmidt numbers, the markers nearly follow the fluid particles and cluster around low speed streaks where the surface divergence is negative (see, Fig. 9 c)). On the other hand, spotty regions with few particles, caused by up-current of the high speed fluid beneath the interface, are also observed. These results are consistent with those obtained by the hybrid scheme, in which the concentration field near the interface is characterized by the low concentration spots caused by splatting, and the high concentration streaks. In Fig. 1, the typical velocity field around a scalar maker which passes downward = 3 plane at Sc = 1 is obtained by linear stochastic estimation. It is observed that quasi-streamwise vortices cause the strong downward motion near the interface and enhance the mass transfer. 3.. 3. Gas transfer rate Finally, the gas transfer rates at high Schmidt numbers obtained by the hybrid DNS/ES scheme 119 5 6 7 8 1 3 4 5 6 7 8 1 Sc Fig. 11: Gas transfer rates at high Sc. present study; :Eulerian, : agrangian, : Hanratty, : De Angelis et al., : Calmet and Magnaudet.

4 th Int. Symp. on Turbulence, Heat and Mass Transfer and the agrangian method are compared in Fig. 11. The previous experimental data obtained by Hanratty [7], and the numerical data by De Angelis et al. [3] and Calmet and Magnaudet [8] are also plotted. The gas transfer rates K = K / u t at Sc = 1 obtained by the hybrid and agrangian methods are.97 and.95, respectively, and show very good agreement. If we assume that the time scale of the quasi-streamwise vortex should be equal to p l v / u t, where l v = 35 is the diameter of the vortex estimated from the spanwise two point correlation of the normal velocity v (see, also Fig. 1), K =.95 Sc -1/ is obtained by using the surface renewal theory. The quantitative agreement between our results and the model prediction indicates that the surface renewal is governed by the quasi-streamwise vortices in liquid. The gas transfer rate is inversely proportional to Sc 1/, and the qualitative agreement between the present and previous data is fairly good. The underestimation in the present study is mainly due to the low Reynolds number flow considered here, since the normal velocity fluctuation near the interface, i.e., splatting, is enhanced with increasing the Reynolds number. At high Reynolds numbers, the effect of capillary waves should be also taken into account. It will be studied in our future work. 4. Conclusions High Schmidt number gas absorption into liquid across an air-water interface was successfully calculated by the hybrid DNS/ES scheme, suggesting that it is a promising tool for analyzing multiple-scale heat and mass transfer phenomena. Due to the wind shear, coherent turbulent structures are generated in liquid, and quasi-streamwise vortices induce impingement of upwelling flows on the interface, i.e., splatting. The splatting motion causes intermittent low concentration spots near the interface and enhances the interfacial scalar flux by replacing the fluids adjacent to the interface with those beneath the interface. With an assumption that the surface renewal period should be equal to the typical time scale of the quasi-streamwise vortex, the gas transfer rate was well predicted. The results indicate that it is the quasi-streamwise vortex scaled by the shear units that governs the mass transfer across a shear-driven air-water interface. References 1. P. ombardi, V. De Angelis and S. Banerjee. Direct Numerical Simulation of Near-Interface Turbulence in Coupled Gas-iquid Flow. Phys. Fluids, 8-6: 1643-1665, 1996.. R. Nagaosa and R. A. Handler. Statistical Analysis of Coherent Vortices near a Free Surface in a Fully Developed Turbulence. Phys. Fluids, 15-: 375-394, 3. 3. V. De Angelis and S. Banerjee. Heat and Mass Transfer Mechanism at Wavy Gas-iquid Interface. In 1st. Int. Symp. TSFP, pp. 149-154, 1999. 4. D. V. Papavassiliou and T. J. Hanratty. The Use of agrangian Methods to Describe Turbulent Transport of Heat From a Wall. Industrial Eng. Chem. Res., 34: 3359-3367, 1995. 5. A. Mitsuishi, Y. Hasegawa and N. Kasagi. arge Eddy Simulation of Mass Transfer across an Air-Water Interface at High Schmidt Numbers. In 6th. ASME/JSME Thermal Engineering Joint Conference, TED- AJ3-31, 3. 6. Y. Na and T. J. Hanratty. imiting Behavior of Turbulent Scalar Transport Close to a Wall. Int. J. Heat Mass Flow, 43: 1749-1758,. 7. T. J. Hanratty. Effect of Gas Flow on Physical Absorption. In Air-Water Gas Transfer, pp. 1-33, Amer Society of Civil Engineers Publ., 1991. 8. I. Calmet and J. Magnaudet. High-Schmidt Number Mass Transfer Through Gas-iquid Interfaces. Int. J. Heat Fluid Flow, 19: 5-53, 1998. 1