• No results found

Turbulent oscillating channel flow subjected to a wind stress

N/A
N/A
Protected

Academic year: 2021

Share "Turbulent oscillating channel flow subjected to a wind stress"

Copied!
6
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

1 Also at Department of Applied Mathematics, University of Twente, The Netherlands 2 Participates in J. M. Burgers Centre, Research school for Fluid Dynamics, The Netherlands

1 INTRODUCTION

In addition to tidal forcing, the stress exerted on the free surface might play a role in driving the flow in estuary regions. Therefore, we estimate the ampli-tude of both effects based on data for the Wester-schelde estuary. Here, the flow is strongly tidal driven with a typical fluid velocity during high tide of U = 0.2 – 1.0 m/s. The Reynolds number based on the depth (h= 10 m) is Reh =Uh/! = 10

6

– 107 with ! the kinematic viscosity. For this range of Re the flow is turbulent for all phases of the tidal cycle. The estimated maximum wall stress due to the tides is 0.08–1 N/m2. A typical wind velocity of 7 m/s yields a wind stress of 0.17 N/m2. The wind stress and the wall stress are thus of the same order. The turbulent flow in the Westerschelde estuary is char-acterized by a large Keulegan–Carpenter number, KC = 100–600, which indicates that the time scale related to the turbulence is much shorter than the tidal period.

Turbulence is known to accelerate mixing and transport of particles. Hence, it is of relevance to sand sedimentation and dispersion of plankton for-mations. This work is part of a project aimed at modeling dispersion of plankton in geophysical flows. The approach is to first investigate particle dispersion at the smallest turbulent scales. This knowledge can then be used for modeling dispersion

statistics in large-scale geophysical flows. To inquire the type of turbulence in such a geophysical flow we performed large-eddy simulations of an oscillating flow subjected to a wind stress at the free surface. In the following sections we first describe the problem and numerical model in Sec. 2 and present the re-sults in Sec. 3–5. A short discussion of the rere-sults and of future work is given in Sec. 6.

2 PROBLEM DESCRIPTION

In this work we study the turbulent oscillating chan-nel flow subjected to a wind stress by means of large eddy simulations (LES). LES only resolves the larg-est scales in the flow, while the effect of the smaller scales are modelled. The separation of the flow field into large and small scales is achieved by filtering. The velocity field

(

u,v,w

) (

= u1,u2,u3

)

can be

writ-ten as u = ˆu+u"" with the filtered velocity and u "" the small-scale part. Applying the filtering to the Navier–Stokes equation yields

j ij i j j i j j i i x f x x u xi p x u u dt u # # $ + # # # + # # $ = # # + # $ % ! & ˆ ˆ ˆ ˆ ˆ ˆ 2 1 , (1)

where & is the density of the fluid and the fil-tered pressure. The term % represents the stresses ij

Turbulent oscillating channel flow subjected to a wind stress

W. Kramer & H. J. H. Clercx

1

Fluid Dynamics Laboratory2, Department of Physics, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands

V. Armenio

Dipartimento di Ingegneria Civile e Ambientale, Università degli studi di Trieste, Trieste, Italy

ABSTRACT: Large eddy simulations of a periodic channel subjected to an oscillating pressure gradient and a wind stress at the free surface are presented. The resulting pulsating flow can be decomposed in a constant part which is the result of the wind stress and an oscillating part due to the pressure gradient. The pressure gradient first accelerates the flow leading to an increased wall stress and higher turbulence levels. Near the bottom a logarithmic layer is observed with strong turbulent streaks as is typical for steady boundary layer turbulence. When the flow is decelerated, streaks are weakened and smoothened by viscosity. After the flow reverses a log layer with turbulent streaks reappear when the wall stress builds up. At the free-surface the wind stress drives a constant production of turbulence and streaks are present throughout the cycle. As bottom and free-surface streaks are aligned mainly one-component turbulence is observed in the interior.

(2)

the small scales of the flow exert on the large scales. As the small scales in the flow are not resolved in LES a model is required for the subgrid stresses % . ij

In the simulations we use a dynamic eddy-viscosity model combined with a scale-similar model (Ar-menio & Piomelli, 2000).

The channel domain is periodic along the hori-zontal x- and y-direction and is bounded in the verti-cal direction by a no-slip bottom (z=0) and a free-surface layer at the top (z =1). For a sketch of the domain see figure 1. The horizontal dimensions are larger than the channel height to capture the largest eddies in the domain (lx×ly×h=2×1.4×1). To

mimic a tidal flow an oscillating pressure gradient

( )

t U

fp =$ 'cos' (2)

with frequency ' =1/80and velocity amplitude

1 =

U is applied over the x-direction. Along the positive x-direction a constant wind stress %wind=10

-3

acts on the free-surface layer. All quantities are made dimensionless using the height of the channel and the velocity amplitude of the tidal oscillation.

Equation (1) is solved using a finite-volume method based on the method by Zang et al. (1994). The Reynolds number and Keulegan–Carpenter number are decreased to Reh =Uh/! = 5123

4 and

80

KC = to make the simulations feasible. As no wall-model is available for this kind of flow, the wall stress must be resolved by the LES. For the no-slip boundary layer to be resolved the first grid cell is set to be equal to one wall unit, which is defined

as * z z z =( ( + with % ! u z =* and & % % w,max

u = . The maximum wall stress %w,max

can be estimated in advance using the maximum wall stress for the purely oscillating case and adding the surface stress. Using data from Jensen et al. (1989) this yields an estimated maximum of

=

max ,

w

% 3.0123-3. Resolving the free-surface layer proved to require a finer resolution. Here, the first grid cell height is (z+ =1 2. The horizontal grid spacing required in the boundary layer for resolved LES is ( + )60

x for the streamwise direction and 30

)

(y+ for the spanwise direction. These re-quirements are reached with a resolution of

128 64

48× × if grid stretching is applied for the ver-tical. For these grid resolutions the numerical model has been used successfully to simulate a turbulent oscillating channel flow as is typical for the Gulf of Trieste (Salon et al., 2007).

3 MEAN VELOCITY

The obtained flow field is strongly turbulent during the complete cycle. The combination of the tidal and wind forcing results in a pulsating mean flow in the x-direction (figure 2). The mean flow u

( )

z is ob-Figure 1. The domain describes a water column of size

h l

lx× y× bounded by a no-slip bottom and a free-surface

at the top. An oscillating pressure gradient fp is applied over the x-direction, while a wind stress %wind is acting on the free surface in the same direction as the pressure gradi-ent.

Figure 2. The mean streamwise velocity for oscillating channel flow subjected to a wind stress for different phases (a) 0°–60° (b) 90°–150° (c) 180°–240° and (d) 270°–330°. The darker colors represent the later phases, e.g. in (a) the profiles go from light gray, to dark gray to black for 0°, 30° and 60°, respectively.

(3)

tained by plane and phase averaging. The wind stress drives a constant mean flow in the interior with strong shear layers at the bottom and free-surface layer. A similar velocity profile is observed, although with a larger velocity, when the oscillating pressure gradient is absent. The pressure gradient will accelerate and decelerate the flow in positive x-direction in the first half period. While in the second half period it will accelerate and decelerate the fluid in the negative x-direction. The mean velocity reaches a maximum velocity of about U = 1.4 in the interior at * = 100°. The flow is then decreases, subsequently reverses and reaches a value of U= -0.6. at * = 250°.

The wall stress is given in figure 3. Integrated over a period the wall stress matches the applied wind stress. An above-average wall stress is ob-served during a shorter period 20°–170° than a be-low average value. The typical Reynolds number

=

= '!

' 2 max

Re U 41236 in the first half cycle and 71235 in the second half cycle. The wall stress pro-files in the first and second half cycle are in agree-ment with the profiles Jensen et al. (1989) observed for different Re (see their figure 9). The Reynolds ' numbers indicate that the first half cycle is in the fully turbulent regime leading to an increased wall stress and increased production of turbulence. The second half cycle is in the intermittent turbulent re-gime. A burst like production of turbulence is then occurring in the deceleration phase (Hino et al., 1983). This leads to a small increase of the wall stress and turbulent kinetic energy around * = 270°. The mean velocity profile in the boundary layers is given in figure 4. If only a wind stress is present we find that a logarithmic boundary layer is present both at bottom and the free surface. In the bottom layer the log law is u+ = $1logz+ +C

+ with the von

Karman constant + =0.41and C=5. In the pres-ence of an oscillating pressure gradient a clear log layer is observed with the constant C ranging be-tween 5 and 7 for the phases 30°–150° when the wall stress is large. The same increased values of C were found by Salon et al. (2007) for the purely os-cillating case. For the phases, when the mean

veloc-ity is either small or reversing, no log layer is pre-sent. If the wall stress then increases again at the phases 270°–300° a log layer reappears.

In the first half cycle the free surface layer does not differ from the constant case, i.e. with only a wind stress acting on the free surface. Then

C z h z u u+ $ +( = )= $1log + +

+ describes the log

layer at the free surface, but with + =0.5 and C =1. The lower value of C relates to a thinner viscous sublayer. Tsai et al. (2005) argued that presence of horizontal fluctuations at the free-surface has a simi-lar effect as surface roughness, which leads to a de-crease of the constantC. Starting at 210° the thick-ness of the viscous layer seems to increase to +

z = 10. Hence, the log layer present at 240°–300° is fur-ther away ( =C 3) from the free surface. When a log layer is present either at the free surface or at the bottom wall, it extends approximately a quarter height in to the domain.

4 REYNOLDS STRESSES

The stresses exerted by the turbulent fluctuations on the mean flow are called the Reynolds stresses. Symmetry prescribes that u ""w is the only non-zero shear stress. This stress distributes the mean streamwise momentum in the vertical direction. Momentum created by the wind stress is subse-quently distributed by the viscous and Reynolds stress and destroyed at the no-slip bottom. Without an oscillating pressure gradient u ""w is uniform and negative in the interior. Figure 5 gives u ""w during different phases of the oscillatory flow. In the first half cycle the wall stress exceeds the wind stress. In the acceleration phase the total stress

w u z u # $ " " # =!

% hence first increase near the Figure 4. The mean streamwise velocity in wall units for (a) 0° (b) 90° (c) 180° and (d) 270°. The lines relate to the theo-retical laws for the viscous sublayer and logarithmic layer. Figure 3. The phase-averaged stress at the wall. Integrated

(4)

bottom and subsequently fills a larger part of the domain. In the deceleration phase a relaxation of the stress towards a uniform distribution is observed. In the second half cycle, when the wall stress reverses, opposite shear stresses occur in the lower half of the domain.

The normal stresses u "i"ui , i.e. twice the kinetic energy of the turbulent fluctuations, are present for all phase throughout the domain. Turbulent kinetic energy of the streamwise fluctuations is directly cre-ated by shear production $ u"w" # u #z. Hence, streamwise fluctuations are particular intense in the boundary layers where the shear is strong. In the in-terior turbulent kinetic energy is about a factor four smaller. Pressure and inertial forces transfer energy from the streamwise fluctuations towards the cross-stream and vertical fluctuations. Turbulent kinetic energy is lower for these directions and more uni-formly distributed over the domain. Overall turbu-lence levels in the bottom boundary layer are vary-ing with changvary-ing wall stress. The fixed wind stress results in more constant levels of turbulence in the upper part of the domain.

5 STRUCTURES OF THE TURBULENT FLOW The varying wall stress has an impact on the in-tensity and structure of turbulence. The turbulent structures are visualized using the fluctuations in the streamwise velocity u"=u$ u . In the first half cy-cle the strong shear causes the formation of turbulent low-speed streaks in the bottom layer (figure 7). When de flow decelerates the streaks are smooth-ened by viscosity. Then in the second half cycle wall stress builds up again and turbulent streaks reappear at * = 270°.

At the free surface high-speed streaks and low-speed spots are present throughout the cycle. The same features were observed by Tsai et al. (2005) for a stress driven free surface flow. Small-scale tur-bulent fluctuations appear at 210° and 300° (not shown here). Then the layer immediate below the free surface is moving in the opposite direction.

To study the turbulent structure in the interior we constructed Lumley triangle (Pope, 2000) or turbu-lence triangle (figure 8). In each horizontal plane the Figure 5. The Reynolds shear stress u ""w (black) for different phases. The grey line corresponds to the shear stress observed when only a wind stress is present.

Figure 6. The square root of the normal Reynolds stress 2 1

i iu

u "" , i.e. the streamwise r.m.s. velocity (black), the spanwise r.m.s. velocity (gray) and the vertical r.m.s. velocity (dashed).

(5)

second II and third invariants b III of the anisot-b

ropic Reynolds stress tensor, bij = ui"u"j $,ij 3, are

calculated. The combination of the two invariants re-late to an anisotropic state of the Reynolds stresses. For most phases we observe one-component turbu-lent structures in the bottom boundary layer, where the streamwise fluctuations are dominant over the spanwise and vertical fluctuations. The structures become more 3D isotropic away from the wall. This specific signature in the Lumley map, compares well with the one observed for steady boundary layer flow (Kim et al., 1987). In the free-surface layer tur-bulence is also nearly one component throughout the cycle.

The laminarization of the flow in the bottom boundary layer in de deceleration phases leads to 2D Reynolds stresses or pancake-like structures at 180°. The region where two-component turbulence is ob-served then further protrudes into the domain at 240°. In the wall layer a return to one-component turbulence is driven by the increasing amplitude of the wall stress.

6 CONCLUSION

The oscillating pressure gradient together with a fixed stress at the free surface in the same direction results in a pulsating flow. The first half cycle is in the turbulent regime with a strongly increased wall stress. Due to the decreased mean velocity in the second half cycle turbulence falls back to the

inter-Figure 7. Color plots of the streamwise velocity fluctuations (a) at the free-surface (z =h) and (b) in the bottom boundary layer (z=10$2). Values are ranging between -0.3 (black) and 0.3 (white).

Figure 8. The anisotropy of the Reynolds stresses is mapped to the turbulence triangle (grey). The lower edge (0,0) relates to 3D isotropic structures, the left edge (-1/6,1/6) to 2D isotropic structures and the right edge (1/3,1/3) to 1D structures. The black line quantifies the structure of the turbulence at different depths within the fluid column. Marked are the first twenty grid points above the bottom wall (plus) and below the free surface (star).

(6)

mediate regime with a burst-like production of tur-bulence in the deceleration phase. However in real estuary flows it can be expected that both half cycles are in the turbulent regime because of the overall higher Reynolds number.

At the bottom a logarithmic layer with strong tur-bulent streaks develops in the acceleration phases. In the deceleration phase a relaminarization takes place as the streaks are smoothened by viscosity. High speed streaks are present at the free surface through-out the cycle. The observed pattern is in agreement with the findings of Tsai et al. (2005). Shear produc-tions leads to intense streamwise fluctuaproduc-tions in the boundary layer, which are factor four more energetic than in the interior. The fixed stress at the free sur-face leads to constant turbulence intensities, while turbulence levels in the bottom boundary layer are varying with the wall stress. Spanwise and vertical fluctuations are less intense and overall one-component turbulence is observed. The relaminari-zation of the bottom boundary layer gives rise to two-component turbulence that protrudes into the in-terior.

The large Keulegan–Carpenter number results in turbulent flow where the turbulence can be consid-ered quasi-stationary for most phases. This is con-firmed by the presence of log layers during accelera-tion, which are related to steady boundary-layer turbulence. This steady state regime was observed by Scotti & Piomelli (2000) for pulsating flows. The penetration depth of the turbulence created at the wall is then much larger than the channel height.

In estuaries density stratification can play an im-portant role. A stable stratification can suppress tur-bulent motion in the vertical direction. This might decouple the interaction between the bottom and free-surface layer turbulence. Additionally the orien-tation of the wind stress has an important role. Both wind stress orientation and density stratification will be investigated in future work. The simulations for these geophysical flows will then be the basis to test plankton dispersion models.

7 ACKNOWLDGEMENT

This programme is funded by the Netherlands Or-ganisation for Scientific Research (NWO) and Technology Foundation (STW) under the Innova-tional Research Incentives Scheme grant ESF.6239.

REFERENCES

Armenio V. & Piomelli, U. 2000. A Lagrangian mixed sub-grid-scale model in generalized coordinates. Flow, Turbu-lence and Combustion. 65:51–81.

Hino, M., Kashiwayanagi, M., Nakayama, A. & Hara, T. 1983 Experiments on the turbulence statistics and the structure of a reciprocating oscillatory flow. 131: 363-400.

Jensen. B. L., Summer, B. M. & Fredsøe, J. 1989. Turbulent oscillatory boundary layers at high Reynolds numbers. Journal of Fluid Mechanics 206: 265-297.

Kim, J., Moin P. & Moser, R. 1987. Turbulence statistics in fully developed channel flow at low Reynolds number. Journal of Fluid Mechanics 177:133–166.

Pope, S. B. 2000 Turbulent Flows. Cambridge University Press.

Salon, S., Armenio, V. & Crise, A. 2007. A numerical investi-gation of the Stokes boundary layer in the turbulent regime. Journal of Fluid Mechanics 570:253–296.

Scotti, A. & Piomelli, U. 2001. Numerical simulation of pul-sating turbulent channel flow. Physics of Fluids 13:1367– 1384.

Tsai, W.-T., Chen, S.-M. and Moeng, C.-H. 2005 A numerical study on the evolution and structure of a stress-driven free-surface turbulent shear flow. Journal of Fluid Mechanics 545:163–192.

Zang. Y., Street, R. L. & Koseff, J. R. 1994. A non-staggered grid, fractional time step method for time-dependent in-compressible Navier–Stokes equation is curvilinear coordi-nates. Journal of Computational Physics 114:18–33.

Referenties

GERELATEERDE DOCUMENTEN

doch ook - maakte Van der Schuere reedsgebruik van de eigen- sÇhap, dat bij eèn opgaande deeling hef deeltal gelijk is aan deeler X quotient (+ rest): ,,Ofte multipliceert

The Reynolds shear stress uw for the channel flow driven by an oscillating pressure gradient and free surface stress during different phases.. For each subsequent phase the profiles

The small increase of the wall stress amplitude and turbulent kinetic energy around ϕ = 270 ◦ is most likely due to a burst like production of turbulence occurring in the

License: Licence agreement concerning inclusion of doctoral thesis in the Institutional Repository of the University of Leiden. Downloaded

4 Differences in blood velocity profiles between young and 55 elderly healthy volunteers in the internal carotid artery;. a study

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

It attempts to determine if the Matthean community refl ected in the Gospel had a two-tier social structure comprised of followers who strictly obeyed the demanding commands of

Comparing our MUSE data and the fiducial wind model (Cols 3 and 4), we find that the model and data match qualitatively for the blue- (red-) shifted absorptions in the quasar