02.03.2014 Views

Ronnie Knikker, Villeurbanne, F - Eurotherm 2008

Ronnie Knikker, Villeurbanne, F - Eurotherm 2008

Ronnie Knikker, Villeurbanne, F - Eurotherm 2008

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

NUMERICAL STUDY OF BUOYANCY EFFECTS<br />

IN A CHANNEL FLOW GENERATED BY<br />

ASYMMETRIC WALL INJECTION<br />

R. <strong>Knikker</strong>, M. Michard, G. Bois<br />

CETHIL, CNRS, INSA-Lyon, Université Lyon1, F-69621, <strong>Villeurbanne</strong>, France.<br />

Abstract<br />

A numerical study is presented on the effects of buoyancy in a two-dimensional channel flow<br />

generated by wall injection in the unstably stratified regime for moderate Reynolds numbers.<br />

Fluid is injected with uniform velocity and temperature through a porous horizontal wall located<br />

in front of an impermeable wall at lower uniform temperature. The channel is closed on one<br />

side with an impermeable adiabatic wall. Buoyancy effects are characterised by the Richardson<br />

number. We show that for zero of low Richardson numbers, the flow is steady and presents<br />

a recirculation zone in the corner of the impermeable walls, growing in size with increasing<br />

buoyancy effects. Unsteady motion appears for Richardson numbers above a critical value,<br />

without external forcing of the flow. Initially, flow perturbations are periodic and remain<br />

confined to the region near the head end wall. At higher Richardson numbers, however, this<br />

unsteady motion triggers the intrinsic instability of this flow, called parietal vortex shedding.<br />

Resulting flow structures are amplified in the streamwise direction and affect in turn the thermal<br />

boundary layer further downstream.<br />

1 Introduction<br />

Natural and mixed convection have been extensively studied both for flows in closed cavities<br />

(Xin and Le Quéré, 2006) and for open channel flows (Nicolas, 2002), while the case of a semiopen<br />

channel flow generated by blowing through a porous wall with imposed temperature<br />

variations is less investigated regarding unsteady effects. The generic geometry of interest in<br />

the present paper is shown in figure 1. Fluid is injected with uniform velocity U in at the<br />

lower porous wall located in front of an impermeable upper wall, with a wall spacing called<br />

h; the channel is closed at x = 0 with an impermeable front end wall, while the exit is free.<br />

A temperature difference is imposed between upper and lower walls, while the front end wall<br />

is adiabatic. Regarding practical applications, this geometry is a very simplified model, called<br />

cold model, of the flow induced by gas expansion due to combustion in solid propellant rocket<br />

motors. Steady laminar flow induced by wall injection without front end wall is often referred to<br />

as a Berman flow, by extension of the primitive case studied by Berman (1953) for the suction<br />

case.<br />

no-slip, uniform temperature T 2<br />

no-slip<br />

adiabatic<br />

y<br />

x<br />

outlet<br />

h<br />

injection velocity U in , uniform temperature T 1<br />

Figure 1: Sketch of geometry, boundary conditions and streamlines.


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

In the quasi-isothermal case, the control parameter is the injection Reynolds number Re in<br />

based on U in and h, and the main features of the flow are quite different from that observed<br />

in an open channel flow. First, due to uniform injection, the flow is accelerated and both bulk<br />

velocity U b and Reynolds number based on U b are increasing with the downstream distance<br />

x; therefore, some transition from a laminar state to a turbulent one is expected inside the<br />

channel and/or in boundary layers, even if acceleration is known to delay the transition. The<br />

second main feature of the flow induced by wall injection is that streamlines exhibit a strong<br />

curvature, due to the finite spacing between porous and impermeable walls. As a consequence,<br />

the flow exhibit some intrinsic instability, called parietal vortex shedding (PVS).<br />

For numerical simulation of the velocity field, the front end wall is generally removed in<br />

order to simplify the computations: the flow is supposed to be symmetric around the x = 0<br />

line, and the axial gradients of pressure and vertical velocity are set to zero. In such a case, for<br />

steady, laminar, incompressible flow, the velocity field show a self-similar solution well known<br />

in the literature.<br />

PVS has been predicted theoretically (Casalis et al., 1998) using stability analysis. The<br />

appearance of unsteady motion due to transition and PVS has also been numerically predicted,<br />

using statistical turbulence models (Kourta, 2004) or Large Eddy Simulations (Apte and Yang,<br />

2003; Chaouat and Schiestel, 2005); in some of these computations, a random forcing is introduced<br />

at the injection wall, either to trigger PVS or to simulate fluctuations of the injection<br />

velocity. Some studies take into account the front wall with a no-slip condition; in that case, the<br />

formation of a secondary vortex in the corner near the porous wall is mentioned (Apte and Yang,<br />

2003), but this eddy is generally considered as a numerically induced unwanted vortex.<br />

The characteristics of the Berman flow with both fluid injection and temperature variations<br />

are much less documented in the literature. For uniform injection and laminar, incompressible<br />

flow without front end wall nor buoyancy, both velocity and temperature fields show self-similar<br />

solutions; Fournier et al. (2007) determined the main properties of the temperature self-similar<br />

solution and heat transfer coefficient as a function of Reynolds and Peclet numbers, both for<br />

symmetric and non-symmetric injection; in Fournier et al. (2005), mass and heat transfer have<br />

been numerically studied using statistical models. Ferro and Gnavi (2002) studied the influence<br />

of viscosity variations due to temperature on the stability of the Berman flow, both for suction<br />

and injection cases, but without buoyancy effects. For studies related to solid-propellant rocket<br />

motors applications, the injection velocity and corresponding Reynolds numbers are generally<br />

relatively high; as a consequence, due to the downstream acceleration of the flow, Mach numbers<br />

inside the channel are not small, and compressibility effects can induce some temperature<br />

variations. Nevertheless, theses effects appear as second order phenomena, and do not modify<br />

the basic features of the velocity field, compared to an isothermal case (Apte and Yang, 2003;<br />

Chaouat and Schiestel, 2005). Finally, some papers (Chu et al., 2003) are devoted to the study<br />

of a reacting case with high temperature variations due to injection of reactive products.<br />

The objective of the present study is to investigate the effect of buoyancy on the flow<br />

dynamics in the unstably stratified case, for moderate Reynolds numbers. More precisely, we<br />

are focused on the characterisation of the vortex in the upper left corner due to the end wall,<br />

the appearance of unsteady motion for increasing buoyancy effects and the coupling between<br />

this unsteady motion and the intrinsic instability of the Berman (or Taylor) flow.<br />

Section 2 give in more details the governing equations and control parameters, and an<br />

overview of the numerical methods. In section 3, we present the main features of the flow without<br />

and with buoyancy, in a range of governing parameters where the flow remain steady. We<br />

then show that unsteady motion can appear for larger Richardson numbers. Final conclusions<br />

are drawn in section 4.


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

2 Methodology<br />

The configuration retained in this study is the two-dimensional channel (figure 1) closed by an<br />

adiabatic impermeable wall at the head end. Fluid at the temperature T 1 is injected homogeneously<br />

through the lower wall and the upper impermeable wall is at constant temperature<br />

T 2 . As mentioned previously, we focus on the case of unstably stratified flow, therefore T 2 < T 1<br />

when gravity forces are pointing in the usual downward direction.<br />

Physical model and governing equations<br />

Radiative heat fluxes and the effect of variable properties are neglected. The fluid motion and<br />

buoyancy effects are modelled by the Boussinesq approximation. Space variables, time, pressure<br />

and velocity components are normalised using the channel height h, the injection velocity U in<br />

and the reference density ρ. Under these assumptions, the continuity, momentum and energy<br />

equations are<br />

∇ · ⃗u = 0<br />

∂⃗u<br />

∂t + ⃗u · ∇⃗u = −∇p + Re−1 ∇ 2 ⃗u + Ri Θ (1)<br />

∂Θ<br />

∂t + ⃗u · ∇⃗ Θ = Pe −1 ∇ 2 Θ<br />

where Θ = (T −T 2 )/(T 1 −T 2 ) is the non-dimensional temperature, Re = U in h/ν is the Reynolds<br />

number based on the injection velocity, channel height and fluid viscosity ν and Pe = Re Pr<br />

is the Péclet number. The control parameter for buoyancy effects is the Richardson number,<br />

defined as<br />

Ri = gβ(T 1 − T 2 )h<br />

U 2 in<br />

where g is the gravitational acceleration and β is the coefficient of volumetric thermal expansion.<br />

Note that the Richardson and Rayleigh numbers are related through Ra = Ri Re 2 Pr. In this<br />

study, the Reynolds number is fixed at Re = 200 and the Prandtl number is Pr = 0.7.<br />

Numerical method<br />

The governing equations (1) are solved on a staggered Cartesian non-uniform grid. Fourthorder<br />

compact discretization (Lele, 1992) is used for all spatial derivatives and the resulting<br />

ODE is advanced in time using the semi-implicit third-order Runge-Kutta/Crank-Nicolson<br />

scheme proposed by Spalart et al. (1991). Similar methods for incompressible flows are used<br />

by Brüger et al. (2005), Kampanis and Ekaterinaris (2006) and Vedy et al. (2003) and have<br />

shown to be accurate and efficient for unsteady incompressible flows. The pressure equation<br />

is discretized in a consistent way using fourth-order schemes and solved using an iterative<br />

technique. A second-order discrete Poisson operator is used at each iteration and solved using<br />

generalized cyclic reduction (Swarztrauber, 1974). A fourth-order accurate solution of the<br />

pressure is obtained after several iterations.<br />

Special attention has been payed to the discretization of the non-linear terms in equations<br />

(1). The advection term in the momentum equation is recast in the so-called skewsymmetric<br />

formulation and discretized using centered compact schemes. In this way, the<br />

method is stable and conserves correctly the total kinetic energy in the limit of isothermal<br />

inviscid flows (Schiestel and Viazzo, 1995). The same procedure applied to the advection term<br />

in the energy equation leads to unphysical oscillations. This term is therefore rewritten in divergence<br />

form and discretized using a third-order upwind interpolation of the temperature followed<br />

by centered compact finite differencing of the energy flux. Thermal energy is then correctly<br />

conserved and the method is stable due to upwinding. The diffusion terms are discretized using


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

centered schemes. Physical boundary conditions are implemented directly into the schemes<br />

whenever possible. Otherwise, third-order one-sided schemes are used for the boundary points.<br />

A constant pressure boundary condition is used to model the outflow conditions at the right<br />

boundary.<br />

The accuracy of the numerical method is verified using the self-similar solution for the<br />

Berman problem by replacing the head wall with an adiabatic slip boundary. A correct convergence<br />

of the numerical solution is obtained on the entire computational domain, except for<br />

the region near the outlet boundary which is affected by the constant pressure condition. To<br />

minimise the influence of the outlet boundary, all other computations are calculated on a computational<br />

domain of length 30 h. A non-uniform grid is used of size 768 ×64. The grid spacing<br />

in the x-direction varies smoothly from 0.011h near the wall to 0.064h at the outlet, and in<br />

the y-direction from 0.011h near the boundaries to 0.017h at the center of the channel. The<br />

inlet velocity near the head end wall is defined by a boundary layer profile to comply with the<br />

no-slip boundary condition at x = 0. This profile covers roughly 6 grid points near the wall and<br />

is found to have negligible influence on the remaining flow structures. All other computations<br />

are calculated with the no-slip boundary condition at the left wall.<br />

3 Results<br />

Steady flow topology<br />

For small values of the Richardson number and a Reynolds number of Re = 200 the flow is<br />

found to be steady. Figure 2 illustrates the temperature distribution and flow pattern in the<br />

range Ri = 0 to 18. In the absence of buoyancy effects (Ri = 0), a small recirculation zone is<br />

visible in the upper left corner. According to Moffatt (1964), the formation of secondary eddies<br />

is expected in the upper left corner where the flow is dominated by viscous effects. Moffat<br />

predicted that any flow near the corner must consist of a sequence of eddies of decreasing size<br />

and rapidly decreasing intensity. For walls forming a right angle, the relative size of the vortices<br />

falls off in geometric progression whose value is around 10, while their relative intensity decay<br />

in a geometric progression with a common ratio near 2000. Therefore, we expect to capture no<br />

more than one counter-clockwise vortex in the upper left corner.<br />

Figure 2: Effect of buoyancy on the temperature field and streamlines in steady flow conditions. The<br />

downstream region for x > 2 is not shown for clarity.


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

In the presence of buoyancy effects, cold fluid falls along the end wall towards the porous<br />

wall. On the other hand, injection of hot fluid at the porous wall induces an upward motion;<br />

consequently the cold fluid flowing along the side wall is deviated towards the streamwise<br />

direction, and then transported upwards and remains trapped inside a vortex. As a result, the<br />

size of the counter-clockwise natural recirculation increases with increasing buoyancy effects.<br />

Unsteady flow<br />

Unsteady motion is observed for increasing buoyancy effects. The transition from a steady<br />

to unsteady flow is defined by a critical Richardson number that lies between Ri = 18 and<br />

Ri = 20. In terms of the Rayleigh number, unsteady flow appears between 5.04 × 10 5 and<br />

5.6 × 10 5 . Snapshots of the temperature distribution for several unsteady cases are shown in<br />

figure 3. Pockets of fresh fluid appear at the upper wall giving rise to buoyancy plumes ejected<br />

from the upper wall. Nevertheless, buoyancy effects are not sufficient to prevent damping of<br />

these plumes and the formation of a thermal boundary layer further downstream. For Ri ≤ 90,<br />

the thermal boundary layer is of constant thickness far downstream and described by the selfsimilar<br />

steady solution of the Berman problem. On the other hand, for Ri = 120 this boundary<br />

layer remains unsteady due to its interaction with parietal vortex shedding.<br />

Ri = 30<br />

Ri = 50<br />

Ri = 70<br />

Ri = 90<br />

Ri = 120<br />

Figure 3: Snapshots of the temperature distribution for different values of the Richardson number in<br />

the unsteady flow regime.<br />

Parietal vortex shedding<br />

Figure 4 shows snapshots of the vorticity field at different values of the Richardson number<br />

on a larger part of the computational domain. For high Richardson numbers, organized flow<br />

structures are observed near the injection wall on the right side of the computational domain.<br />

Similar flow structures, characteristic of parietal vortex shedding, are observed in the isothermal<br />

case by Apte and Yang (2003) after introduction of random perturbations to the injection<br />

velocity. For Ri = 90, these flow structures remain weak and disappear further downstream.


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

For Ri = 120, buoyancy plumes are sufficiently strong to trigger the instability and parietal<br />

vortex shedding is amplified as shown in figure 4. Interestingly, since the temperature field<br />

converges downstream to a boundary layer, the role of buoyancy in the formation of parietal<br />

vortex shedding is limited to the onset of the instability.<br />

Ri = 70<br />

Ri = 90<br />

Ri = 120<br />

Figure 4: Snapshots of the vorticity fields for different values of the Richardson number. An aspect<br />

ratio of 3 : 1 is used to enhance the visibility of flow structures.<br />

The mean fluctuation of temperature and velocity are shown in figure 5. For Ri = 90,<br />

the unsteady flow motion is strongest in the region near the head end vortex but decreases<br />

rapidly further downstream. For Ri = 120, fluctuations of both velocity and temperature<br />

are distributed over a much larger domain. Downstream of the vortex region, large velocity<br />

fluctuations are observed near the injection wall which are directly related to the presence of<br />

parietal vortex shedding. The temperature and velocity fluctuations near the upper boundary<br />

are due to perturbations of the boundary layer and the large gradients of temperature and<br />

velocity at this location.<br />

(a) Ri = 90<br />

Ri = 120<br />

(b) Ri = 90<br />

Ri = 120<br />

Figure 5: Standard deviation of temperature (a) and velocity fluctuation (b) for Ri = 90 and Ri = 120.<br />

The image aspect ratio is 3 : 1.


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

Spectral analysis<br />

Frequency spectra for the heat flux averaged along the upper wall for Ri = 30 and 120 are<br />

shown in figure 6a. At the lower Richardson number, the strong fundamental peak in the<br />

spectrum and the well-defined harmonics indicate that flow structures are essentially periodic.<br />

The power spectrum at Ri = 120 is a broad-band spectrum, despite the regular pattern of<br />

coherent structures resulting from parietal vortex shedding.<br />

Figure 6b plots the fundamental frequency of the periodic signal as a function of the Richardson<br />

number in the range Ri = 20 to 90 for which the flow is essentially periodic. Similar spectra<br />

and the same fundamental frequencies are obtained from the shear force at the head end wall.<br />

The frequency at Ri = 20 is of the order of U in /h and increases with buoyancy force. A jump<br />

in the frequency is observed around Ri = 49.<br />

Power spectral density<br />

10 0 Dimensionless frequency<br />

10 −5<br />

10 −10<br />

Ri=30<br />

Ri=120<br />

Dimensionless frequency<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

(a)<br />

10 −15<br />

0 2 4 6 8 10<br />

(b)<br />

0<br />

0 20 40 60 80 100<br />

Richardson number<br />

Figure 6: Spectral analysis of the heat flux averaged along the upper wall; (a) power spectra for<br />

Ri = 30 and Ri = 120, (b) frequency of the most energetic mode for Richardson numbers between 0<br />

and 90. The frequency is expressed in non-dimensional units using U in /h.<br />

4 Conclusions<br />

The following main conclusions can be drawn from this two-dimensional numerical study. For<br />

zero or weak buoyancy force, the flow is steady and presents a recirculation zone in the upper<br />

left corner. The size of this recirculation zone increases with increasing buoyancy effect. The<br />

flow motion becomes unsteady above a certain critical Richardson number. For values near<br />

the critical one, flow perturbations are periodic and confined to the region near the head end<br />

wall. The fundamental frequency depends on the strength of the buoyancy force. For larger<br />

values of the Richardson number, the unsteady motion triggers the natural instability (PVS)<br />

of this flow. To our knowledge, this work is the first demonstration in this flow configuration of<br />

unsteady effects resulting from buoyancy forces. Note that no artificial forcing is used in this<br />

study and boundary conditions are steady.<br />

However, it is well known that for channel flows of the Poiseuille-Rayleigh-Bénard type threedimensional<br />

effects in the form of longitudinal rolls appear with increasing thermal stratification<br />

(Xin et al., 2006). It remains to be explored under which conditions similar flow structures<br />

appear in the present accelerated channel flow, in particular with respect to the unsteady twodimensional<br />

flow structures observed in this study. Further research is also needed to understand<br />

the physical mechanisms underlying unsteadiness and determine the essential parameters of this<br />

flow as well as the cause of the frequency jump observed in the periodic flow regime.


5th European Thermal-Sciences Conference, The Netherlands, <strong>2008</strong><br />

References<br />

S. Xin and P. Le Quéré. Natural-convection flows in air-filled, differential heated cavities with adiabatic<br />

horizontal walls. Numerical Heat Transfer, Part A, 50:437–466, 2006.<br />

X. Nicolas. Bibliographical review on the Poiseuille-Rayleigh-Bénard flows: the mixed convection flows<br />

in horizontal rectangular ducts heated from below. International Journal of Thermal Sciences, 41:<br />

961–1016, 2002.<br />

A. Berman. Laminar flow in channels with porous wall. J. Appl. Phys., 24 (9):456–475, 1953.<br />

G. Casalis, G. Avalon, and J.-Ph. Pineau. Spatial instability of planar channel flow with fluid injection<br />

through porous walls. Phys. Fluids, 10(10):2558–2568, 1998.<br />

A. Kourta. Instability of channel flow with fluid injection and parietal vortex shedding. Computers<br />

& Fluids, 33(2):155–178, 2004.<br />

S. Apte and V. Yang. A large-eddy simulation study of transition and flow instability in a porouswalled<br />

chamber with mass injection. J. Fluid Mech., 477:215–225, 2003.<br />

B. Chaouat and R. Schiestel. A new partially integrated transport model for subgrig-scale stresses<br />

and dissipation rate for turbulent developing flows. Physics of Fluids, 17(6):065106, 2005.<br />

C. Fournier, M. Michard, and F. Bataille. Heat transfer in a laminar channel flow generated by<br />

injection through porous walls. Journal of Fluids Engineering, 129:1048–1057, 2007.<br />

C Fournier, F Bataille, and M Michard. Transition characteristics of flowfield in a non-isothermal<br />

duct with wall injection. In Proceedings of 4th International Conference on Computational Heat<br />

and Mass transfer, pages 899–904, Paris, France, 2005.<br />

S. Ferro and G. Gnavi. Effects of temperature-dependent viscosity in channels with porous walls. Phys<br />

Fluids, 14(2):839 – 849, 2002.<br />

W.-W. Chu, V. Yang, and J. Majdalani. Premixed flame response to acoustic waves in a porous-walled<br />

chamber with surface mass injection. Combustion and Flame, 133(3):359–370, May 2003.<br />

S.K. Lele. Compact finite difference schemes with spectral-like resolution. J. Comput. Phys., 103(1):<br />

16–42, 1992.<br />

P.R. Spalart, R.D. Moser, and M.M. Rogers. Spectral methods for the Navier-Stokes equations with<br />

one infinite and two periodic directions. J. Comput. Phys., 96(2):297–324, 1991.<br />

A. Brüger, B. Gustafsson, P. Lotstedt, and J. Nilsson. High order accurate solution of the incompressible<br />

Navier-Stokes equations. J. Comput. Phys., 203(1):49–71, February 2005.<br />

N.A. Kampanis and J.A. Ekaterinaris. A staggered grid, high-order accurate method for the incompressible<br />

Navier-Stokes equations. J. Comput. Phys., 215(2):589–613, July 2006.<br />

E. Vedy, S. Viazzo, and R. Schiestel. A high-order finite difference method for incompressible fluid<br />

turbulence simulations. Int. J. Num. Meth. Fluids, 42(11):1155–1188, 2003.<br />

P.N. Swarztrauber. Direct method for the discrete solution of separable elliptic equations. J. Numer.<br />

Anal., SIAM, 11:1136–1150, 1974.<br />

R. Schiestel and S. Viazzo. A Hermitian-Fourier numerical method for solving the incompressible<br />

Navier-Stokes equations. Computers & Fluids, 24(6):739–752, 1995.<br />

H.K. Moffatt. Viscous and resistive eddies near a sharp corner. J. Fluid Mech., 18:1–18, 1964.<br />

S. Xin, X. Nicolas, and P. Le Quéré. Stability analyses of longitudinal rolls of Poiseuille-Rayleigh-<br />

Bénard flows in air-filled channels of finite transversal extension. Numerical Heat Transfer, Part A,<br />

50:467–490, 2006.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!