• No results found

HåligheterStability of Fluids with Shear-Dependent Viscosity in the Lid-driven Cavity

N/A
N/A
Protected

Academic year: 2022

Share "HåligheterStability of Fluids with Shear-Dependent Viscosity in the Lid-driven Cavity"

Copied!
37
0
0

Loading.... (view fulltext now)

Full text

(1)

Stability of Fluids with

Shear-Dependent Viscosity in the Lid-Driven Cavity

Simon Haque

Thesis Submitted to The Royal Institute of Technology for the Master’s Degree

Supervisor:

Luca Brandt

August 26, 2011

(2)

Abstract

The classical problem of the lid-driven cavity extended infinitely in the spanwise direction is considered for non-Newtonian shear-thinning and shear-thickening fluids where the viscosity is modeled by the Carraeu- Yasuda model. Linear stability is used to determine the critical Reynolds number at which the two-dimensional base-flow becomes unstable to three-dimensional spanwise-periodic disturbances. We consider a square cavity, characterized by steady unstable modes, and a shallow cavity of aspect ratio 0.25, where oscillating modes are the first to become unstable for Newtonian fluids. In both cases, the critical Reynolds number first decreases with decreasing power-index n (from shear- thickening to shear-thinning fluids) and then increase again for highly pseudoplastic fluids. In the latter case, this is explained by the thinner boundary layers at the cavity walls and less intense vorticity inside the domain. Interestingly, oscillating modes are found at critical condi- tions for shear-thickening fluids in a square cavity while the shallow cavity supports a new instability of lower frequency for large enough shear-thinning. Analysis of kinetic energy budgets and structural sen- sitivity are employed to investigate the physical mechanisms behind the instability.

(3)

1 Introduction

1.1 The Lid-Driven Cavity

The incompressible and Newtonian flow inside a lid-driven cavity is one of the most studied fluid mechanics problem since the establishment of compu- tational fluid dynamics (CFD). It describes the flow inside a rectangular box due to the tangential translation of one side wall. Much attention has been drawn to the cavity flow, both from an industrial and academic standpoint.

For example, practical applications can be found in abundance in coating and mixing devices [1]. The geometric simplicity of the problem makes it ideal for numerical discretization and computations, while the discontinuous mathematical boundary condition for the velocity leads to singular proper- ties close to the corners of the moving lid at the two neighboring stationary walls. These features has made the lid-driven cavity a popular method of verification for Navier-Stokes solvers [2] and different numerical techniques.

Furthermore, this flow has interesting behavior from the point of view of instabilities and bifurcations in closed system behavior.

Previous work have been conducted both numerically and experimen- tally. Burggraf [3] aimed to investigate the Prandtl-Batchelor theorem for a square cavity. He used a relaxation method to compute solutions for Reynolds number (Re) between 0 and 400. The results suggests that for higher Re an inviscid core vortex is formed, but secondary eddies develop near the bottom corners of the square for all Re.

Later, Pan and Acrivos [4] also used relaxation techniques to find creep- ing flow solutions close to bottom corners of the cavity for aspect ratios, Γ, between 0.25 and 5 where Γ is defined as the ratio between the depth of the cavity and the length of the moving lid. Moreover, through experiments, they studied the flow for Re ranging from 20 to 4000 and concluded that for finite cavities and as Re → ∞ a single inviscid core will form and the corner eddies reduces significantly in size. For infinitely deep cavities the size of the primary vortex hinders the core from becoming fully inviscid even as Re

→ ∞ resulting in a balance between viscous and inertial forces in the cavity.

The main assumption in [4] is two-dimensional flow.

Ghia et al. [5] and Schreiber and Keller [6] focused on numerical methods which they tested on the two-dimensional cavity with Γ= 1 for Re up to 10000. Both studies indicate that at such high velocities the position of the core vortex moves toward the center of the square and that the bottom left (BL) and right (BR) eddies grow in size. The BR is often referred to as the down stream secondary eddy (DSE).

Extensive three-dimensional work was performed by Koseff and Street together with others [7]-[8]. They carried out experiments with spanwise aspect ratio, Λ, between 1 and 3 where Λ is defined as the ratio between the width and length of the moving lid. The results showed that the flow

(4)

becomes turbulent at an Re in the interval 6000-8000 and that fluid motion in the spanwise direction is considerable [7]. In subsequents investigations they also established local three-dimensionality as corner vortices at the end walls are formed as well as longitudinal Taylor-Grtler-like (TGL) vortices due to instabilities close to the DSE [9], [10]. The experimental results were later used to confirm numerical result obtained by the same research group [8]. The main conclusion that can be drawn from the work of Koseff et al.

is that two-dimensional flow (for Λ up to 3 ) at high Re cannot occur, it is strictly fictitious.

Kim and Moin used a time-splitting scheme together with the approximate- factorization technique to investigate the three-dimensional driven cavity with periodic boundary condition in the transversal (spanwise) direction, i.e. without end wall effects [11]. They added small random perturbations in the spanwise direction to the two-dimensional solutions and found that a pair of TGL vortices appeared at Re around 900. This result was impor- tant because it proved the flow becomes three-dimensional at high Re even without the presence of end walls in the spanwise direction.

Chiang et al. set out to map the eddy structure as function of the Re up to 1300 [12]. They used a finite volume method on a cavity with square cross section and Λ = 3. It is shown in the paper that the BL and DSL become developed at around Re = 50. At an Re close to 100, the corner eddies adjacent to the lid are formed. As the Re is increased it becomes evident that the two-dimensional nature of the flow undergoes a transition to a fully three-dimensional character in the form of a bifurcation from the steady state to an oscillating periodic state.

Additional information can be found in a review paper by Shankar and Deshpande [13].

1.2 Stability

After the work of Koseff, Street and their colleagues much attention has been drawn to a new aspect of the lid-driven cavity, namely stability. Since it is generally accepted that the flow goes through a transition a from two- dimensional to a three-dimensional state, research have been conducted to find at what Re this conversion occurs. As Albensoeder et al. describes in [2] the flow undergoes symmetry breaking instabilities before becoming turbulent. The task is to find at what Re the first instability takes place when increasing the Re from 0. This Re is often referred to as the critical Re.

An innovative experiment was carried out by Aidun et al. in which a small amount of through flow was injected into a cavity with Λ = 3 to compensate for the liquid which was removed by the moving lid [14], [15].

They concluded that this addition would not alternate the flow character- istics significantly. Two sets of experiments were set up. In the first one,

(5)

the transition from steady to unsteady state was studied when increasing Re from 100 to 2000. From the results they concluded that the core vortex and DSE is stable up to a Re of around 825, but then small-amplitude time- periodic waves form at the DSE, extending towards the end walls. They estimated a critical Re between 825 and 925. In [15] they also found super- critical bifurcations from steady state to a pair of spiral waves at Re = 966 with a dimensionless frequency of 0.1112. Upon increasing the Re, travel- ing waves developed for Re between 1000 and 1300 leading to much more complex dynamics. As the Re was increased further, non-uniformly sized and irregular spaced mushroom-like spikes appear. In the second set of ex- periments the velocity was decelerated from high speeds (Re ∼ 2000) to low (Re < 500) and showed that the steady state are not unique.

Ramanan and Homsy performed a numerical linear stability analysis of a square cavity with periodic boundary condition in the spanwise direction by applying a finite difference approach [16]. Similar to Kim and Moin, they first solved for the two-dimensional base flow and then perturbed it with three-dimensional disturbances using a uniform mesh of 65×65 grid points. The result indicated a Grtler type instability close to the separating streamline between the core vortex and the DSE. The flow became unstable at Re ∼ 594 to a stationary mode with a transversal wave number κ ∼ 2.12 Adding compressibility effects to the problem, Ding and Kawahara used a finite element method to calculate the critical Re in a square cavity of infinite span [17]. They detected an oscillatory mode with non-dimensional frequency ω = 0.08 at Re = 920 for κ = 7.4.

During the last decade, Kuhlmann and coworkers have produced a sig- nificant amount of work on problems of the lid-driven cavity. For example, utilizing a cavity where two lids move in different direction, they were able to show non-uniqueness in the two-dimensional steady flow [18]. In a following study they employed a finite volume discretization to establish the critical Re for a cavity with different Γ but with Λ → ∞ [2]. With a non-uniform 141×141 grid and by linear stability analysis it was shown that the base flow becomes unstable to four different three-dimensional modes at higher Re depending on the aspect ratio. For example, the square cavity suffers a stationary instability at Re around 786 and a κ of 15.4, hence quite different from results obtained by earlier research. These values were confirmed by experiments in a cavity with dimensions Γ = 1 and Λ = 6.55 and are now generally considered to be correct. Kuhlmann et al. asserted that this mode had been suppressed by end wall effects (due to short spanwise length) in previous work.

To the best of the author’s knowledge, very little work concerning sta- bility of non-Newtonian fluids in the lid-driven cavity exists. Pakdel et al.

experimentally analyzed two-dimensional cavity flow for elastic liquids [1].

With spanwise aspect ratios between 0.25 and 4 they used an ideal elastic fluid and Deborah number (De) ranging from 0 to 35 to conduct their in-

(6)

vestigation. At low De the flow was two-dimensional and the core vortex moved upstreams as Re was increased. For low De the flow became unstable and three-dimensional. When the lid motion was stopped a distinct recoil effect caused the direction of the core vortex to reverse.

The aim of this thesis project is to numerically establish bench mark re- sults for non-Newtonian shear-thinning and, to some extent, shear-thickening fluids in the classical lid-driven cavity, assuming Λ → ∞. A two-dimensional base flow solution is initially acquired and then perturbed with three-dimensional disturbances. Linear stability analysis then leads to an eigenvalue problem which is solved using the Arnoldi iteration method.

2 Problem Formulation

2.1 The Geometry

The geometry and generic flow behavior of the lid-driven cavity is depicted in Figure 1. The lid has a width Lx and is moving with a constant velocity V in the x-direction, hence the the Reynolds number Re is defined as:

Re = ρLxV ˆ µ0

,

where ρ and ˆµ0 is the density and zero shear rate viscosity, respectively, of the fluid. Furthermore, the aspect ratio is given by Γ = Ly/Lx. If a third z-direction is added then the spanwise aspect ratio is Λ = Lz/Lx, where Lz is the length of the cavity in the corresponding direction.

Core Vortex

Corner Vortices

V

Ly

Lx

Figure 1: Geometry and notation of the lid-driven cavity.

(7)

2.2 The Viscosity Model

As mentioned, an important aspect of this work which differentiates it from many earlier investigations is that it concerns Non-Newtonian fluids. To examine the non-Newtonian effects, a simple model where viscosity is de- pendent on shear rat is adopted. Fluid elasticity is neglected. Hence, for the the relation can mathematically be written as:

µ = µ( ˙γ), (1)

where ˙γ is the rate-of-shear tensor and will simply be referred to as ˙γ.

Thus, many empirical relations can be used to fit experimental data. Some examples, including the Cross and Bingham model can be found in [19].

For this project the so called Carreau-Yasuda model is employed from which the connection between viscosity and rate of strain is given by:

ˆ

µ( ˙γ) = ˆµ+ (ˆµ0− ˆµ)[1 + ( ˙γλ)a]n−1a . (2) Here, ˆµ is the infinite shear rate viscosity. In this work, ˆµ and ˆµ0 are set to 0.001 and 1 respectively. Dividing equation (2) by ˆµ0 yields the non- dimensional viscosity:

µ( ˙γ) = µˆ

ˆ µ0

+ (1 −µˆ

ˆ µ0

)[1 + ( ˙γλ)a]n−1a . (3) The exponent a is fixed at 2 in this analysis since then the rheological behav- ior of many polymeric solutions can be fitted to the Carreau-Yasuda model [20]. For a more detailed description of the parameters featured in equation (3) see [21].

2.3 Governing Equations

The incompressible Non-Newtonian lid-driven cavity flow is considered. The flow is induced by the tangential velocity of the top wall in the positive x- direction. All remaining walls are rigid. The cavity length in the transverse z-direction is assumed to be infinite. Thus, the flow is governed by the Navier-Stokes and continuity equation which is expressed in non-dimensional form as:

∂u

∂t + u · ∇u = −∇p + 1

Re · ∇[µ(∇u + ∇uT)] (4a)

∇ · u = 0. (4b)

The no-slip boundary conditions imposed on the problem are the follow- ing:

(8)

u =

(1 ex at y = LLy

x = Γ (5a)

0 at x = 0, x = 1 and y = 0. (5b) Note that for equation (5), the origin of the coordinate system has been placed in the bottom left corner of Figure 1. Also, the boundary conditions are dimensionless in accordance with equation (4).

2.4 Linear Stability Analysis

Since the cavity length is presumed infinite in the z-direction, equation (4) has a steady two-dimensional solution [Ub, pb] = [Ub(x, y), pb(x, y)]. A per- turbation [ ˜U, ˜p] = [ ˜U(x, y, z, t), ˜p(x, y, z, t)] is added to this time indepen- dent state in order to perform the linear stability analysis. Thus, it is possible to decompose the flow variables into:

u = Ub+ ˜U (6a)

p = pb+ ˜p. (6b)

Furthermore, a similar formation is introduced for the viscosity:

µ = µb+ ˜µ, (7)

where µb = µb(x, y) and ˜µ = ˜µ(x, y, z, t) are the base flow (steady state) and perturbation viscosity respectively. According to [22] the perturbation viscosity is equal to the first term of the Taylor series expansion of (3):

˜

µ = ˙γij( ˜U) ∂µ

∂ ˙γij(Ub). (8)

Now, substituting equations (6) and (7) into (4), subtracting the base flow variables and linearizing around [Ub, pb] yields a linear stability problem for the perturbations which can be formulated compactly as:

∂ ˜U

∂t + L(Ub, Re) ˜U + ∇˜p = 0 (9a)

∇ · ˜U = 0. (9b)

Above, L(Ub, Re) ˜U is:

L(Ub, Re) ˜U = ˜U · ∇Ub+ Ub· ∇ ˜U

− 1

Re∇ · [µb(∇ ˜U + (∇ ˜U)T) + ˜µ(∇Ub+ (∇Ub)T)].

(9)

As introduced, for example, by Albensoeder et al. the general solution to equation (9) can be written as complex modes of the form:

U(x, y, z, t) = ˆ˜ U(x, y)exp[σt + iκz] + compl. conj. (10a)

˜

p(x, y, z, t) = ˆp(x, y)exp[σt + iκz] + compl. conj., (10b) where κ is the transverse wavenumber of the disturbance. A two-dimensional instability corresponds to κ = 0. Inserting the ansatz (10) into (9) finally produces a linearized general eigenvalue problem:

σ ˆU + L(Ub, Re) ˆU + ∇ˆp = 0 (11a)

∇ · ˆU = 0. (11b)

The complex eigenvalue σ contains information about the growth rate (RE{σ}) and frequency (IM{σ}) of the instability. Subsequently, ˆq = ( ˆU, ˆp) is the eigenmode. The main task for this analysis is to determine for which Re and κ the growth rate becomes zero for a given Γ, n and λ. These Re and κ will be referred to as the critical Reynolds number (Rec) and wavenumber (κc).

2.5 Structural Sensitivity

Investigating the sensitivity of the unstable modes will give further insight about the origin of the instability. The theoretical framework is based on the work of Giannetti and Pralits [23] who introduce and define the wavemaker of the instability as the region in space where a change in the structure of the problem causes the largest drift in the eigenvalues. Compared to their analysis, terms for the perturbation viscosity are added here. Introducing a small momentum force in the stability equations yields the following eigen- value problem:

σ00+ L(Ub, Re) ˆU0+ ∇ˆp0= δH( ˆU0, ˆp0) (12a)

∇ · ˆU0= 0. (12b)

As explained in [24], δH is a differential operator representing a force pro- portional to the local perturbation velocity:

δH( ˆU0, ˆp0) = δM (x, y) · ˆU0 = δ(x − x0, y − y0)δM0· ˆU0

Here, δM0 is a coupling matrix and δ(x − x0, y − y0) is the Kronecker delta function. The eigenvalue and eigenmode drifts are given as the expansion σ0 = σ + δσ, ˆU0 = ˆU + δ ˆU and ˆp0 = ˆp + δ ˆp. Inserting these into equation

(10)

(12) and excluding higher order terms yields an equation for the eigenvalue drift:

σδ ˆU + L(Ub, Re)δ ˆU + ∇δ ˆp = −δσ ˆU + δM · ˆU (13a)

∇ · δ ˆU = 0 (13b)

The Lagrange identity is introduced as a function of the differentiable direct field q = (u, p) and its corresponding adjoint field g+= (f+, m+). More on adjoint methods can be found in [25] and [26]. By taking the inner product of equation (11) and the adjoint field and using differentiation by parts one acquire:

[(σ ˆU + L(Ub, Re) ˆU + ∇ˆp) · ˆf++ (∇ · ˆU) · ˆm+]

+ [ ˆU · (−σˆf++ L+(Ub, Re)ˆf++ ∇ ˆm+) + ˆp∇ˆf+] = ∇ · J(ˆq, ˆg+), (14) where J is a bilinear concomitant and L+ is the adjoint operator of the linearized Navier-Stokes equation.

J(ˆq, ˆg+) = Ub( ˆU ·ˆf+) + 1

Re[µb(∇ˆf++ (∇ˆf+)T) · ˆU − µb(∇ ˆU + (∇ ˆU)T) ·ˆf+

− ˜µ(∇Ub+ (∇Ub)T) · ˆf+] + ˆm+U + ˆˆ pˆf+, and

L+(Ub, Re)ˆf+= Ub· ∇ˆf+− ∇Ub· ˆf+ + 1

Re[µb(∆ˆf++ (∆ˆf+)T) + (∇Ub+ (∇Ub)T) · ∇ˆf+B(Ub)].

where B(Ub) is an operator on the form:

B(Ub) =

 2∂ ˙∂µγ

11(Ub) ·∂x + 2∂ ˙∂µγ

12(Ub) ·∂y 2∂ ˙∂µγ

21(Ub) ·∂x + 2∂ ˙∂µγ

22(Ub) ·∂y

Additionally, the adjoint mode ˆg+(x, y) = (ˆf+, ˆm+) satisfies the following system of equations:

−σˆf++ L+(Ub, Re) + ∇ ˆm+ = 0 (15a)

∇ · ˆf+ = 0 (15b)

(11)

Considering equations (13) and (14) and integrating over the entire do- main D gives:

−δσ Z

D

ˆf+· ˆUdS + Z

D

ˆf+· δM · ˆUdS = I

DJ(ˆq, ˆg+) · ndl. (16) The boundary conditions are chosen in such a way that the right hand side of equation (16) is zero. The sensitivity tensor S is introduced:

S(x0, y0) =ˆf+(x0, y0) ˆU(x0, y0) R

Dˆf+· ˆUdS

. (17)

Note that ˆf+U represents a dyadic product between the adjoint and directˆ mode. Combining equation (16) and (17) generates the final expression for the eigenvalue drift:

δσ(x0, y0) = R

Dˆf+· δM · ˆUdS R

Dˆf+· ˆUdS

=ˆf+· δM0· ˆU R

Dˆf+· ˆUdS

= S : δM0 = SijδM0ij. (18)

The core of the instability can be found by studying different norms of the tensor S.

2.6 Kinetic Energy Analysis

By performing an energy analysis, additional information about the instabil- ity mechanism can be extracted. Multiplying equation (9a) with the complex conjugate of the perturbation velocity ˜Uproduces the perturbation kinetic energy which in index notation can be written as:

i∂ ˜Ui

∂t + ˜Uij∂Ubi

∂xj

+ ˜UiUbj∂ ˜Ui

∂xj

= − ˜Ui ∂ ˜p

∂xi

+ 2 ReU˜i

∂xj

beij) + 2 ReU˜i

∂xj

(˜µEij). (19) Note here that Ubiand Ubj are the i:th and j:th component of the base flow velocity Ub. Furthermore, eij and Eij are the perturbation and base flow shear rate tensors respectively:

(12)

eij = 1 2(∂ ˜Ui

∂xj

+∂ ˜Uj

∂xi

) (20)

Eij = 1 2(∂Ubi

∂xj

+∂Ubj

∂xi

) (21)

Subsequently, the kinetic energy budget is:

d(Ekin)

dt = ∂

∂xj

[Ubjii+1

2( ˜Ujp+ ˜˜ Uj)+ 1

Reµb( ˜Uieij+ ˜Uieij)+ 1

ReEij( ˜Uiµ+ ˜˜ Uiµ˜)]

−1

2( ˜Uij+ ˜Uij)∂Ubi

∂xj − 2

Reµb(eijeij) − 1

Re(˜µeijEij + ˜µeijEij) (22) Here Ekin= 12( ˜Uii) is the kinetic energy and the superscript * denotes the complex conjugate of the corresponding quantity. The first term on the right hand side of equation (22) can be interpreted as the kinetic energy transfer throughout the domain. However since there is no net flux in a closed flow system like the cavity, it vanishes. The second and third contribution are the production and dissipation of the perturbation kinetic energy, whereas the last expression is an additional term due the non-Newtonian properties of the flow. Note that the three dimensional effects only come into play in the energy dissipation since Ub3= W = 0 and ∂z = 0 for the base flow Ub.

3 Numerical Method

The numerical computations have been performed using a variant of the second order finite difference code developed by Giannetti and Luchini and described in [23]. To begin with the two-dimensional steady base flow is calculated by discretizing the flow variables on a staggered grid. Equations (4)-(5) are then solved using Newton-Raphson iteration, where the linear equations produced are inverted by means of a sparse LU decomposition.

Next, the base flow solution is inserted into the perturbation equations (9) and the linear stability analysis is executed. Eigenvalues and modes of both the direct and adjoint field are computed by employing the Arnoldi shift and invert method. For all calculations a shift of 2 + 0i has proved sufficient in order to find correct results (including oscillatory modes). Moreover, a eigenvalue tolerance of 10−8 has been chosen to guarantee converged re- sults. As mentioned above, the instability occurs when the real part of the eigenvalue is larger than zero. A non-zero imaginary part corresponds to wave propagating in the z-direction, hence, the perturbation mode is non- stationary. Finally, the sensitivity, and thereby core of the instability, is investigated by multiplying the direct and adjoint fields.

(13)

The eigenvalue strongly depends on Re, Γ, κ, the power-law index n and the time constant λ. The critical Re was found by applying the following strategy. Firstly, Γ, n and λ are set. Secondly, a range in κ is selected, often starting from 0 to 20 (the wavenumber is assumed to be less then 30 since modes with higher κ are expected to be strongly damped [2]) and an incremental step size Dκ usually initially selected to be 1. Thirdly, a bisection method is utilized to find an unstable Re. Two values of Re are chosen, one which is sure to be unstable (Reunstable) and one of which is sure to be stable (Restable). Then, the eigenvalues are checked for all κ at Re(1) = (Reunstable+ Restable)/2. If the real part of all eigenvalues are stable then Restable = Re(1) and the next iteration will start with Re(2) = (Re(1) + Reunstable)/2. However, if an unstable eigenvalue is found for a specific κ(k) then Reunstable= Re(1) and the next run will start at Re(2) = (Restable+Re(1))/2 and κ will only be scanned from κ = κ(k)−Dκ. A similar classification of Re(2) is made and so on until Reunstable− Restable ≤ ∆Re where ∆Re is a specified tolerance. After this, the parameters are refined with the intention to narrow down the interval between stable and unstable Re and thus finding the critical value.

3.1 Mesh

The square cavity of adimensional size Lx = Ly = 1 will be mainly in- vestigated in this work. Unless otherwise stated, a mesh size (nx× ny) of 250×250 has been chosen for this case. Furthermore, a parabolic mapping is selected with the mesh being stretched by a ratio of 4 towards all boundaries of the geometry in both the x and y direction. Grid points are more densely clustered at the walls than in the center of the cavity as seen in 2(a). This turned out to be necessary to obtain correct results in these high-gradient regions.

Analysis of a shallow cavity with aspect ratio Γ = 0.25 has also been performed. The dimensions are therefore set to Lx = 1, Ly = 0.25 where the number of grid points used is 300 × 75. Again parabolic stretching is used. Figure 2(b) shows the grid used for these cases. Note that for the sake of clarity only every fourth point is shown in the figure.

(14)

0 0.5 1 0

0.5 1

(a)

0 0.5 1

0 0.125 0.25

(b)

Figure 2: Mesh used square and shallow cavity. (a) Γ = 1 and (b) Γ = 0.25

3.2 Code Validation

The linear stability analysis has been validated by reproducing the results for Newtonian fluid in [2]. It can be seen from Table 1 that the present results based on the stretched 150 × 150 grid are already in good agreement, with an error about or below 1%.

Table 1: Comparison between the critical values presented by Albensoeder et al.

(141 × 141 grid points)[2] and the present results for a square cavity with resolution (150 × 150).

Γ ReC, [2] ReC Error % κC, [2] κC Error % ωC, [2] ωC Error %

0.25 1152.7 1165 1.07 20.63 20.6 -0.15 2.27 2.25 -0.82

1 786.3 789.1 0.36 15.43 15.2 -1.30 0 0 0

(15)

Grid dependence for the non-Newtonian cases is checked by finding the critical Re and κ when Γ = 1, n = 0.5 and λ = 10 for the several mesh sizes, see Table 2. The values reported in the Table confirm that the choice of grid points yields satisfactory results.

Table 2: Grid independence study. The tables reports the critical values at differ- ent resolution for shear-thinning fluids with n = 0.5 and λ = 10.

Grid Resolution ReC Error % κC Error %

150 × 150 330.7 6.23 13.9 -6.08

200 × 200 317.5 1.99 14.5 -2.03

250 × 250 (Reference) 311.3 - 14.8 -

300 × 300 307.5 -1.22 14.8 0

4 Results

In this section we examine the stability of the square and shallow cavity. Re- sults are presented for values of the power-law index between 0.4 < n < 1.4 and time constant λ =1, 10 and 100. As already mentioned, the remaining constants in the constitutive equation (3) are set to a = 2, ˆµ= 0.001 and ˆ

µ0= 1. When n < 1, the fluid is shear thinning and the viscosity decreases with the strain rate. The opposite is true for shear-thickening fluids defined by values n > 1, while Newtonian fluids are retrieved when n = 1 and the viscosity becomes independent of the strain rate.

4.1 The Square Cavity, Γ = 1

The base-flow streamlines computed by Newton iterations has been initially compared with solutions obtained by direct numerical simulations (DNS) with the code Nek5000 [27]; the data reported in Figure 3 reveal good agree- ment. In the figure, we report as example the streamlines for shear-thinning fluid with n = 0.4, λ = 10 and Re = 456.

(16)

Nekton code

0.2 0.4 0.6 0.8 0.2

0.4 0.6 0.8 1

0 0.2 0.4 0.6 0.8

CPL code

0.2 0.4 0.6 0.8 0.2

0.4 0.6 0.8 1

0 0.2 0.4 0.6 0.8

Figure 3: Comparison of the nonlinear base flow computed with direct numerical simulations (Nekton code) and Newton iterations (CPL code) for n = 0.4, λ = 10 and Re = 456. The contours indicates streamlines.

Once the numerical calcualtions of the base flow is further validated, we proceed to the stability analysis. The neutral curve (critical Re vs. n) is reported in Figure 4(a) for the square cavity and different values of the parameter λ. Shear-thickening effects induce a significant increase of ReC, this effect being more pronounced for larger values of λ. The opposite applies when n < 1; in this case however the critical Re first decreases and then increases when increasing the shear-thinning properties (decreasing n).

The viscosity varies locally inside the fluid and one can therefore define a local Reynolds number

Reloc= ρLxV ˆ µ .

This is displayed in Figure 5 at critical condition for 4 different values of the power index n. The plots reveal that the region of largest Reloc, lowest viscosity, are located at the centre of the cavity and at the lower corners for shear-thickening fluids and for values of n close to 1, whereas they are located close to the upper wall and in a ring all around the cavity when the fluid is characterized by significant shear-thinning, n < 0.6. It is also interesting to consider the average Reynolds number Reavg, defined as

Reavg = R Relocdxdy

A (23)

where Reloc is defined above and A is the total area of the cavity. Again, µ( ˙γ) is given by equation (3). This average Reynolds number is shown in Figure 4(b) versus the index n. The large decrease of the critical Re when decreasing n is significantly reduced when considering Reavg instead.

As discussed below, this suggests that the same instability mechanism is at work for 1.4 > n > 0.6. Conversely, the increase of critical Reynolds number at low n is now more evident. Interestingly, the trend is consistent for all

(17)

values of λ considered and the difference almost disappear when re-scaling the neutral curves with the local viscosity.

0.4 0.6 0.8 1 1.2 1.4 500 1000 1500 2000 2500 3000 3500

4000 λ = 1

λ = 10 λ = 100

(a)

Re

n

0.4 0.6 0.8 1 1.2 1.4 500 1000 1500

2000 λ = 1

λ = 10 λ = 100

(b)

Reavg

n

Figure 4: (a) Critical Reynolds number versus the index n for different values of λ. (b) Average Reynolds number Reavg at neutral conditions versus n.

n=1.4, Re

c=2570.31

0 0.5 1

0 0.5 1

500 1000 1500 2000

n=0.9, Re

c=577.344

0 0.5 1

0 0.5 1

500 1000 1500 2000

n=0.6, Re

c=313.77

0 0.5 1

0 0.5 1

500 1000 1500 2000

n=0.4, Re

c=456.055

0 0.5 1

0 0.5 1

500 1000 1500 2000

Figure 5: Distribution of the local Reynolds number along the neutral curve Γ = 1 and λ = 10.

The effect of the shear-dependent viscosity on the base flow is visualized by the streamwise component Ub displayed in Figure 6. The boundary layer at the lid becomes thinner and thinner when decreasing n while the magnitude of the negative counterflow at the lower wall decreases. This is also associated to weaker vorticity in the centre of the cavity.

(18)

n=1.2

0 0.5 1

0 0.5 1

0 0.5 1

n=0.8

0 0.5 1

0 0.5 1

0 0.5 1

n=0.6

0 0.5 1

0 0.5 1

0 0.5 1

n=0.4

0 0.5 1

0 0.5 1

0 0.5 1

Figure 6: Distribution of the x-component of the baseflow velocity Ub at fixed Re = 600 for Γ = 1 and λ = 10 for the indicated values of n.

The circular frequency and the spanwise wavenumber pertaining to the first instability are reported in Figure 7. Steady modes are the first to become unstable in the case of Newtonian fluid and this is still valid for shear-thinning fluids, see Figure 7(a) . However, non-stationary modes are first unstable for n > 1.2 for all λ investigated. These modes have a lower spanwise wavenumber as displayed in Figure 7(b). Interestingly, a mode of even higher frequency and lower spanwise wave-number κ appears as the most dangerous when n = 1.4. The frequency and wave-number are more or less independent of λ and of the power index n when steady modes appear first.

(19)

0.4 0.6 0.8 1 1.2 1.40 0.2 0.4 0.6

0.8 λ = 1

λ = 10 λ = 100

(a)

ω

n

0.4 0.6 0.8 1 1.2 6

8 10 12 14 16

λ = 1 λ = 10 λ = 100 (b)

κ

n

Figure 7: (a) Frequency ω and (b) critical spanwise wavenumber κ of the first instability mode for different values of λ plotted versus the power index n.

Direct and adjoint modes indicate where in the flow field the perturbation amplitude is maximized and the location of highest receptivity, respectively.

The modulus of the unstable steady modes for n = 1 (Newtonian) and n = 0.4 are displayed in Figure 8. For the Newtonian fluid, velocity perturbations are mainly found on the left side of the cavity; a finding common to modes in the range 1.1 > n > 0.4, where the critical Reynolds number Reavg is almost constant. When further decreasing the power-index n we see that the perturbation is now located on a thinner region evident also on the lower wall and partially on the right side. A second peak is formed in the bottom right corner of the cavity for the v-component of the velocity perturbation.

(20)

umode, n=1

0 0.5 1

0 0.5 1

0 0.01

vmode, n=1

0 0.5 1

0 0.5 1

0 0.01

wmode, n=1

0 0.5 1

0 0.5 1

0 0.01

umode, n=0.4

0 0.5 1

0 0.5 1

0 0.01 0.02

vmode, n=0.4

0 0.5 1

0 0.5 1

5 10 15 x 10−3

wmode, n=0.4

0 0.5 1

0 0.5 1

0 0.005 0.01

Figure 8: Magnitude of the x-, y- and z-velocity components (u, v and w) of the first instability mode for (a) n = 1 and (b) n = 0.4, λ = 10.

The adjoint modes concerning the first instability in shear-thinning fluid are shown in Figure 9. Considering the Newtonian fluid, we see that the region of highest receptivity is located on the corner opposite to that where the disturbance is largest. The adjoint v+ and w+ is stronger on the right wall of the cavity. In the case of strong shear-thinning, the flow is most sensitive to forcing in the x-direction at the upper wall, while normal forcing is most efficient on the left side of the cavity.

(21)

umode + , n=1

0 0.5 1 0.2

0.4 0.6 0.8

24 68 1012 x 10−3

vmode + , n=1

0 0.5 1 0.2

0.4 0.6 0.8

0.005 0.01 0.015 0.02

wmode + , n=1

0 0.5 1 0.2

0.4 0.6 0.8

2 46 810 x 10−3

umode + , n=0.4

0 0.5 1 0.2

0.4 0.6 0.8

5 10 15 x 10−3

vmode + , n=0.4

0 0.5 1 0.2

0.4 0.6 0.8

24 68 1012 x 10−3

wmode + , n=0.4

0 0.5 1 0.2

0.4 0.6 0.8

2 4 6 8 x 1010 −3

Figure 9: Magnitude of the x-, y- and z components (u+, v+ and w+) of the adjoint of the first instability mode for (a) n = 1 and (b) n = 0.4, λ = 10.

The distribution of the time-periodic modes observed for shear-thickening fluids is reported in Figure 10. Here we show results for n = 1.2 and n = 1.4, the latter associated to higher frequency and longer spanwise scale, as dis- cussed above. For n = 1.2, the velocity fluctuations are mainly located on the left side of the cavity, as for the unstable steady modes; the area of large disturbance being now wider. When further increasing n = 1.4, the mode appears now as a large vortex in the centre of the cavity. This mode enhances and decreases the amplitude of the base flow vortex periodically in time and spanwise direction.

(22)

umode, n=1.4

0 0.5 1

0 0.5 1

2 4 6 8 10 x 10−3

vmode, n=1.4

0 0.5 1

0 0.5 1

0 0.005 0.01

wmode, n=1.4

0 0.5 1

0 0.5 1

0 5 x 10−3

umode, n=1.2

0 0.5 1

0 0.5 1

0 5 10 x 1015 −3

vmode, n=1.2

0 0.5 1

0 0.5 1

24 68 1012 x 10−3

wmode, n=1.2

0 0.5 1

0 0.5 1

0 5 10 x 10−3

Figure 10: Magnitude of the x-, y- and z-velocity components (u, v and w) of the first instability mode for (a) n = 1.4 and (b) n = 1.2, λ = 10.

The adjoint of the non-stationary modes are reported in Figure 11. In both cases, the receptivity to forcing in the y-direction is strongest and located on the right side of the cavity, as for shear-thinning fluids. The receptivity to momentum forcing in x is instead more diffuse, peaking at the lower or upper wall when varying n. Forcing in the spanwise direction is more effective at the right wall.

(23)

umode + , n=1.4

0 0.5 1 0.2

0.4 0.6 0.8

2 4 6 8 x 10−3

vmode + , n=1.4

0 0.5 1 0.2

0.4 0.6 0.8

0.005 0.01 0.015 0.02

wmode + , n=1.4

0 0.5 1 0.2

0.4 0.6 0.8

2 4 6 8 x 10−3

umode + , n=1.2

0 0.5 1 0.2

0.4 0.6 0.8

2 4 6 8 x 10−3

vmode + , n=1.2

0 0.5 1 0.2

0.4 0.6 0.8

0.005 0.01 0.015 0.02

wmode + , n=1.2

0 0.5 1 0.2

0.4 0.6 0.8

2 4 6 8 x 10−3

Figure 11: Magnitude of the x-, y- and z components (u+, v+ and w+) of the adjoint of the first instability mode for (a) n = 1.4 and (b) n = 1.2, λ = 10.

Next, the structural sensitivity is presented. As introduced above, its distribution provides important information about the instability mecha- nism. Results for both shear-thinning and thickening fluids are presented in Figure 12. As we increase shear-thinning, the region at the core of the instability is getting thinner but still consists of a ring inside the cavity.

In the case of shear-thickening fluid and unsteady modes, we see that the wave-maker is located on the lower-left corner.

(24)

n=1.4, Re

c=2570.31, κ=5.9

0.2 0.4 0.6 0.8 0.2

0.4 0.6 0.8

0 1 2 3 4 5

n=1.2, Re

c=1492.2, κ=7.6

0.2 0.4 0.6 0.8 0.2

0.4 0.6 0.8

0 1 2 3 4 5

n=0.7, Re

c=353.516, κ=13.9

0.2 0.4 0.6 0.8 0.2

0.4 0.6 0.8

0 1 2 3 4 5

n=0.4, Re

c=456.6, κ=15.9

0.2 0.4 0.6 0.8 0.2

0.4 0.6 0.8

0 1 2 3 4 5

Figure 12: Structural sensitivity for the first instability along the neutral curve for λ = 10 and the indicated values of n.

We now present the analysis of the perturbation kinetic energy budget.

As introduced in section 2.6 the kinetic energy budget, i.e. equation (22), is given by the sum of production, dissipation and additional production terms related to the varying viscosity. The density of energy production associated to the base flow shear is reported in Figure 13, again for λ = 10.

We choose here to display 4 representative cases: two unsteady modes in shear-thickening fluids, n = 0.7 indicative of Newtonian and weakly shear- thinning fluids and n = 0.4 representing the region of increasing critical Reynolds number. One can see from the figure that the Newtonian fluid actually has the largest energy production and that its spatial distribution is almost independent on the power index n. The spatial distribution of the additional energy production due to non-Newtonian effects is shown in Figure 14. The additional term is, per definition, positive for shear-thinning fluids and negative for shear-thickening fluids. It is interesting to note that this additional term is most relevant on the left and lower side of the cavity.

(25)

n=1.4, Rec=2570.3, κ=5.9

0 0.2 0.4 0.6 0.8 0

0.5 1

0 5 10 15

n=1.2, Rec=1492.2, κ=7.6

0 0.2 0.4 0.6 0.8 0

0.5 1

0 5 10 15

n=0.7, Re

c=353.5, κ=13.9

0 0.2 0.4 0.6 0.8 0

0.5 1

0 5 10 15

n=0.4, Re

c=456.6, κ=15.9

0 0.2 0.4 0.6 0.8 0

0.5 1

0 5 10 15

Figure 13: Density of production of perturbation kinetic energy for selected values of n and λ = 10 at neutral conditions.

n=1.4, Re

c=2570.3, κ=5.9

0 0.2 0.4 0.6 0.8 0

0.5 1

−0.5 0 0.5 1 1.5

n=1.2, Re

c=1492.2, κ=7.6

0 0.2 0.4 0.6 0.8 0

0.5 1

−0.5 0 0.5 1 1.5

n=0.7, Rec=353.5, κ=13.9

0 0.5

0 0.5 1

−0.5 0 0.5 1 1.5

n=0.4, Rec=456.6, κ=15.9

0 0.5

0 0.5 1

−0.5 0 0.5 1 1.5

Figure 14: Density of the additional energy production term related to varying viscosity. It is reported for the indicated values of n and λ = 10 at neutral conditions.

The sum of all production and dissipation terms is displayed in Figure 15 for the same cases as before and at neutral conditions. Production is dom- inating in the lower left corner for intermediate values of n. In the case of strong shear-thinning, the peak of total production is on the lower right cor- ner, whereas for unsteady modes (n = 1.4) the peak of positive production

(26)

is more diffuse towards the lower left corner. Negative production, dissipa- tion, appears usually in this layers close to the regions of highest positive production and on the upper left corner. It is also interesting to note that in this close configuration the wave-maker of the instability and the region of largest production of perturbation kinetic energy almost overlap; this was not the case for the cylinder flow as shown in [24].

n=1.4, Rec=2570.3, κ=5.9

0 0.2 0.4 0.6 0.8 0

0.5 1

−5 0 5 10 15

n=1.2, Rec=1492.2, κ=7.6

0 0.2 0.4 0.6 0.8 0

0.5 1

−5 0 5 10 15

n=0.7, Re

c=353.5, κ=13.9

0 0.2 0.4 0.6 0.8 0

0.5 1

−5 0 5 10 15

n=0.4, Re

c=456.6, κ=15.9

0 0.2 0.4 0.6 0.8 0

0.5 1

−5 0 5 10 15

Figure 15: Density of the total energy production for the indicated values of n and λ = 10 at neutral conditions.

Finally, we integrate the densities of the different energy terms over the domain. The results are shown in Figure 16 for λ = 10 at neutral condi- tion and for fixed Re = 600 and κ = 15. In 16(a) we see that production associated to the base flow shear and dissipation have maximum around n = 1 and decrease when adding shear-dependent viscosity. The decrease in dissipation magnitude can be associated to the vorticity of the instability mode, while the reduction in production to the localization of the perturba- tion when n < 1 and to the weaker base shear when n > 1. The additional production term becomes relevant when n < 0.6. It is also instructive to study the energy budget at fixed Reynolds number. In this case, the sum of the different terms reveals whether the mode is stable or unstable since it can be related to the real part of the eigenvalue. For the case depicted in the figure, the flow is unstable for 0.9 > n > 0.4. In addition to the expected increase of the additional production for lower values of n, we no- tice a decrease of the classic production with shear-thinning. This is due to the spatial de-correlation between the region of largest disturbance (lower and left wall, cf. figure 8) and the region of largest shear (upper wall, cf.

figure 6).

(27)

0.4 0.6 0.8 1 1.2

−21.4

−1 0 1 2

(a)

E

n

0.2 0.4

0.6 0.8

−31

−2

−1 0 1 2

Production Dissipation Additional Production Sum

(b)

E

n

Figure 16: Budget of production of perturbation kinetic energy for λ = 10. (a) Budget at neutral condition and (b) at fixed Re = 600 and κ = 15.

Combining the results presented above, we can conclude that the in- stability mechanism is not significantly changed for weak shear-thinning and shear-thickening effects. Indeed, the Reavg is almost constant in this range, steady modes are the most unstable and the instability wave-maker is similar. When further increasing the shear-thinning properties, we see a surprising increase of the critical Reynolds number based on the zero-shear- rate viscosity. This can be explained by the fact that the region of largest shear become more and more localized close to the wall and the base-flow vorticity is weaker towards the centre of the cavity and on the left side: in this case very large local Reynolds number (very low local viscosity) is nec- essary to overcome dissipation with the relevant contribution from the extra production terms associated to the shear-thinning effects. Interestingly, we notice that unsteady modes are the first to become unstable when the power index n is above 1.2 with modes significantly longer in the spanwise direc- tion. Finally, we note that the instability characteristics are found to vary with the power index n, while no significant qualitative variations are found with respect to the time constant λ.

4.2 The Shallow Cavity, Γ = 0.25

We now consider the shallow cavity of aspect ratio Γ = 0.25, as in [2], where the first instability mode is time-periodic in Newtonian fluids. Computations have been performed with λ = 10; as shown above this parameter does not seem to significantly affect the physics of the instability, the main variations coming from the power index n.

The neutral curves are first displayed in Figure 17 both in terms of zero-shear rate viscosity, Reynolds number Re, and average viscosity, aver- age Reynolds number Reavg. The critical Reynolds number decreases from

(28)

shear-thickening to shear-thinning fluids (decreasing n) whereas critical val- ues of the average Reynolds number first decrease and then increase with a minimum at n ≈ 0.7, similarly to what observed for the square cavity.

0.4 0.6 0.8 1

1.2 1000 2000 3000 4000

(a)

Re

n

0.4 0.6 0.8 1

1.2 500 1000 1500 2000

(b)

Reavg

n

Figure 17: (a) Critical Reynolds number versus the index n and (b) Average Reynolds number Reavg at neutral conditions versus n for Γ = 0.25 and λ = 10.

The effect of shear-thinning viscosity on the two-dimensional base flow is depicted in Figure 18 where we report the x-component of the baseflow velocity Ub for 4 values of n. As n decreases we clearly see that the main vortex moves to the right side of the cavity. In addition, and in analogy to the case of the square cavity, the boundary layer at the moving lid becomes thinner at lower n.

n=1

0 0.5 1

0 0.1 0.2

−0.5 0 0.5 1 n=1.3

0 0.5 1

0 0.1 0.2

−0.5 0 0.5 1

n=0.8

0 0.5 1

0 0.1 0.2

−0.5 0 0.5 1

n=0.5

0 0.5 1

0 0.1 0.2

−0.5 0 0.5 1

Figure 18: Distribution of the x-component of the baseflow velocity Ub at fixed Re = 500 for Γ = 0.25, λ = 10 and the values of n indicated.

Unlike the square cavity, for the shallow cavity all critical modes are oscillatory. The frequency and spanwise wavenumber of the neutral modes are reported in Figure 19; here one can see that the value for κc and ωc

drops continuously with decreasing power-law index , from about κ = 21.2 and ω = 2.34 respectively, until suddenly at n = 0.4 the first instability

(29)

appears to have a long-wavelength (κ ≈ 11) and low-frequency (ω ≈ 0.7).

0.4 0.6 0.8 1

1.2 1 1.5 2 2.5

(a)

ω

n

0.4 0.6 0.8 1

1.2 12 14 16 18 20 22

(b)

κ

n

Figure 19: (a) Frequency ω and (b) critical spanwise wavenumber κ of the first instability mode plotted versus the power index n for Γ = 0.25 and λ = 10

The shape of the direct modes does is shown in Figure 20 for shear- thinning fluids. The region of largest fluctuations is seen to move from the top-right corner to the centre of the cavity (left side of the big vortex dis- played in Figure 18) when shear-thinning increases and the instability mode has a significantly lower frequency. In all cases, velocity fluctuations are detectable on the right-half of the cavity. The corresponding adjoint modes, depicted in Figure 21, reveal a certain overlap with the direct modes; inter- estingly, forcing in the y-direction is one order of magnitude most effective than forcing in the other two directions when n = 0.9, whereas acting in the x-direction at the lower wall is the most effective way to excite the slow oscillatory mode at n = 0.4.

References

Related documents

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

Från den teoretiska modellen vet vi att när det finns två budgivare på marknaden, och marknadsandelen för månadens vara ökar, så leder detta till lägre

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

I regleringsbrevet för 2014 uppdrog Regeringen åt Tillväxtanalys att ”föreslå mätmetoder och indikatorer som kan användas vid utvärdering av de samhällsekonomiska effekterna av

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating

Re-examination of the actual 2 ♀♀ (ZML) revealed that they are Andrena labialis (det.. Andrena jacobi Perkins: Paxton &amp; al. -Species synonymy- Schwarz &amp; al. scotica while

The ambiguous space for recognition of doctoral supervision in the fine and performing arts Åsa Lindberg-Sand, Henrik Frisk &amp; Karin Johansson, Lund University.. In 2010, a