• No results found

Global stability analysis of three-dimensional boundary layer flows

N/A
N/A
Protected

Academic year: 2021

Share "Global stability analysis of three-dimensional boundary layer flows"

Copied!
42
0
0

Loading.... (view fulltext now)

Full text

(1)

Global stability analysis of three-dimensional

boundary layer flows

by

Mattias Brynjell-Rahkola

October 2015 Technical Reports Royal Institute of Technology

Department of Mechanics SE-100 44 Stockholm, Sweden

(2)

Akademisk avhandling som med tillst˚and av Kungliga Tekniska H¨ogskolan i Stockholm framl¨agges till offentlig granskning f¨or avl¨aggande av teknologie licentiatsexamen fredagen den 30 Oktober 2015 kl 13.00 i sal D3, Kungliga Tekniska H¨ogskolan, Lindstedtsv¨agen 5, Stockholm.

c

Mattias Brynjell-Rahkola 2015

(3)
(4)

Global stability analysis of three-dimensional boundary

layer flows

Mattias Brynjell-Rahkola

Linn´e FLOW Centre and Swedish e-Science Research Centre (SeRC), KTH Mechanics, Royal Institute of Technology

SE-100 44 Stockholm, Sweden

Abstract

This thesis considers the stability and transition of incompressible boundary layers. In particular, the Falkner–Skan–Cooke boundary layer subject to a cylindrical surface roughness, and the Blasius boundary layer with applied lo-calized suction are investigated. These flows are of great importance within the aviation industry, feature complex transition scenarios, and are strongly three-dimensional in nature. Consequently, no assumptions regarding homogeneity in any of the spatial directions are possible, and the stability of the flow is governed by an extensive three-dimensional eigenvalue problem.

The stability of these flows is addressed by high-order direct numerical simulations using the spectral element method, in combination with a Krylov subspace projection method. Such techniques target the long-term behavior of the flow and can provide lower limits beyond which transition is unavoidable. The origin of the instabilities, as well as the mechanisms leading to transition in the aforementioned cases are studied and the findings are reported.

Additionally, a novel method for computing the optimal forcing of a dynam-ical system is developed. This type of analysis provides valuable information about the frequencies and structures that cause the largest energy amplifica-tion in the system. The method is based on the inverse power method, and is discussed in the context of the one-dimensional Ginzburg–Landau equation and a two-dimensional flow case governed by the Navier–Stokes equations.

Descriptors: Hydrodynamic stability, transition to turbulence, global analy-sis, boundary layers, roughness, laminar flow control, Stokes/Laplace precon-ditioner, optimal forcing, crossflow vortices, Ginzburg–Landau, Falkner–Skan– Cooke, Blasius, lid-driven cavity.

(5)

Global stabilitetsanalys av tredimensionella gr¨

ansskikts-str¨

omningar

Mattias Brynjell-Rahkola

Linn´e FLOW Centre and Swedish e-Science Research Centre (SeRC), KTH Mechanics, Royal Institute of Technology

SE-100 44 Stockholm, Sweden

Sammanfattning

Denna avhandling behandlar stabilitet och transition i inkompressibla gr¨ans-skikt. Specifikt studeras ett Falkner–Skan–Cooke gr¨ansskikt med en cylin-drisk ytoj¨amnhet och ett Blasius gr¨ansskikt med lokal sugning. B˚ada dessa str¨omningar ¨ar av stor betydelse inom flygindustrin, uppvisar ett komplext transitionsf¨orlopp, och ¨ar fullst¨andigt tredimensionella i sin karrakt¨ar. Detta medf¨or att inga antaganden g¨allande homogenitet i n˚agon av de spatiella rikt-ningarna kan g¨oras, och att str¨omningens stabilitet best¨ams av ett omfattande tredimensionellt egenv¨ardesproblem.

Stabiliteten hos str¨omningen ¨ar unders¨okt med h¨ogordningsmetoder in-nefattande direkta numeriska simleringar med en spektralelementmetod, i kom-bination med en Krylov projektionsmetod. Med dessa tekniker kan beteendet hos str¨omningen under l˚angsiktiga f¨orh˚allanden studeras, och nedre gr¨ansv¨ar-den best¨ammas f¨or vilka transition ¨ar oundvikligt. Ursprunget till instabilitet-erna, samt bakomliggande mekanismer till transition i de ovann¨amnda fallen studeras, och resultaten rapporteras.

Ut¨over detta, har en ny metod f¨or att best¨amma optimala krafter p˚a ett dynamiskt system utvecklats. Den h¨ar typen av analys ger v¨ardefull informa-tion om vilka frekvenser och kraftdistribuinforma-tioner som ger st¨orst energitillv¨axt i systemet. Metoden ¨ar baserad p˚a inversiterationer med skift och diskuteras med anknytning till den endimensionella Ginzburg–Landau ekvationen och ett tv˚adimensionellt str¨omningsfall beskrivet av Navier–Stokes ekvationer.

Nyckelord: Hydrodynamisk stabilitet, transition till turbulens, global analys, gr¨ansskikt, ytoj¨amnhet, lamin¨ar str¨omningskontroll, Stokes/Laplace prekondi-tionerare, optimal kraft, korsfl¨odesvirvlar, Ginzburg–Landau, Falkner–Skan– Cooke, Blasius, topp-driven kavitet.

(6)

Preface

The present thesis deals with global stability analysis of three-dimensional shear flows. An emphasis is put on boundary layers, which can be found in a wide range of applications. In the first part, concepts that are central to the topic are introduced, along with a discussion on the numerical techniques that have been implemented and/or utilized. In the second part, three papers are ap-pended. In the case of submitted articles, these have been adjusted to comply with the present format, but not modified in content.

Paper 1. M. Brynjell-Rahkola, L. Tuckerman, P. Schlatter, & D. S. Henningson, 2015. A method for computing optimal forcing of convec-tively unstable flows using Laplace preconditioning. Submitted to SIAM J. Sci. Comput.

Paper 2. M. Brynjell-Rahkola, N. Shahriari, P. Schlatter, A. Han-ifi and D. S. Henningson, 2015. Onset of global instability behind distributed surface roughness in a Falkner–Skan–Cooke boundary layer. Internal report Paper 3. M. Brynjell-Rahkola, E. Barman, A. Peplinski, A. Hanifi and D. S. Henningson, 2015. On the stability of flat plate boundary layers subject to localized suction. Internal report

October 2015, Stockholm Mattias Brynjell-Rahkola

(7)

Division of work between authors

The main advisor for the project is Prof. Dan S. Henningson (DH). Dr. Philipp Schlatter (PS) and Prof. Ardeshir Hanifi (AH) act as co-advisors.

Paper 1

Code development, simulations and data post-processing was performed by Mattias Brynjell-Rahkola (MBR) under supervision of Laurette Tuckerman (LT). The paper was written by MBR and LT with input from PS and DH. Paper 2

Code development was done by MBR and simulations were set-up by Nima Shahriari (NS). Simulations and data post-processing was performed by MBR and NS, with feedback from PS, AH and DH. The paper was written by MBR and NS with input from PS, AH and DH.

Paper 3

The simulations were set-up by Adam Peplinski (AP) and run by Emelie Bar-man (EB) under supervision of MBR. Post-processing was done by MBR and EB, with feedback from AH and DH. The paper was written by MBR with input from AH and DH.

(8)
(9)

Contents

Abstract iv

Sammanfattning v

Preface vi

Part I - Overview and summary

Chapter 1. Introduction 3

Chapter 2. Governing equations 5

2.1. Linearized Navier–Stokes equations 5

2.2. State space formulation 5

Chapter 3. Numerical approach 7

3.1. Spectral element method 7

3.2. Fractional step method 9

3.3. Timestepper approach 11

3.4. Stokes preconditioner 12

Chapter 4. Global stability analysis 15

4.1. Modal analysis 15

4.2. Non-modal analysis 17

Chapter 5. Summary of the papers 22

Chapter 6. Conclusions and outlook 25

Acknowledgements 27

Bibliography 28

Part II - Papers

(10)

Paper 1. A method for computing optimal forcing of convectively unstable flows using Laplace preconditioning 37 Paper 2. Onset of global instability behind distributed surface

roughness in a Falkner–Skan–Cooke boundary layer 59 Paper 3. On the stability of flat plate boundary layers subject

to localized suction 95

(11)

Part I

(12)
(13)

CHAPTER 1

Introduction

Since the early work by Rayleigh, the theory of hydrodynamic stability has developed into a refined subject over the past century, owing to important contributions by Reynolds, Orr, Sommerfeld, Tollmien, Squire, and many oth-ers (Drazin & Reid 1982). Thanks to the increased availability of large-scale computers and development of software with good scalability properties, the concepts of hydrodynamic stability originally developed for one-dimensional parallel flows governed by the Orr–Sommerfeld equation, have been extended to problems with several inhomogeneous spatial dimensions governed by the Navier–Stokes equations. In particular, stability analysis of three-dimensional flows (e.g. global stability analysis) began to emerge in literature in the begin-ning of the twenty-first century, e.g. prolate spheroid (Tezuka & Suzuki 2006), three-dimensional airfoil (Mack et al. 2008), jet in crossflow (Bagheri et al. 2009c) and three-dimensional lid-driven cavity flow (Giannetti et al. 2009).

Ever since the appearance of these works, efforts have been devoted to ex-tend the applicability range of global analysis to more complex flows of interest within engineering and industrial applications. Such flows typically place high demands on the computational mesh and cannot be realized in plain rectangu-lar domains subject to appropriate boundary conditions, or by simply extruding a two-dimensional mesh in the spanwise direction. With an increased complex-ity of the problem, the range of transition patterns typically widens due to an involved interaction between various physical phenomena and mechanisms. Accurate reproduction of the transition scenario, and unraveling of instability mechanisms active in such cases raises critical questions such as how numerical treatment of boundaries and finite domains affect the outcome of the analysis. As an example, a seemingly simple flow case like the jet in crossflow has been found to be susceptible to elliptic, Kelvin–Helmholtz and von K´arm´an insta-bilities (Bagheri 2010), as well as to exhibit an extreme sensitivity to spatial resolution and domain size (Peplinski et al. 2015). To date, a complete under-standing of these issues is missing, hence requiring continued research on the topic, alongside development of new algorithms and methods for addressing them. For an account of some of these challenges and possibilities, and a sum-mary of findings from studies involving stability of two- or three-dimensional flows, see the reviews by Theofilis (2011) and G´omez et al. (2012).

(14)

4 1. INTRODUCTION

In order to discuss these topics further, a precise definition of global stability is required. Such a definition is provided by Schmid & Henning-son (2001) based on the kinetic energy of a disturbance u′

k in the domain, EΩ= (1/2)R(u′ku′k)dV . A basic state Uk is stable to perturbations if

lim t→∞

EΩ(t) EΩ(0)

= 0, (1.1)

and furthermore globally stable, if there exists a threshold energy δ > 0 such that (1.1) holds for limδ→∞EΩ(0) < δ. In terms of Reynolds number (Re), a flow is referred to as globally stable if Re < ReGwhere ReG denotes the lowest Reynolds number for which turbulence can be sustained. However, as further argued, such a definition of ReG need not hold for all flows, in which case it is necessary to introduce ReT, for which the flow will relaminarize (Schmid & Henningson 2001).

A different definition of global stability is based on local stability analysis. Here the existence of a region of local absolute instability is concluded to be a necessary but insufficient condition for the existence of a global instability. Consequently, a basic state will be globally stable if there is no region of absolute instability in the domain, i.e. the response to a localized initial impulse (Green’s function) does not tend to infinity with time at any fixed location x in the laboratory frame (Huerre & Monkewitz 1990; Chomaz 2005).

Provided that a certain flow at hand is globally unstable, transition to turbulence in a non-linear direct numerical simulation (DNS) is inevitable given that the domain is sufficiently long and well-resolved. A schematic picture of various transition scenarios is given in figure 1.1, where the different routes to turbulence have been ordered from left to right with increasing disturbance amplitude (Saric et al. 2002). The work presented herein is mainly concerned with the study of primary modes (known to dominate the transition scenario in low-disturbance environments), and growth due to harmonic forcing (associated with transient growth).

Environmental disturbances

Receptivity

Primary modes Transient growth

Secondary mechanisms

Breakdown

Turbulence

amplitude

(15)

CHAPTER 2

Governing equations

2.1. Linearized Navier–Stokes equations

This thesis concerns the incompressible Navier–Stokes equations with constant material properties, ∂u∗ ∂t∗ + u ∗ · ∇u∗=1 ρ∇p ∗+ ν ∇2u+ f(2.1a) ∇ · u∗= 0. (2.1b)

To simplify the comparison between results for different fluids, as well as be-tween experiments and numerics, it is customary to non-dimensionalize these equations using a suitable set of reference quantities. One such set of quanti-ties appropriate for boundary layers is the free-stream velocity U∗

∞, the bound-ary layer displacement thickness δ∗ and the fluid density ρ, i.e. u = u/U

∞, p = p∗/(ρU∗2

∞), x = x∗/δ∗and t = t∗(U∞∗/δ∗).

Equation (2.1) is non-linear in nature due to the presence of the advective term on the left-hand side. However, for transition scenarios involving growth of primary eigenmodes (see figure 1.1), the initial stages of the transition pro-cess will be linear. In fact, the energy growth of finite amplitude perturba-tions is completely governed by linear mechanisms, whereas the non-linearity of Navier–Stokes merely redistributes energy among different flow structures (Henningson 1996). To study whether small perturbations will grow or decay, the fluid velocity and pressure are decomposed into a steady base flow and a small perturbation, i.e. u = U + εu′ and p = P + εp′ where ε≪ 1. These de-compositions inserted into (2.1) with the above non-dimensionalization, yields the linearized Navier–Stokes equations

∂u′ ∂t + (U· ∇)u ′+ (u′ · ∇)U = −∇p′+ 1 Reδ∗∇ 2u+ f(2.2a) ∇ · u′ = 0, (2.2b)

where the non-linear perturbation term (u′· ∇)u′ has been dropped from the right-hand side of (2.2a).

2.2. State space formulation

An important result regarding the incompressible Navier–Stokes equations is the fact that any state of the dynamical system (2.1) is completely determined by the velocity field, while the pressure can be treated as a dependent variable.

(16)

6 2. GOVERNING EQUATIONS

As reviewed by Deville et al. (2002), the pressure serves as a Lagrange multiplier such that the solution to the corresponding Stokes problem satisfies a saddle-point problem (i.e. the inf-sup condition).

According to the Helmholtz–Hodge decomposition theorem (see Chorin & Marsden 2000), a vector field w on some domain Ω, may be uniquely decom-posed in the form w = u′ +∇p′, where ∇ · u= 0 and u· n = 0 on ∂Ω. As a consequence of the theorem it follows that u′ and ∇p′ are orthogonal. By introducing a projection operator P that orthogonally projects the vector field w onto its divergence-free component u′, one has P(w) = P(u′) = u′. Applying this to the incompressible Navier–Stokes equations (volume forces f′ are neglected) gives

P ∂u

∂t +∇p

′= P

−(U · ∇)u′− (u′· ∇)U +Re1 δ∗∇ 2u′ ⇐⇒ ∂u′ ∂t +✘✘ ✘✘✘✿0

P(∇p) = P (−(U · ∇)u− (u· ∇)U)

| {z } ≡NUu′ + P  1 Reδ∗∇ 2u′ | {z } ≡Lu′ . (2.3a)

Thus, the pressure has been eliminated from the governing equations, which can be written as the sum of an advective and a diffusive operator acting on the divergence-free velocity u′,

∂u′

∂t =NUu ′+

Lu′≡ AUu′. (2.4)

This so-called state space formulation is useful when analyzing the linear stabil-ity of a flow. Given a solution to equation (2.4), the perturbation pressure field may be recovered by solving a Poisson problem subject to suitable boundary conditions (see e.g. Kreiss et al. 1994).

(17)

CHAPTER 3

Numerical approach

3.1. Spectral element method

The spectral element method (SEM) was first introduced by Patera (1984) and combines the geometrical flexibility of finite element methods with the high-order accuracy provided by spectral methods, thus making it a suitable method for studying stability and transition in complex flow cases of interest in many engineering applications. Although the original SEM as presented by Patera featured Chebyshev polynomials as basis functions, the implementation adopted in Nek5000 (Fischer et al. 2008) is based on Legendre polynomials.

Since the SEM is a Galerkin method, the Navier–Stokes equa-tions are solved using a weak formulation. Considering the non-dimensionalization of (2.1) discussed in chapter 2, the weak problem reads: Find(u, p)∈ H1 b(Ω)d× L20(Ω) such that ∂ ∂thw, ui + hw, u · ∇ui = h∇ · w, pi − 1 Reδ∗h∇w, ∇ui + hw, fi ∀w ∈ H 1 0(Ω)d (3.1a) −hq, ∇ · ui = 0 ∀q ∈ L2 0(Ω), (3.1b)

whereh·, ·i denotes the inner product, L2

0(Ω) is a space of functions that are L2 -integrable on the domain Ω with a zero average value, and H1

b(Ω)dand H01(Ω)d are Sobolev spaces of vector functions having d components and satisfying in-homogeneous and in-homogeneous boundary conditions, respectively (see Deville et al. 2002; Karniadakis & Sherwin 2013 for further details). The boundary conditions of (3.1) will be discussed in the next section.

Discretization of (3.1) involves dividing the domain Ω into E non-overlapping quadrilateral elements Ωe(e = 1, . . . , E). In order to avoid spuri-ous pressure modes (Canuto et al. 2007), Maday & Patera (1989) suggested a staggered SEM referred to as PN-PN −2, where the velocity is represented on N + 1 Gauss–Lobatto–Legendre (GLL) quadrature points and the pressure on N− 1 Gauss–Legendre (GL) quadrature points. Assuming a three-dimensional flow case (d = 3), the discrete velocity in element Ωe mapped onto the three-dimensional reference element ˆΩ = [−1, 1]3can be written

u(x(ξ, ζ, η))|Ωe = N X i=0 N X j=0 N X k=0 uei,j,khN,i(ξ)hN,j(ζ)hN,k(η) (ξ, ζ, η)∈ ˆΩ, (3.2) 7

(18)

8 3. NUMERICAL APPROACH

where hN,i, hN,j and hN,k are one-dimensional N th-order Lagrange inter-polants based on Legendre polynomials, x is a mapping function of the local geometry and ue

i,j,k are the unknown nodal values in Ωe(Deville et al. 2002). Since the Legendre polynomial is one of the classical orthogonal polyno-mials (the Hermite polynomial considered in paper 1 being another one), the inner products in (3.1) can be efficiently evaluated using GLL- and GL-quadrature rules.

3.1.1. Boundary conditions

In the derivation of (3.1), boundary terms arise from the pressure and the diffusive terms, − I Γ (w· n) pdS +Re1 I Γ w· ∂u ∂ndS = I Γ w· n ·  1 Reδ∗∇u − pI  dS, (3.3) with Γ = ∂Ω. It is desirable to set (3.3) to zero. For boundaries equipped with Dirichlet conditions this is achieved by the choice of test functions. For other boundary conditions e.g. outflow, n· (Re−1δ∗ ∇u − pI) = 0. On the rear outflow boundary, these equations become

1 Reδ∗ ∂u ∂x − p = 0; 1 Reδ∗ ∂v ∂x = 0, (3.4)

and on the free-stream boundary 1

Reδ∗ ∂v

∂y− p = 0; u(x, ytop) = U∞, (3.5) with the second equation of (3.5) replacing Re−1δ∗(∂u/∂y) = 0. Equation (3.4), (3.5) apply to non-accelerating boundary layers such as Blasius, and are used in paper 3. On the other hand, for accelerating (or decelerating) boundary layers such as Falkner–Skan (FS), where the external flow obeys a power law U∞(x) = Cxm, the velocity components read

ufs= U∞(x)df dη, (3.6a) vfs= 1 2 s 2 m + 1 U∞(x) x 1 Reδ∗ 0  (1− m)df dηη− (1 + m)f  , (3.6b)

in which f is obtained from a similarity solution (Falkner & Skan 1931) and η here denotes a dimensionless wall-normal boundary layer coordinate. Now the pressure cannot be expected to balance velocity gradients on the free-stream and outflow boundaries. In order to handle this type of flow, (3.4) and (3.5) need to be modified to enable an ambient pressure pa to be specified, i.e.

1 Reδ∗ ∂u ∂x− p = pa; 1 Reδ∗ ∂v ∂x = 0 (3.7) 1 Reδ∗ ∂v

(19)

3.2. FRACTIONAL STEP METHOD 9 0 0.5 1 1.5 0 1 2 3 4 5 6 ufs, u y

(a) Streamwise velocity component.

-4 -3 -2 -1 0 x 10-3 0 1 2 3 4 5 6 vfs, v y

(b) Wall-normal velocity component.

Figure 3.1: Verification of boundary condition implementation. Profiles extracted from a non-linear DNS at streamwise position x ={0.0, 100.0, 200.0, 300.0} are plotted in order with symbols (×, , ∗, ◦), together with the corresponding FS similarity solution plotted with lines (—–,- - -,· · · ,-·-·-).

From (3.4) the pressure along the wall is seen to be zero since ∂u/∂x = 0. By setting pa = 0 in this point, the pressure will be zero in the lower outflow corner also for (3.7), and hence along the entire outflow boundary following the boundary layer approximation (Schlichting 1979). Taking the upper out-flow corner as a reference point and setting the velocities in this point equal to the corresponding FS solution, i.e. uref = ufs(xoutflow, ytop), the pressure along the free-stream boundary can be estimated using the inviscid Bernoulli equation, ˜p(x) = (1/2)[uT

ref· uref− u(x, ytop)T · u(x, ytop)]. This finally gives an expression for the ambient pressure, which can be plugged into (3.8), pa= (1/Reδ∗

0)(∂vfs/∂y)− ˜p.

As seen in figure 3.1, very good agreement is obtained between the DNS im-plementing (3.7)–(3.8) and the FS similarity solution. The boundary conditions defined by (3.7)–(3.8) are used for the Falkner–Skan–Cooke (FSC) (Falkner & Skan 1931; Cooke 1950) simulations in paper 2. A slight discrepancy in the boundary conditions still exist since ∂v/∂x6= 0 in the free-stream. However, as seen in figure 3.1b the wall-normal component is orders of magnitude smaller than the streamwise component and its rate of change with x can be assumed negligible.

3.2. Fractional step method

Having discussed the spectral element discretization briefly, the fractional step method used to advance the flow in time is considered. In the case of

(20)

10 3. NUMERICAL APPROACH

PN− PN −2, Nek5000 essentially solves the non-linear and linearized Navier– Stokes equations following the same steps. For the purpose of the present discussion, linearized equations will be assumed and only a first order time discretization will be considered, although the derivation readily extends to schemes of higher order.

Applying a first order backward differentiation time-stepping scheme (Euler backward) for the diffusive term and a first order extrapolation time-stepping scheme (Euler forward) for the advective term, the discretized version of (2.2) reads Bβ0u ′n− β 1u′n−1 ∆t = C(U)u ′n−1+ DTp′n+ 1 Reδ∗ Au′n+ Bfn (3.9a) −Du′n= 0. (3.9b)

In (3.9), the operator B is the diagonal spectral element mass matrix contain-ing the integration weights, A is the discrete Laplacian, and DT and D are the gradient and divergence operators, respectively. The gradient and divergence operators are transpose matrices of one another and handle interpolation be-tween the GLL and GL grids (assuming PN − PN −2) (Deville et al. 2002). The advection operator C(U) is evaluated using the convective form.

By reshuffling the terms of (3.9), one can identify the discrete Helmholtz operator H = (β0/∆t)B + Re−1δ∗A. To further simplify notation, let hn= C(U)u′n−1+ Bfn+ (β

1/∆t)Bu′n−1 and ∆p′n= p′n− p′n−1, which gives Hu′n− DT∆p′n= hn+ DTp′n−1 ⇐⇒ Hu′n− DT∆p′n+HQDT∆p′n − HQDT k∆p′n | {z } =0 = hn−1+ DTp′n−1 ⇐⇒ Hu′n− HQDT∆p′n= hn−1+ DTp′n−1 − (HQ − I)DT∆p′n | {z } ≡en . (3.10)

At this stage the operator Q is an unspecified operator. Equation (3.10) is solved by carrying out a block LU-decomposition (Perot 1993) of the left hand side of (3.10) and (3.9b) such that

 H 0 −D −DQDT   u′ ∆p′n  =  hn−1+ DT 1p′n−1 0  +  en 0  (3.11a)  I −QDT 0 I   u′n ∆p′n  =  u′ ∆p′n  . (3.11b)

(21)

3.3. TIMESTEPPER APPROACH 11 As further discussed by Couzy (1995), the simplest choice of Q is H−1, which causes the residual en to vanish and the equations to solve become

Hu′= hn−1+ DTp′n−1 (3.12a) −DH−1DT | {z } ≡ ˜E ∆p′n= Du′ (3.12b) u′n= u′+ H−1DT∆p′n. (3.12c) The solution of Navier–Stokes by equation (3.12) is referred to as the Uzawa method (Fischer 1997; Deville et al. 2002) and involves no splitting error. However, it requires d Helmholtz-solves for every pressure iteration in a d-dimensional problem, i.e. nested iterations. For large Reynolds number and small timesteps, the Helmholtz operator H will be close to the mass matrix B, which for the case of SEM is diagonal and hence easily inverted. Choosing Q= (∆t/β0)B−1 and dropping the term en, the equations become

Hu′= hn−1+ DTp′n−1 (3.13a) −∆tβ 0 DB−1DT | {z } ≡E ∆p′n= Du′ (3.13b) u′n= u′+∆t β0 B−1DT∆p′n, (3.13c) where E is symmetric positive definite and is referred to as a consistent Poisson operator (Fischer 1997). Solution by (3.13) avoids the nested solves of the Uzawa method but contains a non-zero splitting error

en= β 0 ∆tB+ 1 Reδ∗ A ∆t β0 B−1− I  DT∆p′n= 1 Reδ∗ ∆t β0 AB−1DT∆p′n. (3.14) Assuming that the difference in velocity u′nbetween two consecutive timesteps is small, the solution time can be decreased by only solving for the difference in velocity relative to the previous timestep,

Hu′n= hn−1+ DTp′n−1−∆tβ 0 HB−1DT∆p′n ⇐⇒ H∆u′n= hn−1+ DTp′n−1 − Hu′n−1−∆tβ 0 HB−1DT∆p′n. (3.15a)

3.3. Timestepper approach

Considering the state space formulation of Navier–Stokes (2.4), the general so-lution to this differential equation reads u′(x, t) = exp(AUt)u′0 with t = n∆t. This is typically obtained by numerical integration in time from an initial con-dition u′(x, 0) = u′0. The numerical routine approximating the matrix expo-nential T (t) ≡ exp(AUt) is referred to as timestepper and will in addition to advancing the flow in time (i.e. solve (3.15a)), impose boundary conditions and

(22)

12 3. NUMERICAL APPROACH

continuity on the flow field. Repeated application of the linearized timestep-perT (t) amounts to power iteration with the matrix exponential and will be further discussed in section 4.1.

3.4. Stokes preconditioner

For a solution to a differential equation to be integrated stably with an ex-plicit time integration scheme, the timestep is required to be small, and thus the operator exp(∆tAU) will be close to identity. As a consequence, eigen-computations based on numerical integration of (2.4) are bound to require a large amount of steps, which for a contemporary large-scale DNS simulation requiringO(103) processing elements is intractable.

An alternative method introduced by Tuckerman and co-workers (Tucker-man 1989a; Tucker(Tucker-man et al. 2000; Tucker(Tucker-man & Barkley 2000) for the com-putation of steady states and to eigenpairs, and implemented in paper 1 for the Ginzburg–Landau equation, involves obtaining the action of the govern-ing operatorAU directly without time integration. By assuming a first order timestepping scheme as considered in section 3.2, an alternative expression for the timestepperT (∆t) may be obtained from (2.4) as

u′n= (I− ∆tL)−1(I + ∆tN

U)u′n−1≡ T (∆t)u′n−1. (3.16) Given a code that carries out the action ofT (∆t), one may consider the differ-ence between two consecutive flow fields, that is

u′n− u′n−1=T (∆t)u′n−1− u′n−1 = (I− ∆tL)−1(I + ∆tNU)− I  u′n−1 = (I− ∆tL)−1∆t | {z } ≡P(∆t) (NU+L)u′n−1 (3.17) =P(∆t)AUu′n−1. (3.18)

Two important remarks regarding the operatorP(∆t) are:

(i) AU is ill-conditioned due to the presence of the LaplacianL. For large timestepsP(∆t) = (I − ∆tL)−1∆t≈ L−1, and the application ofP(∆t) will counteract the presence ofL and serve as a preconditioner (referred to as the Stokes or Laplace preconditioner);

(ii) in the limit of small timestepsP(∆t) = (I − ∆tL)−1∆t≈ ∆t, and the application ofP(∆t) will be a mere scaling of the operator AU, moving its eigenvalues towards the origin thus making it more ill-conditioned. By equation (3.18), one has a means of realizing the action of the operatorAU using tools that are readily available in an existing forward Euler/backward Euler timestepping code. For a flow solver designed for problems having one or several homogeneous spatial directions, e.g. SIMSON (Chevalier et al. 2007), a Fourier and a Chebyshev representation can be used respectively for discretiza-tion of the homogeneous and inhomogeneous direcdiscretiza-tions (Canuto et al. 2007). The operators arising from such discretizations will be banded or block-diagonal

(23)

3.4. STOKES PRECONDITIONER 13 and inverted by direct methods (see also Tuckerman 1989b). In such cases, the timestep could be chosen solely based on the above remarks. However, given a code such as Nek5000 that is designed for arbitrary geometries, other con-siderations need to be taken into account as the Helmholtz and the consistent Poisson problems are now solved by iterative methods (preconditioned conju-gate gradient and generalized minimal residual, respectively).

Assuming that no volume forces are acting on the flow, and the absolute pressure (p′n) is solved for rather than the pressure difference (∆p′n), one has

H∆u′n+ HQDTp′n= C(U)u′n−1+ β1 ∆tBu ′n−1 − Hu′n−1 =⇒ ∆u′n= H−1  C(U)u′n−1−Re1 δ∗ Au′n−1  − QDTp′n. (3.19a) Following the discussion on the choice of Q in section 3.2, it is clear that a large ∆t, for which H−1 ≈ Re

δ∗A−1 would serve as a good preconditioner, is incompatible with the approximation Q = (∆t/β0)B−1. Accurate repre-sentation ofP(∆t)AU thus requires the use of the Uzawa method, in which (3.12a) needs to be solved for each iteration of (3.12b). To investigate this matter, the two-dimensional lid-driven cavity flow with Re = 100 is consid-ered. Power iterations are carried out to compute the eigenpair with eigenvalue λ =−1.229 + i0.291 (see figure 3.2a) using an eigenvalue shift σ = −1.2 + i0.3,

u′n+1= (AU− σI)−1u′n ⇐⇒

(P(∆t)AU− σP(∆t))u′n+1=P(∆t)u′n. (3.20a) Equation (3.20a) is solved with BiCGSTAB2 (Gutknecht 1993) to a tol-erance of 10−8, and as a convergence criterion for the power iterations, |λn

− λn−1

| < 10−8 is used. As seen in figure 3.2b, the iteration count de-creases monotonously with increasing ∆t, hence showing the desirable effect of the preconditionerP(∆t). However, as shown in figure 3.2c, the iterations re-quired to solve the Helmholtz problem (3.12a) and the Poisson problem (3.12b) increase almost exponentially with increasing ∆t. As a result, the execution time can be expected to have a minimum for intermediate timestep sizes. This minimum exists around ∆t = 5· 10−2for the present test case1(see figure 3.2d). The number of BiCGSTAB2-iterations is limited to 2000, and for the case of ∆t = 10−3 the prescribed convergence criterion could not be reached. As a consequence, the number of power iterations required to reach convergence for this ∆t increased from 8 to 18.

1Execution time was measured using 8 threads on an Intel R CoreTM

(24)

14 3. NUMERICAL APPROACH -3 -2 -1 0 -2 -1 0 1 2 λr λi

(a) Eigenvalue spectrum obtained with a reference implementation using ARPACK (Lehoucq et al. 1997), λ = −1.229 + i0.291 (×) and λ = −0.543 (⋄). 10-3 10-2 10-1 100 0 500 1000 1500 2000 ∆t It er at io n s (b) BiCGSTAB2 iterations vs. ∆t. 10-3 10-2 10-1 100 0 50 100 150 200 250 300 350 ∆t It er at io n s

(c) Helmholtz (—–) and Poisson (- - -) itera-tions vs. ∆t. 10-3 10-2 10-1 100 1 2 3 4 5 6 7x 10 4 ∆t se c.

(d) Execution time (seconds) vs. ∆t.

Figure 3.2: Effect of the preconditionerP(∆t) for different ∆t on the solution of (3.20a).

(25)

CHAPTER 4

Global stability analysis

4.1. Modal analysis

Given a base flow U corresponding to a fixed point in phase space, modal analysis aims at investigating the eigenpairs of the Jacobian1

AU linearized about this base flow. In particular, one is interested in determining whether an infinitesimal perturbation will grow or decay, and hence whether the fixed point can be deemed stable or unstable. A fixed point is said to be stable if all the eigenvalues i} have negative real part, unstable if at least one eigenvalue has a positive real part, and marginally stable if one eigenvalue is purely imaginary. Properties inferred from a linear analysis will generalize to the corresponding non-linear system by the Hartman–Grobman theorem and the stable manifold theorem, provided that none of the eigenvalues has a zero real part (Guckenheimer & Holmes 1983).

Considering three-dimensional flows that are inhomogeneous in all spatial directions, the perturbation ansatz reads

u′(x, t) = ˆu(x)eλt+ c.c., (4.1) where λ is complex-valued following a temporal instability framework (Huerre & Monkewitz 1990; Chomaz 2005), and the frequency and growth rate of the eigenfunction ˆu is given by the imaginary and real part of λ. This ansatz substituted into (2.4) yields the standardized eigenvalue problem

AUu(x) = λˆˆ u(x). (4.2)

For simple one- and two-dimensional dynamical systems, (4.2) can be solved by direct methods e.g. QR-algorithm, where the discretized operator can be reduced to upper Hessenberg form through Householder reflections and then subsequently reduced to a Schur form by shifted inverse iteration and deflation (Trefethen & Bau 1997). However, with increased geometrical complexity an exact representation and a full diagonalization of the operator may be impos-sible to realize due to limited resources in memory, storage and computational power. For such problems, one usually has to resort to matrix-free methods instead, Tuckerman & Barkley (2000); Barkley et al. (2008); Bagheri et al. (2009a).

1The linearized Jacobian is derived from a Taylor expansion of the velocity field about the fixed point U, i.e. ∂

∂t(U + εu ′) = A(U) + ∂A ∂u Uεu ′+ O(ε2u′2) =⇒ ∂u′ ∂t = ∂A ∂u Uu ′, where (∂A/∂u)|U= AU. 15

(26)

16 4. GLOBAL STABILITY ANALYSIS

The solution of an eigenvalue problem using matrix-free methods relies on an iterative subspace projection method, where flow fields obtained from repeated power iterations are used to generate an m-dimensional Krylov space Km= span{u′1, u′2, . . . , u′m}. (4.3) This set of flow fields{u′

j} is orthonormalized and factorized using the Arnoldi algorithm (see e.g. Lehoucq et al. 1997; Trefethen & Bau 1997), such that

SVm= VmHm+ rmeTm, (4.4) where Vm is an m-dimensional orthonormal basis forKm, Hm an upper Hes-senberg matrix representing the projection ofS onto Km, rmthe residual vector of the factorization and emthe mth unit vector. The residual is orthogonal to the basis Vm and if this vanishes, eigenpairs of the upper Hessenberg matrix will be eigenpairs of the operatorS.

The operatorS can be either the timestepping operator T (t), as is the case in paper 2 and paper 3, or the linearized shifted Jacobian (AU− σI)−1 as in paper 1. The eigenvalues ofS are referred to as Ritz-values (here denoted θj), and in the former case withS = T (t), these are related to the eigenvalues ofAU as ℜ(λj) = 1 t ln|θj|; ℑ(λj) = 1 t arg(θj). (4.5)

The temporal difference t = l∆t between two subsequent vectors in the Krylov space (4.3) needs to be chosen such that every timestep ∆t satisfies the CFL-condition, and l such that the generated subspace captures the dynamics of the full system, i.e. satisfies the Nyquist criterion (Bagheri et al. 2009a). In the latter case, whereS = (AU− σI)−1, the Ritz-values are simply related to the eigenvalues as ωj= θ−1j + σ. The eigenfunctions of S are similar to those of AU for both cases.

4.1.1. Energy analysis

To analyze which gradients and structures are responsible for production and dissipation of energy, an energy analysis may be carried out, as is done in paper 2. An expression for the evolution equation of the perturbation energy may be obtained by multiplying the linearized disturbance equation (2.2a) with the complex conjugate velocity perturbation

e u′ k ∂u′ k ∂t =− eu′kUj ∂u′ k ∂xj− e u′ ku′j ∂Uk ∂xj − e u′ k ∂p′ ∂xk + 1 Reδ e u′ k ∂2u′ k ∂xj∂xj ++ eu′ kfk′, (4.6) here expressed using tensor notation with (e· ) representing a complex conjugate. Assuming that the force term is proportional to the velocity, i.e. fk′ = γu′k, forming the complex conjugate of (4.6) and adding these equations gives

(27)

4.2. NON-MODAL ANALYSIS 17 ∂ ∂t e u′ ku′k 2 ! =−12ue′ ku′j+ u′kue′j  ∂Uk ∂xj − 1 Reδ ∂ eu′ k ∂xj ∂u′ k ∂xj +1 2 ∂ ∂xj h − eu′ ku′kUj− eu′kp′δkj− u′kpe′δkj + 1 Reδ e u′ k ∂u′ k ∂xj + u′k ∂ eu′ k ∂xj !# + γ eu′ ku′k. (4.7) By inserting the global mode ansatz (4.1) for velocity together with a similar one for pressure into the above equation, expressions for production, dissipation and transport of energy for every eigenpair (ˆu, λ) can be written (see Appendix B of paper 2 for an equivalent but slightly different derivation). Furthermore, by defining the disturbance kinetic energy in the domain of integration Ω as EΩ= (1/2)

R

Ω( eu′ku′k)dV , and slightly modifying (4.7), an expression for the growth rate of the eigenfunction is obtained, E−1(∂EΩ/∂t) = 2λr.

4.1.2. Structural sensitivity

Another way of analyzing the stability mechanisms, that give complementary information to the energy analysis, is the study of the ’wavemaker’. Huerre & Monkewitz (1990) described this concept as a self-excited low-amplitude phenomenon in a region of local absolute instability, which acts as a source for unstable waves. Due to non-normality of the governing linearized Navier– Stokes operator, there will be a large spatial separation between the direct and adjoint eigenfunctions (Chomaz 2005). Although the adjoint eigenfunctions are found to indicate regions of strong sensitivity to forcing (Hill 1995), it was argued by Giannetti & Luchini (2007) that the instability mechanism cannot be identified by either the direct or adjoint eigenfunctions separately. Instead they proposed to study the product between these eigenfunctions

µ(x) = Rkˆv(x)kkˆu(x)k Ωv(x)ˆ · ˆu(x)dV

, (4.8)

which characterize the drift in an eigenvalue due to a structural perturbation in the linearized Navier–Stokes operator (a similar expression was also derived by Chomaz 2005). Regions where µ(x) is different from zero can hence be expected to determine the characteristics of global oscillations in the flow. This type of analysis is considered in paper 3.

4.2. Non-modal analysis

Non-modal analysis involves the study of transiently amplifying structures, and constitutes the centermost transition path of figure 1.1. Mathematically, transient growth arises from non-normality of the governing operatorAU, or equivalently, a non-orthogonal set of eigenfunctions as illustrated graphically by Chomaz (2005). The study of transient growth is hence closely tied to the study

(28)

18 4. GLOBAL STABILITY ANALYSIS

of ε-pseudospectra defined in (4.16) (Trefethen et al. 1993; Reddy et al. 1993). Due to the size of the problems associated with many three-dimensional flow cases, a full representation of the ε-pseudospectra may not be computationally realizable and an alternative might be to analyze the pseudospectra of a smaller Hessenberg matrix (Toh & Trefethen 1996), as discussed in connection with eigenpairs in section 4.1.

Physically on the other hand, transient growth arises from the lift-up effect (Landahl 1975), by which streamwise vortices transport low-speed fluid into high-speed regions of a shear layer, and simultaneously transport high-speed fluid into low-speed regions, resulting in the formation of streamwise streaks (Trefethen et al. 1993; Reddy et al. 1998). To connect these observations with the mathematical framework outlined above, such streamwise vortices may be viewed as ε-pseudomodes (Trefethen et al. 1993), and several studies e.g. Butler & Farrell (1992), have focused on determining the structures causing the largest transient amplification of perturbations.

In the present thesis another issue, concerned with the maximum response of a dynamical system subject to a harmonic driving force is considered,

∂u′

∂t =AUu

(x) + ˆζ(x)eiωt. (4.9) The forcing function and frequency causing the largest energy amplification is defined as the optimal forcing of the system. In paper 1, a novel method for determining the optimal forcing that is based on the Stokes (or Laplace) preconditioner (chapter 3.4) is presented. It is shown that, the optimal forcing problem can be formulated as an eigenvalue problem of the adjoint and direct resolvent operator R†(iω,AU)R(iω, AU) ˆζ(x) = G(ω)2ζ(x)ˆ ⇐⇒ (4.10) h (AU− iωI)(A†U+ iωI) i−1 ˆ ζ(x) = G(ω)2ζ(x),ˆ (4.11) where the leading eigenvalue G(ω) is seen to represent the energy gain, and the corresponding eigenfunction ˆζ(x) the corresponding forcing function. In order to find the forcing frequency to which the flow is most sensitive, (4.11) needs to be solved for a range of forcing frequencies ω. The method is herein applied to the linearized Navier–Stokes equations.

The adjoint linearized Navier–Stokes equations read −∂v ′ ∂t =∇s + (U · ∇)v ′− (∇U)Tv+ 1 Reδ∗ 0 ∇2v+ g(4.12a) ∇ · v′= 0, (4.12b)

wherein v′ represents the adjoint velocity field, g′ an adjoint force field, and s the adjoint pressure (see e.g. Bagheri et al. 2009b for details on the derivation). Introducing the variable substitutions, τ ≡ −t and ˜s ≡ −s, (4.12) receives the same form as (2.2). Considering a state space formulation of this equation

(29)

4.2. NON-MODAL ANALYSIS 19 similar to that in section 2.2, gives

∂v′ ∂τ = P (U· ∇)v ′− (∇U)Tv′ | {z } ≡N† Uv′ + P  1 Reδ∗∇ 2v′  | {z } =Lv′ . (4.13)

As done for the direct timestepper in section 3.4, a similar procedure can be used to obtain an expression the adjoint timestepper,

v′n= (I− ∆τL)−1(I + ∆τNU†)v ′n−1

≡ T†(∆τ )v′n−1. (4.14) From this, the difference between two consecutive adjoint timesteps reads

v′n− v′n−1= (I− ∆τL)−1∆τ | {z } ≡P(∆τ ) (NU† +L)v ′n−1= P(∆τ)A†Uv ′n−1. (4.15)

Note that operatorL and P(∆τ) are self-adjoint in contrast to the Ginzburg– Landau equation.

Returning to the problem of determining the optimal forcing as formulated in equation (4.11), the action of the governing operator is realized through four steps involving the solution of two equation systems, and two applications of the Stokes preconditioner, namely:

1. ˆζI← P(∆t) ˆζn.

2. ˆζII← [P(∆t)AU− iωP(∆t)]−1ζˆI. 3. ˆζIII← P(∆τ) ˆζII.

4. ˆζn+1← [P(∆τ)A†+ iω∗P(∆τ)]−1ζˆIII.

To solve the systems in steps two and four, both BiCGSTAB (van der Vorst 1992; Gutknecht 1993) and BiCGSTAB2 (Gutknecht 1993) have been imple-mented and adapted for Nek5000. In order to test and verify the validity of the above methodology, the optimal forcing of the two-dimensional lid-driven cavity flow at Re = 100 is computed. To the authors’ knowledge, this has not been done before.

The result is plotted in figure 4.1 and shows that the largest energy ampli-fication in the flow is achieved for a steady forcing with ω = 0.0, corresponding to eigenvalue-resonance with the least stable eigenvalue (cf. figure 3.2a). For a normal governing operatorAU, G(ω)∼ 1/|λr|, i.e. the amplification is propor-tional to the inverse distance of the forcing frequency to the eigenvalue closest to the imaginary axis. In contrast, if AU is non-normal, much larger energy amplification is obtainable, i.e. G(ω)∼ 1/ε0. Here ε0is the smallest value of ε for which the ε-pseudospectra defined as

Λε(AU)≡ {z ∈ C | k(AU− zI)−1k ≥ ε−1}, (4.16) protrudes into the right half of the complex plane (Chomaz 2005). Given λr=−0.543, this would yield G(ω) ∼ 1.842 which is in good agreement with figure 4.1 and shows that the lid-driven cavity at this Reynolds number is governed by a normal operator.

(30)

20 4. GLOBAL STABILITY ANALYSIS -100 -8 -6 -4 -2 0 2 4 6 8 10 0.5 1 1.5 2 ω G (ω )

Figure 4.1: Optimal forcing of two-dimensional lid-driven cavity, Re = 100.

The corresponding optimal forcing function is shown in figure 4.2a-4.2b. Comparison of this with the least stable adjoint eigenfunction (figure 4.2c-4.2d), obtained with a similar method as discussed in section 3.4, shows a strong resemblance. This is also the case for the Ginzburg–Landau operator in paper 1, and is in good agreement with Sipp (2012), who argued that a forcing function equal to the adjoint global mode provides the largest impact on the flow in case of a forcing frequency close to the frequency of a global mode.

(31)

4.2. NON-MODAL ANALYSIS 21

(a) Force profile, x-component. (b) Force profile, y-component.

(c) Adjoint eigenfunction, x-component. (d) Adjoint eigenfunction, y-component.

Figure 4.2: Optimal forcing function corresponding to ω = 0.0, compared to the most unstable adjoint eigenfunction with eigenvalue λ∗=−0.543.

(32)

CHAPTER 5

Summary of the papers

Paper 1

A method for computing optimal forcing of convectively unstable flows using Laplace preconditioning

Common shear flows such as channel flow and boundary layer flows are typically seen to exhibit large transients as a result of the non-normality of the governing operator. For such flows stability predictions based on the leading eigenvalue may fail and more accurate transition estimates may be obtained by studying the resolvent operator governing the output of a system subject to a harmonic driving force.

Common methods for solving this problem involve singular value decompo-sition (SVD) and direct-adjoint time integration using a timestepper. Although existing literature have proven that such methods give valid results for a num-ber of flow cases, both of them become impractical when considering large three-dimensional flow cases. In the present paper, a novel matrix-free method that relies on the solution of linear systems is presented. The method is appli-cable to problems of arbitrary size and complexity, but is herein discussed in relation with the one-dimensional linearized Ginzburg–Landau equation.

The key to the method involves evaluation of the resolvent operator by approximate inversion of the shifted direct and adjoint Jacobians. To this end, the systems are preconditioned using the Laplace preconditioner. For the present model problem, explicit representation of all operators is possible, which enables a detailed study of the properties of the Laplace preconditioner as well as a validation of the results using direct methods.

Paper 2

Onset of global instability behind distributed surface roughness in a Falkner– Skan–Cooke boundary layer

The laminar swept wing boundary layer flow is commonly approximated by the Falkner–Skan–Cooke similarity solutions. Within this approximation, the spanwise free-stream component is assumed to be constant and the streamwise

(33)

5. SUMMARY OF THE PAPERS 23 free-stream velocity component to accelerate with downstream distance from the leading edge. Typical transition scenarios for such boundary layers in low-disturbance environments, involve the excitation of low-disturbances by surface roughness that grow in space and eventually develop into strong non-linear crossflow vortices. Crossflow vortices are susceptible to secondary instabilities and act as noise amplifiers. However, for sufficiently large roughness elements, the flow is seen to support self-sustained oscillations that seems to drive the instabilities that cause transition further downstream.

In this study, this phenomenon is investigated further using a global sta-bility analysis. In order to understand the transition scenario, a global energy analysis is carried out, which identifies the mechanisms responsible for produc-tion and dissipaproduc-tion of perturbaproduc-tion energy, as well as their spatial locaproduc-tion in the domain. Due to the downstream amplification of the crossflow vor-tices, structures of the flow will not be localized, but extend throughout the entire domain. As a consequence, the results of the stability analysis will be very sensitive to the particular domain size. By considering the kinetic energy budget of a perturbation in the flow, a possible explanation to this domain dependency is discussed. In addition, the critical roughness parameters for the Falkner–Skan–Cooke boundary layer are compared to those pertaining to the two-dimensional boundary layers. It is suggested that the onset of global sta-bility for the flow around a discrete roughness element is mainly determined by the local flow conditions around the roughness, and to a lesser extent by the particular details of the boundary layer.

Paper 3

On the stability of flat plate boundary layers subject to localized suction

A substantial amount of the total drag on an airplane is accounted for by skin-friction drag, which in turn is significantly lower for a laminar boundary layer compared to a turbulent one. Therefore, large portions of laminar flow over a vehicle are preferable. One existing technique to extend this region is through application of suction. Suction applied on a boundary layer has been found to yield a fuller (more stable) velocity profile, and a thinner boundary layer (lower Reynolds number), hence delaying the onset of transition. For sufficiently strong suction ratios however, premature transition is observed. As a result, most studies in the field have focused on the underlying breakdown mechanisms and applicable design criteria.

In the present study, the stability of a Blasius boundary layer subject to localized suction is addressed from the point of a global stability analysis. This is most likely the first of its kind for this particular flow case and enables a detailed analysis of the receptivity and sensitivity of the flow. It is shown that the structures inside the pipe are the first to become unstable due to unsteadiness shed from the separation bubble located at the pipe junction.

(34)

24 5. SUMMARY OF THE PAPERS

Eventual instability of the boundary layer is also found to originate from this separation bubble, hence indicating the importance of controlling this structure to stabilize a case subject to supercritical suction.

(35)

CHAPTER 6

Conclusions and outlook

Global stability analysis

In this work the global stability of two boundary layer flows exhibiting consider-ably different behavior have been investigated. Some conclusions and remarks from these studies are summarized below.

For the Falkner–Skan–Cooke boundary layer subject to a discrete surface roughness, two dynamically very different areas have been observed – one im-mediate near-wake of the roughness, and one far-field region associated with the crossflow vortices. It has been noted that structures emanating from the near-wake excite the downstream crossflow vortices and unsteady structures shed from a supercritical roughness element drive a secondary instability down-stream. However, the structures of the near-wake have been seen to decay af-ter sufficient distance from the roughness, whereas the crossflow vortices are amplified and reached their largest amplitude immediately before the outflow boundary. Variations of the box length showed that the stability characteris-tics of the flow were sensitive to this parameter, which complicated the analysis and interpretation of the results. A complete description of this phenomenon is still under development and will be pursued in the near future. It is ten-tatively suggested that the observed change in the eigenvalue spectrum with domain length, is accounted for by a change in the advective transport of the perturbation energy due to a change in the underlying baseflow. Production and dissipation terms are noted to remain unchanged, thus indicating that the mechanisms responsible for the instability also remain unchanged.

Regarding the Blasius boundary layer with applied localized suction, un-stable structures seen within the pipe and the boundary layer were observed to be closely tied to the separation bubble that developed at the pipe junction. For this case, various box lengths have not been studied, but might be con-sidered as well in the near future. The important difference between this flow case and that previously discussed, is the presence of only one dynamical re-gion, namely the near-wake. Therefore, structures are expected to decay after a certain distance from the disturbance source, and the regions critical to the in-stability are expected to be localized at the suction site. It is therefore believed that the results presented herein will be relatively insensitive with respectto the domain dependency, compared to the case with a roughness submerged into a Falkner–Skan–Cooke boundary layer.

(36)

26 6. CONCLUSIONS AND OUTLOOK

Optimal forcing

The novel method for computing optimal forcing developed in this work has suc-cessfully been applied to both the one-dimensional Ginzburg–Landau operator and the two-dimensional lid-driven cavity flow governed by the Navier–Stokes operator. In both cases, the Stokes (or Laplace) preconditioner is shown to work satisfactorily, rendering a lower iteration count with increased timestep.

Future work on this method will strive to extend its range of applicability to more complex flows, such as the dimensional Couette flow, the two-dimensional Blasius boundary layer, and eventually a fully three-two-dimensional flow. The main challenge is the computational cost associated with the method. Some improvements in efficiency already await implementation, but as seen, the main bottleneck is the exponential increase in nested iterations associated with the solution of Poisson and Helmholtz problems. Experience with solvers such as BiCGSTAB and BiCGSTAB2 (Gutknecht 1993; van der Vorst 1992) show very erratic convergence behavior, and below a certain ∆t a solution simply cannot be found (as reflected by the discontinuous step in iteration count and execution time in figure 3.2). A solution that might remedy this issue is implementation of generalized minimal residual (GMRES) (Saad & Schultz 1986), which is known to exhibit monotonic convergence at the expense of larger memory overhead associated with storage of the full Krylov space. It is believed that satisfactory convergence behavior and decrease in computational cost can be obtained by fine-tuning the size of the timestep and frequency with which GMRES is restarted on a case by case basis.

(37)

Acknowledgements

First of all I would like to thank Dan Henningson for giving me the opportu-nity to do a PhD under his supervision. He has always believed in me, and encouraged me to pursue my ideas. Thanks to his generosity I have been able to attend many conferences and workshops, which has broadened my views and perspectives. My gratitude also goes to my co-advisors Philipp Schlatter and Ardeshir Hanifi for their great enthusiasm, and for always taking the time to discuss research when needed.

I am grateful for having the opportunity to meet and work with great re-searchers such as Laurette Tuckerman, from whom I have learned a lot and been inspired. A special thanks goes to Adam Peplinski for sharing his knowl-edge about the research code Nek5000. Without his help much of this work would not have been possible. Luca Brandt is acknowledged for many inter-esting discussions. My appreciation also goes to Shervin Bagheri whose PhD thesis has been a handbook for my daily work.

A great thanks goes to my room mates Azad, Nicol`o and Bobke for making our office ”the place to be in” in Osquars Backe right now... Thanks also to my other fellow PhD students and friends for making it a pleasure to come to work every day – Taras, imanl(!!!), hosse, clisa, nimash, Elektra, Jacopo, U˘gis, Nicolas, Ellinor, Prabal, Erik, Walter and Marco.

Till slut vill jag ocks˚a rikta ett stort tack till mina f¨or¨aldrar och min syster f¨or allt st¨od under min doktorering. Utan er hj¨alp hade jag inte varit d¨ar jag ¨

ar idag.

(38)

Bibliography

Bagheri, S.2010 Analysis and control of transitional shear flows using global modes. PhD thesis, KTH Royal Institute of Technology.

Bagheri, S., ˚Akervik, E., Brandt, L. & Henningson, D. S.2009a Matrix-free methods for the stability and control of boundary layers. AIAA J. 47 (5), 1057– 1068.

Bagheri, S., ˚Akervik, E., Brandt, L. & Henningson, D. S. 2009b Matrix-free methods for the stability and control of boundary layers. AIAA J. 47 (5), 1057– 1068.

Bagheri, S., Brandt, L. & Henningson, D. S.2009c Input–output analysis, model reduction and control of the flat-plate boundary layer. J. Fluid Mech. 620, 263– 298.

Bagheri, S., Schlatter, P., Schmid, P. J. & Henningson, D. S.2009d Global stability of a jet in crossflow. J. Fluid Mech. 624, 33–44.

Barkley, D., Blackburn, H. M. & Sherwin, S. J.2008 Direct optimal growth analysis for timesteppers. Int. J. Numer. Meth. Fl. 57 (9), 1435–1458.

Barkley, D. & Tuckerman, L. S. 1997 Stokes preconditioning for the inverse power method. In Lecture Notes in Physics: Proc. of the Fifteenth Int’l. Conf. on Numerical Methods in Fluid Dynamics (ed. P. Kutler, J. Flores & J.-J. Chattot), pp. 75–76. Springer, New York.

Butler, K. M. & Farrell, B. F.1992 Three-dimensional optimal perturbations in viscous shear flow. Phys. Fluids A 4 (8), 1637–1650.

Canuto, C., Hussaini, Y., Quarteroni, A. & Zang, T. A.2007 Spectral Methods: Evolution to Complex Geometries and Applications to Fluid Dynamics. Springer Berlin Heidelberg.

Chevalier, M., Schlatter, P., Lundbladh, A. & Henningson, D. S.2007 SIM-SON - A Pseudo-Spectral Solver for Incompressible Boundary Layer Flows. Tech. Rep. TRITA-MEK 2007:07. KTH Mechanics.

Chomaz, J.-M.2005 Global instabilities in spatially developing flows: Non-normality and nonlinearity. Annu. Rev. Fluid Mech. 37 (1), 357–392.

Chorin, A. J. & Marsden, J. E.2000 A Mathematical Introduction to Fluid Me-chanics. Texts in Applied Mathematics v. 4. Springer New York.

Cooke, J. C. 1950 The boundary layer of a class of infinite yawed cylinders. Proc. Camb. Phil. Soc. 46 (04), 645–648.

(39)

BIBLIOGRAPHY 29

Couzy, W.1995 Spectral element discretization of the unsteady Navier–Stokes equa-tions and its iterative solution on parallel computers. PhD thesis, ´Ecole Poly-technique F´ed´erale de Lausanne.

Deville, O., Fischer, P. F. & Mund, E. H.2002 High-Order Methods for Incom-pressible Fluid Flow . Cambridge University Press.

Drazin, P. G. & Reid, W. H.1982 Hydrodynamic stability. Cambridge University Press.

Falkner, V. M. & Skan, S. W.1931 Some approximate solutions of the boundary layer equations. Phil. Mag. 12 (80), 865–896.

Fischer, P. F.1997 An overlapping schwarz method for spectral element solution of the incompressible Navier–Stokes equations. . 133 (1), 84–101.

Fischer, P. F., Lottes, J. W. & Kerkemeier, S. G.2008 nek5000 Web page. http://nek5000.mcs.anl.gov.

Giannetti, F. & Luchini, P.2007 Structural sensitivity of the first instability of the cylinder wake. J. Fluid Mech. 581, 167–197.

Giannetti, F., Luchini, P. & Marino, L.2009 Linear stability analysis of three-dimensional lid-driven cavity flow. In Atti del XIX Congresso AIMETA di Mec-canica Teorica e Applicata, pp. 14–17. Aras Edizioni Ancona, Italy.

G´omez, F., Clainche, S. L., Paredes, P., Hermanns, M. & Theofilis, V.2012 Four decades of studying global linear instability: Progress and challenges. AIAA J. 50 (12), 2731–2743.

Guckenheimer, J. & Holmes, P.1983 Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Applied Mathematical Sciences v. 42. Springer New York.

Gutknecht, M. H.1993 Variants of BICGSTAB for matrices with complex spec-trum. SIAM J. Sci. Comp. 14 (5), 1020–1033.

Henningson, D.1996 Comment on ”Transition in shear flows. Nonlinear normality versus non-normal linearity” [Phys. Fluids 7, 3060 (1995)]. Phys. Fluids 8 (8), 2257–2258.

Hill, D. C.1995 Adjoint systems and their role in the receptivity problem for bound-ary layers. J. Fluid Mech. 292, 183–204.

Huerre, P. & Monkewitz, P. A.1990 Local and global instabilities in spatially developing flows. Annu. Rev. Fluid Mech. 22 (1), 473–537.

Karniadakis, G. & Sherwin, S.2013 Spectral/hp Element Methods for Computa-tional Fluid Dynamics: Second Edition. OUP Oxford.

Kreiss, G., Lundbladh, A. & Henningson, D. S. 1994 Bounds for threshold amplitudes in subcritical shear flows. J. Fluid Mech. 270, 175–198.

Landahl, M.1975 Wave breakdown and turbulence. SIAM J. Appl. Math. 28 (4), 735–756.

Lehoucq, R. B., Sorensen, D. C. & Yang, C.1997 ARPACK User’s Guide: Solu-tion of Large Scale Eigenvalue Problems by Implicitly Restarted Arnoldi Methods. Society for Industrial and Applied Mathematics.

Mack, C. J., Schmid, P. J. & Sesterhenn, J. L.2008 Global stability of swept flow around a parabolic body: connecting attachment-line and crossflow modes. J. Fluid Mech. 611, 205–214.

(40)

30 BIBLIOGRAPHY

equations. In State of the Art Surveys in Computational Mechanics (ed. A. K. Noor), pp. 71–143. New York: ASME.

Patera, A. T.1984 A spectral element method for fluid dynamics: Laminar flow in a channel expansion. J. Comput. Phys. 54 (3), 468 – 488.

Peplinski, A., Schlatter, P. & Henningson, D. S. 2015 Global stability and optimal perturbation for a jet in cross-flow. Eur. J. Mech. B-Fluid. 49, 438–447. Perot, J. B. 1993 An analysis of the fractional step method. J. Comput. Phys.

108(1), 51–58.

Reddy, S. C., Schmid, P. J., Baggett, J. S. & Henningson, D. S. 1998 On stability of streamwise streaks and transition thresholds in plane channel flows. J. Fluid Mech. 365, 269–303.

Reddy, S. C., Schmid, P. J. & Henningson, D. S. 1993 Pseudospectra of the Orr–Sommerfeld operator. SIAM J. Appl. Math. 53 (1), 15–47.

Saad, Y. & Schultz, M. H.1986 GMRES: A generalized minimal residual algorithm for solving nonsymmetric linear systems. SIAM J. Sci. Stat. Comput. 7 (3), 856– 869.

Saric, W. S., Reed, H. L. & Kerschen, E. J.2002 Boundary-layer receptivity to freestream disturbances. Annu. Rev. Fluid Mech. 34 (1), 291–319.

Schlichting, H.1979 Boundary-Layer Theory, 7th edn. McGraw-Hill, Inc.

Schmid, P. J. & Henningson, D. S.2001 Stability and Transition in Shear Flows. Applied Mathematical Sciences v. 142. Springer.

Sipp, D.2012 Open-loop control of cavity oscillations with harmonic forcings. J. Fluid Mech. 708, 439–468.

Stroud, A. H.1974 Numerical Quadrature and Solution of Ordinary Differential Equations: A Textbook for a Beginning Course in Numerical Analysis. Applied Mathematical Sciences v. 10. Springer.

Tezuka, A. & Suzuki, K.2006 Three-dimensional global linear stability analysis of flow around a spheroid. AIAA J. 44 (8), 1697–1708.

Theofilis, V.2011 Global linear instability. Annu. Rev. Fluid Mech. 43, 319–352. Toh, K. C. & Trefethen, L. N.1996 Calculation of pseudospectra by the arnoldi

iteration. SIAM J. Sci. Comp. 17 (1), 1–15.

Trefethen, L. N. & Bau, D.1997 Numerical Linear Algebra. SIAM.

Trefethen, L. N., Trefethen, A. E., Reddy, S. C. & Driscoll, T. A.1993 Hydrodynamic stability without eigenvalues. Science 261 (5121), 578–584. Tuckerman, L. S.1989a Steady-state solving via Stokes preconditioning; recursion

relations for elliptic operators. In Lecture Notes in Physics: Proc. of the Eleventh Int’l. Conf. on Numerical Methods in Fluid Dynamics (ed. D. L. Dwoyer, M. Y. Hussaini & R. G. Voigt), pp. 573–577. Springer, New York.

Tuckerman, L. S. 1989b Transformations of matrices into banded form. J. Com-put. Phys. 84 (2), 360 – 376.

Tuckerman, L. S. & Barkley, D.2000 Bifurcation analysis for timesteppers. In Numerical Methods for Bifurcation Problems and Large-Scale Dynamical Systems (ed. E. Doedel & L. S. Tuckerman). Springer, New York.

Tuckerman, L. S., Bertagnolio, F. Daube, O., Le Qu´er´e, P. & Barkley, D.2000 Stokes preconditioning for the inverse Arnoldi method. In Continuation Methods for Fluid Dynamics (ed. D. Henry & B. A.), Notes on Numerical Fluid Mechanics, vol. 74, pp. 241–255. Vieweg.

(41)

van der Vorst, H.1992 CGSTAB: A fast and smoothly converging variant of Bi-CG for the solution of nonsymmetric linear systems. SIAM J. Sci. Stat. Comput. 13(2), 631–644.

(42)

References

Related documents

Den totala bränsleförbrukningen vid stabil flygning för olika hastigheter i intervallet V Pr,min till och med 64, 37 m s studeras sedan, där den

It is shown in this paper that a Falkner–Skan–Cooke (FSC) boundary layer with a roughness element large enough to excite crossflow (CF) vortices, but small enough to avoid an

But on moving further downstream the decay of the vortices lead to reduction of that fluctuation and it can evidenced that even beyond the 530 mm (66% chord), the flow remains

measuremm1ts in the Meteorological Wind Tunnel at Colorado State Univer- sity, Sandborn and Zoric have documented that for a flat plate turbulent boundary layer

These facts point out the importance of the runner cone on the pressure recovery of elbow draft tube.. Early separation on the runner cone may give rise to a vortex rope and

Conclusions and outlook 23 In this work no adjoint method was used, but a resolvent analysis might shed more light on the sensitivity of the damping of the branch of eigenvalues

In the late afternoon, from 2 h before sunset until when the surface buoyancy flux reduces to 0, (1) the TKE decreases more rapidly than during the early AT within the whole PBL,

Davidsson (2005) used analytical methods to study the transient growth of streamwise elongated fluctuations in the streamwise velocity component (streaks) for a flat plate boundary