<<

An experimental study on the wake behind a rectangular forebody with variable inlet conditions

by

Renzo Trip

February 2014 Technical Reports from Royal Institute of Technology KTH Mechanics SE-100 44 Stockholm, Sweden Akademisk avhandling som med tillst˚and av Kungliga Tekniska H¨ogskolan i Stockholm framl¨agges till offentlig granskning f¨oravl¨aggande av teknologie licenciatexamen onsdag den 19 mars 2014 kl 10:15 i sal E3, Osquarsbacke 14, Kungliga Tekniska H¨ogskolan, Stockholm.

TRITA-MEK 2014:06 ISSN 0348-467X ISRN KTH/MEK/TR-14/06-SE

©Renzo Trip 2014 Universitetsservice US–AB, Stockholm 2014 Renzo Trip 2014, An experimental study on the wake behind a rectan- gular forebody with variable inlet conditions Linn´eFlow Centre, KTH Mechanics, SE-100 44 Stockholm, Sweden Abstract The wake behind a rectangular forebody with variable inlet conditions is investigated. The perforated surface of the two-dimensional rectangular fore- body, with a smooth leading edge and a blunt trailing edge, allows for modification by means of wall suction. The test section, of which the rect- angular forebody is the main part, is experimentally evaluated with a series of hot-wire and Prandtl tube measurements in the boundary layer and the wake. For a suction coefficient of Γ 9, corresponding to 0.9% suction of the free stream velocity, the asymptotic suction boundary layer (ASBL) is obtained at 5 the trailing edge of the forebody for> laminar boundary layers (Rex 1.6 10 3.8 105). The key feature of the ASBL, a spatially invariant boundary thick- ness which can be modified independent of the Reynolds number,= is used× to− perform× a unique parametrical study. 5 6 Turbulent boundary layers (Rex 4.5 10 3.0 10 ) subject to wall suction are also investigated. For a critical suction coefficient Γcrit, which depends on Rex, the boundary layer relaminarizes.= × Strong− evidence× is found to support the hypothesis that turbulent boundary layers will ultimately attain the ASBL as well, provided that the wall suction is strong enough. The effect of the modulated laminar and turbulent boundary layers on the wake characteristics is studied. The shape of the mean wake velocity profile, scaled with the velocity deficit U0 and the wake half width ∆y1 2, is found to be independent of x h, for x h 6 and Re 6.7 103. The wake width is h ￿ shown to scale with the effective thickness of the body h 2δ1, where the ratio is expected to vary with￿ the downstream￿ > location.> × A decrease of the displacement thickness leads to a+ decrease of the base pressure, with Cp,b 0.36 in the ASBL limit. The Strouhal number based on the effective thickness becomes Sth 2δ1 0.29 in the ASBL limit and in- dependent of the plate= − thickness (h) Reynolds number, in the range Reh 3 3 + ≈ 2.9 10 6.7 10 . For the turbulent boundary Sth 2δ1 is found to be 25% lower, which shows that the wake characteristics depend on the state of the= + boundary× − layer× at the trailing edge. 4 The total is found to be reduced by as much as 30% for Reh 2.7 10 when a wall normal velocity of only 3.5% of the free stream velocity is applied. Wall suction successively reduces the total drag with increasing wall= suction,× 3 4 at least in the Reynolds number range Reh 8.0 10 5.5 10 . Descriptors: Bluffbodies, wake flow, asymptotic suction boundary layer, flow = × − × control

iii Preface This licentiate thesis within the area of fluid mechanics concerns an experimen- tal study on the wake behind a rectangular forebody with variable inlet condi- tions. The thesis is divided into two parts. The first part gives an introduction, discussing: the relevance of the research, the underlying fluid mechanics and the experimental methodology. A summary of the results and conclusions pre- sented in the second part is also included. The second part consists of three papers that are redacted to meet the present thesis format.

February 2014, Stockholm Renzo Trip

iv Contents

Abstract iii

Preface iv

Part I. Overview and summary

Chapter 1. Introduction 1

Chapter 2. The wake of bluffbody 3 2.1. Definition of a bluffbody 3 2.2. shedding 4 2.3. Drag prediction 4

Chapter 3. Boundary layers 7 3.1. Previous Studies 7 3.2. Boundary layer approximation 8 3.3. Falkner-Skan boundary layer 9 3.4. Asymptotic suction boundary layer 10 3.5. Turbulent boundary layer 12 3.6. Turbulent suction boundary layer 14

Chapter 4. Experimental setup and measurement techniques 18 4.1. Hot-wire anemometry 18 4.2. Pressure measurements 22 4.3. Particle Image Velocimetry 24

Chapter 5. Summary of results and conclusions 26

Chapter 6. Papers and authors contributions 28

Chapter 7. Acknowledgements 29

v References 31

Part II. Papers

Paper 1. A new test section for studies on wake instability and its control 39

Paper 2. An experimental study on the wake of a rectangular forebody with suction applied over the surfaces 79

Paper 3. An experimental study on the wake behind a rectangular forebody with a laminar and turbulent suction boundary layer 93

vi Part I

Overview and summary

CHAPTER 1

Introduction

Let’s start with a little experiment: take this thesis and throw it up in the air! Initially, the spine cuts through the air swiftly. Besides a thin boundary layer close to the thesis surface, the air offers little resistance to deter the thesis from reaching its final destination. But somewhere in mid flight the thesis tumbles and now faces its destiny front cover first. Intuitively, one feels that the altered orientation has changed the dynamics around the thesis significantly. Actually, if one would visualize the air around the thesis now, one would see that a large wake has formed behind the thesis. Where will the thesis, that has served its purpose as an example of a bluffbody, end up? The experiment described above reveals that bluffbodies moving through a fluid, or bluffbodies placed in a moving fluid, have a strong interaction with the fluid surrounding it. The wake that forms behind a is a good example where this interaction becomes visible. Sketches by Leonardo da Vinci making note of the phenomenon date back to the year 1509 (Bingham 2007) and it is no wonder that it has aroused the interest of many scientists ever since. The formation of a wake behind a bluffbody is inherent to the shedding of vortices and an increased drag. In practice, this leads to noise, structural oscillations and an increase in the fuel consumption of road vehicles, for exam- ple. The widespread occurrence of the phenomenon may justify why a better understanding of the underlying fluid dynamics around bluffbodies is to be sought after. This thesis aims at contributing to this understanding by adding to the sparse amount of experimental data available on the relation between the inlet conditions of the wake and the wake characteristics. The inlet conditions of the wake are formed by the boundary layers at the trailing edge a bluffbody. In similar studies the boundary layer is modified by adding some kind of surface roughness (Buresti et al. 1997; Rowe et al. 2001; Durgesh et al. 2013). The number of cases studied is typically small and the possibility to control the boundary layer properties limited. In the present study, the boundary layer is modified by applying uniform suction over the flat surfaces of a rectangular forebody. The laminar and in particular the turbulent boundary layer over a flat plate subject to uniform suction are a relatively untouched topics themselves. The unique futures of

1 2 1. INTRODUCTION the asymptotic suction boundary layer and some characteristics of a turbulent boundary layer with wall suction will therefore be discussed in detail as well. The first part of this thesis will introduce the terminology and fundamen- tal concepts of the wake behind a bluffbody and the laminar and turbulent boundary layer with and without suction. Paper 1 is devoted to the the ex- perimental setup, including an experimental validation comprising wake and boundary layer measurements. In Paper 2 the relation between the properties of a laminar (asymptotic suction) boundary layer and the wake characteristics is discussed. The turbulent boundary layer subject to suction and near wake measurements are included in Paper 3. CHAPTER 2

The wake of bluffbody

In this chapter an introduction to several concepts related to the wake of a bluff body will be given. The concepts introduced here will be discussed in more detail in relation to the inlet conditions of the wake in Paper 2 and Paper 3 of the present thesis.

2.1. Definition of a bluffbody In order to give a proper definition of a bluffbody the concept of drag must be introduced first. One should keep in mind that it is only possible to distinguish between aerodynamic and bluffbodies if the typical Reynolds number of the flow is sufficiently high, as correctly pointed out by Buresti (2012). Drag is defined as the force component in the direction of motion. The forces acting over a body can, based on their origin, be divided into (skin-) forces and pressure forces. The friction forces act tangential to the surface and pressure forces, on the other hand, act normal to the sur- face. Accordingly, the drag can be divided into friction drag, also called viscous drag, and pressure drag, which is also referred to as form drag. The total drag is given by the sum of the viscous drag and the pressure drag. The total drag of an aerodynamic, or streamlined, body is dominated by viscous drag, whereas the pressure drag is much larger for bluffbodies. The pressure drag acting on bluffbodies is larger because of separation, which prevents the static pressure from fully recovering. It should be noted that the pressure at the base of the body is lower than the free-stream pressure. The part of the bluffbody before separation will be referred to as forebody, whereas the part of the body after separation is called base. For a blunt bluff body the point of separation is fixed at the sharp corners at the trailing edge. For a rounded object, e.g. a cylinder, the point of separation is typically a function of the Reynolds number. The drag on a bluffbody is directly related to the wake behind the bluff body that is formed by the separating boundary layers. Before discussing in more detail the complex structure of the fluctuations in this region of disturbed flow, a description can be given in terms of the mean streamwise velocity com- ponent U x, y .

( ) 3 4 2. THE WAKE OF BLUFF BODY

The velocity deficit present in the wake decreases and the wake width in- creases with increasing distance from the body in the downstream direction. In the far wake, x h 30 (Antonia & Rajagopalan 1990), the boundary layer ap- proximation holds and the wake becomes self-similar. By defining the velocity deficit U0 and the￿ > half wake width ∆y1 2 as: 1 U x U U x, 0 ,U￿ x, ∆y U U x , (2.1) 0 1 2 2 0 it can be shown( (Pope) ≡ ∞ 2000)− ( that) the wake( velocity￿ ) = profile∞ − is given( ) by: U U x, y y exp ln 2 , (2.2) U0 x ∆y1 2 ∞ − ( ) where y is the wall-normal direction.= ￿− ( ) ￿ ( ) ￿ 2.2. Vortex shedding The complexity of the fluid structures in the wake can be described by two dimensionless parameters, namely the Reynolds number Reh and the Strouhal number Sth: hU fsh Reh ,Sth , (2.3) ν∞ U with the length scale, h, the free-stream velocity, U , and the kinematic vis- = = ∞ cosity, ν. The vortex shedding frequency, f , is related to the periodicity with s ∞ which edies are being shed from the upper and lower surface of a bluffbody above a critical Reynolds number Reh,crit. For a cylinder Reh,crit 47 (Norberg 1994) and Reh,crit 120 (Buresti 2012) for a body with a blunt trailing edge. ≈ Above Reh 200 the initially laminar wake becomes three-dimensional and starts to transition≈ to a turbulent wake (Bloor 1964). The different stages and types of instabilities≈ involved in the transition of a laminar to turbulent wake are described by e.g. Zdravkovich (1997) and Williamson (1996). Due to the various instabilities that play a role, the relation between Reh and Sth is very complex, a complete overview is presented, among others, by (Norberg 1994). 5 In a large range, for a cylinder approximately given by 250 Reh 2 10 , Sth only varies over a relatively small range. According to equation (2.3) the shedding frequency, fs, is than inversely proportional to< the free-stream< × velocity, if the proper length scale h is chosen. The role that the boundary layer thickness and state play in this respect will be discussed in in Paper 2 and Paper 3 of the present thesis.

2.3. Drag prediction The drag acting on a bluffbody can be obtained from the mean velocity deficit in the wake that forms behind it. The forces acting on a fluid within a control volume enclosing the body must equal the rate of change of the momentum, i.e. the momentum flux. Here, the walls of the wind tunnel form a natural boundary 2.3. DRAG PREDICTION 5

A B

Control Volume Fτ

D U∞ − p y U y p∞ ( ) ( )

0.8 Figure 2.1. A schematic overview of the forces acting upon the fluid within the control volume.

of the control volume, which is completed with the fictitious boundaries A and B upstream and downstream of the body, respectively, as shown in figure 2.1. In a typical setup the mass is conserved within the control volume, but for the bluffbody that will be discussed in this thesis it is possible to remove fluid from the system. However, the amount of fluid removed is typically less than 1% of the inflow and therefore assumed negligible. Assuming that no mass is added or removed within the control volume, i.e. conservation of mass, implies:

lh ρ U U dy 0 , (2.4) 0 ∞ with lh denoting the height of￿ the( wind− tunnel) = and ρ the density of the fluid. Using this expression, the momentum flux can be written as:

lh lh Momentum flux ρ U 2 U 2 dy ρ U U U dy , (2.5) 0 0 ∞ ∞ which has to be equal= to￿ the￿ forces− acting￿ = upon￿ the fluid( − within) the control volume. As shown in figure 2.1, the forces that can be distinguished, besides the drag D exerted by the body, are the friction forces Fτ at the wind tunnel walls, the pressure forces due to pressure difference p p y and the viscous stresses τxx: ∞ − ( ) lh lh lh ρU U U dy p p y dy D Fτ τxx dy . (2.6) 0 0 0 ∞ ∞ ￿If the body( − is thin) compared= ￿ ( to− the( )) height− of− the wind+ ￿ tunnel,( ) the friction forces at the wind tunnel walls can be neglected. The drag coefficient Cd is 6 2. THE WAKE OF BLUFF BODY

1.5

1 Drag D Momentum flux [-]

d 0.5 Pressure term C Viscous stresses

0

-0.5 0510 15 20 25 30

x!h [-]

Figure 2.2. The drag coefficient Cd obtained at several streamwise positions and the other force contributions for con- stant flow conditions. then given by:

lh p p y U U y lh u 2 y ′ Cd 1 2 1 d 2 d , (2.7) 0 ρU 2 U U h 0 U 2 h ∞2 − ( ) where the= ￿ last￿ term is∞ based+ on∞ a￿ simplification− ∞ ￿￿ ￿ of￿ − the￿ viscous∞ stresses:￿ ￿ 2 τxx ρu , (2.8) ′ which, according to Van Dam (1999), can be used for an incompressible flow = − at high Reynolds numbers. The drag obtained from the velocity deficit should be independent of streamwise position where it is measured. In figure 2.2 the drag coefficient Cd is shown for several downstream locations and decomposed into different force contributions. CHAPTER 3

Boundary layers

The aim of this study is to vary the inlet conditions of the wake. In order to do so, a rectangular forebody with a permeable surface that allows for modification of the boundary layer by means of blowing and/or suction is used, a detailed description is given in Paper 1. In this chapter, a description of laminar and turbulent boundary layers with and without wall-suction will be given.

3.1. Previous Studies The idea to modify the boundary layer by means of suction was first discussed in the 1930s as a way to improve the of airplane wings. The advantage of suction in this context is twofold: a thinner boundary layer leads to a reduction of the induced wake, and the skin friction is likely to be reduced if the transition of the initially laminar boundary layer to a turbulent boundary layer is postponed or even prevented. Griffith & Meredith (1936) showed that an exact analytical solution exist for the laminar boundary layer over a flat plate with a uniform suction velocity, V0, applied over the surface. For this so called Asymptotic Suction Boundary Layer (ASBL), the growth of the boundary layer is opposed by the velocity component in the direction of the wall, with the result that the streamwise velocity U is no longer a function of the streamwise position x. The continuity equation then dictates that the wall normal velocity must be constant. As the name suggest, the boundary layer asymptotes towards the ASBL state, which is then only achieved sufficiently far downstream. A numerical solution of the boundary layer with uniform suction over a plate in a state before the ASBL is established is given by Iglish (1944), with suction starting at the leading edge of the plate, and Rheinboldt (1955), with suction starting after an initial entry length. A complete overview of the theory of laminar flow over a flat plate with uniform suction can be appreciated from (Rosenhead 1963). Kay (1953) was the first to show experimentally that the velocity profile does indeed approach the ASBL when uniform suction is applied over the sur- face of a flat plate. The ASBL could only be established and retained if the boundary layer was laminar at the start of the suction region, i.e. after an initial non porous entry length. No conclusive evidence for the existence of an equivalent turbulent asymptotic boundary layer could be given for an initially

7 8 3. BOUNDARY LAYERS turbulent boundary layer. However, it was shown that suction does lead to a considerable thinning of the turbulent boundary layer. A comprehensive experimental study on the existence of a turbulent as- ymptotic suction boundary layer was carried out by Dutton (1958). For a critical ratio V0 U a turbulent asymptotic suction profile exists, but no gen- eral asymptotic solution was found. The critical ratio depends on the state of ∞ the boundary layer￿ at commencement of suction and the surface type: porous or perforated. For values smaller or greater than this critical value there was no tendency, at least not within the length of the plate, for the boundary layer to reach a state in which the boundary layer thickness and the profile shape remained constant. As the suction increases, the velocity profile shape divides into two distinct parts, an inner part which resembles the laminar asymptotic suction profile and an outer layer which takes the form of a long thin tale which only slowly diminishes. Dutton (1958) suggested that the boundary layer might eventually relaminarize and reach the laminar asymptotic state. Two approaches can be distinguished in regard of describing the turbulent boundary layer subject to wall suction: those based on Prandtl’s mixing length theory (Rotta 1970; Simpson 1970) and the ones that seek for similarity laws for the inner and outer region with a common overlap region (Black et al. 1965; Tennekes 1965). The latter is only valid for moderate levels of suction and for turbulent boundary layers that are in equilibrium state. The first method, on the other hand, was proved by Black et al. (1965) to be in satisfactory agreement with the experimental results shown by Dutton (1958) and Kay (1953).

3.2. Boundary layer approximation Based on a scaling analysis of the Navier-Sokes equation the equations of motion can be simplified for the relatively thin boundary layer, see e.g. Kundu & Cohen (2008). For a steady, two-dimensional, incompressible flow the boundary layer equations (3.9, 3.10) and the continuity equation (3.11) read: ∂U ∂U 1 ∂P ∂2U U v ν , (3.9) ∂x ∂y ρ ∂x ∂y2 ∂P + 0 = − , + (3.10) ∂y ∂U ∂V = 0− . (3.11) ∂x ∂y For a boundary layer developing+ over= a solid surface, the above set of equations has to obey the following boundary conditions: U x, 0 0 ,Ux, U ,V x, 0 0 , (3.12) with the density ρ, the kinematic ν, the static pressure P and U and ( ) = ( ∞) = ∞ ( ) = V the velocity components in the streamwise x and wall normal y directions, 3.3. FALKNER-SKAN BOUNDARY LAYER 9 respectively. Equation (3.10) implicates that the pressure is uniform across the boundary layer, i.e. the static pressure at the surface is equal to the static pres- sure outside the boundary layer. Hence, the pressure gradient can be written in terms of the free stream velocity U using the Bernouilli equation as:

1 dP ∞ dU U . (3.13) ρ dx∞ dx∞ ∞ The thickness of the boundary layer= can be expressed as the displacement thickness δ1: U δ1 1 dy , (3.14) 0 ∞ U which is defined as the distance= ￿ by which￿ − the∞ ￿ wall should be moved outward in order to get the same mass flux in a hypothetical frictionless flow. Another expression of the boundary layer thickness can be given if not the mass flux but the momentum loss due to the presence of the boundary layer is considered. The momentum thickness δ2 is given by: U U δ2 1 dy . (3.15) 0 ∞ U U

The ratio of the displacement= ￿ thickness∞ ￿ and− momentum∞ ￿ thickness:

δ1 H12 , (3.16) δ2 is called the shape factor H12. = Hereafter, the laminar boundary layer without suction will be discussed, followed by a discussion on the laminar boundary layer with suction. There- after, the equivalent turbulent boundary layers will be described.

3.3. Falkner-Skan boundary layer Falkner and Skan (1931) showed that equation (3.9) has a similarity solution in case the velocity outside the boundary layer varies in the form: U x axn , (3.17) with n and a being constants. The∞( similarity) = variable η corresponds to: y U y η y Re . (3.18) δ x ￿ xν x x ∞ ￿ = = = If a similarity solution exists,( than) the velocity distribution U U in the wall normal direction at each location x is given by the same function g: ￿ ∞ U g η . (3.19) U

∞ = ( ) 10 3. BOUNDARY LAYERS

In order to solve the system of equations that is formed, a stream function ψ is introduced, with: ∂ψ ∂ψ u ,v , (3.20) ∂y ∂x and which thus obeys the continuity= equation= − (Eq. 3.11). It follows: y η η ψ Udy δ Udη δ U g η dη νxU f η , (3.21) 0 0 0 ￿ where: ∞ ∞ = ￿ = ￿ = ￿ df ( ) = ( ) g η . (3.22) dη Substitution of equation (3.17), equation( ) = (3.20) and equation (3.21) into equa- tion (3.9) with equation (3.13) gives an ordinary differential equation: n 1 f η f η f η nf η 2 0 , (3.23) 2 ′′′ + ′′ ′ with boundary conditions:( ) + ( ) ( ) − ( ) = f 0 0 ,f 0 0 ,f 1 . (3.24) Equation (3.24) is the so called Falkner-Skan′ equation,′ an equation that can be ( ) = ( ) = (∞) = solved numerically with a similarity solution of the velocity distribution. Using equation (3.14), the displacement thickness becomes:

δ1,FS bδ x , (3.25) with: η δ1,FS = ( ) b ∞ 1 f dη η f η . (3.26) η 0 δ x ′ ∞ ∞ = = ￿ = ( − ) = − ( ) 3.4. Asymptotic suction( ) boundary layer

Here, the laminar boundary layer with a uniform suction velocity, V0,willbe considered, starting with the asymptotic limit. Sufficiently far downstream the growth of the boundary layer, promoted by the viscosity, is exactly opposed by the negative velocity V y in the wall-normal direction. Hence, the boundary layer thickness remains constant and the streamwise velocity U is no longer a function of the streamwise( ) position x. The continuity equation (3.11) then dictates that the wall normal velocity must be constant and independent of the wall-normal coordinate y and equation (3.9) becomes: dU d2U V ν , (3.27) 0 dy dy2 with the boundary conditions given by:=

U x, 0 0 ,Ux, U ,V x, 0 V0 . (3.28) This differential equation can readily be shown to have an analytical solution: ( ) = ( ∞) = ∞ ( ) = U y yV 1 exp 0 , (3.29) U ν ( ) ∞ = − ￿ ￿ 3.4. ASYMPTOTIC SUCTION BOUNDARY LAYER 11

which is in the exponential form and known as the Asymptotic Suction Bound- ary Layer (ASBL). The boundary layer length scales and the shape factor are given by: ν 1 ν δ1,ASBL ,δ2,ASBL ,H12,ASBL 2 , (3.30) V0 2 V0 which are independent= of−U , a unique= property− further discussed= in Paper 2. The solution of the laminar∞ boundary layer with a zero pressure gradient and suction applied that has not yet reached the asymptotic state was described by Rheinboldt (1955). A similar variable substitution as shown for the Falkner- Skan solution can be made to find similarity. However, an additional similarity variable ξ is needed: x U ξ V ,η y . (3.31) 0 ￿ ￿U ν x∞ν Again, a stream function is= introduced∞ similar= to equation (3.21), but it is now a function, f, of ξ and η: ψ νxU f ξ,η . (3.32) ￿ Substitution of equation (3.32) using equation∞ (3.20) into equation (3.9) with = ( ) the pressure gradient equal to zero gives: ∂3f 1 ∂2f 1 ∂f ∂2f ∂f ∂2f f ξ , (3.33) ∂η3 2 ∂η2 2 ∂ξ ∂η2 ∂η ∂η∂ξ with the boundary conditions:+ + ￿ − ￿ f ξ,0 ξ,f ξ,0 0 ,f ξ, 1 . (3.34) In order to solve this equation, the′ boundary layer′ is progressively calculated for ( ) = ( ) = ( ∞) = points downstream of an initial position, typically the position where suction is commenced. The corresponding value of ξ0 is given by:

x0 ξ0 V0 , (3.35) ￿U ν

with x0 corresponding to the entry= length∞ from the start of boundary layer to the initial point where suction is applied. The displacement thickness δ1,RB is given by: η δ1,RB ∞ 1 f dη η f ξ,η f ξ,η0 . (3.36) η δ x 0 ′ ∞ ∞ Before the turbulent= ￿ boundary( − ) layer= can− be( discussed,) + ( a final) note on the ASBL should be( made) in regard of the effect of suction on the bypass transi- tion related to free stream . It has been shown both theoretically (Fransson & Alfredsson 2003), numerically (Schlatter & Orl¨u2011;¨ Khapko 2014) and experimentally (Fransson 2010; Yoshioka et al. 2004) that wall suc- tion suppresses the disturbance growth and may significantly delay or inhibit the break down to turbulence. 12 3. BOUNDARY LAYERS

3.5. Turbulent boundary layer The analysis found in most books about turbulent boundary layers, is based on the division of the turbulent boundary into two regions: the inner region where the turbulent stresses are negligible compared to the viscous stresses and the outer region where the viscous stresses are negligible compared to the turbulent stresses. The inner and outer layers have different characteristic length scales. The outer layer, similar to the laminar boundary layer, scales with the boundary layer thickness δ. The inner layer, on the other hand, scales with the viscous length scale ν uτ ,withuτ being the friction velocity given by:

￿ ∂U uτ τw ρ ,where τw µ , (3.37) ∂y y 0 ￿ = ￿ = ￿ = and τw is the wall shear stress. The scaled velocity can consequently be de- scribed by the functions Φ y and Ψ Y : + U yu Inner region:( ) U ( ) Φ τ Φ y , (3.38) + uτ ν + U U y Outer region: = =Ψ( )Ψ= Y( , ) (3.39) uτ δ ∞ − which are referred to as the law of the wall= ( and) = the( velocity) defect law, re- spectively. The inner and outer region overlap for large enough Reynolds numbers. Within this overlap region both laws must hold. The derivative of equation (3.38) and equation (3.39) multiplied with y yields the following expression: 1 ∂U dΦ dΨ y Y const. , (3.40) uτ ∂y + dy dY which must be constant because= the+ = two− relations= on the right depend on different length scales. As a result, the velocity profile must be logarithmic in this overlap region: 1 u ln y C1 , (3.41) + κ + U U 1 = ln Y +C2 , (3.42) uτ κ ∞ − = + with κ being the Von K´arm´anconstant and C1 and C2 other arbitrary con- stants. A self-similar equation can be found starting with the equation of mo- tion, as was done for the laminar cases. However, the flow properties are time dependent now, due to the turbulent fluctuations. By applying Reynolds de- composition, decomposing the instantaneous flow properties into mean and fluctuating components (U u¯ u ), the turbulent boundary layer equation ′ = + 3.5. TURBULENT BOUNDARY LAYER 13 can be derived: ∂u¯ ∂u¯ 1 ∂p¯ ∂2u¯ ∂ u¯ v¯ ν u v , (3.43) ∂x ∂y ρ ∂x ∂y2 ∂y ′ ′ ∂p¯ + 0 = − , + − ￿ ￿ (3.44) ∂y and the continuity equation reads:= − ∂u¯ ∂v¯ 0 (3.45) ∂x ∂y Due to the additional Reynolds stress+ term= u v there is no solution to this set of equations, unless a turbulence model is used′ to′ express it in terms of the mean velocity field. The Prandtl mixing length is such a turbulence model. It assumes that lumps of fluid retain their momentum while being displaced by large eddies over distance lm, referred to as the mixing length. Then the velocity fluctuations u y and v y can be written as: ′ ′ ∂u¯ ( ) ( ) u y v y lm , (3.46) ′ ′ ∂y and consequently: ( ) ∼ ( ) ∼ ∂u¯ 2 u v l2 . (3.47) m ∂y ′ ′ The mixing length lm is assumed− to∼ be￿ proportional￿ to the distance from the wall with the Von K´arm´anconstant κ, introduced earlier, as the proportionality constant:

lm κy . (3.48) Within the overlap region, where the turbulent stress is approximately equal 2 = to the wall shear stress ρuτ , one obtains: u ∂u¯ τ , (3.49) κy ∂y which can readily be shown to yield the= same logarithmic velocity distribution found by the two layer approach. The Prandtl mixing length model is nowadays considered to be obsolete, but has proved to work rather well (Davidson 2004). The reason it is introduced here, is the lack of an equivalent two layer approach for the turbulent boundary layer with suction. As will be shown in the next section, Prandtl’s mixing length model allows to give a good description. The logarithmic velocity distribution within the overlap region can be ex- tended to hold in the outer layer and the viscous sublayer by adding a wake function and applying an appropriate damping function, respectively. Such an example for a so called composite velocity profile can be found in (Nickels 2004) for example. 14 3. BOUNDARY LAYERS

3.6. Turbulent suction boundary layer To describe the turbulent boundary layer with suction, the theory by Rotta (1970) is followed, starting with the turbulent boundary layer equation and the continuity equation (3.43- 3.45). For the asymptotic suction boundary layer the derivatives of the mean flow quantities with respect to x are rendered zero. Although an asymptotic state might not be reached for the turbulent suction boundary layer, it is assumed that these terms will be relatively small and can therefore be neglected. Integrating the remaining terms of equation (3.43) leads to: du¯ ν u v u2 V u,¯ (3.50) dy τ 0 ′ ′ where the invoked boundary layer− conditions,= + which are similar to those stated in equation (3.12), lead to the additional wall shear stress term on the right hand side. To be able to solve this equation, the previously discussed Prandtl’s mixing length model is introduced, i.e. equation (3.47) and equation (3.48). To include the viscous sublayer, the van Driest damping function is applied (Van Driest 1956), leading to: 2 2 2 u V0u¯ ∂u¯ u v κ2y2 1 exp y τ , (3.51) ￿ ￿ νA ￿ ∂y ′ ′ ￿ ￿ ￿ ￿ + ￿￿ − = ￿ − − ￿ ￿ ￿ with A being the van Driest￿ constant.￿ Upon substitution￿￿ of equation (3.51) into equation (3.50) one will￿ recognize a quadratic equation￿ which can readily be shown by invoking the quadratic formula to give: 2 du¯ 2 u V0u¯ τ , (3.52) dy 2 2 2 2 2 2 ν ν 4κ y 1 exp ￿y +u ￿V0u¯ νA u V0u¯ = ￿ τ τ ￿ which can+ be solved+ using￿ − a numerical￿− ( solver.+ )￿( )￿￿ ( + ) In the present study, the van Driest constant A, the friction velocity uτ and the wall position yw are obtained from the experimental data in a least square fit algorithm based on equation (3.52). An initial “guess” of yw is found by a linear extrapolation of points close to the wall and the other parameters are discussed below. Van Driest (1956) treats A as a constant, but notes that it is more likely to be a function of the wall suction. In the present study A is therefore a variable, with the value used by Van Driest (1956) (A 26) as the initial value in the fit algorithm. The variation of A with free stream velocity and wall suction is shown in figure 3.1. An average of A 28.7= is found for the cases without suction (markers colored black) and a decrease is shown for an increase of wall suction. One outlier, which corresponds to= U 10 m s subject to the largest wall suction, is likely caused by relaminarization. ∞ = ￿ 3.6. TURBULENT SUCTION BOUNDARY LAYER 15

100

U∞ = 3m!s 80 U∞ = 5m!s U∞ = 10 m!s U∞ = 15 m!s

A 60 U∞ = 20 m!s

40

20 010002000 3000 4000 5000 6000

Reδ2 [-]

Figure 3.1. The van Driest constant A as function of the

Reynolds number Reδ2 . A lighter shade of gray corresponds a larger wall normal velocity, with a black face color correspond- ing to the case without suction.

1.5 U∞ = 3m!s U 5m s 1.45 ∞ = ! U∞ = 10 m!s U∞ = 15 m!s 1.4

[-] U∞ = 20 m!s 12 H 1.35

1.3

1.25 0 1000 2000 3000 4000 5000 6000

Reδ2 [-]

Figure 3.2. The shape factor H12 as function of the Reynolds

number Reδ2 . 16 3. BOUNDARY LAYERS

10−3 5.5 × U∞ = 3m"s 5 U∞ = 5m"s

4.5 U∞ = 10 m"s

U∞ = 15 m"s 4 [-] U∞ = 20 m"s f C 3.5 Osterlund¨

3

2.5

2 010002000 3000 4000 5000 6000

Reδ2 [-]

Figure 3.3. The skin-friction as function of the Reynolds ¨ number Reδ2 compared to a relation found by Osterlund (1999). A lighter shade of gray corresponds a larger wall nor- mal velocity, with a black face color corresponding to the case without suction.

A higher value of the shape factor is also found for the outlier, as shown in figure 3.2. Without wall suction, the shape factor lies in the range H12 1.3 1.5 and decreases with increasing Reynolds number as expected. As wall suction increases, the shape factor decreases, i.e. an increase in wall suction corresponds= − to a fuller profile. A final check on the validity of the fit algorithm is obtained by analyzing the skin-friction coefficient Cf given by: u2 C 2 τ (3.53) f U 2 = The initial value of uτ is based on the ICET∞ skin-friction relation (Bailey et al. 2013). In figure 3.3 the results for Cf are compared to the relation found by Osterlund¨ (1999). A fairly good agreement is found for the cases without wall suction. With increasing wall suction, the velocity profile becomes fuller and as expected Cf increases (since du¯ dy increases) and the Reynolds number based on δ2 decreases. Suction is known to stabilize￿ the flow. Due to suction the u v is signifi- cantly reduced (Antonia et al. 1994). The reduction in v and to a′ lesser′ extent ′ 3.6. TURBULENT SUCTION BOUNDARY LAYER 17

u can be explained by an increased organization of the flow with less sweep and ejection′ events as well as more steady near-wall streamwise streaks. Although the du¯ dy is increased it must mean that the production is reduced and that dissipation ultimately leads to complete relaminarization. ￿ CHAPTER 4

Experimental setup and measurement techniques

Measurements are carried out in a new test section mounted in the Boundary Layer (BL) wind tunnel at KTH Mechanics. The new test section is described in detail in Paper 1. In the same paper, the corrections of hot-wire and Pitot tube measurements required for an accurate experimental validation of the new test section are given. In this chapter, a complementary description of hot-wire anemometry, pressure measurements and an introduction to particle image velocimetry is presented.

4.1. Hot-wire anemometry The Hot-Wire Anemometry (HWA) technique is based on the forced-convective heat transfer from a heated wire by the flow it is exposed to. How the heat transfer is translated into a measure of the flow velocity depends on type of HWA that is being used. The most obvious procedure, referred to as Constant Current Anemometry (CCA), is to keep the current heating the wire constant and use the change in voltage over wire due to the changing resistance as a mea- sure for the flow velocity. To measure turbulent flow properties it is common to use Constant Temperature Anemometry (CTA), because this alternative al- lows a high sampling frequency, up to kHz in practice. In the CTA mode the wire temperature is kept constant and the voltage required to do so is used as a measure for the velocity.

The wire is operated at a constant temperature, Th, that is higher than the ambient temperature to make the sensor less sensitive to temperature fluc- tuations. Based on the definition:

R T R T OH h cal , (4.54) R Tcal ( ) − ( ) = ( ) with R the resistance of the wire and Tcal the temperature of the surrounding fluid during calibration. The overheat OH is typically chosen between 0.5 and 0.8, where a higher overheat captures turbulent fluctuations better. For boundary layer measurements a lower overheat is sometimes used, as will be discussed shortly.

18 4.1. HOT-WIRE ANEMOMETRY 19

15 s]

! 10 [m U

5 Modified King’s law Calibration data

0 0.6 0.7 0.8 0.9 E [V]

Figure 4.1. Typical hot-wire calibration curve. In the inset, the hot-wire probe used for boundary layer measurements is shown.

One relation between the anemometer output voltage E and the velocity U is given by the modified King’s law (Johansson & Alfredsson 1982):

1 1 2 2 n 2 U k1 E E0 k2 E E0 , (4.55) where E0 is the voltage= at U￿ 0m− s￿ and+ the( coe− fficients) k1, k2 and n are to be found by calibration of the hot-wire against the reading from e.g. a Prandtl tube located in the free stream.= An￿ example of a calibration curve is shown in figure 4.1. The designation ‘modified’ refers to the second term on the right- hand side which is added to account for the effect of free convection. Free convection has a significant contribution at low velocities, for example close to a wall. In figure 4.1 the in-house made boundary layer hot-wire probe is shown. The hot-wire is typically soldered between two prongs at the extremity of the hot-wire probe. Tungsten and platinum are common materials for the hot-wire, because these metals have a high melting point and their resistance varies lin- early with temperature for the temperatures of interest. For increased strength a platinum/rhodium alloy can be used. The diameter of the hot-wire is in the order of micrometers in diameter and has a length to diameter ratio of lh dh 200 so that the heat conduction to the prongs can be neglected.

￿ > 20 4. EXPERIMENTAL SETUP AND MEASUREMENT TECHNIQUES

5

4 Exp. Data Discarded points 3 [-]

1 ASBL δ !

y 2 V0!U∞ 1

0 00.20.4 0.6 0.8 1

U!U∞ [-]

Figure 4.2. Boundary layer measurements with a hot-wire (U 1.0ms, v0 U 0.012).

∞ = ￿ ￿ ∞ = 1

Exp. Data Artificial Points

[-] ASBL st

∞ order u 1 U 0.5 nd − 2 order 1 3thorder 4thorder

0 ] −

[ 0.2 # ∞ U U − 1 ! ∞

U 0.1 U

0 0 2 4 6 y [mm] 0.3 Figure 4.3. Displacement and momentum thickness inte- grand, including the Taylor series expansion. 4.1. HOT-WIRE ANEMOMETRY 21

4.1.1. Near-wall boundary layer measurements Operating a hot-wire within a boundary layer over a porous surface with suction applied in the wall-normal direction raises the question whether the wall normal velocity leads to a significant increase of the forced-convection. Because the wall normal velocity is much smaller than the streamwise velocity component, even close to the wall as indicated in figure 4.2, the additional contribution can be neglected. Figure 4.2 does show that the velocity is overestimated close to the wall. The overestimation is caused by heat conduction from the hot-wire to the wall, an error that can be minimized by keeping the overheat ratio low. Points from the wall up to the point of minimum velocity are discarded. In addition, one more point is omitted if the velocity derivative has not reached a maximum. It is apparent that the boundary layer measurements close to the wall are underresolved. This leads to an underestimation of the momentum thickness and to lesser extent to an overestimation of the displacement thickness, see figure 4.3. To improve the estimate of the momentum thickness a Taylor series expansion of the displacement thickness is used to obtain additional artificial points close to the wall in case of the ASBL. For the ASBL, the Taylor series expansion at the wall of the displacement thickness integrand I is given by: U V V I 1 1 1 exp 0 y exp 0 y , (4.56) U ν ν

= − ∞ = − ￿ − ￿− ￿￿ = ￿− ￿ V 1 V 3 1 V 4 I 0 1 0 y 0 y3 0 y4 O y4 . (4.57) ν 2 ν 6 ν It is clear from( figure) = 4.3− that+ the￿ first￿ four− terms￿ will￿ give+ a( proper) approxima- tion. The additional artificial points do not lead to a considerable improvement in case of the displacement thickness, but it does show a clear effect for the momentum thickness.

4.1.2. Vortex shedding frequency measurements A hot-wire is also used to measure the shedding frequency. Although the shedding frequency can be obtained from measurements in the boundary layer or in the wake, a better signal-to-noise ratio can be obtained with a probe placed downstream of the trailing edge just outside the wake, as shown in figure 4.4. The sampling frequency for most measurements is chosen to be fs 3 kHz, which is more than enough for capturing the shedding frequency considering the Nyquist criteria. = The anemometer offers the possibility to apply a high- and a low- pass filter. The high-pass filter can be used to remove low frequency fluctuations if present, which is not the case in the present wind tunnel. A low-pass filter of 0.3 kHz is used to avoid aliasing and to remove electronic noise. 22 4. EXPERIMENTAL SETUP AND MEASUREMENT TECHNIQUES

0.1

Wake Boundary Layer 0.05 [V] ¯ E 0 − E

-0.05

-0.1 0 0.1 0.2 0.3 0.4 0.5 t [s]

Figure 4.4. Sample of the time signal as recorded by the hot-wire in the boundary layer and in the wake.

4.2. Pressure measurements Pressure measurements typically comprise a Pitot or a Prandtl tube, which are robust compared to the fragile hot-wire probes discussed in the previous section. A number of a-posteriori corrections have to be considered to obtain accurate results. A correction method is described in Paper 1 and will therefore not be repeated here. Examples of a Pitot tube and a Prandtl tube are shown in figure 4.5.

4.2.1. Near-wall boundary layer measurements Velocity measurements in the boundary layer are carried with a Pitot tube. It is common to start the wall normal traverse with the probe touching the wall, or being even slightly bended. As a result, some point close to the wall must be discarded. First, the first derivative of u with respect to y is calculated. All points for which a negative, unphysical, gradient is found are discarded. Then, the second derivative is calculated and all points up to the maximum are discarded as well. This method has been checked extensively and proved to be the best objective method to discard points. An example of a boundary layer measurement and the points that are discarded is shown in figure 4.6. This example is measured in combination with a Furness FC0510 (0 200 2000 Pa, 0.25% FSD) pressure transducer. − ￿ 4.2. PRESSURE MEASUREMENTS 23

1

Figure 4.5. Left: Near wall boundary layer measurement with a Pitot tube. Right: Parallel mounted hot-wire probe (Left) and Prandtl tube (Right) at the extremity of a 1D tra- verse.

5

4 Exp. Data Discarded points 3 [-] ASBL 1 δ !

y 2

1

0 00.20.4 0.6 0.8 1

U U∞ [-] ! Figure 4.6. Boundary layer measurements with a Pitot tube (U 2.5ms, V0 U 0.0057).

∞ = ￿ ￿ ∞ = The reference static pressure in the boundary layer can be obtained from e.g. a Prandtl probe located in the free stream. It follows from the bound- ary layer approximation that the static pressure is constant throughout the boundary layer.

4.2.2. Near-wake pressure measurements In the far wake, x h 30 (Antonia & Rajagopalan 1990), the thin shear layer approximation is valid. Similar to the boundary layer equivalent, the static pressure is constant￿ in> the cross stream direction. However, in the near wake this approximation does not hold which means that the static pressure must be measured locally. Therefore a Prandtl tube is used in the near wake. 24 4. EXPERIMENTAL SETUP AND MEASUREMENT TECHNIQUES

5 [-]

h 0 ! y

-5

6 810 15 25 x h [-] ! Figure 4.7. A relative comparison of the static pressure evo- lution in the streamwise direction (U 10 m s)

∞ = ￿ The design of a static-pressure tube is not straight-forward. In particular the location of the static pressure holes is essential for an accurate pressure measurement (Tropea et al. 2007). At the tip of the probe the flow will accel- erate. As a result the static pressure will be overestimated up to four diameters downstream. The support or stem downstream of the pressure holes, on the other hand, causes the local static pressure to increase due to the flow decel- eration, which extents over twenty probe diameters upstream. Some Prandtl tubes are designed in such a way, L-shaped, that the influence of the tip and support cancel each other out. In the present study, a straight Prandtl tube is used, The static pressure holes are located far enough downstream of the tip and far enough upstream of the traverse stem to avoid influence thereof. In figure 4.7 the evolution of the static pressure in the streamwise direc- tion is shown, measured with a ScaniValve 16 channel Digitial Sensor Array (2500 Pa, 0.2% FSD) pressure transducer. It is clear that the static pressure recovers, but that it has not fully recovered at x h 25. The recovery be- tween x h 6 and x h 25 is about 60%. So the use of a Prandtl tube is not redundant. ￿ = ￿ = ￿ =

4.3. Particle Image Velocimetry Hot-wire anemometry and pressure measurement techniques are not suitable to be used in the recirculation region of a bluffbody due to reversed flow direction and strong three dimensionality of the flow. An appropriate method to use in these type of flow is Particle Image Velocimetry. PIV is non-intrusive and is able to capture the instantaneous flow structure, which can comprise two or three velocity components. 4.3. PARTICLE IMAGE VELOCIMETRY 25

PIV is based on the measurement of the displacement of tracer particles during a short time interval. The tracer particles, e.g. smoke for the application in a wind tunnel, makes the optically transparent fluid visible. Two sheets of particles are illuminated succesively and recorded with a (high-speed) camera. The two images are divided into interrogation areas for which the velocity vector is computed by means of cross correlation. By means of proper orthogonal decomposition (POD) and dynamic mode decomposition (DMD) the velocity fields measured with PIV can be decom- posed into small scale turbulence and cycle-to-cycle variations. Furthermore, the time-averaged flow field can be used to carry out detailed stability analyses, as was done by Camarri et al. (2013). CHAPTER 5

Summary of results and conclusions

The goal of this thesis is to contribute to the fundamental knowledge about the aerodynamics in the wake behind bluffbodies. In order to do so, an exper- imental study has been carried out in a specially designed setup, consisting of a rectangular forebody with a smooth leading edge and a blunt trailing edge. It is shown both theoretically and experimentally in Paper 1 that the main feature of the new test section, to modulate the boundary layer and therewith the inlet conditions of the wake by means of suction, makes it possible to perform unique parametrical studies. This is due to the asymptotic suction boundary layer (ASBL) that is reached at the trailing edge of the rectangular forebody for a suction coefficient Γ 9 in case of an initially laminar boundary layer. The ASBL provides the unique feature that the Reynolds number and the boundary layer thickness can be> altered independent of each other. The experimental setup is experimentally evaluated with a series of hot- wire and Prandtl tube measurements in the boundary layer and wake of the rectangular-based forebody. The ASBL threshold proved to be independent of 3 3 the Reynolds number within the range Reh 2.9 10 6.7 10 . Based on the mean velocity deficit and the corresponding rms-values in the wake it was concluded that the wake is symmetric. Over= the× forebody− × the flow was found to be two dimensional but for low velocities (U 2.5ms) the wake at x h 15 is not, possibly due to the presence of the wind tunnel walls. ∞ ≤ ￿ 5 In Paper 3 the initially turbulent boundary layer (Rex 4.5 10 ￿ =6 3.0 10 ) subject to wall suction is investigated. The Γcrit for which relam- inarization is observed is found to be a function of Reh. Strong= evidence× − is provided× that the turbulent boundary layer will ultimately attain the laminar ASBL as well, provided that the wall suction is strong enough. The influence of a modulated laminar boundary layer on the wake char- acteristics is considered in Paper 2. A decrease of the displacement thickness leads to a decrease of the base base pressure. A similar clear trend is not ob- served for the shedding frequency. In the ASBL limit the shedding frequency becomes Sth 2δ1 0.29, independent of the Reynolds number in the range Re 2.9 103 6.7 103. The base pressure coefficient in the ASBL limit h + is Cp,b 0.36 for≈ all cases but Reh 2877, for which a much lower value is found.= × − × = − = 26 5. SUMMARY OF RESULTS AND CONCLUSIONS 27

In Paper 3 the modulated turbulent boundary layers, as well as near wake measurements, are included in the study on the effect of inlet variations on the wake characteristics. The shape of the mean wake velocity profile, scaled with the velocity deficit U0 and the wake half width ∆y1 2, is found to be independent of x h, for x h 6 and Re 6.7 103. The wake width is shown to scale with h ￿ the effective thickness of the body h 2δ1, where the ratio is expected to vary with￿ the downstream￿ > location.> From× the near wake measurements with a hot- wire it could be concluded that the turbulence+ intensity based on the urms level 3 on the centerline remained unchanged for Reh 6.7 10 . It can be concluded that the state of the boundary layer has a significant effect on the vortex shedding frequency and base> pressure.× The Strouhal num- ber based on the effective plate thickness is Sth 2δ1 0.22 for the turbulent boundary layer is, i.e. about 25% lower than the value found for the laminar + ≈ boundary layer in the ASBL limit (Sth 2δ1 0.29). When the initially turbu- lent boundary layer starts to relaminarize, the vortex shedding frequency and + the base pressure start to approach those found≈ in the ASBL limit. The total drag, obtained from near wake pressure measurements, is also discussed in Paper 3. The maximum drag reduction found is 30% for Reh 2.7 104 with a wall normal suction velocity of only 3.5% of the free stream velocity. Wall suction successively reduces the total drag width increasing wall= 3 4 suction,× at least in the range Reh 8.0 10 5.5 10 . The observations made in this study are an addition to the existing in- sights. They do certainly not provide= enough× − information× to fully understand the governing physics required for an applicable control of the wake, e.g. to accomplish drag reduction for road vehicles. It is believed that an even better understanding can be obtained if also the mean velocity and velocity fluctua- tions very close to the body, in the recirculation region, are known. Particle image velocimetry measurements are therefore now planned. Those measure- ments will make it possible to not only study the flow structures themselves by means of proper orthogonal decomposition (POD) and dynamic mode de- composition (DMD), but also to carry out detailed stability analyses on the experimental data. CHAPTER 6

Papers and authors contributions

Paper 1 A new test section for wind tunnel studies on wake instability and its control. Bengt E. G. Fallenius (BF), Renzo Trip (RT) & Jens H. M. Fransson (JF)

This paper is written by BF and RT with feedback from JF. RT wrote the chapter on the experimental validation of the setup and performed the mea- surements described in this chapter. Parts of this work have been presented at the 9th European Fluid Mechanics Conference 2012, Rome, Italy. Paper 2 An experimental study on the wake of a rectangular forebody with suction ap- plied over the surfaces. Renzo Trip (RT) & J. H. M. Fransson (JF).

This conference paper is written by RT with feedback from JF. The exper- imental data described in the paper are obtained and analyzed by RT with guidance of JF. This work is presented at the 4th International Conference on Jets, Wakes and Separated Flows (ICJWSF-2013), Nagoya, Japan. Paper 2 An experimental study on the wake behind a rectangular forebody with a lami- nar and turbulent suction boundary layers. Renzo Trip (RT) & J. H. M. Fransson (JF).

RT planned, performed and analyzed the experiments shown in this paper under supervision of JF. The paper was written by RT with feedback from JF.

28 CHAPTER 7

Acknowledgements

A PhD, or a licentiate at this stage, is not a project one can bring to a good end without the guidance, the knowledge, the craftsmanship, the support and the company of others. I would therefore like to express my sincere gratitude to those who have contributed to this thesis. The first person I would like to thank is my supervisor prof. Jens Fransson. For his trust in me as a future researcher, and for providing me with the knowledge and support required in the process of becoming. Here, I should also thank my co-supervisor Dr. Bengt Fallenius who has not only provided me with a well thought-out experimental setup but also introduced me to experimental techniques and helped me out numerous times solving problems that seemed insurmountable. Although face-to-face meetings were limited due to their residence abroad, the contributions of Dr. Simone Camarri and prof. Alessandro Talamelli are considered to be of great value. Dr. Ramis Orlu,¨ Dr. Alexandre Suryadi and Dr. Robert Downs have been most helpful by sharing their broad knowledge and experience in the field of (experimental) fluid mechanics. Toolmakers Joakim Karlstr¨omand G¨oran R˚adberg deserve a special word of appreciation, their skills and expertise are indispensable for any experimen- talist. The administrative staffat the department of Mechanics has also been of great support in various matters. Prof. Henrik Alfredsson, Dr. Nils Tillmark and Docent Fredrik Lundell are acknowledged for the pleasant and inspiring working environment, as well as the good collaboration between the different research groups in the Fluid Physics Laboratory. It has been great to work in the midst of my esteemed colleagues Sohrab Sattarzadeh and Marco Ferro. As my roommates, they have often been the first ones to turn to when the dynamics of a fluid confused me. The friendly atmosphere in the office I have also experienced during fikas, lunch breaks and afterworks with my former and current coworkers. Our, of- ten humorous, discussions have inspired me to endeavor, sometimes jointly,

29 innebandy, running, swimming, climbing, baking and swedish lessons. Obvi- ously, not all leading to a resounding success, eller hur?

The Swedish Research Council (VR) is acknowledged for their financial sup- port. References

Antonia, R. & Rajagopalan, S. 1990 Determination of drag of a circular cylinder. AIAA journal 28 (10), 1833–1834. Antonia, R., Spalart, P. & Mariani, P. 1994 Effect of suction on the near-wall anisotropy of a turbulent boundary layer. Phys. Fluids 6 (1), 430–432. Bailey, S. C. C., Hultmark, M., Monty, J. P., Alfredsson, P. H., Chong, M. S., Duncan, R. D., Fransson, J. H. M., Hutchins, N., Marusic, I., McKeon, B. J. et al. 2013 Obtaining accurate mean velocity measurements in high reynolds number turbulent boundary layers using pitot tubes. J. Fluid Mech. 715,642–670. Bearman, P. W. 1965 Investigation of the flow behind a two-dimensional model with a blunt trailing edge and fitted with splitter plates. J. Fluid Mech. 21 (part 2), 241–256. Bingham, R. 2007 Plasma physics: On the crest of a wake. Nature 445 (7129), 721–722. Black, T., Sarnecki, A. & Mair, W. 1965 The turbulent boundary layer with suction or injection. HM Stationery Office. Bloor, M. S. 1964 The transition to turbulence in the wake of a circular cylinder. J. Fluid Mech. 19 (02), 290–304. Buresti, G. 2012 Elements of fluid dynamics. World Scientific. Buresti, G., Fedeli, R. & Ferraresi, A. 1997 Influence of afterbody rounding on the pressure drag of an axisymmetrical bluffbody. J. Wind Eng. Ind. Aerodyn. 69,179–188. Camarri, S., Fallenius, B. E. G. & Fransson, J. H. M. 2013 Stability analysis of experimental flow fields behind a porous cylinder for the investigation of the large-scale wake vortices. J. Fluid Mech. 715,499–536. Davidson, P. A. 2004 Turbulence: An Introduction for Scientists and Engineers: An Introduction for Scientists and Engineers. Oxford University Press. Durgesh, V., Naughton, J. W. & Whitmore, S. A. 2013 Experimental inves- tigation of base-drag reduction via boundary-layer modification. AIAA journal 51 (2), 416–425. Dutton, R. 1958 The effects of distributed suction on the development of turbulent boundary layers . HM Stationery Office.

31 Fransson, J. H. M. 2010 Turbulent spot evolution in spatially invariant boundary layers. Phys. Rev. E 81 (3), 035301. Fransson, J. H. M. & Alfredsson, P. H. 2003 On the disturbance growth in an asymptotic suction boundary layer. J. Fluid Mech. 482,51–90. Griffith, A. & Meredith, F. 1936 The possible improvement in aircraft perfor- mance due to the use of boundary layer suction. Royal Aircraft Establishment Report No. E 3501,12. Jodai, Y., Takahashi, Y., Ichimiya, M. & Osaka, H. 2008 The effects of splitter plates on turbulent boundary layer on a long flat plate near the trailing edge. J. Fluids Eng. 130 (5). Johansson, A. V. & Alfredsson, P. H. 1982 On the structure of turbulent channel flow. Journal of Fluid Mechanics 122,295–314. Kay, J. M. 1953 Boundary-layer flow along a flat plate with uniform suction.HM Stationery Office. Khapko, T. 2014 Transition to turbulence in the asymptotic suction boundary layer. TRITA-MEK Tech. Rep. 2014:03. Licentiate Thesis, Royal Institute of Technol- ogy, Stockholm. Kundu, P. & Cohen, I. 2008 Fluid Mechanics. Elsevier Academic Press. Mariotti, A. & Buresti, G. 2013 Experimental investigation on the influence of boundary layer thickness on the base pressure and near-wake flow features of an axisymmetric blunt-based body. Exp. Fluids 54 (11), 1–16. Nickels, T. 2004 Inner scaling for wall-bounded flows subject to large pressure gra- dients. J. Fluid Mech. 521,217–239. Norberg, C. 1994 An experimental investigation of the flow around a circular cylin- der: influence of aspect ratio. J. of Fluid Mech. 258,287–316. Osterlund,¨ J. M. 1999 Experimental studies of zero pressure-gradient turbulent boundary layer flow .PhDThesis,RoyalInstituteofTechnology,Stockholm. Pope, S. B. 2000 Turbulent flows . Cambridge university press. Rheinboldt, W. 1955 Zur Berechnung station¨arer Grenzschichten bei kontinuier- licher Absaugung mit unstetig ver¨anderlicher Absaugegeschwindigkeit. Rosenhead, L. 1963 Laminar Boundary Layers: An Account of the Development, Structure, and Stability of Laminar Boundary Layers in Incompressible Fluids, Together with a Description of the Associated Experimental Techniques. Claren- don Press. Rotta, J. C. 1970 Control of turbulent boundary layers by uniform injection and suction of fluid. Tech. Rep.. DTIC Document. Rowe, A., Fry, A. & Motallebi, F. 2001 Influence of boundary layer thickness on base pressure and vortex shedding frequency. AIAA J 39 (4). Schlatter, P. & Orl¨ u,¨ R. 2011 Turbulent asymptotic suction boundary layers studied by simulation. In Journal of Physics: Conference Series,,vol.318,p. 022020. IOP Publishing. Simpson, R. L. 1970 Characteristics of turbulent boundary layers at low reynolds numbers with and without transpiration. J. Fluid Mech. 42 (04), 769–802. Tennekes, H. 1965 Similarity laws for turbulent boundary layers with suction or injection. J. Fluid Mech. 21 (part 4), 689–703. Tropea, C., Yarin, A. L. & Foss, J. F. 2007 Springer handbook of experimental fluid mechanics, , vol. 1. Springer. Van Dam, C. 1999 Recent experience with different methods of drag prediction. Progress in Aerospace Sciences 35 (8), 751–798. Van Driest, E. 1956 On turbulent flow near a wall. Journal of the Aeronautical Sciences (Institute of the Aeronautical Sciences) 23 (11), 1007–1011. Williamson, C. 1996 Vortex dynamics in the cylinder wake. Annu. Rev. Fluid Mech. 28 (1), 477–539. Yoshioka, S., Fransson, J. H. M. & Alfredsson, P. H. 2004 Free stream tur- bulence induced disturbances in boundary layers with wall suction. Phys. Fluids 16,3530. Zdravkovich, M. M. 1997 Flow around Circular Cylinders: Volume 1: Fundamen- tals, , vol. 1. Oxford University Press.