+ All Categories
Home > Documents > Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

Date post: 09-Dec-2016
Category:
Upload: vipin
View: 213 times
Download: 1 times
Share this document with a friend
11
Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone Sharon O. Stephen and Vipin Michael University of Birmingham, Birmingham, England B15 2TT, United Kingdom DOI: 10.2514/1.J052082 A theoretical linear-stability analysis is used to consider the effect of a passive porous wall on the first Mack mode of a hypersonic boundary layer on a sharp slender cone. The effects of curvature and of the attached shock are included for axisymmetric and nonaxisymmetric disturbances. The flow in the hypersonic boundary layer is coupled to the flow in the porous layer. The theoretical models of the porous walls developed in previous studies are used for regular microstructures and random microstructures. The physical parameters chosen correspond to those of their studies. The resulting transcendental equations are solved numerically. Neutral solutions are presented, indicating a destabilizing effect of the porous wall. The spatial growth rates determined demonstrate that the porous wall leads to a significant increase in growth rates for nonaxisymmetric modes. The effects of varying the porous-wall parameters are investigated. Nomenclature A = displacement function A y = admittance A Y = scaled admittance a = cone radius ~ a = half-pore width a = tortuosity C D = dynamic compressibility, C D γp d = fiber diameter h = nondimensional porous-layer thickness, h δ Kn = Knudsen number L = length scale M = Mach number P = pressure disturbance Pr = Prandtl number R = Reynolds number, U δ ν Re = Reynolds number, U L ν r p = nondimensional pore radius, r p δ r s = shock location T = temperature tan h = hyperbolic tangent in Eqs. (2) and (7) U = velocity u, v = velocity disturbance x, r, ϕ = orthogonal coordinates Z 0 = characteristic impedance α = disturbance wave number, α r iα i γ = specific-heat ratio δ = boundary-layer displacement thickness θ c = cone angle θ s = shock angle Λ = propagation constant λ = skin friction μ = viscosity ν = kinematic viscosity ρ = density ρ D = dynamic density, ρ D ρ W ϕ 0 = porosity Ω = disturbance frequency ω = angular frequency Subscripts s = shock w = wall = just behind the shock I. Introduction T RANSITION to turbulence in hypersonic flows is associated with amplification of the first and/or second Mack modes. The first Mack mode is the high-speed counterpart of TollmienSchlichting waves, and so a viscous instability, with modes located close to the boundary. The second Mack mode is an inviscid instability. The second mode is believed to be responsible for transition to turbulence on hypersonic slender bodies. These modes correspond to the unstable perturbations occurring in the low- frequency band and the high-frequency band, respectively, of the slow (S) mode identified by Fedorov and Tumin [1]. Recent experiments have shown that a porous coating greatly stabilizes the second Mack mode of the hypersonic boundary layer on sharp slender cones [24]. The effect of the porous coating is to reduce the growth rates of the second mode to a level where they are comparable with those of the first mode (occurring at lower frequencies). In addition, the first mode is observed to be slightly destabilized by the presence of the porous coating. Thus, the first mode may now be more significant in the transition process. The destabilizing effect of these porous walls on the first Mack mode has been confirmed by recent numerical simulations [510]. This destabilization has been shown to be significant for a felt-metal porous coating [6]. A previous work on the linear instability of the first Mack modes in the hypersonic flow over a cone includes the effect of curvature and also the attached shock [11]. The effect of curvature is significant with disturbances only existing over a finite range of wave numbers for a fixed radius. Beyond a critical radius, dependent on the azimuthal wave number, all disturbances are damped. It was demonstrated that the effect of the shock, which allows incoming and outgoing waves, gives rise to multiple modes [11]. This study showed that the effect of the shock should not be neglected; otherwise, the correct effect of curvature will not be realized. The modes that exist in the absence of the shock are now totally destroyed. The effect of nonlinearity and curvature in hypersonic boundary-layer flow has also been considered [12]. The current study investigates the effects of a porous layer on the first Mack mode for axisymmetric and Presented as Paper 2010-4286 at the AIAA Fluid Dynamics Conference, Chicago, 28 June 20101 July 2010; received 11 May 2012; revision received 12 November 2012; accepted for publication 14 November 2012; published online 6 February 2013. Copyright © 2012 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved. Copies of this paper may be made for personal or internal use, on condition that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 1533-385X/13 and $10.00 in correspondence with the CCC. *[email protected]. [email protected]. 1234 AIAA JOURNAL Vol. 51, No. 5, May 2013 Downloaded by UNIVERSITY OF MINNESOTA on April 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.J052082
Transcript
Page 1: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

Effects of Porous Walls on Hypersonic Boundary Layersover a Sharp Cone

Sharon O. Stephen∗ and Vipin Michael†

University of Birmingham, Birmingham, England B15 2TT, United Kingdom

DOI: 10.2514/1.J052082

A theoretical linear-stability analysis is used to consider the effect of a passive porouswall on the firstMackmode of

a hypersonic boundary layer on a sharp slender cone. The effects of curvature and of the attached shock are included

for axisymmetric andnonaxisymmetric disturbances.The flow in the hypersonic boundary layer is coupled to the flow

in the porous layer. The theoretical models of the porous walls developed in previous studies are used for regular

microstructures and random microstructures. The physical parameters chosen correspond to those of their studies.

The resulting transcendental equations are solved numerically. Neutral solutions are presented, indicating a

destabilizing effect of the porouswall. The spatial growth rates determineddemonstrate that the porouswall leads to a

significant increase in growth rates for nonaxisymmetric modes. The effects of varying the porous-wall parameters

are investigated.

Nomenclature

A = displacement functionAy = admittanceAY = scaled admittancea = cone radius~a = half-pore widtha∞ = tortuosityCD = dynamic compressibility, CDγp−d = fiber diameterh = nondimensional porous-layer thickness, h∕δKn = Knudsen numberL = length scaleM = Mach numberP = pressure disturbancePr = Prandtl numberR = Reynolds number, U−δ

∕ν−Re = Reynolds number, U−L

∕ν−rp = nondimensional pore radius, rp∕δrs = shock locationT = temperaturetan h = hyperbolic tangent in Eqs. (2) and (7)U = velocityu, v = velocity disturbancex, r, ϕ = orthogonal coordinatesZ0 = characteristic impedanceα = disturbance wave number, αr iαiγ = specific-heat ratioδ = boundary-layer displacement thicknessθc = cone angleθs = shock angleΛ = propagation constantλ = skin frictionμ = viscosityν = kinematic viscosityρ = densityρD = dynamic density, ρD∕ρW

ϕ0 = porosityΩ = disturbance frequencyω = angular frequency

Subscripts

s = shockw = wall− = just behind the shock

I. Introduction

T RANSITION to turbulence in hypersonic flows is associatedwith amplification of the first and/or second Mack modes.

The first Mack mode is the high-speed counterpart of Tollmien–Schlichting waves, and so a viscous instability, with modes locatedclose to the boundary. The second Mack mode is an inviscidinstability. The second mode is believed to be responsible fortransition to turbulence on hypersonic slender bodies. These modescorrespond to the unstable perturbations occurring in the low-frequency band and the high-frequency band, respectively, of theslow (S) mode identified by Fedorov and Tumin [1]. Recentexperiments have shown that a porous coating greatly stabilizes thesecond Mack mode of the hypersonic boundary layer on sharpslender cones [2–4]. The effect of the porous coating is to reduce thegrowth rates of the secondmode to a level where they are comparablewith those of the first mode (occurring at lower frequencies). Inaddition, the first mode is observed to be slightly destabilized by thepresence of the porous coating. Thus, the firstmodemaynowbemoresignificant in the transition process.The destabilizing effect of these porous walls on the first Mack

mode has been confirmed by recent numerical simulations [5–10].This destabilization has been shown to be significant for a felt-metalporous coating [6].A previous work on the linear instability of the firstMackmodes in

the hypersonic flow over a cone includes the effect of curvature andalso the attached shock [11]. The effect of curvature is significantwith disturbances only existing over a finite range of wave numbersfor a fixed radius. Beyond a critical radius, dependent on theazimuthal wave number, all disturbances are damped. It wasdemonstrated that the effect of the shock, which allows incoming andoutgoingwaves, gives rise tomultiplemodes [11]. This study showedthat the effect of the shock should not be neglected; otherwise, thecorrect effect of curvaturewill not be realized. Themodes that exist inthe absence of the shock are now totally destroyed. The effect ofnonlinearity and curvature in hypersonic boundary-layer flow hasalso been considered [12]. The current study investigates the effectsof a porous layer on the first Mack mode for axisymmetric and

Presented as Paper 2010-4286 at the AIAA Fluid Dynamics Conference,Chicago, 28 June 2010–1 July 2010; received 11May 2012; revision received12 November 2012; accepted for publication 14 November 2012; publishedonline 6 February 2013. Copyright © 2012 by the American Institute ofAeronautics and Astronautics, Inc. All rights reserved. Copies of this papermay be made for personal or internal use, on condition that the copier pay the$10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 RosewoodDrive, Danvers, MA 01923; include the code 1533-385X/13 and $10.00 incorrespondence with the CCC.

*[email protected].†[email protected].

1234

AIAA JOURNALVol. 51, No. 5, May 2013

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 2: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

nonaxisymmetric disturbances. This is to confirm the experimentalobservations with a theory that includes the effect of curvature. Theeffect of curvature is of practical importance and has not beenpreviously investigated theoretically for a porous coating. Also, if thesecond mode (and higher modes) is greatly stabilized, then the firstmode may well lead to transition to turbulence, particularly if thedisturbances are triggered by wall roughness, because this mode islocated close to the wall and governed by viscous effects.We consider the effect of a porous wall on the linear instability of

hypersonic flow over a sharp slender cone. In this theoretical andasymptotic investigation for large Mach number and large Reynoldsnumber, the scales used are appropriate to the first Mack-modeinstability,which is governedby a triple-deck structure. The effects ofcurvature and the attached shock are taken into account. The currentstudy considers axisymmetric and nonaxisymmetric disturbances toinvestigate which modes will be more dominant. The effect of theporouswall changes the boundary condition on the normal velocity atthe interface.The porous model of Fedorov et al. for a regular microstructure

is considered [2]. This can be extended to consider a randommicrostructure [3] and to include gas-rarefaction effects in the porouslayer [4]. The regular microstructure comprising of layers of finemesh of square cross section is also considered, motivated by recentexperiments [13]. These cases are characterized by an admittanceAy,which is a function of the disturbance frequency and depends on thephysical properties of the flow and the porous layer. The physicalparameters are chosen to correspond to the relevant experimentalstudies [2–4].In Sec. II, the theoretical models of the porous walls to be

considered are described. The linear-stability problem is presentedfor axisymmetric and nonaxisymmetric disturbances, resulting in adispersion relation for each case. Section III considers neutral andnonneutral solutions of the dispersion relations for the differentporous-wall models. Maximum spatial growth rates are presentedand comparisons made with available experimental results. Theeffects of varying porosity, pore radius, and pore depth are consid-ered. Finally, maximum spatial growth rates for a porous surfaceoptimized for the secondMack-mode stabilization are comparedwiththose for a regular microstructure comprising circular pores. InSec. IV, we draw some conclusions.

II. Methods, Assumptions, and Procedures

The flow of a compressible viscous fluid over a sharp cone with aporous boundary, of semi-angle θc, is considered at hypersonic speedU0 aligned with its axis. The attached shock makes an angle θs withthe cone surface, a situation that is illustrated in Fig. 1. The dashedlines in this figure indicate the triple-deck structure of the disturbedflow. Spherical polars x; θ;ϕ are the natural coordinate system inwhich to describe the basic flow, and here ϕ denotes the azimuthalangle. Furthermore, the radial distance x has been nondimension-alized with respect to L, the distance from the tip of the cone to thelocation under consideration.

The basic flow has been described previously [11], and so only theimportant features are summarized as follows. Away from the surfaceof the cone, the flow satisfies the (inviscid) Euler equations. Thevelocities are nondimensionalized with respect to U−, in which U−is the magnitude of the fluid velocity just behind the shock.Additionally, the time, pressure, and density are nondimensionalizedwith respect toL∕U−, ρ−U

2−, and ρ−, respectively, inwhich ρ− is the

density just behind the shock. Finally, the basic temperature isnondimensionalized by T−, the temperature just behind the shock.The jump conditions at the shockmust be considered and the velocitycomponents satisfy the Taylor–Maccoll equations [14]. A constant-density approximation may be made to these equations in thehypersonic limit [15]. For the slender cone considered here, thisapproximation agrees well with the exact solution [16].From this hypersonic small-disturbance approximation to the

basic flow [16], an approximate value of the shock angle may beobtained from the expression

sinθs θc sin θc

γ 1

2 1

M2∞ sin2 θc

1∕2

(1)

in which γ is the specific-heat ratio, and M∞ denotes the Machnumber in the freestream. For the weak-shock solutions, relevantto the experiments, this gives excellent agreement with the exactsolution.Note that the termM∞ sin θc isO1. Thus, the shock anglemay be calculated, for a fixed cone angle and freestream Machnumber. A rough indication of the results from Eq. (1) correspondingto the experimental conditions of interest is that θs ≈ θc. This will betaken into account in the formulation of the linear-stability analysis.This solution is not valid close to the surface of the cone, and so a

boundary-layer solution has to be introduced in this region. TheReynolds number of the flow is defined by Re ρ−U−L

∕μ−.Taking the angle of the cone to be small, the governing equations inthe boundary-layer region are given in terms of dimensionlesscoordinates x; r;ϕ [11]. Then, L r is the normal direction tothe cone surface, in which r a on the generator of the cone.The corresponding nondimensional velocities are u; v; w, and thenondimensionalized pressure and density p and ρ, respectively.These satisfy the compressible conservation of mass, Navier–Stokes,and energy equations. The boundary conditions are no slip at thesurface of the cone (coupled to the porous layer) and appropriateconditions at the shock location. The nondimensional temperatureand viscosity at the surface of the cone are taken to be Tw and μw,respectively. The only restriction imposed on the temperatureboundary condition is Tw ≫ 1, which is violated only for situationsinvolving strong cooling on the cone wall [11]. Usually, the walltemperature is taken to be Tw TbTr, in which Tr is the adiabaticwall temperature given by

Tr 1Prp γ − 1

2M2

in which Pr is the Prandtl number and M is the Mach number justbehind the shock. Thus, unless the constant Tb is very small, Tw willbe of Oγ − 1M2 for both adiabatic walls (Tb 1) or isothermalwalls. The analysis is unaffected by the particular choice oftemperature–viscosity law. The choice only affects the bounds placedon various parameters of the problem. Sutherland’s viscosity law[μw ∼ 1 CT1∕2

w , C being a constant] is used henceforth.

A. Porous Boundary

We will present the results corresponding to porous surfaces usedin the previous experimental investigations [3,13,17]. In all cases, theporous-layer admittance Ay can then be expressed in the form

Ay −ϕ0∕Z0 tan hΛh (2)

in which ϕ0 is the porosity of the surface. The porous-layerparameters are nondimensionalized with respect to the boundary-layer displacement thickness δ, and so the nondimensional pore

Fig. 1 Geometry of the cone and shock. The cone is taken to be of semi-

angle θc with the attached shock making an angle θs with the surface of

the cone.

STEPHEN AND MICHAEL 1235

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 3: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

depth h h∕δ and the nondimensional pore radius rp rp∕δ.Z0 and Λ are the characteristic impedance and propagation constantof an isolated pore, respectively. Fedorov et al. [4] give the followingexpressions for the porous-layer characteristics:

Z0 ρD∕CD

pM

Twp and Λ iωM

Twp

ρDCD

p(3)

in which ω is the disturbance frequency. These are functions of thecomplex dynamic density ρD and complex dynamic compressibilityCD. The precise definitions of these quantities depend on thestructure of the porous wall, and are given in Secs. II.A.1–3 for thecases investigated here.The wall boundary condition, in all cases, is then given by

v Ayp − p− (4)

in which v is the nondimensional normal-velocity componentand p− γ−1M−2.

1. Regular Microstructure

Following previous investigations [2,4], we consider the porouslayer on the cone surface to be a sheet of thickness h perforatedwith cylindrical blind holes of radius rp and equal spacings rp

π∕ϕ0

p. This model takes into account gas-rarefaction

effects. We have

ρD 11−FBν;ζ ; CD 1 γ − 1FBE; ζ

Prp

FBν; ζ Gζ1−0.5Bνζ

2Gζ ; FBE; ζPrp Gζ

Prp

1−0.5BEζPrp2Gζ

Prp

9=;(5)

in which

Bν 2α−1ν − 1Kn; BE γ2α−1E − 1∕γ 1PrKn;

Gζ 2J1ζζJ0ζ

(6)

and ζ rpiωρwR∕μw

p. Here, R is the Reynolds number based on

the boundary-layer displacement thickness, J0;1 are Bessel functionsof the first kind, αν and αE are molecular accommodation coeffi-cients, and Kn is the Knudsen number.

2. Mesh Microstructure

Following Lukashevich et al. [13], we consider the porous coatingon the cone surface to comprise of several layers of stainless-steelwire mesh. A similar model to the one described previously for aregular microstructure is employed. Following Kozlov et al. [18], wehave different expressions for the complex dynamic density andcompressibility. The expressions for the porous-layer characteristicsfor a square-mesh microstructure are [18]

ρD 1∕1 − Fζ; CD 1 γ − 1F~ζFζ 1 ζ2

P∞m0

2

γ2mβ2m

1 − tan hβm

βm

F~ζ 1 ~ζ2P∞

m0

2

γ2m ~βm2

1 − tan h ~βm

~βm

9>>>>>=>>>>>;

(7)

in which

γm π

m 1

2

; βm

γ2m − ζ2

q; ~βm

γ2m − ~ζ2

q(8)

The characteristic size of an isolated pore is given by

ζ

iωρw ~a

2

μwR

sand ~ζ

Prp

ζ (9)

Here, ~a is the half-pore width, and the flow parameters are chosen tofit the experimental conditions [13]. Following Lukashevich et al.[13], rarefaction effects are neglected for this porous model.

3. Random Microstructure

Following Fedorov et al. [3], we consider the porous layer on thecone surface to have a random microstructure. This was consideredbecause it closely resembles the surface of thermal protection tilesused to protect reentry vehicles from aerodynamic heating. A similarmodel to the one used for the regular microstructure is employed.Wehave different expressions for the complex dynamic density andcompressibility. Fedorov et al. [3] give the following expressionsfor the porous-layer characteristics for flow over a felt-metalmicrostructure:

ρD a∞h1 gλ1

λ1

i; CD γ − γ−1

1gλ2 λ2

gλi 1 4a∞μwλi

σϕ0r2p

q; λ1 ia∞ρwω

ϕ0σ ; λ2 4Prλ1

9=;(10)

where the characteristic pore size is

rp πd

21 − ϕ02 − ϕ0(11)

Here, d is the fiber diameter; σ is the flow resistivity, and its value ischosen to fit the experimental data for flow over the felt metal. Thetortuosity a∞ is taken to be unity. Following Fedorov et al. [3],rarefaction effects are neglected for this porous model.

B. Linear-Stability Problem

The linear stability of the basic flow described previously for aslender cone with M ≫ 1 and Re≫ 1 is investigated in the weak-interaction region following a triple-deck formulation [19,20]. This iswith respect to the interaction parameter M∞R

−1∕6, in whichM∞R

−1∕6 ≪ 1 [21,22]. This ensures that we are considering alocation far from the tip of the cone to ensure that theviscous–inviscidinteraction between the boundary layer and the inviscid flow is small.The conditions to be satisfied at the shock by a disturbance to thisbasic flow must be specified, and these have been derived in detail[23]. The requisite constraints were obtained by considering thelinearized jump conditions at the shock for infinitesimal wavesbeneath the shock; a similar procedure was adopted for flow over awedge [19]. Although the basic flow is not uniform in the regionbelow the shock, the jump conditions may still be evaluated at theundisturbed position of the shock [23]. The condition satisfied by thepressure amplitudes of the two acoustic waves (which are incidentand reflected from the shock) is found to be similar to that for awedge [19].Attention is focused at a location on the surface of the cone

with nondimensional radius a a∕L. It is assumed thataRe3∕8M1∕4μ−3∕8w T

−9∕8w a ∼O1 denotes the scale of the

radius at this point; thus, we have chosen sin θc ∼ θc∼Re−3∕8M−1∕4μ3∕8w T

9∕8w . We note that, in terms of the interaction

parameter χ M∞a∕L, our analysis corresponds to a moderate

inviscid interactionwith χ ∼O1. Because fromEq. (1), θs ≈ θc, theshock is located in the upper deck of the triple-deck structure. Theangle of the cone and the flow parameters enable the shock angle andthe corresponding value of a to be obtained. Then, if rs denotes thescaled location of the shock, the ratio a∕rs may be obtained simplyfrom geometric arguments. We find for a slender cone that

a

rs≈

sin θctan θs sin θc

(12)

1236 STEPHEN AND MICHAEL

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 4: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

where we have taken cos θc ≈ 1. Because θs ≈ θc, a roughapproximation gives a∕rs ≈ 0.5.Our study is confined to the question of the stability of the flow at a

location on the body where the boundary-layer thickness isORe−1∕2L , which is thin compared to the local radius of the cone,allowing the subsequent analysis to capture the effects of curvature onthe stability problem. For smaller values of radius a, the problemreduces to that of the planar case (flow over a wedge). The analysis issomewhat simplified if nonparallel effects can be neglected, and thisis justifiable if the Newtonian assumption γ − 1 ≪ 1 is made [19].This condition arises from ensuring that the streamwise wavelengthof the disturbances is much less than the distance from the apex of thecone. Thus, for simplicity, this condition is taken to hold in thefollowing analysis, although it can be easily relaxed for moreinvolved studies. This condition leads to restrictions on the lowerbounds of M [11]. The experimental conditions for the poroussurfaces of interest have been shown to be in the range of validity ofour asymptotic analysis [24].It is convenient to scale out some of the parameters in the problem,

namely μw, Tw, and λ, in which the last quantity denotes theboundary-layer skin friction. For axisymmetric flow, the Machnumbermay be scaled out of the linear-stability problem.A summaryof the triple-deck analysis for axisymmetric and nonaxisymmetricdisturbances is given in Michael and Stephen [24].Previous scaling is applied to the resulting linear-disturbance

equations for axisymmetric disturbances [19,25]. Analytic solutionsof these equations yield an eigenrelation relating the streamwisewave number α and frequency Ω of the disturbance, namely

Ai 0ξ0R∞ξ0

Aiξ dξ

−iα1∕3AY iαI0iαrsK0iαa − I0iαaK0iαrsI0iαrsK1iαa I1iαaK0iαrs

(13)

Here, ξ0 −i1∕3Ωα−2∕3, Aiξ is the Airy function, Knz andInz are the usual modified Bessel functions, and Ay Re−1∕8μ1∕8w λ1∕4T3∕8

w M2 − 13∕8AY . The angular frequency ofdisturbance propagation through the pore is ω R∕ReRe1∕4μ−1∕4w λ3∕2T−3∕4

w M2 − 11∕4Ω. The eigenrelation cor-responding to a solid wall is recovered if AY 0.A similar analysis is applied for nonaxisymmetric disturbances

with azimuthal wave number n. The scaling is different in terms ofMach number [11], and in particular for the wall admittance anddisturbance frequency Ay Re−1∕8μ1∕8w λ1∕4T3∕8

w M5∕4AY andω R∕ReRe1∕4μ−1∕4w λ3∕2T−3∕4

w M−1∕2Ω. The resulting eigenrela-tion for nonaxisymmetric disturbances is given by

Ai 0ξ0R∞ξ0

Aiξ dξ

iα1∕3AY

in2

αa2

IniαrsKniαa − IniαaKniαrsIniαrsK 0n iαa − I 0n iαaKniαrs

(14)

III. Results and Discussion

To present our results in regimes of practical interest, we will usethe flow parameters from the relevant experimental studies [2,3,13].The cone angle and Mach number from the experiments willdetermine the shock angle θs and the scaled radius a. Then, the scaledshock location rs is determined as described in Eq. (12). Theexperimental conditions correspond to a∕rs 0.57; thus, we chooseto present our results for this case.The values for flow of a perfect gas were chosen as M 5.3,

Tw Tad, γ 1.4, Pr 0.71, Re1 15.2 × 106, T− 56.4 K,and αν αE 0.9. These values correspond to the ones used inprevious numerical studies [17]. The experiments were conducted ona cone of 0.5 m in length.

A. Neutral Modes

We now consider neutrally stable solutions of the dispersionrelations (13) and (14) corresponding to the axisymmetric case andthe nonaxisymmetric case, respectively. The presence of the shockallows for multiple modes of solutions.

1. Regular Microstructure

The current results for the regular microstructure comprising aregular array of cylindrical pores of circular cross section arecompared to the results for a solid wall for axisymmetric andnonaxisymmetric modes. For the regular microstructure [4], theresults presented as follows are for the porous-layer parametersrp 28.5 μm,ϕ0 0.2, and h ≫ rp. The last relation implies thatΛh → ∞, and so Eq. (2) simplifies to Ay −ϕ0∕Z0. The effectof pore depth is investigated in Sec. III.C. The porous-layercharacteristics chosen correspond to experimental values [4]. Theflow conditions match the experimental conditions [13]. Here, a coneof angle θc 7 degwas usedwith a flow ofMach numberM∞ 6.The effects of rarefaction are included and the Knudsen number iscalculated from the expression

Kn μwM

rpR

2πγTw

p

giving Kn 0.494.Figure 2 shows the neutral values of frequency Ω as a function of

radius a for the first four axisymmetric modes for a∕rs 0.57. Weidentify increasing mode numbers with higher neutral values ofΩ, asindicated by the arrow in Fig. 2 and subsequent figures. The scaledradius varies linearly with L. The range 0 < a < 5 corresponds to0 < L < 0.02 m for axisymmetric modes for all the wall modelsconsidered. The dashed lines are for the porous wall and the solidlines correspond to the results for a solid wall [11]. The flow isunstable in regions above these curves. For the higher modes for thesolutions for α, there is no discernible difference between the results(not shown). However, the neutral solutions for Ω are lower for aporouswall than those for a solidwall, particularly for larger values ofa. Thus, the flow over the porous surface will become unstable forlower frequencies than those for the solid wall.Figure 3 shows the neutral solutions for Ω as a function of a for

nonaxisymmetric disturbances for a∕rs 0.57 and n 1 and n 2 for a solid wall and a regular microstructure. For n 1, the range0 < a < 5 corresponds to 0 < L < 14 m. Again, the porous wall hasonly a small effect on the neutral values of α (not shown). Thebehavior of the first mode is different from the higher modes withΩ → 0 as a → 0. We see from Fig. 3 that the porous wall has a verysmall effect on the neutral values of Ω for this first mode. However,the porous wall has a significant effect on the higher modes, leadingto lower neutral values of Ω. Thus, the effect of the porous wall is

0

10

20

30

40

50

60

0 0.5 1 1.5 2 2.5a

3 3.5 4 4.5 5

Ω

Fig. 2 Neutral values of frequency Ω for the first four axisymmetric

modes as a function of radiusawithn 0 anda∕rs 0.57: – – –, regularporous wall; —— , solid wall.

STEPHEN AND MICHAEL 1237

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 5: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

destabilizing. Moreover, this effect increases for the higher modes.Similar results have been obtained for n 3 (not shown).

2. Mesh Microstructure and Random Microstructure

The neutral results for the felt-metal microstructure are comparedto those for ameshmicrostructure. FollowingLukashevich et al. [13],we consider the mesh microstructure to correspond to ϕ0 0.8and ~a 0.05 mm. In their experiments, the porosity of the meshmicrostructure is larger than that of the regular array of circular pores.

This will lead to a larger destabilizing effect on the first Mack mode(see Sec. III.B). We show our results for h ≫ ~a.Figure 4 shows the neutral values of Ω as a function of a for the

mesh microstructure and the felt-metal microstructure for the firstfour axisymmetric modes. We find that the felt metal has a largerdestabilizing effect than the mesh microstructure, resulting in lowervalues of neutral frequency.Figure 5 shows the neutral solutions for Ω as a function of a for

nonaxisymmetric disturbances forn 1 anda∕rs 0.57 for ameshmicrostructure and a random microstructure for the first five modes.The neutral values of Ω for the mesh microstructure are lower thanthose for the regular microstructure comprising circular pores as aresult of the porosity being larger.We see from Fig. 5 that the randommicrostructure has a significant destabilizing effect, with the neutralvalues much lower than those for the mesh microstructure.

B. Spatial Growth Rates

We now concern ourselves with an examination of the spatialevolution of disturbances so that we concentrate on solutions ofEqs. (13) and (14) with Ω real and α complex. If α αr iαi, thenαi > 0 is indicative of stability, whereas αi < 0 denotes spatialinstability.Figures 6–8 show the dependence of the spatial growth-rate

parameter αi on the mode frequency Ω for a regular microstructurecompared to a solid wall for n 0, n 1, and n 2, respectively.The results are shown for a∕rs 0.57 and a 0.6, 1.0, 1.5, and 2.0.There is a complete family of modes, as we saw in our account ofneutral disturbances, and it is clear that for eachmember of the family,there is a cutoff frequency Ωc such that for Ω < Ωc that particularmode is stable, but it becomes unstable if Ω > Ωc. The effect of theneutral values being altered by the presence of the porous wall for allwall models leads to a variation in the growth rates. The maximumvalues of (−αi) for the porouswall are increased greatly from those ofa solid wall for each mode.In addition, we also observed that the growth rate for the porous

modes does not rapidly approach zero at high frequencies, as is thecase for the solidwall. Figure 6 shows, that for the porouswall and thesolid wall, the first mode has the largest growth rate for axisymmetricdisturbances for smaller values of a. However, for larger values of a,the second mode has the largest maximum growth rate. As aincreases, the growth rates decrease.From Fig. 7, we see that for nonaxisymmetric modes with n 1,

the growth rate of the highermodes is greatly enhanced by the regularporous wall. It is the last mode shown, which has the highest growthrate. The results for αi presented in Fig. 8 for n 2 are similar tothose for n 1.Figure 9 shows the results for αi for the felt-metal microstructure

with ϕ0 0.75. These are compared to the values for a meshmicrostructure with ϕ0 0.8. The results are presented for n 0;

0

10

20

30

40

50

60

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Ω

a

a) n = 1

0

10

20

30

40

50

60

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Ω

a

b) n = 2

Fig. 3 Neutral values of frequencyΩ for the first five nonaxisymmetric

modes as a function of radius a with a∕rs 0.57: – – –, regular porous

wall; —— , solid wall.

0

10

20

30

40

50

60

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Ω

a

Fig. 4 Neutral values of frequency Ω for the first four axisymmetric

modes as a function of radius a with n 0 and a∕rs 0.57: – – –, mesh

microstructure; —— , random microstructure.

0

5

10

15

20

25

30

35

40

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Ω

aFig. 5 Neutral values of frequencyΩ for the first five nonaxisymmetric

modes as a function of radius a with n 1 and a∕rs 0.57: – – –, mesh

microstructure; —— , random microstructure.

1238 STEPHEN AND MICHAEL

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 6: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0 10 20 30 40 50 60 70 80 90 100

α i

Ωa) a = 0.6 b) a = 1.0

c) a = 1.5 d) a = 2.0

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-0.25

-0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

0.2

0.25

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

0.2

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

Fig. 6 Spatial growth-rate parameters αiΩ for the first four nonneutral axisymmetric modes for a∕rs 0.57: – – –, regular porous wall; —— , solid

wall.

-1

-0.5

0

0.5

1

1.5

2

2.5

3

3.5

0 10 20 30 40 50 60

α i

Ω

a) a = 0.6

-0.5

0

0.5

1

1.5

2

0 10 20 30 40 50

α i

Ω

b) a = 1.0

c) a = 1.5 d) a = 2.0

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 10 20 30 40 50

α i

Ω-0.1

-0.05

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0 5 10 15 20 25 30 35 40

α i

Ω

Fig. 7 Spatial growth-rate parameters αiΩ for the first five nonneutral nonaxisymmetricmodes fora∕rs 0.57 andn 1: – – –, regular porouswall;—— , solid wall.

STEPHEN AND MICHAEL 1239

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 7: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

0 10 20 30 40 50 60

α i

Ω

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

0 10 20 30 40 50 60

α i

Ω

-0.2

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 10 20 30 40 50

α i

Ω

-0.2

0

0.2

0.4

0.6

0.8

1

0 5 10 15 20 25 30

α i

Ω

a) a = 0.6 b) a = 1.0

c) a = 1.5 d) a = 2.0

Fig. 8 Spatial growth-rate parameters αiΩ for the first five nonneutral nonaxisymmetricmodes fora∕rs 0.57 andn 2: – – –, regular porouswall;—— , solid wall.

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

a) a = 0.6 b) a = 1.0

c) a = 1.5 d) a = 2.0Fig. 9 Spatial growth-rate parameters αiΩ for the first four nonneutral axisymmetric modes for a∕rs 0.57: – – –, mesh microstructure; —— ,

random microstructure.

1240 STEPHEN AND MICHAEL

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 8: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

a∕rs 0.57; and a 0.6, 1.0, 1.5, and 2.0. We find that the growthrates of all modes are increased greatly for the felt-metalmicrostructure compared to the mesh microstructure. For the felt-metal microstructure, the first mode has the largest growth rate,switching to the secondmode as the radius a increases. This occurs ata larger value of a than for the mesh microstructure. These resultsagree with numerical results for a random microstructure [26].Figure 10 shows the spatial growth rates for n 1 for the mesh

microstructure and the felt-metal microstructure. The effect of thefelt-metal microstructure is to give a very large increase in themaximum growth rates, approximately four times in magnitude,compared to the mesh microstructure.We end this discussion on the effects of a porouswall on the spatial

growth rates by considering the effects of rarefaction for a regularmicrostructure comprising an array of circular pores. In Fig. 11, weshow the neutral values of Ω as a function of a for the first fivenonaxisymmetric modes with n 1 and a∕rs 0.57 for Kn 0andKn 0.494. We see that the effect of nonzero Knudson numberis destabilizing, with lower values of Ω compared to Kn 0,particularly for the higher modes. The effect of rarefaction on thespatial growth rates is illustrated in Fig. 12, showing the resultscorresponding to Fig. 11. The growth rates are increased for nonzeroKnudson number, with larger increases for the higher modes. Wenote, that for the Mack second mode, the effect of rarefaction isstabilizing [4]. However, rarefaction has been shown to give largerfirst-mode destabilization and less second-mode stabilization forhypersonic flow over a flat plate [26].

C. Maximum Spatial Growth Rates

We investigate the effects of changing the properties of the porouswall on the linear stability of hypersonic flow over a sharp cone. Wechoose to concern ourselves with an examination of the effect of thephysical properties of the porous layer on the maximum spatialevolution of disturbances. Thus, we consider the modewhich has the

largest value of (−αi). The effects of porosity, pore radius, and poredepth on the maximum spatial growth rates σmax are considered forazimuthal wave number n 1, a∕rs 0.57, and a 0.8.The effect of increasing the porosity ϕ0 (by decreasing the pore

spacing) for fixed pore radius rp 30 μm is illustrated in Fig. 13.We see, that as the porosity increases, the maximum spatial growthrate σmax increases. Here, σmax is the maximum value of −αi for thedifferent modes for particular values of a∕rs.In Fig. 14,we show the effect of the pore radius rp on themaximum

spatial growth rate σmax for ϕ0 0.25. We see that a larger poreradius leads to larger maximum spatial growth rates.The previous results presented have been obtained by assuming an

infinite pore depth. The effect of finite pore depth is demonstrated inFig. 15 for porosity ϕ0 0.25 and rp 30 μm. We see that as h∕rp

-4

-3

-2

-1

0

1

2

3

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-1.5

-1

-0.5

0

0.5

1

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

-1.2

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0 10 20 30 40 50 60 70 80 90 100

α i

Ω

a) a = 0.6 b) a = 1.0

c) a = 1.5 d) a = 2.0Fig. 10 Spatial growth-rate parameters αiΩ for the first five nonneutral nonaxisymmetric modes for a∕rs 0.57 and n 1: – – –, mesh

microstructure; —— , random microstructure.

0

5

10

15

20

25

30

35

40

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Ω

aFig. 11 Neutral values of frequencyΩ for nonaxisymmetric modes as a

function of radius a for a regular microstructure with n 1 and

a∕rs 0.57: – – –, Kn 0.494; —— , Kn 0.

STEPHEN AND MICHAEL 1241

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 9: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

increases, themaximum spatial growth rate increases, but it levels offforh∕rp > 20. In the experiments, the perforated sheet was relativelythick with h 0.45–0.5 mm and rp 25–30 μm, givingh∕rp 15–20. This verifies our earlier assumption of takingh∕rp ≫ 1 in our calculations. Previous calculations determined thatmaximum stabilization of the second Mack mode was achieved forh∕rp 3 [13].

In Fig. 16, we show the effect of pore depth for a meshmicrostructure. An increase in the depth corresponds to more layersof the fine mesh. We find that for h 0.15 mm (corresponding tothree layers ofmesh), themaximumgrowth rate is amaximum for thefirst Mack mode. Previous results [13] showed that h 0.15 mmgave themaximum stabilization of the secondMackmode. So, this isthe worst case for the first Mack mode.In Fig. 17, we present the maximum spatial growth rates to

correspond to the experimental conditions of Maslov [17] fora regular microstructure with rp 25 μm, ϕ0 0.2, andRe 10 × 106. Here, the dimensional spatial growth rate is shownas a function of the distance along the cone for the mode giving thelargest growth rate for n 0, 1, 2, 3 and a∕rs 0.57. The dashedlines correspond to the porouswall and the solid lines correspond to asolid wall. We see, that for the nonaxisymmetric modes, the growthrates are significantly larger for the porous wall compared to the solidwall. However, the size of the growth rates is much smaller than thoseobtained for the second Mack mode [17].Previous parametric studies of regular porous coatings and mesh

coatings have been carried out with the focus on the stabilization ofMack’s second-mode instability [2,13,27,28]. These studies revealthat the porous-layer performance can be optimized by controllingthe porosity and porous-layer thickness. These parametric studiesindicate that optimal porous coatings have thickness h∕rp ≈ 3–3.5.Our results indicate that the porous-layer thickness in this range forregular porous coatings also provides optimal first Mack-modestabilization. Parametric studies also show that high porosityprovides maximum second Mack-mode stabilization. However,numerical studies [29] reveal that porous coatings with too closely

-1

-0.5

0

0.5

1

1.5

2

2.5

0 5 10 15 20 25 30 35 40 45 50

α i

ΩFig. 12 Spatial growth-rate parameters αiΩ for the first five

nonneutral nonaxisymmetric modes for a regular microstructure for

a∕rs 0.57, n 1, and a 0.6: – – –, Kn 0.494; —— , Kn 0.

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

σ max

, mm

-1

Φ0

Fig. 13 Maximum spatial growth rates σmax for varying porosityϕ0 for

a regular microstructure for n 1, a∕rs 0.57, rp 30 μm, and

a 0.8.

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0.045

0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0.0008 0.0009 0.001

σ max

, mm

-1

rp*, m

Fig. 14 Maximum spatial growth rates σmax for varying pore radius rp

for a regular microstructure for n 1, a∕rs 0.57, ϕ0 0.25, anda 0.8.

0.0015

0.002

0.0025

0.003

0.0035

0.004

0.0045

0.005

2 4 6 8 10 12 14 16 18 20

σ max

, mm

-1

h/rp

Fig. 15 Maximum spatial growth rates σmax for varying pore-depth

ratio h∕rp for a regular microstructure for n 1, a∕rs 0.57,ϕ0 0.25, rp 30 μm, and a 0.8.

0.002

0.0025

0.003

0.0035

0.004

0.0045

0.005

0.05 0.1 0.15 0.2 0.25 0.3 0.35

σ max

, mm

-1

h*, mm

Fig. 16 Maximum spatial growth rates σmax for varying pore depth h

for n 1, a∕rs 0.57, ϕ0 0.25, and a 0.8 for a mesh micro-

structure for ~a 0.05 mm.

1242 STEPHEN AND MICHAEL

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 10: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

spaced pores trigger a new shorter wavelength instability whosegrowth rate can be larger than that of the second Mack mode. Theauthors have attempted to optimize the design of porous coatingsbased on the acoustic-scattering properties of the porous layer. Theypropose a porous coating with fixed low porosity comprising ofspanwise grooves. Each porous cavity has a depth h, half-width b,and spacing that varies along the longitudinal length of the cone. Theregular porous model of Eq. (5) can be used to study this model bymaking the following changes [29]:

ζ biωρwμw

R

s; FB; ζ tanζ

ζ1 − Bζ tanζ

The effect of this new design on the first Mack-mode instability isexamined. In Fig. 18, maximum unstable growth rates of the firstazimuthal mode n 1 are compared using this porousmodel and theregular porous model both with porosity ϕ0 0.2. The regularporous model is assumed to be infinitely thick, and the pore radius isfixed at rp 25 μm. FromFig. 18,we see that, at smaller streamwisedistance, the new design leads to lower amplification of unstabledisturbances and with increasing streamwise distance the differencebetween the growth rates of the twomodels becomes very small. Thisnovel design corresponds to porous coatings with low porosity andlarge cavity aspect ratio (2b∕h) (i.e., thinner coatingswith less pores).These types of coatings are easier tomanufacture and incorporate intothermal protection systems in hypersonic vehicles [29]. Recent

studies have investigated the optimum shape, size, and spacing for theporous holes to obtain maximum reduction in growth rates for thesecond mode [30].

D. Nonlinear Effects

The linear-stability analysis presented will not be valid forlarger disturbances. Thus, it is important to determine the effect ofnonlinearity on the stability of hypersonic boundary-layer flowover asharp slender cone with a porous wall. This has been investigatedexperimentally for second-mode disturbances using the bicoherencemethod [31,32]. We have carried out a lengthy theoretical weaklynonlinear analysis for the first Mackmode for hypersonic flow over aporous surface [24]. The results are presented in [24], and show thatthe effect of nonlinearity is greatly affected by the presence of aporous wall.

E. Wall Admittance

Because it has been observed experimentally and confirmedtheoretically and numerically that a porous microstructuredestabilizes the first Mack mode, it would be desirable to design amicrostructure that stabilizes the second Mack mode while notappreciably destabilizing the first Mack mode. In their efforts todetermine the types of microstructure that will lead to reducedamplification of the first Mack mode, Wang and Zhong investigatedthe effect of the phase angle of the wall admittance on the growth ofthe first Mack mode [8]. For their numerical investigations on a flatplate, they discovered that a smaller phase angle of the walladmittance leads to a weaker destabilization of the first Mack mode.The effect of wall admittance was also investigated theoretically byCarpenter and Porter [33]. They considered a thin porous sheetstretched over a plenum chamber. When the phase of the walladmittance was very close to π∕2, they discovered that Tollmien–Schlichting waves were completely stabilized. Thus, future studiesshould be focused on determining the wall parameters that willminimize the destabilizing effect of a porous wall on the first Mackmode. This will enhance the current knowledge of how the firstMackmode in hypersonic boundary layers may be controlled.

IV. Conclusions

The authors have presented neutral results for the wave numberand frequency of linear disturbances for hypersonic flow over apassive porous wall, with scales appropriate to the first instabilitymode for axisymmetric and nonaxisymmetric disturbances. Theporous-wall parameters were chosen to correspond to previoustheoretical and numerical studies. For the values chosen here, theauthors find that the neutral values of streamwise wave number αare slightly reduced from those corresponding to a solid wall.However, the neutral values of the disturbance frequency Ω fornonaxisymmetric disturbances can be substantially lower than thosefor a solid wall, particularly for larger values of radius a. In addition,the spatial growth rates presented demonstrate that the porous wallhas a destabilizing effect on the nonaxisymmetricmodes. The growthrates are significantly larger than those for the solid wall.The effect of rarefaction for a regular microstructure was shown to

be destabilizing.The authors have also investigated the effect of the porous-wall

parameters on the maximum spatial growth rate. The effect ofporosity, pore radius, and pore depth on the maximum spatial growthrates was also considered. It was shown that increasing each of thesequantities leads to larger growth rates, but they level off.The authors have made some comparisons of the dimensional

growth rates for the first-mode disturbances enhanced by the poroussurface with those obtained from previous numerical results for thesecond mode. It was found that the first-mode growth rates weresmaller than the second-mode ones.The authors have determined the effect of regular microstructures,

randommicrostructures, and mesh microstructures on the first Mackmode in a hypersonic boundary layer. The formulation will allowfor alternative porous walls to be investigated in a straightforward

0

0.002

0.004

0.006

0.008

0.01

0.012

0.1 0.2 0.3 0.4 0.5

σ max

, mm

-1

L*,m

n=1n=2n=3n=0

Fig. 17 Maximum spatial growth rates σmax as a function of L for aregular microstructure for n 0, 1, 2, 3; a∕rs 0.57; ϕ0 0.2; andrp 25 μm: – – –, regular porous wall; —— , solid wall.

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

0.011

0.1 0.2 0.3 0.4 0.5

σ max

, mm

-1

L*,m

Fig. 18 Maximum spatial growth rates σmax as a function of L for

a∕rs 0.57, ϕ0 0.2, and n 1: – – –, regular microstructure for

rp 25 μm; —— , spanwise grooves.

STEPHEN AND MICHAEL 1243

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82

Page 11: Effects of Porous Walls on Hypersonic Boundary Layers over a Sharp Cone

manner. Investigations to determine the thickness of porous coatingstominimise the growth of the firstMackmodes have been carried out.A porous coating comprising spanwise grooves was also considered.

Acknowledgments

Thisworkwas sponsored by theU.S. Air ForceOffice of ScientificResearch, Air ForceMateriel Command, U.S. Air Force, under grantnumber FA8655-08-1-3044. The U.S. government is authorizedto reproduce and distribute reprints for governmental purposesnotwithstanding and copyright notation thereon. Vipin Michaelacknowledges additional financial support from the School ofMathematics, University of Birmingham for his Ph.D. studies.Discussions with A. Ruban are gratefully acknowledged.

References

[1] Fedorov, A. V., and Tumin, A., “Branching of Discrete Modes inHigh-Speed Boundary Layers and Terminology Issues,” AIAA Paper2010-5003, July 2010.

[2] Fedorov, A. V., Malmuth, N. D., Rasheed, A., and Hornung, H. G.,“Stabilization of Hypersonic Boundary Layers by Porous Coatings,”AIAA Journal, Vol. 39, No. 4, 2001, pp. 605–610.doi:10.2514/2.1382

[3] Fedorov, A., Shiplyuk, A., Maslov, A., Burov, E., and Malmuth, N.,“Stabilization of a Hypersonic Boundary Layer Using an UltrasonicallyAbsorptive Coating,” Journal of Fluid Mechanics, Vol. 479,March 2003, pp. 99–124.doi:10.1017/S0022112002003440

[4] Fedorov, A., Kozlov, V., Shiplyuk, A., Maslov, A., and Malmuth, N.,“Stability of Hypersonic Boundary Layer on Porous Wall with RegularMicrostructure,” AIAA Journal, Vol. 44, No. 8, 2006, pp. 1866–1871.doi:10.2514/1.21013

[5] Wang, X., and Zhong, X., “Numerical Simulation and TheoreticalAnalysis on Boundary-Layer Instability Affected by Porous Coating,”AIAA Paper 2009-3679, June 2009.

[6] Wang, X., and Zhong, X., “The Stabilization of a Hypersonic BoundaryLayer Using Local Sections of Porous Coating,” Physics of Fluids,Vol. 24, No. 3, 2012, p. 034105.doi:10.1063/1.3694808

[7] Wang, X., and Zhong, X., “Numerical Simulations on Mode S Growthover Feltmetal and Regular Porous Coatings of a Mach 5.92 Flow,”AIAA Paper 2011-375, Jan. 2011.

[8] Wang, X., and Zhong, X., “Phase Angle of Porous Coating Admittanceand Its Effect on Boundary-Layer Stabilization,” AIAA Paper 2011-3080, June 2011.

[9] De Tullio, N., and Sandham, N. D., “Direct Numerical Simulation ofBreakdown to Turbulence in a Mach 6 Boundary Layer over a PorousSurface,” Physics of Fluids, Vol. 22, No. 9, 2010, p. 094105.

[10] Egorov, I. V., Fedorov, A. V., and Soudakov, V. G., “Receptivity of aHypersonic Boundary Layer over a Flat Plate with a Porous Coating,”Journal of Fluid Mechanics, Vol. 601, April 2008, pp. 165–187.doi:10.1017/S0022112008000669

[11] Seddougui, S. O., and Bassom, A. P., “Instability of HypersonicFlow Over a Cone,” Journal of Fluid Mechanics, Vol. 345, Aug. 1997,pp. 383–411.doi:10.1017/S0022112097006423

[12] Stephen, S.O., “Nonlinear Instability ofHypersonic FlowOver aCone,”Quarterly Journal ofMechanics&AppliedMathematics, Vol. 59,No. 2,2006, pp. 301–319.doi:10.1093/qjmam/hbl003

[13] Lukashevich, S. V., Maslov, A. A., Shiplyuk, A. N., Fedorov, A. V., andSoudakov, V. G., “Stabilization of High-Speed Boundary Layer UsingPorous Coatings of Various Thicknesses,” AIAA Paper 2010-4720,July 2010.

[14] Taylor, G., andMaccoll, J., “TheAir Pressure on aConeMoving atHighSpeeds. I,” Proceedings of the Royal Society of London, Series A:

Mathematical and Physical Sciences, Vol. 139, No. 838, 1933,pp. 278–297.

[15] Hayes, W. D., and Probstein, R. F., Hypersonic Flow Theory, 2nd ed.,Academic Press, New York, 1966, pp. 143–144.

[16] Rasmussen, M., Hypersonic Flow, Wiley, New York, 1994, pp. 68–78.[17] Maslov, A. A., “Experimental and Theoretical Studies of Hypersonic

Laminar Flow Control Using Ultrasonically Absorptive Coatings(UAC),” Rept. ISTC-2172-2001, May 2003.

[18] Kozlov, V. F., Fedorov, A. V., and Malmuth, N., “Acoustic Propertiesof Rarefied Gases Inside Pores of Simple Geometries,” Journal of theAcoustical Society of America, Vol. 117, No. 6, 2005, pp. 3402–3412.doi:10.1121/1.1893428

[19] Cowley, S. J., and Hall, P., “On the Instability of the HypersonicFlow Past a Wedge,” Journal of Fluid Mechanics, Vol. 214, May 1990,pp. 17–42.doi:10.1017/S0022112090000027

[20] Duck, P. W., and Hall, P., “Non-Axisymmetric Viscous Lower-BranchModes in Axisymmetric Supersonic Flows,” Journal of Fluid

Mechanics, Vol. 213, April 1990, pp. 191–201.doi:10.1017/S0022112090002282

[21] Stewartson, K., The Theory of Laminar Boundary Layers in

Compressible Fluids, Oxford Univ. Press, London, 1964, p. 173.[22] Brown, S. N., Khorrami, A. M., Neish, A., and Smith, F. T., “On

Hypersonic Boundary-Layer Interactions and Transition,” Philosophi-cal Transactions of the Royal Society of London, Series A:

Mathematical and Physical Sciences, Vol. 335, No. 1637, 1991,pp. 139–152.doi:10.1098/rsta.1991.0040

[23] Seddougui, S. O., “Stability of Hypersonic Flow over a Cone,”Transition, Turbulence and Combustion, edited by Hussaini, M. Y.,Gatski, T. B., and Jackson, T. L., Kluwer, Dordrecht, The Netherlands,1994, pp. 50–59.

[24] Michael, V., and Stephen, S. O., “Nonlinear Stability of HypersonicFlow over a Cone with Passive Porous Walls,” Journal of Fluid

Mechanics, Vol. 713, Dec. 2012, pp. 528–563.doi:10.1017/jfm.2012.472, 2012

[25] Duck, P. W., and Hall, P., “On the Interaction of Tollmien–SchlichtingWaves in Axisymmetric Supersonic Flows,” Quarterly Journal of

Mechanics and Applied Mathematics, Vol. 42, Feb. 1989, pp. 115–130.doi:10.1093/qjmam/42.1.115

[26] Wang,X., andZhong,X., “The Impact of Porous Surface onHypersonicBoundary Layer Instability,” AIAA Paper 2010-5021, July 2010.

[27] Fedorov, A. V., andMalmuth, N. D., “Parametric Studies of HypersonicLaminar Flow Control Using a Porous Coating of Regular Micro-structure,” AIAA Paper 2008-588, Jan. 2008.

[28] Fedorov,A.V., Brés,G.A., Inkman,M., andColonius, T., “Instability ofHypersonic Boundary Layer on a Wall with Resonating Micro-Cavities,” AIAA Paper 2011-373, Jan. 2011.

[29] Brés, G. A., Inkman, M., and Colonius, T., “Alternate Designs ofUltrasonic Absorptive Coatings for Hypersonic Boundary LayerControl,” AIAA Paper 2009-4217, June 2009.

[30] Sandham, N. D., and Lüdeke, H., “A Numerical Study of Mach 6Boundary Layer Stabilization by Means of a Porous Surface,” AIAAPaper 2009-1288, Jan. 2009.

[31] Chokani, N., Bountin, D. A., Shiplyuk, N., and Maslov, A. A.,“Nonlinear Aspects of Hypersonic Boundary-Layer Stability on aPorous Surface,” AIAA Journal, Vol. 43, No. 1, 2005, pp. 149–155.doi:10.2514/1.9547

[32] Bountin, D., Shiplyuk, A., and Maslov, A., “Evolution of NonlinearProcesses in aHypersonicBoundaryLayer on a SharpCone,” Journal ofFluid Mechanics, Vol. 611, Sept. 2008, pp. 427–442.doi:10.1017/S0022112008003030

[33] Carpenter, P. W., and Porter, L. J., “Effects of Passive Porous Walls onBoundary-Layer Instability,” AIAA Journal, Vol. 39, No. 4, 2001,pp. 597–604.doi:10.2514/2.1381

X. ZhongAssociate Editor

1244 STEPHEN AND MICHAEL

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

INN

ESO

TA

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

82


Recommended