+ All Categories
Home > Documents > Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

Date post: 08-Dec-2016
Category:
Upload: sergey
View: 214 times
Download: 0 times
Share this document with a friend
8
Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface Dmitry Bountin, Timur Chimitov, and Anatoly Maslov Institute of Theoretical and Applied Mechanics, Novosibirsk 630090, Russia Andrey Novikov § and Ivan Egorov Central Aerohydrodynamic Institute (TsAGI), Zhukovsky 140180, Moscow Region, Russia Alexander Fedorov ** Moscow Institute of Physics and Technology, Zhukovsky 140180, Moscow Region, Russia and Sergey Utyuzhnikov †† University of Manchester, Manchester, England M13 9PL, United Kingdom DOI: 10.2514/1.J052044 Stability of a supersonic near-wall flow over a shallow grooved plate in the freestream of Mach 6 is investigated by means of numerical simulations and wind-tunnel experiments. Numerical solutions of two-dimensional NavierStokes equations are used to model propagation of artificial disturbances of several fixed frequencies generated by an actuator placed on the wall. It is shown that the high-frequency forcing excites unstable waves in the flat-plate boundary layer. These waves are relevant to the second-mode instability. The wavy wall damps the disturbances in a high-frequency band while it enhances them at lower frequencies. Stability experiments are conducted in the Institute of Theoretical and Applied Mechanics Tranzit-M shock tunnel under natural freestream conditions. The measured disturbance spectra are similar to those predicted numerically. They contain a peak associated with the second-mode instability. This peak is damped by the wavy wall, while a marginal increase of the disturbance amplitude is observed at lower frequencies. Although the experiments qualitatively confirm the wavy-wall stabilization concept, further stability and transition measurements are needed to clarify its robustness. Nomenclature A = spectral density for disturbance c p = pressure coefficient f = frequency M = Mach number p = pressure T = temperature t = time Re 1 = unit Reynolds number U = streamwise velocity u; v = flow velocity component x = streamwise coordinate measured from the plate leading edge y = vertical coordinate measured from the flat-plate surface γ = specific heat ratio Δ = difference between values from disturbed and steady fields δ = boundary-layer thickness μ = dynamic viscosity ρ = density Subscripts 0 = total w = on the wall = freestream I. Introduction L AMINARTURBULENT transition leads to substantial increase of the aerodynamic drag and surface heating, and it reduces the efficiency of propulsion systems of hypersonic vehicles [1,2]. Smoothing and shaping of the vehicle surface helps to avoid early transition due to roughness, leading-edge contamination, and cross-flow and Görtler instabilities. However, with these measures, the laminar run may still be short because of the amplification of unstable disturbances of the first and/or second mode during the linear phase of the transition process. The wall cooling, which naturally occurs on hypersonic-vehicle surfaces, strongly stabilizes the first mode, while it destabilizes the second mode. In this case, transition can be second-mode-dominated, and laminar flow-control (LFC) concepts should address the second-mode instability. The following are the categories of LFC techniques [3]: 1) passive techniques such as shaping and passive coatings; 2) active techniques such as suction, local cooling, or heating; and 3) reactive techniques such as actuators and microelectromechanical systems. Because of severe environmental conditions of a hypersonic flight associated with large heat fluxes and high temperatures of the boundary-layer flow, it is difficult to use the active and reactive techniques. The Presented as Paper 2012-1105 at the 50th AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition, Nashville, Tennessee, 912 January 2012; received 23 April 2012; revision received 11 October 2012; accepted for publication 5 November 2012; published online 1 March 2013. Copyright © 2012 by Andrey V. Novikov. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. 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. *Researcher, Laboratory for Physical Problems of Hypersonic Flows Control; [email protected]. Researcher, Laboratory for Physical Problems of Hypersonic Flows Control; [email protected]. Deputy Director; [email protected]. Member AIAA. § Researcher, Aerothermodynamics Department; also Associate Professor, Department of Aeromechanics and Flight Engineering, Moscow Institute of Physics and Technology, Zhukovsky 140180; [email protected]. Head of Department, Aerothermodynamics Department; currently Professor, Department of Aeromechanics and Flight Engineering, Moscow Institute of Physics and Technology, Zhukovsky 140180; ivan.egorov@tsagi. ru. **Associate Professor, Department of Aeromechanics and Flight Engineering; [email protected]. Associate Fellow AIAA. †† Senior Research Fellow, School of Mechanical, Aerospace and Civil Engineering; also Head of Laboratory, Laboratory for Mathematical Modeling of Nonlinear Processes in Gas Media, Moscow Institute of Physics and Technology, Dolgoprudny 141700; [email protected]. 1203 AIAA JOURNAL Vol. 51, No. 5, May 2013 Downloaded by NEW YORK UNIVERSITY on April 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.J052044
Transcript
Page 1: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

Stabilization of a Hypersonic Boundary LayerUsing a Wavy Surface

Dmitry Bountin,∗ Timur Chimitov,† and Anatoly Maslov‡

Institute of Theoretical and Applied Mechanics, Novosibirsk 630090, Russia

Andrey Novikov§ and Ivan Egorov¶

Central Aerohydrodynamic Institute (TsAGI), Zhukovsky 140180, Moscow Region, Russia

Alexander Fedorov**

Moscow Institute of Physics and Technology, Zhukovsky 140180, Moscow Region, Russia

and

Sergey Utyuzhnikov††

University of Manchester, Manchester, England M13 9PL, United Kingdom

DOI: 10.2514/1.J052044

Stability of a supersonic near-wall flow over a shallow grooved plate in the freestream ofMach 6 is investigated by

means of numerical simulations and wind-tunnel experiments. Numerical solutions of two-dimensional Navier–

Stokes equations are used tomodel propagation of artificial disturbances of several fixed frequencies generated by an

actuator placed on the wall. It is shown that the high-frequency forcing excites unstable waves in the flat-plate

boundary layer. These waves are relevant to the second-mode instability. The wavy wall damps the disturbances in a

high-frequency bandwhile it enhances them at lower frequencies. Stability experiments are conducted in the Institute

of Theoretical and Applied Mechanics Tranzit-M shock tunnel under natural freestream conditions. The measured

disturbance spectra are similar to those predicted numerically. They contain a peak associatedwith the second-mode

instability. This peak is damped by the wavy wall, while amarginal increase of the disturbance amplitude is observed

at lower frequencies. Although the experiments qualitatively confirm the wavy-wall stabilization concept, further

stability and transition measurements are needed to clarify its robustness.

Nomenclature

A = spectral density for disturbancecp = pressure coefficientf = frequencyM = Mach numberp = pressureT = temperaturet = timeRe1 = unit Reynolds numberU = streamwise velocity�u; v� = flow velocity component

x = streamwise coordinate measured from the plate leadingedge

y = vertical coordinate measured from the flat-plate surfaceγ = specific heat ratioΔ = difference between values from disturbed and steady

fieldsδ = boundary-layer thicknessμ = dynamic viscosityρ = density

Subscripts

0 = totalw = on the wall∞ = freestream

I. Introduction

L AMINAR–TURBULENT transition leads to substantialincrease of the aerodynamic drag and surface heating, and it

reduces the efficiency of propulsion systems of hypersonic vehicles[1,2]. Smoothing and shaping of the vehicle surface helps to avoidearly transition due to roughness, leading-edge contamination, andcross-flow and Görtler instabilities. However, with these measures,the laminar run may still be short because of the amplification ofunstable disturbances of the first and/or second mode during thelinear phase of the transition process. The wall cooling, whichnaturally occurs on hypersonic-vehicle surfaces, strongly stabilizesthe first mode, while it destabilizes the second mode. In this case,transition can be second-mode-dominated, and laminar flow-control(LFC) concepts should address the second-mode instability.The following are the categories of LFC techniques [3]: 1) passive

techniques such as shaping and passive coatings; 2) active techniquessuch as suction, local cooling, or heating; and 3) reactive techniquessuch as actuators and microelectromechanical systems. Because ofsevere environmental conditions of a hypersonic flight associatedwith large heat fluxes and high temperatures of the boundary-layerflow, it is difficult to use the active and reactive techniques. The

Presented as Paper 2012-1105 at the 50th AIAA Aerospace SciencesMeeting including the New Horizons Forum and Aerospace Exposition,Nashville, Tennessee, 9–12 January 2012; received 23 April 2012; revisionreceived 11 October 2012; accepted for publication 5 November 2012;published online 1 March 2013. Copyright © 2012 by Andrey V. Novikov.Published by the American Institute of Aeronautics and Astronautics, Inc.,with permission. Copies of this paper may be made for personal or internaluse, on condition that the copier pay the $10.00 per-copy fee to the CopyrightClearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; includethe code 1533-385X/13 and $10.00 in correspondence with the CCC.

*Researcher, Laboratory for Physical Problems of Hypersonic FlowsControl; [email protected].

†Researcher, Laboratory for Physical Problems of Hypersonic FlowsControl; [email protected].

‡Deputy Director; [email protected]. Member AIAA.§Researcher, Aerothermodynamics Department; also Associate Professor,

Department of Aeromechanics and Flight Engineering, Moscow Institute ofPhysics and Technology, Zhukovsky 140180; [email protected].

¶Head of Department, Aerothermodynamics Department; currentlyProfessor, Department of Aeromechanics and Flight Engineering, MoscowInstitute of Physics and Technology, Zhukovsky 140180; [email protected].

**Associate Professor, Department of Aeromechanics and FlightEngineering; [email protected]. Associate Fellow AIAA.

††Senior Research Fellow, School of Mechanical, Aerospace and CivilEngineering; also Head of Laboratory, Laboratory for MathematicalModeling of Nonlinear Processes in Gas Media, Moscow Institute of Physicsand Technology, Dolgoprudny 141700; [email protected].

1203

AIAA JOURNALVol. 51, No. 5, May 2013

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 2: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

passive techniques are of primary interest for hypersonic LFCstudies.In connection with this, Fedorov et al. [4] showed that a porous

coating providing absorption of disturbance energy can massivelysuppress the second-mode instability. These theoretical results wereconfirmed by the experiments of Rasheed et al. [5], conducted on asharp cone in the Graduate Aerospace Laboratories of the CaliforniaInstitute of Technology T5 high-enthalpy shock tunnel as well as bydirect numerical simulations (DNS) [6]. A review of studies related tothis LFC concept is given in [7].The second-mode instability can also be affected by local shaping

of the body surface. It is well known that increasing Mach numberproduces a stabilization effect on the flow in free shear layers andwakes (e.g., [8,9]). It is natural to assume that a relatively long freeshear layer formed near a streamlined surface may decrease thegrowth rates and damp the second-mode disturbances with a shortwavelength (of the order of the layer thickness). This hypothesis wasmotivated by the numerical studies [10,11] addressing stability of thesupersonic (freestream Mach number 5.373) flow over a 5.5 degcompression corner. It was shown that the second mode growsexponentially in the regions lying upstream and downstream from theseparation bubble, while it remains neutral across the separatedregion. Similar, although less pronounced, stabilization effect ispresent in the numerical simulations [12] of two-dimensional (2-D)disturbances in a Mach 4.8 flow over a 6 deg compression corner.In [11], it is also shown that the second-mode amplitudes decrease

in a separated mixing layer. This is consistent with the numericalinvestigation [13] of the disturbance evolution in a Mach 4.8 flat-plate boundary layer with a localized 2-D roughness element. It wasfound that the disturbance was strongly damped behind theroughness element, where the separation flow mimics the flowstructure over the compression corner considered in [11].It should be noted that acoustic disturbances are effectively excited

in a relatively long separation bubble [11]. They, in turn, generate thesecond-mode waves, which grow rapidly downstream from thereattachment point. It is assumed that, if a long separation bubble isreplaced by a sequence of small ones, it is feasible to exploit theaforementioned stabilization effect of the free shear layer and, at thesame time, avoid detrimental acoustic resonances within the bubbles.This could be achieved with the help of a shallow, grooved, wavysurface, which produces a relatively stable free shear layer bridgingneighboring cavities. This wavy-wall stabilization (WWS) conceptwas confirmed by our 2-D DNS [14]. It was shown that the second-mode waves of high frequency are stabilized by the concave wavywall comprising several shallow cavities of a half-sinus shape. Asexpected, the stabilization effect is not observed for low-frequencydisturbances with the length scale of the order thewaviness period. Itwas also found that the wavy surface weakly affects the acousticcomponent of disturbances in the boundary layer (i.e., the wavinessdoes not produce detrimental effects associated with secondaryreflections of acoustic waves in the separation regions). It waspointed out that naturally occurring wavy surfaces (such as bowed

panels of thermal protection system [15]) do not necessarily lead tomore unstable flow and premature transition. These configurationsneed special treatments of the interaction between instability andmean-flow irregularities.Note that the aforementioned 2-D numerical simulations do

not capture Görtler vortices and other three-dimensional (3-D)disturbances that could grow on concave surfaces and lead topremature transition. The main objective of this paper is to validatethe WWS concept by experiments. Experimental and 2-D DNSresults are presented for unsteady hypersonic flow over a wavy plateat the freestream Mach number 6. The study addresses the second-mode instability of the near-wall flow comprising a set of localseparation bubbles. In accord with our previous DNS results [14] aswell as additional parametric computations, a wavy-plate model wasdesigned and manufactured. The model was tested in the Mach 6shock tunnel with measurements of the natural disturbance spectra.For the freestream parameters related to these experiments, the DNSwas conducted for 2-D disturbances generated by a local forcing(periodic suction–blowing). The DNS results are compared with theexperimental data.

II. Problem Formulation

A. Experimental Setup

The experiments are carried out in a Ludwieg-type short-durationwind tunnel, the Tranzit-M of the Institute of Theoretical andAppliedMechanics, Siberian Branch of the Russian Academy of Sciences, atthe freestream Mach number M∞ � 6. The wind-tunnel layout isshown in Fig. 1. Air is heated up by ohmic heaters and is gathered inthe plenum chamber under pressure up to 200 bar. After opening thefast-acting valve, air flows to the settling chamber and to the testsection through the contoured nozzle of 300mm exit diameter. Then,gas is compressed by a diffuser and flows to thevacuum tank. The runtime is about 300–350 ms, depending on the initial pressure. Duringthe run, the total pressure p0 and the stagnation temperature T0 aremeasured in the plenum chamber with the relative error of 1.5% forp0 and 0.9% for T0.Unfortunately, there are no data on the freestream noise in this

tunnel at M∞ � 6. Measurements of the p 00 pressure pulsations(behind a normal shock) have been conducted at M∞ � 4 in thefrequency band from 1 to 200 kHz. The rms value of these pulsationsis 4.0� 0.2% at the stagnation pressures 4 and 8 bar and stagnationtemperatures 290 and 450K.Although the Tranzit-Mwind tunnel is aconventional noisy facility, the freestream noise is relatively small inthe high-frequency band related to the second-mode instability andits higher harmonics.The experimental model is shown in Fig. 2. It is a stainless steel

plate with a nominally sharp leading edge. The plate has a length of350 mm, width of 200 mm, and thickness of 15 mm. The side andleading edges are beveled from belowwith the angle of 30 deg for theside edges and 20 deg for the leading edge. The plate is polished to thenominal surface finish 0.8 micron. The region of grooved wavysurface comprises nine round arc cavities. Deviations of the actual arc

Fig. 1 Layout of the Tranzit-Mwind tunnel: electric heater (1), fast-acting valve (2), plenum chamber (3), settling chamber (4), frame (5), nozzle (6), testsection (7), optical windows (8), isolated pedestal for model installation (9), diffuser (10), and vacuum tank (11). Dimensions are given in millimeters.

1204 BOUNTIN ETAL.

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 3: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

radii from their nominal value are less than 0.2%. Its surface shape,with dimensions, is shown in Fig. 3. The spanwise length of thegrooved region is approximately one-half of the plate span. Themodel has several technological holes for the sensors: 12 holes of8 mm diameter for the atomic-layer thermopile (ALTP) sensors [16]and two holes of 3.2 mm diameter for the pressure sensors (Fig. 2).Disturbances in the wavy surface region are measured by the

high-frequency integrated circuit-piezoelectric (ICP) pressure sensors(PCB Piezotronics Inc., model 132A31) with the relative error of 1%.These sensors have a diameter of 3 mm and a frequency band from11 kHz to 1 MHz. They are small in size, and they weakly influencethe surface shape. Downstream from thewavy region, the fluctuationsare measured by the fast-response heat-flux ALTP sensors withrelative error of 10%. During each run, the measurements on the flatand grooved regions of the model are performed simultaneously.

B. Numerical Problem

The computations are carried out at the freestream conditionscorresponding to the shock tunnel run 579, which represents theWWS effect in a full manner. These conditions are: Mach numberM∞ � 6.0, unit Reynolds number Re1∞ � 10.5 × 106 m−1 (stag-nation pressurep0 � 7.0 × 105 Pa), and temperatureT∞ � 43.18 K(stagnation temperatureT0 � 354.06 K). Thewall is isothermalwithtemperature Tw � 293 K.The Navier–Stokes partial-differential equations for 2-D viscous

compressible unsteady flows are solved numerically. We use thedimensionless conservative form of these equations written incurvilinear coordinates. Curvilinear coordinate system acts as com-putational space, where a grid is formed uniformly in all directions.The fluid is assumed to be a perfect gas with specific heat ratio

γ � 1.4 and Prandtl number Pr � 0.72 (air). The dynamic viscosityμ is calculated using Sutherland’s formula μ � μ∞ �T3∕2� �S� 1�∕� �S� �T�, where �T � T∕T∞ is the dimensionless temperature, and�S � 110 �K�∕T∞. The second viscosity is assumed to be zero.The Navier–Stokes equations are integrated using the in-house

solver HSFlow, which implements an implicit finite-volume shock-capturing method with the second-order approximation in space andtime.AGodunov-type scheme [17]with aRoe approximateRiemann

solver is used. Reconstruction of dependent variables at the grid cellboundaries is performed using the weighted essentially non-oscillatory (WENO) approach. The system of nonlinear algebraicequations (which approximate partial differential equations) is solvedusing the Newton iteration method. At every iteration step, thecorresponding linear algebraic system is solved using the generalizedminimal residual (GMRes) method.This approach is most efficient if the computational domain

contains shock waves and other strong spatial inhomogeneities of theflow such as boundary-layer separations. Using the HSFlow solver, itwas feasible to perform numerical simulations of boundary-layerreceptivity [18] and stability [17], including configurations withseparation bubbles [11], as well as the laminar flow control usingporous coatings [19]. Note that these simulations were carried outwith the help of the total variation diminishing scheme, which israther dissipative. To reduce the dissipative effects and improveaccuracy, the WENO scheme is used herein.The computations are carried out for the flow over a grooved wavy

plate with nine cavities. The surface shape corresponds to the modeland is shown in Fig. 3. Namely, each cavity has a form of round arcgiven by the following formula:

y�x� �������������������������������������������R2 − �x − x0 − l∕2�2

q− R

R � h2� l2

8h; xs < x < xe

where l � 12 mm, and h � 1.8 mm. Dimensions for the first cavityare x0 � 46 mm, xs � 52 mm, and xe � 58 mm; for the secondone, x0 � xs � 58 mm, xe � 70 mm, and so on. The depth h �1.8 mm is approximately equal to the boundary-layer thicknessδ (for the flat-plate case δ ≈ 2.1 mm at x � 100 mm). Preliminaryparametric studies showed that the cavities of these sizes providegentle reattachment and separation of the shear layer on the top ofeach bump.The length of the computational domain is 200 mm, which is less

than the experimental model length (350mm) to save grid nodes. Thecomputations are performed on a structured curved orthogonal gridwith 3001 × 401 nodes in a single block. Themesh is generated using

M∞ = 6

ALTP #0 ICP

ALTP #1 ALTP #2 ALTP #5

Fig. 2 The plate model with wavy surface and locations of ICP and ALTP sensors.

52 mm 1.8

12 mm

ICP ALTP#1

148 mm

165 mm

ALTP#2

20 mm

ALTP#5

4 * 20 mm = 80 mm

Fig. 3 Scheme of the wavy plate. Triangles denote the sensors’ locations.

BOUNTIN ETAL. 1205

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 4: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

numerical conformal mapping of a rectangle onto the computationaldomain. The mapping method is described in [20] and is imple-mented in the third-party open-source tool. It generates a grid that fitswell to the surface shape and has low cell skewness (Fig. 4). The gridis clustered near the surface so that approximately 200 grid y lines arewithin the boundary layer or in the separation region with the mixinglayer.Several computations were also performed on 6001 × 401 and

3001 × 801 grids to verify convergence. The discrepancy in dis-turbance amplitudes was less than 0.97%.The numerical problem is solved in two steps. First, a steady

laminar flowfield is computed using a time-dependent method. Then,unsteady disturbances are introduced by a local periodic suction–blowing on the plate surface. This ismade via the boundary conditionfor the mass-flow perturbation:

�qw�x; t� � ε sin

�2π

x − x1x2 − x1

�sin�2πft�; x1 ≤ x ≤ x2

where �qw � ρwvw∕�ρ∞U∞� is dimensionless mass-flow in thenormal wall direction (ρ is density, v is vertical velocity component),and x1 � 10 mm and x2 � 15 mm are boundaries of the suction–blowing region.The calculations are conducted at 17 forcing frequencies from

f � 88.28 kHz to f � 176.58 kHz that correspond to the frequencyparameters F � 2π �f∕ReL∞ from 0.6667 × 10−4 to 1.333 × 10−4,where �f � f × L∕U∞ is dimensionless frequency. This range wasidentified using the experimental spectra. To ensure the linearevolution of excited disturbances in front of the wavy region, wechose a small forcing amplitude ε � 10−3.

III. Results

A. Experimental Results

Figure 5 shows the pulsation spectra measured by the ICP sensorson thewavy and flat surfaces. On the flat surface (black lines), a peakwith central frequency f ≈ 170 kHz is related to the second-modeinstability. On thewavy surface (gray lines), this peak is shifted to thelower central frequency f ≈ 135 kHz and has essentially a smaller

level, indicating that the wavy wall stabilizes the second-modewaves. Figures 6 and 7 show the spectra of wall heat-fluxfluctuations, which were measured using the ALTP sensors locateddownstream from the wavy region (see Fig. 2). These spectra alsoclearly show damping of the second-mode waves.The ICP spectrum (Fig. 5) has a new peak around the frequency

f ≈ 75 kHz.With available experimental data, we can only speculateon the physical nature of this peak. Thewavy surface induces a steadyperturbation, which leads to modulation of the flow disturbancespropagating in the near-wall region. With the assumption that thedisturbance speed is close to the speed Ue of the outer mean flow,this modulation should give a spectral peak at the frequencyf ≈Ue∕λ ≈ 78 kHz, where λ is the wavelength of the wavy surface.This estimate correlates well with the experimental data. Such amodulation should vanish downstream from thewavy region. Indeed,Fig. 7 shows that the ALTP 1 spectrum just behind the wavy surface(solid blue line) has a low-frequencypeak ofmuch smaller amplitude.Further downstream (see the ALTP 2 spectrum shown by the dashedblue line), the peak is not observed. This allows us to assume that the

Fig. 4 Fragment of the computational grid; the bold black curve is the model shape.

Fig. 5 Spectra of pressure pulsations on the model wall; disturbancesare measured using a ICP sensor in the wavy region: Re1∞ �15.41 × 106 m−1 (run 583: p0 � 10.188 × 105 Pa, T0 � 352.35 K)and Re1∞ � 12.67 × 106 m−1 (run 584: p0 � 8.3902 × 105 Pa,T0 � 352.70 K).

Fig. 6 Spectra of heat-flux fluctuations on the wall measured by ALTPsensors: high freestream Reynolds numbers Re1∞ � 12.25 × 106 m−1

(run 581: p0 � 8.3547 × 105 Pa, T0 � 358.73 K) and Re1∞ � 12.16 ×106 m−1 (run 505: p0 � 8.5124 × 105 Pa, T0 � 364.19 K).

Fig. 7 Spectra of heat-flux fluctuations on the wall measured by ALTP

sensors: low freestream Reynolds numbers Re1∞ � 10.50 × 106 m−1

(run 579: p0 � 7.0 × 105 Pa, T0 � 354.06 K) and Re1∞ � 10.14 ×106 m−1 (run 507: p0 � 7.0338 × 105 Pa, T0 � 362.27 K).

1206 BOUNTIN ETAL.

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 5: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

low-frequency peak is associated with the modulation of disturbancefield, and it does not represent a new instability.In the locationALTP 5 (far downstream from thewavy region), the

spectra on both halves of the model resemble turbulent spectra.Unfortunately, these data do not provide information of the transitionlocation. Additional experiments are needed to clarify the wavy-walleffect on the transition locations.

B. Numerical Results

The computed steady-flow Mach-number field over a groovedplate is shown in Fig. 8. The viscous-inviscid interaction leads toformation of a weak shock wave in the leading-edge vicinity. Furtherdownstream, the cavities induce oblique shocklets emanating fromthe reattachment regions on the bump tops. Figure 9 shows flowrecirculation inside the cavities (dark regions) as well as a weaklyperturbed shear layer (light gray strip), which smoothly bridges theneighboring cavities.The upper boundary of each separation region is almost a straight

line, which is typical for supersonic flows. As a result, the upper

boundary of the whole mixing layer over the wavy plate remainsalmost unchanged compared to the flat-plate case.The wave (pressure) drag coefficient on the wavy wall is

cDp � 4.04 × 10−4, which is greater than zero for the flat wall.However, the wavy wall leads to the reduction of friction drag fromcDf � 11.60 × 10−4 to cDf � 9.53 × 10−4, and the total drag isrelatively small (cDp � 1.97 × 10−4). The heat flux to the groovedwavy surface is close to that in the flat-plate case. As shown inFig. 10, the cavities cause relatively small periodic heat-fluxperturbations without average overheating.Thus, thewavy surfaceweakly affects the global flowfield over the

boundary layer. At the same time, this surface produces a mixinglayer that bridges the cavities and resembles a free shear layer withalmost parallel edges.Analysis of the numerical unsteady solution is done using the

instantaneous disturbance fields. They are obtained as the differencebetween a disturbed flowfield at some instant and a steady-state field(the initial field before introduced forcing), i.e., the disturbance of acertain value φ�t; x; y� is calculated as Δφ�t; x; y� � φ�t; x; y�−φ�0; x; y�.At first, the disturbance fields were computed for the flat plate at

different forcing frequencies from f � 88.3 to 176.6 kHz. It wasfound that the suction–blowing actuator predominantly excitesacoustic waves in the shock-layer region and the second-modewavesin the boundary layer, e.g., Fig. 11 shows the instantaneousdisturbance field of the pressure coefficient cp � �p − p∞�∕�ρ∞U2

∞∕2� at the forcing frequency 138.74 kHz. The cell-typestructures in the outer flow behind the actuator correspond to theinterference between the rear and fore fronts of acoustic waves. Therear fronts are related to the slow acoustic waves and the fore fronts tothe fast acoustic waves. For x > 80 mm, the disturbance in theboundary layer corresponds to the second-mode wave. This wavepropagates downstream with the phase speed, which is slightlysmaller than the mean-flow velocity at the upper boundary-layeredge; its amplitude grows downstream, its wavelength is ≈2δ, andit induces two-cell structures of the pressure-disturbance field.

Fig. 8 Mach-number field of steady flow.

Fig. 9 Normalized horizontal velocity field of steady flow inside the wall cavities.

Fig. 10 Stanton number distribution along the plate surface. Thevertical lines marked “1” show the boundaries of the wavy region.

Fig. 11 Disturbance field of pressure coefficient at t � 0.315 ms for forcing of f � 138.74 kHz:Re1∞ � 10.50 × 106 m−1, flat surface. Triangles showlocations of ICP, ALTP 1, and ALTP 2 sensors (Fig. 2).

BOUNTIN ETAL. 1207

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 6: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

The region of its instability agrees with that predicted by the linearstability theory for the second mode [21].In the case of thewavy plate (Fig. 12), the acoustic disturbances in

the shock layer (above the mixing layer and behind the bow shock)behave similar to the flat-plate case, because the upper boundary ofthe mixing layer is weakly affected by the separation bubbles withinthe cavities. In the near-wall region, evolution of disturbances iscompletely different. The disturbances interact with the shockletsinduced by the surface bumps andwith the cavity flows.Although thefine structure of the disturbance field is rather complicated, there is noevidence of local amplification of the disturbance amplitude, whichcould be caused by these interactions.Figure 13 shows the streamwise distributions of the pressure

coefficient disturbance on the wall. In the flat-plate case (black line),the second mode amplifies starting from x ≈ 80 mm and reaches itsmaximum amplitude at x ≈ 150 mm. In the wavy-plate case (grayline), the disturbance growth is reduced. Thewall-pressure amplitudeis modulated versus xwith the period of wall waviness. As shown inFig. 14, this damping is not instantaneous but persists with time.By processing the time dependencies of the wall-pressure

disturbances (such as those shown in Fig. 14) for every forcingfrequency, we obtained asymptotic values for the disturbance

amplitudes in the x station related to the ALTP 1 coordinate. Theseamplitudes form the spectra max Δpw � fn�f� shown in the upperpanel of Fig. 15. For comparison, the lower panel shows the spectrathat weremeasured using theALTP 1 sensor located just downstreamfrom the wavy region.Both DNS and the experiment show reduction of the maximal

amplitude that qualitatively confirms the WWS concept.Nevertheless, there are essential differences. In the experiment, thewavy wall shifts the second-mode peak to lower frequencies andcauses a massive reduction of high-frequency disturbances. In theDNS, the peak is not shifted, and high-frequency disturbancesare weakly affected. The relative reduction of the peak maximum isabout 50% in the experiment, whereas it is about 35% in DNS.Furthermore, the peak central frequency predicted by DNS(f ≈ 129 kHz for a flat plate) is lower than in the experiment(f ≈ 143 kHz). Presumably, these discrepancies are due to thedifferences in the external forcing and receptivity mechanisms. In theDNS, the disturbances of all frequencies are excited by the 2-Dsuction–blowing actuator, which is located at a fixed x station and hasthe same shape and amplitude. In the experiment, the second-modewaves are predominantly excited by 3-D acoustic disturbances,which have nonuniform distributions of their amplitude versusfrequency and wave-front angles. Furthermore, receptivity toacoustic disturbancesmay be nonuniformversus x, which also affectsthe shape and locus of the second-mode peak observed in the finalx-station.It should be noted that the damping effect was not observed in the

case of the rounded compression corner [11], where the separationbubblewasmuch longer. This can be explained by comparisons of thepressure disturbance fields inside the separation bubble. In thecompression corner, a waveguide is formed within a long separationzone, and acoustic waves of appreciable amplitudes are excited viathe resonance mechanism. These acoustic waves effectively generatethe boundary-layer disturbances growing downstream from thereattachment point. For the configuration considered herein, theseparation bubbles are relatively short, which prevents the resonantexcitation of acoustic disturbances within cavities. Furthermore, theboundary-layer reattachments on the groove tops are gentle andrelatively short, which also prevents the intensive second-modegrowth in these regions.

Fig. 12 Disturbance field of pressure coefficient at t � 0.315 ms for forcing off � 138.74 kHz:Re1∞ � 10.50 × 106 m−1, wavy surface.Triangles showlocations of ICP, ALTP 1, and ALTP 2 sensors (Fig. 2).

Fig. 13 Disturbances of the calculated pressure coefficient alongthe wall at t � 0.315 ms for forcing of f � 138.74 kHz:Re1∞ � 10.50 × 106 m−1; boundaries of the wavy region (1), andlocations of the sensors ICP, ALTP 1, and ALTP 2 (2).

Fig. 14 Time dependency of the calculated disturbance amplitudes of pressure coefficient on the wall at x � 165 mm (ALTP 1 sensor location) forforcing of f � 126.13 kHz (left) and of f � 151.35 kHz (right): Re1∞ � 10.50 × 106 m−1.

1208 BOUNTIN ETAL.

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 7: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

Using the DNS disturbance fields, we evaluated thewavelength ofdisturbances propagating in the near-wall layer over the groovedsurface. It turned out that the maximal reduction of disturbanceamplitude corresponds to about two disturbance wavelengths perwavelength of the wavy wall. This suggests that there could be aninterference (cancellation) process in the grooves that disrupts thesecond-mode instability. However, the experimental data shown inFig. 15, where the stabilization effect is observed in awide frequencyband from 100 to 200 kHz, do not support this assumption.Furthermore, the counterphase cancellation requires a certainrelationship between the phase of unstable waves and the phase ofwaviness. In the wind-tunnel experiments, the initial phase ofinstability has random distributions, and the probability of phasecancellation should be equal to the probability of phase reinforce-ment. With these arguments, it is natural to assume that thestabilization effect is due to altering the mean flow rather than theinterference process.

IV. Conclusions

The stability of a supersonic near-wall flow over a plate with awavy surface has been investigated by means of numericalsimulations and wind-tunnel experiments for the freestream Machnumber 6.Parametric numerical studies showed that it is feasible to design a

wavy surface that maintains an almost parallel mixing layer, bridgingthe neighboring cavitieswith gentle separations and reattachments onthe groove tops. Such a wavy surface slightly perturbs the outer flowwhile producing dramatic changes of the near-wall flow. Because thefree shear layer is more stable than the boundary layer, the wavy wallmay reduce the disturbance growth and, ultimately, increase thelaminar run. This concept was partially confirmed by two-dimensional numerical simulations. It was shown that the second-mode instability is damped by thewavy surface in the frequency bandfrom 110 to 150 kHz, while there is increase of the disturbanceamplitude for frequencies less than 110 kHz.The flat-plate model, with the grooved region comprising nine

shallow cavities, has been designed and tested in the conventionalshock tunnel Tranzit-M of the Institute of Theoretical and AppliedMechanics. Measurements of the wall-pressure and wall heat-fluxfluctuation spectrawere carried out on the flat andwavy surfaceswith

the help of integrated circuit-piezoelectric and atomic-layerthermopile sensors. The experimental results qualitatively agreewith the direct-numerical-simulation predictions. Namely, it wasfound that the wavy wall leads to significant reduction of the spectralpeak associated with the second-mode instability. However, amarginal increase of the disturbance amplitude is observed in thelow-frequency band (80 < f < 100 kHz). Presumably this peak isassociated with the modulation of the disturbance field produced bythe wavy surface, and it does not represent a new instability. Thenature of this peak as well as its relevance to transition should beclarified by further experimental studies. It is also important to test thewavy surface under quiet freestream conditions typical for actualflights.These findings encourage us to continue the experimental and

numerical studies of stability and transition on wavy surfaces. Inparticular, it is important to investigate the wavy-wall effect on thetransition locus. Also, it is desirable to estimate the range offreestream parameters (Reynolds number and Mach number), wherethe wavy surface of fixed sizes affects the boundary layer instability.Besides the laminar flow-control issue, these studies could be usefulfor predictions of transition on naturally occurring wavy surfacesassociated with bowed panels of thermal protection systems.

Acknowledgments

This work is supported by the Russian Foundation for BasicResearch (grant 10-08-00310-а) as well as by the Russian govern-ment under the grant “Measures to Attract Leading Scientists toRussian Educational Institutions” (contract 11.G34.31.0072).

References

[1] Schneider, S., “FlightData forBoundaryLayerTransition atHypersonicand Supersonic Speeds,” Journal of Spacecrafts and Rockets, Vol. 36,No. 1, 1999, pp. 8–20.doi:10.2514/2.3428

[2] Schneider, S., “Hypersonic Laminar-Turbulent Transition on CircularCones and Scramjets Forebodes,” Progress in Aerospace Sciences,Vol. 40, Nos. 1–2, 2004, pp. 1–50.doi:10.1016/j.paerosci.2003.11.001

[3] Kimmel, R., “Aspects of Hypersonic Boundary Layer TransitionControl,” 41st Aerospace Sciences Meeting and Exhibit, AIAA Paper2003-772, 2003.

[4] 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

[5] Rasheed, A., Hornung, H. G., Fedorov, A. V., and Malmuth, N. D.,“Experiments on Passive Hypervelocity Boundary Layer Control Usingan Ultrasonically Absorptive Surface,” AIAA Journal, Vol. 40, No. 3,2002, pp. 481–489.doi:10.2514/2.1671

[6] Sandham, N. D., and Lüdeke, H., “Numerical Study of Mach 6Boundary-Layer Stabilization by Means of a Porous Surface,” AIAA

Journal, Vol. 47, No. 9, 2009, pp. 2243–2252.doi:10.2514/1.43388

[7] Fedorov, A. V., “Transition and Stability of High-Speed BoundaryLayers,” Annual Review of Fluid Mechanics, Vol. 43, 2011, pp. 79–95.doi:10.1146/annurev-fluid-122109-160750

[8] Lees, L., and Gold, H., “Stability of Laminar Boundary Layers andWakes at Hypersonic Speeds. Part 1. Stability of Laminar Wakes,”Proceedings of the International Symposium on Fundamental

Phenomena in Hypersonic Flows, Cornell University Press, Ithaca,1966

[9] Lysenko, V. I., “Experimental Studies of Stability and Transition inHigh-SpeedWakes,” Journal of Fluid Mechanics, Vol. 392, Aug. 1999,pp. 1–26.doi:10.1017/S0022112099004723

[10] Balakumar, P., “Stability of Hypersonic Boundary-Layers over aCompression Corner,” 32nd AIAA Fluid Dynamics Conference and

Exhibit, AIAA Paper 2002-2848, 2002.[11] Egorov, I. V., Fedorov, A. V., andNovikov, A. V., “NumericalModeling

of the Disturbances of the Separated Flow in a Rounded CompressionCorner,” Fluid Dynamics, Vol. 41, No. 4, 2006, pp. 521–530.doi:10.1007/s10697-006-0070-7

Fig. 15 Calculated amplitudes of wall-pressure disturbances fordifferent forcing frequencies (upperpanel) and the spectrumofmeasuredheat-flux disturbances (lower panel) in the x station of ALTP 1 sensor(x � 165 mm), Re1∞ � 10.50 × 106 m−1.

BOUNTIN ETAL. 1209

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44

Page 8: Stabilization of a Hypersonic Boundary Layer Using a Wavy Surface

[12] Pagella, A., Babucke, A., and Rist, U., “Two-Dimensional NumericalInvestigations of Small-AmplitudeDisturbances in a Boundary Layer atMa � 4.8: Compression Corner Versus Impinging Shock Wave,”Physics of Fluids, Vol. 16, No. 7, July 2004, pp. 2272–2281.doi:10.1063/1.1738414

[13] Marxen, O., Iaccarino, G., and Shaqfeh, E. S. G., “DisturbanceEvolution in a Mach 4.8 Boundary Layer with Two-DimensionalRoughness-Induced Separation and Shock,” Journal of Fluid

Mechanics, Vol. 648, 2010, pp. 435–469.doi:10.1017/S0022112009992758

[14] Egorov, I. V., Novikov, A. V., and Fedorov, A. V., “Direct NumericalSimulation of Supersonic Boundary Layer Stabilization Using GroovedWavy Surface,” 48th AIAA Aerospace Sciences Meeting, AIAA Paper2010-1245, 2010.

[15] Kontinos, D. A., and Palmer, G., “Numerical Simulation of MetallicThermal Protection System Panel Bowing,” Journal of Spacecraft andRockets, Vol. 36, No. 6, 1999, pp. 842–849.doi:10.2514/2.3523

[16] Maslov, A. A., Bountin, D. A., Shiplyuk, A. N., Smorodsky, B. V.,Knauss, H., Gaisbauer, U., Wagner, S., and Betz, J., “ALTP SensorApplication for Boundary Layer Measurements,” Proceedings of

the 10th International Conference on Methods of Aerophysical

Research, Part 2, Parallel, Novosibirsk, Russia, June–July 2004,pp. 137–146.

[17] Egorov, I. V., Fedorov, A. V., and Soudakov, V. G., “Direct NumericalSimulation ofDisturbancesGenerated by Periodic Suction-Blowing in aHypersonic Boundary Layer,” Theoretical and Computational Fluid

Dynamics, Vol. 20, No. 1, 2006, pp. 41–54.doi:10.1007/s00162-005-0001-y

[18] Egorov, I. V., Fedorov, A. V., and Soudakov, V. G., “Direct NumericalSimulation of Supersonic Boundary-Layer Receptivity to AcousticDisturbances,” 43rd AIAA Aerospace Sciences Meeting and Exhibit,AIAA Paper 2005-97, 1997.

[19] Egorov, I. V., Fedorov, A. V., Novikov, A. V., and Soudakov, V. G.,“Direct Numerical Simulation of Supersonic Boundary-LayerStabilization by Porous Coatings,” 45th AIAA Aerospace Sciences

Meeting and Exhibit, AIAA Paper 2007-948, 2007.[20] Driscoll, T. A., and Vavasis, S. A., “Numerical Conformal Mapping

Using Cross-Ratios and Delaunay Triangulation,” SIAM Journal on

Scientific Computing, Vol. 19, No. 6, 1998, pp. 1783–1803.doi:10.1137/S1064827596298580

[21] Fedorov, A. V., and Khokhlov, A. P., “Prehistory of Instability in aHypersonic Boundary Layer,” Theoretical and Computational Fluid

Dynamics, Vol. 14, No. 6, 2001, pp. 359–375.doi:10.1007/s001620100038

M. ChoudhariAssociate Editor

1210 BOUNTIN ETAL.

Dow

nloa

ded

by N

EW

YO

RK

UN

IVE

RSI

TY

on

Apr

il 28

, 201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.J

0520

44


Recommended