11
15 th International Conference on Fluid Control, Measurements and Visualization 27-30 May, 2019, Naples, Italy Impact of a gap on boundary layer transition at Mach 6 Yifei Xue 1 , Z.J. Wang 2 , Song Fu 1,* 1 School of Aerospace Engineering, Tsinghua University, Beijing, China 2 Department of Aerospace Engineering, University of Kansas, Lawrence, Kansas, USA * corresponding author: [email protected] Abstract Laminar-turbulent transition in a hypersonic boundary layer can be influenced by imperfections on the wall surface. Transition can be delayed or accelerated depending on the type, configuration and location of the imperfections. Both natural transition and the transition triggered by the imperfections are poorly understood. This paper studies a Mach 6 transitional flow over a flat plate with a gap, which is a simple geometry of the imperfection. We investigate the interaction between the gap and forcing waves using a high order implicit large eddy simulation tool. The forcing waves are obtained by combining a two dimensional instability wave and a pair of sub-harmonic three dimensional instability waves. First, flow structures including expansion and shock waves near the gap are analyzed to illustrate the basic flow. Second, we compare the behavior of transition processes of two angles of attack, AoA = 0 and AoA = 7 . The skin friction coefficient distribution with a gap at AoA = 0 is almost the same as that without a gap, which indicates that the transition process is not disturbed by the gap. In contrast, a slightly earlier transition is observed with a gap at AoA = 7 than without a gap. At AoA = 7 , a Fourier analysis shows that new disturbances with broadband frequencies are triggered in the gap, propagate downstream and influence the amplification of the instability modes. Keywords: Hypersonic transition, implicit large eddy simulation, gap, high order 1 Introduction Imperfections or roughness elements may trigger an early transition or late transition in a hypersonic boundary layer. A two dimensional (2D) gap is one of the simplest geometric imperfections. The influence of a 2D gap on transition has been well studied under subsonic flow conditions. Many criteria [1] are proposed to predict the effects of imperfections on transition. A gap is favorable to the substantial downstream TS-growth when the gap is located after the neutral point. To trigger an early transition, the Reynolds number based on the characteristic length of the imperfection is always found to be larger than a critical Reynolds number. In transonic flows, the increase in Mach number can stabilize two and three dimensional modes in an open-cavity flow according to Sun[2]. However, there are few studies on hypersonic transitional flow due to imperfections, such as the 2D gap. The free stream turbulence is considered to be composed of slow acoustic waves, fast acoustic waves, vorticity waves and entropy waves. Two discrete modes originating from the slow and fast acoustic branches in the eigenvalue spectrum[3], fast discrete mode (mode F) and slow discrete mode (mode S), are the two major modes in a hypersonic boundary layer. When mode S and mode F have the same phase velocity, a synchronization occurs. We name the point the synchronization point. The boundary layer is table before the neutral point (the neutral position of the lower branch of the neutral curve) because the disturbances are damped outside the neutral curve. An imperfection located before the neutral point does not trigger an amplification of the disturbance. When the imperfection is located between the neutral point and the synchronization point, the growth of the disturbance can be accelerated. Mode S can be amplified when a bump imperfection is placed upstream of the synchronization point[4]. The imperfection has the largest influence when it is close to the synchronization point[5]. In another study, a 2D imperfection located downstream of the synchronization point is shown to decrease the Mack second mode[6]. However, only little influence on instability modes such as mode S[7] can be observed with the analysis of stability theory. Transitional processes interact with two dimensional imperfections (i.e., oblique wave breakdown or subharmonic wave breakdown) in a hypersonic boundary also need to be investigated. Besides the existing instability modes, frequency change[8] or new instabilities can also be triggered by imperfections. For example, the Kelvin-Helmholtz instability of a shear layer[9]. In addition, two kinds of oscillation mechanisms[10] within the gap are observed in supersonic transitions. Gap flows in this work can Paper ID:312 1

Impact of a gap on boundary layer transition at Mach 615th International Conference on Fluid Control, Measurements and Visualization 27-30 May, 2019, Naples, Italy Impact of a gap

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Impact of a gap on boundary layer transition at Mach 6

    Yifei Xue1, Z.J. Wang2, Song Fu1,*

    1School of Aerospace Engineering, Tsinghua University, Beijing, China2Department of Aerospace Engineering, University of Kansas, Lawrence, Kansas, USA

    *corresponding author: [email protected]

    Abstract Laminar-turbulent transition in a hypersonic boundary layer can be influenced by imperfections on the

    wall surface. Transition can be delayed or accelerated depending on the type, configuration and location of the

    imperfections. Both natural transition and the transition triggered by the imperfections are poorly understood. This

    paper studies a Mach 6 transitional flow over a flat plate with a gap, which is a simple geometry of the imperfection.

    We investigate the interaction between the gap and forcing waves using a high order implicit large eddy simulation

    tool. The forcing waves are obtained by combining a two dimensional instability wave and a pair of sub-harmonic

    three dimensional instability waves. First, flow structures including expansion and shock waves near the gap are

    analyzed to illustrate the basic flow. Second, we compare the behavior of transition processes of two angles of

    attack, AoA = 0◦ and AoA = −7◦. The skin friction coefficient distribution with a gap at AoA = 0◦ is almost thesame as that without a gap, which indicates that the transition process is not disturbed by the gap. In contrast,

    a slightly earlier transition is observed with a gap at AoA = −7◦ than without a gap. At AoA = −7◦, a Fourieranalysis shows that new disturbances with broadband frequencies are triggered in the gap, propagate downstream

    and influence the amplification of the instability modes.

    Keywords: Hypersonic transition, implicit large eddy simulation, gap, high order

    1 Introduction

    Imperfections or roughness elements may trigger an early transition or late transition in a hypersonic boundary

    layer. A two dimensional (2D) gap is one of the simplest geometric imperfections. The influence of a 2D gap

    on transition has been well studied under subsonic flow conditions. Many criteria [1] are proposed to predict

    the effects of imperfections on transition. A gap is favorable to the substantial downstream TS-growth when

    the gap is located after the neutral point. To trigger an early transition, the Reynolds number based on the

    characteristic length of the imperfection is always found to be larger than a critical Reynolds number.

    In transonic flows, the increase in Mach number can stabilize two and three dimensional modes in an

    open-cavity flow according to Sun[2]. However, there are few studies on hypersonic transitional flow due to

    imperfections, such as the 2D gap. The free stream turbulence is considered to be composed of slow acoustic

    waves, fast acoustic waves, vorticity waves and entropy waves. Two discrete modes originating from the slow

    and fast acoustic branches in the eigenvalue spectrum[3], fast discrete mode (mode F) and slow discrete mode

    (mode S), are the two major modes in a hypersonic boundary layer. When mode S and mode F have the same

    phase velocity, a synchronization occurs. We name the point the synchronization point. The boundary layer

    is table before the neutral point (the neutral position of the lower branch of the neutral curve) because the

    disturbances are damped outside the neutral curve. An imperfection located before the neutral point does not

    trigger an amplification of the disturbance. When the imperfection is located between the neutral point and the

    synchronization point, the growth of the disturbance can be accelerated. Mode S can be amplified when a bump

    imperfection is placed upstream of the synchronization point[4]. The imperfection has the largest influence

    when it is close to the synchronization point[5]. In another study, a 2D imperfection located downstream of

    the synchronization point is shown to decrease the Mack second mode[6]. However, only little influence on

    instability modes such as mode S[7] can be observed with the analysis of stability theory. Transitional processes

    interact with two dimensional imperfections (i.e., oblique wave breakdown or subharmonic wave breakdown)

    in a hypersonic boundary also need to be investigated.

    Besides the existing instability modes, frequency change[8] or new instabilities can also be triggered by

    imperfections. For example, the Kelvin-Helmholtz instability of a shear layer[9]. In addition, two kinds of

    oscillation mechanisms[10] within the gap are observed in supersonic transitions. Gap flows in this work can

    Paper ID:312 1

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    be classified open cavity flow which has two oscillation mechanisms[11]. The first one is the longitudinal

    oscillation which corresponds to the Rossiter mode[12]. The reflected acoustic wave propagates from the aft

    corner to the front corner. The other one is the transverse oscillation. The transverse mechanism has a feedback

    loop within the cavity. The reflected acoustic wave firstly propagates down to the bottom of the gap and

    then transverses to the shear flow. Phenomenons of longitudinal and transverse mechanisms can be observed in

    experiments at Mach 1.71[10]. In a hypersonic flow, these new instability modes need to be further investigated.

    The main contributions of this study include: the wall-resolved large eddy simulations (LES) of a Mach 6

    hypersonic subharmonic transitional flow over a flat plate with a single 2D gap at two angles of attacks, and

    a detailed analysis of the forcing instability waves on the accelerated transition process. The neutral positions

    of the forcing waves are analyzed with a linear stability theory. The oscillation mechanisms of two angles of

    attack are studied by decomposing the instability waves with a Fourier transformation.

    2 Numerical Method

    This section briefly reviews the numerical method used in the present study. The high order solver, hpMusic, is

    based on the flux reconstruction (FR) or correction procedure via reconstruction (CPR) method. This method is

    originally developed by Huynh [13] for hyperbolic conservation laws, and later extended to mixed meshes by

    Wang and Gao, and Haga et al[14][15][16]. Other developments in the FR/CPR methods are reviewed in [17].

    We choose the FR/CPR method because of its ability in handling unstructured meshes, its high-order accuracy,

    its simplicity like a finite difference method, and its scalability on supercomputers. The unsteady compressible

    Navier-Stocks equations are discretized using the with FR/CPR method. We use the following conservation

    law to introduce the basic idea

    ∂U

    ∂ t+∇ ·F(U) = 0, (1)

    where U is the vector of conservative variables, and F is the flux vector. The computational domain is dis-

    cretized with non-overlapping elements Vi. In each element, the conservation law is transformed into a weighted

    residual form with an arbitrary test function W

    Vi

    (∂U

    ∂ t+∇ ·F(U))WdV = 0. (2)

    The conservative variables U is assumed to be a polynomial of degree k, Ui ∈ Pk. Using integration by parts in

    the second term, we arrive at

    Vi

    ∂Ui∂ t

    WdV +∫

    ∂ViWF(Ui) ·ndS−

    Vi

    ∇W ·F(Ui)dV = 0. (3)

    We replace the normal flux term at element interfaces with a common Riemann flux Fncom to achieve conserva-

    tion. Applying integrating by parts again to∫

    Vi∇W ·F(Ui)dV , we obtain

    Vi

    ∂Ui∂ t

    WdV +∫

    ∂ViW [Fncom −F

    n(Ui)]dS−∫

    Vi

    W∇ ·F(Ui)dV = 0. (4)

    The Riemann flux Fncom is computed with the Roe Riemann solver by the use of Vi and Vi+, in which the subscript

    "i+" denotes the neighbor element

    Fncom = Fn

    com(Ui,Ui+,n). (5)

    In order to simplify (4), we replace the surface integral with a volume integral via a lifting operator, δi ∈ Pk(Vi):

    ∂ViW [Fncom −F

    n(Ui)]dS =∫

    Vi

    WδidV. (6)

    Paper ID:312 2

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Fig. 1 Schematic of the computational domain

    Substituting (6) into (4), we obtain

    Vi

    [∂Ui∂ t

    +∇ ·F(Ui)+δi]WdV = 0. (7)

    The aim of introducing the lifting operator is to derive the differential equation from the integral one. As-

    suming that the flux vector F(Ui) can be approximated with a polynomial, then we obtain the final formulation:

    ∂Ui, j∂ t

    +∏ j(∇ ·F(Ui))+δi, j = 0, (8)

    where ∏ j denotes a projection to the polynomial space and the subscript j represents the projection at the

    solution point j, ∏ j(∇ ·F(Ui)) ∈ Pk.

    The FR/CPR method converts the weighted residual form from an integral one to a differential one. It is

    compact in that the scheme only needs the immediate face neighbors. The viscous flux in the compressible NS

    equations is computed with the Bassi-Rebay 2 (BR2) scheme[18]. No explicit subfilter scale models are used.

    Therefore the present approach is called an implicit LES or ILES. The backward-difference formula (BDF2)

    with a LU-SGS solver[19] and a 3rd Runge-Kutta method are employed for time marching. An accuracy pre-

    serving limiter [20] is adopted to capture shock-waves and maintain the high order accuracy of the simulations

    elsewhere. The LES tool has gone through an extensive verification and validation process[21].

    3 Simulation results and discussions

    3.1 Computational setup

    To investigate whether a 2D gap affects the subharmonic transition of a Mach 6 hypersonic flow, we perform a

    wall-resolved ILES of the transitional flow. In order to avoid the difficulty of simulating a strong leading edge

    shock with a high order method, the computational domain starts behind the leading edge and ends after the

    formation of the turbulent boundary layer. The gap configuration and the computational domain are illustrated

    in Fig. 1.

    A laminar boundary layer profile is prescribed at the inflow boundary. The freestream Mach number is

    M∞ = 6 and the free-stream unit Reynolds number is Re∞ = 1.0×107/m, the static temperature is T∞ = 55K.

    The inflow conditions are of the same order of magnitude as the parameters of the turbulent wind tunnel at

    Peking University. The computational domain is 0.02 ≤ x ≤ 0.7m in the streamwise direction and 0.0 ≤ z ≤0.018m in the spanwise direction. Two complete oblique forcing waves are generated at the disturbance stripin the spanwise direction. The instability waves are introduced to the flow by blowing and suction at the

    disturbance strip. The gap starts at x = 0.2m and ends at x = 0.22m with a depth h = 0.01m. A periodicboundary condition is used in the spanwise direction. Buffer region are adopted at both the right end and the

    top of the domain.

    Next we explain the method of blowing and suction. Simultaneous blowing and suction is used in this work

    to make sure no additional fluid is added into the flow. The temperature on the disturbance strip is replaced

    Paper ID:312 3

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Fig. 2 Slice of mesh for AoA = 0◦ gap flow

    with an isothermal condition[22]. The temperature equals the adiabatic wall temperature of the steady solution.

    Following the blowing and suction function of Huai et al[23] , the wall-normal velocity V is prescribed as:

    v(x,z, t) = A2D f (x)sin(ω2Dt)+A3Dg(z) f (x)sin(ω3Dt), (9)

    where, A2D and ω2D are the two dimensional wave’s amplitude and frequency, respectively. A3D and ω3D are

    the oblique wave’s amplitude and frequency. Then the shape function is defined following Fasel et at[24]:

    | f (ξ )|= 15.1875ξ 5 −35.4375ξ 4 +20.25ξ 3,

    g(z) = cos(2πz/λz).(10)

    ξ can be obtained with xm = (x1 + x2)/2:

    ξ =

    x− x1xm − x1

    x1 ≤ x ≤ xm,

    x2 − x

    x2 − xmxm ≤ x ≤ x2.

    (11)

    The location of disturbance strip in the streamwise direction is 0.035 ≤ x ≤ 0.055m (x1 = 0.035m and x2 =0.055m). To obtain a subharmonic transition in the computational domain, a two dimensional wave and a pairof subharmonic oblique waves are added in the disturbance strip. Frequencies of the oblique waves are one half

    of that of the two dimensional wave. The resonance of the three waves leads to the subharmonic transition[25].

    The disturbance amplitude A and dimensionless frequencies F of AoA = 0◦ are set to:

    A2D = 2%U∞, F2D = 0.6×10−4;

    A3D = 2%U∞, F3D = 0.3×10−4.

    (12)

    A negative angle of attack leads to adverse pressure gradient which contributes to the transition. So the distur-

    bance amplitudes at AoA =−7◦ are set to smaller values:

    A2D = 1%U∞, F2D = 0.6×10−4;

    A3D = 0.05%U∞, F3D = 0.3×10−4.

    (13)

    The ω3D and ω2D can be calculated with:

    ω =FU2∞

    ν. (14)

    The mesh used in the simulation ( i.e., a slice mesh of gap flow at z = 0.009m, AoA = 0◦) is illustrated inFig. 2. Mesh in region x ∈ (0.5,0.7)m is progressively coarsened in the streamwise direction which can reducethe disturbances from outflow boundary. The degree of freedoms (DOFs) in each directions are lists in Table 1

    where DOFsin and DOFsout denote the DOFs in the gap and out of the gap, respectively.

    3.2 Stability analysis of the instability waves

    Paper ID:312 4

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Table 1 Grid resolution of the simulation

    nx ny nz ∆x+max ∆y

    +max ∆z

    +max

    gap DoFsin 70 81 140

    DoFsout 1544 94 140 5.7(AoA = 0◦) 2.5(AoA = 0◦) 2.4(AoA = 0◦)

    /10.3(AoA =−7◦) /4.7(AoA =−7◦) /4.3(AoA =−7◦)

    smooth DoFs 1544 94 140 5.7(AoA = 0◦) 2.5(AoA = 0◦) 2.4(AoA = 0◦)/10.3(AoA =−7◦) /4.7(AoA =−7◦) /4.3(AoA =−7◦)

    Before the implicit large eddy simulation, we need to choose a reasonable inflow disturbance for the hyper-

    sonic boundary layer. Stability characteristics of disturbance strip are investigated by LST in this part. We can

    compare the gap location with the neutral point and the synchronization point through the stability analysis.

    The LST analysis explains why we choose the frequencies of two dimensional wave F2D = 0.6×10−4 and

    the subharmonic oblique waves F3D = 0.3×10−4 at the disturbance strip. Here we start with the introduction

    of LST[26]. The basic flow for the LST analysis is computed from the compressible Blasius boundary layer

    equations. The governing equations derived from compressible Navier-Stocks equations can be separated into

    the basic-state equations and the disturbance equations. The form of the disturbances are assumed to be:

    q̃(x,y,z, t) = q̂(y)exp(i(αx+β z−ωt))+ c.c., (15)

    where c.c. stands for complex conjugate. In spatial stability, α = αr + iαi is the complex streamwise wavenum-ber, β is the spanwise wave number and ω is the angular frequency. We denote the Reynolds number based on

    boundary layer thickness with R, and then the eigenvalue problem can be expressed as:

    α = f (β ,ω,R). (16)

    For β = 0, the neutral curve obtained from LST is shown in Fig. 3(a). Instability wave with a frequency ofF2D = 0.6× 10

    −4 will grow within the computational domain. The synchronization point of two dimensional

    instability wave predicted in Fig. 3(b) is x = 0.3635m. The gap is upstream of the synchronization point.Growth of instability wave F2D = 0.6× 10

    −4 should be affected by the gap according to LST analysis. Gap’s

    effect on transition process is discussed with ILES results in the next part.

    (a) Neutral curve for β = 0

    (b) Phase velocity of mode F and mode S

    Fig. 3 LST results

    Then we fix R = 741.63 (x = 0.055m, end of the disturbance strip), the contour levels of αi can be obtainedwith LST. As shown in Fig. 4, the oblique disturbance (F = 0.3×10−4, β = 0.0405) locates inside the neutralcurve (αi < 0). So the oblique waves will grow in downstream direction.

    We can conclude that the current instability waves will grow monotonously. The oblique transition process

    with current forcing frequencies and wave numbers is expected to be affected by the gap.

    Paper ID:312 5

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Fig. 4 Neutral curve for R = 741.63

    Fig. 5 Overall vortex evolution of the gap flow transition process, AoA = 0◦

    3.3 Transition processes at two angles of attack

    Gap flow of AoA = 0◦

    Complete transition processes of smooth and gap flows are obtained with ILES. Similarity results of compress-

    ible boundary layer equations are imposed at the inflow boundary. Subharmonic transitions are induced by the

    forcing instability waves at disturbance strip. An overview of gap flow transition process at AoA = 0◦ demon-strated with vortices is shown in Fig. 5. Vortices are visualized with Q-criterion colored with non-dimensional

    streamwise velocity. The background is the numerical schlieren. Forcing disturbances grow in boundary layer

    and trigger the transition. Two pairs of λ vortices are obtained in spanwise direction at the transition onset. The

    vortex heads grow in normal direction and interact with the boundary layer when approaching the boundary

    layer, as shown in Fig 6(a). Multiple hairpin vortices are observed in Fig 6(b). The hairpin vortices are hard

    to be distinguished because the interaction between the boundary layer and the hairpin vortices. Complicated

    vortices are observed in the later stage of transition, shown in Fig 6(c). Mixture of large and small vortices

    leads to the turbulent boundary layer.

    (a) First stage (b) Second stage (c) Third stage

    Fig. 6 Three vortex stages of the gap flow transition process, AoA = 0◦

    The transition process of smooth flat plate is almost the same with that on gap case. The gap has little effect

    on the transition process at AoA = 0◦. However, there are some local flow structures occur at the corner of thegap, as shown in Fig. 5. So the flow at the gap is particularly discussed. Averaged pressure and streamlines are

    used to illustrate the flow near the gap in Fig. 7. Three separations are observed in the gap including one main

    separation and two small separations on each side of the main separation. Velocity in the gap is mostly less than

    Paper ID:312 6

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Fig. 7 Averaged pressure distribution and streamwise in the gap at the center z-plane

    Fig. 8 Instantaneous slice of streamwise velocity at x-z-plane

    1% of the velocity outside the boundary layer which indicates that the gap is nearly a dead zone. The maximum

    velocity locates at the upper border of the main separation as a result of the acceleration of the expansion from

    the upstream corner. An expansion is followed by a shock at the downstream corner which decelerates the

    flow to an undisturbed condition. The expansion and shock are weak that the streamlines at y = 0m are nearlyundisturbed by the gap.

    The instantaneous streamwise velocity evolution above the cavity is illustrated in Fig. 8 at the position of

    y = 0.0075x+ 0.00105m. A local acceleration is observed at the gap and the streamwise velocity decreaserapidly after the gap. This confirms the velocity acceleration at the gap in the time-averaged results. No special

    change on the streamwise velocity is observed far behind the gap comparing with the velocity evolution in

    smooth flow.

    For the current forcing waves, transition phenomenons of smooth flow and gap flow are the same. The in-

    fluence of the gap is limited to the vicinity of the gap based on the time-averaged results and the instantaneous

    results. As a result, the distribution of skin friction coefficient of gap flow is nearly same with the smooth one

    except for the vicinity of the gap, as shown in Fig. 9.

    Gap flow of AoA =−7◦

    Hypersonic aircrafts usually have a slender, streamlined fuselage to reduce the drag. In addition, the angle

    of attack of the cruise phase is usually limited to several degrees. We choose a possible angle of attack to inves-

    tigate how the angle of attack influences the transition process. A negative angle of attack is likely to occur on

    Fig. 9 Time-averaged and spanwise-averaged distribution of skin friction coefficient at AoA = 0◦

    Paper ID:312 7

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    Fig. 10 Instantaneous slice of streamwise velocity at x-z-plane

    Fig. 11 Time averaged and spanwise averaged distribution of skin friction coefficient at AoA =−7◦

    the aircraft surface which means the fluid flow toward the the surface. Transition processes of AoA =−7◦ gapflow are studied in this part. Numerical results of laminar flat plate are imposed at the inflow boundary, which

    are derived from a simulation of a complete geometry including the leading edge. Typical flow structures of the

    transition at AoA =−7◦ are nearly the same with AoA = 0◦ case. However, the appearance of the λ vortices isearlier than AoA = 0◦ case. Contour of instantaneous streamwise velocity at the slice of y = 0.005x+ 0.0005illustrated in Fig. 10 shows a higher amplification of three dimensional instability waves after the gap.

    Since the larger wave amplification in the AoA = −7◦ gap flow, a relatively earlier transition is observedcomparing to the smooth flow at AoA =−7◦.

    3.4 Influence of the angle of attack on the transition process

    A gap flow at AoA = 0◦ has almost the same transition process as the corresponding smooth flow. How-ever, when the angle of attack is negative, a gap can trigger an earlier transition compared to the corresponding

    smooth one. In order to find out the mechanism of the interaction between forcing instability wave and gap, we

    perform Fourier transformations at both angles of attack.

    Mode amplitude curves in streamwise direction are plotted in Fig. 12 and Fig. 13. Modes ( f ,k) are denotedby frequency f and spanwise wave number k, where f = 1 denotes the forcing frequencies of the oblique waves(F = 0.3×10−4) and k = 1 denotes the fundamental wave number in spanwise direction (two complete waves).For each mode, Y-axis denotes the maximum streamwise velocity perturbation in boundary layer, which is

    non-dimensionalized with the freestream velocity.

    Modes in the gap flow at AoA = 0◦ ( Fig. 12(a) ) have similar modal developments comparing with thesmooth flow ( Fig. 12(b) ). Different modal developments are observed in the gap flow at AoA = −7◦. Theforcing modes (mode(1,1), mode(2,0)),and their second harmonic modes (mode(4,0), mode(2,2)) have the same

    trend in streamwise direction as the smooth flow, which indicates the forcing modes are almost not disturbed

    by the gap. However, other modes with multiple frequencies and spanwise wave numbers are excited in the gap

    and develop downstream. Not only are the modes in Fig. 13(a), other frequencies and spanwise wave numbers

    increase rapidly. Among all the modes, the forcing modes, mode(1,1) and mode(2,0), always dominate in the

    development which determine the transition onset. This may be the reason why the change of transition onset

    by a gap is not obvious.

    In order to find out the reason for the gap flow behaviors of different modes, two slices of non-dimensional

    Paper ID:312 8

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    (a) Gap, AoA = 0◦(b) Smooth, AoA = 0◦

    Fig. 12 Streamwise velocity disturbance evolution at AoA = 0◦

    (a) Gap, AoA =−7◦(b) Smooth, AoA = 0◦

    Fig. 13 Streamwise velocity disturbance evolution at AoA =−7◦

    pressure fluctuation are plotted in Fig 14. Similar to the longitudinal and transverse mechanisms proposed by

    Zhang[11], two feedback mechanisms are observed in the two gap flows with different angles of attack. In the

    gap flow at AoA = 0◦, the upstream disturbances propagate towards the aft face of the gap. Most disturbancesdamp at the right corner. As a result, there are no strong disturbance flow out the boundary layer. In contrast,

    the disturbance path in the gap flow at AoA = −7◦ is illustrated in Fig. 14(b). There is a single feedback loopin the gap. Disturbances from the boundary layer propagate to the aft face of the gap, flow to the bottom with

    the feedback loop and finally transverse back to the boundary layer. According to the experimental results[10],

    a longitudinal mechanism exists in a shallow gap and a deep gap leads to a transverse mechanism. The results

    in this paper show that a negative angle of attack can also cause a transition from longitudinal to transverse

    mode besides the length-to-depth ratio. This may be the reason for the occurrence of the broadband distur-

    bance observed in the FFT analysis. Unfortunately, the typical frequencies for the two mechanisms can not

    be extracted with the current frequency analysis. A large number of instability waves are produced with this

    nonlinear interaction, causing the sudden rise in mode amplitudes, as shown in Fig.13(a).

    (a) Pressure fluctuation in gap flow at AoA = 0◦ (b) Pressure fluctuation in gap flow at AoA =−7◦

    Fig. 14 Distribution of pressure fluctuation at the slice of z = 0.009m

    Paper ID:312 9

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    4 Conclusions

    Transition processes of a Mach 6 boundary layer with a 2D gap have been studied for two angles of attack,

    AoA = 0◦ and AoA = −7◦ with a high order ILES tool, hpMusic, in the present paper. Both 2D and 3Dinstability waves were introduced at the inflow to obtain a subharmonic transition in the simulations. Some

    conclusions are drawn next.

    In the case of AoA = 0◦, the transition phenomenon of the gap flow is similar to that on smooth one. Theinfluence of the gap occurs mainly in the vicinity of the gap. The flow outside the gap is almost undisturbed.

    There is no extra disturbance observed in the gap, and the evolution of the disturbances downstream the gap

    is similar to that of the smooth flow. Instability modes propagate toward the aft surface of the gap and damp

    rapidly. As a result, little difference is observed in the distribution of skin friction coefficient between gap flow

    and smooth flow.

    The forcing disturbances show a different interaction with the gap at AoA =−7◦. Transition onset is earlierthan the corresponding smooth one. Three dimensional vortex structure in the gap flow occurs earlier than that

    in the smooth flow. According to the FFT analysis, plenty of instability modes other than the forcing modes or

    the modes originated from the forcing modes, are observed in the gap and flow downstream which contribute

    to transition. A single feedback loop is observed within the gap which is considered to be the reason for the

    jump of the broadband modes at the gap.

    References

    [1] Forte M, Perraud J, Seraudie A, Beguet S, Gentili L, Casalis G (2015) Experimental and Numerical Study

    of the Effect of Gaps on Laminar Turbulent Transition of Incompressible Boundary Layers, In:Procedia

    IUTAM, Vol. 14, pp 448-458, doi:10.1016/j.piutam.2015.03.073.

    [2] Sun Y, Taira K, Cattafesta L, Ukeiley L (2017) Biglobal instabilities of compressible open-cavity flows.

    Journal of Fluid Mechanics, vol. 826, pp 270-301. doi:10.1017/jfm.2017.416

    [3] Tumin A (2007) Three-dimensional spatial normal modes in compressible boundary layers. Journal of

    Fluid Mechanics, vol. 586, pp 295-322. doi:10.1017/S002211200700691X

    [4] Duan L, Wang X, and Zhong X (2013) Stabilization of a Mach 5.92 Boundary Layer by Two-Dimensional

    Finite-Height Roughness. AIAA Journal, vol. 51(1), pp 266-27. doi:10.2514/1.J051643

    [5] Fong K D, Wang X, Zhong X (2014) Numerical simulation of roughness effect on the stability of a hy-

    personic boundary layer. Computers & Fluids, vol. 96, pp 350-367. doi:10.1016/j.compfluid.2014.01.009

    [6] Sawaya J, Sassanis V, Yassir S, Sescu A, Visbal M (2018) Assessment of the Impact of Two-Dimensional

    Wall Deformation Shape on High-Speed Boundary-Layer Disturbances. AIAA Journal, vol. 56(12), pp

    4787-4800. doi:10.2514/1.J057045

    [7] Marxen O, Iaccarino G, Shaqfeh E S (2014) Nonlinear instability of a supersonic boundary layer with two-

    dimensional roughness. Journal of Fluid Mechanics, vol. 752, pp 497-520. doi:10.1017/jfm.2014.266

    [8] Tang Q, Zhu Y, Chen X, Lee C (2015) Development of second-mode instability in a mach 6 flat

    plate boundary layer with two-dimensional roughness. Physics of Fluids, vol. 27(6), pp 064105.

    doi:10.1063/1.4922389

    [9] Heller H, Bliss D (1975) The physical mechanism of flow-induced pressure fluctuations in cavities and

    concepts for their suppression. In:2nd Aeroacoustics Conference. doi:10.2514/6.1975-491

    [10] Kumar M, Vaidyanathan A (2018) Oscillatory mode transition for supersonic open cavity flows. Physics

    of Fluids, vol. 30(2), pp 026101. doi.org/10.1063/1.5017269

    [11] Zhang X, Edwards J (1990) An investigation of supersonic oscillatory cavity flows driven by thick shear

    layers. The Aeronautical Journal, vol. 94(940), pp 355-364. doi:10.1017/S0001924000023319

    [12] Rossiter J E (1964) Wind tunnel experiments on the flow over rectangular cavities at subsonic and tran-

    sonic speeds. Ministry of Aviation.

    Paper ID:312 10

  • 15th International Conference on Fluid Control, Measurements and Visualization

    27-30 May, 2019, Naples, Italy

    [13] Huynh H T (2007) A Flux Reconstruction Approach to High-Order Schemes Including Discon-

    tinuous Galerkin Methods. In: 18th AIAA Computational Fluid Dynamics Conference, pp 4079,

    doi:10.2514/6.2007-4079

    [14] Wang, Z J, Gao H (2009) A unifying lifting collocation penalty formulation including the discontinuous

    galerkin, spectral volume/difference methods for conservation laws on mixed grids. Journal of Computa-

    tional Physics, vol. 228(21) pp 8161-8186. doi:10.1016/j.jcp.2009.07.036

    [15] Haga T, Gao H, Wang Z J (2011) A high-order unifying discontinuous formulation for the Navier-Stokes

    equations on 3D mixed grids. Mathematical Modelling of Natural Phenomena, vol. 6(3), pp 28-56. doi:

    doi:10.1051/mmnp/20116302

    [16] Wang Z J, Gao H, Haga T (2011) A Unifying Discontinuous CPR Formulation for the Navier-Stokes

    Equations on Mixed Grids. In: Kuzmin A. (eds) Computational Fluid Dynamics 2010. pp 59-65, Berlin,

    Heidelberg. doi:10.1007/978-3-642-17884-9_5

    [17] Huynh H T, Wang Z J, and Vincent P E (2014) High-order Methods for Computational Fluid Dynamics:

    a Brief Review of Compact Differential Formulations on Unstructured Grids, Computers & Fluids, Vol.

    98(2), pp. 209-220. doi:10.1016/j.compfluid.2013.12.007

    [18] Bassi F, Rebay S (2000) A High Order Discontinuous Galerkin Method for Compressible Turbulent Flows.

    In: Cockburn B, Karniadakis G E, Shu CW (eds) Discontinuous Galerkin Methods, Berlin, Heidelberg pp

    77-88. doi:10.1007/978-3-642-59721-3_4

    [19] Sun Y, Wang Z J, Liu Y (2007) Efficient Implicit Non-linear LU-SGS Approach for Viscous Flow Com-

    putation using High-Order Spectral Difference Method. In: 18th AIAA Computational Fluid Dynamics

    Conference, AIAA Paper, pp 2007-4322. doi:10.2514/6.2007-4322

    [20] Li Y, Wang Z J (2017) A convergent and accuracy preserving limiter for the FR/CPR method. In: 55th

    AIAA Aerospace Sciences Meeting, AIAA SciTech Forum, pp 0756. doi:10.2514/6.2017-0756

    [21] Wang Z J, Li Y, Jia F, Laskowski G M, Kopriva J, Paliath U, Bhaskaran R (2017) Towards indus-

    trial large eddy simulation using the FR/CPR method. Computers & Fluids, vol. 156, pp 579-589.

    doi:10.1016/j.compfluid.2017.04.026

    [22] Egorov I V, Fedorov A V (2006) Soudakov V G. Direct numerical simulation of disturbances generated by

    periodic suction-blowing in a hypersonic boundary layer. Theoretical and Computational Fluid Dynamics,

    vol. 20, pp 41-54. doi:10.1007/s00162-005-0001-y

    [23] Huai X, Joslin R, Piomelli U (1997) Large-Eddy Simulation of Transition to Turbulence in Boundary

    Layers. Theoretical and Computational Fluid Dynamics, vol. 9, pp 149-163. doi:10.1007/s001620050037

    [24] Fasel H F, Konzelmann U (1990) Non-parallel stability of a flat-plate boundary layer us-

    ing the complete Navier-Stokes equations. Journal of Fluid Mechanics, vol. 221, pp 311-347.

    doi:10.1017/S0022112090003585

    [25] Chang C, Malik M (1994) Oblique-mode breakdown and secondary instability in supersonic boundary

    layers. Journal of Fluid Mechanics, vol. 273, pp 323-360. doi:10.1017/S0022112094001965

    [26] Ren J, Fu S (2014) Competition of the multiple Gortler modes in hypersonic boundary layer flows. Science

    China Physics, Mechanics & Astronomy, vol. 57, pp 1178-1193. doi:10.1007/s11433-014-5454-9

    Paper ID:312 11

    IntroductionNumerical MethodSimulation results and discussionsConclusions