Unsteady Motion of Shock Wave For A Supersonic Compression Ramp Flow Based On Large Eddy Simulation
Unsteady Motion of Shock Wave For A Supersonic Compression Ramp Flow Based On Large Eddy Simulation
Unsteady Motion of Shock Wave For A Supersonic Compression Ramp Flow Based On Large Eddy Simulation
the ramp. It was found that when the intensity of the shock wave
is sufficiently strong, the unsteady motion of the shock wave will
become important due to turbulent fluctuation enhancement.
Ganapathisubramani et al. (Ganapathisubramani et al., 2007)
used high-speed particle image velocimetry (PIV) technology to
measure the compression ramp flow. It was found that the strip-
like structures with different momentum exist in the upstream
boundary layer, and these strip structures cause a low-frequency
pulsation of the separation bubble. Wu et al. (Wu et al., 2013)
used nano-tracer planar laser scattering (NPLS) technology to
study the laminar/turbulent SWBLI on the supersonic
compression ramp; the overall flow field was analyzed and
local fine structures were identified. FIGURE 1 | Computational domain for the LES.
Yi Zhuang et al. (Zhuang et al., 2018a) have performed an
experimental investigation on a compression ramp shock wave/
turbulent boundary layer interaction at Ma 2.83. The ice dominant frequency in the vicinity of the shock was exhibited
cluster–based planar laser scattering technique was applied to by the streamwise evolution of the pre-multiplied spectrum of
acquire high spatiotemporal resolution images at the center plane. pressure fluctuations. In the upstream boundary layer, the
Two-dimensional slices of the coherent vertical structure (CVS) spectrum presents mainly high-frequency content (St > 1)
were acquired and extracted from these images with a machine linked to the incoming turbulent eddies. However, in the
learning–based method. By comparing the features CVS acquired vicinity of the shock position, a low-frequency broadband
before and after the interaction, the evolution of CVSs in the range emerges and is centered approximately at St 0.03. As
SWBLI flow was analyzed. the author pointed out, the large gap between these frequency
In fact, a large number of experimental studies were carried scales has been reported in previous investigations by Touber and
out at higher Reynolds numbers. But Bookey et al. (Bookey et al., Sandham (Touber and Sandham 2011) and Dupont et al.
2005) selected a low-density gas as the working medium to reduce (Dupont et al., 2006).
the Reynolds number in the experiment and achieve the In this study, a large eddy simulation is used to study the
numerical simulation (DNS, LES) comparison with SWBLI phenomenon of the 24° compression ramp; detailed
experiments. Adams et al. (Adams 2000) used the DNS analysis and discussion are carried out to understand the
method to study the supersonic compression ramp flow and unsteady features of the shock wave in the SWBLI. The
found that the Reynolds shear stress after the SWBLI increases remainder of this manuscript is organized as follows. Section
more greatly than the Reynolds normal stress. Loginov et al. 2 gives a description of the numerical methods and techniques
(Loginov et al., 2006) used the LES method to study the statistical used in this study, including the generation of the inlet
parameter distributions and fluctuation characteristics of the turbulent boundary layer, geometric model, and mesh
turbulent boundary layer in the compression ramp and distribution. Section 3 compares the simulation results with
compared numerical results with the experimental data. The experiments and numerical results in the literature to verify the
flow field analysis showed that the large-scale three- reliability of the program, and the details of the flow field are
dimensional flow structure downstream of the ramp is the discussed and analyzed. The last section gives some conclusions
principal reason for the spanwise unevenness of the flow field. of this study.
Wu et al. (Wu and Martin 2008) used the DNS method to study
the SWBLI in the compression ramp, with the inlet flow
parameters consistent with the experimental conditions given 2 COMPRESSION RAMP AND NUMERICAL
by Bookey et al. (Bookey et al., 2005). By observing the METHOD
spatiotemporal evolution and using the correlation analysis of
the flow field, it was found that the shock wave motion always lags The work in this study is based on the NUAA-Turbo CFD solver
behind the pulsation of the separation bubble. Therefore, a developed by our research group. In the large eddy simulation
“feedback loop” model consisting of the separation bubble, (LES), the finite volume method is used and the ROE scheme is
shear layer, and shock wave system is proposed to explain the used for the evaluation of convective fluxes, the WENO_ZQ
mechanism of the low-frequency unsteady motion of the scheme (Zhu and Qiu 2017) with fifth-order precision for the
shock wave. interface reconstruction, and the sixth-order central difference
Kenzo S.et al. (Sasaki et al., 2021) have investigated the scheme for the spatial discretization of viscous fluxes. Time
mechanisms of low-frequency unsteadiness in an impinging discretization uses the Runge–Kutta method with a total
shock wave/turbulent boundary layer interaction at a Mach variation reduction property of third-order accuracy (Shu and
number of Ma 2. The Strouhal number St fL/U∞ was Osher 1989). The dynamic sub-grid scale model is considered,
used for the space-time spectral analysis to identify the key and the sub-grid scale viscosity coefficient is determined with the
features of the shock motion, where L is defined as the two consecutive filtering by the method of Germano et al.
interaction length and f is the frequency of fluctuation. The (Germano 1991).
FIGURE 2 | Distribution of averaged wall pressure and skin friction coefficient. (A) Averaged wall pressure. (B) Averaged skin friction coefficient.
near the wall. The flow field of the recycled plane in the auxiliary
computation domain is extracted as the inlet boundary condition
The inlet boundary layer thickness δ of the compression ramp for the primary computation. It is specially noted that when the
is used to make the length scale dimensionless. The computation primary/auxiliary computation domain uses grids of different
domain consists of two parts: the flat plane computation domain resolutions, the process of flow field extraction needs to
(auxiliary computation domain) and the 24° compression ramp interpolate the variables. For details, this method can refer to
computation domain (primary computation domain). The Zhong et al., (2021).
“recycling/rescaling” method was used to generate the For the primary computation domain, the upper boundary of
turbulent boundary layer in the auxiliary computation domain the computational domain and the outlet are set to the subsonic
and the computed turbulence information as the inlet boundary outlet boundary condition. The wall condition of the non-slip
condition of the primary computation domain. In the auxiliary isothermal is adopted to the wall, and the wall isothermal
computation, the distance from the inlet plane “7.3δ” is set as the temperature is 307 K. The spanwise boundary adopts periodic
recycled plane. The schematic of the computation domain is boundary condition, and the inlet turbulent boundary conditions
shown in Figure 1. The coordinate system origin is located at the are dynamically given by the auxiliary computation. For the
ramp, and the coordinate axes “x, y, and z” indicate the auxiliary computation, the boundary layer conditions are
streamwise, spanwise, and normal directions respectively. The consistent with the main computation, and the flow Mach
upstream and downstream lengths of the ramp are both 7.73δ, the number is Ma 2.9, the flow static temperature is 108.1 K,
spanwise width is 2.15δ, and the normal height along the wall is and the flow density is 0.074kg/m3 .
5.23δ. The number of grid points in the three directions is “505 × In order to verify the reliability of the LES software, the
89 × 112”. The grid is evenly distributed in the spanwise and numerical results of the compression ramp will be compared
refined along the streamwise direction at the ramp and in the with the experimental results under the same inlet flow
normal direction to guarantee z+ ≈ 1 in the first layer of the grid conditions.
FIGURE 6 | Distribution of Reynolds normal stresses at different streamwise locations. (A) Streamwise component. (B) Spanwise component. (C) Wall-normal
component.
FIGURE 9 | Spatial evolution of Görtler-like vortices in the compression ramp. (A) Streamwise position of the six planes downstream of the separation point. (B)
Distribution of streamwise vorticity and streamlines in (“y”,η) planes.
point. These six sections are named A ~ F, corresponding to the given in Figure 9A, in which the background is space-
streamwise direction positions: −1.7δ, −1.4δ, −0.6δ, 0.7δ, 3.3δ, time–averaged numerical schlieren. Figure 9B shows the
and 5.5δ, respectively. A schematic diagram of six positions is streamwise direction vorticity ωx contour, and streamline
Number of data Time span Sampling frequency/(U∞ /δ) Minimum frequency/(U∞ /δ)
/N /(δ/U∞ )
turbulent boundary layer, where the pressure fluctuation is not weighted power spectrum density (WPSD) expressed as
disturbed by the downstream interference area flow. It can be WPSD(f) fpPSD(f)df. From this figure, we can see more
seen that the pressure signal around Pw /P∞ presents a random clearly the frequency components of pressure signals at different flow
fluctuation, and the amplitude of fluctuation is very small. The direction positions. For the upstream boundary layer, the fluctuating
red solid line indicates the pressure fluctuation with time at the energy is mainly distributed between 100 ~ 101 , and the peak
average separation point (xsep −2.7δ). Compared with the inlet, frequency is about St 3, which is consistent with the
the pressure signal at the average separation point has larger characteristic frequency of the fully developed turbulent boundary
amplitude of fluctuation due to the interaction between the shock layer. For the signal at the average separation point, there is also a
wave and turbulent boundary layer. In addition, from the red wide small peak area in the high-frequency region, which is similar
solid line, the pressure signal at the separation point not only to the WPSD curve obtained in the upstream turbulent boundary
includes the high-frequency fluctuation but also has the low- layer. At the same time, the red curve shows a more obvious energy
frequency fluctuation component as shown in Figure 11. peak appearing in the low-frequency region, and the corresponding
Next, we will make a strict quantitative analysis of the wall characteristic frequency is St 0.04, which is completely consistent
pressure fluctuation signal from the perspective of spectral analysis. with the result (0.02–0.05) summarized by Dussauge et al. (Dussauge
The Welch method (Barbe et al., 2010) is used to segment the et al., 2006). In addition, Wang Bo et al. (Wang 2015), Kenzo S. et al.
discrete pressure signal, aiming to obtain a smoother and less (Sasaki et al., 2021), Tong et al. (Tong et al., 2017), Touber and
variance power spectrum density (PSD) curve. In this study, the Sanham (Touber and Sandham 2009), and Pasquariello et al.
collected pressure signals are divided into three segments, and the (Pasquariello et al., 2017) obtained similar results in their
coincidence rate between the segments is 50%. The Hanning respective numerical simulations.
window is used to add windows for each segment to improve the In order to study the distribution of low-frequency instability
variance performance. It should be noted that in the Welch method, characteristics in the whole field of the compression corner,
the more segmented the data, the smoother the power spectrum Figure 13 shows the distribution of WPSD on all pressure
curve and the smaller the noise, but at the same time, the resolution probes. The white solid line is used to represent the position of
of the power spectrum will be affected. Therefore, in the process of the average separation point xsep and the average reattachment point
data segmentation, we must consider the balance of noise and xrea and the white dotted line is used to represent the position of the
resolution in the power spectrum curve. Figure 11B shows the corner. Once again, it can be clearly seen that the turbulence
PSD distribution curve corresponding to the abovementioned two generation technology in this study does not introduce any low-
wall pressure signals, where the abscissa is the dimensionless frequency energy at the inlet. By observing the whole flow field, it is
frequency proposed by Dussauge et al. (Dussauge et al., 2006), found that this low-frequency instability mainly exists near the
which is defined as St fLsep /Ue . It can be seen from the figure that average separation point, while in other locations of the flow
the peak value of pressure fluctuation energy in the upstream field, the energy of pressure fluctuation almost exists in the high-
boundary layer is located in the high-frequency region, while the frequency range. Especially when the boundary layer is reattached in
peak value of pressure fluctuation energy at the average separation the downstream, its power spectrum distribution is almost the same
point is located in the low-frequency region. Figure 12 shows the as that in the upstream, and the reattached boundary layer shows a
FIGURE 12 | (A) Frequency weighted power spectrum distribution and (B) collected positions.
FIGURE 13 | (A) WPSD contour map of the wall pressure signal (B) peak map of wall pressure signal.
process of flow recovery. This study considers that this low- The LES simulation represents a pair of streamwise vortices,
frequency motion is caused by the unsteady motion of large-scale called Görtler vortex, occupying almost the entire span width
separation bubbles. Obviously, the scale of the separation bubble is about 2δ on the compression ramp. It is found that the spanwise
larger than that of any flow in the boundary layer. skin friction coefficient distribution at the inlet of the turbulent
boundary layer not only determines the spanwise separation
location of the turbulent boundary but also makes the
4 CONCLUSION downstream Görtler vortex to be fixed at a certain spanwise
position, thus forming a time stable flow structure.
In this study, a large eddy simulation (LES) is conducted to investigate The low-frequency instability in the SWBLI is successfully
the shock wave and turbulent boundary layer interaction in a 24° captured by using the power spectrum analysis method. The
compression ramp with an inflow Mach number of Ma 2.9 and corresponding characteristic frequency is St 0.04 near the
compared with experimental and numerical simulation results in the average separation point. The spectrum analysis indicates that the
literature. The shock wave/turbulent boundary layer interaction shock wave motion is related to the pulsation of the separation
(SWBLI) was analyzed from two aspects: time domain and bubble, while in other locations of flow field, the energy of pressure
frequency domain. The main conclusions are as follows: fluctuation represents almost exclusively the high-frequency
The shock wave/turbulent boundary layer interaction represents characteristics of the fully developed turbulent boundary layer.
some flow structure characteristics. When the large scale vortex
passes the root of the shock wave, the shock wave surface wrinkles DATA AVAILABILITY STATEMENT
due to the intermittence of the large-scale vortex in the spanwise
direction. Under the strong inverse gradient pressure of the The raw data supporting the conclusions of this article will be
compression ramp, the large-scale vortex breaks into a small-scale made available by the authors, without undue reservation.
vortex, the turbulent boundary layer manifests a strong anisotropy
characteristic, and the turbulent kinetic energy rapidly increases in
the outer layer of the boundary layer downstream of the ramp. AUTHOR CONTRIBUTIONS
In the region of interaction, the separated shock wave
represents an unsteady motion along the streamwise direction. XH, L-XW, and D-DZ contributed to the conception and design
This phenomenon is successfully captured by the intermittent of the study. XH provided an idea scheme; XH, L-XW, and D-DZ
factor. The range of the shock wave motion is coded; XH analyzed; and NG and l wrote the first draft. All
−3.2δ ≤ x ≤ − 2.48δ, indicating that the unsteady shock wave authors contributed to manuscript revision and read and
motion is around the average separation point. approved the submitted version.
REFERENCES Smits, A. J., and Muck, K.-C. (1987). Experimental Study of Three Shock Wave/
Turbulent Boundary Layer Interactions. J. Fluid Mech. 182 (182), 291. doi:10.
1017/S0022112087002349
Adams, N. A. (2000). Direct Simulation of the Turbulent Boundary Layer along a Tong, F., Yu, C., and Tang, Z. (2017). Numerical Studies of Shock Wave
Compression Ramp at M = 3 and Re=1685. J. J. Fluid Mech. 420 (1), 47–83. Interactions with a Supersonic Turbulent Boundary Layer in Compression
doi:10.1017/S0022112000001257 Corner: Turning Angle Effects. J. Comput. Fluids 149, 56–69. doi:10.1016/j.
Barbe, K., Pintelon, R., and Schoukens, J. (2010). Welch Method Revisited: compfluid.2017.03.009
Nonparametric Power Spectrum Estimation via Circular Overlap. J. IEEE Touber, E., and Sandham, N. D. (2009). Large-Eddy Simulation of Low-Frequency
Trans. Signal Process. 58 (2), 553–565. doi:10.1109/TSP.2009.2031724 Unsteadiness in a Turbulent Shock-Induced Separation Bubble. J. Theor.
Bookey, P., Wyckham, C., and Smits, A. (2005). New Experimental Data of STBLI Comput. Fluid Dyn. 23 (2), 79–107. doi:10.1007/s00162-009-0103-z
at DNS/LES Accessible Reynolds Numbers. Reno: 43rd AIAA Aerospace Touber, E., and Sandham, N. D. (2011). Low-order Stochastic Modelling of Low-
Sciences Meeting and Exhibit. doi:10.2514/6.2005-309 Frequency Motions in Reflected Shockwave/boundary-Layer Interactions.
Dolling, D. S. (2001). Fifty Years of Shock-Wave/boundary-Layer Interaction J. Fluid Mech. 671, 417. doi:10.1017/S0022112010005811
Research - what Next? AIAA J. 39 (8), 1517–1531. doi:10.2514/3.14896 Wang, Bo. (2015). The Investigation into the Shock Wave/boundary-Layer
Dolling, D. S., and Or, C. T. (1985). Unsteadiness of the Shock Wave Structure in Interaction Flow Field Organization. Changsha: Graduate School of National
Attached and Separated Compression Ramp Flows. J. Experiments Fluids 3 (1), University of Defense Technology.
24–32. doi:10.1007/BF00285267 Wu, M., and Martin, M. P. (2008). Analysis of Shock Motion in Shockwave
Dupont, P., Haddad, C., and Debieve, J. F. (2006). Space and Time Organization in and Turbulent Boundary Layer Interaction Using Direct Numerical
a Shock-Induced Separated Boundary Layer. J. Fluid Mech. 559, 255. doi:10. Simulation Data. J. J. Fluid Mech. 594 (594), 71–83. doi:10.1017/
1017/S0022112006000267 S0022112007009044
Dussauge, J. P., Dupont, P., and Debiève, J. F. (2006). Unsteadiness in Shock Wave Wu, Yu., Shi-He, Yi., and Chen, Zhi. (2013). Experimental Investigations on
Boundary Layer Interactions with Separation. J. Aerospace Sci. Tech. 10 (2), Structures of Supersonic Laminar/turbulent Flow over a Compression Ramp.
85–91. doi:10.1016/j.ast.2005.09.006 J. Acta Phys. 62 (18), 308–319. doi:10.7498/aps.62.184702
Floryan, J. M. (1991). On the Görtler Instability of Boundary Layers. J. Prog. Zhong, Dong-Dong., WangXu, Li., and Ge, Ning. (2021). Reflected Shock/
Aerospace Sci. 28 (3), 235–271. doi:10.1016/0376-0421(91)90006-P Boundary Layer Interaction Structure Analysis Based on Large Eddy
Fu-Lin, Ton., Tang, Zhi-Gong., and Xin-Liang, Li. (2016). Direct Numerical Simulation. Chin. J. aeronautics 34 (5), 364–372. doi:10.1016/j.cja.2020.12.009
Simulation of Shock-Wave and Transitional Boundary Layer Interaction in Zhu, J., and Qiu, J. (2017). A New Type of Finite Volume WENO Schemes for
a Supersonic Compression Ramp. J. Chin. J. Aeronautics 37 (12), 3588–3604. Hyperbolic Conservation Laws. J. Sci. Comput. 73 (2-3), 1338–1359. doi:10.
doi:10.7527/S1000-6893.2016.0096 1007/s10915-017-0486-8
Ganapathisubramani, B., Clemens, N., and Dolling, D. (2007). Effects of Upstream Zhuang, Yi., Tan, Hui-Jun., Li, Xin., Guo, Yun-Jie., and Sheng, Fa-Jia. (2018).
Coherent Structures on Low-Frequency Motion of Shock-Induced Turbulent Evolution of Coherent Vertical Structures in a Shock Wave/turbulent
Separation. Reno: 45th AIAA Aerospace Sciences Meeting and Exhibit. doi:10. Boundary-Layer Interaction Flow, Phys. Fluids 30, 111702. doi:10.1063/1.
2514/6.2007-1141 5058278
Germano, M. (1991). A Dynamic Subgrid-Scale Eddy Viscosity Model. Phys. Fluids Zhuang, Yi., Tan, Hui-Jun., Li, Xin., Sheng, Fa-Jia., and Zhang, Yu-chao. (2018).
A 3, 1760–1765. doi:10.1063/1.857955 Görtler-like Vortices in an Impingingshock Wave/turbulent Boundary
Grilli, M., Hickel, S., and Adams, N. A. (2013). Large-Eddy Simulation of a Supersonic Layerinteraction Flow. Phys. Fluids 30, 061702. doi:10.1063/1.5034242
Turbulent Boundary Layer over a Compression–Expansion Ramp. J. Int. J. Heat
Fluid Flow 42 (8), 79–93. doi:10.1016/j.ijheatfluidflow.2012.12.006 Conflict of Interest: Author XH was employed by the company AECC HUNAN
Lee, C. B., and Wang, S. (1995). Study of the Shock Motion in a Hypersonic Shock Aviation Powerplant Research Institute.
System/turbulent Boundary Layer Interaction. Experiments in Fluids 19 (3),
143–149. doi:10.1007/BF00189702 The remaining authors declare that the research was conducted in the absence of
Loginov, M. S., Adams, N. A., and Zheltovodov, A. A. (2006). Large-eddy any commercial or financial relationships that could be construed as a potential
Simulation of Shock-Wave/turbulent-Boundary-Layer Interaction. J. J. Fluid conflict of interest.
Mech. 565 (1), 135–169. doi:10.1017/S0022112006000930
Pasquariello, V., Hickel, S., and Adams, N. A. (2017). Unsteady Effects of Strong Publisher’s Note: All claims expressed in this article are solely those of the authors
Shock-Wave/Boundary-Layer Interaction at High Reynolds Number. J. J. Fluid and do not necessarily represent those of their affiliated organizations, or those of
Mech. 823 (1), 617–657. doi:10.1017/jfm.2017.308 the publisher, the editors and the reviewers. Any product that may be evaluated in
Sasaki, K., Barros, D. C., Andre, V., and Cavalieri, G. (2021). Causality in the Shock this article, or claim that may be made by its manufacturer, is not guaranteed or
Wave/turbulent Boundary Layer Interaction. J. Phys. Rev. Fluids 6 (6), 064609. endorsed by the publisher.
doi:10.1103/PhysRevFluids.6.064609
Settles, G. S., Fitzpatrick, T. J., and Bogdonoff, S. M. (1979). Detailed Study Copyright © 2022 Huang, Wang, Zhong and Ge. This is an open-access article
of Attached and Separated Compression Corner Flowfields in High distributed under the terms of the Creative Commons Attribution License (CC BY).
Reynolds Number Supersonic Flow. AIAA J. 17 (6), 579–585. doi:10. The use, distribution or reproduction in other forums is permitted, provided the
2514/3.61180 original author(s) and the copyright owner(s) are credited and that the original
Shu, C. W., and Osher, S. (1989). Efficient Implementation of Essentially Non- publication in this journal is cited, in accordance with accepted academic practice.
oscillatory Shock-Capturing Schemes. J. J. Comput. Phys. 77 (2), 439–471. No use, distribution or reproduction is permitted which does not comply with
doi:10.1016/0021-9991(88)90177-5 these terms.