You are on page 1of 8

Chinese Science Bulletin

2009
SCIENCE IN CHINA PRESS

Springer

Effect of Mach number on transonic flow past a circular cylinder


XU ChangYue, CHEN LiWei & LU XiYun
Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, China

The effect of Mach number on transonic flow past a circular cylinder is investigated numerically for the free-stream Mach number M from 0.85 to 0.98 and the Reynolds number 2105 based on the diameter of the cylinder. The work provides an insight into several salient features, including unsteady and quasi-steady flow state, formation of local supersonic zone, and evolution of turbulent shear layer. Results show that there exist two flow states dependent of a critical Mach number Mcr around 0.9. One is an unsteady flow state characterized by moving shock waves interacting with the turbulent flow in the near region of the cylinder for M<Mcr, and the other is a quasi-steady flow state with nearly stationary shock waves formed in the near wake for M>Mcr, suppressing vortex shedding from the cylinder. Some flow behaviors in the unsteady and quasi-steady flow states are revealed. From time evolution of flow structures, local supersonic zones are identified in the wake and generated by two typical processes, i.e. reverse flow behind the cylinder and shed vortices in the near wake. The convective Mach number Mc of turbulent shear layers shed from the cylinder is identified nearly as Mc<1 in the unsteady flow regime and Mc>1 in the quasi-steady flow regime, resulting in different evolutions of the shear layers.
circular cylinder, DES, transonic flow, compressible turbulence, shock wave

The understanding of physical mechanisms in transonic flow past a body is of primary importance in applications and fundamentals. Typically, for transonic flow past a circular cylinder, there exist a variety of complex flow phenomena, such as shock wave and turbulent boundary layer interaction, shock wave and vortices interaction, local supersonic zone and shocklets, and turbulent boundary layer instability. Some work on this problem has been carried out experimentally and numerically. However, the underlying physical mechanisms are not well clear and still highly desirable to be studied. Transonic flow past a cylinder has been carried out experimentally for the Reynolds number around 105 to show the drag force measurements and flow visualizations[13]. These experimental results exhibited the transonic drag rise phenomenon and shock wave structures. Moreover, numerical simulations by solving the two-dimensional (2D) Euler equations were also per-

formed to deal with the inviscid flow past a cylinder[4,5]. The numerical solution indicated a transition of the chaotic flow to an almost stationary state for M>0.9 approximately[4]. Recently, 2D compressible Navier-Stokes equations are numerically solved using a finite volume method to investigate the unsteady forces and flow structures near the cylinder at M=0.8[6]. However, these experiments prevent them from directly acquiring detailed unsteady data in the flow field due to measurement difficulties, and the numerical simulations also have their limitation of the inviscid flow assumption or lack of the flow analysis, thus limiting our comprehension of the physical mechanisms involved. Complex shock wave and turbulence interaction ocReceived November 16, 2008; accepted January 8, 2009 doi: 10.1007/s11434-009-0325-x Corresponding author (email: xlu@ustc.edu.cn) Supported by the National Natural Science Foundation of China (Grant Nos. 90405007, 90605005) and Science and Technology Innovative Foundation of Chinese Academy of Sciences (Grant No. CXJJ-237)

Citation: Xu C Y, Chen L W, Lu X Y. Effect of Mach number on transonic flow past a circular cylinder. Chinese Sci Bull, 2009, 54: 1886-1893, doi: 10.1007/s11434009-0325-x

1 Mathematical formulation and method


1.1 Governing equations and boundary conditions To deal with transonic flow past a circular cylinder, the three-dimensional Favre-averaged compressible NavierStokes equations in generalized coordinates are employed. The equation of state for an ideal gas is used, and the molecular viscosity is assumed to obey Sutherlands law. To non-dimensionalize the equations, we use the free-stream variables including the density , temperature T and speed of sound a, and the cylinder diameter D as characteristic scales. The detailed formulations are not shown here for neatness and can be found in previous paper[11]. In this study, the initial condition is set as the free-stream quantities. The far field boundary conditions

2 Results and discussion


2.1 Simulation overview The choices of flow conditions and computational pa1887

Xu C Y et al. Chinese Science Bulletin | June 2009 | vol. 54 | no. 11

HYDROMECHANICS

curs in transonic flow past a cylinder. Recently, an experimental study of interaction between shock wave and turbulent boundary layer is performed based on the nano-based planar laser scattering method[7], and some complex flow structures are qualitatively observed. To obtain quantitative results, however, a detailed study by means of advanced numerical methods is also needed to understand the mechanisms underlying in shock wave and turbulence interaction. The effect of Mach number plays an important role on the flow behaviors for transonic flow past a cylinder. Experimental results have shown that the drag force is mainly dependent on the Mach number for a certain Reynolds number range, say about 105[13]. The previous work also exhibited that the unsteady or stationary flow state is associated with the Mach number[2,4]. Thus, the intrinsic flow mechanism relevant to the flow states is needed to be revealed. In addition, corresponding to the Reynolds number range in the previous experiments, the flow considered lies in turbulent regime. Based on extensive work by Sagauts group[810], the detached eddy simulation (DES) is a reliable way to deal with such a high Reynolds number flow. In this paper, the effect of Mach number on transonic flow past a circular cylinder is investigated numerically by means of the DES technique. The purpose is to achieve an improved understanding of some of the fundamental phenomena, including the unsteady and quasi-steady flow state, formation of local supersonic zone, and evolution of turbulent shear layer. Special attention is given to different flow behaviors in the unsteady and quasi-steady flow states.

are treated by local one-dimensional Riemann-invariants. No-slip and adiabatic conditions are applied on the cylinder surface. Periodic condition is used in the axial direction of the cylinder. 1.2 Turbulence modeling As the high Reynolds number flow is considered, the methodology used here is the DES introduced by Spalart[12]. The model is based on the Spalart-Allmaras model, which is a one-equation model for the eddy vis% cosity v by solving a transport equation, referring to the original paper for details on the constants and the quantities involved[13]. The DES is provided with a destruction term for the eddy viscosity that contains d or the distance to the % closest wall. This term adjusts the eddy viscosity v to % scale with the local deformation rate S producing an
% % eddy viscosity given by v ~ Sd 2 . Then, Spalart[12] proposed to replace the distance d to the closest wall % % with d , defined by d = min(d , CDES ), where is

the characteristic mesh length, and CDES is a constant and is taken as 0.55[10]. 1.3 Numerical method We briefly summarize the numerical approach here and refer to the given references for details. The governing equation is numerically solved by the finite-volume method[11]. The convective term is discretized using a second-order accurate upwind scheme with the Roes flux-difference splitting for capturing shock waves and a cell-centered discretization with low artificial numerical dissipation for simulating turbulent flow. When the central scheme turns on, the fourth order Jameson-type artificial dissipation is employed to prevent numerical oscillations[11]. The temporal integration is performed using an implicit approximate-factorization method with sub-iterations for keeping second-order accuracy. This numerical strategy has already been applied with success to a wide range of turbulent flows such as the compressible turbulent swirling flows injected into a coaxial dump chamber[11] and transonic flows over a cylinder[14]. We have carefully examined the physical model and numerical approach used in this study and verified that the calculated results are reliable.

ARTICLES

rameters are described. According to previous experiments[13], the free-stream Mach number M is chosen as from 0.85 to 0.98, and the Reynolds number based on the cylinder diameter as 2105. After our careful tests[14], the grid number is 5925928 in the radial, azimuthal, and spanwise direction, respectively, with the outer boundary 50D and the spanwise length 1.0D, and the time step is 0.005D/a. The grid stretches in the radial and azimuthal directions are employed to increase the grid resolutions near the wall and in the wake region behind the cylinder. It has been determined that the calculated results can be reliably obtained using the selected parameters. Given the expensive cost in the present simulations, it is reasonable to employ the high-density grid in the cross-section of the cylinder in order to capture fine flow structures and a few grids in the axial direction. In addition, based on our previous large eddy simulations of turbulent flow past a circular cylinder[15], the same grid resolution in the spanwise direction has been used to reasonably predict the nearwall flow structures, which play a significant role on the flow behaviors[16]. Moreover, the present code is equipped with a multi-block domain decomposition feature to facilitate parallel processing in a distributed computing environment. 2.2 Flow states and behaviors To reveal the effect of Mach number on flow states, time dependent lift and drag coefficients are shown in Figure 1. The curves are oscillatory variation at M=0.85 and 0.88, and become smooth with nearly constant values at M=0.95. It is also interesting to observe that the curves

exhibit intermittent change between the oscillatory and smooth variations over different periods at M=0.9. Based on the force behaviors exerted on the cylinder and the evolution of flow structures shown below, we can identify that there exist two flow states with a critical Mach number Mcr around 0.9, which agrees with the experimental data[2]. The flow states can be described as unsteady flow for M<Mcr and quasi-steady flow for M>Mcr. For neatness, the flow behaviors and underlying physical mechanisms around the critical Mach number Mcr, in particular their differences in the unsteady and quasi-steady flow states, are mainly discussed as follows. To demonstrate the flow structures in both the flow states, instantaneous flow patterns at two instants using iso-contours of the magnitude of density gradient are shown in Figure 2(a)(d) for M=0.88 and 0.95, respectively. In the unsteady flow state, e.g. M=0.88 in Figure 2(a) and (b) and the corresponding animation (not shown here), complex flow phenomena are observed and typically associated with the complex interactions of unsteady shock waves with turbulent boundary layers over the cylinder and with turbulent wake behind the cylinder. The boundary layer separation is induced by the shock waves formed over the upper and lower sides of the cylinder. The separated shear layers then become unstable and the relevant instability mechanism will be discussed below. The relatively strong -shock waves are generated behind the cylinder and interact with the shed vortices in the near wake. The flow structures in the near region of the cylinder are well

Figure 1 1888

Time-dependent lift and drag coefficients for M = 0.85 (a), 0.88 (b), 0.9 (c) and 0.95 (d). www.scichina.com | csb.scichina.com | www.springerlink.com

Figure 2 Instantaneous flow structures using the magnitude of density gradient in a cross-section and the mean flow patterns for M = 0.88 ((a), (b), (e)) and 0.95 ((c), (d), (f)). The time interval of (a) and (b) as well as of (c) and (d) is 1.0D/a. The shocklet and streamlines are plotted in the inset of (c). Here, the lines denote the iso-lines of Mach number with solid lines M>1 and dashed lines M<1, and the arrow-lines represent the streamlines.

consistent with experimental visualizations[3,7]. In the quasi-steady flow state, e.g. M=0.95 in Figure 2(c) and (d), relatively strong shear layers with higher convective Mach number shown below are formed from the upper and lower sides of the cylinder, suppressing the recirculation region behind the cylinder. Meanwhile, both strong oblique shock waves are formed in the near wake. Even though local unsteady flow phenomena occur, the force curves still exhibit smooth variation in Figure 1(d), since the forces exerted on a body are mainly determined by the near-wall flow structures[16]. The iso-lines of Mach number and streamlines based on the mean flow quantities obtained by both the axial space- and time-average are shown in Figure 2(e) and (f) for M=0.88 and 0.95, respectively. The oblique shock waves are well observed by the dense iso-lines of Mach number for M=0.95 and the streamlines reasonably change their direction behind the shock waves following

the Rankine-Hugoniot relations. From the streamlines pattern, the recirculation region looks like a wedgeshaped afterbody behind the cylinder, suppressing the large-scale vortex shedding from the cylinder and resulting in the quasi-steady flow behaviors. Moreover, as shown in Figure 2(e) for M=0.88, extensive subsonic flow region exists in the near wake, where complex flow structures evolve and behave as the shed vortices and shock waves interaction. Time-averaged drag coefficient (C D ) is shown in Figure 3(a). A peak value of CD occurs around the critical Mach number Mcr, consistent with the transonic drag rise phenomenon observed experimentally[2]. The relevant reason has never been given. Here, we can get the understanding of this phenomenon based on the flow structures shown above and the mean separation location in Figure 3(b) varying from = 85 approximately in the unsteady flow regime to = 106 in the quasi-steady
1889

Xu C Y et al. Chinese Science Bulletin | June 2009 | vol. 54 | no. 11

HYDROMECHANICS

ARTICLES

Figure 3 (a) Time-averaged drag force. Here, denotes the present calculated results, --- flight test data , experimental data , [2] and experimental data . (b) Mean separation location. Here, increases from the front-point of the cylinder in the clockwise direction.

[17]

[1]

flow regime. In the unsteady flow regime, the pressure drag due to the recirculation region behind the cylinder in Figure 2(e) contributes major part and the shock wave drag due to the entropy rise increases with M. In the quasi-steady flow regime, the pressure drag decreases as the recirculation zone becomes smaller in Figure 2(f), and the shock waves move to the downstream of the cylinder to decrease the shock wave drag. In addition, the data obtained by experiments[1,2] and flight test[17] are also shown in Figure 3(a). The present calculated results agree well with the flight test data, and discrepancies compared with the experimental data occur; it may be due to the wind tunnel blockage effect. 2.3 Formation of local supersonic zones In the transonic flow around a body, some typical flow phenomena, such as local supersonic zone (LSZ) and small shock or shocklet, are still highly desired to be studied in the understanding of the relevant mechanisms. After we carefully examine the flow evolution for both the flow states, it is revealed that, for the first time, two typical processes, i.e. reverse flow behind the cylinder and shed vortices in the near wake, are associated with

the formation of the LSZs. To exhibit the formation of LSZs induced by the reverse flow behind the cylinder, instantaneous flow pattern using the iso-lines of Mach number and streamlines is shown in Figure 4(a) for M=0.88, lying in the unsteady flow regime. The streamlines which are nearly along the shear layer turn into the recirculation region behind the cylinder. The reverse streamlines exhibit divergent distribution and the flow in the divergent tube between the streamlines is accelerated due to expansion effect to generate an LSZ located at A in Figure 4(a). The streamlines then change their direction again and return into the downstream. We further follow this cluster of streamlines to observe the global evolution of the LSZs. The streamlines become convergent when they leave the recirculation region and cross the other two LSZs located at B and C. After our carefully examining the evolution of LSZs, we identify that the LSZ at B is generated by the reverse flow as described above and convects downstream from the recirculation region, and the LSZ at C is induced by the shed vortex from the cylinder, which will be discussed later.

Figure 4 Instantaneous iso-lines of Mach number (solid and dashed lines) and streamlines (arrow-lines) in a cross-section for M=0.88 (a) and 0.95 (b). The -shocklet and streamlines are plotted in the inset of (a), and the enlarged view of the LSZ in the inset of (b). Here, solid lines represent M>1 and dashed lines M<1.

1890

www.scichina.com | csb.scichina.com | www.springerlink.com

Figure 5 Time evolution of flow structures using the iso-lines of Mach number (solid and dashed lines) and vorticity contours (grey flood) in a cross-section for M=0.88. Each time interval from (a) to (d) is 0.5D/a. Here, solid lines represent M>1 and dashed lines M<1; white region is positive vorticity value and black region negative value.

The subsonic flow state occurs in the recirculation region behind the cylinder in Figure 2(f) for M=0.95, lying in the quasi-steady flow regime. Correspondingly, as shown in Figure 4(b) for the instantaneous flow pattern, the LSZs are detected and, similar to the process in Figure 4(a), induced by the reverse flow. Based on the mean flow pattern in Figure 2(f), it is reasonably identified that the evolution of LSZs is limited in the recirculation region and the LSZs disappear within some elapsed time. The formation of LSZs by the shed vortices in the near wake, which only occurs in the unsteady flow regime, is discussed further. The vortices shed from the cylinder interact with the shock waves, and the relevant flow phenomena and behaviors obtained are well consistent with the findings on the vortices-shock wave interaction[18]. Here, we mainly pay attention to the evolution of the LSZs. Time sequence of flow structures using the iso-lines of Mach number and vorticity contours is shown in Figure 5 for M=0.88. As marked by a circle in Figure 5(a), when the positive vortex in a clockwise rotation sheds downstream to cross the shock wave, the supersonic flow state occurs behind the distorted shock induced by the shed vortex. In the following sequences, as the vortex convects downstream, represented by an arrow in Figure 5, the LSZ induced by the vortex

separates from the distorted shock wave and convects downstream accompanying the vortex. Then, the LSZ becomes smaller and smaller and disappears in the downstream. We further turn to the existence of shocklets which are associated with the LSZs, as shown in Figure 4(a). Usually, the region with strong negative dilatation is referred to as shocklets[19,20]. Additional limitation by utilizing the locus of the zero crossing of 2 is also considered to determine the shocklet location[21]. By means of the above approach to represent the shocklets in Figure 4(a), we have identified the shocklets related to the LSZs at A and B, and confirmed that the change of the flow state across the shocklets satisfies the Rankine-Hugoniot conditions. It is interesting to observe that, as clearly shown in the inset of Figure 4(a), -shocklet is detected in the LSZ at A. In addition, we have not observed the existence of shocklet in the LSZ at C induced by the shed vortices. It means that the flow going out of the LSZ can smoothly change to subsonic flow. 2.4 Evolution of turbulent shear layer The existence of compressible shear layers shed from the cylinder has been exhibited in Figure 2(a)(d). The understanding of the shear layer instability has been
1891

Xu C Y et al. Chinese Science Bulletin | June 2009 | vol. 54 | no. 11

HYDROMECHANICS

ARTICLES

Figure 6

(a) Power spectral density (PSD) of the pressure at probe P in Figure 2(a). (b) Convective Mach number Mc of the shear layers.

widely studied[10]. Here, we mainly reveal some different behaviors of turbulent shear layer in the unsteady and quasi-steady flow states. To demonstrate the shear layer lying in turbulent regime, Figure 6(a) shows the power spectral density of pressure at probe P in Figure 2(a). The spectrum slopes of 1 and 5/3 with respect to a characteristic Strouhal number, defined as St=fD/U, are observed. An intermediate frequency range with St1 scaling occurs, representing the contribution of eddies to the pressure near the wall[22]. An approximate St5/3 scaling in the high frequency range is illustrated; it is also verified that the present calculation has reliably simulated the turbulent flow in the inertial subrange. As shown in Figure 2(a)(d) for the developing shear layers, it is of primary interest to investigate the convective Mach number Mc, which is closely associated with the instability behaviors[10,19]. By means of the treatment[10], the distributions of the convective Mach number Mc are given in Figure 6(b). We have identified nearly as Mc<1 in the shear layer developing period for the unsteady flow state and Mc>1 for the quasi-steady flow state. The Mc varies smoothly for M=0.88, and increases to a maximum at x/D=0.65 approximately and then decreases gradually for M=0.95. Based on the convective Mach number, the evolution of turbulent shear layer has different behaviors. As exhibited in Figure 2(a) and (b), the spanwise eddies originating from the Kelvin-Helmholtz instability can clearly be observed in the shear layer growth for relatively low Mc cases. Then, the detached shear layer interacts with the -shock waves formed in the near wake. Moreover, upstream traveling compression waves emanating from the supersonic flow side of shear layer are demonstrated and accumulate to form a shock wave over the cylinder. With the increase of M, as typically shown
1892

in Figure 2(c) and (d) for M=0.95, the shear layers with Mc>1 are generated from the upper and lower sides of the cylinder. As indicated by Simon et al.[10], oblique modes dominate the shear layer instability process. When Mc>1, the compressibility effects are strong. Thus, as shown in the inset of Figure 2(c), we also identify the existence of shocklets, which can be reasonably detected using the approach described above. The shocklets convect downstream with the their feet along the supersonic flow side of shear layer and merge into the compression shocks formed in the near wake. Similar phenomenon was also observed in the boundary layer along a supersonic compression ramp[23] and compressible mixing layer past an axisymmetric trailing edge[10].

3 Conclusions
The effect of Mach number on transonic flow past a circular cylinder was numerically studied by means of a detached eddy simulation technique. Various fundamental mechanisms dictating the flow evolution were examined systematically for a wide range of the freestream Mach number M from 0.85 to 0.98 with a Reynolds number 2105, in the light of both mean and instantaneous flow data. Two typical flow states, i.e. unsteady and quasi-steady flow state with a critical Mach number Mcr around 0.9, are identified in the M range considered. The unsteady flow state is characterized by moving shock waves interacting with turbulent boundary layer over the cylinder and with the vortices shed from the cylinder in the near wake for M<Mcr. The quasi-steady flow state demonstrates that relatively strong oblique shock waves are formed in the near wake and vortex shedding from the cylinder is obviously suppressed for M>Mcr. Based on the mean flow data, the mean separation location changes from =85 approximately in the unsteady flow regime to 106 in the

www.scichina.com | csb.scichina.com | www.springerlink.com

quasi-steady flow regime. The time-averaged drag coefficient exhibits a peak value around Mcr, consistent with the transonic drag rise phenomenon. Some fundamental flow behaviors, such as LSZs, shocklets, and turbulent shear layer, are discussed, in particular their differences for both the unsteady and quasi-steady flow states. The formation of LSZs is closely associated with two typical processes, i.e. reverse flow behind the cylinder and shed vortices in the near wake. In the recirculation region behind the cylinder, the LSZs are induced by the reverse flow. In the unsteady flow regime, the LSZs may convect downstream and are associated with the existence of shocklets. Correspondingly, in the quasi-steady flow regime, the evolution of LSZs is limited in the recirculation region.
1 2 3 4 5 6 Macha J M. Drag of circular cylinders at transonic mach numbers. J Aircraft, 1977, 14: 605607[DOI] Murthy V S, Rose W C. Detailed measurements on a circular cylinder in cross flow. AIAA J, 1978, 16: 549550[DOI] Rodriguez O. The circular cylinder in subsonic and transonic flow. AIAA J, 1984, 22: 17131718[DOI] Botta N. The inviscid transonic flow about a cylinder. J Fluid Mech, 1995, 301: 225250[DOI] Pandolfi M, Larocca F. Transonic flow about circular cylinder. Comput Fluids, 1989, 17: 205220[DOI] Miserda R F B, Leal R G. Numerical simulation of the unsteady aerodynamic forces over a circular cylinder in transonic flow. AIAA Paper 2006-1408, 2006 7 Zhao Y X, Yi S H, He L, et al. The experimental study of interaction between shock wave and turbulence. Chinese Sci Bull, 2007, 52: 12971301[DOI] 8 9 10 Mary I, Sagaut P. Large eddy simulation of flow around an airfoil near stall. AIAA J, 2002, 40: 11391145[DOI] Sagaut P, Deck S, Terracol M. Multiscale and Multiresolution Approaches in Turbulence. London: Imperial College Press, 2006 Simon F, Deck S, Guillen P, et al. Numerical simulation of the compressible mixing layer past an axisymmetric trailing edge. J Fluid Mech, 2007, 591: 215253 11 Lu X Y, Wang S W, Sung H G, et al. Large eddy simulations of turbulent swirling flows injected into a dump chamber. J Fluid Mech, 2005, 527: 171195[DOI] 12 Spalart P R, Jou W H, Strelets M, et al. Comments on the feasibility of LES for wings and on a hybrid RANS/LES approach. In: Proceedings of 1st AFSOR International Conference on DNS/LES, Ruston, 1997

Moreover, the large-scale vortex shedding occurs in the unsteady flow state and induces the LSZs. Then, the LSZs convect downstream accompanying with the shed vortices and disappear gradually in the downstream. The existence of turbulent shear layers shed from the cylinder has been exhibited. The convective Mach number Mc of turbulent shear layers shed from the cylinder is identified nearly as Mc<1 in the unsteady flow regime and Mc>1 in the quasi-steady flow regime. The Kelvin-Helmholtz instability is observed in the shear layer growth for relatively low Mc cases. The oblique modes dominate the shear layer instability process for Mc>1. The shocklets also occur and convect downstream with their feet along the supersonic flow side of shear layer in the quasi-steady flow state.
13 14 Spalart P R, Allmaras S R. A one equation turbulence model for aerodynamic flows. AIAA Paper 1992-0439, 1992 Xu C Y, Chen L W, Lu X Y. Numerical investigation of shock wave and turbulence interaction over a circular cylinder. Mod Phys Lett B, 2009, 23: 317320[DOI] 15 Lu X Y, Dalton C, Zhang J. Application of large eddy simulation to an oscillating flow past a circular cylinder. J Fluids Eng, 1997, 119: 519525[DOI] 16 17 Wu J Z, Lu X Y, Zhuang L X. Integral force acting on a body due to local flow structures. J Fluid Mech, 2007, 576: 265286[DOI] Welsh C J. The drag of finite length cylinders determined from flight tests at high Reynolds numbers for a Mach number range from 0.5 to 1.3. NACA TN, 1953 18 19 Zhang S H, Zhang Y T, Shu C W. Multistage interaction of shock Freund J B, Lele S K, Moin P. Compressibility effects in a turbulent annular mixing layer I. Turbulence and growth rate. J Fluid Mech, 2000, 421: 229267[DOI] 20 Vreman B, Kuerten H, Geurts B. Shocks in direct numerical simulation of the confined three-dimensional mixing layer. Phys Fluids, 1995, 7: 21052107[DOI] 21 Samtaney R, Pulin D I, Kosovi B. Direct numerical simulation of decaying compressible turbulence and shocklet statistics. Phys Fluids, 2001, 13: 14151430[DOI] 22 Na Y, Moin P. The structure of wall-pressure fluctuations in turbulent boundary layers with adverse pressure gradient and separation. J Fluid Mech, 1998, 377: 347373[DOI] 23 Loginov M S, Adams N A, Zheltovodov A A. Large-eddy simulation of shock-wave/turbulent-boundary-layer interaction. J Fluid Mech, 2006, 565: 135169[DOI] wave and a strong vortex. Phys Fluids, 2005, 17: 116101[DOI]

Xu C Y et al. Chinese Science Bulletin | June 2009 | vol. 54 | no. 11

1893

HYDROMECHANICS

ARTICLES

You might also like