We perform direct numerical simulation of flow past the NACA (National Advisory Committee for Aeronautics) 0012 airfoil with partially covered wavy roughness elements near the leading edge. The Reynolds number Rec based on the freestream velocity (U0) and airfoil chord length (C), and the angle of attack (AoA) is fixed, i.e., Rec=5×104,AoA=10°. The leading edge roughness elements are characterized by streamwise sinusoidal-wave geometry that is uniform in the spanwise direction with fixed height, whereas varying wave numbers (k) from 0 to 12. Based on the validation of the smooth baseline case (k =0), the roughness effects on the aerodynamic performance are evaluated in terms of the lift and drag coefficients. The drag coefficient decreases monotonically with k, while the variation of the lift coefficient with k is similar to the “drag crisis” phenomenon observed in cylinder flows. The sharp variations of lift and drag coefficients from k =6 to 8 indicate that k =8 is a critical case, beyond which massive separation occurs and almost covers the airfoil's suction side and dominates the airfoil aerodynamic performance. To further reveal the underlying mechanism, the flow structures, pressure, skin friction coefficients, shear layer transition onset, and boundary layer separation are quantified to investigate the roughness effects. The roughness elements strongly modify the separation behavior, whereas they have little effect on the transition onset. The unsteady interactions and convections of separation bubbles downward the trailing edge also change the wake evolution. Based on the scaling of the asymmetric wake profiles, we find that the wake defect is gently decreasing with k, but the increase in the wake width is much stronger. This scaling analysis gains a better insight into the roughness effects on the momentum deficit flow rate, which confirms the drag mechanism with different roughness wave numbers.

Flow over rough surfaces occurs in a wide range of engineering applications, such as piping systems, marine vehicles, and urban atmospheric boundary layers.1,2 The roughness of these surfaces can significantly affect the boundary layer development and, thus, influence the surface drag, heat, and momentum transfer, etc.3 The pioneering work of quantifying the roughness effect in pipe flows was led by Nikuradse,4 and Moody5 proposed the Moody diagram that characterized the friction factor as a function of Reynolds number and relative roughness. As the Reynolds number increases, three distinct flow regimes were identified: a laminar flow regime where roughness has a negligible effect on the friction, a transitional regime, and a fully turbulent regime where roughness becomes a significant factor once it exceeds a critical size. In the past decades, many investigations have been devoted to the roughness effects in canonical wall-bounded flows in simple geometries, for both laminar6,7 and turbulent regimes.8,9

The flow past bluff bodies is characterized by laminar-turbulent transition and boundary layer separation, and hence, often, to non-equilibrium effects (curvature and pressure gradient). Thus, the roughness effects on the flow behavior would be more complex. In fact, the roughness elements on the bluff body modify the surrounding flow structure, transition scenario, and separation behavior, affecting the aerodynamic performance10 and yielding significant departures from Nikuradse's initial proposal. Cheng et al.11 performed a wall-resolved large-eddy simulation (LES) of flow past a grooved cylinder up to transcritical Reynolds numbers, with azimuthal sinusoidal-shaped groove height ranging from D/16 to D/640 (D is the diameter of the cylinder). The authors found that the pressure and skin friction coefficients and the near-wall field are significantly affected by the groove, with locally trapped separation bubbles formed as cavity flows within grooves. Thus, it is found that the drag coefficient can be increased or reduced depending on both roughness height and flow regime. Zhou et al.12 experimentally investigated rectangular-shaped grooved cylinders in the subcritical regime with a fixed groove height of 0.05D and found that the roughness can reduce the drag coefficient and narrow the wake. Different types of roughness elements, such as dimples13 and irregular ice14 partially or entirely placed on the cylinder, have been investigated, which show different effects on the aerodynamic performance case by case.

Unlike the cylinder with constant curvature, the airfoil geometry may be considered a combination of large curvature (near the front stagnation part) and small curvature near the trailing edge portion of the airfoil, which will no doubt result in rich flow physics with roughness elements. The study of airfoils with roughness in different configurations is motivated by the analysis of surface imperfections due to manufacturing (gaps between metal plates, rivets, etc.) or ice accretions, which form with various shapes near the leading edge. The latter is typical when an aircraft enters a region containing supercooled water droplets under different meteorological and flight conditions.15 Based on representative ice geometries, Bragg et al.16 qualitatively classified ice accretions into four major categories: roughness ice, horn ice, streamwise ice, and spanwise ridge ice. They also observed that many ice shapes are complex, which do not purely fall into a single category but may have features representative of two or more. Most experimental investigations have been conducted, mainly quantifying aerodynamic performance degradation with ice accretions. The general consensus is that ice roughness, regardless of the ice roughness shape, results in increased drag and reduction in lift.17 However, there are minimal numerical investigations of the impact of accretions on airfoils (refer to the review paper by Stebbins et al.18). The main reason is that the ice shapes can vary widely and result in complex three-dimensional surfaces with intricate geometries, which makes the high-quality mesh generation challenging. Consequently, the mesh resolution would also dramatically increase, and the necessity for more computational resources would also increase. Meshing all the detailed roughness features is generally prohibitive for practical computational studies, especially for DNS, which resolves a full range of turbulent length scales. In industrial simulations, most calculations of the airfoil flow are performed by solving the Reynolds-Averaged Navier–Stokes (RANS) equations with turbulence models.18,19 In those settings, the near-wall grid size is of order 0.010.02C (C is the chord length), which could not even resolve small-scale geometric characteristics.20 

Serson et al.21 investigated the effect of spanwise waviness elements on the leading edge of the NACA0012 airfoil with DNS and found that the wavy roughness elements may decrease or increase the lift coefficient, depending on the Reynolds number and flow regime (pre-stall or post-stall). Kumar et al.20 studied effects of leading-edge roughness designed to mimic ice accretion with LES, and the ice accretion is modeled with cylindrical (uniform in the spanwise direction) and randomly distributed bump-like roughness elements. They found that the roughness elements can accelerate the transition to turbulence and produce different 3D structure development. Kim et al.22 investigated the spanwise wavy roughness effect on the airfoil–turbulence interaction by solving the compressible Euler equations and found that the wavy roughness can reduce the surface pressure fluctuation. Thomas et al.23 investigated the development of Tollmien–Schlichting (TS) wave instabilities with the effect of streamwise sinusoidal surface waviness elements and found that long-wave roughness can stabilize TS disturbance. Goodhand et al.24 investigated the impact of streamwise geometric variation near the leading edge and found that even small-scale variation can modify the separation and reattachment locations drastically and, thus, affect the aerodynamic performance of the airfoil, which is in accordance with the Stebbins et al.,18 i.e., even smallest roughness elements could cause noticeable reductions in aerodynamic capabilities.

Despite many studies on this topic, few have attempted to describe the flow in detail, explaining the mechanisms responsible for the changes in the aerodynamic forces caused by the roughness elements. In the present research, we are focusing on the effect of surface roughness elements on the aerodynamic performance of low Reynolds number airfoil as well as the underlying mechanism. To that end, we perform high-resolution DNS of the NACA0012 airfoil near stall condition, with streamwise wavy roughness elements located near the leading edge. The length of the roughened region and the height of the roughness elements are fixed while varying the wavelength. The Reynolds number is set at Rec=5×104, providing a good balance between the tractability of the underlying mechanisms and the computational cost of the highly resolved simulations.

The paper is organized as follows. The computational details, including the flow configuration, numerical scheme, and boundary conditions, are given in Sec. II. The DNS code is validated in Sec. III. The results are presented and analyzed in Sec. IV, including the lift and drag coefficients, pressure and skin friction coefficients, mean flow field, unsteady separation bubble development, and scaling of wake dynamics. Finally, some conclusions are drawn in Sec. V.

The artificial wavy roughness elements are placed near the leading edge on the suction side, as shown in Fig. 1. The tuple (s0,yn) denote the wall-tangential and wall-normal coordinates to the smooth airfoil surface, with the origin located at x/C=0.037 and the arc length from the leading edge to the origin is 0.05Lu, where Lu is the length of the upper airfoil surface. These roughness elements are characterized by the sinusoidal-wave structure as follows:

(1)

where h and λ are the roughness elements' amplitude (height) and wavelength, respectively. The roughened region starts from x/C=0.037 and ends at x/C=0.136, occupying 10% of the upper surface length (i.e., 0.1Lu). The roughness height is set to h=2×103C, and the roughness wave number (RWN) is k=0.1Lu/λ.

FIG. 1.

Sketch of the NACA0012 airfoil with artificial wavy roughness elements.

FIG. 1.

Sketch of the NACA0012 airfoil with artificial wavy roughness elements.

Close modal

The location of the roughened region follows the experimental work of Fujino et al.25 which is also adopted by Bagade et al.26 This choice is prompted by the fact that the most common surface roughness, such as ice accretion, occurs in this region, which is also susceptible to many important flow phenomena, such as dynamic stall. This small-scale roughness height is larger than Fujino et al.25 but slightly smaller than the cylindrical element (h=3×103C) as adopted by Kumar et al.20 The angle of attack (AoA) and Reynolds number are fixed, i.e., AoA=10° and Rec=5×104. This Reynolds number is smaller than the realistic (high speed) flight (Rec106); however, it is applicable to the small-scale autonomous aircraft, wind turbines, etc.27 The angle of attack considered here is close to the stall condition, where the aerodynamic performance can be strongly modified by roughness elements. We perform DNS of six configurations, i.e., k=0,4,6,8,10,and12, to investigate the effects of different roughness geometries, in which k =0 corresponds to the smooth case.

The incompressible Navier–Stokes equations in the generalized curvilinear coordinates read

(2)

where (ξ1,ξ2,ξ3)=(ξ,η,ζ) denote the generalized curvilinear coordinates, and Um and Fim are given by

(3)

The triplet (x1,x2,x3)=(x,y,z) denote the Cartesian coordinates with corresponding velocity components denoted as (u1,u2,u3)=(u,v,w). The symbol p indicates the pressure, ν is the kinematic viscosity, J1 is the Jacobian of the transformation, Um is the volume flux normal to the surface of constant ξm, and Gmn is the mesh skewness tensor. It should be noted that the ξ3 direction is congruent with the z direction since the spanwise geometry is uniform in the present research.

The computational domain is discretized in hexahedral elements using a C-type collocated body-fitted mesh. The governing equations given by (2) are spatially discretized using the energy-conservative fourth-order finite difference scheme proposed by Morinishi et al.28 Specifically, the convective term is discretized in the skew-symmetric form to minimize the aliasing error.29 The pressure–velocity coupling is achieved by the semi-implicit fractional step method,30,31 which follows the predictor-corrector procedure. The solution is advanced in time by the Adams–Bashforth method for the explicit terms and the Crank–Nicolson scheme for the implicit terms. The incompressibility condition is enforced by solving a Poisson-type equation for the pressure, p. Both the modified Helmholtz equations resulting from the implicit treatment of viscous terms in the predictor step and the pressure Poisson equation are solved using the multigrid solver with line-relaxed Gauss–Seidel smoothers. The DNS code has been parallelized using the standard MPI protocol. To achieve optimal load balancing, the meshes are divided into blocks of equal size, and each of them is assigned to a unique processor. The in-house developed DNS code is described in detail by Zhang et al.32 for DNS of airfoil flows.

The numerical setup and domain size are illustrated in Fig. 2. The computational domain size should be large enough to minimize the effects of the artificially imposed boundary conditions. In the present simulations, the domain size (wake length W=10C and domain radius R=6C) is identical to the choice of Zhang and Samtaney33 for DNS of flow past the NACA0018 airfoil at Rec=5×104. We note that a sufficiently long spanwise domain size (Lz) is important to resolve the largest turbulent structures in the spanwise direction. Here, we use Lz=0.8C, as recommended by Zhang and Samtaney,33 which is larger than most of the DNS and LES simulations of an airfoil flow with similar Rec.34–37 In addition, we impose a uniform flow (u,v,w)=(U0,0,0),U0=1 at the inlet, the convective boundary condition u/t+UBu/x=0 on the outflow plane, where UB is the bulk mean velocity, and a no-slip wall boundary condition on the airfoil surface. Finally, we use periodic boundary condition in the spanwise direction.

FIG. 2.

Sketch of the numerical setup and computational domain.

FIG. 2.

Sketch of the numerical setup and computational domain.

Close modal

All the simulations are performed with a mesh Nξ×Nη×Nz=2048×256×256. The time step is restricted by the advective Courant–Friedrichs–Lewy (CFL) condition. Here, the CFL number is set to CFL=0.8. The mesh is clustered around the airfoil surface, and the maximum wall unit Δη+ is around 0.9 (based on the local shear velocity and first wall-normal grid size) to ensure that the near-wall region is fully resolved. The maximum wall units in ξ and z directions, i.e., Δξ+ and Δz+ are less than 10 in the turbulent region. The mesh resolution has been carefully verified in Refs. 33 and 38 and follows the well accepted criterion for DNS of wall-bounded turbulent flows.39 The total simulation time is approximately 60C/U0, and only the last 45C/U0 time units are collected for statistical analysis. All the simulations are performed on the Cray XC40 Shaheen II supercomputer at KAUST. The total computational cost is around 15 million core hours.

This section presents several statistically averaged quantities to validate the DNS code. All these quantities are averaged in time and spanwise direction. The latter average procedure is used throughout the paper unless stated otherwise.

Figure 3 shows the distribution of time- and spanwise-averaged pressure coefficient Cp for the smooth airfoil case (k =0). Here, Cp is computed as

(4)

where pw is the time- and spanwise-averaged wall pressure, and p is the reference pressure taken at the outlet boundary. It can be found that the DNS results compare well with the experimental results from Yamaguchi et al.40 However, there is some discrepancy near the pressure plateau, which is usually used to identify the transition onset and separation point in experiments.41 

FIG. 3.

Distribution of the pressure coefficient, Cp, around the smooth, k =0, NACA0012 airfoil. ○, experimental results from Yamaguchi et al.;40 ———, present DNS results.

FIG. 3.

Distribution of the pressure coefficient, Cp, around the smooth, k =0, NACA0012 airfoil. ○, experimental results from Yamaguchi et al.;40 ———, present DNS results.

Close modal

Figure 4 shows the mean and root mean square (RMS) velocity profiles along the suction side of the smooth airfoil. The DNS predicted mean velocity profiles (u¯/U0) are in good agreement with the experimental results from Ohtake,42 and the reverse flow near the leading edge, i.e., x/C=0.1 is well captured. For the RMS velocity profiles (u¯/U0), the near-wall gradient of u¯/U0 of the present DNS is slightly smaller than the experimental results within the region x/C[0.4,0.9] but compare well in the outer region. Overall, the DNS predicted results generally agree with the experimental results.

FIG. 4.

(a) Mean velocity profiles, u¯/U0, and (b) RMS velocity profiles, u¯/U0, along the suction side of the smooth, k =0, NACA0012 airfoil. ○, experimental results from Ohtake;42 ———, present DNS results.

FIG. 4.

(a) Mean velocity profiles, u¯/U0, and (b) RMS velocity profiles, u¯/U0, along the suction side of the smooth, k =0, NACA0012 airfoil. ○, experimental results from Ohtake;42 ———, present DNS results.

Close modal

In this section, we investigate the roughness effect on the aerodynamic performance of the airfoil, (i.e., the lift and drag coefficients) and the flow field near the body.

The sectional lift and drag coefficients are computed as

(5)

where the lift and drag forces are computed by integrating the wall pressure and shear stress over the airfoil surface.

Figure 5 shows the mean and RMS values of the lift and drag coefficients for different roughness wave numbers. In what follows, ΔCL denotes the decrement of the lift coefficient and the superscript indicates the roughness wave number, k. The lift coefficient, CL, is decreasing from k =0 to k =8, with a smooth variation from k =0 to k =6 (ΔCL/CL0=1%) but a sharp decrease from k =6 to k =8 (ΔCL/CL6=9.8%). At k =8, CL reaches its minimum value, beyond which we observe only small variations. The correlation of CL with k is similar to the drag crisis phenomenon, as observed in bluff body flows, in which the drag coefficient falls suddenly as the Reynolds number increases beyond the sub-critical regime.43–45 The drag crisis phenomenon is attributed to either transition or dynamics of separation bubbles.45 We will further discuss the behavior of CL in Secs. IV B and IV C.

FIG. 5.

Mean and RMS of (a) the lift coefficient, CL and CL,rms and (b) the drag coefficient, CD and CD,rms, for different roughness wave numbers, k.

FIG. 5.

Mean and RMS of (a) the lift coefficient, CL and CL,rms and (b) the drag coefficient, CD and CD,rms, for different roughness wave numbers, k.

Close modal

The drag coefficient, CD, shows a monotonic correlation with k, which indicates that a larger roughness wave number makes the airfoil's surface rougher and increases the drag. Furthermore, the variation of CD with k is much stronger than that of the lift coefficient, CL. In fact, from k =0 to k =6 we observe that ΔCD/CD0=21.1%. Finally, a sharp increase from k =6 to k =8 is observed (ΔCD/CD6=85.1%), similar to the lift coefficient.

The RMS of the lift and drag coefficients, CL,rms and CD,rms, show similar trend with k. For k8, they grow with k with much stronger variations, i.e., CL,rms6/CL,rms0=4.3 and CD,rms6/CD,rms0=11.7. A sharp increase from k =6 to k =8 is also observed for both CL,rms and CD,rms, i.e., CL,rms8/CL,rms6=4.1 and CD,rms8/CD,rms6=2.1. For k8, the variations in the mean and RMS of lift and drag coefficients are not large.

Figure 6 shows the contribution ratio of the pressure to the lift and drag coefficients, for different roughness wave numbers. We note that the pressure contributes more than 99% to the lift coefficient while the contribution ratio of pressure to the drag coefficient is strongly increasing with k, from 82.9% at k =0 to 96% at k =12, indicating that the form drag is dominant for all k cases. Therefore, it is important to examine the distributions of pressure and skin friction coefficients, since the lift and drag coefficients are based on the integration of them around the airfoil. We emphasize that although the contribution of the skin friction coefficient is small (see Fig. 6), the skin friction behavior remains quite important since it is the principal indication of near-wall dynamics including near-wall separation and reattachment.

FIG. 6.

Contribution ratio of the pressure to the lift, CL, and drag coefficients, CD, for different roughness wave numbers, k.

FIG. 6.

Contribution ratio of the pressure to the lift, CL, and drag coefficients, CD, for different roughness wave numbers, k.

Close modal

Based on the above-mentioned analysis, we plot the distributions of the pressure and skin friction coefficients in Fig. 7. The pressure coefficient, Cp, is computed from Eq. (4) and the skin friction coefficient, Cf, is calculated as

(6)

where τw,s is the time- and spanwise-averaged streamwise (parallel to the airfoil surface) wall shear stress, computed from the wall-normal differentiation using the third-order accuracy one-sided velocity-derivative method.46 

FIG. 7.

(a) Distribution of the pressure coefficient, Cp around the airfoil and (b) the skin friction coefficient, Cf, on the suction surface for different roughness wave numbers, k. The shaded region denotes the roughened region. The horizontal dashed line Cf = 0 in (b) is shown for convenience: zero crossing of this line indicates separation and reattachment.

FIG. 7.

(a) Distribution of the pressure coefficient, Cp around the airfoil and (b) the skin friction coefficient, Cf, on the suction surface for different roughness wave numbers, k. The shaded region denotes the roughened region. The horizontal dashed line Cf = 0 in (b) is shown for convenience: zero crossing of this line indicates separation and reattachment.

Close modal

From the Cp profile shown in Fig. 7(a), we observe that the magnitude of the pressure coefficient decreases with increasing roughness wavenumber, k, of the roughened region. The sudden change in Cp from k =6 to k =8 is in accordance with the lift and drag coefficient behaviors as shown in Fig. 5. Downstream the roughened region, the recovery of Cp for k6 is much faster and stronger than k8, which produces larger adverse pressure gradients for k6 except for the zone near the trailing edge, i.e., x/C>0.9. For x/C>0.3, the Cp-profiles for k6 and k8 cases seem collapse on two different curves, respectively. Moreover, the Cp profile from k =8 is slightly larger than k =10 and 12. Not surprisingly, these differences in the coefficient are not obvious on the pressure side, except for x/C>0.9, which would affect the wake development downstream of the trailing edge (see Sec. IV D for more discussions).

The locations of separation and reattachment can be detected from the plot of the friction coefficient, Cf, as shown in Fig. 7(b). We plot the time- and spanwise-averaged streamlines to further visualize the separation bubbles in Fig. 8. For the smooth airfoil case (i.e., k =0), a small separation bubble is observed near the leading edge, also visible in Fig. 8(a) and in agreement with observations by Eljack et al.47 (compressible DNS with Mach number of 0.4). With the increase in the roughness wave number (k6), the reattachment location is moving downstream, and the primary separation bubble is growing (see Fig. 8). More secondary separation bubbles are formed around the roughened region with increasing k. From k =6 to k =8, the primary separation bubble is suddenly enlarged, and massive separation occurs on the suction side beyond k =8, covering approximately 99% of the suction surface. For k =8, there are two bubble centers detected from the closed streamlines, beyond which the bubbles are relaxed, and one primary bubble is formed for k =10 and 12, with the bubble center moving downstream. The separation strength can be measured by the magnitude of Cf inside the separation zone. We can find that although the size of the primary separation bubble is much larger for k8, the strength of separation is weaker than k6. In the separation zone, the peak magnitude of Cf for k6 is decreasing with k, and “oscillating” Cf-profiles are observed in rough airfoil cases [see Fig. 7(b)], which indicate formations of secondary bubbles inside the wavy grooves. The undershoots of local Cf after the roughened region (x/C>0.24) are enhanced with increasing k. This behavior is consistent with the observations in the rough-to-smooth transition of turbulent boundary layer.48 The memory effect of the upstream roughness condition is minimized for x/C0.5 with attached boundary layer (k6). Still, it does not show much variation for the detached boundary layer (k8) due to massive separation.

FIG. 8.

Contour of the time- and spanwise-averaged off diagonal Reynolds stresses, uv¯/U02[0.001,0.1] superimposed on the mean streamlines for different roughness wave numbers, k. (a) k =0, (b) k =4, (c) k =6, (d) k =8, (e) k =10, and (f) k =12. The transition onset is indicated by the threshold value of 0.001.

FIG. 8.

Contour of the time- and spanwise-averaged off diagonal Reynolds stresses, uv¯/U02[0.001,0.1] superimposed on the mean streamlines for different roughness wave numbers, k. (a) k =0, (b) k =4, (c) k =6, (d) k =8, (e) k =10, and (f) k =12. The transition onset is indicated by the threshold value of 0.001.

Close modal

The sudden growth of the primary separation bubble from k =6 to k =8 may also explain the variations of lift and drag coefficients with roughness wave numbers greater than zero, as shown in Fig. 5. These separation and reattachment behaviors imply strong effects on aerodynamic performance. However, we would like to check if the transition also plays a key role. The contour plot of the time- and spanwise-averaged off diagonal Reynolds stress, uv¯/U02 are also shown in Fig. 8. Here, the threshold value of 0.001 is chosen as an indicator for transition onset, which is commonly used in various numerical49 and experimental investigations50 of flow past bluff bodies. The locations of transition onset do not show obvious differences, occurring at x/C0.09±0.02, on the top of primary separation bubbles for all k cases (see Fig. 8). This feature indicates that the aerodynamic performance (lift and drag coefficients) is dominated by these large-scale separation bubbles rather than transition with different roughness configures.

Figure 9 shows RMS of wall pressure (pw,rms) and wall shear stress (τw,rms) along the suction side of the airfoil for different roughness wave numbers. For k6, both pw,rms and τw,rms are increasing inside the roughened region, and decreasing rapidly to some stable level after the reattachment location. For k8, the k =8 case produces slightly larger pw,rms and τw,rms for x/C0.85. Meanwhile, τw,rms is an order of magnitude smaller than pw,rms, indicating that the contribution of wall pressure fluctuations to the RMS of lift and drag coefficients is much larger than wall shear stress fluctuations. The latter is same as the contribution of wall pressure to lift and drag coefficients, as shown in Fig. 6. The mean level of pw,rms on the suction side for k8 is larger than smaller k cases and produces the sharp increase in CL,rms and CD,rms from k =6 to 8 as shown in Fig. 5.

FIG. 9.

RMS of (a) wall pressure and (b) wall shear stress along the suction side of the airfoil for different roughness wave numbers.

FIG. 9.

RMS of (a) wall pressure and (b) wall shear stress along the suction side of the airfoil for different roughness wave numbers.

Close modal

In Sec. IV B, the mean skin friction coefficient and streamline from the DNS have been discussed, which provide helpful information for interpreting the effects of the separation and reattachment on the aerodynamic performance (in the time-averaged sense) due to the wavy roughness. However, these mean-flow concepts are insufficient to fully understand the near-wall flow physics, including the unsteady load force and dynamics of separation bubbles that important in aerodynamic flows. In the present section, we discuss the unsteady separation and reattachment, first on the convective motion of the separation bubble, and then examine the proportion of reverse flow duration in the near-wall field.

Figure 10 shows the space-time diagram of the instantaneous spanwise-averaged skin friction coefficient for different roughness wave numbers, which also represents the evolution of the multiple vortical structures along the suction side.33 The evolution of separation and reattachment, although shown only for a short duration of 5C/U0, reveals the primary features. It is evident that the location of the first separation point remains virtually unchanged for the entire time sequence, the same as the mean separation point [see Fig. 7(b)]. Meanwhile, unsteady separation and reattachment can be observed near the trailing edge, which are missing in the mean Cf-profiles. The flow reversal is strongest around x/C=0.15, 0.16, and 0.2 for k =0, 4, and 6, respectively, and moves downward the airfoil with increasing k. This is also consistent with the locations of mean negative Cf peak as shown in Fig. 7(b). The local convective motions (around the right boundary of the roughened region) of separation bubbles can be identified by the convective speed Vc, based on the slope of the stripes. This velocity corresponds to the signature of shear layer structures on the wall.51 It is found that the convective velocities for different roughness wave numbers are Vc=(0.6±0.06)U0, which is similar to the convective velocity of the reattaching shear layer in backward-facing step flows (Vc=0.6U0).52 For the smooth airfoil case (k =0), Vc=0.66U0 is larger than the NACA0012 case with the same Rec but AoA=5°, i.e., Vc=0.56U0 as reported by Zhang and Samtaney,33 which confirms the argument that larger AoA can produce larger Vc (see Ref. 53). Dandois et al.51 summarized the convective velocities in flows of different configurations (mixing layer, backward-facing step, mixing layer past a cavity, fence, flat-plate leading edge) and found a clear convective velocity, Vc[0.4,0.65]U0 in these flows. These convective velocities are weakly correlated with the lift coefficients for different roughness wave numbers (see Fig. 5), but the sharp decrease in Vc from k =6 to 8 corresponds to the sharp variation of CL and CD. Meanwhile, the convective velocity is increasing to Vc=0.62U0 at k >8, which is slightly larger than Vc of k =4 and 6 (Vc0.6U0). This acceleration may be attributed to the interactions of secondary bubbles (merging and breakup of bubbles with decreased |Cf|) in the roughened region. For k8 cases, although some clear convective velocities can be observed beyond the right boundary of the roughened region, the convection motions of the shear layer structures are complex due to massive separation, which is linked to the convective phenomenon of different frequencies.51 The interactions of instantaneous separation bubbles are also reported by Gao et al.31 for the NACA0018 case (AoA=5°,Rec=105), which may also contribute to the wavy roughness effects on the aerodynamic performance. The convective separation bubbles downward the trailing edge will also change the evolution of wake flows, which can explain the roughness effects based on scaling analysis (see Sec. IV D for further discussions).

FIG. 10.

Space–time diagrams of instantaneous spanwise-averaged skin friction coefficient for different roughness wave numbers, k. (a) k =0, (b) k =4, (c) k =6, (d) k =8, (e) k =10, and (f) k =12. The black solid lines denote the isoline Cf = 0. The vertical solid lines denote the boundary of the roughened region.

FIG. 10.

Space–time diagrams of instantaneous spanwise-averaged skin friction coefficient for different roughness wave numbers, k. (a) k =0, (b) k =4, (c) k =6, (d) k =8, (e) k =10, and (f) k =12. The black solid lines denote the isoline Cf = 0. The vertical solid lines denote the boundary of the roughened region.

Close modal

Figure 11 shows the ratio of reversed flow for all k cases (time of near-wall reversed flow (Cf(x,z,t)<0) over the total simulation time). The flow is fully attached to the surface upstream of the leading edge detachment point, which is the same as the instantaneous spanwise-averaged skin friction coefficient as shown in Fig. 10. However, the unsteady behavior of the separation and reattachment beyond that detachment point is absent in the mean Cf-profiles. The roughened region is alternately occupied by separation and reattachment, which is obvious near the last mean reattachment lines. For k6, the mean attached zones are nibbled, especially near the trailing edge. For the mean attached zone of k =6 case, the number of reverse flow events is small compared to the total number of events. However, the gentle nibble grows into massive annexation, and most of the suction side is dominated by separation for k8. The spanwise variations of the reverse fraction represent the meandering of the wall streamlines, which consist flow structures of different scales and frequencies. All these unsteady motions will contribute to larger RMS of lift and drag coefficients with growing k.

FIG. 11.

Ratio of the time of near-wall reversed flow [Cf(x,z,t)<0] to the total simulation time for different roughness wave numbers, k. (a) k =0, (b) k =4, (c) k =6, (d) k =8, (e) k =10, and (f) k =12. The vertical dashed lines denote the mean separation and the reattachment lines detect the mean Cf-profiles, as shown in Fig. 7.

FIG. 11.

Ratio of the time of near-wall reversed flow [Cf(x,z,t)<0] to the total simulation time for different roughness wave numbers, k. (a) k =0, (b) k =4, (c) k =6, (d) k =8, (e) k =10, and (f) k =12. The vertical dashed lines denote the mean separation and the reattachment lines detect the mean Cf-profiles, as shown in Fig. 7.

Close modal

The wake downstream the trailing edge of an airfoil is asymmetric due to the effects of pressure gradients and curvature.54,55 Based on the momentum conservation law, the drag force acting on an airfoil is related to the wake structure.56,57 Presently, we try to explore the drag mechanism behind the wavy roughness effects from the perspective of wake evolution, which is strongly affected by different separation behaviors (see Secs. IV B and IV C) on the suction side of the airfoil with partially covered roughness elements.

The first quantity of interest is the mean velocity profile in the wake region, where the classical self-similarity wake profile fails for the asymmetric wake.55,58,59 Following the scaling approach proposed by Hah and Lakshminarayana,59 we define two characteristic length scales ls, lp and two characteristic velocity, Uc and Ud, where Uc is the lowest wake velocity in the x direction; Ud=U0Uc is the maximum velocity defect; lp and ls denote the distance from the wake center (location of Uc) to the points where the velocity defect, u¯Uc, is equal to Ud/2 on the pressure and suction sides, respectively (see Fig. 12). Hah and Lakshminarayana59 proposed the following Gauss profile for the wake defect velocity:

(7)

where yc is the coordinate of the wake center, and Λ is a constant equal to –0.6993.59 

FIG. 12.

Sketch of the mean velocity profile in the wake region.

FIG. 12.

Sketch of the mean velocity profile in the wake region.

Close modal

Figure 13 shows the scaled wake profiles for different roughness wave numbers at two typical locations, i.e., x/C=1 close to the trailing edge and x/C=2 away from the trailing edge. In the very-near-wake region, i.e., x/C=1, it is evident that the mean velocity profiles on the suction side follow the scaling equation (7) well but deviate sharply on the pressure side. For k8, the deviation is even pronounced, which might be caused by the stronger pressure gradient effects at the trailing edge (see Fig. 7). The separation bubbles close to the trailing edge for k8 (see Fig. 8) have substantial effects on the evolution of the wake and make the wake profiles more asymmetric. For x/C=2, the “history effect” from the trailing edge is vanishing, the wake structure is becoming more symmetric, and the wake profiles follow the scaling equation (7) well.

FIG. 13.

Scaling of the wake defect velocity at (a) x/C=1 and (b) x/C=2 for different roughness wave numbers, k. From bottom to top, k is increasing from 0 to 12, shifted by 1, 2,…, 5 along the ordinate; ———, scaled wake profile from Eq. (7).

FIG. 13.

Scaling of the wake defect velocity at (a) x/C=1 and (b) x/C=2 for different roughness wave numbers, k. From bottom to top, k is increasing from 0 to 12, shifted by 1, 2,…, 5 along the ordinate; ———, scaled wake profile from Eq. (7).

Close modal

Based on the above-mentioned scaling approach, the maximum velocity defect, (U0Uc)/U0 and wake width, lw=ls+lp for different roughness wave numbers are shown in Fig. 14. It is evident that the wake maximum velocity defect at the wake center is decaying with increasing downstream distances. In the very-near-wake region, i.e., x/C=1, the defect is sharp and increasing with k, but this trend disappears for x/C1.1. For x/C1.2, the defect is decreasing with k, which means that the velocity recovery at the wake center is faster for smaller roughness wave numbers. The decay rate of velocity defect is increasing with k for x/C1.4, but almost the same for x/C1.5. The wake width is increasing with k, and a sharp increase can be observed from k =6 to k =8, which is in accordance with the drag coefficients as shown in Fig. 5. The growth rate of the wake width is also increasing with k and much faster for k8, which confirms that k =8 is a critical case.

FIG. 14.

(a) The wake width, lw and (b) maximum velocity defect, (U0Uc)/U0 for different roughness wave numbers in the wake region within x/C[1,2].

FIG. 14.

(a) The wake width, lw and (b) maximum velocity defect, (U0Uc)/U0 for different roughness wave numbers in the wake region within x/C[1,2].

Close modal

Qualitatively, Pope60 proposed that the momentum deficit flow rate, which equals the drag on the bluff body, follows:

(8)

It should be noted that although the wake defect is decreasing with k, the variation is much weaker than the wake width, which to some extent, explains the drag mechanism with different roughness wave numbers, as shown in Fig. 5(b).

We perform a direct numerical simulation of flow past the NACA0012 airfoil at 10° angle of attack and Reynolds number of 5×104 with partially covered wavy roughness elements to investigate the roughness effect on the aerodynamic performance and the underlying mechanism. The simulations are performed with an energy-conservative fourth-order parallel code solving the incompressible Navier–Stokes equations in generalized curvilinear coordinates. The length of the roughened domain and roughness height are fixed and located near the leading edge, whereas the roughness wavenumber k varies from 4 to 12. The smooth airfoil is used as a baseline case.

The roughness elements degrade the aerodynamic performance, i.e., reduce the lift coefficient and increase the drag coefficient. The drag coefficient increases monotonically with k, while the variation of lift coefficient with k is similar to the drag crisis phenomenon as observed in cylinder flows. Both the sharp variations of lift and drag coefficients from k =6 to k =8 indicate that k =8 is a critical case. The latter result is also confirmed in the RMS of the lift and drag coefficients.

To reveal the underlying mechanism, we examine the mean pressure, skin friction coefficient, and the mean flow field. The roughness elements can strongly affect the separation behavior but have little effect on the location of transition onset. For 0k6, the main separation bubble stretches around the roughened region, and the secondary separation bubble occurs at the windward side of the roughness elements. The case of k =8 is a critical case. In fact, beyond k =8, a massive separation occurs on the upper surface, drastically affecting the aerodynamic performance.

The evolution of separation and reattachment on the airfoil surface is analyzed to further explore the unsteady load force with the dynamics of separation bubbles. Around the roughened region, clear convection motions of separation bubbles can be observed. The convection velocity is largest for the smooth airfoil, and the roughness elements retard the convection motions, which is strongest for the critical case (k =8). These convective bubbles also affect the wake evolution downstream of the trailing edge. The asymmetric wake is analyzed by employing the scaling law by Hah and Lakshminarayana.59 The wake profiles on the pressure side do not follow in the very-near-wake region for k8 due to the downward convective bubbles, but history effect is vanishing far away from the trailing edge, and wake profiles follow the scaling law well. The wake defect is decreasing with k, but the variation is much weaker than the wake width, showing a sharp increase from k =6 to 8. The wake scaling analysis, to some extent, confirms the drag mechanism induced by the roughness elements.

The present simulations are performed at a specific Reynolds number and angle of attack, and it is possible that some of the underlying mechanism changes if the Reynolds number is increased, or the angle of attack is varied. It would be desirable to verify this conjecture by performing simulations at higher Reynolds number and different angles of attack, which can be potentially done in the future.

The Cray XC40 Shaheen II at KAUST was used for all simulations reported. This research was partially supported under KAUST OCRF URF/1/1394-01 and under baseline research funds of R. Samtaney. W. Gao acknowledges the financial support by Visiting Researcher Fund Program of State Key Laboratory of Water Resources and Hydropower Engineering Science (Grant No. 2020HLG02).

The authors have no conflicts to disclose.

Wei Gao: Conceptualization (equal); Formal analysis (equal); Investigation (lead); Methodology (equal); Validation (lead); Writing – original draft (lead); Writing – review & editing (equal). Ravi Samtaney: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Resources (lead); Supervision (lead); Writing – review & editing (equal). Matteo Parsani: Investigation (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
J.
Jiménez
, “
Turbulent flows over rough walls
,”
Annu. Rev. Fluid Mech.
36
,
173
196
(
2004
).
2.
D.
Chung
,
N.
Hutchins
,
M. P.
Schultz
, and
K. A.
Flack
, “
Predicting the drag of rough surfaces
,”
Annu. Rev. Fluid Mech.
53
,
439
471
(
2021
).
3.
M.
Kadivar
,
D.
Tormey
, and
G.
McGranaghan
, “
A review on turbulent flow over rough surfaces: Fundamentals and theories
,”
Int. J. Thermofluids
10
,
100077
(
2021
).
4.
J.
Nikuradse
, “
Strom¨ungsgestze in rauhen rohre
,”
VDI-Forschungsh.
4
,
1
33
(
1933
).
5.
L. F.
Moody
, “
Friction factors for pipe flow
,”
Trans. Am. Soc. Mech. Eng.
66
,
671
684
(
1944
).
6.
S. G.
Kandlikar
,
D.
Schmitt
,
A. L.
Carrano
, and
J. B.
Taylor
, “
Characterization of surface roughness effects on pressure drop in single-phase flow in minichannels
,”
Phys. Fluids
17
,
100606
(
2005
).
7.
Y.
Liu
,
J.
Li
, and
A. J.
Smits
, “
Roughness effects in laminar channel flow
,”
J. Fluid Mech.
876
,
1129
1145
(
2019
).
8.
N.
Saito
,
D. I.
Pullin
, and
M.
Inoue
, “
Large eddy simulation of smooth-wall, transitional and fully rough-wall channel flow
,”
Phys. Fluids
24
,
075103
(
2012
).
9.
D.
Chung
,
L.
Chan
,
M.
MacDonald
,
N.
Hutchins
, and
A.
Ooi
, “
A fast direct numerical simulation method for characterising hydraulic roughness
,”
J. Fluid Mech.
773
,
418
431
(
2015
).
10.
E.
Sagol
,
M.
Reggio
, and
A.
Ilinca
, “
Issues concerning roughness on wind turbine blades
,”
Renewable Sustainable Energy Rev.
23
,
514
525
(
2013
).
11.
W.
Cheng
,
D. I.
Pullin
, and
R.
Samtaney
, “
Large-eddy simulation of flow over a grooved cylinder up to transcritical Reynolds numbers
,”
J. Fluid Mech.
835
,
327
362
(
2018
).
12.
B.
Zhou
,
X.
Wang
,
W.
Guo
,
J.
Zheng
, and
S. K.
Tan
, “
Experimental measurements of the drag force and the near-wake flow patterns of a longitudinally grooved cylinder
,”
J. Wind Eng. Ind. Aerodyn.
145
,
30
41
(
2015
).
13.
P. W.
Bearman
and
J. K.
Harvey
, “
Control of circular cylinder flow by the use of dimples
,”
AIAA J.
31
,
1753
1756
(
1993
).
14.
Y.
Han
and
J.
Palacios
, “
Heat transfer evaluation on ice-roughened cylinders
,”
AIAA J.
55
,
1070
1074
(
2017
).
15.
T.
Cebeci
, “
Calculation of flow over iced airfoils
,”
AIAA J.
27
,
853
861
(
1989
).
16.
M. B.
Bragg
,
A. P.
Broeren
, and
L. A.
Blumenthal
, “
Iced-airfoil aerodynamics
,”
Prog. Aerosp. Sci.
41
,
323
362
(
2005
).
17.
F. T.
Lynch
and
A.
Khodadoust
, “
Effects of ice accretions on aircraft aerodynamics
,”
Prog. Aerosp. Sci.
37
,
669
767
(
2001
).
18.
S. J.
Stebbins
,
E.
Loth
,
A. P.
Broeren
, and
M.
Potapczuk
, “
Review of computational methods for aerodynamic analysis of iced lifting surfaces
,”
Prog. Aerosp. Sci.
111
,
100583
(
2019
).
19.
N.
Rostamzadeh
,
R.
Kelso
,
B.
Dally
, and
K.
Hansen
, “
The effect of undulating leading-edge modifications on NACA 0021 airfoil characteristics
,”
Phys. Fluids
25
,
117101
(
2013
).
20.
V.
Kumar
,
U.
Piomelli
, and
O.
Lehmkuhl
, “
Large-eddy simulations of the flow on an aerofoil with leading-edge imperfections
,”
J. Turbul.
22
,
735
760
(
2021
).
21.
D.
Serson
,
J. R.
Meneghini
, and
S. J.
Sherwin
, “
Direct numerical simulations of the flow around wings with spanwise waviness
,”
J. Fluid Mech.
826
,
714
731
(
2017
).
22.
J. W.
Kim
,
S.
Haeri
, and
P. F.
Joseph
, “
On the reduction of aerofoil–turbulence interaction noise associated with wavy leading edges
,”
J. Fluid Mech.
792
,
526
552
(
2016
).
23.
C.
Thomas
,
S.
Mughal
, and
R.
Ashworth
, “
Development of Tollmien-Schlichting disturbances in the presence of laminar separation bubbles on an unswept infinite wavy wing
,”
Phys. Rev. Fluids
2
,
043903
(
2017
).
24.
M. N.
Goodhand
,
R. J.
Miller
, and
H. W.
Lung
, “
The impact of geometric variation on compressor two-dimensional incidence range
,”
J. Turbomach.
137
,
021007
(
2015
).
25.
M.
Fujino
,
Y.
Yoshizaki
, and
Y.
Kawamura
, “
Natural-laminar-flow airfoil development for a lightweight business jet
,”
J. Aircraft
40
,
609
615
(
2003
).
26.
P. M.
Bagade
,
Y. G.
Bhumkar
, and
T. K.
Sengupta
, “
An improved orthogonal grid generation method for solving flows past highly cambered aerofoils with and without roughness elements
,”
Comput. Fluids
103
,
275
289
(
2014
).
27.
T. J.
Mueller
and
J. D.
DeLaurier
, “
Aerodynamics of small vehicles
,”
Annu. Rev. Fluid Mech.
35
,
89
111
(
2003
).
28.
Y.
Morinishi
,
T. S.
Lund
,
O. V.
Vasilyev
, and
P.
Moin
, “
Fully conservative higher order finite difference schemes for incompressible flow
,”
J. Comput. Phys.
143
,
90
124
(
1998
).
29.
T. A.
Zang
, “
On the rotation and skew-symmetric forms for incompressible flow simulations
,”
Appl. Numer. Math.
7
,
27
40
(
1991
).
30.
Y.
Zang
,
R. L.
Street
, and
J. R.
Koseff
, “
A non-staggered grid, fractional step method for time-dependent incompressible Navier-Stokes equations in curvilinear coordinates
,”
J. Comput. Phys.
114
,
18
33
(
1994
).
31.
W.
Gao
,
W.
Zhang
,
W.
Cheng
, and
R.
Samtaney
, “
Wall-modelled large-eddy simulation of turbulent flow past airfoils
,”
J. Fluid Mech.
873
,
174
210
(
2019
).
32.
W.
Zhang
,
W.
Cheng
,
W.
Gao
,
A.
Qamar
, and
R.
Samtaney
, “
Geometrical effects on the airfoil flow separation and transition
,”
Comput. Fluids
116
,
60
73
(
2015
).
33.
W.
Zhang
and
R.
Samtaney
, “
Assessment of spanwise domain size effect on the transitional flow past an airfoil
,”
Comput. Fluids
124
,
39
53
(
2016
).
34.
L. E.
Jones
,
R. D.
Sandberg
, and
N. D.
Sandham
, “
Direct numerical simulations of forced and unforced separation bubbles on an airfoil at incidence
,”
J. Fluid Mech.
602
,
175
207
(
2008
).
35.
V.
Kitsios
,
L.
Cordier
,
J.-P.
Bonnet
,
A.
Ooi
, and
J.
Soria
, “
On the coherent structures and stability properties of a leading-edge separated aerofoil with turbulent recirculation
,”
J. Fluid Mech.
683
,
395
416
(
2011
).
36.
O.
Lehmkuhl
,
I.
Rodríguez
,
A.
Baez
,
A.
Oliva
, and
C.
Pérez-Segarra
, “
On the large-eddy simulations for the flow around aerodynamic profiles using unstructured grids
,”
Comput. Fluids
84
,
176
189
(
2013
).
37.
I.
Rodríguez
,
O.
Lehmkuhl
,
R.
Borrell
, and
A.
Oliva
, “
Direct numerical simulation of a NACA0012 in full stall
,”
Int. J. Heat Fluid Flow
43
,
194
203
(
2013
).
38.
W.
Zhang
and
R.
Samtaney
, “
A direct numerical simulation investigation of the synthetic jet frequency effects on separation control of low-Re flow past an airfoil
,”
Phys. Fluids
27
,
055101
(
2015
).
39.
N. J.
Georgiadis
,
D. P.
Rizzetta
, and
C.
Fureby
, “
Large-eddy simulation: Current capabilities, recommended practices, and future research
,”
AIAA J.
48
,
1772
1784
(
2010
).
40.
Y.
Yamaguchi
,
T.
Ohtake
, and
A.
Muramatsu
, “
Pressure distribution on a NACA0012 airfoil at low Reynolds numbers
,” in
The Proceedings of the International Conference on Jets, Wakes and Separated Flows
(
The Japan Society of Mechanical Engineers
,
2013
), p.
1201
.
41.
M. S.
Boutilier
and
S.
Yarusevych
, “
Separated shear layer transition over an airfoil at a low Reynolds number
,”
Phys. Fluids
24
,
084105
(
2012
).
42.
T.
Ohtake
, “
Flow field around NACA-0012 airfoil in low Reynolds numbers: Part 2 characteristics of boundary layer
,”
J. Jpn. Soc. Aeronaut. Space Sci.
64
,
123
130
(
2016
).
43.
B.
Cantwell
and
D.
Coles
, “
An experimental study of entrainment and transport in the turbulent near wake of a circular cylinder
,”
J. Fluid Mech.
136
,
321
374
(
1983
).
44.
P.
Bot
,
M.
Rabaud
,
G.
Thomas
,
A.
Lombardi
, and
C.
Lebret
, “
Sharp transition in the lift force of a fluid flowing past nonsymmetrical obstacles: Evidence for a lift crisis in the drag crisis regime
,”
Phys. Rev. Lett.
117
,
234501
(
2016
).
45.
W.
Cheng
,
D.
Pullin
,
R.
Samtaney
,
W.
Zhang
, and
W.
Gao
, “
Large-eddy simulation of flow over a cylinder with ReD from 3.9×103 to 8.5×105: A skin-friction perspective
,”
J. Fluid Mech.
820
,
121
158
(
2017
).
46.
W.
Gao
,
W.
Cheng
, and
R.
Samtaney
, “
Large-eddy simulations of turbulent flow in a channel with streamwise periodic constrictions
,”
J. Fluid Mech.
900
,
A43
(
2020
).
47.
E.
Eljack
,
J.
Soria
,
Y.
Elawad
, and
T.
Ohtake
, “
Simulation and characterization of the laminar separation bubble over a NACA-0012 airfoil as a function of angle of attack
,”
Phys. Rev. Fluids
6
,
034701
(
2021
).
48.
M.
Li
,
C. M.
de Silva
,
A.
Rouhi
,
R.
Baidya
,
D.
Chung
,
I.
Marusic
, and
N.
Hutchins
, “
Recovery of wall-shear stress to equilibrium flow conditions after a rough-to-smooth step change in turbulent boundary layers
,”
J. Fluid Mech.
872
,
472
491
(
2019
).
49.
Y.
Zhou
and
Z. J.
Wang
, “
Effects of surface roughness on separated and transitional flows over a wing
,”
AIAA J.
50
,
593
609
(
2012
).
50.
N. A.
Buchmann
,
C.
Atkinson
, and
J.
Soria
, “
Influence of ZNMF jet flow control on the spatio-temporal flow structure over a NACA-0015 airfoil
,”
Exp. Fluids
54
,
1485
(
2013
).
51.
J.
Dandois
,
E.
Garnier
, and
P.
Sagaut
, “
Numerical simulation of active separation control by a synthetic jet
,”
J. Fluid Mech.
574
,
25
58
(
2007
).
52.
R. L.
Simpson
, “
Turbulent boundary-layer separation
,”
Annu. Rev. Fluid Mech.
21
,
205
232
(
1989
).
53.
M. S.
Boutilier
and
S.
Yarusevych
, “
Parametric study of separation and transition characteristics over an airfoil at low Reynolds numbers
,”
Exp. Fluids
52
,
1491
1506
(
2012
).
54.
A.
Nakayama
, “
Curvature and pressure-gradient effects on a small-defect wake
,”
J. Fluid Mech.
175
,
215
246
(
1987
).
55.
F. O.
Thomas
and
X.
Liu
, “
An experimental investigation of symmetric and asymmetric turbulent wake development in pressure gradient
,”
Phys. Fluids
16
,
1725
1745
(
2004
).
56.
G. K.
Batchelor
,
An Introduction to Fluid Dynamics
(
Cambridge University Press
,
1967
).
57.
D. D.
Chao
and
C. P.
Van Dam
, “
Airfoil drag prediction and decomposition
,”
J. Aircraft
36
,
675
681
(
1999
).
58.
R.
Raj
and
B.
Lakshminarayana
, “
Characteristics of the wake behind a cascade of airfoils
,”
J. Fluid Mech.
61
,
707
730
(
1973
).
59.
C.
Hah
and
B.
Lakshminarayana
, “
Measurement and prediction of mean velocity and turbulence structure in the near wake of an airfoil
,”
J. Fluid Mech.
115
,
251
282
(
1982
).
60.
S. B.
Pope
,
Turbulent Flows
(
Cambridge University Press
,
2000
).