Appl. Math. Mech. -Engl. Ed.   2019, Vol. 40 Issue (3): 331-342     PDF       
http://dx.doi.org/10.1007/s10483-019-2442-8
Shanghai University
0

Article Information

FAN Yitong, CHENG Cheng, LI Weipeng
Effects of the Reynolds number on the mean skin friction decomposition in turbulent channel flows
Applied Mathematics and Mechanics (English Edition), 2019, 40(3): 331-342.
http://dx.doi.org/10.1007/s10483-019-2442-8

Article History

Received Sep. 5, 2018
Revised Nov. 2, 2018
Effects of the Reynolds number on the mean skin friction decomposition in turbulent channel flows
FAN Yitong1, CHENG Cheng1, LI Weipeng1,2     
1. School of Aeronautics and Astronautics, Shanghai Jiao Tong University, Shanghai 200240, China;
2. Engineering Research Center of Gas Turbine and Civil Aero Engine, Shanghai 200240, China
Abstract: As the Reynolds number increases, the skin friction has been identified as the dominant drag in many practical applications. In the present paper, the effects of the Reynolds number on the mean skin friction decomposition in turbulent channel flows up to Reτ=5 200 are investigated based on two different methods, i.e., the FukagataIwamoto-Kasagi (FIK) identity (FUKAGATA, K., IWAMOTO, K., and KASAGI, N. Contribution of Reynolds stress distribution to the skin friction in wall-bounded flows. Physics of Fluids, 14(11), L73-L76 (2002)) and the Renard-Deck (RD) identity (DECK, S., RENARD, N., LARAUFIE, R., and WEISS, P. É. Large-scale contribution to mean wall shear stress in high-Reynolds-number flat-plate boundary layers up to Reθ=13 650. Journal of Fluid Mechanics, 743, 202-248 (2014)). The direct numerical simulation (DNS) data provided by Lee and Moser (LEE, M. and MOSER, R. D. Direct numerical simulation of turbulent channel flow up to Reτ ≈ 5 200. Journal of Fluid Mechanics, 774, 395-415 (2015)) are used. For these two skin friction decomposition methods, their decomposed constituents are discussed and compared for different Reynolds numbers. The integrands of the decomposed constituents are locally analyzed across the boundary layer to assess the actions associated with the inhomogeneity and multi-scale nature of turbulent motion. The scaling of the decomposed constituents and their integrands are presented. In addition, the boundary layer is divided into three sub-regions to evaluate the contributive proportion of each sub-region with an increase in the Reynolds number.
Key words: drag decomposition    mean skin friction    turbulent channel flow    Reynolds number effect    
1 Introduction

Drag, which acts in opposition to a moving body with respect to the surrounding fluid can be categorized using two essentially different methods, near-field and far-field integrations[1]. The near-field method integrates the pressure and viscous stress over the surfaces of the body, resulting in components of pressure drag and frictional drag, respectively. The frictional drag in a wall-bounded turbulent flow has been determined to be much higher than that of a laminar flow at the same Reynolds number. It contributes to the total drag up to 50% for the commercial aircraft, 90% for underwater vehicles, and almost 100% for long pipe and channel flows[2]. As the Reynolds number increases, the frictional drag plays a key role in many practical applications[3].

The mean skin friction, also termed the mean wall shear stress, can be further decomposed into various constituents according to a variety of mathematical derivations and physical interpretations. By applying a triple integration to the Reynolds averaged Navier-Stokes equation, Fukagata et al.[4] derived a simple, direct relationship between the skin friction coefficient and the Reynolds stress for three canonical wall-bounded turbulent flows, which was subsequently named the Fukagata-Iwamoto-Kasagi (FIK) identity. This method has been generalized to compressible flows[5] and complex wall surfaces[6-7]. Moreover, the FIK identity has been refined to be applicable to flows with unavailable streamwise gradients or incomplete statistics over the entire boundary layer[8-9]. Over the years, the FIK identity has been widely used in numerous studies[3, 10-11]. By performing a similar triple integration on the mean vorticity equation, Yoon et al.[12] related the skin friction to the motions of vortical structures. Recently, another theoretical decomposition of the mean skin friction has been proposed by Renard and Deck[13]. By applying a single integration to the mean streamwise kinetic-energy-budget equation in an absolute reference frame (in which the undisturbed fluid is not moving), the mean skin-friction generation is characterized by physical phenomena that reveal the energy transfer from the wall to the fluid.

The control of the skin friction and its associated near-wall aerodynamic/hydrodynamic phenomena is of fundamental and practical importance, motivated in part by energy saving objectives[14]. Several flow control approaches, such as the addition of long-chain polymers, spanwise wall oscillation, near-wall blowing and suction, superhydrophobic surfaces, and micro-grooved structures (riblets), have been proposed and tested to reduce skin friction in turbulent flows[15-19]. Most of the aforementioned flow control approaches have been attempted to manipulate the turbulent structures within the near-wall region. The near-wall Reynolds stress is of particular importance for the prediction and control of the mean skin friction. However, as the Reynolds number is increased to more pragmatic values, the drag reduction efficiency is reduced[20]. For high practical Reynolds number values, the physical thickness of the near-wall region and hence the size of the structures that populate this region become smaller[2, 21-22]. Hwang[23] noted that with an increase in the Reynolds number, the contribution of the near-wall small-scale structures to the skin friction decays, and the large-scale motion has a greater effect considering the effect of their amplitude modulation on small scales and their energy superposition onto the inner region. A spectral analysis of the linearly weighted Reynolds stress in the FIK identity was carried out by Deck et al.[3]. The outer energy site where large scales were located was shown to have a dominant contribution to the skin friction. However, they also reported that this specific contribution for a given scale at a given height, which originated from a mathematical derivation, should be carefully utilized because of its vague physical interpretation with respect to the skin friction generation mechanism. While in Ref. [13], they showed that the buffer layer dynamics at low Reynolds numbers plays a dominant role, whereas at high Reynolds numbers, the logarithmic layer dynamics is most influential in the skin friction generation process, suggesting new drag reduction strategies in the field of flow control.

With the objective of the effective reduction of skin friction at high Reynolds numbers, it is of great importance to clarify and evaluate the effects of the Reynolds number on skin friction and its components based on different drag decomposition methods. So far, to the best of the authors' knowledge, no results of such an analysis have been published. In this report, Section 2 introduces the direct numerical simulation (DNS) approach for turbulent channel flows and yields skin friction coefficients compared with the empirical results. In terms of specific decomposition formulas for the mean skin friction based on two integration methods, the FIK identity and Renard-Deck (RD) identity are presented in Section 3. In Section 4, the dependence of the mean skin friction generation on the Reynolds number is investigated with the assessment of turbulence statistics and different wall layer dynamics. Finally, concluding remarks are given in Section 5.

2 DNS of turbulent channel flows

We use the DNS data of fully developed incompressible channel flows provided by Lee and Moser[24]. The DNS was performed based on the POONGBACK code which uses the spectral numerical method of Kim et al.[25]. The friction Reynolds number Reτ is defined as ρ*uτ*h*/μ*, where the superscript * denotes the dimensional variables, ρ is the density, h is the channel half height, μ is the dynamic viscosity, and uτ is the friction velocity. Five cases with Reτ =180, 550, 1 000, 2 000, and 5 200 were used. Simulation details and validations can be found in Refs. [25]-[28]. The statistical data used are available on the website http://turbulence.ices.utexas.edu.

The mean skin-friction coefficient Cf is defined as the ratio of the mean wall shear stress to the dynamic pressure. It can be written as

(1)

where Re=ρ*ub*h*/μ* is the Reynolds number based on the bulk velocity , y is the site distance from the wall surface, and 〈u〉 is the mean streamwise velocity non-dimensionalized by the channel flow bulk velocity. The calculated mean skin-friction coefficient is compared with Dean's empirical correlation[29], Cf=0.073(2Re)-1/4, as shown in Fig. 1, which is apparently in good agreement with the theoretical prediction.

Fig. 1 The mean skin-friction coefficient in correlation with the Reynolds number
3 Mean skin friction decomposition methods

In this section, two theoretical methods are introduced for incompressible turbulent channel flows to decompose the mean skin friction.

3.1 FIK identity

In the FIK identity, the following conditions are assumed: (ⅰ) constant flow rate, (ⅱ) no-slip condition at the wall surfaces, (ⅲ) statistical homogeneity in the spanwise direction, and (ⅳ) symmetry with respect to the channel central plane. These assumptions are reasonable for incompressible turbulent channel flows. Considering the periodicity of the flow parameters along the streamwise direction, the streamwise statistical homogeneity is additionally assumed. Under these assumptions, a triple integration by parts performed on the Reynolds averaged Navier-Stokes equation in the streamwise direction gives the simplified FIK identity[4],

(2)

where -〈 u'v'〉 is the Reynolds shear stress. The first term, Cf1, FIK, has been proven to be identical to the well-known laminar solution at the same Reynolds number, and this is referred to as the laminar drag. The second term, Cf2, FIK, is characterized by the turbulent fluctuations, i.e., the linearly weighted Reynolds stress over the boundary layer, and consequently, it is described as the turbulence-induced drag. However, the weight function of (1-y) is probably ill-defined for physical interpretation, because the linear nature of the weight function results from the mathematical derivation of the decomposition and has no clear explanation in terms of physical processes[13]. Although the FIK identity is widely used, it is noteworthy that the linearly weighted Reynolds stress has a clear limitation in terms of identifying the skin-friction generation mechanism.

3.2 RD identity

Renard and Deck[13] gave another theoretical decomposition of the mean skin friction based on the averaged streamwise kinetic energy budget in an absolute reference frame (in which the undisturbed fluid is not moving). We refer to this method as the RD identity (named after its authors). For incompressible channel flows, we have

(3)

where the first term on the right-hand side of Eq. (3) represents a direct molecular viscous dissipation in the boundary layer, and the product of -〈u'v'〉 and in the second term on the right hand side of Eq. (3) characterizes the production of the turbulent kinetic energy. In this regard, the RD identity facilitates the visualization of the skin friction generation from the perspective of energy transfer from the wall to the fluid in an absolute reference frame, by means of the dissipation of molecular viscosity and turbulent energy[30]. The advantage of this method is that the mean skin friction is decomposed based on its physical generation mechanism.

4 Results and discussion

Firstly, the estimated skin-friction coefficients and their constituents at different Reynolds numbers are listed in Table 1. For the FIK identity, the relative error (Cf, FIK-Cf)/Cf is confined within ± 0.25% for all cases studied. For the RD identity, the relative error (Cf, RD-Cf)/Cf is larger than that estimated by the FIK identity. Errors within ±1.58% are observed. Basically, both skin friction decomposition methods are in good agreement with the results directly calculated using the wall-normal mean streamwise velocity gradients at the wall.

Table 1 Estimated skin friction coefficients and their constitutes at different Reynolds numbers

The ratios of the constituents to Cf are plotted in Fig. 2. In general, when the Reynolds number increases, Cf1 /Cf reduces, and Cf2 /Cf increases, which occurs for both the FIK identity and RD identity. Although these two drag decomposition methods share some similarity in curve tendencies, they are essentially different with respect to interpreting their contributions to skin friction generation, which will be discussed in the following section.

Fig. 2 Contributions of each term of Cf1, FIK, Cf2, FIK, Cf1, RD, and Cf2, RD to skin friction generation, where the triangle, circle, square, and asterisk markers represent Cf1, FIK/Cf, Cf2, FIK/Cf, Cf1, RD/Cf, and Cf2, RD/Cf, respectively

In the FIK identity, the laminar drag Cf1, FIK and turbulence induced drag Cf2, FIK are separated, and the latter becomes much more important as the Reynolds number is increased. In order to better assess the contribution of the Reynolds stress, Cf2, FIK/Cf can be written as

(4)

where the superscript + denotes normalization by the viscous variables. We now quantify the dependence of the linearly weighted Reynolds stress on the Reynolds number via the pre-multiplied normalized integrand in Cf2, FIK/Cf, i.e., . Figure 3(a) shows the profiles of the pre-multiplied integrands in a semi-logarithmic plot with the exact DNS solutions, which is possibly a new contribution. The areas below the profiles represent the contribution of Cf2, FIK.

Fig. 3 Distributions of the integrand of Cf2, FIK/Cf in the channel flows (a) as a function of y+ and (b) as a function of y

As the Reynolds number increases, the contribution of the outer layer (y+>30) dynamics is augmented, as illustrated in Fig. 3(a). The contribution of the near-wall region y+ < 30 for instance, is dramatically reduced. The implication is that the role of near-wall turbulent motion in terms of skin friction generation is less important, as previously determined in numerous studies. To some extent, this explains the observation of the decay of the drag reduction efficiency at high Reynolds numbers by controlling the near-wall turbulence. In terms of the slight increase in the peak values at high Reynolds numbers, as shown in Fig. 3(b), this is consistent with the tendency of the normalized Reynolds stress over a wide range of Reynolds numbers[24]. A similarity is also found in the peak-point location in the extrinsic scale.

The RD identity characterizes skin friction generation from the perspective of energy transfer from the moving wall to the fluid. Cf1, RD is ascribed to molecular dissipation, which is expected to be dominant in the near-wall region, especially in the viscous sub-layer where the Reynolds stress is negligible compared with the viscous stress. Cf1, RD/Cf in the intrinsic scale is written as follows:

(5)

where uτ is non-dimensionalized by the bulk velocity. The pre-multiplied integrands of Cf1, RD/Cf, i.e., , are plotted in Fig. 4(a). Regardless of the Reynolds number, the curves peak at y+≈6.5. The areas below the curves reflect the importance of Cf1, RD in a straightforward manner. As the Reynolds number increases, the area becomes smaller indicating a negative correlation between the contribution of the viscous dissipation and the Reynolds number. Almost all contributions of the viscous dissipation originate from the region y+ < 30, which is consistent with the physical phenomenon that the viscous stress is of importance in the near-wall region. Figure 4(b) shows the distributions of . A good coincidence of the distributions is apparently observed because of the universal log-law for the near-wall distribution of 〈 u+ as a function of y+. The differences are limited for the case of Reτ=180, which is consistent with Laadhari's report[31] that this constancy holds for Reτ>500. Since , the areas below the curves of are exactly proportional to Cf1, RD /Cf1.5. Thus, a constancy of Cf1, RD /Cf1.5 is proposed, regardless of the Reynolds number.

Fig. 4 (a) Distributions of the pre-multiplied integrand of Cf1, RD/Cf as a function of y+ and (b) distributions of as a function of y+

In terms of the production of the turbulent kinetic energy Cf2, it is locally balanced with the dissipation, transport, and diffusion[32]. If we perform an integration over the entire boundary layer, a perfect balance would exist between the total production of the turbulent kinetic energy and the total dissipation. As seen in Appendix A, Fig. A1 shows the distributions of the turbulent kinetic energy production and dissipation, and Table A1 represents the total turbulent kinetic energy production and turbulent energy dissipation. Their relative errors are limited within 0.2%. Thus, Cf2, RD can also be interpreted as the total turbulent energy dissipation[30]. In contrast, Cf1, RD is characterized as the direct molecular viscous dissipation.

Fig. A1 Distributions of turbulent kinetic energy production and dissipation
Table A1 Total turbulent kinetic energy production and turbulent energy dissipation

Cf2, RD/Cf is written as

(6)

Its pre-multiplied integrand and the corresponding expression are profiled in Fig. 5. The turbulent kinetic energy production is expressed as or alternatively , where ν is the kinematic viscosity. Thus, or essentially represents the turbulent kinetic energy production weighted by y+ν /uτ3 or y+ν /uτ4. Regardless of the Reynolds number, the curves in Fig. 5 peak at almost the same location of y+≈17.0. To identify the reason why the curves peak at y+≈17.0, the distributions of the viscous stress and Reynolds stress -〈 u'v'〉 are plotted in Fig. 6, as well as the turbulent kinetic energy production. A universal rule is found that when the viscous stress and the Reynolds stress are equal, the maximum turbulent kinetic energy production is obtained. The intersections at y+≈11.5 are slightly closer to the wall than the peak at y+≈17.0 in Fig. 5 because of the weight function, as indicated earlier.

Fig. 5 (a) Distributions of the pre-multiplied integrand of Cf2, RD/Cf in a channel flow as a function of y+ and (b) distributions of as a function of y+
Fig. 6 Distributions of Reynolds stress, viscous stress, and turbulent kinetic energy production as a function of y+, where the dotted lines represent the Reynolds stress, the dashed lines indicate the viscous stress, the solid lines indicate the turbulent kinetic energy production, and the cases with Reτ = 180, 550, 1 000, 2 000, and 5 200 are represented by the circle, square, triangle, asterisk, and plus sign markers, respectively (color online)

For cases with Reτ>500, a secondary peak emerges in the outer layer, which is ascribed to the generation of large-scale turbulent motions at these Reynolds numbers[33-34]. The large-scale turbulent motions contain large amounts of turbulent kinetic energy[35-36], which results in an increasing proportion of outer layer dynamics in terms of the skin friction. In the near-wall region, on the other hand, the small-scale turbulent motions are modulated by large-scale motions, which leads to the redistribution of turbulent kinetic energy from the near-wall region to the outer region. This may possibly explain why the peak value at y+≈17.0 in Fig. 5(a) is reduced as the Reynolds number increases. In Fig. 5(b), the distribution of remains approximately unchanged in the near-wall region (y+ < 30 for instance). The area below the curves in Fig. 5(b) reflects the importance of Cf2, RD divided by Cf1.5, considering that . It indicates that with respect to Cf1.5, the effects of the Reynolds number on the contribution of the near-wall dynamics (y+ < 30) are quite unapparent.

Figure 7 reveals the dependence of Cf1, RD/Cf1.5 and Cf2, RD/Cf1.5 on the Reynolds number. A constancy of Cf1, RD/Cf1.5 approximately at 6.59 is identified, especially at high Reynolds numbers. Laadhari[31] once investigated the viscous energy dissipation rate for channel flows and reported that this rate remains constant for the friction Reynolds number larger than 500. The constancy of Cf1, RD/Cf1.5 in the present study is consistent with Laadhari's conclusion[31], which suggests that the contribution of molecular viscous dissipation to the skin friction maintains the same rate of Cf0.5, i.e., Cf1, RD/Cf1.5 ≈ 6.59 × Cf0.5. In contrast with Cf1, RD/Cf1.5, Cf2, RD/Cf1.5 has a logarithmic relationship with Reτ. It is well fitted by

Fig. 7 The dependence of Cf1, RD/Cf1.5 and Cf2, RD/Cf1.5 on the Reynolds number
(7)

At low Reynolds numbers, for instance, Reτ < 550, the contribution of the molecular viscous dissipation is larger than that of the turbulent energy dissipation. As the Reynolds number increases, turbulent fluctuations are strengthened and gradually dominate skin friction generation. Thus, the turbulent energy dissipation becomes the major cause of skin friction generation, as shown in Fig. 7. In terms of the ratio of (Cf1, RD+Cf2, RD) to Cf1.5, which is inversely proportional to uτ, it changes logarithmically with the friction Reynolds number. This suggests an empirical law for the friction velocity in accordance with the DNS results and experimental data in Ref. [30].

Thus, incorporating Dean's correlation gives the empirical expression to quantify the dependence of the contribution on the Reynolds number from direct viscous dissipation, i.e.,

(8)

Alternatively, it can also be formulated with respect to the friction Reynolds number, that is,

(9)

Equations (8) and (9) offer an explicit way to predict the ratio of Cf1 to the total skin friction coefficient in incompressible turbulent channel flows without any flow-field simulations, especially at very high Reynolds numbers. Apparently, the direct viscous dissipation is negatively correlated with the Reynolds number. Once Re or Reτ reaches infinity, the viscous dissipation vanishes, and the skin friction is entirely generated by the turbulent energy dissipation.

With the aim of guiding the application of drag reduction, the boundary layer is further divided into three sub-regions separated by y+=30 and y=0.3. Figure 8 shows the contributive proportion of each sub-region to Cf2 at different Reynolds numbers, based on the FIK identity and the RD identity.

Fig. 8 Contributions of sub-regions to Cf2 based on (a) FIK identity and (b) RD identity

Figures 8(a) and 8(b) share similarities in terms of the tendencies of Cf2sub/Cf2. As the Reynolds number increases, a dramatic reduction of the contribution from the region y+ < 30 and y>0.3 is found, though they are non-negligible at low Reynolds numbers. This is consistent with the phenomenon that a smaller portion of Cf emerges in the near-wall region when the Reynolds number increases, as presented in Figs. 3 and 5. The differences between these two methods are limited in terms of the relative importance of these two regions, which is due to their essentially different interpretation of skin friction generation, as previously indicated. For the region where y+>30 and y < 0.3, its contribution increases significantly with the Reynolds number, which is ascribed to the energy superposition of large-scale turbulent motions[36].

The contributions from the region where y+>30 and y < 0.3 account for more than 60% of Cf2 at Reτ=5 200. Thus, at high Reynolds numbers, it is critically important to suppress turbulent fluctuations in this region, although this might result in some technical difficulties hitherto. Meanwhile, it is noteworthy that, based on the FIK identity, the near-wall contributions (y+ < 30) are almost negligible at Reτ= 5 200 while based on the RD identity, the near-wall region still accounts for nearly 30% of Cf2. Numerous studies have confirmed that artificial manipulation of the near-wall structures is able to achieve high drag reduction efficiency even when Re is very high[37-38]. As for the decay of the drag reduction efficiency previously proposed, it is likely to be associated with large-scale turbulent motions. Thus, at high Reynolds numbers, the inhibition of the large-scale amplitude modulation on the small scales and the outer site energy superposition onto the inner region should be investigated in depth to achieve more efficient drag reduction via control of the near-wall structures.

5 Conclusions

This paper focuses on decomposition methods for mean skin friction in turbulent channel flows. The effects of the Reynolds number on the mean skin friction generation are investigated using both the FIK identity and RD identity at the friction Reynolds numbers Reτ = 180, 550, 1 000, 2 000, and 5 200. The FIK and RD identities are two classical methods for the skin friction decomposition. Although some similarities can be found in the tendencies of Cf1/Cf and Cf2/Cf with regard to the Reynolds number, they are essentially different, especially in terms of the interpretation of their contributions to skin friction generation.

In the case of the FIK identity, it is confirmed that the linearly weighted Reynolds stress plays a dominant role in terms of skin friction generation. Moreover, at higher Reynolds number, its contribution becomes more important. The near-wall turbulent motions play a less important role in the skin friction generation and can almost be neglected at high Reynolds numbers.

In the case of the RD identity, the mean skin friction is decomposed from the perspective of energy transfer from the moving wall to the fluid in an absolute reference frame, by means of molecular viscous dissipation Cf1, RD and turbulent energy production (or dissipation) Cf2, RD. It is determined that Cf1, RD mostly originates from the region y+ < 30 and is negatively correlated with the Reynolds number. A good coincidence of Cf1, RD/Cf1.5 is observed regardless of the Reynolds number. Once the friction Reynolds number exceeds 550, Cf2, RD becomes the major component, accompanied by a secondary peak of the pre-multiplied integrand of Cf2, RD/Cf, as observed in the outer layer. The increasingly important contributions of Cf2, RD at high Reynolds numbers are ascribed to the generation of large-scale turbulent motions, which contain large amounts of turbulent kinetic energy and possibly lead to the redistribution of the turbulent kinetic energy from the near-wall region to the outer region. In contrast with the coincidence of Cf1, RD/Cf1.5, Cf2, RD/Cf1.5 has a logarithmic relationship with Reτ, and is well fitted by Cf2, RD/Cf1.5=1.87ln(Reτ)-5.29. Thus, empirical expressions for Cf1, RD/Cf are proposed, which allow for prediction of the drag-component without any pre-computation of the flow field.

The boundary layer is divided into three sub-regions, separated by y+=30 and y=0.3, to evaluate the contributive proportion of each sub-region to Cf2 for different Reynolds numbers. The results suggest that as the Reynolds number increases, a dramatic reduction of the contribution from the region y+ < 30 and y>0.3 occurs. For the RD identity, at the largest Reynolds number investigated, there is 30% contribution to Cf2 which originates from the region where y+ < 30. This indicates that at high Reynolds numbers, the large-scale amplitude modulation on the small scales and the outer site energy superposition onto the inner region should be restricted to achieve more efficient drag reduction.

Appendix A

In order to compare the turbulent kinetic energy production with dissipation over the entire boundary layer, Fig. A1 shows the distributions of turbulent production P and dissipation -ε as a function of y+. Integrations of P and -ε over the entire boundary layer give the total turbulent kinetic energy production Pt and the dissipation εt, which are listed in Table A1. It can be determined that the relative errors of (Pt-εt)/Pt are limited within 0.2%.

Acknowledgements  The large-scale computations were supported by the Center for High- Performance Computing, Shanghai Jiao Tong University.

References
[1]
GARIÉPY, M., TRÉPANIER, J. Y., and MALOUIN, B. Generalization of the far-field drag decomposition method to unsteady flows. AIAA Journal, 51(6), 1309-1319 (2013) doi:10.2514/1.J051609
[2]
GAD-EL-HAK, M. Interactive control of turbulent boundary layers:a futuristic overview. AIAA Journal, 32(9), 1753-1765 (1994) doi:10.2514/3.12171
[3]
DECK, S., RENARD, N., LARAUFIE, R., and WEISS, P.É. Large-scale contribution to mean wall shear stress in high-Reynolds-number flat-plate boundary layers up to Reθ=13650. Journal of Fluid Mechanics, 743, 202-248 (2014) doi:10.1017/jfm.2013.629
[4]
FUKAGATA, K., IWAMOTO, K., and KASAGI, N. Contribution of Reynolds stress distribution to the skin friction in wall-bounded flows. Physics of Fluids, 14(11), L73-L76 (2002) doi:10.1063/1.1516779
[5]
GOMEZ, T., FLUTET, V., and SAGAUT, P. Contribution of Reynolds stress distribution to the skin friction in compressible turbulent channel flows. Physical Review E, 79(3), 035301 (2009)
[6]
BANNIER, A., GARNIER, É., and SAGAUT, P. Riblet flow model based on an extended FIK identity. Flow, 95, 351-376 (2015)
[7]
PEET, Y. and SAGAUT, P. Theoretical prediction of turbulent skin friction on geometrically complex surfaces. Physics of Fluids, 21(10), 105105 (2009) doi:10.1063/1.3241993
[8]
MEHDI, F. and WHITE, C. M. Integral form of the skin friction coefficient suitable for experimental data. Experiments in Fluids, 50(1), 43-51 (2011) doi:10.1007/s00348-010-0893-1
[9]
MEHDI, F., JOHANSSON, T. G., WHITE, C. M., and NAUGHTON, J. W. On determining wall shear stress in spatially developing two-dimensional wall-bounded flows. Experiments in Fluids, 55, 1656 (2014) doi:10.1007/s00348-013-1656-6
[10]
LI, F. C., KAWAGUCHI, Y., SEGAWA, T., and HISHIDA, K. Reynolds-number dependence of turbulence structures in a drag-reducing surfactant solution channel flow investigated by particle image velocimetry. Physics of Fluids, 17(7), 075104 (2005) doi:10.1063/1.1941366
[11]
DE GIOVANETTI, M., HWANG, Y., and CHOI, H. Skin-friction generation by attached eddies in turbulent channel flow. Journal of Fluid Mechanics, 808, 511-538 (2016) doi:10.1017/jfm.2016.665
[12]
YOON, M., AHN, J., HWANG, J., and SUNG, H. J. Contribution of velocity-vorticity correlations to the frictional drag in wall-bounded turbulent flows. Physics of Fluids, 28(8), 081702 (2016) doi:10.1063/1.4961331
[13]
RENARD, N. and DECK, S. A theoretical decomposition of mean skin friction generation into physical phenomena across the boundary layer. Journal of Fluid Mechanics, 790, 339-367 (2016) doi:10.1017/jfm.2016.12
[14]
KIM, J. Control of turbulent boundary layers. Physics of Fluids, 15(5), 1093-1105 (2003) doi:10.1063/1.1564095
[15]
WHITE, C. M. and MUNGAL, M. G. Mechanics and prediction of turbulent drag reduction with polymer additives. Annual Review Fluid Mechanics, 40, 235-256 (2008) doi:10.1146/annurev.fluid.40.111406.102156
[16]
TOUBER, E. and LESCHZINER, M. A. Near-wall streak modification by spanwise oscillatory wall motion and drag-reduction mechanisms. Journal of Fluid Mechanics, 693, 150-200 (2012) doi:10.1017/jfm.2011.507
[17]
KAMETANI, Y. and FUKAGATA, K. Direct numerical simulation of spatially developing turbulent boundary layers with uniform blowing or suction. Journal of Fluid Mechanics, 681, 154-172 (2011) doi:10.1017/jfm.2011.219
[18]
BHUSHAN, B. Shark-skin surface for fluid-drag reduction in turbulent flow. Biomimetics, Springer, Cham, 227-265(2016)
[19]
CHOI, H., MOIN, P., and KIM, J. Direct numerical simulation of turbulent flow over riblets. Journal of Fluid Mechanics, 255, 503-539 (1993) doi:10.1017/S0022112093002575
[20]
GATTI, D., QUADRIO, M., and FROHNAPFEL, B. Reynolds number effect on turbulent drag reduction. 5th European Turbulence Conference, University of Delft, Delft (2015)
[21]
ROBINSON, S. K. The Kinematics of Turbulent Boundary Layer Structure, National Aeronautics and Space Administration, Washington, D. C. (1991)
[22]
ABBASSI, M. R., BAARS, W. J., HUTCHINS, N., and MARUSIC, I. Skin-friction drag reduction in a high-Reynolds-number turbulent boundary layer via real-time control of large-scale structures. International Journal of Heat and Fluid Flow, 67, 30-41 (2017) doi:10.1016/j.ijheatfluidflow.2017.05.003
[23]
HWANG, Y. Near-wall turbulent fluctuations in the absence of wide outer motions. Journal of Fluid Mechanics, 723, 264-288 (2013) doi:10.1017/jfm.2013.133
[24]
LEE, M. and MOSER, R. D. Direct numerical simulation of turbulent channel flow up to Reτ ≈ 5200. Journal of Fluid Mechanics, 774, 395-415 (2015) doi:10.1017/jfm.2015.268
[25]
KIM, J., MOIN, P., and MOSER, R. Turbulence statistics in fully developed channel flow at low Reynolds number. Journal of Fluid Mechanics, 177, 133-166 (1987) doi:10.1017/S0022112087000892
[26]
LEE, M., MALAYA, N., and MOSER, R. D. Petascale direct numerical simulation of turbulent channel flow on up to 786 K cores. Proceedings of the International Conference on High Performance Computing, Networking, Storage and Analysis, Association for Computing Machinery, New York (2013)
[27]
LEE, M., ULERICH, R., MALAYA, N., and MOSER, R. D. Experiences from leadership computing in simulations of turbulent fluid flows. Computing in Science & Engineering, 16(5), 24-31 (2014)
[28]
OLIVER, T. A., MALAYA, N., ULERICH, R., and MOSER, R. D. Estimating uncertainties in statistics computed from direct numerical simulation. Physics of Fluids, 26(3), 035101 (2014) doi:10.1063/1.4866813
[29]
DEAN, R. B. Reynolds number dependence of skin friction and other bulk flow variables in twodimensional rectangular duct flow. Journal of Fluids Engineering, 100(2), 215-223 (1978) doi:10.1115/1.3448633
[30]
ABE, H. and ANTONIA, R. A. Relationship between the energy dissipation function and the skin friction law in a turbulent channel flow. Journal of Fluid Mechanics, 798, 140-164 (2016) doi:10.1017/jfm.2016.299
[31]
LAADHARI, F. Reynolds number effect on the dissipation function in wall-bounded flows. Physics of Fluids, 19(3), 038101 (0381)
[32]
HINZE, J. O. Turbulence, McGraw-Hill, New York (1975)
[33]
KIM, K. C. and ADRIAN, R. J. Very large-scale motion in the outer layer. Physics of Fluids, 11(2), 417-422 (1999) doi:10.1063/1.869889
[34]
LIU, Z., ADRIAN, R. J., and HANRATTY, T. J. Large-scale modes of turbulent channel flow:transport and structure. Journal of Fluid Mechanics, 448, 53-80 (2001)
[35]
HWANG, J. and SUNG, H. J. Influence of large-scale motions on the frictional drag in a turbulent boundary layer. Journal of Fluid Mechanics, 829, 751-779 (2017) doi:10.1017/jfm.2017.579
[36]
BALAKUMAR, B. J. and ADRIAN, R. J. Large- and very-large-scale motions in channel and boundary-layer flows. Philosophical Transactions of the Royal Society of London A:Mathematical, Physical and Engineering Sciences, 1852(365), 665-671 (2007)
[37]
IWAMOTO, K., FUKAGATA, K., KASAGI, N., and SUZUKI, Y. Friction drag reduction achievable by near-wall turbulence manipulation at high Reynolds numbers. Physics of Fluids, 17(1), 011702 (2005) doi:10.1063/1.1827276
[38]
DENG, B. Q. Research on Mechanism of Drag-Reduction Control Based on Coherent Structures in Wall-Turbulence (in Chinese), Ph. D. dissertation, Tsinghua University (2014)