Lee 2015
Lee 2015
1. Introduction
Recently, relatively ‘high’-Reynolds-number wall-bounded turbulence has been
investigated with the help of innovations in measurement technologies (Nagib et al.
2004; Kunkel & Marusic 2006; Westerweel, Elsinga & Adrian 2013; Bailey et al.
2014) and computation (Borrell, Sillero & Jiménez 2013; El Khoury et al. 2013;
Lee, Malaya & Moser 2013). Because of the simplicity in geometry and boundary
conditions, pressure-driven turbulent flow between parallel walls (channel flow) is
an excellent vehicle for the study of wall-bounded turbulence via direct numerical
simulation (DNS). Since Kim, Moin & Moser (1987) showed agreement between
DNS and experiments in channel flow in 1987, DNS of channel flow has been used
to study wall-bounded turbulence at ever higher Reynolds numbers, as advances in
computing power have allowed. For example, channel flow DNS has been performed
at Reτ = 590 (Moser, Kim & Mansour 1999), Reτ = 950 (Del Álamo et al. 2004) and
Reτ = 2000 (Hoyas & Jiménez 2006). More recently, simulations with Reτ ≈ 4000
were performed separately by Lozano-Durán & Jiménez (2014) in a relatively small
2. Simulation details
In the discussion to follow, the streamwise, wall-normal and spanwise velocities will
be denoted u, v and w respectively, with the mean velocity indicated by a capital letter
and fluctuations by a prime. Furthermore, h·i indicates the expected value or average.
Thus, U = hui and u = U + u0 .
The simulations reported here are DNS of incompressible turbulent flow between
two parallel planes. Periodic boundary conditions are applied in the streamwise (x) and
spanwise (z) directions, and no-slip/no-penetration boundary conditions are applied at
the wall. The computational domain sizes are Lx = 8πδ and Lz = 3πδ, where δ is the
channel half-width, so the domain size in the wall-normal (y) direction is 2δ. The flow
is driven by a uniform pressure gradient, which varies in time to ensure that the mass
flux through the channel remains constant.
A Fourier–Galerkin method is used in the streamwise and spanwise directions, while
the wall-normal direction is represented using a B-spline collocation method (Kwok,
Moser & Jiménez 2001; Botella & Shariff 2003). The Navier–Stokes equations
are solved using the method of Kim et al. (1987), in which equations for the
wall-normal vorticity and the Laplacian of the wall-normal velocity are time-advanced.
This formulation has the advantage of satisfying the continuity constraint exactly
while eliminating the pressure. A low-storage implicit–explicit scheme (Spalart,
Moser & Rogers 1991) based on third-order Runge–Kutta for the nonlinear terms
and Crank–Nicolson for the viscous terms is used to advance in time. The time
step is adjusted to maintain an approximately constant Courant–Friedrichs–Lewy
(CFL) number of one. A new highly optimized code was developed to solve the
Navier–Stokes equations using these methods on advanced petascale computer
architectures. For more details about the code, the numerical methods and how
the simulations were run see Lee et al. (2013, 2014).
The simulations performed here were conducted with resolution comparable to that
used in previous high-Reynolds-number channel flow simulations, when measured in
wall units. Normalization in wall units, that is with kinematic viscosity ν and √
friction
velocity uτ , is indicated with a superscript ‘+’. The friction velocity is uτ = τw /ρ,
where τw is the mean wall shear stress and ρ is density. For the highest-Reynolds-
number simulation reported here, which is designated LM5200, Nx = 10 240 and Nz =
7680 Fourier modes were used to represent the streamwise and spanwise directions,
which results in an effective resolution of 1x+ = Lx+ /Nx = 12.7 and 1z+ = Lz+ /Nz =
6.4. In the wall-normal direction, the seventh-order B-splines are defined on a set
of 1530 = Ny − 6 knot points (Ny is the number of B-spline basis functions), which
398 M. Lee and R. D. Moser
y sin(ηξ π/2)
= , −1 6 ξ 6 1. (2.1)
δ sin(ηπ/2)
The single parameter η in this mapping controls how strongly the knot points are
clustered near the wall, with the strongest clustering occurring when η = 1. In
LM5200, η = 0.97, resulting in the first knot point from the wall at 1y+ w = 0.498 and
the centreline knot spacing of 1y+c = 10.3. The N y collocation points are determined as
the Greville abscissae (Johnson 2005), in which, for n-degree splines, each collocation
point is the average of n consecutive knot points (n = 7 here), with the knots at the
boundary given a multiplicity of n − 1. As a result, the collocation points are
more clustered near the wall than the knot points. Resolution parameters for all the
simulations discussed here are provided in table 1.
Because the mass flux in the channel remains constant, the bulk Reynolds
number RebR can be specified directly for a simulation, where Reb = Ub δ/ν,
δ
Ub = (1/2δ) −δ U(y) dy and U(y) is the mean streamwise velocity. Four simulations
at four different Reynolds numbers were conducted. Of most interest here is the
highest-Reynolds-number case, LM5200, for which the bulk Reynolds number is
Reb = 1.25 × 105 and the friction Reynolds number is Reτ = 5186 (Reτ = uτ δ/ν).
The three other cases simulated, LM180, LM550 and LM1000, were performed
for convenience to regenerate data for previously simulated cases (Kim et al. 1987;
Moser et al. 1999; Del Álamo et al. 2004) using the numerical methods used here.
The simulation details for each case are summarized in table 1.
In addition, channel flow data from four other sources in the literature are included
here for comparison. The first is a simulation at Reτ = 2000 conducted by Hoyas &
Jiménez (2006, HJ2000), which used the same domain size and a similar numerical
scheme to the current simulations, differing only in the use of high-order compact
finite differences in the wall-normal direction, rather than B-splines. A second
simulation (LJ4200) by Lozano-Durán & Jiménez (2014) was at Reτ = 4179 and
DNS of turbulent channel flow up to Reτ ≈ 5200 399
0.005 0.002
Residual
0 0
–0.005 –0.002
0 0.2 0.4 0.6 0.8 1.0 10 0 10 1 10 2 10 3 10 4
F IGURE 1. (Colour online) Statistical stationarity of simulation. (a) The residual in (2.2),
with the linestyle legend given in table 1; the dotted line is the standard deviation of
the estimated total stress in LM5200. (b) The residual in (2.3) of LM5200 (solid), and
the standard deviation of the estimated statistical error in the sum of the right-hand side
terms in (2.3) (dashed).
used the same numerical methods as HJ2000, but the domain size in x and z was
much smaller. The third simulation (BP04100), which was performed by Bernardini
et al. (2014), was at Reτ = 4079 and used a domain size not much smaller than that
used here, but these simulations were performed using second-order finite differences.
Finally, experimental data from laser Doppler velocimetry measurements at two
Reynolds numbers (Reτ = 4048 and 5895, SF4000 and SF6000 respectively) are
reported by Schultz & Flack (2013) and are also included here for comparison.
Summaries of all these data sources are also included in table 1.
We used the method described by Oliver et al. (2014) to estimate the uncertainty in
the statistics reported here due to sampling noise. For LM5200, the estimated standard
deviation of the mean velocity is less than 0.2 % and the estimated standard deviation
of the variance and covariance of the velocity components is less than 0.5 % in the
near-wall region (y+ < 100 say) and 3 % in the outer region (y > 0.2δ say). We also
used the total stress and the Reynolds stress transport equations to test whether the
simulated turbulence was statistically stationary. In a statistically stationary turbulent
channel, the total stress, which is the sum of Reynolds stress and mean viscous stress,
is linear due to momentum conservation:
∂U + y
+
− hu0 v 0 i+ ≈ 1 − . (2.2)
∂y δ
As shown in figure 1(a), the discrepancy between the analytic linear profile and total
stress profile from the simulations is less than 0.002 (in plus units) in all simulations,
and it is much smaller than the standard deviation of the estimated total stress of
LM5200.
The Reynolds stress transport equations, which govern the evolution of the Reynolds
stress tensor, are given by
Dhu0i u0j i ∂hu0i u0j u0k i ∂ 2 hu0i u0j i
∂Uj ∂Ui
= − hu0i u0k i + hu0j u0k i − +ν
Dt ∂xk ∂xk ∂xk ∂xk ∂xk
0 0 0 0 0
∂hp ui i ∂hp uj i
0 0 0
∂ui ∂uj ∂ui ∂uj
0
+ p + − + − 2ν . (2.3)
∂xj ∂xi ∂xj ∂xi ∂xk ∂xk
400 M. Lee and R. D. Moser
While not reported here, all the terms on the right-hand side of (2.3) have
been computed from our simulations, and the data are available at http://turbulence.
ices.utexas.edu. In a statistically stationary channel flow, the substantial derivative on
the left-hand side of (2.3) is zero. Hence, any deviation from zero of the sum of the
terms on the right-hand side of (2.3) is an indicator that the flow is not stationary.
The residual of (2.3) is shown in figure 1(b) (solid lines) in wall units. The values
for all components of the Reynolds stress are much less than 0.0001 in wall units,
which is less than 0.01 % error in the balance near the wall. The relative error in the
balance increases to of the order of 1 % away from the wall, as the magnitude of the
terms in (2.3) decreases. Across the entire channel, the estimated standard deviation
of the statistical noise (dashed lines) is much larger than these discrepancies.
3. Results
3.1. Mean velocity profile
The multiscale character of wall-bounded turbulence, in which ν/uτ is the length
scale relevant to the near-wall flow and δ applies to the flow away from the wall, is
well known. As first noted by Millikan (1938), this scaling behaviour and asymptotic
matching lead to the logarithmic variation of mean streamwise velocity with the
distance from the wall in an overlap region between the inner and outer flows. This
‘log law’ is given by
1
U + = log y+ + B, (3.1)
κ
where κ is the von Kármán constant. In a log layer, the indicator function
∂U +
β(y+ ) = y+ (3.2)
∂y+
is constant and equal to 1/κ. Hence, the indicator function, β, will have a plateau if
there is a logarithmic layer.
The mean streamwise velocity profile is shown in figure 2 for all the data sets
listed in table 1. The profiles from all the relatively high-Reynolds-number cases
are consistent, as expected. Despite this agreement, the indicator function β shows
some disagreement between the three highest-Reynolds-number simulations, as shown
in figure 3(a). In the LM5200 case, β is approximately flat between y+ = 350 and
y/δ = 0.16 (y+ = 830), indicating a log layer in this region. The LJ4200 case also
appears to be converging towards a plateau in this region, but there is apparent
statistical noise in the profile, which is understandable given the small domain size.
However, the BPO4100 simulation does not have a plateau in β. There is also a small
discrepancy between BPO4100 and the other two cases from y+ ≈ 30 to 100. These
discrepancies are significantly larger than the statistical uncertainty in the value of
β (approximately 0.2 %) in the current simulations, and presumably in the BPO4100
simulations, as the averaging time and domain sizes are comparable.
The values of κ and B in (3.1) were determined by fitting the mean velocity data
from LM5200 in the region between y+ = 350 and y/δ = 0.16 to obtain the values of
0.384 ± 0.004 and 4.27 respectively, with R2 = 0.9999, where R2 is the coefficient of
determination, which is one for a perfect fit. The value of κ agrees with the value
computed by Lozano-Durán & Jiménez (2014), but shows a slight discrepancy with
the values measured by Monty (2005) and Nagib & Chauhan (2008), which are κ =
0.37 and 0.39 respectively. The value of κ obtained here is remarkably similar to those
DNS of turbulent channel flow up to Reτ ≈ 5200 401
30
25
20
15
10
5
0
10 0 10 1 10 2 10 3 10 4
F IGURE 2. (Colour online) Mean streamwise velocity profile for all the cases listed in
table 1, where the legend of line styles and symbols is also given.
F IGURE 3. (Colour online) Log-law indicator function β for (a) the highest-Reynolds-
number simulations, LM5200, LJ4200 and BPO4100, (b) simulations at Reτ = 5186, 2003
and 1000 (LM5200, HJ2000 and LM1000), and (c) simulation LM5200 along with the
expressions ((3.3), star) and ((3.4), diamond). In (a), the horizontal dashed line is at β =
1/κ = 1/0.384, and in (c) parameter values (see table 2) fitted from LM5200 are solid, and
those from Jiménez & Moser (2007), Mizuno & Jiménez (2011) are dashed. The linestyle
legend for (a) and (b) is given in table 1.
reported by Österlund et al. (2000), κ = 0.38, and Nagib & Chauhan (2008), κ =
0.384, in the zero-pressure-gradient boundary layer. However, the value of κ reported
here is smaller than κ = 0.40 measured by Bailey et al. (2014) in pipe flow. It should
be noted that the choice of the range for this curve fit is somewhat arbitrary, since
the indicator function is not exactly flat (figure 3c).
From an asymptotic analysis perspective, the log-law relation (3.1) is the lowest-
order truncation of a matched asymptotic expansion in 1/Reτ (Afzal & Yajnik 1973;
Jiménez & Moser 2007). Several higher-order representations of the mean velocity
in the overlap region have been evaluated based on experimental and DNS data in
boundary layers, channels and pipes (Buschmann & Gad-el-Hak 2003; Jiménez &
Moser 2007; Mizuno & Jiménez 2011). Here, we consider two such applications to
channels in the context of the LM5200 data.
Jiménez & Moser (2007) considered a higher-order truncation in which β has the
form +
1 y+
+ ∂U α1
β =y = + + α2 . (3.3)
∂y+ κ∞ Reτ Reτ
402 M. Lee and R. D. Moser
This formulation essentially allows for an Reτ dependence of κ = (1/κ∞ + α1 /Reτ )−1 ,
and introduces a linear dependence on y/δ = y+ /Reτ . Based on data from a simulation
at Reτ ≈ 1000 by Del Álamo et al. (2004, similar to LM1000) and from HJ2000
(see figure 3b), they determined the parameter values shown in table 2. Further, this
form and these values were found to be consistent with experimental measurements
by Christensen & Adrian (2001) at Reynolds numbers up to Reτ = 2433.
Mizuno & Jiménez (2011) considered a different higher-order asymptotic truncation,
for which β is given by
∂U + y+ y+2
β = y+ = + a2 2 . (3.4)
∂y+ κ(y − a1 )
+ Reτ
The second term was motivated by the form of a wake model that is quadratic
for small y/δ, where the coefficient is related to the wake parameter Π by
a2 = (12Π − 2)/κ. The term a1 is an offset (virtual origin) which accounts for
the presence of the viscous layer (Wosnik, Castillo & George 2000). To first order
in a1 /y+ , the offset is equivalent to including an additive a1 /κy+ term, which is
expected from the matched asymptotics. They fitted the inverse of the mean velocity
derivative to y+ /β from (3.4) using the experimental and DNS data mentioned
above and the experiments of Monty (2005), with up to Reτ = 3945, to determine a
Reynolds-number-dependent value of the parameters. When evaluated for Reτ = 5186,
these yield the parameter values shown in table 2.
These two higher-order truncations have also been fitted to the LM5200 data to
obtain values shown in table 2 that are significantly different from the previously
determined values. The expressions for β from (3.3) and (3.4) are plotted in figure 3(c)
with both sets of parameters. It is clear from this figure that the parameter values
obtained by Jiménez & Moser (2007) and Mizuno & Jiménez (2011) do not fit the
LM5200 data, but the parameters fitted to the LM5200 data in the log region match
the data equally well for both truncation forms. The reason for this disagreement with
the parameters from Jiménez & Moser (2007) is clear, since the Reynolds numbers
used in that study were not high enough to exhibit the logarithmic region observed
in LM5200, since y/δ = 0.16 is at y+ = 320 when Reτ = 2000. They appear to have
been fitting (3.3) to the outer-layer profile, and indeed in LM5200 there is a region
0.16 < y/δ < 0.45 in which β is approximately linear in y+ /Reτ , with a slope of one
(figure 3b), in agreement with α2 = 1.0.
In contrast, the data used by Mizuno & Jiménez (2011) included a channel Reynolds
number as high as Reτ = 3945, which should have exhibited a short nearly constant β
plateau as observed in LM5200. This would have been qualitatively different from the
DNS of turbulent channel flow up to Reτ ≈ 5200 403
lower-Reynolds-number cases. This was not reported in that paper. However, at this
Reynolds number, the plotted values of κ and a1 (figure 5 in Mizuno & Jiménez 2011)
are significantly larger than the parameters for lower Reynolds number. Further, the
values for all three parameters obtained from the LM5200 data are within the indicated
uncertainties of the parameters from the Reτ = 3945 experimental data. Perhaps the
reported Reynolds number dependence of the parameters did not reflect a qualitative
change at the highest Reynolds number, because the fit was dominated by the more
numerous lower-Reynolds-number cases.
The apparent extent of the overlap region in the LM5200 case is not sufficient
to distinguish between the two asymptotic truncations, (3.3) and (3.4). Because a1
is so small, the primary distinction is in the lowest non-zero exponent on y+ /Reτ .
This may be of some importance because it determines the way in which the high-
Reynolds-number asymptote of constant β is approached. Unfortunately, this cannot
be determined using the available data.
hu02 i+
max = 3.66 + 0.642 log(Reτ ), (3.5)
9.5 LM1000
8
HJ1000
LJ4200
9.0 BPO4100
LM5200 6
MWT
8.5 PSP
4
8.0
7.5 2
7.0 0
10 3 10 4 10 5 10 –4 10 –3 10 –2 10 –1 10 0
(e) 8 ( f ) 2.0
6
4 1.5
2
1.0
0
–2 0.5
–4
–6 0
10 –4 10 –3 10 –2 10 –1 10 0 0.70 0.75 0.80 0.85 0.90 0.95 1.00
F IGURE 4. (Colour online) Variance of u: (a) as a function of y+ ; (b) zoom of (a) near the
peak; (c) dependence of maximum on Reτ ; solid line, relation (3.5); MWT, boundary layer
in the Melbourne Wind Tunnel (Hutchins et al. 2009; Kulandaivelu 2011); PSP, Princeton
Superpipe (Hultmark, Bailey & Smits 2010; Hultmark et al. 2012); (d) as a function of
y/δ; (e) test function for Townsend’s prediction; (f ) zoom of (d) near the centre of the
channel. The linestyle and symbol legend is given in table 1.
centreline (y/δ = 1), the variance from LJ4200 is significantly larger than from both
BPO4100 and LM5200, while the other simulations indicate that the variance should
be increasing slowly with Reynolds number. It appears that far from the wall, LJ4200
is affected by its relatively small domain size, as might be expected. Finally, the
experimental data with reported uncertainty, ±2 %, from SF4000 in figure 4(a) are
inconsistent with LJ4200 and BPO4100 in the region near the wall (y+ < 300 say).
The reason for this discrepancy is not clear, but it may be due to the difficulty of
measuring velocity fluctuations near the wall. Far from the wall, the experimental
DNS of turbulent channel flow up to Reτ ≈ 5200 405
F IGURE 5. (Colour online) Variance of v (a) and covariance of u and v (b). The linestyle
and symbol legend is given in table 1.
data for these quantities are consistent with the simulations. Measurements are not
available at small enough y+ to compare peak values of hu02 i+ .
Similar inconsistencies are present among the high-Reynolds-number simulation
cases in the wall-normal and spanwise velocity variances. Around the peaks of both
hv 02 i+ and hw02 i+ , BPO4100 exceeds values from LJ4200, despite its somewhat lower
Reynolds number (figures 5a and 6a), while near the centre, these two cases are
in agreement. Only the Reynolds shear stress hu0 v 0 i+ is in agreement in these cases
across all y (figure 5b). Similarly to hu02 i+ , the experimental data from SF4000 for
hv 02 i+ and hu0 v 0 i+ with uncertainties of ±3 % and ±5 % respectively, are inconsistent
with both simulations in the region near the wall, where the experimental data appear
to be quite noisy. Possible reasons for the minor inconsistencies noted among the
DNS are discussed in § 3.3.
There are several anticipated high-Reynolds-number features to be examined in
the data. In particular, similarly to the log law, Townsend’s attached-eddy hypothesis
(Townsend 1976) implies that in the high-Reynolds-number limit, there is an interval
in y in which the Reynolds stress components satisfy
hu02 i+ = A1 − B1 log(y/δ), (3.6a)
hv 02 i+ = A2 , (3.6b)
hw02 i+ = A3 − B3 log(y/δ), (3.6c)
hu0 v 0 i+ = −1. (3.6d)
Consistent with these relations, both hv 02 i+ and −hu0 v 0 i+ are developing a flat region
as the Reynolds number increases (figure 5), though the maximum Reynolds shear
stress is well below one (0.96). Because the total stress
∂U
τtot = ν − hu0 v 0 i (3.7)
∂y
+
is known analytically (τtot = 1 − y/δ) from the mean momentum equation, and because
β varies little over a broad range of y (β = 2.6 ± 0.4 ≈ 1/κ for 30 < y+ < 0.75Reτ )
(see figure 3), the variation with Reynolds number of the Reynolds shear stress near
its maximum can be deduced easily:
+ y+ 1
τRS = −hu0 v 0 i+ ≈ 1 − − + for 30 < y+ < 0.75Reτ . (3.8)
Reτ κy
406 M. Lee and R. D. Moser
0.5 0.5
0 0
–0.5 –0.5
–1.0 –1.0
–1.5 –1.5
10 0 10 1 10 2 10 3 10 4 10 –4 10 –3 10 –2 10 –1 10 0
From
√ this it is clear that the maximum Reynolds shear
√ stress is given by τRSmax ≈ 1 −
2/ κReτ , and that this maximum occurs at y+ ≈ Reτ /κ, as noted by Afzal (1982),
Morrison et al. (2004), Panton (2007) and Sillero et al. (2013). For the conditions
of LM5200 (Reτ = 5186, κ = 0.384), these estimates yield τRSmax ≈ 0.955 occurring
at y+ ≈ 116, in good agreement with the simulations. Further, the error in satisfying
τRS = 1 is less than for a range of y+ that increases in size like Reτ for large Reτ .
Precisely,
s
Re Re 4
y+ −
τ τ
1 − τRS < provided < 1− 2 . (3.9)
2 2 Reτ κ
Thus, for τRS to be within 5 % of one over a decade of variation of y+ would require
more than twice the Reynolds number of LM5200, and for it to be within 1 % at its
peak requires 20 times greater Reynolds number.
According to (3.6), both the variance of u and the variance of w would have a
logarithmic variation over some region of y. In figure 6, it appears that there is such a
logarithmic variation, even at Reynolds numbers as low as Reτ = 1000. The indicator
function y∂y hw02 i is approximately flat from y+ ≈ 100 to y+ ≈ 200 (figure 6c). The
corresponding curve fit is hw02 i+ = 1.08 − 0.387 log(y/δ), which is somewhat different
from the fit hw02 i+ = 0.8 − 0.45 log(y/δ) obtained by Sillero et al. (2013) in a boundary
DNS of turbulent channel flow up to Reτ ≈ 5200 407
1.0
0.5
–0.5
–1.0
10 0 10 1 10 2 10 3 10 4
layer DNS. On the other hand, there is no apparent logarithmic region in the hu02 i+
profiles. The indicator function y∂y hu02 i+ (figure 4e) is not flat anywhere in the domain.
There is, however, a region (200 < y+ < 0.6Reτ ) where the dependence of the indicator
function on y is linear with relatively small slope, and the slope may be decreasing
with Reynolds number, though extremely slowly. In contrast, Hultmark et al. (2012,
2013) observed a logarithmic region in hu02 i+ over the range 800 < y+ < 0.15Reτ in
pipe flow with Reτ > 2 × 104 . The LM5200 simulation is not at high enough Reynolds
number to exhibit such a region if it occurs over the same range in y, since 0.15Reτ <
800 at Reτ = 5186.
As mentioned in § 2, the terms in the Reynolds stress transport equations are
not reported here, though the data are available at http://turbulence.ices.utexas.edu.
Of interest here, however, is the transport equation for the turbulent kinetic energy
K = hu0i u0i i/2. Hinze (1975) argues that at sufficiently high Reynolds number, there
is an intermediate region between inner and outer layers where the transport terms
in the kinetic energy equation are small compared with production, so that in this
region
PK ≈ , (3.10)
where PK is the production of kinetic energy and is the dissipation. In the
formulation and analysis of turbulence models, (3.10) is often assumed to hold
in an overlap region between inner and outer layers (Durbin & Pettersson Reif
2010). The relative error in the balance of production and dissipation (PK / − 1)
for LM1000, HJ2000 and LM5200 is shown in figure 7. In all three cases, there
is a region y+ > 30 and y/δ < 0.6 in which the mismatch between production and
dissipation is of order 10 % or less, but there is no indication that the magnitude
of this mismatch is decreasing with Reynolds number. Indeed, there appears to be a
stable structure with a local minimum in PK / around y+ = 60 and a local maximum
around y+ = 300 followed by a gradual decline towards PK / = 1 with increasing y.
Presumably this decline will become more gradual with increasing Reynolds number
as y/δ = 0.6 increases in y+ .
(a) (b)
10 3 10 3
10 2 10 2
10 1 10 1
10 0 10 0
10 0 10 1 10 2 10 3 10 0 10 1 10 2 10 3
(c) (d)
10 3 10 3
10 2 10 2
10 1 10 1
10 0 10 0
10 0 10 1 10 2 10 3 10 0 10 1 10 2 10 3
1.5
0.2 1.0
0.5
0
10 0 10 1 10 2 0
10 0 10 1 10 2 10 3
F IGURE 9. (Colour online) The k−1 region: (a) kx Euu /u2τ at y+ = 90–170, (b) kz Euu /u2τ at
y+ = 3–14, (c) kz Euu /u2τ at y+ = 111–141.
REFERENCES
A FZAL , N. 1976 Millikan’s argument at moderately large Reynolds number. Phys. Fluids 19 (4),
600–602.
A FZAL , N. 1982 Fully developed turbulent flow in a pipe: an intermediate layer. Ing.-Arch. 52,
355–377.
A FZAL , N. & YAJNIK , K. 1973 Analysis of turbulent pipe and channel flows at moderately large
Reynolds number. J. Fluid Mech. 61, 23–31.
BAILEY, S. C. C., VALLIKIVI , M., H ULTMARK , M. & S MITS , A. J. 2014 Estimating the value of
von Kármán’s constant in turbulent pipe flow. J. Fluid Mech. 749, 79–98.
B ERNARDINI , M., P IROZZOLI , S. & O RLANDI , P. 2014 Velocity statistics in turbulent channel flow
up to Reτ = 4000. J. Fluid Mech. 742, 171–191.
B ORRELL , G., S ILLERO , J. A. & J IMÉNEZ , J. 2013 A code for direct numerical simulation of
turbulent boundary layers at high Reynolds numbers in BG/P supercomputers. Comput. Fluids
80, 37–43.
B OTELLA , O. & S HARIFF , K. 2003 B-spline methods in fluid dynamics. Intl J. Comput. Fluid Dyn.
17 (2), 133–149.
B USCHMANN , M. H. & G AD - EL -H AK , M. 2003 Generalized logarithmic law and its consequences.
AIAA J. 41 (1), 40–48.
C HRISTENSEN , K. T. & A DRIAN , R. J. 2001 Statistical evidence of hairpin vortex packets in wall
turbulence. J. Fluid Mech. 431, 433–443.
C OMTE -B ELLOT, G. 1963 Contribution à l’étude de la turbulence de conduite. PhD thesis, University
of Grenoble, France.
D EAN , R. B. & B RADSHAW, P. 1976 Measurements of interacting turbulent shear layers in a duct.
J. Fluid Mech. 78, 641–676.
D E G RAAFF , D. B. & E ATON , J. K. 2000 Reynolds-number scaling of the flat-plate turbulent boundary
layer. J. Fluid Mech. 422, 319–346.
D EL Á LAMO , J. C. & J IMÉNEZ , J. 2009 Estimation of turbulent convection velocities and corrections
to Taylor’s approximation. J. Fluid Mech. 640, 5–26.
D EL Á LAMO , J. C., J IMÉNEZ , J., Z ANDONADE , P. & M OSER , R. D. 2004 Scaling of the energy
spectra of turbulent channels. J. Fluid Mech. 500, 135–144.
D IXIT, S. A. & R AMESH , O. N. 2013 On the k1−1 scaling in sink-flow turbulent boundary layers.
J. Fluid Mech. 737, 329–348.
D URBIN , P. A. & P ETTERSSON R EIF , B. A. 2010 Statistical Theory and Modeling for Turbulent
Flows. Wiley.
E L K HOURY, G. K., S CHLATTER , P., N OORANI , A., F ISCHER , P. F., B RETHOUWER , G. &
J OHANSSON , A. V. 2013 Direct numerical simulation of turbulent pipe flow at moderately
high Reynolds numbers. Flow Turbul. Combust. 91 (3), 475–495.
F ERNHOLZ , H. H. & F INLEY, P. J. 1996 The incompressible zero-pressure-gradient turbulent boundary
layer: an assessment of the data. Prog. Aerosp. Sci. 32 (8), 245–311.
G UALA , M., H OMMEMA , S. E. & A DRIAN , R. J. 2006 Large-scale and very-large-scale motions in
turbulent pipe flow. J. Fluid Mech. 554, 521–542.
H INZE , J. O. 1975 Turbulence. McGraw-Hill.
DNS of turbulent channel flow up to Reτ ≈ 5200 413
H OYAS , S. & J IMÉNEZ , J. 2006 Scaling of the velocity fluctuations in turbulent channels up to
Reτ = 2003. Phys. Fluids 18 (1), 011702.
H OYAS , S. & J IMÉNEZ , J. 2008 Reynolds number effects on the Reynolds-stress budgets in turbulent
channels. Phys. Fluids 20 (10), 101511.
H ULTMARK , M., BAILEY, S. C. C. & S MITS , A. J. 2010 Scaling of near-wall turbulence in pipe
flow. J. Fluid Mech. 649, 103–113.
H ULTMARK , M., VALLIKIVI , M., BAILEY, S. C. C. & S MITS , A. J. 2012 Turbulent pipe flow at
extreme Reynolds numbers. Phys. Rev. Lett. 108, 094501.
H ULTMARK , M., VALLIKIVI , M., BAILEY, S. C. C. & S MITS , A. J. 2013 Logarithmic scaling of
turbulence in smooth- and rough-wall pipe flow. J. Fluid Mech. 728, 376–395.
H UTCHINS , N., C HAUHAN , K., M ARUSIC , I., M ONTY, J. & K LEWICKI , J. 2012 Towards reconciling
the large-scale structure of turbulent boundary layers in the atmosphere and laboratory.
Boundary-Layer Meteorol. 145 (2), 273–306.
H UTCHINS , N. & M ARUSIC , I. 2007 Large-scale influences in near-wall turbulence. Phil. Trans. R.
Soc. Lond. A 365 (1852), 647–664.
H UTCHINS , N., N ICKELS , T. B., M ARUSIC , I. & C HONG , M. S. 2009 Hot-wire spatial resolution
issues in wall-bounded turbulence. J. Fluid Mech. 635, 103–136.
J IMÉNEZ , J. & M OSER , R. D. 2007 What are we learning from simulating wall turbulence? Phil.
Trans. R. Soc. Lond. A 365 (1852), 715–732.
J OHANSSON , A. V. & A LFREDSSON , P. H. 1982 On the structure of turbulent channel flow. J. Fluid
Mech. 122, 295–314.
J OHNSON , R. W. 2005 Higher order B-spline collocation at the Greville abscissae. Appl. Numer.
Maths 52 (1), 63–75.
K IM , J., M OIN , P. & M OSER , R. 1987 Turbulence statistics in fully developed channel flow at low
Reynolds number. J. Fluid Mech. 177, 133–166.
K IM , K. C. & A DRIAN , R. J. 1999 Very large-scale motion in the outer layer. Phys. Fluids 11 (2),
417.
K ULANDAIVELU , V. 2011 Evolution and structure of zero pressure gradient turbulent boundary layer.
PhD thesis, University of Melbourne.
K UNKEL , G. J. & M ARUSIC , I. 2006 Study of the near-wall-turbulent region of the high-Reynolds-
number boundary layer using an atmospheric flow. J. Fluid Mech. 548, 375–402.
K WOK , W. Y., M OSER , R. D. & J IMÉNEZ , J. 2001 A critical evaluation of the resolution properties
of B-spline and compact finite difference methods. J. Comput. Phys. 174 (2), 510–551.
L EE , M., M ALAYA , N. & M OSER , R. D. 2013 Petascale direct numerical simulation of turbulent
channel flow on up to 786K cores. In Proceedings of SC13: International Conference for
High Performance Computing, Networking, Storage and Analysis. ACM.
L EE , M., U LERICH , R., M ALAYA , N. & M OSER , R. D. 2014 Experiences from leadership computing
in simulations of turbulent fluid flows. Comput. Sci. Engng 16 (5), 24–31.
L OZANO -D URÁN , A. & J IMÉNEZ , J. 2014 Effect of the computational domain on direct simulations
of turbulent channels up to Reτ = 4200. Phys. Fluids 26 (1), 011702.
M ARUSIC , I., M ATHIS , R. & H UTCHINS , N. 2010a High Reynolds number effects in wall turbulence.
Intl J. Heat Fluid Flow 31 (3), 418–428.
M ARUSIC , I., M C K EON , B. J., M ONKEWITZ , P. A., NAGIB , H. M., S MITS , A. J. & S REENIVASAN ,
K. R. 2010b Wall-bounded turbulent flows at high Reynolds numbers: recent advances and
key issues. Phys. Fluids 22 (6), 065103.
M ARUSIC , I., M ONTY, J. P., H ULTMARK , M. & S MITS , A. J. 2013 On the logarithmic region in
wall turbulence. J. Fluid Mech. 716, R3.
M ILLIKAN , C. B. 1938 A critical discussion of turbulent flows in channels and circular tubes. In
Proceedings of the 5th International Congress for Applied Mechanics, pp. 386–392. Wiley.
M IZUNO , Y. & J IMÉNEZ , J. 2011 Mean velocity and length-scales in the overlap region of wall-
bounded turbulent flows. Phys. Fluids 23 (8), 085112.
M OIN , P. 2009 Revisiting Taylor’s hypothesis. J. Fluid Mech. 640, 1–4.
414 M. Lee and R. D. Moser
M ONTY, J. P. 2005 Developments in smooth wall turbulent duct flows. PhD thesis, University of
Melbourne.
M ONTY, J. P. & C HONG , M. S. 2009 Turbulent channel flow: comparison of streamwise velocity
data from experiments and direct numerical simulation. J. Fluid Mech. 633, 461–474.
M ONTY, J. P., H UTCHINS , N., N G , H. C. H., M ARUSIC , I. & C HONG , M. S. 2009 A comparison
of turbulent pipe, channel and boundary layer flows. J. Fluid Mech. 632, 431–442.
M ORRISON , J. F., M C K EON , B. J., J IANG , W. & S MITS , A. J. 2004 Scaling of the streamwise
velocity component in turbulent pipe flow. J. Fluid Mech. 508, 99–131.
M OSER , R. D., K IM , J. & M ANSOUR , N. N. 1999 Direct numerical simulation of turbulent channel
flow up to Reτ = 590. Phys. Fluids 11 (4), 943–945.
N AGIB , H., C HRISTOPHOROU , C., R EUDI , J.-D., M ONKEWITZ , P., Ö STERLUN , J. & G RAVANTE ,
S. 2004 Can we ever rely on results from wall-bounded turbulent flows without direct
measurements of wall shear stress. In 24th AIAA Aerodynamic Measurement Technology and
Ground Testing Conference, Portland, Oregon, p. 2392. American Institute of Aeronautics and
Astronautics (AIAA).
NAGIB , H. M. & C HAUHAN , K. A. 2008 Variations of von Kármán coefficient in canonical flows.
Phys. Fluids 20 (10), 101518.
N ICKELS , T., M ARUSIC , I., H AFEZ , S. & C HONG , M. S. 2005 Evidence of the k1−1 law in a
high-Reynolds-number turbulent boundary layer. Phys. Rev. Lett. 95 (7), 074501.
N ICKELS , T. B., M ARUSIC , I., H AFEZ , S., H UTCHINS , N. & C HONG , M. S. 2007 Some predictions
of the attached eddy model for a high Reynolds number boundary layer. Phil. Trans. R. Soc.
Lond. A 365 (1852), 807–822.
O LIVER , T. A., M ALAYA , N., U LERICH , R. & M OSER , R. D. 2014 Estimating uncertainties in
statistics computed from direct numerical simulation. Phys. Fluids 26 (3), 035101.
Ö STERLUND , J. M., J OHANSSON , A. V., N AGIB , H. M. & H ITES , M. H. 2000 A note on the
overlap region in turbulent boundary layers. Phys. Fluids 12 (1), 1–4.
PANTON , R. L. 2007 Composite asymptotic expansions and scaling wall turbulence. Phil. Trans. R.
Soc. Lond. A 365 (1852), 733–754.
P ERRY, A. E., H ENBEST, S. & C HONG , M. S. 1986 A theoretical and experimental study of wall
turbulence. J. Fluid Mech. 165, 163–199.
ROSENBERG , B. J., H ULTMARK , M., VALLIKIVI , M., B AILEY, S. C. C. & S MITS , A. J. 2013
Turbulence spectra in smooth- and rough-wall pipe flow at extreme Reynolds numbers.
J. Fluid Mech. 731, 46–63.
S CHULTZ , M. P. & F LACK , K. A. 2013 Reynolds-number scaling of turbulent channel flow. Phys.
Fluids 25 (2), 025104.
S ILLERO , J. A., J IMÉNEZ , J. & M OSER , R. D. 2013 One-point statistics for turbulent wall-bounded
flows at Reynolds numbers up to δ + ≈ 2000. Phys. Fluids 25, 105102.
S MITS , A. J. & M ARUSIC , I. 2013 Wall-bounded turbulence. Phys. Today 66 (9), 25–30.
S MITS , A. J., M C K EON , B. J. & M ARUSIC , I. 2011 High-Reynolds number wall turbulence. Annu.
Rev. Fluid Mech. 43 (1), 353–375.
S PALART, P. R. & A LLMARAS , S. R. 1992 A one-equation turbulence model for aerodynamic flows.
In 30th Aerospace Sciences Meeting and Exhibit, p. 439. American Institute of Aeronautics
and Astronautics.
S PALART, P. R., M OSER , R. D. & ROGERS , M. M. 1991 Spectral methods for the Navier–Stokes
equations with one infinite and two periodic directions. J. Comput. Phys. 96 (2), 297–324.
T OWNSEND , A. A. 1976 The Structure of Turbulent Shear Flow, 2nd edn. Cambridge University
Press.
V REMAN , A. W. & K UERTEN , J. G. M. 2014 Comparison of direct numerical simulation databases
of turbulent channel flow at Reτ = 180. Phys. Fluids 26 (1), 015102.
W EI , T. & W ILLMARTH , W. W. 1989 Reynolds-number effects on the structure of a turbulent channel
flow. J. Fluid Mech. 204, 57–95.
W ESTERWEEL , J., E LSINGA , G. E. & A DRIAN , R. J. 2013 Particle image velocimetry for complex
and turbulent flows. Annu. Rev. Fluid Mech. 45 (1), 409–436.
DNS of turbulent channel flow up to Reτ ≈ 5200 415
W INKEL , E. S., C UTBIRTH , J. M., C ECCIO , S. L., P ERLIN , M. & D OWLING , D. R. 2012 Turbulence
profiles from a smooth flat-plate turbulent boundary layer at high Reynolds number. Exp. Therm.
Fluid Sci. 40, 140–149.
W OSNIK , M., C ASTILLO , L. & G EORGE , W. K. 2000 A theory for turbulent pipe and channel flows.
J. Fluid Mech. 421, 115–145.
W U , X., BALTZER , J. R. & A DRIAN , R. J. 2012 Direct numerical simulation of a 30R long turbulent
pipe flow at R+ = 685: large- and very large-scale motions. J. Fluid Mech. 698, 235–281.
Z ANOUN , E.-S., D URST, F. & NAGIB , H. 2003 Evaluating the law of the wall in two-dimensional
fully developed turbulent channel flows. Phys. Fluids 15 (10), 3079.
Z ANOUN , E.-S., N AGIB , H. & D URST, F. 2009 Refined cf relation for turbulent channels and
consequences for high-Re experiments. Fluid Dyn. Res. 41 (2), 021405.