scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Scaling of the energy spectra of turbulent channels

TL;DR: In this paper, the spectra and correlations of the velocity fluctuations in turbulent channels, especially above the buffer layer, were analyzed using direct numerical simulations with friction Reynolds numbers up to Re at very large ones.
Abstract: The spectra and correlations of the velocity fluctuations in turbulent channels, especially above the buffer layer, are analysed using new direct numerical simulations with friction Reynolds numbers up to Re at very large ones.

Summary (2 min read)

1. Introduction

  • While self-similarity is always a welcome feature in physical problems, simplifying their solution and providing insight into their behaviour, its failure in situations in which in principle it ought to apply is perhaps even more interesting, because it forces us to explain what went wrong.
  • T is the time during which the statistics were collected after discarding initial transients.
  • The results, and their relations to the different scaling arguments, are detailed in § 3, followed by a short concluding section.

2. The numerical experiments

  • The spatial discretization uses dealiased Fourier expansions in the wall-parallel planes, and Chebychev polynomials in y.
  • Figure 2 also shows that the misrepresentation of the large scales in the outer layer of the channel for case S950 does not affect substantially the resolved part of the uv-cospectrum, especially as the authors move toward the wall.

3. Results and discussion

  • A similar square-root behaviour was found for the width of the near-wall streaks by Jiménez et al. (2003) , who showed that its most probable cause was the spreading, under the effect of lateral fluctuations, of wakes formed by compact v-structures.
  • And the details of the spreading are probably somewhat different, the source of the square-root behaviour is in both cases the long-term dispersion by background turbulence (Townsend 1976, p. 337) .
  • The other spectral range in which the scaling is reasonably clear is region C in figure 3 (c), which corresponds to the 'global' modes described by Bullock, Cooper & Abernathy (1978) and by del Álamo & Jiménez (2003) as being correlated across the entire flow.
  • This is because, while the short detached structures are self-similar in the sense of being unaffected by boundary conditions, those in region C represent different levels in global eddies spanning the whole flow height, and have a definite vertical structure.

4. Conclusions

  • By using new results from direct simulations of turbulent channels at higher Reynolds numbers and in larger boxes than those available up to now, the authors have probed the corrections to the similarity assumptions in the logarithmic and outer layers of wall-bounded turbulence.
  • The resulting spectral model can be integrated to obtain a mixed scaling for the total intensity of the fluctuating velocity, which collapses the numerical results well with those of de Graaff & Eaton (2000) in the outer layer, and which predicts that this part of the flow will tend to scale with the mean-stream velocity at very high Reynolds numbers.
  • These results do not extend to the near-wall region.
  • According to this argument, the only influence of the wall region on the dynamics of the outer layer would come from changes of the velocity scale (3.5).
  • In Spain this work was supported in part by the CICYT contract BFM2000-1468.

Did you find this useful? Give us your feedback

Figures (6)

Content maybe subject to copyright    Report

J. Fluid Mech. (2004), vol. 500, pp. 135–144.
c
2004 Cambridge University Press
DOI: 10.1017/S002211200300733X Printed in the United Kingdom
135
Scaling of the energy spectra
of turbulent channels
By JUAN C. DEL
´
ALAMO
1
, JAVIER JIM
´
ENEZ
1,2
,
PAULO ZANDONADE
3
AND ROBERT D. MOSER
3
1
School of Aeronautics, Universidad Polit
´
ecnica de Madrid, 28040 Madrid, Spain
2
Centre for Turbulence Research, Stanford University, Stanford, CA 94305, USA
3
Department of Theoretical and Applied Mechanics, University of Illinois, Urbana, IL 61801, USA
(Received 26 June 2003, and in revised form 27 October 2003)
The spectra and correlations of the velocity fluctuations in turbulent channels,
especially above the buffer layer, are analysed using new direct numerical simulations
with friction Reynolds numbers up to Re
τ
= 1900. It is found, and explained, that
their scaling is anomalous in several respects, including a square-root behaviour of
their width with respect to their length, and a velocity scaling of the largest modes
with the centreline velocity U
c
. It is shown that this implies a logarithmic correction
to the k
1
energy spectrum, and that it leads to a scaling of the total fluctuation
intensities away from the wall which agrees well with the mixed scaling of de Graaff
& Eaton (2000) at intermediate Reynolds numbers, but which tends to a pure scaling
with U
c
at very large ones.
1. Introduction
While self-similarity is always a welcome feature in physical problems, simplifying
their solution and providing insight into their behaviour, its failure in situations in
which in principle it ought to apply is perhaps even more interesting, because it forces
us to explain what went wrong. Some of the best-known similarity arguments in fluid
dynamics are those regarding the overlap layer in wall turbulence, and it has lately
become increasingly clear that, while the logarithmic law for the mean velocity profile
may be a good approximation to the experimental data (Zagarola & Smits 1997;
¨
Osterlund et al. 2000), there are serious similarity failures in the behaviour of the
velocity fluctuations, including the scaling of their intensity in wall units and their
predicted k
1
energy spectrum.
The reason why the intensity of the streamwise velocity fluctuations does not scale
well with the friction velocity was given by Townsend (1976), who noted that wall-
parallel motions are inactive from the point of view of the Reynolds stresses, and do
not have to scale with them. The discovery by de Graaff & Eaton (2000) that they
scale well in mixed units was however unexpected, and presents a clear theoretical
challenge, because it implies that they should be describable by a well-defined model.
Even more interesting is the failure of the k
1
energy spectrum, which can be easily
predicted from an assumed lack of length scale for fluctuations whose wavelengths
are short with respect to the flow thickness, but long with respect to the wall
distance (Townsend 1976). It has been known since Laufer (1954) that the k
1
law
is approximately correct, but recent detailed observations by Hites (1997) and by
Morrison et al. (2002) reveal important corrections which are inconsistent with the
original argument.

136 J. C. del
´
Alamo, J. Jim
´
enez, P. Zandonade and R. D. Moser
Case Re
τ
L
x
/h L
z
/h TU
b
/L
x
x
+
z
+
y
+
c
N
x
N
z
N
y
Series 1 L550 547 8π 4π 10 8.94.56.7 1536 1536 257
L950 934 8π 3π 9.27.63.87.6 3072 2304 385
Series 2 S550 550 ππ/277 8.94.56.7 192 192 257
S950 964 ππ/227 7.83.97.8 384 384 385
S1900 1901 ππ/222 7.83.97.8 768 768 769
Tab le 1 . Parameters of the simulations. T is the time during which the statistics were collected
after discarding initial transients. U
b
is the bulk velocity. x and z are the collocation
resolutions parallel to the wall, using N
x
and N
z
points. y
c
is the wall-normal grid spacing
at the centre of the channel, and N
y
is the number of Chebychev polynomials.
The present paper looks at the reasons for those two failures, which we will find
to be related. We use data from new direct numerical simulations of plane turbulent
channels at moderate Reynolds numbers Re
τ
= u
τ
h/ν 6 1900, where u
τ
is the wall-
friction velocity and h is the channel half-width. The new data show that the original
failure of self-similarity occurs in the relation between the lengths and widths of the
structures in the overlap region, and that this failure results in logarithmic corrections
to the k
1
energy spectrum. These are then demonstrated using experimental data,
for which h will represent the pipe radius or the 99% boundary layer thickness.
We then show, and motivate, that the velocity scale of the largest structures is not
the friction velocity, but the mean velocity at the centreline, and use the resulting
spectral model to derive a modified mixed scaling law for the fluctuations.
The new simulations are described in § 2. The results, and their relations to the
different scaling arguments, are detailed in § 3, followed by a short concluding section.
The companion paper by Jim
´
enez, del
´
Alamo & Flores (2003) extends the present
analysis to the buffer and viscous layers, and that by del
´
Alamo et al. (2004) contains
a more detailed analysis of the structures away from the wall.
2. The numerical experiments
The numerical code integrates the Navier–Stokes equations in the form of evolution
problems for the wall-normal vorticity ω
y
and for the Laplacian of the wall-normal
velocity
2
v, as in Kim, Moin & Moser (1987). The spatial discretization uses dealiased
Fourier expansions in the wall-parallel planes, and Chebychev polynomials in y.The
streamwise and spanwise coordinates and velocity components are respectively x, z
and u, w. The temporal discretization is third-order semi-implicit Runge–Kutta, as in
Moser, Kim & Mansour (1999).
The present simulations, summarized in table 1, are divided into two series. The
runs in the first one use numerical boxes with very long periodicities L
x
and L
z
,to
try to account for all the energetic structures in the flow, including those in the outer
region whose size is proportional to h, and that could not be captured accurately in
previous simulations (Jim
´
enez 1998; Kim & Adrian 1999; del
´
Alamo & Jim
´
enez 2001,
2003).
The second series of experiments focuses on the overlap layer, increasing the
Reynolds number at the expense of the size of the numerical box, especially for the
case S1900. Less emphasis will be put on the other two simulations in this series,
which were performed mainly to assess the effect of small numerical boxes on the
computed statistics. That effect cannot necessarily be neglected, since the fractions

Spectra of turbulence in channels 137
Figure 1. (a) Mean velocity defect as a function of wall distance. (b)DeviationΨ of the
mean velocity from the logarithmic function. ------, L550;
, L950; , S550; , S950;
, S1900. The shaded region in (b) covers the maximum scatter of experimental boundary
layers from Smith (1994), Re
θ
= 4600–13000,
¨
Osterlund et al. (2000), Re
θ
= 2500–27000, and
de Graaff & Eaton (2000), Re
θ
= 1430–31000. The hatched area covers experimental pipes
from Perry et al. (1986), Re
τ
= 1600–3800, Durst et al. (1995), Re
τ
= 271–570 and Zagarola &
Smits (1997), Re
τ
= 1700–5.3 × 10
5
.Onlyy/h < 0.1 has been included in the experimental data
to avoid contamination by points in the outer layer.
Figure 2. Premultiplied one-dimensional uv-cospectra k|E
1D
uv
|
+
as functions of wavelength λ
and wall distance. Shaded contours, case L950; line contours, case S950. The contours are
0.2(0.2)0.8 times the maximum value from case L950. The dashed lines mark the size of the
smaller box.
of the Reynolds stress uv which are contained in the simulations L550 and L950
at scales larger than the size of the smaller boxes (i.e. in wavelengths either longer
than λ
x
= πh or wider than λ
z
= πh/2), are respectively 51% and 47% when averaged
across the whole channel.
This result agrees with recent evidence that there are important large-scale contribu-
tions to uv (Jim
´
enez 1998; del
´
Alamo & Jim
´
enez 2001; Liu, Adrian & Hanratty
2001), and is probably the reason why the mean-velocity defects from cases S550, S950
and S1900 do not scale as accurately as those from cases L550 and L950 (figure 1a).
However, figure 2 suggests that the large-scale contributions to uv take place mostly
in the outer region, in agreement with Townsend’s (1976) idea that the impermeability
of the wall limits the large-scale motions with a wall-normal component. This figure
displays the premultiplied one-dimensional uv-cospectra, k
ˆ
u(k, y)
ˆ
v
(k, y), where
ˆ
u
and
ˆ
v are the Fourier coefficients of u and v and k =2
π/λ are either streamwise or

138 J. C. del
´
Alamo, J. Jim
´
enez, P. Zandonade and R. D. Moser
Figure 3. (a, b) Two-dimensional spectral densities Φ
+
as functions of λ
x
/y and λ
z
/y.The
line contours are Φ
+
uu
=0.17. (a) y
+
= 150 and the symbols are Φ
+
vv
=0.06, (b) y/h=0.15 and
the symbols are Φ
+
ww
=0.08. These levels are roughly 5% of the corresponding intensities.
, , case L550; ——–, , L950; , , S1900. The solid straight line is λ
x
= λ
z
,
and the dashed one is λ
2
z
=2κλ
x
y.(c) Sketch of the organization of Φ
uu
, as discussed in text.
(d) Isocontours of correlation coefficient C
vv
=1/3 from case L950. Symbols, C
vv
(15
+
,y): ,
y
+
= 100; , y
+
= 200; ×, y
+
= 300. Lines, C
vv
(y/2,y): ——–, y
+
= 200; , y
+
= 300. The
correlation increases from left to right, and the straight lines are as in (a).
spanwise wavenumbers. Figure 2 also shows that the misrepresentation of the large
scales in the outer layer of the channel for case S950 does not affect substantially
the resolved part of the uv-cospectrum, especially as we move toward the wall. This
is also true for Re
τ
= 550, and it seems reasonable to expect the same to hold for
Re
τ
= 1900, particularly for the O(y) scales of the overlap region. This assumption is
confirmed by figure 1(b), which displays the deviations of the mean-velocity profiles
from a logarithmic law Ψ = U
+
κ
1
log(y
+
), using κ =0.4. It suggests that the
overlap region of case S1900 has approximately reached the typical behaviour of
high-Reynolds-number experimental wall ows, which are represented in the figure
by the hatched and shaded regions. That conclusion, which will be further confirmed
throughout the paper, is not very sensitive to the precise value chosen for κ.
3. Results and discussion
Figures 3(a)and3(b) show isolines of the premultiplied two-dimensional spectrum
of the streamwise velocity, Φ
uu
(k
x
,k
z
,y)=k
x
k
z
ˆ
u
ˆ
u
,aswellasofΦ
vv
and Φ
ww
.
Because of the range of Re
τ
in the numerical channels, figure 3(a) spans a factor of
more than 3 in y/h, while figure 3(b) does the same for y
+
. Both scale well with u
2
τ
and y in the range y<λ
x
< 10 y.

Spectra of turbulence in channels 139
However, the ridge of Φ
uu
lies along λ
z
(y λ
x
)
1/2
, and needs some discussion
because self-similarity would seem to require that λ
z
λ
x
. A similar square-root
behaviour was found for the width of the near-wall streaks by Jim
´
enez et al. (2003),
who showed that its most probable cause was the spreading, under the effect of
lateral fluctuations, of wakes formed by compact v-structures. Although the outer
flow is less coherent than the near-wall layer, and the details of the spreading are
probably somewhat different, the source of the square-root behaviour is in both cases
the long-term dispersion by background turbulence (Townsend 1976, p. 337).
Consider the random stirring of the mean velocity gradient by active eddies of size
O(y) and intensity O(u
τ
), which are represented by the spectral densities Φ
vv
and Φ
ww
in figures 3(a)and3(b). The expected lateral deviations of the fluid elements caused
by those eddies at long times is λ
z
(2 ν
T
t)
1/2
,whereν
T
κu
τ
y is the turbulent eddy
diffusivity in the logarithmic region. As the wakes lengthen, time is converted into
wavelength as λ
x
= Ut,whereU is the difference between the mean flow and the
advection velocity of the forcing fluctuations. There are few data on the advection
velocities of v and w in the logarithmic layer, but Kim & Hussain (1993) find that they
differ from the local mean velocity by at most a few wall units (see also del
´
Alamo
et al. 2004). This leads to the desired relation, λ
2
z
=2κλ
x
y/U
+
, which is represented
in figures 3(a)and3(b) with U
+
= 1. As the wakes widen they also grow in height,
although the y-dependence of ν
T
leads then to a linear growth.
Refer now to the diagram in figure 3(c). Besides the two square-root segments bc
and ad, Φ
uu
is bounded at short wavelengths by the segment ab along λ
z
λ
x
,which
represents the linear dispersion of fluid elements for separations shorter than the
integral scale of the active eddies (see below). This scale is of the order of λ
x
=10y,
and marks both the transition between the linear and square-root behaviours of Φ
uu
,
and the long-wave cut-offs of Φ
vv
and Φ
ww
. It is also the border between wall-attached
and wall-detached eddies. Consider the isocontours in figure 3(d) of the correlation
coefficient of
ˆ
v at two heights, defined as
C
vv
= |
ˆ
v(λ
x
, λ
z
,y
0
)
ˆ
v
(λ
x
, λ
z
,y)|/(|
ˆ
v(λ
x
, λ
z
,y
0
)|
2
|
ˆ
v(λ
x
, λ
z
,y)|
2
)
1/2
. (3.1)
The lines in figure 3(d) indicate that when y
0
y (y
0
= y/2 for the lines shown), and
both y and y
0
are in the logarithmic layer, the λ dependence of the correlations
scales with y. This breaks down however if λ is much larger than h. Similarly, with
y
0
fixed in the viscous layer (y
+
0
= 15 for the symbols in figure 3d), the wavelength
dependence also scales with y, but the region of significant correlation is shifted to
much larger λ
x
. This result, which was also observed for C
uu
and C
ww
, suggests that
there is a range of wavelengths which are correlated within the overlap layer, but not
with the wall, and which are therefore ‘detached’ in the sense of Townsend (1976).
The detached wavelength range extends between the two sets of contours in figure
3(d), and is bounded for large wavelengths at λ
x
10 y, because for such wavelengths,
correlations reach to the wall. It also contains all the energetic scales for v as shown in
figure 3(a). This is consistent with the idea that blocking by the wall limits the size of
the active wall-normal motions, and explains the transition from linear to square-root
behaviour of Φ
uu
. At long wavelengths the u-spectrum is no longer forced by the
active eddies, and only the incoherent square-root behaviour remains. The ridge of
Φ
uu
, which does not contain Φ
vv
nor Φ
ww
, corresponds to Townsend’s (1976) attached
inactive motions. The limit that separates active and inactive eddies is represented by
the chain-dotted lines in figure 3(c).

Citations
More filters
Journal ArticleDOI
TL;DR: In this article, a publisher's version of an article published in Journal of Fluid Mechanics © 2007 Cambridge University Press, Cambridge, UK. www.cambridge.edu.org/
Abstract: This is a publisher’s version of an article published in Journal of Fluid Mechanics © 2007 Cambridge University Press. www.cambridge.org/

1,197 citations


Cites background or methods from "Scaling of the energy spectra of tu..."

  • ...The data are from the channel flow DNS of del Álamo et al. (2004) and are included for illustrative purposes only....

    [...]

  • ...Figure 7 shows streamwise velocity fluctuations from recent channel flow simulations of del Álamo et al. (2004) at Reτ = 934....

    [...]

  • ...Figure 7 shows streamwise velocity fluctuations from recent channel flow simulations of del Álamo et al. (2004) at Reτ = 934. Plot (a) shows the u fluctuations in the log region at z = 150 (z/h = 0.164, where h is the channel half-height). Clearly visible in this plot are several occurrences of very long regions of negative u fluctuation, visible in the plot as elongated dark regions. (In separate DNS studies Tanahashi et al. 2004 show complex clusters of fine-scale vortices existing within these elongated lowspeed features.) These features are >10h in length and have characteristic length scales that scale on outer variables (as shown on the earlier correlation plots). For boundary layers, the streamwise and spanwise length scales inferred from the peak in kxΦuu and kyΦuu are of the order λx ≈ 6δ and λy ≈ 0.7δ respectively. Different length scales have been reported for channel and pipe flows (Jiménez 1998; Kim & Adrian 1999). Figure 7(b) shows the corresponding streamwise velocity fluctuations at z = 15 (z/h = 0.016). In this near-wall region, inner-scaled streaks dominate with commonly reported characteristic length scales λx ≈ 1000 and λy ≈ 100 (e.g. Jiménez & del Álamo 2004). As well as illustrating these two scales, figure 7 seems to imply some kind of interaction. If we peer through the small inner-scaled structures of plot (b), we can see a faint superimposed footprint of the larger-scale features. This characteristic is highlighted by applying a filter of size (h/2 × h/2) to both planes of data. Here, the aim of the filter is to average away the small-scale features, and for this purpose a simple Gaussian is adequate. Figure 7(c) and (d) shows just the negative streamwise velocity fluctuations as grey-scale contours for the filtered fields, and it is noted that remarkable similarities now appear between the remaining large-scale events for both wall-normal positions. One such feature is highlighted by the dot-dashed contour on figure 7(c), which is drawn at a small negative value (u = −0.2). This feature exceeds the length of the numerical domain (>25h). When the same contour is plotted on figure 7(d), with an appropriate streamwise shift, we note that it encloses almost exactly the same large-scale feature. If we refer back to the original velocity fields of plots (a) and (b) it is possible to discern this large-scale feature even through the unfiltered data. In separate DNS studies, Abe, Kawamura & Choi (2004) have also noted a footprint from the outer-layer structure onto the near-wall region, observing that these large-scale structures contribute to the shearstress fluctuations. Tsubokura (2005) concluded a similar result from LES studies of...

    [...]

  • ...Figure 7 shows streamwise velocity fluctuations from recent channel flow simulations of del Álamo et al. (2004) at Reτ = 934. Plot (a) shows the u fluctuations in the log region at z = 150 (z/h = 0.164, where h is the channel half-height). Clearly visible in this plot are several occurrences of very long regions of negative u fluctuation, visible in the plot as elongated dark regions. (In separate DNS studies Tanahashi et al. 2004 show complex clusters of fine-scale vortices existing within these elongated lowspeed features.) These features are >10h in length and have characteristic length scales that scale on outer variables (as shown on the earlier correlation plots). For boundary layers, the streamwise and spanwise length scales inferred from the peak in kxΦuu and kyΦuu are of the order λx ≈ 6δ and λy ≈ 0.7δ respectively. Different length scales have been reported for channel and pipe flows (Jiménez 1998; Kim & Adrian 1999). Figure 7(b) shows the corresponding streamwise velocity fluctuations at z = 15 (z/h = 0.016). In this near-wall region, inner-scaled streaks dominate with commonly reported characteristic length scales λx ≈ 1000 and λy ≈ 100 (e.g. Jiménez & del Álamo 2004). As well as illustrating these two scales, figure 7 seems to imply some kind of interaction. If we peer through the small inner-scaled structures of plot (b), we can see a faint superimposed footprint of the larger-scale features. This characteristic is highlighted by applying a filter of size (h/2 × h/2) to both planes of data. Here, the aim of the filter is to average away the small-scale features, and for this purpose a simple Gaussian is adequate. Figure 7(c) and (d) shows just the negative streamwise velocity fluctuations as grey-scale contours for the filtered fields, and it is noted that remarkable similarities now appear between the remaining large-scale events for both wall-normal positions. One such feature is highlighted by the dot-dashed contour on figure 7(c), which is drawn at a small negative value (u = −0.2). This feature exceeds the length of the numerical domain (>25h). When the same contour is plotted on figure 7(d), with an appropriate streamwise shift, we note that it encloses almost exactly the same large-scale feature. If we refer back to the original velocity fields of plots (a) and (b) it is possible to discern this large-scale feature even through the unfiltered data. In separate DNS studies, Abe, Kawamura & Choi (2004) have also noted a footprint from the outer-layer structure onto the near-wall region, observing that these large-scale structures contribute to the shearstress fluctuations....

    [...]

  • ...The data are from the channel flow DNS of del Álamo et al. (2004) and are included for illustrative purposes only....

    [...]

Journal ArticleDOI
TL;DR: In this article, a new numerical simulation of a turbulent channel in a large box at Reτ=2003 is described and briefly compared with simulations at lower Reynolds numbers and with experiments.
Abstract: A new numerical simulation of a turbulent channel in a large box at Reτ=2003 is described and briefly compared with simulations at lower Reynolds numbers and with experiments. Some of the fluctuation intensities, especially the streamwise velocity, do not scale well in wall units, both near and away from the wall. Spectral analysis traces the near-wall scaling failure to the interaction of the logarithmic layer with the wall. The present statistics can be downloaded from http://torroja.dmt.upm.es/ftp/channels. Further ones will be added to the site as they become available.

1,018 citations


Cites background or result from "Scaling of the energy spectra of tu..."

  • ...It is therefore comparable with previous simulations at lower Reynolds numbers by our group (Del Álamo & Jiménez 2003; Del Álamo et al. 2004). We integrate evolution equations for the wall-normal vorticity ωy and for the Laplacian of the wall-normal velocity φ = ∇2v, as in Kim, Moin & Moser (1987). The streamwise and spanwise coordinates are x and z, and the corresponding velocity components are u and w....

    [...]

  • ...It is therefore comparable with previous simulations at lower Reynolds numbers by our group (Del Álamo & Jiménez 2003; Del Álamo et al. 2004)....

    [...]

  • ...The subject has been discussed for example in Del Álamo et al. (2004)....

    [...]

  • ...The subject has been discussed for example in Del Álamo et al. (2004). However, except to present the additional data point in this figure, it will not be further discussed here. Two-dimensional spectral energy densities at the height of the near-wall kinetic energy maximum are shown in Fig. 3. Two isolines are given for each case, representing the high-intensity core of the spectrum and its outer border. They confirm the results of simulations at lower Reynolds numbers in Del Álamo & Jiménez (2003). The core isolines scale well in wall units. Because the kinetic energy is the integral of the spectrum, this part of the energy also scales in wall units. The scaling failure appears for u and w in the upper right-hand corner, where a spectral ridge extends roughly along λz = 0.15λx. The ridge gets longer as the Reynolds number increases, reaching up to λx ≈ 10h in the case of φuu. The eddies in this ridge are inactive in the sense of Townsend (1976). They are not found in the v spectrum, nor in the Reynolds stress co-spectra in figure 3(c)....

    [...]

  • ...The subject has been discussed for example in Del Álamo et al. (2004). However, except to present the additional data point in this figure, it will not be further discussed here. Two-dimensional spectral energy densities at the height of the near-wall kinetic energy maximum are shown in Fig. 3. Two isolines are given for each case, representing the high-intensity core of the spectrum and its outer border. They confirm the results of simulations at lower Reynolds numbers in Del Álamo & Jiménez (2003). The core isolines scale well in wall units....

    [...]

Journal ArticleDOI
TL;DR: In this paper, a direct numerical simulation of incompressible channel flow at a friction Reynolds number of 5186 has been performed, and the flow exhibits a number of the characteristics of high-Reynolds-number wall-bounded turbulent flows.
Abstract: A direct numerical simulation of incompressible channel flow at a friction Reynolds number ( ) of 5186 has been performed, and the flow exhibits a number of the characteristics of high-Reynolds-number wall-bounded turbulent flows. For example, a region where the mean velocity has a logarithmic variation is observed, with von Karman constant . There is also a logarithmic dependence of the variance of the spanwise velocity component, though not the streamwise component. A distinct separation of scales exists between the large outer-layer structures and small inner-layer structures. At intermediate distances from the wall, the one-dimensional spectrum of the streamwise velocity fluctuation in both the streamwise and spanwise directions exhibits dependence over a short range in wavenumber . Further, consistent with previous experimental observations, when these spectra are multiplied by (premultiplied spectra), they have a bimodal structure with local peaks located at wavenumbers on either side of the range.

910 citations

Journal ArticleDOI
TL;DR: In this article, the authors review wall-bounded turbulent flows, particularly high-Reynolds number, zero-pressure gradient boundary layers, and fully developed pipe and channel flows.
Abstract: We review wall-bounded turbulent flows, particularly high–Reynolds number, zero–pressure gradient boundary layers, and fully developed pipe and channel flows. It is apparent that the approach to an asymptotically high–Reynolds number state is slow, but at a sufficiently high Reynolds number the log law remains a fundamental part of the mean flow description. With regard to the coherent motions, very-large-scale motions or superstructures exist at all Reynolds numbers, but they become increasingly important with Reynolds number in terms of their energy content and their interaction with the smaller scales near the wall. There is accumulating evidence that certain features are flow specific, such as the constants in the log law and the behavior of the very large scales and their interaction with the large scales (consisting of vortex packets). Moreover, the refined attached-eddy hypothesis continues to provide an important theoretical framework for the structure of wall-bounded turbulent flows.

821 citations


Cites methods from "Scaling of the energy spectra of tu..."

  • ...…– 04 5e – 04 4e – 04 3e – 04 2e – 04 1e – 04 Figure 2 Streamwise u spectra at y+ ≈ 100 from direct numerical simulations (DNS) of channel flow at Reτ = 395 (Moser et al. 1999) and 950 (del Álamo et al. 2004) and turbulent boundary layer experiments at Reτ = 2,800 and 19,000 (Mathis et al. 2009a)....

    [...]

  • ...…the wall, the differences between the temporal experimental streamwise spectra converted using Taylor’s hypothesis and actual spatial spectra from the channel flow simulation by del Álamo et al. (2004) led Monty & Chong (2009) to propose a formulation for a wavelengthdependent convection velocity....

    [...]

  • ...…of the available data by Jiménez & Hoyas (2008) and Buschmann et al. (2009), together with analysis of DNS data by del Álamo & Jiménez (2003), del Álamo et al. (2004), and Hoyas & Jiménez (2006), provide support for the Townsend attached-eddy hypothesis predictions for v and w, but also…...

    [...]

Journal ArticleDOI
TL;DR: In this paper, statistics obtained from seven different direct numerical simulations (DNSs) pertaining to a canonical turbulent boundary layer (TBL) under zero pressure gradient are compiled and compared, and the resulting comparison shows surprisingly large differences not only in both basic integral quantities such as the friction coefficient or the shape factor H12, but also in their predictions of mean and fluctuation profiles far into the sublayer.
Abstract: Statistics obtained from seven different direct numerical simulations (DNSs) pertaining to a canonical turbulent boundary layer (TBL) under zero pressure gradient are compiled and compared. The considered data sets include a recent DNS of a TBL with the extended range of Reynolds numbers Reθ = 500–4300. Although all the simulations relate to the same physical flow case, the approaches differ in the applied numerical method, grid resolution and distribution, inflow generation method, boundary conditions and box dimensions. The resulting comparison shows surprisingly large differences not only in both basic integral quantities such as the friction coefficient cf or the shape factor H12, but also in their predictions of mean and fluctuation profiles far into the sublayer. It is thus shown that the numerical simulation of TBLs is, mainly due to the spatial development of the flow, very sensitive to, e.g. proper inflow condition, sufficient settling length and appropriate box dimensions. Thus, a DNS has to be considered as a numerical experiment and should be the subject of the same scrutiny as experimental data. However, if a DNS is set up with the necessary care, it can provide a faithful tool to predict even such notoriously difficult flow cases with great accuracy.

752 citations


Cites methods from "Scaling of the energy spectra of tu..."

  • ...In some figures, the well-known channel-flow data from Iwamoto, Suzuki & Kasagi (2002), Abe, Kawamura & Matsuo (2004), del Álamo & Jiménez (2003), del Álamo et al. (2004), Hoyas & Jiménez (2006), Kawamura, Abe & Matsuo (1999) and Tsukahara et al. (2005) have been utilized for comparisons....

    [...]

References
More filters
Journal ArticleDOI
TL;DR: In this article, a direct numerical simulation of a turbulent channel flow is performed, where the unsteady Navier-Stokes equations are solved numerically at a Reynolds number of 3300, based on the mean centerline velocity and channel half-width, with about 4 million grid points.
Abstract: A direct numerical simulation of a turbulent channel flow is performed. The unsteady Navier-Stokes equations are solved numerically at a Reynolds number of 3300, based on the mean centerline velocity and channel half-width, with about 4 million grid points. All essential turbulence scales are resolved on the computational grid and no subgrid model is used. A large number of turbulence statistics are computed and compared with the existing experimental data at comparable Reynolds numbers. Agreements as well as discrepancies are discussed in detail. Particular attention is given to the behavior of turbulence correlations near the wall. A number of statistical correlations which are complementary to the existing experimental data are reported for the first time.

4,788 citations


"Scaling of the energy spectra of tu..." refers background in this paper

  • ...There are few data on the advection velocities of v and w in the logarithmic layer, but Kim & Hussain (1993) find that they differ from the local mean velocity by at most a few wall units (see also del ´...

    [...]

Book
19 Dec 1975
TL;DR: In this paper, the authors present a method to find the optimal set of words for a given sentence in a sentence using the Bibliogr. Index Reference Record created on 2004-09-07, modified on 2016-08-08
Abstract: Note: Bibliogr. : p. 413-424. Index Reference Record created on 2004-09-07, modified on 2016-08-08

3,758 citations

Journal ArticleDOI
TL;DR: In this paper, numerical simulations of fully developed turbulent channel flow at three Reynolds numbers up to Reτ=590 were reported, and it was noted that the higher Reynolds number simulations exhibit fewer low Reynolds number effects than previous simulations at Reτ = 180.
Abstract: Numerical simulations of fully developed turbulent channel flow at three Reynolds numbers up to Reτ=590 are reported. It is noted that the higher Reynolds number simulations exhibit fewer low Reynolds number effects than previous simulations at Reτ=180. A comprehensive set of statistics gathered from the simulations is available on the web at http://www.tam.uiuc.edu/Faculty/Moser/channel.

2,618 citations


"Scaling of the energy spectra of tu..." refers methods in this paper

  • ...The spatial discretization uses dealiased Fourier expansions in the wall-parallel planes, and Chebychev polynomials in y .T he streamwise and spanwise coordinates and velocity components are respectively x, z and u, w. The temporal discretization is third-order semi-implicit Runge–Kutta, as in Moser, Kim & Mansour (1999) ....

    [...]

01 Jun 1953
TL;DR: In this paper, a hot-wire anemometer was used to measure the turbulent flow in a 10-inch pipe at speeds of approximately 10 and 100 feet per second, and the results include relevant mean and statistical quantities, such as Reynolds stresses, triple correlations, turbulent dissipation, and energy spectra.
Abstract: Measurements, principally with a hot-wire anemometer, were made in fully developed turbulent flow in a 10-inch pipe at speeds of approximately 10 and 100 feet per second. Emphasis was placed on turbulence and conditions near the wall. The results include relevant mean and statistical quantities, such as Reynolds stresses, triple correlations, turbulent dissipation, and energy spectra. It is shown that rates of turbulent-energy production, dissipation, and diffusion have sharp maximums near the edge of the laminar sublayer and that there exist a strong movement of kinetic energy away from this point and an equally strong movement of pressure energy toward it.

1,053 citations


"Scaling of the energy spectra of tu..." refers background in this paper

  • ...It has been known since Laufer (1954) that the k−1 law is approximately correct, but recent detailed observations by Hites (1997) and by Morrison et al. (2002) reveal important corrections which are inconsistent with the original argument....

    [...]

Journal ArticleDOI
TL;DR: Very large-scale motions in the form of long regions of streamwise velocity fluctuation are observed in the outer layer of fully developed turbulent pipe flow over a range of Reynolds numbers.
Abstract: Very large-scale motions in the form of long regions of streamwise velocity fluctuation are observed in the outer layer of fully developed turbulent pipe flow over a range of Reynolds numbers. The premultiplied, one-dimensional spectrum of the streamwise velocity measured by hot-film anemometry has a bimodal distribution whose components are associated with large-scale motion and a range of smaller scales corresponding to the main turbulent motion. The characteristic wavelength of the large-scale mode increases through the logarithmic layer, and reaches a maximum value that is approximately 12–14 times the pipe radius, one order of magnitude longer than the largest reported integral length scale, and more than four to five times longer than the length of a turbulent bulge. The wavelength decreases to approximately two pipe radii at the pipe centerline. It is conjectured that the very large-scale motions result from the coherent alignment of large-scale motions in the form of turbulent bulges or packets of...

853 citations


"Scaling of the energy spectra of tu..." refers background or methods in this paper

  • ...…long periodicities Lx and Lz, to try to account for all the energetic structures in the flow, including those in the outer region whose size is proportional to h, and that could not be captured accurately in previous simulations (Jiménez 1998; Kim & Adrian 1999; del Álamo & Jiménez 2001, 2003)....

    [...]

  • ...It has been known since Laufer (1954) that the k−1 law is approximately correct, but recent detailed observations by Hites (1997) and by Morrison et al....

    [...]

Frequently Asked Questions (13)
Q1. What have the authors contributed in "Scaling of the energy spectra of turbulent channels" ?

Jiménez, del Álamo, Zandonade and Moser this paper showed that the velocity scale of the largest structures is not the friction velocity but the mean velocity at the centreline, and used the resulting spectral model to derive a modified mixed scaling law for the fluctuations. 

This anomalous scaling results in logarithmic corrections to various ranges of the k−1 spectrum, which have been used to collapse the numerical data with experiments at higher Reynolds numbers. 

A consequence of these scalings is a logarithmic correction to the k−1x onedimensional u-spectrum in the range of the detached eddies. 

Some of the best-known similarity arguments in fluid dynamics are those regarding the overlap layer in wall turbulence, and it has lately become increasingly clear that, while the logarithmic law for the mean velocity profile may be a good approximation to the experimental data (Zagarola & Smits 1997; Österlund et al. 2000), there are serious similarity failures in the behaviour of the velocity fluctuations, including the scaling of their intensity in wall units and their predicted k−1 energy spectrum. 

Because of the range of Reτ in the numerical channels, figure 3(a) spans a factor of more than 3 in y/h, while figure 3(b) does the same for y+. 

When it is scaled with u2τ it increases with Reτ by approximately a factor of 2, but it remains roughly constant when scaled with U 2c , except for the lowest Reynolds number. 

The numerical code integrates the Navier–Stokes equations in the form of evolution problems for the wall-normal vorticity ωy and for the Laplacian of the wall-normal velocity ∇2v, as in Kim, Moin & Moser (1987). 

The other spectral range in which the scaling is reasonably clear is region C in figure 3(c), which corresponds to the ‘global’ modes described by Bullock, Cooper & Abernathy (1978) and by del Álamo & Jiménez (2003) as being correlated across the entire flow. 

This effect, combined with the variation of f , produces the shortening of the peak of the one-dimensional spectrum as y increases through the outer layer, which has been documented by various groups, probably first by Perry & Abell (1975). 

with y0 fixed in the viscous layer (y + 0 = 15 for the symbols in figure 3d), the wavelength dependence also scales with y, but the region of significant correlation is shifted to much larger λx . 

The resulting spectral model can be integrated to obtain a mixed scaling for the total intensity of the fluctuating velocity, which collapses the numerical results well with those of de Graaff & Eaton (2000) in the outer layer, and which predicts that this part of the flow will tend to scale with the mean-stream velocity at very high Reynolds numbers. 

The factor y in the denominator inside the logarithm of (3.3) is due to the displacement with wall distance of the upper bound cd of region C in figure 3(c), which progressively ‘cuts’ 

This is shown in figures 5(a) and 5(b), which display the average energy density of u in that spectral band as a function of Reynolds number.