scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Further observations on the mean velocity distribution in fully developed pipe flow

25 Feb 2004-Journal of Fluid Mechanics (Cambridge University Press)-Vol. 501, Iss: 501, pp 135-147
TL;DR: McKeon et al. as discussed by the authors used a smaller Pitot probe to reduce the uncertainties due to velocity gradient corrections, and showed that the velocity profiles in fully developed turbulent pipe flow are repeated using a smaller pitot probe, which leads to significant differences from the Zagarola & Smits conclusions.
Abstract: The measurements by Zagarola & Smits (1998) of mean velocity profiles in fully developed turbulent pipe flow are repeated using a smaller Pitot probe to reduce the uncertainties due to velocity gradient corrections. A new static pressure correction (McKeon & Smits 2002) is used in analysing all data and leads to significant differences from the Zagarola & Smits conclusions. The results confirm the presence of a power-law region near the wall and, for Reynolds numbers greater than 230×10^3 (R+ >5×10^3), a logarithmic region further out, but the limits of these regions and some of the constants differ from those reported by Zagarola & Smits. In particular, the log law is found for 600

Summary (1 min read)

1. Introduction

  • Perry, Hafez & Chong (2001) attributed the steps in the data (which were exacerbated in their analysis by a different choice of κ and B) to roughness effects combined with inappropriate Pitot tube corrections.
  • Recent experiments at Princeton, however, have demonstrated that these deviations were in fact due to inaccurate static pressure corrections (McKeon & Smits 2002) and the absence of a wall term in the Pitot correction.
  • The new static pressure correction removes the deviations observed in the ZS profiles to within the limits of experimental uncertainty .

1.2. Corrections to the measurements

  • Since an appropriate Pitot correction must surely collapse data from probes of all diameters, the pipe flow mean velocity measurements were analysed in two different ways to ensure the conclusions were not affected by the Pitot correction.
  • First, all points taken within 2d of the wall, that is, where the Pitot probe corrections disagreed, were removed from the data set.
  • The remaining data points were therefore independent of the particular Pitot tube corrections used.
  • To be consistent with ZS the Chue (1975) correction was applied to the data, although any of the other corrections, such as MacMillan (1956) , give the same results.
  • This data set (y > 2d) was used in the analysis of the log-law shown here.

1.4. Experimental considerations

  • In other respects, the experimental apparatus and techniques used for the new dataset were virtually the same as those used by ZS, and their error analysis applies almost unchanged to the measurements reported here.
  • The maximum error in velocity measurement remained 0.35%.
  • The first point was taken with the Pitot probe touching the pipe surface, the 56th point was located at the centre of the pipe, and the 57th point was on the other side of the pipe centre.
  • The symmetry of the flow across the pipe, first noted by Zagarola (1996) was confirmed in the present experiment.
  • The data were sampled using a new PC-based data acquisition system (National Instrument data acquisition boards driven by Labview software) at 500 Hz over a two minute period.

Did you find this useful? Give us your feedback

Figures (10)

Content maybe subject to copyright    Report

J. Fluid Mech. (2004), vol. 501, pp. 135–147.
c
2004 Cambridge University Press
DOI: 10.1017/S0022112003007304 Printed in the United Kingdom
135
Further observations on the mean velocity
distribution in fully developed pipe flow
By B. J. M c KEON, J. LI, W. JIANG,
J. F. MORRISON
AND A. J. SMITS
Department of Mechanical and Aerospace Engineering, Princeton University,
Princeton, NJ 08544-0710, USA
(Received 7 October 2002 and in revised form 19 September 2003)
The measurements by Zagarola & Smits (1998) of mean velocity profiles in fully
developed turbulent pipe flow are repeated using a smaller Pitot probe to reduce the
uncertainties due to velocity gradient corrections. A new static pressure correction
(McKeon & Smits 2002) is used in analysing all data and leads to significant
differences from the Zagarola & Smits conclusions. The results confirm the presence
of a power-law region near the wall and, for Reynolds numbers greater than 230 ×10
3
(R
+
> 5 ×10
3
), a logarithmic region further out, but the limits of these regions and
some of the constants differ from those reported by Zagarola & Smits. In particular,
the log law is found for 600 <y
+
< 0.12R
+
(instead of 600 <y
+
< 0.07R
+
), and the
von K
´
arm
´
an constant κ, the additive constant B for the log law using inner flow
scaling, and the additive constant B
for the log law using outer scaling are found
to be 0.421 ± 0.002, 5.60 ± 0.08 and 1.20 ± 0.10, respectively, with 95% confidence
level (compared with 0.436 ±0.002, 6.15 ±0.08, and 1.51 ±0.03 found by Zagarola &
Smits). The data also confirm that the pipe flow data for Re
D
6 13.6 ×10
6
(as a
minimum) are not affected by surface roughness.
1. Introduction
1.1. Background and previous work
Zagarola & Smits (1998) (referred to hereinafter as ZS) presented measurements in
fully developed pipe flow for Reynolds numbers in the range 31 ×10
3
to 35 ×10
6
to study the scaling of the mean velocity profile. They found two overlap regions: a
power law for 60 <y
+
< 500, and, for Reynolds numbers greater than 400 ×10
3
,a
log law in the region 600 <y
+
< 0.07R
+
. Here, y
+
= yu
τ
, y is the distance from the
wall, u
τ
=
τ
w
, ν is the kinematic viscosity, τ
w
is the shear stress at the wall, and
ρ is the fluid density. Also, R
+
= Ru
τ
where R is the radius of the pipe (=D/2).
These findings were supported by a new scaling argument based on dimensional
analysis. For high Reynolds numbers, the scaling for the mean velocity profile in fully
developed turbulent pipe flow may be expressed in terms of an inner-layer scaling
given by
U = f
(y,u
i
,R) (1.1)
Permanent address: School of the Built Environment, Victoria University of Technology, PO
Box 14428, MCMC, Melbourne, Australia.
Permanent address: CARDC, PO Box 211 Mianyang, Sichuan 621000, P. R. China.
Permanent address: Department of Aeronautics, Imperial College, London SW7 2BY, UK.

136 B. J. McKeon, J. Li, W. Jiang, J. F. Morrison and A. J. Smits
and an outer-layer scaling given by
U
c
U = g
(
y,u
0
,R
)
(1.2)
where U is the mean velocity, U
c
is the centreline velocity, and f
and g
denote
a functional dependence. The inner velocity scale u
i
is always taken to be u
τ
,but
choosing the outer velocity scale u
0
is more controversial, as will be seen below.
Non-dimensionalizing equations (1.1) and (1.2) gives, respectively,
U
+
= f (y
+
,R
+
), (1.3)
U
c
U
u
0
= g(η, R
+
), (1.4)
where U
+
= U/u
τ
and η = y/R. ZS showed that the velocity scaling in the overlap
region (where y
+
1andη 1) can be of two types: complete similarity where u
0
/u
τ
is independent of Reynolds number, and incomplete similarity where u
0
/u
τ
continues
to depend on Reynolds number. In the case of complete similarity, matching of the
velocity gradients in the overlap region leads to a logarithmic velocity profile, which
can be expressed in inner-layer variables as
U
+
=
1
κ
ln y
+
+ B (1.5)
and in outer-layer variables as
U
c
U
u
τ
=
1
κ
ln η + B
(1.6)
where κ (usually called the von K
´
arm
´
an constant), B and B
are constants ind-
ependent of Reynolds number.
Alternatively, in the case of incomplete similarity, matching the velocities as well
as the velocity gradients in the overlap region yields a power-law dependence which
in terms of inner-layer variables may be written as
U
+
= Cy
+γ
(1.7)
where the coefficient C and the exponent γ are independent of Reynolds number.
The study by ZS showed that the power-law constants were given by C =8.70 and
γ =0.137, and the log-law constants were given by κ =0.436 ±0.002, B =6.15 ±0.08,
and B
=1.51 ± 0.03. These values of the log-law constants are different from the
commonly accepted values of about 0.41, 5.0 and 0.8, respectively. for the so-called
‘standard log law’ discussed below.
ZS also proposed a new velocity scale for the outer region, u
0
= U
c
¯
U.They
argued that u
0
= U
c
¯
U was a more representative outer velocity scale than the
friction velocity, which describes the inner flow and is impressed upon the outer
region. Since the ratio (U
c
¯
U)/u
τ
becomes constant at high Reynolds number, there
is no change to the overlap analysis leading to a log law at these Reynolds numbers. At
low Reynolds numbers (high enough that an overlap region still exists), the variation
of (U
c
¯
U)/u
τ
with Reynolds number means that matching velocity gradients does
not lead to a Reynolds-number-independent condition. However matching velocities
and velocity gradients leads to the power law of equation (1.7).
ZS obtained their measurements using a round Pitot probe of 0.9 mm OD. A
number of corrections were made to the data, including corrections to the Pitot tube
pressure for the effects of turbulence, viscosity, and velocity gradient, and corrections
to the static pressure for viscous effects. Subsequent examination of the ZS data

Mean velocity in fully developed pipe flow 137
Figure 1. Velocity profiles for the 0.9 mm (ZS) data at Re
D
=3×10
6
(solid lines) and 10 ×10
6
(dashed lines): (a) results using Shaw’s (1960) correction as reported by ZS; (b) results using
correction according to McKeon & Smits (2002).
revealed slight deviations between the velocity profiles in inner scaling (figure 1a).
The deviations (also called ‘steps’) appear as departures from the logarithmic region
for Re
D
> 3 ×10
6
and increase with Reynolds number.
In ZS, the comparison of the data with the log law and the power law was done
by fitting the scaling laws to the mean velocity profile. As pointed out by W. George
(private communication), the mean velocity itself is not a very sensitive quantity in
distinguishing a log law from a power law because the differences are small. This is
especially true at low Reynolds number. Fractional differences were therefore used, as
suggested by Zagarola, Perry & Smits (1997). The fractional difference E = U
+
/U
+
between a fitted curve and the measured data is defined as
E =
U
+
U
+
=1
U
+
fit
U
+
measured
. (1.8)
Zagarola et al. suggested that a good fit to the experimental data should result in
random relative errors that are less than the mean experimental error.
Traditionally, dimensional analysis using u
0
= u
τ
has yielded a log law (Millikan
1938). The so-called ‘standard log law’ arose from the conclusions of several studies
in low-Reynolds-number boundary layers, including that by Bradshaw (1976), that
the overlap region was best represented by a log law with κ 0.41 and B 5.0
(Schlichting 1979).
¨
Osterlund et al. (2000) in a more recent study of boundary layers found a log law
with κ =0.38 and B =4.1 in the range 200
+
<y/δ<0.15 for Re
θ
> 6000. Although
this log law appears over a rather short region in y
+
(200 to about 1000), we should
note that this includes virtually all boundary layer studies, where few studies have
a large enough Reynolds number range to investigate the overlap region with an
upper extent where y
+
> 1000. This includes the work by
¨
Osterlund et al. where the
maximum value of Re
θ
was 27 000.
In a medium-Reynolds-number channel flow (Re
h
< 1.2 ×10
5
based on half-channel
height and mean velocity, 1000 <Re
τ
< 5000), Zanoun et al. (2002) found a similar
value of the von K
´
arm
´
an constant, κ =0.379, with B =4.05, for Re
τ
> 2000 within
the limits y
+
> 100 and y/h < 0.15. However, they also suggested that a log law may

138 B. J. McKeon, J. Li, W. Jiang, J. F. Morrison and A. J. Smits
be valid for limits reaching y
+
=50 and y/h =0.9, which presumably yields different
log-law constants.
1.2. Corrections to the measurements
Perry, Hafez & Chong (2001) attributed the steps in the data (which were exacerbated
in their analysis by a different choice of κ and B) to roughness effects combined
with inappropriate Pitot tube corrections. Recent experiments at Princeton, however,
have demonstrated that these deviations were in fact due to inaccurate static pressure
corrections (McKeon & Smits 2002) and the absence of a wall term in the Pitot
correction. ZS used the static pressure correction proposed by Shaw (1960), who
suggested that the error, p , depends only on d
+
t
and reaches an asymptotic value
of approximately 3.0τ
w
for d
+
t
> 1000, where d
+
t
= u
τ
d
t
and d
t
is the hole diameter.
McKeon & Smits (2002) have shown instead that the error continues to grow with
increasing Reynolds number as long as the ratio d
t
/D is small, and reaches a value
of greater than 7τ
w
at the highest Reynolds numbers obtained in the pipe, where
d
+
t
6500. Although the new correction corresponds to a change of less than 1% in
the maximum dynamic pressure for all cases presented here, the new static pressure
correction removes the deviations observed in the ZS profiles to within the limits of
experimental uncertainty (see figure 1b).
To understand the effect of Pitot probe corrections on the similarity scaling and
the value of the log-law constants, a further study was undertaken by McKeon
et al. (2003) using four different sized Pitot tubes, measuring 0.3 mm to 1.83 mm OD.
A new correction scheme was suggested that collapses the data markedly better than
other wall corrections methods, including that suggested by MacMillan (1956), for
y<2d,whered is the outer diameter of the Pitot tube, and that agrees well with
MacMillan’s and Chue’s (1975) methods over the entire Reynolds number range for
y > 2d.
1.3. Current study
ZS measured 28 velocity profiles over a range of Reynolds numbers from 31 ×10
3
to 36 ×10
6
using a 0.9 mm Pitot probe. We use these data, in addition to new
measurements encompassing 21 profiles obtained using a Pitot probe of 0.3 mm
replicating the ZS Reynolds numbers over the range from 74 ×10
3
to 35 ×10
6
.Both
sets of data included measurements of the pressure drop to determine the friction
factor. The Reynolds numbers and friction factors for the mean velocity measurements
analysed in this work are shown in table 1, along with the symbols that will be used
in the figures that follow.
Since an appropriate Pitot correction must surely collapse data from probes of all
diameters, the pipe flow mean velocity measurements were analysed in two different
ways to ensure the conclusions were not affected by the Pitot correction. First,
all points taken within 2d of the wall, that is, where the Pitot probe corrections
disagreed, were removed from the data set. The remaining data points were therefore
independent of the particular Pitot tube corrections used. To be consistent with ZS the
Chue (1975) correction was applied to the data, although any of the other corrections,
such as MacMillan (1956), give the same results. This data set (y > 2d) was used in
the analysis of the log-law shown here. Second, to investigate the near-wall behaviour,
all the data points for y > d were used. In this case, the data were corrected using
the method suggested by McKeon et al. (2003), which gives good agreement between
data from probes of different diameters for y > d. It should be noted that analysis of

Mean velocity in fully developed pipe flow 139
Symbol Re
D
λ Symbol Re
D
λ
M 75 × 10
3
0.0193 H 2.3 × 10
6
0.0103
150 × 10
3
0.0167 F 3.1 ×10
6
0.0099
N 230 × 10
3
0.0153 C 4.4 × 10
6
0.0094
310 × 10
3
0.0147 | 6.1 ×10
6
0.0090
410 × 10
3
0.0138 J 7.7 × 10
6
0.0086
540 × 10
3
0.0132 P 10.2 ×10
6
0.0082
750 × 10
3
0.0125 B 13.6 × 10
6
0.0080
1.0 × 10
6
0.0118 18.2 ×10
6
0.0077
O 1.3 × 10
6
0.0113 I 35.3 × 10
6
0.0071
X1.7 × 10
6
0.0108
Tab l e 1. Nomenclature and friction factor data (0.9 mm Pitot probe).
this second data set confirmed all the conclusions derived using the first data set (the
set that was independent of the Pitot tube correction used).
1.4. Experimental considerations
In other respects, the experimental apparatus and techniques used for the new
dataset were virtually the same as those used by ZS, and their error analysis applies
almost unchanged to the measurements reported here. The maximum error in velocity
measurement remained 0.35%. However, to improve the accuracy of the wall distance
measurement, a linear encoder with an accuracy of 5 µm per count was used to
determine the wall distance (ZS reported an accuracy of 25 µm). In addition, the
starting position for the Pitot probe was determined by detecting the electrical contact
between the Pitot probe and the pipe surface to an accuracy of 5 µm, compared to
50 µm for ZS. The accumulated position error of running the present traverse system
forward and backward once over a distance of 71 mm was generally less than 30 µm,
compared with 50 µm accuracy for ZS. The position accuracy was estimated to be
±1.7% for measuring points close to the pipe surface and ±0.05% for data points
taken near the centre of the pipe. Across the pipe radius, 57 data points were taken
with logarithmically uniform spacing. The first point was taken with the Pitot probe
touching the pipe surface, the 56th point was located at the centre of the pipe, and the
57th point was on the other side of the pipe centre. The symmetry of the flow across
the pipe, first noted by Zagarola (1996) was confirmed in the present experiment.
The data were sampled using a new PC-based data acquisition system (National
Instrument data acquisition boards driven by Labview software) at 500 Hz over a two
minute period.
2. Mean velocity results
Figure 2 shows sample mean velocity profiles from the present work (i.e. under the
new corrections) for the ZS data.
Both datasets were processed in three steps. First, the value of κ was determined
from the friction factor data by fitting
1
λ
= C
1
log(Re
λ)+C
2
(2.1)

Citations
More filters
Book ChapterDOI
01 Jan 1997
TL;DR: The boundary layer equations for plane, incompressible, and steady flow are described in this paper, where the boundary layer equation for plane incompressibility is defined in terms of boundary layers.
Abstract: The boundary layer equations for plane, incompressible, and steady flow are $$\matrix{ {u{{\partial u} \over {\partial x}} + v{{\partial u} \over {\partial y}} = - {1 \over \varrho }{{\partial p} \over {\partial x}} + v{{{\partial ^2}u} \over {\partial {y^2}}},} \cr {0 = {{\partial p} \over {\partial y}},} \cr {{{\partial u} \over {\partial x}} + {{\partial v} \over {\partial y}} = 0.} \cr }$$

2,598 citations

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 "Further observations on the mean ve..."

  • ...The best-known proposal involves wall-parallel large-scale modes that reach the wall without being constrained by impermeability (Townsend 1976). It was suggested by Del Álamo & Jiménez (2003), on the basis of lower-Reynolds-number simulations, that the important outer modes were global ones spanning the whole channel and scaling with the boundary layer thickness....

    [...]

  • ...Dots with error bars are pipes from McKeon et al. (2004), with Reτ > 2000....

    [...]

  • ...This parameter is nowhere constant, but that is also the case for the high-Reynolds-number pipe data by McKeon et al. (2004) included for comparison, and for other experimental data not included in the figure....

    [...]

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 background from "Further observations on the mean ve..."

  • ...The Princeton Superpipe measurements (Zagarola & Smits 1998, McKeon et al. 2004) came under particular scrutiny, but a large number of new experiments in fully developed channel flows (Liu et al. 2001; Zanoun et al. 2003, 2009; Monty et al. 2007) and turbulent boundary layers (George & Castillo…...

    [...]

  • ...Third, there is now considerable evidence that the constants in the log law depend on the flow, so that the von Kármán constant for pipe flow, where McKeon et al. (2004) found κ = 0.421, is different from boundary-layer and channel flows, where Chauhan et al. (2007), Zanoun et al. (2003), and…...

    [...]

Journal ArticleDOI
TL;DR: In this paper, the authors distill the salient advances of recent origin, particularly those that challenge textbook orthodoxy, and highlight some of the outstanding questions, such as the extent of the logarithmic overlap layer, the universality or otherwise of the principal model parameters, and the scaling of mean flow and Reynolds stresses.
Abstract: Wall-bounded turbulent flows at high Reynolds numbers have become an increasingly active area of research in recent years. Many challenges remain in theory, scaling, physical understanding, experimental techniques, and numerical simulations. In this paper we distill the salient advances of recent origin, particularly those that challenge textbook orthodoxy. Some of the outstanding questions, such as the extent of the logarithmic overlap layer, the universality or otherwise of the principal model parameters such as the von Karman “constant,” the parametrization of roughness effects, and the scaling of mean flow and Reynolds stresses, are highlighted. Research avenues that may provide answers to these questions, notably the improvement of measuring techniques and the construction of new facilities, are identified. We also highlight aspects where differences of opinion persist, with the expectation that this discussion might mark the beginning of their resolution.

716 citations

Journal ArticleDOI
TL;DR: In this paper, the authors analyse recent experimental data in the Reynolds number range of nominally 2 × 104 < Reτ < 6 × 105 for boundary layers, pipe flow and the atmospheric surface layer, and show that the data support the existence of a universal logarithmic region.
Abstract: Considerable discussion over the past few years has been devoted to the question of whether the logarithmic region in wall turbulence is indeed universal. Here, we analyse recent experimental data in the Reynolds number range of nominally 2 × 104 < Reτ < 6 × 105 for boundary layers, pipe flow and the atmospheric surface layer, and show that, within experimental uncertainty, the data support the existence of a universal logarithmic region. The results support the theory of Townsend (The Structure of Turbulent Shear Flow, Vol. 2, 1976) where, in the interior part of the inertial region, both the mean velocities and streamwise turbulence intensities follow logarithmic functions of distance from the wall. © 2013 Cambridge University Press.

618 citations


Cites background from "Further observations on the mean ve..."

  • ...The value of κ = 0.421 reported by McKeon et al. (2004) in the Superpipe probably exceeds these bounds, and the possible reasons for differences between Pitot tube and hot-wire mean-velocity measurements are the subject of ongoing study....

    [...]

  • ...Very high-Reynolds-number studies in the Princeton Superpipe by McKeon et al. (2004) using Pitot tubes have proposed that κ = 0.421....

    [...]

References
More filters
Book
01 Jan 1955
TL;DR: The flow laws of the actual flows at high Reynolds numbers differ considerably from those of the laminar flows treated in the preceding part, denoted as turbulence as discussed by the authors, and the actual flow is very different from that of the Poiseuille flow.
Abstract: The flow laws of the actual flows at high Reynolds numbers differ considerably from those of the laminar flows treated in the preceding part. These actual flows show a special characteristic, denoted as turbulence. The character of a turbulent flow is most easily understood the case of the pipe flow. Consider the flow through a straight pipe of circular cross section and with a smooth wall. For laminar flow each fluid particle moves with uniform velocity along a rectilinear path. Because of viscosity, the velocity of the particles near the wall is smaller than that of the particles at the center. i% order to maintain the motion, a pressure decrease is required which, for laminar flow, is proportional to the first power of the mean flow velocity. Actually, however, one ob~erves that, for larger Reynolds numbers, the pressure drop increases almost with the square of the velocity and is very much larger then that given by the Hagen Poiseuille law. One may conclude that the actual flow is very different from that of the Poiseuille flow.

17,321 citations

Book ChapterDOI
01 Jan 1997
TL;DR: The boundary layer equations for plane, incompressible, and steady flow are described in this paper, where the boundary layer equation for plane incompressibility is defined in terms of boundary layers.
Abstract: The boundary layer equations for plane, incompressible, and steady flow are $$\matrix{ {u{{\partial u} \over {\partial x}} + v{{\partial u} \over {\partial y}} = - {1 \over \varrho }{{\partial p} \over {\partial x}} + v{{{\partial ^2}u} \over {\partial {y^2}}},} \cr {0 = {{\partial p} \over {\partial y}},} \cr {{{\partial u} \over {\partial x}} + {{\partial v} \over {\partial y}} = 0.} \cr }$$

2,598 citations


"Further observations on the mean ve..." refers background in this paper

  • ...The so-called ‘standard log law’ arose from the conclusions of several studies in low-Reynolds-number boundary layers, including that by Bradshaw (1976), that the overlap region was best represented by a log law with κ ≈ 0.41 and B ≈ 5.0 (Schlichting 1979)....

    [...]

Journal ArticleDOI
TL;DR: In this article, a new friction factor relation is proposed which is within ± 1.2% of the data for Reynolds numbers between 10×103 and 35×106, and includes a term to account for the near-wall velocity profile.
Abstract: Measurements of the mean velocity profile and pressure drop were performed in a fully developed, smooth pipe flow for Reynolds numbers from 31×103 to 35×106. Analysis of the mean velocity profiles indicates two overlap regions: a power law for 60 9×103). Von Karman's constant was shown to be 0.436 which is consistent with the friction factor data and the mean velocity profiles for 600 5%) than those predicted by Prandtl's relation. A new friction factor relation is proposed which is within ±1.2% of the data for Reynolds numbers between 10×103 and 35×106, and includes a term to account for the near-wall velocity profile.

794 citations

Journal ArticleDOI
TL;DR: In this article, two independent experimental investigations of the behavior of turbulent boundary layers with increasing Reynolds number were recently completed, and the results are summarized here, utilizing the profiles of the mean velocity, for Reynolds numbers based on the momentum thickness ranging from 2500 to 27,000.
Abstract: Two independent experimental investigations of the behavior of turbulent boundary layers with increasing Reynolds number were recently completed. The experiments were performed in two facilities, the Minimum Turbulence Level (MTL) wind tunnel at Royal Institute of Technology (KTH) and the National Diagnostic Facility (NDF) wind tunnel at Illinois Institute of Technology (IIT). Both experiments utilized oil-film interferometry to obtain an independent measure of the wall-shear stress. A collaborative study by the principals of the two experiments, aimed at understanding the characteristics of the overlap region between the inner and outer parts of the boundary layer, has just been completed. The results are summarized here, utilizing the profiles of the mean velocity, for Reynolds numbers based on the momentum thickness ranging from 2500 to 27 000. Contrary to the conclusions of some earlier publications, careful analysis of the data reveals no significant Reynolds number dependence for the parameters desc...

332 citations


"Further observations on the mean ve..." refers background or methods in this paper

  • ...Österlund et al. (2000) in a more recent study of boundary layers found a log law with κ = 0.38 and B =4.1 in the range 200/δ+ < y/δ < 0.15 for Reθ > 6000....

    [...]

  • ...It was found that log-laws with the constants determined by Österlund et al. (2000) (κ = 0.38 and B = 4.1) and Zanoun et al. (2002) (κ =0.379 and B = 4.05) fit our pipe flow measurements quite well, but only for the small region where 350 <y+ < 950, compared with their range of 200/δ+ < y/δ < 0.15....

    [...]

Frequently Asked Questions (14)
Q1. What are the contributions in "Further observations on the mean velocity distribution in fully developed pipe flow" ?

The results confirm the presence of a power-law region near the wall and, for Reynolds numbers greater than 230 × 10 ( R > 5 × 10 ), a logarithmic region further out, but the limits of these regions and some of the constants differ from those reported by Zagarola & Smits. 

to improve the accuracy of the wall distance measurement, a linear encoder with an accuracy of 5 µm per count was used to determine the wall distance (ZS reported an accuracy of 25 µm). 

Current studyZS measured 28 velocity profiles over a range of Reynolds numbers from 31 × 103 to 36 × 106 using a 0.9 mm Pitot probe. 

For high Reynolds numbers, the scaling for the mean velocity profile in fully developed turbulent pipe flow may be expressed in terms of an inner-layer scaling given byU = f ′(y, ui, ν, R) (1.1)† 

Ū is used as the velocity scale, the additive constant becomes B∗/ξ (where ξ = (UCL − Ū )/uτ ≈ 4.28, see below) which has a value of 0.280 ± 0.02, and the slope becomes 1/κξ which has a value of 1/(1.798 ± 0.01). 

The accumulated position error of running the present traverse system forward and backward once over a distance of 71 mm was generally less than 30 µm, compared with 50 µm accuracy for ZS. 

Non-dimensionalizing equations (1.1) and (1.2) gives, respectively,U+ = f (y+, R+), (1.3)Uc − U u0 = g(η, R+), (1.4)where U+ = U/uτ and η = y/R. ZS showed that the velocity scaling in the overlap region (where y+ 1 and η 1) can be of two types: complete similarity where u0/uτ is independent of Reynolds number, and incomplete similarity where u0/uτ continues to depend on Reynolds number. 

B∗ has a constant value of 1.20 ± 0.1 (again the majority of the uncertainty lies in the value of κ) for 230 × 103 ReD 13.6 × 106. 

The first point was taken with the Pitot probe touching the pipe surface, the 56th point was located at the centre of the pipe, and the 57th point was on the other side of the pipe centre. 

Since an appropriate Pitot correction must surely collapse data from probes of all diameters, the pipe flow mean velocity measurements were analysed in two different ways to ensure the conclusions were not affected by the Pitot correction. 

The position accuracy was estimated to be ±1.7% for measuring points close to the pipe surface and ±0.05% for data points taken near the centre of the pipe. 

Österlund et al. (2000) in a more recent study of boundary layers found a log law with κ = 0.38 and B =4.1 in the range 200/δ+ < y/δ < 0.15 for Reθ > 6000. 

Although this log law appears over a rather short region in y+ (200 to about 1000), the authors should note that this includes virtually all boundary layer studies, where few studies have a large enough Reynolds number range to investigate the overlap region with an upper extent where y+ 1000. 

The authors will use the correction suggested by McKeon et al. (2003), on the basis that it produces good agreement among different diameter Pitot data sets down to y ≈ d , considerably better than other methods such as MacMillan (1956) and Chue (1975).