Sie sind auf Seite 1von 14

Meas. Sci. Technol. 7 (1996) 13961409.

Printed in the UK

New developments and applications


of skin-friction measuring techniques
H H Fernholz, G Janke, M Schober, P M Wagner and D Warnack

Berlin,
Hermann-Fottinger-Institut
fur
Technische Universitat
Stromungsmechanik,
Strae des 17. Juni 135, 10623 Berlin, Germany
Received 25 March 1996, accepted for publication 12 June 1996
Abstract. This survey covers recent developments and applications of four
skin-friction measurement techniques (oil-film interferometry, wall hot wire, surface
fence and wall pulsed wire). Comparisons of the techniques with each other and
with other methods are presented. Applications in attached and separated fully
turbulent boundary layers and in highly accelerated laminar-like flows will be shown
to demonstrate the application range and the limits of the various techniques.

1. Introduction
One of the most important results of boundary layer theory
has been the determination of the wall shear stress w , a
quantity which largely determines the energy necessary for
moving the flow of liquids and gases over solid walls.
Knowledge of the wall shear stress is very important
for many technical applications and for the understanding
of all wall-bounded shear flows. Therefore one would
like to know the magnitude (mean and fluctuating value)
and the direction of the skin-friction vector w and its
distribution over a surface. The flow may be compressible
or incompressible, laminar or turbulent and can even
reverse its direction in the vicinity of the wall in an adverse
pressure gradient.
The classical survey on skin friction measurements,
though primarily related to the performance of aircraft, was
written by Winter (1977) and there have been two more
recent review papers by Hanratty and Campbell (1983) and
by Haritonidis (1989). This paper differs from the surveys
mentioned above in that it concentrates on measuring
techniques which have been applied in our laboratory for
some years and on their application to standard and complex
flows. A few other methods will be addressed briefly if they
are used for comparison or promise potential for the future.
The discussion will be confined to incompressible flow
of gases along an aerodynamically smooth wall without heat
transfer. Special emphasis is put on a comparison between
the different techniques and on their range of validity.
For a two-dimensional incompressible laminar boundary layer with zero pressure gradient, Blasius (1907) calculated the skin friction, expressed in dimensionless form as
the skin friction coefficient cf ,
cf =

2w
= 0.664(Rex )1/2
u2

(1)

where w is the magnitude of the skin friction, and u


the density and the velocity at the edge of the boundary
c 1996 IOP Publishing Ltd
0957-0233/96/101396+14$19.50

layer, respectively, and Rex the Reynolds number defined


by u , (the kinematic viscosity) and the length x in the
streamwise direction. There is no exact solution for the
equivalent case of a turbulent boundary layer and the semiempirical relationships one finds in the literature are all
based on measurements. For more complex flows even such
relationships do not exist and measurements are a must.
Measuring techniques for the wall shear stress may be
divided into the small group of direct methods (floating
element method and oil film interferometry) and the larger
group of indirect methods which must be calibrated. The
floating element balance is probably the oldest device to
measure the wall shear stress with the implication of a
large area and a very sensitive technique to determine small
forces. Kempf (1929) used such a balance for subsonic
flow and, since the area of the balance could be reduced
due to the high skin friction, the floating element balance
has had its main application in compressible boundary
layers. An excellent survey on the design and application of
these balances was provided by Winter (1977) and a rather
comprehensive list of test cases in compressible boundary
layers was presented by Fernholz and Finley (1977, 1981).
For subsonic flows there are two more recent
developments of floating element balances.
(i) The Bechert balances with relatively large surface
elements with sizes 750 mm 600 mm and 400 mm
500 mm (Bechert et al 1985, 1992). They were designed
mainly to compare the skin friction of aerodynamically
smooth and riblet (small-ribbed) surfaces.
(ii) The balances developed by Dickinson (1965)
are much smaller and can be adapted both to subsonic
and to supersonic boundary layers for local skin-friction
measurements. The latter floating element balances have
the advantage that they were (and possibly still are)
commercially available and their application was described
by Nyguyen et al (1984), for example. These balances can
be used in flows with zero pressure gradient only, since they

Skin-friction measuring techniques

cannot compensate for the pressure difference which would


occur in an adverse pressure gradient flow, for example.
This problem was solved by Thomann and his co-workers
(Frei and Thomann 1980, Hirt et al 1986), who designed
and built a ring-type floating element balance with a small
extension in the streamwise direction and a liquid seal for
the gaps of the floating element.
All floating element balances mentioned so far have
the disadvantage that they give a value of the skin friction
which is integrated over a larger or smaller area and
necessarily fail to provide exact pointwise measurements.
This problem may be overcome by the use of miniature
surfaces such as a silicon micromachined floating-element
shear-stress sensor. Such devices have been developed
at MIT (Schmidt et al 1988, Padmanabhan et al 1995),
Caltech (Jiang et al 1995) and Case Western University
(Jiang et al 1995), for example, but published results do
not describe more than results obtained in a first laboratory
test phase.
For turbulent boundary layers the Preston tube is
the most commonly used instrument for skin-friction
measurements. The validity of this measuring technique is
dependent on the validity of the logarithmic law of the wall.
The dimensionless diameter d + = du / must lie in the
logarithmic region if the calibration curves of Patel (1965)
or Head and Vasanta Ram (1971) are used. More recently,
user-friendly calibration curves were suggested by Zurfluh
(1984) and Bechert (1995). The accuracy of the Preston
tube method is about 3% and slightly less for adverse
pressure gradients Patel (1965) and Hirt and Thomann
(1986). The state of the art of Preston tube applications
in compressible boundary layers was discussed by Finley
and Gaudet (1995). This update of the floating-element
balance and the Preston tube appeared to be necessary since
both techniques were used for comparisons with the four
techniques discussed in section 2 of this paper.
2. The measuring techniques
2.1. General remarks
It seems appropriate to begin this brief description of the
four techniques to measure skin friction by presenting some
of their properties and their range of application in tabular
form. Table 1 shows that they cover a wide spectrum
of wall-bounded shear flows although no single technique
possesses all desired properties. Figure 1 illustrates the
techniques. None of the four techniques is dependent on
the logarithmic law of the wall. The oil-film interferometry
is the only direct method, needs therefore no calibration and
has the potential to determine the whole field of wall shear
stress.
For the application of the other three techniques it is
important to heed that the fence height and the height
of the sensor wires of the probes should lie within the
viscous sublayer and that a probe must be calibrated
against reference measurements at the same position (xo , zo )
because of possible variations of skin friction in the
spanwise direction. For the calibration of the Preston tubes
we used an improved version of Patels calibration curve

similar to the one presented by Head and Vasanta Ram


(1971) which bridged the gaps between the three branches.
2.2. Oil-film interferometry
The feasibility of measuring the skin friction from the
movement of interference fringes of a thin oil film was
first realized by Tanner and Blows (1976). Oil-film
interferometry is the only direct method for skin-friction
measurement apart from floating-element balances. It has
a high spatial resolution and is capable of measuring reverse
flow. It does not require any assumptions to be made
concerning the flow field. It is easy to apply and needs
but little instrumentation.
A brief description of some features of the method
for two-dimensional flow will be given here. The full
theory and details of the application of this technique have
been described by Janke (1993). For another version of
this method, the dual-laser beam skin-friction measuring
technique, the reader is referred to Monson (1983) and to
Kim (1989), who applied this technique to supersonic flow.
A silicon-oil film is applied to a smooth solid wall.
This film is spread out by the flow into an extremely thin
layer whose thickness is typically of the order of several
wavelengths of visible light. The film thickness can be
visualized by Fizeau fringes, which originate from the
interference of light reflected from the top and from the
bottom of the oil film. Under monochromatic light the
height of the film at the kth black fringe is given by
hk = h0 + k1h
with
1h =

2[n2

k = 0, 1, 2, . . .

.
sin2 ()]1/2

(2)

(3)

Here, h0 is the height of the zeroth black fringe at the


film edge (k = 0), 1h the difference in height between
two consecutive fringes, n the refractive index of the oil,
the viewing angle of the observer and the wavelength of
the light. Equations (2) and (3) show that, for a constant
viewing angle, the fringes are contour lines of the oil-film
surface.
Inexpensive equipment (figure 2) is sufficient to obtain
bright fringes. A glass plate blackened on its back side
provides a suitable material for a flat wall, then h0 =
1h/2. For smooth curved walls a piece of video tape
glued on to the surface serves the same purpose as the
glass plate but the ratio h0 /1h will then differ from 0.5
and must be determined for a specific wall material. A
sodium lamp ( = 0.589 m), commercially available at
a marginal price, yields better fringes than does a laser.
The motion of the interference fringes (along a straight
skin-friction line) is recorded in an xt diagram. An
example is shown in figure 3 for the separated flow behind
a backwards-facing step. The diagram was recorded along
the x coordinate by repetitively reading only one line of
a properly adjusted video image into an image processor.
The resulting image consisted of 512 by 512 pixels.
The motion of the oil film will now be analysed. We
assume that the oil layer is so thin that it does not change
1397

H H Fernholz et al
Table 1. Comparison of the four measuring techniques.
Surface fence

Wall hot wire

Wall pulsed wire

Oil-film interferometry

Measured quantity
Calibration necessary
Mean value
Temporal resolution
Cross correlation
Spatial resolution
1x , 1z (mm)
Direction of w
APG

Pressure difference
Yes
Yes
Unclear
No

Heat transfer
Yes
Yes
> 10 kHz
Yes

Time of flight
Yes
Yes
' 20 Hz
No

Movement of fringes
No
Yes
No
No

< 1, 3
Yes
Yes

1.5, 0.5
Yes
Yes

< 1, 1
Possible
Yes

Reverse Flow
FPG
Transitional flow
Laminar flow
3D flow
Accuracy (estimated)

Yes
Yes (but restricted)
Unclear
Yes
Yes
4%

< 0.5, 0.5


Yes
Yes (but no
instantaneous
reverse flow)
No
Yes
Yes
Yes
Yes
4%

Yes
Probably yes
Probably yes
Probably yes
Yes
4%

Yes
Yes
Yes
Yes
Yes
< 4%

Surface fence

Wall hot wire

Wall pulsed wire

Oil-film interferometry

Figure 1. Schematic diagrams of the methods.

the flow on top of it but is only driven by the skinfriction distribution of the flow. In this case the temporal
development of the film height in a two-dimensional flow
is described by
1 ( h2 )
h
=
t
2 x

(4)

where is the dynamic viscosity of the oil. For thicker


films higher order terms in h have to be added. These
terms introduce effects of pressure gradient, gravity and
surface tension Janke (1993). The effects can be taken into
account, but for most practical situations h can be made
small enough that they may be neglected.
It can be shown Janke (1993) that, for a sufficiently
viscous oil, the movement of the film surface will be
The
determined by the mean skin friction only.
fluctuating part of , namely the turbulence structure of
the outer flow, will not affect the movement of the film.
Equation (4) can be used in several ways to deduce the
skin friction from an xt diagram not only for spatially
constant but also for skin friction fields with high
gradients. We shall first discuss briefly three of these
methods, then give a short account of an older method
for constant skin friction and finally provide a detailed
description of a fifth method (including an error estimate)
for constant skin friction.
(i) Integration of equation (4) yields
 2
Z
2 x h 0
h
dx .
(x) = a
2
ha
h xa t
1398

(5)

Figure 2. The experimental set-up for the oil-film


interferometry, methods (i)(v).

Thus, the distribution (x) can be determined, starting


either from zero height ha or from a known a , by spatial
integration of the temporal change of h at one or several
fixed times. The integral can also be written as

Z x
Z x
h 0
x
dx =
dh.
(6)

xa t
xa t h=constant

Skin-friction measuring techniques

A problem with this method may occur since the constant


Ia denotes the centre of the characteristics, which cannot
be determined in each case.
Figure 11 (later) shows the skin-friction distribution
downstream of a backwards-facing step obtained from the
xt diagram in figure 3 with methods (i)(iii). Good
agreement is observed.
If the relative spatial variation of the wall shear stress
(1 / ) is small compared with the slope of the film surface
(h/x), the wall shear stress can be assumed constant.
Two more methods will now be discussed for this case.
(iv) For constant skin friction near the leading edge of
a wetted surface area the similarity solution has the form
h=

Figure 3. Characteristics and similarity solutions in the x t


plane.

This is useful when the slope of the contour lines in the


xt diagram is finite everywhere along a line t = constant.
(ii) It can easily be shown that equation (4) possesses
characteristics in the xt plane which are identical to paths
of the particles that form the oil-film surface (figure 3).
These characteristics can be found by connecting the streaks
crossing the interference fringes in the xt diagram. The
slope of the characteristics in the xt plane is the particle
velocity up ; that is
= up / h.

(7)

This method can be applied only if a streaky structure of


the contour lines is discernible in the xt diagram, as is
the case in figure 3. The corresponding characteristics are
also shown in figure 3. Distinguishing the characteristics
from the contour lines becomes difficult if the skin friction
approaches a constant value, because characteristics and
contour lines coincide for = constant.
(iii) It can be shown that equation (4) has similarity
solutions h(x, t) = g(x)/t for large values of t. Obviously,
the function g(x) happens to have the same shape as the
contour lines h = constant in the xt diagram. These lines
are also shown in figure 3. It can be seen that the contour
lines are spread out (or drawn together) by characteristics.
From g(x) the skin friction can be obtained via
Rx
Ia + xa g(x 0 ) dx 0
.
(8)
(x) = 2
g 2 (x)

x
.
t

(9)

That is the oil film has the shape of a wedge. The


xt diagram shows a fan of fringes originating at the
leading edge of the film (figure 4). This relation was
already used by Tanner and Blows (1976). Later, Monson
(1983) developed a two-beam laser method for this case
which made the whole principle rather complicated and also
needed a rather complex measuring device. Therefore we
did not follow this route in our investigations but started
again from Tanners basic principles.
(v) For constant wall shear stress the equation (4)
becomes a simple advection equation for the contour lines
hk :
hk
hk
+ uk
=0
(10)
t
x
with uk = hk /. Since uk is equivalent to the slope of the
contour lines in the xt plane, namely

x
uk =
t hk =constant

(11)

it can be used to determine . Equations (10), (2) and (3)


yield
2[n2 sin2 ()]1/2
(k + h0 /1h)

(12)

2[n2 sin2 ()]1/2


h0
= uk
.
1h

(13)

= uk
and after rearranging
k +

In equation (13) , n, and are known and an arbitrary


number of fringe velocities uk can be measured from
the xt diagram. Thus equation (13) presents an overdetermined system of linear equations for the unknown
quantities and possibly h0 /1h (if h0 /1h is unknown).
and h0 /1h can then be computed by using a least square
fit. The advantage of this procedure is that the quantity
h0 /1h can be determined from the measurements and does
not need to be known a priori . This was important for
the measurements given below, because wall material other
than glass (video tape) was used.
1399

H H Fernholz et al

Figure 4. The x t diagram showing the development of an oil-film wedge generated in a flow with spatially constant wall
shear stress.

We now focus on some practical details of method


(v). The velocity uk was deduced by measuring its slope
1x/1t in the xt diagram, where 1x is the distance within
which a fringe travels within the time 1t. This slope was
determined by means of a computer program such that lines
could be set parallel to the fringe by hand. The coordinates
of the end points of these lines were then available in pixel
units.
The total duration of the measurement was taken by the
program to determine the conversion factor from time to
pixels. Typical measuring times may lie in the range from
10 min to 3 h, depending on the viscosity of the oil and the
wall shear stress. The speed of the oil film should not be too
high, for the flow needs a minimum of 12 min in order to
attain a steady state. Furthermore, there should be at least
ten fringes in order to obtain a stable mean value. The
conversion factor for the space units and the viewing angle
was determined from an image of an object with known
dimensions placed at the position of measurement. The
viewing angle was obtained from the perspective distortion
with an accuracy of 0.5%.
The physical properties of the oil, oil and noil , were
obtained from tables provided by the manufacturer. Usually
silicon oil is used since its dependence on the temperature
is smaller than that of other oils but still not negligible and
can be interpolated between the tabulated values 0, 25 and
1400

40 C by using

oil = AekT

(14)

with an accuracy better than 1%. For the computation of the


dynamic viscosity the density of the tabulated values was
linearly interpolated with an accuracy of better than within
0.5%. Measuring the density of the oil with a Mohr
balance reproduced the interpolated value within 0.2%. The
dynamic viscosity of the oil was measured with a Hoppler
viscosimeter and resulted in differences below 1% of the
computed value. The refractive index is fairly independent
of the temperature and the absolute pressure and can be
taken directly from the tables. Its value is 1.4 0.02 for
most oils.
The assumption of constant wall shear stress in the
area of measurement was checked by comparing method
(i) with method (v). Agreement within 1% between the
two methods was found.
The reproducibility of method (v) is 1%. The oilfilm method and the Preston tube method in a canonical
boundary layer agreed within 2%. Assuming that the
accuracy of the Preston tube method is 3%, the error
of the oil-film method should lie below 4%. Since the
oil-film method does not need to be calibrated, namely is
independent of the Preston tube measurements, its accuracy
can be better than within 4%. This can be shown by
using a method with higher accuracy than the Preston tube
for comparison.

Skin-friction measuring techniques

2.3. The wall hot-wire probe


Wall hot-wire and hot-film probes have been used to
measure the instantaneous local wall shear stress. Their
range of application is, however, quite different.
Wall-mounted hot films can be used for time-resolved
measurements in situations in which the conductivity of the
fluid is larger than the conductivity of the wall material.
This is possible for water and oil, but not for air. In
air, the transport of heat to the substrate and back into
the fluid leads to a reduced dynamic sensitivity of the
hot film (Tardu et al 1991). This effect decreases when
the thermal conductivity of the substrate is increased, but
even with walls made from aluminium or copper, the
RMS value of the wall shear stress in turbulent boundary
layers is underestimated by up to 25% (Dengel et al 1987,
Alfredsson et al 1987). Therefore we use only wall hotwire probes which retain their dynamic sensitivity and can
easily be corrected for heat transfer effects to the wall
(Wagner 1991).
The determination of the instantaneous skin-friction
from hot-wire measurements of the velocity gradient in
the viscous sublayer is strongly affected by the proximity
of the wall. However, in this case the distortion of
the temperature field affects only the sensitivity of the
probe, not its frequency response. The need for near-wall
measurements led to several investigations of the walldistance dependence of hot-wire probes. Janke (1987)
discussed these in detail and found a single relationship
to correct the mean velocity measured with constanttemperature anemometers (CTA) both in laminar and in
turbulent boundary layers. Thus one can conclude that the
mean response of the wall hot-wire probe is independent of
the turbulence structure of the flow.
The wall hot wire needs to be calibrated in a fixed
position against another instrument which measures the
local wall shear stress. This was done first by Bradshaw
and Gregory (1959), who determined only the mean skin
friction. Alfredsson et al (1987) described a probe in which
the hot wire is mounted several wire diameters above the
wall and calibrated against a Preston tube in a turbulent
zero-pressure-gradient boundary layer. At relatively high
Reynolds numbers, this technique leads to plausible results
both for the mean and for the RMS value of the wall shear
stress. However, as the Reynolds number is decreased the
sensitivity of the probe is reduced. This can be avoided
when the probe distance from the wall is increased but,
in order to preserve the linear relationship between the
velocity and the wall shear stress, the maximum distance
from the wall is limited. Fortunately, the wall conduction
effect scales with y + . Thus, there will always be a wall
distance at which the sensitivity is high enough and the
probe is still within the linear range of the velocity profile.
Criteria to find an optimum position will be given below.
In the near-wall region of both laminar and turbulent
shear flows, the velocity increases linearly with the
wall distance.
For dimensionless wall distances
y + = y u / < 5 with u = (w /)1/2 this is also valid for
instantaneous velocity profiles. Therefore we can determine

the linear relationship


w =

u
y

(15)

between the wall shear stress w and the velocity u at a


distance y above the wall. Due to heat conduction near the
wall the velocity um measured with a hot wire calibrated
in the free stream and the true velocity u in the near-wall
region differ from each other. These two velocities are
related by an empirical relationship (Janke 1987)
um y
uy
=
k1

uy

1/2
+ k2

(16)

which holds for laminar and turbulent boundary layers in


the range y + 5. In the following it is assumed that
equation (16) holds also for instantaneous values of u and
um , and so far we know of no experimental data which
contradict this assumption. For substrates with high thermal
conductivity (such as ceramics) and a hot-wire diameter of
5 m, k1 = 0.55 and k2 = 3.2.
For larger wall distances the true velocity is slightly
underestimated. We now consider a probe mounted at a
fixed distance h above the wall. With equations (15) and
(16) a relationship between m and the true skin friction w
can be given:

m = w k1

2
w
h2

1/2
+ k2

2
.
h2

(17)

The instantaneous voltage e of the CTA bridge must


now be related to the instantaneous wall shear stress w .
This can be done either by using a polynomial or by
applying Kings law. Here we have used Kings law
for the following reason. In all technically interesting
flows, the high skewness of the wall shear stress leads to
instantaneous values well above the highest mean value
that could be realized in the calibration process. Those
values can only be calculated when the calibration function
is extrapolated. Since polynomials tend to deviate strongly
from any physical law outside the calibrated region, we
prefer Kings formula for the linearization.
Kings law reads
e = (A + Bm )1/2 .

(18)

The Preston tube can only measure the mean wall shear
stress. Therefore the calibration function must be timeaveraged. This leads to

m =

e2 A
B

1/
(19)

Since this equation cannot be linearized, it is not possible


to determine the coefficients A and B directly. With a
computer-aided measurement system, the squared mean
values of the bridge voltage
h i
h i
 
e2 = e2 i + e02
(20)
i

1401

H H Fernholz et al

can be determined. A similar technique was used by


Ruderich and Fernholz (1986) for the calibration of wall
pulsed-wire probes. However, since the values
 


m i 6= m i
(21)
cannot be determined from the measurements, an error of
m m
r =
m

"

1/
=1

0
1+ m
m

 #1/
(22)

occurs. For 0.5 the time-averaged term can be


estimated to be
1 02
(23)
r '
4 2
by means of a series expansion. If we assume the
turbulence intensity to be of the order of 40%, the resulting
error for the wall shear stress is about 4%. This error can
be further reduced, when the time-records of the bridge
voltage taken at each calibration position are saved. In this
case, the mean values of the linearized bridge voltage can
be calculated and compared with the Preston-tube readings.
The calibration can be further improved by iteratively
correcting the coefficients A and B until good agreement
between the measured and the calculated values of the
mean wall shear stress is reached (Wagner 1991). This is
achieved by (i) calculating the difference between the wall
shear stress w as measured by a Preston tube and the value
computed from the time series e of the bridge voltage using
the first calibration curve, (ii) modifying the Preston-tube
readings in order to double the distance from the calibration
curve and (iii) re-calculating the coefficients A and B from
the modified values. Usually only one or two iterations are
necessary to achieve agreement within 1%.
The constants k1 and k2 in equation (16) hold only
for a specific combination of wall material and hotwire diameter and the calibration curve can be further
improved by determining k1 and k2 . Since deducing
k1 and k2 from velocity hot-wire measurements is very
time-consuming, they are computed directly from the
calibration curve (Warnack 1996). Re-arrangement of
Kings law and substitution of m = (w + k1 w 1/2 + k2 )
by (w + k1 w 1/2 + k2 ) with k1 = k1 1/2 / h and k2 =
k2 2 / h2 yields
w + k1 w 1/2 + k2 =

1 2 A
e
B
B

1/
.

(24)

After determining A and B for k1 = 0.55 and k2 = 3.2


equation (24) can be used to re-calculate improved values
of k1 and k2 . Usually the correlation coefficient approaches
unity very closely after several iterations.
Finally a few remarks on the design of the wall hot-wire
probe are in order. A sketch is shown in figure 1.
The body of the probe is made of 3 mm diameter steel
tubing with an insert consisting of a ceramic cylinder and
two 0.2 mm diameter Invar wires which serve as prongs.
The sensor wire which is soldered on to the prongs is a
2.5 m diameter platinum-coated tungsten wire with goldplated ends and an active length ` = 0.5 mm. This
1402

Figure 5. The bridge output voltage of a


constant-temperature anemometer driving a near-wall
hot-wire probe as a function of the dimensionless wall
distance .

ensures an `+ <
20 up to a skin-friction velocity smaller

than 0.6 m s1 and a ratio `/d = 200 (Wagner 1991).


The optimal wall distance of the hot wire can
be determined from two considerations. One is the
requirement that the hot wire be situated within the viscous
sublayer. The other one is that its sensitivity E = de/d
be near its optimum value, where e is given by equations
(17) and (18) and = h(w /)1/2 /.
Figure 5 presents e() and E() for a typical probe
(ceramic surface and wire diameter d = 2.5 m). The
sensitivity shows a flat maximum in the range 0.5 < <
2.5 and decreases fast for smaller values of . Therefore
the sensor wire must not be too close to the wall. In order
to maintain both sensitivity and linearity, the wall distance
of the probe must be adapted to the mean skin-friction
value. The best results are obtained, if the dimensionless
wall distance is in the range 0.5 h u / 5.
If the probe is built as described and if the above
calibration method is used, this skin-friction measuring
technique has a reproducibility below 1%. The accuracy
of the mean skin friction value w measured by a wall hot
wire is about 4% if the accuracy of the calibration device
(here a Preston tube) is assumed to be 3%.
2.4. The surface fence
The surface fence or sublayer fence, as a means of
measuring the magnitude and direction of the skin friction,
was first applied by Konstantinov and Dragnysh (1960)
and later by Head and Rechenberg (1962) and described
in detail by Vagt and Fernholz (1973). A sketch of a
surface fence is shown in figure 1. The fence height H
should not exceed a value of H + ' 5 and thus remain
within the viscous sublayer. This makes it a device which
is independent of the validity of the logarithmic law of the
wall.
Surface fences are easier to build than are wall hotwire probes and need only a precise manometer to read
the pressure difference 1p upstream and downstream of
the fence. Due to the small size of the fence and the
manufacturing tolerances, it is necessary in general to
calibrate each fence, although the flow around the fence
can also be calculated in principle.

Skin-friction measuring techniques

For the determination of the calibration function of the


surface fence it is appropriate to determine the relevant
parameters. We only consider the influence of the mean
velocity profile and exclude turbulence effects from our
analysis. We assume that the pressure difference 1p
between the front and the back of the fence located in the
viscous sublayer depends on
1p = 1p(uH , H, , )

(25)

where uH is the velocity at the fence height H . The skin


friction w is given implicitly through the linear relationship
u+ = y +
uH = H + u =

H u
H w
u =
.

(26)

We first investigate the Reynolds number range to


determine whether all terms of equation (25) are of equal
importance. The Reynolds number is defined as
Re =

H uH
2
= H+ .

(27)

The height of surface fences is usually chosen such that


0.5 < H + < 5. For H + < 0.5 the pressure difference
becomes so small that it cannot be measured accurately.
At H + 5 the buffer layer begins and equation (26) no
longer applies. Therefore, the Reynolds number lies in the
range 0.25 < ReH < 25.
At the lower end of the Reynolds number range we
may assume creeping flow and neglect inertia forces in
comparison with viscous forces. Equation (25) then reduces
to
1p = 1p(uH , H, )
(28)
which gives
1p H
= F(1)
uH

(29)

or, inserting equation (26)


1p w .

(30)

This shows that the pressure difference is independent of


the height of the fence, a fact first pointed out by Taylor
(1935) discussing the half-Pitot tube.
For high Reynolds numbers, we may neglect the
viscous forces in comparison with the inertia forces, which
is a crude assumption for ReH ' 25. Equation (25) then
reduces to
1p = 1p(uH , H, )
(31)
1p
= F(1).
u2H

(32)

With equation (26) we obtain


1p

w2 H 2
.
2

(33)

Now that the characteristics of the fence are known


at the lower and upper ends of its range of applicability,

Figure 6. A typical calibration curve of a fence.

we may formulate a calibration function by combining


equations (30) and (33) and obtain
1p = C1 w + C2 w2

H2
.
2

(34)

Introducing the dimensionless pressure difference 1p =


1pH 2 /( 2 ) and w = w H 2 /( 2 ), we write
equation (34) in non-dimensionalized form as
1p = C1 w + C2 (w )2

(35)

where C1 and C2 are constants to be determined from the


calibration process.
The desired relationship of the skin friction as a
function of the pressure difference 1p may now be
calculated:
"  
#1/2
1 C1
1 C1 2
1

+
+
1p
.
(36)
w =
2 C2
4 C2
C2
The advantage of equation (36) over Patels (1965)
power law 1p = wn ; n ' 1.5 is that it exhibits also the
correct asymptotic behaviour at the upper and lower ends
of the Reynolds number range, as suggested by dimensional
analysis.
Figure 6 shows a typical calibration curve obtained
using a least square fit based on equation (35). The
agreement between the fitted curve and the calibration
points is better than 1%. Together with the accuracy
of the Preston tube used as a calibration device, the
accuracy of a surface fence is about 4%. As will be
shown in section 3, the accuracy of the mean skin friction
measurement may decrease if the turbulence structure
of the flow under investigation differs strongly from
the turbulence structure of the flow of calibration. An
explanation could be that the the flow field around the fence,
the tubing and the pressure transducer presents a complex
nonlinear system.
It should be noted that Gur (1993) investigated the
applicability of the surface fence to measure fluctuating
values of the wall shear stress by using microphones built
into the probe.
1403

H H Fernholz et al
Table 2. Applications of the measuring techniques.

Type of flow
(authors)
Turbulent boundary layer
(Dengel et al 1987)
Turbulent boundary layer
(Warnack 1996)
Separation and recovery
regions downstream of a
normal plate with a
splitter plate
(Ruderich and Fernholz 1986,
Dengel et al 1987)
Separation and recovery
regions downstream of a
backwards-facing step
(Janke 1993)
Turbulent boundary layer with
and without reverse flow
(Dengel and Fernholz
1989, 1990)
Turbulent boundary layer with
and without reverse flow
(Gasser 1992,
Gasser et al 1993)

Wall
hot
wire

Wall
pulsed
wire

Pressure
gradient

Surface
fence

Oil-film
interferometry

Adverse

Favourable

Adverse

Adverse

Adverse

Adverse

X
X

Preston
tube

Method (v)

Hot
film

Floating
element
balance

Methods (i)(iii)

Figure 7. A comparison of three different measuring techniques for skin friction in an adverse pressure gradient (APG)
axisymmetric boundary layer layer. Data from Dengel et al (1987).

2.5. The wall pulsed-wire probe


The wall pulsed-wire probe, first investigated by Ginder
and Bradbury (1973), is one of the few probes capable
of measuring the instantaneous wall shear stress in highly
turbulent flows with flow reversal. Methods like oil-film
interferometry or surface fences can be applied to flows
with reverse flow regions but have poor temporal resolution
at their present stage of development.
The principle of the velocity pulsed-wire anemometry
was described in detail by Bradbury and Castro (1971) and
Castro (1992), for example and applications of the wall
pulsed-wire probe were given by Castro and Dianat (1983),
Castro et al (1987) and Dengel et al (1987). Therefore
1404

a brief description will suffice here. A wall pulsed-wire


probe, as shown in figure 1, consists of three parallel wires
mounted in a plane parallel to the wall and at a height
within the viscous sublayer. The central wire is heated by
a very short (5 s) electrical pulse, which generates a heat
tracer. The two sensor wires are operated as temperature
sensors and note the arrival of the heat tracer. The time of
flight T of the heat tracer is a measure of the instantaneous
wall shear stress.
The probe is conveniently calibrated against a Preston
tube. Thereafter, the probe can be used to determine the
mean and fluctuating part of the skin friction in regions
where the use of a Preston tube would lead to erroneous

Skin-friction measuring techniques

results. The range of the skin friction covered by the probe


is limited by the requirement that the probe not extend
the viscous sublayer. Furthermore, the instantaneous shear
stress must not create a time of flight shorter than the
shortest time which the pulsed wire anemometer is able
to measure (typically 100 s). Otherwise the truncation of
these high shear stresses will lead to severe errors.
It is thus necessary to adjust the probe geometry,
namely the wire spacing and wall distance, to the special
requirements of the flow under investigation. Typically,
however, the wires are located at a wall distance of 50 m
and separated by 0.7 mm from each other. The range of
skin friction of such a probe lies within 2 N m2 < w <
2 N m2 . The sensor wires are made of 2.5 m platinumcoated tungsten wires with gold-plated ends and an active
length of 0.5 mm. The pulsed wire is made of 5 or 9 m
platinum-coated tungsten wire. Thick pulsed wires give
strong signals but also produce larger wakes. For details
of the probe design the reader is referred to Dengel et al
(1987) or Castro et al (1987).
An appropriate relation between the wall shear stress
and the time of flight was given by Castro et al (1987):
w = A

 2
 3
 
1
1
1
+B
+C
T
T
T

(37)

where A, B and C are constants determined by the


calibration process. The advantages and disadvantages
of second- or third-order polynomials are discussed by
Handford and Bradshaw (1989). The calibration inside
a turbulent boundary layer makes it important to timeaverage equation (37), as shown by Ruderich and Fernholz
(1986). Due to the turbulent fluctuations, the instantaneous
time of flight recorded by the pulsed-wire probe may vary
widely for one calibration point, that is one mean shearstress value known from the Preston tube measurement.
The corresponding equation used during the least square fit
of the calibration process is
 
 2
 3
1
1
1
+B
w = A
+C
.
T
T
T

Figure 8. The skin-friction distribution measured by four


measuring techniques in a favourable pressure gradient
(FPG) turbulent boundary layer without re-laminarization
FPG. Data from Warnack (1996).

Figure 9. The skin-friction distribution measured by four


measuring techniques in a FPG turbulent boundary layer
with re-laminarization FPG. Data from Warnack (1996).

(38)

Equation (38) takes the turbulence structure into account,


such that the probe may also be used in flows with
turbulence structures that differ strongly from the flow used
for the calibration.
Note, that due to its low sampling rate ( 40 Hz)
the wall pulsed-wire probe is not well suited to obtain
spectra. Besides the error caused by the calibration device
the resolution of the time of flight counter becomes a source
of error for high values of the skin friction, typically 1% for
w > 1.5 N m2 . The combined errors of calibration and
time of flight measurement can accumulate to an overall
error of about 4%.
3. Application and comparison of the measuring
techniques
All measuring techniques for the wall shear stress presented
in section 2 have been applied to turbulent boundary layers

with variable pressure gradients and to wall bounded flows


with strong reverse-flow regions. In table 2 the different
types of flow and the measuring techniques are compared.
We shall discuss first measuring techniques applied to
attached boundary layers with adverse and favourable
pressure gradients and then flows with strong and weak
reverse flow.
The first comparative tests were performed in an
axisymmetric adverse pressure gradient turbulent boundary
layer on a cylinder with its axis in the streamwise direction
Dengel et al (1987). Three techniques were compared
(Preston tube, surface fence and wall pulsed wire) upstream
of separation. The results are presented in figure 7 and show
on the whole good agreement.
Skin-friction measurements in favourable pressure
gradient boundary layers are still scarce and therefore
it seemed to be appropriate to compare four measuring
techniquessurface fence, wall hot-wire, Preston tube
and oil-film interferometry in accelerated flows (Warnack
1405

H H Fernholz et al

Figure 10. A comparison of three measuring techniques for skin friction in a wall-bounded turbulent shear layer with strong
reverse flow (xR = 0.38 m). Data from Dengel et al (1987).

1996, Fernholz and Warnack 1996). Patel (1965) observed


already the breakdown of the standard logarithmic law
of the wall in a highly accelerated boundary layer. In
such a flow the Preston-tube method fails. Therefore
Patel and Head (1968) used a surface fence for their
measurements. In our experiment the favourable pressure
gradient boundary layers were generated in a 0.44 m
Plexiglas pipe by means of two centre bodies and the skin
friction was measured on the pipe wall. The mean skin
friction w was made dimensionless with the reference
quantity u20 at the first measuring station where the
pressure gradient is zero and plotted against the streamwise
distance x in figures 8 and 9. Figure 8 shows case 1
which has the lowest acceleration (k 2 106 , k =
[/(u3 )]dp/dx). It reveals the breakdown of the
Preston-tube method in the favourable pressure gradient
region. The wall hot-wire and the surface-fence data
collapse on each other. Figure 9 presents case 2 which
has a highly accelerated flow region k 4 106 and
shows the deviation of the surface-fence data from those of
the wall hot-wire and the oil-film technique in the region
around x 1.6 m. In this region, the turbulence structure
differs strongly from the zero-pressure-gradient boundary
layer, where the fence was calibrated. This is indicated
by the skewness of the wall shear stress. The surface-fence
data are supposed to be incorrect due to the nonlinear effects
caused by the change in the turbulence structure. Neither
the oil-film nor the wall hot-wire technique is affected by
these effects (see section 2). Obviously, the change in
turbulence structure in case 1 is not large enough to cause
discrepancies between the surface-fence and the wall hotwire methods.
The next two flows are wall-bounded turbulent shear
flows with strong reverse-flow regions (Fernholz 1994).
The first flow is generated by a normal plate with a long
splitter plate Ruderich and Fernholz (1986), whereby the
1406

Figure 11. A comparison of three measuring techniques


for skin friction in a wall-bounded turbulent shear layer with
strong reverse flow. Data from Janke (1993).

oncoming flow is separated at the sharp edge of the normal


plate and re-attached on the splitter plate. Three measuring
techniques (wall pulsed-wire, Preston tube oriented in the
respective flow direction and surface fence) are compared
in the reverse-flow region and in the recovery region
downstream of re-attachment (figure 10). Only the wall
pulsed-wire and the surface-fence methods show the correct
skin-friction distribution. The Preston tube fails because the
logarithmic law does not exist in the entire region presented
here (Ruderich and Fernholz 1986).
The second strong reverse-flow (figure 11) is
downstream of a straight backward-facing step (Janke
1993). Here we have compared the wall pulsed-wire and
the surface-fence with the oil-film method using evaluation
methods (i)(iii) (see section 2). The three oil-film methods
agree well with each other and with the wall pulsed-wire

Skin-friction measuring techniques

Figure 12. The distribution of the pressure coefficient cp , skin-friction coefficient cf and reverse-flow parameter w for three
APG boundary layers. All tagged data (O0 , 0 , 0 ) were measured by a wall-pulsed wire probe. From Dengel and Fernholz
(1990).

Figure 13. A comparison of four measuring techniques for skin friction in an APG-bounded turbulent boundary layer. From
Gasser et al (1993).

technique. Small deviations occur for the surface-fence


measurements close to separation and re-attachment. The
interpolated data from the wall pulsed-wire probe show
re-attachment slightly upstream of those of the oil film.
Both discrepancies could be caused by the probe sizes in
the x direction which were used in a rather small reverseflow region (about 40 mm total length) so that the oil film
technique probably gives the true re-attachment point.
The ability of the wall pulsed-wire probe to distinguish
between small positive and negative values in a highly
turbulent flow, even to determine w = 0, is shown in
figure 12. The three boundary layers (Dengel and Fernholz
1990) which were measured in the same axisymmetric test
section as the one described in figure 7 differ slightly in
their pressure distribution upstream and thus in their skin-

friction distributions, resulting in cases in which w is


slightly positive, negative or zero over a longer distance. It
is important to note that instantaneous reverse flow occurred
in all three cases as shown by the distribution of the reverseflow parameter w (measured with the wall pulsed-wire
method).
A similar adverse pressure gradient turbulent boundary
layer with separation and re-attachment and downstream
recovery of the flow in a mild favourable pressure
gradient was investigated by Gasser (1992). A comparison
among the results of four measuring techniques used
in this boundary layer (surface-fence (the probes were
manufactured at the HFI), wall pulsed-wire (the probes
were manufactured at the HFI), Preston tube and floatingelement balance) was given by Gasser et al (1993).
1407

H H Fernholz et al

The experiments were conducted on the inner wall of


an annular pipe of 200 mm diameter where the pressure
distribution was generated by suction of air through a
perforated concentric inner tube. Thus a family of boundary
layers could be generated and at the same time the locations
of the balance and the other skin-friction probes were fixed
while the pressure distribution was moved. The readings
of the different probes were averaged over 72 points,
distributed along the circumference of the pipe, compared
with the reading of the ring-type floating-element balance
of 10 mm axial length and accuracy 0.01 Pa. Figure 13
shows the comparison of the results obtained with the
four devices in four different pressure distributions (Gasser
et al 1993). The deviations among the four techniques
were not completely explained but the wall pulsed-wire
and the floating-element-balance methods show fairly good
agreement at all measuring positions whereas the Preston
tube reads consistently lower than the balance in the region
xA > 200 mm where the boundary layer is in a mild
favourable pressure gradient. The probes read higher
than the balance in regions with strongly decreasing w
(xA < 100 mm).
A more limited comparison (Preston tube and surface
fence) in the flow downstream of a backwards-facing step
was performed by Kiske et al (1981), who already found
discrepancies between the two methods in disturbed wall
bounded turbulent flows. We have not shown detailed
measurements of the fluctuating shear stress (w02 )1/2 in this
review. Such data exist and may be found, for example, in
the investigation of Ruderich and Fernholz (1986), Dengel
et al (1987), Dengel and Fernholz (1990), Janke (1993),
Wagner (1995) and Warnack (1996). There are hardly any
comparisons between measurements of wall pulsed-wire
and wall hot-wire probes, since the application range of the
two probes lies in separated or attached flows, respectively.

4. Conclusions
Four skin-friction measuring techniques in turbulent wall
bounded shear flows were presented. The techniques have
been developed further, probes were designed and built
and they were used in normal and complex flows. The
range of applicability of these measuring techniques was
investigated and they were compared with each other and
with data from a floating-element balance and from Preston
tubes. The Preston tube was also used to calibrate three of
the four techniques investigated herethe oil-film method
is absolute.
All four techniques can measure the mean skin friction,
wall hot-wire and wall pulsed-wire fluctuating quantities,
and oil-film and wall pulsed-wire skin friction in flows in
which the flow direction reverses. They are all independent
of the validity of the standard logarithmic law of the
wall. The four techniques complement each other and have
ranges in which they overlap, but there are other ranges
within which only one or two may be used.
In complex flows no general rule for an application can
be given but some specific points should be addressed.
1408

(i) Oil-film interferometry and wall hot-wires may


safely be applied in attached boundary layers and wallbounded shear layers with variable pressure gradients.
For wall hot-wires there must be no flow reversal.
In highly accelerated boundary layers the surface-fence
measurements may no longer be correct if the fence was
calibrated in a canonical boundary layer. The breakdown
of the Preston-tube method in highly accelerated boundary
layers, as already found by Patel (1965), was confirmed.
(ii) In flows in which the flow direction reverses
or downstream of re-attachment Preston tubes must not
be used. The wall-pulsed wire, the oil-film and the
surface-fence techniques may be used in such flows but
small deviations occur for the surface-fence measurements
close to separation and re-attachment because of small
asymmetries of the probe.

Acknowledgment
The authors acknowledge the financial support of the DFG.
References
Alfredsson P H, Johansson A V, Haritonidis J H and Eckelmann
H 1987 The fluctuation wallshear stress and the velocity
field in the viscous sublayer Phys. Fluids A 31 102633
Bechert D W 1995 On the calibration of Preston tubes AIAA J.
34 2056
Bechert D W, Hoppe G and Reif W-E 1985 On the drag
reduction of the shark skin AIAA Paper 85-0546
Bechert D W, Hoppe G, van der Hoeven J G H and Makris R
1992 The Berlin oil channel for drag reduction research
Exp. Fluids 12 25160
Blasius H 1907 Grenzschichten in Flussigkeiten mit kleiner
Reibung Dissertation Gottingen.
Bradbury L J S and Castro I P 1971 A pulsed-wire technique for
velocity measurements in highly turbulent flow J. Fluid
Mech. 22 67987
Bradshaw P and Gregory N 1959 The determination of local
turbulent skin friction from observations in the viscous
sub-layer, Reports and Memoranda 3202, ARC, London
Castro I P 1992 Pulsed-wire anemometry Exp. Therm. Fluid Sci.
5 77080
Castro I P and Dianat M 1983 Surface flow patterns on
rectangular bodies in thick boundary layers J. Wind. Eng.
Ind. Aero. 11 10719
Castro I P, Dianat M and Bradbury L J S 1987 The pulsed-wire
skin-friction measurement technique Proc. 5th Symp. on
Turbulence and Shear Flows
Dengel P and Fernholz H H 1989 Generation of and
measurements in a turbulent boundary layer with zero skin
friction Advances in Turbulence vol 2, ed H H Fernholz and
H E Fiedler (Berlin: Springer) pp 4327
1990 An experimental investigation of an incompressible
turbulent boundary layer in the vincinity of separation
J. Fluid Mech. 212 61536
Dengel P, Fernholz H H and Hess M 1987 Skin-friction
measurements in two- and three-dimensional highly
turbulent flows with separation Advances in Turbulence
vol 40, ed G Compte-Bellot and J Mathieu (Berlin:
Springer) pp 4709
Dickinson J 1965 The determination of turbulent skin friction
PhD Thesis Laval University Quebec (now at the
Department de Genie Mecanique)
Fernholz H H 1994 Near-wall phenomena in turbulent separated
flows Acta Mechanica 4 5767

Skin-friction measuring techniques


Fernholz H H and Finley P J 1977 An initial compilation of
compressible turbulent boundary layer data AGARDograph
223 Advisory Group Aerospace Research and Development
1981 A further compilation of compressible turbulent
boundary layer data with a survey of turbulence data
AGARDograph 263 Advisory Group Aerospace Research
and Development
Fernholz H H and Warnack D 1996 The effect of a favourable
pressure gradient and of the Reynolds number on an
incompressible axisymmetric turbulent boundary layer
Part. 1 The turbulent boundary layer, submitted for
publication
Finley P J and Gaudet L 1995 The Preston tube in adiabatic
compressible flow Exp. Fluids 19 13341
Frei D and Thomann H 1980 Direct measurements of skin
friction in a turbulent boundary layer with a strong adverse
pressure gradient J. Fluid Mech. 101 7995
Gasser D 1992 Experimentelle Untersuchung stark verzogerter
turbulenter Grenzschichten Dissertation ETH Zurich.
Gasser D, Thomann H and Dengel P 1993 Comparison of four
methods to measure the wall shear stress in a turbulent
boundary layer with separation Exp. Fluids 15 2732
Ginder R B and Bradbury L J 1973 Preliminary investigation for
skin friction measurements in highly turbulent flows, ARC
Report
Gur Y 1993 Mean and fluctuating wall shear stress
measurements PhD Thesis Massachusetts Institute of
Technology, Mechanical Engineering Department
Handford P M and Bradshaw P 1989 The pulsed-wire
anemometer Exp. Fluids 7 12532
Hanratty T J and Campbell J A 1983 Measurement of wall shear
stress Fluid Mechanics Measurements ed R J Goldstein
(New York: Hemisphere) pp 559615
Haritonidis J H 1989 The measurement of wall shear stress
Advances in Fluid Mechanics Measurements ed
M Gad-el-Hak (Berlin: Springer) pp 22961
Head M R and Rechenberg I 1962 The Preston tube as a means
of measuring skin friction J. Fluid Mech. 14 117
Head M R and Vasanta Ram V 1971 Simplified presentation of
the Preston tube calibration Aeronautical Quarterly XXII
295300
Hirt F and Thomann H 1986 Measurements of wall shear stress
in turbulent boundary layers subject to strong pressure
gradients J. Fluid Mech. 171 54762
Hirt F, Zurfluh U E and Thomann H 1986 Skin-friction balances
for strong pressure gradients Exp. Fluids 4 296300
Janke G 1987 Hot wire in wall proximity Advances in
Turbulence ed G Compte-Bellot and J Mathieu (Berlin:
Springer) pp 48898

Janke G 1993 Uber


die Grundlagen und einige Anwendungen

der Olfilm-interferometrie
zur Messung von
Wandreibungsfeldern in Luftstromungen Dissertation
Technische Universitat Berlin
Jiang F, Gupta B, Tai Y C and Goodman R 1995 Measurement
of instantaneous turbulent shear stress distribution by
MEMS based sensors Bull. APS 40 12, FC 2
Kempf G 1929 Neue Ergebnisse der Widerstandsforschung
Werft, Reederei, Hafen II 2349
Kim K-S 1989 Skin friction friction measurements by laser
interferometry in supersonic flows PhD Thesis Pennsylvania
State University

Kiske S, Vasanta Ram V and Pfarr K 1981 The effect of


disturbed turbulence structure on the Preston-tube method
of measuring wall shear stress Aeronautical Quarterly
XXXII 35467
Konstantinov N I and Dragnysh G L 1960 The measurement of
friction stress on a surface, English translation, DSIR RTS
1499
Monson D J 1983 A nonintrusive laser interferometer method
for the measurement of skin friction Exp. Fluids 1 1522
Nguyen V D, Dickinson J, Jean Y, Chalifour Y, Anderson J,
Lemay J, Haeberle D and Larose G 1984 Some
experimental observations of the law of the wall behind
large-eddy breakup devices using servo-controlled skin
friction balances AIAA Paper 840346
Padmanabhan A, Goldberg H D, Breuer K S and Schmidt M A
1995 A silicon micromachined floating-element shear-stress
sensor with optical position sensoring by photodiodes
(private communication) Bull. APS 40 12, FI 1
Pan T, Hyman D, Mehregang M, Reshotko E and Garverik S
1995 Micro shear stress sensors with direct electrical
readout Bull. APS 40 12, GE 5
Patel V C 1965 Calibration of the Preston tube and limitations
on its use in pressure gradients J. Fluid Mech. 23 185208
Patel V C and Head M R 1968 Reversion of turbulent to laminar
flow J. Fluid Mech. 34 37192
Ruderich R and Fernholz H H 1986 An experimental
investigation of the structure of a turbulent shear flow with
separation, reverse flow, and re-attachment J. Fluid Mech.
163 283322
Schmidt M A, Howe R T, Senturia S D and Haritonidis J H 1988
Design and calibration of a microfabricated floating-element
shear-stress sensor IEEE Trans. Electron Devices 35 7507
Tanner L and Blows L 1976 A study on the motion of oil films
on surfaces in air flow, with application to the measurement
of skin friction J. Phys. E: Sci. Instrum. 9 194202
Tardu S, Pham C T and Binder G 1991 Effects of longitudinal
diffusion in the fluid and of heat conduction to the substrate
on the response of wall hot-film gauges Advances in
Turbulence vol 3, ed A V Johansson and P H Alfredsson
(Berlin: Springer) pp 50613
Taylor G I 1935 Measurements with a half-Pitot tube Proc. R.
Soc. A 166 476
Vagt J D and Fernholz H H 1973 Use of surface fences to
measure wall shear stress in three-dimensional boundary
layers Aeronautical Quarterly XXIV 8791
Wagner P M 1991 The use of near-wall hot-wire probes for
time-resolved skin-friction measurements Advances in
Turbulence vol 3, ed A V Johansson and P H Alfredsson
(Berlin: Springer) pp 5249
Wagner P M 1995 Koharente Strukturen der Turbulenz im
wandnahen Bereich von Ablosegebieten Dissertation
Technische Universitat Berlin
Warnack D 1996 Eine experimentelle Untersuchung
beschleunigter turbulenter Wandgrenzschichten Dissertation
Technische Universitat Berlin
Winter K G 1977 An outline of the techniques available for the
measurement of skin friction in turbulent boundary layers
Prog. Aerospace Sci. 18 157
Zurfluh U E 1984 Experimentelle Bestimmung der
Wandschubspannung in turbulenten Grenzschichten
Dissertation ETH Zurich

1409

Das könnte Ihnen auch gefallen