You are on page 1of 57

P~og. A~rospuee Sci. 1977, Vol. 18, pp. I - 57.

Pergamon Press.

Printed in Great Britain.

AN OUTLINE OF THE TEChnIQUES AVAILABLE FOR THE MEASUREMENT OF SKIN FRICTION IN TURBULENT BOUNDARY LAYERS* K. G. Winter Royal Aircraft Establishment, Farnborough, Hants, U.K.

Summary-The techniques covered include force-mdasur~ment balances, the use of the velocity profile, pressure measurements by surface pitot tubes or about obstacles, and the use of the analogies of heat transfer, mass transfer or surface oil-flow. Hot-wire or laser techniques for determining the shear stress within the fluid are not included. The sources of error and ranges of application of the various techniaues are discussed.

i.

INTRODUCTION

In most applications of fluid mechanics a knowledge of the drag created by fluid flowing over a solid surface is essential to the understanding of the performance of a system whether it be a ship or an aircraft or the flow through a pipe. voted to the measurement of skin friction. Considerable effort has therefore been de-

This brief review concerns itself only with exter-

nal flow and with measurements primarily related to the performance of aircraft.

It was, however, the need to estimate the performance of ships which led to the first measurements at high Reynolds number. Probably the first systematic investigations were made over

iO0 years ago by Froude (1872) who measured the drag of a series of planks towed at various speeds along a tank using the elegant apparatus shown in Fig. I. It is interesting to note

that at that time even the qualitative effect of Reynolds number on skin friction was not generally understood. Froude did apparently have a concept of a boundary layer and states:

The investigation of skin friction may be separated into three primary divisions: (i) the law of the variation of resistance with the velocity; (2) the differences in resistance due to differences in the quality of surface; (3) the differences in the resistance per unit of surface due to differences in the length of surface. The necessity of investigating the latter of these conditions may not be at once apparent, it having been generally held that surface-friction varies directly with the area of the surface, and will be the sanm for a given area, whether the surface be long and narrow or short and broad. It has always seemed to me to be impossible that this should be the case, because the portion of the surface that goes first in the line of motion, in experiencing resistance from the water, must in turn co~mmnicate to the water motion in the direction in which it itself is travelling, and consequently the portion of the surface which succeeds the first will be rubbing not against stationary water, but against water partially moving in its own direction, and cannot therefore experience as much resistance from it. If this reasoning holds good, it is certain that doubling, for instance, the length of a surface, though it doubles the area, would not double the resistance for the resistance of the second half would not be as great as that of the first.

*Notes prepared for von Karmon Institute for Fluid Dynamics Lecture Series on "Compressible Turbulent Boundary Layers", March 1 - 5, 1976~ Copyright ~ HMSO (London) 1976.

~d

v~

"0

AA Bed of carriage. BB Plane, the surface friction of which is to be recorded CC Horizontal beam carrying plane and transmitting resistance of the same to spring by means of looped Connecting line hh. DD Cutwater fixed to Beam CC by horns aa. EE "Parallel motion" supporting beam CC I F Counterbalance to beam CC 8 c GG Lever arrangement for steadying the apparatus taking the strain off the spring while uniform speed is being. obtained. H Spring, extension of which measures resistance. KK !ndex ar m communicating extension of L spring to cylinder. W Counterbalance to index arm. L Fulerum of index olrm carried by bar'bb. "~.~ IV. ~ _ ~ I it" ~ .]1it / I L ~ ~ a ~ r j t~ii i u -;~" ~ MM Lever communicating extension of springto index arm. N Connecting link, medium of communication of extension of Spring to index arm, O0 Towing beam, holding fore end of spring. P Brass cap, about which bband MMl~inge. QQ Bar uniting head of lowing beam and c a p P t o frame car ryinql cylinder. RR Bell crank for extending spring by known weights hung on at c, thereby testing scale. dd Connection of bell crank with spring. e Weight giving initial extension to spring forming zero of Scale. S Pen registering extension of spring. UU Revolving cylinder receiving paper on which is registered motion of pen ~ c. TT Frame carrying cylinder ~ c , c a p o b l e of vert ca ~ adjustment. V Time penworked by clock work. " .'I" .

o~

,
i,~II I~)J ~ p ;~!~:" ~l" ~ 1 t l / / LL //l cylinderUU

/ /

~l~

J l II, _U__

u~

'~1

/i~l ~. 1..~

, /!

~eor for transferring mot,o, of oarrage to

C~

rt

"

\.

- : - ' ~ - " - '~

!'i"

"~

/
Rail

Rail~

e~

Surface
i

Fece of W a t e r

. . . . . . . .

B'~

BI

inchestl21, 6~
i i

,~
I i I i i i 1" i

5
|

Skin friction in turbulent boundary layers

Later K e ~ f

(1929) ~ d e

measure~nts

of local skin friction at several stations along the bot~ u n t e d on balances (Fig. ~e direct ~asurement

tom of a pontoon 77 m long using fairly large panels (309 x I010 m ) 2). These measure~nts achieved Re~olds numbers of up to 5 x 108.

of skin friction by force balance was an essential step in setting ~ laws and these measurements of K e ~ f

the basic skin friction

together with those of others, notably Schoenherr (1932),

formed the basis for the generally-accepted skin friction esti~tion for incompressible flow (see, for exa~le; Goldstein 1938). Because of our limited understanding of turbulent flows

-!

Platte I-

,-

o.~J'~30 ~ IO00mm~

[ t -I
.

,],
I ~Jfh~ngun~ / o

N,

aor

-.-"- ~". . . . . . . . . . ] ~ II

F--- ~ ",Feder spannmol"or

H
i i
F
m

Eichgewichte

Gml~ewicht

~J.O mX,,~ I 0 1 0 - 3 0 9 ~

Bewegliche Versuchs~platte

.LI.O mm~ Uhr-

"x~erk --3chreibtrommel

V
Fig. 2. Apparatus used by Kempfe for measurement of local skin-frlction coefficient.

K.G.

~inter

there has been the need to extend direct skin-friction measurements to compressible flows but the difficulty of applying the technique in many situations such as flows with pressure gradients has led also to the development of indirect means of obtaining skin friction. Aerodyna-

mics is largely an empirical science and many experiments are made in which the pressure distribution over bodies is measured for comparison with calculation, or as an aid to understanding the characteristics of the flow. It is the author's view that the value of many of these

experiments would be considerably enhanced if, in addition, the skin-friction distribution were also to be measured, even if only approximately, by one of the simple techniques discussed in the paper. A good review of the variety of techniques which have been devised to cover

the diversity of situations encountered in practice is given by Brown and Joubert (1967). They present a relative classification of techniques on which the following chart has been based. Classification of techniques for measuring skin friction Wall shear stress-- -- Wall similarity--F-Velocity profiles measurement --Liquid tracers

~Analogies
~ - - F l o w about obstacles

Heat transfer L Mass transfer Preston tubes

-- Momentum balance -- Direct measurement

Stanton tubes Razor blades Steps and fences Static pressure holes

The remainder of the paper discusses the various techniques as outlined in this classification and omits any consideration of hot-wire or laser techniques by which the shear stresses within a fluid may be measured and the wall shearing stress obtained by extrapolation. The use

of the momentum equation is also not discussed, since in principle this is straightforward, but in practice is difficult because it is not easy to take account of the three-dimensionality of the flow which generally occurs even in nominally two-dimensional situations, and also because often the skin-friction term can only be derived as the difference of two large terms.

2. ! 2 defined by a2 _ y - I M~/TWe
-

SYMBOLS

, also speed of sound

+-~-~-1

A,B AI, A2 C

constants in the law of the wall area of pitot tube constant in Spence's law of the wall, alson non-dimensional thickness of viscous sub-layer

u 6L/~

cf C P d

local skin-friction coefficient TW/0U2 specific heat of air at constant pressure diameter of Preston tube diameter of static pressure hole diameter of floating element of skin-friction balance force on skin-friction balance

Skin friction in turbulent boundary layers

g h
i

gap round element of skin-friction balance height of step, height of razor blade, thickness of film of oil, enthalpy current molecular diffusion function in equation (8-12) thermal conductivity of air thermal diffusivity of air k/OCp mass transfer coefficient
m

J J k
K

length,

length of heated element, length of mass transfer element, distance apart

of elements in Section 8.3


m

rate of mass transfer per unit area Mach number friction Mach number uT/aw
T

M M P Pf' Pr q
R

pressure pressure-rise parameters for forward and rearward-facing steps (Ap/T W) rate of heat flow per unit area electrical resistance Reynolds number based respectively on streamwise length, momentum thickness~ Preston tube diameter and length of heated element radius of floating element of skin-friction balance, also temperature recovery factor

Re x, Re 8 ,

developed total velocity, equation (4-16) area of floating element of skin-friction balance thickness of floating element of skin-friction balance, also time temperature velocity at edge of boundary layer velocity in direction of U friction velocity
T =

S t T U
u

(u 2 + w 2)

P
v W w

gj
velocity of wall injection cross flow velocity wake component of velocity profile s treamwise coordinate coordinate normal to wall Apd 2 Twd2 Preston-tube coordinates x* = log 4pv2 , y* = log 40--~/~

w(y/6)
x

Y x*, y*

K.G.

Winter

razor-blade f~ B

Aph 2 coordinates x* = log 7

Twh2 ,y* = log--~-f~

kinematic pressure gradient ~ dp P dx crossflow angle in boundary layer crossflow angle at wall heat transfer parameter q/PxDpTwU T w ratio of specific heats boundary layer thickness thickness of viscous sub-layer thickness of thermal layer

Bo
Bq
Y ( S
6L q A P

pressure gradient parameter ~--~


u3 T

't

shear-stress Y

gradient p a r a m e t e r - - - pu 3 ~y

T1 0

O_ dy
0m

variation of g with displacement of skin-friction balance von Karman constant viscosity kinematic viscosity mass concentration density ~C Prandtl number - ~
(~ - I)M e
e

, also o =

1 + ~(y - I)M 2 e T Subscripts e m edge of boundary layer intermediate-enthalpy conditions shearing stress

adiabatic wall conditions w Superscript i refers to incompressible flow. wall conditions

Skin friction in turbulent boundary layers

3.

DIRECT MEASUREMENT

Apart from the experiments of Schutz-Grunow (1940) interest in the direct measuren~ent of skin friction* lapsed until the increasing speed of aircraft called for precise measurements in compressible flows. As a result there have been numerous skin-friction balances designed over

the past few years, and much ingenuity has been exercised in design to overcome the essential problem of obtaining accurate measurements of the shear forces which are very small, three orders less than the inertial forces.

The problems which have to be considered are listed below. (i) Provision of a transducer for measuring small forces or deflections, and the compromise between the requirement to measure local properties and the necessity of having an element of sufficient size that the force on it can be measured accurately. (2) (3) (4) (5) The effect of the necessary gaps around the floating element. The effects of misalignment of the floating element. Forces arising from pressure gradients. The effects of gravity or of acceleration if the balance is to be used in a moving vehicle. (6) (7) (8) (9) Effects of temperature changes. Effects of heat transfer. Use with boundary-layer injection or suction. Effects of leaks.

(i0) Protection of the measuring system against transient normal forces during starting and stopping if the balance is to be used in a supersonic tunnel.

The basic choice to be made is the size of the floating element which dictates the sensitivity required of the measuring system which may he passive (displacement) or active (force-feedback). To illustrate the sensitivity required the table below shows over a range of Maeh

number the force in milligrams to be measured by a balance with a head of IOnnn diameter in a zero-pressure gradient flow at one atmosphere stagnation pressure and a Reynolds number of IO million. M Force mg 0.I 16 0.5 290 i 680 2 540 3 210 A very sensitive transducer

For adverse pressure-gradient flows the forces will be even less.

is therefore needed and a variety of transducers and sizes of floating element has been used. The table on the following page lists some of the designs developed in the past few years.

The balance used by Schultz-Grunow (1940) is included for historical interest.

In this

balance the floating element was rather large and was mounted on offset torsional pivots and restrained by a torsion bar. With the exception of the balance of Ozarapoglu (1973) (Fig. 5)

in which the floating element is supported on air bearings, the remaining balances have either a parallel-linkage supporting arrangement (Fig. 3) or effectively a pivot below the floating element (Fig. 4). A similar arrangement to that of Schultz-Grunow was used in a small balance In their balance the floating element was pivoted about The most popu-

by Kovalenko and Nesterovich (1973).

an axis normal to its surface with the axis offset to one side of the element.

lar device for detecting the position of the floating element is a Linear Variable Differential

*It might be noted that the classical work of Wieghard (1942) on the drag of surface excrescences made use of direct force measurements.

Reference

Test c o n d i t i o n s

Size of f l o a t i n g element (nmO

Type of s u s p e n s i o n / p o s i t i o n / force t r a n s d u c e r

Force range

Schultz-Grunow 300
x

1940 500* Optical/manually-operated offset t o r s i o n bar Parallel interchangeable LVDT

U = 2Om/s

1.6 x 106 < Re < 60 x I0"


x

x 11.5 63 x < 1.2


x

< 16 x 106

Dhawan 106 2 x 20

1953

L o w speed 6 x 104 < Re

20 m g to 800 m g

S u b s o n i c 0.2 < M < 0.8, 0.3 x 106 < Re

S u p e r s o n i c 1.24 < M < 1.44 x < 9 x 106


x

Coles 6.2 x 37.9 Parallel linkage LVDT T r a n s l a t i o n of e l e m e n t support by m i c r o m e t e r 25 dia 50 dia LVDT

1953

M = 1.97 0.4 x 106 < Re

< I0 x 106

3 g

M = 2.57 0.4 x 106 < Re x < 8 x 106


x

M = 3.70 0.5 x 106 < Re < Re

< 8 x 106

M = 4.54 0.4 x i0

Weiler & H a r t w i g 1952

S u p e r s o n i c w i n d tunnel

Lyons

1957

S u p e r s o n i c flight

D o u b l e p a r a l l e l interc o n n e c t e d linkage to e l i m i n a t e s e n s i t i v i t y to linear and rotational accelerations LVDT

30 g

Smith & 50 dia

0,ii < M < 0.32

P a r a l l e l linkage LVDT Kelvin current b a l a n c e

14 g

Walker

1958

106 < Re

< 40 < 106

MacArthur

1963

Shock tunnel

6.4 dia

Parallel

linkage P i e z o e l e c t r i c beams

5 g

Moulic

1963

M = 6 L o w density

0.25 x 25

Side flexure pivot LVDT 25 dia Parallel linkage LVDT

20 m g

Young

1965

Supersonic flow w i t h heat transfer and surface roughness transfer

Dershin et al.

1966

Supersonic flow w i t h mass

"Pointed ellipse"

Parallel

linkage LVDT

*Estimated from sketch

Moore & Pneumatic position sensor High temperature motor 19 dia LVDT 127 dia LVDT Permanent magnet plus coil 20.3 dia LVDT Motor-driven spring Flexure pivot 2 g Flexure pivot 200 g Parallel linkage 200 mg

High temperature hypersonic flows

Flexure pivot

3 g

McVey

1966

Brown &

Low-speed adverse pressure gradients

1969

Joubert

Fowke

1969

Supersonic speeds

Bruno,

Supersonic speeds including flows with heat transfer

Yanta &

1969

Risher

Var iab ie by changing loading spring 18OO g


o" =

Winter &

0.2 < M < 2.8

368 dia

Parallel linkage Resistance strain gauges

Gaudet 7.9 dia

1970

16 x 106 < Re

< 200 106

Hastings &

M=

Parallel linkage LVDT

500 mg
rt o" 0

Sawyer 9.4 dia

1970

iO x 106 < Re

< 30 x 106

1970

Paros (Kistler)

Used in a wide range of conditions including flight. Cooling system available

Pivoted about crossed-spring IOO mg flexure. Differential capaclty.to Permanent magnet plus coil IO g Parallel linkage LVDT 200 mg

Miller

1971

Low-speed flow.

Favourable pressure gradient

25 dia

Franklin

1973

Subsonic wind tunnel and water channel

16 dia

Pivoted. Variable geometry electronic valve. 50.1 x 3.2 Jewelled pivots. Clock springs LVDT 12 x 12 Parallel linkage LVDT 127 dia Air bearings LVDT

i g

Morsy

1974

Low speed flow past circular cylinder

130 mg

van Kuren

1974

High-temperature hypersonic flows with heat transfer. Floating element water cooled

5 g

Ozarapoglu

1975

Low-speed.

Adverse pressure gradients

I g

10
Dashpot

K.G.

Winter

"~

_ _

Atrflow ~ /Floating element 25rimdia /

,./ -t~

Cover

x///////////j[~L/7~ voT

Fig. 3.

Parallel-linkage balance (D.R.L.). In

Transformer (LVDT) which is capable of a resolution of displacement as small as 0.05 ~m. many balances the force is determined from displacement of the element against a spring as indicated by the signal from the LVDT.

There are disadvantages in the use of a displacement

balance since the necessary gaps round the element vary with the load and this variabion may produce spurious effects. In balances of the nulling type the position transducer is used to

provide the signal to a force system which maintains the floating element at a given position The force system is usually either a Kelvin current balance or a coil and permanent magnet. As an alternative to separate position and force transducers the Kelvin current balance may be arranged to serve both purposes as shown by Franklin (1960).

Airflov _ ~ ~Floatingelement 9ram dla/~ H ~ ; ~


~ V , ~ I

::;:~t:# ~
I I~tL~_.~__

H'clt Sh''ld
I ,-.t.,-,, .i.j

""'..'~TI~/AIlK//~I"~///~Fo~.

p,",','""" I :

zllv//y/ z, IN

]--,,,to"o,-,,,

b,,oo,,

(;tossed -sprinq pivot


Fig. 4. Pivoted balance (Kistler). The difficulty of measuring small forces in a displacement-type balance was overcome by MacArthur (1963) by supporting his floating element on a piezo-electric crystal and Winter and Gaudet (1970) were able to use resistance strain gauges in a large balance. Franklin Moore

(1973) has obtained high sensitivity by using a variable geomatry electronic valve.

and McVey (1966) have investigated a wide variety of position and force transducers for application at high skin temperatures.

Skin-friction balances have been used in flight on rockets (Fenter and Lyons, 1957) and on high-speed aircraft (Garringer and Saltzman, 1967; Fisher and Saltzmann, 1973). To eliminate

their sensitivity to linear accelerations of the flight vehicle it is necessary to arrange a mass balance for the measuring system as is done in the Kistler gauge (Paros, 1970). Weilet

(~954) and Lyons (1957) produced designs which were also insensitive to rotational accelera-

Skin friction in turbulent boundary layers

I!

Position adjuster

Airflov

~ Floating 127 mm d i a / e l e m c n t

Force transdu (spring-mounted

ldjusti.g srevr for floating .lem,t

Fig. 5. tions.

Air-bearing balance (Ozarapoglu).

In wind tunnels measurements have been made with heat transfer (McDill, 1962; Young,

1965; Young and Westkaemper, 1965; Bruno et a~., 1969) and roughness (Gaudet and Winter, 1973; Young, 1965) and also with surface transpiration (Dershin et a~., 1966). Westkaemper (1963)

investigated the effect of a mismatch of the surface temperature of the floating element with that of the surface in which it was mounted, and in his particular conditions found no correlation between the force measured and the temperature difference in contrast to the large effect found on heat transfer.

All balances are subject to the effects of misalignment of the face of the element with the surface in which it is mounted and these effects will be different for parallel-linkage or pivoted systems. Fig. 6 illustrates the way the pressure forces due to misalign~ent will act.

For an element which protrudes the pressure rise and pressure drop produced by the forwardfacing step and rearward-facing step respectively, acting on the edges of the element will result in a moment or a force increasing the balance reading. For a moment-measuring balance

there will be an additional increment produced by the suction in the local separation which is likely over the upstream surface of the element. For a recessed element the effects of

the pressure changes may still be large since, in addition to the forces acting on the edges, which will affect the readings of a parallel-linkage balance, separations will occur both at the upstream and downstream edges of the element and will produce pressure forces acting on the face of the element and contributing to the moment on a pivoted balance. That these re-

gions of influence may be of considerable extent may be judged from the measurements of Good and Joubert (1968) who showed that for a fence of small height h in low speed flow the pressure

12

K.G.

Winter

upstream was influenced

for a distance of up to I00 h. in flows with pressure

A moment-measuring

balance

is also

clearly at a disadvantage

gradients.

///Flff~fffFfFW

Parallel-linkage balance

Pivoted balance

Normal force

Protruding floating element

f
NormQI fore*

P-II
Recessed floating element

Fig. 6.

Effects of misalignment

of skin-friction

balance.

Systematic

investigations

of the effects of misalignmant (1965) and by Allen (1976).

have been made by O'Donnell

(1964),

O'Donnell and Westkaemper the parallel-linkage experiences

The balance used by O'Donnell was of

type and so the errors incurred should derive only from the axial forces It should be possible by making use of available data on the drag Gaudet and Winter (1973) showed that the that the pres-

by the element.

of excrescences

to make an estimate of these forces.

drag of shallow holes or of very short cylinders could be estimated by assuming sure variation on the vertical faces was that of forward or rearward-facing

steps multiplied

by the square of the cosine of the local angle of sweep of the edge of the hole, and assuming also that the skin friction on the flat surface was unchanged. ratio of the axial force on the vertical For a protruding cylinder the

faces to the skin friction on the flat surface be-

comes
AF 4 h ~- = --? (Pf + Pr ) 3~ where h is the height, r the radius and Pf and P on forward and rearward-facing O'Donnell steps. (3-1)

r are

the ratios Ap/T w for the pressure rise

A comparison has been made with the example given by

for a balance disc of 25 mm dlamter with an edge thickness of 0.25 mm at M = 2.67 O'Donnell does not quote a length scale for his boundary layer, and so in or recessed

and Re e = 10,O70.

order to deduce a roughness Reynolds number u h/~ for the height of his protruding

Skin friction in turbulent boundary layers

13

balance a value for boundary layer thickness has been taken from Stalmach's (1958) results for M = 2.73. The values of u h/v are below the range covered by Gaudet and Winter and a plausible Now equation (3-1) gives only the force on

extrapolation has been used to obtain Pf and Pr"

the protruding faces and for a floating element surrounded by a gap there will also be forces on the edges within the gap. It is usual to assume that pressure forces effectively act over Adding half the thickness of the submerged

half the thickness of the edge of the element.

edge (0.125 mm) in fact gives an overestimate of the force and an effective depth of penetration of the pressure of 0.02 nnn has been taken. For a recessed balance the only forces will The compari-

be within the gap and the same value of the effective penetration has been taken.

son given in Fig. 7 shows that though this estimate does not give precise agreement with O'Donnell's measurements their general character is well reproduced, suggesting that the physical model assumed is correct.

30
/ / I /

I 2O Estlmste Error IS I0 $
, 1 i

(Ah: O~)/)~//p O' Z


Moc~suremen.

/III/
t

-o.o

-o.o,

0.02 0"04 ~ 0 06 0"08 Protrusion a ~ I

Fig. 7.

Error in skin-friction measurement due to misalignment of floating element-parallel-linkage balance. O'Donnell: M f i 2.67 f Re 8 = 10070.

The balance investigated by Allen has been described by Fowke (1969) and is of the pivoted type. Some of the measurements made by Allen at M - 2.19 for his floating element, both reThese measurements confirm the flow model

cessed and protruding, have been plotted in Fig. 8.

sketched in Fig. 6 in showing that for his geometry the effect of the moment due to the pressures on the face of the recessed element is to cancel the moment of the thrust on the edges and produce an apparent, and very large, increase in the indicated friction force so that the indicated increment in force is positive for a recessed balance as well as for one which protrudes. Allen also varied the size of the gap, g, round his floating element, and as can be

seen in Fig. 8, increasing the size of the gap reduces the effects of the balance misalignment. An attempt has again been made to estimate the effects. The pressures on the vertical faces

are estimatedfrom the drag of forward and rearward-facing steps as was done for O'Donnell's results. For the protruding balance the pressure forces on the flat face of the element have

been ignored. the i n v i s c i d

For t h e r e c e s s e d b a l a n c e t h e Simple e x p e d i e n t h a s b e e n a d o p t e d o f d e t e r m i n i n g flow d e f l e c t i o n s a p p r o p r i a t e to the p r e s s u r e c o e f f i c i e n t s for two-dimensional taking

s t e p s , which can be done e a s i l y b e c a u s e t h e f l o w i s s u p e r s o m e , and h e n c e f i n d i n g t h e a r e a o f t h e f l a t f a c e on which t h e s e p r e s s u r e s a c t . As i n t h e ' c a s e o f O ' D o n n e l l ' s r e s u l t s ,

14

K.G.

Winter

the forces only on the protruding parts of the faces underestimates depth of the edge on which the pressures measured loads for protruding balances.

the loads.

The effective the in

act has been obtained by matching approximately It was found that for g/r = 0.002 the increment The dotted lines in

depth was approximately

Ah/r = 0.04 and for g/r = 0.02, Ah/r = 0.O1.

Fig. 8 show estimates made with these empirical values of Ah/r. matches the measurements

For g/r = 0.002 the estimate

remarkably well over the whole range of h/r, but for g/r = 0.02 the badly. The implication of this result is This re-

effects for the recessed balance are overestimated that the increase duction is perhaps

in gap size also reduces the pressure not unexpected

forces on the balance face.

since as the size of the gap is increased so will the air the pressure variations.

circulate more freely round the gap and ameliorate

!
r

4.0

\ \\\\

(.+=o.o+) "\-

Estlmote

/ //"

O-OOZ
0-012

",,
\,

3-0

\?\

,
-.

+.o

J / o o2o / /'.d, / /' 47:i // ,,), .G-y (++.+0:o,)


./
I ']'+

/ / / / /

Estimot

-0-020

-O-OIS

-~010

-0"0~

~ r

&O~

0-010

0-015

0-020

Fig. 8.

Error in skin-friction measurement due to misalignment floating element--pivoted balance--Allen: M = 2.19.

of

The results of Allen illustrate

the advantages

of having a large gap round the element from On the other hand a large gap may (1953) made a brief investigation

the point of view of reducing the effect of misalignment. be expected to disturb the flow over the surface. Dhawan

of the effect of a gap at low speed and could measure no change in the velocity profiles on a flat plate downstream of a groove in the plate. However, Gaudet and Winter His gap had a Reynolds number u g/v ~ 60. of the drag of excrescences, IO. found a If a high

(1973), in an investigation

measurable

drag increment for grooves normal to the flow when uTg/v exceeded

standard of accuracy is aimed for it is suggested

that u g/v should not exceed iOO, when acto the flow will be about

cording to Gaudet and Winter the shearing stress across a gap normal three times the undisturbed value of skin friction.

Ozarapoglu

(1973) studied the effect of changing the geometry of the gap when the pressure

in

the balance casing was varied.

In zero pressure gradient he found only a smell effect on the that small posiline

balance reading for casing pressures exceeding ambient pressure and suggested tive pressures

should be used to avoid the large errors which can arise if a stagnation

occurs on the upstream edge of the floating element because of a flow into the balance. Moulic (1963) in a study of the strong interaction region near the leading edge of a flat effect of casing pres-

plate in a low-density

flow at a Mach number of 6 found a significant

sure on his balance reading.

He used a very small element of 0.25 25 rmn, and the effect he cross-sectional shape of the element.

measured was probably due to the asymmetrical

Skin friction in turbulent boundary layers

15

Brown and Joubert (1961) studied the secondary forces arising in adverse pressure gradients with a view to applying their balance in three-dimensional flows. grounds that the force F indicated by a balance should be given by They argued on dimensional

= ~w

, ~

(3-2)

where T is the undisturbed shearing stress; w S the floating element area; D diameter; kinematic pressure gradient i dp. p dx' variation of gap with displacement. A variable guT/~ or g/D

This expression (3-2) is for a given instrument with a given gap.

should be included in general, and the results will also depend upon the thickness of the floating element. They compared the shearing stress indicated'by their balance with that deThough their instrument was not of the self-balancing type they

duced from Preston tubes.

were able by tilting it to take all their readings with the floating element centred, so eliminating the variable X. able pressure gradients. Brown and Joubert. Fig. 9. Miller (1973) later extended the work of Brown and Joubert to favour He ensured compatibility by choosing a value of g/D close to that of

The combined results for the influence of secondary forces are shown in

This figure shows that for the particular balance configurations used the secondary The pressure gradients had values of u~/u 3 T A feature of the results is that even at zero pressure gradient the balance readIt might be surmised that this variation is really dependent upon

forces do not exceed 15% of the friction force. up to 0.03.

ing depended upon Du /u.

gu.r/~.
F
IResuKs of Miller 1'20

I
1.15 I.iO

= Results of 8torn I Joubert I$00

~
~ ~

00
-zo -is
,

,.o5 f
-IO\
,\

,ooo
5
, ,

~-/]v-o

,,1

IO.

I'~ +CD 20

~s

30

o.J
Fig. 9. Effects of pressure gradient on reading of skln-frlctlon balance in low-speed flow. Everett (1958) calibrated a skin-friction balance with four different floating elements in channel flow with three different values of favourable pressure gradient obtained by changing the height of the channel. ing from a pressure gradient He found that the commonly accepted expression for the force aris-

Fp

S~ ~

td

(3-3)

which assumes that the external pressure gradient decays linearly over the thickness.of the

]6

K.G.

Winter

edge of an element, was satisfactory underestimated the force for s m i l e r

for a large edge thickness, values of t/D.

t/D = 0.02, but progressively

Ozarapoglu

(1973)

(and Vinh, 1973) studied the errors in the reading of their large balance at gradient where values of ~ / u 3 were 0.02 and 0.07. 90 and 170, were conthe pressure

two positions

in a strong adverse pressure

Because of the large size of the balance siderably

(127 ~mn) their values of ~D/u$,

in excess of those reached by Brown and Joubert.

As a consequence,

changes across the surface of the balance in a streamwise

direction were very large, and they a large

were able to show, by traverses with a Preston tube over the surface of the balance, reduction

in skin friction over the forward part of the balance when the casing pressure was As the casing pressure was the defect in if the

set equal to the mean static pressure over the floating element. reduced,

reducing the flow out of the forward part of the gap round the element,

skin friction was also reduced. gap was increased.

They showed that the errors in skin friction increased

The choice of the size of the gap round an element will clearly he a compromise between conflicting requirements. If flows in zero pressure gradient only are to be studied a wide gap of the balance reading to misalignment of the floating element level of the readings will uncerthe

will reduce the sensitivity

because of the reduction of the pressure changes, but the absolute

be uncertain because of the change in shear stress across the gap, and the resultant tainty about the effective surface area of the element. are desirable

For flows in pressure gradients

minimum possible gap and edge thickness

to reduce the flow through the gaps and

its effects on the pressure at the edges.

However,

even if the greatest care is taken in the design and use of a skin-friction to establish confidence the results in the accuracy of the results from it.

balance

it is difficult Gaudet

Mabey and the All

(1975) compared

from five different

Kistler balances,

and also compared

results from one of these with the large balance used in the tests of Winter and Gaudet. the tests were made in supersonic flows with nominally zero pressure gradient. deflection.

The Kistler

balances were used generally at ranges up to 25% of full-scale two of the balances were discarded 1% accuracy on skin friction.

The results from

because of lack of repeatability

and two agreed to within

A comparison between the results from one of these two and the The Kistler balance reading showed consider-

balance of Winter and Gaudet is shown in Fig. i0.

2.2

2"1 2"0 1"9

inter t Gnudet

,o'o, I"7
1.6 I'$ 1.4

I I

'~ 104
Fig. iO.

I 2

I Re0

, . i . i 8 I0 5

Comparison between results from Kistler balance and large balance of Winter and Gaudet.

Skin friction in turbulent boundary layers

17

able fluctuation and the mean reading varied by as much as 4% from that of the large balance. Mabey and Gaudet tentatively ascribe the variation with Reynolds number of the difference in reading to changes in the rate of variation of the tunnel temperature with time as the tunnel total pressure was changed. The fifth balance was of a different design, with a vent hole

provided downstream of the floating element and coulnunicating with the balance chamber, to reduce the risk of damage from the pressure transients acting across the floating element which occur in the starting and stopping processes of a supersonic tunnel. Fig. II compares the re-

suits of this balance with one of the two balances found previously to perform acceptably for a range o f conditions in the RAE 3 ft x 4ft tunnel, and for one condition in the NOL BoundaryLayer Channel. The results show that the difference in reading between the two balances is

reduced by sealing the vent hole, and the method of plotting (against a Reynolds number based on wall conditions) implies that the difference in reading may be correlated with the differences in surface geometry between the two elements.
O.Z VENT HOLES 14 OPEN CLOSED
/. ~ Q ~.s I 9 o

Acf.
ef
0"I

4.0 4.5 4.7 20

:l

RA[

NOL

. z

^ **

+~++~++"--+,
e

,o

e e y~X ~

uv Im

-0.1
Fig. 11.

Comparison of two Kistler balances showing effect of vent hole.

4.

VELOCITY PROFILE

It has been accepted for many years that the velocity profiles of turbulent boundary layers, at least in moderate pressure gradients, have inner layers for which the velocity scale is the friction velocity u . Clauser (1954) was the first to point out that the resultant similarity T of the viscous sub-layer, adjacent to the wall, a blending region and a logarithmic region. be derived from measurements of the velocity profile. The inner part of the profile consists

f the viscous sub-layer, adjacent to the wall, a blending region and a logarithmic region. With the assumption of constant shearing stress the velocity distribution in the viscous sublayer is given by _u_u = YuT-. u T (4-I)

and the value of the friction velocity, and hence the skln-friction coefficient can, in principl~be readily determined. However, in most aeronautical applications the laminar sublayer Instead the logarithmic or

is too thin for its velocity profile to be determined accurately. "law-of-the-wall" region is used.

In its original form the Clauser chart for determining skin The velocity profile

friction for incompressible flow was constructed in the following way. is

u u
whence

= A log -yuT B T

(4-2)

uU uU T

A l o g YUv +

A log

-~-

+ B

(4-3)

18

K.G.

Winter

so that a series of lines expressing u/U as a function of yU/~ may be drawn with u U = (cf/2) 1 as a parameter. The chart in the form suggested by Clauser is given as Fig. 12. In use the

chart is superposed on a measured velocity profile and the value of the skin-friction coefficient obtained by matching. values adopted for A and B. The value of the skin friction is of course dependent upon the Clauser chose the values A = 5.6 and B = 4.9. The values adopted 1968),

by the organizers of the Stanford Conference on Turbulent Boundary Layers (Khine et a~.,

as giving the best fit for the data available, were A = 5.62 and B = 5.0, and these values have been used in constructing Fig. 12. Other values have, however, been obtained ranging

from those of Ludwieg and Tillmann (1949) A = 5.0, B = 6.5, to those of Winter and Gaudet (1970), A + 6.05, B = 4.05.

I-0

0.9
0.8

0.7 0'6
tJ

O 0.5 0.4 0..3


Q

0.2 0.1 0
I0

4 5678

I0 z

$67B 4

103

~' yU3 ,4 5 67~ 104

.~ 4 5678105

Fig. 12.

Clauser chart.

The velocity profile obtained from the usual method of using pitot tubes will contain errors arising from the apparent displacement of a pitot tube in a shear flow, from the proximity of the probe to the wall, from the effects of turbulence and from the effects of viscosity. De-

spite the many investigations into these effects, of which a good discussion is given by Chue (1975), they are still not completely understood and their consideration is beyond the scope of this paper. Suffice it to say that generally the experimental, points, if uncorrected, will

fall above the logarithmic line near the wall, but the errors will be sufficiently small further away from the wall for a logarithmic region to be apparent.

Alternative and more convenient forms of the Clauser chart have been suggested; amongst them are those of Bradshaw (1959) and Ozarapoglu (1973), Bradshaw suggests that one suitable reference point on the inner-law curve be taken and gives as an example the point yuT/~ = i00 for which u / u ~ 16 (For the range of values of A and B quoted above u/uTat yu /~ = i00 lies beBy taking a range of values of U / u (that is effectively el) a curve The value

tween 16.1 and 16.5).

of u/U against y can be drawn for the particular Reynolds number of an experiment.

of u/U at the intersection of this curve with the measured velocity profile leads to cf since

c
V

ref

Skin friction in turbulent boundary layers

19

The example shown in Fig. 13 is in fact plotted in axes u/U vs. yU/~.

The experimental points

are for a high Reynolds number and so a reference value of yu /v ffi iOOO has been taken for which (with A ffi 5.62, B = 5.0) (u/UT)re f = 21.86. A curve of cf is also included. As can be

seen the method has the advantage that the reference line crosses the curve of the velocity profile nearly at right angles.
0.70

3.0

",
0.65 u U 0.60

~ I

I.~.-

"-'~o
2.5

Io%,
2.0

/
0.55

'i--t ....................
1.5

I.O

o.~o~ 0
Fig. 13.

I
3 yuu

I
4 x I0 -4

~-J

.~ 7
method.

Skin-friction from velocity profile:

Bradshaw's

Ozarapoglu obtains a direct relationship between u / u


as

and yu/v by expressing the log-law (4-2)

u u
T

+ A log ~ u
T

A log y__~u+ B. ~)

(4-4)

Hence for each point on a velocity profile a value of u/u T can be obtained from the value of yu/v using either a plot such as Fig. 14 or an interpolation process. has been made by Rajaratnam and Froelich (1967). A similar suggestion

19-U 16~
Wi7

,6-/ @
-

14

--

1312 "/I
i0 2 2

I
4 6

I
10 3

I
2

I
yu
v 4 6 i0 4

IJ
4

i0 ~

Fig, 14.

Skin-friction from velocity profile:

Ozarapoglu's method.

A more c o m p l i c a t e d p r o c e d u r e h a s a l s o b e e n p r o p o s e d by D i c k i n s o n (1965) i n v h i c h a v e l o c i t y profile is locally fitted yu ~ t o a p o v e ~ - l a w i n t h e form 1 nl

u E'- = F(n)
T

(4-5)

where n i s o b t a i n e d from n =

d log u "

20

K . G . Winter

For flows in pressure gradients the extent of the logarithmic region is reduced, but unless the flow is either in a strong adverse pressure gradient and close to separation, or in a strong favourable gradient and close to relaminarization, a logarithmic region can usually be identified. Since the velocity profile will be available, if it is to be used to determine

skin friction, it should be self-evident whether the determination for any particular flow is satisfactory. The definition of limits is more important when the Preston-tube method is used However, attention should be drawn to the

and will be discussed further under that heading.

use by Coles (see Kline et al., 1968) of a complete outer velocity profile in the form

u u ~-- = A log y v % where w = 2 sin 2 ~ ~ .

~ Y + B +--~ w--~

(4-6)

In the survey lecture for the Stanford Conference he pointed out that equation (4-6) could be used to obtain the skin-friction coefficient by finding values of u deviation of the data from (4-6) is minimized. and 6 such that the RMS

The strength of the wake, 7, the third paraThe whole of the profile

meter in (4-6), is eliminated by evaluating the equation at y = 6.

cannot be used since the viscous sub-layer is not represented in (4-6) and also a discontinuity in slope at the edge of the boundary layer is indicated. Galbraith and Head (1975) show that

the complete velocity profile may be used if this is represented by Thompson's (1965) profile family, This family includes a representation of the sub-layer and blending region, and by

using the concept of intermittency for the wake region avoids the discontinuity at the edge of the boundary layer.

The discussion so far has been confined to incompressible flows but the method can also be used for compressible turbulent boundary layers, though with somewhat less confidence. Allen

(1968), in connection with his investigations into the use of Preston tubes in compressible flow, also examined some of the forms proposed for the law of the wall describing the logarithmic region. He showed that consistent results could be obtained with the Baronti-Libby (1966)

transformation of the velocity profiles, which is complicated in that the determination of the transformed coordinate normal to the wall involves the running integration that equally good results could be obtained with the Fenter-Stalmach (1957) for adiabatic flow. ~Y o

; law ~

dy, but the wall

With the constants for the law of the wall used previously for incompres-

sible flow the Fenter-Stalmach form is given by

U u Tw

I 7

sin -I

= 5.62 log

~ + 5.0 ~w

(4-7)

where u Tw

yand o = 2

IM2 e
e

I + "~-~2 M 1 2

This method has been used by Mabey et el,

(1974) for Mach numbers between 2.5 and 4.5 in com-

paring the values of skin friction obtained from velocity profiles with measurements by skinfriction balance. Figure 15 shows a sample of their results in which cf was deduced from each The accuracy of the method can in part be inferred from the

point of the velocity profiles.

variation of the derived cf over the region of the profiles where the log law may be e~pected

Skin friction in turbulent boundary layers

21

to hold.

The variation of the skin friction in Fig. 15 may indicate a shortcoming in the transformation used.
2.4 2.3 2.2 2.1 20 ~3) 1.9
1.8 1.7 1.6 -o o o 0 0 0 000 0 0o 0

compressibility

o
oo

oo

Reo = 5 . 9 7 x 1 0 3

"c"P'm=~"~- n ~ o o

Cf b a l a n c e
o

oOCP~o
o

o Reo=17"Sxl03
oo~,-9--cf balance I I
2 3

1.5 _ ~ ~ v o v o ~ ~ I I I I

I
4

I
5 6

I
8 I0

I 20

0.3 0.4 0 . 5 0 . 6 0 . 8 I

ymm
(a) M = 2 . 5 1.7
1.6 o o o o o

1.5

o ^ o Re 0 =4.80x103 Cf b a l a n c e

%o

u~ 0

1.3 - od" 1.2 I.I 1.0 0.9 0.8 I I I i 0.3 0.40.50.6 0.8 I (b) I

0oO%

J _ ~o ol~ee=
o OOooo o ooOO I 2 o o o~"cf balance I I 4 S 6 I I 8 I0 3 ymm M-4.5

28.6x10

20

I I

< 60

Fig. 15.

Skin-friction from velocity profiles in supersonic flow: Mabey et al.

Other forms of the law of the wall have been used. form

For example Spence (1959) suggested the

A log u m

(nuTo
Vm r

+ B

(4-8)

with A = 5.76, B = 5.5 and C = 2.3, and where the density pm used in the definition ~ = P u 2 w mT m and the kinematic viscosity vm are evaluated at "intermediate enthalpy" conditions, for which Spence found Eckert's formula h m = O.5(h w + h e ) + 0.22(h - h ) e (4-9)

to give a satisfactory evaluation.

The extra constant C enables a continuous profile to be

continued into the laminar sub-layer.

A particular simple form was shown by Winter and Gaudet

(1970) to give an accurate representaTheir

tion of their velocity profiles in adiabatic flow at Mach numbers between 0.2 and 2.8. expression is u yu --~ = 6.05 log
1 v u T e

i + 4.05 (4-10)

22

K.C.

Winter

\PeJ

0.2M

For flows with heat transfer, relatively few velocity and temperature profiles have been measured. Hopkins et a~. (1972), in one of a series of papers on skin friction at hypersonic

speeds, compared their measured velocity profiles on non-adiabatic flat plates with various forms of the law of the wall. was the most satisfactory. They concluded that the transformation due to van Driest (1951)

In this transformation the law of the wall is represented as

yu
I U a u w Ywhere a2 2 T w T e in-I 2a 2 ~ - b (b 2 + 4a2)
u 1M2 T

h + sin-I (b~ +
= A

log--

w
v w

+ B

(4-11)

4a 2) 1 1

e '

b = a~ --e ~ +T - I" w

For isoenergetic flow, with a recovery factor of unity so that T /T : I + ~(~ - I)M~ , equaw e tion (4-11) reduces to the expression (4-7) given by Fenter and Stalmach (1957) for adiabatic flow.

It should be noted that all the expressions for the law of the wall given above are for smooth, flat surfaces. For surfaces where the boundary-layer thickness is not small compared with the

radius of curvature of the surface the form of the law will change, and for rough surfaces the constant B'will have a reduced value.

Surface injection or suction will also change the form of the law. flows with transpiration the so-called bilog law may be used

It has been shown that for

2uT vw

I<

- i

= A log

B.

(4-12)

Stevenson (1963) showed by comparison with experiment that A and B are independent of vw and u . Jeromin (1968) showed that it is possible to find transformation parameters such that T velocity profiles in compressible flows with injection can be transformed into Stevenson's law of the wall, but the derivation of the parameters is too complex to be used for the determination of skin friction. tained I 0
u

Squire (1969) applied mixing-length theory to the problem and ob-

p~du' (PwVw u' ~w )~ = A log

YUTw + B Vw

which, with the assumption of the quadratic temperature profile

__T =
Tw

1 +

-T T-T r________wu+ e
Tw U
w

(u) 2 r

Skin friction in turbulent boundary layers

23

becomes
U

~
-

d ~
---

=
+ e r

A log
W

+ B

(4-13)

Tw
Squire compared

Tw

(4-13) with experimental results and found that, although a logarithmic region
W

occurred with a constant value for A, the additive constant B varied both with M e and Vw/U T . Thus, though it may be possible to identify a logarithmic region and so to determine cf, a simple form of Clauser plot cannot be derived from (4-13).

Because of our current inadequate understanding of the shear stress distribution in threedimensional turbulent boundary layers the determination of the skin-friction coefficient from

velocity profiles in three dimensions is somewhat speculative.

East and Hoxey (1969) give a

good review of the possible forms of the law of the wall which might be used:

(a) file:

The velocity profile is treated as being simply a twisted form of a two-dimensional

pro-

u
U T

yu = A log-V

+ B

(4-14)

where u is the resultant velocity.

(b) The velocity profile is resolved in the direction of the external stream and the friction velocity r e s o l v e d i n t h e same d i r e c t i o n is obtained:
U u T X

YUz
= A log - -

(4-15)
+ B .

(c)

A "developed total velocity" is used in the form suggested by Perry and Joubert

(1965):

U T

A log y u

+ B

(4-16)

where s

= IO

1 + \d--u~ /

du'

in which the primes denote running variables, and w is the crossflow velocity.

(d)

As suggested by Johnston

(1960) an effective velocity in the direction of the wall shear-

ing stress is used. streamlines

If the direction of the wall shearing stress relative to the external

is BO, then at the wall, and also in the inner region of the boundary layer as (the Johnston triangular model)

shown by the linear region of the polar plot of velocity, + u sec BO. Therefore at the wall /yu~ =f
T

u sec Bo
U

and it is assumed that this similarity can be extended into the logarithmic region so that u sec u
T

13 0
=

yu T Al o g - - ~ + B.

(4-17)

24

K.G.

Winter

A fifth possibility (e) is also considered by Pierce and Zimmerman (1972), based on a suggestion by Coles (1956), in which the relevant velocity is taken to be that resolved along the wall shearing stress direction. cos (8 - 8) u = T This assumption leads to

yu A Iog--+ v

B.

(4-18)

Figure 16 shows two sample comparisons from East and Hoxey of Clauser plots derived from methods (a) to (d) (method (e) is omitted in Fig. 16b), together with the two-dimensional log lines based on skin friction obtained from Preston tubes and razor-balde surface pitot tubes. Also shown are the profiles of the cross-flow velocity w/U. The profiles were measured in the

boundary layer on a flat surface from which an obstacle protruded in the form of a circular cylinder with a fairing. Figs. 16a and b are respectively for the flow at points upstream and That is to say the crossflow

downstream of the inflexion point in the external streamlines. is respectively increasing and decreasing.

In the latter case the outer parts of the velocity On the basis of compari-

profiles bear little resemblance to forms expected in two dimensions.

sons like Fig. 16 for a large number of profiles East and Hoxey conclude that method (d) is the most satisfactory. This conclusion is in agreement with the outcome of the comparisons of

Pierce and Zimmerman in which method (d) is shown to give the largest apparent logarithmic region. In three dimensions the flow in the outer regions is strongly controlled by pressure

gradients and as Fig. 16 shows a simple interpretation like the two-dimensional law of the wake is not possible.

o.9

~ u u

x,,/u

~Iode \ /

v,,-

o.8 0.7

v usecBd u
+ ,,,u "

v
+x

b,o~,~

Razor I
0.8

vusecBo/u
0.7

/:~bx, x ,~

/Preston,,

',,

v v t

0.5

0.5

So~O" o ~ ~

0,3 0.1
0.2

o o
dA~A,%AAd A ,% ~

0.3 0.4
0.2 A,%~Ad

~L%d

"

,5
'%

02.5

I 3.0

I I 3.5 4.0 4.5 .5.0 tOqio , ;y U/v

i .5.5

6.0

I 02.5 3.0

I 5.5 4.0 tOqlo ,

4.5 U/v Y

5.0

I 5,5

6.0

Fig. 16.

Velocity profiles in three-dimensional flow:

East and Hoxey.

5.

THE PRESTON TUBE

A variant of utilizing a fit of measured velocity profiles to the law of the wall is to measure one velocity only at a known distance from the wall. Preston (1954) realised that this could

be achieved with a circular pitot tube lying on the surface, and demonstrated by measurements for a range of sizes of tube in fully established pipe flow that the unique relationship

Skin friction in turbulent boundary layers

25

pv 2

~ \ Ov 2,J

(5-1)

which was to be expected from considerations of wall similarity, was in fact obtained.

In (5-1)

Ap is the difference between the pitot pressure and wall static pressure; d v T
W

is the outside diameter of the pitot tube; is viscosity; is wall shearing stress.

Despite the close relationship between (5-1) and the law of the wall it is not to be expected that one can be derived simply from the other. The pitot tube reading will be affected by

the various sources of error mentioned previously and the static pressure, to which the pitot pressure is referred, will also be subject to errors as first quantified by Shaw (1960). T h e author is of the opinion that the calibration should be regarded as purely empirical within the framework indicated by the similarity scaling.

Following Preston's work, others, notably (Hsu (1955), the Staff of Aerodynamics Division, NPL (1958) and Walker (1959) checked his calibrations in various forms and doubt was cast on its universality. By comparing the readings of a Preston tube and a sub-layer fence in both

the developing flow in a pipe and the fully developed flow, Head and Rechenberg (1962) were able to show that similarity in the two situations existed. (Because of the small size of a

sub-layer fence (see Section 6.1) its reading should give a reliable measure of the skin friction in both situations) Using essentially the same appartus, Patel (1965) produced what is

currently regarded as a definitive calibration, covering a wide range of flow conditions and sizes of Preston tube. x* = His calibration is given in terms of and y* = log d2 w 40~ 2

log Apd2 40v 2

aS ~

y* < 1.5

y* = Ix* + 0.037 y* -- 0.8287 - 0.1381x* + 0.1437x .2 - 0"0060x'3 x* = y* + 2 log (1.95y* + 4.10) I (5-2)

1.5 < y* < 3.5 3.5 < y* < 5.3

These three ranges correspond roughly to the three regions in the velocity profile, the viscous sub-layer yUT/V < 6, the transition region 6 < yu /~ < 60 and the logarithmic region 60 < yu /v < 500.

As was pointed out subsequently by Head and Vasanta Ram (1971) the expressions (5-2) do not quite match at the changeover and also the expression for the outermost region is inconvenient to use because of its implicit form. Furthermore Twd2/4pv2 varies by more than four orders They therefore suggested the use of

of magnitude over the full range of Patel's calibration. two alternative forms of the calibration, the first

*The 4 in the denominator of the logarithms was originally adopted by Preston, since if the height of the centre-line of the pitot tube is taken as the relevant variable in the wall similarity parameter (y = d/2), ~d2w ~ I ~
4pv 2

26

K . G . Winter

Ap
- v s . -

Apd2
-

Tw

Pv 2

is tabulated in Table I, and the second is in effect a Clauser plot for a Preston tube. the calibration is expressed as

If

j
then
w ~p ~p

cf

~u 2

0u 2

Ud

~ou 2 2

u
This expression is shown as a chart in Fig. 17 where -~ U variables with cf as a parameter. = I d__~__) ~
PU2

a n d -dU are used as the


"o

o8[
i~KillIl~nRi3~isn~iKil
0.~ ~p

V
0.. ~

0.~

0.~

0.2

0.,

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Fig. 17.

Preston tube calibration chart:

Head and Vasanta Ram

Bradshaw and Unsworth (1973) give a further alternative, but implicit, expression,

du AP=Tw 96 + 60 log s--~+ 23.7


du valid over the range 50 < - - < u IOOO.

I log5-~-j; duT~

(5-3)

Patel also investigated the limiting pressure gradients, both favourable and adverse for which his calibration might be expected to apply. parameter His proposed limits are based on values of the

Skin friction in turbulent boundary layers

27

Table

ww

as a function of Apd2/pu2 (Head and Vasanta Ram).

M -pv x IO- 4.0 4.2 44 4.6 48 5.0 5*2 5.4 5.6 5.8 6.0 6.2 6.4 6.6 6.8 7.0 7.2 7.4 76 7.8 8.0 8.2 8.4 8.6 8.8 9.0 92 94 96 98 A& T_ pfi x IO- I.0 I.02 1.06 1.10 l-14 I.18 1.22 I.26

AP
TV 918 9.41 9.63 985 10.06 IO.27 IO.47 10.67 IQ87 llt% II.25 II.44 11.62 11.80 Il.98 It.16 12.33 1250 12-67 12.83 12.99 13.15 13.31 13.47 13.63 13.78 13.93 14.08 14.23 14.38 AP

AP@ -p9 x IO 1.30 I.34 1.38 I.42 1.46 I.50 1.54 1.58 1.62 1.66 1.70 1.74 I.78 1.82 1.86 1.90 I.95 2.00 2.05 2.10 2.15 2.20 2.25 2.30 2.40 2.50 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 I.4 1.5 1.6 1.7 I.8 I.9 1.0 1.2
1.4

AP
Is9 16;56 16.81 17.06 17.31 17.55 1779 18.02 IS.25 IS.48 18.71 18.94 1916 19.38 19.59 19.80 20.01 20.27 20.53 20.79 21.05 21.30 21.54 21.78 22.02 22.49 22.96 23.41 23.86 24.30 24.73 25.08 25.43 25-78 26.13 26.48 26.82 2716 27.50 27.83 28.15 28.46 2907 29.66

ApdJ -pd I( 10 4.6 4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2 6.4 6.6 6.8 7.0 7.2 7.4 7.6 7.8 8.0 8.5 9.0 9.5

AP Tu 30.23 30.79 31.33 31.84 32.34 32.84 33.31 33.78. 34.23 34.68 35-l 1 35.52 35.94 36.34 36.72 37.11 37.50 37.87 33.74 39.58 40.40

APT .p4 < 10-4 2.0 2.1 2.2 23 2.4 2.5 2.6 2.7 i-8 2.9 3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0 5.2 5.5 5.0 5.5 PO 7.5 6.0 5.5 a.0 P5

AP Tw 52.32 53.15 53.93 54.68 55.40 56.62 56.80 5745 58.07 58.68 59.28 60.40 61.46 62.47 63.43 64.34 65.20 66.01 66.80 67.57 68.32 6905 70.00 71.55 73.00 74.35 75.60 76.80 77.95 78.95 79.90

AP# mpg x10+ I -4 1.5 I.6 1.7 I.8 1.9 20 2.2 2.4 2.6 2#0 3.0 3.2 3.5 4-o 4.5 5.0, 5.5 6.0 6.5 7.0 7.5 8.0 9.0

AP 710 86.8 88.0 89.1 90.2 91.2 921 93.0 947 96.2 97.6 98.8 99.9 100.9 102.4 104.5 106.4 108.0 10s 3 1 IO.4 III.4 112.2 113.0 113.7 114.9

AP# jg- 90 60 7.0 8.0 90

AP x 1395 142.7 145.4 147.8 1500

AP@ -P9 I< 10-l


zrzir 1.2 I.4 1.6 1.8 2.0 2.2 2.5 3.0 3.5 4.0 4.5 5.0 6.0 7.0 8.0 9.0

AP TV

r.,
14.53 14.67 14.95 15.23 15.51 15.78 16.04 16.30

I.00 I.05 I.10 I.15 I.20 1.25 I *30 1.35 I.40 i-45 !*50 -55 -60 -65 -70 -8 *9

41.18 41.93 42.65 43.34 44.00 44.54 4527 45.87 46.45 47.01 47.56 48.09 48.61 49.12 49.62 50.56 51.46

155.3 158.1 160.6 1628 164.9 166.8 169.3 172.8 175.8 178.5 180.9 183.1 186.8 1900 182.8 195.3

IP~

AP

r-;; -6 ,*o *05 *IO -15 -20 -30 8@80 81.70 82.55 83.35 84.1 85.5

1.0 I.1 1.2 1.4 1.6 l-8 PO 1.2 1.5 I.0 j-5 I.0 4.5

116.0 117.1 118.1 119.8 121.5 123.1 124.6 126.0 127.8 1307 133.3 135.6 137.7

1.0 I.2 I.4 1.6 I.8 2.0 L5 1.0 3.5

197.5 201.4 204.7 207.5 210.1 2125 2174 221.6 225.0

28

K.G.

Winter

h p

de
Pu~ dx ' B

and are: Adverse pressure gradients:


du

Maximum error 3%

0 < A < 0.01 p

v
du

T ~ 200;

Maximum error 6%

0 < A < 0.015 ---~ < 250. P

Favourable pressure gradients:


du Maximum e r r o r 3% 0 > ~ p > -0.005 dA ~ < 200 d-~x < O;

Maximum error 6%

O > Ap > -0.007

du dA v ~ ~ 200 d~x < O.

As Patel points out these limitations are a rough guide only.


(1962) law o f t h e w a l l in pressure gradients is accepted,

He notes that if Townsend's

u
u T

_ f

YUT, &
~ T

where &

v ~T - -pu~ ~y

and that in general AT, which is equal to &p only at the wall, might therefore be expected to be the controlling parameter, course, not possible. l~s use a prior~ in a boundary layer investigation is, of
T

The limit on du /~ is perhaps also rather sweeping since the limiting

value might be expected to depend on the Reynolds number of the boundary layer, tending t@ increase as a boundary-layer Reynolds number increases. The limitation for favourable gradi-

ents of dAp/dX < 0 was introduced to ensure that a boundary layer, if it is in a condition where it may be subject to relaminarization, should only he approaching that state. The phy-

sical features of the flow which lead to the limitations have been discussed by Patel and Head (1968) and are illustrated in Figs. 18 and 19. Fig. 18 presents profiles from Newman's

(1951) experiments (from the Stanford tabulations) and shows how the velocity profiles depart from the law of the wall for zero pressure gradient as the adverse pressure gradient increase~ This departure is predicted by Townsend's velocity profile for a linear shear-stress gradient away from the wall. On the other hand the recent reassessment by Galbraith and Head (1975)

of eddy - viscosity profiles implies that the mixing length increases in strong adverse pressure gradients and that this delays the departure of velocity profiles from the conventional law of the wall.

The profiles as plotted in Fig. 18 differ from those in Fig. 4 of Patel and Head and add weight to the conclusion of Galbraith and Head. The difference is presumably due to a differ-

ent derivation of the skin-friction coefficient in the Stanford tabulations from that used by Patel and Head. It is interesting to note, on the basis of Fig. 18, that the logarithmic re-

gion is entirely absent from the curve for the largest pressure gradient but the profile appears to follow a blending curve of the form suggested by van Driest (1956) without allowance for the normal shear-stress gradient; hence Patel's calibration, for a sufficiently small Preston tube might be applicable.

Skin friction in turbulent boundary layers

29

I00

--

o o o o o o o

90--

80

o
-

o o

70

:r
+ + +

O o

u 60 Uz
50--

o
-

A~
.
x
+

o
o o
:

2
+ + . ~,

0.0075 0.0214

o 0.2980
40 --

0.0905

.:
/
+

o
0

30-0

0 4-

4-

20
.~X

o . rxo. o.~.-'.. x .
+ "4~'x i O

02

3 4~6

slO

3 ~,6

102

I IIII

3 4 56

8103

lJ

YUr
v

Fig. 18.

Velocity profiles in adverse pressure gradient:

Newman.

Figure 19 shows a series of profiles in a strongly favourable pressure gradient.

As the flow

progresses into the favourable gradine (x0 - x B increasing) the velocity profiles, whilst remaining turbulent, first fall below the line of the law of the wall and then lie above the line as the flow starts to revert to a laminar state. is evident at x 0 - xB = -2 in where Ap of -0.005.
=

The departure from the law of the wall

-0.01, in accord with Patel's limiting value for 5p

Brown and Joubert (1967) also investigated the limitations on the use of Preston tubes in adverse pressure gradients, basing the analysis of their experimental data on the dimensional analysis of Perry et a~. (1966). This analysis indicates that the first departure from the

law of the wall due to a pressure gradient will be in the form of a half-power region, that is

UT

where a = I dp pdx' and that the half-power region will start at a constant value of ay/u 2 = 1.41.
T

They therefore

analysed their results for a range of Preston tubes in terms of ud/u 2 and obtained the following table of errors: Preston tube error %
- -

I 1.35

2 1.74 2.06

~5 2.55

7 2.98

ad u2 T

30

K.O.

Winter

22in. en'try A 30 cf

124in.entry 0 U cf -A 0.0178

(f~/s)
20

(ft/s)

8G.18 0 . 0 0 4 9 2 87.32 0 . 0 0 4 6 7

I0

77.93 0 . 0 0 5 3 7 7 8 . 6 5 0.00521

0.0212

68.91 0 . 0 0 5 7 8

68.32 0.00578

0.0244

0 U Ur

59.65 0.00690

59.72 0 . 0 0 5 9 7

0.0288

53.98 0 . 0 0 5 2 7

54.24 0.00516

0.0275

50.02 0 . 0 0 4 4 9 52.97 0.00:387

0.0204

48.650.00386

52.41 0.00318

0.0103

47.74 0.00351

51.87 0 . 0 0 3 0 5 5 0 . 0 0 3 8

I0

I0 ~ yur
v

10 3

104

Fig. 19.

Velocity profiles in accelerating flow:

Patel and Head. They

which are broadly in agreement with their proposed limit of the log law as ~y/u 2 = 1.41.

also suggest that, since the log law region will vanish when the outer and inner limits become the same, that is when y = 30v/u = 1.41 u~/u or av/u 3 = A 0.05, Preston tubes should not T T p be used for stronger pressure gradients than this. Since, however, Patel's calibration extends to the viscous sub-layer this restriction may not be necessary for very small tubes. based on the results of Brown and Joubert are shown in Fig. 20 in terms of du /v vs. A
T

Criteria In

terms of this mapping Patel's criteria for 3% accuracy are in good agreement at their intersection but are conservative as general criteria, but his criteria for 6% accuracy overestimate the permissible limits. Ozarapoglu measured velocity profiles in strong adverse pressure gradHis critical values are also P than accepted previously. P This is confirmed by recent results of Chu and Young (1975).

ients and determined the outer limit of the logarithmic region.

shown on Fig. 20 and are roughly in agreement with Patel's 6% criterion for low values of A but indicate that Preston tubes may be used up to larger values of d

Patel's calibration expressions have an upper limit of y* = 5.3 which corresponds to a value of duT/9 of approximately iOOO. This should be regarded as an upper limit for which the cali-

bration was obtained in Patel's experiments and not an upper limit for the application of the technique, which will be the outer limit of the log-law region. This limit derived from the

experiments of Winter and Gaudet in zero pressure gradient is shown as a function of the Reynolds number based on momentum thickness in Fig. 21.

Skin friction in turbulent boundary layers

31

-0"12

r--O'lO
l

~
I

....~

Oza r s poglu

-0,08

I
,B-,,

Ap
-0"06

I I
~.,

SO

I00

uT._.dd
I

150

luu

~v~

Pig. 20.

Criteria for limits of Preston tube.

IO40 i 6

+I
2 i03
6 6 4 2 SO2 # 6 4
1

, ,,p**|

, i,,,i|

, ,,,,,|

1(3 102

6 8103

I1104

2 Rel)

6105 8

' ~ ' ~'~'lO 6

Pig. 21.

Outer limit of logarithmic region of flat plate.

Several ways have been proposed for transforming the pressure and friction parameters used in the callbration of a Preston tube to enable the calibration to be applied to compressible flow, The first of these was probably that due to Fenter and Stalmaeh (1957); this was derived from their compressible form of the law of the wall already referred to, that is

U ~ UT w

sin_ 1

= A log

, w

B.

(5-4)

32

K . G . Winter

Fenter and Stalmach assume that (5-4) may be applied in place of the incompressible form at y = d/2* where d is the Preston tube diameter, and recast (5-4) as

dU i w oT
or

( sin-i

i ) o~ u

r w .....
~w

2 A log + B

Ow ~e pI Rede sin-I I ~w o ~

<~ U

Re d

log

~ ~

~w

_~e Re d ~w

+ B

~ee

--~wRed(cf) ~

A log

--~wRed(of) ~

+7-7B

log 23/2

(5-5)

The Preston tube calibration is thus expressed as

~w Redc~
u

vs.

Pe ~w oT

where ~ is obtained from the usual expression for pitot tubes in compressible flow, together with the assumption of constant total temperature through the boundary layer. Sinalla (1965) obtained a simpler expression by evaluating the Preston tube parameters at the

"intermediate temperature" using Monoghan's (1955) expression for adiabatic flow T = I + 0.35r(y - I)M 2
e
o

From the data of Fenter and Stalmach he obtained the empirical equation

w
Omd2"r IJm 0.0290

/p2d2u2\O 873
m m (5-6)

in which the p r e s s u r e r i s e Ap used in i n c o m p r e s s i b l e flow i s r e p i a c e d by tOmU2 where u i s the v e l o c i t y i n d i c a t e d by the P r e s t o n tube.


Hopkins and Keener (1966) also used an intermediate temperature hypothesis but instead of replacing Ap by ~OmU2 they used Ap = ~ou2 = (y/2)PeM2. 0.883 OOmd2U2~ 0.0228 Their calibration expression is

O d2T m w ~2
m

(5-7)

Allen (1973) investigated the accuracy of these various calibrations for adiabatic flow using a wide range of sizes of Preston tubes over the Math number range 2 to 4.6, comparing the resuits from the Preston tubes with skin friction measured by a balance.** He also made use of *They also investigated the transformation to the calibration for incompressible flow that would be obtained if the average p~tot pressure over the opening of the Preston tube were taken as the relevant quantity. They conclude in accordance with Hsu's (1955) result for incompressible flow that the difference from simply taking values at d/2 could be neglected. **Alien has recently discovered that the balance, against which he calibrates his Preston tubes, read erroneously. His revised equation (5-9) is log F 2 = 0.01239 (log Pl) 2 + O.7814 log F I 0.4723.

Skin friction in turbulent boundary layers

33

the data of Fenter and Stalmach (M = 1.7 to 3.7), Hopkins and Keener (M = 2.4 to 3.4) for adiabatic flows, and of Keener and Hopkins 41969) (M = 7) for non-adiabatlc flows. In addi-

tion to the calibration laws listed above he also devised a calibration for compressible flow based on Patel's expression for incompressible flow for his highest Reynolds number range, by using the intermediate temperature hypothesis with Ap = PmU2 to obtain

m 2

2 2

1.95

log

4.1

(5-8)

The general conclusion of this experiment was that of these various calibrations those of Fenter and Stalmach, and of Pateli which are based on logarithmic profiles gave the best resuits over the widest range of parameters. Since both of these calibrations are implicit,

Allen proposed an interpolation formula which gave a slightly improved rms deviation from the
data.

This formula is log F 2 = 0.01659 (log FI) 2 + 0.7665 log F I - 0.4681


Pmd u

(5-9)

where F 1

~m

Pm }ae u Pe Vm Re d

F2 = 42Pm~ w)~

0(~/
,e,

IIe

The Sommer and Short (1955) value for the intermediate temperature is used, that is

= i + 0.035M 2 + 0.45 Te and Sutherland's formula is used for the dependence of Viscosity on temperature.

The last word written so far on Preston tube calibrations in compressible flow appears to be that of Bradshaw and Unsworth (1973, 1974). They express justifiable doubts about placing re-

liance either on the assumption that the pressure read by a Preston tube may be taken as that registered by a small pitot tube placed at the position of its centre, or on the concept of calculating fluid properties at an intermediate temperature. bration law based entirely on dimensional grounds as The propose an empirical cali-

for compressible adiabatic flow.

In 45-10) fi represents a calibration for incompressible is the speed of sound at wall conditions). Equa-

flow and fc a compressibility correction 4 ~

tion 45-10) is, of course, an implicit form of the calibration but is used because the functional relationships are clear in this form so that subsequent ~ e n d m e n t on the basis of new e~eri~ntal data can easily be ~ d e . As noted earlier Bradshaw and Unsworth devised a fit to as 2 = 96 + 60 log ~ Tw and expressed the compressibility correction based on Allen's results as + 23.7 log ~ (5-3)

Patel's results over the range 50 < du /~ < I ~ T

34

K.G.

Winter

f
u T

= IO~M~

(5-11)

where M = Y

a w ~v
TW/ W

They tentatively claim an accuracy of about 2% at low speed and small du about 10% for du w/~ w up to i000 and M up to 0.I.

decreasing to

It is the author's view that the table of

Head and Vasanta Ram provides the most satisfactory calibration for incompressible flow; in particular it extends to lower Reynolds numbers than the formula of Bradshaw and Unsworth. For compressible flow the use of the Fenter-Stalmach functions appear to result in somewhat less scatter than the Bradshaw-Unsworth formula, though it is admitted that only a small sample of data has been examined.

One aspect of the calibration of Preston tubes which has not been adequately explained is its sensitivity to the Reynolds number of the static pressure hole to which the Preston-tube pressure is referred.

Equation (5-10) is for adiabatic flow.

For flow with heat transfer a further parameter There David-

8q = q/0wCpTwU w where q is the heat flow per unit area, will enter into the equation. is a need for further systematic experiments to determine the effect of heat transfer.

son (1961) attempted this for M = 5 but found inconsistent results; Holmes and Luxton (1967) in an experiment at low speed and Yanta et a~. (1969) at M = 4.8 found that their results were

best correlated by use of the intermediate temperature concept.

The measurements at supersonic speeds of Yanta et al. for favourable pressure gradients and of Naleid (1958) and Hill (1963) for adverse pressure gradients give satisfactory results but there do not appear to be any general criteria for the use of Preston tubes in pressure gradients in compressible flow.

The Preston tube has also been used in flows with transpiration.

Stevenson (1964) developed

a calibration equation using a power law approximation to the velocity profile (4-12) to find a mean dynamic pressure over the opening of the tube, as had been done by Hsu for impermeable surfaces. If the law of the wall is taken locally as

T vw

flu) I
+ - I (Apd2~ ~

= C

then an approximate expression for a Preston tube calibration can be derived as = C 2 vl +( d u ~ l ~ 2 n + cIdUTl l+n

(5-12)

where C and n will depend upon the Reynolds number of the boundary layer but should have the same values as for an impermeable wall. Typical values are C = 8.4, n = 1/7. Stevenson found

fair agreement between the skin friction deduced from Preston tube readings by means of (5-12) and that from the law of the wall and from momentum traverses for the boundary layer growing on a porous cylinder.

Skin friction in turbulent boundary layers

35

Simpson and Whitten (1968) suggest an alternative form of a Preston tube calibration for a transpired boundary layer in which in addition to the usual shearing-stress and pressure-difference parameters they introduce a third parameter Vw(0w/Ap)l to take account of the transpiration, but they do not give an explicit result.

For three-dimensional flows the Preston tube has been used, for example by East and Hoxey, see Fig. 16. They set the tubes in the flow direction obtained from surveys of the boundary

layer by a yawmeter - in fact in the flow direction at a height ~ times the diameter of the Preston tube. Their results, evaluated with Patel's calibration, appear to be satisfactory.

In two experiments on the flow past obstacles on a flat plate, one a circular cylinder and the other an inclined fence, Prahlad (1968) showed that in the powerful favourable pressure gradients which may develop in a three-dimensional flow the Preston tube should only be used with caution. He used a vectorial pressure grad p and resolved this in the wall shear-stress as P high as 0.04 were 0tbained, considerably in excess of Patel's criterion.

direction to show that at the points where the Preston tube failed, negative values of A

Prahlad (1972) also studied the yaw characteristics of Preston tubes, of surface tubes with openings cut off obliquely at 4 ~ and of double tubes in the form of a 90 Conrad probe and R a j a r a t n a m a n d Muralidhar

gives detailed charts of their characteristics in low-speed flow.

(1968) present calibration curves for a three-tube yawmeter used as a Preston tube.

Other variants of the Preston tube have been suggested, notably that of Rao et al. (1970) who used two tubes of different diameter, though the tubes were not mounted in contact with the wall. They argued that if it is assumed, as has been done so often, that the pitot tubes read

a mean dynamic pressure over their faces of area A], A 2 the difference between the reading of two tubes in a two-dimensional flow is

Ap

~i AI

(log y)2dA] - ~ 2 A2

(log y)2dA2

0u~BA + 0u~A 2 log

i At

log y dA 1

I - ~2

log y dA 2

(5-13)

Thus if two tubes are chosen so that

A~I

if
A1

log y dA I

ffi A-~

if
A2

log y dA 2 ,

a calibration may be obtained which is dependent of B and also of a static pressure measurement. Rao et aZ. give an example for the flow on a flat plate in which the skin friction ob-

tained from the dual-pitot-tube calibration was within 5% of that obtained from either of the tubes separately. known a priori. The technique might be applied to rough wall flows in which B will be un-

An improvement to the dual pitot tube of Rao et al. has been explored by Gupta (1975) in which two probes of different diameters rest on the wall slde-by-side and t h e m o u t h of the smaller one is chamfered at 45 away from the larger one, as shown in Fig. 22. gation a t low s p e e d s Oupta showed t h a t d e v i c e ( a ) , w l t k a d i a m e t e r r a t i o In a limited investi-

o f 0 . 7 8 , gave a p r e s -

s u r e d i f f e r e n c e o f 71% o f t h a t f o r t h e l a r g e r t u b e u s e d

as

a Preston

t u b e , and d e v i c e ( b ) ,

36

K.G.

Winter

with a diameter ratio of 0.64, gave 80%.

Device

(c) is suggested

for use in three-dimensional

flows in which the pressure difference between

the outer tubes is first used to align the de-

vice with the flow and, when the probe is aligned with the flow, the pressure difference between the centre tube and the outer tubes may be used to determine (1976) has explored the use of a combined pitot-statie the skin friction. tube. Bertelrud

tube as a Preston

0.7 mm

A PPreston

,,

"

O.9mm
Ca)

0.45ram
45"

--

~ 0.80

A PPrestn

I'

/, "~

0.7mm 0.7mm

(b)

0.9mm
Fig. 22. Modified Preston tubes: Gupta.

(c)

6. Because of the wall-similarity

OBSTACLES

IN TWO-DIMENSIONAL

FLOW layer the pressures around

of the flow in a turbulent boundary

almost any obstacle can be used to derive skin friction. used are illustrated height or diameter.

Some of'the devices which have been Ap/~ where is a representative and it

in Fig. 23 which shows their sensitivity The sensitivity

of a Preston tube is also shown for comparison,

can be seen that the sensitivities

are all of the same order as that of the Preston tube and The fence or the square ridge have they create is enhanced

cover a range of about 2:1, at a given Reynolds number. higher sensitivities by utilizing face. 6.1.

than a Preston tube since the pressure difference

the suction at the downstream face as well as the pressure rise at the upstream

The devices are considered below in more detail. Fence (1955) and has been used by

The sub-layer fence was probably first suggested by Konstantinov Head and Rechenhert (1962), and by Vagt and Fernholz it has the advantage (1973).

As well as giving a relatively small to re-

large pressure difference main within the viscous pressure gradients.

that it may be made sufficiently

sub-layer and hence to be used with confidence

in flows with strong

Because of its fore-and-aft

symmetry it can also be used quantitatively

Skin friction

in turbulent boundary

layers

37

in flows which are separating. ficult to define,

Its small size has the disadvantage technique

that its geometry

is dif-

and the recommended

is to calibrate each particular calibration

fence against

a Preston tube in a well-behaved

flow, thus using a Preston-tube However,

as the primary

standard giving the surface shearing stress.

the calibration

curve for the fence for

hu /~ up to ii shown in Fig. 23, and taken from Rechenberg (1962), can be joined by a speculaT tlve fairing between hu /9 = ii and I00 to the line representing the results of Good and
T

Joubert

(1968) who investigated

the characteristics

of large fences.

The calibration

for the

square ridge in Fig. 23 is taken from the excrescence (1973).

drag measurements

of Gaudet and Winter that

The fact that this is quite close to the results of Good and Joubert suggests

the geometry of a large fence is not too critical,

but the sensitivity of the drag of ridges that the radius of as

shown by Gaudet and Winter to the rounding of the upper corners indicates the corners should be made as small as possible. a device for compressible

The sub-layer fence is also attractive

flows since its small height will minimize

the effect of variations

of density away from the wall.


4 x I0z 2

M'e

8 6 4

~/

5...i"" " ~ l x k with


cut-cut

2 tO 8 6 4 Z

Fence

,i

~'~

/~-Submerged . - - - step
Static -hole / pair /

| d

6 810

2 u~t 4
Y __

6 8102

,4

8 tO3

Fig. 23. 6.2 Razor blade

Sensitivities

of various obstacles

as skin-friction

meters.

In order to explore the flow in a viscous (1920) used pitot tubes with a rectangular

sub-layer of very small thickness

Stanton et al.

opening of width much greater than height with the They determined the effective height of the

test wall forming the inner surface of the tube.

tubes which would make their readings consistent with a linear velocity profile at the wall and thus produced a skin-friction meter which became known as a Stanton tube. subsequently sufficiently used larger tubes and showed that it was not necessary small to be within the sub-layer Other workers

that the tube s h o u l d b e A re-

for it to act as a skin-friction meter.

view of the investigations

up to 1954 covering both experimental (1955). Hool

and theoretical work was tube could hole and this

given by Trilling and H~kkinen

(1955) suggested that a Stanton-type

readily be formed by attaching a portion of a razor blade over a static-pressure simple device has since become widely used.

A detailed

saudy of the use of razor-blades flows.

at low speeds was made by East

(1966) with a view

to their use in three-dimensional tion determined

He calibrated against a Preston tube with skin fricTh~ ,,~e nf a pnTe~nn of a razor blade has obvious

from Patel's calibration.

38

K . G . Winter

attractions since the skin friction can readily be measured in any experiment in which static pressure holes are provided to measure the pressure distribution.* East simplified the tech-

nique by forming his pressure holes in magnets so that the portion of the razor blade could simply be placed over the hole, obviating the need for adhesive, and enabling the height of the opening formed by the blade to be determined from previous measurements of its thickness. The calibration obtained by East is shown in Fig. 24 and has the equation y* = -0.23 + O.61x* + O.O165~ .2 ph2T where x* = log ph2Ap u2 and y* = log u2 4.0 w (6-1)

3.0 f
/
*

""
O.O0~in

blade

2.0 -

fz~ *-"0.23*0.~18~
+0.0165x* ~,
1 I I

2.0

3.0

4)~phZ ~ x= Lq' k - ~ - )

5.0

6.0

-Fig. 24.

Calibration of Standard razor-blade surface-pitot tubes:

East.

Equation (6-1) is for the "standard" position of the razor blades that is with the edge of the razor blade over the leading edge of the static hole. various fore-and-aft positions is shown in Fig. 25. The change in the pressure rise for An additional error can arise from the

dependence of the static pressure on the size of hole used, and East suggested as a standard that d/h = 6, and that the breadth of the blade should exceed 30 times the height to avoid end effects. Smith et al. (1962) had assumed in their calibrations at supersonic speeds that re-

sults would be unaffected by grinding away the upper portion of the blade to minimize the interactions which might occur between blades in an array. East showed that the removal of the

1.2

,o
Position ,9 h A B ~ ,_ 0.60.O020inx +
O.O0~Oin 0.4 o

"%4

x ..
x

0.8 -

~ _ '

x + T~k.z~

x =~"~

-0"0153in ~'

-d(approx) 0.2 0 ~r -3
I II I I 2 I 3 I 4 5 I I

-2

-t

Ax/h
Fig. 25. The effect on Ap of varying the razor-blade longitudinal position relative to the static hole; East. (1971) should however be noted.

*The use of "plug-in" Preston tubes by Peake, et al.

Skin friction in turbulent boundary layers

39

upper portion of the blade in fact altered the calibration.

Inspection of a typical blade of

thickness 0.2 uln indicates that this is to be expected since the edge is composed of two chamfers, the first being of total angle about 20 and length of 0.25 mm, and the second of angle about 12 and length 0.85 mm, so that the flow over a blade is not likely to attach before the first shoulder and perhaps not before the second shoulder. Gaudet (unpublished) reanalysed

East's results and other unpublished results and found an empirical expression for the effective height of a blade hef f = h + Ah, Ah h h t - h h t

where

'

(6-2)

h being the height to the edge of the blade, and h t the total height. East's calibration then becomes y* = -0.459 + O.962x* - 0.O813x .2 + O.OO8x .3 where hef f is used in the expressions for x* and y*.

His expression for

(6-3)

Pal and Whir,law (1969) found a similar As a secondary experiment during

effect from adhesive tape used to hold down razor blades.

the course of the direct skin friction measurements of Winter and Gaudet, measurements were made of the sensitivity of razor blades for a wide range of heights, achieved by inserting packing under the blades. It was found that an empirically satisfactory, if conceptually un-

satisfying, fit to the results for compressible flow was obtained if ~ in (6-3) was replaced w by FcT w , where F c = (I + O.2M~) is the compressibility factor on skin friction empirically determined by Winter and Gaudet. Equation (6-3) modified in this way was also found to fall

between the two sets of results of Hopkins and Keener (1966) for forward and rearward positions of blades relative to the static pressure hole; these authors used large specially-made blades .ith a single-sided chamfer only. Their static pressure holes were very much smaller relative

to the height of the blades than those of East and Smith et aZ., so that qualitatively the mean of their results for the two positions of the blades corresponds to the results for the standard position of East and Smith et al. 6.3. Forward-facing step (see also Nituch, 1971), appreciating the difficulty of specifying

Nituch and Rainbird (1973)

readily the geometry of chamfered blades, set out to determine an alternative form which could be easily specified and which should give as large a value of pressure rise as possible. Their

starting point was a plain rectangular block mounted downstream of the static pressure hole but they discovered that provision of a semi-cylindrical cut-out in the upstream face of a

block, fitting the downstream half of the static pressure hole, led to an increase of about 40% in the pressure rise. This effect is illustrated in Fig. 23 where their line for the

block with cut-out is shown and also results for a plain block derived from pressure measurements made during the tests of Gaudet and Winter, over the range of hur/v < i00 and from those of Nash-Webber and Oates*(1971) over the range 6 < huT/~ < 160. For a block with cut-out "

diameter 1/3 times the height, of width-to-height ratio 1.5, and length (in the stream direction)-to-height ratio of 3, Nituch and Rainbird give the calibcation equation log Ap = 1.307 + O.117 log oh2Ap ~ w u2 for 3.1 10 6 < ph2Ap < 8.7 x 10 9 . u2

(5-4)

*Note that there are typographical errors in the paper of Nash-Wehher and Oates a power of 2 being omitted from both the terms Re d and Ms/M ~.

40

K.G.

Winter

They found that for zero pressure gradient the calibration was valid within 1% for blocks of height up to 6 times the momentum thickness, region of their boundary 6.4. Submerged step and Sivaselayers. that is extending well beyond the wail-similarity

The submerged step shown in Fig. 23 was suggested by Bradshaw and used by Edwards garam (1968). 0.04 mm. This device is very small,

the length of the ramp being 0.5 r~n and the depth

It was made relatively wide, 4 mm, so that a slot O.i m~ wide, could be used to A similar slot alongside the ramp was used to is apparently

measure the pressure at the base of the step. measure static pressure.

The advantage of the device is that its calibration effects.

independent of compressibility

The line in Fig. 23 is derived from the calibration

of Edwards and Sivasegaram obtained both at low speed and at M = 2.2, the length scale being the height of the step. 6.5. Static-hole pair shown in Fig. 23 is the use of a pair of static pressure holes. in the measurement Shaw

The other possibility

(1960) showed that there is an error incurred

of static pressure because (1968) realized simulT

of the shearing stresses acting across the face of the hole.

Duffy and Norbury

that this effect could be used to measure skin friction if the pressure were measured taneously at two holes of different and thus, sizes.

The pressure error varies nonlinearly with du /~ small, say du /~ < 50, the calibration is very

if the smaller hole is sufficiently

nearly dependent only on the size of the large hole.

The curve shown in Fig. 23 is derived

from the paper of Duffy and Norbury with the length scale taken as the diameter of the larger hole of a pair. The pressure signal available is some two orders less than that from a Pres-

ton tube of comparable diameter. Coleman forwards dependent

A variant on the static-hole

pair was explored by Green and

(1973) which used the pressure and the other 45 backwards. entirely on viscous forces,

difference between a pair of slots one inclined 45 Their aim was to produce a device the reading of which so that the calibration could be universal for any flow in prin-

provided the temperature

and density at the wall were known.

In this they succeeded to


w

ciple by showing that the pressure


T

difference was directly proportional for ~u /~ < iOO.


T

and independent

of ~u /~, where ~ is the slot width, gradients and inaccuracies

However,

the effects of local pressure pre-

in their device in which the slots were displaced streamwise, being obtained. They suggest alternative

vented an absolute calibration

designs which would

be worth studying if an instrument

is required for their purpose

in future.

7.

OBSTACLES IN THREE-DIMENSIONAL

FLOW

7.1.

Fence (1973) reconm~end the use of a surface fence in three-dimenslonal boundary

Vagt and Fernholz

layers by determining difference

its alignment with the surface flow by rotating it so that zero pressure and then setting it at 90 to this direction to find ~w" They also so that the

is obtained,

give an expression calibration 7.2.

for the change in pressure

difference with angle of inclination

for two-dimensional

flow may be used if the inclination

of the flow is known.

Razor blade flows East (1966) suggested a very simp]e way of using razor blades. of the flow to a East He

For three-dimensional investigated

the change in pressure with change in the angle of inclination shown in Fig. 26.

blade and found the unique calibration observed that if measurements known amount,

Using this unique dependence

were made for two different

settings of a blade differing by a

say 30 , then both the magnitude

and direction of the skin friction could be

Skin friction in turbulent boundary layers

41

deduced by means of a curve such as shown in Fig. 27 in combination with Fig. 26 and the calibration for two-dimensional flow (Fig. 24). This technique could be used with other devices.

1.0 ~ll~el~ll:l~" \ o+ ~o,= 0.8 -+'~\

!. / h 36 60 O.O05in x + O.OISin o n

0.6
A,. 0.4 --

~ \ \ ~Z~pp = C OS2

"t't~\\x\

"

'\\
0.2 x~\\\,

20

40

60

I~

I/~1 ~e~
Fig. 26. The effect of razor-blade yaw in a two-dimensional boundary layer: East.

40

30 20
I0 I a \l i

0.5

hO

"~.5

2.0

A PB/A PB':s -I0

Fig. 27. 7.3. Shaped block

The relationship between 8 and ApS/Ap( 8 30) deduced from empirical curve of Fig. 26: East.

A brief investigation of the possibility of utilizing the principle of the cut-out block of Rainbird and Nituch has been made by Dexter (1974). The triangular prism adopted is shown at The in-

the top of Fig. 28, and the shape should readily be accommodated in a circular plug.

tention was effectively to use pressure Pl, as equivalent to the pressure measured for a single block, but replacing the static pressure by the mean of the two pressures P2 and P3 on the

42

K.G.

Winter

rearward surfaces,

and also to determine

the direction of the shear stress indicated by the low frontal aspect for

difference of these pressures.

Dexter found that blocks of the relatively because

ratio suggested by Rainbird and Nituch were unsatisfactory the two rear faces did not vary unambiguously

the pressure difference presumably

with the angular setting, However,

as a result

of the nature of the separation from the front corners.

a satisfactory

calibration This

was found, as shown in Fig. 28, for a block having an aspect ratio of 9 for its faces. is not necessarily a lower limit since the next smaller aspect ratio tested was 3.

The tests

were made only over a very limited range of Reynolds before the calibration is accepted.

number and further work should be done

However, Fig. 28a shows how the angle of the flow relaand Fig. 28b that the angle can be used to determine the

tive to the block can be obtained,

pressure difference between the front face and the mean of that at the rear faces for zero angle. A calibration for zero angle should then give the skin friction. Dexter did not prodifference was in-

duce a calibration

in this form but his results showed that the pressure static pressure.

creased by about 30% compared with using a reference for a block with an equilateral

It should be noted that only over 60

triangular planform a calibration

is required

angular setting to deal with a full 3 6 ~ finite area,

variation of flow direction.

As with any device of

the calibration will be sensitive of the potentialities

to pressure gradients but it is clearly worth

further investigation

of the instrument.

4 =0.3 h h
1.0 0-8

PL

PJ" P,'Pz

P3

0"6 0,4 0,2 D dqroes 0 a I0 20 40 50 60 Determination of sheor-stress direction IO 20

.30 30

o
pz.p3, ~

40

50

60

O 0.6~ 0.4 O.Z

Determination datumpressurerise of
flows: Dexter.

Fig. 28.

Surface block for three-dimensional

Skin friction in turbulent boundary layers

43

8. 8.1. Heated film for two-dimensional Flow

ANALOGIES

Though Fage and Falkner had investigated the relatlonship between local skin friction coefficient and heat transfer from platinum strips embedded in a surface as long ago as 1931, the first practical instrument using the analogy between skin friction and heat transfer was designed by Ludwieg in 1949, and subsequently used in the classical experiments of Ludwieg and Tillmann (1949) on the variation of turbulent skin friction in pressure gradients. Ludwieg

showed that, if a heated surface could be made of sufficiently small stremnwise length that its thermal layer lay within the linear velocity region near the wall, then the heat flow was proportional to the wall-shearing stress to the power of 1/3. tion 1
q kAT =

He deduced the following equa-

0.807

'

(8-1)

with the condition i I 6_~q < 1 (8-2)

1.86 /c f_~6(_Re ~ _h

where q k

is the heat flow per unit area; is the thermal conductivity of air;

AT temperature difference between the heated element of length and the unheated surface; o C Prandtl number; uT~ L the non-dimensional thickness of the sub-layer;

6 the thickness of the thermal layer. q Ludwieg's analysis was repeated by Diaconis (1954) for compressible flows and he showed that equation (8-1) could be used provided the quantities appearing were evaluated at wall conditions.

The instrument designed by Ludwieg consisted of a block of copper 2 ~m long in the stream direction, by 9 mm wide, by 6 mm deep, and heated by an electrical current. The block was mount-

ed beneath a thin sheet of celluloid which formed the airswept surface, and which in turn was attached to a cylindrical housing. The celluloid and an air gap round the remaining sides of The temperature difference between the heated block The calibration performed in the turbu-

the block provided thermal insulation.

and the unheated wall was measured by thermocouples.

lent boundary layer of a low-speed blower wind tunnel confirmed the relationship (8-1) when the substantial heat losses from the block other than to the airstream were taken into account.

Liepmann and Skinner (1954) devised an alternative form of Ludwieg's instrument with a much reduced effective streamwise length. Their instrument consisted of a platinum wire of 13 um They reassessed the de-

diameter cemented into a groove in the surface of a piece of ebonite.

rivation of the form of the calibration of the instrument including the effect of pressure gradients. They also drew attention to the existence of a limitation on the minimum length For the boundary-layer type analysis used to be valid the thickness of the This was expressed

of an element.

thermal layer should be small compared with the length of the element. as

q~ kAT

>

(8-3)

44

K.G.

Winter

Bellhouse and Schultz make resistance

(1966) realised

that the thin-film techniques, measurements

which had been used to flows, could Their

thermometers

for surface temperature

in hypersonic

be applied to measure skin friction in place of the hot wire of Liepmann and Skinner. first investigations at low speeds revealed however that different calibrations

were obtained

in laminar and turbulent flow.

In a subsequent paper (8-1) in the form

(1965), which is concerned with compressible

flows,

they recast equation

q--~ = AT They noted, the element,

T3C o 3 (p~)3
w p that over the thickness of the thermal

(8-4) layer of

following Liepmann and Skinner, C P

and o could be taken as constant and that the factor so that the calibration should be independent

(p~) is almost independent effects. They

of temperature confirmed

of compressibility

this experimentally

for Mach numbers up to 3 using platinum films about 2 mm 0.2 nun The resistance of the heated film was used to measure temperature. its For

fired onto a pyrex glass substrate. own temperature

and that of a similar unheated film to measure

the reference

a film heated to a given temperature, flow,

so that C , ~ and a are independent P they showed that the calibration took the form

of the external

i2R AT

i A(OTW) 3 +

B of the element and p is evaluated

(8-5) at the

where i and R are the heating current and resistance film temperature.

As well as being simple the thin-film gauge has a high frequency response used to measure fluctuating quantities (Bellhouse and Schultz,

and hence can be with

1968) and in facilities

short running times.

Brown and Davey

(1971) describe

the use of a simple apparatus

for calibrating hot-film skin from a rotating

friction gauges in which the gauge is mounted in a stationary plate separated plate by a small air gap.

In a general determined

theoretical

investigation

of heat transfer in shear flows Spence and Brown

(1968)

series expansions temperature

relating the heat convected distribution,

into a stream from a heated element, gradient and ap-

with a "top-hat" proximated

to the skin friction and pressure

these by

i9 ~2~
T

5 dp kAT
18 dx q
'

(8-6)

I0

0o

under the conditions (i) Lu T V ~u (2) T < 640 for a unique calibration in laminar and turbulent flows to be obtained > 6.6 7 i to ensure that the T3-1aw apply;

Brown

(1967) describes

experiments

in which he showed that equation layer approaching

(8-6) gave accurate

re-

sults for skin friction in a laminar boundary

separation

in which circum-

stances a Stanton tube gave erroneously high readings. tion could be applied in laminar and turbulent

He also showed that the same calibra(2) above was met. In

flow provided condition

Skin

friction

in

turbulent

boundary

layers

45

order, 0.15 resin this Pope the

however,

to

ensure the

this

he made a special film clear the was mounted, round the

gauge

in which

the

glass

slide of

of

thickness by an epoxy of

mm, on which with a small

platinum left limit

was embedded slide of on the

in a block

copper

groove

airswept

surface.

The purpose Later, from however, the film to

arrangement (1972) adjacent

was to

effective the that

length

the heated of heat

element. conduction for instead the

investigated surfaces in

theoretically and concluded

consequences was not

this

responsible

differences the source so that

in caliof the

bration

obtained is profile

laminar

and turbulent of effective rather

flows. viscosity than linear

He suggested through form,

that

discrepancy velocity imply

a variation is of

the viscous and showed

sub-layer that velocity

the

quadratic

profiles

ueff Using (8-7)

= uW

he was able which results

(
is

1 + 0.042

?
)

.
calibrations than for laminar of and turbulent Spence flow.

(8-7)
The

to obtain

identical

condition and is

much more stringent

the secondone

and Brown above,

that 3

Ilr ?Y where to AT
W

<

0.51

X 106 (

AT $
W )

is

the

error so

in

r
W

consequent for aur/ov

upon applying = 64 an error

a calibration of 5% in rw is

obtained predicted.

in

a laminar

flow

a turbulent

flow

that

Rubesin the

et

al.

(1975) which the

have

recently

re-examined the losses plug. pipe

the length into

errors of the the

arising element.

from

thermal

conduction (1969) also film

into drew set on

substrate to

effectively of

increases the heat into

Guitton He used

attention the end of

importance rod it fitted

substrate. that

a thin of to used of of the

a glass with

a brass

He noted

a calibration when applied substrate that

his the

instrusame in-

ment obtained strument

mounted

in a brass plate. high

required out

correction that the glass

mounted has

in a perspex a relatively

They point thermal

by Bellhouse ebonite used

and Schultz by Liepmann conductivity. material wire resin. gave also wires of

conductivity, the of of use of

some 14 times a plastics a narrow, and Skinner of the

and Skinner. Because reverted of to

They investigated the the difficulty arrangement was laid

substrate thin-film and used

low thermal onto the

depositing Liepmann surface

element

they

a platinunrrhodium of this Rubesin epoxy method

25 urn diameter, The surface

which

on the to

substrate Gauges

in a layer produced lengths. by

was then handworked losses at zero flow

expose

the wire< effective wire

reduced showed

heat that

and small from

streamwise could readily unit

et al.

the heat upstream

convected

the heated so that

be detected

by unheated separation (1970) surface (but of a

installed

and downstream means of

a triple-wire is described wire

made a sensitive and Railly into the

indicator. see small 8.2. also

An alternative Railly, of 1972) perspex. for

manufacture

by Singh is pressed

in which

a Sum diameter

tungsten

block Heated

films

three-dimensional properties

flow and may be used the heat film of for finding which to the surface flow the flat.) insert. was a direcfilm

The single tion is

film

has directional it the of and finding flow. flow (1966)

by rotating aligned along

a minimum in

transfer, is normal

indicates flow is

that very

(The maximum when the direction used this by this means, with

the

The determination Drinkuth heated and Pierce wire (50

course,

requires

an adjustable the element Mylar

technique rhodium)

an instrument beneath

in which

urn diameter

platinum

mounted

a 25 urn thick

film.

46

K.G.

Winter

McCroskey and Durbin (1971) investigated

the possibilities

of producing a dual-film gauge with the aim of studying of films is of the films

which would give both the surface shearing stress and its direction, the boundary extremely layers on helicopter blades. to the properties

They pointed out that the calibration

sensitive

of the material and the physical dimensions

so that matching a pair of gauges would be very difficult with laboratory methods of manufacture. They therefore obtained films made commercially using the techniques developed for the

production of thin-film strain gauges with two elements 5 ~m thick, 0.05 mm wide and 5.5 mm long. temperature 40 between (Q1 + Q2). system for operating

set mutually perpendicular, twin-bridge

of nickel constant-

They also developed a special

the gauges,

and found that a linear relationship

existed over

the flow angle and the heat transfer from the two elements expressed as (QI - Q2)/ 1/3 They confirmed the T relationship for the heat transfer but noted that the absodepended upon the nature of the suhstrate material.

lute calibration 8.3.

Pulsed heated film friction gauge, analogous (1973). to the pulsed hot-wire technique, has

A variant on the heated-film

been studied by Ginder and Bradbury flow.

They use three parallel

films set normal to the separated by gaps 0.5 mm is to heat the centre the heat

The films are about 12 mm long with the centre film I m m w i d e

wide from the outer films, which are O.I mm wide. film by a short duration electrical pulse and the occurrence elements being provided

The mode of operation

current and to measure

the time interval between

of a change in resistance of one of the outer elements,

two outer flows.

in order that the gauge may be used in separating or attaching to make an estimate of the time interval t.

They use a very simple model that the process

The model assumes

takes place entirely within the linear velocity region at the wall, and that a distance y, is convected along the stream a distance and then diffuses back to the wall. Z (the separaThe time is then

heat diffuses outwards

tion of the heated and receiver elements) t =y2 + 2K

~u

where K is the thermal diffusivity,

K = k/0Cp.

The first detection will occur for the value of y which gives a minimum for t, which leads to

Kt ~2

3 2

K~g_.~ui

(8-8)

The conditions

on the use of the expression

(8-8) are that the time must be less than the small For

direct diffusion

time between the two films, but the distance must also be sufficiently

that the height, y, to which the heat diffuses must be such as to give yum/~ < 5 say. Prandtl number near unity this leads to ~u 4 < Ginder and Bradbury < 125 .

calibrated gauges in a laminar-flow channel and obtained an expression 2

Kt - 2.4 + 0.82
which agrees in form with (8-8). ture rise of the passive The signals to he measured were extremely small,

(8-9)
the tempera-

element being of the order of IO -~ times that of the active element, Ginder and Bradbury applied their

and the time interval being of the order of a millisecond.

Skin friction in turbulent boundary layers

47

gauges in a separated flow behind an aft-facing step and obtained the result shown in Fig. 29, which demonstrates the potential of their instrument for investigating the unsteady properties of separated flows. However, they found that in attached turbulent boundary layers their basic

calibration gave values for skin friction some 20% higher than indicated by a Preston tube, and further study is needed before the instrument can be considered for general use.
.I.0 o

-0.8

0.6

o
0.4

e>

"~
0.2

~ 7.6mm

Gauge span o 3.3mm

I
-4 -3 -2 -I

I
0
' I

I
2 3

I0-4du
dy Fig. 29.

s"

Measurements in separated flow behind a step with "time-offlight gauge": Ginder and Bradbury.

8.4.

Mass transfer

The use qualitatively of a surface coating which will subl~e is a common technique for determining transition from laminar to turbulent flow and is based on the increased rate of s~limation due to the increased skin friction in the turbulent region. (1951) made a quantitative investigation of the s~limation Owen and Ormerod

technique and obtained a relation-

ship of form similar to that for heated films, for the mass transfer from a small finite region, which can be e~ressed as i 3 mR ojA~ where m j A~ ~nerally pressure. = (~2Tw 1 0.54 \ - ~ / , (8-I0)

is the rate of mass transfer per unit area; is the molecular diffusion coefficient, and is the difference in concentration between the surface and the airstream. the concentration at the surface may be taken to correspond to the saturation vapour Konstantinov (1955) also derives this equation but with the constant having a value

0.807 as in the corresponding equation (8-1) for heat transfer, which can be e~ressed in the same form by replacing k the thermal conductivity by the thermal diffusivity K = k/pCp to give 1

0.807 -F~-/
Equation (8-i0) holds only when the diffusion process remains w i ~ i n the s~-layer.

(8-11)
When the

diffusion layer has a thickness of the same order as the turbulent b o u n d a ~

layer the relation-

ship becomes the same as the yon Karman e~ression for heat transfer, that is

48

K.G.

Winter

where Km = oUA$ and J is an analytic function which is zero for j/~ = I. Konstantinov tested diffusion gauges in a low speed wind tunnel. His gauges used water as the

working fluid which was evaporated from the surface of a small porous ceramic block of exposed area 2 mm 18 ~ml. A capillary tube was used to measure the rate of flow of the water to the block. From the results given it is difficult to assess the accuracy.

Murphy and Smith (1956) showed that the mass-transfer analogy could be applied to determine the skin friction on a flat plate. They used a film of silicone fluid and measured the time

history of the variation of the film thickness by an interferometric technique on a flat plate in a wind tunnel at low speed. It is not clear how they distinguished between changes in the

thickness of the film due to evaporation and due to streamwise movement under the shear stresses. Wazzan et a~. (1965) extended the investigation to supersonic speed, and limited

their measurement to the film front where the film was very thin so that convection could be neglected. They concluded on the basis of a comparison of their results with calculations

for a laminar boundary layer that mass transfer was proportional to skin friction.

A more detailed study of the possibilities of the technique was made by Sherwood and Tr~ss (1960) using a subliming material, naphthalene. at subsonic and supersonic speeds. Their experiments were made on a flat plate

A smooth surface on a layer of up to I n~n thickness was The thickness was measured before and after

obtained by applying naphthalene in molten form. a wind-tunnel run by mechanical means.

were in good m agreement with the analysis of Deissler and Loeffler (1958), extended to mass transfer, which indicates that the analogy factor for mass transfer becomes increasingly smaller than that for heat transfer as Mach number increases.

They showed that their results for K

9. 9.1. Oil flow

LIQUID TRACERS

Squire (1960) investigated theoretically the movement of an oil film under a boundary layer and showed that the oil follows the surface streamlines except near separation. did not attempt to produce a quantitative method for determining skin friction. However, he Meyer (1966)

described a variant of the oil-flow technique, the "oil-dot" technique, in which rows of spots of a suitable oil and pigment are placed on a model in a wind tunnel. When the air flow is Meyer discovered

started the dots run into streaks defining the surface streamline direction.

that the length of the streaks depended little on the dot size, as shown in Fig. 30, and that the length of the streak was proportional to the surface shear or heat transfer as shown in Fig. 31 for the flow over a half-cone delta-wing model at M = 12.6. Though the technique is

used regularly for qualitative measurements it has not been fully exploited quantitatively.

Tanner and Blows (1976) suggest a quantitative method of utilizing the temporal variation of the thickness of a liquid film on a surface due only to convection to find the skin friction of air flowing over the surface. They show that the variation with time of the thickness h

of a liquid film in a two-dimensional flow is given by

~t~h

3~I

~x~

(h3 ~

2~I ~x~ (h2Tw) .

(9-1)

If oil is deposited on a surface in a line normal to the flow then, except near separation, the second term in equation (9-1) may be expected to be dominant and the oil will flow down-

Skin friction in turbulent boundary layers

49

oRun697 x Run 698 z~Run699


015

A
/ I
x

\&~
\

010 ,(E

I
Ao

o"

C3 0 05

__
0

I
0.1

I
0.2 0.3

I
0.4

I
0.5

I
0.6

Streak length, Fig. 30.

in

The dependence of streak length on dot size for a fixed condition of surface flow: Meyer.
40

A Heat transfer measurements o FI0w indicator measurements


~
c

3o

&

A
e~
"-

~&
O

Curves matched at A and B

20

i--~r O
B

0.5

1,0

S/S T

Fig.

31.

A comparison distribution

of streak length distribution and heat for an incidence of 15: Meyer.

transfer

50

K G. Winter

stream.

Once the film has spread over the surface to be investigated Since the pressure-gradient

the thickness

h w[l] determ h 2 togical a

crease with time.

term depends upon h 3 and the friction It is therefore

the latter will become increasingly to look for a solution initially correction. By evaluating

dominant with increase of time.

ignoring the pressure-gradient

term and then to determine

the development

of the oil film following a surface particle Tanner x. Hence from equation (9-1)

and Blows show that ht = constant at given

x
Tw 2U 2 . O I t hdx - h (ht) 2 2~ 0 I

x
(ht)dx (9-2)

Tanner and Blows show that, if in a pressure h = ho(l + c) then

gradient a correction

term ~ is introduced

so that

= . ho. dp . + . 2U l
3 Tw dx 3t

I dp
T2 w

(9-3)

They measured and confirmed

the time history of flows using an interferometric the relationship

technique on a glass surface gradient. They also exa-

ht = const for flows in zero pressure

mined a flow approaching friction distribution.

separation

caused by a spoiler on a flat plate and deduced the skinat various times in

Figure 32 shows the film thickness distribution The results are replotted

terms of the number of fringes. is not independent

in Fig. 33 as the product ht, which gradient. The curve labelled

of t because of the effect of the pressure

t = ~ is obtained by applying the pressure-gradient friction can be calculated 350 300 /


2.50

correction.

From this curve the skin

by use of equation

(9-2).

/+,
/

J~

/ / /
/ l/

~ +---+ =T , Y-dp

~ 6~ '~'
~

200 L. h
c

t:5miy 150
I00

/ "\

_
,,

//
/.._~.-----2.0

'5 /
.--~v_.~ ~ 40 ---'--"

,/"'+
bU 80 ~

09.

5O

J
IOO Glass plate No oil

x m m 60

-50

Fig. 32.

Distribution separation:

of oil-film thickness under flow approaching Tanner and Blows.

The technique needs a special environment but clearly could be valuable methods that no knowledge

to make the interferometric

measurements

possible

in specialized

cases, and has the great advantage over many of the test medium is required. It can be extended (1976). As

of the properties

to three-dimensional

flows as shown in a subsequent paper by Tanner and Kulkarni

Tanner and Blows point out care must be taken in flows near separation of oil does not modify the flow. shear stress and film thicknesses There may, however, be limitations

that the accumulation of

on the combinations

for which the film will remain stable.

Figure 34 shows a

Skin friction in turbulent boundary layers

5]

photograph taken by the present author of a very thick oil film on a flat plate at a speed of 35 m/s. The photograph shows the flow in the region of the transition front near the centre

of the plate where the wavelength of the pattern in the oil is clearly decreasing through the transition region. In the lower right of the photograph the turbulent wedge spreading from

the intersection of the plate and the sidewall is indicated by the shorter wavelength of the disturbances compared with those in laminar flow.

70--

IT-oo
80- 5 min + IOmin o 20 min z~ 4 0 m i n {~ 6 0 m i n

50 - -

6u--J/

40-E g -~ 3 0 - -

/
T- 5 min--~. ~''

2O _

I0

I
o 20 40 x mrn 60

80

~oo Tanner and

Fig. 33.

Limiting form of oil-film thickness distribution: Blows.

Camera ! Surface of + polished wood/ Surface 1 painted black Flat plate

3S role

Tunne window Laminar

TurNlent wedge from ~ junction of plate leading edge end tunnel sidewall I 1.7 I 1.6 I I.S I m from leading edge 1.4

~-

Trop-~ition region us indicated by surface pitot tube


Fig. 34. Surface waves in thick film of oil.

52

K.G.

Winter

These difficulties scribed by Tanner

can be avoided by the use of a simplified version of the technique as de(1976), which may be applied to polished metal surfaces. In this version

only the leading edge of an oil film is used where the thickness varies stream distance w so that

linearly with down-

~x = -ut

(9-4) the shearing stress can be x, one to define

Hence by finding the gradient with distance of the film thickness obtained.

This is done by using two laser beams with a known separation,

the leading edge of the film and the other to record the evolution of the thickness by counting fringes as indicated by zero crossings of the output of a photocell light is focused. 9.2. Liquid crystals suggestion for a surface-coating technique is that of Klein and Margozzi (1969, 1970) apon which the reflected

Another

who have made an exploratory pear to be viscous these properties

investigation

of the use of liquid crystals.

Liquid crystals One of

liquids and yet show many of the properties light scattering

of solid crystals.

is selective

so that when illuminated with unpolarized

white

light incident at a given angle liquid crystals at each viewing angle. reflected

reflect strongly only one light wave length can cause a shift in the wavelength of the

Small changes in conditions

light in a reversible way.

Klein and Margozzi

showed that a mixture of liquid cryupon the shearing to temperature

stals could be produced,

the properties although

of which were primarily dependent the mixture also exhibited

stress imposed on the mixture,

sensitivity

and to the angle between the specimen and the direction of illumination Figure 35 shows a calibration ing annulus. The calibration obtained

and scattering.

in the shear flow between a fixed annulus and a rotat-

shows a reversal of colour change for ~w greater than about 300 this reversal would not often be of significance since

N/m 2 (3.06 x 10-2 g/mm2).

In practice

shearing stresses would rarely exceed the critical value.

(For example in a wind tunnel at a a skin-friction coefficient

Mach number of unity and a stagnation pressure of one atmosphere of 0.002 gives a shearing stress of about 75 N/m2.) ments in a pipe that a film of sufficient thickness

However, it was found on making experito exhibit the light-scattering properties

?
I l

IHIII IIIIII
\ O \
o

o~

\ \ \ \

%
4

I 2
550
560 570 580 590 nm 600 Woveleng'l'h,

Fig. 35.

Wavelength of light scattered stress: Klein and Margozzi.

from liquid crystals under shear

Skin friction in turbulent boundary layers

53

produced marked ridges normal to the flow, and that the scattered light signal fluctuated considerably making it extremely difficult to measure the wavelength. A great deal of further

development is therefore required before the technique can be considered for routine application.

IO.

CONCLUDING REMARKS

Of all the techniques reviewed it is apparent that none can be considered an absolute and reliable standard. The obvious technique of directly measuring the surface shearing stress by

a force balance is beset by many pitfalls which may be overcome in particular cases but the possibility of specifying a priori the requirements to be met in general seems remote. An

analysis is given of the errors arising from various causes and this may provide some guidance in design. The most reliable device at present seems to be the Preston tube because of its However, there is

simple geometry and because it has been investigated the most thoroughly.

still room for further work on the effects of pressure gradient, of flow unsteadiness and of heat transfer and in three-dimensional flow. Potentially, sub-layer fences hold most promise

for devices of the pressure-measuring type if a design can be found with a geometry easy to manufacture repeatably, which is very difficult because of their smell size. There is an

opportunity for the exercise of some ingenuity in devising instruments for use in three-dimensional flows. liable. In the long run the heated-element instrument is likely to prove the most re-

For general application the discovery of a simple shear-sensitive surface-coating

agent would be most welcome.

REFERENCES Allen, J . M . (1968) Use of Baronti-Libby transformations and Preston tube calibrations to determine skin friction from turbulent velocity profiles. NASA TN D-4853. Allen, J . M . (1973) Evaluations of compressible-flow Preston tube calibrations. NASA TN D-7190. Allen, J . M . (1976) Systematic study of error sources in skin-friction balance measurements. NASA TN D-8291. Baronti, P. O. and Libby, P. A. (1966) Velocity profiles in turbulent compressible boundary layers. AIAA J. ~, 193-202. Bellhouse, B. T. and Schultz, D. L. (1965) The measurement of skin friction in supersonic flow by means of heated thin film gauges. ARC R & M 3490. Bellhouse, B. J. and Schultz, D. L. (1966) Determination of mean and dynamic skin friction, separation and transition in low-speed flow with a thln-film heated element. J. Fluid Mech. 24, 2, 379-400. Bellhouse, B. J. and Schultz, D. L. (1968) The measurement of fluctuating skin friction in air with heated thin-film gauges. J. Fluid Mech. 32, 4, 675-680. Bertelrud, A. (1976) A surface total head/static tube for skin friction measurement. Technical Note FFA AU-II33. Bradshaw, P. (1959) A simple method for determining skin friction from velocity profiles. J. Aerospace Sci. 26, 841. Bradshaw, P. and Unsworth, K. (1973) A note on Preston tube calibrations in compressible flow. IC Aero. Rep. 73-07. AIAAJ 12, 9, 1293-4 (1974). Brown, G. L. (1967) Theory and application of heated films for skin friction measurement. Proc. Heat Transfer & Fluid Mechanics Institute, Stanford Univ. Press, pp. 361-381. Brown, G. L. and Davey, R. F. (1971) The calibration of hot films for skin friction measurement. Rev. Sci. Instrum. 42, II, 1729-31. Brown, K. C. and Joubert, P. N. (1967) Measurement of skin friction in turbulent boundary layers with adverse pressure gradients. J. Fluid Mech. 35, 4, 737-757. Bruno, J. R., Yanta, W. J. and Risher, D. B. (1969) Balance for measuring skin friction in the presence of heat transfer. Final Report. USA Naval Ordnance Lab. NOLTR-69-56. Chu, J. and Young, A. D. (1975) Measurements in separating two-dimensional turbulent boundary layers. Paper 13, AGARD CPP 168. Chue, S . H . (1975) Pressure probes for fluid measurement. Prog. Aerospace Sci. 16, 147-223. Clauser, F. H. (1954) Turbulent boundary, layers in adverse pressure gradients. J. Aeronaut. Sci. 21, 2, 91-108. Coles, D. (1953) Measurements in the boundary layer on a smooth flat plate in supersonic flow. II. Instrumentation and experimental technique at the Jet Propulsion Laboratory (1953) JPL Report 20-70. Coles, D. E. (1956) The law of the wake in the turbulent boundary layer. J. Fluid Mech. ~, 2, 191-226.

54

K.G.

Winter

Davidson, J. K. (1961) The experimental evaluation of the surface impact probe technique of measuring local skin friction under conditions of moderate heat transfer. Univ. of Texas DRL-468. Deissler, R. G. and Loeffler, A. F. (1958) Analysis of turbulent flow and heat transfer on a flat plate at high Mach numbers with variable fluid properties. NACSI TN 4262. Dershin, H., Leonard, C. A., Gallaher, W. H. and Palmer, J . P . (1966) Direct measurement of compressible turbulent boundary layer skin friction on a porous flat plate with mass injection. NASA CR 79095. Dexter, P. (1974) Evaluation of a skin-friction vector-measuring instrument for use in 3-d turbulent incompressible boundary layers. Univ. of Southampton Dept. of Astronautics & Aeronautics Undergraduate Project. Dhawan, S. (1953) Direct measurements of skin friction. NACA Report 1121. Diaeonis, N. S. (1954) The calculation of wall shearing stresses from heat-transfer measurements in compressible flows. J. Aeronaut Sci. 21, 3, 201-2. Errata Ibid 7, 499 (1954). Dickinson, J. (1965) The determination of skin friction in two-dimensional turbulent flows. Doctoral thesis Univ. Laval. van Driest, E . R . (1951) Turbulent boundary layer in compressible fluids. J. Aeronaut. Sci. 18, 3, 145-160 ~ 216. van Driest, E . R . (1956) On turbulent flow near a wall. J. Aeronaut Sci. 23, 1007-1Oll & 1036. Drinkuth, R. H. and Pierce F. J. (1966) Directional heat meter for wall shear stress measurements in turbulent boundary layers. Rev. Sci. Instrum. 37, 6, 740-741. Duffy, J. and Norbury, J. F. (1968) The measurement of skin friction in pressure gradients using a static hole pair. Thermodynamics & Fluid Mechanics Convention IME. East, L. F. (1966) Measurement of skin friction at low subsonic speeds by the razor-blade technique. ARC R & M 3525. East, L. F. and Hoxey, R. P. (1969) Low-speed three-dimensional boundary layer data, parts 1 and 2. ARC R & M 3653. Edwards, B. and Sivasegaram, S. (1968) An experimental investigation of M = 2.2 turbulent boundary layers in nominally zero pressure gradients. IC Dept. of Mech. Eng. BL/TN/3. Everett, H. U. (1958) Calibration of skin friction balance discs for pressure gradient. Univ. of Texas DRL-426. Fage, A. and Falkner, V . M . (1930) An experimental determination of the intensity of friction on the surface of an aerofoil. ARC R & M 1315, also Proc. R. Soc. A, 129. Fenter, F. W. and Lyons, W. C. (1957) The direct measurement of local skin friction on Aerobee-Hi rockets in free flight. Univ. of Texas DRL-391. Eenter, F. W. and Stalmach, C. J. (1957) The measurement of local turbulent skin friction at supersonic speeds by means of surface impact-pressure probes. Univ. of Texas DRL-392. J. Aerospace Sci. 25, 12, 793-794 (1958). Fisher, D. F. and Saltzman, E. J. (1973) Local skin friction coefficients and boundary-layer profiles obtained in flight from the XN-70-I airplane at Mach numbers up to 2.5. NASA TN D7220. Fowke, J. G. (1969) Development of a skin friction balance to investigate sources of error in direct skin friction measurements. NASA TM X 61905, N69-40-054. Franklin, R. E. (1960) A force-displacement indicator for a drag balance. ARC CP 549. Franklin, R . E . (1973) Some measurements with simple skin-friction balances. Univ. of Oxford Report 1058/73, ARC 34447. Froude, W. (1872) Experiments on the surface-friction experienced by a plane moving through water. 42nd British Association Report pp. 118-124. Galbraith, R. A. McD. and Head, M. R. (1975) Eddy viscosity and mixing length from measured boundary layer developments. Aeronaut. Q XXVI, 2, 134-154. Garringer, D. J. and Saltzman, E. J. (1967) Flight demonstration of a skin-friction gauge to a local Mach number of 4.9. NASA TN D-3830. Gaudet, L. and Winter, K. G. (1973) Measurements of the drag of some characteristic aircraft excrescences immersed in turbulent boundary layers. Paper No. 4, AGARD CP 124. Ginder, R. B. and Bradbury, L. J. S. (1973) A preliminary investigation of a pulsed gauge techniqu for skin friction measurements in highly turbulent flows. ARC 34448. Goldstein, S. (Ed.) (1938) Modern developments in fluid dynamics Vol. II. Clarendon Press 361. Good, M. C. and Joubert, P . N . (1968) The form drag of two-dimensional bluff plates immersed in turbulent boundary layers. J. Fluid Mech. 31, 3, 547-582. Green, J. E. and Coleman, G. T. (1973) An exploratory study of shear-stress-induced pressures in inclined slots beneath a turbulent boundary layer. RAE Technical Memorandum Aero 1513. Guitton, D . E . (1969) A transient technique for measuring skin friction using a flushmounted heated film and its application to a wall jet in still air. McGill University Report 69-9. Gupta, R. P. (1975) A new device for skin-friction measurement in three-dimensional flows. AIAA J. 13, 2, 236-238. Hastings, R. C. and Sawyer, W. G. (1970) Turbulent boundary layers on a large flat plate at M = 4. ARC R & M 3678. Head, M. R. and Rechenberg, I. (1962) The Preston tube as a means of measuring skin friction. ARC 23459. J. Fluid Mech., 14, 1-17 (1962). Head, M. R. and Vasanta Ram V. (1971) Improved presentation of Preston tube calibration. India, Indian Inst. Tech. Kanpur Dept. Aero. Eng. AE-IO/1970. Aeronaut. Q. XXII, 3.

Skin friction in turbulent boundary layers

55

Hill, O. (1963) Experimental investigation of the impact probe method for determining local skin friction in the presence of an adverse pressure gradient for a range of Mach numbers from 1.70 to 2.75. Univ. of Texas DRL-498. Holmes, W. E., Luxton, R. E. (1967) Measurement of turbulent skin friction by a Preston tube in the presence of heat transfer. J. Mech. Eng. Sci., 9, 3, 159-166. Hool, J . H . (1955) Measurement of skin friction using surface tubes. ARL Aero Note 144. Aircraft Engng. 28 (1956). Hopkins, E. J. and Keener, E. R. (1966) Study of surface pitot tubes for measuring turbulent skin friction at supersonic Maeh numbers - adiabatic wall. NASA TN D-3478. Hopkins, E. J., Keener, E. R., Polek, T. E. and Dwyer, H. A. (1972) Hypersonic turbulent skin friction and boundary-layer profiles on non-adiabatic flat plates. AIAA J., I_~0, I, 40-48. Hsu, E . Y . (1955) The measurement of local turbulent skin friction by means of surface pitot tubes. David Taylor Model Basin Report 957. Jeromin, L. O. F. (1968) A transformation for compressible turbulent boundary layers with air injection. J. Fluid Mech., 31, i, 65-94. Johnston, J . P . (1960) On the three-dimensional turbulent boundary-layer generated by secondary flow. J. Bas. Eng. ASME Ser. D 82. Keener, E. R. and Hopkins, E. J. (1969) Use of Preston tubes for measuring hypersonic turbulent skin friction. NASA TN D-5544, also AI2~4 Paper 69-345. Kempf, G. (1929) Neue Ergebnisse der Widerstandsforschung. Werft, Reederei, Hafen II, 234-239. Klein, E. J. and Margozzi, A. P. (1969) Exploratory investigation of the measurement of skin friction by means of liquid crystals. NASA TM X-1774. Klein, E. J. and Margozzi, A. P. (1970) Apparatus for the calibration of shear sensitive liquid crystals. Rev. Sci. Ins~um, 41, 2, 238-239. Proceedings. Computation of turbulent boundary-layers (1968). AFOSR-IFP - Stanford Conference. Kline, S. J., Morkovin, M. V., Savron, G. and Cockrell, D. J. (Eds.) Vol. I Methods, predictions, evaluation and flow structure. Coles, D. E. and Hirst, E. A. (Eds.) Vol. II Compiled data. Konstantinov, N. I. and Dragnysh, G. L. (1960) The measurement of friction stress on a surface. DSIR RTS 1499. (Energomashinostroenie, Tr~dy LPI 176, 191-2OO, 1955). Konstantinov, N. I. (1961) Comparative investigation of the friction stress on the surface of a body. DSIR RTS 1500. (Energomashinostroenie, Tr~dy LPI 176, 201-213, 1955). Kovalenko, V. M. and Nesterovich, N. I. (1973) The drag of ridges of finite length in a a turbulent boundary layer. Izvest. Sibirsk Otdel. A. N. SSSR Ser Tekhn Nauk 13, 3, 107-114. IL4E Lib. Trans. 1852 (1975). van Kuren, J . T . (1974) A skin friction gauge for hyperthermal flow. AFFDL TR-74-60. Liepmann, H. W. and Skinner, G. T. (1954) Shearing-stress measurements by use of a heated element. NACA TN 3268. Ludwieg, H. (1949) Ein Ger~t zur Messung der Wandschubspannung turbulenter Reibungsschichten. Ing. Arch., 17, 207-218. Instrument for measuring the wall shearing stress of turbulent boundary layers. NACA TM 1284 (1950). Ludwieg, H. and Tillmann, W. (1949) Untersuehungen Uber die Wandschubspannung in turbulenten Relbungsschiehten. Ing. Arch., 17, 288-299. Investigations of the wall shearing stress in turbulent boundary layers. NACA TM 1285 (1950). Lyons, W. C. (1957) The design of an acceleration insensitive skin friction balance for use in free flight vehicles at supersonic speed. Univ. of Texas DRL-397. Mabey, D. G. and Gaudet, L. (1973) Some applications of small skin friction balances at supersonic speeds. RAE Technical Memorandum 1514. J. Aircraft 12, iO, 819-825 (1975). Mabey, D. G., Meier, H. U. and Sawyer, W. G. (1974) Experimental and theoretical studies of the boundary layer on a flat plate at Math numbers from 2.5 to 4.5. RAE Technical Report 74127. MaeArthur, R. C. (1963) Transducer for direct measurement of skin friction in the hypersonic shock tunnel. CAL Report 129. McCroskey, W. J. and Durbin, E . J . (1971) Flow angle and shear stress measurements using heated films and wires. USA ASME Paper 71-WA/FE 17. J. bas. En~ng. 94, I, 46-52 (1972). McDill, P. L. (1962) The design and experimental evaluation of a skin f-~iction balance for measuring local turbulent shear stress in the presence of heat transfer at a Mach number of 5. Univ. of Texas DRL-453. Meyer, R. F. (1966) A note on a technique of surface flow visualisation. NRC Aero Report LR-457. Miller, B . L . (1971) A floatlng-element skin-friction meter. Univ. of Leicester Eng. Dept. Report 71-19. Miller, B. L. (1973) Direct measurement of skin friction. ARC 34431. Monaghan, R. J. (1955) On the behaviour of boundary layers at supersonic speeds. Fifth International Aeronautical Conference, IAS-RAeS Proc., pp. 277-315. IAS Preprint 557. Moore, J. W. and McVey, E. S. (1966) Investigation of systems and techniques for multicomponent mieroforce measurements on wind tunnel models. (Progress report July 1965 to January 1966 on a skin friction drag sensor.) Virginia Univ. Report EME-4029-102-66U. NASA CR74385. Morsy, M. G. (1974) An instrument for the direct measurement of the local wall shear stress on circular cylinders. J. Phys. E !, 83-86.

56

K.G.

Winter

Moulic, E. S. (1963) Flat plate skin friction in low density hypersonic flow - preliminary results. ARL Report 63-24. Murphy, J. S. and Smith, A. M. O. (1956) Measurement of wall shearing stress in the boundary layer by means of an evaporating liquid film. J. appl. Phys. 27, 9, 1097-1103. Naleid, J. F. (1958) Experimental investigation of the impact pressure probe method of measuring local skin friction at supersonic speeds in the presence of an adverse pressure gradient. Univ. of Texas DRL-432. Nash-Webber, J. L. and Oates, G. C. (1971) An instrument for skin friction measurements in thin boundary layers. J. Bas. Eng. 93, Series D, 4, 571-575. Newman, B. G. (1951) Some contributions to the study of the turbulent boundary layer. Australian Dept. Supply Report ACA-53. Nituch, M. J. (1971) The use of contruent obstacle blocks for the indirect measurement of turbulent skin friction on smooth surfaces. Master's thesis. Carleton University. Nituch, M. J. and Rainbird, W. J. (1973) The use of geometrically-similar obstacle blocks for the measurement of turbulent skin friction. ARC 34423. O'Donnell, F. B. (1964) A study of the effect of floating element misalignment on skin friction accuracy. Univ. of Texas DRL-515. O'Donnell, F. B. and Westkaemper, J. C. (1965) Measurement of errors caused by misalignment of floating element skin-friction balances. AIAA J., ~, I, 163-165. Owen, P. R. and Ormerod, A. O. (1951) Evaporation from the surface'of a body in an airstream. (With particular reference to the chemical method of indicating boundary-layer transition.) ARC R & M 2875. Ozarapoglu, V. (1973) Measurements in incompressible turbulent flows. Thesis Universit& Laval. Paros, J . M . (1970) Application of the force-balance principle to pressure and skin friction sensors. 16th Annual Technical Meeting, Proc. Inst. Environmental Science, 363-368. Patel, V. C. (1965) Calibration of the Preston tube and limitations on its use in pressure gradients. J. Fluid Mech., 23, i, 185-208. Patel, V. C. and Head, M. R. (--i-968) Reversion of turbulent to laminar flow. J. Fluid Mech., 34, 2, 371-392. Pai, B. R. and Whitelaw, J . H . (1969) Simplification of the razor blade technique and its application to the measurement of wall-shear stress in wall-jet flows. Aeronaut. Q. X X, 355-364. Peak, D. J., Brakmann, G. and Romeskie, J . M . (1971) Comparisons between some high Reynolds number turbulent boundary-layer experiments at Mach 4, and various recent calibration procedures. Paper Ii, AGARD CP 93. Perry, A. E. and Joubert, P. N. (1965) A three-dimensional turbulent boundary layer. J. Fluid Mech., 22, 2, 285-304. Perry, A. E., Bell, J. B. and Joubert, P. N. (1966) Velocity and temperature profiles in adverse pressure gradient turbulent boundary layers. J. Fluid Mech., 25, 2, 299-320. Pierce, F. J. and Zimmerman, B. B. (1972) Wall shear stress inference from two and threedimensional turbulent boundary layer velocity profiles. ASME paper 72-EA/FE4. Pope, R. J. (1972) Skin-friction measurements in laminar and turbulent flows using heated thin-film gauges. AIAA J. iO, 6, 729-730. Prahlad, T. S. (1968) Wall similarity in three-dimensional turbulent boundary layers. AIAA J. 6, 9, 1772-1774. Prahlad, T. S. (1972) Yaw characteristics of Preston tubes. AIAA J. 10, 3, 357-359. Preston, J . H . (1953) The determination of turbulent skin friction by means of pitot tubes. ARC 15758. J1. R. Aeronaut. S. 58, 109-121 (1954). Railly, J . W . (1972) Embedded hot-w---ire method of wall shear measurement. J. Phys. E. 5, 5, 496. Rajaratnam, N. and Froelich, C. R. (1967) Boundary shear stress in turbulent boundary layers on smooth surfaces. J1. R. Aeronaut. Soc., 71, 673, 52-53. Rajaratnam, N. and Muralidhar, D. (1968) Yaw probe used as Preston tube. Aeronaut. J., 72, 1059-60. Rao, G. N. V., Kedravan, N. R. and Mahadevan, R. (1970) Estimation of turbulent boundary layer skin friction by means of a dual pitot tube. J. Aeronaut. Soc. India, __22, 4, 269-271. Rechenberg, I. (1963) The measurement of turbulent wall shear stress. ZFW ii, Ii. ARA Library Translation ii (1965). Rubesin, M. W., Okuno, A. F., Matur, G. G. and Brosh, A. (1975) Flush mounted hot-wire gauge for skin friction and separation detection measurements. 6th International Cong. on Instrumentation in Aerosp. Fac. Schoenherr, K. E. (1932) Resistance of flat surfaces moving through a fluid. Trans. Soc. nay. Archit mar Engrs. NY 40, 279-313. Schultz-Grunow, F. (1940) Ne--ues Reiungswiderstandsgesetz f~r glatte Platten. Luftfahrtforschung, 17, 239-246. New frictional law for smooth plates. NASA TM 986 (1941). Shaw, R. (1960) The influence of the hole dimensions on static pressure measurements. J. Fluid Mech., 7, 4, 550-564. Sherwood, T. K. and Tr~ss, Olev. (1960) Sublimation mass transfer through compressible boundary layers on a flat plate. J. Heat Transfer, 82, C4, 313-324. Sigalla, A. (1965) Calibration of Preston tubes in supersonic flow. AfAA J., ~, 8, 1531. Simpson, R. L. and Whitten, D. G. (1968) Preston tubes in the transpired boundary layer. AIAA J., 6, 9, 1776-1777.

Skin friction in turbulent boundary layers

57

Singh, G. and Railly, J . W . (1970) Embedded hot wire method of wall shear measurement in air flow. J. Phys. E. 3, 405-407. Smith, D. W. and Walker, J. H. (1959) Skin friction measurements in incompressible flow. NASA TR R26. Smith, K. G., Gaudet, L. and Winter, K. G. (1962) The use of surface pitot tubes as skinfriction meters at supersonic speeds. ARC R & M 3351. Sormner, S. C. and Short, Barbara J. (1955) Free-flight measurements of turbulent-boundarylayer skin friction in the presence of severe aerodynamic heating at Mach numbers from 2.8 to 7.0 NACA TN 3391. Spence, D. A. (1959) Distributions of velocity, enthalpy and shear stress in the compressible turbulent boundary layer on a flat plate. RAE Report Aero 2631. J. Fluid Mech., 8, 368-387 (1960). Spence, D. A. and Brown, G. L. (1968) Heat transfer to a quadratic shear profile. J. Fluid Mech., 33, 4, 753-773. Squire, L. C. (1960) The motion of a thin oil sheet under the boundary layer on a body. RAE Report Aero 2636 (1960). J. Fluid Mech., II, 2, 161-179 (1961). Squire, L. C. (1969) A law of the wall for compressible turbulent boundary layers with air injection. J. Fluid Mech., 37, 3, 449-456. Staff of Aerodynamics, Division, NPL. (1958) On the measurement of local surface friction on a flat plate by means of Preston tubes. ARC R & M 3185. Stalmach, C. J. (1958) Experimental investigation of the surface impact pressure probe method of measuring local skin friction at supersonic speeds. Univ. of Texas DRL-410. Stanton, T. E., Marshall, D. and Bryant, C . N . (1920) On the conditions at the boundary of a fluid in turbulent motion. Proc. R. Soc. Series A, 97, 413-434. Stevenson, T . N . (1963) A law of the wall for turbulent boundary layers with suction or injection. Coll. of Aeronautics Report 166. Stevenson, T . N . (1964) The use of Preston tubes to measure the skin friction in turbulent boundary layers with suction or injection. Coll. of Aeronautics Report Aero 173. Tanner, L. H. A skin friction meter, using the viscosity balance principle, suitable for use with flat or curved metal surfaces. (To be published in J. Phys. E.). Tanner, L. H. and Blows, L. G. (1976) A study of the motion of oil films on surfaces in air flow, with application to the measurement of skin friction. J. Phys. E. k, 3, 194-202. Tanner, L. H. and Kulkarni, V. G. The viscosity balance method of skin friction measurement. Further developments including applications to three-dimensional flow. (To be published in J. Phys. E.). Thompson, B. G. J. (1965) A new two-parameter family of mean velocity profiles for incompressible boundary layers on smooth walls. ARC R & M 3463. Townsend, A. A. (1962) The behaviour of a turbulent boundary layer near separation. J. Fluid Mech., 12, 536-554. Trilling, L. and H~kkinen, R. J. (1955) The calibration of the Stanton tube as a skinfriction meter. 50 Jahre Grenzschichtforschung, 201-209. Vagt, J. D. and Fernholz, H. (1973) Use of surface fences to measure wall shear stress in three-dimensional boundary layers. Aeronaut. Q .xxIv, 2, 87-91. Vinh, N. D. (1973) Errors in wall friction measurements performed with the aid of a floating element balance. C. R. Acad. Sci. Paris S$rie A, 277, 1115-7. Wazzan, A. R., Smith, A. M. O. and Lind, R. C. (1965) Mass-transfer method of measuring wall shear stress in supersonic flow. Phys. Fluids, 8, 9, 1641-1646. Wieghardt, K. (1946) Increase of the turbulent frictional resistance caused by surface irregularities. MAP R & T 103. (Translation of FB 1563 ZWB, 1942). Weiler, J. E. and Hartwig, W. H. (1952) The direct measurement of local skin friction coefficient~ Univ. of Texas DRL-295. Weiler, J. E. (1954) Design of an accelaration insensitive skin friction balance for flight testing. Univ. of Texas DRL-342. Westkaemper, J. C. (1963) Step-temperature effects on direct measurements of drag. AIAA J.= !, 7, 1708-1710. Winter, K. G. and Gaudet, L. (1970) Turbulent boundary-layer studies at high Reynolds numbers at Mach numbers between 0.2 and 2.8. ARC R & M 3712. Yanta, W. J., Brott, D. L., and Lee, R. E. (1969) An experimental investigation of the Preston probe including effects of heat transfer, compressibility and favourable pressure gradient. AIAA Paper 69-648. Young, F . L . (1965) Experimental investigation of the effects of surface roughness on compressible turbulent boundary layer skin friction and heat transfer. Univ. of Texas DRL-532. Young, F. L. and Westkaemper, J. C. (1965) Experimentally determined Reynolds analogy factor for fiat plates. AIAA J., 3, 6, 1201-1202.

You might also like