Bradshaw - Fernholz - Turbulence
Bradshaw - Fernholz - Turbulence
Bradshaw - Fernholz - Turbulence
External Flows
H.-H. FERNHOLZ
With 7 Figures
2.1 Introduction
We begin with a brief survey of some general turbulent shear flow
configurations which, merged or distorted, make up the more complex
flow patterns which are often found in practice but which are often not
accessible to experiment or calculation. The greater part of this chapter
will be concerned with a discussion of steady two- and three-dimensional
turbulent boundary layers at low Mach numbers. Boundary layers are
the commonest members of the class of "thin shear layers" (Chapt. 1)
and provide a convenient framework within which to discuss the effects
of boundary conditions on shear layers. If the wall or free stream
boundary conditions are changed, the shear layer will feel this change
as a more or less strong perturbation. Most shear layers in real life
undergo such perturbations, and the difference between this chapter
and some of the existing reviews of th~ subject is that we have tried to
give perturbed shear layers a share of space proportional to their im-
portance. Unperturbed free shear layers--the classical wake, jet and
mixing layer-- are rather rare in practice and the data on growth rate,
etc., which are normally needed can be summarized quite briefly (Sect.
5.9). Free shear layers near solid boundaries are usually perturbed and
46 H.-H. FERNHOLZ
-3
A_p
%0
T-1
1;o -011-~'5~x/~ ~ 10
0 ~.~
1~-"~-~" xll 1'.0
I ~ ," x/l IJ v h
Fig. 2.1. Attached and separating flow (after THWAITES[2.446], Fig. V.13)
I ~' I/ I I II ~~ iI \\
Fig. 2.3a--c. Some complex turbulent shear flow configurations. (a) boundary layer flow
over a step, (b) duct flow in an internal ejector, (c) flow over an aerofoil with a front slot
(from HORTON 1-2.267])
According to the closure model chosen: Vg, ~'w, v-0,q2 etc. (see Chapt. 5).
For three-dimensional flow it is necessary to know a further function,
being a combination of 0, U, V, and W which is determined from a
compatibility condition (TIN(; [2.8]).
Fig. 2.4a and b. Visualization of a boundary layer by means of smoke. Photograph a from
[2.282], photograph b by courtesy of P. BRAOSr~AW.(a) Streamwise view, (b) side view, flow
right to left
1.8. We now seek more details of the turbulent motion, and of the charac-
teristic velocity and length scales of the outer layer. Despite the
difficulties of the task, obvious from the smoke photographs of Fig. 2.4,
some progress can be made, though the outer-layer scales are much more
complicated than those of the inner layer.
In turbulent boundary layers one boundary is formed by the wall,
where the turbulent shear flow must accommodate itself to the wall
through a viscous sublayer and the large velocity gradients associated
52 H.-H. FERNHOLZ
with it. Very close to the wall the fluctuating velocities and therefore the
Reynolds shear stresses tend to zero, while the total shear stress remains
constant. Thus the viscous shear stress pOU/Oy, and the velocity gradient
OU/Sy, are much larger than elsewhere in the boundary layer. The other
boundary is the turbulent/non-turbulent interface where the shear
flow with high vorticity accommodates to an essentially irrotational
outer flow through the "viscous superlayer" (Sect. 1.9). This latter
irregular but very distinct boundary is common to all turbulent thin
shear layers with at least one free boundary, but its quantitative proper-
ties depend on the characteristic scales and the boundary conditions
of the specific shear layer. It has been found useful (see Fig. 1.6) to divide
the inner layer at the wall into the linear sublayer and the buffer layer
(together comprising the viscous sublayer) and the Jog-law region.
The outer layer which covers about 80 % of the boundary layer is bounded
Table 2.1. Multi-layer model of turbulent boundary layer: see text for symbols
Range within O< y< 4Ov/u~ 40v/u,< y<0.2~ 0.2 < y/fi Inhabits
boundary layer y>0.46
Characteristic (v/uOi kr (v/u), kr 6 or A (v3/%)~
length or v/Vo
Characteristic u, u, u,, Uo V~~ K(v~op
velocity
The Inner Layer (0 < y/t~ < O.1-0.2 roughly; 0 < U/U~ < 0.7 very roughly)
Initially we consider two-dimensional constant-density flow over an
impermeable wall, as outlined in Section 1.8, in a region close enough
to the wall for the total shear stress to be nearly equal to the wall value
zw and for all derivatives with respect to x to be negligible. The behavior
of the mean velocity profile, and of the turbulence, depends on the
relative size of the terms on the right-hand side of
Zw=pdU/Oy_off~ (2.1)
hence the name "linear sublayer". Here u~ is the friction velocity (Zw/Qw)~,
and we recall from Section 1.8 that u,y/v=y is a typical Reynolds
number of the energy-containing turbulence. Eq. (2.2) applies for y + < 3
approximately. Support for the neglect of the Reynolds stresses in the
viscous part of the inner layer, based on experiments, has been given
recently by ECKELMANN[2.17"]. Experimental evidence shows that the
pressure gradient has little effect on the sublayer and that (2.2) applies
for flows in ducts [2.18] and pipes [2.19] also.
If one measures the instantaneous velocity profile instead of the
time-mean profile the sublayer is revealed to have a longitudinally
4 This multi-layermodel for turbulent boundary layers has been developedover the
past fifty years. Referring to a few papers only one should mention especially:for the
sublayer PRANDTL[2.10], KLIN~et al. [2.111; for the logarithmiclaw PRAND'rL[2.10, 12],
LUDWIEOand TILLMANN[2.13]; for the outer layer VONKARt,IAN [2.14], CLAUSlm[2.15],
COL~s[2.16]; for the superlayerCORRSINand KISrLERi-1.30].
54 H.-H. FERNHOLZ
where x=0.40 to 0.41 and C=5.0 to 5.2 are determined from experi-
ments s. This is one of the most important relationships for turbulent
boundary layers since, apart from the floating element, almost all
5 The possibilitythat x does in fact depend on the bulk Reynoldsnumberis discussed
in [1.15, 28].
External Flows 55
measuring techniques for skin friction are based on its validity, as well
as most semi-empirical relationships for the wall shear stress (Sect. 5.7).
For further references see, for example, PRESTON [2.33], CLAUSER[2.5],
HEAD and RECHENBERG [2.34] and PATELand HEAD [2.35]. There is
good experimental evidence that the logarithmic law is valid also in wall
layers with a streamwise pressure gradient--though the assumptions
used for its derivation did not include pressure gradients--but that its
range of validity becomes considerably smaller in an adverse pressure
gradient compared with that in zero pressure gradient (cf. the Stanford
Data Catalogue by COLES and HIRST [2.7]). Deviations from the log-
arithmic law occur, however, in wall layers with strong favorable
pressure gradients (PATEL and HEAD [2.35]). A range of validity for the
application of Preston tubes [2.33] to measure skin friction was given
by PATEL [2.36].
Analyses of the effect of the pressure gradient on the velocity distribu-
tion near the wall were performed by TOWNSEND[1.26], MELLOR[2.37]
and MCDONALD [2.38] using (1.50) or refinements. For z-----w+ ~Y, inte-
grating (1.50) and requiring compatibility with (2.3) in the limit ~ 0
gives, with z = oty/-c w,
2.,[([~1+ u2 ] ] .,y
1 = __lxIn --v + C (2.6)
T h e 14scous Superlayer
d ~ o U d y = d [peUe(6_6.)]
#eVE= ~xxSO (2.7)
where the suffix e denotes conditions at the edge, y=fi(x), and where c5"
is the displacement thickness as defined by (2.22).
Note that VE is orders of magnitude more than the propagation
velocity of the interface because the latter is highly convoluted [1.33].
A semi-empirical relationship between the entrainment velocity VE
and the velocity profile has lead HEAD [2.60] to a simple closure for the
58 H.-H. FERNI-IOLZ
A j~ U e - U dY (2.9)
where w(y/6), with w(1)= 2, is Coles' wake function, and the wake factor
B is an absolute constant for ufi/v > 2000.
Equation (2.10) applies only to constant-pressure flows (or, with a
change of B or w, to the self-preserving boundary layers discussed in the
next section). However COLIn [2.16] showed that the velocity profile
outside the viscous sublayer in almost any turbulent boundary layer
could be represented by (2.10) with w(y/6) taken as a universal function,
independent of pressure gradient or Reynolds number, and B (replaced
by H to avoid confusion) taken as a free parameter. It is important to
realize that this is merely a convenient data correlation, the effect of
pressure-gradient history being submerged in /I. Equation (2.10), on
the other hand, is based on the assumption that pressure-gradient
history is unimportant--as indeed it is in a constant-pressure flow.
COLES refers to the function w(y/8) as the "law of the wake", analogous
to the "law of the wall" function f~(u~y/v) in (1.47a), and writes, for the
velocity profile outside the viscous sublayer in any turbulent boundary
layer at low Mach number,
~q)=39~3--125q4+183qs--133q6+38~ 7 (2.12)
where in each case ~/= y/& For a different formulation of the wake law
giving dU/dy = 0 at y = ~ see Section 3.2, and for skin-friction laws derived
from the inner and outer laws see Chapter 5.
A family of velocity profiles covering the whole width of the boundary
layer has been given by THOMPSON [2.66], taking into account the
intermittency in the outer layer and the inner layer by two universal
functions. This approach is readily adaptable to account for the effects
of wall roughness, or of distributed suction or injection, by adapting
the velocity profile in the inner region. It may be regarded as a numerically
specified extension of the simple analytic data fit of COL~. The fact that
60 H.-H. FERNIIOLZ
these are merely data fits and not laws of nature is illustrated by the
quite large differences between Thompson's "intermittency function"
and the true behavior of the intermitteney. The main use of these profile
families is in checking experimental data and in "integral calculation
methods" (Chapt. 5).
the scale and the coordinate the same makes this rather confusing
and we have therefore considered local equilibrium separately (Sect. 1.8).
Local equilibrium is the more important; it is believed that almost any
inner layer has a local-equilibrium region, whereas truly self-preserving
flows are rare.
TOWNSEND [1.26, 2.74, 76, 77] and Ro'rrA [1.5, 2.75], for example,
have investigated self-preserving flows. ROTTA determined the condi-
tions under which self-preserving flows can exist by substituting self-
similar distributions of mean velocity and Reynolds shear stress.
fl = (A/zw)o~dp/dx- (6*/zw)dp/dx )
n =ylA (2.16)
= uJUe
Unfortunately, symbol H is often used for fl, as well as for the wake
parameter. Equation (2.15) is only independent of x if fl, o9 and A or
dA/dx are constant. The condition fl=constant is fulfilled by velocity
distributions of the type Ue~x" or Ue,-~expx and A---constant. ~o
equals constant can be satisfied only if the following relationship holds
This gives G = 6.5 at fl=0, while the actual value is about 6.8.
It is by no means easy to verify self-preserving boundary layers
experimentally. CLAUSER [2.15] kept the velocity profile in the outer
layer (2.8) independent of x by adjusting the adverse pressure gradient
distribution in the x-direction accordingly. In these experiments G
was nearly constant while fl varied, especially in the boundary layer
where the pressure gradient was kept high. Both HERRINGand NORBURY
[2.80] and LAUNDERand STINCHCOMBE[2.81] set up self-preserving
flows with favorable pressure gradients. But there are doubts about the
applicability of their method to determine skin friction in highly ac-
celerated flows. Therefore it is difficult to decide whether self-preserva-
tion was reached at all (cf. COLESand HmST [2.7]). BRADSHAWand
FERRISS [2.82] and BRAI)SI-tAW [2.83] have investigated two self-pre-
serving flows in adverse pressure gradients (Ue,,,x -'z55 and x -'is)
measuring mean velocity profiles and turbulence structure. They found
that Reynolds shear stress and fluctuating velocities show self-preserva-
tion in the outer layer on the average, without being able to specify
the degree of accuracy reached in their measurements of turbulence
quantities.
Similarity laws for turbulent boundary layers with suction or injec-
tion were given by COLES [2.39] and STEVENSON[2.84].
wall region than in the outer part of the boundary layer, evidently for
similar reasons. As far as downstream stability is concerned CLAUSER
[2.15] reported difficulties in establishing self-preserving boundary
layers in strong adverse pressure gradients where a small change in
pressure gradient upstream produced large changes in the flow down-
stream. Problems of this sort are often linked to secondary flows or to
the unsteadiness of a separation region, but Clauser's results cannot
be ruled out completely. A plausible argument for downstream stability
was put forward by BRAOSHAW [2.83] and a more extensive discussion
of the stability of self-preserving flow was given by TOWNSEND [1.26].
or, in more common forms, valid also for compressible boundary layers,
d
(Qov~o) = ~w(1+P) (2.21a)
dr
External Flows 67
or
dO 0 dUe(H+2_ 2 2 _ Cf (2.21b)
dx + U dx M~)=(Zw/&U,)= --~
where
0 or 62= S (QU/QcUe)[1-(U/Ue)]dY
o
(momentum loss thickness) (2.23)
H or H12=61/62 or 6*/0 (shape parameter) (2.24)
and fl was defined in (2.16). Eq. (2.21a) shows that fl measures the relative
effects of pressure gradient and surface shear stress on the rate of loss
of momentum. Since the mean pressure gradient does not appear in the
transport equations for the fluctuating quantities (see Chapt. 1), the
changes in turbulent intensity, shear stress, etc., which are characteristic
of pressure gradient flow are caused by changes of the mean velocity
profile. This again is influenced by the pressure gradient and implicitly
by the skin friction. It is this rather complex relationship between mean
velocity and turbulence quantities which causes the difficulties inherent
in the closure problem of the boundary layer equations. The structure
of the turbulent boundary layer with longitudinal pressure gradients
was investigated by SCnRAtra and KLINE [2.279].
We have seen in Subsection 2.3.1 that the logarithmic law is valid
also in boundary layers with adverse pressure gradients (Ltroww.G and
TmLMANN [2.13] if the flow is not too close to separation, where its
range of validity vanishes. It does not hold, however, for highly ac-
celerated boundary layers (PAnt, [2.36]). The outer law (2.8) is strongly
influenced by the pressure gradient, exhibiting under certain conditions
similar (self-preserving) velocity profiles with uJU and fl as parameters
(see RoTrA or CLAUSER). Typical Reynolds stress ~ , kinetic energy
and mean velocity profiles are shown in Fig. 2.5. At the wall the gradient
of the shear stress profile is determined by the following relation
~u~
u
Ue 0,8 .... -~ ~-- - .... 002
0 0,2
T0,4- [ ...... ~
0.6 y/5 1,0 0
.... 0,4" -
0,2 O.B ~ y/6 1.0
0.012 t 1
.... ~=o
d
0.008
~>0
Ue dx
O.OOl,
Ue
0.2 0,4 0.6 b y/(5 1,0
Wall Roughness
The effect of wall roughness on a turbulent shear layer was first investi-
gated in pipe flow by NmtmADSE [2.153] in 1933. Since then a wide
variety of experiments has been performed in shear flows, ranging between
the extremes of flows in natural watercourses and over aircraft wings or
External Flows 73
where ~ is called the error in origin and is the distance below the crest
of the roughness which locates the origin of y in (2.28). C + is a constant
for the specific roughness geometry.
3) Fully developed roughness (70 < kruJV). The velocity distribution
in the log-law region is now independent of viscosity because the Reynolds
number characteristic of the flow over the rough wall is large. It takes the
following form, of which (2.27) yields a special case for large k,u~/v,
layer is confined to the outer part of the composite boundary layer, and,
due to the intermittent behavior of the interface between the two layers
of the composite boundary layer, the flow switches between regions of
the "new" and the "old" turbulence. A self-preserving state is reached
about 15 6 o downstream from the first roughness element (LUXTON
[2.168]). The fast growth together with too crude assumptions about
velocity and shear stress distributions may account for the unsatisfactory
theoretical results (e.g., TOWNSEND [2.169, 170] and TAYLOR [2.171]).
RAO et al. [2.172] obtained better results from a higher-order closure
method.
In a fully developed rough-wall boundary layer the turbulent energy
production is significantly higher than in a smooth-wall boundary layer
and the shear stress has a positive gradient at the wall (ANrO~A and
LUXTON [2.173]). The reason for the nonzero shear stress gradient in the
absence of acceleration is not wholly clear. The only plausible explanation
is that U and Vvary cyclically with x as the flow passes over the roughness
elements, leading to an apparent "Reynolds stress".
The response of a turbulent boundary layer to a step change in
surface roughness from rough to smooth was investigated experimentally
by ANTONIAand LUXTON [2.87]. Measurements of the drag of some
characteristic aircraft excrescences immersed in turbulent boundary
layers were performed by GAtJDETand WINTER [2.174]. For the discus-
sion of three-dimensional and of nonuniformly distributed roughness
the reader is referred to SCHUCHaaNG [2.157].
Wavy Walls
There are three main areas where one is interested in the flow over wavy
walls: aeronautical, civil, and chemical engineering. Wave generation by
wind and the interaction of wind and sea are outside our scope here. In
aerodynamics even very small ratios of amplitude to wavelength of a
wavy wall, say 0.005, are of interest, since a surface waviness of such a
magnitude can provide a significant contribution to the total drag of a
transonic aircraft (ROGERS[2.175]). INGER and WILLIAMS[2.176] showed
that large changes in the phase of the pressure and temperature per-
turbation occur across the highly nonuniform flow of a turbulent
boundary layer if the boundary layer thickness is comparable to the
surface wavelength. As a result, a wall pressure distribution typical of
subsonic flow can exist in the presence of supersonic external inviscid
flow (see also KENDALL[2.283] and DAVIS[2.284]). The reader is referred
to ASHTON [2.177] and OWEN and THOMSON [2.178] for heat transfer
problems along wavy walls. Civil engineers are interested in the formation
of bed forms in alluvial channels. For flow over rigid wavy boundaries
76 H.-H. FERNHOLZ
the objective has been to evaluate the interactions between the boundary
layer and the external flow as reflected in pressure distribution and flow
resistance (e.g., Ho and GELHAR [2.179]). Turbulent flow in wavy pipes
was investigated by Hso and KEr,rN~v [2.180] who measured both
mean flow and fluctuating quantities. Besides boundary layers along
wavy walls as such, there exists considerable interest for gas-liquid
boundary layer flow where the phase changing interface shows a wavy
form (e.g., KOTAKE [2.181]). In this area little research on turbulent
flow has become known.
Dealing with the effect of wall curvature on turbulent shear layers one
must distinguish between curvature in the transverse (cross-sectional)
and in the longitudinal (camber) direction (the flow along a body of
revolution as investigated by WINTER et at. [2.182] contains both effects).
If for a body of constant radius the boundary layer thickness is small
compared with the body radius, then the flow can be treated as if it were
two dimensional. This assumption does not hold for long slender bodies
of revolution, and the boundary layer equations and turbulence models
must contain terms which take into account the transverse curvature
(e.g., CEBECIand SMITH [ 1.4]). Effects of transverse curvature on turbulent
boundary layers were considered by RICHMOND [2.183] and GINEVSKn
and SOLODKIN [2.184] who derived relationships for the mean velocity
and the Reynolds shear stress profile. For the flow in the vicinity of the
wall they obtained for the shear stress distribution
where rw is the radius of the transverse curvature and where the positive
sign refers to a convex and the negative sign to a concave surface. The
reader is referred to discussions of the inner law by WILLMARTHand YANG
[2.185], BRADSHAW and PATEL [2.186] and of the eddy-viscosity distri-
bution by C~BECIand SMm-I [1.4]. The boundary layer on air-drawn
artificial fibers was investigated by WALZ and MAY~R [2.187] and that
near the tail of a body of revolution by PATEL et al. [2.188].
Effects on the turbulent boundary layer due to longitudinal curvature
are much more difficult to cope with since the structure of the turbulence
is strongly affected, in apparent contrast to the case of transverse
curvature. Details are given in Subsection 3.1.1; here we give a few basic
references to work on boundary layers. PRANOTL [2.189] explained why
the flow of a turbulent boundary layer along a convex surface is more
stable than that along a concave wall, a result consistent with the in-
External Flows 77
7 6 is the thickness of the boundary layer and R the radius ofthe longitudinal curvature.
78 H.-H. FERNHOLZ
The analogy in its simplest form applies only to boundary layers with
zero pressure gradient along a smooth isothermal wall. S t denotes the
Stanton number, cf the skin friction coefficient and F the Reynolds
analogy factor (see RoTTA [2.205]). Simple transfer problems could be
solved by boundary layer methods but most of the technical flow
configurations with heat transfer are extremely complex (e.g., heat ex-
changers) and many of them in addition ill-posed problems. The Reynolds
analogy factor is fairly well documented for constant-pressure boundary
layers, or duct flows, with small temperature differences. Data for flows
in pressure gradients, and plausible methods of correlating the data, are
rarer, and in the case of large temperature differences little information
is available about the turbulence structure, which is now influenced also
by the fluctuations of density and temperature. In the face of these prob-
lems, the usual approach is to transfer first-order closure concepts, such
as those of mixing length or eddy viscosity, from the momentum equation
to the energy equation and to introduce an eddy diffusivity Yx:
U = u , f ( y u J v w, fl), (2.34)
criteria and other research results, usually obtained under quite specific
conditions. MORKOVlN[2.235] gives an admirable survey of the problems
and of current knowledge, while RESI4OTKO [2.236] outlines the most
urgent problems and suggests further work. It is generally accepted
now that transition is preceded by instability of the laminar shear layer.
Since the theory of hydrodynamic stability in shear layers, with boundary
layers as a special case, can be considered as a field of its own, it must
suffice here to refer to the monograph by LIN [1.10] and the papers by
MACK [2.237, 238], CRAm [2.239] and GASTER [2.240] to name but a
few. The more recent stability calculation schemes deal with non-
parallel flow models and take into account nonlinear effects.
In the "ideal case" of an incompressible two-dimensional laminar
boundary layer with zero pressure gradient which is free of "artifical"
disturbances due to boundary conditions, the process of instability and
"natural" transition was summarized by STUART [2.241] and MOLLO-
CHRISTENSEN [2.242]: "Beyond a certain Reynolds number travelling
waves of velocity fluctations develop and grow with downstream
distance. Next, spanwise disturbances cause local free shear layers to
form which again cause bursts of high-frequency instability fluctuations
with a very high growth rate. The flow then erupts locally into the
Emmon's spots (i.e., patches of turbulent flow in a laminar surrounding)
which further downstream cause such a confused flow that we call it
turbulent". It is obvious from this description, from the basic ex-
perimental investigations by SCHUBAUERand KLEaANOFF [2.243] and
KLEBANOFFet al. [2.244], and from the smoke photographs by KNAPP
and ROACHE [2.245] that transition does not occur instantaneously
and that both the location and the extend of the transition region in a
specific flow configuration are required. For this ideal case--even with
variable pressure gradient--the critical point (denoted by a critical
Reynolds number (Reo)crlt. as the upstream boundary) of the transition
region can be reasonably well predicted (e.g., SMITH and GAMBERONI
[2.246], JArFE et al. [2.247]). A general prediction of transition is
impossible at present as we have to cope with an ill-posed problem.
It is rarely possible to specify the details of the initial disturbances
that exist in a real flow because they arise from a number of boundary
conditions such as free-stream turbulence, surface roughness, surface
curvature, surface temperature, three-dimensional effects (secondary
flow), compressibility, noise and structural vibrations. This means
that besides the model and flow conditions, the disturbance environment,
i.e., the wind tunnel or flight path conditions, must also be known.
Furthermore, though it is generally known whether the above-mentioned
parameters have a stabilizing or a destabilizing effect on the boundary
layer, little information is available about the mutual interaction of
External Flows 81
are collateral, i.e., the velocity vector points in one direction throughout
the boundary layer. The departure from two dimensionality can be ac-
counted for by a single term in the momentum integral equation, adding
O/(x o - x)to the right-hand side of (2.21b). Here x o is the virtual origin
of the streamlines as seen in plan view. For details, see PIERCE [2.289"1.
b) The secondary flow due to the lateral curvature of the potential
flow. Because of the turning of the external streamlines the velocity
vectors in the boundary layer, subjected to the same radial pressure
gradient but of smaller size than in the outer flow, follow a turning angle
which is a function of y and is usually largest at the surface, that is, the
velocity profiles are skewed.
Three-dimensional boundary layers belong to the class of three-
dimensional "thin shear layers" where 8/tgy~> ~/dx ,,~ 8/9z when operating
on any velocity component. The boundary layer equations are presented
here for reasons of simplicity in a cartesian coordinate system where the
x-direction can be either surface orientated or be given by the direction
of the external streamlines. Then the coordinates on the wall, x and z,
are the projections of the external streamlines and their orthogonal
trajectories. As for the choice of more general, i.e., curvilinear coordinate
systems for three-dimensional boundary layers and a detailed discussion
thereof, see for example HAYES [2.290], SQUIRE [2.291], RAKICH and
MATEER [2.292], CEBECI et al. [2.293], KRAUSE [2.294] and the review
papers by EICHELBRENNER [2.295] and BLOTTNER [2.296], the latter
especially on calculation methods.
The second kind of three-dimensional shear flow along a wall is
typified by the flow in the corner of a duct or of a wing-body junction
where t3/dy, ~/az >3>d/ax (slender shear flow). The shear stress gradients in
both the y- and z-directions are important in the Reynolds equations.
The flow field as mapped by GESSNERand JONES [2.297] or EICHEL-
BRENNER [2.298] for example shows a corner vortex in the case of
turbulent flow which is called a secondary flow of Prandtl's second kind.
With corner-flow configurations we encounter one of the rare cases where
laminar and turbulent flow show completely different tendencies in their
general behavior. F r o m the considerations and results of ZAMm [2.299],
ZAMIR and YOUNG [2.300], PERKINS [2.301] and BRAGG [2.302] it ap-
pears that the secondary flow along the bisector of the corner is directed
away from the corner in laminar flow s and towards the corner in turbu-
lent flow (e.g., EICBELBRENNER and NGUYEN [2.303], GESSNER and
JONES [2.297]). This latter effect has been found to be linked to the
turbulence characteristics of the flow and has been investigated by
s It is just possible that the secondary flow that appears in the laminar case results
from a localizedinstabilityin the corner, wherethe velocityprofileshave points of inflection.
External Flows 85
~(ffu)+~(Sv)+~(~w)=o, (2.36)
86 H.-H. FERNHOLZ
_ ou Off
(2.37)
y = 6: U = U e ; W= We . (2.40)
_-~ P
- u - ~ = (vT), 8c~x2
U, (2.41)
where the index 2 denotes the coordinate normal to the wall and where
may be either 1 or 3. There are several reasons for criticism here. It is
well known that first-order closures have been successful in two-
dimensional flow only if the boundary layer was always close to equili-
brium. Neither eddy viscosity nor mixing length correlations hold for
flows which are strongly influenced by their upstream history. Equa-
tion (2.41) implies that the direction of the Reynolds shear stress vector
is always the same as the direction of the velocity-gradient vector. This is
often expressed by another auxiliary equation
It has been shown, however (cf. JOHNSTON [2.351] and EAST [2.337]),
that the angle through which the shear stress has turned is generally
intermediate between the angle of the mean velocity and the mean
velocity gradient. Johnston found angular differences between the
direction of the turbulent shear stress and the direction of the mean
velocity gradient of about 20 . There is a further group of calculation
methods where an extended version of HeRd's entrainment concept
is used for three-dimensional turbulent boundary layers. A typical
method is described in Subsection 5.7.3. Suggestions for a second-
order closure procedure have been made by BRADSHAW [2.352] and
SPALDING [2.353]. A version of the law of the wall generalized to three-
dimensional boundary layers was derived by VAN DEN BERG [2.354]
and EAST [2.355].
There exist few measurements of turbulence quantities which is one
of the reason why many of the concepts mentioned above could not be
checked properly (see for example ASI4KENAS[2.356], BRADSHAWand
TERRELL[2.343], JOHNSTON[2.351] and ELSENAARand BOELSMA[2.336])
One of the unsolved questions remains whether and how much the local
flow angle varies down to the wall (PIERCE [2.347], PIERCE and
KROMMENHOEK [2.357] and PRAHLAD [2.358]). VERMEULEN [2.339]
90 H.-H. FERNHOLZ
reported rates of change of the crossflow angle as large as 0.2 deg per
unit of u~y/v in the immediate vicinity of the wall. For measurements
in compressible turbulent boundary layers in three dimensions the reader
is referred to HALL and DICKENS [2.359], RAINBIRD [2.360], FISHER
and WEINSTEIN [2.361, 362] and COUSTEIXand MICHEL [2.363].
Three-dimensional turbulent boundary layers on rotating bodies
have been dealt with for example by PARR [2.364], STEINHEUER[2.365]
and CHAM and HEAD [2.366].
Supersonic flight of aircraft and rockets has undoubtedly been the main
incentive for the research on compressible turbulent boundary layers,
and one of the most important aims has been to determine shear stress
and heat transfer at the wall. It is not surprising therefore, to find a large
External Flows 91
+ a~oRe~ -ff~~ ~ ] /
2,5 . . . .
yxlO2 [ml
S
T 2,0'
1,5
\ , u_.
U6
PaU6
' / \\
0.5
r
T,
Fig. 2.7. Mean velocity, temperature and mass-flow profile for a Mc=7.2 boundary layer
with cooling (Tw/T,=0.51) at Reo= 2074 (taken from [2.393-1. Suffix 6 denotes conditions
at y=6
layer authors showed that the intermittency at the outer edge is more
sharply defined in a Mach 7 boundary layer, but does not start until y/6
much larger than in the low-speed case.
The foundations for fluctuation measurements in compressible
boundary layers were laid by KOVASZNAV[2.373, 375] and MORKOVIN
I-2.376, 377], but there are still only a few data sets available, mainly
obtained with hot-wire anemometers (for a recent investigation see
DEMErRIADES and LADERMAN [2.378]). In these investigations it could
be shown that the three types of disturbance fields or modes (vorticity,
entropy and sound wave), all of which obey the Navier-Stokes equa-
tions, are only weakly dependent on each other when the fluctuation
intensities are small, but interact at larger intensities when linearization
is no longer permissible. As long as the temperature fluctuations can be
assumed isobaric, the sound wave mode (though important in itself) can be
neglected compared with the rms mass-flow fluctuations. It is a lucky
coincidence that the hot wire responds mainly to stagnation-temperature
fluctuations at low overheat ratios and to mass-flow fluctuations at high
overheat ratios, so that these two fluctuation components can be
measured almost independently.
It must not be forgotten, however, that--possibly at Mach numbers
around lO---the sound mode and its interaction with the other modes
mentioned above become important, forming a barrier against the
measuring techniques applied so far, so that electron beam or laser-
Doppler measuring techniques may have to supplement the hot-wire
technique (WALLACE[2.379], HARVEYand BUSHNELL [2.380], I~YNER
[2.381], and YANTAand LEE [2.382]).
KISTLER [2.383] found that mass-flow fluctuations and total tempera-
ture fluctuations (0t) increased with Mach number in adiabatic boundary
layers "dp/dx=0 and Me ~4.7), whereas OWEN et al. [2.374] state that for
moderately cooled boundary layers (dp/dx=O, M e ~ 7 ) the mass-flow
fluctuations appear to be independent of Mach number while the 0 t-
fluctuations decrease with rising Mach number. A relationship between
the fluctuating quantities was derived by MORKOVIN [2.384] if second-
order terms can be neglected
T I T e = 1 + (;~- 1 ) r M z [1 - ( u I u o ) 2] (2.48)
For an isothermal wall, zero pressure gradient and cr--~1 VAN DRIEST
obtained the following relationship:
(2.49)
St = Fzw/Oe U 2 (2.50)
where St is the Stanton number and Fthe Reynolds analogy factor, which
unfortunately has been found to vary between 0.8 and 1.2 (CARY [2.398]
and COLEMANet al. [2.399]).
Equations (2.48) and (2.49) are often written for the total temperature
or the total enthalpy but it is little known that the form given above
permits a more realistic comparison of measured and calculated data.
Attempts have been made to extend these relations to flows with pressure
gradients and heat transfer but the results are not yet convincing (WALZ
[2.372] and KOSTER [2.400] for example). In order to gain more in-
formation about the mean velocity distribution, which must be known
to evaluate the energy integrals, it is natural to revert to the multi-layer
concept of the turbulent boundary layer, especially since the basic
features of the turbulence structure in compressible and incompressible
flow can be assumed to be approximately similar, by virtue of Morkovin's
hypothesis. Many attempts have been made to scale or transform both
abscissa and ordinate of the log law (2.3) or only one of them (COLES
[2.401], ROTTA [2.402], BARONTIand LIBBY [2.403] and TENNEKES
[2.404] to mention but a few). The best agreement with measurements
over a wide range of Mach numbers, heat transfer ratios and even
pressure gradients (GRAN et al. [2.405]) has been obtained if U*/u, is
plotted against y u J v w, where U* is the mean velocity transformed ac-
cording to VAN DRIEST [2.396]. It is, of course, important to heed the
validity range of (2.48) and (2.49) which are used in the transformation,
but the log-law is probably relatively insensitive to adverse pressure
gradients, as in incompressible flow. Deviations from the log-law must,
however, be expected if the upstream history of the boundary layer--be it
a temperature history (e.g., FELLER[2.445]) or the uncompleted transition
process (FERNHOLZ [2.406J)--still affects the boundary layer. These
deviations must not be counted as a failure of the transformation (RoTTA
[2.4071 FERNHOLZ [2.408] and KEENERand HOPKINS [2.409]).
As for the log-law it has been advantageous to use van Driest's
transformation for the outer law, too (FERNrlOLZ [-2.408]). This relation
96 H.-H. FERNHOLZ
U,*- U*
- - - - 4.70 ln(y/A*) - 6.74 (2.51)
U~r
References
2.1 G.N.ABRAMOVICH: The Theory of Turbulent Jets. (MIT Press, Cambridge, Mass,
1963)
2.2 S.F.BIRCH, D.H.RuDv, D.M.BuSHNELL (Eds.): NASA SP 321 (1972) (2 vols.)
2.3 S.N.B. MuRTHY (Ed.): Turbulent Mixing in Non-Reactive and Reactive Flows.
(Plenum Press, New York 1975)
2.4 C. DuP. DoNALDSON, A.J. BILANIN: AGARDograph 204 (1975)
2.5 F.H. CLAUSER: The turbulent boundary layer, in: Advances in Applied Mechanics 1V.
(Academic Press, New York 1956)
2.6 I. TANl: Proceedings, Computation of Turbulent Boundary Layers--1968 AFOSR-
IFP-Stanford Conference, S.J.KLINE et al. (Eds.) (Thermosciences Div., Stanford
Univ. 1969), Vol. I, p. 483
2.7 D.E.CoLES, E.A.HIRST (Eds.): Proceedinys, Computation of Turbulent Boundary
Layers--1968 AFOSR-IFP-Stanford Conference, Vol. II. (Thermosciences Divn.,
Stanford Univ. 1969)
2.8 L.TING: J. Math. Phys. 44, 353 (1965)
2.9 J.E. GREEN : RAE TR 72050 (1972)
2.10 L. PRANDTL: Ergebnisse der Aerodynamischen Versuchsanstalt zu G6ttingen (Olden-
bourg, Miinchen-Berlin 1927), vol. III, p. 1
See also W.TOLLMIEN, H. SCHLICHTING, H. GORTLER (Eds.): Ludwig Prandtl Ge-
sammelte Abhandlunqen. (Springer, Berlin-Gfttingen-Heidelberg 1961). For all
Prandtl references
2.11 S.J. KLINE, W.C. REYNOLDS, F.A. SCHRAUB, P.W. RuNSTADLER: J. Fluid Mech. 30,
741 (1967)
2.12 L. PRANDTL: Ergebnisse der Aerodynamischen Versuchsanstalt zu G6ttingen (Olden-
bourg, Miinchen-Berlin 1932), vol. IV, p. 18
2.13 H.LuDWIEG, W.TILLMANN: Ingr.-Arch. 17, 288 (1949), translated as NACA TM
1285
2.14 TH. YON KARMAN: NACA TM 611 (1931)
2.15 F.H.CLAUSER: J. Aeronaut. Sci. 21, 91 (1954)
2.16 D. CoLES: J. Fluid Mech. 1,191 (1956)
2.17 H.ECKELMANN:J. Fluid Mech. 65, 439 (1974)
2.18 H.ECKELMANN: Mitt. No. 48, Max-Planck-lnst. f/ir Str6mungsforschung (1970)
2.19 H.P. BAKEWELL: Thesis, Dept. Aerospace Sci., Penn. State Univ. (1966); see also
Phys. Fluids 10, 1880 (1967)
2.20 S.J. KLINE, P.W. RUNSTADLER: Trans. ASME 81E, 166 (1959)
2.21 E.R. CORINO, R. S. BRODKEY: J. Fluid Mech. 37, 1 (1969)
2.22 H.Z. KIM, S. J. KLINE, W. C. REYNOLDS: J. Fluid Mech. 50, 133 (1971)
2.23 A.K. GUPTA, J. LAUEER, R. E. KAPLAN: J. Fluid Mech. 50, 493 (1971)
2.24 J.M.WALLACE, H.ECKELMANN, R.S. BRODKEY: J. Fluid Mech. 54, 39 (1972)
2.25 G.R. OFFEN, S.J. KLINE: J. Fluid Mech. 62~ 223 (1974)
2.26 H. UEDA, J.O. HINZE: J. Fluid Mech. 67, 125 (1975)
2.27 S. CORRSIN: Naval Hydrodynamics, Publication 515, National Academy of Sciences
- - National Research Council (1957), Chapt. 15
2.28 J. LAUEER: NACA Rept. 1174 (1955)
2.29 P.S. KLEBANOFF: NACA Rept. 1247 (1955)
2.30 W.W. WILLMARTH, S.S. LU: AGARD-CP-93, 3.1 (1972)
2.31 R.F. BEACKWELDER, R.E. KAPLAN: AGARD-CP-93, 5.1 (1972)
2.32 J.ROTTA: Prog. Aeronaut. Sci. 2, 1 (1962)
2.33 J.H.PR~STON: J. Roy. Aeronaut, Soc. 58, 109 (1954)
2.34 M.R. HEAD, I. RECHENBERG: J. Fluid Mech. 14, 1 (1962)
External Flows 99
2.35 V.C. PATEL, M.R. HEAD: J. Fluid Mech. 34, 371 (1968)
2.36 V.C. PATEL: J. Fluid Mech. 23, 185 (1965)
2.37 G.L.MELLOR: J. Fluid Mech. 24, 255 (1966)
2.38 H. MCDONALD: J. Fluid Mech. 35, 311 (1969)
2.39 D.E.CoLES: AGARD-CP-93, 25.1 (1972)
2.40 R.J. BAKER, B.E. LAUNDER: Intern. J. Heat Mass Transfer 17, 275 and 293 (1974)
2.41 M.LANDAHL: J. Fluid Mech. 29, 441 (1967)
2.42 F.H. BARK: J. Fluid Mech. 70, 229 (1975)
2.43 O.M. PHILEIPS: J. Fluid Mech. 27, 131 (1967)
2.44 H.K. MOFFATT: Proc. AFOSR-IFP Stanford Conf. 1,495 (1969)
2.45 M.J.LIGHTHILL: Proc. AFOSR-IFP Stanford Conf. 1,511 (1969)
2.46 A . K . M . F . HUSSAIN, W.C. REYNOLDS: J. Fluid Mech. 41,241 (1970)
2.47 A . K . M . F . HUSSAIN,W. C. REYNOLDS: J. Fluid Mech. 54, 241 (1972)
2.48 L.S.G. KOVASZNAY: AGARD-CP-93, D1 (1972)
2.49 S.CORRSIN: NACA WR W-94 (1943)
2.50 A.A. ToWNSEND: Aust. J. Sci. Res. Set. A t , 161 (1948)
2.51 H. FIEDLER, M.R. HEAD: J. Fluid Mech. 25, 719 (1966)
2.52 R.E.KAPLAN, J. LAUFER: Proc. 12th Int. Congr. AppL Mech. (Springer, Berlin-
Heidelberg-New York 1969), p. 236
2.53 R.A.ANTONIA: J. Fluid Mech. 56, 1 (1972)
2.54 C.W. VANATTA: T.N. F-54, Mech. Engng. Dept., Sydney Univ. (1973)
2.55 J.C. LARuE: Phys. Fluids 17, 1513 (1974)
2.56 R.A. ANTONIA, J. D. ATgINSON : J. Fluid Mech. 64, 679 (1974)
2.57 P. BRADSHAW, J. MURLIS: Aero Rept. 74-04, Imperial College, London (1974)
2.58 R.F. BLACKWELDER, L.S.G. KOVASZNAY: Phys. Fluids 15, 1545 (1972)
2.59 O.M.PHILLIPS: J. Fluid Mech. 51, 97 (1972)
2.60 M.R. HEAD:ARC R. and M. 3152 (1960)
2.61 M.R. HEAD, V. C. PATEL: ARC R. and M. 3643 (1969)
2.62 A.A. TowNSEND: J. Fluid Mech. 26, 689 (1966)
2.63 R.E.FALco: AIAA Paper 74-99 (1974)
2.64 J. ROTTA: Ingr.-Arch. 19, 31 (1951)
2.65 J.C.RoTTA:FluidMechanicsoflnternalFlow(G.SOVRAN, Ed.)(Elsevier, Amsterdam
1967)
2.66 B.G.J.THOMPSON: ARC R. and M. 3463 (1967)
2.67 D.E.COLES: Rand Corp. R-403-PR (1962)
2.68 J.H.PRESTON: J. Fluid Mech. 3, 373 (1958)
2.69 R.L. SIMPSON: J. Fluid Mech. 42, 769 (1970)
2.70 G.D. HUFFMAN, P. BRADSI-IAW:J. Fluid Mech. 53, 45 (1972)
2.71 J.MuRus: PhD thesis, Imperial College, London (1975)
2.72 V.M.FALKNER, S.W. SKAN: Phil. Mag. 12, 865 (1931)
2.73 L.G.LoITSIANSKI: Laminare Grenzschichten (Akademie-Verlag, Berlin 1976)
2.74 A.A. ToWNSEND: J. Fluid Mech. 22, 773 (1965)
2.75 J.C. RoTTA: J. Aeronaut. Sci. 22, 215 (1955)
2.76 A.A.TowNSEND: J. Fluid Mech. 1,561 (1956)
2.77 A.A.ToWNSEND: J. Fluid Mech. 23, 767 (1965)
2.78 G.L. MELLOR, D. M. GmsoN: J. Fluid Mech. 24, 225 (1966)
2.79 J.F. NASH: AGARDogmph 97, 245 (1965)
2.80 H.J.HERRING, J.F. NORBURY: J. Fluid Mech. 27, 541 (1967)
2.81 B.E. LAUNDER, H.S. STINCHCOMBE: Rept. TWF/TN/21, Dept. Mech. Engng.,
Imperial College, London (1967)
2.82 P. BRADSHAW, D. H.FERRISS : NPL Aero Rept. 1145 (1965)
2.83 P. BRADSHAW: J. Fluid Mech. 29, 625 (1967)
100 H.-H. FERNHOLZ
2.129 G. LACHMANN (Ed.): Boundary Layer and Flow Control. London: Pergamon Press
1961
2.130 B. M. JONES: J. Roy. Aeronaut. Soc. 38, 753 (1934)
2.131 A.D. YOUNG, H. P. HORTON : AGARD-CP-4, 779 (1966)
2.132 H.P. HORTON: PhD thesis, Queen Mary College, London (1968)
2.133 H.P. HORTON : Aeronaut, Quart. 23, 211 (1972)
2.134 H.T. DELPAK: PhD thesis, Queen Mary College, London (1973)
2.135 I.TAut: Progr, Aeronaut. Sci. 5, 70. London: Pergamon Press 1964
2.136 M. GASTER: ARC R, and M. 3595 (1967)
2.137 W.R.BRILEY, H. McDoNALD: J. Fluid Mech. 69, 631 (1975)
2.138 H.J.LuGT, H.J.HAUSSL1NG: Rept. 3748, Naval Ship Res. Devel. Center (1972)
2.139 U.B. MEHTA, Z. LAVAN: J. Fluid Mech. 67,227 (1975)
2.140 I. McGREGOR: PhD thesis, Queen Mary College, London (1954)
2.141 E.A. EICHELaRENNER: Jahrbuch der WGL I959, (Vieweg, Braunschweig 1960),
p. 119
2.142 E.A. EICHELBRENNER: New York Acad, Sci. Ann. 154, 655 (1968)
2.143 D.R.CHAPMAN: NACA TN-2t37 (1950)
2.144 H.H. KoRsT: Trans. ASME 78, 593 (1956)
2,145 L.CRocCo, L. LEES: 3. Aeronaut. Sci. 19, 649 (1952)
2.146 J.H. SMITH, J. P. LAMB: Intern. J. Heat Mass Transl. 17, 1571 0974)
2.147 P. BRADSl-IAW,F. Y. F. WONG : J. Fluid Mech. 52, 113 (1972)
2.148 D.E.ABBOTT, S.J.KLINE: Trans. ASME 84D, 317 (1962)
2.149 M.ARIE, H. ROUSE: J. Fluid Mech. I, 129 (1956)
2.150 T.J. MOELLER, J. M. ROBERTSON: Modem Develop. Theor. Appl. Mech. 1,326 (1963)
2.151 J.C. LE BALLEUR, J. MIRANDE: ONERA T.P. No. 1975-16(1975)
2.152 T.J. MUELLER, H.H.KoRST, W.L. CHOW: Trans. ASME 86D, 221 (1964)
2.153 J. NIKURADSE: VDI Forschungshefl 361 (1933)
2.154 D.H.WooD, R.A.ANTON1A: Aeronaut. Quart. 26, 202 (1975)
2.155 R.L. SIMPSON: AIAA J. I1,242 (1973)
2.156 L.PRANDTL: Ergebnisse der AVA Gbttingen 4 (Oldenbourg, Miinchen 1932) (see also
Gesammelte Abhandlungen I-2.10])
2.157 H.SCHLICHT1NG: Ingr.-Arch. 7, 1 (1936) and NACA TM 823
2.158 F.R. HAMA: Trans. Soc. Nay. Architects Marine Engrs. 62, 333 (1954)
2.159 V.A. IOSELEVICH,V.I.PILIPENKO: Sov. Phys. Dokl. 213, 1266 (1973)
2.160 A.E. PERRY, W. H. SCHOF1ELD, P. N. JOUBERa': J. Fluid Mech. 37,383 (1969)
2.161 C.K.L1u, S.J. KLINE, J.P.JOHNSTON: Rept. MD-15, Mech. Dept. Engng., Stanford
Univ. (1966)
2.162 H. W. TOWNES, R.H. SABERSKY; Intern. J. Heat Mass Transf. 9, 729 (1966)
2.163 R.A.ANTONIA, D.H. WooD: TN F-71, Mech. Engng. Dept., Sydney Univ. (1974)
2.164 A.J.GRASS: J. Fluid Mech. 50, 233 (1971)
2.65 A.E.PERRY, P,N.JoUBERT: J. Fluid Mech, 17, 193 (1963)
2.166 R.A.ANTONIA, R. E. LUXTON: TN F-I, Mech. Engng. Dept., Sydney Univ. (1969)
2,167 R.A.ANTONIA, R.E. LUXTON: Advan. Geophysics 18A, 263 (1974)
2.168 R.E.LuxTON: TN F-12, Mech. Engng. Dept., Sydney Univ. (1970)
2.169 A.A. TowNsEND: J. Fluid Mech. 22, 799 (1965)
2.170 A.A.ToWNSEND: J. Fluid Mech. 26, 255 (1966)
2.171 R.J. TAYLOR: J. Fluid Mech. 13, 529 (1962)
2.172 K.S. RAO, J. C. WYNGAARD, O. R. COTr~: J. Atmosph. Sci. 31,738 (1974)
2.173 R.A. ANTONIA, R. E. LUXTON; Phys. Fluids 14, 1027 (1971)
2.174 L.GAUDET, K.G. WINTER: RAE TM Aero 1538 (1973)
2.175 K.H.ROGERS: J. Aircraft 11,382 (1974)
2.176 G.R. INGER, E.P. WILLIAMS: AIAA J. 10, 636 (1972)
102 H.-H. FERNHOLZ
2.177 G.D. ASHTON: Proc. 1972 Inst. Heat Mass Transfer (1972)
2.178 P.R. OWEN,W.R. THOMSON: J. Fluid Mech. 15, 321 (1963)
2.179 R.T.Ho, L.W.GELHAR; J. Fluid Mech. 58, 403 (1973)
2.180 S.T.Hsu, J.F. KENNEDY: J. Fluid Mech. 47, 481 (1971)
2,181 S. KOTAKE: Intern. J. Heat Mass Transl. 17, 885 (1974)
2.182 K.G. WINTER, J.C. ROTTA, K.G. SMXTH:ARC R. and M. 3633 (1970)
2.183 R. RICHMOND: PhD thesis, Caltech, Pasadena (1957)
2.184 A. S. GINEVSKII,E. E. SOLODKIN:J. Appl. Math. Mech. 22, 1169 (1958)
2.185 W.W. WILLMARTH,C. S. YANG: J. Fluid Mech. 41, 47 (1970)
2.186 P. BRADSHAW,V. C. PATEL: AIAA J. 11,893 (1973)
2.187 A. WALZ, M. MAYER: Glastechn. Ber. 39, 359 and 409 (1966)
2.188 V.C. PATEL, A. NAKAYAMA,R. DAMIAN: J. Fluid Mech. 63, 345 (1974)
2.189 L. PRANDTL: NACA TM-625 (1929)
2.190 G.I.TAYLOR: Phil. Trans. Roy. Soc. A 223, 289 (1923)
2.191 H.GtSRTLER: Nachr. dr. Wiss. Ges. G6ttingen, Math. Phys. 1, 1 (1940)
2.192 I.TANI: J. Geophys. Sci. 67, 3075 (1962)
2.193 P. BRADSHAW:J. Fluid Mech. 36, 177 (1969)
2.194 R.M.C. So, G. L. MELLOR: J. Fluid Mech. 60, 43 (1973)
2.195 R.M.C. So, G. L. MELLOR: Aeronaut. Quart. 26, 25 (1975)
2.196 B.G.J. THOMPSON: AGARDograph 97, 159 (1965)
2.197 R.N. MERONEY, P. BRADSHAW:AIAA J. 13, 1448 (1975)
2.198 H.THOMANN: J. Fluid Mech. 33, 283 (1968)
2.199 V.C. PATEL: ARC R. and M. 3599 (1969)
2.200 B.G. NEWMAN: Canadian Aero. Space J. 15, 288 (1969)
2.201 H.H.FERNHOLZ: DLR FB 66-21 (1966)
2.202 H.H.FERNHOLZ: Z. Flugwiss. 15, 136 (1967)
2.203 F.A. DvORAK: AIAA J. 11,517 (1973)
2.204 H.P.A.H. IRWIN, P. Arnot SMITH: Phys. Fluids !8, 624 (1975)
2.205 J.C. ROTTA: W~rrne Stofffibertragung 7, 133 (1974)
2.206 H.U. MEIER, D. F. GATES, R. L. P. VOISINET: AIAA Paper 74-596 (1974)
2.207 A.I. LEONT'EV: Advances in Heat Transfer (Academic Press, New York 1966),
vol. 3, p. 33
2.208 S.V. PATANKAR, D.B. SPALDING: Heat and Mass Transfer in Boundary Lasers
(Morgan-Grampian, London 1967)
2.209 A. D. GOSMAN,W.M. PUN, A. K. RUNCHAL,D. B. SPALDING,M. WOLFSHSTEIN: Heat
and Mass Transfer, in Recirculating Flows (Academic Press, New York 1968)
2.210 T. CEBECI, A. M. O. SMITH, G. MOSINSKIS:Trans. ASME 92C, 133 (1970)
2.211 R.J. FLAHERTY:J. Aircraft 11,293 (1974)
2.212 M.E. CRAWFORD,W. M. KAYS: Rept. HMT-23, Thermosciences Div., Mech. Engng.
Dept., Stanford Univ. (1975)
2.213 W.J. KELNHOFER:DFVLR-FB 70-66 (1970) (see also Trans. ASME 91A, 281 (1969)
2.214 L.H. BACK, F. CUFFEL, P. F. MASSIER: Intern. J. Heat Mass Transf. 13, 1029 (1970)
2.215 D.R. BOLDMAN,J. F. SCHMIDT, R. C. EHLERS: Trans. ASME 89C, 341 (1967)
2.216 L.H. BACK, R. F. CUEFEL:Trans. ASME 93C, 397 (1971)
2.217 D.W. KEARNEY,W. M. KAYS,R. J. MOFFAT:Int. J. Heat Mass Transl. 16, 1289 (1973)
2.218 A.P.HATTON, V.A.EuSTACE: Proc. 3rd Int. Heat Transfer Conf. Chicago 2, 34
(1966)
2.219 W.H.THIELBAHR, W.M.KAYS, R.J. MoFFAT: Rept. HMT-5, Dept. Mech. Engng.,
Stanford Univ. (1969)
2.220 F.A. DVORAK, M. R. HEAD: Intern. J. Heat Mass Transl. 10, 61 (1967)
2.221 R.E.CmLCOTT: Intern. J. Heat Mass Transf. 10, 783 (1967)
2.222 L. S. FLETCHER, D. G. BRIGGS, R. H. PAGE: AIAA Paper 70-767 (1970)
External Flows 103