Goldstein 1971
Goldstein 1971
.
RICHARD J GOLDSTEIN
I . Introduction . . . . . . . . . . . . . . . . . . . . . . . . 321
I1. Adiabatic Wall Temperature and Film Cooling Effectiveness . . .
. 326
A. Incompressible Flow . . . . . . . . . . . . . . . . . . . 326
B. High-speed Flow . . . . . . . . . . . . . . . . . . . . 327
C . Impermeable Wall Concentration . . . . . . . . . . . . . 329
111. Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 330
A . General Remarks . . . . . . . . . . . . . . . . . . . . 330
B. Two-Dimensional Incompressible Flow Film Cooling-Heat Sink
Model . . . . . . . . . . . . . . . . . . . . . . . . . 330
C . Energy Balance in the Boundary Layer . . . . . . . . . . . 331
D . Two-Dimensional Incompressible Flow Film Cooling-Other
Models . . . . . . . . . . . . . . . . . . . . . . . . 338
E. Two-Dimensional Film Cooling in a High-speed Flow . . . . 340
F. Injection through Discrete Holes-Three-Dimensional Film
Cooling . . . . . . . . . . . . . . . . . . . . . . . . 341
IV . Experimental Studies . . . . . . . . . . . . . . . . . . . . 342
A . General Remarks . . . . . . . . . . . . . . . . . . . . 342
B. Two-Dimensional Film Cooling-Incompressible Flow . . . . 351
C . Two-Dimensional Film Cooling-Compressible Flow . . . . . 361
D . Three-Dimensional Film Cooling . . . . . . . . . . . . . 369
V. Concluding Remarks . . . . . . . . . . . . . . . . . . . . . 315
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . 316
References . . . . . . . . . . . . . . . . . . . . . . . . 371
.
I Introduction
( 0 )
4
ZFR
MAINSTREAM
( C )
COOLANT /-
FIG.2. Film cooling with injection through inclined tubes: (a) injection through
single tube inclined at angle a to mainstream, (b) injection through single row of discrete
tubes inclined at angle a to mainstream.
FILMCOOLING 325
and so interrupted slots and even rows of multiple holes have been
used.
Although film cooling has primarily been used to reduce the convective
heat transfer rate from a hot gas stream to an exposed wall, it could
also be used to shield a surface from thermal radiation if the radiation
absorbtivity of the injectant is high. This can be effectively accomplished
with gas particle suspensions or a liquid coolant. In this review, however,
all the fluids are considered transparent to radiation, and the convective
and radiation heat transfer are then independent and can be treated
separately. Only the effect of film cooling on convective heat transfer
will be considered.
T h e introduction of secondary fluid into the boundary layer with
film cooling may be considered to produce an insulating layer (film)
between the wall to be protected and a gas stream flowing over it.
Alternatively the injected fluid can be considered as a heat sink that
effectively lowers the mean temperature in the boundary layer. As will
be discussed, the secondary fluid usually serves both functions. T h e
introduction of the secondary fluid into the boundary layer at a tempera-
ture lower than the mainstream and its resultant mixing with the fluid
in the boundary layer reduces the temperature in the region downstream
of injection. Note that there is usually considerable mixing of the
injected fluid and the mainstream flow downstream of injection. T h u s
the concept of a film of secondary fluid maintaining its structure for
some distance downstream and isolating the solid surface from the hot
mainstream is not strictly valid, especially with a gas coolant. Although
a separate discrete insulating film is not produced, injection of the
secondary gas can increase the boundary layer thickness and the mass
of fluid entrained into the boundary layer from the free stream. T h e
increased boundary layer thickness tends to decrease the heat transfer
to the wall. However, the increased mainstream flow entrained in the
boundary layer causes increased dilution of the secondary fluid with
a resulting decrease in its effectiveness as a heat sink. T h e significance
and relative importance of these two opposing effects will be discussed
subsequently.
This review is restricted to film cooling with both the mainstream
fluid and the secondary fluid being gases, although not necessarily
the same gas, and with a turbulent boundary layer downstream of
injection. Both two-dimensional and three-dimensional secondary flow
geometries are considered. I n addition, film cooling in compressible
flows as well as in incompressible flows is examined. Although the
emphasis is on adiabatic wall temperatures and uniform mainstreams,
the effects of heat transfer and variable free stream velocity are discussed.
326 RICHARD J. GOLDSTEIN
FLOW
A. INCOMPRESSIBLE
In most film cooling applications the heat transfer from the hot gas
to the surface to be protected is not zero. There is usually some type
of internal cooling, or, in a transient problem, the heat capacity of the
wall material itself is used to take up the heat transferred. T h e general
problem in film cooling is to predict or measure for a given geometry,
mainstream, and secondary flows the relationship between the wall
temperature distribution and heat transfer. Conversely, for a given
mainstream and allowable wall heat transfer the requirement may be
to predict the secondary flow needed to maintain the surface temperature
below some critical value.
With constant property flows the velocity distribution is independent
of the temperature field and it is convenient to use the concept of a
heat transfer coefficient. Thus
= h AT = h(Tw - Taat) (1)
where T , is the local wall temperature. A question arises as to the
datum (i.e., base or reference) temperature Tdat to use in Eq. (1).
I n the limiting case of a perfectly insulated (i.e., adiabatic) surface
the heat flux would be zero and the resulting surface temperature
(distribution) is called the adiabatic wall temperature Taw. Thus the
adiabatic wall temperature could be used as the datum temperature.
T h e heat flux with film cooling would then be
q = h(Tw - Taw) (2)
Use of Eq. (2) yields a heat transfer coefficient that is independent of
the temperature difference for a constant property flow. Note that in
the absence of blowing, Tawwould be equivalent to the free stream
temperature or in the case of high speed flow the recovery temperature.
Most film cooling studies have treated the determination of the
heat transfer coefficient and adiabatic wall temperature distribution
separately with primary emphasis on the latter. Often the heat transfer
coefficient is found to be relatively close to the value without secondary
flow, i.e., dependent primarily on the mainstream boundary layer flow.
On the other hand, the adiabatic wall temperature distribution can vary
considerably and is thus harder (and more important) to predict.
I n addition the adiabatic wall temperature is significant in that it is
the limiting value of wall temperature that can be obtained without
internal wall cooling.
Primary emphasis is given to prediction and measurement of the
FILMCOOLING 327
adiabatic wall temperature distribution even when the assumption of a
constant property flow is invalid, for example in high speed compressible
flow. For this case the use of a reference temperature or reference
enthalpy as with normal boundary layer flow in the absence of injection
will be found useful in predicting heat transfer and will permit application
of Eq. (2).
T h e adiabatic wall temperature is not only a function of the geometry
and the primary and secondary flow fields but also the temperatures
of the two gas streams. T o eliminate this temperature dependence a
dimensionless adiabatic wall temperature, q , called the film cooling
effectiveness is used. For low speed, constant property flow the film
cooling effectiveness is given by
’=
m m
law - 1,
T, - T, (3)
FLOW
B. HIGH-SPEED
For high-speed flow the film cooling effectiveness must be defined
somewhat differently. At the point of injection the wall temperature
T,, would be expected to be the recovery temperature of the secondary
flow or possibly the total temperature of the secondary flow. Far down-
stream the wall temperature might be expected to approach the main-
stream recovery temperature T , (evaluated in the absence of secondary
flow). An expression often used for compressible flow film cooling is
Note that the film cooling effectiveness T,I~ reduces to Eq. (3) when
compressibility effects can be neglected.
An alternate convenient definition of effectiveness for high speed
flows employs the isoenergetic-injection wall temperature distribution,
328 RICHARDJ. GOLDSTEIN
F Tozi
TWF-
" """""'T~ Tawi
secondary flow, T,, and Tawiis much smaller than the difference between
either of these and the expected temperature of the secondary fluid.
Then the recovery temperature T , may be used as a reasonable approxi-
mation for TaWi.
T h e enthalpy can also be used in defining the film cooling effectiveness
in high-speed flows. This would be useful at large temperature differ-
ences. T h e proper enthalpies could then be substituted (for the tempera-
tures) directly into the above expressions for effectiveness.
WALLCONCENTRATION
C. IMPERMEABLE
I t is often difficult to design test systems with walls that sufficiently
approximate adiabatic surfaces. This is particularly apparent when the
adiabatic wall temperature distribution has large gradients and with a
very high temperature mainstream. I n such cases a mass transfer
process can be used as an analogue to film cooling. Thus instead of
injecting a gas at a different temperature from the mainstream, a gas
of different composition would be injected isothermally. This might
apply particularly in a study of the effects of three-dimensional film
cooling with large density differences (between the primary and secondary
flows). T h e mass transfer analogy is also useful for two-dimensional
film cooling with large temperature (density) differences. T h e injected
gas can be completely different in composition from the free stream,
or only a tracer gas might be used in the secondary flow. If the secondary
fluid is otherwise the same as the mainstream, the use of a tracer gives
results comparable to low density (temperature) differences.
T h e mass transfer process is analogous to the heat transfer process
(neglecting thermal diffusion phenomena) if the equivalent dimensionless
parameters of the flow are the same in the two cases and if the Lewis
number is unity. (The Lewis number is the ratio of the Schmidt number
for the mass transfer process to the Prandtl number for the corresponding
heat transfer process.) T h e turbulent Lewis number as well as the
molecular Lewis number should be unity for the analogy to hold.
If the flow is sufficiently turbulent, variations in the molecular Lewis
number from unity may not play an important role, but in all cases
studied to date the value of the turbulent Lewis number should be
considered.
When using the mass transfer analogy with foreign gas injection,
the quantity analogous to the adiabatic wall temperature is the concentra-
tion of the injected gas at an impermeable wall. Although there is
some question as to the proper concentration to use, the mass fraction C
is the most widely used. Equivalent to the film coolii g effectiveness
330 RICHARDJ. GOLDSTEIN
v c = Ciw (7)
In. Analysis
A. GENERAL
REMARKS
A number of theoretical correlations and predictions have been
developed for the film cooling effectiveness. Since interest is chiefly
in turbulent film cooling, the analyses are at least partly empirical yet
often suggestive of the significant features of the flow. Much of the
interest has centered on relatively simple heat sink models in which
the added secondary flow is considered as a sink of heat at the point
of injection reducing the temperature in the downstream boundary
layer and thus the temperature of the wall. The models have been
applied to two-dimensional incompressible and compressible flows and
lately to three-dimensional film cooling. Other analyses use some of
the recent numerical techniques for predicting two-dimensional turbulent
boundary layers and separated flows to obtain predictions of film
cooling effectiveness.
In this section some of the theoretical analyses will be developed
and differences between them discussed. Comparison with experimental
results will be deferred till the next section in which the various experi-
mental studies are described.
where
,t = ( X / ~ S ) [ ( P ~ / CRe21-0.25
L~) (9)
T h e dimensionless blowing rate parameter M is the ratio of the mass
velocity of the injected fluid to the mainstream mass velocity. T h e
distance x is measured downstream from the point of injection. Tribus
and Klein compare their analysis to results for air injected into an air
mainstream. For p2 = pm, C,, = C,, , and Pr M 0.72,
C. ENERGYBALANCE
IN THE LAYER
BOUNDARY
T h e mass flowing within the boundary layer is considered to be
composed of two different fluids from two different streams-the
injected gas (ni2) and the mass which enters (is entrained into) the
boundary layer from the mainstream (&). These gases are assumed
to be well mixed in the boundary layer. At any position downstream
of injection the mass flowing per unit time (ni) in the boundary layer
is given by (see Fig. 5a)
332 RICHARDJ. GOLDSTEIN
c
E
c
m
.- x - 1
u)
! I
u)
W
z
Y
0
r
I-
I, 0.I
5
L
0.08
z
0
0.051 I I I I I I I I
2 4 6 810 20 40 60 80100
[R. J. Goldstein, G . Shavit, and T. S. Chen, J. Heat Transfer 87, 353 (1965).]
, A
CONTROL m
VOLUME
--1
--
-:A=
--
ENTRAINED MASS
-----
I I
I --t"m= m,+m,
FLOW RATE
ENTRAINED ENTHALPY
CONTRF VOLUME /kioocpmT~ Fu)w RATE
(b)
FIG. 5. Control volume when performing (a) mass and (b) energy balances.
These analyses all use essentially the same method to predict m m . They
assume a +th power turbulent velocity profile and a boundary layer
thickness given by
SIX' = 0.376 Re;?'5 (17)
to predict the entrained flow rate & . The parameter x' is the distance
from the point at which the boundary layer starts. Some of the original
derivations use a slightly different value of the constant (0.376), but for
334 RICHARDJ. GOLDSTEIN
comparison purposes they have all been recalculated using Eq. (17).
Equation (17) is valid in the absence of injection and the analyses
assume that it is still valid with injection. T h e primary difference
between the three analyses is the assumed location where the total
mass flow in the boundary layer starts (i.e., where x' = 0). For a 3th
power velocity profile starting at x' = 0, the mass in the boundary layer
from the mainstream at some distance downstream is
mm = pp,U,6 = 0.329pmUmx'
Re;?'5 (18)
Librizzi and Cresci ( 4 ) assume that the boundary layer starts at the
point of injection (x' = x) and at injection (x = 0)
m = m, + mm = m, (19)
Using fi, = pzUzs
mm/& = 0.329(~/Ms)O.~
[Rez(p2/pm)]-0.2 (20)
Putting this into Eq. (16) and using Eq. (9)'
'
1
= I + 0.32950.'
Kutateladze and Leont'ev (5) assume the boundary layer downstream
of injection grows as if it had started upstream of injection at some
distance x". I n order to calculate the distance, x", they assume that the
upstream (fictitious) boundary layer grows as a turbulent boundary
layer which started sufficiently far upstream to have a net mass flow
in it at the point of injection equal to the secondary mass flow rate.
Calculation of the mass flow rate from Eqs. ( 1 1 ) and (18) using the
above assumptions gives
m,/& = 0.329(4.01 + [)"" -1 (23)
Inserting this into Eq. (16) gives
' = 1
1
+ (C,,/C,,)[O.329(4.0I + no."- 11 (24)
' = 1
1
+ 0.249[O.' (25)
Leont'ev (9) extended the model to include the effect of film cooling
FILMCOOLING 335
on heat transfer and also the effect of rough surfaces on film cooling
performance.
Stollery and El-Ehwany (6, 7) assume the boundary layer starts at
injection (x = x') and also that the total mass flow in the boundary
layer is zero at the point of injection. Thus at x = 0,
m=0
and for x >0
m = gp,u,s
Thus using Eq. (1 1)
m, = ipP,U,6 - m2
They also indicate in their analysis the effect of foreign gas injection
(when C,, # CPm)and suggest an approach for determining the film
cooling effectiveness with a variable free stream velocity.
Since the above heat sink models assume complete mixing of the
secondary fluid in the mainstream boundary layer, their validity would
be expected to suffer when applied near the point of injection. Both
the Tribus and Klein prediction (Eq. 10) and the Stollery and El-Ehwany
prediction (Eq. 31) essentially assume no mass flow in the boundary
layer at injection yet a finite heat source. Consequently they predict
a value of infinity for the effectiveness at x = 0. T h e Librizzi and Cresci
and the Kutateladze and Leont'ev correlations, by their assumption
of rh = riz, at the point of injection yield an effectiveness of unity at x = 0.
This is a convenience which should not be overlooked. Note that three
of the correlations (Eqs. (22), (25), and (31)) approach the same predic-
tion far downstream where t is large.
It is of interest that these last three correlations (Eq. (22), (25),
and (31)) are in better agreement with experimental data, as will be
shown below, than the Tribus and Klein correlation (Eq. (8) or (lo)),
even though they use the additional major assumption that the tempera-
ture in the boundary layer is constant. That this latter assumption is
336 RICHARDJ. GOLDSTEIN
not true was known from the earliest test results of Wieghardt (10) as
is shown in Fig. 6. The temperature profile is observed to be similar
for different positions downstream of injection.
The reason the correlations (Eq. (22), (25), and (31)) work so well
follows from the unwritten law that sometimes two invalid assumptions
are better than one. Thus the assumption that the boundary layer is
unaffected by the secondary flow would indicate less flow into the
boundary layer from the mainstream (i.e., less dilution) than actually
occurs, reducing m, in Eqs. (1 1) and (16), and thus predict a larger
effectiveness than occurs. However, the mean temperature in reality is
significantly different from the adiabatic wall temperature; T being
between Tawand T, . Thus the assumption used to get Eq. (16) that
’ =
T - T,
T,- T ,
gives a lower effectiveness than the true value
Y/8,
FIG. 6. Dimensionless boundary layer temperature profiles at various positions down-
stream of injection: M = 0.74, s = 10 mm, Ta- T, = 28T,(-) e~p[-O.768(y/6~)’~/~],
(- - -) e~p[-O.785(y/6~)~]. Distance from injection in meters: ( 0 )0.05, (+) 0.2, (a) 1.0,
(v) 2.0, ( 0 ) 4.0. [K. Wieghardt, AAF Translation No. F-TS-919-RE (1946).]
FILMCOOLING 337
km/km0
=1 + 1.5 x lo-* Re2(p2W,/p,W2) sin a (37)
5 -
-
4- -
ma0
0 3-
-
8
.€
\
8
.€
2 -
I -
0.9 I I
I 2 3 4 5 6 7 8 910
FIG. 7. Ratio of boundary layer entrained mass flow rate with secondary injection kirm
to entrained mass flow rate for zero injection +zmo as a function of secondary flow injection
angle and flow rate.
Data: (0, . , r , D ) , R . J.Goldstein,G.Shavit,andT.S.Chen,J.Heat Transfer87,353(1965);
( 0 , A), K. Wieghardt, AAF Translation No. F-TS-919-RE (1946);
(v), J.P.Harnett, R. C. Birkebak, and E. R. G. Eckert, J.Heat Transfer 83,293 (1 961);
(D), R. A. Seban and L. H. Back, J. Heat Transfer 84, 45 (1962).
[R. J. Goldstein and A. Haji-Sheikh, in Japan SOC.Mech. Engr. 1967 Semi-Intern. Symp.,
213-218, Tokyo (1967).]
D. TWO-DIMENSIONAL
INCOMPRESSIBLE
FLOWFILMCOOLING-
OTHERMODELS
Hatch and Papell (ZZ) use a theoretical model for tangential injection,
in which they envision that the injected gas remains in a separate film
apart from the free stream and try to calculate the heat exchange between
this film and a turbulent boundary layer atop it. I t would be expected
that such a correlation would fit best very close to the region of injection.
Saarlas (12) uses a boundary layer model to predict film cooling
effectiveness. T h e analysis also permits an approximate calculation of
the heat transfer with film cooling and the effect of a variable mainstream
velocity.
FILMCOOLING 339
Seban and Back (13, 14, 25) use the similarity of the temperature
profiles to predict film cooling effectiveness based on a uniform eddy
diffusivity across the boundary layer. They divide the flow (tangential
injection) into three regions: a wall-jet-like flow near the slot where
the secondary flow is preserved, a mixing region, and finally a normal
turbulent boundary layer region. They observe (for M < 0.8) that the
initial region of the flow is defined by XIS < 56M1.5 (14). T h e model
uses a linearized form of the energy equation and an upstream effective
starting point of the boundary layer and predicts reasonable values of
effectiveness for tangential secondary flow injection.
T h e wall jet region is particularly significant at a velocity ratio,
U , / U , , greater than unity. Note that it is apparently the velocity ratio
that plays the key role in determining the approach to wall jet behavior
rather than the mass velocity ratio or blowing rate M . T h e significance
of the velocity ratio might be expected since it indicates (for tangential
injection) whether the secondary fluid will tend to accelerate the main-
stream ( U , > U,) or be accelerated by the mainstream ( U , < U,).
Spalding (16) proposes relations for film cooling with a tangential
slot through which a fluid similar to the mainstream fluid is injected.
Although basically empirical the model reduces to a relation similar
to the heat sink models at low blowing rates and behaves similar to
what would be expected for a wall jet at large blowing rates. He predicts,
for 6' < 7,
7)=l (40)
and for 8' 3 7,
?1 = 715' (404
where
6' = O . ~ ~ ( X / I M SRe,0.2
)~.~ + 1.41{[1 - ( U J U J ] x/s}O.~ (40b)
All of the analyses described so far use models that employ considerable
empirical input. I n an attempt to use a more analytical approach
Whitelaw and co-workers (17, 18, 19) have tried to solve the turbulent
flow boundary layer equations for film cooling with tangential injection.
They use the Patankar and Spalding (20) approach in which a mixing
length and effective turbulent Prandtl number distribution are assumed.
Although this method has many difficulties and there are questions
about its validity and accuracy, it does offer the hope of future solutions
valid for the region close to injection point as well as predicting the
heat transfer coefficient. Other references using this approach include
(21)-(23).
340 RICHARDJ. GOLDSTEIN
E. TWO-DIMENSIONAL
FILMCOOLING
IN A HIGH-SPEEDFLOW
Film cooling in a two-dimensional high-speed flow has been
analyzed (24, where the reference temperature (enthalpy) method was
combined with some of the incompressible flow analyses to obtain the
film cooling effectiveness. As mentioned above, an effectiveness using
the isoenergetic wall temperature distribution as a reference appears
to work best for compressible flows. T h e reference temperature used is
T* = T, + 0.12(Tr - Tm) (441)
and all properties in the boundary layer are evaluated at this temperature.
Thus
t* = (x/Ms)(Rez P ~ ~ P * ) (P*/P,)
- ~ . ~ ~ (42)
T h e local wall temperature T , or Tawwould be used in place of T ,
in Eq. (41) if very large temperature differences are encountered.
Corresponding to the Kutateladze and Leont’ev model (Eq. 24)
71s = {I + (Cgm/C,2)[0.33(4.00 + E*)o’8 - 1]}-’.’ (43)
When the injected fluid is the same as the mainstream fluid the
relations derived for high-speed flow are:
Kutateladze and Leont’ev model;
qis = (1 + 0.25&)-0*8 (44)
Librizzi and Cresci model;
q,, = (1 + 0.3340;8)-1 (45)
Stollery and El-Ehwany model;
qi, = 3.03(;0.8 (46)
and Goldstein and Haji-Sheikh model;
qi, = 1.9 Pr2I3(1 + 0.33[0;8p)-1 (47)
where
/3 = 1 + 1.5 x Re,(pz/p*) sin 01 (48)
I n deriving these expressions for high-speed film cooling the constant
used in the boundary layer growth equation comes from a best fit to
experimental skin friction data proposed by Schlichting (25).
Laganelli (26) uses a similar analysis based on reference properties
to predict film cooling performances in supersonic flow. His results
are similar to those given above. He also extends his results to an
FILMCOOLING 34 1
axisymmetric coordinate system. Librizzi and Cresci (4) have also
considered film cooling in an axisymmetric supersonic flow.
F. INJECTIONTHROUGH DISCRETE
HOLES-THREE-DIMENSIONAL
FILM
COOLIN G
The heat sink concept has been applied to film cooling following
injection through discrete holes (27). With such a geometry there is
little hope of getting a relatively exact analytic description of the velocity
and temperature distributions.
At relatively low mass injection rates the mass addition through a
single hole can be considered to act as a localized heat sink on the
film-cooled surface. T h e transfer process in the boundary layer is
approximated by the conduction equation, the problem being equivalent
to determining the temperature distribution in a semi-infinite solid
medium along whose surface a point heat source is moving in a straight
line with constant velocity. The medium is the mainstream gas, the
strength of the source is determined from the net enthalpy flow added
through the hole and the velocity of the source in that of the free stream,
though in the reverse direction. A major difficulty (and approximation)
is to evaluate an effective thermal conductivity or thermal diffusivity
of the mainstream.
T h e resulting temperature distribution in the mainstream is,
(49)
Along the adiabatic surface Y = 0,
rl(x' )' = +
IMUmD
8r(X/D 0.5)
exp [--0.693 (T)2 ' ]
112
Either Eq. (50) or the direct experimental results for single hole
injection can be used to predict the film cooling performance of a row
of holes using the principle of superposition. As long as the flows from
the individual jets do not interact greatly, superposition appears to
work reasonably well (28). At large blowing rates and far downstream
the jets come together and superposition of single hole results to predict
film cooling from a number of holes cannot be used.
FILM COOLING-INCOMPRESSIBLE
B. TWO-DIMENSIONAL FLOW
-
T,
T,
- e~p[-0.768(y/6,)~~/~]
Chin, Shirvin,
Hayes, Silver
(35)
@m
?,; Air 0.83
to
1.17
0.85
to
1.20
0.25
to
2.5
0.056
to
0.145
7
-
Papell, T r o u t 0.52 0.32 0.0 0.1 5
(39) to to to to
3.54 1.8 13.9 0.80
Hatch, Papell
(11)
7 He 2.45
to
0.34
to
0.018
to
0.53
to
6.5 0.67 1.55 0.57
__.c
1.14 0.76
u,_
Chin, Skirvin, 1.15 0.87 0.0512 0.0822
Hayes, Burggraf to to to to
(38) 1.13 0.887 1.026 0.152
MULTIPLE SLOTS
" Unless otherwise noted, mainstream is air. Tests also for single and double rows of holes.
to slot of 10.25 mm and 5.65 m m . ' Values assume same as Seban (33).
EXPERIMENTAL I N FILMCOOLING
STUDIES
Estimated. Also considered injection through multiple rows of discrete holes. Holes equivalent
TABLE I
Density Temp. Blowing
Free
Injection ratio, ratio, stream
rate, M =
Ref. Geometry gas' PslPm Ta/Tm PSUpIpmUm Mach NO.
Hartnett,
Eckert
Birkebak,
(46) #
7B
. .. .
Air 0.875
to
0.935
1.07
to
1.14
0.28
to
1.23
0.1 185
/M-///L
~ ~ ~
r&7m
1.00 1.15 0.9 0.1 10
&
Eckert, 0.87 1.15 0.19 0.14
to
Birkebak (31) Air 0.93
Samuel,
4% Air 1.1 I 0.78 0.25 0.040
Joubert (37) to to to to
1.28 0.905 3.18 0.085
Th
to
Haji-Sheikh (63) 2.04 1.25 0.408
He 0.3 0.31 0.01 3.01
to to to
0.4 0.39 0.02
Velocity accelerated downstream of injection slot to 2.5 and 1.6 times initial values, Values
measured velocitv data.
EXPERIMENTAL STUDIES I N FILMC O O L I N G (Costinrted)
Velocity Velocity Slot Starting length Measured Slot
ratio, urn size, Reynolds No., parameters,Effec- Reynolds Range of
U,/Um mbec mm Re,, &*is tiveness
q/Taw/Cw No., Re, r/s
0.013 30.5 35.6 0.59 ~-.lo5 0.0359 Taw 0.85 850 1.5
to to to to to to to
0.042 55.0 1.06 x lo5 0.0403 0.05 5800 17.44
taken from Seban (33). ' Based on total step height. ' Based on tunnel dimension rather than
TABLE I
Density Temp. Blowing Free
Injection ratio, ratio, rate, M = stream
Ref. Geometry gas" pJpm T2/Tm p z U z / p m U m Mach No.
yTgr/
I I- Air 0.84 to 1.12 to 0.013 to 0.095 to
Goldstein, Rask, 0.88 1.20 0.052 0.16
Eckert (42) He 0. i 2 1. I2 to 0.0022 to 0.095 to
1.20 0.0076 0.16
' k rm
~~
e$j)3
Whitelaw (18) to
He tracer 2.26
e-m
Metzger, Air 0.96 1.05 0.25 0.04
Carper, to to
Swank (54) 1.49 0.07
k 1.o I .o
T-
Whitelaw (40) Air with 0.47 0.06
to
2.24
He tracer
Carlson, *&& NP
2.76 0.363 0.5
to
1.98
0.1
to
0.5
______
* From Whitelaw (40). As temperatures are presented for only one case, assumed same for all
given boundary thickness. Nitrogen used as mainstream gas. Laminar boundary layer ahead of
EXPERIMENTAL
STUDIFSIN FILMCOOLING
(Continued)
Velocity Velocity Slot Starting length Measured Slot
ratio, Urn size, Reynolds No., parameters, Effec- Reynolds Range of
U,/Um m/sec mm Re,, S*/s q/Tax/Cwtiveness No., Re, x/s
0.015 to 33.2 to 2.54 8.0 / lo5to 0.037 to Taw 0.80 to 700 to 0.78 to
0.0605 55.0 14.0 lo5
/ 0.046 0.04 4000 31.0
0.018 to 33.2 to 2.54 8.0 lo” to 0.037 to Taw 0.80to 150to 0.78 to
0.063 55.0 14.0 lo5
/ 0.046 0.03 300 31.0
0.12 30.5 23.5 6.0 lo5 0.033 Taw 0.85 10,000 2.0
to to to to to to to
2.38 61.0 21.0 lo” 0.058 0.005 100.000 40.0
runs. ’“ Average heat transfer over section from s = 0 to x = I was measured. “ Calculated from
injection. ‘I Hole diameter. ‘ From kk’hitelaw (40).
TABLE I
Density Temp. Blowing Free
Injection ratio, ratio, rate, M = stream
Ref. Geometry gas" pr/p~ T,/Tm prUp/pmVmMach No.
I .o I .o
@m
Kacker, Air with 0.288 0.06
He to
Whitelaw (50) tracer 2.66
k
rm
-
Argor (Refrig. 12)
/,m,
urn_ Arcton 12
(Refrig. 12)
4.17 1.0 2.21
to
16.7
0.017
to
0.050
Burns,
Stollery (29) He 0.14 1.0 0.071 to 0.050
0.236
Psi,
Whitelaw (48) @% Hydrogen
Arcton 12
0.069
to
4.17
1.0 0.021
to
6.85
0.03
to
0.06
to to to to
3.07 0.42 2.99 2.5
~~ ~ ~~
Goldstein,
Ramsey Eriksen,
Eckert, (28) &
& Air
0.85 1.18 0.1
to
2.0
0.088
to
0.176
* Value for S taken form Whitelaw (40). Based on maximum possible velocity. " Average heat
gradients. Air-hydrogen combustion products form mainstream gas. Accelerated flow. y Hole
EXPERIMENTAL
STUDIES
IN FILMCOOLING(Continued)
transfer over section from x = 0 to x = I was measured. " Favorable and nonfavorable pressure
diameter.
352 RICHARDJ. GOLDSTEIN
0.2 0 -4 08 0.8 I 2
relatively good agreement with the latter two analyses. In Fig. 13,
Eq. (38) is observed to compare favorably with the film cooling effec-
tivenesses obtained for tangential injection. The same relation (Eq. 38)
when integrated to predict average values over the length of a wall (44),
gives good agreement with the average film cooling effectiveness
measured for injection through angled slots by Metzger and Fletcher (45).
I I I I I
.02
10 20 40 60 8ODO 200 400 1000
x/Ms
F 0.8
u)
y
u)
0.6
W
L
k-
0.4
LL
LL
W
I3
z-I 0.2
0
0
0
5
LL 0.01
10 20 40 60 80 100 200 400 600
- X
Ms
FIG. 11. Comparison of effectiveness for tangential slot geometry with analysis of
Kutateladze and Leont’ev. (-) Eq. (25); ( 0 ) R. A. Seban, J. Heat Transfer 82,
303 (1960); ( v ) S. Papell and A. M. Trout, NASA Tech. Note TN D-9 (1959). [S. S.
Kutateladze and A. I. Leont’ev, Thennal physics of high temperatures 1, No.2, 281-290
(1963).]
iy -~
P .lo
9 GEOMETRY BOTH REFERENCES
---
__--
TRIBUS AND KLEIN EON. 10
L l B R l t Z l AND CRESCI EQN. 22
GOLDSTIEN AND HAJI-SHEIKH EQN. 38
SOURCE Re2 M
z V REF.42 982 0.0127
4 A4
LL A REF.42 81 6 0.0155
t
0 REF.42 4444 0.0517
o REF.3 4361 0.0400
*02
I 2 4 6 810 20 40 60 80 100
FIG. 12. Film cooling effectiveness with injection of air through a porous section
including comparison with several analyses. [R. J. Goldstein and A. Haji-Sheikh, in
Japan SOC.Mech. Engr. 1967 Semi-Intern. Symp., pp. 213-218, Tokyo (1967).]
3. Slot Geometry
For tangential injection the ratio of lip thickness t to slot opening s
can influence the film cooling effectiveness particularly when the velocity
FILMCOOLING 357
I .o
.8
.6
F
.4
-
GEOMETRY BOTH REFERENCES
I
ln
ln
w .2
2
W
>
F EOUATlON 38
U
w
h.10
w
-
.o I
I 2 4 6 8 1 0 20 40 60 80 100
FIG. 13. Film cooling effectiveness with tangential injection including comparison
with analysis of Goldstein and Haji-Sheikh. [R. J. Goldstein and A. Haji-Sheikh, in
Japan SOC.Mech. Engr. 1967 Semi-Intern. Symp., pp. 213-218, Tokyo (1967).]
ratio U,/U, is near unity. This has been demonstrated using the mass
transfer analogy by Kacker and Whitelaw (50,51) and Burns and
Stollery (29). Figure 14 shows the effectiveness for different values of t / s .
For a lip thickness less than about 40% of the slot opening, the effects
are small. T h e influence of lip thickness also diminishes as the velocity
ratio U,l U , is decreased. Similar phenomena are reported by Sivasegaram
and Whitelaw (52).
T h e significant reduction of film cooling effectiveness that occurs
for large lip thicknesses is probably due to the pronounced separated
and reverse flow region at the lip edge. Under those conditions the
simple heat sink models cannot be used directly, though Eq. (25) and
358 RICHARDJ. GOLDSTEIN
I.o
0.8
0.6
’ 0.4
0.2
10 20 30 40 50 100
x /s
FIG. 14. Effect of increasing slot lip thickness on impermeable wall effectiveness for
0.126, ( A ) 0.38, ( v ) 0.63,
tangential injection at p 2 / p m m 1 and U,/U.Z= 1.07: t / s : (0)
( 0 ) 0.89, (+) 1.14, ( 0 ) 1.90. [S. C. Kacker and J. H. Whitelaw, J. Mech. Engr. Sci. 11,
22(1969).]
.o I 1 I I 1 I I I
FIG. 15. Film cooling effectiveness with injection of He through a porous section
into a mainstream of air including comparison with analysis of Goldstein and Haji-Sheikh.
[R. J. Goldstein and A. Haji-Sheikh, in Japan SOC.Mech. Engr. 1967 Semi-Intern.
Symp., pp. 213-218, Tokyo (1967).]
It should be noted that the lip thickness t was about 60% of the slot
opening in these tests. Their results were in reasonable agreement
with calculations made using the turbulent boundary-layer equations.
6. Heat Transfer
Measurements have been made of the heat transfer with film cooling
on a surface over which a uniform mainstream flowed (30) and on a
surface with a pressure gradient (46). Except at large blowing rates, they
reported that shortly downstream of injection the heat transfer coefficient
FILMCOOLING 361
reduces to the heat transfer coefficient one would expect with no blowing
(Fig. 18). Near the injection region the blowing usually causes a slight
increase in heat transfer coefficient. T h e driving force in defining the
heat transfer coefficient is the difference between the actual wall tempera-
ture and the adiabatic wall temperature. Scesa (60) and Seban and
co-workers (24,32,33) found similar results in that the heat transfer
coefficient was not significantly altered by blowing, although in these
studies the heat transfer coefficient was sometimes found to be reduced
slightly by the blowing. T h e difference in injection geometry used in (30)
as compared to that used in (33) may account for this different trend.
Metzger and co-workers (45,54) observed a slightly larger effect
(increase) on heat transfer coefficient than the other studies, particularly
at large blowing rates and close to the injection location.
C. TWO-DIMENSIONAL
FILMCOOLING
COMPRESSIBLE
FLOW
Many applications of film cooling occur in high-speed flows. Although
the incompressible flow results can often be used for compressible
flow problems, this transformation must be checked experimentally.
This is particularly true if the wall geometry is such as to produce
shock interactions in the film cooled region.
I n several reports (62-63) measurements have been made of the
adiabatic wall temperature distribution downstream of a step-down
slot in supersonic flow. Either air or helium could be injected tangentially
into an air mainstream which had a Mach number of approximately
362 RICHARDJ. GOLDSTEIN
1.0 2.0 3 .Q
VELOCITY RATIO U2IUm
FIG. 17. Impermeable wall effectiveness for air injection with He tracer as a function
of velocity ratio U , / U m . [S. C. Kacker and J. H. Whitelaw, J. Heat Transfer 90, 469
(1 968).]
three. Both heated and cooled secondary flows were used. Due to the flow
over the edge of the splitter plate (separating the secondary and main-
stream flows) there is an expansion fan, a lip shock, a separated region,
and a reattachment shock, whose magnitudes are dependent on the rate
of secondary mass addition. T h e effect of blowing rate on the flow field
is shown by schlieren photography in Fig. 19a and 19b. At the larger
secondary flow rates choking occurs in the injection slot.
T h e results were correlated, using a film cooling effectiveness based
on the isoenergetic flow conditions as described earlier. I n measuring
the isoenergetic film cooling effectiveness two test runs are required
FILMCOOLING 363
2.0
I .8 GEOMETRY
I. 2
I. 0
I I I
FILMCOOLING
D. THREE-DIMENSIONAL
I n many applications of film cooling, design considerations prevent
the use of continuous slots for introduction of the coolant. Discrete
holes may be used for injection, or a slot with discontinuities (due to
structural supports) may be used. If the mainstream is essentially
two-dimensional in the absence of injection, blowing through discrete
openings will result in a nonuniform flow across the span of the film
cooled wall. This is the type of three-dimensional film cooling that will
be discussed and reviewed in this section. T h e film cooling effectiveness
for an adiabatic wall is still of interest, but now the effectiveness is a
function of lateral position as well as downstream distance.
T h e film cooling effectiveness for injection through discrete holes is
usually considerably less than for slot injection at the same rate of
secondary flow per unit span. I n addition, as the blowing rate M is
increased past a relatively low value (perhaps M -. 0.5 for pz m pa),
the effectiveness for injection through discrete holes falls off rapidly.
These phenomena can be understood qualitatively by considering
the interaction of a nontangential jet and a mainstream. There is usually
ample room across the span for mainstream air to flow between the
individual secondary flow entrances. At low blowing rates the jets
entering the flow are quickly turned toward the surface by the main-
stream. As the blowing rate is increased, the jets penetrate into the
mainstream permitting mainstream gas to flow around and under the
entering secondary flow jets. This separates the injected fluid from
the wall and results in relatively low values of film cooling effectiveness.
At still higher blowing rates the jets penetrate further and mix more
with the mainstream.
It should be noted that the dynamic head or dynamic pressure ratio
(p2U22/pmU,2), rather than the blowing rate M is probably the parameter
to use, for a given geometry, in predicting the secondary flow for which
significant penetration of the jet (and reduced effectiveness) occurs.
T h e dynamic head ratio would be important in predicting results for
an application where the densities of the secondary and mainstream
flows are quite different from test results for approximately constant
density studies.
Another important parameter would be the geometry of the hole
through wbich the secondary fluid enters. A geometry which turns
the secondary fluid (and thus the jet momentum) towards the wall as it
enters the mainstream would be desirable in terms of optimizing the
film cooling performance.
370 RICHARDJ. GOLDSTEIN
L
TUBES
FIG. 22a. Injection section and coordinate system for a row of inclined jets. Detail
and flow field are shown for only a single jet interacting with a mainstream. [R. J. Gold-
stein, E. R. G. Eckert, V. L. Eriksen, and J. W. Ramsey, Israel J. Technol. 8, 145 (1970).]
FILMCOOLING 37 1
encountered. The data was correlated using the same parameters as
were used for film cooling through a number of two-dimensional
slots (38). Far downstream the louvers were almost as effective as slots
in protecting the surface.
Several publications (28, 71, 72) have appeared from the University
of Minnesota on measurements of film cooling with injection through
circular tubes (ending flush to the surface) inclined at various angles
to the main flow. Both single tubes and a transverse row of tubes were
used. The general flow configuration is presented in Fig. 22, which
shows qualitatively the flow of the jet entering the mainstream.
Figure 23 shows the film cooling effectiveness downstream of a single
hole through which air enters at an angle 01 of 35" to the main flow.
Even along the hole centerline (2 = 0) the effectiveness is considerably
less than what would be expected for injection through a continuous
slot as shown by the top two curves (Eq. 52). Off centerline (2# 0)
UNNEL FLOOR
FIG.22b. Flow field and coordinate system associated with laterally inclined jet
interacting with a mainstream. [R. J. Goldstein, E. R. G. Eckert, V. L. Eriksen, and
J. W. Rarnsey, Isruel J. Technol. 8, 145 (1970).]
372 RICHARDJ. GOLDSTEIN
0.9 I I I I I I I I
I
0.8 3 =21,8&-io: ASSUMING 30 S W I N G ACROSS SPAN
6 0.7
0.6
2
w
2
6 0.5
w
B 0.4
(3
z
c 0.3
2
LL
0.2
0.I
0
0 5 10 I5 20 25 30 35 40 45
DIMENSIONLESS DISTANCE DOWNSTREAM, X/D
FIG. 23. Axial effectiveness distributions for injection through a single hole at an
injection angle of 35" and M = 0.5. [R. J. Goldstein, E. R. G. Eckert, and J. W. Ramsey,
J . Eng. Power 90, 384 (1968).]
the effectiveness is even less. T h e results shown in this figure are for
M = 0.5, which is approximately the optimum blowing rate to maximize
the film cooling through a single tube at an angle of 35".
Figure 24 shows how the effectiveness varies with blowing rate at
different downstream positions. Data for a single hole and a row of
holes inclined at 35" to the mainstream are presented here. Note that
for a single row of holes the effectiveness reaches a maximum at a
blowing rate M 0.5. This could be interpreted as the blowing rate
(for p z w pm) above which the jet is no longer turned by the mainstream
to hug the wall along which it enters; above that value it increasingly
penetrates into the main flow. At higher blowing rates not only is the
effective protection per unit mass of coolant reduced, but the absolute
value of effectiveness is reduced as well.
At low blowing rates the flows of the individual jets from a row
of holes appear to be independent of one another. T h e two-dimensional
adiabatic wall temperature distribution can then be approximated by
FILMCOOLING 373
0.7
I I I
GEOMETRY
0
0 .5 1.0 1.5 2 .o
BLOWING R A T E , M
FIG. 24. Comparison of the centerline film cooling effectiveness for single hole and
multiplehole injectionat an injection angle of 35"with the flow for various blowing rates M .
[R. J. Goldstein, E. R. G. Eckert, V. L. Eriksen, and J. W. Ramsey, Israel I. Technol. 8,
145 (1970).]
GEOMETRY AY
-2
' o
.I0
b
.I5
I
c
20
d
25
c
.30
Y'I.0
-
I I
-
~ I
-I - Q=W. U.35' ~
C
0
I - I
I I I
FIG.25. Lines of constant film cooling effectiveness for single hole injection at
M = 1.0 for various angles of injection. [R. J. Goldstein, E. R. G. Eckert, V. L. Eriksen,
and J. W. Ramsey, Israel J. Technol. 8, 145 (1970).]
FILMCOOLING 375
V. Concluding Remarks
Considerable understanding of film cooling processes has developed
in the last twenty-five years. Recent important applications indicate
that there are still significant advances to be made.
Further work on numerical solutions to the equations for turbulent
flows should enhance our ability to predict two-dimensional film
cooling phenomena. Accurate predictions for film cooling injected at
an angle to the mainstream with a relatively thick splitter plate, with
high-speed flow or with large density differences may, however, prove
elusive.
For secondary flow through discrete holes or even interrupted slots,
the difficulties in predicting film cooling performance are even greater.
The resulting three-dimensional flow is not yet accessible to anything
but simplified analysis. Much work must still be done experimentally
to understand the effects of hole geometry, density differences, and the
interaction of individual jets on the adiabatic wall temperature distribu-
tion. In addition, information on the effect of the mass addition on the
local heat transfer is required.
ACKNOWLEDGMENT
Several colleagues were of great aid during the preparation of Table I and in reviewing
the manuscript for errors. Particular thanks are due to D. R. Pedersen, who also offered
invaluable assistance in preparing the figures and text for publication.
RICHARDJ. GOLDSTEIN
NOMENCLATURE
REFERENCES
I. E. R. G. Eckert and J. N. B. Livingood, NACA Rept. 1182 (1954).
2. M. Tribus and J. Klein, Heat Transfer, Symp. Univ. Mich 1952, 21 1 (1953).
3. R. J. Goldstein, G. Shavit and T. S. Chen, J. Heat Transfer 87, 353 (1965).
4. J. Librizzi and R. J. Cresci, AIAA (Am. Inst. Aeron. Astronaut.) J. 2, 617 (1964).
5. S. S. Kutateladze and A. 1. Leont’ev, Thermal physics of high temperutures 1, No. 2,
281-290 (1963).
6. J. L. Stollery and A. A. M. El’Ehwany, Intern. /. Heat Mass Transfer 8, 55 (1965).
7. J. L. Stollery and A. A. M. El-Ehwany, Intern. J. Heat Muss Transfer 10, 101 (1967).
8. R. J. Goldstein and A. Haji-Sheikh, in Japan Soc. Mech. Engr. 1967 Semi-Intern.
Symp., 213-218, Tokyo (1967).
9. A. I. Leont’ev, “Advances in Heat Transfer” (T. F. Irvine, Jr. and J. P. Hartnett,
eds.), Vol. 3, p. 33-100. Academic Press, New York, 1966.
10. K. Wieghardt, AAF Translation No. F-TS-919-RE (1946).
11. J. E. Hatch and S. S. Papell, NASA Tech. Note. 2“-130 (1959).
12. M. Saarlas, Ph.D. Thesis, Univ. of Cincinnati (1967).
13. R. A. Seban and L. H. Back, /. Heat Transfer 84, 45 (1962).
14. R. A. Seban and L. H. Back, J. Heat Trunsfer 84, 235 (1962).
15. R.A. Seban and L. H. Back, Intern. J. Heat Mass Transfer 3, 255 (1961).
16. D. B. Spalding, AIAA (Am. Inst. Aeron. Astronaut.) J. 3, 965 (1965).
378 RICHARDJ. GOLDSTEIN
17. S. C. Kacker, B. R. Pai, and J. H. Whitelaw, “Progress in Heat and Mass Transfer”
(T. F. Irvine, Jr., W. Ibele, J. P. Hartnett, and R. J. Goldstein, eds.), Vol. 2, p.
163-1 86. Macmillan (Pergamon), New York, 1969.
18. W. B. Nicoll and J. H. Whitelaw, Intern. J. Heat Mass Transfer 10, 623 (1967).
19. S. C. Kacker, W. B. Nicoll, and J. H. Whitelaw, Imperial College, Dept. of Mech.
Engr. Rep. TWF/TN/30, London, 1967.
20. S. V. Patankar and D. B. Spalding, “Heat and Mass Transfer in Boundary Layers.”
Morgan-Grampian Press, London, 1967.
21. M. Wolfshtein, Imperial College, Dept. of Mech. Engr. Rep. SF/TN/7, London,
1967.
22. E. H. Cole, D. B. Spalding and J. L. Stollery, Imperial College, Dept. of Mech.
Engr. Rep. EHTITNI11, London, 1968.
23. B. R. Pai, Imperial College, Dept. of Mech. Engr. Rep. EHT/TN/9, London, 1968.
24. R. J. Goldstein, E. R. G. Eckert and D. J. Wilson, J. Eng. Ind. 90, 584 (1968).
25. H. Schlichting, “Boundary Layer Theory,” 6th ed., p. 600. McGraw-Hill, New
York, 1968.
26. A. L. Laganelli, Intern. Heat Transfer Conf., 4th, Versailles/Paris, 1970 Pap. No.
69-IC-191 (to be presented).
27. J. W. Ramsey, R. J. Goldstein, and E. R. G. Eckert, Intern. Heat Transfer Conf.,
4th, VersaillesiParis, 1970 Pap. No. 69-IC-136 (to be presented).
28. R. J. Goldstein, E. R. G. Eckert, V. L. Eriksen, and J. W. Ramsey, Israel J. Technol.
8, 145 (1970) (cf. NASA CR-72612;also Univ. of Minnesota, Heat Transfer Lab.
Rept. H T L T R 91 (1969)).
29. W. K. Burns and J. L. Stollery, Intern. J. Heat Mass Transfer 12, 935 (1969).
30. J. P. Hartnett, R. C. Birkebak, and E. R. G. Eckert, J. Heat Transfer 83, 293 (1961).
3 1. E. R. G. Eckert and R. C. Birkebak, in “Heat Transfer, Thermodynamics and Educa-
tion, Boelter Anniversary Volume” (H. A. Johnson, ed.), p. 150-163. McGraw-Hill,
New York, 1964.
32. R. A. Seban, H. W. Chan and S. Scesa, Am. SOC.Mech. Engrs. Pap. 57-A-36 (1957).
33. R. A. Seban, J. Heat Transfer 82, 303 (1960).
34. R. A. Seban, J. Heat Transfer 82, 392 (1960).
35. J. H. Chin, S. C. Skirvin, L. E. Hayes, and A. H. Silver, Am. SOC.Mech. Engrs.
Pap. 58-A-107 (1958).
36. S. C. Kacker and J. H. Whitelaw, Intern. J. Heat Mass Transfer 10, 1623 (1967).
37. A. E. Samuel and P. N. Joubert, Am. SOC.Mech. Engrs. Pap. 64-WAIHT-48 (1964).
38. J. H. Chin, S. C. Skirvin, L. E. Hayes, and F. Burggraf, J. Heat Transfer 83, 281
(1961).
39. S. Papell and A. M. Trout, Nasa Tech. Note TN D-9(1959).
40. J. H. WhiteIaw, Aeron. Research Council, London, Current Pap. No. 942, 1967.
41. N. Nishiwaki, M. Hirata, and A. Tsuchida, in “International Developments in Heat
Transfer,” part IV,p. 675. ASME, New York (1961).
42. R. J. Goldstein, R. B. Rask, and E. R. G. Eckert, Intern. J. Heat Mass Transfer 9,
1341 (1966).
43. I. Mabuchi, JSME (Bulletin of Japanese Soc. of Mech. Engr.) 8, 406 (1965).
44. E. R. G. Eckert, R. J. Goldstein, and D. R. Pedersen, A Discussion of AIAA Pap.
69-523 by D. E. Metzger and D. D. Fletcher. (cf. Reference 45).
45. D. E. Metzger and D. D. Fletcher, AIAA ( A m . Inst. Aeron. Astronaut.) Paper 69-
523, to published in J. Aircraft (1969).
46. J. P. Hartnett, R. C. Birkebak, and E. R. G. Eckert, in “International Developments
in Heat Transfer,” Part IV, p. 682. ASME, New York, 1961.
FILMCOOLING 379
47. M. P. Escudier and J. H. Whitelaw, Intern. J. Heat Mass Transfer 1 1 , 1289 (1968).
48. B. R. Pai and J. H. Whitelaw, Imperial College, Dept of Mech. Engr. Rep. E H T
TN/A/l5, London, 1969.
49. L. W. Carlson and E. Talmor, Intern. J. Heat Mass Transfer 11, 1695 (1969).
50. S. C. Kacker and J. H. Whitelaw, J. Heat Transfer 90, 469 (1968).
51. S . C. Kacker and J. H. Whitelaw, Intern. J . Heat Mass Transfer 12, 1196 (1969).
52. S. Sivasegaram and J. H. Whitelaw, /. Mech. Engr. Sci. 11, 22 (1969).
53. S. S. Papell, N A S A Tech. Note TN D-299 (1960).
54. D. E. Metzger, H. J. Carper, and L. R. Swank, (1.Engr. Power) 90, 157 (1968).
55. C. M. Milford and D. M. Spiers, in “International Developments in Heat Transfer,”
Part IV, p. 669. ASME, New York, 1961.
56. J. G. Lucas and R. L. Golladay, Nasa Tech. Note TN N-1988 (1963).
57. J. G. Lucas and R. L. Golladay, N A S A Tech. Note TN D-3836 (1967).
58. J. J. Williams, Ph. D. Thesis, Univ. of California, Davis, California, 1969.
59. B. R. Pai and J. H. Whitelaw, Aero. Research Council, London, Paper 29928, H.M.T.
182, 1967. Also Imperial College Dept. of Mech. Engr. EHT/TN/8, London, 1967.
60. S. Scesa, Ph.D. Thesis, Univ. of California (1954).
61. R. J. Goldstein, F. K. Tsou and E. R. G. Eckert, Univ. of Minnesota, Heat Transfer
Lab. Rep., H T L T R 54, 1963.
62. R. J. Goldstein, E. R. G. Eckert, F. K. Tsou, and A. Haji-Sheikh, Univ. of Minnesota,
Heat Transfer Lab. Rept. H T L T R 60 (1965).
63. R. J. Goldstein, E. R. G. Eckert, F. K. Tsou, and A. Haji-Sheikh, A I A A ( A m .
Inst. Aeron. Astronaut.) J. 4, 981 (1966).
64. T. Mukerjee and B. W. Martin, in “Proceedings of the 1968 Heat Transfer and Fluid
Mechanics Institute” (A. F. Emery and C. A. Depew, eds.), p. 221. Stanford Univ.
Press, Stanford California, 1968.
65. K. Parthasarathy and V. Zakkay, Aerospace Research Lab. Tech. Rep., Contract
F33615-68-C-1184 Project 7064, Wright Patterson Air Force Base, Ohio, 1968.
66. R. E. Dannenberg, N A S A Tech. Note TN D-1550 (1962).
67. B. H. Lieu, U. S . Naval Ordance Lab. NOLTR No. 224, White Oak, Maryland,
1964.
68. E. Redeker and D.S. Miller, in “Proceedings of the 1966 Heat Transfer and Fluid
Mechanics Institute” (M. A. Saad and J. A. Miller, eds.), p. 387. Stanford Univ.
Press, Stanford, California, 1966.
69. L. W. Woodruff and G. C. Lorenz, A I A A ( A m . Inst. Aeron. Astronaut) J . 4, 969
(1966).
70. F. Burggraf, J. H. Chin, and L. E. Hayes, J. Heat Transfer 83, 286 (1961).
71. R. J. Goldstein, E. R. G. Eckert, and J. W. Ramsey, J. Eng. Power 90,384 (1968).
72. R. J. Goldstein, E. R. G. Eckert, and J. W. Ramsey, N A S A CR-54604; Also Univ.
of Minnesota, Heat Transfer Lab. Rep. H T L T R 82, 1968.