Sei sulla pagina 1di 82

0082-05.

fm Page 279 Wednesday, August 23, 2000 10:09 AM

5 Radiation Heat Transfer


in Electronic Equipment

5.1 INTRODUCTION
Radiation cooling of electronic components and boxes is not usually a concern to
thermal engineers. The three factors that affect cooling by radiation are the temper-
ature difference between an object and its surroundings, the surface characteristics
of the object and its surroundings, and the view that the object has of its surroundings.
Typically, the temperature difference is the delta between the device case and an
outer chassis, or an outer chassis and the walls of a room. Since radiation heat
transfer is based on this temperature difference, when the delta between a component
and its surroundings becomes high enough for radiation cooling to matter, the device
is most likely already above its maximum junction temperature. At room tempera-
ture, the total emissive power of a perfect emitter is about 450 W/m2. This is less
than 10% of the heat that convection in air can transfer.
The characteristics of the surface that covers the device or the chassis in question is
another important variable. Materials used in electronics are normally opaque to radia-
tion. Therefore, the surface characteristics are important to a depth of only about 2.5 
106 m for metals and about 0.5  103 m for nonmetals. In the range of temperatures
used for electronics, the color of the surface does not affect the radiation emittance.
The third variable is the view factor. This is the fraction of the radiation that
leaves one surface and is intercepted by another surface. This can be as high as 1.0
for the case of a sphere inside a larger sphere, or quite low, such as two plates at
an angle approaching 180°. Algebraic equations can calculate the view factor, and
this is usually done with the aid of a computer.
Although radiation is not usually a concern when trying to cool an electronic
package, the package may absorb radiation heat by being near a high-temperature
source. Such a case may occur in the engine compartment of a car, where electronic
modules are exposed to the radiative heat of hot engine components and exhaust
manifolds. Although the color of an object is not important in radiative cooling, color
is important when the object may absorb heat energy from a wide-band radiation
source. This case may occur when we expose an electronic package to the sun.

5.1.1 THE ELECTROMAGNETIC SPECTRUM


Radiation heat transfer occurs by electromagnetic waves traveling at the speed of
light, about 3  108 m/s. These waves may travel through a vacuum or through a
gas. Some gases absorb radiation and this must be considered in exact calculations.

© 2001 by CRC PRESS LLC


0082-05.fm Page 280 Wednesday, August 23, 2000 10:09 AM

Examples of gases that affect the radiation transfer are water vapor, carbon dioxide,
and air containing a large quantity of particulate matter. Standard air does not affect
the results enough to be considered in the power range and temperature range of
electronic cooling.
The electromagnetic spectrum extends from a fraction of a cycle per second
(Hz) having a wavelength of 3  104 m, to waves having frequencies greater than
3  1019 Hz and wavelengths shorter than 1011 m. At the lower-frequency end of
the spectrum are radio waves. As the frequency increases, the bands of infrared
radiation, visible light, ultraviolet radiation, X-rays, and gamma rays are encoun-
tered. Heat waves are primarily found between radio waves and visible light in the
infrared spectrum, but they extend into both regions. Figure 5.1 shows the electro-
magnetic spectrum and the portion of the spectrum that concerns radiation heat transfer.
As the temperature of an object increases the radiated heat energy increases, with
more of this energy being emitted in the visible light range. The human eye does not
respond to electromagnetic waves having a wavelength greater than 0.7 m. This
wavelength corresponds to a temperature of about 800 K or about 530°C. Therefore,
if the temperature of an object is below 800 K, we cannot see the heat.

5.2 RADIATION EQUATIONS


Radiation heat transfer involves emittance, absorptance, reflectance, and transmit-
tance at the surface of an object. In most engineering problems, transmittance is not
a factor. The reflected energy is very small and is accounted for in the emittance
and absorptance of an object in the radiation exchange.
We call a theoretical surface that absorbs all incident radiation and that also emits
all radiation a blackbody. We can find the rate of energy emitted by a blackbody by
the equation

4
qr   A1 T 1

where:

qr  rate of heat flow (W)


  Stefan-Boltzmann constant, 5.66961  108 W/m2 K4
A1  surface area (m2)
T1  absolute temperature of object 1 (K)

This equation shows that heat is radiated proportional to the fourth power of the
absolute temperature. If the blackbody is in an enclosure, and we know the temper-
ature of the enclosure, the net rate of heat transfer is found by

qr   A1 ( T 1  T 2 )
4 4

where T2 is the absolute temperature of the enclosure.

© 2001 by CRC PRESS LLC


0082-05.fm Page 281 Wednesday, August 23, 2000 10:09 AM
© 2001 by CRC PRESS LLC

FIGURE 5.1 (a) The electromagnetic spectrum. (b) The thermal radiation portion of the electromagnetic spectrum, generally
considered to be 0.1 to 100 m.
0082-05.fm Page 282 Wednesday, August 23, 2000 10:09 AM

If we are working with real surfaces, which are not perfect emitters of energy
and are called graybodies, the equation for the net rate of heat transfer is changed to

qr   A1 1 ( T 1  T 2 )
4 4

where 1 is the emittance of the graybody.


Emittance is the ratio of the emission from a graybody compared to a blackbody
at the same temperature. If both the object and the enclosure are graybodies, the net
rate of radiation transfer is found by

q r   A 1  1  1,2 ( T 1  T 2 )
4 4

where  1,2 is a dimensionless modulus, called the transfer factor, that modifies the
equation for perfect radiators to account for the emittances and relative geometries
of graybody 1 to graybody 2.
Sometimes it is convenient to express the radiation heat loss as a radiation heat
transfer coefficient, hr . We find this coefficient by manipulating the radiation heat
transfer equation:

 1 ( T 1  T 2 )
4 4
qr
h r  --------------------------------  ---------------------------------
A1 ( T 1  T 2 ) T1  T2

5.2.1 STEFAN-BOLTZMANN LAW


We call the rate that radiation is emitted per unit area at all possible wavelengths,
in every possible direction, the total hemispherical emissive power, E, which has
units of W/m2. Therefore, we can show this in the form


E  0 E  (  ) d
The spectral distribution of energy emitted by a blackbody has been well docu-
mented. In 1900, Planck1 derived the equation as

2
2hc o
I , b ( , T )  -----------------------------------------
hc o
-
 [ exp ( ---------
5
-)  1]
 kT

where:

h  6.6256  1034 J s/molecule (Planck’s constant)


k  1.3805  1023 J/K molecule (Boltzmann’s constant)
co  2.9979  108 m/s (speed of light in a vacuum)
T  absolute temperature (K)

© 2001 by CRC PRESS LLC


0082-05.fm Page 283 Wednesday, August 23, 2000 10:09 AM

The blackbody is a theoretical diffuse emitter. That is, the surface emits radiation
equally, no matter the direction. The total output from a blackbody, based on Planck’s
theories is

Eb  T 4

where we see that the term  is the Stefan-Boltzmann constant, which depends on
C1 and C2, and has the value 5.66961  108 W/m2 K4.
J. Stefan discovered the formula for radiation heat transfer in 1879. L. Boltzmann
derived the relationship theoretically in 1884. The Stefan-Boltzmann constant allows
the designer to calculate how much radiation is emitted in all directions and over
all wavelengths due to the temperature of the blackbody.

5.3 SURFACE CHARACTERISTICS


A hot object emits heat as electromagnetic energy waves. How much energy is
radiated is related to an emittance factor, . The emittance factor concerns only the
surface characteristics of an object. For example, if we paint an object, the material
we construct the object from does not affect its emittance factor. If we frost a clear
glass window, the glass can absorb heat and the temperature of the glass will increase.
This is because we have altered the surface characteristics.
The emittance factor of a surface is rated by the percentage of energy it can
radiate when compared with a perfect radiator, which we call a blackbody. A
blackbody is an object that emits 100% of its energy; therefore,   1.0. Since this
is a theoretical idea, all real objects have an emittance less than 1.0.
When the waves strike another object, the object partially absorbs and partially
reflects them. In a somewhat transparent object, the waves may be transmitted
through the object. This is shown in Figure 5.2. The radiation energy balance can
then be written

G  G,reflec
G,absorb
G,trans

and therefore,

1  



where:

G  irradiation (W/m2)
  wavelength (m)
 reflectance
 absorptance
 transmittance

The specific amount of energy distributed to these variables is determined by


the temperature difference, the surface characteristics, the object’s geometry, the

© 2001 by CRC PRESS LLC


0082-05.fm Page 284 Wednesday, August 23, 2000 10:09 AM

FIGURE 5.2 A semitransparent body displays the three characteristics of radiation heat
transfer when energy, G, is applied: absorption, G, abs; reflection, G, ref; and transmission,
G, tr.

material, and the electromagnetic wavelength. If the values of reflectance, absorp-


tance, and transmittance are averaged over the entire spectrum, the energy balance
becomes

1

Since we do not normally have transparent surfaces in electronic cooling, we


can neglect transmittance, , which results in

1 

Of course, this shows that if we know one value the other is easily found. A
perfect blackbody not only emits 100% of its energy but also absorbs 100% of the
energy it receives. We call the ability of an object to absorb heat energy absorptance,
which we represent by . The efficiency of a real object to absorb heat energy is
always less than 100%. Similar to emittance, the amount of energy the object absorbs
is based on the surface characteristics and the temperature.
Usually, dull surfaces have good emittance and good absorptance. When we
include visible light energy, dark surfaces have good emittance and absorptance. Shiny
surfaces are usually poor emitters and absorbers of radiation. In visible light the same

© 2001 by CRC PRESS LLC


0082-05.fm Page 285 Wednesday, August 23, 2000 10:09 AM

is true of light-colored surfaces. As an example, a polished metal such as aluminum has


an emittance of about 0.06. This increases to about 0.09 when the aluminum has a
commercial finish and increases to 0.10 when iridited. If the aluminum has a clear,
dull, anodized finish, the emittance increases to about 0.80.
Again, the color of the surface does not have an appreciable effect in the
temperatures that are common in electronics cooling. This is because the wavelength
of infrared heat, about 7m, is well below that of visible light. Therefore, if the
color is not visible it does not have an effect. In the temperature range of concern,
black paint will have the same characteristics as white paint. Because of the effect
of color in visible light such as the sun, we will classify solar absorptance as s.

5.3.1 EMITTANCE
The actual spectral radiation emitted and the direction of the radiation emitted by a
real surface differs from the Planck distribution. The actual emittance can also have
different values when measured in different directions and at different wavelengths.
Figure 5.3 shows the emission from a differential element into a theoretical
hemisphere. We define the total hemispherical emittance as the radiation emittance
over all possible directions and wavelengths:

E(T )
 ( T )  --------------
Eb(T )

FIGURE 5.3 Radiation from a differential area dA into a surrounding hemisphere having a
center at the differential area dA.

© 2001 by CRC PRESS LLC


0082-05.fm Page 286 Wednesday, August 23, 2000 10:09 AM

FIGURE 5.4 Spectral dependence of the spectral normal emissivity ,n of selected materi-
als. (Adapted from Incropera, F. P. and DeWitt, D. P., Fundamentals of Heat and Mass
Transfer, 3rd ed., John Wiley & Sons, 1995. With permission.)

Researchers can use these formulas to categorize the emittance of a variety of


surface finishes and materials. Figure 5.4 shows the spectral emittance dependence
on wavelength for several materials.2 If we want to find the total emissive radiation
power at any temperature, and we know the value of  (T), we can use the integration
of Planck’s distribution,

Eb  T 4

along with the equation for total hemispherical emittance

E(T )
 ( T )  --------------
Eb(T )

If we want to find the spectral emissive radiation power at any wavelength and
temperature, and we know the value of (, T), we use the formula for spectral
emissive power

C1
B , b ( , T )   I , b ( , T )  ------------------------------------
 ˙T
  e 2  1
5 C

© 2001 by CRC PRESS LLC


0082-05.fm Page 287 Wednesday, August 23, 2000 10:09 AM

with the formula for spectral hemispherical emittance

E  ( , T )
  ( , T )  -----------------------
-
E  ,b ( , T )

Table 5.1 contains the emittance and absorptance values for a variety of surfaces.

TABLE 5.1
Total Emittance and Solar Absorptance at ~27oC (~300 K)
Solar
Surface Absorptance, s Emittance,  s 
Aluminum, anodized hard 0.23 0.80 0.288
Aluminum, anodized soft 0.55 0.76 0.724
Aluminum film, evaporated 0.09 0.03 3.0
Aluminum foil, bright side 0.07
Aluminum, oxidized 0.11
Aluminum, polished 0.04
Aluminum, 6061 commercial 0.37 0.04 9.25
finish
Aluminum oxide 0.33
Brass, polished 0.10
Brass, oxidized 0.61
Carbon graphite 0.96 0.88 1.09
Chrome, black deposited on metal 0.96 0.15 6.40
Chromium, blued 0.78 0.18 4.33
Chromium, plating 0.10
Chromium, polished 0.08
Copper, polished 0.04
Copper, oxidized black 0.91 0.16 5.69
Copper, oxidized chemical 0.87 0.13
conversion
Copper, electroplated 0.47 0.03 15.7
Glass 0.90
Glass, second surface mirror 0.13 0.81 0.161
Gold, polished 0.16 0.03 5.33
Gold foil, bright side 0.07
Iron, cast oxidized 0.63
Iron, new galvanized 0.23
Iron, old galvanized 0.28
Iron oxide 0.96
Iron, polished 0.06
Magnesium 0.07
Molybdenum 0.43 0.03 14.3
Nickel, black deposited on metal 0.90 0.15 6.0
Nickel, electroplated 0.22 0.03 7.33
(Continued)

© 2001 by CRC PRESS LLC


0082-05.fm Page 288 Wednesday, August 23, 2000 10:09 AM

TABLE 5.1 (continued)


Total Emittance and Solar Absorptance at ~27oC (~300 K)
Solar
Surface Absorptance, s Emittance,  s  
Paint, aluminized 0.65
Paint, black lacquer 0.96 0.96 1.0
Paint, red 0.96
Paint, yellow 0.95
Paint, black epoxy 0.95 0.87 1.09
Paint, black silicone 0.94 0.90 1.04
Paint, white acrylic 0.26 0.90 0.289
Paint, white zinc oxide 0.95
Paint, white epoxy 0.25 0.85 0.294
Palladium 0.41 0.03 13.7
Platinum 0.33 0.03 11.0
Rhodium 0.28 0.02 14.0
Rubber 0.90 0.90 1.0
Silver, polished 0.07 0.01 7.0
Snow, fresh 0.13 0.82 0.159
Steel, stainless polished 0.17
Stainless steel, dull 0.50 0.21 2.38
Tantalum 0.59 0.02 29.5
Teflon 0.12 0.85 0.141
Water 0.98 0.90
Zinc, galvanized 0.25
Zinc, polished 0.02

5.3.1.1 Emittance Factor

The Stefan-Boltzmann law strictly applies to only blackbodies. Usually, we have


an interchange of thermal energy between two or more bodies. In real engineering
problems we commonly use a transfer factor, 1,2, to represent the interactive emit-
tances and views of the bodies. Calculating this factor involves a complex series of
integrals. Table 5.2 contains the equations needed to find a number of transfer factors.

5.3.1.2 Emittance from Extended Surfaces

Extended surfaces such as plate fin arrays are often used to increase the surface area
of a product. The additional surface area may offer a substantial improvement in
the temperature of the part. Fins also offer a benefit in radiative cooling. The channels
between the fins act as cavities that act on radiant energy and increase the emittance
of a part. The deeper the channel compared to the channel width, the more the
channel acts as a deep cavity, increasing the emittance. Bilitzky8 obtained values of
the increased emittance, Ê, for longitudinal fins of rectangular profile when q 
ÊA( T 1  T 2 ). His results are shown in Figures 5.5 through 5.9. Harper9 suggests
4 4

© 2001 by CRC PRESS LLC


0082-05.fm Page 289 Wednesday, August 23, 2000 10:09 AM

TABLE 5.2
Transfer Factors for Various Geometries
Surface Geometry 
Infinite parallel plates

1
----------------------
1 1
---- + ----  1
 1  2

Concentric spheres, surface A1 surrounded by surface A2.

1
-----------------------------------------
1 A 1
----
------1  ----  1
1 A2  2 

Concentric cylinders, surface A1 surrounded by surface A2.

1
-----------------------------------------
1 A 1
----
------1  ----  1
1 A2  2 

Surface A1 surrounded by very large surface A2 1

General case for two surfaces 1  2

© 2001 by CRC PRESS LLC


0082-05.fm Page 290 Wednesday, August 23, 2000 10:09 AM

3.0

b/z = 10.0

2.5
Effective channel emittance, E

2.0

b/z = 5.0 b/z = 2.0

1.5

b/z = 1.0

1.0

0.5

L/b = 1.0

0.0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Channel surface emissivity, 

FIGURE 5.5 Channel radiative emittance for L/b  1.0. (Adapted from References 8 and 19.)

2.0

b/z = 10.0

1.8

1.6

b/z = 2.0
Effective channel emittance, E

1.4

b/z = 5.0

1.2

1.0

b/z = 1.0

0.8

0.6

0.4

0.2

L/b = 2.0

0.0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Channel surface emissivity, 

FIGURE 5.6 Channel radiative emittance for L/b  2.0. (Adapted from References 8 and 19.)

© 2001 by CRC PRESS LLC


0082-05.fm Page 291 Wednesday, August 23, 2000 10:09 AM

1.6

1.4
b/z = 10.0

1.2
Effective channel emittance, E

b/z = 2.0

b/z = 5.0

1.0

b/z = 1.0

0.8

0.6

0.4

0.2

L/b = 5.0

0.0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Channel surface emissivity, 

FIGURE 5.7 Channel radiative emittance for L/b  5.0. (Adapted from References 8 and 19.)

1.4

1.2
b/z = 10.0
Effective channel emittance, E

1.0

b/z = 5.0

0.8

b/z = 2.0
b/z = 1.0

0.6

0.4

0.2

L/b = 10.0

0.0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Channel surface emissivity, 

FIGURE 5.8 Channel radiative emittance for L/b  10.0. (Adapted from References 8 and 19.)

© 2001 by CRC PRESS LLC


0082-05.fm Page 292 Wednesday, August 23, 2000 10:09 AM

1.2

b/z = 10.0

1.0
Effective channel emittance, E

b/z = 5.0

0.8

b/z = 2.0

0.6

b/z = 1.0

0.4

0.2

L/b = 100
0.0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Channel Surface emissivity, 

FIGURE 5.9 Channel radiative emittance for L/b  100.0. (Adapted from References 8 and 19.)

that the effective emittance of a plate fin array is found by

1   flat
 eff   flat
 ------------------
2 

where:

eff  effective emittance


flat  emittance of the flat surface

5.3.2 ABSORPTANCE
When radiation is intercepted by an object the radiation is partially reflected, partially
absorbed, and sometimes partially transmitted. If we define the spectral irradiation
an object receives at wavelength  as G, then in units of W/m2 the energy balance
has the form
G  G,reflec
G,absorb
G,trans
Usually, the object is opaque and we can neglect G,trans. Reflected radiant energy
does not affect the object. The irradiation that is absorbed will raise the energy level
of the object, causing an increase in temperature.
Absorptance, like directional vectors and wavelength, affects emittance. The
formulas for working with absorptance are similar to those previously discussed for
emittance. The spectral directional absorptance, ,θ (, , ), of a surface is the ratio

© 2001 by CRC PRESS LLC


0082-05.fm Page 293 Wednesday, August 23, 2000 10:09 AM

of the intensity of the radiation absorbed at the individual wavelength, , and in the
direction of  and  to the intensity of the radiation absorbed by a blackbody at the
identical wavelength. Since the effect of temperature on absorptance is exceedingly
small in real applications, we neglect the term T. Therefore, the formula for spectral
directional absorptance is

I  ,i,abs ( , ,  )
 , (  ,  ,  )  -----------------------------------
I  ,i ( , ,  )
where:

, (, , )  spectral directional absorptance (dimensionless)


  wavelength (m)
  zenith angle (rad)
  azimuth angle (rad)
I  spectral intensity (W/m2)

As we learned about emittance, absorption in real objects may occur at different


intensities depending on the wavelength and directional vector of the radiation. To
use the absorptance for real engineering problems, we define a term representing
the directional average of the spectral irradiation, which we call the spectral hemi-
spherical absorptance, ():

G , abs (  )
 (  )  ----------------------
G (  )

5.3.3 REFLECTANCE
The definition and the type of reflectance may take one of several forms. This is
because reflectance is a bidirectional phenomenon. That is, the reflectance depends
not only on the direction of the irradiation, but also the direction of the reflected
radiation.10 Figure 5.10 shows the difference between diffuse and specular reflection.

FIGURE 5.10 Two types of reflected radiation from a plane surface: diffuse reflection and
specular reflection.

© 2001 by CRC PRESS LLC


0082-05.fm Page 294 Wednesday, August 23, 2000 10:09 AM

Because reflectance is such a small part of heat transfer in electronics cooling, we


will present only the most basic form, which is an integrated average over a
hemisphere.

5.3.3.1 Specular Reflectance

The spectral directional reflectance, , (,,), is the fraction of the spectral incident
radiation in the direction  and , reflected by the surface. Therefore,

I , i, reflec ( , ,  )
 , (  , , )  ---------------------------------------
I  ,i ( , ,  )

The spectral hemispherical reflectance, designated (), is the fraction of the


spectral irradiation reflected:

G , reflec (  )
 (  )  -------------------------
G (  )

The total hemispherical reflectance, , is then

G reflec
 -------------
G

In most occurrences of reflectance, is a combination of diffuse and specular.


Occasionally, the reflectance is specular, resembling a mirror-like surface.

5.3.4 TRANSMITTANCE
Although we display the formula for transmittance here, note that this is not one
value but may change with depth and with the wavelength. For example, glass and
water are transparent at the wavelengths of visible light but become opaque at longer
wavelengths. The hemispherical transmittance of a material can be found by

G , trans (  ) G trans
  -----------------------
- and  -----------
-
G (  ) G

5.4 VIEW FACTORS


The view factor, F1,2, is the portion of the radiation leaving surface A1 that surface A2
intercepts. If both surfaces are blackbodies, the energy transfer can be from surface A1
to surface A2, F1,2, or from surface A2 to surface A1, F2,1. The net rate of radiation between

© 2001 by CRC PRESS LLC


0082-05.fm Page 295 Wednesday, August 23, 2000 10:09 AM

FIGURE 5.11 The area of radiation importance of some complex structures can be found
by measuring the projected area of the irregular surface.

FIGURE 5.12 Geometry and nomenclature used for deriving the shape factor formula of
two surfaces.

two blackbodies can be found by determining the radiation from either of the surfaces
to the other surface. After determining the radiation we replace its emissive power
with the difference between the emissive powers of the two surfaces. Since the result
does not depend on which emitting surface is chosen, the surface that is easiest to
calculate is usually selected. If the surface is irregular, such as that shown in Figure
5.11, the projected area is used, that is, the area that would result if a string were
drawn across the surface. Figure 5.12 shows the geometry for derivation of the view
factor.
To learn how much radiation is interchanged, we use differential surface dA1 and
dA2. To find the radiation surface dA2 intercepts, we evaluate

cos  1 cos  2 dA 2
dq 1,2  E b1 dA 1  ------------------------------------
-
 L
2 

© 2001 by CRC PRESS LLC


0082-05.fm Page 296 Wednesday, August 23, 2000 10:09 AM

where:

Eb1  intensity of radiation from surface dA1


cos1  area of surface dA1 seen from surface dA2
cos2  area of surface dA2 seen from surface dA1
L  distance between surfaces

The net rate of radiation heat transfer between surface dA1 and dA2 is then

cos  1 cos  2 dA 1 dA 2
dq 1, 2  ( E b1  E b2 )  ---------------------------------------------
-
 L
2 

To find the radiation transferred between the surfaces, we integrate the term in
the right parentheses over both surfaces; therefore,

cos  1 cos  2 dA 1 dA 2
dq 1, 2  ( E b1  E b2 )    ---------------------------------------------
-
A1 A2 L
2 

We can write this double integral for radiation from dA1 to dA2 as A1F1,2.
Researchers have derived many complex equations for determining the double inte-
gral and the view factor. Table 5.3 shows view factor equations and illustrations of
the geometries. Howell11 gathered the most useful view factors into a single volume.
In cases where the surfaces are significantly different from those in the tables, it is
best to use a computer program that can calculate the view factors of multiple
complex surfaces, such as TRASYS®.12*

5.4.1 CALCULATION OF ESTIMATED DIFFUSE VIEW FACTORS


Since the calculation of the double integrals is so complex, other methods estimate
the view factor and may yield acceptable accuracy. These estimated calculation
methods do not apply to specular reflection, only to diffuse view factors.
The contour integral method is based on the application of Stokes’ theorem to
reduce surface integrals to contour integrals. By superpositioning and reciprocity
the view factor may be expressed as a function of simple shapes view factors that
are already known by a method called Diffuse View Factor Algebra. If we have an
infinitely long triangle enclosure whose surfaces are flat or convex, we may express
the following system of six equations with six unknown variables:
N

 F 1, 2  1.0 i  1,2,3
j1

A i F i, j  A j  F j,i i, j  1,2,3

* TRASYS is a registered trademark of Lockheed Martin.

© 2001 by CRC PRESS LLC


0082-05.fm Page 297 Wednesday, August 23, 2000 10:09 AM

TABLE 5.3
View Factors for Two- and Three-Dimensional Geometries
Ref.

Infinite parallel surfaces of


2
same width F 1, 2  F 2, 1  1
(h  w )  (h  w) 13

Infinite parallel surfaces of different 1


F 1, 2  ------
b
-
widths on the same centerlines 2 --a- 14

2 2
 b c
4
---
---
c b
---  ---
4
a a a a

1 h  2
Infinite 90° surfaces of different h
-  1
 ---
F 1, 2  ---  1
---
 w 
widths with a common edge w - 10
2

(Continued)

© 2001 by CRC PRESS LLC


0082-05.fm Page 298 Wednesday, August 23, 2000 10:09 AM

TABLE 5.3 (continued)


View Factors for Two- and Three-Dimensional Geometries
Ref.

Infinite surface to a row of cylinders a 2 0.5 13


F 1, 2  1   1   --- 
  b 
2 0.5
 1   a--- 
 a

--- tan
1 
--------------------
b
 b   a 2 
  --b- 

n
For n rows of in-line cylinders F 1, n rows  1  ( 1  F 1, 2 )

Infinite surfaces of same width with 10



one common edge, at an angle F 1, 2  F 2, 1  1  sin ---
2

Infinite triangle A1
A2  A3 15
F 1, 2  ---------------------------------
2 A1
-

© 2001 by CRC PRESS LLC


0082-05.fm Page 299 Wednesday, August 23, 2000 10:09 AM

TABLE 5.3 (continued)


View Factors for Two- and Three-Dimensional Geometries
Ref.

Rectangle at an angle to an infinite 1  cos  15


surface F 1, 2  ----------------------
2

Right triangle to congruent triangle, 1 11


F 1, 2  ---  1  -------
1
common short edge
2 3

Infinite concentric cylinders for F 1, 2  1 10


concentric spheres r1  A1 and r1
r2  A2 F 2, 2  1  ----
r2
r1
F 2, 1  ---
r2
-

Exterior of infinite cylinder to 1 15


interior of concentric semicylinder F 1, 2  --2-

(Continued)

© 2001 by CRC PRESS LLC


0082-05.fm Page 300 Wednesday, August 23, 2000 10:09 AM

TABLE 5.3 (continued)


View Factors for Two- and Three-Dimensional Geometries
Ref.

Interior surface of infinite 15


2 2 0.5
semicylinder to itself when a F 1, 1  1  ----  1   r----1 
r----1 sin 1 r----1
   r 2  r2 r2
concentric cylinder is obstructing

Interior of infinite semicylinder 1 to 15


2 r 1 2 0.5  r 1
F 1, 2  ----  1   --- 1 r  r -1
interior of semicylinder 2 when
  -
---- sin  ----1   ---
r 2
 r 2   r 2  r 2
concentric parallel cylinder 3
is present

Infinite parallel cylinders, same 10


1 s 2
  1
diameter F 1, 2  F 2, 1  ----   1
-----
  2r  
 1  s

sin  -----------------   1
-----
1

 s  2r

-----
 1 2r 

© 2001 by CRC PRESS LLC


0082-05.fm Page 301 Wednesday, August 23, 2000 10:09 AM

TABLE 5.3 (continued)


View Factors for Two- and Three-Dimensional Geometries
Ref.

Inside surface of a right circular h 15


h 2 0.5
cylinder to itself F 1, 1   1
----- 
2r   1
-----
 2r 

Base of a cylinder to inside surface h 11


h 2 0.5
F 1, 2  2 -----  1
 ----- 
h
of cylinder  -----
2r   2r 2r

Radius in cylinder base to inside r 2 11


h 2
F 1, 2  ---  1   ----2   ----
1
surface of cylinder 2   r1   r 1
r 2 h 2 2 r 2 0.5

 1
 ----2
 ----   4  ----2 
  r 1  r 1   r 1 

(Continued)

© 2001 by CRC PRESS LLC


0082-05.fm Page 302 Wednesday, August 23, 2000 10:09 AM

TABLE 5.3 (continued)


View Factors for Two- and Three-Dimensional Geometries
Ref.

Annular ring on cylinder base or top  11


1 h r 2 h 2 0.5
-  ----  4  ----2
 ---- 
to inside of right circular cylinder 1
F 1, 2  ---  1
----------------------
2 r 2 r   r   r  
 ----2  1  1 1 1
  r 1

r 2 h 2 2 r 2 0.5

  1
 ----2
 -----1    4  ----2 
  r 1  R 1   r 1


FIGURE 5.13 A depiction of Hottel and Sarofim’s16 “string rule” for shape factors of two-
dimensional configurations.

By solving this equation, we find the six unknown variables of Fi,j. We call a
variation of diffuse view factor algebra the cross-string method, or Hottel and
Sarofim’s rule.16 To use this method, we must first measure the sum of the lengths
of the strings attached to the ends of lines that represent the surfaces. We then
subtract the sum of the lengths of the strings that are not crossed when attached to
the same ends. Next we divide this value by two. The result will be the view factor
between the two surfaces. Figure 5.13 shows the lines used for Hottel and Sarofim’s
rule; therefore,

1
F 1, 2  -------- [ AD
BC  AC  BD ]
2L 1

© 2001 by CRC PRESS LLC


0082-05.fm Page 303 Wednesday, August 23, 2000 10:09 AM

5.5 ENVIRONMENTAL EFFECTS


Some electronic component chassis are exposed to the earth’s ambient environment.
Designers often seal these enclosures to prevent the hazards associated with rain
and airborne contaminants. Since these boxes are not inside a building, they are
exposed to solar radiation and the radiation effects of the atmosphere.
Another class of component is mounted on space vehicles. These electronic boxes
are exposed to solar radiation flux unattenuated by our atmosphere. In earth orbit, a
space vehicle can be exposed to radiation from the sun, the moon, and the earth. During
interplanetary travels the component enclosure may be exposed to planetary radiation
plus solar radiation. Much of this radiation is in the ultraviolet and the visible light
range, but a significant amount contains wavelengths in the infrared region.

5.5.1 SOLAR RADIATION


We call the average heat flux from solar radiation the solar constant, designated G0.
The solar constant is measured unattenuated by the earth’s atmosphere at a mean distance
from the sun of 1.495  1011 m, a distance called one astronomical unit. Under
these conditions the solar constant has a mean value of 1353 W/m2. Because the
earth moves in an elliptical orbit around the sun, the solar constant varies from about
1310 W/m2 in June to 1400 W/m2 in January.
When we measure the solar radiation on the earth’s surface, scattering and
absorption by gas molecules and dust have reduced the solar constant from G0 to Gs
(see Figure 5.14). The solar flux at the earth’s surface is then the sum of direct and

FIGURE 5.14 Schematic of clear sky absorption and scattering of solar radiation, and the
components of irradiation incident on the earth’s surface.

© 2001 by CRC PRESS LLC


0082-05.fm Page 304 Wednesday, August 23, 2000 10:09 AM

scattered radiation, called Gs, and has a value of about 1300 W/m2. The actual value
varies by time of day, time of year, and atmospheric conditions. Large nearby
surfaces such as buildings can also affect the value of Gs. When engineers require
exact values, they should consult NASA publications.
Table 5.1 shows values for the ratio of the absorptance of a surface to its
emittance, /. At infrared wavelengths color is not important. In the visible light
region color is very important. Therefore, the steady-state temperature of a surface
is based on the ratio of solar absorptance, , to the infrared emittance, . A high
ratio of / shows that the surface will absorb a high level of solar energy, resulting
in a higher temperature in space. Therefore, we can control the temperature of a
surface in space by controlling the ratio /.
A surface in space absorbs solar radiation and radiation from the planets and
emits radiation to the vacuum of space, which is at absolute zero, 273.15°C. If
we call the internal energy developed by an electronic box qi then we can express
the total of the absorbed energy and the internal energy as

sG0 As
qi

Therefore,

sG0 As
qi   AsT 4

Rearranging the equation to find the temperature of a surface in space exposed to


the sun, we have

s G 0 A s 0.25
T   -----------------
-
  A s 

5.5.2 ATMOSPHERIC RADIATION


Dust particles and gas molecules, mostly H2O and CO2, absorb and scatter solar
radiation and the radiation from the earth’s surface (see Figure 5.15). Because of
the absorption of radiation at different wavelengths the radiation within the atmo-
sphere does not follow blackbody distribution and is therefore quite complex. Nev-
ertheless, we can estimate the emittance of the atmosphere that corresponds to the
effective temperature of the air at the earth’s surface by

J sky   sky  T e
4

An approximate correlation for clear sky emittance recommended by Brunt17


accounts for the effects of H2O, CO2, and dust particles:

P H2O 0.5
 sky  0.55
1.8  ----------
 P 

© 2001 by CRC PRESS LLC


0082-05.fm Page 305 Wednesday, August 23, 2000 10:09 AM

FIGURE 5.15 Solar spectra outside the earth’s atmosphere and at sea level. (Adapted from
References 7 and 20.)

where:

P H 2 O  partial pressure of water vapor in atmosphere


P  total atmosphere pressure

Berdahl and Fromberg18 correlated equations to account for a variety of clear


sky emittance measurements made in the U.S. For a nighttime sky,

nite sky  0.741


0.0062ΤDP

and for a daytime sky,

day sky  0.727


0.0060ΤDP

where TDP  dewpoint temperature (oC).

REFERENCES
1. Planck, M., The Theory of Heat Radiation, Dover Publishers, New York, 1959.
2. Incropera, F. P. and DeWitt, D. P., Fundamentals of Heat and Mass Transfer, 3rd ed.,
John Wiley & Sons, New York, 1990.
3. Touloukian, Y. S. and Ho, C. Y., Eds., Thermophysical Properties of Matter, Vol. 7,
Plenum Press, New York, 1972.

© 2001 by CRC PRESS LLC


0082-05.fm Page 306 Wednesday, August 23, 2000 10:09 AM

4. Mallory, J. F., Thermal Insulation, Van Nostrand Reinhold, New York, 1969.
5. Gubareff, G. G., Janssen, J. E., and Torborg, R. H., Thermal Radiation Properties
Survey, Minneapolis-Honeywell Regulator Co., Minneapolis, MN, 1960.
6. Kreith, F. and Kreider, J. F., Principles of Solar Energy, Hemisphere, New York, 1978.
7. Mills, A. F., Heat and Mass Transfer, Irwin, Chicago, 1995, 1148.
8. Bilitzky, A., The Effect of Geometry on Heat Transfer by Free Convection from a
Fin Array, M.S. thesis, Ben-Gurion University, Israel, 1986.
9. Harper, C. Handbook of Electronic Packaging, McGraw-Hill, New York, 1995.
10. Siegel, R. and Howell, J. R., Thermal Radiation Heat Transfer, 2nd ed., McGraw-
Hill, New York, 1981.
11. Howell, J. R., A Catalog of Radiation Configuration Factors, McGraw-Hill, New
York, 1985.
12. TRASYS User’s Manual, Version P22, Manual prepared for NASA Johnson Space
Center by Lockheed Engineering Manual Services, April 1988. National Aeronautics
and Space Administration, Washington, D.C., 1988.
13. Hottel, H. C., Radiant Heat Transmission Between Surfaces, Separated by Non-
Absorbing Media, Trans. ASME, FSP-53-196, 265, 1931.
14. Wong, H. Y., Handbook of Essential Formulae and Data on Heat Transfer for
Engineers, Longman, London, 1977.
15. Martin, J. G. and Muriel, M. J. B., Radiation Heat Transfer, in Handbook of Applied
Thermal Design, Guyer, E. C., Ed., McGraw-Hill, New York, 1989.
16. Hottel, H. C. and Sarofim, A. F., Radiative Transfer, McGraw-Hill, New York, 1967.
17. Brunt, D., Radiation in the Atmosphere, Q. J. R. Meteorol. Soc., 66 (Suppl.), 34, 1940.
18. Berdahl, P. and Fromberg, R., The Thermal Radiance of Clear Skies, Sol. Energy,
29, 299, 1982.
19. Kraus, A. D. and Bar-Cohen, A., Design and Analysis of Heat Sinks, John Wiley &
Sons, New York, 1995.
20. Garg, H. P., Treatise on Solar Energy, Vol. 1, John Wiley & Sons, New York, 1982.

© 2001 by CRC PRESS LLC


0082-06.fm Page 307 Wednesday, August 23, 2000 3:54 PM

6 Heat Transfer with Phase


Change

6.1 INTRODUCTION
Heat transfer of a phase change coolant is much more complex than the previous
modes of heat transfer that we have studied. By phase change we denote the following
processes:

a. Solid changing to a liquid—fusion, or melting,


b. Liquid changing to a vapor—evaporation, also boiling,
c. Vapor changing to a liquid—condensation,
e. Liquid changing to a solid—crystallization, or freezing,
f. Solid changing to a vapor—sublimation,
g. Vapor changing to a solid—deposition.

Because these processes involve a fluid medium, we generally classify them as


convection processes. Beyond all of the parameters that we previously examined in
convection, we must now add the variables of surface tension, ; ambient pressure,
P; and either the enthalpy or latent heat of evaporation, hfg; the latent heat of
solidification, hfs; or the latent heat of sublimation, hsg. The latent heat is the amount
of heat required to convert a unit mass of a substance from one phase to another
phase. There is also a difference in density between the liquid phase and the vapor
phase that induces a buoyancy force,   l  v , when bubbles are present.
Also, we will note that heat transfer during phase change does not always occur
with a change in temperature of the media. In fact, a very large rate of heat transfer
can be achieved with very little change in temperature. This is one of the attractions
of phase change heat transfer. Furthermore, in contrast to natural or forced convec-
tion, increasing the T may result in a decrease in the heat transfer coefficient.
Because of the number of variables, there are no accurate general equations or
correlations to use. Of the usable equations, most have an empirical value that
changes with the surface characteristics and must be evaluated by experimentation.
The accuracy of these correlations without experimental verification may be
50%.
Although heat transfer by phase change is not yet widely used in electronics
cooling, as the component heat flux rises the laws of physics dictate that high-end
cooling technologies will progress from air-cooled to liquid-cooled to phase change.
Our studies in this section will be largely theoretical.

© 2001 by CRC PRESS LLC


0082-06.fm Page 308 Wednesday, August 23, 2000 3:54 PM

6.1.1 DEFINITIONS OF PHASE CHANGE PARAMETERS


Vapor pressure—At any temperature above absolute zero, the molecules in a liquid
are in constant motion. Some of these molecules will have a higher velocity than
the average. If the energy in one of these high-speed molecules is greater than the
cohesive forces, the molecule can “escape” through the surface of the liquid. We
know this as evaporation. Since these higher-speed molecules contain more energy
than the remaining molecules, the reduction in energy causes the bulk liquid to cool.
If the liquid is in an airtight enclosure, the escaped molecules will fill the airspace.
Some “escaped” liquid molecules will even reenter the liquid. Eventually, the number
of molecules escaping the liquid equals the number of molecules reentering the liquid.
When this occurs, we call the air space a saturated vapor. Since the molecules exert
a pressure within this enclosure, we call this the saturated vapor pressure. If we increase
the pressure within the airspace, or decrease the volume of the airspace, the vapor will
contain more fluid particles than it can hold. We can then say that we have supersat-
urated the vapor. Supersaturation conditions can occur only temporarily.
Phase change—In the previous discussion, the liquid molecules changed phase,
from a liquid to a vapor. If the walls of the enclosure (from the previous discussion)
are suddenly cooled, we would again supersaturate the vapor. Randomly moving liquid
molecules that strike the cool enclosure walls would leave their vapor state and return
to the liquid state, forming condensation. This is also a change in phase—from vapor
phase to liquid phase. We see an identical process when water, in the solid form of
ice, melts and becomes a liquid.
Another process, known as sublimation, occurs when a material changes from
a solid phase directly to a vapor phase. The opposite of sublimation is deposition,
where the vapor phase changes directly to a solid.
The triple point of a material is the combination of temperature and pressure
conditions at which the material can exist in a solid, liquid, and vapor state simul-
taneously. Water has a triple point of 0.01°C at 4.58 torr. The sublimation point is
the temperature and pressure at which the material can exist as solid and vapor. The
critical point is the temperature at which no amount of pressure will cause the vapor
to liquefy. This occurs when the molecules are moving so fast that the internal
cohesive forces are not strong enough to form a surface.
Figure 6.1 shows the phase diagram for H2O. A is the triple point. B is the critical
point. Curve AB shows the points at which H2O can exist in both the liquid and vapor
phase. Line AC shows where H2O can exist as a solid and a liquid, and AD shows
where it can exist as both a solid and a vapor. Standard atmospheric pressure is 760
torr (760 mmHg). Moving along the 760 torr line we see that H2O has a phase change
from solid to liquid at 0°C, and changes from a liquid to a gas at 100°C. Sublimation
occurs along line efg, where the solid may change directly into a vapor without turning
into a liquid first. At point B, temperatures of 374°C and above, H2O cannot liquefy
at any pressure. This is called the critical temperature and occurs when the gas
molecules are moving so fast that cohesive forces are unable to form a surface.
Technically, a vapor is called a gas when it exceeds the critical temperature.
As previously noted, a liquid will evaporate if the vapor pressure is lower than
saturation pressure. Saturation pressure increases with temperature. If the saturation

© 2001 by CRC PRESS LLC


0082-06.fm Page 309 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.1 The phase diagram for H2O.

pressure is equal to the surrounding atmospheric pressure, the liquid will form vapor
bubbles throughout and boil.
Latent heat—As an object absorbs energy, the object will display a temperature
increase, defined by the specific heat of the object. We call this temperature increase
sensible heat. Another form of heat is latent heat, by which an object will absorb
energy but will not increase in temperature. The latent heat of vaporization is higher
than the latent heat of fusion because the molecules are spaced farther apart in a
gas than in a liquid.

6.2 DIMENSIONLESS PARAMETERS IN BOILING


AND CONDENSATION
For boiling, the Jakob (Ja) number is the ratio of maximum sensible heat absorbed
by the liquid to the latent heat absorbed by the vapor, or the reverse for condensation.
The Jakob number can also be described in terms of temperature difference as

c p ( T s  T sat )
Ja  -------------------------------
-
h fg

This dimensionless group characterizes the heat transfer during a phase change.
The Jakob number is usually quite small. For example, we can let Ts equal the
temperature of the liquid surface of an ice block Ti equal the internal temperature
of the ice block. If we say the temperature difference between an ice block and a

© 2001 by CRC PRESS LLC


0082-06.fm Page 310 Wednesday, August 23, 2000 3:54 PM

liquid surface is 10°C, then Ja  0.058. The Bond number is the ratio of the
gravitational buoyancy force to the surface tension.

6.3 MODES OF BOILING LIQUIDS


Boiling occurs when the temperature of a surface wall, Tw, exceeds that of the liquid
saturation temperature, Tsat. The heat transferred from the heated wall to the liquid
can be found by

qs  h ( T w  T sat )  h  T e

where qs is the rate of heat generation per unit of surface (W/m2) and Te is the excess
temperature. As we know from experience, bubbles within the liquid characterize
the boiling process. The dynamics of the bubbles reaching the surface affect the
fluid motion and, therefore, the convection.
The boiling heat transfer process consists of two basic types: pool boiling, which
occurs in an initially stagnant liquid, and flow boiling, which occurs in the presence
of liquid velocity. Boiling may occur in both process when no bubbles are visible.
We call this subcooled boiling. In this process, the temperature of the liquid is below
the saturation temperature. In subcooled boiling, bubbles form in the superheated
liquid at the wall but are condensed when they grow large enough to extend into the
subcooled bulk liquid. When the vapor extends into the subcooled liquid, it loses its
heat and collapses. Figure 6.2 shows the phases in subcooled boiling. Saturated boiling

FIGURE 6.2 Fluid flow pattern induced by a bubble in a subcooled liquid in different stages
of development collapse.

© 2001 by CRC PRESS LLC


0082-06.fm Page 311 Wednesday, August 23, 2000 3:54 PM

is the type of boiling with which we are most familiar: the temperature of the liquid
exceeds the saturation temperature, bubbles form, and they rise to the liquid surface.

6.3.1 BUBBLE PHENOMENON


Bubbles are an important characteristic of saturated boiling. Bubbles form individ-
ually in microscopic surface imperfections on the heated body. The bubbles grow
until they separate from the heated surface and reach the surface of the liquid. At
higher heat fluxes bubbles form, separate, and reform so quickly that continuous
streams or vapor columns are seen. At still higher heat flux levels, the process occurs
so quickly that the liquid cannot reach the surface of the heated body. In this case
the heated body is continually blanketed by a vapor film.
A vapor bubble forms and grows in a liquid as long as the bubble vapor pressure
is higher than the ambient liquid pressure, that is, p
pl. Stability is reached when

2
p  p l  -------
Rb

where Rb is the radius of the vapor bubble. If we assume that the bubble and the
liquid are at identical temperatures, the  p between the vapor and the liquid can be
translated into a temperature by the Clausius-Clapeyron relation. Therefore, from
Hetsroni1 we see that

2  fg 
T  T l  T s 1 ------- ------ -----------------
R b h fg  l  

While this equation is for an isothermal condition, Bergles and Rohsenow2 found
the theoretical temperature difference between the wall surface and the liquid satu-
ration temperature that will form the first bubble, or the incipience of boiling

8qT 
 T b  T w  T sat  -----------------
 v h fg k f

This equation assumes that temperature decreases linearly with increasing distance
from the heated wall. Since this is not strictly true in actual occurrences of wall
superheat, the equation underpredicts the wall superheat, Tb.
The familiar boiling process is actually called nucleate boiling. When a bubble
becomes large enough to detach from a heated surface, the bubble is said to have
nucleated. In nucleate boiling the characteristic length is the size of the bubble when
it separates from the surface, Lb. To find this dimension, we balance the bubble surface
tension to the bubble buoyancy force, as shown in Figure 6.3; therefore,

Surface Tension  Buoyancy Force


2 3
2 R b   --- R b (  l   )g
3

where Rb is the bubble radius.

© 2001 by CRC PRESS LLC


0082-06.fm Page 312 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.3 Force balance on vapor bubbles in a fluid: (a) unattached forces, (b) attached
vapor bubble forces.

The bubble radius is therefore found by

3  gc 0.5
R b  ------------------------
-
(  l   )g

and the characteristic length is then related to the radius by

 gc 0.5
L b  ------------------------
-
(  l   )g

Cole and Rohsenow3 correlated vapor bubble departure diameters in a saturated


water pool by a dimensionless Eustis number

4 1.25 2
E 0  [ 1.5 10 ( Ja ) ]

and in saturated pools of other liquids by

4 2
E 0  [ 4.65 10 ( Ja ) ]
1.25

where:
g (  l   )D b
2

E0  Eustis number, --------------------------------


gc 
 cT s
Ja*  modified Jakob number ------------
v h f s
-

© 2001 by CRC PRESS LLC


0082-06.fm Page 313 Wednesday, August 23, 2000 3:54 PM

Now that we have correlated the bubble departure diameter, we can find the
frequency of departure. Although current experimental data are irregular, Hetsroni1
found that bubbles that grow very slowly depart at a frequency of about

0.5625  Ja
2

f  -------------------------------
2
-
Db

Cole4 roughly correlated the frequency of departure for rapidly growing bubbles as

0.5
g ( l   )
4
---
f  ---------------------------
3
C D l Db

where CD is the dimensionless drag coefficient.


Theoretically, now that we can calculate the diameter of the bubbles, Db, and
the frequency of bubble departure, f, we can determine the bubble, or vapor flux,
Qv, if we can find the number and size of nucleation sites per unit area. Since this
value is usually unknown, another way to find the vapor flux is by the equation

q
Q   ------------
h fg 

where  is the dimensionless fraction of the heat flux that will result in a net generation
of bubbles. Unfortunately,  is not a constant. Graham and Hendricks5 found that 
varies in a complex manner from about 0.01 to 0.02 at low heat fluxes, to
0.5 at
about 20% of the critical heat flux, and approaches 1.0 at the critical heat flux.
If the vapor flux can be determined, then we can find the amount of agitation
caused by nucleate boiling, which is described as the bubble Reynolds number

Db Q
Re b  -------------
l

A good estimate of the maximum bubble velocity can be found by balancing


the vapor kinetic energy against the buoyancy force of the bubble in terms of the
characteristic length. Therefore,

Vapor Kinetic Energy  Buoyancy


1
---  V max  g (  l   )L
2
2

Since we know the characteristic length is

 gc 0.5
L b  ------------------------
-
(  l   )g

© 2001 by CRC PRESS LLC


0082-06.fm Page 314 Wednesday, August 23, 2000 3:54 PM

the maximum bubble velocity can be described as

 (  l   )gg c 0.25
V max ~  ---------------------------------
-
 
2 

The maximum bubble velocity can also be described using what is known as
the Helmholtz* theory of instability. That is, the bubble vapor column is disrupted
by a wavelength disturbance of Lb, the characteristic length, and becomes unstable
at VH, or

2  g 0.5
V H ~  ----------------c
  Lb 

If the vapor bubble diameter is limited to an experimental range found by


Bromley et al.,6 that is

gc  gc 
-  D b  5.45 ------------------------
3.14 ------------------------ -
g ( l   ) g ( l   )

then we may use a quick estimate of the velocity of an undisturbed bubble given
by Zuber et al.7

1
V b  --- 4gr
3

Similar to bubble growth in nucleate boiling, vapor bubbles collapse when


subcooled. Florschuetz and Chao8 used a temperature integral developed by Plesset
and Zwick9 to determine the rate of collapse, H, of a bubble:

 ˙ 
4 2  1  -----------
2
-  Rb  2 
H -t  ---   R˙  --------  3
 ---- Ja -------
 
R b, i 2 3   ---------
b

R b, i 
  R b, i 

where Rb is the bubble radius and Rb,i is the initial bubble radius.
If we define the bubble collapse period as the time when the bubble volume is
1% of the departure bubble volume, then the ratio Rb/Rb,i  0.2. We now define the
collapse period, c, as

 
1  -----------
2
-  Rb  2 
 c  ---   R˙   --------  3 Rb
 2.32
3  ---------
b R b, i 
  ----------  0.2
  R b, i 
R b, i

* Hermann von Helmholtz (1821–1894) described the theory of instability and also provided a physical
proof of Fourier’s Theorem by producing complex musical tones using individual tuning forks.

© 2001 by CRC PRESS LLC


0082-06.fm Page 315 Wednesday, August 23, 2000 3:54 PM

and substituting and rearranging, we see that

2.32 R b, i R b, i
2 2
 c  ---------- ------------
-  0.580 ------------
-
4 Ja  2
Ja 
2

We can now determine the relationship among the rise velocity, Vb, the rate of
collapse, c, and the departure diameter, Db , to the collapse length, Lc, as

c

Lc  0 V b dt

Substituting the equations for the collapse rate, H, and the undisturbed bubble
velocity, Vb, and letting the ratio of the bubble radius, Rb /Rb,i, equal the ratio of the
bubble diameter, Db /Db,i, we have an equation for the collapse length, Lc, as

D b 2g
2.5 2.5
D b 2g
L c  1.4 -------------------
- ------  0.0292 -----------------------
-
Ja  Ja 
2 48 2

6.3.2 POOL BOILING


There are two different types of boiling: pool, and flow. The type of boiling known
as saturated pool boiling is depicted in Figure 6.4. Note that the temperature of the
boiling liquid is nearly constant except at the surface of the heated wall. At the
heated surface the liquid temperature increases sharply.

FIGURE 6.4 Temperature distribution of the heated solid surface, Ts, and the boiling liquid,
Tsat , during saturated pool boiling.

© 2001 by CRC PRESS LLC


0082-06.fm Page 316 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.5 The boiling curve generated by Nukiyama10 for saturated water. Adapted from
Incropera, F. P., and DeWitt, D. P., Fundamentals of Heat and Mass Transfer, 3rd ed., John
Wiley & Sons, New York, 1995, 722, using the data of Nukiyama, S., “The Maximum and
Minimum Values of Heat Transmitted from Metal to Boiling Water Under Atmospheric
Pressure,” J. Japan Soc. Mech. Eng., 37, 367, 1934. (Translation: Int. J. Heat Mass Transfer,
1419, 1966).

6.3.2.1 Pool Boiling Curve

In 1934, Nukiyama10 published experiments defining the different regimes of pool


boiling. Nukiyama used a nichrome wire immersed in a bath of water. As he
increased the power to the wire, he noticed bubbles forming at an excess temperature,
 Te , of about 5°C. As he increased power further, he noticed that the power supplied
to the wire could be increased greatly without a large increase in temperature. As
he increased the power further still, the temperature suddenly increased dramatically,
and the wire reached its melting point. After experimenting with a platinum wire
with a higher melting point, he noted that the heat flux and temperature were related,
as shown in Figure 6.5. Nukiyama believed that a method of temperature control of
Te, instead of a power-controlled method, would yield a better curve. In 1937, Drew
and Mueller11 performed the experiment using a steam pipe and obtained a curve
similar to that shown in Figure 6.6. This figure shows the relationship of heat flux
and temperature for water at sea-level atmosphere. The relationship is similar for
most other liquids.
We call the range when  Te   Te, A, free convection boiling. Bubbles do not
form in this region because there is not enough vapor in contact with the liquid. Fluid
motion in this range is caused by natural convection. Nucleate boiling occurs when
individual bubbles form on the heated surface and rise to the liquid surface. This point
on the curve is designated A. The portion of the curve designated AB is characterized
by isolated bubbles. As the temperature of the heated surface is increased further, the
bubbles are generated faster and sometimes merge to form vapor columns. This region
is shown by the portion of the curve designated BC. Nucleate boiling occurs in the

© 2001 by CRC PRESS LLC


0082-06.fm Page 317 Wednesday, August 23, 2000 3:54 PM

Boiling
regimes
Free convection Nucleate transition Film

{
{

{
{
Isolated Jets and
bubbles columns
107


qmax
 max
qmax
106 C Critical heat flux, q"
P
q"s (W / m2 )

B
105


qmin
D  min
Leindenfrost point, qq"min
A
10 4
ONB
∆Te,A ∆Te,B ∆Te,C ∆Te,D

10 3
1 5 10 30 120 1000

∆Te = Ts Tsat( C)
_ o

FIGURE 6.6 Characteristics of the boiling curve for a heated horizontal surface in water.

temperature range of  Te,A   Te   Te,C where  Te,C is about 30°C. This is the
range of operation for most heat transfer work. High levels of power can be dissipated
without a large increase in temperature. Point P indicates the point of the maximum
heat transfer coefficient. Ideally, equipment should operate at this point. In water, the
convection coefficient in this region can exceed 104 W/m2 K.
The region when  Te,C   Te   Te,D, where  Te,D is about 120°C, is called
the transition boiling, partial film boiling, or unstable film boiling range. In the
previous range, as each bubble left the surface, liquid covered the surface until a
new bubble formed. In the transition boiling range, new bubbles are formed before
the liquid can reach the surface. A continuous vapor film forms on the surface. The
entire heated surface oscillates between a liquid and vapor blanket. As the temper-
ature differential  Te increases, the entire surface is more often covered by a vapor
layer than the liquid layer. Also, as  Te increases, hc and therefore qs decrease,
because the thermal conductivity of the vapor layer is much lower than when the
adjacent layer was a liquid.
The region when  Te
 Te,D is called film boiling. Point D of the boiling curve
is called the Leidenfrost point. In 1756, Leidenfrost noticed that when water droplets
are placed on a hot surface, the droplets dart about the surface, supported by a vapor
layer. We know that during the transition phase, as  Te increases, a higher percentage
of the surface is covered by vapor at any point in time. The Leidenfrost point occurs
when the entire surface is covered by a vapor layer, and the heat flux reaches a minimum,
qs,D  qmin. In the film boiling range, heat transfer can only occur by conduction through
the vapor layer. After the Leidenfrost point, radiation heat transfer through the vapor
layer becomes more important, and the heat flux increases with increasing  Te.

© 2001 by CRC PRESS LLC


0082-06.fm Page 318 Wednesday, August 23, 2000 3:54 PM

Researchers have experimented with the region after point C, but in actual
engineering applications this region is difficult to control. Any increase in the heat
flux after point C creates a marked increase in temperature. The size of this increase
may cause destruction of the heat flux surface. For this reason, point C is often
called the burnout point or, more commonly, the Critical Heat Flux (CHF).

6.3.2.2 Pool Boiling Correlations

Researchers have calculated a number of correlations to explain the actions of pool


boiling, based on the pool boiling curve. For Nusselt numbers up to the Te, A point,
standard free convection correlations can be used. In the region of nucleate boiling,
where  Te,A   Te   Te,C and where  Te,C is about 30°C, the Nusselt number is
highly dependent upon the rate of bubble formation. Although no exact models are
available to describe this phenomenon, Yamagata et al.12 correlated the influence of
nucleation sites on the heat flux by

qs  C  T e n
a b

where:

qs  surface heat flux (W/m2)


C  constant dependent upon liquid/surface combination
Te  excess temperature (Ts  Tsat) (°C)
a  1.2
n  bubble site density (N/m2) (   T e )
6

b  1/3

Currently, the Rohsenow13 correlation is the most useful:

g ( l   ) 0.5
c p, l  T e  3
 ----------------------
qs   l h fg ------------------------
-
 gc  C h Pr n
sf fg l

where:

  absolute viscosity (N s/m2) cp  specific heat (J/kg K)


hfg  latent heat of vaporization Cs f  coefficient of liquid/surface
(J/kg) combination
g  gravitational acceleration Pr  Prandtl number, cp/k
(9.806 m/s2) n  exponent for liquid
  mass density (kg/m3) l  liquid phase
  surface tension (N/m)  vapor phase
gc  gravitational constant
(1.0 kg m/N s2)

Figure 6.7 shows pool boiling data points for water that were correlated by the
Rohsenow method. Collier14 recommends the following correlation as being simpler

© 2001 by CRC PRESS LLC


0082-06.fm Page 319 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.7 Pool boiling data for water correlated by the Rohsenow method. (Adapted from
References 13 and 36.)

to use than the Rohsenow correlation:


3.33
p 0.17 p 1.2 p 10
qs  0.000481  T 1.8  ------ 4  ------ 10  ------
3.33 2.3
p
e cr
 p cr  p cr  p cr

where:

Te  excess temperature, °C


pcr  critical pressure, atmosphere (101,325 N/m2)
p  operating pressure, atmosphere (101,325 N/m2)

The Rohsenow correlation13 can be manipulated to find the nucleate boiling


Nusselt number in terms of the Jakob number
2
c p, l ( T w  T sat )
------------------------------------- 2
h fg Ja
3 m
-  ---------------
Nu  ----------------------------------- 3
-
m
C sf Pr l C sf Pr l

where the exponent m is 2.0 for water and 4.1 for other liquids.
Danielson et al.15 report a number of values for Csf . The value of Csf can be
assumed to equal ~0.013 when the experimental value is unknown. Hetsroni1 reports
that this value correlates a wide spectrum of experimental data to within 20%.
The most important variables that affect Csf are the surface roughness of the heated

© 2001 by CRC PRESS LLC


0082-06.fm Page 320 Wednesday, August 23, 2000 3:54 PM

TABLE 6.1
Values of the Surface/Liquid Coefficients
Liquid/Surface Combination Csf

Benzene–Chromium 0.010
Carbon Tetrachloride–Copper 0.013
Carbon Tetrachloride–Polished Copper 0.007
Ethyl Alcohol–Chromium 0.0027
Isopropyl Alcohol–Copper 0.0023
n-Pentane–Chromium 0.015
n-Pentane–Polished Copper 0.0154
n-Pentane–Lapped Copper 0.0049
Water–Brass 0.006
Water–Copper 0.013
Water–Scored Copper 0.0068
Water–Polished Copper 0.013
Water–Nickel 0.006
Water–Chemically Etched Stainless Steel 0.0133
Water–Mechanically Polished Stainless Steel 0.0132
Water–Ground and Polished Stainless Steel 0.008

surface and the angle of contact between the vapor bubble and the heated surface.
The surface roughness affects the number of nucleation sites, and the angle of contact
is a measure of the wettability of the surface. Smaller contact angles represent greater
wettability. A totally wetted surface has the least amount of vapor and represents
the greatest heat transfer coefficient. Values of the coefficients of liquid/surface
combinations are shown in Table 6.1.
Table 6.2 presents important boiling point thermophysical data for coolants
commonly used in electronic cooling.

6.3.2.3 Pool Boiling Critical Heat Flux Correlations

The critical heat flux, point C in Figure 6.6, is determined by the maximum rate
that vapor bubbles can leave the heated surface. This can also be described as the
maximum speed that the liquid can re-wet the heated surface after a bubble leaves
the wall. We can now relate the bubble velocity to the critical heat flux by

qmax   V max h fg

or

qmax
-  C max
----------------------
 V max h fg

© 2001 by CRC PRESS LLC


0082-06.fm Page 321 Wednesday, August 23, 2000 3:54 PM
© 2001 by CRC PRESS LLC

TABLE 6.2
Fluid Properties at Respective Boiling Points (1.0 atm)

Coolant
Property FC-72 FC-77 FC-84 FC-87 L-1402 R-12 R-113 Water

Boiling Point, °C 52.0 100.0 83.0 30.0 51.0 30.0 48.0 100.0
Density, Liquid, l, kg/m3 1592 1590 1575 1633 1635 1487 1511 958.0
Density, Vapor,  , kg/m3 12.68 14.31 13.28 11.58 11.25 6.34 7.40 0.59
Absolute Viscosity, , N s/m2 0.00045 0.00045 0.00042 0.00042 0.00052 0.00036 0.00050 0.00027
Specific Heat, cp, J/kg K 1088 1172 1130 1088 1059 – 979 4184
Heat Vaporization, hfg, J/kg 87927 83740 79553 87927 104675 165065 146824 2257044
Thermal Conductivity, k, W/m K 0.0545 0.0570 0.0535 0.0551 0.0596 0.0900 0.0702 0.683
Surface Tension, , N/m 0.0085 0.0080 0.0077 0.0089 0.0109 0.0118 0.0147 0.0589
Coefficient of Expansion, , 1/K 0.0016 0.0014 0.0015 0.0016 0.0016 – 0.0017 0.0002
0082-06.fm Page 322 Wednesday, August 23, 2000 3:54 PM

where:

qmax  critical heat flux (W/m2 )


v  vapor density (kg/m3)
Vmax  maximum bubble velocity (m/s)
hfg  latent heat of vaporization (J/kg)
Cmax  coefficient of critical heat flux geometry

Kutateladze17 defined the equation for the critical heat flux in 1948 by dimen-
sional analysis. Later, in 1959, Zuber18 found the equation for critical heat flux by
a hydrodynamic stability analysis. The critical heat flux can be found by

0.25
qmax  C max h fg [  v (  l   v )gg c ]
2

where Cmax is  /24, also called the Zuber constant. Table 6.3 presents the exper-
imental data of Lienhard and Dhir19 for the value of Cmax using a dimensionless
parameter called L*. This parameter is the ratio of the characteristic length of the
heated surface, Ls, to the characteristic bubble dimension, Lb.
Pressure affects the maximum heat flux because of the influence on vapor density
and the boiling point of the liquid. As the boiling point changes, so does the heat
of vaporization and the surface tension. Therefore, each fluid has a specific pressure
that will yield a maximum heat flux. Cichelli and Bonilla20 have experimentally
demonstrated that the critical heat flux increases with pressure up to 1/3 of the critical
pressure. After this peak, the critical heat flux falls to zero at the critical pressure, as
shown in Figure 6.8.

TABLE 6.3
Values of Cmax (Zuber Constant) for Critical Heat Flux
Surface Geometry Cmax Characteristic Length Range of Applicability

Infinite flat heater 0.15 Width or Diameter L*


27
Small flat heater 12 L
2
Width or Diameter 9  L*  20
0.15 ----------------b
As
Large horizontal 0.12 Radius L
1.2
cylinder
Small horizontal 0.12L* 0.25 Radius 0.15  L*  1.2
cylinder
Large sphere 0.11 Radius 4.26  L*
Small sphere 0.227L* 0.5 Radius 0.15  L*  4.26
Large arbitrary body 0.12 — —
Ls Ls
Note: L*  ----
Lb
-  -------------------------------------
 gc 0.5
-
--------------------------
(  l   )g

© 2001 by CRC PRESS LLC


0082-06.fm Page 323 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.8 Critical heat flux in nucleate boiling as a function of pressure. (From Cichelli,
M. T. and Bonilla, C. F., Trans. AIChE, 41, 755, 1945. With permission.)

When the liquid is subcooled, Zuber et al.7 have found that the critical heat flux
is correlated to a reasonable estimation by

 2k l ( T sat  T l ) 24 
2

0.25
qmax  qmax sat  1 ---------------------------------
- ---------------- ---------------------------------- 
  l  h fg  v g  g c (  l   ) 

where:

qmax sat  critical heat flux in saturated pool boiling (W/m2)


kl  thermal conductivity of the liquid (W/m K)
Tl  temperature of the bulk liquid (C)
l  thermal diffusivity of the liquid, k/cp

 gc 
2 0.25
0.5
  ---- 2 ------------------------
- ---------------------------------
-
3 g ( l   )  gg c (  l   v )

The effect of subcooling on the critical heat flux was studied by Ellion.21 All of
the previous correlations are for clean horizontal surfaces. Bernath22 has shown that
a vertical surface may have a critical heat flux as much as 25% less than a horizontal
surface.

© 2001 by CRC PRESS LLC


0082-06.fm Page 324 Wednesday, August 23, 2000 3:54 PM

6.3.2.4 Pool Boiling Minimum Heat Flux Correlations

There are no adequate theories that describe the transition regime between critical heat
flux and vapor film boiling. This region corresponds to qs , D of Figure 6.6 and is called
the minimum heat flux, or the Leidenfrost point. Fortunately, this area of the boiling
curve has little practical value in electronic cooling. At lower temperatures within this
region, fluid motion is characterized by periodic, unstable liquid and heater contact. At
the higher-temperature region, a stable vapor film forms and the heat flux reaches a
minimum. If the heat flux drops below the minimum required to initiate a stable vapor
film, the liquid will again contact the heater surface and nucleate boiling will be rees-
tablished. Zuber23 used stability theory to derive an equation for minimum heat flux, qmin
gc  g ( l   ) 0.25
qmin  C  h fg ----------------------------------
( l v )
2

Kovalev24 notes that this correlation is accurate to only about 50% at moderate
pressures and has even less accuracy at high pressures. Nevertheless, Berenson,25
for a large flat horizontal heated surface, and Lienhard and Wong,26 for a horizontal
heated cylinder, found that C  ~0.09.

6.3.2.5 Pool Boiling Vapor Film Correlations

The region of vapor film boiling, like minimum heat flux boiling, is not of great
concern in applications involving electronic cooling. Above the Leidenfrost point
the heat flux increases, but at a dramatic increase in surface temperature. In this
range the heated body is enveloped in a continuous blanket of vapor; therefore, heat
transfer is by conduction through the vapor. At temperature ranges much higher than
those seen in electronics, radiation heat transfer across the vapor film becomes
important. Because the vapor film is thin, the Rayleigh number is low, which indicates
heat transfer by conduction only. Because of this, analysis of the heat flux for vapor
film boiling is relatively simple. The minimum wall superheat that sustains the vapor
film was found by Berenson25 to be

 , f h fg g (  l   ) 23
gc  0.5
 , f 13
T min  0.127 ---------------
- ------------------------- ----------------- ------------------------------
-
k , f l  g ( l  ) gc  ( l   )

where the subscript f is that property at the film temperature. For film boiling, the
following correlation for the average heat transfer coefficient is commonly used:

g (  l   )  k ( h fg 0.35c p,  T e )
3 0.25
h c  C fb -------------------------------------------------------------------------------------
-
D   T e

where the constant Cfb is 0.62 for a horizontal cylinder (Bromley27), 0.67 for a sphere
(Frederking and Daniels28), and replacing D with L is 0.71 for a plane vertical surface
(Mills29). When the vapor flow rate is high, that is, when

L (  l   )g ( h fg 0.50c p, v  T e )
3
--------------------------------------------------------------------------------
-
5 10 7
k   T e

© 2001 by CRC PRESS LLC


0082-06.fm Page 325 Wednesday, August 23, 2000 3:54 PM

Frederking and Clark30 recommend the following equation for the average film
boiling heat transfer coefficient
13

h  0.15 --------------------------------------------------------------------------------------
(  l   )g ( h fg 0.50c p,  T e )k v
2
-
  T e

6.3.3 FLOW BOILING


Similar to normal convection, boiling has two modes: pool boiling and forced
convection boiling. Pool boiling is dominated by fluid motion caused by buoyancy
forces, both liquid and vapor, within the fluid. In forced convection boiling, fluid
motion is characterized by an external force such as a pump. Also, similar to normal
convection, flow can be divided into two major categories: external flow and internal
flow. In electronics cooling, external flow usually occurs in large enclosures directly
on the surfaces of hot components. Internal flow occurs through tubes and ducts
and requires a complex analysis because the bulk flow is part liquid and part vapor.
In instances of pool boiling, thermal overshoot is a problem. Thermal overshoot is
the initial resistance of a liquid to boil. The heated surface may reach 30°C above
the normal boiling temperature before actual nucleate boiling occurs. This problem
does not normally occur in forced convection boiling.
In two-phase flow, a variable known as the vapor mass quality, , is often used.
The vapor mass quality describes the ratio of the vapor mass flow to the total mass
flow and is found by


  -------------------
ṁ ṁ l
-

where:

ṁ  mass vapor flow rate (kg/s)


ṁ l  mass liquid flow rate (kg/s)

Another term often used in two-phase internal flow is the mass velocity, G,
which is the mass flow per cross-sectional area. The mass velocity is further divided
into vapor mass flow, Gv , and liquid mass flow, Gl. The quantities can be shown as

G  G Gl

where:


G v  -----
Ac
-

ṁ l
G l  -----
Ac

© 2001 by CRC PRESS LLC


0082-06.fm Page 326 Wednesday, August 23, 2000 3:54 PM

6.3.3.1 External Forced Convection Boiling

In electronic cooling external flow is usually a relative statement. The flow may be
external over hot components, but it is still within a conduit. The pressure drop in
this type of flow is not at all well understood. Some correlations are available, but
the range of applicability is so narrow as to be nearly useless for engineering design.
In cases of conduit flow which passes over exposed components, the engineer should
always build a scale model to determine pressure drop.
At temperatures before boiling incipience, standard correlations for forced con-
vection may be used, with attention to temperature-induced property variations within
the single phase. The maximum heat flux can be increased substantially by using
forced convection in the boiling regime. Researchers have recorded heat flux values
of 35 MW/m2. Lienhard and Eichhorn31 developed a correlation for the maximum
heat flux using the Weber number, which is similar to the Reynolds number. Their
correlation is usually accurate to about 20%. The Weber number, We, is the ratio of
inertia to surface tension and can be described as

 V D
2

We  ---------------

-

For a heated cylinder in a low-velocity liquid cross flow, the critical heat flux
is estimated by

1 4 13
qmax   h fg V ---- 1  ----------
-
 We D

where low velocity is described as:

qmax 
 ---------------- 0.275  0.5
 -------------  -----l  1
  h fg V    

For a heated cylinder in a high-velocity liquid cross flow, the critical heat flux
is estimated by

l 
 ----
0.75 l 
 ----
0.5
- -
     
q m ax   h fg V -------------- ------------------------------
169 19.2 We 1D 3

where high velocity is described as:

qmax 
 ---------------- 0.275  0.5

-------------  -----l  1
  h fg V    

© 2001 by CRC PRESS LLC


0082-06.fm Page 327 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.9 The fluid flow regime progression from liquid to vapor for two-phase vertical
internal flow.

6.3.3.2 Internal Forced Convection Boiling

Forced convection boiling in a tube is quite complex. Researchers have identified


six regions of heat transfer in a vertical tube. These regions may blend into each
other and are not well defined. Approximate views of the modes of vertical internal
two-phase flow are shown in Figure 6.9. When a liquid flow enters a vertical tube
having a surface temperature above Tsat, the liquid initially is heated to Tsat and
initiates nucleate boiling. When nucleate boiling occurs, the single-phase liquid
(mode 1) flow becomes a two-phase bubbly flow (mode 2), composed of vapor
bubbles and liquid. As the fluid temperature increases, the vapor bubbles merge and
form large bubbles called slug flow (mode 3). Churn flow (mode 4) is highly irregular
and unsteady, and is composed of large bubbles that continually break apart and
then merge together. Higher in the tube, annular flow (mode 5) occurs. Annular flow
occurs when liquid is only in contact with the tube walls, and a core of vapor and
a liquid mist fills the center of the tube. Still higher in the tube, flow is composed
of pure vapor and the liquid mist called droplet flow (mode 6), until finally the mist
evaporates and only the single-phase vapor (mode 7) is present. Horizontal tubes
show the same modes of heat transfer but have more of a sloshing effect which tends
to blend the modes even more, as shown in Figure 6.10.
The calculation of a two-phase pressure drop within a tube is extremely complex.
Recently, this has been an area of extensive research, with only some success. The
most generally applicable model is based on the assumption of homogeneous flow.
That is, this model ignores the mode of flow and treats the flow as an average of
vapor and liquid. For a simplistic case of flow through a round tube, the total pressure
gradient is the sum of the pressure gradients related to wall friction, gravitational
resistance, and momentum changes:

   ( 1   )  l

© 2001 by CRC PRESS LLC


0082-06.fm Page 328 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.10 The fluid flow regime progression from liquid to vapor for two-phase hori-
zontal internal flow.

Researchers have proposed many different and quite complex heat transfer
correlations to describe the regimes of internal forced boiling flow. Klimenko32
proposed a methodology which has sufficient accuracy, is not overly complex, and
applies to heat transfer stages before mist flow. To use the Klimenko method, we
must first determine if the heat transfer is dominated by film evaporation or by
nucleate boiling, by evaluating the dimensionless parameter Φ.
l
-  1  ----- 
13
Gh
  ----------fg- 1   ----

q    l 

If Φ
1.6 104, film evaporation is the dominant mode of heat transfer; if Φ 
1.6 104, nucleate boiling dominates. The next step in the Klimenko method is to
determine the Nusselt number resulting from the two-phase heat transfer coefficient,
hTP , in either film evaporation or by nucleate boiling. The characteristic length for the
Nusselt number is based on bubble size, which is found by
 gc 0.5
L b  ------------------------
-
g ( l   )
If the heat transfer mode is nucleate boiling, Φ  1.6 104, the Nusselt number
is found by
0.15
Nu  7.4 10 q* P* Pr l
3 0.6 0.5 1  3  k----w-
 kl 
where:
qL b
q*  ----------------
h fg   l
-

p pL
p*  --------------------------------------
-
0.5 
---------b
[  g ( l   ) ] 

kw  wall material thermal conductivity (W/m K)


© 2001 by CRC PRESS LLC
0082-06.fm Page 329 Wednesday, August 23, 2000 3:54 PM

If the heat transfer mode is film evaporation , Φ


1.6 104, the Nusselt number
is found by

 -----   k----w-
0.2 0.09
0.6 1  6
Nu  0.087Re Pr l
 l   k l 

where:

 V Lb
Re  ------------
l
-

G l
V  ---- 1   -----  1
pl   

Finally, knowing the Nusselt number for the heat transfer mode, we can calculate
a single-phase heat transfer coefficient, which is based on the Reynolds number of
the liquid portion of the flow. Usually the single-phase heat transfer coefficient, hFC,
is so much smaller than the two-phase coefficient, hTP , that it can be ignored.
However, if it is significant, the overall heat transfer coefficient can be found by

3 3 13
h c  ( h TP h FC )

Klimenko’s correlations,32 shown in Figure 6.11 and 6.12, match experimental


data with a mean absolute deviation of about 13% for most common coolants used
in electronic cooling. The limitations are

6.1 103 N/m2  p  3.04 106 N/m2


50 kg/m2 s  G  2690 kg/m2 s
0.017    1.00
0.00163 m  D  0.0413 m

6.4 EVAPORATION
Liquid may exist on a surface if the surface is below the saturation temperature, Tsat.
Evaporation, like boiling, is considered a convection heat transfer process. Evapora-
tion occurs when the molecules in a liquid gain enough energy to escape the liquid
binding energy and enter the vapor state. Since the molecules leaving the liquid
contain excess energy, the net effect is a loss of energy of the liquid. Since temperature
is a measure of energy, the temperature of the liquid decreases when evaporation
occurs. The convection of vapor from a surface is related to the mass transfer coef-
ficient, hm. The magnitude of evaporative cooling can be expressed as the equation

h
T   T s  h fg  -----m [  A, sat ( T s )   A,  ]
 hc 

© 2001 by CRC PRESS LLC


0082-06.fm Page 330 Wednesday, August 23, 2000 3:54 PM
© 2001 by CRC PRESS LLC

FIGURE 6.11 Klimenko data for the effect of wall conductivity on the Nusselt number for internal forced convection boiling flow. (From Klimenko, V. V.,
Int. J. Heat Mass Transfer, 31, 541, 1988. With permission.)
0082-06.fm Page 331 Wednesday, August 23, 2000 3:54 PM
© 2001 by CRC PRESS LLC

FIGURE 6.12 Klimenko data showing the transition from nucleate boiling to film evaporation for internal forced convection flow. (From Klimenko,
V. V., Int. J. Heat Mass Transfer, 31, 541, 1988. With permission.)
0082-06.fm Page 332 Wednesday, August 23, 2000 3:54 PM

where:

hm  mass transfer convection coefficient (m/s)


hc  convection heat transfer coefficient (W/m2 K)
A,sat(Ts)  saturated vapor density (kg/m 3)
A,∞  infinite vapor density (kg/m3)

The ratio of the heat transfer coefficient to the mass transfer coefficient, hc/hm,
can also be expressed in terms of a dimensionless Lewis number, Le,

h k
-----c  ------------------n   c Le l  n
hm D AB Le p

where:

DAB  binary mass diffusion coefficient (m2/s)


Le  Lewis number, /DAB
n  13

Using this relationship we see that the cooling effect can also be described as

 A h fg p A, sat ( T s ) p A, 
T   T s  --------------------------
23
-  ----------
- ----------------------
 c p Le Ts T

where:

  molecular weight of species (kg/kmol)


  universal gas constant (8.315 kJ/kmol)
pA,sat(Ts)  saturated vapor pressure (N/m2)
pA,  infinite vapor pressure (N/m2)

Gilliland and Sherwood33 correlated the mass transfer coefficient from liquids
to air in a wetted-wall column as

D AB 0.83 0.44


h m  0.023  --------
- Re Sc
 D L

where:

D  tube diameter (m)


Sc  Schmidt number, /DAB 
2000  Re  35,000
0.6  Sc  2.5

© 2001 by CRC PRESS LLC


0082-06.fm Page 333 Wednesday, August 23, 2000 3:54 PM

Rohsenow and Choi34 indicate that by using the Chilton-Colburn factor, j, the
mass transfer coefficient of evaporation from a flat, wetted surface in laminar flow
can be found by

V
h m  0.664 -------------------------
0.67
-
Re L Sc

and for turbulent flow by

V
h m  0.037 -----------------------
0.2 0.67
-
Re L Sc

where V is the velocity in m/s, and the Reynolds number is based on the length of
the surface.
A Reynolds number based on the thickness of a falling evaporative film has also
been proposed. In this case, the falling film Reynolds number, Re, is defined in
terms of the hydraulic diameter, DH, and the mean velocity, um, of the film as


 l  ------
- 4
l um DH  l  4
Re   -----------------
-  ---------------------
-  -------
l l l

where:

4 Ac 4
D H  --------
-
P  P
------


u m  ------
l 
-

  mass flow per unit film width (kg/m s)


g l  ( l  v )
 3

  0 l u dy  ----------------------------------
3 l
-

In evaporation from falling films of water, Chun and Seban35 determined the
following correlations for the Nusselt number

1  3
Nu   --- Re 
3
0  Re   30
4 
 0.22
Nu  0.822Re  30  Re   Re tr
3 0.4 0.65
Nu  3.8 10 Re Pr  l Re tr  Re 

where Retr is the turbulent Reynolds number, which is equal to 5800Prl1.06.

© 2001 by CRC PRESS LLC


0082-06.fm Page 334 Wednesday, August 23, 2000 3:54 PM

6.5 CONDENSATION
Condensation will form on a surface if the vapor contacting the surface is saturated,
and the surface temperature is lower than the temperature of the saturated vapor.
When this occurs, the latent heat within the vapor is transferred to the cooler surface.
Normally, condensation occurs as drops on a surface. If the conditions that caused
the condensation are steady state, and the surface is clean, the droplets will coalesce
and form a condensate film on the surface. In a gravity field, the laminar fluid film
will flow. If the surface is long enough, the laminar flow of condensate may become
wavy and then turbulent.
Contrary to other forms of heat transfer, as T increases the heat transfer coef-
ficient decreases. This is because as the temperature difference increases, more vapor
becomes condensate, which causes a thicker layer of film. Because condensation heat
transfer is a conduction process, the thicker film impedes the transfer of heat.
Because of the heat transfer impedance caused by the thick fluid film, applica-
tions using condensation often have short flow paths or use small horizontal cylin-
ders. The short flow paths do not allow the film to become very thick. Because of
the action of film condensation, droplet condensation provides a higher heat transfer
coefficient. Heat transfer rates of droplet condensation can be more than a magnitude
higher than film condensation. Therefore, most applications use some type of coating
on the condensation surfaces which promotes droplet formation.
The condensation film begins at point x  0 and continues down a surface. The
thickness of the film increases as x increases, as shown in Figure 6.13. Because the

FIGURE 6.13 The geometry and nomenclature used to describe film condensation on a
vertical surface.

© 2001 by CRC PRESS LLC


0082-06.fm Page 335 Wednesday, August 23, 2000 3:54 PM

flow moves at a constant velocity, and because the film is thicker at larger values
of x, the mass flow rate also increases with x.
The following correlations for condensation heat transfer are most accurate when
evaluated at the film temperature recommended by Addoms36:

T eval  T w 0.33 ( T sv  T w )

Nusselt37 obtained the first relationships for the condensation variables in 1916.
Nusselt used simplifying assumptions to arrive at the basic relationship of the
thickness of the film, ,
0.25
4  l k l x ( T sat  T w )
  -------------------------------------------
-
g  l (  l   )hf g

the local heat transfer coefficient, hx,


0.25
 l (  l   )hf g k l
3
h x  g--------------------------------------------
-
4  l x ( T sat  T w )

and the local Nusselt number, Nu,


0.25
h x x g  (    )h x 3
Nu  -------
- l l fg
k l --------------------------------------------
4  l k l ( T sat  T w )
-

For a vertical surface of unit width and a height of L, the average heat transfer
coefficient can be written as
0.25
g  l (  l   )hf g k l sin 
3
h c  0.943 --------------------------------------------------------
-
 l L ( T sat  T w )

In the previous equations by Nusselt, the modified latent heat of vaporization is


hf g  hfg0.375cp,l(Tsat  Tw). Rohsenow38 modified the relationship to match exper-
imental data by equating the hf g  hfg0.68cp,l(Tsat  Tw), for the range when Pr

0.5 and when cp,l(Tsat  Tw) hf g  1.0. Sadasivan and Lienhard39 calculated the
modified latent heat without Nusselt's simplifying assumptions as hf g 
hfg (0.6830.228/Prl)cp,l(Tsat  Tw). Chen40 further modified Nusselt's equation to
account for interfacial shear and momentum forces. Chen corrected the heat transfer
coefficient as

0.25
p, l ( T sat  T w )
 1 0.68 c---------------------------- c p, l ( T sat  T w ) k l ( T sat  T w )
- 0.02 ---------------------------- - --------------------------
 h fg h fg  l h fg 
hc  hc  -----------------------------------------------------------------------------------------------------------------------------
k l ( T sat  T w ) c p, l ( T sat  T w ) k l ( T sat  T w )
-
 1 0.85 --------------------------  0.15 ----------------------------- -------------------------- 
  l h fg h fg  l h fg 

© 2001 by CRC PRESS LLC


0082-06.fm Page 336 Wednesday, August 23, 2000 3:54 PM

where:

c p, l ( T sat  T w )
------------------------------------  2.0
h fg
k l ( T sat  T w )
--------------------------------  20
 l h fg
1  Pr l  10.05

Chun and Seban35 found that the following correlations for the local and the
average Nusselt number are applicable for water condensation flowing down a flat
surface. When the Reynolds number, based on the film thickness, Reδ , represents
laminar film condensation, and 0  Reδ  30,

1  3
Nu   --- Re 
3
4 

When the Reynolds number indicates wavy laminar film condensation, and 30 
Reδ  Retr,
 0.22
Nu  0.822Re 

When the Reynolds number indicates turbulent film condensation, and Retr  Reδ,

3 0.4 0.65
Nu  3.8 10 Re  Pr l

 1.06
where Retr is the turbulent Reynolds number, which is equal to 5800 Pr l .
To find the average Nusselt number, we use an equation based on the distance
the condensate has traveled down the surface. Using the Reynolds number based on
that length, we have the equation for laminar film condensation, when ReL  30,
written as

0.25

4 Pr  ------ 2-l  1  3
Nu  --- l g
3 --------- ---------------
4Ja l L

When the Reynolds number indicates wavy laminar condensation flow, that is,
when 30  ReL  Retr

0.18
2l- 1  3
 ------
Nu  --------
Pr l  g 
- ---------------
4Ja l L

© 2001 by CRC PRESS LLC


0082-06.fm Page 337 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.14 Local Nusselt number as a function of the Reynolds number for laminar and
turbulent film condensation of water. Adapted from References 29 and 35.

When the Reynolds number indicates turbulent film condensation, that is, when
ReL
Retr, then the average Nusselt number is described as

10  6
-l 
2 13
 ------ 3
9.12 10 Ja l ( L  x tr )
Pr  g ----------------------------------------------------------
- Re 0.6
Nu  --------l- --------------- 13
 ˙2l 
tr
4Ja l L 0.35
 ------- Pr l
 g

Figure 6.14 shows the local Nusselt number as a function of the Reynolds number
for laminar and turbulent film condensation data gathered by Chun and Seban.35
The average heat transfer coefficient of a pure saturated vapor condensing on a
horizontal tube can be found by

0.25
g  l (  l   )hf g k l
3
h c  0.725 --------------------------------------------
-
D  l ( T sat  T w )

Some applications have vertical rows of horizontal tubes; that is, the tubes are
arranged so that the film is falling from an upper horizontal tube to a lower horizontal
tube. If the condensation flow is continuous, the tube diameter, D, in the previous
equation can be replaced by DN, where N is the number of tubes. This method gives
conservative results because the condensation flow is rarely continuous.
Chen40 suggested that since the condensation flow is subcooled, additional con-
densation occurs when the condensate is between the horizontal tubes. Assuming
that all of the subcooling is used for additional condensation, Chen obtained the

© 2001 by CRC PRESS LLC


0082-06.fm Page 338 Wednesday, August 23, 2000 3:54 PM

average heat transfer coefficient as

0.25
g  l (  l   )hf g k l
3
c p ( T sv  T w )
h c  0.728 1 0.2 ------------------------------- ( N  1 ) ---------------------------------------------
h fg DN  l ( T sat  T w )

This correlation is very accurate when the following condition is met:

( N  1 )c p ( T sv  T w )  2.0
---------------------------------------------------
-
h fg

6.6 MELTING AND FREEZING


Phase change materials are used in many specialized electronic cooling applications.
Most often, transient power applications such as those used in missiles realize these
benefits. As we have seen, during a phase change a material may absorb great
amounts of power with only a small increase in temperature. For an application such
as a missile, an onboard package of phase change material may absorb the heat
given off by an electronic package, without the need for a dedicated cooling system.
The phase change material absorbs the heat and melts. Depending on the quantity of
heat and the mass of phase change material, the cooling effect will last until all of
the material has melted. The opposite effect, freezing, occurs when the material gives
up its latent heat to the surroundings. Applications for the freezing effect might be
seen in missiles that travel outside the atmosphere for a short time, where the elec-
tronics package may need to be protected from the extreme cold of space until re-entry.
The quantity of heat that a phase change material can absorb can be found by
the equation

q  m a m  h m c p, s ( T m  T i ) c p, l ( T 2  T m )

where:

q  quantity of heat stored (W)


m  mass of phase change material (kg/m3)
am  fraction melted (%)
hm  latent heat of fusion (J/kg)
T  absolute temperature (K)
Ti  initial absolute temperature (K)
Tm  melting point absolute temperature (K)
T2  final absolute temperature (K)
cp  specific heat (J/kgK)
cp,s  average specific heat between Ti and Tm (J/kgK)
cp,l  average specific heat between Tm and T2 (J/kgK)

© 2001 by CRC PRESS LLC


0082-06.fm Page 339 Wednesday, August 23, 2000 3:54 PM

FIGURE 6.15 Temperature distribution for ice forming on water, with ambient air as a heat
sink.

When a phase change material freezes, the problem can be considered to be one
of boundary condition values, as shown in Figure 6.15. We can see that the solution
to this problem is quite complex. Researchers have obtained only solutions for simple
cases. We can instead use an approximate solution that was obtained by assuming
that the heat capacity of the subcooled solid phase is negligible relative to the latent
heat of solidification. Another simplifying assumption is that the physical properties
are uniform and that the conductance and the heat sink temperature are constant
throughout the process. To find the time to form a solid phase of a specific thickness
we have:

 -----
L
2
L
 2k- --- U
-
t  T T -
---------------------
 -------------------
s o
-
  h fs 

REFERENCES
1. Hetsroni, G., Ed., Handbook of Multiphase Systems, Hemisphere, Washington, D.C.,
1982.
2. Bergles, A. E. and Rohsenow, W. M., The Determination of Forced Convection
Surface Boiling Heat Transfer, J. Heat Transfer, 86, 365, 1964.
3. Cole, R. and Rohsenow, W. M., Correlation of Bubble Diameters of Saturated Liquids,
Chem. Eng. Prog. Symp. Ser., 65(92), 211, 1969.
4. Cole, R., Photographic Study of Boiling in Region of Critical Heat Flux, AIChE J.,
6, 533, 1960.

© 2001 by CRC PRESS LLC


0082-06.fm Page 340 Wednesday, August 23, 2000 3:54 PM

340 Thermal Design of Electronic Equipment

5. Graham, R. W. and Hendricks, R. C., Assessment of Convection, Conduction and


Evaporation in Nucleate Boiling, NASA TND-3943, National Technical Information
Service, Springfield, VA, 5/1967.
6. Bromley, L. A., Leroy, N. R., and Robbers, A., Heat Transfer in Forced Convection
Film Boiling, Ind. Eng. Chem., 45, 2639, 1953.
7. Zuber, N., Tribus, M., and Westwater, J. W., The Hydrodynamic Crisis in Pool Boiling
of Saturated and Subcooled Liquids, Paper 27, Int. Dev. in Heat Transfer, ASME,
New York, 1961.
8. Florschuetz, L. W. and Chao, B. T., On the Mechanics of Vapor Bubble Collapse, J.
Heat Transfer, 87, 209, 1965.
9. Plesset, M. S. and Zwick, S. A., A Non-Steady Heat Diffusion Problem with Spherical
Symmetry, J. Appl. Phys., 23, 95, 1952.
10. Nukiyama, S., The Maximum and Minimum Values of Heat Transmitted from Metal
to Boiling Water Under Atmospheric Pressure, J. Jpn. Soc. Mech. Eng., 37, 367, 1934.
(Translation: Int. J. Heat Mass Transfer, 9, 1419, 1966.)
11. Drew, T. B. and Mueller, C., Boiling, Trans. AIChE, 33, 449, 1937.
12. Yamagata, K., Kirano, F., Nishiwaka, K., and Matsuoka, H., Nucleate Boiling of Water
on the Horizontal Heating Surface, Mem. Fac. Eng. Kyushu Imp. Univ., 15, 98, 1955.
13. Rohsenow, W. M., A Method of Correlating Heat Transfer Data for Surface Boiling
of Liquids, Trans. ASME, 74, 969, 1952.
14. Collier, J. G., Convective Boiling and Condensation, 2nd ed., McGraw-Hill, New
York, 1981.
15. Danielson, R. D., Tousignant, L., and Bar-Cohen, A., Saturated Pool Boiling of
Commercially Available Perfluorinated Inert Liquids, Proc. ASME-JJME Therm. Eng.
Joint Conf., 1987, 419.
16. Hetsroni, G., Ed., Handbook of Multiphase Systems, Hemisphere, Washington, D.C.,
1982.
17. Kutateladze, S. S., On the Transition to Film Boiling Under Natural Convection,
Kotloturbostroenie, 3, 10, 1948.
18. Zuber, N., Hydrodynamic Aspects of Nucleate Boiling, Ph.D. dissertation, University
of California, Los Angeles, 1959.
19. Lienhard, J. H. and Dhir, V. K., Hydrodynamic Prediction of Peak Pool Boiling Heat
Fluxes from Finite Bodies, J. Heat Transfer, 95, 152, 1973.
20. Cichelli, M. T. and Bonilla, C. F., Heat Transfer to Liquids Boiling Under Pressure,
Trans. AIChE, 41, 755, 1945.
21. Ellion, M. E., A Study of the Mechanism of Boiling Heat Transfer, Mem. 20–88, Jet
Propulsion Laboratory, Pasadena, CA, 3/1954.
22. Bernath, L., A Theory of Local-Boiling Burnout and Its Application to Existing Data,
Chem. Eng. Prog. Symp. Ser., 56(30), 95, 1960.
23. Zuber, N., On the Stability of Boiling Heat Transfer, Trans. ASME, 80, 711, 1958.
24. Kovalev, S. A., An Investigation of Minimum Heat Fluxes in Pool Boiling of Water,
Int. J. Heat Mass Transfer, 9, 1219, 1966.
25. Berenson, P. J., Film Boiling Heat Transfer for a Horizontal Surface, J. Heat Transfer,
83, 351, 1961.
26. Lienhard, J. H. and Wong, P. T. Y., The Dominant Unstable Wavelength and Minimum
Heat Flux During Film Boiling on a Horizontal Cylinder, J. Heat Transfer, 86, 220,
1964.
27. Bromley, L. A., Heat Transfer in Stable Film Boiling, Chem. Eng. Prog., 46, 221,
1950.

© 2001 by CRC PRESS LLC


0082-06.fm Page 341 Wednesday, August 23, 2000 3:54 PM

Heat Transfer with Phase Change 341

28. Frederking, T. H. K. and Daniels, D. J., The Relation Between Bubble Diameter and
Frequency of Removal From a Sphere During Film Boiling, J. Heat Transfer, 88, 87,
1966.
29. Mills, A. F., Heat and Mass Transfer, Irwin, Chicago, 1995, 640.
30. Frederking, T. H. K. and Clark, J. A., Natural Convection Film Boiling on a Sphere,
Adv. Cryog. Eng., 8, 501, 1962.
31. Lienhard, J. H. and Eichhorn, R., Peak Boiling Heat Flux on Cylinders in a Cross
Flow, Int. J. Heat Mass Transfer, 19, 1135, 1976.
32. Klimenko, V. V., A Generalized Correlation For Two-Phase Forced Flow Heat Trans-
fer, Int. J. Heat Mass Transfer, 31, 541, 1988.
33. Gilliland, E. R. and Sherwood, T. K., Diffusion of Vapors into Air Streams, Ind. Eng.
Chem., 26, 516, 1934.
34. Rohsenow, W. M. and Choi, H., Heat, Mass, and Momentum Transfer, Prentice-Hall,
Englewood Cliffs, NJ, 1961.
35. Chun, K. R. and Seban, R. A., Heat Transfer to Evaporating Liquid Films, J. Heat
Transfer, 93, 391, 1971.
36. Addoms, J. N., Heat Transfer at High Rates to Water Boiling Outside Cylinders,
D. Sc. thesis, Massachusetts Institute of Technology, Cambridge, MA, 1948.
37. Nusselt, W., Die Oberflachenkondensation des Wasserdampfes, Z. Vergle. D-Ing., 60,
541, 1916.
38. Rohsenow, W. M., Heat Transfer and Temperature Distribution in Laminar Film
Condensation, Trans. ASME, 78, 1645, 1956.
39. Sadasivan, P. and Lienhard, J. H., Sensible Heat Correction in Laminar Film Boiling
and Condensation, J. Heat Transfer, 109, 545, 1987.
40. Chen, M. M., An Analytical Study of Laminar Film Condensation. I. Flat Plates.
II. Single and Multiple Horizontal Tubes, Trans. ASME, Ser. C., 83, 48, 1961.

© 2001 by CRC PRESS LLC


0082-06.fm Page 342 Wednesday, August 23, 2000 3:54 PM

© 2001 by CRC PRESS LLC


0082-07.fm Page 343 Wednesday, August 23, 2000 10:12 AM

7 Combined Modes of
Heat Transfer for
Electronic Equipment

7.1 INTRODUCTION
Although we have examined the three modes of heat transfer in great detail, in
engineering practice we usually see cases where modes are combined simultaneously.
Radiation is not usually a significant factor, but to achieve the best result we should
calculate the importance of each mode. In a computer chip, for example, heat is conducted
in parallel paths from the junction to the case and leads. Heat is then conducted from
the leads to the circuit board, and from the case to a heat sink. Simultaneously, the
heat in the leads and in the heat sink is convected to the air and radiated to the ambient
environment.
Table 7.1 is a reminder of the equations used for heat transfer and thermal resistance
in the three modes.
The simplest way to solve most problems of combined modes is to set up a
resistance network. In this manner we can graphically examine the paths of each
mode for simultaneous, parallel, and series heat transfer.

7.2 CONDUCTION IN SERIES AND IN PARALLEL


Because combined modes of heat transfer often occur in parallel, reviewing the
theory of series and parallel combinations may be helpful as applied to conduction.
When heat is conducted through a single wall of a single material, the rate of heat
conduction and the thermal gradient are constant. However, when heat is conducted
in a series path of different materials, the temperature gradient is different for each
material. Examining the composite wall of three materials in series, as shown in
Figure 7.1, the rate of heat conduction through the entire block, qk, can be found by

T1  T4
q k  -----------------------------------------------------
------   ------   ------
L L L
  kA A   kA B  kA C

© 2001 by CRC PRESS LLC


0082-07.fm Page 344 Wednesday, August 23, 2000 10:12 AM

TABLE 7.1
Thermal Resistances and Heat Transfer Rates
Heat Transfer Mode Rate of Heat Transfer, q Thermal Resistance, 

Conduction

k L
q k  ( T 1  T 2 ) ---  k  -----
kA
-
L

Convection

1
qc  hc A ( T s  T  )  c  ---------
hc A

Radiation

T1  T2
q r  A 1  1, 2  ( T 1  T 2 )  r  ---------------------------------------------
-
4 4
A 1  1, 2  ( T 1  T 2 )
4 4

If we have an indeterminate number, N, of individual materials, n, in series, the


rate of heat conduction is
T T1  TN  1
q k  -----------
-  ------------------------------
-
n  1
L n N  L 
------ ------
 kA n  kA n

Since  T is the overall temperature difference across the entire composite, we


often call it the potential temperature. If we describe the thermal resistance to heat

© 2001 by CRC PRESS LLC


0082-07.fm Page 345 Wednesday, August 23, 2000 10:12 AM

FIGURE 7.1 Thermal conduction through a three-layer system in series.

conduction as  k  L/ Ak , then we can rewrite the equation as

T1  TN  1 T
q k  --------------------------
-  --------------------------
-
 n  1 k,n  n  1 k,n
n N n N


From this equation we see that the flow of heat is proportional to the temperature
potential. From this basic understanding of series heat flow and resistance, we can
determine the characteristics of parallel heat flow through different materials. We
know that the total rate of heat flow, qk, through two different materials, A and B,
in a parallel path is the sum of the heat flows, qk  qA  qB. More comprehensively,
to account for each area of heat flow we have

T1  T2 T1  T2 T1  T2
q k  -------------------
 ------
L  -------------------  --------------------
 ------
L 1 2 
 kA A  kA B
 ------------------
-
  1   2

For the more common problem of combinations of series and parallel heat flows,
such as shown in Figure 7.2, we see that the thermal resistance for the material in
parallel, Section 2, is

B C
 2  ------------------
B  C
-

© 2001 by CRC PRESS LLC


0082-07.fm Page 346 Wednesday, August 23, 2000 10:12 AM

FIGURE 7.2 Thermal conduction through a wall consisting of series and parallel paths of
heat flow.

and the rate of heat flow is found by


T
q k  ----------------------
-
 n  1 n
n3

For the composite shown in Figure 7.2, N  3; for the number of layers in series,
n is the thermal resistance of the nth layer; and T is the overall temperature
difference across the exterior walls.

7.3 CONDUCTION AND CONVECTION IN SERIES


In electronics cooling problems involving conduction and convection modes of heat
transfer in series, such as shown in Figure 7.3, we must usually determine the
temperature increase of a device when we know the heat rate. We can easily add
convection to a series conduction model by using the convection resistance term
1
 c  ---------
hc A
Figure 7.4 depicts an example of heat being transferred from a hot fluid to a
cold fluid, through a wall. Both sides of the wall receive a different rate of convective
heat transfer. We can describe the rate of heat transfer as
Th  Tc T
q  ---------------------
-  -------------------------------
  2  3
n  1 i
n3
 1

© 2001 by CRC PRESS LLC


0082-07.fm Page 347 Wednesday, August 23, 2000 10:12 AM

FIGURE 7.3 Conduction and convection in an electronic module. The silicon die is encapsu-
lated in an epoxy foam insulator case. The majority of the heat transfer is through the die surface.

FIGURE 7.4 Heat transfer from a hot gas to a cold gas through a plane wall.

where:

1
 1  -------------------
( h c A ) hot
L
 2  -----
kA
-

1
 3  --------------------
( h c A ) cold

© 2001 by CRC PRESS LLC


0082-07.fm Page 348 Wednesday, August 23, 2000 10:12 AM

qk

Wall

FIIGURE 7.5 Heat transfer from a hot gas to a cold gas through a plane wall, showing a
parallel path for radiation.

7.4 RADIATION AND CONVECTION IN PARALLEL


Many problems in electronic cooling involve the combined effect of simultaneous
radiation and convection heat transfer. The total rate of heat transfer for the system
shown in Figure 7.5 is the sum of the radiation and convection effects, q  qr  qc,
which we can also write as

q  hr A ( T 1  T 2 )  hc A ( T 1  T 2 )  ( hr  hc ) A ( T 1  T 2 )

The radiation heat transfer coefficient is the equation

(T 1  T 2)
4 4
qr
hr  -------------------------
-   1, 2 ---------------------------
-
A1 T 1  T 2 T1  T2

In most systems, determining the radiant heat transfer coefficient directly is very
difficult. Since the temperature factor T contains the temperatures of the emitter
and the receiver, we can evaluate it only when we know both. In electronics cooling,
the temperature of the emitter usually varies with power; therefore, we must estimate
a value for the emitter and then reiterate until the solution converges in steady state.

7.5 OVERALL HEAT TRANSFER COEFFICIENT


The overall heat transfer coefficient, U, is used to describe the result of multiple
convection coefficients as a single value. A common form of heat transfer used in heat
exchangers is to transfer heat from a higher-temperature fluid to a lower-temperature

© 2001 by CRC PRESS LLC


0082-07.fm Page 349 Wednesday, August 23, 2000 10:12 AM

fluid when the fluids are separated by a wall. If we know that the wall is plane and
there is only convection on both sides, we can find the rate of heat transfer using
the following equation:

Th  Tc T
q  -----------------------------------------------------
 --------
1 
-  -------------------------------
1  2  3
-   ------   ---------
L 1
 
hc A   kA  hc A 
h c

This equation describes the flow of heat only in terms of a temperature potential
difference and the thermal transfer characteristics of each section in the heat flow
path. In some instances it is more helpful to describe the heat flow as a single value
in terms of the resistances (or conductances). Rewriting the equation in terms of an
overall value, we obtain

q  UA  T

where:

T  total temperature difference (˚C)


UA  -------------------------------
1 -  ----------
1-
1  2  3  TOT

The area that UA is based on should be stated in the problem to avoid uncertainty.
If the chosen area is small, it may not reflect the overall value of the system but,
instead, of only a local area.
We can describe the overall heat transfer coefficient on the outside area, Ao, of
a tube by the equation:

1
U o  -----------------------------------------------------------------
r
A o ln  ----o-
Ao ri 1
------------  ------------------- -  ------
Ai hi 2 kL ho

Conversely, we can determine the heat transfer coefficient on the inside area, Ai,
of a tube with the equation:

1
U i  -----------------------------------------------------------------
r
A i ln  ----o-
Ai ri 1
--------------  -------------------  -----
Ao ho 2 kL hi

The overall heat transfer coefficient can also be used to describe radiation heat
transfer. In this case, the fluid would be a gas. If we examine a plane wall that
separates a hot gas from a cold gas, as shown in Figure 7.4, we see that heat is
transferred into the wall by both convection and radiation. We can describe this

© 2001 by CRC PRESS LLC


0082-07.fm Page 350 Wednesday, August 23, 2000 10:12 AM

section of the heat transfer model as

q  q c  q r  h c A ( T g  T sg )  h r A ( T g  T sg )
T g  T sg
 ( h c1  h r1 ) A ( T g  T sg )  --------------------
1
-

where:

Tg  T1  temperature of hot gas (˚C)


Tsg  T2  temperature
4 4
of hot wall surface (˚C)
 A
( T g  T sg )
hr1  -------------------------------------
T g  T sg  radiation heat transfer coefficient (W/m2 K), where
1.0
hc1  convective coefficient from the gas to the wall surface ( W/m 2 K )
1
 1  ----------------------------
(h  h ) A
-  combined thermal resistance of Section 1, Figure 7.2 (K/W)
r cl

After the heat is transferred into the wall, it is transferred through the wall by
conduction. The heat transfer through this second section can be written as

kA T sg  T sc
q  q k  ------ ( T sg  T sc )  ----------------------
2
-
L

where:

Tsc  temperature of the cool wall surface (˚C)


1  thermal resistance of Section 2 (K/W)

After the heat is transferred through the wall, it is transferred into the cooler gas
by convection. The convection coefficient through this third section can be written as

T sc  T c
q  q c  h c3 A ( T sc  T c )  --------------------
3
-

where:

Tc  temperature of the cooler gas (˚C)


3  thermal resistance of Section 3 (K/W)

© 2001 by CRC PRESS LLC


0082-APP Page 351 Wednesday, August 23, 2000 9:54 AM

Appendix

© 2001 by CRC PRESS LLC


0082-APP Page 353 Wednesday, August 23, 2000 9:54 AM
© 2001 by CRC PRESS LLC

353
TABLE A1
Air at Sea-Level Atmospheric Pressure
Temp. Density Coef. Exp. Specific Heat Thermal Cond. Absolute Viscosity Kinematic Viscosity Prandtl Number
T   103 cp k  106 v 106 Pr
°F °C kg/m3 1/K J/kg K W/m K N s/m2 m2/s –

32 0 1.293 3.664 1003.9 0.02417 17.17 13.28 0.7131


41 5 1.269 3.598 1004.3 0.02445 17.35 13.67 0.7127
50 10 1.242 3.533 1004.6 0.02480 17.58 14.16 0.7122
59 15 1.222 3.470 1004.9 0.02512 17.79 14.56 0.7118
68 20 1.202 3.412 1005.2 0.02544 18.00 14.98 0.7113
77 25 1.183 3.354 1005.4 0.02577 18.22 15.40 0.7108
86 30 1.164 3.298 1005.7 0.02614 18.46 15.86 0.7103
95 35 1.147 3.244 1006.0 0.02650 18.70 16.30 0.7098
104 40 1.129 3.193 1006.3 0.02684 18.92 16.76 0.7093

Thermal Design of Electronic Equipment


113 45 1.111 3.142 1006.6 0.02726 19.19 17.27 0.7087
122 50 1.093 3.094 1006.9 0.02761 19.42 17.77 0.7082
131 55 1.079 3.048 1007.3 0.02801 19.68 18.24 0.7077
140 60 1.061 3.003 1007.7 0.02837 19.91 18.77 0.7072
149 65 1.047 2.957 1008.0 0.02876 20.16 19.26 0.7067
158 70 1.030 2.914 1008.4 0.02912 20.39 19.80 0.7062
167 75 1.013 2.875 1008.8 0.02945 20.60 20.34 0.7057
176 80 1.001 2.834 1009.3 0.02979 20.82 20.80 0.7053
185 85 0.986 2.795 1009.8 0.03012 21.02 21.32 0.7048
194 90 0.972 2.755 1010.3 0.03045 21.23 21.84 0.7044
203 95 0.959 2.718 1010.7 0.03073 21.41 22.33 0.7041
212 100 0.947 2.683 1011.2 0.03101 21.58 22.79 0.7038
0082-APP Page 354 Wednesday, August 23, 2000 9:54 AM
© 2001 by CRC PRESS LLC

TABLE A2
Water at Sea-Level Atmospheric Pressure
Temp. Density Coef. Exp. Specific Heat Thermal Cond. Absolute Viscosity Kinematic Viscosity Prandtl Number
T    103 cp k  106 v 106 Pr
°F °C kg/m3 1/K J/kg K W/m K N s/m2 m2/s –

32 0 999.9 0.068 4217.5 0.5580 1794 1.794 13.56


41 5 1000 0.018 4202.7 0.5677 1530 1.530 11.33
50 10 999.7 0.095 4192.4 0.5774 1296 1.296 9.410
59 15 999.1 0.16 4185.8 0.5870 1136 1.137 8.101
68 20 998.2 0.22 4181.7 0.5967 993 0.995 6.959
77 25 997.1 0.26 4179.5 0.6064 880.6 0.883 6.069
86 30 995.7 0.31 4178.6 0.6155 792.4 0.796 5.380
95 35 994.1 0.35 4178.5 0.6243 719.8 0.724 4.818
104 40 992.2 0.39 4179.0 0.6325 658.0 0.663 4.348
113 45 990.2 0.42 4179.9 0.6401 605.1 0.611 3.951
122 50 988.1 0.45 4181.1 0.6472 555.1 0.562 3.586
131 55 985.8 0.48 4182.6 0.6536 512.6 0.520 3.280
140 60 983.5 0.51 4184.5 0.6594 470.0 0.478 2.983
149 65 980.8 0.54 4186.8 0.6643 436.0 0.445 2.748
158 70 978 0.57 4189.5 0.6686 402.0 0.411 2.519
167 75 974.9 0.60 4192.9 0.6724 376.6 0.386 2.348
176 80 971.7 0.63 4196.6 0.6753 350.0 0.361 2.175
185 85 968.5 0.66 4201.0 0.6778 330.5 0.341 2.048
194 90 965 0.69 4205.7 0.6797 311.0 0.322 1.924
203 95 961.7 0.72 4210.6 0.6811 294.3 0.306 1.819
212 100 958.4 0.75 4215.5 0.6822 277.5 0.290 1.715
0082-APP Page 355 Wednesday, August 23, 2000 3:58 PM
© 2001 by CRC PRESS LLC

TABLE A3
Perflurocarbon FC-72 at Atmospheric Pressure (Boils at 56°C)
Temp. Density Coef. Exp. Specific Heat Thermal Cond. Absolute Viscosity Kinematic Viscosity Prandtl Number
T    103 cp k  106 v  106 Pr
°F °C kg/m3 1/K J/kg K W/m K N s/m2 m2/s –

32 0 1740 1.601 1005.0 0.0600 1009.5 0.5802 16.91


41 5 1727 1.611 1016.2 0.0595 932.4 0.5399 15.93
50 10 1714 1.619 1025.6 0.0590 861.6 0.5027 14.98
59 15 1701 1.626 1033.2 0.0585 799.5 0.4700 14.12
68 20 1688 1.633 1039.8 0.0580 743.0 0.4402 13.32
77 25 1675 1.640 1046.6 0.0575 693.8 0.4142 12.63
86 30 1662 1.647 1053.5 0.0570 648.2 0.3900 11.98
95 35 1649 1.654 1060.8 0.0565 610.1 0.3700 11.46
104 40 1636 1.662 1068.7 0.0560 574.3 0.3510 10.96
113 45 1623 1.670 1077.5 0.0555 543.9 0.3351 10.56
122 50 1610 1.680 1087.0 0.0550 514.8 0.3198 10.17
131 55 1597 1.689 1096.5 0.0545 486.0 0.3043 9.778
0082-APP Page 356 Wednesday, August 23, 2000 9:54 AM
© 2001 by CRC PRESS LLC

TABLE A4
Perflurocarbon FC-77 at Atmospheric Pressure (Boils at 97°C)
Temp. Density Coef. Exp. Specific Heat Thermal Cond. Absolute Viscosity Kinematic Viscosity Prandtl Number
T   103 cp k  106 v 106 Pr
°F °C kg/m3 1/K J/kg K W/m K N s/m2 m2/s –

32 0 1838 1.399 1005 0.0649 2356 1.282 36.48


41 5 1826 1.407 1016 0.0646 2117 1.159 33.30
50 10 1814 1.414 1025 0.0643 1905 1.052 30.37
59 15 1802 1.421 1033 0.0640 1719 0.9539 27.75
68 20 1789 1.429 1041 0.0637 1554 0.8686 25.40
77 25 1777 1.436 1048 0.0634 1413 0.7592 23.36
86 30 1765 1.443 1056 0.0631 1288 0.7298 21.56
95 35 1753 1.451 1063 0.0628 1178 0.6720 19.94
104 40 1740 1.458 1071 0.0625 1083 0.6224 18.56
113 45 1728 1.466 1079 0.0621 1001 0.5793 17.39
122 50 1716 1.473 1087 0.0617 927.0 0.5402 16.33
131 55 1704 1.481 1096 0.0613 862.4 0.5061 15.42
140 60 1691 1.489 1105 0.0609 805.0 0.4761 14.61
149 65 1679 1.497 1114 0.0604 753.2 0.4486 13.89
158 70 1667 1.504 1123 0.0600 706.1 0.4236 13.22
167 75 1655 1.512 1131 0.0595 662.3 0.4002 12.59
176 80 1642 1.520 1140 0.0590 622.1 0.3789 12.02
185 85 1630 1.527 1147 0.0585 584.0 0.3583 11.45
194 90 1618 1.534 1154 0.0580 548.0 0.3387 10.90
203 95 1605 1.541 1159 0.0575 513.2 0.3198 10.34

Source: Data from FluorinertTM Liquids Product Manual, 3M, St. Paul, MN. With permission.
0082-APP Page 357 Wednesday, August 23, 2000 9:54 AM

TABLE A5
Thermophysical Properties of Nonferrous Metals at 20°C
Density Coef. Exp. Specific Heat Thermal Cond.
   106 cp k
Materials kg/m3 1/K J/kg K W/m K

Aluminum (1100) 2713 23.6 921 222


Aluminum (2014) 2796 23.0 921 192
Aluminum (2024) 2768 23.2 921 189
Aluminum (5052) 2685 23.8 921 139
Aluminum (6061) 2713 23.4 963 180
Aluminum (7075) 2796 23.6 963 121
Aluminum (356) 2685 21.4 935 159
Beryllium 1855 11.5 1884 151
Brass (C36000) 8498 20.5 380 116
Bronze (C22000) 8802 18.4 377 189
Copper (C11000) 8913 17.6 383 391
Copper (C12200) 8941 17.6 385 339
Copper (C22000) 8802 18.4 377 189
Copper (alloy MF 202) 8862 17.0 382 150
Glass seal (alloy Ni 50) 8332 8.46 482 10.4
Gold 19,321 14.2 129 313
Inconel (625) 8442 12.8 410 9.82
Kovar 8343 4.30 439 16.0
Lead 11,349 29.3 130 33.9

© 2001 by CRC PRESS LLC


0082-APP Page 358 Wednesday, August 23, 2000 9:54 AM

TABLE A6
Thermophysical Properties of Ferrous Metals at 20°C
Density Coef. Exp. Specific Heat Thermal Cond.
   106 cp k
Materials kg/m3 1/K J/kg K W/m K

Carbon steel (AISI 1010) 7830 6.60 434 64.0


Carbon steel (AISI 1042) 7840 6.50 460 50.0
Cast iron (ASTM A-48) 7197 10.8 544 50.2
Cast iron (ASTM A-220) 7363 13.5 544 51.1
Cast steels (carbon and alloy) 7834 14.7 440 46.7
Stainless steel (4130) 7833 13.5 456 43.3
Stainless steel (17-4 PH) 7778 10.8 461 18.0
Stainless steel (304) 8027 17.3 477 16.3
Stainless steel (316) 2685 16.0 468 16.3
Stainless steel (440) 7750 10.1 461 24.2

TABLE A7
Thermophysical Properties of Plastics at 20°C
Coef. Exp. Specific Heat Thermal Cond.
Density    106 cp k
Materials kg/m3 1/K J/kg K W/m K

ABS (acrylonitrile 1058 72.0 1466 2.70


butadiene styrene)
Acetal 1415 82.8 1465 3.01
Acrylic 1178 81.0 1466 2.49
Alkyd 2206 36.0 1256 9.87
Cellulose acetate 1257 121 1508 3.01
Epoxy (cast) 1148 59.4 1884 4.15
Epoxy (IC molding) 1820 17.0 984 4.00
Fluorocarbon (PTFE) 2196 90.9 1047 2.91
Polyamide (nylon type 6) 1247 89.1 1675 2.08
Phenolic 1387 37.4 1570 1.74
Polycarbonate 1203 67.5 1256 2.39
Polybutylene 1307 72.0 1905 1.90
Terephthalate (PBT)
Polyester 1287 85.5 1780 2.29
Polyimide 1427 47.7 1214 8.05
Polyamide-imide 1397 36.0 – 2.94
Polyetherimide 1277 54.0 1090 2.2
Polyesteretherketone 1317 40.5 – 2.95
Polyetherketone – 103 – –
Polystyrene 1039 72.9 1361 1.54
Polyethylene 933 225 2261 3.95
Polypropylene 903 86.4 1884 2.22
Polyvinyl chloride (PVC) 1447 54.0 1050 1.77

© 2001 by CRC PRESS LLC


0082-APP Page 359 Wednesday, August 23, 2000 9:54 AM

TABLE A8
Thermophysical Properties of Ceramics at 20°C
Density Coef. Exp. Specific Heat Thermal Cond.
   106 cp k
Materials kg/m3 1/K J/kg K W/m K

Aluminum oxide 3982 5.67 879 30.0


Aluminum nitride 3200 4.40 711 200
Beryllium oxide 2900 7.00 1030 300
Boron nitride (cubic) 2200 3.80 709 1300
Diamond (film) 3500 2.00 510 1200
Fused quartz 2200 0.50 745 1.60
Glass (die attach) 2900 50.0
Silicon 2300 4.20 664 83.7
Silicon nitride 3300 2.00 624 21.0

TABLE A9
Properties of Common Gases
 pc  106 Tc (K)   106 T

Gas MW cp/cv atm K N s/m2 Å K

Acetylene C2H2 26.038 1.260 6.1404 308.3 – 4.033 231.8


Air (a) 28.966 1.400 3.6883 132.0 19.3 3.711 78.60
Ammonia NH3 17.031 1.310 11.2777 405.6 32.7 2.900 558.3
Argon Ar 39.948 1.660 4.8738 150.8 26.4 3.542 93.30
Butane C4H10 58.124 1.090 3.7998 425.2 25.0 4.687 531.4
Carbon CO2 44.010 1.285 7.3766 304.2 34.3 3.941 195.2
dioxide
Carbon CO 28.010 1.399 3.4958 132.9 19.0 3.690 91.70
monoxide
Chlorine Cl2 70.096 1.355 7.7009 417.0 42.0 4.217 316.0
Ethane C2H6 30.070 1.183 4.8840 305.4 21.7 4.443 215.7
Ethylene C2H4 28.054 1.208 5.0360 282.4 21.7 4.163 224.7
Fluorine F2 37.997 – 5.2183 144.3 27.5 3.357 112.6
Freon-12 CC12F2 120.914 1.139 4.1240 385.0 – – –
Helium He 4.003 1.667 0.2270 5.190 2.54 2.551 10.22
Hydrogen H2 2.016 1.404 1.2970 33.20 3.47 2.827 59.70
Methane CH4 16.043 1.320 4.6003 190.6 15.9 3.758 148.6
Methanol CH4O 32.042 1.203 8.0960 512.6 39.3 3.626 481.8
Neon Ne 20.183 1.667 2.7561 44.40 16.3 2.820 32.80
Nitrogen N2 28.013 1.400 3.3945 126.2 18.0 3.798 71.40
Nitrous N2O 44.013 1.303 7.2449 309.6 33.2 3.828 232.4
oxide
Octane C8H18 114.232 1.044 4.2845 563.4 24.1 – –
Oxygen O2 32.00 1.395 5.0461 154.6 25.0 3.467 106.7
Pentane C5H12 72.151 1.086 3.3742 469.6 25.0 5.784 34.10
Propane C3H8 44.097 1.124 4.2456 369.8 23.3 5.118 237.1

© 2001 by CRC PRESS LLC


0082-APP Page 360 Wednesday, August 23, 2000 9:54 AM

TABLE A9 (continued)
Properties of Common Gases
 pc  106 Tc (K)   106 T

Gas MW cp/cv atm K N s/m2 Å K

Steam H2O 18.015 1.329 22.049 647.3 54.1 2.641 809.1


Toluene C7H8 92.140 – 4.1139 591.7 127.0 – –
Xenon Xe 131.30 1.660 5.8364 289.7 53.7 4.047 231.0

Note:

Pc  critical pressure (Pa)

MW  molecular weight

Tc  critical temperature (K)

T  effective temperature characteristic force potential (K)

  ratio of specfic heat at constant pressure to specific heat at constant volume at 15°C to 25°C

  molecular collision diameter (Å)

c  viscosity at critical pressure and critical temperature (N s/m2)


a Properties of air based on:
78.084% N2
20.946% O2
0.934% Ar
0.033% CO2
1.0% other, by volume.

TABLE A10
U.S. Standard Atmosphere
Altitude Temperaure Density Pressure Viscosity
H (m) T (K)  (kg/m3) P (N/m2)   106 (N/m2)

100 288.80 1.237 102,534 17.92


0 288.15 1.225 101,327 17.89
100 287.50 1.213 100,131 17.86
200 286.85 1.202 98,948 17.83
400 285.55 1.179 96,613 17.76
600 284.25 1.156 94,324 17.70
800 282.95 1.134 92,079 17.64
1000 281.65 1.112 89,877 17.57
1500 278.40 1.058 84,558 17.41
2000 275.15 1.007 79,497 17.25
2500 271.90 0.9567 74,684 17.09
3000 268.65 0.9091 70,110 16.93
3500 265.40 0.8632 65,766 16.77

© 2001 by CRC PRESS LLC


0082-APP Page 361 Wednesday, August 23, 2000 9:54 AM

TABLE A10 (continued)


U.S. Standard Atmosphere
Altitude Temperaure Density Pressure Viscosity
H (m) T (K)  (kg/m3) P (N/m2)   106 (N/m2)

4000 262.15 0.8191 61,642 16.61


4500 258.90 0.7768 57,730 16.44
5000 255.65 0.7361 54,022 16.28
6000 249.15 0.6597 47,183 15.95
7000 242.65 0.5895 41,062 15.60
8000 236.15 0.5252 35,601 15.26
9000 229.65 0.4664 30,380 14.92
10,000 223.15 0.4127 26,437 14.57
11,000 216.65 0.3639 22,633 14.21
12,000 216.65 0.3108 19,331 14.21
13,000 216.65 0.2655 16,511 14.21
14,000 216.65 0.2268 14,102 14.21
15,000 216.65 0.1937 12,045 14.21
16,000 216.65 0.1654 10,288 14.21
17,000 216.65 0.1413 8,787.1 14.21
18,000 216.65 0.1207 7,505.2 14.21
19,000 216.65 0.1031 6,410.4 14.21
20,000 216.65 0.08803 5,475.2 14.21
25,000 221.65 0.03946 2,511.2 14.48
30,000 226.65 0.01801 1,172.0 14.76
35,000 237.05 0.00821 558.97 15.31
40,000 251.05 0.00385 277.55 16.04
45,000 265.05 0.00188 143.15 16.75
50,000 270.65 0.00097 75.954 17.03
60,000 245.45 0.00028 20.317 15.75
70,000 217.45 0.00007 4.6348 14.26

© 2001 by CRC PRESS LLC

Potrebbero piacerti anche