Logo Studenta

F05_03SI

¡Este material tiene más páginas!

Vista previa del material en texto

blackbody emissive power Wb is given by the
 relation
The preparation of this chapter is assigned to TC 1.3, H
Fluid Flow.
Copyright © 2005, ASHRAE
s not a factor. blackbody. The 
Stefan-Boltzmann
eat Transfer and
motion of free electrons. Thermal energy transfer occurs in the
direction of decreasing temperature. In opaque solid bodies, ther-
mal conduction is the significant heat transfer mechanism because
no net material flows in the process and radiation i
the emissive power depends on the surface characteristics. At a
defined surface temperature, an ideal surface (perfect emitter) emits
the highest amount of energy. Such a surface is also a perfect
absorber (i.e., it absorbs all incident radiant energy) and is called a
energy between parts of a continuum by transfer of kinetic energy
between particles or groups of particles at the atomic level. In
gases, conduction is caused by elastic collision of molecules; in liq-
uids and electrically nonconducting solids, it is believed to be
caused by longitudinal oscillations of the lattice structure. Thermal
conduction in metals occurs, like electrical conduction, through the
3.1
power of the absolute temperature and increase rapidly with tem-
peratures. Unlike conduction and convection, no medium is
required to transmit electromagnetic energy.
Every surface emits energy. The rate of emitted energy per unit
area is the emissive power of the surface. At any given temperature,
temperature) can also serve as point functions.
Thermal Conduction. This heat transfer mechanism transports
tion heat transfer rates are driven primarily by temperature gradients
and somewhat by temperature because of temperature-dependent
properties, radiative heat transfer rates are driven by the fourth
CHAPT
HEAT TR
Heat Transfer Processes ........................................................... 3.1
Thermal Conduction ................................................................. 3.2
Thermal Radiation .................................................................... 3.8
Thermal Convection ................................................................ 3.13
EAT transfer is energy in transit because of a temperature dif-Hference. The thermal energy is transferred from one region to
another by three modes of heat transfer: conduction, radiation, and
convection. Heat transfer is among a group of energy transport phe-
nomena that includes mass transfer (see Chapter 5), momentum
transfer (see Chapter 2), and electrical conduction. Transport phe-
nomena have similar rate equations, in which flux is proportional to
a potential difference. In heat transfer by conduction and convec-
tion, the potential difference is the temperature difference. Heat,
mass, and momentum transfer are often considered together
because of their similarities and interrelationship in many common
physical processes.
This chapter presents the elementary principles of single-phase
heat transfer with emphasis on HVAC applications. Boiling and
condensation are discussed in Chapter 4. More specific information
on heat transfer to or from buildings or refrigerated spaces can be
found in Chapters 25 through 32 of this volume and in Chapter 12
of the 2002 ASHRAE Handbook—Refrigeration. Physical proper-
ties of substances can be found in Chapters 18, 22, 24, and 39 of this
volume and in Chapter 8 of the 2002 ASHRAE Handbook—Refrig-
eration. Heat transfer equipment, including evaporators, condens-
ers, heating and cooling coils, furnaces, and radiators, is covered in
the 2004 ASHRAE Handbook—HVAC Systems and Equipment. For
further information on heat transfer, see the Bibliography.
HEAT TRANSFER PROCESSES
In the applications considered here, the materials are assumed to
behave as a continuum: that is, the smallest volume considered
contains enough molecules so that thermodynamic properties (e.g.,
density) are valid. The smallest length dimension in most engineer-
ing applications is about 100 µm (10–1 mm). At a standard pressure
of 101.325 kPa at 0°C, there are about 3 × 1010 molecules of air in
a volume of 10–3 mm3; even at a pressure of 1 Pa, there are 3 × 107
molecules. With such a large number of molecules in a very small
volume, the variation of the mass of gas in the volume resulting
from variation in the number of molecules is extremely small, and
the mass per unit volume at that location can be used as the density
of the material at that point. Other thermodynamic properties (e.g.,
Related Commercial Resources 
ER 3
ANSFER
Overall Heat Transfer Coefficient ............................................ 3.18
Heat Transfer Augmentation ................................................... 3.19
Heat Exchangers ..................................................................... 3.28
Symbols ................................................................................... 3.34
Fourier’s law for conduction is
(1a)
where
q″ = heat flux (heat transfer rate per unit area A), W/m2
∂t/∂x = temperature gradient, K/m
k = thermal conductivity, W/(m·K)
The magnitude of heat flux q″ in the x direction is directly pro-
portional to the temperature gradient ∂t/∂x. The proportionality fac-
tor is thermal conductivity k. The minus sign indicates that heat
flows in the direction of decreasing temperature. If temperature is
steady, one-dimensional, and uniform over the surface, integrating
Equation (1a) over area A yields
(1b)
Equation (1b) applies where temperature is a function of x only. In
the equation, q is the total heat transfer rate across the area of cross
section A perpendicular to the x direction. Thermal conductivity val-
ues are sometimes given in other units, but consistent units must be
used in Equation (1b).
If A (e.g., a slab wall) and k are constant, Equation (1b) can be
integrated to yield
(2)
where L is wall thickness, t1 is the temperature at x = 0, and t2 is the
temperature at x = L.
Thermal Radiation. In conduction and convection, heat transfer
takes place through matter. In thermal radiation, energy is emitted
from a surface and transmitted as electromagnetic waves, and then
absorbed by a receiving surface. Whereas conduction and convec-
q″ k–
∂t
∂x
-----=
q kA–
∂t
∂x
-----=
q
kA t1 t2–( )
L
--------------------------
t1 t2–
L kA( )⁄
------------------= =
Wb = σT
4
3.2 2005 ASHRAE Handbook—Fundamentals (SI)
 
where σ = 5.67 × 10–8 W/(m2·K4) is the Stefan-Boltzmann con-
stant. Wb is in W/m
2, and T is in K.
The ratio of the emissive power of a nonblack surface to the
blackbody emissive power at the same temperature is the surface
emissivity ε, which varies with wavelength for some surfaces (see
discussion in the section on Actual Radiation). Gray surfaces are
those for which radiative properties are wavelength-independent.
When a gray surface (area A1, temperature T1) is completely
enclosed by another gray surface (area A2, temperature T2) and sep-
arated by a transparent gas, the net radiative heat transfer rate from
surface 1 is
where Wb and ε are the blackbody emissive power and emissivity of
the surfaces. The radiative heat transfer coefficient hr is defined
as
(3a)
For two common cases, Equation (3a) simplifies to 
(3b)
(3c)
Note that hr is a function of the surface temperatures, one of which
is often unknown.
Thermal Convection. When fluid flows are produced by exter-
nal sources such as blowers and pumps, the solid-to-fluid heat trans-
fer is called forced convection. If fluid flow is generated by density
differences caused solely by temperature variation, the heat transfer
is called natural convection. Free convection is sometimes used to
denote natural convection in a semi-infinite fluid.
For convective heat transfer from a solid surface to an adjacent
fluid, the convective heat transfer coefficient h is defined by
where
= heat fluxfrom solid surface to fluid
ts = solid surface temperature, °C
tref = fluid reference temperature for defining convective heat transfer 
coefficient
If the convective heat transfer coefficient, surface temperature,
and fluid temperature are uniform, integrating this equation over
surface area As gives the total convective heat transfer rate from the
surface:
(4)
Combined Heat Transfer Coefficient. For natural convection
from a surface to a surrounding gas, radiant and convective heat
transfer rates are usually of comparable magnitudes. In such a case,
the combined heat transfer coefficient is the sum of the two heat
transfer coefficients. Thus,
h = hc + hr
where h, hc, and hr are the combined, natural convection, and radi-
ation heat transfer coefficients, respectively.
q1
A1 Wb1 Wb2–( )
1 ε1 A1+⁄ A2 1 ε2–( )⁄ ε2⁄
---------------------------------------------------------------=
hr
q1 A1⁄
T1 T2–
-----------------
σ T1
2 T2
2+( ) T1 T2+( )
1 ε1 A1+⁄ A2 1 ε2–( )⁄ ε2⁄
--------------------------------------------------------------- = =
A1 A2: hr
σ T1
2 T2
2+( ) T1 T2+( )
1 ε1 1+⁄ ε2 1–⁄
--------------------------------------------------==
A2 >> A1: hr σε1 T1
2 T 2
2+( ) T1 T2+( )=
q″ h ts tref–( )=
q″
qc hAs ts tref–( )
ts tref–
1 hAs⁄
-----------------= =
Thermal Resistance R. In Equation (2) for conduction in a
slab, Equation (3a) for radiative heat transfer rate between two sur-
faces, and Equation (4) for convective heat transfer rate from a sur-
face, the heat transfer rate can be expressed as a temperature
difference divided by a thermal resistance R. Thermal resistance is
analogous to electrical resistance, with temperature difference and
heat transfer rate instead of potential difference and current,
respectively. All the tools available for solving series electrical
resistance circuits can also be applied to series heat transfer cir-
cuits. For example, consider the heat transfer rate from a liquid to
the surrounding gas separated by a constant cross-sectional area
solid, as shown in Figure 1. The heat transfer rate from the fluid to
the adjacent surface is by convection, then across the solid body by
conduction, and finally from the solid surface to the surroundings
by convection and radiation, as shown in the figure. A series circuit
using the equations for the heat transfer rates for each mode is also
shown.
From the circuit, the heat transfer rate is computed as
For steady-state problems, thermal resistance can be used
• With several layers of materials having different thermal conduc-
tivities, if temperature distribution is one-dimensional
• With complex shapes for which exact analytical solutions are not
available, if conduction shape factors are available
• In many problems involving combined conduction, convection,
and radiation
Although the solutions are exact in series circuits, the solutions
with parallel circuits are approximate because a one-dimensional
temperature distribution is assumed.
Further use of the resistance concept is discussed in the sections
on Thermal Conduction and Overall Heat Transfer Coefficient.
THERMAL CONDUCTION
Steady-State Conduction
One-Dimensional Conduction. Solutions to steady-state heat
transfer rates in (1) a slab of constant cross-sectional area with
parallel surfaces maintained at uniform but different temperatures,
(2) a hollow cylinder with heat transfer across cylindrical surfaces
only, and (3) a hollow sphere are given in Table 1.
Mathematical solutions to a number of more complex heat con-
duction problems are available in Carslaw and Jaeger (1959). Com-
plex problems can also often be solved by graphical or numerical
methods such as described by Adams and Rogers (1973), Croft and
Lilley (1977), and Patankar (1980).
Two- and Three-Dimensional Conduction: Shape Factors.
There are many steady cases with two- and three-dimensional tem-
perature distribution where a quick estimate of the heat transfer rate
Fig. 1 Thermal Circuit
Fig. 1 Thermal Circuit
q
tf 1 tf 2–
1
hA
-------
L
kA
------
1 hc A⁄( ) 1 hr A⁄( )
1 hc A⁄ 1 hr A⁄+
------------------------------------------+ +
---------------------------------------------------------------------=
Heat Transfer 3.3
 
is desired. Conduction shape factors provide a method for getting
such estimates. Heat transfer rates obtained by using conduction
shape factors are approximate because one-dimensional tempera-
ture distribution cannot be assumed in those cases. Using the con-
duction shape factor S, the heat transfer rate is expressed as
q = Sk(t1 – t2) (5)
where k is the material’s thermal conductivity, and t1 and t2 are the
temperatures of two surfaces. Conduction shape factors for some
common configurations are given in Table 2. When using a conduc-
tion shape factor, the thermal resistance is
R = 1/Sk (6)
Transient Conduction
Often, heat transfer and temperature distribution under tran-
sient (varying with time) conditions must be known. Examples are
(1) cold-storage temperature variations on starting or stopping a
refrigeration unit, (2) variation of external air temperature and
solar irradiation affecting the heat load of a cold-storage room or
wall temperatures, (3) the time required to freeze a given material
under certain conditions in a storage room, (4) quick-freezing
objects by direct immersion in brines, and (5) sudden heating or
cooling of fluids and solids from one temperature to another.
For slabs of constant cross-sectional areas, cylinders, and
spheres, analytical solutions in the form of infinite series are avail-
able. For solids with irregular boundaries, use numerical methods.
Lumped Mass Analysis. One elementary transient heat transfer
model predicts the rate of temperature change of a body or material
with uniform temperature, such as a well-stirred reservoir of fluid
whose temperature is a function of time only and spatially uniform
at all instants. Such an approximation is valid if 
Table 1 Heat Transfer Rate and Thermal Resistance for 
Sample Configurations
Configuration
Heat Transfer 
Rate
Thermal 
Resistance
Constant 
cross-
sectional 
area slab
Hollow 
cylinder 
with 
negligible 
heat transfer 
from end 
surfaces
Hollow 
sphere
qx kAx
t1 t2–
L
--------------=
L
kAx
---------
qr
2πkL ti to–( )
ln
ro
ri
----
⎝ ⎠
⎛ ⎞
--------------------------------= R
ln ro ri⁄( )
2πkL
----------------------=
qr
4πk ti to–( )
1
ri
---
1
ro
----+
----------------------------= R
1 ri⁄ 1 ro⁄–
4πk
----------------------------=
where
Bi = Biot number [ratio of (1) internal temperature difference 
required to move energy within the solid of liquid to (2) 
temperature difference required to add or remove the same 
energy at the surface]
h = surface heat transfer coefficient
 V = material’s volume
As = surface area exposed to convective heat transfer
 k = material’s thermal conductivity
The temperature is given by
(7)
where
M = body mass
cp = specific heat at constant pressure
qgen = internal heat generation
qnet = net heat transfer rate to substance (into substance is positive, and 
out of substance is negative)
Equation (7) is applicable when pressure around the substance is
constant; if the volume is constant, replace cp with the constant-
volume specific heat cv. Note that with the density of solids and liq-
uids being substantially constant, the two specific heats are almost
equal. The term qnet may include heat transfer by conduction, con-
vection, or radiation and is the difference between the heat transfer
rates into and out of the body. The term qgen may include a chemical
reaction (e.g., curing concrete) or heat generation from a current
passing through a metal.
A common case for lumped analysis is a solid bodyexposed to a
fluid at a different temperature. The time taken for the solid temper-
ature to change to tf is given by
(8)
where
M = mass of solid
cp = specific heat of solid
A = surface area of solid
h = surface heat transfer coefficient
τ = time required for temperature change
tf = final solid temperature
to = initial uniform solid temperature
t
∞
= surrounding fluid temperature 
Example 1. A 1 mm diameter copper sphere is to be used as a sensing ele-
ment for a thermostat. It is initially at a uniform temperature of 21°C. It
is then exposed to the surrounding air at 20°C. The combined heat
transfer coefficient is h = 60.35 W/(m2·K). Determine the time taken
for the temperature of the sphere to reach 20.9°C. The properties of
copper are
ρ = 8933 kg/m3 cp = 385 J/(kg·K) k = 401 W/(m·K)
Bi = hR /k = 60.35(0.001/2)/401 = 7.5 × 10–5, which is much less than
1. Therefore, lumped analysis is valid.
 M = ρ(4πR3/3) = 4.677 × 10–6 kg
Using Equation (8), τ = 2.778 × 10–4 h = 1 s.
Nonlumped Analysis. In cases where the Biot number is greater
than 0.1, the variation of temperature with location within the mass
must be accounted for. This requires solving multidimensional par-
tial differential equations. Many common cases have been solved
and presented in graphical forms (Jakob 1957; Myers 1971; Sch-
neider 1964). In other cases, it is simpler to use numerical methods
(Croft and Lilley 1977; Patankar 1980). When convective boundary
conditions are required in the solution, values of h based on steady-
Bi
h V As⁄( )
k
---------------------= 0.1≤
Mcp
dt
dτ
----- qnet qgen+=
ln 
tf t∞–
to t∞–
---------------
hAτ
Mcp
----------–=
3.4 2005 ASHRAE Handbook—Fundamentals (SI)
 
Table 2 Conduction Shape Factors
Configuration Shape Factor S, m Restriction
Edge of two adjoining walls 0.54W W > L/5
Corner of three adjoining walls (inner surface at T1 and 
outer surface at T2)
0.15L L << length 
and width of 
wall
Isothermal rectangular block embedded in semi-
infinite body with one face of block parallel to surface 
of body
L > W
L >> d, W, H
Thin isothermal rectangular plate buried in semi-
infinite medium d = 0, W > L
d >> W
W > L
d > 2W
W >> L
Cylinder centered inside square of length L L >> W
W > 2R
Isothermal cylinder buried in semi-infinite medium
L >> R
L >> R
d > 3R
d >> R
L >> d
Horizontal cylinder of length L midway between two 
infinite, parallel, isothermal surfaces L >> d
Isothermal sphere in semi-infinite medium
Isothermal sphere in infinite medium 4πR
2.756L
1 d
W
-----+⎝ ⎠
⎛ ⎞ln
0.59
---------------------------------------
H
d
-----
⎝ ⎠
⎛ ⎞
0.078
πW
ln 4W L⁄( )
-------------------------
2πW
ln 4W L⁄( )
-------------------------
2πW
ln 2πd L⁄( )
---------------------------
2πL
ln 0.54W R⁄( )
---------------------------------
2πL
cosh 1– d R⁄( )
-------------------------------
2πL
ln 2d R⁄( )
-------------------------
2πL
ln
L
R
--- 1
ln L 2d⁄( )
ln L R⁄( )
-----------------------–
------------------------------------------------
2πL
ln
4d
R
------
⎝ ⎠
⎛ ⎞
------------------
4πR
1 R 2d⁄( )–
----------------------------
Heat Transfer 3.5
 
state correlations are often used. However, this approach may not be
valid when rapid transients are involved.
Estimating Cooling Times for One-Dimensional Geometries.
Cooling times for materials can be estimated (McAdams 1954) by
Gurnie-Lurie charts (Figures 2, 3, and 4), which are graphical solu-
tions for heating or cooling of constant cross-sectional area slabs
with heat transfer from only two parallel surfaces, solid cylinders
with heat transfer from only the cylindrical surface, and solid
spheres. These charts assume an initial uniform temperature distri-
bution and no change of phase. They apply to a body exposed to a
constant-temperature fluid with a constant surface heat transfer
coefficient of h.
Using Figures 2, 3, and 4 for constant cross-sectional area solids,
long solid cylinders, and spheres, it is possible to estimate both the
temperature at any point and the average temperature in a homoge-
neous mass of material as a function of time in a cooling process.
The charts can also be used for fluids, which, absent motion, behave
as solids.
Figures 2, 3, and 4 represent four dimensionless quantities: tem-
perature Y, distance n, inverse of the Biot number m, time, and Fo, as
follows:
where
tc = temperature of fluid medium surrounding solid
t1 = initial uniform temperature of solid
t2 = temperature at a specified location n, and dimensionless time Fo
r = distance from midplane (constant cross-sectional area slab), or 
radius (solid cylinder and sphere)
rm = half-thickness (constant cross-sectional area slab), or outer 
radius (solid cylinder and sphere)
k = thermal conductivity of solid
α = thermal diffusivity of solid = k/ρcp
ρ = density of solid
cp = constant pressure specific heat of solid
τ = time
Fo = Fourier number
Approximate Solution for Fo > 0.2. For Fo > 0.2, the transient
temperature is very well approximated by the first term of the series
solution. The single-term approximations for the three cases are:
Constant cross-sectional area solid (thickness 2L):
(9)
Solid cylinder (radius R):
(10)
where J0 is the Bessel function of the first kind, order zero.
Solid sphere (radius R):
(11)
Y
tc t2–
tc t1–
--------------= n r
rm
-----= m k
hrm
---------= Fo ατ
rm
2
------=
Bi
hL
k
------ Fo ατ
L
2
------==
Y c1exp µ1
2
Fo–( )cos µ1n( ) c1
2sin µ1( )
µ1 sin µ1( )cos µ1( )+
--------------------------------------------------==
Bi
hR
k
------- Fo ατ
R
2
------==
Y c1exp µ1
2
Fo–( )J0 µ1n( ) c1
2 Bi sin µ1( )
µ1
2
Bi
2
+( )J0 µ1( )
------------------------------------------==
Bi
hR
k
------- Fo ατ
R
2
------==
Y
c1exp µ1
2
Fo–( )sin µ1n( )
µ1n
--------------------------------------------------------- c1
2 Bi sin µ1( )
µ1 sin– µ1( )cos µ1( )
-----------------------------------------------==
where c1 and µ1 are coefficients in the first term of the series solutions.
Values of c1 and µ1 are given in Table 3 for a few values of Bi.
From a regression analysis (Couvillion 2004), the values of µ1 in
Table 3 can be expressed as
(12)
where the values of curve fit constants a, a1, a2, and a3 are given in
Table 4.
Finding µ1 allows calculation of the constant c1 from Equations
(9), (10), or (11). Note that µ1 is in radians. The Bessel functions in
Equation (10) are available in many software applications.
Example 2. Apples, approximated as .60 mm diameter solid spheres and
initially at 30°C, are loaded into a chamber maintained at 0°C. If the
surface heat transfer coefficient is 14 W/(m2·K), estimate the time
required for the center temperature to reach 1°C.
Properties of apples are
ρ = 830 kg/m3 k = 0.42 W/(m2·K) 
cp = 3600 J/(kg·K) d = 60 mm = 0.06 m 
 h = 14 W/(m2·K)
Assuming that it will take a long time for the center temperature to
reach 1°C, use the one-term approximation Equation (11). From the
values given,
From Equation (11) with lim(sin 0/0) = 1, Y = c1 exp(–µ
2
1 Fo). For
Bi = 1, from Table 3, c1= 1.2732 and µ1 = 1.5708. Thus,
Note that using Equation (12) gives a value of 1.57 for µ1, which is
only 0.05% less than the tabulated value of 1.5708. Also note that Fo =
0.2 corresponds to an actual time of 1280 s.
Although the one-term approximation for temperature distri-
bution is satisfactory for Fo > 0.2, the total heat transferred from
Fo = 0 to Fo = 0.2 as a fraction of the heat transferred from Fo = 0
to Fo = ∞ becomes significant as Bi increases. In those cases,
calculating additionalterms in the series solution may be needed for
satisfactory results.
Multidimensional Temperature Distribution. One-dimensional
transient temperature distribution charts can be used to find the tem-
peratures with two- and three-dimensional temperatures of solids.
For example, consider a solid cylinder of length 2L and radius rm
exposed to a fluid at tc on all sides with constant surface heat transfer
coefficients h1 on the end surfaces and h2 on the cylindrical surface,
as shown in Figure 5.
The two-dimensional, dimensionless temperature Y(x1,r1,τ) can
be expressed as the product of two one-dimensional temperatures
Y1(x1,τ) × Y2(r1,τ), where
Y1 = dimensionless temperature of constant cross-sectional area slab 
at (x1,τ), with surface heat transfer coefficient h1 associated with 
two parallel surfaces
Y2 = dimensionless temperature of solid cylinder at (r1,τ) with 
surface heat transfer coefficient h2 associated with cylindrical 
surface
1
µ1
2
------ a0
a1
Bi
----- a2 exp
a3
Bi
-----
–⎝ ⎠
⎛ ⎞+ +=
 Y
tc t2–
tc t1–
--------------
0 1–
0 30–
---------------
1
30
------= = = n r
rm
-----
0
0.03
---------- 0= = =
Bi
hrm
k
---------
14 0.03×
0.42
---------------------- 1= = =
 α k
ρcp
--------
0.42
830 3600×
--------------------------- 1.406 10 7– m2 s⁄×= = =
Fo 1
µ1
2
-----– ln
Y
c1
-----
1
1.57082
------------------– ln
1 30⁄
1.2732
---------------- 1.476 ατ
rm
2
------
1.406 10 7– τ×
0.032
---------------------------------= = = = =
τ 9448 s=
3.6 2005 ASHRAE Handbook—Fundamentals (SI)
 
Fig. 3 Transient Temperatures for Infinite Cylinder
Fig. 2 Transient Temperatures for Infinite Slab
Fig. 2 Transient Temperatures for Infinite Slab
Fig. 3 Transient Temperatures for Infinite Cylinder
Heat Transfer 3.7
 
Fig. 4 Transient Temperatures for Spheres
From the charts or Equations (9) to (11), if Fo > 0.2, determine Y1 at
(x1/L, ατ/L
2, k/h1L) and Y2 at (r1/L, ατ/r
2
m, k/h2L).
Example 3. A 70 mm diameter by 125 mm high soda can, initially at 30°C,
is cooled in a chamber where the air is at 0°C. The heat transfer coeffi-
cient is 20 W/(m2·K). Determine the maximum temperature in the can
1 h after starting the cooling. Assume the properties of the soda are
those of water, and that the soda inside the can behaves as a solid body.
Because the cylinder is short, the temperature of the soda is affected
by the heat transfer rate from the cylindrical surface and end surfaces. The
slowest change in temperature, and therefore the maximum temperature,
is at the center of the cylinder. Denoting the dimensionless temperature
by Y, 
Y = Ycyl × Ypl
where Ycyl is the dimensionless temperature of an infinitely long 70 mm
diameter cylinder, and Ypl is the dimensionless temperature of a
125 mm thick slab. Each of them is found from the appropriate Biot
and Fourier number. For evaluating the properties of water, choose a
temperature of 15°C and a pressure of 101.35 kPa. The properties of
water are
ρ = 999.1 kg/m3 k = 0.5894 W/(m·K) cp = 4184 J/(kg·K)
α = k/ρ = 1.41 × 10–7 m2/s τ = 3600 s
tc = 0°C t1 = 30°C
1. Determine Ycyl at n = 0.
Table 3 Values of c1 and µ1 in Equations (9) to (11)
Bi
Constant Cross 
Section [Equation (9)]
Solid Cylinder 
[Equation (10)]
Solid Sphere 
[Equation (11)]
c1 µ1 c1 µ1 c1 µ1
0.5 1.0701 0.6533 1.1143 0.9408 1.1441 1.1656
1.0 1.1191 0.8603 1.2071 1.2558 1.2732 1.5708
2.0 1.1785 1.0769 1.3384 1.5995 1.4793 2.0288
4.0 1.2287 1.2646 1.4698 1.9081 1.7202 2.4556
6.0 1.2479 1.3496 1.5253 2.0490 1.8338 2.6537
8.0 1.2570 1.3978 1.5526 2.1286 1.8920 2.7654
10.0 1.2620 1.4289 1.5677 2.1795 1.9249 2.8363
30.0 1.2717 1.5202 1.5973 2.3261 1.9898 3.0372
50.0 1.2727 1.5400 1.6002 2.3572 1.9962 3.0788
Table 4 Curve Fit Constants for µ1
a0 a1 a2 a3
Constant cross-
sectional slab
0.355 172 0.995 309 0.050 421 3.550 265
Solid cylinder 0.136 699 0.497 324 0.036 019 3.963 059
Solid sphere 0.073 707 0.331 704 0.027 464 4.483 979
Fig. 4 Transient Temperatures for Spheres
Fig. 5 A Solid Cylinder Exposed to Fluid
Fig. 5 Solid Cylinder Exposed to Fluid
3.8 2005 ASHRAE Handbook—Fundamentals (SI)
 
Bicyl = hR/k = 20 × 0.035/0.5894 = 1.188
Focyl = ατ/R
2 = (1.41 × 10–7) × 3600/0.0352 = 0.4144
Focyl > 0.2, so use the one-term approximation with Equations
(12) and (10).
From Equation (10), Ycyl = c1 exp(–µ
2
1Focyl)J0(0), so
 
Using the constants from Table 4,
a0 = 0.136 699 a1 = 0.497 324 a2 = 0.036 019 a3 = 3.963 059
µcyl = 1.3042 J0(0) = 1 Ycyl = 0.572
2. Determine Ypl at n = 0.
Bipl = hL/k = 20 × 0.0625/0.5894 = 2.121
Fopl = 1.41 × 10
–7 × 3600/0.06252 = 0.1299
Fopl < 0.2, so the one-term approximation is not valid. Using
Figure 2 (or a commercial software package for more accuracy),
Ypl = 0.9705. Thus,
Note: The solution may not be exact because convective motion of the
soda during heat transfer has been neglected. The example illustrates
the use of the technique.
THERMAL RADIATION
Radiation, one of the basic mechanisms for energy transfer
between different temperature regions, is distinguished from con-
duction and convection in that it does not depend on an intermediate
material as a carrier of energy but rather is impeded by the presence
of material between the regions. Radiant energy transfer occurs
because of energy-carrying electromagnetic waves that are emitted
by atoms and molecules when their energy content changes. The
amount and characteristics of radiant energy emitted by a quantity
of material depend on the material’s nature, microscopic arrange-
ment, and absolute temperature. Although rate of energy emission is
independent of the surroundings, the net energy transfer rate
depends on the temperatures and spatial relationships of the surface
and its surroundings.
Blackbody Radiation
The rate of thermal radiant energy emitted by a surface depends
on its absolute temperature and the surface characteristics. The total
energy emitted per unit time per unit area of a black surface is called
the blackbody emissive power Wb and is given by the Stefan-
Boltzmann law:
Wb = σT 
4 (13)
where σ is the Stefan-Boltzmann constant [5.670 × 10−8 W/
(m2·K4)].
The energy emitted by a body comprises electromagnetic waves
of many different frequencies or wavelengths. Planck showed that
the spectral distribution of the energy radiated by a blackbody is
(14)
where
Wbλ = blackbody spectral (monochromatic) emissive power, W/m
3
λ = wavelength, m
T = temperature, K
c1
2Bicyl sin µcyl( )
µcyl
2 Bicyl
2
+( ) J0 0( )
-----------------------------------------------=
1
µcyl
2
--------- a0
a1
Bicyl
----------- a2 exp
a3
Bicyl
-----------
–⎝ ⎠
⎛ ⎞+ +=
Y 0.572 0.9705× 0.5551
tc t2–
tc t1–
--------------
0 t2–
0 30–
--------------- t2⇒ 16.7°C= = = = =
Wbλ
C1
λ
5
e
C2 λT⁄ 1–( )
------------------------------------=
C1 = first Planck’s law constant = 3.742 × 10
–16 W·m2
C2 = second Planck’s law constant = 0.014 388 m·K
The blackbody spectral emissive power Wbλ is the energy
emitted per unit time per unit surface area at wavelength λ per unit
wavelength interval around λ; that is, the energy emitted per unit
time per unit surface area in the wavelength interval dλ is equal to
Wbλdλ . The Stefan-Boltzmann law can be obtained by integrating
Equation (14) over all wavelengths:
Wien showed that the wavelength λmax, at which the monochro-
matic emissive power is a maximum (not the maximum wave-
length), is given by
(15)
Equation (15) is Wien’s displacement law: the maximum spectral
emissive power shifts to shorter wavelengths as temperature
increases, such that, at very high temperatures, significant emission
eventually occurs over the entire visible spectrum as shorterwave-
lengths become more prominent. For additional details, see Incrop-
era and DeWitt (2002).
Actual Radiation
The blackbody emissive power Wb and blackbody spectral emis-
sive power Wbλ are the maxima at a given surface temperature.
Actual surfaces emit less and are called nonblack. The emissive
power W of a nonblack surface at temperature T radiating to the
hemispherical region above it is given by
W = εσT 4 (16)
where ε is the total hemispherical emissivity. The spectral emis-
sive power Wλ of a nonblack surface is given by
(17)
where ελ is the spectral hemispherical emissivity. The relationship
between ε and ελ is given by
or
(18)
If ελ does not depend on λ, then, from Equation (18), ε = ελ, and
the surface is called gray. Gray surface characteristics are often
assumed in calculations. Several classes of surfaces approximate
this condition in some regions of the spectrum. The simplicity is
desirable, but use care, especially if temperatures are high. Grayness
is sometimes assumed because of the absence of information relat-
ing ελ as a function of λ .
When radiant energy reaches a surface, it is absorbed, reflected,
or transmitted through the material. Therefore, from the first law of
thermodynamics,
α + ρ + τ = 1
where
α = absorptivity (fraction of incident radiant energy absorbed) 
Wbλ λd
0
∞
∫ σT
4 Wb= =
λmaxT 2.898 µm·K=
W
λ
ε
λ
 
C1
λ
5(eC2 λT⁄ 1– )
-------------------------------------- =
W εσT 4 W
λ
λd
0
∞
∫ ελWbλ λd
0
∞
∫= = =
ε
1
σT
4
--------- ε
λ
Wbλ λd
0
∞
∫=
Heat Transfer 3.9
 
ρ = reflectivity (fraction of incident radiant energy reflected) 
τ = transmissivity (fraction of incident radiant energy transmitted) 
For an opaque surface, τ = 0 and ρ + α = 1.
Emissivity is a function of the material, its surface condition, and
its surface temperature. Table 5 lists selected values; Modest (2003)
and Siegel and Howell (2002) have more extensive lists.
Kirchhoff’s law relates emissivity and absorptivity of any
opaque surface from thermodynamic considerations; it states that,
for any surface where incident radiation is independent of angle or
where the surface is diffuse, ελ = αλ. If the surface is gray, or the inci-
dent radiation is from a black surface at the same temperature, then
Table 5 Emissivities and Absorptivities of Some Surfaces
Surface
Total Hemispherical 
Emissivity
Solar 
Absorptivity*
Aluminum
Foil, bright dipped 0.03 0.10
Alloy: 6061 0.04 0.37
Roofing 0.24
Asphalt 0.88
Brass
Oxidized 0.60
Polished 0.04
Brick 0.90
Concrete, rough 0.91 0.60
Copper
Electroplated 0.03 0.47
Black oxidized in Ebanol C 0.16 0.91
Plate, oxidized 0.76
Glass
Polished 0.87 to 0.92
Pyrex 0.80
Smooth 0.91
Granite 0.44
Gravel 0.30
Ice 0.96 to 0.97
Limestone 0.92
Marble
Polished or white 0.89 to 0.92
Smooth 0.56
Mortar, lime 0.90
Nickel
Electroplated 0.03 0.22
Solar absorber, electro-
oxidized on copper
0.05 to 0.11 0.85
Paints
Black
Parsons optical, silicone high 
heat, epoxy
0.87 to 0.92 0.94 to 0.97
Gloss 0.90
Enamel, heated 1000 h at 
650 K
0.80
Silver chromatone 0.24 0.20
White
Acrylic resin 0.90 0.26
Gloss 0.85
Epoxy 0.85 0.25
Paper, roofing or white 0.88 to 0.86
Plaster, rough 0.89
Refractory 0.90 to 0.94
Sand 0.75
Sandstone, red 0.59
Silver, polished 0.02
Snow, fresh 0.82 0.13
Soil 0.94
Water 0.90 0.98
White potassium zirconium 
silicate
0.87 0.13
Source: Mills (1999)
*Values are for extraterrestrial conditions, except for concrete, snow, and water.
ε = α as well, but many surfaces are not gray. For most surfaces listed
in Table 5, the total absorptivity for solar radiation is different from
the total emissivity for low-temperature radiation, because the wave-
length distributions are different, and ελ varies with wavelength.
Example 4. In outer space, the solar energy flux on a surface is 1150 W/m2.
Two surfaces are being considered for an absorber plate to be used on
the surface of a spacecraft: one is black, and the other is specially
coated for a solar absorptivity of 0.94 and infrared emissivity of 0.1.
Coolant flowing through the tubes attached to the plate maintains the
plate at 340 K. The plate surface is normal to the solar beam. For each
surface, determine the (1) heat transfer rate to the coolant per unit area
of the plate, and (2) temperature of the surface when there is no coolant
flow.
In space, there is no convection. From an energy balance on the sur-
face, we obtain
• Incident energy flux – energy flux leaving the surface – heat flux to the
coolant = 0
• Energy flux leaving the surface = emitted energy flux + reflected
energy flux
• Heat flux to the coolant = Incident energy flux – emitted energy
flux – reflected energy flux
• Heat flux to the coolant = absorbed energy flux – emitted energy flux
1. For the black surface: 
ε = α = 1, ρ = 0
Absorbed energy flux = 1150 W/m2
At Ts = 340 K, emitted energy flux = Wb = 5.67 × 10
–8 × 3404 =
757.7 W/m2
Heat flux to coolant = 1150 – 757.7 = 392.3 W/m2
For the special surface, use the solar aborptivity to determine the
absorbed energy flux and infrared emissivity to compute the emitted
energy flux.
Heat flux to coolant = 0.94 × 1150 – 0.1 × 5.67 × 10–8 × 3404 =
319.1 1005 W/m2
2. Surface temperature with no coolant flow
Without coolant flow, heat flux to the coolant is zero. Therefore,
absorbed energy flux = emitted energy flux. For the black surface,
For the special surface,
If the material is opaque, as most solids are in the infrared range,
τ = 0 and α + ρ = 1. For a black surface, α = 1, ρ = 0, and τ = 0.
Platinum black and gold black are almost perfectly black and have
absorptivities of about 98% in the infrared region. A large cavity
with a small opening approaches a black body; most of the incident
energy is absorbed by repeated reflection within it and very little
escapes the cavity, so the absorptivity and therefore the emissivity of
such a cavity are close to unity. Certain flat black paints also exhibit
emissivities of 98% over a wide range of conditions. They provide
a much more durable surface than gold or platinum black and are
frequently used on radiation instruments and as standard reference
in emissivity or reflectance measurements.
The foregoing discussion relates to total hemispherical radiation
from surfaces. Energy distribution over the hemispherical region
above the surface also has an important effect on the rate of heat
transfer in various geometric arrangements.
In estimating heat transfer rates between surfaces of different
geometries, radiation characteristics, and orientations, it is usually
assumed that
• All surfaces are gray or black
• Emission and reflection are diffuse
• Properties are uniform over the surfaces
• Absorptivity equals emissivity and is independent of the temper-
ature of the source of incident radiation
1150 5.67 10 8– Ts
4×× Ts⇒ 377.1 K= =
0.94 1150× 0.1 5.67× 10 8– Ts
4×× Ts⇒ 660.8 K= =
3.10 2005 ASHRAE Handbook—Fundamentals (SI)
Fig. 6 Radiation Angle Factors for Various Geometries
 
• The material in the space between the radiating surfaces neither
emits nor absorbs radiation
These assumptions greatly simplify problems, and give good ap-
proximate results in many cases.
Angle Factor
The ratio of radiant energy leaving a surface i and directly inci-
dent on surface k to the total radiant energy leaving i in all directions
is the angle factor Fik; it is also known as view factor, shape factor,
and configuration factor. The angle factor from area Ak to area Ai, Fki,
is similarly defined, merely by interchanging the roles of i and k.
Some of the relations for the angle factor are given below.
Reciprocity relation.
Fik Ai = Fki Ak (19a)
Decomposition relation. For threesurfaces i, j, and k, with Aij
indicating one surface with two parts denoted by Ai and Aj, 
AijFij-k = AiFi-k + AjFj-k (19b)
AkFk-ij = AkFk-i + AkFk-j (19c)
Law of corresponding corners. This law is discussed by Love
(1968) and Suryanarayana (1995). Its use is shown in Example 5.
Summation rule. For an enclosure with N surfaces, some of
which may be inside the enclosure,
(19d)
Note that a concave surface may “see itself,” and Fii = 0 for such a
surface.
Numerical values of the angle factor for common geometries are
given in Figure 6. For equations to compute angle factors for many
configurations, refer to Siegel and Howell (2002).
Example 5. A picture window, 3 m long and 1.8 m high, is installed in a
wall as shown in Figure 7. The bottom edge of the window is on the
floor, which is 6 by 10 m. Denoting the window by 1 and the floor by
234, find F234-1.
Fig. 6 Radiation Angle Factors for Various Geometries
Fi-k
k =1
N
∑ 1=
Heat Transfer 3.11
 
From decomposition rule,
A234 F234-1 = A2F2-1 + A3F3-1 + A4F4-1
By symmetry, A2F2-1 = A4F4-1 and A234-1 = A3F3-1 + 2A2F2-1.
A23F23-15 = A2F2-1+ A2F2-5 + A3F3-1 + A3F3-5
From the law of corresponding corners, A2F2-1 = A3F3-5, so there-
fore A23F23-5 = A2F2-5 + A3F3-1 + 2A2F2-1. Thus,
A234 F234-1 = A3F3-1 + A23F23-15 – A2F2-5 – A3F3-1 = A23F23-15 – A2F2-5
A234 = 60 m
2 A23 = 45 m
2 A2 = 15 m
2
From Figure 6A with Y/X = 10/6 = 1.67 and Z/X = 1.8/4.5 = 0.4,
F23-15 = 0.061. With Y/X = 10/1.5 = 6.66 and Z/X = 1.8/1.5 = 1.2, F2-5 =
0.041. Substituting the values, F234-1 = 1/60 (45 × 0.061 – 15 × 0.041) =
0.036.
Radiant Exchange Between Opaque Surfaces
A surface radiates energy at a rate independent of its surround-
ings and absorbs and reflects incident energy at a rate dependent on
its surface condition. The net radiant heat flux from a surface is
denoted by q″ or q″i, and the net heat transfer rate from a surface by
q or qi. The net heat transfer rate is the difference between the total
radiant energy leaving the surface and radiant incident energy; it is
the rate at which energy must be supplied from an external source to
maintain the surface at a constant temperature.
Several methods have been developed to solve specific radiant
exchange problems. The radiosity method and thermal circuit
method are presented here.
Consider the heat transfer rate from a surface of an n-surface
enclosure with an intervening medium that does not participate in
radiation. All surfaces are assumed gray and opaque. To calculate
the radiation exchange at each surface of the enclosure, two terms
must be defined: the irradiation G (total rate of incident radiant
energy on a surface per unit area) and the radiosity J (total rate of
radiant energy leaving a surface per unit area).
Radiosity Ji is the sum of the energy flux emitted and the energy
flux reflected:
Ji = εiWb + ρiGi (20)
Note that for a black surface with ε = 1 and ρ = 0, J = Wb . Be-
cause transmissivity is zero and the surfaces are gray, reflectivity is
Fig. 7 Diagram for Example 5
Fig. 7 Diagram for Example 5
ρi = 1 – αi = 1 – εi
Thus,
Ji = εiWbi + (1 – εi) Gi (21)
The net radiant energy transfer qi is the difference between the total
energy leaving the surface and the total incident energy:
qi = Ai (Ji – Gi) (22)
Eliminating Gi between Equations (21) and (22),
(23)
Equation (23) can be used only for nonblack surfaces; for a black
surface, the equation is indeterminate. See Equation (27) for black
surfaces.
Radiosity Method. Consider an enclosure of n isothermal sur-
faces with areas of A1, A2, …, An, emissivities of ε1, ε2, …, εn, and
reflectivities of ρ1, ρ2, …, ρn, respectively. Some of them may be at
uniform but different known temperatures and the remaining
surfaces with a uniform but different and known heat fluxes. The
radiant energy flux incident on a surface Gi is the sum of the radiant
energy reaching it from each of the n surfaces:
(24)
Substituting Equation (24) into Equation (21),
(25)
Combining Equations (22) and (24),
(26)
Note that in Equations (25) and (26), the summation includes
surface i.
Equation (25) is written for surfaces with known temperatures,
and Equation (26) for those with known heat fluxes. The resulting
set of simultaneous, linear equations can be solved manually for the
unknown Js if the number of surfaces is small (three or four). The
solution for a larger number of surfaces is more easily accomplished
using a computer.
Once the radiosities (Jis) are known, the net radiant energy trans-
fer by each surface is determined from Equation (26) or, for a non-
black surface, from Equation (23). For a black surface, Ji = Wbi and,
from Equation (26),
(27)
Note that in using the equations, for all black surfaces, J is replaced
by the corresponding blackbody emissive power Wb.
All diffuse radiation processes are included in the aforemen-
tioned enclosure method, and surfaces with special characteristics
are assigned consistent properties. An opening is treated as an
equivalent surface area Ae with a reflectivity of zero. If energy
qi
Wbi Ji–
1 εi–( ) εiAi⁄
-------------------------------=
GiAi Fki Jk Ak Fik Jk Ai
k=1
n
∑=
k=1
n
∑= or Gi Fik Jk
k=1
n
∑=
Ji εiWbi 1 εi–( ) Fik Jk
k=1
n
∑+=
Ji
qi
Ai
----- Fik Jk
k=1
n
∑+=
qi
Ai
----- Wbi Fik Jk
k=1
n
∑–=
3.12 2005 ASHRAE Handbook—Fundamentals (SI)
 
enters the enclosure diffusely through the opening, Ae is assigned
an equivalent temperature and radiation properties; otherwise, its
temperature is taken as zero. If the loss through the opening is
desired, qe is found. A window in the enclosure is assigned its
actual properties.
Thermal Circuit Method. Another method to determine the
heat transfer rate is by the thermal circuits for radiative heat transfer
rates. Heat transfer rates from surface i to surface k and surface k to
surface i are respectively given by
 qi-k = AiFi-k (Ji – Jk) and qk-i = AkFik-i(Jk – Ji)
Using the reciprocity relation AiFi-k = AkFk-i , the net heat transfer
rate from surface i to surface k is
(28)
 Equations (23) and (28) are analogous to the current in a resis-
tance, with the numerators representing the thermal potential differ-
ence, and the denominator representing the thermal resistance. This
analogy can be used to solve radiative heat transfer rates among sur-
faces, as illustrated in Example 6.
A surface in radiant balance is one for which radiant emission is
balanced by radiant absorption; heat is neither removed from nor
supplied to the surface. Reradiating surfaces (insulated surfaces
with qnet = 0) can be treated in Equation (23) as being perfectly
reflective (i.e., ε = 0) and Wbi
 = Ji. After solving for the radiosities,
the equilibrium temperature of such a surface can be found from
Using angle factors and radiation properties as defined assumes
that the surfaces are diffuse radiators, which is a good assumption
for most nonmetals in the infrared region, but poor for highly pol-
ished metals. Subdividing the surfaces and considering the variation
of radiation properties with angle of incidence improves the approx-
imation but increases the work required for a solution. Also note that
radiation properties, such as absorptivity, have significant uncer-
tainties, for which the final solutions should account.
Example 6. Consider a 4 m wide, 5 m long, 2.5 m high room as shown in
Figure 8. Heating pipes, embedded in the ceiling (1), keep its tempera-
ture at 40°C. The floor (2) is at 30°C, and the side walls (3) are at 18°C.
The emissivity of each surface is 0.8. Determine the radiative heat
transfer rate from each surface.
Consider the room as a three-surface enclosure. The corresponding
thermal circuit is also shown. The heat transfer rates are found after
finding the radiosityof each surface. The radiosities are determined by
qik qi-k qk-i– AiFi-k Ji Jk–( )
Ji Jk–
1 AiFi-k⁄
----------------------= = =
Ti
Ji
σ
----
⎝ ⎠
⎛ ⎞
1/4
=
Fig. 8 Diagrams for Example 6
Fig. 8 Diagrams for Example 6
solving the thermal circuit. Note that, at any surface, the algebraic sum
of the heat transfer rates is zero. From Figure 6B,
F1-2 = F2-1 = 0.376
From the summation rule, F1-1 + F1-2 + F1-3 = 1. With F1-1 = 0, 
F1-3 = 1 – F1-2 = 0.624 = F2-3
R1 = = = 0.0125 m
–2 = R2
R3 = = = 0.005 556 m
–2
R12 = = = 0.133 m
–2
R13 = = = 0.080 13 m
–2 = R23
The net heat transfer rate to each surface = 0. Therefore,
Surface 1:
Surface 2:
Surface 3:
Substituting the values and solving for J1, J2, and J3,
J1 = 524.5 W/m
2 J2 = 475.1 W/m
2 J3 = 418.9 W/m
2
Radiation in Gases
Monatomic and diatomic gases such as oxygen, nitrogen, hydro-
gen, and helium are essentially transparent to thermal radiation.
Their absorption and emission bands are confined mainly to the
ultraviolet region of the spectrum. The gaseous vapors of most
compounds, however, have absorption bands in the infrared region.
Carbon monoxide, carbon dioxide, water vapor, sulfur dioxide,
ammonia, acid vapors, and organic vapors absorb and emit signifi-
cant amounts of energy.
Radiation exchange by opaque solids is considered a surface
phenomenon, although radiant energy does penetrate the surface of
all materials. The absorption coefficient gives the rate of exponen-
tial attenuation of the energy. Metals have large absorption coeffi-
cients, and radiant energy penetrates less than 100 nm at most.
Absorption coefficients for nonmetals are lower. Radiation may be
considered a surface phenomenon unless the material is transparent
or translucent. Gases have small absorption coefficients, so the path
length of radiation through gas is very significant.
Beer’s law states that the attenuation of radiant energy in a gas
is a function of the product pg L of the partial pressure of the gas and
the path length. The monochromatic absorptivity of a body of gas of
thickness L is then given by
(29)
1 ε1–
A1ε1
--------------
1 0.8–
20 0.8×
-------------------
1 ε3–
A3ε3
--------------
1 0.8–
45 0.8×
-------------------
1 0.8–
484.4 0.8×
---------------------------
1
A1F1-2
----------------
1
20 0.376×
-------------------------
1
A1F1-3
----------------
1
20 0.624×
-------------------------
Wb1 J1–
R1
---------------------
J2 J1–
R12
----------------
J3 J1–
R13
----------------+ + 0=
Wb2 J2–
R2
---------------------
J1 J2–
R12
----------------
J3 J2–
R23
----------------+ + 0=
Wb3 J3–
R3
---------------------
J1 J3–
R13
----------------
J2 J3–
R23
----------------+ + 0=
Wb1 5.67 10
8–× 313.24× 545.6/m2= =
Wb2 479.2 W/m
2 Wb3 407.7 W/m
2
==
q1
Wb1 J1–
R1
---------------------
545.6 524.5–
0.0125
--------------------------------- 1688 W= = =
q2 328 W q3 2016 W–==
α
λL 1 e
α
λ
L–
–=
Heat Transfer 3.13
 
Because absorption occurs in discrete wavelengths, the absorp-
tivities must be summed over the spectral region corresponding to
the temperature of the blackbody radiation passing through the gas.
The monochromatic absorption coefficient αλ is also a function of
temperature and pressure of the gas; therefore, detailed treatment of
gas radiation is quite complex.
Estimated emissivity for carbon dioxide and water vapor in air at
24°C is a function of concentration and path length (Table 6). The
values are for a hemispherically shaped body of gas radiating to an
element of area at the center of the hemisphere. Among others, Hot-
tel and Sarofim (1967), Modest (2003), and Siegel and Howell
(2002) describe geometrical calculations in their texts on radiation
heat transfer. Generally, at low values of pg L, the mean path length
L (or equivalent hemispherical radius for a gas body radiating to its
surrounding surfaces) is four times the mean hydraulic radius of the
enclosure. A room with a dimensional ratio of 1:1:4 has a mean path
length of 0.89 times the shortest dimension when considering radi-
ation to all walls. For a room with a dimensional ratio of 1:2:6, the
mean path length for the gas radiating to all surfaces is 1.2 times the
shortest dimension. The mean path length for radiation to the 2 by
6 face is 1.18 times the shortest dimension. These values are for
cases where the partial pressure of the gas times the mean path
length approaches zero ( pg L ≈ 0). The factor decreases with increas-
ing values of pg L. For average rooms with approximately 2.4 m ceil-
ings and relative humidity ranging from 10 to 75% at 24°C, the
effective path length for carbon dioxide radiation is about 85% of
the ceiling height, or 2 m. The effective path length for water vapor
is about 93% of the ceiling height, or 2.3 m. The effective emissivity
of the water vapor and carbon dioxide radiating to the walls, ceiling,
and floor of a room 4.9 by 14.6 m with 2.4 m ceilings is in Table 7.
The radiation heat transfer from the gas to the walls is then
(30)
The examples in Tables 6 and 7 and the preceding text indicate
the importance of gas radiation in environmental heat transfer prob-
lems. In large furnaces, gas radiation is the dominant mode of heat
transfer, and many additional factors must be considered. Increased
pressure broadens the spectral bands, and interaction of different ra-
diating species prohibits simple summation of emissivity factors for
the individual species. Non-blackbody conditions require separate
calculations of emissivity and absorptivity. Hottel and Sarofim
(1967) and McAdams (1954) discuss gas radiation more fully.
The examples in Tables 6 and 7 and the preceding text indicate
the importance of gas radiation in environmental heat transfer prob-
lems. In large furnaces, gas radiation is the dominant mode of heat
transfer, and many additional factors must be considered. Increased
pressure broadens the spectral bands, and interaction of different
radiating species prohibits simple summation of emissivity factors
Table 6 Emissivity of CO2 and Water Vapor in Air at 24°C
Path Length,
m
CO2, % by Volume Relative Humidity, %
0.1 0.3 1.0 10 50 100
3 0.03 0.06 0.09 0.06 0.17 0.22
30 0.09 0.12 0.16 0.22 0.39 0.47
300 0.16 0.19 0.23 0.47 0.64 0.70
Table 7 Emissivity of Moist Air and CO2 in Typical Room
Relative Humidity, % εg
10 0.10
50 0.19
75 0.22
q σAwεg Tg
4 Tw
4
–( )=
for the individual species. Non-blackbody conditions require sepa-
rate calculations of emissivity and absorptivity. Hottel and Sarofim
(1967) and McAdams (1954) discuss gas radiation more fully.
THERMAL CONVECTION
Convective heat transfer involves energy transfer by fluid move-
ment and molecular conduction (Burmeister 1983; Kays and Craw-
ford 1993). Consider heat transfer to a fluid flowing inside a pipe. If
the Reynolds number is large enough, three different flow regions
exist. Immediately next to the wall is a laminar sublayer, where
heat transfer occurs by thermal conduction; next to that is a transi-
tion region called the buffer layer, where both eddy mixing and
conduction effects are significant; the final layer, extending to the
pipe’s axis, is the turbulent region, where the dominant mechanism
of transfer is eddy mixing.
In most equipment, the main body of fluid is in turbulent flow,
and the laminar layer exists at the solid walls only. For low-velocity
flow in small tubes, or highly viscous liquids such as glycol (i.e., at
low Reynolds numbers), the entire flow may be laminar with no
transition to turbulent.
Convection is broadly divided into two categories, forced (fluid
flow is caused by external means) and natural (fluid flow is caused
solely by temperature differences inthe fluid).
Although some generalized heat transfer coefficient correlations
have been mathematically derived from fundamentals, they are
often obtained from correlations of experimental data. Normally,
the correlations use appropriate dimensionless numbers, shown in
Table 8, that are derived from dimensional analysis or analogy.
Forced Convection
Forced-air coolers and heaters, forced-air- or water-cooled con-
densers and evaporators, and liquid suction heat exchangers are exam-
ples of equipment that transfer heat primarily by forced convection.
External Flow. When fluid flows over a flat plate, a boundary
layer forms adjacent to the plate. The velocity of the fluid at the
plate surface is zero and increases to its maximum free stream
value just past the edge of the boundary layer (Figure 9). Boundary
layer formation is important because the temperature change from
plate to fluid (thermal resistance) is concentrated there. Where the
boundary layer is thick, thermal resistance is great and the heat
transfer coefficient is small. Flow within the boundary layer im-
mediately downstream from the leading edge is laminar. As flow
Table 8 Dimensionless Numbers Commonly Used in 
Convective Heat Transfer
Name Symbol Value Application
Nusselt 
number
Nu hD/k, hL /k, q″L /∆tk Natural or forced 
convection, boiling or 
condensation
Reynolds 
number
Re GD/µ , ρVL/µ , ρVD/µ Forced convection, boiling 
or condensation
Prandtl 
number
Pr cpµ/k Natural or forced 
convection, boiling or 
condensation
Stanton 
number
St h/Gcp or Nu/Re Pr Forced convection
Grashof 
number
Gr gβρ2∆tL3/µ2 Natural convection.
For ideal gases, β = 1/T.
Peclet 
number
Pe GDcp /k or Re Pr Forced convection
Graetz 
number
Gz GD2cp /kL or Re Pr D/L Laminar forced convection
Rayleigh 
number
Ra gβ ∆tL3/να = Gr Pr Natural convection
3.14 2005 ASHRAE Handbook—Fundamentals (SI)
 
proceeds along the plate, the laminar boundary layer increases in
thickness to a critical value. Then, turbulent eddies develop within
the boundary layer, except for a thin laminar sublayer adjacent to
the plate.
The boundary layer beyond this point is turbulent. The region
between the breakdown of the laminar boundary layer and estab-
lishment of the turbulent boundary layer is the transition region.
Because turbulent eddies greatly enhance heat transport into the
main stream, the heat transfer coefficient begins to increase rapidly
through the transition region. For a flat plate with a smooth leading
edge, the turbulent boundary layer starts at distance xc from the
leading edge where the Reynolds number Re = Vxc/ν is in the range
300 000 to 500 000 (in some cases, it can be higher). In a plate with
a blunt front edge or other irregularities, it can start at much smaller
Reynolds numbers.
Internal Flow. For tubes, channels, or ducts of small diameter
at sufficiently low velocity, the laminar boundary layers on each
wall grow until they meet. This happens when the Reynolds num-
ber based on tube diameter, Re = Vavg D/ν, is less than 2000 to
2300. Beyond this point, the velocity distribution does not
change, and no transition to turbulent flow takes place. This is
called fully developed laminar flow. When the Reynolds number
is greater than 10 000, the boundary layers become turbulent
before they meet, and fully developed turbulent flow is estab-
lished (Figure 10).
The characteristic dimension for internal flow in pipes and tubes
is the inside diameter. For noncircular tubes or ducts, the hydraulic
diameter Dh is used to compute the Reynolds and Nusselt numbers.
It is defined as
(31)
Fig. 9 Boundary Layer Buildup (Vertical Scale Magnified)
Fig. 9 Boundary Layer Build-up (Vertical Scale Magnified)
Fig. 10 Boundary Layer Buildup in Entrance Region of Tube
or Channel
Fig. 10 Boundary Layer Build-up in Entrance Region of 
Tube or Channel
Dh 4
Cross-sectional area for flow
Total wetted perimeter
---------------------------------------------------------------------×=
Inserting expressions for cross-sectional area and wetted perim-
eter of common cross sections shows that the hydraulic diameter is
equal to 
• The diameter of a round pipe
• Twice the gap between two parallel plates
• The difference in diameters for an annulus
• The length of the side for square tubes or ducts
Table 9 lists various forced-convection correlations. In the gen-
eralized, dimensionless correlation, Equation (33), the Nusselt
number (dimensionless heat transfer coefficient) is determined by
flow conditions and by fluid properties as indicated by the Reynolds
and Prandtl numbers. One functional form of Equation (33) is Equa-
tion (34), known as Colburn’s analogy. In that equation, f is the
Fanning friction factor. The Colburn j-factor, Nu/Re Pr1/3, is related
to the friction factor by the interrelationship of the transport of
momentum and energy. These factors are plotted in Figure 11.
Simplified correlations for atmospheric air are also given in
Table 9. Figure 12 gives graphical solutions for water.
With a uniform tube surface temperature and heat transfer coef-
ficient, the exit temperature can be calculated using
(32)
where ti and te are the inlet and exit bulk temperatures of the fluid,
and ts is the pipe surface temperature.
In many cases, the convective heat transfer coefficient varies sig-
nificantly in the direction of flow because of the temperature depen-
dence of the fluid properties. In such cases, it is common to use an
average value of h in Equation (32). Such an average value may be
computed either as the arithmetic mean of h evaluated at the inlet
and exit fluid temperatures or evaluated at the arithmetic mean of
the inlet and exit temperatures. In many cases, the convective heat
transfer coefficient is approximately a linear function of the fluid
bulk temperature. In such cases, the inlet and exit temperatures are
related by
where hi and he are the convective heat transfer coefficients at the
inlet and bulk temperatures respectively, and hs is evaluated at the
surface temperature.
With uniform surface heat flux q″, the temperature of fluid at any
section can be found by applying the first law of thermodynamics:
(33)
Fig. 11 Typical Dimensionless Representation of Forced-Con-
vection Heat Transfer
Fig. 11 Typical Dimensionless Representation of Forced-
Convection Heat Transfer
ln
ts te–
ts ti–
-------------
hA
m· cp
---------–=
ln
hi ts te–( )
he ts ti–( )
-----------------------
hs A
m· cp
---------
–=
m· cp t ti–( ) q″A=
Heat Transfer 3.15
 
Table 9 Forced-Convection Correlations
I. General Correlation Nu = f (Re, Pr) (T9.1)
II. Internal Flows for Pipes: Characteristic length = D, pipe diameter, or Dh, hydraulic diameter.
 where = mass flow rate, Q = volume flow rate, Pwet = wetted perimeter, 
Ac = cross-sectional area, and ν = kinematic viscosity (µ/ρ).
Colburn’s analogy (T9.2)
Laminar : Re < 2300 (T9.3)a
Developing (T9.4)
Fully developed, round Nu = 3.66 Uniform surface temperature
Nu = 4.36 Uniform heat flux
Turbulent: 
Fully developed
Evaluate properties at bulk 
temperature tb except µs and 
ts at surface temperature
Nu = 0.023 Re4/5Pr 0.4 Heating fluid 
Re ≥ 10 000
(T9.5)b
Nu = 0.023 Re4/5Pr 0.3 Cooling fluid 
Re ≥ 10 000
(T9.6)b
(T9.7)c
For fully developed flows, set D/L = 0. Multiply Nu by (T/Ts)
0.45 for gases 
and by (Pr/Prs)
0.11 for liquids
For viscous fluids (T9.8)a
For noncircular tubes, use hydraulic mean diameter Dh in the equations for Nu for an approximate value of h.
III. External Flows for Flat Plate: Characteristic length = L = length of plate. Re = VL/ν.
All properties at arithmetic mean of surface and fluid temperatures.
Laminar boundary layer: 
Re < 5 × 105
Nu = 0.332 Re1/2Pr1/3 Local value of h (T9.9)Nu = 0.664 Re1/2Pr1/3 Average value of h (T9.10)
Turbulent boundary layer:
Re > 5 × 105
Nu = 0.037 Re4/5Pr1/3 Local value of h (T9.11)
Laminar-turbulent boundary 
layer : Re > 5 × 105
Nu = (0.337 Re4/5 – 871)Pr1/3 Average value Rec = 5 × 10
5 (T9.12)
IV. External Flows for Cross Flow over Cylinder: Characteristic length = D = diameter. Re = VD /ν.
All properties at arithmetic mean of surface and fluid temperatures.
Average value of h (T9.13)d
V. Simplified Approximate Equations: h is in W/(m2·K), V is in m/s, D is in m, and t is in °C.
Flows in pipes
Re > 10 000
Atmospheric air (0 to 200°C): h = (3.76 – 0.00497t)V 0.8/D 0.2
Water (3 to 200°C): h = (1206 – 23.9t)V 0.8/D 0.2
(T9.14)e
Flow over cylinders Atmospheric air: 0°C < t < 200°C, where t = arithmetic mean of air 
and surface temperature.
h = 2.755V 0.471/D 0.529 35 < Re < 5000
h = (4.22 – 0.002 57t)V 0.633/D 0.367 5000 < Re < 50 000
Water: 5°C < t < 90°C, where t = arithmetic mean of water and 
surface temperature.
h = (461.8 – 2.01t)V 0.471/D 0.529 35 < Re < 5000
h = (1012 – 9.19t)V 0.633/D 0.367 5000 < Re < 50 000 (T9.15)f
Sources: aSieder and Tate (1936), bDittus and Boelter (1930), cGnielinski (1990), dChurchill and Bernstein (1977), eBased on Nu = 0.023 Re4/5Pr1/3, fBased on Morgan (1975).
Re
ρVavgDh
µ
----------------------
m· Dh
Acµ
-----------
QDh
Acν
-----------
4m·
µPwet
--------------
4Q
νPwet
--------------= = = = = m·
Nu
Re Pr1/3
-------------------
f
2
---=
Nu 1.86
 Re Pr 
L D⁄
----------------
⎝ ⎠
⎛ ⎞
1/3 µ
µs
-----
⎝ ⎠
⎛ ⎞ 0.14
=
L
D
----
 Re Pr 
8
----------------
µ
µs
-----
⎝ ⎠
⎛ ⎞ 0.42<
Nu 3.66 0.065 D L⁄( )Re Pr
1 0.04 D L⁄( )Re Pr[ ]2/3+
--------------------------------------------------------------+=
Nu
fs 2⁄( ) Re 1000–( )Pr
1 12.7 fs 2⁄( )
1/2 Pr2/3 1–( )+
--------------------------------------------------------------------- 1
D
L
----
⎝ ⎠
⎛ ⎞
2/3 
+= fs
1
1.58 ln Re 3.28–( )2
-------------------------------------------------=
Nu 0.027 Re4/5Pr1/3
µ
µs
-----
⎝ ⎠
⎛ ⎞ 0.14
=
Nu 0.3
0.62 Re1/2Pr1/3
1 0.4 Pr⁄( )2/3+[ ]
1/4
------------------------------------------------ 1
Re
282 000
-------------------
⎝ ⎠
⎛ ⎞
5/8 
+
4/5
+=
3.16 2005 ASHRAE Handbook—Fundamentals (SI)
 
where is the fluid mass flow rate, A is the surface area up to the
point where the bulk temperature is t, and ti is inlet temperature. The
surface temperature can be found using
q″ = h(ts – t) (34)
With uniform surface heat flux, surface temperature increases in the
direction of flow along with the fluid.
Natural Convection
Heat transfer with fluid motion resulting solely from temperature
differences (i.e., from temperature-dependent density and gravity)
is natural (free) convection. Natural-convection heat transfer coef-
ficients for gases are generally much lower than those for forced
convection, and it is therefore important not to ignore radiation in
calculating the total heat loss or gain. Radiant transfer may be of the
same order of magnitude as natural convection, even at room tem-
peratures, and, therefore, both modes must be considered when
computing heat transfer rates from people, furniture, and so on in
buildings (see Chapter 8).
Natural convection is important in a variety of heating and refrig-
eration equipment, such as (1) gravity coils used in high-humidity
Fig. 12 Heat Transfer Coefficient for Turbulent Flow of Water
Inside Tubes
Fig. 12 Heat Transfer Coefficient for Turbulent Flow of 
Water Inside Tubes
m·
Fig. 13 Regimes of Free, Forced, and Mixed Convection—
Flow in Horizontal Tubes
Fig. 13 Regimes of Free, Forced, and Mixed Convection—
Flow in Horizontal Tubes
cold-storage rooms and in roof-mounted refrigerant condensers,
(2) the evaporator and condenser of household refrigerators, (3) base-
board radiators and convectors for space heating, and (4) cooling pan-
els for air conditioning. Natural convection is also involved in heat
loss or gain to equipment casings and interconnecting ducts and pipes.
Consider heat transfer by natural convection between a cold fluid
and a hot vertical surface. Fluid in immediate contact with the sur-
face is heated by conduction, becomes lighter, and rises because of
the difference in density of the adjacent fluid. The viscosity of the
fluid resists this motion. The heat transfer rate is influenced by grav-
itational acceleration g, coefficient of thermal expansion β, viscos-
ity µ, density ρ, specific heat cp, thermal conductivity of the fluid k,
temperature difference ∆t, and characteristic dimension L. These
variables are included in the dimensionless numbers given in the
correlating equations in Table 10. The Nusselt number Nu is a func-
tion of the product of the Prandtl number Pr and Grashof number Gr.
The product of Nu and Pr is the Raleigh number Ra.
Convection from horizontal plates facing downward when
heated (or upward when cooled) is a special case. Because the hot
air is above the colder air, theoretically no convection should occur.
Some convection is caused, however, by secondary influences such
as temperature differences on the edges of the plate. As an approx-
imation, a coefficient of somewhat less than half the coefficient for
a heated horizontal plate facing upward can be used.
Other information on natural convection is available in the Bib-
liography under Heat Transfer, General.
Comparison of experimental and numerical results with existing
correlations for natural convective heat transfer coefficients indicate
that caution should be used when applying coefficients for (isolated)
vertical plates to vertical surfaces in enclosed spaces (buildings).
Altmayer et al. (1983) and Bauman et al. (1983) developed
improved correlations for calculating natural convective heat trans-
fer from vertical surfaces in rooms under certain temperature
boundary conditions.
Natural convection can affect the heat transfer coefficient in the
presence of weak forced convection. As the forced-convection
effect (i.e., the Reynolds number) increases, “mixed convection”
(superimposed forced-on-free convection) gives way to pure forced
convection. In these cases, consult other sources (e.g., Grigull et al.
1982; Metais and Eckert 1964) describing combined free and forced
convection, because the heat transfer coefficient in the mixed-
convection region is often larger than that calculated based on the
natural- or forced-convection calculation alone. Metais and Eckert
(1964) summarize natural-, mixed-, and forced-convection regimes
for vertical and horizontal tubes. Figure 13 shows the approximate
limits for horizontal tubes. Other studies are described by Grigull
et al. (1982).
Example 7. Chilled water at 5°C flows inside a freely suspended 20 mm OD
pipe at a velocity of 2.5 m/s. The surrounding air is at 30°C, 70% rh. The
pipe is to be insulated with cellular glass having a thermal conductivity of
0.045 W/(m·K). Determine the radial thickness of the insulation to pre-
vent condensation of water on the outer surface. 
Solution. In Figure 14,
tf i = 5°C tfo = 30°C di = OD of tube = 0.02 m
ki = thermal conductivity of insulation material = 0.045 W/(m·K)
From the problem statement, the outer surface temperature to of the
insulation should not be less than the dew-point temperature of air. The
dew-point temperature of air at 30°C, 70% rh = 23.93°C. To determine
the outer diameter of the insulation, equate the heat transfer rate per
unit length of pipe (from the outer surface of the pipe to the water) to
the heat transfer rate per unit length from the air to the outer surface:
(35)
to tfi–
1
hi di
----------
1
2ki
------- ln 
do
di
-----+
---------------------------------------
tfo to–
1
hot do
-------------
---------------=Heat Transfer 3.17
 
Heat transfer from the outer surface is by natural convection to air,
so the surface heat transfer coefficient hot is the sum of the convective
heat transfer coefficient ho and the radiative heat transfer coefficient hr .
With an assumed emissivity of 0.7 and using Equation (3c), hr =
4.3 W/(m2·K). To determine the value of do, the values of the heat
transfer coefficient associated with the inner and outer surfaces (hi and
ho, respectively) are needed. Compute the value of hi using Equation
(T9.7). Properties of water at an assumed temperature of 5°C are
ρw = 1000 kg/m
3 µw = 0.001 518 (N·s)/m
2 cpw = 4197 W/(m·K)
kw = 0.5708 W/(m·K) Prw = 11.16
Using Equation (T9.7),
f s = 0.023 11 Nud = 205.6 hi = 5869 W/(m
2·K)
To compute ho using Equation (T10.10), the outer diameter of the
insulation material must be found. Determine it by iteration by assum-
ing a value of do, computing the value of ho, and determining the value
of do from Equation (35). If the assumed and computed values of do are
Characteristic dimension: L = height 109 < Ra < 1012 (T10.3)a
Properties at (ts + t∞)/2 except β at t∞
q″s = constant
Characteristic dimension: L = height
Properties at ts, L/2 – t∞ except β at t∞
Equations (51) and (52) can be used for vertical cylinders if D/L > 
35/Gr1/4 where D is diameter and L is axial length of cylinder
10–1 < Ra < 1012 (T10.4)a
III. Horizontal plate
Characteristic dimension = L = A/P, where A is plate area and P is perimeter
Properties of fluid at (ts + t∞)/2
Downward-facing cooled plate and upward-facing heated plate Nu = 0.96 Ra1/6 1 < Ra < 200 (T10.5)
b
Nu = 0.59 Ra1/4 200 < Ra < 104 (T10.6)
b
Nu = 0.54 Ra1/4 2.2 × 104 < Ra < 8 × 106 (T10.7)
b
Nu = 0.15 Ra1/3 8 × 106 < Ra < 1.5 × 109 (T10.8)
b
Downward-facing heated plate and upward-facing cooled plate Nu = 0.27 Ra1/4 105 < Ra < 1010 (T10.9)
b
IV. Horizontal cylinder
Characteristic length = d = diameter 109 < Ra < 1013 (T10.10)c
Properties of fluid at (ts + t∞)/2 except β at t∞
V. Sphere
Characteristic length = D = diameter Ra < 1011 (T10.11)d
Properties at (ts + t∞)/2 except β at t∞
VI. Horizontal wire
Characteristic dimension = D = diameter 10–8 < Ra < 106 (T10.12)e
Properties at (ts + t∞)/2
VII. Vertical wire
Characteristic dimension = D = diameter; L = length of wire Nu = c (Ra D/L)0.25 + 0.763 c(1/6)(Ra D/L)(1/24) c(Ra D/L)0.25 > 2 × 10–3 (T10.13)e
Properties at (ts + t∞)/2
In both Equations (T10.12) and (T10.13), and
VIII. Simplified equations with air at mean temperature of 21°C: h is in W/(m2·K), L and D are in m, and ∆t is in °C.
Vertical surface 105 < Ra < 109 (T10.14)
Ra > 109 (T10.15)
Horizontal cylinder 105 < Ra < 109 (T10.16)
Ra > 109 (T10.17)
Sources: aChurchill and Chu (1975a), bLloyd and Moran (1974), Goldstein et al. (1973), cChurchill and Chu (1975b), dChurchill (1990), eFujii et al. (1986).
Nu 0.825 0.387Ra
1/6
1 0.492 Pr⁄( )9/16+[ ]
8/27
-----------------------------------------------------------+
⎩ ⎭
⎨ ⎬
⎧ ⎫2
=
Nu 0.825 0.387Ra
1/6
1 0.437 Pr⁄( )9/16+[ ]
8/27
-----------------------------------------------------------+
⎩ ⎭
⎨ ⎬
⎧ ⎫2
=
Nu 0.6 0.387 Ra
1/6
1 0.559 Pr⁄( )9/16+[ ]
8/27
-----------------------------------------------------------+
⎩ ⎭
⎨ ⎬
⎧ ⎫2
=
Nu 2 0.589 Ra
1/4
1 0.469 Pr⁄( )9/16+[ ]
4/9
--------------------------------------------------------+=
2
Nu
------- ln 1 3.3
cRan
------------+⎝ ⎠
⎛ ⎞
=
c 0.671
1 0.492 Pr⁄( ) 9/16( )+[ ]
4/9( )
---------------------------------------------------------------=
n 0.25 1
10 5 Ra( )0.175+
-------------------------------------+=
h 1.33 t∆
L
-----
⎝ ⎠
⎛ ⎞
1/4
=
h 1.26 t∆( )1/3=
h 1.04
T∆
D
-------
⎝ ⎠
⎛ ⎞
1/4
=
h 1.23 t∆( )1/3=
Red
ρvd
µ
--------- 32 944= =
Table 10 Natural Con
I. General relationships Nu = f (Ra
Characteristic length depends on geometry
Ra = Gr Pr
II. Vertical plate
ts = constant Nu 0.6=
vection Correlations
) (T10.1)
10–1 < Ra < 109 (T10.2)a
Gr
gβρ2 T∆ L3
µ2
-------------------------------= Pr
cpµ
k
--------- t∆ ts t∞–= =
8 0.67Ra
1/4
1 0.492 Pr⁄( )9/16+[ ]
4/9
--------------------------------------------------------+
3.18 2005 ASHRAE Handbook—Fundamentals (SI)
 
close to each other, the correct solution has been obtained. Otherwise,
recompute ho using the newly computed value of do and repeat the
process. 
Assume do = 0.05 m. Properties of air at tf = 27°C and 101.325 kPa
are
ρ = 1.176 kg/m3 k = 0.025 66 W/(m·K) µ = 1.858 × 10–5 (N·s)/m2
Pr = 0.729 β = 0.003 299 (at 273.15 + 30 = 293.15 K)
Ra = 71 745 Nu = 7.157 ho = 3.67 W/(m
2·K)
hot = 3.67 + 4.3 = 7.97 W/(m
2·K)
From Equation (35), do = 0.044 28 m. Now, using the new value
of 0.044 28 m for the outer diameter, the new values of ho and hot are
3.78 W/(m2·K) and 8.07 W/(m2·K), respectively. The updated value
of do is 0.044 03 m. Repeating the process, the final value of do =
0.044 01 m. Thus, an outer diameter of 0.045 m (corresponding to an
insulation radial thickness of 12.5 mm) keeps the outer surface tem-
perature at 24.1°C, higher than the dew point. (Another method to
find the outer diameter is to iterate on the outer surface temperature
for different values of do .)
OVERALL HEAT TRANSFER COEFFICIENT
In most steady-state heat transfer problems, more than one heat
transfer mode is involved. The various heat transfer coefficients may
be combined into an overall heat transfer coefficient so that the total
heat transfer rate can be calculated from the terminal temperatures.
The solution to this problem is achieved using thermal circuits.
Consider the steady-state heat transfer from one fluid to another
separated by a solid wall: from the warmer fluid at t1, to the solid
wall, through the wall, then to the colder fluid at t2. The wall surface
temperatures are ts1 and ts2. An overall heat transfer coefficient U
based on the difference between bulk temperatures t1 – t2 of the two
fluids is defined as
q = UA(t1 – t2) (36)
where A is the surface area. Because Equation (36) is a definition of
U, the surface area A on which U is based is arbitrary and should
always be specified in referring to U.
In steady state, heat transfer rates are equal from the warmer fluid
to the solid surface, from one solid surface to the next, and then to
the cool fluid. Temperature drops across each part of the heat flow
path are related to the resistances as
t1 – ts1 = qR1 t1 – ts2 = qR2 ts2 – t2 = qR3
Adding the three equations gives
(37)
The equations are analogous to those for electrical circuits; for
thermal current flowing through n resistances in series, the resis-
tances are additive.
Fig. 14 Diagram for Example 7
Fig. 14 Diagram for Example 7
t1 t2–
q
--------------
1
UA
-------- R1 R2 R3+ += =
Ro = R1 + R2 + R3 + ··· + Rn
Similarly, conductance is the reciprocal of resistance, and for
heat flow through resistances in parallel, the conductances are
additive:
This expression has limited applicability because, with heat
transfer in parallel paths, the assumption of one-dimensional tem-
perature distribution is rarely satisfied.
Thermal resistances for convection and radiation are written as
However, application of these resistances is limited to those
cases where the radiation heat transfer is only between two surfaces
and the temperature differences are not very great. The radiation
coefficient hr is a function of the temperatures, radiation proper-
ties, and geometrical arrangement of the enclosure and the body in
question.
Overall Resistance
Analysis using resistances can be illustrated by considering heat
transfer from outside air to cold water inside an insulated pipe. The
temperature gradients and nature of the resistance analysis are
shown in Figure

Continuar navegando

Materiales relacionados

14 pag.
F05_02SI

User badge image

Apuntes Generales

17 pag.
F05_04SI

User badge image

Apuntes Generales

28 pag.
A03_33IP

User badge image

Apuntes Generales