Transferencia de Calor Analitico PDF
Transferencia de Calor Analitico PDF
Transferencia de Calor Analitico PDF
Heat
Transfer
Je-Chin Han
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts
have been made to publish reliable data and information, but the author and publisher cannot assume
responsibility for the validity of all materials or the consequences of their use. The authors and publishers
have attempted to trace the copyright holders of all material reproduced in this publication and apologize to
copyright holders if permission to publish in this form has not been obtained. If any copyright material has
not been acknowledged please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmit-
ted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented,
including photocopying, microfilming, and recording, or in any information storage or retrieval system,
without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.
com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood
Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and
registration for a variety of users. For organizations that have been granted a photocopy license by the CCC,
a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used
only for identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
Preface
Je-Chin Han
1
Heat Conduction Equations
1.1 Introduction
1.1.1 Conduction
Conduction is caused by the temperature gradient through a solid mate-
rial. For example, Figure 1.1 shows that heat is conducted from the high-
temperature side to the low-temperature side through a building or a
container wall. This is a one-dimensional (1-D) steady-state heat conduction
problem if T1 and T2 are uniform. According to Fourier’s conduction law, the
temperature profile is linear through the plane wall.
dT T1 − T 2
q = −k =k (1.1)
dx L
and
q
q ≡ or q = q Ac
Ac
where q is the heat flux (W/m2 ), q the heat rate (W or J/s), k the thermal
conductivity of solid material (W/m K), Ac the cross-sectional area for
conduction, perpendicular to heat flow (m2 ), and L the conduction length (m).
One can predict heat rate or heat loss through the plane wall by knowing T1 ,
T2 , k, L, and Ac . This is the simple 1-D steady-state problem. However, in real-
life application, there are many two-dimensional (2-D) or three-dimensional
(3-D) steady-state heat conduction problems; there are cases where heat gen-
eration occurs in the solid material during heat conduction; and transient
heat conduction problems take place in many engineering applications. In
addition, some special applications involve heat conduction with moving
boundary. All these more complicated heat conduction problems will be
discussed in the following chapters.
1
2 Analytical Heat Transfer
T1 Ac
T2
q"
x
L
FIGURE 1.1
1-D heat conduction through a building or container wall.
1.1.2 Convection
Convection is caused by fluid flow motion over a solid surface. For example,
Figure 1.2 shows that heat is removed from a heated solid surface to cooling
fluid. This is a 2-D boundary-layer flow and heat transfer problem. According
to Newton, the heat removal rate from the heated surface is proportional
to the temperature difference between the heated wall and the cooling fluid.
The proportional constant is called heat transfer coefficient; and the same
heat rate from the heated surface can be determined by applying Fourier
Conduction Law to the cooling fluid.
Also,
−kf dT
q dy y=0
h= = (1.3)
Ts − T ∞ Ts − T ∞
Velocity or thermal
Air f low boundary layer
T∞
U ∞ T∞ U∞
q′
Ts
Heated surface, As
FIGURE 1.2
Velocity and thermal boundary layer.
Heat Conduction Equations 3
and
q
q ≡ or q = q As
As
1.1.3 Radiation
Radiation is caused by electromagnetic waves from solids, liquid surfaces, or
gases. For example, Figure 1.3 shows that heat is radiated from a solid surface
TABLE 1.1
Typical Values of Heat Transfer Coefficient
Type of convection h, W /m2 · K
Natural convection
Caused by ΔT : air 5
Caused by ΔT : water 25
Forced convection
Caused by fan, blower: air 25–250
Caused by pump: water 50–20,000
Boiling or condensation
Caused by phase change
Water Steam 10,000–100,000
Freon Vapor 2500–50,000
4 Analytical Heat Transfer
′′
qrad
FIGURE 1.3
Radiation from a solid surface.
q = σTs4
and
q
q ≡ or q = q As
As
where ε is the emissivity of the real surface, ε = 0 − 1 (ε metal < nonmetal ε),
Ts the absolute temperature of the surface, K (K = ◦ C + 273.15), σ the Stefan–
Boltzman constant, σ = 5.67 × 10−8 W/m2 K4 , and As the surface area for
radiation (m2 ).
Air flow
T∞ h
′′
qconv ′′
qrad,net
Machine or equipment
Surface A at Ts , ε, As
FIGURE 1.4
Heat transfer between two surfaces involving radiation and convection.
where
qconv = h(Ts − T∞ )
qrad,net = εσTs4 − αεsur σTsur
4
= εσ(Ts4 − Tsur
4
), if ε = α, εsur=1
= εσ(Ts2 + Tsur
2
)(Ts + Tsur )(Ts − Tsur )
= hr (Ts − Tsur )
where
hr = εσ(Ts2 + Tsur
2
)(Ts + Tsur ) (1.7)
Also
q
q ≡ or q = q As
As
where α is the absorptivity, Tsur the surrounding wall temperature (◦ C or
K), εsur the emissivity of the surrounding wall, hr the radiation heat transfer
coefficient (W/m2 K), and As the surface area for radiation (m2 ). Total heat
transfer rate can be determined by knowing Ts , Tsur , h, ε, εsur , σ, and As .
6 Analytical Heat Transfer
where Ein − Eout is the net heat conduction, Eg is the heat generation, and Est
is the energy stored in the control volume.
Figure 1.6 shows the conservation of energy in a differential control
volume in a 3-D Cartesian (rectangular) coordinate. If we consider energy
TH
T(x, y, z, t)
y
k, α
x
z
TL
FIGURE 1.5
Heat conduction through a solid medium.
Heat Conduction Equations 7
∂q
z qz + z
dz ∂qy
∂z qy + dy
∂y
Est ∂qx
qx y qx + dx
Eg ∂x
x
qy q
z
FIGURE 1.6
The volume element for deriving the heat conduction equation.
∂T
qx = −kAx (1.10)
∂x
Eg = q̇ dx dy dz (1.12)
∂(ρ dx dy dz · Cp · T)
Est = (1.13)
∂t
Dividing out the dimensions of the small control volume dx dy dz, Equa-
tion 1.14 is simplified as
∂ ∂T ∂(ρCp T)
k + q̇ = (1.15)
∂x ∂x ∂t
z (b)
(a)
T(r, θ, ϕ)
r
T(r, z, ϕ) θ
r
ϕ
ϕ
FIGURE 1.7
(a) Cylindrical coordinate system. (b) Spherical coordinate system.
Heat Conduction Equations 9
Ts
T(x, t)
FIGURE 1.8
Boundary conditions—given surface temperature.
10 Analytical Heat Transfer
(a) (b)
T T
T(x, t)
T(x, t)
q"s
x x
FIGURE 1.9
(a) Finite heat flux. (b) Adiabatic surface.
∂ 2T
1-D =0
∂x2
∂ 2T ∂ 2T
2-D + =0
∂x2 ∂y2
∂ 2T ∂ 2T ∂ 2T
3-D + + =0
∂x2 ∂y2 ∂z2
T
T∞, h T(0, t)
T(x, t)
FIGURE 1.10
Convective boundary conditions.
T(x, 0) = Ti
FIGURE 1.11
Initial condition.
Heat Conduction Equations 11
∂ 2T q̇
+ =0
∂x2 k
∂ 2T q̇
− =0
∂x2 k
∂ 2T 1 ∂T
1-D =
∂x2 α ∂t
∂ 2T ∂ 2T 1 ∂T
2-D + =
∂x 2 ∂y 2 α ∂t
∂ 2T ∂ 2T ∂ 2T 1 ∂T
3-D + + =
∂x 2 ∂y 2 ∂z 2 α ∂t
The above-mentioned three kinds of BCs can be applied to the 1-D, 2-D,
or –3-D heat conduction problems, respectively. For example, as shown in
Figure 1.12,
∂T(0, y, t)
x = 0, −k = 0 (adiabatic surface)
∂x
∂T(a, y, t)
x = a, −k = h[T(a, y, t) − T∞ ] (surface convection)
∂x
∂T(x, 0, t)
y = 0, −k = q (surface heat flux)
∂y
y = b, T(x, b, t) = Ts surface temperature
Remarks
In general, heat conduction problems, regardless of 1-D, 2-D, 3-D, steady
or unsteady, can be solved analytically if thermal conductivity is a given
constant and thermal BCs are known constants. The problems will be ana-
lyzed and solved in Chapters 2 through 4. However, in real-life applications,
there are many materials whose thermal conductivities vary with tempera-
ture and location, k(T) ∼ k(x, y, z). In these cases, the heat conduction equation
shown in Equation 1.16 becomes a nonlinear equation and is harder to solve
analytically.
12 Analytical Heat Transfer
Ts
y
Insulated T∞ , h
0 a x
q"s
FIGURE 1.12
Heat conduction in 2-D system with various boundary conditions.
PROBLEMS
1.1 Derive Equation 1.18.
1.2 Derive Equation 1.19.
1.3 a. Write the differential equation that expresses transient heat
conduction in 3-D (x, y, z coordinates) with constant heat
generation and constant conductivity.
b. Simplify the differential equation in (a) to show steady-state
conduction in one dimension, assuming constant conductivity.
c. If the BCs are: T = T1 . . . at . . . x = x1 , T = T2 . . . at . . . x = x2 ,
solve the second-order differential equation to yield a temper-
ature distribution (T). Express the answer (T) in terms of (T1 ,
T2 , x, x1 , x2 ).
d. Using the Fourier law and the results from (c), develop an
expression for the heat rate per unit area, assuming constant
conductivity. Express your answer in terms of (k, T1 , T2 , x1 , x2 ).
Reference
1. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
2
1-D Steady-State Heat Conduction
∂ 2T
=0 (2.1)
∂x2
dT
= C1
dx
T = c1 x + c2 (2.2)
at x = 0, T = Ts1 = c1 · 0 + c2 = c2
at x = L, T = Ts2 = c1 L + c2
Ts2 − Ts1
c1 = , c2 = Ts1 ,
L
Ts,1 − Ts,2
T(x) = Ts,1 − x (2.3)
L
Applying Fourier’s Conduction Law, one obtains the heat transfer rate
through the plane wall
T∞,1 h1
0 x L
FIGURE 2.1
Conduction through plane wall and thermal–electrical network analogy.
The heat transfer rate divided by the cross-sectional area of the plane wall,
that is, heat flux is
q ∂T Ts,1 − Ts,2
q = = −k =k (2.5)
A ∂x L
At the convective surfaces, from Newton’s Cooling Law, the heat transfer
rates are
T∞,1 − Ts,1
q = Ah1 (T∞,1 − Ts,1 ) = (2.6)
(1/Ah1 )
and
Ts,2 − T∞,2
q = Ah2 (Ts,2 − T∞,2 ) = (2.7)
(1/Ah2 )
Applying the analogy between the heat transfer and electrical network, one
may define the thermal resistance like the electrical resistance. The thermal
resistance for conduction in a plane wall is
L
Rcond = (2.8)
kA
The thermal resistance for convection is then
1
Rconv = (2.9)
Ah
The total thermal resistance may be expressed as
1 L 1 1
Rtot = + + = (2.10)
Ah1 kA Ah2 UA
where U is the overall heat transfer coefficient.
1-D Steady-State Heat Conduction 15
T(r) = c1 ln r + c2
at r = r1, T = Ts,1 = c1 ln r1 + c2
at r = r2 , T = Ts,2 = c1 ln r2 + c2
Contact surface
T∞,1 h1
Ts,1
k2, α2
Hot fluid
Ta ΔT Cold fluid
k1, α1 Tb
Ts,2
T∞,2 h2
0
L1 L2
FIGURE 2.2
Temperature drop due to thermal contact resistance between surface a and surface b.
16 Analytical Heat Transfer
r2
r1
T∞,1 h1
Inside
hot fluid Ts,1
Ts,2 T∞,2 h2
Outside
cold fluid
FIGURE 2.3
Conduction through circular tube wall.
If we consider radiation heat loss between the tube outer surface and
surrounding wall,
4 4
qradiation = A2 εσ(Ts,2 − Tsur )
2 2
= A2 εσ(Ts,2 + Tsur )(Ts,2 + Tsur )(Ts,2 − Tsur )
= A2 hr (Ts,2 − Tsur )
where
2 2
hr = εσ(Ts,2 + Tsur )(Ts,2 + Tsur )
and
q = qconv + qrad
= A2 h2 (Ts,2 − T∞,2 ) + A2 hr (Ts,2 − Tsur )
1 ln(r2 /r1 ) 1
Rtot = + + (2.17)
h1 2πr1 l 2πkl (h2 + hr )2πr2 l
k
(ro )critical = (2.18)
h
18 Analytical Heat Transfer
Ts,1
. Ts,2
T∞,1 h1 q
T∞,2 h2
x
O
−L L
FIGURE 2.4
Flat plate heat conduction with internal heat generation and asymmetrical boundary conditions.
∂ 2T q̇
2
+ =0
∂x k
(2.19)
dT q̇
= − x + c1
dx k
The general solution is
q̇ 2
T=− x + c1 x + c 2
2k
with asymmetrical boundary conditions [2] shown in Figure 2.4,
q̇ 2
at x = L, T = Ts,L = − L + c1 L + c 2
2k
q̇
at x = −L, T = Ts,1 = − (−L)2 + c1 (−L) + c2
2k
Solve for c1 and c2 , one obtains the temperature distribution
q̇L2 x2 Ts,2 − Ts,1 x Ts,2 + Ts,1
T(x) = 1− 2 + +
2k L 2 L 2
1-D Steady-State Heat Conduction 19
(a) (b)
T0 T0
. .
Ts q Ts q Ts
T∞ h T∞ h T∞ h
x x
L L
FIGURE 2.5
Flat plate heat conduction with internal heat generation. (a) Symmetric boundary conditions.
(b) Adiabatic surface at midplane.
q̇L2
T0 = Ts + (2.21)
2k
dT
q = −k = h(Ts − T∞ ) (2.22)
dx x=L
dT q̇L
=−
dx x=L k
q̇L
Ts = T∞ + (2.23)
h
20 Analytical Heat Transfer
T0
T∞ h T∞ h
Ts . Ts
q
r r0
FIGURE 2.6
Cylindrical rod heat conduction with internal heat generation.
q̇r02
T0 = Ts + (2.26)
4k
1-D Steady-State Heat Conduction 21
q̇ro
Ts = T∞ + (2.28)
2h
qf
qx ∂qx
qx + dx
x ∂x
h, T∞ Cold fluid
FIGURE 2.7
One-dimensional conduction through thin fins with uniform cross-section area.
at a given fin geometry and working conditions. We assume that heat conduc-
tion through the fin is 1-D steady state because the fin is thin. The temperature
gradient in the other two dimensions is neglected. We will begin with the con-
stant cross-sectional area fins and then consider the variable cross-sectional
area fins. The following is the energy balance of a small control volume of
the fin with heat conduction through the fin and heat dissipation into cooling
fluid, as shown in Figure 2.7. The result of temperature distributions through
fins of different materials can be seen from Figure 2.8.
dqx
qx − qx + dx − hAs (T − T∞ ) = 0 (2.29)
dx
d dT
− −kAc dx − hP dx(T − T∞ ) = 0 (2.30)
dx dx
Tb
Copper
Aluminum
Steel
Plastic
T∞
x
FIGURE 2.8
Temperature distributions through fins of different materials.
1-D Steady-State Heat Conduction 23
d2 T hP
− (T − T∞ ) = 0
dx2 kAc
(2.31)
d2 (T − T∞ ) hP
− (T − T∞ ) = 0
dx2 kAc
Let θ(x) = T(x) − T∞ , Equation 2.31 becomes
d2 θ hP
2
− θ=0 (2.32)
dx kAc
d2 θ
− m2 θ = 0 (2.33)
dx2
The general solution is
x = 0, θ(0) = θb = c1 + c2
x=L
emx − e−mx
sinh mx =
2
emx + e−mx
cosh mx =
2
d sinh mx
= cosh mx · m dx
dx
d cosh mx
= sinh mx · m dx
dx
For case 2: dθ/dx|x=L = 0, that is, c1 memL − c2 me−mL = 0, solve for c1 and
c2 , one obtains the temperature distribution through the fin base (qf ) as
For case 3: θ(L) = θL , that is, c1 emL + c2 e−mL = θL , solve for c1 and c2 , one
obtains the temperature distribution and heat transfer through the fin base
(qf ) as
θ
= e−mx (2.40)
θb
dθ(0)
It should be noted that the above results can be applied to any fins with
uniform cross-sectional area. This includes the fins with circular, rectangu-
lar, triangular, and other cross sections as shown in Figure 2.7. One should
know how to calculate the fin cross-sectional area Ac and fin perimeter P
(circumferential length) for a given uniform cross-sectional area fin geometry.
The fin effectiveness must be greater than unity in order to justify using
the fins. Normally, it should be greater than 2 in order to include the material
and manufacturing costs. In general, the fin effectiveness is greater than 5 for
most of the effective fin applications. For example, for the long fins (case 4 fin
tip boundary conditions), the fin effectiveness is
√
hPkAc θb kP
ηε = = >1∼5 (2.43)
hθb Ac hAc
qfin
ηf = (2.44)
qmax
By definition, the fin efficiency is between 0 and 1. However, the fin effi-
ciency is around 0.9–0.95 for most of the efficient fin application. For example,
for long fins (case 4 fin tip boundary conditions), the efficiency is
√
hPkAc θb
ηf = ≥ 90% (2.45)
hPLθb
26 Analytical Heat Transfer
In order to get the higher fin effectiveness and fin efficiency, we need to have
a thin fin (larger P/Ac ratio) with larger thermal conductivity k (aluminum or
copper) and low working fluid heat transfer coefficient h (air cooling).
q = qfin + qnon-fin
q = Nηf qmax + h(Tb − T∞ )Anon-fin
where qr = radiation gain from solar = constant, or qr = radiation loss =
4 ). If we consider q = constant, the solution of above equation
−εσ(T 4 − Tsur r
can be obtained by Equation 2.34 by setting
qr
θ = T − T∞ −
h
If k is constant, we have
d dT hP
Ac − (T − T∞ ) = 0
dx dx k
(2.46)
d d(T − T∞ ) hP
Ac − (T − T∞ ) = 0
dx dx k
(a) (b)
h, T∞
hAs(T−T∞) h, T∞
qx dqx
qf x qx + dx
dx
t
qf
dx r1
dr
r2
(c) h, T∞ (d)
h, T∞
o ro
qf
r
t
dx
x 0
FIGURE 2.9
One-dimensional heat conduction through thin fins with variable cross-section area. (a) Conical
fin; (b) annular fin; (c) taper fin; (d) disk fin.
Ac = 2πr · t. The fin perimeter including the top and the bottom, P = 2 · 2πr.
Let θ = T − T∞ , Equation 2.46 becomes
d dθ h · 4πr
2πrt − θ=0
dr dr k
d dθ 2 hr
r − θ=0
dr dr kt
(2.47)
d dθ 2 hr2
r r − θ=0
dr dr kt
d dθ
r r − m2 r 2 θ = 0
dr dr
where m2 = 2h/kt.
1. Convective boundary
∂T ∂θ(r2 ) −h
−k = h(Tr2 − T∞ ) or = θr
∂r r=r2 ∂r k 2
T|r=r2 = T∞ or θ(r2 ) = T∞ − T∞ = 0
At r = r1 , θ = θb = a0 I0 (mr1 ) + a1 K0 (mr1 ),
∂θ(r2 ) dI0 (mr) dK0 (mr)
At r = r2 , = a0 + a
dr r=r2 dr r=r2
1
∂r
= a0 mI1 (mr2 ) − a1 mK1 (mr2 ) = 0,
30 Analytical Heat Transfer
d d
I0 (mr) = mI1 (mr) and K0 (mr) = −mK1 (mr)
dr dr
Solve for a0 and a1
K1 (mr2 )
a0 = θb
I0 (mr1 )K1 (mr2 ) + I1 (mr2 )K0 (mr1 )
I1 (mr2 )
a1 = θb
I0 (mr1 )K1 (mr2 ) + I1 (mr2 )K0 (mr1 )
J I
J0 (x)
1 J1 (x)
I0 (x)
0
I1 (x)
1
0
x x
Y K
Y0 (x)
K1 (x)
Y1 (x)
K0 (x)
0 0
x x
FIGURE 2.10
The characteristics of Bessel functions.
Examples
R″tc,A, R″tc,B
r3
r2
r1 ho, T∞,o
q″h, (2πr2)
T∞,i Th T∞,o
q′i q′o
1 ln(r2/r1) ln(r3/r2) 1
R″tc,B , R″tc,A
hi 2πr1 2πkB 2πkA ho 2πr3
FIGURE 2.11
Thermal circuit of a composite cylindrical wall and all resistance.
32 Analytical Heat Transfer
the composite. The outer surface is exposed to ambient air, which is at T∞,o
and provides a convection coefficient of ho . Under steady-state conditions,
a uniform heat flux of qh is dissipated by the heater.
a. Sketch the equivalent thermal circuit of the system and express all
resistances in terms of relevant variables.
b. Obtain an expression that may be used to determine the heater tempera-
ture, Th .
c. Obtain an expression for the ratio of heat flows to the outer and inner
fluids, qo /qi . How might the variables of the problem be adjusted to
minimize this ratio?
SOLUTION
a. See the sketch shown in Figure 2.11.
b. Performing an energy balance for the heater, Ėin = Ėout , it follows that
Th − T∞,i
qh (2πr2 ) = qi + qo =
(hi 2πr1 )−1 + (ln(r2 /r1 )/2πkB ) + Rtc,B
Th − T∞,o
+
(ho 2πr3 )−1 + (ln(r3 /r2 )/2πkA ) + Rtc,A
−1
qo (T − T∞,o ) (hi 2πr1 ) + (ln (r2 /r1 ) /2πkB ) + Rtc,B
= h ·
qi (Th − T∞,i ) (ho 2πr3 )−1 + (ln (r3 /r2 ) /2πkA ) + Rtc,A
SOLUTION
a. q̇ = q̇o 1 − (x/L)2
x 2
d2 T q̇o
= − 1 −
dx 2 k L
dT q̇o 1 x3
=− x− + C1
dx k 3 L2
q̇o 1 2 1 x4
T =− x − + C1 x + C2
k 2 12 L2
1-D Steady-State Heat Conduction 33
dT
Boundary conditions: x = 0, = 0 (symmetry); x = L,
dx
dT
−k = h (T − T∞ )
dx
b. q̇ = a + b(T − T∞ )
d2 T b a
+ T − T ∞ − =0
dx 2 k b
1/2
a h cos b/k a/b x
T − T∞ − = 1/2 1/2 1/2
b h cos b/k L + b/k sin b/k L
2.3. A long gas turbine blade (Figure 2.12) receives heat from combustion gases
by convection and radiation. If reradiation from the blade can be neglected
Ts T∞ , determine the temperature distribution along the blade. Assume
a. The blade tip is insulated.
b. The heat transfer coefficient on the tip equals that on the blade sides.
The cross-sectional area of the blade may be taken to be constant.
SOLUTION
If the blade is at a much lower temperature than the combustion gases, radi-
ation emitted by the blade will be much smaller than the absorbed radiation
and can be ignored to simplify the problem. An energy balance on an element
of fin Δx long gives
dT dT
−kAc + kA − hP Δx(T − T∞ ) + qrad P Δx = 0
dx x dx x+Δx
c
d2 T hP q P
2
− (T − T∞ ) + rad = 0
dx kAc kAc
34 Analytical Heat Transfer
Combustion
gases
L
T∞, h
Tb
FIGURE 2.12
A turbine blade modeled as a fin with constant cross-sectional area.
d2 T hP q
2
− T − T∞ + rad = 0
dx kAc h
= T + q /h. Then
which defines an “effective” ambient temperature T∞ ∞ rad
.
the solutions can be obtained by replacing T∞ with T∞
a. Insulated blade tip.
T − T∞ + qrad /h cosh m (L − x)
=
Tb − T∞ + qrad /h cosh mL
2.4. The attached Figure shows a straight fin of triangular profile (Figure 2.13).
Assume that this is a thin fin with w t . Derive the heat conduction equation
of fin: determine the temperature distributions in the fin analytically; and
determine the fin efficiency.
SOLUTION
From Figure 2.13,
d dT h dAs
Ac − (T − T∞ ) = 0 (2.55)
dx dx k dx
tw d2 T tw dT 2hw
x + − (T − T∞ ) = 0
L dx 2 L dx k
1-D Steady-State Heat Conduction 35
T∞, h
t
W
dx
L
x
FIGURE 2.13
A triangular straight fin with variable cross-sectional area.
d2 T dT 2h
x + − L (T − T∞ ) = 0
dx 2 dx kt
d2 θ dθ
x2 +x − m2 Lxθ = 0 (2.56)
dx 2 dx
where
t
Ac = w x
L
dAs = 2w dx
θ ≡ T − T∞
2h
m2 ≡
kt
As
2wL
dz √ 1 √ 1 2m2 L
= 2m L · · x −1/2 = m L √ =
dx 2 x z
2
dθ dθ dz d θ d dθ dz
= · ; = · ;
dx dz dx dx 2 dz dx dx
36 Analytical Heat Transfer
d2 θ dθ
z2 +z − z 2θ = 0 (2.57)
dz 2 dz
θ = C1 Io (z) + C2 Ko (z)
Boundary conditions
at x = 0, Ko → ∞, C2 = 0,
at x = L, θ = θb ,
√
I0 2m xL
θ = θb ·
I0 (2 mL)
dT I (2 mL)
qf = +kAc = θb ktwm 1
dx x=L I0 (2 mL)
qf 1 I1 (2 mL)
ηf = =
hθb 2wL mL I0 (2 mL)
2.5. A hollow transistor (Figure 2.14) has a cylindrical cap of radius ro and height
L, and is attached to a base plate at temperature Tb . Show that the heat
dissipated is
I (mro ) sinh mL + I1 (mro ) cosh mL
qf = 2πkro t Tb − T∞ m 0
I0 (mro ) cosh mL + I1 (mro ) sinh mL
where the metal thickness is t , m = (h/kt )1/2 , and the heat transfer coef-
ficient on the sides and top is assumed to be the same, h, and outside
temperature, T∞ .
t ro
T∞, h
L hi = 0
Tb
FIGURE 2.14
A hollow transistor modeled as a fin.
1-D Steady-State Heat Conduction 37
SOLUTION
The cap is split into two fins, (1) a straight fin of width 2πro and length L, and
(2) a disk fin of radius ro .
For the straight fin with T = Tb at x = 0,
1/2
T1 − T∞ = C1 sinh mx + (Tb − T∞ ) cosh mx; m = h/kt (2.58)
Since kAc is the same for both fins at the junction. Substituting Equa-
tions 2.58 and 2.59 into Equation 2.60 gives
C2 I0 (mro ) = C1 sinh mL + Tb − T∞ cosh mL
−C2 I1 (mro ) = C1 cosh mL + Tb − T∞ sinh mL
Solving,
Remarks
This chapter deals with 1-D steady-state heat conduction through the plane
wall with and without heat generation, cylindrical tube with and without
heat generation, and fins with constant and variable cross-sectional area.
Although it is a 1-D steady-state conduction problem, there are many engi-
neering applications. For example, heat losses through building walls, heat
transfer through tubes, and heat losses through fins. In undergraduate heat
transfer, we normally ask you to calculate heat transfer rates through the plane
38 Analytical Heat Transfer
walls, circular tubes, and fins, with given dimensions, material properties, and
thermal boundary conditions. Therefore, you can pick up the right formulas
and plug in with these given numbers and obtain the results.
However, in this intermediate heat transfer level, we are more focused on
how to solve the heat conduction equation with various thermal boundary
conditions and how to obtain the temperature distributions for a given phys-
ical problem. For example, how to solve the temperature distributions for the
plane walls, circular tubes, with and without heat generation, and fins with
constant and variable cross-sectional area, with various thermal boundary
conditions. In particular, we have introduced one of very powerful mathe-
matical tools, Bessel function, to solve the fins with variable cross-sectional
area with various thermal boundary conditions. This is the only thing new as
compared to the undergraduate heat transfer.
PROBLEMS
2.1. The performance of gas turbine engines may be improved by
increasing the tolerance of the turbine blades to hot gases
emerging from the combustor. One approach to achieving high
operating temperatures involves application of a thermal barrier
coating (TBC) to the exterior surface of a blade, while passing
cooling air through the blade. Typically, the blade is made from
a high-temperature superalloy, such as Inconel (k ≈ 25 W/m K)
while a ceramic, such as zirconia (k ≈ 1.3 W/m K), is used as
a TBC.
Consider conditions for which hot gases at T∞,o = 1700 K and
cooling air at T∞,i = 400 K provide outer- and inner-surface con-
vection coefficients of ho = 1000 W/m2 K and hi = 500 W/m2 K,
respectively. If a 0.5-mm-thick zirconia TBC is attached to a 5-
mm-thick Inconel blade wall by means of a metallic bonding
=
agent, which provides an interfacial thermal resistance of Rt,c
10−4 m2 K/W, can the Inconel be maintained at a temperature that
is below its maximum allowable value of 1250 K? Radiation effects
may be neglected, and the turbine blade may be approximated as
a plane wall. Plot the temperature distribution with and without
the TBC. Are there any limits to the thickness of the TBC?
2.2. 1-D heat conduction through a circular tube, as shown in Fig-
ure 2.3. Determine heat loss per tube length as the following
conditions:
Given:
Steam Inside the Pipe Air Outside the Pipe Steel Pipe AISI 1010
Find: q/L =?
1-D Steady-State Heat Conduction 39
T•, h2 T•, h2
T•, h2 h1 = 0 T•, h2
Tb ri
ro
Transistor
FIGURE 2.15
A hollow cylindrical copper tube fin.
40 Analytical Heat Transfer
b. If during the air cooling, the annular fin has also received radi-
ation energy from the surrounding environment, Tsur , can you
sketch, compare, and comment on the fin temperature profile
T(r) with that in (a)?
2.7. Both sides of a very thin metal disk, as shown in Figure 2.9d, are
heated by convective hot air at temperature T∞ . The convection
heat transfer coefficient h can be taken constant over the disk. The
periphery at r = R is maintained at a uniform temperature TR .
a. Derive the steady-state heat conduction equation of the disk.
b. Propose a solution method and the associated boundary con-
ditions that can be used to determine the disk temperature
distributions. Sketch the disk temperature profile T(r). Do
you think the disk temperature is hotter at the center or the
periphery? Why?
c. If during the air heating, the disk has also emitted a net uniform
radiation flux qrad to the surrounding environment, derive the
steady-state heat condition equation of the disk and propose
a solution method to determine the disk temperature distri-
butions. Can you sketch, compare, and comment on the disk
temperature profile with that in (b)?
2.8. The front surface of a very thin metal disk is cooled by convective
air at temperature T∞ while the back surface is perfectly insulated,
as shown in Figure 2.9d. The convection heat transfer coefficient h
can be taken to be constant over the front surface of the disk. The
periphery at r = R is maintained at a uniform temperature TR by
a heat source.
a. Derive the steady-state heat conduction equation of
disk.
b. Determine the disk temperature distributions with the associ-
ated boundary conditions. Sketch the disk temperature profile
T(r).
c. If during the air cooling, the disk has also emitted a net uniform
radiation flux qrad to the surrounding environment, derive
the steady-state heat conduction equation of the disk and
determine the disk temperature distributions. Can you sketch,
compare, and comment on the disk temperature profile with
that in (b)?
2.9. A thin conical pin fin is shown in Figure 2.9a. Determine analyt-
ically the temperature profile in the pin fin. Also determine the
heat flux through the pin fin base.
Given:
Pin fin tip temperature: TR > T∞ Pin fin height: l
Pin fin base diameter: d Pin fin base temperature: Tb
Cooling air at T∞ , h Hot wall at Tb
2.10. The wall of a furnace has a height L = 1 m and is at a uni-
form temperature of 500 K. Three materials of equal thickness
t = 0.1 m, having the properties listed in the table attached, are
placed in the order shown in the Figure to insulated the furnace
1-D Steady-State Heat Conduction 41
Air
Furnace
A B C
500 K T∞ = 300 K
V∞ = 1.5 m/s
1 2 3 4
FIGURE 2.16
A furnace composite plane wall model.
wall (Figure 2.16). Air at T∞ = 300 K blows past the outer layer
of insulation at a speed of V∞ = 1.5 m/s as shown in the figure.
a. Assuming 1-D, steady conduction through the insulation lay-
ers, calculate the heat flux from the wall to the surroundings,
q. For this, choose the most appropriate of the following two
correlations:
h̄L 1/2
= 0.664 ReL Pr1/3 (ReL < 105 ; laminarflow)
k
h̄L
= (0.037ReL0.8 − 850)Pr1/3 (ReL > 105 ; turbulentflow)
k
A 100
B 10
C 1
Air 0.026 1.177 1.846 × 10−5 1.006
T0 TL
Copper rod
V
x=0 x=L
FIGURE 2.17
A moving fin model.
2.15. A thin conical pin fin is attached to a hot base plate at Tb . The
cooling air has temperature T∞ and convection heat transfer coef-
ficient h. Determine analytically the temperature profile in the pin
fin. Also, determine the heat flux through the pin fin base.
2.16. Solve temperature profile for the annulus fin geometry as shown
in Figure 2.9b, assume that the fin tip is exposed to convection
fluid with same given temperature and convection heat transfer
coefficient.
2.17. Solve the temperature profile for the annulus fin geometry as
shown in Figure 2.9b, assume that the fin tip is fixed at a given
temperature between the fin base and the convection fluid.
2.18. Determine the solutions shown in Equations 2.36 and 2.37.
2.19. Determine the solutions shown in Equations 2.38 and 2.39.
2.20. Determine the solutions shown in Equations 2.40 and 2.41.
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
44 Analytical Heat Transfer
2. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
3. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
4. V. Arpaci, Conduction Heat Transfer, Addison-Wesley Publishing Company, Read-
ing, MA, 1966.
3
2-D Steady-State Heat Conduction
∂ 2T ∂ 2T
2
+ 2 =0 (3.1)
∂x ∂y
∂ 2θ ∂ 2θ
+ = 0, if let θ = T − T0 (3.2)
∂x2 ∂y2
Boundary conditions:
x = 0, T = 0 or x = 0, θ = 0 homogeneous BC
x = a, T = 0 or x = a, θ = 0 homogeneous BC
y = 0, T = 0 or y = 0, θ = 0 homogeneous BC
y = b, T = Ts or y = b, θ = Ts − T0 = θs nonhomogeneous BC
Here, we defined a homogeneous BC as T = 0, or ∂T/∂x = 0, ∂T/∂y = 0;
θ = 0, or ∂θ/∂x = 0, ∂θ/∂y = 0, that is, temperature or temperature gradient
at a given boundary surface (in the x- or y-direction) equals 0. In contrast,
we define a nonhomogeneous BC as T = 0, ∂T/∂x = 0, ∂T/∂y = 0; or θ = 0,
∂θ/∂x = 0, ∂θ/∂y = 0, that is, temperature or temperature gradient at a given
boundary surface (in the x- or y-direction) does not equal 0.
Equations 3.1 and 3.2 can be solved by the method of separation of variable.
By means of this, we can separate the temperature from depending on two
directions, T(x, y), to one direction each, T(x) and T(y), respectively. The final
2-D temperature distribution is a product of each 1-D temperature solution,
that is, T(x, y) = T(x) · T(y). The following outlines the method of separation
45
46 Analytical Heat Transfer
y
T = Ts Isofluxes
b
T=0 T=0
Isotherm
x
T=0 a
y
Ts θ = θs
b
T0 T0
θ = T –T0
θ=0 θ=0
x
T0 θ=0 a
FIGURE 3.1
2-D heat conduction with three homogeneous and one nonhomogeneous boundary conditions.
of variable [1–4]. We need four BCs, two in the x-direction and two in the
y-direction, to solve the 2-D heat conduction problem. One important note is
that, among four BCs, only one nonhomogenoeus BC is allowed in order to
apply the principal of separation of variable method. For a given problem,
we need to make sure that only one nonhomogeneous BC exists either in the
x- or in the y-direction. The principal of superposition will be used for the
problems having two, three, or four nonhomogeneous BCs. We will begin with
the simplest case, as shown in Figure 3.1, with given surface temperatures as
BCs. Then we will move to more complicated cases with surface heat flux and
surface convection BCs as well as the problems required in the principal of
superposition.
T(x, y) = X(x)Y(y) (3.3)
∂T ∂X dX
=Y =Y
∂x ∂x dx
∂ 2T ∂ 2X d2 X
= Y = Y
∂x2 ∂x2 dx2
2-D Steady-State Heat Conduction 47
∂T ∂Y dY
=X =X
∂y ∂y dy
∂ 2T ∂ 2Y d2 Y
= X = X
∂y2 ∂y2 dy2
d2 X d2 Y
Y + X =0
dx2 dy2
(3.4)
1 d2 X 1 d2 Y
− 2
=
X dx Y dy2
The equality can hold only if both sides are equal to a constant as each side
of Equation 3.4 is a function of an independent variable. The constant can
be positive, negative, or zero. However, the positive number of the constant
is the only possibility for the BCs in this case. The readers may note that
zero and negative constants do not satisfy the BCs. Therefore, we express
Equation 3.4 as
1 d2 X 1 d2 Y
− 2
= = λ2 (3.5)
X dx Y dy2
Then we have
at x = 0, T = 0, ⇒ X = 0, then C1 = 0,
at x = a, T = 0, ⇒ X = 0, then C2 sin λa = 0 ⇒ λn a = nπ, n = 0, 1, 2, 3, . . . and
then,
nπ
λn =
a
Therefore,
nπx
X(x) = C2 sin
a
48 Analytical Heat Transfer
With sinh(x) = (ex − e−x )/2, let C5 = C3 /2, the above equation can be
written as
Y( y) = C5 sinh(λn y)
Cn = 0, if m = n;
Ts sin(nπx/a) dx
Cn = 2
, if m = n.
sin (nπx/a) sinh(nπb/a) dx
How does one perform integration? From the integration table, one
obtains
a
nπx a nπx a a
1. dx = −
sin cos = (1 − cos(nπ))
a nπ a 0 nπ
0
⎧
a ⎨0, n = even
= 1 − (−1)n = 2a
nπ ⎩ , n = odd
nπ
a
nπx a nπx 1 2nπx a a
2. sin2 dx = − sin =
a 2nπ a 2 a 0 2
0
or
a a
nπx
2 1 − cos(2nπx/a) a 1 a 2nπx a a
sin dx = dx = − sin =
a 2 2 2 2nπ a 0 2
0 0
2-D Steady-State Heat Conduction 49
(2/nπ) [1 − (−1)n ] Ts
Cn =
sinh(nπb/a)
The second-order partial differential equations (PDEs) T(x, y) are split into
two second-order ordinary differential equations (ODEs) T(x) and T(y). The
second-order ODE with two homogeneous BCs is the so-called eigenvalue
equation. The solution of the eigenfunctions, sine and cosine, depend on
the two homogeneous BCs. The eigenvalues, λ, can be determined by one
of the two homogeneous BCs (either both in the x-direction, or both in the
y-direction). The solution of the other second-order ODE is a decay curve
of combining e(x) and e(−x) for an infinite-length problem, or sinh(x) and
cosh(x) for a finite-length problem. The only nonhomogeneous BC will be
used to solve the final unknown coefficient Cn . The integrated value Cn can
be determined by performing integration of sin, sin-square or cos, cos-square,
depending on the given BCs, by using the characteristics of the orthogonal
functions.
If we let θ = T − T0 , follow the same procedure, the 2-D temperature
distribution becomes
nπx nπy
θ(x, y) = θ(x)θ(y) = Cn sin sinh (3.9)
a a
∞
(2/nπ) [1 − (−1)n ] θs nπx nπy
θ(x, y) = sin sinh (3.10)
sinh(nπb/a) a a
n=1
∂θ
y q " = −k = given
s ∂y
T1 T1
θ = T−T1
θ=0 θ=0
T1 θ=0 a x
FIGURE 3.2
2-D heat conduction with three homogeneous and one heat flux nonhomogeneous boundary
conditions.
but one. For example, in Figure 3.2, the three temperature BCs become homo-
geneous by setting θ = T − T1 and the only nonhomogeneous heat flux BC
becomes qs = −k(∂θ/∂y). The solution will be the product of sine and cosine
in the x-direction (two homogeneous BCs) and sinh cosh in the y-direction
(one nonhomogeneous BC). As before, the nonhomogeneous heat flux BC will
be used to solve the final unknown integrated value Cn .
Given a long rectangular bar with a constant heat flux along one edge,
other edges are isothermal. In order to obtain homogeneous BCs on the three
isothermal edges, let θ = T − T1 . Laplace’s equation, Equation 3.2, applies to
this steady 2-D conduction problem.
∂ 2θ ∂ 2θ
+ =0
∂x2 ∂y2
Boundary conditions:
at x = 0, y = 0, x = a, θ=0
at y = b, −k(∂θ/∂y) = qs
y ∂θ
T∞ , h h (θ− θ∞) = −k b
∂y
b
T1 θ = T−T1 T1
θ=0 θ=0
T1 θ=0 a x
FIGURE 3.3
2-D heat conduction with one convective boundary conditions.
Therefore, we obtain
∞
2qs a[1 − (−1)n ] nπx nπy
T(x, y) = T1 − sin sinh (3.11)
kn π cosh(nπb/a)
2 2 a a
n=1
x = 0, 0<y<b: θ = 0,
y = 0, 0<x<a: θ = 0,
x = a, 0<y<b: θ = 0,
52 Analytical Heat Transfer
⎧
⎪
⎪ ∂T
⎪
⎪h(T − T∞ ) = −k
⎪
⎪ ∂y
⎨ ∂(T − T1 )
y = b, 0 < x < a : h[(T − T1 ) − (T∞ − T1 )] = −k
⎪
⎪ ∂y
⎪
⎪
⎪
⎪
∂θ
⎩h(θ − θ∞ ) = −k
∂y
∞
nπx nπy
θ(x, y) = Cn sin sinh
a a
n=1
∞
∂θ nπx nπ nπb h nπx nπb
= Cn sin cosh =− Cn sin sinh − θ∞
∂y a a a k n a a
n=1
nπx nπ nπb h nπb h
Cn sin cosh + sinh = θ∞
a a a k a k
n=1
a
(h/k)θ∞ 0 sin(nπx/a) dx
Cn =
a
((nπ/a) cosh(nπb/a) + (h/k) sinh(nπb/a)) sin2 (nπx/a) dx
0
θ∞ (2/nπ) [1 − (−1)n ]
=
(sinh(nπb/a) + (nπ/a)(h/k) cosh(nπb/a))
Therefore, we obtain
∞
(2/nπ) [1 − (−1)n ] sin(nπx/a) sinh(nπy/a)
T − T1
= (3.12)
T∞ − T 1 sinh(nπb/a) + (nπ/a)(h/k) cosh(nπb/a)
n=1
∂θ
hθ = −k
h, T∞ ∂y
∇2 θ = 0 Superposition
T( y ) ∇2 T = 0 T0 ⇒ θ( y ) θ0 ⇒
θ =T–T∞ θ = θ1 + θ2 + θ3
q″ ∂θ
q″ = −k
∂y
∇2 θ1 = 0 θ1 = 0 θ2 = θ2 ( y ) ∇2 θ 2 = 0 θ2 = 0 θ3 = 0 ∇2 θ3 = 0 θ3 = θ0
θ1= 0
⊕ ⊕
FIGURE 3.4
Principle of superposition for two-dimensional heat conduction with four nonhomogenous
boundary conditions.
T(r,ϕ)
ϕ
r
z z
T(r,z,ϕ) T(r,z)
r r
FIGURE 3.5
2-D heat conduction in cylindrical coordinates.
∂ 2T ∂ 2T ∂ 2T
+ + =0 (3.15)
∂x2 ∂y2 ∂z2
y y
T1
T1
θ0 T0 T1 T1 T1
T1 z x
x T1
FIGURE 3.6
3-D heat conduction in Cartesian coordinates.
2-D Steady-State Heat Conduction 55
Let θ = T − T1 .
The above governing equation and the associated BCs become
∂ 2θ ∂ 2θ ∂ 2θ
+ + 2 =0
∂x2 ∂y2 ∂z
Let θ = X(x)Y(y)Z(z). Put its derivatives in the above 3-D heat conduction
equation and obtain
1 d2 X 1 d2 Y 1 d2 Z
− 2
= 2
+ = λ2 (3.16)
X dx Y dy Z dz2
This implies
d2 X
+ λ2 X = 0 (3.17)
dx2
1 d2 Y 1 d2 Z
− = − λ2 = μ2
Y dy2 Z dz2
⎧ 2
⎪ d Y
⎪
⎨ 2 +μ Y =0
2
dy
(3.18)
⎪
⎪ 2
⎩ d Z − (λ2 + μ2 )Z = 0
dz2
⎧ ∼
⎪ X = c1 cos λx + c2 sin λx
⎪
⎨ ∼
Y = c3 cos μy + c4 sin μy
⎪
⎪ √ √
⎩Z ∼= c5 e− λ +μ z + c6 e λ +μ z
2 2 2 2
56 Analytical Heat Transfer
Examples
x = 0, 0<y <b: T = T0
x = a, 0<y <b: T = T0
y = 0, 0<x <a: T = T0
y = b, 0<x <a: T = cx
SOLUTIONS
a.
∂ 2θ ∂ 2θ
2
+ 2 =0 (3.21)
∂x ∂y
Let θ x, y = X (x) · Y y ;
1 ∂ 2X ∂ 2Y
= − = λ2
X ∂x 2 ∂y 2
∂ 2X
− λ2 X = 0 (3.22)
∂x 2
∂ 2Y
+ λ2 Y = 0 (3.23)
∂y 2
2-D Steady-State Heat Conduction 57
Y = C3 e −λy + C4 e λy
θ = C1 cos (λx) + C2 sin (λx) C3 e−λy + C4 eλy (3.24)
Boundary conditions:
i. x = 0: θ = 0
ii. x = a: θ = 0
iii. y = 0: θ = 0
iv. y = b: θ = cx − T0
Applying BC (i) into Equation 3.24, C1 = 0
Applying BC (ii) into Equation 3.24, C2 sinh (λa) = 0, λn = nπ/a.
Applying BC (iii) into Equation 3.24, C3 = −C4 .
nπx
θ x, y = C2 · C4 sin eλn y − e−λn y
a
∞ nπx nπy
(3.25)
θ x, y = Cn sin sinh
a a
n=1
a
Cn = 0 (cx − T0 ) sin (nπx/a) dx (3.26)
a
sinh nπb/a sin2 (nπx/a) dx
0
a nπx
(cx − T0 ) sin dx
a
0
nπx
a
a 2 nπx ax T ·a nπx a
=c sin − cos + 0 cos .
nπ a nπ a 0 nπ a 0
a nπx ca2 T0 · a
(cx − T0 ) sin dx = (− cos (nπ)) + (cos (nπ) − 1)
a nπ nπ
0
a nπx ca2 T0 · a
(cx − T0 ) sin dx = (−1)n+1 + (−1)n − 1
a nπ nπ
0
a nπx a T0 · a
(cx − T0 ) sin dx = (−1)n (−ca + T0 ) −
a nπ nπ
0
58 Analytical Heat Transfer
cx
0 0
Isotherms 0 Isofluxes
FIGURE 3.7
Sketch for the isotherms and isofluxes.
a nπx
x 1 2nπx a a
sin2 dx = − sin =
a 2 4nπ a 0 2
0
Hence,
∞ nπx nπy
2 (−1)n+1 (ca − T0 ) − T0
θ x, y = sin sinh
π n sinh nπb/a a a
n=1
∂T
h(T−T∞) = −k
y ∂y y
∂θ
hθ = −k
h, T∞ ∂y
b b
θ = T–T∞
∂T ∂θ ∂θ
qs =0 qs= −k =0
∂x ∂x ∂x
0 ∂T a x 0 ∂θ a x
=0 =0
∂y ∂y
FIGURE 3.8
A long rectangular bar with heat flux as one nonhomogenous boundary condition.
2-D Steady-State Heat Conduction 59
SOLUTIONS
a.
∂ 2θ ∂ 2θ
2
+ 2 =0 (3.27)
∂x ∂y
Let θ x, y = X (x) · Y (y ):
1 ∂ 2X ∂ 2Y
= − = λ2
X ∂x 2 ∂y 2
∂ 2X
− λ2 X = 0 (3.28)
∂x 2
∂ 2Y
+ λ2 Y = 0 (3.29)
∂y 2
X = C1 sinh (λx) + C2 cosh (λx)
Y = C3 sin λy + C4 cos λy
θ = C1 sinh (λx) + C2 cosh (λx) C3 sin λy + C4 cos λy (3.30)
Boundary conditions:
i. x = 0: qs = −k (∂θ/∂x)
ii. x = a: ∂θ/∂x = 0
iii. y = 0: ∂θ/∂y = 0
iv. y = b: hθ = −k (∂θ/∂y )
Applying BC (iii) into Equation 3.30, C3 = 0.
Applying BC (ii) into Equation 3.30, C1 λ cosh (λa) + C2 λ sinh (λa) = 0,
C2 = −C1 coth (λa).
θ = Cn sinh (λx) − coth (λa) cosh (λx) cos λy (3.31)
hCn sinh (λx) − coth (λa) cosh (λx) cos λb
= +k λCn sinh (λx) − coth (λa) cosh (λx) sin λb
h
λ tan λb =
k
∞
θ= Cn sinh (λn x) − coth (λn a) cosh (λn x) cos λn y (3.32)
n=1
where λn tan λn b = (h/k ).
60 Analytical Heat Transfer
∞
qs = −k Cn λn cosh (λn x) − coth (λn a) λn sinh (λn x) cos λn y x=0
n=1
∞
qs = −k Cn λn cos λn y
n=1
b b
qs
− cos λn y dy = Cn cos2 λn y dy
k λn
0 0
b
qs 1 b y 2 sin λn y cos λn y
− sin λn y 0 = Cn +
k λn λn 2 4λn
0
2qs sin λn b
Cn = −
k λn sin λn b cos λn b + bλn
∞
2qs sin λn b
θ=
k λn sin λn b cos λn b + bλn
n=1
× − sinh (λn x) + coth (λn a) cosh (λn x) cos λn y
y
T = T3
b
T = T1 T = T2
x
0
a
q″s = 0
FIGURE 3.9
A thin rectangular plate with two one nonhomogenous boundary condition.
2-D Steady-State Heat Conduction 61
SOLUTIONS
a.
∂ 2T ∂ 2T
2
+ =0
∂x ∂y 2
Let θ = T − T1 , then
∂ 2θ ∂ 2θ
2
+ 2 =0
∂x ∂y
Let θ = θ1 + θ2 ,
∂ 2 θ1 ∂ 2 θ1
+ =0
∂x 2 ∂y 2
∂ 2 θ2 ∂ 2 θ2
+ =0
∂x 2 ∂y 2
θ1 = C1 cos(λx) + C2 sin(λx) C3 cosh(λy ) + C4 sinh(λy )
x = 0, θ1 = 0, ⇒ C1 = 0
x = a, θ1 = 0, ⇒ λa = nπ; λn = (nπ/a) n = 1, 2, 3, . . .
y = 0, ∂θ1 /∂y = 0; ⇒ C4 = 0
∞
θ1 = Cn cosh(λn y ) sin(λn x)
n=1
∞
nπx
nπb
θ1 (x, b) = (T3 − T1 ) = Cn1 cosh sin
a a
n=1
a
(T3 − T1 ) 0 sin ((nπx/a)) dx
Cn1 =
cosh (nπb/a) 0a sin2 ((nπx/a)) dx
(T3 − T1 )a · (1 − cos(nπ)/nπ) 2(T3 − T1 ) 1 − (−1)n
= =
(a/2) cosh (nπb/a) nπ cosh (nπb/a)
∞ nπy nπx
2(T3 − T1 ) 1 − (−1)n
θ1 = · cosh sin
πn cosh (nπb/a) a a
n=1
Solution for θ2
θ2 = C1 cosh(λx) + C2 sinh(λx) C3 cos(λy ) + C4 sin(λy )
62 Analytical Heat Transfer
y = 0, ∂θ/∂y = 0; ⇒ C4 = 0
y = b, θ2 = 0; ⇒ λn = nπ/2b
x = 0, θ2 = 0; ⇒ C1 = 0
∞
θ2 = Cn2 sinh(λn x) cos(λn y )
nodd
∞
θ2 (a, y ) = (T2 − T1 ) = Cn2 sinh(λn a) cos(λn y ),
nodd
(T2 − T1 ) 0b cos (nπy /2b) dy
Cn2 =
sinh (nπa/2b) 0b cos2 (λn y ) dy
∞ nπx nπy
4(T2 − T1 )(−1)(n−1)/2
θ2 = sinh cos
n nπ sinh (nπa/2b) 2b 2b
odd
Finally,
⎧
∞
⎨ 2(T − T ) nπy nπx
3 1 1 − (−1)n
T (x, y ) = T1 + cosh sin
⎩ nπ cosh nπb/a a a
n=1
⎫
∞
4(T2 − T1 ) (−1)(n−1)/2 nπx nπy ⎬
+ sinh cos
nπ n sinh (nπa/2b) 2b 2b ⎭
odd
x = 0, 0 < y < b : T = To
y = 0, 0 < x < a : T = T0
y
T = T0 + T2 sin(π x/a)
b
T = T0
T = T0 + T1 sin(π y/b)
x
0 T = T0 a
FIGURE 3.10
A long rectangular rod with two nonhomogenous boundary conditions.
SOLUTIONS
a. Let θ = T − T0 , then
∂ 2θ ∂ 2θ
+ =0
∂x 2 ∂y 2
x = 0; θ=0
x = a; θ = T1 sin (πy /b)
y = 0; θ = 0,
πx
y = b; θ = T2 sin
a
x = 0, θ1 = 0, θ2 = 0
πy
x = a, θ1 = 0, θ2 = T1 sin
b
y = 0, θ1 = 0, θ2 = 0
πx
y = b, θ1 = T2 sin θ2 = 0
a
and T = T0 + θ
T = T0 + θ1 + θ2
64 Analytical Heat Transfer
Remarks
In this chapter, we have introduced a very powerful mathematical tool, the
method of separation of variables, to solve typical 2-D heat conduction prob-
lems with various thermal BCs. In the undergraduate-level heat transfer, we
normally employ the finite-difference energy balance method to solve the
2-D heat conduction problems with various thermal BCs. The finite-difference
numerical methods and solutions will be discussed in Chapter 5. Here we
are more focused on the analytical methods and solutions for various 2-D
heat conduction problems. In general, all kinds of 2-D heat conduction prob-
lems with various BCs can be solved analytically by using superposition of
separation of variables.
The most important thing for applying separation of variable is that you
have to set up your problem where only one nonhomogeneous BC is allowed.
If you have more than one nonhomogeneous BC, you have to employ the
superposition principle to split into two or three subproblems in order to
use separation of variables. The problems become more complicated if you
work with 3-D heat conductions with heat generation and with complex BCs,
but they are still workable. However, if you are interested in solving for the
2-D and 3-D cylindrical coordinate systems and for the 2-D and 3-D spherical
coordinate systems with complex BCs, they are beyond the intermediate-level
heat transfer, you need to look at the advanced heat conduction textbook for
solutions.
Another popular method is using the finite-difference method to be dis-
cussed in the later chapter. It is particularly true when you deal with
complicate BCs such as convection. In real-life engineering applications, the
convection heat transfer coefficients normally are varied along the solid sur-
face. This will cause additional complexity for the separation of variables
because we normally assume the uniform convection BCs to simplify the prob-
lem. This will not cause any complexity at all by using the finite-difference
numerical method.
PROBLEMS
3.1. A long rectangular bar 0 ≤ x ≤ a, 0 ≤ y ≤ b, and a, b << L, the
bar length, is heated at y = o and y = b, respectively, to a uniform
temperature To and is insulated at x = 0. The side of x = a loses
heat by convection to a fluid at temperature T∞ with a convection
coefficient h.
a. Write down, step by step, a solution method and associated
BCs, which can be used to determine the bar steady-state
temperature distributions.
b. Sketch the heat flows and the isothermal profiles in the
rectangular bar.
3.2. An infinitely long rod of square cross section (LXL) floats in a
fluid. The heat transfer coefficient between the rod and the fluid
2-D Steady-State Heat Conduction 65
is relatively large compared to that between the rod and the ambi-
ent air, that is, hf >> h or hf ∼= ∞. Determine the steady-state
temperature distributions in the rod with the associated BCs.
a. Use the analytical approach.
b. Sketch the isotherms and isoflux in the rod, if Tf < T∞ and
h = a constant value.
3.3. A long fin of rectangular cross section (2LXL) with a thermal con-
ductivity k is subjected to the BCs is shown on the sketch (the
left side is kept at To , the right side is perfectly insulated, the
upper side is exposed to a constant flux, and the lower side is
exposed to a convection air flow). q = constant; steam To , h = ∞;
air h = constant, T∞ .
a. Determine the temperature distribution in the fin.
b. Approximately plot the temperature and heat flow profiles in
the fin, if To > T∞ .
3.4. Refer to Figure 3.1, and determine the temperature distributions
for 2-D heat conduction with the following BCs:
(1) x = 0, T = To (2) x = 0, T = To
x = a, T = To x = a, T = To
y = 0, T = To y = 0, T = Ts
y = b, T = Ts y = b, T = To
(3) x = 0, T = To (4) x = 0, T = Ts
x = a, T = Ts x = a, T = To
y = 0, T = To y = 0, T = To
y = b, T = To y = b, T = To
(1) x = 0, T = T1 (2) x = 0, T = T1
x = a, T = T1 x = a, T = T1
y = 0, T = T1 y = 0, qs = −k(∂T/∂y)
y = b, qs = −k(∂T/∂y) y = b, T = T1
(3) x = 0, T = T1 (4) x = 0, qs = −k(∂T/∂x)
x = a, qs = −k(∂T/∂x) x = a, T = T1
y = 0, T = T1 y = 0, T = T1
y = b, T = T1 y = b, T = T1
(1) x = 0, T = T1 (2) x = 0, T = T1
x = a, T = T1 x = a, T = T1
y = 0, T = T1 y = 0, −k(∂T/∂y)
= h(T − T∞ )
y = b, −k(∂T/∂y) = h(T − T∞ ) y = b, T = T1
66 Analytical Heat Transfer
(1) x = 0, T = T1 (2) x = 0, T = T1
x = a, T = T1 x = a, T = T1
y = 0, T = T1 y = 0, T = T1
y = b, T = T1 y = b, T = T1
z = 0, T = To z = 0, T = T1
z = c, T = T1 z = c, T = To
W
1
[cos2 (aw)] dw = a[2aW + sin (2aW)]
4
0
W
and [cos(aw) · cos(bw)] dw = 0, when a = b.
0
3.15. Given a very long and wide fin with a height of 2H. The base of
the fin is maintained at a uniform temperature of Tb . The top and
bottom surfaces of the fin are exposed to a fluid whose temper-
ature is T∞ (T∞ < Tb ). The convective heat transfer coefficient
between the fin surfaces and the fluid is h.
a. Derive an expression for the steady 2-D local temperature in
the fin, in terms of the thermal conductivity of the fin, k, the
convective heat transfer coefficient, h, the half-height of the fin,
H, and the base and fluid temperature, Tb and T∞ .
b. Sketch the steady 2-D temperature and heat flux distribution
in the fin.
Note that
x 1
cos2 (ax) dx = a [2ax + sin(2ax)] and
4
0
x
[cos(ax) · cos(bx)] dx = 0 when a = b.
0
References
1. V. Arpaci, Conduction Heat Transfer, Addison-Wesley Publishing Company, Read-
ing, MA, 1966.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
4. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
4
Transient Heat Conduction
The temperature in a solid material changes with location as well as with time,
and this is the so-called transient heat conduction problem. We may have
1-D, 2-D, or 3-D transient heat conduction problem depending on the real
applications. However, some problems can be modeled as zero-dimensional
(0-D) because the temperature in a solid material uniformly changes only with
time and does not depend on location. This is a special case of the transient
heat conduction problem. The finite-length solid material of 1-D, 2-D, or 3-D
transient problem can be solved by separation of variable method and the 0-D
transient problem can be solved by the lumped capacitance method. Figure 4.1
shows typical finite-length solid materials of the 1-D transient problem for the
slab (or the plane wall), cylinder, and sphere coordinates. The semiinfinite
solid material of the 1-D transient problem can be solved by the similarity
method, the Laplace transform method, or the approximate integral method.
The unsteady 3-D heat conduction equation with heat generation is
∂ 2T ∂ 2T ∂ 2T q̇ 1 ∂T
+ + + = (4.1)
∂x2 ∂y2 ∂z2 k α ∂t
The simplified case of the unsteady 1-D heat conduction equation without
heat generation becomes
∂ 2T 1 ∂T
= (4.2)
∂x 2 α ∂t
We need both the initial and two BCs in order to solve the temperature
depending on (x, t) as sketched in Figure 4.2. The solution of separation of
variable method is
T = T(x, t)
Initial condition:
t = 0, T(x, 0) = Ti .
69
70 Analytical Heat Transfer
–L L r
x
θ x θ r θ = fn B , F , r
= fn Bi , Fo , ; = fn Bi , Fo , ;
θ0 L θ0 ro θ0
i o
ro
FIGURE 4.1
1-D transient heat conduction and the characteristic length.
Boundary conditions:
dT
x = 0, =0
dx
∂T(L, t)
x = L, T(L, t) = Ts or −k = h[T(L, t) − T∞ ]
∂x
Ti T (x,t) Ti
T (x,t)
t q″ q″
t
Ts
h, T∞
x x
–L 0 L –L 0 L
FIGURE 4.2
1-D transient problems for a slab or a plane wall.
Transient Heat Conduction 71
(b) hAs
(a) θ – t
Ti 1 = e ρVc
θi
T (t)
θ
t θ
= fn(Bi, Fo) θi
θ0
0.368
x hAs
–L L t
0-D Transient problem–lumped 1 ρVc
capacitance method Thermal time constant
(c)
1 ρVc
= τt
θ hAs
θi
0.368
τ1 τ 2 τ3 t
The solution of 0-D heat conduction
FIGURE 4.3
Method of lumped capacitance.
capacitance method. Note that the lumped method can be applied to any irreg-
ular geometry as long as the assumption of the entire material temperature
uniformly changing with time is valid during the transient. Therefore, we do
not need to solve for 1-D, 2-D, or 3-D transient conduction equations.
Consider the energy balance on the solid material during the cooling (or
heating) process as shown in Figure 4.3:
d(ρVCT)
= −hAs (T − T∞ ) (4.3)
dt
Let θ = T − T∞ then
dθ
ρVC = −hAs θ
dt
dθ hAs
=− θ
dt ρVC
hAs
dθ = − θ dt
ρVC
θ T − T∞
= e−(hAs /ρVC)t = e−(hLc /k)·(αt/Lc ) = e−Bi·Fo = f (Bi, Fo)
2
= (4.4)
θi Ti − T ∞
The above temperature decay solution is plotted in Figure 4.3. Therefore, the
solid temperature can be predicted with time for a given material with certain
geometry under the cooling or heating condition. The material with a smaller
thermal time constant (τt = (ρVc/hAs )) can quickly reach the environment
temperature.
Now, the question is under what condition the lumped capacitance solu-
tion can be used. The answer is that the Biot (Bi) number must be less than
0.1. Therefore, to use 0-D solution, the condition Bi = (hLc /k) < 0.1 must be
satisfied. The Biot number is defined as the ratio of surface convection (h, the
convection heat transfer coefficient from the solid surface; Lc , the characteristic
length of the solid material) to solid conduction (k, the thermal conductivity
of the solid material). The smaller Bi number implies a small-sized mate-
rial with high-conductivity exposure to a low convection cooling or heating
fluid. For the case of smaller Bi, the temperature inside the solid material
changes uniformly (independent of location) with environmental cooling or
heating during the transient. Of course, Bi = 0.1 implies that we may have
10% error by using the lumped solution. The smaller Bi is better for using
the lumped solution. Another point is that the solution can be applied to any
geometry if the condition of Bi < 0.1 is valid. For a given solid geometry,
the characteristic length Lc = (volume/surface area) = (V/As ). For example,
as shown in Figure 4.1, the characteristic length Lc = L is for a 2L-thick slab
(plane wall),
1 1
Lc = Ro for the cylinder and Lc = Ro for the sphere coordinate.
2 3
d(ρVCT)
= −hAs (T − T∞ ) + q̇V + As qr
dt
where qr = radiation gain from solar flux = constant, or qr = radiation loss =
εσ(T 4 − Tsur
4 ); q̇ = heat generation = I 2 R due to electric current and resistance
heating = constant.
If we consider qr = constant and q̇ = constant, the solution of the above
equation can be obtained by Equation 4.4 by setting
q̇V + As qr
θ = (T − T∞ ) −
hAs
Transient Heat Conduction 73
d(ρVCT)
= −hAs (T − T∞ ) − εσAs (T 4 − Tsur
4
)
dt
∂ 2θ 1 ∂θ
= (4.5)
∂x 2 α ∂t
d2 X
+ λ2 X = 0
dx2
dτ
+ λ2 ατ = 0
dt
X = c1 sin λx + c2 cos λx
τ = c3 e−λ
2 αt
74 Analytical Heat Transfer
(a) (b)
Ti Ti
t t
h, T∞
Ts
–L 0 L –L 0 L
Constant surface
Convective boundary condition temperature boundary condition
FIGURE 4.4
1-D transient heat conduction.
Initial condition:
θ(x, 0) = θi = Ti − T∞
Boundary conditions:
⎧
⎪ ∂θ(0, t) !
⎨ =0
∂x c =0
⇒ 1
⎪
⎩−k ∂θ(L, t) −kc 2 (− sin λL)λ = hc2 cos λL
= hθ(L, t)
∂x
k sin(λL) · λ = h cos(λL)
h
λn = cot(λn L),
k
hL
λn L = cot(λn L) = Bi cot(λn L).
k
λn is determined by the convection BC.
where cn = c2 c3 .
Applying the initial condition
θi = cn cos(λn x) e0 ,
where λ n L = λ n
Transient Heat Conduction 75
cn cos(λn x) e−λn αt ,
2
⇒ θ = T − T∞ =
∂ 2θ ∂ 2θ ∂ 2θ 1 ∂θ
+ + 2 = (4.10)
∂x 2 ∂y 2 ∂z α ∂t
θ(x, y, z, t) = θx (x, t) · θy (y, t) · θz (z, t) (4.11)
⎧
⎪
⎪ ∂ 2 θx 1 ∂θx
⎪
⎪ =
⎪
⎪ ∂x 2 α ∂t
⎪
⎪
⎨ ∂ 2θ
y 1 ∂θy
=
⎪ ∂y
⎪ 2 α ∂t
⎪
⎪
⎪
⎪
⎪
⎪ ∂ θz
2 1 ∂θz
⎩ =
∂z 2 α ∂t
76 Analytical Heat Transfer
Ts
Ts Ti Ts
Ts
x
z
FIGURE 4.5
Multidimensional transient heat conduction.
where θx (x, t), θy (x, t), and θz (x, t) can be solved by the aforementioned
method of separation of variables with a given surface temperature or
convection BCs.
∂ 2 θ q̇ 1 ∂θ
+ = (4.12)
∂x 2 k α ∂t
The temperature profile can be solved by separation of variables. For
example, to use the separation of variables method, we define
d2 θ1 q̇
+ =0
dx2 k
where θ1 can be solved by 1-D heat conduction with internal heat generation
and θ2 can be solved by the aforementioned method of separation of variables
with a given convection or surface temperature BCs.
For 3-D transient heat conduction with heat generation, as shown in
Figure 4.6, the following equations can be used to solve temperature profiles
T(x, y, z, t) or θ(x, y, z, t):
∂ 2θ ∂ 2 θ ∂ 2 θ q̇ 1 ∂θ
+ + 2 + = (4.14)
∂x2 ∂y2 ∂z k α ∂t
.
q
–L x L
z
FIGURE 4.6
Multidimensional transient heat conduction with heat generation.
where
∂ 2 θ1 q̇
2
+ =0
∂x k
∂ 2 θ2x 1 ∂θ2x
=
∂x2 α ∂t
∂ 2 θ2y 1 ∂θ2y
=
∂y2 α ∂t
∂ 2 θ2z 1 ∂θ2z
=
∂z2 α ∂t
A general form of the solution of 1-D transient heat conduction with the
convection BC applicable to a slab, a cylinder, and a sphere, as shown in
Figure 4.1, can be written [2] as
∞
θ = Cn f (λn η) e−λn Fo
2
(4.16)
θi
n=1
∞
Q
Cn F(λn ) e−λn Fo
2
=1− (4.17)
Qo
n=1
Cylinder (1/r)(∂/∂r)(r(∂θ/∂r)) (2J1 (λn )/(λn [Jo2 (λn ) Jo (λn (r/ro )) (2J1 (λn )/λn )
= (1/r)(∂θ/∂t) +J12 (λn )]))
Sphere (1/r2 )(∂/∂r)(r2 (∂θ/∂r)) (2[sin λn − λn cos λn ]) (sin(λn (r/ro )) 3(sin λn
= (1/r)(∂θ/∂t) (λn − sin λn cos λn ) (λn (r/ro )) −λn cos λn )/λ3n
78 Analytical Heat Transfer
x hL αt
Bi cos λn − λn sin λn = 0 for the slab, η =, Bi = , Fo = 2
L k L
r hro αt
λn J1 (λn ) − BiJo (λn ) = 0 for the cylinder, η = , Bi = , Fo = 2
ro k ro
r hro αt
λn cos λn + (Bi − 1) sin λn = 0 for the sphere, η = , Bi = , Fo = 2
ro k ro
Ts Heating
T (x,t)
q″ t
Ti
x
q″
q″
t Cooling
Ts
FIGURE 4.7
1-D transient heat conduction in a semiinfinite solid material.
Transient Heat Conduction 79
given by
∂ 2T 1 ∂T
=
∂x2 α ∂t
The surface BC is
T(0, t) = Ts
T(∞, t) = Ti
∂ 2θ 1 ∂θ
=
∂x 2 α ∂t
1 d2 θ 1 −x dθ
=
4αt dη2 α 2t(4αt)1/2 dη
80 Analytical Heat Transfer
α d2 θ −x dθ
2
= 1/2
4αt dη 2t(4αt) dη
d2 θ −2x dθ dθ
2
=√ = −2η (4.19)
dη 4αt dη dη
x = 0, η = 0, θ(0) = 1
x = ∞, η = ∞, θ(∞) = 0
dP dθ
= c1 e−η
2
= −2ηP ⇒ P ≡
dη dη
⎧
⎪ dP
⎪
⎪ = −2η dη
⎨ P
dP
⎪
⎪ = −2η dη
⎪
⎩ P
ln P = −η2 + c
η
θ = c1 e−η dη + c2
2
∞ √
−u2 π 2
0 = c1 e du + 1 = c1 + 1 ⇒ c1 = − √
2 π
0
η
T − Ti 2 −u2 x x
θ= =1− √ e du = 1 − erf (η) = erfc √ = 1 − erf √
Ts − Ti π 4αt 4αt
0
(4.20)
Transient Heat Conduction 81
T−Ti
θ=
Ts−Ti
1
θ = erfc (η)
x
η=
4αt
FIGURE 4.8
Solution of 1-D transient heat.
and
∂T ∂θ
qs = −k = −k(Ts − Ti )
∂x x=0 ∂x x=0
η
∂ 2 −u2
= −k(Ts − Ti ) 1− √ e du
∂x π 0 x=0
(4.21)
2 1
= −k(Ts − Ti ) − √ e−η √
2
π 4αt η=0
k(Ts − Ti )
= √
παt
∂ 2T 1 ∂T
=
∂x 2 α ∂t
T(0, t) = Ts
T(∞, t) = Ti
Initial condition:
T(x, 0) = Ti .
82 Analytical Heat Transfer
Let θ = T − Ti , therefore,
∂ 2θ 1 ∂θ
=
∂x2 α ∂t
θ(0, t) = θs = Ts − Ti
θ(x, 0) = 0
θ(∞, t) = 0
∞
∂ 2 θ̄ 1
= [sθ̄ − θ̄(x, 0)] (4.24)
∂x 2 α
d2 θ̄ s √ √
− s/αx s/αx
− θ̄ = 0 ⇒ θ̄ = c1 e + c 2 e
dx2 α
Ts
T (x,t)
Ti
0 x
FIGURE 4.9
Solution of 1-D transient heat conduction with given surface temperature boundary condition.
So
θs −√s/αx
θ̄ = e
s
Applying the inverse Laplace transform from a given table, we obtain the
following results as sketched in Figure 4.9.
θ T − Ti x
= = erfc √ (4.25)
θs Ts − T i 4αt
T(x, 0) = Ti
T(∞, t) = Ti
∂T(0, t)
−k = qs
∂x
Let θ = T − Ti , then
θ(x, 0) = 0
θ(∞, 0) = 0
∂θ(0, t)
−k = qs
∂x
qs 1 √
θ̄ = √ − e− s/αx
k s s/α
84 Analytical Heat Transfer
qs″
T (x,t)
0 T (x,t)
FIGURE 4.10
Solution of 1-D transient heat conduction with constant surface heat flux boundary condition.
Applying the inverse Laplace transform from a given table, we obtain the
following results as sketched in Figure 4.10.
√
qs 4αt −(x2 /4αt) x
T − Ti = √ e − x · erfc √ (4.26)
k π 4αt
√
q 4αt
at x = 0, Ts − Ti = s √ (4.27)
k π
Case 3: Convective surface BC:
T(x, 0) = Ti
T(∞, t) = Ti
∂T(0, t)
−k = h[T∞ − T(0, t)]
∂x
Let θ = T − Ti
θ(x, 0) = 0
θ(∞, 0) = 0
∂θ(0, t)
−k = h[θ∞ − θ(0, t)]
∂x
Applying the inverse Laplace transform from a given table, we obtain the
following results as sketched in Figure 4.11.
θ T − Ti x 2 2 x h√
= = erfc √ − e(x(h/k)+α(h /k )t) erfc √ + αt
θ∞ T∞ − T i 4αt 4αt k
(4.28)
∂T(0, t)
q = −k = h[T∞ − T(0, t)] (4.29)
∂x
Transient Heat Conduction 85
h, T∞
T (x,t)
0 T (x,t)
FIGURE 4.11
Solution of 1-D transient heat conduction convective surface boundary condition.
θi = Ti − Ts , t=0
∂θ(δ, t)
θ(0, t) = 0, = 0, t>0
∂x
∂θ ∂ 2θ
=α 2 (4.30)
∂t ∂x
δ δ 2
∂θ ∂ θ
dx = α 2 dx
∂t ∂x
0 0
∂θ ∂θ
=α −α
∂x x=δ ∂x x=0
where
δ δ δ
∂θ d dδ dδ d dδ
dx ≡ θ dx − θx=δ + θx=0 = θ dx − · θi + 0
∂t dt dt dt dt dt
0 0 0
δ
d dδ ∂θ ∂θ
∴ θ dx− · θi = α −α
dt dt ∂x x=δ ∂x x=0
0
⎡δ ⎤
d ⎣ ∂θ
⇒ θ dx − θi δ⎦ = −α
dt ∂x x=0
0
θ(x, t) = a + bx + cx2
86 Analytical Heat Transfer
Ts
T (x,t)
t
Ti
x δ (t)
Penetration depth
FIGURE 4.12
Solution of 1-D transient conduction for semiinfinite solid material using approximate integral
method.
subject to BC:
θ x x 2
=2 − ≡ 2η − η2
θi δ δ
x
η = , dx = δ dη
δ
Then
⎡1 ⎤
d ⎣ 2αθi
θi (2η − η2 )δ dη − θi δ⎦ = −
dt δ
0
dδ2
⇒ − 12α = 0 at t = 0, δ(0) = 0
dt
√
⇒ δ = 12αt
θ x x 2 2x x2
=2 − =√ − (4.31)
θi δ δ 12αt 12αt
T∞
Ice
Tm Water
qs″ x
Moving boundary
1 2
Ts
0 x x = δ(t)
FIGURE 4.13
Heat conduction with moving boundary problem—freezing.
θ = T − Tm (4.32)
∂ 2 θ1 1 ∂θ1
=
∂x2 α1 ∂t
∂ 2 θ2 1 ∂θ2
=
∂x2 α1 ∂t
Boundary conditions (4.33):
x = 0, θ1 = θs (4.33a)
x = ∞, θ2 = θ∞ (4.33b)
x = δ(t), θ1 = 0 (4.33c)
θ2 = 0 (4.33d)
∂θ1 ∂θ2 dδ
k1 − k2 = Lρ1 (4.33e)
∂x ∂x dt
where L represents the latent heat of melting.
Initial conditions:
t = 0, δ = 0, T = T∞
Assume ρ1 = ρ2 .
Neumann applied the similarity method:
x
θ1 = c1 + c2 erf √ (4.33f)
4α1 t
x
θ2 = c1 + c2 erf √ (4.33g)
4α2 t
88 Analytical Heat Transfer
√ √
At x = δ, δ ∼ t to obtain θ1 = 0, δ = b t.
Insert Equations 4.33a through 4.33d into Equations 4.33f and 4.33g
c1 = θs
− θs
c2 = √
erf (b/ 4α1 )
θ∞
c2 = √
erfc(b/ 4α2 )
θ∞
c1 = θ∞ − √
erfc(b/ 4α2 )
η
2 ∂η d
e−η dη
2
=√
π ∂x dη
0
2 1 −x2
=√ √ exp
π 4α1 t 4α1 t
θ1 can be obtained from Equation 4.33f, θ2 can be obtained from Equation 4.33g
∂θ1
q1 = k1 can be determined
∂x 0
Tm
Ice T∞
Water
qs″ x
Moving boundary
Linear
Ts
approximation
0 x x = δ (t)
FIGURE 4.14
Slow freezing: T∞ ∼
= Tm .
Let θ = T − Tm ,
∂ 2θ 1 ∂θ
= (4.35)
∂x 2 α ∂t
⎧
⎪ θ(0, t) = Ts − Tm = θs
⎪
⎨
θ(δ, t) = 0
(4.36)
⎪
⎪ ∂θ dδ
⎩−k = Lρ
∂x δ dt
Ts Liquid Solid
Moving boundary
Tm
Ti
x x = δ (t)
FIGURE 4.15
Heat conduction with moving boundary problems.
90 Analytical Heat Transfer
Liquid Solid
Ts Moving boundary
qs″
Tm = T (δ ,t)
0 x Ti
x = δ (t)
FIGURE 4.16
Slow melting: Tm ≡ Ti .
δ
d dδ ∂θ
θ dx = −α ρL +
dt dt ∂x 0
0
Boundary conditions:
x = 0, θs = −c1 δ + c2 δ2 (4.37a)
dθ ∂θ dδ ∂θ
x = δ, =0= + , that is, θ(δ, t) = 0
dt δ ∂x δ dt ∂t δ
2
∂θ k ∂θ
=− +
∂x δ ρL ∂t δ
2
∂ 2θ ∂θ k
∴α =
∂x 2 ∂x δ ρL
δ
k
α · 2c2 = c12 . (4.37b)
ρL
From Equations 4.37a and 4.37b, we obtain
αρL
c1 = [1 − 1 + μ]
δk
αδ + θs
c2 =
δ2
where μ = (2θs Cp /L) is the Stefan number.
Transient Heat Conduction 91
Va
Tm
qs″
T∞
T i = T∞
x=0 x = L, ∞
FIGURE 4.17
Ablation at surface of flat wall.
δ ∼ b t ∼ 2αμt
4.4.2.1 Ablation
Ablating heat shields have been successful in satellite and missile reentry to
the earth’s atmosphere as a means of protecting the surface from aerodynamic
heating. In this application, the high heat flux generated at the surface first
causes an initial transient temperature rise until the surface reaches the melt-
ing temperature, Tm . Ablation (melting of the surface) begins and follows
a second short transient period, and then a steady-state ablation velocity is
reached. The melted material is assumed to run off immediately. The prob-
lem can be simplified to 1-D transient heat conduction with moving boundary
due to ablation. Figure 4.17 shows ablation at the surface of the flat plate with
an imaging ablation velocity, Va , moving to the left [5]. The heat conduction
equation for this reference frame is Equation 4.38 with an added enthalpy
flux term associated with the moving velocity Va :
∂ ∂T ∂T ∂T
k + ρcVa = ρc (4.38)
∂x ∂x ∂x ∂t
The problem can be solved by three stages: (1) the initial transient before
the surfaces reach Tm , (2) the second transient period during ablation, and
92 Analytical Heat Transfer
(3) after the steady-state ablation velocity has been reached, the temperature
distribution in the material is steady. The initial transient problem has been
solved before (i.e., Va = 0, given the surface heat flux BC). The second tran-
sient problem, Equation 4.38, can be solved by the finite-difference method.
For the steady-state problem with constant ablation velocity, Equation 4.38
becomes
∂ ∂T ∂T
k = −ρcVa (4.39)
∂x ∂x ∂x
x = 0, T = Tm
x = ∞, T = T∞ = Ti
∂T
x = ∞, =0 (4.40)
∂x
For constant properties k, ρ, c, and for the case of constant surface heat flux
qs , Equation 4.39 is solved by integrating twice and evaluating the integration
constants with Equation 4.40. Let θ = (dT/dx), then
d Va
θ=− θ
dx α
dθ Va
=− dx
θ α
dθ Va
= − dx
θ α
Va
ln θ = − x+C
α
dT
θ= = C1 e−(Va /α)x + C2
dx
where at x = ∞, (dT/dx) = 0 = θ, ∴ C2 = 0.
Then (dT/dx) = C1 e−(Va /α)x ,
dT = C1 e−(Va /α)x dx
α −(Va /α)x
T = −C1 e + C3
Va
Transient Heat Conduction 93
at x = ∞, T = T∞ = C3
α
at x = 0, T = Tm = −C1 + C3
Va
Va Va
−C1 = (Tm − C3 ) = (Tm − T∞ )
α α
Therefore,
The total heat conducted into the solid material evaluated with the
temperature distribution, Equation 4.41, is
∞
k(Tm − T∞ )
qc = ρC (T − T∞ ) dx = (4.44)
Va
0
The total heat transferred to the surface in time t is qs · t. Then for this period
of time t, the fraction of the total heat transferred which was conducted into
the solid material is obtained by substituting Equation 4.43 into Equation 4.44:
Examples
SOLUTION
1 ∂ ∂θ 1 ∂θ ∂ 2 θ 1 ∂θ 1 ∂θ
r = or + =
r ∂r ∂r α ∂t ∂r 2 r ∂r α ∂t
Boundary conditions:
i. r = 0, ∂θ∂r r =0 = 0
ii. r = ro , −k ∂θ ∂r = hθro
r =ro
Separation of variables:
θ = R(r )τ(t )
∂τ 2
+ λ2 τ = 0; τ = C3 e−αλ t
∂t
∂ 2R 1 ∂R
+ + λ2 R = 0
∂r 2 r ∂r
R(r ) = C1 J0 (λr ) + C2 Y0 (λr )
Applying BCs
∂R
= −C1 λJ1 (0) + C2 λY1 (0) = 0, J1 (0) = 0, ⇒ C2 = 0
∂r r =0
∂R h
−k = kC1 λJ1 (λro ) = hθR , λJ1 (λro ) = J0 (λro ), λn = λro
∂r r =ro k
where Bi = (hro /k ),
∞
2 2 r
θ= C n e−(α/ro )λn t J0 λn
ro
n=1
at t = 0, θ=1
∞
r
1= Cn J0 λn
ro
n=1
SOLUTION
1 ∂2 1 ∂θ
(r θ) =
r ∂r 2 α ∂t
Boundary conditions:
i. r = ro − k ∂θ
∂r = hθro
ii. t = 0 θ = 1
Let U(r , t ) = r θ(r , t ),
∂ 2U 1 ∂U
=
∂r 2 α ∂t
⎧
⎨U = 0 at r = 0
BCs ∂U h 1
⎩ + − U=0 at r = ro
∂r k ro
U=r for t = 0
U = R(r )τ(t )
∂τ 2
+ αλ2 τ = 0; ⇒ τ(t ) = C3 e−αλ t
∂t
∂ 2R
+ λ2 R(r ) = 0
∂r 2
R(r ) = C1 sin(λr ) + C2 cos(λr )
U = 0, ⇒ C2 = 0
at r = 0,
h 1
C1 λ cos(λro ) + − C1 sin(λro ) = 0
k ro
λn = λro
λn cos(λn ) + (B i − 1)sin(λn ) = 0
where Bi = (hro /k ),
∞
r 2
U= Cn sin λn e−λn F0
ro
n=1
At t = 0, θ = 1, U = r ,
4.3. Solve transient temperature profiles in a semiinfinite solid body using the
Laplace transform for the following BCs of constant surface heat flux.
If at time t = 0 the surface is suddenly exposed to a constant heat flux
qs —for example, by radiation from a high-temperature source—the resulting
temperature response is
q 4αt 1/2 −x 2 /4αt x
T − Ti = s e − x erfc
k π (4αt )1/2
SOLUTION
1-D transient:
∂ 2T 1 ∂T
=
∂x 2 α ∂t
Initial condition:
t = 0; T = Ti .
Boundary conditions:
i. x = 0; −k ∂T
∂x = qs
ii. x → ∞; T = T i
Let θ = T − T i ,
∂ 2θ 1 ∂θ
=
∂x 2 α ∂t
Initial condition:
t = 0; θ = 0.
Boundary conditions:
i. x = 0; ∂θ = q
−k ∂x s
ii. x → ∞; θ = 0
Transient Heat Conduction 97
∂ 2 θ̃ s
− θ̃ = 0 (1)
∂x 2 α
Boundary conditions:
∂ θ̃ = sq
i. x = 0; −k ∂x s
ii. x → ∞; θ̃ = 0.
Solving,
√ √
θ̃ = C1 e− s/αx + C2 e s/αx (2)
Applying BC (ii), C2 = 0. √
Applying BC (i), C1 = (qs /s)(1/k s/α).
Substituting this into above (2),
q 1 √
θ̃ = s √ e− s/αx (3)
s k s/α
Rearranging,
√
q /k α √ √
θ̃ = s 3/2 e−(x/ α) s
s
4.4. Solve transient temperature profiles in a semiinfinite solid body using the
Laplace transform for the following BCs of convective heat transfer to the
surface.
If at time t = 0 the surface is suddenly exposed to a fluid at temperature
T∞ , with a convective heat transfer coefficient h, the resulting temperature
response is
T − Ti x 2 x h
= erfc − ehx/k +(h/k ) αt erfc + (αt )1/2
T∞ − Ti (4αt )1/2 (4αt )1/2 k
SOLUTION
1-D transient
∂ 2T 1 ∂T
=
∂x 2 α ∂t
98 Analytical Heat Transfer
Initial condition:
t = 0; T = Ti .
Boundary conditions:
i. x = 0; ∂θ = h(θ − θ )
−k ∂x ∞ 0
ii. x → ∞; T = Ti
Let θ = T − Ti ,
∂ 2θ 1 ∂θ
=
∂x 2 α ∂t
Initial condition:
t = 0; θ=0
Boundary conditions:
i. x = 0; ∂θ = h(θ − θ )
−k ∂x ∞ 0
ii. x → ∞; θ=0
Applying the Laplace transform,
∂ 2 θ̃ s
− θ̃ = 0 (1)
∂x 2 α
Boundary conditions:
i. x = 0; −k dθ̃(0,t
dx
)
= h θs∞ − θ̃(0, t )
ii. x → ∞; θ̃ = 0
By solving,
√ √
θ̃ = C1 e− s/αx + C2 e s/αx (2)
Applying BC (ii), C2 = 0.
Applying BC (i),
(h/k )θ∞
C1 = √
((h/k ) + ( s/α))s
By rearranging,
√ √ √
(h/k ) α√ −(x/ α) s
θ̃ = θ∞ √ e
(h/k ) α+ s s
Transient Heat Conduction 99
Remarks
There are many engineering problems involving 0-D, 1-D, 2-D, or 3-D tran-
sient heat conduction with various thermal BCs. In the undergraduate-level
heat transfer, we have focused mainly on how to apply the lumped capaci-
tance solutions to solve relative simple engineering problems. For 1-D and
multidimensional transient conduction problems, we normally do not go
through the detailed mathematical equations and solutions. Instead, students
are expected to apply these formulas to solve many engineering relevant prob-
lems by giving solid material geometry with appropriate thermal properties
and thermal BCs.
In the intermediate-level heat transfer, Chapter 4 we have introduced sev-
eral very powerful mathematical tools such as similarity method, Laplace
transform method, and integral approximate method, in addition to the sep-
aration of variables method already mentioned in Chapter 3. Specifically, the
separation of variables method is convenient to solve the transient conduc-
tion problems with finite-length dimensions such as the plate, cylinder, and
sphere with various thermal BCs. However, the similarity method, Laplace
transform method, or integral approximate method is more appropriate to
solve the transient conduction problems with semiinfinite solid material for
various thermal BCs.
1-D transient heat conduction with moving boundaries belongs to advanced
heat conduction material. (Readers can skip these topics.)
PROBLEMS
4.1. The wall of a rocket nozzle is of thickness L = 25 mm and is made
from a high-alloy steel for which ρ = 8000 kg/m3 , c = 500 J/kg K,
and k = 25 W/m K. During a test firing, the wall is initially at
Ti = 25◦ C and its inner surface is exposed to hot combustion gases
for which h = 500 W/m2 K and T∞ = 1750◦ C. The firing time is
limited by the nozzle inner-wall temperature when it reaches
1500◦ C. The outer surface is well insulated.
a. Write the transient heat conduction equation, the associated
BCs, and determine the nozzle wall temperature distributions.
100 Analytical Heat Transfer
(Note: The diameter of the nozzle is much larger than its thick-
ness. No need to perform integration of orthogonal functions
if you run out of time.)
b. Sketch several nozzle wall temperature profiles during tran-
sient heating.
c. To increase the firing time, changing the wall thickness L is
considered. Should L be increased or decreased? Why? The
value of the firing time could also be increased by select-
ing a wall material with different thermophysical properties.
Should materials of larger or smaller values of ρ, c, and k be
chosen?
4.2. A one-side-insulated metal plane wall with a thickness of L is
initially at temperature Ti and suddenly the other side is heated
by forced convection water at temperature T∞ with a convection
heat transfer coefficient h.
a. Outline, step by step, the procedures and the associated initial
and BCs that may be used to solve the temperature distribu-
tions in the plane wall. You do not need to solve the transient
temperature distribution.
b. Sketch the temperature profiles in the plane wall during the
heating process. Also, estimate the surface temperature at the
final steady-state condition.
4.3. A long metal plane wall with a thickness of 2L is initially at tem-
perature Ti and suddenly both sides are heated by convection
fluid flow at temperature T∞ with a convection heat transfer coef-
ficient h. Outline the procedures that may be used to solve the
temperature distributions in the plane wall and sketch the tem-
perature profiles in the plane wall during the heating process for
two different cases.
a. Fluid flow is natural convection air.
b. Fluid flow is forced convection water.
c. Also, estimate the surface temperature at the final steady-state
condition. Which fluid flow will reach the steady temperature
faster? Why? Make appropriate assumptions in order to justify
your answers.
4.4. A large flat plate (with a thickness of 2L) initially at Ti is suddenly
plunged into a liquid bath at T∞ . Derive an expression for the
instantaneous temperature distribution in the plate, if,
a. The heat transfer coefficient between the two surfaces of the
plate and the liquid, h, is given as constant and finite.
b. The heat transfer coefficient h between the two surfaces of the
plate and the liquid is very large so that the temperatures on the
two surfaces of the plate may be assumed to change abruptly
to the temperature of the liquid (i.e., T(L,t) = T∞ , for t > 0).
c. Sketch the instantaneous temperature distribution in the plate
for both (a) and (b) if T∞ > Ti .
4.5. A semiinfinite solid initially at a uniform temperature Ti and
suddenly exposed at its surface to a constant heat flux q”.
a. Determine the temperature history in the solid.
Transient Heat Conduction 101
(1) x = ±a, T = Ts
y = ±b, T = Ts
z = ±c, T = Ts
References
1. V. Arpaci, Conduction Heat Transfer, Addison-Wesley Publishing Company, Read-
ing, MA, 1966.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
4. B. Mikic, Conduction Heat Transfer, Class Notes, MIT, MA, 1974.
5. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
5
Numerical Analysis in Heat Conduction
y T (x, y)
T1
Grids cells
T0 T0
x
T0
FIGURE 5.1
Finite difference method to solve 2-D heat conduction problem.
4
qi→0 + q̇(Δx · Δy · 1) = 0 (5.1)
i=1
T 1 − T0 T 2 − T0 T 3 − T0
kΔy · 1 · + kΔy · 1 · + kΔx · 1 ·
Δx Δx Δy
T 4 − T0
+ kΔx · 1 · + q̇(Δx · Δy) = 0 (5.2)
Δy
If Δx = Δy,
1 q̇ΔxΔx
T0 = T1 + T2 + T3 + T4 + (5.3)
4 k
T 1 − T0 Δx T 4 − T0 Δx T 3 − T0
kΔy · 1 · +k ·1· +k ·1·
Δx 2 Δy 2 Δy
Δx
+ hΔy(T∞ − T0 ) + q̇ · Δy = 0 (5.4)
2
Numerical Analysis in Heat Conduction 107
1
Δx
2
T3
Insulated h, T∞
wall
Δy T1 T0
h, T∞
Δy T4
Convection
y
x qx'
Uniform heat flux
T3
1 T1 T2
Δy
2
T0 qx'
Δx
T3 m, n + 1 i, j+1
T0 T2 m – 1, n m + 1, n i – 1, j Δy i + 1, j
T1 Δy Δy
Δx Δx Δx
m, n – 1 i, j–1
T4
FIGURE 5.2
Finite difference method to solve 2-D heat conduction problem.
Δy T 1 − T0 T 3 − T0 Δy T 2 − T0
k ·1· + kΔx · 1 · +k ·1·
2 Δx Δy 2 Δx
Δy
+ qs Δx · 1 + q̇ · Δx = 0 (5.5)
2
Another approach is to make grid nodes directly from the heat conduc-
tion equation. The 2-D steady-state heat conduction equation with heat
generation is
∂ 2T ∂ 2T q̇
2
+ 2
+ =0
∂x ∂y k
q̇
(Tm−1,n + Tm+1,n + Tm,n−1 + Tm,n+1 ) + (Δx)2 = 4Tm,n (5.7)
k
where m = 1, 2, 3, . . ., n = 1, 2, 3, . . . .
The above linear equation can be applied to any interior nodes. Theoreti-
cally, one would obtain m × n linear equations, and therefore, temperature
T(x, y) = Tm,n can be solved using the matrix method [1,2]. For example,
let Tm,n = T1 , T2 , T3 , . . . , TN , and the above linear equations can be applied
T1 , T1 , T3 , . . . , TN . Rearranging the equation, one obtains
where
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
a11 a12 ··· a1N T1 C1
⎢ a21 a22 ··· a2N ⎥ ⎢ T2 ⎥ ⎢ C2 ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
[A] = ⎢ . ⎥, [T] = ⎢ .. ⎥ , [C] = ⎢ .. ⎥
⎣ .. ⎦ ⎣ . ⎦ ⎣ . ⎦
aN1 aN2 ··· aNN TN CN
Numerical Analysis in Heat Conduction 109
Example 5.1
We use Figure 5.3 as an example to demonstrate how to solve the 2-D heat
conduction problem by using the finite-difference method. Let Δx = Δy , q̇ = 0.
Rearrange the temperatures from the energy balance on node 1, 2, 3, ….
1
T1 = (Ts + Ts + T2 + T3 ) ⇒
4
−4T1 + T2 + T3 + 0 + 0 + 0 + 0 + 0 = −2Ts
1
T2 = (T1 + T4 + T1 + Ts )
4
2T1 − 4T2 + 0 + T4 + 0 + 0 + 0 + 0 = −Ts
1
T3 = (Ts + T5 + T4 + T1 )
4
1
T4 = (T3 + T6 + T3 + T2 )
4
1
T5 = (Ts + T7 + T6 + T3 )
4
1
T6 = (T5 + T8 + T5 + T4 )
4
1 hΔx
T7 = 2T5 + T8 + Ts + 2 T∞
4 + 2(hΔx/k ) k
1 hΔx
T8 = T 6 + T7 + T∞
2 + (hΔx/k ) k
110 Analytical Heat Transfer
Ts
1 2 1
Ts 3 4 3 Ts
5 6 5
7 8 7
T∞, h
FIGURE 5.3
Example of using finite difference method to solve 2-D heat conduction problem.
Place temperatures on the left-hand side of the equation and the constants on the
right-hand side. We can form a coefficient matrix [A], temperature matrix [T ], and
column matrix [C ]. The linear equations of the finite-difference energy balance
on each grid point can be represented by the product of [A][T ] = [C ]. Therefore,
the temperature distribution can be obtained if we know how to solve [T ] from
[A] and [C ]. So the main job for this method is how to obtain [A] and [C ]. The
solution for [T ] is
[T ] = [A]−1 [C ]
⎡ ⎤
−4 1 1 0 0 0 0 0
⎢2 −4 0 1 0 0 0 0 ⎥
⎢ ⎥
⎢1 0 −4 1 1 0 0 0 ⎥
⎢ ⎥
⎢0 1 2 −4 0 1 0 0 ⎥
⎢ ⎥
⎢0 0 1 0 −4 1 1 0 ⎥
[A] = ⎢⎢0
⎥
⎥
⎢ 0 0 1 2 −4 0 1 ⎥
⎢ 2h ⎥
⎢0 0 0 0 2 0 − 4+ Δx 1 ⎥
⎢ ⎥
⎢ k ⎥
⎣ h ⎦
0 0 0 0 0 1 1 − 2 + Δx
k
⎡ ⎤
−Ts ⎡ ⎤
⎢ −Ts ⎥ T1
⎢ ⎥ ⎢T2 ⎥
⎢ −T ⎥ ⎢ ⎥
⎢ s ⎥ ⎢T3 ⎥
⎢ 0 ⎥ ⎢ ⎥
⎢ ⎥ ⎢T4 ⎥
⎢ −Ts ⎥ ⎢ ⎥
[C ] = ⎢ ⎥ [T ] = ⎢T5 ⎥
⎢ 0 ⎥ ⎢ ⎥
⎢ ⎥ ⎢T6 ⎥
⎢ ⎥ ⎢ ⎥
⎢−Ts − 2h ΔxT∞ ⎥ ⎣T7 ⎦
⎢ k ⎥
⎣ h ⎦
T8
− ΔxT∞
k
Numerical Analysis in Heat Conduction 111
Example 5.2
⎡ ⎤
T1,1
⎢T ⎥
⎢ 2,1 ⎥
⎢ ⎥
⎢ .. ⎥
⎢ . ⎥
⎢ ⎥
⎡ ⎤ ⎢ ⎥
T1,1 T2,1 T3,1 T4,1 T5,1 ⎢T1,1 ⎥
⎢ ⎥
⎢T1,2 T5,2 ⎥ ⎢ ⎥
⎢ T2,2 T3,2 T4,2 ⎥ ⎢T2,2 ⎥
[T ] = ⎢
⎢T1,3 T2,3 T3,3 T4,3 ⎥ ⎢
T5,3 ⎥ = ⎢ . ⎥
.. ⎥
⎣T1,4 T2,4 T3,4 T4,4 T4,5 ⎦ ⎢ ⎢
⎥
⎥
T1,5 T2,5 T3,5 T4,5 T5,5 ⎢T ⎥
⎢ 1,5 ⎥
⎢ ⎥
⎢T2,5 ⎥
⎢ ⎥
⎢ . ⎥
⎢ .. ⎥
⎣ ⎦
T5,5
⎡ ⎤
C1,1
⎢C ⎥
⎢ 2,1 ⎥
⎢ ⎥
⎢ .. ⎥
⎢ . ⎥
⎢ ⎥
⎡ ⎤ ⎢ ⎥
C1,1 C2,1 C3,1 C4,1 C5,1 ⎢C1,1 ⎥
⎢ ⎥
⎢C1,2 C5,2 ⎥ ⎢ ⎥
⎢ C2,2 C3,2 C4,2 ⎥ ⎢C2,2 ⎥
[C ] = ⎢
⎢C1,3 C2,3 C3,3 C4,3 C5,3 ⎥ = ⎢ . ⎥
⎥ ⎢ . ⎥
⎣C1,4 C2,4 C3,4 C4,4 C4,5 ⎦ ⎢ ⎢
. ⎥
⎥
C1,5 C2,5 C3,5 C4,5 C5,5 ⎢ ⎥
⎢C1,5 ⎥
⎢ ⎥
⎢C2,5 ⎥
⎢ ⎥
⎢ . ⎥
⎢ .. ⎥
⎣ ⎦
C5,5
⎡ ⎤
a11 a21 a31 a41 a51
⎢a12 a22 a32 a42 a52 ⎥
⎢ ⎥
[A] = ⎢
⎢a13 a23 a33 a43 a53 ⎥
⎥
⎣a14 a24 a34 a44 a54 ⎦
a15 a25 a35 a45 a55
4
qs" = 0 h, T∞
insulated 3 Convection
wall
2
n = 1, ..., 5
y
2 3 4 5
x
qs˝ uniform heat flux
m = 1, ..., 5
FIGURE 5.4
Example of using finite difference method to solve 2-D heat conduction problem with various
boundary conditions.
k = 1,2,3, ..., k
T (x, y, x)
i = 1,2,3, ..., i
j = 1,2,3, ..., j
FIGURE 5.5
Finite difference method to solve 3-D heat conduction problem.
For the case of a curved boundary [3], the interior nodes can be determined as
shown in Figure 5.6.
Unknown :Ti,j
Ti, j+ 1
bΔy Δy
Ti – 1, j Ti , j Ti+ 1, j
cΔy Ti, j– 1
Δx
aΔx
FIGURE 5.6
Finite difference method to solve the interior nodes next to the curved boundary.
114 Analytical Heat Transfer
∂ 2T q̇ 1 ∂T
+ = .
∂x2 k α ∂t
This equation is a parabolic equation. As discussed in Chapter 4, this equa-
tion can be solved analytically by using the separation of variables method,
similarity method, Laplace transform method, or integral method. In this
chapter, we would like to solve 1-D and 2-D transient conduction prob-
lems with various BCs by using the finite-difference explicit method and the
finite-difference implicit method [1].
Example 5.3
Energy balance at the interior nodes, for example, node 2, as shown in Figure 5.7:
q = energy storage
Lower bond:
1 P +1
T1P + T3P − 2T2P = (T − T2P )
Fo 2
T2P +1 = Fo(T1P + T3P ) + (1 − 2Fo)T2P (5.15)
Numerical Analysis in Heat Conduction 115
Δx
Ts
p+1
T1
p+1
T2
p+1
T3 p+1
p
T4
T1 p + 1 step
y p
T2 p p step
T3 p
T4
Ti
x 0 1 2 3 4
1
Δx
2
FIGURE 5.7
Finite difference energy balance method for one-dimensional transient heat conduction with
given surface temperature boundary condition.
Example 5.4
p p p+1 p
p T − T0 T − T0 Δx
h(T∞ − T0 ) + k 1 = 0 ρCp (5.16)
Δx Δt 2
p+1 p 2h Δt p 2αΔt p p
T0 = T0 + (T∞ − T0 ) + (T − T0 )
ρCp Δx Δx 2 1
p p
= 2Fo T1 + Bi T∞ + (1 − 2 Fo − 2 Bi Fo) T0 (5.17)
where
p+1
h, T∞ T0
p+1
T1
p+1
p T2 p+1
T0 T3
p p + 1 step
y T1
p
T2 p p step
T3
Ti
1
x Δx
2
FIGURE 5.8
Finite difference energy balance method for one-dimensional transient heat conduction with
convection boundary conditions.
Example 5.5
p p p+1 p
T − T0 T − T0 Δx
qs + k 1 = 0 ρCp (5.18)
Δx Δt 2
p+1 Δt k Δt p p
T0 = q + T + (1 − 2Fo)T0 (5.19)
(Δx/2)ρCp s (Δx/2)ρCp 1
p+1
T0
p+1
q"s T1 p+1
T2 p+1
p
T3
T0 p + 1 step
y p
T1 p p step
T2 p
T3
Ti
1
x Δx
2
FIGURE 5.9
Finite difference energy balance method for 1-D transient heat conduction with surface heat flux
boundary condition.
Numerical Analysis in Heat Conduction 117
q = energy storage
p+1 p+1 p+1 p+1 p+1 p
T1 − T2 T3 − T2 T2 − T2
k +k + q̇ · Δx · 1 = ρCp Δx (5.20)
Δx Δx Δt
∂ 2T ∂ 2T q̇ 1 ∂T
+ + =
∂x 2 ∂y 2 k α ∂t
Ta
x
1 2 3
T∞, h 4 5 6 T∞, h
7 8 9
Tb
FIGURE 5.10
Finite difference method to solve a 2-D conduction problem.
Example 5.6
Figure 5.10 shows a long, square bar with opposite sides maintained at Ta and Tb ,
and the other two sides lose heat by convection to a fluid at T∞ . The conductivity
of the bar material is k convective heat transfer coefficient is h. For the given mesh,
use the finite-difference energy balance method to obtain a coefficient matrix [A],
temperature matrix [T ], and a column matrix [C ].
SOLUTION
Figure 5.10 shows the prescribed surface conditions and the nodes. Symmetry
allows us to consider just nine nodes. For the interior, nodal temperatures are
1
T2 = (Ta + T1 + T3 + T5 )
4
1
T3 = (Ta + T2 + T2 + T6 )
4
1
T5 = (T2 + T4 + T6 + T8 )
4
1
T6 = (T3 + T5 + T5 + T9 )
4
120 Analytical Heat Transfer
1
T8 = T5 + T7 + T9 + Tb
4
1
T9 = T6 + T8 + T8 + Tb
4
k Δy k Δx k Δy k Δx k Δx
− + + hΔy T1 + T2 + T4 = − Ta − hΔyT∞
Δx Δy Δx 2Δy 2Δy
k Δx k Δy k Δx k Δy k Δx
T − + + hΔy T4 + T5 + T7 = −hΔyT∞
2Δy 1 Δx Δy Δx 2Δy
k Δx k Δy k Δx k Δy k Δx
T4 − + + hΔy T7 + T8 = − T − hΔyT∞
2Δy Δx Δy Δx 2Δy b
⎡
k Δy k Δx k Δy k Δx
⎢− + + hΔy
⎢ Δx Δy Δx 2Δy
⎢ −4
⎢ 1 1
⎢
⎢ 2 −4
⎢
⎢
⎢ k Δx k Δy k Δx
⎢ − + + hΔy
⎢ 2Δy Δx Δy
⎢
[A] = ⎢ 1 1
⎢
⎢
⎢ 1
⎢
⎢ k Δx
⎢
⎢
⎢ 2Δy
⎢
⎢
⎣
⎤
⎥
1 ⎥
⎥
⎥
1 ⎥
⎥
k Δy k Δx ⎥
⎥
Δx ⎥
2Δy ⎥
⎥
−4 1 1 ⎥
⎥
2 −4 1⎥
⎥
⎥
k Δy k Δx k Δy ⎥
− + + hΔy ⎥
Δx Δy Δx ⎥
⎥
⎥
1 1 −4 1⎥
⎦
1 2 −4
Numerical Analysis in Heat Conduction 121
⎡ ⎤
⎡ ⎤ k Δx
T1 −Ta − hΔyT∞
⎢ 2Δy ⎥
⎢ ⎥ ⎢ ⎥
⎢T2 ⎥ ⎢ −T a ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢T3 ⎥ ⎢ −Ta ⎥
⎢ ⎥ ⎢ ⎥
⎢T ⎥ ⎢ −hΔyT∞ ⎥
⎢ 4⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
[T ] = ⎢T5 ⎥ [C ] = ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢T6 ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
⎢T7 ⎥ ⎢ k Δx ⎥
⎢ ⎥ ⎢− Tb − hΔyT∞ ⎥
⎢ ⎥ ⎢ 2Δy ⎥
⎢T8 ⎥ ⎢ ⎥
⎣ ⎦ ⎢ −Tb ⎥
T9 ⎣ ⎦
−Tb
Remarks
The finite-difference method is a very powerful numerical technique to solve
many engineering application problems. As long as you know how to perform
the basic energy balance at the interior nodes as well as at the boundary nodes,
this method essentially can solve all kinds of heat conduction problems with
complex thermal BCs. In the undergraduate-level heat transfer, students are
normally required to perform simple energy balance at any specified node
inside a 2-D steady-state solid material and on the boundary.
In the intermediate-level heat transfer, we are more focused on how to per-
form simple energy balance as well as how to discretize the heat conduction
equation in order to solve the 1-D and 2-D steady-state heat conduction prob-
lems with various BCs by using the matrix inverse method. We also put in
effort to solve the 1-D and 2-D transient heat conduction problems with var-
ious BCs by using the finite-difference implicit method and explicit method.
In general, the same technique can be used to solve heat conduction problems
with cylindrical and spherical coordinates.
PROBLEMS
5.1. Refer to Figure 5.4, show the matrices [A], [T], and [C], with the
following grid distributions:
W
[cos2 (aw)] dw = (1/4a)[2aW + sin (2aW)]
0
W
and [cos(aw) · cos(bw)] dw = 0, when a = b.
0
5.10. Derive Equations 5.23 and 5.25 for 3-D transient heat conduction
problems.
124 Analytical Heat Transfer
References
1. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. K.-F. Vincent Wong, Intermediate Heat Transfer, Marcel Dekker, Inc., New York, NY,
2003.
6
Heat Convection Equations
The flow is a laminar flow if Re < 300 × 103 , and a turbulent flow if Re >
300 × 103 . √
Flow boundary-layer thickness δ(x) ∼ x
Dynamic viscosity μ,
Kinematic viscosity ν = μ/ρ
125
126 Analytical Heat Transfer
U∞ ,V∞ U∞ U∞
u (x, y)
y δ (x)
x Laminar Turbulent
flow flow
FIGURE 6.1
Hydrodynamic and thermal boundary layer over a flat plate (if Pr = 1).
Shear stress
∂u 1 2
τw = μ = Cf · (ρU∞ − 0)
∂y y=0 2
Friction coefficient:
1
Cf = aReb ; Cf = aReb ∼ ;
Re
Reynolds number is defined as the fluid inertia force against viscous force
(i.e., the fluid particle tries to move but viscosity tries to resist it from moving)
and is a combination of velocity, viscosity, and length (distance measured from
the leading edge of the plate). When Reynolds number is approximately less
than 300 × 103 , the fluid particle moves like laminar, layer to layer from free-
stream velocity to zero velocity on the solid surface, and creates shear stress
over the solid surface. When the Reynolds number is greater than 300 × 103 ,
the fluid particle tends to become unstable (random motion) and gradually
transitions into the turbulent flow boundary layer. In the laminar boundary
layer, the velocity profile gradually changes from free-stream value to zero
on the surface as a parabolic shape. But, in the turbulent boundary layer,
the velocity profile remains fairly uniform as the free-stream value till near
the surface and then suddenly changes to zero on the surface. This is due to
turbulent mixing (the particle moves up and down, back and forth) so that
free-stream velocity is able to move closer to the surface. From the application
point of view, shear stress (viscosity × velocity gradient at the surface, i.e.,
τw = μ(∂u/∂y)|y=0 ) decreases with decreasing velocity gradient and viscosity.
Heat Convection Equations 127
τw δ U∞
U∞
δ
U∞ δ
Laminar Turbulent
flow flow
0 Transition x 0
FIGURE 6.2
Hydrodynamic boundary layer, friction factor, and shear stress profile.
T∞ T∞ T∞
y
T (x, y)
δ T (x)
x
Laminar Tw Turbulent
0 flow flow
FIGURE 6.3
Thermal boundary layer over a heated flat plate.
128 Analytical Heat Transfer
plate as shown in Figure 6.1, due to conductivity of fluid and velocity distri-
bution, the temperature gradually decreases from its free-stream maximum
value to that at the flat plate. The hot fluid particle conducts heat from the
free-stream into the cold surface through the velocity boundary layer. There-
fore, a thermal boundary-layer thickness is developed over a solid surface
due to fluid flow.
The temperature profile and associated thermal boundary-layer thickness
over a flat plate is solved in Chapter 7. In an ideal case (assume Pr = 1),
the thermal boundary layer is identical to hydrodynamic boundary layer as
shown in Figure 6.1 or 6.3. In this ideal case, the temperature profile is the
same as the velocity profile through the entire boundary layer over the flat
plate. Once we determine the temperature profile over a flat plate, T(y) at a
given distance x, the thermal boundary-layer thickness, the heat flux on the
surface, and the heat transfer coefficient (or Nusselt number) can be obtained
as follows: √
Thermal boundary-layer thickness δT (x) ∼ x
If Pr = 1, then δ(x) = δT (x).
At the body surface, the heat flux is
∂T ∂T
qw = −k = −kf ≡ h(Tw − T∞ )
∂y y=0 ∂y y=0
The heat transfer coefficient h with the unit of W/m2 k can be expressed as
−kf (∂T/∂y)y=0 −kf ((T∞ − Tw )/δT ) kf kf
h= ∼ ∼ ∼ ∼ k f U∞
Tw − T ∞ Tw − T ∞ δT δ
δ h
q"w
δT
Pr > 1 U∞
δT
Laminar Turbulent
δ flow flow
0 Transition x
Pr < 1
FIGURE 6.4
Thermal boundary layer, heat transfer coefficient, and heat flux profile.
130 Analytical Heat Transfer
∂ρ
+ ∇ · (ρV) = 0 (6.1)
∂t
where
V = iu + jv + kw
and
∂ ∂ ∂
∇=i +j +k
∂x ∂y ∂z
are the velocity vector and the del operator for unit vectors, i, j, and k in the
x-, y-, and z-direction, respectively.
Conservation of momentum:
DV
ρ = −∇P + μ∇ 2 V + ρg (6.2)
Dt
Conservation of energy:
Dh DP
ρ = + ∇ · k∇T + μΦ + q̇ (6.3)
Dt Dt
∂ ∂ ∂
− (ρu dy) dx − (ρv dx) dy = (ρ dx dy) (6.5)
∂x ∂y ∂t
ρv dx + ∂ (ρv dx) dy
∂y
U∞
U∞
ρu dy + ∂ (ρu dy) dx
ρu dy ∂x
Δy
y
Δx
x
ρv dx
FIGURE 6.5
Conservation of mass.
132 Analytical Heat Transfer
∂ ∂
− (ρu dy) dx − (ρv dx) dy = 0
∂x ∂y
∂(ρu) ∂(ρv)
+ =0 (6.6)
∂x ∂y
Conservation of momentum:
From Newton’s second law, net force exerting on a body equals the
momentum change.
Fx = max
Fy = may
∂ ∂ ∂Px ∂ 2 ∂ ∂
σx + τxy − = ρu + (ρuv) + (ρu) (6.8)
+∂x
,- . +∂y,- . ∂x
+,-. ∂x
+ ,-
∂y ∂t
. + ,- .
normal pressure unsteady
stress shear convective term
stress gradient term
∂u
σx = 2μ (6.9)
∂x
∂u ∂v
τxy = τyx = μ + (6.10)
∂y ∂x
∂
τxy + (τ ) dy
∂y yx
∂
σx σx + (σ ) dx
∂x x
Px Px + ∂ (Px) dx
∂x
τyx
∂
ρv dx dz.u + (ρv dx dz.u) dy
∂x
∂
ρu.u dy dz + (ρu.u dy dz) dx
ρu.udydz ∂x
ρv dx dz.u
U∞ U∞
FIGURE 6.6
Conservation of momentum.
Substitute the mass conservation equation into and we will obtain the
x-direction momentum equation
∂u ∂u 1 ∂P ∂ 2u ∂ 2u
u +v =− +υ 2 + υ 2 . (6.11)
∂x ∂y ρ ∂x ∂x ∂y
+ ,- . + ,- . + ,- .
convection pressure stress
∂v ∂v 1 ∂P ∂ 2v ∂ 2v
u +v =− +υ 2 +υ 2 (6.12)
∂x ∂y ρ ∂y ∂x ∂y
Conservation of energy:
Unsteady state:
Perform energy balance shown in Figure 6.7,
∂qx ∂qy ∂ ∂
− dx − dy − ρu dy · Cp · T dx − ρv dx · Cp · T dy
∂x ∂y ∂x ∂y
∂ ρ dx dy · Cp · T
= (6.13)
∂t
∂ ρCp T ∂ ∂
+ ρuCp T + ρvCp T
∂t
+ ,- . ∂x ∂y
+ ,- .
unsteady steady convection
∂ ∂T ∂ ∂T q̇
= k + k + + μΦ (6.14)
∂x ∂x ∂y ∂y k +,-.
+ ,- . +,-.
heat dissipation
heat diffusion heat source due to fricion
generation
∂
qy + (qy) dy ρvi dx dz + ∂ (ρvi dx dz) dy
∂y ∂y
qx ∂
qx + (qx) dx
∂x
Econd, x Econd,x + dx
Δy
Econv, x Econv,x + dx
Δx
ρ ui dy dz ρui dy dz + ∂ (ρui dy dz) dx
∂x
Econd,y Econv,y
i = enthalpy
qy ρvi dx dz P
= e + = C pT
ρ
Internal energy
FIGURE 6.7
Conservation of energy.
T∞ U∞
v,y
δ, δT
u,x 0
Tw L
FIGURE 6.8
Boundary-layer approximations.
136 Analytical Heat Transfer
uv
∂u ∂u ∂v ∂v
, ,
∂y ∂x ∂x ∂y
∂T ∂T
∂y ∂x
∂u ∂v 1 ∂P ∂ 2u
u +u =− +ν 2 (6.17)
∂x ∂y ρ ∂x ∂y
1 ∂P
− =0 (6.18)
ρ ∂y
For an incompressible flow, that is, M < 0.2, Φ ∼ 0, and so the energy
equation becomes
∂T ∂T ∂ 2T
u +v =α 2 (6.19)
∂x ∂y ∂y
∂U∞ 1 ∂P
U∞ =− (6.20)
∂x ρ ∂x
Note that Equation 6.18 implies that there is no pressure change in the
y-direction within the boundary layer. Equation 6.20 implies that pressure
change in the x-direction within the boundary layer can be predetermined
from velocity and its velocity change in the x-direction outside of the bound-
ary layer. Therefore, Equation 6.20 can be substituted into Equation 6.17 to
solve for the velocity profiles inside the boundary layer.
x y u v
Let x∗ = , y∗ = , u∗ = , v∗ = ,
L L U∞ U∞
T − Tw P
T∗ = , P∗ = , (6.21)
T∞ − T w ρU∞2
∂u∗ ∂v∗
∗
+ ∗ =0
∂x ∂y
∂u∗ ∂u∗ ∂P∗ υ ∂ 2 u∗
u∗ ∗
+ v∗ ∗ = − ∗ + · ∗2
∂x ∂y ∂x U∞ L ∂y
∂P∗
=0
∂y∗
∂T ∗ ∂T ∗ α ∂ 2T∗
u∗ ∗
+ v∗ ∗ = · ∗2
∂x ∂y U∞ L ∂y
α 1 1
= = (6.22)
U∞ L (U∞ L/υ) · (υ/α) Re · Pr
where
∗
∂u∗ dP
= f2 x∗ , ReL ,
∂y∗ y∗ =0 dx∗
τw m(U∞ /L) ∗ dP∗
Cf = = f
2 2
x , ReL ,
(1/2)ρV∞2 (1/2)ρU∞ dx∗
∗
2 ∗ dP
Cf = f2 x , ReL , (6.23)
ReL dx∗
138 Analytical Heat Transfer
∗
Special case: when flow over a flat plate dP /dx∗ = 0, the average friction
factor can be determined from Reynolds number as
2
C̄f = f2 (ReL ) = aReLm (6.24)
ReL
hL
NuL ≡ = f5 (ReL , Pr) = aReLm Prn (6.26)
k
The above analysis concludes that, for flow over a flat plate, the local friction
factor (at a given location x) is a function of Reynolds number only, and the
local heat transfer coefficient or Nusselt number (at a given location x) is a
function of Reynolds number as well as Prandtl number.
2
Cfx = f2 x∗ , ReL (6.27)
ReL
Nux = f4 x∗ , ReL , Pr (6.28)
If f2 = f4 , Pr = 1,
ReL
Cf = f2 = f4 = Nu
2
1 Nu (hL/k) h
Cf = = St = = (6.29)
2 Re · Pr (ρVL/μ) · (μCp /k) ρCp V
Heat Convection Equations 139
Reynolds analogy:
1
Cf = St (6.30)
2
Experimentally, we obtained
1
Cf Pr−2/3 = St (6.31)
2
Remarks
This chapter provides the basic concept of boundary-layer flow and heat
transfer; it focuses on how to derive 2-D boundary-layer conservations
for mass, momentum, and energy; boundary-layer approximations; non-
dimensional analysis; and Reynolds analogy. Students have come across
these equations in their undergraduate-level heat transfer. However, in the
intermediate-level heat transfer, students are expected to fully understand
how to obtain these equations.
PROBLEMS
6.1. For hot-gas flow (velocity V∞ , temperature T∞ ) over a cooled
convex surface (surface temperature Ts ), answer the following
questions:
a. Sketch the “thermal boundary-layer thickness” distribution on
the entire convex surface and explain the results.
b. Sketch the possible local heat transfer coefficient distribution
on the convex surface and explain the results.
c. Define the similarity parameters (dimensionless parameters)
that are important to determine the local heat transfer coeffi-
cient on the convex surface.
d. Write down the relationship among those similarity parame-
ters and give explanations.
e. Write down how to determine the local heat flux from the
convex surface.
6.2. For cold-gas flow (velocity V∞ , temperature T∞ ) over a heated
convex surface (surface temperature Ts ), answer the following
questions:
140 Analytical Heat Transfer
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
3. H. Schlichting, Boundary-Layer Theory, Sixth Edition, McGraw-Hill Book
Company, New York, NY, 1968.
7
External Forced Convection
∂Ψ
u= (7.1)
∂y
∂Ψ
v=− (7.2)
∂x
∂ ∂Ψ ∂ ∂Ψ
+ − =0 (7.3)
∂x ∂y ∂y ∂x
∂Ψ ∂ 2Ψ ∂Ψ ∂ 2 Ψ ∂ 3Ψ
− =υ (7.4)
∂y ∂x∂y ∂x ∂y2 ∂y3
Ψ(x, y) ⇒ Ψ(η)
141
142 Analytical Heat Transfer
U∞
U∞
Streamline
FIGURE 7.1
Stream lines for flow over a flat plate.
where
∂f ∂η C2 ∂η 1 η
f = , =√ , = − C2 yx−(3/2) = − ,
∂η ∂y x ∂x 2 2x
Similarly,
∂ 2Ψ ∂ ∂Ψ ∂ c2 c2 ∂f ∂η c2 f
= = f = = 2√
∂y2 ∂y ∂y ∂y c1 c1 ∂η ∂y c1 x
External Forced Convection 143
∂ 3Ψ ∂ ∂ 2Ψ ∂ C22 f C2 ∂ f ∂η
3
= = √ = 2 √ ·
∂y ∂y ∂y2 ∂y C1 x C1 ∂η x ∂y
at
∂Ψ c2
y = ∞, u = U∞ = = f
∂y y=∞ c1
Let
c2
= U∞ . (7.9)
c1
Therefore,
'
y y U∞ y
η = C2 √ = √ = Rex (7.10)
x x υ x
Ψ Ψ
f = C1 √ = √ (7.11)
x υU∞ x
∂Ψ C2
u= = f = U∞ f (7.12)
∂y C1
'
∂Ψ 1 1 1 υU∞
v=− =− √ f − f η = f η−f (7.13)
∂x 2 C1 x 2 x
144 Analytical Heat Transfer
df u
f = = = velocity profile (7.14)
dη U∞
d2 f d(u/U∞ )
f = 2
= = velocity gradient (7.15)
dη dη
Boundary conditions:
at
u
y = 0, u = v = 0, ⇒ η = 0, f = = 0; v = 0, ⇒ f = 0 ⇒ f (0) = 0 (7.16)
U∞
at y = ∞, u = U∞ , ⇒ f (∞) = 1
1
f = − ff (7.17)
2
Equation 7.17 may be solved numerically by expressing f (η) in a power
series (1908 Blasius Series Expansion) with the above-mentioned BCs as
where C3 = 0 at η = 0, f = 0.
External Forced Convection 145
Therefore,
η η η
− 0 1/2( f dη)
0 0e dη dη
f = ∞ − η 1/2(f dη) (7.20)
0 e dη
0
For example, use of the trapezoidal rule for the numerical integrations:
Choose ηmax = 5 = Δη · N, when N = 50 (the number of integration steps),
Δη = 0.1
Let ηi+1 = ηi + Δη for i = 0, 1, 2, . . . , with ηo = 0
Initial guess fi = ηi
η ηi η ηi η η ηi
Calculate 0 i f dη, e− 0 f dη , 0 i e− 0 f dη dη, and 0 i 0 i e− 0 f dη dη dη
η ηi
Calculate C1 = 1/ 0 max e− 0 f dη dη , fi , fi , fi
Check convergence 1 − finew /fiold < ε?, for i = 0, 1, 2, . . . , N. If not, go back
again.
From the tabulated data shown in Table 7.1 [2] or Figure 7.2, for given
Rex = ρU∞ x/μ, the velocity profile u x, y at any location (x, y) and the shear
stress f at the wall (y = 0, η = 0) can be determined.
TABLE 7.1
Flat Plate Laminar Boundary Layer Functions [2]
η=y U∞
vx f f = Uu f
∞
0 0 0 0.332
0.4 0.027 0.133 0.331
0.8 0.106 0.265 0.327
1.2 0.238 0.394 0.317
1.6 0.420 0.517 0.297
2.0 0.650 0.630 0.267
2.4 0.922 0.729 0.228
2.8 1.231 0.812 0.184
3.2 1.569 0.876 0.139
3.6 1.930 0.923 0.098
4.0 2.306 0.956 0.064
4.4 2.692 0.976 0.039
4.8 3.085 0.988 0.022
5.2 3.482 0.994 0.011
5.6 3.880 0.997 0.005
6.0 4.280 0.999 0.002
6.4 4.679 1.000 0.001
6.8 5.079 1.000 0.000
146 Analytical Heat Transfer
u
f ′=
U∞ Universal
velocity profile
0.5
Experimental data
y 5
η= x Rex
FIGURE 7.2
Graphical sketch of velocity profile from similarity.
∂Ψ ∂θ ∂Ψ ∂θ ∂ 2θ
− =α 2 (7.22)
∂y ∂x ∂x ∂y ∂y
U∞,T∞ T∞
x
Tw
FIGURE 7.3
Concept of thermal boundary layer.
External Forced Convection 147
Performing
∂θ ∂θ ∂η C2
= = θ √
∂y ∂η ∂y x
∂ 2θ ∂ ∂θ ∂ C2 ∂ C2 ∂η C22
= = √ θ = √ θ · = θ
∂y2 ∂y ∂y ∂y x ∂η x ∂y x
∂θ ∂θ ∂η −η
= · = θ ·
∂x ∂η ∂x 2x
1
θ + Pr f θ = 0 (7.23)
2
Boundary conditions:
θ (0) = 0
(7.24)
θ (∞) = 1
at η = 0, θ(0) = 0, C2 = 0
at η = 0, θ(∞) = 1,
1
C1 = η
η − Pr/2( f dη) dη
0 e 0
Therefore,
η −
η
Pr/2( f dη) dη
0e
0
θ= η (7.25)
∞
0 e− 0 Pr/2( f dη) dη
η η
Pr
θ = θ (0) exp − f dη dη (7.26)
0 0 2
148 Analytical Heat Transfer
where
1
θ (0) = ∞ η (7.27)
0 exp − 0 Pr/2( f dη) dη
From the tabulated data or Figure 7.4, the temperature profile T(x, y) at any
location (x, y) and the heat flux at the wall (y = 0, η = 0) can be determined.
From η, we obtain θ and T(x, y) for the given Prandtl number.
T − Tw T − T∞
θ= θ=
T∞ − Tw Tw − T∞
1 1
Pr > 1
Pr = 1
Pr < 1
Pr < 1
Pr = 1
Pr > 1
5 5
U∞
η=y η=y
U∞
vx vx
FIGURE 7.4
Graphical sketch of temperature profile from similarity solutions.
External Forced Convection 149
√
Let u/U∞ = 0.99 at η = 5, that is, η = √
(y/x) Rex = 5, where y = δ, and the
edge of the boundary layer, δ = y = 5x/ Rex
τw 2f (0) 0.664 1
Cfx = = √ =√ ∼√ (7.29)
2
(1/2)ρU∞ ρU∞ x/μ Rex x
1.328
C̄fx =
ReL
when
ρU∞ L
ReL =
μ
'
∂T ∂T ∂η ∂θ U∞
q = −k = −k = k(Tw − T∞ )
∂y y=0 ∂η ∂y y=0 ∂η υx
η=0
'
U∞
= −k(T∞ − Tw )θ (0)
υx
θ (0) = 0.332 Pr1/3 from the Table 7.1 (for Pr = 1) or Figure 7.4
'
q −k (∂T/∂y)y=0 U∞ 1
h= = =k · θ (0) ∼ √ (7.30)
Tw − T ∞ Tw − T ∞ υx x
hx
Remarks
There are many engineering applications involving external laminar flow
heat transfer such as electronic components cooling and plate-type heat
exchangers design. In the undergraduate-level heat transfer, there are many
heat transfer relations between Nusselt numbers and Reynolds and Prandtl
150 Analytical Heat Transfer
ρvdx or d ∫ ρu dy dx δ
dx
d
∫ ρu dy ∫ ρu dy + ∫ (ρu dy)dx
dx
Conservation of mass
d
ρvU∞dx or U∞ ∫ ρu dy dx δ
dx
d
∫ ρuu dy + ∫ (ρuu dy)dx
∫ ρuu dy dx
Momentum change
δ
P dδ
∂
∫ P dy ∫ P dy + ∫ P dy dx
∂x
τw dx
Net force
FIGURE 7.5
Integral method.
⎛δ ⎞
dδ d ⎝ d d
−τw dx + P dx − P dy⎠dx = (ρuu dy) dx − U∞ (ρu dy) dx
dx dx dx dx
0 + ,- .
+ ,- . momentum change
net force
(7.32)
where
d d dδ dP
− P dy = − P dy = −P −δ
dx dx dx dx
d dU∞ d
= ρuu dy − ρU∞ dy − U∞ ρu dy
dx dx dx
dU∞ dU∞
+ ρu dy − ρu dy
dx dx
d dU∞ d
= ρuu dy + ρ (u − U∞ ) dy − ρuU∞ dy
dx dx dx
d dU∞
= ρu (u − U∞ ) dy + ρ (u − U∞ ) dy
dx dx
d dU∞
τw = ρu (U∞ − u) dy + ρ (U∞ − u) dy (7.33)
dx dx
For a flat plate flow, dU∞ /dx = 0
∂u d
τw = μ = ρu(U∞ − u) dy (7.34)
∂y y=0 dx
⎛δ ⎞
δT
d d ⎝
qs dx = ρCp uT dy dx − Cp T∞ ρu dy⎠ dx (7.35)
dx 0 dx
0
δT
d
= ρCp u(T − T∞ ) dy dx
dx
0
δT
∂T d
qs = −k = ρCp u(T − T∞ ) dy (7.36)
∂y y=0 dx
0
d δ
CpT∞ ∫ ρudydx δT
dx 0
δT
δT
∫ 0 ρCpuT dy + (
d δT
∫ ρCpuT dy dx )
∫ 0 ρCpuT dy dx 0
q ″s dx
FIGURE 7.6
Conservation of energy.
External Forced Convection 153
u = a + by
u = a + by + cy2
u = a + by + cy2 + dy3
u = a + by + cy2 + dy3 + ey4
τw μ (∂u/∂y)0
Cfx = 2
= 2
can be determined
(1/2)ρU∞ (1/2)ρU∞
T = a + by
T = a + by + cy2
T = a + by + cy2 + dy3
T = a + by + cy2 + dy3 + ey4
qw −k(∂T/∂y)0
h= = can be determined.
Tw − T ∞ Tw − T ∞
Examples
7.1 Assume third-order velocity and temperature profiles for the boundary layer
flow to satisfy the BCs:
154 Analytical Heat Transfer
u = a + by + cy 2 + dy 3
T = a + by + cy 2 + dy 3
u−0 3 y 1 y 3 u
= − =
U∞ − 0 2 δ 2 δ U∞
Put the above velocity profile into momentum integral to solve for δ(x)
δ
31 d 2 3 y 1 y 3 3 y 1 y 3
μU∞ = ρU∞ − 1− + dy
2δ dx 2 δ 2 δ 2 δ 2 δ
0
d 39 2
= δρU∞
dx 280
dδ 140 υ
δ =
dx 13 U∞
1 2 140 υ
δ = x +C
2 13 U∞
at x = 0, δ = 0, C = 0 √
Therefore, δ (x) = 4.64 (υx/U∞ )
Put δ(x) back to velocity profile to obtain the final velocity profile. And the
friction factor can be calculated as
or
T − T∞ 3 y 1 y 3
=1− +
Tw − T∞ 2 δT 2 δT
Put the above temperature profile into the energy integral to solve for δT (x).
If Pr = 1, δ = δT , u = T − Tw
δT 4.64
= √
x Rex
The coefficient 4.64 is from momentum integral. It equals 5.0 from the
similarity solution.
External Forced Convection 155
δT
∂T 3 1 d 3 y
−k = k (T − T ∞ ) = ρC
∂y y =0
w p
2 δT dx 2 δ
0
1 y 3 3 y 1 y 3
− U∞ (Tw − T∞ ) 1 − + dy
2 δ 2 δT 2 δT
Then,
/ 3
3 1 δT 1 δT 1 d δT
k = ρCp U∞ 3δ −
2 δT δ 10 δ 70 dx δ
4 0
δT 2 1 δT 1 dδ
+3 −
δ 20 δ 280 dx
Let
4
δ δT 3 δT
r = T < 1, ∼ ∼0
δ δ δ
We obtain
dr dδ α
2r 2 δ2 + r 3δ = 10
dx dx U∞
From above
280 υx dδ 140 υ
δ2 = , δ =
13 U∞ dx 13 U∞
We obtain
280 υx dr 140 υ α
2r 2 + r3 = 10
13 U∞ dx 13 U∞ U∞
Then,
dr 13
4r 2 x + r3 =
dx 14Pr
Let s = r 3
dr 1 d 3 1 ds
r2 = r =
dx 3 dx 3 dx
4 ds 13
s+ x =
3 dx 14Pr
156 Analytical Heat Transfer
13
s = r 3 = C1 x −3/4 +
14 Pr
13
at x = 0, C1 = 0 , r 3 =
14Pr
δ
13 1/3 ∼
r = T = = 0.975 ( Pr ) −1/3 ∼ Pr −1/3
δ 14Pr
4.64
δT = δ Pr−1/3 = √ · Pr −1/3
Rex
−k (∂T /∂y )y =0 3 k
h= =
Tw − T∞ 2 δT
3 k 3 k
= −1/3
= √
2 δPr 2 x(4.64/ Rex )Pr −1/3
For this typical case,
hx 1/2
Nux = = 0.323 Rex Pr 1/3 (7.38)
k
1 13 13 3/4
0 = C1 3/4 + , C1 = − x
xo 14Pr 14Pr o
Therefore,
x 3/4
1/3
δ
r = T = 0.975Pr −1/3 1 −
o
δ x
U∞,T∞
δ
δT
Tw
0 x Apply heat
U∞,T∞
δ
δT
x0
Apply heat
x
0
FIGURE 7.7
Integral approximation method.
External Forced Convection 157
δT
= 0.975Pr −1/3
Pr −1/3
δ
1/2
hx 0.323 Rex Pr 1/3
Nux = =
3
(7.39)
k 1 − (x0 /x)3/4
Remarks
For the integral method, students are expected to know how to sketch and
derive momentum and energy integral equations from mass, force, and heat
balance across the boundary layer. Students are also expected to know how
to solve velocity boundary-layer thickness and the friction factor from the
derived momentum integral equation by assuming any velocity profile to
satisfy velocity BCs across the boundary layer; as well as how to solve ther-
mal boundary thickness and the heat transfer coefficient from the derived
energy integral equation by assuming any temperature profile to satisfy ther-
mal BCs (given surface temperature or surface heat flux) across the thermal
boundary layer. Note that one will get a slightly different velocity boundary-
layer thickness (and friction factor) by using different velocity profiles across
the boundary layer, and a slightly different thermal boundary-layer thick-
ness (and a heat transfer coefficient) by using different temperature profiles
across the thermal boundary layer. This is the nature of the integral method.
Another note is that velocity and thermal boundary-layer thickness is the
same if Prandtl number unity is assumed.
PROBLEMS
7.1. Consider a steady, incompressible, low-speed 2-D laminar
boundary-layer flow (at U∞ , T∞ ) over a flat plate at a uni-
form wall temperature TW . Assume that there exist no body
force and constant thermal and fluid properties. The similarity
momentum and energy equations are listed here for reference:
f + (1/2)ff = 0 and θ + (1/2)Pr f θ = 0.
a. From the Blasius solution of the above similarity equations,
sketch the relations between the similarity functions (f , θ) and
the similarity variable (η) for both water and air (i.e., sketch f
versus η for both water and air on the same plot; and θ versus
η for both water and air on the same plot). Explain why they
158 Analytical Heat Transfer
u = a + by + cy2 + dy3
T = a + by + cy2 + dy3
164 Analytical Heat Transfer
Nu = 0.564 Pe1/2
c. Determine the local heat transfer coefficient (hx ) and the Nus-
selt number Nux . Express the answer in terms of the thermal
boundary-layer thickness δt .
7.21. Consider a 2-D laminar air flow over a friction less plate. The flow
approaches the leading edge of the plate with uniform velocity
U∞ and temperature T∞ . The plate is subjected to a constant wall
temperature, Ts (> T∞ ).
a. Qualitatively sketch hydrodynamic (δ) and thermal boundary
layer (δt ) growth as a function of distance x from the leading
edge.
b. Qualitatively sketch velocity and temperature profiles at a
distance x from the leading edge.
c. A third-order polynomial of the form
T − Ts 3
= a0 + a1 (y/δt ) + a2 (y/δt )2 + a3 y/δt
T∞ − Ts
is used to describe the temperature profile. Determine the
constants a0 , a1 , a2 , and a3 using appropriate BCs.
d. Express local Nusselt numbers (Nux ) in terms of local thermal
boundary-layer thickness δt .
e. Set up an integral energy balance equation. Do not attempt to
solve the equation.
7.22. Consider a 2-D, steady, incompressible laminar flow over a flat
plate. The flow approaches the leading edge with free-stream
velocity of U∞ and temperature T∞ . The flat plate is frictionless
and it is kept at a uniform temperature of Ts (> T∞ ).
a. State clearly all the boundary-layer assumptions.
b. If the temperature distribution at any axial distance x is
approximated by a linear profile (T − Ts )/(T∞ − Ts ) = y/δt ,
derive an expression for the local Nusselt number distribution.
7.23. Consider a steady laminar viscous fluid with a free-stream veloc-
ity V∞ and temperature T∞ flows over a flat plate at a uniform
wall temperature Tw . Assume that the thermal fluids properties
are constant.
a. If the fluid has a Prandtl number of one (i.e., Pr = 1.0), deter-
mine the local heat transfer coefficient along the plate. You may
use the method of integral approximation with the assump-
tions of the linear velocity and temperature profiles across the
boundary layers, that is, u = a + by and T = c + dy, where a, b,
c, and d are constants.
b. If the fluid’s Prandtl number is not equivalent to one (i.e., Pr >
1 or Pr < 1), outline the methods (no need to solve) in order
to determine the surface heat transfer. You may use the same
assumptions as in part (a). Does the heat transfer coefficient
increase or decrease with the fluid Prandtl number? Explain
your answers.
7.24. A flat horizontal plate has a dimension of 10 cm × 10 cm. The plate
is maintained at a constant surface temperature of 300◦ K with a
water jacket.
166 Analytical Heat Transfer
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
3. H. Schlichting, Boundary-Layer Theory, Sixth Edition, McGraw-Hill, New York,
NY, 1968.
4. W. M. Kays and M. E. Crawford, Convective Heat and Mass Transfer, Second
Edition, McGraw-Hill, New York, NY, 1980.
5. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
6. E. Levy, Convection Heat Transfer, Class Notes, Lehigh University, 1973.
8
Internal Forced Convection
R
r
U max x
τw
ReD
0 x
FIGURE 8.1
Velocity profile and shear stress distribution in a circular tube or between two parallel plates.
Xf,h Xf,t
Xf,h
= 0.05ReD
D Lam
Xf,h
≅ 20 +
D Turb
Xf,t
= 0.05ReD Pr
D Lam
FIGURE 8.2
Hydraulic entrance length and thermal entrance length in a circular tube or between two parallel
plates.
Internal Forced Convection 169
For the laminar flow, the thermal entrance length to tube diameter ratio is
about 5% of Reynolds number (based on the tube diameter) times Prandtl
number. This implies that the thermal entrance length increases with increas-
ing Reynolds number (because a thinner boundary layer requires longer
distance for the boundary layer to merge) and Prandtl number (because lower
thermal conductivity requires longer distance to merge) [1–4].
Figure 8.2 also shows that the heat transfer coefficient decreases from the
entrance along the tube and becomes a constant value when thermal boundary
layer reaches the fully developed condition, and the heat transfer coefficient
increases with Reynolds number (because of a thinner boundary layer from
the entrance and the longer entrance length). It is noted that the thermal
entrance length is identical to the hydrodynamic entrance length if Pr = 1.
For the turbulent flow, the thermal entrance length is harder to determine;
just like the hydrodynamic entrance length, the thermal entrance length is
around 10–20 tube diameter. It is hard to distinguish whether the turbu-
lent flow is thermally fully developed or not from 10 to 20 tube diameter
downstream.
ρVD VD 4ṁ
ReD = = = (8.1)
μ ν πDμ
Laminar flow is observed if ReD ≤ 2300.
At a certain distance from the entrance, the velocity profile u(r) remains
unchanging along the tube (if the fluid properties remain constant). Corre-
spondingly, there will be no velocity component in the radial direction. Also,
the axial pressure gradient required to sustain the flow against the viscous
forces will be constant along the tube (no momentum change). The flow is
hydrodynamically fully developed. The differential governing equations for
the flow inside a circular tube are [1–4]
∂u 1 ∂(vr)
+ =0 (8.2)
∂x r ∂r
∂u ∂u 1 ∂P 1 ∂ ∂u
u +v =− +υ r (8.3)
∂x ∂y ρ ∂x r ∂r ∂r
∂ (uT) ∂ (vT) 1 ∂ ∂T
+ =α r (8.4)
∂x ∂y r ∂r ∂r
170 Analytical Heat Transfer
v=0 (8.5)
∂u
=0 (8.6)
∂x
1 ∂P 1 ∂ ∂u
0=− +υ r (8.7)
ρ ∂x r ∂r ∂r
1 dP 1 d du
=υ r
ρ dx r dr dr
dP du
r dr = μd r
dx dr
dP 1 2 du
r = μr + C1
dx 2 dr
at r = 0, du/dr = 0, C1 = 0
dP 1
r dr = μ du
dx 2
dP 1 2
r = μu + C2
dx 4
at r = R, u = 0, C2 = (1/4)R2 (dP/dx)
1 2 dP 1 dP
r = μu + R2
4 dx 4 dx
1 dP 2
u= r − R2
4μ dx
u r 2
=1− (8.8)
Umax R
Internal Forced Convection 171
R
1
V= u2πr dr (8.9)
πR2
0
R r 2
1
= 1− umax 2πr dr
πR2 R
0
2umax 1 2 1 4 R
= r − r
R2 2 4R2 0
1
V= Umax (8.10)
2
ΔPAc + τw πDΔx = 0
τw πDΔx
ΔP = −
(1/4)πD2
ΔP −4τw 4 1 2
= = − f ρU (8.13)
Δx D D 2
4 1 2 16 32μU 32μ ρ D μ 32μ2 ρDU
=− ρU =− = − U = −
D2 ReD D2 D2 ρ D μ D3 ρ μ
32μ2
=− ReD ∼ ReD (8.14)
D3 ρ
And pumping power can be obtained as P ∼ = ΔP (volume flow). Figure 8.4
shows that the friction factor decreases and the pressure drop increases with
Reynolds number.
172 Analytical Heat Transfer
1U
Xf,h V=
2 max
Umax
τwπDΔx
PAc (P + ΔP)Ac
∫ ρu · u · d Ac ∫ ρu · u · dAc
2πr dr 2πr dr
Δx
FIGURE 8.3
Force balance in fully developed flow region.
f ΔP
Δx
0 ReD
FIGURE 8.4
Friction factor and pressure drop versus Reynolds number in fully developed flow region.
Internal Forced Convection 173
(a) qw = constant
Ti
Tw
(b)
q ″w
Temp Tw
T(r)
Tb
Ti
x
(c)
15
hD
k
hD
4.3 = Nu =
k
x
FIGURE 8.5
Laminar flow in a circular tube with uniform surface heat flux condition. (a) Thermal boundary
layer; (b) temperature; and (c) heat transfer coefficient.
qw
qw = = h(Tw − Tb ) (8.15)
As
qw
qw = = h(Tw − Tb ) (8.16)
+,-. As + ,- .
const. const.
∂ T − Tw
=0 (8.20)
∂x Tb − T w
T − Tw
= f (r) = f (x)
Tb − T w
∂T ∂Tw T − Tw ∂Tb ∂Tw
− − − =0
∂x ∂x Tb − Tw ∂x ∂x
+ ,- .
=0
∂T ∂Tw ∂Tb
= = = constant (8.21)
∂x ∂x ∂x
∂T 1 ∂ ∂T
ρCp u =k r = T(r) only (8.22)
∂x r ∂r ∂r
r 2
ρCp qw dT
r 2V 1 − dr = d r
k R ṁC /πD dr
+ ,- . + p,- .
u ∂T
∂x
∂T
r = 0, T = Tc or =0
∂r
r = R, T = Tw
ρCp dT r2 r4
T − Tc = U −
k dx max 4 16R2
ρCp ∂T R2 R4 ρCp ∂T 3
Tw − Tc = Umax − = Umax R2
k ∂x 4 16R2 k ∂x 16
where
R
ρuT · 2πr dr 7 ρCp dT
Tb = 0R = Tc + Umax R2
ρu · 2πr dr 96 k dx
0
ρCp dT 3 2 7 2
Tw − T b = Umax R − R (8.23)
k dx 16 96
Internal Forced Convection 175
k 96 k
= =
(44/96)R 22 D
hD 96
NuD ≡ = = 4.314 (8.24)
k 22
∂T ∂Tb T − Tw
= (8.25)
∂x ∂x Tb − Tw
hD
NuD = = 3.66 (8.27)
k
Examples
8.1. Internal flow, fully developed laminar forced convection: Consider a low-
speed, constant-property, fully developed laminar flow between two parallel
plates at y = ±H, as shown in Figure 8.6. The plates are electrically heated
to give a uniform wall heat flux. Determine the velocity profile, the friction
factor, and the Nusselt number.
Assumptions:
Low speed ⇒ Φ = 0
Constant properties:
Fully developed ⇒ du/dx = 0
Thermally fully developed ⇒ dT /dx = const
176 Analytical Heat Transfer
q″s
x 2H
q″s
FIGURE 8.6
Two parallel plates at a uniform wall heat flux.
a. Velocity profile
Continuity equation:
∂u ∂υ ∂u
+ =0 ⇒ =0
∂x ∂y ∂x
Momentum equation:
∂u ∂u 1 ∂P ∂ 2u ∂ 2u
u +υ =− +ν + 2
∂x ∂y ρ ∂x ∂x 2 ∂y
υ=0
∂u ∂ 2u
=0 ⇒ =0
∂x ∂x2
1 ∂P ∂ 2u
= ν 2 ⇒ Governing equation (8.28)
ρ ∂x ∂y
with solution
1 ∂P 2
u= y + c1 y + c 2 (8.29)
2μ ∂x
where ν = μ/ρ
Boundary conditions:
At y = 0, ∂u/∂y = 0 (maximum velocity)
At y = H, u = 0
c1 = 0
1 ∂P 2
c2 = − H
2μ ∂x
Internal Forced Convection 177
H 2 ∂P y2
u=− 1− 2 (8.30)
2μ ∂x H
umax at y = 0
H 2 ∂P
umax = −
2μ ∂x
y2
u = umax 1 − 2 (8.31a)
H
y2
2 0H umax 1 − 2 dy
ρudA H
um = =
ρA 2H
H
umax y − (y 3 /3H 2 )
um = 0 = umax ((2/3)H)
H H
2
um = umax
3
Thus,
3 y2
u = um 1 − 2 (8.31b)
2 H
b. Friction factor:
− (∂P /∂x) Dh
f ≡ 2 /2
ρum
ΔPAc + τw (2w )Δx = 0
ΔP −2τw w −2τw w τw
= = =−
ΔX Ac 2wH H
H
∂u 3 2y um
τw = μ =μ um = 3μ
∂y y =H 2 H2 0 H
Therefore,
ΔP −3μum
=−
Δx H2
4Ac 4(2wH)
DH = = = 4H
P 2w
178 Analytical Heat Transfer
Thus,
But,
ρum DH 4u m Hρ
ReDh = =
μ μ
Finally,
96
f = (8.32)
ReDh
c. Energy equation:
∂T ∂T ∂ 2T ∂ 2T
ρCp u +υ =k + + q̇ + Φ
∂x ∂y ∂x 2 ∂y 2
dT d2 T
= const ⇒ =0
dx dx 2
∂ 2T u ∂T
= (8.33)
∂y 2 α ∂x
Boundary conditions:
y = 0, dT /dy = 0 → c1 = 0
y = b, T = Ts
5 um 2 dT
c2 = Ts − H
8 α dx
3 um 2 y 2 y4 5 dT
T = H − − + Ts (8.34)
2 α 2H 2 12H 4 12 dx
Internal Forced Convection 179
ρc v uT dA
Tm ≡
ρcv um A
/ 0 / 0
H 3 y2 3 um 2 y 2 y4 5 dT
2 ρcv um 1 − 2 H − − + T s dy
0 2 H 2 α 2H 2 12H 4 12 dx
Tm =
ρcv um (2H)
2 !
1 ! 1
9 um H 1 1 5 1 1 5 dT 3 Ts
Tm = − − − + + + Ts −
4 α 6 60 12 10 84 36 dx 2 3
17 um H 2 dT
Tm = − + Ts (8.35)
35 α dx
qs = h(Ts − Tm )
∂T
qs = − k
∂y y =H
k (3/2)(um /α) · (2/3)H (dT /dx)
h=
Ts + (17/35)(um /α)H 2 (dT /dx) − Ts
35 k
h=
17 H
h(4H) 35 k 4H
NuD = =
k 17 H k
140
NuD = = 8.235
17
Remarks
There are many engineering applications such as electronic equipments, mini-
scale channels, and compact heat exchangers that required laminar flow heat
transfer analysis and design. In the undergraduate-level heat transfer, stu-
dents are expected to know many heat transfer relations between Nusselt
numbers and Reynolds and Prandtl numbers for developing and fully devel-
oped flows inside circular tubes at various surface thermal BCs. Students are
expected to calculate heat transfer coefficients from these relations by giving
Reynolds and Prandtl numbers.
In the intermediate-level heat transfer, this chapter focuses on how to solve
fully developed heat transfer problems for flow between two parallel plates or
inside circular tubes at uniform surface heat flux BCs. Students are expected
to know how to analytically determine the velocity profile, the friction factor,
the temperature profile, and the Nusselt number for these cases. Here we
do not include how to analytically determine the heat transfer coefficient at
uniform surface temperature BCs.
In advanced heat convection, students will learn how to analytically predict
heat transfer in both developing flow and thermal entrance regions; with
180 Analytical Heat Transfer
various thermal BCs such as variable surface heat flux as well as variable
surface temperature BCs; for flow in rectangular channels with various aspect
ratios and flow in annulus with various thermal BCs. They require more
complex mathematics and are beyond the intermediate-level heat transfer.
PROBLEMS
8.1. Consider a steady constant-property laminar flow between two
parallel plates at y = ±. The plates are electrically heated to give a
uniform wall heat flux. The differential equations for momentum
and energy are listed here for reference:
∂u ∂u 1 ∂P ∂ 2u ∂ 2u
u +v =− +ν +
∂x ∂y ρ ∂x ∂ x2 ∂ y2
∂T ∂T ∂ 2T ∂ 2T ν ∂u 2
u +v =α + +
∂x ∂y ∂ x2 ∂ y2 cp ∂ y
a. Assume
a low-speed, linear velocity profile (i.e., u = um
1 − y/ with maximum velocity at y = 0, and zero velocity
at y ± ) between two parallel plates, and also assume a ther-
mally, fully developed condition, and write down the simpli-
fied equations for momentum and energy and the associated
BCs that can be used for this problem.
b. Under the assumption in (a), determine the Nusselt number
on the plate.
c. Consider a fully developed velocity profile (i.e., a parabolic
velocity profile) between two parallel plates and a thermally,
fully developed condition, and comment on whether the Nus-
selt number on the plate will be higher, the same, or lower than
those of symmetry linear velocity profile in (a)? Explain why.
8.7. Consider an incompressible laminar 2-D flow in a parallel plate
channel as shown below. The top plate is pulled at a constant
velocity UT . The top and bottom plates are maintained at constant
heat flux qs . Flow is both hydrodynamically and thermally fully
developed. Assume that the pressure gradient is zero in a parallel
plate channel.
a. Obtain differential equations governing the velocity U(Y) and
temperature T(Y) fields.
b. Use appropriate BCs to evaluate U(Y). Obtain T(Y). Do not
attempt to evaluate constants of integration for the tempera-
ture field.
8.8. Find the Nusselt number for the following problems.
a. Fully developed Couette flow (i.e., assume that velocity and
temperature profiles do not change along the channel) with
the lower plane wall at uniform wall temperature T0 and the
upper plane wall at T1 . If the velocity profile is a linear profile
(U = 0 at the lower plane wall, U = V at the upper plane wall),
find the temperature profile from the energy equation.
b. Fully developed Poiseuille flow (i.e., assume that velocity and
temperature profiles do not change along the channel) with
Internal Forced Convection 183
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
3. W. M. Kays and M. E. Crawford, Convective Heat and Mass Transfer, Second
Edition, McGraw-Hill, New York, NY, 1980.
4. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
9
Natural Convection
gβ(Tw − T∞ )x3
Grx = (9.1)
ν2
Rax = Grx · Pr (9.2)
Continuity
∂u ∂v
+ =0 (9.3)
∂x ∂y
185
186 Analytical Heat Transfer
ρw, Tw = C
T∞, ρ∞
τw 0
dP
dx
g
x
FIGURE 9.1
Natural convection boundary layer from a heated vertical wall.
Momentum
∂u ∂u ∂ 2u
u +v = gβ(T − T∞ ) + ν 2 (9.4)
∂x ∂y ∂y
Energy
∂T ∂T ∂ 2T
u +v =α 2 (9.5)
∂x ∂y ∂y
The following shows how to derive the above natural convection momen-
tum equation from the original momentum equation:
∂u ∂u 1 ∂P ∂ 2u
X-momentum u +v =− −g+ν 2
∂x ∂y ρ ∂x ∂y
From the above momentum equation, one can see that natural convection is
due to temperature difference between the surface and fluid and the gravity
force. This implies that there is no natural convection if there exists no temper-
ature gradient or no gravity force. The larger delta T and gravity (means larger
Grashof number) will cause larger natural circulation and results in thinner
boundary-layer thickness and higher friction (shear) and higher heat transfer
coefficient. The Grashof number in natural convection plays a similar role
as Reynolds number does in forced convection; the larger Grashof number
causes higher heat transfer in natural convection as the greater Reynolds num-
ber has higher heat transfer in forced convection. Prandtl number plays the
same role in both natural and forced convection, basically the fluid property.
Just like in forced convection, both similarity and integral methods can be
used to solve natural convection boundary-layer equations. The following
only outline the similarity method from Ostrach in 1953.
Similarity variable:
' 1/4
4 gβ(Tw − T∞ ) y Grx
η=y 2
= (9.6)
4ν x x 4
where
gβ(Tw − T∞ )x3
Grx =
ν2
Ψ(x, η)
f (η) = (9.7)
4ν (Grx /4)1/4
T − T∞
θ= (9.8)
Tw − T ∞
Put them into the above momentum equation and energy equation,
respectively:
∂Ψ
u= = ···
∂y
∂Ψ
v=− = ···
∂x
∂T
= ···
∂x
∂T
= ···
∂y
188 Analytical Heat Transfer
0.28 Pr = 0.73
u
f ′=
1
2v 1.0
Gx 2
x
0.12
10
100
1000
0
0 1 2 3 4
1
y Gx 4
η=
x 4
T−T∞
θ=
Tw−T∞ Pr = 0.73
1.0
10
100
1000
0
1 2 3 4
1
y Gx 4
η=
x 4
FIGURE 9.2
Dimensionless velocity and temperature profiles from heated vertical wall.
Natural Convection 189
temperature profiles from the heated vertical wall, for different Prandtl
fluids.
u T∞
f = √
2 gx Tw − T ∞
ux −1/2
= Gx
2ν
u
= (9.13)
1/2
Gx (2ν/x)
∂T k Grx 1/4 dθ
qw = −k = − (Tw − T∞ ) (9.14)
∂y 0 x 4 dη y=0
where
dθ
= θ (0) = f (Pr)
dη y=0
qw −k(∂T/∂y)0
∴h= =
Tw − T ∞ Tw − T ∞
1/4
hx Grx dθ
Nu = =− (9.15)
k 4 dη y=0
hx 1/4
Nux = = 0.359Grx (9.16)
k
hx L 4 1/4
Nux = = NuL = 0.478GrL (9.17)
k 3
190 Analytical Heat Transfer
where
L 1/4
1 4k GrL dθ 4
hx = hx dx = − = hL
L 3L 4 dη 0 3
0
gβ(Tw − T∞ )L3
GrL = (9.18)
ν2
1/4
3 2Pr
Nux = (Grx Pr)1/4 (9.19)
4 5 1 + 2 Pr1/2 + 2 Pr
In general,
gβ(Tw − T∞ )x3
Grx Pr = Rax =
να
Note: The following is a simple guideline whether the problem can be solved
by forced convection, natural convection, or mixed (combined forced and
natural) convection.
If Grx /Rex2 < 1, the problem can be treated as forced convection.
If Grx /Re2 ∼
x = 1, the problem can be treated as mixed convection.
If Grx /Rex2 > 1, the problem can be treated as natural convection.
Tw, ρw
T
q″w
T∞, ρ∞
∂
∫ ρuu dy + ∂x ∫ ρuu dy dx
∂
∫ Pdy + ∫ P dy dx
∂x
dP∞
dx τw
U∞ = 0
∫ Pdy
∫ ρuu dy
g
x
y
FIGURE 9.3
Integral method.
∂P ∂
−τw − ρg dy − dy = ρuu dy
∂x ∂x
δ δ
τw μ ∂u d
= = gβ (T − T∞ ) dy − u2 dy (9.22)
ρ∞ ρ∞ ∂y 0 dx
0 0
δT
∂T d
q = −k = ρcp u (T − T∞ ) dy (9.23)
∂y 0 dx
0
u (x, δ) = 0, T (x, δT ) = T∞
∂u (x, δ) ∂T (x, δT )
= 0, =0
∂y ∂y
192 Analytical Heat Transfer
1 1
m= ; n=
2 4
Put this into momentum and energy integral equations:
1/2 1/2
80 Gx ν
u1 (x) =
3 (20/21) + Pr x
1/4
δ (x) 240 (1 + (20/21Pr)) 1/4 x
= → δ (x) ∼
x Pr · Gx Tw − T ∞
∂T 2k (Tw − T∞ )
qw = −k = = h (Tw − T∞ )
∂y 0 δ (x)
1/4
hx x 2x Pr 1/4
Nux = = = · Rax (9.26)
k δ (x) 15 ((20/21) + Pr)
Nux ∼
1/4
= 0.413Rax for Pr = 0.733 (9.27)
1/4
Note: Nux = 0.359Rax for Pr = 0.733 by using the exact similarity solu-
tion.
gβ (Tw − T∞ ) x3
Rax = < 108 − 109 − Laminar natural convection
να
Remarks
In the undergraduate-level heat transfer, we have heat transfer correlations
of external natural convection for a vertical plate, an inclined plate, a hor-
izontal plate, a vertical tube, and a horizontal tube as well as heat transfer
correlations of internal natural convection for a horizontal tube, between two
parallel plates, and inside a rectangular cavity with various aspect ratios.
These correlations are important for many real-life engineering applications
such as electronic components.
In the intermediate-level heat transfer, this chapter focuses on how to
analytically solve the external natural convection from a vertical plate at
Natural Convection 193
PROBLEMS
9.1. Consider the system of boundary-layer equations
∂u ∂v
+ =0
∂x ∂y
∂u ∂v ∂ 2u
u +v = R (T − T∞ ) + ν 2
∂x ∂y ∂y
∂ (T − T∞ ) ∂ (T − T∞ ) ∂ 2 (T − T∞ )
u +v =α
∂x ∂y ∂ y2
y=0 u = 0, v = 0, qw = constant
y=∞ u=0 T = T∞
hx x
= 0.443 (Grx Pr)1/4
k
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
3. W. M. Kays and M. E. Crawford, Convective Heat and Mass Transfer, Second Edition,
McGraw-Hill, New York, NY, 1980.
4. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
10
Turbulent Flow Heat Transfer
and
t
1
u = lim u (t) dt ∼
=0
t→∞ t
0
195
196 Analytical Heat Transfer
U∞,T∞
δ
v'
−u' u'
y −v'
x Tw δL
u>v
u' ≅ −v'
FIGURE 10.1
Typical turbulent boundary-layer flow velocity and temperature profile.
FIGURE 10.2
Instantaneous time-dependent velocity and temperature profiles inside a turbulent boundary
layer.
Turbulent Flow Heat Transfer 197
u = u + u ≡ u + u
v = v + v ≡ v + v
T = T + T ≡ T + T
P = P + P ≡ P + P
ρ = ρ + ρ ≡ ρ + ρ
Since
t+Δt
u dt = 0
t
t+Δt
v dt = 0
t
t+Δt
P dt = 0
t
t+Δt
T dt = 0
t
Therefore,
v = T = ρ = P ∼
=0
Therefore,
∂ ∂
(ρu) + (ρv) = 0 (10.2)
∂x ∂y
and
∂ ∂
ρu + ρv =0 (10.3)
∂x ∂y
Therefore,
∂ ∂ ∂ ∂ ∂ 2 ∂ ∂
ρu2 + ρuv + ρu2 + ρu v + ρu + ρuv + 2uρ u
∂x ∂y ∂x ∂y ∂x ∂y ∂x
∂
∂P
∂ 2u ∂ 2u
+ uρ v + vρ u = − + ρfx + ρ fx + μ + 2
∂y ∂x ∂x2 ∂y
Rearranging,
∂u ∂u ∂P ∂ 2u ∂ 2u
ρu + ρ u + ρv + ρ v =− + ρfx + ρ fx + μ + 2
∂x ∂y ∂x ∂x2 ∂y
(10.5)
⎫
∂ ⎪
+ ρu2 + ρ u2 + uρ u ⎪ ⎬
∂x
6 Reynolds Stress
ρu v + ρ u v + vρ u ⎪
∂
+ ⎪
⎭
∂y
Y-Momentum:
∂v ∂v ∂v 1 ∂P μ ∂ 2v ∂ 2v ∂u v ∂v2
u +v + =− + + 2 − − + fy
∂x ∂y ∂t ρ ∂y ρ ∂x2 ∂y ∂x ∂y
+ ,- . + ,- .
viscous forces Reynolds stress, Turbulent stress
(10.7)
From boundary-layer approximation, Y-Momentum equation is not impor-
tant as compared with X-Momentum equation.
Therefore,
∂T ∂T
cp (ρu + ρ u ) + ρv + ρ v
∂x ∂y
∂ ∂T ∂ ∂T ∂P ∂P
= k + k +u +v + Φ + ρε (10.9)
∂x ∂x ∂y ∂y ∂x ∂y
∂ ∂u 2
+ cp ρu T + ρ u T + uρ T + + ...
∂x ∂x
2
∂
∂u
+ cp ρv T + ρ v T + vρ T + + ...
∂y + ,- . ∂x
6 Reynolds Flux
∂u ∂u 1 ∂ ∂u
1 ∂
u +v = μ − ρu v = (τviscous + τturb. ) (10.11)
∂x ∂y ρ ∂y ∂y ρ ∂y
∂T ∂T 1 ∂ ∂T 1 ∂
u +v = k − ρCp v T = qmolecular + qturb.
∂x ∂y ρCp ∂y ∂y ρCp ∂y
(10.12)
∂u ∂u
τtotal = τm + τt = μ + ρεm (10.13)
∂y ∂y
∂T ∂T
qtotal = qm + qt = k + ρCp εH (10.14)
∂y ∂y
∂u
τtotal = ρ (ν + εm ) (10.15)
∂y
∂T
qtotal = ρCp (α + εH ) (10.16)
∂y
υt εm
Prt = = ∼1 (10.17)
αt εH
• Turbulent viscosity μt
• Turbulent diffusivity υt = (μt /ρ) = εm
• Turbulent diffusivity for heat εH = αt
dP 1 ∂
= (rτ)
dx r ∂r
dP d
r = (rτ)
dx dr
dP d
r dr = (rτ) dr
dx dr
Therefore,
1 dP
τ= r ∼r
2 dx
From BCs,
r = 0, τ = 0,
r = R, τ = τw
One obtains
r
τw τ= (10.18)
R
r R−y y ∂u
τ = τw = τw = 1 − τw = ρ(ν + νt )
R R R ∂y
∂u νt y τw
ν 1+ = 1−
∂y ν R ρ
y+ νt ∂u ν νt du+
1− + = 1+ √ √ = 1+
R ν ∂y (τw /ρ) τw /ρ ν dy+
where
'
u yu∗ Ru∗ τw
u+ = , y+ = , R+ = , u∗ =
u∗ v v ρ
ΔP
y
r P1 P2
R
τw
FIGURE 10.3
Force balance in a circular tube.
Turbulent Flow Heat Transfer 203
U∞,T∞
δ
Fully
turbulent
Turbulent δL
flow Very small
FIGURE 10.4
Concept of 2-D turbulent boundary layer flow.
Therefore,
vt εm 1 − (y+ /R+ )
= = −1 (10.19)
v ν (du+ /dy+ )
vt εm 1 − (y+/δ+ )
= = −1 (10.20)
v ν (du+/dy+ )
where
δu∗
δ+ =
v
τw 1 Δu
= (10.22)
qw Cp ΔT
qw τw
=
Cp (Tw − T∞ ) U∞
204 Analytical Heat Transfer
where
1 2
τw = ρU · Cf
2 ∞
1 ρU∞ x μCp
Nu = Cf ·
2 μ k
Nu 1 1
= Cf ⇒ Cf = St (10.23)
Re · Pr 2 2
1
Cf = St · Pr2/3 (10.24)
2
Nu (hx/k) h qw
St = = = =
Re · Pr (ρU∞ x/μ) · (μCp /k) ρCp U∞ ρCp U∞ (Tw − T∞ )
(10.25)
There are two kinds of problems:
0.046
Cf = 0.2
(10.26)
ReD
1 0.046 Nu
= Pr2/3
2 ReD0.2 Re · Pr
hD 0.8 1/3
NuD = = 0.023 ReD Pr for cooling (10.27)
k
hD 0.8 0.4
NuD = = 0.023 ReD Pr for heating (10.28)
k
0.0592
Cf = (10.29)
Rex0.2
1 0.0592 Nux
= Pr2/3
2 Rex0.2 Rex Pr
hx
Nux = = 0.0296 Rex0.8 Pr1/3 (10.30)
k
Turbulent Flow Heat Transfer 205
u
u+ = (10.31)
u∗
yu∗
y+ = (10.32)
ν
' '
∗ τw (1/2)Cf ρU∞
2 1
u = = = U∞ Cf (10.33)
ρ ρ 2
−0.2
Cf = 0.046 ReD , for turbulent flow in a tube
Cf = 0.0592 Rex−0.2 , for the turbulent flow over a flat plate
1
u
f' =
U∞ Blausius
Laminar flow
0
y 5
η= Re x
x
Laminar buffer
u sublayer zone
u+ = Turbulent core
u*
20 Turbulent wake
(core flow)
5 30 500 1000
yu*
ln y + = ν
FIGURE 10.5
Analytical universal velocity profile for laminar boundary layer and semiempirical law of wall
velocity profile for turbulent boundary layer.
Consider a laminar sublayer region, very close to the wall region where
viscosity is dominated and turbulence is damped at the wall.
du
τ = ρν (10.34)
dy
τw du u
=ν ≈ν
ρ dy w y
' '
τw τw du
=ν
ρ ρ dy
du
u∗ u∗ = ν
dy
Therefore,
u yu∗
=
u∗ ν
Turbulent Flow Heat Transfer 207
and
u+ = y + (10.35)
v'
−u' u'
−vv''
FIGURE 10.6
Concept of turbulence in 2-D turbulent boundary-layer flow.
208 Analytical Heat Transfer
1
u+ = ln y+ + C (10.40)
κ
Or
u+ = 2.44 ln y+ + 5.5 (10.42)
Then, consider the buffer zone between the laminar sublayer and the
turbulence region, 5 ≤ y+ ≤ 30, the velocity profile can be obtained as
u+ = 5 ln y+ − 3.05 (10.43)
Δx
R
y
x
ΔTb
q"w
FIGURE 10.7
Turbulent flow heat transfer in a circular tube.
Assume
u=V
∂T ∂Tb
=
∂x ∂x
1 dTb dT C
(R − y)V = (α + εH ) +
2 dx −dy R − y
where C = 0 at R − y = 0, (dT/dy) = 0.
Therefore,
1 dTb y − R
dT = V dy
2 dx α + εH
T y
1 dTb y−R
dT = V dy
2 dx α + εH
Tw 0
y y+
1 ∂Tb y−R q 1 − (y+ /R+ )
T − Tw = V dy = √w dy+
2 ∂x α + εH ρCp τw /ρ (1/Pr) + (εH /ν)
0 0
Therefore,
y+
Tw − T 1 − (y+ /R+ )
T + ≡ = dy+ (10.45)
(qw /ρCp u∗ ) (1/Pr) + (εH /ν)
0
εm 1 − (y+ /R+ )
= −1
υ (du+ /dy+ )
du+ εm
0 ≤ y+ ≤ 5, y + R+ , u+ = y+ , = 1, =0
dy+ ν
du+ 5
5 ≤ y+ ≤ 30, u+ = 5 ln y+ − 3.05, = +
dy+ y
du+ 2.5
y+ ≥ 30, u+ = 2.5 ln y+ + 5.0, = +
dy+ y
Turbulent Flow Heat Transfer 211
The above energy integral equation 10.45 can be obtained in the following
three-regions and the Prandtl number effect is shown in Figure 10.8.
y+
+ 1 − (y+ /R+ )
T = dy+
(1/Pr) + (εH /ν)
0
εH εm
=
ν ν
For the region 0 ≤ y+ ≤ 5 with y+ ≈ 0 and (εm /ν) ≈ 0,
y+
T + = Pr dy+ = Pr y+
0
+ +
u =y , T + = Pr y+ (10.46)
εm εH y+
= = −1
ν ν 5
y+
1
T + − T5+ = dy+
(1/Pr) + (y+ /5) − 1
5
y+
1 1 y+
=5 d + − 1
(1/Pr) + (y+ /5) − 1 Pr 5
5
Pr
u
= u+
u* 20
(Tw −T )
= T+ If Pr = 1,
⎛ "
qw ⎞ u+ ≅ T +
⎝ ρCpu* ⎠ 5
5 30 500 1000
yu*
ln y+ =
ν
FIGURE 10.8
Law of wall temperature profile for turbulent boundary layer.
212 Analytical Heat Transfer
Therefore,
1 y+ 1 5
T + − T5+ = 5 ln + − 1 − 5 ln + −1
Pr 5 Pr 5
(1/Pr) + (y+ /5) − 1
= 5 ln
(1/Pr)
Pr y+
= 5 ln 1 + − Pr (10.47)
5
εH εm 1 − (y+ /R+ )
= = −1
ν ν (2.5/y+ )
{[ ] }−1
Therefore,
+
T + − T30 = 2.5 ln y+ − 2.5 ln 30 (10.48)
The next question is how to determine heat transfer coefficient from the law
of wall for temperature profile. This is shown in the following.
One assumes the simple velocity and temperature profiles as
u y 1/7 r 1/7
∼
= = 1−
Umax R R
T − Tw ∼ y 1/7 r 1/7
= = 1−
Tc − T w R R
Therefore,
∫ u2πr dr
ub or V= = 0.82Umax
∫ 2πr dr
∫R
0 (Tw − Tc )(1 − (r/R))
1/7 U
max (1 − (r/R))
1/7 · r dr
Tw − T b =
∫R
0 Umax (1 − (r/R))
1/7 · r dr
∼ 15
= (Tw − Tc )
18
= 0.833(Tw − Tc )
Turbulent Flow Heat Transfer 213
where
Therefore,
qw qw
h= =
Tw − T b 0.833(q /ρCp u∗ )[5Pr + 5 ln(5Pr + 1) + 2.5 ln(Ru∗ /v30)]
The final heat transfer coefficient and the Nusselt number are expressed as
√
hD Re · Pr Cf /2
NuD ≡ = √ (10.49)
k 0.833 5Pr + 5 ln(5Pr + 1) + 2.5 ln (ReD Cf /2)/60
where
0.046 0.079
Cf = 0.2
or Cf = 0.25
ReD ReD
For given ReD , and Pr, the above prediction is fairly close to the following
experimental correlation:
∂T ∂Tw
∼ ≈0
∂x ∂x
y+
1
T+ = dy+
(1/Pr) + (εH /ν)
0
214 Analytical Heat Transfer
Follow a similar procedure as in the previous case; one can obtain three-region
temperature profile.
∂T −qw
(α + εH ) =
∂y ρCp
∂T − qw /ρCp
=
∂y α + εH
y
− qw 1 u∗
T − Tw = dy ∗
ρCp υ((1/Pr) + (εH /υ)) u
0
y
qw 1
Tw − T = dy+ (10.51)
ρCp u∗ ((1/Pr) + (εH /υ))
0
where εH
εm .
And from Equation 10.20,
εm 1 − (y+/δ+ )
= −1
υ (du+/dy+ )
0 < y+ < 5, u+ = y+
T + = Pr y+ (10.52)
5 < y+ ≤ 30,u+ = 5 ln y+ − 3.05
+ + Pr y+
T − T5 = 5 ln 1 + − Pr (10.53)
5
30 ≤ y+ , u+ = 2.5 ln y+ + 5.0
+
T + − T30 = 2.5 ln y+ − 2.5 ln 30 (10.54)
Turbulent Flow Heat Transfer 215
where
u∞ 1
u+
30 = 14, u+
∞ = ∗
=√
u Cf /2
Therefore,
The final heat transfer coefficient and the Stantan number can be obtained as
1/5
where Nux = (hx/k), Rex = (ρU∞ x/μ), Cf = 0.0592/Rex .
For given Rex and Pr, the above predict Nux value is very close to the
following experimental correlation:
Remarks
In undergraduate heat transfer, students are expected to know how to
calculate heat transfer coefficients (Nusselt numbers) for turbulent flows over
a flat plate at uniform surface temperature and inside a circular tube at uni-
form surface heat flux, by using heat transfer correlations from experiments,
that is, Nusselt numbers relate to Reynolds numbers and Prandtl numbers.
There are many engineering applications involving turbulent flow conditions.
These turbulent flow heat transfer correlations are very useful for basic heat
transfer calculations such as for heat exchangers design.
In intermediate-level heat transfer, this chapter focuses on how to derive
RANS equation; introduce the concept of turbulent viscosity and turbulent
Prandtl number; Reynolds analogy; Prandtl mixing length theory; law of
wall for velocity and temperature profiles; and turbulent flow heat transfer
coefficients derived from law of wall velocity and temperature profiles and
216 Analytical Heat Transfer
PROBLEMS
10.1. Consider a steady low-speed, constant-property, fully turbu-
lent boundary-layer flow over a flat surface at constant wall
temperature. Based on the Reynolds time-averaged concept,
the following momentum and energy equations are listed for
reference:
∂u ∂u ∂ ∂u
u +v = (ν + εM )
∂x ∂y ∂y ∂y
∂T ∂T ∂ ν ε ∂T
u +v = + M
∂x ∂y ∂y Pr Prt ∂y
√
τw /ρ, friction temperature T ∗ = 3◦ C ∼
= qw /(ρcp u∗ ), pipe wall
temperature Tw =100◦ C, air flow Prandtl Pr = 0.7
10.3. Consider the Von Kaŕman–Martinelli heat–momentum analogy
for a turbulent pipe flow:
a. If a two-layer universal velocity profile will be employed,
that is,
u R − r 1/7
=
umax R
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. F. Incropera, and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
3. W.M. Kays and M.E. Crawford, Convective Heat and Mass Transfer, Second Edition,
McGraw-Hill, New York, NY, 1980.
4. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
5. H. Schlichting, Boundary-Layer Theory, Sixth Edition, McGraw-Hill, New York,
NY, 1968.
6. E. Levy, Convection Heat Transfer, Class Notes, Lehigh University, 1973.
11
Fundamental Radiation
dq
Iλ (θ, ϕ, λ, T) = (11.1)
dAn dω
Visible
Yellow
Green
Violet
Blue
Red
X-rays Infrared
Ultraviolet
Microwave
Thermal radiation
Gamma
rays
0.4 0.7
FIGURE 11.1
Conceptual view of hemispheric radiation.
and
dAn = dA cos θ
Wavelength
FIGURE 11.2
(a) Spectral radiation varies with wavelength. (b) Directional distribution.
Fundamental Radiation 223
(a) n (b) r
dAn = r2 sin θ dθ dφ dl
r dθ
r sin
θ dα = dl/r
r r sin θ dφ (c)
θ r
dθ
dA dAn
dφ dω = dAn/r2
FIGURE 11.3
(a) Conceptual view of hemispheric radiation from a differential element area. (b) Definition of
plane angle. (c) Definition of solid angle.
∞
π/2
Eλ = 2π Iλ (θ, λ, T) sin θ cos θ dθ
0
For the isotropic and black surface (independent of vertical direction angle),
π/2
Eb,λ = 2πIb,λ (λ, T) sin θ cos θ dθ = πIb,λ (11.4)
0
Eb = πIb = σT 4 (11.5)
224 Analytical Heat Transfer
E = πI = εσT 4 (11.6)
E(θ, ϕ, λ, T)
ελ,θ (θ, ϕ, λ, T) = (11.7)
Eb (λ, T)
E(T)
ε(T) = =0∼1 (11.8)
Eb (T)
εθ
ελ Non conductor
ε 0.9 ε
Conductor
0.1 ε
λ θ
0 45° 90°
Gray surface ε ≅ ελ = const Diffuse surface ε ≅ εθ = const
FIGURE 11.4
Definition of gray surface and diffuse surface.
Fundamental Radiation 225
G, irradiation ρG irradiation
αG absorption
τG transmission
FIGURE 11.5
Radiation energy balance on a surface.
about zero (except window glasses); therefore, reflectivity can be found from
emissivity too.
Absorptivity α = Ga /G
Reflectivity ρ = Gρ /G
Transmissivity τ = Gτ /G
From radiation energy balance, α + ρ + τ = 1
If τ = 0, and assume α = ε, then ρ = 1 − α ≈ 1 − ε
Blackbody radiation is the maximum radiation from an ideal surface.
Blackbody is a diffuse surface and can emit the maximum radiation and can
absorb the maximum radiation (i.e., α = 1and ε = 1). Therefore, blackbody
radiation intensity is not a function of direction ( = (θ, ϕ)), but a function of
wavelength and temperature (= (λ, T)). Planck obtained blackbody radiation
intensity from quantum theory as
2hC02
Iλ,b (λ, T) = (11.9)
λ5 exp (hC0 /λkT) − 1
C1 λ−5
Eλ,b = πIλ,b =
e 2 /λT) − 1
(C
Therefore, Planck emissive power for the black surface can be shown as
2hC02 C1
Eλ,b (λ, T) = πIλ,b (λ, T) = π = 5
λ [exp (hC0 /λkT) − 1]
5 λ [exp(C2 /λT) − 1]
226 Analytical Heat Transfer
∞
∞
However, for a real surface the emissive power is lower than Planck’s black-
body radiation (because the emissivity for the real surface is less than unity).
Therefore, the emissive power for the real surface is
E = εσT 4 (11.11)
Figure 11.6 shows emissive power versus wavelength over a wide range of
temperatures [1–4]. In general, emissive power increases with absolute tem-
perature; emissive power for the black surface (solid lines) is greater than that
for the real gray surface (lower than the solid lines, depending on emissivity)
for a given temperature.
108
lmaxT=2898 mm K
Spectral emissive power, El,b (W/m2 mm)
1000 K
102 800 K
100
300 K
10–2
10–4
10–1 100 101 102
Wavelength, l (mm)
FIGURE 11.6
Spectral blackbody emissive power.
Fundamental Radiation 227
ε(λ)
Ts = 500 K 0.8 ε3
Brick ε2
wall 0.5
Tc = 2000 K
0.1 ε1
0 1.5 10 λ, μm
λ1 λ2
Coal bed
FIGURE 11.7
Radiation between hot coal bed and cold brick wall with nongray.
From Figure 11.6 we see that the blackbody emissive power distribution has
a maximum and that the corresponding wavelength λmax depends on tem-
perature. Taking a derivation on Equation 11.9 with respect to λ and setting
the result as equal to zero, we obtain Wien’s displacement law as
The focus of Wien’s displacement law is also shown in Figure 11.5. According
to this result, the maximum emissive power is displaced to shorter wave-
lengths with increasing temperature. For example, the maximum emission
is in the middle of the visible spectrum (λmax ≈ 0.5 μm) for solar radiation
at 5800 K; the peak emission occurs at λmax = 1 μm for a tungsten filament
lamp operating at 2900 K emitting white light, although most of the emission
remains in the IR region.
There are many engineering surfaces with diffuse but not gray behaviors. In
this case, surface emissivity is a function of wavelength and is not the same as
absorptivity. Figure 11.7 shows the radiation problem between a hot coal bed
and a cold brick wall with an emissivity function of wavelength [4]. To deter-
mine emissive power from the cold brick wall, one needs to determine the
average emissivity from the brick wall first. The following outlines a method
to determine average emissivity and absorptivity.
The following shows how to determine ε(Ts ), E(Ts ), and α(Ts ). Average
emissivity can be determined by adding three regions of wavelength shown
in Figure 11.7. Then treat each region as a product of constant emissivity and
fraction of blackbody emissive power to total blackbody emissive power as
shown in Figure 11.8. The fraction value is a function of wavelength and
temperature, and can be obtained from integration in each region (e.g., see
Figure 11.9 or Table 11.1) [4].
∞
0 ε(λ)Eb dλ
ε(Ts ) = (11.13)
Eb
λ1 λ2 λ3
0 Eb dλ λ1 Eb dλ λ2 Eb dλ
= ε1 + ε2 + ε1 (11.14)
Eb Eb Eb
228 Analytical Heat Transfer
E(λ,b) λ
(λT ) ∫Eλ,b dλ
0
0 λ λ1 λ2 λ3
FIGURE 11.8
Concept of fraction method from a blackbody.
where
λ
Eλ,b dλ λ
λT
0 0 Eλ,b dλ Eλ,b d(λT)
F0−λ = = = = f (λT) (11.16)
∞ σT 4 σT 5
Eλ,b dλ 0
0
λ2 λ
0 Eλ,b dλ − 0 1 Eλ,b dλ
Fλ1 −λ2 = = F0−λ2 − F0−λ1 (11.17)
σT 4
From Table 11.1 or from Figure 11.9, F0−λ is a function of λT(μm K), with
T = Ts = 500 K, emission from the brick wall.
1.0
0.8
0.6
F(0→λ)
0.4
0.2
0
0 4 8 12 16 20
λT × 10−3 (μm K)
FIGURE 11.9
Fraction of the total blackbody emission in the spectral band from 0 to λ as a function of λT.
Fundamental Radiation 229
TABLE 11.1
Blackbody Radiation Functions
λT(μm · K) F(0→λ)
200 0.000000
400 0.000000
600 0.000000
800 0.000016
1000 0.000321
1200 0.002134
1400 0.007790
1600 0.019718
1800 0.039341
2000 0.066728
2200 0.100888
2400 0.140256
2600 0.183120
2800 0.227897
2898 0.250108
3000 0.273232
3200 0.318102
3400 0.361735
3600 0.403607
3800 0.443382
4000 0.480877
4200 0.516014
4400 0.548796
4600 0.579280
4800 0.607559
5000 0.633747
5200 0.658970
5400 0.680360
5600 0.701046
5800 0.720158
6000 0.737818
6200 0.754140
6400 0.769234
6600 0.783199
6800 0.796129
7000 0.808109
7200 0.819217
7400 0.829527
continued
230 Analytical Heat Transfer
7600 0.839102
7800 0.848005
8000 0.856288
8500 0.874608
9000 0.890029
9500 0.903085
10,000 0.914199
10,500 0.923710
11,000 0.931890
11,500 0.939959
12,000 0.945098
13,000 0.955139
14,000 0.962898
15,000 0.969981
16,000 0.973814
18,000 0.980860
20,000 0.985602
25,000 0.992215
30,000 0.995340
40,000 0.997967
50,000 0.998953
75,000 0.999713
100,000 0.999905
ε(Ts ) = 0.61
The brick wall is not a gray surface, and hence α(Ts ) = ε(Ts ). But it is a diffuse
surface, and hence α(λ) = ε(λ). The irradiation from the black coal bed (at
temperature Tc = 2000 K) to the brick wall is G(λ) ∝ Eb . The following shows
a similar way to determine brick wall absorptivity.
∞ ∞
0 α(λ)G(λ) dλ ε(λ)Eb dλ
α(Ts ) ≡ ∞ = 0 (11.18)
0 G(λ) dλ
Eb
= ε1 F0−λ1 + ε2 [F0−λ2 − F0−λ1 ] + ε3 [1 − F0−λ2 ] (11.19)
= 0.1 × 0.275 + 0.5 × (0.986 − 0.273) + 0.8 × (1 − 0.986)
where the fraction value can be found from Table 11.1, or from Figure 11.9,
with T = Tc = 2000 K, irradiation from the black coal bed.
Therefore, average absorptivity can be calculated as
2000
Solar spectrum outside
the Earth’s atmosphere
Monochromatic solar energy flux (W/m2 μm)
1600 Solar spectrum on ground
under a clear atmosphere
O3
1200
H2O
800 O2
O3
H2O, CO2
400 H2O, CO2
H2O
0
0.3 0.5 1.0 1.5 2.0 2.5
Wavelength (μm)
FIGURE 11.10
Solar spectra outside the Earth’s atmosphere and on the ground.
It is known that H2 O vapor and CO2 gas in the atmosphere not only absorb
solar radiation but also absorb radiation from the Earth’s surface at around
300 K that give radiation of wavelengths from 10 to 20 μm with an emissivity
about 1.0. In addition, H2 O vapor and CO2 gas in the atmosphere (sky) can
also emit energy at wavelengths of 5–10 μm at the effective sky temperature
around 250–270 K (assume an emissivity of about 0.8–1.0).
Figure 11.11 shows a typical setup for a solar collector. A special glass is
used as a cover for the collector and a specialized coating is used on the
collector plate and tubes where the solar energy is collected to maximize the
performance of the collector.
Solar energy
Glass
cover
FIGURE 11.11
A typical setup for a solar collector.
Fundamental Radiation 233
q '' Gs
radiation
window−sky
q ''
convection
outside air
q ''
radiation
window−walls
q ''
convection
room air
FIGURE 11.12
A typical design for a house with the skylight.
Figure 11.12 shows a typical design for a house with the skylight. The thin
glass of the skylight of a house has a specific spectral emissivity or absorptivity
distribution. For a given solar flux, atmospheric emission flux, interior surface
emission flux, inside and outside house convection conditions, the thin glass
temperature or the inside house temperature can be predicted.
Examples
11.1. A simple solar collector plate without the cover glass has a selective absorber
surface of high absorptivity α1 (for λ < 1 μm) and low absorptivity α2 (for
λ > 1 μm). Assume that solar irradiation flux = Gs , the effective sky tem-
perature =Tsky , the absorber surface temperature =Ts , and the ambient air
temperature =T∞ ; determine the useful heat removal flux (quseful ) from the
collector under these conditions. What is the correspondent efficiency (η)
of the collector?
SOLUTION
Performing an energy balance on the absorber plate per unit surface area,
we obtain
quseful
= αs Gs + αsky Gsky − qconv −E
q
η = useful
Gs
where αs = α1
αsky = α2
4
Gsky = σTsky
E = εσTs4
234 Analytical Heat Transfer
ε∼
= α2
qconv = h(Ts − T∞ )
11.2. A thin glass is used on the roof of a greenhouse. The glass is totally trans-
parent for λ < 1 μm, and opaque with an absorptivity α = 1 for λ > 1 μm.
Assume that solar flux = Gs , atmospheric emission flux = Gatm , thin glass
temperature = Tg , and interior surface emission flux=Gi , where Gatm and Gi
are concentrated in the far IR region (λ > 10 μm), and determine the tem-
perature of the greenhouse ambient air (i.e., inside room air temperature,
T∞,i ).
SOLUTION
Performing an energy balance on the thin glass plate per unit surface
area, and considering two convection processes (inside and outside green-
house), two emissions (inside and outside the glass plate), and three
absorbed irradiations (from solar, atmospheric, interior surface), we obtain
T∞,i from
= α1 F0−1 μm + α2 [1 − F0−1 μm
Remarks
This chapter covers the same topics as in the undergraduate-level heat trans-
fer. These include spectrum thermal radiation intensity and emissive power
for a blackbody as well as a real surface at elevated temperatures; surface radi-
ation properties such as spectral emissivity and absorptivity for real-surface
radiation; how to obtain the total emissivity or absorptivity from the fraction
method; how to perform energy balance from a flat surface including radi-
ation and convection; and solar and atmospheric radiation problems. This
chapter provides fundamental thermal radiation and surface properties that
are useful for many engineering applications such as surface radiators, space
vehicles, and solar collectors.
Fundamental Radiation 235
PROBLEMS
11.1. A diffuse surface having the following spectral distributions
(ελ = 0.3 for 0 ≤ λ ≤ 4 μm, ελ = 0.7 for 4 μm ≤ λ) is maintained
at 500 K when situated in a large furnace enclosure whose walls
are maintained at 1500 K. Neglecting convection effects,
a. Determine the surface’s total hemispherical emissivity (ε) and
absorptivity (α).
b. What is the net heat flux to the surface for the prescribed
conditions?
Given: σ = 5.67 × 10−8 (W/m2 K4 )
11.2. An opaque, gray surface at 27◦ C is exposed to an irradiation of
1000 W/m2 , and 800 W/m2 is reflected. Air at 17◦ C flows over the
surface, and the heat transfer convection coefficient is 15 W/m2 K.
Determine the net heat flux from the surface.
11.3. A diffuse surface having the flowing spectral characteristics (ελ =
0.4 for 0 ≤ λ ≤ 3 μm, ελ = 0.8 for 3 μm ≤ λ) is maintained at
500 K when situated in a large furnace enclosure whose walls are
maintained at 1500 K:
a. Sketch the spectral distribution of the surface emissive power
Eλ and the emissive power Eλ,b that the surface would have
if it were a blackbody.
b. Neglecting convection effects, what is the net heat flux to the
surface for the prescribed conditions?
c. Plot the net heat flux as a function of the surface temperature
for 500 ≤ T ≤ 1000 K. On the same coordinates, plot the heat
flux for a diffuse, gray surface with total emissivities of 0.4 and
0.8.
d. For the prescribed spectral distribution of ελ , how do the
total emissivity and absorptivity of the surface vary with
temperature in the range 500 ≤ T ≤ 1000 K?
11.4. The spectral, hemispherical emissivity distributions for two dif-
fuse panels to be used in a spacecraft are as shown.
For panel A: ελ = 0.5 for 0 ≤ λ ≤ 3 μm, ελ = 0.2 for 3 μm ≤ λ.
For panel B: ελ = 0.1 for 0 ≤ λ ≤ 3 μm, ελ = 0.01 for 3 μm ≤ λ.
Assuming that the backsides of the panels are insulated and that
the panels are oriented normal to the solar flux at 1300 W/m2 ,
determine which panel has high steady-state temperature.
11.5. From a heat transfer and engineering approach, explain how
a glass greenhouse, which is used in the winter to grow veg-
etables, works. Include sketches of both the system showing
energy flows and balances, and of radiation property data (radio-
active properties versus wavelengths) for greenhouse compo-
nents (glass and the contents inside the greenhouse). When appli-
cable, show the appropriate equations and properties to explain
the greenhouse phenomenon. When finished with the above for
a glass greenhouse, extend your explanation to global warm-
ing, introducing new radioactive properties and characteristics if
needed.
236 Analytical Heat Transfer
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. K.-F. Vincent Wong, Intermediate Heat Transfer, Marcel Dekker, Inc., New York, NY,
2003.
4. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
12
View Factor
cos θj
dqi−j = Ii dAi cos θi dwj−i = Ii cos θi dAj dAi (12.1)
R2
cos θi cos θj
= πIi dAi dAj
πR2
Performing integration over surface area i and surface area j, one obtains
radiation rate from surface i to surface j as
cos θi cos θj
qi−j = Ji dAj dAi (12.2)
πR2
Ai Aj
239
240 Analytical Heat Transfer
Aj
nj
θj
R
dAj
ni
dωj-i
θi
Ai
dAi
FIGURE 12.1
Radiation exchange between two diffuse isothermal surfaces.
Similarly,
1 cos θi cos θj
Fji = dAi dAj (12.4)
Aj πR2
Aj Ai
n
Fij = 1 (12.6)
j=1
2 3
1 N surfaces
FIGURE 12.2
View factor for a N-surface enclosure.
View Factor 241
That is,
Figure 12.3 shows the differential view factor between two differential areas
i and j, and between area i and differential area j. The differential view factor
between differential area i and differential area j can be obtained as
dAj
θj
n
R
n
θi
dAj
dAi
θj
n
R
n
θi
dAi
Ai
FIGURE 12.3
Concept of view factor between two differential areas.
242 Analytical Heat Transfer
Similarly, the differential view factor between area i and differential area j is
Ji (cos θi cos θj /πR2 ) dAi dAj dAj cos θi cos θj
dFAi−dAj = = dAi
J i Ai Ai πR2
Ai
Example 12.1
Determine the view factor between two parallel discs as shown in Figure 12.4.
Assume that Ai Aj , the distance between two surfaces is L, and the larger disc
has a diameter D.
dAj
Aj
θj
L R
θi
Ai
FIGURE 12.4
View factor between two parallel disks.
View Factor 243
where Ai = Ai dAi ,
with θi = θj = θ, R 2 = r 2 + L2 , cos θ = L/R, and dAj = 2πr dr ,
cos2 θ
Fij = dAj
πR 2
Aj
D/2
r dr D2
= 2L2 =
(r 2 + L2 )2 D 2 + 4L2
0
(1)
L1 + L2 − Lac
F1−2 = (12.7)
2L1
244 Analytical Heat Transfer
d 3 c
4 2
a b
1
FIGURE 12.5
Concept of Hotell’s cross-string method for 2-D geometry.
(2)
L1 + L4 − Lbd
F1−4 = (12.8)
2L1
(3)
Lac + Lbd − (L2 + L4 )
F1−3 = (12.9)
2L1
(1) F1−2 + F1−ac = 1
F1−2 = 1 − F1−ac
Lac
=1− Fac−1
L1
Lac
=1− (1 − Fac−2 )
L1
Lac Lac L2
=1− + · · F2−ac
L1 L1 Lac
Lac L2
=1− + · (1 − F2−1 )
L1 L1
Lac L2 L2 L1
=1− + − · · F1−2
L1 L1 L1 L2
Lac L2
=1− + − F1−2
L1 L1
L1 + L2 − Lac
∴ F1−2 =
2L1
(2) Similarly, F1−4 = ((L1 + L4 − Lbd )/2L1 )
(3) F1−2 + F1−3 + F1−4 = 1
Example 12.2
Determine the view factor between two parallel plates with partial blockages as
shown in Figure 12.6.
Example 12.3
Determine the view factor between two opposite circular tubes as shown in Figure
12.7. From Hottel’s cross-string method, the view factor is
2L1 − 2L2 L − L2
F1−2 = = 1
2A1 πR
where
A1
b b c
2
c
A2
FIGURE 12.6
View factor between two parallel plates with partial blockages.
246 Analytical Heat Transfer
e d
θ
R c
θ
b a
1 D 2
FIGURE 12.7
View factor between two opposite circular tubes.
Let X = 1 + D/(2R)
2 1
F1−2 = (x 2 − 1)1/2 + sin−1 −X
π X
2 π 1
F1−2 = (x 2 − 1)1/2 + − cos−1 −X
π 2 X
Example 12.4
Determine the view factor between two circular tubes with partial blockage as
shown in Figure 12.8. From Hottel’s cross-string method, the view factor can be
determined as follows:
Therefore,
A B E F
α D
α
β h
1 2
R h
C G
H I
d d
FIGURE 12.8
View factor between two circular tubes with partial blockage.
View Factor 247
TABLE 12.1
View Factors for 2-D Geometries
Geometry Relation
j
wj
α
Inclined parallel plates of equal width Fij = 1 − sin
2
and a common edge
j
w
α
i
w
1 + (wj /wi ) − [1 + (wj /wi )2 ]1/2
Perpendicular plates with a common Fij =
2
edge
j
wj
i
wi
wi + wj − wk
Three-sided enclosure Fij =
2wi
wj
wk
k j
i
wi
!
Parallel cylinders of different radii 1
Fij = π + [C2 − (R + 1)2 ]1/2 − [C2 − (R − 1)2 ]1/2
2π
R 1
rj +(R − 1) cos−1 −
ri C C
1
R 1
−(R + 1) cos−1 +
C C
i j R = rj /ri , S = s/ri
S C =1+R+S
continued
248 Analytical Heat Transfer
j
L
i
s2
s1
1/2 1/2
D 2 D s2 − D 2
Infinite plane and row of cylinders Fij = 1 − 1 − + tan−1
s s D2
s
D
j
Concentric cylinders A1
F12 = 1; F21 =
A2 A2
A1
F22 = 1 − F21 = 1 −
A2
A1
1
Long duct with equilateral triangular F12 = F13 =
2
cross-section
1 2
3
c 2
1/2 c
Long parallel plates of equal width F12 = F21 = 1 + −
a a
1
c
2
continued
View Factor 249
s
Long adjacent parallel cylinders of Let X = 1 + , then
d
equal diameters 1 1
F12 = F21 = (X 2 − 1)1/2 + sin−1 −X
1 2 π X
d d
s
Concentric spheres A1
F12 = 1; F21 =
A2
A2 A1
F22 = 1 − F21 = 1 −
A2
A1
1
Regular tetrahedron F12 = F13 = F14 =
3
3
2
1
1
Sphere near a large plane area F12 =
2
1
continued
250 Analytical Heat Transfer
1
A2
Area on the inside of a sphere F12 =
4πR2
A1
R A2
Source: Data from A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992; F. Incropera
and D. Dewitt, Fundamentals of Heat and Mass Transfer, John Wiley & Sons, Fifth Edition, 2002.
dAj
θj
Aj θi y
a
Ai dAi b
c
x
FIGURE 12.9
View factor between two adjacent surfaces.
⎡ ⎤
1 ⎢ ⎥
FAi−Aj = ⎣ ln R dxi dxj + ln R dyi dyj + ln R dzi dzj ⎦ (12.11)
2πAi
ci cj
⎡b ⎧b
⎨ 1/2
1 ⎣
FAi−Aj = dxj ln (xi − xj )2 + a2 dx
2πbc ⎩
0 0
⎫
0 1/2 ⎬
+ ln (xi − xj )2 + c2 + a2 dxi
⎭
b
⎧b
0 ⎨ 1/2
+ dxj ln (xi − xj )2 + 02 dx
⎩
b 0
⎫⎤
0 1/2 ⎬
+ ln (xi − xj )2 + c2 + 02 dxi ⎦ (12.12)
⎭
b
252 Analytical Heat Transfer
The following shows how to use Stokes’ theorem to determine the view factor
between two opposite surfaces [3] as shown in Figure 12.10.
⎡ ⎤
1 ⎢ ⎥
FAi−Aj = ⎣ ln R dxi dxj + ln R dyi dyj + ln R dzi dzj ⎦
2πAi
ci cj
⎧ ⎫
⎨c 1/2 ⎬
1
FAi−Aj = ln xi2 + (yj − yi )2 + a2 dyj dyi
2πbc ⎩ ⎭
ci 0
⎧ ⎫
⎨b 1/2 ⎬
1
+ ln (xj − xi )2 + (c − yi )2 + a2 dxj dxi
2πbc ⎩ ⎭
ci 0
⎧ ⎫
⎨0 1/2 ⎬
1
+ ln (b − xi )2 + (yj − yi )2 + a2 dyj dyi
2πbc ⎩ ⎭
ci c
⎧ ⎫
⎨0 1/2 ⎬
1
+ ln (xj − xi )2 xi2 + yi2 + a2 dxj dxi
2πbc ⎩ ⎭
ci b
0 c ! 1/2 1/2 1
1 2 2 2 2 2
= ln (yj − yi ) + a + ln b + (yj − yi ) + a dyj dyi
2πbc
c 0
+ other integrals
/
2a2 (1 + (b/a)2 )(1 + (c/a)2 )
= ln
πbc 1 + (b/a)2 + (c/a)2
b 2 1/2 −1 b/a
+ [1 + (c/a) ] tan
a [1 + (c/a)2 ]1/2
2 1/2
c b −1 c/a
+ 1+ tan 1/2
a a 1 + (b/a)2
1
b −1 b c −1 c
− tan − tan (12.13)
a a a a
Table 12.2 shows several useful view factors for 3-D geometries [4] that
can be determined by using Stoke’s theorem to transform area-to-line
integration.
View Factor 253
z c
b
A2(x2, y2, c)
a
c
y
b
A1(x1, y1, 0)
FIGURE 12.10
View factor between two opposite surfaces.
Example 12.5
Determine view factors F1−2 and F2−1 for the following geometries shown in
Figure 12.11.
For geometries (a) and (b),
where Fj−i , Fj−3 , F4−i , and F4−3 , are available from formulas or charts.
Example 12.6
Determine view factors F1−4 and F4−1 from Figures 12.13a and b:
A1 F1−4 = A2 F2−3
Ai Fi−j = A1 F1−j + A2 F2−j
TABLE 12.2
View Factors for 3-D Geometries
Geometry Relation
X = X/L, Y ⎧= Y/L
1/2
2 ⎨
Aligned parallel rectangles
(1 + X 2 )(1 + Y 2 )
Fij = ln
πXY ⎩ 1 + X2 + Y2
X
j +X(1 + Y 2 )1/2 tan−1
(1 + Y 2 )1/2
L Y
+Y(1 + X 2 )1/2 tan−1
i (1 + X 2 )1/2
Y ⎫
⎬
X −1 −1
−X tan X − Y tan Y
⎭
Source: F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, John Wiley & Sons,
Fifth Edition, New York, NY, 2002.
Hence,
1
A1 F1−4 = [A F − A1 F1−3 − A2 F2−4 ] (12.15)
2 i i−j
where Fi−j , F1−3 , and F2−4 , are available from Table 12.2 formulas or charts.
And, A1 F1−4 = A4 F4−1 .
View Factor 255
(a) D (b)
L 1 (top)
1 (top)
L
L
2 (side)
L=D
2 (side)
3 (bottom) 3 (bottom)
FIGURE 12.11
(a) A cylindrical furnace. (b) A cubic furnace.
A1
A3
Ai
A4 A2
Aj
FIGURE 12.12
Algebraic method.
A4
A3
A3 A4
A1 A1 A2
A2
FIGURE 12.13
Applications of shape factor algebra to opposing and adjacent rectangles.
Remarks
This chapter covers the same information as in the undergraduate-level heat
transfer. In the undergraduate-level heat transfer, students are expected to
know how to use those view factors available from tables or charts in order to
256 Analytical Heat Transfer
calculate radiation heat transfer between two surfaces for many engineering
applications. However, at the intermediate-level heat transfer, students are
expected to focus more on how to derive view factors instead of how to use
them. In particular, students are expected to know how to determine the view
factors by using Hottel’s string method for many 2-D geometries. For 3-D
geometry view factors, we do not go into much detail because of the required
double-area integrations that belong to advanced radiation heat transfer.
PROBLEMS
12.1 Determine the view factors for Examples 1, 2, 3, and 4, respectively.
12.2 Determine the view factors shown in Table 12.1.
12.3 Determine the view factors shown in Table 12.2.
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. K.-F.V. Wong, Intermediate Heat Transfer, Marcel Dekker, Inc., New York, NY, 2003.
4. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, John Wiley &
Sons, Fifth Edition, New York, NY, 2002.
13
Radiation Exchange in a Nonparticipating
Medium
Perform energy balance on the i surface: net energy = energy out (radios-
ity) − energy in (irradiation)
qi = Ai (Ji − Gi ) (13.1)
257
258 Analytical Heat Transfer
q Gi
A i
N
Ai
ρiGi
Ti i
Ei
}Ji
εi
FIGURE 13.1
Radiation heat transfer between N surfaces in an enclosure.
also
Ji = εi Ebi + (1 − εi )Gi (13.2)
Therefore,
Ji − εi Ebi Ji − εi Ji − Ji + εi Ebi
qi = Ai Ji − = Ai
1 − εi 1 − εi
εi (Ebi − Ji )
= Ai
1 − εi
Ebi − Ji
⇒ qi = (13.3)
(1 − εi )/(εi Ai )
Then perform energy exchange between surface i and the rest of surfaces j:
net energy = energy out (radiosity) − energy in (irradiation)
qi = Ai (Ji − Gi )
where
N
N
A i Gi = Fji Aj Jj = Fij Ai Jj (13.4)
j=1 j=1
Therefore,
⎛ ⎞
N
qi = Ai ⎝ Ji − Fij Jj ⎠
j=1
Radiation Exchange in a Nonparticipating Medium 259
N
By using j=1 Fij = 1, and multiplying to Ji ,
⎛ ⎞
N
N
qi = Ai ⎝ Fij Ji − Fij Jj ⎠
j=1 j=1
So that energy transfer between surface i and the rest of enclosure surfaces
j becomes
N
N
qi = Ai Fij (Ji − Jj ) = qij (13.5)
j=1 j=1
Ebi − Ji N N
qi = = Ai Fij (Ji − Jj ) = qij
(1 − εi )/εi Ai
j=1 j=1
Ebi − Ji N
Ji − J j
N
qi = = = qij (13.6)
(1 − εi )/(εi Ai ) (1/Ai Fij )
+ ,- . j=1 + ,- . j=1
surface resistance geometrical resistance
due to emissivity
due to view factor
N
Ji = εi Ebi + (1 − εi ) Fij Jj (13.7)
j=1
σ(T14 − T24 )
q1 = q12 = −q2 = (13.8)
(1 − ε1 )/(A1 ε1 ) + 1/(A1 F12 ) + (1 − ε2 )/(A2 ε2 )
qi1
J1
qi2
J2
( Ai Fi1 )
−1
FIGURE 13.2
Network representation of the radiative exchange between surface i and the remaining surfaces
of an enclosure.
–q2 A2 T2 ε2 –q2 A2 T2 ε2 2
2 2
q1 A1 T1 ε1 q1 A1 T1 ε1 1
–q2
r2 A2 T2 ε2
r1
q
r1
q1
q A1 T1 ε1
r2
Eb1 = σT 14 Eb2 = σT 24
Eb1 J1 J2 Eb2
q1 –q2
1– ε1 1 1– ε2
A1ε1 A1F12 A2ε2
FIGURE 13.3
Radiation between a two-surface enclosure.
Radiation Exchange in a Nonparticipating Medium 261
Special case 2—Radiation between two parallel surfaces with middle shields as
shown in Figure 13.4:
A1 σ(T14 − T24 )
q1 = q12 =
(1 − ε1 )/ε1 + (1/F13 ) + (1 − ε31 )/ε31
+(1 − ε32 )/ε32 + (1/F32 ) + (1 − ε2 )/ε2 (13.10)
A1 σ(T14 − T24 )
= = −q2
(1/ε1 ) + (1 − ε31 )/ε31 + (1 − ε32 )/ε32 + (1/ε2 )
q12
Radiation
shield
q1
ε31 ε32
1 3 2
A1 A3 A2
ε1 ε2
σT 14 σT 14
Eb1 J1 J31 Eb3 J31 J2 Eb2
q1 −q2
1– ε1 1 1– ε31 1– ε32 1 1– ε2
A1ε1 A1F13 A3ε31 A3ε32 A3F32 A2ε2
FIGURE 13.4
Radiation between two parallel surfaces with middle shield.
262 Analytical Heat Transfer
A2 T2 ε2 A2 T2 ε2
AR TR εR A2 T2 ε2
AR AR
AR TR 2
TR TR
εR q1 1
q2
εR
εR
A1 T1 ε1 A2 T2 ε2
A1 T1 ε1
A1 T1 ε1 A1 T1 ε1
EbR
1– εR
qR = 0
ARεR
JR = EbR
1 1
σT 14 A1F1R qR 2 ARFR2 σT 24
q1R
q1 Eb1 J1 q12 J2 Eb2
1– ε1 1 1– ε2 −q2
A1ε1 A1F12 A 2 ε2
FIGURE 13.5
Electric furnaces with a reradiating surface.
∵ qR = 0, therefore, q1 = −q2
σT14 − σT24
q1 = −q2 =
(1 − ε1 )/A1 ε1 + 1/(A1 F12 + 1/[(1/A1 F1R ) + (1/A2 F2R )])
+(1 − ε2 )/A2 ε2
(13.11)
where T1 , and T2 are given
AR FR2 = A2 F2R
And surface emissivity, area, and view factors are also given or predeter-
mined.
If q1 = −q2 is determined from above and if qR = 0, how to determine
reradiation surface temperature TR =?
From energy balance on the reradiation surface,
J1 − JR JR − J2
q1R = = qR2 =
1/A1 F1R 1/AR FR2
and
Eb1 − J1 1 − ε1
q1 = ⇒ J1 = σT14 − q1
(1 − ε1 )/A1 ε1 A 1 ε1
J2 − Eb2 1 − ε2
q2 = ⇒ J2 = q2 + σT24
(1 − ε2 )/A2 ε2 A 2 ε2
Radiation Exchange in a Nonparticipating Medium 263
1
d s
2 2
FIGURE 13.6
A radiant heater panel model.
from
J1 − JR JR − J2 A1 F1R J1 + AR FR2 J2
= ⇒ JR =
1/A1 F1R 1/AR FR2 A1 F1R + AR FR2
Therefore,
1/4
JR
JR = Eb,R = σTR4 ⇒ TR = (13.12)
σ
Special case 4—A radiant heater panel problem: A long radiant heater panel
consists of a row of cylindrical electrical heating elements, as shown in
Figure 13.6.
The above Equations 13.11 and 13.12 can be used to determine heat trans-
fer rate q1 = −q2 , and TR , respectively. However, we need to calculate view
factors F11 , F12 , and F1R (or F13 ). In Table 12.1,
1 1
F11 = (X 2 − 1)1/2 + sin−1 − X
π X
with X = 1 + (s/d).
Assume F12 ∼ = F13 for symmetry and F11 + F12 + F13 = 1.
Therefore, F12 = 1/2(1 − F11 ).
Case A—Given each surface temperature to determine its heat flux,Ti given, ⇒
qi = ?
Use energy balance on surface i and energy exchange between surface i and
the rest of surfaces j, from Equation 13.6,
Ebi − Ji
N
Ji − J j
= (13.13)
(1 − εi )/(εi Ai ) 1/Ai Fij
j=1
Applying the above equation to each surface (1, 2, 3,…, and N), respectively,
one obtains the following N radiosity linear equations (after rearranging
them).
Then coefficient matrix [A], column matrix [ J], and column matrix [C] can
be formed to satisfy the N linear equations. Therefore, the unknown radiosity
matrix [ J] can be determined by solving the given inverse matrices [A] and [C].
[A][ J] = [C]
[J] = [A]−1 [C] = · · ·
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
a11 a12 · · · a1N J1 c1
⎢ a21 a22 · · · a2N ⎥ ⎢ J2 ⎥ ⎢ c2 ⎥
[A] = ⎢
⎣ ···
⎥ [J] = ⎢ ⎥ [C] = ⎢ ⎥
··· ··· ··· ⎦ ⎣ · · ·⎦ ⎣ · · ·⎦
aN1 aN2 · · · aNN JN cN
Once the unknown radiosity J from each surface i has been solved from the
aforementioned matrix relation, radiation heat transfer from each surface can
be shown from Equation 13.6 as
Ebi − Ji σTi4 − Ji
qi = = (13.14)
(1 − εi )/εi Ai (1 − εi )/εi Ai
Special example for a three-surface enclosure problem: If surface temperatures
shown in Figure 13.7 are given (T1 , T2 , T3 ), how to determine surface heat
transfer rates (q1 , q2 , q3 )?
From Equation 13.13,
Ebi − Ji
N
Ji − J j
=
(1 − εi )/Ai εi 1/Ai Fij
j=1
Radiation Exchange in a Nonparticipating Medium 265
2 2
2 2
3 3 3
3
1 3
1
1 1
FIGURE 13.7
Radiation heat transfer among a three-surface enclosure.
where a11 = (A1 ε1 /(1 − ε1 ) + A2 F12 + A1 F13 ), a12 = −A2 F12 , a13 = −A1 F13 ,
c1 = (A1 ε1 /(1 − ε1 ))σT14 .
Apply for surface 2:
where a21 = −A2 F21 , a22 = ((A2 ε2 /1 − ε2 ) + A2 F21 + A2 F23 ) , a23 = −A2 F23 ,
c2 = (A2 ε2 /(1 − ε2 ))σT24 .
266 Analytical Heat Transfer
where a31 = −A3 F31 , a32 = −A3 F32 , a33 = (A3 ε3 /(1 − ε3 ) + A3 F31 + A3 F32 ),
c3 = (A3 ε3 /(1 − ε3 ))σT34 .
From the above three linear equations, the following matrix can be formed:
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
a11 a12 a13 J1 C1
A = ⎣ a21 a22 a23 ⎦ J = ⎣ J2 ⎦ C = ⎣ C2 ⎦
a31 a32 a33 J3 C3
[A][ J] = [C]
[J] = [A]−1 [C]
Eb1 − J1 σT14 − J1
⇒ q1 = =
(1 − ε1 )/(A1 ε1 ) (1 − ε1 )/(A1 ε1 )
Eb2 − J2 σT24 − J2
⇒ q2 = =
(1 − ε2 )/A2 ε2 (1 − ε2 )/A2 ε2
Eb3 − J3 σT34 − J3
⇒ q3 = =
(1 − ε3 )/A3 ε3 (1 − ε3 )/A3 ε3
Radiation Exchange in a Nonparticipating Medium 267
Case B—Given each surface heat flux to determine its temperature, qi given, ⇒
Ti = ?
Use energy exchange between surface i and the rest of surfaces j, from
Equation 13.6,
N
Ji − J j
qi = (13.15)
1/Ai Fij
j=1
Apply the above to each surface (1, 2, 3, …, and N) and obtains N radiosity
linear equations as
a11 J1 + a12 J2 + · · · + a1N JN = c1
a21 J1 + a22 J2 + · · · + a2N JN = c2
..
.
aN1 J1 + aN2 J2 + · · · + aNN JN = cN
Once matrix [J] has been solved, use Equation 13.14 on each surface i to
determine surface temperature on each surface
Ebi − Ji σTi4 − Ji
⇒ qi = =
(1 − εi )/εi Ai (1 − εi )/εi Ai
or
1 − εi
Ebi = σTi4 = qi + Ji
A i εi
Therefore,
1/4
qi (1 − εi )/Ai εi + Ji
Ti = (13.16)
σ
Case C—Combined Case A and Case B
• Some surfaces are given temperatures but heat fluxes are unknown.
• Some surfaces are given heat fluxes but temperatures are unknown.
• Use the same procedure shown for Case A and Case B in order to
form matrix [A], column matrix [J], and column matrix [C].
• After determining matrix [J], either heat fluxes or temperatures can
be determined.
Special case for the blackbody radiation problem: Use the aforementioned results
for any blackbody surface with unity emissivity (εi = 1).
268 Analytical Heat Transfer
dFdAi −dAj
K(r̄i , r̄j ) ≡
dAj
Therefore,
N
Ji (ri ) = εi σTi4 (r̄i ) + (1 − εi ) Jj (r̄j )K(r̄i , r̄j ) dAj (13.18)
j=1 A
j
Ai Aj
Ai Aj
dAi dAj
dAi ri rj
dAj
FIGURE 13.8
Radiation exchange between nonisothermal surfaces.
Radiation Exchange in a Nonparticipating Medium 269
When Ji (r̄i ) is determined by using the matrix method, and for given Ti (r̄i ),
then q εi 4
(r̄i ) = σTi (r̄i ) − Jj (r̄j ) (13.19)
A i 1 − εi
can be solved.
Example
Apply the numerical method—Simpson’s rule (Trapezoidal rule) for the nonisother-
mal surfaces shown in Figure 13.9.
Given: εa = 0.9, Ta = 1000−1500◦ C
εb = 0.2, Tb = 300◦ C
1. q̄a =?
2. Compare qa = qa1 + qa2 =?
q N
= Ji (r̄i ) − Jj r̄j K r̄i , r̄j dAj
A i
j=1
when Ji (ri ) is determined by the matrix method, and for given q/A i (ri ), then
1 − εi q
σTi4 (r̄i ) = (r̄ ) + Ji (r̄i )
εi A i i
a b ya yb
1000C
1m 2 300C
1250C
1m 1
1500C
1m
FIGURE 13.9
Radiation between nonisothermal surfaces.
270 Analytical Heat Transfer
a. Integral model
q
∞
q
= dλ (13.20)
A i A i
0
b. Band model
q N
q
= Δλk (13.21)
A i A i
k=1
ελ ελ
Nonmetal
Metal
Δλ1 Δλ2 Δλ3
λ λ
FIGURE 13.10
Radiation exchange among diffuse, isothermal, nongray surfaces.
Radiation Exchange in a Nonparticipating Medium 271
4(3) 4
1(3) 1
2(3) 2
Specular surface
FIGURE 13.11
Radiation exchange among diffuse and specular surfaces.
Assume all surfaces shown in Figure 13.12 are diffuse emitting, gray, and
isothermal; then
Ji = εi σTi4 + (1 − εi )Gi
Nd
Gid = Jj EAi−Aj (13.25)
j=1
where
EAi−Aj = FAi−Aj + ρsk FAi(k)−Aj
k
q Gi
A
i
N
Ai
ρiGi
Ti i
Ei
}Ji
εi
FIGURE 13.12
Radiation heat transfer between N surfaces (including diffuse and specular) in an enclosure.
272 Analytical Heat Transfer
N
Gis = εj σTj4 EAi−Aj (13.26)
j=Nd+1
Remarks
In undergraduate-level heat transfer, students are expected to know how to
calculate radiation heat transfer between two surfaces or between two sur-
faces with a third reradiating surface by using the electric network analogy
method for many engineering applications such as electric heaters, radia-
tion shields, and electric furnaces with insulating side walls, and so on. In
intermediate-level heat transfer, this chapter focuses on how to analyze and
solve radiation heat transfer problems in an N-surfaces enclosure by using the
matrix linear equations method for more complicated electric or combustion
furnaces applications. Students are expected to know how to set up a matrix
from linear equations by applying energy balance on each of N-surfaces
with given surface temperatures or surface heat fluxes BCs. Here we assume
that each N-surface has gray and diffuse properties and keeps at isother-
mal condition. We do not go into much details for any N-surface behaving
as nongray, nondiffuse (specular), or nonisothermal condition. These require
more complex mathematics and belong to advanced radiation topics.
PROBLEMS
13.1. A rectangular oven is 1 m wide, 0.5 m tall, and 2 m deep into the
paper and is used to bake a carbon-fiber cloth with an electric
heater at the top. All vertical walls are reradiating (reflectory and
insulated). Take ε1 = 0.7, ε2 = 0.9, and ε3 = 0.8. The heater tem-
perature is 650◦ C when 20 kW of power is supplied. Convection
is negligible.
Given:
W
σ = 5.67 × 10−8 2 4
m K
Radiation Exchange in a Nonparticipating Medium 273
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
4. R. Siegel and J. Howell, Thermal Radiation Heat Transfer, McGraw-Hill, New York,
NY, 1972.
5. J. Chen, Conduction and Radiation Heat Transfer, Class Notes, Lehigh University,
1973.
14
Radiation Transfer through Gases
At 1 atm total pressure, CO2 has partial pressure Pc and water vapor has
partial pressure Pw . L, geometric mean bean length, is defined as
4V ∼ 4V
L= = 0.9 (14.3)
As As
where V is the volume and As is surface area.
275
276 Analytical Heat Transfer
100
H2O
CO2
L = 100 cm
αλ,x-%
50
0
1 2 3 5 10 15 20 25 30
Wavelength - μm
FIGURE 14.1
Band emission of carbon dioxide and water vapor.
D d
CO2 + N2
a b a
L = 1.26R
1.0 ft-atm
1.0 ft-atm
PwL PcL
0.2
εw 0.1 ε c 0.1 0.2
0.01
0.01
FIGURE 14.2
Gas radiation properties for water vapor and carbon dioxide.
Radiation Transfer through Gases 277
dAj
dAi
FIGURE 14.3
Hemispherical gas radiation to an element area at the center of base.
The concept of geometric mean beam length for other gas mass geometries
will be discussed in a later section.
The total gas emissivity for combined CO2 and H2 O can be obtained as
εg = εc + εw − Δε (14.4)
dIλ (x)
= −kλ dx
Iλ (x)
278 Analytical Heat Transfer
T1 T2
Iλ,0 Iλ(x) Iλ L
0 dx L
x
FIGURE 14.4
Absorption in a gas.
ln Iλ (x) = −Kλ x + C1
Iλ (x) = e(−Kλ x+C1 ) = C e−Kλ x
at x = 0, C = Iλ,0
Iλ (x) = Iλ,0 e−Kλ x
at x = L
Iλ,L = Iλ,0 e−Kλ L (14.7)
This exponential decay is called Beer’s law. One can define the transmis-
sivity as
Iλ,L
τλ = = e−Kλ L (14.8)
Iλ,0
The absorptivity is
αλ = 1 − τλ = 1 − e−Kλ L .
where Kλ Ibλ is intensity gained due to gas emission, −KIλ (x) is the intensity
attenuated due to gas absorption.
Radiation Transfer through Gases 279
at x = 0, C = Iλ,0 − Ibλ
at x = L, and we obtain
Ibλ = Ib = σTg4
where Ii− is the intensity approaching surface dAi , using the concept devel-
oped in Equation 14.10, replacing L by R = Ij+ e−KR + Ibg (1 − e−KR ), Ij+ the
intensity leaving surface dAj = Jj /π, Jj the radiosity leaving surface dAj =
emission and reflection from surface dAj , Ibg the intensity from blackbody
280 Analytical Heat Transfer
dAj
θj
R
dGi
θi
dAi
FIGURE 14.5
Elemental surface for radiation in an enclosure containing an isothermal gray gas.
gas emission = Ebg /π = σTg4 /π, Ebg the blackbody gas emission = σTg4 , dwj =
dAj cos θj /R2 , R the distance (beam length) between surface dAi and dAj , and
K the gas absorption coefficient.
Therefore,
cos θi cos θj
dGi = Jj e−KR + Ebg (1 − e−KR ) dAj dAi
πR2
Aj
1
Gi = dGi dAi
Ai
cos θi cos θj
1
= Jj e−KR + Ebg (1 − e−KR ) dAj dAi (14.12)
Ai πR2
Ai Aj
The distance (beam length) R varies over the surface. For convenience, we
define a mean surface (mean beam length) Lij , such that
1 cos θi cos θj
−KLij −KLij
Gi = Jj e + Ebg (1 − e ) dAj dAi
Ai πR2
Ai Aj
= Jj e−KLij + Ebg (1 − e−KLij ) Fij (14.13)
Radiation Transfer through Gases 281
If KLij is small, for the optically thin gases low pressure, e−KLij ≈ 1 − KLij ,
Equation 14.14 becomes
1 cos θi cos θj
Lij = dAj dAi (14.15)
Ai Fij πR
Ai Aj
4V
= (14.19)
As
where V is the volume of the gas in the enclosure and As is the enclosure
surface area.
In general, the geometric mean beam length (Lij ) between surfaces i and j of
an enclosure should be determined from Equation 14.4, and can be deter-
mined by Equation 14.15 for the optically thin gas (i.e., small absorption
coefficient K, low pressure, and small enclosure Lij , KLij is small). In addi-
tion, it can be determined from Equations 14.18 and 14.19, respectively, for
a single-surface enclosure with a uniform temperature and emissivity. How-
ever, in some problems, for the optically thick gases (i.e., KL is not small,
high pressure), the geometric mean beam length (L) is less than the above-
mentioned values. From experience, the geometric mean beam length has
been proved to be a good approximation for the actual mean beam length.
For practice, Equation 14.3, L ∼
= 0.9(4V/As ), can be used.
282 Analytical Heat Transfer
qi = Ai (Ji − Gi ) (14.22)
Tg
i Gi
Ebg
Ji N
eg
Ti Ai qi
ag
FIGURE 14.6
Radiation heat transfer through gases in an enclosure.
Radiation Transfer through Gases 283
and
Ji = εi Ebi + (1 − εi )Gi (14.23)
Therefore,
Ebi − Ji
qi = (14.24)
(1 − εi )/(Ai εi )
qi = Ai (Ji − Gi )
= A i Ji − Fij Jj (1 − αij,g ) − Ebg εi,g + Ai εi,g Ji − Ai εi,g Ji
= Ai εi,g (Ji − Ebg ) + Ai Fij (1 − αij,g )( Ji − Jj ) (14.25)
where
Ai Gi = Aj Fji Jj (1 − αg ) + Ai εi,g Ebg
= Ai Fij Jj (1 − αg ) + Ai εi,g Ebg (14.26)
Therefore,
Ji − Ebg
N
Ji − J j
qi = + (14.27)
1 j=1 1
Ai εi,g Ai Fij (1 − αij,g )
+ ,- . + ,- .
resistance due to resistance due to view factor
gas emissivity and gas absorptivity
N
Ji − J j
qi =
1/Ai Fij
j=1
284 Analytical Heat Transfer
Therefore,
[A] [ J] = [C]
[J] = [A]−1 [C]
Similarly, Equation 14.28 can be used to form the matrix [A][J] = [C] as
follows:
Once matrix [ J] has been solved, then surface heat transfer rate can be
determined from Equation 14.24 as
Ebi − Ji
⇒ qi =
(1 − εi )/(εi Ai )
Radiation Transfer through Gases 285
N
N
Ebg − Ji
qg = Ai εi,g (Ebg − Ji ) = (14.29)
1/Ai εi,g
i=1 i=1
However, we still need Equations 14.24 and 14.27 or Equation 14.28 to solve
Ji using the matrix method. The aforementioned gas radiation problems can
also be solved by method 1—the electric network analogy method.
(a) 2
1 2
2
Gases Gases
Gas 2 2
1 or coal
2 1
2
2
2 Gases 2
Gases Gases
1 1
or coal or coal
(b) Ebg
1 1
A1ε1–g A2ε2–g
Eb1 J1 J2 Eb2
1 – ε1 1 1 – ε2 –q2
A1ε1 A1F12 (1 – α1–2,g) A2ε2
FIGURE 14.7
(a) Radiation between hot gases and two-surface enclosures; (b) Electric network for radiation
from hot gas to two-surface enclosure.
286 Analytical Heat Transfer
14.7b shows the associated electric network from hot gas to surfaces 1 and 2.
Hot gases release energy to surfaces 1 and 2 through their resistances (with
gas emissivity less than unity); each surface has its own resistance (with sur-
face emissivity less than unity). There is a reduced view factor between two
surfaces because gas cannot be completely seen through between two sur-
faces (due to the gas absorption effect). If gas absorptivity is zero, the view
factor is the same as the one with nonparticipating gases (such as air). Energy
balance on surfaces 1 and 2 must be performed in order to solve for radiosities
J1 and J2 , respectively. Then, energy releases from hot gases, and heat transfer
to surfaces 1 and 2 can be determined.
Special case 2: Figure 14.8 shows that several combustion furnaces can be
modeled as heat transfer between hot gases and a single gray surface enclo-
sure (assume an enclosure at a uniform temperature). The simple electric
network can be used to solve this type of problem.
From Equations 14.24 and 14.28, solve for q1 as
Special case 3—Net radiation heat transfer between nongray gases and a single
black enclosure: To further simplify the problem, assume that the whole furnace
1
Natural 1 1
gases
1
1 1
FIGURE 14.8
Radiation between hot gases and single-surface enclosure.
Radiation Transfer through Gases 287
Black six-surface
Ts As
εs = αs = 1
Gray gases Tg
CO2 + H2O
FIGURE 14.9
Radiation between hot gases and single black enclosure.
εg = εc + εw − Δε
αg = αc + αw − Δα
Δα = Δε
R Ebg
1 1
A1eg1 ARegR 1 – εR
AReR
Eb1 JR
R g R
1 – ε1 J1 1 EbR
A1F1t1gR qR = 0
A1e1
1
FIGURE 14.10
A gray enclosure and a refractory surface filled with a gray gas.
Special case 4—Gray enclosure filled with a gray gas: Figure 14.10 shows a
furnace consisting of a hot or a cold gray surface (1), a refractory surface R, and
a gray gas, g, where each element is assumed to be at a uniform temperature
−T1 , TR , and Tg ; determine the radiation heat transfer between the surface (1)
and gas as
σ(T14 − Tg4 )
q1g = (14.36)
(1 − ε1 )/A1 ε1 + 1/{A1 εg1 + 1/[1/(AR εgR ) + 1/(A1 F1R τ1gR )]}
Special case 5—Two gray surfaces with a gray gas: Figure 14.11 shows a furnace
consisting of a hot gray surface (1), cold gray surface (2), a refactory surface
R, and a gray gas g; determine the radiation heat transfer [1].
Real furnace applications—The zone method: Figure 14.12 shows the concept of
the zone method for real-furnace applications proposed by Hottle (MIT) [5].
In a real furnace, combustion gases as well as furnace surface temperatures are
nonuniform. The problem can be solved by dividing gases and surfaces into
a number of gas zones and surface zones, respectively. Energy balance can
be performed on each subsurface (each zone) and between each subsurface
and the rest of subsurfaces (zones) through gas zones. View factors need
to be calculated between subsurfaces too. The solution procedures are quite
complicated.
Ebg
2
1 1 1
A1εg1 A2εg2
R R ARε gR
g 1 1
A1F1Rτ1gR A2F2Rτ2gR
JR
Eb1 Eb 2
1 − ε1 J 1
1 J2 1 − ε 1
A1ε1 1 A1F12τ1g2 A1ε1
FIGURE 14.11
An enclosure of a gray hot surface, a gray cold surface, and a refractory surface.
Radiation Transfer through Gases 289
FIGURE 14.12
Concept of zone method for real furnace heat transfer problem.
The problem will be even more complicated if convection effects are con-
sidered, which is up to 20% of heat transfer rate for a circulation well-mixed
furnace. In reality, soot formation and radiation can further make the problem
harder to model.
+
Gases L T2
In cos θ
θ Iλ qλ
Tg(x) ds dy = ds cos θ
y y dw
τ θ
o T1
T1 x T2
T1 > T2
FIGURE 14.13
Radiation heat transfer through gases with varying temperature.
290 Analytical Heat Transfer
gray gas temperature changes from gray diffuse surface 1 to surface 2 [5,6].
∂I
=0
∂φ
∂I
= 0
∂θ
ebλ (y) rλ Gλ (y)
dIλ+ (θ) = −kλ Iλ+ ds − rλ Iλ+ ds + kλ ds + ds (14.37)
π 4π
y
τλ = (1/λp ) dy ⇒ τL = (L/λp ) = Lβλ
0
Boundary conditions:
Radiosity at surface 1,
J1 ε1 σT14 + (1 − ε1 )G1
I + (0) = =
π π
J2 ε2 σT24 + (1 − ε2 )G2
I − (L) = =
π π
From the above (1), (2), and BCs, a solution can be achieved by an
exponential or numerical method [6].
Radiation Transfer through Gases 291
1
2 Thick τL= ∞ Optically thick
Numerical
0 τ τL 1 0 τL 10
FIGURE 14.14
Nondimensional temperature and heat flux profiles.
For 1-D, gray gases, gray and diffuse surface, radiation only, the heat flux
σ T14 − T24 Q
q= (14.38)
1 + Q ((1/ε1 ) + (1/ε2 ) − 2)
q 1
Q≡ ≡ (14.39)
J1 − J 2 1 + (3/4)τL
σT 4 (τ) − J2
φ(τ) ≡ (14.40)
J1 − J 2
T 41
(a)
(b)
T 42
τ/τL
FIGURE 14.15
Temperature profile between two black surfaces through participating gas.
Remarks
In the undergraduate-level heat transfer, from charts, students are expected
to know how to read the emissivity and absorptivity values of water vapor
and carbon dioxide in a furnace at given size, temperature, and pressure,
and then apply these properties to calculate radiation transfer between these
gases and the furnace wall, assuming the blackbody furnace wall at a uniform
temperature.
In the intermediate-level heat transfer, this chapter focuses on how to
derive volumetric absorption; geometric mean beam length; radiation trans-
fer between gray gases at a uniform temperature and an N-surfaces furnace
with each surface at different gray diffuse uniform temperature conditions.
Students are expected to know how to analytically solve gas radiation prob-
lems by using the matrix linear equations method for an N-surfaces furnace
with various surface thermal BCs. By using electric network analogy, this
chapter has also provided several relevant engineering applications such as
radiation transfer between gas at a high uniform temperature and one-surface
furnace assuming the gray diffuse surface at a low uniform temperature;
or between gas at a high uniform temperature and two-surfaces furnace
assuming gray diffuse surfaces with each surface at different low uniform
temperatures.
This chapter does not go into much detail on real-furnace applications
with varying gas temperature and surface temperature by using the zon-
ing method. We only deal with the 1-D varying gas temperature, steady-state,
and gray diffuse surface problem. However, in real-life applications, there are
many gas radiation problems involving flow convection; 2-D or 3-D varying
Radiation Transfer through Gases 293
PROBLEMS
14.1. A hemispherical furnace is shown in Figure 14.7. If the furnace
contains CO2 + N2 gases at 1 atm pressure and temperature Tg ,
determine the total radiation heat transfer from gases to sur-
faces 1 and 2 (assume Tg > T1 > T2 , and make other necessary
assumptions).
a. Based on the analogy of electric resistance network, draw a
radiation heat transfer network from gases to surfaces 1 and 2.
b. Determine the total radiation from gases to surfaces 1 and 2.
The final solutions should be the function of given tempera-
tures, surface area, and radiation properties.
14.2. A hemispherical furnace is shown in Figure 14.7.
a. If the furnace contains N2 gas at 5 atm pressure, determine the
net radiation heat transfer from surface 1 to surface 2 (assume
T1 > T2 , and make other necessary assumptions).
b. If the furnace contains CO2 + N2 gases at 5 atm pressure and
temperature Tg , determine total radiation heat transfer from
gases to surfaces 1 and 2 (assume Tg > T1 > T2 , and make
other necessary assumptions).
c. Reconsider (b), if surface 1 now is a reradiating surface, deter-
mine the total radiation heat transfer to the surface 2 of the
furnace. In this new condition, comment on whether the radi-
ation transfer to surface 2 will be higher, the same, or lower
than that of (b) (make necessary assumptions).
14.3. A long hemicylindrical furnace is shown.
a. Determine the net radiation heat transfer from surface 1 to
surface 2, q12 .
b. If T2 = T2 (θ), describe how to determine q12 .
c. Consider combustion gray gas with a uniform temperature Tg
and emissivity εg inside the furnace, and determine the total
radiation heat transfer from gas to surfaces 1 and 2, qg . Assume
T1 , T2 constant.
14.4. A hemispherical furnace, with a reradiating floor and a water-
cooled ceiling, contains CO2 and N2 gases at 1 atm pressure
and 1000◦ C. Take ε1 = 0.8, ε2 = 0.7, D = 1 m, and T2 = 500◦ C.
Determine the radiant heat transfer to the ceiling of the furnace.
Assume gray gases.
Given: σ = 5.67 × 10−8 (w/m2 K4 ).
Volume of a sphere = (4/3)π((1/2)D)2
Surface of a sphere = 4π((1/2)D)2
14.5. A hemispherical furnace, with a reradiating floor and a water-
cooled ceiling, contains 2CO2 and 8N2 gases at 1 atm pressure
294 Analytical Heat Transfer
References
1. W. Rohsenow and H. Choi, Heat, Mass, and Momentum Transfer, Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1961.
2. A. Mills, Heat Transfer, Richard D. Irwin, Inc., Boston, MA, 1992.
3. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
4. R. Siegel and J. Howell, Thermal Radiation Heat Transfer, McGraw-Hill, New York,
NY, 1972.
5. H. Hottel and A. Sarofim, Radiative Transfer, McGraw- Hill, New York, NY, 1967.
6. J. Chen, Conduction and Radiation Heat Transfer, Class Notes, Lehigh University,
1973.
Appendix A: Mathematical Relations
and Functions
x x2 x3 x4
ex = 1 + + + + + ···
1! 2! 3! 4!
x3 x5 x7
sin x = x − + − + ···
3! 5! 7!
x2 x4 x6 x8
cos x = 1 − + − + − ···
2! 4! 6! 8!
x3 x5 x7
sinh x = x + + + + ···
3! 5! 7!
x2 x4 x6
cosh x = 1 + + + + ···
2! 4! 6!
ex − e−x
sinh x =
2
ex + e−x
cosh x =
2
d sin x = cos x dx; d cos x = − sin x dx
d sinh x = cosh x dx; d cosh x = + sinh x dx
sin x dx = − cos x + c; cos x dx = sin x + c
1 1 1 1
sin2 x dx = − sin x cos x + x + c = − sin 2 x + x + c
2 2 4 2
1 1 1 1
cos2 x dx = sin x cos x + x + c = sin 2x + x + c
2 2 4 2
sinh x dx = cosh x + c; cosh x dx = sinh x + c
297
298 Appendix A
1 1
sinh2 x dx = sinh x cosh x − x + c
2 2
1 1
cosh2 x dx = sinh x cosh x + x + c
2 2
(continued)
x sinh x cosh x tanh x
2.80 8.1919 8.2527 0.99263
2.90 9.0596 9.1146 0.99396
(x/2)2 (x/2)4
J0 (x) = 1 − 2
+ − ···
(1!) (2!)2
x (x/2)3 (x/2)5
J1 (x) = − + − ···
2 1!2! 2!3!
..
.
/ 0
(x/2)v (x/2)2 (x/2)4
Jv (x) = 1− + − ···
Γ(v + 1) 1!(v + 1) 2!(v + 1)(v + 2)
(x/2)2 (x/2)4
I0 (x) = 1 + + + ···
(1!)2 (2!)2
x (x/2)3 (x/2)5
I1 (x) = − + − ···
2 1!2! 2!3!
..
.
/ 0
(x/2)v (x/2)2 (x/2)4
Iv (x) = 1+ + + ···
Γ(v + 1) 1!(v + 1) 2!(v + 1)(v + 2)
300 Appendix A
d d
[J0 (mx)] = −mJ1 (mx), [Y0 (mx)] = −mY1 (mx)
dx dx
d d
[I0 (mx)] = mI1 (mx), [K0 (mx)] = −mK1 (mx)
dx dx
x J0 (x) J1 (x)
0.0 1.0000 0.0000
0.1 0.9975 0.0499
0.2 0.9900 0.0995
0.3 0.9776 0.1483
0.4 0.9604 0.1960
0.5 0.9385 0.2423
0.6 0.9120 0.2867
0.7 0.8812 0.3290
0.8 0.8463 0.3688
0.9 0.8075 0.4059
1.0 0.7652 0.4400
1.1 0.7196 0.4709
1.2 0.6711 0.4983
1.3 0.6201 0.5220
1.4 0.5669 0.5419
1.5 0.5118 0.5579
1.6 0.4554 0.5699
1.7 0.3980 0.5778
Appendix A 301
(continued)
x J0 (x) J1 (x)
1.8 0.3400 0.5815
1.9 0.2818 0.5812
2.0 0.2239 0.5767
2.1 0.1666 0.5683
2.2 0.1104 0.5560
2.3 0.0555 0.5399
2.4 0.0025 0.5202
A.3.3 Modified Bessel Functions of the First and Second Kinds [1]
(continued)
x e−x I0 (x) e−x I1 (x) ex K0 (x) ex K1 (x)
5.6 0.1728 0.1565 0.5188 0.5633
5.8 0.1696 0.1542 0.5101 0.5525
6.0 0.1666 0.1520 0.5019 0.5422
6.4 0.1611 0.1479 0.4865 0.5232
6.8 0.1561 0.1441 0.4724 0.5060
7.2 0.1515 0.1405 0.4595 0.4905
7.6 0.1473 0.1372 0.4476 0.4762
8.0 0.1434 0.1341 0.4366 0.4631
8.4 0.1398 0.1312 0.4264 0.4511
8.8 0.1365 0.1285 0.4168 0.4399
9.2 0.1334 0.1260 0.4079 0.4295
9.6 0.1305 0.1235 0.3995 0.4198
10.0 0.1278 0.1213 0.3916 0.4108
η
2
e−u du
2
erf η = √
π
0
erfcη ≡ 1 − erf η
x
η= √
4αt
References
1. F. Incropera and D. Dewitt, Fundamentals of Heat and Mass Transfer, Fifth Edition,
John Wiley & Sons, New York, NY, 2002.
2. V. Arpaci, Conduction Heat Transfer, Addison-Wesley Publishing Company, Read-
ing, MA, 1966.
Mechanical Engineering
Analytical
Heat
Transfer
Je-Chin Han
“… it will complete my library … [and] complement the existing
literature on heat transfer. It will be of value for both graduate
students and faculty members.”
—Bengt Sunden, Lund University, Sweden
K12869
6000 Broken Sound Parkway, NW
Suite 300, Boca Raton, FL 33487
711 Third Avenue
an informa business New York, NY 10017
2 Park Square, Milton Park
Abingdon, Oxon OX14 4RN, UK