You are on page 1of 97

Eng6901 - Heat Transfer I 1

1 Introduction
1.1 What is Heat Transfer?
• Heat transfer is the transmission of energy due to a temperature difference.
• Thermodynamics deals with equilibrium processes, and it can be used to determine the
temperatures at end states of a process, or the total amount of heat transferred during a
process. Information concerning the nature of a heat transfer process, the rate at which
it occurs, or intermediate temperature values is not available.
• Heat transfer analyses study the mechanisms by which heat is transferred, and permit
calculation of instantaneous heat transfer rates and temperatures, as well as initial and
final temperatures, and total amounts of heat transfer.
• There are three modes of heat transfer:

1. Conduction (solids and fluids) – Transfer of heat in a medium due to a temperature


differential (or gradient).
2. Convection (fluids) – Transfer of heat in a fluid medium due to bulk motion of the
fluid and diffusion of energy from a surface.
3. Radiation (solids, fluids, and vacuum) – Transfer of heat due to emission of energy
in the form of electromagnetic waves.

1.2 Conduction Heat Transfer


• Conduction heat transfer occurs due to a temperature differential in a medium.
• Heat will be transferred from high temperature (high energy) to low temperature (low
energy) regions of the medium.
• Molecules and atoms in the higher temperature region of a gas will have higher kinetic
and vibrational energy levels. Collisions between these particles and those in lower en-
ergy regions will transfer energy to the lower energy regions. This energy transfer will
increase the kinetic and vibrational energy levels of the lower energy region (→ increase
in temperature), and it is called a diffusion process.
• A similar process occurs in liquids. The molecules are more densely packed, however,
and collisions are more frequent, therefore, higher heat transfer rates can be obtained in
liquids than gases (for a given temperature difference).
• In solids, lattice vibrations, and the motion of free electrons are responsible for the trans-
fer of heat due to conduction. Solids with a very strong crystalline structure are poor
conductors (e.g. ceramics), while solids with many free electrons and a less regular crystal
lattice structure are good heat conductors (e.g. copper).
• The rate of conduction heat transfer is dependent on the temperature gradient within the
medium. From experiment it was found that:
dT
qx00 ∝ (1)
dx
where qx00 (W/m2 ) is the heat flux (i.e. rate of heat transfer per unit area) in the x (m)
direction, and dT /dx (◦ C/m) is the temperature gradient in the x direction.
Eng6901 - Heat Transfer I 2

• The constant of proportionality is the thermal conductivity, k (W/m·◦ C). Substitution of


k into Eq. (1) gives Fourier’s Law of heat conduction:

dT
qx00 = −k (2)
dx

• The thermal conductivity is a material property, and it is often temperature dependent.

• The negative sign is necessary, because temperature flows from high to low temperatures,
i.e. if heat transfer occurs in the positive x direction, then the temperature gradient is
negative. The negative sign is included to give a positive value of heat transfer in a
positive co-ordinate direction.

• Consider steady state, one-dimensional conduction heat transfer through a plane wall of
thickness L, and uniform k; Eq. (2) can be integrated to give:

T2 − T1 T1 − T2 ∆T
qx00 = −k =k =k (3)
L L L
i.e. the temperature profile is linear.

• To determine the total rate of heat transfer, qx (W), the area normal to the temperature
gradient (or the direction of heat transfer) must be included.

dT
qx = qx00 Ax = −kAx (4)
dx

1.3 Convection Heat Transfer


• Convection heat transfer is characterized by the transfer of energy by the random motion
of molecules (diffusion), and the bulk motion of a fluid medium.

• Consider the steady flow of a fluid with freestream temperature and velocity T∞ and
u∞ , respectively, over a heated flat plate which is maintained at a uniform and constant
temperature Ts .
Eng6901 - Heat Transfer I 3

• A hydrodynamic boundary layer will form on the plate due to the viscosity of the fluid.

• A thermal boundary layer will also form on the plate due to the difference in temperature
between the plate and the freestream.

• The hydrodynamic and thermal boundary layers may be of different thicknesses.

• The velocity of the fluid at the surface of the plate is zero (due to the no-slip boundary
condition), therefore, heat is transferred from the plate to the fluid by conduction (i.e.
a diffusion process). This heat is carried downstream from the plate due to the motion
of the fluid (advection). The combination of these two mechanisms of heat transfer is
convection.

• The thermal boundary layer thickens with distance along the plate, and if the plate tem-
perature is uniform, then the heat transfer rate decreases with distance (the temperature
gradient is reduced at the surface of the plate, therefore, the rate of diffusion of energy
from the surface is reduced).

• There are three types of convection heat transfer:

1. Forced convection – fluid motion is induced by an external means, e.g. pump, fan,
or pressure differential.
2. Free (or natural) convection – fluid motion is induced by buoyancy forces caused by
density gradients in the fluid due to temperature differentials.
3. Phase change – boiling and condensation heat transfer.
Eng6901 - Heat Transfer I 4

• Mixed convection exists when free and forced convection are of the same order of magni-
tude (e.g. low velocities and large buoyancy forces).

• Convection heat transfer rates are calculated using Newton’s Law of Cooling:

q 00 = h(Ts − T∞ ) (5)

where h is the convection heat transfer coefficient (W/m2 ·◦ C).

• The convection coefficient is dependent on the thermodynamic properties of the fluid, and
the properties of the fluid flow (i.e. laminar or turbulent flow, and the flow geometry).
The determination of h is nontrivial.

• Equation (5) is written assuming positive heat transfer from the surface to the fluid.

1.4 Radiation Heat Transfer


• Thermal radiation is energy emitted by matter in the form of electromagnetic waves.

• It is, perhaps, the most important form of heat transfer, as it is the mode by which the
earth receives heat from the sun.

• Radiation heat transfer can occur through fluids, solids, and a vacuum (e.g. space).

• The rate at which energy is emitted by a surface is called its emissive power, E (W/m2 ).
The maximum emissive power is called the blackbody emissive power, Eb , and is defined
by the Stefan-Boltzmann Law:
Eb = σTs4 (6)
where Ts is the absolute temperature of the surface (K), and σ is the Stefan-Boltzmann
constant, 5.67 × 10−8 W/m2 ·K4 .

• The heat flux emitted by a real body is less than that emitted by a blackbody (ideal
radiator):
E = σTs4 (7)
where  is the emissivity of the surface (0 ≤  ≤ 1) or fraction of energy emitted by the
surface compared to that emitted by a blackbody. The emissivity is a function of the
surface material and finish.

• Radiation may also be incident on a surface from its surroundings (e.g. the sun, a furnace,
a lamp). The total rate at which radiation is incident on a surface is called irradiation,
G (W/m2 ).

• Only a fraction of the incident radiation may be absorbed by a surface. The remainder
may be reflected and/or transmitted. The fractions absorbed, reflected, and transmit-
ted are defined by the absorptivity (α), the reflectivity (ρ), and the transmissivity (τ ),
respectively.
1=α+ρ+τ (8)
where the values of α, ρ, and τ are all between 0 and 1.

• The rate of absorption of incident radiation by a surface is Gabs (W/m2 ):

Gabs = αG (9)
Eng6901 - Heat Transfer I 5

• The values of α, , ρ, and τ are all dependent on the material, and the surface finish of
the material. Further, the properties may be dependent on the wavelength of radiation
(e.g. glass is transparent to visible light, but opaque to ultraviolet light).

• Later, it will be shown that for a gray body ( = const) α = .

• The radiant exchange of energy between two bodies can be written as follows:

qrad = FG σA(T14 − T24 ) (10)

where FG is a term related to the geometry of the two surfaces, and is called the shape
factor. The shape factor takes into account that not all energy from one surface will reach
another (e.g. only a small fraction of the energy emitted by the sun reaches the earth).

• For the special case where one body is surrounded by another (e.g. a person in a room),
the shape factor becomes 1 and the radiant exchange is:

qrad = σA(Ts4 − Tsur


4
) (11)

• Equation (11) may be written using a heat transfer coeffcient.

qrad = hr A(Ts − Tsur ) (12)

where
hr ≡ σ(Ts + Tsur )(Ts2 + Tsur
2
) (13)
i.e. the expression for the rate of radiation heat transfer has been linearized.

• Note: the heat transfer coefficient, hr , is a nonlinear function of temperature, so this


expression is only useful when Ts and Tsur are known.

• The advantage of writing the radiation heat transfer rate in the form of Eq. (12) is that
it will be possible to use a thermal resistance analogy similar to that used for convection
heat transfer.
Eng6901 - Heat Transfer I 6

• Often convection and radiation coexist. The steady state energy balance for a surface at
Ts exposed to a convection environment (u∞ , T∞ ), and a surroundings at Tsur is:

q = qconv + qrad = hA(Ts − T∞ ) + σA(Ts4 − Tsur


4
) (14)
or
q = qconv + qrad = hA(Ts − T∞ ) + hr A(Ts − Tsur ) (15)
• One of these modes of heat transfer may be negligible relative to the other and may be
neglected (this simplifies analysis).
• e.g. Assume air is the fluid and  = 0.8 is the surface emissivity for the case given above.
Compare the relative magnitudes of h and hr for the following environmental conditions.
Can convection or radiation be neglected?

1. Natural convection, h = 6 W/m2 ·◦ C, Ts = 127◦ C, T∞ = Tsur = 27◦ C.


2. Forced convection, h = 50 W/m2 ·◦ C, Ts = 127◦ C, T∞ = Tsur = 27◦ C.
3. Forced convection, h = 200 W/m2 ·◦ C, Ts = 127◦ C, T∞ = Tsur = 27◦ C.
4. Forced convection, h = 200 W/m2 ·◦ C, Ts = 600◦ C, T∞ = Tsur = 500◦ C.

Case h (W/m2 ·◦ C) hr (W/m2 ·◦ C) hr /h


1 6 7.94 1.32
2 50 7.94 0.16
3 200 7.94 0.04
4 200 101.5 0.51

Note:

1. Radiation is significant for natural convection situations, even for low temperatures.
2. Radiation is often not significant in the presence of forced convection, except in cases
where high temperatures are involved, therefore, radiation is often neglected under
forced convection conditions.
3. The concept of an effective heat transfer coefficient, including the effects of both
convection and radiation will be introduced later.
Eng6901 - Heat Transfer I 7

1.5 The 1st Law of Thermodynamics and Heat Transfer


• Consider application of the 1st Law of thermodynamics to the control volume (cv) shown
below:

• The potential exists for energy to be transported into the cv at the rate Ėin , transported
out at the rate Ėout , generated within the cv (e.g. chemical reaction) at the rate Ėg , and
the amount of energy stored in the cv can be changing at the rate Ėst .

Ėin + Ėg − Ėout = Ėst (16)

or integrating over a time interval ∆t:

Ein + Eg − Eout = ∆Est (17)

• Heat transfer is concerned with energy transports across the cv boundary due to temper-
ature differences (by conduction, convection, and radiation), generation of thermal energy
within the cv, and changes in the energy storage related to changes in temperature. So
the problem becomes to identify the relevant modes of heat transfer, generation term(s),
and storage term(s), and then evaluate them.

• Frequently in heat transfer analyses the 1st Law is used to give a surface energy balance
(i.e. at the boundary of a cv) to determine a balance between the different modes of heat
transfer. In this case, the cv has no mass or volume, i.e. it is a control surface, in which
the energy storage and generation terms are nonexisent, and only the surface phenomena
are important (i.e. the heat transfers into and out of the control surface).
Eng6901 - Heat Transfer I 8

• e.g. The flat, uninsulated, concrete slab roof of a house in Arizona:


Eng6901 - Heat Transfer I 9

2 The Governing Equations of Conduction Heat Transfer


2.1 Fourier’s Law
• Fourier’s Law:
dT
qx = −kAx (18)
dx
was obtained from generalization of experimental observations. It states that the con-
duction heat transfer rate in a material is proportional to the cross-sectional area of the
material (i.e. the area perpendicular to the heat transfer direction), and the local tempera-
ture gradient in the material. The constant of proportionality is the thermal conductivity,
k, which is a thermophysical property of the material. Remember that the negative sign
is included because heat flows in the direction of decreasing temperature.

• Division of Fourier’s Law by the area, Ax , defines the heat flux, qx00 :

dT
qx00 = −k (19)
dx

• Note: both the heat transfer rate, qx , and the heat flux, qx00 , use the subscript x, implying
heat transfer in the x direction and that heat transfer is directional, i.e. the conduction
heat transfer rate and flux are vector quantities:

~q 00 = qx00~ı + qy00~ + qz00~k (20)


∂T ~ ∂T ~ ∂T ~
 
= −k i+ j+ k (21)
∂x ∂y ∂z
= −k ∇T ~ (22)

2.2 The Heat Diffusion Equation


2.2.1 Derivation of the Heat Diffusion Equation
• The heat diffusion equation is the governing equation for conduction heat transfer.

• This equation facilitates the determination of the temperature field in a material. Knowl-
edge of the temperature field permits calculation of heat fluxes at any location, and
evaluation of other important information, e.g. evaluation of thermal stresses, volume,
phases...

• The equation is derived by application of the 1st Law to a differential control volume of
dimensions dx, dy, and dz:
Eng6901 - Heat Transfer I 10

• Consider the cv to consist of a homogeneous material, in which there is no advection.


In the presence of temperature gradients, conduction heat transfer will exist at the six
faces of the control volume. Energy may be generated within the control volume, and the
amount of energy stored within the control volume may change.
• Applying the 1st Law to this control volume:
Ėin + Ėg − Ėout = Ėst (23)
where Ėin and Ėout are due to conduction heat transfer at the surfaces of the control
volume.
• The rates of conduction heat transfer into the control volume are:
∂T
qx = −kdydz (24)
∂x
∂T
qy = −kdxdz (25)
∂y
∂T
qz = −kdxdy (26)
∂z
• Using a Taylor’s series expansion, neglecting second and higher order terms, the rates of
conduction heat transfer out of the control volume are:
∂qx
qx+dx = qx + dx (27)
∂x
∂qy
qy+dy = qy + dy (28)
∂y
∂qz
qz+dz = qz + dz (29)
∂z
Eng6901 - Heat Transfer I 11

• Energy may be generated within the control volume (due to chemical and/or nuclear
reaction, current flow):
Ėg = q̇∆V = q̇dxdydz (30)
where q̇ (W/m3 ) is a volumetric source or generation term.

• The rate of change of energy storage within the control volume is:
∂T ∂T
Ėst = ρcp ∆V = ρcp dxdydz (31)
∂t ∂t
where cp is the specific heat of the material within the control volume.

• The 1st Law for the control volume, Eq. (23), can be written as:
∂T
qx + qy + qz + q̇dxdydz − qx+dx − qy+dy − qz+dz = ρcp dxdydz (32)
∂t

• Substituting Eqs. (27) to (29), for qx+dx , qy+dy , and qz+dz , into Eq. (32) gives:

∂qx ∂qy ∂qz ∂T


− dx − dy − dz + q̇dxdydz = ρcp dxdydz (33)
∂x ∂y ∂z ∂t
Substituting the definitions of qx , qy , and qz , Eqs. (24) to (26), into Eq. (33), and dividing
by the volume of the control volume, gives:
∂ ∂T ∂ ∂T ∂ ∂T ∂T
     
k + k + k + q̇ = ρcp (34)
∂x ∂x ∂y ∂y ∂z ∂z ∂t
which is the heat diffusion equation governing three-dimensional, unsteady heat conduc-
tion in Cartesian co-ordinates, in a material with nonisotropic k.

• If the amount of energy stored within the control volume is invariant with time, i.e. a
steady state process, the heat diffusion equation reduces to:
∂ ∂T ∂ ∂T ∂ ∂T
     
k + k + k + q̇ = 0 (35)
∂x ∂x ∂y ∂y ∂z ∂z

• If k is isotropic (uniform) then the heat diffusion equation can be reduced to:

∂2T ∂2T ∂2T q̇ 1 ∂T


2
+ 2
+ 2
+ = (36)
∂x ∂y ∂z k α ∂t
or
q̇ 1 ∂T
∇2 T +
= (37)
k α ∂t
where α (m2 /s) is the thermal diffusivity of the material (α = k/ρcp ). Considering steady
state conditions, Eq. (37) reduces to Poisson’s equation:

∇2 T + =0 (38)
k
and for no source term the governing equation reduces to Laplace’s equation:

∇2 T = 0 (39)
Eng6901 - Heat Transfer I 12

• Considering one-dimensional heat conduction in Cartesian co-ordinates:


∂ ∂T ∂T
 
k + q̇ = ρcp (40)
∂x ∂x ∂t
under steady state conditions:
∂ ∂T
 
k + q̇ = 0 (41)
∂x ∂x
with no source term:
∂ ∂T
 
k =0 (42)
∂x ∂x
and uniform k:
∂2T
=0 (43)
∂x2
i.e. for steady state one-dimensional heat conduction with no source term, the heat flux
is constant in the direction of heat transfer, and if k is uniform, the temperature gradient
is constant.

• Using Eq. (37), and the definition of the Laplacian operator, ∇2 , one can easily write the
heat diffusion equation for isotropic k in the polar cylindrical (r, θ, z) co-ordinate system:

∂2T 1 ∂T 1 ∂2T ∂2T q̇ 1 ∂T


2
+ + 2 2
+ 2
+ = (44)
∂r r ∂r r ∂θ ∂z k α ∂t
Eng6901 - Heat Transfer I 13

and the spherical (r, θ, φ) co-ordinate system:

1 ∂ ∂T 1 ∂ ∂T 1 ∂2T q̇ 1 ∂T
   
r2 + sin θ + 2 + = (45)
r2 ∂r ∂r r2 sin θ ∂θ ∂θ r2 sin θ ∂φ2 k α ∂t

2.2.2 Boundary Conditions for the Heat Diffusion Equation


• Solution of the heat diffusion equation will require the application of appropriate boundary
conditions.

• Two boundary conditions are required for each spatial co-ordinate, since the heat diffusion
equation is second order with regard to the co-ordinate directions. The governing equation
is first order with regard to time, therefore, only the initial conditions are required.

• The following are the four spatial boundary conditions used in this course.

1. Specified surface temperature


T (0, t) = Ts (46)
Eng6901 - Heat Transfer I 14

2. Constant surface heat flux


(a) Finite heat flux
∂T
−k = qx00 (47)
∂x x=0

(b) Adiabatic, or insulated surface



∂T
=0 (48)
∂x x=0

3. Convection surface condition



∂T
−k = h (T∞ − T (0, t)) (49)
∂x x=0
Eng6901 - Heat Transfer I 15

2.3 Thermal Properties


2.3.1 Thermal Conductivity (k)
• Thermal conductivity is a thermophysical property of a material.

• In general, k is a function of temperature, and it is very important to choose the appro-


priate value of k for the temperature, or temperature range, that will be present in a
problem.

• Values for the thermal conductivity, and other properties are given in Table A.1 for
metallic solids, Table A.2 for nonmetallic solids, Table A.3 for common materials, Table
A.4 for gases, Table A.5 for saturated liquids, and Table A.6 for saturated water.

• A good electrical conductor is a good thermal condutor, therefore, a material with a low
resitivity, ρe (Ω·m2 /m), has a high thermal conductivity.

• The behaviour of k with variation in T depends on the substance:

– Gases – T increase, k increase (more frequent and violent molecular collisions).


– Liquids – T increase, k increase or decrease (dependent on the temperature variation
of the intermolecular forces).
– Solids – T increase, k increase or decrease (depends on the temperature variation of
intermolecular and interlattice forces).

2.3.2 Thermal Diffusivity (α)


• Thermal diffusivity is the ratio of the ability of a material to conduct thermal energy to
its ability to store thermal energy:
k
α= (50)
ρcp

• A material with a high thermal capacity (ρcp ) will respond slowly to variations in the
thermal conditions of its surroundings, whereas a material with a low thermal capacity
will respond quickly to transient thermal conditions (e.g. water (ρcp ∼ 4 MJ/m3 ·◦ C and
air (ρcp ∼ 1 kJ/m3 ·◦ C).

• From the definition of α, a material with a low α will respond slowly to variations in the
thermal conditions of its surroundings. This can also be deduced by studying the heat
diffusion equation, Eq. (37).

• Note that the tabular values of α have been mulitplied by 10n , so α for pure aluminium
at 300K is 97.1 × 10−6 m2 /s, Table A.1.
Eng6901 - Heat Transfer I 16

3 Steady State, One-Dimensional Heat Conduction


3.1 Cartesian Co-ordinates
• The equations governing steady state, one-dimensional heat conduction in a plane wall
with constant and uniform k and boundary conditions, and no source term are:

d2 T
=0 (51)
dx2
the heat diffusion equation, with boundary conditions (in this case):

1. T (0) = Ts,1
2. T (L) = Ts,2

and Fourier’s Law:


dT
qx = −kA (52)
dx
• Integrating Eq. (51) twice gives:
T = C1 x + C2
Applying the boundary conditions gives:

C2 = Ts,1
Ts,2 − Ts,1
C1 =
L
Therefore, the temperature distribution in the wall is:
Ts,2 − Ts,1
T = x + Ts,1 (53)
L
i.e. the T (x) profile is linear.
Eng6901 - Heat Transfer I 17

• Substituting the temperature gradient:


dT Ts,2 − Ts,1
= (54)
dx L
into Fourier’s Law gives:
dT kA
qx = −kA = − (Ts,2 − Ts,1 )
dx L
kA
= (Ts,1 − Ts,2 )
L
kA
= ∆T
L
∆T
= (55)
L/kA

3.1.1 Thermal Resistance


• Note the similarity of Eq. (55) with:

V1 − V2
i= (56)
R
i.e. a flow rate is equal to a potential (to cause a flow rate) divided by the resistance to
the flow rate.

• This leads us to the concept of thermal resistance (Rth ):

∆T
q= (57)
Rth

i.e. for the 1-D Cartesian system (steady state, uniform k, no source) the thermal resis-
tance is:
L
Rth ≡ (58)
kA
Eng6901 - Heat Transfer I 18

• A thermal resistance for convection heat transfer is derived as follows:


Ts − T∞
q = hA(Ts − T∞ ) = (59)
1/hA

therefore
1
Rth ≡ (60)
hA

• A thermal resistance for radiation heat transfer is defined as:


Ts − Tsur
q = hr A(Ts − Tsur ) = (61)
1/hr A

therefore
1
Rth ≡ (62)
hr A
Eng6901 - Heat Transfer I 19

3.2 Cylindrical Systems


• The steady state conduction heat transfer in a long hollow cylinder (e.g. a pipe) with con-
stant and uniform k and boundary conditions, and no source is governed by the following
equations:

d2 T 1 dT
+ =0 (63)
dr2 r dr
the heat diffusion equation, with boundary conditions:

1. T (r1 ) = Ts,1
2. T (r2 ) = Ts,2

and Fourier’s Law:


dT
qr = −k2πrL (64)
dr
• The shape of the T (r) profile can be deduced from Fourier’s Law (A increases with
increasing r, therefore, since qr and k are constant, dT /dr decreases with increasing r).
Eng6901 - Heat Transfer I 20

• The T (r) profile can be evaluated by integrating Eq. (63) twice and applying the temper-
ature boundary conditions to give (see Incropera and DeWitt):
Ts,1 − Ts,2 r
 
T (r) = ln + Ts,2 (65)
ln(r1 /r2 ) r2

• The temperature gradient can be found from this equation and substituted into Fourier’s
Law to determine the heat transfer rate and thermal resistance. Knowledge of the T (r)
profile is not required to define the thermal resistance, however, as Fourier’s Law can be
rearranged:
dr k2πL
=− dT
r qr
and integrated: Z r2
dr k2πL Ts,2
Z
=− dT
r1 r qr Ts,1
to give:
r2 2πkL
 
ln = (Ts,1 − Ts,2 )
r1 qr
or
Ts,1 − Ts,2 ∆T
qr = ln(r2 /r1 )
= (66)
Rth
2πkL
therefore the thermal resistance is:
ln(r2 /r1 )
Rth ≡ (67)
2πkL

3.3 Spherical Co-ordinates


• The equations governing steady state heat conduction through a spherical shell with
constant and uniform k and boundary conditions, and no source are:

1 d dT
 
2
r2 =0 (68)
r dr dr
the heat diffusion equation, with boundary conditions:
Eng6901 - Heat Transfer I 21

1. T (r1 ) = Ts,1
2. T (r2 ) = Ts,2

and Fourier’s Law:


dT
qr = −k4πr2 (69)
dr
• Again, the shape of the T (r) profile can be deduced from Fourier’s Law (A increases with
increasing r, therefore, dT /dr must decrease with increasing r).

• Integration of Fourier’s Law can be used to determine the thermal resistance for 1-D
spherical heat conduction: Z r2
dr 4πk Ts,2
Z
2
= − dT
r1 r qr Ts,1
which gives:
1 1 4πk
− = (Ts,1 − Ts,2 )
r1 r2 qr
or
Ts,1 − Ts,2 ∆T
qr =   = (70)
1 1 1 Rth
4πk r1 − r2

therefore, the thermal resistance is:


1 1 1
 
Rth ≡ − (71)
4πk r1 r2
Eng6901 - Heat Transfer I 22

3.4 Using Thermal Resistance(s) (Composite Walls)


• Consider a planar wall with uniform k, and no source, exposed to two known convection
environments, h1 , T∞,1 and h2 , T∞,2 . Determine the steady state heat transfer rate
through the wall, qx .

• The convection heat transfer rate on the left side of the wall (in the positive x direction)
is:
q1 = h1 A(T∞,1 − Ts,1 ) (72)
The conduction heat transfer rate through the wall is:
kA
qx = (Ts,1 − Ts,2 ) (73)
L
The convection heat transfer rate on the right side of the wall (in the positive x direction)
is:
q2 = h2 A(Ts,2 − T∞,2 ) (74)
• Application of steady state energy balances to control surfaces about surfaces 1 and 2
gives:

q1 = qx (75)
qx = q2 (76)
i.e.
kA
h1 A(T∞,1 − Ts,1 ) = (Ts,1 − Ts,2 ) (77)
L
kA
(Ts,1 − Ts,2 ) = h2 A(Ts,2 − T∞,2 ) (78)
L
Eng6901 - Heat Transfer I 23

i.e. two equations with two unknown temperatures, Ts,1 and Ts,2 . These equations can be
solved for the unknown temperatures and then the heat transfer rate, qx , can be evaluated.

• Using thermal resistances to solve the problem, i.e. determine qx :

T∞,1 − Ts,1 Ts,1 − Ts,2 Ts,2 − T∞,2


qx = = = (79)
Rc,1 Rw Rc,2

where
1
Rc,1 = (80)
h1 A
L
Rw = (81)
kA
1
Rc,2 = (82)
h2 A
Note: Eq. (79) implies use of the two surface energy balances.

• Equation (79) can be solved to show:

T∞,1 − T∞,2
qx = (83)
Rc,1 + Rw + Rc,2

i.e.
∆T
qx = P (84)
Rth
just as in series electric circuits.

• So for this problem the heat transfer rate, qx , can be written in several ways:
T∞,1 − T∞,2
qx = (85)
Rc,1 + Rw + Rc,2
T∞,1 − Ts,1
= (86)
Rc,1
Ts,1 − Ts,2
= (87)
Rw
Ts,2 − T∞,2
= (88)
Rc,2
T∞,1 − Ts,2
= (89)
Rc,1 + Rw
Ts,1 − T∞,2
= (90)
Rw + Rc,2
Eng6901 - Heat Transfer I 24

• What about a multilayer wall (i.e. composite wall) exposed to two known convection
environments (k’s constant and uniform, steady state, no source)?

1
Rc,1 = (91)
h1 A
LA
RA = (92)
kA A
LB
RB = (93)
kB A
LC
RC = (94)
kC A
1
Rc,2 = (95)
h2 A
The thermal resistance network gives the following expressions for qx :
T∞,1 − T∞,2
qx = (96)
Rc,1 + RA + RB + RC + Rc,2
T∞,1 − Ts,1
= (97)
Rc,1
Ts,1 − Ts,2
= (98)
RA
Ts,2 − Ts,3
= (99)
RB
Ts,3 − Ts,4
= (100)
RC
Ts,4 − T∞,2
= (101)
Rc,2
T∞,1 − Ts,2
= (102)
Rc,1 + RA
T∞,1 − Ts,3
= (103)
Rc,1 + RA + RB
T∞,1 − Ts,4
= (104)
Rc,1 + RA + RB + RC
Eng6901 - Heat Transfer I 25

• A composite system in cylindrical co-ordinates can be treated in a similar manner (steady


state, no source, constant and uniform k’s, long pipe).

1
Rc,1 = (105)
h1 2πr1 L
ln(r2 /r1 )
RA = (106)
2πkA L
ln(r3 /r2 )
RB = (107)
2πkB L
ln(r4 /r3 )
RC = (108)
2πkC L
1
Rc,2 = (109)
h2 2πr4 L
The thermal resistance network gives the following balance:
T∞,1 − T∞,2
qr = (110)
Rc,1 + RA + RB + RC + Rc,2
Eng6901 - Heat Transfer I 26

• What about a composite wall such as that used in building construction?

1
Rc,1 = (111)
h1 A
LA
RA = (112)
kA A
LB
RB = (113)
k B AB
LC
RC = (114)
k C AC
LD
RD = (115)
kD A
1
Rc,2 = (116)
h2 A
The result is a series-parallel thermal circuit:
T∞,1 − T∞,2
qx = (117)
Rc,1 + RA + Req + RD + Rc,2

where Req is an equivalent resistance to the parallel circuit:

1
Req = 1 1 (118)
RB + RC

Note: the circuit, as drawn, assumes that the surface temperatures of materials B and C
are equivalent. This is required for the assumption of one-dimensional heat conduction,
but it is not strictly true, especially as the difference in the magnitudes of kB and kC
increases.
Eng6901 - Heat Transfer I 27

3.5 Thermal Contact Resistance


• Thus far, perfect thermal contact has been assumed between materials in a composite
system, i.e. the temperatures at the surfaces of two materials in contact are identical.

• Consider steady state, one-dimensional, no source heat conduction between two materials
that are brought into contact. The temperature distribution appears as follows:

• The heat transfer rate is:


T1 − T2,A T2,A − T2,B T2,B − T3
qx = = = (119)
LA /kA A 1/hc A LB /kB A

where 1/hc A is the thermal contact resistance, Rt,c , and hc (W/m2 ·◦ C) is the contact
coefficient.

• When materials are mechanically joined, an extra resistance arises due to the joint, i.e.
the roughness of the contact surfaces.
Eng6901 - Heat Transfer I 28

• Note: continuous contact is not maintained between materials A and B. If continuous


contact existed the T profile would be continuous, the slope would be discontinuous if
kA 6= kB .

• Heat transfer through the joint occurs due to:

1. conduction between A and B at the points of contact; and


2. conduction, and radiation at high temperatures, through the gaps.

• The value of hc will depend on the thermal conductivity of the fluid in the voids, kf , the
thermal conductivities of the two materials in contact, the contact area between the two
materials, Ac , the void area, Av , and the thickness of the joint, Lg . The thermal circuit
for the joint is:

T2,A − T2,B T2,A − T2,B T2,A − T2,B


qx = = + (120)
1/hc A Rm Rf
where
Lg
Rf = (121)
kf Av
Lg Lg
Rm = + (122)
2kA Ac 2kB Ac
therefore
1 1
1 1 (123)
Lg L + Lg hc A
2kA Ac
+ 2k gA kf Av
B c

and
1 Ac 2kA kB Av
   
hc = + kf (124)
Lg A kA + kB A

• To decrease the thermal contact resistance, i.e. increase hc :

– increase Ac relative to Av (smoother surface or increased joint pressure)


– decrease Lg (smoother surface or increased joint pressure)
– increase kf (fill the void with thermal grease)
00 (m2 ·◦ C/W), i.e. 1/h the
• Tables 3.1 and 3.2, Incropera and DeWitt, list values of Rt,c c
contact conductance, for representative joints and void materials. More data is available
from manufacturers of thermal greases.
Eng6901 - Heat Transfer I 29

3.6 Conduction with Thermal Energy Generation


• Energy generation in materials may occur due to resistance heating (i.e. I 2 R), nuclear
reaction, and chemical reaction. The absorption of radiation may also be treated as a
source term. This generation of energy is treated as a volumetric source term, q̇ (W/m3 ).

• Consider steady state, one-dimensional heat conduction in a plane wall with uniform k,
and q̇:

• The governing equation is:


d2 T q̇
2
+ =0 (125)
dx k
with boundary conditions T (−L) = Ts,1 and T (L) = Ts,2 .

• Integrating Eq. (125) twice:


q̇ 2
T =− x + C1 x + C2 (126)
2k
and using the boundary conditions to evaluate C1 and C2 gives:
 2 !
q̇L2 x Ts,2 − Ts,1 x Ts,1 + Ts,2
 
T (x) = 1− + + (127)
2k L 2 L 2

• If Ts,1 = Ts,2 = Ts :
 2 !
q̇L2 x
T (x) = 1− + Ts (128)
2k L
and the maximum temperature occurs at x = 0:

q̇L2
T (0) = T0 = + Ts (129)
2k
Using T0 the temperature distribution can be written as:
 2
T (x) − T0 x
= (130)
Ts − T0 L
Eng6901 - Heat Transfer I 30

• Note: dT /dx|x=0 = 0 when the maximum temperature is located at x = 0, and dT /dx = 0


at the surface of a perfectly insulated material, therefore, Eq. (128) can be used for the
temperature distribution in a material with uniform source in contact with a perfectly
insulated surface.

• When the surfaces of the wall are exposed to convection environments, Eq. (127) can
be used for the temperature distribution when steady state energy balances are used on
control surfaces at surfaces 1 and 2 to determine Ts,1 and Ts,2 .


dT
−k = h2 (Ts,2 − T∞,2 ) (131)
dx L

dT
−k = h1 (T∞,1 − Ts,1 ) (132)
dx −L
but
dT q̇x Ts,2 − Ts,1
=− + (133)
dx k 2L
Eng6901 - Heat Transfer I 31

therefore, the energy balances are


Ts,2 − Ts,1
q̇L − k = h2 (Ts,2 − T∞,2 ) (134)
2L
Ts,2 − Ts,1
−q̇L − k = h1 (T∞,1 − Ts,1 ) (135)
2L
If, h1 = h2 = h, and T∞,1 = T∞,2 = T∞ , then:

q̇L
Ts,1 = Ts,2 = Ts = T∞ + (136)
h

• Consider a composite wall consisting of three materials. Material A is resting on a per-


fectly insulated surface, material B has a uniform source term q̇B , and material C is
exposed to a convection environment.

• Note: the temperature distribution in material B is given by a form of Eq. (128):


2 !
q̇B L2B x

T (x) = 1− + Ts,3 (137)
2kB LB

where Ts,3 is found from an energy balance applied to the control surface at surface 3, or
by using an overall energy balance:

Ėg = Ėout (138)

i.e.
Ts,3 − T∞
q̇B LB = (139)
LC /kC + 1/h
therefore
LC 1
 
Ts,3 = T∞ + q̇B LB + (140)
kC h
Eng6901 - Heat Transfer I 32

Also
q̇B LB
Ts,4 = T∞ + (141)
h
And
q̇B L2B
Tmax = Ts,1 = Ts,2 = + Ts,3 (142)
2k
• NOTE: The use of thermal resistances requires qx = const, and in a material with a
source qx 6= const at each cross-section. DO NOT use resistances in any material with
a source.

3.6.1 Cylindrical Co-ordinates


• Consider steady state, one-dimensional conduction heat transfer in a solid cylinder with
uniform k and q̇.

• The governing equation is:


1 d dT q̇
 
r + =0 (143)
r dr dr k
with boundary conditions:

dT
=0 or T (0) is finite (144)
dr r=0
and
T (ro ) = Ts (145)

• Integrating Eq. (143) twice:

q̇ 2
T =− r + C1 ln r + C2 (146)
4k
and applying the boundary conditions gives:
2 !
q̇r2 r

T (r) = o 1− + Ts (147)
4k ro
Eng6901 - Heat Transfer I 33

• If the outside of the cylinder is exposed to a convection environment, an overall energy


balance can be used to determine Ts :

q̇(πro2 L) = h(2πro L)(Ts − T∞ ) (148)

or
q̇r
Ts = T∞ + (149)
2h

3.6.2 Summary
• General expressions for the temperature distribution, heat flux, and heat transfer rate
for Cartesian, radial and spherical co-ordinate systems are summarized in Appendix C of
Incropera and DeWitt.

3.7 Heat Transfer from Extended Surfaces (Fins)


• Consider convection heat transfer from a surface area A. The convection resistance is
1/hA. To promote heat transfer from the surface the thermal resistance may be decreased
by increasing h or increasing the surface area exposed to the convection environment. Fins
are used as a means to increase the surface area, and promote heat transfer. Examples of
fins are the finned tube used in a car radiator, an air-cooled motorcycle engine, electric
and hotwater baseboard heaters, computer chip heat sinks, and furnace walls.

• In the best possible scenario, the increased area will remain at the original temperature
of the unaugmented surface area. In general, this does not occur, due to the conduction
resistance in the material that has been added. The ideal case can be approached for
short fins of very high thermal conductivity.

• To develop a means of analysing fins, the 1st Law will be applied to the control volume
in the fin shown below.

• Assumptions:

1. steady state;
Eng6901 - Heat Transfer I 34

2. no source;
3. uniform h and k; and
4. one-dimensional conduction in the fin. This implies that at any location x along the
length of the fin the temperature is uniform across the cross-sectional area Ac . This
assumption is most appropriate for thin fins of high thermal conductivity.

• Applying the 1st Law to the control volume of length dx:


" # " # " #
Energy conducted Energy conducted Energy convected from
− − =0
in left face out right face the exposed surface area

i.e.
qx − qx+dx − dqconv = 0 (150)
where
dT
qx = −kAc (151)
dx
dT d dT
 
qx+dx = −kAc −k Ac dx (152)
dx dx dx
and
dqconv = hdAs (T − T∞ ) (153)

• Substituting Eqs. (151), (152), and (300) into Eq. (150) gives:
d dT h dAs
 
Ac − (T − T∞ ) = 0 (154)
dx dx k dx

• Defining the surface area:


dAs = P dx
where P is the perimeter, and assuming uniform cross-sectional area, Ac :
d2 T hP
2
− (T − T∞ ) = 0 (155)
dx kAc

• Defining a temperature difference:


θx = Tx − T∞ (156)
Eq. (155) can be written as:
d2 θ
− m2 θ = 0 (157)
dx2
where
hP
m2 = (158)
kAc
• Equation (157) governs the steady-state, one-dimensional temperature distribution in a
fin with uniform cross-sectional area, and uniform thermal conductivity and convection
heat transfer coefficient. The general solution is:
θx = C1 emx + C2 e−mx (159)
where the constants C1 and C2 depend on the two boundary conditions imposed on the
fin.
Eng6901 - Heat Transfer I 35

• One boundary condition is that the base temperature of the fin is fixed at Tb :

θx=0 = Tb − T∞ = θb (160)

Substitution of Eq. (160) into Eq. (159) gives:

θb = C1 + C2 (161)

3.7.1 Case A: Convection Heat Loss from the Tip of the Fin
• This is the most realistic tip boundary condition. It is derived by performing a steady
state energy balance on a control surface around the tip of the fin, i.e. [heat conducted
in the left face] = [heat convected away from the right face]:


dT
hAc (TL − T∞ ) = −kAc
dx x=L
or, using θ:

hθL = −k (162)
dx x=L
The tip boundary condition gives:

h(C1 emL + C2 e−mL ) = km(C2 e−mL − C1 emL ) (163)

• Solving Eqs. (161) and (163) for C1 and C2 gives the following temperature distribution
in the fin.
θ cosh(m(L − x)) + (h/mk) sinh(m(L − x))
= (164)
θb cosh(mL) + (h/mk) sinh(mL)
• The rate of heat loss from the fin would be of interest. This may be determined by
integrating the convection heat flux over the surface area of the fin.
Z Z
qf = h(Tx − T∞ ) dAs = hθx dAs
Af Af

Fortunately, there is an easier way to calculate qf . An energy balance at the base of the
fin gives:
dT dθ
qf = −kAc = −kAc (165)
dx x=0 dx x=0
Eng6901 - Heat Transfer I 36

Using the temperature distribution given by Eq. (164) to determine the temperature
gradient at the base of the fin results in:
p sinh(mL) + (h/mk) cosh(mL)
qf = hP kAc θb (166)
cosh(mL) + (h/mk) sinh(mL)

3.7.2 Case B: The Insulated Tip


• The insulated tip boundary condition implies that the tip of the fin is perfectly insulated,
therefore, the temperature gradient at the tip is zero.

dT dθ
= =0 (167)
dx x=L
dx x=L
This boundary condition would apply when a fin butts against an insulated surface, or in
a symmetry problem. It is also very useful in approximating Case A.

• Application of the boundary condition gives:

C1 emL − C2 e−mL = 0 (168)

which when solved along with the base boundary condition, Eq. (161), results in the
following temperature distribution:

θ cosh(m(L − x))
= (169)
θb cosh(mL)

• Applying Eq. (165) results in the following expression for the heat transfer rate:
p
qf = hP kAc θb tanh(mL) (170)

3.7.3 Case C: Specified Tip Temperature


• When the tip temperature of a fin is specified, the tip boundary condition is:

θx=L = TL − T∞ = θL (171)

which when used with Eq. (161) gives:

θ (θL /θb ) sinh(mx) + sinh(m(L − x))


= (172)
θb sinh(mL)

• Applying Eq. (165) results in:


p cosh(mL) − θL /θb
qf = hP kAc θb (173)
sinh(mL)
Eng6901 - Heat Transfer I 37

3.7.4 Case D: Very Long (or Infinite) Fin


• If a fin is infinitely long, the tip temperature will be equivalent to the environment tem-
perature:
TL→∞ − T∞ = θL→∞ = 0 (174)
Which when solved with the base boundary condition, Eq. (161), gives:
θ
= e−mx (175)
θb
and the heat transfer rate is: p
qf = hP kAc θb (176)

• An infinite fin may not seem very practical, however, infinity is a relative term. For a fin
with a low thermal conductivity, the fin tip temperature may approach the environment
temperature in a length of several centimeters. The advantage of being able to use this
boundary condition is the simple solution, as opposed to that obtained for Case A. This
was of importance before the days of the calculator.

3.7.5 Fin Efficiency


• A fin efficiency can be defined as follows:
Actual heat transfer rate
ηf =
Heat transfer rate if the whole fin is at Tb
or
qf qf
ηf = = (177)
qmax hAf θb
If the fin efficiency is known, the fin heat transfer rate can be calculated easily:
qf = ηf hAf θb (178)

• The Case A fin efficiency is:


1 sinh(mL) + (h/mk) cosh(mL)
 
qf = (179)
mL cosh(mL) + (h/mk) sinh(mL)

• The case B fin efficiency is:


tanh(mL)
ηf = (180)
mL
• The case D fin efficiency is:
1
ηf = (181)
mL
• Note: Even in these advanced times, the calculation of qf for Case A is a tedious affair.
If only it could be appoximated with a simpler solution. It can! When a corrected length,
Lc , is used, the Case B solution can be used to approximate Case A. The corrected length
is defined as:
Lc = L + Atip /P (182)
and the fin surface area would be calculated as:
Af = P Lc (183)
Note:
Eng6901 - Heat Transfer I 38

1. The length correction is defined such that the tip area will be included automatically
when the fin area is calculated in this manner.
2. For a pin fin:
πD2 /4 D
Lc = L + =L+
πD 4
3. For a rectangular fin (t << z):
zt t
Lc = L + =L+
2(z + t) 2

• The Case A solution is approximated by the Case B (insulated tip) solution when Lc is
used to replace L in the insulated tip solutions. Then:

tanh(mLc )
ηf = (184)
mLc

• To make things even easier, ηf may be plotted against a nondimensional parameter,


therefore, for a constant cross-sectional area, one ηf curve may be generated.

• Consider a thin rectangular cross-section fin.


1/2 1/2
hP 2h
 
mLc = Lc = Lc
kAc kt

where P/Ac = 2/t, see Eq. (3). Defining a profile area:

Ap = Lc t

mLc may be written as follows:


!1/2
2h
mLc = Lc3/2 (185)
kAp

The fin effciency for all rectangular cross-section fins is plotted in Fig. 3.18, Incropera
and DeWitt. This diagram or Eq. (184) may be used to determine the efficiency of any
rectangular cross-section fin. Note: mLc would be evaluated using Eq. (3.7.5) or (185).

• The fin solutions discussed thus far are for constant cross-section fins, e.g. cylindrical pin,
or straight rectangular fins. Fins may also be triangular, conical, annular (circumferential
or circular), or other shapes. The cross-sectional area of all these fins is variable. Equa-
tion (154), which accounts for the variation of cross-sectional area with x would have to
be solved to obtain the temperature distribution and heat transfer rate for these fins.
Unfortunately, the solution to this equation may be tedious and involve Bessel functions,
so it would be time consuming to evaluate (see Section 3.6.4, Incropera and DeWitt).
Fortunately, fin efficiency may be used. Equations for the fin efficiency of several types of
fin geometry are given in Table 3.5, Incropera and DeWitt, and ηf is plotted in Fig. 3.18
for parabolic and triangular fins, and in Fig. 3.19 for annular rectangular profile fins. The
appropriate corrected lengths, Lc and profile areas, Ap are included in Table 3.5 and the
figures. The fin heat transfer rate would again be calculated from Eq. (178).
Eng6901 - Heat Transfer I 39

3.7.6 Fin Resistance


• The heat transfer rate for a fin may be written as follows:
θb
qf = ηf hAf θb = (186)
Rt,f
where Rt,f is the thermal resistance of the fin.
1
Rt,f = (187)
ηf hAf

• The fin thermal resistance becomes useful when the total heat loss from a surface is of
interest. For example, consider a finned surface with N fins. The total rate of heat

transfer from the surface would be equivalent to the rate of heat transfer from all the fins,
and all of the exposed surface area:

q = N (qf + qb ) (188)

where qb is the rate of heat transfer from the exposed surface area, Ab , between two fins.
θb θb
qb = = (189)
Rt,b 1/hAb
Note: the fin and base resistance are in parallel.
Eng6901 - Heat Transfer I 40

• Equation (188) may be written as follows:

q = (N ηf hAf + N hAb )θb (190)

• If there is a contact resistance between the fin and the base surface (e.g. due to a me-
chanical joining process), an extra resistance must be included.

• Defining the contact resistance:


1
Rt,c = (191)
hc Ac
the heat transfer rate from one fin would be:
θb
qf = (192)
Rt,f + Rt,c

and the total heat transfer rate from a finned surface may be written as:
θb
q= (193)
Req

where Req is the equivalent resistance for the parallel circuit:

1
Req = 1 1 (194)
Rt,f +Rt,c + Rt,b
Eng6901 - Heat Transfer I 41

4 Forced Convection: External Flows


4.1 Flow Over Flat Surfaces
4.1.1 Non-Dimensional form of the Equations of Motion
• Consider a two-dimensional, steady state, incompressible flow of a constant property,
Newtonian fluid, with freestream velocity u∞ , past a flat plate of length L.

• The governing equations for this flow are the continuity equation:
∂u ∂v
+ =0 (195)
∂x ∂y
and the Navier-Stokes equations:
!
∂u ∂u ∂p ∂2u ∂2u
ρu + ρv =− +µ + 2 (196)
∂x ∂y ∂x ∂x2 ∂y
!
∂v ∂v ∂p ∂2v ∂2v
ρu + ρv =− +µ + (197)
∂x ∂y ∂y ∂x2 ∂y 2

• Define the following non-dimensional variables:

x∗ = x/L u∗ = u/u∞ p∗ = p/ρu2∞


(198)
y ∗ = y/L v ∗ = v/v∞

• Substitution of these non-dimensional variables into Eq. (195) results in the following
form of the continuity equation:
u∞ ∂u∗ u∞ ∂v ∗
+ =0
L ∂x∗ L ∂y ∗
which can be written in the following non-dimensional form:
∂u∗ ∂v ∗
+ =0 (199)
∂x∗ ∂y ∗

• Substitution of the non-dimensional variables into the x-component of the Navier-Stokes


equations, Eq. (196), gives:
!
ρu2∞ ∗ ∂u∗ ρu2∞ ∗ ∂u∗ ρu2∞ ∂p∗ µu∞ ∂ 2 u∗ ∂ 2 u∗
u ∗
+ v ∗
= − ∗
+ 2 + ∗2
L ∂x L ∂y L ∂x L ∂x∗2 ∂y
Eng6901 - Heat Transfer I 42

Multiplying this equation by L/ρu2∞ gives:


!
∂u∗ ∂u∗ ∂p∗ µ ∂ 2 u∗ ∂ 2 u∗
u∗ ∗ + v ∗ ∗ = − ∗ + + ∗2
∂x ∂y ∂x ρu∞ L ∂x∗2 ∂y

But ReL = ρu∞ L/µ, therefore, the non-dimensional form of the x-component of the
Navier-Stokes equations can be written as:
!
∂u∗ ∂u∗ ∂p∗ 1 ∂ 2 u∗ ∂ 2 u∗
u∗ ∗ + v∗ ∗ =− ∗ + + ∗2 (200)
∂x ∂y ∂x ReL ∂x∗2 ∂y

• Similarly, the non-dimensional form of the y-component of the Navier-Stokes equations


is: !
∗ ∗ ∂p∗ ∂ 2v∗ ∂ 2v∗
∗ ∂v ∗ ∂v 1
u +v =− ∗ + + ∗2 (201)
∂x∗ ∂y ∗ ∂y ReL ∂x∗2 ∂y

• Equations (199), (200), and (201) are the non-dimensional forms of the equations that
govern the steady state, two-dimensional, incompressible flow of a constant property,
Newtonian fluid, with freestream velocity u∞ , past a flat plate of length L.

• Note: the only parameter in these equations is the Reynolds number, therefore, Re should
appear as a parameter in the solutions for the hydrodynamic boundary layer on a flat
plate.

4.1.2 Order of Magnitude Analysis for a Boundary Layer Flow


• Consider the order of magnitude of the non-dimensional variables that appear in Eqs.
(199), (200), and (201):

O(x∗ ) = 1 O(u∗ ) = 1 O(v ∗ ) =?


(202)
O(y ∗ ) = δ/L = δ ∗  1 O(p∗ ) =? O(ReL ) =?

• Now, consider the order of magnitude of each term in Eqs. (199), (200), and (201):

∂u∗ ∂v ∗
+ =0
∂x∗ ∂y ∗

!
∂u∗ ∂u∗ ∂p∗ 1 ∂ 2 u∗ ∂ 2 u∗
u∗ ∗ + v∗ ∗ =− ∗ + + ∗2
∂x ∂y ∂x ReL ∂x∗2 ∂y
Eng6901 - Heat Transfer I 43

!
∗ ∂p∗ ∗ ∂ 2v∗ ∂ 2v∗
∗ ∂v ∗ ∂v 1
u ∗
+ v ∗
= − ∗
+ + ∗2
∂x ∂y ∂y ReL ∂x∗2 ∂y

• Note:

1. From the continuity equation O(v ∗ ) = δ ∗  1 (since both terms must be of the same
magnitude).
2. For the viscous terms to be the same order as the inertia terms in the x-component
of the Navier-Stokes equations O(ReL ) = 1/δ ∗2  1.
3. For the pressure term to be the same order as the inertia terms in the x-component
of the Navier-Stokes equations (to prevent infinite accelerations) O(p∗ ) = 1.

• The non-dimensional forms of the equations governing the flow in the boundary layer are:
∂u∗ ∂v ∗
+ =0 (203)
∂x∗ ∂y ∗
!
∂u∗ ∂u∗ ∂p∗ 1 ∂ 2 u∗
u∗ ∗ + v∗ ∗ =− ∗ + (204)
∂x ∂y ∂x ReL ∂y ∗2
∂p∗
0=− (205)
∂y ∗
• Or in dimensional form:
∂u ∂v
+ =0 (206)
∂x ∂y
!
∂u ∂u ∂p ∂2u
ρu + ρv =− +µ (207)
∂x ∂y ∂x ∂y 2
∂p
0=− (208)
∂y
• One term has been eliminated from the x-component of the Navier-Stokes equations.

• The y-component of the Navier-Stokes equations has been reduced to a hydrostatic pres-
sure distribution.

• By performing the order of magnitude analysis one equation has been simplified, and the
number of equations that must be solved has been reduced by one.
Eng6901 - Heat Transfer I 44

4.1.3 Non-Dimensional form of the Energy Equation


• Consider a two-dimensional, steady state, incompressible flow of a constant property,
Newtonian fluid at freestream temperature T∞ and velocity u∞ past a flat plate of length
L, maintained at a constant temperature Ts .

• The energy equation may be written as:


!
∂T ∂T ∂2T ∂2T
ρcp u + ρcp v =k 2
+ + µΦ (209)
∂x ∂y ∂x ∂y 2
where 2 2 2
∂u ∂v ∂u ∂v
  
Φ=2 +2 + + (210)
∂x ∂y ∂y ∂x
• Define the following non-dimensional variables
x∗ = x/L u∗ = u/u∞ θ = (T − Ts )/(T∞ − Ts )
(211)
y ∗ = y/L v ∗ = v/u∞

• Substituting these non-dimensional variables into Eqs. (209) and (210) gives:
!
ρcp u∞ (T∞ − Ts ) ∂θ ∂θ k(T∞ − Ts ) ∂2θ ∂2θ µu2∞ ∗
 
u∗ ∗ + v ∗ ∗ = + + Φ (212)
L ∂x ∂y L2 ∂x∗2 ∂y ∗2 L2
where 2 2 2
∂u∗ ∂v ∗ ∂u∗ ∂v ∗
  
Φ∗ = 2 +2 + + (213)
∂x∗ ∂y ∗ ∂y ∗ ∂x∗
• Simplifying Eq. (212):
!
∂θ ∂θ k ∂2θ ∂2θ µu∞
u∗ ∗ + v ∗ ∗ = + + Φ∗ (214)
∂x ∂y ρcp u∞ L ∂x∗2 ∂y ∗2 ρcp L(T∞ − Ts )

• But !
k k µ k µ 1 1
  
= = = = (215)
ρcp u∞ L ρcp u∞ L µ µcp ρu∞ L P rReL Pe

µu∞ µu∞ u∞
 
=
ρcp L(T∞ − Ts ) ρcp L(T∞ − Ts ) u∞
!
u2∞ µ

=
cp (T∞ − Ts ) ρu∞ L
!
u2∞ µ

= −
cp (Ts − T∞ ) ρu∞ L
Ec
= − (216)
ReL
Eng6901 - Heat Transfer I 45

• Note:
ρu∞ L inertia forces
Reynolds # = ReL = ∝ (217)
µ viscous forces
µcp ν rate of diffusion of momentum
Prandtl # = P r = = ∝ (218)
k α rate of diffusion of thermal energy
convective transport of thermal energy
Peclet # = P e = P rReL ∝ (219)
conductive transport of thermal energy
u2∞ kinetic energy/unit volume of flow
Eckert # = Ec = ∝ (220)
cp (Ts − T∞ ) thermal energy/unit volume of flow

• The non-dimensional form of the energy equation can be written as follows:


!
∗∂θ ∂θ 1 ∂2θ ∂2θ Ec ∗
u + v∗ ∗ = + − Φ (221)
∂x∗ ∂y Pe ∂x∗2 ∂y ∗2 ReL

• Note: the only parameters in the thermal problem are P r, ReL , and Ec.

4.1.4 Order of Magnitude Analysis for a Thermal Boundary Layer


• Similar to the analysis of the hydrodynamic boundary layer, consider the order of mag-
nitude of each term in Eq. (221):

O(x∗ ) = 1 O(u∗ ) = 1 O(Ec) =?


∗ ∗ ∗ ∗
O(y ) = δT /L = δT  1 O(v ) = δ  1 O(P r) =?
(222)
O(θ) = 1 O(ReL ) = 1/δ ∗2 O(P e) =?

• Expanding all terms of the non-dimensional form of the energy equation:


! 2 2 2 !
∂θ ∂θ 1 ∂2θ ∂2θ Ec ∂u∗ ∂v ∗ ∂u∗ ∂v ∗
  

u + v∗ ∗ = + − 2 +2 + +
∂x∗ ∂y Pe ∂x∗2 ∂y ∗2 ReL ∂x∗ ∂y ∗ ∂y ∗ ∂x∗

• Note:

1. For the conduction terms to be of the same order as the convection terms O(1/P e) =
δT∗2 . This is sensible, since O(ReL ) = 1/δ ∗2 , and O(P r) ≈ 1 for most common fluids.
2. For any of the viscous dissipation terms to be of the same order as the remainder of
the equation, the only possibility is for O(Ec) ≈ 1, then the (∂u∗ /∂y ∗ )2 term will
remain
Eng6901 - Heat Transfer I 46

• The non-dimensional form of the energy equation for the thermodynamic boundary layer
on a flat plate (steady state, constant property, Newtonian fluid, incompressible flow) is:
2
∂θ ∗ ∂θ 1 ∂2θ Ec ∂u∗


u ∗
+ v ∗
= ∗2
− (223)
∂x ∂y P e ∂y ReL ∂y ∗
But Ec ≥ 1 only at high velocities,

e.g. for air, (cp ≈ 1000 J/kg·o C, (Ts − T∞ ) ≈ 100o C → u∞ ≈ 316 m/s for Ec =
u2∞ /(cp (Ts − T∞ )) = 1. The speed of sound at 300K is 347 m/s.

therefore, at low velocities, viscous dissipation is negligible. Viscous dissipation is very im-
portant at high velocities, e.g. the space shuttle, and SR-71 Blackbird spy plane. Viscous
dissipation gives rise to frictional heating.

• Neglecting viscous dissipation, the equation governing the thermal boundary layer on a flat
plate for steady, two-dimensional, incompressible flow of a constant property, Newtonian
fluid is:
∂T ∂T ∂2T
ρcp u + ρcp v =k 2 (224)
∂x ∂y ∂y
or
∂T ∂T ∂2T
u +v =α 2 (225)
∂x ∂y ∂y

4.1.5 Skin Friction and Heat Transfer Coefficients


• We would be interested in the frictional drag due to the hydrodynamic boundary layer.
The frictional drag is due to the shear stress at the plate surface, i.e.:

∂u
τs = µ (226)
∂y y=0

• The shear stress is often written in terms of a skin friction coefficient, Cf :

ρu2∞
τs = Cf (227)
2

• Since the velocity gradient ∂u/∂y at y = 0 varies with x, the skin friction coefficient will
also be a function of x. Further, the non-dimensional Navier-Stokes equations, Eqs. (200)
and (201) illustrate that the only parameters that would influence the solution for the
velocity gradient at the wall (i.e. y = 0) are Re and ∂p/∂x. But, the pressure gradient is
only a function of x, and it is determined by the geometry of the flow, therefore, for flows
of different fluids past the same geometry, only the Reynolds number and position on the
body should influence the skin friction coefficient:

Cf = Cf (x, Rex ) (228)

• To solve for Cf we need the velocity distribution in the boundary layer, therefore, we
need to solve Eqs. (206) and (207).
Eng6901 - Heat Transfer I 47

• The heat flux at the surface of the plate exposed to the convection environment, qs00 , can
be written as follows:
00 ∂T
qs = h(Ts − T∞ ) = −k (229)
∂y y=0
Heat is transferred from the wall to the fluid by conduction (since the molecules of fluid
next to the wall have zero velocity relative to the plate).

• The heat transfer coefficient is defined as follows:



−k ∂T

∂y y=0
h= (230)
Ts − T∞
To determine h we need the temperature gradient at the wall, i.e. the temperature dis-
tribution, therefore, we need to solve the energy equation for the boundary layer, i.e. Eq.
(224), which will require a prior solution for the hydrodynamic boundary layer.

• The non-dimensional form of the energy equation, Eq. (223), illustrates that the only
non-dimensional parameters that should appear in the thermal boundary layer solution
are x∗ , y ∗ , P e (or Re and P r), and Ec.

• Instead of working with the heat transfer coefficent, it is common to use a non-dimensional
variable called the Nusselt number (N u):
hL Actual heat transfer in the presence of flow
N uL = ∝ (231)
k Heat transfer if only conduction occurs
or locally:
hx
N ux = (232)
k
• Since only P r, Re, and Ec are the parameters of the flow:

h = h(P r, Re, Ec)


N u = N u(P r, Re, Ec)

• If Ec is small, h and N u are only functions of P r and Re and:

h = h(P r, Re) (233)


N u = N u(P r, Re) (234)

4.1.6 Solutions for Laminar Boundary Layer Flow


• Blausius (1908) developed an analytical solution for the hydrodynamic boundary layer
(see Section 7.2.1, Incropera and DeWitt):
5x
δ = 1/2
(235)
Rex
Cf,x = 0.664Re−1/2
x (236)

• Polhausen (1921) developed an analytical solution for the thermal boundary layer:
hx x
N ux = = 0.332Rex1/2 P r1/3 P r ≥ 0.6 (237)
k
Eng6901 - Heat Transfer I 48

• Note:

1. These correlations are for local values, i.e. they are functions of x.
2. The boundary layer thickens at the rate of x1/2 . For a given x, the thickness decreases
with increasing Re (as the influence of viscous forces decreases).
3. The Nusselt number increases at the rate of x1/2 but hx decreases at the rate of
x1/2 (i.e. the thermal boundary layer thickens with increasing x, decreasing the
temperature gradient and the heat transfer rate at the surface of the plate).
4. The expected non-dimensional parameters have arisen in the analytical solutions.

• Often, we are not interested in the local skin friction, or heat transfer rate, but in the
total frictional drag over a surface, or the total heat transfer rate from (or to) the surface.

• The total frictional drag force, D, would be defined as follows:

ρu2∞
Z L Z L
D= τs w dx = Cf,x w dx (238)
0 0 2
Since ρu2∞ /2 is constant:
ρu2∞
D = Cf A (239)
2
where Z L
1
Cf = Cf,x dx (240)
L 0
Substituting Eq. (236) into Eq. (240) and performing the integral gives:
−1/2
C f = 1.328ReL (241)

• The total heat transfer rate from the plate, q, can be evaluated as follows:
Z L
q= hx (Ts − T∞ )w dx (242)
0

but (Ts − T∞ ) is constant for this problem, therefore,

q = hA(Ts − T∞ ) (243)

where Z L
1
h= hx dx (244)
L 0
Substituting hx from Eq. (237) into Eq. (244) gives:
k 1/2
hL = 0.664 ReL P r1/3 P r ≥ 0.6 (245)
L
or an average Nusselt number can be defined as:
1/2
N uL = 0.664ReL P r1/3 P r ≥ 0.6 (246)

• Note: Cf , N uL , and hL are twice the corresponding value at the position L from the
leading edge (due to the exponent on Rex ).
Eng6901 - Heat Transfer I 49

• The above correlations for h and N u have restrictions on the Prandtl number. Churchill
and Ozoe experimentally obtained the following correlation for laminar flow over a flat
isothermal plate which is valid for all Prandtl numbers:
1/2
0.3387Rex P r1/3
N ux =  1/4 P ex ≥ 100 (247)
1 + (0.0468/P r)2/3

where N uL = 2N uL .

• The correlations developed thus far are based on the assumption that the fluid properties
are constant and uniform. When there is a significant difference between the plate and
freestream temperatures, the fluid properties are evaluated at a film temperature:
Ts + T∞
Tf = (248)
2

• When the plate is not heated over its entire length (e.g. heating starts at a location ξ
from the leading edge of the plate) Eq. (237) is modified to give:

"  3/4 #−1/3


ξ
N ux = 0.332Rex1/2 P r1/3 1− (249)
x

Constant heat flux


• For a laminar boundary layer flow over a flat plate which has a constant and uniform wall
heat flux (qs00 , W/m2 ) the local Nusselt number is:

N ux = 0.453Rex1/2 P r1/3 P r ≥ 0.6 (250)

Note: only the constant has changed between Eq. (237) and Eq. (250).

• If the heat flux is given, we would be interested in the local wall temperature, Ts,x , or the
mean temperature difference Ts − T∞ :
k
qs00 = hx (Ts,x − T∞ ) = N ux (Ts,x − T∞ ) (251)
x
therefore
qs00 x
Ts,x − T∞ = (252)
N ux k
Eng6901 - Heat Transfer I 50

So: Z L 00
1 qs x q 00 L/k
Ts − T∞ = dx = s (253)
L 0 N ux k N uL
Using Eq. (250) to determine N ux :
qs00 L/k
Ts − T∞ = 1/2
(254)
0.680ReL P r1/3
and
3
qs00 = hL (Ts − T∞ ) (255)
2
Note:

1. The mean Nusselt number for laminar flow over a flat plate with qs00 = const is:
1/2
N uL |qs00 =const = 0.680ReL P r1/3 (256)
which is only 2% larger than the mean value for the constant Ts boundary condition,
Eq. (246), therefore, it is acceptable to use any of the N uL correlations for constant
Ts to determine Ts − T∞ for the constant qs00 boundary condition.
2. The correlation developed by Ozoe and Churchill may also be used for the constant
heat flux boundary condition by replacing the constants 0.3387 and 0.0468 in Eq.
(247) with 0.4637 and 0.0207, respectively.

4.1.7 Reynolds-Colburn (or Chilton-Colburn) Analogy


• The skin friction coefficient for laminar boundary layer flow over a uniform temperature
flat plate is given by Eq. (236). This equation can be rearranged as follows:
Cf,x
= 0.332Rex−1/2 (257)
2
• The expression for the local Nusselt number for a laminar boundary layer on a uniform
temperature flat plate is Eq. (237). Dividing Eq. (237) by Rex P r:
N ux
= 0.332Rex−1/2 P r−2/3 (258)
Rex P r
the left hand side of this equation is called the Stanton number, Stx , therefore:
Stx P r2/3 = 0.332Re−1/2
x (259)

• Comparison of Eqs. (257) and (259) gives:


Cf,x
Stx P r2/3 = (260)
2
or
Cf
StP r2/3 = (261)
2
• This is the Reynolds-Colburn analogy between fluid friction and heat transfer. For ex-
ample, if experimental measurements are made of the frictional drag on a body → C f →
St → h → heat transfer rate (or vice versa).
• The Reynolds-Colburn analogy applies for laminar and turbulent boundary layers on flat
plates, and, in a modified form, for turbulent tube flow.
Eng6901 - Heat Transfer I 51

4.1.8 Turbulent Boundary Layers


• Consider a steady, incompressible flow of a Newtonian fluid past a flat plate. The
freestream conditions are constant at u∞ and T∞ , and the plate temperature is a uniform
and constant Ts .

• As the Reynolds number increases, the inertia forces begin to dominate the viscous forces
and instabilities in the flow can no longer be damped out by viscous effects. The flow will
go through a transition from a laminar boundary layer to a turbulent boundary layer.

• The turbulent portion of the boundary layer is thicker than the laminar portion, and
instead of smooth lamina, it consists of eddies of varying size.

• The effect of these eddies is to give a mean velocity profile that is fuller than in the laminar
portion of the boundary layer. This will result in higher velocity gradients at the surface
of the plate → higher wall shear stress (and Cf,x ) → higher frictional drag.

• Similarly, the mean temperature profile in the boundary layer becomes fuller, and the
temperature gradient at the wall will increase. Since the temperature gradient at the wall
increases, then the heat transfer rate will increase, therefore, by Eq. (230), the convection
heat transfer coefficient will be higher than for laminar flow.

• In laminar flow, the shear stress is a function of the fluid properties (i.e. µ) and the velocity
gradient. In turbulent flow, the shear stress is a function of µ, the velocity gradient, and
the flow properties. This occurs, because the eddies will cause a transport of momentum
through the boundary layer, and this transport is modelled as a shear stress:
∂u
τt = ρM (262)
∂y
where M is the eddy viscosity, which is due to the fluid motion.

• An analytical solution cannot be found for turbulent flows, due to the dependence of the
flow on the flow properties, therefore, experimental measurements are used to develop
empirical correlations for Cf and N u in turbulent flows.
Eng6901 - Heat Transfer I 52

• The local skin friction coefficient is evaluated by the following equation:

Cf,x = 0.0592Re−1/5
x Rex ≤ 107 (263)

and this equation may be used for Rex < 108 to within 15% accuracy.

• Using the Reynolds-Colburn analogy:

N ux = 0.0296Rex4/5 P r1/3 0.6 < P r < 60, Rex < 108 (264)

• For heating starting at a position ξ from the leading edge of the plate:

N ux = 0.0296Re4/5
x Pr
1/3
[1 − (ξ/x)9/10 ]−1/9 0.6 < P r < 60, Rex < 108 (265)

• For uniform wall heat flux:

N ux = 0.0308Rex4/5 P r1/3 0.6 ≤ P r ≤ 60 (266)

i.e. 4% higher than for the constant wall temperature boundary condition.

4.1.9 Mixed Boundary Layer Conditions


• A laminar boundary layer will eventually go through a transition to a turbulent boundary
layer. Define the plate length as L and the position where the critical Reynolds number,
Rex,c , occurs as xc . When 0.95 < xc /L ≤ 1, then the mean laminar convection coefficient,
Eq. (245), can be used to determine the total heat transfer rate from the plate. When
xc /L ≤ 0.95 then the laminar and turbulent sections of the boundary layer should be
accounted for when determining the total heat transfer rate from the plate:
Z xc Z L !
1
hL = hlam dx + hturb dx (267)
L 0 xc

where the transition is assumed to occur abruptly at xc .

• Substituting Eqs. (237) and (264) into Eq. (267) gives:


4/5
N uL = (0.037ReL − A)P r1/3 (268)

where
4/5 1/2
A = 0.037Rex,c − 0.664Rex,c (269)
is dependent upon the critical Reynolds number.

• The critical Reynolds number for flow over a smooth flat plate is 5 × 105 , therefore:
4/5
N uL = (0.037ReL − 871)P r1/3 0.6 < P r < 60, 5 × 105 < ReL ≤ 108 (270)

and
0.074 1742
Cf = 1/5
− 5 × 105 < ReL ≤ 108 (271)
ReL ReL
Eng6901 - Heat Transfer I 53

• If the boundary layer is completely turbulent (e.g. it is tripped at the leading edge of the
plate) A = 0:
4/5
N uL = 0.037ReL P r1/3 (272)
−1/5
Cf = 0.074ReL (273)
4/5
Note: these equations would also be appropriate when xc /L  1, since A  0.037ReL .

• All of the foregoing correlations are to be used with properties evaluated at the film
temperature.

• These correlations are acceptable for engineering calculations, but they may be up to 25%
in error due to freestream turbulence, and surface roughness.

4.1.10 How to Evaluate Convection Heat Transfer


• The calculation of convection heat transfer rates uses the following procedure:

1. Identify the geometry of the flow. All of the convection heat transfer correlations
are dependent upon the geometry involved.
2. Specify the reference temperature, and evaluate all fluid properties at this reference
temperature. Usually the film temperature, or a mean bulk temperature is used,
however, there are exceptions.
3. Calculate the Reynolds number. Is the flow laminar or turbulent, or both? Is there
anything to cause the flow to be completely turbulent, e.g. a very rough surface, or
freestream turbulence?
4. Determine the boundary condition on the surface, i.e. constant temperature or uni-
form flux.
5. Decide if local or mean values are required.
6. Pick an appropriate correlation based on the previous steps.

4.2 Flow Across Cylinders and Spheres


• The total drag force, FD , acting on a cylinder in a cross flow is a function of a frictional
component, due to the shear stress in the fluid at the surface of the cylinder, and a
component due to the pressure differential acting on the cylinder due to the formation of
the wake. These two drag components are called frictional and form (or pressure drag),
respectively.

• For ReD < 2 the flow remains attached to the cylinder, and the drag is mainly due to
friction.

• As the Reynolds number is increased boundary layer separation and wake formation
become important, and form drag dominates frictional drag.

• As the freestream is brought to rest at the stagnation point on the cylinder, a maximum
pressure is attained. As the flow expands about the cylinder, the pressure decreases,
and the boundary layer develops in a favourable pressure gradient (dp/dx < 0). As the
flow passes the maximum height of the cylinder, however, it will begin to decelerate, and
Eng6901 - Heat Transfer I 54

consequently the pressure will increase, producing an adverse pressure gradient (dp/dx >
0).

• Since the pressure in a boundary layer is constant at any x location, this adverse pressure
gradient will decelerate the flow within the boundary layer. If the flow has insufficient
momentum to overcome the adverse pressure gradient it will separate from the surface of
the cylinder and create a recirculation zone, or a wake. Thereafter, the pressure cannot
increase, and this gives rise to a large pressure differential between the front and back of
the cylinder → high form drag.

• A laminar boundary layer carries less momentum near the surface of the cylinder than a
turbulent boundary layer, therefore, it will separate earlier (θ = 80o ) than the turbulent
boundary layer (θ = 140o ) → higher form drag.
Eng6901 - Heat Transfer I 55

• The transition Reynolds number for a cylinder is approximately 2 × 105 .

• Figure 7.9, Incropera and DeWitt, above, shows the variation of N uD as a function of
angular position from the stagnation point on the cylinder, i.e. 0◦ . The Nusselt number
decreases from the stagnation point as the laminar boundary layer grows. Between 80◦
and 100◦ N u increases rapidly, due to the transition from laminar to turbulent flow. The
Nusselt number decreases as the turbulent boundary layer is established. The Nusselt
number increases as 140◦ is reached, due to boundary layer separation, and increased
mixing in the wake region.

• Due to the large changes that occur in the flow over a cylinder, depending on position
and Reynolds number, empirical correlations are used to determine overall heat transfer
coefficients for cylinders in cross flow.
Eng6901 - Heat Transfer I 56

• The correlation developed by Hilpert:

hD
N uD = = CRem
DP r
1/3
P r ≥ 0.6 (274)
k
is widely used for gases, and liquids. The properties used in this equation are evaluated
at the film temperature, and N uD and ReD are based on the characteristic dimension of
the cylinder, i.e. its diameter. The constants C and m are determined from Table 7.2,
Incropera and Dewitt.

• Equation (274) may also be used for cylinders with noncircular cross-sections when Table
7.3, Incropera and DeWitt, is used to define C and m.

• Zhukauskas developed the following correlation:


1/4
Pr

N uD = CRem
DP r
n
0.7 < P r < 500, 1 < ReD < 106 (275)
P rs
where C and m are obtained from Table 7.4, Incropera and DeWitt, and n = 0.37 for
P r ≤ 10 and n = 0.36 for P r > 10. All properties are evaluated at T∞ except for P rs .

• Churchill and Bernstein have developed the following correlation, which is valid for
ReD P r > 0.2:
1/2
" 5/8 #4/5
0.62ReD P r1/3 ReD

N uD = 0.3 +  1/4 1 + (276)
1 + (0.4/P r)2/3 282, 000

where all properties are evaluated at the film temperature.

• Note: these correlations may be in error by as much as 20% in engineering calculations.

• The behaviour of fluid flow about a sphere is similar to that about a cylinder, however,
the drag coefficient is reduced due to the three-dimensional nature of the flow.

• McAdams has proposed a correlation for the convection heat transfer coefficient on a
sphere in a gas:
N uD = 0.37Re0.6
D 17 < ReD < 70, 000 (277)
where all properties are evaluated at the film temperature.

• Whitaker has developed the following correlation:


1/4
µ

1/2 2/3
N uD = 2 + (0.4ReD + 0.06ReD )P r0.4 (278)
µs

which is valid for 0.71 < P r < 380, 3.5 < ReD < 7.6 × 104 , and 1 < µ/µs < 3.2. All
properties in this correlation are evaluated at T∞ except µs which is evaluated at Ts .

• Ranz and Marshall developed the following correlation for convection heat transfer from
freely falling liquid drops:
1/2
N uD = 2 + 0.6ReD P r1/3 (279)

• Note: Eqs. (278) and (279) reduce to N uD = 2 when ReD → 0, which is the value
obtained for conduction from the surface of a sphere in a stationary infinite medium.
Eng6901 - Heat Transfer I 57

5 Forced Convection: Internal Flow


5.1 Hydrodynamic Fundamentals
• Consider steady state flow in a circular tube of radius ro :

• The fully developed velocity profile is parabolic for laminar flow. The profile is flatter for
turbulent flow.

• The friction and heat transfer rate are highest in the developing flow region, and asymptote
to a constant value in the fully developed region.

• The length of the developing flow region (or hydrodynamic entry region), xf d,h , is a
function of the Reynolds number for laminar flows:
xf d,h
 
≈ 0.05ReD (280)
D lam

where
ρum D
ReD = (281)
µ
and um is a mean velocity of the flow. The transition Reynolds number, ReD,c , is 2300
for tube flow.

• The hydrodynamic entry length for turbulent flow is independent of Reynolds number:
xf d,h
 
10 ≤ ≤ 60 (282)
D turb

• To derive an expression for the laminar velocity profile in a circular tube of radius ro ,
consider the steady, incompressible flow of a constant property Newtonian fluid in the
tube. In the fully developed region v = 0, and ∂u/∂x = 0. The Navier-Stokes equations
written in Polar-Cylindrical co-ordinates can be solved to give:
" 2 #
1 dp 2 r
  
u=− r 1− (283)
4µ dx o ro

and the mean velocity is:


ro2 dp
um = − (284)
8µ dx
Eng6901 - Heat Transfer I 58

• Defining the Moody (or Darcy) friction factor as:


−(dp/dx)D
f= (285)
ρu2m /2
the pressure drop in a tube of length L can be written as:
ρu2m L u2
Z L
∆p = − f dx = f ρ m (286)
0 2D D 2
where L replaces dx.
• Substituting the expression for the mean velocity, Eq. (284), and the definition of the
Reynolds number into Eq. (285) gives:
64
f= (287)
ReD

• Experimental data must be used to determine the friction factor for turbulent flows in
rough tubes, i.e. the Moody diagram, Fig. 8.3, Incropera and DeWitt. The following
correlations exist for fully developed turbulent flow in smooth tubes:
−1/4
f = 0.316ReD ReD ≤ 2 × 104 (288)
−1/5 4
f = 0.184ReD ReD ≥ 2 × 10 (289)
−2 6
f = (0.79 ln ReD − 1.64) 3000 ≤ ReD ≤ 5 × 10 (290)

5.2 Thermodynamic Fundamentals


• Consider a fluid entering a tube with a uniform velocity and temperature.

• The temperature profile in the fully developed region depends on the boundary condition
applied at the tube wall (Ts = const, or qs00 = const)
• The thermal entry length for laminar flow is:
xf d,t
 
= 0.05ReD P r (291)
D lam

i.e. if P r > 1, xf d,h < xf d,t and vice versa. The thermal entry length for turbulent flow is
the same as the hydrodynamic entry length.
Eng6901 - Heat Transfer I 59

• A bulk (or mean) temperature is used in the calculation of heat transfer rates for internal
flows. It is determined from the rate of transport of thermal energy through a cross-
section, Ac : Z
Ėt = ṁcv Tm = ρucv T dAc (292)
Ac
i.e. R
Ac ρucv T dAc
Tm = (293)
ṁcv
• The heat flux is defined as:
qs00 = h(Ts − Tm ) (294)
Note: Tm is a function of x (Tm increases if Ts > Tm and decreases if Ts < Tm ). It will
be shown that h is constant in the fully developed region, therefore, if Ts is constant then
qs00 decreases with x, and if qs00 is constant then Ts − Tm is a constant.

• In the thermally fully developed region ∂T /∂x 6= 0, but:

∂ Ts (x) − T (r, x)
 
=0 (295)
∂x Ts (x) − Tm (x) f d,t

therefore:
∂ Ts − T −∂T /∂r|r=ro
 
= 6= f (x) (296)
∂r Ts − Tm r=ro
Ts − Tm
But
∂T
qs00 =k = h(Ts − Tm ) (297)
∂r r=ro
therefore
h
6= f (x) (298)
k
So in the thermally fully developed region the convection coefficient is constant.

• Since h is constant, the constant qs00 boundary condition gives:



dTs dTm
= (299)
dx f d,t
dx f d,t

so (Ts − Tm ) is constant.

• Applying the 1st Law to an element of fluid flowing in a tube (i.e. an open system of
constant volume fixed in space):
Eng6901 - Heat Transfer I 60

• Neglecting ∆ek , ∆ep , and all forms of work except for flow work:

dqconv = ṁdh (300)

and if the fluid is assumed to have constant specific heats:

dqconv = ṁcp dTm (301)

• This equation can be integrated from the inlet to the exit of the tube to give:

qconv = ṁcp (Tm,o − Tm,i ) (302)

• Defining dqconv = qs00 P dx, Eq. (301) can be rearranged to give:

dTm q 00 P P
= s = h(Ts − Tm ) (303)
dx ṁcp ṁcp

• For constant wall heat flux Eq. (303) can be integrated to give:

qs00 P
Tm (x) = Tm,i + x (304)
ṁcp

so the bulk temperature varies linearly with x. Note: Ts − Tm increases until the fully
developed region is reached (due to the higher h in the entrance region) and (Ts − Tm )
becomes constant.

• For the constant wall temperature boundary condition, Eq. (303) can be rewritten
as:
dTm d(∆T ) P
=− = h∆T (305)
dx dx ṁcp
where ∆T = Ts − Tm . Rearranging this equation, and integrating from inlet to outlet:
Z ∆To Z L
d(∆T ) P
=− h dx (306)
∆Ti ∆T ṁcp 0
Eng6901 - Heat Transfer I 61

or Z L !
∆To PL 1 PL
ln =− h dx =− h (307)
∆Ti ṁcp L 0 ṁcp
which can be rearranged to give:
!
∆To Ts − Tm,o PL
= = exp − h (308)
∆Ti Ts − Tm,i ṁcp

So the temperature difference (Ts − Tm ) decays exponentially.

• Using Eq. (307):


(Ts − Tm,o ) − (Ts − Tm,i )
qconv = ṁcp ((Ts − Tm,o ) − (Ts − Tm,i )) = −P Lh (309)
ln(∆To /∆Ti )
or
qconv = ṁcp (Tm,o − Tm,i ) = hA∆Tlm (310)
where
∆To − ∆Ti
∆Tlm = (311)
ln(∆To /∆Ti )
i.e. the log mean temperature difference.

5.3 Internal Flows: Correlations


5.3.1 Fully Developed Laminar Flow
• An analytical solution can be obtained for fully developed laminar flow in a circular tube
of diameter D:
hD
N uD = = 4.36 qs00 = const (312)
k
= 3.66 Ts = const (313)

Table 8.1, Incropera and DeWitt, lists N uD values for fully developed laminar flow in a
variety of noncircular cross-sections.
Eng6901 - Heat Transfer I 62

5.3.2 Laminar Flow: Entry Region


• For a thermal entry region occurring in a fully developed velocity field, Hausen obtained
the following relation for constant Ts :
hD 0.0668(D/L)ReD P r
N uD = = 3.66 + (314)
k 1 + 0.04[(D/L)ReD P r]2/3
which gives a mean h over the entry region. Note: this value asymptotes to the fully
developed value of N uD = 3.66.
• The correlation above is not generally applicable, since it assumes a fully developed veloc-
ity profile. For the combined entry region (i.e. hydrodynamic and thermal entry region)
Seider and Tate obtained the following relation:
1/3  0.14
ReD P r µ

N uD = 1.86 (315)
L/D µs
which is valid for Ts = const, 0.48 < P r < 16, 700, 0.0044 < (µ/µs ) < 9.75, and
ReD P r(D/L) > 10.
• The properties in both of these correlations are evaluated at the mean bulk temperature
of the fluid, except µs .

5.3.3 Turbulent Flow in Circular Tubes


• The Fanning friction coefficient is defined as:
τs
Cf = (316)
ρu2m /2
and the shear stress at the wall of a circular tube is:
du
 
τs = −µ (317)
dr r=ro

Using the fully developed (laminar) velocity profile, and the mean velocity obtained from
that profile:
f
Cf = (318)
4
Using the Chilton-Colburn analogy:
Cf f N uD
= = StP r2/3 = P r2/3 (319)
2 8 ReD P r
and the friction factor, Eq. (290):
4/5
N uD = 0.023ReD P r1/3 (320)

• The Dittus-Boelter correlation is the preferred form of the above correlation:


4/5
N uD = 0.023ReD P rn (321)

where n = 0.4 when Ts > Tm , and n = 0.3 when Ts < Tm . This correlation is valid for
0.7 ≤ P r ≤ 160, ReD ≥ 10, 000, and (L/D) ≥ 10. All properties should be evaluated at
Tm . This correlation should only be used for moderate Ts − Tm .
Eng6901 - Heat Transfer I 63

• When the temperature difference between the wall and the bulk fluid conditions becomes
large, there can be a large variation in properties in the fluid. For these conditions, Seider
and Tate have developed the following correlation:
0.14
µ

4/5
N uD = 0.027ReD P r1/3 (322)
µs
which is valid for 0.7 ≤ P r ≤ 16, 700, ReD ≥ 10, 000, and (L/D) ≥ 10. All properties
should be evaluated at Tm except µs .
• Both Eqs. (321) and (322) are valid for constant Ts and qs00 .
• To reduce the errors (which may be as large as 25%) that may be induced by Eqs. (321)
and (322), Petukhov developed the following correlation:
(f /8)ReD P r
N uD = (323)
1.07 + 12.7(f /8)1/2 (P r2/3 − 1)
which can produce errors of up to 10%. This correlation is valid for 0.5 < P r < 2000, and
104 < ReD < 5 × 106 , and for constant Ts and qs00 boundary conditions. The friction factor
may be obtained from a Moody diagram, or an appropriate smooth tube correlation. Fluid
properties are evaluated at the fluid bulk temperature, except for µs .
• Gnielinski modified the Petukhov correlation for use at lower Reynolds numbers:
(f /8)(ReD − 1000)P r
N uD = (324)
1 + 12.7(f /8)1/2 (P r2/3 − 1)
This correlation is valid for 0.5 < P r < 2000, and 3000 < ReD < 5 × 106 , and for constant
Ts and qs00 boundary conditions. The friction factor may be obtained from a Moody
diagram, or an appropriate smooth tube correlation. Fluid properties are evaluated at
the fluid bulk temperature, except for µs .

5.3.4 Turbulent Flow: Entry Region


• The turbulent entry region is usually small 10 < (xf d,t /D) < 60, and it is often acceptable
to assume that the fully-developed N uD can also be used in the developing region.
• Nusselt developed the following correlation:
D 0.055 L
 
4/5
N uD = 0.036ReD P r1/3 10 < < 400 (325)
L D
where the properties are evaluated at the mean bulk temperature of the fluid.

5.3.5 Flows in Noncircular Tubes


• The previously listed correlations may be used for tubes of noncircular cross-sections when
the diameter D is replaced by the hydraulic diameter, Dh :
4Ac
Dh = (326)
P
where Ac is the cross-section of the tube, and P is the wetted perimeter.
• Correlations for flow in a concentric tube annulus are given in Section 8.7, Incropera
and DeWitt, however, the Dittus Boelter correlation, Eq. (321), may be used with the
appropriate hydraulic diameter for fully developed turbulent flow in the annulus.
Eng6901 - Heat Transfer I 64

6 Radiation Heat Transfer


6.1 Introduction
• Radiation heat transfer occurs due to the propagation of electromagnetic waves from a
body at a temperature T .

• The existence of radiation heat transfer explains why:

– frost exists when the ambient air temperature is greater than 0◦ C;


– rooms with windows may be hotter than interior rooms that are served by the same
air conditioning system;
– black car interiors are hotter than light coloured interiors;
– dark cars rust faster than light cars;
– white is the favourite colour of car in Colorado;
– we associate blue with cold temperatures (even though a blue flame is hotter than a
red flame);
– we are here.

• As discussed earlier, the net rate of radiation heat transfer between two bodies is governed
by:
q = F FG Aσ(T14 − T24 ) (327)
where F is the emissivity of a surface (property of the surface), FG is the shape factor
(property of the geometry), and σ is the Stefan-Boltzmann constant. This section will
concentrate on the evaluation of F , FG , and the calculation of radiation heat transfer
rates between two or more bodies.

6.2 Physical Mechanism of Radiation


• Electromagnetic waves are propagated at the speed of light:

c = λν (328)

where λ is the wavelength, ν is the frequency, and c = 3 × 108 m/s is the speed of light.

• Electromagnetic waves exist for a wide range of wavelengths. Thermal radiation exists in
the wavelength range 0.1 µm < λ < 100 µm.

• Note: visible light occurs in the range 0.35 µm < λ < 0.75 µm, i.e. blue to red, therefore,
it is possible to “see” some thermal radiation. See Fig. 12.3, Incropera and DeWitt.

• The magnitude of radiation varies with wavelength, due to the nature of a surface and its
temperature, i.e. radiation is spectral.

• Radiation is also directional: it may be diffuse (equal intensity in all directions), or


specular (e.g. a mirror).

– Note: a mirror is specular only to visible light, it may be diffuse for other wavelengths.
Eng6901 - Heat Transfer I 65

– We will assume diffuse radiation (makes life easier), therefore, radiation will be
emitted and reflected equally in all directions.

• Three properties (functions of the material and its surface finish) of interest are the
absorptivity, α, reflectivity, ρ, and transmissivity, τ . Where α, ρ, and τ are the fractions
of radiation energy incident on a surface that are absorbed, reflected, and transmitted,
respectively, and:
α+ρ+τ =1 (329)

• Another property of interest is the emissivity, , which is the ratio of the energy emitted
by a surface to that emitted by an ideal radiator. It will be shown that α = .

• These four properties are also functions of wavelength, but they are usually assumed
constant within certain wavelength ranges to ease analysis.

• The ideal radiator is called a blackbody.

– A blackbody absorbs all incident radiation (i.e. α = 1).


– For a given temperature and wavelength, no body can emit more energy than a
blackbody.
– A blackbody is a diffuse emitter.
– They are called black, because they are black to the human eye (i.e. all visible light
would be absorbed). Nothing is truly black, but a surface may be black in certain
wavelength ranges (e.g. carbon black (soot) at 20◦ C).

6.2.1 The Planck Distribution


• From statistical thermodynamics, the energy density of radiation per unit volume, per
unit wavelength is:
2hc2
Iλ,b (λ, T ) = 5 hc/λkT (330)
λ (e − 1)
where h = 6.6256 × 10−34 J·s is Planck’s constant, and k = 1.3805 × 10−23 J/K is
Boltzmann’s, constant. The temperature T is in Kelvin, and c (m/s) is the speed of
light in a vacuum. The energy density has units W/m3 · µm. This energy density is the
intensity of radiation emitted by a blackbody.
Eng6901 - Heat Transfer I 66

• The spectral emissive power of a blackbody, Eλ,T (λ, T ), can be obtained from the intensity,
Iλ,b (λ, T ), by integrating over a hemisphere above a surface to give the energy emitted
per unit area, at a given wavelength and temperature:
C1
Eλ,T (λ, T ) = (331)
λ (e /λT
5 C2 − 1)

where C1 = 3.742 × 108 W·µm4 /m2 , and C2 = 1.439 × 104 µm·K, and the units of Eλ,b are
W/m2 · µm. Eλ,b is the energy density emitted at a given temperature and wavelength.

• A plot of Eλ,b for different temperatures gives.

• The magnitude of Eλ,b increases with increasing temperature, and the maximum value
shifts to shorter wavelengths, i.e. higher frequencies, as the temperature increases. The
wavelength with the maximum Eλ,b is found from Wien’s displacement law:

λmax T = 2897.8 µm · K (332)

• For sunlight (T = 5800K) λmax = 0.5 µm, which is in the middle of the visible light
region. For T = 1000K λmax = 2.8976 µm, which is in the infrared region, i.e. invisible,
however, some energy is emitted in the visible (red) region, see Fig. 12.12, Incropera and
DeWitt. Formula 1 brake disks have an operating temperature of 800◦ C, therefore, they
can be seen to glow red on dark days (when the ambient light intensity has been reduced
enough such that the radiation emitted by the disks may be seen).
Eng6901 - Heat Transfer I 67

• Figure 12.13 also explains why daylight colour film gives false colours (skewed to oranges
for incandescent light, and greens for flourescent light) when used inside without a flash.
• If the spectral blackbody emissive power (or density) is integrated over all wavelengths,
the result is the blackbody emissive power:
Z ∞
Eb = Eλ,b dλ = σT 4 (333)
0
where
2C1 π 5 W
σ= = 5.669 × 10−8 2 (334)
15C24 m · K4
is the Stefan-Boltzmann constant.
• Band Emission
– The fraction of the total energy emitted by a blackbody between wavelengths of 0
and λ is: Rλ Rλ
0 Eλ,b dλ Eλ,b dλ
F0−λ = R ∞ = 0 (335)
0 Eλ,b dλ σT 4
– Dividing the Planck distribution by T 5 :
Eλ,b C1
5
= 5 5 (336)
T λ T (exp(C2 /λT ) − 1)
i.e. Eλ,b /T 5 = f (λT ), and:
Z λT
Eλ,b
F0−λ = d(λT ) (337)
σT 50
is tabulated in Table 12.1, Incropera and DeWitt.
– To determine the amount of energy emitted between two wavelengths λ1 and λ2 :
Eb,λ1 −λ2 = σT 4 (F0−λ2 − F0−λ1 ) (338)
Eb,0−λ1 Eb,0−λ1
 
= σT 4 − (339)
σT 4 σT 4
Z λ2 T2 Z λ1 T1 !
4 Eλ,b Eλ,b
= σT d(λT ) − d(λT ) (340)
0 σT 5 0 σT 5

6.2.2 Emissivity
• Consider a “black” enclosure (i.e. it emits energy according to the Stefan-Boltzmann Law,
Eb = σT 4 ). Place a body inside the enclosure:
Eng6901 - Heat Transfer I 68

• If qi is the radiant flux (W/m2 ) arriving at the surface of the body, then the body will
absorb qi As α (W).

• When the body reaches thermal equilibrium with the enclosure (energy emitted = energy
absorbed):
EAs = qi As α (341)

• If the body is replaced by a blackbody of the same size and shape, and it attains thermal
equilibrium with the enclosure:
Eb As = qi As (342)
since α = 1 for a blackbody.

• Then:
E
==α (343)
Eb
i.e. the ratio of the emissive power of a body to the emissive power of a blackbody is the
emissivity of the body, and  = α. This equation is Kirchoff’s Law.

• In general,  = (λ, T ), and the monochromatic or spectral emissivity is defined as:



λ = (344)
Eλ,b

• For a blackbody  = 1. For a gray body the emissivity is assumed to be independent of


wavelength, i.e. λ = .

•  depends on the material, surface finish, λ and T .

– 0.039 <  < 0.057 polished Aluminium (440K < T < 1070K)
Eng6901 - Heat Transfer I 69

– 0.20 <  < 0.31 oxidized aluminium (299K < T < 940K).
– The  of an oxidized surface is much larger than  for a clean surface.
–  = 0.906 for snow-white enamel on a rough iron plate.
– 0.80 <  < 0.95 for black or white lacquer

• e.g. A solar collector utilizes special coatings to effectively give α = 1 for sunlight and
 = 0 for infrared radiation (i.e. the emissivity is effectively zero in the wavelength range
where the maximum emissive power of the collector is located).
Eng6901 - Heat Transfer I 70

• Many surfaces behave as gray bodies, i.e.  = constant within certain wavelength ranges.
If the incident or emitted radiation has the highest energy density between λ1 and λ2
shown below, then the body is behaving as an approximate gray body.

• A blackbody is an idealization, however, effective blackbodies can be created:

– An intrinsic blackbody (due to the material and its surface finish) behaves as a
blackbody for a given temperature and wavelength range (e.g. a black surface at
room temperature in the visible light region (0.35 µm < λ < 0.7 µm)).
– A geometric blackbody may be created by constructing an enclosure with a small
hole. Any energy that is radiated in at the hole will be absorbed as the beam is
reflected around the enclosure before it escapes.

• Emissivity/absorptivity help explain why black car interiors are hotter than light coloured
interiors, and why dark cars rust faster.
Eng6901 - Heat Transfer I 71

6.3 Radiation Shape (or View) Factor


• To begin, consider the radiant heat exchange between two blackbodies (i.e. α =  = 1).

• What is the rate of energy exchange between the two bodies shown above when they are
maintained at T1 and T2 ? i.e. How much energy leaves 1 and reaches 2, and how much
energy leaves 2 and reaches 1?

• This calculation will require the radiation shape factor, Fij , which is the fraction of energy
leaving body i that reaches body j.

• Using the shape factor, the rate of energy leaving 1 and reaching 2 is:

Eb1 A1 F12 (345)

and the rate of energy leaving 2 and reaching 1 is:

Eb2 A2 F21 (346)

• Since bodies 1 and 2 are both blackbodies, all incident radiation will be absorbed, and
the net rate of energy transfer is:

q12 = Eb1 A1 F12 − Eb2 A2 F21 (347)

Note: the signs, directions, and nomenclature.

• If T1 = T2 then q12 = 0 and:


Eb1 A1 F12 = Eb2 A2 F21 (348)
therefore
A1 F12 = A2 F21 (349)
Equation (349) is the reciprocity relation. This relation is valid for any diffuse emitter,
and is used extensively in radiation shape factor algebra. In general:

Ai Fij = Aj Fji (350)

So
q12 = A1 F12 (Eb1 − Eb2 ) = A2 F21 (Eb1 − Eb2 ) (351)
Eng6901 - Heat Transfer I 72

• The problem is to determine F12 or F21 . Fortunately, the shape factors for many common
geometries are tabulated, and shape factor algebra can be used to obtain other shape
factors from the tabulated data.
• Assume diffuse radiation from the two areas shown in the previous figure. The energy
emitted by surface dA1 in the direction of φ1 is:
Ib dA1 cos φ1 (352)
where Ib is the blackbody intensity (radiation emitted per unit area, per unit solid angle)

• The radiation that reaches an area element dAn at a distance r from A1 is:
dAn
Ib dA1 cos φ1 (353)
r2
where dAn /r2 is the solid angle subtended by area dAn (normal to r). See Fig. 12.6,
Incropera and DeWitt. The solid angle is:
dAn = r dφ r dψ sin φ = r2 sin φ dφ dψ (354)
But the total emitted power (or hemispherical emitted power) Eb dA1 is:
Z 2π Z π/2
Eb dA1 = Ib dA1 dψ sinφ cos φ dφ (355)
0 0
= πIb dA1 (356)
Eng6901 - Heat Transfer I 73

or
Eb
Ib = (357)
π
• Considering the radiant energy exchange between the two bodies, define:
dAn = cos φ2 dA2 (358)
i.e. the projection of dA2 onto the perpendicular to r.
• The energy leaving 1 and reaching 2 is:
dA1 dA2
dq1→2 = Eb1 cos φ1 cos φ2 (359)
πr2
and the energy leaving 2 and reaching 1 is:
dA2 dA1
dq2→1 = Eb2 cos φ2 cos φ1 (360)
πr2
The net rate of heat energy exchange is:
dA1 dA2
Z Z
q12 = (Eb1 − Eb2 ) cos φ1 cos φ2 (361)
A2 A1 πr2
i.e. the integral is A1 F12 (or A2 F21 ), which can be used to determine the rate of radiation
heat transfer from surface 1 to surface 2:
q12 = A1 F12 σ(T14 − T24 ) (362)
This expression is for radiant heat exchange between blackbodies, for gray bodies an
appropriate  is required.
• Shape factors are tabulated in Tables 13.1 and 13.2, and Figs. 13.4, 13.5 and 13.6 of
Incropera and DeWitt. Thermal Radiation Heat Transfer, by R. Siegel and J.R. Howell
tabulates many other shape factor relations.

6.3.1 Shape Factor Algebra


• Many shape factors for common geometries may be found from tabulated data. Shape
factors for more complicated geometries may be developed from the tabulated data and
the use of some basic relations, such as reciprocity:
Ai Fij = Aj Fji (363)
and n
X
Fij = 1 (364)
j=1
e.g. for three surfaces:
F11 + F12 + F13 = 1 (365)
Shape factors may also be split up, e.g.:
F3−1,2 = F31 + F32 (366)

• A common example used to demonstrate shape factor algebra is perpendicular plates.


Figure 13.6 (see Appendix D) and Table 13.2 give the shape factor Fij as a function of
Z/X and Y /X. Determine the shape factor F14 for surfaces 1 and 4 shown in the figure
below:
Eng6901 - Heat Transfer I 74

– Shape factors F32 , F3,4−2 , F3,4−1,2 and F3−1,2 can be found from Fig. 13.6 or Table
13.2 (or vice versa using reciprocity). Then:

A1 F14 = A1,2 F1,2−3,4 − A2 F2−3,4 − A1,2 F1,2−3 + A2 F23 (367)

– i.e. try to write the desired shape factor as a function of known factors. Here, 1 can
see 3 and 4, and 2 can see 3 and 4, so eliminate 1-3, 2-3, and 2-4 from the total.
– This is the negative view: start with the whole and eliminate unnecessary factors.
– Be careful of situations where Fii 6= 0 (concave surfaces).
– An alternate method is to be positive, and construct the total from the sum of its
parts.
A1,2 F1,2−3,4 = A1 F1−3,4 + A2 F2−3,4 (368)
and
A1 F1−3,4 = A1 F13 + A1 F14 (369)
but
A1,2 F1,2−3 = A1 F13 + A2 F23 (370)
therefore
A1,2 F1,2−3,4 = A1,2 F1,2−3 − A2 F23 + A1 F14 + A2 F2−3,4 (371)
or
A1 F14 = A1,2 F1,2−3,4 − A2 F2−3,4 − A1,2 F1,2−3 + A2 F23 (372)
Eng6901 - Heat Transfer I 75

6.4 Radiation Heat Exchange between Nonblackbodies


• Once the shape factor has been determined, the evaluation of radiation heat transfer is
easy for blackbodies. In general, however,  6= 1 (then ρ 6= 0 and τ 6= 0) and things
become more complicated due to reflections between surfaces or one (concave) surface.

• The following assumptions will be used to simplify the analysis of radiation heat transfer
between nonblackbodies:

1. diffuse radiation;
2. uniform surface temperatures;
3.  and ρ are constant over a surface; and
4. τ = 0.

• Define the irradiation, G, which is the total radiation incident on a surface per unit time
per unit area. Irradiation is assumed to be uniform over a surface.

• Define the radiosity, J, as the total radiation that leaves a surface per unit time per unit
area, i.e. the amount of energy emitted and reflected. Radiosity is assumed to be uniform
over a surface. The radiosity is:
J = Eb + ρG (373)
Since τ is assumed to be zero, and α = , then 1 = ρ + τ + α gives:

ρ=1− (374)

and the radiosity is defined as:

J = Eb + (1 − )G (375)

• The net rate of energy leaving a surface is:

q 00 = J − G (376)
= Eb + (1 − )G − G (377)

but from Eq. (375):


J − Eb
G= (378)
1−
then
J − Eb
q 00 = Eb + (J − Eb ) − (379)
1−
J − J − J + Eb
= (380)
1−
(Eb − J)
= (381)
1−
Therefore, the total rate of energy transport is:
Eb − J
q= (382)
(1 − )/A
Eng6901 - Heat Transfer I 76

i.e. a net loss if Eb > J. Note the similarity with q = potential/Rth . A “surface” resistance
can be defined as:
1−
(383)
A
which will allow calculation of heat loss (or gain) from a surface.

• What about the energy exchange between two surfaces? The amount of energy leaving 1
and reaching 2 is:
J1 A1 F12 (384)
and the amount of energy leaving 2 and reaching 1 is:

J2 A2 F21 (385)

then
q12 = J1 A1 F12 − J2 A2 F21 (386)
and using reciprocity (A1 F12 = A2 F21 ):

q12 = A1 F12 (J1 − J2 ) = A2 F21 (J1 − J2 ) (387)

or
J1 − J2 Potential
q12 = = (388)
1/A1 F12 “Space”resistance

• Note: the space resistance used here is purely due to the shape factor (i.e. geometry), it
does not include any resistance due to a participating medium (e.g. a gas) between the
two surfaces.

• Two resistance elements are defined to be used in thermal radiation network analyses
Eng6901 - Heat Transfer I 77

• e.g. Two surfaces exchanging heat with each other and nothing else.

• So
Eb1 − Eb2
q12 = 1−1 1−2 (389)
1 A1 + A11F12 + 2 A2
where
Eb1 − Eb2 = σ(T14 − T24 ) (390)

• e.g. Three bodies exchanging energy with each other and nothing else.

J1 − J2
q12 = 1 (391)
A1 F12

J1 − J3
q13 = 1 (392)
A1 F13
J2 − J3
q23 = 1 (393)
A2 F23
Eng6901 - Heat Transfer I 78

• Note: even if the three temperatures are known, three equations are still required to
solve for the unknown J’s before the heat transfer rates may be solved. How to obtain
the equations? Place a control volume boundary around each node and use the 1st Law
to give: heat transfer in = heat transfer out, or sum of heat transfer in = 0. Note the
similarity to electric circuits: current in = current out, or sum of current in = 0.

• For node 1: q1 = q12 + q13


Eb1 − J1 J1 − J2 J1 − J3
1−1 = 1 + 1 (394)
1 A1 A1 F12 A1 F13

• For node 2: q2 = q21 + q23


Eb2 − J2 J2 − J1 J2 − J3
1−2 = 1 + 1 (395)
2 A2 A2 F21 A2 F23

• For node 3: q3 = q31 + q32


Eb3 − J3 J3 − J1 J3 − J2
1−3 = 1 + 1 (396)
3 A3 A3 F31 A3 F32

Note:

q1 − q12 − q13 = 0 (397)


q1 + q21 + q31 = 0 (398)

• e.g. Four bodies exchanging heat with each other and nothing else.
Eng6901 - Heat Transfer I 79

• For node 1: q1 = q12 + q13 + q14


Eb1 − J1 J1 − J2 J1 − J3 J1 − J4
1−1 = 1 + 1 + 1 (399)
 1 A1 A1 F12 A1 F13 A1 F14

• For node 2: q2 = q21 + q23 + q24


Eb2 − J2 J2 − J1 J2 − J3 J2 − J4
1−2 = 1 + 1 + 1 (400)
 2 A2 A2 F21 A2 F23 A2 F24

• For node 3: q3 = q31 + q32 + q34


Eb3 − J3 J3 − J1 J3 − J2 J3 − J4
1−3 = 1 + 1 + 1 (401)
 3 A3 A3 F31 A3 F32 A3 F34

• For node 4: q4 = q41 + q42 + q43


Eb4 − J4 J4 − J1 J4 − J2 J4 − J3
1−4 = 1 + 1 + 1 (402)
 4 A4 A4 F41 A4 F42 A4 F43

Solve for the unknowns.

• e.g. Two surfaces exchanging heat connected by a third surface which is perfectly insulated
(or reradiating).

• There is no surface resistance on surface 3, since no heat is removed (or added) by surface
3 (perfect insulation), therefore, everything that is absorbed by 3 is reradiated back to
the other surfaces. The temperature of surface 3 effectively floats. This circuit can be
solved using the three nodal equations (394) through (396), where the LHS of Eq. (396)
is set equal to zero. But, since all heat leaving one reaches 2 (or vice versa):

Eb1 − Eb2
q12 = 1−1 1 1−2 (403)
1 A1 + A1 F12 +1/(1/A1 F13 +1/A2 F23 ) +  2 A2
Eng6901 - Heat Transfer I 80

• e.g. Two surfaces with given temperatures, exchanging heat by radiation in a large enclo-
sure with a known surface temperature.

• The surface resistance on 3 reduces to zero, because of the large area of 3. The space
resistance is eliminated, and J3 is replaced by Eb3 . Only equations for nodes 1 and 2, Eqs.
(394) and (395), are required to solve for this problem (J1 and J2 are the only unknowns.

• Note: the similarity between the circuit for this problem and the previous example. The
two problems cannot be solved in the same manner, however, because heat may be re-
moved (or added) by surface 3 when it is a blackbody, but no heat may be added (or
removed) by surface 3 when it is perfectly insulated.
Eng6901 - Heat Transfer I 81

6.4.1 Radiation Shields


• Consider radiation heat exchange between two gray infinite plates of known temperature

• The radiation heat flux between surfaces 1 and 2 is:


σ(T 4 − T 4 )
q 00 = 1−1 11−2 2 1 (404)
1 + 2 + F12
σ(T14 − T24 )
= (405)
1/1 + 1/2 − 1
(406)
since F12 = 1 (infinite plates).
• If a third infinite plate is placed between the original two plates:

• Assuming 1 = 2 = 3 :
σ(T14 − T24 )
q 00 = 1− (407)
 + 1 + 1− 1−
 +  +1+
1−

σ(T14 − T24 )
= (408)
4/ − 2
Eng6901 - Heat Transfer I 82

which is 1/2 of the flux with no shield.

• If n shields are used between the two surfaces, this will give n + 1 space resistances, and
2n + 2 surface resistances, so:
X
00 1−
Rth = (2n + 2) + (n + 1)(1) (409)
 
2

= (n + 1) −1 (410)


with no shield (n = 0):


X
00 2
Rth = −1 (411)

Therefore, the radiation heat flux with n shields is 1/(n + 1) of the flux without shields
when the overall temperature difference is fixed. Examples of radiation shields are the heat
shields used above catalytic converters on cars, and the shields on motorcycle exhausts.
Radiation shields are a relatively cheap and easy means of decreasing heat transfer rates,
and temperatures.
Eng6901 - Heat Transfer I 83

7 Transient Heat Conduction


7.1 Introduction
• Thus far only steady state heat transfer, or equilibrium conditions, have been studied. If
a body is subjected to a change in thermal environment, a certain amount of time will be
required for it to reach a new thermal equilibrium.

• e.g. How long does it take to chill a can of beer? How can that time be reduced?

• e.g. How long must one wait before drinking a British beer after it has been removed from
a refrigerator?

• e.g. How long will it take to thaw a piece of frozen chicken? How can that time be
reduced?

• To study transient behaviour, the rate of change of internal energy of the body must
be accounted for in the governing equation, and the boundary conditions must be mod-
ified appropriately to match the new thermal environment. For a material with uniform
thermal conductivity the governing equation may be written as follows:
q̇ 1 ∂T
∇2 T + = (412)
k α ∂t
where α = k/ρc is the thermal diffusivity.

• Note: if α is small then ∂T /∂t is small, and vice versa. If ρc ↑, then α ↓, and ∂T /∂t ↓,
because the material has a high capacity to store thermal energy (i.e. large thermal mass).
If k ↓, then α ↓, and ∂T /∂t ↓, because the material will have a high resistance to heat
conduction, therefore, more time will be required to conduct a given amount of heat. If
two materials with different α’s are subjected to the same variable thermal environment,
the material with the lower α will require more time to heat up, or cool down (e.g. compare
water and air (α = 1.5 × 10−7 and 2.2 × 10−5 m2 /s, at 300K, respectively)).

• Consider one-dimensional (Cartesian), unsteady, constant and uniform property, no source


heat conduction in a material:
∂2T ∂T
k 2 = ρc (413)
∂x ∂t
with the following specified temperature boundary conditions:

1. T = Ti at t = 0, or T (x, 0) = Ti
2. T = T1 for x = x1 , t > 0, or T (x1 , t) = T1
3. T = T2 for x = x2 , t > 0, or T (x2 , t) = T2

• In general, there are five types of boundary conditions:

1. specified temperature;
2. specified flux;
3. convection heat transfer;
4. radiation heat transfer; and
Eng6901 - Heat Transfer I 84

5. convection and radiation heat transfer.

The boundary conditions may also change with time → numerical solution. Only constant
boundary conditions will be considered here.

• Solution methods:

1. Analytical–separation of variables, Laplace transforms → limited applications due


to geometry and boundary condition restrictions
2. Lumped Capacity Analysis–the temperature of a body is assumed to be uniform at
any instant in time (simple but very useful method)
3. Heisler Charts–simpified analytical solutions presented in chart form (permits eval-
uation of multidimensional transient temperature distributions without the tedium
of evaluating series solutions)
4. Numerical–finite element, finite difference, finite volume, boundary element methods
(permits simulation of problems with complex geometries and boundary conditions,
in materials with time (or temperature) variant properties)

7.2 Lumped Capacity Analysis


• Idealized method based on the assumption that the temperature of a body is uniform at
any instant in time. This is not strictly correct, since for heat to be transferred to or from
a body a temperature gradient must exist within the body (from Fourier’s Law).

• The analysis is developed in the context of a body exposed to a convection environment,


i.e. the body is heated or cooled by a fluid.

• The justification for using lumped capacity analysis (LCA) is if the convection resistance
on the surface of the body is much larger than the conduction resistance within the body,
then the major portion of the temperature change occurs in the fluid

• e.g. Consider a thin copper sheet with a fixed temperature on one side, and the other side
is exposed to a convection environment.

To − T∞
q 00 = (414)
L/k + 1/h
Eng6901 - Heat Transfer I 85

if k ∼ 380 W/m·◦ C, L ∼ 10−3 m, and h ∼ 35 W/m2 ·◦ C (air at 35 m/s) then L/k =


2.6 × 10−6 and 1/h = 2.9 × 10−2 , i.e. the resistance in the copper is negligible compared
to that in the air, therefore, it may be neglected. This implies that the temperature in
the copper may be assumed uniform at To , and the temperature drop To − T∞ occurs in
the fluid.

• LCA assumes the internal resistance of the body is negligible compared to the external
resistance.

• With this assumption, an unsteady energy balance for a body exposed to a convection
environment can be written as follows:
dT
hAs (T − T∞ ) = −ρcV (415)
dt
i.e. the rate of convection heat transfer from the body is equivalent to the time rate of
decrease of internal energy within the body (1st Law).

• The initial condition of the body is: T = Ti at t = 0 (or T (0) = Ti ).

• Defining:
θ = T − T∞ (416)
the initial condition becomes: θ(0) = θi , and the energy balance can be written as:


hAs θ = −ρcV (417)
dt
which has solution:
hAs
 
θ = C1 exp − t (418)
ρcV
Applying the initial condition gives C1 = θi , so the solution to the transient temperature
of the body is:
θ T − T∞ hAs
 
= = exp − t (419)
θi Ti − T∞ ρcV
• Note:

1. As hAs /ρcV increases the body will cool (or heat) faster, therefore, to accelerate
cooling/heating: increase h and As (decrease convection resistance); decrease ρ, c
and V (decrease the amount of stored energy).
2. When comparing two bodies exposed to the same thermal environment a comparison
of hAs /ρcV values can be used to identify the body that will cool/heat quicker.

• Applicability of the Lumped Capacity Analysis

– Valid when the surface convection resistance is large relative to the internal convec-
tion resistance, which begs the question how large is large?
Eng6901 - Heat Transfer I 86

– Consider one-dimensional steady state constant property conduction through a plane


wall exposed to a convection environment.

kA
(Ts,1 − Ts,2 ) = hA(Ts,2 − T∞ ) (420)
L
or writing the ratio of the temperature drops:

Ts,1 − Ts,2 L/kA Rcond hL


= = = (421)
Ts,2 − T∞ 1/hA Rconv k

For LCA to apply, (Ts,1 − Ts,2 ) << (Ts,2 − T∞ ), i.e. Rcond << Rconv , or hL/k < 1.
– It has been found that LCA is applicable when:
hLc
Bi ≡ < 0.1 (422)
k
Where Lc = V /A is a characteristic dimension used in the Biot number, Bi.
– Note: When solving a transient problem, check Bi. It shows understanding, and it
can considerably simplify analyses.

• Note: the similarity of the LCA solution with electric circuits.


Eng6901 - Heat Transfer I 87

– Note: hAs /ρcV = 1/τ where τ is the time constant.

hAs 1 1
= = (423)
ρcV (1/hAs )(ρcV ) Rth Cth

i.e.
θ t
 
= exp − (424)
θi Rth Cth
where Rth is the thermal resistance, and Cth is the thermal mass (or capacitance),
and the product is a time constant (RC) just as in electric circuits.

• Note: Rth can be replaced with


P
Rth , e.g. a copper sphere surrounded by asbestos
exposed to a natural convection environment. Here RCu << (RAs + Rconv ), and LCA
would apply.

• Note: What if convection and radiation heat transfer occur from the surface of a body?

– The energy balance would be:


∂T
qrad + qconv = −ρcV (425)
∂t
or
∂T
σAs (T 4 − Tsur
4
) + hAs (T − T∞ ) = −ρcV (426)
∂t
– If Tsur = T∞ , then the radiation coefficient, hr , can be used:
∂T
(hr + h)As (T − T∞ ) = −ρcV (427)
∂t
or an effective heat transfer coefficient can be defined, hef f = h+hr , and the transient
temperature in the body would be given by:
θ T − T∞ hef f As
 
= = exp − t (428)
θi Ti − T∞ ρcV
which would apply when:
hef f Lc
Bi = < 0.1 (429)
k
– Note: hr = f (T ), therefore, it would vary with time, and the quality of solution
would depend upon the temperature used to evaluate hr .

• The total amount of heat lost (or gained) by a body during a time interval may be found
as follows:
Z t Z t
t
  
Q= q dt = hAs θ dt = (ρcV )θi 1 − exp − (430)
0 0 Rth Cth
Eng6901 - Heat Transfer I 88

A Eulerian and Lagrangian Viewpoints


• The field of motion of a fluid is described in terms of the velocities (~u), and accelerations
(~a) of fluid particles at various positions in the fluid space.

• The two methods of studying the motion of groups of particles in a continuum are the
Lagrangian and Eulerian viewpoints.

A.1 Lagrangian View


• The co-ordinates of a moving particle are functions of time, i.e. attention is focussed on
a particle and this particle is followed in space.

x = x(a, b, c, t) (431)
y = y(a, b, c, t) (432)
z = z(a, b, c, t) (433)

Where a, b, and c are used to define the co-ordinates of the particle at an initial time t0 .

• Since the position of the particle is a function of time only, the velocity and acceleration
of the particle can be defined as follows:
dx~ dy ~ dz ~
~u = i+ j+ k (434)
dt dt dt
d2 x~ d2 y ~ d2 z ~
~a = i+ 2j+ 2k (435)
dt2 dt dt

• The Lagrangian view is used in Mechanics, Solid Mechanics, and Kinematics.

• The number of particles that would have to be tracked to give an adequate description
of a flow field makes the Lagrangian view unwieldy for fluid flow. Further, one is usually
interested in the velocity, acceleration, temperature, and pressure at certain locations in
a fluid flow field, rather than the motion of a particular particle.

A.2 Eulerian View


• An alternate means of obtaining a description of a flow field is to focus attention on a
fixed point in space, and observe the flow characteristics at that point as particles pass
by. When enough points are used, an instantaneous picture of ~u and ~a throughout the
flow field is obtained.

• In the Eulerian view x, y, and z are independent variables (i.e. the location of a fixed
point in space), whereas in the Lagrangian view x, y, and z are dependent variables (i.e.
the location of a moving particle in space at a time t).

• In the Eulerian view, the velocity and acceleration are functions of position and time:

~u = ~u(x, y, z, t) (436)
~a = ~a(x, y, z, t) (437)
Eng6901 - Heat Transfer I 89

• The change in the x-component of velocity, u, is:


∂u ∂u ∂u ∂u
du = dx + dy + dz + dt (438)
∂x ∂y ∂z ∂t
Dividing by dt:
du ∂u dx ∂u dy ∂u dz ∂u dt
= + + + (439)
dt ∂x dt ∂y dt ∂z dt ∂t dt
gives the total, substantial or material derivative:
Du ∂u ∂u ∂u ∂u
= +u +v +w (440)
Dt ∂t ∂x ∂y ∂z
which is the total rate of change of the x-component of velocity of a particle as it passes
by a fixed point in space.

• The first term on the R.H.S. of Eq. (440), ∂u/∂t, is called the local acceleration, and the
three remaining terms are the convective change, i.e. the rate of change due to motion of
a particle into an area of higher or lower velocity.

• The acceleration may be written in vector notation as follows:

D~u ∂~u  ~ 
~a = = + ~u · ∇ ~u (441)
Dt ∂t

• A flow is defined to be steady when all local accelerations are zero, i.e. ∂~u/∂t = 0.
 
~ ~u = 0, which
• A flow is uniform when all convective accelerations are zero, i.e. ~u · ∇
would mean the flow is parallel.

B Conservation of Mass (Continuity Equation)


• Consider fluid flow through a control volume of dimensions ∆x, ∆y, and ∆z that is fixed
in space.
Eng6901 - Heat Transfer I 90

• Application of the principle of mass conservation to this control volume gives:


     
Total rate at which Total rate at which Net rate of decrease
mass exits a cv  −  mass enters a cv  =  of mass within the  (442)
     

across its boundary across its boundary control volume
Eng6901 - Heat Transfer I 91

or
∂(ρu) ∂(ρv)
   
ρu + ∆x ∆y∆z − ρu∆y∆z + ρv + ∆y ∆x∆z
∂x ∂y
∂(ρw) ∂ρ
 
−ρv∆x∆z + ρw + ∆z ∆x∆y − ρw∆x∆y = − ∆x∆y∆z (443)
∂z ∂t
Cancelling like terms and dividing by volume gives the continuity equation:

∂ρ ∂(ρu) ∂(ρv) ∂(ρw)


+ + + =0 (444)
∂t ∂x ∂y ∂z
which may be written using the total derivative of ρ:
Dρ ~ · ~u = 0
+ ρ∇ (445)
Dt
or in the following form:
∂ρ ~
+ ∇ · (ρ~u) = 0 (446)
∂t
If the flow is steady:
~ · (ρ~u) = 0
∇ (447)
And if the flow is incompressible the continuity equation reduces to:
~ · ~u = 0
∇ (448)
Eng6901 - Heat Transfer I 92

C Conservation of Momentum (Navier-Stokes Equations)


• The Navier-Stokes equations define the motion of all fluids. They are a direct result of
the application of Newton’s Second Law to fluids.

• The Navier-Stokes equations may be derived using one of the following views of Newton’s
Second Law:

1. The acceleration of a fluid particle of mass ∆m results from the application of body
and surface forces.
2. The rate of change in momentum of the flow through a control volume is caused by
body and surface forces.

• Using view (1):


∆m~a = F~surface + F~body (449)
where the acceleration is defined as follows in the Eulerian view:
D~u ∂~u  ~ 
~a = = + ~u · ∇ ~u (450)
Dt ∂t
therefore:
D~u
∆m = F~surface + F~body (451)
Dt
• To use view (2), consider the rate of change of momentum as fluid flows through a control
volume of dimensions ∆x, ∆y, and ∆z that is fixed in space.
Eng6901 - Heat Transfer I 93

• Note: only fluxes of x momentum are shown.

• The mass of this control volume will be:

∆m = ρ∆x∆y∆z (452)

• The rate of change in x momentum in the control volume is:


     
Total rate at which Total rate at which Rate of change
 momentum exits a cv  −  momentum enters a cv  +  of momentum within the 
     
across its boundary across its boundary control volume
(453)
or
∂(ρu) ∂u
  
ρu + ∆x u+ ∆x ∆y∆z − ρuu∆y∆z
∂x ∂x
∂(ρv) ∂u
  
+ ρv + ∆y u+ ∆y ∆x∆z − ρvu∆x∆z
∂y ∂y
∂(ρw) ∂u
  
+ ρw + ∆z u+ ∆z ∆x∆y − ρwu∆x∆y
∂z ∂z
∂(ρu)
+ ∆x∆y∆z (454)
∂t
Cancelling like terms and higher-order terms (e.g. ∆x2 ∆y∆z), and dividing by volume:

∂(ρu) ∂(ρu) ∂(ρv) ∂(ρw) ∂u ∂u ∂u


+u +u +u + ρu + ρv + ρw (455)
∂t ∂x ∂y ∂z ∂x ∂y ∂z
The first term can be written as:
∂ρ ∂u
u+ρ (456)
∂t ∂t
Using the continuity equation, Eq. (444), the rate of change of x momentum per unit
volume is:
∂u ∂u ∂u ∂u Du
ρ + ρu + ρv + ρw =ρ (457)
∂t ∂x ∂y ∂z Dt
and view (2) gives:
D~u
ρ∆x∆y∆z = F~surface + F~body (458)
Dt
• Note: both views (1) and (2) give the same equation, i.e. Newton’s Second Law.

• The body forces are attributed to the mass of the body, e.g. gravity and coriolis forces
(only gravity forces will be considerd here):

F~body = (ρ∆x∆y∆z)~g (459)

where
~g = gx~i + gy~j + gz~k (460)
The x-component of the gravity force is:

Fx,body = (ρ∆x∆y∆z)gx (461)


Eng6901 - Heat Transfer I 94

• The surface forces are due to the normal and shear stresses on the surfaces of the control
volume:

∂σx ∂σy ∂σz


σx0 = σx + ∆x σy0 = σy + ∆y σz0 = σz + ∆z
∂x ∂y ∂z

0 ∂τxy 0 ∂τxz 0 ∂τyx


τxy = τxy + ∆x τxz = τxz + ∆x τyx = τyx + ∆y
∂x ∂x ∂y
0 ∂τyz 0 ∂τzx 0 ∂τzy
τyz = τyz + ∆y τzx = τzx + ∆z τzy = τzy + ∆z (462)
∂y ∂z ∂z
• The net surface force acting in the x direction is:
∂σx
 
Fx,surface = σx + ∆x ∆y∆z − σx ∆y∆z
∂x
∂τyx
 
+ τyx + ∆y ∆x∆z − τyx ∆x∆z
∂y
∂τzx
 
+ τzx + ∆z ∆x∆y − τzx ∆x∆y
∂z
∂σx ∂τyx ∂τzx
= ∆x∆y∆z + ∆x∆y∆z + ∆x∆y∆z (463)
∂x ∂y ∂z

• Substituting Eqs. (461) and (463) into Eq. (458) and dividing by the volume of the control
volume gives:
Du ∂σx ∂τyx ∂τzx
ρ = + + + ρgx (464)
Dt ∂x ∂y ∂z
Eng6901 - Heat Transfer I 95

Similarly, in the y and z directions:


Dv ∂σy ∂τxy ∂τzy
ρ = + + + ρgy (465)
Dt ∂y ∂x ∂z
Dw ∂σz ∂τxz ∂τyz
ρ = + + + ρgz (466)
Dt ∂z ∂x ∂y
• The stress-rate of strain relations for a Newtonian fluid are defined in Handout 1:
∂u 2 ~ ∂v ∂u
 
σx = −p + 2µ − µ(∇ · ~u) τxy = τyx = µ +
∂x 3 ∂x ∂y
∂v 2 ~ ∂w ∂v
 
σy = −p + 2µ − µ(∇ · ~u) τyz = τzy = µ +
∂y 3 ∂y ∂z
∂w 2 ~ ∂u ∂w
 
σz = −p + 2µ − µ(∇ · ~u) τzx = τxz = µ + (467)
∂z 3 ∂z ∂x

• Substitution of Eq. (467) into Eq. (464) gives:


Du ∂p ∂ ∂u 2 ∂ ∂u ∂v ∂w
    
ρ = − +2 µ − µ + +
Dt ∂x ∂x ∂x 3 ∂x ∂x ∂y ∂z
∂ ∂v ∂u ∂ ∂u ∂w
     
+ µ + + µ + + ρgx (468)
∂y ∂x ∂y ∂z ∂z ∂x
If viscosity is assumed constant (valid for small changes in temperature):
!
Du ∂p ∂2u ∂2u ∂2u µ ∂ ~
ρ =− +µ + 2 + 2 + (∇ · ~u) + ρgx (469)
Dt ∂x ∂x2 ∂y ∂z 3 ∂x

In vector notation:
D~u ~ + µ∇2 ~u + µ ∇(
~ ∇~ · ~u) + ρ~g
ρ = −∇p (470)
Dt 3
If the flow is incompressible, the continuity equation equation gives:
~ · ~u = 0
∇ (471)

therefore, the vector form of the Navier-Stokes equations may be written as follows for
an incompressible flow of a constant viscosity fluid:
D~u ~ + µ∇2 ~u + ρ~g
ρ = −∇p (472)
Dt

~ where h is the vertical direction (measured positive up), the three


• Defining ~g = −g ∇h,
components of the Navier-Stokes equations for incompressible flow of a constant viscosity
fluid are:
!
∂u ∂u ∂u ∂u ∂p ∂h ∂2u ∂2u ∂2u
ρ + ρu + ρv + ρw = − − ρg +µ + 2 + 2 (473)
∂t ∂x ∂y ∂z ∂x ∂x ∂x2 ∂y ∂z
!
∂v ∂v ∂v ∂v ∂p ∂h ∂2v ∂2v ∂2v
ρ + ρu + ρv + ρw = − − ρg +µ + + (474)
∂t ∂x ∂y ∂z ∂y ∂y ∂x2 ∂y 2 ∂z 2
!
∂w ∂w ∂w ∂w ∂p ∂h ∂2w ∂2w ∂2w
ρ + ρu + ρv + ρw = − − ρg +µ + + (475)
∂t ∂x ∂y ∂z ∂z ∂z ∂x2 ∂y 2 ∂z 2
Eng6901 - Heat Transfer I 96

D Radiation Shape Factors


Eng6901 - Heat Transfer I 97

You might also like