You are on page 1of 78

MET302 Heat and Mass Transfer

Module I - Conduction
What is heat transfer?
• Heat transfer (or heat) is thermal energy in transit due
to a spatial temperature difference.
• Whenever a temperature difference exists in a
medium or between media, heat transfer must occur.
Why study heat transfer?
• A thermodynamic analysis simply tells us how much
heat must be transferred to realize a specified change
of state to satisfy the conservation of energy principle.
• In practice we are more concerned about the rate of
heat transfer (heat transfer per unit time) than we are
with the amount of it.
An example…
• The amount of heat transferred from a
thermos bottle as the hot coffee inside
cools from 90°C to 80°C, can be
determined by a thermodynamic
analysis alone.
• But a typical user or designer of a
thermos is primarily interested in how
long it will be before the hot coffee
inside cools to 80°C.
Thermodynamic analysis cannot answer Fig 1

this question….
Modes of heat transfer
Conduction - heat transfer that occur across a stationary
medium (solid or fluid) driven by a temperature gradient.
Convection - heat transfer processes effected by the flow of
fluids (eg. – heat transfer between a surface and a fluid moving
over the surface)
Radiation - All surfaces at finite temperature emit energy in the
form of electromagnetic waves.

Fig 2
Some application areas of heat transfer

Fig 3
Categories of heat transfer problems
Rating problems
The rating problems deal with the determination of the
heat transfer rate for an existing system at a specified
temperature difference.

Sizing problems
The sizing problems deal with the determination of the
size of a system in order to transfer heat at a specified
rate for a specified temperature difference.
Conduction
• Conduction is the transfer of energy from the more
energetic to the less energetic particles of a substance
due to interactions between the particles.
• Conduction can take place in solids, liquids, or gases.
• In gases and liquids, conduction is due to the collisions
and diffusion of the molecules during their random
motion.
• In solids, it is due to the combination of vibrations of
the molecules in a lattice and the energy transport by
free electrons.
Some examples…
• A cold canned drink in a warm room
eventually warms up to the room
temperature as a result of heat transfer from
the room to the drink through the aluminum
can by conduction.
• The exposed end of a metal spoon suddenly
immersed in a cup of hot coffee is eventually
warmed due to the conduction of energy
through the spoon.
• On a winter day, there is significant energy
loss from a heated room to the outside air.
This loss is principally due to conduction heat
transfer through the wall that separates the
room air from the outside air.
Fig 4 a - b
The rate equation – Fourier law
• For the one-dimensional plane wall shown in Figure on
the right, having a temperature distribution T(x), the rate
equation is expressed as,

dT
q  k
''
x
dx Fig 5
''
• The heat flux, q x is the heat transfer rate in the x-direction
per unit area perpendicular to the direction of transfer,
and it is proportional to the temperature gradient, dT/dx,
in this direction.
• The parameter k is a transport property known as the
thermal conductivity (W/m K) and is a characteristic of
the wall material.
• The minus sign indicates that heat is transferred in the
direction of decreasing temperature.
• Recognizing that the heat flux is a vector quantity, we
can write a more general statement of the conduction
rate equation (Fourier’s law) as follows:
 T T T 
q  kT  k  i
''
j k 
 x y z 

q  k iq  jq  kq
'' ''
x
''
y
''
z 
• It is also implicit in the above equation that the
medium in which the conduction occurs is isotropic
(thermal conductivity is independent of the coordinate
direction).
Key features of Fourier’s law
• It is not an expression that may be derived from first
principles; it is instead a generalization based on
experimental evidence.
• It defines an important material property, the thermal
conductivity.
• The heat flux is normal to an isotherm and in the
direction of decreasing temperature.

Fig 6

• The Fourier’s law applies for all matter, regardless of


its state (solid, liquid, or gas).
Thermal conductivity
Solids
• In solids, the transport of thermal energy may be due to
two effects: the migration of free electrons and lattice
vibrational waves (phonons).
• In pure metals, the electron contribution to conduction
heat transfer dominates, whereas
c
in nonconductors and
semiconductors, the phonon contribution is dominant.
• Kinetic theory yields the following expression for the
thermal conductivity
C – specific heat per unit volume
1
k  C c mfp c – average speed or electrons/phonons
3
mfp – mean free path of electrons/phonons
The temperature dependence of the thermal
conductivity of selected solids.

Fig 7
Fluids
• In fluids, the intermolecular spacing is much larger
and the motion of the molecules is more random
than for the solid state.
• Therefore, thermal energy transport is less effective.
• The effect of temperature, pressure, and chemical
species on the thermal conductivity of a gas may be
explained in terms of the kinetic theory of gases.
cv – specific heat per unit volume
1
k  cv  c mfp c – average speed or electrons/phonons
3
mfp – mean free path of electrons/phonons
 – density of the gas
• The thermal conductivity of nonmetallic liquids
generally decreases with increasing temperature.
• Water, glycerine, and engine oil are notable exceptions.

Fig 8
• The thermal conductivity is independent of pressure
except in extreme cases as, for example, when
conditions approach that of a perfect vacuum.

Fig 9
The Heat Diffusion Equation
Consider a homogeneous medium within which there is no bulk
motion (advection) and the temperature distribution T(x, y, z) is
expressed in cartesian coordinates.

We define an infinitesimally
Small (differential) control
volume, dx.dy.dz, as shown
in figure.

Fig 10
• Apply the rate form of the first law to the control
volume.
• Only thermal forms of energy will be considered.
• In the absence of motion (or with uniform motion),
there are no changes in mechanical energy and no work
being done on the system.
• If there are temperature gradients, conduction heat
transfer will occur across each of the control surfaces.
• The conduction heat rates perpendicular to each of the
control surfaces at the x-, y-, and z-coordinate locations
are indicated by the terms qx, qy, and qz, respectively.
• The conduction heat rates at the opposite surfaces can
then be expressed as a Taylor series expansion where,
neglecting higher-order terms,
qx
qx  dx  qx  dx
x
q y
q y  dy  q y  dy (1)
y
qz
qz  dz  qz  dz
z
• The above equation for qx+dx, state that the x-
component of the heat transfer rate at x+dx is equal
to the value of this component at x plus the amount by
q x
which it changes with respect to x (i.e. ) times dx.
x
• Within the medium there may also be an energy
source term associated with the rate of thermal
energy generation. This term is represented as,
The energy generation term represents some energy
E g  q dx dy dz conversion process involving thermal energy on one
hand and some other form of energy, such as chemical,
(2) electrical, or nuclear, on the other.

• where, q is the rate at which energy is generated per


unit volume of the medium (W/m3).
• If the material is not experiencing a change in phase,
latent energy effects shall be neglected, and the
energy storage term may be expressed as,

 T
Est  c p dx dy dz (3)

t
• On a rate basis, the general form of the conservation
of energy requirement is
E in  E out  E g  E st (4)

T
(qx  q y  qz )  (qx  dx  q y  dy  qz  dz )  q dx dy dz  c p dx dy dz
   

 
E in
 
E out Eg  t 
E st

 q x q y q z  T
  dx  dy  dz   q dx dy dz  c p dx dy dz
   


x y

z  E
g
  
t  (5)
 Est
E in  E out
• The heat transfer in x, y and z directions (qx, qy, and
qz respectively) can be expressed in terms of the
corresponding temperature gradients by invoking
Fourier’s law of heat conduction (assuming an
isotropic medium).
T
q x  k dy dz
x
T
q y  k dz dx (6)
y
T
q z  k dx dy
z
• Substituting for qx, qy and qz in eq. 5,
  T    T    T  T
k    k    k   q  c p
x  x  y  y  z  z  t
(6)

• Equation 6 is the general form, in cartesian coordinates,


of the heat diffusion equation or heat equation.
• From its solution, we can obtain the temperature
distribution T(x, y, z) as a function of time.
• The heat equation, Equation 6, states that at any point in
the medium the net rate of energy transfer by conduction into a
unit volume plus the volumetric rate of thermal energy
generation must equal the rate of change of thermal energy
stored within the volume.
• If the thermal conductivity is constant, the heat
equation can be simplified to,
 T  T  T q 1 T
2 2 2
 2  2   (7)
x 2
y z k  t
k
where   is the thermal diffusivity (m2/s).
c p

• It measures the ability of a material to conduct


thermal energy relative to its ability to store thermal
energy.
• Materials of large  will respond quickly to changes
in their thermal environment, whereas materials of
small  will respond more sluggishly, taking longer to
reach a new equilibrium condition.
The heat equation in cylindrical coordinates

Fig 11

The general form of the heat flux


vector and hence of Fourier’s law in  T 1 T T 
cylindrical coordinates is, q  kT  k  i
''
j k 
 r r  z 
• Components of heat transfer into and out of the
control volume along r ,  , z directions can be
expressed as,
T qr
qr  k (rd dz ) qr  dr  qr  dr
r r
1 T 1 q
q  k (dr dz ) q  d  q  rd
r  r 
T q z
q z  k (rd dr ) q z  dz  q z  dz
z z
• Proceeding as in the case of Cartesian coordinates, the
heat equation in cylindrical coordinates is:
1   T  1   T    T  T
 kr  2 
 k 
  k   q  c p
 (8)
r r  r  r     z  z  t
The heat equation in spherical coordinates

Fig 12

The general form of the heat flux


vector and hence of Fourier’s law in  T 1 T 1 T 
spherical coordinates is, q  kT  k  i
''
j k 
 r r  r sin   
• Components of heat transfer into and out of the
control volume along r , ,  directions can be
expressed as,
T qr
qr  k (r sin  d d )
2 qr  dr  qr  dr
r r
1 T 1 q
q  k (r sin  d dr ) q  d  q  rd
r  r 
1 T 1 q
q  k (rd dr ) q  d  q  r sin d
r sin   r sin  

• Proceeding as in the case of Cartesian coordinates,


the heat equation in spherical coordinates is:
1   2 T  1   T  1   T  T
 kr  2 2  k   2  k sin    q  c p

r r 
2
r  r sin      r sin      t
(9)
Boundary and Initial conditions
Table 1

Dirichlet boundary condition

Neumann boundary condition

Robin boundary condition


One dimensional heat conduction without
heat generation
Plane wall
• The temperature distribution in the wall can be
determined by solving the heat equation with the
proper boundary conditions.
• For steady-state conditions with no distributed source
or sink of energy (no heat generation or heat
absorption) within the wall, the appropriate form of
the heat equation is

d  dT 
k 0 (10)
dx  dx 
Fig 13
• If the thermal conductivity of the wall material is
assumed to be constant, the equation may be integrated
twice to obtain the general solution.

T ( x)  C1 x  C2 (11)

• To obtain the constants of integration, C1 and C2,


boundary conditions must be introduced.
• Applying the specified temperature boundary condition
(Dirichlet boundary condition) at x=0 and x=L.
T ( x  0)  Ts ,1
T ( x  L)  Ts , 2
• At x=0, T=Ts,1
Ts ,1  C1.0  C2 C2  Ts ,1

• at x=L, T=Ts,2.
Ts , 2  Ts ,1
Ts , 2  C1L  C2 C1 
L
• Therefore
x
T ( x)  (Ts , 2  Ts ,1 )  Ts ,1 (12)
L
• From this result it is evident that, for one-dimensional,
steady-state conduction in a plane wall with no heat generation
and constant thermal conductivity, the temperature varies linearly
with x.
• Once the temperature distribution in the wall is
known, the heat transfer through the wall can be
obtained by using Fourier’s law.

dT kA
qx  kA  (Ts ,1  Ts , 2 ) (13)
dx L
• Equation 13 indicates that the heat transfer qx (for a
plane wall) is a constant, independent of x.
Thermal resistance
• For the special case of one-dimensional heat transfer
with no internal energy generation and with constant
properties, an analogy exists between the diffusion of
heat and electrical charge.
• Defining resistance as the ratio of a driving potential
to the corresponding transfer rate, it follows from
Equation 13 that the thermal resistance for
conduction in a plane wall is

Ts ,1  Ts , 2 L
Rt ,cond   (14)
qx kA
• A thermal resistance may also be associated with
heat transfer by convection at a surface. The
convection heat transfer from a surface to a fluid is,

qconv  hA(Ts  T ) (15)

• The thermal resistance for convection is then,


Ts  T 1
Rt ,conv   (16)
qx hA

• The equivalent thermal circuit for the plane wall with


convection surface conditions is shown in Figure 14.
Fig 14
The composite wall
The heat transfer
through the composite
wall is;

T ,1  T , 4
qx  (17)

R t

R t
- total thermal
resistance

1 LA LB LC 1
Fig 15
 Rt  h A  k A  k A  k A  h A
1 A B C 4
• With composite systems, it is often convenient to
work with an overall heat transfer coefficient U,
which is defined by an expression analogous to
Newton’s law of cooling.

qx  UA(T,1  T, 4 ) (18)

• Comparing equations 17 and 18

1 1 LA LB LC 1
 Rt  UA  h A  k A  k A  k A  h A (19)
1 A B C 4
Contact resistance
In composite systems, there will be a temperature drop across the
interface between materials. This temperature drop is due to what
is known as the thermal contact resistance, Rt,c.

For a unit area of the interface, the resistance is defined as

TA  TB
R ''
t ,c ''
(20)
qx

The existence of a finite contact


resistance is due principally to
surface roughness effects.

Fig 16
• Contact spots are interspersed with gaps that are, in
most instances, air filled. Heat transfer is therefore
due to conduction across the actual contact area and
to conduction and/or radiation across the gaps.
• The contact resistance may be reduced by
o increasing the area of the contact spots.
• by increasing the joint pressure, and/or
• by reducing the roughness of the mating surfaces.
o by selecting an interfacial fluid of large thermal conductivity.

Table 2
Radial systems – Cylinder
• Radial systems (cylindrical/spherical) often experience
temperature gradients in the radial direction only and
may therefore be treated as one-dimensional.
• A common example is the hollow cylinder whose
inner and outer surfaces are exposed to fluids at
different temperatures.
• For steady-state conditions with no heat generation,
the appropriate form of the heat equation is,
1 d  dT 
 kr 0 (21)
r dr  dr 
Fig 17
• The temperature distribution in the cylinder can be
determined by solving equation 21and applying
appropriate boundary conditions.
• Assuming the value of k to be constant, equation 21
may be integrated twice to obtain the general
solution.
T (r )  C1 ln r  C2 (22)

• Applying the specified temperature boundary


condition (Dirichlet boundary condition) at r=r1 and
r=r2.
Ts ,1  C1 ln r1  C2 Ts , 2  C1 ln r2  C2 (23)
• Solving for C1 and C2 and substituting into the
general solution,
Ts ,1  Ts , 2 r
T (r )  ln    Ts , 2 (24)
 r1   r2 
ln  
 r2 
• The temperature distribution associated with radial
conduction through a cylindrical wall is logarithmic,
not linear.
• Heat transfer through the cylindrical wall can now be
determined by applying Fourier’s law.
dT
qr  k (2r1 L) (25)
dr r  r1
Ts ,1  Ts , 2 1 1
qr  k (2r1 L)
 r1  r r2
ln    
 r2   r2  r  r1

2kL(Ts ,1  Ts , 2 )
qr  (26)
 r2 
ln  
 r1 

• From this result it is evident that, for radial conduction


in a cylindrical wall, the thermal resistance is of the
form
 r2 
ln  
Rt ,cond   r1  (27)
2kL
Composite cylindrical wall
The heat transfer
through the composite
cylindrical wall is;

T ,1  T , 4
qx 
R t (28)

Fig 18
R t
- total thermal
resistance

 r2   r3   r4 
ln   ln   ln  
1  r1   r2   r3  1
 Rt  h (2r L)  2k L  2k L  2k L  h (2r L) (29)
1 1 A B C 4 4
• When expressed in terms of an overall heat transfer
coefficient,
qr  UA(T,1  T, 4 ) (30)

• If U is defined in terms of the inside area, A1  2r1L


equations 29 and 30 may be equated to yield,

1 1 r1  r2  r1  r3  r1  r4  r1 1
 Rt  U A  h  k ln r   k ln r   k ln r   r h (31)
1 1 1 A  1 B  2 C  3 4 4

• This definition is arbitrary, and the overall coefficient


may also be defined in terms of A4 or any of the
intermediate areas
Radial systems – Sphere
• Radial systems (cylindrical/spherical) often experience
temperature gradients in the radial direction only and
may therefore be treated as one-dimensional.
• A common example is the hollow sphere whose inner
and outer surfaces are exposed to fluids at different
temperatures.
• For steady-state conditions with no heat generation,
the appropriate form of the heat equation is,
1 d  2 dT 
 kr 0 (32)
r dr 
2
dr 
• The temperature distribution in the spherical shell
can be determined by solving equation 32 and
applying appropriate boundary conditions.
• Assuming the value of k to be constant, equation 32
may be integrated twice to obtain the general
solution. C
T (r )   1
 C2 (33)
r
• Applying the specified temperature boundary
condition (Dirichlet boundary condition) at r=r1 and
r=r2.
C1 C1
Ts ,1    C2 Ts , 2    C2 (34)
r1 r2
• Solving for C1 and C2 and substituting into the
general solution,
Ts ,1  Ts , 2 1 r2Ts , 2  r1Ts ,1
T (r )   (35)
(r2  r1 ) r (r2  r1 )
r1r2

• The temperature distribution associated with radial


conduction through a cylindrical wall is hyperbolic.
• Heat transfer through the spherical wall can now be
determined by applying Fourier’s law.

dT
qr  k (4r ) 1
2
(36)
dr r  r1
Ts ,1  Ts , 2  1 
qr  k (4r )2
 2 
1
(r2  r1 )  r  r r
1
r1r2
Ts ,1  Ts , 2
qr  (4r1r2 k ) (37)
(r2  r1 )
• From this result it is evident that, for radial conduction
in a spherical wall, the thermal resistance is of the
form
(r2  r1 )
Rt ,cond  (38)
4r1r2 k
Composite spherical wall
The heat transfer
through the composite
spherical wall is;

T ,1  T , 4
qx  (39)
R t

Fig 19
 t
R - total thermal
resistance

1 r2  r1 r3  r2 r4  r3 1
 Rt  h (4r 2 )  4r r k  4r r k  4r r k  h (4r 2 )
(40)

1 1 1 2 A 2 3 A 3 4 A 4 4
• When expressed in terms of an overall heat transfer
coefficient,
qr  UA(T,1  T, 4 ) (41)

• If U is defined in terms of the inside area, A1  4r12 L


equations 40 and 41 may be equated to yield,
1 r2  r1 2 r3  r2 2 r4  r3 2 1 1
2
1 r
 Rt  U A  h  r r k r1  r r k r1  r r k r1  r 2 h (42)
1 1 1 1 2 2 3 3 4 4 4

• This definition is arbitrary, and the overall coefficient


may also be defined in terms of A4 or any of the
intermediate areas
One dimensional heat conduction with heat
generation
• This section considers the additional effect of
processes in which thermal energy is being generated
due to conversion from some other energy form, on
the temperature distribution.
• A common thermal energy generation process
involves the conversion from electrical to thermal
energy in a current-carrying medium (Ohmic, or
resistance, or Joule heating).
• The rate at which energy is generated by passing a
current I through a medium of electrical resistance,
Re is,

Eg  I R e
2 (43)
• If this power generation (W) occurs uniformly
throughout the medium of volume V, the volumetric
generation rate (W/m3) is then
E g I 2 Re (44)
q  
V V
• Energy generation may also occur as a result of the
deceleration and absorption of neutrons in the fuel
element of a nuclear reactor or exothermic chemical
reactions occurring within a medium.
• A conversion from electromagnetic to thermal
energy may occur due to the absorption of radiation
within the medium.
Plane wall
• Consider the plane wall of figure 20, in which there is
uniform energy generation per unit volume (is
constant) and the surfaces are maintained at Ts,1 and
Ts,2.
• For constant thermal conductivity k, the appropriate
form of the heat equation is,
d 2T q
2
 0 (45)
dx k
• The general solution is
q 2
T  x  C1 x  C2 (46)
2k
Fig 20
Fig 21
• Applying the specified temperature boundary
conditions at the boundaries.

T ( L)  Ts ,1 T ( L)  Ts , 2

• Solving for C1 and C2,


Ts , 2  Ts ,1 q 2 Ts , 2  Ts ,1
C1  C2  L 
2L 2k 2

• The temperature distribution is,


qL2  x 2  Ts , 2  Ts ,1 x Ts , 2  Ts ,1
T ( x)  1  2    (47)
2k  L  2 L 2
• The preceding result simplifies when both surfaces are
maintained at a common temperature, Ts,1=Ts,2=Ts.
• The temperature distribution is then symmetrical
about the midplane (figure 21), and is given by,
qL2  x 2 
T ( x)  Ts  1  2  (48)
2k  L 
• The maximum temperature exists at the midplane, x=0
qL2
T ( x  0)  T0  Ts  (49)
2k
• It is important to note that at, x=0 (fig 21) the
temperature gradient is zero (ie no heat transfer across the
plane – adiabatic surface) dT (50)
0
dx x 0
• To use the foregoing results, the surface temperature(s)
Ts must be known.
• However, a common situation is one for which it is the
temperature of an adjoining fluid, T∞ , and not Ts, which
is known.
• A relation between T∞ and Ts may be developed by
applying a surface energy balance.
dT
k  h(Ts  T ) (51)
dx x  L  
  heat convected to the surrounding fluid at T
heat conductedto the surface, x  L at Ts

• Obtaining the temperature gradient at x=L from


equation 48 and substituting into equation 51,
qL
Ts  T  (52)
h
Radial systems - Cylinder
• Consider the long, solid cylinder of figure 21, which
could represent a current-carrying wire or a fuel
element in a nuclear reactor.
• To determine the temperature distribution in the
cylinder, the appropriate form of the heat equation
must be solved. For constant thermal conductivity k,
the heat equation is,
1 d  dT  q
r   0 (53)
r dr  dr  k
• This expression may be integrated twice to obtain
q 2
T (r )   r  C1 ln r  C2 (54)
4k
Fig 22
• Applying boundary conditions
dT T (r  ro )  Ts
0 (55)
dr r 0
• Solving for C1 and C2,
qro2
C1  0 C2  Ts 
4k
• The temperature distribution is,
qro2  r 2 
T (r )  1  2   Ts (56)
4k  ro 
• The temperature at the centre of the rod (r=0) is,
qro2
T (r  0)  To  Ts  (57)
4k
• The relation between Ts and T∞ can be developed by
applying a surface energy balance,
dT
k  h(Ts  T ) (58)
dr r ro  
  heat convected to the surrounding fluid at T
heat conductedto the surface, r  ro at Ts
• Substituting the value of temperature gradient at r=ro
in equation 58,
qro
Ts  T  (59)
2h
Radial systems - Sphere
• Consider a solid sphere of radius ro. Assuming the
thermal conductivity to be constant, the heat
equation is,
1 d  2 dT  q
r   0 (60)
r dr  dr  k
2

• This expression may be integrated twice to obtain,


q 2 2C1
T (r )   r   C2 (61)
6k r
• Applying boundary conditions
dT T (r  ro )  Ts (62)
0
dr r 0
• Solving for C1 and C2,
qro2
C1  0 C2  Ts 
6k
• The temperature distribution is,

qro2  r 2 
T (r )  1  2   Ts (63)
6k  ro 
• The temperature at the centre of the sphere (r=0)
is,
qro2
T (r  0)  To  Ts  (64)
6k
• The relation between Ts and T∞ can be developed by
applying a surface energy balance,
dT
k  h(Ts  T ) (65)
dr r ro  
  heat convected to the surrounding fluid at T
heat conductedto the surface, r  ro at Ts
• Substituting the value of temperature gradient at r=ro
in equation 65,
qro
Ts  T  (66)
3h
Variable thermal conductivity

• The thermal conductivity of a material, in general,


varies with temperature.
• However, this variation is mild for many materials in
the range of practical interest and can be disregarded.
• When the variation of thermal conductivity with
temperature in a specified temperature interval is
large, it may be necessary to account for this variation
to minimize the error.
• If the variation of thermal conductivity with
temperature k(T) is known, the average value of the
thermal conductivity in the temperature range
between T1 and T2 can be determined from
T2

 k (T )dT (67)
k ave 
T1

T2  T1

• The variation in thermal conductivity of a material


with temperature in the temperature range of interest
can often be approximated as a linear function and
expressed as
k (T )  ko (1  T ) (68)

• Where  is called the temperature coefficient of


thermal conductivity.
• The average value of thermal conductivity in the
temperature range T1 to T2 in this case can be
determined from,
T2

 k (1  T )dT
o
 T2  T1 
k ave   k o 1     k (Tave )
T1

T2  T1  2 

Note that the average thermal


conductivity in this case is equal
to the thermal conductivity value
at the average temperature.

Fig 23
Conduction shape factor
• Heat transfer in simple geometries such as large plane
walls, long cylinders, and spheres can be approximated
as one-dimensional, making such problems relatively
easier to solve.
• But many problems encountered in practice are two-
or three-dimensional and involve rather complicated
geometries for which no simple solutions are available.
• An important class of heat transfer problems for
which simple solutions are obtained include those
involving two surfaces maintained at constant
temperatures T1 and T2.
• The steady rate of heat transfer between these two
surfaces can be expressed as,
q  Sk (T1  T2 ) (69)

• where S is the conduction shape factor, which has the


dimension of length, and k is the thermal conductivity of
the medium between the surfaces.
• The conduction shape factor depends on the geometry
of the system only.

Isothermal cylinder of length L Two parallel isothermal cylinders The edge of two adjoining
buried in a semi-infinite medium placed in an infinite medium walls of equal thickness
(L>>D and z >1.5D) (L>>D1, D2, z)
Fig 24
• For a three-dimensional wall, as in a furnace, separate
shape factors are used to calculate the heat flow
through the edge and corner sections, with the
dimensions shown in figure 25.
• When all the interior dimensions are greater than
one-fifth of the wall thickness,
A
S wall  Sedge  0.54 D Scorner  0.15L
L

A – area of wall

L – wall thickness

D – length of edge

Fig 25
Critical radius of insulation
• Adding more insulation to a plane wall always
decreases heat transfer.
• Since the heat transfer area A is constant, and adding
insulation always increases the thermal resistance of
the wall without increasing the convection resistance.
• Adding insulation to a cylindrical pipe or a spherical
shell, however, is a different matter.
• The additional insulation increases the conduction
resistance of the insulation layer but decreases the
convection resistance of the surface because of the
increase in the outer surface area for convection.
• Consider a cylindrical pipe
of outer radius r1 whose
outer surface temperature
T1 is maintained constant.
• The pipe is now insulated
with a material whose
thermal conductivity is k and
outer radius is r2. Fig 26

• The rate of heat transfer from the insulated pipe to


the surrounding air (h, T∞) can be expressed as
T1  T T1  T
q 
Rins  Rconv  r2  (70)
ln  
 r1   1
2Lk h(2r2 L)
• The value of r2 at which q reaches a maximum is
determined from the requirement that, dq  0
dr2

• Performing the differentiation and solving for r2


yields the critical radius of insulation for a cylindrical
body to be k
rcr ,cylinder  (71)
h
The rate of heat transfer from
the cylinder increases with the
addition of insulation for r2<rcr,
reaches a maximum when r2=rcr, Fig 27
and starts to decrease for r2 >rcr.
• For a spherical container of outer radius r1 and
temperature T1 with insulation of material having
thermal conductivity k and radius r2, the rate of heat
transfer to the surrounding air can be expressed as,
T1  T T1  T
q  (72)
Rins  Rconv r2  r1 1

4r1r2 k h(2r2 L)
• The critical radius of insulation for the spherical
container can be obtained as in the previous case for
a cylindrical pipe.
2k
rcr ,cylinder  (73)
h

You might also like