You are on page 1of 77

AME 365

HEAT TRANSFER
& COMBUSTION
DERIVATION OF GENERAL HEAT
CONDUCTION EQUATION
• Illustration of heat flow in three dimensions

• From energy balance:[Heat conducted into the volume +


Heat generated within the volume] = {heat conducted
away from the volume + change in internal energy}
CONTINUATION
• For unsteady state, change in internal energy is not zero
and temperature is a function of time
• Heat conducted into the volume is given by Q and during
time dt , total energy is (Qx + Qy +Qz)dt
• If is the rate of generation of heat energy per unit
volume
• Total energy generated is given by dxdydzdt
• Heat conducted away from the volume is

Q x  dx 
 Q y  dy  Q y  dz dt
• Change in internal energy mcdT  cdT dxdydz 
• Where,
m  d  dxdydz
General Heat Conduction Equation
• From the statement of energy balance
   
 Q x  Q y  Q z dt  q g dxdydzdt  Q x  dx  Q y  dy  Q z  dz dt  cdT dxdydz 
• Now from Taylor’s series expansion:
  T    T 
    dydz  dxdt     dxdydzdt
x  x  x  x 
• And Substituting for the area term  A  dydz in the equation we obtain:
x


  
Q x  Q x  dx dt  
x
 
  
Q x dxdt    Ax

T 
 dxdt
x  x 
  T    T 
   dydz  dxdt     dxdydzdt
x  x  x  x 
• This is only the simplification for the x-component but the same analysis is
carried out for the y and z-components
General heat conduction equation in
three dimensions.
• Simplifying the y and z terms as previously carried out
the x-component we finally obtain the equation:

   T    T    T 
            dvdt  q g dvdt  cdTdv
 x  x  y  y  z  z 
• And dividing by dvdt we obtain the general heat conduction
equation before specific simplifications as:

  T    T    T   T
            q g 
x  x  y  y  z  z   d t
Case 1 :Fourier-Biot equation
if only λ is assumed constant
• λ can be taken out and the differential
operation carried out resulting in as:

•  T  T  T q g 1 T
2 2 2
   
x y
2 2
z 2
  d t
Case 2: If λ is constant and
• These conditions result in the FOURIER’S
equation:

 T  T  T 1 T
2 2 2
  
 y z  d t
2 2 2
Case 3:
λ assumed constant at steady-state with

• These conditions result in the POISSON


equation

 T  T  T q g
2 2 2
    0
x 2
y 2
z 2

Case 4:
Steady state-conditions with λ assumed constant and

• These conditions result in the LAPLACE


Equation

T T T
2 2 2
   0
x y z
2 2 2
General/Combined One-Dimensional Heat Conduction Equation

• An examination of the one-dimensional transient heat


conduction equations for the plane wall, cylinder, and
sphere reveals that all three equations can be expressed in
a compact form as:

• Note: r=x and n = 0 for a plane wall, r= r and n= 1 for a


cylindrical wall, and r= r and n = 2 for a spherical wall.
Boundary and Initial Conditions
•The heat conduction equations were developed using an
energy balance on a differential element inside the
medium, and they remain the same regardless of the
thermal conditions on the surfaces of the medium.

•The heat flux and the temperature distribution in a


medium depend on the conditions at the surfaces, and the
description of a heat transfer problem in a medium is not
complete without a full description of the thermal
conditions at the bounding surfaces of the medium.

•The mathematical expressions of the thermal conditions at


the boundaries are called the boundary conditions.
Boundary and Initial Conditions Condt.
•The temperature at any point on a medium at a specified time also depends
on the condition of the medium at the beginning of the heat conduction
process. Such a condition, which is usually specified at time t = 0, is called
the initial condition, which is a mathematical expression for the temperature
distribution of the medium initially.

•In rectangular coordinates, the initial condition can be specified in the


general form as

•where the function f(x, y, z) represents the temperature distribution


throughout the medium at time t = 0. When the medium is initially at a
uniform temperature of T i , the initial condition can be expressed as T(x, y, z,
0) = T i .
•Note that under steady conditions, the heat conduction equation does not
involve any time derivatives, and thus we do not need to specify an initial
Boundary and Initial Conditions Contd.
•The heat conduction equation is first order in time, and
thus the initial condition cannot involve any derivatives (it
is limited to a specified temperature).

•However, the heat conduction equation is second order in


space coordinates, and thus a boundary condition may
involve first derivatives at the boundaries as well as
specified values of temperature. Boundary conditions
most commonly encountered in practice are the specified
temperature, specified heat flux, convection, and radiation
boundary conditions.
1. Specified Temperature Boundary

Condition
For one-dimensional heat transfer through a plane wall of thickness L, for
example, the specified temperature boundary conditions can be expressed as

• where T 1 and T 2 are the specified temperatures at surfaces at x = 0 and x = L,


respectively.
• The specified temperatures can be constant, which is the case for steady heat
conduction, or they may vary with time.
2. Specified Heat Flux Boundary Condition
• The heat flux in the positive x-direction anywhere in the medium,
including the boundaries, can be expressed by Fourier’s law of
heat conduction as

For a plate of thickness L subjected to heat flux


of 50 W/ m 2 into the medium from both sides,
for example, the specified heat flux boundary
conditions can be expressed as

• Note that the heat flux at the surface at x = L is in


the negative x-direction, and thus, it is − 50 W/ m2. The direction of heat flux
arrows at x = L in this case would be reversed.
3. Special Case: Insulated Boundary
• The boundary condition on a perfectly insulated surface (at x = 0,
for example) can be expressed as:

• NOTE: On an insulated surface, the first derivative of temperature


with respect to the space variable (the temperature gradient) in
the direction normal to the insulated surface is zero.
4. Another Special Case: Thermal
Symmetry
5. Convection Boundary Condition
• The convection boundary condition is based on a surface energy balance
expressed as

•where h 1 and h 2 are the convection heat transfer coefficients and T ∞1 and
T ∞2 are the temperatures of the surrounding media on the two sides of the
6. Radiation Boundary Condition
• Using an energy balance, the radiation boundary condition on a
surface can be expressed as

• where ε 1 and ε 2 are the emissivities of the boundary surfaces, σ


= 5.67 × 10 −8 W/ m2 ⋅ K 4 is the Stefan–Boltzmann constant, and T
surr, 1 and T surr, 2 are the average temperatures of the surfaces
surrounding the two sides of the plate, respectively. Note that the
temperatures in radiation calculations must be expressed in K or R
(not in °C or °F ).
7. Interface Boundary Conditions
• The boundary conditions at the interface of two bodies A and B in
perfect contact at x = x 0 can be expressed as

where k A and k B are the


thermal conductivities of the
layers A and B, respectively.
8. Generalized Boundary Conditions
• In general, a surface may involve convection, radiation, and specified heat flux
simultaneously. The boundary condition in such cases is obtained from a
surface energy balance, expressed as

All of the quantities in the above relations


are known except the temperatures and their
derivatives at r = 0 and r 0 . Also, the radiation
part of the boundary condition is often ignored
for simplicity by modifying the convection heat
transfer coefficient to account for the contribution
of radiation. The convection coefficient h in that
case becomes the combined heat transfer coefficient.
SPECIAL CASE
Combined Convection, Radiation, and Heat Flux

• The heat transfer through the wall is


given to be steady and one-
dimensional, and thus the temperature
depends on x only and not on time.
That is, T = T(x).

• The boundary condition on the


outer surface can be expressed
as

where is the incident solar heat flux.


HEAT GENERATION IN A SOLID
• Many practical heat transfer applications involve the conversion
of some form of energy into thermal energy in the medium.
• Such media are said to involve internal heat generation, which
manifests itself as a rise in temperature throughout the
medium.

• Some examples of heat generation are resistance heating in


wires, exothermic chemical reactions in a solid, and nuclear
reactions in nuclear fuel rods where electrical, chemical, and
nuclear energies are converted to heat, respectively
HEAT GENERATION IN A SOLID
Heat generation is usually expressed per unit volume of
the medium and is denoted by , whose unit is W/m3. For
example, heat generation in an electrical wire of outer
radius r o and length L can be expressed as

where I is the electric current and Re is the electrical


resistance of the wire
HEAT GENERATION IN A SOLID
• The maximum temperature T max in a solid that involves uniform
heat generation occurs at a location farthest away from the outer
surface when the outer surface of the solid is maintained at a
constant temperature T s .
• Under steady conditions, the energy balance for this solid can be
expressed as
HEAT GENERATION IN A SOLID
condt.

• For a large plane wall of thickness 2L with both sides of the wall
maintained at the same temperature T s , a long solid cylinder of radius
r o, and a solid sphere of radius r o, the equation above is expressed as:
HEAT GENERATION IN A SOLID
contd.
The heat generated within this inner cylinder must be equal to the
heat conducted through its outer surface. That is, from Fourier’s law
of heat conduction,

where A r = 2πrL and V r = π r 2 L at any location r.

Integrating from r = 0 where T(0) = T 0 to r = r o


where T( r o ) = T s yields
HEAT GENERATION IN A SOLID contd.

where T 0 is the centerline temperature of the cylinder, which is the maximum


temperature, and Δ T max is the difference between the centerline and the surface
temperatures of the cylinder, which is the maximum temperature rise in the
cylinder above the surface temperature.

• Once Δ T max is available, the centerline


temperature can easily be determined from
VARIABLE THERMAL
CONDUCTIVITY, k(T)
When the variation of thermal conductivity with
temperature k(T) is known, the average value of the
thermal conductivity in the temperature range between
T 1 and T 2 can be determined from

• Note that in the case of constant thermal conductivity


k(T) = k, the equation reduces to k avg = k, as expected.
VARIABLE THERMAL CONDUCTIVITY,
k(T)
• The rate of steady heat transfer through a plane wall, cylindrical
layer, or spherical layer for the case of variable thermal
conductivity can be determined by replacing the constant thermal
conductivity k with the k avg expression (or value)
VARIABLE THERMAL CONDUCTIVITY,
k(T)
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

where β is called the temperature coefficient of thermal conductivity.


The average value of thermal conductivity in the temperature range
T 1 to T 2 in this case can be determined from

Note that the average thermal conductivity


in this case is equal to the thermal conductivity
value at the average temperature
STEADY-STATE CONDUCTION HEAT TRANSFER
INTRODUCTION
In this unit, steady state heat conduction problems involving one-dimensional temperature
distribution are treated. The concept of thermal resistance for cases with one-dimensional
steady state temperature distribution is introduced and the use of equivalent thermal
circuits for analysing problems with multiple layers of different materials. The application
of heat transfer to extended surfaces is also discussed. A general heat conduction equation
in a three dimensional approach is also derived.

LEARNING OBJECTIVES
After studying this unit, you should be able to:
1. Solve steady- state conduction problems involving one-dimensional temperature distribution.
2. Understand the concept of thermal resistance for cases with one-dimensional steady temperature
distribution.
3. Demonstrate the use of equivalent thermal circuits for analyzing problems with multiple layers of
different materials.
4. Analytically determine heat transfer rates in extended surfaces.
5. Understand the concept of contact thermal resistance in heat transfer analysis.
6. Examine the effect of internal heat generation on temperature distribution and heat transfer.
7. Solve transient problems with uniform temperature distribution.
DEFINITION OF CONDUCTION HEAT TRANSFER

 Heat conduction is that mode of energy transfer between


solids brought into direct contact having different
temperatures or within solids as a result of temperature
difference or with liquids or gases as a result of temperature
difference without an appreciable movement of matter.

 This takes places as a result of kinetic motion or direct


impact of molecules, as in the case of fluid at rest, and by
the drift of electrons as in the case of metals. In a solid,
which is a good electric conductor a large number of free
electrons move about in the lattice; hence materials that
are good electric conductors are generally good heat
conductors (e.g. copper, silver etc)
ONE DIMENSIONAL STEADY STATE CONDUCTION

• The term one-dimensional refers to the fact that only one co-
ordinate is needed to describe the spatial variation of the
dependent variables.
• Temperature gradients exist along only a single space coordinate
direction, and heat transfer occurs exclusively in that direction.
• We begin our consideration of one-dimensional, steady state
conduction by discussing heat transfer with no internal generation
in common geometries.
• The concept of thermal resistance (analogous to electrical
resistance) is introduced as an aid to solving conduction heat
transfer problems. The effect of internal heat generation on the
temperature distribution and heat rate is then treated. The general
heat conduction equation is introduced. Finally, conduction analysis
is used to describe the performance of extended surfaces or fins,
wherein the role of convection at the external boundary must be
DIRECTION OF HEAT FLOW IN RESPONSE TO A TEMPERATURE GRADIENT

• The figures below illustrates the direction of heat flow,


in a one-dimensional coordinate system when the
gradient, dT/dx, is positive or negative. Note that
conduction heat transfer is determined using Fourier’s
Law 
Q dT
q   
"
x
A dx
THERMAL DIFFUSIVITY

• The product c p called heat capacity is frequently used in heat


transfer analysis and it represents the heat storage capacity of the
material per unit volume.
• In heat transfer analysis, the ratio of the thermal conductivity to the
heat capacity is an important property termed the thermal diffusivity ,
which has units of m2/s: Heat conducted 
d   (m 2 s)
Heat stored c p

• The thermal conductivity λ represents how well a material conducts


heat, and the heat capacity c p represents how much energy a
material stores per unit volume.
• The larger the thermal diffusivity, the faster the propagation of heat
into the medium. A small value of thermal diffusivity means that heat
is mostly absorbed by the material and a small amount of heat will be
conducted further. Thermal diffusivity ranges from  d  0.14  10 6 m 2 /s
for water to 149  10 6 m 2 /s for silver.
HEAT CONDUCTION IN A PLANE WALL

For one-dimensional heat conduction in a plane wall,


temperature is a function of the x coordinate only and
heat is transferred exclusively in this direction.

In the figure, a plane wall


two fluids of different temperatures.
Heat transfer occurs by convection
from the hot fluid at to one surface
of the wall at T∞1 , by conduction
through the wall, and by convection
from the outer surface of the wall at
to the cold fluid at T∞2.
One-Dimensional steady-state conduction equation
The temperature distribution in the wall can be determined by
solving the heat equation with the proper boundary conditions
imposed. The heat equation for a plane wall of constant cross-
section is: d  dT 
  0
dx  dx 
If the thermal conductivity, λ, of the wall material is assumed
constant then,
T x   C1 x  C 2
To find values of the constants we apply the boundary conditions of
the first kind at x = 0 and x =L, in which case
T 0   Ts ,1 and T L   Ts , 2
Applying the conditions to the general solution, we have:

    x
T x  Ts , 2  Ts ,1  Ts ,1
L
Solving the established differential equation
• From equation, it is evident that for one-dimensional, steady
state heat conduction in a plane wall with no heat generation
and constant thermal conductivity, the temperature varies
linearly with x.
• Now that we have the temperature distribution, we can use
the Fourier's law to determine the conduction heat transfer
rate. That is,
dT  Ax

Qx   Ax   Ts ,1  Ts ,2 
dx L
Notes:
a. Ax is the area of the wall normal to the direction of heat
transfer and, for the plane wall, it is a constant independent
of x. q x Q x
b. The heat flux, is obtained by dividing by Ax
THERMAL RESISTANCE

Defining resistance as the ratio of a driving potential difference to the corresponding


transfer rate, the thermal resistance for conduction is,
driving potential difference Ts ,1  Ts , 2 L
Rt  Rt ,cond  
corresponding transfer rate 
Qx  Ax

Similarly, for electrical conduction in the same system, Ohm's law provides an
electrical resistance of the form
E1  E 2 L
Re   where k is the electrical conductivity
I kA

Comparing the two equations, the heat flow rate is analogous to electric current I,
while the temperature difference (Ts,1 – Ts,2) is analogous to the potential difference E1
– E 2.

A thermal resistance may also be associated with heat transfer by convection at a


Q x  ATs  T 
surface. From Newton’s law of cooling, Ts  T 1
Rt ,conv  

Qx A
THERMAL RESISTANCE

• The heat transfer rate may be determined from separate


considerations of each element in the network. Since is constant
throughout the network, it follows that

 T ,1  Ts ,1 Ts ,1  Ts , 2 Ts , 2  T , 2
Qx   
1 / 1 A L / A 1/  2 A
• In terms of the overall temperature difference, , and the total
thermal resistance, , the heat transfer rate may also be expressed
as
 T ,1  T , 2
Qx 
Rtot
• ecause the conduction and convection resistances are in series and
may be summed, it follows that 1 L 1
Rtot   
 1 A A  2 A
Thermal Resistance for Radiation Effect
Another resistance may be pertinent if a surface is
separated from large surroundings by a gas. In
particular, the radiation exchange between the
surface and its surroundings may be required. lt
follows that a thermal resistance for radiation
may be defined as:

Ts  Tsur 1
Rt ,rad  
Q rad r A
HEAT CONDUCTION IN A COMPOSITE WALL.
 Equivalent thermal circuits may also be used for more complex systems, such
as composite walls. Such walls may involve any number of series and parallel
thermal resistances due to layers of different materials.

 Consider the series composite wall.


The one-dimensional heat transfer rate
for this system may be expressed as

 T ,1  T , 4
Q
 Rt
T ,1  T , 4
Q 
1 /  1 A  L A /  A A  LB /  B A  LC / C A  1 /  4 A

 T ,1  Ts ,1 Ts ,1  T2 T2  T3
Q    ...
1 /  1 A L A /  A A LB /  B A
Overall heat transfer coefficient, U
• With composite system it is convenient to work with an overall
heat transfer coefficient, U, which is defined by an expression
analogous to Newton’s law of cooling.

Q  U A T
where U is the overall heat transfer coefficient.
• The overall heat transfer coefficient is related to the total
thermal resistance as
1 1
U 
Rtot A 1 /  1   L A /  A   LB /  B   LC / C   1 /  4 

T 1
Rtot   Rt  
• In general, we may write Q UA
Possible representation of series-parallel
composite wall configurations
Composite walls may also be characterised by
series-parallel configurations:
Thermal contact resistance
• The contact resistance for the system shown is
given by

" T A  TB
Rt ,c 
q

• The heat transfer through the interface is the sum of


the heat transfers through the solid contact spots and
the gaps in the non-contact areas and can be expressed
as:
Q  Q contact  Q gap   c ATinterface
Factors affecting contact resistance
 The value of thermal contact resistance depends on
the surface roughness and the material properties as
well as the temperature and pressure at the interface
and type of fluid trapped at the interface.

 Thermal contact resistance is observed to decrease


with decreasing surface roughness and increasing
interface pressure.

 Thermal contact resistance can be minimised by


applying thermal conducting liquid called thermal
grease such as silicon oil.
RADIAL SYSTEMS
Cylindrical and spherical systems often
experience temperature gradients in the radial
direction only and therefore may be treated as
one-dimensional.
Under steady state conditions with no heat
generation, such systems may be analysed by
using the standard method, which begins with
the appropriate form of heat equation, or the
alternative method, which begins with the
appropriate form of Fourier's law.
Fourier’s Law in Cylindrical coordinates
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 one-dimensional heat conduction equation in cylindrical
coordinates is: 1 d  dT 
 r 0
r dr  dr 
From the Fourier's law, the rate at which energy is conducted across
any cylindrical surface in the solid may be expressed as:

dT dT

Qr  Ar   2rL 
dr dr
Solution of Equation
dT dT

Qr  Ar   2rL 
dr dr
• Assuming the value of to be constant, the
Equation may be integrated twice to obtain the
general solution
T r   C1 Inr   C 2

• To evaluate the constants we introduce the


boundary conditions: T r1   Ts ,1and T r2   Ts , 2
Ts ,1  Ts , 2 r 
T r   In   Ts , 2
Inr1 / r2   r2 
Heat transfer rate in a cylindrical shell
• If the temperature distribution is now used
with Fourier’s law we obtain the following for
the heat transfer rate:
2L Ts ,1  Ts , 2 
Q r 
Inr2 / r1 
• the thermal resistance developed in
comparison with the ohm’s law is of the form:

Inr2 / r1 
Rt ,cond 
2L
Composite cylindrical wall
• Composite cylindrical wall is treated as follows:
Heat transfer rate in composite cylindrical wall
• Rate of heat transfer is evaluated as:
T ,1  T , 4
Q r 
 1 Inr2 / r1  Inr3 / r2  Inr4 / r3  1 
     
 2r1 L 1 2 A L 2 B L 2C L 2r4 L 4 

T ,1  T , 4

Qr   UAT ,1  T , 4   U 1 A1 (T )  U 2 A2 (T )  ....
Rtot
1
U1 
1 r r r r r r r 1
 1 In 2  1 In 3  1 In 4  1
 1  A r1  B r2 C r3 r4  4

1
U1A1  U 2 A 2  U 3 A 3  U 4 A 4 
Rt
Spherical Shells
• For the differential control volume, energy
conservation requires that q r  q r  dr .For steady-
state, one-dimensional conditions with no heat
generation, the appropriate form of the heat
conduction equation in spherical coordinates is:

1 d 2 dT
2
( r )0
r dr dr
Thermal resistance for spherical shell
• From Fourier’s law we can write:
 dT 2 dT
Qr  Ar   (4r )
dr dr

• Which is integrated yielding:


 4 (Ts ,1  Ts , 2 ) (Ts ,1  Ts , 2 )
Qr  
(1 r1 )  (1 r2 ) Rt ,cond
• The thermal resistance for a spherical shell is
given as:
1 1 1 r2  r1
Rt ,cond  (  )
4 r1 r2 4r2 r1
Critical Radius of Insulation
• Consider a cylindrical pipe of outer radius whose outer surface
temperature is maintained constant.
• The pipe is now insulated with a material whose thermal
conductivity is and outer radius is , Heat is lost from the pipe to
the surrounding medium at temperature , with a convection heat
transfer coefficient .
• The rate of heat transfer from the insulated pipe to the
surrounding air can be expressed as

T1  T T1  T
Q  
Rins  Rconv Inr2 / r1  1

2L  2r2 L 
Variation of thermal resistance with
cylinder radius
• The variation of heat transfer with the outer radius of insulation
is plotted in the Figure below. The variation of thermal resistance
with radius is shown alongside. At the critical radius the thermal
resistance is a minimum resulting in maximum heat transfer from
the system .
HEAT TRANSFER FROM EXTENDED SURFACES
(CONDUCTION-CONVECTION SYSTEMS)

The term extended surface is commonly used in reference


to a solid that experiences energy transfer by conduction
within its boundaries, as well as energy transfer by
convection (and/or radiation) between its boundaries and
the surroundings.

Although there are many different situations that involve


combined conduction-convection effects, the most
frequent application is one in which an extended surface is
used specifically to enhance the heat transfer rate
between a solid and an adjoining fluid. Such an extended
surface is termed a fin.
FIN ANALYSIS
Fin of rectangular cross section
Energy balance on differential element

• Applying the conservation energy requirement to a differential element of


the Figure, we obtain

 Q x  Q  x  dx   dQ conv
Derivation of the general differential equation

 dT
• From Fourier’s law, we know that Q x  Ac
dx
• From Taylor’s series expansion we write:
dQ dT d  dT 
Q x  dx 
 Qx  x
dx 
Q x  dx  Ac    Ac dx
dx dx dx  dx 
dQ conv  dAs T  T 

Substituting the foregoing rates into the energy balance, we obtain


d  dT   dAs
 Ac  T  T   0
dx  dx   dx

d 2T  1 dAc  dT

or  1  dAs 
 T  T   0
 
 Ac  dx 
2
dx  Ac dx  dx
Fins of uniform constant cross-section
• We consider straight rectangular and pin fins of
uniform cross section. Each fin is attached to a
base surface of temperature T 0  Tb and extends
into a fluid of temperature T . For the prescribed
fins, A c is a constant and A s is the surface area
measured from the base to x and P is the fin
perimeter. Accordingly, with A s  Px. and
• dAc / dx , 0 Equation reduces to: dAs / dx  P
defining an excess temperature as d T  P T  T   0
2


dx 2 A

 x   T x   T d 2
2
 m 2
 0 2
m =
hP
dx k Ac
Solving the constitutive fin differential
d 
equation  m --------(2.48)
2
0 2
2

dx
2 hP
m =
k Ac

• Equation 2.48 is a linear, homogeneous, second-order differential


equation with constant coefficients. Its general solution is of the form:

•  x   C1e mx
 C2e  mx
---------------(2.49a)

• We seek to equation 2.49a for the following three boundary conditions:


Boundary conditions
1. The fin is infinitely long (Case 1)
2. Insulated at the end (Case 2)
3. Convection at the end (Case 3)

CASE 1: INFINITELY LONG FIN

• Case 1: (Infinitely long fin, L = x = ∞)


 x   C e  C e
mx  mx

• 1 2

 0  Tb  T   b  Tb  T  C1e 0  C 2 e 0

Tb  T  C1  C 2   b
 C1e   C 2 e   0
• θ=0@x=∞
 mx
e
• . Note that this condition will be satisfied by mx
e
but not by the other prospective solution function
• Since it turns to infinity as x getslarger.  x   mx
  x    b e mx
or e
• C1 = 0 and C2 = θb b
CASE 1 CONCLUDED
• The temperature variation along the fin from
the base is given below. The heat transfer rate
for the infinitely long fin is found from Fourier’s
law as provided below:

− 𝑚𝑥 − 𝑥 √ h 𝑃 / 𝑘 𝐴𝑐
𝑇 (𝑥)−𝑇 ∞=(𝑇 𝑏 −𝑇 ∞ )𝑒 =𝑒 (𝑇 𝑏 − 𝑇 ∞ )

𝑄˙ long fin =− 𝑘 𝐴 𝑐
𝑑𝑇
𝑑𝑥 |
𝑥=0
= √ h 𝑃 /𝑘 𝐴𝑐 ( 𝑇 𝑏 − 𝑇 ∞ )
CASE 2 INSULATED FIN TIP

• Negligible heat loss from fin tip (Insulated fin


tip, ( )
•  0  T  T that is C1  C 2  Tb  T
b 

d mL
dx
0 C1e mL
 C2e 0
x L

• From which we obtain T x   T  cosh mLhence


 x  the heat
transfer rate is: Tb  T cosh mL

˙𝑄 insulated tip =− 𝜆 𝐴𝑐 𝑑𝑇
𝑑𝑥 |
𝑥=0
=√ h 𝑃 /𝑘 𝐴𝑐 ( 𝑇 𝑏 −𝑇 ∞ ) tanh 𝑚𝐿
Case 3: Convection at the end (or combined convection and radiation)

• Invoking the general solution and substituting


the boundary conditions we obtain:
dT
T  T  Tb  T   0  Tb  T q cond   A at  x  L
dx

C1  C 2  Tb  T q conv  A T  T at  x  L

dT
 A  A T L   T 
dx xL

C1  C 2  Tb  T
SOLUTIONS FOR CASE 3
• The temperature profile and heat transfer rate
are given by:

 cosh mL  x    / m sinh mL  x 



b cosh mL  a / m sinh mL

sinh mL   / m cosh mL

Qconvection from fin tip  PAc Tb  T 
cosh mL   / m sinh mL
FIN EFFECTIVENESS

• Fins are used to increase the heat transfer from a surface


by increasing the effective surface area.
• However, a fin itself represents a conduction resistance to
heat transfer from the original surface.
• For this reason, there is no assurance that the heat
transfer rate will be increased through the use of fins. An
assessment of this matter may be made by evaluating the
fin effectiveness εf.
• It is defined as the ratio of the fin heat transfer rate to the
heat transfer rate that would exist without the fin.
is the fin cross-sectional area 
Q fin
ε fin 
αA c,b (Tb  T )
FIN EFFECTIVENESS
Q̇ fin √ h Pk Ac (T b −T ∞ ) k P

tanh mL
ε longfin = = =
Comments: Q̇ nofin h A c (T b −T ∞ ) h Ac
• Fin effectiveness is enhanced by the choice of a material of high thermal conductivity.
Aluminium alloys are mostly common. Copper and iron are other options. Although
copper is superior from the stand-point of thermal conductivity, aluminium alloys are the
more common choice because of additional benefits related to lower cost and weight.
• Fin effectiveness is also enhanced by increasing the ratio of the perimeter to the cross-
sectional area. For this reason the use of thin plate and slender pin, but closely spaced
fins, is preferred, with the proviso that the fin gap must not be reduced to a value for
which flow between the fins is severely impeded, thereby reducing the convection
coefficient.
• The use of fins is most effective in applications involving a low convection heat transfer
coefficient. Thus, the use of fins is more easily justified when the medium is a gas
instead of a liquid and the heat transfer is by natural convection instead of by forced
convection
FIN EFFICIENCY

• It is defined as the ratio of the heat transferred from


the fin surface to that which would be transferred if
the whole of the fin surface were at the temperature
of the base surface
• The fin efficiency is defined as
Q fin Actual heat transfer rate from the fin surface
 fin  
Q fin max Ideal heat transfer from the fin if the entire fin surface were at the base temperature

˙ 𝑓𝑖𝑛
𝑄
𝜼 𝒇𝒊𝒏=
h 𝐴 𝑓𝑖𝑛 ( 𝑇 𝑏 − 𝑇 ∞ )
RELATION BETWEEN FIN EFFICIENCY AND
EFFECTIVENESS

= PL
Where P = Perimeter of the fin
L = Length of the fin

For fins, if then the fin is said to be classified as a


long fin
Fin efficiency for straight fins
• Rectangular, triangular and parabolic fins
Fin efficiency chart for annular fins
• Annular fins of rectangular profile
Overall effectiveness- a better measure of
performance of finned surfaces
• Total heat transfer from a surface will consider
the unfinned portion of the surface as well as
the fins Q  Q  Q total fin unfin fin

 Aunfin (Tb  T )   finA fin (Tb  T )


  ( Aunfin   fin A fin )(Tb  T )
• Overall effectiveness is defined as:

Q total , fin  ( Aunfin   fin A fin )(Tb  T )


 fin ,overall  

Q A (T  T )
total , no fin no fin b 
Overall surface efficiency
• The overall surface efficiency ηo characterises
an array of fins and the base surface to which
they are attached. It is defined as
Q total Q t
o  
Q max At b

NA f
o  1  (1   f )
At
END OF LECTURES
• AS RECORDED IN LUKE 19:13
• WORK WITH THE MATERIALS
PROVIDED AND “OCCUPY TILL
I COME YOUR WAY LATER”
• THANK YOU

You might also like