Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

HEAT TRANSFER

Heat is defined as energy transferred by virtue of a temperature


difference. It flows from regions of higher temperature to regions of lower
temperature. It is customary to refer to different types of heat transfer
mechanisms as modes. The basic modes of heat transfer are conduction,
radiation, and convection.

1. Conduction Heat Transfer:

Conduction is the transfer of heat from one part of a body at a higher


temperature to another part of the same body at a lower temperature, or from
one body at a higher temperature to another body in physical contact with it
at a lower temperature. The conduction process takes place at the molecular
level and involves the transfer of energy from the more energetic molecules
to those with a lower energy level. This can be easily visualized within
gases, where we note that the average kinetic energy of molecules in the
higher-temperature regions is greater than that of those in the lower-
temperature regions. The more energetic molecules, being in constant and
random motion, periodically collide with molecules of a lower energy
level and exchange energy and momentum. In this manner there is a
continuous transport of energy from the high-temperature regions to those of
lower temperature. In liquids the molecules are more closely spaced than in
gases, but the molecular energy exchange process is qualitatively similar to
that in gases. In solids that are nonconductors of electricity (dielectrics), heat
is conducted by lattice waves caused by atomic motion. In solids that are
good conductors of electricity, this lattice vibration mechanism is only a
small contribution to the energy transfer process, the principal contribution
being that due to the motion of free electrons, which move in a similar way
to molecules in a gas.

When a temperature gradient exists in a body, experience has shown that


there is an energy transfer from the high-temperature region to the low-
temperature region. We say that the energy is transferred by conduction and
that the heat-transfer rate per unit area is proportional to the normal
temperature gradient:

1
When the proportionality constant is inserted,

…………… (1)

Where qx is the heat-transfer rate and ∂T/∂x is the temperature gradient


in the direction of the heat flow. The positive constant k is called the thermal
conductivity of the material, and the minus sign is inserted so that the second
principle of thermodynamics will be satisfied; i.e., heat must flow downhill
on the temperature scale, as indicated in the coordinate system of Figure 1.
Equation (1) is called Fourier’s law of heat conduction after the French
mathematical physicist Joseph Fourier, who made very significant
contributions to the analytical treatment of conduction heat transfer. It is
important to note that Equation (1) is the defining equation for the thermal
conductivity and that k has the units of watts per meter per Celsius degree in
a typical system of units in which the heat flow is expressed in watts.

Examples of conduction heat transfer are legion. The exposed end of a metal
spoon suddenly immersed in a cup of hot coffee will eventually be 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

2
principally due to conduction heat transfer through the wall that separates
the room air from the outside air.

It is possible to quantify heat transfer processes in terms of appropriate rate


equations. These equations may be used to compute the amount of energy
being transferred per unit time. For heat conduction, the rate equation is
known as Fourier’s law. For the one-dimensional plane wall shown in
Figure 2, having a temperature distribution T(x), the rate equation is
expressed as:

The heat flux (W/m2) 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 is a consequence of the fact that heat is
transferred in the direction of decreasing temperature. Under the steady-state
conditions shown in Figure 2, where the temperature distribution is linear,
the temperature gradient may be expressed as:

3
and the heat flux is then

………….. (2)

Note that this equation provides a heat flux, that is, the rate of heat transfer
per unit area. The heat rate by conduction, qx (W), through a plane wall of
area A is then the product of the flux and the area, qx = q”x . A .

Example 1: The wall of an industrial furnace is constructed from 0.15-m-


thick fireclay brick having a thermal conductivity of 1.7 W/m K.
Measurements made during steady state operation reveal temperatures of
1400 and 1150 K at the inner and outer surfaces, respectively. What is the
rate of heat loss through a wall that is 0.5 m by 1.2 m on a side?

Solution:
Known: Steady-state conditions with prescribed wall thickness, area,
thermal conductivity, and surface temperatures.
Find: Wall heat loss.
Schematic:

4
Assumptions:
1. Steady-state conditions.
2. One-dimensional conduction through the wall.
3. Constant thermal conductivity.

Analysis: Since heat transfer through the wall is by conduction, the heat flux
may be determined from Fourier’s law. Using Equation 2, we
have:

The heat flux represents the rate of heat transfer through a section of unit
area, and it is uniform (invariant) across the surface of the wall. The heat
loss through the wall of area is then

Comments: Note the direction of heat flow and the distinction between heat
flux and heat rate.

Example 2: The thermal conductivity of a sheet of rigid, extruded insulation


is reported to be k = 0.029 W/m oK. The measured temperature difference
across a 20-mm-thick sheet of the material is T1- T2 = 10 oC.

(a) What is the heat flux through a 2 m 2 m sheet of the insulation?


(b) What is the rate of heat transfer through the sheet of insulation?

Solution:
Known: Thermal conductivity, thickness and temperature difference across
a sheet of rigid extruded insulation.
Find: (a) The heat flux through a 2 m × 2 m sheet of the insulation, and (b)
The heat rate through the sheet.

5
Schematic:

Assumptions: (1) One-dimensional conduction in the x-direction, (2)


Steady-state conditions, (3) Constant properties.

Analysis: From Equation 1.2 the heat flux is

Solving,

The heat rate is

Comments: (1) Be sure to keep in mind the important distinction between


the heat flux (W/m2) and the heat rate (W). (2) The direction of heat flow is
from hot to cold. (3) Note that a temperature difference may be expressed in
kelvins or degrees Celsius.

6
We now set ourselves the problem of determining the basic equation
that governs the transfer of heat in a solid, using Equation (1) as a starting
point.

Consider the one-dimensional system shown in Figure 1. If the system


is in a steady state, i.e., if the temperature does not change with time, then
the problem is a simple one, and we need only integrate Equation (1) and
substitute the appropriate values to solve for the desired quantity. However,
if the temperature of the solid is changing with time, or if there are heat
sources or sinks within the solid, the situation is more complex. We consider
the general case where the temperature may be changing with time and heat
sources may be present within the body. For the element of thickness dx, the
following energy balance may be made:

Energy conducted in left face + heat generated within element = change in


internal energy + energy conducted out right face

These energy quantities are given as follows:

7
Where
q˙ = energy generated per unit volume, W/m3
c = specific heat of material, J/kg · ◦C
ρ = density, kg/m3

Combining the relations above gives

or

This is the one-dimensional heat-conduction equation. To treat more than


one-dimensional heat flow, we need consider only the heat conducted in and
out of a unit volume in all three coordinate directions, as shown in Figure 4-
a. The energy balance yields:

And the energy quantities are given by

8
Figure 4: Elemental volume for three-dimensional heat-conduction analysis:
(a) Cartesian coordinates; (b) cylindrical coordinates; (c) spherical
coordinates.

So that the general three-dimensional heat-conduction equation is

………….. (3)

9
For constant thermal conductivity, Equation (3) is written

…………….. (3a)

Where the quantity α = k/ρc is called the thermal diffusivity of the material.
The larger the value of α, the faster heat will diffuse through the material.
This may be seen by examining the quantities that make up α. A high value
of α could result either from a high value of thermal conductivity, which
would indicate a rapid energy-transfer rate, or from a low value of the
thermal heat capacity ρc. A low value of the heat capacity would mean that
less of the energy moving through the material would be absorbed and used
to raise the temperature of the material; thus more energy would be available
for further transfer. Thermal diffusivity α has units of square meters per
second.

In the derivations above, the expression for the derivative at x + dx has


been written in the form of a Taylor-series expansion with only the first two
terms of the series employed for the development.

Equation (3a) may be transformed into either cylindrical or spherical


coordinates by standard calculus techniques. The results are as follows:

Cylindrical coordinates:

......................... (3b)

Spherical coordinates:

………. (3c)

The coordinate systems for use with Equations (3b) and (3c) are indicated in
Figure 3b and c, respectively. Many practical problems involve only special
cases of the general equations listed above. As a guide to the developments
in future chapters, it is worthwhile to show the reduced form of the general
equations for several cases of practical interest.

10
Steady-state one-dimensional heat flow (no heat generation):

………….…….(4)

Note that this equation is the same as Equation (1) when q = constant.

Steady-state one-dimensional heat flow in cylindrical coordinates (no


heat generation):

…………………. (5)

Steady-state one-dimensional heat flow with heat sources:

……………….. (6)

Two-dimensional steady-state conduction without heat sources:

…..................... (7)

11
2. Thermal Conductivity:

To use Fourier’s law, the thermal conductivity of the material must be


known. This property, which is referred to as a transport property, provides
an indication of the rate at which energy is transferred by the diffusion
process. It depends on the physical structure of matter, atomic and
molecular, which is related to the state of the matter. In this section we
consider various forms of matter, identifying important aspects of their
behavior and presenting typical property values.

From Fourier’s law, the thermal conductivity associated with conduction in


the x-direction is defined as:

Similar definitions are associated with thermal conductivities in the y- and


z-directions, but for an isotropic material the thermal conductivity is
independent of the direction of transfer. kx =k y =k z =k

From the foregoing equation, it follows that, for a prescribed


temperature gradient, the conduction heat flux increases with increasing
thermal conductivity. In general, the thermal conductivity of a solid is larger
than that of a liquid, which is larger than that of a gas. As illustrated in
Figure 2.1, the thermal conductivity of a solid may be more than four orders
of magnitude larger than that of a gas. This trend is due largely to
differences in intermolecular spacing for the two states.

12
2.1

Equation (1) is the defining equation for thermal conductivity. On the


basis of this definition, experimental measurements may be made to
determine the thermal conductivity of different materials. For gases at
moderately low temperatures, analytical treatments in the kinetic theory of
gases may be used to predict accurately the experimentally observed values.
In some cases, theories are available for the prediction of thermal
conductivities in liquids and solids, but in general, many open questions and
concepts still need clarification where liquids and solids are concerned.

The mechanism of thermal conduction in a gas is a simple one. We


identify the kinetic energy of a molecule with its temperature; thus, in a
high-temperature region, the molecules have higher velocities than in some
lower-temperature region. The molecules are in continuous random motion,
colliding with one another and exchanging energy and momentum.
The molecules have this random motion whether or not a temperature
gradient exists in the gas. If a molecule moves from a high-temperature
region to a region of lower temperature, it transports kinetic energy to the
lower-temperature part of the system and gives up this energy through
collisions with lower-energy molecules.

13
Table 2-1 lists typical values of the thermal conductivities for several
materials to indicate the relative orders of magnitude to be expected in
practice. In general, the thermal conductivity is strongly temperature-
dependent.

Table 2.1

14
Other Relevant Properties:

In our analysis of heat transfer problems, it will be necessary to use


several properties of matter. These properties are generally referred to as
thermophysical properties and include two distinct categories, transport and
thermodynamic properties. The transport properties include the diffusion rate
coefficients such as k, the thermal conductivity (for heat transfer), and v, the
kinematic viscosity (for momentum transfer). Thermodynamic properties, on
the other hand, pertain to the equilibrium state of a system. Density ( ) and
specific heat (cp) are two such properties used extensively in thermodynamic
analysis. The product cp (J/m3.K), commonly termed the volumetric heat
capacity, measures the ability of a material to store thermal energy. Because
substances of large density are typically characterized by small specific
heats, many solids and liquids, which are very good energy storage media,
have comparable heat capacities ( ). Because of their very small
densities, however, gases are poorly suited for thermal energy storage
( ).

In heat transfer analysis, the ratio of the thermal conductivity to the


heat capacity is an important property termed the thermal diffusivity α ,
which has units of m2/s:

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, while materials of small α will
respond more sluggishly, taking longer to reach a new equilibrium
condition.

Example 2:1 The thermal diffusivity is the controlling transport property


for transient conduction. Using appropriate values of k, , and cp , calculate
for the following materials at the prescribed temperatures: pure aluminum,
300 and 700 K; silicon carbide, 1000 K; paraffin, 300 K.

15
Solution:

Known: Definition of the thermal diffusivity α.

Find: Numerical values of α for selected materials and temperatures.

Properties: Table A.1, pure aluminum (300 K):

Table A.1, pure aluminum (700 K):

Hence

Table A.2, silicon carbide (1000 K):

Table A.3, paraffin (300 K):

16
3. Convection Heat Transfer

The convection heat transfer mode is comprised of two mechanisms.


In addition to energy transfer due to random molecular motion (diffusion),
energy is also transferred by the bulk, or macroscopic, motion of the fluid.
This fluid motion is associated with the fact that, at any instant, large
numbers of molecules are moving collectively or as aggregates. Such
motion, in the presence of a temperature gradient, contributes to heat
transfer. Because the molecules in the aggregate retain their random motion,
the total heat transfer is then due to a superposition of energy transport by
the random motion of the molecules and by the bulk motion of the fluid. It is
customary to use the term convection when referring to this cumulative
transport and the term advection when referring to transport due to bulk fluid
motion.

We are especially interested in convection heat transfer, which occurs


between a fluid in motion and a bounding surface when the two are at
different temperatures. Consider fluid flow over the heated surface of Figure
3.1. A consequence of the fluid–surface interaction is the development of a
region in the fluid through which the velocity varies from zero at the surface
to a finite value u∞ associated with the flow. This region of the fluid is
known as the hydrodynamic, or velocity, boundary layer. Moreover, if the
surface and flow temperatures differ, there will be a region of the fluid
through which the temperature varies from Ts at y= 0 to T∞ in the outer flow.
This region, called the thermal boundary layer, may be smaller, larger, or
the same size as that through which the velocity varies. In any case, if Ts T∞
convection heat transfer will occur from the surface to the outer flow.

3.1

17
The convection heat transfer mode is sustained both by random
molecular motion and by the bulk motion of the fluid within the boundary
layer. The contribution due to random molecular motion (diffusion)
dominates near the surface where the fluid velocity is low. In fact, at the
interface between the surface and the fluid (y=0), the fluid velocity is zero
and heat is transferred by this mechanism only. The contribution due to bulk
fluid motion originates from the fact that the boundary layer grows as the
flow progresses in the x direction. In effect, the heat that is conducted into
this layer is swept downstream and is eventually transferred to the fluid
outside the boundary layer. Appreciation of boundary layer phenomena is
essential to understanding convection heat transfer. It is for this reason that
the discipline of fluid mechanics will play a vital role in our later analysis of
convection.

Convection heat transfer may be classified according to the nature of


the flow. We speak of forced convection when the flow is caused by external
means, such as by a fan, a pump, or atmospheric winds. As an example,
consider the use of a fan to provide forced convection air cooling of hot
electrical components on a stack of printed circuit boards (Figure 3.3a). In
contrast, for free (or natural) convection the flow is induced by buoyancy
forces, which are due to density differences caused by temperature variations
in the fluid. An example is the free convection heat transfer that occurs from
hot components on a vertical array of circuit boards in air (Figure 3.3b). Air
that makes contact with the components experiences an increase in
temperature and hence a reduction in density. Since it is now lighter than the
surrounding air, buoyancy forces induce a vertical motion for which warm
air ascending from the boards is replaced by an inflow of cooler ambient air

While we have presumed pure forced convection in Figure 3.3a and


pure natural convection in Figure 3.3b, conditions corresponding to mixed
(combined) forced and natural convection may exist. For example, if
velocities associated with the flow of Figure 3.3a are small and/or buoyancy
forces are large, a secondary flow that is comparable to the imposed forced
flow could be induced. In this case, the buoyancy-induced flow would be
normal to the forced flow and could have a significant effect on convection
heat transfer from the components. In Figure 3.3b, mixed convection would
18
result if a fan were used to force air upward between the circuit boards,
thereby assisting the buoyancy flow, or downward, thereby opposing the
buoyancy flow

3.3

We have described the convection heat transfer mode as energy


transfer occurring within a fluid due to the combined effects of conduction
and bulk fluid motion. Typically, the energy that is being transferred is the
sensible, or internal thermal, energy of the fluid. However, there are
convection processes for which there is, in addition, latent heat exchange.
This latent heat exchange is generally associated with a phase change
between the liquid and vapor states of the fluid. Two special cases of interest
in this text are boiling and condensation. For example, convection heat
transfer results from fluid motion induced by vapor bubbles generated at the
bottom of a pan of boiling water (Figure 3.3c) or by the condensation of
water vapor on the outer surface of a cold water pipe (Figure 3.3d).

19
Regardless of the particular nature of the convection heat transfer process,
the appropriate rate equation is of the form

………………… (3.1a)

Where, the convective heat flux (W/m2), is proportional to the difference


between the surface and fluid temperatures, Ts and T∞, respectively. This
expression is known as Newton’s law of cooling, and the parameter
h (W/m2.K) is termed the convection heat transfer coefficient. It depends on
conditions in the boundary layer, which are influenced by surface geometry,
the nature of the fluid motion, and an assortment of fluid thermodynamic
and transport properties.

Any study of convection ultimately reduces to a study of the means by


which h may be determined. Convection heat transfer will frequently appear
as a boundary condition in the solution of conduction problems. In the
solution of such problems we presume h to be known, using typical values
given in Table 3.1.

3.1

20
When Equation 3.1a is used, the convection heat flux is presumed to be
positive if heat is transferred from the surface (Ts T∞) and negative if heat is
transferred to the surface (T∞ Ts) . However, if (T∞ Ts) , there is nothing
to preclude us from expressing Newton’s law of cooling as:

…………. (3.1 b)

In which case heat transfer is positive if it is to the surface.

4. Radiation Heat Transfer

Thermal radiation is energy emitted by matter that is at a nonzero


temperature. Although we will focus on radiation from solid surfaces,
emission may also occur from liquids and gases. Regardless of the form of
matter, the emission may be attributed to changes in the electron
configurations of the constituent atoms or molecules. The energy of the
radiation field is transported by electromagnetic waves (or alternatively,
photons). While the transfer of energy by conduction or convection requires
the presence of a material medium, radiation does not. In fact, radiation
transfer occurs most efficiently in a vacuum.

Consider radiation transfer processes for the surface of Figure 4.1a.


Radiation that is emitted by the surface originates from the thermal energy of
matter bounded by the surface, and the rate at which energy is released per
unit area (W/m2) is termed the surface emissive power E. There is an upper
limit to the emissive power, which is prescribed by the Stefan–Boltzmann
law.

………………..(4.1)

Where Ts is the absolute temperature (K) of the surface and is the Stefan–
Boltzmann constant . Such a surface is called an
ideal radiator or blackbody.

The heat flux emitted by a real surface is less than that of a blackbody
at the same temperature and is given by

21
……………..(4.2)

4.1

Where is a radiative property of the surface termed the emissivity. With


values in the range , this property provides a measure of how
efficiently a surface emits energy relative to a blackbody. It depends
strongly on the surface material and finish.

Radiation may also be incident on a surface from its surroundings.


The radiation may originate from a special source, such as the sun, or from
other surfaces to which the surface of interest is exposed. Irrespective of the
source(s), we designate the rate at which all such radiation is incident on a
unit area of the surface as the irradiation G (Figure 4.1a).

A portion, or all, of the irradiation may be absorbed by the surface,


thereby increasing the thermal energy of the material. The rate at which
radiant energy is absorbed per unit surface area may be evaluated from
knowledge of a surface radiative property termed the absorptivity α. That is,

………………..(4.3)

where . If and the surface is opaque, portions of the


irradiation are reflected. If the surface is semitransparent, portions of the
irradiation may also be transmitted. However, while absorbed and emitted
radiation increase and reduce, respectively, the thermal energy of matter,
reflected and transmitted radiation have no effect on this energy. Note that
22
the value of α depends on the nature of the irradiation, as well as on the
surface itself. For example, the absorptivity of a surface to solar radiation
may differ from its absorptivity to radiation emitted by the walls of a
furnace.

In many engineering problems (a notable exception being problems


involving solar radiation or radiation from other very high temperature
sources), liquids can be considered opaque, and gases can be considered
transparent, to radiation heat transfer. Solids can be opaque (as is the case
for metals) or semitransparent (as is the case for thin sheets of some
polymers and some semiconducting materials).

A special case that occurs frequently involves radiation exchange


between a small surface at Ts and a much larger, isothermal surface that
completely surrounds the smaller one (Figure 4.1b). The surroundings
could, for example, be the walls of a room or a furnace whose temperature
Tsur differs from that of an enclosed surface (Tsur ≠ Ts). The irradiation may
be approximated by emission from a blackbody at Tsur, in which case
. If the surface is assumed to be one for which (a gray
surface), the net rate of radiation heat transfer from the surface, expressed
per unit area of the surface, is

……………….(4.4)

This expression provides the difference between thermal energy that


is released due to radiation emission and that which is gained due to
radiation absorption.

There are many applications for which it is convenient to express the


net radiation heat exchange in the form

……………..(4.5)

Where, from Equation 1.7, the radiation heat transfer coefficient hr is

………………… (4.6)

23
Here we have modeled the radiation mode in a manner similar to convection.
In this sense we have linearized the radiation rate equation, making the heat
rate proportional to a temperature difference rather than to the difference
between two temperatures to the fourth power. Note, however, that hr
depends strongly on temperature, while the temperature dependence of the
convection heat transfer coefficient h is generally weak.

The surfaces of Figure 4.1 may also simultaneously transfer heat by


convection to an adjoining gas. For the conditions of Figure 4.1b, the total
rate of heat transfer from the surface is then

………….(4.7)

Example: An uninsulated steam pipe passes through a room in which the air
and walls are at 25°C. The outside diameter of the pipe is 70 mm,
and its surface temperature and emissivity are 200°C and 0.8,
respectively. What are the surface emissive power and
irradiation? If the coefficient associated with free convection heat
transfer from the surface to the air is 15 W/m2.K, what is the rate
of heat loss from the surface per unit length of pipe?

Solution:
Known: Uninsulated pipe of prescribed diameter, emissivity, and surface
temperature in a room with fixed wall and air temperatures.
Find:

1. Surface emissive power and irradiation.


2. Pipe heat loss per unit length.

24
Schematic:

Assumptions:

1. Steady-state conditions.
2. Radiation exchange between the pipe and the room is between a small
surface and a much larger enclosure.
3. The surface emissivity and absorptivity are equal.

Analysis:

1. The surface emissive power may be evaluated from Equation 4.2,


while the irradiation corresponds to . Hence

2. Heat loss from the pipe is by convection to the room air and by
radiation exchange with the walls. Hence, q=qconv.+qrad. and from
Equation 4.7, with A=π DL,

The heat loss per unit length of pipe is then

25
Example 2: Air at 20◦C blows over a hot plate 50 by 75 cm maintained at
250◦C. The convection heat-transfer coefficient is 25 W/m2·◦C.
Calculate the heat transfer.

Solution: From Newton’s law of cooling

Example 3: Two infinite black plates at 800◦C and 300◦C exchange heat by
radiation. Calculate the heat transfer per unit area.

Solution: So we find immediately

Example 4: A horizontal steel pipe having a diameter of 5 cm is maintained


at a temperature of 50◦C in a large room where the air and wall
temperature are at 20◦C. The surface emissivity of the steel
may be taken as 0.8. Using the data of Table 1-3, calculate the
total heat lost by the pipe per unit length.

26
5. Steady-State Conduction - One Dimension

We now wish to examine the applications of Fourier’s law of heat


conduction to calculation of heat flow in some simple one-dimensional
systems. Several different physical shapes may fall in the category of one-
dimensional systems: cylindrical and spherical systems are one-dimensional
when the temperature in the body is a function only of radial distance and is
independent of azimuth angle or axial distance. In some two-dimensional
problems the effect of a second-space coordinate may be so small as to
justify its neglect, and the multidimensional heat-flow problem may be
approximated with a one-dimensional analysis. In these cases the differential
equations are simplified, and we are led to a much easier solution as a result
of this simplification.

5.1. The Plane Wall

First consider the plane wall where a direct application of Fourier’s


law [Equation (1)] may be made. Integration yields

(5.1)

When the thermal conductivity is considered constant. The wall thickness is


x, and T1and T2 are the wall-face temperatures. If the thermal conductivity
varies with temperature according to some linear relation k = k0(1 + βT ),
the resultant equation for the heat flow is

(5.2)

If more than one material is present, as in the multilayer wall shown in


Figure 5-1, the analysis would proceed as follows: The temperature
gradients in the three materials are shown, and the heat flow may be written:

Note that the heat flow must be the same through all sections.

27
Solving these three equations simultaneously, the heat flow is written

5.1

(5.3)

At this point we retrace our development slightly to introduce a different


conceptual viewpoint for Fourier’s law. The heat-transfer rate may be
considered as a flow, and the combination of thermal conductivity, thickness
of material, and area as a resistance to this flow. The temperature is the
potential, or driving, function for the heat flow, and the Fourier equation
may be written

(5.4)

a relation quite like Ohm’s law in electric-circuit theory. In Equation (5-1)


the thermal resistance is x/kA, and in Equation (5-3) it is the sum of the
three terms in the denominator. We should expect this situation in Equation
(5-3) because the three walls side by side act as three thermal resistances in
series. The equivalent electric circuit is shown in Figure 5-1b.

The electrical analogy may be used to solve more complex problems


involving both series and parallel thermal resistances. A typical problem and

28
its analogous electric circuit are shown in Figure 5-2. The one-dimensional
heat-flow equation for this type of problem may be written

(5.5)

Where the Rth are the thermal resistances of the various materials. The units
for the thermal resistance are ◦C/W or ◦F·h/Btu.

5.2

5.2. Insulation and R valued

In Section 1 we noted that the thermal conductivities for a number of


insulating materials are given in Appendix. In classifying the performance of
insulation, it is a common practice in the building industry to use a term
called the R value, which is defined as

29
(5.6)

The units for R are ◦C·m2/W or ◦F·ft2·h/Btu. Note that this differs from the
thermal resistance concept discussed above in that a heat flow per unit area
is used.

5.3. Radial Systems

Cylinders

Consider a long cylinder of inside radius ri, outside radius ro, and
length L, such as the one shown in Figure 5-3. We expose this cylinder to a
temperature differential Ti − To and ask what the heat flow will be. For a
cylinder with length very large compared to diameter, it may be assumed
that the heat flows only in a radial direction, so that the only space
coordinate needed to specify the system is r. Again, Fourier’s law is used by
inserting the proper area relation. The area for heat flow in the cylindrical
system is

So that Fourier’s law is written

(5.7)

30
5.3

5.4

31
With the boundary conditions

The solution to Equation (5-7) is

(5.8)

And the thermal resistance in this case is

The thermal-resistance concept may be used for multiple-layer cylindrical


walls just as it was used for plane walls. For the three-layer system shown in
Figure 5-4 the solution is

(5.9)

The thermal circuit is also shown in Figure 5-4.

Spheres

Spherical systems may also be treated as one-dimensional when the


temperature is a function of radius only. The heat flow is then

(5.10)

The derivation of Equation (5-10) is left as an exercise.

32
Example 5.1:

An exterior wall of a house may be approximated by a 4-in layer of common


brick [k = 0.7 W/m·◦C] followed by a 1.5-in layer of gypsum plaster
[k = 0.48 W/m·◦C]. What thickness of loosely packed rock-wool insulation
[k =0.065 W/m·◦C] should be added to reduce the heat loss (or gain) through
the wall by 80 percent?

Solution: The overall heat loss will be given by

Because the heat loss with the rock-wool insulation will be only 20 percent
(80 percent reduction) of that before insulation

We have for the brick and plaster, for unit area,

so that the thermal resistance without insulation is

Then

and this represents the sum of our previous value and the resistance for the
rock wool

So that

33
Example 5.2:

A thick-walled tube of stainless steel [18% Cr, 8% Ni, k = 19 W/m·◦C] with


2-cm inner diameter (ID) and 4-cm outer diameter (OD) is covered with a 3-
cm layer of asbestos insulation [k =0.2W/m·◦C]. If the inside wall
temperature of the pipe is maintained at 600◦C, calculate the heat loss per
meter of length. Also calculate the tube–insulation interface temperature.

Solution: Figure Example 5-2 shows the thermal network for this problem.
The heat flow is given by

5.2

This heat flow may be used to calculate


the interface temperature between
the outside tube wall and the insulation.
We have

Where Ta is the interface temperature,


which may be obtained as

The largest thermal resistance clearly results from the insulation, and thus
the major portion of the temperature drop is through that material.

34
Summary of One-Dimensional Conduction Results

Many important problems are characterized by one-dimensional,


steady state conduction in plane, cylindrical, or spherical walls without
thermal energy generation. Key results for these three geometries are
summarized in Table 5.3, where T refers to the temperature difference, Ts,1-
Ts,2, between the inner and outer surfaces identified in Figures 5.6, 5.7, and
5.8. In each case, beginning with the heat equation, you should be able to
derive the corresponding expressions for the temperature distribution, heat
flux, heat rate, and thermal resistance.

5.3

35
5.6

5.7

5.8

36
Convection Boundary Conditions

We have already seen in Section 1 that convection heat transfer can be


calculated from

(5.11)

An electric-resistance analogy can also be drawn for the convection process


by rewriting the equation as

Where the 1/hA term now becomes the convection resistance.

5.4. The Overall Heat-Transfer Coefficient

Consider the plane wall shown in Figure 5.9 exposed to a hot fluid A
on one side and a cooler fluid B on the other side. The heat transfer is
expressed by

The heat-transfer process may be represented by the resistance network in


Figure 5.9, and the overall heat transfer is calculated as the ratio of the
overall temperature difference to the sum of the thermal resistances:

(5.12)

5.9

37
Observe that the value 1/hA is used to represent the convection resistance.
The overall heat transfer by combined conduction and convection is
frequently expressed in terms of an overall heat-transfer coefficient U,
defined by the relation

(5.13)

Where A is some suitable area for the heat flow. In accordance with
Equation (5-12), the overall heat-transfer coefficient would be

The overall heat-transfer coefficient is also related to the R value of


Equation (5-6) through

For a hollow cylinder exposed to a convection environment on its inner and


outer surfaces, the electric-resistance analogy would appear as in Figure
5-10 where, again, TA and TB are the two fluid temperatures. Note that the
area for convection is not the same for both fluids in this case, these areas
depending on the inside tube diameter and wall thickness. The overall heat
transfer would be expressed by

(5.14)

in accordance with the thermal network shown in Figure 5-10. The terms Ai
and Ao represent the inside and outside surface areas of the inner tube. The
overall heat-transfer coefficient may be based on either the inside or the
outside area of the tube. Accordingly,

(5.15)

(5.16)

38
The general notion, for either the plane wall or cylindrical coordinate
system, is that

5.10

Example 5.3:

“Two-by-four” wood studs have actual dimensions of 4.13 × 9.21 cm and a


thermal conductivity of 0.1 W/m·◦C. A typical wall for a house is
constructed as shown Figure Example 5-3. Calculate the overall heat-
transfer coefficient and R value of the wall.

5.3

39
Solution: The wall section may be considered as having two parallel heat-
flow paths: (1) through the studs, and (2) through the insulation. We will
compute the thermal resistance for each, and then combine the values to
obtain the overall heat-transfer coefficient.

1. Heat transfer through studs (A = 0.0413 m2 for unit depth). This heat flow
occurs through six thermal resistances:

a. Convection resistance outside of brick

b. Conduction resistance in brick

c. Conduction resistance through outer sheet

d. Conduction resistance through wood stud

e. Conduction resistance through inner sheet

f. Convection resistance on inside

40
The total thermal resistance through the wood stud section is

2. Insulation section (A = 0.406 − 0.0413 m2 for unit depth). Through the


insulation section, five of the materials are the same, but the resistances
involve different area terms, i.e., 40.6 − 4.13 cm instead of 4.13 cm, so that
each of the previous resistances must be multiplied by a factor of 4.13/(40.6
− 4.13) = 0.113. The resistance through the insulation is

and the total resistance through the insulation section is

The overall resistance for the section is now obtained by combining the
parallel resistances in Equations (a) and (b) to give

This value is related to the overall heat-transfer coefficient by

Where A is the area of the total section = 0.406 m2. Thus,

As we have seen, the R value is somewhat different from thermal resistance


and is given by

41
Example 5.4:

A small metal building is to be constructed of corrugated steel sheet walls


with a total wall surface area of about 300 m2. The air conditioner consumes
about 1 kW of electricity for every 4 kW of cooling supplied1. Two wall
constructions are to be compared on the basis of cooling costs. Assume that
electricity costs $0.15/kWh. Determine the electrical energy savings of using
260 mm of fiberglass batt insulation instead of 159 mm of fiberglass
insulation in the wall. Assume an overall temperature difference across the
wall of 20◦C on a hot summer day in Texas.

Solution:

Consulting Table 5.2 (Numbers 19 and 20) we find that overall heat transfer
coefficients for the two selected wall constructions are

The heat gain is calculated from q = UA T, so for the two constructions

The energy consumed to supply this extra cooling is therefore

Extra electric power required = (840)(1/4) = 210 W

and the cost is

Cost = (0.210kW)(0.15$/kWh) = 0.0315 $/hr

Assuming 10-h/day operation for 23 days/month this cost becomes

(0.0315)(10)(23) = $7.25/month

42
Both of these cases are rather well insulated. If one makes a comparison to a
2 × 4 wood stud wall with no insulation (Number 4 in Table 5.2) fill in the
cavity (U = 1.85 W/m2·◦C), the heating load would be

and the savings compared with the 260-mm fiberglass insulation would be

11,100 − 1020 = 10,080 W

Producing a corresponding electric power saving of $0.378/h or


$86.94/month. Clearly the insulated wall will pay for itself. It is a matter of
conjecture whether the 260-mm of insulation will pay for itself in
comparison to the 159-mm insulation.

43
5.2

44
Example 5.5:

Water flows at 50◦C inside a 2.5-cm-inside-diameter tube such that hi =


3500 W/m2·◦C. The tube has a wall thickness of 0.8 mm with a thermal
conductivity of 16 W/m·◦C. The outside of the tube loses heat by free
convection with ho=7.6 W/m2·◦C. Calculate the overall heat-transfer
coefficient and heat loss per unit length to surrounding air at 20◦C.

Solution:

There are three resistances in series for this problem, as illustrated in


Equation (5-14). With L = 1.0 m, di = 0.025 m, and do = 0.025 + (2)(0.0008)
= 0.0266 m, the resistances may be calculated as

Clearly, the outside convection resistance is the largest, and overwhelmingly


so. This means that it is the controlling resistance for the total heat transfer
because the other resistances (in series) are negligible in comparison. We
shall base the overall heat-transfer coefficient on the outside tube area and
write

or a value very close to the value of ho =7.6 for the outside convection
coefficient. The heat transfer is obtained from Equation (a), with

45
5.5. Critical Thickness of Insulation

Let us consider a layer of insulation which might be installed around a


circular pipe, as shown in Figure 5-11. The inner temperature of the
insulation is fixed at Ti, and the outer

5.11

Surface is exposed to a convection environment at T∞. From the thermal


network the heat transfer is

(5.17)

Now let us manipulate this expression to determine the outer radius of


insulation ro, which will maximize the heat transfer. The maximization
condition is

Which gives the result

(5.18)

Equation (5-18) expresses the critical-radius-of-insulation concept. If the


outer radius is less than the value given by this equation, then the heat
transfer will be increased by adding more insulation. For outer radii greater

46
than the critical value an increase in insulation thickness will cause a
decrease in heat transfer. The central concept is that for sufficiently small
values of h the convection heat loss may actually increase with the addition
of insulation because of increased surface area.

Example 5.6:

Calculate the critical radius of insulation for asbestos [k = 0.17 W/m·◦C]


surrounding a pipe and exposed to room air at 20◦C with h = 3.0 W/m2·◦C.
Calculate the heat loss from a 200◦C, 5.0-cm-diameter pipe when covered
with the critical radius of insulation and without insulation.

Solution:

From Equation (5-18) we calculate ro as

The inside radius of the insulation is 5.0/2 = 2.5 cm, so the heat transfer is
calculated from Equation (5-17) as

Without insulation the convection from the outer surface of the pipe is

So, the addition of 3.17 cm (5.67 − 2.5) of insulation actually increases the
heat transfer by 25 percent.

As an alternative, fiberglass having a thermal conductivity of 0.04 W/m·◦C


might be employed as the insulation material. Then, the critical radius would
be

47

You might also like