Heat Transfer

You might also like

Download as pptx, pdf, or txt
Download as pptx, pdf, or txt
You are on page 1of 87

HEAT TRANSFER

•Heat transfer is the science that seeks to predict the energy transfer that may take place

between material bodies as a result of a temperature difference. Thermodynamics teaches that

this energy transfer is defined as heat.

•The science of heat transfer seeks not merely to explain how heat energy may be transferred, but also

to predict the rate at which the exchange will take place under certain specified conditions. The fact

that a heat-transfer rate is the desired objective of an analysis points out the difference between heat

transfer and thermodynamics.

•Thermodynamics deals with systems in equilibrium; it may be used to predict the amount of energy

required to change a system from one equilibrium state to another; it may not be used to predict how

fast a change will take place since the system is not in equilibrium during the process.
• Certainly, in designing a plant in which heat must be exchanged with the surroundings,
the size of heat transfer equipment, the materials of which it is to be constructed, and the
auxiliary equipment required for its utilization are all important considerations for
engineer. Not only must the equipment accomplish its required mission but it must also
be economical to purchase and to operate.

• Considerations of an engineering nature such as these require both a familiarity with the
basic mechanisms of energy transfer and an ability to evaluate quantatively these rates
as well as the important associated quantities. Our immediate goal is to examine the
basic mechanisms of energy transfer and to consider the fundamental equations for
evaluating the rate of energy transfer.
HEAT TRANSFER
REFERENCES :

1. W. McCabe, J. Smith, P. Harriott, Unit Operations of Chemical Engineering, 7th Ed.,

McGraw Hill, 2003.

2. C. Geankoplis, A. Hersel, D. H. Lepek, Transport Processes and Separation Process

Principles, Prentice-Hall, 5th Ed., 2018

3. A. K. Ray (Editor), Coulson and Richardson’s Chemical Engineering: Volume 2B: Separation

Processes,  6th Ed., 2018.

4. C. O. Bennett, J.E. Myers, Momentum, Heat and Mass Transfer, 3rd Ed., McGraw-Hill, New

York, 1982.

 
5. D. Green,  Marylee Z. Southard, Perry's Chemical Engineers' Handbook, 9th Ed.,
McGraw-Hill, New York, 2007.

6. Yunus Cengel and Afshin Ghajar, Heat and Mass Transfer: Fundamentals and Applications

7. R. B. Bird,  W. E. Stewart, E. N. Lightfoot, Transport Phenomena, 2nd Ed., Wiley, 2006.

8. A. S. Lavine, T. L. Bergman, F.P. Incropera, D.P. DeWitt, Fundamentals


of Heat and Mass Transfer, Wiley, 2019.

9. J.R. Welty, C.E. Wicks, R.E. Wilson, Fundamentals of Momentum, Heat and Mass
Transfer, 3rd Ed., Wiley, New York, 1984.

10. D. Green, M. Z. Southard,  Perry's Chemical Engineers' Handbook, 9th Ed., 2009.
HEAT TRANSFER AND ITS APPLICATIONS

• Practically all the operations that are carried out by the chemical engineer
involve the production and absorption of energy in the form of heat.

• The laws governing the transfer of heat and types of apparatus that have
for their main object the control of heat flow are therefore of great
importance. This section deals with heat transfer and its applications in
process engineering.

• When two objects at different temperatures are brought into contact, heat
flows from the object at the higher temperature to that at the lower
temperature. The net flow is always in the direction of the temperature
decrease.
The mechanisms by which the heat may flow are three : conduction, convection, and
radiation.
PRINCIPLES OF STEADY-STATE HEAT TRANSFER
INTRODUCTION AND MECHANISMS OF HEAT TRANSFER

• The transfer of energy in the form of heat occurs in many chemical


and other type of processes.

• Heat transfer often occurs in combination with other separation


processes, drying of lumber or foods, alcohol distillation, evaporation,
condensation, burning of fuel.

• Heat transfer occurs because of a temperature-difference driving force


and heat flows from the high to the low-temperature region.
• Writing for a general-property balance of momentum, thermal energy, or mass at
unsteady-state for specifically heat transfer :
- (1)
• Making an unsteady-state heat balance for the x direction only on the volume element
or control volume in Fig. 1, with the cross-sectional area A :
Δx.A) =  + ρ Δx.A) ……….. (1)

Area A
𝒒𝒙 ∨ 𝒙 𝒒 𝒙 ∨ 𝒙 +  Δx Fig. 1 Unsteady-state balance for
heat transfer in control volume

x 𝒙+ Δ x
where is rate of heat generated per volume. Assuming no heat generation and also
assuming steady-state heat transfer, where the rate of accumulation is zero. Eqn (1)
becomes :

= ……. (2)

This means that the rate of heat input by conduction equals to the rate of heat output by
conduction; or is a constant with time.

In this part we are concerned with a control volume where the rate of accumulation of
heat is zero and we have steady-state heat transfer. The rate of heat transfer is then
constant with time, the temperature at various points in the system do not change with
time.
BASIC MECHANISMS OF HEAT TRANSFER

• Heat Transfer may occur by any one or more of the three basic
mechanisms of heat transfer : conduction, convection, and radiation.

1. Conduction : In conduction, heat can be conducted through solids, liquids,


and gases. The heat is conducted by the transfer of the energy of motion
between adjacent molecules. In a gas the ‘’hotter’’ molecules, which have
greater energy and motions, impart energy to the adjacent molecules at
lower energy levels. This type of transfer is present to some extent in all
solids, gases, or liquids, in which a temperature gradient exists.
• Heat flow of this kind is called conduction, and according to Fourier’s law, heat flux is proportional to
the temperature gradient and opposite to it in sign. For one dimensional heat flow, Fourier’s law is :
= -k
where q : rate of heat flow in direction normal to the surface
A : surface area
T : tempeature
x : distance normal to surface
k :proportionality constant or thermal conductivity

2. Convection : The transfer of heat by convection implies the transfer of heat by bulk transport
and mixing of macroscopic elements of warmer portions with cooler portions of a gas or liquid. It
also often refers to the energy exchange between a solid surface and a fluid. The second meaning is
more important for unit operations, as it includes heat transfer from metal walls, solid particles,
and liquid surfaces.
• The convective flux is usually proportional to the difference between the surface
temperature and the temperature of the fluid, as stated Newton’s law of cooling :

h …… (4)
where : surface temperature; : bulk temperature of fluid, far from surface
h : heat transfer corfficient

Unlike thermal conductivity, the heat transfer coefficient is not intrinsic property of the
fluid, but depends on the flow patterns determined by fluid mechanics as well as on the
thermal properties of fluid.

• A distinction must be made between forced-convection heat transfer, where a fluid is


forced flow past a solid surface by a pump, fan, or other mechanical means, and natural
or free convection, where warmer or cooler fluid next to the solid surface causes
a circulation because of a density difference resulting from the temperature differences in
the fluid. Examples of heat transfer by convection are loss of heat from a car radiator
where the air is circulated by a fan, cooking of foods in a vessel being stirred, cooling of
a hot cup of coffee by blowing over the surface, and so on.

3. Radiation : Radiation differs from heat transfer conduction and convection in that no
physical medium is needed for its propagation. Radiation is a term given to the transfer of
energy through space by electromagnetic waves. If radiation is passing through empty
space, it is not transformed to heat or any other form of energy, nor is it diverted from its
path. If, however, matter appears in its path, radiation will be transmitted, reflected, or
absorbed. It is the only absorbed energy that appears as heat, and this transformation is
quantitative. For example, fused quartz transmits practically all the radiation
that strikes it; a polished opaque surface or mirror will reflect most of the radiation impinging
on it; a black or matte surface will absorb most of the radiation received by it and will
transform such absorb energy quantitatively to heat.

The energy emitted by a black body is proportional to the fourth power of the absolute
temperature :

….. (5)
where : rate of radiant energy emission per unit area
: Stefan-Boltmann constant
: absolute temperature

Monatomic and most diatomic gases are trandparent to thermal radiation, and it is quite
common to find that heat is flowing through masses of such gases both by radiation and by
conduction-convection.
• Some examples to the radiation are :
• The loss of heat from a radiator or uninsulated steam pipe to the air of
a room

• Heat transfer in furnaces and other high-temperature heating equipment

• The transport of heat to the earth from the sun

• Cooking of food when passed below red-hot electric heaters

• Heating of fluids in coils of tubing inside a combustion furnace


FOURIER’S LAW OF HEAT CONDUCTION

For the general molecular transport equation, all three main types of rate-
transfer processes-momentum transfer, heat transfer, and mass transfer-are
characterized by the same general type of equation. The transfer of electric
current can also be included in this category. This basic equation is :

Rate of a transfer process = … (6)

• This equation states what we know intuitively : that in order transfer

a property such as heat or mass, we need a driving force to

overcome a resistance.
The transfer of heat by conduction also follows this basic equation and is
written Fourier’s law for heat conduction in fluids or solids:

= -k … (7)
where : the heat transfer rate in the x direction, in watts (W); A : the cross-

sectional area normal to the direction of flow of heat in (); T : temperature in K;

x : distance in m; k : thermal conductivity in W/m.K (SI). The quantity is called

the heat flux in W/dT/dx : temperature gradient in the x direction. The minus

sign (-) in eqn (7) is required, because if the heat flow is positive in a given

direction, the temperature decreases in this direction.


CONVERSION FACTORS :
T; hermal conductivity (k) : W/m.K; Btu/h.ft. F; kcal/h.m.C
= 1.73073 W/m.K
For the heat flux and power :
1 Btu/h. = 3.1546 W/

Fourier’s law, eqn (7), can be integrated for the case of steady-state heat
transfer through a flat wall of constant cross-sectional area A, where the
inside temperature is at point 1 and at point 2, a distance ofaway.
Rearranging eqn (7),
= -k … (8)
Integrating, assuming that k is constant and does not vary with temperature and dropping
the subscript x on
… (9)
Example 1. Heat Loss Through an Insulating Wall.
Calculate the heat loss per of surface area for an insulating wall composed of
25.4 mm thick fiber insulating board, where the inside temperature is 352.7 K
and outside temperature is 291.7 K. Thermal conductivity of fiber insulating
board is 0.048 W/m.K.
Solution : The thickness of insulating board -=0.0254 m. Substituting into the following
equation, = = 105.1; q/A= 105.1 W/
Thermal Conductivity
As seen in Table 1, gases have quite low of thermal conductivities, liquid intermediate
values, and solid metals very high values.
Table 1. Thermal Conductivities of Some Materials at 101.325 kPa (1 atm) Pressure (k in W/m.K)
Substance T (K) k Substance T (K) k
GASES SOLIDS
Air 273 0.0242 Ice 273 2.25
373 0.0316 Fire claybrick 473 1.00
273 0.167 Paper - 0.130
n-Butane 273 0.0135 Hard rubber 273 0.151
LIQUIDS Cork board 303 0.043
Water 273 0.569 Asbestos 311 0.168
366 0.680 Rock wool 266 0.029
Benzene 303 0.159 Steel 291 45.3
Biological mat., foods 373 45
333 0.151 Copper 273 388
373 377
Thermal conductivity of gases
1. In gases, the mechanism of thermal conduction is relatively simple. The molecules are in
continuous random motion, colliding with one another and exchanging energy and
momentum. If a molecule moves from a high-temperature region to a region of lower
temperature, it transports kinetic energy to this region and gives up this energy through
collisions with lower-energy molecules. Since smaller molecules move faster, gases such
as hydrogen should have higher thermal conductivities, as shown in Table 1. Theories to
predict thermal conductivities of gases are reasonably accurate and are given elsewhere
(R1). However, for completeness, the predictive model developed by Chapman–Enskog
for gases is given below,

2. (10)
k : thermal conductivity (W/m.K)
T : temperature (K)
, )
: collision diameter (dimensionless)

2. Thermal conductivity of liquids

The physical mechanism of energy conduction in liquids is somewhat similar to that of gases,
where higher-energy molecules collide with lower-energy molecules. However, the molecules
are packed so closely together that molecular force fields exert a strong effect on the energy
exchange. Since an adequate molecular theory of liquids is not available, most correlations to
predict the thermal conductivities are empirical.
… (11)
where T : temperature; a and b are empirical constants. Thermal conductivities (k) of liquids are
essentially independent of pressure. The thermal conductivity of liquids varies moderately with
temperature and often can be expressed as a linear variation.

Water has a high thermal conductivity compared to organic-type liquids such as benzene. As
shown in Table 1, the thermal conductivities of most unfrozen foodstuffs, such as skim milk and
applesauce, which contain large amounts of water, have thermal conductivities near that of pure
water.

3. Solids : The thermal conductivity of homogeneous solids varies quite widely, as may be seen for
some typical values in Table 1. The metallic solids of copper and aluminum have very high thermal
conductivities, while some insulating nonmetallic materials such as rock wool and corkboard have
very low conductivities.
Thermal conductivities of insulating materials such as rock wool approach that of air since
the insulating materials contain large amounts of air trapped in void spaces. Superinsulations
for insulating cryogenic materials such as liquid hydrogen are composed of multiple layers of
highly reflective materials separated by evacuated insulating spacers.

Values of thermal conductivity are considerably lower than for air alone.

Convection Heat Transfer Coefficient

It is well known that a hot piece of material will cool faster when air is blown or forced
past the object. When the fluid outside the solid surface is in forced or natural convective
motion, we express the rate of heat transfer from the solid to the fluid, or vice versa, by the
following equations:
where q is the heat-transfer rate in W, A is the area in , is the temperature of the solid
surface in K, is the average or bulk temperature of the fluid flowing past in K, and h is
the convective heat-transfer coefficient in K. In English units, h is in Btu/h · °F.

• The coefficient h is a function of the system geometry, fluid properties, flow


velocity, and temperature difference. In many cases, empirical correlations are
available to predict this coefficient, since it often cannot be predicted theoretically.
Because we know that when a fluid flows past a surface there is a thin, almost
stationary layer or film of fluid adjacent to the wall that presents most of the
resistance to heat transfer, we often call the coefficient h a film coefficient.
Table 12.4-1. Approximate Magnitude of Some Heat-Transfer Coefficients
Ranges of h values
Mechanism Btu/h. .F W/K
Condensing steam 1000-5000 5700-28.000
Condensing organics 200-500 1100-2800
Boiling liquids 300-5000 1700-28000
Moving water 50–3000 280–17000
Moving hydrocarbons 10–300 55–1700
Still air 0.5–4 2.8–23
Moving air 2–10 11.3–55
STEADY-STATE HEAT CONDUCTION

In this section the Fourier conduction equation will be used in the analysis of one-
dimensional, steady-state heat flow in solid systems that are geometrically simple. The flat
wall, hollow cylinder and hollow sphere, in addition being simple, are the commonest
shapes with which engineers must deal.
Conduction Through a Flat Slab or Wall
In the application of the Fourier conduction equation
(13)
to a flat wall, it is apparent that if heat is flowing normal to the principal surfaces, area
term A is constant. Furthermore, if the conductivity is assumed to be constant, then the q
at any cross section is proportional to the temperature gradient dT/dx. If energy is neither
generated nor accumulated in the wall, q is identical at all cross sections and so is dT/dx.
A sketch of the system and graphical representation of its temperature profile are given in
Fig. 2. Distance, x

Temperature, T
𝑇1 𝑇1

𝑇2 𝑇2

Fig. 2 Heat conduction in a flat wall


Δx 0 Δ𝒙
Distance,
If the thermal conductivity (k) varies with temperature, dT/dx is not constant. If heat flow and area
remain constant at all cross sections and k increases with decreasing temperature, the temperature
gradient must diminish in the direction of decreasing temperature. Thus curve representing temperature in
steady-state flow for the system concave upward.
The integration of equation dx is readily performed when q, k, and A are constant, and
gives
(14)
The equation (11) can also be integrated and solved for the temperature T at a point x.
(15)
This equation indicates the linearity of temperature with distance when the conditions are
as stated.

If the thermal conductivity is not constant but varies linearly with temperature, then
substituting Eqn. (11) into Eqn. (13) and integrating,

()) … (16)
=)/2 … (17)
This means that the mean of k (i.e., ) to use in Eqn. (16) is the value of k evaluated at the
linear average of and ).
As stated in the introduction to transport processes in Eqn. (14) can be written in that
form :
= = ... (18)
Where R = and is the resistance in K/W or h °F /Btu.
EXAMPLE : Flat Wall Conduction : A layer of pulverized cork 152 mm thick is
used as a layer of thermal insulation in a flat wall. The temperature of the cold side of the
cork is 4.4 °C and that of the warm side is 82.2 °C. The thermal conductivity of the cork at
0 °C is 0.036 W/m.K, and that at 93.3 °C is 0.055 W/m.K. The area of the wall 2.32 . What
is the rate of heat flow through wall in Watts and Btu/h?
Solution : The arithmetic average temperature of the cork is : +)/2=(4.4+82.2)/2=43.3 °C
By linear interpolation the thermal conductivity at 43.3 °C is :
= = 0.0448; = 0.0448 W/m.K
Also, A= 2.32 ; = 0.152 m; Substituting in Eqn. (18) :
= 53.2 W
53.2W= 53.2

Multilayer Flat Wall


When heat is conducted is conducted through a flat wall consisting of layers of different
substances, a situation is produced comparable with an electrical system in which a
number of resistances are connected in series. The thermal system is shown in Fig. 3.
Fig. 3 Heat flow a multilayer flow

The Fourier equation can be written for each of these


layers, and for steady-state the heat flow q is the same for
all three walls. By analogy with Eq.(14) we can write,

(19)

The overall temperature difference is usually specified in engineering systems, so an equation


for heat flow as a function of will be derived. Rewriting Eq. (19) we get,
=q
=q
=q
Adding these equations gives
=q or

or (20)
If the integrated form of the Fourier equation (14) is considered analogous to Ohm’s law for
electric conduction, then the quantity is a measure of the resistance to heat flow. The
denominator of Eq. (20) is the overall resistance, which is the sum of the individual resistances.
Figure 3 shows temperature gradients for the different layers of solids. Because q/A is the
same for all layers, it follows that is the same for all layers; thus is inversely proportional to
the thermal conductivity.

An important application of heat conduction through multilayer walls occurs in the use of
fins which are attached by crimping, soldering, or welding to the outer surface of a tube to
increase the heat transfer rate.

EXAMPLE : Multilayer Flat Wall: A flat furnace wall is constructed of a 114-mm layer of Sil-o-cel
brick, with a thermal conductivity of 0.138 W/m.K backed by 229-mm layer of common brick, of
conductivity 1.38 W/m.K. The temperature of the inner face of the wall is 760 °C, that of the outer face is
76.6 °C. a) What is the heat loss through the wall? b) What is the temperature of the interface between the
refractory brick and the common brick? c) Supposing that the contact between the two brick layers is poor
and that a ‘’contact resistance’’ of 0.088 K.m2/W is present, what would be the heat loss?
a) ; = 0.114 m; = 0.229 m; = 0.138 W/m.K; = 1.38 W/m.K;
Substituting in Eqn. (20) :
= = 688.9 W/m2
b) = =688.9; =688.9()=569.1
= 190.9 (The interface temperature between the two bricks)
c) The thermal resistance : [m/W/(m.K)]= [m2.K/W]=0.088 m2.K/W
The total resistance which now includes a contact resistance R’, is : +R’
+R’= +0.088=1.080 m2.K/W

= = 632.78; = 632.8 W/m2


EXAMPLE : Multilayer Flat Wall-Heat Flow Through an Insulated Wall of Cold Room
A cold-storage room is constructed of an inner layer of 12.7 mm of pine, a middle layer of 101.6
mm of cork board, an outer layer of 76.2 mm of concrete. The wall surface temperature is 255.4 K
inside the cold room and 297.1 K at the outside surface of the concrete. Thermal conductivities for
pine, 0.151, for cork board, 0.0433; and for concrete, 0.762 W/m.K. Calculate the heat loss in 1 m 2
and the temperature at the interface between the wood and cork board.

Solution : Calling T1=255.4 K, T4= 297.1K, pine as material A, cork as B, and concrete as C, a
tabulation of the properties and dimensions are as follows:

= 0.151; = 0.0433; = 0.762;

RA = = 0.0841 K/W

RB = = 2.346 K/W
RC = = 0.100 K/W
Substituting into Eq. (20),
==-16.48 W

-16.48 W; since the answer is negative, heat flows in from the outside.
• To calculate the temperature at the interface between the pine wood and cork,
q= = =-16.48; = -16.48(0.0841)=-1.38597;
= 255.4+1. 38597=256.79
=
• An alternative procedure for calculating is to use the fact that the temperature drop
is proportional to the resistance :
=; substituting Eq. (21),

==-1.39 K; =256.8 K
Hollow Cylinder
The condition of heat in the walls of a hollow cylinder can be described mathematically by the
Fourier equation written in rectangular coordinates, but the equation is usually written in
cylindrical coordinates for convenience. For the example shown in Fig. 4, the equation is,
(22)

Fig. 4 Heat conduction in a hollow cylinder

The area normal to heat flow is

The equation can be integrated and solved for the heat transfer
rate q in terms of the boundary conditions:
q
(24)
This equation (24) can be multiplied by - in both numerator and denominator to produce
(25)
(26) : thermal conductivity;

The area term in Eq. (26), Alm, is the log mean area . The temperature difference T is T1-T2

instead of T2-T1. This usage is contrary to matematical practice but is conventional in


writing integrated heat-transfer equations for which the direction of heat flow is usually
clearly understood. Equation (26) is similar in form to Eq.(14) for conduction in a flat wall,
the principal difference being in the area terms. In most engineering application (e.g., pipes),
In these circumstances the arithmetic mean area may be used in Eq. (26), with a consequent area in q less than
4 per cent.

Integration of Eqn. (22) to give a relation for the temperature T at any radial position r shows that the T is
linear function of lnr, rather than of r as in the of a flat wall.

Multilayer Cylinder

In the process industries, heat transfer often occurs through multilayers of cylinders, as for
example when heat is being transferred through the walls of an insulated pipe. Figure 13.1-3
shows a pipe with two layers of insulation around it, that is, a total of three concentric hollow
cylinders. The temperature drop is T1 – T2 across material A, T2 – T3 across B, and T3 – T4 across
C.

The heat-transfer rate q will, of course, be the same for each layer, since we are at steady state.
Writing an equation similar to Eq. (13.1-9) for each concentric cylinder,
== (1)
where
=; (2)

Using the same method of combining the equations to eliminate T 2 and

T3 as was done for the flat walls in series, the final equations are
q= (3)

q= = (4)

Hence, the overall resistance is again the sum of the individual


resistances in series.
Example 1: Multilayer Cylinder
A thick-walled tube of stainless steel [18% Cr, 8% Ni, k = 19 W/m.K] with 2-cm inner diameter
(ID) and 4-cm outer diameter (OD) is covered with a 3-cm layer of asbestos insulation [k = 0.2
W/m.K]. 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. The outside
temperature of the asbestos insulation is 100 ◦C.
Solution : Figure Example 2-2 shows the thermal network for this problem.
The heat flow is given by

q = 683.0 W
r1=20 mm/2=10 mm=0.010 m; r2=40 mm/2=20 mm
r3=r2+30 mm=20 mm+30 mm=50 mm=0.050 m
A1=1L=2
A2=2L=2
A3=3L=2
=(A2-A1)/ln(A2/A1) = (ln()=0.090647 m2
=(A3-A2)/ln(A3/A2) = ( ln()=0.205716
= 0.205716 m2
q/L=683.0 W/1 m = 683.0 W/m
= = 683.0; = 683.0() = 498.0
= 498.0 =100+498.0 =598.0 ◦C; = 598.0 ◦C ( 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.
Example : Heat Loss from an Insulated Pipe (Multilayer Cylinder)
A thick-walled tube of stainless steel (A) having a k = 21.63 W/m.K with dimensions of
0.0254 m ID and 0.0508 m OD is covered with a 0.0254-m-thick layer of an insulation (B),

k = 0.2423 W/m.K. The inside-wall temperature


of the pipe is 811 K and the outside surface of the insulation is at 310.8 K. For a 0.305-m
length of pipe, calculate the heat loss and also the temperature at the interface between the
metal and the insulation.
Solution : Calling T1 = 811 K, T2 the interface, and T3 = 310.8 K, the dimensions are
From Eq. (2), the log mean areas for the stainless steel (A) and insulation (B) are :

From Eq. (3), the resistances are :

Hence, the heat-transfer rate is

Solving, 811 – T2 = 5.5 K and T2 = 805.5 K. Only a small


To calculate the temperature T2, temperature drop occurs across the metal wall because of its high
thermal conductivity.
THE PLANE WALL

First consider the plane wall where a direct application of Fourier’s law may be made.
Integration yields.

If more than one material is present, as in the multilayer wall shown in Figure 2-1, 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.
INSULATION AND R VALUES
In Chapter 1 we noted that the thermal conductivities for a number of insulating materials
are given in Appendix A. 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
Figure 2-2 Series and parallel
one-dimensional heat transfer
through a composite wall and
electrical analog.
Conduction Heat Transfer Through a Hollow Sphere
Steady-state conduction normal to the walls of a hollow sphere can be readily analyzed
and expressed in an equation analogous to the Eq. for flat wall, and equation The
equation is used as the starting point in the analysis; area normal to heat flow is 4πr2. For
the case of a hollow sphere with temperature T1 at inner radius r1, and temperature T2 at

outer radius r2.

q = -4πk
T1
r1
r2 T2 =4;πk; ;

q= : the geometric mean area


q
Integration of the eqn. q = -4πk to give the temperature T at any radial position r inside
the wall of the sphere shows that of the sphere shows that the temperature is a linear
function of 1/r.
Multilayer Spere
Analysis of a multilayer, spherical system is made in the same manner as for cylinder and
flat surfaces. For a system of three concentric spherical layers of materials A, B, and C, we
obtain

, , and : geometrical mean areas for materials A, B, and C.


=; =; =;
Combined Convection, Conduction, and Overall Coefficients
In many practical situations, the surface temperatures (or boundary conditions at the surface)
are not known, but there is a fluid on both sides of the solid surfaces. Consider the plane wall in
Fig. 13.2-1a, with a hot fluid at temperature T1 on the inside surface and a cold fluid at T4 on the

outside surface. The convective coefficient on the outside is ho W/m2.K, and hi on the inside.
(Methods for predicting the convective h will be given in Chapter 15.)
The heat-transfer rate using Eqs. (12.4-1) and (12.6-1) is given as

Expressing 1/hiA, ΔxA/kAA, and 1/hoA as resistances, and combining the equations as
before,
A more important application is heat transfer from a fluid outside a cylinder, through a
metal wall, to a fluid inside the tube, as often occurs in heat exchangers. In Fig. 13.2-1b,
such a case is shown.
where Ai represents 2πLri, the inside area of the metal tube; AA lm the log mean area of the

metal tube; and Ao the outside area. The overall heat-transfer coefficient U for the

cylinder may be based on the inside area Ai or the outside area Ao of the tube. Hence,
Using the same procedure as before, the overall heat-transfer rate through the cylinder is
Figure 13.2-1. Heat flow with convective boundaries: (a) plane wall, (b) cylindrical wall.
where Ai represents 2πLri, the inside area of the metal tube; AA lm the log mean area of the

metal tube; and Ao the outside area. The overall heat-transfer coefficient U for the cylinder

may be based on the inside area Ai or the outside area Ao of the tube. Hence,
Overall Heat Transfer Coefficients
The problem of conduction through a series of resistances has already been considered. It is
now appropriate to study systems in which heat is transferred by a series of conduction and
convection mechanisms.

Figure 21.5 illustrates a system in which heat flows from a fluid at a bulk temperature T1,

through the pipe Wall, through a layer of insulation, and finally into a fluid at bulk

temperature T5. The interfacial temperatures T2, T3, and T4 are all as shown.
T3 T2
T4

T1

T5

Figure 21.5 Heat transfer through an insulated pipe

Heat transfer coefficients hi and ho refer to the inside and outside coefficients,
respestively. The steady-state flow rate is
=
The individual temperature drops can be found by rearranging (21.12),

These equations are added and rearranged to give


(21.13)

Each term in the denominator can be considered as a resistance, so that Eq. (21.13) can be
written as
q= (21.14)
However, more common procedure is to multiply the right-hand side of Eq. (21.13) by
either / or /. Using the latter term, we obtain
(21.15)

The quantity is called the overall coefficient of heat transfer based


on the outside area and is usually designated as Uo. If Eq. (21.13) had been multiplied
by /, a quantity called Ui, the overall coefficient of heat transfer based on the inside area,
would have resulted. This quantity is defined as
Ui = (21.16)
The definitions of these quantities, Uo and Ui, make it possible to write
(21.17)
(21.18)

The ratios , etc., can be replaced by the ratios of the radii . In many

cases these ratios are nearly unity and can be taken as such. This is about the only basis
for deciding whether to calculate an outside area or an inside coefficient. For example, if
the inside resistance is much greater than the other three resistances (as shown by hi being

much smaller than ho, kB/ , and kC/ ), one can write
= (21.19)
as a satisfactory approximation. The basis for doing this is that the last three terms of the

denominator are small in comparison with 1/hi , so that the ratios of the radii would

represent negligible corrections. In certain problems it is reasonable to assume that


all the terms but one can be ignored, and we write or , depending on which term remains.

The illustration chosen here contains four resistances in series. However the overall
coefficient can be used to represent as many resistances in series as desired. Such cases can
be analyzed using the procedures followed here, but the nature of additinal terms in the
overall coefficient is rather obvious, so that these terms can be written by analogy with
those given in Eqs. (21.5) and (21.6).

If the system is one of flat, parallel walls, the areas cancel, so that the expression

(21.20)
is rigorous.

The overall coefficient could be defined in terms of one of the log mean areas, but this is
seldom done. In fact, when tabulations of overall coefficients are encountered, they
frequently do not state whether the values given are for Uo or Ui because the uncertainty in

the individual coefficients is usually assumed to be Uo.

Tabulations of overall coefficients in Perry and Nelson. The sample in Table 21.2 is
taken from Perry.
Fouling Coefficients
Tendency of certain fluids to form deposits on heat transfer surfaces is a serious problem in
the design of heat-exchange equipment. It can happen that the thermal resistance due to
the deposit is greater that the sum of all the other resistances.
In some cases a hard scale is deposited on heat-transfer surfaces in a boiler or
evapoarator. Coke is often deposited inside the tubes of oil heaters in a refinery. The
resistance of such layers can be represented using a heat transfer coefficient equal to the
thermal conductivity of the scale divided by its thickness. This type of scale can be removed
by sandblasting, by the use of pneumatic cleaning tools, and occasionally by pumping a
chemical cleaning solution through the equipment.

Another type of fouling is the porous deposit formed from mud, soot, and even vegetable
matter. The thermal conductivities of all these materials may be high, but the fluid contained
in the pores frequently has a much lower thermal conductivity. The effective thermal
conductivity, therefore, may be almost as low as that of the fluid. These deposits can
sometimes be removed by blowing with steam, air, or hot water. The growth of the vegetable
material in condensers isinhibited by chlorinated water.
Table 21-2 Overall Coefficients of Heat Transfer
U, W/m2.K

Stabilizer reflux condenser 534 (94 Btu/h.ft2.oF)


Oil preheater 613 (108 Btu/h.ft2.oF)
Reboiler (condensing steam to boiling 1700-4540 (300-800 Btu/h.ft2.oF)
water)
Air heater (molten salt toheater) 34 (6 Btu/h.ft2.oF)
Steam-jacketed vessel evaporating milk 2840 (500 Btu/h.ft2.oF)

The presence of fouling deposits is anticipated in the design of heat exchangers by the
use of a fouling coefficient which has the form of heat transfer coefficient. An equation for
the heat flux through the scale is written
(21.21)
If the fouling resistance is incorporated into the anlysis of series of resistances, the final
equation for heat transfer is the same as Eqs. (21.17) and (21.18). However, the overall
coefficient contains an extra term and is written
(21-22)
The fouling coefficient for the inside surface is hdi. If fouling occurred simultaneously on
the outer surface, the term would need to be added to the denominator.

Tabulation of fouling coefficients are given in many references. A sample tabulation of


fouling coefficients ise given in Table 21.3.

It is apparent that the fouling resistance will usually increase with time until cleaning is
necessary; however, there are some cases for which the resistance ceases to increase after
some time because rate of deposition is balanced by the rate of removal of scouring. No
indication of the effect of time is given in Table 21-3.
Table 21-3 Fouling Coefficients
hd, W/m2K (Btu/h.ft2.oF)
Overhead vapors from crude-oil 5680 (1000)
distillation
Dry crude-oil (150-38 oC)
• Velocity 0.61 m/s 1420 (250)
• Velocity 0.61-1.2 m/s 1874 (330)
• Velocity 1.2 m/s 1874 (330)
Air 2840 (500)
Steam (non-oil bearing) 1136 (200)
Water, Great Lakes, over 52oC 2840 (500)
Example 21-2 Overall Heat Transfer Coefficients
A reflux condenser contains ¾ in., 16 gauge copper tubes (ID = 0.620 in., OD = ¾ in.) in
which cooling water circulates. Hydrocarbon vapors condense on the exterior surfaces of
the tubes. Find the overall heat transfer coefficient Uo. The inside convective coefficient
can be taken as 4540 W/m2K, and the outside coefficient as 1420 W/m2K. The appropriate
fouling coefficients from Table 21-3 are hdo = 5680 and hdi =2840 W/m2K.
Solution : The following quantities can be calculated for use in Eq. (21-22)
(21-22)
(21-22a)
= 19.05 mm;
Critical Thickness of Insulation for a Cylinder
In Fig. 13.2-3, a layer of insulation is installed around the outside of a cylinder whose
radius r1 is fixed and with a length L. The cylinder has a high thermal conductivity and the

inner temperature T1 at point r1 outside the cylinder is fixed. An example is the case where
the cylinder is a metal pipe with saturated steam inside. The outer surface of the insulation
at T2 is exposed to an environment at T0 where convective heat transfer occurs. It is not
obvious if adding more insulation with a thermal conductivity of k will decrease the heat-
transfer rate.

At steady state, the heat-transfer rate q through the cylinder and the insulation equals
the rate of convection from the surface:
As insulation is added, the outside area, which is A = 2πr2L, increases, but T2 decreases.
However, it is not apparent whether q increases or decreases. To determine this, an
equation similar to Eq. (13.2-5), with the resistance of the insulation represented by Eq.
(13.1-11), is written using the two resistances:
(13.2-24)
To determine the effect of the thickness of insulation on q, we take the derivative of q
with respect to r2, equate this result to zero, and obtain the following for maximum heat
flow:
=0 (13.2-25)
Solving,
k/ho (13.2-26)

where (r2)cr is the value of the critical radius when the heat-transfer rate is a maximum. Hence, if

the outer radius r2 is less than the critical value, adding more insulation will actually increase the
heat-transfer rate q. Also, if the outer radius is greater than the critical, adding more insulation will
decrease the heat-transfer rate. Using values of k and ho typically encountered, the critical radius is
only a few mm. As a result, adding insulation on small electrical wires could increase the heat loss.
Adding insulation to large pipes decreases the heat-transfer rate.

Figure 13.2-3. Critical radius for insulation of a cylinder or a pipe.


EXAMPLE 13.2-3. Insulating an Electrical Wire and Critical Radius
An electric wire having a diameter of 1.5 mm and covered with a plastic insulation
(thickness = 2.5 mm) is exposed to air at 300 K and ho = 20 W/m2K. The insulation has a
k of 0.4 W/m.K. It is assumed that the wire’s surface temperature is constant at 400 K and
is not affected by the covering.
a. Calculate the value of the critical radius.
b. Calculate the heat loss per m of wire length with no insulation.
c. Repeat (b) for the presence of insulation.
Solution: For part (a), using Eq. (13.2-26),

For part (b), L = 1.0 m, r2 = 1.5/(2 × 1000) = 0.75 × 10–3 m, A = 2πr2L

Substituting into Eq. (13.2-23),

=(30 W/m2K)(2)(1.0 m)(400-300)K=9.42 W


For part (c) with insulation, r1 = 1.5/(2 × 1000) = 0.75x10-3 m, r2 = (2.5 + 1.5/2)/1000 = 3.25x10–3 m.
Substituting into Eq. (13.2-24),

Hence, adding insulation greatly increases the heat loss.


Problem : A metal steam pipe having an outside diameter of 30 mm has a surface temperature of 400 K
and is to be insulated with an insulation having a thickness of 20 mm and a k of 0.08 W/m.K. The pipe is
exposed to air at 300 K and a convection film coefficient of 30 W/m2K.
a) Calculate the critical radius and the heat loss per m of length for the bare pipe.
b) Calculate the heat loss for the insulated pipe assuming that the surface temperature of the pipe remains
constant.
Example :
Derive a relation for the critical radius of insulation for a sphere.
r1 : Internal radius of insulation (sphere); r2 : outside radius of insulation (sphere); ho : Film heat transfer

coefficient; To : Bulk fluid temperature surrounding insulation; k : Thermal conductivity of insulation

Solution :
The Critical Radius of an Insulation for a Sphere
HEAT EXCHANGERS
The design of heat exchanger apparatus is discussed in Chapter 27. However, a brief look
at the design of one of the simpler pieces of equipment, the double-pipe heat exchanger, is
appropriate at this point.
The determination of the required heat transfer area is of the principal objectives in the
design of heat exchangers. The basic equationfor determining area is Eq. (21-3)

It is, however, more common to write Eq. (21-17) for a differential segment of the

= (21-17)

exchanger for which the outside tube area is dAo and integrate as
=o (21-23)

This equation contains the overall coefficient of heat transfer Uo, based on the outside

tube area Ao, and the difference the mixing-cup temperatures of the hot and cold fluids, .

The reason for not us Eq. (21-17) directly is that both and Uo may vary from one of the

exchanger to the other. Eq. (21-18) which defined Ui in terms of Ai, can also be expressed in
differential form and integrated in the manner shown above. We shall limit our subsequent
analysis to equations based on the outside area; however, it should be remembered that a
similar treatment exists in terms of Ai and Ui.

Under steady-state conditions the mixing-cup temperatures of the hot and cold fluidsin
a heat exchanger are assumed to be fixed at any cross-sections normal to the flow. If we
designate these temperatures as Th and Tc. For convenience, this quantity will be written as
=Th-Tc (21-24)

The double-pipe heat exchanger consists of two concentric tubes through which the hot and
cold fluids flow. The fluids may flow in the same direction, i.e., concurrent, or they may
flow in opposite , i.e., countercurrent. The double-pipe heat exchanger can be made from a
concentric pair of single lengths of pipe or from a number of lengths arranged in a vertical
row with the sections connected at alternate ends, as shown in Fig. 21-6.

Fig. 21-6 Double-pipe heat exchanger


The exchanger shown in Fig. 21-6 can be regarded in mathematically as long run of two
concentric pipes. The analysis below will be based on this simplification. Two fluids with
constant but different inlet temperatures enter opposite ends of the exchanger shown in
Fig. 21-7.

T2
In this system, the hot fluid enters the inner pipe at a constant rate of kg/h and the cold

fluid enters the annulus at a constant rate of kg/h. The choice of path for each fluid
depends on considerations such as corrosion, fluid pressure, and permissible pressure
drop. The temperatures of the fluids as a function of length are shown in Fig. 21-7. The
fluid temperatures usually do not vary linearly with distance from the inlet.

A heat balance can be written for the differential section of the exchanger shown in Fig.
21-7.

ddAo= dAi (21-25)

Equation (21-25) shows that if the specific heats are constant, the individual mixing-cup
temperatures are linear with respect to q, as illustrated in Fig. 21-7. The difference
between the mixing-cup temperatures, , is also linear with respect to q.
Consequently, we can express the derivative of with respect to q in terms of the overall
change in and the total heat transfer rate q in the exchanger. The derivative can be written
as
= (21-26)
From Eq.(21-25) we obtain an expression for dq in terms of the overall coefficient and
substitute it in Eq. (21-26) to give
= (21-27)
This equation is rearranged to give
= (21-28)
(21-29)
The quantity in brackets is the logarithmic mean temperature difference, lm, where
lm = (21-29a)
= (21-29b)

The same equation would have been obtained if the flow of the two streams in Fig.
21-7 had been concurrent or parallel. If a are equal, that is, = Eq.(21-25) shows that
any change in temperature of the hot fluid along the exchanger is accompanied by equal
change in the temperature of the cold fluid. In such a case, with countercurrent flow, is
the same at all cross sections in the exchanger, so that and are equal, making the right-
hand side of Eq. (21-29) indeterminate. This presents no problem, however, because if
and are both constant, Eq. (21-25) can be integrated directly and gives for whole
exchanger.

= (21-17)
Example : Double Pipe Heat Exchanger
a) Oil flowing at the rate of 7 258 kg/h with a Cpm = 2.01 kJ/kg · K is cooled from 394.3 K
to 338.9 K in a countercurrent flow heat exchanger by water entering at 294.3 K and
leaving at 305.4 K. Calculate the flow rate of the water and the overall heat transfer
coefficient Ui if the Ai is 5.11 m2.
b) Calculate the flow rate of the water and the overall heat transfer coefficient Ui if the Ai
is 5.11 m2 for concurrent flow
Solution :
If the overall coefficient varies within the exchanger, Eq. (21-29) is not rigorous,
although it is often used as an approximation. In some circumstances the convective
coefficient of one of the fluids can be expressed as a linear function of the mixing-cup
temperature of the fluid. If this fluid offers the major resistance to heat transfer, the overall
coefficient is approximately equal to the convective coefficient (as was shown earlier in this
chapter), so that Uo is a linear function of the mixing-cup temperature of that fluid. Since
the temperature of each fluid and the temperature difference are linear function of q.
Under these circumstances the overall coefficient is a linear function of expression,

in Eq.(21-28), we obtain

You might also like