21.09.2015 Views

Course No Course Title Credits

Heat & Mass Transfer - acharya ng ranga agricultural university

Heat & Mass Transfer - acharya ng ranga agricultural university

SHOW MORE
SHOW LESS
  • No tags were found...

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

For Class use only<br />

ACHARYA N.G. RANGA AGRICULTURAL UNIVERSITY<br />

<strong>Course</strong> <strong>No</strong>. FDEN- 221<br />

<strong>Course</strong> HEAT AND MASS TRANSFER<br />

<strong>Title</strong>:<br />

<strong>Credits</strong>: 2 (1 + 1)<br />

Prepared by<br />

Er. B. SREENIVASULA REDDY<br />

Assistant Professor (Food Engineering)<br />

College of Food Science and Technology<br />

Chinnarangapuram, Pulivendula – 516390<br />

YSR (KADAPA) District, Andhra Pradesh


DEPARTMENT OF FOOD ENGINEERING<br />

1 <strong>Course</strong> <strong>No</strong> : FDEN - 221<br />

2 <strong>Title</strong> : Heat and Mass Transfer<br />

3 Credit hours : 2 (1+1)<br />

4 General Objective : To impart knowledge to students on different<br />

modes of heat transfer through extended surfaces,<br />

study of heat exchanges and principles of mass<br />

transfer<br />

5 Specific Objectives :<br />

a) Theory : By the end of the course, the students will acquire<br />

knowledge from different modes of heat transfer,<br />

extended surfaces, boiling and condensation<br />

process and principles of heat exchangers which<br />

are very essential in dairy and food industries<br />

b) Practical By the end of the course, the students will learn<br />

efficient design of heat exchangers on the basis of<br />

overall heat transfer coefficient and LMTD<br />

A) Theory Lecture Outlines<br />

1 Introduction to Heat Transfer- Basic Mechanisms of Heat Transfer -<br />

Conduction Heat transfer - Fourier's Law of Heat Conduction- Convection<br />

Heat Transfer- Radiation Heat Transfer<br />

2 The basic equation that governs the transfer of heat in a solid - Thermal<br />

conductivity<br />

3 One-dimensional steady-state conduction of heat through some simple<br />

geometries - Conduction Through a Flat Slab or Wall - Conduction Through<br />

a Hollow Cylinder - Conduction Through a Hollow Sphere<br />

4 Conduction Heat Transfer Through A Composite Plane Wall - Conduction<br />

Heat Flow Through A Composite Cylinder<br />

5 The Overall Heat-Transfer Coefficient - Critical Thickness of Insulation<br />

6 Heat Source Systems: One-dimensional steady state heat conduction with<br />

heat generation : Heat Flow through slab / Plane Wall<br />

7 Steady state heat conduction with heat dissipation to environment-<br />

Introduction to extended surfaces (FINS) of uniform area of cross section -<br />

Different fin configurations - General Conduction Analysis Equation<br />

8 Equation of temperature distribution with different boundary conditions, Fin<br />

Performance and Overall surface efficiency of FINS<br />

9 Principles Of Unsteady-State Heat Transfer: Derivation of Basic Equation;


Simplified Case For Systems With Negligible Internal Resistance ; Total<br />

Amount of Heat Transferred ; Dimensional Analysis in Momentum Transfer<br />

10 Some important empirical relations used for determination of heat transfer<br />

coefficient: Nusselt’s number, Prandtl number, Reynold’s number, Grashoff<br />

number<br />

11 Radiation - heat transfer, Radiation Properties, radiation through black and<br />

grey surfaces, determination of shape factors<br />

12 Introduction to condensing and boiling heat transfer, Condensation Heat-<br />

Transfer Phenomena, Film Condensation Inside Horizontal Tubes , Boiling<br />

Heat Transfer<br />

13 Heat exchangers- general introduction; Double-pipe heat exchanger; Shelland-tube<br />

heat exchanger; Cross-flow exchanger; fouling factors, LMTD<br />

14 Design problems on heat exchangers: Calculation of heat exchanger size<br />

from known temperatures, Problem on Shell-and-tube heat exchanger ,<br />

Design of shell-and-tube heat exchanger<br />

15 Introduction To Mass Transfer: A Similarity of Mass, Heat, and Momentum<br />

Transfer Processes; Fick's Law for Molecular Diffusion<br />

16 Molecular Diffusion In Gases : Equimolar Counter diffusion in Gases<br />

B) Practical Class Outlines<br />

1 Determination of thermal conductivity of milk and dairy products<br />

2 Tutorials on heat conduction through slab, cylinder and sphere<br />

3 Tutorials on heat conduction through slab, cylinder and sphere<br />

4 Tutorials on heat conduction through slab, cylinder and sphere<br />

5 Tutorials on extended surfaces (FINS)<br />

6 Tutorials on unsteady state heat conduction<br />

7 Determination of specific heat of food materials<br />

8 Tutorials on determination of Nusselt’s number by dimensional analysis<br />

9 Tutorials on LMTD and NTU method of analysis of heat exchangers<br />

10 Study of shell and tube heat exchanger<br />

11 Study of plate heat exchanger<br />

12 Study on temperature distribution and heat transfer in HTST pasteurizer<br />

13 Study on temperature distribution and heat transfer in HTST pasteurizer<br />

14 Design problems on heat exchangers - I<br />

15 Design problems on heat exchangers - II<br />

16 Design problems on heat exchangers - III<br />

References<br />

1 Geankoplis, C.J. 1978. Transport Processes and Unit Operations. Allyn and<br />

Bacon Inc., Newton, Massachusetts.<br />

2 Holman, J. P. 1989. Heat Transfer. McGraw Hill Book Co., New Delhi.<br />

3 Incropera, F. P. and De Witt, D .P. 1980. Fundamentals of Heat and Mass


Transfer. John Wiley and Sons, New York.<br />

4 Gupta, C. P. and Prakash, R. 1994. Engineering Heat Transfer. Nem Chand<br />

and Bros., Roorkee


LECTURE NO.1<br />

INTRODUCTION TO HEAT TRANSFER- BASIC MECHANISMS OF HEAT<br />

TRANSFER - CONDUCTION HEAT TRANSFER - FOURIER'S LAW OF<br />

HEAT CONDUCTION- CONVECTION HEAT TRANSFER- RADIATION<br />

HEAT TRANSFER<br />

Introduction<br />

Heat transfer is the science that seeks to predict the energy transfer<br />

that may take place between material bodies as a result of a temperature<br />

difference. Thermodynamics deals with systems in equilibrium; it may be used<br />

to predict the amount of energy required to change a system from one<br />

equilibrium state to another; it may not be used to predict how fast a change<br />

will take place since the system is not in equilibrium during the process. Heat<br />

transfer supplements the first and second principles of thermodynamics by<br />

providing additional experimental rules which may be used to establish<br />

energy-transfer rates.<br />

As an example of the different kinds of problems that are treated by<br />

thermodynamics and heat transfer, consider the cooling of a hot steel bar that<br />

is placed in a pail of water. Thermodynamics may be used to predict the final<br />

equilibrium temperature of the steel bar-water combination. Thermodynamics<br />

will not tell us how long it takes to reach this equilibrium condition or what the<br />

temperature of the bar will be after a certain length of time before the<br />

equilibrium condition is attained. Heat transfer may be used to predict the<br />

temperature of both the bar and the water as a function of time.<br />

Basic Mechanisms of Heat Transfer<br />

Heat transfer may occur by anyone or more of the three basic<br />

mechanisms of heat transfer: conduction, convection, and radiation.<br />

1. Conduction Heat transfer. In conduction, heat can be conducted through<br />

solids, liquids, and gases. The heat is conducted by the transfer of the energy<br />

of motion between adjacent molecules. In a gas the "hotter" molecules, which<br />

have greater energy and motions, impart energy to the adjacent molecules at<br />

lower energy levels. This type of transfer is present to some extent in all<br />

solids, gases, or liquids in which a temperature gradient exists. In conduction,


energy can also be transferred by "free" electrons, which is quite important in<br />

metallic solids. Examples of heat transfer mainly by conduction are heat<br />

transfer through walls of exchangers or a refrigerator, heat treatment of steel<br />

forgings, freezing of the ground during the winter, and so on.<br />

Fourier's Law of Heat Conduction<br />

The basic rate transfer process equation for processes such as<br />

momentum transfer, heat transfer, mass transfer and electric current is as<br />

follows:<br />

driving force<br />

rate of a transfer process = ----------- (1)<br />

resis tan ce<br />

This equation states what we know intuitively: that in order to transfer a<br />

property such as heat or mass, we need a driving force to overcome a<br />

resistance.<br />

The transfer of heat by conduction also follows this basic equation and<br />

is written as Fourier's law for heat conduction in fluids or solids:<br />

where<br />

q x<br />

A<br />

dT = − k<br />

--------------- (2)<br />

dx<br />

qx<br />

is the heat-transfer rate in the x direction in watts (W), A is<br />

the cross-sectional area normal to the direction of flow of heat in m 2 , T is<br />

temperature in K, x is distance in m, and k is the thermal conductivity in<br />

W/m·K in the SI system. The quantity<br />

The quantity<br />

dT / dx<br />

q x<br />

/ A is called the heat flux in W/m 2 .<br />

is the temperature gradient in the x direction. The minus<br />

sign in eq. (2) is required because if the heat flow is positive in a given<br />

direction, the temperature decreases in this direction.<br />

Fourier's law, eq.(2), can be integrated for the case of steady-state<br />

heat transfer through a flat wall of constant cross-sectional area A, where the<br />

inside temperature is T 1 at point 1 and T 2 at point 2, a distance of x2 −x<br />

1 m<br />

away. Rearranging eq. (2),<br />

qx<br />

A<br />

x2<br />

∫<br />

x1<br />

dx<br />

= −<br />

k<br />

∫<br />

T2<br />

T1<br />

dT<br />

-------(3)<br />

Integrating, assuming that k is constant and does not vary with<br />

temperature and dropping the subscript x on<br />

qx<br />

for convenience,


q<br />

A<br />

k<br />

= ( T ) 1<br />

−T<br />

2 -------- (4)<br />

x −x<br />

2<br />

1<br />

2. Convection Heat Transfer. The transfer of heat by convection implies the<br />

transfer of heat by bulk transport and mixing of macroscopic elements of<br />

warmer portions with cooler portions of a gas or liquid. It also often refers to<br />

the energy exchange between a solid surface and a fluid. A distinction must<br />

be made between forced-convection heat transfer, where a fluid is forced<br />

to flow past a solid surface by a pump, fan, or other mechanical means, and<br />

natural or free convection, where warmer or cooler fluid next to the solid<br />

surface causes a circulation because of a density difference resulting from<br />

the temperature differences in the fluid. Examples of heat transfer by<br />

convection are loss of heat from a car radiator where the air is being<br />

circulated by a fan, cooking of foods in a vessel being stirred, cooling of a hot<br />

cup of coffee by blowing over the surface, and so on.<br />

Convective Heat-Transfer Coefficient<br />

It is well known that a hot piece of material will cool faster when air is<br />

blown or forced past the object. When the fluid outside the solid surface is in<br />

forced or natural convective motion, we express the rate of heat transfer from<br />

the solid to the fluid or vice versa, by the following equation:<br />

q= hA ( T w<br />

−T<br />

f<br />

) -------------- (5)<br />

where q is the heat-transfer rate in W, A is the area in m 2 , T w is the<br />

temperature of the solid surface in K, T f is the average or bulk temperature of<br />

the fluid flowing past in K, and h is the convective heat-transfer coefficient in<br />

W/m 2 .K.<br />

The coefficient h is a function of the system geometry, fluid properties,<br />

flow velocity, and temperature difference. In many cases, empirical<br />

correlations are available to predict this coefficient, since it often cannot be<br />

predicted theoretically. Since we know that when a fluid flows past a surface<br />

there is a thin, almost stationary layer or film of fluid adjacent to the wall which<br />

presents most of the resistance to heat transfer, we often call the coefficient h<br />

a film coefficient.<br />

3. Radiation Heat Transfer:- Radiation differs from heat transfer by<br />

conduction and convection in that no physical medium is needed for its


propagation. Radiation is the transfer of energy through space by means of<br />

electromagnetic waves in much the same way as electromagnetic light<br />

waves transfer light. The same laws that govern the transfer of light govern<br />

the radiant transfer of heat. Solids and liquids tend to absorb the radiation<br />

being transferred through them, so that radiation is important primarily in<br />

transfer through space or gases. The most important example of radiation is<br />

the transport of heat to the earth from the sun. Other examples are cooking of<br />

food when passed below red-hot electric heaters, heating of fluids in coils of<br />

tubing inside a combustion furnace, and so on.<br />

The rate of energy emitted by a black body is proportional to the fourth<br />

power of the absolute temperature of the body and directly proportional to its<br />

surface area. Thus<br />

q emitted<br />

= σAT<br />

4<br />

where σ is the proportionality constant and is called the Stefan-<br />

Boltzmann constant with the value of 5.669 x 10 -8 W/m 2 .K 4 . The equation is<br />

called the Stefan-Boltzmann law of thermal radiation, and it applies only to<br />

blackbodies.<br />

Problem Heat Loss Through an Insulating Wall<br />

Calculate the heat loss per m 2<br />

of surface area for an insulating wall<br />

composed of 25.4 -mm-thick fiber insulating board, where the inside<br />

temperature is 352.7 K and the outside temperature is 297.1 K. The thermal<br />

conductivity of fiber insulating board is 0.048 W/m.K<br />

Solution:<br />

The thickness x2 − x1<br />

= 0.0254 m.<br />

Substituting into the eq.<br />

q<br />

A<br />

= k<br />

0.048<br />

( T ) 1<br />

−<br />

2<br />

=<br />

x −x<br />

T 0.0254<br />

2<br />

1<br />

= 105.1 W/m 2<br />

(352 .7−297<br />

.1)


LECTURE NO.2<br />

THE BASIC EQUATION THAT GOVERNS THE TRANSFER OF HEAT IN A<br />

SOLID - THERMAL CONDUCTIVITY<br />

The basic equation that governs the transfer of heat in a solid<br />

Fig.2.1 Elemental volume for one-dimensional heat conduction analysis<br />

Consider (Fig. 2.1) the general case where the temperature may be<br />

changing with time and heat sources may be present within the body. For the<br />

element of thickness dx the following energy balance may be made:<br />

Energy conducted in left face + heat generated within element<br />

= change in internal energy + energy conducted out right face<br />

These energy quantities are given as follows:<br />

Energy in left face =<br />

q x<br />

∂T<br />

=−kA<br />

∂x<br />

Energy generated within element =<br />

q .<br />

Adx<br />

Change in internal energy =<br />

ρcA<br />

∂T<br />

dx ∂τ<br />

Energy out right face =<br />

q<br />

x+<br />

dx<br />

=− kA<br />

∂T<br />

∂x<br />

x+<br />

dx<br />

⎡ ∂T<br />

∂ ⎛ ∂T<br />

⎞ ⎤<br />

= − A⎢k<br />

+ ⎜k<br />

⎟dx<br />

⎥<br />

⎣ ∂x<br />

∂x<br />

⎝ ∂x<br />

⎠ ⎦


(In the derivations, the expression for the derivative at<br />

x + dx<br />

has been<br />

written in the form of a Taylor-Series expression with only the first two terms<br />

of the series employed for the development.)<br />

where<br />

.<br />

q = energy generated per unit volume, W /m 3<br />

c = specific heat of material, J /kg .°C<br />

ρ = density, kg/m 3<br />

Combining the relations above gives<br />

or<br />

heat source.<br />

∂<br />

− kA T<br />

+<br />

∂x<br />

q .<br />

Adx =<br />

∂T<br />

ρcA<br />

dx ∂τ<br />

⎡ ∂T<br />

− A⎢k<br />

⎣ ∂x<br />

.<br />

∂ ⎛ ∂T<br />

⎞ ∂T<br />

⎜k<br />

⎟dx<br />

+ q=ρc<br />

∂x<br />

⎝ ∂x<br />

⎠ ∂ τ<br />

∂ ⎛ ∂T<br />

⎞ ⎤<br />

+ ⎜k<br />

⎟dx<br />

∂x<br />

⎥<br />

⎝ ∂x<br />

⎠ ⎦<br />

This is the one-dimensional steady state heat-conduction equation with<br />

Figure 2.2 Elemental volume for three-dimensional heat-conduction analysis:<br />

(a) cartesian coordinates; (b) cylindrical coordinates;<br />

(c) spherical coordinates.


The general three-dimensional heat-conduction equation is<br />

∂ ⎛ ∂T<br />

⎜k<br />

∂x<br />

⎝ ∂x<br />

⎞<br />

⎟+<br />

⎠<br />

∂ ⎛ ∂T<br />

⎜k<br />

∂y<br />

⎝ ∂y<br />

⎞<br />

.<br />

∂ ⎛ ∂T<br />

⎞ ∂T<br />

⎟+ ⎜k<br />

⎟+ q=ρc<br />

⎠ ∂z<br />

⎝ ∂z<br />

⎠ ∂ τ<br />

For constant thermal conductivity equation can be written as<br />

where the quantity<br />

2 2 2<br />

∂ T ∂ T ∂ T q 1 ∂T<br />

+ + + =<br />

2 2 2<br />

∂x<br />

∂x<br />

∂x<br />

k α ∂τ<br />

.<br />

α = k / ρc<br />

is called the thermal diffusivity of the material.<br />

The larger the value of α , the faster heat will diffuse through the material. A<br />

high value of α could result either from a high value of thermal conductivity,<br />

which would indicate a rapid energy-transfer rate, or from a low value of the<br />

thermal heat capacity ρ c . A low value of the heat capacity would mean that<br />

less of the energy moving through the material would be absorbed and used<br />

to raise the temperature of the material; thus more energy would be available<br />

for further transfer. α has the units of m 2 /s.<br />

The three-dimensional heat-conduction equation with heat generation<br />

for cylindrical or spherical co-ordinates is<br />

Cylindrical coordinates:<br />

Spherical coordinates:<br />

2<br />

2 2<br />

∂ T 1 ∂T<br />

1 ∂ T ∂ T q 1 ∂T<br />

+ + + + =<br />

2<br />

2 2 2<br />

∂r<br />

r ∂r<br />

r ∂φ<br />

∂z<br />

k α ∂τ<br />

.<br />

1 ∂<br />

r ∂r<br />

2<br />

2<br />

( rT )<br />

+<br />

r<br />

2<br />

1 ∂ ⎛ ∂T<br />

⎞<br />

⎜sin<br />

θ ⎟+<br />

sin θ ∂θ<br />

⎝ ∂θ<br />

⎠ r<br />

2<br />

1<br />

sin<br />

2<br />

2<br />

∂ T q 1 ∂T<br />

+ =<br />

2<br />

θ ∂φ<br />

k α ∂τ<br />

The reduced form of the general equations for several cases of practical<br />

interest.<br />

Steady-state one-dimensional heat flow (no heat generation):<br />

2<br />

d T<br />

2<br />

dx<br />

= 0<br />

Steady-state one-dimensional heat flow in cylindrical coordinates (<strong>No</strong><br />

heat generation):<br />

.<br />

2<br />

d T<br />

2<br />

dr<br />

1 dT<br />

+<br />

r dr<br />

= 0


Steady-state one-dimensional heat flow with heat sources:<br />

2<br />

d T<br />

2<br />

dx<br />

.<br />

q<br />

+ = 0 k<br />

Thermal conductivity<br />

Thermal conductivity, k, is the property of a material that indicates its<br />

ability to conduct heat. Thermal conductivity is measured in W/m·K. The<br />

thermal conductivity predicts the rate of energy loss (in watts, W) through a<br />

piece of material.<br />

Thermal conductivity, k, also defined as the quantity of heat Q that flows<br />

per unit time through a food of unit thickness and unit area having unit<br />

temperature difference between faces.<br />

The reciprocal of thermal conductivity is thermal resistivity. In general,<br />

the thermal conductivity is strongly temperature-dependent.<br />

Thermal energy may be conducted in solids by two modes: lattice<br />

vibration and transport by free electrons. In good electrical conductors a<br />

rather large number of free electrons move about in the lattice structure of the<br />

material. Just as these electrons may transport electric charge, they may also<br />

carry thermal energy from a high-temperature region to a low-temperature<br />

region, as in the case of gases. In fact, these electrons are frequently referred<br />

to as the electron gas. Energy may also be transmitted as vibrational energy<br />

in the lattice structure of the material. In general, however, this latter mode of<br />

energy transfer is not as large as the electron transport, and for this reason<br />

good electrical conductors are almost always good heat conductors, viz.,<br />

copper, aluminum, and silver, and electrical insulators are usually good heat<br />

insulators. A notable exception is diamond, which is an electrical insulator, but<br />

which can have a thermal conductivity five times as high as silver or copper. It<br />

is this fact that enables a jeweler to distinguish between genuine diamonds<br />

and fake stones. A small instrument is available that measures the response<br />

of the stones to a thermal heat pulse. A true diamond will exhibit a far more<br />

rapid response than the non genuine stone.


LECTURE NO.3<br />

ONE-DIMENSIONAL STEADY-STATE CONDUCTION OF HEAT THROUGH<br />

SOME SIMPLE GEOMETRIES - CONDUCTION THROUGH A FLAT SLAB<br />

OR WALL - CONDUCTION THROUGH A HOLLOW CYLINDER -<br />

CONDUCTION THROUGH A HOLLOW SPHERE<br />

CONDUCTION HEAT TRANSFER<br />

One-dimensional steady-state conduction of heat through some simple<br />

geometries<br />

Fig. 3.1 Heat conduction in a flat wall: (a) geometry of wall, (b) temperature<br />

plot.<br />

Conduction Through a Flat Slab or Wall<br />

Consider a flat slab or wall (Fig. 3.1) where the cross-sectional area A<br />

and k in are constant,<br />

q k<br />

The eq. = ( T1<br />

−T<br />

2)<br />

can be rewrite as<br />

A x −x<br />

2<br />

1<br />

q<br />

A<br />

=<br />

k<br />

∆x<br />

where ∆ x=<br />

x 2<br />

−x1<br />

.<br />

( T 1<br />

−T<br />

2)<br />

The above indicates that if T is substituted for T 2 and x for x<br />

2 , the<br />

temperature varies linearly with distance, as shown in Fig.3.1(b).<br />

If the thermal conductivity is not constant but varies linearly with<br />

temperature, then substituting<br />

integrating,<br />

k = a=<br />

bT into the above equation and


q<br />

A<br />

T 2 .<br />

where<br />

T1<br />

+ T<br />

q 2<br />

a + b<br />

2<br />

k<br />

=<br />

( T1<br />

−T2<br />

) =<br />

m ( T1<br />

T2<br />

)<br />

A ∆x<br />

∆x<br />

−<br />

b T 1<br />

+ T<br />

k a<br />

2<br />

m<br />

= +<br />

2<br />

This means that the mean value of k (i.e., k m ) to use in<br />

k<br />

= ( T ) 1<br />

−T<br />

2 is the value of k evaluated at the linear average of T 1 and<br />

∆x<br />

The rate of a transfer process equals the driving force over the<br />

q k<br />

resistance and the equation = ( T 1<br />

−T<br />

2)<br />

can be rewritten in that form<br />

A ∆x<br />

as:<br />

where<br />

R ∆x<br />

/ kA<br />

T −T2<br />

T1<br />

−T2<br />

q = =<br />

∆x<br />

/ kA R<br />

1 =<br />

driving force<br />

resis tan ce<br />

= A and is the resistance in K/W.<br />

Fig. 3.2 Heat conduction in a cylinder<br />

Conduction Through a Hollow Cylinder<br />

In many instances in the process industries, heat is being transferred<br />

through the walls of a thick-walled cylinder, such as a pipe that may or may<br />

not be insulated. Consider the hollow cylinder in Fig.3.2 with an inside radius<br />

of r 1 , where the temperature is T 1 , an outside radius of r 2 having a temperature<br />

of T 2 , and a length of L m. Heat is flowing radially from the inside surface to<br />

the outside. Rewriting Fourier's law, with distance dr instead of dx,<br />

q x<br />

A<br />

= −<br />

k<br />

dT<br />

dr


The cross-sectional area normal to the heat flow is<br />

A =2πrL<br />

Substituting A value and rearranging, and integrating,<br />

q<br />

2πL<br />

r2<br />

∫<br />

r1<br />

dr<br />

r<br />

= −<br />

k<br />

∫<br />

T2<br />

T1<br />

dT<br />

q=<br />

k<br />

2πL<br />

ln( r<br />

2<br />

/ r1<br />

)<br />

( T −T<br />

)<br />

1<br />

2<br />

Conduction Through a Hollow Sphere<br />

Heat conduction through a hollow sphere is another case of onedimensional<br />

conduction. Using Fourier's law for constant thermal conductivity<br />

with distance dr, where r is the radius of the sphere,<br />

the radius.<br />

q x<br />

A<br />

= −<br />

k<br />

dT<br />

dr<br />

The cross-sectional area normal to the heat flow is<br />

2<br />

A = 2πr<br />

Substituting A value and rearranging, and integrating,<br />

q<br />

4π<br />

r2<br />

∫<br />

r1<br />

dr<br />

2<br />

r<br />

= −<br />

k<br />

∫<br />

T2<br />

T1<br />

( T −T<br />

)<br />

4πk<br />

1 2<br />

q =<br />

1/ r −1/<br />

r<br />

1<br />

2<br />

=<br />

dT<br />

T −T<br />

( 1/ r −1/<br />

r ) / 4πk<br />

It can easily be shown that the temperature varies hyperbolically with<br />

1<br />

1<br />

2<br />

2


LECTURE NO.4<br />

CONDUCTION HEAT TRANSFER THROUGH A COMPOSITE PLANE WALL<br />

- CONDUCTION HEAT FLOW THROUGH A COMPOSITE CYLINDER<br />

CONDUCTION HEAT TRANSFER THROUGH A COMPOSITE PLANE<br />

WALL<br />

Fig. 4.1 Heat flow through multilayer wall<br />

Consider the heat flow through composite wall made of several<br />

materials of different thermal conductivities and thicknesses. An example is a<br />

wall of a cold storage, constructed of different layers of materials of different<br />

insulating properties. All materials are arranged in series in the direction of<br />

heat transfer, as shown in the above Figure.<br />

The thickness of the walls are x 1, x 2 , and x 3 and the thermal<br />

conductivites of the walls are K 1 , K 2 , and K 3 , respectively. The temperatures at<br />

the contact surfaces are T 2 , T 3 , and T 4.<br />

From Fourier’s Law,<br />

q<br />

=−K A<br />

A<br />

This may be written as<br />

dT<br />

dx<br />

∆T<br />

= T<br />

x1<br />

T<br />

1<br />

−T<br />

2<br />

= . q<br />

K A<br />

x2<br />

T<br />

2<br />

−T<br />

3<br />

= . q<br />

K A<br />

1<br />

2<br />

2<br />

−T<br />

x3<br />

T<br />

4<br />

−T<br />

3<br />

= . q<br />

K A<br />

3<br />

1<br />

q.<br />

∆x<br />

=−<br />

KA


Total temperature difference, ∆T<br />

=∆T<br />

1<br />

+∆T<br />

2<br />

+ ∆T<br />

3<br />

⎡ x1<br />

x2<br />

x3<br />

T1 −T4<br />

= q.<br />

⎢ + +<br />

⎣K1<br />

A K2<br />

A K31<br />

where, T1 − T4<br />

= thermal potential responsible for heat flow. The<br />

⎤<br />

⎥<br />

A⎦<br />

⎡ x1<br />

x2<br />

x<br />

⎢ + +<br />

⎣K1<br />

A K2<br />

A K31<br />

3<br />

⎤<br />

⎥<br />

A⎦<br />

is known as the total thermal resistance of the<br />

composite ass. It is similar to the electrical resistance in series.<br />

The thermal circuit for multilayer rectangular system is shown in the<br />

following figure.<br />

Fig 4.2 Electrical analog of one dimensional heat transfer through composite<br />

wall<br />

T1<br />

−T4<br />

q =<br />

R + R + R<br />

CONDUCTION HEAT FLOW THROUGH A COMPOSITE CYLINDER<br />

1<br />

2<br />

3<br />

Fig.4.3 One-dimensional heat flow through multiple cylindrical sections and<br />

electrical analog<br />

Consider a long cylinder of inside radius r i , outside radius r o , and length<br />

L, such as the one shown in above Figure 4.3. We expose this cylinder to a


temperature differential T i - T o and determine what the heat flow will be. For a<br />

cylinder with length very large compared to diameter, it may be assumed that<br />

the heat flows only in a radial direction, so that the only space coordinate<br />

needed to specify the system is r. Again, Fourier's law is used by inserting the<br />

proper area relation. The area for heat flow in the cylindrical system is<br />

A =2πrL<br />

so that Fourier's law is written<br />

with the boundary conditions<br />

q<br />

r<br />

dT<br />

= −k<br />

Ar<br />

or<br />

dr<br />

q r<br />

=−2πk rL<br />

T = T i at r = r i ,<br />

dT<br />

dr<br />

T = T o at r = r o<br />

The solution to equation<br />

q<br />

r<br />

dT<br />

= −k<br />

Ar<br />

is<br />

dr<br />

q=<br />

k<br />

2πL<br />

ln( r<br />

2<br />

/ r1<br />

)<br />

( T −T<br />

)<br />

1<br />

2<br />

and the thermal resistance in this case is<br />

R<br />

th<br />

ln( ro<br />

/ ri<br />

)<br />

=<br />

2πkL<br />

The thermal-resistance concept may be used for multiple-layer<br />

cylindrical walls just as it was used for plane walls. For the three-layer system<br />

shown in Figure the solution is<br />

2π<br />

L ( T1<br />

−T4<br />

)<br />

q =<br />

ln( r2<br />

/ r1<br />

) ln( r3<br />

/ r2<br />

) ln( r4<br />

/ r2<br />

)<br />

+ +<br />

K K K<br />

a<br />

The thermal circuit is also shown in Figure.<br />

B<br />

B


LECTURE NO.5<br />

THE OVERALL HEAT-TRANSFER COEFFICIENT - CRITICAL THICKNESS<br />

OF INSULATION<br />

5.1 THE OVERALL HEAT-TRANSFER COEFFICIENT<br />

Fig. 5.1 Heat flow with convective boundaries in plane wall<br />

Fig. 5.2 Heat flow with convective boundaries in plane wall- Electrical<br />

Analog<br />

In many practical situations the surface temperatures (or boundary<br />

conditions at the surface) are not known, but there is a fluid on both sides of<br />

the solid surfaces. Consider the plane wall in the Figure 5.1, with a hot fluid at<br />

temperature T 1 on the inside surface and a cold fluid at T 4 on the outside<br />

surface. The convective coefficient on the outside is h o W/m 2 .K and h i on the<br />

inside.<br />

The heat transfer is expressed by<br />

q=<br />

h A<br />

i<br />

K<br />

A<br />

A<br />

( T −T<br />

) = ( T −T<br />

) = h A T − )<br />

1 2<br />

2 3 o<br />

(<br />

3<br />

T4<br />

∆x<br />

A<br />

The heat-transfer process may be represented by the resistance<br />

network in electrical analog Figure, and the overall heat transfer is calculated


as the ratio of the overall temperature difference to the sum of the thermal<br />

resistances:<br />

T1<br />

−T4<br />

T1<br />

−T<br />

q=<br />

=<br />

1 ∆x<br />

A<br />

1<br />

+ +<br />

∑R<br />

h A K A h A<br />

∴<br />

4<br />

i<br />

A<br />

The overall heat transfer by combined conduction and convection is<br />

frequently expressed in terms of an overall heat-transfer coefficient U, defined<br />

by the relation<br />

q=<br />

U A∆<br />

T overall<br />

o<br />

where ∆ T overall<br />

= T 1<br />

−T<br />

4 and U is<br />

U<br />

1<br />

∆x<br />

1<br />

A<br />

= + + W/m<br />

hi<br />

K<br />

A<br />

h<br />

2 .K<br />

o<br />

value as<br />

The overall heat-transfer coefficient is also related to the R<br />

U =<br />

1<br />

R value<br />

A more important application is heat transfer from a fluid outside a<br />

cylinder, through a metal wall, to a fluid inside the tube, as often occurs in<br />

heat exchangers.<br />

boundaries.<br />

Figure 5.3 Resistance analogy for hollow cylinder with convection


<strong>No</strong>te that the area for convection is not the same for both fluids in this<br />

case, these areas depend on the inside tube diameter and wall thickness. In<br />

this case the overall heat transfer would be expressed by<br />

q =<br />

1<br />

h A<br />

i<br />

i<br />

T1<br />

−T<br />

ln<br />

o<br />

+<br />

2πKL<br />

4<br />

( r / r )<br />

i<br />

1<br />

+<br />

h A<br />

o<br />

o<br />

The terms A i and A o represent the inside and outside surface areas of<br />

the inner tube. The overall heat-transfer coefficient may be based on either<br />

the inside or the outside area of the tube. Accordingly,<br />

Ui<br />

=<br />

1<br />

h<br />

i<br />

1<br />

Ai<br />

ln<br />

o<br />

+<br />

2πKL<br />

( r / r )<br />

i<br />

Ai<br />

+<br />

A<br />

o<br />

1<br />

h<br />

o<br />

U<br />

o<br />

=<br />

Ao<br />

A<br />

i<br />

1<br />

h<br />

i<br />

+<br />

1<br />

Ao<br />

ln<br />

o<br />

2πKL<br />

( r / r )<br />

i<br />

1<br />

+<br />

h<br />

o<br />

The general notion, for either the plane wall or cylindrical coordinate<br />

system, is that<br />

U<br />

A<br />

1<br />

=<br />

∑R<br />

th<br />

=<br />

R<br />

1<br />

th , Overall<br />

5.2 CRITICAL THICKNESS OF INSULATION<br />

Fig. 5.4 Critical radius for insulation of cylinder or pipe<br />

Consider a layer of insulation is installed around the outside of a<br />

cylinder whose radius r 1 , is fixed and with a length L. The cylinder has a high<br />

thermal conductivity and the inner temperature T 1 at point r 1 outside the<br />

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 T 2 is<br />

exposed to an environment at T o where convective heat transfer occurs. It is<br />

not obvious if adding more insulation with a thermal conductivity of k will<br />

decrease the heat-transfer rate.<br />

At steady state the heat-transfer rate q through the cylinder and the<br />

insulation equals the rate of convection from the surface:<br />

( T − )<br />

q = h A<br />

2 -------------- (1)<br />

o<br />

T o<br />

T1<br />

−T4<br />

q =<br />

1 ln<br />

o i<br />

+<br />

h A 2πKL<br />

i<br />

i<br />

( r / r )<br />

1<br />

+<br />

h A<br />

As insulation is added, the outside area, which is<br />

o<br />

o<br />

--------- (2)<br />

A<br />

2<br />

= 2πr<br />

L , increases,<br />

but T 2 decreases. However, it is not apparent whether q increases or<br />

decreases. To determine this, an equation similar to Eq. (2) with the<br />

resistance of the insulation represented by Eq.(3) is written using the two<br />

resistances:<br />

r2<br />

−r<br />

R =<br />

K A<br />

1<br />

lm<br />

⎛r<br />

⎞<br />

2<br />

ln<br />

⎜<br />

r<br />

⎟<br />

1<br />

=<br />

⎝ ⎠<br />

2πKL<br />

A<br />

lm = log mean area<br />

2<br />

q =<br />

ln<br />

2<br />

K<br />

πL<br />

( T1<br />

−To<br />

)<br />

( r / r ) 1<br />

1<br />

+<br />

r h<br />

2<br />

o<br />

As the outside radius, r 2 , increases, then in the denominator, the first<br />

term increases but the second term decreases.<br />

Thus, there must be a critical radius, r c , that will allow maximum rate of<br />

heat transfer, q.<br />

To determine the effect of the thickness of insulation on q, we take the<br />

derivative of q with respect to r 2 , equate this result to zero, and obtain the<br />

following for maximum heat flow. The maximization condition is<br />

d<br />

q<br />

dr<br />

=<br />

−2πL<br />

( T −T<br />

⎡<br />

⎢<br />

⎣<br />

1<br />

( r / r )<br />

⎛ 1<br />

)<br />

⎜<br />

⎝ r2<br />

K<br />

2 ln<br />

2 1<br />

1<br />

K<br />

o<br />

+<br />

r h<br />

2<br />

1<br />

−<br />

r h<br />

0<br />

⎤<br />

⎥<br />

⎦<br />

2<br />

2<br />

2<br />

o<br />

⎞<br />

⎟<br />


T 1 ,T 0 , K, L, r o , r i are constant terms.<br />

Therefore,<br />

⎛ 1<br />

⎜<br />

⎝r2<br />

K<br />

1<br />

−<br />

2<br />

r<br />

1 1 =<br />

r K r<br />

2<br />

2<br />

h o<br />

2<br />

2<br />

h o<br />

⎞<br />

⎟=<br />

0<br />

⎠<br />

When outside radius becomes equal to critical radius, or r 2 = r c , we get<br />

( r 2<br />

) =<br />

cr<br />

k<br />

h<br />

o<br />

Where (r 2 ) cr is the value of the critical radius when the heat transfer rate<br />

is a maximum. Hence, if the outer radius r 2 is less than the critical value,<br />

adding more insulation will actually increase the heat- transfer rate q. Also, if<br />

the outer radius is greater than the critical, adding more insulation will<br />

decrease the heat transfer rate. Using values of K and h o typically<br />

encountered, the critical radius is only a few mm. As a result, adding<br />

insulation on small electrical wires could increase the heat loss. Adding<br />

insulation to large pipes decreases the heat transfer rate.<br />

radius of r 2 ,<br />

When no insulation is provided then for a metal pipe with an outside<br />

q<br />

bare<br />

= 2πr<br />

2<br />

L ho<br />

2<br />

( T −T<br />

)<br />

The rate of heat transfer from an insulated pipe, where the annular<br />

insulating shell has an inside radius of r 2 and an outer radius of r 3 ,<br />

o<br />

q<br />

insulated<br />

Then,<br />

2 r3<br />

L ho<br />

= π<br />

r3<br />

h<br />

1+<br />

K<br />

( T −T<br />

)<br />

o<br />

2<br />

r<br />

ln<br />

r<br />

3<br />

2<br />

o<br />

q<br />

q<br />

insulated<br />

bare<br />

r<br />

=<br />

r<br />

3<br />

2<br />

⎡<br />

⎢ 1<br />

⎢<br />

⎢ r3<br />

ho<br />

1+<br />

⎢<br />

⎣ K<br />

ln<br />

⎤<br />

⎥<br />

⎥<br />

r3<br />

⎥<br />

r ⎥<br />

2 ⎦


LECTURE NO.6<br />

HEAT SOURCE SYSTEMS: ONE-DIMENSIONAL STEADY STATE HEAT<br />

CONDUCTION WITH HEAT GENERATION: HEAT FLOW THROUGH<br />

SLAB / PLANE WALL<br />

HEAT-SOURCE SYSTEMS<br />

A number of interesting applications of the principles of heat transfer<br />

are concerned with systems in which heat may be generated internally.<br />

Nuclear reactors are one example; electrical conductors and chemically<br />

reacting systems are others.<br />

Fig.6.1 Sketch illustrating one- dimensional conduction problem with<br />

heat generation.<br />

One-dimensional steady state heat conduction with heat generation:<br />

Heat Flow through slab / Plane Wall<br />

Consider the plane wall with uniformly distributed heat sources shown<br />

in the above Figure. The thickness of the wall in the x direction is 2L, and it is<br />

assumed that, the dimensions in the other directions are sufficiently large that<br />

the heat flow may be considered as one-dimensional. The heat generated per<br />

unit volume is q • , and we assume that the thermal conductivity does not vary<br />

with temperature. (This situation might be produced in a practical situation by<br />

passing a current through an electrically conducting material)<br />

The differential equation which governs the heat flow is


2<br />

d T<br />

2<br />

dx<br />

.<br />

q<br />

+ = 0 k<br />

2<br />

d T q<br />

=−<br />

2<br />

dx k<br />

Integrating twice with respect to x results in<br />

dT<br />

dx<br />

.<br />

.<br />

q<br />

=− x+<br />

c<br />

k<br />

.<br />

q 2<br />

T = − x + c1x<br />

+ c -------------- (1)<br />

2<br />

2k<br />

For the boundary conditions we specify the temperatures<br />

on either side of the wall, i.e.,<br />

T = T w at x = ± L<br />

1<br />

Since the temperature must be the same on each side of the wall, c 1<br />

must be zero.<br />

dT<br />

= 0,<br />

q<br />

• = 0⇒c<br />

1<br />

= 0<br />

The temperature at the mid plane (x = 0) is denoted by T o and from<br />

Equation (1)<br />

At mid plane = x=0 and T= T 0<br />

T o = c 2<br />

The temperature distribution equation (1) becomes<br />

2<br />

T =− x +<br />

.<br />

q<br />

2k<br />

T<br />

0<br />

T<br />

.<br />

q<br />

2k<br />

x<br />

2<br />

−T<br />

0<br />

= − --------------- (2)<br />

Assumed T=T w at x= L<br />

.<br />

q<br />

2k<br />

L<br />

2<br />

T w<br />

−T<br />

0<br />

= − ----------------- (3)<br />

(2)<br />

(3)<br />

T −T0<br />

⇒<br />

T −T<br />

w<br />

0<br />

2<br />

⎛ x ⎞<br />

= ⎜ ⎟<br />

⎝ L ⎠<br />

------------ (4)


An expression for the midplane temperature T o may be obtained through an<br />

energy balance. At steady-state conditions the total heat generated must<br />

equal the heat lost at the faces. Thus<br />

⎛<br />

2⎜−<br />

KA<br />

⎝<br />

dT<br />

dx<br />

⎤<br />

⎥<br />

⎦<br />

x=<br />

L<br />

⎞ •<br />

⎟=<br />

q.<br />

A.2L<br />

⎠<br />

where A is the cross-sectional area of the plate. The temperature gradient at<br />

the wall is obtained by differentiating Equation (4):<br />

T −T0<br />

T −T<br />

w<br />

0<br />

⎛<br />

= ⎜<br />

⎝<br />

x<br />

L<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

T −T<br />

x<br />

⎝L<br />

⎛ ⎞<br />

0<br />

= ( Tw<br />

−T0<br />

) ⎜ ⎟ ⎠<br />

2x<br />

dT =<br />

w 0<br />

.<br />

2<br />

L<br />

( T −T<br />

) dx<br />

dT<br />

⎤<br />

⎦<br />

at x= L ( T T )<br />

dx ⎥<br />

=<br />

w−<br />

o<br />

. L<br />

x=<br />

L<br />

2<br />

2<br />

Since,<br />

dT<br />

−K.<br />

dx<br />

= q<br />

•<br />

L<br />

dT<br />

dx<br />

•<br />

q L<br />

= − K<br />

•<br />

qL<br />

T w<br />

− 2 0<br />

=−<br />

L K<br />

( T )<br />

•<br />

2<br />

qL<br />

T w<br />

−T<br />

0<br />

=−<br />

2K<br />

•<br />

2<br />

qL<br />

T<br />

0<br />

= + T w<br />

2K<br />

----- (5)


LECTURE NO.7<br />

STEADY STATE HEAT CONDUCTION WITH HEAT DISSIPATION TO<br />

ENVIRONMENT- INTRODUCTION TO EXTENDED SURFACES<br />

(FINS) OF UNIFORM AREA OF CROSS SECTION - DIFFERENT FIN<br />

CONFIGURATIONS - GENERAL CONDUCTION ANALYSIS<br />

EQUATION<br />

Steady State heat conduction with heat dissipation to environment:<br />

Introduction to extended surfaces (FINS) of Uniform area of cross<br />

section<br />

The term extended surface is commonly used to depict an important<br />

special case involving heat transfer by conduction within a solid and heat<br />

transfer by convection (and/or radiation) from the boundaries of the solid.<br />

Heat transfer from the boundaries of a solid to be in the same direction as<br />

heat transfer by conduction in the solid. In contrast, for an extended surface,<br />

the direction of heat transfer from the boundaries is perpendicular to the<br />

principal direction of heat transfer in the solid.<br />

Fig. 7.1 Combined conduction and convection in a structural element.<br />

Consider a strut that connects two walls at different temperatures and<br />

across which there is fluid flow (Figure 7.1). With T 1 >T 2 , temperature gradients<br />

in the x-direction sustain heat transfer by conduction in the strut. However,<br />

with T 1 >T 2 > T∞there is concurrent heat transfer by convection to the fluid,<br />

causing q x , and hence the magnitude of the temperature gradient,<br />

decrease with increasing x.<br />

dT / dx<br />

An extended surface is used specifically to enhance heat transfer<br />

between a solid and adjoining fluid. Such an extended surface is termed a fin.<br />

, to


The heat transfer rate may be increased by increasing the surface area<br />

across which the convection occurs. This is done by employing fins that<br />

extend from the wall into the surrounding fluid. The thermal conductivity of the<br />

fin material can have a strong effect on the temperature distribution along the<br />

fin and therefore influences the degree to which the heat transfer rate is<br />

enhanced. Ideally, the fin material should have a large thermal conductivity to<br />

minimize temperature variations from its base to its tip. In the limit of infinite<br />

thermal conductivity, the entire fin would be at the temperature of the base<br />

surface, thereby providing the maximum possible heat transfer enhancement.<br />

Fig. 7.2 Use of fins to enhance heat transfer from a plane wall (a) Bare<br />

surface (b) Finned surface.<br />

Examples of fin applications are easy to find. Consider the<br />

arrangement for cooling engine heads on motorcycles and lawn mowers or for<br />

cooling electric power transformers. Consider also the tubes with attached fins<br />

used to promote heat exchange between air and the working fluid of an air<br />

conditioner. Two common finned tube arrangements are shown in Figure 7.3<br />

Fig.7.3 Schematic of typical finned-tube heat exchangers.


Different fin configurations<br />

Different fin configurations are illustrated in Figure 7.4. A straight fin is<br />

any extended surface that is attached to a plane wall. It may be of uniform<br />

cross-sectional area, or its cross-sectional area may vary with the distance x<br />

from the wall. An annular fin is one that is circumferentially attached to a<br />

cylinder, and its cross section varies with radius from the wall of the cylinder.<br />

The foregoing fin types have rectangular cross sections, whose area may be<br />

expressed as a product of the fin thickness t and the width w for straight fins<br />

or the circumference<br />

2 πr<br />

for annular fins. In contrast a pin fin, or spine, is an<br />

extended surface of circular cross section. Pin fins may also be of uniform or<br />

non-uniform cross section. In any application, selection of a particular fin<br />

configuration may depend on space, weight, manufacturing, and cost<br />

considerations, as well as on the extent to which the fins reduce the surface<br />

convection coefficient and increase the pressure drop associated with flow<br />

over the fins.<br />

FIGURE 7.4 Fin configurations. (a) Straight fin of uniform cross section. (b)<br />

Straight fin of non-uniform cross section. (c) Annular fin. (d) Pin fin / spine.<br />

General Conduction Analysis Equation<br />

As engineers we are primarily interested in knowing the extent to which<br />

particular extended surfaces or fin arrangements could improve heat transfer<br />

from a surface to the surrounding fluid. To determine the heat transfer rate<br />

associated with a fin, we must first obtain the temperature distribution along<br />

the fin.


Fig.7.5 Energy balance for an extended surface<br />

A general form of the energy equation for an extended surface is as<br />

follows:<br />

2<br />

d T<br />

2<br />

dx<br />

⎛ 1<br />

⎜<br />

⎝ Ac<br />

dA<br />

dx<br />

⎞dT<br />

⎟<br />

⎠ dx<br />

⎛ 1<br />

−<br />

⎜<br />

⎝ Ac<br />

h dA<br />

k dx<br />

⎞<br />

⎟<br />

⎠<br />

( T −T<br />

) = 0<br />

c<br />

s<br />

+<br />

∞<br />

Its solution for appropriate boundary conditions provides the<br />

temperature distribution, which may be used with Fourier’s equation (<br />

dT<br />

q x<br />

= −kA ) to calculate the conduction rate at any x.<br />

dx


LECTURE NO.8<br />

EQUATION OF TEMPERATURE DISTRIBUTION WITH DIFFERENT<br />

BOUNDARY CONDITIONS, FIN PERFORMANCE AND OVERALL<br />

SURFACE EFFICIENCY OF FINS<br />

Fins of Uniform Cross-Sectional Area<br />

To solve general form of fin energy equation it is necessary to be more<br />

specific about the geometry. We begin with the simplest case of straight<br />

rectangular and pin fins of uniform cross section (Figure 8.1). Each fin is<br />

attached to a base surface of temperature T(O) = T b and extends into a fluid<br />

of temperature<br />

T<br />

∞.<br />

Fig. 8.1 Straight fins of uniform cross section (a) Rectangular fin (b) Pin<br />

fin<br />

For the prescribed fins, A c is a constant and A s = P x , where A s is the<br />

surface area measured from the base to x and P is the fin perimeter.<br />

Accordingly, with dA c /dx = 0 and dA s /dx = P, general form of the energy<br />

equation for an extended surface reduces to<br />

2<br />

d T<br />

2<br />

dx<br />

hP<br />

−<br />

kA<br />

c<br />

( T − T ) = 0<br />

∞<br />

--------(1)<br />

To simplify the form of this equation, we transform the dependent<br />

variable by defining an excess temperature θ as<br />

where, since<br />

into<br />

∞<br />

Equation (3), we then obtain<br />

θ ( x ) = T ( x)<br />

−T<br />

-------------(2)<br />

∞<br />

T is a constant, d θ / dx = dT / dx . . Substituting Equation (2)


where<br />

2<br />

d θ 2<br />

− θ= 0<br />

2 m<br />

dx<br />

hP<br />

m 2 =<br />

-------------(4)<br />

kA c<br />

--------(3)<br />

Equation (3) is a linear, homogeneous, second-order differential equation with<br />

constant coefficients. Its general solution is of the form<br />

mx −mx<br />

θ ( x)<br />

= C1<br />

e + C2e<br />

---------------- (5)<br />

By substitution it may readily be verified that Equation (5) is indeed a<br />

solution to Equation (3).<br />

To evaluate the constants C 1 and C 2 of Equation (5), it is necessary to<br />

specify appropriate boundary conditions. One such condition may be specified<br />

in terms of the temperature at the base of the fin (x = 0)<br />

θ =<br />

( 0)= Tb<br />

−T<br />

∞<br />

θb<br />

----------------- (6)<br />

The second condition, specified at the fin tip (x = L), may correspond to<br />

one of four different physical situations.<br />

The first condition, case A, considers convection heat transfer from the<br />

fin tip. Applying an energy balance to a control surface about this tip (Figure<br />

8.2), we obtain<br />

hA<br />

c<br />

[ T ( L)−T<br />

]<br />

∞<br />

= −kA<br />

dT<br />

dx<br />

x=<br />

L<br />

or<br />

h<br />

dT<br />

( L)=−<br />

kA<br />

dx<br />

θ --------------- (7)<br />

x=<br />

L<br />

Figure 8.2 Conduction and convection in a fin of uniform cross section.


That is, the rate at which energy is transferred to the fluid by<br />

convection from the tip must equal the rate at which energy reaches the tip by<br />

conduction through the fin. Substituting Equation (5) into Equations (6) and<br />

(7), we obtain, respectively,<br />

And<br />

θb = C 1<br />

+ C 2 ----------- (8)<br />

mL −mL<br />

−mL<br />

h( C1e<br />

+ C2e<br />

) = km(<br />

C2e<br />

−C1<br />

e<br />

Solving for C 1 , and C 2 , it may be shown, after some manipulation, that<br />

θ cosh m(<br />

L − x)<br />

+ ( h / mk )sinh m(<br />

L − x)<br />

=<br />

θ cosh mL + ( h / mk ) sinh mL<br />

b<br />

mL<br />

)<br />

------- (9)<br />

The form of this temperature distribution is shown schematically in<br />

Figure 8.2. <strong>No</strong>te that the magnitude of the temperature gradient decreases<br />

with increasing x. This trend is a consequence of the reduction in the<br />

conduction heat transfer q x<br />

(x)<br />

with increasing x due to continuous<br />

convection losses from the fin surface.<br />

We are particularly interested in the amount of heat transferred from<br />

the entire fin. From Figure 8.2 it is evident that the fin heat transfer rate<br />

may be evaluated in two alternative ways, both of which involve use of the<br />

temperature distribution. The simpler procedure, and the one that we will use,<br />

involves applying Fourier's law at the fin base. That is,<br />

q<br />

f<br />

= q<br />

b<br />

=− kA<br />

c<br />

dT<br />

dx<br />

=− kA<br />

dθ<br />

c<br />

x=0 dx<br />

x=0<br />

q<br />

f<br />

----- (10)<br />

Hence, knowing the temperature distribution, θ (x)<br />

, q<br />

f may be<br />

evaluated, giving<br />

q<br />

f<br />

sinh mL + ( h / mk )cosh mL<br />

= hPkA<br />

cθ b<br />

------- (11)<br />

cosh mL + ( h / mk )sinh mL<br />

However, conservation of energy dictates that the rate at which heat is<br />

transferred by convection from the fin must equal the rate at which it is<br />

conducted through the base of the fin. Accordingly, the alternative formulation<br />

for<br />

q<br />

f is<br />

q<br />

q<br />

f<br />

f<br />

∫<br />

A f<br />

[ ( x T∞] dA<br />

s<br />

= h T )−<br />

∫A<br />

f<br />

= h )<br />

θ ( x dA<br />

s -------------- (12)


where<br />

Af<br />

Af is the total, including the tip, fin surface area. Substitution of<br />

Equation (9) into Equation (12) would yield Equation (11):<br />

The second tip condition, case B, corresponds to the assumption that<br />

the convective heat loss from the fin tip is negligible, in which case the tip may<br />

be treated as adiabatic and<br />

dθ<br />

dx<br />

x =L<br />

=0<br />

-------- (13)<br />

Substituting from Equation (5) and dividing by m, we then obtain<br />

C<br />

mL −mL<br />

1<br />

e −C2e<br />

=<br />

Using this expression with Equation (8) to solve for C 1 and C 2 and<br />

substituting the results into Equation (5), we obtain<br />

θ cosh m(<br />

L −x)<br />

=<br />

θ cosh mL<br />

b<br />

0<br />

--------- (14)<br />

Using this temperature distribution with Equation (10), the fin heat<br />

transfer rate is then<br />

q<br />

f<br />

= hPkA<br />

cθb<br />

tanh mL ----- (15)<br />

In the same manner, we can obtain the fin temperature distribution<br />

and, heat transfer rate for case C, where the temperature is prescribed at the<br />

fin tip. That is, the second boundary condition is θ ( L)<br />

= θ , and the resulting<br />

expressions are of the form<br />

θ<br />

=<br />

θ<br />

b<br />

( θ θ )<br />

L<br />

/<br />

b<br />

sinh mx + sinh m(<br />

L − x)<br />

sinh mL<br />

L<br />

------- (16)<br />

q<br />

f<br />

=<br />

cosh mL −θ<br />

L<br />

/ θb<br />

hPkA<br />

cθb<br />

sinh mL<br />

The very long fin, case D, is an interesting extension of these results.<br />

In particular, as L →∞, θ →0<br />

and it is easily verified that<br />

L<br />

θ<br />

=e −<br />

θ<br />

b<br />

mx<br />

q<br />

f<br />

=<br />

hPkA<br />

c<br />

θ<br />

b<br />

Fin Performance<br />

Fins are used to increase the heat transfer from a surface by<br />

increasing the effective surface area. However, the fin itself represents a<br />

conduction resistance to heat transfer from the original surface. For this


eason, there is no assurance that the heat transfer rate will be increased<br />

through the use of fins. An assessment of this matter may be made by<br />

evaluating the fin effectiveness ε<br />

f . It is defined as the ratio of the fin heat<br />

transfer rate to the heat transfer rate that would exist without the fin. Therefore<br />

q<br />

f<br />

εf<br />

=<br />

hA θ<br />

c,<br />

b<br />

b<br />

where<br />

A<br />

c , b<br />

is the fin cross-sectional area at the base. In any rational design<br />

the value of εf<br />

should be as large as possible, and in general, the use of fins<br />

may rarely be justified unless ε<br />

f<br />

≥ 2.<br />

Fin effectiveness is enhanced by the choice of a material of high<br />

thermal conductivity. Aluminum alloys and copper come to mind. However,<br />

although copper is superior from the stand point of thermal conductivity.<br />

aluminum alloys are the more common choice because of additional benefits<br />

related to lower cost and weight. Fin effectiveness is also enhanced by<br />

increasing the ratio of the perimeter to the cross-sectional area. For this<br />

reason, the use of thin, but closely spaced fins, is preferred.<br />

Another measure of fin thermal performance is provided by the fin<br />

efficiencyη f . The maximum driving potential for convection is the<br />

temperature difference between the base (x = 0) and the fluid, θ b<br />

= T b<br />

−T<br />

∞.<br />

Hence the maximum rate at which a fin could dissipate energy is the rate that<br />

would exist if the entire fin surface were at the base temperature. However,<br />

since any fin is characterized by a finite conduction resistance, a temperature<br />

gradient must exist along the fin and the above condition is an idealization. A<br />

logical definition of fin efficiency is therefore<br />

q<br />

f<br />

η<br />

f<br />

= =<br />

qmax<br />

q<br />

hA<br />

f<br />

f<br />

θ<br />

b<br />

where A f is the surface area of the fin.<br />

Overall Surface Efficiency<br />

In contrast to the fin efficiency<br />

η<br />

f , which characterizes the<br />

performance of a single fin, the overall surface efficiency η<br />

o characterizes an<br />

array of fins and the base surface to which they are attached. Representative<br />

arrays are shown in Figure 8.3, where S designates the fin pitch. In each case<br />

the overall efficiency is defined as


where<br />

q<br />

f<br />

η<br />

f<br />

= =<br />

qmax<br />

qt<br />

hA θ<br />

t<br />

b<br />

q<br />

t is the total heat rate from the surface area A t associated with both<br />

the fins and the exposed portion of the base (often termed the prime surface).<br />

If there are N fins in the array, each of surface area A f , and the area of the<br />

prime surface is designated as A b , the total surface area is<br />

A = NA + A<br />

t<br />

f<br />

b<br />

The maximum possible heat rate would result if the entire fin surface,<br />

as well as the exposed base, were maintained at<br />

T<br />

b .<br />

Figure 8.3 Representative fin arrays. (a) Rectangular fins. (b) Annular fins.


LECTURE NO.9<br />

PRINCIPLES OF UNSTEADY-STATE HEAT TRANSFER: DERIVATION OF<br />

BASIC EQUATION; SIMPLIFIED CASE FOR SYSTEMS WITH<br />

NEGLIGIBLE INTERNAL RESISTANCE; TOTAL AMOUNT OF HEAT<br />

TRANSFERRED ; DIMENSIONAL ANALYSIS IN MOMENTUM TRANSFER<br />

PRINCIPLES OF UNSTEADY-STATE HEAT TRANSFER: DERIVATION OF<br />

BASIC EQUATION<br />

Introduction<br />

In steady state heat-transfer systems the temperature at any given<br />

point and the heat flux were always constant over time. In unsteady state or<br />

transient processes the temperature at any given point in the system changes<br />

with time. Before steady-state conditions can be reached in a process, some<br />

time must elapse after the heat-transfer process is initiated to allow the<br />

unsteady-state conditions to disappear.<br />

Unsteady-state heat transfer is important because of the large<br />

number of heating and cooling problems occurring industrially. In metallurgical<br />

processes it is necessary to predict cooling and heating rates for various<br />

geometries of metals in order to predict the time required to reach certain<br />

temperatures. In food processing, for example, the canning industry,<br />

perishable canned foods are heated by immersion in steam baths or chilled by<br />

immersion in cold water. In the paper industry wood logs are immersed in<br />

steam baths before processing. In most of these processes the material is<br />

suddenly immersed in a fluid of higher or lower temperature.<br />

Fig. 9.1 Unsteady State conduction in one direction


Derivation of Unsteady-State Conduction Equation<br />

To derive the equation for unsteady-state conduction in one direction in<br />

a solid, refer to Fig. 9.1. Heat is being conducted in the x direction in the<br />

∆ x, ∆y<br />

, ∆z<br />

in size. For conduction in the x direction, write<br />

q x<br />

∂T<br />

= −kA<br />

------- (1)<br />

∂x<br />

∂T<br />

The term means the partial or derivative of T with respect to x,<br />

∂x<br />

with the other variables, y, z, and time t, being held constant. Next, making a<br />

heat balance on the cube, we can write<br />

rate of heat input + rate of generation = rate of heat output + rate of heat accumulati on<br />

---------- (2)<br />

The rate of heat input to the cube is<br />

rate of heat input =<br />

q<br />

∂T<br />

x / x<br />

= −k<br />

( ∆y<br />

∆z<br />

) ----------(3)<br />

∂x<br />

x<br />

is<br />

Also, rate of heat output =<br />

The rate of accumulation of heat in the volume<br />

q<br />

∂T<br />

x / x +∆ x<br />

= −k<br />

( ∆y<br />

∆z<br />

)<br />

------- (4)<br />

∂x<br />

x+∆<br />

x<br />

, in time ∂ t<br />

∆ x ∆y<br />

, ∆z<br />

∂T<br />

rate of heat accumulation = ( ∆x , ∆y<br />

, ∆z<br />

) ρc<br />

------(5)<br />

p<br />

∂t<br />

The rate of heat generation in volume ∆ x, ∆y<br />

, ∆z<br />

is<br />

rate of heat generation =<br />

( ∆x<br />

∆y<br />

∆z<br />

) q ------(6)<br />

.<br />

Substituting Eqs. (3)-(6) into (2) and dividing by<br />

Letting<br />

respect to x or<br />

⎛∂T<br />

−k<br />

⎜<br />

x<br />

q<br />

⎝ ∂<br />

+<br />

.<br />

x<br />

∂T<br />

−<br />

∂x<br />

∆x<br />

x+∆x<br />

⎞<br />

⎟<br />

⎠<br />

= ρc<br />

p<br />

∂T<br />

∂t<br />

∆ x,<br />

∆y<br />

, ∆z<br />

------- (7)<br />

∆ x approach zero, we have the second partial of T with<br />

2<br />

∂T / ∂x<br />

2<br />

on the left side. Then, rearranging,<br />

∂T<br />

∂t<br />

k<br />

=<br />

ρc<br />

p<br />

2<br />

∂ T<br />

2<br />

∂x<br />

.<br />

q<br />

+<br />

ρc<br />

p<br />

2<br />

∂ T<br />

= α<br />

2<br />

∂x<br />

.<br />

q<br />

+<br />

ρc<br />

p<br />

--------(8)


where α is<br />

k / ρc<br />

p , thermal diffusivity. This derivation assumes constant<br />

k ,and<br />

,ρ c p . In SI units, α = m 2 /s, T = K, t = s, k = W/m.K, ρ = kg/m 3 , q . =<br />

W/m 3 , and<br />

c<br />

p = J/kg.K.<br />

For conduction in three dimensions, a similar derivation gives<br />

2<br />

∂T<br />

⎛∂<br />

T<br />

= α<br />

2<br />

t<br />

⎜<br />

∂ ⎝ ∂x<br />

2<br />

∂ T<br />

+<br />

2<br />

∂y<br />

2<br />

∂ T ⎞ q<br />

+<br />

2<br />

z<br />

⎟<br />

∂<br />

+<br />

⎠ ρ<br />

.<br />

c p<br />

--------- (9)<br />

In many cases, unsteady-state heat conduction is occurring but the<br />

rate of heat generation is zero. Then Eqs. (8) and (9) become<br />

∂T<br />

∂t<br />

2<br />

∂ T<br />

= α ------(10)<br />

2<br />

∂x<br />

2 2 2<br />

∂T ⎛∂<br />

T ∂ T ∂ T ⎞<br />

= α ⎜ + +<br />

⎟<br />

2 2 2 ----- (11)<br />

∂t<br />

⎝ ∂x<br />

∂y<br />

∂z<br />

⎠<br />

z and time t.<br />

Equations (10) and (11) relate the temperature T with position x, y, and<br />

SIMPLIFIED CASE FOR SYSTEMS WITH NEGLIGIBLE INTERNAL<br />

RESISTANCE<br />

Basic Equation<br />

Consider a solid which has a very high thermal conductivity or very low<br />

internal conductive resistance compared to the external surface resistance,<br />

where convection occurs from the external fluid to the surface of the solid.<br />

Since the internal resistance is very small, the temperature within the solid is<br />

essentially uniform at any given time.<br />

An example would be a small, hot cube of steel at T o K at time t = 0,<br />

suddenly immersed in a large bath of cold water at<br />

T ∞which is held constant<br />

with time. Assume that the heat transfer coefficient h in W/m 2 .K is constant<br />

with time. Making a heat balance on the solid object for a small time interval of<br />

time dt s, the heat transfer from the bath to the object must equal the change<br />

in internal energy of the object:<br />

hA<br />

( T T ) dt = c V dT<br />

∞ −<br />

p<br />

ρ<br />

where A is the surface area of the object in m 2 , T the average<br />

temperature of the object at time t in s, ρ the density of the object in kg/m 3 ,


and V the volume in m 3 . Rearranging the equation and integrating between<br />

the limits of T = T o when t = 0 and T = T when t = t,<br />

T = T<br />

∫<br />

dT<br />

T −T<br />

T = T0<br />

∞<br />

hA<br />

=<br />

c ρV<br />

p<br />

∫t<br />

t = t<br />

= 0<br />

dt<br />

⎛ hA<br />

t<br />

c p V ⎟ ⎞<br />

−⎜<br />

⎝ ρ ⎠<br />

T −T<br />

∞<br />

= e<br />

T0<br />

−T<br />

∞<br />

This equation describes the time-temperature history of the solid<br />

object. The term<br />

c p<br />

ρ V is often called the lumped thermal capacitance of the<br />

system. This type of analysis is often called the lumped capacity method or<br />

Newtonian heating or cooling method.<br />

Equation for Different Geometries<br />

In using the above equation the surface / volume ratio of the object<br />

must be known. The basic assumption of negligible internal resistance was<br />

made in the derivation. This assumption is reasonably accurate when<br />

N Bi<br />

hx<br />

= 1<br />


calculate the temperature of the ball after 1 h (3600 s). The average physical<br />

properties are k = 43.3 W/m.K, ρ = 7849 kg/m3 , and c p = 0.4606 kJ/kg.K.<br />

Solution: For a sphere characteristic dimension<br />

V 25 .4<br />

x1 = =<br />

r = = 8.47 × 10<br />

A 3 1000 × 3<br />

The Biot Number<br />

hx 11 .36 (8.47 × 10<br />

N Bi<br />

= =<br />

k 43.3<br />

−3<br />

m<br />

−3<br />

1 =<br />

)<br />

0.00222<br />

Then,<br />

This value is


q( t)<br />

= hA ( T0<br />

−T<br />

∞<br />

) e<br />

⎛ hA ⎞<br />

−⎜<br />

⎟t<br />

c p V<br />

⎝ ρ ⎠<br />

To determine the total amount of heat Q in W.s or J transferred from<br />

the solid from time t = 0 to t = t, we can integrate the above equation:<br />

Q=<br />

Q=<br />

c<br />

t<br />

∫<br />

= t<br />

t = 0<br />

p<br />

q(<br />

t)<br />

d(<br />

t)<br />

=<br />

t<br />

∫<br />

= t<br />

t = 0<br />

hA ( T<br />

⎛<br />

⎜<br />

hA<br />

0<br />

−T<br />

⎞<br />

⎟<br />

− t<br />

⎜c<br />

V ⎟<br />

⎝ pρ<br />

⎠<br />

ρV<br />

( T 0<br />

−T<br />

∞<br />

)[1 −e<br />

]<br />

∞<br />

) e<br />

⎛ ⎞<br />

⎜<br />

hA<br />

− ⎟t<br />

⎝<br />

cpρV<br />

⎠<br />

dt<br />

Problem 9.2 Total Amount of Heat in Cooling<br />

For the conditions in problem 9.1, calculate the total amount of heat<br />

removed up to time t = 3600 s.<br />

Solution:<br />

From Problem 9.1,<br />

hA<br />

cpρV<br />

11 .36<br />

−4<br />

−1<br />

= 3.71 × 10<br />

− 3<br />

= (0.4606 × 1000 )(7849 )(8.47 × 10 )<br />

s<br />

Q=<br />

c<br />

3<br />

3<br />

−5<br />

3<br />

Also, V = 4πr<br />

/3=<br />

4( π)(0.0254<br />

) / 3=<br />

6.864 × 10 m<br />

Substituting the values into<br />

p<br />

⎛<br />

⎜<br />

hA<br />

⎞<br />

⎟<br />

− t<br />

⎜c<br />

V ⎟<br />

⎝ pρ<br />

⎠<br />

ρV<br />

( T 0<br />

−T<br />

∞<br />

)[1 −e<br />

]<br />

−4<br />

−5<br />

−(3.71<br />

× 10 )( 3600 )<br />

( 0.4606 × 1000 )( 7849 )( 6.864 × 10 )(699<br />

.9−394 .3) × [1 −<br />

]<br />

Q = e<br />

= 5.589X10 4 J<br />

DIMENSIONAL ANA LYSIS IN MOMENTUM TRANSFER<br />

Dimensional Analysis of Differential Equations<br />

Dimensional homogeneity requires that every term in a given equation<br />

have the same units. Then, the ratio of one term in the equation to another<br />

term is dimensionless. Knowing the physical meaning of each term in the<br />

equation, we are then able to give a physical interpretation to each of the<br />

dimensionless parameters or numbers formed. These dimensionless<br />

numbers, such as the Reynolds number and others, are useful in correlating<br />

and predicting transport phenomena in laminar and turbulent flow.<br />

Often it is not possible to integrate the differential equation describing a<br />

flow situation. However, we can use the equation to find out which


dimensionless numbers can be used in correlating experimental data for this<br />

physical situation.<br />

Systems that are geometrically similar are said to be dynamically<br />

similar if the parameters representing ratios of forces pertinent to the situation<br />

are equal. This means that the Reynolds, Euler, or Froude numbers must be<br />

equal between the two systems.<br />

This dynamic similarity is an important requirement in obtaining<br />

experimental data for a small model and extending these data to scale up to<br />

the large prototype. Since experiments with full-scale prototypes would often<br />

be difficult and / or expensive, it is customary to study small models. This is<br />

done in the scale-up of chemical process equipment and in the design of<br />

ships and airplanes.<br />

Dimensional Analysis Using the Buckingham Method<br />

The method of obtaining the important dimensionless numbers from<br />

the basic differential equations is generally the preferred method. In many<br />

cases, however, we are not able to formulate a differential equation which<br />

clearly applies. Then a more general procedure is required, known as the<br />

Buckingham method. In this method the listing of the important variables in<br />

the particular physical problem is done first. Then we determine the number of<br />

dimensionless parameters into which the variables may be combined by using<br />

the Buckingham pi theorem.<br />

The Buckingham theorem states that the functional relationship among<br />

‘q’ quantities or variables whose units may be given in terms of ‘u’<br />

fundamental units or dimensions may be written as (q - u) independent<br />

dimensionless groups, often called<br />

π ' s . [This quantity u is actually the<br />

maximum number of these variables that will not form a dimensionless group.]<br />

Let us consider the following example, to illustrate the use of this<br />

method. An incompressible fluid is flowing inside a circular tube of inside<br />

diameter D. The significant variables are pressure drop∆ p , velocity v,<br />

diameter D, tube length L, viscosity µ , and density ρ . The total number of<br />

variables is q = 6.


The fundamental units or dimensions are u = 3 and are mass M, length L, and<br />

2<br />

time t. The units of the variables are as follows: ∆pin M / Lt , v in L/t, D in L,<br />

µ in M/Lt, and ρ in M/L 3 . The number of dimensionless groups or π ' s is<br />

q− u , or 6-3 = 3. Thus,<br />

)<br />

π1 = f ( π2<br />

, π3<br />

Next, we must select a core group of u (or 3) variables which will<br />

appear in each π group and among them contain all the fundamental<br />

dimensions. Also, no two of the variables selected for the core can have the<br />

same dimensions. In choosing the core, the variable whose effect one desires<br />

to isolate is often excluded (for example,<br />

∆ p ). This leaves us with the<br />

variables v, D, µ and ρ to be used. (L and D have the same dimensions.)<br />

We will select D, v, and p to be the core variables common to all three<br />

groups. Then the three dimensionless groups are<br />

π<br />

a b c 1<br />

1<br />

= D v ρ ∆p<br />

π =<br />

π<br />

d e f 1<br />

2<br />

D v ρ L<br />

i 1<br />

3<br />

=D<br />

g v h ρ µ<br />

To be dimensionless, the variables must be raised to certain exponents<br />

a, b, c, and so forth.<br />

First we consider the π1<br />

group:<br />

π<br />

a b c 1<br />

1<br />

= D v ρ ∆p<br />

To evaluate these exponents, we write the above equation dimensionally by<br />

substituting the dimensions for each variable:<br />

M<br />

o<br />

o<br />

L t<br />

o<br />

= 1=<br />

L<br />

a<br />

b<br />

c<br />

⎛ L ⎞ ⎛ M ⎞<br />

⎜ ⎟ ⎜ 3 ⎟<br />

⎝ t ⎠ ⎝ L ⎠<br />

Next we equate the exponents of L on both sides of this equation, of M,<br />

and finally of t:<br />

(L) 0 = a + b - 3c - 1<br />

(M) 0 = c + 1<br />

(t) 0 = -b - 2<br />

Solving these equations, a = 0, b = -2, and c = -1.<br />

a b c 1<br />

Substituting these values into Eq. π = D v ρ ∆ ,<br />

1<br />

p<br />

M<br />

Lt<br />

2<br />

= ∆p =<br />

v ρ<br />

. π 1 2<br />

N<br />

Eu


Repeating this procedure for π<br />

2 and π<br />

3 ,<br />

L<br />

π 2<br />

=<br />

D<br />

Dv ρ<br />

π<br />

3<br />

= = N Re<br />

µ<br />

Finally, substituting π<br />

1 , π<br />

2 and π<br />

3 into equation π1 = f ( π2<br />

, π3<br />

)<br />

∆p<br />

⎛<br />

= f ⎜<br />

2<br />

v ρ ⎝<br />

L<br />

D<br />

Dvp ⎞<br />

, ⎟<br />

µ ⎠<br />

This type of analysis is useful in empirical correlations of data.<br />

However, it does not tell us the importance of each dimensionless group,<br />

which must be determined by experimentation, nor does it select the variables<br />

to be used.


LECTURE NO.10<br />

SOME IMPORTANT EMPIRICAL RELATIONS USED FOR DETERMINATION<br />

OF HEAT TRANSFER COEFFICIENT: NUSSELT’S NUMBER, PRANDTL<br />

NUMBER, REYNOLD’S NUMBER, GRASHOFF NUMBER<br />

The Nusselt Number<br />

For forced convection of a single-phase fluid with moderate<br />

temperature differences, the heat flux per unit area<br />

q<br />

"<br />

w<br />

is nearly<br />

proportional to the temperature difference ∆ T = Tw −T<br />

* . From the Newton’s<br />

law of cooling:<br />

q<br />

"<br />

w<br />

=<br />

h<br />

( T −T* )<br />

w<br />

where h is called the heat transfer coefficient, with units of W/m 2 .<br />

But h is dimensional and thus its value depends on the units used. The<br />

traditional dimensionless from of h is the Nusselt number, Nu, which may be<br />

defined as the ratio of convection heat transfer to fluid conduction heat<br />

transfer under the same conditions. Consider a layer of fluid of width L and<br />

temperature difference ( T w<br />

− T*<br />

) . Assuming that the layer is moving so that<br />

convection occurs, the heat flux would be,<br />

q<br />

"<br />

w<br />

=<br />

h<br />

( T −T* )<br />

w<br />

If, on the other hand, the layer were stagnant, the heat flux would be<br />

entirely due to fluid conduction through the layer:


q<br />

k ( Tw<br />

T* =<br />

)<br />

L<br />

" −<br />

w<br />

The Nusselt number is defined as the ratio of these two:<br />

N<br />

"<br />

qw(<br />

convection<br />

=<br />

"<br />

q ( conduction<br />

hL<br />

k<br />

uL<br />

=<br />

w<br />

)<br />

)<br />

A Nusselt number of order unity would indicate a sluggish motion little<br />

more effective than pure fluid conduction: for example, laminar flow in a long<br />

pipe. A large Nusselt number means very efficient convection: For example,<br />

turbulent pipe flow yields Nu of order 100 to 1000.<br />

Prandtl Number(Pr)<br />

The Prandtl number Pr is a dimensionless number approximating the<br />

ratio of momentum diffusivity (kinematic viscosity) and thermal diffusivity. It is<br />

named after the German physicist Ludwig Prandtl. It can be expressed as<br />

Pr<br />

υ<br />

=<br />

α<br />

where<br />

Pr = Prandtl's number<br />

υ = momentum diffusivity (m 2 /s)<br />

α = thermal diffusivity (m 2 /s)<br />

The Prandtl number can alternatively be expressed as<br />

µc<br />

Pr =<br />

p<br />

k<br />

where


μ = absolute or dynamic viscosity (kg/m. s, cP)<br />

c p = specific heat capacity (J/kg K,)<br />

k = thermal conductivity (W/m K)<br />

The Prandtl Number is often used in heat transfer and free and forced<br />

convection calculations.<br />

<strong>No</strong>te that whereas the Reynolds number and Grashof number are<br />

subscripted with a length scale variable, Prandtl number contains no such<br />

length scale in its definition and is dependent only on the fluid and the fluid<br />

state. As such, Prandtl number is often found in property tables alongside<br />

other properties such as viscosity and thermal conductivity.<br />

Typical values for Pr are:<br />

• around 0.015 for mercury<br />

• around 0.16-0.7 for mixtures of noble gases or noble gases with<br />

hydrogen<br />

• around 0.7-0.8 for air and many other gases,<br />

• between 4 and 5 for R-12 refrigerant<br />

• around 7 for water (At 20 degrees Celsius)<br />

• between 100 and 40,000 for engine oil<br />

• around 1 × 10 25 for Earth's mantle.<br />

For mercury, heat conduction is very effective compared to convection:<br />

thermal diffusivity is dominant. For engine oil, convection is very effective in<br />

transferring energy from an area, compared to pure conduction: momentum<br />

diffusivity is dominant.<br />

In heat transfer problems, the Prandtl number controls the relative<br />

thickness of the momentum and thermal boundary layers. When Pr is small, it<br />

means that the heat diffuses very quickly compared to the velocity<br />

(momentum). This means that for liquid metals the thickness of the thermal<br />

boundary layer is much bigger than the velocity boundary layer.


The mass transfer analog of the Prandtl number is the Schmidt number.<br />

Boundary Layer<br />

In physics and fluid mechanics, a boundary layer is that layer of fluid<br />

in the immediate vicinity of a bounding surface where effects of viscosity of<br />

the fluid are considered in detail. In the Earth's atmosphere, the planetary<br />

boundary layer is the air layer near the ground affected by diurnal heat,<br />

moisture or momentum transfer to or from the surface. On an aircraft wing the<br />

boundary layer is the part of the flow close to the wing. The boundary layer<br />

effect occurs at the field region in which all changes occur in the flow pattern.<br />

The boundary layer distorts surrounding non-viscous flow. It is a phenomenon<br />

of viscous forces. This effect is related to the Reynolds number.<br />

Fig.10.1 Boundary layer visualization, showing transition from laminar to<br />

turbulent condition<br />

Reynolds number<br />

In fluid mechanics, the Reynolds number Re is a dimensionless<br />

number that gives a measure of the ratio of inertial forces ρV 2 /L to viscous<br />

forces μV/L 2 and consequently quantifies the relative importance of these two<br />

types of forces for given flow conditions. The concept was introduced by<br />

George Gabriel Stokes in 1851, but the Reynolds number is named after<br />

Osborne Reynolds (1842–1912), who popularized its use in 1883.<br />

Reynolds numbers frequently arise when performing dimensional<br />

analysis of fluid dynamics problems, and as such can be used to determine


dynamic similitude between different experimental cases. They are also used<br />

to characterize different flow regimes, such as laminar or turbulent flow:<br />

laminar flow occurs at low Reynolds numbers, where viscous forces are<br />

dominant, and is characterized by smooth, constant fluid motion, while<br />

turbulent flow occurs at high Reynolds numbers and is dominated by inertial<br />

forces, which tend to produce chaotic eddies, vortices and other flow<br />

instabilities.<br />

Grashof number Gr<br />

The Grashof number Gr is a dimensionless number in fluid dynamics<br />

and heat transfer which approximates the ratio of the buoyancy to viscous<br />

force acting on a fluid. It frequently arises in the study of situations involving<br />

natural convection. It is named after the German engineer Franz Grashof.<br />

Gr<br />

L<br />

3<br />

g β(<br />

T −T∞<br />

) L<br />

= for vertical flat plates<br />

s<br />

2<br />

υ<br />

Gr<br />

L<br />

3<br />

g β(<br />

T −T∞<br />

) D<br />

= for pipes<br />

s<br />

2<br />

υ<br />

where the L and D subscripts indicates the length scale basis for the Grashof<br />

Number.<br />

g = acceleration due to Earth's gravity<br />

β = volumetric thermal expansion coefficient<br />

(equal to approximately 1/T, for ideal fluids, where T is absolute<br />

temperature)<br />

T s = surface temperature<br />

T ∞ = bulk temperature<br />

L = length<br />

D = diameter<br />

υ = kinematic viscosity


The transition to turbulent flow occurs in the range 10 8 < Gr L < 10 9 for<br />

natural convection from vertical flat plates. At higher Grashof numbers, the<br />

boundary layer is turbulent; at lower Grashof numbers, the boundary layer is<br />

laminar.<br />

LECTURE NO.11<br />

RADIATION - HEAT TRANSFER, RADIATION PROPERTIES, RADIATION<br />

THROUGH BLACK AND GREY SURFACES, DETERMINATION OF<br />

SHAPE FACTORS<br />

Introduction<br />

Thermal radiation is that electromagnetic radiation emitted by a body<br />

as a result of its temperature.<br />

PHYSICAL MECHANISM<br />

There are many types of electromagnetic radiation; thermal radiation is<br />

only one. Regardless of the type of radiation, we say that it is propagated at<br />

the speed of light, 3 x 10 8<br />

wavelength and frequency of the radiation,<br />

where<br />

c = speed of light<br />

λ= wavelength<br />

υ = frequency<br />

m/s. This speed is equal to the product of the<br />

c=λυ<br />

The unit for A may be centimeters, angstroms (1<br />

ο<br />

A = 10 -8 cm), or<br />

micrometers (1 µ m = 10<br />

-6<br />

m). A portion of the electromagnetic spectrum is<br />

shown in Figure 11.1. Thermal radiation lies in the range from about 0.1 to<br />

100 µ m, while the visible-light portion of the spectrum is very narrow,<br />

extending from about 0.35 to 0.75 µ m.


Fig.11.1 Electromagnetic spectrum<br />

The propagation of thermal radiation takes place in the form of discrete<br />

quanta, each quantum having an energy of<br />

E=hυ<br />

where h is Planck's constant and has the value<br />

h = 6.625 X 10 -34 J.s<br />

The total radiation energy emitted is proportional to absolute<br />

temperature to the fourth power:<br />

4<br />

E b<br />

= σT<br />

The above equation is called the Stefan-Boltzmann law, E b is the<br />

energy radiated per unit time and per unit area by the ideal radiator, and σ is<br />

the Stefan-Boltzmann constant, which has the value<br />

σ = 5.669 x 10 -8 W/m 2 . K 4<br />

where E b , is in watts per square meter and T is in degrees Kelvin.<br />

We are interested in radiant exchange with surfaces-hence the reason<br />

for the expression of radiation from a surface in terms of its temperature. The<br />

subscript b in Equation denotes that this is the radiation from a blackbody. We<br />

call this blackbody radiation because materials which obey this law appear<br />

black to the eye; they appear black because they do not reflect any radiation.<br />

Thus a blackbody is also considered as one which absorbs all radiation<br />

incident upon it. E b is called the emissive power of a blackbody.<br />

It is important to note at this point that the "blackness" of a surface to<br />

thermal radiation can be quite deceiving in so far as visual observations are<br />

concerned. A surface coated with lampblack appears black to the eye and<br />

turns out to be black for the thermal-radiation spectrum. On the other hand,<br />

snow and ice appear quite bright to the eye but are essentially "black" for


long-wavelength thermal radiation. Many white paints are also essentially<br />

black for long-wavelength radiation.<br />

Fig. 11.2 Sketch showing effects of incident radiation.<br />

RADIATION PROPERTIES<br />

When radiant energy strikes a material surface, part of the radiation is<br />

reflected, part is absorbed, and part is transmitted, as shown in Figure 11.2.<br />

We define the reflectivity ρ as the fraction reflected, the absorptivity α as<br />

the fraction absorbed, and the transmissivity τ as the fraction transmitted.<br />

Thus<br />

ρ+ α+<br />

τ=1<br />

Most solid bodies do not transmit thermal radiation, so that for many<br />

applied problems the transmissivity may be taken as zero. Then<br />

ρ +α=1 ( τ=0<br />

)<br />

Two types of reflection phenomena may be observed when radiation<br />

strikes a surface. If the angle of incidence is equal to the angle of reflection,<br />

the reflection is called specular. On the other hand, when an incident beam is<br />

distributed uniformly in all directions after reflection, the reflection is called<br />

diffuse. These two types of reflection are depicted in Figure 11.3.<br />

Figure 11.3 (a) Specular ( φ=<br />

1<br />

φ2<br />

) and (b) diffuse reflection.


<strong>No</strong>te that a specular reflection presents a mirror image of the source to<br />

the observer. <strong>No</strong> real surface is either specular or diffuse. An ordinary mirror<br />

is quite specular for visible light, but would not necessarily be specular over<br />

the entire wavelength range of thermal radiation. Ordinarily, a rough surface<br />

exhibits diffuse behavior better than a highly polished surface. Similarly, a<br />

polished surface is more specular than a rough surface. The influence of<br />

surface roughness on thermal-radiation properties of materials is a matter of<br />

serious concern and remains a subject for continuing research.<br />

The emissive power of a body E is defined as the energy emitted by<br />

the body per unit area and per unit time.<br />

The ratio of the emissive power of a body to the emissive power of a<br />

blackbody at the same temperature is equal to the absorptivity of the body.<br />

This ratio is defined as the emissivity ε of the body,<br />

E<br />

ε=<br />

E b<br />

so that<br />

ε=α<br />

The above equation is called Kirchhoff's identity.<br />

The Gray Body<br />

A gray body is defined such that the monochromatic emissivity ε λ of<br />

the body is independent of wavelength. The monochromatic emissivity is<br />

defined as the ratio of the monochromatic emissive power of the body to the<br />

monochromatic emissive power of a blackbody at the same wavelength and<br />

temperature. Thus<br />

ε λ<br />

E<br />

=<br />

λ<br />

E b λ<br />

The glass, which is essentially transparent for visible light, is almost<br />

totally opaque for thermal radiation emitted at ordinary room temperatures.


Figure 11.4 Method of constructing a blackbody enclosure.<br />

Construction of a Blackbody<br />

The concept of a blackbody is an idealization; i.e., a perfect blackbody<br />

does not exist. All surfaces reflect radiation to some extent, however slight. A<br />

blackbody may be approximated very accurately, however, in the following<br />

way. A cavity is constructed, as shown in Figure 11.4, so that it is very large<br />

compared with the size of the opening in the side. An incident ray of energy is<br />

reflected many times on the inside before finally escaping from the side<br />

opening. With each reflection there is a fraction of the energy absorbed<br />

corresponding to the absorptivity of the inside of the cavity. After the many<br />

absorptions, practically all the incident radiation at the side opening is<br />

absorbed. It should be noted that the cavity of Figure 11.4 behaves<br />

approximately as a blackbody emitter as well as an absorber.<br />

Fig.11.5 Sketch showing area elements used in deriving radiation shape<br />

factor<br />

RADIATION SHAPE FACTOR


Consider two black surfaces A 1 and A 2 , as shown in Figure 11.5. We<br />

wish to obtain a general expression for the energy exchange between these<br />

surfaces when they are maintained at different temperatures. The problem<br />

becomes essentially one of determining the amount of energy which leaves<br />

one surface and reaches the other. To solve this problem the radiation shape<br />

factors are defined as<br />

F 1-2 = fraction of energy leaving surface 1 which reaches surface 2<br />

F 2-1 = fraction of energy leaving surface 2 which reaches surface 1<br />

F m-n = fraction of energy leaving surface m which reaches surface n<br />

Other names for the radiation shape factor are view factor, angle<br />

factor, and configuration factor. The energy leaving surface 1 and arriving at<br />

surface 2 is<br />

E b1 A 1 F 12<br />

and the energy leaving surface 2 and arriving at surface 1 is<br />

E b2 A 2 F 21<br />

Since the surfaces are black, all the incident radiation will be absorbed, and<br />

the net energy exchange is<br />

E b1 A 1 F 12 - E b2 A 2 F 21 = Q 1-2<br />

If both surfaces are at the same temperature, there can be no heat<br />

exchange, that is, Q 1-2 = 0. Also,<br />

E b1 = E b2<br />

so that<br />

A 1 F 12 = A 2 F 21<br />

The net heat exchange is therefore<br />

Q 1-2 = A 1 F 12 (E b1 - E b2 ) = A 2 F 21 (E b1 - E b2 )<br />

The above equation is known as a reciprocity relation, and it applies in<br />

a general way for any two surfaces m and n:<br />

A m F mn = A n F nm<br />

Although the relation is derived for black surfaces, it holds for other<br />

surfaces also as long as diffuse radiation is involved.


LECTURE NO.12<br />

INTRODUCTION TO CONDENSING AND BOILING HEAT TRANSFER,<br />

CONDENSATION HEAT-TRANSFER PHENOMENA, FILM<br />

CONDENSATION INSIDE HORIZONTAL TUBES, BOILING HEAT<br />

TRANSFER<br />

Introduction<br />

The convection processes are associated with a change of phase of a<br />

fluid. The two most important examples are condensation and boiling<br />

phenomena.<br />

In many types of power or refrigeration cycles one is interested in<br />

changing a vapor to a liquid, or a liquid to a vapor, depending on the particular<br />

part of the cycle under study. These changes are accomplished by boiling or<br />

condensation, and the engineer must understand the processes involved in<br />

order to design the appropriate heat-transfer equipment. High heat-transfer<br />

rates are usually involved in boiling and condensation, and this fact has also<br />

led designers of compact heat exchangers to utilize the phenomena for<br />

heating or cooling purposes not necessarily associated with power cycles.<br />

CONDENSATION HEAT-TRANSFER PHENOMENA<br />

Consider a vertical flat plate exposed to a condensable vapor. If the<br />

temperature of the plate is below the saturation temperature of the vapor,<br />

condensate will form on the surface and under the action of gravity will flow<br />

down the plate. If the liquid wets the surface, a smooth film is formed, and the<br />

process is called film condensation. If the liquid does not wet the surface,<br />

droplets are formed which fall down the surface in some random fashion. This<br />

process is called dropwise condensation.<br />

In the film-condensation process the surface is blanketed by the film,<br />

which grows in thickness as it moves down the plate. A temperature gradient<br />

exists in the film, and the film represents a thermal resistance to heat transfer.<br />

In dropwise condensation a large portion of the area of the plate is<br />

directly exposed to the vapor; there is no film barrier to heat flow, and higher<br />

heat-transfer rates are experienced. In fact, heat-transfer rates in dropwise


condensation may be as much as 10 times higher than in film<br />

condensation.<br />

Because of the higher heat-transfer rates, dropwise condensation<br />

would be preferred to film condensation, but it is extremely difficult to maintain<br />

since most surfaces become wetted after exposure to a condensing vapor<br />

over an extended period of time. Various surface coatings and vapor additives<br />

have been used in attempts to maintain dropwise condensation, but these<br />

methods have not met with general success. Measurements indicate that the<br />

drop conduction is the main resistance to heat flow for atmospheric pressure<br />

and above. Nucleation site density on smooth surfaces can be of the order of<br />

10 8 sites per square centimeter, and heat-transfer coefficients in the range of<br />

170 to 290 kW/m 2 .°C.<br />

FILM CONDENSATION INSIDE HORIZONTAL TUBES<br />

Condensation inside tubes is of considerable practical interest because<br />

of applications to condensers in refrigeration and air-conditioning systems, but<br />

unfortunately these phenomena are quite complicated and not amenable to a<br />

simple analytical treatment. The overall flow rate of vapor strongly influences<br />

the heat transfer rate in the forced convection condensation system, and this<br />

in turn is influenced by the rate of liquid accumulation on the walls.<br />

BOILING HEAT TRANSFER<br />

When a surface is exposed to a liquid and is maintained at a<br />

temperature above the saturation temperature of the liquid, boiling may occur,<br />

and the heat flux will depend on the difference in temperature between the<br />

surface and the saturation temperature. When the heated surface is<br />

submerged below a free surface of liquid, the process is referred to as pool<br />

boiling. If the temperature of the liquid is below the saturation temperature,<br />

the process is called sub-cooled, or local, boiling. If the liquid is maintained<br />

at saturation temperature, the process is known as saturated, or bulk,<br />

boiling.<br />

The different regimes of boiling are indicated in Figure 12.1, where<br />

heat-flux data from an electrically heated platinum wire submerged in water


are plotted against temperature excess T w – T sat . In region I free-convection<br />

currents are responsible for motion of the fluid near the surface. In this region<br />

the liquid near the heated surface is superheated slightly, and it subsequently<br />

evaporates when it rises to the surface. The heat transfer in this region can be<br />

calculated with the free-convection relations. In region II bubbles begin to form<br />

on the surface of the wire and are dissipated in the liquid after breaking away<br />

from the surface. This region indicates the beginning of nucleate boiling. As<br />

the temperature excess is increased further, bubbles form more rapidly and<br />

rise to the surface of the liquid, where they are dissipated. This is indicated in<br />

region III. Eventually, bubbles are formed so rapidly that they blanket the<br />

heating surface and prevent the inflow of fresh liquid from taking their place.<br />

At this point the bubbles coalesce and form a vapor film which covers the<br />

surface. The heat must be conducted through this film before it can reach the<br />

liquid and effect the boiling process. The thermal resistance of this film causes<br />

a reduction in heat flux, and this phenomenon is illustrated in region IV, the<br />

film-boiling region. This region represents a transition from nucleate boiling to<br />

film boiling and is unstable. Stable film boiling is eventually encountered in<br />

region V. The surface temperatures required to maintain stable film boiling are<br />

high, and once this condition is attained, a significant portion of the heat lost<br />

by the surface may be the result of thermal radiation, as indicated in region VI.<br />

Figure 12.1 Heat-flux data from an electrically heated platinum wire


in<br />

An electrically heated wire is unstable at point a since a small increase<br />

∆ Tx<br />

, at this point results in a decrease in the boiling heat flux. But the wire<br />

still must dissipate the same heat flux, or its temperature will rise, resulting in<br />

operation farther down to the boiling curve. Eventually, equilibrium may be<br />

reestablished only at point b in the film-boiling region. This temperature<br />

usually exceeds the melting temperature of the wire, so that burnout results. If<br />

the electric-energy input is quickly reduced when the system attains point a, it<br />

may be possible to observe the partial nucleate boiling and unstable film<br />

region.<br />

In nucleate boiling, bubbles are created by the expansion of entrapped<br />

gas or vapor at small cavities in the surface. The bubbles grow to a certain<br />

size, depending on the surface tension at the liquid-vapor interface and the<br />

temperature and pressure. Depending on the temperature excess, the<br />

bubbles may collapse on the surface, may expand and detach from the<br />

surface to be dissipated in the body of the liquid, or at sufficiently high<br />

temperatures may rise to the surface of the liquid before being dissipated.<br />

When local boiling conditions are observed, the primary mechanism of heat<br />

transfer is thought to be the intense agitation at the heat-transfer surface,<br />

which creates the high heat-transfer rates observed in boiling. In saturated, or<br />

bulk, boiling the bubbles may break away from the surface because of the<br />

buoyancy action and move into the body of the liquid. In this case the heattransfer<br />

rate is influenced by both the agitation caused by the bubbles and the<br />

vapor transport of energy into the body of the liquid.


LECTURE NO.13<br />

HEAT EXCHANGERS- GENERAL INTRODUCTION; DOUBLE-PIPE HEAT<br />

EXCHANGER; SHELL-AND-TUBE HEAT EXCHANGER; CROSS-FLOW<br />

EXCHANGER; FOULING FACTORS, LMTD<br />

Introduction<br />

In the process industries the transfer of heat between two fluids is<br />

generally done in heat exchangers. The most common type is one in which<br />

the hot and cold fluids do not come into direct contact with each other but are<br />

separated by a tube wall or a flat or curved surface. The transfer of heat from<br />

the hot fluid to the wall or tube surface is accomplished by convection,<br />

through the tube wall or plate by conduction, and then by convection to the<br />

cold fluid.<br />

Different Types<br />

Double-pipe heat exchanger<br />

The simplest exchanger is the double-pipe or concentric pipe<br />

exchanger. This is shown in Fig. 13.1, where one fluid flows inside one pipe<br />

and the other fluid flows in the annular space between the two pipes. The<br />

fluids can be in cocurrent or countercurrent flow. The exchanger can be made<br />

from a pair of single lengths of pipe with fittings at the ends or from a number<br />

of pairs interconnected in series. This type of exchanger is useful mainly for<br />

small flow rates.<br />

Fig.13.1 Flow in a double pipe heat exchanger<br />

Shell-and-tube exchanger<br />

If larger flows are involved, a shell-and-tube exchanger is used, which<br />

is the most important type of exchanger in use in the process industries. In


these exchangers the flows are continuous. Many tubes in parallel are used,<br />

where one fluid flows inside these tubes. The tubes, arranged in a bundle, are<br />

enclosed in a single shell and the other fluid flows outside the tubes in the<br />

shell side. The simplest shell-and-tube exchanger is shown in Fig. 13.2(a) for<br />

one shell pass and one tube pass, or a 1-1 counterflow exchanger. The cold<br />

fluid enters and flows inside through all the tubes in parallel in one pass. The<br />

hot fluid enters at the .other end and flows counterflow across the outside of<br />

the tubes. Cross baffles are used so that the fluid is forced to flow<br />

perpendicular across the tube bank rather than parallel with it. The added<br />

turbulence generated by this cross-flow increases the shell-side heat-transfer<br />

coefficient.<br />

Fig.13.2. Shell-and-tube heat exchangers:<br />

(a) 1 shell pass and 1 tube pass (1-1 exchanger);<br />

(b) 1 shell pass and 2 tube passes (1-2 exchanger).<br />

Fig. 13.2(b) a 1-2 parallel-counterflow exchanger is shown. The liquid<br />

on the tube side flows in two passes as shown and the shell-side liquid flows<br />

in one pass. In the first pass of the tube side, the cold fluid is flowing


counterflow to the hot shell-side fluid; in the second pass of the tube side, the<br />

cold fluid flows in parallel (cocurrent) with the hot fluid. Another type of<br />

exchanger has two shell-side passes and four tube passes. Other<br />

combinations of number of passes are also used sometimes, with the 1-2 and<br />

2-4 types being the most common.<br />

Cross-flow exchanger<br />

When a gas such as air is being heated or cooled, a common device<br />

used is the cross-flow heat exchanger shown in Fig. 13.3 (a). One of the<br />

fluids, which is a liquid, flows inside through the tubes, and the exterior gas<br />

flows across the tube bundle by forced or sometimes natural convection. The<br />

fluid inside the tubes is considered to be unmixed, since it is confined and<br />

cannot mix with any other stream. The gas flow outside the tubes is mixed,<br />

since it can move about freely between the tubes, and there will be a<br />

tendency for the gas temperature to equalize in the direction normal to the<br />

flow. For the unmixed fluid inside the tubes, there will be a temperature<br />

gradient both parallel and normal to the direction of flow.<br />

Fig. 13.3 Cross-flow heat exchanger<br />

A second type of cross-flow heat exchanger shown in Fig. 13.3 (b) is<br />

typically used in air-conditioning and space-heating applications. In this type<br />

the gas flows across a finned-tube bundle and is unmixed, since it is confined<br />

in separate flow channels between the fins as it passes over the tubes. The<br />

fluid in the tubes is unmixed.<br />

Fouling Factors and Typical Overall U Values<br />

In actual practice, heat-transfer surfaces do not remain clean. Dirt,<br />

soot, scale, and other deposits form on one or both sides of the tubes of an


exchanger and on other heat-transfer surfaces. These deposits offer<br />

additional resistance to the flow of heat and reduce the overall heat-transfer<br />

coefficient U. In petroleum processes coke and other substances can deposit.<br />

Silting and deposits of mud and other materials can occur. Corrosion products<br />

which could constitute a serious resistance to heat transfer may form on the<br />

surfaces. Biological growth such as algae can occur with cooling water and in<br />

the biological industries.<br />

To avoid or lessen these fouling problems, chemical inhibitors are often<br />

added to minimize corrosion, salt deposition, and algae growth. Water<br />

velocities above 1 m/s are generally used to help reduce fouling. Large<br />

temperature differences may cause excessive deposition of solids on surfaces<br />

and should be avoided if possible.<br />

The effect of such deposits and fouling is usually taken care of in<br />

design by adding a term for the resistance of the fouling on the inside and<br />

outside of the tube in Equation as follows:<br />

where,<br />

U<br />

i<br />

=<br />

1 1<br />

+<br />

h h<br />

i<br />

di<br />

1<br />

( r0<br />

−ri<br />

) Ai<br />

+<br />

k A<br />

A<br />

Alm<br />

Ai<br />

+<br />

A h<br />

o<br />

o<br />

Ai<br />

+<br />

A h<br />

hdi<br />

is the fouling coefficient for the inside and<br />

hdo<br />

the fouling coefficient for the outside of the tube in W/m 2 . K.<br />

THE LOG MEAN TEMPERATURE DIFFERENCE<br />

Consider the double-pipe heat exchanger shown in Figure 13-1. The<br />

fluids may flow in either parallel flow or counter-flow, and the temperature<br />

profiles for these two cases are indicated in Figure 13-4. The heat transfer in<br />

this double-pipe arrangement can be calculated with the following equation<br />

q<br />

= UA ∆T m ----- (13.1)<br />

where<br />

U = overall heat-transfer coefficient<br />

A = surface area for heat transfer consistent with definition of U<br />

∆ T m = suitable mean temperature difference across heat<br />

exchanger<br />

The above Fig.13.4 shows that the temperature difference between the<br />

hot and cold fluids varies between inlet and outlet, and the average value has<br />

o<br />

do


to be calculated in the above equation. For the parallel- flow heat exchanger<br />

shown in Figure 13.4 (a), the heat transferred through an element of area dA<br />

may be written<br />

dq =−m c<br />

h<br />

h<br />

.<br />

dT<br />

h<br />

.<br />

= m<br />

c<br />

c<br />

c<br />

dT<br />

c<br />

------ (13.2)<br />

Fig. 13.4 Temperature Profiles for (a) parallelflow and (b) counterflow<br />

in double-pipe heat exchanger<br />

where the subscripts h and c designate the hot and cold fluids, respectively.<br />

The heat transfer could also be expressed<br />

From Equation (13.2-6)<br />

dT<br />

h<br />

dq = U ( Th −T<br />

c)<br />

dA ------------ (13.3)<br />

.<br />

= −<br />

.<br />

.<br />

m<br />

dq<br />

h<br />

c<br />

h


where<br />

Thus<br />

dT<br />

c<br />

.<br />

= −<br />

.<br />

.<br />

m<br />

dq<br />

c<br />

c<br />

c<br />

ṁ represents the mass-flow rate and c is the specific heat of the fluid.<br />

1 1<br />

dT<br />

h<br />

− dT<br />

c<br />

= d( Th<br />

−Tc<br />

) =−dq<br />

( + )<br />

. . ---------- (13.4)<br />

m c m c<br />

h<br />

h<br />

c<br />

c<br />

Solving for dq from Equation (13.3) and substituting into Equation (13.4)<br />

gives<br />

⎛ ⎞<br />

d(<br />

Th<br />

−T<br />

c)<br />

u<br />

⎜ 1 1<br />

=−<br />

⎟<br />

dA<br />

Th<br />

T ⎜<br />

+<br />

.<br />

. ⎟ ------------ (13.5)<br />

−<br />

c<br />

⎝mh<br />

ch<br />

mc<br />

cc<br />

⎠<br />

This differential equation may now be integrated between conditions 1<br />

and 2 as indicated in Figure 13.4.<br />

The result is<br />

⎛ ⎞<br />

Th<br />

2<br />

−T<br />

c2<br />

=−<br />

⎜ 1 1<br />

ln uA<br />

⎟<br />

⎜<br />

+<br />

. .<br />

−<br />

⎟ --------------- (13.5)<br />

Th<br />

1<br />

Tc<br />

1<br />

⎝mh<br />

ch<br />

mc<br />

cc<br />

⎠<br />

Returning to Equation (13.2), the products<br />

ṁ cc c<br />

and<br />

.<br />

mhc h<br />

may be<br />

expressed in terms of the total heat transfer q and the overall temperature<br />

differences of the hot and cold fluids. Thus<br />

.<br />

m<br />

h<br />

c<br />

h<br />

=<br />

T<br />

h1<br />

q<br />

−T<br />

h2<br />

.<br />

q<br />

mc<br />

cc<br />

=<br />

Tc<br />

2<br />

−T<br />

c1<br />

Substituting these relations into Equation (13.5) gives<br />

( Th<br />

2<br />

−T<br />

c2)<br />

−(<br />

Th<br />

1−T<br />

c1)<br />

q= UA<br />

------- (13.6)<br />

ln [( T −T<br />

) / ( T −T<br />

)]<br />

h2<br />

c2<br />

h1<br />

c1<br />

Comparing Equation (13.6) with Equation (13.1), the mean temperature<br />

difference is the grouping of terms in the brackets. Thus<br />

( Th<br />

2<br />

−T<br />

c2)<br />

−(<br />

Th<br />

1−T<br />

c1)<br />

∆ Tm<br />

=<br />

--------<br />

ln[( Th<br />

2<br />

−T<br />

c2) / ( Th<br />

1−T<br />

c1)]<br />

(13.7)<br />

This temperature difference is called the log mean temperature<br />

difference (LMTD). It is the temperature difference at one end of the heat<br />

exchanger less the temperature difference at the other end of the exchanger


divided by the natural logarithm of the ratio of these two temperature<br />

differences.<br />

The above derivation for LMTD involves two important assumptions:<br />

(1) the fluid specific heats do not vary with temperature, and (2) the<br />

convection heat-transfer coefficients are constant throughout the heat<br />

exchanger.<br />

If a heat exchanger other than the double-pipe type is used, the heat<br />

transfer is calculated by using a correction factor applied to the LMTD for a<br />

counterflow double-pipe arrangement with the same hot and cold fluid<br />

temperatures. The heat-transfer equation then takes the form<br />

q= UAF ∆T m -------- (13.8)<br />

values of the correction factor F according are plotted in Figures 13.5 to 13.8<br />

for several different types of heat exchangers. When a phase change is<br />

involved, as in condensation or boiling (evaporation), the fluid normally<br />

remains at essentially constant temperature and the relations are simplified.<br />

For this condition, P or R becomes zero and we obtain<br />

F = 1.0 for boiling or condensation<br />

Figure 13.5 Correction-factor plot for exchanger with one shell pass and<br />

two, four or any multiple of tube passes.


Figure 13.6 Correction-factor plot for exchanger with two shell passes and<br />

four, eight, or any multiple of tube passes.


Figure 13.7 Correction-factor plot for single-pass cross-flow<br />

exchanger, both fluids unmixed<br />

Figure 13.8 Correction-factor plot for single-pass cross-flow<br />

exchanger one fluid mixed the other unmixed.


LECTURE NO.14<br />

DESIGN PROBLEMS ON HEAT EXCHANGERS: CALCULATION OF HEAT<br />

EXCHANGER SIZE FROM KNOWN TEMPERATURES, PROBLEM ON<br />

SHELL-AND-TUBE HEAT EXCHANGER, DESIGN OF SHELL-AND-TUBE<br />

HEAT EXCHANGER<br />

Problems on Design of Heat Exchangers<br />

Problem 14.1 Calculation of heat exchanger size from known<br />

temperatures<br />

Water at the rate of 68 kg/min is heated from 35 to 75 °C by an oil<br />

having a specific heat of 1.9 kJ /kg. °C. The fluids are used in a counterflow<br />

double-pipe heat exchanger, and the oil enters the exchanger at 110 °C and<br />

leaves at 75 °C. The overall heat-transfer coefficient is 320 W/m 2 .°C.<br />

Calculate the heat exchanger area.<br />

Solution<br />

The total heat transfer is determined from the energy absorbed by the<br />

water:<br />

.<br />

q = m c ∆T<br />

= (68) (4180) (75 - 35) = 11.37 MJ /min (a)<br />

w<br />

w<br />

w<br />

= 189.5 kW<br />

Since all the fluid temperatures are known, the LMTD can be calculated by<br />

using the temperature scheme in the following Figure.<br />

(110 −75)<br />

−(75<br />

−35)<br />

∆T m<br />

=<br />

= 37 .44 C<br />

ln[( 110 −75) / (75 −35)]


Then, Since q= UA ∆T<br />

m ,<br />

5<br />

1.895 × 10<br />

A = = 15 .82 m<br />

(320 )(37 .44 )<br />

Problem 14.2 Shell-and-tube heat exchanger<br />

Instead of the double-pipe heat exchanger of problem 14.1, it is desired to<br />

use a shell-and-tube exchanger with the water making one shell pass and the<br />

oil making two tube passes. Calculate the area required for this exchanger,<br />

assuming that the overall heat-transfer coefficient remains at 320 W/m 2 . °C.<br />

Solution<br />

To solve this problem, determine a correction factor from Figure 13.5 to be<br />

used with the LMTD calculated on the basis of a counterflow exchanger. The<br />

parameters according to the nomenclature of Figure 13.5 are<br />

T 1 = 35 °C T 2 = 75 °C t 1 = 110 °C t 2 = 75 °C<br />

t<br />

P=<br />

T<br />

1<br />

T<br />

R=<br />

t<br />

−t<br />

−t<br />

1<br />

75 −110<br />

=<br />

35 −110<br />

2 1<br />

=<br />

2<br />

−T<br />

−t<br />

1<br />

35 −75<br />

=<br />

75 −110<br />

1 2<br />

=<br />

0.467<br />

1.143<br />

2<br />

Hence, the correction factor is<br />

and the heat transfer is<br />

F = 0.81<br />

q=<br />

UAF<br />

∆<br />

T m<br />

so that<br />

5<br />

1.895 × 10<br />

A =<br />

= 19 .53 m<br />

(320 )(0.81)(37 .44 )<br />

2


Problem 14.3 Design of shell-and-tube heat exchanger<br />

Water at the rate of 3.8 kg/s is heated from 38 to 55 °C in a shell-andtube<br />

heat exchanger. On the shell side one pass is used with water as the<br />

heating fluid, 1.9 kg/s entering the exchanger at 93 °C. The overall heattransfer<br />

coefficient is 1419 W / m 2 . °C, and the average water velocity in the<br />

1.9 cm diameter tubes is 0.366 m/s. Because of space limitations the tube<br />

length must not be longer than 2.5 m. Calculate the number of tube passes,<br />

the number of tubes per pass, and the length of the tubes, consistent with this<br />

restriction.<br />

Solution<br />

First assume one tube pass and check to see if it satisfies the<br />

conditions of this problem. The exit temperature of the hot water is calculated<br />

from<br />

q=<br />

m c<br />

h<br />

h<br />

.<br />

dT<br />

h<br />

.<br />

= m<br />

c<br />

c<br />

c<br />

dT<br />

(a)<br />

(3.8)(4.18 )(55 −38 )<br />

∆T h<br />

=<br />

= 34 C<br />

(1.9)(4.18 )<br />

c<br />

so<br />

T<br />

h . exit = 93 - 94 = 59 °C<br />

The total required heat transfer is obtained from Equation (a) for the cold fluid:<br />

q = (3.8)(4.18)(55 - 38) = 270 kW<br />

For a counterflow exchanger, with the required temperature<br />

LMTD<br />

=∆T<br />

m<br />

(93 −55)<br />

−(59<br />

−38)<br />

=<br />

= 28 .66<br />

ln[( 93 −55) / (59 −38)]<br />

<br />

C<br />

q=<br />

UA<br />

∆<br />

T m<br />

(b)<br />

3<br />

270 × 10<br />

A = = 6.639<br />

(1419 )(28 .66 )<br />

m<br />

2<br />

Using the average water velocity in the tubes and the flow rate, calculate the<br />

total flow area with<br />

.<br />

m c<br />

= ρAu


3.8<br />

A = = 0.0104<br />

(1000 )(0.366 )<br />

m<br />

2<br />

This area is the product of the number of tubes and the flow area per tube:<br />

0.0104<br />

=n πd<br />

4<br />

(0.0104 )(4)<br />

n= π (0.019 )<br />

2<br />

2<br />

=<br />

36 .7<br />

or n = 37 tubes. The surface area per tube per meter of length is<br />

2<br />

πd =π (0.019 ) = 0.0597 m / tube . m<br />

Recall that the total surface area required for a one-tube-pass exchanger was<br />

calculated in Equation (b) as 6.639 m 2 . Thus compute the length of tube for<br />

this type of exchanger from<br />

nπdL<br />

=6.639<br />

6.639<br />

L = = 3m<br />

(37 )(0.0597 )<br />

This length is greater than the allowable 2.438 m, so we must use<br />

more than one tube pass. When we increase the number of passes,<br />

correspondingly increase the total surface area required because of the<br />

reduction in LMTD caused by the correction factor F. Next try two tube<br />

passes. From Figure 13.5, F = 0.88, and thus<br />

A<br />

q<br />

2.70 × 10<br />

3<br />

total<br />

= =<br />

=<br />

UF ∆T<br />

m<br />

(1419 )(0.88)(28 .66 )<br />

7.54 m<br />

The number of tubes per pass is still 37 because of the velocity<br />

requirement. For the two-tube-pass exchanger the total surface area is now<br />

related to the length by<br />

so that<br />

A total<br />

=2 nπdL<br />

2<br />

7.54<br />

L = = 1. 707 m<br />

(2)(37 )(0.0597 )<br />

This length is within the 2.438 m requirement, so the final design choice is<br />

Number of tubes per pass = 37<br />

Number of passes = 2


Length of tube per pass = 1.707 m


LECTURE NO.15<br />

INTRODUCTION TO MASS TRANSFER: A SIMILARITY OF MASS, HEAT, AND<br />

MOMENTUM TRANSFER PROCESSES; FICK'S LAW FOR MOLECULAR<br />

DIFFUSION<br />

INTRODUCTION TO MASS TRANSFER<br />

A Similarity of Mass, Heat, and Momentum Transfer Processes<br />

The various separation processes have certain basic principles which<br />

can be classified into three fundamental transfer (or "transport") processes:<br />

momentum transfer, heat transfer, and mass transfer. The fundamental<br />

process of momentum transfer occurs in such operations as fluid flow,<br />

mixing, sedimentation, and filtration. Heat transfer occurs in conductive and<br />

convective transfer of heat, evaporation, distillation, and drying.<br />

The third fundamental transfer process, mass transfer, occurs in<br />

distillation, absorption, drying, liquid-liquid extraction, adsorption, ion<br />

exchange, crystallization, and membrane processes. When mass is being<br />

transferred from one distinct phase to another or through a single phase, the<br />

basic mechanisms are the same whether the phase is a gas, liquid, or solid.<br />

This was also shown for heat transfer, where the transfer of heat by<br />

conduction followed Fourier's law in a gas, solid, or liquid.<br />

General molecular transport equation. All three of the molecular transport<br />

processes - momentum, heat, and mass-are characterized by the general<br />

type of equation,<br />

rate of<br />

atransfer<br />

process<br />

driving force<br />

= --------(1)<br />

resis tan ce<br />

Molecular diffusion equations for momentum, heat, and mass transfer<br />

Newton's equation for momentum transfer for constant density can be<br />

written as follows<br />

µ d( v x<br />

ρ)<br />

τzx<br />

= −<br />

------- (2)<br />

ρ dz<br />

where<br />

τzx<br />

is momentum transferred / s.m 2 ,<br />

µ<br />

ρ is kinematic viscosity in m2 /s,


ρ and c p :<br />

z is distance in m, and<br />

v xρ is momentum/m 3 , where the momentum has units<br />

of kg.m/s.<br />

Fourier's law for heat conduction can be written as follows for constant<br />

where<br />

q d(ρc<br />

pT<br />

)<br />

z<br />

= −α<br />

--------- (3)<br />

A dz<br />

q z<br />

/ A is heat flux in W/m 2 ,<br />

α is the thermal diffusivity in m 2 /s, and<br />

ρc p<br />

T is J/m 3 .<br />

The equation for molecular diffusion of mass is Fick's law and can be<br />

written as follows for constant total concentration in a fluid:<br />

where,<br />

J<br />

*<br />

Az<br />

= −D<br />

AB<br />

dc<br />

dz<br />

A<br />

------------ (4)<br />

*<br />

J<br />

Az is the molar flux of component A in the z direction due to<br />

molecular diffusion in kg mol A/s·m 2 ,<br />

D<br />

AB the molecular diffusivity of the molecule A in B in m 2 /s,<br />

cA<br />

the concentration of A in kg mol / m 3 , and<br />

z the distance of diffusion in m.<br />

The similarity of Eqs. (2), (3), and (4) for momentum, heat, and mass<br />

transfer should be obvious. All the fluxes on the left-hand side of the three<br />

equations have as units transfer of a quantity of momentum, heat, or mass<br />

per unit time per unit area. The transport properties µ/ ρ, α and DAB<br />

have units of m 2 /s, and the concentrations are represented as momentum/m 3 ,<br />

J/m 3 , or kg mol/m 3 .<br />

Examples of Mass- Transfer Processes<br />

Mass transfer is important in many areas of science and engineering.<br />

Mass transfer occurs when a component in a mixture migrates in the same<br />

phase or from phase to phase because of a difference in concentration<br />

between two points. Many familiar phenomena involve mass transfer. Liquid<br />

in an open pail of water evaporates into still air because of the difference in<br />

all


concentration of water vapor at the water surface and the surrounding air.<br />

There is a "driving force" from the surface to the air. A piece of sugar added to<br />

a cup of coffee eventually dissolves by itself and diffuses to the surrounding<br />

solution. Many purification processes involve mass transfer. In uranium<br />

processing, a uranium salt in solution is extracted by an organic solvent.<br />

Fick's Law for Molecular Diffusion<br />

Molecular diffusion or molecular transport can be defined as the<br />

transfer or movement of individual molecules through a fluid by means of the<br />

random, individual movements of the molecules. Imagine the molecules<br />

traveling only in straight lines and changing direction by bouncing off other<br />

molecules after collision. Since the molecules travel in a random path,<br />

molecular diffusion is often called a random-walk process.<br />

Fig 15.1 Schematic diagram of molecular diffusion process<br />

In Fig. 15.1 the molecular diffusion process is shown schematically. A<br />

random path that molecule A might take in diffusing through B molecules from<br />

point (1) to (2) is shown. If there are a greater number of A molecules near<br />

point (1) than at (2), then, since molecules diffuse randomly in both directions,<br />

more A molecules will diffuse from (1) to (2) than from (2) to (1). The net<br />

diffusion of A is from high- to low-concentration regions.<br />

As another example, a drop of blue liquid dye is added to a cup of<br />

water. The dye molecules will diffuse slowly by molecular diffusion to all parts<br />

of the water. To increase this rate of mixing of the dye, the liquid can be<br />

mechanically agitated by a spoon and convective mass transfer will occur.


The two modes of heat transfer, conduction and convective heat transfer are<br />

analogous to molecular diffusion and convective mass transfer.<br />

First, we will consider the diffusion of molecules when the whole bulk<br />

fluid is not moving but is stationary. Diffusion of the molecules is due to a<br />

concentration gradient. The general Fick's law equation can be written as<br />

follows for a binary mixture of A and B:<br />

where<br />

J<br />

*<br />

Az<br />

= −cD<br />

AB<br />

dx<br />

dz<br />

A<br />

-------- (5)<br />

c is total concentration of A and B in kg mol A+B/m 3 , and<br />

xA<br />

is the mole fraction of A in the mixture of A and B.<br />

If c is constant, then since<br />

c = cx ,<br />

A<br />

A<br />

c dx = d( cx ) = dc -------------- (6)<br />

A<br />

A<br />

A<br />

Substituting into Eq. (5) obtain the following equation for constant total<br />

concentration:<br />

J<br />

*<br />

Az<br />

= −D<br />

AB<br />

dc<br />

dz<br />

A<br />

------------ (7)<br />

This equation is the one more commonly used in many molecular diffusion<br />

processes.<br />

Problem Molecular Diffusion of Helium in Nitrogen<br />

A mixture of He and N 2 gas is contained in a pipe at 298 K and 1 atm<br />

total pressure which is constant throughout. At one end of the pipe at point 1<br />

the partial pressure pA<br />

1 of He is 0.60 atm and at the other end 0.2 m (20 cm)<br />

p<br />

A2 = 0.20 atm. Calculate the flux of He at steady state if DAB<br />

of the He-N 2<br />

mixture is 0.687 X 10 -4 m 2 /s (0.687 cm 2 /s). Use SI units.<br />

Solution:<br />

Since total pressure P is constant, then c is constant, where c is as<br />

follows for a gas according to the perfect gas law:<br />

where<br />

PV = nRT ----------------------- (8)<br />

n<br />

V<br />

n is kg mol A+B,<br />

V is volume in m 3 ,<br />

P<br />

= = c<br />

-------------- (9)<br />

RT


T is temperature in K,<br />

R is 8314.3 m 3 Pa / kg mol.K or<br />

R is 82.057 x 10 -3 m 3 . atm / kg mol.K, and<br />

c is kg mol A+B/m 3 .<br />

For steady state the flux in Eq.(4) is constant. Also,<br />

constant. Rearranging Eq. (4) and integrating,<br />

DAB<br />

for a gas is<br />

J<br />

*<br />

Az<br />

z2<br />

∫<br />

z1<br />

dz<br />

= −D<br />

AB<br />

CA 2<br />

∫<br />

cA 1<br />

dc<br />

A<br />

J<br />

*<br />

Az<br />

( cA<br />

1<br />

−c<br />

A2)<br />

= DAB<br />

----------- (10)<br />

( z −z<br />

)<br />

2<br />

1<br />

Also, from the perfect gas law, pA V = nA<br />

RT , and<br />

pA<br />

nA<br />

cA = 1<br />

1<br />

=<br />

------------------- (11)<br />

RT V<br />

Substituting Eq. (11) into (10),<br />

J<br />

*<br />

Az<br />

( pA<br />

1<br />

−pA2)<br />

= DAB<br />

----------------- (12)<br />

RT ( z −z<br />

)<br />

2<br />

1<br />

This is the final equation to use, which is in a form easily used for<br />

gases. Partial pressures are p<br />

A1<br />

= 0.6 atm = 0.6 X 1.01325 X 10 5 = 6.08 X<br />

10 4 Pa and p<br />

A2<br />

using SI units,<br />

= 0.2 atm = 0.2 X 1.01325 X 10 5 = 2.027 X 10 4 Pa. Then,<br />

*<br />

J AZ<br />

(0.687<br />

=<br />

−4<br />

4<br />

× 10 )(6.08 × 10 −2.027<br />

8314 (298 )(0.20 −0)<br />

× 10<br />

4<br />

)<br />

= 5.63 X 10 -6 kg. mol A/s. m 2<br />

If pressures in atm are used with SI units,<br />

*<br />

J AZ<br />

(0.687 × 10<br />

(82.06 × 10<br />

=<br />

−3<br />

4<br />

)(0.60 −0.20)<br />

)(298 )(0.20 −0)<br />

−<br />

= 5.63 X 10 -6 kg. mol A/s. m 2<br />

Other driving forces (besides concentration differences) for diffusion<br />

also occur because of temperature, pressure, electrical potential, and other<br />

gradients.


LECTURE NO.16<br />

MOLECULAR DIFFUSION IN GASES: EQUIMOLAR COUNTER DIFFUSION<br />

IN GASES<br />

MOLECULAR DIFFUSION IN GASES: Equimolar Counter diffusion in<br />

Gases<br />

In Fig. 16.1 a diagram is given of two gases A and B at constant total<br />

pressure P in two large chambers connected by a tube where molecular<br />

diffusion at steady state is occurring. Stirring in each chamber keeps the<br />

concentrations in each chamber uniform. The partial pressure<br />

p > p and p > p . Molecules of A diffuse to the right and B to the left.<br />

A1 A2<br />

B2<br />

B1<br />

Since the total pressure P is constant throughout, the net moles of A diffusing<br />

to the right must equal the net moles of B to the left. If this is not so, the total<br />

pressure would not remain constant. This means that<br />

J<br />

= −<br />

------- (13)<br />

* *<br />

AZ<br />

J BZ<br />

FIGURE 16.1 Equimolar counter diffusion of gases A and B.<br />

The subscript z is often dropped when the direction is obvious. Writing<br />

Fick's law for B for constant c,<br />

<strong>No</strong>w since<br />

J<br />

*<br />

B<br />

B<br />

= −<br />

D<br />

BA<br />

dc<br />

dz<br />

P = p A<br />

+ p = constant, then<br />

Differentiating both sides,<br />

B<br />

c = c A<br />

+ c B ------- (15)<br />

dc<br />

A<br />

dc B<br />

------(14)<br />

= −<br />

---------(16)


Equating Eq. J<br />

J<br />

*<br />

Az<br />

= −D<br />

AB<br />

dc<br />

dz<br />

A<br />

to<br />

J<br />

*<br />

B<br />

= −<br />

D<br />

BA<br />

dc<br />

dz<br />

dc<br />

A * dc<br />

B<br />

= −DAB<br />

= − J<br />

B<br />

=−(<br />

− DBA<br />

------- (17)<br />

dz<br />

dz<br />

*<br />

A<br />

)<br />

Substituting Eq. (16) into (17) and canceling like terms,<br />

coefficient<br />

D<br />

AB<br />

= D BA<br />

This shows that for a binary gas mixture of A and B, the diffusivity<br />

DAB<br />

for A diffusing into B is the same as DBA<br />

for B diffusing into A.<br />

Problem Equimolar Counter diffusion<br />

Ammonia gas (A) is diffusing through a uniform tube 0.10 m long<br />

containing N 2 gas (B) at 1.0132 X 10 5 Pa pressure and 298 K. The diagram is<br />

similar to Fig. 16.2. At point 1, p<br />

A1<br />

= 1.013 X 10 4 Pa and at point 2, p<br />

A2<br />

=<br />

B<br />

,<br />

0.507 X 10 4 Pa. The diffusivity D<br />

AB<br />

= 0.230 X 10 -4 m 2 /s.<br />

(a) Calculate the flux<br />

*<br />

J<br />

A at steady state.<br />

(b) Repeat for<br />

*<br />

J B<br />

Solution:<br />

J<br />

*<br />

Az<br />

where<br />

= D<br />

AB<br />

( pA<br />

1<br />

−pA2)<br />

RT ( z −z<br />

)<br />

2<br />

1<br />

P = 1.0132 X 10 5 Pa,<br />

z 2 – z 1 = 0.10 m, and<br />

T = 298 K. Substituting into above equation<br />

J<br />

*<br />

A<br />

= D<br />

AB<br />

−4<br />

4<br />

( pA<br />

1−pA2)<br />

(0.23 × 10 )(1.013 × 10 −0.507<br />

=<br />

RT ( z −z<br />

) 8314 (298 )(0.10 −0)<br />

2<br />

1<br />

× 10<br />

4<br />

)<br />

= 4.70 X 10 -7 kg mol A/s.m 2<br />

Rewriting the above equation for component B for part (b) and noting that<br />

p<br />

p<br />

= P−p<br />

= 1.0132 X 10 5 - 1.013 X 10 4 = 9.119 X 10 4 Pa and<br />

B1 A1<br />

= P −p<br />

= 1.0132 X 10 5 - 0.507 X 10 4 = 9.625 X 10 4 Pa,<br />

B2 A2<br />

J<br />

*<br />

B<br />

= D<br />

AB<br />

−4<br />

4<br />

( pB<br />

1−pB2)<br />

(0.23 × 10 )(9.119 × 10 −9.625<br />

× 10<br />

=<br />

RT ( z −z<br />

) 8314 (298 )(0.10 −0)<br />

2<br />

1<br />

4<br />

)<br />

=- 4.70 X 10 -7 kg mol B/s.m 2


The negative value for<br />

*<br />

J<br />

B means the flux goes from point 2 to point 1.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!