You are on page 1of 800

This page is intentionally left blank.

Fundamentals
of
Heat and Mass Transfer

M. Thirumaleshwar
Professor and Head
Department of Mechanical Engineering
Vivekananda College of Engineering and Technology
Puttur, Karnataka
and
Former Senior Scientific Officer and Head
Cryogenics Section
Bhabha Atomic Research Centre
Trombay, Mumbai
Mathcad is a trademark of Mathsoft Engineering & Education, Inc. in the United States and/or countries.

Copyright © 2006 Dorling Kindersley (India) Pvt. Ltd


Licensees of Pearson Education in South Asia

No part of this eBook may be used or reproduced in any manner whatsoever without the publisher’s prior
written consent.

This eBook may or may not include all assets that were part of the print version. The publisher reserves the
right to remove any material present in this eBook at any time.

ISBN 9788177585193
eISBN 9788131798652

Head Office: A-8(A), Sector 62, Knowledge Boulevard, 7th Floor, NOIDA 201 309, India
Registered Office: 11 Local Shopping Centre, Panchsheel Park, New Delhi 110 017, India
This page is intentionally left blank.
To Sri Sathya Sai Baba
Preface

Heat and mass transfer is one of the core subjects in the engineering syllabus of most Indian universities.
This book is written to meet the needs of students of mechanical and automobile engineering. It should also
serve as a reference book for professional engineers working in the areas of thermal engineering, metal-
lurgy, refrigeration and air conditioning, insulation, etc.
This book is an outcome of my teaching notes. As such, the material presented here has been class
tested and continuously improved through the feedback received from the students. My aim is to present
the subject in a simple, lucid manner with an emphasis on the understanding of the physical phenomena
and principles involved. This is reinforced with numerous worked-out examples culled from the question
papers of various Indian universities. Adequate care has been taken to solve as many types of problems as
possible and to add a sufficient number of theory questions and numerical problems at the end of each
chapter.
Chapter 1 provides an introduction to the subject and presents an overview of the three principal
modes of heat transfer, namely, conduction, convection and radiation and the corresponding rate equations.
Chapter 2 explains Fourier’s Law, and, as a corollary to it, the topic of thermal conductivity is discussed.
The concept of thermal resistance is also introduced. The general differential equation of conduction is
derived in Chapter 3 and the application of different types of boundary conditions is explained. Chapter 4
deals with one-dimensional steady state conduction; here, heat transfer through three of the most popular
and practically useful geometries—namely, plane slab, cylindrical shell and spherical shell—are considered.
The effect of variable thermal conductivity and the concept of critical thickness of insulation are also dealt
with. The use of conduction shape factors in the analysis of two-dimensional conduction is discussed. Con-
duction in the aforementioned geometries with internal heat generation is considered in Chapter 5. In
Chapter 6, heat transfer from extended surfaces (that is, fins) with different boundary conditions is studied;
fin performance and application of fin theory to practical cases of temperature measurement are explained.
The topic of transient or unsteady state heat conduction is taken up in Chapter 7. Numerical methods of
solving steady- and unsteady-state conduction in one-dimensional and two-dimensional conduction are
dealt with in Chapter 8. Forced convection and natural convection are explained in Chapters 9 and 10
respectively; here mathematics is kept to a minimum and a large number of practical (empirical) correla-
tions are presented with suitable examples. In Chapter 11, boiling and condensation are explained; various
boiling regimes and the associated heat transfer correlations are presented. Chapter 12 covers design and
analysis of heat exchangers; formulas are derived for the LMTD as well as effectiveness–NTU approach.
Problems involving sizing and rating of heat exchangers are worked out. Radiation heat transfer is pre-
sented in Chapter 13; radiation properties and fundamental laws of radiation are explained first, and then,
the radiation heat transfer between surfaces and enclosures is studied. Chapter 14 is devoted to the study of
mass transfer; here, both diffusion and convective mass transfers are considered and their analogy with
diffusion and convective heat transfers is pointed out.
I have liberally drawn material from available textbooks on the subject while writing this book. Grate-
ful acknowledgements have been to all those authors.
I would also like to thank my students for persuading me to publish my notes in a book form. Without
their active involvement and encouragement, this work would not have been possible.
Sincere thanks are also due to the authorities of Fr. Conceicao Rodrigues Institute of Technology, Vashi,
and to my former colleagues in the Department of Mechanical Engineering, who encouraged me in this
venture.
My publishers, Pearson Education, have shown great patience and extended full cooperation through-
out this project. I would like to thank, in particular, Mr. Sanjay Singh, Mr. K. Srinivas, and Mr. Alok Kumar
Singh, who have consistently been encouraging and helpful. I am grateful to them.
Finally, I would like to express my sincere thanks and appreciation to my wife, Kalavathi, for her
constant encouragement and for putting up with many inconveniences during the period of writing this
book.
All comments and suggestions for improving the utility of this text are welcome.

M. THIRUMALESHWAR

viii PREFACE
About the Author

M. Thirumaleshwar graduated in mechanical engineering from Karnataka Regional Engineering College,


Surathkal, in 1965. He obtained his M.Sc. (cryogenics) from the University of Southampton, U.K., and Ph.D.
(cryogenics) from the Indian Institute of Science, Bangalore. He is a Fellow of the Institution of Engineers
(India), and is also a Foundation Fellow of the Indian Cryogenics Council.
Prof. Thirumaleshwar has worked in India and abroad in the fields of heat transfer, fluid flow, vacuum
system design and cryopumping. He has vast experience in the design and execution of large systems for
the Department of Space and Department of Atomic Energy, and also in building and motivating a scien-
tific team. He was associated with renowned universities, public sector industries, and government depart-
ments such as the Department of Science and Technology and ISRO.
Prof. Thirumaleshwar was the head of the cryogenics department at Bhabha Atomic Research Centre,
Trombay, and Centre for Advanced Technology, Indore from 1966 to 1992. He worked as Guest Collabora-
tor at the Superconducting Super Collider Laboratory of the Universities Research Association, in Dallas,
USA, from 1990 to 1993. He was also associated with the Institute of Cryogenics, Southampton, U.K. as a
Visiting Research Fellow from 1993 to 1994.
Prof. Thirumaleshwar was Professor and Head of Mechanical Engineering at Fr. Conceicao Rodrigues
Institute of Technology, Vashi, for eight years. He is currently Professor and Head of Mechanical Engineer-
ing, Vivekananda College of Engineering and Technology, Puttur, Karnataka.
Prof. Thirumaleshwar has more than 50 publications to his credit and has attended many national and
international conferences and seminars.
This page is intentionally left blank.
Brief Contents

Preface vii
Abouth the Author ix
About Mathcad® xxi
Nomenclature xxxi
1 Introduction and Basic Concepts 1
2 Fourier’s Law and Its Consequences 13
3 General Differential Equations for Heat Conduction 26
4 One-dimensional Steady State Heat Conduction without Heat Generation 47
5 One-dimensional Steady State Heat Conduction with Heat Generation 147
6 Heat Transfer from Extended Surfaces (FINS) 221
7 Transient Heat Conduction 266
8 Numerical Methods in Heat Conduction 329
9 Forced Convection 382
10 Natural (or Free) Convection 477
11 Boiling and Condensation 529
12 Heat Exchangers 578
13 Radiation 641
14 Mass Transfer 723
Appendix 761
Bibliography 763
Index 764
This page is intentionally left blank.
Contents

Preface vii
About the Author ix
About Mathcad® xxi
Nomenclature xxxi

1 Introduction and Basic Concepts 1


1.1 Introduction 1
1.2 Thermodynamics and Heat Transfer 1
1.3 Applications of Heat Transfer 1
1.4 Fundamental Laws of Heat Transfer 2
1.5 Analogies with Other Transport Processes 2
1.6 Modes of Heat Transfer 2
1.6.1 Conduction 3
1.6.2 Convection 5
1.6.3 Radiation 6
1.6.4 Combined Heat Transfer Mechanism 8
1.7 Steady and Unsteady Heat Transfer 10
1.8 Heat Transfer in Boiling and Condensation 10
1.9 Mass Transfer 11
1.10 Summary 11
Questions 11
Problems 12

2 Fourier’s Law and Its Consequences 13


2.1 Introduction 13
2.2 Fourier’s Law of Heat Conduction 13
2.3 Thermal Conductivity of Materials 14
2.3.1 Thermal Conductivity of Solids 14
2.3.2 Thermal Conductivity of Liquids 18
2.3.3 Thermal Conductivity of Gases 19
2.3.4 Insulation Systems 21
2.4 Concept of Thermal Resistance 22
2.4.1 Conduction 22
2.4.2 Convection 23
2.4.3 Radiation 23
2.4.4 Practical Applications of Thermal Resistance Concept 23
2.4.5 Limitations for the Use of Thermal Resistance Concept 24
2.5 Thermal Diffusivity (a) 24
2.6 Summary 24
Questions 25

3 General Differential Equations for Heat Conduction 26


3.1 Introduction 26
3.2 General Differential Equation for Heat Conduction in Cartesian Coordinates 26
3.3 General Differential Equation for Heat Conduction in Cylindrical Coordinates 30
3.4 General Differential Equation for Heat Conduction in Spherical Coordinates 31
3.5 Boundary and Initial Conditions 31
3.5.1 Prescribed Temperatures at the Boundaries (B.C. of the First Kind) 32
3.5.2 Prescribed Heat Flux at the Boundaries (B.C. of the Second Kind) 32
3.5.3 Convection Boundary Condition (B.C. of the Third Kind) 33
3.5.4 Interface Boundary Condition (B.C. of the Fourth Kind) 34
3.6 Summary of Basic Equations 42
3.7 Summary 44
Questions 44
Problems 45

4 One-dimensional Steady State Heat Conduction 47


4.1 Introduction 47
4.2 Plane Slab 47
4.3 Heat Transfer through Composite Slabs 50
4.4 Overall Heat Transfer Coefficient, U (W/(m2C)) 53
4.5 Thermal Contact Resistance 63
4.6 Conduction with Variable Area 66
4.7 Cylindrical Systems 74
4.8 Composite Cylinders 79
4.9 Overall Heat Transfer Coefficient for the Cylindrical System 82
4.10 Spherical Systems 91
4.11 Composite Spheres 95
4.12 Overall Heat Transfer Coefficient for the Spherical System 97
4.13 Critical Thickness of Insulation 101
4.14 Optimum (or Economic) Thickness of Insulation 109
4.15 Effect of Variable Thermal Conductivity 113
4.15.1 Plane Slab with Variable Thermal Conductivity 113
4.15.2 Hollow Cylinder with Variable Thermal Conductivity 121
4.15.3 Hollow Sphere with Variable Thermal Conductivity 129
4.16 Two-dimensional Conduction—Shape Factor 134
4.17 Summary of Basic Conduction Relations 139

xiv CONTENTS
4.18 Summary 140
Questions 141
Problems 142

5 One-dimensional Steady State Heat Conduction with Heat Generation 147


5.1 Introduction 147
5.2 Plane Slab with Uniform Internal Heat Generation 147
5.2.1 Plane Slab with Uniform Internal Heat Generation—
Both the Sides at the Same Temperature 148
5.2.2 Plane Slab with Uniform Internal Heat Generation—
Two Sides at Different Temperatures 152
5.2.3 Plane Slab with Uniform Internal Heat Generation—
One Face Perfectly Insulated 155
5.3 Cylinder with Uniform Internal Heat Generation 166
5.3.1 Solid Cylinder with Internal Heat Generation 167
5.3.2 Hollow Cylinder with Heat Generation 175
5.4 Sphere with Uniform Internal Heat Generation 197
5.4.1 Solid Sphere with Internal Heat Generation 197
5.4.2 Alternative Analysis 199
5.4.3 Analysis with Variable Thermal Conductivity 200
5.5 Applications 204
5.5.1 Dielectric Heating 204
5.5.2 Heat Transfer through a Piston Crown 207
5.5.3 Heat Transfer in Nuclear Fuel Rod (without cladding) 208
5.5.4 Heat Transfer in Nuclear Fuel Rod with Cladding 212
5.6 Summary of Basic Conduction Relations, with Heat Generation 215
5.7 Summary 216
Questions 218
Problems 219

6 Heat Transfer from Extended Surfaces (FINS) 221


6.1 Introduction 221
6.2 Fins of Uniform Cross Section (Rectangular or Circular)—
Governing Differential Equation 222
6.2.1 Infinitely Long Fin 224
6.2.2 Fin of Finite Length with Insulated End 226
6.2.3 Fin of Finite Length Losing Heat from Its End by Convection 229
6.2.4 Fin of Finite Length with Specified Temperature at Its End 231
6.2.5 Summary of Fin Formulae 233
6.3 Fins of Non-uniform Cross section 248
6.4 Performance of Fins 250
6.4.1 Fin Efficiency 250
6.4.2 Fin Effectiveness (ef) 255
6.4.3 Thermal Resistance of a Fin 256
6.4.4 Total Surface Efficiency (or, Overall Surface Efficiency, or
Area-weighted Fin Efficiency), ht 257

CONTENTS xv
6.5 Application of Fin Theory for Error Estimation in Temperature Measurement 261
6.6 Summary 263
Questions 264
Problems 264

7 Transient Heat Conduction 266


7.1 Introduction 266
7.2 Lumped System Analysis (Newtonian Heating or Cooling) 267
7.3 Criteria for Lumped System Analysis (Biot Number and Fourier Number) 269
7.4 Response Time of a Thermocouple 271
7.5 Mixed Boundary Condition 276
7.6 One-dimensional Transient Conduction in Large Plane Walls, Long
Cylinders and Spheres when Biot Number > 0.1 281
7.6.1 One Term Approximation Solutions 281
7.6.2 Heisler and Grober Charts 284
7.7 One-dimensional Transient Conduction in Semi-infinite Solids 300
7.8 Transient Heat Conduction in Multi-dimensional Systems—Product Solution 308
7.8.1 Temperature Distribution in Transient Conduction in
Multi-dimensional Systems 308
7.8.2 Heat Transfer in Transient Conduction in Multi-dimensional Systems 310
7.9 Summary of Basic Equations 317
7.10 Summary 319
Questions 320
Problems 320
Appendix 323

8 Numerical Methods in Heat Conduction 329


8.1 Introduction 329
8.2 Finite Difference Formulation from Differential Equations 330
8.3 One-dimensional, Steady State Heat Conduction in Cartesian Coordinates 331
8.4 Methods of Solving a System of Simultaneous, Algebraic Equations 344
8.5 One-dimensional, Steady State Conduction in Cylindrical Systems 348
8.6 One-dimensional, Steady State Conduction in Spherical Systems 352
8.7 Two-dimensional, Steady State Conduction in Cartesian Coordinates 356
8.8 Numerical Methods for Transient Heat Conduction 363
8.8.1 One-dimensional Transient Heat Conduction in a Plane Wall 365
8.8.2 Two-dimensional Transient Heat Conduction 372
8.9 Accuracy Considerations 378
8.10 Summary 378
Questions 379
Problems 379

9 Forced Convection 382


9.1 Introduction 382
9.2 Physical Mechanism of Forced Convection 382
9.3 Newton’s Law of Cooling and Heat Transfer Coefficient 384

xvi CONTENTS
9.4 Nusselt Number 384
9.5 Velocity Boundary Layer 385
9.6 Thermal Boundary Layer 388
9.7 Differential Equations for the Boundary Layer 390
9.7.1 Conservation of Mass—The Continuity Equation for The Boundary Layer 390
9.7.2 Conservation of Momentum Equation for The Boundary Layer 391
9.7.3 Conservation of Energy Equation for The Boundary Layer 392
9.8 Methods to Determine Convective Heat Transfer Coefficient 393
9.8.1 Dimensional Analysis 394
9.8.2 Exact Solutions of Boundary Layer Equations 402
9.8.3 Approximate Solutions of Boundary Layer Equations—
Von Karman Integral Equations 408
9.8.4 Analogy Between Momentum and Heat Transfer 429
9.9 Flow Across Cylinders, Spheres and Other Bluff Shapes and Packed Beds 431
9.9.1 Flow Across Cylinders and Spheres 432
9.9.2 Flow Across Bluff Objects 436
9.9.3 Flow Through Packed Beds 436
9.9.4 Flow Across a Bank of Tubes 440
9.10 Flow Inside Tubes 445
9.10.1 Hydrodynamic and Thermal Boundary Layers for Flow in a Tube 445
9.10.2 Velocity Profile for Fully Developed, Steady, Laminar Flow 446
9.10.3 Heat Transfer Considerations in a Pipe 448
9.10.4 Fully Developed Laminar Flow Inside Pipes of Non-circular Cross-sections 453
9.10.5 Turbulent Flow Inside Pipes 454
9.11 Summary of Basic Equations for Forced Convection 469
9.12 Summary 473
Questions 473
Problems 474

10 Natural (or Free) Convection 477


10.1 Introduction 477
10.2 Physical Mechanism of Natural Convection 477
10.3 Dimensional Analysis of Natural Convection—Grashoff Number 478
10.4 Governing Equations and Solution by Integral Method 480
10.5 Empirical Relations For Natural Convection Over Surfaces and Enclosures 484
10.5.1 Vertical Plate at Constant Temperature, Ts 484
10.5.2 Vertical Cylinders at Constant Temperature, Ts 485
10.5.3 Vertical Plate with Constant Heat Flux 485
10.5.4 Horizontal Plate at Constant Temperature, Ts 491
10.5.5 Horizontal Plate with Constant Heat Flux 493
10.5.6 Horizontal Cylinder at Constant Temperature 494
10.5.7 Free Convection from Spheres 498
10.5.8 Free Convection from Rectangular Blocks and Short Cylinders 499
10.5.9 Simplified Equations for Air 501
10.5.10 Free Convection in Enclosed Spaces 501

CONTENTS xvii
10.5.11 Free Convection in Inclined Spaces 504
10.5.12 Natural Convection Inside Spherical Cavities 505
10.5.13 Natural Convection Inside Concentric Cylinders and Spheres 506
10.5.14 Natural Convection in Turbine Rotors, Rotating Cylinders, Disks and Spheres 508
10.5.15 Natural Convection from Finned Surfaces 512
10.6 Comprehensive Correlations from Russian Literature 514
10.7 Combined Natural and Forced Convection 516
10.7 Summary of Basic Equations for Natural Convection 519
10.8 Summary 525
Questions 525
Problems 526

11 Boiling and Condensation 529


11.1 Introduction 529
11.2 Dimensionless Parameters in Boiling and Condensation 529
11.3 Boiling Heat Transfer 530
11.3.1 Boiling and Evaporation 530
11.3.2 Boiling Modes 530
11.3.3 Origin and Growth of Bubbles 530
11.3.4 Boiling Regimes and Boiling Curve 531
11.3.5 Burnout Phenomenon 532
11.3.6 Heat Transfer Correlations for Pool Boiling 533
11.3.7 Simplified Correlations for Boiling with Water 538
11.3.8 Flow Boiling 542
11.4 Condensation Heat Transfer 550
11.4.1 Introduction 550
11.4.2 Film Condensation and Flow Regimes 551
11.4.3 Nusselt’s Theory for Laminar Film Condensation on Vertical Plates 552
11.4.4 Film Condensation on Inclined Plates, Vertical Tubes,
Horizontal Tubes and Spheres, and Horizontal Tube Banks 558
11.4.5 Effect of Vapour Velocity, Nature of Condensing
Surface and Non-condensable Gases 560
11.4.6 Simplified Calculations for Water 560
11.4.7 Film Condensation inside Horizontal Tubes 573
11.4.8 Drop-wise Condensation 574
11.5 Summary 575
Questions 575
Problems 576

12 Heat Exchangers 578


12.1 Introduction 578
12.2 Types of Heat Exchangers 578
12.3 Overall Heat Transfer Coefficient 581
12.4 The LMTD Method for Heat Exchanger Analysis 589
12.4.1 Parallel Flow Heat Exchanger 589
12.4.2 Counter-flow Heat Exchanger 591
xviii CONTENTS
12.5 Correction Factors for Multi-pass and Cross-flow Heat Exchangers 600
12.6 The Effectiveness–NTU Method for Heat Exchanger Analysis 604
12.6.1 Effectiveness–NTU Relation for a Parallel-flow Heat Exchanger 605
12.6.2 Effectiveness–NTU Relation for a Counter-flow Heat Exchanger 606
12.7 The Operating-line/Equilibrium-line Method 620
12.8 Compact Heat Exchangers 622
12.9 Hydro-mechanical Design of Heat Exchangers 629
12.10 Summary 630
Questions 631
Problems 631
Appendix 633

13 Radiation 641
13.1 Introduction 641
13.2 Properties and Definitions 642
13.3 Laws of Black Body Radiation 645
13.3.1 Planck’s Law for Spectral Distribution 645
13.3.2 Wein’s Displacement Law 647
13.3.3 Stefan–Boltzmann Law 648
13.3.4 Radiation from a Wave Band 649
13.3.5 Relation between Radiation Intensity and Emissive Power 649
13.3.6 Emissivity, Real Surface and Grey Surface 651
13.3.7 Kirchhoff’s Law 653
13.4 The View Factor and Radiation Energy Exchange between Black Bodies 657
13.5 Properties of View Factor and View Factor Algebra 659
13.6 Methods of Determining View Factors 661
13.6.1 By Direct Integration 661
13.6.2 By Analytical Formulas and Graphs 663
13.6.3 By Use of View Factor Algebra 667
13.6.4 By Graphical Techniques 673
13.7 Radiation Heat Exchange between Grey Surfaces 675
13.7.1 Radiation Exchange between Small, Grey Surfaces 676
13.7.2 The Electrical Network Method 676
13.7.3 Radiation Heat Exchange in Two-zone Enclosures 679
13.7.4 Radiation Heat Exchange in Three-zone Enclosures 688
13.7.5 Radiation Heat Exchange in Four-zone Enclosures 691
13.8 Radiation Shielding 698
13.9 Radiation Error in Temperature Measurement 708
13.10 Radiation Heat Transfer Coefficient (hr) 711
13.11.1 Volumetric Absorption and Emissivity 712
13.11.2 Gaseous Emission and Absorption 712
13.11 Radiation from Gases, Vapours and Flames 712
13.12 Solar and Atmospheric Radiation 717
13.13 Summary 719
Questions 719
Problems 720
CONTENTS xix
14 Mass Transfer 723
14.1 Introduction 723
14.2 Concentrations, Velocities and Fluxes 723
14.2.1 Concentrations 724
14.2.2 Velocities 724
14.2.3 Fluxes 725
14.3 Fick’s Law of Diffusion 725
14.4 General Differential Equation for Diffusion in Stationary Media 729
14.5 Steady State Diffusion in Common Geometries 731
14.5.1 Steady State Diffusion Through a Plain Membrane 731
14.5.2 Steady State Diffusion through a Cylindrical Shell 732
14.5.3 Steady State Diffusion through a Spherical Shell 734
14.6 Equimolal Counter-diffusion in Gases 740
14.7 Steady State Uni-directional Diffusion—Diffusion of Water Vapour through Air 744
14.8 Steady-state Diffusion in Liquids 747
14.8.1 Steady-state Equimolal Counter-diffusion in Liquids 747
14.8.2 Steady-state Uni-directional Diffusion in Liquids 748
14.9 Transient Mass Diffusion in Semi-infinite, Stationary Medium 748
14.10 Transient Mass Diffusion in Common Geometries 751
14.11 Mass Transfer Coefficient 751
14.12 Convective Mass Transfer 752
14.13 Reynolds and Colburn Analogies for Mass Transfer 754
14.14 Summary 758
Questions 758
Problems 759

Appendix 761

Bibliography 763

Index 764

xx CONTENTS
About Mathcad®

All the problems in this text are solved with Mathcad® 7 Professional. Therefore, it was felt that including a short
note on use of Mathcad would be useful. However, this short note is not a tutorial on Mathcad; many specialised
books are available for that purpose (e.g., Introduction to Mathcad for Scientists and Engineers by Sol Wieder,
McGraw-Hill, 1993), in addition to the instruction manual supplied along with the software. Mathcad software
itself contains a tutorial on its use.
Purpose of this note is to make the reader comfortable with the Mathcad worksheets, using which problems
have been solved in this textbook.

What is Mathcad?
Mathcad is a very powerful and popular problem-solving tool for students of science and engineering. It turns
your computer screen into a ‘live Maths note pad’, and has a ‘free form interface’, i.e. you can add equations, text
and graphs in a single document. One great advantage of Mathcad is that equations are entered in ‘real Math’
notation (i.e., as you would enter in a note pad by hand) and not in a single line, complicated manner as in
programming languages such as FORTRAN. This makes it very easy to see if there is any mistake committed
while entering the equation. There are built-in functions and formulas and there is facility for user-defined func-
tions, too. Unlimited vectors and matrices, ability to solve problems numerically and symbolically, root finding,
quick and very easy 2-D and 3-D graphics, click selecting of Greek and other symbols from palettes are some
other highlights. All this is done without any programming, but, just with a few clicks in Windows.
Symbols in Mathcad Worksheet
Mathcad uses usual math notations. +, –, * and / have usual meaning: addition, subtraction, multiplication and
division. One advantage in Mathcad is that you can assign a value to a variable and use that variable subse-
quently throughout your worksheet. Symbol for assignment is := i.e., a colon combined with ‘equal’ sign.
Consider the following example. Let variables A, B and C be assigned values of 3, 5 and 7, respectively.
Then, the product A ´ B ´ C is obtained by simply typing A× B × C = , i.e., result is obtained by typing the desired
mathematical operation, followed by = (i.e., equals sign of maths). Some typical calculations using A, B and C are
shown below:
A := 3 B := 5 C := 7 (assigning values to variables A, B and C)
A × B × C = 105 (multiplication)
2× A + 8× B – 4× C = 18 (multiplication, addition and subtraction)
A ×B
= 2.143 (division)
C
B 2 – 4 × A× C = – 59 (exponentiation)

A3 + B3 + C 3 = 22.249 (taking square root)


F A I = 1.089
exp GH B×C JK (using ‘built-in’ exponential function)

Note that typing the equals sign ( ‘ = ’) after typing the mathematical operation, gives the final result imme-
diately and accurately.
‘What–if ’ Analysis in Mathcad
If a phenomenon depends on many variables, estimating the effect of varying one variable on the phenomenon,
while rest of the variables are held constant, is known as ‘what–if’ analysis. Such an analysis is carried out very
easily in Mathcad.
Consider, for example, the heat flow by conduction through a rod.
Heat flow rate Q, through the rod is given by:
(T1 - T2 )
Q = k×A× ,W
L
where,
k = thermal conductivity of the material, (W/(mK)
A = area of cross section of the rod, m2
(T1 – T2) = temperature difference between the two ends of rod, (where T1 > T2), and
L = Length of rod, m.
Now, suppose that we are interested to find out the value of Q for rods made up of different materials, say,
copper, aluminium and stainless steel, i.e. we would like to study the variation of Q with k, rest of the variables
being held constant. This is done very easily and quickly in Mathcad, as follows: Let T1 = 300 K, T2 = 200 K, L =
05 m, A = 0.785 ´ 10 –4 m 2.
First, define Q as a function of all variables. Then, write the data, assigning values for T1, T2, L and A. Next,
assign the first value of k (i.e. for copper), and type ‘Q(k) = ‘ (i.e. Q(k) followed by an ‘equals’ sign), and the value
of Q appears immediately. Now, to see the change in Q for the next value of k, again, assign the new value for k,
followed by ‘Q(k) = ‘, and the new value of Q appears immediately. Similarly, repeat for other values of k. Entire
worksheet of these calculations is shown below:
(T1 - T2 )
Q := k × A× W (heat transfer rate by conduction)
L
T1 := 300 k, T2 = 200 K, L = 0.5 m, A = 0.785 ´ 10 –4 m2
(T1 - T2 )
Q(k, A, T1, T2, L) := k × A× (define Q as a function of variables involved)
L
Copper: k := 407 W/(mK) (mean value of k between 300 K and 200 K)
Then, Q(k, A, T1, T2, L) := 6.39 W
Aluminium: k := 237 W/(mK) (mean value of k between 300 K and 200 K)
Then, Q(k, A, T1, T2, L) := 3.721 W
S.S (AISI 304): k := 13.75 W/(mK) (mean value of k between 300 K and 200 K)
Then, Q(K, A, T1, T2, L) := 0.216 W.
In a similar manner, by individually changing other values, namely, area of cross section (A), end tempera-
tures (T1, T2) and length (L), effect on the heat transfer rate (Q) can be studied.
Producing the Results in Tabular Form
Many times, we need the results to be presented in a tabular form. This is done very easily in Mathcad. Let us
say, we need to produce a table of values for Gaussian error function. Gaussian error function is defined as

z
follows:
2 y
erf(y) = exp (– V 2 ) dV
× (Gaussian error function...defined)
p 0
(Note: In the above definition, integral sign is obtained by clicking on the appropriate button on the calculus
palette.)

xxii ABOUT MATHCAD®


To present the values of erf(y) for values of y ranging from zero to 1, first define a range variable y, varying
from 0 to 1, with an increment of 0.05. Then, typing ‘y = ’ immediately gives the values of y one below the other;
similarly, type ‘erf(y) = ’, and values of erf(y) appear one below the other. Arrange these two sets side by side,

z
and we have the required results in a tabular form. This worksheet procedure is shown below:
2 y
erf(y) := × exp (– V 2 ) dV (Gaussian error function...defined)
p 0
y := 0, 0.05, ... , 1 (define range variable y, varying from 0 to 1 with an increment of 0.02)

y erf(y)
0 0
0 0.0564
0.1 0.1125
0.15 0.168
0.2 0.2227
0.25 0.2763
0.3 0.3286
0.35 0.3794
0.4 0.4284
0.45 0.4755
0.5 0.5205
0.55 0.5633
0.6 0.6039
0.65 0.642
0.7 0.6778
0.75 0.7112
0.8 0.7421
0.85 0.7707
0.9 0.7969
0.95 0.8209
1 0.8427

Graphing in Mathcad
Graphing in Mathcad is very easy. Let us say, we would like to produce a graph of the effectiveness (e ) of a
parallel flow heat exchanger, which is a function of number of transfer units (N) and the capacity ratio (C).
Mathematical expression for the effectiveness of parallel flow heat exchanger is:
1 - exp (- N × (1 + C ))
e=
1+ C
Then, first express e as a function of N and C; this is done in Mathcad by simply typing:
1 - exp (- N × (1 + C ))
e (N, C ) := (express e as a function of N and C)
1+ C
Let us draw a graph of variation of e with N for a value of C = 1, say:
First step is to define a ‘range variable’ N, varying from say, 0 to 6, in steps of 0.1. In Mathcad, it is written
in the form:
N := 0. 0.1, ... , 6 (define a range variable N, varying from 0 to 6 in increments of 0.1)
Then, click on the graphing palette, and select the x–y graph. A graphing area appears with two ‘place
holders’, one on the x-axis and the other on the y-axis. Fill in the x-axis place holder with N. On the y-axis place
holder, fill in e (N, 1). Click anywhere outside the graph and immediately the graph appears. If we desire to draw
in the same graph, the next curves for C = 0.8, 0.4 and zero, just type a comma after the already typed e (N, 1) and
type e (N, 0.8), e (N, 0.4), e (N, 0), and click anywhere outside the graph area, and immediately the graph is
redrawn with all the 4 curves. Further, there are simple mouse-click commands for giving titles for the graph, x-
axis and y-axis, and also for showing grid lines and legend. Logarithmic scaling also can be applied by simple
mouse-click commands. Entire worksheet is shown below:

ABOUT MATHCAD® xxiii


1 - exp (- N × (1 + C ))
e (N, C) := (express e as a function of N and C)
1+ C
N := 0, 0.1, ... , 6 (define a range variable N, varying from 0 to 6 in increments of 0.1)

Effectiveness of parallel-flow HX
1

0.8

e(N, 1)
0.6
e(N, 0.8) C=1 C = 0.4
e(N, 0.4) C = 0.8 C=0
0.4
e(N, 0)

0.2

0
0 1 2 3 4 5 6
N

FIGURE 1 Example of graphing in Mathcad

By following the procedure already explained, we can produce a table of NTU vs. e for, say, C = 1, 0.8, 0.4
and 0; this worksheet is shown below:
1 - exp (- N × (1 + C ))
e (N, C) := (express e as a function of N and C)
1+ C
N := 0, 0.1, ... , 6 (define a range variable N, varying from 0 to 6 in increments of 0.2)

N e (N, 1) e (N, 0.8) e (N, 0.4) e (N, 0)


0 0 0 0 0
0.2 0.165 0.168 0.174 0.181
0.4 0.275 0.285 0.306 0.33
0.6 0.348 0.367 0.406 0.451
0.8 0.399 0.424 0.481 0.551
1 0.432 0.464 0.538 0.632
1.2 0.455 0.491 0.581 0.699
1.4 0.47 0.511 0.614 0.753
1.6 0.48 0.524 0.638 0.798
1.8 0.486 0.534 0.657 0.835
2 0.491 0.54 0.671 0.865
2.2 0.494 0.545 0.681 0.889
2.4 0.496 0.548 0.689 0.909
2.6 0.497 0.55 0.696 0.926
2.8 0.498 0.552 0.7 0.939
3 0.499 0.553 0.704 0.95
3.2 0.499 0.554 0.706 0.959
3.4 0.499 0.554 0.708 0.967
3.6 0.5 0.555 0.71 0.973
3.8 0.5 0.555 0.711 0.978
4 0.5 0.555 0.712 0.982
4.2 0.5 0.555 0.712 0.985
Contd.
xxiv ABOUT MATHCAD®
Contd.
4.4 0.5 0.555 0.713 0.988
4.6 0.5 0.555 0.713 0.99
4.8 0.5 0.555 0.713 0.992
5 0.5 0.555 0.714 0.993
5.2 0.5 0.556 0.714 0.994
5.4 0.5 0.556 0.714 0.995
5.6 0.5 0.556 0.714 0.996
5.8 0.5 0.556 0.714 0.997
6 0.5 0.556 0.714 0.998

Solving Equation with One Variable (Root Finding)


To solve an equation with one variable, we can use the ‘root’ function:
Let us say, we would like to solve:
x + ln (x) =0. 10
This is a transcendental equation and solution requires a trial
and error procedure. We first define a function: y(x) = x + ln (x);
5
then, assume a guess value for x, and apply the root function to get
the root in a single step. Of course, guess value for x must be as- y( x )
sumed carefully, to facilitate a correct solution, since many times, 0
there is a possibility that more than one root may exist. Quickly
drawing a graph of y(x) for some values of x will help in choosing
a ‘good’ guess value. In the above case, let us draw the graph of –5
y(x) for x = 0 to 5, with an increment of 0.1, see Fig. 2. We can see 0 2 4 6
from the graph, that the curve crosses y(x) = 0 at around a x value x
of 0.5. So, let us assume the guess value of x as 0.5. Then, apply the
root function, i.e. simply type: ‘root (y(x), x) = ’, and the solution FIGURE 2 y(x) vs. x, to get approximate
solution of y(x) = 0
appears immediately. We get x = 0.567 as the solution. Entire
worksheet of this solution is shown below:
y(x) := x + ln (x) (define the function y(x))
x := 0, 0.1, ... , 5 (define a range variable x, varying from 0 to 5, in increments of 0.1)
Draw the graph, to guess the approximate root of y(x) = 0:
x := 0.5 (guess value of root, after seeing Fig. 2)
root ( y (x), x) := 0.567 (correct value of root from the root function)
Remember that numerical methods are used by Mathcad in the above solution. Calculations are terminated
by the computer when a set value of ‘tolerance’ is achieved. Built-in tolerance is 0.001. You can easily change this
value of tolerance by re-assigning its value, say: TOL := 0.0001, for example.
Solving a Set of Simultaneous Equations (Both Linear and Non-linear)
To solve a set of simultaneous equations, we use the “solve block” of Mathcad. Again, the procedure is very
simple: start with guess values for the variables involved say x1, x2 . Then, type ‘Given’ and immediately below
it, type the constraints, i.e. the set of equations to be solved. Here, while typing the constraints, take care to use
the ‘ = ’ sign, and not the assignment sign, ‘ : = ‘. Then, type ‘Find(x1, x2) = ‘, and immediately, the answer
appears, in vector form, giving values of x1, x2 , in that order. Entire worksheet of solving a set of two equations
is shown below:
Solve the following set of equations:
2
4 ×x1 – 2× x 2 = 2
x1 + x2 = 3
Start with guess values for x1 and x2, type ‘Given’, and below that write the two constraint equations, finally
type ‘Find (x1, x2) = ’, and the value of the two variables appear in that order:
x1 := 1 x2 := 1 (guess values for x1, x2)
Given
2
4 × x1 – 2× x 2 = 2
ABOUT MATHCAD® xxv
x1 + x2 = 3

Find (x1, x2) =


LM1.172OP
N1.828Q
i.e. x1 = 1.172 x2 = 1.828
Take one more example of using the solve block:
Solve the following set of equations
3× x1 – x2 + 3 × x3 = 0
– x1 + 2× x2 + x3 = 3
2× x1 – x2 – x3 = 2
Start with guess values for x1, x2, and x3, type ‘Given’, and below that write the three constraint equations,
finally, type ‘Find (x1, x2, x3) = ’, and the values of the three variables appear in that order:
x1 := 1 x2 := 1 x3 := 1 (guess values for x1, and x2 and x3)
Given
3 × x1 – x2 + 3× x3 = 0
– x1 + 2× x2 + x3 = 3
2× x1 – x2 – x3 = 2

LM 2 OP
Find (x1, x2, x3) = MM 3 PP
N- 1Q
i.e. x1 = 2 x2 = 3 x3 = – 1.
Note that to solve equation with one variable also we can use the solve block, instead of ‘root’ function.
Differentiation in Mathcad
Differentiation of a function, y(x), is done easily in Mathcad. On the calculus palette, click on the d/dx button and
a format for the first derivative appears, as shown:
d
I
dI
Now, fill in the place holders with y(x) and x as shown:
d
y(x)
dx
As an example, let us say, we would like to find the value of the first derivative of the following function at
x = 2:
y(x) =4× x 3 + 8 × x2 – 5× x + 6.
First, define the function which has to be differentiated; next, define the first derivative, y‘(x) using the
calculus palette, as explained above. Then, simply type: y‘(x) = , and the result appears immediately. See the
following worksheet:
y(x) := 4× x 3 + 8× x2 – 5× x + 6 (define a function)
d
Then, let y ¢(x) := y(x) (define the first derivative of y(x))
dx
i.e. y¢(2) := 75 (value of first derivative at x = 2)
Finding the Maxima or Minima of a Function
For this purpose, we equate the first derivative to zero, get the value of x and then substitute in the function, i.e.
we set y ‘(x) = 0, and get the value of x. To do this, we can easily use the root function. To check if the value of x
so obtained gives maximum or minimum, we have to determine if the value of second derivative is positive (for
a minimum) or negative (for a maximum). The second derivative is defined simply as the derivative of the first
derivative. See the procedure below:
y(x) := 4× x 3 + 8× x 2 – 5× x + 6 (define a function)

xxvi ABOUT MATHCAD®


d
Then, let y ¢(x) := y (x) (define the first derivative of y(x))
dx
i.e. x := 1 (trial value of x)
root (y ¢(x), x) = 0.261 (use the root function to get the value of x at which y¢(x) = 0)
d
y ¢¢(x) := y ¢(x) (define the second derivative)
dx
Therefore, y ¢¢(0.261) = 22.264 (this is > 0; therefore, y(x) is a minimum)
and, y(0.261) = 5.311 (value of y(x) at x = 0.261)
Confirm it by drawing the graph of y(x) vs. x:
x = – 2, – 1.99, ... , 3 200
In general, a minimum may occur not only at x = 0.261, but,
there may be other values of x also at which the function passes
through a minimum. The value of x given by Mathcad is the one
nearest to the trial value of x chosen for the ‘root’ function. There- y(x) 100
fore, in such problems, it is always advisable to draw the graph first
and examine if there is more than one minimum (or maximum).
Integration
0
Integration between given limits is an operation required very often –2 0 2 4
while solving heat transfer problems. Again, this is done simply by x
clicking the appropriate button from the calculus palette, and a for-

z
mat for integration appears as shown:
I
I dI
I

Now, just fill in the place holders, type ‘ = ‘ (i.e. ‘equals’ sign) and the result appears immediately.
For example, let us say, we would like to integrate the function y(x) = 1+ sin (x) between the limits x = 0 and
x = p /2. We proceed as shown in the following worksheet:
y(x) := 1 + sin (x) (define the function)
Click the appropriate button on the calculus palette, fill in the place holders, and type ‘= ’, and the result
appears immediately:

z
p
2
y (x) dx = 2.571.
0
Take one more example of integration within given limits:
Integrate the following function between the limits x = 2 and x = 5:
y(x) = x3 + 4 × x 2 – 3 × x + 6
We proceed as shown in the following worksheet:
y(x) = x 3 + 4× x 2 – 3× x + 6 (define the function)
Click the appropriate button on the calculus palette, fill in the place holders, and type ‘= ’, and the result

z
appears immediately:
5
y(x) dx = 294.75.
2

Programming in Mathcad
Mathcad-7 Professional version has programming capability, too. Just as in the case of other programming lan-
guages, there is facility for conditional branching, looping constructs, error handling, using other programs as
sub-routines, etc.
A Mathcad program is a special kind of expression, which returns a value—a scalar, vector, array, nested
array or string. An ‘expression’ in Mathcad is only a simple statement, whereas a ‘program’ can consist of as
many statements as required to compute the answer.

ABOUT MATHCAD® xxvii


Programs are written using the ‘programming palette’. Programming palette has only 10 buttons: add line,
¬ (assignment), if, while, for, break, otherwise, return, on error, continue. However, with its very wide math-
ematical and graphing functionality, coupled with programming capability and the convenience of Windows
platform, Mathcad is a very powerful and versatile tool for engineering and scientific calculations.
It is impossible to illustrate all the programming capabilities of Mathcad in this short introduction.
However, we shall give only two small examples:
Consider the problem of calculating the friction factor for flow of fluid in a smooth tube. Friction factor
depends on the Reynolds number (ReD) based on tube diameter (D). If the Reynolds number is less than 2300, the
flow is termed ‘laminar’ and the friction factor is given by: f = (64/ ReD); if ReD > 2300, flow is turbulent, and the
expression for friction factor is: f = 0.184. (ReD)– 0.2.
We would like to write a Mathcad program to return the value of f for any input value of ReD i.e. f (ReD).
Worksheet for this program is developed as explained below:
We start with the definition of friction factor as a function of ReD on the LHS; then, click on ‘add line’ button
in the programming palette. Two place holders appear as shown:
I
ffactor(ReD) :=
I

Now, position the cursor in the top place holder and click on the ‘if’ button in the programming palette. We
see:
I if I
ffactor(ReD) :=
I

Fill in the place holders on the left and right of ‘if’ by 64/ReD and ReD < 2300, respectively as shown:
64
if ReD < 2300
ffactor(ReD) := ReD
I

Next, position the cursor on the bottom place holder and click on the ‘otherwise’ button in programming
palette. We see:
64
if ReD < 2300
ffactor(ReD) := ReD
I otherwise
Now, fill in the remaining place holder by 0.184.ReD–0.2. We get, finally:
64
if ReD < 2300
ffactor(ReD) := ReD
- 0.2
0.184 × ReD otherwise
Entire worksheet is given below:
Program to compute the friction factor (ffactor) for a smooth tube as a function of Reynolds number
(ReD):
64
if ReD < 2300
ffactor(ReD) := ReD
- 0.2
0.184 × ReD otherwise
Now, for any value of ReD, we can get the value of f by simply typing ffactor(ReD) =.
For example,
ReD = 2000 ffactor(ReD) = 0.032 (friction factor when ReD = 2000)
ReD = 4000 ffactor(4000) = 0.035 (friction factor when ReD = 4000)
ReD = (2 ´ 106) ffactor(2 ´ 106) = 0.01 (friction factor when ReD = 2 ´ 10 6)
Consider one more example of programming in Mathcad:
This program to find the sum of the series, S = 1 + 2 + 3 + 4 +…..+ N, illustrates the use of ‘for’ loop:

xxviii ABOUT MATHCAD®


We denote the sum of the N terms as Sum (N). Type Sum (N) on the LHS and put the definition sign. We see
the following:
Sum(N) := I
Position the cursor on the place holder and click on ‘add line’ button in the programming palette. We get:
I
Sum (N) :=
I

In the first place holder, type S ¬ 0; this initialises the Sum, S. Note that the left arrow ( ¬ ) denotes an
assignment symbol inside the program, and it must be typed by clicking the left arrow button in the program-
ming palette. Positioning the cursor in the other place holder, click on ‘for’ button in the programming palette.
Then, we see:

S¬0
Sum (N) := for I Î I
I

After ‘for’, now, fill in i (the counter which varies through the ‘for’ loop) and after the e sign, fill in the range
1, ... , N. We see:

S¬0
Sum (N) := for i Î1, ..., N
I

In the last place holder, type the command S ¬ S + i; this command updates the value of Sum after each
pass through the loop. Loop will stop when the counter ‘i ’ reaches the value of N, passed in the input. We get:

S¬0
Sum (N) := for i Î1, ..., N
S¬S+i
Next, select the last line and click on ‘add line’ button in the programming palette. We see:

S¬0
for i Î1, ... , N
Sum (N) :=
S¬S+i
I

Now, fill in the place holder by S; it means that when i reaches the value of N, the loop will stop and the last
value of S will be returned as Sum (N):

S¬0
for i Î1, ..., N
Sum (N) :=
S¬S+i
S
Worksheet containing the entire program described above is shown below:
Problem. Write a Mathcad program to find the sum of the series: S = 1 + 2 + 3 + ... + N

S¬0
for i Î1, ... , N
Sum (N) :=
S¬S+i
S
Examples:
Sum (2) = 3 (sum of first two terms, i.e. S = 1 + 2)

ABOUT MATHCAD® xxix


Sum (3) = 6 (sum of first three terms, i.e. S = 1 + 2 + 3)
Sum (5) = 15 (sum of first five terms, i.e. S = 1 + 2 + 3 + 4 + 5)
Sum (10) = 55 (sum of first ten terms)
Sum (50) = 1.275 ´ 10 3 (sum of first fifty terms)
Sum (100) = 5.05 ´ 10 3 (sum of first hundred terms.)
Note how brief and succinct is the program.
Two programs given above are simple enough; but, they illustrate the way the program is built-up in Mathcad. For
longer programs, more lines are added simply by clicking on ‘add line’ button, as and when required. Few more simple
programs are given in the text.
Mathcad has several features, such as sequences, series, sums, products, factorials, derivatives and integrals, vectors
and matrices, capability to draw x–y, bar, scatter, polar, surface and contour plots, etc., all by just a few clicks on the
mouse. Only the most essential features, used in this text, are described above.

xxx ABOUT MATHCAD®


Nomenclature

The following is not a complete list of symbols used. In fact, symbols are explained whenever they appear in the text.

Notation Meaning
A, Ac Area, Area of cross section
C Specific heat, mass concentration
Ch, Cc Capacity rates of hot and cold fluids in a heat exchanger
cp, cv Specific heat at constant pressure and constant volume
D, d Diameter
D Diffusion coefficient
E Energy or emissive power
Eb Emissive power of black body
Ebl Monochromatic emissive power
F Force
F12 View factor from surface-1 to surface-2
f Friction coefficient or function
g(or a) Acceleration due to gravity (or, acceleration)
h Heat transfer coefficient or head of fluid
hfg Latent heat
I Intensity of radiation
J Joule’s constant, or radiosity
k Thermal conductivity
x, L Distance, length
m Mass
N RPM
n Number
P, p Pressure, perimeter
Q Quantity of heat
q Heat flux
qg Heat generation rate per unit volume
R Thermal resistance
Ru Universal gas constant
R, r Radius
S Distance, Conduction shape factor
T, t Temperature (K or deg.C)
t Time
U, u Overall heat transfer coefficient or velocity
V Volume, velocity
Contd.
Contd.

u, v, w Velocities
w, W Weight, width
x, y, z Cartesian coordinates, velocities
a Thermal diffusivity or absorptivity
b Coefficient of volumetric expansion
d Boundary layer thickness
e Effectiveness of heat exchanger, or emissivity of surface
f Angle
g Ratio of specific heats
h Efficiency
l Wavelength or coefficient used in approximate solution of transient con-
duction problems
m Dynamic viscosity
n Kinematic viscosity
p Mathematical constant (= 3.14), or terms in Buckingham -p theorem
q Temperature, Angle
r Mass density
s Stefan–Boltzann constant, or surface tension
t Shear stress, or transmissivity, or time
w Angular velocity
AMTD Arithmetic Mean Temperature Difference
LMTD Logarithmic Mean Temperature Difference
NTU Number of Transfer Units

Dimensionless parameters

Name Symbol Formula

h ×L
Biot number Bi ... k for solid
k
u2
Eckert number Ec
c p × DT

DT
Euler number Eu
1
× r ×u 2
2

a ×t
Fourier number Fo
L2
g × b × D T × L3
Grashoff number Gr
n2
Modified Grashoff number Grm or Gr‘ Gr × Nu
Graetz number Gz Re.Pr.D / L
2
Colburn j factor j St × Pr 3
c p × DT
Jakob number Ja
hf g

u
Mach number M
g ×R × T

Contd.

xxxii NOMENCLATURE
Contd.

h ×L
Nusselt number Nu ...k for fluid
k
u ×L
Peclet number Pe = Re × Pr
a
cp ×m
Prandtl number Pr
k
Rayleigh number Ra Gr . P r
r ×u ×L
Reynolds number Re
m

h Nu
Stanton number St =
r ×u × c p Re × P r
Weber number We
r ×u 2 ×L
s
Mass transfer parameters:
a
Lewis number Le
D
hm ×L
Sherwood number Sh
D
n
Schmidt number Sc
D

NOMENCLATURE xxxiii
CHAPTER

1
Introduction and Basic
Concepts

1.1 Introduction
We are embarking on a study of heat and mass transfer. Heat is defined as energy in transit. Heat itself cannot be
seen, but its effect can be felt and measured as a property called temperature. Heat transfer occurs whenever two
bodies at different temperatures are brought in contact with each other or, whenever there is a temperature gradient
within a body. Science of heat transfer involves the study of principles that govern and the methods that determine
the rate of heat transfer. Often, we are also interested in the spatial temperature distribution within a body causing
that heat transfer.
In this chapter, we will primarily give an introduction to the three different modes of heat transfer, viz., con-
duction, convection and radiation and the corresponding rate equations that govern these processes. We will
mention about the fundamental laws generally applied in heat transfer analysis, and analogies with other transport
processes. We will also indicate a few areas where the science of heat transfer and mass transfer finds its
applications.

1.2 Thermodynamics and Heat Transfer


In the course on thermodynamics you have studied the interaction of heat and work and the laws of thermodynam-
ics. First law deals with energy balances and leads to the concept of enthalpy, whereas second law deals with
availability balances and determines the direction in which heat energy will flow and leads to the concept of en-
tropy. In fact, it is the second law which says that heat flows from a location of high temperature to a location of low
temperature. Then, you may be wondering as to what is the need for a separate science of heat transfer. The answer
is, thermodynamics deals with equilibrium (quasistatic) processes; total heat transferred from one equilibrium state
to another equilibrium state can be easily calculated by the laws of thermodynamics. However, the rate of heat
transfer and the temperature variation with time and position cannot be calculated by these laws alone and to do
this, we need the laws of heat transfer. As already stated, heat transfer requires a temperature gradient, i.e.
essentially we need to know the temperature distribution within a body and the laws connecting the rate of heat
transfer with this temperature gradient.
In the chapters to follow, we will study these laws as well as their applications.

1.3 Applications of Heat Transfer


Heat transfer is an important branch of thermal science which has applications in diverse fields of engineering.
(a) Mechanical engineering In boilers, heat exchangers, turbine systems, internal combustion engines etc.
(b) Metallurgical engineering In furnaces, heat treatment of components etc.
(c) Electrical engineering Cooling systems for electric motors, generators, transformers etc.
(d) Chemical engineering In process equipments used in refineries, chemical plants etc.
(e) Nuclear engineering In removal of heat generated by nuclear fission using liquid metal coolants, design of
nuclear fuel rods against possible burnout etc.
(f) Aerospace engineering & space technology In the design of aircraft systems and components, rockets, mis-
siles etc.
(g) Cryogenic engineering In the production, storage, transportation and utilisation of cryogenic liquids (at
very low temperatures ranging from 100 K to 4 K or even lower) for various industrial, research and
defence applications.
(h) Civil engineering In the design of suspension bridges, railway tracks, airconditioning and insulation of
buildings etc.
There are numerous other applications where principles and methods of heat transfer are widely applied and
directly affect our lives.

1.4 Fundamental Laws of Heat Transfer


Science of heat transfer, of course, operates within the limits of the laws of thermodynamics. Additionally, subsid-
iary laws relating to fluid flow and rate equations for different modes of heat transfer are also required for a
complete solution.
Fundamental laws governing heat transfer are enumerated below:
(i) First law of thermodynamics—gives conservation of energy,
(ii) Second law of thermodynamics—gives direction of heat flow,
(iii) Equation of continuity—gives conservation of mass,
(iv) Equation of flow—Newton’s Second law of motion—Navier Stokes’ equation,
(v) Rate equations governing the three modes of heat transfer,
(a) Conduction—Fourier’s law of conduction
(b) Convection—Newton’s law of cooling
(c) Radiation—Stefan Boltzmann’s law
(vi) Empirical relations for fluid properties such as specific heat, thermal conductivity, viscosity etc.,
(vii) Equation of state for the fluid.
As we go along, we will have occasion to see how these laws are applied to the problems at hand. Of course,
which laws have to be applied in a given situation depends on the specific problem; aim of this course is specifically
to give such an insight to the student.

1.5 Analogies with Other Transport Processes


We will take a little diversion from our main stream of thought here. It is apparent that there is so much symmetry
in nature and natural processes that occur, that sometimes, by observing one process, one may be able to predict the
outcome of another similar process. For example, consider the three important processes of transport of energy
(heat), transport of momentum and transport of mass. We know that in nature, flow occurs spontaneously from a
higher potential to a lower potential: electricity flows from a higher voltage potential to a lower voltage potential
and water flows from a higher datum level (pressure) to a lower datum level. Similarly, for the aforementioned
three transport processes, we can observe following:
(i) Transport of heat energy—occurs from a higher temperature level to a lower temperature level.
(ii) Transport of momentum—occurs from a higher velocity level to a lower velocity level.
(iii) Transport of mass—occurs from a higher concentration level to a lower concentration level.
In other words, we can say that the driving potential for heat transfer is the temperature gradient, driving
potential for momentum transfer is the velocity gradient and the driving potential for mass transfer is the concen-
tration gradient. Therefore, we can feel that the governing equations for these processes must have some similarity.
In fact, it will be seen later, when we study conduction, convection and mass transfer, that such a similarity does
exist in the governing equations for these processes. Therefore, knowing the solutions for one transport process, we
will be able to predict the solutions for another transport process by analogy.

1.6 Modes of Heat Transfer


Generally, for convenience of analysis, we consider heat transfer in three different modes: conduction, convection
and radiation.

2 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1.6.1 Conduction
Conduction is a microscopic phenomenon. Here, more energetic particles of a substance transfer their energy to
their less energetic neighbours. Conduction can occur in a solid, liquid or gas. In a solid, transfer of energy occurs by
lattice vibrations and/or free electrons. In a liquid or gas, the transfer of energy occurs by collisions and diffusion of
molecules. It should be noted that in solids energy transfer occurs only by conduction whereas in liquids and gases
other modes of energy transfer are also possible. Consider, for example, a copper rod, well insulated on its surface,
held with its one end inside a furnace at a high temperature; its other end is open to atmosphere at room
temperature. It can easily be observed that after some time, the end open to atmosphere will get warmer and reach
a temperature higher than ambient. We say that heat is transferred along the rod by conduction.
Governing rate equation for conduction is given by Fourier’s law. This is an empirical law based on experi-
mental observations of Biot, but formulated by the French mathematical physicist, Fourier in 1822. It states that the
rate of heat flow by conduction in a given direction is proportional to the area normal to the direction of heat flow and to the
gradient of temperature in that direction.
Referring to Fig. 1.1, for heat flow in the X-direction, Fourier’s law k
states,
dT
Qx = –kA ,W ...(1.1)
dx T1 Slope = dT/dx
i.e. Qx = –kA (T2 – T1)/L (For a plane slab, in
steady state) Q
= kA (T1 – T2)/L
dT
or, qx = –k , W/m 2 ...(1.2) T2
dx
i.e. qx = k (T1 – T2)/L
Here, Q x is the rate of heat transfer in the positive X-direction, A is
the area normal to direction of heat flow, dT/dx is the temperature
gradient in the X-direction and k is a proportionality constant. Note that L
if the heat flow has to be in the positive X-direction, the temperature
must go on decreasing in the X direction, i.e. the temperature will X
decrease as X increases which means that the temperature gradient is
negative; therefore, we insert a negative sign in Eqs. 1.1 and 1.2 to make FIGURE 1.1 Fourier’s law
the heat flow positive in the positive X-direction. qx in Eq. 1.2 is known as
heat flux, which is nothing but the heat flow rate per unit area.
The proportionality constant k in Eqs. 1.1 and 1.2 is known as “thermal conductivity”, a property dependent on
the material. Thermal conductivity of materials varies over a wide range, by about 4 to 5 orders of magnitude. For
example, at 20°C, thermal conductivity of air is 0.022 W/(mC), of water 0.51 W/(mC), that of asbestos 0.095 W/
(mC), that of stainless steel 19.3 W/(mC) and that of pure silver, about 407 W/(mC). Thermal conductivity, essen-
tially depends upon the material structure (i.e. crystalline or amorphous), density of material, moisture content,
pressure and temperature of operation.
We will study more about thermal conductivity in the next chapter.
Example 1.1. Asbestos layer of 10 mm thickness (k = 0.116 W/mK) is used as insulation over a boiler wall. Consider an area
of 0.5 m2 and find out the rate of heat flow as well as the heat flux over this area if the temperatures on either side of the
insulation are 300°C and 30°C.
Solution. See Fig. Example 1.1(a). Here, dT/dx is linear i.e. the temperature gradient is linear.
Heat flux q is determined from Eq. 1.2
dT
qx = –k , m2
dx
= – 0.116 ´ (30 – 300)/0.01
= 3132 W/m2
Rate of heat flow Q is given by,
Q = Heat flux (q) ´ Area
= 3132 ´ 0.5
= 1566 W.

INTRODUCTION AND BASIC CONCEPTS 3


k Now, let us demonstrate working out this problem in Mathcad. A sample
worksheet from Mathcad calculation is shown below. Explanatory notes are in-
cluded; please read them carefully.
First data values are entered. Here, you assign the values of variables, L, k, A,
300°C etc. as shown below. Note the assignment symbol, i.e., :=
T(x) Data:
Q L := 0.01 m k := 0.116 W/(mK) A := 0.5 m2
T1 := 300°C T2 := 30°C
Solution:
30°C For asbestos:
(T1 - T2 )
q := k × i.e., q = 3.132 ´ 10 3 W/m2
L
Q := q ×A i.e., Q= 1.566 ´ 10 3 W.
Note: This was a simple problem. Now, suppose, you want to see the effect of
0.01 m using another insulation, say, glasswool with k = 0.038 W/(mC). Then, re-enter only
this data for k and copy the equations again; immediately, you will see that the
X answers are updated with this new value of k. (There is no need to re-enter the other
FIGURE Example 1.1(a) data again). See below,

For glasswool:
k := 0.039 W/(mK)
k value for glasswool is re-entered. In subsequent calculations, this value of k will be used. Rest of the data values will be as
entered earlier.
(T1 - T2 )
q := k× i.e., q = 1.053 ´ 10 3 W/m 2
L
Q := q×A i.e., Q = 526.5 W.
Here is another example of the versatality of Mathcad: Suppose, we want to use asbestos as the insulation, and one surface
maintained at 30°C, but the other surface temperature is varied from 300°C to 350°C, say in steps of 10°C; we wish to
compute the corresponding values for q and Q.
For asbestos:
k := 0.116 W/(mC)
re-enter this value to update. Mathcad uses the latest value of variable entered before doing calculation.
T 1 := 300, 310, ..., 350 (Define T1 as a range variable from 300 to 350 in
increments of 10)
(T1 - T2 )
q (T1) := (Define q as a function of T1 only, since we are keeping
L other parameters constant)

Q (T 1) := q (T1)×A (Define Q as a function of T1, since we are keeping other


parameters as constant)
Now, simply type T = . Immediately, a table is produced showing different values of T1 as shown below. Next, enter
q (T 1) and Q(T1), each followed by ' = ' mark, and tables showing the computed values of q and Q at respective values of T1
are produced immediately. Arrange them side by side for easy readability:

T1 q (T 1) Q (T 1)
4
300 2.7 ´ 10 1.35 ´ 10 4
310 2.8 ´ 104 1.4 ´ 104
320 2.9 ´ 104 1.45 ´ 104
330 3 ´ 104 1.5 ´ 104
340 3.1 ´ 104 1.55 ´ 104
350 3.2 ´ 104 1.6 ´ 104

4 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Graphing. Graphing—the function is very easy in Mathcad. In the graph pallete, press the x–y graph pallete; a blank x–y
graph is presented with place holders for x-axis and y-axis variables; just fill them up and the graph is drawn automatically.
You can easily add title, grid lines, legend, x-axis and y-axis titles, etc. (See Fig. Example 1.1(b)).

4 Variation of q and Q with T1


4 ´ 10

2
4 Note: q(W/m ) and
3 ´ 10 Q(W) plotted against
T1(C) are straight
q ( T 1)
lines, as expected.
Q ( T 1)
4
2 ´ 10

4
1 ´ 10
300 320 340 360
T1
heat flux
heat transfer rate

FIGURE Example 1.1(b)

1.6.2 Convection
Convection is a macroscopic phenomenon. It occurs only in fluids. When a fluid flows over a body that is at differ-
ent temperature than itself, heat transfer occurs by convection; the direction of heat transfer, of course, depends on
the relative magnitude of the temperatures of the fluid and the surface. In addition, if two fluids at different tem-
peratures are mixed together, heat transfer occurs by convection. Boiling and condensation also involve convective
heat transfer, but with phase change.
In convection, the fluid particles themselves move and thus carry energy from a high temperature level to a
low temperature level. As an example, consider a hot copper plate held hanging in air. The air layer in the
immediate vicinity of the plate gets heated up, its density decreases (since the room air pressure is constant) and
therefore rises up, thus carrying away heat with it; the cooler air takes the place of the displaced hot air, gets heated,
rises up and this process continues till the plate attains equilibrium with room temperature.
In case of convection, fluid motion may occur by density differences caused by temperature differences, as
mentioned in the above example. Such a case is known as natural (or free) convection. When fluid motion is caused
by an external agency such as a pump, fan or atmospheric winds, that case is known as forced convection. One can
intuitively feel that heat transfer in the case of forced convection is higher as compared to free convection.
In the case of convective heat transfer, determining the amount of
heat transfer analytically is a little complicated since fluid motion is
Ts
involved and the equations of fluid flow have to be coupled to the
equation of energy.
Governing rate equation for convection is given by Newton’s Law
of Cooling. Fig. 1.2 shows a situation of natural convection.
Here, a flat plate at a surface temperature of Ts is held vertically; the
ambient is at a temperature T f . Then, the rate of heat transfer is given by h, Tf
Newton’s law as follows,
Q = hA(Ts – Tf), W ...(1.3)
or, q = h(Ts – Tf ), W/m 2 ...(1.4)
where, Ts is the surface temperature (°C) , Tf is the fluid temperature
(°C), A is the surface area (m2) exposed to the fluid, Q is the rate of heat FIGURE 1.2 Newton’s Law of Cooling
transfer (W) from the surface to the fluid, q is the heat flux (W/m2) and for convection
h is coefficient of heat transfer for convection.
INTRODUCTION AND BASIC CONCEPTS 5
Of course, if the fluid temperature is higher than the surface temperature, heat will be transferred from the
fluid to the surface and in that case, the heat transfer rate is given by,
Q = hA(T f – T s ), W
Here, a few words about h are appropriate. The convective heat transfer coefficient, h, is not a property of the
surface material nor that of the fluid. Instead, h is a complicated function of the type of flow (i.e. whether the flow is
laminar or turbulent), geometry and orientation of the body, fluid properties (such as specific heat, thermal
conductivity, viscosity), average temperature and the position along the surface of the body. In normal practice,
even though h varies along the length of the body, it is customary to take an average (or mean) value of h over the
entire body, h m, and use it in Eq. 1.3 to calculate the total heat transfer rate.
Note that Eq. 1.3 does not give any insight into the nature of h and should therefore be considered only as a
definition of h.
Typical values of convective heat transfer coefficient, h, for a few situations are given in Table 1.1.
1.6.3 Radiation
All bodies above the temperature of 0 K emit radiation. There are two theories of radiation, i.e. (i) Maxwell’s wave
theory which states that radiation is emitted as electromagnetic waves, and (ii) Planck’s corpuscular theory which
states that radiation is emitted in discrete quanta or packets of energy. In practice both these theories are used.
Electromagnetic waves travel at the speed of light and generally obey all laws of light. Radiation is emitted over all
the wavelengths. However, the radiation emitted over the wavelength range of 0.1 mm to 100 mm is known as
thermal radiation since radiation in this particular range gets converted to heat when absorbed by a body. Higher the
temperature, smaller the wavelength of radiation emitted and deeper its penetration through a body.

TABLE 1.1 Typical values of convective heat transfer coefficient, h

Situation h, W/(m2K)
Air (1 bar, free convection) 6 – 30
Air (1 bar, forced convection) 10 – 200
Water (free convection) 500 – 1000
Water (forced convection) 600 – 8000
Boiling water 2500 – 100000
Condensing steam 2500 – 70000

Thermal radiation is a volume phenomenon, i.e. the radiation is the result of excitation of all the particles of a
body. However, the radiation travels to the surface and is then emitted from the surface.
When radiation falls on a body, it may be attenuated within a short distance from the surface (of the order of a
few angstroms), or get reflected from the surface or just pass through the body. One or more of these phenomena
may occur simultaneously. In vacuum, radiation propagates without any attenuation. For practical purposes,
atmospheric air is considered to be transparent to thermal radiation.
Governing rate equation for emission of radiation flux from a body is given by the Stefan–Boltzmann law:
E b = sT 4 , W/m 2 ...(1.5)
where, s = Stefan–Boltzmann constant
= 5.6697 ´ 10 –8 W/(m2 K4)
T = temperature in Kelvin
Eb = black body emissive power.
Note that Eq. 1.5 defines the emissive power of a black body, i.e. an ideal emitter.
Radiation flux emitted by a real body is less than that of the black body and is given by,
E = e E b = e s T 4 W/m2 ...(1.6)
where, e is known as Emissivity, lies between zero and unity. Emissivity depends on the surface material,
surface finish, temperature and the wavelength of radiation.
By definition, a black body is also an ideal absorber, i.e. it absorbs all the radiation falling on it. However, a real
body absorbs only a part of the radiation falling on it.

6 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Hence, we can write,
q abs = a q inc W/m 2 ...(1.7)
where, a is absorptivity and lies between zero and unity.
In general, a and e are different, but for practical purposes, we assume them to be equal to each other.
Radiation exchange between bodies. Consider the case of two surfaces at different temperatures T1 and T2 facing
each other, separated by a medium like air, which is transparent to radiation. Surface 1 emits radiation at T1 and
surface 2 emits radiation at T2; there will be a net radiation heat transfer between the two surfaces depending on the
magnitudes of T1 and T2, emissivities of the surfaces, relative orientation of the surfaces and the distance between
them. In general, this calculation is complicated.
However, consider the special case of a small body (A1, T1, e 1) surrounded by a large enclosure (A 2, T 2) as
shown in Fig. 1.3.

4 A2, T2
q2 = sT2

4
q1 = e1 sT1 e1, A1, T1

FIGURE 1.3 Radiation from a small surface to a very large surrounding

Let A 1 << A 2 and T 1 > T 2. Also, the large enclosure can be approximated as a black body with respect to the
small surface A 1. Then,
Radiation energy emitted by surface A 1 = A 1 e 1 s T 14
Radiation energy flux emitted by black surface A 2 = s T 24
Out of this energy falling on it, the energy absorbed by surface A 1 = a 1 A 1 s T 24
Therefore, net radiation energy leaving the surface A 1 is given by
Q 1 = A 1 e 1 s T 14 – A 1 a 1 s T 24
For a 1 = e1, we get
Q1 = A1 e1 s (T 14 – T 24), W …(1.8)
[Note that T 1 and T 2 must be expressed in Kelvin.]
If Q 1 is positive, heat is lost from the surface and if Q 1 is negative, heat is gained by the surface.
Consider, the case of two finite surfaces A 1 and A 2 facing each other, as shown in Fig. 1.4.
Let the temperatures be T 1 and T 2 and the emissivities e 1 and e 2 , respectively. Assuming that radiation
exchange occurs only between the two surfaces, the net radiation exchange between them is given by
Q 1 = F 1 A 1 s (T 14 – T 24), W ...(1.9)
where, F1 is known as shape factor or view factor, which includes the effect of orientation, emissivities and the
distance between the surfaces. So, determination of F1 becomes important and we will study about the analysis of
such problems in the chapter on radiation.
Radiation heat transfer coefficient. We define a radiation heat transfer coefficient, h r . This is particularly useful
in cases where convection and radiation occur simultaneously. Analogous to convection, we write,
q 1 = h r (T1 – T2), W/m2, (where h r is the radiation heat transfer
coefficient)
Considering Eq. 1.8, we can write,
Q 1 = A 1 e 1 s (T 14 – T 24) = A 1 hr (T1 – T2)
i.e. hr = e1 s (T 12 – T 22) (T1 + T2), W/(m2K) ...(1.10)

INTRODUCTION AND BASIC CONCEPTS 7


Surroundings

A2, Î2, T2

A1, Î1, T1

FIGURE 1.4 Radiation heat exchange between two finite areas

1.6.4 Combined Heat Transfer Mechanism


So far, we dealt with the three modes of heat transfer, namely, conduction, convection and radiation separately, for
the sake of analysis. However, in practice, one or more of these modes of heat transfer may occur simultaneously.
For example, if a hot casting is removed from its mould and kept open in a room, it will lose heat to the
surroundings by convection as well as radiation. Similarly, a roof heated up by Sun will lose heat to the
surroundings by convection and radiation. In a heat exchanger, if hot and cold fluids are flowing on either side of
the separating wall, the heat transfer involves convection on either side of the wall and conduction through the
wall. In all these cases of combined heat transfer, rate equations for the respective modes of heat transfer and the
First law of thermodynamics will be used to solve the problems. Example 1.2 will make the procedure clear.
Example 1.2. A small metallic sphere of emissivity 0.9 loses heat to the surroundings at a rate of 450 W/m2 by radiation
and convection. If the ambient temperature is 300 K and the convective heat transfer coefficient between the sphere and
ambient air is 15 W/m2K, find out the surface temperature of the sphere.
Solution.

Surroundings
Ta

qrad

A, e, Ts
qconv

Air

Ta, h

FIGURE Example 1.2

8 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Let the surface area of the sphere be A and surface temperature, Ts. Ambient temperature, Ta = 300 K.
Writing an energy balance (i.e. applying the First law):
Total heat lost = heat lost by convection + heat lost by radiation
i.e. 450 ´ A = h A (Ts – Ta) + A e s (Ts4 – Ta4)
i.e. 450 = 15 ´ (Ts – 300) + 0.9 ´ 5.67 ´ 10 – 8 ´ (Ts4 – 300 4)
i.e. 15 T s + 5.103 ´ 10 – 8 ´ Ts4 – 5363.343 = 0
Solving by trial and error, we get Ts = 321.3 K.
[Note: It is important that Ts and Ta are expressed in Kelvin.]
Comment: Many times, combined effect of convection and radiation in a given situation is accounted for by specifying a
combined heat transfer coefficient, h comb.
Then, total heat transferred is given by: Q tot = h comb A(Ts – Ta).
We shall solve this problem now in Mathcad. Our purpose here is to show the ease with which trial and error
solution is done in Mathcad using the ‘solve block’.
Let surface area of the sphere be A, heat transfer flux to surroundings, q. Also, surface temperature is Ts and ambient
temperature Ta. Heat transfer coefficient for convection is h. Then, from energy balance,
q.A = h.A.(Ts – Ta) + A.e.s.(Ts4 – T a4)
i.e. q = h .(Ts – Ta ) + e.s.(Ts4 – T a4)
Data:
q := 450 W/m 2 h := 15 W/(m 2K) Ta := 300 K e := 0.9 s := 5.67×10 – 8 W/(m 2K 4).
Now, Ts is the unknown. We will apply the energy balance as mentioned above and solve for Ts. However, since a 4th
order Ts term is there, we will have to solve it by trial and error.
We use the ‘solve block’ of Mathcad. Here, firstly, we assume a trial value for the unknown quantity, i.e. for T s. Let us
assume say, T s = 400 K. Then, we write the solve block, starting with Given as shown below. Below Given we write the
constraint condition, namely, the energy balance equation in our case. End the solve block by typing Find (T¢s) = and value
of Ts appears immediately.
Ts = 400 K (Trial value of Ts)
Given
450 = h ×(Ts – Ta) + e×s×(Ts4 – T a4) (Constraint equation)
Find (Ts) = 321.3 K.
Example 1.3. Electronic power devices are mounted to a heat sink having an exposed surface area of 0.045 m 2 and an
emissivity of 0.8. When the devices dissipate a total power of 20 W and air and surroundings are at 27°C, the average sink
temperature is 42°C. What average temperature will the heat sink reach when the devices dissipate 30 W for the same
environmental conditions?
Solution. Let us solve this problem in Mathcad:
Let, the surface area of the heat sink be A, and heat dissipated to surroundings, Q. Also, surface temperature is Ts and
surrounding temperature and air temperature Ta. Heat transfer coefficient for convection is h. Then, from energy balance:

Surroundings
Tsurr
Heat sink

As, e, Ts

Air

Ta, h

FIGURE Example 1.3

INTRODUCTION AND BASIC CONCEPTS 9


Heat dissipated by devices = heat lost by sink by convection + heat lost by radiation to surroundings
i.e. Q = h.A.(Ts – Ta) + A×e×s×(Ts4 – T a4)
[Important: Express all temperature in Kelvin since radiation calculations are involved.]
Data:
Case 1: h has to be calculated first using the energy balance; all other quantities are known.
Q := 20 W A := 0.045 m 2 T s := 42 + 273 K Ta := 27 + 273 K Tsurr := 27 + 273 K
e := 0.8 s := 5.67 ´ 10 – 8 W/m 2 K 4
From the energy balance:
Q - s × e × A × (Ts4 - Tsurr
4
)
h=
A × (Ts - Ta )
i.e. h = 24.351 W/m 2 K.
Now, for case 2: If heat dissipated is changed to 30 Watt, what will be the new equilibrium temperature attained by the
sink surface? Other conditions remain the same, i.e. now, h value is known, but Ts has to be calculated.
We will apply the energy balance as mentioned above and solve for Ts. However, since a 4th order Ts term is there, we
will have to solve it by trial and error.
We use the solve block of Mathcad. Here, firstly, we assume a trial value for the unknown quantity, i.e. for Ts . Let us
assume, say, Ts = 450 K. Then we write the solve block, starting with Given as shown below. Below Given, we write the
constraint condition, namely, the energy balance equation in our case. End the solve block by typing Find(Ts)= and value of
Ts appears immediately.
Q := 30 W (It is necessary to update the value of Q, since the value
will be used in energy balance below in solve block)
Ts := 450 K
(Trial value)
Given
Q = h×A×(Ts – Ta) + e×s ×A×(Ts4 – T 4surr)
Find (Ts) = 322.353 K
That is, the new equilibrirm temperature of the heat sink surface is 322.353 K when the amount of heat dissipated
changes to 30 Watt.

1.7 Steady and Unsteady Heat Transfer


In general, temperature distribution within a body depends on position as well as time, i.e. T = T(x, y, z, t). When
the temperature depends only on spatial coordinates and is independent of time coordinate, we call it steady state
heat transfer, i.e. T = T(x, y, z) and T ¹ T (t). However, if the temperature also depends on time coordinate in addition
to the spatial coordinates, then we call it unsteady state heat transfer, i.e. T = T(x, y, z, t). For example, a heat
exchanger, when just started, has its wall temperature changing with both position and time, i.e. it undergoes
unsteady state heat transfer; after sufficient time elapses, it reaches steady state, i.e. temperature becomes inde-
pendent of time and is a function of only position. Similarly, start up of a boiler, quenching of a billet, etc., are
examples of unsteady state heat transfer. In unsteady state heat transfer, internal energy of the system changes.
Changing of temperature with time can also happen in a cyclic manner, i.e. the same temperature occurs at the
same position at definite intervals of time; for example, variation of temperature at a location on earth over a 24-
hour cycle, variation of temperature on the cylinder head of an internal combustion engine, etc. This is known as
periodic or quasi-steady state heat transfer process; in such a case, rate of heat flow and rate of energy storage undergo
periodic variation.
We will study more about this topic of unsteady state heat transfer in the chapter on transient conduction.

1.8 Heat Transfer in Boiling and Condensation


In boiling and condensation, there is a phase change involved during heat transfer. During melting, a solid absorbs
heat and gets converted to liquid; during boiling, the liquid absorbs heat and gets converted to vapour. Reverse
processes occur in solidification and condensation, respectively. In all these cases of heat transfer with phase
change, temperature remains constant during the process. Many practical applications of heat transfer with boiling
and condensation can be cited: different types of chemical process plants, cryogenic separation of gases, distillation

10 FUNDAMENTALS OF HEAT AND MASS TRANSFER


columns, condensers, reboilers, refrigeration and air conditioning applications etc. Heat transfer in boiling and
condensation is characterised by very high values of heat transfer coefficients and should therefore be preferable
from the heat transfer point of view. Further, practically isothermal conditions occurring during these processes
makes them thermodynamically desirable. However, since the mechanism of boiling and condensation are not
amenable to straightforward mathematical treatment, mostly we have to resort to empirical relations to calculate
the heat flux. Further, if there is a flow, we would rather like to avoid boiling/condensation and the resulting two-
phase flow since precise calculation of pressure drops in such cases is difficult. In the chapter on boiling and
condensation, we shall present many useful empirical relations for heat transfer coefficient and heat flux.

1.9 Mass Transfer


Mass transfer is defined as mass in transit caused as a result of species concentration difference in a mixture. Just as the
temperature difference is the driving potential in case of heat transfer, concentration difference is the driving
potential in case of mass transfer.
In a stationary medium, mass transfer occurs purely by diffusion from a plane of high species concentration to
a plane of low species concentration. This is analogous to heat transfer by diffusion in case of conduction heat
transfer.
Governing rate equation for diffusion mass transfer is given by Fick’s law which states that for a binary
mixture of species B and C, the diffusion mass flux of the species B is given by,
mb dCb
Nb = = – Dbc ...(1.11)
A dx
where, Nb = mb/A = Mass flux of species B, kg/(sm2)
A = Area through which mass transfer occurs, m 2
Cb = Concentration of species B, kg/m 3
dCb /dx = Concentration gradient
Dbc = Diffusion coefficient or mass diffusivity, m 2/s
Note the similarity between Fick’s law for mass diffusion and Fourier’s law for heat conduction.
Mass transfer occurs by diffusion as well as convection when a fluid flows over a surface and if there is a
concentration difference of a given species in the fluid and the surface. Relations for convective mass transfer with
relatively low mass concentration levels in the fluid, are similar to those of convective heat transfer, i.e. an analogy
exists between convective heat transfer and convective mass transfer.
There are several practical applications of principles of mass transfer; absorption, desorption, distillation, sol-
vent extraction, drying, humidification, sublimation, etc. In many cases, it is interesting to note that heat and mass
transfer occur simultaneously. We will give an introduction to diffusion and convective mass transfer in the chapter
on mass transfer.

1.10 Summary
In this chapter, we took a bird’s eye view of the three important modes of heat transfer, namely, conduction, con-
vection and radiation. We also mentioned about the topics of heat transfer in boiling and condensation and mass
transfer. We studied that there are three rate equations that govern the three modes of heat transfer, namely,
Fourier’s law for conduction, Newton’s law of cooling for convection and Stefan–Boltzmann law for radiation. We
briefly mentioned about the difference between the science of heat transfer and thermodynamics, steady state and
unsteady state heat transfer and the fundamental laws applied while solving heat transfer problems. Few
application areas of heat transfer and mass transfer were listed.
In the subsequent chapters, we will study in detail about the different modes of heat transfer as well as mass
transfer and their applications.

Questions
1. In what way is the science of heat transfer different from thermodynamics?
2. Explain with examples the three modes of heat transfer.
3. Explain the respective rate equations governing conduction, convection and radiation and mass diffusion.
4. How is boiling heat transfer different from other modes of heat transfer?
5. Differentiate between steady and unsteady state heat transfer.
6. Mention a few industrial applications of mass transfer.

INTRODUCTION AND BASIC CONCEPTS 11


Problems
1. Determine the heat transfer rate per square metre of surface of a cork board, 3 cm thick, when a temperature
difference of 75°C is applied across the board. Take the value of thermal conductivity (k) of cork board as 0.04 W/
(mC).
2. What thickness of glasswool (k = 0.038 W/mC) should be used to limit the heat leak rate to 50 W/m 2 when the
temperature difference across the layer is 50°C?
3. A fluid at 10°C flows over a 2 cm OD, 2 m long tube whose surface is maintained at 150°C. If the heat transfer
coefficient between the fluid and the tube surface is 250 W/(m2C), find out the heat transfer from the tube to the
fluid.
4. An 8 cm diameter sphere is heated internally with a 100 W electric heater. Assuming that the sphere dissipates heat
only by convection, calculate the surface temperature of the sphere if the convection heat transfer coefficient is 15
W/(m2 C). Assume ambient temperature as 25°C.
5. A small plate (5 cm x 5 cm) has its bottom surface insulated and the top surface maintained at 600 K by electric
heating. Emissivity of the surface = 0.85. Find out the heat lost by radiation to the surroundings at a temperature of
300 K.
6. A flat plate is heated at a rate of 750 W/m2 by exposing its one surface to sun and the other surface is insulated. It
loses heat by radiation and convection to surroundings at a temperature of 300 K. If the emissivity of the surface is
0.85 and the convective heat transfer coefficient between the plate and the surroundings is 12 W/(m2 K), determine
the temperature of the plate.
7. A metal casting of size (0.25 m ´ 0.25 m ´ 1.0 m high) and at a temperature of 1200 K is removed from its mould and
kept standing on its end in a large room. Emissivity of the exposed surfaces is 0.85. Find out the heat lost by
radiation and convection if the surrounding is at 300 K and the convective heat transfer coefficient between the
casting surfaces and surrounding air is 12 W/(m 2 K).

12 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

2
Fourier’s Law and Its
Consequences

2.1 Introduction
While studying the subject of heat transfer, one of our objectives is to calculate the rate of heat transfer. From the
second law of thermodynamics, we know that there must be a temperature gradient for heat transfer to occur, i.e.
heat flows from a location of high temperature to a location of low temperature. Fourier’s law gives the relation
between the rate of heat flow and temperature gradient and is therefore considered to be the fundamental law of
conduction.
In this chapter, we will first study Fourier’s law and the assumptions behind this law. Then, follow two impor-
tant consequences of Fourier’s law; the first one being the definition of thermal conductivity—an important trans-
port property of matter, and the second one being the concept of thermal resistance. We will study about the
thermal conductivity of solids, liquids and gases and the variation of this property with temperature. Thermal
resistance concept simplifies the solution of many practical problems of steady state heat transfer with no internal
heat generation, but involving heat transfer through multiple layers or when different modes of heat transfer occur
simultaneously.

2.2 Fourier’s Law of Heat Conduction


This is the basic rate equation for heat conduction which gives a relation between the rate of heat transfer and the
temperature gradient.
Fourier’s law states that one-dimensional, steady state heat flow rate between two isothermal surfaces is proportional to
the temperature gradient causing the heat flow and the area normal to the direction of heat flow.
Referring to Fig. 2.1, we get,
dT
Q µA ...(2.1)
dx
dT
Q = – kA ...(2.2)
dx
where, Q = heat flow rate in X-direction, W
A = area normal to the direction of heat flow (note this carefully), m2
dT/dx = temperature gradient, deg./m
k = thermal conductivity, a property of the material, W/(mC) or W/(mK)
This is the differential form of Fourier’s equation written for heat transfer in the X-direction.
Negative sign in Eq. 2.2 requires some explanation. We know that heat flows from a location of higher temperature
to a location of lower temperature. Referring to Fig. 2.1, if the heat flow rate Q has to occur in the positive X-
direction, temperature has to decrease in the positive X-direction, i.e. temperature must decrease as X increases; this
k means that temperature gradient dT/dx is negative. Since we would
like to have the heat flowing in the positive X-direction to be consid-
T1 Slope = dT/dx ered as positive, a negative sign is inserted in Eq. 2.2, so that Q
becomes positive.
Let us state succinctly the assumptions and other salient points
Q regarding the Fourier’s law:
(i) Fourier’s law is an empirical law, derived from experimental
observations and not from fundamental, theoretical consid-
T2 erations.
(ii) Fourier’s law is defined for steady state, one-dimensional heat
flow.
(iii) It is assumed that the bounding surfaces between which heat
flows are isothermal and that the temperature gradient is
constant, i.e. the temperature profile is linear.
L (iv) There is no internal heat generation in the material.
(v) The material is homogeneous (i.e. constant density) and
X isotropic (i.e. thermal conductivity is the same in all direc-
tions).
FIGURE 2.1 Fourier’s law (vi) Fourier’s law is applicable to all states of matter, i.e. solid,
liquid or gas.
(vii) Fourier’s law helps to define thermal conductivity i.e. from Eq. 2.2 we can write, for steady state heat trans-
fer through a slab of thickness L and its two surfaces at constant temperatures of T1 and T2, (T1 > T2).
Q = – kA dT/dx
= –kA (T2 – T1)/L
= kA (T1 – T2 )/L
We can say that
Q = <k> when A = 1 m 2, dT = 1 deg., dx = 1 m,
i.e. thermal conductivity of a material is numerically equal to the heat flow rate through an area of one m2 of a slab
of thickness 1 m with its two faces maintained at a temperature difference of one degree celcius.
Therefore, the unit of thermal conductivity is obtained from:
L
k =Q , W/(mC) or W/(mK) …(2.3)
A (T1 - T2 )
Note that W/(mC) and W/(mK) mean the same thing in Eq. 2.3, (T1 – T2) is the temperature difference which
is the same whether it is deg.C or deg.K.

2.3 Thermal Conductivity of Materials


We state Fourier’s law again:
Q = – kA dT/dx
= kA (T1 – T2)/L.
Here, k is the thermal conductivity, a property of the material. Its units: W/(mC) or W/(mK). Thermal conduc-
tivity, essentially depends upon the material structure (i.e. crystalline or amorphous), density of material, moisture
content, pressure and temperature of operation.
Thermal conductivity of materials varies over a wide range, by about 4 to 5 orders of magnitude. For example,
thermal conductivity of Freon gas is 0.0083 W/(mC) and that of pure silver is about 429 W/(mC) at normal pressure
and temperature.
Fig 2.2 shows the range of variation of thermal conductivity of different classes of materials:
Table 2.1 gives values of thermal conductivities for a few materials at room temperature.
2.3.1 Thermal Conductivity of Solids
Thermal conductivity of solids is made up of two components,

14 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Solid metals
1000
Liquid metals
Non-metallic solids
100 Non-metallic liquids

Thermal conductivity (w/m°C)


Insulating materials
Non-metallic gases
10
Evacuated insulations

0.1

0.01

FIGURE 2.2 Range of thermal conductivities of various materials

(i) due to flow of free electrons, and


(ii) due to lattice vibrations.
First effect is known as electronic conduction and the second effect is known as phonon conduction.
2.3.1.1 Metals and alloys. In case of pure metals and alloys,
(a) There is an abundance of free electrons and the electronic conduction predominates. Since free electrons are
also responsible for electrical conduction, it is observed that good electrical conductors are also good
thermal conductors, e.g. copper, silver etc.
(b) Any effect which inhibits the flow of free electrons in pure metals reduces the value of thermal conductiv-
ity. For example, with a rise in temperature, the lattice vibration increases and this offers a resistance to the
flow of electrons and therefore, for pure metals thermal conductivity decreases as temperature increases
(uranium and aluminium are exceptions). Fig. 2.3 shows the variation of thermal conductivity with
temperature for a few metals.

TABLE 2.1 Thermal conductivity of a few materials at room temperature

Material k, W/mC
Diamond 2300
Silver 429
Copper 401
Gold 317
Aluminium 237
Iron 80.2
Mercury (l) 8.54
Glass 0.78
Brick 0.72
Water (l) 0.613
Wood (oak) 0.17
Helium (g) 0.152
Refrigerant-12 0.072
Glass fibre 0.043
Air (g) 0.026

FOURIER’S LAW AND ITS CONSEQUENCES 15


(c) Alloying decreases the value of thermal conductivity since
400 Cu the foreign atoms cause scattering of free electrons, thus
380 Cu(99.9%) impeding their free flow through the material. For example,
360
350 thermal conductivity of pure copper near about room
Thermal conductivity, W/m×K

~
~ ~
~ temperature is 401 W/(mC) while presence of traces of
240 Al(99 arsenic reduces the value of thermal conductivity to 142 W/
220 .5%)
(mC).
200
(d) Heat treatment, mechanical forming and cold working
180 Mg reduce the value of thermal conductivity of pure metals.
160
Mg(99.6%) (e) Thermal conductivity of alloys generally increases as
140
temperature increases. Fig. 2.4 below confirms this trend for
120 Zn a few alloys.
100
(f) Since the phenomenon of electron conduction is responsible
80 Pt
for both thermal conduction and electrical conduction, it is
60 Fe(99.2%) reasonable to presume that there must be relation between
40 Pb
Hg these two quantities. In fact, Weidemann–Franz law gives
20
0 this relation. This law, based on experimental results, states
–100 0 100 200 300 400 500 the ratio of thermal and electrical conductivities is the same for all
Temperature, °C metals at the same temperature and this ratio is directly
proportional to the absolute temperature of the metal.
FIGURE 2.3 Variation of thermal conductivity k
with temperature for a few metals µT
s
k
=C ...(2.4)
sT

1 1. Nichrome
2. Brass-30
120
Thermal conductivity, W/m°C

3. Bronze
4. Manganese bronze
2 5. Sn–Zn alloy
6. Gun bronze
4
80 7. Phosphor bronze
6 8. Manganin
3
9. Constantan
7 10 Monel metal
5
11. Liquid Sn–Zn Alloy

40 12. Nickel steel


8
9 11
10
12
0
–200 0 200 400
Temperature, °C
FIGURE 2.4 Variation of thermal conductivity with temperature for a few alloys

16 FUNDAMENTALS OF HEAT AND MASS TRANSFER


where k = thermal conductivity of metal, W/(mK)
s = electrical conductivity of metal, (ohm.m)– 1
C = Lorentz number, a constant for all metals
= 2.45 ´ 10 –8 W Ohms/K 2
An important practical application of Weidemann–Franz law is to determine the value of thermal conductivity
of a metal at a desired temperature, knowing the value of electrical conductivity at the same temperature. Note that
it is easier to measure experimentally the value of electrical conductivity than that of thermal conductivity.
2.3.1.2 Non-metallic solids. In case of non-metallic solids,
(a) For dielectrics, there are no free electrons and the thermal conductivity values are much lower than those of
metals. For heat insulating materials, general range of values of k are from 0.023 W/(mC) to 2.9 W/(mC).
Thermal conductivity increases with temperature for insulating materials as shown in Fig. 2.5.

9 1. Air
0.80
2. Mineral wool
8
3. Slag wool
Thermal conductivity, W/m°C

4. 85% Magnesia
0.60 7 5. Sovelite
6. Diatomaceous brick
7. Red brick
0.40 8. Slag-concrete brick
~
~
9. Fireclay brick
6
0.16

5
4
0.08 3
2
1

0 100 200 300


Temperature, °C

FIGURE 2.5 Variation of thermal conductivity with temperature for insulating materials

(b) For porous heat insulating materials (brick, concrete, asbestos, slag, etc.), thermal conductivity depends
greatly on density of the material and the type of gas filling the voids. For example, k of asbestos increases
from 0.105 to 0.248 W/(mC) as density increases from 400 to 800 kg/m3; this is due to the fact that thermal
conductivity of air filling the voids is much less than that of the solid material.
(c) Thermal conductivity of porous materials also depends on the moisture content in the material; k of a damp
material is much higher than that of the dry material and water taken individually.
(d) Thermal conductivity of granular materials increases with temperature since with increasing temperature,
radiation from the granules also comes into picture along with conduction of medium filling the spaces.
(e) Variation of thermal conductivity of solids with temperature: In heat transfer calculations, generally we assume
k to be constant when the temperature range is small; however, when the temperature range is large, it is
necessary to take into account the variation of k with temperature.
Usually, for solids, a linear variation of thermal conductivity with temperature can be assumed without
loss of much accuracy.
k(T) = k0 (1 + b T) ...(2.5)
where, k(T) = thermal conductivity at desired temperature T, W/(mC)
k0 = thermal conductivity at reference temperature of 0°C, W/(mC)

FOURIER’S LAW AND ITS CONSEQUENCES 17


1.1
70–30 Brass 1. Al
2. Ag
1 3. Cu
1.0 2
3 4. Mg
4 5. Zn
5 6. Pb
0.9
6 7. W
8. Fe
7
k (T) /ko

8 9. Ni
0.8

9
0.7

0.6
0 50 100 150 200 250 300
Temperature, °C

FIGURE 2.6 Variation of thermal conductivity with temperature for a few pure metals

TABLE 2.2 Representative values of k0 and b in Eq. 2.5

Material k0 (W/mC) b ´ 10 4, (1/C)

Metals and alloys


Aluminium 246.985 – 2.227
Chromium 97.123 – 5.045
Copper 401.5275 – 1.681
Stainless steel 14.695 + 10.208
Uranium 26.679 + 8.621
Insulators
Fireclay brick 0.76 0.895
Red brick 0.56 0.66
Sovelite 0.092 0.12
85% Magnesia 0.08 0.101
Slag wool 0.07 0.101
Mineral wool 0.042 0.07

b = a temperature coefficient, 1/C


T = temperature, °C
Fig. 2.6 shows the variation of k with temperature for a few pure metals. It may be noted that the variation
is linear as indicated in Eq. 2.5.
In Eq. 2.5, value of b may be positive or negative. Generally, b is negative for metals (exception being uranium)
and positive for insulators and alloys. Table 2.2 gives representative values of k 0 and b for a few materials.
2.3.2 Thermal Conductivity of Liquids
2.3.2.1 Non-metallic liquids. Heat propagation in liquids is considered to be due to elastic oscillations. As per this
hypothesis, the thermal conductivity of liquids is given by,

18 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4
Ac p r 3
k= 1
...(2.6)
M3
where, cp = specific heat of liquid at constant pressure
r = density of liquid
M = molecular weight of liquid
A = constant depending on the velocity of elastic wave propagation in the liquid; it does not depend on
nature of liquid, but on temperature.
It is noted that the product A.c p is nearly constant. As temperature rises, density of a liquid falls and as per Eq.
2.6 the value of thermal conductivity also drops for liquids with constant molecular weights. (i.e. for non-associated
or slightly associated liquids). This is generally true as shown in Fig. 2.7.
Notable exceptions are water and glycerin, which are heavily associated liquids. With rising pressure, thermal
conductivity of liquids increases. For liquids, k value ranges from 0.07 to 0.7 W/(mC).

0.70
1. Vaseline
0.66 2. Benzene
8 3. Acetone
0.62 4. Castor oil
Thermal conductivity, W/m°C

5. Ethyl alcohol
0.58
6. Methyl alcohol
0.54 7. Glycerine
~
~ ~
~ 8. Water
7
0.30

0.26

0.22
6
5
0.18
3 4
2
0.14
1
0.10
0 20 40 60 80 100 120 140 160
Temperature, °C
FIGURE 2.7 Thermal conductivity of non-metallic liquids

2.3.2.2 Liquid metals. Liquid metals like sodium, potassium etc. are used in high flux applications as in nuclear
power plants where a large amount of heat has to be removed in a small area. Thermal conductivity values of liquid
metals are much higher than those for non-metallic liquids. For example, liquid sodium at 644 K has k = 72.3 W/
(mK); liquid potassium at 700 K has k = 39.5 W/(mK); and liquid bismuth at 589 K has k = 16.4 W/(mK).
2.3.3 Thermal Conductivity of Gases
(a) Heat transfer by conduction in gases at ordinary pressure and temperature is explained by the Kinetic
Theory of Gases. Temperature is a measure of kinetic energy of molecules. Random movement and
collision of gas molecules contribute to the transport of kinetic energy, and, therefore, to transport of heat.
So, the two quantities that come into picture now are: the mean molecular velocity, V and the mean free
path, l. Mean free path is defined as the mean distance travelled by a molecule before it collides with
another molecule.

FOURIER’S LAW AND ITS CONSEQUENCES 19


Thermal conductivity of gases is given by,
1
k= V lc v r ...(2.7)
3
where, V = mean molecular velocity
l = mean free path
c v = specific heat of gas at constant volume
r = density
(b) As pressure increases, density r increases, but the mean free path l decreases almost by the same proportion
and the product l r remains almost constant, i.e the thermal conductivity of gases does not vary much with
pressure except at very low (less than 20 mmHg) or very high (more than 20,000 bar) pressures.
(c) As to the effect of temperature on thermal conductivity of gases, mean molecular velocity V depends on
temperature as follows,
3GT
V=
M
where G = Universal gas constant = 8314.2 J/kmol K
M = molecular weight of gas
T = absolute temperature of gas, K
i.e. mean molecular velocity varies directly as the square root of absolute temperature and inversely as the
square root of molecular weight of a gas. Specific heat, cv also increases as temperature increases. As a
result, thermal conductivity of gases increases as temperature increases.
(d) For the reason stated above, gases with lower molecular weights such as helium and hydrogen have higher
values of thermal conductivities (almost by 5 to 10 times) as compared to gases with higher molecular
weights such as air.
(e) Generally, thermal conductivity values for gases vary in the range 0.006 to 0.6 W/(mC)
(f) Thermal conductivity of steam and other imperfect gases depend very much on pressure unlike that of
perfect gases.
Fig. 2.8 and Fig. 2.9 show the variation of k with temperature for a few gases.

0.15
1. Water vapour
Thermal conductivity, W/m°C

0.13 2. Carbon dioxide


1 3. Air
0.11
4. Argon
0.09 2 5. Oxygen
6. Nitrogen
0.07
3
0.05
4
0.03

0.01
0 200 400 600 800 1000°C
0.05 5
0.04 6

0.03
0 100 200 300 400 500 600

Temperature, °C
FIGURE 2.8 Variation of k with temperature for a few gases

20 FUNDAMENTALS OF HEAT AND MASS TRANSFER


0.16

en
Thermal conductivity, W/m°C

og
dr
Hy
m
liu
He
0.12

0.08

0.04
–200 –100 0 +100
Temperature, °C
FIGURE 2.9 Variation of k with temperature for hydrogen and helium

2.3.4 Insulation Systems


It is appropriate here to consider insulation systems generally used. In industries where huge amount of thermal
energy is dealt with, be it for high temperature or low temperature application, it is necessary to see that the most
suitable insulation is adopted. This has become particularly important now, since there is widespread awareness
about the energy crunch and the cost of energy.
Insulation is required for high temperature systems as well as low temperature systems. In high temperature
systems, any leakage of heat from boilers, furnaces or piping carrying hot fluids represents an energy loss.
Similarly, in low temperature/cryogenic systems, any heat leakage into the low temperature region represents an
energy loss since from thermodynamics we know that to pump out a given amount of heat from a low temperature
region would need a disproportionately large amount of work to be put in at room temperature.
Insulation systems may be classified as,
(i) fibrous
(ii) cellular k
(iii) powder
(iv) reflective.
T1
Since in non-homogeneous insulation materials, a T(x)
combination of conduction, convection or radiation is Q
involved, they are characterised by an “effective thermal
conductivity”. Solid materials have cells of spaces formed
inside them by foaming. There may be air or some other gas T2
inside these voids. Type of gas used affects the property of
the material. Obviously, density of these systems plays an
important role in determining the effective thermal
conductivity. Sometimes, the intervening spaces are
evacuated to reduce the convection losses. To get L
X
extremely low values of thermal conductivity—of the
order of a few mW/(mK)—multiple layers of highly T1 T2
reflective materials are introduced in between the Q Q
insulation layers. These are called superinsulations and are Rcond = L/kA
used in cryogenic and space applications.
Table 2.3 gives details about some of the common FIGURE 2.10 Conduction heat flow through a
insulations used in industry. slab— thermal resistance

FOURIER’S LAW AND ITS CONSEQUENCES 21


TABLE 2.3 Common Insulations used in Industry

Insulation material Density Temperature Thermal conductivity Application


(kg/m 3) range, °C (mW/(mC)
Mineral wool blankets 175 –290 450 – 1000 50 – 130 Vessels, pipes
Mineral fibre blankets 125 Up to 750 37 – 80 Vessels, pipes
Fibre glass mats 10 – 50 60 – 360 30 – 55 Vessels, tanks
Mineral fibre block 210 Up to 1100 50 – 130 Boilers and tanks
Calcium silicate, board/block 25 – 100 230 – 1000 32 – 85 Boilers, chimney liners
Fibre glass board 25 – 100 20– 450 32 – 52 Boilers, tanks, heat
exchangers
Fibre glass blanket 10 – 50 –160 to 230 24– 86 Chillers, tanks, vessels
Expanded polystyrene 20 – 50 –100 to 40 22 – 25 Chilled vessels
Polyurethane foam 25 – 50 – 180 to –150 16 – 20 Low temp. piping
Polyurethane foam 32 – 170 to 110 16 – 20 Tanks, vessels (cold/hot)
Fine perlite, evacuated 180 – 200 to 30 0.95 Cryogenic service
to < 0.001 torr
0.006 mm Al foil + 0.15 20 layers/cm Less than – 150 0.037 Cryogenic service
mm fibreglass paper,
vacuum 10–5 torr
0.0087 mm Al foil 30 layers/cm Less than – 150 0.014 Cryogenic service
+ carbon loaded glass-
fibre paper, vacuum
10 –5 torr

2.4 Concept of Thermal Resistance


2.4.1 Conduction
Consider a slab of thickness L, constant thermal conductivity k, with its left and right faces maintained at tempera-
tures T1 and T2. If T1 is greater than T2, we know that heat will flow from left to right and the heat flow rate is given
by Fourier’s law,
Q = kA(T1 – T2)/L …(2.9)
Now, consider this: in a pipe carrying a fluid, the flow occurs under a driving potential of a pressure difference
and there is resistance to flow due to pipe friction; in an electrical conductor, flow of electricity occurs under the
driving potential of a voltage difference and there is a resistance to the flow of electric current. Similarly,
considering Eq. 2.9, we can say that flow of heat Q occurs in the slab by conduction under a driving potential of a
temperature difference (T1 – T2) and the material offers a thermal resistance to the flow of heat. So, we can write Eq.
2.9 as,
T - T2 DT
Q= 1 =
L L
kA kA
DT Temperature difference
= = ...(2.10)
Rth Thermal resistance
Rth = L/(kA) is known as Thermal resistance of the slab for conduction.
It is seen that there is a clear analogy between the flow of heat and flow of electricity, as shown below,

Situation Driving force Flow Resistance


Electric circuit Voltage, D V Current, I Electric Resistance R = D V/I
Thermal circuit Temp. Difference, D T Heat flow rate, Q Thermal Resist., Rth = D T/Q

22 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Fig. 2.10 above shows the thermal circuit for the situation of flow of heat through a plane slab by conduction.
For the slab, we write,
R cond = L/(kA) …(2.11)
Note that units of thermal resistance is (C/W) or, K/W.
2.4.2 Convection
Consider the case of a fluid flowing with a free stream velocity U and free stream temperature Tf, over a heated
surface maintained at a temperature Ts . Let the heat transfer coefficient for convection between the surface and the
fluid be h. Then, the heat transfer rate from the surface to the fluid is given by Newton’s rate equation,
Q = hA (Ts – Tf)
This can be written as,
Ts - Tf DT Temperature difference
Q= = = ...(2.12)
1 1 Convective thermal resistance
hA hA
Again, note the analogy between flow of electricity and
U, Tf, h
the flow of heat (see Fig. 2.11).
Q Ts
So, for heat transfer by convection, we write,
1
R conv = …(2.13)
( hA)
Note that the units are (C/W) or (K/W). Ts Tf
2.4.3 Radiation Q Q
For the case of heat transfer between two finite surfaces, at
Rconv = 1/hA
temperatures T1 and T2 (Kelvin), net radiation heat transfer
between them is given by equation, FIGURE 2.11 Convection heat transfer—
Q1 = F1 A1 s (T14 – T24), W thermal resistance
where, F 1 is known as shape factor or view factor, which
includes the effects of orientation, emissivities and the distance between the surfaces. s is the Stefan–Boltzmann
constant.
Write the above equation in the following form,
T1 - T2 DT Temperature difference
Q= = = …(2.14)
1 Rrad Radiation thermal resistance
F1 A1 s (T1 + T2 )(T12 + T22 )
Clearly, the radiation thermal resistance may be written as,
1
R rad = , C/W …(2.15)
F1 A1 s (T1 + T2 )(T12 + T22 )

2.4.4 Practical Applications of Thermal Resistance Concept


There are two important practical application of the thermal resistance concept:
(i) To analyse the problems where one or more modes of heat transfer occur simultaneously. For example, in a heat
exchanger plate when a hot fluid flows on one side and a cold fluid on the other side, we have heat transfer
occurring at either surface by convection and through the plate itself by conduction. Obviously, the thermal
resistances in this case are all in series and the rules of series resistances in an electrical circuit apply, i.e.
total thermal resistance is the sum of the three resistances.
i.e. Reff = Rconv1 + Rcond + Rconv2
But in some other cases, the thermal resistances may be in parallel; for example, a heated wall of a furnace
may lose heat to ambient by convection as well as radiation, i.e. heat transfer occurs from the wall by these
two modes simultaneously in parallel. Then we apply the rule for parallel resistances, i.e. effective
resistance is given by,

FOURIER’S LAW AND ITS CONSEQUENCES 23


1 1 1
= +
Reff Rconv Rrad
(ii) To analyse the problems where multiple layers of materials of different thermal conductivities are used; e.g.
in furnace walls which are lagged with 2 or 3 layers of insulation, insulation of walls of houses in cold
weather, lagging of pipes, etc. Since the thermal conductivities and thicknesses of materials used may be
different, thermal resistances of individual layers are different and it becomes convenient to use the thermal
resistance concept to find out the total resistance and hence the heat flow rate.
2.4.5 Limitations for the Use of Thermal Resistance Concept
Thermal resistance concept can be used only when all the following conditions are satisfied.
(i) One-dimensional conduction
(ii) Steady state conduction
(iii) No internal heat generation.
Note: In this chapter, we have just introduced the concept of thermal resistance. We will study more about this
concept and apply it to analyse heat transfer in composite slabs, cylinders and spheres and also to situations where
more than one mode of heat transfer exist simultaneously, in Chapter 4. Therein, we shall also solve several
numerical problems to illustrate the applications of this concept.

2.5 Thermal Diffusivity (a )


Often, during heat transfer analysis, particularly while dealing with transient conduction problems, we come
across a quantity called Thermal diffusivity, defined as,
k m2
a= , ...(2.16)
r cp s
where, k = thermal conductivity of the material, W/(mC)
r = density, kg/m3
cp = specific heat at constant pressure, J/(kg.C)
Note that unit of a is m2/s.
Let us consider the physical significance of thermal diffusivity, a : Thermal conductivity (k) of a material is a
transport property and denotes its ability to conduct heat; higher the value of k, better the ability of material to
conduct heat. The product (r c p) is known as volumetric heat capacity, has units of J/(m3 K), and denotes the ability of
the material to store heat. Higher the value of (rcp), larger the heat storage capacity. Generally, solids and liquids
which are good storage media have higher volumetric heat capacity (> 1 MJ/m3 K ) as compared to gases ( about 1
kJ/m3 K), which are poor heat storage media. Therefore, thermal diffusivity, i.e. the ratio of k to (r cp) gives the
relative ability of the material to conduct heat as compared to its ability to store heat. Larger the value of a, faster the
propagation of heat into the material. In other words, a represents the ability of the material to respond to changes
in the thermal environment; larger the value of a, quicker the material will come into thermal equilibrium with its
surroundings. Values of a for materials vary over a wide range. For example, for copper at room temperature, its
value is approx. 113 ´ 10 –6 m2/s, whereas for glass it is about 0.34 ´ 10 – 6 m2/s.
Table 2.4 shows typical values of thermal diffusivity for a few materials.

2.6 Summary
In this chapter, we studied Fourier’s law for one-dimensional conduction. This is a very important topic and stu-
dent must be clear about the assumptions behind this law; particularly, you should note that the area used in
applying this law is the area normal to the direction of heat flow. Fourier’s law opens the door for further learning
about conduction; we will use it immediately in the next chapter to derive the general differential equation for
conduction heat transfer. In this chapter, we also studied two important consequences of Fourier’s law: firstly,
definition of thermal conductivity—an important transport property of material—and, secondly, concept of
thermal resistance. We studied in some detail about the thermal conductivity of solids, liquids and gases and the
variation of thermal conductivity with temperature. Thermal diffusivity—a significant property while studying
transient conduction—was mentioned and its physical significance explained.

24 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 2.4 Typical values of thermal diffusivity (a ) for a few materials at room temperature

Material a ´ 10 6, (m2/s)
Silver 149
Gold 127
Copper 113
Aluminium 97.5
Iron 22.8
Mercury (l) 4.7
Marble 1.2
Ice 1.2
Concrete 0.75
Brick 0.52
Glass 0.34
Glasswool 0.23
Water (l) 0.14
Beef 0.14
Wood (oak) 0.13

In the next chapter, we shall derive the general differential equation for conduction which, when solved, will
give the temperature distribution in a material; knowing the temperature distribution, we can easily determine the
heat transfer rate by applying the Fourier’s law.

Questions
1 State and explain Fourier’s law for one-dimensional conduction. What are the underlying assumptions?
2. What are the important consequences of Fourier’s law?
3. Define ‘thermal conductivity’. What are the factors affecting the thermal conductivity of a material?
4. Write a short note on thermal conductivity of solids, liquids and gases.
5. How does thermal conductivity vary with temperature for metals, alloys and insulators?
6. Name the different insulations used in industry and mention the specific purpose for which each is used.
7. Explain the analogy between flow of heat and flow of electricity.
8. Explain the concept of thermal resistance. What are the practical uses of this concept?
9. What do you understand by the term ‘thermal diffusivity’? Explain its physical significance.
10. On a cold, winter morning, the aluminium handle of the front door of your house feels cold to touch as compared
to the wooden door frame, even though both were exposed to the same cold environment throughout the night.
Explain why?
11. The inner and outer surfaces of a 5 m ´ 6 m brick wall of thickness 30 cm and thermal conductivity 0.69 W/(mC),
are maintained at temperatures of 20°C and 35°C, respectively. Determine the rate of heat transfer through the
wall.
12. In an experiment to find out the thermal conductivity of a material, an electric heater is sandwiched between two
identical samples, each of size (10 cm ´ 10 cm) and thickness 0.5 cm, and all the four outer edges are well insulated.
At steady state, it is observed that the electric heater draws 35 W of power and the temperature of each sample was
90°C on the inner surface and 82°C on the outer surface. Determine the thermal conductivity of the material at the
average temperature.
13. By conduction, 3 kW of energy is transferred through 0.5 m2 section of a 5 cm thick insulating material of thermal
conductivity 0.2 W/(mC). Determine the temperature difference across the layer.

FOURIER’S LAW AND ITS CONSEQUENCES 25


CHAPTER

3
General Differential
Equations for Heat
Conduction

3.1 Introduction
In heat transfer analysis, one of our objectives is to determine the temperature distribution within the body at any
given instant, i.e. the temperature at every point in the body, taken as a continuum. Then, we can calculate the heat
transfer rate at any point in a given direction by applying the Fourier’s law. Knowledge of temperature distribution
is also required in other fields of engineering analysis; e.g. in calculation of thermally induced stresses, thermal
expansion , optimum thickness of insulation, etc.
General technique to obtain the temperature distribution over the entire body is to consider a differential
control volume within the body and apply the law of conservation of energy to this differential control volume. This
results in a differential equation. Solution of this differential equation with appropriate initial and boundary
conditions gives the temperature field, i.e. the temperature at any point within the body.
In this chapter, first, general differential equations for conduction is derived in Cartesian (i.e. rectangular or
x-y-z) coordinates. This is useful to analyse problems of heat transfer in rectangular-shaped bodies such as squares,
rectangles, parallellopiped, etc. Next, the general differential equation of conduction is stated in cylindrical and
spherical coordinates; these are useful to solve heat transfer problems in cylinders and spheres. Simplifications to
the general differential equations for different, possible practical conditions are presented. Typical boundary
conditions encountered in practice and the methods to represent them mathematically are explained. A summary
of the equations is given at the end of the chapter for ready reference.

3.2 General Differential Equation for Heat Conduction in


Cartesian Coordinates
This is also known as heat diffusion equation or, simply heat equation. Consider a homogeneous body within
which there is no bulk motion and heat transfer occurs in this body by conduction. Temperature distribution within
the body at any given instant is given by: T (x, y, z, t ). The coordinate system used in this derivation is given in
Fig. 3.1.
Consider a differential volume element dx.dy.dz from within the body as shown. It has six surfaces and each
surface may be assumed to be isothermal since the differential element is very small. Further, the body is assumed
to be rigid, i.e. negligible work is done on the body by external mechanical forces.
Let us make an energy balance on this differential element. Let us list out the various energy terms involved:
first, there is energy conducted into the element; second, there is energy conducted out of the element; third, for
generality, let there be energy generated within the element, say, due to joule heating, chemical reaction or nuclear
fission, etc. Net heat conducted into the element in conjunction with the heat generated within the element, will
Qz+dz

Q y+dy

C
G

B F
z
Qx D dz Q x+dx
y H

dy
x
Qy A dx E

Qz

FIGURE 3.1 Nomenclature for derivation of general differential equation for heat conduction
in Cartesian coordinates

obviously cause an increase in the energy content (or the internal energy) of the element. We can write it math-
ematically as
E in – Eout + Egen = E st ...(3.1)
where, E in = energy entering the control volume per unit time
E out = energy leaving the control volume per unit time
E gen = energy generated within the control volume per unit time
E st = energy storage within the control volume per unit time.
Let us calculate these quantities, one by one.
To calculate E in . Energy enters the differential control volume from all the three sides by conduction only, since the
control volume is embedded within the body considered.
Let the energy entering the control volume in the X-direction through face ABCD be Q X. Similarly, QY and QZ
enter the control volume from the faces ABFE and DAEH as shown in the Fig. 3.1.
E in = Qx + Q y + Q z ...(3.2)
To calculate E out . Energy entering the control volume in the X-direction at face ABCD leaves the control volume at
the opposite face EFGH. This is designated as Q x +dx. Similarly, Q y +dy and Q z+dz leave the control volume from the
surfaces opposite to the ones at which they entered. Therefore, we write,
E out = Q x+dx + Q y +dy + Q z+dz ...(3.3)
Now, from calculus, we know that Q x + dx etc. can be expressed by a Taylor series expansion, where, neglecting
the higher order terms, we can write,
¶ Qx
Q x+dx = Q x + ×dx ...(3.4a)
¶x
¶Q y
Q y+dy = Q y + ×dy ...(3.4b)
¶y
¶Q z
Q z+ dz = Q z + ×dz ...(3.4c)
¶z
To calculate E gen . Let there be uniform heat generation within the volume at a rate of qg (W/m3). Heat generation is
a volume phenomenon, i.e. heat is generated throughout the bulk of the body—so, note its units (W/m3). As

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 27


mentioned earlier, heat may be generated within the body due to passage of an electric current, a chemical reaction,
nuclear fission, etc. Then, for the differential control volume dx.dy.dz, we can write,
E gen = qg dx dy dz ...(3.5)
To calculate E st . As a result of the net energy flow into the control volume from all the three directions and the heat
generated within the control volume itself, internal energy of the control volume increases. This will manifest itself
as an increase in the temperature of the control volume. Let the temperature of the control volume increase by dT in
time dt. Then, if r is the density and cp , the specific heat of the material of the control volume, rate of increase of
internal energy of control volume is given by,
¶T
E st = r dx dy dz× cp ...(3.6)
¶t
Now, substituting for all terms in Eq. 3.1, we get,
E in – Eout + Egen = E st
¶T
i.e. (Q x + Q y + Q z) – (Q x+dx + Q y +dy + Q z +dz) + qg.dx.dy.dz = r dx dy dz× cp
¶t

F ¶Q × dx + ¶Q × dy + ¶Q × dzI ¶T
i.e. – GH ¶ x
x
¶y ¶z
y
JK z
+ q g dx dy dz = r cp
¶t
dx dy dz ...(3.7)

Now, let us bring in Fourier’s law of heat conduction. If, for generality, we assume k x, k y, k z to be the thermal
conductivities of the material in the x, y and z-directions respectively, and A x, A y and A z to be the areas normal to
the respective heat flow directions, we can write for the heat flow rates,
¶T ¶T
Q x = –k x A x = – k x dy dz ...(3.8a)
¶x ¶x
¶T ¶T
Q y = –k y A y = – k y dx dz ...(3.8b)
¶y ¶y
¶T ¶T
Q z = –k z A z = – k z dx dy ...(3.8c)
¶z ¶z
Substituting Eq. (3.8) in (3.7), and dividing throughout by dx.dy.dz, we obtain,

¶ FG
¶T IJ + ¶ FG k ¶T I
JK
¶ ¶T FG IJ ¶T
¶x
kx
H
¶x K ¶y H y
¶y
+
¶z
kz
¶z H K
+ q g = r cp
¶t
...(3.9)

This is the general form of heat diffusion equation in Cartesian coordinates, for time dependent (i.e. unsteady
state) heat conduction, with variable thermal conductivity and uniform heat generation within the body. This is a very
important basic equation for conduction analysis. It has to be solved with appropriate initial and boundary
conditions to get the temperature distribution within the body as a function of spatial and time coordinates. Of
course, the heat transfer rate is calculated applying the Fourier’s law, once the temperature distribution is known.
Now, if the material is isotropic, i.e. the thermal conductivity is the same in all the three directions, i.e. k x = k y
= k z = k say, then we can write,

¶T F ¶T ¶TI F¶T I F I

¶x
k
¶x
+ GH

¶y
k
¶y
+

¶z
k
¶zJK
+ q g = r cpGH
¶t JK GH JK ...(3.10)

Eq. 3.10 is the general form of heat diffusion equation in Cartesian coordinates, for time dependent (i.e.
unsteady state) conduction, when thermal conductivity varies with temperature (i.e. with position) and uniform heat
generation occurs within the body.
If k is constant and does not vary with temperature, i.e. k does not change with position, the Eq. 3.10 can be
written as,
F¶ T + ¶ T + ¶ TI + q
2 2 2
¶T
k GH ¶x ¶y ¶z JK
2 2 2 g = r cp
¶t

28 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F¶ T + ¶ T + ¶ TI + q
2 2 2 r cp ¶T 1 ¶T
i.e. GH ¶x ¶y ¶z JK k
2 2 2
g
=
k ¶t
=
a ¶t
...(3.11)

qg 1 ¶T
i.e. Ñ 2T + = ...(3.12)
k a ¶t
where, a = k/(r cp ) is thermal diffusivity, and
Ñ = Laplacian operator.
Solution of general form of heat diffusion equation as given in Eq. 3.10 or 3.12 is rather complicated. However,
in many practical applications, we make simplifying assumptions and the resulting equations are easily solved.
For example:
¶T
(i) Steady state This means that the temperature at any position does not change with time, i.e. = 0. So,
¶t
Eq. 3.12 becomes:

qg
Ñ 2T + =0 ...(3.13)
k
This is known as Poisson equation and is for steady state, three-dimensional heat conduction with heat
generation, with constant thermal conductivity, in Cartesian coordinates.
(ii) With no internal heat generation This means that qg term is zero. So, Eq. 3.12 becomes,
1 ¶T
Ñ2 T = ...(3.14)
a ¶t
This is known as Diffusion equation, and it represents time dependent, three-dimensional heat conduc-
tion, with no internal heat generation, and with constant thermal conductivity, in Cartesian coordinates.
¶T
(iii) Steady state, with no internal heat generation This means that qg and are zero. So, Eq. 3.12
¶t
becomes,
Ñ 2T = 0 ...(3.15)
This is known as Laplace equation, and it represents steady state, three-dimensional heat conduction with
no internal heat generation, with constant thermal conductivity, in Cartesian coordinates.
(iv) One-dimensional, steady state, with no internal heat generation This means that,

¶2 T ¶2 T ¶T
2
= = 0; qg = 0 and =0
¶x ¶y 2 ¶t
So, Eq. 3.12 becomes,

d 2T
= 0. ...(3.16)
dx 2
Note that now, partial derivative is written as full derivative since temperature is dependent on one coordinate
only.
You may be wondering why we have to consider one-dimensional heat flow when we are dealing with three-
dimensional bodies. You will be surprised to know that solution of this simplified version of heat conduction
equation for cases of simple geometries gives results with acceptable accuracy for engineering applications. One-
dimensional conduction implies that temperature gradient is considerable only in one direction and is relatively
negligible in the other three directions; consequently, heat flow is also in only one direction. Examples of such
practical cases are: large slab (or wall) where length in one dimension (say, its thickness) is small compared to the
other two-dimensions—then, temperature varies only along its thickness; long cylinder, whose temperature may
be assumed to vary only along its radius; sphere, whose temperature may be assumed to vary only along its radius.

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 29


3.3 General Differential Equation for Heat Conduction in Cylindrical
Coordinates
Eq. 3.10 derived earlier is suitable to analyse heat transfer in regular bodies of rectangular, square or parallelopiped
shapes. But, if we have to analyse heat transfer in cylindrical-shaped bodies (which are commonly used in practice),
then, working with cylindrical coordinates is more convenient, since in that case, the coordinate axes match with
the system boundaries.
Nomenclature for cylindrical coordinate system is shown in Fig. 3.2.

T(r, f, z)

f
r

f r

FIGURE 3.2 Nomenclature for derivation of general differential equation for heat conduction in cylindrical
coordinate system

Differential equation for heat conduction in cylindrical coordinates may be derived by considering an elemen-
tal cylindrical control volume of thickness dr and making an energy balance over this control volume, as was done
in the case of Cartesian coordinates, or, coordinates transformation can be adopted; for this purpose, transforma-
tion equations are,
x = r cosf
y = r sinf
z =z
f = tan –1(y/x)
The resulting general differential equation in cylindrical coordinates is,
1 ¶ FG¶T IJ + 1 ¶ T + ¶2T2 + qg
2
1 ¶T
r ¶r
r
H¶r K r ¶f ¶z k
2 2
=
a ¶t
...(3.17)

Eq. 3.17 is the general differential equation in cylindrical coordinates, for time dependent, three-dimensional
conduction, with constant thermal conductivity and with internal heat generation.
For one-dimensional conduction in r direction only, we get from Eq. 3.17,
1 ¶ ¶T FG IJ + q g 1 ¶T
r ¶r
r
¶r H K k =
a ¶t
...(3.18)

¶2T 1 ¶T qg 1 ¶T
i.e. 2
+ + = ...(3.19)
¶r r ¶r k a ¶t
Eq. 3.19 represents one-dimensional, time dependent conduction in r direction only, with constant k and
uniform internal heat generation, in cylindrical coordinates.
And, for steady state, one-dimensional heat conduction in r direction only, with constant k and uniform
heat generation, Eq. 3.19 reduces to,

d 2T 1 dT qg
2
+ + = 0. ...(3.20)
dr r dr k

30 FUNDAMENTALS OF HEAT AND MASS TRANSFER


3.4 General Differential Equation for Heat Conduction in Spherical
Coordinates
To analyse heat transfer in spherical systems, working with spherical coordinates is more convenient, since the
coordinate axes match with system boundaries. Nomenclature for the spherical coordinates is shown in Fig. 3.3.

T(r, f, q)
q
q

r r
O y

f f

FIGURE 3.3 Nomenclature for derivation of general differential equation for heat conduction in spherical
coordinate system

Differential equation for heat conduction in spherical coordinates may be derived by considering an elemental
spherical control volume and making an energy balance over this control volume, as was done in the case of
Cartesian and cylindrical coordinates, or, coordinate transformation can be adopted using the following
transformation equations,
x = r sin q sin f
y = r sin q sin f
z = r cos q
The resulting general differential equation in spherical coordinates is,
FG
1 ¶ 2 ¶T IJ + 1 ¶ FG
¶T 1 IJ
¶ 2T qg 1 ¶T
r ¶r
r
H¶r K r 2
sin q ¶q
sin q
H
¶q
+ 2 2 K
r sin q ¶f 2
+
k
=
a ¶t
...(3.21)

Eq. 3.21 is the general differential equation in spherical coordinates, for time dependent, three-dimensional
conduction, with constant thermal conductivity and with internal heat generation.
For one-dimensional conduction in r direction only, we get from Eq. 3.21,
1 ¶ 2 ¶T FG IJ + qg 1 ¶T
r 2 ¶r
r
¶r H K k =
a ¶t
...(3.22)

¶ 2T 2 ¶T q g 1 ¶T
i.e. 2
+ + = ...(3.23)
¶r r ¶r k a ¶t
Eq. 3.23 represents one-dimensional, time dependent conduction in r direction only, with constant k and
uniform internal heat generation, in spherical coordinates.
And, for steady state, one-dimensional heat conduction in r direction only, with constant k and uniform
heat generation, Eq. 3.23 reduces to,
d 2T 2 dT qg
2
+ + = 0. ...(3.24)
dr r dr k

3.5 Boundary and Initial Conditions


Let us rewrite the heat diffusion eqn. i.e. Eq 3.11 in Cartesian coordinates,

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 31


F¶ T + ¶ T + ¶ TI + q
2 2 2
1 ¶T
GH ¶x ¶y ¶z JK k
2 2 2
g
=
a ¶t
Of course, temperature distribution in the body is obtained by solving this differential equation.
We observe that this is a second order differential equation in spatial coordinates; therefore, its solution will
need two boundary conditions to eliminate the two constants of integration. Also, this equation is of first order in
time coordinate; so, it will require one initial condition, i.e. the condition at t = 0.
Commonly encountered Boundary Conditions (B.C.’s) are:
(i) Prescribed temperature conditions at the boundaries—known as B.C. of the first kind or Dirichlet
condition
(ii) Prescribed heat flux condition at the boundaries—known as B.C. of the second kind or Neumann
condition
(iii) Convection boundary condition—known as B.C. of the third kind
(iv) Interface boundary condition—known as B.C. of the fourth kind.
We will study below the method of representing the boundary conditions mathematically.
3.5.1 Prescribed Temperatures at the Boundaries (B.C. of the First Kind)
This situation is represented in Fig. 3.4. Here, it is granted that the temperatures at the boundaries are specified and
are constant. For example, temperature at a surface is
constant if that surface is in contact with a melting solid
or boiling liquid. Or, in more general case, the variation
T(x, t)|x=0 = T1 T(x, t)|x=L = T2 of temperature at the surface may be specified as a
function of position and time. Referring to Fig. 3.4, the
surface at x = 0 is maintained at a constant temperature
T 1 and the surface at x = L is maintained at constant
temperature T2.
Mathematically, these conditions are represented
as:
T(x, t)|x =0 = T(0, t) = T1 ...(3.25 a)
T(x, t)|x =L = T(L, t) = T2. ...(3.25 b)
0 L
3.5.2 Prescribed Heat Flux at the
X Boundaries (B.C. of the Second
Kind)
FIGURE 3.4 Prescribed temperatures at the Here, the heat flux at the boundaries is assumed to be
boundaries (B.C. of the first kind) known. For example, if a surface is heated by an electric
heater, the heat flux entering the surface is known. This
condition is depicted in Fig. 3.5 (a).
In Fig. 3.5 (a), a plate of thickness L is shown. At x = 0, i.e. at the left face, a heat flux q0 is supplied; this is
conducted into the material as shown. At x = L, i.e. at the right face, a heat flux qL is supplied and this is also
conducted into the material. This situation is mathematically represented as follows, remembering that the conduc-
tion flux at a surface is equal to the heat flux supplied. Also, note clearly that –k.(¶T/¶x) represents the heat flux in
the positive X-direction, i.e. from left to right; if the direction of heat flux in a slab is from right to left, obviously, it
is equal to +k (¶ T/¶ x).
At x = 0: q 0 = – k.(¶T/¶ x)|x = 0 …(3.26 a)
At x = L: q L = +k.(¶T/¶ x)|x = L …(3.26 b)
Note again that in Eq. 3.26 b, RHS is positive since the heat flux at x = L is in the negative X-direction, i.e. from
right to left as shown in the Fig. 3.5 (a). Similar relations can be written if the geometry is cylindrical or spherical.
There are two special cases of this boundary condition,
(i) Insulated boundary Many times, to reduce the heat loss (or gain), the boundary is insulated with an
appropriate insulating material. Even though theoretically heat loss will be reduced to zero only with an
infinitely thick insulation thickness, heat loss may be practically assumed to be zero with a sufficiently thick
insulation; we call this as perfect insulation.

32 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Heat flux, qo Conduction Insulation
¶T/¶x|x=0 = 0
q0 = –k ∂T/∂x|x=0

Conduction Heat flux, qL


+k ¶T/¶x|x=L = qL

0 L 0 L

X X

FIGURE 3.5(a) Prescribed heat flux at FIGURE 3.5(b) Insulated boundary at x = 0


the boundaries (B.C. of second kind)

So, for a perfectly insulated boundary at x = 0,


shown in Fig. 3.5 (b), the heat flux across the bound-
ary is zero and we represent this condition math- Centre plane
ematically as follows,
k.(¶T(0, t)/¶x) =0
Zero slope
or, (¶T(0, t)/¶x) = 0 ...(3.26c) Symmetric
(ii) Thermal symmetry In many cases, there is thermal temperature
symmetry over a plane inside the system being ana- distribution
lysed; for example, consider a copper plate, initially
heated to a high temperature and then hung in air
for cooling. It is intuitively clear that heat flow is
from the centre of the plate to the two sides and the 0 L
centre plane will be the plane of symmetry. In other
words, no heat will cross this plane, i.e. this plane is
equivalent to an insulated boundary. X
So, for the centre plane, we can write,
FIGURE 3.5(c) Thermal symmetry at x = L/2
[¶ T(L/2, t)/¶ x] = 0 …(3.26d)
3.5.3 Convection Boundary Condition (B.C. of the Third Kind)
This is a more common practical situation, where heat transfer occurs at the boundary surface to or from a fluid
flowing on the surface at a known temperature and a known heat transfer coefficient, e.g. in heat exchangers,
condensers, reboilers etc.
Consider, again, a slab of thickness L as shown in Fig. 3.6.
At the left surface (x = 0), a hot fluid of temperature T1 is flowing with a heat transfer coefficient h 1, supplying
heat into the body. At the right surface (x = L), a cold fluid at a temperature T2 is flowing on the surface, removing
heat from the body with a heat transfer coefficient h 2.
Equating the conduction heat flux to the convection heat flux on either surface and remembering to note the
direction of heat flow (i.e. whether it is in the positive X-direction or negative X-direction), we can represent this
boundary condition mathematically as follows,
At X = 0: h1(T1 – T|X = 0) =– k.(¶ T/¶ x)|X =0 …(3.27 a)
At X = L: h2(T|X = L – T2) =– k.(¶T/¶ x)|X =L …(3.27 b)
In Eq. 3.27 b, we write for the conduction heat flux at x = L: – k.(¶T/¶x)|X =L since heat is flowing in positive X-
direction.
Using the same principles, expressions can be written for convection boundary conditions at surfaces of
cylindrical and spherical geometries.

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 33


Fluid flow,
T2, h2

Convection Conduction
h2(Tx|=L–T2) = –k ¶T/¶x|x=L

Conduction Convection
h1(T1–T|x=0) = –k ¶T/¶x|x=0

Fluid flow,
T1, h1
0 L

FIGURE 3.6 Convection boundary condition (B.C. of the third kind)

3.5.4 Interface Boundary Condition (B.C. of the Fourth Kind)


When a system is made up of one or more layers of different materials, solution of the problem requires that the
conditions at the interface between the layers A and B is specified. Perfect thermal contact at the interface
presupposes the following requirements:
(i) both the bodies must have the same tempera-
ture at the interface
TA(x) = TB(x)
(ii) heat flux on both the sides of the interface
A B –kA ¶T/¶x = –kB¶T/¶x must be same.
Interface boundary condition is depicted in
Fig. 3.7.
We write, at the interface,
TA = T B …(3.28a)
– k A.(¶ T/¶x) =–k B.(¶ T/¶ x) …(3.28b)
Conduction Conduction Of course, the four boundary conditions ex-
plained above do not cover all the possible boundary
conditions that may be encountered in practice. How-
0 L ever, in any given situation, correct B.C. can be
derived by applying the energy balance at the surface
(i.e. to a control volume of zero volume—which
X means that no energy storage is possible at the control
FIGURE 3.7 Interface boundary condition surface—and, heat entering IN = heat going OUT), as
(B.C. of the fourth kind) was done in deriving Eq. 3.27.

As a further example of this technique, consider a slab of thickness L; at its left surface, it receives heat by
radiation and at its right face, loses heat by radiation. This situation is represented mathematically as shown in
Fig. 3.8.
On the left hand side, energy impinging on the surface by radiation is equated to the energy conducted into the
slab; since the heat is conducted in the positive X-direction the conduction term (flux) has a negative sign as per
Fourier’s law. Similarly, on the right hand face, radiation impinging on the surface is conducted into the slab from
right to left, i.e. in the negative X-direction; therefore, we put a positive sign in the conduction term, as shown in the
Fig. 3.8.

34 FUNDAMENTALS OF HEAT AND MASS TRANSFER


e1 e2

Conduction
Radiation at T1
4 4
e1s(T1 –T |x=0) = –k ¶T/¶x|x=0

Conduction Radiation at T2
4 4
e1s(T2 –T |x=L) = +k ¶T/¶x|x=L

0 L
X

FIGURE 3.8 Radiation boundary conditions at the surfaces

Example 3.1. Temperature variation in a slab is given by: T(x) = 100 + 200 x – 500 x 2, where x is in metres; x = 0 at the left
face and x = 0.3 m at the right face. Thermal conductivity of the material k = 45 W/(mC). Also, c p = 4 kJ/(kgK) and r = 1600
kg/m3. Determine:
(i) Temperature at both surfaces
(ii) Heat transfer at left face and its direction
(iii) Heat transfer at right face and its direction
(iv) Is there any heat generation in the slab? If so, how much?
(v) Maximum temperature in the slab and its location
(vi) Time rate of change of temperature at X = 0.1 m if the heat generation rate is suddenly doubled
(vii) Draw the temperature profile in the slab
(viii) Average temperature of the slab. k, qg
Solution. Temperature profile is given; so, temperatures at the Temperature profile
left and right faces are easily determined by substituting x = 0 and
x = 0.3 m. Maximum temperature is determined by first differenti-
ating T (x) w.r.t. x and equating to zero to get the position (x max) Qleft Qright
where the maximum occurs and then substituting this x max in
T(x). Temperature profile is graphed using Mathcad. Time rate of
change of temperature at X = 0.1 m is found by applying the time
dependent, one dimensional, heat conduction equation in Carte-
sian coordinates. Procedure to determine the average temperature
of the slab is explained at the end. We shall solve this problem in
Mathcad, with suitable comments at each step.
Data: L
L := 0.3 m k := 45 W/mC cp := 4000 J/kg K
k
r := 1600 kg/m3 a := FIGURE Example 3.1(a)
r × cp
i.e. a = 7.031 ´ 10– 6 m2/s
T(x) := 100 + 200x – 500x 2 (Define T(x)... i.e.
temperature as a function of x)
Temperature at left face, i.e. at x = 0 T(0) = 100°C
Temperature at right face, i.e. at x = 0.3 m: T(0.3) = 115°C
To find max. temperature
d
Define the first derivative of T(x): T¢(x): = T(x)
dx
d
Also, define the second derivative of T(x): T²(x): = T²(x)
dx
By hand calculation: we get: T ¢(x) = 200 – 1000.x

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 35


We set T ¢(x) equal to zero to get the position x max where temperature is maximum
i.e. 200 – 1000.x = 0.
This gives x = 0.2 m. Substitute this value of x max in T(x) to get the value of Tmax.
So, Tmax = T(0.2) = 100 + 200 ´ 0.2 – 500 ´ (0.2)2 = 120°C.
However, in Mathcad, all this procedure is very simple. Read the comments in Mathcad solution below.
Set T ¢(x) = 0 and find out the value of x max. To do this, use the root function, which solves the root of T ¢(x) = 0. First,
assume a trial value of x; then use the root function which gives the true value of x
x := 0.15 (Trial value of x)
x max := root(T ¢(x), x) x max is obtained from the root function)
i.e. x max = 0.2 m (value of x where T is a max.)
To get Tmax: Substitute this value of x max in T(x):
T(x max) = 120°C
To Sketch the temperature distribution in the slab:
x := 0, 0.01, ..., 0.3 (Define the range variable x, i.e. x to vary from 0
to 0.3 m in steps of 0.01 m)
To draw the graph:
Just select the x – y plot from pallete, plug in x and T(x) in the place holders:
x is in metres and T(x) in deg. C. Click anywhere outside the graph region; immediately the graph appears.
Note from the graph that the maximum temperature occurs at x = 0.2 m and its value is 120°C, as already calculated.
To calculate the heat fluxes at the left and right faces:
Variation of T(x) with x for Slab
Apply the Fourier’s law at x = 0 and x = 0.3 m, remembering that
120
temperature gradient is given in T ¢(x), already defined.
q left := – k×T¢(0) (applying Fourier’s law
at left face, i.e. at x = 0)
115
i.e. qleft =– 9 ´ 10 3 (Heat flux at the left
face (W/m 2); note that –ve sign
indicates heat flowing from right to left)
T(x) 110 qright := – k×T¢(0.3) (applying Fourier’s law
at right face, i.e. at x = 0)
i.e. q right = 4.5 ´ 10 3 (Heat flux at the right
105 face (W/m 2); note that +ve
sign indicates heat flowing
from left to right)
100 q total := |q left| + |qright | (Total heat generated per
0 0.1 0.2 0.3 m 2 of surface)
x i.e. qtotal = 1.35 ´ 10 W/m (Total heat generated/m 2)
4 2

FIGURE Example 3.1(b) Therefore, q g, the volumetric heat generated rate is given by
total heat generated per unit volume:
qtotal
qg = ; i.e. qg = 4.5 ´ 10 4 W/m 3 (volumetric heat generation rate in the slab)
1× 0 . 3

To calculate the time rate of change of temperature at x = 0.1 m when qg is suddenly doubled:
We have the time dependent differential equation for heat conduction in Cartesian coordinates
¶ 2T q g r cp ¶T ¶T 1 ¶T
+ = =
¶x 2
k k ¶t a ¶t
Therefore,
¶T ¶ 2 T qg
=a 2 + a
¶t ¶x k
¶ 2T ¶T
From the given equation for temperature distribution, it is clear that does not depend on x, i.e. depends only
¶x 2 ¶t
on qg:

36 FUNDAMENTALS OF HEAT AND MASS TRANSFER


dT 2×qg
:= a×T ²(x) + a× (define dT/dt as a function x. Now, we can get
dτ (x) k
dT/dt at any x by simply substituting that value of
x in the function defined)

dT
i.e. = 7.031 ´ 10 –3 C/s (time rate of change of temp.)
dτ(0.1)
Note that this is true for all x since T²(x) does not
depend on x for the temperature distribution given
To determine the average temperature of the slab:
For a differential element of thickness dx, amount of heat energy contained in the element is equal to A.dx.r.c p.T(x). Total
amount of energy in the slab is obtained by integrating this from x = 0 to x = 0.3. Now, if the average temperature of slab is
Tav , amount of energy in the slab can also be written as: r.A.L.cp.Tav. Equating these two expressions, we get

z
L

rALc pTav = (Adx) r cp T(x)


0

z z
L 0. 3
1 1
i.e. Tav = T (x) dx = (100 + 200 x - 500 x2 ) dx
L 0. 3
0 0

Tav =
1 LM (0. 3)2 500
(0.3)3 =
1 OP
i.e.
0. 3 MN
30 + 200 ´
2
-
3 0. 3 PQ
[39 – 4.5] = 115°C

In Mathcad, evaluating the integral within given limits is very easy. First, define Tavg and then just plug in the limits;

z
Mathcad automatically evaluates the integral and gives the value.
L
1
Tavg := × T (x) dx (Mathcad easily does the integration of T(x) within
L 0
the limits specified)

i.e. Tavg = 115°C (Average temperature of the slab.)


Note that Mathcad directly gives the value of the integral within the limits specified; there is no need to expand the
integral and write down as you do in hand calculations.
Example 3.2. Uniform internal heat generation at qg = 5 ´ 10 7 W/m3 occurs in a cylindrical nuclear reactor fuel rod of 50
mm diameter, and under steady state conditions the temperature distribution is of the form:
T(r) = 800 – 4.167 ´ 10 5 r 2, where T is in deg. Celsius and r is in metres. The fuel rod properties are: k = 30 W/(mK), r =
1100 kg/m3 and cp = 800 J/(kgK)
(a) What is the rate of heat transfer per unit length of the rod at r = 0 (i.e. at the centre line) and at r = 25 mm (i.e. at the
surface)?
(b) Sketch the temperature distribution along the radius.
(c) If the reactor power is suddenly increased to 10 8 W/m3, what is the initial time rate of temperature change at r = 0
and r = 25 mm?
(d) Find the average temperature of the rod in the first case.
Solution. Here, temperature distribution is given; so, heat flux can be calculated at any radius r from Fourier’s law: q = – k
(dT/dr). Temperature distribution along the radius is easily graphed with Mathcad. Time rate of change of temperature
when qg changes is found out by applying the time dependent, one-dimensional heat conduction equation in cylindrical
coordinates. Average temperature of the cylinder is obtained from first principles as done in the case of slab in Example 3.1.
Data:
R := 0.025 m
k := 30 W/mC
cp := 800 J/kgK
r := 1100 kg/m 3
L := 1 m

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 37


Temperature profile T(r) := 800 – 4.167 ´ 105 × r 2 (Define T(r)... i.e.
temperature as a
function of r )
d
T ¢(r) := T(r) (Define first derivative of T(r))
dr
d
k, qg T¢¢(r) := T ¢(r) (Define second derivative of T(r))
Q dr
Q centre =0 (heat transfer rate at the centre is zero
since temperature at centre is maximum
and dT/dr = 0 at r = 0.)
To find the heat transfer rate at the surface (i.e. at r = 0.025 m):
Apply Fourier’s law: Q surface =– k A s (dT/dr)| at r = R
T¢(R) = – 2.083 ´ 104 C/m
R
(dT/dr at the surface, i.e. at r = R)
FIGURE Example 3.2(a) Q surface := k ×(2 × p ×R × L) × T ¢(R)
(heat transfer rate at the surface is obtained by applying
Fourier’s law at the surface, i.e. at r = R; T¢(R) is the
temperature gradient at r = R)
i.e. Q surface = 9.818 ´ 10 4 W/metre length (heat transfer at the surface)
Temperature distribution:
r := 0, 0.001, ..., 0.025 (define a range variable, i.e. r varies from 0 to 0.025 m in
steps of 0.001 m)
Then, select the x–y graph from pallete and fill in the place holders in both the axes. On x-axis, fill in r and on y-axis,
fill in T(r). Click anywhere outside the graph region and immediately, the graph appears.
To calculate the time rate of change of temperature at x = 0.1 m:
We have the one-dimensional, time dependent differential equation, with constant k, for heat conduction in cylindrical
coordinates:
d 2T 1 dT q g 1 dT
+ + =
dr 2 r dr k a dt
Therefore, time rate of change of temperature is given by:
¶T F
¶ 2T 1 ¶T qg I
¶t
=a GH
+
¶r 2 r ¶r
+
k JK
Temperature (C) vs. radius (m) for cylinder
800

700

T(r)

600

500
0 0.01 0.02 0.03
r
FIGURE Example 3.2(b)

38 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Our aim is to find out dT/dt when qg changes suddenly to 10 8 W/m3
q g := 108 W/m3
k
a := i.e. a = 3.409 ´ 10 –5 m 2/s (thermal diffusivity.)
r × cp

d
T² (r) := T¢(r) ...Define second derivative of T(r) w.r.t. r = d 2 T/dr 2
dr

dT F qg I ¶T
dt (r) GH 1
:= a× T ¢¢ (r) + × T ¢(r ) +
r k JK (define
¶t
the desired time rate of change of temperatire
as a function of r)

At the surface i.e. at r = R:

F qg I ¶T
GH
dT by dt (r) := a× T ¢¢ (r) +
1
r
× T ¢(r ) +
k JK (define
¶t
, the desired time rate of change)

i.e. dT by dt (0.025) = 56.814 C/s


At the centre, i.e. at r = 0:

F qg I
GH
dT by dt (r) := a× T ¢¢ (r) +
k JK (since at r = 0, dt/dr = 0)

i.e. dT by dt (0) = 85.225 C/s.


To determine the average temperature of the cylinder:
For a differential element of thickness dr, amount of heat energy contained in the element is equal to 2 p.r.dr.L.r.cp.T(r).
Total amount of energy in the cylinder is obtained by integrating this from r = 0 to r = R =0.025 m. Now, if the average
temperature of cylinder is Tav, amount of energy in the cylinder can also be written as: r.p.R2.L.cp .Tav. Equating these two
expressions, we get,

z
R

pR2 Lr c p Tav = (2 prdr) L r cp T (r )


0

z
R
2
i.e. Tav = T (r ) r dr
R2
0

z
0 . 025
2
i.e. Tav = ( 800 - 4.167 ´ 10 5 r 2 ) r dr
R2
0

Tav =
2 LM
(0. 025)2
- 4.167 ´ 10 5 ´
(0.025)4 OP
i.e.
(0.025) 2 ´ 800 ´
MN
2 4 PQ
i.e. Tav = 669.78°C.

z
All the above calculations are done just in one step easily in Mathcad:
R
2
Tavg := × T(r) rdr (define Tavg Mathcad easily does the integration of T(r) within
R2 0 the limits specified)

i.e. Tavg = 669.781°C (Average temperature of the cylinder)


Note that Mathcad directly gives the value of the integral within the limits specified; there is no need to expand the
integral and write down as you do in hand calculations.

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 39


Example 3.3. Consider an orange, assumed to be a sphere of 8 cm diameter, producing an average internal heat generation
of 2.25 ´ 104 W/m 3 during its ripening. Thermal conductivity of the material is 0.15 W/(mK) and its centre temperature is
observed to be 50°C. Assuming one-dimensional, steady state conduction, find out:
(i) temperature distribution along the radius, Temperature profile
(ii) surface temperature,
(iii) heat transferred at the surface of the sphere,
(iv) draw the temperature profile along the radius, and
(v) average temperature of the sphere.
k, qg Q
Solution.
Data:
R := 0.04 m k := 0.15 W/mC qg := 2.25 ´ 10 4 W/m3
Tc := 50 C...centre temperature
(i) Temperature distribution For steady state, one-dimensional conduc- R
tion, for a sphere, we have the controlling differential equation: FIGURE Example 3.3
d 2T 2 dT q g
+ + =0 ...(a)
dr 2 r dr k
To get the temperature distribution, we have to solve Eq. (a) with the following Boundary Conditions (B.C.’s):
(i) at r = 0, (dT/dr) = 0, since the temperature has to be maximum at the centre because the heat flows from centre to
periphery and symmetry considerations.
(ii) Given: Tc = 50°C at the centre, i.e. at r = 0
Multiplying Eq. (a) by r 2:
2
d 2T dT q g r
r2 2
+ 2r + =0
dr dr k

d 2 dT FG
qg r 2 IJ
i.e.
dr
r
dr
+
H k
=0
K
Integrating, we get,
3
dT q g r
or, r2 + = C1
dr 3k

dT qg r C1
or, =- + ...(b)
dr 3 k r2
Integrating again,
- qg r 2 C1
T(r) = - + C2 ...(c)
6k r
Applying B.C. (i) to Eq. b: C 1 = 0
- qg r 2
Then, T(r) = + C2
6k
Applying B.C. (ii) to Eq. c: C2 = 50
Substituting values of C1 and C2 in Eq. c, we get the temperature distribution in the sphere:
- qg r 2
T(r) = + 50. ...(d)
6k
(ii) Surface temperature Now, temperature at the surface is obtained simply by putting r = R = 0.04 m in Eq. d. It is easier
to work in Mathcad; first, define the function T(r):
- qg r 2
T(r) := + 50 (Define T(r)...i.e. temperature as a function of r)
6k

40 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Surface temperature:
T(R) = 10°C (surface temperature, i.e. at r = 0.04 m)
(iii) Next, heat transfer at the surface This is determined by Fourier’s law since we already have the relation for the
temperature distribution:
i.e. Q(r) = –k.(4 pR 2).(dT/dr)| r =R . Gradients such as dT/dr and d 2 T/dr 2 etc. are easily found in Mathcad, once the
function T(r) is defined:
Heat transfer at the surface:
d
T¢(r) := T(r) (Define derivative of T(r))
dr
Q(r) := k×(4×p×R 2)×T ¢(r) (Q(r) is the heat transfer rate at radius r, by Fourier’s law)
i.e. Q(R) = 6.032 W (heat transfer rate at the surface, i.e. at r = R = 0.04 m)
Check: this must be equal to heat generated inside the orange in steady state Qgen.
4
Q gen = qg × ×p×R3
3
i.e. Q gen = 6.032 W...checks.
Sketch the temperature profile along the radius This is done very easily and conveniently in Mathcad. First, define a
range variable r from 0 to R = 0.04 m, in steps of say, 0.001 m; then, select the x – y graph from the pallete and just fill in the
place holders, i.e. fill in r in the place holder on the x-axis and T(r) in the place holder on the y-axis. Click anywhere outside
the graph region and immediately, the graph appears:
Temperature profile along the radius:
r := 0, 0.001.. 0.04 (define the range variable, i.e. r to vary from 0 to 0.04 m,
in steps of 0.001 m)

T(r) vs. r for sphere


60

40
T(r)...in Celsius
T(r)
r...in metres
20

0
0 0.01 0.02 0.03 0.04
r

FIGURE Example 3.3(b)

Note from the graph that maximum temperature occurs at the centre (r = 0); slope of the temperature curve, (dT/dr)
tends to zero (i.e. the curve becomes almost horizontal) as it aproaches the y-axis at r = 0.
(v) Average temperature of the sphere For a differential element of thickness dr, amount of heat energy contained in the
element is equal to 4pr 2.dr.r.cp.T(r). Total amount of energy in the sphere is obtained by integrating this from r = 0 to r = R.
Now, if the average temperature of sphere is Tav, amount of energy in the sphere can also be written as: r.(4/3)p.R3.cp .Tav.
Equating these two expressions we get,

z
R
4
p R 3 rc p Tav = ( 4 pr 2 dr ) r c p T (r)
3
0

z
R
3
i.e. Tav = T (r ) r 2 dr
R3
0

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 41


z FGH I
0 . 04
r2
i.e. Tav =
3
(0.04)3
- qg
6k JK
+ 50 r 2 dr
0

Tav =
3 LM
(0.04)3
- 2. 25 ´ 104 ´
(0.04)5 OP
3 50 ´
i.e.
(0.04) MN3 5 ´ 6 ´ 0.15 PQ
i.e. Tav = 26°C.

z
All the above calculations are done just in one step easily in Mathcad:
R
3
Tavg := × T (r) r 2 dr (define T avg Mathcad easily does the integration of T(r) within the
R3 0
limits specified)

i.e. Tavg = 26°C (Average temperature of the sphere)


Note that Mathcad directly gives the value of the integral within the limits specified; there is no need to expand the
integral and write down as you do in hand calculations.

3.6 Summary of Basic Equations

TABLE 3.1

Sl. No. Equation Remarks

¶ FG
¶T IJ + ¶ FGk ¶T IJ + ¶ FGk ¶T IJ + q ¶T
Three-dimensional, time dependent heat conduc-
1
¶x
kx
H
¶x K ¶y H y
¶y K ¶z H z
¶z K g = r cp
¶t
tion equation with heat generation and tempera-
ture dependent k, in Cartesian coordinates.

F¶T + ¶T + ¶ TI + q
2 2 2
r c p ¶T 1 ¶T
Three-dimensional, time dependent heat conduc-
2 GH ¶x ¶y ¶z JK k
2 2 2
g
=
k ¶t
=
a ¶t
tion equation with heat generation and constant k,
in Cartesian coordinates.

F¶T + ¶T + ¶ TI + q
2 2 2
Poisson equation, i.e. three-dimensional, steady
3 GH ¶x ¶y ¶z JK k
2 2 2
g
=0 state heat conduction equation with heat genera-
tion and constant k, in Cartesian coordinates.

F ¶ T + ¶ T + ¶ T I = 1 ¶T
2 2 2
Diffusion equation, i.e. three-dimensional, time
4 GH ¶x ¶y ¶z JK a ¶t
2 2 2
dependent heat conduction equation with no heat
generation and constant k , in Cartesian coordi-
nates.

Laplace equation, i.e. three-dimensional, steady


¶ 2T ¶ 2T ¶ 2T
5 2
+ 2
+ =0 state heat conduction equation. with no heat gen-
¶x ¶y ¶z 2 eration and with constant k, in Cartesian coordi-
nates.

1 ¶ ¶T FG IJ + 1 ¶ T + ¶ T + q
2 2
1 ¶T
Three-dimensional, time dependent heat conduc-

H K r ¶f ¶z k
g
6 r = tion equation with heat generation and constant.
2 2 2
r ¶r ¶r a ¶t k, in cylindrical coordinates.

1 ¶ 2 ¶T FG IJ + 1 ¶ ¶T FG IJ + Three-dimensional, time dependent heat conduc-


7
r 2 ¶r
r
¶r H K r 2
sin q ¶q
sin q
¶q H K tion equation with heat generation and constant k,
in spherical coordinates.

1 ¶ 2T qg 1 ¶T
+ =
r sin q ¶f 2
2 2
k a ¶t

Contd.

42 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.

¶ FG¶T IJ + q ¶T
One-dimensional, time dependent heat conduc-
8
¶x
k
H¶x K g = r cp
¶t
tion equation with heat generation and tempera-
ture dependent k, in Cartesian coordinates.

1 ¶ FG
¶T IJ + q ¶T One-dimensional, time dependent heat conduc-
9
r ¶r
rk
H
¶r K g = r cp
¶t
tion equation with heat generation and tempera-
ture dependent k, in cylindrical coordinates.

1 ¶ 2 ¶T FG IJ + q ¶T One-dimensional, time dependent heat conduc-


10
r 2 ¶r
r k
¶r H K g = r cp
¶t
tion equation with heat generation and tempera-
ture dependent k, in spherical coordinates.

11 Equations 8, 9, 10 are compactly written as, Compact form of one-dimensional, time


1 ¶ n ¶T FG IJ + q ¶T dependent, heat conduction equation with heat
r n ¶r
r k
¶r H K g = r cp
¶t
generation and temperature dependent k

where, n = 0 for Cartesian coordinates, use x as


variable instead of r
n = 1 for cylindrical coordinates.
n = 2 for spherical coordinates.

1 ¶ n ¶T FG IJ + q 1 ¶T Compact form of one-dimensional, time


H K k
g
12 r = dependent, heat conduction equation with heat
r n ¶r ¶r a ¶t
generation and constant k
where, n = 0 for Cartesian coordinates, use x as
variable instead of r
n = 1 for cylindrical coordinates.
n = 2 for spherical coordinates.

1 ¶ n ¶T FG IJ + q Compact form of one-dimensional, steady


state, heat conduction equation with heat gen-
H K k
g
13 r =0
r n ¶r ¶r eration and constant k
where, n = 0 for Cartesian coordinates, use x as variable
instead of r
n = 1 for cylindrical coordinates.
n = 2 for spherical coordinates.
Alternate form of one-dimensional, steady
d 2T q g state, heat conduction equation with heat gen-
14 + =0
dx 2 k eration and constant k, in Cartesian coordi-
nates.

d 2T 1 dT FG IJ + q Alternate form of one-dimensional, steady

H K k
g
15 + =0 state, heat conduction equation with heat gen-
dr 2 r dr eration and constant k, in cylindrical coordi-
nates.

d 2T 2 dT FG IJ + q Alternate form of one-dimensional, steady


H K k
g
16 + =0 state, heat conduction equation with heat gen-
dr 2 r dr
eration and constant k, in spherical coordi-
nates.

d 2T q g 1 dT FG IJ Alternate form of one-dimensional, time de-


17
dx 2
+
k
=
a dt H K pendent, heat conduction equation with heat
generation and constant k, in Cartesian
coordinates.

Contd.

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 43


Contd.

d 2T FG IJ + q
1 dT FG IJ
1 dT Alternate form of one-dimensional, time depend-
H K k H K
g
18 + = ent, heat conduction equation with heat genera-
dr 2 r dr a dt
tion and constant k, in cylindrical
coordinates.

d 2T 2 dTFG IJ + q FG IJ
1 dT Alternate form of one-dimensional, time depend-
H K k H K
g
19 + = ent, heat conduction equation with heat genera-
dr 2 r dr a dt
tion and constant k, in spherical coordinates.

3.7 Summary
This chapter lays the foundation for the study of heat transfer by conduction. First, general differential equation for
conduction was derived in Cartesian (or, rectangular) coordinates. This equation has to be solved for a given sys-
tem applying the appropriate boundary and initial conditions to get the temperature field. To do this, mathematical
representation of more common types of boundary and initial conditions are explained. Once the temperature
distribution within the body is known, rate of heat transfer (or heat flux) at any point is calculated easily by apply-
ing Fourier’s law. Cartesian coordinates are used while dealing with rectangular geometries such as squares, rec-
tangles, walls, parallelopipes, etc; these geometries find applications in furnaces, boiler walls, walls of buildings, air
conditioning ducts, etc. Next, general differential equations for conduction in cylindrical and spherical systems are
stated. They are useful in solving heat transfer problems involving cylindrical tanks, pipes, spherical storage ves-
sels, reactors, etc. Summary of the basic relations is given in Tabular form for ready reference.
In engineering practice, we ordinarily deal with three-dimensional objects; however, solution of three-dimen-
sional general differential equation is rather complicated. So, a simplifying assumption is made sometimes, of one-
dimensional conduction, i.e. temperature variation is substantial only in one-dimension and the temperature
variation is considered to be negligible in the other two-dimensions. Many practical problems fit into this category:
e.g. walls whose thicknesses are small compared to other dimensions, long cylinders, spheres, etc. In such cases,
analytical solutions for one-dimensional heat transfer problems are very much simplified.
In the next chapter, we shall study one-dimensional, steady state conduction as applied to a few regular
geometries such as slabs, cylinders and spheres.

Questions
1. Derive the general differential equation in rectangular coordinates (i.e. Cartesian coordinates). Therefrom, write
down the governing differential equations for the following cases:
(i) 3-dimensional, constant k, unsteady state conduction with heat generation
(ii) 3-dimensional, constant k, steady state conduction without heat generation
(iii) 3-dimensional, temperature dependent k, steady state conduction with heat generation
(iv) One-dimensional, constant k, unsteady state conduction with heat generation
(v) One-dimensional, temperature dependent k, unsteady state conduction with heat generation
(vi) One-dimensional, constant k, steady state conduction without heat generation
(vii) One-dimensional, constant k, steady state conduction with heat generation.
2. Derive the general equation for the 3-dimensional unsteady state heat conduction with uniform rate of heat gen-
eration in an isotropic solid. Hence, deduce Laplace’s equation.
[V.T.U., Aug. 2001]
3. Write down the two-dimensional, steady state heat conduction equation in x and y variables in rectangular coor-
dinate system, for the case of temperature dependent k and with uniform heat generation in the body.
4. Write down the one-dimensional, time dependent heat conduction equation in spherical and cylindrical
coordinate systems, in the r variable, with temperature dependent k and with uniform heat generation in the body.
5. In a medium, heat conduction equation is given in the following form:

1 ¶ FG
¶T ¶ IJ ¶T FG IJ + q
r ¶r
rk
H
¶r
+
¶z K
k
¶z H K g =0

(a) Is the heat transfer steady or transient?


(b) Is heat transfer one, two or three-dimensional?

44 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(c) Is there heat generation in the medium?
(d) Is the thermal conductivity of the medium constant or variable with temperature?
6. In a medium, heat conduction equation is given in the following form:

FG
1 ¶ 2 ¶T IJ = 1 ¶T
r 2 ¶r H
r
¶r K a ¶t
(a) Is the heat transfer steady or transient?
(b) Is heat transfer one, two or three-dimensional?
(c) Is there heat generation in the medium?
(d) Is the thermal conductivity of the medium constant or variable with temperature?
7. Explain what do you understand by ‘one-dimensional heat conduction’.
8. State the general differential equation for steady state heat conduction in cylindrical and spherical coordinates.
9. What is the need to have the general differential equation for heat conduction in three separate coordinate sys-
tems? Give their applications.
10. What is meant by ‘Initial condition’ and ‘Boundary Condition’?
11. Explain the B.C.’s. of first, second and third kinds. Represent them mathematically.
12. Write down the mathematical formulation of the B.C.’s for heat conduction in a rectangular region 0 £ x £ a, 0 £ y
£ b, for:
(i) Boundary at x = 0: heat removed at constant rate of q0 (W/m2)
(ii) Boundary at x = a: heat dissipation by convection with heat transfer coefficient h a into the ambient air at constant
temperature Ta
(iii) Boundary at y = 0: maintained at a constant temperature T0
(iv) Boundary at y = b: heat supplied into the medium at a rate of q b (W/m 2)
13. Write down the B.C. for the case of a cylindrical wall with inside radius r 1 and outside radius r2, when the inside
surface is heated uniformly at a rate of q (W/m2) and the outside surface dissipates heat by convection with a heat
transfer coefficient h 2 (W/(m2 C) into the ambient air at zero deg.C.
14. A spherical shell, inside radius r1, outside radius r2, is heated at the inner surface electrically at a rate of q1 (W/m2
); outside surface dissipates heat by convection with a convection heat transfer coefficient h 2, into ambient at
temperature Ta Write down the B.C.’s.

Problems
1. A wall, 1.5 m thick has the following temperature distribution:
T(x) = 60 + 18x – 6 x 3 where x is in metres and T(x) is in deg. C. Determine the location of maximum temperature
and the heat flow per m2 area at both the faces. Take k = 25 W/(mC). Also, find out the average temperature of the
wall.
2. Consider a plane wall 2 cm thick, with uniformly distributed heat sources (qg, W/m3) inside its volume; its left and
right faces are maintained at temperatures T1 and T2 , respectively. Steady state temperature distribution in this
wall is given by:
T(x) = 160 – 1000x – 105 x2 . If q g = 40 MW/m3, determine:
(i) Temperatures T1 and T2
(ii) Heat flux at the left face
(iii) Heat flux at the right face
(iv) Heat flux at the centre of the plate
(v) Average temperature of the plate.
3. Temperature distribution in a slab of 1 m thickness is given by:
T(x) = 900 – 300x – 50x2. Heat transfer occurs across an area of 10 m2 and there is uniform heat generation at a rate
of qg = 1000 W/m3. Assume density r = 1600 kg/m3, thermal conductivity k = 45 W/(mK) and specific heat cP = 4
kJ/(kgK). Calculate:
(i) Maximum temperature in the slab
(ii) Energy entering the left face (i.e. at x = 0)
(iii) Energy leaving the wall at right face (i.e at x = 1 m)
(iv) Rate of change of energy storage in the slab , and
(v) Time rate of temperature change at x = 0.5 m in the slab.
4. The temperature distribution across a large concrete slab 50 cm thick, heated from one side, as measured by ther-
mocouples approximates to the relation:

GENERAL DIFFERENTIAL EQUATIONS FOR HEAT CONDUCTION 45


T(x) = 60 – 50x + 12x2 + 20x3 – 15x4 , where T is in deg. C and x is in metres. Considering an area of 5 m2, compute:
(i) heat entering and leaving the slab in unit time
(ii) heat energy stored in unit time
For concrete, take k = 1.2 W/(mK). [V.T. U., Jan./Feb. 2003]
5. A hollow cylinder of inner radius r1 and outer radius r2 has temperature variation along the radius given by:
T(r) = 400 – 400. ln(r/r 1). Thermal conductivity of the material, k = 45 W/(mC).
If r 1 = 5 cm and r 2 = 10 cm, determine the direction and rate of flow of heat at the two surfaces for 1 m length of
pipe.
6. A hollow sphere of inner radius r1 and outer radius r 2 has the temperature along the radius varying as:
T(r) = 400 + 400. ln (r/r 2). Assume k = 45 W/(mK). If r 1 = 5 cm and r 2 = 10 cm, determine the direction and rate of
flow of heat at the two surfaces. Also, find out the average temperature of the sphere.
7. A 5 cm diameter cylindrical rod (k = 15 W/(mC)), with a uniform heat generation rate of q g (W/m 3) inside it, has
a radial temperature distribution given by:
T(r) = 315 – 2.1 ´ 10 4 r 2 where T is in deg. C, r in metres. Determine:
(i) Maximum temperature in the rod
(ii) Volumetric rate of heat generation
(iii) Average temperature of the cylinder.
8. The steady state radial temperature profile in a 10 cm diameter solid sphere is given by:
T(r) = 101.4 – 1390 r 2, where T is in deg. C and r, in metres. Its k = 10 W/(mC). The sphere is placed in an ambient
of 30°C.
(a) What is the maximum temperature in the sphere?
(b) Is there heat generation in the sphere? If yes, at what rate?
(c) Calculate the convection coefficient at the outer surface.

46 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

4
One-dimensional Steady
State Heat Conduction
without Heat Generation

4.1 Introduction
In this chapter, we shall take up the study of one-dimensional, steady state heat conduction, without heat genera-
tion in a few common geometries such as plane slab, cylindrical shell and spherical shell. By one-dimensional
conduction, we mean that temperature variation is significant only in one-dimension and is negligible in other
dimensions; and, steady state means that the temperature does not vary with time at any location. Obviously,
solution of differential equation governing one-dimensional conduction will be much easier than that of the
general differential equation for three-dimensional conduction.
There are many practical instances where the heat conduction may be considered to be one-dimensional, e.g.
a plane slab whose thickness is small as compared to its length and breadth may be considered to have its
temperature varying only along its thickness; temperature in a long, cylindrical shell may be considered to be
varying only along its radius etc.
Solution of the governing differential equation along with the boundary conditions gives the temperature
field within the material and then, by applying Fourier’s law, we can calculate the heat flux at any point.
We shall, first, study heat transfer in three common, important geometries, namely plane slab, cylindrical
and spherical systems, with thermal conductivity of the material remaining constant. Plane slab is an important
case, applicable to analysis of heat transfer in boiler walls, furnace walls, and walls of buildings etc. Cylindrical
geometry is extremely popular for piping, containers etc. along with their insulations. Similarly, sphere is a
popular geometry used in industry to store hot/cold liquids, gases, chemicals etc.
We will also study heat transfer through multiple layers in these three geometries applying the thermal
resistance concept, already mentioned in the second chapter. We shall derive expressions for overall heat transfer
coefficient which is very useful in study of heat exchangers. We shall also present the concept of critical thickness of
insulation and optimum thickness of insulation and study their practical applications.
We shall, next, examine the heat transfer and temperature distribution in these geometries when the thermal
conductivity of the material varies with temperature.
Finally, we shall study briefly about two-dimensional conduction and present values of shape factors for a
few common situations.

4.2 Plane Slab


Consider a plane slab as shown in Fig. 4.1. Let the thickness be L. Temperatures at the two faces are constant and
uniform, i.e. T = T 1 at x = 0 and T = T2 at x = L.
k Assumptions:
(i) One-dimensional conduction, i.e. thickness L is small
T1 compared to the dimensions in the y and z-
directions.
T(x)
Q (ii) Steady state conduction, i.e. temperature at any point
within the slab does not change with time; of course,
temperatures at different points within the slab will
T2 be different.
(iii) No internal heat generation.
(iv) Material of the slab is homogeneous (i.e. constant
density) and isotropic (i.e. value of k is same in all
directions).
L
Our problem is to find out the temperature field within
X the slab and then the heat flux at any point.
T1 T2 We start with the general differential equation in Carte-
Q Q sian coordinates, since the geometry under consideration is a
Rslab = L/(kA) slab, we have from eqn. (3.9):

FIGURE 4.1 Plane slab and thermal circuit ¶ FG


¶T IJ + ¶ FG k ¶T I + ¶ FG k ¶T IJ + q
JK ¶z H ¶z K ¶T
¶x
kx
H
¶x K ¶y H y
¶y
z g = r cp
¶t
¶T ¶T
In this case, = = 0, since one-dimensional conduction, i.e. temperature gradients are zero in y and
¶y ¶z
z-directions
¶T
= 0, since steady state conduction
¶t
qg = 0, since there is no internal heat generation
k x = k y = k z = k, say, since the material is isotropic and not dependent on temperature.
So, the governing equation for the plane slab with the above-mentioned assumptions becomes:
d FGdT IJ
dx
k
Hdx K =0 ...(4.1)

d 2T
i.e. = 0, since k is a constant ...(4.2)
dx 2
Temperature field is obtained by solving Eq. 4.2.
Integrating Eq. 4.2 once:
dT
= C1
dx
Integrating again:
T(x) = C 1 x + C 2 ...(4.3)
Eq. 4.3 is the general solution for the temperature distribution. Values of the two integration constants C1
and C2 are obtained from the two boundary conditions, namely,
B.C.(i): T = T1 at x = 0
B.C.(ii): T = T2 at x = L
From B.C.(i) and Eq. 4.3: T(0) = T1 = C2
From B.C.(ii) and Eq. 4.3: T(L) = T2 = C1 L + C2
= C 1 L + T1
Therefore, C1 = (T2 – T1)/L
Substituting values of C 1 and C 2 in Eq. 4.3, we get,

48 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T2 - T1
T(x) = x + T1 ...(4.4)
L
From Eq. 4.4, we immediately observe that
(i) temperature distribution is linear in the slab
(ii) temperature distribution is independent of k
Eq. 4.4 can be written in non-dimensional form as follows,
T ( x ) - T1 x
= ...(4.5)
T2 - T1 L
Next, to find the heat flux, apply Fourier’s law,
dT T - T1 T - T2
q = –k = –k 2 =k 1 , W/m 2 ...(4.6)
dx L L
Again, note that q is independent of x, i.e. heat flux is the same at every point within the slab.
Now, it is a simple matter to find the heat flow rate, since Q = q.A
kA (T1 - T2 )
i.e. Q= ,W ...(4.7)
L
Also, recollect immediately that the thermal resistance of the plane slab for conduction is given by,
DT DT
R slab = = ...from Eq. (2.7)
Q kA (T1 - T2 )
L
L
i.e. Rslab = ...(4.8)
kA
Note: Many times, weight of insulation may be a criterion while selecting insulation. In such cases, we have,
W = r A L = r A (Rslab k A) = (r k) A2 (Rslab)
i.e. for a specified thermal resistance, material with smallest product of r and k will be the lightest. The analysis
shown above to determine the temperature profile and heat flux is the standard approach to solve a heat
conduction problem, i.e. write down the general differential equation in the appropriate coordinate system,
simplify it with the assumptions applicable to the problem at hand and then solve it in conjunction with the
boundary conditions to get the temperature distribution; once the temperature distribution is known, heat flux is
calculated by the application of Fourier’s law.
Alternatively:
For steady state heat conduction with no heat generation, let us apply the First law, namely,
dEst
E in – Eout + Egen = ...(4.9)
dt
Since there is no heat generation, Egen = 0 and RHS = 0 since it is steady state.
i.e. E in = Eout ...(4.10)
This means that energy entering the system (i.e. slab, in this case) is equal to the energy leaving the system,
i.e. Q is a constant. Then, we can directly integrate the Fourier’s equation. Even though Q is not known to start
with, we know that it is a constant and therefore, Q can be taken out of the integral sign. We proceed as follows:
Refer to Fig. 4.1. From Fourier’s law, we have,
dT
Q = – kA ...(4.11)
dx
Separating the variables and integrating from x = 0 to x = L (with T = T1 to T = T2), we get,

z z
L T2

Q dx = – kA dT , since Q, k, A are constants for the slab.


0 T1

i.e. QL = – kA(T 2 – T 1) = kA (T1 – T 2)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 49


kA (T1 - T2 )
i.e. Q= ,W ...(4.12)
L
Note that Eq. 4.12 for heat transfer rate is the same as Eq. 4.7.
To get temperature distribution: Integrating Eq. 4.11 between x = 0 and x = x. (with correspondingly, T = T 1
and T = T(x)), we get,

z z
x T
Q dx = – kA dT
0 T1

kA (T1 - T ( x))
i.e. Q= ,W ...(4.13)
x
Now, remember that in steady state, Q is the same through each layer of the slab. So, equating Eqs. 4.12 and
4.13, we get,
kA (T1 - T2 ) kA (T1 - T ( x))
i.e. Q= =
L x
(T2 - T1 )
i.e. T(x) = x + T1 ...(4.14)
L
Note that Eq. 4.14 is the same as Eq. (4.4). From Eq. 4.14, we can write the temperature distribution in the
slab in non-dimensional form as follows,
T ( x ) - T1 x
= ...(4.15)
T2 - T1 L
Note that Eq. 4.15 is the same as Eq. 4.5
Note: Above alternative analysis is applicable only for steady state conduction, with no internal heat generation.

4.3 Heat Transfer through Composite Slabs


Heat transfer through a composite slab, consisting of 2 or 3 layers of materials of different thermal conductivities,
is considered next. This is a very common application, e.g. in the case of insulation of furnace walls, insulation of
walls of buildings, refrigerators, cold storage plants, hot water tanks, etc.
While solving heat transfer problems in composite slabs under steady state conditions, it is convenient to
use the thermal resistance concept.
Consider a composite slab consisting of three layers 1, 2 and 3 as shown in Fig. 4.2. Let the thicknesses of the
three layers be L1, L 2 and L 3, respectively; also, the respective thermal conductivities are k 1, k 2 and k 3.
On the LHS of the composite slab, a fluid at a temperature Ta flows on the surface with a convective heat
transfer coefficient of ha and on the RHS of the slab, a fluid at a temperature of Tb flows with a convective heat
transfer coefficient of h b , as shown. Let Ta be higher than Tb, so that steady state heat transfer rate Q is from left
to right as indicated in the Fig. 4.2.
Assumptions:
(i) Steady state, one-dimensional heat conduction.
(ii) No internal heat generation.
(iii) Constant thermal conductivities k1, k2 and k3.
(iv) There is perfect thermal contact between layers, i.e. there is no temperature drop at the interface and the
temperature profile is continuous.
Since it is a case of steady state conduction with no internal heat generation, it is clear from the First law that
heat flow rate Q, through each layer is the same. Referring to Fig. 4.2, it may be seen that heat flows from the
fluid at temperature Ta to the left surface of slab 1 by convection, then by conduction through slabs 1, 2 and 3,
and then, by convection from the right surface of slab 3 to the fluid at temperature Tb.
Let the area of the slab normal to the heat flow direction be A(m2). Now, considering each case by turn,

50 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T1 T2 T3 T4
Fluid flow
1 2 3 Tb, hb

Q Q

Ta, ha
Temperature profile
k1 k2 k3
Fluid flow

L1 L2 L3
x

Ta T1 T2 T3 T4 Tb
Q Q
Ra R1 R2 R3 Rb

FIGURE 4.2 Composite slab with three layers and the thermal resistance network

Convection at the left surface of slab 1


Q = h a A(Ta – T1), from Newton’s Law of Cooling
Q
i.e. Ta – T1 = ...(a)
ha A
Conduction through slab 1
k1A (T1 - T2 )
Q= , from Fourier’s law
L1
QL1
i.e. T1 – T2 = ...(b)
k1 A
Conduction through slab 2
k2 A (T2 - T3 )
Q= , from Fourier’s law
L2
QL2
i.e. T2 – T3 = ...(c)
k2 A
Conduction through slab 3
k3 A (T3 - T4 )
Q= , from Fourier’s law
L3
QL3
i.e. T3 – T4 = ...(d)
k3 A
Convection at the right surface of slab 3
Q = hb A(T4 – Tb), from Newton’s Law of Cooling
Q
i.e. T4 – Tb = ...(e)
hb A

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 51


On adding Eqs. a, b, c, d and e, we get

Ta – Tb = Q
LM 1 + L + L + L + 1 OP
1 2 3
...(f)
Nh A k A k A k A h AQ
a 1 2 3 b

i.e. Ta – Tb = Q[R a + R 1 + R 2 + R 3 + R b] ...(g)


where, R a = convective resistance at left surface of slab 1,
R1 = conductive resistance of slab 1,
R2 = conductive resistance of slab 2,
R3 = conductive resistance of slab 3, and
Rb = convective resistance at right surface of slab 3.
So, we write Eq. g as:

Q=
LM Ta - Tb OP ...(4.16)
NR a + R1 + R2 + R3 + Rb Q
Now, observe the analogy with Ohm’s law. Refer to the Fig. 4.2 for the equivalent thermal circuit. It is clear
that (Ta – Tb) is the total temperature potential, Q is the heat current flowing and the total resistance is the sum of
the individual five resistances which are in series.
For thermal resistances in series, we have,
R tot = S R …(4.17)
For thermal resistances in parallel
Thermal resistances may be arranged in parallel too, as shown in Fig. 4.3.
Here, the main assumption is that the left hand and right
L Insulated hand faces of the composite slab are at uniform and isothermal
temperatures T1 and T2, respectively, as shown in the Fig. 4.3.
T1 T2 Also, the lateral surfaces are insulated so that the heat flow can
be considered as one-dimensional, in the X-direction only.
1 k1 From the analogy with the electrical circuit, when the
resistances are in parallel, the total resistance is given by:
Q Q 1 1 1 1 1
= + = +
2 k2 Rtot R1 R2 L L
T1 T2 k1 A k2 A
R1 R2
X i.e. R tot = ...(4.18)
R1 + R2
R1 = L/(k1A) For thermal resistances in series and parallel: General case of
T1 T2
thermal resistances arranged in series and parallel is shown in
Q Q Fig. 4.4.
Again, remember that one-dimensional heat flow is
assumed; strictly, this is possible only when all the materials of
R2 = L/(k2A) the composite slab have the same value of thermal
conductivity. If the thermal conductivities of materials 1, 2 and
FIGURE 4.3 Composite slab with parallel
3 differ greatly, then obviously, the heat flow will not be one-
resistances
dimensional since the heat will tend to flow through the path of
least resistance. Therefore, it is necessary that for practical purposes, for one-dimensional flow to be applicable,
the thermal conductivities do not vary drastically.
Applying the rules of electrical circuit for series and parallel resistances, we have,
DT T1 - T4
Q= = ...(4.19)
R1 + Reff + R5 R1 + Reff + R5

52 FUNDAMENTALS OF HEAT AND MASS TRANSFER


where, R eff is the effective resistance of the three T2 T3
resistances R2, R3 and R4 in parallel, as shown in Fig. 4.5. T1 T4

1 2 5
1 1 1 1
i.e. = + + ...(4.20)
Reff R2 R3 R4
Q 3 Q
Note: Observe that the concept of thermal resistance is
very useful in solving heat transfer problems in multiple
layers of different thermal conductivities and also when
multimodes of heat transfer are present. Only 4
conditions to be satisfied to apply this concept are (i)
steady state heat transfer, and (ii) no internal heat Insulation
generation. T2 T3

4.4 Overall Heat Transfer R2


T1 T4
Coefficient, U (W/(m2C)) Q Q
Consider the case of a furnace where heat is transferred R1 R3 R5
by the hot gases to the inside surface by convection,
then by conduction through one, two or three layers of
R4
brick and insulation, and finally to ambient air by
convection at the outermost surface. This situation is FIGURE 4.4 Composite slab with series–parallel
represented in Fig. 4.2. resistances and the equivalent thermal circuit
Now, in most of the practical cases, temperature of the hot gases (Ta) and that of the ambient (Tb) are known;
intermediate temperatures are not known. We would like to have the heat transfer given by a simple relation of
the form
Q = U A (Ta – Tb) = UA DT …(4.21)
where, Q is the heat transfer rate (W), A is the area of heat transfer perpendicular to the direction of heat transfer,
and (Ta – Tb) = DT is the overall temperature difference.
Our problem is to derive a relation for U.
Now, we have from Eq. 4.16,

Q=
LM Ta - Tb OP ...(4.16)
NR a + R1 + R2 + R3 + Rb Q
Comparing Eq. 4.16 and Eq. 4.21, we can write

Q = UA(Ta – Tb) =
LM Ta - Tb OP = T - T
a b
NR a + R1 + R2 + R3 + Rb Q åR th

1
i.e. UA = ...(4.22)
åR th

1 W
i.e. U= , ...(4.23)
A åR th m2C

1
Or, U= ...(4.24)
1 L1 L2 L3 1
+ + + +
ha k1 k2 k3 hb
Remember the expression for U as given by Eq. 4.23; it is easier and is applicable when we deal with other
geometries, too.
Concept of overall heat transfer coefficient is particularly useful in heat exchanger designs. Consider a heat
exchanger where a hot fluid flows on one side of a heat exchanger wall and a cold fluid flows on the other side.
Then, heat transfer is by convection on the hot side, by conduction across the separating wall and again by

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 53


convection on the cold side. In such a case, overall heat transfer coefficient is obtained by applying Eq. 4.23.
Values of overall heat transfer coefficients for many practical cases are tabulated in handbooks (see the chapter
on heat exchangers).
Example 4.1. Determine the steady state heat transfer through a double pane window, 0.8 m high, 1.5 m wide, consisting
of two 4 mm thick glass layers (k = 0.78 W/(mC)), separated by a 10 mm thick stagnant layer of air (k = 0.026 W/(mC)).
Inside temperature of room air is maintained at 20°C with a convective heat transfer coefficient of h a = 10 W/(m2C).
Outside air temperature is – 10°C and the convective heat transfer coefficient on the outside is h b = 40 W/(m2 C). Also,
determine the overall heat transfer coefficient.
Solution. The schematic diagram and the equivalent thermal circuit is shown in Fig. Example. 4.1.

T1 T2 T3 T4 Fluid flow
Tb = – 10°C
2
hb = 40 W/m °C

Q Q

Temperature profile
k1 k2 k1
Fluid flow
Ta = 20°C
2
ha = 10 W/m °C
0.01 m
0.004 m 0.004 m

x
Ta T1 T2 T3 T4 Tb
Q Q
Ra R1 R2 R3 Rb

FIGURE Example 4.1 Double pane window and equivalent thermal circuit

This is the case of steady state, one-dimensional conduction, without internal heat generation, through a composite
slab. Therefore, we can conveniently apply the thermal resistance concept. Note that heat transfer occurs from left to
right, i.e. from the warm, inside air to the glass surface on the left by convection, then by conduction through the glass
layer, then again by conduction through the stagnant air layer (no convection here since the air layer is stagnant), and by
conduction through the second glass layer and finally, by convection to the outside cold air.
Let us solve this problem in Mathcad:
Data:
L 1 := 0.004 m, L2 := 0.01 m, L3 := 0.004 m, Ta := 20°C Tb := 10°C ha := 10 W/(m2 C)
hb :=40 W/(m2 C) k1 := 0.78 W/(mC) k2 := 0.026 W/(mC) k3 := k 1 W/(mC)
A :=1.5 ´ 0.8 m 2 i.e. A = 1.2 m2
Convective resistance on the inside, Ra :
1
R a := i.e. R a = 0.083 C/W
ha × A
Conductive resistance through first glass layer, R 1:
L1
R 1 := i.e. R 1 = 4.274 ´ 10 –3 C/W
k1 × A

54 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Conductive resistance through stagment air layer, R2 :
L2
R 2 := i.e. R 2 = 0.321 C/W
k2 × A
Conductive resistance through second glass layer, R3 :
L3
R 3 := i.e. R 3 = 4.274 ´ 10 –3 C/W
k3 × A
Convective resistance on the outside, Rb:
1
R b := i.e. R b = 0.021 C/W
hb × A
Since all the resistances are in series, we write,
R tot := Ra + R1 + R 2 + R3 + R b
i.e. R tot = 0.433 C/W ... total themal resistance
Therefore, heat transfer rate through the double pane window is given by,
Ta - Tb
Q :=
Rtot
i.e. Q = 69.248 W
Overall heat transfer coefficient, U:
1
U := (define U...refer to the derivation of Eq. 4.23)
A × Rtot
i.e. U = 1.924 W/(m2 C).
Note the magnitudes the thermal resistances offered by the glass layer and the air layer. Resistance of air layer is
much more because of its poor thermal conductivity.
It is instructive to see what will be the heat flow rate if a window with a single glass layer is used. In such a case,
If a single layer glass window is used:
R tot = R a + R 1 + Rb i.e. R tot = 0.108 C/W
Ta - Tb
And, Q := i.e. Q = 276.65 W (heat transfer rate with single glass layer)
Rtot
Observe the difference in heat flow rates for a single glass window and a double pane window. This is the reason
why double pane windows are used, particularly in cold weather. Of course, an additional advantage is that it shields
the residents from outside noise, too.
Example 4.2. Find the heat flow rate through the composite wall shown below, in Fig. 4.7. Assume one-dimensional
conduction. Thermal conductivities of slabs A, B, C and D are 150, 30, 65 and 50 W/(mC), respectively. (M.U. Dec.
1997)
Solution. This is a case of steady state, one-dimensional heat conduction in composite slabs with no internal heat gen-
eration. Therefore, thermal resistance concept may be used very conveniently.
Referring to the Fig. Ex. 4.2 we can write,
Q = DT/(RA + RBC +RD)
where, D T = Total temperature difference
RA = thermal resistance of slab A
RBC = effective thermal resistance of slabs B and C, which are in parallel
RD = thermal resistance of slab D.
Calculations are done using Mathcad:
Data:
AA:= 100 ´ 10 –4 m2 AB := 30 ´ 10 –4 m2 AC := 70 ´ 10 –4 m2 AD := 100 ´ 10 –4 m2 LA := 0.03 m
LB := 0.08 m LC := 0.08 m LD := 0.05 m kA := 150 W/(mC) kB := 30 W/(mC) kC := 65 W/(mC)
kD:= 50 W/(mC) DT := (400 – 60) deg.C

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 55


60°C
Q

3 cm

7 cm
D
400°C B

C
A 5 cm
8 cm

Q 10 cm 3 cm

RB
400°C 60°C

Q Q
RD
RB

FIGURE Example 4.2 Composite slab with series–parallel resistances

Calculate the thermal resistances:


LA
R A := i.e. R A = 0.02 C/W
k A × AA
LB
R B := i.e. RB = 0.889 C/W
kB × AB

LC
R C := i.e. R C = 0.176 C/W
kC × AC
LD
R D := i.e. R D = 0.1 C/W
k D × AD
Now, resistances R B and R C are in parallel; Let their effective resistance be RBC :
RB × RC
Then, R BC := i.e. RBC = 0.147 C/W
RB + RC
Total thermal resistance: Rtot := RA + RBC + RD
Adding: R tot = 0.267 C/W
DT
Therefore, heat transfer rate: Q :=
Rtot
i.e. Q = 1.274 ´ 10 3 W.
Example 4.3. A composite wall consists of a 10 cm layer of building brick (k = 0.7 W/(mC)) and 3 cm thick plaster (k =
0.5 W/(mC)). An insulation material of k = 0.08 W/(mC) is to be added to reduce the heat transfer through the wall by
70%. Determine the thickness of the insulating layer.
Solution. In this case the temperatures on either side of the composite wall are not given. So, we assume that the overall
temperature difference D T remains the same in both the cases. Also, since it is the case of steady state heat transfer with
no internal heat generation, we can apply the thermal resistance concept. We have:
Case (i): steady state heat transfer for the composite wall consisting of building brick and plaster. Let the steady state
heat transfer rate be Q; let the thermal resistances of the building brick be R1 and that of plaster be R 2.

56 FUNDAMENTALS OF HEAT AND MASS TRANSFER


DT DT

Q Q 0.3Q 0.3Q

k1 k2 k1 k2 k3

0.1 m 0.03 m 0.1 m 0.03 m L3


X X
Case (i) Case (ii)
FIGURE Example 4.3 Composite wall without and with insulation

Case (ii): steady state heat transfer for the composite wall consisting of building brick and plaster plus the insulation
layer. Now, from problem statement, the steady state heat transfer rate will be 0.3 Q; thermal resistances of the building
brick, plaster and insulation are R 1, R 2 and R3 , respectively.
These two cases are depicted in Fig. Example 4.3.
We have, for case (i): Q = DT/(R1 + R2), and
for case (ii): 0.3 Q = DT/(R1 + R 2 + R 3)
Dividing: (R1 + R2)/(R 1 + R 2 + R 3) = 0.3
Considering a heat transfer area of A = 1 m 2, let us do the calculations in Mathcad:
Data:
L 1 := 0.1 m L2 := 0.03 m k1 := 0.7 W/(mC) k2 := 0.5 W/(mC) k3 := 0.08 W/(mC)
A := 1 m 2 L 3, thickness of insulation layer is to be found out
Thermal resistances:
L1
R1 := i.e. R 1 = 0.143 C/W (thermal resistance of brick layer)
k1 × A
L2
R2 := i.e. R 2 = 0.06 C/W (thermal resistance of plaster layer)
k2 × A
R 3 = L3/(k 3 A) (thermal resistance of insulation layer)
Therefore, R1 + R 2 = 0.203 C/W
And, (R 1 + R 2)/(R1 + R 2 + R 3) = 0.3
R1 + R2
i.e. R3 := - ( R1 + R2 ) ...define R3
0. 3
i.e. R 3 = 0.473 C/W ...define L3
Therefore, thickness of insulation layer:
L3 := R3 ×k 3 ×A
i.e. L 3 = 0.038 m (thickness of insulation layer required to
reduce the heat transfer rate by 70%)
Example 4.4. A square plate heater (size: 15 cm ´ 15 cm) is inserted between two slabs. Slab A is 2 cm thick (k = 50 W/
(mK)) and slab B is 1 cm thick (k = 0.2 W/(mK)). The outside heat transfer coefficients on both sides of A and B are 200
and 50 W/(m2 K), respectively. Temperature of surrounding air is 25°C. If the rating of heater is 1 kW, find:
(i) maximum temperature in the system
(ii) outer surface temperatures of two slabs.
Draw the equivalent circuit for the system (P.U. Nov. 1994)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 57


Q Square plate heater
T1 T0
T2
A B

Q1 Q2
2
2 hb = 50 W/m C
hb = 200 W/m C

Ta = 25°C kA kB
Ta = 25°C

0.02 m
0.01 m

Ta T1 T0 T2 Ta
Q1 Q2
Ra R1 R2 Rb

FIGURE Example 4.4 Two slabs with a plate heater in between and the thermal circuit

Solution. This is a case of steady state conduction, with no internal heat generation within the slabs. So, thermal resist-
ance concept can be applied. Note that, obviously, the maximum temperature, T 0, will occur at the heater in between the
slabs; and, the total heat supplied, Q, is divided into two portions: Q 1 flowing out to the left and Q 2 flowing out to the
right, as shown in Fig. Example. 4.4.
We use the condition: Q = Q 1 + Q 2 = 1000 W
Consider one m2 area of heat transfer, i.e. A = 1 m 2
Then, Q1 = (T0 – Ta)/(R1 + Ra), and
Q2 = (T0 – Ta)/(R2 + Rb)
where, R1 = thermal resistance of slab A
R2 = thermal resistance of slab B
Ra = convective resistance on the left face of A, and
Rb = convective resistance on the right face of slab B.
Let us get the solution in Mathcad:
Data:
L A := 0.02 m L B := 0.01 m k A := 50 W/(mK) k B := 0.2 W/(mK) ha := 200 W/(m2 K)
h b := 50 W/(m 2 K) Ta := 25°C A := 0.15×0.15 m2
i.e. A = 0.022 m2 Q := 1000 W (rating of the heater)
Let the temperature at the heater be T 0...to be found out
Thermal resistances:
LA
R 1 := i.e. R 1 = 0.018 C/W (thermal resistance of slab A)
kA × A
LB
R 2 := i.e. R 2 = 2.222 C/W (thermal resistance of slab B)
kB × A
1
R a := i.e. R a = 2.222 C/W (convective resistance on the left face of slab A)
ha × A
1
R b := i.e. R b = 0.889 C/W (convective resistance on the right face of slab B)
hb × A

58 FUNDAMENTALS OF HEAT AND MASS TRANSFER


To get Q1 and Q2, We have,
(T0 - Ta ) (T0 - Ta )
Q1 = and Q 2 =
( R1 + Ra ) ( R2 + Rb )

FQ I ( R2 + Rb )
i.e. GH Q JK
1

2
=
(R1 + Ra )
...(a)

and, Q1 + Q2 = 1000 ...(b)


Solve Eqs. a and b simultaneously to get Q1 and Q2:
R2 + Rb
Q 1 by Q 2 := i.e. Q 1 by Q 2 = 12.963
R1 + Ra
Therefore, Q1 = 12.963 Q 2
Now, since Q1 + Q2 = 1000, we get
12.963 Q2 + Q2 = 1000
i.e. 13.963 Q 2 = 1000
1000
i.e. Q2 := i.e. Q2 = 71.618 W
13 .963
And, Q1 = 1000 – Q 2 i.e. Q1 = 928.382 W
To calculate maximum temperature, T0 , We have
Q 1 = (T0 – T a)/(R1 + R a)
Therefore:
T0 := Q1 ×(R 1 + Ra) + Ta i.e. T0 = 247.812°C ...maximum temperature in the system
Verify:
T0 - Ta
Q 2 := i.e. Q2 = 71.618 W ...verified.
R2 + Rb

To get T1 and T2, we calculate T1 and T2 applying the Fourier’s law to slab A and B separately:
i.e. Q1 = (T0 – T1)/R1, and Q2 = (T0 – T2)/R2
i.e. T1 := T0 – Q1 ×R1
i.e. T1 = 231.307°C ...temperature on left face of A
And, T2 := T0 – Q2 ×R2
i.e. T2 = 88.66°C ...temperature on right face of B
Example 4.5. A composite insulating wall has three layers of material held together by 3 cm diameter aluminium rivet
per 0.1 m2 of surface. The layers of material consist of 10 cm thick brick, with hot surface at 200°C, 1 cm thick wood with
cold surface at 10°C. These two layers are interposed with a third layer of insulation material 25 mm thick. Thermal
conductivities of materials are: k (brick) = 0.93 W/(mK), k (ins) = 0.12 W/(mK), k (wood) = 0.175 W/(mK) and k(Al) =
204 W/(mK). Assuming one dimensional heat flow, calculate the percentage increase in heat transfer rate due to rivets.
Solution. Consider 0.1 m 2 area of the wall. By data, this area contains one alumilium rivet of 3 cm diameter.
Since this is a case of steady state conduction with no internal heat generation, we can apply the thermal resistance
concept.
There are two cases of heat transfer, (a) with the rivet let the heat transfer rate be Q1, and (b) without the rivet let the
heat transfer rate be Q2. Now, the total temperature drop is the same for both the cases, i.e (200 – 10) = 190 deg.C.
Therefore,
(Q 2/Q1) = (Total thermal resistance with the rivet)/(Total thermal resistance without the rivet)
Mathcad solution to this problem is given below:
Data:
L brick := 0.1 m Lins := 0.025 m L wood := 0.01 m d := 0.03 m L := Lbrick + Lins + L wood
i.e. L = 0.135 m k brick := 0.93 W/(mK) k ins := 0.12 W/(mK) k wood := 0.175 W/(mK)
k Al := 204 W/(mK) A := 0.1 m 2
p × d2
Area of rivet: Arivet :=
4

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 59


Brick Insuation Wood

200 °C 10 °C

Al. rivet,
3 cm diameter

10 cm
2.5 cm 1 cm

R1 R2 R3
200 °C 10 °C

Q1 Q1
RA1

(a) Thermal circuit with the rivet

200 °C R1 R2 R3 10 °C
Q2 Q2
(b) Thermal circuit without the rivet

FIGURE Example 4.5 Composite slab with aluminium rivet

i.e. Arivet = 7.069 ´ 10 –4 m2


Thermal resistances: Let the thermal resistances of brick, insulation and wood, without the rivet in position, be R 1, R2
and R3 , respectively. Then,
Lbrick
R 1 := i.e. R1 = 1.075 C/W (thermal resistance of brick layer)
kbrick × A

Lins
R 2 := i.e. R2 = 2.083 C/W (thermal resistance of insulation layer)
kins × A
Lwood
R 3 := i.e. R3 = 0.571 C/W (thermal resistance of wood layer)
kwood × A
L
R rivet := i.e. R rivet = 0.936 C/W (thermal resistance of rivet)
k AI × Arivet
Now, without the rivet, total thermal resistance is = (R 1 + R 2 + R 3), since all the resistances are in series.
When, the aluminium rivet is in place, strictly speaking, while calculating the thermal resistances of brick,
insulation and wood, area used must be = (0.1 m2 minus the area of rivet); however, note that area of rivet is very small
(i.e. 0.0007068 m2) compared to 0.1 m2. Therefore, with the rivet also, we use the same area of 0.1 m2, i.e. we use the same
R 1, R 2 and R 3.
Without the rivet, total thermal resistance, Rtot:
R tot = R 1 + R2 + R 3, i.e. Rtot = 3.730031 C/W
With the rivet, total thermal resistance, Reff:
Now, refer to Fig. Example 4.5 R tot is in parallel with R rivet. Therefore,
Rtot × Rrivet
Reff := i.e. Reff = 0.748 C/W (effective resistance when rivet is in place)
Rtot + Rrivet
Let Q 1 = heat transfer rate with the rivet and Q 2 = heat transfer rate without the rivet

60 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Let Q 2 by Q 1 = Q 2/Q 1. Since temperature difference is same, i.e. (200 – 10) = 190°C in both cases,
Q 2/Q 1 is equal to the ratio, (Rwith rivet)/(Rwithout rivet)
Reff
Q 2 by Q 1 := i.e. Q2 by Q 1 = 0.201
Rtot

Therefore, % increase in heat transfer,


Increase = (Q 1 – Q 2) ´ 100/Q 1 = [ 1 – (Q 2/Q 1)] ´ 100
Increase := (1 – Q 2 by Q 1) × 100
i.e. Increase = 79.937% (percentage increase in heat transfer rate
due to aluminium rivet)
Note: To be accurate, while calculating the thermal resistances of brick, insulation and wood, if we consider the area as
0.1 m2 minus the area of rivet, i.e. 0.0992932 m2 instead of 0.1 m 2, we get the following results:
0.1
R totnew := Rtot× C/W (new value of Rtot)
0.0992932
i.e. Rtotnew = 3.756582 C/W (new value of Rtot)
With the rivet in place:
Rtotnew × Rrivet
Reff := i.e. Reff = 0.749 C/W (effective resistance when rivet is in place)
Rtotnew + Rrivet
Note that new value of R eff has become 0.749 as compared to earlier value of 0.748.
And, to calculate the increase in heat transfer,
Reff
Q 2 by Q 1 := i.e. Q2 by Q 1 = 0.199
Rtotnew
Increase := (1 – Q 2 by Q 1)×100
i.e. Increase = 80.05% (Note that this is not much different from earlier value of 79.937%.)
Example 4.6. The inside temperature of a furnace wall, 200 mm thick, is 1350°C. The mean thermal conductivity of wall
material is 1.35 W/(mC). The heat transfer coefficient of outside surface is a function of temperaturte difference and is
given by:
h = 7.85 + 0.08 ´ DT where DT is the temperature difference between outside wall surface and surroundings. Deter-
mine the rate of heat transfer per unit area, if the surrounding temperature is 40°C. (M.U., May 2000)
Solution. Refer to Figure Example 4.6.

T1 = 1350°C T2

Q Q
ha = 7.85 + 0.085 DT

Ta = 40°C

0.2 m

x
T1 T2 Ta
Q Q
Rwall Ra

FIGURE Example 4.6 Furnace wall with convection on outside surface

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 61


Data:
A := 1 m2 T 1 := 1350°C L := 0.2 m Ta := 40°C k := 1.35 W/mC h (T2) := 7.85 + 0.08×(T 2 – T a)
Since we have steady state heat transfer with no internal heat generation, we apply the thermal resistance concept.
Also, the heat transfer rate, Q is the same through each layer, i.e. through the furnace wall as well as through the
convective layer adjacent to the outside surface of the furnace wall.
Thermal resistances:
L
R wall := i.e. Rwall = 0.148 C/W
k×A
1
Ra (T2) := (define Ra, the outside convective resistance as a function of T2)
h (T2 )× A
Now, the heat transfer rate through the wall is equal to the heat transfer rate through the outside convective layer.
i.e. Q = (T 1 – T 2)/Rwall, and
Q = (T2 – T a)/Ra
When we equate these two equations and simplify, we get a quadratic equation in T2. Then solve it for T2. Once we
get T2, we can easily calculate Q by applying any one of the above two equations.
We have:
T1 - T2 T - Ta
= 2
Rwall Ra

1350 - T2 T2 - 40
i.e. =
0.148 1
7 . 85 + 0.08 (T2 - 40)
Simplifying, we get,
0.01184 T22 + 1.2146 T2 – 1377.528 = 0
This is a quadratic equation in T2, whose roots are given by,

- b ± b 2 - 4 ac
T2 =
2a
where, a = 0.01184, b = 1.2146 and c = – 1377.528
On substituting, the values of a, b and c, we get

- 1. 2146 ± (1.2146) 2 + 4 ´ 0.01184 ´ 1377 .528


T2 :=
2 ´ 0.01184
Solving, we get
T2 = 293.637°C
And Q = (1350 – 293.637)/0.148
= 7137.59 W.
Note: The above procedure is, however, cumbersome. But, with Mathcad, the problem is easily solved using the solve
block. Here, we start with a trial value of T2 and then, write the constraint, i.e. the equality of the above two equations,
within the solve block, just after Given. Then, the command Find(T2) immediately gives the value of T2.
T2 := 1000°C (Trial value)
Given
T1 - T2 T - Ta
= 2
Rwall Ra (T2 )
Find(T2) = 293.508
T2 := 293.508°C (temperature of outside surface of furnace wall)
To find heat transfer rate, Q:
T1 - T2
Q := (considering the wall only)
Rwall

62 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Q = 7.131 ´ 10 3 W (the heat transfer rate per m 2.)
Verify:
Considering the convective layer:
T2 - Ta
Q := i.e. Q = 7.131 ´ 10 3 W (verified.)
Ra (T2 )

Note: The values of Q obtained by the two methods are almost same, as it should be.

4.5 Thermal Contact Resistance


So far, while dealing with composite slabs, we assumed that there is perfect thermal contact at the interface, which
means that there is no temperature drop at the interface. However, in many cases, particularly when the mating
surfaces are rough, this may not be true and there will be a temperature drop at the interface. This temperature
drop at the interface is due to what is known as contact resistance.
Physical reasoning for contact resistance is explained with reference to Fig. 4.5.

Air gap
Interface
A B
T1 T2

A B
T
Q Q
T1
Tc1

Tc2 T2
X
x

T1 Tc1 Tc2 T2 Ta
Q Q
R1 Rcontact R2

FIGURE 4.5 Thermal contact resistance

Figure. 4.5 shows an enlarged view of the interface between two slabs A and B. It may be observed that
though the surfaces are smooth, physical contact between A and B occurs only at a few points, i.e. at the peaks as
shown. Therefore, heat transfer occurs by conduction through this solid contact area and also by gas conduction
through the gas filling the interfacial voids. Note that there is no convection in the interfacial gas since the space
of interfacial voids is very small; Also, at the temperatures normally encountered, radiation is negligible. So, in
effect, resistance to heat transfer is by two mechanisms:
(i) by solid conduction at the peaks, and
(ii) by gas conduction through the interfacial gas in the voids.
Of these two, solid conduction is usually negligible. Note that there is a temperature drop at the interface,
(Tc1 – Tc2) and the temperature profile is not continuous. Thermal contact resistance is defined as the temperature
drop at the interface divided by the heat transfer rate per unit area.
DT T - Tc 2 m 2 C
Rc = = c1 , ...(4.25)
Q Q W
A A
Interface thermal contact conductance is defined as the inverse of the contact resistance, and is given by:
1 W
h contact = , ...(4.26)
Rc m 2 C

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 63


Thermal contact resistance depends on:
(i) surface roughness—smoother the surface, lesser the resistance
(ii) interface temperature—higher the temperature, lesser the resistance
(iii) interface pressure—higher the pressure, lesser the resistance
(iv) type of material—softer the material, lesser the resistance
Thermal contact resistance may be reduced by:
(i) making the mating surfaces very smooth
(ii) inserting a layer of conducting grease (such as silicon based thermal grease, whose thermal conductivity
is about 50 times that of air) at the interface
(iii) inserting a ‘shim’ (thin foil) made of a soft material such as indium, lead, tin or silver between the
surfaces
(iv) filling the interstitial voids with a gas of higher thermal conductivity than that of air (e.g. helium)
(v) increasing the interface pressure
(vi) in case of permanently bonded joints, contact resistance can be reduced by using an epoxy or soft solder
rich in lead, or a hard solder of gold/tin alloy
Table 4.1 below gives thermal contact resistance for metallic interfaces under vacuum conditions, at
different externally applied pressures.
Table 4.2 illustrates the effect of interfacial fluid on the thermal resistance, for the specific case of an
aluminium interface, with the surfaces having 10 mm roughness and the pressure being 105 N/m 2.
Thermal contact resistance of some typical solid/solid interfaces is given Table 4.3.

TABLE 4.1 Thermal contact resistance (Rc ´ 104, m2 K/W), with vacuum at interface

Contact pressure 100 kN/m2 10,000 kN/m2


Stainless steel 6 – 25 0.7 – 40
Copper 1 – 10 0.1 – 0.5
Magnesium 1.5 – 3.5 0.2 – 0.4
Aluminium 1.5 – 5.0 0.2 – 0.4

TABLE 4.2 Thermal contact resistance, Rc , for AI interface with surface roughness of 10 m m
(Pressure = 10 5 N/m2)

Interfacial fluid R c ´ 104, (m2 K/ W)


Air 2.75
Helium 1.05
Hydrogen 0.72
Silicon oil 0.525
Glycerine 0.265

TABLE 4.3 Thermal contact resistance, Rc , for some solid/solid interfaces

Interface R c ´ 10 4, (m2 K/ W)

Silicon chip/lapped Al in air (27 – 500 kN/m 2) 0.3 – 0.6


Al/Al with indium foil filter (~100 kN/m2) ~0.07
S.S/S.S with indium foil filter (~3500 kN/m2) ~0.04
Al/Al with Dow corning 340 grease (~100 kN/m2) ~0.07
S.S/S.S with Dow corning 340 grease (~3500 kN/m2) ~0.04
Silicon chip/Al with 0.02 mm epoxy 0.2 – 0.9
Brass/Brass with 15 m m tin solder 0.025 – 0.14

64 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Values of Rc are given in heat transfer handbooks.
Generally, contact resistance is in series with other resistances and is taken into account by adding the same
to other resistances. If the area of contact is given (= A) and R c (in m2 C/W) is known, then, R c /A will give the
contact resistance (in C/W) for the given area.
Following problem will clarify the use of this concept.
Example 4.7. Consider a plane composite wall that is composed of two materials of thermal conductivities kA = 0.1 W/
(mK) and kB = 0.04 W/(mK) and thicknesses LA = 10 mm and LB = 20 mm. The contact resistance at the interface between
the two materials is known to be 0.3 m2 K/W. Material A adjoins a fluid at 200°C for which h = 10 W (m2 K) and material
B adjoins a fluid for which h = 20 W/(m2 K).
(i) What is the rate of heat transfer through a wall that is 2 m high and 2.5 m wide?
(ii) Determine the overall heat transfer coefficient.
(iii) Sketch the temperature distribution.
Solution. Refer to Fig. Example 4.7.

T1 Interface T2

A B

Q Q
2
2 hb = 20 W/m C
ha = 10 W/m C

Ta = 200 C Tb = 40 C

T x

Ta
T1 Temperature profile
Tc1

Tc2

T2
Tb

X
Ta T1 Tc1 Tc2 T2
Q Q
Ra R1 Rc R2 Rb

FIGURE Example 4.7 Two slabs with contact resistance at the interface

Data:
L A := 0.01 m L B := 0.02 m k A := 0.01 W/(mK) k B := 0.04 W/(mK) Ta := 200°C Tb := 40°C
ha := 10 W/(m2 K) hb := 20 W/(m2 K) A := 2 ×2.5 m2 Rinterface := 0.3 (m 2 K)/W
Since it is steady state heat transfer and there is no internal heat generation, we can apply the thermal resistance
concept. Overall temperature drop, i.e. the temperature potential, DT = (200 – 40) = 160 deg. C. And, the resistances

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 65


involved are: two convective resistances, two conductive resistances of slabs A and B, and the contact thermal resistance
at the interface. All these resistances may be calculated for a heat transfer area of 5 m 2 and then added up together since
all of them are in series. Then, rate of heat transfer is calculated as, Q = DT/R tot
Thermal resistances:
1
R a := i.e. Ra= 0.02 K/W (convective thermal resistance for fluid adjoining slab A)
ha × A
1
R b := i.e. R b = 0.01 K/W (convective thermal resistance for fluid adjoining slab B)
hb × A
LA
R 1 := i.e. R 1 = 0.02 K/W (conductive thermal resistance of slab A)
kA × A
LB
R 2 := i.e. R 2 = 0.1 K/W (conductive thermal resistance of slab B)
kB × A
Rinterface
R c := i.e. R c = 0.06 K/W (contact resistance at the interface for A = 5 m2)
A
Therefore, total thermal resistance, Rtot:
Rtot := R A + R 1 + Rc + R 2 + Rb i.e. Rtot = 0.21 K/W
Heat transfer rate, Q:
Ta - Tb
Q := i.e. Q = 761.905 W (heat transfer rate)
Rtot
Overall heat transfer coefficient, U:
1
We have: U :=
A × Rtot
i.e. U = 0.952 W/(m2 K)
Temperature drops:
DTa := Q ×R a i.e. DTa = 15.238°C (temperature drop in the convective layer adjoining to A)
Therefore, T 1 := T a – DTa i.e. T 1 = 184.762°C (temperature at left face of A)
DT 1 := Q ×R 1 i.e. DT 1 = 15.238°C (temperature drop in the layer A)
Therefore, Tc 1 := T1 – DT1 i.e. Tc 1 = 169.524°C
DT c := Q ×R c i.e. DT c = 45.714°C (temperature drop at the interface)
Therefore, Tc 2 := Tc1 – DTc i.e. Tc 2 = 123.81°C
DT 2 := Q ×R 2 i.e. DT 2 = 76.19°C (temperature drop in the layer B)
Therefore, T 2 := Tc 2 – DT2 i.e. T 2 = 47.619°C (temperature at right face of B)
DT b := Q ×R b i.e. DT b = 7.619°C (temperature drop in the convective layer adjoining to B)
Check: the temperature Tb must equal 40°C
Therefore, T b := T 2 – DTb i.e. T b = 40°C (temperature of fluid adjoining to B matches with data given)

4.6 Conduction with Variable Area


While considering Fourier’s law, namely,
dT
Q = – kA
dx
the area A is normal to the direction of heat transfer. In case of plane slabs studied so far, the area normal to the
direction of heat flow was constant. However, this need not be the case always. In practice, many times, we come
across solid shapes like truncated cones, truncated wedges, developed cylinder, etc. that may be used as
structural members, struts or supports. Analysis of heat transfer through such members is easily done by direct
integration of Fourier’s equation, under the following conditions:

66 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i)One-dimensional conduction QX QX+dx
(ii)Steady state heat transfer
T1 T2
(iii) No internal heat generation
(iv) Variation of area can be expressed mathematically
as a function of x.
Q Q
Consider a truncated cone as shown in Fig. 4.6.
The left face of the truncated solid is at the
coordinate position x 1 and the right face is at coordinate
X1
position x 2, as shown. Let the left face be at temperature
T 1 and the right face at T 2. Let T 1 > T 2. We would like to X2
find out the heat transfer rate Q and the temperature
distribution in the solid, under steady state corditions. X
Consider a differential control volume between x = x
dX
and x = x + dx.
By applying the First law, it is clear that heat transfer FIGURE 4.6 Conduction through variable area
rate, Q x = Q x + dx, since there is steady state, one-
dimensional heat transfer with no internal heat generation, i.e. the heat transfer rate is the same through all
sections and is a constant, Q. Further, let us say that area A is a function of x and can be expressed mathemati-
cally as A = A(x).
To start with, we do not know the value of Q, but we do know that it is a constant; so, we can take Q out of
the integral sign, while directly integrating Fourier’s equation,
dT
Q = – k(T) A(x) ...(4.27)
dx
for the general case when thermal conductivity, k, is a function of temperature, T.
Separating the variables,
Q dx = – k(T) A(x)dT ...(4.28)
Now, integrating between x = x1 and x = x2, with corresponding T = T1 and T = T2,
we write,

z z
x2 T2
dx
Q = – k (T ) dT ...(4.29)
A ( x)
x1 T1

If k is a constant and does not vary with temperature, we can write,

z z
x2 T2
dx
Q = –k dT ...(4.30)
A ( x)
x1 T1

Now, if T1 and T2 at any two corresponding x values of x1 and x2 are known, then Q can be calculated from
the Eq. 4.30. To obtain the temperature distribution in the solid, we use the condition that Q is the same through
each layer in steady state, i.e. get Qx by integrating between x1 and any x (i.e. temperature varying from T1 to
T(x)) and equate this to the already obtained value of Q.
If k is varying with temperature, use Eq. 4.29 and follow the same procedure.
We illustrate the method by solving a problem.
Example 4.8. A conical cylinder of length L and radii R 1 and R2, (R1 < R 2) is fully insulated on the outer surface. The two
ends are maintained at T1 and T 2, (T 1 > T1). Considering one-dimensional steady state heat flow, derive expressions for
heat flow and temperature distribution.
As a numerical example, taking: R 1 = 1.25 cm, R 2 = 2.5 cm, L = 20 cm, T 1 = 227°C, T 2 = 27°C, k = 40 W/(mC), find:
(i) steady state heat transfer rate, Q
(ii) temperature at mid-plane
(iii) temperature at a plane 15 cm from the small end
(iv) draw the temperature profile in the solid
Solution. Refer to Fig. Example 4.8.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 67


x Assumptions:
(i) steady state
T1 T2
(ii) one-dimensional conduction, in X-direction
(since sides are insulated)
Rx (iii) no internal heat generation
Q Q (iv) constant k
To determine heat transfer rate:
R1 R2 At any distance x from the origin, heat transfer rate Qx is
given by:
dT
L Q x = – kAx
dx
Insulated
where, Ax is the area normal to the direction of heat
X flow at x.
Radius at x is given by:
FIGURE Example 4.8 Conduction with variable area
R2 - R1
R x = R1 + x = R1 + C.x,
L
R2 - R1
where C= = constant
L
Now,
Ax = p Rx2 = p .(R1 + C.x)2
Therefore,
dT
Q = –kp (R 1 + C.x) 2 ...(a)
dx
Separating the variables and integrating from x = 0 to x = L and correspondingly, T = T1 to T = T2, we get:

z
L

z
T2
dx
Q = - k p dT ...(b)
( R1 + C. x) 2
0 T1

Note that Q is taken out of the integral sign since it is a constant and is the same at all sections.
Now, LHS of Eq. b is given by:

i.e. Q
z
L

0
( R1
dx
+ C . x) 2 zL

0
LM ( R + C.x) OP
= Q ( R1 + C. x) - 2 dx = Q
MN (- 1).C PQ
1
-1 L

L OP
QL 1O QL M
i.e. LHS = - M
1
-
C MN R + CL R PQ
P R - R MM R1 - R
= – -
1
R1
PP ...(substituting for C)
MN R + L L
PQ
1 1 2 1 2 1
1

i.e. LHS =
QL L 1
R - R NR
M - R1 OPQ = RQLR
2 1 1 2 1 2

since, C = (R2 – R1)/L


RHS of Eq. b is given by:
RHS = kp (T1 – T2)
On equating LHS and RHS, we get

QL LM R - R OP = kp (T – T )
2 1
R2 - R1 MN R R PQ
1 2
1 2

k p (T1 - T2 ) R1 R2
i.e. Q= ...(c)
L
Eq. c is the desired expression for the heat transfer rate.

68 FUNDAMENTALS OF HEAT AND MASS TRANSFER


To determine temperature distribution:
Let temperature at any x be T(x).
General procedure is to equate the expression for heat transfer at any x obtained by integrating Eq. b between x = 0
and x = x (i.e. between T = T1 and T = T(x)), and the expression c derived above for Q between the two sections at known
temperatures of T1 and T2.
So, Eq. b becomes:

z z
x T ( x)
dx
Q = - k p dT ...(d)
( R1 + C . x)2
0 T1

Now, LHS of Eq. d becomes:

z LM (R + C.x) OP LM OP
x x
-1
Q 1 1
LHS = Q (R1 + C. x)- 2 dx = Q 1

0
MN (- 1) C PQ 0
= -
MN
C R1 R1 + Cx PQ
Remembering that C = (R 2 – R1)/L and Rx = R1 + C.x, we get:

LHS =
QL LM
1 1 OP = QL LM (R - R ) OPx 1

Q (R - R ) MN R R PQ
- ...(e)
N
( R2 - R1 ) R1 Rx 2 1 1 x

RHS of Eq. d becomes:

z
T ( x)

- k p dT = k p (T 1 – T(x)) ...(f)
T1

Equating Eqs. e and f, we get


R1 Rx ( R2 - R1 ) k p (T1 - T ( x))
Q= ...(g)
L ( Rx - R1 )
Now, equate the two expressions for Q, given by Eqs. c and g,
k p (T1 - T2 )R1 R2 k p (T1 - T (x)) R1 Rx (R2 - R1 )
=
L L ( Rx - R1 )

T1 - T (x) R ( Rx - R1 )
i.e. = 2
T1 - T2 Rx (R2 - R1 )

R
1- 1
T1 - T (x) Rx
i.e. = ...(h)
T1 - T2 R
1- 1
R2
Eq. h gives the expression for temperature distribution in the solid as a function of x. Again, remember that Rx is the
radius at any x and is given by:
Rx = R1 + (R2 – R1).(x/L).
Now, let us solve the numerical example in Mathcad. Refer to Fig. Example 4.8.
Data:
R 1 := 0.0125 m R2 := 0.025 m T1 := 227°C T2 := 27°C L := 0.2 m k := 40 W/(mC)
Heat transfer rate, Q:
k × p × (T1 - T2 ) × R1 R2
Q := (define Q, as given by the Eq. c, derived above)
L
On substituting values, we get
Q = 39.27 W (heat transfer rate through the section)
Tempearature at mid-plane, i.e. at x = 0.1 m:
x := 0.1 m (at mid-plane of the section)
We have, at x = 0.1 m, Rx = R1 + (R2 – R1).(x/L)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 69


FG x IJ
Therefore: R x (x) := R 1 + (R2 – R1)×
H LK (define R x as a function of x)

i.e. R x (0.1) = 0.1 m (radius at x = 0.1 m)


Now, temperature distribution is given by Eq. h derived above:

F1 - R I
GH R (x) JK x
1

From Eq. (h): T (x) := T 1 – (T1 – T )×


2
F1 - R I (define T as a function of x)
GH R JK 1

Therefore, T (0.1) = 93.667°C (temperature at mid-plane.)


Temperature at x = 15 cm from LHS:
Simply substitute x = 0.15 in T(x):
T(0.15) = 55.571°C (temperature at a plane 15 cm from LHS)
To draw the temperature profile within the solid:
In Mathcad, this is very easy. Define a range variable x, varying from x = 0 to x = 0.2 m. Temperature T(x) as a function
of x is already defined. From the graph pallette, choose x – y graph and fill in the place holders on the x-axis and y-axis.
Click anywhere outside the graph region and immediately the graph appears: (see Fig. Ex. 4.8(b) below)
x := 0.0.01, ... , 0.2 (define the range variable x, varying from 0
to 0.2 m, at an interval of 0.01 m)

Temperature distribution in truncated cone


250 x is in metres and
temperature in deg.C

200

150
T(x)

100

50

0
0 0.05 0.1 0.15 0.2
x

FIGURE Example 4.8(b)

Note from the graph that temperature at the mid-plane, i.e. at x = 0.1 m is 93.67°C, as calculated earlier.
Exercise: If the RHS, i.e. the larger diameter end, is at 227°C and the LHS is at 27°C, how does the heat transfer and the
temperature distribution change?
Example 4.9. A structural support has the shape of a truncated cone (see Fig. Example 4.9) of length 0.2 m and its area
varies with x as A = (p/4).x 3. Its circumference is perfectly insulated. Thermal conductivity of the material varies with
temperature and is given by: k(T) = 14.695(1 + 0.0010208 T), where T is in deg. C and k is in W/(mC). What is the steady
state heat transfer rate through this strut if the two ends are maintained at 400°C and 150°C, as shown? Also, find the
temperature at the mid-plane. Draw the temperature profile in the solid.
Solution.
p 3 2
Data: A(x) := ×x m x 1 := 0.08 m x2 := 0.28 m L := x 2 – x 1 i.e. L = 0.2 m
4
k(T) := 14.695 (1 + 0.0010208T) W/(mC) T1 := 400°C T2 := 150°C

70 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Insulated k(T) = 14.695 (1 + 0.0010208 T)

400 C 150 C

Q Q

X1 = 0.08 m

X2 = 0.28 m

X
L = 0.2 m
FIGURE Example 4.9 Conduction with variable area and variable thermal conductivity

Assumptions:
(i) steady state conduction
(ii) one-dimensional conduction
(iii) no internal heat generation
(iv) k varying with temperature.
To find heat transfer rate, Q:
Since this is a case of steady state, one-dimensional conduction, with no internal heat generation, Q is constant through
each section in the solid; so, we can directly integrate Fourier’s equation, keeping the Q outside the integral. Integrating
between two known temperatures at x = x1 and x = x2, we calculate Q.
Q = – k (T ) A(x) dT/dx …(a)
Substituting for k(T) and A(x):
FG p x IJ dT
Q = –14.695 ´ (1 + 0.0010208 T) ´
H4 K
3
´ ...(b)
dx
dx p
i.e. Q = –14.695 ´ (1 + 0.0010208 T) ´ ´ dT ...(c)
x3 4
Since Q is constant, separating the variables and integrating between x = 0.08 and x = 0.28 (with T1 = 400°C and T2
= 150°C), we get:

Q
z
0 . 28

0 . 08
dx
x3
= –11.5356
z
150

400
(1 + 0.0010208 T) ´ dT ...(d)

LM
Q -
1 OP LM
0 . 28
T O
P
= – 11.5356 ´ T + 0. 0010208 ´
2
150

i.e.
MN 2 x2 PQ 0 . 08N 2 Q 400

L
Q [78.125 – 6.378] = – 11.5356 ´ M(150 - 400) +
0.0010208 2
- 400 2 )
OP
i.e.
N 2
´ (150
Q
On simplifying, we get
Q ´ 71.747 = 3692.315
3692. 315
i.e. Q= = 51.47 W ...(e)
71.747
i.e. the steady state heat transfer rate for the solid is 51.47 W.
In Mathcad, above calculation is done just in one step; there is no need to expand the integral and substitute the
values. We write directly from Eq. a:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 71


z
T2

- k (T ) dT

z
T1
Q= x2 i.e. Q = 51.505 W (heat transfer rate through the section)
1
dx
A (x )
x1

This value of Q matches with the value obtained earlier.


To find temperature at mid-plane, i.e. at x = 0.18 m:
Let the temperature at mid-plane be Tm. Since, Q is a constant, which is already calculated, integrate between x = 0.08
and x = 0.18 with corresponding T = 400°C and T = Tm; we get a quadratic equation in Tm and solving it, get Tm:
Integrating Eq. c between x = 0.08 and x = 0.18, with T = 400°C and T = Tm, we get,

Q -
LM 1 OP 0 . 18
LM
= – 11.5356 ´ T + 0. 0010208 ´
T2 OP Tm

MN 2 x2 PQ 0 . 08 N 2 Q 400
...(f)

i.e.

51.47 ´ [78.125 – 15.432] = – 11.5356 ´ (Tm - 400 ) +


LM 0.0010208
´ (Tm2 - 400 2 )
OP
N 2 Q
On simplifying, we get,
58.878 ´ 10 –4 ´ Tm2 + 11.5356 ´ Tm – 2329.48 = 0 ...(g)
Eq. g is a quadratic in Tm and its roots are given by,

- b ± b 2 - 4 ac
Tm = ...(g)
2a
–4
where, a = 58.878 ´ 10 , b = 11.5356 and c = – 2329.48
Substituting:

- 11. 5356 ± (11. 5356) 2 + 4 ´ 58.878 ´ 10 - 4 ´ 2329 . 48


Tm =
2 ´ 58.878 ´ 10 - 4
On solving, we get,
Tm = 184.599°C
i.e. temperature at the mid-plane is 184.599°C.
Above calculation is rather laborious to perform by hand. In Mathcad, it solved very easily using the solve block:
Temperature at the mid-plane, i.e. at x = 0.18 m:
Let the temperature at mid-plane (x = 0.18 m) be Tm. Integrate between x = 0.08 m and x = 0.18 m and use the condition
that heat flow rate is the same (already calculated) through all sections. Use the solve block starting with a trial value of
Tm:
x := 0.18 m (at mid-plane)
T m := 250°C (trial value of Tm)

z z
Given
x Tm
1
Q. dx = k (T ) dT
x1 A (x ) T1

Temp(x) := Find (Tm) (temperature at any x is defined using the solve book)
Note that instead of just finding Tm by typing Tm =, we have defined it as equal to a function Temp(x) within the
solve block. Advantage of doing this is that the same solve block repeatedly does the calculations for all values of x as
desired, taking each time the starting trial value of temperature as 250. This facility is a great advantage and is used
below while drawing the temperature profile.
Therefore, Temp(0.18) = 184.514 (temperature at x = 0.18 m)
i.e. T m = 184.514°C (temperature at mid-plane)

72 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Verify temperature at x = 0.08 m and x = 0.28 m:
Temp(0.08) = 400°C (verified)
Temp(0.28) = 150°C (verified)
Note that the mid-plane temperature obtained now matches the value obtained earlier.
To draw the temperature profile:
Define the range variable x from 0.08 m to 0.28 m at an interval of 0.01 m. Then, choose the x – y plot from the graph
palette and fill in the place holders on the x-axis and y-axis. Click anywhere outside the graph region and the graph
appears immediately: See Fig. Ex. 4.9(b)
x := 0.08, 0.09, ..., 0.28 (define range variable x, varying from 0.08 m to 0.28 m,
at interval of 0.01 m)

Temperature profile
(varying area & thermal conductivity)
400 Temperature in deg.C
and x in metres

300

Temp (x)

200

100
0.1 0.15 0.2 0.25
x

FIGURE Ex. 4.9(b)

Note: Verify from the graph that the temperature at the two ends and the mid-plane match with the values obtained
earlier.
Also, observe the ease with which the temperature profile is drawn—calculation of temperature at each value of x
and drawing the graph is done in one step; if this has to be done in hand calculation, to determine T at each x, one will
have to solve a quadratic equation for T at each point. Obviously, it is a very tedious job.
If k is a constant, how does heat transfer and temperature distribution change?
Let k = 18.82 W/(mC)
Again, since Q is constant through each section in the solid, we can directly integrate Fourier’s equation, keeping
the Q outside the integral. Integrating between two known temperatures at x = x 1 and x = x 2, we calculate Q.
Q = – kA(x) dT/dx

z z
x2 T2
dx
i.e. Q = -k dT
A (x )
x1 T1

In Mathcad, this is solved very easily:


From the above equation, we write for Q:

z
- k ×(T2 - T1 )
Q := x2 i.e. Q = 51.504 W (heat transfer rate through the section)
1
dx
x1 A ( x)
Note that this value of Q is the same as obtained earlier with variable k.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 73


Temperature profile:
Let the temperature at any x be T(x). Integrate between x = 0.08 m and x = x and use the condition that heat flow rate is
same (already calculated) through all sections.
We have:

z z
x2 T2
dx
Q = -k dT
A (x )
x1 T1

z z
x T (x )
dx
i.e. Q = -k dT = –k [T(x) – T 1]
A (x )
x1 T1

z
i.e.
x
Q 1
T(x) := T 1 – × dx (defines temperature as a function of x)
k x1 A (x )

Temperature at mid-plane:
Put x = 0.18 m in T(x):
T(0.18) = 181.55°C (temperature at midplane)
Note that with constant k, the mid-plane temperature is 181.55°C whereas with variable k, it was 184.514°C.
To draw the temperature profile:
Define the range variable x from 0.08 m to 0.28 m at an interval of 0.01 m. Then, choose the x – y plot from the graph
palette; type x in the place holder on the x-axis and Temp(x), T(x) in the place holder on the y-axis. Click anywhere
outside the graph region; graphs for temperature distribution with variable k as well as with constant k, appear
immediately: See Fig. Ex. 4.9(c)
x := 0.08, 0.09, ... , 0.28 (define range variable x, varying from 0.08 m to
0.28 m, at an interval of 0.01 m)

Temperature profile (varying & constant k)


400
Temperature in deg.C
and x in metres
350

Temp (x) 300


T(x)

250

200

150
0.1 0.15 0.2 0.25
Variable k x
Const. k

FIGURE Ex. 4.9(c)

It may be observed from the graph that with constant k assumption, temperature is lower throughout, as compared
to the case of variable k.

4.7 Cylindrical Systems


Cylindrical systems are practically important because of their common application in varied industries, power
plants, refineries, etc. Cylindrical geometry is popular for applications in heat exchangers, condensers, storage
tanks, etc.

74 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Here, we analyse the cylindrical shell for heat transfer in one-dimensional conduction, i.e. it is assumed that
temperature gradients are significant only in the radial direction; so, heat flow occurs only in the radial direction.
Now, it is important to remember that for a cylindrical system with heat flow in the radial direction, the area
normal to the direction of heat flow is not a constant, but varies with r; this was not the case for a plane slab,
where A remained constant for all x.
Consider a long cylinder of length L, inside radius ri and outside radius ro. Inner and outer surfaces are at
uniform temperatures of Ti and To , respectively, see Fig. 4.7.

Q
k
To

Q
Ti

Temperature profile,
Ti
L logarithmic

To
ri ri
Ti To
r0 Q Q r0
Rcy1 = ln(ro/ri)/(2pkL)
(a) Cylindrical system and the equivalent thermal circuit (b) Variation of temperature along the radius

FIGURE 4.7 Heat transfer through a cylindrical shell

Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) No internal heat generation.
Now, this is a cylindrical system; so, it is logical that we start with the general differential equation for one-
dimensional conduction, in cylindrical coordinates. So, we have, from Eq. 3.17:
1 ¶ FG ¶T IJ
1 ¶ 2T ¶ 2T q g 1 ¶T
r ¶r H
r
¶r
+ 2
K +
r ¶f 2 ¶z 2
+
k
=
a ¶t
In this case:
¶T/¶t = 0, since it is steady state conduction
¶T/¶f = 0 = ¶T/¶z, since it is one-dimensional conduction, in the r direction only
qg/k = 0, since there is no internal heat generation.
Therefore, the controlling differential equation for the cylindrical system, under the above mentioned
stipulations, becomes:
d 2T 1 ¶T
+ =0 ...(4.31)
dr 2 r ¶r
Note that now, it is not partial derivative, since there is only one variable, r.
We have to solve this differential equation to get the temperature distribution along r and then apply
Fourier’s law to calculate the heat flux at any position.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 75


Multiplying Eq. 4.31 by r, we get
d 2T ¶T
r + =0
dr 2 ¶r
d ¶T FG IJ = 0
i.e.
dr
r
¶r H K
Integrating,
¶T
r = C1
¶r
¶T C
or, = 1 ...(4.32)
¶r r
Integrating again,
T(r) = C1 ln(r) + C 2 ...(4.33)
where, C1 and C2 are constants of integration.
Eq. 4.33 gives the temperature distribution as a function of radius.
C 1 and C 2 are found out by applying the two B.C.’s:
(i) at r = ri, T = Ti
(ii) at r = ro, T = To
B.C. (i) gives, Ti = C1 ln (ri) + C 2 ...(a)
B.C. (ii) gives, To = C1 ln(ro) + C 2 ...(b)
Subtracting Eq. b from Eq. a:
Ti – To = C 1 ln(ri/ro)
Ti - To To - Ti
i.e. C1 =
Fr I
ln G J i
=
FG r IJ
o
Hr K o
ln
Hr K
i
and, from Eq. a:
To - Ti
C 2 = Ti –
FG r IJ ln(r )
o
i
ln
Hr K
i
Substituting C 1 and C 2 in Eq. 4.33, we get
To - Ti To - Ti
T(r) =
FG r IJ
o
ln(r) + T i –
FG r IJ ln(r )
o
i
ln
Hr K
i
ln
Hr K
i

To - Ti
ln G J
F rI
F r I Hr K
i.e. T(r) = Ti + ...(4.34)
ln G J o i
HrK i
Eq. 4.34 is the desired equation for temperature distribution along the radius. Note that the temperature
distribution is logarithmic for the cylindrical system, whereas it was linear for a plane slab. Temperature
distribution for the cylindrical system is sketched in Fig. 4.7(b).
Eq. 4.34 is written in non-dimensional form as follows:

FG r IJ
T (r ) - Ti Hr K
ln
i
To - Ti
=
Fr I
ln G J o
...(4.35)

Hr K i

76 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For thin cylinders, ri » ro, and then the temperature distribution within the shell is almost linear.
Next, to find the heat transfer rate, Q:
We apply the Fourier’s law.
Considering the inner radius ri,
dT C
Q = –k A r = –k 2p ri L 1 (using Eq. (4.32))
dr r = ri ri

To - Ti
i.e. Q = –k 2p ri L
ri ln
FG r IJ
HrK
o

2p kL(Ti - To )
i.e. Q=
Fr I ...(4.36)
ln GH r JK
o
i
Eq. 4.36 gives the desired expression for rate of heat transfer through the cylindrical system.
Note that Q is dependent on ln(ro/ri) rather than on (ro – ri). Implication of this is as follows:
For the same DT, k and L, heat transfer rate through a cylindrical shell of 5 cm ID and 10 cm OD is the same
as that through a shell of 10 cm ID and 20 cm OD, though in the first case, shell thickness is 5 cm and in the
second case, the shell thickness is 10 cm.
Now, writing Eq. 4.36 in a form analogous to Ohm’s law:
DT T - To
= i
Q=
Rcyl
ln o
r FG IJ
ri H K
2p kL
Immediately, we observe that thermal resistance for conduction for a cylindrical shell is given by,

FG r IJ
o

R cyl =
ln
Hr K
i
...(4.37) To
2p kL dr
Alternatively
For steady state, one-dimensional conduction, with no heat generation,
since the heat flow rate is the same and constant at every crosssection,
we can directly integrate the Fourier’s equation between the two known
temperatures (and the corresponding, known radii), keeping Q out of
the integral sign; this will give us Q. Then, at any r, the temperature T(r) Ti
is calculated by integrating between r = ri and r = r (with T = Ti and
T = T(r)), and equating the Q obtained now to the expression for Q
obtained earlier. Refer to Fig. 4.8.
At any radius r, consider a thin cylindrical shell of thickness dr; let
the temperature differential across this thin layer be dT. Then, in steady
state, rate of heat transfer through this layer Q, can be written from
Fourier’s law, to be equal to: ri

dT ro
Q = –kA r , where Ar = 2p rL
dr
FIGURE 4.8 Cylindrical system
dr
i.e. Q = – 2p kLdT ...(4.38)
r
Integrating Eq. 4.38 from ri to ro (with temperature from Ti to To),

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 77


F r I = –2 p kL(T
Q ln GH r JK
o
i
o – T i)

2p kL(Ti - To )
or, Q=
Fr I
ln G J o
(same as Eq. 4.36)

HrK i
To get the temperature profile within the cylindrical shell:
At any radius r, let the temperature be T(r).
Integrating Eq. 4.38 from ri to r,
2p kL (Ti - T ( r ))
Q=
F rI
ln G J
...(4.39)

HrK i
Now, apply the principle that in steady state, Q is the same through each layer, i.e. equate Eqs. 4.36 and 4.39:
2p kL(Ti - To ) 2p kL(Ti - T ( r ))
Fr I
ln G J
=
F rI
o
HrK
i
ln GH r JK
i

FG r IJ
T (r ) - Ti
ln
Hr K i
i.e.
To - Ti
=
F r I
ln G J o
...same as Eq. 4.35

Hr K i

Concept of log mean area. For a plane slab, we have the simple relation for the thermal resistance, i.e. R slab =
L/(kA), where L is the thickness of the slab, k is the thermal conductivity and A is the cross-sectional area normal
to the direction of heat flow.
Now, many times, it is desirable to write the thermal resistance of the cylindrical system also in a form
analogous to this form of R slab. Then, we take L as equivalent to (ro – ri), and k is the thermal conductivity, and let
Am be an equivalent area of the cylindrical system, which is to be found out. Then, we write:

FG r IJ
o

Rcyl =
ln
Hr Ki r -r
= o i
2p kL kAm

2p L ( ro - ri ) Ao - Ai Ao - Ai
Therefore, Am =
FG r IJ =
F 2p r I =
FG A IJ ...(4.40)
ln
HrK
o
i
ln GH 2 p r JK
o
i
ln o
HA K
i

where, Ao = 2p ro = outside surface area of the cylindrical shell


A i = 2p ri = inside surface area of the cylindrical shell
Am is known as the log mean area for the cylindrical system.
Note that physically, Am is the area of an equivalent slab having the same thermal conductivity k, whose
thickness is equal to the thickness of the cylindrical shell, and has the same heat flow rate as for the cylindrical
shell.
In practice, concept of log mean area is useful in analysing the lagging (i.e. insulation) of steam pipes.
If (Ao/Ai) < 2, then Am can be approximated to be the arithmetic average of Ao and Ai
i.e. A m = (Ao + Ai)/2
Example 4.10. A cylindrical insulation for a steam pipe has an inside radius ri = 6 cm, outside radius ro = 8 cm and k =
0.5 W/(mC). The inside surface of the insulation is at a temperature Ti = 430°C and the outside surface is at To = 30°C.
Determine:

78 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i) heat loss per metre length of this insulation k = 0.5 W/(mC)
(ii) temperature at mid-thickness of the insulation, and To = 30 C
(iii) draw the temperature profile along the radius.
Q
Solution. Data:
ri := 0.06 m r o := 0.08 m L := 1 m Ti :=
430°C To := 30°C k := 0.5 W/(mC)
Ti = 430 C
Heat transfer rate, Q:
Since it is a case of steady state, one-dimensional conduction,
with no internal heat generation, we can apply the thermal re-
sistance concept, i.e. Q = DT/Rcyl .
Therefore, ri = 0.06 m
Fr I
ln G J o

R cyl :=
HrK i
i.e. R cyl = 0.092 C/W
ro = 0.08 m
Ti To
2 ×p × k × L Q Q
(thermal resistance of the cylindrical shell) Rcyl = ln(ro/ri)/(2pkL)
Ti - To
Therefore, Q := FIGURE Example 4.10 Heat transfer through
Rcyl cylindrical insulation
i.e. Q = 4.368 ´ 10 3 W (heat transfer rate/m length.)
Temperature at mid-thickness of insulation, i.e. at r = 0.07 m:
Temperature distribution in the cylindrical shell is given by Eq. 4.35,

FG r IJ
T (r ) - Ti
ln
Hr K i
To - Ti
:=
F r I
ln G J o
Hr K i

FG r IJ
Hr K ln
i
i.e. T(r) := T + (T – T )×
i o
Fr I
i

ln G J o
(define temperature as a function of r)

Hr K i

i.e. T(0.07) = 215.665°C (temperature at mid-thickness of insulation)


Temperature profile in the insulation:
In Mathcad, this is very easy to draw. We already have expression for T(r), i.e. temperature as a function of r. Let us
define the range variable r, starting from r = 0.06 m to r = 0.08 m, varying in steps of 0.001 m. Then, from the graph
palette choose x – y graph, fill in r in the place holder of x-axis and T(r) in the place holder of y-axis. Click anywhere
outside the graph region and the graph appears immediately: See Fig. Ex. 4.10(b)
r = 0.06, 0.061, ..., 0.08 (define range variable, r; first value = 0.06, next
value = 0.061 m, last value = 0.08 m)
Obviously, temperature distribution is logarithmic, as seen from the expression for T(r).

4.8 Composite Cylinders


Composite cylinders belong to a practically important category, e.g. lagged pipes carrying steam or other high
temperature fluids, insulated pipes carrying coolant or cryogenic fluids, insulated tanks, etc. and, these are ana-
lysed as composite cylinders.
Consider a system of composite cylinders as shown in Fig. 4.9.
A cylinder of inner radius r1, outer radius r2 and thermal conductivity k1 is covered with another layer (say,
insulation) of radius r3 and thermal conductivity k2. There is perfect thermal contact at the interface between the
two layers, i.e. there is no temperature drop at the interface. Let T2 be the interface temperature. Further, let a hot

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 79


Temperature profile in cylindrical system
500

400

300
T(r) T(r) in deg.C
r in metres
200

100

0
0.06 0.065 0.07 0.075 0.08
r
FIGURE Ex. 4.10(b)

T2
k2 T3
k1 Q
T2
T3
Q
hb
hb Tb
T1
Tb ha
T1
Ta
ha
Ta
r1
r1 r2
r2
r3
r3 Ta Temperature profile
T1
Ta Tb T2
Q Q T3
Ra R1 R2 Rb Tb
(a) Composite cylinders and (b) Composite cylinders and
equivalent thermal circuit temperature profile

FIGURE 4.9 Composite cylinders, equivalent themal circuit and temperature profile

fluid at a temperature Ta flow through the inner pipe with a heat transfer coefficient ha. On the outside, let the
heat be lost to a cold fluid at a temperature Tb flowing with a heat transfer coefficient of h b. Let L be the length of
the cylindrical system.
Assumptions:
(i) Steady state heat flow
(ii) One-dimensional conduction in the r direction only
(iii) No internal heat generation
(iv) Perfect thermal contact between layers.

80 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Heat transfer occurs from inside to outside in the positive r direction, i.e. heat transfer occurs by convection
from the hot fluid at Ta to the inner wall of inner cylindrical layer at T1, then by conduction through the inner
cylindrical layer, again by conduction through the outer cylindrical layer and finally by convection from outer
wall of outer cylindrical layer at T3 to the cold fluid at Tb.
Under the given stipulations, it is clear that heat flow rate, Q through each layer is the same.
Let us write separately the heat flow equations for these 4 layers:
Convection from the hot fluid to inner wall at T 1:
Q = h a (2p r1)L (T a – T1), ... from Newton’s Law of Cooling
Q
i.e. (T a – T1) = = QRa ...(a)
ha ( 2 p r1L )
Conduction through first cylindrical layer:
2 p k1 L (T1 - T2 )
Q=
FG r IJ 2
Hr K
ln
1

Fr I
Q ln G J 2

i.e. (T 1 –T)=
2
H r K = QR
1
1 ...(b)
2 p k1 L
Conduction through second cylindrical layer:
2 p k2 L (T2 - T3 )
Q=
FG r IJ 3
Hr K
ln
2

Fr I
Q ln G J 3

i.e. (T 2 –T)=
3
H r K = QR
2
2 ...(c)
2 p k2 L
Convection from the outer wall at T3 to cold fluid at Tb:
Q = hb (2p r 3)L (T 3 – T b), (from Newton’s Law of Cooling)
Q
i.e. (T 3 – Tb) = = QR b ...(d)
hb ( 2 p r3 L )
Adding Eqs. a, b, c and d,
(Ta – Tb) = Q (Ra + R 1 + R 2 + Rb)
Ta - Tb T - Tb
i.e. Q= = a ...(4.41)
Ra + R1 + R2 + Rb R å
2 p L (Ta - Tb )
i.e. Q=
Fr I Fr I
ln G J ln G J
2 3
...(4.42)

1
+
1
+
Hr K + Hr K
1 2
ha r1 hb r3 k1 k2
If there are N concentric cylinders, we can write,
2 p L (Ta - Tb )
Q= N
FrN + 1 I ...(4.43)
1
+
1
ha r1 hb rN + 1
+ å 1
kN
lnGHrN JK
1

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 81


Basically, remember that in the composite cylindrical system just studied, the various resistances such as the
two convective resistances and the two conductive resistances are all in series. Then, by analogy with the rules of
electrical circuit, total thermal resistance is the sum of the individual resistances. Once these individual
resistances are identified and calculated, it is a simple matter to calculate the heat flow rate by analogy with
Ohm’s law, i.e. Q = DT/R total. Temperatures at the interfaces are calculated by using the fact that Q is the same
through each layer and by applying the analogy of Ohm’s law for each layer by turn.

4.9 Overall Heat Transfer Coefficient for the Cylindrical System


Referring to Fig. 4.9, it is clear that heat transfer occurs from hot fluid at Ta to the inner cylinder by convection,
then through the inner and outer cylindrical shells by conduction and then to the outer cold fluid at Tb by
convection. In many practical applications, the interface temperatures are not known but only the hot and cold
fluid temperatures, i.e. Ta and Tb are known. Here, (Ta – Tb) is the overall temperature difference because of which
heat flow occurs. We would rather like to write the heat transfer rate in terms of the known overall temperature
difference, as follows,
Q = UA DToverall = UA (Ta – Tb)
where U is an overall heat transfer coefficient and A is the area normal to the direction of heat flow. In the case of
a plane slab, A was a constant with x; however, in the case of a cylindrical system, area normal to the direction of
heat flow is 2 p r L, and clearly, this varies with r. Therefore, while dealing with cylindrical systems, we have to
specify as to which area U is based on, i.e. whether it is based on inside area or outside area. (Generally, U is
based on outside area since pipes are specified on outside diameters.)
We write,
Q = U i A i (Ta – Tb) = Uo Ao (Ta – Tb) ...(4.44)
where, Ui = overall heat transfer coefficient based on inside area
Uo = overall heat transfer coefficient based on outside area
A i = heat transfer area on inside
Ao = heat transfer area on outside
Comparing Eq. 4.44 with Eq. 4.41, we get,
Ta - Tb
Q= = U i A i (Ta – Tb) = Uo Ao (Ta – Tb) ...(4.45)
åR
1
i.e. U i A i = Uo Ao = ...(4.46)
åR
Therefore,
1
Ui = ...(4.47a)
Ai åR
1
Uo = ...(4.47b)
Ao åR
We can also write,

1 1
Ui =
Ai åR
=
LM FG r IJ ln FG r IJ
2 3
OP
2p r L ´ M
1
ln
Hr K + Hr K + 1 PP
MM 2p r L h
1 2
1 +

MN
1 a 2 p k1 L 2 p k2 L 2 p r3 L hb
PP
Q

82 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1
i.e. Ui =
LM 1 + F r I ln F r I + F r I ln F r I + F r I 1 OP ...(4.48a)

MN h GH k JK GH r JK GH k JK GH r JK GH r JK h PQ
1 2 1 3 1
a 1 1 2 2 3 b

And,

1 1
Uo =
Ao åR
=
LM FG r IJ ln FG r IJ
2 3
OP
2p r L ´ M
1
ln
Hr K + Hr K + 1 PP
MM 2p r L h
1 2
3 +

MN
1 a 2 p k1 L 2 p k2 L 2 p r3 L hb
PP
Q
1
i.e. Uo =
LMF r I 1 + F r I ln F r I + F r I ln F r I + 1 OP ...(4.48b)

MNGH r JK h GH k JK GH r JK GH k JK GH r JK h PQ
3 3 2 3 3
1 a 1 1 2 2 b

Note: Eqs. 4.48 a and 4.48 b give Ui and Uo in terms of the inside and outside radii. You need not memorise
them. To calculate Ui or Uo while solving numerical problems, just remember Eq. 4.46, namely:
1 Asbestos, k = 0.082 W/(mK)
Ui Ai = Uo Ao =
å R Magnesia, k = 0.07 W/(mK)
T2
Once the total thermal resistance SR is calculated, Ui or Q
Uo is easily found out from Eq. 4.46. T3 = 20 C
The concept of overall heat transfer coefficient in
cylindrical systems is often useful in heat exchanger designs, T1 = 195 C
since cylindrical geometry is a popular choice in heat
exchangers.
Example 4.11. A 10 cm OD pipe carrying saturated steam at a r1 = 0.05 m
temperature of 195°C is lagged to 20 cm diameter with magnesia
and further lagged with laminated asbestos to 25 cm diameter. r2 = 0.10 m
The entire pipe is further protected by a layer of canvas. If the
temperature under the canvas is 20°C, find the mass of steam r3 = 0.125 m
condensed in 8 hrs on a 100 m length of pipe. Take thermal con-
T1 T2 T3
ductivity of magnesia as 0.07 W/(mK) and that of asbestos as
Q Q
0.082 W/(mK). Neglect the thermal resistance of the pipe
material. Rmag Rasb
(M.U. Dec. 1997) FIGURE Example 4.11 Heat transfer in
Solution. lagged pipe and the equivalent thermal circuit
Data:
r 1 := 0.05 m r2 := 0.10 m r3 := 0.125 m L := 100 m T1 := 195°C T3 := 20°C
k asb := 0.082 W/mK kmag := 0.07 W/mK
Thermal resistances:

FG r IJ2

Rmag :=
ln
Hr K 1
(define thermal resistance of cylindrical magnesia shell)
2 ×p × kmag × L
i.e. R mag = 0.016 C/W (thermal resistance of magnesia)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 83


Fr I
ln GH r JK
3
2
R asb := (define thermal resistance of cylindrical asbestos shell)
2 ×p × kasb × L
i.e. R asb = 4.331 ´ 10 –3 C/W (thermal resistance of asbestos)
Rtot := Rasb + Rmag i.e. Rtot = 0.02 C/W (total thermal resistance)
Heat transfer rate, Q:
T1 - T3
Q := i.e. Q = 8.71 ´ 10 3 W (heat transfer rate)
Rtot
To calculate steam condensed in 8 hrs:
hfg := 1951 kJ/kg (latent heat of evaporation for steam at 195°C...from Steam Tables)
Therefore,

Q ´ 10- 3
m := × 3600.8
hfg
i.e. m = 128.581 kg (condensation of steam in 8 hrs.)
Note: If we need to calculate the interface temperature T 2 , apply the equivalent Ohm’s law, keeping in mind that heat
transfer rate through each layer is constant. For magnesia layer, wer can write: Q = (T1 – T 2)/Rmag.
To calculate interface temperature, T 2:
From Q = (T 1 – T 2)/Rmag, we get:
T 2 := T 1 – Q×R mag, C (temperature at the interface)
i.e. T 2 = 57.725°C (temperature at the interface)
Check: with reference to the asbestos layer, Q = (T 2 – T3)/R asb. Verify this:
Heat transfer through asbestos layer:
(T2 - T3 )
= 8.71 ´ 10 3 W (verified)
Rasb
Example 4.12. If in Example 4.11, there is a contact resistance of 0.02 (m2 K)/W between the pipe surface and magnesia,
and 0.05 (m2 K)/W at the interface between magnesia and asbestos, calculate the new value of heat transfer rate. Also,
calculate the temperature drops at the two interfaces.
Solution.
Data:
r 1 := 0.05 m r2 := 0.10 m r3 := 0.125 m L := 100 m T1 := 195°C T3 := 20°C k asb := 0.082 W/mK
kmag := 0.07 W/mK Rcont1 := 0.02 m2 K/W (contact resistance between pipe and magnesia)
R cont 2 := 0.05 m 2 K/W (contact resistance between magnesia and asbestos)
Equivalent thermal circuit and temperature profiles are shown in Fig. Example 4.12.
Thermal resistances:
Fr I
ln GH r JK
2

1
Rmag := (define thermal resistance of cylindrical magnesia shell)
2 × π × kmag × L
i.e. R mag = 0.016 C/W (thermal resistance of magnesia)
FG r IJ
Hr K
3
ln
2
Rasb := (define thermal resistance of asbestos shell)
2 × π × k asb × L
i.e. Rasb = 4.331 ´ 10 –3 C/W (thermal resistance of asbestos)

84 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T1 Tc1 T2 Tc2 T3 Asbestos,
Q Q k = 0.082 W/(mK)
Rc1 Rmag Rc2 Rasb Magnesia,
k = 0.07 W/(mK)
FIGURE Example 4.12(a) Equivalent thermal circuit T2
including contact resistances Q
T1 = 195 C
Contact resistances: T3 = 20 C
Between the pipe surface and magnesia, contact resistance
is given as Rcont1 = 0.02 (m2 K)/W; note that this resistance r1 = 0.05 m
is per m2 of surface. Actual surface area is (2p r1) L. There-
fore, contact resistance Rc1 = 0.02/(2 p r1 L), K/W. r2 = 0.10 m
Similarly, at the interface between magnesia and as- r3 = 0.125 m
bestos, contact resistance is given as 0.05 (m 2/K)/W and
the surface area at the interface is (2p r2 L) and therefore, DT1 = 5.176 C
contact resistance R c2 = 0.05/(2pr2 L), K/W.
Rcont 1
Rc1 := C/W (define contact DTmag = 128.139 C
2× p × r1 × L resistance between
pipe and magnesia)

i.e. Rc1 = 6.366 ´ 10 –4 C/W (contact resistance


between pipe and DT2 = 6.47 C
magnesia) DTasb = 35.215 C
Rcont 2
Rc2 := C/W (define contact resistance
2× p × r2 × L between magnesia and
FIGURE Example 4.12(b) Temperature profile in
asbestos)
the layers

i.e. Rc2 = 7.958 ´ 10 –4 C/W (contact resistance between magnesia and asbestos)
Therefore,
Rtotal = Rasb + Rmag + Rc1 + Rc2 C/W (total thermal resistance)
i.e. Rtotal = 0.022 C/W
Heat transfer rate, Q:
T1 - T3
Q := W (heat transfer rate)
Rtotal
i.e. Q = 8.131 ´ 10 3 W (heat transfer rate)
Note that as a result of including the thermal contact resistances, obviously the total resistance to heat flow
increases and the new value of heat transfer rate is reduced to 8131 W from the earlier value of 8710 W.
Temperature drops at the interface:
Let DT1 be the temperature drop at interface 1, i.e. between pipe surface and magnesia and DT2, the temperature drop at
interface 2, i.e. between magnesia and asbestos. From analogy with Ohm’s law, we have:
DT1 := Rc 1 ×Q°C (temperature drop at interface between pipe and magnesia)
i.e. DT1 = 5.176°C (temperature drop at interface between pipe and magnesia)
And,
DT2 := Rc 2×Q°C (temperature drop at interface between magnesia and asbestos)
i.e. DT2 = 6.47°C (temperature drop at interface between magnesia and asbestos)
Also,
DTmag := Rmag ×Q°C (temperature drop in magnesia layer)
i.e. DTmag = 128.139°C (temperature drop in magnesia layer)
DTasb := Rasb ×Q°C (temperature drop in asbestos layer)
i.e. DTasb = 35.215°C (temperature drop in asbestos layer)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 85


Check: Total temperature drop must be equal to (195 – 20) = 175°C. Verify this:
DTtotal = DT1 + DTmag + DT2 + DTasb (define total DT
i.e. DTtotal = 175°C (total temperature drop...verified.)
Example 4.13. A metal (k = 45 W/(mC)) steam pipe of 5 cm ID and 6.5 cm OD is lagged with 2.75 cm radial thickness of
high temperature insulation having thermal conductivity of 1.1 W/(mC). Convective heat transfer coefficients on the
inside and outside surfaces are h i = 4650 W/(m2 K) and h o = 11.5 W/(m2 K), respectively. If the steam temperature is
200°C and the ambient temperature is 25°C, calculate:
(i) heat loss per metre length of pipe
(ii) temperature at the interfaces, and
(iii) overall heat transfer coefficients referred to inside and outside surfaces (i.e. calculate Ui and Uo).
Solution. This is a case of steady state, one-dimensional heat transfer with no internal heat generation in any of the
layers. So, thermal resistance concept is applicable and is used to find out the rate of heat transfer, Q. Also, the principle
that Q is the same through each layer is used along with the equivalent Ohm’s law to determine the temperatures at the
interfaces. Overall heat transfer coefficients are determined by applying Eq. 4.46, namely,
Ui Ai = Uo Ao = 1/SR
See Figure Example 4.13.

Insulation, k = 1.1 W/(mK)


Pipe

T2
Q 2
hb = 11.5 W/(m C)
T1 Tb = 25 C
2
ha = 4650 W/(m C
T3
Ta = 200 C

r1 = 0.025 m Ta T1 T2 T3 Tb
Q Q
r2 = 0.0325 m Ra Rpipe Rins Rb
r3 = 0.06 m

FIGURE Example 4.13(a) Lagged steam pipe, FIGURE Example 4.13(b) Equivalent thermal circuit
with convection

Data:
r1 := 0.025 m r2 := 0.0325 m r3 := 0.06 m kpipe := 45 W/(mK) kins := 1.1 W/(mK) L := 1 m
Ta := 200°C Tb := 25°C hi := 4650 W/(m2 K) ho := 11.5 W/(m2K)
Thermal resistances:
Heat transfer occurs from the inside to outside, as shown in Fig. Example 4.14. Starting from inside, first there is convec-
tive resistance between steam and the pipe surface, then conductive resistances through the pipe material and insulation,
then again, convective resistance between the outer surface and the ambient. Let us calculate these resistances, by turn:
Fr I
ln GH r JK
2

1
Rpipe := , C/W (define thermal resistance of pipe)
2 × p × kpipe × L

i.e. R pipe = 9.279 ´ 10 –4 C/W (thermal resistance of pipe)


FG r IJ
Hr K
3
ln
2
Rins := , C/W (define thermal resistance of insulation)
2 × p × kins × L

86 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. R ins = 0.089 C/W (thermal resistance of insulation)
1
Ra := , C/W (define convective resistances on the inside surface)
hi 2 ×p × r1 × L
i.e. R a = 1.369 ´ 10 –3 C/W (convective resistance on the inside surface)
1
R b := , C/W (define convective resistance on the outside surface)
ho 2 × p × k 3 × L
i.e. Rb = 0.231 C/W (convective resistance on the outside surface)
Therefore, total resistance:
Rtot := Rpipe + R ins + Ra + R b, C/W (since all resistances are in series)
i.e. R tot = 0.322 C/W (total thermal resistance)
Heat transfer rate, Q is equal to overall temperature difference divided by total thermal resistance, by analogy with
Ohm's law
Ta - Tb
Q := ,W (define Q.)
Rtot
i.e. Q = 544.046 W (heat transfer rate)
Temperatures at the interfaces:
For each layer, Q is the same and is equal to the temperature drop through that layer divided by the thermal resistance
of that particular layer. Apply this for the inside convective layer, the two conductive layers through pipe and
insulation, and then again, the convective layer on the outside, by turn:
(Ta – T1) = Q.R a
(T1 – T2) = Q.Rpipe
(T2 – T3) = Q.Rins
(T3 – Ta) = Q.Rb
We get:
T1 := Ta – Q×R a°C (define T1)
i.e. T1 = 199.255°C (temperature of inside surface of pipe)
T2 := T1 – Q×R pipe°C (define T2)
i.e. T2 = 198.75°C (temperature of interface of pipe and insulation)
T3 := T2 – Q×R ins°C (define T3)
i.e. T3 = 150.489°C (temperature of outer surface of insulation)
Finally, check for value of Tb:
Tb := T3 – Q×R b°C (define Tb)
i.e. Tb = 25°C (matches with the data...verified)
Overall heat transfer coefficients, Ui and Uo :
Overall heat transfer coefficients are determined by applying Eq. (4.46), namely, Ui Ai = Uo Ao = 1/S R: Remember:
Ai = 2p r1.L and Ao = 2 pr 3.L
1
Ui := W/(m2 C) (define Ui)
( 2 × p × r1 × L) × Rtot
i.e. Ui = 19.791 W/(m2 C) (overall heat transfer coefficient based on inside area)
And,
1
Uo := W/(m2 C) (define U o)
( 2 × p × r3 × L) × Rtot
i.e. Uo = 8.246 W/(m2 C) (overall heat transfer coefficient based on outside area)
Example 4.14. A 160 mm dia pipe carrying saturated steam is covered by a layer of lagging of thickness 40 mm (k = 0.8
W/(mC)). Later, an extra layer of lagging of 10 mm thickness (k = 0.12 W/(mC)) is added. If the surrounding tempera-
ture remains constant and heat transfer coefficient for both lagging materials is 10 W/(m2K), determine the percentage
change in rate of heat loss due to the extra lagging layer.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 87


Lagging 2, k2 = 1.2 W/(mK)
Lagging, k1 = 0.8 W/(mK)
Pipe Lagging 1, k1 = 0.8 W/(mK)
Pipe
2
ha = 10 W/(m C) 2
ha = 10 W/(m C)

r1 = 0.08 m
r1 = 0.08 m
r2 = 0.12 m r2 = 0.12 m
r3 = 0.13 m
DT DT

Q Q Q Q
R1 Ra R1 R2 Ra

FIGURE Example 4.14(a) Pipe with one FIGURE Example 4.14(b) Pipe with two
layer of lagging layer of lagging

Solution. See Fig. Example. 4.14.


Data:
r 1 := 0.08 m r2 := 0.12 m r3 := 0.13 m k1 := 0.8 W/(mC) k2 := 0.12 W/(mC)
ha := 10 W/(m2 C) L := 1 m
Since this is a case of steady state, one-dimensional conduction with no internal heat generation, thermal resistance
concept is applicable.
In case (i): Thermal resistance is the sum of conduction resistance in lagging layer number 1 and convective resistance
over its surface. Conduction resistance of the pipe material and the convective resistance between steam and inner
surface of pipe are neglected, since no data is given. See Fig. Example. 4.14a.
In case (ii): Thermal resistance is the sum of conduction resistances in lagging layers number 1 and number 2 and the
convective resistance over the surface of lagging layer number 2.
Obviously, Rtotal for case (ii) is more than that for case (i); accordingly, heat transfer rate for the second case, Q 2 is
less than that for first case, Q 1.
From analogy with Ohm’s law, we write:
Q 1 = DT/R tot1 and Q 2 = DT/Rtot2 where DT is the overall temperature difference, which is the same for both cases.
Therefore, (Q 2/Q 1) = (Rtot1/R tot2).
And, % change in heat flow rate = (Q 1 – Q 2) ´ 100/Q 1 = [1 – (Q 2/Q 1)] ´ 100
Thermal resistances:
FG r IJ
Hr K
2
ln
1
R1 := C/W (define thermal resistance of lagging layer 1)
2 × p × k1 × L
i.e. R1 = 0.081 C/W (thermal resistance of lagging layer 1)
FG r IJ
Hr K
3
ln
2
R2 := C/W (define thermal resistance of lagging layer 2)
2 × p × k2 × L
i.e. R 2 = 0.106 C/W (thermal resistance of lagging layer 2)
1
R a1 := C/W (define convective resistance over surface of lagging)
ha × ( 2 × p × r2 × L)

88 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Ra 1 = 0.133 C/W (convective resistance over surface of lagging layer 1)
1
R a2 := C/W (define convective resistance over surface of lagging)
ha × ( 2 × p × r3 × L)
i.e. Ra 2 = 0.122 C/W (convective resistance over surface of lagging layer 2)
Total resistances:
Case (i): with lagging layer 1 only:
R tot 1 := R1 + Ra1 i.e. Rtot 1 = 0.213 C/W (total resistance for case (i))
Case (ii): with lagging layer 1 and 2:
R tot 2 := R1 + R2 + Ra2 i.e. R tot 2 = 0.309 C/W (total resistance for case (ii))
Percentage change in heat transfer rate:
First, find out (Q 2/Q 1) from: (Q 2/Q 1) = (Rtot1/R tot 2)
Rtot 1
Q 2 by Q 1 := (define Q 2 /Q 1)
Rtot 2
Therefore, Q 2 by Q 1 = 0.6897 (value of Q 2 /Q 1)
And,
Per cent change := (1 – Q 2 by Q 1)×100 (define % change)
i.e. Per cent change = 31.029 i.e. change in heat transfer rate is 31.029%.
Example 4.15. A 3.3 cm OD steel pipe, outside surface of which is at 500 K, is surrounded by still air at 300 K. The heat
transfer coefficient by natural convection is 10 W/(m2 K). It is proposed to reduce the heat loss to half by applying
magnesia insulation (k = 0.07 W/(mK) on the outside surface of the pipe. Determine the thickness of the insulation.
Assume pipe surface temperature and convective heat transfer coefficients remain the same.
Solution. Thermal resistance concept is applicable since it is a case of steady state, one-dimensional conduction, with no
internal heat generation.
There are two cases:
Case (i): Without insulation, i.e. bare pipe–now, the heat transfer occurs only by natural convection on the pipe surface
and the heat transfer rate, Q1 is given by Newton’s Law of Cooling, namely, Q1 = h a (2 pr1. L).DT, or, Q 1 = DT/Ra1, where
Ra1 is the convective resistance and DT = (500 – 300) deg.
Case (ii): With insulation: Now, the heat transfer rate, Q2 is given to be half of Q1. Thermal resistances involved are: the
conductive resistance of the cylindrical insulation layer (= R1) and the convective resistance over the insulation surface
(= Ra 2).
i.e. Q 2 = DT/(R1 + Ra2). Write the expression for Q2 and solve the resulting transcendental equation by trial and error to
get the outer radius of insulation.
Situations of case (i) and (ii) are depicted in Fig. Example. 4.15(a) and (b).
Data:
r 1 := 0.0165 m ha := 10 W/(m2 K) kins := 0.07 W/(mC) T1 := 500 K T2 := 300 K L := 1 m
Let rins be the outer radius of insulation
Case (i): bare pipe:
1
Ra1 := C/W (convective thermal resistance on bare pipe)
ha × ( 2 × p × r1 × L)
i.e. Ra1 = 0.965 C/W (convective thermal resistance on bare pipe)
Therefore,
T1 - T2
Q 1 := W (define heat transfer rate, Q 1)
Ra 1
i.e. Q 1 = 207.345 W (heat transfer rate for bare pipe)
Case (ii): pipe with insulation:
Q1
Now, Q 2 := i.e. Q 2 = 103.673 W (heat transfer rate, with insulation)
2
We have: Q 2 = DT/SR,
i.e. Q 2 = DT/(R1 + Ra2)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 89


Pipe Insulation, kins = 0.07 W/(mK)

Pipe Q2 = Q1/2

Q1 2 2
ha = 10 W/(m C) ha = 10 W/(m C)

r1 = 0.0165 m

r1 = 0.0165 m

rins = ?
DT DT

Q1 Q1 Q2 Q2
Ra1 R1 Ra2

FIGURE Example 4.15(a) Pipe without insulation FIGURE Example 4.15(b) Pipe with insulation

So, we get:
DT T1 - T2 500 - 300
Q2 =
å
=
F r IJ
ln G
=
F r I
ln G
...(a)

H 0.0165 JK
R
Hr K+
ins ins

1 1 1
+
2 p kins L ha (2 p rins L) 2 ´ p ´ 0 .07 ´ 1 10 ´ ( 2 ´ p ´ rins ´ 1)
Simplifying, we get:
F r I + 0.015915 = 1.92914
2.27366 ´ ln GH 0.0165 JK
ins

rins
This equation has to be solved to get rins ; and, trial and error solution is required since it is a transcendental
equation. Solve it by hand, as an exercise.
However, it is easily solved in Mathcad, using solve block. Start with a trial value of r ins and write the constraint
(i.e. Eq. (a)) immediately below ‘Given’; then the command Find(r ins ) gives the value of r ins: Note that you need not even
simplify Eq. (a).
r ins := 0.05 m (trial value of r ins)
Given
T1 - T2
Q2 =
F r IJ
ln G
HrK
ins

1 1
+
2 ×p × kins × L ha × 2 ×p × rins × L
Find (r ins) = 0.030683
i.e. r ins := 0.0307 m (outer radius of insulation)
Therefore, thickness of insulation:
tins := r ins – r 1 m (define thickness of insulation)
i.e. tins = 0.014 m (thickness of insulation)

90 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4.10 Spherical Systems
Spherical system is one of the most commonly used geometries in industry. It finds its applications as storage
tanks, reactors, etc. in petrochemical, refineries and cryogenic industries. Sphere has minimum surface area for a
given volume and material requirement to manufacture a sphere is minimum compared to other geometries.
Here, let us analyse the spherical shell for heat transfer in one-dimensional conduction, i.e. it is assumed that
temperature gradients are significant only in the radial direction; so, heat flow occurs only in the radial direction.
Now, here also, as in the case of a cylindrical system, the area normal to the direction of heat flow is not a
constant, but varies with r.
Consider a spherical shell, inside radius ri and outside radius ro. Inner and outer surfaces are at uniform
temperatures of Ti and To , respectively. See Fig. 4.10.

k To Q Q

Ti

Temperature profile,
Ti hyperbolic

ri
To
r0 ri
Ti To
r0
Q Q
Rsph = (ro – ri)/(4pkrori)

FIGURE 4.10(a) Spherical system and the FIGURE 4.10(b) Variation of temperature
equivalent thermal circuit along the radius

Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) No internal heat generation.
Now, since we are considering a spherical system, it is logical that we adopt spherical coordinates. General
differential equation for conduction in spherical coordinates is given by Eq. 3.21. For the above mentioned
assumptions, Eq. 3.21 reduces to:
d 2T 2 dT
+ =0 ...(4.49)
dr2 r dr
Note that now, it is not partial derivative, since there is only one variable, r.
We have to solve this differential equation to get the temperature distribution along r and then apply
Fourier’s law to calculate the heat flux at any position.
Multiplying Eq. 4.49 by r2,
d 2T dT
r2 + 2r =0
dr 2 dr
d FG r dT IJ
H K
2
i.e. =0
dr dr

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 91


Integrating,
dT
r2 = C1
dr
dT C
or, = 21 ...(4.50)
dr r
Integrating again,
C1
T(r) =– + C2 ...(4.51)
r
where, C 1 and C 2 are constants of integration.
Eq. 4.51 gives the temperature distribution in the spherical shell as a function of radius.
C 1 and C 2 are found out by applying the two B.C.’s:
(i) at r = ri, T = Ti
(ii) at r = ro, T = To
B.C. (i) gives: Ti = – C1/ri + C2 ...(a)
B.C. (ii) gives: To = – C1/ro + C2 ...(b)
Subtracting Eq. b from Eq. a:
Ti – To = C1 .[(1/ro) – (1/ri)]
Ti - To
i.e. C1 =
1 1
-
ro ri
and, from Eq. a:

LM OP
C 2 = Ti + MM
1 Ti - To
ri 1 - 1
PP
ro riMN PQ
Substituting C1 and C 2 in Eq. 4.51, we get

LM O
T - T P F 1 1I
T(r) = T – M
MM 1 - 1 PPP ´ GH r - r JK
i o
i ...(4.52)
i
Nr r Q
o i

Eq. 4.52 is the desired equation for temperature distribution along the radius.
Eq. 4.52 is written in non-dimensional form as follows:
1 1
T (r ) - Ti
-
r ri r F r-r I
To - Ti
=
1 1
-
= o ´
r GH r - r JK
o
i
i
...(4.53)

ro ri
Temperature distribution for the spherical system is shown in Fig. 4.10 (b). Note that the temperature
distribution is a hyperbola.
Next, to find the heat transfer rate, Q:
We apply the Fourier’s law. Since it is steady state conduction, with no heat generation, Q is the same through
each layer.
Considering the outer surface, i.e. at r = ro
dT Ti - To 1
Q = – kAr = –k ´ 4p ro2 ´ ´ 2
dr r = ro
1 1 r
- o
ro ri
92 FUNDAMENTALS OF HEAT AND MASS TRANSFER
4 p k (Ti - To ) 4 π k ri ro (Ti - To )
i.e. Q= = ...(4.54)
1 1 ro - ri
-
ri ro
Eq. 4.54 gives the desired expression for rate of heat transfer through the spherical system.
Now, writing Eq. 4.54 in a form analogous to Ohm’s law:
DT T - To
Q= = i
Rsph ro - ri
4p kro ri
Immediately, we observe that thermal resistance for conduction for a spherical shell is given by:

Rsph =
ro - ri
=
1 1 1 LM
-
OP ...(4.55)
4p kro ri N
4 p k ri ro Q
Alternatively
Since it is steady state, one-dimensional conduction, with no heat generation, heat flow rate, Q is constant at
every cross-section; so, we can directly integrate Fourier’s equation between the two known temperatures (and
the corresponding, known radii), keeping Q out of the integral sign; this will give us Q. Then, at any r, the
temperature T(r) is calculated by integrating between r = ri and r = r (with
T = Ti and T = T(r)), and equating the Q obtained now to the expression for To
Q obtained earlier. dr
Refer to Fig. 4.11.
At any radius r, consider an elemental spherical shell of thickness dr;
let the temperature differential across this thin layer be dT. Then, in steady
state, rate of heat transfer through this layer Q, can be written from
Fourier’s law, to be equal to:
Ti
dT
Q = –kAr , where Ar = 4 p r 2
dr
dr
i.e. Q 2 = – 4pkdT ...(4.56)
r
Integrating Eq. 4.56 from ri to ro (with temperature from Ti to To),

z z
ro To ri
dr
Q = – 4pk dT
r2 ro
ri Ti

Q
LM - 1OP ro
= 4pk(Ti – To)
FIGURE 4.11 Spherical system
i.e.
NrQ ri

i.e. Q
LM 1 - 1 OP = 4pk(T – T )
Nr r Q
i o
i o

4pk (ri ro )(Ti - To )


or, Q= ...same as Eq. 4.54
ro - ri
To get the temperature profile within the spherical shell:
At any radius r, let the temperature be T(r).
Integrating Eq. 4.56 from ri to r, i.e. replace ro by r and To by T(r) in Eq. 4.54,
4p kri r (Ti - T (r ))
Q= ...(4.57)
r - ri
Now, apply the principle that Q is the same through each layer, i.e. equate Eqs. 4.54 and 4.57:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 93


4pk (ri ro )(Ti - To ) 4pkri r (Ti - T ( r ))
=
ro - ri r - ri

T (r ) - Ti r F
r - ri I
i.e.
To - Ti
= o ´
r GH
ro - ri JK (same as Eq. 4.53)

Concept of “geometric mean area” for a spherical system:


As in the case of the cylindrical system, it is convenient to think of an equivalent slab for the spherical system, i.e.
we would like to express the thermal resistance of the sphere in the form of the thermal resistance for a slab (i.e.
R = L/(kA)). If we define L as the thickness of the spherical shell, i.e. Lsph = (ro – ri), we can write from Eq. 4.55:
ro - ri ro - ri
R sph = =
4p kro ri k ( 4 p ro2 )( 4 p ri2 )

ro - ri Lsph
i.e. R sph = = ...(4.58)
k Ao Ai k Ag
i.e. thermal resistance of spherical shell, R sph is expressed in a form analogous to that of a plane slab. Here, the
equivalent area, Ag = Ao Ai , is known as “geometric mean area”. It represents the area of an equivalent slab of
the same material, with a thickness equal to that of the spherical shell and transfers the same amount of heat per
unit time under the same temperature potential as for the spherical shell. Note that that Ao and Ai are the outer
and inner surface areas, respectively, of the spherical shell.
Note: It is very common that containers have hemispherical ends. Then, remember that thermal resistance of a
hemispherical spherical shell is half that of a spherical shell as given by Eq. 4.55.
Example 4.16. Consider an aluminium hollow sphere of inside radius r i = 2 cm, outside radius ro = 6 cm and k = 200 W/
(mC). The inside surface is kept at an uniform temperature of Ti = 100°C and outside surface dissipates heat by
convection with h = 80 W/(m2 C) into ambient air at a temperature of Ta = 20°C. Determine:
(i) outside surface temperature of the sphere in steady state
(ii) rate of heat transfer
(iii) temperature within the aluminium sphere at a radius r = 3 cm
(iv) sketch the temperature distribution along the radius.
Solution. See Fig. Example. 4.16.
Hollow sphere Data:
k = 200 W/(mC) ri := 0.02 m ro := 0.06 m k := 200 W/(mC)
Q T := 100°C T := 20°C h := 80 W/(m2 C)
To i a
2 Let To be the temperature of outside surface of spherical
ha = 80 W/(m C)
shell.
Ta = 20 C Since it is a case of steady state, one-dimensional heat
Ti = 100 C transfer with no internal heat generation, thermal resistance
concept is applicable.
Heat transfer through the sphere is by conduction and
then, from the outer surface of the sphere to ambient, it is by
convection. So, calculate these resistances, i.e. Rsph is given by
ri = 0.02 m Eq. 4.55 and convective resistance, Ra = 1/(h.Ao), where Ao =
(4 p r o2) is the outer surface area of the spherical shell. Now,
ro = 0.06 m
apply the equivalent Ohm’s law, i.e. Q = DT/R tot to get the
heat transfer rate, Q. See Fig. Example 4.16 for equivalent
DT = Ti Ta thermal circuit.
Ti To Ta Thermal resistances:
Q Q
ro - ri
Rsph Ra Rsph := C/W
4 × p × k × ro × ri
FIGURE Example 4.16 Hollow sphere (define the thermal resistance of spherical shell)
with convection

94 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Rsph = 0.013 C/W (thermal resistance of spherical shell)
1
Ra := C/W (define convective thermal resistance on the outer
h × ( 4 × p × ro2 ) surface of spherical shell)

i.e. Ra = 0.276 C/W (convective thermal resistance on the outer surface


of spherical shell)
Total thermal resistances:
Therefore, Rtot := Rsph + Ra C/W (define total thermal resistance)
i.e. Rtot = 0.29 C/W (total thermal resistance)
Heat thermal rate, Q:
Apply: Q = DT/R tot
Ti - Ta
i.e. Q := ,W (define Q)
Rtot
i.e. Q = 276.268 W (rate of heat transfer)
Temperature of outer surface of spherical shell:
Apply the equivalent Ohm’s law only to the convective layer, remembering that Q is same through each layer in steady
state, i.e. Q = (To – Ta)/Ra . Therefore,
To = Ta + Q.Ra
We have: To := Ta + Q.Ra , (define To)
i.e. To = 96.336 C (temperature of outer surface of shell)
Verify with reference to the spherical shell:
i.e. (Ti – To)/Rsph must be equal to Q:
Ti - To
We get: = 276.268 W (verified)
Rsph
Temperature at a radius of r = 3 cm:
Temperature distribution along the radius is given by Eq. 4.53, namely,
1 1
T (r) - Ti
-
r ri r F r-r I
To - Ti
=
1 1
-
= o ´
r GH r - r JK
o
i

i
...(4.58)
ro ri
Therefore, we get:
LM r × F r - r I OP °C
MN r GH r - r JK PQ
o i
T(r): = T i + (To – Ti)× (define temperature as a function of radius)
o i

Now, substitute r = 0.03 in T(r) to get temperature at that radius:


i.e. T(0.03) = 98.168°C (temperature at a radius of 0.03 m)
Temperature profile along the radius:
This is drawn very easily in Mathcad. First, define a range variable r varying from inner radius to outer radius, i.e. from
0.02 m to 0.06 m, say at an interval of 0.001 m. Then, choose the x – y graph from the graph palette. Fill in the place
holders on the x-axis and y-axis with r and T(r), respectively. Click anywhere outside the graph region and ilmmediately
the graph appears: See Fig. Ex. 4.16(b)
r := 0.02, 0.021, ... , 0.06 ...define range variable r; starting value of r = 0.02,
next value = 0.021 and last value of r = 0.06 m
Note: verify from the graph that temperature at r = 3 cm is, indeed, 98.168 deg.C.

4.11 Composite Spheres


Assumptions:
(i) Steady state heat flow
(ii) One-dimensional conduction in the r direction only

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 95


Temperature profile in spherical shell
100 Temperature in deg.C and
radius in metres

99

T(r)
98

97

96
0.01 0.02 0.03 0.04 0.05 0.06
r
FIGURE Ex. 4.16(b)

k2 T2
k1 T3
T2 Q
T3
Q hb
hb Tb
T1
Tb
T1 hb
Ta
ha
Ta
r1
r1
r2
r2 Temperature profile
r3
r3 Ta
T1
Ta Tb
T2
Q Q T3
Ra R1 R2 Rb Tb

FIGURE 4.12(a) Composite spheres FIGURE 4.12(b) Composite spheres


and equivalent thermal circuit and temperature profile

(iii) No internal heat generation


(iv) Perfect thermal contact between layers.
Consider a system of composite cylinders as shown in Fig. 4.12.
A hollow sphere of inner radius r1, outer radius r2 and thermal conductivity k 1 is covered with another layer
(say, insulation) of radius r 3 and thermal conductivity k 2. There is perfect thermal contact at the interface between
the two layers, i.e. there is no temperature drop at the interface. Let T2 be the interface temperature. Further, let a
hot fluid at a temperature Ta transfer heat to the inner sphere with a heat transfer coefficient ha. On the outside, let
the heat be lost from the surface at a temperature of T3 to a cold fluid at a temperature Tb flowing with a heat
transfer coefficient of hb .
Under the given stipulations, it is clear that heat flow rate, Q through each layer is the same. Let us write
separately the heat flow equations for the 4 layers:

96 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Convection from the hot fluid to inner wall at T 1:
Q = ha (4 p r12)(Ta – T1), (from Newton’s Law of Cooling)
Q
i.e. (Ta – T1) = = QR a ...(a)
ha ( 4 p r12)
Conduction through first spherical layer:
4 p k1 r2 r1 (T1 - T2 )
Q=
r2 - r1
Q (r2 - r1 )
i.e. (T1 – T2) = = QR 1 ...(b)
4 p k1 r2 r1
Conduction through second spherical layer:
4 p k1 r3 r2 (T2 - T3 )
Q=
r3 - r2
Q (r3 - r2 )
i.e. (T2 – T3) = = QR 2 ...(c)
4 p k2 r3 r2
Convection from the outer wall at T 3 to cold fluid at Tb:
Q = hb (4p r32)(T 3 – Tb), ( from Newton’s Law of Cooling)
Q
i.e. (T3 – Tb) = = QR b ...(d)
hb ( 4 p r 23)
Adding Eqs. a, b, c and d:
(Ta – Tb) = Q (Ra + R 1 + R 2 + Rb)
Ta - Tb T - Tb
i.e. Q= = a ...(4.59)
Ra + R1 + R2 + Rb R å
4 p (Ta - Tb )
i.e. Q= ...(4.60)
1 1 r -r r -r
+ + 2 1+ 3 2
ha r12 hb r32 k1 r1 r2 k2 r3 r2
If there are N concentric spheres, we can write:
4 p (Ta - Tb )
Q= N
Fr - rN I ...(4.61)
1
+
1
ha r12 hb rN2 + 1
+ å GH k
1
N +1

N rN rN + 1
JK
Basically, remember that in the composite spherical system just studied, the various resistances such as the
two convective resistances and the two conductive resistances are all in series. Then, by analogy with the rules of
electrical circuit, total thermal resistance is the sum of the individual resistances. Once these individual
resistances are identified and calculated, it is a simple matter to calculate the heat flow rate by analogy with
Ohm’s law, i.e. Q = DT/R total. Temperatures at the interfaces are calculated by using the fact that Q is the same
through each layer and by applying the analogy of Ohm’s law for each layer by turn.

4.12 Overall Heat Transfer Coefficient for the Spherical System


As in the case of cylindrical systems, we define an overall heat transfer coefficient for the spherical systems also.
Again, in this case too, the area normal to the direction of heat flow varies with the radius and it is necessary to
specify as to on which area the overall heat transfer coefficient is based.
Accordingly, we write:
Q = U i A i (Ta – Tb) = Uo Ao (Ta – Tb)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 97


where, Ui = overall heat transfer coefficient based on inside area
Uo = overall heat transfer coefficient based on outside area
A i = heat transfer area on inside
Ao = heat transfer area on outside
Therefore, we get:
Ta - Tb
Q= = U i A i (Ta – Tb) = Uo Ao (Ta – Tb)
åR
1
i.e. Ui Ai = Uo A o =
åR
Therefore,
1
Ui = and
Ai åR
1
Uo =
Ao åR
We can also write:
1 1
Ui =
Ai å R
=
4 p r12
L 1
´M +
1 r -r
+ 2 1 + 3 2
r -r OP
MN 4 p h r a 1
2 2
4 p hb r3 4 p k r r
1 1 2 4 p k2 r3 r2 PQ
1
i.e. Ui =
LM 1 1 F r I 2
r (r - r ) r 2 ( r - r ) OP ...(4.62 a)

MN h + h ´ GH r JK + 1 2 1 + 1 3 2
PQ
1
a b 3 k1 r2 k2 r3 r2

And,
1 1
Uo =
Ao åR
=
4 p r32 ´
LM 1 +
1 r -r
+ 2 1 + 3 2
r -r OP
MN 4 p h ra 1
2
4 p hb r32 4 p k1 r1 r2 4 p k2 r3 r2 PQ
1
i.e. Uo =
LM 1 F r I 2
r32 ( r2 - r1 ) r3 (r3 - r2 ) OP ...(4.62 b)

MN h ´ GH r JK
3 1
a 1
+
hb
+
k1 r1 r2
+
k2 r2 PQ
Note: Eqs. 4.62 a and 4.62 b give U i and Uo in terms of the inside and outside radii. You need not memorise
them. To calculate U i or Uo while solving numerical problems, just remember Eq. 4.46, namely
1
Ui Ai = Uo Ao = ...(4.46)
åR
Example 4.17. A spherical vessel containing hot fluid at 160°C is of 0.4 m OD and is made of titanium of 25 mm
thickness. The thermal conductivity is 20 W/(mK). The vessel is insulated with two layers of 5 cm thick insulations of
thermal conductivities 0.06 and 0.12 W/(mK). There is a contact resistance of 6 ´ 10 –4 and 5 ´ 10 –4 (m2 C)/W between the
metal and first insulation and between the insulating layers. The outside is exposed to surrounding at 30°C with a
convection coefficient of 15 W/(m2 K). Determine the rate of heat loss, the interface temperatures and the overall heat
transfer coefficients based on inside surface area as well as outside surface area (i.e. calculate U i and U o).
Solution. See Fig. Ex. 4.17(a) & (b).

98 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T1 T2 T2 T3 T3 T4 Ta
Q Q
RTit Rc1 R1 Rc2 R2 Ra

FIGURE Example 4.17(a) Equivalent thermal circuit including contact resistances

Data:
r 1 := 0.175 m r2 := 0.20 m r3 := 0.25 m r4 := 0.30 m kTit := 20 W/(mK)
k 1 := 0.06 W/(mK) k2 := 0.12 W/(mK) T1 := 160°C Ta := 30°C
ha := 15 W/(m 2 K) Rcont1 := 6 ´ 10– 4 m2 C/W Rcont2 := 5 ´ 10 –4 m2 C/W
Thermal resistances:
Thermal resistance network is shown in Fig. Example 4.17(a).
Conductive resistances:
r2 - r1
R1 := (define thermal resistance of spherical titanium shell)
4 ×p × kTit × r2 × r1
i.e. R 1 = 2.842 ´ 10 –3 C/W (thermal resistance of spherical titanium shell)
r3 - r2
R2 := (define thermal resistance of first insulation shell)
4 × p × k1 × r3 × r2
i.e. R 2 = 1.326 C/W (thermal resistance of first insulation shell)
r4 - r3
R3 := (define thermal resistance of second insulation shell)
4 × p × k2 × r4 × r3
i.e. R 3 = 0.442 C/W (thermal resistance of second insulation shell)
Convective resistances:
1
Ra := C/W (define the convective resistance between outer insulation surface and the ambient air)
ha × ( 4 ×p × r42 )
i.e. Ra = 0.059 C/W (convective resistance between outer insulation surface and the ambient air)
Contact resistances:
Between the titanium surface and first insulation, contact resistance is given as Rcont1 = 6 ´ 10 – 4 (m2 C)/W; note that this
resistance is per m2 of surface. Actual surface area is (4 p r 22). Therefore, contact resistance R c1 = 6 ´ 10 –4/(4 p r 22), C/W
Similarly, at the interface between the two insulation layers, contact resistance is given as 5 ´ 10 –4 (m2 C)/W and the
surface area at the interface is (4 p r 32) and therefore, contact resistance R c 2 = 5 ´ 10– 4/(4 p r 32), C/W
Rcont 1
R c1 := C/W (define contact resistance between titanium and first layer of insulation)
( 4 × p × r22 )
i.e. Rc 1 = 1.194 ´ 10 –3 C/W (contact resistance between titanium and first of insulation)
Rcont 2
R c2 := C/W (define contact resistance between the two layers of insulation)
( 4 × p × r32 )
i.e. Rc 2 = 6.366 ´ 10 –4 C/W (contact resistance between the two layers of insulation)
Therefore,
Rtotal := R1 + R 2 + R3 + R c1 + R c2 + Ra, C/W (total thermal resistance)
i.e. R total = 1.832 C/W
Heat transfer rate, Q:
T1 - Ta
Q := ,W (heat transfer rate)
Rtotal
i.e. Q = 70.96 W (heat transfer rate)
Interface temperatures:
To calculate the interface temperatures, let us calculate the temperature drop through each layer, starting from the inner
titanium layer, i.e. DT = Q.R, from Ohm’s law. Also, remember that Q is the same through each layer.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 99


Temperature drops at the interfaces:
Let DTTit be the temperature drop in the titanium metal layer and DTc1, the temperature drop at the interface 1 between
titanium and first insulation layer, and D T1, the temperature drop through the first insulation layer, and D Tc2, the
temperarture drop at the interface 2 between the two insulation layers, and DT2 the temperature drop through the
second insulation layer, and D T3, the temperature drop in the outer convection layer.
From analogy with Ohm’s law, we have:
DTTit := R 1 ×Q°C (temperature drop through titanium metal shell)
i.e. DTTit = 0.202°C (temperature drop through titanium metal shell)
Therefore, T2 := T1 – D TTit
i.e. T2 = 159.798°C (temperature of outer surface of titanium shell)
And,
DTc1 := R c 1 ×Q°C (temperature drop at interface between metal
and first insulation due to contact resistance)
i.e. D Tc1 = 0.085°C (temperature drop at interface between metal
and first insulation layer)
Therefore, T ¢2 := T2 – DTc1
i.e. T¢2 = 159.714°C (temperature of inner surface of first insulation shell)
And,
DT1 := R 2 ×Q°C (temperature drop through the first insulation layer)
i.e. DT1 = 94.114°C (temperature drop through the first insulation layer)
Therefore, T 3 := T2 + DT1
i.e. T 3 = 65.599°C (temperature of outer surface of first insulation shell)
And,
DTc2 := R c2 ×Q°C (temperature drop at interface between the two
insulation layers due to contact resistance)
i.e. D Tc2 = 0.045°C (temperature drop at interface between the two
insulation layers)
Therefore, T ¢3 := T3 – DTc2
i.e. T¢3 = 65.554°C (temperature of inner surface of second
insulation shell)
And,
DT2 := R 3×Q°C (temperature drop through the second insulation layers)
i.e. DT2 = 31.371°C (temperature drop through the second insulation layer)
Therefore, T 4 := T3 – DT2
i.e. T4 = 34.183°C (temperature of outer surface of second insulation shell)
Check: Considering the outer convective layer, and applying Ohm’s law, we should get Q = (T4 – Ta)/Ra. This should
equal 70.96 W, obtained earlier. Verify this:
T4 - Ta
Q := i.e. Q = 70.96 W (verified.)
Ra
Temperature profile is shown in Fig. Example 4.17(b):
Overall heat transfer coefficients, Ui and Uo :
Remember that: Ui Ai = Uo .Ao = 1/SR
where, Ai and Ao are the inner and outer surface areas of spherical shells, respectively.
Now, Ai := 4×p r 12 i.e. Ai = 0.385 m2 (inner surface area of spherical shell)
and, Ao := 4×p r 42 i.e. Ao = 1.131 m2 (outer surface area of spherical shell)
1
Therefore, Ui := W/(m2 C) (overall heat transfer coefficient based on inner
Rtotal × Ai surface area)

i.e. Ui = 1.418 W/(m2 C) (overall heat transfer coefficient based on inner surface area)
1 2
And, Uo := W/(m C) (overall heat transfer coefficient based on outer surface area)
Rtotal × Ao

100 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T4
Q
2
ha = 15 W/(m C)

T1 = 160 C Ta = 30 C

r1 = 0.175 m
r2 = 0.20 m
r3 = 0.25 m
r4 = 0.30 m Titanium shell
T1 = 160 C Insulation-1
T2 = 159.798 C Insulation-2
T2 = 159.714 C Convective layer

T3 = 65.599 C
T3 = 65.554 C
T4 = 34.183 C
Ta = 30 C

FIGURE Example 4.17(b) Temperature profile in the layers

i.e. Uo = 0.483 W/(m2 C) (overall heat transfer coefficient based on outer surface area)
Note: We can also write: Uo = Ui ´ (Ai /Ao).

4.13 Critical Thickness of Insulation


Insulation is added to a surface to reduce the heat loss from the surface to the ambient, if the surface is hot, or to
reduce the heat loss into the surface from the ambient, if the surface is cold. Either way, the aim is to reduce the
heat loss. Generally, addition of insulation does reduce the heat loss; however, there are some interesting cases
where this may not be so, and the addition of insulation, in fact, increases the heat loss!
To get an insight into such a possibility, let us consider following two cases:
Case (i): Insulating a cubical furnace. Let the furnace wall be at a high temperature of T1. Insulation is provided
over this wall to reduce the heat loss to the ambient, which is at a temperature of Ta. Furnace wall loses heat to
the surroundings by conduction through the insulation layer and by convection from the outer surface of
insulation. So, the resistance to heat transfer is made up of two components, namely, conductive resistance of the
insulation slab (= L/(kA)) and convective resistance between the wall surface and the surroundings (= 1/(h.A)),
where L is the thickness of the insulation slab, k its thermal conductivity and h is the heat transfer coefficient for
convection. A is the area normal to the direction of heat flow, which is a constant for a slab configuration. Obvi-
ously, as the insulation thickness is increased, its conductive resistance increases and the convective resistance
remains constant and therefore, the total resistance increases; as a result, the heat loss goes on decreasing as the
insulation thickness goes on increasing.
Case (ii): Insulating a pipe carrying a hot fluid. Let the pipe wall be at a high temperature of T1. Insulation is
provided over this wall to reduce the heat loss to the ambient, which is at a temperature of Ta. Pipe wall loses
heat to the surroundings by conduction through the insulation layer and by convection from the outer surface of
insulation. So, the resistance to heat transfer is made up of two components, namely, conductive resistance
through the cylindrical insulation layer [= ln(r 2/r1)/(2p kL)] and convective resistance between the wall surface
and the surroundings [= 1/(h.Ao)], where r1 is the inner radius of insulation layer (or, outer radius of pipe), r2 is
the outer radius of insulation layer, k its thermal conductivity, L is length of pipe, and h is the heat transfer
coefficient for convection. Ao is the area of outer surface of insulation. Obviously, as the insulation thickness is

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 101


increased, i.e. as insulation radius r 2 is increased, conductive resistance of insulation increases; however,
convective resistance, given by [1/(h.Ao)] goes on decreasing since Ao, the outside surface area goes on increasing
with increasing radius. Therefore, the total resistance may increase or decrease, depending on the relative rates of
change of these two resistances. And, there are situations where the total resistance does decrease as the
insulation thickness increases i.e. the heat transfer rate Q increases as the thickness is increased! Let us analyse
when this happens.
For the above case, the equivalent thermal resistance circuit is shown below in Fig. 4.13 (a).
Consider any radius r of the insulation. Let us investigate the variation of the two resistances and therefore,
of Q, as the insulation radius r varies:
The total temperature potential for heat flow is (T1 – Ta). The resistances involved are:
(i) Rins = conductive resistance of the cylindrical insulation layer
i.e.
T1 Ta F I
r
Q Q R ins =
1
2 p kL
ln GH JK
r1
Rins Rconv
and,
(ii) R conv = convective resistance on the surface of
FIGURE 4.13(a) Equivalent thermal circuit
insulation, i.e. at radius r.
for a cylinder with insulation
1 1
i.e. R conv = =
R ha Ao ha (2 p rL)
Rtot As stated earlier, conduction resistance, Rins increases
as r increases, and, convection resistance, Rconv decreases as
r increases. Variation of Rins and Rconv with r are shown in
Fig. 4.13(b); this figure also shows the variation of the total
Rins resistance, Rtot given by:
R tot = R ins + R conv
Note that R tot passes through a minimum. The insula-
tion radius at which the resistance to heat flow is minimum
is called ‘critical radius’, rc; i.e. the heat flow is a maximum
Rconv at the critical radius.
Correspondingly, the variation of heat flow per unit
r1 rc r length, (Q/L), with r is shown in Fig. 4.13(c):
FIGURE 4.13(b) Variation of resistances In Fig. 4.13 (c), r1 is the radius of the bare pipe and the
with insulation radius for a cylinder value of Q/L at this point is the heat transfer rate per unit
length for the bare pipe. Insulation is added over the pipe
Qmax and till the insulation radius r reaches the value of rc, Q/L
Q/L
goes on increasing and reaches a maximum at r = rc. As the
insulation radius is increased further, Q/L decreases, but is
still at value higher than that for the bare pipe. As can be
x y seen from the figure, at point y, value of Q/L is the same as
at point x, i.e. the value of Q/L for the bare pipe. Beyond the
value of r corresponding to point y, value of Q/L decreases
with r and the insulation becomes really effective. It should
be noted that when the radius of the pipe r1 is less than that
of critical radius rc, the insulation is not really effective in
the radius range of r 1 to rc , since adding the insulation
actually increases the heat flow rate.
Mathematically, to find out at what insulation radius r
the R tot becomes a minimum for the cylindrical system, let
r1 rc r us differentiate the expression for R tot and equate to zero;
then, to confirm that at that rc, the R tot indeed goes through
FIGURE 4.13(c) Heat transfer per unit
a minimum, verify that the second derivative of R tot w.r.t. r
length vs. insulation radius for a cylinder
is positive:

102 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Now,
F I
r
R tot = R ins + R conv =
1
2 p kL
lnGH JK
r1
+
1
ha ( 2 p rL)
...(4.62)

In the above expression for R tot, r the external radius of insulation, is the only variable. So, to find out at what
value of r the R tot is minimum, differentiate R tot w.r.t r and equate to zero:
d 1 r 1 1 ( - 1)
(R tot) = ´ 1´ + ´ 2 =0
dr 2 p kL r r1 ha ( 2 p L ) r
1 1 1 1
i.e. ´ = ´
2 p kL r ha ( 2 p L ) r 2
k
i.e. r = rc = ...(4.63)
ha
Eq. 4.63 gives the expression for critical radius, rc for the cylindrical system.
To confirm that at r = rc , R tot indeed is minimum, let us find out the value of (d 2 R tot/dr2 ) at r = rc :

i.e.
d2
( Rtot ) =
-1 LM
´
1
+
1
´
2 OP
dr 2 r=r c
N
2 p k L r 2 ha ( 2 p L ) r 3 Q r = rc

Substituting r = rc = k/ha in the above expression:


d2 -1 h2 1 h3 ha2
2
(R tot) = ´ a2 + ´ a3 = ...(4.64)
dr 2 p kL k ha p L k 2p k3 L

It is clear from Eq. 4.64 that (d 2 R tot/dr2 ) at r = rc is +ve i.e. Q/L Qmax
at r = rc, the R tot is a minimum. There are two cases of practical
interest, as shown in Fig. 4.14:
Case (i): r1 < rc: This situation is shown in Fig. 4.14 (a).
Here, r1, the radius of the pipe happens to be less than the a b
critical radius. Therefore, addition of insulation increases the
heat flow rate (Q/L) till the insulation radius reaches a value
of rc, the critical radius (i.e. range a – b in the figure) and
beyond this point, further addition of insulation decreases the
value of (Q/L). In practice, such a case is likely to occur if in-
sulation material of poor quality is applied to pipes or wires of
small radius. This situation is profitably utilised in insulating
current carrying wires, where the electrical insulation pro-
vided is a material of poor thermal conductivity; here, the r1 rc r
radius of the current carrying wire is small and is generally
FIGURE 4.14(a) Heat transfer per unit
less than the value of critical radius. Thus, addition of
length vs. insulation radius for a cylinder
electrical insulation actually helps to dissipate more heat from
when r1 < rc
the wire and results in cooling it.
Case (ii): r1 > rc: This situation is shown in Fig. 4.14(b).
Here, r1, the radius of the pipe happens to be more than the critical radius. Therefore, addition of insulation
decreases the heat flow rate (Q/L) as shown in range (a – b) in the figure. In practice, such a case is applicable in
insulation of steam pipes and refrigerant lines. However, it may not be necessary to check for critical radius
while insulating steam lines due to the following reason: generally, the value of k for insulations used in those
applications is of the order of 0.05 W/(mC), and ha for natural convection is of the order of 5 W/(m2 C) and thus
the critical radius is of the order of rc = (0.05/5) = 0.01 m, i.e. 1 cm. Often, the pipe radius is more than this value,
and addition of insulation will decrease the rate of heat transfer as desired.
Critical thickness of insulation for a sphere. Case of a sphere is similar to that of cylinder since here also, as the
radius of insulation increases, the surface area increases. So, as the insulation radius is increased, the conduction
resistance of insulation increases and the convection resistance decreases.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 103


Q/L Qmax Let r1 be the radius of the sphere on which insulation is
applied, and let r be the outer radius of insulation. See Fig.
4.15. We would like to investigate the change of R tot as
insulation radius r is varied:
We have, for the spherical system:
a b r - r1 1
Rtot = Rins + Rconv = + ...(4.65)
4p k rr1 ha ( 4 p r 2 )
Differentiating R tot w.r.t. r and equating to zero:

d d LM F I OP
MN GH JK
1 1 1 1
- +
PQ
(R tot) = =0
dr dr 4 p k r1 r ha ( 4 p r 2 )

1 F
G 0 + r1 IJK - h 42p r = 0
rc r1 r

4p k H
i.e. 2 3
FIGURE 4.14(b) Heat transfer per unit a
length vs. insulation radius for a cylinder
when r1 > rc 1 2
i.e. - =0
4 p k r 2 ha 4 p r 3
Sphere
2k
i.e. r = rc = ...(4.66)
ha
Insulation
r
Eq. 4.66 gives the expression for critical radius, rc for the
spherical system. (To confirm that at r = rc, R tot indeed is mini-
mum, check that the value of (d 2 R tot/dr 2 ) at r = rc. is positive.
r1 This is left as an exercise to the student.) Therefore, critical
thickness of insulation for spherical system = (rc – r 1).
Example 4.18. A refrigerant suction line of 25 mm OD is to be
insulated using a material of thermal conductivity k = 0.25 W/
(mK). The surface heat transfer coefficient h a is 10 W/(m2 K). Verify
if the insulation is effective or not. What should be the maximum
FIGURE 4.15 Critical radius for a sphere value of thermal conductivity of insulation to reduce the heat
transfer? (M.U. 2000)
kins = 0.25 W/(mK) Solution.
Data:
25 mm OD
refrigerant line r 1 := 0.0125 m kins := 0.25 W/(mK) h a := 10 W/(m2 K)
Therefore,
Critical radius, rc :
kins
r c :=
ha
2
ha = 10 W/(m K) (define critical radius for cylindrical system, from Eq. 4.64)
i.e. rc = 0.025 m (...critical radius)
This value of rc is more than r 1, i.e. starting from the refriger-
FIGURE Example 4.18 Critical radius for ant line surface at radius r 1, as we go on increasing the thickness of
a cylinder insulation, Q goes on increasing (instead of decreasing) till we
reach a radius of rc.
Therefore, insulation is not effective.
Max value of k to reduce heat transfer:
As the insulation radius exceeds the value of rc, Q starts decreasing. So, the maximum value of the thermal conductivity
required for rc to be equal to r 1 is given by: k max = r 1 ´ h a (from rc = k/ha)
i.e. kmax := r 1 ×h a
i.e. kmax = 0.125 m (maximum value of thermal conductivity of insulation to reduce heat transfer)

104 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Example 4.19. In Example 4.18, calculate the heat loss per metre length for every 2.5 mm increase in thickness of insula-
tion, up to a radius of 34.5 mm. Draw the graph of (Q/L) vs. radius of insulation. Given: temperature of refrigerant line
surface, T1 = – 20°C and ambient temperature, Ta = 25°C.
Solution.
Data:
r 1 := 0.0125 m kins := 0.25 W/(mK) ha := 10 W/(m2 K) T1 := 20°C Ta := 25°C L := 1 m
Heat transfer per unit length, Q/L:
Overall temperature difference is DT = (T1 – Ta). This value will be negative since T 1 is less than Ta, and the heat flow is
from outside ambient to inside line surface. So, we write DT as (Ta – T1) to make Q positive. The total resistance, Rtot =
R cond + R conv, where R cond is the conductive resistance of cylindrical layer of insulation and R conv is the convective
resistance between the insulation surface and the ambient.
We have from Eq. (4.62):
r F I
Rtot = R ins + R conv =
1
2 p kL
ln
r1 GH JK
+
1
ha (2 p rL)
...(4.62)

And, for L = 1 m, Q = DT/Rtot


Ta - T1
i.e. Q(r) =
F rI
ln G J
W/m (define heat transfer rate per metre length as a function of insulation radius, r)

Hr K1
+
1
2 ×p × kins × L ha × (2 ×p × r × L)
Calculate Q for various values of r:
Let r vary from 12.5 mm to 34.5 m at an increment of 2 mm.
So, define a range variable r to vary through this range. Then, in Mathcad, just give the command ‘r = ’ and a Table
of r values appears; also type the command ‘Q(r) = ’ and a Table of Q values at the defined values of r appears. Arrange
them side by side:
r = 0.0125, 0.0145, ... , 0.0345 ...define the range variable r; first value = 0.0125 m,
next value = 0.0145 m and last value = 0.345 m

r Q (r)
0.0125 35.343
0.0145 37.748
0.0165 39.428
0.0185 40.545
0.0205 41.235
0.0225 41.607
0.0245 41.743
0.0265 41.707
0.0285 41.546
0.0305 41.296
0.0325 40.983
0.0345 40.627

Above Table gives the values of r and corresponding values of Q(r), side by side.
Observe that starting from the bare refrigerant line, (Q/L) goes on increasing as insulation is applied, reaches a
peak at r = 25 mm (i.e. the critical thickness) and then goes on decreasing. It is interesting to note that even with 34.5 mm
thickness of insulation, heat transfer per metre is larger than that with no insulation!
Graph of Q(r) vs. r:
This is drawn easily in Mathcad. Range variable, r and heat flow rate per unit length, Q(r) are already defnied. Now,
from the Graph palette, choose x – y graph, fill in the place holder on the x-axis with r and that on the y-axis with Q(r)
and click outside the graph region. Immediately the graph appears: See Fig. Ex. 4.19

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 105


Q/L = vs. r for cylindrical system
42

40

Q(r)
38

36

34
0.01 0.015 0.02 0.025 0.03 0.035
r

FIGURE Example 4.19

Note that Q(r) passes through a maximum at r = 0.025 m, i.e. at the value of critical radius. Observe the similarity
with Fig. 4.13(c).
Example 4.20. A 25 mm OD pipe line is to be thermally insulated with a material of k = 0.25 W/(mK). Heat transfer
coefficient for surroundings, h a = 12 W/(m2 K). Check whether the insulation would be effective or not. What should be
the maximum value of k for the insulating material to effectively reduce the heat transfer? Also, find the thickness of
insulation if an alternative material with k = 0.04 W/(mK) is employed and it is desired to reduce the heat transfer to
20.7% of that of bare pipe.
Solution.
Data:
r 1 := 0.0125 m (outside radius of pipe line)
kins := 0.25 W/(mK) (thermal conductivity of insulation material)
ka lt := 0.04 W/(mK) (thermal conductivity of alternative insulation material)
ha := 12 W/(m2 K) (heat transfer coefficient on the outside surface of insulation)
L := 1 m (length of pipe line)
Critical radius, rc:
kins
r c := (define critical radius for cylindrical system, from Eq. 4.64)
ha
i.e. r c := 0.021 m (critical radius)
This value of rc is more than r 1 , i.e. starting from the pipe surface at radius r 1, as we go on increasing the thickness
of insulation, Q goes on increasing (instead of decreasing) till we reach a radius of rc . Therefore, insulation is not
effective.
Max value of k to reduce heat transfer:
As the insulation radius exceeds the value of rc, Q starts decreasing. So, the maximum value of the thermal conductivity
required for rc to be equal to r 1 is given by: k max = r 1 ´ ha. (from r c = k/ha)
i.e. kmax := r 1 h a
i.e. kmax := 0.15 m (maximum value of thermal conductivity of insulation to reduce heat transfer.)
Thickness of alternative insulation when its k = 0.04 W/(mK):
It is also stated that with this alternative insulation heat transfer rate must be 20.7% of that of the bare pipe. Now, note
that for the bare pipe, there is heat transfer only by convection at its surface. Let the convective resistance to heat transfer
on the bare surface be R bare. For the insulated pipe, let the total resistance be R tot. Obviously, R tot is made up of
conductive resistance of the cylindrical insulation material (R ins) and the convective resistance over its surface (Rconv).,
i.e. Rtot = Rcond + R conv.
Since the heat transfer with insulation is 20.7% of that for the bare pipe, for the same DT, we write:
Q ins = 0.207 ´ Qbare

106 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. DT/Rtot = 0.207 ´ DT/R bare
i.e. Rtot = R bare/0.207
Now, Rba re = 1/{(2 p r1L).ha}
Rcond = ln(ra lt /r1)/(2pka lt L), where ra lt is the radius of alternative insulation
R conv = 1/(2pra lt ha )
Thermal resistances:
1
Rbare := C/W (define thermal resistance of bare pipe/m)
( 2 × p × r1 × L) × ha
i.e. Rbare = 1.061 C/W (thermal resistance of bare pipe)
Therefore,

Rbare
Rtot := C/W (define total thermal resistance of insulated pipe)
0. 207
i.e. Rtot = 5.126 C/W (total thermal resistance of the insulated pipe)
Now, Rtot = Rcond + Rconv
FG r IJ
HrK+
alt
ln
1 1
i.e. Rtot = Rcond + Rconv = ...(a)
2 p k alt L ha ( 2 p ralt L)

F r I
ln GH 0.0125 JK
alt

1
i.e. Rtot = +
2 ´ p ´ 0.04 ´ 1 12 ´ 2 ´ p ´ ralt ´ 1

F r I
ln GH 0.0125 JK
alt

1
i.e. 5.126 = +
2 ´ p ´ 0.04 ´ 1 12 ´ 2 ´ p ´ ralt ´ 1

F r I + 0.013263 – 5.126 = 0
i.e. 3.9789 ´ ln GH 0.0125 JK
alt
r alt
...(b)

Eq. b is a transcendental equation and has to be solved by trial and error. This is done easily in Mathcad, using
solve block. Start with a trial value of ralt, then type ‘Given’; and immediately below ‘Given’ type the constraint viz. Eq.
b. Then, the command ‘Find (ralt)’ gives the value of ralt. as shown below:
r alt := 0.05 (trial value of r alt)
Given
F r I 0.013263
3.9789 ln GH 0.0125 JK
alt
+
ralt
– 5.12 = 0

Find (r alt) = 0.04186


i.e. r alt := 0.04186 m (radius of alternate insulation)
i.e. we get: r a lt, the radius of the alternate insulation = 0.04186 m i.e. 41.86 mm.
Thickness of alternate insulation:
Therefore, thickness of alternate insulation, t, is given by:
t := r alt – r1 m (define thickness of alternate insulation)
i.e. t = 0.02936 m (thickness of alternate insulation.)
Note: While solving Eq. a, it was first simplified to Eq. b and then Solve block was used. This was done just for clarity.
However, while using Mathcad, Eq. a can directly be solved in the Solve block as shown below: (Also see Example 4.15)
r alt := 0.1 (trial value of r alt)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 107


Given
FG r IJ
HrK
alt
ln
1 1
Rtot = +
2 ×p × k alt × L ha × 2 ×p × ralt × L
Find(r alt) = 0.04186
i.e. ralt = 0.04186 m (radius of alternate insulation...same value as obtained above.)
Note: In this case, we took a different trial value of ralt to start with, just to show that it makes no difference in the final
value of ralt.
Example 4.21. A wire of 8 mm diameter at a temperature of 60°C is to be insulated by a material having k = 0.174 W/
(mC). Heat transfer coefficient on the outside, h a = 8 W/(m2K). Ambient temperature, Ta = 25°C. For maximum heat loss,
what is the minimum thickness of insulation and the heat loss per metre length? Find the increase in heat dissipation
due to insulation. Also, calculate the increase in current carrying capacity due to insulation. (M.U. 2000)
Solution. If r 1, the radius of the wire is less than the critical radius rc , then, as the insulation is added on the bare wire,
heat loss goes on increasing, and becomes a maximum at the critical radius. So, the problem is to determine the critical
radius.
Data:
r 1 := 0.004 m kins := 0.174 W/(mC) ha := 8 W/(m2 K) T1 := 60°C Ta := 25°C L := 1 m
Critical radius:
k ins
r c := m (define critical radius for cylindrical system)
ha
i.e. rc = 0.022 m (critical radius.)
Thickness of insulation for maximum heat transfer:
Maximum heat transfer occurs at the critical radius. Therefore, thickness of insulation at the critical radius is:
tins := rc – r 1 m (define tins)
i.e. tins = 0.01775 m.. = 17.75 mm (thickness of insulation)
Increase in heat transfer rate due to insulation:
Calculate heat loss for the bare wire and insulated wire separately and compare them. Remember: heat loss = DT/R
Heat loss/m without insulation:
T1 - Ta
Q 1 := W/m (i.e. overall DT divided by the convective resistance between
1
wire surface and ambient)
ha × 2 × p × r1 × L

i.e. Q 1 = 7.037 W/m (heat loss for bare wire)


Heat loss/m when insulated up to critical thickness:
T1 - Ta
Q 2 :=
Fr I
ln G Jc
W/m (overall DT divided by the sum of convective and
conductive resistances)
Hr K
1
+
1
2 ×p × kins × L ha × 2p × rc × L

i.e. Q 2 = 14.207 W/m (heat loss, with insulation up to critical thickness)


Therefore, percentage increase in heat dissipation:
Q2 - Q1
×100 = 101.888% (percentage increase in heat transfer rate due to insulation.)
Q1
Increase in current carrying capacity due to insulation:
Now, heat dissipation with bare wire, Q 1 = I 12 ´ R, and,
heat dissipation with insulated wire, Q 2 = I 22 ´ R,
where, I1 and I 2 are currents for bare wire and wire with insulation, respectively, and R is the electrical resistance of the
wire. Therefore,
I2
= (Q2/Q1)(1/2)
I1

108 FUNDAMENTALS OF HEAT AND MASS TRANSFER


And,
Increase in current carrying capacity is given by:
Increase = (I2 – I1)/I1
= {(I2 /I1) – 1}
And, percentage increase = {(I2/I 1) – 1} ´ 100

i = o(Q /Q
2 1
1/ 2
t
- 1 ´ 100

LMF Q I 1
OP
Per cent increase := MG J
2
i.e.
MNH Q K
2
PP
- 1 × 100 (define% increase in current carrying capacity)
1
Q
i.e. Per cent increase = 42.087 (% increase in current carrying capacity.)

4.14 Optimum (or Economic) Thickness of Insulation


As mentioned earlier, insulation on a hot or cold surface reduces the heat loss and thus effects a considerable
saving in energy. Therefore, choice of suitable insulation, its application to the surface and maintenance over the
lifetime of insulation are very important aspects in industry.
Consider the case of an insulation being applied on a Cost/yr
hot surface. As the thickness of insulation goes on
increasing, the amount of heat loss from the surface goes
on decreasing, i.e. the cost of energy lost goes on
decreasing; but, at the same time, the first cost of material
and labour to apply the insulation goes on increasing. Total cost
Therefore, there are two opposing factors coming into
consideration while determining the combined cost, Cost of
namely, increased saving in cost of fuel (or energy) lost insulation
and increased expenditure for material and labour
towards the insulation.
‘Optimum’ or ‘Economic’ thickness of insulation is
that thickness for which the combined cost of the amount
of energy lost through the insulation and the total Cost of lost heat
(material + labour) cost of insulation is a minimum.
Obviously, optimum thickness of insulation depends topt tins
on many factors: fixed cost such as material cost of
FIGURE 4.16(a) Determination of optimum
insulation, and varying costs such as: cost of energy,
thickness of insulation
interest and depreciation, taxes, maintenance costs, etc.
Fig. 4.16 (a) shows cost of heat loss, cost of insulation and the total cost on an annualised basis, plotted
against thickness of insulation:
From the figure we note that the cost of insulation increases with thickness almost linearly, whereas the cost
of heat lost through insulation decreases exponentially. So, the total cost, which is the sum of these two costs,
decreases initially, reaches a minimum and then increases again. Thickness at which total cost is a minimum is
the optimum thickness. Mathematically, this is found out by differentiating the expression for total cost w.r.t.
thickness and equating to zero.
If we wish to compare three or four insulations for the same job, we can draw similar ‘Total cost curves’ for
those insulations and the thickness of the insulation having the lowest total cost is the optimum thickness. This is
shown in Fig. 4.16 (b), where insulation D has optimum thickness.
Generally, optimum thickness of insulation is calculated in one of the following ways:
(i) Combined cost of heat lost plus the insulation (including material) on an yearly basis should be
minimum, or
(ii) The insulation should pay for itself in two or three years, i.e. the cost of material and labour of insulation
should be equal to the cost of energy saved (by way of reducing the heat lost through the insulation) over
two or three years, or

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 109


Cost/yr (iii) Minimise the total present cost, considering an
interest rate, but neglecting other factors such as
depreciation, taxes and maintenance costs.
A In method (iii) mentioned above, the procedure to
B find optimum thickness of insulation is as follows:
Let Q be the heat loss through the insulation per
C D year. Let Ch be the cost of heat loss per unit (Rs/kJ). So,
the cost of heat loss/year = Q.Ch. Then, if i is the com-
pound interest rate, compounded annually, the total
present value of heat loss, P1 over the service life of n
years is given by:
n QCh
P1 = å j = 1 (1 + i) j
, Rs ...(a)

topt tins And, let the present material + labour cost for
insulation be C ins Rs/m 3. Then, if the volume of insula-
FIGURE 4.16(b) Comparison of insulations tion applied is V (m 3), the present value of insulation, P 2
is given by:
P 2 = V ´ Cins, Rs ...(b)
Then, from Eqs. a and b,
Total present value or cost , PT is given by:
n QCh
PT = å j = 1 (1 + i) j
+ VC ins Rs ...(c)

In Eq. c, first term on the RHS is dependent on radius (or thickness, L for a slab) of insulation, and by
differentiating PT w.r.t. radius (or L) and equating to zero, we obtain the value of r (or L) that gives minimum PT.
And then, optimum thickness is easily calculated.
Following examples illustrate the procedure of finding out the optimum thickness:
Example 4.22. A reactor, heated with saturated steam at 7.917 bar (Tsa t = 170°C) is 1.5 m in diameter and 2 m long,
operates 5840 hrs per year. Assume that surface of the reactor is at 170°C and the ambient is at 30°C. It is insulated with
an insulation of k = 0.038 W/(mC) which costs Rs. 16,000 per m3 of insulation (including cost of material, labour,
cladding, etc.). Heat transfer coefficient on the outer surface is 30 W/(m2 C). Cost of steam is Rs. 700 per ton. Latent heat
of steam at the given conditions is 2050 kJ/kg. Efficiency of the steam heating may be taken as 80%. Determine the
optimum thickness of insulation and the money saved per year. Assume that surface temperature of the reactor and the
heat transfer coefficient remain the same for the reactor with and without insulation.
Solution.
Data:
D := 1.5 m L := 2 m T1 := 170°C Ta := 30°C ha := 30 W/(m2 C) kins := 0.038 W/(mC)
700
Cins := 16,000 Rs/m 3 (cost of insulation per m3) Ch := Rs/kJ (cost of heat energy in steam)
1000 ´ 2050
i.e. Ch := 3.41463 ´ 10 –4 (cost of heat energy in steam) hoven := 0.8 (efficiency of oven)
First, we find out the cost of heat lost for the bare reactor (= Cost1). Then, for the insulated reactor, find the costs of
energy lost through the insulation (i.e. CostP1) and the cost of insulation itself (i.e. CostP2) as a function of insulation
thickness. Adding them together gives total cost (i.e. CostTotal). Find out the thickness at which the CostTotal is
minimum by differentiating the expression for CostTotal and equating to zero, or graphically. This thickness is the
‘optimum thickness of insulation’. Now, find cost of heat lost through this insulation of optimum thickness per year (=
Cost2).
Cost of heat lost from the surface of bare reactor (Cost1):
p ×D 2
A := 2× + p×D×L, m 2 (define total surface area of the cylindrical reactor)
4
i.e. A = 12.959 m2 (total surface area of reactor)
Q bare := ha × A×(T1 – Ta), W (define heat loss from surface of bare reactor)

110 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Q bare = 5.443 ´ 10 4 W (heat loss from the surface of bare furnace)
Q × 5840 ´ 3600
Therefore, Q := bare kJ (heat loss per year (of 5840 hrs))
1000
9
i.e. Q = 1.144 ´ 10 kJ (heat loss per year for bare reactor)
Efficiency of steam heating is 80%. Therefore, to dissipate Q amount of energy, the oven must consume an energy
Qin = Q/0.8

Q
Therefore, Q in := kJ/year (define energy consumed by reactor/year)
h oven
i.e. Q in = 1.43 ´ 10 9 kJ/year (energy consumed by bare reactor/year)
Now, find annual cost for bare reactor from the given data that 1 kJ costs Rs. 3.41463 ´ 10 –4.
Cost 1 := Qin ×Ch Rs/year (define annual cost of heat for bare reactor)
5
Cost 1 := 4.884 ´ 10 Rs/year (annual cost of heat for bare reactor.)
Cost of heat loss through insulation, when reactor is insulated (CostP1):
When insulation is installed, the heat loss is determined from:
Q ins = DT/Rtot where DT = (T1 – Ta) and Rtot = (R cond + R conv).
Further, since the reactor diameter is large (more than 1 m), we assume the surface to be flat, i.e. we will consider it
as a slab to calculate the thermal resistance.
Let tins be the thickness of insulation.
Then, we can write:
T1 - Ta
Q ins = W
tins 1
+
k ins A ha A
And, now, heat loss per year for furnace with insulation will be,
Q 1 = Q ins ´ 5840 ´ 3600/1000, kJ/yr.
Efficiency of steam heating is 80%. Therefore, to dissipate Q 1 amount of energy, the reactor must consume an
energy Q 1in = Q 1/0.8, kJ/yr.
Now, find annual cost of heat loss for insulated reactor as a function of tins from the given data that 1 kJ costs Rs.
3.41463 ´ 10 –4
CostP1(tins ) = Q1in ´ Ch, ... Rs./yr.
Therefore,
T1 - Ta 1
CostP1(tins) := ×(5840 ´ 3600 ´ 10 –3)× ×Ch (Rs/yr...define cost of lost heat)
tins 1 h
+ oven
k ins × A ha × A
Cost of insulation (material + labour etc.)...(i.e. CostP2):
Cost of insulation = Volume of insulation ´ cost per unit volume
i.e. CostP2(tins) := (A ´ tins) ´ Cins Rs
Therefore,
CostP2(tins) := A×tins ×Cins ...Rs (define cost of insulation)
Total cost of (heat loss through insulation/yr + insulation)...(i.e. CostTotal):
CostTotal(tins ) := CostP1(tins) + CostP2(tins) Rs/yr (define cost of lost heat)
To find optimum thickness of insulation...(i.e. topt):
Differentiate the expression for CostTotal w.r.t. t ins and equate to zero. Root of the resulting equation is the value of topt .
In Mathcad, this procedure is very easy: First, assume a trial value of tins . Next, define the derivative of CostTotal
as: CostTotal’. Then, use the ‘root function’ to find the root of CostTotal¢ = 0:
tins := 0.1 m (trial value of optimum insulation thickness)
d
CostTotal¢(tins) := CostTotal(tins ) (define the first derivative of CostTotal)
dtins

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 111


topt := root(CostTotal¢(tins), tins ) ...use root function to get the root = t opt
Therefore,
topt = 0.05336 m...= 5.34 cm (thickness of optimum insulation where total cost is minimum.)
To find t opt graphically:
Plot the three costs, namely, CostP1, CostP2 and CostTotal as a function of tins. Optimum thickness is that thickness at
which the CostTotal is a minimum.: This is very easily done in Mathcad: First, define a range variable tins varying from
an initial value of say, 1 cm up to a final value of say, 20 cm with an increment of 0.5 cm. Then, select the x–y graph from
the graph palette and fill in tins in the place holder on the x-axis. On the y-axis place holder, fill in CostP1(tins),
CostP2(tins), CostTotal (tins). Click outside the graph area and the three graphs appear immediately: See Fig. Ex. 4.22.
tins := 0.01, 0.015... 0.2 (define range variable t ins : first value = 0.01 m,
next value = 0.015 m and last value = 0.2 m)

4
Optimum thickness of insulation
6 ´ 10

Cost of heat lost


Cost P1 (tins) 4
4 ´ 10
Cost P2 (tins) Cost of insulation

Cost Total (tins) Total cost

4
2 ´ 10

0
0 0.05 0.1 0.15 0.2
tins

FIGURE Example 4.22

Note: Observe from the graph that the minimum of CostTotal is at 0.053 m, i.e. the optimum thickness of insulation =
0.053 m, as obtained mathmatically earlier.
Money saved per year due to insulation:
For the bare reactor, cost of heat lost from the surface per year is Cost1.
For the insulated reactor, insulated with optimum thickness of insulation, cost of heat lost per year is: CostP1(t opt).
Difference between these two costs is the money saved per year:
Cost1 = 4.884 ´ 10 5 Rs (already calculated)
4
CostP1(t opt ) = 1.133 ´ 10 Rs/yr (cost of heat lost through optimum thickness of insulation)
Saving per year is given by:
Saving = Cost1 – CostP1 (topt ).
Therefore,
Saving := Cost1 – CostP1 (topt) (define Saving...money saved per year due to insulation)
i.e. Saving = 4.771 ´ 10 5 Rs/yr (money saved due to insulation.)
It may be noted that method of finding out ‘optimum thickness’ of insulation is rather involved. In practice, it is
more convenient to select the optimum thickness of insulation from the charts and tables prepared by TIMA (Thermal
Insulation Manufacturers’ Association) and their member companies.

112 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4.15 Effect of Variable Thermal Conductivity
In the cases considered so far, we assumed thermal conductivity, k to be constant, i.e. k did not change with
temperature. However, this assumption may not be always true. When the k of a material varies rapidly with
temperature or when the temperature range of operation is large, it becomes necessary to take into account the
variation of k with temperature.
In general, analysis of heat transfer with variable k is complicated. However, for the special case of one-
dimensional steady state heat conduction with no internal heat generation, solutions for simple geometries such
as slabs, cylinders and spheres are obtained quite easily as explained below.
Generally, k varies with temperature linearly as follows:
k(T) = ko (1 + bT) ...(4.67)
where, ko = thermal conductivity at 0 deg. C
b = temperature coefficient of thermal conductivity
T = temperature above 0 deg. C.
4.15.1 Plane Slab with Variable Thermal Conductivity
Consider a plane slab as shown in Fig. 4.17. Let the thickness be L. k is a linear function of temperature, given by
Eq. 4.67. Temperatures at the two faces are constant and uniform, i.e. T = T1 at x = 0 and T = T2 at x = L.
Assumptions:
(i) One-dimensional conduction, i.e. thickness L is small compared to the dimensions in the y and z-
directions
(ii) Steady state conduction, i.e. temperature at any point within the slab does not change with time; of
course, temperatures at different points within the slab will be different.
(iii) No internal heat generation
(iv) k varies linearly with temperature, i.e. k(T) = ko(1 + b T). dX
X k = k(T)
Our problem is to first, find out the temperature field within
the slab and then, the heat flux at any point. T1 T2
For the above assumptions, the governing differential equa-
tion reduces to:
d FG dTIJ Q

dx H
k (T )
dx K = 0 in 0 £ x £ L

with k(T) = ko (1 + b T)
B.C.’s: T = T1 at x = 0
T = T 2 at x = L L
X
Solution to the above governing equation with the B.C.’s T1 T2
shown, gives the temperature profile and then, by applying Q Q
Fourier’s law we can get the heat flux any point.
Rlab = L/(kmA)
Alternative, simple method:
For heat transfer rate, Q: FIGURE 4.17 Plane slab with k = k(T)
Remember that as far as there is steady state, one-dimensional and the thermal circuit
heat transfer with no internal heat generation, Q flowing through
each layer is a constant, as a consequence of First law. Then, we can directly integrate the Fourier’s equation.
keeping the Q outside the integral sign, since it is a constant, though its value is yet unknown. Performing the
integration within the limits of B.C.’s given, gets the value of Q. Then, using the fact that Q is the same between
any two layers, we get the temperature profile. This method is outlined below:
Consider a differential element of thickness dx at a distance x from the origin as shown in Fig. 4.17. If dT is
the temperature differential across this element, then we can write from Fourier’s law:
dT
Q = –k(T)A
dx

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 113


where, k (T) is given by Eq. 4.67
A = area normal to the direction of heat flow (same at all sections for a slab)
dT/dx is the temperature gradient
Substituting for k(T), separating the variables and integrating from x = 0 to x = L (with T = T1 to T = T2),
we get:

z z
L T2
Q
dx = - ko (1 + bT ) dT ...(a)
A
0 T1

QL LM T2 T2 F I OP
i.e.
A
= + ko (T1 - T2 ) + ko b 1 - 2
MN 2 2 GH JK PQ ...(b)

QL LM FG
(T1 + T2 ) ´ (T1 - T2 ) IJ OP
i.e.
A
= ko (T1 - T2 ) + k o b
N H 2 KQ
QL LM
T + T2
= (T1 – T2) ko + ko b 1
OP
i.e.
A N 2 Q
QL
i.e. = (T1 – T2) ´ ko (1 + b Tm) = (T1 – T2) ´ k m
A
where, k m = k o (1 + b Tm ) is the mean value of k at the mean value of temperature, Tm.
And, T m = (T 1+T 2)/2
Therefore,
km A (T1 - T2 )
Q= ,W ...(c)
L
Eq. c gives the heat transfer rate for the plane slab, with variable thermal conductivity, k varying linearly
with temperature.
Eq. c is important since in most of the practical cases of insulation for furnaces or lagging of hot pipes,
thermal conductivity, indeed, varies linearly with temperature.
Writing Eq. c in a form analogous to Ohm’s law:
DT (T - T2 )
Q= = 1 ,W ...(4.68)
Rslab L
km A
From Eq. 4.68, it is clear that expression for Q for a slab with thermal conductivity varying linearly with
temperature, is of the same form as for a slab with constant k, except that k is replaced by k m.
To get temperature distribution within the slab:
In Eq. a above, integrate between x = 0 and x = x (with correspondingly, T = T1 and T = T(x)), i.e. result is easily
obtained by replacing L by x, T2 by T(x) in Eq. b:

Q=
Ak o LM b
(T1 - T (x )) + (T12 - T ( x )2 )
OP
i.e.
x N 2 Q ...(d)

We write from from Eq. b:

Q=
Ak o LM b
(T1 - T2 ) + (T12 - T22 )
OP
L N 2 Q ...(e)

Equate Eqs. d and e, since in steady state, Q is same through each layer of the slab. We get:

b T ( x)2 x (T1 - T2 )(1 + b Tm ) b T1 FG IJ = 0


2
+ T(x) +
L
– T1 1 +
2 H K ...(f)

114 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eq. f is a quadratic equation in T, whose roots are given by:

- b ± b 2 - 4 ac
T(x) =
2a
where, a = b/2,
b=1
x (T1 - T2 )(1 + bTm ) FG
bT1 IJ
c=
L
– T1 1 +
H
2 K
Therefore,

LM1 - 4 ´ b ´ RS x (T - T )(1 + b T FG 1 + b T IJ UVOP


H 2 K WPQ
1
-1 ± m ) - T1
T(x) =
MN 2 TL
1 2

b

2

-1 1 2 RS FG b T1 IJ
x (T1 - T2 )(1 + b Tm ) UV
T H K
i.e. T(x) = + + ´ T1 1 + -
b b2 b 2 L W
-1 FG 1 + T IJ 2
2 x
i.e. T(x) =
b
+
Hb K 1 - ´ ´ (T1 - T2 )(1 + bTm )
b L
...(4.69)

Eq. 4.69 gives the temperature distribution within the slab, with the thermal conductivity varying linearly
with temperature.
Temperature profile is shown graphically in Fig. 4.18: k = ko (1 + bT)
Note that: for b > 0, temperature profile is convex b = + ve
for b < 0, temperature profile is concave , and b=0
T1
for b = 0, temperature profile is linear (i.e. for constant k, b = ve
temperature profile is linear, as already shown).
Q
Shape of the temperature profile can easily be deduced as follows:
From k = ko (1 + b T), we can write:
dT T2
= ko b
dx
Therefore, for positive value of b: dk/dT is positive, i.e. k in-
creases with increasing temperature or decreases with decreasing L
temperature. Now, from Fourier’s law:
X
FG Q IJ = k RS-FG dT IJ UV
H A K T H dx K W FIGURE 4.18 Temperature profile in
a slab with variable k

As x increases, T decreases and, therefore, k also decreases. Then, to keep the heat flux, (Q/A), constant,
– (dT/dx) must increase; i.e. (dT/dx) must decrease. So, the curve is convex. See the upper curve for b > 0 in
Fig. 4.18.
For negative value of b, by similar argument, the curve will be concave as shown in Fig. 4.18.
Example 4.23. A plane wall of fire clay brick of thickness 25 cm has temperatures of 1350°C and 50°C on its two sides. k
for fireclay brick varies as:
k(T) = 0.838 (1 + 0.0007 T), W/(mC) where T is in deg.C. Calculate:
(i) the rate of heat flow
(ii) temperature at mid-plane
(iii) distance of the plane at 400°C from LHS
(iv) sketch the temperature distribution in the wall.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 115


k = k(T) Solution. See Fig. Ex. 4.23.
Data:
T1 = 1350°C T1 := 1350°C T2 := 50°C L := 0.25 m A := 1 m2
k(T) := 0.838 ×(1 + 0.0007 ×T ) W/(mC)
Q Note that k varies linearly with temperature since the above
relation is of the form:
k(T) = ko (1 + b T), where:
T2 = 50°C
ko := 0.838 (k value at zero deg.C)
b := 0.0007 (temperature coefficient conductivity)
Since k is varying linearly with temperature, Q, the heat
transfer rate is determined by the same formula as for plane slab
L = 0.25 m
with constant k, except that instead of k we have to use k m, the
X
mean value of k(T). Remember that for a plane slab with constant k,
T1 T2 we have:
Q Q
Q = DT/R sla b and R sla b =L/(kA)
Rlab = L/(kmA)
So, first, determine Tm = (T 1 + T 2)/2, then, km and then, Rsla b
FIGURE Example 4.23 Plane wall—k varying (T1 + T2 )
linearly with T Tm := °C (define mean temperature Tm)
2
i.e. Tm = 700°C (mean temperature)
km := ko×(1 + b ×Tm), W/(mC) (define km, mean thermal conductivity)
i.e. km = 1.249 W/(mC) (mean value of thermal conductivity)
Thermal resistance:
L
Rwa ll := C/W ...define thermal resistance of wall
km × A
i.e. Rwa ll = 0.2 C/W ...thermal resistance of wall
Heat transfer rate, Q:
Therefore,
T1 - T2
Q := W ...define heat transfer rate through the wall)
Rwall
i.e. Q := 6.49282 ´ 103 W (heat transfer rate through the wall.)
Temperature at mid-plane, i.e. at x = 0.125 m:
We can use Eq. 4.69 and substitute x = 0.125 in that equation.
However, let us work this out from fundamentals. Remember Q is same through each layer in steady state, since
there is no internal heat generation. And, Q is already worked out to be 6492.82 W.
From Fourier’s law, at any x, we can write:
(Q/A) = –k(dT/dx)

z z z
0 . 125 T (x ) T ( x)
Q
i.e. dx = – k (T ) dT = - 0.838 (1 + 0.0007 T ) dT
A
0 1350 1350

1
i.e. 6492.82 ´ 0.125 = 0.838 ´ (1350 – T(x)) + 0.838 ´ 0.0007 ´ ´ (1350 2 – T(x)2)
2
i.e. 0.0002933 T 2 + 0.838 T – 854.2367 = 0
Solving for root of this quadratic equation in T, we get the value of T at x = 0.125 m, i.e. at mid-plane:
1
- 0 .838 + [( 0 .838) 2 + 4 × 0.0002933854 . 23675] 2
T= (Root of quadratic equation)
2 × 0 .0002933

i.e. T = 797.034°C (temperature at mid-plane, i.e. at x = 0.125 m)

116 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Verify: verify this value of T from direct formula 4.69:
We have:

T(x) =
-1 FG 1 + T IJ 2
2 x
b
+
Hb K 1 - ´ ´ (T1 - T2 )(1 + b Tm )
b L
... eqn. (4.69)

Put x = 0.125
x := 0.125 m (at mid-plane)

LMF 1 I OP
1
2 2
2 x
MNGH b + T JK
-1
× × (T1 - T2 ) × (1 + b Tm )
T(x) =
b
+ 1 -
b L PQ (define T(x))

Therefore,
T(0.125) = 797.033°C (temperature at mid-plane...verified)
Note the ease with which Eq. 4.69 is evaluated in Mathcad.
Distance of the plane at 400°C from LHS:
Again, in Mathcad, we can solve this easily by Eq. 4.69. But, first let us solve it by conventional method: Let the distance
of the plane at 400°C from origin (i.e. LHS) be x. Then,
Form Fourier’law, at any x, we can write:
(Q/A) = – k(dT/dx)

z z z
x 400 400
Q
i.e. dx = - k (T ) dT = - 0. 838 (1 + 0.0007 T ) dT
A
0 1350 1350

Here, Q is known, already calculated to be 6492.82 W; and, A = 1 m2


Performing the integration:
(1350 2 - 400 2 )
6492.82x = 0.838×(1350 – 400) + 0.838 ×0.0007×
2
(1350 2 - 400 2 )
0.838 × (1350 – 400) + 0.838 × 0.0007 ×
i.e. x := 2 m (define x)
6492. 82
i.e. x = 0.198 m, ...distance of the plane at 400°C from the LHS.
Verify: verify this value of x from direct formula 4.69:
We use the solve block. Start with a trial value of x. Put the constraint of Eq. 4.69 with T(x) as 400°C, immediately
below ‘Given’. Then, the command ‘Find (x) =’ gives the value of x:
x := 0.24 m (trial value of x)
Given

LMF 1 I OP
1
2 2
2 x
MNGH b + T JK
-1
- × × (T1 - T2 ) × (1 + b Tm )
400 =
b
+ 1
b L PQ (Write the constraint that T(x) = 400 in 4.69)

Find (x) = 0.198 m (value of x where T = 400°C...verified.)


To draw temperature profile inside the wall:
Mathcad is ideally suited to do this. First, define a range variable x varying from 0 to 0.25 m, at an interval of 0.01 m.
Then, select x – y graph from the graph palette, fill in x in the place holder on the x-axis and fill in T(x) in the place
holder on y-axis. Click anywhere outside the graph, and immediately the graph is drawn: See Fig. Ex. 4.23(b).
x := 0, 0.01, ... , 0.25 ...define a range variable x, with starting value = 0,
next value = 0.01 and last value = 0.25 m
It may be verified from the above graph that temperatures at x = 0.125 m (i.e. mid-plane) and at x = 0.198 m are
797°C and 400°C, respectively.
Example 4.24. In a furnace, the gas temperature is maintained at 1300°C and the surrounding temperature is 30°C.
The furnace walls are made of a layer of refractory material of thickness 30 cm and thermal conductivity k 1 = (0.28 +

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 117


Temperature profile in wall with variable k
1500

1000

T(x) T(x) in deg.C.


x in metres

500

0
0 0.05 0.1 0.15 0.2 0.25
x

FIGURE Example 4.23(b)

T1 T2 T3 3
k1 = 0.28 + 0.23324 ´ 10 T, W/(mC)
3
k2 = 0.113 + 0.023278 ´ 10 T, W/(mC)

2
Q Q = 750 W/m
2
2
Refractory brick hb = 10 W/(m C)
ha = 30 W/(m C)

Ta = 1300°C Tb = 30°C

L1 = 0.3 m
L2
x
Ta Tb
Q Q
Ra R1 R2 Rb

FIGURE Example 4.24 Furnace with insulations of variable k

0.23324 ´ 10 –3 T), W/(mC) and a layer of bricks with a thermal conductivity k 2 = (0.113 + 0.023278 ´ 10 –3 T), W/(mC).
The heat transfer coefficient from gases to refractory wall is 30 W/(m2 C) and that from brick to surroundings is 10 W/
(m2 C). What should be the thickness of brick layer of the setting so that the loss of heat to surroundings should not
exceed 750 W/m2?
Solution. See Fig. Example 4.24.
Data:
Ta := 1300°C (temperature of hot gases in furnace) Tb := 30°C(temperature of surroundings)
L 1 := 0.3 m(thickness of refractory material)
Let L 2 be the thickness of brick layer.
k 1 : = 0.28 + 0.2334 ´ 10 –3 ×T W(mC) (Thermal conductivity of refractory material)
k 2 := 0.113 + 0.023278 ´ 10 –3 ×T W(mC) ha := 30 W(mC) hb := 10 W(m2 C) Q := 750 Wm 2 A := 1 m2

118 FUNDAMENTALS OF HEAT AND MASS TRANSFER


This is the case of steady state, one-dimensional conduction with no internal heat generation in the walls. There-
fore, thermal resistance concept is applicable. Also, use the fact that in steady state, heat flow rate through each layer is
the same.
Fig. Example 4.24 shows the set up.
First, find temperatures T1 and T3, temperatures of exposed surfaces of refractory and the brick wall, respectively.
Convective thermal resistances Ra and Rb:
1
Ra := C/W (define convective thermal resistance between hot gases and refractory material)
ha × A
i.e. Ra := 0.033 C/W (convective thermal resistance between hot gases and refractory material)
1
Rb := C/W (define convective thermal resistance between surroundings and brick wall)
hb × A
i.e. Rb := 0.1 C/W (convective thermal resistance between surroundings and brick wall.)
Temperatures T1 and T3:
Apply the fact that heat flow rate, Q, through both convective layers is the same, i.e. Q = 750 W/m 2. Therefore, from
Q = (Ta – T1)/Ra and Q = (T3 – Tb)/Rb , we write
T 1 = Ta – Q.Ra, and
T 2 = Tb + Q.Rb
T1 := Ta – Q.Ra°C (define T1 , temperature of exposed surface of refractory layer)
i.e. T 1 = 1.275 ´ 103 °C (temperature of exposed surface of refractory layer.)
T3 := Tb + Q×Rb °C (define T 3, temperature of exposed surface of brick layer)
i.e. T 3 = 105°C (temperature of exposed surface of brick layer.)
To find the interface temperature T2, between the two layers:
Apply the fact that Q through the refractory layer is 750 W/m2. Here, we note that thermal conductivity varies linearly.
So, expression for the thermal resistance is the same as for a wall with constant thermal conductivity [i.e. R = L/(kA) ]
except that k is replaced by km.
So, R 1 = L 1/(kmA); km is obtained by substituting T = Tm = (T 1 + T2)/2 in the given expression for k(T).
Since T2 is not yet known, let us write:
T1 - T2
Q=
L1
LM0.28 + 0.23324 ´ 10 ×FG T + T IJ OP × A
H 2 KQ
-3 1 2
N
In the above equation, all values except T 2 are known. So, simplifying, we get a quadratic equation in T 2 and by
solving that equation we get T 2. This is left as an exercise to the student.
Instead of following that laborious procedure, let us use the solve block of Mathcad to solve for T2 easily:
Start with a trial value for T2 (say = T 2 = 100) and write the above mentioned constraint immediately below ‘Given’.
Then, the command ‘Find(T 2) = ‘ instantly gives the value of T 2:
T2 := 100°C (trial value of T2)
Given
T1 - T2
Q=
L1
LM0.28 + 0.23324 ´ 10 ×FG T + T IJ OP × A
H 2 KQ
-3 1 2
N
Find(T2) = 848.583
i.e. T2 := 848.583°C (temperature of exposed surface of refractory layer.)
Thickness of the brick layer, L2:
Thickness of brick layer L2 is obtained by aplying the equivalent Ohm’s law to the brick layer, i.e. Q = DT/R2, where
DT = (T2 – T3), and
R 2 = thermal resistance of brick layer = L 2/(km A); now, mean value of thermal conductivity, km has to be used since
the thermal conductivity of brick layer varies linearly with temperature.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 119


First, calculate km. For this, substitute the value of Tm in the given expression for k of brick layer. Tm = (T2 + T3)/2:
(T2 + T3 )
Tm := °C (define mean temperature of brick layer)
2
i.e. Tm = 476.791°C (mean temperature of brick layer)
Therefore, mean value of thermal conductivity for brick layer is given by:
k m := 0.113 + 0.023278 ´ 10 –3 × Tm W/(mC) (define mean temperature of brick layer)
i.e. k m = 1.124 W/(mC) (mean thermal conductivity of brick layer)
Now, applying equivalent Ohm’s law to the brick layer:
T2 - T3
Q= (Ohm’s law for brick layer)
L2
km × A

(T2 - T3 ) × k m × A
i.e. L2 := m (define L 2, the thickness of brick layer)
Q
i.e. L 2 = 0.123 m = 12.3 cm (thickness of brick layer.)
Example 4.25. Thermal conductivity of a plane wall varies with temperature according to the relation k(T) = ko (1 + bT 2 ),
where ko and b are constants.
(a) Develop an expression for the heat flow through the slab per unit area if the surfaces at x = 0 and x = L are
maintained at uniform temperatures T1 and T2 . respectively.
(b) Develop a relation for the thermal resistance of the wall if the heat transfer surface is A
(c) Calculate the heat transfer rate through A = 0.1 m2 of the plate for T1 = 200°C, T2 = 0°C, L = 0.4 m, ko = 60 W/
(mC), and b = 0.25 ´ 10 –4 C –2.
Solution. (a) Expression for heat transfer rate, Q:
First of all, note that in this case, thermal conductivity varies with temperature in a non-linear fashion; therefore, the
relations derived earlier for mean thermal conductivity, km and thermal resistance, R, cannot be used.
However, still, it is a fact that there is one dimensional, steady state conduction with no heat generation in the wall.
Therefore, from First law, Q, the heat transfer rate is same through any section of the slab. So, we can directly integrate
Fourier’s equation, taking the Q outside the integral:
dT
Then, we write: Q = – k(T)×A× (Fourier’s law)
dx
Now, substitute for k(T):
dT
i.e. – Q = k o ×(1 + b T 2)×A×
dx
Separating the variables and integrating from x = 0 to x = L, (and T = T 1 to T = T2),

z z
we get:
L T2
Q
dx = – k o × (1 + b × T 2 ) dT
A 0 T1

Q×L ko ×b
i.e. = k o ×(T 1 – T 2) + ×(T 13 – T23)
A 3
Q×L ko ×b
i.e. = k o ×(T 1 – T 2) + × (T1 – T2)×(T12 + T1 ×T 2 + T 22)
A 3
ko × A × (T1 - T2 )
Q=
b LM
× 1 + × (T12 + T1 × T2 + T22 )
OP
i.e.
L 3 N Q ...(a)

Eq. a is the required expression for the heat transfer rate, Q.


(b) Expression for thermal resistance:
Let us write Eq. a in the form analogous to the Ohm’s law:
T1 - T2 T1 - T2
Q= =
R L
LM
ko × A × 1 +
b
× (T 2 + T1 × T2 + T22 )
OP
N 3 1 Q
120 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Therefore, it is clear that the thermal resistance is given by
L
L b iOPQ
Rth = ...(b)
k × A × M1 + × dT2
+ T1 × T2 + T22
o
N 3 1

Eq. b gives the expression for thermal resistance for the wall with k varying with temperature as: k(T) = ko (1 + b T 2 ).
(c) Numerical problem:
Data:
A := 0.1 m 2 T 1 := 200°C T2 := 0.0°C L := 0.4 m ko := 60 W/(mC) b := 0.25 ´ 10 –4 C –2
Therefore, heat transfer rate is given by Eq. a derived above:
ko × A × (T1 - T2 )
Q :=
LM
b
× 1 + × (T12 + T1 × T2 + T22 ) ..W
OP (define heat transfer rate, Q)
L 3N Q
Substituting the values and simplifying, we get:
Q = 4000 W (heat transfer rate through the plate.)

4.15.2 Hollow Cylinder with Variable Thermal Conductivity


Consider a long, hollow cylinder as shown in Fig. 4.19. Let length of cylinder be L, inside radius ri and outside
radius ro. Inner and outer surfaces are at uniform temperatures of Ti and To , respectively. Let k vary linearly with
temperature as given by Eq. 4.67.

T + dT
k(T)
To To
Q dr

Ti
T
L
Ti r

ri
r0 ri
r0
T1 To
Q Q
Rcyl = In(ro/ri)/(2pkmL)

FIGURE 4.19(a) Cylindrical system with variable k and FIGURE 4.19(b) Elemental volume of
the equivalent thermal circuit thickness dr

Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Thermal conductivity varies linearly with temperature
i.e. k(T) = ko (1 + bT)
(iv) No internal heat generation.
Now, since this is a cylindrical system, we start with the general differential equation for one-dimensional
conduction, in cylindrical coordinates (see Eq. 3.17). For the stipulated conditions, the governing equation
reduces to:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 121


d FG dT IJ
dr H
rk (T )
dr K
= 0 in ri £ r £ ro

with k(T) = ko (1 + b T)
B.C.’s: T = Ti at r = ri
T = To at r = ro
Solution to the above governing equation with the B.C.’s shown, gives the temperature profile and then, by
applying Fourier’s law we can get the heat flux any point.
Alternatively:
For heat transfer rate, Q:
Since there is steady state, one-dimensional heat transfer with no internal heat generation, Q flowing through
each layer is a constant, as a consequence of First law. Then, we can directly integrate the Fourier’s equation
keeping Q outside the integral sign, since it is a constant, though its value is yet unknown. Performing the
integration within the limits of B.C.’s given, gets the value of Q. Then, using the fact that Q is the same between
any two layers, we get the temperature profile. This method is outlined below:
Consider a differential element of thickness dr at a distance r from the origin as shown in Fig. 4.19. If dT is
the temperature differential across this element, then we can write from Fourier’s law:
dT
Q = – k(T) Ar
dr
where, k(T) is given by Eq. 4.67
A r = area at radius r, normal to the direction of heat flow = 2 p rL
dT/dr is the temperature gradient
Substituting for k(T), separating the variables and integrating from r = ri to r = ro (with T = Ti to T = To), we
get:

z z
ro To
dr
Q = – 2p k o L (1 + bT ) dT
r
ri Ti

Fr I LM OP
i.e. Q ln GH r JKo
i
= 2p k o L (Ti - To ) +
N
b
2
´ (Ti2 - To2 )
Q
Fr I
Q ln G J
L (T + T ) O
i.e.
Hr K
o
i
= 2p L(T – T )k M1 + b ´
i
N o
2
o PQ = i o
2p L(Ti – To)km ...(a)

i.e. Q = 2p km L (Ti - To ) ...(4.70)


ln o
r FG IJ
ri H K
where,
k m = k o (1 + b Tm ) = mean value of thermal conductivity and,
Tm = (Ti + To)/2 = mean value of temperature.
Note that Eq. 4.70 for heat transfer Q for a cylindrical system with linearly varying k, is of the same as form
as for a cylindrical system with constant k, except that k is replaced by k m.
Eq. 4.70 is important since in most of the practical cases of insulation for lagging of hot pipes, thermal
conductivity, indeed, varies linearly with temperature.
Writing Eq. 4.70 in a form analogous to Ohm’s law, i.e. Q = DT/R, it is clear that thermal resistance of a
cylindrical system with linearly varying k is given by:

FG r IJ
o

Rcyl =
ln
HrK
i
...(4.71)
2p km L
122 FUNDAMENTALS OF HEAT AND MASS TRANSFER
To get temperature distribution within the cylindrical shell:
Integrate the Fourier’s equation. between r = ri and r = r. (with correspondingly, T = Ti and T = T(r)), i.e. result is
easily obtained by replacing ro by r, To by T(r) in Eq. a:
2 p L (Ti - T (r )) ko LM1 + b ´ (T + T (r)) OP i
i.e. Q=
F rI
ln G J
N 2 Q ...(b)

Hr K i
We write from Eq. a:
2 p L (Ti - To ) ko LM1 + b ´ (T + T ) OP i o
Q=
Fr I
ln G J o N 2 Q ...(c)

Hr Ki
Equate Eqs. b and c, since in steady state, Q is the same through each layer of the cylinder:

LM1 + b ´ (T + T (r)) OP = 2 p L (T - T ) k LM1 + b ´ (T + T ) OP


2 p L (Ti - T (r )) ko i i o o i o
F rI
ln G J
N 2 Q Fr I N
ln G J
2 Q o
HrKi H K
r i

F rI
ln G J
L
(T – T(r)) M1 + b ´
(T + T (r )) O Hr K
i.e. i
N 2
i
PQ = F r I (T – T )[1 + b ´ T ] i
i o m
ln G J o
Hr K i

F rI
ln G J
b 2 H r K (T – T ) [1 + b ´ T )
2 o
i.e. (T –T(r)) +
i
2
T - T (r ) =
i
Fr I
ln G J o
i o m

Hr K i

FG r IJ
bT (r ) 2
bTi2 H r K (T – T )[1 + b ´ T ] = 0
ln
i
i.e.
2
+ T (r) –
2
–T +
i
F
ln G J
r I o
i o m ...(d)

Hr K i

Eq. d is a quadratic in T(r). Its roots are given by:

- b ± b 2 - 4 ac
T(r) =
2a

where, a = b/2,
b=1

FG r IJ
bTi2
ln
H r K (T – T )[1 + b ´ T ]
i
c=–
2
– Ti +
Fr I
ln G J
o
i o m

Hr K
i

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 123


Therefore,

LM R| FG r IJ U|OP
MM1 - 4 ´ b ´ |S- b T i
2
H r K (T - T )[1 + b ´ T
ln
i |P
]V P
-1 ±
MM 2 | 2
-T + i
F
ln G J
r I
i o m
||P
N |
T HrK
o
i WPQ
T(r) =
b

2

R| FG r IJ U
-1 1 2 b Ti2 |S ln
H r K (T - T )[1 + b ´ T ]||V
i
i.e. T(r) =
b
+ -
b2 b
´ -
2
- Ti +
|| Fr I
ln G J o
i o
|| m

T HrK i W
F rI
-1 FG 1 + T IJ - 2 ´ GH r JK (T - T )[1 + b ´ T ]
2 ln
i
i.e. T(r) =
b
+
H b K b ln FG r IJ
i
o
i o m ...(4.72)

HrK i

Eq. 4.72 gives the temperature distribution within the cylindrical shell, with the thermal conductivity
varying linearly with temperature.
Compare this with Eq. 4.69 for a slab with the k varying linearly with temperature.
Example 4.26. The inner and outer radii of a hollow cylinder are 5 and 10 cm, respectively. The inside surface is main-
tained at 300°C and the outside surface at 100°C. The thermal conductivity varies with temperature over this
temperature range as: k(T) = 0.5 ´ (1 + 10 –3 T), where T is in deg.C and k(T) is in W/(mC). Determine:
(i) heat flow rate per metre length of cylinder
(ii) temperature at mid-thickness of shell, and
(iii) sketch the temperature profile within the shell.
Solution. See Fig. Ex. 4.26. k(T) T = 100°C
Data: o

ri := 0.05 m ro := 0.10 m L := 1 m
Q
Ti := 300°C To := 100°C k(T) := 0.5 ×(1 + 10–3 ×T)
Therefore, comparing with k(T) = ko (1 + bT), we write:
k o := 0.5 W/(mC) (k at zero deg.C)
b := 10 –3 C–1 (temperature coefficient of thermal conductivity.) Ti = 300°C
Recognise immediately that the thermal conductivity varies
with temperature linearly.
Therefore, expression for the heat transfer rate, Q, for a
cylindrical shell is of the same form as that for the case of ri = 0.05 m
constant k, except that k is replaced by k m, the mean value of
thermal conductivity. See Eq. 4.70. ro = 0.1 m
Heat transfer rate, Q:
First, find Tm and then, km. Then use Eq. 4.70: T1 To
T + To Q Q
Tm := i °C (define mean (average) temperature Tm)
2 R1 = In(ro/ri)/(2pkmL)
i.e. Tm = 200°C (mean (average) temperature Tm.)
FIGURE Example 4.26 Cylinder with variable k
and equivalent thermal circuit

124 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore, k m := ko ×(1 + b × Tm) (define mean value of thermal conductivity, k m)
i.e. km = 0.6 W/(mC) (mean value of thermal conductivity)
Therefore,
Thermal resistance:
Fr I
ln GH r JK
o

i
Rcyl := C/W (define thermal resistance of cylinder)
2 × p × km × L
i.e. Rcyl = 0.184 C/W (thermal resistance of cylinder)
And,
Ti - To
Q := W (define the heat transfer rate)
Rcyl
i.e. Q = 1087.766 W (heat transfer rate per m length.)
Temperature at mid-thickness of shell, i.e. at r = 0.075 m:
This can be found out directly from Eq. (4.72) by substituting r = 0.075. But, let us first work this out from fundamentals
and then verify the result from Eq. 4.72. If T is the temperature at r, then use the fact that Q is the same through each
layer of the shell. Remember, Q is already calculated above.
From Fourier’s equation, we have:
dT
Q = – k(T)×Ar × (Fourier’s law)
dr
dT
i.e. Q = – k(T)×2×p r×L×
dr
dT
i.e. Q = – k o ×(1 + b×T)×2×p×r×L ×
dr
Separating the variables and integrating from r = r 1 to r = r (and T = T1 to T = T) and keeping Q outside the integral,

z z
since it is a constant):
r T
1
Q× dr = – 2×p×ko ×L× (1 + b × T ) dT
ri r T i

F r I = – 2×p×k ×L× LMF T + b ×T I - F T + b ×T I OP


2 2
i.e. Q× lnGH r JK i MNGH 2 JK GH 2 JK PQ
o i
i

F 7.5 IJ = – 2×p×0.5×1× LMFG T + b ×T IJ - FG 300 + 0.001 ´ 300


1087.766 ln G
2 2
I OP
JK PQ
i.e.
H5K MNH 2 K H 2
Simplifying, we get:
b ×T 2
+ T – 204.609 = 0
2
This is a quadratic equation in T, whose positive root is:

b
-1 + 1 + 4× × 204 .609
2
T= °C ...define T at r = 0.075 m
b

2
i.e. T = 187.105°C ...temperature at r = 0.075 m.
The above procedure is the conventional procedure where you get the quadratic equation in T and then, solve for
its roots. However, when you use Mathcad, there is no need for all that labour; just use the solve block of Mathcad. Start
with a trial value of T (say, 120°C) and in the solve block, immediately below ‘Given’, write the constraint, given by Eq.
a above. There is no need to perform the integration, since Mathcad does it internally. Then, the command ‘Find(T) = ’
immediately gives the value of T:
r := 0.075 m (radius at mid-thickness)
T := 120°C (trial value of temperature at r = 0.075 m)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 125


z z
Given
r T
1
Q× dr = – 2 ×p×ko ×L× (1 + b × T ) dT
ri r T i

Find(T) = 187.105°C (temperature at r = 0.075 m.)


Verify: Now, verify this result from Eq. 4.72:
Rewriting Eq. 4.72:

F rI
-1 FG 1 + T IJ 2
2
ln GH r JK
i
T(r) :=
b
+
Hb K i - ×
b F r I × (T - T ) × (1 + b × T )
ln G J
o
i o m (define T(r))

HrK
i

i.e. T (0.075) = 187.105°C (temperature at mid-thickness of shell.)


To sketch the temperature distribution in the shell:
First, define the range variable, r, varying from r = 0.05 m to r = 0.1 m, with an increment of 0.001 m. Then, choose the
x – y graph from the graph palette, fill in ‘r’ and ‘T(r)’ in the place holders on the x-axis and y-axis respectively. Click
anywhere outside the graph region and immediately the graph appears:
r := 0.05, 0.051, ..., 0.1 (define the range variable r, with first value = 0.05,
next value = 0.051, and last value = 0.1 m.)

F rI
-1 FG 1 + T IJ 2
2
GH r JK
ln
i
T (r) :=
b
+
Hb K i - ×
b F r I × (T - T ) × (1 + b × T )
ln G J o
i o m

HrK i

T(r) for cylindrical shell with variable k


300

r in metres
250 and T(r) in
deg.C

T(r)
200

150

100
0.04 0.05 0.06 0.07 0.08 0.09 0.1
r

FIGURE Example 4.26(b)

It may be seen from the graph that at the inside and outside surfaces of the cylindrical shell, the temperatures are
100°C and 300°C, respectively, as given in data.
Example 4.27. A steam pipe, 20 cm OD carries steam at 260°C and is insulated with a material whose thermal conductiv-
ity varies linearly with temperature. Insulation thickness is 6 cm. Outer surface of insulation is at 60°C and the heat flow
rate in steady state is measured to be 230 W/m. Reported value of k for this insulation is 0.081 W/(mC) at 100°C. Find
out the expression for k(T). Also, find the temperature at mid-thickness of insulation and sketch the temperature profile
in the insulation.

126 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Solution.
Data.
ri := 0.10 m ro := 0.16 m L := 1 m Ti := 260°C To := 60°C Q := 230 W/m
Since it is given that the thermal conductivity varies with temperature linearly, expression for the heat transfer rate,
Q, for a cylindrical shell is of the same form as that for the case of constant k, except that k is replaced by km, the mean
value of thermal conductivity. See Eq. 4.70.
From Eq. 4.70, we have:
2× p × km × L × (Ti - To )
Q=
Fr I
ln GH r JK
o

Therefore,
Fr I
Q × ln GH r JK
o

i
km := W/(mC) (define mean thermal conductivity, k m)
2 × p × L ×(Ti - To )
i.e. km = 0.08602 W/(mC) (mean thermal conductivity, k m)
Now, for linear variation of k with T, we have the variation of the form:
k (T) = ko ×(1 + b × T )
Here, use k o = 0.081 W/(mC) and for T substitute (T – 100):
ko = 0.081 W/(mC) (thermal conductivity at 100°C)
Therefore,
LM (( 260 - 100) + ( 60 - 100) OP
0.08602 = 0.081× 1 + b ×
N 2 Q
i.e. 1.062 = 1 + b × 60
1.062 - 1 –1
i.e. b := C (temperature coefficient of thermal conductivity)
60
i.e. b = 1.033 ´ 10 –3 C –1 (temperature coefficient of thermal conductivity)
Therefore, k(T) is of the form:
k(T) := 0.081×[1 + 1.033 ´ 10 –3 ×(T – 100)] W/(mC)..Eq. (A) (expression for linear variation of k(T ).)
Temperature at mid-thickness, i.e. at r = 0.13 m:
Temperature at r = 0.13 m is calculated by integrating Fourier’s equation from r = r 1 to r = r (and T = T 1 to T = T). While
doing so, Q is constant and can, therefore, be taken out of the integral sign:
From Fourier’s equation we have:
dT
Q = – k(T)×Ar (Fourier’s law)
dr
dT
i.e. Q = – k(T)×2× p×r×L ×
dr
dT
i.e. Q = – ko × (1 + b × (T – 100))× 2× p ×r × L
dr
Separating the variables and integrating from r = r i to r = r (and T = Ti to T = T) and keeping Q outside the integral,

z z
since it is a constant):
r T
1
Q× dr = – 2×p×ko ×L× (1 + b × (T - 100)) dT ...(b)
ri r Ti

F rI LMF T + b ×T - 100 × b × TI - F T + b ×T - 100 × b ×T I OP


2 2
i.e. GH r JK
Q × ln
i
= – 2×p×ko ×L×
MNGH 2 JK GH 2 i
i
JK PQ i

F 0.13 I LF b × T - 100 × b × TI – 268.057 O


= – 2×p×0.081 ´ 1× MG T +
2
i.e. 230× ln G
H 0.1 JK MNH 2 JK PQ
ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 127
LMF T + b ×T 2
I OP
i.e. – 118.554 =
MNGH 2 JK
- 100 × b × T - 268 .057
PQ
1. 033 ´ 10 - 3 × T 2
i.e. + 0.897×T – 149.503 = 0
2
This is a quadratic equation in T, whose positive root is given by:

1. 033 ´ 10 - 3
- 0 .897 + 0 .897 2 + 4 × × 149. 503
2
T := -3
(define T)
1.033 ´ 10

2

i.e. T = 153.162°C (temperature at r = 0.13 m.)


Note: Above-mentioned procedure is, however, laborious. Instead, let us solve for T using solve block of Mathcad. Start
with a trial value of T (say, 100°C) and in the solve block, immediately below ‘Given’, write the constraint, given by Eq.
(B). Then, the command ‘Find(T) = ‘immediately gives the value of T; But, we will define T(r) = Find(T), so that the same
solve block will repeatedly calculate T for any r. This will be useful to draw the temperature profile, i.e. T(r) vs. r.
r := 0.13 m (value of r at mid-thickness of insulation)
T := 100°C (trial value of temperature at radius r)
Given


z
ri
r
1
r z T
dr =– 2×p×ko ×L× (1 + b × (T - 100) dT

T(r) := Find(T)
T i

...define T(r), temperature at any radius


Therefore,
T (0.13) = 153.197°C (Temperature at r = 0.13 m.)
Once again, note the great advantage of using the solve block. For any r, now the temperature T(r) can be calculated
by just putting the value of r in T(r). This will be used to sketch the temperature profile in the insulation:
To sketch the temperature profile in the insulation:
This is done easily in Mathcad. First, define a range variable r, varying from r = 0.1 m to r = 0.16 m, with an increment
of 0.001 m. Then choose x–y graph from the graph palette and fill in the place holders on the x-axis and y-axis with ‘r’
and ‘T(r)’, respectively. Click anywhere outside the graph region and immediately the graph appears: See. Fig. Ex. 4.27.
r := 0.1, 0.101, ... , 0.16 (define the range variable r, with first value = 0.1,
next value = 0.101, and last value = 0.16 m.)

T(r) in cylindrical system with variable k


300

r in metres
250
and T(r) in
deg.C
200
T(r)

150

100

50
0.08 0.1 0.12 0.14 0.16
r

FIGURE Example 4.27

128 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Check: check the value of T(0.13) obtained above by finding out the value of Q between r = 0.1 m and r = 0.13 m; Since
Q should be the same through each layer, we should get Q = 230 W: Between r = 0.1 m (Ti = 260°C) and r = 0.13 m (T =
153.197°C), find k m and then apply Eq. 4.70:
From Eq. A:
LM
k m := k o × 1 + b ×
( 260 - 100) + (153 × 197 - 100) OP (define k m)
N 2 Q
i.e. k m = 0.08992 W/(mC) (mean value of k between r = 0.1 m and r = 0.13 m)
Then, from Eq. 4.70:
r := 0.13 m (radius at mid-thickness)
T := 153.197°C (temperature at r = 0.13 m)
2× p × k m × L × (Ti - T )
Q :=
F rI W (define Q)
ln GH r JK
i

i.e. Q = 229.999 W/m (checks)

4.15.3 Hollow Sphere with Variable Thermal Conductivity


Consider a hollow sphere as shown in Fig. 4.20. Let the inside radius be ri and outside radius ro. Inner and outer
surfaces are at uniform temperatures of Ti and To, respectively. (Ti > To ). Let k of the material vary with linearly
with temperature as given by Eq. 4.67 i.e. k(T) = ko (1 + bT).

k(T) To
T + dT
Q
To
dr

Ti
T
Ti r

ri ri

r0 r0

T1 To
Q Q
Rsph = (ro ri)/(4pkmrori)

FIGURE 4.20(a) Spherical system with variable k FIGURE 4.20(b) Elemental volume of
and the equivalent thermal circuit thickness dr

Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Thermal conductivity varies linearly with temperature, i.e. k(T) = ko (1 + b T )
(iv) No internal heat generation.
Now, since this is a spherical system, we start with the general differential equation for one dimensional
conduction, in spherical coordinates (see Eq. 3.21). For the stipulated conditions, the governing equation reduces
to:
FG
d 2 dT IJ = 0 in r £ r £ r
dr H
r k (T )
dr K i o

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 129


with k(T) = ko (1 + b T)
B.C.’s : T = Ti at r = ri
T = To at r = ro
Solution to the above governing equation with the B.C.’s shown, gives the temperature profile and then, by
applying Fourier’s law we can get the heat flux any point.
Alternatively:
For heat transfer rate, Q:
Since there is steady state, one-dimensional heat transfer with no internal heat generation, Q flowing through
each layer of the spherical shell is a constant, as a consequence of First law. Then, we can directly integrate the
Fourier’s equation keeping the Q outside the integral sign, since it is a constant, though its value is yet unknown.
Performing the integration within the limits of B.C.’s given, we get the value of Q. Then, using the fact that Q is
the same between any two layers, we get the temperature profile. This method is outlined below:
Consider a differential element of thickness dr at a distance r from the origin as shown in Fig. 4.20. If dT is
the temperature differential across this element, then we can write from Fourier’s law:
dT
Q = – k (T)A r
dr
where, k(T) is given by Eq. 4.67
Ar = area at radius r, normal to the direction of heat flow = 4pr2
dT/dr is the temperature gradient
Substituting for k(T), separating the variables and integrating from r = ri to r = ro (with T = Ti to T = To), we
get:
dT
Q = – ko(1 + bT)(4p r2 )
dr

z z
ro To
dr
Q = – 4p ko (1 + bT ) dT
r2
ri Ti

Q
LM - 1OP ro
LM OP
= 4p ko (Ti - To ) +
b
´ (Ti2 - To2 )
i.e.
NrQ ri N Q 2

F 1 1I L
Q G - J = 4p (T – T ) k M1 + b ´
(T + T ) O
i.e.
Hr r K
i o N i
2 PQ = 4p k
o o
i o
m (Ti – To) ...(a)

where,
k o = k m (1 + b Tm ) = mean value of thermal conductivity and,
T m = (Ti + To)/2 = mean value of temperature
4 p km (Ti - To ) 4p km ri ro (Ti - To )
i.e. Q=
F 1 - 1I =
ro - ri
...(4.73)
GH r r JK
i o

Note that Eq. 4.73 for heat transfer Q for a spherical system with linearly varying k, is of the same as form
as for a spherical system with constant k, except that k is replaced by km.
Eq. 4.73 is important since in most of the practical cases, thermal conductivity varies linearly with
temperature. Writing Eq. 4.73 in a form analogous to Ohm’s law, i.e. Q = DT/R, it is clear that thermal resistance
of a spherical system with linearly varying k is given by:
ro - ri
i.e. R sph = ...(4.74)
4p km ri ro
To get temperature distribution within the spherical shell:
Integrate the Fourier’s equation between r = ri and r = r (with correspondingly, T = Ti and T = T(r)), i.e. result is
easily obtained by replacing ro by r, To by T(r) in Eq. a:

130 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4 p ko LM(T - T (r)) + b ´ (T i
2
- T ( r )2 ) OP
i.e. Q=
FG 1 - 1IJ MN i
2 PQ
...(b)

H r rK
i
We write from Eq. a:

L 2
- To2 ) OP
FG 1 - 1 IJ MMN(T - T ) + b ´
4 p ko (T i
Q=
PQ
i o ...(c)
2
Hr r K
i o
Equate Eqs. b and c, since in steady state, Q is the same through each layer of the sphere. Simplifying, we get
a quadratic equation in T(r). Its solution gives the value of T(r), the temperature at radius r. This is left as an
exercise to the student.
Final expression for T(r) is:

-1 FG 1 + T IJ 2
2 ro (r - ri )
T(r) =
b
+
Hb K i - ´ ´
b r (ro - ri )
´ (Ti - To ) ´ [1 + b ´ Tm ] ...(4.75)

Eq. 4.75 gives the temperature distribution within the spherical shell, with the thermal conductivity varying
linearly with temperature. Compare this with Eq. 4.69 for a slab, and 4.72 for a cylinder, with the k varying
linearly with temperature.
Example 4.28. The inside and outside surfaces of a hollow sphere, a < r < b, at r = a and r = b are maintained at uniform
temperatures T1 and T 2 , respectively. The thermal conductivity varies with temperature as:
k(T) = k o (1 + a T + b T 2 )
(a) Develop an expression for the total heat flow rate Q through the sphere.
(b) Develop a relation for the thermal resistance of the hollow sphere.
Solution. Note that now, the variation of k with temperature is not
linear; however, the method to solve the problem is the same as T + dT
adopted earlier, i.e. consider an elemental volume of thickness dr and T2
directly integrate the Fourier’s equation from r = a to r = b, remember- dr
ing that in steady state, one-dimensional conduction, with no internal 2
k(T) = ko(1 + aT + bT )
heat generation, Q is the same through each layer and is, therefore,
taken out of the integral sign.
See Fig. Example 4.28. T
So, for an elemental volume at a radius r and of thickness dr, we Ti r
have:
dT
Q = – k(T)×Ar × (Fourier’s law) a
dr
b
dT
i.e. Q = – k o (1 + a×T + b×T 2)×4×p×r 2×
dr
Separating the variables and integrating from r = a to r = b (and
FIGURE Example 4.28 Sphere with
corresponding, T = T 1 to T = T2), we get: non-linear variation of k(T)


za
b
1
r2 z T2
dr = – 4×p×ko × (1 + a × T + b × T 2 ) dT
T 1

FG 1 - 1 IJ = 4×p×k × LM(T - T ) + a ×(T - T ) + b ×(T - T )OP


H a bK
2 2 3 3

i.e.
No 1
2 3
2
Q 1 2 1 2

Q=
4× p × k × a × b
o L a b
×(T – T )× M1 + ×(T + T ) + × (T 2
+ T1 × T2 + T22 )
OP
i.e.
(b - a) N 2 1
3
2 1 2 1
Q ...(a)

Eq. a is the desired expression for heat transfer rate, Q.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 131


Thermal resistance:
Writing Eq. a in a form analogous to Ohm’s law, i.e. Q = DT/R, we get:
1
R sphere =
F 1 1I L
4 × p × k × G - J × M1 + a × T
b OP ...(b)
o
H a bK N m +
3
× (T12 + T1 × T2 + T22 )
Q
where, Tm = (T 1 + T 2)/2 ...mean temperature.
Eq. b gives the required expression for thermal resistance of the sphere with temperature dependent k.
Example 4.29. An insulated spherical container to store liquid nitrogen, is overall 0.5 m OD and the insulation is 12 cm
thick, whose k varies with temperature as:
k(T) = 0.028 (1 + 5 ´ 10 –3 T), where T is in deg.C.
If surface temperature of the sphere is 90 K, and the outside
k(T) = 0.028 (1 + 0.005 T) surface is at 20°C, find out:
To= 20°C (i) heat transfer rate in steady state
Q (ii) temperature at mid-thickness of insulation, and
(iii) sketch the temperature profile.
Solution. See Fig. Ex. 4.29
Recognise immediately that the thermal conductivity varies
with temperature linearly.
Ti = – 183°C Therefore, expression for the heat transfer rate, Q, for a
spherical shell is of the same form as that for the case of constant
ri = 0.13 m k, except that k is replaced by k m, the mean value of thermal con-
r = 0.25 m ductivity. See Eq. 4.73.
o

Ti To Data:
Q Q ri := 0.13 m ro := 0.25 m Ti := – 183°C
R1 = (ro – ri)/(4pkmrori) To := 20°C k(T) := 0.028 × (1 + 5 ´ 10 –3 ×T)
Therefore, comparing with k(T) = ko (1 + b T), we write:
FIGURE Example 4.29 Sphere with variable k ko := 0.028 W/(mC) b := 5 ´ 10–3 C– 1
Heat transfer rate, Q:
First, find Tm and then, km . Then use Eq. 4.73:
Ti + To
Tm := °C (define mean (average) temperature Tm)
2
i.e. Tm := – 81.5°C (mean (average) temperature Tm.)
Therefore, km := ko ×(1 + b × Tm) (define mean value of thermal conductivity, km)
i.e. k m = 0.01659 W/(mC) (mean value of thermal conductivity)
Therefore,
Thermal resistance:
ro - ri
Rsph := C/W (define thermal resistance of sphere)
4 × p × km × ri × ro
i.e. Rsph = 17.711 C/W (thermal resistance of sphere)
And,
Ti - To
Q := W (define the heat transfer rate)
Rsph
i.e. Q = – 11.462 W (heat transfer rate)
Note: Negative value of Q indicates that heat transfer is from outside to inside (i.e. in the direction opposite to the
positive r direction).
Temperature at mid-thickness of shell i.e. at r = 0.19 m:
This can be found out directly from Eq. 4.75 by substituting r = 0.075.
But, let us first work this out from fundamentals and then verify the result from Eq. 4.75.
If T is the temperature at r, then use the fact that Q is the same through each layer of the shell. Remember, Q is
already calculated above.

132 FUNDAMENTALS OF HEAT AND MASS TRANSFER


From Fourier’ equation, we have:
dT
Q = – k(T)×Ar × (Fourier’s law)
dr
dT
i.e Q = – k(T)×4×p×r 2 ×
dr
dT
i.e Q = – k o ×(1 + b×T)×4×p×r 2×
dr
Separating the variables and integrating from r = r 1 to r = r (and T = Ti to T = T) and keeping Q outside the integral,

z z
since it is a constant):
r T
1
Q× dr = – 4×p×k o × (1 + b × T ) dT ...(a)
ri r2 T i

F 1 - 1I LMF T + b ×T I - F T + b ×T I OP 2 2
i.e. GH r r JK

i MNGH 2 JK GH 2 JK PQ
= 4×p×k o × i
i

F 1 - 1I LL
= 4× p × 0.028× M M - 183 +
0 .005 ) - 183) O F b ×T I O
PQ H 2 JK PPQ
P
2 2
–11.462× G
i.e.
H 0.13 0.19 JK MNMN 2
- GT +

– 79.099 = 99.278 – G T +
F 0.005× T I 2
i.e.
H 2 JK
0.005 2
i.e. ×T + T + 20.179 = 0
2
This is a quadratic equation in T, whose root is:

0.005
-1 + 1 - 4× ´ 20.179
2
T := °C (define T at r = 0.19 m)
0. 005

2
i.e. T = – 21.315°C (temperature at r = 0.19 m.)
Note: When we use Mathcad, there is no need to adopt the above tedious procedure. Instead, use the solve block. Start
with a trial value of T (say, 10°C) and in the solve block, immediately below ‘Given’, write the constraint, given by Eq.
a above. There is no need to perform the integration, since Mathcad does it internally. Then, the command ‘Find(T) = ’
immediately gives the value of T:
r := 0.19 m (radius at mid-thickness of shell)
T := 10°C (trial value of T)

z z
Given
r T
1
Q× dr =– 4×p×k o × (1 + b × T ) dT
ri r2 T i

Find(T) = – 21.28
i.e. T := – 21.28°C (Value of temperature at r = 0.19 m, i.e. mid- thickness of shell )
Verify: Now, verify this result from Eq. 4.75 too:
Rewriting Eq. 4.75:

-1 FG 1 + T IJ 2
LM
2 ro ( r - ri ) OP (define Temp(r))
Temp(r) :=
b
+
Hb K i - × ×
MN
b r ( ro - ri )
× (Ti - To ) × (1 + b × Tm )
PQ
i.e. Temp(0.19) = – 21.28°C (temperature at mid-thickness of shell)
To sketch the temperature distribution in the shell:
First, define the range variable, r, varying from r = 0.13 m to r = 0.25 m, with an increment of 0.001 m. Then, choose the
x–y graph from the graph palette, fill in ‘r’ and Temp(r) in the place holders on the x-axis and y-axis respectively. Click
anywhere outside the graph region and immediately the graph appears:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 133


r := 0.13, 0.131, ... , 0.25 (define range variable r, with first value = 0.13 m ,
next value = 0.131 m, and last value = 0.25 m)

Temperature profile in sphere with variable k


100

r in metres
50 and T(r) in
deg.C
0

Temp (r)
– 50

– 100

– 150

200
0.13 0.15 0.17 0.19 0.21 0.23 0.25
r

FIGURE Example 4.29(b)

Note how the temperature within the spherical insulation shell increases from – 183°C at r = 0.13 m to 20°C at
r = 0.25 m; temperature at mid-thickness (r = 0.19 m) is – 21.28°C, as calculated earlier.

4.16 Two-dimensional Conduction—Shape Factor


So far, we have considered heat transfer in different geometries assuming that heat transfer is one-dimensional.
This is a very good approximation and gives closely accurate results for simple geometries such as slabs,
cylinders and spheres. However, there are many practical cases where this is not a reality as in the case of
irregular shapes or when the temperatures along the boundaries are non-uniform. In such cases, heat transfer
will be in more than one-dimension. Practical examples are: heat treatment of engineering components of
irregular shapes, heat transfer in I.C. engine blocks, chimneys, air conditioning ducts, etc.
To solve multidimensional heat transfer problems, basically, there are four methods:
Analytical method. In this method start with the general differential equation for conduction in the required
coordinate system and solve it in conjunction with the given initial and boundary conditions to get the
temperature field; then apply the Fourier’s equation and get the heat flux at any desired point. However, this
method is suitable only to simple geometries. Otherwise, the solutions are quite cumbersome and require
knowledge of infinite series, Bessel functions, Legendre polynomials, Laplace transform methods and complex
variable theory.
Graphical method. Graphical methods are used for two-dimensional problems with isothermal and adiabatic
boundaries. This is an approximate method. Here, temperature and heat flow lines are drawn by free hand,
remembering that isothermal and heat flow lines are orthogonal, thus forming curvilinear squares. Once such a
‘flux plot’ is drawn, heat flow is easily calculated by applying Fourier’s law to each ‘heat flow lane’. Again, this
method is suitable to simple geometries only and was popular in the early days when computing techniques
were primitive. This method is now almost obsolete.
Analogical method. This method makes use of the electrical analogy between the governing equations of electro-
statics and heat conduction to plot the potential field:
i.e.
¶2E ¶ 2E
+ =0
¶x 2 ¶y 2

134 FUNDAMENTALS OF HEAT AND MASS TRANSFER


and,
¶ 2T ¶ 2T
+ =0
¶x 2 ¶y 2
i.e. temperature and voltage are analogous.
Special conducting paper (or, conducting solution in a bath) is used to make a model of the geometry being
investigated and the isothermal (equipotential) lines are traced using a probe. Then, heat flow lines are drawn
normal to the isothermal lines. This method is more accurate than free hand plotting.
Numerical method. Here, the body is divided into a number of discrete sub-volumes; centre of each sub-volume
is called a ‘node’ and the nodes are connected by fictitious ‘conducting rods’. By making a heat balance on each
node a set of algebraic equations are obtained and these are solved by standard methods to get the temperature
field.
Of the above-mentioned methods, numerical methods have taken over other methods because of availability
of high speed computers and the ability to analyse complex shapes and deal with complicated boundary
conditions. We shall explain this technique in a later chapter.
Shape factors for two-dimensional conduction:
Here, we will explain an approximate, but simple method to analyse a particular type of 2-D conduction
problems where steady state heat transfer occurs between two surfaces at fixed temperatures, T 1 and T 2, with an
intervening solid medium in between. If Q is the rate of heat transfer between two temperature potentials T 1 and
T 2, with the thermal conductivity of intervening material being k, with no heat generation in the medium, we
write:
Q = kS (T 1 – T 2) ...(4.76)
where, S is known as shape factor and has dimension of length. Note that Eq. 4.76 is applicable only when there
is conduction, i.e. in solids. For liquids and gases, where convection is generally the predominant phenomenon,
this equation is not applicable.
From Eq. 4.76, immediately it follows that thermal resistance of the medium is given by:
Rth = 1/(kS) ...(4.77)
Now, recall that thermal resistance of a plane wall, cylinder and sphere are given, respectively, by:
L
Rwall =
k×A

FG r IJ
o

Rcyl =
ln
Hr K
i
2 ×p × k × L

ro - ri
and, Rsph =
4 ×p × k × ri × ro
From this, since we can write: S = 1/(R.k), we get:
A
Swall =
L
2×p × L
Scyl =
FG r IJ
o
ln
Hr K
i

4×p × ri × ro
and, Ssph =
ro - ri
One important application of the concept of shape factor is in calculation of heat transfer in a furnace. Here,
separate shape factors are used to calculate the heat flow through the walls, edges and corners. When all the
interior dimensions are greater than one-fifth of the wall thickness, we get:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 135


A
S wall = , S edge = 0.54×D, and Scorner = 0.15×L
L
where, A = inside area of the wall, L = wall thickness, and D = length of edge.
S has been computed by researchers using electrical analogy or numerical methods, for several cases of practical
interest.
Figs. 4.21 to 4.30 give conduction shape factors for a few selected two-dimensional systems:

Isothermal cylinder of length L buried in a semi- Vertical Isothermal cylinder of length L buried in a
infinite medium (L >> D and z > 1.5 D) semi-infinite medium (L >> D)
S = 2 p L /ln(4z/D) S = 2 p L/ln(4L/D)
T2 T2

z T1 L
T1

L D

FIGURE 4.21 FIGURE 4.22

Two parallel isothermal cylinders placed in an Disk buried parallel to the surface in a semi-
infinite medium (L >> D 1, D 2, z) infinite medium (z >> D)
S = 2 p L/cos h –1{4z 2 – D12 – D22)/(2D1.D2)} S=4D
S = 2 D when z = 0
T1 T2
T2

T1 z
D2

D1
L

D
z

FIGURE 4.23 FIGURE 4.24


Isothermal sphere at T1 buried in an infinite Isothermal sphere buried in a semi-infinite medium
medium at T2 2p D
S=
S = 2/D 1 - 0.25 D/z

T1 T2

Medium at T2 z T1

D
D

FIGURE 4.25 FIGURE 4.26

136 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Isothermal sphere buried in a semi-infinite medium A square flow passage
at T2, whose surface is insulated For a/b > 1.4: S = 2pL/{0.93 ln (0.948 a/b)}
2p D For a/b < 1.41: S = 2pL/{0.785 ln (a/b)}
S=
1 - 0.025 D/z
T2
Insulated

T1 T1
z T2–medium

D a L
b
FIGURE 4.27
FIGURE 4.28

Circular isothermal cylinder of length L at the centre Eccentric circular hole in a cylindrical solid of length
of a square solid bar of same length (L >> w) L, (L > D2)
S = 2pL/ln (1.08 w/D) S = 2 pL/cos h –1 {D1 – D2 – 4z 2)/(2D1.D2)}

D1

T2

Diameter = D T1
T2
z

T1 L
L

w D2

FIGURE 4.29 FIGURE 4.30

Example 4.30. A spherical tank of diameter D = 2 m containing radio-active material is buried in the earth. The distance
between earth’s surface and the tank’s centre is 5 m. Heat release resulting from radioactive decay in the tank is 700 W.
Calculate the steady state temperature of tank’s surface if
the earth’s surface is at 10°C. The value of k of earth at this T2 = 10°C
location may be taken as: k = 1 W/(mC).
Solution. Refer to Fig. Example 4.30.
Data: Q = 700 W
z=5m T1 = ?
D := 2 m z := 5 m T2 := 10°C
Q := 700 W k := 1 W(mC)
For this situation, Shape factor is given in Fig. 4.26.
We have:
2 ×p × D
S := (define shape factor)
D D=2m
1 - 0 .25 ×
z
i.e. S = 13.963 m (Shape factor for given configuration) FIGURE Example 4.30 Isothermal sphere buried
in a semi-infinite medium

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 137


Also,Q= k×S ×(T1 – T2) where T1 is the temperature of tank’s surface.
Therefore,
Q
T1 := + T 2°C (define T1)
k ×S
i.e. T 1 = 60.134°C (temperature of tank’s surface.)
Example 4.31. Inside dimensions of a small cubical furnace constructed of fireclay bricks (k = 1.04 W/(mC)) are: 0.6 m ´
0.6 m ´ 0.6 m, walls being 0.1 m thick. The temperatures on the inside and outside surfaces are 550°C and 50°C, respec-
tively. Determine the heat lost from the furnace.
Solution. Recognise that this problem can be solved by use of ‘shape factors’. Also, recall that when the interior dimen-
sions of the furnace are greater than one-fifth of the wall thickness, we have, for Shape factors:
A
Swall = , Sedge = 0.54×D, and S corner = 0.15×L
L
where, A = inside area of the wall, L = wall thickness, and D = length of edge.
See Fig. Example 4.31.
Note that for a cubical structure, there are 6 wall sections, 12
Q edges and 8 corners. Calculate the Shape factors and compute the
total Shape factor by adding all of them.
T2 = 50°C Data:
Size of furnace: 0.6 m ´ 0.6 m, ´ 0.6 m i.e. dimension of each wall, D =
0.6 m
0.1 m
D := 0.6 m L := 0.1 m A := D×D m2
i.e. A = 0.36 m2 k := 1.04 W/(mC) T1 := 550°C T2 := 50°C
0.6 m S for Walls:
T1 = 550°C 0.6 m S for a single wall section is given by:
A
S := m (define S for single wall section)
0.6 m L
i.e. S = 3.6 m (Shape factor for single wall section.)
FIGURE Example 4.31 Cubical furnace Therefore, for 6 wall sections:
Swalls := S×6 m (S for 6 wall sections)
i.e. Swalls = 21.6 m (S for 6 wall sections)
S for Edges:
S for a single edge is given by:
S := 0.54 ×D m (define S for single edge)
i.e. S = 0.324 m (Shape factor for single edge.)
Therefore, for 12 edges:
Sedges := S×12 m (S for 12 edges)
i.e. Sedges = 3.888 m (S for 12 edges)
S for Corners:
S for a single corner is given by:
S := 0.15 ×L m (define S for single corner)
i.e. S = 0.015 m (Shape factor for single corner.)
Therefore, for 8 corners:
Scorners := S×8 m (S for 8 corners)
i.e. Scorners = 0.12 m (S for 8 corners)
Total Shape factor:
Therefore, total shape factor is obtained by summing up the shape factors for all the walls, edges and corners:
Stotal := Swalls + Sedges + Scorners m (total shape factor)
i.e. Stotal = 25.608 m (total shape factor)
Heat transfer rate, Q:
Therefore, total heat loss from the furnace is given by:
Q := k×Stotal ×(T1 – T2), W (define Q)
i.e. Q = 1.332 ´ 10 4 W = 13.32 kW (total heat loss rate from furnace.)

138 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4.17 Summary of Basic Conduction Relations
We have analysed steady state, one-dimensional heat transfer, with no internal heat generation, in three impor-
tant geometries, namely, plane slab, cylinder and sphere and derived relations for temperature distribution and
rate of heat transfer. We also studied the effect of variable thermal conductivity on these results. Since all these
relations are practically important, they are tabulated in Table 4.4 and Table 4.5, for easy reference.

TABLE 4.4 Relations for steady state, one-dimensional conduction with no internal heat generation, and
constant k

Relation Plane wall Cylindrical shell Spherical shell

d 2T 1 d FG
dT IJ = 0 1 d 2 dTFG IJ
Governing differential equation
dx 2
= 0 ×
r dr

H
dr K ×
r 2 dr
r ×
Hdr K = 0

FrI
ln G J
1 1
T (x ) - T1 x T (r ) - Ti Hr K
i T (r ) - Ti r ri
-
Temperature distribution
T2 - T1
=
L To - Ti
=
F
ln G J
r I To - Ti
=
1
-
1
Hr K
o
ro ri
i

k × A × (T1 - T2 ) 2× p × k ×L × (Ti - To ) 4×p × k ×ri ×ro × (Ti - To )


Heat transfer rate, Q, (W)
L Fr I
ln G J
ro - ri
Hr K
o

Fr I
ln G J
Hr K
o
L i ro - ri
Thermal resistance, R, (C/W)
k ×A 2 ×p × k ×L 4 × p × k × ri × ro

k 2× k
Critical radius, rc , (m) –
h h

TABLE 4.5(a) Relations for steady state, one-dimensional conduction with no internal heat generation and k
varying linearly with temperature as:
k (T ) = ko (1 + bT )
k m = ko (1 + b Tm ); Tm = (T1 + T2)/2

Relation Plane wall

d FG dT IJ
Governing differential equation
dx H
k (T )×
dx K = 0

-1 FG 1 + T IJ 2
2 x
Temperature distribution, T(x)
b
+
Hb K 1 - × ×(T - T2 ) ×(1 + b ×Tm )
b L 1

km × A × (T1 - T2 )
Heat transfer rate, Q, (W)
L

L
Thermal resistance, R, (C/W)
km ×A

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 139


TABLE 4.5(b) Relations for steady state, one-dimensional conduction with no internal heat generation, and k
varying linearly with temperature as:
k (T ) = k o (1 + b T )
km = ko (1 + b Tm ); Tm = (Ti + To)/2

Relation Cylindrical shell

d FG dT IJ
Governing differential equation
dr H
r × k (T )×
dr K = 0

FrI
-1 FG 1 + T IJ 2
2
ln GH r JK
i
Temperature distribution, T (r)
b
+
Hb K i - ×
b F r I ×(T - T )×(1 + b ×T
ln G J
i o m)

Hr K
o

2× p × k m ×L × (Ti - To )
Heat transfer rate, Q, (W)
Fr I
ln GH r JK
o

FG r IJ
Hr K
o
ln
i
Thermal resistance, R, (C/W)
2 × p × k m ×L

TABLE 4.5(c) Relations for steady state, one-dimensional conduction with no internal heat generation,
and k varying linearly with temperature as:
k (T ) = ko (1 + b T )
km = ko (1 + b T m); Tm = (Ti + To)/2

Relation Spherical shell

d 2 FG dT IJ = 0
Governing differential equation
dr H
r × k (T )×
dr K
-1 FG 1 + T IJ 2
LM
2 ro (r - ri ) OP
Temperature distribution, T (r)
b
+
Hb K i - × ×
M
b r (ro - ri )
× (Ti - To ) ×(1 + b ×Tm )
P
4× p × k m × ri × ro × (Ti - To )
Heat transfer rate, Q, (W)
ro - ri

ro - ri
Thermal resistance, R, (C/W)
4 × p × k m ×ri × ro

4.18 Summary
In this chapter, we studied the application of general differential equation for conduction to the cases of steady
state, one-dimensional conduction, with no internal heat generation, in three simple, but important geometries,
namely, plane slab, cylinder and sphere. Expressions for temperature distribution and rate of heat transfer were
derived in these cases by two approaches:
(i) starting with the appropriate differential equation for the problem under consideration, and
(ii) by direct integration of Fourier’s equation.

140 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Despite the mathematical simplicity of the relations derived, note that these relations give reasonably accu-
rate results in practical situations.
Concept of thermal resistance was used to solve steady state, one-dimensional heat transfer problems in
composite systems with no internal heat generation. Concept of contact resistance, and how to include its effect
on heat transfer rate was explained.
Critical radius of insulation was studied for cylindrical and spherical systems; how to find out the optimum
thickness of insulation was also indicated.
Next, the effect variable conductivity on rate of heat transfer and temperature distribution was explained
with reference to the above-mentioned three geometries.
Finally, different methods of solving multi-dimensional heat transfer problems were briefly mentioned and
Shape factors for two-dimensional heat transfer for many cases of practical importance were tabulated.
In the next chapter, we shall continue the study of steady state, one-dimensional heat conduction in simple
geometries, but with internal heat generation.

Questions
1. Explain what is meant by ‘one-dimensional conduction’.
2. Explain ‘log mean area’ for a hollow cylindrical system and ‘geometric mean area’ for a hollow spherical sys-
tem.
[M.U.]
3. Explain the concept of ‘thermal resistance’. What are its applications?
4. Derive an expression to determine the heat flow through a composite cylindrical shell with two layers. Assume
no heat generation and that steady state is reached. [M.U.]
5. Derive the following expression for loss of heat from a lagged pipe per square metre of metal surface per degree
temperature difference between the metal and lagging surface:
k
q² =
FG R IJ
r × log
H rK
where, k is the thermal conductivity of lagging material; r and R are the radii of metal and lagging surface.
Neglect thermal resistance due to metal surface. [V.T.U.]
6. Derive an expression for steady state heat transfer through a composite spherical shell with two layers. [M.U.]
7. Prove that steady state heat transfer rate through the walls of spherical container is given by:

FG k + k IJ ×FG T - T IJ
Q = 4 × p ×R 1 ×R 2 ×
H 2 K HR -R K
1 2 2

2 1

where, k = k 1 + (k 2 – k 1) (T – T1)/(T2 – T1) and T1, T2, k 1, k 2,, R 1, R 2 are all constants. [M.U.]
8. Show that the heat transfer in a steady state unidirectional conduction through a spherical wall is given by:
Q = p kd 1d 2 DT/d
where, d 1 and d 2 are the inner and outer diameter of the sphere, respectively and d is the wall thickness.
[M.U.]
9. What do you mean by ‘overall heat transfer coefficient’? Derive an expression for the same for the case of a
composite cylinder of two layers, based on inside surface as well as outside surface (i.e. for Ui and Uo).
10. Derive an expression for critical thickness of insulation in case of an electric cable. Explain the significance of
critical thickness. [V.T.U.]
11. Show that for a sphere, critical radius is given by: rc = 2 k ins/h.
12. The thermal conductivity of a certain material varies according to the following relation: k = k 0 (1+aT),where k 0
and a are constants. Prove that the heat transfer at steady state condition through a plane wall of thickness L is
given by
Q = k m A (T 1 – T2)/L
where, k m = k o [1 + a(T 1 + T 2)/2]
Also, derive the equation for temperature distribution. [M.U.]
13. Steady one-dimensional heat conduction takes place through the slab of a material whose thermal conductivity
varies linearly with T as: k(T) = k 0 (1 + aT) where a is a constant. The slab is of thickness s and the two faces of
the slab are maintained at temperatures T1 and T2. There is no heat generation within the slab. Solving basic

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 141


differential equation for one-dimensional, steady state conduction, find out the expression for temperature dis-
tribution within the slab. [M.U.]
14. When thermal conductivity varies linearly with T as: k(T) = k 0 (1 + bT), show that heat transfer rate through a
cylindrical shell is given by:
DT
Q=
Fr I
ln GH r JK
o

2 ×p × k m × L
All notations have usual meaning.
15. When thermal conductivity varies linearly with T as: k(T) = k 0 (1 + bT), show that heat transfer rate through a
spherical shell is given by:
DT
Q=
ro - ri
4 × p × k m × ri × ro
All notations have usual meaning.
16. A spherical shell of radii r 1 and r 2 is made of material with thermal conductivity K(T) = k0T 2. Derive an expres-
sion for the heat transfer rate if the surfaces are held at temperatures T1 and T2 , respectively. [M.U.]
17. What do you mean by ‘optimum thickness’ of insulation? How is it determined?
18. Enumerate different methods of solving two-dimensional heat conduction problems.
19. What is ‘conduction shape factor’? How is it related to thermal resistance? Explain its applications.

Problems
Plane slabs and composite slabs:
1. A brick wall (k = 0.7 W/(mC)) is 0.3 m thick. Inner surface is maintained at 45°C and outside surface, at 25°C.
Calculate the heat transfer rate per m2 of area. Also, find the temperature at the mid-plane.
2. A large window glass of thickness 4 mm (k = 0.78 W/(mC)) is exposed to warm air at 20°C at its inner surface,
with a convective heat transfer coefficient between inner surface and air being 15 W/(m2 C). Outside air is at –
10°C and associated heat transfer coefficient is 45 W/(m2 C). Find out the temperatures of inner and outer sur-
faces of the glass and the overall heat transfer coefficient.
3. A furnace wall is made up of 12 cm thick fireclay (k = 0.93 W/(mK)), 20 cm thick red brick (k = 0.7 W/(mK))
with covering of 6 mm thick mild steel plate (k = 39 W(mK)). 18 steel bolts, each of 20 mm diameter are used per
m2 for fixing the steel plate and composite wall together. Find the heat transfer per m2 of furnace wall (length of
bolt 32.6 cm). [M.U.]
4. A house wall may be approximated as two 1.2 cm layers of fibre insulating board, a 8 cm layer of loosely packed
asbestos and a 10 cm layer of common brick. Assuming convection heat transfer coefficients of 15 W/(m2 K) on
both sides of the wall, calculate the overall heat transfer coefficient for this arrangement. (k fibre board = 0.033 W/
(mK), k asbestos = 0.17 W/(mK), k brick = 0.65 W/(mK). [M.U.]
5. A composite wall consists of 15 cm thick layer of material A and a 30 cm thick layer of material B. Thermal
conductivity of the two materials are different and unknown but constant. The outer surface temperature of
material A is 250°C and the outer surface temperature of material B is 50°C. An insulation of k = 0.05 W/(mK)
and thickness 2 cm is added to the outer face of B. It is observed that outer surface of A acquires a temperature
of 330°C and the junction between B and insulation is at 230°C. The outer surface of insulation is at 30°C.
Estimate the rate of heat flow per m2.
(i) before the addition of insulation.
(ii) after the addition of insulation. [M.U.]
6. A furnace wall 30 cm thick has k = 1.4 W/(mK). The heat transfer coefficient of the outer surface is given as h =
(8.1 + 0.09. DT) where DT is the temperature difference between outside wall surface and surrounding. If the
inner surface temperature is 1450°C, calculate the rate of heat loss per unit area. The furnace wall is insulated
such that the heat losses do not exceed 500 W/m2. This is done by putting two layers of insulation on the outer
surface; the first one is of heat resisting brick of k = 0.6 W/(mK) and the second one is of silica brick of k = 0.15
W/(mK). If the thickness of silica brick is 30 cm, find the thickness of heat resisting brick. Assume the surround-
ing temperature as 40°C. [M.U]
7. The inside temperature of furnace wall, 200 mm thick, is 1350°C. The mean thermal conductivity of wall
material is 1.35 W/(mC). The heat transfer coefficient of outside surface is a function of temperature difference

142 FUNDAMENTALS OF HEAT AND MASS TRANSFER


and is given by: h = 7.85 + 0.08 DT where DT is the temperature difference between outside wall surface and
surroundings. Determine the rate of heat transfer per unit area, if the surrounding temp. is 40°C. [M.U.]
8. (a) Calculate the rate of heat flow through 1 m2 area of a clean heating surface of a steam boiler if the flue gas
temperature is 1500°C and the boiling water temperature is 250°C. The heat transfer coefficients from gas to
the wall and from the wall to the water are 120 W/(m2C) and 4500 W/(m2 C)m respectively. The thermal
conductivity of boiler wall material is 52 W/(mC) and its thickness is 12 mm.
(b) If the heating surface exposed to the gas side is covered with soot layer of 1 mm (k = 0.07 W/(mC)) and
water side surface is covered with scale of 1.5 mm thickness (k = 0.8 W/(mC)), calculate the rate of heat
flow and surface temperatures of the corresponding layers. [M.U.]
9. In order to reduce heat loss from a furnace wall, the thickness of its brickwork is increased by 100%. The tem-
perature of the inner surface of the brickwork is 660°C. The temperature of the outer surface, before increasing
the wall thickness was 235°C. Calculate the percentage decrease in the heat loss due to increase in brickwork
thickness. Assume that the thermal conductivity and heat transfer coefficient remain constant. Take atmospheric
temperature as 35°C. [M.U.]
10. A furnace wall is made of composite wall of total thickness 55 cm. The inside layer is made of refractory
material of k = 2.3 W/(mK) and the outside layer is made of an insulating material of k = 0.2 W/(mK). The mean
temperature of the gases inside furnace is 900°C and the interface temperature is 520°C. The heat transfer
coefficient between the gases and inner surface can be taken as 230 W/(m2 K) and between outer surface and
atmosphere as 46 W/(m2 K). Assuming the temperature of surrounding air as 30°C, calculate:
(i) required thickness of each layer
(ii) rate of heat loss per unit area, and
(iii) the temperature of the surface exposed to gases and of the surface exposed to atmosphere. [M.U.]
Contact resistance:
11. A plane composite wall is made of two materials A and B with thermal conductivities kA = 0.1 W(mK) and k B =
0.04 W/(mK). Thicknesses are LA = 10 mm and LB = 20 mm. Contact resistance between the layers is 5 ´ 10 –4
m2 K/W. A fluid at 400°C flows over the free surface area of A with ha = 12 W/(m2 K) and a fluid at 30°C flows
over the surface of B with hb = 25 W/(m2 K). Determine the rate of heat transfer per m2 of the wall surface,
temperature drop at the interface and overall heat transfer coefficient.
12. Two 5 cm diameter, 15 cm long aluminium bars (k = 176 W/(mC)), with ground surfaces are held against each
other at a pressure of 20 bar and the thermal contact conductance at the interface, h c = 11,000 W/(m2 C). Bars are
insulated along their length. Top and bottom surfaces of this two-bar system are maintained at temperature of
250°C and 30°C, respectively. Calculate the rate of heat transfer along the bar in steady state and also the tem-
perature drop at the interface.
Variable area:
13. Ends A and B of a tapered rod, 250 mm long are 50 mm and 25 mm in diameter. The rod is insulated along its
lateral surface. If A and B are maintained at temperatures of 300°C and 27°C, respectively, and k of the material
is 40 W/(mC), find, in steady state:
0.5
(i) heat flow rate through the rod, and D = 0.5 (x)
(ii) temperature at mid-point. Insulated
14. Circular cross section of a cone like solid (k = 25 W/
(mK)), shown in Fig. Problem 4.14, varies as: D = 0.5 T1 = 700 K T2 = 500 K
x 0.5 × x 1 = 25 mm from origin and x 2 = 125 mm.
Temperatures T1 and T2 are maintained at 700 K and 500
K, respectively. Lateral surface is well-insulated. Q
Calculate:
(i) steady state heat transfer rate
(ii) temperature at mid-section. x = 0.025 m
1
Cylinders and composite cylinders: x2 = 0.125 m
15. A steam pipe, 0.12 m OD is insulated with 25 mm thick x
layer of an insulation of k = 0.08 W/(mK). Temperature
FIGURE Problem 4.14 Tapered cone-like solid
of inner and outer surfaces of the insulation are 400°C
and 40°C, respectively. Find the steady state rate of heat loss per metre length of pipe. Also, find the tempera-
ture of insulation at its mid-thickness.
16. A stainless steel tube (k = 19 W/(mK)), 2 cm ID and 4 cm OD is covered with a 4 cm layer of asbestos insulation
(k = 0.2 W/(mK)). If the steady state heat loss rate per metre length of tube is measured to be 750 W/m, and the
outside surface temperature is limited to 50°C, what is the temperature of the inside wall of the tube?

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 143


17. A metal (k = 45 W/(mK)) steam pipe of 5 cm ID and 6.5 cm OD is lagged with 2.75 cm radial thickness of high
temperature insulation having thermal conductivity of 1.1 W/(mK). h i and ho are 4650 W/(m2 K) and 11.5 W/
(m2 K), respectively. If the steam temperature is 200°C and the ambient temperature is 25°C, calculate:
(a) heat loss per metre length of pipe
(b) temperatures at the interfaces
(c) overall coefficient of heat transfer referred to inside and outside surfaces (i.e. calculate Ui and Uo). [M.U.]
18. A steel pipe having an external diameter of 8 cm carries steam at 40 bar and is lagged with a layer, 4 cm thick of
material of k = 0.04 W/(mK). Ambient temperature is 20°C and the surface of lagging has h = 10 W/(m2 K).
What thickness of lagging of k = 0.06 W/(mK) must be added to reduce the steam condensation by 50% if the
surface coefficient remains unchanged? Neglect resistance of the pipe material and also of the steam film on the
inside of the steam pipe. [M.U.]
19. A steam pipe (k = 45 W/(mK)) having 70 mm. I.D. and 85 mm O.D. is lagged with two insulation layers; the
layer in contact with the pipe is 35 mm thick asbestos (k = 0.15 W/(mK)) and it is covered with 25 mm thick
magnesia insulation (k = 0.075 W/(mK)). The heat transfer coefficients for inside and outside surfaces are 220
W/(m2K) and 6.5 W/(m2 K) respectively. If the temperature of steam is 350ºC and ambient temperature is 30ºC,
calculate: (i) the steady state heat loss per metre length of pipe. (ii) overall heat transfer coefficients based on: (a)
inside surface of pipe, and (b) outside surface of pipe. [M.U.]
20. A 200 mm ID pipe carries superheated steam at 210°C. The value of k of pipe material = 13 W/(mK). The outside
insulating layer has k = 0.2 W/(mK). Mean temperature at the interface is 195°C. h between steam and pipe wall
= 60 W/(m2 K) and between outer surface and ambient air is 35 W/(m2 K). Assuming the total thickness of pipe
(including pipe material) is 100 mm, ambient air at 30°C, calculate:
(i) required thickness of each layer
(ii) rate of heat transfer per unit outer area, and
(iii) inner and outer surface temperatures.
Spheres and composite spheres:
21. A hollow sphere of 5 cm ID, 10 cm OD is made of a material of k = 45 W/(mC). What is the heat flux required
at the inner surface to maintain a steady state, inside surface temperature of 200°C when the outside surface
temperature is 15°C? What is the temperature at the mid-thickness of the shell?
22. A hollow sphere of 10 cm ID, 30 cm OD, of material k = 35 W/(mK), contains a liquid chemical. Inner and outer
surface temperatures are 450°C and 150°C. Determine the heat flow rate through the sphere. Also, estimate the
temperature at a point quarter of the way between the inner and outer surfaces.
23. A spherical tank, 1 m in diameter is maintained at temperature of 120°C and exposed to a convection environ-
ment with h = 25 W/(m2K) and temperature of ambient is 15°C. What thickness of urethane foam (k = 20 mW/
(mK)) should be added to ensure that the outer temperature of the insulation does not exceed 40°C? What
percentage reduction in heat loss results from installing this insulation?
24. A hollow spherical form is used to determine thermal conductivity (k) of an insulator. It has an ID of 20 cm, OD
of 50 cm. A heater located at the centre of the sphere dissipates 30 W in steady state operation. Under steady
conditions, temperatures at radii of 15 cm and 20 cm were measured to be 80°C and 60°C, respectively. Deter-
mine the k. Also, find the outer surface temperature. If the surrounding is at 30°C, what is the heat transfer
coefficient over the surface? Plot the temperature profile along the radius.
25. A 600 mm OD sphere storing liquid is provided with two insulating layers, a high temperature insulation of k =
0.35 W/(mK) and a low temperature insulation of k = 0.07 W/(mK). The thickness of the former is 100 mm. The
temperature drop across the high temperature insulation is required to be 2.5 times that across the low tempera-
ture insulation. What should be the thickness of the latter? [V.T.U.]
‘Critical radius’ and ‘optimum thickness’ of insulation:
26. Determine the critical radius for a pipe covered with a layer of asbestos (k = 0.2 W/(mC)), exposed to atmos-
phere, if the outside heat transfer coefficient is 12 W/(m2C).
27. An electrical conductor of diameter 1.5 mm is covered with an insulation of k = 0.15 W/(mC).
(a) If the external convective heat transfer coefficient is 40 W/(m2C), what is the thickness of insulation that
will cause maximum heat transfer?
(b) If the surface of the conductor is maintained at 150°C and ambient temperature is 30°C, and convective
heat transfer coefficient remains the same, what is the rate of heat transfer (i) for bare wire, and (ii) with
optimum insulation?
28. A sphere of diameter 5 cm is heated electrically from inside. It is exposed to an ambient at 30°C with a heat
transfer coefficient of 25 W/(m2C). If the surface of the sphere is to be maintained at 150°C, calculate the rate of
heat loss from the sphere:

144 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i) if the sphere is bare
(ii) if the sphere is covered with insulation of k = 1 W/(mC) to a thickness that will cause maximum heat
transfer.
29. A hollow copper sphere has outer radius of 5 cm. The internal temperature gradient can be neglected. The
temperature of the sphere is to be maintained at 100ºC by an embedded electric heater. The surrounding tem-
perature is 20ºC and the outside heat transfer coefficient is 10 W/(m2K). If this sphere is covered by 5 cm thick
insulation (k = 0.5 W/(mK)), what will happen to heat loss? Calculate the percentage change in heat loss.
[M.U.]
30. A 1 mm diameter wire is maintained at a temperature of 350°C and exposed to convective environment at 30°C
with h = 120 W/(m2K). Calculate the thermal conductivity which will just cause an insulation thickness of 0.2
mm to produce a critical radius. How much of this insulation must be added to reduce the heat transfer to 75%
of that which would be experienced by the bare wire?
31. An electric cable of 10 mm diameter is to be laid in the atmosphere at 25°C. The estimated surface temperature
of the cable due to heat generation is 65°C. Find the maximum percentage increase in heat dissipated, when the
wire is insulated with rubber of k = 0.16 W/(mK). Take h o = 10 W/(m2K). [M.U.]
32. A wire of 8 mm diameter at a temperature of 60°C is to be insulated by a material having k = 0.174 W/(mK).
Given: ho = 8 W/mK, ambient temperature = 25°C. For maximum heat loss, what is the minimum thickness of
insulation and heat loss per metre length? Also, find the increase in heat dissipation due to insulation. [M.U.]
33. A cylindrical vessel, heated with saturated steam at 6.18 bar (Tsat = 160°C) is 1.0 m in diameter and 1.6 m long,
operates 16 hrs. per day, 365 days a year. Assume that surface of the reactor is at 160°C and the ambient is at
25°C. It is insulated with an insulation of k = 0.038 W/(mC) which costs Rs. 17,000 per m 3 of insulation
(including cost of material, labour, cladding, etc.). Heat transfer coefficient on the outer surface is 25 W/(m2C).
Cost of steam is Rs. 750 per ton. Latent heat of steam at the given conditions is 2083 kJ/kg. Determine the
optimum thickness of insulation and the money saved per year. Assume that surface temperature of the vessel
and the heat transfer coefficient remain the same for the reactor with and without insulation.
34. A 85 mm diameter pipe will carry steam at 220°C. It is proposed to lag it with an insulation of k = 0.15 W/(mC)
such that heat transfer rate is reduced to 25% of that from the bare pipe. The outside heat transfer coefficient
with air for lagged pipe is 10 W/(m2 C) and, that from the bare pipe at 220°C is 25 W/(m2C). The ambient
temperature is 30°C. If the estimated insulation cost is Rs. 16,000/m3 (including installation and other expenses)
and the steam cost is Rs. 800 per ton, calculate the pay-back period, assuming a rate of interest of Re. 0.18/
(yr)(Re). Latent heat of steam is 1859 kJ/kg.
Variable thermal conductivity (slabs, cylinders and spheres):
35. The wall of a furnace consists of two layers, one of fireclay of thickness 12.25 cm and the other of red brick of
thickness 48 cm. The thermal conductivity of fireclay is a function of temperature and its given by k 1 = (0.28 +
0.00023 T) W/(mK) where T is in degree Celsius. k 2 = thermal conductivity of red brick =0.7 W/(mK). The inside
surface temperature is 1150°C and outside red brick wall temperature is 55°C. Calculate the amount of heat lost
per m2 area of the furnace wall and interface temperature. [M.U.]
36. Thermal conductivity of a plane slab varies with temperature as: k(T) = ko (1 + b T 2), where ko and b are
constants. If the slab is 0.3 m thick and ko = 60 W/(mC), b = 0.25 ´ 10 –4 C –2 and the two faces of the slab are
maintained at 250°C and 30°C, respectively, find out the steady state heat transfer rate per m2 of the area. Also,
calculate the temperature at mid-thickness. Plot the temperature profile along the thickness.
37. A hollow cylinder whose k varies with temperature as: k(T) = 0.5 (1 + 0.001 T), where T is in deg.C, has an ID of
7.5 cm and OD of 12.5 cm. If the inside and outside surfaces are at uniform temperatures of 250°C and 100°C,
respectively, find out the steady state heat transfer rate per metre length of pipe. Also, find out the temperature
at a radius of 10 cm.
38. A long, hollow cylinder is constructed from a material whose k varies with temperature as: k(T) = (0.01 + 0.001
T), where k(T) is in W/(mK) and T is in deg.C. The inner and outer radii of the cylinder are 125 mm and 250
mm respectively. Under steady state conditions, the inner and outer surface temperatures are 698 K and 363 K,
respectively. Determine:
(i) the rate of heat flow per metre length
(ii) the temperature of the air on the outside of the cylinder, if the surface heat transfer coefficient on the
exterior surface is 14.5 W/(m2 K). [V.T.U.]
39. Thermal conductivity of a sphere, 6 cm ID, 10 cm OD, varies as: k(T) = ko (1 b T 3 ), W/(mC), where T is in deg.C.
Calculate the steady state heat transfer rate if ko = 60 W/(mC), b = 0.25 ´ 10– 4 ,C –3, and Ti = 200°C and To = 0°C.
Also, calculate the temperature at mid-thickness. Plot the temperature along the radius.
40. A cylindrical pressure vessel of 1 m ID and wall thickness 15 cm has the thermal conductivity of its material
varying as: k(T) = 44 ´ (1 – 0.00042 T), W/(mC) where T is in deg.C. If the inside surface temperature is 500°C

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITHOUT HEAT GENERATION 145


and the outer surface temperature is 300°C, what is the steady state heat transfer rate? Also, calculate the tem-
perature at mid-thickness. Plot the temperature variation along the radius.
41. A spherical vessel, 0.3 m ID and 12 cm thick, is made of a material with k(T) = 51 ´ (1 – 0.0008 T), W/(mK)
where T is in deg.C. Inside surface is at – 160°C and outside surface at – 50°C. Calculate the heat loss. What is
the temperature at mid-thickness? Plot the temperature along the radius.
Conduction shape factors:
42. An isothermal sphere of radius r = 8 cm is buried in a large body of earth. The sphere is maintained at an
uniform temperature of 150°C; temperature of earth at large distance from the sphere is 15°C. Calculate the rate
of heat loss from the sphere, in steady state if k of earth is 1.2 W/(mC).
43. Two parallel pipes are buried in earth with a spacing of 0.5 m between their centre lines. One of the pipes has a
radius r 1 = 8 cm and its surface is at 80°C; the other pipe has a radius of r 2 = 5 cm and its surface is at 200°C.
Calculate the heat transfer rate between the pipes per metre length of pipes if k of earth is 1.2 W/(mC).
44. A sphere of 1 m OD containing a radioactive material is buried such that its upper-most point is 1.5 m below the
earth’s surface. If the outside surface temperature of the sphere is 450°C and k of the soil is 1.2 W/(mC), deter-
mine the rate of heat loss from the sphere. Surface temperature of soil is 30°C.
45. Inside dimensions of a furnace are: 3 m ´ 2.5 m ´ 2 m and walls are 0.2 m thick, with a k = 1.3 W/(mC).
Temperature of inner and outer surfaces are 400°C and 50°C, respectively. Calculate the rate of heat loss from
the furnace.

146 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

5
One-dimensional Steady
State Heat Conduction
with Heat Generation

5.1 Introduction
In the previous chapter, we studied one-dimensional, steady state heat conduction for a few simple geometries.
In those cases, there was no internal heat generation in the medium, i.e. the term qg appearing in the general
differential equation was zero. So, the temperature distribution was determined purely by the boundary
conditions. However, there are many practical cases where there is energy generation within the system and we
would be interested to find out the temperature distribution within the body and the heat flux at any location, in
such cases.
Examples of situations with internal heat generation are:
(i) Joule heating in an electrical conductor due to the flow of current in it
(ii) Energy generation in a nuclear fuel rod due to absorption of neutrons
(iii) Exothermic chemical reaction within a system (e.g. combustion), liberating heat at a given rate through-
out the system
(iv) Heat liberated in ‘shielding’ used in nuclear reactors due to absorption of electromagnetic radiation such
as gamma rays
(v) Curing of concrete
(vi) Magnetisation of iron
(vii) Ripening of fruits and in biological decay processes.
Temperature distribution and heat flux are of special interest in some cases where safety of the system or
personnel is involved, e.g. ‘burn-out’ of nuclear fuel rods may occur due to excessive heat, causing a catastrophe,
if suitable precautions for adequate cooling are not taken. Also, analysis of electrical machinery, transformers and
electrical heaters would require that the generation of internal energy is taken into consideration.
Energy generation rate within the system is a volumetric phenomenon; so, its units are: W/m3.
In this chapter, we shall examine the heat transfer in simple geometries (i.e. plane slabs, cylinders and
spheres), with uniform internal energy generation. Several possible boundary conditions will be considered. We
will study the cases with constant thermal conductivity as well as temperature dependent thermal conductivity.
Finally, we will also analyse a few practical applications in the light of the theory studied with reference to these
simple geometries.

5.2 Plane Slab with Uniform Internal Heat Generation


Case of a plane slab with internal heat generation has practical applications in nuclear shielding, fuel rods in
nuclear reactors, electrical conductors, dielectric heating, etc.
While analysing a plane slab with internal heat generation, we shall consider three cases of boundary condi-
tions:
(i) both the sides of the slab are at the same temperature
(ii) two sides of the slab are at different temperatures, and
(iii) one of the sides is insulated.
5.2.1 Plane Slab with Uniform Internal Heat Generation—Both the Sides at the
Same Temperature
Consider a plane slab of thickness 2L as shown in Fig. 5.1. Other dimensions of the slab are comparatively large,
so that heat transfer may be considered as one-dimensional in the x-direction, as shown.
The slab has a constant thermal conductivity k, and a
uniform internal heat generation rate of qg (W/m3). Both
To = Tmax k, qg the sides of the slab are maintained at the same, uniform
temperature of T w. Then, it is intuitively clear that
Temperature
distribution (parabolic) maximum temperature will occur at the centre line, since
the heat has to flow from the centre outwards. Therefore,
Tw Tw it is advantageous to select the origin of the rectangular
coordinate system on the centre line, as shown.
Let us analyse this case for temperature distribution
within the slab and the heat transfer to the sides.
Assumptions:
(i) One-dimensional conduction, i.e. thickness L is
small compared to the dimensions in the y and z-
L L directions.
X (ii) Steady state conduction, i.e. temperature at any
point within the slab does not change with time;
FIGURE 5.1 Plane slab with internal heat
of course, temperatures at different points within
generation—both sides at the same
the slab will be different.
temperature
(iii) Uniform internal heat generation rate, qg (W/m3).
(iv) Material of the slab is homogeneous (i.e. constant density) and isotropic (i.e. value of k is same in all
directions).
We wish to find out the temperature field within the slab and then the heat flux at any point.
We start with the general differential equation in Cartesian coordinates, namely, Eq. 3.9, since the geometry
under consideration is a slab. For the above-mentioned stipulations, Eq. 3.9 reduces to:
d 2T qg
2
+ =0 ...(5.1)
dx k
Solution of Eq. 5.1 gives the temperature profile and then, by using Fourier’s equation we get the heat flux at
any point.
Two B.C.’s are required to solve this second order differential equation.
B.C.’s:
(i) a t x = 0, dT/dx = 0, since temperature is maximum at the centre line.
(ii) At x = ± L, T = Tw
Integrating Eq. 5.1 once,
dT - qg × x
= + C1 ...(a)
dx k
Integrating again,
- qg × x 2
T= + C1 × x + C 2 ...(5.2)
2×k

148 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eq. 5.2 is the general solution for temperature distribution; this is an important equation for the slab with
heat generation. Whatever may be the boundary conditions, solution is given by Eq. 5.2; only the values of
integration constants C1 and C2 change depending on the B.C.’s.
For the present case,
applying B.C. (i) to Eq. a:
C1 = 0
applying B.C. (ii) to Eq. 5.2:
- qg × L2
Tw = + C2
2× k

qg × L2
i.e. C2 = Tw +
2×k
Substituting for C1 and C2 in Eq. 5.2:
- qg × x 2 qg × L2
T(x) = + Tw +
2×k 2×k
+ qg
i.e. T(x) = Tw + × (L 2 – x 2) ...(5.3)
2×k
where, L is half-thickness of the slab. (Remember this)
Note that the temperature, when there is internal heat generation, is not independent of k as in the case of a
slab with no internal heat generation.
Also, by observation, T = Tmax at x = 0. (You can show this easily by differentiating Eq. 5.3 w.r.t. x and
equating to zero.)
Then, putting x = 0 in Eq. 5.3:
qg × L2
Tmax = Tw + ...(5.4)
2×k
Then, from Eqs. 5.3 and 5.4, we get:

T - Tw L2 - x 2 x FG IJ 2

Tmax - Tw
=
L2
=1–
L H K ...(5.5)

Eq. 5.5 gives the non-dimensional temperature distribution in a slab of half-thickness L, with heat genera-
tion. Note that the temperature distribution is parabolic, as shown in Fig. 5.1.
Make two important observations:
(i) From Eq. a, it is clear that temperature gradient (and, therefore, heat flux) for a slab with heat generation
depends on x, whereas it was independent of x in case of a slab with no heat generation.
(ii) From Eq. 5.3, we note that temperature distribution for a slab with heat generation depends on k,
whereas it was independent of k in case of a slab with no heat generation
Convection boundary condition:
In many practical applications, heat is carried away at the boundaries by a fluid at a temperature Tf flowing on
the surface with a convective heat transfer coefficient, h (e.g. current carrying conductor cooled by ambient air or
nuclear fuel rod cooled by a liquid metal coolant). Then, mostly, it is the fluid temperature that is known and not
the wall temperature of the slab. In such cases, we relate the wall temperature and fluid temperature by an
energy balance at the surface, i.e. heat conducted from within the body to the surface is equal to the heat
convected away by the fluid at the surface.
In the case of a plane slab, with both sides at the same temperature, it is clear from consideration of
symmetry that half the amount of heat generated travels to the surface on the right and the other half, to the left.
If A is the surface area of the slab (normal to the direction of heat flow),
we have, from energy balance at the surface:
qg × A × L = h ×A × (Tw – Tf )

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 149


qg × L
i.e. Tw = T f + ...(5.6)
h
Substituting Eq. 5.6 in Eq. 5.3,
qg × L qg
T(x) = Tf + + × (L2 – x 2 ) ...(5.7)
h 2× k
Eq. 5.7 gives temperature distribution in a slab with heat generation, in terms of the fluid temperature, Tf .
Remember, again, that L is half-thickness of the slab.
Heat transfer:
In the case of a slab with no internal heat generation, heat flux was the same at every point within the slab, since
dT/dx was a constant and independent of x. However, when there is heat generation, dT/dx is not independent
of x (see Eq. a), and obviously, heat flux , q ( = – k A dT/dx) varies from point to point along x. But, by observation,
we know that the heat transfer rate from either of the surfaces must be equal to half of the total heat generated
within the slab, for the B.C. of Tw being the same at both the surfaces.
i.e. Q = qg A L ...(5.8)
This is easily verified by applying the Fourier’s law at the surface, i.e. at x = L, since we now have the
temperature distribution given by Eq. 5.3. We get,
Q = – kA (dT/dx)|x = L

i.e. Q = – kA[– qg × 2x/(2k)]|x = L


i.e. Q = + qg AL (same as obtained in Eq. 5.8)
5.2.1.1 Alternative analysis. In the alternative method, which is simpler, instead of starting with the general
differential equation, we derive the above equations from physical considerations.
Let us consider a plane inside the slab at a distance x from the origin, as shown in Fig. 5.2.
We know from observation that maximum temperature oc-
k, qg X curs on the centre line, i.e. centre line is the line of symmetry and
To = Tmax no heat passes across the centre line.
Temperature So, making an energy balance for the surface at a distance x
distribution from the centre line, we can write:
(parabolic) (Heat generated in the volume from x = 0 to x = x ) = (heat
Tw Tw leaving surface at x by conduction)
Then,
dT
qg × A × x = – k × A × ...(a)
dx
Separating the variables and integrating,
- qg × x 2
T(x) = +C ...(b)
L L 2×k
X
Now, at x = 0, Qx = 0 and at x = L, Q L = qg AL, reaches a
FIGURE 5.2 Plane slab with internal heat maximum.
generation—both sides at the same At x = L, T = Tw:
temperature Then, from Eq. b:

- qg × L2
Tw = +C
2×k

+ qg × L2
i.e. C = Tw + ...(c)
2× k
Substitute C from Eq. c in Eq. b:

150 FUNDAMENTALS OF HEAT AND MASS TRANSFER


qg
T(x) = Tw + × (L 2 – x 2) ...(d) (same as Eq. 5.3)
2× k
Eq. d gives the temperature distribution in the slab with heat generation, in terms of the wall temperature,
Tw.
At x = 0, T = Tmax:
Then, from Eq. d:
qg × L2
Tmax = Tw + ...(e)
2×k

qg × L2
i.e. Tmax – Tw = ...(5.9)
2×k
Eq. 5.9 gives the maximum temperature difference within the slab (L is the half-thickness), when tempera-
tures on both sides of the slab are the same. From this equation, Tmax can be calculated, after having determined
Tw from Eq. 5.6. Eq. 5.9 is, therefore, important, since in many cases, we would be interested to know the
maximum temperature reached within the material, to ensure that the material will not melt in a given situation.
Remember this equation.
5.2.1.2 Analysis with variable thermal conductivity. In the above analysis, thermal conductivity of the material was
assumed to be constant. Now, let us make an analysis when the thermal conductivity varies linearly with
temperature as:
k(T) = ko (1 + b T ),
where, ko and b are constants.
Again, considering Fig. 5.2, we have from heat balance (see Eq. a above):
dT
qg × x = – k(T) ×
dx
dT
i.e. – qg × x = ko × (1 + b ×T) ×
dx
Separating the variables and integrating,

z (1 + b ×T ) dT =
- qg
ko z x dx

b ×T 2 - qg x 2
i.e. T+ = × +C ...(f)
2 ko 2
where, C is a constant, determined from the boundary condition:
Now, at x = 0, T = To
Then, from Eq. f,

b ×To2
C = To +
2
Substituting value of C in Eq. f,

b ×T 2 - qg x 2 b ×T02
T+ = × + To +
2 ko 2 2

b ×T 2 qg × x 2 F b × T02 I
i.e. +T+ GG
- To - JJ = 0 ...(g)
2 2 × ko H 2 K

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 151


Eq. g is a quadratic in T. Its solution is:

b qg × x
2 F b ×T02 I
- 1 + 1 - 4× ×
2 2 × ko
GGH
- To -
2
JJK
T(x) =
b

2

-1 F 1 + 2 ×T + T I - q × x 2
i.e. T(x) =
b
+ GH b b 2 JK b × k o o
2 g

-1 FG T + 1 IJ - q × x 2
g
2
i.e. T(x) =
b
+
H b K b ×k
o
o
...(5.10)

Eq. 5.10 gives T(x) in terms of To (i.e. Tmax at x = 0).


Remember that x is measured from the centre line.
If we need T(x) in terms of Tw, put the B.C.: at x = L, T = Tw in Eq. f, get the value of C and then substitute C
in Eq. f to get a quadratic in T. Its solution is:

-1 FG T 1 IJ 2
q g × (L2 - x 2 )
T(x) =
b
+
H w +
b K -
b × ko
..(5.11)

Remember again, that L is the half-thickness of the slab and both the sides of the slab are maintained at the
same temperature, Tw.
5.2.2 Plane Slab with Uniform Internal Heat Generation—
Two Sides at Different Temperatures
Consider a plane slab of thickness L, with constant thermal conductivity k, and temperatures at the two faces
being T1 and T2 as shown in Fig. 5.3. Coordinate system and the origin is chosen as shown.
Let T1 > T2. Now, Tmax must occur somewhere within the slab since heat is being generated in the slab and
is flowing from inside to outside, both to the left and right faces. Let Tmax occur at a distance xmax from the origin,
as shown in Fig. 5.3.
Our aim is to find out the temperature profile in the slab, position where the maximum temperature occurs
in the slab, and the heat transfer rates to the left and right faces.
Assumptions:
k, qg (i) One-dimensional conduction, i.e. thickness L
is small compared to the dimensions in the y
Tmax and z directions.
(ii) Steady state conduction i.e. temperature at
Temperature any point within the slab does not change
T1 distribution with time; of course, temperatures at different
points within the slab will be different.
(iii) Uniform internal heat generation rate, qg (W/
T2 m3).
(iv) Material of the slab is homogeneous (i.e.
constant density) and isotropic (i.e. value of k
Xmax is same in all directions).
Under these assumptions, as shown in section
L
X
5.2.1, the general solution for temperature distribution
is given by Eq. 5.2, i.e.
FIGURE 5.3 Plane slab with internal heat generation,
- qg × x 2
two sides at different temperature T = + C1 × x + C 2
2×k

152 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eq. 5.2 is the general solution for temperature distribution; C1 and C2 are obtained by applying the boundary
conditions. For the present case, B. C.’s are:
B.C. (i): at x = 0, T = T1
B.C. (ii): at x = L, T = T2
Then, from B.C.(i) and Eq. 5.2, we get: C2 = T1 and, from B.C.(ii) and Eq. 5.2, we get:
- qg × L2
T2 = + C1 × L + T1
2× k

T2 - T1 qg × L
i.e. C1 = +
L 2×k
Substituting for C1 and C2 in Eq. 5.2,
- qg × x 2 FT -T I×x + T
qg × L
T(x) =
2×k
+ GH L
2
J
2×k K
1
+ 1

L q (T - T ) O
i.e. T(x) = T + M(L - x) ×
N
1
2 × k
+
g
L
PQ × x
2 1
...(5.12)

Eq. 5.12 gives the temperature distribution in the slab of thickness L, with heat generation and the two sides
maintained at different temperatures of T1 and T2.
Location and value of maximum temperature:
To find out where the maximum temperature occurs, differentiate Eq. 5.12 w.r.t. x and equate to zero; solving, let
the value of x obtained be xmax ; substitute the obtained value of xmax back in Eq. 5.12 to get the value of Tmax. This
procedure will be demonstrated while solving a problem.
Heat transfer to the two sides:
Total heat generated within the slab is equal to:
Qtot = qg AL
Part of this heat moves to the left and gets dissipated at the left face; remaining portion of the heat generated
moves to the right and gets dissipated from the right face.
Applying Fourier’s law:
Qright = – k A (dT/dx)|x = L
Qleft = – k A (dT/dx)|x = 0 (this will be negative since heat flows from
right to left, i.e. in negative x-direction)
Of course, sum of Qright and Qleft must be equal to Qtot.
Convection boundary condition:
Let heat be carried away at the left face by a fluid at a temperature Ta flowing on the surface with a convective
heat transfer coefficient, ha, and on the right face, by a fluid at a temperature Tb flowing on the surface with a
convective heat transfer coefficient, hb . In such cases, we relate the wall temperature and fluid temperature by an
energy balance at the surfaces, i.e. heat conducted from within the body to the surface is equal to the heat
convected away by the fluid at the surface.
Further, the maximum temperature occurs at x = xmax , already calculated. Then, heat generated in the slab in
the volume between x = 0 and x = x max has to move to the left face and the heat generated in the volume between
x = x max and x = L has to move to the right face, since no heat can cross the plane of maximum temperature.
Then, we have, from energy balance at the two surfaces:
On the left face:
qg × A× xmax = ha ×A × (T1 – Ta ) ...(a)
On the right face:
qg × A× (L – xmax) = hb×A × (T2 – Tb ) ...(a)
From Eqs. a and b, we get T1 and T2 in terms of known fluid temperatures Ta and Tb, respectively. Thus after
obtaining T1 and T2, substitute them in Eq. 5.12 to get the temperature distribution in terms of fluid temperatures
Ta and Tb.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 153


5.2.2.1 Effect of variable k , for a slab with heat generation, and two sides at different temperatures. For the assump-
tions of one-dimensional, steady state conduction with uniform heat generation and k varying with temperature
linearly as: k(T) = ko (1 + b T ), the controlling differential equation is (see Chapter 3):
d FG dT IJ + q
dx H
k(T )×
dx K g =0

dT
Integrating k(T)× + qg × x = C1 where C1 is a constant.
dx

z z
Separating the variables and integrating:

k(T) dT = (C1 – qg × x)dx

Substituting for k(T):


z [ko ×(1 + b×T )]dT =
z (C1 – qg × x )dx

b ×T 2 F qg × x 2 I
i.e. T+ =
1
GG
× C1 × x - + C2 JJ ((A)...where C1 and C2 is a constant of integration)
2 ko H 2 K
C1 and C2 are found out by applying the B.C.’s in Eq. A:
B.C.(i): at x = 0, T = T1
B.C.(ii): at x = L, T = T2
From B.C.(i) and Eq. A, we get:
F b ×T12 I
C2 = ko× T1 + GH 2 JK
From B.C.(ii) and Eq. A, we get:

b ×T22 LM
q g × L2 b × T12 F I OP
T2 + =
1
× C1 × L -
MN + ko × T1 + GH JK P
2 ko 2 2
Q
k F b ×T22 I
q g × L ko b ×T12 F I
L GH JK GH JK
Therefore, C1 = × T o + + - × T1 +
2
2 2 L 2
Substituting values of C1 and C2 in Eq. A:

b ×T 2 x F
b × T22 qg × L × x x I
b × T12 qg × x 2 b × T12 F I F I
i.e. T+
2
= × T2 +
L GH 2
+
2 × ko
- × T1 +
L JK
2
-
2 × ko
+ T1 +
2 GH JK GH JK
b ×T 2 F b × T22 x qg × x I F
b × T12 x I FG IJ
i.e. T+
2
GH
= T2 +
2
× +
L 2 × ko JK
× (L - x ) + T1 +
2 GH × 1-
L JK H K
T2 LF
+ T – MG T
b × T I x q ×x
2 F b × T I × FG 1 - x IJ OP = 0 2
i.e. b×
2 MNH 2 + J 2
× +
2 K L 2×k
× (L - x ) + G T +
H
g
2 K H
J L K PQ
o
1
1

This is a quadratic in T. Its positive root is:

LMF
b × T22 x qg × x I
b × T12 x F I FG IJ OP
MNGH JK GH JK H
b
- 1 + 1 + 4 × × T2 +
2 2
× +
L 2 × ko
×( L - x) + T1 +
2
× 1-
L K PQ
T(x) =
b

2

154 FUNDAMENTALS OF HEAT AND MASS TRANSFER


LMFb × T12 x I
b × T12 Fb × T22 qg × x I OP
MNGH JK GH JK
-1 1 2
T(x) = × T1 + - × T1 + - T2 - ×( L - x )
i.e.
b
+
b2
+
b 2 L 2 2
+
2 × ko PQ
-1 F 1 +T I
2 × T1 2×x LM b OP
qg × x
i.e. T(x) =
b
+ GH b 2 1
2
+
b JK-
b ×L
× (T1 - T2 ) + × (T1 - T2 ) ×(T1 + T2 ) +
N 2 Q
b × ko
× ( L - x)

-1 FG 1 + T IJ 2
2×x qg × x
i.e. T(x) =
b
+
Hb K 1 -
b ×L
× (T1 - T2 ) × (1 + b × Tm ) +
b × ko
× (L - x )

T1 + T2
where, Tm = (mean temperature.)
2
Eq. 5.12a gives the temperature distribution in a slab of thickness L, with heat generation, with the two faces
maintained at different temperatures, when the k varies linearly with temperature.
5.2.3 Plane Slab with Uniform Internal Heat Generation—
One Face Perfectly Insulated
Consider a plane slab of thickness L, with constant thermal conductivity k, and one of the faces (say, left face) is
insulated as shown in Fig. 5.4. Other face of the slab is at a temperature of Tw. Coordinate system and the origin
is chosen as shown.
Now, Tmax must occur on the insulated left surface of the k, qg
slab since heat is being generated in the slab and is constrained Insulated
to flow from left face to right face.
Temperature
Our aim is to find out the temperature profile in the slab, distribution
and the heat transfer rate. Tmax
Assumptions:
(i) One-dimensional conduction, i.e. thickness L is small Tw
ha
compared to the dimensions in the y and z directions.
(ii) Steady state conduction, i.e. temperature at any point Ta
within the slab does not change with time; of course,
temperatures at different points within the slab will be
different.
(iii) Uniform internal heat generation rate, qg (W/m3). L
(iv) Material of the slab is homogeneous (i.e. constant X
density) and isotropic (i.e. value of k is same in all
directions). FIGURE 5.4 Plane slab with internal heat
Under these assumptions, as shown in section 5.2.1, the generation, one side insulated
general solution for temperature distribution is given by Eq. 5.2,
i.e.
- qg × x 2
T= + C1× x + C2 ...(5.2)
2×k
Eq. 5.2 is the general solution for temperature distribution; C1 and C2 are obtained by applying the boundary
conditions. For the present case, B. C.’s are:
B.C.(i): at x = 0, dT/dx = 0, since perfectly insulated. ( Note: ‘perfectly insulated’ means that Q = 0, i.e. – k A
(dT/dx) = 0, and since k and A are not zero, dT/dx must be zero).
B.C.(ii): at x = L, T = Tw
From Eq. 5.2:
dT - qg × x
= + C1
dx k

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 155


Then applying B.C.(i), we get: C1 = 0
From B.C.(ii) and Eq. 5.2:
qg × L2
C 2 = Tw +
2×k
Substituting for C1 and C2 in Eq. 5.2:
qg
T(x) = Tw + × (L 2 – x2 ) ...(5.13)
2× k
Eq. 5.13 gives the temperature distribution in a slab of thickness L, with heat generation when one side is
perfectly insulated. Fig. 5.4 shows the temperature distribution in the slab; note that temperature curve should
approach the left face horizontally, since (dT/dx) = 0 at x = 0.
Note that Eq. 5.13 is the same as Eq. 5.3, which was derived for a slab with heat generation, when both the
sides were maintained at the same temperature, except that now, L is the thickness of the slab and not half-
thickness.
In case of convection boundary condition:
Let the heat be lost from the un-insulated surface to a fluid at Ta , flowing on the surface with a heat transfer
coefficient of ha. Then, we relate Tw and Ta by making an energy balance at the right face. Since the left face is
insulated, all the heat generated in the slab travels to the surface on the right and gets convected away to the
fluid.
Heat generated in the slab:
Qgen = qg × A × L
Heat convected at surface:
Qconv = ha× A × (Tw – Ta)
Equating the heat generated and heat convected, we get:
qg × L
Tw = Ta + ...(a)
ha
Substituting from (a) in Eq. 5.13,
qg × L qg
T(x) = Ta + + × (L 2 – x 2 ) ...(5.14)
ha 2×k
Eq. 5.14 gives the temperature distribution in a slab with heat generation and constant k, insulated at one
face and losing heat at the other face to a fluid by convection, in terms of the fluid temperature.
Note: If the convection resistance is zero, which means that the heat transfer coefficient is infinity, the wall
temperature and the fluid temperature are the same, i.e. Ta = Tw , and Eq. 5.14 reduces to Eq. 5.13.
Maximum temperature:
Obviously, maximum temperature occurs at the insulated surface. This can be easily verified by differentiating
the expression for temperature distribution, Eq. 5.13, w.r.t. x and equating to zero. Putting x = 0 in Eqn. 5.13:
qg × L2
Tmax = Tw + ...(5.15)
2×k
Eq. 5.15 gives Tmax in terms of wall temperature, Tw.
Substituting for Tw from Eq. a in Eq. 5.15:
qg × L q g × L2
T max = Ta + + ...(5.16)
ha 2×k
Eq. 5.16 gives Tmax in terms of fluid temperature, Ta .
From Eq. 5.13 and 5.15, we can write:

T ( x ) - Tw L2 - x 2 x FG IJ 2

Tmax - Tw
=
L2
=1–
L H K ...(5.17)

156 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eqn. 5.17 gives non-dimensional temperature distribution for a slab with heat generation, and one face insu-
lated. This equation is the same as Eq. 5.5 for a slab with heat generation and both faces at the same temperature,
except that, now L is the thickness of the slab (In the case of Eq. 5.5, L was the half-thickness).
Example 5.1. Heat is generated uniformly in a stainless steel plate having k = 20 W/(mK). The thickness of the plate is 1
cm and heat generation rate is 500 MW/m3. If the two sides of the plate are maintained at 100°C and 200°C, respectively,
calculate:
(i) the temperature at the centre of the plate
(ii) the position and value of maximum temperature
(iii) heat transfer at the left and right faces
(iv) sketch the temperature profile in the slab.
Solution.
Data:
L := 0.01 m A := 1 m2 k := 20 W/(mc) qg := 500 ´ 106 W/m3 T1 := 200 C T2 := 100 C
See Fig. Example 5.1.
This is the case of one-dimensional, steady state conduction through a plate with heat generation, when the two
faces of the plate are maintained at different temperatures. So, we can directly apply Eq. 5.12 to get T(x) at any position
x.
However, let us solve this problem from first principles, and then verify the result from Eq. 5.12.
For this situation, governing differential eqution is:
d 2T q g k = 20 W/(mC)
+ = 0 ...(a) 6 3
dx 2 k Tmax qg = 500 ´ 10 W/m

dT q g × x Temperature distribution
Integrating: + = C1 ...(b)
dx k

- qg × x 2 T1 = 200°C
Integrating again: T(x) = + C1 × x + C2 ...(c)
2× k
Apply the B.C.’s: i.e. T2 = 100°C
(i) at x = 0: T1 = 200°C
(ii) at x = L = 0.01 m: T2 = 100°C
From B.C. (i) and Eq. c: C2 = 200 Xmax
From B.C. (ii) and Eq. c:
L
- qg × L2 X
T2 = + C1 × L + C 2
2×k FIGURE Example 5.1 Plane slab with internal heat
FT + q ×L 2
I generation, two sides at different temperature
GH 2 × k
2
g
- C2 JK
i.e. C1 := ...define C1
L
i.e. C1 = 1.15 ´ 10 5
Substituting for C1 and C2 in Eq. c:
- 500 ´ 106 × x 2
T(x) := + 1.15 ´ 10 5 × x + 200 ((d) (define T(x))
2 ´ 20
Eq. d gives the temperature profile.
Temperature at the centre of the plate:
Put x = 0.005 m in Eq. d:
i.e. T(0.005) = 462.5°C (temperature at the centre of plate.)
Verify: from direct formula Eq. 5.12:

T(x) := T1 + (L - x) ×
LM qg
+
OP
T2 - T1
x
MN 2× k L PQ ...(5.12)

Put x = 0.005 m: T(0.005) = 462.5 °C (verified.)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 157


Position and value of maximum temperature:
We have the relation for T(x) in Eq. d. Differentiate it w.r.t. x and equate to zero. Root of the resulting equation gives the
position, xmax of the location of maximum temperature. Then, substitute xmax back in Eq. d to get Tmax .
We have:
- 500 ´ 106 × x 2
T(x) := + 1.15 ´ 105 × x + 200 (Eq. d...define T(x))
2 × 20

FG d T(x)IJ
Let T¢(x) =
H dx K
Then,
- 500 ´ 10 6 × 2 × x
T¢(x) = + 1.15 ´ 105 ...(d)
2 × 20
Putting T¢(x) = 0 and solving:
1.15 ´ 10 5 ´ 2 ´ 20
x=
500 ´ 10 6 ´ 2
i.e. x = 4.6 ´ 10 – 3 m = 4.6 mm = xmax...position of maximum temperature from LHS.
i.e. xmax = 0.0046 m.
Verify: In Mathcad, there is no need to do the labour of differentiation, equating to zero and then solving for x, as done
above.
Instead, define T’(x) as the first derivative of T(x) w.r.t. x and use the ‘root function’ to find the root of T’(x) = 0: For
this, first, assume a trial (guess) value of x:
FG d T(x)IJ
T¢(x) :=
H dx K (define T¢(x))

x := 0.002 (trial value of x)


xmax := root (T¢(x), x) (define xmax as the root of equation T ¢(x) = 0)
i.e. xmax = 4.6 ´ 10 – 3 m (position of maximum temperature from LHS...verified.)
Value of maximum temperature:
This is obtained by putting the value of xmax in Eq. d:
i.e. put x = xmax in T(x):
Tmax := T(0.0046) °C (define Tmax)
i.e. Tmax = 464.5°C (value of maximum temperature.)
Heat transfer to left and right faces:
Knowing T(x), it is easy to find T’(x) = (dT/dx) at x = 0 and at x = L; We have already found out, in Eq. d, T’(x) – just put
x = 0 or x = L, as required. Then, apply Fourier’s law to get Q at x = 0 and x = L:
Heat transfer from left face, Q1:
Q1 := – k×A×T ¢(0) (define Q1...Fourier’s law)
i.e. Q1 = – 2.3 ´ 106 W/m2 = 2300 kW/m2 ...heat transfer from left face.
Note that negative sign indicates that heat flow is in a direction opposite to the positive X-direction, i.e. heat flow is
from right to left, as far as the left face is concerned. Heat is flowing from centre to left side in steady state.
Heat transfer from right face, Q2:
Q2 := – k×A×T¢(0.01)
i.e. Q2 = 2.7 ´ 106 W/m2 = 2700 kW/m2 (heat transfer from right face.)
Verify: Sum of Q1 and Q2 must be equal to the total heat generated, Qgen
Qgen := qg × A×L, W (define total heat generated)
i.e. Qgen =5 ´ 106, W (total heat generated)
Also, |Q1| + |Q2| = 5 ´ 106, W = Qgen (verified)
Note: Remember that absolute values of Q1 and Q2 are to be used, disregarding the signs, since the sign only indicates
the direction of heat flow.
To plot the temperature profile in the slab:
This is done very easily in Mathcad. First, define a range variable x, varying from 0 to 0.01 m, with an increment of
0.0005 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with x
and T(x), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig. Ex. 5.1(b).

158 FUNDAMENTALS OF HEAT AND MASS TRANSFER


x := 0, 0.0005, ... , 0.01 (define a range variable x starting value = 0,
next value = 0.0005 m and last value = 0.01 m)

Temperature profile-slab with heat generation


500

450
x in metres and
400
T(x) in deg.C
350
T(x)
300

250

200

150

100
0 0.002 0.004 0.006 0.008 0.01
x

FIGURE Example 5.1(b)

Note: It may be observed from the graph that the maximum k = 20 W/(mC) 6 3
temperature is 464.5°C and it occurs at x = 0.0046 m. qg = 500 ´ 10 W/m
Example 5.2. If in Example 5.1, the temperatures on either Tmax = 412.5°C
side of the plate are maintained at 100°C, calculate: Temperature
(i) the temperature on the centre line distribution (parabolic)
(ii) temperature at one-quarter of the thickness from the
surface Tw Tw = 100°C
(iii) draw the temperature profile.
Solution.
Data:
2L = 0.01 m L := 0.005 m A := 1 m2
k := 20 W/(m.c) qg := 500 ´ 106 W/m3 Tw := 100°C L L
See Fig. Example 5.2. X
This is the case of one-dimensional, steady state conduc-
tion through a plate with heat generation, when the two FIGURE Example 5.2 Plane slab with internal heat
faces of the plate are maintained at the same temperature. So, generation, both sides at the same temperature
we can directly apply Eq. 5.3 to get T(x) at any position x.
qg
i.e. T(x) := Tw + × (L2 – x 2 ) ...(5.3)
2× k
where, L is half-thickness of the slab.
Temperature at the centre line of plate:
At mid-plane, x = 0; therefore, substitute x = 0 in Eq. 5.3:
T(0) = 412.5°C (temperatue at the centre line of plate.)
T(0) = 412.5ºC is also the maximum temperature in the plate.
Temperature at one-quarter the thickness from the surface:
i.e. at x = 0.00025 m from the centre line. Put x = 0.00025 in Eq. 5.3:
T(0.00025) = 334.375°C (temperature at 1/4 of the thickness from surface.)
To draw the temperature profile:
We shall draw the temperature profile for the right half; by symmetry, temperature profile on the left half is the mirror
image of that on the right half.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 159


First, define a range variable x, varying from 0 to 0.005 m, with an increment of 0.00025 m. Then, choose x–y graph
from the graph palette, and fill up the place holders on the x-axis and y-axis with x and T(x), respectively. Click
anywhere outside the graph region, and immediately the graph appears. See Fig. Ex. 5.2(b).
x := 0, 0.00025, ... , 0.005 (define a range variable x..starting value = 0,
next value = 0.00025 m and last value 0.0005 m)

T(x) for plate with heat generation


500

450

400
x in metres and
350 T(x) in deg. C
T(x)
300

250

200

150

100
0 0.001 0.002 0.003 0.004 0.005
x

FIGURE Example 5.2(b)

Note: Above graph shows the temperature profile for the right half of a plate with internal heat generation, when both
the sides are maintained at 100°C. For the left side of the plate, temperature profile is identical, mirror image of this
graph.
Example 5.3. In Example 5.1, if the thermal conductivity of the material varies as: k(T) = ko(1 + b T), (W/(mC) where ko =
14.695 W/(mC) and b = 10.208 ´ 10 –4, (C –1), and T is in deg.C.
(i) calulate the temperature on the centre line
(ii) find location and value of maximum temperature in the plate
(iii) find heat transfer rate to the left and right sides, and
(iv) draw the temperature profile.
Solution. See Fig. Example 5.3.

k(T) = 14.695 (1 + 0.0010208 T) W/(mC)


Tmax = 479.9°C 6 3
qg = 500 ´ 10 W/m

Temperature distribution

T1 = 200°C

T2 = 100°C

Xmax
L
X

FIGURE Example 5.3 Plane slab with internal heat generation, variable k, with two sides at different
temperature

160 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Data:
L := 0.01 m A := 1 m2 k(T) := ko ×(1 + b ×T) ko := 14.965 W/(mC) b := 10.208 ´ 10 –4 (1/C)
qg := 500 ´ 10 6 W/m3 T1 := 200°C T2 := 100°C
We can directly use Eq. 5.13 to get the temperature at any location; however, let us work out this problem from
fundamentals and verify the result from Eq. 5.13.
We start with the governing differential equation for the case of a slab in steady state, one dimensional conduction
with heat generation and variable k, and integrate it twice in conjunction with the B’C.’s, to get the temperature profile:
We have:
d F dT I +q
dx GH
k(T )×
dx JK g = 0

dT
Integrating: k(T) × + qg × x = C1 where C1 is a constant.
dx

z z
Separating the variables and integrating:

k(T)d T = (C1 – qg × x)d x

Substituting for k(T):


z [ko × (1 + b×T)]d T =
zF (C1 – qg × x)d x

I
b ×T2 qg × x 2
i.e. T+
2
=
1
ko
GH
× C1 × x -
2
+ C2 JK ...(a)

Eq. a is the general equation for temperature distribution. Constants C1 and C2 are obtained by applying the B.C.’s:
B.C.(i): at x = 0, T = T1
B.C.(ii): at x = L, T = T2
F b × T12 I
From B.C.(i) and Eq. a C2 := ko × T1 + GH 2
JK
substituting C2 = 3.239 ´ 10 3
From B.C.(ii) and Eq. a

F b × T22 I
qg × L2
GH
k o × T2 +
2
+ JK2
- C2
C1 :=
L
substituting, we get C1 = 2.331 ´ 10 6.
Substituting value of C1 and C2 in Eq. a and simplifying, we get:
b ×T 2
+ T – (1.58595 ´ 105 × x – 1.70126 ´ 10 7×x 2 + 220.416) = 0 ...(b)
2
Eq. b is a quadratic in T. Its positive root is given by:

10. 208 ´ 10- 4


- 1 + 1 + 4× × (1. 58595 ´ 105 × x - 1.70126 ´ 107 × x 2 + 220. 416)
2
T(x) := ...(c)
10. 208 ´ 10- 4

2
Eq. c gives the variation of temperature with x.
Temperature at the centre line:
Put x = 0.005 m in Eq. c
T(0.005) = 473.597°C (temperature at centre line.)
Verify: verify this result from direct formula, eqn. 5.12, a.
T1 + T2
Tm = i.e. Tm = 150°C (mean value of temperature)
2

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 161


T(x) :=
-1 FG 1 + T IJ 2
2 ×x qg × x
i.e.
b
+
Hb K 1 -
b ×L
× (T1 - T2 ) × (1 + b × Tm ) +
b × ko
× (L - x ) ...5.12(a)

i.e. T(0.005) = 473.596°C (verified.)


Location and value of maximum temperature in the plate:
Differentiate Eq. c w.r.t. x and equate to zero and get xmax, the position of Tmax ; substitute this value of xmax back in Eq.
c to get value of Tmax.
In Mathcad, we do not have to go through the labour of differentiation, equating to zero, then solving etc. We use
the ‘root function’. First, define T’(x) = d(T(x))/dx. Then, assume a trial value of x and type the command ‘root(T’(x),
xtrial) = ‘. This gives root of T’(x) = 0.

10.208 ´ 10 - 4
-1+ 1 + 4× × (1. 58595 ´ 10 5 × x - 1.70126 ´ 107 × x 2 + 220 . 416)
2
T(x) := ...define T(x)...(c)
10. 208 ´ 10 - 4

2

d
T¢(x) := T(x) ...define T¢(x)
dx
x := 0.002 m (trial value of x)
xmax := root (T ¢(x), x) (define xmax)
i.e. xmax = 4.661 ´ 10 – 3 m = 4.661 mm (location of maximum temperature...distance from LHS.)
Value of maximum temperature is obtained by putting x = xmax in Eq. c
i.e. T(xmax ) = 474.913°C (value of maximum temperature)
Heat transfer to left and right faces:
Knowing T(x), it is easy to find T’(x) = (dT/dx) at x = 0 and at x = L; We have already defined T’(x)—just put x = 0 or x
= L, as required. Then, apply Fourier’s law to get Q at x = 0 and x = L:
Remember k(T) := ko × (1 + b ×T) (define k(T))
Heat transfer from left face, Q1:
Q1 := – k(T1) × A ×T ¢(0) (define Q1...Fourier’s law)
i.e. Q1 = – 2.331 ´ 106 W/m 2 = – 2331 kW/m2 (heat transfer from left face)
Note: Negative sign indicates that heat is flowing from right to left, i.e. in the negative X-direction.
Check: This should equal the amount of heat generated between x = 0 and x = xmax
Heat generated between x = 0 and x = xmax :
Qgen1 := qg × A× (xmax – 0)
i.e. Qgen1 = 2.331 ´ 10 6 W/m 2 (verified.)
Heat transfer from right face, Q2:
Q2 := – k(T2) × A× T ¢(0.001) (define Q2...Fourier’s law)
i.e. Q2 = 2.669 ´ 106 W/m2 = 2669 kW/m2 (heat transfer from left face)
Check: This should equal the amount of heat generated between x = xmax and x = L.
Heat generated between x = 0 and x = x max and x = L
Qgen2 := q g × A × (L – x max)
i.e. Qgen2 = 2.669 ´ 106 W/m 2 (verified.)
To plot the temperature profile in the plate:
This is done very easily in Mathcad. First, define a range variable x, varying from 0 to 0.01 m, with an increment of
0.0005 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with x
and T(x), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig. Ex. 5.3(b).
x := 0, 0.0005, ... , 0.01 (define a range variable x..starting value = 0,
next value = 0.0005 m and last value = 0.01 m)
Note: It may be observed from the graph that the maximum temperature is 474.9°C and it occurs at x = 0.00466 m.
Example 5.4. A plane wall of thickness 0.1 m and k = 25 W/(mK), having uniform volumetric heat generation of 0.3
MW/m3 is insulated on one side and is exposed to a fluid at 92°C. The convective heat transfer coefficient between the
wall and the fluid is 500 W/(m2K). Determine:
(i) the maximum temperature in the wall
(ii) temperature at the surface exposed to the fluid
(iii) Draw the temperature profile.

162 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Temperature profile-slab with heat generation.
Variable k(T)
500
x in metres and
450 T(x) in deg.C

400

350
T(x)
300

250

200

150

100
0 0.002 0.004 0.006 0.008 0.01
x

FIGURE Example 5.3(b)

Solution. See Fig. Example 5.4. X k = 25 W/(mK)


Data: 3
L := 0.01 m A := 1 m2 k := 25 W/(mK) Insulated qg = 0.3 MW/m
qg := 0.3 ´ 106 W/m3 Ta := 92°C ha := 500 W/(m 2K)
Temperature
By observation, we know that maximum temperature Tmax distribution
occurs on the insulated wall; this is so because, the heat gener-
ated in the wall is constrained to flow from left to right since the 2
left face is insulated and for this to occur, temperature on the Tw ha = 500 W/(m K)
left must be higher than that on the right.
We can directly apply Eq. 5.14 and put x = 0 in that
equation to get Tmax . Ta = 92°C
We have, from eqn. 5.14
qg × L qg
T(x) := Ta + + ×(L 2 – x 2) ...(5.14) L = 0.1 m
ha 2× k
X
Maximum temperature in the wall: (occurs at the insulated
surface i.e. at x = 0) FIGURE Example 5.4 Plane slabe with internal
Put x = 0 in Eq. (5.14): heat generation, one side is insulated
T(0) = 212°C (maximum temperature in the wall,
occurs on the insulated left surface.)
Temperature at the surface exposed to the fluid:
Put x = 0.1 in Eq. (5.14):
T(0.1) = 152°C (temperature at the surface exposed to the fluid.)
To draw the temperature profile:
Using Mathcad, this is very easy. First, define a range variable x, varying from 0 to 0.1 m, with an increment of 0.005 m.
Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with x and T(x),
respectively. Click anywhere outside the graph region, and immediately the graph appears. Fig. Ex. 5.4(b).
x := 0,. 0.005, ... , 0.1 (define a range variable x...starting value = 0,
next value = 0.005 m and last value = 0.1 m)
Note: It may be observed from the graph that the maximum temperature is 212°C and it occurs at x = 0 and at x = 0.1 m
the temperature is 152°C.
Example 5.5. The exposed surface (x = 0) of a plane wall of thermal conductivity k, is subjected to microwave radiation
that causes volumetric heating to vary as: qg (x) = qo (1 – x/L) where qo (W/m3) is a constant. The boundary at x = L is

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 163


Temperature profile-slab with heat generation
220

210

200 x in metres and


T(x) in deg.C
190
T(x)
180

170

160

150

140
0 0.02 0.04 0.06 0.08 0.1
x

FIGURE Example 5.4(b)

k 3 perfectly insulated, while the exposed surface is main-


qg = qo(1 – x/L), W/m tained at a constant temperature, To. Determine the tem-
perature distribution T(x) in terms of x, L, k, q o and To.
To
Insulated Solution. See Fig. Example 5.5.
Here, the heat generation rate is not uniform
Microwave throughout the volume, but varies with position.
For the assumption of one-dimensional, steady state
conduction with constant k, and the internal heat genera-
tion at the specified rate, the governing differential equa-
tion is:
d 2 T qg
+ = 0 ...(a)
L dx 2 k
d 2T q o FGx IJ
X Substitute for qg:
dx 2
+ × 1-
k HL K = 0

FIGURE Example 5.5 Plane slab with variable heat


dT qo × x qo × x 2
generation rate, one side insulated Integrating, + – = C1 ...(b)
dx k 2 ×k ×L
qo × x 2 q × x3
Again, integrating, T(x) + – o = C1 x + C2 ...(c)
2× k 6 ×k × L
where, C 1 and C2 are constants of integration.
Eq. c gives the temperature distribution. C 1 and C2 are obtained by applying the boundary conditions:
B.C. (i): at x = 0, T = To
B.C. (ii): at x = L, dT/dx = 0, since right face is insulated
B.C. (i) and Eq. c gives: C2 = To
qo × L
B.C. (ii) and Eq. b gives: C1 =
2×k
Substituting C 1 and C2 back in Eq. c:
qo × x 3 qo × x 2 qo × L × x
T(x) = - + + To
6 ×k ×L 2 ×k 2 ×k

F
qo × L × x x 1 x2 I
i.e. T(x) = To +
2× k GH× 1- + × 2
L 3 L JK ...(d)

164 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eq. d is the desired relation for temperature distribution as a function of x, L, k, qo and To.
Example 5.6. A copper conductor (k = 380 W/(mC), r = 2 ´ 10 –8 ohm ´ m), 8 mm diameter and 1 m long, connects two
large plates. One face is maintained at 30°C and the other face, at 50°C. Space between the plates is filled with an
insulation.
(i) What is the maximum temperature and its location if the maximum current flowing is 150 A?
(ii) Calculate the heat dissipation to LHS and RHS
(iii) Draw the temperature profile.
Solution. See Figure Example 5.6.

Insulated
d = 8 mm, k = 380 W/(mC)

T1 = 30°C
T2 = 50°C

L=1m

FIGURE Example 5.6 Rod connected between two plates

Data:
L := 1.0 m d := 0.008 m T1 := 30°C T2 := 50°C k := 380 W/(mC) r := 2 ´ 10 – 8 Ohm ´ m
p × d2
I := 150 Amp A := m 2 i.e. A := 5.027 ´ 10 – 5 m 2
4
Obviously, maximum temperature will occur at a location nearer to the end at 50°C.
Since the bar is laterally insulated, it is a case of one-dimensional conduction in the X-direction, in steady state,
with heat generation and constant k.
So, the controlling differential equation is:
d 2T q g
+ = 0 ...(a)
dx 2 k

dT qg × x
Integrating: + = C1
dx k

- qg × x 2
Integrating again: T(x) = + C1 × x + C2 ...(b)
2× k
Eq. b gives the temperature distribution in the bar.
Apply the B.C.’s to get C1 and C2, the constants of integration.
B.C. (i): at x = 0, T = 30°C
B.C. (ii): at x = 1 m, T = 50°C
To calculate q g :
Q I2 ×R
qg = = , where Q is the heat generated.
Volume Volume
r ×L
Resistance R: R :=
A
i.e. R = 3.979 ´ 10 – 4 Ohm (resistance of the rod)
I ×R
2
Therefore, qg := W/m3 (heat generation rate due to Joule heating)
A×L

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 165


i.e. qg = 1.78104 ´ 10 5 W/m 3 (heat generated rate)
B.C.(i) and Eq. b gives: C2 = 30
qg × L2
50 + - C2
2× k
B.C.(ii) and Eq. b gives: C1 :=
L
Substituting and simplifying C1 = 254.347
Substituting C1 and C2 back in Eq. b
- (1.78104 ´ 10 5 ) × x 2
T(x) = + 254.347 × x + 30
2 ´ 380
i.e. T(x) = – 234.347× x 2 + 254.347×x + 30 ...(c)
Eq. c is the desired eqn. c for temperature distribution.
Location and value of maximum temperature:
Location of maximum temperature is obtained by differentiating Eq. c w.r.t. x and equating to zero:
i.e. – 234.347× (2x) + 254.347 = 0
254.347
i.e. x :=
234. 347 ´ 2
i.e. x = 0.543 m (location of maximum temperature...(this is the distance from LHS))
Value of maximum temperature:
Substitute this value of x in Eq c
T(0.543) = 99.013°C (value of maximum temperature)
Heat dissipated to LHS and RHS:
Since the temperature profile is known, get T’(x) = dT(x)/dx at x = 0 and x = L, and then apply Fourier’s law at x = 0 and
x = L, to get Qleft and Qright:
d
T¢(x) := T(x) (define T ¢(x), the first derivative of T (x) w.r.t. x)
dx
Therefore, T¢(0) = 254.347 C/m (dT/dx at x = 0, ...i.e. at LHS)
and, T ¢(1) = – 214.347 C/m (dT/dx at x = 1 m, ...i.e. at RHS)
So, we have:
Qleft := – k×A×T ¢(0) W (define Qleft)
i.e. Qleft = – 4.858 w (heat dissipated from left end.)
Note: Negative sign of Q indicates that heat is flowing from right to left, i.e. in negative X-direction.
And, Qright := – k × A × T ¢(1), W (define Qright)
i.e. Qright = 4.094 W (heat dissipated from right end.)
Check: Sum of the heat dissipated from left and right ends must be equal to the total heat generated in the bar:
Qtot := |Qleft| + |Qright | W (define Qtot)
i.e. Qtot = 8.952 W (total heat dissipated)
Now, Qgen := I 2 R W (heat generated by Joule heating)
i.e. Qgen = 8.952 W (checks with Qtot.)
To draw the temperature profile:
First, define a range variable x, varying from 0 to 1 m, with an increment of 0.01 m. Then, choose x–y graph from the
graph palette, and fill up the place holders on the x-axis and y-axis with x and T(x), respectively. Click anywhere outside
the graph region, and immediately the graph appears. See Fig. Ex. 5.6(b).
x := 0, 0.01, ... , 1 (define a range variable x...starting value = 0,
next value = 0.01 m and last value = 1 m)
Note from the graph that maximum temperature of 99.01°C is reached at x = 0.543 m, i.e. beyond the mid-point,
towards the right end.

5.3 Cylinder with Uniform Internal Heat Generation


There are several applications of cylindrical geometry with internal heat generation, e.g. current carrying
conductors, nuclear fuel rods, chemical reactors, etc. We shall consider solid cylinders as well as hollow cylinders
with different types of boundary conditions.

166 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T(x) for a bar with heat generation
100
x in metres and
T(x) in deg.C
80

60
T(x)

40

20

0
0 0.2 0.4 0.6 0.8 1
x

FIGURE Example 5.6(b)

5.3.1 Solid Cylinder with Internal Heat Generation


Consider a solid cylinder of radius, R and length, L. There is uniform heat generation within its volume at a rate
of qg (W/m3). Let the thermal conductivity, k be constant.
See Fig. 5.5.
We would like to analyse this system for temperature distribution and maximum temperature attained.

Temperature profile,
parabolic
To
k, qg
Tw
Q

R
L

FIGURE 5.5(a) Cylindrical system with heat FIGURE 5.5(b) Variation of temperature along the
generation radius

Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Uniform internal heat generation rate, qg (W/m3).
With the above stipulations, the general differential equation in cylindrical coordinates (see Eq. 3.17) reduces
to:
d 2T 1 dT qg
2
+ × + =0 ...(a)
dr r dr k

d 2T dT q g × r
Multiplying by r: r × 2
+ + =0
dr dr k

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 167


d F dT I - qg ×r
i.e.
dr GH

dr JK =
k

dT - qg × r 2
Integrating: r× = + C1
dr 2×k

dT - qg × r C1
i.e. = + ...(b)
dr 2×k r

- qg × r 2
Integrating again: T(r) = + C1 × ln (r) + C2 ...(5.18)
4×k
Eq. 5.18 is the general relation for temperature distribution along the radius, for a cylindrical system, with
uniform heat generation.
C1 and C2, the constants of integration are obtained by applying the boundary conditions.
(Remember Eq. 5.18, since the same equation will be the starting point in the analysis of hollow cylin-
ders too, with different boundary conditions.)
In the present case, B.C.’s are:
B.C. (i): at r = 0, dT/dr = 0, i.e. at the centre of the cylinder, temperature is finite and maximum (i.e. To =
Tmax) because of symmetry (heat flows from inside to outside radially).
B.C. (ii): at r = R, i.e. at the surface, T = Tw
From B.C. (i) and Eq. b, we get: C1 = 0
From B.C. (ii) and Eq. 5.18, we get:
- qg × R 2
Tw = + C2
4× k

qg × R 2
i.e. C2 = Tw +
4×k
Substituting C1 and C2 in Eq. 5.18
- qg × r 2 qg × R 2
T(r) = + Tw +
4×k 4×k
qg
i.e. T(r) = Tw + × (R2 – r 2) ...(5.19)
4× k
Eq. 5.19 is the relation for temperature distribution in terms of the surface temperature, Tw . Note that this is
a parabolic temperature profile, as shown in Fig. 5.11(b).
Maximum temperature:
Maximum temperature occurs at the centre, because of symmetry considerations (i.e. heat flows from the centre
radially outward in all directions; therefore, temperature at the centre must be a maximum.)
Therefore, putting r = 0 in Eq. 5.19:
qg × R 2
Tmax = Tw + ...(5.20)
4×k
From Eq. 5.19 and 5.20,

T - Tw r FG IJ 2

Tmax - Tw
=1–
R H K ...(5.21)

Eq. 5.21 is the non-dimensional temperature distribution for the solid cylinder with heat generation.

168 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Convection boundary condition:
In many practical applications, heat is carried away at the boundaries by a fluid at a temperature Ta flowing on
the surface with a convective heat transfer coefficient, h (e.g. current carrying wire cooled by ambient air). Then,
mostly, it is the fluid temperature that is known and not the surface temperature, Tw , of the cylinder. In such
cases, we relate the wall temperature and fluid temperature by an energy balance at the surface, i.e. heat
generated and conducted from within the body to the surface is equal to the heat convected away by the fluid at
the surface.
i.e. p × R 2 × L × qg = h × (2 × p × R × L) × (Tw – Ta )
qg × R
i.e. Tw = Ta + ...(c)
2× h
Substituting c in Eq. 5.19:
qg × R qg
T(r) = Ta + + × (R2 – r 2 ) ...(5.22)
2 ×h 4× k
Again, for maximum temperature put r = 0 in Eq. 5.22:
qg × R qg × R 2
Tmax = Ta + + ...(5.23)
2×h 4× k
Eq. 5.23 gives maximum temperature in the solid cylinder in terms of the fluid temperature, Ta .
5.3.1.1 Alternative analysis. In the alternative method,
which is simpler, instead of starting with the general Tw
differential equation, we derive the above equations from dr
physical considerations. See Fig. 5.6.
Let us write an energy balance with an understanding Q
r
that at any radius r, the amount of heat generated in the
volume within r = 0 and r = r, must move outward by To
conduction.
i.e. at any radius r, we write the energy balance:
dT
qg × p × r 2 ×L = – k ×(2 ×p ×r × L) × ...(a)
dr
- qg R
dT = × r × dr
2× k
FIGURE 5.6 Solid cylinder with heat generation

z z
Integrating:
- qg
dT = rdr
2×k

- qg × r 2
i.e. T(r) = +C ...(b)
4×k
Eq. b gives the temperature distribution along the radius.
Get the constant of integration, C from the B.C.: at r = R, T = Tw
qg × R 2
i.e. C = Tw +
4× k
Substituting C back in Eq. b:
- qg × r 2 qg × R 2
T(r) = + Tw +
4×k 4× k

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 169


qg
i.e. T(r) = T w + × (R 2 – r 2) ...(c)
4× k
Eq. c gives the temperature distribution along the radius, in terms of the surface temperature of the cylinder.
Note that Eq. c is the same as Eq. 5.19 derived earlier.
In many applications, temperature drop between the centre line (where maximum temperature occurs) and
the surface is important (e.g. in nuclear fuel rods, to ensure that the fuel rod does not melt). Then, from Eq. c,
putting r = 0:
qg × R 2
To – T w = ...(5.24)
4× k
Eq. 5.24 is important; it gives the maximum temperature difference in a solid cylinder with heat generation.
Knowing Tw , one can easily find out To (= Tmax).
Compare Eq. 5.24 with Eq. 5.9, derived earlier for the maximum temperature difference in a slab with uni-
form heat generation.
5.3.1.2 Analysis with variable thermal conductivity. In the above analysis, thermal conductivity of the material was
assumed to be constant. Now, let us make an analysis when the thermal conductivity varies linearly with tem-
perature as:
k(T) = ko (1 + b T ),
where, ko and b are constants.
Again, considering Fig. 5.6, we have from heat balance (see Eq. a above):
dT
qg × p × r 2 × L = – k(T) × (2 × p ×r ×L) × ...(a)
dr
- qg
i.e. k(T) × dT = × r × dr
2
Substituting for k(T) and integrating:

z ko ×(1 + b× T)dT =
- qg
2
×
z r dr

b ×T 2 - qg × r 2
i.e. T+ = +C ...(e)
2 4 × ko
C is determined from the B.C.: at r = 0, T = To
We get:

b ×To2
C = To +
2
Substituting C in Eq. e:

b ×T 2 qg × r 2 b ×To2
+T+ – To – =0 ...(f)
2 4 × ko 2
Eq. f is a quadratic in T. Its positive root is given by:

F
b qg × r
2
b × To2 I
-1+ 1 - 4× × GGH
2 4 × ko
- To -
2
JJK
T(r) =
b

2

170 FUNDAMENTALS OF HEAT AND MASS TRANSFER


-1 F 1 +T 2 × To I qg × r 2
i.e. T(r) =
b
+ GH b 2 o
2
+
b JK
-
2 × b × ko

-1 FG 1 + T IJ 2
qg ×r 2
i.e. T(r) =
b
+
Hb K o -
2 × b × ko
...(5.25)

Eq. 5.25 gives temperature distribution in a solid cylinder with internal heat generation and linearly varying
k. Compare this equation with that obtained for a slab, with temperature at either side being the same, i.e. Eq.
5.10.
Eq. 5.25 gives T(r) in terms of To (i.e. Tmax at r = 0).
If we need T(r) in terms of Tw : in Eq. e, C is determined from:
B.C.: at r = R, T = Tw
we get:

b ×Tw2 qg × R 2
C = Tw + +
2 4 × ko
Substituting C in Eq. e, we get a quadratic in T, and solving we get, for temperature distribution:

-1 FG 1 + T IJ 2
(R 2 - r 2 )
T(r) =
b
+
Hb K w + qg ×
2 × b × ko
...(5.26)

5.3.1.3 Current carrying conductor. This is a very important practical application. Cooling of current carrying
conductors enhances their current carrying capacity. Knowledge of temperature distribution is required to make
sure that temperatures leading to ‘burn out’ of the conductor are not reached. Conductors have to operate safely
in superconducting magnets, transformers, motors and electrical machinery, since sudden failure of conductor
may lead to conditions that are unsafe to the operator as well as the machine.
In the case of current carrying conductors, uniform internal heat generation occurs due to Joule heating.
Consider a conductor of cross-sectional area, Ac and length, L. Let the current carried be I (A). Let the
electrical resistivity of the material be r (Ohm ´ m).
Then, heat generated per unit volume = Qg/Volume of conductor,
where, Qg is the total heat generated (W).
Qg = I 2× R where R = electrical resistance of wire, (Ohms)
r ×L
But, R=
Ac

FG r × L IJ
I2 ×
I2 × R H A K = FG I IJ
c
2
Therefore, qg =
Ac × L
=
A ×Lc HA K c
× r, W/m3

i = I/Ac , is known as the ‘current density’. Note its units: A/m2


i2 1
i.e. qg = i 2×r = where ke = = electrical conductivity, (Ohm m) – 1
ke r
Therefore, temperature distribution in a current carrying wire (of solid, cylindrical shape) is given by Eq.
5.19, viz.
qg
T(r) = Tw + ×(R2 – r 2 ) ...(5.19)
4× k

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 171


Substituting for qg , we get:
i2 × r
T(r) = Tw + × (R2 – r 2) ...(5.19a)
4 ×k
Eq. 5.19a gives the temperature distribution in the current carrying wire, in terms of the surface temperature,
Tw . Maximum temperature, which occurs at the centre, is obtained by putting r = 0 in Eq. 5.19 a. i.e.
i 2 × r × R2
Tmax = Tw + ....(5.20a)
4×k
And, from Eqs. 5.19a and 5.20a, we get:

T - Tw r FG IJ 2

Tmax - Tw
=1–
R H K
Note that the above equation for non-dimensional temperature distribution in a current carrying wire is the
same as Eq. 5.21.
Example 5.7. (a) A 3.2 mm diameter stainless steel wire, 30 cm long has a voltage of 10 V impressed on it. The outer
surface temperature of the wire is maintained at 93°C. Calculate the centre temperature of the wire. Take the resistivity
of the wire as 70 micro-ohm ´ cm and the thermal conductivity as 22.5 W/(mK).
(b) The heated wire in the above example is submerged in a fluid maintained at 93°C. The convection heat transfer
coefficient is 5.7 kW/(m2K). Calculate the centre temperature of the wire.
Solution. See Figure Example 5.7.

d = 3.2 mm, k = 22.5 W/(mC) d = 3.2 mm, k = 22.5 W/(mC)


–8 –8
r = 70 ´ 10 Ohm ´ m 2 r = 70 ´ 10 Ohm ´ m
ha = 5700 W/(m K)
Tw = 93°C
Ta = 93°C

L = 0.3 m L = 0.3 m
E = 10 V E = 10 V

FIGURE Example 5.7(a) Wire with an impressed FIGURE Example 5.7(b) Wire with an impressed
voltage, Tw is known voltage, Ta is known

Data:
do
do := 0.0032 m R := m i.e. R = 1.6 ´ 10 –3 m L := 0.3 m r := 70 ´ 10 –8 Ohm ´ m k := 22.5 W/(mC)
2
do2 L
Tw := 93°C Ta := 93°C h := 5700 W/(m 2 C) E := 10 V A := p × , m2 Resistance := r × Ohm
4 A
i.e . Resistance = 0.026 Ohm (electrical resistance of the wire)
E2
P := W (define power generated due to current flow)
Resistance
3
i.e. P = 3.83 ´ 10 W ...power generated
P
qg := W/m3 (define the internal heat generation rate)
A ×L
i.e. qg := 1.587 ´ 10 9 W/m3 (the internal heat generation rate)
Case (a): Wire surface temperature is given;
To calculate centre temperature (i.e. maximum temperature):
We have, from Eq. 5.20:

172 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1
Tmax := Tw + qg × R 2 ×
4×k
i.e. Tmax = 138.15°C ...centre temperature of wire
Case (b): Wire submerged in a fluid;
To calculate the centre temperature (i.e. maximum temperature):
We have, from Eq. 5.23:
qg × R qg
Tmax := Ta + + × R2
2× h 4×k
i.e. Tmax = 360.929°C ...centre temperature of wire
Example 5.8. Meat rolls of 25 mm diameter, k = 1 W/(mC) are heated by microwave heating. Centre temperature of the
roll is 90°C. Surrounding temperature is at 30°C. Heat transfer coefficient at the surface is 25 W/(m2C). Find the micro-
wave heating capacity required in W/m3.
Solution.
Data:
R := 0.0125 m k := 1 W/(mC) To := 90°C Ta := 30°C h := 25 W/(m 2 C)

2 R = 12.5 mm, k = 1 W/(mC)


ha = 25 W/(m K)

To = 90°C Ta = 30°C

FIGURE Example 5.8 Microwave heating of meat roll

Remember that for the cylindrical roll, maximum temperature occurs at the centre.
We have, from Eq. 5.23:
qg × R qg × R 2
To = Ta + +
2× h 4 ×k

FR +R I 2
i.e. To = Ta + qg × GH 2× h 4 × k JK
To - Ta
i.e. qg :=
FR +R I 2
W/m 3 (define qg)
GH 2 × h 4 × k JK
i.e. qg = 2.07 ´ 105 W/m3 (= 207.6 kW/m3 ...required microwave heating capacity.)
Example 5.9. A long cylindrical rod of diameter 200 mm with k = 0.5 W/(mK) experiences uniform volumetric heat
generation of 24,000 W/m3. The rod is encapsulated by a circular sleeve having an outer diameter of 400 mm and k of 4
W/(mK). Outer surface of the sleeve is exposed to cross flow of air at 27°C with convection coefficient of 25 W/(m2K).
(i) Find the temperature at the interface between the rod and the sleeve and on the outer surface.
(ii) What is the temperature at the centre of the rod?
(iii) What is the temperature at mid-radius of the rod?
(iii) Sketch the temperature distribution.
Solution. See Figure Example 5.9.
Data:
R1 := 0.1 m R2 := 0.2 m L := 1 m k1 := 0.5 W/(mK) k2 := 4 W/(mK) Ta := 27°C
ha := 25 W/(m2 K) qg := 24000 W/m 3
Let T0, T1, and T2 be the centre temperature of the rod, interface temperature between the rod and the sleeve, and
the outer surface temperature of sleeve, respectively, as shown in Fig. 5.16.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 173


k2 = 4 W/(mK) For the inner cylindrical rod with heat generation, we have, from
Eq. 5.20:
k1 = 0.5 W/(mK)
Q 3 qg × R12
qg = 24 kW/m To = T1 + ...(a)
4 × k1
2
ha = 25 W/(m K) Here, however, T1 is not presently known. But T1 is related to the
Ta = 27°C known ambient temperature Ta by considering the steady state heat
transfer from the inner cylinder through the outer sleeve by conduction
and then to the ambient fluid by convection.
T1
T2 T1 - Ta
i.e. Q= ...(b)
Rsleeve + Rconv
d1 = 0.2 m where, Rsleeve = thermal resistance of sleeve, and Rconv = convective
resistance on outer surface of sleeve
d2 = 0.4 m First, find Q, the steady state heat transfer rate = heat generation
rate in the cylinder
FIGURE Example 5.9 Encapsulated rod 2
with heat generation i.e. Q := p× R1 ×L ×qg W (define Q, the total heat generated)
i.e. Q = 753.982 W (total heat generated rate.)
Thermal resistances:
FG R IJ
HR K
2
ln
1
Rsleeve := C/W (define thermal resistance of sleeve)
2 × p × k2 × L
i.e. Rsleeve = 0.028 C/W (thermal resistance of sleeve)
1
and, Rconv = C/W (convective resistance on the outer surface of sleeve)
ha × ( 2 × p × R2 × L)
i.e. Rconv = 0.032 C/W (convective resistance on the outer surface of sleeve.)
Temperatures T1, T2 and To :
From Eq. b:
T1 := Q× (Rsleeve + Rconv) + Ta °C (define T1 the interface temperature
between cylinder and sleeve)
i.e. T1 = 71.794°C (the interface temperature between
cylinder and sleeve.)
To find T2:
T1 - T2
We have: Q= W ...applying Ohm’s law to the sleeve
Rsleeve
i.e. T2 := T1 – Q ×Rsleeve, °C ...define T2
i.e. T2 = 51°C (temperature on the outer surface of sleeve.)
To find To :
From Eq. a:
qg × R12
To := T1 + °C (define To, the centre temperature)
4 × k1
i.e. To = 191.794°C (centre temperature of cylinder.)
Temperature at the mid-radius of the rod, i.e. at r = 0.05 m:
For a cylinder with heat generation, temperature distribution is given by Eq. 5.19:
qg
i.e. T(r) = Tw + × (R2 – r 2) ...(5.19)
4 ×k
For the present case, this equation becomes:
qg
× ( R1 – r 2)
2
T(r) := T1 + (define T(r))
4 × k1
Therefore, T(0.05) = 161.794°C (temperature at mid-radius of the rod.)

174 FUNDAMENTALS OF HEAT AND MASS TRANSFER


To sketch the temperature profile:
Temperature profile for the rod with heat generation is given by:
qg 2
T(r) = T1 + × ( R1 – r 2 )
4 × k1
And, temperature profile for the cylindrical shell of sleeve (with no heat generation) is given by Eq. 4.34, i.e.

T2 - T1 FrI
F R I GH R JK
s
t(rs) = T1 + × ln where rs = any radius within the sleeve.
ln G J
1

HR K
2

To sketch the temperature profile in the rod, define a range variable r, varying from 0 to 0.1 m, with an increment
of 0.005 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with r
and T(r), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig. Ex. 5.9(b).
r := 0, 0.005, ... , 0.1 (define a range variable r..starting value = 0,
next value = 0.005 m and last value = 0.1 m)
To sketch the temperature profile in the sleeve, define a range variable r s, varying from 0.1 to 0.2 m, with an
increment of 0.005 m. Then, in the above graph, on the x-axis place holder, put a comma after r, and enter rs and on the
y-axis place holder, put a comma after T(r) and enter t(rs). Click anywhere outside the graph region, and immediately
both the graphs appear.
rs := 0.1, 0.105, ... , 0.2 (define a range variable rs..starting value = 0.1,
next value = 0.105 m and last value = 0.2 m)

T(r) for rod with heat generation and sleeve


200
190
180
170 r, rs in metres and
160 T(r), t(rs) in deg.C
150
T(r) 140
t(rs) 130
120
110
100
90
80
70
60
50 0 0 0 0
0 .02 .04 .06 .08 0.1 0.12 0.14 0.16 0.18 0.2

r, rs
FIGURE Example 5.9(b)

In the above figure from r = 0 to r = 0.1 m, the graph shows the temperature profile within the solid rod with
internal heat generation; from the radius of 0.1 m to 0.2 m, the graph shows the temperature profile within the
cylindrical sleeve placed over the rod.
Note that at r = 0, To = 191.8°C, at r = 0.1 m, T1 = 71.8°C and at r = 0.2 m, T2 = 51°C.

5.3.2 Hollow Cylinder with Heat Generation


Hollow cylinder geometry has significant practical applications. Many times, nuclear fuel rods are made of hol-
low cylinder geometry where the heat generated is carried away by a (liquid metal) coolant flowing either on the
inside or outside the tubes. Hollow electrical conductors of cylindrical shape are used for high current carrying
applications, where again, cooling is done by a fluid flowing on the inside. There are annular reactors, insulated
either from inside or outside, used in chemical processes.
We shall study heat transfer in a hollow cylindrical system, with different boundary conditions.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 175


5.3.2.1 Hollow cylinder with the inside surface insulated. Consider steady state, one-dimensional heat transfer in a
hollow cylinder of length L, inside radius ri and outside radius ro , with a uniform internal heat generation rate of
qg (W/m3). Thermal conductivity, k is constant. Let the inside surface be perfectly insulated; that means, all the
heat generated in the cylindrical shell has to move only outwards, in the positive r-direction. Let the
temperatures on the inside and outside surfaces be T i and To respectively. See Fig. 5.7.
Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Uniform internal heat generation rate, qg (W/m3).
With the above stipulations, the general differential equation in cylindrical coordinates (see Eq. 3.17) reduces
to:
d 2T1 dT qg
+ ×
2
+ =0 ...(a)
dr r dr k
Integrating Eq. a twice, we get the general solution for temperature distribution, namely, Eq. 5.18, as done in
section 5.3.1:
- qg × r 2
T(r) = + C1 × ln(r) + C2 ...(5.18)
4×k

Insulated k, qg Insulated
Q k, qg Q
dr

Ti To Ti To

ri ri

ro ro

FIGURE 5.7 Hollow cylinder with heat FIGURE 5.8 Hollow cylinder with heat
generation, inside surface insulate generation, inside surface insulated

Eq. 5.18 is the general relation for temperature distribution along the radius, for a cylindrical system, with
uniform heat generation.
C1 and C2, the constants of integration are obtained by applying the boundary conditions.
In the present case the B.C.’s are:
B.C.(i): at r = ri T = Ti and dT/dx = 0 (since inner surface is insulated), and
B.C.(ii): at r = ro T = To
Get C1 and C2 from these B.C.’s and substitute back in Eq. 5.18 to get the temperature distribution. This is left
as an exercise for the student (See Example 5.11 for procedure of working out a numerical problem).
We shall, however, derive the expression for temperature distribution by a simpler method of physical
consideration and heat balance:
Alternative Method
See Fig. 5.8.
Consider any radius r in the cylindrical shell as shown.

176 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Since the inside surface is insulated, heat generated within the volume between r = ri and r = r, must travel
only outward; and, this heat must be equal to the heat conducted away from the surface at radius r.
Writing this heat balance,
dT
qg × p ×(r2 – ri2 )×L = – k×2×p×r× L×
dr
where, dT/dr is the temperature gradient at radius r.
qg × ri2 dr qg
i.e. dT = × – r × dr
2×k r 2× k

qg × ri2 qg × r 2
Integrating T(r) = × ln(r) – +C ...(b)
2× k 4×k
Eq. b is the general solution for temperature distribution.
The integration constant C is obtained by the B.C.:
At r = ro , T = To
Applying this B.C. to Eq. b:
qg × ro2 qg × ri2
C = To + - ×ln (ro)
4×k 2× k
Substituting value of C back in Eq. b we get,
qg × ri2 qg × r 2 qg × ro2 q g × ri2
T(r) = × ln( r ) - + To + - × ln(ro )
2×k 4×k 4×k 2×k

qg × ri2 LMF r I 2
FG r IJ - FG r IJ 2
OP
MNGH r JK H rK HrK
o o
i.e. T(r) = To +
4× k
×
i
- 2 × ln
i PQ ...(5.27)

Eq. 5.27 gives the temperature distribution in a hollow cylinder with heat generation, insulated on the inside
surface, in terms of the outer wall temperature, To.
Putting r = ri and T = Ti in Eq. 5.27, we get,
LMF r I Fr I O
2

- 2 × ln G J - 1P
qg × ri2
Ti = To + G
4 × k MH r K
× J o
H r K PQ
o

N i i

LF r I 2
Fr I O
× MG J - 2 × ln G J - 1P
qg × ri2 o o
i.e. Ti – To =
4× k MNH r Ki H r K PQ i
...(5.28)

Eq. 5.28 is important, since it gives the maximum temperature drop in the cylindrical shell, when there is
internal heat generation and the inside surface is insulated.
If either of To or Ti is given in a problem, then the other temperature can be calculated using Eq. 5.28.
Convection boundary condition:
If heat is carried away at the outer surface by a fluid at a temperature Ta flowing on the surface with a convective
heat transfer coefficient, ha , then, it is the fluid temperature that is known and not the surface temperature, To . In
such cases, we relate the surface temperature and fluid temperature by an energy balance at the surface, i.e. heat
generated within the body and conducted to the outer surface is equal to the heat convected away by the fluid at
the surface.
i.e. e j
qg × p × ro2 - ri2 L = ha ×2×p×ro × L × (To – Ta )

i.e. To = Ta +
e
qg × ro2 - ri2 j ...(c)
2 × ha × ro

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 177


Substituting the value of To from Eq. c in Eq. 5.27, we get:

e
qg × ro2 - ri2 j + q × r × LMF r I 2 2
FG r IJ - FG r IJ OP
2

4 × k MGH r JK
g i
H r K H r K PQ
o o
T(r) = Ta + - 2 × ln ...(5.29)
2 × ha × ro
N i i

Eq. 5.29 gives the temperature distribution in the cylindrical shell with heat generation, inside surface
insulated, when the heat generated is carried away by a fluid flowing on the outer surface.
5.3.2.2 Analysis with variable thermal conductivity. In the above analysis, thermal conductivity of the material was
assumed to be constant. Now, let us make an analysis when the thermal conductivity varies linearly with
temperature as:
k(T) = ko (1 + b T),
where, ko and b are constants.
Again, considering Fig. 5.8, we have, from heat balance:

d i
qg ×p × r 2 - ri2 × L = – k(T)×2 ×p×r×L ×
dT
dr
where, dT/dr is the temperature gradient at radius r.

i.e. k(T)×dT =
e
qg × ri2 - r 2 j × dr
2×r
Substituting for k(T),
qg × ri2 dr qg
ko×(1 + b× T)×dT = × - r × dr
2 r 2

b ×T 2 qg × ri2 qg ×r 2
Integrating, T+ = × ln(r) - +C ...(d)
2 2 × ko 4 × ko
In Eq. d, C is the integration constant. It is obtained by applying the B.C.,
At r = ri, T = Ti
Applying this B.C. to Eq. d:

b ×Ti2 qg × ri2 qg × ri2


C = Ti + – × ln (ri) +
2 2 × ko 4 × ko
Substituting value of C back in Eq. d:
F q ×r 2 2 2 2 I
b ×T 2
GG 2 × k × ln(r) – q4 ××kr + T + b ×2T - q2 ××kr × ln(r ) + q4 ××kr
2
+T–
g i g i g i g i
JJ = 0
2 H o o
i
o
i
o K
L - q × r LF r I F rI O b × T OP
+T– M × MG J - 2 × ln G J - 1P + T +
2 2
b ×T 2 g i
2

MN 4 × k MNH r K 2 P
i
i.e.
2 o H r K PQ
i
Q
=0
i
i ...(e)

Eq. e is a quadratic in T. Its positive root is given by:

b - q g × ri
LM
2 LMF r I 2
F r I - 1OP + T + b × T 2 OP
MN MNGH r JK GH r JK P PQ
- 1+ 1 + 4× × × - 2 × ln i
i

T(r) =
2 4 × ko i i
Q 2
b

2

178 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F1 2 × TiI q g × ri2 LMF r I 2
F r I - 1OP
GH b JK MNGH r JK GH r JK P
-1
i.e. T(r) = + + Ti2 + - × - 2 × ln
Q
2 b 2 × b × ko
b i i

FG 1 + T IJ 2
qg × ri2 LMF r I 2
F r I - 1OP
MNGH r JK GH r JK P
-1
i.e. T(r) =
b
+
Hb K i -
2 × b × ko
×
i
- 2 × ln
i
Q
...(5.30)

Eq. 5.30 gives the temperature distribution in a hollow cylinder with internal heat generation when the
inside surface is insulated and the thermal conductivity varies linearly with temperature.
5.3.2.3 Hollow cylinder with the outside surface insulated. Consider steady state, one-dimensional heat transfer in a
hollow cylinder of length L, inside radius ri and outside radius ro, with a uniform internal heat generation rate of
qg (W/m3). Thermal conductivity, k is constant. Let the outside surface be perfectly insulated; that means, all the
heat generated in the cylindrical shell has to move only inwards, in the negative r-direction. Let the temperatures
on the inside and outside surfaces be Ti and To, respectively. See Fig. 5.9.
Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k Ti Insulated
(iv) Uniform internal heat generation rate, qg (W/m3). k, qg
Q
With the above stipulations, the general differential equation in
cylindrical coordinates (see Eq. 3.17) reduces to:
To
d 2T 1 dT qg
2
+ × + =0 ...(a)
dr r dr k
Integrating Eq. a twice, we get the general solution for tem-
perature distribution, i.e. Eq. 5.18, as done in section 5.3.1:
- qg × r 2
T(r) = + C1 ×ln(r) + C2 ...(5.18)
4×k ri
Eq. 5.18 is the general relation for temperature distribution
ro
along the radius, for a cylindrical system, with uniform heat
generation. FIGURE 5.9 Hollow cylinder with heat
C1 and C 2, the constants of integration are obtained by apply- generation, outside surface insulated
ing the boundary conditions.
In the present case, the B.C.’s are: Insulated
Ti
B.C.(i): at r = ri T = Ti , and dr
B.C.(ii): at r = ro T= To , and dT/dx = 0 (since outer surface is Q
k, qg
insulated).
Get C1 and C2 from these B.C.’s and substitute back in Eq. 5.18
to get the temperature distribution. This is left as an exercise for the r To
student (see Example 5.12. for procedure of working out a
numerical problem).
We shall, however, derive the expression for temperature
distribution by a simpler method of physical consideration and heat
balance:
Alternative Method:
See Fig. 5.10. ri
Consider any radius r in the cylindrical shell as shown. ro
Since the outside surface is insulated, heat generated within
the volume between r = ro and r = r, must travel only inward; and, FIGURE 5.10 Hollow cylinder with heat
this heat must be equal to the heat conducted from the surface at generation, outside the surface insulated
radius r.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 179


Writing this heat balance,

e j
qg × p × ro2 - r 2 × L = k×2×p×r× L ×
dT
dr
where, dT/dr is the temperature gradient at radius r.
Note that the term on the RHS has positive sign, since, now, the heat transfer is from outside to inside, i.e. in
the negative r-direction (because the outside surface is insulated).
qg
i.e. dT =
2× k × r
er2
o j
- r 2 × dr ...(a)

Integrating:
qg × ro2 qg × r 2
T(r) = × ln(r) – +C ...(b)
2×k 4×k
Eq. b is the general solution for temperature distribution. The integration constant C is obtained by the B.C.:
At r = ri , T = Ti
Applying this B.C. to Eq. b:
qg × ro2 qg × ri2
C = Ti – × ln(ri) +
2×k 4×k
Substituting value of C back in Eq. b:
qg × ro2 qg × r 2 qg × ro2 qg × ri2
T(r) = × ln(r) – + Ti – × ln(ri) +
2×k 4×k 2×k 4×k

L F rI F r I F rI O 2 2
× M2 × ln G J + G J - G J P
qg × ro2 i
i.e. T(r) = Ti +
4× k MN H r K H r K H r K PQ
i o o
...(5.31)

Eq. 5.31 gives the temperature distribution in a hollow cylinder with heat generation, insulated on the
outside surface, in terms of the inner wall temperature, Ti .
Putting r = ro and T = To in Eq. 5.31, we get,

qg × ro2 LM F r I F r I 2 OP
MN GH r JK + GH r JK
o i
To – Ti =
4× k
× 2 × ln
i o
-1
PQ ...(5.32)

Eq. 5.32 is important, since it gives the maximum temperature drop in the cylindrical shell, when there is
internal heat generation and the outside surface is insulated.
If either of To or Ti is given in a problem, then the other temperature can be calculated using Eq. 5.32.
Convection boundary condition:
If heat is carried away at the inner surface by a fluid at a temperature Ta flowing on the surface with a convective
heat transfer coefficient, ha , then, it is the fluid temperature that is known and not the surface temperature, Ti . In
such cases, we relate the surface temperature and fluid temperature by an energy balance at the surface, i.e. heat
generated within the body and conducted to the inner surface is equal to the heat convected away by the fluid at
the surface.
i.e. qg × p × (r o2 – r i2)× L = ha× 2× p× ri× L× (Ti – Ta)

i.e. Ti = Ta +
e
qg × ro2 - ri2 j ...(c)
2 × ha × ri
Using Eq. c in Eq. 5.31, we get:

e
qg × ro2 - ri2 j + q × r × LM2× ln F r I + F r I - F r I OP
2 2 2

4×k M
g o
G J G J G J i

N H r K H r K H r K PQ
T(r) = Ta + ...(5.33)
2 × ha × ri i o o

180 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eq. 5.33 gives the temperature distribution in a hollow cylinder with heat generation, insulated on the out-
side surface, cooled by a fluid on the inside, in terms of the fluid temperature, Ta.
5.3.2.4 Analysis with variable thermal conductivity. In the above analysis, thermal conductivity of the material was
assumed to be constant. Now, let us make an analysis when the thermal conductivity varies linearly with tem-
perature as:
k(T) = ko (1 + b T),
where, ko and b are constants.
Again, considering Fig. 5.10, we have, from heat balance:

qg ×p× ro - re 2 2
j ×L = k(T)×2 × p × r × L × dTdr
where, dT/dr is the temperature gradient at radius r.
Separating the variables and substituting for k(T),
qg qg × ro2 dr qg
ko×(1 + b ×T) × dT =
2× r
er 2
o j
- r 2 × dr =
2
×
r

2
× r × dr

b ×T 2 qg × ro2 qg × r 2
Integrating: T+ = × ln(r) – +C ...(d)
2 2 × ko 4 × ko
In Eq. d get the integration constant, C from the B.C.: at r = ro , T = To
2
b × To2 qg × ro qg × ro2
i.e. C = To + - × ln(ro ) +
2 2 × ko 4 × ko
Substitute value of C back in Eq. d:

b ×T 2 F q ×r 2
qg × r 2 b × To2 qg × ro
2
qg × ro2 I =0
2
+T– GH 2× k
g o

o
× ln(r ) -
4 × ko
+ To +
2
-
2 × ko
× ln ( ro ) +
4 × ko
JK ...(e)

Eq. e is a quadratic in T. Its positive root is given by:

b qg × ro
2
F qg ×r 2 b × To2 qg × ro
2
qg × ro2 I
- 1+ 1 + 4× ×
2 2 × ko
GH
× ln (r ) -
4 × ko
+ To +
2
-
2 × ko
× ln(ro ) +
4 × ko
JK
T(r) =
b

2
After some manipulation, we get:

FG 1 + T IJ 2
q g × ro2 LM F r I - F r I 2 OP k, qg

MN GH r JK GH r JK
-1 o o
T(r) =
b
+
Hb K o -
2 × b × ko
× 2 × ln -1
PQ ...(5.34)
Tm

Eq. 5.34 gives the temperature distribution in a hollow cylinder


rm To
with internal heat generation when the outside surface is insulated Ti
and the thermal conductivity varies linearly with temperature.
5.3.2.5 Hollow cylinder with both the surfaces maintained at constant
temperatures. Consider steady state, one-dimensional heat transfer
in a hollow cylinder of length L, inside radius ri and outside radius
ro , with a uniform internal heat generation rate of qg (W/m3).
Thermal conductivity, k is constant. Let the temperatures on the in- ri
side and outside surfaces be Ti and To , respectively. The cylinder is
losing heat from both the surfaces. ro
See Fig. 5.11.
FIGURE 5.11 Hollow cylinder with heat
generation, losing heat from both surfaces

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 181


Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Uniform internal heat generation rate,
qg (W/m3).
With the above stipulations, the general differential equation in cylindrical coordinates (see Eq. 3.17 reduces
to:
d 2T1 dT qg
+ ×
2
+ =0 ...(a)
dr r dr k
Integrating Eq. a twice, we get the general solution for temperature distribution, i.e. Eq. 5.18, as done in
section 5.3.1:
- qg × r 2
T(r) = + C1 × ln(r) + C2 ...(5.18)
4×k
Eq. 5.18 is the general relation for temperature distribution along the radius, for a cylindrical system, with
uniform heat generation. C1 and C2, the constants of integration are obtained by applying the boundary
conditions.
In the present case, the B.C.’s are:
B.C.(i): at r = ri T = Ti, and
B.C.(ii): at r = ro T= To
Get C1 and C2 from these B.C.’s and substitute back in Eq. 5.18 to get the temperature distribution.
From B.C.(i) and Eq. 5.18:
- qg × ri2
Ti = + C1 × ln(ri) + C2 ...(a)
4×k
From B.C.(ii) and Eq. 5.18:
- qg × ro2
To = + C1 ×ln(ro) + C2 ...(b)
4×k
Subtracting Eq. a from Eq. b:
qg
e j Fr I
To – T i =
4×k
× ri2 - ro2 + C1 × ln GH r JK o
i

qg
(To - Ti ) + e
4×k
× ro2 - ri2 j
i.e. C1 =
Fr I
ln G J o
HrK i

And, from Eq. a:


qg
qg × ri2
(To - Ti ) + e 4×k
× ro2 - ri2 j
C2 = Ti +
4×k

Fr I
ln G J o
× ln (ri )

HrK i

Temperature distribution in the cylindrical shell is obtained by substituting C1 and C2 in Eq. 5.18.
After lengthy algebraic manipulations, we get,

182 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG r IJ LM F r I F r I 2 OP
T (r ) - Ti
ln
H r K + q × er - r j × MM ln GH r JK - GH r JK
i g
2
o i
2
i i
-1
PP
To - Ti
=
F r I 4× k (T - T ) M ln F r I F r I
ln G J
2
P
....(5.35)

MN GH r JK GH r JK
o i
- 1P
o o
HrK i i
o
i Q
Position and value of maximum temperature:
Position of maximum temperature must lie somewhere between ri and ro , since heat is flowing to both inside and
outside surfaces. Let the position be at a radius of rm . Then, rm is found out by differentiating the expression for
T(r) given by Eq. 5.35 w.r.t. r and equating to zero. Then, this value of rm is substituted back in Eq. 5.35 to obtain
Tmax. The procedure will be illustrated in an example, later.
Heat transfer to both surfaces:
Knowing the temperature distribution, heat transfer rate is easily determined by applying the Fourier’s law.
Heat transfer rate at the inner surface, Q|r = ri = – k(2p ri L) (dT/dr)|r = ri

Heat transfer rate at the outer surface, Q|r = ro = – k(2pro L) (dT/dr)|r = ro


Note that heat transfer to inner surface will be negative since the heat flow is from outside to inside, i.e. in
the negative r-direction.
Check: Sum of the amount of heats flowing to the inner and outer surfaces must be equal to the total amount of
heat generated in the cylindrical shell.
Convective boundary conditions:
If heat is carried away at the inner surface by a fluid at a temperature Ta flowing on the surface with a convective
heat transfer coefficient, ha , and on the outer surface, by a fluid at a temperature Tb flowing on the surface with a
convective heat transfer coefficient, hb then, the surface temperatures can be related to the fluid temperatures by
making an energy balance at the surfaces. i.e. heat generated within the body and conducted to the inner and
outer surfaces must be equal to the heat convected away by the fluid at the respective surfaces.
See Example 5.10 for procedure of working out a numerical problem.
Alternative Method:
See Fig. 5.11.
Since heat is transferred from both the inside and outside surfaces, maximum temperature, Tm must occur
somewhere in the shell. Let it occur at a radius rm. Obviously, rm lies in between ri and ro . Now, note that surface
at rm is an isothermal surface; also, since maximum temperature occurs at rm , no heat will cross the surface at rm
i.e. dT/dr at r = rm will be zero. This also means that surface at rm may be considered as representing an insulated
boundary condition.
So, the cylindrical shell may be thought of as being made up of two shells; the inner shell, between r = ri and
r = rm , insulated on its ‘outer periphery’ and, an outer shell, between r = rm and r = ro , insulated at its ‘inner
periphery’.
Then, maximum temperature difference for the inner shell and outer shell can be written from Eq. 5.32 and
5.28, respectively. So, we write:
For the ‘inner shell’ (insulated on the ‘outer’’ surface):

qg × rm2 LM F r I F r I 2 OP
MN GH r JK + GH r JK
m i
Tm – T i =
4× k
× 2 × ln
i m
-1
PQ ...(a)

Eq. a is obtained by replacing ro by rm and To by Tm in Eq. 5.32.


For the ‘outer shell’ (insulated on the ‘inner’ surface):

qg × rm2 LMF r I 2
F r I - 1OP
MNGH r JK GH r JK P
o o
Tm – To = × - 2 × ln ...(b)
4 ×k m m
Q
Eq. b is obtained by replacing ri by rm and Ti by Tm in Eq. 5.28.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 183


Subtracting Eq. a from b:

qg × rm2 LMF r I F I
2
F I F I 2 OP
Ti – To = × GMH r JK - 2× ln GH rr JK - 1 - 2× ln GH rr JK - GH rr JK
o o m i
+1
PQ
4×k
N m m i m

LF r I F r I 2
Fr I 2
F r IO
× MG J - G J + 2 × ln G J - 2 × ln G J P
qg × rm2 o i m m
i.e. Ti – To =
4× k MNH r K H r K
m Hr K
m H r K PQ o i
...(c)

Eq. c must be solved for rm . We get:


qg
e j Fr × r I
qg × rm2
Ti – To =
4×k
× ro2 - ri2 +
4×k GH r r JK 2 × ln m
o
i
m

qg
× dr - r i +
2
q ×r 2 Fr I
× 2 × ln G J
g
2
m i
i.e. Ti – To =
4× k
o
4×k
i
Hr K o

q L F r IO
× Mer - r j + 2r × ln G J P
g 2 2 2 i
i.e. Ti – To =
4 × k MN
o i
H r K PQ m
o

(Ti - To ) × 4 × k L 2 2 F r IO
= Mer - r j + 2 × r × ln G J P 2 i
i.e.
qg MN o i
H r K PQ m
o

(T - T ) × 4 × k
i o
er - r j o
2
i
2
i.e. r m2 =
Fr I -
q × 2 × ln G J 2 × ln G J
iFr I i
g
Hr K oHr K o

e j
qg × ro2 - ri2 - 4 × k ×(Ti - To )
i.e. r m2 =
FG r IJ
o
q g × 2 × ln
Hr K
i

e j
qg × ro2 - ri2 - 4 × k × (Ti - To )
i.e. rm =
Fr I ...(5.36)
qg × 2 × ln GH r JK o
i

Substituting the value of rm from Eq. 5.36 in either of Eqs. a or b, we get the maximum temperature in the
shell.
Then, temperature distribution in the inner shell is determined from Eq. 5.32 and that in the outer shell is
determined from Eq. 5.28.
When Ti and To are equal:
When the cooling on the surfaces is such that both Ti and To are the same, an interesting situation develops: then,
it is seen from Eq. 5.36 that, position of maximum temperature in the shell is given by:

ro2 - ri2
rm =
FG r IJ
o
2 × ln
HrK i

i.e. rm depends only on the physical dimensions of the cylindrical shell and not on the thermal conditions.

184 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For example, for a hollow cylinder with ri = 5 cm and ro = 10 cm, when the inside and outside surfaces are
maintained at the same temperature (Ti = To ), the maximum temperature in the shell occurs at a radius of:

102 - 52
rm =
10 FG IJ
2 × ln
5 H K
i.e. rm = 7.355 cm.
This result is valid, whatever may be the value of uniform heat generation.
Example 5.10. A hollow cylinder 6 cm ID, 9 cm OD, has a heat generation rate of 5 ´ 106 W/m3. Inner surface is
maintained at 450°C and outer surface at 350°C. k of the material is 3 W/(mK).
(i) Determine the location and value of maximum tem-
perature. k = 3 W/(mK)
3
(ii) What is the temperature at mid-thickness of the qg = 5 MW/m
shell? Tm
(iii) Determine the fraction of heat generated going to the
To = 350°C
inner surface, and
(iv) Sketch the temperature profile. Ti = 450°C
Solution. See Figure Example 5.10.
Data:
ri := 0.03 m ro := 0.045 m L := 1 m
Ti := 450°C To := 350°C k := 3 W/(mK) ri = 0.03 m
qg := 5 ´ 10 6 W/m 3
Position of maximum temperature can be immediately rm
determined from Eq. 5.36 and then, the value of maximum
ro = 0.045 m
temperature may be determined from Eq. 5.28 or 5.32.
However, let us work out this problem from first princi- FIGURE Example 5.10 Hollow cylinder with heat
ples and then, verify the results from the formulas already generation, losing heat from both surfaces
derived.
Temperature distribution:
For the assumption of one-dimensional, steady state conduction with heat generation in a cylindrical geometry, we have
the governing differential equation:
d 2T 1 dT q g
+ × + = 0 ...(a)
dr 2 r dr k

d 2T dT qg × r
Multiplying by r : r× + + = 0
dr 2 dr k

d F dT I - qg × r
i.e.
drGH

dr JK =
k

dT - qg × r 2
Integrating: r× = + C1
dr 2 ×k

dT - qg × r C
i.e. = + 1 ...(b)
dr 2× k r

- qg × r 2
Integrating again: T(r) = + C1 × ln(r) + C2 ...(5.18)
4×k
Eq. 5.18 gives the temperature distribution. C1 and C2 are determined from the B.C.’s:
B.C.(i): at r = ri , we have T = Ti
B.C.(ii): r = ro , we have T = To
From B.C.(i) and Eq. 5.18:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 185


- qg × ri2
Ti = + C1 × ln (r i) + C2 ...(c)
4×k
From B.C.(ii) and Eq. 5.18:
- qg × ro2
To = + C1 × ln (ro) + C2 ...(d)
4×k
Subtracting Eq. c from Eq. d:
qg
d i Fr I
To – Ti =
4×k
× ri2 - ro2 + C1 × ln GH r JK
o

qg
(To - Ti ) +
4× k
d
× ro2 - ri2 i
i.e. C1 :=
Fr I (define integration constant C1)
ln GH r JK
o

i.e. C1 = 909.449 (value of C1, after substituting numerical values from data)
and, from Eq. c:
qg
qg × ri2
(To - Ti ) +
4× k
d
× ro2 - ri2 i
C2 := Ti +
4× k
-
Fr I
ln G J o
× ln(ri) (define integration constant C2)

HrK i

i.e. C2 = 4.01404 ´ 103 (value of C2, after substituting numerical vaues from data)
Substituting C1 and C2 in Eq. 5.18, we get the temperature distribution as:
- qg × r 2
T(r) = + C1 × ln(r) + C2
4×k
i.e. T(r) := – 4.16667 ´ 105× r 2 + 909.449 ln (r) + 4.01404 ´ 10 3 ...(e)
Eq. e is the desired temperature distribution in the shell as a function of radius, r.
Position and value of maximum temperature:
To get the position where maximum temperature occurs, differentiate Eq. e w.r.t. r and equate to zero. Let the location
be at a radius rm . Then, substitute rm back in Eq. e to get value of maximum temperature, Tm .
Differentiating Eq. e w.r.t. r and equating to zero:
d 909. 449
T(r) = – 4.16667 ´ 105 × 2 × r + = 0
dr r

909. 449
i.e. rm =
2 × 4 ×16667 ´ 105
i.e. rm = 0.033 m (position of maximum temperature)
And, substituting rm in Eq. e, we get Tmax
T(rm) = 457.935°C (value of maximum temperature)
Note: In Mathcad, there is no need to actually differentiate Eq. e and equate to zero, then solve etc. First, define T’(r) as
the first derivative of T(r) = dT(r)/dr and then use the solve block to get the root of T’(r) = 0. For doing this, assume a trial
value of r to start with. Procedure is shown below:
d
T¢(r) := T(r) (define first derivative of T(r))
dr
r := 0.03 m (trial value of r)
Given
T¢(r) = 0
rmax := Find (r) (define r max)
i.e. rmax = 0.033 m (position of maximum temperature...verified.)

186 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Check: check also from the direct formula Eq. (5.36) for rm :

d i
qg × ro2 - ri2 - 4 × k (Ti - To )
rm :=
Fr I ...(5.36)
q g × 2 × ln GH r JK
o

i.e. rm = 0.033 m (position of maximum temperature...checks.)


Check the maximum temperature from Eq. 5.28:
We have, for a shell insulated on the inner surface:

qg × ri2 LMF r I 2
F r I - 1OP
MNGH r JK GH r JK P
o o
Ti – To = × - 2 × ln ...(5.28)
4× k i i
Q
Apply this formula for the ‘outer shell’, i.e. between r = rm and r = ro ,: Now, replacing Ti by the maximum tempera-
ture Tm and ri by rm , we get:
LMF r I Fr I O
- 2 × ln G J - 1P
2
qg × rm2
Tm – To = G
4 × k MH r K
× J o

H r K PQ
o

N m m

q × r LF r I Fr I O
× MG J - 2 × ln G J - 1P
2 2
g m o o
4 × k MH r K H r K PQ
i.e. Tm := T + (define Tm , the maximum temperature)
N
o
m m

i.e. Tm = 457.931°C (value of maximum temperature...checks.)


Temperature at mid-thickness i.e. at r = 0.0375 m:
Substitute r = 0.0375 in Eq. e for T (r):
i.e. T(0.0375) = 442.004°C (temperature at mid-thickness of shell)
Fraction of heat generated going to inner surface:
First, find the total heat generated in the shell:

d i
Qtot := p × ro2 - ri2 × L× qg, W ...define the total heat generation in the shell = (Volume ´ qg)
i.e. Qtot = 1.767 ´ 104 W (total heat generated in the shell)
Heat going to inner surface is equal to the amount of heat generated between r = r i and r = rm , since no heat crosses
the isothermal surface at rm .

d i
Qinner := p × rm2 - ri2 × L ×qg W (define the heat going to inner surface of the shell)
3
i.e. Qinner = 3.006 ´ 10 W (heat going to inner surface of the shell)
Therefore, fraction of heat going to inner surface:
Qinner
Fraction := ...define Fraction
Qtot
i.e. Fraction = 0.17 (i.e. 17% of the total heat generated goes to the inner surface.)
Note: Heat removed at the inner surface can also be found out by applying the Fourier’s law at r = ri: Remember,
temperature gradient is given by T’(r).
Qinner := – k × (2 × p × ri × L) × T¢(ri) W (define heat flow at inner surface...Fourier’s law)
i.e. Qinner = – 3.006 ´ 103 W (negative sign indicates that heat flow is radially inwards...verified.)
To sketch the temperature distribution:
To sketch the temperature profile in the shell, define a range variable r, varying from 0.03 to 0.045 m, with an increment
of 0.001 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with r
and T(r), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig. Ex. 5.10(b).
r := 0.03, 0.031, ... , 0.045 (define a range variable r..starting
value = 0.03, next value = 0.031 m and
last value = 0.045 m)
Note from the graph that maximum temperature occurs at r = 0.033 m.
Example 5.11. A high temperature, gas cooled nuclear reactor consists of a composite cylindrical wall for which a
thorium fuel element (k = 57 W/(mK)) is encased in graphite (k = 3 W/(mK)) and gaseous helium flows through an

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 187


T(r) in a cylindrical shell with heat generation
500
r in metres and
485
T(r) in deg.C
470
455
440
T(r)
425
410
395
380
365
350
0.03 0.033 0.036 0.039 0.042 0.045
r

FIGURE Example 5.10(b)

Helium coolant channel annular coolant channel, as shown in Fig. 5.23. Consider conditions
Ta = 600 K, ha = 2000 W/(m2K) for which helium temp Ta = 600 K and convective coefficient ha at
the outer surface of graphite = 2000 W/(m2K). If r1 = 8 mm, r2 = 11
Graphite, k2 = 3 W/(mK)
mm, r3 = 14 mm and qg = 108 W/m3, find out temperatures T1 and
Thorium, k1 = 57 W/(mK) T2 , i.e. at inner and outer surfaces of the fuel element. Also, draw
T2
qg = 108 W/m3 the temperature profile in the fuel element and graphite.
T1 Solution. See Figure Example 5.11.
Data:
Insulated r1 := 0.008 m r2 := 0.011 m r3 := 0.014 m
qg := 10 8 W/m3 k1 := 57 W/(mK) k2 := 3 W/(mK)
Ta := 600 K ha := 2000 W/(m2K) L := 1 m
Find out T1 and T2.
r1 = 0.008 m
r2 = 0.011 m Note that heat generation is only in thorium. Inside surface of
thorium is insulated. So, in steady state, all the heat generated in
r3 = 0.014 m thorium flows out by conduction through graphite shell and then
by convection to helium gas.
FIGURE Example 5.11 Hollow cylindrical
Total heat generation rate, Q:
fuel element, encased in graphic and cooled
Q := qg × Volume
by helium gas externally
i.e. d i
Q = qg × p × r22 - r12 × L, W (define q, total heat generated)
4
i.e. Q = 1.79071 ´ 10 W (total heat generated in thorium fuel element)
Now, this Q is transferred to helium gas coolant. Thermal resistances involved are:
Rcyl = thermal resistance of the cylindrical graphite shell, and
Rconv = convective resistance between helium gas and outer surface of graphite shell.
These two resistances are in series. Total temperature potential, DT = (T2 – Ta )
Thermal resistances:
FG r IJ
Hr K
3
ln
2
Rcyl := C/W (define thermal resistance of graphite shell)
2 × p × k2 × L
i.e. Rcyl = 0.013 C/W (thermal resistance of graphite shell)
1
Rconv := C/W (define convective resistance on outside surface of graphite shell)
ha × 2 × p × r3 × L

188 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Rconv = 5.684 ´ 10 –3 C/W (convective resistance on outside surface of graphite shell)
T2 - Ta
Now, we have: Q=
Rcyl + Rconv
Therefore, T2 := Q× (Rcyl + Rconv) + Ta K (define T2)
i.e. T2 = 930.89 K (temperature at outer surface of fuel element.)
To get temperature profile in thorium and then, find T1:
T1 is the temperature at the inner surface of the fuel element. It is insulated on the inner surface. So, we can apply Eq.
5.27, derived earlier for a hollow cylinder with heat generation, with the inner surface insulated. However, first we will
solve it from first principles and then verify the results from Eq. 5.27:
For the assumption of one-dimensional, steady state conduction with heat generation in a cylindrical geometry, we
have the governing differential equation:
d 2T 1 dT q g
+ × + = 0 ...(a)
dr 2 r dr k

d 2T dT qg × r
Multiplying by r: r× + + = 0
dr 2 dr k

d FGdT IJ - qg × r
i.e.
dr

Hdr K =
k

dT - qg × r 2
Integrating: r× = + C1
dr 2 ×k

dT - qg × r C
i.e. = + 1 ...(b)
dr 2× k r

- qg × r 2
Integrating again: T(r) = + C1 × ln (r) + C2 ...(5.18)
4×k
Eq. 5.18 gives the temperature distribution. C1 and C2 are determined from the B.C.’s:
B.C.(i): at r = r1, we have dT/dr = 0, since inner surface is insulated.
B.C.(ii): at r = r2, we have T = T2 ,
From B.C.(i) and Eq. b
- qg × r1 C1
0= +
2 × k1 r1

+ qg × r12
i.e. C1 := (define C1)
2 × k1
i.e. C1 = 56.14035 (value of integration constant C1)
From B.C.(ii) and Eq. 5.18
- qg × r22
T2 = + C1 × ln(r2) + C2
4 × k1

- q g × r22
i.e. C2 := T2 + – C1 × ln(r2) (define C2)
4 × k1
i.e. C2 = 1.23714 ´ 10 3 (value of integration constant C2)
Substituting C1 and C2 in Eq. 5.18,
T(r) := – 4.38596 ´ 105 × r 2 + 56.14035 ln(r) + 1237.14 ...(c)...defines T(r)
Eq. c gives temperature profile in the thorium fuel element.
Temperature T1 on the inner surface of thorium:
Put r = 0.008 m in Eq. c
T(0.008) = 938.007
i.e. T1 = 938.007 K (temperature on the inner surface of thorium fuel element.)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 189


Verify from Eq. 5.27:
We have:

+ qg × ri2 LMF r I 2 OP
FG r IJ - FG r IJ 2

MNGH r JK H rK HrK
o o
T(r) = To +
4×k
×
i
- 2 × ln
PQ i
...(5.27)

LF r I Fr I F rI O
× MG J - 2 × ln G J - G J P
2 2
+ qg × r12
MNH r K H r K H r K PQ
2 2
i.e. T(r) := T2 + (Eq. 5.27, with notations of this problem)
4 × k1 1 1

i.e. T(0.008) = 938.012 K (temperature at the inner surface of the fuel element...verified.)
Check:
Heat flux at the interface must be the same for thorium as well as graphite:
dT (r ) dt (rs )
i.e. – k× = – k×
dr dr
at r = rs = 0.011 m, where T(r) is temperature profile for thorium and t(rs) is temperature profile for graphite.
Temperature profile for thorium is already obtained as:
T(r) = – 4.38596 ´ 10 5 × r 2 + 56.14035 ln (r) + 1237.14
Temperature profile for graphite is, from Eq. 4.34 for a cylindrical shell:
T3 - T2 Fr I
× ln G Js
t(rs) = T2 +
F r
ln G J
I Hr K
2
(Eq. 4.34, with notations of this problem)

H K
3
r2

where, rs = any radius within the graphite shell


Now, define their first derivatives w.r.t. radius:
d
T¢ (r) := T(r) and,
dr

d
t¢(rs) := t(r s )
dr
Therefore, heat flux in thorium at r = 0.011 m:
qthorium = – k1 × T¢(0.011) ...define heat flux, from Fourier’s law
i.e. qthorium = 2.591 ´ 10 5 W/m2 ...heat flux in thorium at the interface
Heat flux in graphite at r = 0.011 m:
qgraphite = – k2 × t¢(0.011) ...define heat flux, from Fourier‘s law
i.e. qgraphite = 2.591 ´ 10 5 W/m 2 ...heat flux in graphite at the interface
Therefore, we observe that heat fluxes are same for thorium and graphite at the interface
(verified.)
To sketch the temperature profiles:
For temperature profile in graphite, we use Eq. 4.34, which was derived for a cylindrical shell with no heat generation.

T3 - T2 Fr I
× ln G Js
t(rs) = T2 +
F r
ln G J
I Hr K
2
...4.34, with notations of this problem

Hr K
3

where, rs = any radius within the graphite shell.


To sketch the temperature profile in the thorium fuel element, define a range variable r, varying from 0.008 to 0.011 m,
with an increment of 0.0001 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis
and y-axis with r and T(r), respectively. Click anywhere outside the graph region, and immediately the graph appears.
r := 0.008, 0.0081, ... , 0.011 (define a range variable r..starting
value = 0.008, next value = 0.0081 m and
last value = 0.011 m)

190 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T(r) for thorium, insulated on inside
940
939
r in metres and
938 T(r) in K
937
936
T(r)
935
934
933
932
931
930
0.008 0.0085 0.009 0.0095 0.01 0.0105 0.011
r

FIGURE Example 5.11(b)

It may be seen from the graph that at r = 0.008 m, temperature T1 is 938.01 K and, at r = 0.011 m temperature T2 is
930.9 K.
Similarly, to sketch the temperature profile in the graphite shell, define a range variable rs, varying from 0.011 to
0.014 m, with an increment of 0.00001 m. Then, choose x–y graph from the graph palette, and fill up the place holders on
the x-axis and y-axis with rs and t(rs), respectively. Click anywhere outside the graph region, and immediately the graph
appears. Fig. Ex. 5.11(c)
rs := 0.011, 0.01101, ..., 0.014 (define a range variable r..starting
value = 0.011, next value = 0.01101 m
and last value = 0.014 m)

T(r) for graphite shell (no heat generation)


1000
975
950
925 rs in metres and
900 t(rs) in K
875
t(rs)
850
825
800
775
750
725
700
0.011 0.0115 0.012 0.0125 0.013 0.0135 0.014
rs

FIGURE Example 5.11(c)

It may be seen from the graph that at at r = 0.011 m, temperature T2 is 930.9 K, and at r = 0.014 m, temperature T3
is 701.8 K.
Example 5.12. A hollow conductor with ri = 0.6 cm, ro = 0.8 cm is made up of metal of k = 20 W/(mK) and electrical
resistance per metre of 0.03 ohms. Find the maximum allowable current if the temperature is not to exceed 50°C
anywhere in the conductor. The cooling fluid inside is at 38°C. (Conductor is insulated on the outside).
Solution. See Figure Example 5.12.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 191


Insulated k = 20 W/(mK) Data:
R = 0.03 W/m ri := 0.006 m ro := 0.008 m k := 20 W/mK
R := 0.03 Ohms/m length Ti := 38°C L := 1 m
Note: temperature on inside surface is assumed as that of
the fluid flowing, since heat transfer coefficient between the
To = Tmax = 50°C surface and the fluid is not given.
To := 50°C (temperature of outer surface ...this
Ti = 38°C
is also maximum temperature in
conductor, since conductor is insulated
on outside)
This is the case of a hollow cylinder with heat genera-
ri = 0.006 m tion, cooled from inside and insulated on the outside
surface. Both the inside and outside temperatures are
known. So, one can use Eq. 5.32 directly to get qg . Once qg is
r = 0.008 m known, the current, I can easily be calculated.
We shall, however, solve the problem from first princi-
FIGURE Example 5.12 Hollow conductor with heat ples and then check the result with Eq. 5.32:
generation, insulated on outside surface, cooled on As shown in the earlier two examples, the general
inside equation for temperature distribution in a hollow cylinder
with heat generation is given by Eq. 5.18, i.e.
- qg × r 2
T(r) = + C1 × ln (r) + C2 ...(5.18)
4×k
Eq. 5.18 gives the temperature distribution. C1 and C2 are determined from the B.C.’s:
B.C.(i) at r = ri , we have T = Ti, known temperature
B.C.(ii) at r = ro , we have: dT/dr = 0, since insulated.
From 5.18, we have:
dT - qg × r C1
= + ...(a)
dr 2× k r
From B.C.(ii) and Eq. a:
- qg × ro C1
0= +
2 ×k ro

qg × ro2
i.e. C1 = (integration constant C1)
2× k
From B.C.(i) and Eq. 5.18:
- qg × ri2 qg × ro2
Ti = + × ln(ri) + C 2
4 ×k 2 ×k

qg × ri2 LM F r I 2 OP
MN GH r JK
o
or, C2 = Ti –
4× k
× 2×
i
× ln (ri ) - 1
PQ
Substituting C1 and C2 in Eq. 5.18:

- qg × r 2 qg × ro2 q g × ri2 LM F r I 2
OP
MN GH r JK
o
T(r) =
4 ×k
+
2 ×k
× ln (r ) + Ti -
4 ×k
× 2×
i
× ln (ri ) - 1
PQ
- qg × r 2
i.e. T(r) = + qg (1.6 ´ 10 –6) × ln (r) + 38 + 8.63559 ´ 10 –6 ×q g ...(b)
80
Eq. b is the desired expression for temperature distribution in a hollow cylinder with heat generation, when the
outer surface is insulated.
Now, by data: at r = 0.008 m (i.e. at r = ro), T = 50°C
Put this in Eq. b and solve to get qg:

192 FUNDAMENTALS OF HEAT AND MASS TRANSFER


- qg × (0. 008)2
i.e. 50 = + qg × (1.6 ´ 10 –6)× ln(0.008) + 38 + 8.63559 ´ 10–6 × qg
80
50 - 38
i.e. qg := W/m3 (define qg )
-6 -6 (0. 008)2
(1.6 ´ 10 ) × ln (0.008) + 8.63559 ´ 10 -
80
i.e. qg = 1.088 ´ 108 W/m3 ...heat generation rate
Verify: Verify this result using Eq. 5.32, already derived, for a hollow cylinder with heat generation when the outer
surface is insulated.

qg × ro2 LM Fr I +Fr I 2
OP
To – T1 =
4× k MN
× 2 × ln GH r JK GH r JK
o

i
i

o
-1
PQ ...(5.32)

4 × k × (To - Ti )
Therefore qg :=
L Fr I Fr I O W/m 3 (define qg)
r × M2 × ln G J + G J - 1P
2
o i

MN H r K H r K PQ
2
o
i o

8 3
i.e. qg = 1.088 ´ 10 W/m (heat generation rate...verified.)
Maximum allowable current in conductor:
Let the current be I (A). Then,
Qgen I2 ×R
qg = =
Volume p × ( ro2 - ri2 ) × L

qg × p × ( ro2 - ri2 ) × L
Therefore, I := A (define I)
R
i.e. I = 564.824 A (maximum allowable current in conductor)
For completeness, let us draw the temperature profile too:
To sketch the temperature profile in the shell, define a range variable r, varying from 0.006 to 0.008 m, with an
increment of 0.0001 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-
axis with r and T(r), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig.
Ex. 5.12(b).

T(r) :=
LM - q × r
g
2

+ qg × (1.6 ´ 10 - 6 ) × ln ( r ) + 38 + 8. 63559 ´ 10 - 6 × qg
OP ...define T(r)...(b)
MN 80 PQ
r := 0.006, 0.0061, ... , 0.008 (define a range variable r..starting value = 0.006,
next value = 0.0061 m and last value = 0.008 m)

T(r) for hollow cond ins. on outside


50

47
r in m and
44 T(r) in deg.C
T(r)

41

38

35
6 6.2 6.4 6.6 6.8 7 7.2 7.4 7.6 7.8 8
r ´ 103

FIGURE Example 5.12(b)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 193


It may be verified from the graph that on the inside and outside surfaces, the temperature are 38°C and 50°C,
respectively.
Note: In Examples 5.10, 5.11 and 5.12, we have solved from first principles, problems with all the three types of B.C.’s,
namely, cylindrical shell losing heat from both the surfaces, cylindrical shell insulated on the inside surface, and cylin-
drical shell insulated on the outside surface. Temperature profiles are also drawn for all the three cases. Student is
advised to study the procedure followed carefully.
Example 5.13. A long, hollow cylinder has inner and outer radii as 5 cm and 15 cm, respectively. It generates heat at the
rate of 1.0 kW/m3. If the maximum temperature occurs at the radius 10 cm and the temperature of the outer surface is
50°C, find:
(i) temperature of the inner surface
To = 50°C (ii) maximum temperature in the cylinder.
Assume k = 0.5 W/(mC).
Tm = ? Soution. See Figure Example 5.13.
Data:
k = 0.5 W/(mK)
ri := 0.05 m ro := 0.15 m rm := 0.1 m
qg = 1 kW/m3
L := 1 m To := 50°C
k := 0.5 W/(mK) qg := 1000 W/m3
Ti , temperature of inner surface is to be found out; also,
the value of maximum temperature, Tm .
Ti = ?
This is a problem of cylindrical shell with heat genera-
tion and losing heat from both surfaces; position of
ri = 0.05 m maximum temperature is at a radius of 10 cm and it is
equivalent to insulated surface since no heat crosses the
rm = 0.1 m
position of maximum temperature So, the whole shell be-
ro = 0.15 m tween the radii of 5 cm and 15 cm may be thought of as
being made of two shells: one inner shell, between radii of 5
FIGURE Example 5.13 Hollow conductor with heat cm and 10 cm, insulated on the outer surface, and the other,
generation, losing heat on both surfaces, location of an outer shell, between the radii of 10 cm and 15 cm,
maximum temperature given insulated on the inner surface.

To find heat generation rate, qg :


We have, for a shell insulated on the inner surface:

qg × ri2 LMF r I 2
F r I - 1OP
MNGH r JK GH r JK P
o o
Ti – To = × - 2 × ln ...(5.28)
4× k i i
Q
Apply this formula for the ‘outer shell’, i.e. between r = rm and r = ro ,: Now, replacing Ti by the maximum tempera-
ture Tm and ri by rm , we get:
LMF r I Fr I O
- 2 × ln G J - 1P
2
qg × rm2
4 × k MGH r JK
o o
Tm – T o =
N
×
m H r K PQ m

q × r LF r I Fr I O
×M - 2 × ln G J - 1P
2 2

4 × k MGH r JK
g m o o
i.e. Tm := T +
o
N m H r K PQ m
(define Tm , the maximum temperature)

i.e. Tm = 52.195°C (value of maximum temperature)


Temperature at the inner surface, Ti:
This is easily determined from Eq. 5.36, i.e.

q g × ( ro2 - ri2 ) - 4 × k × (Ti - To )


rm =
Fr I ...(5.36)
q g × 2 × ln GH r JK
o

194 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore,
Fr I
qg × ( ro2 - ri2 ) - rm2 × qg × 2 × ln GH r JK
o

i
Ti := + To (define Ti)
4× k
i.e. Ti = 49.014°C (temperature on the inner surface)
Alternatively:
Instead of applying direct formulas 5.28 and 5.36, which are rather complicated, it is, perhaps, easier to work from first
principles:
As explained earlier, the general Eq. for temperature distribution in a cylindrical shell with heat generation is given
by Eq. 5.18:
- qg × r 2
T(r) = + C1 ×ln(r) + C 2 ...(5.18)
4×k

dT (r) - qg × r C
and, = + 1 ...(a)
dr 2× k r
Eq. 5.18 gives the temperature distribution. C1 and C2 are determined from the B.C.’s:
Position of maximum temperature is given as at a radius of 10 cm, i.e. at the radius of 10 cm, we have an isothermal
surface and no heat crosses this surface i.e. it is equivalent to the B.C. dT/dr = 0 at r = 10 cm.
B.C.(i) at r = 0.015 m, we have: T = 50°C
B.C.(ii) at r = 0.01 m, we have dT/dr = 0
Applying these two B.C.’s, we get C 1 and C 2; then substitute C1 and C 2 back in Eq. 5.18 to get the temperature
profile. Then, maximum temperature is found out by simply putting r = 0.1 m in the equation for temperature profile.
From B.C.(ii) and Eq. a:
qg × rm2
C1 := (define C1)
2× k
i.e. C 1 = 10 (value of C 1, integration constant)
From B.C.(i) and Eq. 5.18:
- qg × ro2
50 = + C1 ×ln(ro) + C2
4× k

qg × ro2
i.e. C2 := 50 + – C1 ×ln(ro) (define C2)
4 ×k
i.e. C 2 = 80.2212 (C 2...integration constant)
Therefore, temperature distribution is given by:
- qg × r 2
T(r) = + C1 ×ln(r) + C 2
4×k
i.e. T(r) := – 500 ×r2 + 10×ln(r) + 80.2212 ((b) ...equation for temperature distribution.)
Maximum temperature
Put r = 0.1 m in Eq. b:
T(0.1) = 52.195 ...same as obtained earlier...verified.
Temperature at inner surface, Ti:
Put r = 0.05 m in Eq. b:
T(0.005) = 49.014 ...same as obtained earliler ...verified.
Example 5.14. A thin, hollow tube with 4 mm inner diameter and 6 mm outer diameter, carries a current of 1000
amperes. Water at 30°C is circulated inside the tube for cooling the tube. Taking heat transfer coefficient on the water
side as 35,000 W/(m2C) and k for the material as 18 W/(mC), estimate the surface temperature of the tube if its outer
surface is insulated. Electrical resistance of the tube is 0.0065 ohms per metre length.
Solution. See Figure Example 5.14.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 195


To Data:
Insulated ri := 0.002 m ro := 0.003 m k := 18 W/mK
I := 1000 A R := 0.0065 Ohms/m length
Ta := 30°C ha := 35000 W/(m 2K) L := 1 m
Problem is to find temperature on outer surface, To.
k = 18 W/(mK)
R = 0.0065 Ohms/m First, find qg, heat generation rate per unit volume:
Ti
Qg
qg =
Ta = 30°C Volume
where Qg is total heat generated in the conductor volume.
ha = 35 kW/(m2K) I 2 ×R
i.e. qg = W/m 3 (define qg)
p × ( ro2 - ri2 ) × L
ri = 0.002 m i.e. qg = 4.18 ´ 108 W/m 3 (heat generation rate per unit volume)
To find outer surface temperature To:
ro = 0.003 m
This is the case of a hollow cylinder with heat generation,
FIGURE Example 5.14 Hollow conductor with insulated on the outside. Therefore, Eq. 5.32 is applicable:
heat generation, losing heat on inside surface by qg × ro2 LM F r I F r I 2
OP
MN GH r JK + GH r JK
o i
convection, insulated on outside To – Ti =
4× k
× 2 × ln
i o
-1
PQ ...(5.32)

In the above equation RHS can be calculated, since all quantities on RHS are known.
However, Ti is not known yet; it is calculated by making an energy balance on the inner surface, remembering that
all the heat generated in the shell flows only to the inner surface, since the outer surface is insulated:
i.e. heat generated in the shell = heat transferred to water from inner surface by convection
i.e. I2 ×R = ha × (2×p×ri ×L)×(Ti – Ta)
I2 ×R
i.e. Ti := + Ta°C (define inner surface temperature Ti)
ha × ( 2 × p × r1 × L)
i.e. Ti = 44.779°C (temperature of inner surface)
Now, from Eq. 5.32:

qg × ro2 LM F r I F r I 2
OP
MN GH r JK + GH r JK
o i
To = Ti +
4× k
× 2 × ln
i o PQ
- 1 °C (define To , temperature of outer surface)

To = 57.988°C (temperature of outer surface.)


Exercise: Solve this problem, from fundamentals, i.e. starting from Eq. 5.18.

k = 50 W/(mK) Example 5.15. A nuclear fuel element is in the form of a hollow


qg = ? cylinder insulated at the inner surface. Its inner and outer radii
are 5 cm and 10 cm, respectively. The outer surface gives heat
ha = 100 W/(m2K) to a fluid at 50°C where the unit surface conductance is 100 W/
(m 2K). k of the material is 50 W/(mK). Find the rate of heat
generation so that maximum temperature in the system will not
Ta = 50°C exceed 200°C.
Solution. See Fig. Example 5.15.
Insulated Data:
ri := 0.05 m ro := 0.1 m k := 50 W/mK
Ta := 50°C ha := 100 W/(m2 K)
L := 1 m Ti := 200°C
ri = 0.05 m To find rate of heat generation, qg:
This is the case of a hollow cylinder with heat generation,
ro = 0.01 m insulated on the inside, losing heat on the outer surface to a
fluid flowing at temperature Ta, with heat transfer coefficient
FIGURE Example 5.15 Hollow cylinder with heat ha. Therefore, Eq. 5.29 is applicable:
generation, losing heat on outer surface by convec-
tion, inner surface insulated

196 FUNDAMENTALS OF HEAT AND MASS TRANSFER


e
qg × ro2 - ri2 j+q × ri2 LMF r I 2
FG r IJ - FG r IJ 2 OP
MNGH r JK
g
H rK Hr K
o o
T(r) = Ta +
2 × ha × ro 4×k
×
i
- 2 × ln
i PQ ...(5.29)

Here, Ti is known to be 200°C, since maximum temperature occurs on insulated inner surface, at r = 0.05 m. So, in
Eq. 5.29 replace r by ri and T(r) by Ti ; then, the only unknown, qg can be calculated:

d
qg × ro2 - ri2 i + q × r × LMF r I 2 2
F r I - F r I OP 2

4 × k MGH r JK GH r JK GH r JK P
g i o o i
i.e. Ti = Ta + - 2 × ln
2 × ha × ro
N i i i
Q
Ti - Ta
Therefore, qg :=
d i+ LMF r I Fr I O
(define qg)
- 2 × ln G J - 1P
2
ro2 - ri2 ri
2

MNGH r JK
o o

2 × ha × ro 4×k
×
i H r K PQ i

qg = 3.79582 ´ 10 W/m = 379.582 kW/m3


5 3
(heat generation rate/unit volume.)
Exercise: Solve this problem, from fundamentals, i.e. starting from Eq. 5.18.

5.4 Sphere with Uniform Internal Heat Generation


Spherical geometry is popular for many applications, such as reactors for chemical processes, storage of radio-
active wastes, experimental nuclear fuel elements, etc. We shall consider heat transfer in a solid sphere with
different types of boundary conditions.
5.4.1 Solid Sphere with Internal Heat Generation
Consider a solid sphere of radius, R. There is uniform heat generation within its volume at a rate of qg (W/m3).
Let the thermal conductivity, k be constant. See Fig. 5.12.
We would like to analyse this system for temperature distribution and maximum temperature attained.

k, qg Temperature profile,
To parabolic
Q

Tw
Solid sphere

R
R

FIGURE 5.12(a) Spherical system with heat FIGURE 5.12(b) Variation of temperature along
generation the radius

Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Uniform internal heat generation rate, qg (W/m3).
With the above stipulations, the general differential equation in spherical coordinates (see Eq. 3.21) reduces
to Eq. 3.24, i.e.
d 2T 2 dT qg
+ × + =0 ...(a)
dr 2 r dr k
We have to solve Eq. a to get the temperature profile; then, by applying Fourier’s law, we can get the heat
flux at any point.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 197


d 2T dT qg × r 2
Multiplying Eq. a by r 2: r 2× + 2× r × + =0
dr 2 dr k

F
d 2 dT I = - q ×r 2
i.e.
dr GH
r ×
dr JK k
g

dT - qg × r 3
Integrating: r2 = + C1
dr 3× k

dT - q g × r C1
i.e. = + 2 ...(b)
dr 3×k r

- qg ×r 2 C1
Integrating again: T(r) = - + C2 ...(5.37)
6× k r
Eq. 5.37 is the general relation for temperature distribution along the radius, for a spherical system, with
uniform heat generation. C1 and C2, the constants of integration are obtained by applying the boundary
conditions.
In the present case, B.C.’s are:
B.C. (i): at r = 0, dT/dr = 0, i.e. at the centre of the sphere, temperature is finite and maximum (i.e. To = Tmax)
because of symmetry (heat flows from inside to outside radially, in all directions).
B.C. (ii): at r = R, i.e. at the surface, T = Tw
From B.C. (i) and Eq. b, we get: C1 = 0
From B.C. (ii) and Eq. 5.37, we get:
- qg × R 2
Tw = + C2
6× k

qg × R 2
i.e. C2 = Tw +
6× k
Substituting C1 and C2 in Eq. 5.37:

- qg × r 2 qg × R 2
T(r) = + Tw +
6×k 6× k
qg
i.e. T(r) = Tw + × (R 2 – r 2) ...(5.38)
6× k
Eq. 5.38 is the relation for temperature distribution for a solid sphere, in terms of the surface temperature,
Tw . Note that this is a parabolic temperature profile, as shown in Fig. 5.12(b).
Maximum temperature:
Maximum temperature occurs at the centre, because of symmetry considerations (i.e. heat flows from the centre
radially outward in all directions; therefore, temperature at the centre must be the maximum).
Therefore, putting r = 0 in Eq. 5.38:
qg × R 2
Tmax = Tw + ...(5.39)
6× k
From Eqs. 5.38 and 5.39,

T (r ) - Tw FG r IJ 2

Tmax - Tw
=1–
H RK ...(5.40)

198 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eq. 5.40 is the non-dimensional temperature distribution for the solid sphere with heat generation.
Heat flow at the surface:
Of course, in steady state, heat transfer rate at the surface must be equal to the heat generation rate in the sphere,
FG 4 IJ p R q
H 3K
3
Qg = g ...(c)

Heat transfer by conduction at the outer surface of sphere is given by Fourier’s law:
i.e. Qg = – kA(dT/dr)|at r = R
F - q ×R I
i.e. Qg = – k×4× p × R 2× GH 3× k JK
g
(using Eq. 5.38 for T(r))

4
i.e. Qg = × p × R 3× qg ...(5.41)
3
Eq. 5.41 and Eq. c are the same, as expected.
Convection boundary condition:
When heat is carried away at the outer surface by a fluid at a temperature Ta flowing on the surface with a
convective heat transfer coefficient, h a . (e.g. hot spherical ball cooled by ambient air), then, mostly, it is the fluid
temperature that is known and not the surface temperature, Tw , of the sphere. In such cases, we relate the wall
temperature and fluid temperature by an energy balance at the surface, i.e. heat generated and conducted from
within the body to the surface is equal to the heat convected away by the fluid at the surface.
4
i.e. × p × R 3 × qg = ha × (4× p × R 2 )× (Tw – Ta )
3
qg × R
i.e. Tw = Ta + ...(d)
3 × ha
Substituting in Eq. 5.38:
qg × R qg
T(r) = Ta + + × (R2 – r 2) ...(5.42)
3 × ha 6× k
Again, for maximum temperature put r = 0 in Eq. 5.42:

qg × R qg × R 2
Tmax = Ta + + ...(5.43)
3 × ha 6× k
Eq. 5.43 gives the centre temperature of the sphere with heat generation, in terms of the fluid temperature,
when the heat generated is carried away at the surface by a fluid. k, qg
Tw
5.4.2 Alternative Analysis dr
In the alternative method, which is simpler, instead of starting with
the general differential equation, we derive the above equations from
Q
physical considerations. See Fig. 5.13. r
Let us write an energy balance with an understanding that at any To
radius r, the amount of heat generated in the volume within r = 0 and
r = r, must move outward by conduction.
4 dT
× p × r 3 × qg = – k× (4× p × r 2)× ...(a)
3 dr
- qg
i.e. dT = × r× dr
3×k

z z
R
- qg
Integrating: dT = r dr
3×k FIGURE 5.13 Solid sphere with
heat generation

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 199


- qg × r 2
i.e. T(r) = +C ...(b)
6×k
Eq. b gives the temperature distribution along the radius.
Get the constant of integration, C from the B.C.: at r = R, T = Tw
Then, from Eq. b:
qg × R 2
i.e. C = Tw +
6× k
Substituting back in Eq. b:

- qg × r 2 qg × R 2
T(r) = + Tw +
6×k 6× k
qg
i.e. T(r) = Tw + × (R 2 – r 2 ) ...(c)
6× k
Eq. c gives the temperature distribution along the radius, in terms of the surface temperature of the cylinder.
Note that Eq. c is the same as Eq. 5.38 derived earlier.
In many cases, temperature drop between the centre of the sphere (where maximum temperature occurs)
and the surface is important. Then, from Eq. c, putting r = 0:
qg × R 2
Tmax = Tw + ...(d)
6× k
Eq. d is the same as Eq. 5.39.
And, from Eqs. c and d, we can write:

T (r ) - Tw r FG IJ 2

Tmax - Tw
=1–
R H K ...(e)

Eq. e is the same as Eq. 5.40, and gives the non-dimensional temperature distribution in the sphere with heat
generation. If heat generated in the sphere is carried away by convection, by a fluid flowing on the surface of the
sphere, the wall temperature and fluid temperature are related by an energy balance at the surface, as done
earlier.
5.4.3 Analysis with Variable Thermal Conductivity
In the above analysis, thermal conductivity of the material was assumed to be constant. Now, let us make an
analysis when the thermal conductivity varies linearly with temperature as:
k(T) = ko (1 + b T ),
where, ko and b are constants.
Again, considering Fig. 5.13, we have from heat balance (see Eq. a above):
4 dT
× p × r 3 × qg = – k (T) × (4× p × r 2) × ...(a)
3 dr
- qg
i.e. k(T)× dT = × r× dr
3
Substituting for k(T) and integrating:

z ko × (1 + b × T) dT =
- qg
3 z rdr

b ×T 2 - qg × r 2
i.e. T+ = +C ...(f)
2 6 × ko

200 FUNDAMENTALS OF HEAT AND MASS TRANSFER


C is determined from the B.C.: at r = 0, T = To
We get:

b ×To2
C = To +
2
Substituting C in Eq. f:

b ×T 2 qg × r 2 b ×To2
+T+ – To – =0 ...(g)
2 6 × ko 2
Eq. g is a quadratic in T. Its positive root is given by:

F
b qg × r
2
b × To2 I
- 1+ 1 - 4× × GGH
2 6 × ko
- To -
2
JJK
T(r) =
b

2

-1 FG 1 2 ×ToIJ qg × r 2
i.e. T(r) =
b
+
Hb 2
+ To2 +
b K
-
3 × b × ko

F 1 +T I 2
qg × r 2
i.e. T(r) =
-1
b
+ GH b JK o -
3 × b × ko
...(5.44)

Eq. 5.44 gives temperature distribution in a solid sphere with internal heat generation and linearly varying k.
Compare this equation with that obtained for a slab, with temperature at either side being the same, i.e. Eq. 5.10
and that for a solid cylinder, i.e. Eq. 5.25.
Eq. 5.44 gives T(r) in terms of To (i.e. Tmax at r = 0).
If we need T(r) in terms of Tw , then in Eq. f, use the B.C: at r = R, T = Tw
Then we get:
2
b ×Tw2 qg × R
C = Tw + +
2 6 × ko
Substitute this in Eq. f and get a quadratic in T.
Solving, we get, for temperature distribution:

-1 FG 1 + T IJ 2
qg × (R 2 - r 2 )
T(r) =
b
+
Hb K w +
3 × b × ko
...(5.45)

Example 5.16. A solid sphere of radius, R = 10 mm and k = 18 W/(mC) k = 18 W/(mC), qg = 2 ´ 106 W/m3
has a uniform heat generation rate of 2 ´ 106 W/m3. Heat is conducted
away at its outer surface to ambient air at 20°C by convection, with a Q
heat transfer coefficient of 2000 W/(m2C).
(i) Deternine the steady state temperature at the centre and outer Ta = 20°C
surface of the sphere.
(ii) Draw the temperature profile along the radius. R ha = 2000 W(m2.K)
Solution. See Figure Example 5.16. To
Data:
R := 0.01 m ha := 2000 W/(m2 K) k := 18 W/mK) Tw
Ta := 20°C qg := 2 ´ 106 W/m3
To calculate Ta and Tw :
From Eq. 5.39, we have FIGURE Example 5.16 Solid sphere
with heat generation
qg × R 2
Tmax = Tw + ...(5.39)
6× k
ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 201
And, from heat balance on the surface of the sphere,
4
× p × R 3 × qg = ha × (4× p × R 2) × (Tw – Ta)
3
qg × R
i.e. Tw := Ta +
3 × ha
i.e. Tw = 23. 333°C (surface temperature of sphere.)
qg × R qg × R 2
Therefore, Tmax := Ta + +
3 × ha 6× k
i.e. Tmax = 25.185°C (centre temperature of sphere.)
To sketch the temperature profile:
Temperature distribution is given by Eq. 5.42, i.e.
qg × R qg
T(r) := Ta + + × (R 2 – r 2) ...(5.42)
3 × ha 6×k
To sketch the temperature profile in the sphere, define a range variable r, varying from 0 to 0.01 m, with an
increment of 0.0005 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis
and y-axis with r and T(r), respectively. Click anywhere outside the graph region, and immediately the graph
appears: See Fig. Ex. 5.16(b)
r := 0, 0.0005, ... , 0.01 (define a range variable r..starting value = 0,
next value = 0.0005 m and last value = 0.01 m)

T(r) for solid sphere with heat generation


26
r in metres and
T(r) in deg.C
25.5

25

T(r)
24.5

24

23.5

23
0 01 .002 .003 .004 .005 .006 .007 .008 .009 0.01
0.0 0 0 0 0 0 0 0 0
r

FIGURE Example 5.16(b)

It may be verified from the graph that temperature of the centre and outside surface of the sphere are 25.19°C and
23.33°C, respectively.
Example 5.17. In a sphere of radius R, heat generation rate varies with the radius as: qg = qo [1-(r/R)2]. If the thermal
conductivity k, is constant, derive an expression for the variation of temperature with radius.
Solution. This is a case of solid sphere with variable rate of heat generation.
See Figure Example 5.17.
The method is, as usual, to start with the governing equation for the assumtions of the problem, namely, one-
dimensional, steady state conduction with heat generation, with constant k, in spherical coordinates:
d 2 T 2 dT q g
i.e. + × + = 0 ...(a)
dr 2 r dr k

202 FUNDAMENTALS OF HEAT AND MASS TRANSFER


k qg = qo{1 – (r/R)2} Temperature profile,
Q To parabolic

Tw
Solid sphere

R
R

FIGURE Example 5.17(a) Solid sphere with variable FIGURE Example 5.17(b) Variation of temperature
heat generation along the radius

d 2T dT qg × r 2
Multiplying Eq. a by r 2 : r 2 × + 2 × r × + = 0
dr 2 dr k

F
d 2 dT I
qg × r 2
i.e.
drGH
r ×
dr
+
k JK = 0

LM FG r IJ OP × r
2

MN H R K QP
qo × 1 - 2

d FdT I
Substituting for qg:
dr
r2 ×GH
dr
+ JK k
= 0

dT qo × r 3 qo × r 5
Integrating: r2× + – = C1
dr 3× k 5× k × R2

dT q ×r q ×r3 C
i.e. + o - o 2 = 21 ...(b)
dr 3× k 5×k × R r

- qo × r 2 qo × r 4 C
Integrating again: T(r) = + - 1 + C2 ...(c)
6×k 20 × k × R 2 r
Eq. c gives the temperature distribution. Obtain C1 and C2 by applying the B.C.’s:
B.C. (i): at r = 0, dT/dr = 0, since temperature is maximum at the centre due to symmetry.
B.C. (ii): at r = R, T = Tw
From B.C. (i) and Eq. b, we get C1 = 0
From B.C. (ii) and Eq. c
qo × R 2 qo × R 2
C2 = Tw + -
6× k 20 × k
Substituting for C1 and C2 in Eq. c:
qo × r 2 qo × r 4 q × R 2 qo × R 2
T(r) = Tw – + + o -
6×k 20 × k × R 2
6×k 20 × k

qo qo
i.e. T(r) = Tw + × (R2 – r 2) – × (R 4 – r 4) ...(d)
6× k 20 × k × R 2
Eq. d gives the desired temperature distribution in the sphere.
When r = 0, T = To = Tmax. Then, from Eq. d:
qo q × R2 7 qo × R 2
To – Tw = × (R2) – o = × ...(e)
6× k 20 × k 60 k
Eq. e gives the maximum temperature difference in the sphere with heat generation varying with position as:
qg = qo ×[1 – (r/R)2].

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 203


Example 5.18. In Example 5.17, if qo = 106 W/m3, R = 0.04 m, k = 12 W/(mC), and if the centre temperature is 200°C,
determine the surface temperature. Also, find the heat flow rate at the surface. Draw the temperature profile.
Solution.
Data:
LM F r I OP 2
R := 0.04 m
MN GH R JK PQ
qg := qo × 1 - qo := 106 W/m3 k := 12 W/(mC) To := 200°C

To find the surface temperature:


Apply the Eq. d developed in the previous Example:
qo qo
i.e. T(r) = Tw + × (R2 – r 2) – × (R 4 – r 4) ...(d)
6× k 20 × k × R 2
Temperature is maximum when r = 0:
qo qo
i.e. To = Tw + × R2 – × R2
6× k 20× k

qo qo 2
Therefore, Tw := To – × R2 + R
6× k 20× k
i.e. Tw = 184.444°C (temperature in the surface of sphere)
Heat flow rate at the surface, Q:
Apply Fourier’s law at the surface, since, now, we have equation for temperature distribution:
dT (r )
i.e. Q =– k× 4× p × R2×
dr r=R

Now, we have:
LM
T(r) := Tw +
qo
× (R 2 - r 2 ) -
qo OP
×(R4 - r 4 )
MN 6×k 20 × k × R 2 PQ
In Mathcad, we do not have to actually differentiate and expand the expression.
But, define T’(r) = dT(r)/dr and find out T’(r) at r = R:
d
T¢(r) := T(r) (define the first derivative of T(r) w.r.t. T)
dr
Therefore, T¢(R) = – 444.444 (value of T¢(r) at r = R = 0.04 m)
Therefore, Q := – k×4× p × R 2 ×T¢(R), W (define heat transfer rate at the surface)
i.e. Q = 107.233 W (heat transfer rate at the surface.)
To sketch the temperature profile in the sphere, define a range variable r, varying from 0 to 0.04 m, with an increment
of 0.001 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with r
and T(r), respectively. Click anywhere outside the graph region, and immediately the graph appears.
r := 0, 0.001, ..., 0.04 (define a range variable r..starting value = 0,
next value = 0.001 m and last value = 0.04 m)
It may be verified from the graph that temperature at the centre and on the outside surface of the sphere are 200°C
and 184.44°C, respectively.

5.5 Applications
In this chapter, so far, we studied the steady state, one-dimensional heat transfer, with internal heat generation in
simple geometries such as slabs, cylinders and spheres. Now, we shall analyse some practical examples based on
these geometries.
5.5.1 Dielectric Heating
Dielectric heating is a very popular, industrial method of heating adopted to heat insulating materials such as
wool, rubber, plastics and textiles. Here, a high frequency, high voltage alternating current is applied to the plates
of a condenser; the insulating material to be heated is placed between the plates. Heat is generated within the
volume at an uniform rate.

204 FUNDAMENTALS OF HEAT AND MASS TRANSFER


200
198 r in metres and
T(r) in deg.C
196
194
192
T(r) 190
188
186
184
182
180
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
r

FIGURE Example 5.17(b) T(r) vs. r for sphere-variable heat generation

q Wool (Insulation material)


T1 T2
q1 q2
Electrode-1 Electrode-2

Ta
Ta
h2
h1

Plate-1 X Plate-2
dX
L
X

FIGURE 5.14 Dieletric heating

Refer to Fig. 5.14. The insulation material of thickness L is placed between the two electrodes 1 and 2 and
high frequency, high voltage, alternating current is applied. Plates 1 and 2 will also get heated up while the
insulation material is uniformly heated up at a rate of qg (W/m3). Both the plates lose heat to the ambient air at
temperature Ta, with heat transfer coefficients of h1 and h2, respectively. Let the plate temperatures be T1 and T2,
as shown.
It is clear that this situation is similar to a plane wall with uniform heat generation and we shall use the
general differential equation for conduction in Cartesian coordinates, with the following assumptions:
Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the x direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Uniform internal heat generation rate, qg (W/m3).
Consider any section within the volume at a distance x from the origin. Let the temperature at this section
be T.
Now, with the above assumptions, the controlling differential equation is:

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 205


d 2T qg
2
+ =0 ...(a)
dx k
Let us define excess temperature, q = T – Ta , where Ta is constant ambient temperature.
Then, q 1 = T1 – Ta, and
q 2 = T2 – Ta
Also,
dq dT
= , and
dx dx
d 2q d 2T
2
=
dx dx 2
Then, Eq. a may be written as:
d 2q qg
2
+ =0 ...(b)
dx k

dq qg × x
Integrating, + = C1
dx k

qg × x 2
Integrating again, q+ = C1 × x + C2 ...(c)
2× k
Eq. c gives the temperature distribution in the medium.
Integration constants, C1 and C2 are obtained from the B.C’s.
B.C.(i): at x = 0, heat conducted must be equal to the heat removed by convection from Plate 1 to the
ambient.

F dq I
i.e. + kA GH dx JK = h1 A (T1 – Ta ) = h1 Aq 1
x=0
(Note that positive sign is used on the LHS of Fourier’s equation above, since the heat flow on the left plate is
from right to left, i.e. in the negative x-direction).

dq
i.e. k × A × C 1 = h1 × A× q 1 ...since = C1
dx x=0

h1 ×q 1
or, C1 = ...(d)
k
B.C.(ii): at x = 0, q = q1
Therefore, from Eq. c:
C2 = q 1 ...(e)
Substituting C1 and C2 in Eq. c:
- qg × x 2 h1 ×q 1 × x
q(x) = + + q1 ...(5.46)
2×k k
Eq. 5.46 gives the temperature distribution in the medium, in terms of q 1.
q 2 for the plate on the right is obtained by putting x = L and q = q 2 in Eq. 5.46.
- q g × L2 h1 ×q 1 × L
i.e. q2 = + + q1 ...(f)
2×k k

206 FUNDAMENTALS OF HEAT AND MASS TRANSFER


It is obvious that in steady state, total heat generated within the medium must be equal to the sum of heat
convected away at the left and right plates:
i.e. qg × L × A = h1 × A× q 1 + h2×A×q 2
or, qg × L = h1 × q 1 + h2 × q 2 ...(g)
Electrode temperatures T1 and T2 are obtained by solving Eqs. f and g simultaneously.
5.5.2 Heat Transfer through a Piston Crown
Cylinder and piston arrangement is shown in Fig. 5.15.
Piston crown is subjected to a uniform heat flux due to convection and radiation from the hot gases and
cylinder walls. Let this heat flux be qg (W/m2). Let the outside radius of the piston crown be R and its thickness,
b. Let To be the temperature of outer surface of the crown, and k, the thermal conductivity of the crown material.

Piston crown Cylinder dr


To

Heat flux, qg (W/m2)


R
r

FIGURE 5.15 Heat transfer through piston crown

To derive the differential equation governing the temperature distribution in the crown, let us follow the
usual procedure of writing an energy balance on an infinetisimal control volume:
Consider an elemental volume at radius r and of width dr as shown.
Heat conducted into the element at radius, r :
dT
Qr = – k× (2× p × r× b) ×
dr
(Remember that area in Fourier’s equation is the area normal to the direction of heat flow = (2p rb)).
Heat given by gases to the element:
Qg = qg × (2× p × r × dr)
Heat conducted out of the element at radius (r + dr) = Q|r + dr =
d
Qr +
× (Qr)× dr
dr
Then, in steady state, writing an energy balance:
Qr + Qg = Q|r + dr
F dQ I .dr
= Qr + GH dr JKr

d
Therefore, Qg = ×(Qr)× dr
dr

d F dTI
i.e. qg ×2× p ×r× dr =
dr
GH
× - k × 2 ×p × r × b ×
drJK
× dr

d F dT I qg × r
i.e.
dr
GH

dr JK +
k ×b
=0 ...(a)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 207


dT qg × r C
Integrating, + = 1 ...(b)
dr 2 × k × b r

qg × r 2
Integrating again, T(r) + = C1 × ln(r) + C2 ...(c)
4 × k ×b
Eq. c gives the temperature distribution along the radius for the piston crown. C1 and C2 are determined
from the B.C.’s:
dT
B.C.(i): at r = 0, = 0 since temperature is a maximum at the centre by symmetry (i.e. heat flows from
dr
centre to periphery radially).
B.C.(ii): at r = R, T = To
From B.C.(i) and Eq. b: C1 = 0
From B.C.(ii) and Eq. c:
qg × R 2
To + = C2
4 × k ×b
Substituting C1 and C2 in Eq. c
qg × r 2 qg × R 2
T(r) + = To +
4 × k ×b 4 × k ×b
qg
i.e. T(r) = To + × (R2 – r 2) ...(5.47)
4× k × b
Eq. 5.47 gives the temperature distribution along the radius for the piston crown.
Note that the temperature distribution is parabolic.
Maximum temperature:
Maximum temperature occurs at the centre, i.e. at r = 0.
Putting r = 0 in Eq. 5.47, we get:
qg × R 2
Tmax = To + ...(5.48)
4 × k ×b
If Q is the total heat given by gases to the piston crown, then,
Q = p × R 2× qg
Q
i.e. qg =
p × R2

Q R2
Therefore, Tmax = T0 + 2
× ...(5.49)
p ×R 4×k ×b
And, thickness of piston crown:
Q
b= ...(5.50)
4×p × k ×(Tmax - To )
Eq. 5.50 is important, since it gives the thickness required for the piston crown in terms of Q, Tmax and To .
5.5.3 Heat Transfer in Nuclear Fuel Rod (without cladding)
In a nuclear fuel rod, heat is generated by slowing down of neutrons in a fissionable material; however, this heat
generated is not uniform throughout the material, but, varies with position according to the following relation:
LM F r I OP 2
qg = qo × 1 -
MN GH R JK PQ ...(a)

208 FUNDAMENTALS OF HEAT AND MASS TRANSFER


where, qo = heat generation rate per unit volume at the k qg = qo{1 – (r/R)2}
centre (i.e. at r = 0), and Tw
R = outer radius of the solid fuel rod.
We would like to get an expression for the temperature Q
distribution in the fuel rod, maximum temperature in the
rod and, of course, the total heat transferred. See Fig. 5.16. To ha
Assumptions:
Ta
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Internal heat generation at a varying rate: qg = qo.{1
– (r/R)2}, (W/m3).
For these assumptions, the controlling differential
R
equation in cylindrical coordinates becomes:
FIGURE 5.16 Cylindrical fuel rod with heat
d 2T 1 dT qg
2
+ × + =0 ...(b) generation varying with position
dr r dr k
d 2T dT qg × r
Multiplying by r: r × 2
+ + =0
dr dr k

d F
I + q ×r = 0
dT
i.e.
drGH
JK kr×
dr
g

d F dT I q ×r L F r I O
2
G r× J + × M1 - G J P = 0
o
dr H dr K k MN H R K PQ
i.e.

dT q Fr r I 2 4
r× ×G - o
J =C
k H 2 4× R K
Integrating: + 2 1 ...(c)
dr

dT q Fr r I C 3

k GH 2 4 × R JK
o 1
i.e. + × - = 2
dr r

q Fr r I 2 4

k GH 4 16 × R JK
o
Integrating again, T(r) + × - = C × ln(r) + C 2 1 2 ...(d)

Eq. d gives temperature profile within the fuel rod. C1 and C2 are obtained from the B.C.’s:
dT
B.C.(i) at r = 0, = 0 since temperature is a maximum at the centre of the rod.
dr
B.C.(ii) Also, at r = 0, T = Tmax
Then, from Eq. c, C1 = 0
And, from Eq. d: C2 = Tmax
Substituting C1 and C2 in Eq. d:
F
qo r 2 r4 I =T
T(r) + GH
× -
k 4 16× R2 JK max

- qo r 2 F r4 I
i.e. T(r) – Tmax =
k
× GH
-
4 16× R2 JK ...(5.51)

Eq. 5.51 gives the temperature distribution in terms of the centre temperature of the fuel rod.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 209


Surface temperature Tw :
Surface temperature of the rod is obtained by replacing r by R and T(r) by Tw in Eq. 5.51:
F
qo R 2 R4 I
i.e. Tmax – Tw =
k
× GH
4
-
16× R 2
JK
3 × qo × R 2
i.e. Tmax – Tw = ...(5.52)
16 × k
Eq. 5.52 is an important relation, since it gives the maximum temperature drop in the fuel rod. It is necessary
to know this quantity to ensure that sufficient cooling is provided, so that the fuel rod does not get overheated or
melt down.
Heat flow from the surface:
Knowing the temperature distribution, heat flow rate at any point is obtained by applying the Fourier’s law:
At the surface, i.e. at r = R:

F dT I
Q = – kA GH dr JK
r=R

L q F R R I OP
Q = k × A× M × G -
3

MN k H 2 4 × R JK PQ
o
i.e. 2

qo × A × R
i.e. Q= ...(5.53)
4
Convection boundary conditions:
If the heat generated is carried away at the surface by a fluid at temperature Ta , flowing with a convective heat
transfer coefficient of ha , we write the energy balance in steady state, i.e.
Heat generated in the rod = Heat carried away by convection at the surface.
qo × A × R
i.e. = ha ×A× (Tw – Ta )
4
qo × R
i.e. Tw = Ta +
4 × ha
Substituting this value of Tw in Eq. 5.52, we get:

qo × R 3 × qo × R 2
Tmax – Ta = +
4 × ha 16 × k

F
qo × R 1 3 × R I
i.e. Tmax – Ta =
4
× GH
+
ha 4 × k JK ...(5.54)

Eq. 5.54 gives the maximum temperature (i.e. at the centre) in the fuel rod, in terms of the fluid temperature.
Example 5.19. A cylindrical fuel rod is of 20 cm diameter and has k = 40 W/(mK). Surface temperature of the rod is
75°C. Heat generation rate in the rod is given by:
qg = qo {1 – (r/R)2}, where qo = 5.25 ´ 106 W/m3. Determine the temperature at the centre of the rod, and the heat transfer
rate per metre length of rod. Also, draw the temperature profile.
Soution. See Figure Example 5.19.
Data:
LM F r I OP W/m
2

MN GH R JK PQ
3
R := 0.1 m Tw := 75°C k := 40 W/(mK) qg = qo × 1 -

qo := 5.25 ´ 106 W/m 3 L := 1 m

210 FUNDAMENTALS OF HEAT AND MASS TRANSFER


6 2 Temperature at the centre of rod:
k = 40 W/(mK) qg = 5.25 ´ 10 {1 – (r/R )}
We use Eq. 5.52:
Tw = 75°C
3 × qo × R 2
i.e. Tmax – Tw = ...(5.52)
Q 16 × k

3 × qo × R 2
To i.e. Tmax := Tw +
16 × k
i.e. Tmax = 321.094°C (temperature at the centre of the rod.)
Heat transfer rate per metre length of rod:
We use Eq. 5.53:
qo × A × R
i.e. Q= ...(5.53)
4
qo × (2 × p × R × L) × R
R = 0.1 m Therefore, Q := W/m (define Q)
4
4
FIGURE Example 5.19 Cylindrical fuel rod i.e. Q = 8.24668 ´ 10 W ... = 82.4668 KW/m
with heat generation varying with position (heat transfer rate/m.)

To draw temperature profile:


We use Eq. 5.51:
F
- qo r 2 r4 I
i.e. T(r) – Tmax =
k GH
× -
4 16 × R 2 JK ...(5.51)

F
qo r 2 r4 I
i.e. T(r) := Tmax –
k GH
× -
4 16 × R 2 JK ...define T(r).

To sketch the temperature profile in the cylinder, define a range variable r, varying from 0 to 0.1 m, with an increment
of 0.005 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis with r
and T(r), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig. Ex. 5.19(b).
r := 0, 0.005, ... , 0.1 (define a range variable r..starting value = 0,
next value = 0.005 m and last value = 0.1 m)

Temp. distribution in a nuclear fuel rod.


350

r in metres and
300 T(r) in deg.C

250

T(r) 200

150

100

50
0 0.02 0.04 0.06 0.08 0.1
r

FIGURE Example 5.19(b)

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 211


Cladding, kc Fuel rod, kf, qg Note from the graph that temperature at the centre and
surface are 321.1°C and 75°C, respectively.
Q 5.5.4 Heat Transfer in Nuclear Fuel Rod
with Cladding
Generally, fuel rod for use in a nuclear reactor is ‘lagged’
with a tight fitting cladding material, to prevent
oxidation of the surface of the fuel rod by direct contact
with the coolant. Usually, aluminium is used as the
cladding material. We would like to analyse the
temperature distribution and heat transfer in the
Rf combined system of (fuel rod + cladding). Remember
Rc that heat generation occurs only in the fissile material of
T Temperature profile the fuel rod and, in the cladding there is no heat genera-
Tf tion.
Tc In steady state, heat generated in the fuel rod is
conducted through the cladding and then, dissipated to
the coolant flowing around the cladding by convection.
It is assumed that there is no contact resistance between
r the fuel rod and the cladding, i.e. there is continuity of
FIGURE 5.17 Cylindrical fuel rod with cladding heat flux and temperature profile at the interface. See
Fig. 5.17.
Let R f = outer radius of fissionable fuel rod
k f = thermal conductivity of fuel rod
Rc = outer radius of cladding material
kc = thermal conductivity of cladding
material.
Let heat generation rate in the fuel rod vary with position according to the following relation:
LM F r I OP2

MN GH R JK PQ
qg = qo × 1 -

where, qo = heat generation rate per unit volume at the centre (i.e. at r = 0), and
R = outer radius of the solid fuel rod.
Assumptions:
(i) Steady state conduction
(ii) One-dimensional conduction, in the r direction only
(iii) Homogeneous, isotropic material with constant k
(iv) Internal heat generation at a varying rate: qg = qo.{1–(r/R)2 }, (W/m3).
For these assumptions, the controlling differential equation in cylindrical coordinates becomes:
d 2T 1 dT qg
+ × + =0 ...(a)
dr 2 r dr k

d 2T dT qg × r
Multiplying by r: r× 2
+ + =0
dr dr k

d FdT I + q ×r
i.e.
dr
r×GH
dr JK k g
=0 ...(b)

dT -q
Now, = (from Fourier’s law, where q is the heat flux)
dr k

212 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG
d - r ×q qg ×r IJ
Therefore,
dr k H+
k
=0
K
d
or, (r× q) = qg× r ((c) since k is a constant.)
dr
Let us denote fuel rod and cladding materials by subscripts f and c, respectively.
Then,
for fuel rod:
d
(r ×qf) = qg × r
dr

LM F I OP 2

MM GH JK PP × r
d r
i.e. (r× q f ) = qo × 1 - ...(d)
dr
N
Rf
Q
for cladding:
d
(r × qc) = 0 ((e)...since there is no heat generation in cladding)
dr

Fr r I 2 4
Integrating Eq. d GG 2 - 4 × R JJ + C
r× q f = qo ×
H K 2
f
1

Fr r I C 3
q = q ×G -
GH 2 4 × R JJK + r
1
i.e. f o 2
...(f)
f

Integrating Eq. e r× qc = C2
C2
i.e. qc = ...(g)
r
Now, apply the B.C.’s:
B.C. (i): q f = finite, at r = 0
B.C.(ii): q f = qc , at r = Rf , i.e. at the interface
Then, from Eq. f and B.C.(i) C1 = 0
And, from Eqs. f and g, and B.C.(ii):

Rf R 3f F I
C2
= q f = qo × - GG JJ
Rf 2 4 × R 2f H K
C2 qo × R f
i.e. =
Rf 4

qo × R 2f
i.e. C2 =
4
Therefore, heat flux through the fuel rod and cladding may be written as:

dTf Fr r 3 I
q f = – kf × = qo× GG 2 - 4 × R JJ ((h)...from Eq. f)
dr H 2
f K

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 213


dTc qo × R 2f
and, qc = – kc ×= ((i)...putting value of C2 in Eq. g)
dr 4×r
Next, temperatures Tf and Tc in fuel rod and cladding are obtained by integrating Eqs. h and i, respectively:
Fr4 r2 I
Tf =
qo
× GG - JJ + C ...(j)
H 2
k f 16 × R f 4 K 3

- qo × R2f
and, Tc = ×ln(r) + C4 ...(k)
4 × kc
Get C3 and C4 by applying the B.C.’s:
B.C.(iii): Tc = Tw , at r = Rc..i.e. at outer surface of cladding
B.C.(iv): Tc = Tf , at r = Rf , i.e. at the interface
Then, from Eq. k and B.C. (iii):
qo × R 2f
C4 = T w + × ln(Rc)
4 × kc
Immediately substituting C4 in Eq. k, we get:

- qo × R2f F qo × R 2f I
Tc =
4 × kc
× ln(r) + Tw + GH 4 × kc
JK
× ln ( Rc )

qo × R 2f FG R IJ
HrK
c
i.e. Tc = Tw + × ln ...(5.55)
4 × kc

qo × R 2f FG R IJ
HrK
c
and, Tc – Tw = × ln ...(5.56)
4 × kc
Eq. 5.55 gives the temperature distribution in the cladding.
Eq. 5.56 gives the temperature drop across the cladding.
And, from Eq. j and B.C.(iv):
F
R 4f R 2f I
Tf =
qo
× GG - JJ + C3 = Tc
H
k f 16 × R 2f 4 K
3 × qo × R2f
i.e. Tf = C3 – = Tc ...(l)
16 × k f
Then, from Eq. 5.55, and Eq. l:

3 × qo × R 2f FR I
qo × R 2f
C3 =
16× k f
+ Tw +
4×k
GH R JK
c
× ln c
f

qo × R 2f F 3 1 F R II
×G
GH 4 × k + k ×ln GH R JK JJK
i.e. C3 = Tw + c
4 f c f

214 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Substituting C3 in Eq. j:
F
r4 r2 I
i.e. Tf =
qo
× GG - JJ + C ....(j)
H
k f 16 × R 2f 4 K 3

we get:
F I qo × R 2f F 3 F II
Tf =
qo
×
r4
GG -
r2
JJ + T + × GG 4 × k +
1
× lnGH
Rc
JK JJK ...(5.57)
H 2
k f 16 × R f 4 K w
4 H f kc Rf

Eq. 5.57 gives the temperature distribution in the fuel rod.


Maximum temperature in the fuel rod:
This occurs at the centre of the rod.
Putting r = 0 in Eq. 5.57, we get:

qo × R f2 F 3 F II
Tmax = Tw + × GG 4 × k +
1
× ln
Rc
GH JK JJK ...(5.58)
4 H f kc Rf

i.e. Eq. 5.58 gives the maximum temperature in the fuel rod.

5.6 Summary of Basic Conduction Relations, with Heat Generation


In this chapter, we have analysed steady state, one-dimensional heat transfer, with internal heat generation, in
three important geometries, namely, plane slab, cylinder and sphere and derived relations for temperature
distribution, maximum temperature difference and rate of heat transfer. We also studied the effect of variable
thermal conductivity on some of these results. Since all these relations are practically important, they are
tabulated in Table 5.1 to Table 5.7, for easy reference.

TABLE 5.1 Relations for steady state, one-dimensional conduction with internal heat generation, and
constant k

Relation Plane wall Plane wall Plane wall


(both sides at Tw ) (sides at T1 and T2 ) (insulated on one side)
Governing d 2T q g
+ =0 d 2T qg d 2T qg
differential + =0 + =0
dx 2 k dx 2 k dx 2 k
equation

Temperature
T (x) = T w +
qg
×(L2 – x 2) T (x) = T1 + (L - x ) ×
LM qg
+
(T2 - T1) OP T (x) = Tw +
qg
(L2 – x 2)
distribution 2×k N 2 ×k L Q 2× k

Heat transfer
rate at the dT (x )
Q = q g×A ×L Q left = – k × A × at x = 0 qg × A × L
surface, Q , (W) dx

dT (x )
Q right = – k × A × at = x = L
dx

q g ×L2 Equate dT ( x )/ dx to zero; substitute q g ×L2


Tmax – T w , (C)
2 ×k resulting x in T (x ) to get Tmax 2 ×k
Comments L is half-thickness L is the thickness of slab L is the thickness of slab;
of slab; Maximum maximum temperature
temperature occurs occurs on the insulated
on the centre line surface

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 215


TABLE 5.2 Relations for steady state, one-dimensional conduction with internal heat generation and k
varying linearly with temperature as:
k (T ) = ko (1 + b T )
km = ko(1 + b Tm ); Tm = ( T1 + T2)/2

Relation Plane wall of thickness, L (sides at T1 and T2 )

d FG dT IJ
Governing differential equation
dx H
k (T )×
dx
+ qg = 0
K
-1 FG 1 + T IJ 2
2×x qg ×x
Temperature distribution T (x) =
b
+
Hb K 1 -
b× L
×(T1 - T2 ) ×(1 + b ×Tm ) +
b ×ko
×(L - x )

dT (x )
Heat transfer rate at the surface, Q, (W) Q left = – k × A× at x = 0
dx

dT (x )
Q right = – k × A × at x = L
dx

Tmax , (C) Equate dT(x)/dx to zero; Subst. resulting x in T(x) to get Tmax

TABLE 5.3 Relations for steady state, one-dimensional conduction with internal heat generation, and constant k

Relation Solid cylinder Hollow cylinder (inside surface insulated)

dT2
1 dT q g d 2T 1 dT q g
Governing differential equation + × + = 0 + × + = 0
dr 2 r dr k dr 2 r dr k

qg q g × ri 2 LMF r I 2
FG r IJ - FG r IJ 2
OP
× (R 2 – r 2)
MNGH r JK H r K Hr K PQ
o o
Temperature distribution T (r) = T w + T (r) = T o + × - 2 × ln
4× k 4 ×k i i

Heat transfer rate


at the surface Q , (W)
qg ×p × R 2 × L d
qg ×p × ro2 - ri 2 × L i
q g ×R 2 q g × ri 2 Fr I 2
Fr I -1
Tmax – Tw , (C)
4 ×k 4 ×k
× GH r JK
o

i
- 2 × ln GH r JK
o

Comments L is length of cylinder; L is length of cylinder;


maximum temperature occurs maximum temperature occurs
at the centre on the inside surface

5.7 Summary
In this chapter, we studied one-dimensional, steady state heat transfer through simple geometries of a plane slab,
cylinder (both solid and hollow) and sphere, with internal heat generation. Whether the heat generation rate is
uniform or varying with position, the solution technique is, always, to start with the appropriate general
differential equation and solve it by applying the boundary conditions. Once the temperature distribution is
known, rate of heat transfer at any location is easily calculated by applying Fourier’s law.
Problems of heat transfer when the thermal conductivity varies with temperature were also studied.
Applications of these techniques to some practical cases with internal heat generation, such as dielectric
heating, current carrying conductor, nuclear fuel rods with and without cladding, etc. were discussed.

216 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 5.4 Relations for steady state, one-dimensional conduction with internal heat generation, and
constant k
Relation Hollow cylinder Hollow cylinder
(outside surface insulated) (surfaces at T1 and T2 )
Governing
differential d 2T 1 dT q g d 2T 1 dT q g
+ × + = 0 + × + = 0
equation dr 2 r dr k dr 2 r dr k

FrI
L FrI Fr I F r I O2 2 ln GH r JK
× M2 × ln G J + G J - G J P
Temperature q g × ro2 T (r ) - Ti i

MN H r K H r K H r K PQ Fr I
i
T (r) = Ti + =
distribution 4 ×k To - Ti
i o o
ln G J
Hr K
o

LM F r I F r I 2
OP
q dr - r i M GH r JK GH r JK PP
2ln 2 -1
×M
g o i i i

4 × k (T - T ) M F r I F r I
+ × -

MN ln GH r JK GH r JK
o i o o
2

- 1P
P
i i Q
Heat transfer
rate at the
surface, Q , (W)
d
qg × p × ro2 - ri 2 × L i Qinner = – k × Ai ×
dT (r )
dr
at r = ri

dT (r )
Qouter = – k × Ao× at r = ro
dr

q g × ro2 Fr I +F r I 2
q g × ri 2 Fr I 2
Fr I -1
Tmax – Tw , (C)
4 ×k
× 2 × ln GH r JK GH r JK
o

i
i

o
-1
4 ×k
× GH r JK
o

i
- 2 × ln GH r JK
o

Comments L is length of cylinder; L is length of cylinder;


maximum temperature occurs on the Position of maximum temperature
outside surface. is given by:

d i
q g × ro2 - ri 2 - 4 × k × (Ti - To )
rm =
FG r IJ
Hr K
o
q g × 2 × ln
i

TABLE 5.5 Relations for steady state, one-dimensional conduction with internal heat generation and k
varying linearly with temperature as:
k(T) = ko (1 + b T)
km = ko (1 + b Tm ); Tm = (T1 + T2)/2
Geometry Temperature distribution, T(r)

-1 FG 1 + T IJ 2
q g ×(R 2 - r 2 )
Solid cylinder T (r) =
b
+
Hb K w +
2 b ×ko

FG 1 + T IJ LMF r I F I O
GMH r JK - 2 ×ln GH rr JK - 1PP
2 2
-1 q g ×ri 2
Hollow cylinder with inside surface insulated T (r) =
b
+
Hb K i -
2 × b × ko
×
N i
Q i

F 1 +T I 2
q g ×ro2 L Fr I Fr I O
× M2 ×ln G J - G J - 1P
2
Hollow cylinder with outside surface insulated T (r) =
-1
b
+ GH b JK o -
2 × b ×k o MN H r K H r K PQ
o o

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 217


TABLE 5.6 Relations for steady state, one-dimensional conduction with internal heat generation, and
constant k

Relation Solid sphere

d 2T 2 dT q g
Governing differential equation + × + = 0
dr 2 r dr k
qg
Temperature distribution T (r) = Tw + (R 2 – r 2)
6× k

4
Heat transfer rate at the surface, Q, (W) × p × R 3 × qg
3

q g ×R 2
Tmax – Tw , (C)
6 ×k
Comments Maximum temperature occurs at the centre.

TABLE 5.7 Relations for steady state, one-dimensional conduction with internal heat generation and k
varying linearly with temperature as:
k(T) = ko (1 + b T )
km = ko (1 + b Tm ) ; Tm = (T1 + T2)/2

Relation Solid sphere

d 2T 2 dT q g
Governing differential equation + × + = 0
dr 2 r dr k

-1 FG 1 + T IJ 2
q g ×(R 2 - r 2 )
Temperature distribution T (r) =
b
+
Hb K w +
3 b ×ko

4
Heat transfer rate at the surface, Q, (W) × p × R 3 × qg
3
Comments Maximum temperature occurs at the centre.

Several problems were solved and graphical representation of temperature distribution using Mathcad was
highlighted.
Finally, at the end of the chapter, the basic relations developed in this chapter for the aforesaid three
geometries, are tabulated for easy reference.
In the next chapter, we will study an important application of combined heat transfer of conduction and
convection, namely fins or extended surfaces.

Questions
1. Why are the cases with heat generation analysed? Give some practical examples.
2. Derive an expression for temperature distribution under one-dimensional steady state heat conduction with
heat generation of qg (W/m 3) for the following system:
Plate of wall thickness L, thermal conductivity k, temperature being T1 and T2 at the two faces.
3. Pressure vessel for a nuclear reactor is approximated as a large flat plate of thickness L. Inside surface at x = 0 is
insulated. Outside surface at x = L is maintained at a uniform temperature T2. Gamma ray heating of the plate is
represented by:
qg (x) = qo exp (– ax), (W/m3) where qg, qo and a are constants.
(a) Develop an expression for temperature distribution in the plate.
(b) Develop an expression for temperature at the insulated surface (x = 0)
(c) Develop an expression for the heat flux at the outer surface, i.e. at x = L.

218 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4. In the above problem, if the surface at x = 0 is insulated and the surface at x = L dissipates heat by convection
with a convective heat transfer coefficient of h to a fluid at a temperature of Ta, develop an expression for the
temperature distribution in the wall and the temperature of the insulated surface.
5. For solid cylinder of radius R, with heat generation qg (W/m3), surface temperature Tw and centre temperature
T0, show that temperature distribution is given by:
(T – Tw)/ (T0 – Tw ) = 1 – (r/R)2
6. Determine the one-dimensional temperature distribution T(r) for a solid cylinder of radius R, constant thermal
conductivity k, when the heat generation rate varies as:
qg (x) = qo {1 – (r/R)}, (W/m3) where qg , qo are constants. Boundary surface at r = R is kept at zero deg.C.
7. A hollow cylinder of inside radius ri , outside radius ro , has its inner and outer surfaces maintained at uniform
temperature T1 and T2. Inside surface is insulated. Thermal conductivity k is constant and there is uniform heat
generation rate qg (W/m3). Show that:

q g × ri 2 LMF r I 2
F r I - 1OP
MNGH r JK GH r JK P
o o
Tmax – T2 = × - 2 ×ln
4 ×k i i
Q
8. Derive an expression for the variation of temperature along the radius for a solid sphere of constant k when
there is uniform heat generation in the solid. Temperature of the surface (r = R) is Tw .
9. How does the temperature distribution change if the thermal conductivity varies linearly with temperature as: :
k(T) = ko (1 + b T ), where ko and b are constants.
10. In a solid sphere of radius R, heat is generated at a rate of qg = qo{1 – (r/R)2 }, W/m3, where qo is a constant.
Boundary surface at r = R is maintained at a constant temperature Tw . Develop an expression for the steady state
temperature distribution, T(r).

Problems
Plane slab:
1. A plane wall 6 cm thick generates heat internally at the rate of 0.30 MW/m3. One side of the wall is insulated,
and the other is exposed to an environment at 93°C. The convection heat transfer coefficient between the wall
and the environment is 570 W/m2K. Thermal conductivity of the wall is k = 21 W/(mK). Calculate the
maximum temperature in the wall.
2. A large, 3 cm thick plate (k = 18 W/(mK)) has a uniform heat generation rate of 5 MW/m3. Both the sides of the
plate are exposed to an ambient at 25°C. Find out the maximum temperature in the plate and where it occurs.
Draw the temperature profile in the plate.
3. A 4 cm thick brass plate (k = 110 W/(mC)), has uniform internal heat generation rate of 2 ´ 10 5 W/m3. Its one
face is insulated and the other face is exposed to a stream of cooling air at 20°C flowing with a heat transfer
coefficient of 45 W/(m2C). Find the maximum temperature in the plate and where it occurs. Draw the
temperature profile.
4. A steel plate 25 mm thick, (k = 50 W/(mK)) has uniform volumetric heat generation rate of 50 MW/m3. Its two
surfaces are maintained at 150°C and 100°C. Neglecting end effects, determine:
(i) position and value of maximum temperature
(ii) heat flow rate from each surface.
Cylinder:
5. A S.S. rod of 2 cm diameter carries an electric current of 900 A. Thermal and electrical conductivities of the rod
are 16 W/(mC) and 1.5 ´ 104 (Ohm cm) –1, respectively. What is the temperature difference between the centre
line and periphery in steady state?
6. A copper wire 1 mm in diameter is insulated with a plastic to an outer diameter of 3 mm and is exposed to an
environment at 40°C. Find the maximum current carried by the wire in amperes without heating any point of
plastic above 90°C. Heat transfer coefficient from the outer surface of the plastic to the surrounding is 10 W/
(m2K), k of plastic = 0.4 W/(mK), electrical conductivity of copper is 5 ´ 10 7 ohm– 1 m –1. Also, find the maximum
temperature of the wire. Given: k of copper = 380 W/(mK).
7. An electric cable of k = 20 W/(mC), 3 mm in diameter and 1 m long, has resistivity r = 70 ohm.cm. A current of
190 A flows through it and the wire is submerged in a fluid at a temperature of 90°C with a heat transfer
coefficient of 4000 W/(m2C). Calculate the centre temperature of the wire.
8. A chemical reaction takes place in a packed bed (k = 0.5 W/(mC)) between two coaxial cylinders of radii 10 cm
and 35 cm. The inner surface is insulated and is maintained at 500°C. If the reaction produces a uniform heat
generation of 500 kW/m 3, find the temperature of the outer surface.

ONE-DIMENSIONAL STEADY STATE HEAT CONDUCTION WITH HEAT GENERATION 219


9. An electric resistance wire of radius 1.5 mm has k = 25 W/(mC). It is heated by passing a current and heat
generation rate is 2 ´ 109 W/m3. Determine the difference between the temperature at the centre line and the
surface if the surface is maintained at a constant temperature.
10. Consider a copper rod of radius 6 mm, k = 380 W/(mK) wherein heat is generated at a uniform rate of 4.5 ´ 108
W/m3. It is cooled by convection from its surface to ambient air at 30°C with h = 1800 W/(m2C). Determine the
surface temperature of the rod.
11. A hollow S.S. tube with ri = 25 mm, ro = 35 mm, k = 15 W/(mK), electrical resistivity r = 0.7 ´ 10-6 Ohm.m, has
uniform heat generation inside it, induced by an electric current. The heat is transferred by convection to air
flowing through the tube. If the air temperature is 400 K and the convective heat transfer coefficient is 150 W/
(m2K) and the maximum allowable temperature anywhere in the tube is 1400 K, determine the maximum allow-
able electric current. Assume that the tube surface is perfectly insulated. Draw the temperature profile in the
cylindrical shell.
12. A cylindrical rod, 6 cm radius, generates heat at a rate of 2.5 MW/m3. k of the material is 20 W/(mK). It is clad
with a stainless steel layer of 6 mm thickness (k = 14 W/(mK), whose outer surface is cooled by a fluid at 180°C
with a heat transfer coefficient of 600 W/(m2K). Determine the temperature at the centre of the rod and also on
the outer surface and interface. Draw the temperature profile in both the rod and the cladding.
13. Rate of heat generation in a cylindrical fuel rod is given by:
qg = qo {1 – (r/R)2}, W/m 3, where R is the radius of the fuel rod.
(a) Calculate the temperature drop from the centre line to the surface of the rod, for the following data:
diameter of fuel rod = 25 mm, qo = 80 ´ 106 W/m3, k = 20 W/(mK).
(b) If the heat removal rate from the outer surface of the rod is 0.2 MW/m2, what would be the temperature
drop from the centre to the surface?
Sphere:
14. A homogeneous sphere of 9 cm diameter has a uniform heat generation rate of 5 ´ 107 W/m3. k of the material
is 15 W/(mK). If the surface temperature is maintained at 75°C,
(i) determine the temperature at the centre of the sphere
(ii) draw the temperature profile along the radius.
Assume steady state, one dimensional conduction.
15. A solid sphere of radius R = 6 mm, k = 25 W/(mC), has a uniform heat generation rate of 2500 W/m3. Heat is
carried away by convection at its outer surface to ambient air at 30°C with a heat transfer coefficient of 25 W/
(m2C). Determine the steady state temperature at the centre and outer surface of the sphere.
16. Average heat generation during ripening of oranges is estimated as 325 W/m3. Assuming the orange to be a
sphere of diameter 10 cm, and k = 0.15 W/(mC), find out the centre temperature of the orange if the surface is
maintained at 10°C. Draw the temperature profile along the radius.
17. A hollow sphere of 10 cm ID, 20 cm OD, is made of a material of k = 18 W/(mK). Heat is generated internally at
a uniform rate of 3 MW/m3. Inside surface of the sphere is insulated. Develop an expression for the temperature
profile in the sphere and determine the maximum temperature in the material, if the outside surface
temperature is maintained at 300°C. Draw the temperature profile in the shell.

220 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

6
Heat Transfer from
Extended Surfaces (FINS)

6.1 Introduction
In this chapter, we shall discuss heat transfer from extended surfaces, also called Fins.
Fins are generally used to enhance the heat transfer from a given surface.
Consider a surface losing heat to the surroundings by convection. Then, the heat transfer rate Q, is given by
Newton’s Law of Cooling:
Q = h As (Ts – Ta),
where, h = heat transfer coefficient between the surface and the ambient
As = exposed area of the surface
Ts = temperature of the surface, and
Ta = temperature of the surroundings.
Now, if we need to increase the heat transfer rate from the surface, we can:
(i) increase the temperature potential, (Ts – Ta); but, this may not be possible always since both these
temperatures may not be in our control
(ii) increase the heat transfer coefficient h; this also may not be always possible or it may need installing an
external fan or pump to increase the fluid velocity and this may involve cost consideration, or
(iii) increase the surface area As ; in fact, this is the solution generally adopted. Surface area is increased by
adding an ‘extended surface’ (or, fin) to the ‘base surface’ by extruding, welding or by simply fixing it
mechanically.
Addition of fins can increase the heat transfer from the surface by several folds, e.g. an automobile radiator
has thin sheets fixed over the tubes to increase the area several folds and thus increase the rate of heat transfer.
Generally, fins are fixed on that side of the surface where the heat transfer coefficient is low. Heat transfer
coefficients are lower for gases as compared to liquids (see Table 1.1). Therefore, one can observe that fins are
fixed on the outside the tubes in a car radiator, where cooling liquid flows inside the tubes and air flows on the
outside across the fins.
Likewise, in the condenser of a household refrigerator, freon flows inside the tubes and the fins are fixed on
the outside of these tubes to enhance the heat transfer rate.
Typical application areas of fins are:
(i) Radiators for automobiles
(ii) Air-cooling of cylinder heads of internal combustion engines (e.g. scooters, motor cycles, aircraft
engines), air compressors, etc.
(iii) Economizers of steam power plants
(iv) Heat exchangers of a wide variety, used in different industries
(v) Cooling of electric motors, transformers, etc.
(vi) Cooling of electronic equipments, chips, I.C. boards,
etc.
(vii) Fin theory is also used to estimate error in
temperature measurement while using thermometers
(a) (b) (c) (d) or thermocouples.
Types of fins:
There are innumerable types of fins used in practice. Some of
the more common types are shown in Fig. 6.1.
A straight fin or spine is an extended surface added to a
(e) (f) (g)
plane wall. Annular fin is attached circumferentially to a
cylinder to increase its surface area. Fins of rectangular,
circular, triangular, trapezoidal and conical sections are some
of the types commonly used.
Fig. 6.1(a)…longitudinal fin of rectangular profile
Fig. 6.1(b)…cylindrical tube with fins of rectangular
profile
(h) (i)
Fig. 6.1(c)….longitudinal fin of trapezoidal profile
Fig. 6.1(d)….longitudinal fin of triangular profile
FIGURE 6.1 Different types of fins
Fig. 6.1(e)….longitudinal fin of parabolic profile
Fig. 6.1(f)….cylindrical pin fin
Fig. 6.1(g)….truncated conical spine
Fig. 6.1(h)….parabolic spine
Fig. 6.1(i)….cylindrical tube with radial fin of rectangular or truncated conical profile.
Cross-sectional areas of annular fins vary with the radius; in contrast, rectangular or cylindrical spines have
constant cross-sectional area. Triangular or parabolic fins are used when one optimizes the fins from the view
point of weight or volume.
Determination of heat transfer in fins requires information about the temperature profile in the fin. We get
the differential equation describing the temperature distribution in the fin by the usual procedure of writing an
energy balance for a differential volume of the fin. We shall start by doing this for a fin of uniform cross section.

6.2 Fins of Uniform Cross Section (Rectangular or Circular)—Governing


Differential Equation
Let us analyse heat transfer in a fin of rectangular cross section. Same analysis will be valid for a fin of circular
cross section also.
Consider a fin of rectangular cross section attached to the base surface, as shown in Fig. 6.2. Let L be the
length of fin, w, its width and t, its thickness. Let P be the perimeter = 2 (w + t). Let Ac be the area of cross section,
and, To , the temperature at the base, as shown.
Assumptions:
(i) Steady state conduction, with no heat generation in the fin
(ii) Thickness t is small compared to length L and width w, i.e. one-dimensional conduction in the X-
direction only.
(iii) Thermal conductivity, k of the fin material is constant.
(iv) Isotropic (i.e. constant k in all directions) and homogeneous (i.e. constant density) material.
(v) Uniform heat transfer coefficient h, over the entire length of fin.
(vi) No bond resistance in the joint between the fin and the base wall, and
(vii) Negligible radiation effect.
Base temperature, To is higher than the ambient temperature, Ta. Temperature will drop along the fin from
the base to the tip of the fin, as shown in Fig. 6.2(b). Heat transfer will occur by conduction along the length of
the fin and by convection, with a heat transfer coefficient h, from the surface of the fin to the ambient.
Our aim is to derive a differential equation governing the temperature distribution in the fin. Once we get
the temperature field, heat flux at any point can easily be obtained by applying Fourier’s law.

222 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Qconv
Ac
h, Ta
t Y
To To Temperature profile

Qx Qx+dx w TL
X
Ta
X
Z

X dx
L

FIGURE 6.2(a) Rectangular fin of uniform FIGURE 6.2(b) Temperatre profile along
cross section length of fin

Consider an elemental section of thickness dx at a distance x from the base as shown. Let us write an energy
balance for this element:
Energy going into the element by conduction = (Energy leaving the element by conduction + Energy leaving
the surface of the element by convection)
i.e. Qx = Qx + dx + Q conv ...(a)
where,
Qx = heat conducted into the element at x
Qx + dx = heat conducted out of the element at x + dx, and
Qconv = heat convected from the surface of the element to ambient
We have:
dt
Qx = – k ×Ac ×
dx
d dTFG IJ
Qx + dx = – k ×Ac×
dx
T+
dx H
. dx
K
dT d 2T
i.e. Qx + dx = – k ×Ac – k × Ac 2 × dx
dx dx
and, Qconv = h × As (T – Ta)
i.e. Qconv = h × (P × dx) × (T – Ta)
where, As is the surface area of the element P, its perimeter.
Substituting the terms in Eq. a,

dT F dT d 2T I
– k × Ac ×
dx GH
= - k ×Ac
dx dx
JK
- k. Ac 2 ×dx + h × (P ×dx) × (T – Ta)

d 2T
i.e. k × Ac × × dx – h × (P × dx) × (T – Ta) = 0
dx 2
d 2T
i.e. – m2 × (T – Ta) = 0 ...(b)
dx 2

h×P
where m=
k× Ac

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 223


Note that m has units of: (m –1) and is a constant, since for a given operating conditions of a fin, generally h
and k are assumed to be constant.
Now, define excess temperature,
q = T – Ta
dq dT d 2q d 2T
Therefore, = and, =
dx dx dx 2 dx 2
since Ta is a constant.
Substituting in Eq. b,
d 2q
– m2 × q = 0 ...(6.1)
dx 2
Eq. 6.1 is the governing differential equation for the fin of uniform cross section considered.
Eq. 6.1 is a second order, linear, ordinary differential equation. Its general solution is given by calculus
theory, in two equivalent forms:
q (x) = C1 × exp (–m × x) + C2 × exp(m × x) ...(6.2a)
where, C1 and C2 are constants
and,
q(x) = A × cos h (m × x) + B × sin h (m × x) ...(6.2b)
where A and B are constants, and cos h and sin h are hyperbolic functions, defined in Table 6.1.
Eq. 6.2a or 6.2b describes the temperature distribution in the fin along its length.
To calculate the set of constants C1 and C2, or A and B, we need two boundary conditions:
One of the B .C.’s is that the temperature of the fin at its base, i.e. at x = 0, is To, and this is considered as
known.
i.e. B.C. (i): at x = 0, T = To
Regarding the second boundary condition, there are several possibilities:
Case (i): Infinitely long fin,
Case (ii): Fin insulated at its end (i.e. negligible heat loss from the end of the fin),
Case (iii): Fin losing heat from its end by convection, and
Case (iv): Fin with specified temperature at its end.
It may be remarked here, that while for case (i), it is convenient to choose the solution in the form given by
Eq. 6.2a and for cases (ii) and (iii), choosing the solution in the form given by Eq. 6.2b makes the analysis easy.
Before we proceed further, let us tabulate a few useful relations for hyperbolic functions: (See Table 6.1).
6.2.1 Infinitely Long Fin
This simply means that the fin is very long. Consequence of this assumption is that temperature at the tip of the
fin approaches that of the surrounding ambient as the fin length
approaches infinity. See Fig. 6.3 (a).
h, Ta
To determine the temperature distribution:
Ac
To The governing differential equation, as already derived, is given by
Eq. 6.1, i.e.

Q t d 2q
– m2 × q = 0 ...(6.1)
dx 2
And, we shall choose for its solution for temperature distribu-
TL = Ta tion Eq. (6.2a), i.e.
q (x) = C1 × exp (– m × x) + C2 × exp (m × x) ...(6.2a)
L®¥ C1 and C2 are obtained from the B’C.’s:
B.C. (i): at x = 0, T = To
X B.C. (ii): as x ® ¥, T ® Ta, the ambient temperature.
From B.C. (i):
FIGURE 6.3(a) Infinitely long fin of at x = 0, q (x) = To – Ta = qo
uniform cross section

224 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 6.1 Relations for hyperbolic functions
Sl. No. Relation
exp(b ) + exp(- b )
(a) cos h (b ) =
2
exp(b ) - exp(- b )
(b) sin h (b ) =
2
(c) exp (b ) + exp(– b ) = 2 cos h (b )
(d) exp (b ) – exp(– b ) = 2 sin h (b )
(e) exp (b ) = cos h (b ) + sin h (b )
(f) exp (– b ) = cos h (b ) – sin h (b )
(g) sin h (0) = 0
(h) cos h (0) = 1
d
(i) (sin h (m ×x) = m ×cos h (m × x)
dx
d
(j) cos h (m ×x) = m × sin h (m ×x)
dx
(k) cos h (– x) = cos h (x)
(l) sin h (– x) = – sin h (x)
(m) cos h (x + y) = cos h (x) ×cos h (y) + sin h (x) ×sin h (y)
(n) cos h (x – y) = cos h (x) ×cos h (y) – sin h (x) ×sin h (y)
(o) sin h (x + y) = sin h (x) ×cos h (y) + cos h (x) ×sin h (y)
(p) sin h (x – y) = sin h (x) ×cos h (y) – cos h (x) ×sin h (y)

From B.C. (ii): q/qo


at x = ¥, q (x) = Ta – Ta = 0
1
From B.C. (ii) and Eq. 6.2a: C2 = 0 Temperature profile
From B.C. (i) and Eq. 6.2a: C1 = q o
m1 < m 2 < m 3
Substituting C1 and C2 back in eqn. 6.2a, we get:
q (x) = qo × exp (– m ×x) m1
m2
q ( x) m3
i.e. = exp (– m ×x)
qo 0 X
FIGURE 6.3(b) Dimensionless temperature
T ( x) − Ta
i.e. = exp (– m ×x) ...(6.3) profile along length of fin
To − Ta
Eq. 6.3 gives the temperature distribution in an infinitely long fin of uniform cross section, along the length.
This is shown graphically in Fig. 6.3b. Note that temperature distribution is exponential.
It may be observed from the graph that as the parameter m increases, dimensionless temperature ratio falls
steeply. As the fin length tends to infinity, dimensionless temperature ratio approaches zero, as shown in the Fig.
6.3(b)
To determine the heat transfer rate:
Heat transfer rate from the fin may be determined by either of the two ways:
(a) by the application of Fourier’s law at the base of the fin, i.e. in steady state, the heat transfer from the fin
must be equal to the heat conducted into the fin at its base.
i.e. Qfin = – k AcdT(x)/dx|x = 0 = – k Ac dq (x)/dx|x = 0 ...(c)

z z
(b) by integrating the convective heat transfer for the entire surface of the fin, i.e.
L L
Qfin = h ×P ×(T - Ta ) dx = h × P ×q dx ...(d)
0 0

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 225


Of course, the results obtained by both the methods must be the same; but applying method (a) is easier.
By method (a):
FG d T (x)IJ FG d q (x)IJ
Qfin = – k ×Ac ×
H dx K x =0
= – k × Ac ×
H dx K x =0
...(c)

= – k×A × M
L d eq ×e - m×x
jOPQ
i.e. Qfin c
N dx o
x =0

= – k × A × (– m) × q .e e j
- m× x
i.e. Qfin c o
x=0
i.e. Qfin = k × Ac × m × qo ...(6.4)
Substituting for m:

h. P
Qfin = k × Ac × ×q o
k. Ac

i.e. Qfin = h × P × k × A c ×q o = c
h × P × k × A c × To - Ta h ...(6.5)
Eq. 6.4 or 6.5 gives the heat transfer rate through the fin.
Let us verify this result from method (b):

z z
By method (b):
L L
Qfin = h × P ×(T - Ta ) dx = h × P ×q dx ...(d)

z
0 0
¥
i.e. Qfin = h × P ×q o × e - m. x dx
0

1
i.e. Qfin = × h × P ×q o
m
i.e. Qfin = h × P × k × A c ×q o = h ×P ××
k A c × To − Ta b g ...(same as Eq. 6.5)

6.2.2 Fin of Finite Length with Insulated End


End of a fin is generally not insulated; so, here, what we mean is that the heat transfer from the end of the fin is
negligible as compared to the heat transfer from the surface of the fin. Mostly, this is true, since the area of the
end of fin is negligible compared to the exposed surface area of the fin; in fact, this is the most important case. See
Fig. 6.4.

h, Ta Ac
To

Q t To
Temperature profile
(dT/dx)x = L = 0

L Ta TL

X X

FIGURE 6.4(a) Fin of finite length, end insulated FIGURE 6.4(b) Temperature profile for fin
insulated at its end

226 FUNDAMENTALS OF HEAT AND MASS TRANSFER


To determine the temperature distribution:
The governing differential equation., as already derived, is given by Eq. 6.1, namely,

d2 q
– m2 × q = 0 ...(6.1)
dx 2
And, we shall choose for its solution for temperature distribution, Eq. 6.2b, i.e.
q (x) = A × cos h (m × x) + B × sin h (m × x) ...(6.2b)
Constants A and B are obtained from the B’C.’s:
B.C.(i):
at x = 0, q (x) = To – Ta = qo
B.C. (ii):
dT dq
at x = L, = = 0 since the end is insulated.
dx dx
From B.C. (i) and Eq. 6.2b:
A = qo
From B.C. (ii) and Eq. 6.2b:
FG dq IJ
H dx K x=L
=0

i.e. A× m× sin h(m× L) + B× m× cos h(m× L) = 0 (using relations in Table 6.1)


Substituting fo A: qo× (m× sin h(m × L) + (B × m × cos h (m × L) = 0)
sin h( m×L)
i.e. B = – qo ×
cos h(m ×L)
Substituting for A and B in Eq. 6.2b
sin h( m×L)
q (x) = qo × cos h (m × x) – qo × × sin h (m × x)
cos h(m ×L)
q (x) cos h (m×L )×cos h( m×x ) - sin h( m×L)×sin h( m. x )
i.e. =
qo cos h (m×L )
q (x) cos h(m×(L - x ))
i.e. = ((6.6)...using relation no. (n) from Table 6.1)
qo cos h(m×L)
T ( x ) - Ta cos h(m×(L - x ))
i.e. = ...(6.7)
To - Ta cos h(m×L)
Eq. 6.6 or 6.7 gives the temperature distribution in the fin with negligible heat transfer from its end.
Same relations are obtained if we start with the general solution for temperature distribution as given by Eq.
6.2a; however, algebraic manipulations required are rather lengthy.
Temperature at the end of the fin:
This is easily determined by putting x = L in Eq. 6.6 or 6.7:
q (L ) 1
i.e. = (6.6a ... since cos h (0) = 1)
qo cos h(m ×L)

TL - Ta 1
and, = ...(6.7a)
To - Ta cos h(m ×L)
To - Ta
or, TL = + Ta ...(6.7b)
cos h(m ×L)
Eq. 6.7b gives the temperature at the end of a fin (i.e. at x = L), when the end of the fin is insulated.

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 227


To determine the heat transfer rate:
Heat transfer rate from the fin may be determined by the application of Fourier’s law at the base of the fin, i.e. in
steady state, the heat transfer from the fin must be equal to the heat conducted into the fin at its base.
i.e. Qfin = – k Ac dT(x)/dx|x = 0 = – k Ac dq(x)/dx|x = 0
Therefore,

Qfin = – k × Ac × qo ×
LM - m.sin h(m.(L - x)) OP
N cos h(m.L) Q x=0

i.e. Qfin = k × Ac × m × qo × tan h (m × L) ...(6.8)


i.e. Qfin = h × P × k × Ac ×q o × tan h( m×L ) ...(6.9)
Remember: qo = (To – Ta).
Eq. 6.8 or 6.9 gives the heat transfer rate from the fin, insulated at its end.
Comparing Eq. 6.8 with that obtained for heat transfer from an infinitely long fin, i.e. Eq. 6.4, we see that a
fin with insulated end becomes equivalent to an infinitely long fin when tan h (m . L) = 1.
Table 6.2 below shows values of tan h (m . L) for values of (m . L) ranging from 0 to 5; same table is also shown
in graphical form on the right, for easy visualisation.
It is observed from the Table 6.2 that when (m × L) for the insulated-end fin reaches a value of about 2.8, heat
transfer rate becomes about 99% of that obtained for an infinitely long fin. And, beyond a value of (m × L) more

TABLE 6.2 Values of tan h(X) for


different values of X

X tan h(X)
0 0
0.2 0.19738
0.4 0.37995
X versus tan h(X) and {1/cos h(X)}
0.6 0.53705 1
0.8 0.66404
0.9
1 0.76159
1.2 0.83365 0.8
1.4 0.88535 0.7
1.6 0.92167 tan h(X) 0.6
1.8 0.94681 1 0.5
2 0.96403
cos h(X) 0.4
2.2 0.97574
2.4 0.98367 0.3
2.6 0.98903 0.2
2.8 0.99263 0.1
3 0.99505
0
3.2 0.99668 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
3.4 0.99777 X
3.6 0.99851
3.8 0.999 Note: X = m×L
4 0.99933 h◊p
4.2 0.99955 m=
k◊Ac
4.4 0.9997
4.6 0.9998
4.8 0.99986
5 0.99991

228 FUNDAMENTALS OF HEAT AND MASS TRANSFER


than 5, the fin with insulated end can be considered as infinitely long. Therefore, from the heat transfer point of
view, there is no great advantage in having a fin with (m × L) greater than 2.8 or 3.
6.2.3 Fin of Finite Length Losing Heat from its End by Convection
This is a more realistic case, though the relations developed are a little more complicated, as we shall see pres-
ently. See Fig. 6.5.

h, Ta Ac
To

Q t To
Temperature profile
–k(dT/dx)x = L = h.qL
Convection
TL
L Ta

X X

FIGURE 6.5(a) Fin of finite length, end losing FIGURE 6.5(b) Temperature profile for fin losing
heat by convection heat at its end

Here, heat conducted to the tip of the fin must be equal to the heat convected away from the tip to the
ambient, i.e.
F dT I
– k×Ac × GH dx JK x=L
= h × Ac × (TL – Ta)

– k× G
F dT I
i.e.
H dx JK x=L
= h× q L

To determine the temperature distribution:


The governing differential equation., as already derived, is given by Eq. 6.1, i.e.
d 2q
– m2 × q = 0 ...(6.1)
dx 2
And, we shall choose for its solution for temperature distribution, Eq. 6.2b, i.e.
q (x) = A × cos h (m × x) + B × sin h (m × x) ...(6.2b)
Constants A and B are obtained from the B’C.’s:
B.C.(i):
at x = 0, q(x) = To – Ta = qo
Applying B.C.(i) to Eq. 6.2b:
A = qo
B.C. (ii): at x = L,
heat conducted to the end = heat convected from the end
FG dq (x) IJ
i.e. – k × Ac ×
H dx K x=L
= h × Ac × q (L) where q (L) = TL – Ta

k× G
F dq (x) IJ
i.e.
H dx K x=L
+ h × q (L) = 0

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 229


i.e. k× [A × m × sin h(m × L) + B × m × cos h (m × L)] + h × [A × cos h (m × L) + B × sin h (m . L)] = 0
(using relations in Table 6.1)
Substituting for A:
A× {(m × k × sin h (m × L) + h × cos h (m × L)} + B× {(m × k × cos h (m × L) + h.sin h (m × L)}

F h I
GH
- q o × sin h( m×L) +
m ×k
× cos h(m×L ) JK
i.e. B= , since A = qo.
h
cos h( m×L ) + × sin h( m×L )
m× k
Now, substitute for A and B in the general solution given by Eq. 6.2:

F sin h(m×L) + h ×cos h(m×L)I


q ( x) GH m× k JK
= cos h (m × x) – × sin h (m × x)
qo h
cos h( m×L) + ×sin h( m×L)
m. k

F cos h(m×x)×cos h(m×L) + h ×cos h(m×x)×sin h(m×L)I - sin h(m×x)×sin h(m×L) - h ×cos h(m×L)×sin h(m×x)
q ( x) GH m. k JK m. k
i.e. =
qo h
cos h( m ×L) + ×sin h ( m ×L)
m ×k

h
[cos h(m ×L).cos h(m×x) - sin h(m×L)×sin h(m×x)] + ×[sin h(m× L)×cos h(m× x) - cos h(m×L).sin h (m×x )]
q ( x) m×k
i.e. =
qo h
cos h(m×L) + ×sin h(m×L)
m×k

h
cos h ( m×( L - x)) + ×sin h ( m ×( L - x))
q ( x) m ×k
i.e. = ((6.10)...using relations (n) and (p) from Table 6.1)
qo h
cos h ( m×L) + ×sin h ( m ×L)
m ×k
Eq. 6.10 gives the temperature distribution in a fin losing heat by convection at its end.
Remember again that:
q(x) = T(x) – Ta
and, q o = To – Ta
Note that when h = 0, i.e. for negligible heat transfer at the tip of the fin, Eq. 6.10 reduces to Eq. 6.6, for a fin
with insulated tip.
To determine the heat transfer rate:
Heat transfer rate from the fin may be determined by the application of Fourier’s law at the base of the fin, i.e. in
steady state, the heat transfer from the fin must be equal to the heat conducted into the fin at its base.
i.e. Qfin = – k Ac [dT(x)/dx]x = 0 = – k Ac dq(x)/dx]x = 0

F - m×sin h(m×L) - h ×m×cos h(m×L)I


GH m×k JK
i.e. Qfin = – k×A ×q ×
c o
F cos h(m.L) + h ×sin h(m×L)I
GH m× k JK
F sin h(m×L) + h ×cos h(m×L)I
GH m×k JK
i.e. Qfin = k×A ×m×q ×
c
F cos h(m×L) + h ×sin h(m×L)I
o

GH m× k JK
230 FUNDAMENTALS OF HEAT AND MASS TRANSFER
F tan h(m.L) + h I
GH m. k JK
i.e. Qfin = k×A ×m×q ×
c
F 1 + h ×tan h(m×L)I
o ...(6.11)
GH m.k JK
Eq. 6.11 gives the heat transfer rate from a fin losing heat by convection at its tip.
Note: Eq. 6.11 is important since it represents the heat transfer rate for a practically important case of a fin
losing heat from its end. However, it is rather complicated to use. So, in practice, even when the fin is losing heat
from its tip, it is easier to use Eq. 6.8 or 6.9 obtained for a fin with insulated tip, but with a corrected length, Lc
rather than the actual length, L, to include the effect of convection at the tip. In that case, only to evaluate Q, L is
replaced by a corrected length Lc, in Eq. 6.8 or 6.9, as follows:
t
For rectangular fins: Lc = L + where t is the thickness of fin
2
r
For cylindrical (round) fins: Lc = L + where r is the radius of the cylindrical fin.
2
6.2.4 Fin of Finite Length with Specified Temperature at its End
This type of problem occurs very often in practice, e.g. when a structural member is used as a heat shunt between
two heat reservoirs. Then, the problem is to find out the heat transfer through that member.
Let us formulate the problem as follows:
Problem. A thin fin of length L has its two ends attached to two Qconv
parallel walls, maintained at temperatures T1 and T2, as shown in
h, Ta
Fig. 6.6. The fin loses heat by convection to the ambient air at Ta.
Assuming one-dimensional conduction, derive an expression for T1 T2
temperature distribution in the fin. Then, deduce an expression for Q1 Q2
the heat lost by the fin.
To determine the temperature distribution:
The governing differential equation., as already derived, is given by L
Eq. 6.1, i.e.
d 2q X
– m2 × q = 0 ...(6.1)
dx 2
And, we shall choose for its solution for temperature distribution, Temperature profile
T1
Eq. 6.2b i.e.
q (x) = A ×cos h(m ×x) + B × sin h(m × x) ...(6.2b) T2
Constants A and B are obtained from the B’C.’s:
B.C.(i): at x = 0: T = T1 i.e. q = q1
B.C.(ii): at x = 0: T = T2 i.e. q = q2 X
From B.C.(i) and Eq. 6.2b: FIGURE 6.6 Fin of finite length, with speci-
A = q1 fied temperature at two ends and the
From B.C.(ii) and Eq. 6.2b: temperature profile along the length
q2 = A × cos h (m × L) + B× sin h (m× L)
i.e. q2 = q1× cos h(m× L) + B× sin h (m× L)
q 2 - q 1 ×cos h(m×L)
Therefore, B=
sin h(m×L)
Substituting for A and B in Eq. 6.2b:
q 2 - q 1 ×cos h(m×L)
q (x) = q1 × cos h(m × x) + × sin h(m × x)
sin h(m×L)
q 1 ×sin h(m×L)×cos h(m× x) - q 1 ×cos h(m×L)×sin h(m×x) + q 2 × sin h(m× x)
i.e. q (x) =
sin h(m × L)

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 231


q 1×sin h(m×(L - x)) + q 2 ×sin h(m× x)
i.e. q (x) = ((6.12)...using relation (p) from Table 6.1)
sin h(m×L)
Eq. 6.12 gives the temperature distribution along the length of fin, when its two ends are maintained at two speci-
fied temperatures.
To determine the heat transfer rate:

z z
Total heat transfer rate from the fin is determined by integrating the convection heat transfer over the length of the fin:
L L
Qfin = h × (P × dx) × (T(x) – Ta) = h × (P × dx) × q (x)

z
0 0
L
i.e. Qfin = h×P ×q (x) dx

z
0

L
q 1 sin h( m ×( L - x)) + q 2 ×sin h( m ×x )
i.e. Qfin = h × P × dx
0 sin h ( m×L)

Qfin =
hP LM - q cos h(m(L - x)) + q cos h(mx) OP
1 2
L

i.e.
sin h ( mL) N m m Q 0

h ×P L - q ×(1 - cos h(m×L)) + q ×(cos h(m×L) - 1)OP


sin h(mL) MN m
Qfin = 1

Q
2
i.e. ×
m

h× P
i.e. Qfin = × [(q1 + q2) × (cos h (m × L) –1)]
m×sin h(m× L)

h× P
But, m=
k × Ac
Therefore, substituting for m:
F cos h(m×L) - 1I
Qfin = h× P ×k × Ac × (q1 + q2) × GH sin h(m×L) JK
F cos h(m×L) - 1I
+ q )× G
i.e. Qfin = k × Ac× m × (q 1 2
H sin h(m×L) JK ...(6.13)

Eq. 6.13 gives the heat transfer rate for a fin with specified temperatures at its both ends.
To find the minimum temperature in the fin:
Differentiate the expression for q (x), i.e. Eq. 6.12 w.r.t. x and equate to zero; solving it, we get xmin, the position where
minimum temperature occurs. Then, substitute this value of xmin back in Eq. 6.12 to get the value of Tmin. (Remember:
q (x) = T (x) – Ta).
When both the ends of fin are at the same temperature:
Now, T1 = T2 (i.e. q 1 = q 2 ), and obviously, the minimum temperature will occur at the centre, i.e. at x = L/2.
Then, substituting q 1 = q 2 and x = L/2 in Eq. 6.12, we get for minimum temperature:
q 1 ×sin h(m×(L - x)) + q 2 ×sin h(m× x)
q (x) = ...(6.12)
sin h(m×L)

LM FG L IJ OP + q ×sin h FG m× L IJ
Therefore, qmin =
MN H
q 1 ×sin h m × L -
2 KQ 1
H 2K
sin h( m ×L)

FG m× L IJ
i.e. qmin =
2×q 1 ×sin h
H 2K ...(6.14)
sin h(m× L)
Remember: qmin = Tmin – Ta

232 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Let us make two important notes: Qconv
Note 1: Pin fin (or, spine) of uniform cross section:
See Fig. 6.7. All the above analysis for a fin of rectangular cross
section shown in Fig. 6.2, is valid for a fin of uniform circular h, Ta
cross section too.
To
Note 2: Fin parameter, m:
It may be observed from all the derivations for fins done so far, D
that the parameter m occurs in all the equations. By definition,
h× P
m= L
k × Ac
where, h is the heat transfer coefficient between the fin surface
and the ambient, Ac is the cross-sectional area of the fin, P is the
perimeter of the fin section and, k is the thermal conductivity of
the fin material. Units of m is: m –1 . X
(a) For rectangular fin of Fig. 6.2:
We get:
FIGURE 6.7 Fin of circular (round) cross section
Ac = w × t
P = 2 × (w + t)
where, w = width of fin and t = thickness of fin
Therefore,
h× P
m=
k × Ac

2×h×(w + t)
i.e. m=
k ×w ×t
Then, for thin fins, i.e. w << t, we can write:
2×h
m= (for thin fins)
k ×t
(b) For round fin (or pin fin) of Fig. 6.7:
In this case,
p ×D 2
Ac =
4
P = p×D
where, D is the diameter of the fin.
Therefore,

h× P h×p ×D
m= =
k × Ac p ×D2

4

4× h
i.e. m= (for round (or pin) fins)
k ×D

6.2.5 Summary of Fin Formulae


The foregoing results, derived for fins with different boundary conditions at the tip, are summarized in Table 6.3
for easy reference.
Example 6.1. (a) A very long, 25 mm diameter copper rod (k = 380 W/(mC)), extends horizontally from a plane heated
wall at 150°C. Temperature of surrounding air is 30°C and heat transfer coefficient between the surface of the rod and
the surroundings is 10 W/(m2K).
(i) Determine the rate of heat loss from the rod
(ii) How long the rod should be to be considered as infinite?
(iii) Draw the temperature profile along the length of the rod.

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 233


TABLE 6.3 Temperature distribution and heat transfer rate for fins of uniform cross section
*(x) = (T(x) – Ta ), m = /{h × P/(k × Ac)}

Case Tip condition (x = L) Temperature distribution, q(x)/qo Heat transfer rate, Qfin
q (x )
1 Infinitely long = exp (– m ×x) Qfin = k ×Ac ×m ×qo
qo
L ® ¥, q (L) = 0
Insulated at the tip q (x ) cos h (m ×(L - x ))
2 = Qfin = k × Ac × m× qo × tan h (m × L)
(dq / d x )|x = L = 0 qo cos h (m ×L )

q (x )
3 Convection from tip = Qfin = k × Ac × m × q o x
qo

h FG tan h(m ×L) + h IJ


H m ×k K
cos h (m ×(L - x )) + ×sin h (m ×(L - x ))
dq m ×k
–k
dx
= h×q(L)
cos h (m ×L ) +
h
.sin h (m ×L ) FG1 + h ×tan h (m ×L)IJ
H m ×k K
x =L
m ×k

q 1 sin h ((m ×(L - x )) + q 2 ×sin h (m ×x )


4(a) Prescribed temperatures q (x) = Qfin = k × Ac × m × (q 1 + q 2 )x
sin h (m ×L )
at the tip, ends
FG cos h (m×L) - 1IJ
x = 0 ® q = q1 H sin h (m×L) K
x = L ® q = q2
When temperatures at both
ends are equal, q 1 sin h(m ×(L - x )) + q 1 sin h (m ×x )
4(b) q (x) = Qfin = k × Ac × m × (2 × q 1).
T1 = T2 or, q 1 = q 2 sin h (m ×L)

FG cos h (m×L) - 1IJ


H sin h (m×L) K
Minimum temperature is given by:
FG m ×L IJ
qmin =
2 ×q 1 × sin h
H 2K
sin h (m ×L )

To = 150°C (b) Compare the temperature distribution in the rod if the materials
2 were: (i) copper (k = 380 W/(mC)), (ii) aluminium (k = 200 W/(mC))
h = 10 W/(m C)
and, (iii) steel (k = 55 W/(mC)). Other data is the same as in part (a).
Ta = 30°C
2 Solution. Since it is stated that it is a very long rod, we will take L
k = 380 W/(m C) as ¥. So, relations derived for an infinitely long fin apply.
See Fig. Example 6.1.
Data:
D = 0.025 m D := 0.025 m L := ¥ m k := 380 W/(mC) To := 150°C
Ta := 30°C h := 10 W/(m2C)
Heat transfer rate from the rod:
L®¥
First, let us calculate the parameter m:
h× P
We have: m = where, P is the perimeter and Ac is the
k × Ac
X
area of cross section.
FIGURE Example 6.1 Fin of circular (round) p ×D 2
Then, Ac: = m2 (define the area of cross section of the rod)
cross section 4

234 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Ac = 4.909 ´ 10–4 m2 (area of cross section of the rod)
and, P: = p × D, m (define the perimeter of the rod)
i.e. P = 0.079 m (perimeter of the rod)
h× P
Therefore, m: = m–1 (define the parameter m.)
k × Ac
i.e. m = 2.052 m–1 (parameterm.)
Now, apply Eq. 6.4 for heat transfer from a a very long fin:
Qfin = k× Ac × m × qo ...(6.4)
i.e. Qfin : = k× Ac × m × (To – Ta)
Substituting values: Qfin = 45.931 W (heat loss rate from the fin.)
Length of rod required to consider it as infinitely long:
Read the discussion under section 6.2.2.
An infinitely long fin has no heat transfer from its end since the end temperature tends to the ambient temperature as the
length tends to infinity. Therefore, comparing the expressions for Q for an infinitely long fin and a fin with insulated at
its end, i.e.
Qfin = k× Ac × m × qo (for infinitely long fin)
Qfin = k × Ac × m× qo × tan h (m×L) (for a fin with insulated end,)
we see that they are equivalent when tan h (m × L) is equal to 1.
From Table 6.2, it is seen that at (m × L) = 5, tan h (m × L) is almost equal to 1.
Therefore, the rod can be considered as infinitely long, if: m × L > 5, or L > (5/m):
5
i.e. L: = m (define L)
m
i.e. L = 2.437 m (length of rod rquired to consider it as infinitely long.)
To draw the temperature profile in the rod:
We need the equation for temperature profile.
Eq. 6.3 gives the temperature profile for a very long fin:
T (x) - Ta
i.e. =exp (– m× x) ...(6.3)
To - Ta
Therefore, T(x) := Ta + (To – Ta) × exp(– m × x) (equation for temperature profile in the rod)
We use Mathcad to draw the temperature profile. First, define a range variable x, varying from 0 to say, 2.5 m, with
an increment of 0.1 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-
axis with x and T (x), respectively. Click anywhere outside the graph region, and immediately the graph appears.
(Fig. Ex. 6.1, b)
x := 0, 0.1, ... , 2.5 (define a range variable x.. starting value = 0,
next value = 0.1 m, and last value = 2.5 m)
Observe from the graph that at x = 0, the temperature is 150°C and at the end of the fin, the temperature is 30°C
which is that of the ambient. This matches with the boundary conditions of the problem.
(b) Compare the temperature distribution if the fin materials are aluminium (k = 200 W/(mC)) and stainless steel
(k = 55 W/(mC)):
For aluminium: kA1 := 200 W/(m × C) (thermal conductivity of aluminium)
h× P
Therefore, mA1 := m–1 (define fin parameter m for aluminium)
kA1 × Ac
i.e. mA1 = 2.828 m–1
and corresponding length required to be considered as infinitely long is:
5
L := i.e. L = 1.768 m
mA 1
For steel: ks := 55 W/(m× C) (thermal conductivity of steel)
h× P
Therefore, ms := m–1 (define fin parameter m for steel.)
ks × Ac
i.e. ms = 5.394 m–1

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 235


Temperature distribution for infinitely long fin
150
140
130
120
110
100
90
80 x in metres and
T(x)
70 T(x) in deg.C
60
50
40
30
20
10
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
x

FIGURE Example 6.1(b)

and, corressponding length required to be considered as infinitely long is:


5
L: = i.e. L = 0.927 m
ms
To draw the temperature profiles for the three materials:
First, define T as a function of x and m:
i.e. T (x, m) := Ta + (To – Ta) × exp (m × x) (equation for temperature profile in the rod)
Then, we have: mcu := 2.052 ...parameter m for copper...calculated in part (a)
mA1 = 2.828 ...parameter m for aluminium
ms = 5.394 ...parameter m for steel
We use Mathcad to draw the temperature profiles. First, define a range variable x, varying from 0 to say, 3.0 m,
with an increment of 0.1 m. Then, choose x–y graph from the graph palette, and fill up the place holder on the x-axis
with x; and in the place holder on the y-axis, fill up T(x, mcu ), T(x, m Al ), T(x, ms ) separated by commas as shown. Click
anywhere outside the graph region, and immediately the graphs appear. (Fig. Ex. 6.1, c)
x := 0, 0.1, ... , 3 (define a range variable x.. starting value = 0, next
value = 0.1 m, and last value = 3.0 m)
It may be observed from the above graph that:
(i) higher the thermal conductivity, higher is the steady state temperature attained at a given location.
(ii) to attain the same temperature on the rod, longer length is required for a material of higher thermal
conductivity.
(iii) it is verified that fin can be considered as infinitely long if the lengths are 2.437, 1.768 and 0.927 m for copper,
aluminium and steel, respectively, i.e. the end temperature becomes equal to the ambient temperature at these
lengths.
Example 6.2. To determine the thermal conductivity of a long, solid 2.5 cm diameter rod, one half of the rod was
inserted to a furnace while the other half was projecting into air at 27°C. After steady state had been reached, the
temperatures at two points 7.6 cm apart were measured and found to be 126°C and 91°C, respectively. The heat transfer
coefficient over the surface of the rod exposed to air was estimated to be 22.7 W/(m2K). What is the thermal conductivity
of the rod?
Solution. See Fig. Example 6.2.
Data:
D := 0.025 m h := 22.7 W/(m2 K) Ta := 27°C
Since it is a very long (or, infinitely long) rod, we can take any point as the origin. But, taking point A as origin as
shown in the Fig. 6.9 simplifies the solution. Then, we write:

236 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Temperature distribution for infinitely long fin
150
140
130
120
110
100
T(x, mcu) 90
Copper
80
T(x, mAl) Aluminium
70 Steel
T(x, ms) 60
50
40
30
20
10
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
x

FIGURE Example 6.1(c)

Furnace
2
h = 22.7 W/(m C)
Ta = 27 C

k=?
126 C 91 C

A B D = 0.025 m

0.076 m
L®¥

FIGURE Example 6.2 Very long fin of circular cross section

x := 0.076 m (distance of point B from origin (i.e. point A))


To := 126°C (temperature at the origin (point A))
T (x): = 91°C (temperature at point B.)
Temperature distribution in a long fin is given by Eq. 6.3:
T (x) - Ta
i.e. =exp (– m × x) ...(6.3)
To - Ta

F T ( x) - T I
- ln GH T - T JK
o a
a

Therefore, m := m–1 (define fin parameter m)


x
–1
i.e. m = 5.74 m
But, we have, by definition of m:
h× P
m= (Eq. (A) definition of fin parameter m)
k × Ac

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 237


In the present case,
P := p × D, m (define P, the perimeter)
i.e. P = 0.079 m (perimenter)
p ×D 2
and, Ac := , m2 (define area of cross section Ac)
4
i.e. Ac = 4.909 ´ 10–4 m2 (area of cross section Ac)

h× P
Therefore, from Eq. (A): k: = W/(mK) define thermal conductivity
Ac ×m 2
i.e. k = 110.237 W/(mK) ...value of thermal conductivity.
Example 6.3. Aluminum square fins (0.5 mm ´ 0.5 mm) of 1 cm length are provided on the surface of an electronic
semiconductor device to carry 46 mW of energy generated by the electronic device and the temperature at the surface of
the device should not exceed 80°C. The temperature of the surrounding medium is 40°C. Thermal conductivity of alu-
minium = 190 W/(mK) and heat transfer coefficient h = 12.5 W/(m2K). Find number of fins required to carry out the
above duty. Neglect the heat loss from the end of the fin. [M.U.]
Solution. This is the case of fin, insulated at its end, since by data, there is no heat loss from the end of the fin. Therefore,
Eq. 6.7 for temperature distribution and Eq. 6.8 for heat transfer rate, are applicable.
See Fig. Example 6.3.

To = 80 C
2
h = 12.5 W/(m C) k = 190 W/(mK)
Ta = 40 C
0.5 mm sq.

Insulated
L = 0.01 m

Figure Example 6.3 Finite fin insulated at its tip

Date:
Qtot := 0.046 W L: = 0.01 m w := 0.0005 m t := 0.0005 m k := 190 W/(m K) To := 80°C
Ta := 40°C h := 12.5 W/(m2 K)
Let us first calculate heat transferred from one fin; then, knowing the total amount of heat to be transferred, we can
find out the total number of fins required.
Fin parameter m:
h× P
We have m= (fin parameter)
k × Ac
Now, Ac := w × t, m2 (define area of cross section of fin)
i.e. Ac =2× 5 ´ 10–7 m2 (area of cross section of fin)
and, P := 2 × (w + t), m (define perimeter of fin sectior)
i.e. P =2 ´ 10–3 m (perimeter of fin section)
h× P
Therefore, m := m–1 (define fin parameter m)
k × Ac
i.e. m = 22.942 m–1 (fin parameter m)
Also, qo := To – Ta °C (define excess temperature at the base)
i.e. qo = 40°C (q at the base, i.e. at x = 0)

238 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Applying Eq. 6.8 for heat transfer rate from a fin with insulated end:
i.e. Qfin := k × Ac × m × qo × tan h (m× L) ...(6.8)
we get: Qfin = 9.82818 ´ 10–3 W (heat transfer per fin)
Therefore, number of fins required to carry 46 mW:
Qtot
N= = 4.68 (i.e. say, 5 fins.)
Qfin
Example 6.4. A cylinder 5 cm diameter and 50 cm long, is provided with 14 longitudinal straight fins of 1 mm thick and
2.5 mm height. Calculate the heat loss from the cylinder per second if the surface temperature of the cylinder is 200°C.
Take h = 25W/(m2 K), k = 80 W/(mK), and Ta = 45°C. [M.U.]
Solution. See Fig. Example 6.4.

t = 0.001 m

L = 0.0025 m
w = 0.5 m

To = 200 C

D = 0.05 m Total 14 no. of fins


k = 80 W/(mK)

2
h = 25 W/(m K)
Ta = 45 C

FIGURE Example 6.4 Longitudinal fins on a cylinder, losing heat from tip

This is the case of a fin with convection from its end. Therefore, Eq. 6.10 for temperature distribution and Eq. 6.11
for heat transfer rate, are applicable. However, Eq. 6.11 is a little complicated to use; so, as remarked earlier, we shall use
the Eq. 6.8 for a fin with insulated end, but with the modification that the corrected length, Lc is used instead of L. Then,
we will check the result thus obtained, by applying Eq. 6.11.
Total heat transfer is calculated as the sum of heat transferred from all the 14 fins and the convective heat transfer
from the unfinned base surface of the cylinder, which is at a temperature of 200°C.
Data:
L := 0.0025 m w := 0.5 m t := 0.001 m N := 14 D := 0.05 m k := 80 W/(m K) To := 200°C
Ta := 45°C h := 25 W/(m2K) qo := To – Ta °C i.e. qo = 155°C
Fin parameter m:
h× P
We have: m=
k × Ac
Now, Ac := w × t, m2 (define area of cross section of fin)
i.e. Ac =5 ´ 10–4 m2 (area of cross section of fin)
and, P := 2 . (w + t), m (define perimeter of fin section)
i.e. P = 1.002 m (perimeter of fin section)
h× P
Therefore, m := m–1, (define fin parameter m)
k × Ac
i.e. m = 25.025 m–1 (fin parameter m)
Corrected length, Lc:
For rectangular cross section:

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 239


t
Lc := L + ,m (define corrrected length)
2
–3
i.e. Lc =3 ´ 10 m (corrected length.)
Heat transfer from fins:
Applying Eq. 6.8 for heat transfer rate from a fin with insulated end, using Lc instead of L:
i.e. Qfin := k × Ac × m × qo × tan h (m× Lc) ...(6.8)
we get: Qfin = 11.62642 W (heat transfer per fin)
Therefore,
heat transfer from 14 fins: Qtot1 := Qfin × 14 W (define heat transfer from 14 fins)
i.e. Qtot1 = 162.77 W (heat transfer from 14 fins.)
Heat transfer from unfinned surface of cylinder:
Unfinned surface area: Aunfin := (p ×D – N× t) × w, m2 = subtracting the area occupied by 14 fins
from the surface area of the cylinder
i.e. Aunfin = 0.072 m 2 ...unfinned area
Therefore, Qtot2 := h × Aunfin × (To – Ta) W ...heat transfer from unfinned base area
i.e. Qtot2 = 277.217 W ...heat transfer from unfinned base area
Total heat transfer rate:
Qtotal := Qtot1 + Qtot2 W ...define total heat transfer
i.e. Qtotal = 439.987 W ....total heat transfer.
Verify: Let us check the result obtained by using Eq. 6.8 by comparing it with the result that would be obtained if we
use the accurate relation for heat transfer for fin with convection from its end, i.e. Eq. 6.11:

F tan h(m×L) + h I
GH m× k JK
Qfin := k × A × m × q ×
c o
F 1 + h ×tan h(m×L)I ...(6.11)
GH m×k JK
i.e. Qfin = 11.62266 W (heat transfer per fin)
This value compares very well with the result obtained from Eq. 6.8, i.e. 11.626 W.
Example 6.5. Two ends of a copper rod (k = 380 W/(mK)), 15 mm diameter and 300 mm long are connected to two
walls, each maintained at 300°C. Air is blown across the rod with a heat transfer coefficient of 20 W/(m 2K). Air
temperature is 40°C. Determine:
(i) mid-point temperature of the rod
(ii) net heat transfer to air
(iii) heat transferred from the first 0.1 m of the rod from LHS. [M.U.]
Solution. See Fig. Example 6.5.

D = 0.015 m, k = 380 W/(mK)

2
h = 20 W/(m C)
T1 = 300 C Ta = 40 C T2 = 300 C

L = 0.3 m

FIGURE Example 6.5 Fin with equal temperature at both ends

240 FUNDAMENTALS OF HEAT AND MASS TRANSFER


This is the case of a fin with specified temperatures at its both ends. So, we can directly use Eq. 6.12 for temperature
distribution and Eq. 6.13 for heat transfer rate.
However, easier method to solve, is as follows: since the temperatures at both ends are same, it is immediately
clear that the minimum temperature occurs at the mid-point. i.e. at the mid-point, dT/dx is equal to zero; but, this is also
the condition for an insulated end. Therefore, the given rod of length L may be considered as made up of two fins, each
of length L/2, insulated at its end. So, for one half of the rod, we can apply the simpler Eq. 6.7 for temperature
distribution and Eq. 6.8 for heat transfer, for a fin with insulated end.
We will verify the result later, by using Eq.6.12 and 6.13.
Data:
D := 0.015 m L := 0.3 m k := 380 W/(m K) T1 := 300°C T2 := 300°C
Ta := 40°C h := 20W/(m2K)
Fin parameter m:
h× P
We have: m=
k × Ac

p ×D 2
Now, Ac := , m2 (define area of cross section of fin)
4
i.e. Ac = 1.767 ´ 10 –4 m2 (area of cross section of fin)
and, P := p × D, m (define perimeter of fin section)
i.e. P = 0.047 m (perimeter of fin section)
h× P
Therefore, m:= m–1, (define fin parameter m)
k × Ac
i.e. m = 3.746 m–1, (fin parameter m.)
Mid-point temperature of rod:
Now, left half of the rod can be considered as a fin of length L/2, with its end insulated.
So, for temperature distribution, apply Eq. 6.7, putting L = L/2, To = T1
T ( x) - Ta cos h ( m ×( L - x))
= ...(6.7)
To - Ta cos h ( m ×L)
Putting L = L/2:

LM FG L - xIJ OP
N H 2 KQ
cos h m×
T (x) := (T1 – Ta) ×
F LI
cos h G m× J
+ Ta (Eq. A...temperature distribution in the rod)

H 2K
Therefore, mid-point temperature is obtained by putting x = L/2:
i.e. T(0.15) = 263.734°C (mid-point temperature.)
To draw the temperature profile:
We use Mathcad to draw the temperature profile. First, define a range variable x, varying from 0 to 0.3 m, with an
increment of 0.01 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis
with x and T(x), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig. Ex.
6.5(b)
x := 0, 0.01, ... , 0.3 (define a range variable x..starting value = 0,
next value = 0.01 m, and last value = 0.3 m)
It may be verified from the graph that temperature at both the ends is 300°C and the minimum. temperature occurs
at mid-point (i.e. x = 0.15 m), with Tmin = 263.73°C.
Also, note that temperature distribution as given in Eq. A plots the temperature distribution over the whole length
since beyond x = L/2 = 0.15 m, in the numerator of first term in Eq. A, the relation cos h (– x) = cos h (x) applies, and
beyond the mid-point, we get a mirror image of the graph on the left.
Heat transfer:
Heat transfer for the first half of the rod is given by Eq. 6.8. Total heat transfer from the rod is, of course, twice this value:
i.e. Qfin = k × Ac × m× qo × tan h (m × L) ...(6.8)
Note that in Eq. 6.8, we have to put L = L/2, q o = (T1 – Ta)

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 241


T(x) for fin–both ends at same temperature
300
295
290
285
T(x) x in metres and
280 (T(x) in deg.C
275
270
265
260
0 0.05 0.1 0.15 0.2 0.25 0.3
x

FIGURE Example 6.5(b)

Therefore, total heat transfer from the rod:


FG L IJ
Qtotal := 2× k × Ac × m × (T1 – Ta) × tan h m×
H 2K ...(B)

i.e. Qtotal = 66.642 W (heat transfer from the rod)


Verify: Let us verify the results now from Eqs. 6.12 and 6.13.
For temperature profile: use Eq. 6.12:
q 1 ×sin h(m×(L - x)) + q 2 ×sin h(m×x)
i.e. q (x) = ...(6.12)
sin h(m×L)
Put here: q 1 = q 2 = T1 – Ta
(T1 - Ta ) × sin h(m × (L - x)) + (T1 - Ta ) × sin h(m × x)
q(x) := ...(C)
sin h (m × L)
L
and, at mid-point, x= = 0.15 m
2
Therefore, q (0.15) = 223.734°C (excess temperature = T(x) – T(a))
i.e. T(0.15) = 223.734 + 40 = 263.734°C (temperature at mid-point...same as obtained earlier.)
or, directly from Eq. 6.14:

FG m×L IJ
i.e. qmin =
2×q 1 ×sin h
H2K ...(6.14)
sin h(m×L)

FG m×L IJ
i.e. Tmin: = Ta +
2×(T1 - Ta )×sin h
H 2K
sin h(m×L)
i.e. Tmin = 263.734 C (temperature at mid-point
...same as obtained above.)
For heat transfer: use Eq. 6.13:
F cos h(m×L) - 1 I
Qfin = k× Ac × m × (q1 + q2) × GH sin h(m×L) JK ...(6.13)

Put here: q1 = q2 = T1 – Ta = 260

F cos h(m×L) - 1I
i.e. Qfin := k× Ac × m × (260 + 260) × GH sin h(m×L) JK ...(D)

242 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Qfin = 66.642 W (heat transfer from rod...same as obtained earlier.)
Heat transfer from the first 0.1 m from LHS:
To get this, integrate the convective heat transfer from x = 0 to x =0.1 using for T(x), Eq. A.

LM FG L - xIJ OP
N H 2 KQ
cos h m×
T(x) := (T1 – Ta) ×
F LI
cos h G m× J
+ Ta (eq. (A)...temperature distribution in the rod)

H 2K

i.e.
Q := h × p × D ×

Q = 22.716 W
z
0
0.1
(T ( x) - Ta ) dx (eq. (E)...define Q, heat transfer from a length of 0.1 m from LHS)

(heat transfer from a length of 0.1 m from LHS.)


Note: In Eq. E, we have used Newton’s Law of Cooling, i.e. Q = h A DT. Elemental area involved was P. dx = p . D. dx
and DT = (T1 – Ta). Also, note that while using Mathcad, the calculation of integral within the prescribed limits is
returned directly; there is no need to do the labour of expanding the integral and substituting the limits.
Example 6.6. (a) In Example 6.6, if the two ends of the rod are maintained at 300°C and 260°C, respectively, determine:
(i) location and value of minimum temperature in the rod
(ii) mid-point temperature of the rod
(iii) draw the temperature profile
(iv) net heat transfer to air
(v) heat transferred from the first 0.1 m length of the rod from LHS
(vi) heat transferred from the left end (i.e. at x = 0)
(b) If in this example, if there is an uniform heat generation qg = 1.5 ´ 105 W/m3 in the rod, determine:
(i) location and value of minimum temperature in the rod
(ii) mid-point temperature of the rod
(iii) draw the temperature profile.
Solution. See Fig. Example 6.6.

D = 0.015 m, k = 380 W/(mK)

2
h = 20 W/(m C)
T1 = 300 C Ta = 40 C T2 = 260 C

L = 0.3 m

FIGURE Example 6.6a Fin wth different temperture at the two ends

Data:
T2 := 260°C ...temperature on RHS
Rest of the data is same as given in Example 6.6.
This is the case of a fin with specified temperatures at its both ends. So, we can use Eq. 6.12 for temperature
distribution and Eq. 6.13 for heat transfer rate.
However, let us work out this problem from fundamentals, and then verify the result from Eqs. 6.12 and 6.13.
Fin parameter m:
h× P
We have: m=
k × Ac

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 243


This is already worked out to be: m = 3.746 m–1 (fin parameter m)
Now, as shown in section 6.2.1, the controlling differential equation for this problem is:
d 2q
– m2 × q =0 ...(6.1)
dx 2
And, for its solution, let us choose Eq. 6.2b:
q (x) = A × cos h (m × x) + B × sin h(m × x) ...(6.2b)
Eq. 6.2b gives the temperature profile in the fin. Constants A and B are determined from the boundary conditions:
B.C. (i): at x = 0, q (0) = q1
B.C. (ii): at x = L, q (L) = q 2
Then, from B.C. (i) and Eq. 6.2b:
A = q 1 = T1 – Ta
i.e. A = 260
and, from B.C.(ii) and Eq. 6.2b:
q2 = (T2 – Ta) = A × cos h (m× L) + B × sin h(m × L)
(T2 - Ta ) - A×cos h(m×L)
i.e. B := (define constant B)
sin h(m×L)
i.e. B = – 161.52177 (value of constant B)
Therefore, equation for temperature distribution is:
q (x): = 260 × cos h (m × x) – 161.5218 sin h (m× x) ...(A)
Location and value of minimum temperature:
Differentiate Eq. A w.r.t. x and equate to zero; solving, we get the location xmin of minimum temperature from LHS.
Then, substitute x = xmin in Eq. A to get the value of minimum temperature
Let dq (x)/dx be defined as q ¢(x).
dq( x)
q ¢(x) = = 260 × m× sin h(m × x) – 161.5218 m × cos h (m× x) = 0
dx
161.5218
Solving, tan h (m × x) = = 0.621
260
i.e. m × x = a tan h (0.621) = 0.727 (a tan h means inverse of tan h)
0.727 0.727
i.e. xmin = = = 0.194 m (location of minimum temperature in the rod.)
m 3.746
Now, substitute this value of xmin in Eq. A to get value of Tmin:
i.e. q (0.194) = 203.742°C ...value of q (min)
i.e. Tmin := 203.742 + Ta ...since q (x) = T(x) – Ta
i.e. Tmin = 243.742°C (minimum temperature in the rod.)
Temperature at mid-point:
Put x = 0.15 m in Eq. A:
q (0.15) = 206.524 (excess temperature at mid-point)
Therefore, Tmid: = 206.524 + Ta °C (since q (x) = T (x) – Ta )
i.e. Tmid = 246.524°C (temperature at mid-point of rod.)
To draw the temperature profile:
We use Mathcad to draw the temperature profile. First, define a range variable x, varying from 0 to 0.3 m, with an
increment of 0.01 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis
with x and T(x), respectively. Click anywhere outside the graph region, and immediately the graph appears. See Fig.
below.
We have:
q (x) := 260 × cos h (m × x) – 161.5218 sin h (m × x) ...(A)
Therefore, T(x) = q (x) + Ta (define T (x), temperature at any point in the rod)
x := 0, 0.01, ..., 0.3 (define a range variable x...starting value = 0,
next value = 0.01 m, and last value = 0.3 m)
Note from the graph that the end temperatures are 300°C and 260°C, and temperature at mid-point is 246.52°C;
also, the minimum temperature occurs at x = 0.194 m, and its value is 243.74°C.
Net heat transfer from the rod:
This is obtained by integrating the convective heat transfer over the entire surface of the rod:

244 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T(x) for fin-ends at different temperature
300
290
280
270
260
T(x) 250 x in metres and
T(x) in deg.C
240
230
220
210
200
0 0.05 0.1 0.15 0.2 0.25 0.3
x

i.e.
Q :=
z
0

Q = 61.515 W
0. 3
h×P ×q(x) dx W (from Newton’s law, Q = h A DT ; A = P. dx, and q(x) = T(x) – Ta)

...heat transfer from the rod.


Heat transfer from the first 0.1 m of length of rod:

z
This is obtained by integrating the convective heat transfer from x = 0 to x = 0.1 m:
0.1
Q := h×P ×q(x) dx W (from Newton’s law, Q = h AD T ; A = P. dx, and q (x) = T(x) – Ta)
0

i.e. Q = 22.197 W (heat transfer from first 0.1 m of the rod.)


Now, verify from Eqs. 6.12 and 6.13:
We have, for temperature distribution:
q 1 ×sin h(m×(L - x)) + q 2 ×sin h(m× x)
q (x) := ...(6.12)
sin h(m×L)
q 1 := T1 – Ta i.e. q1 = 260
q 2 := T2 – Ta i.e. q2 = 220
Therefore, q (0.15) = 206.524°C (excess temperature at mid-point)
i.e. Tmid := 206.524 + Ta
i.e. Tmid = 246.524°C (temperature at mid-point...verified.)
And, for heat transfer:
F cos h(m×L) - 1I
Qfin := k× Ac × m × (q 1 + q 2) × GH sin h(m×L) JK ...(6.13)

i.e. Qfin = 61.515 W (heat transfer from the rod...verfied.)


Heat transferred from the left end (i.e. at x = 0):
Qleft is calculated by applying the Fourier’s law at x = 0.
We already have the equation for temperature distribution, i.e.
q (x) = 260 × cos h (m× x) – 161.5218 sin h (m× x) ...(A)
F dq ( x ) I
Qleft = – k × Ac × GH dx JK x=0
(applying Fourier’s law at the left end)

i.e. Qleft := – k × Ac × (0 – 161.5218 m× 1) W (since sin h (0) = 0 and cos h (0) = 1)


i.e. Qleft = 40.634 W (heat transferred from left end.)
Check: Let us check this result also by calculating heat transferred from the right end; sum of heat transferred from left
and right ends must be equal to 65.515 W, calculated earlier.
d
Let q ¢(x) := q (x) (define the first derivative of q (x) w.r.t. x)
dx
Qright := – k × Ac × q ¢(0.3) (applying Fourier’s law at x = L, i.e. x = 0.3 m)
i.e. Qright = – 20.881 W (heat transferred from the right end.)

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 245


5 3
D = 0.015 m, k = 380 W/(mK), qg = 1.5 ´ 10 W/m
2
h = 20 W/(m C)
Ta = 40 C
T1 = 300 C T2 = 260 C

dx
X

L = 0.3 m

FIGURE Example 6.6(b) Fin with heat generation, two ends at prescribed temperature

Remark: Negative sign indicates that heat transfer is from right to left, i.e. in negative X-direction. Also, note that now,
we have performed the differentiation and put x = 0.3 m, directly in Mathcad, instead of doing it by long hand.
Adding, |Qleft| + |Qright| = 61.515 W (total heat transferred…checks with earlier result.)
(b) If there is uniform heat generation, qg (W/m3) in the rod:
See Fig. Example 6.6(b).
Let us derive the governing differential equation by the usual method of making an energy balance on a differential
element of the rod of length dx at a distance x from the origin, as shown in the Fig. 6.6(b).
We write:
Energy into the element from left face + heat generated in the element =
Energy out at the right face + Energy lost by convection from the surface of the element
i.e.
dT F - k ×A dT d 2T I
– k× Ac ×
dx
+ qg × Ac × dx = GH c
dx dx JK
- k× Ac × 2 × dx + h × (P × dx) × (T – Ta)

d 2T
i.e. – k× Ac × × dx + h × (P × dx) × (T – Ta) – qg × Ac × dx = 0
d x2

d 2T
i.e. k× Ac × – h × P × (T – Ta) + qg × Ac = 0
d x2

d 2T qg
i.e. – m 2 × (T – Ta) + = 0 ...(a)
d x2 k

h× P
where, m=
k × Ac
Substitutingq = T – Ta we get:
d 2q qg
– m2 × q + = 0 ...(b)
d x2 k
qg
Now, make another subsitution: q ¢ = q –
k ×m 2
Then, Eq. b becomes:
d 2q ¢
– m 2× q ¢ = 0 ...(c)
dx 2
General solution of Eq. c is:
q ¢(x) = A × cos h (m × x) + B × sin h (m × x) ...(d)
A and B constants, determined from Boundary Conditions, i.e.

246 FUNDAMENTALS OF HEAT AND MASS TRANSFER


B.C. (i): at x = 0, q¢(0) = q¢1
B.C. (i) and Eq. d gives: q¢1 = A
B.C. (ii): at x = L, q¢(L) = q¢2
B.C. (ii) and Eq. d gives:
q ¢2 = q ¢1 × cos h (m × L) + B × sin h (m × L)
q ¢2 - q 1¢ ×cos h(m×L)
i.e. B=
sin h(m×L)
Substituting for A and B in Eq. d:
q ¢2 - q 1¢ ×cos h(m×L)
q ¢(x) = q ¢1 cos h× (m× x) + × sin h (m × x)
sin h(m×L)
q 1¢ ×(cos h(m×x )×sin h(m×L) - sin h(m× x)×cos h(m×L)) + q 2¢ ×sin h(m× x)
i.e. q ¢(x) =
sin h(m×L)
q 1¢ ×(sin h(m×(L - x))) + q ¢2 ×sin h(m×x)
i.e. q ¢(x) = (using eqn. (p) from Table 6.1...(e))
sin h(m×L)
qg
Remembering that: q ¢ = q – and, q = T – Ta
k ×m 2

qg q ¢1 ×(sin h(m×(L - x))) + q ¢2 ×sin h(m×x)


we get: T(x) = Ta + + ...(f)
k ×m 2 sin h(m×L)
Eq. f gives the desired temperature profile in the rod when there is uniform heat generation in the rod and the ends
are maintained at prescribed temperatures.
Now, for the present case:
qg = 1.5 ´ 105 W/m3 (heat generation rate)
qg
q ¢1 = T1 – Ta – i.e. q ¢1 = 231.875
k ×m 2
qg
q¢2 = T2 – Ta – i.e. q ¢2 = 191.875
k ×m 2
Therefore, from Eq. f:
qg q ¢1 ×(sin h(m×(L - x))) + q ¢2 ×sin h(m×x)
Temp (x) := Ta + + ...(f)...define Temp (x)
k ×m 2 sin h(m×L)
Minimum temperature in the rod:
Differentiate Eq. f w.r.t. x and equate to zero: solving, get the location of minimum temperature, xmin. Substitute this
value of x back in Eq. f to get Tmin:
d
Let Temp¢(x): = Temp (x) (define Temp’(x) as the first derivative of Temp(x) w.r.t. x)
dx
Use solve block to solve Temp¢(x) = 0. Start with a trial value of x, write the constraint immediately below ‘Given’;
then the typing Find (x) immediately gives the value of xmin.
x := 0.1 m (trial value of x)
Given
Temp¢(x) = 0
Find(x) = 0.2
i.e. xmin := 0.2 m (location of minimum temperature in the rod.)
Substitute this xmin back in Eq. f to get Tmin:
Temp(xmin) = 247.289°C (minimum temperature in the rod, with heat generation.)
Temperature at mid-point:
At mid-point, x := 0.15:
Temp(0.15) = 250.447°C (temperature at mid-point, when there is heat generation in the rod.)

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 247


To draw temperature profile in the rod:
We use Mathcad to draw the temperature profile. First, define a range variable x, varying from 0 to 0.3 m, with an
increment of 0.01 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis with x
and on y-axis, with T(x), Temp(x), respectively. Click any where outside the graph region, and immediately the graph
appears. It contains two curves, one for x vs. T(x) and the other x vs. Temp(x). See fig. below.
We have:
q (x) := 260 × cos h (m × x) – 161.5218 sin h (m × x). ...(A)
Therefore, T(x) := q (x) + Ta (define T(x), tempreature at any point in the rod, without heat generation.)

qg q ¢1 ×(sin h(m×(L - x))) + q ¢2 ×sin h(m×x)


Temp(x) := Ta + + (define Temp(x) with heat generation.)
k ×m 2 sin h(m×L)

x := 0, 0.01, ... , 0.3 (define a range variable x...starting value = 0,


next value = 0.01 m, and last value = 0.3 m)

Temperature distribution for fin,


ends at T1 and T2
300
290
280
270
260
T(x) x in metres and
250 Temperature in deg.C
Temp (x)
240
230
220
210
200
0 0.05 0.1 0.15 0.2 0.25 0.3
x
without heat generation
with heat generation

Note: In the above graph, temperature distribution in the rod is drawn for both the cases i.e. with and without heat
generation, for comparison. Temperature in the rod is everywhere higher with heat generation, as would be expected.
With heat generation, minimum temperature occurs at 0.2 m from LHS, whereas without heat generation, minimum
temperature occurs at 0.194 m from LHS. It can also be seen that left end is at 300°C and the right end at 260°C, as
specified.

6.3 Fins of Non-uniform Cross Section


So far, we considered fins of uniform cross section. But, very often in practice, we find that fins of nonuniform
cross section are also used. See Fig. 6.8, (b), (c), and (e). For example, annular fins are provided over a circular
tube, as shown in Fig. 6.8(c), to enhance heat transfer. Here, the fin thickness may be constant, but its area for
heat transfer along its radius, i.e. (2p r t ), varies with the radius.
In such cases, the general differential equation governing the temperature distribution is derived by making
an energy balance across an elemental volume, just as we did in the case of fin of uniform cross section.
Consider a fin of nonuniform cross section as shown in Fig. 6.9.
Consider an elemental section of thickness dx at a distance x from the base as shown. Let us write an energy
balance for this element:
Energy going into the element by conduction = (Energy leaving the element by conduction + Energy leaving
the surface of the element by convection)
i.e. Qx = Qx + dx + Qconv ...(a)

248 FUNDAMENTALS OF HEAT AND MASS TRANSFER


t
t

t w

w r1 L
L
r2
L

(a) Straight rectangular fin (b) Straight triangular fin (c) Circular fin of rectangular section

D D

L
L

(d) Pin fin, circular section (e) Pin fin, conical section

FIGURE 6.8 Typical Fins: (a) and (d) of uniform crosssection, and (b), (c) and (e): of non-uniform crosssection

where, dQconv
Qx = heat conducted into the element at x
Qx + dx = heat conducted out of the element at x + dx, and h, Ta dAs
Qconv = heat convected from the surface of the element
To
to ambient.
We have, from Fourier’s law:
dT Qx Qx+dx
Qx = – k × Ac ×
dx
Note that here, Ac, the cross-sectional area varies with x.
dx
And, X

dQx
Qx + dx = Qx + × dx X
dx
dT d dT FG IJ FIGURE 6.9 Fin of non-uniform cross-section
i.e. Qx + dx = – k × Ac ×
dx
– k×
dx
Ac ×
dx H
× dx
K
Convection heat transfer rate from the elemental volume is given by:
Qconv = h × dAs × (T – Ta)
where, dAs is the surface area of the elemental volume.
Substituting the terms in Eq. a,
d FG
dT IJ – h × dAs × (T – T ) = 0
dx
× Ac ×
H
dx K k dx a

d 2T F 1 × dA I × dT – F 1 × h × dA I × (T – T ) = 0.
+G
H A dx JK dx GH A k dx JK
c s
i.e. a ...(6.15)
dx 2 c c

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 249


Equation 6.15 is the controlling differential equation for one-dimensional conduction in a fin of non-uniform
cross section. Remember again that both Ac and As vary with x.
Solution of Eq. 6.15 gives the temperature distribution in the fin, and then, by applying Fourier’s law, we can
get the heat flux at any point.
However, solution of Eq. 6.15 is rather complicated and involves Bessel Functions. So, in practice, heat
transfer in fins of non uniform cross sections are calculated by resorting to ‘Fin efficiency’ graphs, as explained in
the next section.

6.4 Performance of Fins


Recollect that purpose of attaching fins over a surface is to increase the heat transfer rate. How well this purpose
is achieved is characterised by two performance parameters:
(i) Fin efficiency, h f and
(ii) Fin effectiveness, e f .
6.4.1 Fin Efficiency
Fin transfers heat to the surroundings from its surface, by convection. For convection heat transfer, the driving
force is the temperature difference between the surface and the surrounding. However, temperature drops along
the length of the fin because of the finite thermal conductivity of the fin material; so, heat transfer becomes less
effective towards the end of the fin. Obviously, in the ideal case of the entire fin being at the same temperature as
that of the base wall, the heat transferred from the fin will be maximum. So, fin efficiency is defined as the
amount of heat actually transferred by a given fin to the ideal amount of heat that would be transferred if the
entire fin were at its base temperature, i.e.
Qfin
hf = ...(6.16)
Qmax
where,
Qfin = actual amount of heat transferred from the fin, and
Qmax = maximum (or ideal) amount of heat that would be transferred from the fin, if the entire fin
surface were at the temperature of the base.
(a) For an infinitely long fin:
For an infinitely long fin, actual heat transferred is given by Eq. 6.5:
i.e. Qfin = h×P ×k × Ac ×q o = h ×P ×k × Ac × (To – Ta) ...(6.5)
To calculate Qmax, if the entire fin surface were to be at a temperature of To, the convective heat transfer from
the surface would be:
Qmax = h × P × L × (To – Ta) ...(A)
where, P is the perimeter of the fin section and (P.L) is the surface area of the fin.
Dividing Eq. 6.5 by Eq. A:

h× P×k× Ac ×(To - Ta )
hf =
h×P×L×(To - Ta )

1
i.e. hf =
h×P
×L
k × Ac

1
i.e. hf = ((6.17)...fin efficiency for very long fin.)
m× L
(b) For a fin with insulated end:
For the case of a fin with an insulated end, we get actual heat transferred Qfin from Eq. 6.7:
i.e. Qfin = h ×P ×k × Ac × (To – Ta) × tan h (m × L) ...(6.7)
and, fin efficiency is given by:

250 FUNDAMENTALS OF HEAT AND MASS TRANSFER


h× P×k× Ac ×(To - Ta )×tan h( m×L)
hf =
h×P×L×(To - Ta )

tan h( m×L)
i.e. hf =
h× P
×L
k × Ac

tan h ( m×L)
i.e. hf = ((6.18)...fin efficiency for a fin with insulated end)
m× L
Note: For the more realistic case of a fin losing heat from its end, as stated earlier, to calculate heat transfer,
Eq.6.9 itself may be used, but , with a corrected length Lc in place of L.
It is instructive to represent Eq. 6.18 in graphical form:
Let X = m×L
X := 0.1, 0.2.. 5 (let (m × L) vary from (m × L) = 0.1 to 5 with an increment of 0.1)
The graph looks as follows:

Fin efficiency graph–insulated end


1
0.9
0.8
0.7
tan h(X) 0.6
X 0.5 Note: X = m.L
0.4
0.3
0.2
0.1
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
X
It may be noted from the graph that:
(i) Fin efficiency is maximum for the trivial case of L = 0, i.e. when there is no fin. So, fin efficiency is not
maximised w.r.t. the fin length, but generally, w.r.t. volume or weight of material, which also has cost
implications.
(ii) Nearer to the base of the fin, fin efficiency is high and it goes on decreasing as we move towards the end
of the fin; this is because, the surface temperature of the fin falls as we move away from the base towards
the end. Again , it is clear that there is not much gain if (m × L) is increased beyond a value of about 3.
Table 6.4 gives the values of fin efficiency for a few fin shapes:
Note: In Table 6.4:
I0 = modified zero order Bessel function of first kind
K0 = modified zero order Bessel function of second kind
I1 = modified first order Bessel function of first kind
K1 = modified first order Bessel function of second kind.
Fin efficiency graphs:
As can be seen from Table 6.4, expressions for fin efficiency of fins of non-uniform cross sections are rather
complicated and involve Bessel functions. In practice, to find out the heat transfer from such fins, we use ‘Fin
efficiency charts’. Once the fin efficiency is obtained from these graphs, actual heat transferred from the fin is
calculated using the definition of fin efficiency, i.e.
Qfin
hf = ...(6.16)
Qmax
where,
Qmax = h × P ×L × (To – Ta)

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 251


TABLE 6.4 Fin efficiency (h f ) for a few fin shapes
Ac = area of cross section, Af = total fin surface area, Lc = corrected length, P = perimeter of fin section, h = heat
transfer coefficient, m = {h ×P /(k ×Ac )}

Sl.No. Description Parameters Fin efficiency (nf)


Straight fin of
rectangular section. tan h (m ×Lc )
1 Af = 2 × w× Lc hf =
See Fig. 6.8(a) m ×Lc

t
Lc = L +
2

2×h
m= ...thin fins, w >
k ×t

L FtI O
1
Straight fin of triangular 2

A = 2 × w × ML + G J P
2
section. 1 I1(2 × m × L )
MN H 2 K PQ
2
2 f hf = ×
See Fig. 6.8(b) m ×L I o (2×m ×L )

2×h
m=
k ×t

3
Circular fin of
rectangular section d
Af = 2 × p × r22c - r12 i h f = C2 ×
LM (K (m ×r )×I (m ×r
1 1 1 2c ) - I1(m ×r1)×K 1(m ×r2c )) OP
See Fig. 6.8(c) N (I (m×r )×K (m ×r
o 1 1 2c ) + K o (m ×r1)×I1(m ×r2c )) Q
FG 2×r IJ
HmK
1
t
r2c = r2 + C2 =
2 dr - r i
2
2c 1
2

2×h
m=
k ×t

Pin fin, circular


section tan h (m ×Lc )
4 A f = p × D × Lc hf =
See Fig. 6.8(d) m ×Lc

D
Lc = L +
4

4×h
m=
k ×D

LM FG IJ OP
1
Pin fin, conical 2
section p ×D 2 D 2
2 I 2 (2×m ×L )
5
See Fig. 6.8(e)
Af =
2
×L +
2 MN H K PQ hf = ×
m ×L I1(2×m ×L)

4×h
m=
k ×D

252 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Fin efficiency for rectangular & triangular fins
1
0.95
0.9
0.85
0.8
0.75

Fin efficiency
0.7
0.65
0.6
0.55
0.5
0.45
0.4
0.35
0.3
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
rectangular fins mLc or mL
triangular fins

100 t rectangular fin


L+
Lc = 2
L triangular fin
80
Fin efficiency hf per cent

t
tL rectangular fin
L Am = t
L triangular fin
60 2

40
t

20 L

0
0.5 1.0 1.5 2.0 2.5

Lc 3 / 2 (h k . Am )1/ 2

FIGURE 6.10 Efficiency of straight, rectangular (or, cylindrical, pin fins) and triangular fins (Ref. Fig.6.8, a & b)

Qmax is easily calculated from the given data.


Fig. (6.10) gives fin efficiency values for fins of rectangular and triangular sections.
It may be noted that graph for rectangular fins is also valid for cylindrical, pin fins since equation for fin
efficiency is the same (see Table 6.4); of course, m and Lc will be different for pin fins.
For straight, rectangular fins and For straight, triangular fins:
cylindrical, pin fins:
tan h (m×Lc ) 1 I1 ( 2×m×L)
hf = hf = ×
m ×Lc m× L I o ( 2 × m× L )
Let X = m × Lc Let Z = m ×L
tan h(X ) 1 I 1 ( 2× Z )
i.e. hf = i.e. hf = ×
X Z I o ( 2× Z )

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 253


TABLE 6.5 Efficiency vs. mLc for straight, rectangu- TABLE 6.6 Efficiency vs. mL for straight, triangular
lar fins, or cylindrical pin fins X = m×L c fins Z = m× L
tan h ( X ) 1 I1(2 × Z )
Effcy(X) = X = 0.1, 0.2, ... , 3 Effcy (Z) = × Z = 0.1, 0.2, ... , 3
X Z I 0 (2 × Z )
tan h (X ) 1 I1(2 × Z )
X Z ×
X Z I 0 (2 × Z )
0.1 0.997 0.1 0.995
0.2 0.987 0.2 0.981
0.3 0.971 0.3 0.958
0.4 0.95 0.4 0.928
0.5 0.924 0.5 0.893
0.6 0.895 0.6 0.855
0.7 0.863 0.7 0.815
0.8 0.83 0.8 0.775
0.9 0.796 0.9 0.736
1 0.762 1 0.698
1.1 0.728 1.1 0.662
1.2 0.695 1.2 0.628
1.3 0.663 1.3 0.597
1.4 0.632 1.4 0.567
1.5 0.603 1.5 0.54
1.6 0.576 1.6 0.515
1.7 0.55 1.7 0.492
1.8 0.526 1.8 0.47
1.9 0.503 1.9 0.45
2 0.482 2 0.432
2.1 0.462 2.1 0.415
2.2 0.444 2.2 0.399
2.3 0.426 2.3 0.384
2.4 0.41 2.4 0.37
2.5 0.395 2.5 0.357
2.6 0.38 2.6 0.345
2.7 0.367 2.7 0.334
2.8 0.355 2.8 0.323
2.9 0.343 2.9 0.313
3 0.332 3 0.304

Note: On x-axis use: (m× Lc) for rectangular fins and, (m× L) for triangular fins
Since these two types of fins are used very much in practice, efficiency values are also given in tabular form
above:
Fig. 6.11 gives fin efficiency values for circumferential fins of rectangular profile.
Note that X-axis in Fig. 6.10 is:
3
h
Lc2 ×
k × Am
where, Am is the profile area of the fin.
(Am = L.t for rectangular section and (L.t/2) for a triangular section).
Rationale of using this complicated expression in the X-axis is as follows:

254 FUNDAMENTALS OF HEAT AND MASS TRANSFER


100
Straight fin
90

80 t

Fin efficiency hf, %


L

70
t

60 ro
ri
1.2
50 5
2.0 1.5
0 0
3.0
40 FG ro + t IJ 0
H 2K ri =
30
0 0.2 0.4 0.6 0.8 1.0 1.4 1.6 1.8 2.0 2.2 2.4

FG r t IJ 3/2
FG t IJ 3/2

H o +
2
- ri
K 2h / k t (ro - ri ) or L +
H 2 K 2h / k t L

FIGURE 6.11 Efficiency of circumferential rectangular fins

Considering a fin of rectangular cross section, insulated at its end, we can write:
tan h( m×L)
hf =
m× L

h×P h×( 2×w + 2×t )


Now, m×L = ×L = ×L
k× Ac k×w×t
For a very wide fin: i.e. w >> t, we can write:
3
2×h×w 2×h 2×h
m×L = ×L = ×L = × L2
k ×w×t k×t k ×t×L
3
2×h
i.e. m×L = × L2 ...(6.19)
k × Am
where, Am = (L.t), is the profile area for the rectangular section. So, on the X-axis, instead of (m.L), what is plotted
is:
3
h
Lc2 ×
k × Am
where, Lc is the corrected length, to take into account convection from the end.
6.4.2 Fin Effectiveness (ef )
Consider a fin of uniform cross-sectional area Ac, fixed to a base surface. Purpose of the fin is to enhance the heat
transfer. If the fin were not there, heat would have been transferred from the base area Ac, by convection. By
attaching the fin, area for convection increases i.e. convective resistance ( = 1/(h.A)) decreases; however, it is
obvious that a conduction resistance due to the solid fin is now introduced and the total heat transfer would
depend upon the net thermal resistance. As we go on increasing the length of fin, convection resistance will go
on decreasing, but conduction resistance will go on increasing. This means that attaching a fin may not
necessarily result in effectively increasing the heat transfer. Therefore, how effective the fin is in enhancing the
heat transfer is characterised by a parameter called fin effectiveness.

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 255


Fin effectiveness is defined as the ratio of the heat transfer rate with the fin in place, to the heat transfer that
would occur if the fin were not there, from the area of the base surface where the fin was originally fixed.
i.e. e f = (heat transfer rate with fin)/(heat transfer rate without fin)
Qfin
i.e. ef = ...(6.20)
h× Ac ×(To - Ta )
Fin effectiveness equal to 1 means that there is no enhancement of heat transfer at all by using the fin; if the
fin effectiveness is less than 1, that means that the fin actually reduces the heat transfer by adding additional
thermal resistance! Obviously, ef should be as large as possible. Use of fins is hardly justified unless fin
effectiveness is greater than about 2, i.e. e f ³ 2.
To get an insight into the physical implications of fin effectiveness, let us consider an infinitely long fin:
Then, we have:

h× P×k× Ac ×(To - Ta )
ef = (fin effectiveness for very long fin)
h× Ac ×(To - Ta)

k ×P
i.e. ef = ...(6.21)
h×Ac
Eq. 6.21 is an important equation. Following significant conclusions may be derived from this equation:
(i) Thermal conductivity, k should be as high as possible; that is why we see that generally, fins are made up
of copper or aluminium. Of course, aluminium is the preferred material from cost and weight
considerations.
(ii) Large ratio of perimeter to area of cross section is desirable; that means, thin, closely spaced fins are
preferable. However, fins should not be too close as to impede the flow of fluid by convection.
(iii) Fins are justified when heat transfer coefficient h is small, i.e. generally on the gas side of a heat
exchanger rather than on the liquid side. For example, the car radiator has fins on the outside of the tubes
across which air flows.
(iv) Requirement that ef ³ 2, gives us the criterion:
k ×P
>4 ...(6.22)
h×Ac
These two important parameters, namely, h f and e f are related to each other as follows:
Qfin Qfin h f ×h×A f ×(To - Ta )
ef = = =
Qbase h× Ac ×(To - Ta ) h× Ac ×(To - Ta )
Af
i.e. ef = × hf ...(6.23)
Ac

6.4.3 Thermal Resistance of a Fin


Consider a fin of cross-sectional area Ac fixed on a base surface. Then, the convective thermal resistance of the
base area is:
1
Rb = ((6.24a)...convective thermal resistance of base area)
h×Ac
When fin is attached, we compute a thermal resistance for the fin, from the usual definition, i.e.
DT T - Ta
Rfin = = o ((6.24b)...thermal resistance of fin)
Qfin Qfin
Values of Qfin depend on the conditions at the tip of the fin and may be obtained from Table 6.3. Dividing
equation 6.24a by Eq. 6.24b, we get:
Rb Qfin
= = ef ((6.25)...from the definition of e f in Eqn. 6.20)
Rfin h× Ac ×(To - Ta )

256 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Fin effectiveness may be considered as a ratio of thermal resistances and clearly, to achieve higher fin effec-
tiveness, the conductive resistance of the fin must be smaller than the convective resistance, calculated with
reference to the base cross-sectional area.
Concept of fin resistance is very useful to represent a finned surface in a thermal circuit, remembering that
the conductive resistance along the fin and the convective resistance from the surface of the fin are in parallel.
6.4.4 Total Surface Efficiency (or, overall surface efficiency, or area-weighted fin
efficiency), h t
What we have analysed so far, is a single fin. However, in practice, a single fin is seldom used; it is always an
array of fins fixed on a base surface.
In a heat exchanger, where use of fins is most prevalent, fins serve the purpose of increasing the amount of
heat transferred.
Total heat exchange area (At) may be considered as made up of two areas:
(i) the base or prime surface area, Ap, on which there are no fins, and
(ii) the total fin surface area (N . Af)
where, N is the total number of fins, and Af is the surface area of each fin.
Now, the prime surface (or, un-finned surface) is 100% effective; but, all the fin surface area provided is not
100% effective, since there is always a temperature gradient along the fin. From the definition of fin efficiency, we
know that effective area of the fin surface is h f . Af .
Therefore, considering the total area of the array, i.e. (Ap + N. A f), we can define an total or overall surface
efficiency, ht , such that:
h t × At = 1 × Ap + h f × N × Af
But A t = Ap + N × Af
Therefore, h t × At = (At – N × Af) + h f ×N× Af
N ×A f
i.e. ht = 1 – × (1 – h f) ...(6.26)
At
Eq. 6.26 gives the value of overall or total surface efficiency (or, area –weighted fin efficiency) for a fin array.
In other words, effective heat transfer area of the array is = (ht At), where At is the total area of the prime surface
plus all the fin area.
Concept of overall surface efficiency is very useful in calculating the heat transfer rates in heat exchangers
where fins may be provided on one or both sides of the wall. In such a case, overall heat transfer coefficient may
be obtained from:
1
Uo × Ao = Ui × Ai =
SR
1
Uo × Ao = Ui × Ai = ...(6.27)
ho ×h to × Ao + Rwall + hi ×h ti ×hi
where, Uo = overall heat transfer coefficient based on total outer surface area
Ui = overall heat transfer coefficient based on total inner surface area
Ao = total outer surface area
A i = total inner surface area
h to = total surface efficiency for outer surface
hti = total surface efficiency for inner surface
ho = average heat transfer coefficient on the outer surface
hi = average heat transfer coefficient on the inner surface
For the popular case of a tubular heat exchanger, with fins on the outside and no fins on the inside, we have:

FG r IJ
o

h ti = 1 Ai = 2× p×r i×L and, Rwall =


ln
Hr K
i
2×p ×k ×L
where, ri and ro are inside and outside radii of the tube, k is the thermal conductivity of tube material and L is the
tube length.
HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 257
Example 6.7. A steel rod (k = 30 W/(mC)), 10 mm in diameter and 50 mm long, with an insulated end is to be used as a
spine. It is exposed to surroundings with a temperature of 65°C and a heat transfer coefficient of 50 W/(m2C). The
temperature of the base is 98°C. Determine:
(i) fin efficiency (ii) temperature at the end of spine, and (iii) heat dissipation. [M.U.]
Solution. See Fig. Example 6.7.

2
h = 50 W/(m C)

To = 98°C k = 30 W/(mC)
Ta = 65°C

Q D = 0.01 m

(dT/dx)x=L = 0

L = 0.05 m

FIGURE Example 6.7 Fin of finite length, end insulated

Data:
D := 0.01 m L := 0.05 m k := 30 W/(mC) To := 98°C Ta := 65°C h := 50 W/(m2C)
Fin efficiency:
Fin efficiency for a fin with insulated end is given by Eq. 6.18:
tan h (m×L)
i.e. hf = ((6.18)...fin efficiency for a fin with insulated end)
m×L
First, let us calculate the parameter m:
h× P
We have: m= where, P is the perimeter and Ac is the area of cross section.
k × Ac

p ×D 2
Then, Ac: = m2 (define the area of cross section of the rod)
4
i.e. Ac = 7.854 ´ 10–5 m2 (area of cross section of the rod)
and, P: = p ×D m (define the perimeter of the rod)
i.e. P = 0.031 m (perimeter of the rod)

h× P
Therefore, m: = m–1 (define the parameter m.)
k × Ac
i.e. m = 25.82 m–1 (parameter m.)
Therefore, from Eq. 6.18:
tan h(m×L)
hf: =
m×L
i.e. h f = 0.666 = 66.6% (fin efficiency.)
Temperature at the end of the spine, i.e. at x = L:
We use Eq. 6.7 for the temperature distribution in a fin with insulated tip:
T( x) - Ta cos h(m×(L - x))
i.e. = ...(6.7)
To - Ta cos h(m×L)

258 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Putting x = L in Eq. 6.7, we get:
TL - Ta 1
=
To - Ta cos h( m×L)

To - Ta
i.e. TL := + Ta
cos h(m×L)
i.e. TL = 81.874°C (temperature at the end of the spine.)
Heat dissipation from the spine:
We use Eq. 6.8 for heat dissipation from a fin with insulated end:
i.e. Qfin = h× P ×k × Ac × q o ×tan h (m × L) ...(6.9)
Here, q o := To – Ta °C (excess temperature at the base of fin)
i.e. q o = 33°C (excess temperature at the base of fin)
And, Qfin := h× P ×k × Ac × q o ×tan h (m × L) W (define heat transfer from the fin)
i.e. Qfin = 1.725 W (heat dissipated from the spine.
Example 6.8. Circular aluminium fins of constant rectangular profile are attached to a tube of outside diameter D = 5
cm. The fins have thickness t = 2 mm, height L = 15 mm, thermal conductivity k = 200 W/(mC), and spacing 8 mm (i.e.
125 fins per metre length of tube). The tube surface is maintained at a uniform temperature To = 180°C, and the fins
dissipate heat by convection into the ambient air at Ta = 25°C, with a heat transfer coefficient ha = 50 W/(m2C).
Determine the net heat transfer per metre length of tube.
Solution. See Fig. Example 6.8.

Tube

To = 180°C r1 = 0.025 m

Fins, 125 nos./m, k = 200 W/(mC) r2 = 0.04 m

2 L = 0.015 m
t = 0.002 m ha = 50 W/(m C)
Ta = 25°C

L
r1

r2

FIGURE Example 6.8 Circular fin of rectangular section

D := 0.05 m L := 0.015 m r1 := 0.025 m r2 := 0.040 m t := 0.002 m k := 200 W/(mC)


To := 180 C Ta := 25 C ha := 50 W/(m2C) N := 125
This is the case of heat transfer in a fin array. So, we will use ‘total surface efficiency’ concept. First, let us find out
the fin parameter m: (See Table 6.4)
ha × P
We have: m= where, P is the perimeter and Ac is the area of cross section.
k × Ac

P 2
But, = for thin fins
Ac t

2×ha
Therefore, m := m–1 (define parameter m)
k ×t
i.e. m = 15.811 m–1 (parameter m.)

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 259


Fin efficiency:
Fin efficiency for circular fins is obtained from graphs in Fig. 6.11. To use those graphs, we need to calculate the follow-
ing:
t
r2c := r2 + m (define corrected radius)
2
i.e. r2c = 0.041 m (corrected radius)
t
Lc := L + m (define corrected length)
2
i.e. Lc = 0.016 m (corrected length for fin)

2 ha
and, = 129.1
k ×t ×(r2 - r1 )

2 ha
Therefore, L3c/2 × = 0.261 (factor to be used on X-axis of Fig. 6.11)
k ×t×(r2 - r1 )

r2 c
and, = 1.64 (factor for use in Fig. 6.11)
r1
Now, with the value of 0.261 enter the X-axis of Fig. 6.11. See where the ordinate cuts the curve for r2c/r1 = 1.64.
Move to the left and read on the Y-axis the value of h f .
From the Fig. 6.11 we read: hf = 0.97 = 97% ( fin efficiency.)
Alternatively:
From Table 6.4, we have:
For circular fins of rectangular section:

FG 2×r IJ
H m K × LM (k (m×r )×I (m×r ) - I (m×r )×K (m×r )) OP
1

1 1 1 2c 1 1 1 2c
h f (m, r1, r2c) = ...define hf as a function of m, r1 and r2c
dr - r i MN (I (m×r )×K (m×r ) + K (m×r )× I (m×r )) PQ
2
2c 1
2
0 1 1 2c 0 1 1 2c

In the present case, m := 15.811 r1 := 0.025 r2c := 0.041


therefore, h f (m, r1, r2c) = 0.973 (fin efficiency...almost same as obtained from the graph)
(Note the ease with which Mathcad calculates the Bessel functions in the above equation.)
Total surface efficiency:
This is given by Eq. 6.26:
N ×Af
i.e. ht = 1 – × (1 – h f ) ...(6.26)
At
where h f = 0.973 as already calculated.
Surface area of each fin:

d
Af := 2× p × r22c - r12 m2 i (factor 2 is used to consider both upper and lower areas of the fin)
–3 2
i.e. A f = 6.635 ´ 10 m (surface area of each fin)
Prime (or base) surface area: (This is unfinned area)
Ap := 2× p × r1 × (1 – N × t) m2 (prime surface area for 1 m length of tube)
i.e. Ap = 0.118 m2 (prime surface area per metre length of tube)
Therefore, total area:
At := Ap + N×Af m2 (prime area plus total fin area)
i.e. At = 0.947 m2 (total area)
Applying Eq. 6.26:
N ×Af
h t := 1 – × (1 – h f ) (define total surface efficiency)
At
i.e. h t = 0.976 (total surface efficiency)
Heat transfer rate for the fin array:
Therefore, heat transfer rate is given by:

260 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Q := ha × h t × At × (To – Ta) W/m (define total heat transfer rate in the fin array)
i.e. Q =7.167 ´ 103 W/m = 7.167 kW/m (heat transfer rate.)
Alternatively:
You need not memorize the Eq. 6.26 for the total surface efficiency. Just remember that prime surface is 100% effective,
whereas out of the total fin area of (N.Af), only (hf . N. Af ) is effective. So, the total heat transfer from the fin array can be
written as:
Q := (Ap + h f × N× A f ) × h a × ( To – Ta) W (define heat transfer rate from the array)
i.e. Q = 7.167 ´ 103 W/m (same as obtained above.)
Heat transfer rate if there are no fins:
It is interesting to compare the heat transfer rate obtained above, with the heat transfer rate that would be obtained if
there were no fins:
If there are no fins, heat transfer will be by convection from the surface of the bare tube. Applying Newton’s Law of
Cooling, we get:
Qtube := ha × (2 × p × r1 × 1) × (To – Ta) W (define heat transfer by convection from tube surface)
i.e. Qtube = 1.217 ´ 103 W (heat transfer rate from the bare tube)
Q
We get: = 5.887
Qtube
i.e. heat transfer increases by nearly 6 times because of addition of fins.

6.5 Application of Fin Theory for Error Estimation in


Temperature Measurement
Temperature of a fluid flowing in a pipe is generally measured with a thermometer placed in a thermowell
welded radially or obliquely to the pipe wall. Thermowell is a thin tube, generally of a material of low thermal
conductivity, such as stainless steel, filled with oil, for better thermal contact with the thermometer bulb. See
Fig. 6.12.

Thermometer

To
Thermowell, thickness = d

Fluid, L
Ta, ha TL
d

FIGURE 6.12 Error estimation in temperature measurement

Let L = length of thermowell


d = diameter of thermowell
d = thickness of thermowell wall
To = temperature at the root of thermowell, i.e. on the pipe surface
Ta = temperature of the fluid flowing, and
T L = temperature measured by the thermometer
ha = heat transfer coefficient between the thermowell and the fluid.
If the temperature of a hot fluid flowing in the pipe is Ta, obviously, the temperature indicated by the ther-
mometer, TL will not be equal to Ta, but less than Ta, because of heat loss along the wall of the thermowell from
its tip to the root (and, vice versa, for a cold fluid flowing in the pipe).
Our aim is to estimate the error in the thermometer reading.

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 261


We apply the fin theory. Considering the thermowell to be a fin protruding from the pipe wall, with an
insulated tip (i.e. no heat transfer from its tip), we can write, from Eq. 6.7:
T ( x ) - Ta cos h(m×(L - x ))
= ...(6.7)
To - Ta cos h(m×L)
At the tip, i.e. at x = L, T (x) = TL.
Substituting x = L in Eq. 6.7:
TL - Ta 1
= ...(a)
To - Ta cos h(m ×L)
And, the error in thermometer reading is given by:
To - Ta
TL – Ta = ...(b)
cos h(m ×L)
From Eq. b, we observe that to reduce the temperature error, we should have the factor 1/cos h(m. L) as
small as possible. To achieve this, looking at the graph of 1/cos h(m.L) vs. (m . L) given in section 6.2.2, it is clear
that

h×P
×L
k ×Ac
must be as large as possible.
This leads to the following conclusions:
(i) value of heat transfer coefficient, h should be large
(ii) value of thermal conductivity, k should be small
(iii) thermowell should be long and thin-walled. (thermowell may be placed obliquely inside the pipe, to
make it long).
Again, for the thermowell, treated as a fin, we have:

h×P
m=
k× Ac

P p ×d
and, = for d << d
Ac p ×d ×d
P 1
i.e. =
Ac d
i.e. fin parameter, m does not depend upon thermowell pocket diameter, when the wall thickness is very small
compared to its diameter.
Example 6.9. The temperature of air in an air stream in a tube is measured by a thermometer placed in a protective well
filled with oil. The thermowell is made of steel tube of 1.5 mm thick sheet of length 120 mm. The thermal conductivity of
steel = 58.8 W/(mK). and ha = 23.3 W/(m2K). If the air temperature recorded was 84°C, estimate the measurement error,
if the temperature at the base of the well was 40°C. [M.U.]
Solution. See Fig. Example 6.9.
Data:
L := 0.12 m d := 0.0015 m k := 58.8 W/(mK) To := 40°C TL := 84°C ha := 23.3 W/(m2C)
Let Ta be the temperature of air flowing.
Let us first calculate fin parameter m:
ha × P
m=
k × Ac

P p ×d 1
Again, = = (where, P is the perimeter, d is the diameter of thermowell.)
Ac p ×d×d d

262 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Thermometer
To = 40°C

Thermowell, thickness = d = 0.0015 m


k = 58.8 W/(mK)

L = 0.12 m Ta
TL = 84° C 2
ha = 23.3 W/(m K)
d

FIGURE Example 6.9 Error estimation in temperature measurement

ha
i.e. m: =
k ×d
or, m = 16.253 m–1 (fin parameter m)
and, cos h(m × L) = 3.5869
Considering the thermowell as a fin with insulated end, we have:
TL - Ta 1
=
To - Ta cos h( m×L)

84 - Ta 1
i.e. =
40 - Ta 3.5869
( 3.5869 ´ 84) - 40
or, Ta: =
2.5869
i.e. Ta = 101.009°C (temperature of air flowing.)
Error in temperature measurement:
Actual temperature of air is 101.009°C, while recorded temperature is 84°C.
i.e. Ta – TL = 17.009°C (error in temperature measurement)

Ta - TL
or, ×100 = 16.839 (Percentage error in measurement of temperature = 16.8%.)
Ta

6.6 Summary
Fins are widely used in industry to enhance heat transfer from surfaces. In this chapter, first, we derived the
general differential equation governing the temperature distribution in a fin, from an energy balance on a
differential element of the fin. Subsequently, solution of this differential equation with different boundary
conditions was obtained to get temperature distribution in the fin. Once the temperature distribution is known,
heat transfer rate through the fin is easily calculated by applying Fourier’s law. Four important cases considered
were:
Case (i): Infinitely long fin,
Case (ii): Fin insulated at its end (i.e. negligible heat loss from the end of the fin),
Case (iii): Fin losing heat from its end by convection, and
Case (iv): Fin with specified temperature at its two ends.
Performance of fins was discussed with reference to parameters such as fin efficiency and fin effectiveness;
concepts of thermal resistance of fins and total surface efficiency, or, area weighted fin efficiency of a fin array
was explained.
Graphs and tables for practically important fin geometries were presented.
Finally, application of fin theory to correction of error in temperature measurement was studied.

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 263


In this text, in our study of heat conduction so far, one of the assumptions in our analysis was steady state
heat conduction, i.e. temperature at any given location in the solid was assumed to be constant and did not
change with time. In the next chapter, we shall study the cases of transient conduction, i.e. when the temperature
at a given location in the solid changes with time.

Questions
1. Explain why fins are widely used. Discuss a few commonly used types of fins.
2. ‘Addition of fins may not necessarily increase the heat transfer from a surface; it may even decrease the heat
transfer’—comment on this statement.
3. Define ‘fin efficiency’ and ‘fin effectiveness’. Explain, as a corollary, why thin, closely spaced fins of a material
of good thermal conductivity are preferable.
4. Explain why fins are generally used on the gas side in a gas-to-liquid heat exchanger.
5. For an infinitely long fin, with usual notations, prove that heat dissipated is given by:
Qfin = h× P ×k × Ac × q o = h× P ×k × Ac × (To – Ta)
6. Using usual notations and starting from basics, derive and solve a differential equation for heat flow through a
moderately long pin fin (dT/dx at x = L is zero) to get an expression for the non-dimensional temperature
distribution along the length of the fin as:
q ( x) cos h (m×(L - x))
=
qo cos h(m×L)
and also show that heat transferred in the fin is given by:
Qfin = h× P ×k × Ac × q o ×tan h (m × L).
7. A thin fin of length L, has its two ends attached to two parallel walls which have temperatures, T1 and T2. The
fin loses heat by convection to ambient air at T¥. Obtain an analytical expression for the one dimensional tem-
perature distribution along the length of the fin.
8. The end of a very long cylindrical rod is attached to a heated wall and its surface is in contact with a cold fluid.
If the rod diameter were doubled, by what percentage would the heat transfer rate change?

Problems
1. A copper pin fin, 0.25 cm diameter, protrudes from a wall at 95°C into ambient air at 25°C. The heat transfer is
mainly by free convection with heat transfer coefficient, h = 10 W/(m2K). Calculate the heat loss assuming that
the fin is infinitely long. For copper, take k = 395 W/(mK).
2. Calculate the rate of heat loss from a rectangular fin of length 2 cm, on a plane wall. Thickness of fin is 2 mm
and its breadth is 20 cm. Take q 1 = 200°C, h = 17.5 W/(m2K), k = 52.5 W/(mK). Assume that heat loss from the
tip is negligible.
3. Aluminum square fins (0.5 mm x 0.5 mm) of 10 mm length are provided on the surface of an electronic device to
carry 45 mW of energy generated by the device. The temperature at the surface of the device should not exceed
80°C, while temperature of the surrounding medium is 40°C. Assume k for aluminium = 190 W/(mK), h = 12W/
(m2K). Find the number of fins required, neglecting heat loss from the end of the fin.
4. An aluminum fin, 0.5 mm square and 1cm long, is attached to a semiconductor device to provide additional
cooling. The base of the fin can be assumed to be at the inside temperature of 80°C. Find the cooling capacity
provided by the fin. Ambient temperature = 40°C, k = 177 W/(mK), h = 12.44 W/(m2K).
5. One end of a copper rod, 15 cm long and 0.6 cm diameter, is connected to a wall at 200°C while the other end
protrudes into a room whose air temperature is 21°C. If the tip of the rod is insulated, estimate the heat lost by
the rod, assuming the heat transfer coefficient between its surface and surrounding air as 28 W/(m2K). Also,
calculate the efficiency of the fin. Take k for copper = 370 W/(mK). State the assumptions made.
6. A cylinder 5 cm diameter and 1 m long, is provided with 12 longitudinal, straight fins of 1 mm thick and 2.5 cm
height. k of fin material is 75 W/(mK). Calculate the heat lost from the cylinder if the surface temperature of the
cylinder is 200°C and that of the surrounding is 40°C. Given: heat transfer coefficient between the cylinder and
fins and surrounding air = 25 W/(m2K).
7. Circumferential fins of constant thickness of 1 mm (k = 190 W/(mK)), are attached on a 50 mm OD pipe at a
pitch of 5 mm. Fin length is 20 mm. Wall temperature is 150°C. Convection heat transfer coefficient is 45 W/
(m2K). Determine heat flow rate from 1 m length of pipe. Compare the heat flow with fins to that without the
fins.

264 FUNDAMENTALS OF HEAT AND MASS TRANSFER


8. A hot plate (1 m ´ 1 m), at 150°C is to be cooled by attaching on its surface, 10,000 number of cylindrical, pin
fins of each, 3 mm diameter and 3 cm long. Surrounding air is at 25°C. Heat transfer coefficient between the fin
surfaces and the surroundings is 30 W/(m2C). Determine:
(i) overall surface effectiveness
(ii) heat transfer rate, with the fins in place
(iii) heat transfer rate from the plate, if there were no fins
(iv) decrease in thermal resistance due to attaching the fins.
9. An aluminium fins are fixed on one side (size: 1 m x 1 m), of an electronic device to increase the heat
dissipation. Fins are of rectangular cross section, 0.2 cm thick and 3 cm long. There are 100 fins per metre.
Convection heat transfer coefficient for both the plate and the fins is 30 W/(m2K). Determine the percentage
increase in the rate of heat transfer due to attaching the fins.
10. An iron bar, 15 mm in diameter, spans the distance between two plates, 50 cm apart. Air at 25°C flows in the
space between the plates resulting in heat transfer coefficient of 15 W/(m2K). Calculate the heat transfer and
temperature at the middle of the bar, if the plates are maintained at 125°C each. For iron, k = 45 W/(mK).
11. Two ends of a 6 mm diameter copper rod (U–shaped) having k = 330 W/(mK), are rigidly connected to a vertical
wall as shown in Fig. Problem 6.11. Wall temperature is constant at 100°C. Developed length of the rod is 50 cm
and is exposed to air at 30°C. Combined convective and radiative heat transfer coefficient is 30 W/(m2 K).
Calculate:
(i) the temperature at the centre of the rod
(ii) net heat transfer from the rod to air.

D = 0.006 m

TL = 100°C
2
h = 30 W/(m C)

To = 100°C Ta = 30°C

0.25 m

FIGURE Problem 6.11 U-shaped rod, both ends fixed to a wall

12. A steel rod (k = 55 W/(mK)), of length 50 cm, diameter 2.5 cm, has its two ends maintained at 150°C and 60°C.
Ambient air, to which heat is dissipated by the rod, is at 25°C and the heat transfer coefficient is 20 W/(m2 K).
Determine:
(i) minimum temperature in the rod
(ii) temperature at the mid-point of the rod, and
(iii) heat transfer rates from the left and right ends.
13. A Hg-thermometer placed in a well filled with oil, is required to measure the temperature of compressed air
flowing in a pipe. The well is 14 cm long and is made of steel 1.5 mm thick. The temperature indicated by the
thermometer is 100°C. The pipe wall temperature is 50°C. The film coefficient outside the wall is 30W/(m2C).
Estimate the % error in measurement of temperature of air. k for steel = 40W/(mC).

HEAT TRANSFER FROM EXTENDED SURFACES (FINS) 265


CHAPTER

7
Transient Heat
Conduction

7.1 Introduction
In chapter 3, we derived the general differential equation for conduction and then applied it to problems of
increasing complexity, e.g. first, we studied heat transfer in simple geometries without heat generation and then
we studied heat transfer when there was internal heat generation. In all these problems, steady state heat transfer
was assumed, i.e. the temperature within the solid was only a function of position and did not depend on time,
i.e. mathematically, T = T(x, y, z). However, all the process equipments used in engineering practice, such as
boilers, heat exchangers, regenerators, etc. have to pass through an unsteady state in the beginning when the
process is started, and, they reach a steady state after sufficient time has elapsed. Or, as another example, a billet
being quenched in an oil bath, goes through temperature variations with both position and time before it attains
a steady state. Conduction heat transfer in such an unsteady state is known as transient heat conduction or,
unsteady state conduction, or time dependent conduction. Obviously, in transient conduction, temperature
depends not only on position in the solid, but also on time. So, mathematically, this can be written as T = T(x, y,
z, t), where t represents the time coordinate.
Naturally, solutions for transient conduction problems are a little more complicated compared to steady
state analysis, since now, an additional parameter, namely time (t) is involved.
Typical examples of transient conduction occur in:
(i) heat exchangers
(ii) boiler tubes
(iii) cooling of cylinder heads in I.C. engines
(iv) heat treatment of engineering components and quenching of ingots
(v) heating of electric irons
(vi) heating and cooling of buildings
(vii) freezing of foods, etc.
Two types of transient conduction may be identified:
(a) periodic heat flow problems, where the temperatures vary on a regular, periodic basis, e.g. in I.C. engine
cylinders, alternate heating and cooling of earth during a 24 hr cycle (by sun) etc.
(b) non-periodic heat flow problems, where temperature varies in a non-linear manner with time.
To solve a given one-dimensional, transient conduction problem, one could start with one of the relevant
general differential equations discussed in chapter 3 and by solving it in conjunction with appropriate boundary
conditions, and get the temperature distribution as a function of position and time. For example, for one-
dimensional conduction, in Cartesian coordinates, we have:
d 2T 1 dT
= × -without heat generation)
dx 2 a dt
d 2T qg 1 dT
and, + = × (with heat generation.)
dx 2 k a dt
However, there is a set of problems encountered in practice, where the temperature gradients within the
solid are very small, (i.e. the internal resistance to conduction is negligible) which can be solved simply by
applying the energy balance principle. Consider for example, a small body made of, say, copper, at a high
temperature, being quenched in a medium like oil. Then, the body loses heat to the medium. Heat flows by
conduction from within the body to the surface and then, by convection to the medium. When the body is very
small or when the thermal conductivity of the material of the body is very large, temperature gradients within
the body will be very small and may be neglected. In such a case, temperature within the body is only a function
of time and is independent of spatial coordinates, i.e. the whole body acts as lump and temperatures of all points
within the body decrease (or increase if the object is being heated) uniformly en-mass. Heat transfer process from
the body, in this case, is controlled by the convection resistance at the surface rather than by the conduction
resistance in the solid. Such an analysis, where the internal resistance of the body for heat conduction is
negligible and the whole body may be treated as a lump as far as temperature increase or decrease is concerned,
is known as lumped system analysis.
In this chapter, first, we shall study the lumped system analysis; then, we shall present analytical and chart
solutions for some of the practically important transient conduction problems for the cases of a large slab, long
cylinder, sphere and a semi-inifinite medium. Finally, product solution method of solving multidimensional
transient conduction problems will be explained.

7.2 Lumped System Analysis (Newtonian Heating or Cooling)


As mentioned above, in lumped system analysis, the internal conduction resistance of the body to heat flow (i.e.
L/(k. A)) is negligible compared to the convective resistance (i.e. 1/(h . A)) at the surface. So, the temperature of
the body, no doubt, varies with time, but at any given instant, the temperature within the body is uniform and is
independent of position, i.e. T = T (t) only. Practical exam-
ples of such cases are: heat treatment of small metal pieces, A
measurement of temperature with a thermocouple or ther-
mometer, etc., where the internal resistance of the object for h, Ta
heat conduction may be considered as negligible.
Solid body
Analysis: (m, V, r, Ti)
Consider a solid body of arbitrary shape, volume V, mass m,
density r, surface area A, and specific heat Cp. See Fig. 7.1.
To start with, at t = 0, let the temperature throughout the Q = hA (Ta T(t))
body be uniform at T = Ti. At the instant t = 0, let the body
be suddenly placed in a medium at a temperature of Ta, as FIGURE 7.1 Lumped system analysis
shown. For the sake of analysis, let us assume that Ta > Ti;
however, same analysis is valid for Ta < Ti too. Then, heat will be transferred from the medium to the body and
the temperature of the body will increase with time. Let the temperature of the body rise by a differential amount
dT in a differential time interval dt, thus increasing the internal energy of the solid.
Writing an energy balance for this situation:
Amount of heat transferred into the body in time interval dt =
Increase in the internal energy of the body in time interval dt
i.e. h × A × (Ta – T(t)) × d t = m ×Cp × dT = r ×Cp ×V × dT ...(7.1)
since m = r×V
Now, since Ta is a constant, we can write:
dT = d(T(t) – Ta)
Therefore,
d(T (t ) - Ta ) - h×A
= × dt ...(7.2)
T (t ) - Ta r ×Cp ×V

TRANSIENT HEAT CONDUCTION 267


Integrating between t = 0 (i.e. T = Ti) and any t, (i.e. T = T(t)),
F (T(t ) - T ) I = -h×A×t
ln GH T - T JK r×C ×V
i a
a
p

F -h×A×t I
= exp G
H r×C ×V JK
T (t ) - T a
i.e. ...(7.3)
T -T
i a p

Now, let:
r×Cp ×V
=t
h× A
where, t is known as thermal time constant and has units of time.
Therefore, Eq. 7.3 is written as:
T (t ) - Ta -t FG IJ
Ti - Ta
= exp
t H K ...(7.4)

Now denoting q = (T(t) – Ta), we write Eq. 7.4 compactly as:


q T (t ) - Ta -t FG IJ
qi
=
Ti - Ta
= exp
t H K ...(7.5)

Eq. 7.5 gives the temperature distribution in a solid as a function of time, when the internal resistance of the
solid for conduction is negligible compared to the convective resistance at its surface.
Eq. 7.5 is represented graphically in Fig. 7.2.
From Eq. 7.5 and Fig 7.2, we note:
(i) temperature distribution is exponential, i.e. temperature changes rapidly initially and approaches that of
the medium exponentially.
(ii) either the time required by the body to reach a certain temperature or the temperature attained by the
body after a certain time interval, can be found out from Eq. 7.5.
(iii) larger the value of time constant t, longer is the time required for the body to reach a particular
temperature.
(iv) time required for the body to attain 36.8% of the applied temperature difference is indicated in the Fig.
7.2(a). This is known as one time period and is of importance in connection with measurement of
temperatures with thermocouples. Larger the value of time constant, larger is the time period. We shall
comment on this later in this chapter.

q/qi T

Tg
1 Exponential heating

t = (rCp V)/(hA)
Ti t

Exponential cooling
0.368
T1

FIGURE 7.2(a) Temperature variation with time FIGURE 7.2(b) Newtonian heating and cooling
in a lumped system

268 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Instantaneous heat transfer:
At any instant t, heat transfer between the body and the environment is easily calculated since we have the
temperature distribution from Eq. 7.4:
dT(t )
Q(t) = m × Cp × ,W ...(7.6a)
dt
At that instant, heat transfer must also be equal to:
Q(t) = h × A × (T(t) – Ta), W ...(7.6b)
Total heat transfer:
Total heat transferred during t = 0 to t = t, is equal to the change in internal energy of the body:
Qtot = m × Cp × (T(t) – Ti), J ...(7.7a)
Qtot may also be calculated by integrating Eq. 7.6a:

Maximum heat transferred:


Qtot =
z
0
t
Q(t ) dt , J ...(7.7b)

When the body reaches the temperature of the environment, obviously, maximum heat has been transferred:
Qmax = m × Cp × (Ta – Ti), J ...(7.8)
If Qmax is negative, it means that the body has lost heat, and if Qmax is positive, then body has gained heat.

7.3 Criteria for Lumped System Analysis


(Biot Number and Fourier Number)
For the simple analysis made above, we had the fundamental assumption that the internal conductive resistance
of the body was negligible as compared to the convective resistance at its surface. This was stated in a rather
qualitative way. Now, let us study the criteria required for the lumped system analysis to be applicable.
Consider a plane slab in steady state, transferring heat to a fluid
on its surface with a heat transfer coefficient of h, as shown in Bi << 1
Fig. 7.3. (The criterion arrived at is readily extended to transient Bi = 1
conditions later.)
Bi >> 1
Let the surface on the left be maintained at temperature T1 and T1
the surface on the right is at a temperature of T2 as a result of heat T2
h, Ta
being lost to a fluid at temperature Ta, flowing with a heat transfer T2
coefficient h. Writing an energy balance at the right hand surface,
k ×A T2
× (T1 – T2) = h × A × (T2 – Ta)
L Ta
Rearranging, Qcond Qconv

FLI
T1 - T2
GH k×A JK Rcond h×L
L
X
T2 - Ta
=
F1I =
Rconv
=
k
= Bi ...(7.9)
GH h×A JK FIGURE 7.3(a) Biot number and
temperature distribution in a plane wall
The term, (h . L)/k, appearing on the RHS of Eq. 7.9 is a dimensionless number, known as Biot number.
Biot number is a measure of the temperature drop in the solid relative to the temperature drop in the con-
vective layer. It is also interpreted as the ratio of conductive resistance in the solid to the convective resistance at
its surface. This is precisely the criterion we are looking for. Note from Fig. 7.3(a) the temperature profile for Bi
<< 1. It suggests that one can assume a uniform temperature distribution within the solid if Bi << 1.
Situation during transient conduction is shown in Fig. 7.3(b). It may be observed that temperature distribu-
tion is a strong function of Biot number. For Bi << 1, temperature gradient in the solid is small and temperature
can be taken as a function of time only. Note also that for Bi >> 1, temperature drop across the solid is much
larger than that across the convective layer at the surface.
Therefore, to fix the criterion for which lumped system analysis is applicable, let us define Biot number, in
general, as follows:
TRANSIENT HEAT CONDUCTION 269
T(x, 0) = Ti T(x, 0) = Ti

h, Ta

Bi << 1 Bi = 1 Bi >> 1

FIGURE 7.3(b) Biot number and transient temperature distribution in a plane wall

h×Lc
Bi = ...(7.10)
k
where, h is the heat transfer coefficient between the solid surface and the surroundings, k is the thermal
conductivity of the solid, and Lc is a characteristic length defined as the ratio of the volume of the body to its
surface area, i.e.
V
Lc =
A
With this definition of Bi and Lc, for solids such as a plane slab, long cylinder and sphere, it is found that
transient temperature distribution within the solid at any instant is uniform, with the error being less than about
5%, if the following criterion is satisfied:
h×Lc
Bi = < 0.1 ...(7.11)
k
In other words, if the conduction resistance of the body is less than 10% of the convective resistance at its
surface, the temperature distribution within the body will be uniform within an error of 5%, during transient
conditions.
Lc for common shapes:
A ×2×L
(i) Plane wall (thickness 2L): Lc = = L = half-thickness of wall
2× A

p ×R 2 ×L R
(ii) Long cylinder, radius R: Lc = =
2×p ×R ×L 2
4
×p × R3
3 R
(iii) Sphere, radius, R: Lc = 2
=
4×p ×R 3

L3 L
(iv) Cube, side L: Lc = =
6×L2 6
Therefore, we can write Eq. 7.3 as:

- h× A×t F I
q
qi
=
T (t ) - Ta
Ti - Ta
= exp
r ×Cp ×V
GH JK if Bi < 0.1 ...(7.12)

Eq. 7.12 is important. Its appliation to a given problem is very simple and solution of any transient
conduction problem must begin with examining if the criterion, Bi < 0.1 is satisfied to see if Eq. 7.12 could be
applied.

270 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Now, the term (h×A×t)/(r× Cp ×V) can be written as follows:

h×A×t FG h×L IJ × FG k×t IJ = FG h×L IJ × FG a ×t IJ = Bi × Fo V


H k K GH r×C ×L JK H k K H L K
c c
= ...since = Lc
r ×Cp ×V p
2
c
2
c A

a ×t
where, Fo = = Fourier number, or relative time.
L2c
Fourier number, like Biot number, is an important parameter in transient heat transfer problems. It is also
known as ‘dimensionless time’. Fourier number signifies the degree of penetration of heating or cooling effect
through a solid. For small Fo, large t will be required to get significant temperature changes.
With the aforesaid definitions of Biot number and Fourier number, now, we can rewrite Eq. 7.12 as:
q T (t ) - Ta
= = exp (–Bi × Fo) if Bi < 0.1 ...(7.13)
qi Ti - Ta
Eq. 7.13 is plotted in Fig. 7.4 below. On the X-axis, (Bi . Fo) is plotted against q/qi on Y-axis. As expected, the
graph is a straight line, with a negative slope when the Y-axis has logarithmic scale. Remember that this graph is
for the cases where lumped system analysis is applicable, i.e. Bi < 0.1.

Transient temperature distribution


in solids, Bi < 0.1
1

Note: X = Bi×Fo
0.1

exp ( X)

0.01

3
1◊10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
X

FIGURE 7.4 Dimensionless temperature distribution in solids during transient heat transfer, (Bi < 0.1), for
lumped system analysis

7.4 Response Time of a Thermocouple


Lumped system analysis is usefully applied in the case of temperature measurement with a thermometer or a
thermocouple. Obviously, it is desirable that the thermocouple indicates the source temperature as fast as
possible. If the thermocouple is measuring changing temperatures, then also, it should follow the temperature
changes at a rate faster than the rate of temperature change. ‘Response time’ of a thermocouple is defined as the
time taken by it to reach the source temperature.
Consider Eq. 7.12:
F
- h× A×t I
q
qi
=
T (t ) - Ta
Ti - Ta
= exp GH
r ×Cp ×V
JK if Bi < 0.1 ...(7.12)

For rapid response, the term (h A t )/(r C p V ) should be large so that the exponential term will reach zero
faster. This means that:
(i) increase (A/V), i.e. decrease the wire diameter

TRANSIENT HEAT CONDUCTION 271


(ii) decrease density and specific heat, and
(iii) increase the value of heat transfer coefficient h.
As mentioned earlier, the quantity (r C p V )/(h ×A ) is known as ‘thermal time constant’, t, of the measuring
system and has units of time. At t = t, i.e. at a time interval of one time constant, we have:
T (t ) - Ta
= e –1 = 0.368 ...(7.14)
Ti - Ta
From Eq. 7.14, it is clear that after an interval of time equal to one time constant of the given temperature
measuring system, the temperature difference between the body (thermocouple) and the source would be 36.8%
of the initial temperature difference, i.e. the temperature difference would be reduced by 63.2%.
Time required by a thermocouple to attain 63.2% of the value of initial temperature difference is called its
sensitivity.
For good response, obviously, the response time of thermocouple should be low. As a thumb rule, it is
recommended that while using a thermocouple to measure temperatures, reading of the thermocouple should be
taken after a time equal to about four time periods has elapsed.
Example 7.1. A steel ball of 5 cm diameter initially at a uniform temperature of 450°C is suddenly placed in an environ-
ment at 100°C. Heat transfer coefficient h, between the steel ball and the fluid is 10 W/(m2K). For steel, cp = 0.46 kJ/
(kgK), r = 7800 kg/m3, k = 35 W/(mK). Calculate the time required for the ball to reach a temperature of 150°C. Also,
find the rate of cooling after 1 hr. Show graphically how the temperature of the sphere falls with time. [M.U.]
Solution.
Data:
R := 25 ´ 10–3 m r := 7800 kg/m3 Cp := 460 J/(kgK) k := 35 W/(mK) Ti := 450°C Ta := 100°C
4
h := 10 W/(m2K) T := 150°C A := 4 × p × R2, m2 V := × p × R3, m3
3
First, calculate the Biot number:
4 LM OP
h×Lc h V FG IJ h 3
×(p )× R 3
N Q
Bi =
k
= ×
k A H K = ×
k 4 ×p × R 2

h R
i.e. Bi: = × (define Biot number)
k 3
–3
i.e. Bi = 2.381 ´ 10 (Biot number.)
Since Bi < 0.1, lumped system analysis is applicable, and the temperature variation within the solid will be within
an error of 5%. Applying Eq. 7.12, we get:

T(t ) - Ta - h× A ×t F I
q
qi
=
Ti - Ta
= exp
r ×Cp ×V
GH JK if Bi < 0.1 ...(7.12)

T - Ta -t FG IJ
i.e.
Ti - Ta
= exp
t H K where, t is the time constant.

And, time constant is given by:


r×c p ×V r×cp R
t= = × (since for sphere, V/A = R/3)
A× h h 3
r×cp R
i.e. t := × (define time constant, t)
h 3
i.e. t = 2990 s (time constant)
Therefore, we write:
150 - 100 -t FG IJ
450 - 100
=exp
2990 H K where, t is the time required to reach 150°C

FG 50 IJ = -t
i.e. ln
H 350 K 2990

272 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG 50 IJ s
or, t := – 2990 ln
H 350 K (define t, the time required to reach 150°C)

i.e. t = 5.818 ´ 103 s (time required to reach 150°C.)


i.e. t = 1.616 hrs.
Rate of cooling after 1 hr.:
i.e. t := 3600 s
From Eq. 7.12, we have:
LM F - h×A×t I + T OP
MN
T(t ) := (Ti - Ta )×expGH r ×c ×V JK P (define T (t))
Q
a
p

dT L - h×A OP × exp F - h×A×t I C/s


=(T – T ) × M
i.e.
dt
i
MN r ×V ×c PQ GH r ×c ×V JK
a
p p
(rate of cooling)

d
i.e. T(t ) = – 0.035 C/s (rate of cooling after 1 hr.)
dt
negative sign indicates that as time increases, temperature falls.
Note that in Mathcad, there is no need to separately differentiate and substitute the values. All that is done in one
step as shown above.
To sketch the fall in temperature of sphere with time:
Temperature as a function of time is given by Eq. 7.12:

T(t ) - Ta - h × A ×t F I
q
qi
=
Ti - Ta
= exp
r ×c p ×V
GH JK if Bi < 0.1 ...(7.12)

F - h×A×t I
i.e. T(t) := Ta + (Ti – Ta) × exp GH r ×c ×V JK
p
...(A)

We will plot Eq. A against different times, t :


We use Mathcad to draw the graph. First, define a range variable t, varying from 0 to say, 4 hrs, with an increment
of 0.1 hrs. Then, choose x–y graph from the graph palette, and fill up the place holders on the X-axis and Y-axis with t
and T (t), respectively. Click anywhere outside the graph region and immediately the graph appears:
t := 0, 0.1, ... , 4 (define a range variable, t varying from zero to 4 hrs, with an increment of 0.1 hrs.)

Transient cooling of sphere-lumped system


500

450

400

350
T(t×3600) 300 t in hrs. and T(t) in deg. C

250

200
150
100
0 0.5 1 1.5 2 2.5 3 3.5 4
t
FIGURE Example 7.1 Transient cooling of a sphere considered as a lumped system

Note from the Fig. 7.4 how the cooling progresses with time. After about 4 hrs duration, the sphere approaches the
temperature of the ambient. You can also verify from the graph that the time required for the sphere to reach 150°C is
1.616 hrs, as calculated earlier.

TRANSIENT HEAT CONDUCTION 273


Example 7.2. A 50 cm x 50 cm copper slab, 6 mm thick, at a uniform temperature of 350°C, suddenly has its surface
temperature lowered to 30°C. Find the time at which the slab temperature becomes 100°C. Given: r = 9000 kg/m3, cp =
0.38 kJ/(kgK), k = 370 W/(mK), h = 100 W/(m2K). Also, find out the rate of cooling after 60 seconds.
Solution.
Data:
L := 0.05 m B := 0.05 m (breadth) d := 0.006 m (thickness) r := 9000 kg/m3 cp := 380 J/(kgK)
k := 370 W/(mK) Ti := 350°C Ta := 30°C h := 100 W/(m2K) T := 100°C
A := 2 × L × B, m2 V := L × B × d , m3
First, calculate the Biot number:
V
Characteristic length: Lc := (define characteristic length for the plate)
A
i.e. Lc =3 ´ 10–3 m (characteristic length for plate = half the thickness )
h×Lc
Therefore, Biot number: Bi: = (define Biot number)
k
–4
i.e. Bi = 8.108 ´ 10 (Biot number)
Since Bi < 0.1, lumped system analysis is applicable, and the temperature variation within the solid will be within
an error of 5%. Applying Eq. 7.12, we get:

T(t ) - Ta F
- h× A ×t I if Bi < 0.1
q
qi
=
Ti - Ta
= exp GH
r ×Cp ×V
JK ...(7.12)

T - Ta F -t I
= exp G J where, t is the time constant.
i.e.
Ti - Ta HtK
And, time constant is given by:
r×c p ×V r×cp
t= = × Lc (since for plate, V/A = Lc)
A× h h

r×cp
i.e. t: = × Lc (define time constant, t)
h
i.e. t = 102.6 s (time constant)
Therefore, we write:
100 - 30 -t FG IJ
350 - 30
=exp
102.6 H K where, t is the time required to reach 100°C

FG 70 IJ = -t
i.e. ln
H 320 K 102.6
FG 70 IJ s
or, t := – 102.6 ln
H 320 K ...define t, the time required to reach 100°C

i.e. t = 155.934 s ...time required to reach 100°C.


i.e. t = 0.043 hrs.
Rate of cooling after 60 s:
i.e. t = 60 s
From Eq. 7.12, we have:
LM F - h×A×t I + T OP
MN
T(t) := (Ti - Ta )×expGH r ×V ×c JK P ...define T (t)
Q
a
p

dT F h×A I × exp F - h×A×t I C/s


=(T – T ) × G
H r ×V ×c JK GH r ×V ×c JK
i.e. i a ...rate of cooling
dt p p

d
i.e. T(t) = – 1.738 C/s ...rate of cooling after 60 s.
dt
Negative sign indicates that as time increases, temperature falls.

274 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Note again, that in Mathcad, there is no need to separately differentiate and substitute the values. All that is done
in one step as shown above.
Example 7.3. A carbon steel (AISI 1010) shaft of 0.2 m diameter is heat treated in a gas-fired furnace whose gases are at
1200 K and provide a convection coefficient of 80 W/(m2K). If the shaft enters the furnace at 300 K, how long must it
remain in the furnace to achieve a centre line temperature of 900 K? Given thermophysical properties of AISI 1010
carbon steel: r = 7854 kg/m3, k = 48.8 W/(mK), cp = 559 J/(kgK).
Solution.
Data:
R := 0.1 m r := 7854 kg/m3 cp := 559 J/(kg K) k := 48.8 W/(m K) Ti := 300 K Ta := 1200 K
V
h := 80 W/(m2K) T := 900 K A := 2 × p ×R × L, m2 V := p × R2 × L, m3 Lc :=
A
R
i.e. Lc = m (characteristic length)
2
i.e. Lc = 0.05 m (characteristic length.)
First, calculate the Biot number
h ×Lc
Bi: = (define Biot number)
k
i.e. Bi = 0.082 (Biot number)
Since Bi < 0.1, lumped system analysis is applicable, and the temperature variation within the solid will be within
an error of 5%. Applying Eq. 7.12, we get:

T(t ) - Ta - h× A ×t F I
q
qi
=
Ti - Ta
= exp
r ×Cp ×V
GH JK if Bi < 0.1 ...(7.12)

T - Ta -t FG IJ
i.e.
Ti - Ta
= exp
t H K where, t is the time constant.

And, time constant is given by:


r×c p ×V r×cp R
t= = × (since for cylinder, V/A = R/2)
A× h h 2

r×cp R
i.e. t: = × (define time constant, t)
h 2
i.e. t = 2.74399 ´ 103 s (time constant)
Therefore, we write:
900 - 1200 -t FG IJ
300 - 1200
=exp
2743.99 H K where, t is the time required to reach 900 K

FG 300 IJ = -t
i.e. ln
H 900 K 2743.99
FG 300 IJ
or, t := – 2743.99 ln
H 900 K s (define t, the time required to reach 900 K)
3
i.e. t = 3.015 ´ 10 s (time required to reach 900 K.)
i.e. t = 0.838 hrs.
Example 7.4. A thermocouple (TC) junction is in the form of 8 mm sphere. Properties of the material are: cp = 420 J/
(kgK), r = 8000 kg/m3, k = 40 W/(mK), and heat transfer coefficient, h = 45 W/(m 2K). Find, if the junction is initially at
a temperature of 28°C and inserted in a stream of hot air at 300°C:
(i) the time constant of the TC
(ii) The TC is taken out from the hot air after 10 s and kept in still air at 30°C. Assuming ‘h’ in air as 10 W/(m 2K),
find the temperature attained by the junction 15 s after removing from hot air stream. [M.U.]
Solution.
Data:
R := 4 ´ 10–3 m r := 8000 kg/m3 cp := 420 J(kg k) k := 40 W/(mK) Ti := 28°C Ta := 300°C
4
h := 45 W/(m2K) A := 4 × p × R2, m2 V := × p × R3, m3
3
TRANSIENT HEAT CONDUCTION 275
First, calculate the Biot number:
4 LM OP
h×Lc h V FG IJ h 3
×(p )× R 3
N Q
Bi =
k
= ×
k A H K = ×
k 4 ×p × R 2

h R
i.e. Bi: = × (define Biot number)
k 3
–3
i.e. Bi = 1.5 ´ 10 (Biot number)
Since Bi < 0.1, lumped system analysis is applicable, and the temperature variation within the solid will be within
an error of 5%.
See Fig. Example 7.4 (a).
Time constant is given by:
Thermocouple, D = 8 mm
Ti = 28°C r×c p ×V r×cp R
t= = × (since for sphere, V/A = R/3)
A×h h 3
r×cp R
i.e. t := × ...define time constant, t)
h 3
Ta = 300°C Air i.e. t = 99.556 s (time constant.)
h = 45 W/(m K)
2 Temperature of TC after 10s:
t := 10 s (time duration for which TC is kept in
the stream at 300°C)
We use Eq. 7.12, i.e.
FIGURE Example 7.4 (a) Temperature meas-
urement, with thermocouple placed in the air T(t ) - Ta - h× A ×t F I
stream
q
qi
=
Ti - Ta
= exp
r ×Cp ×V
GH JK if Bi < 0.1 …(7.12)

T - Ta -tFG IJ
i.e.
Ti - Ta
= exp
t H K where, t is the time constant.

FG -t IJ + T °C
Therefore, T: =(Ti – Ta) × exp
HtK a (define temperature of TC after 10 s in the stream)

i.e. T = 53.994°C (temperature of TC 10 s after it is placed in the stream at 30°C)


(b) Now, this TC is removed from the stream at 300°C and placed in still air at 30°C. So, the temperature of 53.994°C
becomes initial temperature Ti for this case:
i.e. new Ti: Ti := 53.994°C (initial temperature when the TC is placed in still air)
And, new t : t := 15 s (duration for which TC is kept in still air)
And, new Ta: Ta := 30°C (new temperature of ambient)
And, new h: h := 10 W/(m2K) (heat transfer coefficient in still air.)
See Fig. Example 7.4 (b).
And, new time constant:
r×cp R
Thermocouple, D = 8 mm i.e. t := h × 3 (define time constant, t)
Ti = 53.994°C
i.e. t = 448 s (time constant)
FG -t IJ + T
Still air Therefore, T: = (Ti –Ta) × exp
HtK a °C

...define temperature
Ta = 300°C
2 of TC after 15 s in still air
h = 10 W/(m K)
i.e. T = 53.204°C (temperature of TC 15 s after it
is placed in still air at 30°C.
FIGURE Example 7.4 (b) Temperature
measurement, with thermocouple placed in
7.5 Mixed Boundary Condition
still air
In the cases studied so far, transient conduction was induced
in the solid by subjecting it to convection on all its sides.

276 FUNDAMENTALS OF HEAT AND MASS TRANSFER


However, this need not be always so. Transient conditions
can also be induced in a body by having radiation on any
of its surfaces, or by subjecting any of its surfaces to
electrical heating, or by internal heat generation caused by
flow of electric current. Heat flux
2 Slab, Convection h, Ta
In the general case, where transient conditions are q, W/m
T(t)
induced in a body by the combined effect of convection,
radiation and heat generation, the controlling differential
equation can be derived in the usual way, by writing an
energy balance on the body taken as a control volume, i.e.
net energy entering into the body results in an increase in
the internal energy of the body. However, the resulting
differential equation will be a nonlinear one, and is not
amenable to exact analytic solution, and has to be solved L
by approximate finite difference methods.
FIGURE 7.5 Transient conduction in a slab with
Let us analyse one particular case (which is quite
mixed boundary conditions
common), where one boundary surface is subjected to a
uniform heat flux and the other boundary surface is subjected to convection. See Fig. 7.5.
As shown in the Fig. 7.5, a slab of thickness L, is subjected to a uniform heat flux q (W/m2) at its left face and
loses heat by convection on its right face to a fluid at a temperature Ta, with a heat transfer coefficient, h. Then,
applying energy balance for this case, we write:
(Energy going into the slab – Energy leaving the slab) = Increase in internal energy of the slab

i.e. q × A – h × A × (T(t) – Ta) = r × Cp ×A ×L ×


bg
dT t
dt

i.e.
dT t
+
bg
h×(T (t ) - Ta )

q
=0 ...(7.15)
dt r ×Cp ×L r×C p ×L

Substituting: q = T(t) – Ta i.e.


dq
=
dT t bg
dt dt
h× A
and, a=
r ×V ×C p

q×A
and, b= (remember: A/V = 1/L)
r ×V ×C p

dq
Eq. 7.15 becomes: + a×q – b = 0 ...(7.16)
dt
b
Now, introduce the transformation: q ¢ = q – ...(a)
a
dq ¢ dq
then, =
dt dt
and, substituting Eq. a in 7.16:
dq ¢
+ a×q ¢ = 0 ...(7.17)
dt
Seperating the variables and intergrating from t = 0 to t = t, (and, q ¢ = qi ¢ to q ¢ = q ¢)

=exp (–a × t) ...(7.18)
q i¢

TRANSIENT HEAT CONDUCTION 277


Substituting now for q ¢ and q :

FG b IJ
T (t ) - Ta -
H a K =exp (–a × t )
F
-G J
bI
...(7.19)
Ti - Ta
H aK
b
T (t ) - Ta a × (1 – exp (–a × t))
i.e. =exp (–a × t) + ...(7.20)
Ti - Ta Ti - Ta

LM T(t ) - T - FG b IJ OP
−1
MM
a
H aK P
MN T - T - GH a IJK PPQ
F
and, also from Eq. 7.19: t= × ln ...(7.21)
a b
i a

Note that for t = ¥, Eq. 7.20 reduces to:


b q
T(t) = Ta + = Ta + ...(7.22)
a h
Eq. 7.22 gives the steady state temperature in the slab.
Example 7.5. An aluminium plate (r = 2707 kg/m3, Cp = 0.896 kJ/(kgC), and k = 200 W/(mC)) of thickness 3 cm is at an
initial, uniform temperature of 60°C. Suddenly, it is subjected to uniform heat flux q = 8000 W/m2, on one surface while
the other surface is exposed to an air stream at 25°C, with a heat transfer coefficient of h = 50 W/(m2C).
(i) Is lumped system analysis applicable to this case?
(ii) If yes, plot the temperature of the plate as a function of time, and
(iii) What is the temperature of the plate in steady state?
Solution. See Fig. Example 7.5.

3
k = 200 W/(mC), r = 2707 kg/m ,
Cp = 896 J/(kgK)

Convection
Heat flux Slab, 2
2 h = 50 W/(m C)
q = 8 kW/m T(t)
Ta = 25°C

L = 0.03 m

Figure Example 7.5 Transient conduction in a slab with mixed boundary conditions

Data:
L := 0.03 m r := 2707 kg/m3 Cp := 896 J/(kgC) k := 200 W/(mC) Ti := 60°C Ta := 25°C
h := 50 W/(m 2C) q := 8000 W/m2
First, claculate the Biot number:
L
h ×Lc h F I=
V h×
Bi =
k
= × GH JK
k 2× A k
2 (definition of Biot number)

278 FUNDAMENTALS OF HEAT AND MASS TRANSFER


h ×L
i.e. Bi: = i.e. Bi = 3.75 ´ 10–3 ...Biot number
2 ×k
Since Bi < 0.1, lumped system analysis is applicable, and the temperature variation within the solid will be within
an error of 5%.
To plot the temperature of plate as a function of time:
Clearly, this is a case of mixed boundary conditions, wherein at the left surface there is heat input by uniform heat flux
impinging on that surface, and on the right surface, there is heat loss by convection. So, we can directly apply Eq. 7.20.
b
T(t ) - Ta a × (1 – exp (–a × t))
i.e. = exp (–a× t ) + ..(7.20)
Ti - Ta Ti - Ta

h× A
Here, a=
r ×V ×Cp

q× A
and, b= (remember: A/V = 1/L)
r ×V ×Cp

h
i.e. a := i.e. a = 6.872 ´ 10–4
r×Cp ×L

q
and, b := i.e. b = 0.11
r×Cp ×L
Therefore, from Eq. 7.20:

LM b OP
T(t) := Ta + (T – T ) × Mexp( - a×t ) + a ×(1 - exp( - a×t ))P ...define T(t)
i a
MMN T -T
i a PPQ
To plot T(t) against time, let us define a range variable t, from say 0 s to 10,000 s, at an interval of 50 s. Then, select
the x–y plot from the graph palette, and fill up the place holders on the x-axis and y-axis with t and T(t), respectively.
Click anywhere outside the graph and immediately the graph appears: See Fig. Ex. 7.5(b)
t := 0, 50, ... , 10,000 (define a range variable t , such that initial value = 0,
next value = 50 and last value = 1000 s.)

Transient conductance in a plate (mixed B.C.)


200
190
180
170
160
150
140
T(t) 130 t in seconds, and T(t) in deg.c
120
110
100
90
80
70
60
50 4
0 2000 4000 6000 8000 1 ¥ 10
t

Figure Example 7.5(b)

TRANSIENT HEAT CONDUCTION 279


Temperature of plate in steady state:
We directly use Eq. 7.22 for t = ¥, i.e. steady state condition:
q
i.e. Tsteady = Ta + ...(7.22)
h
i.e. Tsteady = 185°C (steady state temperature of plate.)
Note from the above graph, that at large t (= beyond about 8,000 s), the temperature of the plate, indeed, tends to
a value of 185°C.
Example 7.6. A household electric iron has an aluminium base (r = 2700 kg/m3, Cp = 0.896 kJ/(kgC), and k = 200 W/
(mC)), and weighs 1.5 kg. Total area of iron is 0.06 m2 and it is heated with a 500 W heating element. Initially, the iron
is at ambient temp. of 25°C. How long will it take for the iron to reach 110°C once it is switched on? Take heat transfer
coefficient between iron and the ambient air as 15 W/(m2K).
Solution.
Data:
A := 0.06 m2 r := 2700 kg/m3 Cp := 896 J/(kgC) k := 200 W/(mC) Ti := 25°C Ta := 25°C
2
h := 15 W/(m C) T := 110°C M := 1.5 kg Q := 500 W
First, calculate the Biot number:
M
V := i.e. V = 5.556 ´ 10–4, m3 (volume of iron)
r
h × Lc h V FG IJ
Bi =
k
= ×
k A H K (definition of Biot number)

M FG IJ
Bi :=
h
×
r H K (since Volume of iron = Mass/density )
k A
i.e. Bi = 6.944 ´ 10–4 (Biot number.)
Since Bi < 0.1, lumped system analysis is applicable, and the temperature variation within the solid will be within
an error of 5%.
Now, writing the energy balance for the iron at any time t,
Rate of total heat generated – Rate of heat lost by convection = Rate of increase of internal energy
dT (t )
i.e. Q – h × A × (T(t) – Ta) = r×V×Cp × ...(a)
dt
dT (t ) h ×A ×(T (t ) - Ta ) Q
i.e. + – = 0 ...(b)
dt r ×V ×Cp r×V ×C p

dq dT (t )
Substituting: q = T(t) – Ta i.e. =
dt dt
h× A
and, let: a=
r ×V ×Cp

Q
and, b=
r×V ×C p

dq
Eq. b becomes: + a× q – b = 0 ...(c)
dt
Note that Eq. c is the same as Eq. 7.16, derived earlier. And, the solution for t is obtained as Eq. 7.21, with the
definition of ‘a’ and ‘b’ as follows:
h× A
a := i.e. a = 6.696 ´ 10–4
r ×V ×Cp

Q
and, b := i.e. b = 0.372
r×V ×C p

280 FUNDAMENTALS OF HEAT AND MASS TRANSFER


LM
T - Ta -
b OP
And, t: =
-1
a
×ln MM a
b PP ...(7.21)
MN
Ti - Ta -
a PQ
i.e. t = 247.975 s ...time required for iron to reach 110°C.

7.6 One-dimensional Transient Conduction in Large Plane Walls, Long


Cylinders and Spheres when Biot Number > 0.1
There are many situations in practice, where the temperature gradient in the solid is not negligible (i.e. Bi > 0.1)
and the lumped system analysis is not applicable. In such situations, we start with the general differential
equation for time dependent, one-dimensional conduction in the appropriate coordinate system and solve it in
conjunction with the boundary conditions.
In this section, we shall analyse one-dimensional transient conduction in large plane walls, long cylinders
and spheres when Bi > 0.1.
7.6.1 One Term Approximation Solutions
Fig. 7.6 shows schematic diagram and coordinate systems for a large, plane slab, long cylinder and a sphere.

T = Ti at t = 0 T = Ti at t = 0

T = Ti at t = 0

Convection h, Ta Convection
h, Ta Convection
h, Ta

L R R
X r r
(a) Large, plane slab (b) Long cylinder (c) Sphere

FIGURE 7.6 One-dimensional transient conduction in simple geometries

Consider a plane slab of thickness 2L, shown in Fig. 7.6(a) above. Initially, i.e. at t = 0, the slab is at an
uniform temperature, Ti. Suddenly, at t = 0, both the surfaces of the slab are subjected to convection heat transfer
with an ambient at temperature Ta, with a heat transfer coefficient of h, as shown. Since there is geometrical and
thermal symmetry, we need to consider only half the slab, and that is the reason why we chose the origin of the
coordinate system on the mid-plane. Then, we can write the mathematical formulation of the problem for plane
slab as follows:
d 2T 1 dT
= × in 0 < x < L, for t > 0 ...(7.23, a)
dx 2 a dt
dT
=0 at x = 0, for t > 0 ...(7.23, b)
dx
dT
–k × = h ×(T – Ta) at x = L, for t > 0 ...(7.23, c)
dx
T = Ti for t = 0, in 0 < x < L ...(7.23, d)
The solution of the above problem, however, is rather involved and consists of infinite series. So, it is more
convenient to present the solution either in tabular form or charts.

TRANSIENT HEAT CONDUCTION 281


For this purpose, we define the following dimensionless parameters:
(i) Dimensionless temperature:
T ( x , t ) - Ta
q (x, t) =
Ti - Ta
(ii) Dimensionless distance from the centre:
x
X=
L
(iii) Dimensionless heat transfer coefficient:
h×L
Bi = (Biot number)
k
(iv) Dimensionless time:
a ×t
Fo = (Fourier number)
L2
Non-dimensionalisation of the results with the above-mentioned dimensionless numbers enables us to
present the results practically over a wide range of operating parameters, either in tabular or graphical forms.
To deal with a long cylinder or a sphere, we do exactly what we did with the plane slab, i.e. start with the
appropriate differential equation for one-dimensional, time dependent conduction in cylindrical or spherical
coordinates. Boundary conditions will be the same as in Eq. 4.23 except that x is replaced by r and L is replaced
by R. Again, results are non-dimensionalised with the dimensionless parameters as mentioned above; however,
note one important difference in defining Biot number now, while using the tabular or chart solutions:
Characteristic length in Biot number is taken as half-thickness L for a plane wall, Radius R for a long
cylinder and sphere instead of being calculated as V/A, as done in lumped system analysis.
For all these three geometries, as mentioned earlier, the solution involves infinite series, which are difficult
to deal with. However, it is observed that for Fo > 0.2, considering only the first term of the series and neglecting
other terms, involves an error of less than 2%. Generally, we are interested in times, Fo > 0.2. So, it becomes very
useful and convenient to use one term approximation solution, for all these three cases, as follows:
T ( x , t ) - Ta - l2 × Fo l ×x FG IJ
Plane wall: q (x, t ) =
Ti - Ta
= A1× e 1 × cos 1
L H K ...Fo > 0.2 ...(7.24, a)

T (r , t ) - Ta - l2 × Fo l ×r FG IJ
Long cylinder: q (x, t ) =
Ti - Ta
= A1× e 1 × J0 1
R H K ...Fo > 0.2 ...(7.24, b)

FG l ×r IJ
HRK
1
sin
T (r , t ) - Ta - l2 × Fo
Sphere: q (x, t ) = = A1 × e 1 × ...Fo > 0.2 ...(7.24, c)
Ti - Ta l 1 ×r
R
In the above equations, A1 and l1 are functions of Biot number only.
A1 and l1 are calculated from the following relations:
For Plane wall:
l 1 × tan(l1) = Bi
4 × sin( l 1 )
A1 =
2×(l 1 ) + sin 2×(l 1 )
For Long cylinder:
J (l )
l1 × 1 1 = Bi
J0 (l 1 )
2 × J 1 (l 1 )
A1 =
l 1 ×[( J 0 ( l 1 )) 2 + ( J1 ( l 1 ))2 ]

282 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For Sphere:
1 – l1 × cot (l1) = Bi
4× sin( l 1 ) _ ( l 1 )×cos (l 1 )
A1 =
2×( l 1 ) - sin 2×(l 1 )
Values of A1 and l1 are given in Table 7.1. (See Appendix at the end of this chapter for Mathcad functions to
calculate these parameters). Function J0 is the zeroth order Bessel function of the first kind and J1 is the first order
Bessel function of the first kind. Values of J0 and J1 can be read from Table 7.2. (Obtained directly from Mathcad).
Now, at the centre of a plane wall, cylinder and sphere, we have the condition x = 0 or r = 0. Then, noting
that cos(0) = 1, J0 (0) = 1, and limit of sin(x)/x is also 1, Eq. 7.24 becomes:
at the centre of plane wall, cylinder and sphere:
T0 - Ta - l2 × Fo
Centre of plane wall: q0 = = A1 × e 1 ...(7.25, a)
(x = 0) Ti Ta
-

T0 - Ta - l2 × Fo
Centre of long cylinder: q0 = = A1 × e 1 ...(7.25, b)
(r = 0) Ti - Ta

T0 - Ta - l2 × Fo
Centre of sphere: q0 = = A1 × e 1 ...(7.25, c)
(r = 0) Ti - Ta

Therefore, first step in the solution is to calculate the Biot number; once the Biot number is known, constants
A1 and l1 are found out from Tables 7.1 and 7.2, and then use relations given in Eqs. 7.24 and 7.25 to calculate the
temperature at any desired location.
The one-term solutions are presented in chart form in the next section. But, generally, it is difficult to read
charts accurately. So, relations given in Eqs. 7.24 and 7.25 along with Tables 7.1 and 7.2, should be preferred to
the charts.
Calculation of amount of heat transferred, Q:
Many times, we need to calculate the amount of heat lost (or gained) by the body, Q, during the time interval t =
0 to t = t, i.e. from the beginning up to a given time. Again, we non-dimensionalise Q by dividing it by Qmax, the
maximum possible heat transfer. Obviously, maximum amount of heat has been transferred when the body has
reached equilibrium with the ambient, i.e.
Qmax = r × V × Cp ×(Ti – Ta) = m ×C p× (Ti – Ta) ...(7.26)
where r is the density, V is the volume, (rV) is the mass, Cp is the specific heat of the body.
If Qmax is positive, body is losing energy; and if it is negative, body is gaining energy.
Based on the one-term approximation discussed above, (Q/Qmax) is calculated for the three cases, from the
following:
Q sin (l 1 )
Plane wall: = 1 – q0 × ...(7.27, a)
Qmax l1

Q J (l )
Cylinder: = 1 – 2 × q0 × 1 1 ...(7.27, b)
Qmax l1

Q F sin(l ) - l ×cos(l ) I
Sphere:
Qmax
= 1 – 3×q 0 × GH 1
l 3
1
1 1
JK ...(7.27, c)

Note:
(i) Remember well the definition of Biot number i.e. Bi = (hL/k), where L is half-thickness of the slab, and Bi
= (hR/k), where R is the outer radius of the cylinder or the sphere.
(ii) Foregoing results are equally applicable to a plane wall of thickness L, insulated on one side and
suddenly subjected to convection at the other side. This is so because, the boundary condition dT/dx = 0

TRANSIENT HEAT CONDUCTION 283


TABLE 7.1 Transient heat conduction in a plane wall, long cylinder and sphere-coefficients for one-term
approximation

Plane wall Cylinder Sphere


Bi, l1, A1 l1 A1 l1 A1
0.01 0.0998 1.0017 0.1412 1.0025 0.1730 1.0030
0.02 0.1410 1.0033 0.1995 1.0050 0.2445 1.0060
0.04 0.1987 1.0066 0.2814 1.0099 0.3450 1.0120
0.06 0.2425 1.0098 0.3438 1.0148 0.4217 1.0179
0.08 0.2791 1.0130 0.3960 1.0197 0.4860 1.0239
0.1 0.3111 1.0161 0.4417 1.0246 0.5423 1.0298
0.2 0.4328 1.0311 0.6170 1.0483 0.7593 1.0592
0.3 0.5218 1.0450 0.7465 1.0712 0.9208 1.0880
0.4 0.5932 1.0580 0.8516 1.0931 1.0528 1.1164
0.5 0.6533 1.0701 0.9408 1.1143 1.1656 1.1441
0.6 0.7051 1.0814 1.0184 1.1345 1.2644 1.1713
0.7 0.7506 1.0918 1.0873 1.1539 1.3525 1.1978
0.8 0.7910 1.1016 1.1490 1.1724 1.4320 1.2236
0.9 0.8274 1.1107 1.2048 1.1902 1.5044 1.2488
1.0 0.8603 1.1191 1.2558 1.2071 1.5708 1.2732
2.0 1.0769 1.1785 1.5995 1.3384 2.0288 1.4793
3.0 1.1925 1.2102 1.7887 1.4191 2.2889 1.6227
4.0 1.2646 1.2287 1.9081 1.4698 2.4556 1.7202
5.0 1.3138 1.2403 1.9898 1.5029 2.5704 1.7870
6.0 1.3496 1.2479 2.0490 1.5253 2.6537 1.8338
7.0 1.3766 1.2532 2.0937 1.5411 2.7165 1.8673
8.0 1.3978 1.2570 2.1286 1.5526 2.7654 1.8920
9.0 1.4149 1.2598 2.1566 1.5611 2.8044 1.9106
10.0 1.4289 1.2620 2.1795 1.5677 2.8363 1.9249
20.0 1.4961 1.2699 2.2880 1.5919 2.9857 1.9781
30.0 1.5202 1.2717 2.3261 1.5973 3.0372 1.9898
40.0 1.5325 1.2723 2.3455 1.5993 3.0632 1.9942
50.0 1.5400 1.2727 2.3572 1.6002 3.0788 1.9962
100.0 1.5552 1.2731 2.3809 1.6015 3.1102 1.9990
¥ 1.5708 1.2732 2.4048 1.6021 3.1416 2.0000

at x = 0 for the mid-plane of a slab of thickness 2L (see Eq. 7.23, b), is equally applicable to a slab of
thickness L, insulated at x = 0.
(iii) These results are also applicable to determine the temperature response of a body when temperature of
its surface is suddenly changed to Ts. This case is equivalent to having convection at the surface with a
heat transfer coefficient,
h = ¥; now, Ta is replaced by the prescribed surface temperature, Ts.
(iv) Again, remember that these results are valid for the situation where Fourier number, Fo > 0.2.
7.6.2 Heisler and Grober Charts
The one term approximation solutions (Eq. 7.25) were represented in graphical form by Heisler in 1947. They
were supplemented by Grober in 1961, with graphs for heat transfer Eq. 7.27. These graphs are shown in Fig. 7.7,
7.8 and 7.9, for plane wall, long cylinder and a sphere, respectively.

284 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 7.2 Zeroth and first order Bessel functions of the first kind
x := 0, 0.1, ..., 3.2 ...define range variable x from 0 to 3.2, with an increment of 0.1

x J0(x) J1(x)
0 1 0
0.1 0.9975 0.04994
0.2 0.99002 0.0995
0.3 0.97763 0.14832
0.4 0.9604 0.19603
0.5 0.93847 0.24227
0.6 0.912 0.2867
0.7 0.8812 0.329
0.8 0.84629 0.36884
0.9 0.80752 0.40595
1 0.7652 0.44005
1.1 0.71962 0.4709
1.2 0.67113 0.49829
1.3 0.62009 0.52202
1.4 0.56686 0.54195
1.5 0.51183 0.55794
1.6 0.4554 0.5699
1.7 0.39798 0.57777
1.8 0.33999 0.58152
1.9 0.28182 0.58116
2 0.22389 0.57762
2.1 0.16661 0.56829
2.2 0.11036 0.55596
2.3 0.05554 0.53987
2.4 0.00251 0.52019
2.5 – 0.04838 0.49708
2.6 – 0.0968 0.47082
2.7 – 0.14245 0.4416
2.8 – 0.18504 0.40971
2.9 – 0.22431 0.37543
3 – 0.26005 0.33906
3.1 – 0.29206 0.30092
3.2 – 0.32019 0.26134

How to use these charts?


First chart in each of these figures gives the non-dimensionalised centre temperature T0. i.e. at x = 0 for the slab
of thickness 2L, and at r = 0 for the cylinder and sphere, at a given time t. Temperature at any other position at
the same time t, is calculated using the next graph, called ‘position correction chart’. Third chart gives Q/Qmax.
Procedure of using these charts to solve a numerical problem is as follows:
(i) First of all, calculate Bi from the data, with the usual definition of Bi, i.e. Bi = (h.Lc)/k, where Lc is the
characteristic dimension, given as: Lc = (V/A) i.e. Lc = L, half-thickness for a plane wall, Lc = R/2 for a
cylinder, and Lc= R/3 for a sphere. If Bi < 0.1, use lumped system analysis; otherwise, go for one-term
approximation or chart solution.

TRANSIENT HEAT CONDUCTION 285


1.0
0.7
0.5
0.4
0.3
0.2 14
1. 12
0
0.1 10
50

0. 0
T(0, t) – Ta

8 .7 0 0 0. 0. 0.2
0.07 9
Ti – Ta

7 45
0.05
0.04 40

.6 .5 4 3

6
0.03 35 h×L
Bi =

5
0.02 k
qo =

30
4
0.01

3
0.007

25
2.5
0.1 0.0 0

0.005
0.004

2.0
1.4 1.2

20 18

10
0.003

80 70

0
1
1.6
5

90
0.002 Bi
1.8

60
16
0.001
0 1 2 3 4 8 12 16 20 24 28 40 60 80 100 120 140 200 300 400 500 600
a×t
Fo = 2
L

x/L = 0.2
1.0 1.0
0.9 0.9 hL
0.4
0.8 Bi = k
T(0, t) – Ta
T(x, t) – Ta

0.8
0.5 0.7 1
0.7
0.00
Q/Qmax

0.6 2
5
0.01
0.02
0.05
0.00
0.00

0.1
0.6 0.6

0.2
0.5

10
20

50
1
2
5
0.5 0.7
Bi =

0.5
0.4 0.8 0.4
q=

0.3 0.3
0.2 0.9 0.2
0.1 1.0 0.1
0 0 –5 –4 –3 –2 –1 2 3 4
0.01 0.1 1.0 10 100 10 10 10 10 10 1 10 10 10 10
–1 k 2
2
h a×t
Bi = (Bi) (Fo) =
h×L 2
k
(b) (c)
FIGURE 7.7 Dimensionless transient temperatures and heat flow in an infinite plate of width 2L

(ii) If Bi > 0.1, i.e. if we have to go for one-term approximation or chart solution, calculate the Biot number
again with the appropriate definition, i.e. Bi =(hL/k) for a plane wall where L is half-thickness, and Bi =
(hR/k) for a cylinder or sphere, where R is the outer radius. Also, calculate Fourier number, Fo = a.t/L2
for the plane wall, and Fo = a . t/R2 for a cylinder or sphere.
(iii) To calculate the centre temperature, use chart (a) from Figs. 7.7, 7.8 or 7.9, depending upon the geometry
being considered. Enter the chart on the x-axis with the calculated Fo and draw a vertical line to intersect
the (1/Bi) line; from the point of intersection, move horizontally to the left to the y-axis to read the value
of qo = (To – Ta)/(Ti – Ta). Here, To is the centre temperature, which can now be calculated since Ti and Ta
are known.
(iv) To calculate the temperature at any other position, use Fig. b of Fig. 7.7, 7.8 or 7.9, as appropriate. Enter
the chart with 1/Bi on the x-axis, move vertically up to intersect the (x/L) or (r/R) curve as the case may
be, and from the point of intersection, move to the left to read on the y-axis, the value of q = (T –Ta)/(To
– Ta). Here, T is the desired temperature at the indicated position. We multiply q and qo to get:
F T -T I×FT -T I = T - T
i.e. q×q0 = GH T - T JK GH T - T JK T - T
0
a
a
0
i
a
a i
a
a
...(7.28)

286 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1.0
0.7
0.5 h×R
0.4 Bi =
k
0.3 5.0
0.2 3.0 4.
0
2.5 25
0.1 2. 20
T(0, t) – Ta

0 18
Ti – Ta

0.07 1. 1 10
0.05 6 146 90 0
0.01 80

1. .2 0
0.03 12

4
70

1 1.
0.02
qo =

10 9
60
0.01

8
0.8 .7

7
0.007
0

50 45
0.6 .5

6
0.005 1
0 0.4

0.004

40
0.003 Bi
0.3

35
0.20.1

0.002

30
0

0.001
0 1 2 3 4 8 12 16 20 24 28 40 60 80 100 120 140 200 300
a×t
Fo = 2
R
(a)

1.0 1.0
0.9 0.9
0.8 0.4 0.8
T(0, t) – Ta

1
T(r, t) – Ta

0.7 0.5
0.00

0.7
Q/Qmax

0.6 0.6 0.6


2
0.015
Bi =

0.02
0.05

0.5
0.00
0.00

0.5 0.7
0.1
0.2
0.5

10
20

30
1
2
5
0.4 k 0.4
0.8 Bi - 1 =
q=

0.3 h×R 0.3


0.2 0.9 0.2
0.1 1.0 0.1
0 0 –5 –4 –3 –2 –1 2 3 4
0.01 0.1 1.0 10 100 10 10 10 10 10 1 10 10 10 10
k 2
h ×a×t
Bi - 1 = 2
(Bi) (Fo) =
h×R 2
k
(b) (c)

FIGURE 7.8 Dimensionless transient temperatures and heat flow for a long cylinder

From Eq. 7.28, we can easily calculate T, the desired temperature at the given position, since Ti and Ta are
known.
(v) To find out the amount of heat transferred Q, during a particular time interval t from the beginning (i.e.
t = 0), use Fig. c from Figs. 7.7, 7.8 or 7.9, depending upon the geometry. Enter the x-axis with the value
of (Bi2. Fo) and move vertically up to intersect the curve representing the appropriate Bi, and move to the
left to read on the y-axis, the value of Q/Qmax. Q is then easily found out since Qmax = mCp(Ti – Ta). And,
Q = (Q/Qmax ). Qmax.
Note the following in connection with these charts:
(i) These charts are valid for Fourier number Fo > 0.2.
(ii) Specifically, remember that while calculating Biot number, characteristic length (Lc) used is L, the half-
thickness for a plane wall, and outer radius, R for the cylinder and the sphere (Lc is, now, not equal to:
(V/A)).

TRANSIENT HEAT CONDUCTION 287


1.0
0.7
0.5
0.4
0.3
0.2 3.5 14
3. 12
0
0.1 2. 10
Ti – Ta

8
q(0, t)

50 4
0.07 2.
6
9
0.05 8

5 40 35
0.04

2. 2.2 .0
7

4
0.03
qo =

6
0.02

2 .8 1.6 .4 1.2
1

30
5
0.01

25
1
0.007

100 90
0.005

20
0.004 1.0 1
0.7

18
0.003
0 .5

5
0.35

Bi
0.05 0

80 0
16
0.002
0.2 0.1

7
60
0.001
0 0.1 1.0 1.5 2.0 2.5 3 4 5 6 7 8 9 10 20 30 40 50 90 130 170 210 250
a×t
Fo = 2
R
(a)

r/r0 = 0.2
1.0 1.0
0.9 0.9 Bi = h × R
0.8 0.4 k
T(0, t) – Ta

0.8
T(r, t) – Ta

0.7 0.5 0.7


1
Q/Qmax

0.00

2
5

0.6 0.6
0.01
0.02
0.05
0.00
0.00

0.1
0.6
0.2
0.5

10

20

50
1
2
5
0.5 0.5
Bi =

0.7
0.4 0.4
q=

0.3 0.8 0.3


0.2 0.9 0.2
0.1 0.1
0 1.0 0 –5
0.01 0.1 1.0 10 100 10 10
–4
10
–3
10
–2
10
–1
1 10 10
2
10
3
10
4

–1 k 2
h ×a×t
Bi = 2
h×R (Bi) (Fo) = 2
k
(b) (c)

FIGURE 7.9 Dimensionless transient temperatures and heat flow for a sphere

(iii) In these graphs, (1/Bi) = 0, corresponds to h ® ¥, which means that at t = 0, the surface of the body is
suddenly brought to a temperature of Ta and thereafter maintained at Ta at all times.
(iv) To calculate Q up to a given time, first find out Q/Qmax from the Grober’s chart and calculate Qmax from
Qmax = mCp (Ti – Ta). (See Eq. 7.26). Then, Q is calculated as: Q = (Q/Qmax).Qmax.
(v) Note from the ‘position correction charts’ that at Bi < 0.1 (i.e. 1/Bi > 10), temperature within the body can
be taken as uniform, without introducing an error of more than 5%. This was precisely the condition for
application of ‘lumped system analysis’.
(vi) As stated earlier, it is difficult to read these charts accurately, since logarithmic scales are involved; also,
the graphs are rather crowded with lines. So, use of one-term approximation with tabulated values of A1
and l1 should be preferred. However, these graphs are extremely useful for a quick estimation of values
required.

288 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Use of one-term approximation solutions and the transient conduction charts is illustrated in the following
examples.
Example 7.7. A steel plate (a = 1.2 ´ 10 –5 m2/s, k = 43 W/(mC)), of thickness 2L = 10 cm, initially at a uniform tempera-
ture of 250°C is suddenly immersed in an oil bath at Ta = 45°C. Convection heat transfer coefficient between the fluid
and the surfaces is 700 W/(m2C).
(i) How long will it take for the centre plane to cool to 100°C?
(ii) What fraction of the energy is removed during this time?
(iii) Draw the temperature profile in the slab at different times.
Solution.
Data:
L := 0.05 m a := 1.2 × 10–5 m2/s k := 43 W/(mC) Ti := 250°C Ta := 45°C
2
h := 700 W/(m C) T0 := 100°C
To calculate: the time t, surface temperature and fraction of heat transferred Q/Qmax.
First, check if lumped system analysis is applicable:
h×L
Bi := (define Biot number)
k
i.e. Bi = 0.814 (Biot number.)
It is noted that Biot number is > 0.1; so, lumped system analysis is not applicable.
We will adopt Heisler chart solution and then check the results from one-term approximation solution.
To find the time required for the centre to reach 100°C:
For using the charts, Bi = hL/k, which is already calculated.
a ×t
Fourier number: Fo =
L2
Centre temperature is given as 100°C. Therefore, calculate q 0:
T0 - Ta
q0 := (define q0)
Ti - Ta
i.e. q0 = 0.268 (value of q0)
1
Also, = 1.229 (value of 1/Bi)
Bi
Now, with this value of q0, enter the y-axis of Fig. 7.7,a. Move horizontally to intersect the 1/Bi = 1.229 line; from
the point of intersection, move vertically down to x-axis to read Fo = 2.4.
So, we get: Fo := 2.4
Fo ×L2
Then, t := s (define t, the time required for the centre to reach 100°C)
a
i.e. t = 500s (time required for the centre to reach 100°C.)
Surface temperature:
At the surface, x/L = 1. Enter Fig. 7.7, b on the x-axis with a value of 1/Bi = 1.229, move up to intersect the curve of
x/L = 1, then move to left to read on y-axis the value of q = 0.7
T - Ta
i.e. q= = 0.7
T0 - Ta
Therefore, T := 0.7 × (T0 – Ta) + Ta °C (temperature on the surface)
i.e. T = 83.5°C (temperature on the surface.)
Fraction of maximum heat transferred, Q/Qmax:
We will use Grober’s chart, Fig. 7.7, c:
We need Bi2Fo to enter the x-axis:
We get: Bi2 × Fo = 1.59
With the value of 1.59, enter the x-axis of Fig. 7.7, c, move vertically up to intersect the curve of Bi = 0.814, then
move horizontally to read Q/Qmax = 0.8
Q
i.e. from Fig. 7.7c, we get: = 0.8
Qmax
i.e. 80% of the energy is removed by the time the centre tamperature reached 100°C.

TRANSIENT HEAT CONDUCTION 289


Verify by one-term approximation solution:
Time required for the centre to reach 100°C:
From Eq. 7.25, a, we have:
T0 - Ta - l2 × Fo
Centre of plane wall: q0 = = A1× e 1 ...(7.25, a)
Ti - Ta
(x = 0)
A1 and l1 have to be found from Table 7.1, against Bi = 0.814
0.8274 - 0.7910
Interpolating: l1 = 0.7910 + × 1.4
10
i.e. l1 = 0.796
1.1107 - 1.1016
and, A1 = 1.1016 + × 1.4
10
i.e. A1 = 1.103
T0 - Ta
Now, = 0.26829
Ti - Ta
Therefore, Eq. 7.25, a becomes:
0.26829 2
⋅ Fo
= e −0.796
1.103
FG 0.268 IJ
i.e. Fo :=
- ln
H 1.103 K
( 0.796 ) 2
i.e. Fo = 2.233
Fo ×L2
Then, t :=s (define t, the time required for the centre to reach 100°C)
a
i.e. t = 465.188 s (time required for the centre to reach 100°C.)
Compare this value with the one got from Heisler’s chart, i.e. 500 s. The error is in reading the chart.
Surface temperature when the centre has reached 100°C:
From Eq. 7.24, a, we have:
T (x, t ) - Ta - l2 × Fo
= A1× e 1 × cos
l 1 ×x FG IJ
q (x, t ) =
Ti - Ta L H K ...Fo > 0.2 ...(7.24, a)

Here, x/L = 1, at the surface of the plate. So, we get:

e
T: =(Ti – Ta)× A1 × e - l
2
1 × Fo
j
×cos(l 1 ) + Ta (define T (x, t))
T = 83.413°C (temperature at the surface.)
Compare this with the value of 83.5°C obtained earlier. They are quite close.
Fraction of maximum heat transferred, Q/Qmax:
From Eq. 7.27, a, we have:
Q sin(l 1 )
Plane wall: = 1 – q 0× ...(7.27, a)
Qmax l1

T0 - Ta sin (l 1 )
i. e. Fraction: = 1 – × ...define Fraction, Q/Qmax
Ti - Ta l1
i.e. Fraction = 0.759
i.e. 75.9% of the energy is removed by the time the centre temperature has reached 100°C.
Compare this with the value of 80% obtained earlier; again, the error is in reading the charts.
Note: It is apparent from this example that the error involved in reading the graphs can be substantial; this is because
logarithmic scales are involved and also the lines are rather crowded in the graph. So, one-term approximation with
table of values of A1 and l1 against Bi should be preferred.

290 FUNDAMENTALS OF HEAT AND MASS TRANSFER


To draw temperature profile in the plate at different times:
We have, for temperature distribution at any location:
T (x, t ) - Ta × Fo FG l ×x IJ
H LK
2
Plane wall: q (x, t) = = A1 × e - l 1
× cos 1
...Fo > 0.2 ...(7.24, a)
Ti - Ta

T - Ta 2
× Fo
And, centre of plane wall: q 0 = 0 = A1 × e - l 1
...(7.25, a)
(x = 0) Ti - Ta

a ×t
Fourier number as a function of t : Fo(t) := ...for slab
L2
By writing Fourier number as a function of t, and including it in Eq. A below as shown, it is ensured that for each
new t, the corresponding new Fo is calculated.

e
Ta + (Ti - Ta )× A1 ×e - l
2
1 × Fo (t )
j if x = 0
Then, T(x, t) := F
T + (T - T )×G A ×e - l × Fo (t )
×cos G
F l ×x IJ IJ otherwise ...(A)
H L KK
2

H
1 1
a i a 1

For a given t, we will plot Eq. A against x; then, we will repeat for different times, t:
We use Mathcad to draw the graph. First, define a range variable x, varying from 0 to say, 0.05 m, with an
increment of 0.001 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-
axis with x and T (x,30), respectively. Since our aim is to plot T(x, t) for different values of x for given t, start with t = 30
s; immediately, this graph is drawn, when we click anywhere outside the graph region. To get the graph for next value
of t =120, on the y-axis, next to the earlier entry, type a comma and enter T(x,120) and click anywhere outside the graph
region. Repeat this for different values of t, as shown. See Fig. Example 7.7.
x := 0, 0.001, ... , 0.05 (define a range variable x varying from zero to
0.05 m, with an increment of 0.001 m)

Transient cooling of a large plate


300
280
260
240
220
Temperature (deg.C)

200
180
160
140
120
100
80
60
40
20
0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
Distance from centre line (m)
after 30 sec. after 2 min. after 5 min.
after 10 min. after 25 min.

FIGURE Example 7.7 Transient cooling of a large plate, one-term approximation solution

Note:
(i) Note that the above graph shows temperature distribution for one half of the plate; for the other half, the tem-
perature distribution will be identical.

TRANSIENT HEAT CONDUCTION 291


(ii) See from the above Fig. Example 7.7 how cooling progresses with time. After a time period of 25 min the
temperatures in the plate are almost uniform at 45°C.
(iii) Eq. A illustrates a small piece of Mathcad programming. It uses the “if...otherwise” condition, i.e. if x = 0, the
temperature at the centre is given by Eq.7.25, a; otherwise, temperature distribution is given by Eq. 7.24, a.
Example 7.8. A long, 15 cm diameter cylindrical shaft made of stainless steel 304 (k = 14.9 W/(mC), r = 7900 kg/m3,
Cp = 477 J/(kgC), and a = 3.95 ´ 10-6 m2/s), comes out of an oven at an uniform temperature of 450°C. The shaft is then
allowed to cool slowly in a chamber at 150°C with an average heat transfer coefficient of 85 W/(m2C).
(i) Determine the temperature at the centre of the shaft 25 min after the start of the cooling process.
(ii) Determine the surface temperature at that time, and
(iii) Determine the heat transfer per unit length of the shaft during this time period.
(iv) Draw the temperature profile along the radius for different times.
Solution.
Data:
L := 1 m R := 0.075 m a := 3.95 ´ 10–6 m2/s k := 14.9 W/(mC) Cp := 477 J/(kg C) r := 7900 kg/m3
Ti := 450°C Ta := 150°C h := 85 W/(m2C) t := 1500 s
To calculate: the centre temperature afer time t, surface temperature and amount of heat transferred during this
period.
First check if lumped system analysis is applicable:
R

Bi := 2 (define Biot number...for a cylinder, Lc = (V/A) = R/2)
k
i.e. Bi = 0.214 (Biot number.)
It is noted that Biot number is > 0.1; so, lumped system analysis is not applicable.
We will adopt Heisler chart solution and then check the results from one-term approximation solution.
To find the centre temperature after a time period of 1500 s:
For using the charts, now, remember that Bi is defined as:
h×R
Bi := (define Biot number)
k
i.e. Bi = 0.428 (Biot number)
a ×t
Fourier number: Fo := (define Fourier number)
R2
i.e. Fo = 1.053 (Fourier number)
1
Also, = 2.337 ...value of 1/Bi
Bi
Now, with the value of Fo = 1.053, enter the x-axis of Fig. 7.8,a. Move vertically up to intersect the 1/Bi = 2.337 line;
from the point of intersection, move horizontally to left, to read on the y-axis q 0 = 0.49.
T0 - Ta
So, we get: q0 = 0.49 =
Ti - Ta
i.e. T0 := Ta + 0.49 × (Ti – Ta) (define centre temperature)
i.e. T0 = 297°C (centre temperature after 25 min duration.)
Surface temperature:
At the surface, r/R = 1. Enter Fig. 7.8, b on the x-axis with a value of 1/Bi = 2.337, move up to intersect the curve of
r/R = 1, then move to left to read on y-axis th value of q = 0.76
T - Ta
i.e. q= = 0.76
T0 - Ta
Therefore, T := 0.76×(T0 – Ta) + Ta°C (Temperature on the surface)
i.e. T = 261.72°C (temperature on the surface.)
Amount of heat transferred, Q:
We will use Grober’s chart, Fig. 7.8, c:
We need Bi2Fo to enter the x-axis:
We get: Bi2×Fo = 0.193

292 FUNDAMENTALS OF HEAT AND MASS TRANSFER


With the value of 0.193, enter the x-axis of Fig. 7.8, c, move vertically up to intersect the curve of Bi = 0.428, then
move horizontally to read Q/Qmax = 0.55
Q
i.e. from Fig. 7.8, c, we get: = 0.55
Qmax
Now, Qmax = r ×V × Cp × (Ti – Ta) (maximum heat transfer possible)
i.e. Qmax := r×(p×R2×L)×Cp×(T1 – Ta) (define Qmax)
i.e. Qmax = 1.998 ´ 107 J (maximum heat transfer)
Therefore, Q = Q max × 0.55 J (define Q)
i.e. Q = 1.099 ´ 107 J (heat transferred during 25 min)
Verify by one-term approximation solution:
Centre temperature reached after 25 min:
From Eq. 7.25, b, we have:
T - Ta - l2 × Fo
Centre of long cylinder: q0 = 0 = A1 × e 1 ...(7.25, b)
(r = 0) Ti - Ta

A1 and l1 have to be found from Table 7.1, against Bi = 0.428


0.9408 - 0.8516
Interpolating: l1 := 0.8516 + × 2.8
10
i.e. l1 = 0.877
1.1143 - 1.0931
and, A1 := 1.0931 + × 2.8
10
i.e. A1 = 1.099
- l21 × Fo
Therefore, q 0 := A1 × e ...(7.25, b)
i.e. q 0 = 0.489 (dimensionless centre tamperature)
Then, again from Eq. 7.25, b:
T0 := Ta + 0.489 × (Ti – Ta) (define centre temperature)
i.e. T0 = 296.7°C ...centre temperature of cylinder after 25 min.
Note that this value compares well with the value of 297°C obtained by reading Heisler charts.
Surface temperature after 25 min:
From Eq. 7.24, b, we have:
T ( r , t ) - Ta × Fo FG l ×r IJ ...Fo > 0.2
HRK
2
Long cylinder: q (x, t) = = A1 × e - l 1
× J0 1
...(7.24, b)
Ti - Ta
In Eq. 7.24, b, J0 is the zeroth order Bessel function of the first kind. Its value can be read from Table 7.2. However,
while using Mathcad, J0 can be got directly by typing ‘J0 (l1)=’.
i.e. J0(l1) = 0.817
And, while using Mathcad, it is not even necessary to separately obtain the value of J0(l1).
See below the expression for T. While calculating the expression for T, value of J0(l1) is returned and substituted
automatically, and we get the final value of T as shown.
Here, r/R = 1, at the surface of the cylinder. So, we get:

e
T := (Ti – Ta)× A1 × e - l
2
1 × Fo
j
× J o ( l 1 ) + Ta (define T(x, t ))
i.e. T = 269.899°C (temperature at the surface.)
Compare this with the value of 261.72°C obtained earlier from the charts. The error is in reading the charts.
Amount of heat transferred, Q:
From Eq. 7.27, b we have:
Q J (l )
Cylinder: =1 – 2 × q0 × 1 1 ...(7.27, b)
Qmax l1

T0 - Ta J1 (l 1 )
i.e. Fraction : = 1 – 2× × (define Fraction, Q/Qmax)
Ti - Ta l1
i.e. Fraction = 0.556

TRANSIENT HEAT CONDUCTION 293


Now, Qmax = 1.998 ´ 107 J (already calculated)
Therefore, Q = Qmax × 0.556 J (heat transferred in 25 min)
i.e. Q = 1.111 ´ 107 J (heat transferred in 25 min.)
Note again that this value of Q is quite close to that obtained from Grober’s chart.
To draw radial temperature distribution at different times:
Let us draw radial temperature distribution at t = 15 min, 25 min, 1 hr., etc.
We have, for temperature distribution at any location:
T (r , t ) - Ta × Fo FG l ×r IJ
HRK
2
Long cylinder: q (x, t) = = A1 × e - l 1
× J0 1
...Fo > 0.2 ...(7.24, b)
Ti - Ta

T0 - Ta - l2 × Fo
Centre of long cylnder: q0 = = A1 × e 1 ...(7.25, b)
(r = 0) Ti - Ta

a ×t
Fourier number as a function of t : Fo(t ) := ...for cylinder
R2

e
Ta + (Ti - Ta )× A1 ×e - l
2
1 × Fo ( t )
j if r = 0
Then, T(r, t) := F
T + (T - T )×G A ×e - l21 × Fo (t )
×J G
F l ×r IJ IJ otherwise ...(A)
H H R KK
1
a i a 1 0

For a given t, we will plot Eq. A against r; then, we will repeat for different times, t:
We use Mathcad to draw the graph. First, define a range variable r, varying from 0 to say, 0.075 m, with an
increment of 0.001. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis
with r and T(r, 900), respectively. Since our aim is to plot for different values of r for given t, start with t = 900 s;
immediately, this graph is drawn, when we click anywhere outside the graph region. To get the graph for next value of
t =1500, on the y-axis, next to the earlier entry, type a comma and enter T(r,1500) and click anywhere outside the graph
region. Repeat this for different values of t as shown. See Fig. Ex. 7.8.
r := 0, 0.001, ... , 0.075 (define a range variable r varying from zero
to 0.075 m, with an increment of 0.001 m)

Transient cooling of a cylinder


400
375
350
Temperature (deg.C)

325
300
275
250
225
200
175
150
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Radial distance from centre (m)
after 15 min. after 25 min.
after 1 hr. after 2 hrs.
after 3 hrs.

FIGURE Example 7.8 Transient cooling of a cylinder, one-term approximation solution

Note:
(i) See from the figure how cooling progresses with time. After a time period of 2 hrs the temperatures along the
radius are almost uniform, but is yet to reach ambient temperature of 150°C. After about 3 hrs; the body has
almost come to equilibrium with the ambient
294 FUNDAMENTALS OF HEAT AND MASS TRANSFER
(ii) Eq. A illustrates a small piece of Mathcad programming. It uses the “if...otherwise” condition, i.e. if r = 0, the
temperature at the centre is given by Eq. 7.25, b; otherwise, temperature distribution is given by Eq. 7.24, b.
(iii) Observe from the graph that after 25 min, temperature at the centre (r = 0) is 296.7°C and at the surface (r = 0.075
m), the temperature is 269.9°C as already calculated.
Example 7.9. An apple, which can be considered as a sphere of 8 cm diameter, is initially at a uniform temperature of
25°C. It is put into a freezer at –15°C and the heat transfer coefficient between the surface of the apple and the surround-
ings in the freezer is 15 W/(m2C). If the thermophysical properties of apple are given to be: r = 840 kg/m3, Cp = 3.6 kJ/
(kgC), k = 0.513 W/(mC), and a = 1.3 ´ 10–7 m2/s, determine:
(i) centre temperature of the apple after 1 hour,
(ii) surface temperature of apple at that time, and
(iii) amount of heat transferred from the apple.
(iv) draw the temperature profile along the radius for different times.
Solution.
Data:
R := 0.04 m a := 1.3 × 10–7 m2/s k := 0.513 W/(mC) Cp := 3600 J/(kgC) r := 840 kg/m3 Ti := 25°C
2
Ta := – 15°C h := 15 W/(m C) t := 3600 s
To calculate: the centre temperature after time t, surface temperature and amount of heat transferred during this
period.
First check if lumped system analysis is applicable:
R

Bi = 3 (define Biot number...for a sphere, Lc = (V/A) = R/3)
k
i.e. Bi = 0.39 (Biot number)
It is noted that Biot number is > 0.1; so, lumped system analysis is not applicable.
We will adopt Heisler chart solution and then check the results from one-term approximation solution.
To find the centre temperature after a time period of 3600 s:
For using the charts, now, remember that Bi is defined as:
h×R
Bi := (define Biot number)
k
i.e. Bi = 1.17 (Biot number)
a ×t
Fourier number: Fo := (define Fourier number)
R2
i.e. Fo = 0.292 (Fourier number)
1
Also, = 0.855 (value of 1/Bi.)
Bi
Now, with the value of Fo = 0.292, enter the x-axis of Fig. 7.9,a. Move vertically to intersect the 1/Bi = 0.855 line;
from the point of intersection, move horizontally to left, to read on the y-axis q 0 = 0.45
So, we get:
T0 - Ta
q 0 = 0.45 =
Ti - Ta
i.e. T0 := Ta + 0.45 × (Ti – Ta) (define centre temperature)
i.e. T0 = 3°C (centre temperature after 1 hr duration.)
Surface temperature:
At the surface, r/R =1. Enter Fig. 7.9, b on the x-axis with a value of 1/Bi = 0.855, move up to intersect the curve of r/R
= 1, then move to left to read on y-axis the value of q = 0.6
T - Ta
i.e. q= = 0.6
T0 - Ta
Therefore, T := 0.6 × (T0 – Ta) + Ta °C (temperature on the surface)
i.e. T = – 4.2°C (temperature on the surface.)
Amount of heat transferred, Q:
We will use Grober’s chart, Fig. 7.9, c:

TRANSIENT HEAT CONDUCTION 295


We need Bi2Fo to enter the x-axis:
We get Bi2 × Fo = 0.4
With the value of 0.4, enter the x-axis of Fig. 7.9, c, move vertically up to intersect the curve of Bi = 1.17, then move
horizontally to read Q/Qmax = 0.56
Q
i.e. from Fig. 7.9, c, we get: = 0.56
Qmax
Now, Qmax = r ×V × Cp × (T1 – Ta) (maximum heat transfer possible)
F4 I
:= r × G ×p ×R J × C × (T – T )
Qmax
H3 K (define Q max)
3
i.e. p i a

i.e. Qmax = 3.243 ´ 104 J (maximum heat transfer)


Therefore, Q = Qmax ´ 0.56 J (define Q)
i.e. Q = 1.816 ´ 104 J (heat transferred during 1 hr.)
Verify by one-term approximation solution:
Centre temperature reached after 1 hr:
From Eq. 7.25, c, we have:
T0 - Ta 2
× Fo
Centre of sphere: q0 = = A1× e - l 1
...(7.25, c)
(r = 0) Ti - Ta

A1 and l1 have to be found from Table 7.1, against Bi = 1.17


2.0288 - 1.5708
Interpolating: l1 := 1.5708 + × 1.7
10
i.e. l1 = 1.649
1.4793 - 1.2732
and, A1 := 1.2732 + × 1.7
10
i.e. A1 = 1.308
2
× Fo
Therefore, q0: = A1× e - l 1
...from Eq. (7.25, c)
i.e. q0 = 0.591 ...dimensionless centre temperature
Then, again, from Eq. 7.25, c:
T0: = Ta + 0.591× (Ti – Ta) ...define centre temperature
i.e. T0 = 8.64°C ...centre temperature of sphere after 1 hr.
Compare this value with the value of 3°C obtained by reading the graph; error is due to the error in reading the
graph.
Surface temperature after 1 hr:
From Eq. 7.24, c, we have:

FG l ×r IJ
HRK
1
sin
T ( r , t ) - Ta 2
× Fo
Sphere: q (x, t) = = A1 × e - l 1
× ...Fo > 0.2 ...(7.24, c)
Ti - Ta l 1 ×r
R
Here, r/R = 1, at the surface of the sphere. So, we get:
F b g IJ
sin l 1
GH × Fo
2

T: =(Ti – Ta)× A1 × e - l × + Ta ...define T (x, t)


K
1

l1
i.e. T = – 0.711°C ...temperature at the surface.
Compare this value with the value of – 4.2°C obtained by reading the graph; error is due to the error in reading the
graph.
Amount of heat transferred, Q:
From Eq. 7.27, c, we have:

Q F
sin(l 1 ) - l 1 ×cos(l 1 ) I
Sphere:
Qmax
= 1 – 3× q0 ×
l31 GH JK ...(7.27, c)

296 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F
T0 - Ta sin(l 1 ) - l 1 ×cos(l 1 ) I
i.e. Fraction := 1 – 3×
Ti - Ta
× GH l31 JK (define Fraction, Q/Qmax)

i.e. Fraction = 0.555


Now, Qmax = 3.243 ´ 10 4 J (already calculated)
Therefore, Q = Qmax × 0.555 J (heat transferred in 1 hr)
i.e. Q = 1.8 ´ 104 J (heat transferred in 1 hr.)
Note again that this value of Q compares well with 1.811 ´ 104 J, obtained from Grober’s chart.
To draw temperature profile along the radius at different times:
We have, for temperature distribution at any location:

FG l ×r IJ
HRK
1
sin
T ( r , t ) - Ta 2
× Fo
q (x, t) = = A1× e - l 1
× ...Fo > 0.2 ...(7.24, c)
Ti - Ta l 1 ×r
R

T0 - Ta 2
× Fo
And, at centre of sphere: q0 = = A1× e - l 1
...(7.25, c)
Ti - Ta
a ×t
Fourier number as a function of t : Fo (t ) := (for sphere)
R2

e
Ta + (T1 - Ta )× A1 ×e - l
2
1 × Fo (t )
j if r = 0
LM F l ×r IJ OP
sin G
H R K P otherwise
1

T + (T - T )× M A ×e
Then, T(r, t ) := ...(A)
MM PP
- l 1 × Fo (t )
2

a ×
1 1 1
l ×r1

N R Q
For a given t , we will plot Eq. A against r; then, we will repeat for different times, t:
We use Mathcad to draw the graph. First, define a range variable r, varying from 0 to say, 0.04 m, with an
increment of 0.001. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-axis
with r and T(r, 1800), respectively. Since our aim is to plot for different values of r for given t, start with t = 1800 s;
immediately, this graph is drawn, when we click anywhere outside the graph region. To get the graph for next value of
t =3600, on the y-axis, next to the earlier entry, type a comma and enter T(r, 3600) and click anywhere outside the graph
region. Repeat this for different values of t as shown. See Fig. Ex. 7.9.
r := 0, 0.001, ... , 0.04 (define a range variable r varying from zero to
0.04 m, with an increment of 0.001 m)
Note:
(i) See from the Fig. Example 7.9 how cooling progresses with time. After a time period of 5 hrs the temperature
along the radius is almost uniform, but is yet to reach ambient temperature of – 15°C.
(ii) Eq. A illustrates a small piece of Mathcad programming. It uses the “if...otherwise” condition, i.e. if r = 0, the
temperature at the centre is given by Eq. 7.25, c; otherwise, temperature distribution is given by Eq. 7.24, c.
Example 7.10. A large concrete slab, one side of which is insulated, is 60 cm thick and is initially at 70°C. The other side
is suddenly exposed to hot combustion gases at 1000°C with a heat transfer coefficient of 30 W/(m2C). Determine:
(i) time required for the insulated surface to reach 500°C
(ii) temperature distribution in the wall at that instant
(iii) amount of heat transferred during that time period.
Take average properties of concrete as follows:
r = 500 kg/m3, Cp = 837 J/(kgC), k = 1.25 W/(mC), and a = 0.3 ´ 10 –5 m2/s.
Solution.
Data:
L := 0.6 m a := 0.3 × 10–5 m2/s k := 1.25 W/(mC) Cp := 837 J/(kgC) r := 500 kg/m3 Ti := 70°C
Ta := 1000°C h := 30 W/(m2C) T0 := 500°C
To calculate: the time t , at which temperature of insulated surface will reach 500°C, temperature distribution in the
slab at that instant, and amount of heat transferred during this period.

TRANSIENT HEAT CONDUCTION 297


Transient cooling of a spherical apple
30
25
20
15

Temperature (C)
10
5
0
5
10
15
20
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
Radial distance from centre (m)
after 30 min. after 1 hr.
after 2 hrs. after 3 hrs.
after 5 hrs.

FIGURE Example 7.9 Transient cooling of a sphericl apple, one-term approximation solution

Time required for the insulated surface to reach 500°C:


First of all, recognise that boundary condition at the insulated surface is the same as at the midplane of a slab of half-
thickness, L i.e. dT/dx at x = 0 is zero.
Therefore, for the present case, we take the thickness of the slab as L.
We will solve this problem by one-term approximation solution:
From Eq. 7.25, a, we have:
T0 - Ta 2
× Fo
Centre of plane wall: q0 = = A1 × e - l 1
...(7.25, a)
(x = 0) Ti - Ta

Eq. 7.25, a is also valid for the insulated surface of a wall of thickness L, as explained above.
h ×L
Bi := (define Biot number)
k
i.e. Bi = 14.4 (Biot number.)
A1 and l1 have to be found from Table 7.1, against Bi = 14.4
1.4961 - 1.4289
Interpolating: l 1 := 1.4289 + × 4.4
10
i.e. l1 = 1.458
1.2699 - 1.2620
and, A1 := 1.2620 + × 4.4
10
i.e. A1 = 1.265
T0 - Ta
Now, = 0.53763
Ti - Ta
Therefore, Eq. 7.25, a becomes:
0.53763 2
⋅ Fo
= e −1.458
1.265

298 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG 0.53763 IJ
i.e. Fo :=
- ln
H 1.265 K
(1.458 )2
i.e. Fo = 0.403 (Fourier number)
Fo×L2
Then, t := s (define t, the time require for the insulated surface to reach 500°C)
a
4
i.e. t = 4.8302 ´ 10 s (time required for the insulated surface to reach 500°C.)
i.e. t = 13.417 hrs
To plot the temperature distribution in the slab when t = 13.417 hrs:
We have to draw temperature as a function of position (i.e. x) for given t of 13.417 hrs.
We use Eq. 7.24, a, i.e.
T (x, t ) - Ta × Fo FG l ×x IJ
H LK
2
Plane wall: q (x, t) = = A1 × e - l 1
× cos 1
...Fo > 0.2 ...(7.24, a)
Ti - Ta
Therefore, we write:
F FG l ×x IJ IJ
GH × Fo
H L KK
2
T(x, t ) := Ta + (Ti – Ta) × A1 × e - l 1
×cos 1

To plot the T(x, t) using Mathcad, we first define a range variable x from zero to L = 0.6 m, and then select the x–y
graph from the graph palette. On the place holder on the x-axis, fill in x and on the place holder on the y-axis, fill in T(x,
t). Click anywhere outside the graph region and the graph appears.
x := 0, 0.01, ... , 0.6 (range variable x from zero to 0.6 m with an
increment of 0.01 m)

Transient temperature distribution in insulated slab


1000
950
900
850
800
750
T(x, t) 700 x in metres,
temperature in deg.C
650
600
550
500
450
400
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6
x

FIGURE Example 7.10 Transient temperature distribution in an insulated slab at the instant t = 13.417 hrs

Remember: x is measured from the insulated surface.


Amount of heat transferred per unit surface area, Q:
From Eq. 7.27, a, we have:
Q sin (l 1 )
Plane wall: = 1 – q0 × ...(7.27, a)
Qmax l1

TRANSIENT HEAT CONDUCTION 299


Q T - Ta sin (l 1 )
i.e. =1 – 0 ×
Qmax Ti - Ta l1
Now, Qmax :=r × L ×Cp × (Ta – Ti) (Joules per unit surface area)
i.e. Qmax = 2.335 ´ 108 (Joules per unit surface area)
F T0 - Ta sin(l 1 ) I
Therefore, GH
Q := Qmax × 1 -
Ti - Ta
×
l1 JK J/m2 ...define Q

i.e. Q = 1.48 ´ 108 J/m2 (heat transferred per unit surface area during the process)
Positive value of Q indicates that heat is transferred into the slab.

7.7 One-dimensional Transient Conduction in Semi-infinite Solids


A solid which has one exposed surface and extends to infinity in other directions is known as a semi-infinite
solid. So, change in boundary condition at the exposed surface initiates temperature transients in the solid. One-
dimensional transient heat conduction in semi-infinite solids, without heat generation, is of interest because of
many practical applications. Common example is that of earth’s surface subjected to changes in the ambient
conditions, thus causing transient conditions in the soil at some depth from the surface; or, in the case of a thick
slab, when the exposed surface is subjected to a temperature variation, in the early stages when the effect is not
felt at the distant surface, it can be idealised as a semi-infinite solid, to solve the transient conditions near the
surface.
Consider a semi-infinite solid, extending from x = 0 to x = ¥, initially at a uniform temperature, Ti. There is
no internal energy generation. Now, if there is a change in the thermal conditions at the exposed surface at x = 0,
transient conditions will be induced in the solid. Fig. 7.10 illustrates three possible boundary conditions at the
surface:
Case (i): Constant surface temperature:
See Fig. 7.10 (a). The solid is initially at a uniform temperature Ti and for times t > 0, the boundary surface at x =
0 is maintained at temperature T o. Starting with the differential Eq. for one-dimensional, time dependent
conduction, for these boundary conditions, the non-dimensional temperature distribution in the solid is obtained
as:
F x I
T ( x , t ) - T0
Ti - T0
= erf GH
2 a ×t
JK ...(7.29)

where, erf (z ) is the Gaussian error function defined as:

Initially, solid at Ti Initially, solid at Ti


Heat flux,
T = To, for t > 0 qo, for t > 0

O X O X

(a) (b)

Initially, solid at Ti

Convection, for t > 0

Ta, h O X

(c)

FIGURE 7.10 One-dimensional transient conduction in semi-infinite solids

300 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 7.3 Values of ‘error function’

z1 erf (z1) x1 erf (x1)


0 0 1 0.8427
0.03 0.0338 1.05 0.8624
0.06 0.0676 1.1 0.8802
0.09 0.1013 1.15 0.8961
0.12 0.1348 1.2 0.9103
0.15 0.168 1.25 0.9229
0.18 0.2009 1.3 0.934
0.21 0.2335 1.35 0.9438
0.24 0.2657 1.4 0.9523
0.27 0.2974 1.45 0.9597
0.3 0.3286 1.5 0.9661
0.33 0.3593 1.55 0.9716
0.36 0.3893 1.6 0.9763
0.39 0.4187 1.65 0.9804
0.42 0.4475 1.7 0.9838
0.45 0.4755 1.75 0.9867
0.48 0.5027 1.8 0.9891
0.51 0.5292 1.85 0.9911
0.54 0.5549 1.9 0.9928
0.57 0.5798 1.95 0.9942
0.6 0.6039 2 0.9953
0.63 0.627 2.05 0.9963
0.66 0.6494 2.1 0.997
0.69 0.6708 2.15 0.9976
0.72 0.6914 2.2 0.9981
0.75 0.7112 2.25 0.9985
0.78 0.73 2.3 0.9989
0.81 0.748 2.35 0.9991
0.84 0.7651 2.4 0.9993
0.87 0.7814 2.45 0.9995
0.9 0.7969 2.5 0.9996
0.93 0.8116 2.55 0.9997
0.96 0.8254 2.6 0.9998
0.99 0.8385 2.65 0.9998
2.7 0.9999
2.75 0.9999
2.8 0.9999

erf (z ) =
2
p
×
z
0
z
exp ( - u 2 ) du

Error function is a standard mathematical function. It is integrated numerically and the values are tabulated
...(7.30)

in Table 7.3.
Gaussian error function is also shown plotted in Fig. 7.11.

TRANSIENT HEAT CONDUCTION 301


Gaussian error function
1

0.9

0.8

0.7

0.6
erf (z) 0.5

0.4

0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
z

FIGURE 7.11 Gaussian error function erf(z ) vs. z

Remember again that in Fig. 7.11, quantities plotted on x-axis and y-axis are respectively x and erf (x ), with
the definition:
x T ( x , t ) - T0
z= and, erf(z ) = from Eq. 7.29
2× a ×t Ti - T0
Then, from Eq. 7.29, we have the temperature distribution as:

z
x
2
T(x, t ) = T0 + (Ti – T0)× × 4 ×a ×t exp( - u2 ) du ...(7.31)
p 0
Once the temperature distribution is known, heat flux at any point is obtained by applying Fourier’s law, i.e.
dT ( x, t )
Qi = – k ×A× W (instantaneous heat flow rate at a given x location)
dx
Performing the differentiation on T (x, t ) given by Eq. 7.31 by Leibnitz’s rule, we get,

dT T - T0 - x2 F I
dx
= i
p ×a ×t
× exp
4×a ×tGH JK
Substituting this in Fourier’s law, we get:

F -x I 2
exp GH 4×a ×t JK
i.e. Qi = –k × A × (Ti – T0) × W ...(7.32)
p ×a ×t
Heat flow rate at the surface (x = 0):
Putting x = 0 in Eq. 7.32,
(T0 - Ti )
Qsurface = k×A× W ...(7.33)
p ×a ×t
Total heat flow during t = 0 to t = t :

z
This is obtained by integrating Eq. 7.33 from t = 0 to t = t.
k ×A ×(T0 - Ti ) t 1
Qtotal = × dt
p ×a 0 t

302 FUNDAMENTALS OF HEAT AND MASS TRANSFER


t
i.e. Qtotal = k × A × (T0 – Ti)× 2×
p ×a

t
i.e. Qtotal = 1.13 × k × A × (T0 – Ti)× J ...(7.34)
a
Criterion to apply these relations for a finite slab:
For a slab of finite thickness L, above relations for a semi-infinite slab can be applied if:
L
³ 0.5
2× a ×t
Penetration depth and penetration time:
‘Penetration depth’ is the distance from the surface where the temperature change is within 1% of the
change in the surface temperature, i.e.
T - T0
= 0.99
Ti - T0
From Table 7.3, this corresponds to:
x
= 1.8
2× a ×t
i.e. penetration depth, ‘d’ is given by:
d = 3.6 × a ×t
‘Penetration time’ is the time taken for the surface perturbation to be felt at that depth. Therefore,
d d2
= 1.8 or, tp =
2× a ×t p 13×a
Case (ii): Constant surface heat flux:
See Fig. 7.10 (b). The solid is initially at a uniform temperature Ti, and for times t > 0, the boundary surface at x
= 0 is subjected to a constant heat flux q o .(W/m2). Then, the temperature distribution in the solid is given as:

2×q0 ×
a ×t
F -x I 2 F F x II
T(x, t ) = Ti +
p
× exp GH 4×a ×t JK q ×x
– 0 × 1 - erf GG GH 2× a ×t JK JJK ...(7.35)
k k H
where, erf is the error function defined ealier.
Case (iii): Convection at the exposed surface:
See Fig. 7.10 (c). The solid is initially at a uniform temperature Ti, and for times t > 0, the boundary surface at x
= 0 is subjected to convection with a fluid at temperature Ta and heat transfer coefficient, h.
Then, an energy balance at the surface gives:
FG dT IJ
– k×
H dx K x =0
= h × (Ta – T(0, t))

and, the non-dimensional temperature distribution in the solid is given as:


F I – F exp F h×x + h ×a ×t I I × F 1 - erf F x + h× a ×t I I
2
T ( x , t ) - Ti
Ta - Ti
= 1 – erf
x
GH
2× a ×t
JK GH GH k k JK JK GGH GH 2× a ×t k JK JJK
2
...(7.36)

To represent Eq. 7.36 in graphical form:

x h× a ×t
Put z= and, h =
2× a ×t k

TRANSIENT HEAT CONDUCTION 303


h×x
then, = 2×z× h and, Eq. 7.36 can be written as:
k
q (h, z ) = 1 – erf (z ) – (exp(2 × z × h + h2)) × (1 – erf(z + h)) ...(7.37)
Let us plot q (h, z ) against z for various values of h:
We use Mathcad to draw the graph. First, define a range variable z, varying from 0 to say, 1.8, with an
increment of 0.05. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and
y-axis with z and q (0.05, z ), respectively. Since our aim is to plot q (h, z ) for different values of z for given h,
start with h = 0.05; immediately, this graph is drawn, when we click anywhere outside the graph region. To get
the graph for next value of h = 0.1, on the y-axis, next to the earlier entry, type a comma and enter q (0.1, z ), and
click anywhere outside the graph region to get the next graph, Repeat this for different values of h as shown.
z := 0, 0.05, ... , 1.8 (define a range variable z varying from zero to
1.8, with an increment of 0.05)

Semi-infinite solid convection at surface


1

¥
3
2
1

q(z, h) 0.1
0.5
0.2

0.1

h = 0.05

0.01
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
z

FIGURE 7.12 Non-dimensional temperature for a semi-infinite solid with convection on its surface

Remember again, that in the above graph, definition of V and h are as follows:

x h× a ×t
z= and, h =
2× a ×t k
T ( x , t ) - Ti
and, q (h, z ) =
Ta - Ti
The uppermost curve in the graph is for very large h, and can be taken as for h = ¥. It signifies h = ¥, and
this implies that convection resistance is equal to zero and the temperature of the surface is equal to that of the
fluid; in other words, this case is equivalent to the case (i) already studied, where surface temperature was
suddenly changed to To, and then maintained at that temperature for all times t > 0.
Example 7.11. A thick copper slab (a = 1.1 ´ 10–4 m2/s, k = 380 W/(mC)) is initially at a uniform temperature of 250°C.
Suddenly, its surface temperature is lowered to 60°C.
(i) How long will it take the temperature at a depth of 3 cm to reach 100°C?
(ii) What is the heat flux at the surface at that time?
(iii) What is the total amount of heat removed from the slab per unit surface area till that time?
Solution.
Data:
a := 1.1 ´ 10–4 m2/s k := 380 W/(mC) Ti := 250°C T0 := 60°C x := 0.03 m T := 100°C

304 FUNDAMENTALS OF HEAT AND MASS TRANSFER


To find the time t required to reach 100°C at a depth of 0.03 m, surface heat flux and amount of heat transferred
during this period.
Time required to reach 100°C at a depth of 3 cm from the surface:
Since this is a very large slab, we will consider it as a semi-infinite medium, with the surface suddenly brought to and
maintained at a constant temperature, To. This belongs to case (i), refer to Fig. 7.10 (a).
So, Eq. 7.29 is applicable, to get temperature variation as function of position and time, i.e.
F I
T (x, t ) - T0
=erf
x
GG JJ ...(7.29)
Ti - T0 2× a ×t H K
T - T0
Now, we get: = 0.211 since all temperatures are given.
Ti - T0
From Table 7.3 for values of error function, or from Fig. 7.11, it is seen that:
erf(0.189) = 0.211
x
i.e. = 0.189
2× a ×t

x2
Therefore, t := s (time required or temperature to reach 100°C at
4 × 0.189 2 ×a a depth of 3 cm from the surface)

i.e. t = 57.262 s (time required for temperature to reach 100°C at a


depth of 3 cm from the surface.)
Heat flux at the surface:
This is obtained from Eq. 7.33, i.e.
(T0 - Ti )
Qsurface = k × A × W ...(7.33)
p ×a ×t

(T0 - Ti )
Therefore, heat flux: qsurface = k × W/m2 ...since q = Q/A
p ×a ×t
i.e. qsurface = – 5.133 ´ 105 W/m2
Note: negative sign indicates that energy is leaving the surface, which is true, since the slab is being cooled.
Total amount of heat removed, per unit surface area:
This is obtained by integrating Eq. 7.33 from t = 0 to t = t, and is given by Eq. 7.34, i.e.
A := 1 m2 ...surface area
t
Qtotal := 1.13 × k × A ×(T0 – Ti) × J ...(7.34)
a
i.e. Qtotal = – 5.886 ´ 107 J/m2 ...total heat removed from the slab.
Note: again, negative sign indicates that heat is leaving the slab.
Example 7.12. A large block of steel (a = 1.4 ´ 10–5 m2/s, k = 45 W/(mC)) is initially at a uniform temperature of 25°C.
Suddenly, its surface is exposed to a constant heat flux of 3 ´ 105 W/m2. Calculate the temperature at a depth of 3 cm
after a period of 1 min.
Solution.
Data:
a := 1.4 ´ 10–5 m2/s k := 45 W/(mC) Ti := 25°C q0 := 3 ×105 W/m2 x := 0.03 m t := 60 s
To find the temperature after a period of time t = 60 s, at a depth of 0.03 m.
Temperature at a depth of 3 cm, after a time period of 60 s:
This is the case of a semi-infinite slab, with constant heat flux conditions at its exposed surface. So, this is case (ii), refer
Fig. 7.10 (b).
So, Eq. 7.35 is applicable, to get temperature variation as function of position and time, i.e.

2×qo ×
a ×t
F -x I F F II
qo × x
GG GG x
JJ JJ
2
T(x, t ) := Ti +
k
p
× exp GH 4×a ×t JK –
k
× 1 - erf
H H 2× a ×t KK
...(7.35)

TRANSIENT HEAT CONDUCTION 305


Substituting and calculating, we get,
i.e. T(x, t ) = 98.949°C (temperature at a depth of 3 cm, after a time period of 60 s.)
Note: In Mathcad, there is no need to separately find out erf() and substitute, etc. All calculations are done in one step,
since error function is one of the built-in functions in Mathcad.
Example 7.13. A thick concrete slab (a = 7 ´ 10–7 m2/s, k = 1.37 W/(mC)) is initially at a uniform temperature of 350°C.
Suddenly, its surface is subjected to convective cooling with a heat transfer coefficient h = 100 W/(m2C) into an ambient
at 30°C. Calculate the temperature 8 cm from the surface, 1 h after the start of cooling.
Solution.
Data:
a := 7 ×10–7 m2/s k := 1.37 W/(m.C) Ti := 350°C Ta := 30°C h := 100 W/(m2C)
x := 0.08 m t := 3600 s
To find the temperature after a period of time t = 3600 s, at a depth of 0.08 m.
Temperature at a depth of 8 cm, after a time period of 3600 s:
This is the case of a semi-infinite slab, with convection conditions at its exposed surface. So, this is case (iii), refer Fig.
7.10 (c).
So, Eq. 7.36 is applicable, to get temperature variation as function of position and time, i.e.

F I F F h×x h ×a ×t I I F F x h× a ×t I I
T (x, t ) - Ti
GG x
JJ – GH exp GH k + k JK JK × GG 1 - erf GG 2× a ×t + k JJ JJ
2
= 1 – erf ...(7.36)
Ta - Ti H
2× a ×t K 2
H H KK
Therefore,
LM F x I F F h×x h ×a ×t I I F F x I I OP
GG 2× a ×t JJ - GH exp GH k + k JK JK × GG 1 - erf GG 2× a ×t h × a ×t
JJ JJ P
2

MM
T(x, t): = Ti + (Ta – Ti)× 1 - erf
H K H H
+
K K PQ
N k
2

i.e. T(x, t) = 287.811°C (temperature at a depth of 8 cm, after a time period of 3600 s.)
Again, note the ease with which above expression is calculated in Mathcad.
Exercise: Check this result from Fig. 7.12.
To show graphically the progress of cooling at various times:
It is interesting to see how the cooling of the slab progresses with time. So, let us calculate the temperatures reached by
the same point, i.e. at a depth of 8 cm from the surface, for different time periods:
t := 0.1, 0.2, ..., 15 (define a range variable t , varying from 0.1 hr
to say, 15 hr at an interval of 0.1 hr)

Progress in cooling a semi-infinite slab


400
380
360
340
320
300
280
t in hrs.
T(x, t×3600) 260
240 T(X, t) in deg.C
220
200
180
160
140
120
100
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
t

FIGURE Example 7.13 Semi-infinite slab with convection at its surface—Temperature of a point 8 cm below
the surface for various time periods, t

306 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Note: See from the above graph that after about 15 hrs the temperature of the point 8 cm below the surface is approach-
ing a temperature of 100°C.
Example 7.14. In areas where ambient temperature drops to sub zero temperatures and remains so for prolonged peri-
ods, freezing of water in underground pipelines is a major concern. It is of interest to know at what depth the water
pipes should be buried so that the water does not freeze.
At a particular location, the soil is initially at a uniform temperature of 15°C and the soil is subjected to a sub zero
temperature of –20°C continuously for 50 days.
(i) What is the minimum burial depth required to ensure that the water in the pipes does not freeze?, i.e. pipe
surface temperature should not fall below 0°C.
(ii) Plot the temperature distributions in the soil for different times i.e. after 1 day, 1 week, etc. Properties of soil
may be taken as: a = 0.138 ´ 10–6 m2/s, r = 2050 kg/m3, k = 0.52 W/(mK), Cp = 1840 J/kg.K.
Solution.
Data:
a := 0.138 ×10 –6 m2/s k := 0.52 W/(mC) Ti := 15°C T0 := – 20°C T := 0.0°C t := 50 × 24 × 3600 s
i.e. t = 4.32 ´ 106 s (time duration of exposure of soil to sub zero temperature)
To find the depth x required to reach 0°C under these conditions.
Depth at which temperature reaches 0°C:
We shall consider earth’s surface as a semi-infinite medium, with the surface suddenly brought to and maintained at a
constant temperature, To. This belongs to case (i), refer to Fig. 7.10 (a).
So, Eq. 7.29 is applicable, to get temperature variation as function of position and time, i.e.

T (x, t ) - T0 F I
= erf GG
x
JJ ...(7.29)
Ti - T0 H
2× a ×t K
T - T0
Now, we get: = 0.571 since all temperatures are given.
Ti - T0
From Table 7.3 for values of error function, or from Fig. 7.11, it is seen that:
erf(0.559) = 0.571
x
i.e. = 0.559
2× a ×t

Therefore, x: = 0.559×2× a ×t m (define x)


i.e. x = 0.863 m (depth at which pipes should be buried to
prevent water from freezing.)
To plot the temperature distributions in the soil at a depth of 1 m for different times, t :
Again, we use Eq. 7.29. From this equation temperature as a function of x and t is written as:
F x I
T(x, t): = T0 + (Ti – T0) × erf GG 2× a ×t JJ ...(A)
H K
To plot Eq. A against x for different t, in Mathcad, first of all define a range variable x varying from 0 to 1 m at an
interval of, say, 0.01 m. Then, select x–y graph from the graph palette and fill in the place holder on the x-axis with x and
the place holder on the y-axis with T(x, t1), T(x, t2), T(x, t3)... etc. where t1, t2 ... are different times, as desired. Take care
that t is entered in s. Then click anywhere outside the graph region and the graph appears immediately.
x := 0, 0.01, ... , 1 (define a range variable x, varying from 0 to 1 m,
with an increment of 0.01 m)
Note from the Fig. Example 7.14 that:
(i) even after a period of 50 days of exposure of the surface to an ambient at – 20°C, temperature at a depth of 1 m
has reached only about 2.5°C.
(ii) after 50 days, freezing temperature of 0°C is reached at a depth of 0.863 m, as calculated.
(iii) slope of the temperature curve, dT/dx, at the surface (i.e. at x = 0) decreases as time increases; this means that,
heat extracted from the surface decreases as time increases.

TRANSIENT HEAT CONDUCTION 307


Semi-infinite medium Temperature profile on x
20
15

Temperature (deg.C)
10
5
0
5
10
15
20
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Depth from surface (m)
after 1 day after 7 days
after 15 days after 30 days
after 50 days

FIGURE Example 7.14 Semi-infinite medium—Temperature variation in 1 m depth for sudden change in
surface temperature after different times

7.8 Transient Heat Conduction in Multi-dimensional Systems—


Product Solution
In sections 7.6 and 7.7 we considered one-term approximate solutions and Heisler chart solutions for infinite
plates, long cylinders, spheres and also for a semi-infinite medium. Underlying assumptions throughout were:
one-dimensional conduction and no internal heat generation. However, there are many practical cases where
assumption of one-dimensional conduction may not be valid, i.e. temperature gradients may be significant in
more than one-dimension. For example, in a ‘short cylinder’ whose length is comparable to diameter, it is
intuitively clear that temperature gradients will be significant in both the longitudinal and radial directions, i.e.
the heat transfer will be two-dimensional. Similarly, for a long rectangular bar, it is reasonable to say that heat
transfer will be significant in both the x and y directions, and in a parallele piped, heat transfer will be three-
dimensional.
7.8.1 Temperature Distribution in Transient Conduction in Multi-dimensional Systems
Some of the common two-dimensional geometries of interest are: a short cylinder, semi-infinite cylinder, infinite
rectangular bar, etc. These geometries can be imagined to be obtained by the intersection of any two of the one-
dimensional systems studied above and for which one-term approximate solutions or chart solutions are
available. Just to give an example, a short cylinder of radius R and length 2L can be imagined to be obtained by
the intersection of a long cylinder of radius R and an infinite plate of thickness 2L; an infinite rectangular bar of
sides 2L1 and 2L2 is obtained by the intersection of two infinite plates of thickness 2L1 and 2L2 respectively, etc.
Now, in such cases, it has been shown (proof is beyond the scope of this book) that for a two-dimensional
system, with no internal heat generation, it is possible to construct the solutions for dimensionless temperature
distribution in transient heat conduction, by combining the solutions of dimensionless temperature distributions
obtained for one-dimensional transient conduction, i.e. the desired two-dimensional solution is given as a
product of the one-dimensional solutions of the individual systems which form the two-dimensional body by
their intersection. So, in general, we write:
FqI FqI FqI FqI
GH q JK
i solid
= GH q JK
i system 1
× GH q JK
i system 2
× GH q JK
i system 3
...(7.38)

LHS of Eq. 7.38 refers to the two or three-dimensional body under consideration and system1, system2 etc.
are the one-dimensional systems which by their intersection form the body. (q/qi) is the dimensionless

308 FUNDAMENTALS OF HEAT AND MASS TRANSFER


temperature distribution of the one-dimensional system, which is available from Heisler charts or one-term ap-
proximation solutions.
Some of the combinations of such one-dimensional systems and the resulting two-dimensional bodies are
shown in Fig. 7.13. Remember that for a semi-infinite solid, x coordinate is measured from the surface, and for
the plane wall, it is measured from the mid-plane. In Fig.7.13, for convenience, we use the following notations:
F T ( x, t ) - T I FqI
qwall(x, t ) = GH T - T JK i a
a

wall
= GH q JK
i wall
...(7.39, a)

(r, t ) = G
F T (r , t ) - T I FqI
H T - T JK = G J
a
qcyl
i a long_cyl Hq K
i long_cyl
...(7.39, b)

F T (x, t ) - T I FqI
= G J
(x, t ) = G
H T - T JK
a
qsemi_inf
i a semi_inf
Hq K
i semi_ inf
...(7.39, c)

With this notation, two-dimensional solution for a long, rectangular bar is given by:
F T (x , y , t ) - T I
GH T - T JK
i a
a

rect_ bar
= qwall (x, t ) × qwall (y, t) ...(7.40)

And, two-dimensional solution for a short cylinder is given by:

q(x, y, t) = qwall(x, t).qsemi-inf(y, t)

q(x, y, t) = qwall(x, t).qwall(y, t)

y
y
x

2L x

FIGURE 7.13(a) Semi-infinite plate FIGURE 7.13(b) Infinite rectangular bar

q(x, y, z, t) = qwall(x, t).qwall(y, t).qsemi-inf(z, t) q(x, y, z, t) = qwall(x, t).qwall(y, t).qwall(z, t)

z z
y y

x x

FIGURE 7.13(c) Semi-infinite rectangular bar FIGURE 7.13(d) Rectangular parallelepiped

TRANSIENT HEAT CONDUCTION 309


x x
q(x, r, t) = qcyl(r, t).qwall(x, t)
q(x, r, t) = qcyl(r, t).qsemi-inf(x, t)

FIGURE 7.13(e) Semi-infinite cylinder FIGURE 7.13(f) Short cylinder

F T (r , x , t ) - T I
GH T - T JK
i a
a

short_cyl
= qwall (x, t ) × qcyl (r, t ) ...(7.41)

Important Note:
(i) Dimensionless temperatures for the one-dimensional systems used to form the product solution for the
two/three-dimensional body, must be chosen at the correct locations. In doing so, always remember that
for a semi-infinite plate, x is measured from the surface and for an infinite plate, x is measured from the
mid-plane.
(ii) If temperature is to be calculated after a given time for the multidimensional body, the solution is
straightforward, as shown; however, if the time is to be calculated to attain a given temperature, then, a
trial and error solution will be required.
7.8.2 Heat Transfer in Transient Conduction in Multi-dimensional Systems
It has been shown that heat transfer in a multidimensional body in transient conduction can be obtained by using
the Grober charts (see Figs. 7.7, 7.8 and 7.9) for Q/Qmax for the one-dimensional systems constituting the given
multidimensional body.
For a body formed by the intersection of two one-dimensional systems 1 and 2, we have:

F Q I F Q I + F Q I × LM1 - F Q I OP
GH Q JK
max total
= GH Q JK GH Q JK MN GH Q
max 1 max 2 max 1
JK PQ ...(7.42)

For a body formed by the intersection of three one-dimensional systems 1, 2 and 3, we have:

F Q I F Q I + F Q I × LM1 - F Q I OP + F Q I × LM1 - F Q I OP × LM1 - F Q I OP


GH Q JK = GH Q JK GH Q JK M GH Q JK P GH Q JK MN GH Q JK P MN GH Q JK PQ
max total max 1 N max 2 max 1 Q max 3 max 1 Q max 2

...(7.43)
Example 7.15. A rectangular aluminium bar 8 cm ´ 5 cm (a = 8.4 ´ 10–5 m2/s, k = 200 W/(mC), Cp = 890 J/(kgC), r =
2700 kg/m3), is initially at a uniform temperature of Ti = 200°C. Suddenly, the surfaces are subjected to convective
cooling into an ambient at Ta = 20°C, with the convection heat transfer coefficient between the fluid and the surfaces
being 300 W/(m2C). Determine the centre temperature of the bar after 1 min from the start of cooling
Solution. Recognise that this is the case of an infinite rectangular bar (Fig. 7.13b), formed by the intersection of two
infinite plates, one of thickness 2L1 = 8 cm and the other, 2L2 = 5 cm.
Therefore, product solution can be adopted to get dimensionless temperature distribution.
Data:
L1 := 0.04 m L2 := 0.025 m a := 8.4 × 10–5 m2/s k := 200 W/(mC) r: = 2700 kg/m3
2
Cp := 890 J/(kgC) Ti := 200°C Ta := 20°C h := 600 W/(m C) t := 60 s
To find: the centre temperature T0, after time t, surface temperature and amount of heat transferred
Centre temperature of the slab:
Solution q is given as the product of the solutions for two infinite slabs 1 and 2:

310 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For slab 1:
h ×L1
Bi := (define Biot number)
k
i.e. Bi = 0.12 (Biot number.)
a ×t
Fourier number: Fo := 2
L1
i.e. Fo = 3.15
For dimensionless temperature at the centre of the wall, we use Eq. 7.25, a:
T - Ta 2
× Fo
Centre of plane wall: q0 = 0 = A1 × e - l 1
...(7.25, a)
(x = 0) Ti - Ta

A1 and l1 have to be found from Table 7.1, against Bi = 0.12


0.4328 - 0.3111
Interpolating: l1 := 0.3111 + × 1.2
10
i.e. l 1 = 0.326
1.0311 - 1.0161
and, A1 := 1.0161 + × 1.2
10
i.e. A1 = 1.018
2
× Fo
Therefore, q01 := A1 × e - l
1
(dimensionless centre temperature for slab 1)
i.e. q01 = 0.729 (dimensionless centre temperature for slab 1)
For slab 2:
h ×L2
Bi := (define Biot number)
k
i.e. Bi = 0.075 (Biot number)
a ×t
Fourier number: Fo := 2
L2
i.e. Fo = 8.064
For dimensionless temperature at the centre of the wall, we again use Eq. 7.25, a:
A1 and l1 have to be found from Table 7.1, against Bi = 0.075
0.2791 - 0.2425
Interpolating: l1 := 0.2425 + × 15
20
i.e. l 1 = 0.27
1.013 - 1.0098
and, A1 := 1.0098 + × 15
20
i.e. A1 = 1.012
- l2 × Fo
Therefore, q02 := A1 × e 1 (dimensionless centre temperature for slab 2)
i.e. q02 = 0.562 (dimensionless centre temperature for slab 2)
Therefore, dimensionless centre temperature for the two-dimensional slab is given by the product solution:
q 0 := q 01 × q 02 (define q0, dimensionless centre temperature for given slab)
i.e. q 0 = 0.41 (demensionless centre temperature for given slab)
Centre temperature of given slab:
T0 - Ta
We have: q0 =
Ti - Ta

Therefore, T0 := Ta + q 0 × (Ti – Ta) ...define centre temperature of slab


i.e. T0 = 93.775°C (centre temperature of two-dimensional slab.)

TRANSIENT HEAT CONDUCTION 311


Exercise: Find out the amount of heat transferred per metre length, Q. Also solve this problem, using Heisler and
Grober charts. see Fig. 7.7.
Example 7.16. A short, brass cylinder (k = 110 W/(mC), r = 8530 kg/m3, Cp = 389 J/(kgC), and a = 3.39 ´ 10–5 m2/s), of
8 cm diameter and height 15 cm is initially at a uniform temperature of Ti = 200°C. The cylinder is placed in a convective
environment at 40°C for cooling with an average heat transfer coefficient of 500 W/(m2C).
(i) Determine the temperature at the centre of the cylinder 2 min after the start of the cooling process.
(ii) Determine the centre temperature of the top surface at that time, and
(iii) Determine the heat transfer from the cylinder during this time period.
(iv) Draw the temperature–time history for the centre of the short cylinder
Solution.
Data:
L := 0.075 m R := 0.04 m a := 3.39 ´ 10 – 5 m2/s k := 110 W/(mC) Cp := 389 J/(kgC)
r := 8530 kg/m3 Ti := 200 C Ta := 40 C h := 500 W/(m2C) t := 120 s
Recognise that this short cylinder can be considered to be formed by the intersection of a long cylinder of radius
R = 4 cm and a plane wall of thickness 2L = 15 cm. See Fig. 7.13 (f).
Therefore, product solution can be used. We apply Eq. 7.41, i.e.
F T (r , x , t ) - T I
GH T - T JK
i a
a

short_cyl
= qwall (x, t ) × qcyl(r, t ) ...(7.41)

Temperature at the centre of cylinder:


q (0, 0, t ) = q wall (0, t ) × q cyl (0, t)
For dimensionless centre temperature of plane wall:
a ×t
Fo := (Fourier number)
L2
i.e. Fo = 0.7232 (this is > 0.2)
h×L
Bi := (Biot number)
k
i.e. Bi = 0.34091 (Biot number)
For dimensionless temperature at the centre of the wall, we use Eq. 7.25, a:
T0 - Ta 2
× Fo
Centre of plane wall: q0 = = A1× e - l 1
...(7.25, a)
Ti - Ta
A1 and l1 have to be found from Table 7.1, against Bi = 0.341
0.5932 - 0.5218
Interpolating: l1 := 0.5218 + × 4.1
10
i.e. l1 = 0.55107
1.0580 - 1.0450
and, A1 := 1.0450 + × 4.1
10
i.e. A1 = 1.05033
2
Fo
Therefore, q 01 := A1 × e - l × 1
(dimensionless centre temperature for slab 1)
i.e. q 01 = 0.84323 (dimensionless centre temperature for slab)
For dimensionless centre temperature of cylinder:
a ×t
Fo := (Fourier number)
R2
i.e. Fo = 2.5425 (this is > 0.2)
h×R
Bi := (Biot number)
k
i.e. Bi = 0.18182 (Biot number)
A1 and l1 have to be found from Table 7.1, against Bi = 0.182
0.6170 - 0.4417
Intepolating: l1: = 0.4417 + × 8.2
10

312 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. l1 = 0.58545
1.0483 - 1.0246
and, A1 := 1.0246 + × 8.2
10
i.e. A1 = 1.04403
2
Fo
Therefore, q02 := A1 × e - l × 1
(dimensionless centre temperature for clylinder)
i.e. q 02 = 0.43677 (dimensionless centre temperature for cylinder)
i.e. qcyl (0, t ) = 0.43677
F T (0 , 0 , t ) - T I
Therefore, GH T - T JK
i a
a

short_cyl
= (0.84323) (0.43677) = 0.3683

Let T(0, 0, t ) = Tcentre


i.e. Tcentre := Ta + 0.3683 × (Ti – Ta) °C ...temperature at the centre
i.e. Tcentre = 98.928°C (temperature at the centre of short cylinder.)
Temperature at the centre of top surface of cylinder:
q (0, L, t) short_cyl = q wall (L, t)× q cyl(0, t)
Note that centre of top surface of the short cylinder is still at the centre of the long cylinder (r = 0) and at the outer
surface of intersecting plane wall (x = L). First, find the surface temperature of the plane wall: x = L = 0.075 m
x
Now, = 1 and, we use Eq. 7.24, a:
L
T ( x , t ) - Ta × Fo FG l ×x IJ
H LK
2
Plane wall: q (x, t ) = = A1 × e - l 1
× cos 1
...Fo > 0.2 ...(7.24, a)
Ti - Ta
Fo := 0.723 l1: = 0.55107 A1: = 1.05033 (for slab, already calculated.)
- l21 × Fo
Therefore, A1× e ×cos(l1) = 0.71845
T (L , t ) - Ta
i.e. = 0.71845
Ti - Ta
q (0, L, t) short_cyl = qwall(L, t) × q cyl (0, t)
F T ( 0, L, t ) - T I
i.e. GH T - T JK
i a
a

short_cyl
= qwall(L, t) × q cyl (0, t) = (0.71845) (0.43677) = 0.3138

Let T(0, L, t ) = T topsurface_centre


i.e. Ttopsurface_centre := Ta + (Ti – Ta) × 0.3138
i.e. Ttopsurface_centre = 90.208°C (temperature at the centre of top surface.)
Heat transfer from the short cylinder:
We use Eq. 7.42:

F QI F Q I + F Q I × LM1 - F Q I OP
GH Q JKmax total
= GH Q JK GH Q JK MN GH Q
max 1 max 2 max
JK PQ
1
...(7.42)

First, determine Qmax :


Qmax = r × (p ×R2 × 2 × L) × Cp × (Ti – Ta) J (maximum heat transfer = m×Cp× DT)
i.e. Qmax = 4.00295 ´ 105 J (maximum heat transfer)
Now, dimensionless heat transfer ratio Q/Qmax is determined for both the geometries from eqns. (7.27), or from
Grober’s charts, i.e. Figs. 7.7 (c) and 7.8 (c).
For the plane wall:
Bi := 0.341 Fo := 0.723 (already calculated)
l1 := 0.55107 q 0 := 0.84323
Q sin (l 1 )
Plane wall: = 1 – q0× ...(7.27, a)
Qmax l1

sin (l 1 )
Therefore, 1 – q0× = 0.19881
l1

TRANSIENT HEAT CONDUCTION 313


F QI
i.e. GH Q JK
max 1
= 0.19881

For the long cylinder:


Bi := 0.18182 Fo := 2.5425 (already calculated)
l1 := 0.58545 q 0 := 0.43677
Q J (l )
Cylinder: =1 – 2×q 0 × 1 1 ...(7.27, b)
Qmax l1
J1 (l 1 )
Therefore, 1 – 2× q 0 × = 0.58168
l1

F QI
i.e. GH Q JK
max 2
= 0.58168

Now, apply Eq. 7.42:

F Q I = F Q I + F Q I × LM1 - F Q I OP
GH Q JK GH Q JK GH Q JK MN GH Q JK PQ
max total max 1 max 2 max 1

F Q I = 0.19881 + 0.58168 (1 – 0.19881)


i.e. GH Q JK
max total

F Q I = 0.66485
i.e. GH Q JK
max total

i.e. Q := Qmax × 0.66485 J (define Q)


i.e. Q = 2.66136 ´ 105 J (heat transferred from the short cylinder
during the time period of 120 s.)
Exercise: Work out this problem using the Heisler charts & Grober’s charts.
To draw temperature–time history for centre of short cylinder:
Let us rewrite the values of l1 and A1 for wall and cylinder as follows:
For infinite wall:
Bi := 0.34091
Therefore: lwall := 0.55107
and, Awall := 1.05033
For infinite cylinder
Bi := 0.18182
Therefore: lcyl := 0.58545
and, Acyl := 1.04403
a ×t
Fourier number of wall as a function of t : Fo wall(t ) :=
L2
a ×t
Fourier number of cylinder as a function of t : Focyl (t) :=
R2
We have for dimensionless centre temperatures:
T0 - Ta 2
× Fo
Centre of plane wall: q0 = = A1 × e - l1
...(7.25, a)
(x = 0) Ti - Ta

T0 - Ta 2
× Fo
Centre of long cylinder: q0 = = A1 × e - l1
...(7.25, b)
(r = 0) Ti - Ta

314 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Then, centre temperature of the short cylinder is given as a function of time as follows:
2
× Fowall (t )
q c_wall ¬ Awall ×e - l wall

- l2cyl × Fo cyl ( t )
q c_cyl ¬ Acyl ×e
Tcentre(t) := ...(A)
q centre ¬ q c_wall ×q c_cyl
Ta + (Ti - Ta )×q centre

Therefore, Tcentre(120) = 98.92705 (checks with earlier result)


t := 0, 10, ... , 1000 (define a range variable t, varying from
0 to 1000 s, with an increment of 10 s)

Centre temperature of short cylinder vs. time


260
240
220
200
180
160
Tcentre(t) 140
120
100
80
60
40
20
0 100 200 300 400 500 600 700 800 900 1000
t
FIGURE Example 7.16 Temperature–time history for centre temperature of a short cylinder

Note:
(i) Note from the graph that centre temperature reaches the ambient temperature after about 600 s.
(ii) Eq. A is a piece of Mathcad programming. LHS defines the function T(t); on the RHS, there are 4 lines. First line
defines dimensionless centre temperature of infinite wall, next line defines dimensionless centre temperature of
long cylinder; third line defines dimensionless centre temperature of short cylinder and the last line defines the
temperature at the centre of short cylinder.
(iii) By defining Fourier number as a function of t, we ensure that for each new t, new values of Fo are calculated for
the wall as well as the cylinder.
(iv) Above graph is important, particularly when the problem is to find the time required for the centre of the short
cylinder to reach a given temperature. Then, construct the above graph and then read off the value of time
against the desired temperature. For example, from the graph, we see that time required for the centre tempera-
ture to reach 85°C is about 150 s.
(v) We can also use the solve block to find accurately the time required for the centre temperature to reach 85°C, as
shown below.
(vi) In the above graph, a t = 0, centre temperature is shown as 215.5°C and not 200°C; this error is due to the fact
that two one-term approximation solutions are multiplied together.
t := 100 s (trial value of t)
Given
Tcentre (t) = 85
Find(t) = 149.65588 s ...time required for the centre temperature to reach 85°C.
Interpolation with Mathcad:
In all the above examples, A1 and l1 for given Bi were found out by manual interpolation from Table 7.1. However, this
interpolation can be done easily and accurately in Mathcad, as follows: First, prepare Table 7.1 as an ASCII file, with the
name :Coeff.prn. Then, read this file into a matrix M by the command READPRN, as follows:
M := READPRN(“Coeff.prn”)

TRANSIENT HEAT CONDUCTION 315


Then, extract the columns of this matrix to get Biot number and values of l1and A1 for plane wall, cylinder and
sphere. Remember that columns of the matrix are generally numbered starting from zero. Matrix M has 7 columns: 0,
1,...6. 0th column gives a vector of Biot numbers, 1st column gives l1 values for wall, 2nd column gives A1 values for
wall, 3rd and 4th columns give l1 and A1 values for cylinder and, 5th and 6th columns give l1 and A1 values for sphere,
respectively.
Biot := M<0> l1wall := M < 1 > A1wall := M <2> l1cyl := M < 3 >
A1cyl := M < 4 > l1sph := M < 5 > A1sph := M < 6 >
Then, use the ‘linterp’ function for linear interpolation. Here, each column must have the same number of values. If
there are two vectors X and Y giving a series of x and y values, for any given x-value, y-value is obtained by: linterp(X,
Y, x-value). This command performs the linear interpolation to give the y-value corresponding to desired x-value.
Let us define functions to quickly get l1 and A1 for wall, cylinder and sphere, for given Biot number:
l1_Wall(Bi) := linterp(Biot, l1 wall, Bi) (defines l1 for wall, for given Bi)
example: l1_wall (0.341) = 0.55107
A1_wall (Bi) := linterp (Biot, A1 wall, Bi) (defines A1 for wall, for given Bi)
example: A1_wall (0.341) = 1.05033
l1_cyl (Bi) := linterp(Biot, l1 cyl, Bi) (defines l1 for cylinder, for given Bi)
example: l1_cyl(0.18182) = 0.58513
Al_cyl (Bi) := linterp (Biot, A1 cyl, Bi) (defines A1 for cylinder, for given Bi)
example: A1_cyl (0.18182) = 1.04399
l1_sph (Bi) := linterp (Biot, l1 sph, Bi) (defines l 1 for sphere, for given Bi)
example: l1_sph (0.25) = 0.84005
Al_sph (Bi) := linterp (Biot, Al sph, Bi) (defines A1 for sphere, for given Bi.)
example: A1_sph (0.25) = 1.0736.
Compact Mathcad program to find the centre temperature of short cylinder:
Above problem can be solved in a single step by the following Mathcad program:

h×L
Bi _ wall ¬
k
h ×R
Bi _ cyl ¬
k
l wall ¬ l 1 _ wall( Bi _ wall)
Awall ¬ A1 _ wall( Bi _ wall)
l cyl ¬ l 1 _ cyl( Bi _ cyl)
Acyl ¬ A1 _ cyl( Bi _ cyl)
Tcentre (L, R, h, k, Ti, Ta, t, a) = a ×t
Fowall ¬ 2
L
a ×t
Focyl ¬ 2
R
- l2 × Fo
q c_wall ¬ Awall ×e wall wall
- l2 × Focyl
q c_cyl ¬ Acyl ×e cyl

q centre ¬ q c_wall ×q c_cyl


Ta + (Ti - Ta )×q centre

LHS of the above program defines the centre temperature of the short cylinder as a function of the variables L, R, h,
k, Ti, Ta, t and a. RHS has 12 lines. First two lines define the Biot number for wall and cylinder, respectively. In 3rd and
4th lines, we get the l1 and A1 for the wall using the interpolation functions defined earlier. In 5th and 6th lines l1 and
A1 are calculated for the long cylinder. In 7th and 8th lines, Fourier numbers are calculated for wall and cylinder,
respectively. Centre temperatures of wall and long cylinder are calculated in lines 9 and 10, respectively. In 11th line,
dimensionless centre temperature of short cylinder is calculated as a product solution. Finally, the last line gives the
temperature at the centre of the short cylinder.
Advantage of this program is that it is quick and gives accurate calculation of the final result, i.e. the centre
temperature of the short cylinder. However, the disadvantage is that values calculated in the intermediate steps are not
available outside the program.

316 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For the above problem:
L := 0.075 m R := 0.04 m a := 3.39× 10–5 m2/s k := 110 W/(mC) Ti := 200°C Ta := 40°C
h := 5000 W/(m2C) t := 120 s
Therefore, we get:
Tcentre (L, R, h, k, Ti, Ta , t, a ) = 98.984°C (centre temperature of short cylinder.)
Compare this with the value of 98.93°C obtained earlier. Difference is due to the truncation errors crept in to the
solution in the earlier case.

7.9 Summary of Basic Equations


Basic relations derived in this chapter are summarised below in Table 7.4, for convenience and ready reference.

TABLE 7.4 Basic relations for transient conduction


Relation Comments
d 2T 1 dT Governing differential equation in Cartesian cords. for
= × one-dimensional, transient cond. without heat genera-
dx 2 a dt
tion.

q T (t ) - Ta - h × A×t F I
qi
=
Ti - Ta
= exp
r ×C p ×V
GH JK if Bi < 0.1 Lumped system analysis,
h ×Lc V
Bi = and Lc =
k A
q T (t ) - Ta a ×t
= = exp (– Bi × Fo) if Bi < 0.1 Fo = = Fourier number, or relative time
qi Ti - Ta L2c

r×C p ×V
= t Time constant (seconds)
h×A
dT (t )
Q (t ) = m × Cp × ,W Instantaneous heat transfer rate
dt
Q (t ) = h × A × (T (t ) – Ta), W

z
Qtot = m × Cp × (T(t ) – Ti), J
t Total heat transfer from time = 0 to t
Qtot = Q (t ) dt , J
0

Qmax = m × Cp × (Ta – Ti), J Maximum heat transfer

b
Temperature distribution when transient condition is
T (t ) - Ta a × (1 – exp(– a × t ))
= exp(– a × t ) + induced by mixed B.C. (e.g. a slab with constant heat
Ti - Ta Ti - Ta flux, q, at one surface and convection at the other sur-
face)
h×A
a=
r ×V ×C p

q ×A
b=
r ×V ×C p

LMT (t ) - T - FG b IJ OP Time required to attain a given temperature in the


1
MM H aK P
a
above case

MN T - T - GH ba IJK PPQ
F
t = – × ln
a
i a

b q Steady state temperature for the above case


T (t ) = Ta + = Ta + (obtained by putting t = ¥, in Eq. 7.20
a h
Contd.

TRANSIENT HEAT CONDUCTION 317


Contd.

T (x , t ) - Ta FG l ×x IJ ...Fo > 0.2 One-term approximation solution for plane wall


H L K
2
q (x, t) = = A1 × e - l ×Fo × cos 1 1
Ti - Ta

T (r , t ) - Ta
×J G
F l ×r IJ ...Fo > 02. One-term approximation solution for long cylinder
HRK
2
q (x, t) = = A1 × e - l ×Fo 1
0
1
Ti - Ta

F l ×r IJ
sin G
One-term approximation solution for a sphere
H R K ...Fo > 0.2
1
T (r , t ) - Ta 2
q (x, t ) = = A1 × e - l ×Fo 1
×
Ti - Ta l 1×r
R

T0 - Ta 2
One-term approximation-centre temperature for plane
q0 = = A1 × e - l ×Fo 1

Ti - Ta wall

T0 - Ta 2 One-term approximation-centre temperature for long


q0 = = A1 × e - l ×Fo 1

Ti - Ta cylinder

T0 - Ta
One-term approximation-centre temperature for sphere
2
q0 = = A1 × e - l ×Fo 1

Ti - Ta
Q sin(l 1)
= 1 – q0 × Dimensionless heat transfer for large, plane wall
Qmax l1

Q J (l )
= 1 – 2 ×q 0 × 1 1 Dimensionless heat transfer for long cylinder
Qmax l1

Q F
sin(l 1) - l 1×cos(l 1) I
Qmax
= 1 – 3 ×q 0 × GH l31 JK Dimensionless heat transfer for a sphere

Semi-infinite slab:

F I Dimensionless temperature distribution in a semi-infi-


T (x , t ) - T0
Ti - T0
= erf
x
GH
2× a ×t
JK nite slab, surface temperature suddenly changed to T0

T(x, t ) = T0 + (T1 – T0) ×

(T0 - Ti )
2
p
×
z0
x
4 ×a ×t
exp(- u 2 )du
Temperature distribution in a semi-infinite slab, surface
temperature suddenly changed to T0

Q surface = k × A × ,W Heat flow rate at the surface, for above case


p ×a ×t
Total heat flow during time period t for the above case
t
Q total = 1.13 × k × A × (T0 – Ti) × ,J
a

Semi-infinite slab:
Temperature distribution in a semi-infinite slab, surface is subjected to constant heat flux, q0:

2×q 0 ×
a ×t
F -x I F F II
GG JK JJ
2
T(x, t ) = T i +
p
× exp GH 4×a ×t JK –
q 0 ×x
GH
× 1 - erf
x
k k H 2× a ×t K
Contd.

318 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.

Semi-infinite slab:
Temperature distribution in a semi-infinite slab, surface is subjected to convection at its surface:

F I – F F h ×x h a ×t I I × F1 - erf F x II
T (x , t ) - Ti x
GH JK GH exp GH k + k JK JK GG GH 2× a ×t h × a ×t
JK JJ
2
= 1 – erf +
Ta - Ti 2× a ×t H 2
k K
Multidimensional transient conduction:
Temperature distribution for a body formed by intersection of three bodies:
FqI FqI FqI FqI
GH q JK
i solid
= GH q JK
i system1
× GH q JK
i system 2
× GH q JK
i system 3

Temperature distribution in long, rectangular bar:


F T (x , y , t ) - T I
GH T - T JK i a
a

rect_ bar
= q wall (x, t ) × q wall (y, t )

Temperature distribution in short cylinder:


F T (x , y , t ) - T I
GH T - T JK i a
a

short_ cyl
= q wall (x, t ) × q cyl(r, t)

Heat transfer in two-dimensional transient conduction:

FQI F Q I + F Q I × LM1 - F Q I OP
GH Q JK max total
= GH Q JK GH Q JK MN GH Q
max 1 max 2 max
JK PQ
1

Heat transfer in three-dimensional transient conduction:

FQI F Q I + F Q I × LM1 - F Q I OP + F Q I × LM1 - F Q I OP × LM1 - F Q I OP


GH Q JK max total
= GH Q JK GH Q JK MN GH Q
max 1 max 2 max
JK PQ GH Q JK MN GH Q
1 max 3 max
JK PQ MN GH Q
1 max
JK 2
PQ

7.10 Summary
In this chapter, we dealt with transient conduction, i.e. time dependent conduction, for three important, simple
geometries, namely, plane slab, long cylinder and sphere. In general, in transient conduction, temperature within
the body depends both on time and spatial coordinates. However, when the resistance for conduction within the
body is negligible as compared to the convective resistance at the surface of the body, analysis becomes simpler
and we adopt ‘lumped system analysis’, i.e. the whole body heats up or cools down as a ‘lump’, and the
temperature within the body is uniform, and is a function of time only. This is characterised by the value of non-
dimensional Biot number (Bi) being less than 0.1. When Biot number is more than 0.1, results for temperature
distribution become more complicated and are obtained as infinite series. However, if the non-dimensional time,
Fourier number (Fo) is more than 0.2, it is found that considering only the first term of the infinite series and
neglecting rest of the terms, introduces an error of no more than 2%. Such an approximate solution is known as
‘one-term approximation’. Coefficients for use in the one-term approximation have been tabulated. Now, the
same results are presented in graphical form too, known as ‘Heisler charts’ for all the three geometries consid-
ered. However, these graphs are subject to reading errors and, whenever better accuracy is desired, relations for
one-term approximation should be used.
Dimensionless heat transfer during transient conduction may be obtained either from one- term
approximaion solutions, or from the ‘Grober’s charts’, also given for the three geometries.
‘Product solution’ was explained for multidimensional transient conduction, when the temperature varia-
tion in a given body cannot be considered as one-dimensional, if the body in question could be considered as

TRANSIENT HEAT CONDUCTION 319


having been formed by the intersection of two or more one-dimensional systems for which solutions are avail-
able.
Just as in the case of steady state conduction, in transient conduction too analytical methods have their
limitation, i.e. difficulty in taking into account complex shape of the body, varying boundary conditions, or
accounting for varying thermophysical properties and heat transfer coefficients. In such cases, numerical
methods should be preferred since it is simple to handle such problems with numerical methods.
In the next chapter, we shall study numerical methods, as applied to steady state and transient conduction.

Questions
1. Differentiate between transient conduction and steady state conduction.
2. What do you understand by ‘lumped system analysis’? What are the underlying assumptions? What is the
criterion to apply lumped system analysis?
3. Explain the importance and physical significance of: Biot number and Fourier number, in transient conduction.
4. In which situation is lumped system analysis likely to be applicable—in water or in air? Why?
5. With usual notations, show that temperature distribution in a body during Newtonian heating or cooling is
given by:

T(t ) - Ta F
- h ×A ×t I
Ti - Ta
= exp GH
r ×Cp ×V
JK ...[V.T.U.]

6. For transient conduction with negligible internal resistance, prove that:


q T(t ) - Ta
= = exp (– Bi × Fo) ...[M.U.]
qi Ti - Ta
7. Discuss the effect of Biot number and Fourier number on ‘time constant’ of a thermocouple. ...[M.U.]
8. What are Heisler charts? Explain their significance in solving transient conduction problems. ...[V.T.U.]
9. What is meant by ‘one-term approximation solution’? When is it applicable?
10. What is the use of Grober’s charts?
11. What do you mean by a ‘semi-infinite medium’? In what situations the assumption of semi-infinite medium
appropriate?
12. Explain the ‘product solution method’ for multidimensional transient conduction problems. What is the main
precaution to be taken while using this method?

Problems
Lumped system analysis:
1. A large copper slab, 5 cm thick at a uniform temperature of 350°C, suddenly has its surface temperature lowered
to 30°C. Find the time at which the slab temperature becomes 100°C. Given: r = 9000 kg/m3, cp = 0.38 kJ/(kgK),
k = 370 W/(mK), h = 100 W/(m2K). Also, find out the rate of cooling after 60 seconds.
2. An aluminium plate (r = 2707 kg/m3, Cp = 0.896 kJ/(kgC), and k = 200 W/(mC)) of thickness 3 cm is at an
initial, uniform temperature of 40°C. Suddenly, it is subjected to uniform heat flux q = 7000 W/m2, on one
surface while the other surface is exposed to an air stream at 20°C, with a heat transfer coefficient of h = 60 W/
(m2C).
(i) Is lumped system analysis applicable to this case?
(ii) If yes, plot the temperature of the plate as a function of time, and
(iii) What is the temperature of the plate in steady state?
3. A household electric iron has an aluminium base (r = 2700 kg/m3, Cp = 0.896 kJ/(kgC), and k = 200 W/(mC)),
which weighs 1.4 kg. Total area of iron is 0.05 m2 and is heated with a 500 W heating element. Initially, the iron
is at ambient temperature of 20°C. How long will it take for the iron to reach 120°C once it is switched on? Take
heat transfer coefficient between iron and the ambient air as 18 W/(m2K).
4. A copper ball of 8 cm diameter, initially at a uniform temperature of 350°C is suddenly placed in an
environment at 90°C. Heat transfer coefficient h, between the ball and the fluid is 100 W/(m2K). For copper, cp =
0.383 kJ/(kgK), r = 8954 kg/m3, k = 386 W/(mK). Calculate the time required for the ball to reach a temperature
of 150°C. Also, find the rate of cooling after 1 hr. Show graphically how the temperature of the sphere falls with
time.
5. A 12 mm diameter, mild steel sphere initially at a uniform temperature of 540°C is suddenly placed in an air
stream at 27°C, with a heat transfer coefficient h of 114 W/(m2C). For mild steel, cp = 0.475 kJ/(kgK), r = 7850
kg/m3, k = 42.5 W/(mK), a = 0.043 m2/hr.

320 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i) Calculate the time required for the ball to reach a temperature of 95°C.
(ii) Also, find the instantaneous heat transfer rate two minutes after the commencement of cooling....[V.T.U.]
6. A steel bar of diameter 6 cm is to be annealed by cooling it slowly from 850°C to 150°C in an ambient at 30°C.
Heat transfer coefficient between the surface of the bar and the ambient is 40 W/(m2C). Determine the time
required for annealing. For steel, cp = 0.5 kJ/(kgK), r = 7800 kg/m3, k = 50 W/(mK).
7. An egg with a mean diameter of 40 mm and initially at 20°C is placed in boiling water for 4 min and found to
be boiled to the consumer’s taste. For how long should a similar egg for the same consumer be boiled when
taken from a refrigerator at 5°C? Take the following properties for the egg: cp = 2.0 kJ/(kgK), r = 1200 kg/m3, k
= 10 W/(mK).
Take value of heat transfer coefficient h = 100 W/(m2C). ...[M.U.]
8. A thermocouple junction is in the form of 4 mm diameter sphere. Properties of the material are Cp = 420 J/(kgK),
r = 8000 kg/m 3, k = 40W/(mK), This junction, initially at 40°C, is inserted in a stream of hot air at 300°C, with
h = 45W/(m2K). Find:
(i) time constant of the thermocouple.
(ii) thermocouple is taken out from hot air after 10 sec and is kept in still air at 30°C
Assuming heat transfer coefficient in air as 10 W/(m2K), find the temperature attained by the junction 20 sec
after removing from hot air stream. ...[M.U.]
9. A thermocouple junction is in the form of 3 mm diameter sphere. Properties of the material are: Cp= 400 J/(kgK),
r = 8600 kg/m3, k = 30W/(mK). This junction, is inserted in a gas stream to measure temperature, with a heat
transfer coefficient of h = 45 W/(m2K). How long will it take for the thermocouple to record 98% of the applied
temperature difference?
One-term approximate solution and Heisler charts:
10. A large plate of aluminium 5 m thick, is initially at 250°C, and it is exposed to convection with a fluid at 75°C
with a heat transfer coefficient of 500 W/(m2K). Calculate the temperature at a depth of 1.25 cm from one of the
faces, one minute after the plate is exposed to the fluid. What is the amount of heat removed from the plate
during this time?
Take themophysical properties of aluminium as: cp = 0.9 kJ/(kgK), r = 2700 kg/m3, k = 215 W/(mK), a =
8.4 ´ 10 –5 m2/s.
11. A steel plate (a = 1.2 ´ 10–5 m2/s, k = 43 W/(mC)), of thickness 2L = 8 cm, initially at a uniform temperature of
200°C is suddenly immersed in an oil bath at Ta = 40°C. Convection heat transfer coefficient between the fluid
and the surface is 700 W/(m2C). How long will it take for the centre plane to cool to 90°C? What fraction of the
energy is removed during this time?
12. A long, 15 cm diameter cylindrical shaft made of stainless steel 304 (k = 14.9 W/(mC), r = 7900 kg/m3, Cp = 477
J/(kgC), and a = 3.95 ´ 10–6 m2/s), is initially at a temperature of 250°C. The shaft is then allowed to cool slowly
in an ambient at 40°C, with an average heat transfer coefficient of 85 W/(m2C).
(i) Determine the temperature at the centre of the shaft 15 min after the start of the cooling process.
(ii) Determine the surface temperature at that time, and
(iii) Determine the heat lost per unit length of the shaft during this time period.
13. A solid brass sphere (k = 60 W/(mC), a = 1.8 ´ 10–5 m2/s) of 18 cm diameter is initially at 150°C. It is cooled in
an environmemt at 20°C with a heat transfer coefficient of 600 W/(m2C).
(i) How long will it take for the centre of the sphere to reach 50°C?
(ii) Also, calculate the fraction of energy removed from the sphere during this time.
(iii) Draw the radial temperature profile after different time durations at intervals of 15 min.
14. A heavily insulated steel pipe line is 1 m in diameter and is 40 mm thick. Initially, the wall is at a uniform
temperature of – 15°C. Suddenly, a hot fluid at 75°C enters the pipe with a heat transfer coefficient of 600
W/(m2C) between the fluid and the inner surface.
(i) Calculate the temperature on outer metal surface 10 min after the hot fluid is let in to the pipe.
(ii) What is the heat flux from the fluid to the pipe at that time?, and
(iii) How much energy is transferred per metre length of pipe during this time interval?
[Hint: Since diameter >> thickness of pipe, the pipe wall may be considered as a plane slab. This is a plane slab
of thickness L, insulated at one surface; therefore, its insulated surface is equivalent to the mid-plane of a plane
slab of thickness 2L. (See Example 7.10) Find Bi and Fo, and apply the one-term approximation solution formulas
for temperature distibution and heat transferred. Heat flux at the inner surface is obtained by first calculating
the temperature Ti at the inner surface (i.e. at x/L = 1), and then, by Newton’s equation i.e. q = h(Ti – Ta). You
may also check your results by Heisler and Grober charts.]

TRANSIENT HEAT CONDUCTION 321


Semi-infinite medium:
15 A thick aluminium slab, (a = 8.4 ´ 10–5 m2/s, k = 200 W/(mC)) initially at 250°C, has its surface temperature
suddenly lowered to and maintained 40°C.
(i) How long will it take the temperature at a depth of 4 cm to reach 100°C?
(ii) What is the heat flux at the surface at that time?
(iii) What is the total amount of heat removed from the slab per unit surface area till that time?
16. A thick concrete slab, (a = 7 ´ 10–7 m2/s, k = 1.37 W/(mC)) initially at 350°C, has its surface suddenly exposed
to a convection environment at 30°C, with a heat transfer coefficient of 100 W/(m2C). What is the temperature at
a depth of 8 cm from the surface after a period of 1 hour?
17. A large block of steel (a = 1.4 ´ 10–5 m2/s, k = 45 W/(mC)) is initially at a uniform temperature of 20°C.
Suddenly, its surface is exposed to a constant heat flux of 3.5 ´ 10 5 W/m2. Calculate the temperature at a depth
of 4 cm after a period of 2 min.
18. In areas where ambient temperature drops to sub-zero temperatures and remains so for prolonged periods,
freezing of water in underground pipelines is a major concern. It is of interest to know at what depth the water
pipes should be buried so that the water does not freeze.
At a particular location, the soil is initially at a uniform temperature of 15°C and the soil is subjected to a sub-
zero temperature of – 15°C continuously for 60 days.
(i) What is the minimum burial depth required to ensure that the water in the pipes does not freeze? (i.e. pipe
surface temperature should not fall below 0°C.)
(ii) Plot the temperature distributions in the soil for different times i.e. after 1 day, 1 week, etc.
Properties of soil may be taken as: a = 0.138 ´ 10–6 m2/s, r = 2050 kg/m3, k = 0.52 W/(mK), Cp = 1840 J/kgK.
19. A motor car weighing 1350 kg is moving at a speed, u = 50 km/h. It is stopped in 5 sec by 4 brakes with brake
bands of 250 cm2 area each, pressing against steel drums. Assuming that the brake lining and the drum surfaces
are at the same temperature and that the heat is dissipated by flowing across the surface of the drums (assumed
t be very thick), find the maximum temperature rise.
[Hint: K.E. of the vehicle, {(1/2)m.u2} is dissipated in a time of t = 5 sec. i.e. heat flow rate Q = {(K.E.)/t} is
known. Then, considering the drum surface as semi-infinite slab, apply Eq. 7.33 to get (To – Ti)].
Product solution:
20. A rectangular aluminium bar 6 cm ´ 3 cm (a = 8.4 ´ 10–5 m2/s, k = 200 W/(mC), Cp = 890 J/(kgC), r = 2700 kg/
m3), is initially at a uniform temperature of Ti = 150°C. Suddenly the surfaces are subjected to convective cooling
into an ambient at Ta = 20°C, with a convection heat transfer coefficient between the fluid and the surfaces being
250 W/(m 2C).
(i) Determine the centre temperature of the bar after 1 min from the start of cooling
(ii) What is the heat transferred per metre length of the bar during this period?
21. A short aluminium cylinder (k = 200 W/(mC), r = 2700 kg/m3, Cp = 890 J/(kgC), and a = 8.4 ´ 10 –5 m2/s), of 8
cm diameter and height 4 cm is initially at a uniform temperature of Ti = 200°C. The cylinder is subjected to
convective cooling with a fluid at 20°C, with an average heat transfer coefficient of 300 W/(m2C).
(i) Determine the temperature at the centre of the cylinder 1 min after the start of the cooling process.
(ii) Determine the centre temperature of the top surface at that time, and
(iii) Determine the heat transfer from the cylinder during this time period.
22. A 20 cm long, 15 cm diameter aluminium block (a = 9.75 ´ 10–5 m 2/s, k = 236 W/(mC), Cp = 896 J/(kgC), r =
2700 kg/m3), is initially at a uniform temperature of 25°C. The block is heated in a furnace at 1100°C till the
centre temperature reaches 250°C. If the heat transfer coefficient on all surfaces of the block is 60 W/(m2C),
determine how long the block should remain in the furnace.
[Hint: This short cylinder is considered as obtained by the intersection of an infinite plate and an infinite
cylinder. Solution involves trial and error method: For a range of times, calculate the centre temperature of the
short cylinder and plot a graph of time vs. centre temperature. From this graph, read the time corresponding to
a centre temperature of 250°C. While selecting the time range, be careful to see that the desired centre
temperature of 250°C is bracketed by the results obtained for the time range.]
23. A solid lead cylinder 0.5 m in diameter and 0.5 m in length, initially at a uniform temperature of 150°C, is
dropped into a medium at 20°C in which the heat transfer coefficient is 1200 W/(m2C). Plot the temperature–
time history of the centre of this cylinder.

322 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Appendix
Mathcad functions for Transient conduction for Sab, Cylinder and Sphere ...One term approximation
(Fo > 0.2):
1. Plane wall:
Values of l 1:
l1 := 1.5 (guess value)
Given
l1×tan (l1) = Bi
l1 wall (Bi) := Find (l1) ((A7.1)...Function to determine l1 as a function of Bi)
l1wall (40) = 1.5325 (Example)
Values of A1:
4 ×(l 1 wall ( Bi))
A1wall (Bi) := ((A7.2)...Function to determine A1)
2 × l 1 wall (Bi) + sin( 2 × l 1 wall (Bi))
A1 wall (100) = 1.2731 (Example)
Centre temp. of plane wall:
T0 - Ta
q0 = (T0 = centre temp., Ti = initial temp., Ta = ambient temp.)
Ti - Ta
q 0wall (Bi, Fo) := A1wall (Bi)× exp (– l1wall (Bi)2 × Fo) ((A7.3)...Function to determine
centre temp, of plane wall)
q0 wall (1, 3) = 0.121 (Example)
Temp. at any location in a plane wall:
T ( x , t ) - Ta x
q (x, t) = xbyL =
Ti - Ta L
q wall (Bi, Fo, xbyL) := A1wall (Bi)× exp (– l1wall (Bi)2 × Fo) × cos (l1wall (Bi) × xbyL)
((A7.4)...Function to determine temp. at any location in plane wall)
q wall (1, 3, 0) = 0.121 (Example)
Heat transfer in a plane wall:
Q sin (l 1 )
=1 – q0 × ...where Qmax = mCp (Ta – Ti)
Qmax l1
sin (l 1 wall (Bi))
Q by Qmax wall(Bi, Fo) := 1 – q0 wall(Bi , Fo) ×
l 1 wall (Bi )
Example A7.1. An Aluminium slab 10 cm thick, is initially at an uniform temperature of 600°C. It is suddenly immersed
in a liquid at 90°C and heat is transferred with a heat transfer coeff. of 1100 W/(m2.K). Determine;
(i) temperature at the centre line after 1 min.
(ii) temperature at the surface after 1 min.
(iii) total energy removed per unit area of the slab during this time period
Thermophysical data for Aluminium are: a = 8.85 ´ 10 –5 m2/s, k = 215 W/(m.K), r = 2700 kg/m3, Cp = 900 J/(kg.K)
Solution.
Date:
L := 0.05 (heat thickness) a := 8.85 × 10 – 5 m2/s k := 215 W/(m.C) r := 2700 kg/m3
Cp := 900 J/(kg.K) Ti := 600°C Ta := 90°C h := 1100 W(m2.C) t := 60 s
To calculate: the centre line temp., surface temp. and energy transferred per unit surface area of slab.
First check if lumped system analysis is applicable:
h×L
Bi := ...define Biot number
k
i.e. Bi = 0.256 ...Biot number.
It is noted that Biot number is > 0.1; so, lumped system analysis is not applicable. We will adopt one-term
approximation solution.
TRANSIENT HEAT CONDUCTION 323
To find the centre line temp.:
a ×t
Fourier number: Fo := i.e. Fo = 2.124 (> 0.2...therefore, one term approx. is applicable)
L2
T0 - Ta
q0 = (T0 = centre temp., Ti = initial temp., Ta = ambient temp.)
Ti - Ta
q 0 wall (Bi, Fo) := A1wall (Bi)× exp (– l1wall (Bi)2 × Fo) (Function to determine centre temp. of plane wall)
Therefore,
q 0wall (Bi, Fo) = 0.63
And, T0 := 0.63 × (Ti – Ta) + Ta
i.e. T0 = 411.3°C (Centre line temp....Ans.)
Surface temperature:
At the surface, x/L = 1.
T ( x , t ) - Ta x
q (x, t) = xbyL =
Ti - Ta L
We have:
q wall (Bi, Fo, xbyL) := A 1wall (Bi)× exp (l1wall (Bi)2 × Fo) × cos (l1 wall (Bi) × xbyL (Function to determine temp.
at any location in plane wall)
Therefore,
q wall (Bi, Fo, 1) = 0.557 (at the surface, since x/L =1)
And, T := 0.557× (Ti – Ta) + Ta
i.e. T = 374.07°C (Surface temp...Ans.)
Amount of heat transferred, Q, in one minute:
Qmax r × Cp ×V × (Ti - Ta )
= = r × Cp × (2 × L) × (Ti – Ta) J/m2 (max. heat trans. per unit area)
A A
We have:
sin ( l 1 wall ( Bi ))
q by Qmax wall (Bi, Fo) := 1 – q wall (Bi, Fo) × (Function to determine Q/Qmax)
l 1 wall ( Bi )
i.e. Q by Qmax wall (Bi, Fo) = 0.394
Therefore, Q by A := 0.394 ×[ r ×Cp × (2 ×L) × (Ti – Ta )]
i.e. Q by A = 4.883 ×107 J/m2 (heart tr. per unit area from the slab in one min...Ans.)
2. Infinite Cylinder:
Values of l 1:
l1 := 1.5 (guess value)
Given
J (l )
l1 × 1 1 = Bi
J0 (l 1 )
l1cyl (Bi) := Find (l1) ((A7.6)...Function to determine l1 as a function of Bi)
l 1cyl (10) = 2.1795 (Example)
Values of A1:
2 × J1 (l 1 cyl ( Bi))
A 1cyl (Bi) := ((A7.7)...Function to determine A1)
l cyl (Bi ) ×( J 0 (l 1 cyl (Bi ))2 + J1 (l 1 cyl (Bi)2 )
A 1cyl (10) = 1.5677 (Example)
Centre temp. of long cylinder:
T0 - Ta
q0 = (T0 = centre temp., Ti = initial temp., Ta = ambient temp.)
Ti - Ta
q 0 cyl (Bi, Fo) := A1 cyl (Bi)× exp (–l1 wall (Bi)2 × Fo) ((A7.8)...Function to determine centre
temp. of long cylinder)
q 0 cyl (0.1, 18) = 0.031 (Example)

324 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Temp. at any radius in a long cylinder:
T ( r , t ) - Ta
q(r, t ) = (r = any radius, r0 = outeer radius of cyl.)
Ti - Ta
r
rby r0 :=
r0
q cyl (Bi, Fo, rby r0) := A1cyl (Bi)×exp(–l1 cyl (Bi)2 × Fo)×J0(l 1cyl (Bi)× rby r0)
q cyl (0.1, 18, 0) = 0.031 (Example...(A7.9)...Function to determine temp. at any location)
Heat transfer in a long cylinder:
Q J (l )
=1 – 2× q0 × 1 1 (where Qmax = mCp (Ta – Ti))
Qmax l1

J1 ( l 1 cyl (Bi))
Qby Qmax cyl (Bi, Fo) := 1 – 2× q0 cyl (Bi, Fo) × ((A7.10)...Function to determine Q/Qmax)
l 1 cyl (Bi )
Qby Qmax cyl (1, 1) = 0.797 (Example)
Example A7.2. A long stainless steel shaft 10 cm in diameter is initially at an uniform temperature of 25°C. It is placed in
a furnace at 950°C and the heat transfer coeff. is 150 W/(m2.K).
(i) Calculate the time required for the axis temperature to reach 700 C
(ii) what is the temperature at a radial position of 3 cm from the centre at that time?
(iii) what is the amount of heat transferred per unit length during this time period?
For steel, a = 3.954 ´ 10 –6 m2/s, k = 14.9 W/(m.C), r = 7900 kg/m3, Cp = 477 J/(kg.C)
Solution.
Data:
L := 1 m r0 := 0.05 m a := 3.954 ×10 – 6 m2/s k := 14.9 W/(m.C) Cp := 477 J/(kg.C) Ti := 25°C
Ta := 950°C h := 150 W/(m2.C) T0 := 700°C (axis temp.)
To calculate: the time t, temp. at a rad. of 3 cm, and amount of heat transferred during this period.
First check if lumped system analysis is applicable:
r0

Bi :=2 (define Biot number...for a cylinder, Lc = (V/A) = r0/2)
k
i.e. Bi = 0.252 (Biot number.)
It is noted that Biot number is > 0.1; so, lumped system analysis is not applicable. We will adopt one term
approximation solution.
To find the time reqd. for the centre line temp. to reach 700°C:
For one term approximation, now remember that Bi is defined as:
h × r0
Bi := (define Biot number)
k
i.e. Bi = 0.503 (Biot number)
a ×t
Fourier number: Fo = (define Fourier number)
r02
We have:
T0 - Ta
q0 = (T0 = centre temp., Ti = initial temp., Ta = ambient temp.)
Ti - Ta
2
q 0 cyl (Bi, Fo) := A 1 cyl (Bi) × exp(–l1 cyl (Bi ) ×Fo) (Function to determine centre temp. of long cylinder)
T0 - Ta
i.e. = 0.27027
Ti - Ta
i.e. the function q0 cyl is equal to 0.27027. Let us calculate the fourier no. to satisfy this requirement. We use the Solve
block of Mathcad:
Fo := 0.2 (guess values)

TRANSIENT HEAT CONDUCTION 325


Given
q0 cyl (Bi, Fo) = 0.27027
Find (Fo) = 1.592
i.e. Fourier number: Fo := 1.592
Note: Observe the ease with which above calculation is perfomed with Mathcad.
Fo × r02
i.e. t :=
a
i.e. t = 1.007 × 10 3 s (time reqd. for the centre line to reach 700°C...Ans.)
Temperature at a radial distance of 3 cm from centre:
At the required posotion, r/r0 = 3/5.
i.e. rby r0 := 0.6
We have:
q 0 cyl (Bi, Fo, rby r0) := A1 cyl (Bi) ×exp(l1 cyl (Bi)2× Fo)× J0(l 1 cyl (Bi) × rby r0) (Function to determine temp.
at any location)
i.e. qcyl (Bi, Fo, rby r0) = 0.249
And,
T ( r , t ) - Ta
q(r, t) = (r = any radius, r0 = outer radius of cyl.)
Ti - Ta
i.e. T := 0.249×(Ti – Ta) + Ta
i.e. T= 719.675°C (temp. at radial distance of 3 cm...Ans.)
Amount of heat transferred, Q:
Now, Qmax = r× V ×Cp ×(Ta – Ti) (max. heat transfer possible)
i.e. Qmax := r(p× r 02 × L) × Cp × (Ta – Ti) (define Q max)
i.e. Qmax = 2.738 × 107 J (max. heat transfer)
We have:
J 1 ( l 1 cyl ( Bi ))
Qby Qmax cyl (Bi, Fo) := 1 – 2 ×q0 cyl (Bi, Fo) × (Function to determine Q/Qmax)
l 1 cyl ( Bi)
i.e. Qby Qmax cyl (Bi, Fo) = 0.759
Therefore, Q := Qmax × 0.759
i.e. Q = 2.078 ×107 J (amount of heat transferred when the centre
line reached 700°C, i.e. in 1007 seconds...Ans.)
3. Spheres
Values of l 1:
l1 := 2.5 (guess value)
Given
1 – l1 × cot(l1) = Bi
l1 sph (Bi) := Find (l1) ((A7.11)...Function to determine l1 as a function of Bi)
l 1sph (10) = 2.8363 (Example)
Values of A1:
4 ×(sin ( l 1 sph (Bi )) - l 1 sph (Bi ) × cos ( l 1 sph ( Bi))
A1 sph (Bi) := ((A7.120)...Function to determine A1)
2 × l 1 sph (Bi ) - sin ( 2 × l 1 sph ( Bi))
A1 sph (100) = 1.999 (Example)
Centre temp. of Sphere:
T0 - Ta
q0 = (T0 = centre temp., Ti = initial temp., Ta = ambient temp.)
Ti - Ta
q0 sph (Bi, Fo) := A1 sph (Bi) ×exp(– l1 sph (Bi) 2 × Fo) ((A7.13)...Function to determine
centre temp. of sphere)
q0 sph (0.02, 30) = 0.167 (Example)

326 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Temp. at any location in a sphere:
T (r , t ) - Ta
q (r, t ) = (r = any radius, r0 = outer radius of cyl.)
Ti - Ta
r
rby r0 :=
r0

A1 sph (Bi) × exp( - l 1 sph (Bi )2 × Fo if rby r0 = 0


q sph(Bi, Fo, rby r0) := sin(l 1 sph (Bi ) × rby r0 )
A1 sph (Bi) × exp( - l 1 sph (Bi )2 × Fo) × otherwise
(l 1 sph (Bi) × rby r0 )
((A7.14)...Function to determine temp. at any location)
q sph (0.02, 30, 0) = 0.167 (Example)
Heat transfer in a sphere:

Q F sin(l ) - l ×cos(l ) I
Qmax
=1 – 3q 0× GH l
1
3
1
1
JK 1
(where Qmax = mCp (Ta – Ti))

F sin(l (Bi)) - l 1 sph (Bi) × cos(l 1 sph (Bi )) I


(Bi, Fo) × G JJ
1 sph
Qby Qmax sph (Bi, Fo) := 1 – 3× q 0 sph
GH l 1 sph (Bi)
3
K
((A7.15)...Function to determine Q/Qmax)
Qby Qmax sph (1, 1) = 0.916 (Example)
Example A7.4. A stainless steel sphere, 10 mm in diameter is initially at an uniform temperature of 450°C. It is suddenly
placed in a water bath at 25°C and the heat transfer coeff. is 6000 W/(m2.K).
(i) Calculate the time required for the centre temperature to reach 50 C
(ii) what is the temperature at the surface of the sphere at that time?
(iii) what is the amount of heat transferred during this time period?
For steel, a = 3.954 ´ 10 –6 m2/s, k = 14.9 W/(m.C), r = 7900 kg/m3, Cp = 477 J/(kg.C)
Solution.
Data:
r0 := 0.005 m a := 3.954 ×10 –6 m2/s k := 14.9 W/(m.C) Cp := 477 J/(kg.C) r := 7900 kg/m 3
Ti := 450°C Ta := 25°C h := 6000 W/(m2.C) T0 := 50°C (centre temp.)
To calculate: the time t, temp. at the surface, and amount of heat transferred during this period.
First check if lumped analysis is applicable:
r0

Bi := 3 (define Biot number...for a sphere, Lc = (V/A) = r0/3)
k
i.e. Bi = 0.671 (Biot number.)
It is noted that Biot number is > 0.1; so, lumped system analysis is not applicable. We will adopt one term
approximation solution.
To find the time required for the centre to reach 50°C:
For one term solution, now, remember that Bi is defined as:
h × r0
Bi := (define Biot number)
k
i.e. Bi = 2.013 (Biot number)
a ×t
Fourier number: Fo =
r 20
We have:
T0 - Ta
q0 = (T0 = centre temp. T i = initial temp., Ta = ambient temp.)
Ti - Ta

TRANSIENT HEAT CONDUCTION 327


q0 sph (Bi, Fo) := A1 sph (Bi) ×exp(–l1 sph (Bi)2 × Fo) (Function to determine centre temp. at sphere)
T0 - Ta
i.e. = 0.05882
Ti - Ta
i.e. the function q0 sph is equal to 0.05882. Let us calculate the Fourier no. to satisfy this requirement. We use the Solve
block of Mathcad:
Fo := 0.2 (guess number)
Given
q0 sph(Bi× Fo) = 0.05882
Find(Fo) = 0.78
i.e. Fourier number: Fo := 0.78
Note: Observe the ease with which above calculation is performed with Mathcad.
Fo × r 02
i.e. t :=
a
i.e. t = 4.932 s (time reqd. for the centre temp. to reach 50°C...Ans.)
Temperature at the surface of sphere:
At the surface, r/r0 = 1.
i.e. rby r0 := 1
We have:
T (r, t ) - Ta
q (r, t) = (r = any radius, r0 = outer radius of cyl.)
Ti - Ta

A1 sph (Bi ) × exp( - l 1 sph ( Bi) 2 × Fo ) if rby r0 = 0


q sph (Bi, Fo, rby r0) := sin ( l 1 sph ( Bi) × rby r0 = 0
A1 sph (Bi ) × exp( - l 1 sph ( Bi) 2 × Fo ) × otherwise
( l 1 sph ( Bi) × rby r0 = 0)

(Function to determine temp. at any location)


i.e. q sph (Bi, Fo, l) = 0.026
And, T := 0.026× (Ti – Ta) + Ta
i.e. T = 36.05°C (temp. at the surface of sphere...Ans.)
Amount of heat transferred, Q:
Now, Qmax = r × V × Cp × (Ti – Ta) (max. heat transfer possible)
F 4 ×p × r 30 I × C × (T – T )
i.e. Qmax := r× GH 3
JK p i a (define Q max)

i.e. Qmax = 838.558 J (max. heat transfer)


We have:
F sin (l (Bi ) - l 1 sph ( Bi) × cos( l 1 sph ( Bi)) I
Qby Qmax sph (Bi, Fo) := 1 – 3× Q0 sph (Bi, Fo) × GG 1 sph
JJ
H l 1 sph (Bi ) 3 K
(Function to determine Q/Qmax)
i.e. Qby Qmax sph (Bi, Fo) = 0.962
Therefore, Q := Qmax × 0.962
i.e. Q = 806.693 J (amount of heat transferred when the centre of
sphere reaches 50°C, i.e. in 4.932 seconds...Ans.)

328 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

8
Numerical Methods in
Heat Conduction

8.1 Introduction
In chapter 3, we derived the general differential equation for heat conduction in cartesian, cylindrical and spheri-
cal coordinates. Subsequently, considering one-dimensional conduction, we solved these differential equations,
with appropriate boundary conditions, for cases of simple geometries such as a plane wall, cylinder and sphere
and obtained temperature distribution in those geometries; then, by applying Fourier’s law, heat transfer rate
was obtained. The analytical solutions obtained for temperature distribution are known as ‘exact solutions’ since
temperature at any point in the solid is obtained by applying the equations derived. While getting an exact
solution is always preferable, following points in connection with the analytical solutions must be noted:
(i) Analytical solutions are suitable for simple geometries such as a plane wall, cylinder or sphere, where the
surface of the body and the coordinate surfaces coincide, i.e. surfaces of a plane wall are completely
bounded by the coordinate surfaces of a cartesian coordinate system, surfaces of a cylinder and sphere
are completely bounded by a cylindrical and spherical coordinate system respectively.
(ii) However, for irregular geometries, analytical solutions become difficult. For example, if there is a handle
on a cylindrical cup, finding out the temperature distribution in the system becomes very difficult or
impossible by analytical methods.
(iii) Further, even in simple geometries, if there is variation of thermal conductivity with temperature, or if
the heat transfer coefficient varies over the surface, or if there is radiation heat transfer involved at the
surfaces, severe non-linearities are introduced and analytical solutions become highly complicated or
impossible.
(iv) Many times, analytical solutions, even if available for certain problems, are so complicated with the pres-
ence of infinite series, Bessel functions etc. that the user gets intimidated from using them.
In such cases, popular alternative method is ‘numerical solution’. Here, the differential equation is substi-
tuted by a set of algebraic equations and simultaneous solution of these algebraic equations gives the tempera-
tures at selected, ‘discrete points’ in the system. So, the important difference to be noted is that while in an
analytical solution, temperature is obtained at any point in the body, in a numerical solution temperatures are
obtained only at selected, discrete points or ‘nodes’. By selecting these nodes close enough, sufficiently accurate
results are obtained.
Advantages of numerical methods are:
(i) easy to apply, with the availability of high speed computers
(ii) desired accuracy can be obtained by controlling the number of nodes or ‘mesh size’.
(iii) variation in area, thermal conductivity or heat transfer coefficients, and complicated boundary conditions
can be easily taken into account.
(iv) mathematical model for a numerical solution is more likely to be a better representative of the actual
system
(v) parametric study to observe the effect of variation of different parameters on the solution, or ‘what-if’
analysis, is easier with numerical methods in conjunction with high speed computers.
Generally used numerical techniques are ‘finite difference’, ‘finite element’, ‘boundary element’ and ‘energy
balance or control volume’ methods. We will adopt energy balance method since it is intuitively easier to apply
energy balance on control volumes and does not involve complicated mathematical formulations.
In this chapter, we shall learn to formulate set of algebraic equations from the differential equations in
cartesian, cylindrical and spherical coordinates and solve them for one-dimensional, steady state conduction.
Then, we shall study the finite difference representation and solution of two-dimensional, steady state conduc-
tion problems. Since the numerical solution essentially involves solving a set of algebraic equations simultane-
ously, we shall study the different methods of solving simultaneous algebraic equations. Finally, numerical
solution of one-dimensional and two-dimensional transient conduction problems will be described.

8.2 Finite Difference Formulation from Differential Equations


As mentioned earlier, in this book, we shall formulate finite difference equations by making energy balance on
differential control volumes. However, as an introduction and as an example, for one case, let us obtain the finite
difference form of equation directly from the differential equation mathematically, starting with the definition of
first derivative and second derivatives.
Consider the governing equation for one-dimensional, steady state heat conduction with heat generation:

d 2T ( x ) qg
2
+= 0 in 0 < x < L ...(8.1)
dx k
Now, le us divide the region 0 < x < L into M sub-regions. Then, size of each sub-region is:
L
Dx = ...(8.2)
M
So, there are M + 1 nodes, starting from m = 0 to m = M, as shown in Fig. 8.1.
Coordinate of node m is x = m.Dx. and let temperature of node m be Tm.
Now, in Eq. 8.1, we need second derivative of T. To represent it in terms of finite differences, we proceed as
follows:
Consider locations (m + ½) and (m – ½) as shown in Fig. 8.1. First derivative of temperature dT/dx at these
locations is written in terms of finite differences as:
FG dT (x) IJ Tm + 1 - Tm
H dx K m+
1
2
=
Dx
...(8.3a)

and,
FG dT (x) IJ Tm - Tm - 1
H dx K m-
1
2
=
Dx
...(8.3b)

Then, the second derivative d2T/dx2 at node m is approximated as:

FG dT (x) IJ FG dT (x) IJ
F d T ( x) I
2 H dx K m+
1
-
H dx K m-
1

GH dx JK 2
= 2
Dx
2

T0 T1 T2 Tm 1 Tm Tm + 1 TM 1 TM
x

0 1 2 m 1 m m+1 M 1 M
Dx
m 1/2 m + 1/2
x=0 x=L

FIGURE 8.1 Finite difference representation of derivatives

330 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F d T ( x) I
2 Tm - 1 - 2 ×Tm + Tm + 1
i.e. GH dx JK
2
=
( D x )2
...(8.4)
m
Substituting Eq. 8.4 in Eq. 8.1:

( D x)2 × qm
(Tm – 1 – 2×Tm + Tm + 1) + =0 ...(8.5)
k
where, qm is the energy generation rate at node m, and m = 1, 2, 3.....M – 1
Eq. 8.5 is the finite difference form of representation of the differential equation given by Eq. 8.1. It is valid
for the ‘interior nodes’ i.e. nodes 1, 2….M – 1. Since qm, k and Dx are known quantities, Eq. 8.5 provides (m – 1)
simultaneous algebraic equations for temperature. But, there are M + 1 nodes, and we need two more equations
to solve M + 1 node temperatures; these two equations are obtained by finite difference representation of bound-
ary conditions at nodes m = 0 and m = M, as will be shown later.

8.3 One-dimensional, Steady State Heat Conduction in Cartesian Coordinates


Now, we shall develop the finite difference formulation using the ‘energy balance’ approach. In this method, the
medium in question is sub-divided into many sub-volumes; centre of each sub-volume is known as a ‘node’ and
each node represents the average properties of the sub-volume around it. Thus, at node ‘m’, temperature Tm of
that node represents the average temperature of the sub-volume around node ‘m’. It is imagined that these nodes
are connected to each other by ‘conducting rods’ i.e. in effect, the total volume is replaced by a network of nodes
with conducting rods. It is further assumed that temperature between adjacent nodes varies linearly.
Consider one-dimensional, steady state heat conduction in a plane wall of thickness L, with heat generation
rate qg (x) and constant thermal conductivity k. Now, let us divide the region 0 < x < L into M sub-regions. Then,
thickness of each sub-region is:
Dx = L/M. So, there are totally (M + 1) nodes, starting from m = 0 to m = M, as shown in Fig. 8.2. Coordinate
of node m is x = m.Dx. and let temperature of node m be Tm . Remembering that each node represents the sub-
volume around it (of thickness Dx), it is clear that interior nodes 1, 2…M – 1 represent full sub-volumes whereas
boundary nodes 0 and M represent half volumes (of thickness Dx/2).
To get the difference equation for the interior nodes, let us write an energy balance for the volume element
represented by node m. Assuming that all heat conduction is into the element, we can write, for steady state
conditions:
Rate of heat conduction from left + Rate of heat conduction from right + Rate of heat generation inside the
element = 0. i.e.
Q left + Q right + qm ×A×Dx = 0 ...(8.6)
where qm is the heat generation rate per unit volume for sub-volume represented by node m (assumed constant
for the entire wall), A is the heat transfer area perpendicular to the direction of heat flow (constant for the wall),
and A. Dx is the volume of the element. Now, note that for a wall with heat generation, temperature distribution
is not linear. However, we make an approximation that the temperature variation between two nodes is linear;
and this assumption is valid for small values of Dx. So, writing the energy balance, with the direction of all heat
flow into the element,
Tm - 1 - Tm Tm + 1 - Tm
k×A× + k×A× + qm ×A×Dx = 0 ...(8.7)
Dx Dx

qm ×( D x)2
i.e. (Tm – 1 – 2×Tm + Tm + 1) + =0 ...(8.8)
k
where, q m is the energy generation rate at node m, and m = 1,2,3.....M – 1
Note that Eq. 8.8 is identical to Eq. 8.5 derived earlier mathematically by consideration of definition of first
and second derivatives.
Again, Eq. 8.8 is applicable only to M – 1 interior nodes; we will need two more equations to solve M
unknown node temperatures. These two equations are obtained by writing energy balance at the two boundary
nodes 0 and M.

NUMERICAL METHODS IN HEAT CONDUCTION 331


Volume element of node ‘m’
Plane wall
Volume element of qm Volume element of
node ‘0’ node ‘M’
Qleft Qright
Dx

T0 T1 T2 Tm 1 Tm Tm + 1 Tm 1 Tm
x
0 1 2 m 1 m m+1 M 1M
m 1/2 m + 1/2 Dx
x=0 x=L

FIGURE 8.2 Finite difference formulation in a plane wall by energy balance

It is convenient to assume while writing the energy balance, that all heat flows are towards the node in
question; the signs of heat flow adjust themselves when the set of coupled algebraic equations so obtained are
solved simultaneously.
Writing in terms of thermal resistances, Eq. 8.7 can be written as:
Tm - 1 - Tm Tm + 1 - Tm
+ + qm ×A×Dx = 0 ...(8.9)
Rm - 1, m Rm + 1, m

F Dx I
where Rm – 1, m = GH k × A JK m - 1, m
= thermal resistance between nodes m – 1 and m

= G
F Dx I
Rm + 1, m
H k × A JK m + 1, m
= thermal resistance between nodes m + 1 and m

qm = energy generation rate at node m


A× Dx = volume of element about node m
Eq. 8.9 is more general and allows for the variation of thermal conductivity and cross-sectional area with
position. When k and A are constants, Eq. (8.9) reduces to eqn. (8.8).
Boundary conditions:
Eq. 8.8 or 8.9 developed above are applicable to internal nodes. For nodes at the boundaries (i.e. for nodes 0 and
M), difference equations are developed again by writing the energy balance for the volume elements containing
these nodes. While doing so, the boundary conditions prescribed in the problem must be taken into account.
Also, note that volume elements for nodes ‘0’ and ‘M’ for a plane wall are half-volumes as shown in Fig. 8.2.
Most commonly encountered boundary conditions are: prescribed temperature, prescribed heat flux, con-
vection and radiation boundary conditions.
(i) Prescribed temperatures at the boundaries This is the simplest of the boundary conditions. Let the tempera-
tures at x = 0 and x = L be given as To and TM respectively. Then, T(o) = To and T(L) = TM, give the two additional
equations required to solve for M + 1 unknown node temperatures. In this case, there is no need to write energy
balance for volume elements at the boundaries, since the temperatures at the boundaries are known.
To develop finite difference equations for the other boundary conditions, we apply energy balance to the
volume elements of nodes at the boundaries, i.e. nodes 0 and M (See Fig. 8.3). Remember that these volume
elements are only half volumes of thickness Dx/2, each. Also, while writing energy balance, consider all energy
flows as flowing into the element. Heat flux into the element is considered as positive and out of the element, it
is negative.
Then, energy balance for the volume element for node ‘0’ on the left boundary of the wall is given by:
(T1 - T0 ) A×D x FG IJ
Q left + k×A×
Dx
+ qo ×
2
=0
H K ...(8.10)

Let us now apply Eq. 8.10 to get difference equations for boundary nodes ‘0’ and ‘M’:

332 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Volume element of node ‘0'
heat generation qo
Dx/2

Qleft Qright = kA(T1 T0)/Dx

T0 T1 T2 Tm 1 Tm Tm + 1 TM 1 TM
x

0 1 2 m 1 m m+1 M 1 M
Dx
x=L
x=0
FIGURE 8.3 Finite difference formulation for left boundary of a plane wall

(ii) Prescribed heat flux at the boundaries Let q left and q right be the heat flux at nodes ‘0’ and ‘M’ respectively.
Then, from Eq. 8.10:
For node ‘0’:
(T1 - T0 ) FG
A×D x IJ = 0
q left ×A + k×A×
Dx
+ qo ×
2 H K ...(8.11)

( D x)2 × qo 2× D x × qleft
i.e. 2×T 1 – 2×T 0 + + =0 ...(8.12)
k k
For node ‘M’: Replace the subscript ‘0’ by ‘M’ and subscript ‘1’ by ‘M – 1’:
(TM - 1 - TM ) FG A × D x IJ = 0
q right ×A + k×A×
Dx
+ q M×
H 2 K ...(8.13)

( D x )2 × qM 2× D x × qright
i.e. 2×TM – 1 – 2×TM + + =0 ...(8.14)
k k
Eqs. 8.12 and 8.14 are finite difference representation of the prescribed heat flux conditions at nodes ‘0’ and
‘M’ respectively.
For insulated boundary condition and for a plane of thermal symmetry:
This is a special case of prescribed heat flux condition. Now, q left = q right = 0. Then, Eqs. 8.12 and 8.14 become:

( D x)2 × qo
2×T1 – 2×T0 + =0 ...(8.15)
k

( D x )2 × qM
2×TM – 1 – 2×TM + =0 ...(8.16)
k
Eq. 8.15 and 8.16 for an insulated boundary or a plane of thermal symmetry can be obtained more easily by
applying the ‘mirror image concept’. In this simple method, the insulated boundary or the plane of thermal
symmetry is considered as a mirror. Thus, for the node ‘0’, insulated left face becomes a mirror and reflects node
1; then, node ‘0’ has the reflected node ‘1’ on its left and node ‘1’ on its right and we write the difference equation
as if the node ‘0’ is an internal node. Then, applying Eq. 8.8 for an internal node, putting m = 0, we get:

qm ×( D x)2
(Tm – 1 – 2×Tm + Tm + 1) + =0 ...(8.8)
k
Put m = 0 and T –1 = T 1:

qo ×( D x)2
i.e. 2×T 1 – 2×T0 + =0 ...(8.17)
k
Equation (8.17) is the same as eqn. (8.15).

NUMERICAL METHODS IN HEAT CONDUCTION 333


Similarly, for node ‘M’, right hand surface which is insulated becomes the mirror and node ‘M – 1’ is re-
flected further to the right of node ‘M’ and now, considering node ‘M’ as an internal node, Eq. 8.8 becomes:

qm ×( D x) 2
(Tm – 1 – 2×Tm + Tm + 1) + =0 ...(8.8)
k
Put m = M, and M + 1 = M – 1:

(D x) 2 × q M
i.e. 2×T M – 1 – 2×T M + =0 ...(8.18)
k
Eq. 8.18 is the same as Eq. 8.16.
Thus, note that for an insulated boundary condition, or for a plane of thermal symmetry, it is very conven-
ient to use the ‘mirror image concept’ and write the difference equation as if the boundary node is an internal
node.
(iii) Convection boundary condition Let the boundaries at x = 0 and x = L be subjected to convection to a fluid
at a temperature of Ta with a heat transfer coefficient of h.
Then, Eq. 8.10 becomes:
For node ‘0’:
(T1 - T0 ) FG
A× D x IJ = 0
h×A×(Ta – T0) + k×A×
Dx
+ q 0×
H
2 K ...(8.19)

FG
h×D x IJ
( D x )2 × qo 2× h × D x
i.e. 2×T1 – 2×T0 × 1 +
kH +
kK +
k
×Ta = 0 ...(8.20)

For node ‘M’: Replace the subscript ‘0’ by ‘M’ and subscript ‘1’ by ‘M – 1’:
We get:
(TM - 1 - TM ) FG A × D x IJ = 0
h×A×(Ta – TM) + k×A×
Dx
+ q M×
H 2 K ...(8.21)

FG h×D x IJ( D x )2 × qM 2× h × D x
i.e.
H
2×TM – 1 – 2×TM × 1 +
k
+
K k
+
k
×Ta = 0 ...(8.22)

Eq. 8.20 and 8.22 are finite difference representations for convective boundary conditions at nodes ‘0’ and
‘M’ respectively.
(iv) Radiation boundary condition Let the surrounding temperature be Ta, emissivity of the surface e, and s , the
Stefan– Boltzmann constant. Then, Eq. 8.10 becomes:
For node ‘0’:
(T1 - T0 ) A× D x FG IJ = 0
e×s×A×(Ta4 – T04) + k×A×
Dx
+ q 0×
2 H K ...(8.23)

For node ‘M’: Replace the subscript ‘0’ by ‘M’ and subscript ‘1’ by ‘M – 1’:
We get:
TM - 1 - TM FG A × D x IJ = 0
e×s×A×(Ta4 – TM4) + k×A×
Dx
+ q M×
H 2 K ...(8.24)

We generally try to avoid radiation boundary condition even with numerical methods, since as can be seen
easily from Eqs. 8.23 and 8.24, finite difference equations now become highly non-linear and are difficult to solve.
(v) Combined convection and radiation boundary condition Let there be radiation as well as convection at the
surfaces, giving a combined heat transfer coefficient of h comb and let the fluid temperature be Ta. Then, Eq. 8.10
becomes:
For node ‘0’:
(T1 - T0 ) FG
A×D x IJ = 0
h comb ×A×(Ta – T0) + k×A×
Dx
+ q 0×
2 H K ...(8.25)

334 FUNDAMENTALS OF HEAT AND MASS TRANSFER


h FG
×D x ( D x) × q0 IJ
2× hcomb × D x 2
i.e. 2×T1 – 2×T0 × 1 + comb
k H +
k
+
Kk
×Ta = 0 ...(8.26)

For node ‘M’: Replace the subscript ‘0’ by ‘M’ and subscript ‘1’ by ‘M – 1’:
We get:
(TM - 1 - TM ) FG A × D x IJ = 0
h comb ×A×(Ta – TM) + k×A×
Dx
+ q M×
H 2 K ...(8.27)

hFG ×D x ( D x ) × qMIJ 2
2× hcomb × D x
i.e.
H
2×TM – 1 – 2×TM × 1 + comb
k
+
k K
+
k
×Ta = 0 ...(8.28)

(vi) Interface boundary condition With no contact resistance:


Node ‘m’ is at the interface between two solids in ‘perfect thermal contact’, i.e. there is no contact resistance and
both the surfaces are at the same temperature at the interface node ‘m’. This situation is shown in Fig. 8.4.

Interface
qA, m qB, m
Medium A Medium B
kA kB
Q left Qright
x
0 1 2 m 1m m+1

Dx
Qleft = KAA(Tm 1 Tm)/Dx
Qright = KBA(Tm + 1 Tm)/Dx

FIGURE 8.4 Finite difference formulation for interface boundary condition

So, the finite difference formulation for this boundary condition is given by:
Tm - 1 - Tm Tm + 1 - Tm FG A × D x IJ + q × FG A × D x IJ = 0
kA ×A×
Dx
+ k B ×A×
Dx
+ q A, m ×
H 2 K B, m
H 2 K ...(8.29)

In the above relation, subscripts A and B refer to materials A and B, k is the thermal conductivity, q is the
heat generation rate and A is the area of cross-section normal to the direction of heat flow.
With contact resistance:
If there is a contact resistance R c at the interface, we use the resistance concept to write the difference equation.
(See Eq. 8.9). Now, at the interface, there is a temperature drop. Let the temperature at the interface drop from Tc1
to Tc 2.
Then, we can write:
Tm - 1 - Tc1 Tm + 1 - Tc2 FG A × D x IJ + q × FG A × D x IJ = 0
kA ×A×
Dx
+ k B ×A×
Dx
+ q A, m×
H 2 K B, m
H 2 K ...(8.30)

And, temperature drop at the interface is calculated as:


Rc
DTc = (Tc 1 – Tc2 ) = Q × ...(8.31)
A
where Q is the heat flow rate through the interface (i.e. between nodes (m – 1) and (m + 1)) and (R c/A) is the
interface thermal resistance.
Example 8.1. One face of a slab of thickness 1 cm (k = 20 W/(mC)), is maintained at 40°C and the other surface is
subjected to a convection heat transfer with a fluid at 100°C with a heat transfer coefficient of 4000 W/(m2 C). There is
uniform internal heat generation in the slab at a rate of 8 ´ 10 7 W/m3.
(a) Dividing the slab into 5 equally spaced sub-regions, find the temperatures at the different nodes. Assume one-
dimensional, steady state conduction.

NUMERICAL METHODS IN HEAT CONDUCTION 335


T0 T1 T2 T3 T4 T5 (b) If the left surface is insulated, what is the temperature on that sur-
x face in steady state?
Solution.
0 1 Dx 2 3 4 5
Data:
x=0 x=L L := 0.01 m M := 5 k := 20 W/(mC) T0 := 40°C
Ta := 100°C h := 4000 W/(mC) qg := 8 ´ 107 W/m 3
FIGURE Example 8.1 Finite L 0.01
difference nodes for Example 8.1 D x := = i.e. Dx = 0.002 m
M 5
Note that there are 6 nodes, numbered as: 0, 1, 2, 3, 4, and 5. Out of these, nodes ‘0’ and ‘5’ are boundary nodes and
the nodes 1, 2, 3 and 4 are internal nodes. Temperature of node ‘0’ is given, i.e. T0 = 40°C, for the first part of the
problem.
Fig. Example 8.1 shows the schematic of finite difference nodes for this problem.
Apply Eq. 8.8 for interior nodes, 1, 2, 3 and 4:
qg ×( D x)2
(Tm – 1 – 2×Tm + Tm + 1) + = 0 (8.8)
k
qg ×( D x)2
We have: = 16
k
Node 0: T 0 = 40°C (by data...(a))
Node 1: T0 – 2×T1 + T2 + 16 = 0 ...(b)
Node 2: T1 – 2×T2 + T3 + 16 = 0 ...(c)
Node 3: T2 – 2×T3 + T4 + 16 = 0 ...(d)
Node 4: T3 – 2×T4 + T5 + 16 = 0 ...(e)
For Node 5: here, we have convection boundary condition. So, apply Eq. 8.22:
FG h×D x IJ + ( D x) × q
2
2× h × D x
H K
M
2×TM – 1 – 2×TM × 1 + + ×Ta = 0 ...(8.22)
k k k
Here, q M = qg
Then, for M = 5, we get:
2×T4 – 2 .8×T5 + 16 + 80 = 0 (f)
Eq. a to f have to be solved simultaneously to get 6 nodal temperatures. Of course, in this case temperature at node
‘0’ is already known.
We shall discuss the different methods of solving coupled algebraic equations, later. But, now, we will use ‘Solve
block’ of Mathcad to solve these 6 equations simultaneously.
We start with assumed or trial values for all the variables i.e. for the temperatures at nodes 1 to 5. Then, in the solve
block, immediately below ‘Given’ write all the constraint equations. Then, the command ‘Find (T0, T 1, T 2 ... T5)’ immedi-
ately gives a vector of temperature values:
T 1: = 50 T2: = 50 T3: = 50 T4: = 50 T5: = 50 (trial values of temperatures)
Given T 0 = 40°C (by data...(a))
T0 – 2×T 1 + T 2 + 16 = 0 ...(b)
T1 – 2×T 2 + T 3 + 16 = 0 ...(c)
T2 – 2×T 3 + T 4 + 16 = 0 ...(d)
T3 – 2×T 4 + T 5 + 16 = 0 ...(e)
2×T 4 – 2×8 ×T5 + 16 + 80 = 0 ...(f)
Temp: = Find(T0, T1, T2, T3, T4 , T5) (‘Temp’ is the vector containing values of temperatures T0, T1 ... T5)
Therefore,

LM40 OP
MM93.333 PP
Temp = M
MM152 PPP
130. 667

MM157 .333
P
N 667PQ
146 .

336 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Temperatures at different nodes are:
T 0 = 40°C T1 = 93.333°C T2 = 130.667°C T3 = 152°C T4 = 157.333°C T5 = 146.667°C
To draw the temperature distribution:
In the above, temperatures at various nodes are contained in vector ‘Temp’. In Mathcad, elements of the vector are
generally counted starting from zero. i.e. zeroth element of vector Temp gives value of T0 = 40, element numbered 1
gives value of T1 = 93.333, and so on.
To draw the graph, first we define a range variable i = 0 to 5. Then choose x–y graph from the graph pallete and on
the x-axis place holder fill up i and in the y-axis place holder fill up Tempi. Click anywhere outside the graph and
immediately the graph appears.
i := 0, 1, ..., 5 (define the range variable i, varying
from 0 to 5 with an increment of 1)

200

150

Tempi 100

50

0
0 1 2 3 4 5
i

FIGURE Example 8.1(b) Temperature at different nodes in the slab

In the Fig. Example 8.1 (b) ‘i’ is the node no. on the x-axis and on the y-axis, Tempi, the corresponding node
temperature is plotted.
(b) When left surface is insulated:
Now, the node ‘0’ is on an insulated boundary. Difference equation for node ‘0’ is obtained now, treating it as an
internal node if the insulated surface is imagined to be a mirror i.e. node ‘1’ extends to the left of node ‘0’ and Eq. 8.8 is
applicable.
qm ×( D x ) 2
i.e. (T m – 1 – 2×Tm + Tm + 1 + = 0 ...(8.8)
k
qg ×( D x)2
For m = 0: T –1 – 2×To + T1 + = 0
k
From mirror image concept: T –1 = T 1
Therefore, for node ‘0’, we get:
2×T 1 – 2×T 0 + 16 = 0 ...(a¢)
Equations for other nodes remain unchanged.
Therefore, solving Eq. a¢ alongwith b, c, d, e and f simultaneously will give the temperatures at nodes 0 to 5.
Use ‘solve block’ to solve the set of algebraic Eqs. a¢ to f simultaneously, in Mathcad. Start with assumed or trial
values of temperatures:
T 0 : = 50 T1: = 50 T2: = 50 T3: = 50 T4: = 50 T5: = 50 (trial values of temperatures)
Given
2×T 1 – 2×T 0 + 16 = 0 ...(a¢)
T0 – 2×T 1 + T 2 + 16 = 0 ...(b)
T1 – 2×T 2 + T 3 + 16 = 0 ...(c)
T2 – 2×T 3 + T 4 + 16 = 0 ...(d)
T3 – 2×T 4 + T 5 + 16 = 0 ...(e)
2 ×T 4 – 2.8×T5 + 16 + 80 = 0 ...(f)
Temp := Find(T0, T1, T2, T3, T4, T5) (‘Temp’ is the vector containing values of temperatures T0, T1 ... T5)

NUMERICAL METHODS IN HEAT CONDUCTION 337


Therefore,

LM500OP
MM492PP
Temp = M
MM428PPP
468

MM372 P
N300PQ
i.e. Temperatures at different nodes are:
T 0 = 500°C T1 = 492°C T2 = 468°C T3 = 428°C T4 = 372°C T5 = 300°C
Let us compare these values with those obtained from analytical solution. Analytical solution to this problem is
easily obtained from the following mathematical formulation of the problem:
d 2T ( x ) q g ( x)
+ = in 0 < x < L (eqn. (h)...for a slab)
dx 2 k
dT ( x)
= 0 at x = 0 ((i)...since insulated)
dx
FG dT (x) IJ
h ×A×(T5 – Ta) = – k×A×
H dx K x=L
...(j)...convection at the right surface, i.e. at x = L

Solving the above governing differential Eq. h with the boundary conditions i and j at x = 0 and x = L, we get the
following analytical solution for temperature distribution:
qg × L2 LM FG x IJ OP + q × L
2
g
T(x): =
2× k MN H L K PQ h
× 1- + Ta ...(k)

Then, temperatures at nodes ‘0’ to 5 are obtained by putting corresponding x values in T(x):
Temperatures from Analytical solution Temperatures from numerical soluion
Node 0: T(0) = 500°C T 0 = 500°C
Node 1: T(0.002) = 492°C T 1 = 492°C
Node 2: T(0.004) = 468°C T 2 = 468°C
Node 3: T(0.006) = 428°C T 3 = 428°C
Node 4: T(0.008) = 372°C T 4 = 372°C
Node 5: T(0.01) = 300°C T 5 = 300°C
So, we see that values of temperatures obtained by numerical methods match extremely well with the values ob-
tained by ‘exact’ analytical solution, i.e. even with only 5 equal divisions of the slab, we get very accurate solution by
numerical method. Hence its popularity.
Example 8.2. Consider a slab of thickness, L = 1 cm. Thermal conductivity of the slab material varies linearly with
temperature as: k(T) = 26.679 (1 + 8.621 ´ 10 –4 T), W/(mC), where T is in deg. C. Surface at x = 0 is insulated and the
other surface at x = L is subjected to a convection heat transfer with a fluid at 100°C with a heat transfer coefficient of
4000 W/(m2 C).There is uniform internal heat generation in the slab at a rate of 8 ´ 10 7 W/m3. Dividing the slab into 5
equally spaced sub-regions, find the temperatures at the different nodes. Assume one-dimensional, steady state conduc-
tion.
Solution.
Data.
L : = 0.01 m M := 5 k(T) = 26.679(1 + 8.621 ´ 10 –4 ×T) W/(mC)
This is of the form:k(T) := k 0 (1 + b×T)
where,k 0 := 26.679 W/(mC) and, b := 8.621 ´ 10 –4 1/C Ta := 100°C h := 4000 W/(m2 C) qg := 8 ´ 10 7 W/m3

L 0.01
Dx = = i.e. D x := 0.002 m
M 5
Note that there are 6 nodes, numbered as: 0, 1, 2, 3, 4, and 5. Out of these, nodes ‘0’ and ‘5’ are boundary nodes and
the nodes 1, 2, 3 and 4 are internal nodes.
Fig. Example 8.2 shows the schematic of finite difference nodes for this problem.

338 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Now, for the interior nodes, Eq. 8.8 is not applicable since the thermal conductivity varies with temperature.
Let us derive the difference equation for the interior nodes first. Consider any internal node ‘m’ and apply the
energy balance for the differential volume around node ‘m’. Remember to consider that all energy flows are into the
control volume. Using the thermal resistance concept, we get:
Tm - 1 - Tm Tm + 1 - Tm
+ + qm ×(A×D x) = 0
Dx Dx
LM F IJ OP × A
Tm - 1 + Tm LM F Tm + 1 + Tm IJ OP × A
G
MN H
k0 × 1 + b ×
K PQ
2 MN GH k0 × 1 + b ×
2 K PQ
L F T + T I OP + (T – T )×k × LM1 + b × F T + T IO
i.e. (T m – 1 – T )×k × M1 + b × G
m 0
MN H 2 JK PQ
m-1 m

MN GHm+1 m 0
m+1

2
m
JK PPQ + q ×(Dx)
m
2
= 0

b q ×( D x ) 2
i.e. (T m – 1 – 2×T m + T m + 1) + ×[(Tm – 1) 2 – 2×(Tm)2 + (Tm + 1)2] + m = 0 ...(A)
2 k0
qm
Eq. A gives the difference equation for the interior nodes 1, 2,
3, and 4. It is seen that this equation is non-linear and solving the Qleft Qright
set of non-linear equations by conventional methods is difficult.
But, as we shall presently see, in Mathcad, it is very easy to get T0 T1 Tm 1 Tm Tm + 1 T5
solution using the ‘solve block’. x

qg ×( D x) 2
0 1 2 3 4 5
We have: q m = qg and, = 11.994
k0
x=0 Dx x=L
In Eq. A, let us put m =1, 2, 3 and 4 to get the difference equa-
tions for the respective nodes:
FIGURE Example 8.2 Finite difference nodes
for Example 8.2
b
Node 1: (T 0 – 2×T1 + T2) + ×[(T 0)2 – 2×(T1)2 + T2)2] + 11.994 = 0 ...(b)
2
b
Node 2: (T 1 – 2×T2 + T3) + ×[(T1)2 – 2×(T2) 2 + T3) 2] + 11.994 = 0 ...(c)
2
b
Node 3: (T 2 – 2×T3 + T4) + ×[(T2)2 – 2×(T3) 2 + T4) 2] + 11.994 = 0 ...(d)
2
b
Node 4: (T 3 – 2×T4 + T5) + ×[(T3)2 – 2×(T4) 2 + T5) 2] + 11.994 = 0 ...(e)
2
Difference equations for boundary nodes:
For node ‘0’: Apply the energy balance to the half-volume around the node ‘0’; all heat lines flowing into the
volume.
There is no heat flowing from the left side of the control volume into node ‘0’ since the surface is insulated. Writing
other terms, we get:
T1 - To FG Dx IJ
Dx H
+ qg × A ×
2 K = 0

LMk × L1 + b × FG T + T IJ O × AOP
MN MN H 2 K PQ PQ
0 1
0

L F T + T IJ OP + q ×(D x)
(T – T )× k × M1 + b × G
g
2

N H 2 KQ 2
0 1
i.e. 1 0 0 = 0

b q g × ( D x) 2
i.e. (T 1 – T 0) + ×(T 12 – T 02) + = 0
2 2 ×k0
b
i.e. ×(T 12 – T 02) + 5.997 = 0
(T 1 – T0) + ...(a)
2
Eq. a is the difference equation for node ‘0’. This equation is also a non-linear equation

NUMERICAL METHODS IN HEAT CONDUCTION 339


For node 5: Apply the energy balance to the half-volume around the node 5; all heat lines flowing into the volume.
There is convection condition on the right surface. Writing the energy balance, we get:
T4 - T5 FG Dx IJ
Dx H
+ h ×A (Ta – T 5) + q g × A ×
2 K = 0

LMk × L1 + b × FG T + T5 IJ OP × AOP
MN MN H K Q PQ
4
0
2

h ×(Ta - T5 ) × D x q g × ( D x)
2
b
i.e. (T 4 – T 5) + × (T 42 – T 52) + + = 0
2 k0 2× k0

b
i.e. ×(T 42 – T52) – 0.3×T5 + 29.986 + 5.997 = 0
(T 4 – T5) + ...(f)
2
Eq. f is the difference equation for node 5. This equation is also non-linear.
Now, we have got 6 equations, namely Eqs. a, b...f and there are 6 unknown node temperatures. So, solving these 6
coupled equations simultaneously, we get the temperatures T 0, T 1 ... T6.
We use ‘solve block’ of Mathcad to solve these equations.
We start with assumed or trial values for all the variables i.e. for the temperatures at nodes 0 to 5. Then, in the solve
block, immediately below ‘Given’ write all the constraint equations. Then, the command ‘Find (T0, T1, T2 ... T 5)’ immedi-
ately gives a vector of temperature values:
T0 := 50 T1 := 50 T2 := 50 T3 := 50 T4 := 50 T5 := 50 (trial values of temperatures)
Given
b
(T1 – T 0) + ×(T12 + T02) + 5.997 = 0 ...(a)
2
b
(T 0 – 2×T1 + T2) + ×[(T0)2 – 2×(T1) 2 + T2) 2] + 11.994 = 0 ...(b)
2
b
(T 1 – 2×T2 + T3) + ×[(T1)2 – 2×(T2) 2 + T3) 2] + 11.994 = 0 ...(c)
2
b
(T 2 – 2×T3 + T4) + ×[(T2)2 – 2×(T3) 2 + T4) 2] + 11.994 = 0 ...(d)
2
b
(T 3 – 2×T4 + T5) + ×[(T3)2 – 2×(T4) 2 + T5) 2] + 11.994 = 0 ...(e)
2
b
(T 4 – T5) + × (T42 – T52) – 0.3×T 5 + 29.986 + 5.997 = 0 ...(f)
2
Temp: = Find(T0, T1, T2, T3, T 4, T5) (‘Temp’ is the vector containing values of temperatures T0, T1, ... T5)
Therefore,

LM414.482OP
MM 410.058PP
Temp = M
MM .203PPP
396.709
374

MM299
342 .128
P
N .853 PQ
i.e. Temperatures at different nodes are:
T 0 = 414.482°C T1 = 410.058°C T2 = 396.709°C T3 = 374.203°C T4 = 342.128°C T5 = 299.853°C
When there is no analytical solution to compare the results obtained by numerical methods, the number of sub-
divisions can be increased and the results obtained with the increased sub-divisions may be compared with the earlier
results; and this process may be continued till the difference between the successive results converges to a pre-deter-
mined accuracy. Better alternative is to make a heat balance check: In this case, since the left side is insulated, all the heat
generated in the slab must be dissipated at the right surface to the fluid by convection. Heat generated per 1 m2 of area
= (8 ´ 10 7) ´ (1 ´ 0.01) = 8 ´ 10 5 W, and the heat transferred by convection from the right face to the fluid = h.A.DT = 4000
´ 1 ´ (299.853 – 100) = 7.994 ´ 105 W; i.e. heat generated = heat dissipated by convection.

340 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Example 8.3. A straight fin of rectangular cross-section has length L = 3 cm, thickness t = 0.5 cm and width w = 10 cm.
Thermal conductivity of fin material, k = 20 W/(mC). Temperature at the base of the fin is T 0 = 200°C and there is
negligible heat transfer from the tip of the fin. The fin dissipates heat from its surfaces into the surroundings at 25°C with
a heat transfer coefficient of 15 W/(m2 C). Using the finite difference method with 10 equally spaced sub-divisions, each
of length Dx = 0.3 cm, determine:
(a) temperatures at the nodes
(b) rate of heat transfer from the fin, and
(c) fin efficiency
Solution.
Data.
L := 0.03 m w := 0.1 m t := 0.005 m M := 10 k : = 20 W/(mC) T0 := 200°C Ta := 25°C
L 0 .03
h := 15 W/(m2 C) Ac := w×t i.e. Ac = 5 ´ 10 –4 m2 Dx := = i.e. Dx = 0.003 m
M 10
Note that there are 11 nodes, numbered as: 0, 1, 2, 3, 4, …10. Out of these, nodes ‘0’ and ‘10’ are boundary nodes
and the other nodes are internal nodes. Temperature of node ‘0’ is given, i.e. T0 = 200°C.
Fig. Example 8.3 shows the schematic of finite difference nodes for this problem.

h = 15 W/(mC) Insulated
Ta = 25°C
To
t = 0.005 m

L = 0.03 m
qconv

Qleft Qright
T0 T1 Tm 1 Tm Tm + 1 T10
x

Dx 10
0 1 2 Dx

x=0 x=L

FIGURE Example 8.3 Finite difference nodes for Example 8.3

Difference equations for internal nodes:


Consider a typical internal node ‘m’ and write an energy balance for the differential volume represented by node ‘m’.
Remember that all heat flows are into the control volume. There is conduction from nodes (m – 1) and (m + 1) into node
m and also there is heat flow by convection from the ambient into the control volume:
FT - Tm I + k× A × F T - Tm I + h ×((2×w + 2×t)×D x)×(T
k×Ac × GH
m-1

Dx JK GH
c
m+1

Dx JK a – Tm) = 0

2 × h × ( w + t ) ×( D x ) 2 × (Ta - Tm )
i.e. Tm – 1 – 2 ×Tm + Tm + 1 + = 0 ...eqn. (A)
k × Ac
Eq. A gives the finite difference equation for the internal nodes, m = 1, 2, ... 9.
We have:
2 × h × ( w + t ) ×( D x ) 2 × (Ta - Tm )
Tm – 1 – 2 ×Tm + Tm + 1 + = 0 ...eqn. (A)
k × Ac
Putting m = 1 etc.,
Node 1: T0 – 2×T1 + T2 – 2.835 ´ 10 –3 × T1 + 0.071 = 0 ...(b)
Node 2: T1 – 2×T2 + T3 – 2.835 ´ 10 –3 × T2 + 0.071 = 0 ...(c)
Node 3: T2 – 2×T3 + T4 – 2.835 ´ 10 –3 × T3 + 0.071 = 0 ...(d)
Node 4: T3 – 2×T4 + T5 – 2.835 ´ 10 –3 × T4 + 0.071 = 0 ...(e)
Node 5: T4 – 2×T5 + T6 – 2.835 ´ 10 –3 × T5 + 0.071 = 0 ...(f)

NUMERICAL METHODS IN HEAT CONDUCTION 341


Node 6: T 5 – 2×T6 + T7 – 2.835 ´ 10 –3 × T6 + 0.071 = 0 ...(g)
Node 7: T 6 – 2×T7 + T8 – 2.835 ´ 10 –3 × T7 + 0.071 = 0 ...(h)
Node 8: T 7 – 2×T8 + T9 – 2.835 ´ 10 –3 × T8 + 0.071 = 0 ...(i)
Node 9: T 8 – 2×T9 + T10 – 2.835 ´ 10 –3 ×T 9 + 0.071 = 0 ...(j)
Difference equations for boundary nodes:
For node ‘0’: By data, temperature of node ‘0’ is the temperature of base surface:
i.e. T 0 = 200°C ...(a)
For node 10: Consider the half-volume surrounding node ‘10’ and write the energy balance, with all heat flow lines
into the volume. Remember that heat flow from right of control volume into the node 10 is zero, since the surface is
considered as insulated:
F T - T I + h× L(2 × w + 2 ×t)× D x O × (T
k×Ac × GH D x JK MN
9 10

2 PQ
a – T 10) = 0

h × ( w + t ) × ( D x) 2
i.e. T9 – T 10 + × (Ta – T 10) = 0
k × Ac
i.e. T9 – T 10 – 1.418 ´ 10 –3 ×T 10 + 0.035 = 0 ...(k)
Eqs. a to k give 10 equations for the 10 node temperatures. Solving these equations simultaneously, we get the node
temperatures.
We use ‘solve block’ of Mathcad to solve these equations.
We start with assumed or trial values for all the variables i.e. for the temperatures at nodes 0 to 10. Then, in the
solve block, immediately below ‘Given’ write all the constraint equations. Then, the command ‘Find (T 0, T 1, T 2 ... T10)’
immediately gives a vector of temperature values:
T 0 := 200 (by data)
T1 := 50 T2 := 50 T3 := 50 T4 := 50 T5 := 50 (trial values of temperatures)
T6 := 50 T7 := 50 T8 := 50 T9 := 50 T10 := 50 (trial values of temperatures)
Given
T 0 = 200 C ...(a)
T 0 – 2× T1 + T 2 – 2.835 ´ 10 –3 ×T 1 + 0.071 = 0 ...(b)
T 1 – 2× T2 + T 3 – 2.835 ´ 10 –3 ×T 2 + 0.071 = 0 ...(c)
T 2 – 2× T3 + T 4 – 2.835 ´ 10 –3 ×T 3 + 0.071 = 0 ...(d)
T 3 – 2× T4 + T 5 – 2.835 ´ 10 –3 ×T 4 + 0.071 = 0 ...(e)
T 4 – 2× T5 + T 6 – 2.835 ´ 10 –3 ×T 5 + 0.071 = 0 ...(f)
T 5 – 2× T6 + T 7 – 2.835 ´ 10 –3 ×T 6 + 0.071 = 0 ...(g)
T 6 – 2× T7 + T 8 – 2.835 ´ 10 –3 ×T 7 + 0.071 = 0 ...(h)
T 7 – 2× T8 + T 9 – 2.835 ´ 10 –3 ×T 8 + 0.071 = 0 ...(i)
T 8 – 2×T 9 + T 10 – 2.835 ´ 10 –3 ×T 9 + 0.071 = 0 ...(j)
T9 – T 10 – 1.418 ´ 10 –3 ×T 10 + 0.035 = 0 ...(k)
Temp := Find(T0, T1, T2, T3, T4, T5, T6, T7, T 8, T9, T 10) ...(a)
Therefore, (Temp. is the vector containing values of temperatures T0, T1, ...T10)

0
0 200
1 195.707
2 191.899
3 188. 563
4 185.691
5 183. 274
6 181. 306
Temp =
7 179.78
8 178.694
9 178.043
10 177 .826

342 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Temperatures at different nodes are:
T0 = 200°C T6 = 181.306°C
T1 = 195.707°C T7 = 179.78°C
T2 = 191.899°C T8 = 178.694°C
T3 = 188.563°C T9 = 178.043°C
T4 = 185.691°C T10 = 177.826°C
T5 = 183.274°C
Analytical solution: For the case of a fin, with an insulated end, analytical expression for temperature distribution along
the length is:
h. p
For fin: P := 2×w + 2×t ...perimeter, and m = i.e. m = 17.748 1/m ...fin parameter
k . Ac

cos h (m × (L - x))
T(x) := Ta + (T0 – Ta )× (temperature distribn. in a fin with insulated end)
cos h (m × L)
Temperatures from Analytical solution Temperatures from numerical soluion
Node 0: T(0) = 200 T 0 = 200°C
Node 1: T(0.003) = 195.706 T 1 = 195.707°C
Node 2: T(0.006) = 191.896 T 2 = 191.899°C
Node 3: T(0.009) = 188.559 T 3 = 188.563°C
Node 4: T(0.012) = 185.686 T 4 = 185.691°C
Node 5: T(0.015) = 183.269 T 5 = 183.274°C
Node 6: T(0.018) = 181.3 T 6 = 181.306°C
Node 7: T(0.021) = 179.775 T 7 = 179.78°C
Node 8: T(0.024) = 178.688 T 8 = 178.694°C
Node 9: T(0.027) = 178.038 T 9 = 178.043°C
Node 10: T(0.03) = 177.821 T 10 = 177.826°C
We observe that values of temps. obtained by numerical methods match extremely well with the values obtained by
‘exact’ analytical solution, i.e. with only 10 equal divisions of the fin length, we get very accurate solution by numerical
method.
Heat transferred by the fin, Q fin:
Q fin must be equal to the amount of heat entering into the fin at its base.
Write the heat balance for the half-volume around node ‘0’:

LM O
T -T P
+ M
MM D x PPP + h×(2×w + 2×t)× 2 × (T – T ) = 0
Dx
Q fin 1 0
a 0

N k×A Q
c

LM O
T -T P
:= M
MM D x PPP – h×(2× w + 2×t)× 2 (T – T )
Dx
i.e. Q fin 0 1
a 0

N k×A Q
c

i.e. Q fin = 15.137 W.


Fin efficiency, h f :
Fin efficiency is the ratio of actual heat transferred by the fin to the maximum heat that would be transferred if the entire
fin surface were at the base temp.
Qfin
hf =
Qmax
Q max: = h×(2 ×w + 2 ×t)×L×(T0 – Ta) W (maximum heat transfer, if the entire fin were at base temperature)
i.e. Q max = 16.538 W (maximum heat transfer by fin)
Qfin
Therefore, hf: =
Qmax
i.e. h f = 0.915 = 91.5% (fin efficiency)

NUMERICAL METHODS IN HEAT CONDUCTION 343


8.4 Methods of Solving a System of Simultaneous, Algebraic Equations
From what we have studied so far, it is clear that while solving steady state heat conduction problems by finite
difference formulation, we get a set of algebraic equations by applying the energy balance to the various nodes
and this set has to be solved simultaneously to obtain the temperatures at different nodes. There are many equa-
tion solvers and powerful software using which one can obtain the solution easily without knowing the intrica-
cies of the methods involved. While solving problems, you might have noticed that we used Mathcad, in which
solving even non-linear algebraic equations was extremely simple. Also, ready computer programs and sub-
routines are available for users, who have to simply change one or two lines of the program to adapt the solution
for their particular problem. For solving simultaneous, linear algebraic equations, subroutines such as LEQT1F,
LEQT1B, LEQT2F, LEQT2B supplied by the International Mathematical and Statistical Libraries (IMSL) have
been popular in scientific community.
Still, it is worthwhile to know the basics of different methods involved in solving a set of algebraic equa-
tions. We shall briefly present a few methods:
(i) Relaxation method
(ii) Direct methods: (a) Gaussian elimination , and (b) Matrix inversion
(iii) Iterative methods, e.g. Gauss – Siedel iteration method
(i) Relaxation method This is basically a trial and error solution and does not require a computer. But, it is
practicable to use only when the number of equations is small, say, less than 10. As an example, consider a set of
following three algebraic equations:
a 1 ×x + b 1 ×y + c 1 ×z = 0
a 2 ×x + b 2 ×y + c 2 ×z = 0
a 3 ×x + b 3 ×y + c 3 ×z = 0
The coefficients a 1, b 1, …, c 1, etc. are known, and our aim is to solve this set for x, y and z. Then, the ‘Relaxa-
tion technique’ consists of the following steps:
(a) To start with, assume values for x, y and z.
(b) Since the assumed values are certainly likely to be in error, each of the above equations will not be zero,
but equal to some residual values R 1, R 2 and R 3:
a 1 ×x + b 1 ×y + c 1 ×z = R 1
a 2 ×x + b 2 ×y + c 2 ×z = R 2
a 3 ×x + b 3 ×y + c 3 ×z = R 3
(c) Our aim is to reduce R 1, R 2 and R 3 to zero by suitably varying the assumed values of x, y and z, by trial
and error. This is done systematically, by first setting up a ‘unit change table’, i.e. a table showing the
change in the values of residuals for unit change in x, y and z.
(d) Set up a ‘Relaxation table’ wherein you begin with the initially assumed values of x, y and z and the
resultant residuals. Then, start ‘relaxing’ the largest residual by suitably changing the value of x, y or z,
taking guidance from the ‘unit table’ already set up.
(e) Continue the procedure till all the residuals are relaxed to zero.
Obviously, this procedure is slow and time consuming and cannot be used when the number of equations to
be solved is large.
(ii) Direct methods Direct methods have a fixed number of well defined steps to systematically solve the equa-
tions for the unknown values. However, they consume more of computer memory and time compared to itera-
tive methods, and are suitable for comparatively small number of equations. Under ‘direct methods’, we shall
study two methods: (a) Gaussian elimination method, and (b) Matrix inversion method:
(a) Gaussian elimination method In this method, used for solution of a system of linear algebraic equa-
tions, one of the unknowns is eliminated systematically in each step, and at the end of the elimination
process, the last equation involves only one unknown, and then the remaining unknowns are obtained
one by one by ‘back substitution’. To make the process clear, let us consider a simple system of three
algebraic equations, as given below:
x + 2×y + 3×z = 33 ...(a)
x – 4×y + z = – 11 ...(b)
3×x + y + z = 18 ...(c)
Now, we ‘triangularize’ the given set of equations by repeated application of three basic row operations:

344 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i) multiplication of a row by a constant (ii) adding one row to another row, and (iii) interchange of two
rows.
In the above, use Eq. a to elimlnate x from Eqs. b and c, by adding –1 times (a) to (b) and by adding –3
times (a) to (c). We get:
x + 2×y + 3×z = 33 ...(a¢)
–6×y – 2×z = 44 ...(b¢)
– 5×y – 8×z = 81 ...(c¢)
Next, eliminate y from Eq. c¢ by multiplying Eq. b¢ by –5/6 and adding to Eq. c¢:
We get:
x + 2×y + 3×z = 33
6×y + 2×z = 44
z=7
Above set of equations is known as ‘triangularized set’ of equations.
Having obtained the value of z, now back-substitute in the previous equations to get value of y as y = 5,
and one more ‘back-substitution’ in the preceding equation gives the value of x as x = 2.
Since we had only three equations in the above set, we could do the elimination or triangularization by
hand. However, Gaussian elimination method for a system of large number of equations is done with a
computer, using matrix notation to represent the equation. Coefficients constitute a square matrix called
‘coefficient matrix’ and the constant terms are stored in a vector called ‘right hand side vector’. Compu-
tation sub-routines normally combine these two into a single ‘augmented matrix’ and the above proce-
dure is done by the computer program to eliminate the terms below the main diagonal of the augmented
matrix. This results in a matrix of ‘upper diagonal form’. Then, back-substitution is performed by the
program systematically to get the solution.
So, for example, in the above set of equations, the augmented matrix will be:

LM1 2 3 33 OP
MM1 - 4 1 - 11 PP
N3 1 1 18 Q
Now, resorting to already mentioned row operations on this matrix, elements under the main diagonal
are eliminated and the upper diagonal form of the matrix is obtained as:

LM1 2 3 33 OP
MM0 6 2 44 PP
N0 0 1 7 Q
Last row means that z = 7. Now, back-substitution is done to get values of y and x. Gaussian elimination
method is conveniently programmed in a computer and ready subroutines are available to solve a set of
N linear algebraic equations simultaneously.
(b) Matrix inversion method In this method, the set of equations is written in the following matrix form:
[A] [T] = [B], where
[A] is the coefficient matrix, [T] is the vector of temperatures to be found out, and [B] is the vector of
constants (RHS) of the equations. Solution of this system by matrix inversion method is given by:
[T] = [A]–1 [B], where [A] –1 is the inverse of matrix [A].
Matrix inversion is performed generally by using readily available computer subroutines. In Mathcad,
inverse of a matrix A is obtained in a single step by the command A –1 =.
For the problem illustrated above, we have:

LM1 2 3 OP LM 33 OP
A = M1 -4 1 PP B = M - 11P
MN3 1 1 Q MN 18 PQ

NUMERICAL METHODS IN HEAT CONDUCTION 345


and,

LM - 5 1 7 OP
MM 381 38 19
-4 1
PP
A –1 ®
MM 19 19 19 PP (inverse of A, from Mathcad)

MM 13 5 -3 PP
N 38 38 19 Q
Therefore,
T := A –1 ×B (T is the vector containing x, y, z as its elements)
LM2OP
i.e. T= 5 MM PP
N7 Q
which means that x = 2, y = 5 and z = 7.
This result is the same as obtained earlier.
Once again, when the number of equations is relatively large, this is not a preferred method, from the
point of view of computer memory and storage.
(iii) Gauss–Siedel iteration method Iteration methods are used when the number of algebraic equations to be
solved is relatively large. Gauss–Siedel iteration (also called Liebmann iteration) method is one of the most popu-
lar iteration methods because of its simplicity. The method involves the following steps:
(a) Solve each equation for one of the unknowns, i.e. write each unknown in terms of other unknowns
(b) Assume guess values for all unknowns, and from the equations developed in step (a), compute the un-
knowns, each time using the most recently computed values for the unknowns in each equation
(c) Repeat this procedure until the successive values of an unknown converge to a specified accuracy.
To illustrate this procedure, let us consider the example given below. We have a set of equations as follows:
3×x – y + 3×z = 0 ...(a)
–x + 2×y + z = 3 ...(b)
2×x – y – z = 2 ...(c)
Now, write each equation for one of the unknowns. i.e.
y - 3× z
x=
3
(3 + x - z)
y=
2
z = – 2 + 2×x – y
Now, assume guess values for x, y and z. Say, x = 1, y = 1 and, z = 1. These are the ‘zeroth’ iteration values.
With these guess values, begin the iteration and in each equation, use the latest values of unknowns as
available. So, after ‘first’ iteration we have:
x=1 y=1 z=1 (inital guess values)
y - 3× z
x= i.e. x = – 0.667 (with y = 1, z = 1)
3
(3 + x - z)
y= i.e. y = 0.667 (with x = – 0.667, z = 1)
2
z = – 2 + 2×x – y i.e. z = – 4 ...with x = – 0.667, y = 0.667
Now, for the ‘second’ iteration, continue the procedure, with the latest values of unknowns. We get:
x = – 0.667 y = 0.667 z=–4 (next guess values from previous iteration)
y + 3× z
x= i.e. x = 4.222 (with y = 0.667, z = – 4)
3

346 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(3 + x - z)
y= i.e. y = 5.611 (with x = 4.222, z = – 4)
2
z = –2 + 2×x – yi.e. z = 0.833 (with x = 4.222, y = 5.611)
For the ‘third’ iteration, take x = 4.222, y = 5.611 and z = 0.833, and continue. This process is programmed
easily in a computer and the results normally converge within about 100 iterations. Of course, we can also in-
struct the program to stop when the difference between successive values of unknowns converge to a pre-deter-
mined accuracy.
A simple Mathcad program to perform the above iteration is shown below. It does the iteration 100 times.
Final values of x, y and z are returned as a vector R.

R : = x0 ¬ 1
y0 ¬ 1
z0 ¬ 1
for i Î 0.. 100
y - 3 × zi
xi + 1 ¬ i
3
(3 + xi + 1 - zi )
yi + 1 ¬
2
z i + 1 ¬ ( - 2 + 2 × xi + 1 - yi + 1 )

LMxi + 1 OP
MMyi + 1 PP
N zi + 1 Q
LM 2 OP
And, R= MM 3 PP
N- 1Q
which means that x = 2, y = 3 and z = –1.
In the above program, LHS defines a vector R. On the RHS, there are 10 lines. First three lines assign the
initial guess values for x, y and z. Next 4 lines show the ‘for loop’, for 100 iterations, where in x, y and z are
calculated, each time using the latest available values of unknowns. Next 3 lines constitute the latest values of x,
y and z which are stored as the elements of the vector R.
It is interesting to note that in the above program, if iteration is carried out only for 5, 10, 20, 50 and 100
loops (by changing the 4th line), following are the results:
After 5 After 10 After 20 After 50 After 100
iterations iterations iterations iterations iterations

LM 1.755 OP LM 1.987 OP LM 2 OP LM 2 OP LM 2 OP
R = M 2.718 P R = M 2.982 P R= M3P R= M3P R= M3P
MN- 1.208PQ MN- 1.008PQ MN- 1PQ MN- 1PQ MN- 1PQ
i.e. even with only 10 iterations we are very close to the final result. By the time 20 iterations are over,
solution has already converged to the final result.
It is stated that for steady state heat conduction problems, Gauss–Siedel iteration process is inherently stable
and always converges into a solution.
Note: Of course, above program can be further refined to stop when the successive values of x, y and z differ by
a pre-determined small value e . (say, e = 0.001).

NUMERICAL METHODS IN HEAT CONDUCTION 347


Above program in Mathcad is shown only to illustrate the procedure of iterative solution. While actually
using Mathcad, we would use the ‘solve block’ (which also follows an iterative algorithm), as follows:
x := 0 y := 0 z := 0 (guess values)
Given
3×x – y + 3×z = 0 ...(a)
–x + 2×y + z = 3 ...(b)
2×x – y – z = 2 ...(c)

LM2OP
Find (x, y, z) = M 3P
MN1PQ
You may put any guess value to start with; it makes no difference on the final result. However, it is essential
that each unknown is assigned some guess value to start with.
Accuracy of the solutions Some comments on the accuracy of finite difference solutions are appropriate. We
noted earlier that in solving heat conduction problems by finite difference methods, accuracy improves as the
number of nodes is made larger. However, this would mean that a larger number of algebraic equations have to
be solved simultaneously. This situation has following inherent drawbacks: the computer memory required in-
creases and also, more importantly, the round off errors in successive calculations increase since they are cumu-
lative. Therefore, one should start with a coarse mesh and then gradually refine it depending upon the accuracy
of final results required. Note that for the normal problems encountered in practice, a coarse mesh generally
gives results of acceptable accuracy; remember that, anyway, there are uncertainties in the values of thermal
properties and heat transfer coefficients available to the designer.

8.5 One-dimensional, Steady State Conduction in Cylindrical Systems


We shall now develop finite difference formulation by energy balance method, for one-dimensional, steady state
heat conduction in cylindrical coordinates.
Consider a long, solid cylinder of radius R in which the heat flow is only in the radial direction. Let the rate
of internal heat generation be qg (W/m3). The region from r = 0 to r = R is divided into M sub-regions, each of
thickness Dr = R/M. Therefore, there are
(M + 1) nodes, numbered as 0, 1, 2, …, M. See Fig. 8.5.
Writing an energy balance for the volume element around node ‘m’, remembering that all heat flows are into
the volume, we get; in steady state:
Rate of energy flowing into the volume from left + Rate of energy flowing into the volume from right + Rate
of heat generated in the volume = 0.
Substituting the values,
Tm - 1 - Tm Tm + 1 - Tm
+ + (2×p×m×Dr×Dr)×L×qm = 0
Dr Dr
FG IJ
Dr FG Dr IJ
H
2 ×p × m × D r -
2 K×L×k
H
2 ×p × m × D r +
2
×L×k
K
First term in the above equation is the heat flowing into node ‘m’ from node ‘(m – 1)’. Denominator of the
first term is the thermal resistance between ‘m’ and ‘(m – 1)’; it is written in the form (L/k A) where A is the mean
area i.e. area of the plane mid-way between nodes ‘m’ and ‘(m – 1)’. This form of thermal resistance (as if for a
plane wall), is alright for the cylindrical system when Dr << R, which is generally the case. Also, this makes the
equation simpler. Similarly, the second term in the above equation is the heat flowing into node ‘m’ from node
‘(m + 1)’. The third term gives the heat generated in the elemental volume. L is the length of the cylinder and qm
is the heat generation rate per unit volume for the elemental volume (= qg, generally).
Simplifying the above equation, we get:
F 1 - 1 I ×T F I ( D r )2 × qm
GH 2× m JK m–1 GH
– 2×Tm + 1 +
1
2× m JK
×Tm + 1 +
k
=0 ...(8.32)

Eq. 8.32 is the finite difference equation for internal nodes i.e. for nodes 1, 2, …,(M – 1), with constant
thermal conductivity and internal heat generation.
348 FUNDAMENTALS OF HEAT AND MASS TRANSFER
TM

TM 1
Volume element of node m

Cylinder
Tm + 1

Tm qm.DV
Qright
Tm 1
Dr
T3

T2
Qlef t

T1

T0
r
0 1 2 3 m 1 m m+1 M 1 Dr/2
Dr/2
m 1/2 m + 1/2 Dr M
r=0 r=R

FIGURE 8.5 Finite difference formulation in a cylindrical/spherical system

At the centre: i.e. at r = 0:


Writing the energy balance for the half-volume (of thickness Dr/2) around node ‘0’, we get:
T1 - T0 FG IJ
Dr
2

Dr
+ p×
2 H K ×L×q0 = 0
Dr
2 ×p × × L× k
2
In the above, first term is the heat conduction rate from node ‘1’ to node ‘0’ and the second term is the heat
generation term. q 0 is the heat generation rate per unit volume at node ‘0’ (= qg, generally). Simplifying the above
equation, we get:

( D r )2 × q0
4×(T1 – T0) + =0 ...(8.33)
k
Eq. 8.33 gives the finite difference equation for the centre node ‘0’, with constant thermal conductivity and
internal heat generation.
At the periphery: i.e. at node ‘M’:
As in the earlier cases, here too, finite difference equation is obtained by applying the energy balance to the half-
volume around node ‘M’. Of course, the nature of equation depends on the boundary condition, i.e. if it is pre-
scribed temperature, or prescribed heat flux or convection boundary condition. For convection boundary
conditions, where heat transfer from the periphery is with an ambient at temperature Ta with a heat transfer
coefficient of h, energy balance around node ‘M’, gives:
TM - 1 - TM Dr
+ (2×p ×M×Dr×L)×h×(Ta – TM) + 2×p ×M×Dr× ×L×qM = 0
Dr 2
FG Dr IJ
H
2 ×p × M × D r -
2
×L×k
K
In the above, first term is the heat conduction rate from node ‘(M – 1)’ to node ‘M’ and the second term is the
convective heat transfer between the periphery and the ambient, and the third term is the heat generation term.

NUMERICAL METHODS IN HEAT CONDUCTION 349


qM is the heat generation rate per unit volume at node ‘M’ ( = qg , generally). Simplifying the above equation, we
get:

F 1 - 1 I ×T LMF 1 - 1 I + D r × h OP ×T D r×h ( D r )2 × q M
GH 2× M JK M-1 –
MNGH 2 × M JK k PQ M +
k
×Ta +
2×k
=0 ...(8.34)

Eq. 8.34 gives the finite difference equation for the boundary node ‘M’, with convection conditions, constant
thermal conductivity and internal heat generation.
Example 8.4. A nuclear fuel element is in the form of a hollow cylinder insulated at the inner surface. Its inner and outer
radii are 5 cm and 10 cm respectively. The outer surface gives heat to a fluid at 50°C where the unit surface conductance
is 100 W/(m2 K). k of the material is 50 W/(mK). Dividing the shell into 5 equal sub-regions, find the temperatures at
different nodes. What is the maximum temperature in the
system? Given: Rate of heat generation in the fuel element
k = 50 W/(mK) is 3.796 ´ 10 5 W/m3.
5
qg = 3.796 ´ 10 W/m
3 Solution. See Fig. Example 8.4 (a).
Data:
To = ? ri := 0.05 m ro := 0.1 m k := 50 W/mK
Ta := 50°C h := 100 W/(m2 K) L := 1 m
qg := 3.796 ´ 105 W/m3 M := 10
ro - ri
2 Dr := i.e. Dr = 0.01 m
ha = 100 W/(m K) 5
Ta = 50°C Fig. 8.4 (b) shows the schematic for finite difference
representation:
Insulated
Note: Temperature on inner surface is the maximum
temperature since inner surface is insulated.
ri = 0.05 m Here, nodes 5 and 10 are boundary nodes and nodes 6, 7, 8
and 9 are internal nodes. Node 5 is on the insulated surface
ro = 0.1 m
and node 10 is on the outer surface with convection bound-
ary condition. By this system of numbering, r coordinate of
FIGURE Example 8.4(a) Hollow cylinder with heat node ‘m’ is m.Dr and M = 10 = no. of subdivisions in the
generation, losing heat on outer surface by convection, outer radius.
inner surface insulated Finite difference equations for internal nodes:
i.e. for nodes 6 to 9, we can apply Eq. 8.32, viz.

T10

T9 Volume element of node m


T8
Cylinder
T7
Tm qm.DV

T5 Qright
Dr

Qleft

r
0 5 m 7 8 Dr 9 10
ri = 0.05 m Dr/2
r=0 ro = 0.1 m r=R

FIGURE Example 8.4(b) Finite difference formulation in a cylindrical system

350 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F 1 - 1 I ×T F 1 I ( D r ) 2 × qm
GH 2 × m JK m–1 – 2 ×Tm + 1 + GH 2×m JK
×Tm + 1 +
k
= 0 ...(8.32)

Putting m = 6.....9, in the above equation we get difference equations for those respective nodes:
m = 6: 0.917×T5 – 2×T6 + 1.083×T7 + 0.759 = 0 ...(b)
m = 7: 0.929×T6 – 2 ×T7 + 1.071×T8 + 0.759 = 0 ...(c)
m = 8: 0.938×T7 – 2 ×T8 + 1.063×T9 + 0.759 = 0 ...(d)
m = 9: 0.944×T8 – 2×T9 + 1.056×T10 + 0.759 = 0 ...(e)
Finite difference equations for boundary nodes:
For node 5 Temperature on the insulated surface is maximum. Difference equation for node 5 is obtained by writing
heat balance on the half-volume around node 5, with all the heat flow lines going into the volume. From the left there is
no heat flow since the surface is insulated. So, there is heat flow only from right and there is heat generation term:
LMF IJ 2
OP
MNGH
T6 - T5 Dr
i.e.
Dr
+ p × 5× D r +
2 K - (5 × D r)2 × L × q g = 0
PQ
FG Dr IJ
H
2 ×p × 5 × D r +
2
×L×k
K
In the above equation first term gives the heat flow into the half-volume and the second term is the heat generation
term for the half-volume.
Simplifying, we get:
LM FG Dr IJ OP LMF 5 × D r + D r I 2
OP
N H
(T 6 – T 5)× 2 × p × 5 × D r +
K Q MNGH 2 K
J
× k + Dr×p× - ( 5 × D r ) 2 ×qg = 0
2 PQ
i.e. (T6 – T5)×17.279 + 6.261 = 0 ...(a)
For node 10 We can use Eq. 8.34; but, in this case, we can straightaway get the temperature at node 10 by a heat
balance at the outer surface, as follows:
All heat generated in the shell goes to the outer surface since the inner surface is insulated. Writing the heat balance
at the outer surface,
Heat generation in the shell = heat transferred by convection at the outer surface
i.e. p×(ro2 – ri2)×L×qg = h ×(2×p×ro ×L)×(T10 – Ta)
p × ( ro2 - ri2 ) × L × qg
i.e. T10 = + Ta
h × ( 2 ×p × ro × L)
i.e. T10 = 192.35°C ...eqn. (f) ...temperature at node 10
Now, we have 6 equations a to f. Solving them simultaneously, we get temperatures at 6 nodes, i.e. nodes 5 to 10.
We use ‘solve block’ of Mathcad to solve this set of equations. Start with guess values for all unknown temperatures
and immediately below ‘Given’, type the constraint equations. Then, the command ‘Find (T5, ..., T10)’ gives the tempera-
tures immediately:
T5 := 50 T6 := 50 T7 := 50 T8 := 50 T9 := 50 T10 := 192.35 (guess values of temperatures)
Given
(T6 – T 5)×17.279 + 6.261 = 0 ...(a)
0.917×T 5 – 2×T 6 + 1.083×T 7 + 0.759 = 0 ...(b)
0.929×T 6 – 2×T 7 + 1.071×T 8 + 0.759 = 0 ...(c)
0.938×T 7 – 2×T 8 + 1.063×T 9 + 0.759 = 0 ...(d)
0.944×T8 – 2×T 9 + 1.056×T10 + 0.759 = 0 ...(e)
T 10 = 192.35 ...(f)
Temp. := Find(T 5, T 6, T 7, T8, T9, T10) (define vector Temperature in which node temps. are stored)

LM200.371OP
MM200.008PP
Temp = M
MM197 .418PPP
199 .001
Therefore,

MM196 .122
P
N192.35 PQ
NUMERICAL METHODS IN HEAT CONDUCTION 351
i.e. The temperatures at different nodes are:
T 5 = 200.371°C T6 = 200.008°C T7 = 199.001°C T8 = 197.418°C T9 = 195.122°C T10 = 192.35°C
Check: We can check these results with the analytical solution. This is a problem of a cylindrical shell with the inside
surface insulated and outside surface losing heat by convection; in that case, temperature distribution is given by:

LM d
qg × ro2 - ri2 i + q × r × LMF r I
2 2
FG r IJ - FG r IJ 2
OPOP
MN
T(r): = Ta +
g i
G J
4 × k MH r K
o
- 2 × ln
H rK HrK
o

PQPQ ...(5.29)
2 × h × ro
N i i

(Analytical solution for temperature distribution)


Substitute different values for r and get temperatures at corresponding nodes:
Analytical solution Numerical soluion with 5 nodes
Node 5: T(0.05) = 200.007 T 5 = 200.371°C
Node 6: T(0.06) = 199.649 T 6 = 200.008°C
Node 7: T(0.07) = 198.645 T 7 = 199.001°C
Node 8: T(0.08) = 197.065 T 8 = 195.418°C
Node 9: T(0.09) = 194.956 T 9 = 195.122°C
Node 10: T(0.1) = 192.35 T 10 = 192.35°C
Just for comparison, if the shell is divided into 10 nodes, (numbered from 10 to 20), each of radial thickness 0.005
m, following will be the result, in comparison to the analytical results:
LM d
qg × ro2 - ri2 i + q × r × LMF r I
2 2
FG r IJ - FG r IJ 2
OPOP
MN
T(r) = Ta +
g i
G J
4 × k MH r K
o
- 2 × ln
H rK HrK
o
PQPQ (Analytical solution for temperature distribution)
2 × h × ro
N i i

Substitute for r and get temperatures at different nodes:


Analytical solution Numerical soluion with 10 nodes
Node 10: T(0.05) = 200.007 200.024°C
Node 11: T(0.055) = 199.915 199.931°C
Node 12: T(0.06) = 199.649 199.665°C
Node 13: T(0.065) = 199.223 199.237°C
Node 14: T(0.07) = 198.645 198.658°C
Node 15: T(0.075) = 197.924 197.936°C
Node 16: T(0.08) = 197.065 197.076°C
Node 17: T(0.085) = 196.075 196.083°C
Node 18: T(0.09) = 194.956 194.962°C
Node 19: T(0.095) = 193.714 193.717°C
Node 20: T(0.01) = 192.35 192.35°C
It may be noted that the values by finite difference methods match extremely well with the analytical results.

8.6 One-dimensional, Steady State Conduction in Spherical Systems


We shall now develop finite difference formulation by energy balance method, for one-dimensional, steady state
heat conduction in spherical coordinates.
Consider a solid sphere of radius R in which the heat flow is only in the radial direction. Let the rate of
internal heat generation be q g (W/m3). The region from r = 0 to r = R is divided into M sub-regions, each of
thickness Dr = R/M. Therefore, there are (M + 1) nodes, numbered as 0, 1, 2, …, M. See Fig. 8.5.
Writing an energy balance for the volume element around node ‘m’, remembering that all heat flows are into
the volume, we get; in steady state:
Rate of energy flowing into the volume from left + Rate of energy flowing into the volume from right + Rate
of heat generated in the volume = 0.
Substituting the values,
Tm - 1 - Tm Tm + 1 - Tm
+ + [4×p×(m×D×r)2 Dr]×q m = 0 ...(8.35)
Dr Dr
FG Dr IJ 2
FG Dr IJ 2

H
4 ×p × m × D r -
2 K ×k
H
4 ×p × m × D r +
2 K ×k

352 FUNDAMENTALS OF HEAT AND MASS TRANSFER


First term in the above equation is the heat flowing into node ‘m’ from node ‘(m – 1)’. Denominator of the
first term is the thermal resistance between ‘m’ and ‘(m – 1)’; it is written in the form (L/kA) where A is the mean
area i.e. area of the plane mid-way between nodes ‘m’ and ‘(m – 1)’. This form of thermal resistance (as if for a
plane wall), is alright for the spherical system when Dr << R, which is generally the case. Also, it makes the
equation simpler. Similarly, the second term in the above equation is the heat flowing into node ‘m’ from node
‘(m + 1)’. The third term gives the heat generated in the elemental volume. qm is the heat generation rate per unit
volume for the elemental volume (= qg, generally).
Simplifying the above equation, we get:

F1 - 1 I 2
F I 2
( D r )2 × qm
GH 2 × m JK ×(Tm – 1 – Tm ) + 1 + GH 1
2 ×m JK ×(T m + 1 – Tm ) +
k
=0 ...(8.36)

Eq. (8.36) gives finite difference equations for internal nodes, i.e. nodes 1, 2, …, (M – 1).
At the centre, r = 0:
Applying the energy balance to the half-volume around node ‘0’,

T1 - T0 4 FG IJ
Dr
3

Dr
+
3
×p ×
2H K × q0 = 0

FG D r IJ 2
4 ×p ×
H 2K ×k

Simplifying,

( D r )2 × q0
6×(T1 – T0) + =0 ...(8.37)
k
Eq. 8.37 gives finite difference equation for the centre i.e. node ‘0’.
For the boundary node ‘M’:
Difference equation for boundary node ‘M’ is written in the same manner as was done for other nodes, i.e. by
writing an energy balance on the half-volume around node ‘M’. The nature of relation obtained will depend
upon the boundary condition i.e. prescribed temperature, prescribed heat flux, or convection boundary condi-
tion.
As an example, let us write the difference equation for node ‘M’ when the convection conditions prevail at
the boundary. Let there be heat transfer at the boundary with a fluid flowing at a temperature of Ta with a heat
transfer coefficient of ‘h’. Then, writing an energy balance for the half-volume around node ‘M’, we get:

TM - 1 - TM Dr
+ 4×p (M×D×r) 2 ×h(Ta – TM) + 4×p×(M×D r)2 × ×q M = 0 ...(8.38)
Dr 2
FG Dr IJ 2

H
4 ×p × M × D r -
2 K ×k

In the above, first term is the heat conduction rate from node ‘(M – 1)’ to node ‘M’ and the second term is the
convective heat transfer between the outer surface and the ambient, and the third term is the heat generation
term. qM is the heat generation rate per unit volume at node ‘M’ (= qg, generally). Simplifying the above equation,
we get:

F1 - 1 I 2 LMF 1 I 2
D r×h OP D r×h ( D r )2 × qM
GH 2 × M JK ×TM – 1 –
MNGH 1 - 2× M JK +
k PQ
× TM +
k
× Ta +
2×k
=0 ...(8.39)

Eq. 8.39 gives the difference equation for the boundary node ‘M’ when convection conditions prevail at the
boundary.
Example 8.5. A solid sphere of radius, R = 10 mm and k = 18 W/(mC) has an uniform heat generation rate of 2 ´ 10 6 W/
m3. Heat is conducted away at its outer surface to ambient air at 20°C by convection, with a heat transfer coefficient of
2000 W/(m2 C). Using numerical method and dividing the radius into 10 equal sub-divisions,

NUMERICAL METHODS IN HEAT CONDUCTION 353


3
(i) deternine the steady state temperature at the centre and outer
k = 18 W/(mC), qg = 2000 W/m surface of the sphere.
Q (ii) draw the temperature profile along the radius.
Solution. Let us now solve this problem numerically, dividing the ra-
dius into 10 equal sub-divisions, i.e. M = 10. So, there are 11 nodes,
Ta = 20°C marked as 0, 1, ..., 10.
Data:
To R ha = 25 W(m.C)
R : = 0.01 m ha := 2000 W/(m2K) k := 18 W/(mK)
Ta := 20°C qg := 2 ´ 10 6 W/m 3 M := 10 Dr := 0.001 m
Refer to Fig. 8.5 for finite difference scheme of nodes for this prob-
Tw lem.
Node ‘0’ is the centre node, and node ‘10’ is the boundary node.
FIGURE Example 8.5 Solid sphere with Nodes 1, 2, ..., 9 are the internal nodes. There is convection condition at
heat generation the boundary.
For internal nodes 1, 2...,9: We apply Eq. 8.36 for internal nodes:

F1 - 1 I 2
F I 2
( D r )2 × q m
GH 2 × m JK ×(Tm – 1 – Tm) + 1 + GH 1
2×m JK ×(T m + 1 – Tm) +
k
= 0 ...(8.36)

Putting m = 1, 2,...9, we get difference equations for those respective nodes:


m = 1, Node 1: 0.25×(T0 – T1) + 2.25×(T2 – T1) + 0.111 = 0 ...(b)
m = 2, Node 2: 0.563×(T1 – T2) + 1.563×(T3 – T2) + 0.111 = 0 ...(c)
m = 3, Node 3: 0.694×(T2 – T3) + 1.361×(T4 – T3) + 0.111 = 0 ...(d)
m = 4, Node 4: 0.766×(T3 – T4) + 1.266×(T5 – T4) + 0.111 = 0 ...(e)
m = 5, Node 5: 0.81×(T4 – T5) + 1.21×(T6 – T5) + 0.111 = 0 ...(f)
m = 6, Node 6: 0.84×(T5 – T6) + 1.174×(T7 – T6) + 0.111 = 0 ...(g)
m = 7, Node 7: 0.862×(T6 – T7) + 1.148×(T8 – T7) + 0.111 = 0 ...(h)
m = 8, Node 8: 0.879×(T7 – T8) + 1.129×(T9 – T8) + 0.11 = 0 ...(i)
m = 9, Node 9: 0.892×(T8 – T9) + 1.114 ×(T10 – T9) + 0.11 = 0 ...(j)
For centre node ‘0’: Apply Eq. 8.37:
( D r ) 2 × q0
6×(T1 – T0) + = 0 ...(8.37)
k
i.e. 6×(T1 – T0) + 0.111 = 0 ...(a)
For boundary node 10: Since there is convection condition at the surface, Eq. 8.39 can be applied. However, in this
problem, since all the heat generated in the sphere has to be dissipated at the surface by convection, we can make an
energy balance at the surface and get the temperature at node 10 directly:
Making an energy balance at the surface of the sphere,
4
× p × R 3 × qg = ha ×(4×p× R2)× (T10 – Ta)
3
4
× p × R 3 × qg
i.e. T10 := Ta + 3 (define T 10)
ha × ( 4 × p × R 2 )
T10 = 23.333°C ...(k)
So, now we have 11 Eqs. a to k for 11 node temperatures T0 to T10 and by solving these equations simultaneously,
we get the temperatures at the different nodes.
We use ‘solve block’ of Mathcad to solve this set of equations. Start with guess values for all unknown temperatures
and immediately below ‘Given’, type the constraint equations. Then, the command ‘Find (T0,...T10)’ gives the tempera-
tures immediately:
T0 := 50 T2 := 50 T3 := 50 T4 := 50
T5 := 50 T6 := 50 T7 := 50 T8 := 50 T9 := 50 T10 := 23.333 (guess values of temperatures)
Given
6 ×(T1 – T0) + 0.111 = 0 ...(a)
0.25×(T0 – T1) + 2.25×(T2 – T1) + 0.111 = 0 ...(b)
0.563×(T1 – T2) + 1.563×(T3 – T2) + 0.111 = 0 ...(c)
0.694×(T2 – T3) + 1.361×(T4 – T3) + 0.111 = 0 ...(d)

354 FUNDAMENTALS OF HEAT AND MASS TRANSFER


0.766×(T3 – T4) + 1.266×(T5 – T4) + 0.111 = 0 ...(e)
0.81×(T4 – T5) + 1.21×(T6 – T5) + 0.111 = 0 ...(f)
0.84×(T5 – T6) + 1.174×(T7 – T6) + 0.111 = 0 ...(g)
0.862×(T6 – T7) + 1.148×(T8 – T7) + 0.111 = 0 ...(h)
0.879×(T7 – T8) + 1.129×(T9 – T8) + 0.11 = 0 ...(i)
0.892×(T8 – T9) + 1.114×(T10 – T9) + 0.11 = 0 ...(j)
T10 = 23.333°C ...(k)
Temp: = Find(T0, T 1, T 2, T 3, T 4, T 5, T 6, T 7, T 8, T 9, T 10)

0
0 25.163
1 25.144
2 25.093
3 25.003
4 24.876
5 24.712
6 24. 51
i.e. Temp =
7 24. 271
8 23. 994
9 23.682
10 23. 333

i.e. the node temperatures are: corresponding radius at the nodes


T0 := 25.163°C rad0 := 0.0 m
T1 := 25.144°C rad1 := 0.001 m
T2 := 25.093°C rad2 := 0.002 m
T3 := 25.003°C rad3 := 0.003 m
T4 := 24.876°C rad4 := 0.004 m
T5 := 24.712°C rad5 := 0.005 m
T6 := 24.51°C rad6 := 0.006 m
T7 := 24.271°C rad7 := 0.007 m
T8 := 23.994°C rad8 := 0.008 m
T9 := 23.682°C rad9 := 0.009 m
T10 := 23.333°C rad10 := 0.01 m
To compare numerical results with analytical temperature distribution:
For a sphere with internal heat generation, analytical expression for temperature distribution is:
qg × R qg
Temp(r): = Ta + + ×(R2 - r 2 ) ...(5.42)
3 × ha 6×k
Substitute different values for r and get temperatures at corresponding nodes:
Analytical solution Numerical soluion with 11 nodes
Node 0: T(0.0) = 25.185 T 0 = 25.163°C
Node 1: T(0.001) = 25.167 T 1 = 25.144°C
Node 2: T(0.002) = 25.111 T 2 = 25.093°C
Node 3: T(0.003) = 25.019 T 3 = 25.003°C
Node 4: T(0.004) = 24.889 T 4 = 24.876°C
Node 5: T(0.005) = 24.722 T 5 = 24.712°C
Node 6: T(0.006) = 24.519 T 6 = 24.51°C
Node 7: T(0.007) = 24.278 T 7 = 24.271°C
Node 8: T(0.008) = 24 T 8 = 23.994°C
Node 9: T(0.009) = 23.685 T 9 = 23.682°C
Node 10: T(0.01) = 23.333 T 10 = 23.333°C

NUMERICAL METHODS IN HEAT CONDUCTION 355


To sketch the temperature profile in the sphere, define a range variable r, varying from 0 to 0.01 m, with an
increment of 0.0005 m. Then, choose x–y graph from the graph palette, and fill up the place holders on the x-axis and y-
axis with r and Temp(r) respectively. Click anywhere outside the graph region, and immediately the graph appears.
Also, to sketch the temperature profile from numerical results in the same graph for comparison, define a range
variable i = 0 to 10 (nodes) and on the x-axis, next to r put a comma and fill up radi and on the y-axis fill up Ti after
putting a comma next to Temp(r), as shown.
Click anywhere outside the graph and both the graphs appear immediately: See Fig. Ex. 8.5(b)
r := 0, 0.0005, ..., 0.01 (define a range variable r..starting value = 0,
next value = 0.0005 m and last value = 0.01 m)
i := 0, ..., 10 (define a range variable, i varying from
0 to 10, with an increment of 1)

Temp. distr. for sphere with heat gen.


26

25.5

25 r in metres and
T(r) in deg. C
Temp (r)
Ti 24.5

24
analytical temperature distribution
temperature distribution by finite difference
23.5

23
0.01
0.001

0.002
0

0.004
0.003

0.006

0.007

0.008
0.009
0.005

r, radi

FIGURE Example 8.5(b) Temperature distribution in a solid sphere with heat generation

It is seen that temperature distribution obtained by numerical methods is very close to the analytical results.

8.7 Two-dimensional, Steady State Conduction in Cartesian Coordinates


So far, we considered numerical procedures for one-dimensional, steady state conduction, i.e. temperature gradi-
ents were significant only in one direction as compared to other directions. However, some times we come across
problems where temperature gradients are significant in more than one direction, familiar examples being in
large chimneys and L–shaped bars etc. In this section, we shall study the procedure of solving two-dimensional
steady state problems (with heat generation) by numerical methods.
Consider a two-dimensional system in which temperature gradients are significant in the x and y directions.
Let the x–y plane be subdivided into rectangular mesh of nodes, with spacing of Dx and Dy in x and y directions
respectively. Then, the nodes are numbered with a double subscript notation. i.e. a typical node, T m,n is the node
with a x-coordinate of (m.Dx) and y-coordinate of (n.Dy). Node count is m = 0, 1, …, M in the x direction and n =
0, 1, …, N in the y direction. See Fig. 8.6.
We see that there are basically three types of nodes: internal nodes, surface nodes, and corner nodes, marked
1, 2 and 3 respectively in the Fig. 8.6 (a).
Difference equations for different nodes are written in the usual manner by making an energy balance for
the elemental volume around the node in question, with all the heat flow lines going into the volume. Elemental
volumes for the internal node, surface node and corner nodes are shown by dotted lines around the nodes, in the
Fig. 8.6.

356 FUNDAMENTALS OF HEAT AND MASS TRANSFER


N 3

Tm, n Tm, n + 1
Tm, n
Dy
1 Tm Tm + 1, n
2 1, n

2
Tm, n 1
1
Typical internal node, Tm, n and
the surrounding nodes
0 1 2 M
Dx

(a) (b)

FIGURE 8.6 Finite difference representation for two-dimensional conduction-nodal network

Difference equations for internal nodes:


Consider a typical internal node, Tm,n in the x–y plane, with unit depth perpendicular to the plane of paper, as
shown in Fig. 8.6 (b). It is surrounded by 4 nodes: Tm-1, n, Tm, n +1, Tm +1, n, and Tm, n –1. Let us make an energy
balance on the elemental volume surrounding the node Tm, n . It is observed that heat flows into the node from all
the four directions, i.e. left, right, up and down. In addition, let there be heat generation in the volume at a rate of
(DV.qg), W, where qg, (W/m3), is the uniform volumetric heat generation rate in the system.
Writing the energy balance, in steady state,
Q left + Q right + Q up + Q down + DV×qg = 0 ...(8.40)
i.e.
Tm - 1, n - Tm , n Tm + 1, n - Tm , n Tm , n + 1 - Tm , n
k×Dy× + k×Dy× + k×Dx×
Dx Dx Dy
Tm , n - 1 - Tm , n
+ k×Dx× + qg ×Dx×Dy = 0
Dy
Simplifying, we get,
Tm - 1, n - 2 × Tm , n Tm + 1, n Tm , n - 1 - 2 × Tm , n + Tm , n + 1 qg
2
+ 2
+ =0 ...(8.41)
(D x) (D y ) k
Eq. 8.41 gives the difference equation for internal nodes, i.e. for m = 1, 2, …, (M – 1), and n = 1, 2, …, (N – 1).
Now, generally a square mesh is used i.e. Dx = Dy = (Dx, say). Then, the Eq. 8.41 simplifies to:
qg ×( D x)2
Tm – 1, n + Tm + 1, n + Tm, n + 1 + Tm, n – 1 – 4×Tm, n + =0 ...(8.42)
k
Eq. 8.42 is the finite difference equation for the internal nodes, with Dx = Dy.
Note that the first 4 terms in Eq. 8.42 are the temperatures of the surrounding 4 nodes, and the last term is
the heat generation term.
When there is no heat generation in the body, the difference equation for the node reduces to:
Tm - 1, n + Tm + 1, n + Tm , n + 1 + Tm , n - 1
Tm, n = ...(8.43)
4
i.e. when there is no heat generation, and a square mesh is used in the analysis, temperature of an internal node
is given as the arithmetic average of the surrounding four temperatures. It will be useful to remember this.
Difference equations for boundary nodes:
Boundary nodes may be on the surface or on the corners. Difference equations are developed for boundary nodes
in a similar manner as for interior nodes, i.e. by making an energy balance on the elemental volume surrounding

NUMERICAL METHODS IN HEAT CONDUCTION 357


the node. In Fig. 8.6 (a), we can see that the surface node 2 is surrounded by a half-volume and the corner node
3 has a quarter volume attached to it. Exact form of the difference equation will depend upon the boundary
conditions i.e. prescribed temperature, prescribed heat flux, insulated, convection or radiation boundary condi-
tions.
Fig. 8.7 shows some common boundary conditions encountered in practice.

Dx
Tm, n + 1
Tm, n Tm, n

Dy
Tm + 1, n Tm 1, n h, Ta
Tm 1, n
h, Ta
Tm, n Tm, n 1
1

(a) node at internal (b) node at plane surface


corner with convection with convection

Tm, n Tm, n + 1 Tm, n


Tm 1, n h, Ta

Tm 1, n q
Tm, n 1

Tm, n 1

(c) node at external (d) node at plane surface with


corner with convection uniform heat flux

FIGURE 8.7 Finite difference representation for two-dimensional conduction-different boundary conditions

Example 8.6. Develop finite difference equations for an interior corner node with convection conditions, using the en-
ergy balance method. See Fig. 8.7 (a).
Solution. As shown in the Fig. 8.7, elemental volume around the node is ¾ of full volume. Writing an energy balance for
this volume, we apply Eq. 8.40:
Q left + Q right + Q up + Q down + DV×qg = 0 ...(8.40)
i.e.
Tm - 1, n - Tm, n D y Tm + 1, n - Tm, n Tm , n + 1 - Tm, n
k×Dy× + k× × + k×Dx×
Dx 2 Dx Dy

D x Tm , n - 1 - Tm, n Dx Dy FG IJ 3 FG IJ
+ k×
2
×
Dy
+ h×
2
+
2 H K
× (Ta - Tm , n ) + qg ×
4
×D x×D y
H K = 0

Remembering that Dx = Dy = (Dx, say), we get on simplification,


FG 2× h× D x IJ 3 qg 2× h × D x
Tm, n – 1 + 2×Tm – 1, n + 2 ×Tm, n + 1 + Tm + 1, n – 6 +
H k K
×Tm, n + ×(Dx)2 ×
2 k
+
k
×Ta = 0

And, if there is no internal heat generation,


FG 2× h× D x IJ 2× h × D x
Tm, n – 1 + 2 ×Tm – 1, n + 2 ×Tm, n + 1 + Tm + 1, n – 6 +
H k K
×Tm, n +
k
×Ta = 0 ...(8.44)

Finite difference equations for the boundary conditions shown in Fig. 8.7 are summarized in Table 8.1.
Note: In Eqs. 8.42 and 8.44, put h = 0 or q = 0, to get difference equations for an insulated surface or a surface with
thermal symmetry.

358 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 8.1 Summary of steady state, finite difference equations for different boundary conditions
(q = heat flux, h = convection heat transfer coefficient k = thermal conductivity, no internal heat generation, and Dx = Dy)

Situation Finite difference equation (with D x = D y, no heat generation)


(1) Node at an internal corner with convection, Fig. 8.7,a:

FG IJ
2 ×h × D x 2× h × D x
Tm, n – 1 + 2 × Tm – 1, n + 2 × Tm, n + 1 + Tm + 1, n – 6 +
H k K× Tm, n +
k
× Ta = 0 ...(8.44)

(2) Node at a plane surface with convection, Fig. 8.7,b:

2× h × D x FG
h×D x IJ
2 × Tm – 1, n + Tm, n + 1 + Tm, n – 1 +
k
× Ta – 2 ×
Hk K
+ 2 ×Tm, n = 0 ...(8.45)

(3) Node at an external corner with convection, Fig. 8.7,c:

2× h × D x FG h×D x IJ
(Tm, n – 1 + Tm – 1, n) +
k H
× Ta – 2×
k K
+ 1 × Tm, n = 0 ...(8.46)

(4) Node at a plane surface with uniform heat flux, Fig. 8.7,d:
2×q × D x
(2 × Tm – 1, n + Tm, n + 1 + Tm, n – 1) + – 4× Tm, n = 0 ...(8.47)
k

Example 8.7. Develop difference equation for a tool tip, shown in Fig. Ex- q
ample 8.7. There is uniform heat flux, q (W/m2) on the upper surface. As-
sume a constant thickness, t for the tool tip.
Solution. Note that the node (m, n) is enclosed by surfaces AB, BC and CA.
A Tm + 1, n
Finite difference equation is developed by writing the energy balance B
Tm, n
on the elemental volume surrounding node (m, n) i.e. on the volume en-
closed by AB, BC and CA, remembering that we assume all heat flow lines
into the volume element.. We get: C Dy
Dx Dy Tm + 1, n - Tm, n ( D x)2 + ( D y)2 h, Ta
×t × q + k × ×t× + h× t × × (Ta - Tm, n ) = 0
2 2 Dx 2
In the above equation, first term is the heat flux through surface AB, Tm + 1, n 1
second term is the conduction across surface BC, and third term is convec-
tion at surface CA.
Dx
Simplifying the above equation, with Dx = Dy, we get:
F k I ×T Fk + I FIGURE Example 8.7 Finite difference
GH D x JK m + 1, n - GH D x JK
2 × h × Tm , n + (q + 2 × h × Ta ) = 0 representation for a tool tip

Above equation gives the desired difference equation for the tool tip.
Example 8.8. For the two-dimensional region shown in Fig. Example 8.8, with constant k (= 20 W/(mC)) and no internal
heat generation, and with the indicated boundary conditions, formulate the finite difference equations and solve for
unknown temperatures. Use Dx = Dy = 1 cm.
Solution.
Data:
Dx :=0.01 m Dy := 0.01 m Ta := 20°C h := 50 W/(m 2 C) k := 20 W/(mC)
Nodes are represented by numbers 1, 2, ...., 7. Elemental volume pertinent to each node is also marked around it
and numbered a, b, ..., r.
We shall develop finite difference equation for each node by writing the energy balance for the corresponding
elemental volume around that node.
For node 1:
Elemental volume to be considered is 1/4 volume, 1-a-b-c-1.
For this elemental volume, considering unit depth, heat transfers into the volume are:
From left surface, there is no heat transfer, since it is insulated.

NUMERICAL METHODS IN HEAT CONDUCTION 359


Convection Dx = Dy = 1 cm

y
2
h = 50 W/(m C)
Insulated a Ta = 20°C
1 3

c b d
2 4 g 5 h 6 k 7 Insulated

Dy r j p
e f i
x
T = 150°C
Dx

FIGURE Example 8.8 Finite difference representation for two-dimensional conduction-nodal network

i.e. Q left = 0
Form top, i.e surface 1-a: there is convection:
FG D x ×1IJ ×(T
i.e. Q top = h×
H2 K a – T1)

From right, there is conduction from node 3 through surface a-b:


FG D y ×1IJ × T - T
Q right = k×
H 2 K Dx
3 1
i.e.

From down, there is conduction from node 2 through surface b-c:


FG D x ×1IJ × T - T
Q down = k×
H 2 K Dy
2 1
i.e.

There is no heat generation term in this problem.


So, heat balance on the elemental volume for node 1 gives:
FG D x ×1IJ ×(T FG D y ×1IJ × T - T FG D x ×1IJ × T - T

H2 K – T1) + k×
H 2 K Dx + k×
H 2 K Dy
3 1 2 1
a = 0

i.e. 0.25×(Ta – T1) + 10×(T3 – T1) + 10×(T2 – T1) = 0 ...(a)


Eq. a is the difference equation for node 1.
For node 2:
Here, elemental volume to be considered is 1/2 volume, c-b-e-r. and, energy balance can be written as we did for
node 1.
However, since the surface is insulated, it is easier to use the mirror image concept and consider the node 2 as an
internal node. So, to the left of node 2, we have T4, mirror image of temperature of node 4. Then, considering 2 as
internal node, we get difference equation for node 2:
T4 + T1 + T4 + 150 – 4 ×T2 = 0 ...(b)
For node 3:
This is a corner node with convection. Elemental volume to be considered is 1/4 volume, a-3-d-b.
We can directly apply Eq. 8.46, viz.
2× h × D x h×D x FG IJ
(Tm, n – 1 + Tm – 1, n) +
k
×Ta – 2 ×
k H
+ 1 ×Tm, n = 0
K ...(8.46)

i.e. T4 + T1 + 0.05 ×Ta – 2.05×T3 = 0 ...(c)


For node 4:
This is an internal corner node with convection. Elemental volume to be considered is 3/4 volume, g-f-e-b-d-4.
Again, we can directly apply Eq. 8.44, viz.
FG 2× h× D x IJ 2× h × D x
Tm, n – 1 + 2×Tm – 1, n + 2× Tm, n + 1 + Tm + 1, n – 6 +
H k K
×Tm, n +
k
×Ta = 0 ...(8.44)

360 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. 150 + 2 ×T2 + 2 ×T3 + T5 – 6.05 ×T4 + 0.05×Ta = 0 ...(d)
Eq. d is the difference equation for node 4.
For node 5:
This is a surface node with convection. Elemental volume to be considered is 1/2 volume, g-f-i-h-g.
Again, we can directly apply Eq. 8.45, viz.
2 ×h × D x FGh× D x IJ
2 ×Tm – 1, n + Tm, n + 1 + Tm, n – 1 +
k
×Ta – 2 ×
H k K
+ 2 ×Tm, n = 0 ...(8.45)

Remembering that Eq. 8.45 was developed for a vertical surface, and in the present case, we are dealing with a
horizontal surface, we can write:
2×150 + T4 + T6 + 0.05×Ta – 4.05 ×T5 = 0 ...(e)
For node 6:
This is identical to node 5. So, we get:
2×150 + T5 + T7 + 0.05×Ta – 4.05 ×T6 = 0 ...(f)
For node 7:
This is a corner node with conduction from left, convection on the top, insulated on the right, and conduction from
down. Elemental volume to be considered is 1/4 volume, k-7-p-j.
Writing the energy balance:
FG D y ×1IJ × T - T FG D x ×1IJ ×(T FG D x ×1IJ × 150 - T

H 2 K Dx + h×
H2 K – T7) + 0 + k ×
H 2 K Dy
6 7 7
a = 0

i.e. 10×(T6 – T7) + 0.25× (Ta – T7) +10×(150 – T7) = 0 ...(g)


Temperatures at nodes 1 to 7 are obtained by simultaneously solving 7 Eqs. a to g.
We use ‘solve block’ of Mathcad to solve this set of equations. Start with guess values for all unknown temperatures
and immediately below ‘Given’, type the constraint equations. Then, the command ‘Find (T1, ..., T7)’ gives the tempera-
tures immediately:
T 1 := 50 T2 := 50 T3 := 50 T4 := 50 T5 := 50 T6 := 50 T7 := 50 (guess values of temperatures)
Given
0.25 ×(Ta – T1) + 10×(T3 – T 1) + 10×(T2 – T 1) = 0 ...(a)
T4 + T 1 + T4 + 150 – 4×T2 = 0 ...(b)
T4 + T1 + 0.05×Ta – 2.05×T3 = 0 ...(c)
150 + 2×T2 + 2 ×T3 + T5 – 6.05×T4 + 0.05 ×Ta = 0 ...(d)
2.150 + T 4 + T6 + 0.05 ×Ta – 4.05×T5 = 0 ...(e)
2.150 + T 5 + T7 + 0.05 ×Ta – 4.05×T6 = 0 ...(f)
10×(T6 – T 7) + 0.25× (Ta – T7) + 10 ×(150 – T 7) = 0 ...(g)
Temp := Find(T 1, T 2, T3, T4, T5, T6, T7) (node temperatures are stored in vector ‘Temp’.)

LM138.552 OP
MM142.929 PP
i.e.
M137 .139 P
Temp = M 141. 582 P
MM145.437 PP
MM146.438 PP
N146.636 Q
i.e. The node temperatures are:
T 1 = 138.552°C T2 = 142.929°C T3 = 137.139°C T4 = 141.582°C
T 5 = 145.437°C T6 = 146.438°C T7 = 146.636°C
Note: In Mathcad, while using the solve block, a great advantage is that the equations can be written in any order; also,
there is no need to collect the coefficient of each variable separately. Equations can be entered without simplification, as
in the form we get after making the energy balance.
Example 8.9. A very long bar of square cross-section has its four sides held at constant temperatures as shown in
Fig. Example 8.9. Determine the temperatures at the internal nodes. Compare the results with analytical solution.
Solution. There are 9 internal nodes. Difference equations for these nodes are obtained by applying Eq. 8.42, viz.

NUMERICAL METHODS IN HEAT CONDUCTION 361


y qg ×( D x)2
Tm – 1, n + Tm + 1, n + Tm, n + 1 + Tm, n – 1 – 4×Tm, n + = 0
k
T = 200 C ...(8.42)
2 In the present case, there is no internal heat generation. So, the
last term of the above equation will be zero. Therefore, we get:
1 2 3
Node 1: 150 + 200 + T2 + T4 – 4×T1 = 0 ...(a)
Node 2: T1 + 200 + T3 + T5 – 4×T2 = 0 ...(b)
4 5 6 T = 150 C
T = 150 C Node 3: T2 + 200 + 150 + T6 – 4×T3 = 0 ...(c)
Node 4: 150 + T1 + T5 + T7 – 4×T4 = 0 ...(d)
7 8 9
Node 5: T4 + T2 + T6 + T8 – 4×T5 = 0 ...(e)
Node 6: T5 + T3 + 150 + T9 – 4×T6 = 0 ...(f)
x Node 7: 150 + T4 + T8 + 150 – 4×T7 = 0 ...(g)
0 T = 150 C 2 Node 8: T7 + T5 + T9 + 150 – 4×T8 = 0 ...(h)
Node 9: T8 + T6 + 150 + 150 – 4×T9 = 0 ...(i)
FIGURE Example 8.9 Finite difference
Eqs. a to i give 9 difference equations for 9 internal nodes. By
representation for two-dimensional conduc-
solving these equations simultaneously, we get the temperatures at
tion-nodal network
nodes 1 to 9.
We use ‘solve block’ of Mathcad to solve this set of equations. Start with guess values for all unknown temperatures
and immediately below ‘Given’, type the constraint equations. Then, the command ‘Find (T1, ..., T9)’ gives the tempera-
tures immediately:
T 1 := 50 T2 := 50 T3 := 50 T4 := 50 T5 := 50
T 6 := 50 T7 := 50 T8 := 50 T9 := 50 (guess values of temperatures)
Given
150 + 200 + T2 + T4 – 4×T1 = 0 ...(a)
T1 + 200 + T3 + T 5 – 4×T2 = 0 ...(b)
T2 + 200 + 150 + T6 – 4×T3 = 0 ...(c)
150 + T1 + T5 + T 7 – 4×T4 = 0 ...(d)
T4 + T2 + T6 + T 8 – 4×T5 = 0 ...(e)
T5 + T3 + 150 + T 9 – 4×T6 = 0 ...(f)
150 + T4 + T8 + 150 – 4×T7 = 0 ...(g)
T7 + T5 + T9 + 150 – 4×T8 = 0 ...(h)
T8 + T6 + 150 + 150 – 4×T9 = 0 ...(i)
Temp := Find(T 1, T 2, T 3, T 4, T 5, T 6, T7, T8, T 9) (node temperatures are stored in vector ‘Temp’.)

LM171.429 OP
MM176.339PP
MM171 . 429
P
159 . 375P
Temp = M162 . 5 P
i.e.
MM159.375PP
MM153.571PP
MM154.911PP
MN153.571PQ
i.e. The node temperatures are:
T1 = 171.429°C T2 = 176.339°C T3 = 171.429°C T4 = 159.375°C T5 = 162.5°C T6 = 159.375°C
T7 = 153.571°C T8 = 154.911°C T9 = 153.571°C
Comparison with analytical solution:
Analytical solution for this problem is a little complicated and is given in terms of an infinite series, as follows:

FG y IJ
¥
( - 1)n + 1 H L K × sin FG n ×p × x IJ
sin h n × p ×
2
q = qc × ×
p å n
×
F WI
sin h G n × p × J
H LK
n=1
H LK
362 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Nomenclature for the above equation for the present problem is as follows:
q = T – 150 (T = temperature at the desired point; 150°C is the constant temperature on three sides)
qc := 200 – 150 (temperature difference between the temperature of fourth side and the consant temperature)
n (number of terms considered in the infinite series)
x, y (coordinates of the point where temperature is desired)
L := 2 m (length along x-axis)
W := 2 m (length along y-axis)
Above equation is solved very easily in Mathcad:
Let us re-define q as a function of (x, y), and consider only 6 terms of the infinite series (n = 6) as shown below, for
convenience:

FG y IJ
6
( - 1)n + 1 + 1 H L K × sin FG n ×p × x IJ
sin h n × p ×
2
q (x, y) := qc × ×
p å n
×
F WI
sin h G n × p × J
H LK ...(A) (define q as a function of x and y)
n=1
H LK
Now, substitute (x, y) corresponding to different nodes and get the analytical temperature at those nodes immedi-
ately:

Node q (x, y) Temperature of node (deg. C) Temperature by


numerical method
1 q (0.5, 1.5) = 21.623 T1 = q (0.5, 1.5) + 150 T1 = 171.623 171.429
2 q (1.0, 1.5) = 27.059 T2 = q (1.0, 1.5) + 150 T2 = 177.059 176.339
3 q (1.5, 1.5) = 21.623 T3 = q (1.5, 1.5) + 150 T3 = 171.623 171.429
4 q (0.5, 1.0) = 9.102 T4 = q (0.5, 1.0) + 150 T4 = 159.102 159.375
5 q (1.0, 1.0) = 12.5 T5 = q (1.0, 1.0) + 150 T5 = 162.5 162.5
6 q (1.5, 1.0) = 9.102 T6 = q (1.5, 1.0) + 150 T6 = 159.102 159.375
7 q (0.5, 0.5) = 3.399 T7 = q (0.5, 0.5) + 150 T7 = 153.399 153.571
8 q (1.0, 0.5) = 4.771 T8 = q (1.0, 0.5) + 150 T8 = 154.771 154.991
9 q (1.5, 0.5) = 3.399 T9 = q (1.5, 0.5) + 150 T9 = 153.399 153.571

We make following important observations:


(i) Even with a crude mesh of 4 ´ 4, we get values of temperatures at the nodes very close to the analytical
results.
(ii) Note that the analytical relation to find the temperature at any point is very complicated, and to solve it
without a computer is rather laborious and time consuming. But, with Mathcad, even this analytical
solution is easy to perform.
(iii) Numerical method of formulating difference equations by energy balance method is easy and straight
forward, only labour being in solving the set of simultaneous equations. But, with Mathcad, this is also
very easy.
Irregular boundaries:
Very often, in practice, we have to analyse bodies with irregular boundaries, e.g. engine blocks, turbine blades,
aerofoil sections etc. In such cases, the whole volume cannot be entirely filled up by the square mesh used for
numerical analysis. Still, easiest and simple approach to deal with such geometries is to fill the geometry by
approximating the irregular boundary with simple square mesh elements, as shown in Fig. 8.8. This method,
generally gives results of acceptable accuracy, particularly when the mesh size is small and the nodes are quite
close to each other. Many commercially available specialized software for numerical analysis employ more so-
phisticated methods.

8.8 Numerical Methods for Transient Heat Conduction


In transient conduction, temperature varies with both position and time. So, to obtain finite difference equations
for transient conduction, we have to discretize both space and time domains. This scheme is illustrated in
Fig. 8.9.

NUMERICAL METHODS IN HEAT CONDUCTION 363


Irregular Here, the space increment is Dx and the time increment is Dt.
boundary At a given node ‘m’, x – coordinate is (m.Dx) and at a given time
Approximation step ‘i’, time from start up is (i.Dt), as is clear from the Fig. 8.9.
Starting from initial temperature at t = 0, at each node we calculate
the temperature at a successive time interval of Dt till we reach the
desired time at which temperature has to be calculated. Therefore,
obviously, number of calculations required in case of transient con-
duction is much more. Time step is shown in superscript, i.e. Tmi is
the temperature of node ‘m’ at time step ‘i’ (at time = i.Dt from start
up) and the notation Tmi + 1 means the temperature of node ‘m’ at the
time step (i + 1) ( at time = (i + 1)Dt from start up).
Formulation of finite difference equations in transient conduc-
tion is done by an energy balance on the elemental volumes con-
taining the nodes, just as was done in the case of steady state
FIGURE 8.8 Approximating an irregular conduction; however, now, on the RHS, there appears a term repre-
boundary senting the change in energy content of the elemental volume, with
time. Also, as in the earlier case, while writing the energy balance,
it is assumed that all heat lines flow into the elemental volume.
t We write, for a given volume element:
(Heat transferred into the volume element from all sides, per unit
Tm
i+1 time) + (Heat generated within the volume element per unit time) =
i+1 (Change in energy content of the volume element per unit time).

i i i i (Qleft + Qup + Qright + Qdown + Qg = r×V element ×C p ×


LM T
( i + 1)
m - Tmi OP
Tm – 1 Tm Tm + 1
MN Dt PQ
...(8.48)
Dt In the above equation, as already mentioned, Tmi is the tem-
perature of node ‘m’ at time step ‘i’ (i.e. at time = i.Dt from start up)
1
and Tmi +1 is the temperature of node ‘m’ at the time step (i + 1) (i.e.
x
0 1 m–1 m m+1 M at time = (i + 1)Dt from start up).Cp is the specific heat and r is the
Dx density of the medium. (Tmi +1 – Tmi )/Dt is the finite difference ap-
proximation of the term dT/dt.
Figure 8.9 Finite difference repre- Now, regarding the terms on the LHS of Eq. 8.48, the question
sentation for one-dimensional, transient arises as to whether we should consider the temperatures of the
conduction-nodal network nodes at step ‘i’ or step ‘(i + 1)’. In fact, both the methods are
adopted in practice. While applying Eqs. 8.48 to write the finite dif-
ference equation for a node, if the terms on the LHS of the equation are considered at time step ‘i’, then, the
method is known as explicit method of approach; if the terms on the LHS of the equation are considered at time
step ‘(i + 1)’, then, the method is known as implicit method of approach. To summarize:
Explicit method:

(Qlefti + Qupi + Qrighti + Qdowni + Qgi) = r×V element ×C p ×


LM T
( i + 1)
m - Tmi OP ...(8.49)
MN Dt PQ
Implicit method:

(Qlefti +1 + Qupi+1 + Qrighti+1 + Qdowni +1) + Qgi +1 = r×V element ×C p ×


LM T
( i + 1)
m - Tmi OP ...(8.50)
MN Dt PQ
In the explicit method, time derivative is calculated in ‘forward difference’ form, and in implicit method, the
time derivative is in the ‘backward difference’ form.
Eqs. 8.49 and 8.50 are applicable in any coordinate system and for multidimensional systems too; however,
when more than one-dimension is involved, number of surfaces through which heat flows into the elemental
volume are more and there will be correspondingly more terms on the LHS of Eqs. 8.49 and 8.50.

364 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Explicit method is called so, because temperature of the node ‘m’ at time step (i + 1) is calculated explicitly in
terms of the temperatures calculated at the previous time step ‘i’; therefore, the calculations are quite straight
forward; however it suffers from a serious limitation that the time increment cannot be independently fixed, but
has an upper limit because of stability considerations. But in case of implicit method, this limitation on time
duration is not there and we can choose any time step; but the implicit method requires that at each time step,
nodal temperatures have to be solved simultaneously.
8.8.1 One-dimensional Transient Heat Conduction in a Plane Wall
Consider one-dimensional, transient heat conduction in a plane wall of thickness L, with heat generation rate
qg (x, t) and constant thermal conductivity k. Now, let us divide the region 0 < x < L into M sub-regions. Then,
thickness of each sub-region is:
Dx = L/M. So, there are totally (M + 1) nodes, starting from m = 0 to m = M, as shown in Fig. 8.10. Coordi-
nate of node ‘m’ is x = m.Dx. and let temperature of node ‘m’ be Tm. Remembering that each node represents the
sub-volume around it (of thickness Dx), it is clear that interior nodes 1, 2…M – 1 represent full sub-volumes
whereas boundary nodes 0 and M represent half volumes (of thickness Dx/2). Volume of element surrounding
node ‘m’ is A.Dx.

Volume element of node ‘m’


qm
Plane wall
Dx/2
Qleft Qright

Dx
Ta
ha T0 T1 T2 Tm – 1 Tm Tm + 1 TM – 1 TM
x

0 1 2 m–1 m m+1 M–1 M


Dx
m – 1/2 m + 1/2
x=0 x=L

FIGURE 8.10 Finite difference formulation in a plane wall by energy balance for transient heat conduction

To get the finite difference formulation, we apply the general energy balance, i.e. Eq. 8.48:
Tm - 1 - Tm Tm + 1 - Tm T i + 1 - Tmi
k×A× + k×A× + qm ×(A×Dx) = rA×Dx×Cp × m ...(8.51)
Dx Dx Dt
Simplifying,

qm ×( D x)2 ( D x )2
Tm – 1 – 2×Tm + Tm + 1 + = ×(Tmi + 1 – Tmi) ...(8.52)
k a × Dt
k
where, a= = thermal diffisivity of the material.
r × Cp

a ×Dt
Now, the term is the finite difference form of the Fourier number, Fo
( D x )2
So, Eq. 8.52 reduces to:

qm ×( D x)2 (T )i + 1 - (Tm )i
Tm – 1 – 2×Tm + Tm + 1 + = m ...(8.53)
k Fo
Now, as we mentioned earlier, in the LHS of Eq. 8.53, we can use the temperatures of the nodes at the
‘previous time step, i’, or temperatures at the ‘next time step, i + 1’. If we use temperatures at time step ‘i’, it is the
‘explicit method’ and if the temperatures at time step ‘i + 1’ are used, then, it is the ‘implicit method’.

NUMERICAL METHODS IN HEAT CONDUCTION 365


Explicit method:
i
qm ×( D x)2 (T )i + 1 - (Tm )i
(Tm – 1) i – 2×(Tm) i + (Tm + 1) i + = m ...(8.54)
k Fo
Now, the new temperature Tmi +1 can be explicitly solved since the other terms involved at the previous time
step ‘i’, are already known. So, we write for Tmi+ 1:

( qm )i ×( D x)2
(Tm)i +1 = Fo×[(Tm – 1)i + (Tm + 1)i] + (1 – 2×Fo)×(Tm)i + Fo× ...(8.55)
k
Eq. 8.55 is the explicit difference equation valid for all interior nodes, 1, 2, …, (M – 1), when there is internal
heat generation.
When there is no heat generation, Eq. 8.55 reduces to:
(Tm)i +1 = Fo×[(Tm – 1)i + (Tm + 1) i] + (1 – 2×Fo)×(Tm)i ...(8.56)
Implicit method:
If in the LHS of Eq. 8.53, we use the values at time step (i + 1), we get the implicit relation for the node tempera-
tures:

( qm )i + 1 × ( D x)2 (T )i + 1 - (Tm )i
i.e. (Tm – 1) i +1 – 2×(Tm)i + 1 + (Tm + 1)i + 1 + = m ...(8.57)
k Fo
Eq. 8.57 is simplified to:
LM
(1+ 2×Fo)×(Tm) i + 1 – Fo (Tm - 1 )i + 1 + (Tm + 1 )i + 1 +
(qm )i + 1 ×( D x )2 OP = + (T ) i
...(8.58)
MN k PQ m

Eq. 8.58 is the implicit difference equation valid for all interior nodes, 1,2,…,(M – 1), when there is internal
heat generation.
When there is no heat generation, Eq. 8.58 reduces to:
(1 + 2×Fo)×(Tm)i + 1 – Fo×[(Tm – 1)i + 1 + (Tm + 1) i + 1] – (Tm )i = 0 ...(8.59)
With the use of either the explicit or the implicit equations given above, we get M – 1 nodal equations.
Unless the temperatures at the boundaries are specified in the problem, we need two more equations for the
boundary nodes ‘0’ and ‘M’. These are obtained by applying the energy balance for the half-volumes around
these nodes. See Fig. 8.10. Exact nature of the difference equations depends on the specific boundary condition.
For example:
For node ‘0’ with convection boundary condition:
Explicit formulation:

(T1 )i - (T0 )i Dx Dx (T )i + 1 - (T0 )i


h×A×[Ta – (T0)i] + k×A× + (q 0)i ×A× = r×A× ×Cp × 0 ...(8.60)
Dx 2 2 Dt
Simplifying:
LM
(T0) i + 1 = (1 – 2×Fo – 2×Fo×Bi)×(T0) i + Fo× 2 ×(T1 )i + 2 × Bi ×Ta +
(q0 )i ×( D x )2 OP ...(8.61)
MN k PQ
h ×D x
where Bi = = Biot number
k
When there is no heat generation, Eq. 8.61 for explicit formulation becomes:
(T0)i + 1 = (1 – 2×Fo – 2×Fo×Bi)×(T0) i + Fo×[2×(T1) i + 2×Bi×Ta] ...(8.62)
For other types of boundary conditions, difference equations are developed in a similar manner, by applying
the energy balance on the elemental volume containing the node and considering all the heat flows to be into the
volume.
Once the difference equations are developed for all nodes with suitable Dx, next step is to choose a suitable
time increment Dt. Then, starting with the initial conditions at t = 0, solve the difference equations for the tem-
peratures Tmi +1 at all the nodes at the next time step t = Dt. Now, using these values of temperatures as ‘previous

366 FUNDAMENTALS OF HEAT AND MASS TRANSFER


values’, again get the nodal temperatures at the next time step t = 2. Dt, using the same difference equations.
Thus, continue to march in time till the solution is obtained for the desired time interval.
Stability criterion:
We said that once the explicit difference equations are developed, suitable time interval Dt has to be chosen. This
is done keeping the stability criterion in mind, since the explicit method is not unconditionally stable. That
means, above a certain value of Dt, the solution will not converge. This limit on Dt is determined from math-
ematical and thermodynamic considerations (see good text books on numerical methods) as follows:
“Coefficients of all Tmi in the Tmi + 1 expressions (called ‘primary coefficients’) must be greater than or equal to
zero for all nodes ‘m’”.
Considering Eq. 8.55 for interior nodes, we see that coefficient of Tmi is (1 – 2.Fo) and applying the above
mentioned criterion for stability, we get:
1 – 2×Fo ³ 0
a ×D t 1
i.e. Fo = 2
£ ( for interior nodes, one-dimensional conduction...(8.63))
( D x) 2
Now, Dt must be fixed from Eq. 8.63. qg
Plane wall
However, generally, boundary nodes with convec-
Dx/2
tion conditions are more restrictive and in such cases,
coefficient of Tmi from the most restrictive eqn. must be
considered for the stability criterion and the time step
Dt must be determined with respect to that coefficient. Ta Ta
Example 8.10. A large uranium plate of thickness L = 10 h h
cm, (k = 28 W/(mC), a = 12.5 ´ 10 –6 m2/s) is initially at an T0 T1 T2 T3 T4 T5
uniform temperature of 100°C. Heat generation rate in the x
plate is 5 ´ 10 6 W/m3. At time t = 0, both the left and right 0 1 2 3 4 5
sides of the plate are subjected to convection with a fluid at
at temperature of 0°C and a heat transfer coefficient of 1500
W/(m2C). Using a uniform nodal spacing of 2 cm, develop x=0
the explicit finite difference formulations for all nodes, and
FIGURE Example 8.10 Finite difference formula-
determine the temperature distribution in the plate after 5
min. Also, find out how long it will take for steady condi-
tion in a plate by energy balance for transient heat
tions to be reached in the plate. conduction
(b) Also, solve this problem by implicit finite difference formulation.
Solution.
Data:
L := 0.1 m k := 28 W/(mC) a := 12.5 ´ 10 –6 m 2/s qg := 5 ´ 10 6 W/m 3 T := 100°C Ta := 0°C
2
Dx : = 0.02 m h := 1500 W/(m C) (convective heat transfer coeff.) M := 5 t := 300 s
Difference equations for interior nodes:
Nodes 1, 2, 3 and 4 are interior nodes. Finite difference equations for these nodes by explicit method are obtained from
Eq. 8.55, by setting m = 1, 2, 3, 4. i.e.
i
( q m ) × ( D x) 2
(Tm)i +1 = Fo×[(Tm – 1) i + (Tm + 1) i] + (1 – 2×Fo)×(Tm)i + Fo× ...(8.55)
k
We get:
i
(q g ) × (D x )2
Node 1: T1i +1 = Fo[(T 0i + T2i ) + (1 – 2 ×Fo)×T1i + Fo× ...(b)
k
i
(q g ) × (D x )2
Node 2: T2i +1 = Fo[(T 1i + T 3i ) + (1 – 2×Fo)×T 2i + Fo× ...(c)
k
i
(q g ) × (D x )2
Node 3: T3i +1 = Fo[(T 2i + T 4i ) + (1 – 2×Fo)×T 3i + Fo× ...(d)
k
i
(q g ) × (D x )2
Node 4: T4i +1 = Fo[(T 3i + T 5i ) + (1 – 2×Fo)×T 4i + Fo× ...(e)
k

NUMERICAL METHODS IN HEAT CONDUCTION 367


Difference equation for boundary nodes:
For node ‘0’:
Node ‘0’ is on the left surface, subjected to convection. Applying the Eq. 8.61 directly:
LM
(T0) i + 1 = (1 – 2×Fo – 2×Fo×Bi)×(T0)i + Fo× 2 × (T1 ) + 2 × Bi × Ta +
i (q0 )i × (D x)2 OP
MN k PQ ...(8.61)

h× D x
where Bi = = Biot number
k
LM
(T0) i + 1 = (1 – 2×Fo – 2×Fo×Bi)×(T0)i + Fo× 2 × (T1 ) + 2 × Bi × Ta +
i ( qg ) × ( D x ) 2 OP
i.e.
MN k PQ ...(a)

For node 5:
This is a node with convection boundary condition. So, applying the energy balance to the half-volume around node 5,
with all the heat lines flowing into the element, we get:
F T - T I + q ×A× D x = r×A× D x ×C × T
i i i+1 i
- T5
h ×A×(Ta – T5i ) + k×A× GH D x JK
4 5
2
g
2
p
5
Dt

L ( q ) × ( Dx) O
– 2× F × B )× (T ) + F × M 2 × (T ) + 2 × Bi × T + PPQ
2

T5i
+1 i i a
= (1 – 2× Fo
i.e. o i 5
MN o 4 a
k
...(f)

Now, we have to fix the upper limit of Dt from stability criterion. To do that, we observe that in Eqs. a to f, the
smaller coefficient of Tmi is in Eq. f, i.e. (1 – 2. Fo – 2.Fo.Bi) must be greater than or equal to zero. Putting this condition,
we get:
h ×D x
1 – 2×Fo – 2×Fo× ³0
k
1
i.e. Fo £
F h × D x IJ
2 ×G 1 +
H k K
( D x)2
i.e. Dt £
F h × D x IJ
2 ×a ×G 1 +
H k K
i.e. Dt £ 7.724 s
This means that a time step less than 7.724 s has to be employed from stability criterion.
Let us choose:
Dt : = 5 s
a ×D t
Then, Fo: = i.e. Fo = 0.1563
(D x )2
Substituting all relevant numerical values in Eqs. a to f, we get the explicit difference equations as:
T0i + 1 = 0.353×T0i + 0.1563×(2 ×T1i + 71.429) ...(a)
T1i + 1 = 0.1563× (T0i + T2i) + 0.688×T1i + 11.161 ...(b)
T2i + 1 = 0.1563× (T1i + T3i) + 0.688×T2i + 11.161 ...(c)
T3i + 1 = 0.1563× (T2i + T4i) + 0.688×T3i + 11.161 ...(d)
T4i + 1 = 0.1563× (T3i + T5i ) + 0.688×T4i + 11.161 ...(e)
T5i + 1 = 0.363×(T5)i + 0.1563×[2×(T4)i + 71.429] ...(f)
Initial temperature of the plate at t = 0 and i = 0, is given as 100°C.
i.e. T00= T10 = T30 = T40 = T50 = 100°C
Therefore, at the next time step i = 1, i.e. at Dt = 5 s, temperatures at nodes 0 to 5 can be explicitly calculated from
Eqs. a to f. Then, calculate temperatures at the nodes for next time step of Dt = 10 s, using the same Eqs. a to f, since the
temperatures at the previous time step are already calculated. Thus, march in time till we reach the time limit specified
in the problem, 5 min, i.e. there are 60 time steps of 5 s each.

368 FUNDAMENTALS OF HEAT AND MASS TRANSFER


This calculation is easily done in Mathcad. We slightly change the notation for convenience in calculation: we write
the superscripts as subscripts to work with matrix notation, as shown below.
In the small Mathcad program given below, LHS defines a function ‘Temp(n)’ where n is the no. of time steps,
which we can specify. Output is a vector containing step no., total time elapsed, and node temperatures T0, T1, ..., T5.
On the RHS, first 6 lines define the initial temperatures at the nodes, all equal to 100°C.
Then, a ‘for loop’ evaluates the finite difference Eqs. a to f, each ‘new’ node temperature being calculated in terms
of the temperatures calculated in the previous time step. Here, the number of time steps, ‘n’ can be changed since it is
included in function definition on the LHS.

Temp ( n) = T 0 0 ¬ 100
T 10 ¬ 100
T 2 0 ¬ 100
T 3 0 ¬ 100
T 4 0 ¬ 100
T 5 0 ¬ 100
for i Î 0.. n
T 0 i + 1 ¬ 0 .353 × T 0 i + 0.1563 × ( 2 × T 1i + 71. 429)
T 1i + 1 ¬ 0.1563 × (T 0 i + T 2 i ) + 0 .688 × T1i + 11.161
T 2 i + 1 ¬ 0 .1563 × (T 1i + T 3 i ) + 0.688 × T 2 i + 11.161
T 3 i + 1 ¬ 0 .1563 × (T 2i + T 4 i ) + 0.688 × T 3i + 11.161
T 4 i + 1 ¬ 0.1563 × (T 3 i + T 5i ) + 0.688 × T 4 i + 11.161
T 5 i + 1 ¬ 0 . 353 × T 5i + 0 .1563 × ( 2 × T 4 i + 71. 429)
[ i 5 × i T 0 i T 1i T 2 i T 3 i T 4 i T 5 i ]

Temp(0) = [0 0 100 100 100 100 100 100] (starting at time = 0)


i = step no.; Dt= one time step = 5 s; t = time duration from beginning = i. Dt , s
i t T0 T1 T2 T3 T4 T5
Temp(2) = [2 10 73.369 117.213 122.449 122.449 117.213 73.369]
Temp(4) = [4 20 75.447 127.666 142.472 142.472 127.666 75.447]
Temp(12) = [12 60 95.576 169.805 202.468 202.468 169.805 95.576]
Temp(18) = [18 90 108.991 196.256 236.479 236.479 196.256 108.991]
Temp(24) = [24 120 120.036 217.958 264.179 264.179 217.958 120.036]
Temp(36) = [36 180 136.45 250.194 305.285 305.285 250.194 136.45]
Temp(48) = [48 240 147.409 271.714 332.724 332.724 271.714 145.409]
Temp(60) = [60 300 154.724 286.079 351.041 351.041 286.079 154.724]
Temp(120) = [120 600 167.466 311.102 382.942 382.946 311.102 167.466]
Temp(180) = [180 900 169.155 314.419 387.176 387.176 314.419 169.155]
Temp(250) = [250 1.25 ´ 10 3 169.388 314.878 387.761 387.761 314.878 169.388]
Temp(260) = [260 1.3 ´ 10 3 169.395 314.891 387.778 387.778 314.891 169.395]
Temperature distribution after 5 min.:
Above Table of results gives node temperatures at different time steps.
Temp(60) corresponds to 60th time step, i.e. 300 s from beginning.
We note that after 5 min. the node temperatures are:
T 0 = T5 = 154.724°C; T1 = T 4 = 286.07°C; T2 = T 3 = 351.041°C;
Time to reach steady state:
It may be seen from the Table that from about 240th step, the temperatures at the nodes do not vary much as we advance
in time, i.e. steady state is reached at about 20 min. from start up.
To draw the temperatures at the nodes at different times:
It is instructive to graphically represent the manner in which the plate proceeds to attain steady state temperature. First
represent the node temperatures at different time steps as vectors:

NUMERICAL METHODS IN HEAT CONDUCTION 369


LM100OP
MM100PP
Step0 = M
MM100PPP
100
(intial temperature distribution in nodes 0, 1, ..., 5)

MM100 P
N100PQ
Similarly, temperature distributions after 1, 5, 10, 20 and 30 min. are given Step1, Step5,...,etc., below:

LM 95.58 OP LM154.72OP LM167.47 OP


MM 169.81PP MM286.08PP MM 311.1 PP
Step1 = M Step5 = M Step10 = M
MM202.47PPP MM351.04PPP MM382.95PPP
202. 47 351. 04 382. 95

MM 169 .81
P MM286 .08
P MM167
311.1
P
N 95.58 PQ N154.72PQ N .47 PQ
LM169.38OP LM169.41OP
MM314.86PP MM314.92PP
Step20 = M Step30 = M
MM387.74PPP MM387.82PPP
387 .74 387 .82

MM169
314.86
P MM314 . 92
P
N .38PQ N169.41PQ
To draw the graph, first define a range variable i = 0 to 5 with an increment of 1. This represents nodes 0, 1, …, 5.
Select the x–y graph from the graph palette and fill up ‘i’ in the place holder on the x-axis. In the place holder of y-axis,
fill up above shown temperature vectors, with a comma between each. Click anywhere outside the graph region and the
graphs appear:
i := 0, ..., 5 (define a range variable ‘i’ varying from 0 to 5, with an increment of 1)
It is seen from the graph that steady state is reached at about 20 min. from start up.

Transient temperature distribution in a plate


400

350
Temperature (deg. C)

300

250

200

150

100

50
0 1 2 3 4 5
Node number
Initial temperature distribution After 10 min.
After 1 min. After 20 min.
After 5 min. After 30 min.

370 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(b) Implicit method:
Difference equations for interior nodes:
Nodes 1, 2, 3 and 4 are interior nodes. Finite difference equations for these nodes by implicit method are obtained from
Eq. 8.58, by setting m = 1, 2, 3, 4. i.e.
LM
(1 + 2×Fo)×(Tm) i + 1 – Fo× (Tm - 1 )
i+1
+ (Tm + 1 )i + 1 +
(qm )i + 1 × (D x)2 OP – (T ) = 0 i
MN k PQ m ...(8.58)

LM
(1 + 2 ×Fo)×T 1i + 1 – Fo× (T0 )
i+1
+ (T2 )
i+1
+
OP – T
( q g ) × ( D x) 2 i
Node 1:
MN k PQ 1 = 0 ...(b)

(1 + 2×Fo)×T 2i + 1
L
– Fo× M(T )
i +1
+ (T3 )
i +1
+
( q ) × ( D x) O
g
PP – T
2
i
Node 2:
MN 1
k Q
2 = 0 ...(c)

(1 + 2 ×Fo)×T 3i + 1
L
– Fo× M(T )
i +1
+ (T4 )
i +1
+
(q ) × ( D x ) O
g
PP – T
2
i
Node 3:
MN 2
k Q
3 = 0 ...(d)

(1 + 2 ×Fo)×T 4i + 1
L
– Fo× M(T )
i +1
+ (T5 )
i +1
+
(q ) × (D x ) O
g
PP – T
2
i
Node 4:
MN 3
k Q
4 = 0 ...(e)

Difference equations for boundary nodes:


Nodes 0 and 5 are boundary nodes, with convection conditions.
For node ‘0’:
Writing the energy balance for the half-volume around node ‘0’, with all heat flow lines going into the volume element,
with the LHS of Eq. 8.60 expressed at time step (i + 1), we get:
i+1 i +1 i i
(T1 ) - (T0 ) Dx Dx (T ) + 1 - (T0 )
h×A×[Ta – (T0)i + 1] + k×A× + q g ×A× = r×A× ×Cp × 0
Dx 2 2 Dt

2× Fo × h × D x Fo × q g ×( D x)2
i.e. ×[Ta – (T0)i + 1] + 2×Fo×[(T 1)i + 1 – (T0)i + 1] + = (T 0)i + 1 – (T 0) i ...(a)
k k
Eq. a is the implicit finite difference formulation for node ‘0’, with convection conditions.
For node ‘5’:
Writing the energy balance for the half-volume around node ‘5’, with all heat flow lines going into the volume element,
with the LHS of energy balance equation expressed at time step (i + 1), we get:

FT i +1
- T5i + 1 I + q ×A× D x Dx i
T + 1 - T5
i
h ×A×(Ta – T5i + 1) + k×A× GH 4
Dx JK 2
g = r×A×
2
×Cp × 5
Dt

2× Fo × h × D x Fo × q g ×( D x)2
i.e. ×[Ta – (T5)i + 1] + 2×Fo×[(T 4)i + 1 – (T5)i + 1] + = (T 5)i + 1 – (T 5) i ...(f)
k k
Eq. f is the implicit finite difference formulation for node 5, with convection conditions.
Now, we can choose any Dt, since there is no problem of stability in implicit formulation.
Let us choose:
Dt := 10 s
a ×Dt
Therefore, Fo := i.e. Fo = 0.3125
( D x )2
Inserting numerical values, Eqs. a to f are written as:
0.67×[Ta – (T0)i + 1] + 0.625 ×[(T1)i + 1 – (T0)i + 1] + 22.321 = (T0)i + 1 – (T0)i ...(a)
1.625×T1i + 1 – 0.3125×[(T0)i + 1 – (T2)i + 1 + 71.429] – T1i = 0 ...(b)
1.625×T2i + 1 – 0.3125×[(T1)i + 1 + (T3)i + 1 + 71.429] – T2i = 0 ...(c)
1.625×T3i + 1 – 0.3125×[(T2)i + 1 + (T4)i + 1 + 71.429] – T3i = 0 ...(d)
1.625×T4i + 1 – 0.3125×[(T3)i + 1 + (T5)i + 1 + 71.429] – T4i = 0 ...(e)
0.67×[Ta – (T5)i + 1] + 0.625×[(T4)i + 1 + (T5)i + 1] + 22.321 = (T5)i + 1 – (T5)i ...(f)

NUMERICAL METHODS IN HEAT CONDUCTION 371


Now, to start with, i.e. at t = 0, all the node temperatures T0, T1, ..., T5 are known. Then, at the next time step, solve
Eqs. a to f simultaneously to get the node temperatures at that time step. Using these results, solve the Eqs. a to f at the
next time step, etc. till you reach the given time limit.
This part of the problem is left as an exercise for the student. Write a computer program to accomplish this task. Use the Gauss–
Siedel iteration technique for the solution of simultaneous equations.

8.8.2 Two-dimensional Transient Heat Conduction


Fig. 8.11 shows a rectangular region where the heat transfer in x and y directions are significant, and heat transfer
in the z direction is negligible. Divide the rectangular region into a nodal network of thicknesses Dx and Dy as
shown. Let the thickness in the z direction be unity.
Finite difference equations are developed by writing the energy balance for an elemental volume surround-
ing the node under consideration. All heat flows are considered to be flowing into the volume.
Difference equations for interior nodes:
A typical interior node, Tm,n and the elemental volume surrounding it, and immediate neighbours of this node
are shown in Fig. 8.11 (b). Node Tm, n is surrounded by 4 nodes: Tm – 1, n, Tm, n + 1, Tm + 1, n, and Tm, n – 1. Let us make
an energy balance on the elemental volume surrounding the node Tm, n. It is observed that heat flows into the
node from all the four directions, i.e. left, up, right and down. In addition, let there be heat generation in the
volume at a rate of (DV.qg), W, where qg, (W/m3), is the uniform heat generation rate in the system.

Tm, n Tm, n + 1 Tm, n


Dy
Tm – 1, n Tm + 1, n

2 Tm, n – 1
1
Typical internal node, Tm, n and
0 1 2 M the surrounding nodes
Dx
(a) (b)

FIGURE 8.11 Finite difference representation for two-dimensional conduction-nodal network

Writing the energy balance,


dT
Q left + Q up + Q right + Q down + DV×qg = m×Cp × ...(8.64)
dt
i.e.
Tm - 1, n - Tm , n Tm , n + 1 - Tm , n Tm + 1, n - Tm , n
k×Dy× + k×Dx× + k×Dy×
Dx Dy Dx

Tm , n - 1 - Tm , n T i + 1 - Tmi
+ k×D x × + q g ×Dx×Dy = r× Dx×Dy×Cp × m ...(8.65)
Dy Dt
For Dx = Dy (i.e. a square mesh), we get:
qg ×( D x )2 Tmi+ 1 - Tmi
Tm – 1, n + Tm + 1, n + Tm, n + 1 + Tm, n – 1 – 4×Tm, n + = ...(8.66)
k Fo
a ×D t
where Fo = = Fourier number, and a is thermal diffusivity.
( D x )2
Now, on the LHS of Eq. 8.66, if we use the ‘previous’ time step ‘i’, we get the explicit formulation of finite
difference equation for interior nodes:

372 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e.
qg ×( D x) 2 (Tm , n )i + 1 - (Tm , n )i
(Tm – 1, n)i + (Tm + 1, n)i + (Tm, n + 1)i + (Tm, n – 1)i – (4×Tm, n)i + = ...(8.67)
k Fo
i.e.
q g ×( D x)2
(Tm, n)i + 1 = Fo×[(Tm – 1, n)i + (Tm + 1, n)i + (Tm, n + 1)i + (Tm, n – 1)i] + (1 – 4×Fo)×(Tm, n)i + Fo × ...(8.68)
k
Eq. 8.68 is valid for all interior nodes, when there is heat generation.
If there is no heat generation, Eq. 8.68 simplifies to:
(Tm, n)i + 1 = Fo×[(Tm – 1, n)i + (Tm + 1, n)i + (Tm, n + 1)i + (Tm, n – 1)i] + (1 – 4×Fo)×(Tm, n)i ...(8.69, a)
i
As mentioned earlier, stability criterion in the explicit method requires the coefficient of (Tm, n) to be positive
and this condition gives the upper limit on the time increment Dt, as follows:
a ×D t 1
Fo = £ (stability criterion for interior nodes...(8.70))
( D x )2 4
Now, on the LHS of Eq. 8.66, if we use the ‘future’ time step ‘i + 1’, we get the implicit formulation of
finite difference equation for interior nodes. So, we get:
i+1 i i+1 i+1 i+1 i+1
(Tm, n ) = (1 + 4× Fo) ×Tm+, n1 - Fo (Tm + 1, n + Tm - 1, n + Tm, n + 1 + Tm, n - 1 ) ...(8.69, b)
Difference equations for boundary nodes:
Boundary nodes may be on the surface or on the corners. Difference equations are developed for boundary nodes
in a similar manner as for interior nodes, i.e. by making a heat balance on the elemental volume surrounding the
node. Exact form of the difference equation will depend upon the boundary conditions i.e. prescribed tempera-
ture, prescribed heat flux, insulated, convection or radiation boundary conditions.
Fig. 8.12 shows some common boundary conditions encountered in practice:
Finite difference equations for the boundary situations shown in Fig. 8.12 are given in Table 8.2.

Dx
Tm, n + 1
Tm, n Tm, n
Tm, n
Dy
Tm – 1, n h,Ta Tm – 1, n h, Ta
Tm + 1, n
Tm – 1, n
h,Ta
Tm, n – 1
Tm, n – 1 Tm, n – 1
(a) node at internal (b) node at plane surface (c) node at external
corner with convection with convection corner with convection

FIGURE 8.12 Finite difference representation for two-dimensional conduction-different boundary


conditions

Example 8.11. Consider the L-bar shown in Fig. Example 8.11, with constant k (= 20 W/(mC)) and no internal heat
generation. Its left and right sides are insulated and the bottom surface is maintained at 150°C at all times. If at time t =
0, the top surface is suddenly exposed to a fluid at 20°C with a convection coefficient of 50 W/(m2C), determine the
temperature at the node 3 after 1, 2, 5, 10, 15 and 20 min. Use explicit formulation and take Dx = Dy = 1 cm. Take thermal
diffusivity of the body as 3.2 ´ 10 –6 m2/s.
Solution.
Data:
Dx := 0.01 m Dy := 0.01 m Ta := 20°C h := 50 W/(m2 C) k := 20 W/(mC) a : = 3.2 ´ 10 –6 m 2/s
Nodes are represented by numbers 1, 2, ..., 7. Elemental volume pertinent to each node is also marked around it and
numbered a, b, ..., r.
We shall develop finite difference equations for each node by writing the energy balance for the corresponding
elemental volume around that node.

NUMERICAL METHODS IN HEAT CONDUCTION 373


TABLE 8.2 Summary of transient, finite difference equations for different boundary conditions
(q = heat flux, h = convection heat transfer coefficient k = thermal conductivity, no internal heat generation,
and Dx = Dy)
Situation Finite difference equation
(with D x = D y, no heat generation)
(1) Node at an internal corner with convection, Fig. 8.12,a:
Explicit method:
2 4 FG IJ
× Fo ×[2×(Tm, n + 1)i + 2×(Tm – 1, n)i + (Tm + 1, n)i + (Tm, n – 1)i + 2 × Bi × Ta] + 1 - 4 ×Fo - ×Fo ×Bi ×(Tm, n )i ...(8.71)
(Tm, n)i + 1 =
3 3 H K
Stability criterion for above:
3
Fo ×(3 + Bi ) £ ...(8.72)
4
Implicit method:
LM1 + 4 ×Fo × FG1 + Bi IJ OP × (T 2 ×Fo 4
H 3 KQ
i+1
m, n) – ×[(Tm + 1, n)i + 1 + (Tm, n – 1)i + 1 + 2 ×(Tm, n – 1 )i + 1 + 2×(Tm + 1, n)i + 1] – ×Bi ×Fo ×Ta
N 3 3
– (Tm, n)i = 0 ...(8.73)
2. Node at a plane surface with convection Fig. 8.12b:
Explicit method:
(Tm, n)i + 1 = Fo ×[2×(Tm – 1, n)i + ( Tm, n + 1)i + (Tm, n – 1 )i + 2× Bi × Ta ] + (1 – 4× Fo – 2× Fo × Bi )×(Tm, n)i ...(8.74)
Stability criterion for above:
1
Fo ×(2 + Bi ) £ ...(8.75)
2
Implicit method:
[1 + 2× Fo ×(2 + Bi )]2×(Tm, n)i + 1 – Fo ×[2×(Tm – 1, n)i + 1 + (Tm, n + 1)i + 1 + (Tm, n – 1 )i + 1] = (Tm, n)i + 2× Bi × Fo × Ta ...(8.76)
3. Node at a plane surface, insulated:
To obtain finite difference equation or stability criterion for an insulated surface (or a surface of thermal symmetry), set
B i = 0 (i.e. h = 0) in Eqs. 8.74, 8.75 or 8.76.
4. Node at exterior corner, with convection Fig. 8.12c:
Explicit method:
(Tm, n)i + 1 = 2 × Fo ×[(Tm – 1, n)i + (Tm, n – 1)i + 2× Bi × Ta ] + (1 – 4× Fo – 4× Fo × Bi )×(Tm, n)i ...(8.77)
Stability criterion for above:
1
Fo ×(1 + Bi ) £ ...(8.78)
4
Implicit method:
(1 + 4× Fo ×(1 + Bi ))×(Tm, n)i + 1 – 2× Fo ×[(Tm – 1, n)i + 1 + (Tm, n – 1)i + 1] = (Tm, n)i + 4× Bi × Fo × Ta ...(8.79)

Convection Dx = Dy = 1 cm

y
2
h = 50 W/(m C)
Insulated a Ta = 20 C
1 3

c b d
2 4 g 5 h 6 k 7 Insulated

Dy r j p
e f i
x
T = 150 C
Dx
FIGURE Example 8.11 Finite difference representation for two- dimensional conduction-nodal network

374 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For node 1:
Elemental volume to be considered is 1/4 volume, 1-a-b-c-1. This node is subjected to convection from top and conduc-
tion from right and conduction from bottom.
From left surface, there is no heat transfer, since it is insulated.
Considering all heat flows to be flowing into the elemental volume, and writing an energy balance, we get:
FG D x ×1IJ ×(T FG D y ×1IJ × T - Ti i
FG D x ×1IJ × T - T
i i
Dx Dy i
T + 1 - T1
i

H2 K – T 1i) + k×
H 2 K Dx + k×
H 2 K Dy
3 1 2 1
a = r× × ×Cp × 1
2 2 Dt

Remembering that Dx = Dy, and dividing by k/4 and simplifying:


i i
2 ×h × D x T + 1 - T1
×(Ta – T 1i) + 2×(T3i – T1i ) + 2×(T2i – T 1i) = 1
k Fo
FG
T 1i + 1 = 1 - 4 × Fo - 2 × Fo ×
h×Dx IJ
×T 1i + 2×Fo× T3 + T2 +
i i h× D xFG IJ
i.e.
H k K k H
× Ta
K ...(a)

Eq. a is the explicit difference equation for temperature of node 1 at (i + 1)st time step.
For node 2:
Here, elemental volume to be considered is 1/2 volume, c-b-e-r. and, energy balance can be written as we did for node 1.
However, since the surface is insulated, it is easier to use the mirror image concept and consider the node 2 as an
internal node. So, to the left of node 2, we have T4, mirror image of temperature of node 4. Then, considering 2 as
internal node, we get difference eqn for node 2, from Eq. 8.69:
(Tm, n)i + 1 = Fo×[(Tm – 1, n)i + (Tm + 1, n)i + (Tm, n + 1)i + (Tm, n – 1)i ] + (1 – 4 ×Fo)×(Tm, n)i ...(8.69)
i.e. T2i + 1 = (1 – 4×Fo)×T2i + Fo ×(T4i + T4i + T1i + 150) ...(b)
Eq. b is the explicit difference equation for temperature of node 2 at (i + 1)st time step.
For node 3:
This is corner node with convection. Elemental volume to be considered is 1/4 volume, a-3-d-b.
Applying energy balance, with all heat flow lines into the volume:

FG D x + D y IJ ×(T F D y T1i - T3i I + F k× Dx × T - T I i i


Dx D y i
T + 1 - T3
i

H2 2K a GH
– T3i ) + k ×
2
×
Dx JK GH 2 D y JK 4 3
= r× ×
2 2
×Cp × 3
Dt

Dividing by k/4 and simplifying,


FG h× D xIJ i i FG
h× D x IJ
H
T 3i + 1 = 1 - 4 × Fo - 4 × Fo ×
k K
× T3i + 2 ×Fo× T1 + T4 + 2 ×
k H × Ta
K ...(c)

Eq. c is the explicit difference equation for temperature of node 3 at (i + 1)st time step.
For node 4:
This is an internal corner node with convection on two sides. Elemental volume to be considered is 3/4 volume, g-f-e-b-
d-4.
Again, applying energy balance, with all heat flow lines into the volume:
LMh × F D x + D y I ×(T - T ) + F k × D y × T - T I + F k × D x × 150 - T I + F k × D y × T - T I + F k × D x × T - T I OP
i i i i i i i

MN GH 2 2 JK GH 2 D x JK GH D y JK GH D x JK GH D y JK PQ
i 5 4 4 2 4 3 4
a 4
2
i i
3 ×D x ×D y T + 1 - T4
= r× × Cp × 4 Dividing by 3k/4 and simplifying,
4 Dt

F h× D xI Fo FG h× D x IJ
GH
T 4i + 1 = 1 - 4 × Fo - 4 × Fo ×
3× k JK
×T 4i +
3
i
H
i i
× 4 × T2 + 4 ×150 + 2 × T5 + 2 × T3 + 4 ×
k
× Ta
K ...(d)

Eq. d is the explicit difference equation for temperature of node 4 at (i + 1)st time step.
For node 5:
This is a surface node with convection. Elemental volume to be considered is 1/2 volume, g-f-i-h-g.

NUMERICAL METHODS IN HEAT CONDUCTION 375


Again, applying energy balance,
F k × D y × T - T I + F k × D x × 150 - T I + F k × D y × T - T I = r×Dx × D y ×C × T
i i i i i i+1 i
- T5
h×Dx ×(Ta – T5i ) + GH 2 D x JK GH
6 5
D y JK GH 2
5
D x JK
4 5
2
p
5
Dt
Dividing by k/2 and simplifying,
FG
T 5i + 1 = 1 - 4 × Fo - 2 × Fo ×
h×Dx IJ FG
×T 5i + Fo × T4 + T6 + 2 × 150 + 2 ×
i i h×Dx IJ
H k K H k
× Ta
K ...(e)

Eq. e is the explicit difference equation for temperature of node 5 at (i+1)st time step.
For node 6:
This is identical to node 5. So, we get, by shifting node numbers by 1:
FG h×Dx IJ i i FG h×Dx IJ
H
T 6i + 1 = 1 - 4 × Fo - 2 × Fo ×
k K H
×T 6i + Fo × T5 + T7 + 2 × 150 + 2 ×
k
× Ta
K ...(f)

Eq. f is the explicit difference equation for temperature of node 6 at (i+1)st time step.
For node 7:
This is corner node with conduction from left, convection on the top, insulated on the right, and conduction from down.
Elemental volume to be considered is 1/4 volume, k-7-p-j.
Writing the energy balance, remembering that right surface is insulated, and all heat flow lines into the volume, we
get:
D y T6i - T7i Dx D x 150 - T7i Dx Dy i
T + 1 - T7
i
k× × + h× ×(Ta – T7i) + 0 + k× × = r× × × Cp × 7
2 Dx 2 2 Dy 2 2 Dt
Dividing by k/4 and simplifying,
FG
T 7i + 1 = 1 - 4 × Fo - 2 × Fo ×
h×Dx IJ
×T 7i + 2× Fo × T6 + 150 +
i FG
h× D x IJ
H k K Hk
× Ta
K ...(g)

Eq. g is the explicit difference equation for temperature of node 7 at (i+1)st time step.
Now, we have to fix the upper limit of Dt from stability criterion. To do that, we observe that in Eqs. a to f, the
smallest coefficient of Tmi is in Eq. c, i.e. (1 - 4. Fo -4.Fo.h.Dx/k) must be positive. Putting this condition, we get:
h ×D x
1 – 4 ×Fo – 4×Fo× ³0
k
1
i.e. Fo £
F h × D x IJ
4 ×G1 +
H k K
( D x )2
i.e. Dt £
FG h× D x IJ
H
4 ×a × 1 +
k K
i.e. Dt £ 7.622 s
This means that a time step less than 7.622 s has to be employed from stability criterion. Let us choose:
Dt := 5 s
a ×Dt
Then, Fo := i.e. Fo = 0.16
( D x )2
Substituting all relevant numerical values in Eq. a to f, we get the explicit difference equations as:
T1i + 1 = 0.352×T1i + 0.32×(T3i + T2i + 0.5) ...(a)
T2i + 1 = 0.36 ×T2i + 0.16 ×(T4i + T4i + T1i + 150) ...(b)
T3i + 1 = 0.344×T3i + 0.32×(T1i + T4i + 1) ...(c)
T4i + 1 = 0.35467×T4i + 0.05333×(4 ×T2i + 2×T5i + 2×T3i + 602) ...(d)
T5i + 1 = 0.352×T5i + 0.16×(T4i + T6i + 301) ...(e)
T6i + 1 = 0.352×T6i + 0.16×(T5i + T7i + 301) ...(f)
T7i + 1 = 0.352×T7i + 0.32×(T6i + 150.5) ...(g)
Initial temperature of the plate at t = 0 and i = 0, is given as 150°C.
i.e. T10 = T20 = T30 = T40 = T50 = T60 = T70 = 150°C

376 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore, at the next time step i = 1, i.e. at Dt = 5 s, temperatures at nodes 1 to 7 can be explicitly calculated from
Eqs. a to g. Then, calculate temperatures at the nodes for next time step of Dt = 10 s, using the same Eqs. a to g, since the
temperatures at the previous time step are already calculated. Thus, march in time till we reach the time limit specified
in the problem. Note that for 1 min. there are 12 time steps of 5 s each.
This calculation is easily done in Mathcad. We slightly change the notation for convenience in calculation: we write
the superscripts as subscripts to work with matrix notation, as shown below.
In the small Mathcad program given below, LHS defines a function Temp(n) where n is the no. of time steps, which
we can specify. Output is a vector containing step number total time elapsed, and node temperatures T1, T2, ..., T7.
On the RHS, first 7 lines define the initial temperatures at the nodes, all equal to 150°C.
Then, a ‘for loop’ evaluates the finite difference Eqs. a to g, each ‘new’ node temperature being calculated in terms
of the temperatures calculated in the previous time step. Here, the no. of time steps, ‘n’ can be changed since it is
included in function definition on the LHS.

Temp (n) = T 1o ¬ 150


T 2 o ¬ 150
T 3 o ¬ 150
T 4 o ¬ 150
T 5 o ¬ 150
T 6 o ¬ 150
T 7 o ¬ 150
for i Î 0 , ... , n
T 1i + 1 ¬ 0. 352 × T 1i + 0 . 32 × (T 3 i + T 2 i + 0 . 5)
T 2 i + 1 ¬ 0 . 36 × T 2 i + 0 .16 × (T 4 i + T 4 i + T 1i + 150)
T 3 i + 1 ¬ 0 . 344 × T 3 i + 0 .32 ×(T 1i + T 4 i + 1)
T 4 i + 1 ¬ 0. 35467 × T 4 i + 0 .05333 × ( 4 × T 2 i + 2 × T 5 i + 2 × T 3 i + 602 )
T 5 i + 1 ¬ 0 . 352 × T 5 i + 0.16 ×(T 4 i + T 6 i + 301)
T 6 i + 1 ¬ 0. 352 × T 6 i + 0.16 × (T 5 i + T 7 i + 301)
T 7 i + 1 ¬ 0. 352 × T 7 i + 0 .32 × (T 6 i + 150 . 5)
[i 5 × i T 1i T 2 i T 3i T 4 i T 5 i T 6 i T 7 i ]

Check: Temp(0) = [0 0 150 150 150 150 150 150 150] (starting at time = 0)
i = step no.; Dt= one time step = 5 s; t = time duration from beginning = i. Dt , s
i t T1 T2 T3 T4 T5 T6 T7
Temp(12) = [12 60 142.823 146.402 141.661 145.968 146.68 146.827 146.85]
Temp(24) = [24 120 141.573 145.591 140.488 145.326 146.442 146.726 146.78]
Temp(60) = [60 300 141.331 145.434 140.261 145.203 146.398 146.708 146.729]
Temp(120) = [120 600 141.33 145.434 140.26 145.202 146.398 146.708 146.769]
Temp(180) = [180 900 141.33 145.434 140.26 145.202 146.398 146.708 146.769]
Temp(240) = [240 1.2 × 10 3 141.33 145.434 140.26 145.202 146.398 146.708 146.769]
In the above Table, first column gives step number second column gives the time elapsed (seconds), and 5th column
gives the temperature of node 3. We get:
Time (min) Temperature of node 3 (deg.C)
0 150
1 141.661
2 140.488
5 140.261
10 140.26
15 140.26
20 140.26
i.e. steady state is reached after about 5 min. from start up.

NUMERICAL METHODS IN HEAT CONDUCTION 377


8.9 Accuracy Considerations
As mentioned earlier, numerical methods yield approximate values as compared to ‘exact analytical solutions’.
This is due to the following errors inherent in numerical methods:
(i) Discretization error This is due to the approximation used in formulation of numerical solutions i.e. due
to the error involved in writing the derivatives in terms of differences. Remember that we assume the
temperature variation between adjacent nodes to be linear, which may not be so in practice. This is
equivalent to considering only first two terms in the Taylor series expansion of the temperature at the
given node. Generally, discretization error is cumulative; but if the function changes sign, it is possible
that the errors may cancel. Discretization error is proportional to the square of the time step Dt (or Dx).
Therefore, smaller the mesh size, smaller the discretization error.
(ii) Round off error This is due to the fact that computer retains only 15 digits accuracy in a calculation (in
double precision) and the rest of the digits are either chopped off or rounded off. When this is done
continuously for a large number of calculations, error is carried over to successive calculations and the
cumulative error can be significant. Obviously, the round off error is proportional to the total number of
computations performed, and reduces as the mesh size increases.
Therefore, while aiming at reducing the error involved in numerical methods, we note that we have to deal
with two opposing effects: if the mesh size Dx (or time step size Dt) is decreased, discretization error is reduced,
but the round off error increases since the total number of calculations increases. So, practical way of approach-
ing the solution is to start with a coarse mesh and then gradually refine the mesh size and observe if the results
converge.
As a note of caution, it should be pointed out that getting an accurate solution of the nodal equations may
not necessarily mean that an accurate solution to the physical problem has been obtained, if the formulation of
nodal equations itself is erroneous. Therefore, as a check, some sort of energy balance using the final solution is
also recommended.

8.10 Summary
While considering heat transfer in solids with complicated geometries and boundary conditions, and tempera-
ture dependent thermal properties, it is difficult to formulate ‘exact’ analytical solutions. In such cases, numerical
methods are adopted to determine the temperature distribution and heat transfer rates.
In this chapter, we first considered the numerical solution of one-dimensional steady state conduction in
cartesian, cylindrical and spherical coordinates. Then, two-dimensional conduction in cartesian coordinates was
studied. Finally, numerical solutions for one-dimensional and two-dimensional transient conduction in cartesian
coordinates was explained.
‘Finite difference method’ involves converting the partial differential equations of heat transfer into a set of
coupled algebraic equations and then solving them. Analytical solution gives the temperature at any point in the
medium; however, in numerical method, we divide the volume into discrete subvolumes and each subvolume is
represented by a ‘node’ and the temperatures are determined at these discrete nodes.
Method adopted to convert the differential equations into a set of algebraic equations is to write an energy
balance at each node. As a rule, all heat flow lines are considered to be flowing into the node considered. While
writing the energy balance for a steady state problem, sum of all heat flows into the node must be equal to zero.
Nodes at the boundaries for different boundary conditions are also handled in the same way, i.e. by writing
energy balances at the boundary nodes. Care must be taken to see that at any boundary node, the volume consid-
ered must be the one appropriate to that node (i.e. half volume for a surface node, ¼ volume for an external
corner node, ¾ volume for an internal corner node etc.)
Solution of the set of algebraic equations may be obtained by ‘direct methods’ or by ‘iteration methods’.
Direct methods are: Gaussian elimination method and Matrix inversion methods. Example of iteration method is
the popular ‘Gauss – Siedel iteration method’ where one starts with the guess values of temperatures. These
methods were explained briefly in this chapter.
While considering the numerical method for transient conduction, we again adopt the technique of writing
the energy balances at the nodes; however, now we say that net energy flowing into a node results in a variation
of the energy content of the subvolume represented by that node during the time interval Dt. In the ‘explicit
method’, heat transfer and heat generation terms are considered at the ‘previous time step’ i, whereas in the
‘implicit method’, these terms are considered at the ‘new time step’, i + 1. In the explicit formulation, tempera-
tures are obtained in a straightforward manner in terms of the values obtained at the previous time step. How-

378 FUNDAMENTALS OF HEAT AND MASS TRANSFER


ever, explicit method suffers from a disadvantage that we cannot use any time step we like and the solution
becomes unstable unless the time step is below a particular value as dictated by ‘stability criterion’. In implicit
formulation, there is no such limitation on the time step i.e. any larger time step can be used resulting in smaller
number of total calculations; however, at each step, all the equations have to be solved simultaneously.
While discussing the accuracy of the numerical solution, it was pointed out that smaller the mesh size, better
the accuracy; but, now the total number of computations will be more and this introduces larger round off errors.
Further, from a practical point of view, when convection boundary conditions are involved (which is invariably
the case), the uncertainty in the value of heat transfer coefficients itself may be of the order of 20% and it is quite
likely that the thermal properties may also be in error by 10 to 15%. Therefore, there is no point in having an
unduly fine mesh. So, practical way of approaching the solution is to start with a coarse mesh and then progres-
sively refine the mesh size, observing that the temperature values at given nodes go on converging.

Questions
1. When is a numerical solution adopted for a problem? What are its advantages and limitations?
2. Mention the methods used to convert partial differential equations of conduction heat transfer into finite differ-
ence equations.
3. Explain the energy balance procedure to obtain the finite difference formulation of one-dimensional conduction
problem in cartesian coordinates.
4. Explain the energy balance procedure to obtain the finite difference formulation of one-dimensional conduction
problem in cylindrical and spherical coordinates.
5. Explain the procedure of writing finite difference equation for an insulated boundary.
6. Explain the ‘direct’ and ‘iterative’ methods used for the solution of a system of algebraic equations.
7. ‘Heat transfer problems involving variable thermal conductivity and radiation boundary conditions are difficult
to handle’ – explain this statement.
8. Give two examples of two-dimensional conduction where numerical methods are employed conveniently.
9. Finite difference formulation for a general interior node in a medium is given by:
qg ×( D x)2 Tm
i +1
- Tm
i
Tm – 1, n + Tm + 1, n + Tm, n + 1 + Tm, n – 1 – 4×Tm, n + =
k Fo
(i) Is the heat transfer in this medium steady or transient?
(ii) Is there heat generation in the medium?
(iii) Is the heat transfer one, two or three-dimensional?
(iv) Is the nodal spacing constant or variable?
(v) Is the thermal conductivity of the medium constant or variable?
10. Explain the method of handling an irregular boundary while writing finite difference equations.
11. How does the procedure of finite difference formulation for transient conduction differ from that for steady state
conduction?
12. Explain the principle of getting ‘explicit’ and ‘implicit’ formulations for transient conduction.
13. Explain the ‘stability criterion’ when using explicit formulation for one-dimensional and two-dimensional tran-
sient conduction.
14. What are the relative advantages and disadvantages of explicit and implicit formulations?
15. Explain the types of errors inherent in numerical methods. How to reduce these errors?
16. How does the step size influence the discretization and round off errors?

Problems
One-dimensional steady state conduction:
1. A large plane wall of thickness L = 0.5 m, thermal conductivity k = 14 W/(mC), and surface area A = 20 m2, has
its left face maintained at a constant temperature of 150°C and the right face is exposed at ambient at 20°C, with
a heat transfer coefficient of h = 20 W/(m2C). Assuming a nodal spacing of 10 cm, and steady one-dimensional
heat transfer, formulate the finite difference equations for all nodes and solve them to find the temperatures at
all nodes. What is the rate of heat transfer through this wall?
2. A plane wall of thickness 0.1 m and k = 20 W/(mC) has uniform heat generation of 0.35 MW/m3. It is insulated
on one side and the other side is subjected to convection heat transfer with a fluid at 90°C flowing with a heat
transfer coefficient of 550 W/(m2C). Determine the temperature distribution in the wall by finite difference
method.

NUMERICAL METHODS IN HEAT CONDUCTION 379


3. A 30 mm diameter copper cable carries 200 A and has an electrical resistance of 5 milli-ohms per metre. Cable
loses heat to the ambient air at 15°C with a convection coefficient of 20 W/(m2C). Determine the temperature
distribution by numerical method. Compare the results with exact solution.
4. A 0.6 cm diameter solid aluminium sphere (k = 200 W/(mC) has an energy generation rate of 108 W/m3. Sphere
loses heat from its outside surface to an ambient at 25°C by convection with a heat transfer coefficient of 150 W/
(m2C). Calculate the steady state radial temperature distribution by the finite difference method by dividing the
region into 6 elements, each of radial thickness Dr = 0.05 cm. Compare the results with exact solution.
5. Consider a straight fin of circular cross-section, 5 mm in diameter and 50 mm long (k = 14 W (mC)). Surface of
the fin is exposed to ambient air at 20°C with a convection heat transfer coefficient of 80 W/(m2C). Base of the
fin is maintained at 150°C. Assuming the tip of the fin to be insulated, determine the temperature distribution in
the fin, heat transferred and the fin efficiency by finite difference method. Use 5 equal subdivisions along the
length. Compare your results with the exact solution.
6. Consider an aluminium straight fin of square cross-section (4 mm ´ 4 mm), 2 cm long (k = 200 W (mC)). Surface
of the fin is exposed to ambient air at 20°C with a convection heat transfer coefficient of 25 W/(m2 C). Base of the
fin is maintained at 150°C. Assuming that the tip of the fin is also losing heat by convection, determine the
temperature distribution in the fin, heat transferred and the fin efficiency by finite difference method. Use 4
equal subdivisions along the length. Compare your results with the exact solution.
Two-dimensional steady state conduction:
7. A long rod of square cross-section (3 cm ´ 3 cm), has its top and bottom surfaces maintained at 0°C while the left
and right surfaces are maintained at 50°C and 100°C respectively. Determine the steady state temperature distri-
bution in the rod, using a node spacing of 1 cm.
8. A long rod of square cross-section (2 cm ´ 2 cm), has its top surface maintained at 120°C while each of the other
three surfaces is maintained at 80°C. Determine the steady state temperature distribution in the rod, using a
node spacing of 0.5 cm. Check the results with analytical solution.
9. Refer to the L-bar shown in Fig. Problem 8.9.
If the thermal conductivity of the material is 15 W/(mC), find out the temperatures at all the nodes. (Note: Dx =
Dy = 1 cm)

Convection Dx = Dy = 1 cm

y
2
h = 30 W/(m C)
Insulated a Ta = 20 C
1 3

c b d
2 4 g 5 h 6 k 7 Insulated

Dy r j
e f i p
x
T = 200 C
Dx

FIGURE Problem 8.9 Two-dimensional conduction in a L-bar

10. If in the L-bar shown in Fig. Problem 8.9 above, there is a heat generation at a rate of 1 MW/m3, all the other
data remaining the same, determine the temperatures at all nodes.
11. If in the L-bar shown in Fig. Problem 8.9 above, if the right face is also subjected to convection conditions of the
top surface, all the other data remaining the same, determine the temperatures at all nodes.
12. Consider a long bar of rectangular cross-section (6 cm wide ´ 9 cm height), with a thermal conductivity of 14
W/(mC). Top surface of the bar (with 60 mm width) is exposed to air at 90°C with a convection coefficient of 80
W/(m2C), while the other three surfaces are maintained at 35°C. Using a nodal spacing of 1.5 cm, determine the
steady state temperature distribution in the bar and the heat transfer rate per unit length of the bar.
13. A gas duct made of fire brick, (k = 1 W/(mC)), has outer dimension of 4 m ´ 4 m. Gas passage area is 2 m ´ 2 m,
centrally located. Inner walls are at a temperature of 900°C and the outer walls are at 40°C. Determine the
temperature distribution in the wall and the heat transfer rate per metre length.

380 FUNDAMENTALS OF HEAT AND MASS TRANSFER


14. Consider two-dimensional steady state conduction in a region, y
2 cm ´ 2 cm, with the boundary conditions as shown in Fig.
Problem 8.14. For the material, k = 60 W/ (mC) and there is
internal heat generation at a rate of 10 7 W/m3. Using finite dif- T=0 C
2
ference method, calculate the unknown node temperatures.
One-dimensional transient conduction: 1 2 3
15. A very thick copper plate (k = 386 W/(mC), aa = 11 ´ 10 –5 m2/
s) is initially at 400°C. Suddenly, its surface temperature is 4 5 6 T=0 C
T=0 C
lowered to 20°C. Considering the plate as semi-infinite plate
and using a mesh size Dx = 1 cm, calculate the temperature at x 7 8 9
= 5 cm from the surface, 2 min. after lowering the surface tem-
perature.
10 11 12 x
16. A water main is buried below the surface of soil which is ini-
0 2
tially at an uniform temperature of 25°C. Suddenly, the surface
temperature drops to –30°C and is maintained so for a period Insulated
of 60 days. Determine the depth at which the water mains FIGURE Problem 8.14 Two-dimensional
must be placed to avoid freezing of water. Take properties of steady state conduction
soil as: r = 2050 kg/m3, k = 0.52 W/(mC), Cp = 1840 J/(kgK), a
–6 2
= 0.138 ´ 10 m /s. (Hint: Consider the soil as semi-infinite medium; calculate temperatures at distances upto 6
m below the surface and find the depth at which the temperature would be 0°C, by interpolation).
17. A 6 cm thick steel plate (a = 1.6 ´ 10 –5 m2/s, k = 60 W/(mC)), is initially at an uniform temperature of 250°C. It
is suddenly exposed to a cold air stream at 20°C on both the surfaces, with a heat transfer coefficient of 350 W/
( m2C). Determine the centre plane temperature at t = 5, 10 and 15 min. from starting of cooling. Use explicit
formulation with a mesh size of Dx = 1 cm.
18. Two ends of a steel rod 1.2 cm diameter and 2.5 m long, are maintained at 250°C and 50°C and the curved
surface of the rod is perfectly insulated. Suddenly, an electric current is passed through the rod, causing heat
generation in the rod at an uniform rate of 3000 W/m3. Find the temperature distribution in the rod for the first
five time increments. Take k = 35 W/(mC) and a = 1.5 ´ 10 –5 m2/s.
Two-dimensional transient conduction:
19. The L-bar shown in Fig. Problem 8.9 is initially at an uniform temperature of 200°C. Its top surface is suddenly
exposed to convection with an air stream at 20°C with a convection coefficient of 80 W/(m2C). Bottom surface is
maintained at 200°C throughout and the left and right surfaces are insulated as shown. Taking k = 15 W/(mC)
and a = 3.2 ´ 10 –6 m2/s, calculate the temperature of node 3 after 1, 3, 5, 10 and 30 min. Use explicit formulation.
20. A steel bar of 3 cm ´ 3 cm cross-section is initially at an uniform temperature of 500°C. (a = 1.0 ´ 10– 5 m2/s, k =
35 W/(mC)). Suddenly, all the 4 surfaces of the bar are exposed to an air stream at 20°C with a heat transfer
coefficient of 120 W/(m 2C). Using explicit formulation and a mesh size of Dx = Dy = 0.5cm, calculate the centre
temperature at t = 1, 5 and 10 min. after the start of cooling. (Hint: Use symmetry consideration—consider only
a quarter of the cross-section).

NUMERICAL METHODS IN HEAT CONDUCTION 381


CHAPTER

9
Forced Convection

9.1 Introduction
In the previous chapters, we studied about conduction heat transfer, where heat transfer was a molecular phe-
nomenon and was considered mainly in solids; convection was mentioned only in passing and was considered
only as a boundary condition while analysing conduction heat transfer.
In convection heat transfer, there is a flow of fluid associated with heat transfer and the energy transfer is
mainly due to bulk motion of the fluid. When the flow of fluid is caused by an external agency such as a fan or
pump or due to atmospheric disturbances, the resulting heat transfer is known as ‘Forced convection heat trans-
fer’; when the flow of fluid is due to density differences caused by temperature differences, the heat transfer is
said to be by ‘Natural (or free) convection’. For example, if air is blown on a hot plate by a blower, heat transfer
occurs by forced convection, whereas, a hot plate simply hung in air will lose heat by natural convection.
In this chapter, we shall study about forced convection heat transfer. Since there is a flow of fluid involved
in convection heat transfer, it is clear that the flow field will influence the heat transfer greatly. Mathematical
solution of convection heat transfer will, therefore, require the simultaneous solution of differential equations
resulting by the application of conservation of mass, conservation of momentum and conservation of energy,
under the constraints of given boundary conditions. For a three-dimensional fluid flow, mathematical solution of
the resulting differential equations is a formidable task and it is usual to make many simplifying assumptions to
get a mathematical solution. Still, it must be stated that exact mathematical solutions, even for simple convection
heat transfer cases, are rather complicated and it is common practice to resort to empirical relations for solutions
of problems involving convection heat transfer. These empirical relations are obtained by researchers after per-
forming large number of experiments for several practically important situations and are presented in terms of
non-dimensional numbers.
In this chapter, we shall first describe the physical mechanism of forced convection and then mention about
the convective heat transfer coefficient and various factors affecting the same. Then, we shall discuss about veloc-
ity and thermal boundary layers. Application of conservation of mass, momentum and energy in respect of the
boundary layer will be demonstrated next. We shall not rigorously solve these equations, but will only mention
the methods of solution, since our emphasis will be on practical solutions with the use of empirical relations.
Then, we present several empirical relations to determine friction and heat transfer coefficients for flow over
different geometries such as a flat plate, cylinder and sphere for flow under laminar and turbulent conditions.
Finally, flow inside tubes will be considered and determination of heat transfer coefficient by analogy with the
mechanism of fluid flow will be explained.

9.2 Physical Mechanism of Forced Convection


Consider a hot iron block whose surface is at a temperature Ts. Let this surface be cooled by a fluid at a tempera-
ture Ta, flowing over its surface at a velocity U, as shown in Fig. 9.1.
We know that heat will be carried away from the hot iron block by the flowing fluid and the block will cool.
We also know that if the velocity of the fluid is increased, more heat is carried away and the block will be cooled
faster. For the purpose of analysis, we quantify the preceding statement by a dimensionless number called,
Fluid flow
U, Ta
y

U Surface of block
Qc Ty
uy
¶T/¶y|y=0

Ta Ts

FIGURE 9.1 Temperature and velocity distribution in laminar, forced convection over a hot block

‘Reynolds Number’, in honour of Osborne Reynolds, an English scientist. Reynolds number is defined as fol-
lows:
x×r
Re = U …(9.1)
m
where U = mean velocity of flow, m/s
r = densiy of fluid, kg/m 3
m = dynamic viscosity of fluid, kg/(ms), and
x = a characteristic dimension of the flow passage, equal to the linear distance along the flow direction
in the case of a flat plate or the pipe diameter in the case of a flow through a pipe. For non-circular passages (such
as a square or rectangular passage), the characteristic dimension in Eq. 9.1 is the ‘equivalent diameter’, defined
as:
4×Ac
de = …(9.2)
P
where de = equivalent diameter, m
Ac = area of cross-section, m 2, and
P = wetted perimeter, m
For a rectangular cross-section of breadth ‘a’ and height ‘b’, we get from Eq. 9.2:
4×Ac 4 × a ×b 2× a × b
de = = = …(9.2,a)
P 2 × ( a + b) ( a + b)
And, for an annulus formed by a tube of outer diameter d 1 placed within a tube of inner diameter d2, equiva-
lent diameter is calculated as:
p 2
× (d2 - d12 )
de = 4× 4 = d2 – d1 …(9.2,b)
p ×( d1 + d2 )
Note that Eq. 9.2 is used in connection with the calculation of pressure drop for flow through an annulus;
but, for the case of heat transfer, say from a hot fluid flowing through the inner tube to a cold fluid flowing
through the outer tube, since the heat transfer occurs only through the surface of the inner tube, we use for the
equivalent diameter:
p 2
× (d2 - d12 )
4 d 2 - d12
de = 4× = 2 …(9.2c)
p × d1 d1
If the Reynolds number is below a certain value, as determined by experiments, the flow is laminar; i.e. the
fluid layers move parallel to each other in an orderly manner. As the velocity of flow increases, i.e. as the value

FORCED CONVECTION 383


of Reynolds number increases, there is more disorder in the fluid and the fluid flow becomes ‘turbulent’; fluid
‘chunks’ move at random and obviously the heat transfer increases, since these chunks of fluid carry the heat
with them. Transition from laminar to turbulent flow occurs not at a fixed value of Reynolds number, but occurs
in a range called ‘transition range of Reynolds numbers’. For example, for flow through a pipe, at values of
Reynolds number below 2300 the flow is laminar, for values above 4000 the flow is turbulent and in between is
the transition range. Value of Reynolds number is affected by fluid properties, dimension of flow passage and
also by surface conditions.
Fig. 9.1 also shows the velocity profile and the temperature profile for laminar flow. The velocity profile is
parabolic. As the flowing fluid comes in contact with the surface of the block, a thin layer adheres to the surface
and essentially remains stationary with zero velocity; this phenomenon is known as ‘no slip condition’ in the
terminology of fluid mechanics. The fluid layer adjacent to this layer has its velocity retarded as compared to the
free stream velocity due to the effect of viscosity of the fluid, and the next layer has slightly higher velocity, etc.
till the free stream velocity is attained at a layer farther away from the surface. The point we are trying to make
here is that immediately next to the solid surface, there is essentially a stationary layer of fluid and the heat
transfer through this fluid layer is by ‘pure conduction’; subsequently, since the next layers of fluid are in motion
convection heat transfer occurs.
For this stationary fluid layer, the heat flux is given by Fourier’s law:
q cond = – kf (dT/dy)|y = 0 …(9.3)
where kf is the thermal conductivity of the fluid and (dT/dy)|y = 0 is the temperature gradient at y = 0 i.e. at the
surface.

9.3 Newton’s Law of Cooling and Heat Transfer Coefficient


Governing rate equation for convection heat transfer is given by ‘Newton’s Law of Cooling’ (also known as
‘Newton–Rikhman Law’). According to this law, the heat flux in convection heat transfer is given by:
q conv = h×(Ts – Ta ) …(9.4)
where h is the convective heat transfer coefficient and (Ts – Ta) is the temperature difference between the hot
surface and the flowing fluid. Unit of h is: W/(m2C) so that the heat flux has units of W/m 2.
Though Eq. 9.4 looks very simple, it is very subtle. The reason is: heat transfer coefficient, h, depends on
several factors such as:
(i) the fluid properties like density, viscosity, thermal conductivity and specific heat,
(ii) type of flow (laminar or turbulent),
(iii) shape of fluid passage (circular, rectangle or a flat surface),
(iv) nature of the surface (rough/smooth) and
(v) orientation of the surface
In fact, entire thrust in determining the heat transfer rate in convection is to find out this value of ‘h’ in a
reliable manner.

9.4 Nusselt Number


Since we know that adjacent to the solid surface the fluid layer is stationary and the heat transfer in this fluid
layer is by conduction, and the heat transferred by convection subsequently must be equal to this fluid layer, we
can equate Eqs. 9.3 and 9.4:
We can write:
h = {–k f (dT/dy)|y = 0}/(Ts – Ta) ...(9.5)
i.e. the problem of finding out the value of ‘h’ reduces to the task of finding out the temperature gradient (dT/dy)
at y = 0 i.e. at the surface.
Since the heat transfer coefficient depends on flow conditions, its value on a surface varies from point to
point. However, we generally take an average value of ‘h’ by properly averaging the local value of heat transfer
coefficient over the entire surface.
It is also common practice to non-dimensionalise the heat transfer coefficient with ‘Nusselt number’. Nusselt
number is defined as:
h ×d
Nu = …(9.5)
kf

384 FUNDAMENTALS OF HEAT AND MASS TRANSFER


where d is a characteristic dimension and k f is the fluid thermal conductivity.
To get a physical interpretation of the Nusselt number, consider a thin layer of fluid with thickness d and with a
temperature difference of DT between the two surfaces. Then, we have:
qconv h× DT h ×d
= = = Nu …(9.6)
qcond D T kf
kf ×
d
In other words, Nusselt number tells us how much the heat transfer is enhanced due to convection as com-
pared to only conduction. Or, higher the Nusselt number, larger the heat transfer by convection. If Nu = 1, it
means that heat transfer is by conduction alone.
Example 9.1. Air at 25°C flows over a flat surface maintained at 65°C. Temperature measured at a location of 0.3 mm
from the surface is 45°C. Find out the value of the local heat transfer coefficient. Thermal conductivity of air at the
average temperature may be assumed as 0.027 W/(mC).
Solution.
Data:
Ta = 25°C Ts = 65°C d = 0.0003 m DT = 45 – 65 = – 20°C kf = 0.027 W/(mC)
Now, to find out heat transfer coefficient, apply Eq. 9.5:
h = {–k f (dT/dy)y = 0}/(Ts – Ta) ...(9.5)
Now,
dT - 20
:= C/m (temperature gradient at the surface i.e. at y = 0)
dy 0 .0003

dT
i.e. = – 6.667 ´ 104 C/m
dy
(Note that temperature gradient is negative since, starting from the plate surface, as we proceed in the y direction,
T decreases as y increases.)
Therefore,
0 .027 ´ 6.667 ´ 10 4
h := ...from eqn. (9.5)
40
i.e. h = 45.002 W/(m2C).

9.5 Velocity Boundary Layer


Concept of ‘boundary layer’ was introduced by Ludwig Prandtl in the year 1904. According to this concept,
when a fluid flows over a surface, the flow field can be considered to be divided into two regions: one, a thin
layer adjacent to the solid surface, called the ‘boundary layer’, where the viscosity effects are predominant and
velocity and temperature gradients are very large, and second, a layer beyond the boundary layer where the
velocity and temperature gradients are equal to their free stream values. The boundary layer thickness (d) is
arbitrarily defined as that distance from the surface in the y-direction at which the velocity reaches 99% of the
free stream velocity, U. Boundary layer concept helps in simplification of momentum equations and, in particu-
lar, solution of viscous flow problems was greatly facilitated by this concept.
Let us first study the development of boundary layer for a flow over a flat plate. Flow over a flat plate is
important from a practical point of view, since flow over turbine blades and aerofoil sections of air plane wings
can be approximated as flow over a flat plate. See Fig. 9.2.
Consider a thin, flat plate. The leading edge and the trailing edge of the plate are shown in the Fig. 9.2. Let
a fluid approach the flat plate at a free stream velocity of U. The fluid layer immediately in contact with the plate
surface adheres to the surface and remains stationary, and in fluid mechanics, this phenomenon is known as ‘no
slip’ condition. Then, the fluid layer next to this stationary layer has its velocity retarded because of the viscosity
effects i.e. due to the frictional force or ‘drag’ exerted between the stationary and the moving layers. This effect
continues with subsequent layers upto some distance in the y-direction till the velocity equals the free stream
velocity U. This region of fluid layer in which the viscosity effects are predominant is known as the ‘velocity (or
hydrodynamic) boundary layer’, or simply the ‘boundary layer’. Thickness of the boundary layer is arbitrarily
defined as that distance in the y-direction from the plate surface at which the velocity is 99% of the free stream
velocity. Note the following points in connection with the boundary layer:

FORCED CONVECTION 385


y

x
Transition region

Laminar region Turbulent region


Fluid flow Velocity profile
U
U
Buffer layer
U
u Laminar sublayer
u

Leading edge Flat plate Trailing edge

FIGURE 9.2 Development of boundary layer over a flat plate

(i) The boundary layer divides the flow field into two regions: one, ‘the boundary layer region’ where the
viscosity effects are predominant and the velocity gradients are very steep, and, second, ‘the inviscid
region’ where the frictional effects are negligible and the velocity remains essentially constant at the free
stream value.
(ii) Since the fluid layers in the boundary layer travel at different velocities, the faster layer exerts a drag
force ( or frictional force) on the slower layer below it; the drag force per unit area is known as ‘shear
stress (t)’. Shear stress is proportional to the velocity gradient at the surface. This is the reason why, in
fluid mechanics, the velocity profile has to be found out to determine the frictional force exerted by a
fluid on the surface. Shear stress is given by:

F dU I
t s = m× GH dy JK y=0
N/m2 …(9.7)

where m is ‘dynamic viscosity’ of the fluid; its unit is kg/(ms) or N.s/m2. Viscosity is a measure of
resistance to flow. For liquids, viscosity decreases as temperature increases, whereas for gases, viscosity
increases as the temperature increases. Viscosities of a few fluids at 20°C are given in Table 9.1. It may be
observed that viscosity varies by several orders of magnitude for different fluids.
(iii) Use of Eq. 9.7 to determine the surface shear stress is not very convenient, since it requires a mathemati-
cal expression for the velocity profile; so, in practice, surface shear stress is determined in terms of the
free stream velocity from the following relation:

rU 2
t s = Cf , N/m2 …(9.8)
2

TABLE 9.1 Dynamic viscosity of a few fluids at 20°C

Fluid m (kg/( m.s)

Glycerin 1.49
Engine oil 0.80
Ethyl alcohol 0.00120
Water 0.00106
Freon-12 0.000262
Air 0.0000182

386 FUNDAMENTALS OF HEAT AND MASS TRANSFER


where Cf is a ‘friction coefficient’ or ‘drag coefficient’. r is the density of the fluid. C f is determined
experimentally in most cases. Drag coefficient varies along the length of the flat plate. Average value of
drag coefficient (Cfa) is calculated by suitably integrating the local value over the whole length of the plate
and then the drag force over the entire plate surface is determined from:

rU 2
FD = Cfa ×A× ,N …(9.9)
2
2
where A = surface area, m .
(iv) Starting from the leading edge of the plate, for some distance along the length of the plate, the flow in the
boundary layer is ‘laminar’ i.e. the layers of fluid are parallel to each other and the flow proceeds in a
systematic, orderly manner. However, after some distance, disturbances appear in the flow and beyond
this ‘transition region’, flow becomes completely chaotic and there is complete mixing of ‘chunks’ of fluid
moving in a random manner i.e. the flow becomes ‘turbulent’.
(v) Transition from laminar to turbulent flow depends primarily on the free stream velocity, fluid properties,
surface temperature and surface roughness, and is characterized by ‘Reynolds number’. Reynolds
number is a dimensionless number, defined as:
Re = (Inertia forces/Viscous forces)
Or,
U ×x
Re = …(9.10)
n
where U = free stream velocity, m/s
x = characteristic length i.e. for a flat plate it is the length along the plate in the
flow direction, from the leading edge, and
n = kinematic viscosity of fluid = m/r, m2/s, where r is the density of fluid.
When the Reynolds number is low, i.e. when the flow is laminar, inertia forces are small compared to
viscous forces and the velocity fluctuations are ‘damped out’ by the viscosity effects and the layers of
fluid flow systematically, parallel to each other. When the Reynolds number is large, i.e. when the flow is
turbulent, inertia forces are large compared to the viscous forces and the flow becomes chaotic. For a flat
plate, in general, for practical purposes, the ‘critical Reynolds number, Rec’ at which the flow changes
from laminar to turbulent is taken as 5 ´ 10 5. It should be understood clearly that this is not a fixed value
but depends on many parameters including the surface roughness.
(vi) There is intense mixing of fluid particles in turbulent region; therefore, heat transfer is more in turbulent
flow as compared in laminar flow. This is the reason why special efforts are made in the design of heat
exchangers to increase turbulence. However, one has to pay a premium of increased pressure drop i.e.
increased power to pump the fluid through the heat exchanger.
(vii) Velocity profile in the laminar flow is approximately parabolic.
(viii) Turbulent region of boundary layer is preceded by transition region as shown in Fig. 9.2.
(ix) Turbulent boundary layer itself is made of three layers: a very thin layer called ‘laminar sub-layer’, then,
a ‘buffer layer’ and, finally, the ‘turbulent layer’.
(x) Velocity profile in the laminar sub-layer is approximately linear, whereas in the turbulent layer the veloc-
ity profile is somewhat flat, as shown.
(xi) Thickness of the boundary layer, d, increases along the flow direction; as we shall see later, d is related to
the Reynolds number as follows: in the laminar flow region:
5 ×x
d lam = …(9.11)
( Rex )0. 5
and for turbulent flow region:
0.376 ×x
d turb = …(9.12)
( Rex )0. 2
where Rex is the Reynolds number at position x from the leading edge.

FORCED CONVECTION 387


9.6 Thermal Boundary Layer
When the temperature of a fluid flowing on a surface is different from that of the surface, a ‘thermal boundary
layer’ develops on the surface, in a manner similar to the development of the velocity boundary layer. Let us
illustrate the development of the thermal boundary layer with reference to a flat plate. See Fig. 9.3.
Consider a fluid at an uniform velocity of U and a uniform temperature of Ta approach the leading edge of
a thin, flat plate as shown. Let the flat plate be at a uniform temperature of Ts. Let Ta > Ts. Then, the first layer that
comes in contact with the surface will adhere to the surface (no slip condition) and reach thermal equilibrium
with the surface and attain a temperature of Ts. Then, the fluid particles in this layer will exchange energy with
the particles in the adjoining layer, which in turn will exchange energy with the subsequent layer, and so on.
Thus a temperature profile will develop in the flow field and the temperature will vary from Ts at the surface to
Ta at the free stream. In Fig. 9.3, the term (T – Ts) is plotted against y as shown. Thus at the surface, (T – Ts) = 0
and at the free stream condition, (T – Ts) = (Ta – Ts). The region in which the temperature variation in the y-
direction is significant is known as ‘thermal boundary layer’. Thickness of the thermal boundary layer (d t) at any
location is defined as that distance from the plate surface in the y-direction where (T – Ts) = 0.99 ´ (Ta – Ts). i.e
where the temperature difference between the fluid and the surface has reached 99% of the maximum possible
temperature difference of (T a – Ts ). In other words, at the outer edge of the thermal boundary layer, the
dimensionless temperature ratio, (T – Ts)/(Ta – Ts) is equal to 99%.
Thickness of the thermal boundary layer increases with increasing distance along the plate; this is due to the
fact that effects of heat transfer are felt more, further downstream.
If the approaching fluid stream temperature Ta is less than the plate surface temperature, then the tempera-
ture profile in the thermal boundary layer will be as shown below, in Fig. 9.4:

Ta > Ts

Fluid flow T – Ts Ta – Ts
U, Ta Temperature
Ta – Ts
T – Ts profile
d
d

Leading edge Flat plate at Ts Trailing edge

FIGURE 9.3 Development of thermal boundary layer over a flat plate

Ta < Ts

Temperature profile

Ts – T

dt dt
y y
Ta

x x
Ts Ts – Ta

FIGURE 9.4 Thermal boundary layer over a flat plate when Ta < Ts

388 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Temperature of the fluid changes from a maximum at the plate surface to the free stream temperature, as we
proceed from the surface upwards in the y-direction. Vertical distance from the plate surface where the ratio (Ts
– T)/(Ts – Ta) is equal to 99% represents the thickness of the thermal boundary layer.
Velocity profile in the hydrodynamic boundary layer depends on the viscosity of the fluid, whereas tem-
perature profile in the thermal boundary layer depends on the viscosity, specific heat and thermal conductivity
of the fluid, in addition to the velocity.
Relative magnitudes of the thicknesses of the hydrodynamic boundary layer (d ) and thermal boundary layer (d t)
depend on the dimensionless parameter ‘Prandtl number’ defined as:
Pr = (Molecular diffusivity of momentum)/(Molecular diffusivity of heat)
Or,
n m ×Cp
Pr = = …(9.13)
a k
where m is dynamic viscosity, Cp is the specific heat and k is the thermal conductivity of the fluid.
Also, n is kinematic viscosity = m/r, and a is the thermal diffusivity.
Prandtl number is of the order of 1 for gases, less than 0.01 for liquid metals and more than 100,000 for
heavy oils. See Table 9.2.

TABLE 9.2 Range of Prandtl numbers for fluids


Fluid Pr

Liquid metals 0.004 – 0.030


Gases 0.7 – 1.0
Water 1.7 – 13.7
Light organic fluids 5 – 50
Oils 50 – 100,000
Glycerin 2000 – 100,000

Regarding the relative growth of velocity and thermal boundary layers in a fluid, we may note the follow-
ing:
(i) For gases, where Pr = (n/a) is of the order of 1, we see that the momentum and heat dissipate almost at
the same rate i.e. thicknesses of the hydrodynamic and thermal boundary layers are of the same order;
(ii) for liquid metals since Pr << 1, it means that heat diffuses at a much higher rate than the momentum for
liquid metals i.e. the thermal boundary layer is much thicker than hydrodynamic boundary layer for
liquid metals (See Fig 9.5,a), and,
(iii) for heavy oils (Pr >> 1), momentum diffuses at a faster rate than heat through the medium and this is
evident from Fig. (9.5,b); thus, the thermal boundary layer is much thinner than hydrodynamic boundary
layer.
For laminar conditions, thickness of thermal boundary layer is related to hydrodynamic boundary layer,
approximately as follows:

Pr << 1 Pr >> 1

dt d

d dt

(a) Liquid metals (b) Oils


FIGURE 9.5 Thermal and velocity boundary layers over a flat plate for liquid metals and oils

FORCED CONVECTION 389


dt 1
= …(9.14)
d Pr 0. 33
where Pr is the Prandtl number.

9.7 Differential Equations for the Boundary Layer


In convection studies, since there is a fluid flow, we are interested in the shear stress and the friction coefficient;
to determine these we need the velocity gradient at the surface. Similarly, to determine the convection coefficient,
we need the temperature gradient at the surface. To determine the velocity gradient at the surface, we apply the
equation of conservation of momentum (in conjunction with the equation of conservation of mass) to a differen-
tial volume element in the boundary layer. And, to determine the temperature gradient at the surface, we apply
the equation of conservation of energy to a differential volume element in the boundary layer. We start with the
application of equation for conservation of mass:
9.7.1 Conservation of Mass–The Continuity Equation for The Boundary Layer
Consider a differential control volume, of section (dx.dy) and unit depth, within the boundary layer, as shown in
Fig. 9.6.

X
r[v + (¶v/¶y).dy]

Elemental control
volume, (dx.dy.l)

ru dy r[u + (¶u/¶x).dx]
dx

rv

FIGURE 9.6 Elemental control volume in the boundary layer over a flat plate for conservation of mass

Assumptions:
(i) Flow is steady, incompressible
(ii) Constant fluid properties
(iii) Pressure variation is only in the X-direction
(iv) Shear in Y-direction is negligible
(v) Continuity in space and time
Let u and v be the velocity components in the X and Y-directions. Then, remembering that the mass flow
rate is given by (density x velocity x area) and that the depth is unity in the Z-direction, we can write:
Mass flow into the control volume in X-direction = r.u.(dy.1)
Mass flow out of the control volume in X-direction = r.[u + (¶u/¶x).dx].(dy.1)
Therefore, net mass flow into the element in the X-direction = –r.(¶u/¶x).dx.dy
Similarly, net mass flow into the control volume in the Y-direction is = –r.(¶v/¶y).dy.dx
Since the net mass flow into control volume, in steady state, must be equal to zero, we write:
–r.{(¶u/¶x) + (¶v/¶y)}.dx.dy = 0
i.e. for a two-dimensional flow in the boundary layer, equation of conservation of mass is given by:
(¶u/¶x) + (¶v/¶y) = 0 ….(9.15)
Eq. 9.15 is known as ‘continuity equation’ for two-dimensional, steady flow of an incompressible fluid.

390 FUNDAMENTALS OF HEAT AND MASS TRANSFER


9.7.2 Conservation of Momentum Equation for The Boundary Layer
This is obtained by the application of Newton’s second law of motion to the differential element, which states
that the net force on the element in the X-direction is equal to the net momentum efflux from the control volume
in the X-direction. Fig. 9.7 shows the momentum fluxes and forces acting on the differential control volume.

r[v + (¶v/¶y).dy].[u + (¶u/¶y).dy]

(r.u).u r[u + (¶u/¶x).dx].[u + (¶u/¶x).dx]

dx
(ru).v
(a) Momentum fluxes

ty+dy = m.{(¶u/¶y) + (¶/¶y) (¶u/¶y).dy}

p dy [p + (¶p/¶x).dx]
dx
ty = m.(¶u/¶y)

(b) Forces

FIGURE 9.7 Conservation of momentum in a two-dimensional, incompressible boundary layer

For no pressure gradients in the Y-direction and with the assumption that viscous shear in the Y-direction is
negligible,
Momentum flow in X-direction into left face = r.u 2.dy
Momentum flow in X-direction out of right face = r.[u+ (¶u/¶x).dx] 2.dy
= r.u2.dy + 2.r.u. (¶u/¶x).dx.dy
x-momentum flow entering bottom face = r.u.v.dx
x-momentum flow leaving upper face = r.[v + (¶v/¶y).dy].[u + (¶u/¶y).dy].dx
= r.u.v.dx + r.u.(¶v/¶y).dx.dy + r.v.(¶u/¶y).dx.dy
Therefore, net momentum change in the X-direction =
[momentum flux out of the right and top faces] – [momentum flux into the left and bottom faces]
= [r.u2.dy + 2.r.u.(¶u/¶x).dx.dy] + [r.u.v.dx + r.u.(¶v/¶y).dx.dy + r.v.(¶u/¶y).dx.dy]
– r.u2.dy – r.u.v.dx
= 2.r.u.(¶u/¶x).dx.dy + r.u.(¶v/¶y).dx.dy + r.v.(¶u/¶y).dx.dy
= r.{u.(¶u/¶x) + v.(¶u/¶y)}.dx.dy + r.u.{(¶u/¶x) + (¶v/¶y)}.dx.dy
Now, from continuity Eq. 9.15, we have: (¶u/¶x) + (¶v/¶y) = 0; Therefore, net momentum transfer in the X-
direction = r.{u.(¶u/¶x) + v.(¶u/¶y)}.dx.dy …(a)
Now, let us calculate the forces acting on the control volume in the X-direction:
Pressure forces:
Pressure force on the left face is p.(dy.1) and over the right face is
–[p + (¶p/¶x).dx].(dy.1)
Therefore, net pressure force in the direction of motion is: –(¶p/¶x).dx.dy
And,
Viscous shear forces :
Viscous shear force at the bottom face is: m(¶u/¶y).(dx.1)
Viscous shear force at the top face is: [m(¶u/¶y) + m(¶ 2u/¶y2).dy].(dx.1)

FORCED CONVECTION 391


Therefore, net viscous force in the direction of motion =
[m(¶u/¶y) + m(¶ 2u/¶y 2).dy].(dx.1) – m(¶u/¶y).(dx.1) = m(¶ 2u/¶y2).dx.dy
Therefore,
Fx = Resultant applied force in the X-direction =
Net pressure force in the X-direction + net viscous force in the X-direction
i.e. Fx = – (¶p/¶x).dx.dy + m(¶ 2u/¶y2).dx.dy …(b)
Equating Eqs. a and b as per Newton’s second law, and neglecting second order differentials, we get:
r.{u.(¶u/¶x) + v.(¶u/¶y)} = m(¶ 2u/¶y2) – (¶p/¶x). …(9.16)
Eq. 9.16 is known as ‘conservation of momentum equation’ for two-dimensional, steady flow of an incom-
pressible fluid.
If the pressure variation in the X-direction is negligible, (which is true for flow over a flat plate since (¶U/¶x)
= 0), Eq. 9.16 reduces to:
u.(¶u/¶x) + v.(¶ u/¶ y) = n.(¶ 2u/¶y2). …(9.17)
where n = m/r = kinematic viscosity
9.7.3 Conservation of Energy Equation for The Boundary Layer
Assumptions:
(i) steady, incompressible flow
(ii) conduction is only in the Y-direction
(iii) temperature change in the X-direction is small i.e. negligible conduction in flow direction
(iv) specific heat (Cp) of the fluid is constant
(v) negligible viscous heating
(vi) negligible body forces
Fig. 9.8 shows the rate at which energy is conducted and convected into and out of the differential control
volume.
Note that in addition to the conductive terms, there are four convective terms.
Let us write the different energy terms and apply the energy balance which states that net rate of conduction
and convection should be equal to zero:
Convective terms:
For the X-direction:
Energy into the control volume = r.Cp.u.T.dy
Energy out of the control volume = r.Cp.{u + (¶u/¶x).dx}.{T + (¶T/¶x) dx}.dy
Therefore, neglecting the product of differentials, net energy convected into the control volume in the X-
direction is given by: –r. Cp .{u.(¶T/¶x) + T.(¶u/¶x)}.dx.dy
Similarly, net energy convected into the control volume in the Y-direction is given by:
–r.Cp.{v. (¶T/¶y) + T.(¶v/¶y)}.dx.dy
Conductive terms:
Conduction in Y-direction.

2 2
–k.dx.{(¶T/¶y) + (¶ T/¶y ).dy} r.Cp.{v + (¶v/¶y).dy}.{T + (¶T/¶y).dy}.dx

r.Cp.u.T.dy r.Cp.{u + (¶u/¶x).dx}.{T + (¶T/¶x).dx}.dy


dy
2 2
–k.dy.(∂T/∂x) –k.dy.{(¶T/¶x) + (¶ T/¶x ).dx}
dx

–k.dx.(¶T/¶y) r.Cp.v.T.dx

FIGURE 9.8 Conservation of energy in a two-dimensional, incompressible boundary layer

392 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Net conduction into the control volume in the Y-direction is given by:
– k.dx.(¶T/¶y) – [–k.dx.{(¶T/¶y) + (¶ 2T/¶y2).dy} = k. (¶ 2 T/¶y2).dx.dy
Similarly, for completeness, net conduction into the control volume in the X-direction is given by:
k. (¶ 2T/¶x2).dx.dy
When the viscous work is neglected, making an energy balance, we have:
Algebraic sum total of heat flow to the control volume due to conduction and convection must be equal to
zero.
i.e.
–r.Cp.{u.(¶T/¶x) + T.(¶u/¶x)}.dx.dy – r.Cp .{v.(¶T/¶y) + T.(¶v/¶y)}.dx.dy + k.(¶ 2T/¶x 2).dx.dy
+ k. (¶ 2T/¶y2).dx.dy = 0
i.e.
–r.Cp .{u. (¶T/¶x) + T.(¶u/¶x) + v.(¶T/¶y) + T.(¶v/¶y)}.dx.dy +
k.{(¶ 2T/¶x2) + (¶ 2T/¶y2)}.dx.dy = 0
i.e.
– r.Cp.{u. (¶T/¶x) + v.(¶T/¶y) + T.[(¶u/¶x) + (¶v/¶y)]}+ k.{(¶ 2T/¶x2) + (¶ 2T/¶y2)} = 0
Now, from continuity equation, (¶u/¶x) + (¶v/¶y) = 0; also, since the boundary layer is very thin, (¶T/¶y)
>> (¶T/¶x). (i.e. conduction in X-direction is negligible).
Therefore, energy balance equation becomes:
u. (¶T/¶x) + v.(¶T/¶y) = (k/r.Cp). (¶ 2T/¶y2)
or,
k
u.(¶T/¶x) + v. (¶T/¶y) = a.(¶ 2T/¶y2), where a = = thermal diffusivity …(9.18)
r ×Cp
This is the energy equation for a two-dimensional, steady incompressible flow, when the viscous dissipa-
tion is neglected, i.e. for very low velocities of flow.
Observe the similarity between Eq. 9.17 for momentum balance and the Eq. 9.18 for energy balance.
In Eq. 9.17, n = m/r = kinematic viscosity, also known as momentum diffusivity. In Eq. 9.18, a is the diffu-
sivity of heat. Their ratio is known as Prandtl number and is equal to:
Pr = n/a = (m/r)/(k/r.Cp) = Cp .m/k …(9.19)
If n = a, then Pr = 1 and the momentum and energy equations are identical; thus, Prandtl number controls
the relation between the velocity and temperature distributions.
When the viscous dissipation cannot be neglected, as in the case of very viscous fluids (e.g. in journal bear-
ings), or when the fluid shear rate is extremely high, an additional term for ‘viscous dissipation, j ‘ appears on
the LHS of the energy balance. j is given by:
j = m.{[(¶u/¶y) + (¶v/¶x)] 2 + 2.[(¶u/¶x) 2 + (¶v/¶y) 2] – (2/3).[(¶u/¶x) + (¶v/¶y)] 2 …(9.20)
We shall not consider viscous dissipation in our discussions.

9.8 Methods to Determine Convective Heat Transfer Coefficient


As stated earlier, in convection heat transfer analysis, the primary problem is to determine the heat transfer
coefficient. Once this quantity is determined, heat transfer rate from the surface is easily determined by applying
Newton’s law.
There are generally, five methods available to determine the convective heat transfer coefficient:
(i) dimensional analysis in conjunction with experimental data
(ii) exact mathematical solutions of boundary layer equations
(iii) approximate solutions of boundary layer equations by integral methods
(iv) analogy between heat and momentum transfer, and
(v) numerical analysis
Of course, none of them can by itself, solve all the problems we come across in practice, since each method
has its own limitation.
Of the above mentioned methods, ‘dimensional analysis’ is mathematically simple, but has the disadvan-
tage that it does not give any insight into the phenomenon occurring; also, it does not give any equation that can

FORCED CONVECTION 393


be solved, but requires experimental data to get the coefficients in the equations. However, this method helps in
the interpretation of the experimental data and extends the range of applicability by expressing the data in terms
of dimensionless groups.
‘Exact solutions of boundary layer equations’ involve simultaneous solutions of differential equations de-
rived for the boundary layer. These are rather complicated and solutions are available for a few simple flow
situations, such as flow over a flat plate, an airfoil, or a circular cylinder, in laminar flow. Describing the turbu-
lent flow mathematically is rather difficult. We shall only give an outline of this method, since our emphasis is on
practical solutions to convection heat transfer problems by using empirical relations.
‘Approximate solutions for boundary layer equations’ consider a finite control volume for analysis, rather
than an infinitesimal control volume, and integral equations are derived; however, solution requires assuming
equations to describe the velocity and temperature profiles satisfying the boundary conditions. This method is
relatively simple, and it is possible to get solutions to problems that cannot be treated by exact method of analy-
sis. This method can be applied to turbulent flow also.
‘Analogy between heat and momentum transfer’ is a very useful tool to deduce the convective heat transfer
coefficient by the knowledge of flow friction data only, particularly for turbulent flows, without actually conduct-
ing heat transfer experiments. This method utilizes the fact that the momentum and energy equations have the
same form, under certain conditions, and therefore, the solutions also must have the same form. Further, it is
simple to conduct flow (friction) experiments, as compared to heat transfer experiments.
‘Numerical methods’ involves discretizing the differential equations and are therefore approximate. Solu-
tions are obtained at discrete points in time and space rather than continuously; however, accuracy can be im-
proved to acceptable levels by taking sufficiently close grids. Main advantage of numerical methods is that
variation in fluid properties and boundary conditions can be easily handled.
9.8.1 Dimensional Analysis
Dimensional analysis considers the various quantities that contribute to the phenomenon and reduces these vari-
ables into dimensionless groups; however, dimensional analysis alone is not of much use, and this method must
always be supplemented by experimental data since to determine the coefficients in the functional relationships
between the dimensionless groups we need actual, practical data. Also, it is necessary to have some insight into
the problem before we start the analysis, since we have to first list the pertinent variables that influence the
phenomenon. Once this is done, mathematics involved is minimum, and the method can be applied routinely to
most of the problems.
9.8.1.1 Primary dimensions and dimensional formulas. Fundamental axiom of dimensional analysis is that equa-
tions describing a physical phenomenon must be dimensionally homogeneous (i.e. dimensions of the two sides
of the equation are identical) and units therein must be consistent.
‘Dimension’ is a qualitative expression whereas unit is quantitative. For example, when the distance be-
tween two points is spoken of as ‘length’ it is qualitative; instead, if we say that the distance is so many metres or
kilometres or miles, we are speaking in terms of ‘Units’.
In S.I. system, there are four ‘primary dimensions’ viz. Length (L), Mass (M), Time (t) and Temperature (T).
Other derived quantities can be expressed in terms of these primary dimensions. Dimensional formula for a
physical quantity is obtained from its definition or from physical laws involved. For example,
Dimension of length of a bar: [L]
Dimension of velocity: Distance/time: [L/t] = L.t–1
Dimension of Force: Mass x acceleration = [M.L/t2] = [M. L. t –2]
Dimension of Work : Force x distance : [M. L. t– 2].L = [M. L2. t – 2]
Dimension of Power : Work/time : = [M. L2.t– 3], …, etc.
Table 9.3 shows a few physical quantities, their symbols, units and dimensional formulas.
9.8.1.2 Buckingham p theorem. This theorem is used to determine the number of independent dimensionless
groups that can be obtained from a set of physical quantities that govern a given phenomenon.
According to this rule, if the number of pertinent physical variables governing a phenomenon is ‘n’, and the
number of primary dimensions to express the dimensional formulas of these n quantities is ‘m’, then, the number
of independent dimensionless groups that can be formed by combining these physical quantities is given by (n –
m). If these dimensionless groups are designated by p 1, p 2, …, etc. then, relation between them can be expressed
as:
F(p 1, p 2, p 3, …) = 0 …(9.21)

394 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 9.3 Some physical quantities of importance in heat transfer and their dimensions

Quantity Symbol Units Dimensions


Mass M kg M
Length L, d, D, x m L
Time t, t s t
Temperature T, q K T
Area A m2 L2
Volume V m3 L3
Velocity V,U m/s Lt –1
Acceleration a m/s 2 Lt –2
Force F N M Lt–2
Work W Nm(=J) M L2 t–2
Power P W(=J/s) M L2 t–3
Density r kg/m3 ML –3
Pressure, stress p, s N/m2 M L–1t–2
Viscosity m kg/(ms) M L–1t–1
Kinematic viscosity n(= m/r) m2 /s L2 t–1
Specific heat cp J/kgK L2 t–2 T –1
Thermal conductivity k W/(mK) MLt –3 T –1
Thermal diffusivity a m2 /s L2 t –1
Heat transfer coefficient h W/(m2 K) Mt –3 T –1
Coefficient of volume expansion b 1/K T –1

For example, if in a problem, there are 5 physical quantities which are described by 3 primary dimensions only,
then there are (5 – 3 ) = 2 dimensionless groups and the solution is of the form:
F(p 1, p 2) = 0. …(9.22)
Or,
p 1 = f (p 2) …(9.23)
Then experimental data can be presented by plotting p 1 against p 2.
If there are 3 dimensionless groups in another problem, the solution is of the form:
F(p 1, p 2, p 3) = 0. …(9.24)
Or,
p 1 = f(p 2, p 3) …(9.25)
Now, experimental data can be presented by plotting p 1 against p 2 for different values of p 3.
While applying the Buckingham method, after determining the number of p terms that can be formed, further
procedure is as follows: Of the total of ‘n’ number of variables, select a ‘core group’ of ‘m’ number of variables,
which repeat for each p term; these are known as ‘repeated variables’; then, each p term is formed by the core
group plus one of the remaining (n – m) variables. Each of the variables in the core group is raised to a suitable
power to maintain dimensional homogeneity. Selection of the core group should be done as per the following
thumb rules:
(a) variables in the core group must contain among themselves all the fundamental dimensions involved in
the phenomenon.
(b) the repeating variables must not form dimensionless groups among themselves
(c) invariably, dependent variable should not be incuded in the core group
(d) no two variables in the core group should have the same dimensions
(e) in general, repeating variables should be chosen such that one variable contains a geometric property
(e.g. length ‘l’, diameter ‘D’ or height ‘h’), other variable contains a flow property (e.g. velocity ‘V’, accel-
eration ‘a’ etc.), and the other variable contains a fluid property (e.g. density ‘r’, dynamic viscosity ‘m’
etc.). In most of the cases, repeated variables or the core group consist of : (l, V, r), (d, V, r), (l, V, m), or (d,
V, m).
Procedure of applying the Buckingham method is illustrated below:

FORCED CONVECTION 395


9.8.1.3 Dimensional analysis for forced convection. Now let us illustrate the application of Buckigham’s p theorem
to the case of convection heat transfer for a fluid flowing across a heated tube; of course, same approach is
applicable for heat transfer for a fluid flowing inside a tube or flowing over a plate.
First, it is necessary to list the pertinent parameters influencing the physical phenomenon. From the descrip-
tion of the problem, it appears reasonable to assume that the physical quantities listed below (along with their
dimensional formulas) are relevant to this problem:
1. Tube diameter (D)…[L]
2. Fluid density (r)…[ML –3]
3. Fluid velocity (V)…[Lt –1]
4. Fluid viscosity (m)…[ML –1t –1]
5. Specific heat (Cp)…[L2 t –2T –1]
6. Thermal conductivity (k)…[MLt –3 T–1], and
7. Heat transfer coefficient (h)…[Mt –3 T– 1]
Thus, we see that there are 7 pertinent variables affecting the physical phenomenon and they contain 4
fundamental dimensions L, M, t and T.
Then, from Buckingham’s p theorem, we deduce that (7 – 4) = 3 independent dimensionless groups would
be formed to correlate experimental data.
Now, let us form the ‘core group’ of 4 variables, keeping in mind the principles enumerated above. Let us
choose d, V, r, and h for the core group. They contain among themselves all the primary dimensions; they do not
form dimensionless groups among themselves; no two variables have same dimensions; and, one variable(D) is
a geometric property, one variable(V) is a flow property, and r is a fluid property. Then, the different p terms are
obtained by combining the core group with each one of the remaining (7 – 4) properties:
p 1 = ha.rb.Dc.V d. m
p 2 = hm.rn. D p.V q.Cp
p 3 = hw.rx.D y.V z. k
Exponents of terms in p-terms are chosen so as to make the p terms dimensionless. So, we start with p 1 and
write the dimensional formulas of each quantity and apply the requirement of dimensional homogeneity:
For p 1:
p 1 = h a.rb.D c.V d.m
M 0 L0 t0 T 0 = 1 = [Mt –3T –1]a. [ML – 3]b.[L]c. [Lt – 1]d. [ML – 1t–1]
Equating the exponents of M, L, t and T on either side, for dimensional homogeneity:
Exponents of M: 0= a+b+1
Exponents of L: 0 = –3b + c + d – 1
Exponents of t: 0 = –3a – d – 1
Exponents of T: 0 = –a
Solving the above equations, we get:
a = 0; b = –1; c = –1; d = –1
Therefore,
p 1 = r–1.D–1.V – 1.m
i.e. p 1 = m/(r.V.D)
Since p 1 is dimensionless anyway,
we shall choose: p 1 = r.V.D/m
Recognize that p 1 is the dimensionless Reynolds number (Re).
For p 2:
p 2 = hm.rn.D p.V q.Cp
Then, M L t T = 1 = [Mt–3 T–1]m. [ML– 3]n.[L]p. [Lt– 1]q. [L2t–2 T– 1]
0 0 0 0

Equating the exponents of M, L, t and T on either side, for dimensional homogeneity:


Exponents of M: 0= m+n
Exponents of L: 0 = – 3n + p + q + 2
Exponents of t: 0 = –3 m – q – 2
Exponents of T: 0 = –m – 1

396 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Solving the above equations, we get:
m = –1; n = 1; p = 0; q = 1
Therefore,
p 2 = h –1.r.V.Cp
i.e. p 2 = (Cp .r.V)/h
Since the dimensions of h and k/D are same, we write:
p 2 = (Cp .r.V.D)/k
Dividing this by another dimensionless number, i.e. Reynolds number gives again another dimensionless
number; so, we get:
p 2 = {(Cp.r.V.D)/k}/{r.V.D/m} = m.Cp/k
Recognise that p 2 is the dimensionless Prandtl number (Pr).
For p 3 :
p 3 = hw.rx.Dy.V z.k
Then, M L t T = 1 = [Mt–3 T–1]w. [ML– 3]x.[L]y. [Lt –1]z. [MLt –3T –1]
0 0 0 0

Equating the exponents of M, L, t and T on either side, for dimensional homogeneity:


Exponents of M: 0= w+x+1
Exponents of L: 0 = –3 x + y + z + 1
Exponents of t: 0 = –3 w – z – 3
Exponents of T: 0 = –w – 1
Solving the above equations, we get:
w = –1; x = 0; y = –1; z = 0
Therefore,
p 3 = h –1.D– 1.k = k/(h.D)
Since k/(h.D) is dimensionless, (h.D)/k is also dimensionless. So, we choose:
p 3 = (h.D)/k
Recognize that p 3 is the dimensionless Nusselt number (Nu).
Then, according to the Buckingham p theorem,
p 3 = F(p 1, p 2)
Or,
Nu = C. Re m.Prn. ...(9.26)
where C, m and n are constants evaluated experimentally.
Eq. 9.26 is the desired relation among the various physical quantities affecting forced convection across a
tube, expressed in terms of dimensionless numbers Nu, Re and Pr.
Note:
(a) If we had taken (D, r, m, k) for the core group (or, repeating variables), then combining the core group with V,
cp and h in turn, we would have got, respectively:
p 1 = (r V D)/m = Re
p 2 = m Cp/k = Pr, and
p 3 = h D/k = Nu
i.e. the same result as obtained earlier.
(b) If, instead, we choose (V, m, r, Cp) as the core group, then the dimensionless terms obtained are:
Re = (r V D)/m
Pr = mCp/k, and
St = h/(r V Cp) = h/(G Cp) = Stanton number,
where G = r.V = mass velocity
In fact, another way of expressing heat transfer correlations is:
St = F(Re, Pr). …(9.27)
9.8.1.4 Advantages and limitations of dimensional analysis
Advantages:
(i) It is mathematically quite simple.

FORCED CONVECTION 397


(ii) When a given physical phenomenon depends on a large number of variables, dimensional analysis re-
duces the number of variables for experimentation by getting the dimensionless numbers with suitable
combination of those variables. Advantage of having lesser number of variables for experimentation is
obvious.
(iii) Dimensional analysis helps in interpretation of experimental data and in deriving suitable empirical,
design equations.
(iv) It also helps in planning the experimental work for a particular problem.
(v) It helps to extend the range of experimental results; for example, if a particular set of results for air in
forced convection is expressed in terms of Nusselts number, Reynolds number and Prandtl numbers,
then the same results can be applied to another fluid, say, water, if the corresponding dimensionless
numbers are the same.
(vi) It helps in getting a partial solution to problems, when the mathematical solution is too complicated.
Limitations:
(i) It does not give any insight into the physical phenomenon occurring.
(ii) Selection of variables has to be done with care; if it is wrongly done, results will be erroneous.
(iii) It does not give an exact functional relation which can be solved; dimensional analysis requires experi-
mental data to get the coefficients in the functional relationship.
(iv) If it is required to get the effect of one particular variable on the rest of the variables in a particular
problem, it is difficult to get this information by dimensional analysis.
Application of dimensional analysis to the case of heat transfer by natural convection will be described in
the next chapter.
9.8.1.5 Dimensionless numbers and their physical significance. There are many dimensionless numbers that we
come across in heat transfer studies. Their physical significance must be clearly understood and this is is facili-
tated by expressing these dimensionless numbers as the ratios of two forces. This requires a little explanation:
Many times, in fluid mechanics and heat transfer studies, it becomes impossible or impracticable to conduct
experiments on the actual prototype size of the system. Then, studies are done on a model of reduced size. Then,
the question arises as to how to relate the results of the experiments done on the model to the actual prototype.
To be able to do so, certain criteria have to be satisfied. These criteria, known as ‘criteria for similitude’ are the
following:
(a) Geometrical similarity Two objects are geometrically similar if the ratios of corresponding linear dimen-
sions are equal.
(b) Kinematic similarity This represents similarity of motion , i.e. if the ratios of velocities of corresponding
particles are equal, there is said to be kinematic similarity.
(c) Dynamic similarity This represents similarity of forces. If there is kinematic similarity and in addition,
the ratios of homologous forces in the systems are also the same, there is said to be dynamic similarity.
If all the above criteria are satisfied, then there is complete correspondence or similarity between the model
and the prototype.
Further, in an incompressible flow, if the conditions of geometrical similarity and dynamic similarity are
satisfied, then kinematic similarity is automatically achieved.
Geometric similarity can be easily achieved by constructing the model of the actual system to a certain
reduced scale. One way of ensuring dynamic similarity is by making sure that some relevant dimensionless
numbers are the same for both the model and the prototype, since these dimensionless numbers can be expressed
as ratios of certain forces. Let us illustrate this by considering different forces that are relevant to fluid mechanics
and heat transfer:
(1) Inertia force (Fi):
Fi = mass x acceleration , i.e.

dV r× L3 ×V
F i = r×L3 × = = r ×L2 ×V 2 …(a)
dt L
V
(2) Viscous force (Fv):
Fv = shear stress x area, i.e.

398 FUNDAMENTALS OF HEAT AND MASS TRANSFER


dV 2
Fv = t×L2 = m× ×L = m×V ×L …(b)
dy
(3) Gravity force (Fg):
Fg = mass x gravitational acceleration, i.e.
Fg = r×L3 ×g …(c)
(4) Surface tension (F t):
F t = s ×L …(d)
where s is the coefficient of surface tension (units: Force/unit length)
(5) Elasticity force (Fe):
Fe = Ev ×L2 …(e)
where Ev is the bulk modulus of elasticity of the fluid.
(6) Pressure force (Fp):
Fp = pressure x area, i.e.
Fp = Dp×L2 ...(f)
Now, let us form the ratios of inertia force with other forces:
Ratios of forces:

Inertia force r ×V 2 ×L2 r ×V × L


(a) = = = Reynolds Number = Re
Viscous force m × V ×L m

r ×V 2 × L2 V2 F I 2

(b)
Inertia force
Gravity force
=
r × g × L3
=
g×L
= GH V
g×L
JK = (Fr)2

Here, Fr is known as ‘Froude number’.

Inertia force r ×V 2 ×L2 r ×V 2 × L


(c) = = = Weber Number = Wn
Surface Tension force s ×L s

Inertia force r ×V 2 × L2 V2 V2
(d) = = = 2 = (Ma)2
Elasticity force Ev × L 2 Ev Vs
r

Ev
where Vs = = Sonic velocity and,
r
Ma = Mach number

Inertia force r ×V 2 × L2 r ×V 2
(e) = 2
= = Euler Number = En
Pressure force D P ×L DP
Dimensionless numbers mentioned above occur frequently in fluid mechanics.
Some of the dimensionless numbers occurring in heat transfer are:
Reynolds number:
We have:

r ×V × L r ×V 2 ×L2
Re = =
m m × V ×L
i.e. Re = Inertia force/Viscous force
i.e. Reynolds number is a measure of relative magnitudes of inertial and viscous forces occurring in a given flow
situation. At low velocities, Reynolds number is low, i.e. viscous effects are large and any flow disturbances are
easily damped by viscous effects and the different layers in the flow move systematically, parallel to each other;
this is called laminar flow. If, on the other hand, the Reynolds number is large, effect of inertial forces are pre-
dominant and the flow pattern is completely random, with the chunks of particles moving in all directions; this
is called turbulent flow. Thus, Reynolds number denotes the type of flow i.e. if the flow is laminar or turbulent.

FORCED CONVECTION 399


Prandtl number:
We have:
m ×Cp r ×n ×Cp n n kinematic viscosity
Pr =
k
=
k
=
F k I =
a
=
thermal diffusivity
GH r ×C JK
p

i.e. Prandtl number is the ratio of kinematic viscosity of the fluid to its thermal diffusivity. n represents the
diffusion of momentum through the fluid whereas a represents the diffusion of heat (energy) through the fluid.
Therefore Pr is a measure of relative effectiveness of momentum and energy transport in the medium by diffu-
sion. For oils Pr >> 1, and this signifies that in oils, momentum transport is more rapid than the transport of
energy; for gases, Pr ~ 1 and this means that in gases, momentum and energy are transported by diffusion at
almost the same rate. For the case of liquid metals, where Pr << 1, the energy transport is many times more rapid
as compared to the transport of momentum.
Prandtl number is also the significant parameter which influences the relative growth of velocity and tem-
perature profiles. Hydrodynamic and thermal boundary layer thicknesses are related by:
d
= Prn
dt
where ‘n’ is a positive exponent.
For gases (Pr ~ 1), d ~ d t, for oils (Pr >> 1), d >> d t, and for liquid metals (Pr <<1), d << d t.
Nusselt number:
We have:
h×L
Nu =
k
Consider a plate at a temperature Ts, over which a fluid at a temperature Ta is flowing; then, immediately
adjacent to the surface there will be a stationary layer of fluid. In this layer, heat transfer is, obviously by conduc-
tion and then the heat is transferred to the stream by convection. Making an energy balance and equating these
two quantities,
F dT I
Q = –k×A× GH dy JK y=0
= h×A×(Ts – Ta)

F dT I
- k ×G J
i.e. h=
H dy K y=0
Ts - Ta

F dT I
h×L
- GH dy JK y=0
i.e. =
k (Ts - Ta )
L
i.e. Nusselt number may be interpreted as a ratio of temperature gradient at the surface to an overall, reference
temperature gradient.
Looking at it in another way, multiplying both the numerator and denominator of the expression for Nu by
DT, we can write:
h×L h× DT convective heat flux
Nu = = =
k D T conductive heat flux

L
i.e. Nusselts number is an indication of the enhancement of heat transfer by convection.

400 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Stanton number:
We have:
h h
St = =
G × Cp r ×V × Cp
This can be written as:
h×L
h h k Nu
St =
G × Cp
=
r ×V × Cp
=
FG IJ FG
r × V ×L m × Cp I
JK
=
Re × Pr
H m KH×
k
Stanton number is expressed in terms of other three dimensionless numbers, namely Nusselts number,
Reynolds number and Prandtl number. Note that Stanton number comes into picture only in connection with the
forced convection heat transfer, since the term for velocity (V) is contained in the expression for Nu.
In another interpretation, if the temperature difference between the wall surface and the bulk of the fluid is
DT,
convective heat flux = h. DT; and
energy transported by the bulk fluid flow per unit cross-section of flow area =
mass flow rate x Cp x DT = (V.r).Cp.DT
Therefore, taking their ratio:
h× DT h
= = Nu
(V × r ) × Cp × D T V × r × Cp
In other words, Nusselt number may also be interpreted as the ratio of convective heat flux to the rate of
energy transport by the bulk flow of the fluid per unit area of flow cross-section.
Peclet number:
We have:
r ×V ×L ×C p r ×V × L Cp × m
Pe = = × = Re×Pr
k m k
i.e. Peclet number may be expressed as the product of Reynolds number and Prandtl numbers. Again, as we have
shown above, energy transported by the bulk fluid flow per unit cross-section of flow area =
mass flow rate x Cp x DT = (V.r).Cp.DT, and
heat flux due to conduction across a distance L for the same DT = kDT/L. Taking their ratio:
(V × r ) × Cp × D T r ×V ×L ×C p
FG k × DT IJ =
k
= Pe

H L K
i.e. Peclet number may be interpreted as the ratio of rate of heat transfer by bulk flow to the rate of heat transfer
by conduction.
Graetz number:
This dimensionless number is related to the heat transfer to a fluid flowing through a circular pipe. By definition,
it is the ratio of heat capacity of the fluid flowing per unit length of the pipe to the thermal conductivity of the
pipe i.e.
F m×C I
GH L JK p p × D2
4
× r ×V × Cp
p r × V × D m × Cp D p D
Gz = = = × × × = × Re × Pr ×
k k ×L 4 m k L 4 L
where D is the diameter and L is the length of pipe. Therefore, Graetz number is similar to Peclet number, but is
used in connection with heat transfer analysis of laminar flow in pipes.

FORCED CONVECTION 401


Grashoff number:
Grashoff number occurs only in connection with heat transfer in natural convection (we shall study this later).
We have, by definition:

L3 × r 2 × b × g × D T
Gr =
m2
This can be written as:

Gr =
L3 × r 2 × b × g × D T
= (r×L3 ×b×g×DT)×
r
= (r ×L 3 ×b×g×DT)×
LM r ×V ×L OP
2 2

m 2
m 2
MN (m ×V ×L ) PQ
2

In other words,
Inertia force
G 1 = Buoyant force×
(Viscous force)2
Role of Grashoff number in natural (free) convection is similar to that of Reynolds number in forced convec-
tion.
9.8.2 Exact Solutions of Boundary Layer Equations
Here, we shall illustrate the method in connection with the heat transfer for flow on a flat plate. However, we
shall only give an outline of the method, since, as we stated earlier, our focus is to enumerate the empirical
relations useful for practical calculations.
Recollect that the equations of continuity, momentum and energy for the boundary layer on a flat plate are
given, respectively, by:
(¶u/¶x) + (¶v/¶y) = 0 …(9.15)
u.(¶u/¶x) + v.(¶u/¶y) = n.(¶ 2u/¶y2) …(9.17)
u.(¶T/¶x) + v.(¶T/¶y) = a.(¶ 2T/¶y2) …(9.18)
Now, solving the momentum equation in conjunction with the continuity equation gives the velocity distri-
bution, boundary layer thickness and shear stress (or friction force) at the surface. Exact mathematical solution is
rather complex; its outline is given below:
Since the velocity profiles at different distances from the leading edge of the plate are similar, they can be
considered to differ from each other only by a ‘stretching factor’ in the y-direction. So, the dimensionless velocity
u/U can be expressed at any location x as a function of dimensionless distance y/d from the wall.
Define:
U
h = y× = stretching factor
n ×x
Also, a stream function y (x, y) is defined such that it satisfies the continuity equation, and letting
y= n × x ×U × f (h) where u = ¶ y/¶ y; v = ¶ y/¶ x
Substituting for the terms in the momentum equation in terms of h gives an ordinary, nonlinear, third order
differential equation:

d 2 f (h) d 3 f (h)
f(h)× + 2×
=0
dh 2 dh 3
Solution of this differential equation was obtained numerically, by Blasius. The result is shown in Fig. 9.9.
In Fig. 9.9, abscissa is a dimensionless distance (y/x).Rex0.5 and the ordinate is a dimensionless velocity (u/
U), where u is the local velocity in the x-direction and U is the free stream velocity.
Two important observations are to be made from Fig. 9.9:
(a) first, when the x-coordinate reaches a value of 5 , the y-coordinate is 0.99 i.e. the local velocity reaches 99% of
the stream velocity value when (y/x).Rex0.5 reaches a value of 5. However, from the definition of the boundary
layer thickness d, we know that y = d when u/U = 99%. Therefore, we can write:

402 FUNDAMENTALS OF HEAT AND MASS TRANSFER


u/U
0. 5
At (u/U) = 0.99, (y/x).Rex = 5.0

1.0 Slope = 0.332

0.5

0. 5
(y/x).Rex
0 1.0 2.0 3.0 4.0 5.0 6.0

FIGURE 9.9 Velocity ratio in laminar boundary layer, as per Blasius

5×x
d= …(9.28)
Rex

r ×U × x
where Rex = , local value of Reynolds number.
m
(b) second observation is that the slope at y = 0 is 0.332, i.e.

LM d FG u IJ OP
MM F y H U K I PP = 0.332
MN d GH x × Re JK PQ
x
y=0

We get:
F du I U
GH dy JK y=0
= 0.332×
x
× Rex …(9.29)

Then, the wall shear stress, t is given by:


F du I U
t = m× GH dy JK y=0
= 0.332×m ×
x
× Rex …(9.30)

And, the friction coefficient (or drag coefficient), is by definition:


t 0.664
Cfx =
F r×U I 2
=
Rex
…(9.31)

GH 2 JK
This is the local value of friction coefficient.
Average value of friction coefficient (C fa) over a plate length of L is obtained by integrating Eq. 9.31 between
x = 0 and x = L. i.e.

i.e.
Cfa =
1 L
L 0
Cfa = 2×CfL
z
× C fx dx =
1. 328
ReL
…(9.32)

…(9.33)
Thus, for laminar flow over a flat plate, average friction coefficient is twice the value of local friction coeffi-
cient at x = L.
Solution of the energy Eq. 9.18 gives the value of convective heat transfer coefficient.

FORCED CONVECTION 403


Observe the similarity between the equation of momentum 9.17 and equation of energy 9.18. This fact led
Pohlhausen to follow Blasius assumption of a similarity parameter and stream function as follows:
U
h = y× = similarity parameter
n ×x

y= n × x ×U × f (h)
and, the following ordinary differential equation is obtained:
d2 q Pr dq T - Ts
+ ×f× =
dh 2 2 dh Ta - Ts
T - Ts
where q=
Ta - Ts
Observe that now the ratio, (n/a), i.e. Prandtl number, enters the solution. If we draw a graph of excess
temperature ratio (T – Ts)/(Ta – Ts) against (y/x).Rex0.5 we get different curves for different Prandtl numbers;
however if the excess temperature ratio is plotted against (y/x).Re x0.5Pr0..333, we get a single curve for all Prandtl
numbers and the plot is similar to that in Fig. 9.9. This plot is shown in Fig. 9.10.

q = (T – Ts)/(Ta – TS) 0. 5 333


At q = 0.99, (y/x).Rex .Pr = 5.0

1.0 Slope = 0.332

0.5

0. 5 0.333
(y/x).Rex .Pr
0 1.0 2.0 3.0 4.0 5.0 6.0

FIGURE 9.10 Dimensionless temperature ratio in laminar boundary layer, for flow over a flat plate

Again, there are two important observations to be made from Fig. 9.10:
(a) first, when the X-coordinate reaches a value of 5 , the Y-coordinate is 0.99 i.e. the local excess temperature
reaches 99% of the value of total temperature difference between the free stream temperature and the plate sur-
face temperature, when (y/x).Rex0.5.Pr0.333 reaches a value of 5. However, from the definition of the thermal
boundary layer thickness d t , we know that y = dt when (T – Ts)/(Ta – Ts) = 99%. Therefore, we can write:
5×x
dt = …(9.34)
Rex × Pr 0 . 333
Therefore, immediately, using Eq. 9.28, we can write for the relationship between the thicknesses of hydro-
dynamic and thermal boundary layers:
d
= Pr 0.333 …(9.35)
dt
(b) second observation is that the slope at y = 0 is 0.332, i.e.

LM d F T - T I OP
MM GH T - T JK PP
s
a s

MM d FGH yx × Re × Pr IJK PP
= 0.332
0 . 333
x
N Q y=0

404 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore,
F dT I (Ta - Ts )
GH dy JK y=0
= 0.332×
x
× Re x × Pr 0. 333 …(9.36)

Then, local heat transfer flux (considering the stationary layer):


F dT I (Ta - Ts )
q = –k× GH dy JK y=0
= – k × 0.332×
x
× Re x × Pr 0. 333 …(9.37)

Then, we have for convective heat transfer coefficient:

F dT I
- k× GH dy JK y=0
h=
Ts - Ta
and, we can write, using Eq. 9.37:
h×x
= Nux = 0.332× Re x × Pr 0 . 333 …(9.38)
k
And the local heat transfer coefficient is:
q k 0 . 333
hx = = 0.332× × Rex × Pr …(9.39)
Ts - Ta x
Average value of heat transfer coefficient is obtained by integrating Eq. 9.39 between x = 0 and x = L. We get:
ha = 2×hx = L …(9.40)
i.e. average value of heat transfer coefficient is twice the local value at x = L.
And, then, average Nusselt number is given by:
Nua = 0.664 Re L × Pr 0 . 333 …(9.41)
Eq. 9.41 is valid for Pr ³ 0.6.
In the above equations, properties of the fluid are evaluated at the mean temperature between the free
stream temperature and the plate surface temperature i.e. at the ‘film temperature’ given by:
Ts + Ta
Tf = …(9.42)
2
Eq. 9.41 is not valid for liquid metals (Pr << 1); for liquid metals, following correlation is suggested by Kays:
Nux = 0.565 ×Pe x0.5 …(Pr <.005)…(9.43)
where Pex = Rex × Pr = Peclet number
Example 9.2. Dry air at atmospheric pressure and 20°C is flowing with a velocity of 3 m/s along the length of a long, flat
plate, 0.3 m wide, maintained at 100°C.
(a) Calculate the following quantities at x = 0.3 m:
(i) boundary layer thickness (ii) local friction coefficient (iii) average friction coefficient (iv) local shear stress due to
friction (v) thickness of thermal boundary layer (vi) local convection heat transfer coefficient (vii) average heat transfer
coefficient (viii) rate of heat transfer from the plate between x = 0 and x = x, by convection, and (ix) total drag force on
the plate between x = 0 and x = 0.3 m.
(b) Also, find out the value of xc.(i.e. the distance along the length at which the flow turns turbulent, Rec = 5 ´ 10 5).
Solution.
Data:
100 + 20
W := 0.3 m Ts := 100°C Ta := 20°C U := 3 m/s Tf :=
= 60°C
2
Properties of air are to be taken at the film temperature of 60°C. We get, from the data tables:
r := 1.025 kg/m3 Cp := 1017 J/(kgK) m := 19.907×10 –6 kg/(ms) k := 0.0279 W/(mK) Pr := 0.71
(a) At x = 0.3 m:
x := 0.3 m

FORCED CONVECTION 405


r ×U × x
Rex := (local Reynolds number at x = 0.3 m)
m
i.e. Rex := 4.634 ´ 10 4
Note that Re x is less than 5 ´ 105. Therefore boundary layer is laminar and the equations derived above are applica-
ble.
(i) Boundary layer thickness, d:
5× x
We have: d := ...(9.28)
Re x
i.e. d = 6.968 ´ 10 –3 m (thickness of boundary layer)
(ii) Local friction coefficient C fx :
0 .664
We have: Cfx := ...(9.31)
Re x
i.e. Cfx = 3.085 ´ 10 –3 (Local friction coefficient.)
(iii) Average friction coefficient Ca :
1.328
We have: Cfa :=
Re x
i.e. Cfa = 6.16904 ´ 10 – 3 (Average friction coefficient.)
Or: from Eq. 9.33
Cfa := 2×Cfx i.e. Cfa = 6.16904 ´ 10–3
(iv) Local shearing stress, t:
We have
U
t := 0.332×m× × Re x …(9.30)
x
2
i.e. t = 0.014 N/m (Local shearing stress.)
(v) Thickness of thermal boundary layer:
We have:
d
= Pr 0 . 333 …(9.35)
dt
d
i.e. d t :=
Pr 0 . 333
i.e. d t = 7.81 ´ 10 –3 m (thickness of thermal boundary layer.)
(vi) Local convection heat transfer coefficient:
We have
k
hx := 0.332×
× Re x × Pr 0 . 333 …(9.39)
x
i.e. hx = 5.93 W/(m2c) (local heat transfer coefficient.)
(vii) Average heat transfer coefficient:
From Eq. 9.40, average heat transfer coefficient between x = 0 and x = x is equal to twice the value of local heat
transfer coefficient at x = x
i.e. h a := 2× hx
i.e. h a = 11.86 W/(m2C) (average heat transfer coefficient.)
(viii) Rate of heat transfer from the plate between x = 0 and x = x, by convection:
Area := W ´ 0.3 m2 (area of heat transfer)
i.e. Area = 0.09 m2
Q := ha ×Area×(Ts – Ta)
i.e. Q = 85.395 W (heat transfer rate from the plate between x = 0 and x = 0.3 m.)
(ix) Total drag force on the plate between x = 0 and x = 0.3 m
FD := t ×Area, N (drag force)
i.e. FD = 1.28 ´ 10 –3 N.

406 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(b) Distance at which flow turns turbulent:
r × U × xc
We have: Rec = = 5 ´ 10 5 (critical Reynolds number)
m

5 × 105 × m
Therefore, xc :=
r ×U
i.e. xc = 3.237 m (distance at which flow becomes turbulent.)
Example 9.3. Dry air at atmospheric pressure and 20°C is flowing with a velocity of 3 m/s along the length of a flat
plate, (size: 0.5 m ´ 0.25 m), maintained at 100°C. Using Blasius exact solution, calculate the heat transfer rate from: (i)
the first half of the plate (ii) full plate, and (iii) next half of plate.
Solution.
Data:
100 + 20
L := 0.5 m W := 0.25 m Ts := 100°C Ta := 20°C U := 3 m/s Tf := = 60°C
2
Properties of air are to be taken at the film temperature of 60°C. We get, from data tables:
r := 1.025 kg/m 3 Cp := 1017 J/(kgk) m := 19.907 ´ 10 –6 kg/(m×s) k := 0.0279 W/(mK) Pr := 0.71
(i) Heat transfer rate from the first half of plate:
Now, characteristic dimension to calculate Reynolds number is half the length of plate:
x := 0.25 m
r ×U × x
Therefore, Rex := (local Reynolds no. at x = 0.5/2 = 0.25 m)
m
i.e. Rex = 3.862 ´ 10 4
5
This value is less than 5 ´ 10 ; so, the boundary layer is laminar and the equations derived above are applicable:
We have:
k
hx := 0.332× × Re x × Pr 0 . 333 ...(9.39)
x
i.e. hx = 6.496 W/(m2 C) (local heat transfer coefficient)
Therefore average heat transfer coefficient between x = 0 and x = 0.25 m:
ha := 2× hx
i.e. ha = 12.992 W/(m2C) (average heat transfer coefficient)
Area := 0.25 ×0.25 m 2 (area of half of plate)
i.e. Area = 0.0625 m2 (area of half of plate)
Therefore, heat transferred from first half of plate:
Q 1 := ha × Area ×(Ts – Ta) W
i.e. Q 1 = 64.962 W
(ii) Heat transfer rate from the entire plate:
For the full plate x = L = 0.5 m
L := 0.5 m
r ×U × L
Therefore, Re L := (local Reynolds no. at L = 0.5 m)
m
4
i.e. ReL = 7.723 ´ 10
This value is less than 5 ´ 105; so, the boundary layer is laminar and the equations derived above are applicable:
We have:
k
hL := 0.332×
× Re L × Pr 0 . 333 …(9.39)
L
i.e. hL = 4.594 W/(m2C) (local heat transfer coefficient)
Therefore average heat transfer coefficient between x = 0 and x = 0.5 m:
h a := 2× h L
i.e. ha = 9.187 W/(m2 C) (average heat transfer coefficient)
Area := 0.5×0.25 m2 (area of full of plate)
i.e. Area = 0.125 m2 (area of half of plate)

FORCED CONVECTION 407


Therefore, heat transferred from entire plate:
Q 2 := ha × Area×(Ts – Ta) W
i.e. Q 2 = 91.87 W
(iii) Heat transfer rate from next half of plate:
This is equal to heat transferred from the entire plate minus the heat transferred from the first half of plate = Q 2 – Q 1
i.e. Q 2 – Q 1 = 26.908 W (heat transferred from next half of plate)

9.8.3 Approximate Solutions of Boundary Layer Equations–Von Karman Integral


Equations
It may be observed that Blasius solution to the momentum equation, though exact, is quite cumbersome even for
the simple case of a flat plate; further, much ingenuity is required in selecting a suitable similarity parameter h
y for the solution. In the approximate method of Von
Karman, instead of developing the differential equa-
tions starting from an infinitesimal control volume, a
finite control volume is selected and integral equations
x are developed; this may be done either by directly inte-
Finite control grating the momentum (or energy) equation or by writ-
volume of interest ing a momentum (or energy) balance for the finite
B C control volume. This latter approach is shown below.
Velocity profile Consider a finite control volume, A-B-C-D, that
U
extends from the wall surface in the Y-direction well
u into the free stream (i.e. beyond the boundary layer),
d has a thickness of dx in the X-direction and has unit
width in the Z-direction, as shown. Let the height of
A D AB be H (> d). Now, let us write the momentum bal-
dx

z
ance:
H
FIGURE 9.11 Finite control volume in the boundary
Mass flow rate entering face AB = r × u dy
layer over a flat plate, for integral approach

z z
0
H
r× u dy +
d FG H IJ
Mass flow rate leaving face CD =
0 dx H 0
r × u dy × dx
K
Since no mass can enter the control volume from face AD, it is clear from the mass balance that the incre-
mental mass, i.e.
d
dx
FG
H z
0
H
r × udy × dx
IJ
K
must have entered the control volume through face BC, with the free stream velocity U.
The x-momentum fluxes are:

z
Influx through face AB:
H
r× u2 dy
0
efflux through face CD =

influx through BC =
z0
H
r× u2 dy +
d
dx
F
GH z0
H I
JK
r × u 2 dy × dx


d
dx
FG
H z
0
H IJ
K
r × udy × dx

Assuming that there are no pressure forces (i.e. pressure gradient in X-direction is zero) and no body forces,
and also that there is no shear force on the upper face BC since it is outside the boundary layer, we write the
momentum balance:
Drag or shear force at the plate surface = net momentum change for the control volume.

408 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e.
t w ×dx = U×
d
dx
F
GH z 0
H I
JK
r × udy × dx –
d
dx
F
GH z 0
H I
JK
r × u 2 dy × dx

GH z GH z
d F d I d F d I
t w ×dx = U×
dx 0
JK
r × udy × dx –
dx 0
JK
r × u2 dy × dx

Note that upper limit of integration is replaced by d since the integrand is zero for y > d, i.e. outside the
boundary layer.
Simplifying, we get:

tw =
d
dx
LM
N z
0
d OP
r × (U - u) × udy
Q
i.e. tw = r×U ×
dx N H
2
z
M G1 - Uu IJK × Uu dy OPQ
d L F
0
d

Eq. 9.44 is known as Von Karman integral momentum equation for the boundary layer. It expresses wall
...(9.44)

shear stress tw as a function of non-dimensional velocity distribution (u/U). It is clear from Eq. 9.44 that if we
know the velocity distribution in the boundary layer, we can calculate the wall shear stress easily.
Now, method of solution is to assume a velocity distribution in the boundary layer to start with. At first
sight, this looks ridiculous to assume the velocity distribution, but since the boundary layer is very thin, assum-
ing a velocity profile which satisfies the boundary conditions does not introduce much error. This is verified
from practical results and also, as shown in Table 9.4, assumption of different velocity profiles does not give
much variation in calculated values of boundary layer thickness d or the friction coefficient Cf.
Since, from the experiments, it is observed that velocity distributions in the boundary layer at different x-
locations are geometrically similar, we can say that the dimensionless velocity distribution (u/U) is a function of
dimensionless distance from the wall (y/d). So, let us assume a velocity profile as follows:

u y FG IJ = a + b× FG y IJ + c× FG y IJ 2
FG y IJ 3

U
= f
d H K Hd K Hd K + d×
Hd K …(9.45)

Eq. 9.45 must satisfy the following boundary conditions:


At y = 0: u = 0 and

d2 u
=0
dy 2
At y = d: u = U and
du
=0
dy
Applying these boundary conditions, we get the constants a, b, c and d in Eq. 9.45 and the velocity profile
becomes:

u 3 y FG IJ – 1 ×FG y IJ 3

U
= ×
2 d H K 2 Hd K …(9.46)

Let us now introduce this cubic velocity profile into the Von Karman momentum integral Eq. 9.44. Simplify-
ing, we get:
39 dd
tw = × r ×U 2 × …(9.47)
280 dx
At the solid surface, Newton’s law of viscosity gives:

F du I LM d L 3 F y I 1 F y I 3 OP OP
tw = m× GH dy JK y=0
= m×
MN dy MMN 2 ×U ×GH d JK - 2 ×U ×GH d JK PQ PQ y=0

FORCED CONVECTION 409


3 m ×U
i.e. tw = × …(9.48)
2 d
Equating Eqs. 9.47 and 9.48, we get:
39 dd 3 m ×U
× r ×U 2 × = ×
280 dx 2 d
140 m
i.e. d×dd = × ×dx
13 r ×U
Since d is a function of x only, we integrate the above equation and applying the condition that at x = 0, d =
0, we get:
d2 140 m × x
= ×
2 13 r ×U
Or, in non-dimensional form, this may be written as:
d 140. 2 m 4.64
= × = …(9.49)
x 13 x × r ×U Rex
where Rex is the Reynolds number with characteristic distance x from the leading edge of the plate.
Eq. 9.49 gives the boundary layer thickness d, at a distance x from the leading edge.
To calculate the shear stress at the wall, let us insert this value of d in the expression 9.48 for tw :
3 m ×U 3 m ×U
tw = × = ×
2 d 2 4.64 × x
Rex

r ×U 2 0.646
i.e tw = ×
2 Rex
Now, from the definition of local skin friction coefficient, we have:
tw 0.646
Cfx = = …(9.50)
1 2 Rex
× r ×U
2
Average skin friction coefficient is given by:

Cfa =
1
L
×
z
0
L
C fx dx =
1
L
×
z
0
L 0.646
r ×U
m
× x
dx

m 1. 292
i.e. Cfa = 1.292× = …(9.51)
L × r ×U ReL
where ReL is the Reynolds number based on length L of the plate.
Note that values of boundary layer thickness and skin friction coefficient obtained above with the approxi-
mate, integral method, match reasonably well with the values obtained by exact analysis of Blasius.
Further, if we assume a velocity profile other than the cubic velocity profile assumed above (satisfying the
boundary conditions), it is observed that the results obtained do not differ greatly. Table 9.4 demonstrates this
fact for some velocity profiles, including linear, parabolic and cubic. Blasius exact results are shown for compari-
son:
Note that above results are valid for laminar boundary layer conditions only.
Mass flow through the boundary:
If we consider a section at any distance x from the leading edge, mass flow through that section is given by:
m x = ò [Area ´ Velocity ´ density]; integration is performed within the limits 0 to d.

410 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Table 9.4 Boundary layer thickness ( d ) and skin friction coefficient (Cfa) for different velocity profiles
d
Velocity profile Boundary conditions × Re x C f a × ReL
x
At y = 0 At y = d

u y
1. = u= 0 u= U 3.46 1.155
U d

u y FG y IJ 2
2.
U
= 2×
d

Hd K u= 0 u= U 5.47 1.462

du
=0
dy

u 3 y 1 y
3
FG IJ
3.
U
= ×
2 d
– ×
2 d H K u= 0 u= U 4.64 1.292

d 2u du
= 0 = 0
2 dy
dy

u FG p × y IJ
4.
U
= sin
H2 dK u= 0 u= U 4.78 1.310

5. Blasius exact solution 5.0 1.328

z
i.e.
d
mx = r × udy
0
Assuming the cubic velocity profile as done earlier, substituting for u, we get:

mx =
z0
LM L 3 F y I 1 F y I
MN MMN 2 ×GH d JK - 2 × GH d JK
d
r× U ×
3 OPOP dy
PQPQ
mx
L 3 y - 1 ´ y OP
= r×U× M ´
2 4 d

MN 4 d 8 d PQ 3
0
i.e.
5
mx = × r ×U ×d …(9.52)
8
Mass entrained between two sections at x1 and x 2 can be calculated using Eq. 9.52 as:
5
dm = × r ×U × (d 2 – d 1) …(9.53)
8
where d 1 and d 2 are the thicknesses of boundary layer at sections x 1 and x 2 respectively.
Integral energy equation:
Von Karman integral technique may be applied to get an approximate solution for the energy equation of the
boundary layer, as shown below:
Consider a finite control volume that encloses both the hydrodynamic and thermal boundary layers,
(laminar and incompressible) as shown in Fig. 9.12. Assume that the fluid properties do not vary with tempera-
ture and are constant; let the heating of the plate commence at a distance x 0 from the leading edge of the plate.
That means that thermal boundary layer develops only beyond x 0 from the leading edge.
Le us make an energy balance on the control volume ABCD.

FORCED CONVECTION 411


Velocity boundary layer
Finite control
volume of interest Thermal boundary layer
B C
U, Ta

d
dt

A D
x0 dx

FIGURE 9.12 Finite control volume in the boundary layer over a flat plate, for integral energy equation

Energy enters the control volume by convection at face AB, leaves by convection at face CD; also, energy
enters the control volume by conduction through face AD and by convection through face BC. Let us write the

z
various terms involved:
H
Fluid mass entering face – AB = r × udy

z z
0
H
r × udy +
d FG H IJ
Fluid mass leaving through face – CD =
0 dx H r × udy × dx
0 K
From continuity consideration, mass increment viz.

must enter the control volume from top face BC.


d
dx
F
GH z0
H
r × udy × dx
I
JK
Heat fluxes through the four faces are:
Heat influx through AB = Qx = mass X specific heat X temperature

z
i.e.
H
Qx = r× udy × Cp × T …(a)

z z
0
H d F H I
Heat efflux through CD = Qx + dx =
0
r× udy × Cp × T +
dx GH 0
JK
r × udy ×C p × T × dx …(b)

Heat influx through upper face BC: Since face BC is well outside the thermal boundary layer, its temperature
is equal to free stream temperature, Ta. So, we have:

QBC =
d
dx
Heat conducted into the control volume through lower face AD =
F
GH z 0
H
r × udy × Cp ×Ta
I
JK …(C)

F dt I
QAD = – k×A× GH dy JK y = 0
F dT I
i.e QAD = – k× dx× GH dy JK y=0
...(d)

Writing the energy balance:


Heat flow into the control volume = Heat flow out of the control volume

412 FUNDAMENTALS OF HEAT AND MASS TRANSFER


or,
Qx + Q BC + QAD = Qx + dx
Substituting from Eqs. a, b, c and d and simplifying, we get:
d
dx
LM
N z
0
H
(Ta - T )× udy =
OP k ×F dT I
Q r ×C GH dy JK
p y=0

i.e.
d L
dx N
M z
0
H
(Ta
O F dT I
- T )× udy P = a ×G J
Q H dy K y=0
...(9.54)

Eq. 9.54 is the integral energy equation for the boundary layer, with constant thermo-physical properties and
constant free stream temperature.
Note that we have neglected viscous dissipation in the element since it is very small for low velocities.
To solve the integral energy equation we have to assume the velocity and temperature profiles; let us as-
sume cubic velocity profile and cubic temperature profiles.
Cubic velocity profile, as shown earlier, is:

u 3 y 1 y FG IJ 3

U
= × – ×
2 d 2 d H K ...(9.55)

Temperature distribution must satisfy the boundary conditions:

d2 T
At y = 0: T = Ts and =0
dy 2
dT
At y = dt : T = Ta and =0
dy
These boundary conditions are of the form as required for the velocity profile; therefore, temperature distri-
bution is also of the form:

T - Ts 3 y 1 y F I F I 3
q
qa
=
Ta - Ts
= ×
2 dt
- ×
2 dt
GH JK GH JK ...(9.56)

where Ts is the plate surface temperature, Ta is the free stream temperature and dt is the thickness of thermal
boundary layer at a given section.
Now, the Eqs. 9.55 and 9.56 are inserted in the integral Eq. 9.54 and simplified. For most gases (Pr @ 1) and
oils (Pr > 1), thermal boundary layer is thinner than hydrodynamic boundary layer, i.e. d t < d; so, upper limit of
integration is changed to d t instead of H because the integrand becomes zero beyond y = d t.
Final solution for the thermal boundary layer thickness is:

L OP
1

0. 976 M F x I
3 3

× M1 - G J
dt
PP
4
H xK
0
= ...(9.57)
Pr MN
1
d
3 Q
Remember that Eq. 9.57 is for the case when the heating of the plate starts at a distance of x 0 from the leading
edge. Instead, if the heating starts from the leading edge itself, putting x 0 = 0, we get:
dt 0.976
= 1
...(9.58)
d
Pr 3
Observe that this value of d t is close to the value obtained with exact analysis.
Local heat transfer coefficient (hx ):
We obtain hx from the relation:

FORCED CONVECTION 413


F dT I
- k× GH dy JK y=0
hx =
Ts - Ta
Getting dT/dy from Eq. 9.56, and taking the values of d and (d t/d) from Eqs. 9.49 and 9.57 respectively, we
get:
1 1
k 1
hx = 0.332× × Re x2 × Pr 3 × ...(9.59)
LM F x I OP
x 1
3 3

MM1 - GH x JK PP
0 4

N Q
and, in terms of non-dimensional Nusselt number, we write:
1 1
h ×x 0. 332 × Re x2 × Pr 3
Nux = x = ...(9.60)
LM F I OP
1
k 3 3

MM GH JK PP
x 4
1- 0
x
N Q
If the plate is heated over the entire length, x 0 = 0, and we get:
1 1
k
hx = 0.332× × Rex2 × Pr 3 ...(9.61)
x
and,
1 1
hx × x
Nux = = 0.332× Rex2 × Pr 3 ...(9.62)
k
Note that Eq. 9.62 is in excellent agreement with the value obtained with exact analysis.

z
Average value of the heat transfer coefficient is obtained by integrating the local value over the entire plate:
1 L
ha = × hx dx
L 0
Performing the integration after substituting for h x , we get:
ha = 2×h L ...(9.63)
Similarly, average value of Nusselt number is obtained as:
1 1
ha × L
Nua = = 0.664× ReL2 × Pr 3 ...(9.64a)
k
or, Nua = 2×Nu L ...(9.64b)
r ×U × L
where ReL =
m
Note that all the above analysis is for laminar boundary layer conditions; property values are taken at film
temperature (i.e. mean value of surface and free stream temperatures), given by:
Ts + Ta
Tf =
2
Eq. 9.61 is valid for fluids with Prandtl numbers varying from 0.6 to 50 i.e. it is not applicable to liquid
metals for whom Pr << 0.6 and for heavy oils or silicones for whom Pr >> 50.
For a wide range of Prandtl numbers Churchill and Ozoe have given the following correlation, for laminar
flow on an isothermal flat plate:

414 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1 1
0. 3387 × Rex2 × Pr 3
Nu x = ...for Rex Pr > 100 ...(9.65)
LM F 0.0468 I OP
1
2 4

MM1 + GH Pr JK PP
3

N Q
For constant heat flux conditions:
All the above relations were derived for laminar flow over a flat plate, temperature of the plate being maintained
constant. However, there are many practical cases where the heat flux over the surface is constant (e.g. when the
surface is heated by electrical heaters).
For the case of constant heat flux, it is shown that local Nusselt number is given by:
1 1
h×x
Nux = = 0.453× Rex2 × Pr 3 ...Pr ³ 0.6...(9.66)
k
In terms of surface heat flux and temperature difference, this is written as:
qs × x
Nux = ...(9.67)
k × (Ts - Ta )

z
Average temperature difference along the plate for this case is obtained by performing the integration:
1 L
(Ts – Ta )avg = × (Ts - Ta ) dx
L 0
Substituting for (Ts – Ta) from Eq. 9.67 and performing the integration, we get:
L
qs ×
(Ts – Ta)avg = k ...(9.68)
1 1
0.6795 ReL2 × Pr 3

3
and, qs = × hL × (Ts - Ta )avg ...(9.69)
2
2
In the above equations, q s is the heat flux per unit area with units: W/m .
Again, for the constant heat flux case, Eq. 9.65 for very wide range of Prandtl numbers, is modified as:
1 1
0. 4637 × Rex2 × Pr 3
Nux = ...for Rex Pr > 100 ...(9.70)
LM F 0.02052 I OP
1
2 4

MM1 + GH Pr JK PP
3

N Q
Fluid properties are still evaluated at the film temperature.
In all cases, average Nusselt number is Nua = 2 × NuL ... (9.70a)
Example 9.4. Air at 20°C and atmospheric pressure is flowing with a velocity of 3 m/s along the length of a flat plate,
maintained at 60°C. Calculate: (i)hydrodynamic boundary layer thickness at 20 cm and 40 cm from the leading edge, by
the approximate method (ii) mass entrainment rate between these two sections assuming a cubic velocity profile, and
(iii) heat transferred from the first 40 cm of the plate.
Solution.
Data:
60 + 20
Ts := 60°C Ta := 20°C U := 3.0 m/s x1 := 0.2 m x2: = 0.4 m Tf =
= 40°C
2
Properties of air are to be taken at the film temperature of 40°C. We get, from data tables:
r := 1.092 kg/m3 Cp := 1014 J/(kgK) m := 19.123 ´ 10 –6 NS/m2 k := 0.0265 W/(mK) Pr := 1.01

FORCED CONVECTION 415


(i) Hydrodynamic boundary layer thickness at section-1 (i.e. x = 0.2 m) of plate:
Now, characteristic dimension to calculate Reynolds number is length x1
r ×U × x1
Therefore, Rex1 := (local Reynolds no. at x 1 = 0.2 m)
m
i.e. Re x1 = 3.426 ´ 104
This value is less than 5 ´ 105; so, the boundary layer is laminar and the equations derived above are applicable:
Hydrodynamic boundary layer thickness, d 1:
4 .64 × x1
We have: d 1 := ...(9.49)
Re x 1
i.e. d 1 = 5.013 ´ 10 –3 m (thickness of hydrodynamic boundary layer at x 1 = 0.2 m.)
Hydrodynamic boundary layer thickness at section-2 (i.e. x = 0.4 m) of plate:
Now, characteristic dimension to calculate Reynolds number is length x 2:
r ×U × x2
Therefore, Rex2 := (local Reynolds no. at x 2 = 0.4 m)
m
i.e. Rex2 = 6.852 ´ 10 4
This value is less than 5 ´ 10 5; so, the boundary layer is laminar and the equations derived above are applicable:
Hydrodynamic boundary layer thickness d 2:
4 .64 × x2
We have: d 2 := ...(9.49)
Re x 2
i.e. d 2 = 7.09 ´ 10 –3 m (thickness of hydrodynamic boundary layer at x 2 = 0.2 m.)
(ii) Mass flow entrained between sections 1 and 2:
For a cubic velocity profile, mass flow entrained between section 1 and 2 is already shown to be:
5
dm : = ×r×U×(d 2 – d 1) ...(9.53)
8
i.e. dm = 4.252 ´ 10 –3 kg/s (mass entrained between sections 1 and 2.)
(iii) heat transferred from the first 40 cm of the plate:
Now, we have: Rex2 = 6.852 ´ 10 4
1 1
Nux := 0.332× Re x22 × Pr 3 ...(9.62)
i.e. Nu x = 87.197 (Nusselt number)
hx × x
But, Nu x =
k
Nux × k
Therefore, hx :=
x2
i.e. hx = 5.777 W/(m2C) (local heat transfer coefficient)
To calculate the heat transferred from first 40 cm of the plate, we need the average value of heat transfer coefficient
over this length. It is given by twice the value of local heat transfer coefficient at x = 0.4 m. i.e.
ha := 2× hx ...(9.63)
i.e. ha = 11.554 W/(m2C) (average heat transfer coefficient over 40 cm length)
Area := 0.4.1 m2 (heat transfer area for unit width)
Therefore, heat transferred over 40 cm length of plate:
Q := ha × Area×(Ts – Ta) W
i.e. Q = 184.858 W ...heat transferred over 40 cm length of plate.
Example 9.5. Engine oil at 30°C is flowing with a velocity of 2 m/s along the length of a flat plate, maintained at 90°C.
Calculate, at a distance of 40 cm from the leading edge: (i) hydrodynamic and thermal boundary layer thicknesses by the
exact method (ii) local and average values of friction coefficient (iii) local and average values of heat transfer coefficient,
and (iv) heat transferred from the first 40 cm of the plate for unit width.
Solution.
Data:
90 + 30
Ts := 90°C Ta := 30°C U := 2.0 m/s x := 0.4 m Tf = = 60°C
2
416 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Properties of engine oil are to be taken at the film temperature of 60°C. We get, from data tables:
r := 864 kg/m3 Cp := 2047 J/(kgK) m := 72.5 ´ 10– 3 Ns/m2 k := 0.140 W/(mK) Pr := 1050
(i) Hydrodynamic and thermal boundary layer thickness at 0.4 m from leading edge of plate:
Now, characteristic dimension to calculate Reynolds number is: x = 0.4 m
r ×U × x
Therefore, Rex := (local Reynolds no. at x = 0.4 m)
m
i.e. Rex = 9.534 ´ 10 3
This value is less than 5 ´ 105; so, the boundary layer is laminar and the equations derived earlier are applicable:
Hydrodynamic boundary layer thickness d:
5× x
We have: d := ...(9.28)
Re x
i.e. d = 0.02 m (thickness of hydrodynamic boundary layer.)
Thickness of thermal boundary layer:
We have:
d
= Pr 0.333 ...(9.35)
dt
d
i.e. d t :=
Pr 0 . 333
i.e. d t = 2.02 ´ 10 –3m (thickness of thermal boundary layer.)
Note that thermal boundary layer thickness is very small compared to that of hydrodynamic boundary layer, since
Pr >> 1.
(ii) Local and average values of friction coefficient:
We have:
0 .664
Cfx := ...(9.31)
Re x
i.e. Cfx = 6.8 ´ 10 –3 (value of local friction coefficient.)
And,
Cfa := 2× Cfx ...(9.33)
i.e. Cfa = 0.014 (value of average friction coefficient)
(iii) Local and average values of heat transfer coefficient:
Since Prandtl number is very high and Rex ×Pr = 1.001 ´ 107 > 100,
we shall use Eq. (9.65), i.e.
1 1
0. 3387 × Re x2 × Pr 3
Nux := (for Rex Pr > 100...(9.65))
LM F 0.0468 I OP
1
2 4

MM1 + GH Pr JK
3

PP
N Q
i.e. Nu x = 336.027 (Nusselt number)
Nux × k
And, hx := (local value of heat transfer coefficient)
x
2
i.e. hx = 117.61 W/(m C) (value of local heat transfer coefficient.)
Therefore, average value of heat transfer coefficient
ha := 2× hx
i.e. ha = 235.219 W/m2C) (value of average heat transfer coefficient.)
(iv) Heat transferred from the first 40 cm of the plate for unit width.
Area := 0.4 m2 (area of heat transfer for unit width)
Therefore, Q := ha Area × (Ts – Ta)
i.e. Q = 5.645 ´ 103 W (heat transfer rate from the plate between x = 0 and x = 0.4 m

FORCED CONVECTION 417


Example 9.6. An air stream at 20°C and atmospheric pressure, flows with a velocity of 5 m/s over an electriacally heated
flat plate (size: 0.5 m ´ 0.5 m), heater power being 1 kW. Calculate:
(i) the average temperature difference along the plate (ii) heat transfer coefficient, and (iii) temperature of the plate at the
trailing edge
Solution.
Data:
1000
Ta := 20°C U := 5.0 m/s L := 0.5 m W := 0.5 m qs := i.e. qs = 4 ´ 103 W/m2
0 . 5 × 0. 5
Note that properties have to be evaluated at the film temperature; however, since the temperature of the plate is not
constant, but varies along the length, we shall start the analysis taking the properties at 20°C and then refine the values
later.
At 20°C and atmospheric pressure, properties of air are:
r := 1.205 kg/m3 Cp := 1005 J/(kgK) n := 15.06 ´ 10 –6 m2/s k := 0.02593 W/(mK) Pr := 0.703
(i) the average temperature difference along the plate
First check Reynolds number for laminar flow:
U ×L
ReL := i.e. ReL = 1.66 ´ 10 5 < 5 ´ 10 5
n
Therefore, flow is laminar.
For constant heat flux conditions, we use Eq. 9.68, to calculate the average temperature difference:
L
qs ×
(Ts – Ta) avg = k ...(9.68)
1 1
0 .6795 × Re L2 × Pr 3
i.e. (Ts – Ta)avg = 313.325 deg. C
Now, find the properties again at a film temperature of: (20 + 313.3)/2 = 161.5°C
We get:
n := 30.1 ´ 10– 6 m2/s k := 0.0365 W/(mK) Pr := 0.682
Now, using Eq. 9.68 again, we get:
U ×L
ReL := i.e. ReL = 8.306 ´ 10 4
n
and, (Ts – Ta)avg = 317.882 deg.C
Therefore, film temperature:
317 . 88 + 20
= 168.94°C
2
Now, again, find the properties again at a film temperature of: 169°C
We get:
n := 31.25 ´ 10 –6 m2/s k := 0.0371 W/(mK) Pr := 0.6815
Now, using Eq. 9.68 again, we get:
U ×L
ReL := i.e. ReL = 8 ´ 104
n
and, (Ts – Ta)avg = 318.737 deg. C
Therefore, film temperature:
318.737 + 20
= 169.369°C
2
Therefore, the film, temperature has not changed much. So, we conclude:
(Ts – Ta)avg = 318.737 deg. C (Average value of temperature difference over the plate length.)
(ii) Convection heat transfer coefficient:
We have, for the case of constant heat flux:
1
h× x 1
Nux = = 0.453 × Re L2 × Pr 3 ...(9.66)
k

418 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Now,
U ×L
ReL := i.e. ReL = 8 ´ 10 4 (Note: taking v at 169°C)
n
1 1
From Eq. 9.66: NuL := 0.453× Re L2 × Pr 3
i.e. NuL = 112.754
NuL × k
Therefore, hL :=
L
i.e. hL = 8.366 W/(m2C) (local heat transfer coefficient at the end of plate, i.e. at x = L)
Therefore, average heat transfer coefficient over the whole length of plate:
havg := 2× hL
i.e. havg = 16.733 W/(m2C) (average heat transfer coefficient over the plate.)
(iii) Temperature of the plate at the trailing edge:
From the basic heat flow equation., we have:
qs × L NuL × k
Ts – Ta = as h =
NuL × k L
i.e. Ts – Ta = 478.1
i.e. Ts = 498.1°C ...temperature of plate at trailing edge.
Example 9.7. Sodium–Potassium alloy (25% + 75 %), at 300°C , flows with a velocity of 0.4 m/s over a flat plate (size: 0.3
m ´ 0.1 m), maintained at 500°C. Calculate (i) the hydrodynamic and thermal boundary layer thicknesses (ii) local and
average value of friction coefficient (iii) heat transfer coefficient, and (iv) total heat transfer rate
Solution.
Data:
300 + 500
Ts := 500°C Ta := 300°C U := 0.4 m/s L := 0.3 m W := 0.1 m Tf =
= 400°C
2
Properties of Na–K alloy are to be taken at the film temperature of 400°C. We get, from data tables:
n := 0.308 ´ 10 –6 m2/s k := 22.10 W/(mK) Pr := 0.0108
(i) Hydrodynamic and thermal boundary layer thickness at 0.3 m from leading edge of plate:
Now, characteristic dimension to calculate Reynolds number is: x = 0.3 m
U ×L
Therefore, ReL := (local Reynolds no. at L = 0.3 m)
n
5
i.e. ReL = 3.896 ´ 10
This value is less than 5 ´ 105; so, the boundary layer is laminar and the equations derived earlier are applicable:
Hydrodynamic boundary layer thickness, d :
5 ×L
We have: d := ...(9.28)
Re L
i.e. d = 2.403 ´ 10 –3 m (thickness of hydrodynamic boundary layer.)
Thickness of thermal boundary layer, d t :
We have:
d
= Pr 0.333 ...(9.35)
dt
d
i.e. d t :=
Pr 0 . 333
i.e. d t := 0.011 m (thickness of thermal boundary layer.)
Note that thermal boundary layer thickness is very large compared to that of hydrodynamic boundary layer, since
Pr << 1.
(ii) Local and average values of friction coefficient:
We have:
0 .664
CfL := ...(9.31)
Re L

FORCED CONVECTION 419


i.e. CfL = 1.064 ´ 10 –3 (local value of friction coefficient)
And,
Cfa := 2×CfL ...(9.33)
i.e. Cfa = 2.128 ´ 10 –3 (average value of friction coefficient)
(iii) Local and average values of heat transfer coefficient:
Since Prandtl number is very low (liquid metal) and ReL ×Pr = 4.208 ´ 10 3 > 100,
we shall use Eq. 9.65 i.e.
1 1
0. 3387 × Re x2 × Pr 3
Nu x := (for Rex Pr > 100...(9.65))
LM F 0.0468 I OP
1
2 4

MM1 + GH Pr JK
3

PP
N Q
i.e. Nu L = 33.791 (Nusselt number)
NuL × k
And, hL := (local value of heat transfer coefficient)
L
3 2
i.e. hL = 2.489 ´ 10 W/(m C) (local value of heat transfer coefficient)
Therefore, average value of heat transfer coefficient
ha := 2× hL
i.e. ha = 4.978 ´ 103 W/(m2C) (average value of heat transfer coefficient)
(iv) Heat transferred from the plate:
Area := 0.03 m2 (area of heat transfer)
Therefore, Q := ha × Area×(Ts – Ta)
i.e. Q = 2.987 ´ 10 4 W (heat transfer rate from the plate between x = 0 and x = 0.4 m)
Note: Alternatively, for liquid metals, we can also use Eq. 9.43 to get local Nusselt number:
Nux := 0.565 Pe x0.5 ((Pr < 005)...(9.43))
where Pe is the Peclet number = Re.Pr
i.e. NuL := 0.565 × (ReL × Pr)0.5
i.e. Nu L = 36.65
Compare this value of Nusselt number with the value of 33.791, obtained from Eq. 9.65.
NuL × k
Then, hL := (local value of heat transfer coefficient)
L
3 2
i.e. hL = 2.7 ´ 10 W/(m C) (local value of heat transfer coefficient)
Therefore, average value of heat transfer coefficient
ha := 2× hL
i.e. ha = 5.4 ´ 103 W/(m2c) (average value of heat transfer coefficient)
And, Q := ha × Area×(Ts – Ta)
i.e. Q = 3.24 ´ 104 W (heat transfer rate from the plate between x = 0 and x = 0.4 m.)
Value of Q thus obtained is about 8.5% higher than the value obtained by using Eq. 9.65.
9.8.3.1 Turbulent boundary layer flow over a flat plate. Consider a flat plate over which a fluid flows with a free
stream velocity of U. At the leading edge the fluid comes in contact with the surface and then along the length a
boundary layer develops, as explained earlier. For a certain distance from the leading edge the flow in the
boundary layer is ‘laminar’, i.e. the flow is regular and the layers of fluid are all parallel to each other; however,
after this distance, called ‘critical distance’ (xc), the flow becomes ‘turbulent’, i.e. the flow becomes highly irregu-
lar and there is completely random motion of fluid chunks. The transition from laminar to turbulent is not sud-
den, but there is a transition region in between. The dimensionless number characterizing the type of flow i.e.
whether it is laminar or turbulent, is the Reynolds number, Re (= r.U.L/m). For a flat plate, generally accepted
value of Re at which flow becomes turbulent is 5 ´ 10 5; however, it should be understood that this value is not a
fixed value, but depends on the surface conditions i.e. if the surface is smooth or rough.
The turbulent boundary layer itself is thought of as subdivided into three sections viz. a laminar sub-layer,
a buffer layer and lastly, a turbulent region. See Fig. 9.2.
Now, one could easily imagine that because of the nature of random motion of fluid in turbulent flow, an
exact mathematical analysis of this phenomenon is rather difficult. Models have been proposed by many re-

420 FUNDAMENTALS OF HEAT AND MASS TRANSFER


search workers to explain the observed phenomenon: Reynolds conducted his famous ‘dye experiment’ in 1883
to visually demonstrate the transition from laminar to turbulent flow. In turbulent flow it is observed that sec-
ondary motions of the fluid are superimposed on the main flow and there are irregular fluctuations of local
velocity. Chunks of fluid, called ‘eddies’ move across the line of motion causing mixing of the fluid, thus causing
the transport of momentum as well as energy. Therefore, in turbulent flow, heat transfer is enhanced; also, there
is increased ‘drag force’ or pressure drop. Prandtl (1925) suggested that the ‘eddies’ moving across the fluid
layers cause the transport of momentum, and the average transverse distance moved by an eddy before it gets
mixed with other particles and loses its identity is called as ‘mixing length’. This mixing length is akin to the
‘mean free path’ appearing in the kinetic theory of gases.
Turbulent flow is important in heat transfer applications, since there is increased heat transfer in turbulent
flow; of course, this is achieved with a penalty of increased pressure drop. It is usual to introduce ‘turbulence
promoters’ in applications where increased heat transfer is the primary consideration.
We shall not go into the theories of turbulence, but give here the important results useful for practical appli-
cations.
Velocity distribution:
Boundary layer thickness is more in turbulent flow as compared to that in laminar flow. Also, the velocity distri-
bution is more uniform across the thickness of boundary layer as shown in Fig. 9.2. It is observed experimentally
that the velocity distribution in turbulent flow follows the one-seventh power law:

FG IJ
1
u y 7
U
=
d H K ...(9.71)

Surface shear stress:


Surface shear stress is given by:

FG y IJ
7 1
4
tw = 0.0225r× U 4 ×
HdK ...(9.72)

Hydrodynamic boundary layer thickness:


This is obtained by solving the integral momentum equation, i.e.

tw =
d
dx
LM
N z
0
d
r × (U - u) × udy
OP
Q
Substituting for u(y) and tw from Eqs. 9.71 and 9.72 respectively, and solving, we get:
-1
d
= 0.371× Rex5 ...(9.73)
x
Thermal boundary layer thickness:
In turbulent flow, since the effects of physical movement of eddies predominates over the diffusion effects,
Prandtl number does not have much influence on the thermal boundary layer thickness, d t and is of the same
order as the hydrodynamic boundary layer thickness, d.
Local skin friction coefficient:
Remembering that local skin friction coefficient is defined as:
tw
Cfx =
1
× r ×U 2
2
and using Eq. 9.72 for tw and Eq. 9.73 for d, we get:
-1
Cfx = 0.0576 Rex5 ...(9.74)
Average value of skin friction coefficient:
Average value of Cfx over length L is given by:

Cfa =
1
L
×
z0
L
C fx dx

421
FORCED CONVECTION
Substituting for Cfx from Eq. 9.74 and performing the integration, we get:
-1
Cfa = 0.072× ReL5 (for 5 ´ 10 5 < ReL < 107...(9.75))
5 7
Eq. 9.75 is valid for 5 ´ 10 < ReL < 10 and 0.6 < Pr < 60.
For values of ReL between 107 and 109 following equation is suggested by Prandtl and Schlichting:
0. 455
Cfa = (for 10 7 < ReL < 10 9...(9.76))
d b gi
log ReL
2. 58

where log (ReL) is the logarithm to base 10.


Local and average Nusselt numbers:
Local Nusselt number is calculated by applying Colburn analogy (which we shall study in the next section). We
get:
1
hx × x
Nu x = = 0.0288× Rex0 .8 × Pr 3 (0.6 < Pr < 60 ...(9.77))
k
and,
1
ha × L 0 .8
Nuavg = = 0.036× ReL × Pr 3 ...(9.78)
k
For Eqs. 9.77 and 9.88, remember: 5 ´ 10 5 < ReL < 10 7 and 0.6 < Pr < 60
Local and average heat transfer coefficients:
These are determined from:

FG k IJ × Re 1

H xK
0 .8
hx = 0.0288× x × Pr 3 ...(9.79)
and,

FG k IJ × Re 0 .8
1
ha = 0.036×
H LK L × Pr 3 ...(9.80)

For an unheated starting length of x0:


In turbulent flow, when heating starts from an initial length of x 0 , i.e. thermal boundary layer begins at x = x 0:
1
0. 8
0.0288 × Re x × Pr 3
Nux = ...(9.81)
LM F I OP
1
9 9

MM1 - GH x JK
x0 10
PP
N Q
Note: both x 0 and x are measured from the leading edge of the plate.
Some comments on the variation of local heat transfer coefficient and local friction coefficient along the length x
from the leading edge of the plate, in laminar and turbulent flow are appropriate:
(a) In laminar flow, we have:
t 0.664
Cfx =
F r×U I 2
=
Rex
...(9.31)

GH 2 JK
and
h×x
= Nux = 0.332× Re x × Pr 0 . 333 ...(9.38)
k
i.e. in laminar flow, local friction coefficient varies as x – 1/2; likewise, from Eq. 9.38, it is clear that local heat
transfer coefficient also varies as x –1/2. Of course, at the leading edge (i.e. at x = 0), both these values are infinite
and then decrease along the length of the plate according to x –1/2.
422 FUNDAMENTALS OF HEAT AND MASS TRANSFER
(b) In turbulent flow, we have:
-1
Cfx = 0.0576× Rex5 ...(9.74)
and,

FG k IJ × Re 1

H xK
0 .8
hx = 0.0288× x × Pr
3 ...(9.79)

i.e. in turbulent flow, both the local friction coefficient and the local heat transfer coefficient vary as x-0.2. So, as we
proceed along the length of the plate, initially, starting from the leading edge, the flow is laminar where both the
local friction and heat transfer coefficients vary as x–1/2; then, the flow turns turbulent when the critical distance
is reached, and both the local friction and heat transfer coefficients reach their highest values at this point and
then they decrease along the distance according to: x –0.2. This is shown graphically in Fig. 9.13. In Fig. 9.13, the
transition region is also shown.
For uniform heat flux conditions:
Local Nusselt number increases by about 4% over the value for constant wall temperature, and is given by:
1
Nux = 0.0308× Rex0 .8 × Pr 3 ...(9.81a)
Also, in the above equations, it is assumed that flow over the plate is turbulent over the entire plate from the
leading edge itself, or alternatively, region of laminar flow is too small compared to the region of turbulent flow.
9.8.3.2 Combined laminar and turbulent flow over a flat plate. As explained earlier, for a flow over a flat plate, the
flow at the leading edge starts as laminar and after a critical distance xc the flow becomes turbulent. If the dis-
tance over which the flow is laminar is not negligible as compared to the distance over which the flow is turbu-
lent (i.e. the plate is long enough to cause the boundary layer to become turbulent, but not long enough to neglect
the length over which the flow is laminar), average friction coefficient and average Nusselt number over the
entire plate are determined by integrating the respective local values over two regions, i.e. the laminar region,
0 < x < xc and, the turbulent region, xc < x < L, as shown below:

and,
Cfa =
1
L
×
LM
N z 0
xc
C fx laminar dx +
z L

xc
C fx turb dx
OP
Q
...(9.82)

Nz z
h=
1
×
LM xc
hx laminar dx +
L
hx turb dx
OP ...(9.83)
L 0 xc Q
µ 0. 2
µ 0. 5
hx, Cfx x
hx, Cfx x

Transition region

Laminar Turbulent
Fluid flow region region
U

Leading edge Flat plate Trailing edge

FIGURE 9.13 Variation of local friction and heat transfer coefficients for flow over a flat plate

FORCED CONVECTION 423


If we perform the integration taking the value of critical Reynolds number, Rec as 5 ´ 105, we get for the
average friction coefficient and average Nusselt number, the following relations:
0.074 1742
Cfa = 1
– ...(5 ´ 105 < ReL < 107)...(9.84a)
ReL
ReL5
Another, more general relationship used for critical Reynolds number other than 5 ´ 10 5 is:
0. 455 A
Cfa = – ×8.28 ...(9.84b)
c log( ReL h 2 . 584 Re L
where value of A is 1050, 1700, 3300 and 8700 respectively for values of Rec equal to 3 ´ 105, 5 ´ 105, 1 ´ 106, and
3 ´ 106.
and, for critical Reynolds number of 5 ´ 10 5, average Nusselt number over the entire plate is
F 4 I 1
Nuavg =
h×L
GG
= 0.036 × ReL5 - 836 × Pr 3 JJ ...(0.6 < Pr < 60), and
k H K (5 ´ 105 < ReL < 107)...(9.85a)

and, more generally, for critical Reynolds numbers other than 5 ´ 105:
1
e
Nuavg = Pr 3 × 0.036 × ReL - A
0 .8
j ...(9.85b)

where A = 0.036×Rec0.8 – 0.664×Rec0.5


Example 9.8. A refrigerated truck is moving at a speed of 85 km/h where ambient temperature is 50°C. The body of the
truck is of rectangular shape of size 10 m (L) ´ 4 m(W) ´ 3 m(H). Assume the boundary layer is turbulent and the wall
surface temperature is at 10°C. Neglect heat transfer from vertical front and backside of truck and flow of air is parallel
to 10 m long side. Calculate heat loss from the four surfaces.
For turbulent flow over flat surfaces: Nu = 0.036.Re 0.8 . Pr 0.33
Average properties of air at 30°C: r = 1.165 kg/m3, Cp = 1.005 kJ/kgK, n = 16.10 –6 m2/s, Pr = 0.701
(M.U. Dec. 1999).
Solution.
Data:
Ts := 10°C Ta := 50°C L := 10 m W := 4 m H := 3 m A := L×(W + H)×2 m2 i.e. A = 140 m2
Ts + Ta
Tf := + 273 Tf = 303 K (film temperature)
2
Truck is moving at a speed of 85 km/h, i.e.
85000
U := i.e. U = 23.611 m/s (velocity of air over the surfaces)
3600
Properties at Tf by data:
n := 16 ´ 10 –6 m2/s r := 1.165 kg/m3 Cp := 100 5 J/kgK Pr := 0.701
n × r ×Cp
k := k = 0.02672 W/mK (thermal conductivity of air)
Pr
Check if flow is laminar or turbulent:
Reynolds number:
L ×U
ReL :=
n
i.e. ReL = 1.476 ´ 107 ...> 5 ´ 10 5
i.e. flow is turbulent since Reynolds number is more than 5 ´ 105
Heat transfer:
For turbulent flow, we have:
1
Nu := 0.036× Re L0 . 8 × Pr 3 (Nusselts number)
i.e. Nu = 1.738 ´ 104

424 FUNDAMENTALS OF HEAT AND MASS TRANSFER


k × Nu
h := W/(m2K) (heat transfer coefficient)
L
i.e. h = 46.448 W/(m2K) (heat transfer coefficient)
Therefore,
Q := h×A×(Ta – Ts) W (total heat transfer rate from all the four surfaces)
i.e. Q = 2.60 ´ 105 W (total hat transfer rate from all the four surfaces.)
Also, find out the power required to overcome wind resistance:
We have, average skin friction coefficient given by:
0 . 455
Cfa := (for 10 7 < Re L < 109...(9.76))
dlog bRe gi
L
2 . 58

i.e. Cfa = 2.82 ´ 10 –3 (average skin friction coefficient)


r × A ×U 2
Therefore, Drag force: FD := Cfa × (N)
2
i.e. FD = 128.406 N (Drag force)
Therefore, Power: P := FD ×U, W
i.e. P = 3.032 ´ 10 3 W (Power required to overcome air resistance)
i.e. P = 3.032 kW (Power required to overcome air resistance.)
Example 9.9. A flat plate, 1 m wide and 1.5 m long is maintained at 90°C in air with free stream temperature of 10°C,
flowing along 1.5 m side of the plate. Determine the velocity of air required to have a rate of energy dissipation as 3.75
kW. Use correlations:
NuL = 0.664 Re 0.5 Pr 1/3 for Laminar flow, and
NuL = [0.036 Re 0.8 – 836].Pr 1/3 for turbulent flow.
Take average properties of air at 50°C: r = 1.0877 kg/m3, Cp = 1.007 kJ/kgK,
m = 2.029.10 –5 kg/m.s, Pr = 0.703, k = 0.028 W/mK [P.U.; 1995]
Solution.
Data:
Ts := 90°C Ta := 10°C L := 1.5 m W := 1 m Q := 3750 W A := L×(W)×2 m2 i.e. A = 3 m2
Ts + Ta
Tf : = + 273 Tf = 323 K (film temperature)
2
Properties at Tf : by data:
m := 2.029 ´ 10– 5 kg/(ms) r := 1.0877 kg/m3 Cp := 1007 J/kgK Pr := 0.703 k := 0.028 W/mK
Nusselt number:
We have, for convection heat transfer:
Q = ha × A×(Ts – Ta) W (from Newton’s Law of Cooling)
Q
i.e. ha := W/(m2C) (average heat transfer coefficient)
A ×(Ts - Ta )
i.e. ha = 15.625 W/(m2C) (average heat transfer coefficient)
Therefore, Nusselt number:
ha × L
NuL :=
k
i.e. Nu L = 837.054
Now, we do not know if the flow is laminar or turbulent. To determine this, we need the Reynolds number. But, we
do not know the velocity to determine the Reynolds number. So, we shall first assume the flow to be laminar and then
check if the Reynolds number works out to be less than the critical Reynolds number (i.e. 5 ´ 105):
For Laminar flow:
1
Nu L = 0.664× Re L0 . 5 × Pr 3

LM Nu OP 2

i.e. ReL := M L
PP
MN 0.664 × Pr 1
3
Q
FORCED CONVECTION 425
i.e. ReL = 2.01 ´ 106
This value of Reynolds number is greater than the critical Reynolds number of 5 ´ 10 5. Therefore, the assumption
that the flow is laminar is wrong.
Then, for turbulent flow, we use the relation:
1

d
NuL := 0. 036 × ReL0 . 8 - 836 × Pr 3 i
Therefore,

LM L Nu O
MM M + 836OPP PP
1
0. 8
L

M N Pr Q PP
1

:= M
3
ReL
MM 0 .036
PP
MN PQ
i.e. ReL = 7.36 ´ 105 > 5 ´ 105 (Therefore, assumption of turbulent flow is correct.)
To find the velocity of air:
r ×U × L
We have: ReL = (Reynolds number, by definition)
m

Re L × m
i.e. U := m/s (velocity of air)
r ×L
i.e. U = 9.152 m/s (velocity of air.)
Example 9.10. Air at 30°C flows over a flat plate, 0.4 m wide and 0.75 m long with a velocity of 20 m/s. Determine the
heat flow rate from the surface of the plate assuming that the flow is parallel to the 0.75 m side. Plate is maintained at
90°C. Use correlations:
NuL = 0.664 Re 0.5 Pr 1/3 for Laminar flow, and
NuL = [0.036 Re 0.8 – 836].Pr 1/3 for turbulent flow.
Take average properties of air at 60°C: r = 1.06 kg/m3, Cp = 1.008 kJ/kgK,
n = 18.97×10 –6 m2/s, Pr = 0.708, k = 0.0285 W/mK [M.U.]
Solution.
Data:
Ts := 90°C Ta := 30°C U := 20 m/s L := 0.75 m W := 0.4 m
Ts + Ta
Tf := + 273 Tf = 333 K (film temperature)
2
Properties at Tf by data:
n := 18.97 ´ 10 –6 m2/s r := 1.06 kg/m3 Cp := 1008 J/kgK Pr := 0.708 k := 0.0285 W/mK
First, let us find out the distance from the leading edge at which the flow turns turbulent, assuming the critical
Reynolds number to be 5 ´ 105, i.e. Lc at which the critical Reynolds number is reached:
Rec := 5 ´ 105 (critical Reynolds number)
Lc ×U
Rec =
n
Re c ×n
i.e. Lc :=
U
i.e. Lc = 0.474 m ...length from leading edge, at which flow turns turbulent.
i.e. along the length of the plate, for a distance of 0.474 m, the flow is laminar. This distance can not be neglected as
compared to the total length of the plate of 0.75 m. Therefore, combined effect of laminar and turbulent boundary layer
flow has to be considered.
For the case of combined laminar and turbulent boundary layers, we have:
L ×U
ReL := i.e. Re L = 7.907 ´ 105 (Reynolds number at the end of plate)
n

426 FUNDAMENTALS OF HEAT AND MASS TRANSFER


ha × L F 4
I 1
Nu avg =
k GH
= 0. 036 × ReL5 - 836 × Pr 3 JK ...(0.6 < Pr < 60), and (5 ´ 105 < ReL < 107) ...(9.85a)

F 0.036 × Re 4
I 1
Therefore, Nuavg := GH 5
L JK
- 836 × Pr 3 (average Nusselt number over the entire plate)

i.e. Nu avg = 932.666


Nuavg × k
or, h a := W/(m2K) (average heat transfer coefficient over the entire plate)
L
2
i.e. ha = 35.441 W/(m K) (average heat transfer coefficient over the entire plate)
Therefore, heat transfer rate:
Q := ha (L×W) × (Ts – Ta) W (heat transfer rate from the entire plate)
i.e. Q = 637.944 W (heat transfer rate from the entire plate)
Alternatively, we can calculate the heat transferred by the laminar and turbulent regions separately, and then add
them up, to get the total heat transfer rate for the whole plate:
For laminar flow region (i.e. upto a distance of 0.474 m along the length):
1
Nu lam := 0.664× Re c0 . 5 × Pr 3 (Nusselts number for Laminar region)
i.e. Nu lam = 418.47
Nulam × k
Therefore, ha_lam := W/(m2K) (average heat transfer coefficient over the laminar region)
Lc
i.e. ha_lam = 25.148 W/(m2k) (average heat transfer coefficient over the laminar region)
Therefore, heat transfer rate for the laminar region, Q 1:
Q 1 := ha_lam ×(W×Lc) (Ts – Ta) W (heat transfer from laminar region)
i.e. Q1 = 286.233 W (heat transfer from laminar region)
For turbulentr flow region (i.e. from a distance of 0.474 m upto the end of plate):
Local Nusselt number for the turbulent region is given by:
hx × x 1
Nux = = 0.0288× Re x0 . 8 × Pr 3 ...(9.77)
k
FG k IJ ×FG U × x IJ 0 .8 1
i.e. hx = 0.0288×
H xK H n K × Pr 3

= 0.0288 ×k× Pr × G J
FUI 1 0. 8

i.e. hx
HnK 3 × x - 0. 2

Therefore, average value of heat transfer coefficient for turbulent region is obtained as

i.e.
h a_turb = 0.0288 ×k× Pr 3 ×
1
FG U IJ 0.8 × 1 × Lx- 0.2 dx
H n K (L - Lc ) L z c

1
FG U IJ 0. 8
1 ( L0 . 8 - L0c . 8 )
h a_turb = 0.0288 ×k× Pr 3 ×
HnK ×
( L - Lc )
×
0. 8
i.e.
LMF U × L I 0. 8
FG (U × L ) IJ OP
0. 8

MNGH n JK
1
1
H n K PQ
c
h a_turb = 0.036×k× Pr 3 × × -
( L - Lc )
i.e.
1
1
h a_turb = 0.036×k× Pr 3 × ×[( Re L ) 0 . 8 - ( Rec ) 0 . 8 ]
( L - Lc )

ha _turb × ( L - Lc ) 1
i.e. = 0.036× [( Re L ) 0 . 8 - ( Re c ) 0 . 8 ] × Pr 3
k
FORCED CONVECTION 427
ha_turb × ( L - Lc )
Note that is the average Nusselts number for turbulent region
k
Heat transfer rate for the turbulent region Q 2
1
1
h a_turb := 0.036×k× Pr 3 × ×[( Re L ) 0 . 8 - ( Rec ) 0 . 8 ]
( L - Lc )
i.e. ha_turb = 53.229 W/(m2K) (average heat transfer coefficient over the turb. region)
Q 2 := ha_turb ×(W)×(L – Lc)×(Ts – Ta) W (heat transfer rate for turbulent region)
i.e. Q 2 = 352.269 W (heat transfer rate for turbulent region)
Therefour, total heat transfer rate, Q:
Q := Q 1 + Q 2
i.e. Q = 638.502 W (total heat transfer rate for the plate)
This value matches with the value obtained earlier by direct formula.
To show graphically the variation of local heat transfer coefficient over the entire length of plate:
We have stated earlier that the local heat transfer coefficient for the laminar region varies as x –0.5 and that for the
turbulent region varies as x– 0.2. Let us illustrate this graphically, using Mathcad.
For laminar region, i.e. from x 1 = 0 to x 1 = 0.474 m along the length of plate, local heat transfer coefficient as a
function of x is written as:
FG
k x1 ×U IJ 0.5 1
hx_lam (x 1) := 0.332×
x1
×
n H K × Pr 3

For turbulent region, i.e. from x 2 = 0.474 m to x 2 = 0.75 m along the length of plate, local heat transfer coefficient as
a function of x is written as:
FG
k x 2 ×U IJ 0 .8 1
hx_turb(x 2) := 0.088×
x2
×
n H K × Pr 3

Now, for the first case, let us define a range variable x1 varying from x 1 = 0 to x 1 = 0.474 m and draw the graph by
choosing the x–y graph from the graph pallete, and filling up the place holder on the x-axis with x1 and the place holder
on the y-axis with hx-lam (x 1); then for the second case, again define a range variable x 2 varying from x 2 = 0.474 m to x2 =
0.75 m and in the place holder on the x-axis, put a comma after x 1 and type x 2 and in the place holder on the y-axis put
a comma after hx-lam (x 1) and then type h x-turb (x 2). Click anywhere outside the graph region and immediately the graphs
appear. See Fig. Ex. 9.10
x1 := 0, 0.01, ..., 0.47 (define range variable x 1 varying from 0 to 0.47 m, with an increment of 0.01 m)
x 2 := 0.47, 0.48, ..., 0.75 (define range variable x 2 varying from 0.47 to 0.75 m, with an increment of 0.01 m)

hx against x for laminar & turbulent flow


100
90
Heat transfer coefficient (W/m C)
2

80
70
60
50
40
30
20
10
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Dist. from leading edge (m)
hx for laminar flow
hx for turbulent flow

FIGURE Example 9.10 Variation of local heat transfer coefficient along the length of a flat plate for laminar
and turbulent boundary layer heat transfer

428 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Note:
(i) In the above Fig. Example 9.10 first portion of the curve is for laminar boundary layer heat transfer and the
second portion is for turbulent boundary layer heat transfer. In laminar portion, the local heat transfer falls
steeply, as x –0.5 , along the length upto the critical distance; once the critical distance is reached, the boundary
layer turns turbulent and the local heat transfer coefficient suddenly increases to a high value and then, with
increasing x the local heat transfer coefficient drops more gradually (as x –0.2) as compared to the laminar por-
tion.
(ii) In the laminar region, the heat transfer coefficient varies from an infinite value at x = 0 to about 12 W/(m2C) at
x = 0.47 m. And, average heat transfer coefficient for the laminar region, as already calculated, is 25.148 W/
(m2 C).
(iii) In the turbulent region, the heat transfer coefficient varies from a value of about 56 W/(m2C) at x = 0.47 m to
about 51 W/(m2C) at x = 0.75 m. And, average heat transfer coefficient for the turbulent region, as already
calculated, is 53.229 W/(m2C).
(iv) Average heat transfer coefficient over the entire plate, for the combined laminar and turbulent regions, is 35.441
W/(m2 C).

9.8.4 Analogy Between Momentum and Heat Transfer


We have shown that the two-dimensional equations for the momentum transport and energy transport have
identical forms. It is reasonable to assume that their solutions also must have some correspondence to each other.
Solution of momentum equation leads us to a relation for the skin friction coefficient and the drag force;
similarly, solution of the energy equation leads us to an expression for the heat transfer coefficient. So, we seek an
analogy or relation between the fluid friction and heat transfer coefficients:
9.8.4.1 Relation between the fluid friction and heat transfer coefficient in laminar flow for a flat plate. Recollect that
the average Nusselt number for laminar flow over a flat plate is given by:
Nu a = 0.664× Re L × Pr 0 . 333 ...(9.41)
This can be rewritten as:
-1 -2
Nua
= 0.664× ReL2 × Pr 3 ...(a)
ReL × Pr
Now, the LHS of Eq. a is a dimensionless number known as “Stanton number”, Sta.
Substituting for Nua, Re L and Pr from their respective definitions, we get:
ha ha
Sta = = ...(b)
r×U × Cp G × Cp
where G = r.U (kg/(m2s), is known as mass velocity.
Therefore, we write:
-1 -2
Sta = 0.664× ReL2 × Pr 3 ...(c)
2 -1
i.e. Sta ×Pr 3 = 0.664× ReL2 ...(d)

z
However, we have already shown that:
1 L 1. 328
Cfa = × C fx dx = (9.32)
L 0 ReL
i.e.
-1
C fa
= 0.664× Re L2 ...(e)
2
Then, comparing Eqs. d and e, we can write:
2
C fa
Sta ×Pr 3 = ...(9.86)
2

FORCED CONVECTION 429


This relation is known as ‘Colburn analogy’ and it gives a simple relation between the heat transfer coeffi-
cient and the friction coefficient. Eq. 9.86 is valid for values of Pr between 0.6 and 50. LHS of Eq. 9.86 is also
known as ‘Colburn j-factor’, and is generally used to correlate heat transfer coefficient with Reynolds number.
Note the important significance of this analogy: just by knowing the friction coefficient, one can predict the
heat transfer coefficient for that situation; and conducting experiments to determine friction coefficient is, many
times, practically much easier than conducting experiments to determine heat transfer coefficients.
9.8.4.2 Reynolds and Colburn analogies for turbulent flow over a flat plate. Considering the laminar sub-layer adja-
cent to the plate surface, we have the relation for shear stress, along the X-direction and at y = 0:
du
t x = m× ...(a)
dy
and, heat flux at the surface in the y-direction is:
dT
q = – k× ...(b)
dy
Combining Eqs. a and b:
k dT
q = – t×
× ...(c)
m du
Now, if Prandtl number is unity, i.e. if Cp = k/m, we replace (k/m) in Eq. c by Cp, and separating the vari-
ables, we write, assuming q and t to be constant:
qs
×du = – dT ...(d)
t s × Cp
In Eq. d, subscript s indicates that q and t are considered at the surface of the plate.
Integrating Eq. d between the limits u = 0 when T = Ts and u = U when T = Ta, gives:
qs
×U = (Ts – Ta) ...(e)
t s × Cp
However, by definition, the local heat transfer and friction coefficients are given by:

qs r×U 2
hx = and, tx = Cfx ×
Ts - Ta 2
U2
Then, Eq. e can be written as: hx × U = Cfx × r × × Cp
2

hx Nux C fx
i.e. Stx = = = ...(9.87)
r×U × Cp Rex × Pr 2
Eq. 9.87 is known as ‘Reynold’s analogy’ and it gives a relation between Nusselts number (i.e. heat transfer
coefficient) and the friction coefficient. Note that Reynolds analogy was derived with the assumption that Pr = 1
and is valid for most of the gases.
However, when the Prandtl number is different from unity, Colburn’s analogy, i.e.
2
C fx
Stx × Pr 3 = ...(9.88)
2
is applied. This is valid for values of Pr between 0.6 and 50.
In practice, to apply the analogy between momentum and heat transfer, it is necessary to know the friction
coefficient Cfx. For turbulent flow over a flat plate, we have the empirical relation for local friction coefficient:
-1
Cfx = 0.0576× Rex5 ...(9.74)
Eq. 9.74 is valid for: 5 ´ 10 5 < Rex < 10 7.
Example 9.11. Air at 27°C and 1 atm flows over a flat plate at a speed of 2 m/s. Assuming that the plate is heated over
its entire length to a temperature of 60°C, calculate the heat transfer for the first 0.4 m of the plate. Also, compute the

430 FUNDAMENTALS OF HEAT AND MASS TRANSFER


drag force exerted on the first 0.4 m of the plate using Reynolds analogy. Assume air to be a perfect gas with R = 287
J/kgK and Cp = 1.006 kJ/kgK. (M.U. May 1999).
Solution.
Data:
Ts := 60°C Ta := 27°C U := 2 m/s L := 0.4 m B := 1 m
Ts + Ta
Tf := + 273 Tf = 316.5 K (film temperature)
2
R := 287 J/(kgK) (Gas constant for air)
P := 1.01325 ´ 10 5 Pa (atmospheric pressure)
Properties at Tf :
P
n := 17.2 ´ 10 –6 m2/s Cp := 1006 J/kgK Pr := 0.71 k := 0.0271 W/mK r := kg/m 3
R × Tf
i.e. r := 1.115 kg/m3
Reynolds number:
L ×U
ReL :=
n
i.e. ReL = 4.65 ´ 10 4 (less than Recr = 5 ´ 105)
i.e. flow is laminar.
Therefore, we have for average Nusselts number:
1
Nua := 0.664× Re L × Pr 3 ...(9.41)
i.e. Nua = 127.752
Average heat transfer coefficient:
Nua × k
ha := W/(m2 K) (average heat transfer coefficient)
L
i.e. ha = 8.655 W/(m2 K) (average heat transfer coefficient)
Heat transfer rate:
Q := ha × (L×B)×(Ts – Ta) W (heat transfer rate for the first 0.4 m length)
i.e. Q = 114.249 W (heat transfer rate for the first 0.4 m length.)
To calculate the drag force:
We have, for mass velocity:
G := r×U kg/(sm2) (mass velocity)
Therefore, Stanton number, by definition:
ha
St :=
G × Cp
i.e. St = 3.856 ´ 10 –3 (Stanton number)
Nua
Check: St = = 3.869 ´ 10– 3 (checks.)
ReL × Pr
By Reynolds Analogy: St = Cf /2
i.e. Cf := 2×St
i.e. Cf = 7.713 ´ 10 –3 (skin friction coefficient)
Drag force:
U2
FD: = Cf ×r × × ( L × B)
2
i.e. FD = 6.883 ´ 10 –3 N (drag force exerted on first 0.4 m length.)

9.9 Flow Across Cylinders, Spheres and Other Bluff Shapes and Packed Beds
So far, we studied external flow over a flat plate. Next, we shall consider flow across cylinders, spheres and other
bluff shapes such as disk or half cylinder. These cases are of considerable practical importance. Case of single
cylinder in cross flow is identical to the case of cooling of an electrical cable by forced convection by air flowing

FORCED CONVECTION 431


across it; also determination of local velocities in a flow by ‘hot wire anemometer’ involves the heat transfer from
a single platinum wire maintained at a constant temperature (or by passing a constant current through it) and
correlating the change in current (or change in resistance) to the velocity of flow. Heat transfer from a sphere is
important when we are interested in performance of systems where clouds of particles are heated or cooled in a
stream of fluid. Such an understanding is generally required when we correlate data for heat transfer in fluid
beds, especially in the field of chemical engineering. If the particle is of an irregular shape, then an equivalent
diameter is used in place of sphere diameter, i.e. D is taken as the diameter of an equivalent sphere that has the
same surface area as that of the irregular shape. Front portion of an aeroplane wing can be approximated as a
half cylinder while calculating the local heat transfer coefficients over the forward portion of the wing.
9.9.1 Flow Across Cylinders and Spheres
Now, the characteristic length taken to calculate the Reynolds number is the external diameter D of the cylinder
or sphere. And the Reynolds number is defined, as usual:
U ×D
ReD =
n
where U is the uniform velocity of flow as it approaches the cylinder or sphere.
The critical Reynolds number for flow across cylinder or sphere is:
Recr = 2 ´ 10 5
5
i.e. upto Re = 2 ´ 10 , the boundary layer remains laminar and beyond this value, the boundary layer becomes
turbulent.
Flow patterns for a flow across a cylinder are shown in Fig. 9.14. Fluid particles at the mid-plane of a stream
approaching the cylinder strike the cylinder at the ‘stagnation point’ and come to a halt, thus increasing the
pressure. Rest of the fluid branches around the cylinder forming a boundary layer that embraces the cylinder
walls. Pressure decreases in the flow direction and the velocity increases. At very low free stream velocities (Re <
4), the fluid completely wraps around the cylinder; as the velocity increases, boundary layer detaches from the
surface at the rear, forming a wake behind the cylinder. This point is called ‘separation point’. Flow separation
occurs at about q = 80 deg. when the boundary layer is laminar and at about q = 140 deg. when the boundary
layer is turbulent.
Drag coefficient (CD): Drag force for a cylinder in cross flow is primarily due to two effects: one, ‘friction
drag’ due to the shear stress at the surface, and the other, ‘pressure drag’ due to the pressure difference between
the stagnation point and the wake. At low Reynolds numbers (< 4), friction drag is predominant, and at high
Reynolds numbers (> 5000), pressure drag is predominant. At the intermediate values of Re, both the effects
contribute to the drag.

Wake

q
U

Stagnation point
Separation point

Laminar boundary layer

Turbulent boundary layer


Laminar boundary layer

Separation point

FIGURE 9.14 Flow patterns for cross flow over a cylinder

432 FUNDAMENTALS OF HEAT AND MASS TRANSFER


100
80
60
40
a b c d e
20

10
8
6
4
Cp
2
Cylinders
1
0.8
0.6 Spheres
0.4

0.2

0.1
2 3 4 5 6
0.1 0.2 0.5 1 2 5 10 20 50 10 10 10 10 10
Rep

FIGURE 9.15 Drag Coefficient Versus Reynolds Number for Long Circular Cylinders and Spheres in
Cross-Flow

Average drag coefficient CD for cross flow over a cylinder and sphere are shown in Fig. 9.15. Then, the drag
force acting on the body in cross flow is obtained from:

r×U 2
F D = CD ×AN × ,N
2
where AN is the ‘frontal area’ i.e. area normal to the direction of flow.
A N = L.D ...for a cylinder of length L

p ×D2
and, AN = ...for a sphere
4
In Fig. 9.15, there are 5 sections, a, b, c, d and e shown. Comments corresponding to these sections of the
figure are given below:
(a) At Re < 1, inertia forces are negligible and the flow adheres to the surface and drag is only by viscous
forces. Heat transfer is purely by conduction.
(b) At Re = about 10, inertia forces become appreciable; now, pressure drag is about half of the total drag.
(c) At Re of the order of 100, vortices separate and the pressure drag predominates.
(d) At Re values between about 1000 and 100,000, skin friction drag is negligible compared to the pressure
drag. Point of separation is at about q = 80 deg. measured from the stagnation point.
(e) At Re > 100,000, flow in the boundary layer becomes turbulent and the separation point moves to the
rear.
Heat transfer coefficient: Because of the complex nature of flow, most of the results are empirical relations
derived from experiments.
Variation of local Nusselt number around the periphery of a cylinder in cross flow is given in Fig. 9.16. Nu
is high to start with at the stagnation point, then decreases as q increases due to the thickening of laminar bound-
ary layer. For the two curves at the bottom, minimum is reached at about q = 80 deg., the separation point in
laminar flow. For the rest of the curves, there is a sharp increase at about q = 90 deg. due to transition from
laminar to turbulent flow; Nu reaches a second minimum at about q = 140 deg. due to flow separation in turbu-
lent flow, and thereafter increases with q, due to intense mixing in the turbulent wake region.
Between q = 0 and 80 deg. empirical equation for local heat transfer coefficient is:

FG IJ 0.5
LM F q I OP 3

MN GH 90 JK QP
hc (q ) × D r ×U × D
× Pr 0 . 4 × 1 -
Nu(q) =
k
= 1.14×
mH K ...(9.89)

FORCED CONVECTION 433


800 While calculating heat transfer coefficient for a cylinder
in cross flow, of practical interest is the average heat transfer
700 coefficient over the entire surface. A comprehensive relation
for cross flow across a cylinder is given by Churchill and
600 Bernstein:
Re =
219,0
00

L F Re I OP
186 4
500 17 .000 1 1 5 5

× MM1 + G
0.0 00
h×D 0.62 × Re 2 × Pr 3
J 8
PP
LM F 0.4 I OP NM H 28200 K
140.0
Nu cyl
Nu(q)

00 = = 0.3 + 1
400 k
101.300
2 4
Q
MM1 + GH Pr JK PP
3
300 70.8
00

200
N Q
...(9.90)
100 Eq. 9.90 is valid for 100 < Re < 10 7, and Re.Pr > 0.2 and
correlates very well all available data. Fluid properties are
0
40 80 120 160
evaluated at ‘film temperature’, Tf = (Ts + Ta)/2 = average of
q = Degrees from stagnation point
surface and free stream temperatures.
In the mid-range of Reynolds numbers, i.e. 20,000 < Re <
FIGURE 9.16 Circumferential Variation of the 400,000, it is suggested that following equation be used:
Heat Transfer Coefficient at High Reynolds
Numbers for a Circular Cylinder in Cross Flow
1
F F 1
IJ
1 I
(W.H. Giedt) Nucyl =
h ×D
= 0.3 + GG GH
0.62 × Re 2 × Pr 3
× 1+
Re
K
2 JJ
LM F I OP H K
1
k 2 4
28200

MM GH JK PP
0. 4 3
1+
Pr
N Q ...(9.91)
for 20,000 < Re < 400,000, and Re ×Pr > 0.2
Below Pe = (Re.Pr) = 0.2, following relation is recommended by Nakai and Okazaki:

F F 1 II -1

= G 0.8237 - ln G Pe JJ JJ
Nu cyl
GH GH 2
KK
(for Pe < 0.2...(9.92))

For Eqs. 9.91 and 9.92 also, properties are evaluated at the film temperature.
For heat transfer from a single cylinder in cross flow, for liquid metals, following relation is recommended
by Ishiguro et. al.:
Nu cyl = 1.125×(Re×Pr)0.413 (for 1 < Re.Pr < 100...(9.93))
However, note that Eq. 9.90 is quite comprehensive and is also valid for liquid metals.
For circular cylinder in cross flow, for gases, following relation is widely used:
1
Nu = C×Re n × Pr 3 ...(9.94)
where, values of C and n are given in Table 9.5:
TABLE 9.5 Values of C and n in Eq. 9.94
Re C n
0.4–4 0.989 0.330
4–40 0.911 0.385
40–4,000 0.683 0.466
4,000–40,000 0.193 0.618
40,000–400,000 0.0266 0.805

434 FUNDAMENTALS OF HEAT AND MASS TRANSFER


All fluid properties are taken at film temperature.
For non-circular cylinders:
Again, Eq. 9.94 is applicable.
For non-circular cylinders, Fig. 9.17 below, gives the values of C and n to be used in Eq. 9.94. This figure also
shows the characteristic dimension ‘D’ used to calculate the Reynolds number, for each geometry.
Flow across spheres:
For gases, McAdams recommends following relation:
Nusph = 0.37×Re0.6 (for 25 < Re < 100,000...(9.95))
For flow of liquids past spheres, Kramers suggests following relation:
Nu sph ×Pr–0.3 = 0.97 + 0.68×Re0.5 (for 1 < Re < 2000...(9.96))
In Eq. 9.95 and 9.96, fluid properties are evaluated at film temperature.
A comprehensive equation for gases and liquids flowing past a sphere is given by Whitaker:
F 1 2 I Fm I
GG
Nu sph = 2 + 0. 4 × Re 2 + 0.06 × Re 3 × Pr 0 . 4 × JJ GH m JK
a ...(9.97)
H K w

Eq. 9.97 is valid for: 3.5 < Re < 80,000 and 0.7 < Pr < 380. Here, fluid properties are evaluated at free stream
temperature.
A special case is that of heat and mass transfer from freely falling liquid drops and the following correla-
tion of Ranz and Marshall is applicable:
1 1
Nu avg = 2 + 0.6× Re 2 × Pr 3 ...(9.97a)

Re C n
3 5
U D 5 ¥ 10 – 10 0.246 0.588

Square

3 5
U D 5 ¥ 10 – 10 0.102 0.675

Square
3 4
5 ¥ 10 – 1.95 ¥ 10 0.16 0.638
U D 4 5
1.95 ¥ 10 – 10 0.0385 0.782
Hexagon

3 5
U D 5 ¥ 10 – 10 0.153 0.638

Hexagon

3 4
U D 4 ¥ 10 – 1.5 ¥ 10 0.228 0.731

Vertical plate
3 4
U D 2.5 ¥ 10 – 1.5 ¥ 10 0.224 0.612

Ellipse

3 4
U D 3 ¥ 10 – 1.5 ¥ 10 0.085 0.804

Ellipse

FIGURE 9.17 Constants C and n for cross flow over noncircular cylinders

FORCED CONVECTION 435


For heat transfer from a sphere to a liquid metal, following correlation is recommended:
1
Nu sph = 2 + 0.386×(Re×Pr) 2 (for 36,000 < Re < 200,000...(9.98))
In Eq. 9.98, fluid properties are to be evaluated at film temperature.
9.9.2 Flow Across Bluff Objects
Normal flat plate (width D):
2
NuD = 0.20×ReD 3 (for 1 < Re < 4 ´ 10 5...(9.99))
Half round cylinder of dia. D, with flat surface at rear:
2
NuD = 0.16×ReD 3 (for 1 < Re < 4 ´ 105...(9.100))
In Eqs. 9.99 and 9.100, fluid properties are evaluated at film temperature.
9.9.3 Flow Through Packed Beds
Here, a gas or liquid flows through a bed packed with solid particles (such as spheres, cylinders or commercial
packings like Raschig rings, ceramic saddles etc.). During the ‘charging cycle’, the hot fluid, while passing
through the bed, gives up its ‘heat’ to the solid particles, and during the ‘discharge cycle’, the incoming cooler
fluid picks up the stored heat from the solid particles.
Packed beds are used in catalytic reactors, grain dryers, storage of solar thermal energy, gas chromatogra-
phy, regenerators and desiccant beds.
Reynolds number in the correlations is based on a ‘superficial velocity’ Us , i.e. the fluid velocity that would
exist if the bed were empty. Characteristic length used is the equivalent diameter of the packing, Dp. Another
parameter that appears in some correlations is the void fraction, e, i.e. the fraction of bed volume that is empty.
Whitaker recommends following relation for heat transfer between the gas and packings (including cylin-
ders with diameter equal to height, spheres, or several types of commercial packings such as Raschig rings,
partition rings or Berl saddles):
F 1 2 I 1
ha × Dp
=
1- e
GG JJ
× 0. 5 × ReD2 p + 0.02 × ReD3 p × Pr 3 ...(9.101)
k e H K
where ha is the average heat transfer coefficient
Eq. 9.101 is valid for: 20 < ReDp < 10,000, and 0.34 < e < 0.78.
Packing diameter Dp is defined as six times the volume of the particle divided by the particle surface area;
for a sphere, Dp = diameter of sphere. All properties are evaluated at bulk fluid temperature (One may use the
average of inlet and outlet temperature of the heat exchanger). In the above correlation, Reynolds number is
defined as:
Dp ×U s
ReDp =
n × (1 - e )
Eq. 9.101 does not correlate data well for cube packings.
To determine the heat transfer from the wall of a packed bed to a gas, Beek recommends the following
relation, for particles like cylinders, which can pack next to the wall:
1 1
ha × Dp 0. 8
0. 4
3 × Pr 3 + 0.094× Re
= 2.58× ReDp Dp × Pr ...(9.102)
k
and, for particles like spheres, which contact the wall at one point:
1 1
ha × Dp 0. 8
0. 4
3 × Pr 3 + 0.220× Re
= 0.203× ReDp Dp × Pr ...(9.103)
k
In Eqs. 9.102 and 9.103, properties of fluid are evaluated at the film temperature. Also, the Reynolds number
is:
U s × Dp
40 < Re Dp =
< 2000
n
436 FUNDAMENTALS OF HEAT AND MASS TRANSFER
where Dp is the diameter of sphere or cylinder. For other types of packings, Whitaker’s definition of Dp may be
used.
Beek also gives the correlation for the friction factor:

Dp F I
f=
L
×
Dp
r ×U s2
=
1- e
e 3 GH
× 1.75 + 150×
1-e
ReDp
JK ...(9.104)

where Dp is the pressure drop over a length L of the packed bed.


Example 9.12. Air at 35°C flows across a cylinder of 50 mm diameter at a velocity of 50 m/s. The cylinder surface is
maintained at 145°C. Find the heat loss per unit length. Properties at mean temperature of 90°C are: r = 1 kg/m3, m = 20
´ 10 – 6 kg/(ms), k = 0.0312 W/(mC), Cp = 1.0 kJ/(kgC).
Use the relation: Nu D = 0.027.(ReD) 0.805.(Pr)1/3 [M.U.]
Solution.
Data:
Ta + Ts
Ts := 145°C Ta := 35°C V := 50 m/s D := 0.05 m L := 1 m (assumed) Tf := i.e. Tf = 90°C
2
Properties at Tf :
m := 20 ´ 10 –6 kg/(ms) k := 0.0312 W/(mC) Cp := 1000 J/(kgC) r := 1 kg/m3 A := p×D×L m2
i.e. A = 0.157 m2
Reynolds number:
D ×V ×r
Re :=
m
i.e. Re = 1.25 ´ 10 5
Prandtl number:
Cp × m
Pr :=
k
i.e. Pr = 0.641
Nusselts number:
We have:
1
Nu D : = 0.027×Re 0.805 ×Pr 3
i.e. Nu D = 295.122
Therefore, heat transfer coefficient:
k × NuD
h :=
D
i.e. h : = 184.156 ...W/(m2 C).
Heat transferred, Q:
Q := h×A×(Ts – Ta)
i.e. Q = 3.182 ´ 103 W
Example 9.13. A hot wire probe is 5 mm in length, 10 mm diameter wire with an electrical resistance of 150 ohms/m. The
wire is maintained at a constant temperature of 50°C. If the the probe is kept in an air stream flowing at a velocity of 10
m/s and at 1 bar and 25°C, determine the current required to maintain the wire temperature at 50°C.
Solution.
Data:
Ta + Ts
Ts := 50°C Ta := 25°C U := 10 m/s D := 10 ´ 10 –6 m L := 0.005 m Tf := i.e. Tf = 37.5°C
2
Properties at Tf :
n := 16.7 ´ 10 –6 m2/s k := 0.02704 W/(mK) Pr := 0.706
Reynolds number:
U ×D
Re := i.e. Re = 5.988
n
Therefore, Re×Pr = 4.228

FORCED CONVECTION 437


Since Re.Pr > 0.2, we can use the correlation of Churchill and Bernstein, viz.

LM OP
MM 1 1
L F Re I OP
4 PP
MM0.3 + 0.62 × Re × Pr × M1 + G PP
5 5

MM H 28200 JK
2 3 8
Nu cyl: =
MM LM F 0.4 I OP PP PP
...(9.90)
N Q
1
2 4

MM MMN1 + GH Pr JK PPQ
3

PP
N Q
i.e. Nucyl = 1.491 (Nusselt number)
Therefore, heat transfer coefficient:
Nucyl × k
h :=
D
i.e. h = 4.031 ´ 10 3 W/(m 2C) (heat transfer coefficient)
Heat transferred Q:
Q := h×(p×D×L)×(Ts – Ta) W
i.e. Q = 0.016 W (heat dissipated = 16 mW)
This is also equal to the value of electrical power dissipated; Q = I2.R
R := 150.0.005 ohms (electrical resistance of the wire)
i.e. R = 0.75 ohms
Therefore, current flow required:
Q
I := Amp
R
i.e. I = 0.145 A (current flow required.)
Alternatively:
To calculate Nu we can also use Eq. 9.94:
1
Nu = C×Re n ×Pr 3 ...(9.94)
Then, for circular cylinder, we get for Re = 5.988, from the Table 9.5:
C := 0.911 and n = 0.385
1
Therefore, Nu := C×Re n ×Pr 3
i.e. Nu = 1.616
And,
Nu × k
h :=
D
i.e. h = 4.369 ´ 10 3 W/(m 2C) (heat transfer coefficient)
Therefore, Q: = h×(p×D×L)×(Ts – Ta) W
i.e. Q = 0.017 W (heat dissipated = 17 mW)
This value is almost the same as obtained by the correlation of Churchill and Bernstein.
Therefore, current flow required:

Q
I := Amp
R
i.e. I = 0.151 A (current flow required.)
Example 9.14. Air at 25°C flows across an elliptical tube 6 cm ´ 12 cm size, perpendicular to the minor axis with a
velocity of 3 m/s. Tube surface is maintained at 55°C. Determine the value of convection coefficient.
Solution.
Data:
Ta + Ts
Ts := 55°C Ta := 25°C U :=3 m/s D1: = 0.06 m D2 := 0.12 m Tf := i.e. Tf = 40°C
2
Properties at Tf = 40°C:

438 FUNDAMENTALS OF HEAT AND MASS TRANSFER


n := 17.6 ´ 10 –6 m2/s k := 0.0265 W/(mK) Pr := 0.71
Reynolds number:
See Fig. 9.17 for the case of flow across an ellipse.
U × D1
Re := i.e. Re = 1.023 ´ 10 4
n
Then, we use Eq. 9.94, viz.
1
Nu = C×Re n × Pr 3 ...(9.94)
Values of C and n are obtained from Fig. 9.17 as:
C := 0.224 and, n: = 0.612
1
Therefore, Nu := C×Re n × Pr 3
i.e. Nu = 56.838
Heat transfer coefficient:
Nu × k
Therefore, h=
D1
i.e. h = 25.104 W/(m2 K) ...heat transfer coefficient.
Example 9.15. In a packed bed heat exchanger, air is heated from 40°C to 360°C by passing it through a 10 cm diameter
pipe, packed with spheres of 8 mm diameter. The flow rate is 18 kg/h. Pipe surface temperature is maintained at 400°C.
Determine the length of bed required.
Solution.
Data:
Ts := 400°C Tin := 40°C Tout : = 360°C D: = 0.008 m dpipe : = 0.1 m
18
m air := i.e. mair = 5 ´ 10 – 3 kg/s
3600
Average air temperature = (40 + 360)/2 = 200°C
Therefore, average film temperature = (200 + 400)/2 = 300°C
Taking properties of air at 300°C:
r := 0.596 kg/m 3 Cp := 1047 J/(kgK) k := 0.0429 W/(mK) n := 49.2 ´ 10 –6 m 2/s Pr := 0.71
Equivalent particle diameter = 6 ´ volume/surface area = D for a sphere
i.e. Dp := 0.008 m (equivalent particle diameter.)
Therefore, superficial velocity:
mair
Us :=
p × dpipe

4
i.e. Us = 1.068 m/s (superficial velocity)
Reynolds number:
Therefore,
U s × Dp
ReDp :=
n
i.e. ReDp = 173.684
Nusselts number:
We use Eq. 9.103, viz.
h a × Dp 1 1
3 × Pr 3 + 0.220× Re 0 . 8 ×Pr 0.4
:= 0.203× ReDp ...(9.103)
Dp
k
U s × Dp
for 40 < ReDp = < 2000
n
h a × Dp
i.e. = 12.888
k

FORCED CONVECTION 439


Heat transfer coefficient
12. 888 × k
ha := W/(m2 K) (heat transfer coefficient)
Dp
i.e. h a = 69.112 W/(m2K) (heat transfer coefficient)
Now, heat gained by air, Q = heat transfer between the wall surface and air
i.e. Q := mair ×Cp ×(Tout – Tin) W
i.e. Q = 1.675 ´ 10 3 W (heat gained by air)
This should be equal to heat transfer between the wall surface and air = ha x pipe surface area x LMTD
Here, LMTD is the ‘log mean temperature difference’ between the pipe surface and the air stream. Since the tem-
perature of air stream goes on changing along the length of heat exchanger, we use a mean temperature difference
between the pipe surface and this air stream, given by LMTD.
LMTD is defined as follows:
D Tmax - D Tmin
LMTD =
F DT I (see chapter on Heat exchangers for derivation of this equation for LMTD)
ln GH DT JK
max

min

DTmax := Ts – Tin i.e. DTmax = 360°C


DTmin := Ts – Tout i.e. DTmin = 40°C
D Tmax - D Tmin
Therefore, LMTD :=
F DT I
ln GH DT JK
max

min

i.e. LMTD = 145.638°C


Now, writing the heat balance, Q = ha X pipe surface area x LMTD , we get:
Q = ha ×(p×dpipe×L)×LMTD
where L is the length of pipe (= height of bed)
Q
i.e. L: = m (length of bed)
ha × p × dpipe × LMTD
i.e. L = 0.53 m (length of bed.)

9.9.4 Flow Across a Bank of Tubes


Flow across a bank of tubes is practically a very important case. In many industrial heat exchangers, one of the
fluids flows inside the tubes in a shell and the second fluid flows through the shell, across the tubes. Typical
applications are: in water tube boilers where water flows through the tubes and hot flue gases flow across these
tubes, waste heat recovery systems, air conditioning applications and common ‘shell and tube’ heat exchangers
used in numerous industrial applications.
Tubes in a tube bank may be arranged either in an ‘in-line’ configuration or in a ‘staggered’ configuration, as
shown in Fig. 9.18. In the figure, SL is the ‘longitudinal pitch’, ST is the ‘transverse pitch’ and SD is the ‘diagonal
pitch’.
Zhukauskas (1972) proposed the following correlation for Nusselts number, based on a large amount of
experimental data:

ha × D F I 0. 25
Nu a =
k
= C×(ReD)m ×Pr0.36 × GH JK
Pr
Prw
...(9.105)

where Nua is the average Nusselts number


ha is the average heat transfer coefficient
ReD = (d r U max)/m
Pr is the bulk Prandtl number
Prw is the wall Prandtl number
And,
ST
Umax = ×U (for aligned arrangement...(9.106))
ST - D

440 FUNDAMENTALS OF HEAT AND MASS TRANSFER


SL SL

D
ST ST
U U
SD D
Ti Ti

Transverse row: 1 2 3 4 Transverse row: 1 2 3 4


(a) Aligned arrangement of tubes (a) Staggered arrangement of tubes

FIGURE 9.18 Flow across a tube bank

ST
and, U max = ×U (for staggered arrangement...(9.107))
2×(SD - D)
Note: (i) While calculating Umax for the staggered arrangement, calculate with both the Eqs. 9.106 and 9.107 and
adopt the larger value so obtained. U is the velocity of fluid as it approaches the tube bank.
(ii) For gases, Prandtl number ratio may be dropped, since it does not have much influence
(iii) All properties (except Prw) are evaluated at free stream temperature
Eq. 9.105 gives very good prediction when the number of tube rows in the bank,
N > 20, and 0.7 < Pr < 500, and 1000 < ReD_max < 2 ´ 10 6. However, the equation can be used even when
N < 20, with a correction factor applied. If N = 4, error involved in prediction is about 25%.
Eq. 9.105 takes the following forms for various flow regimes:
For Laminar flow ( i.e. 10 < ReD < 100):

F Pr I 0 . 25
0.4
Nua = 0.8× ReD × Pr 0. 36 × GH Pr JK w
(for In-line tubes...(9.108))

F Pr I 0 . 25
×G
H Pr JK
0.4
and, Nua = 0.9× ReD × Pr 0. 36 (for staggered tubes...(9.109))
w
These equations have been validated also in the range: 50 < Re D < 1000.
For transition regime (i.e. 1000 < ReD < 2 ´ 105):

F Pr I 0 . 25
Nua = 0.27× ReD
0 .63
× Pr 0 . 36 × GH Pr JK
w
(for In-line tubes, ST/SL > 0.7...(9.110))

Note: ST/SL < 0.7 for in-line tubes, gives very ineffective heat exchanger and should not be used.

FS I 0.2
F Pr I 0 . 25
Nua = 0.35× GH S JK
T
L
× ReD
0 .60
× Pr 0. 36 × GH Pr JK
w
(for staggered tubes, ST/SL < 2...(9.111))

F Pr I 0 . 25
and, Nua = 0.40× ReD
0 .60
× Pr 0 . 36 × GH Pr JK
w
(for staggered tubes, ST/SL greater than or equal to 2...(9.112))

For turbulent regime (i.e. ReD > 2 ´ 105):

FORCED CONVECTION 441


F Pr I 0. 25
Nua = 0.021× ReD
0 .84
× Pr 0 . 36 × GH Pr JK w
...for In-line tubes...(9.113))

F Pr I 0. 25
×G
H Pr JK
0 .84
Nua = 0.022× ReD × Pr 0 . 36 (for staggered tubes, Pr > 1...(9.114))
w
0. 84
and, Nua = 0.019× ReD (for staggered tubes, Pr = 0.7...(9.115))
For staggered arrangement, with ST/D = 2 and SL /D = 1.4, we have the relation due to Achenbach:
0 .883
Nua = 0.0131× ReD ×Pr 0.36 ...(9.116)
5 6
Eq. 9.116 is valid in the range 4.5 ´ 10 < Re D < 7 ´ 10 .
If the number of tube rows is < 20, a correction factor is applied to the calculated Nusselts number as
follows:
Nu a_N = Nua ×C 2 ...(9.117)
where Nua_N is the Nusselts number for the actual tube bank with N < 20, and
Nua is the value of Nusselts number calculated for N > 20, using one of the appropriate relations
given above
C 2 is the correction factor taken from Table 9.6.

TABLE 9.6 Correction factor C2 in Eq. 9.117 for N < 20

N 1 2 3 4 5 7 10 13 16

Alinged 0.70 0.80 0.86 0.90 0.92 0.95 0.97 0.98 0.99
Staggered 0.64 0.76 0.84 0.89 0.92 0.95 0.97 0.98 0.99

Pressure drop: Pressure drop (in Pascals) for flow of gases over a bank of tubes is given by:
2
2 × f ¢×Gmax ×N mw F I 0.14
Dp =
r
×
mb GH JK Pa...(9.118).

where Gmax = mass velocity at minimum flow area = r.Umax


r = density, evaluated at free stream conditions
N = number of transverse rows
mb = average free stream viscosity
Friction factor, f‘ is given by:

LM OP
f ¢ = M0.25 + PP× Re
0.118
MM L (S - D O
- 0 .16
D (for staggered tubes...(9.119))
PP
1. 08

MN MN D PQ
T
Q
and,

LM FS I OP
M 0.08 × G J PP × Re
H DK
L

f ¢ = M0.044 +
- 0.15
(for in-line tubes...(9.120))
MM LM (S - D OP T
0 . 43 + 1.13 ×
D
SL
PP D

NM N D Q PQ

442 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Example 9.16. Air at 1 bar and 20°C flows across a bank of tubes 10 rows high and 4 rows deep; air velocity is 8 m/s,
measured at the entry to the tube bank. Diameter of the tubes is 25 mm and surface temperature of the tubes is main-
tained at 80°C. Tubes are arranged in an in - line manner. ST = SL = 37.5 mm. Calculate the total heat transfer per unit
length of the tube bank, and also the exit air temperature. Also, find out the pressure drop.
Solution.
Data:
Ts := 80°C Ti := 20°C U := 8 m/s ST := 0.0375 m SL := 0.0375 m D := 0.025 m
Reyonlds number:
This is based on Umax . We have:
ST
Umax := ×U (for aligned arrangement...(9.106))
ST - D
i.e. Umax = 24 m/s (maximum velocity)
Taking properties of air at free stream temperature of 20°C:
r := 1.164 kg/m3 Cp := 1012 J/(kgK) k := 0.0251 W/(mK) n := 15.7 ´ 10 –6 m2/s Pr := 0.71
U ×D
Therefore, ReD := max
n
i.e. ReD = 3.822 ´ 10 4 (Reynolds number)
Nusselts number:
Since Reynolds number is between 1000 and 200,000 which is in the transition regime, the appropriate equation for
average Nusselts number is:
F Pr I 0 . 25

Nua = 0.27× ReD0.63 ×Pr0.36 × GH Pr JK


w
...for in-line tubes, ST/SL > 0.7...(9.110)

The last term, i.e. the ratio of Prandtl numbers can be neglected for gases: So, we have:
Nua := 0.27 ×ReD0.63 ×Pr0.36
i.e. Nua = 183.923 (Nusselts number)
Therefore, average heat transfer coefficient is:
Nua × k
ha := W/(m2C) (average heat transfer coefficient)
D
2
i.e. h a = 184.658 W/(m C) (average heat transfer coefficient (N > 20))
This is the value of heat transfer coefficient that would be obtained if there were 20 rows of tubes in the direction of
flow. But, in the present case, there are only 4 rows in the direction of flow. So, from the Table, we get the correction
factor as:
C 2 = 0.90
Therefore, actual heat transfer coefficient is = 184.658 ´ 0.9 (actual average heat transfer coefficient)
i.e. ha = 166.193 W/(m2 C) (actual average heat transfer coefficient)
Surface area for heat transfer for unit length of tubes is:
A := (10 ×4)×(p×D×1) m 2/m (for 10 rows high, 4 rows deep)
i.e. A = 3.142 m 2/m.
Total heat transfer rate Q:
Now, total heat transfer rate is given by Newton’s law:
Q = ha ×A×D×T
Here DT is the average temperature difference between the wall and the air stream. However, temperature of air
stream goes on changing from entry to exit in the heat exchanger. So, we use a ‘mean temperature difference’ called
LMTD (log mean temperature difference). Expression for LMTD is derived in the chapter on heat exchangers. For the
present, let us take for LMTD:
(Ts - Ti ) - (Ts - To )
LMTD =
FT -T I
ln GH T - T JK
s

s
i

We need the exit temperature To of the air stream. This is calculated by a heat balance:
Q = ha ×A.(LMTD)
mass_flow: = r×U×10×ST ...kg/s (mass flow rate; 10 rows high ST is transverse distance)

FORCED CONVECTION 443


i.e. mass_flow = 3.492 kg/s
Then, we can write the heat balance:
h×A×LMTD = mass_flow×Cp × (To – Ti )
Substitute for LMTD and solve for To .
Use Solve block of Mathcad; assume a guess value for To to start with, say To = 70°C. Then type ‘Given’ and write
the constraint; then type Find (To) and get the answer:
To := 70 (guess value)
Given
(Ts - Ti ) - (Ts - To )
ha ×A×
FT -T I = mass_flow×Cp × (To – Ti )
ln GH T - T JK
s

s
i

Find (To) = 28.241


i.e. To = 28.241°C (exit air temperature)
Therefore, heat transfer rate, Q:
(Ts - Ti ) - (Ts - To )
Q := ha ×A×
FT -T I
ln GH T - T JK
s
s i

i.e. Q = 2.912 ´ 10 4 W/m = 29.12 kW/m


Alternatively, we can use the arithmetic average value of air stream between the inlet and outlet temperature, since
this is simpler to calculate and error involved will not be much:
LM FG To + T IJ OP = mass_flow×C ×(T
N H 2 KQ
i
Then, Q = ha × A× Ts - p o – Ti)

Using Solve block as earlier, to obtain To:


To := 70 (guess value)
Given
LM FG To + T IJ OP = mass_flow×C ×(T
N H 2 KQ
i
Q = ha × A× Ts - p o – Ti)

Find (To) = 28.255


i.e. To = 28.255°C (exit air temperature)
i.e. we get practically the same value for To as obtained earlier.

LM FG T + T IJ OP W/m
N H 2 KQ
o i
And, Q := ha × A× Ts -

i.e. Q = 2.917 ´ 10 4 W/m = 29.17 kW/m.


Pressure drop:
We have:
2
2 × f ¢× Gmax ×N mw F I 0.14 Pa
Dp =
r
×
mb GH JK ...(9.118)

Gmax := r×Umax kg/s.m2 (mass velocity)


i.e. Gmax = 27.936 kg/s.m2 (mass velocity)
N =4 (number of transverse rows)

LM FS I OP
0. 08 × G J
M
f ¢ := M0 .044 +
H DK
L
PP × Re - 0 . 15
(for in-line tubes...(9.120))
MM LM S - D OP
T
0 . 43 + 1.13 ×
D
SL PP D

MN N D Q PQ
i.e. f ¢ = 0.065 (friction factor)

444 FUNDAMENTALS OF HEAT AND MASS TRANSFER


mw := 20.79 ´ 10 –6 kg/ms (dynamic viscosity of air at 80°C)
mb := 18.46 ´ 10 –6 kg/ms (dynamic viscosity of air at average free stream temperature of 24.5°C)
Therefore,
2
2 × f ¢× Gmax F I 0.14 Pa
×N mw
Dp :=
r GH JK
×
mb
i.e. Dp = 354.613 Pa = 0.003546 bar

9.10 Flow Inside Tubes


Circular tubes are the most commonly used geometry for cooling and heating applications, in industry. Often,
tubes of other geometries such as square or rectangle are also used. We are interested in heat transfer in such
cases; pressure drop occurring during flow is also of interest since it has a direct bearing on the pumping power
required to cause the flow.
Observe the major difference between the external flows just studied and the internal flow through pipes: in
the external flow, say over a flat plate, there was a free surface of fluid and the boundary layer was free to grow
indefinitely; however, in a pipe flow, the flow is confined within the pipe and the boundary layer growth is
limited to grow only upto the centre of the pipe.
9.10.1 Hydrodynamic and Thermal Boundary Layers for Flow in a Tube
Consider a fluid entering into a circular pipe, with a uniform velocity U (See Fig. 9.19). Fluid layer coming in
contact with the pipe surface comes to a complete halt and the adjacent layers slow down gradually due to
viscosity effects. Since the total mass flow in a section must remain constant, velocity in the central portion
increases. As a result, a ‘velocity boundary layer’ develops along the pipe. Thickness of the velocity boundary
layer increases along the flow length until the entire pipe is filled up with the boundary layer, as shown. ‘Hydro-
dynamic entry length (Lh)’ is the distance from the entry point to the point where the boundary layer has devel-
oped upto the centre. In the region beyond the hydrodynamic entry length, the velocity profile is fully developed
and remains unchanged; this is the ‘hydrodynamically developed region’. As will be shown later, velocity profile
in the fully developed region, in laminar flow, is parabolic; in turbulent flow, the velocity profile is a truncated
one.
Similarly, when a fluid at an uniform temperature enters a pipe whose wall is at different temperature, a
‘thermal boundary layer ‘ develops along the pipe. Thickness of thermal boundary layer also increases along the
flow length till the boundary layer reaches the centre of the pipe. ‘Thermal entry length (Lt)’ is the distance from
the entry to the point where the thermal boundary layer has reached the centre, and is shown in the Fig. 9.19.
Beyond this point, along the length, we have the ‘fully developed flow’ i.e. the flow is both hydrodynamically
and thermally fully developed.
Temperature profile may vary with x even in the thermally developed region. However, the dimensionless
temperature profile expressed as (T – Ts)/(Tm – Ts) remains constant in the thermally developed region, whether
the temperature of the pipe surface remains constant or the heat flux at the surface remains constant. (Tm is the
bulk or mean temperature at a given section).

Velocity boundary layer Velocity profile Thermal boundary layer Temperature profile
TS

U Ti

Lt
Lh
Hydrodynamic entry Hydrodynamically Thermal entry region Thermally
region developed region developed region
(a) Development of velocity boundary layer (b) Development of thermal boundary layer

FIGURE 9.19 Flow inside a pipe

FORCED CONVECTION 445


Relative growth of hydrodynamic and thermal boundary layers is controlled by the dimensionless Prandtl
number. For gases, Pr =1, and the hydrodynamic and thermal boundary layers essentially coincide; for oils Pr >>
1 and the hydrodynamic boundary layer outgrows the thermal boundary layer, i.e. hydrodynamic entry length is
smaller for oils. For fluids with Pr << 1, such as liquid metals, thermal boundary layer outgrows the hydrody-
namic boundary layer and consequently, the thermal entry length is shorter than the hydrodynamic entry length.
Reynolds number is the dimensionless number that characterizes the flow inside a tube as laminar or turbu-
lent. Reynolds number is defined as:
Re = (Um.D)/n
where Um is the mean velocity in the pipe, and n is the kinematic viscosity of the fluid. Flow regimes are defined
as follows, depending upon the Reynolds number:
Re < 2300 (Laminar flow)
2300 £ Re £ 4000 (Transition flow)
Re > 4000 (Turbulent flow)
Hydrodynamic and thermal entry lengths:
In laminar flow:
Lh_lam = 0.05×Re×D ...(9.121a)
Lt_lam = 0.05×Re×Pr×D ...(9.121b)
In turbulent flow, hydrodynamic and thermal entry lengths are independent of Re and Pr and are generally
taken to be:
Lh_turb = Lt_turb = 10×D ...(9.122)
The friction coefficient or shear stress at the surface is related to the slope of the velocity profile at the
surface. Since the velocity profile remains essentially constant in the hydrodynamically developed region, the
friction factor and the shear stress remain constant in the hydrodynamically developed region. By a similar argu-
ment, heat transfer coefficient also remains constant in the thermally developed region.
At the entry to the tube, thickness of the boundary layer is practically zero; so velocity and temperature
gradients at the surface are almost infinite at the entry, which means that the heat transfer coefficient and pres-
sure drop are the highest in the entry region and go on decreasing along the length.
Generally, in practice, turbulent flows prevail in heat transfer applications; length of pipes is also generally
much larger as compared to the hydrodynamic and thermal entrance lengths. Therefore, flow through pipes is
generally assumed to be fully developed over the entire length.
9.10.2 Velocity Profile for Fully Developed, Steady, Laminar Flow
Consider a fully developed, steady, laminar flow in a pipe. Consider a fluid element of length L and radius r, as
shown in Fig. 9.20.
We are interested to get the velocity profile and the pressure drop (or friction factor) during flow. This is
obtained by making a force balance on a cylindrical fluid element as shown in Fig. 9.20. Forces acting on the
element are: pressure forces at the ends and the shear forces on the surface; there is no change in momentum
since the velocities are same at both sections 1 and 2. So, writing a force balance:
(p 1 – p 2)×p×r 2 = t×(2×p ×r×L) ...(a)
du
But, t = – m× ...(b)
dr
2 2
(u/umax) = [1 – (r /R )]

u
t R
umax r
p1 p2
t

FIGURE. 9.20 Laminar flow through a pipe

446 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(negative sign, since r is measured opposite to the direction of y).
So,
du - ( p1 - p2 )
= ×r ...(c)
dr 2 × m ×L
Separating the variables and integrating,

i.e.
z 0
1 du =
u
- ( p1 - p2 ) R
2 × m ×L
× rdr
r z ...(d)

1 ( p1 - p2 )
u= × ×(R2 – r2) ...(e)
4 ×m L
This can also be written as:
- 1 dp
u= × ×(R2 – r2) ...(9.123)
4 × m dx

dp - ( p1 - p2 )
since, in differential form, =
dx L
Negative sign in Eq. 9.123 indicates that pressure decreases in the flow direction. Also, note that the velocity
profile is parabolic.
Now, maximum velocity occurs at r = 0, i.e. at the centre:
- 1 dp 2
i.e. u max = × ×R ...(9.124)
4 × m dx
Eq. 9.124 gives the maximum velocity in the pipe.
From Eqs. 9.123 and 9.124, we get:

u FG r IJ 2

umax
=1–
H RK ...(9.125)

Average or mean velocity, um, is obtained by equating the volumetric flow to the integrated paraboloidal

z
flow:
R
um ×p×R 2 = u ×( 2 × p × r ) dr
0

umax - 1 dp 2
i.e. um = = × ×R ...(9.126)
2 8 × m dx
Now, friction factor is defined by:
2
- dp f r × um
= × ...(9.127)
dx D 2
2
r ×um
where D is the pipe diameter and is the dynamic pressure.
2
Integrating Eq. 9.127, we get, ‘Darcy – Weisbach equation’ for pressure drop:

Dp f r × m 2m
= × ...(9.128)
L D 2
where Dp = p 1 – p 2 and, L = x 2 – x1
From Eqs. 9.126 and 9.127, we get:
64
f= ...(9.129)
ReD

FORCED CONVECTION 447


Eq. 9.129 gives the friction factor for laminar flow (Re < 2000), in a pipe flow.
Since volumetric flow rate, Q = A.um, we can write for head loss:
Dp Q ×L× m
hL = = 128 × ...(9.130)
r p × D4 × r

p R4
Or, Q= × ×(p1 – p 2) ...(9.131)
8×m L
Eq. 9.131 is known as ‘Hagen – Poiseuille equation’.
Darcy-Weisbach Eq. 9.128 is applicable to non-circular ducts also, if D is replaced by ‘hydraulic diameter
(Dh)’, defined by:
4×A
Dh = ...(9.132)
P
where A is the area of cross-section and P is the wetted perimeter.
Values of product of friction factor and Reynolds number for two important duct configurations (viz. annu-
lar ducts and rectangular ducts) are given Tables 9.7 and 9.8 below:

TABLE 9.7 Annular ducts

Ratio of radii f.Re


0.001 74.68
0.01 80.11
0.05 86.27
0.10 89.37
0.20 92.35
0.40 94.71
0.60 95.59
0.80 95.92
1.00 96.00

TABLE 9.8 Rectangular ducts

Ratio of sides f.Re


0.05 89.91
0.10 84.68
0.125 82.34
0.166 78.81
0.25 72.93
0.40 65.47
0.50 62.19
0.75 57.89
1.00 56.91

9.10.3 Heat Transfer Considerations in a Pipe


Most of the practical cases of heat transfer involving a pipe flow fall under two categories:
(a) surface heat flux on the pipe is constant, e.g. when the pipe is subjected to radiation or heated electrically
by winding an electric tape, or
(b) pipe surface temperature is constant, e.g. when there is condensation or boiling occurring on the surface
of the pipe.

448 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(a) Constant surface heat flux, qs:
Let a fluid enter a pipe subjected to a constant surface heat flux qs, with a mean inlet temperature of Ti and let the
mean exit temperature of fluid be Te . Then, the heat transfer rate can be written as:
Q = qs ×A = m×Cp ×(Te – Ti) ...(9.133)
Or, the mean exit temperature of fluid may be written as:
qs × A
Te = Ti + ...(9.134)
m× Cp
where A is the surface area of the pipe, m is the mass flow rate of fluid and Cp is its mean specific heat. Mean
fluid temperature, Tm increases linearly in the flow direction. The surface temperature is determined from:
qs = h×(Ts – Tm) ...(9.135)
When h is constant, for constant surface heat flux, (Ts – Tm) is constant, i.e. the surface temperature also
increases linearly in the flow direction. This situation is shown graphically in Fig. 9.21:
Now, we are interested to get the temperature profile and the heat transfer coefficient during flow. This is
obtained by making an energy balance on a cylindrical fluid element shown in Fig. 9.22. Here, the surface heat
flux along the length is constant, i.e.
dqs
=0
dx
Heat flows to be considered are: conduction in and out of the element at the ends and the heat convected in
and out by virtue of flow.

T
Entry region Fully developed region

dx
Ts

Te
Tm
qr
DT = qs/h
dr

x qr+dr
0 L r2pr×duCp×T

FIGURE 9.21 Tube surface and mean fluid tem- FIGURE 9.22 Control element for energy balance
peratures for a pipe with constant surface heat flux in pipe flow

So, writing an energy balance:


Heat flow into the element by conduction =
dT
dQr = – k×2×p×r×dx×
dr
Heat flow out of the element by conduction =
F dT + d T × drI
2
dQr + dr = – k×2×p ×(r + dr)×dx× GH dr dr JK 2

Net heat convected out of the element is:


dT
dQConv = 2×p×r×dr×r ×Cp×u× ×dx
dx
By energy balance:
Net energy convected out = net energy conducted in

FORCED CONVECTION 449


i.e. dQConv = dQr – dQr + dr
Substituting for the above terms and simplifying neglecting higher order differentials,
1 d FG
dT IJ 1 dT
we get: ×
u × r dr

H
dr K = ×
a dx
...(9.136)

As already discussed, with constant heat flux at the wall, average fluid temperature must increase linearly
with x, so that dT/dx = constant i.e. temperature profiles will be similar at different locations along the length.
To solve Eq. 9.136, we have to insert the expression for the velocity profile given by Eq. 9.125, with the
boundary conditions:
dT
= 0 at r = 0
dr
FG dT IJ
and k×
H dr K r=R
= qs = constant

So, Eq. 9.136 becomes:

d FG
dT IJ 1 dT r2 F I
dr

H
dr K = ×
a dx R
GH
× umax × 1 - 2 × r JK
Integrating,

dT 1 dT r2 r4 F I +C

dr
= ×
a dx
× umax ×
2
-
4 × R2
GH JK 1

Integrating again,

1 dT r2 F
r4 I + C ×ln(r) + C
T= ×
a dx
× umax × - GH
4 16 × R 2 JK 1 2

Applying the first B.C., we get: C 1 = 0. Also, T = Tc at r = 0, at centre of the pipe, i.e. C2 = TC
Therefore, temperature distribution in terms of temperature at the centre of the pipe is:
2 LMF r I 2
FG IJ OP 4

MNGH R JK
1 dT umax × R 1 r
T – Tc = ×
a dx
×
4
× - ×
H K PQ
4 R
...(9.137)

Bulk temperature:
For convection heat transfer in a pipe, we have:
local heat flux, q = h×(Ts – Tb) ...(9.138)
where Ts is the wall temperature, and Tb the ‘bulk temperature’, which is an energy averaged temperature across
the pipe, calculated from:

zR
r × 2 ×p × r × u × Cp ×Tdr

z
Tb = 0 ...(9.139)
R
r × 2 ×p × r × u × Cp × dr
0
Again, we have already shown that bulk temperature is a linear function of x for constant heat flux at the
wall. Performing the calculation in Eq. 9.139, (using Eq. 9.137), we get:

7 umax × R2 dT
T b = Tc + × × ...(9.140)
96 a dx
And, wall (or, surface) temperature is given by:
2
3 umax × R dT
Ts = Tc + × × ...(from eqn. 9.137, with r = R)...(9.141)
16 a dx

450 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Now, the heat transfer coefficient is given by:

FG dT IJ
Q = h×A×(Ts – Tb) = k×A×
H dr K r=R

FG dT IJ

H dr K r=R
i.e. h= ...(9.142)
Ts - Tb
Now, the numerator in Eq. 9.142 is the temperature gradient and is given by:

FG dT IJ umax dT r r3 F I umax × R dT
H dr K r=R
=
a
× × -
dx 2 4 × R 2GH JK =
4 ×a
×
dx
...(9.143)
r=R
Substituting Eqs. 9.140, 9.141 and 9.143 in 9.142, we get
24 k 48 k
h= × = × ...(9.144)
11 R 11 D
Or, in terms of Nusselts number:
h ×D
NuD = = 4.364 ...(9.145)
k
Note the interesting result that for, steady, fully developed laminar flow in a pipe whose walls are subjected
to a constant heat flux, the Nusselts number is a constant = 4.364. Of course, at the entrance region, value of
Nusselts number will be somewhat higher.
(b) Constant surface temperature, Ts:
Let a fluid enter a pipe whose surface is maintained at a constant temperature Ts , with a mean inlet temperature
of Ti and let the mean exit temperature of fluid be Te. Then, the mean temperature of the fluid Tm approaches the
surface temperature asymptotically, as shown in Fig. 9.23.
Now, the temperature of the surface is constant and the fluid temperature varies continuously from Ti at the
inlet to Te at the exit. To determine the heat transfer rate, we have the Newton’s rate equation, Q = hA DTm, where
DTm is a mean temperature difference between the surface and the fluid. In the chapter on heat exchangers, it will
be shown that this mean temperature difference, also known as ‘log mean temperature difference (LMTD)’, is
given as:
D Te - D Ti
DT m = LMTD =
F DT I
ln G
...(9.146)

H DT JK
e
i

where, DTi and DTe are the temperature differences at the inlet and outlet, as shown

- h×A F I
Also,
D Te
D Ti
= exp
m × Cp
GH JK ...(9.147)

Here, m is the mass flow rate (kg/s), A is the area of heat transfer and Cp is the specific heat of the fluid.
From Eq. 9.147, one can calculate the mean fluid temperature at the exit. The term h.A/(m.Cp) is known as
‘Number of Transfer Units (NTU)’ and is a measure of the size of the heat exchanger.
D Te
i.e. = exp(– NTU) ...(9.148)
D Ti
By making an analysis similar to the one as we did in the case of constant heat flux at the walls, we can show
that for the case of constant wall temperature, for steady, laminar flow, the Nusselts number is a constant, given
by:
h ×D
NuD = = 3.656 ...(9.149)
k

FORCED CONVECTION 451


T Nud

Ts = constant
DTe
Uniform heat flux

DTi Tm Constant wall temperature


DT = Ts – Tm
4.364
3.656
Ti

x (x/D)/(Re.Pr)
0 L
FIGURE 9.23 Variation of mean fluid temperature FIGURE 9.24 Variation of Nusselts number
for a pipe with constant surface temperature with (x/D)/(Re.Pr), for laminar flow in a pipe

Again, note that this is for fully developed flow and in the entrance region the values will be higher.
Nature of variation of Nusselts number with the dimensionless number (x/D)/(Re.Pr) is shown in the fol-
lowing graph (Fig. 9.24).
Note that for fully developed flows, Nusselts number approaches the asymptotic values of 4.364 and 3.656
for the cases of uniform heat flux and constant wall temperature, respectively.
For short pipes (L/D is small, < 60), with constant wall temperature, fully developed velocity profile (para-
bolic), average Nusselt number is given by Hausen as:

FG D IJ × Re × Pr
Nu avg = 3.66 +
H LK
0.0668
...Pr > 0.7...(9.150a)
LF D I O
2

1 + 0.04 × MG J × Re × Pr P
3

NH L K Q
This equation gives the average Nusselt number over the length of tube, including the entry region. Here, Re
= (D.um .r)/m. Also, in the above expression, the dimensionless group in the denominator is known as Graetz
number, i.e.
D
Gz = Re×Pr×
L
For oils, or other fluids in which viscosity varies with temperature considerably, the constant 0.0668 in equa-
tion 9.150a must be multiplied by (m/m s)0.14.
Another correlation for the above conditions is:

F I 0 . 333

Nu avg
G Re× Pr JJ
= 1.67× G (for(L/D)/(Re.Pr) < 0.01, constant wall temperature...(9.150b))
GH DL JK
In Eq. 9.150, property values are taken at mean bulk temperature. If the outlet temperature is not specified,
iterative working will be required.
Another correlation to take care of the property variations is suggested by Sieder and Tate:

F I F I
1
3
G Re × Pr JJ ×G m J
= 1.86× G
0 .14

GH DL JK H m K
Nu avg ...(9.150c)
s

452 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For short pipes (L/D is small, < 60), with constant wall temperature, velocity profile still developing, aver-
age Nusselt number is given by Hausen as:

FG D IJ × Re × Pr
Nu avg = 3.66 +
0.104 ×
H LK (Pr = 0.7...(9.150d))
LF D I
1 + 0.016 × MG J × Re × Pr P
O 0. 8

HN L K Q
For oils, or other fluids in which viscosity varies with temperature considerably, the constant 0.104 in Eq.
9.150d must be multiplied by (m/m s)0.14.
For long lengths, at constant wall temperature, Nusselt number asymptotically approaches the value 3.66.
For short pipes with constant wall heat flux, with fully developed parabolic velocity profile, Hausen’s corre-
lation for local Nusselt number is:

FG D IJ × Re × Pr
H LK
0.023 ×
Nu = 4.36 +
LF D I O
1 + 0.0012 × MG J × Re × Pr P
(Pr > 0.7...(9.151a))

NH L K Q
Another relation recommended for above conditions is:

F I 0 . 33

Nuavg
G Re× Pr JJ
= 1.30× G (for (L/D)/(Re.Pr) < 0.01, constant wall heat flux...(9.151b))
GH DL JK
For short pipes with constant wall heat flux, with developing velocity profile, Hausen’s correlation for local
Nusselt number is:

FG D IJ × Re × Pr
H LK
0.036 ×
Nu = 4.36 +
LF D I
1 + 0.0011× MG J × Re × Pr P
O (Pr = 0.7...(9.151c))

HN L K Q
For long pipes with constant wall heat flux, average Nusselt number approaches the value 4.364, as already
discussed.
9.10.4 Fully Developed Laminar Flow Inside Pipes of Non-circular Cross-sections
Nusselts number and friction factor for fully developed laminar flow inside pipes of non-circular cross-sections
are given in Table 9.9. Here, Reynolds number and Nusselts number are based on the hydraulic diameter, which
was defined earlier, as:
4×A
Dh = ...(9.132)
P
where A is the area of cross-section and P is the wetted perimeter.
Flow through an annulus: Practically important case is the flow through an annulus with the outer surface
insulated, and the inside surface maintained at either a constant temperature or constant heat flux.
In the case of an annulus, the hydraulic diameter as given by Eq. 9.132 viz.
Dh = (Do – Di). For fully developed laminar flow, Nusselt number varies with (Di/Do) as shown in Table 9.10.
Here, Nu T is the Nusselt number with the inner wall maintained at constant temperature and Nu H is the Nusselt
number with the inner surface maintained at constant heat flux. Outside surface is insulated for both the cases.
In laminar flow, surface roughness of the pipe does not have much effect on Nusselts number or friction
factor.

FORCED CONVECTION 453


TABLE 9.9 Nusselts number and friction factor for fully developed Laminar flow in pipes of various cross-
sections

Cross-section of pipe a/b or, q, deg. Nu(T s = const.) Nu(qs = const.) Friction factor, f
Circle (dia. = D) - 3.66 4.36 64/Re
Hexagon - 3.35 4.00 60.20/Re
Square - 2.98 3.61 56.92/Re
Rectangle of a/b = 1 2.98 3.61 56.92/Re
width ‘a’ and 2 3.39 4.12 62.20/Re
height ‘b’ 3 3.96 4.79 68.36/Re
4 4.44 5.33 72.92/Re
6 5.14 6.05 78.80/Re
8 5.60 6.49 82.32/Re
¥ 7.54 8.24 96.00/Re
Ellipse, of major a/b = 1 3.66 4.36 64.00/Re
axis ‘a’ and 2 3.74 4.56 67.28/Re
minor axis ‘b’ 4 3.79 4.88 72.96/Re
8 3.72 5.09 76.60/Re
16 3.65 5.18 78.16/Re
Triangle, with q = 10 1.61 2.45 50.80/Re
apex angle q = (deg.) 30 2.26 2.91 52.28/Re
60 2.47 3.11 53.32/Re
90 2.34 2.98 52.60/Re
120 2.00 2.68 50.96/Re

TABLE 9.10 Nusselt numbers for fully developed laminar flow in an annulus, insulated on the outside
Di / Do 0.05 0.10 0.25 0.50
NuT 17.46 11.56 7.37 5.74
NuH 17.81 11.91 8.5 6.58

9.10.5 Turbulent Flow Inside Pipes


9.10.5.1 Velocity profile and pressure drop. Experimental results of Nikuradse for turbulent flow in smooth pipes
indicated a power–law form for velocity profile:

F yI
1

= G J
u n
umax H RK ...(9.152)

where u is the local time-average velocity, umax is the time–average velocity at the centre, R is the radius of the
pipe and y = (R – r), is the distance from the pipe wall. Values of index n are given in Table 9.11 for different
values of Reynolds numbers:
Pressure drop for turbulent flow in pipes is also given by the Darcy – Weisbach equation i.e.
2
Dp f r × um
= × ...(9.128)
L D 2
TABLE 9.11 Values of index ‘n’ in Eq. 9.152 for turbulent flow in pipes
Re n
3
4 ´ 10 6.0
2.3 ´ 104 6.6
1.1 ´ 105 7.0
1.1 ´ 106 8.8
2 ´ 106 10.0
3.2 ´ 106 10.0

454 FUNDAMENTALS OF HEAT AND MASS TRANSFER


However, friction factor f must be determined experimentally. (Note that in case of laminar flow equation
for friction factor was derived analytically as f = 64/Re).
Average or mean velocity, um over the cross-section is easily calculated for the power-law profile as:

z R 1
( R - r ) n × m 2max × 2 ×p × rdr

z
0
um = R
2 ×p × rdr
0
Performing the integration we get the result as:

2 × n2
um = (average or mean velocity...(9.153))
( 2 × n + 1) × (n + 1)
Friction factor ‘f’ for smooth pipes is given by the following empirical relations:
f = 0.316×Re–0.25 (for 2 ´ 104 < Re < 8 ´ 10 4...(9.154))
f = 0.184×Re–0.2 (for 104 < Re < 10 5...(9.155))
–2
f = (0.79×ln(Re) – 1.64) (for 3000 < Re < 5 < 106...(9.156))
Eq. 9.156 for friction factor, developed by Petukhov, covers a wide range of Reynolds numbers.
Friction factor ‘f’ for commercial or ‘rough’ pipes is given by Colebrook’s formula (1939) or from the
Moody’s diagram. Here, surface imperfections on the internal surface extend beyond the laminar sub-layer and
are characterized by a ‘roughness height’ ‘e ’ and the ‘relative roughness’ (e/D) is a parameter in the Moody’s
diagram. See Fig. 9.25. Note that in the region of complete turbulence, friction factor is mainly dependent on the
relative roughness. Values of ‘e’ for commercial piping are given in Table 9.12.
Colebrook formula:
LMF e I + 18.7 OP
MNGH R JK Re × f PQ
1
= 1.74 – 2×log ...(9.156a)
f
Here, logarithm is to base 10. This equation is slightly difficult to calculate since f occurs on both sides of the
equation and an iterative solution will be required. Instead, following formula for f is relatively easier to calcu-
late:
1. 325
f= ...(9.156b)
F ln F e I + 5.74 I 2

GH GH 3.7 D JK Re JK 09

Losses in pipe fittings:


Fittings, valves, etc. are part of the piping system and they also offer resistance to flow of fluid. Losses through
fittings can be quite considerable in large, industrial piping systems. Generally, head loss through a valve or
fitting is expressed in the following form:
u2
h L = kL × m ...(9.157)
2
Values of ‘loss coefficient’, kL for some common valves and fittings are given in Table 9.13.
In practice, while calculating pressure drop in a piping system, for each valve and fitting, an ‘equivalent
length Leq‘is found out and added to the straight length of piping and then the Darcy – Weisbach equation is
applied. Equivalent length for a valve or fitting is calculated from:
kL × D
Leq = ...(9.158)
f
9.10.5.2 Heat transfer coefficient for turbulent flow inside pipes. Analytical treatment of turbulent flow is rather
complicated as compared to that of laminar flow; therefore, empirical relations based on extensive experimental
data have been suggested. Reynold’s analogy between momentum and heat transfer supplies the simplest corre-
lation:

FORCED CONVECTION 455


0.1
0.09
0.08
0.07 0.05
0.04
0.06
0.03

0.05 0.02
0.015
0.04
0.01

D
Relative roughness e
0.008
f, friction factor

0.006
0.03
0.004
0.025
Equation 9.155 0.002
Laminar flow
64
0.02 f= 0.001
ReD
0.0008
0.0006
0.0004
0.015
0.0002
0.0001
Laminar Transition
flow zone Complete turbulence, rough pipes 0.000.05
0.1
0.009 Critical
zone 0.000.01
0.008 3 4 5 6 7 8
10 2 3 4 5 6 78 910 2 3 4 5 6 7 8 910 2 3 4 5 6 7 8 9 10 2 3 4 5 6 7 8 910 2 3 4 5 6 7 8 9 10
e
Reynolds number ReD = ru D/u = 0.000.005
D
e
= 0.000.001
D

FIGURE 6.25 Moody’s diagram for friction factor for flow through pipes

TABLE 9.12 Roughness height ‘e’ for commercial piping

Type of piping e, mm

Drawn tubing 0.0015


Brass, lead, glass, spun cement 0.0075
Commercial steel or wrought iron 0.05
Cast iron (asphalt dipped) 0.12
Galvanized iron 0.15
Wood stave 0.2 to 1.0
Cast iron (uncoated) 0.25
Concrete 0.3 to 3.0
Riveted steel 1 to 10

Reynold’s analogy between momentum and heat transfer for turbulent flow in a pipe:
In laminar flow, we have the expression for shear stress and heat transfer as follows:
t du
= n× (in laminar flow)
r dy
q dT
= a× (in laminar flow)
r ×Cp dy

456 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 9.13 Loss coefficient (kL) for some common valves and fittings

Item kL

Angle valve, fully open 3.1 to 5.0


Ball check valve, fully open 4.5 to 7.0
Gate valve, fully open 0.19
Globe valve, fully open 10
Swing check valve, fully open 2.3 to 3.5
Regular radius elbow, screwed 0.9
Regular radius elbow, flanged 0.3
Long radius elbow, screwed 0.6
Long radius elbow, flanged 0.23
Close return bend, screwed 2.2
Flanged return bend, two elbows, regular radius 0.38
---do---, long radius 0.25
Standard Tee, screwed, flow through run 0.6
---do---flow through side 1.8

Here n and a represent momentum and thermal diffusivity, respectively. It is a molecular phenomenon i.e.
in laminar flow, momentum is transported between layers of fluid at a molecular level. However, in turbulent
flow, there is an additional factor of ‘eddy transport’ i.e. chunks of fluid, called, ‘eddies’ also physically move
between layers and contribute to the transport of momentum and heat. This is represented for momentum and
heat transfer, respectively, as follows:
t du
= (n + e M)× (in turbulent flow...(9.159))
r dy
q dT
= (a + eH)× (in turbulent flow...(9.160))
r × Cp dy
Now, let us assume that momentum and heat are transported at the same rate i.e. e M = eH, and that the
Prandtl number, Pr = 1. Then, dividing Eq. 9.160 by 9.159, we get:
q
×du = dT ...(9.161)
Cp ×t
Now, integrate Eq. 9.161 from the surface to the mean bulk conditions, i.e. from T = Ts, u = 0 to T = Tb and
u = um, assuming that q/t is a constant at the surface = qs /ts :
qs
z z
um
× 1 du =
Cp × t s 0 Ts
Tb
- 1dT

qs × um
i.e. = Ts – Tb ...(9.162)
Cp ×t s
Now, heat flux at the wall can be written as:
qs = h×(Ts – Tb) ...(9.163)
And, the shear stress at the wall = (shear force)/surface area

F p ×D I 2
DP× GH 4 JK DP D
ts = = ×
p ×D × L 4 L

L u2
where, the pressure drop = DP = f× ×r× m
D 2

FORCED CONVECTION 457


So, we get:
f 2
ts = × r × um ...(9.164)
8
Substituting Eqs. 9.163 and 9.164 in Eq. 9.162, we get:
h NuD f
St = = = ...(9.165)
r × Cp × um Red × Pr 8
Eq. 9.165 is called ‘Reynold’s analogy’ for fluid flow in a pipe and is valid for both laminar and turbulent
flows. Note the restriction that Pr = 1, in Reynold’s analogy i.e. it holds good for most of the gases.
For fluids with Prandtl number much different from unity, we have the ‘Colburn analogy’ expressed as
follows:
2
f
St×Pr 3 = ...(9.166)
8
All fluid properties in Eq. 9.166 are evaluated at (Tb + Ts)/2, except Cp in Stanton number, which is evaluated
at the bulk temperature of the fluid.
Note that by analogy between momentum and heat transfer, we get a relation between heat transfer coeffi-
cient (h) and friction coefficient (f), and by knowing any one of them, the other quantity can be calculated.
There are two more analogies, more refined than the ones already mentioned. We shall just state them:
Prandtl analogy:
f
St = 2 (Prandtl analogy...(9.167))
f
1 + 5× × ( Pr - 1)
2
Prandtl analogy reduces to Reynold’s analogy when Pr = 1.
Von Karman analogy:

FG f IJ × Re × Pr
Nu =
H 2K (Von Karman analogy...(9.168))
f L L 5 OO
2 MN
× ( Pr - 1) + ln M1 + ×( Pr - 1)P P
1 + 5×
N 6 QQ
Substituting the f relation from Eq. 9.155 in the Colburn analogy, i.e. 9.166, we get the following relation for
Nusselt number for fully developed turbulent flow in smooth tubes:
1
Nu = 0.023×Re0.8 ×Pr 3 (for 0.7 < Pr < 160, Re > 10,000...(9.169))
This is known as ‘Colburn equation’.
9.10.5.3 Design equations. However, more popularly used design equation for fully developed (L/D > 60), tur-
bulent flow in pipes is the ‘Dittus–Boelter equation’. (1930), given below:
Nu = 0.023×Re 0.8 ×Pr n (for 0.7 < Pr < 160, Re > 10,000...(9.170))
where n = 0.4 for heating and n = 0.3 for cooling of the fluid flowing through the pipe. Here, fluid properties are
evaluated at the bulk mean temperature of fluid i.e. at Tb = (Ti + Te )/2 , where Ti is the temperature of fluid at
pipe inlet and Te is the temperature of fluid at pipe outlet.
If the temperature difference, (Ts – Tb) is significant, then variations in physical properties have to be taken
into account, and in such situations correlation of Sieder and Tate (1936) is recommended:
1
Fm I 0 .14
Nu = 0.027×Re0.8 ×Pr 3 × GH m JK
b
s
(for 0.7 < Pr < 10,000, 6000 < Re < 107...(9.171))

A more recent relation (1970) which fits experimental results better is the following:

458 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG f IJ × Re × Pr
H 8K Fm I
×G J
b
n
Nu =
F fI
1.07 + 12.7 × G J ×(Pr
0.5
Hm K
s
...(9.171a)

H 8K
0 .67
- 1)

mb
where n = 0.11 for heating of fluids, n = 0.25 for cooling of fluids, n = 0 for constant heat flux and is to be
ms
Ts
replaced by for gases, temperature in Kelvin
Tb
Above equations can be used for the cases of heat transfer with constant wall temperature as well as uniform
heat flux at the wall surface.
Also, relations for turbulent flow in circular pipes can be used for non-circular tubes as well, by replacing
pipe diameter D in evaluating Reynolds number by the hydraulic diameter, Dh = 4.A/P.
Correlation for thermal entry region:
For the range of L/D from 10 to 400, Nusselt recommended the following relation for turbulent flow in
pipes:
1
FG D IJ 0 . 055
Nu = 0.036×Re 0.8 × Pr 3 ×
H LK (for 10 < (L/D) < 400...(9.172))

Here, fluid properties are evaluated at mean bulk temperature.


9.10.5.4 Turbulent flow in a long, smooth annulus. For Nusselt number, the correlations for circular pipes are used,
with the hydraulic diameter taken as Dh = Do – Di.
For friction factor, following relation is proposed:
fannulus = 0.085×(Re) –0.25 (Re based on hydraulic diameter...(9.173))
9.10.5.5 Correlations for liquid metals. For fully developed turbulent flow of liquid metals in smooth circular tubes
[(L/D) > 30], with constant surface heat flux, Skupinski et.al. recommend the following correlation:
Nu = 4.82 + 0.0185×Pe 0.827 (3600 < Re < 9.05 ´ 10 5, 100 < Pe < 10,000...(9.174))
Note that Pe = Re.Pr
More recent (1972) correlation, which fits the available data well for flow of liquid metals in pipes with
constant heat flux, is due to Notter and Sleicher:
Nu = 6.3 + (0.0167×Re 0.85 ×Pr 0.93) ...(9.175)
Similarly, for constant surface temperature conditions, for flow of liquid metals, Seban and Shimazaki
recommend the following correlation for Pe > 100, and [(L/D) > 30],
Nu = 5.0 + 0.025×Pe0.8 (for Ts = constant Pe > 100...(9.176))
9.10.5.6 Helically coiled tubes. Coiled tubes are used to enhance the heat transfer coefficient and also to accom-
modate a larger heat exchange surface in a given volume. Heat transfer in a coiled tube is more compared to that
in a straight tube due to the contribution of secondary vortices formed as a result of centrifugal forces.
Here, we define a new dimensionless number, called ‘Dean number, Dn’ as follows:

F DI
1

= Re× G J
2
Dn
Hd K c
...(9.177)

where D is the diameter of the tube and dc is the diameter of the coil.
For laminar flow, following equations are recommended, depending upon the Dean number:
(a) When Dn < 20:
1
Nuavg = 1.7×(Dn 2 × Pr) 6 (Dn < 20, Dn2 ×Pr > 10,000...(9.178))
(b) When 20 < Dn < 100:
1
Nuavg = 0.9×(Re2 ×Pr) 6 (20 < Dn < 100,...(9.179))

FORCED CONVECTION 459


(c) When Dn > 100:
1
F DI 0 .07
Nuavg = 0.7×Re0.43 ×Pr 6 × GH d JK c
(100 < Dn < 830,...(9.180))

All the above three Eqs. viz. 9.178, 9.179 and 9.180 are valid for 10 < Pr < 600.
Also, for coiled tubes, there is not much difference in values of average Nusselt numbers whether the sur-
face temperature is kept constant or the surface heat flux is maintained constant.
In laminar flow, friction factor for a coiled tube is obtained from:
FG 64 IJ × 21.5× Dn × 5.73
fcoiled =
H Re K (1.56 + log(Dn)) (2000 > Dn > 13.5...(9.181))

Here, logarithm is to base 10.


Critical Reynolds number at which flow becomes turbulent in a coiled pipe is given as:

F DI 0 . 32
Recr = 2× GH d JK
c
×10 4 (for 15 < (D/dc) < 860...(9.182))

For values of (D/dc) > 860, critical Reynolds number for a curved pipe is the same as that for a straight pipe.
For turbulent flow in forced convection in helically coiled tubes, Hausen has proposed the following corre-
lation:
Nua _ helical FG 21 IJ ×FG D IJ
Nu a _ straight
=1+
H Re 0 .14 K Hd K c
...(9.183)

Here, LHS is the ratio of average Nusselt numbers for helical and straight tubes, D is the diameter of the
tube and dc is the diameter of the coil.
Example 9.17. Water is heated in the annular section of a double pipe heat exchanger by electrical heating of the inner
pipe. Outer pipe is insulated. Mean bulk temperature of water is 60°C. For the annulus, Di = 2.5 cm and Do = 5 cm.
Determine the convection coefficient and pressure drop/metre length for:
(i) flow rate of 0.04 kg/s, and
(ii) flow rate of 0.5 kg/s
Solution.
Data:
Ta := 60°C Di := 0.025 m Do := 0.05 m L := 1 m m1 := 0.04 kg/s (Case (i)) m2 := 0.5 kg/s
First, we need the properties of water at average temperature of 60°C:
r := 983.3 kg/m3 m := 0.467 ´ 10 –3 kg/(ms) Cp := 4185 J/(kgC) k := 0.654 W/(mC) Pr := 2.99
Case (i): Flow rate is 0.04 kg/s:
Since there is electrical heating of the inside tube, it is a case of constant heat flux at the wall; and, the outside surface is
insulated.
Reynolds number:
To calculate Re, we need hydraulic diameter, since this is annular duct:
4 × Ac
We have, for hydraulic diameter: Dh =
P
i.e. Dh := Do – Di
i.e. Dh = 0.025 m (hydraulic diameter)
m1
Velocity of flow: U1 :=
L p ×(D - D OP
r×M
2
o
2
i
m/s

N 4 Q
i.e. U 1 = 0.028 m/s
Dh × r ×U1
Therefore, Re: =
m
i.e. Re = 1.454 ´ 10 3 < 2300 (laminar flow)

460 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Heat transfer coefficient:
Therefore, this is the case of laminar flow in an annular duct, insulated from outside and subjected to constant heat
flux at the inner wall. We assume fully developed flow.
Then, from Table 9.10 we get:
NuH = 6.58 (for Di /Do = 0.5)
h × Dh
i.e. = 6.58
k
6 . 58 × k
i.e. h := W/(m2 C) (heat transfer coefficient)
Dh
i.e. h = 172.133 W/(m2C) (heat transfer coefficient.)
Pressure drop:
Friction factor for fully developed laminar flow in an annulus, is read from Table 9.7.
We get: f×Re = 95.15 (for ratio of radii = 0.5)
95.15
Therefore, f := i.e. f = 0.065 (friction factor)
Re
Therefore, pressure drop is given by:
Dp f r × um2
= × ...(9.128)
L D 2
L r × u12
i.e. Dp := ×f× (Pa)
Dh 2
i.e. Dp = 0.982 (Pa/metre length...pressure drop.)
Case (ii): Flow rate is 0.5 kg/s:
Reynolds number:
m2
Velocity of flow: U 2 :=
L p ×(D - D ) OP m/s
r×M
2
o
2
i

N 4 Q
i.e. U 2 = 0.345 m/s
DH × r ×U 2
Therefore, Re :=
m
i.e. Re = 1.818 ´ 10 4 > 2300 (turbulent flow)
Heat transfer coefficient:
Therefore, this is the case of turbulent flow in an annular duct, insulated from outside and subjected to constant
heat flux at the inner wall. We assume fully developed flow. And the Dittus-Boelter correlation can be used with the
hydraulic diameter substituted for tube diameter D.
Nu = 0.023× Re 0.8 × Pr n (for 0.7 < Pr < 160, Re > 10,000...(9.170))
Here, n = 0.4, since the fluid is being heated.
i.e. Nu := 0.023×Re 0.8 ×Pr 0.4
i.e. Nu = 91.117 (Nusselt number)
Nu × k
Therefore, h := W/(m2 C) (heat transfer coefficient)
Dh
i.e. h = 2.384 ´ 10 3 W/(m2 C) (heat transfer coefficient)
Pressure drop:
Friction factor for fully developed turb. flow in an annulus, can be read from Moody’s diagram, or we can use Eq.
9.154:
i.e. f := 0.316× Re – 0.25 (for 2 ´ 10 4 < Re < 8 ´ 104...(9.154))
We get: f = 0.027 (friction factor)
Therefore, pressure drop is given by:
Dp f r × um2
= × ...(9.128)
L D 2

FORCED CONVECTION 461


L r ×U 22
i.e. Dp: = ×f× (Pa)
Dh 2
i.e. Dp = 63.814 (Pa/metre length...pressure drop.)
Example 9.18. Water at 20°C flows through a 2.5 cm ID, 1 m long pipe, whose surface is maintained at a constant
temperature of 50°C, at velocity of 5 cm/s. Determine the outlet temperature of water, assuming fully developed hydro-
dynamic boundary layer.
Solution.
Data:
Ti := 20°C Ts := 50°C D := 0.025 m L := 1 m U := 0.05 m/s
We need the properties of water at mean bulk temperature. But, as yet, we do not know the exit temperature of
water. So, let us assume the mean bulk temperature as 30°C and proceed with the calculations; later, we will check this
assumption and refine our calculations, if required.
Properties of water at Tb = 30°C:
r := 996.0 kg/m3 m :=0.798 ´ 10– 3 kg/(ms) Cp := 4178 J/(kgC) k := 0.615 W/(mC) Pr := 5.42
Reynolds number:
D ×U × r
Re :=
m
i.e. Re = 1.56 ´ 10 3 (< 2300...therefore, laminar flow)
L
Now, = 40 and 0.4×Re = 62.406
D
Therefore, flow is in the entrance region
L
D = 4.73 ´ 10 –3 <0.01
Re × Pr
Therefore, we use Eq. 9.150b, viz

F I 0 . 333

Nu avg
G Re × Pr JJ
:= 1.67 × G (for (L/D)/(Re.Pr) < 0.01, constant wall temperature...9.150b)
GH DL JK
i.e. Nu avg = 9.931 (average Nusselt number)
Nuavg × k
Therefore, h :=
D
i.e. h = 244.293 W/(m2 C) (heat transfer coefficient)
Now, determine the outlet temperature by an energy balance:
p ×D 2 FG
Ti + To IJ
i.e.
4
×r×U×Cp ×(To – Ti) = p×D×L×h × Ts -
2H K
In the above equation we have assumed that the mean temperature difference between the water stream and the
surface is the difference between the surface temperature and the arithmetic mean of water temperature at inlet and exit.
Strictly speaking, we should consider the LMTD; however, the assumption of arithmetic mean is good enough and the
error is not much.
Let us solve this easily by Mathcad. Assume a guess value for To to start with, and then write the constraint after
typing ‘Given’. Then the command ‘Find (To)’ gives the value of To immediately:
To := 100 (guess value of To)
Given
p ×D 2 FG
T + To IJ
4
×r×U×Cp ×(To – Ti) = p×D×L×h × Ts - i
2 H K
Find (To) = 25.152
i.e. To = 25.152°C (exit water temperature)
Therefore, mean temperature of water is: (20 + 25.152)/2 = 22.5°C, whereas we had assumed a mean value of 30°C.
Taking the properties of water at 22.5°C, calculations can now be repeated:

462 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Properties of water at Tb = 22.5°C:
r := 997.5 kg/m3 m := 0.95 ´ 10 – 3 kg/(ms) Cp := 4181 J/(kgC) k := 0.602 W/(mC) Pr := 6.575
Therefore,
D ×U × r
Re :=
m
i.e. Re = 1.313 ´ 10 3 (< 2300...therefore, laminar flow)
L
D = 4.635 ´ 10 –3 < 0.01
Re × Pr
Therefore, we use Eq. 9.150b i.e.

F I 0 . 333

Nu avg
G Re × Pr JJ
:= 1.67 × G (for (L/D)/(Re.Pr) < 0.01, constant wall temperature....9.150b)
GH DL JK
i.e. Nu avg = 9.998 (average Nusselt number)
Nuavg × k
Therefore, h :=
D
i.e. h = 240.754 W/(m2 C) (heat transfer coefficient)
Now, determine the outlet temperature by an energy balance, and using Solve block of Mathcad:
To := 100 (guess value of To)
Given
p ×D 2 FG
T + To IJ
4
×r×U×Cp ×(To – Ti) = p×D×L×h × Ts - i
2 H K
Find (To) = 25.073
i.e. To = 25.073°C (exit water temperature.)
Therefore, Tb = (20 + 25.073)/2 = 22.573°C, which is very close to Tb = 22.5°C at which properties of water were
taken. So, To = 25.073°C ...(Ans).
Example 9.19. Air at 1 bar and 20°C flows through a 6 mm ID, 1 m long smooth pipe, whose surface is maintained at a
constant heat flux, with velocity of 3 m/s. Determine the heat transfer coefficient if the exit bulk temperature of air is
80°C. Also determine the exit wall temperature and the value of h at the exit.
Solution.
Data:
Ti := 20°C To := 80°C D := 0.006 m L := 1 m U := 3.0 m/s
Therefore, mean bulk temperature is (20 + 80)/2 = 50°C
i.e. Tb = 50°C (mean bulk temperature of air)
Properties of air at Tb = 50°C:
r := 1.093 kg/m 3 m := 19.61 ´ 10 –6 kg/(ms) Cp := 1005 J/(kgC) k := 0.02826 W/(mC) Pr := 0.698
Reynolds number:
D ×U × r
Re :=
m
i.e. Re = 1.003 ´ 10 3 (< 2300...therefore, laminar flow)
Since the tuube length is short, entrance effect must be considered.
L
We have: = 166.667
D
L
and, D = 0.238 > 0.01
Re × Pr
Nusselt number:
Therefore, we shall use following equation assuming developing velocity profile:

FORCED CONVECTION 463


FG D IJ × Re × Pr
H LK
0 .036 ×
Nu := 4.36 +
LF D I
1 + 0. 0011× MG J × Re × Pr P
O (Pr = 0.7...(9.151c))

HN L K Q
i.e. Nu = 4.511 (Nusselt number)
Heat transfer coefficient:
Nu × k
Therefore, h := W/(m2 C) (heat transfer coefficient)
D
i.e. h = 21.245 W/(m2C) (heat transfer coefficient)
Exit wall temperature:
Since the wall heat flux is constant, we have the relation for h:
qw
h= ...(a)
(Ts - Tb )
Also,
F p × D I ×U
2
Mass flow rate: m := r× GH 4 JK
i.e. m= 9.27 ´ 10 –5 kg/s
and, Q := m×Cp ×(To – Ti) W (total heat transfer rate)
i.e. Q= 5.591 W (total heat transfer rate)
But, Q= qw ×p× D× L (where qw is the constant surface heat flux)
Q
Therefore, qw := W/m2 (surface heat flux)
p ×D× L
i.e. qw = 296.586 W/m2 (surface heat flux)
Therefore, from Eq. a:
qw
Tw_exit := To + °C (surface temperature at exit)
h
i.e. Tw_exit = 93.96°C (surface temperature at exit.)
Example 9.20. Water (under pressure) is heated in an economiser from a temperature of 30°C to 150°C. Tube wall is
maintained at a constant temperature of 350°C. If the water flows at a velocity of 1.5 m/s and the tube diameter is 50
mm, determine the length of tube required.
Solution.
Data:
Ti := 30°C To := 150°C Ts := 350°C D := 0.05 m U := 1.5 m/s
Therefore, mean bulk temperature is (30 + 150)/2 = 90°C
i.e Tb := 90°C ...mean bulk temperature of water
Properties of water at Tb = 90°C:
r := 965.3 kg/m 3 m := 0.315 ´ 10 – 3 kg/(m/s) Cp := 4206 J/(kgC) k := 0.675 W/(mC)
Pr := 1.96
Reynolds number:
D ×U × r
Re :=
m
i.e. Re = 2.298 ´ 10 5 (> 2300...therefore, turbulent flow)
Heat transfer coefficient
Using more recent correlation,

FG f IJ × Re × Pr
H 8K Fm I n

Nu =
F fI 0. 5 ×GH m JK
b
...(9.171a)
1. 07 + 12.7 × G J ×( Pr
s

H 8K
0 . 67
- 1)

464 FUNDAMENTALS OF HEAT AND MASS TRANSFER


mb
where n = 0.11 for heating of fluids, n = 0.25 for cooling of fluids, n = 0 for constant heat flux and is to be replaced
ms
Ts
by for gases, temperature in Kelvin
Tb
We have:
f := (0.79 ×ln (Re) – 1.64) –2 ...(9.156)
i.e. f = 0.015
and, dynamic viscosity of water at wall temperature of 350°C is:
m s = 0.065 ´ 10 –3 kg/(m/s)
Therefore,

FG f IJ × Re × Pr
H 8K F mI 0 . 11

Then, Nu :=
F fI 0. 5 GH m JK
1.07 + 12.7 × G J × (Pr
s

H 8K
0 . 67
- 1)

i.e. Nu = 734.689
Nu × k
Therefore, h :=
D
i.e. h = 9.918 ´ 10 3 W/(m 2C) (heat transfer coefficient)
Length of tube required:
Water temperature varies continuously from 30°C at inlet to 150°C at exit, tube surface temperature remainin con-
stant at 350°C. So, mean temperature difference in Newton’s equation is LMTD, to be very accurate.
(Ts - Ti ) - (Ts - To )
LMTD :=
ln
LM (T - T ) OP
s i

N (T - T ) Q
s o

i.e. LMTD = 255.317°C (log mean temperature difference)


Applying energy balance:
F p × D I ×U×C ×(T
2
r× GH 4 JK p o – Ti) = h ×(p×D×L)×LMTD

Therefore,
F p × D I ×U × C ×(T - T )
2
r× GH 4 JK p o i

L :=
h × (p × D) × LMTD
i.e. L = 3.607 m (length of tube required.)
Note: We could have taken the mean temperature difference as the difference between surface temperature and the
arithmetic mean between inlet and exit of water i.e. DT = 350 – 90 = 260 whereas LMTD was 255.7°C. Then, L would
have been 3.542 m, not much different from 3.6 m; however, using LMTD is accurate method.
Alternatively, if we had used Dittus–Boelter equation, viz.
Nu := 0.023×Re 0.8 ×Pr 0.4 (Pr 0.4 since fluid is being heated)...(9.170)
i.e. Nu = 585.815
Nu × k
and h :=
D
i.e. h = 7.909 ´ 10 3 W/(m 2C) ...heat transfer coefficient.
And, using LMTD we would have got L = 4.524 m
Example 9.21. Sodium potassium alloy (25:75), flowing at a rate of 3 kg/s, is heated in a tube of 5 cm ID from 200°C to
400°C. Tube surface is maintained at constant heat flux and the temperature difference between the tube surface and the
mean bulk temperature of fluid is 40°C. Determine the heat transfer coefficient, heat flux at the surface and length of
tube required.

FORCED CONVECTION 465


Solution.
Data:
Ti := 200°C To := 400°C DT := 40°C D := 0.05 m m := 3.0 kg/s
Therefore mean bulk temperature is (200 + 400)/2 = 300°C
i.e. Tb = 300°C ...mean bulk temperature of Na–K alloy
Properties of Na–K alloy at Tb = 300°C:
r := 799 kg/m 3 n := 0.366 ´ 10 –6 m2/s Cp := 1038.3 J/(kgC) k := 22.68 W/(mC) Pr := 0.0134
Reynolds number:
m
G :=
F p ×D I 2
kg/(sm2) (mass velocity)
GH 4 JK
i.e. G = 1.528 x 103 kg/(sm2) (mass velocity)
G× D
Re := (Reynolds number)
n×r
i.e. Re = 2.612 x 105 (> 2300...therefore, turbulent flow)
Heat transfer coefficient
Using the recent correlation of Notter and Sleicher, fo constant heat flux conditions
Nu := 6.3 + (0.0167×Re 0.85 ×Pr 0.93) ...(9.175)
i.e. Nu = 18.473 (Nusselt number)
Nu × k
Therefore, h := W/(m2 C) (heat transfer coefficient)
D
i.e. h = 8.379 ´ 10 3 W/(m2C) ...heat transfer coefficient
Heat flux at surface:
Now, heat flux is determined from its definition:
qs
h= (where qs is the surface heat flux and DT is the temperature
DT diffierence between surface and the bulk temperature = 40°C,
a constant for constant heat flux conditions.)

i.e. qs := h×DT W/m2 (surface heat flux)


i.e. qs = 3.352 ´ 105 W/m2 (surface heat flux.)
Length of tube required:
This is obtained by a heat balance:
qs ×(p×D×L) = m×Cp ×(To – Ti)
m × Cp × (To - Ti )
i.e. L :=
qs × p × D
i.e. L = 11.833 m (length of tube required.)
Alternatively:
h ×(p×D×L)×D T = m×Cp ×(To – Ti)
m × Cp × (To - Ti )
L :=
h × (p × D) × D T
i.e. L = 11.833 m (same as earlier.)
Also, if we use Eq. 9.174 to determine heat transfer coefficient:
Nu := 4.82 + 0.0185×(Re×Pr)0.827 (3600 < Re < 9.05 ´ 105, 100 < Pe < 10,000...(9.174))
i.e. Nu = 20.602
Nu × k
Therefore, h := W/(m2 C) (heat transfer coefficient)
D
i.e. h = 9.345 ´ 10 3 W/(m 2C) (heat transfer coefficient)
Compare this value of h with that obtained earlier using Eq. 9.175.

466 FUNDAMENTALS OF HEAT AND MASS TRANSFER


And,
m × Cp × (To - Ti )
L :=
h × (p × D) × D T
i.e. L = 10.61 m (length of tube required)
Example 9.22. 180 kg/h of air at one atm. pressure is cooled from 100°C to 20°C while passing through a 3 cm ID pipe
coil bent into a helix of 0.7 m diameter. Calculate the air side heat transfer coefficient.
Solution.
Data:
180
Ti := 200°C To := 20°C D := 0.03 m dc := 0.7 m m := kg/s i.e. m = 0.05 kg/s
3600
Therefore mean bulk temperature is (100 + 20)/2 = 60°C
i.e. Tb := 60°C (mean bulk temperature of air)
Properties of air at Tb = 60°C:
r := 1.06 kg/m3 m := 20.10 ´ 10 –6 kg/(ms) Cp := 1005 J/(KgC) k := 0.02896 W/(mC) Pr := 0.696
Reynolds number:
m
G :=
F p ×D I2
kg/(sm2) (mass velocity)
GH 4 JK
i.e. G = 70.736 kg/(sm2) (mass velocity)
G× D
Re := (Reynolds number)
m
5
i.e. Re = 1.056 ´ 10 ...> 2300...therefore, turbulent flow
Nusselt number for straight tube:
Using the Dittus-Boelter equation for turbulent flow:
Nu = 0.023× Re 0.8 × Prn (for 0.7 < Pr < 160, Re > 10,000...(9.170))
i.e. Nu := 0.023×Re 0.8 ×Pr 0.3 (n = 0.3 since air is being cooled.)
i.e. Nu = 215.457 (Nusselt number...for straight tube)
Nusselt number for helical coil:
We have:
Nua _helical FG 21 IJ ×FG D IJ
Nua _straight
= 1+
H Re0 . 14
K Hd Kc
...(9.183)

Therefore,
LM FG 21 IJ × F D I OP
Nua_helical := Nu × 1 +
MN H Re K GH d JK QP
0 . 14
c

i.e. Nua_helical = 253.855


Heat transfer coefficient:
Nua_helical × k
Therefore, h := W/(m2 C) (heat transfer coefficient)
D
2
i.e. h = 245.054 W/(m C) (heat transfer coefficient.)
Example 9.23. In a long annulus (3.125 cm ID, 5 cm OD), air is heated by maintaining the temperature of outer surface of
the inner tube at 50°C. The air enters at 16°C and leaves at 32°C and its flow velocity is 30 m/s. Estimate the heat
transfer coefficient between the air and the inner tube. Use Dittus - Boelter equation, viz.
Nu D = 0.023.(ReD)0.8.Pr 0.4 ; Average properties of air at 24°C are:
r = 1.614 kg/m3, Cp = 1007 J/(kgC), k = 0.0263 W/(mC), Pr = 0.7 n = 15.9 ´ 10 – 6 m2/s (M.U. 1999)
Solution.
Data:
Ti := 16°C To := 32°C Ts := 50°C Di := 0.03125 m Do := 0.05 m L := 1 m U := 30 m/s

FORCED CONVECTION 467


Reynolds number:
To calculate Re, we need hydraulic diameter, since this is annular duct:
4 × Ac
We have, for hydraulic diameter: Dh =
P
i.e. Dh := Do – Di
i.e. Dh = 0.019 m (hydraulic diameter)
Dh ×U
Therefore, Re :=
n
i.e. Re = 3.538 ´ 10 4 (> 2300...turbulent flow)
Heat transfer coefficient:
We have:
Nu := 0.023×Re 0.8 ×Pr0.4 (Dittus–Boelter equation n = 0.4, since air is being heated)
i.e. Nu = 86.846 (Nusselt number)
Nu × k
Therefore, h :=
Dh
i.e. h = 121.815 W/(m2 C) (heat transfer coefficient)
Also, calculate the pressure drop per metre length:
Friction factor:
We have, from Eq. 9.155:
f := 0.184×Re –0.2 (for 104 < Re < 105...(9.155))
i.e. f = 0.023 (friction factor)
Pressure drop:
L r ×U 2
Therefore, DP := f× × Pa ( = N/m 2) (pressure drop per meter length)
Dh 2
i.e. D P = 877.38 Pa (pressure drop per meter length.)
Example 9.24. Water at 20°C flows flows through a tube, 4 cm diameter 9 m length, tube surface being maintained at
90°C. Temperature of water increases from 20°C to 60°C. Find the mass flow rate. Use Dittus–Boelter equation, viz. Nu D
= 0.023.(ReD)0.8 .Pr0.4; Take properties of water at mean bulk temperature of 40°C as:
r = 993 kg/m3, Cp = 4170 J/kgC), k = 0.64 W/(mC), n = 0.65 ´ 10 –6 m2 s (M.U., 1996)
Solution.
Data:
Ti := 20°C To := 60°C Ts := 90°C D := 0.04 m L := 9 m m := n×r i.e. m = 6.455 ´ 10 –4 kg/(ms)
m ×Cp
Therefore, Pr := i.e. Pr = 4.206
k
Now, from Dittus–Boelter equation we get Nusselt number, hence the heat transfer coefficient h; then writing a heat
balance:
Let m be the mass flow rate (kg/s) of water.
Heat gained by water = heat transferred between the pipe surface and the bulk of water
i.e. m×Cp ×(To – Ti) = h ×As ×LMTD
LM k F m× D I 0 .8
OP
i.e. m×Cp ×(To – Ti) =
MN D × 0.023 GH A × m JK
c
× Pr 0 . 4 × As × LMTD
PQ
p ×D 2
where Ac := i.e. Ac = 1.257 ´ 10 –3 m2 (area of cross-section)
4
As := p×D×L i.e. As = 1.131 m2 (surface area of heat transfer)
(Ts - Ti ) - (Ts - To )
LMTD :=
ln
LM (T - T ) OP
s i
i.e. LMTD = 47.209°C

N (T - T ) Q
s o

468 FUNDAMENTALS OF HEAT AND MASS TRANSFER


LM L k OO
1

F D I
MM MMN D × 0.023 GH A × m JK × Pr × A × LMTD PPQ PPP
0 .8 0. 2
0.4
s
c
Therefore, m :=
MM [C × (T - T )]
p o i PP
MN PQ
i.e. m = 2.373 kg/s (mass flow rate of water)

9.11 Summary of Basic Equations for Forced Convection

Geometry/details Correlation Restrictions


Flat Plate, laminar flow:

Hydrodynamic boundary 5×x


dlam = Re < 5 ´ 10 5
layer thickness (Re x )0 . 5

t 66.4
Re < 5 ´ 10 5
Local friction coefficient Cfx =
F r ×U I 2
=
Re x
GH 2 JK
h×x
Local Nusselt number = Nu x = 0.332× Re x ×Pr 0 . 333 Re < 5 ´ 105, Pr > 0.5
k

Average Frction
coefficient
Cfa =
1 L
z
× Cfx dx =
L 0
1.328
ReL
Re < 5 ´ 10 5

Average Nusselt number Nua = 0.664× ReL ×Pr 0 . 333 Re < 5 ´ 105, Pr > 0.5

Local Nusselt number Nux = 0.565× Pex0.5 ...(Pr <.005) Re < 5 ´ 10 5


for liquid metals Pe = Re. Pr
Flat Plate, turbulent flow:

0.371× x
Hydrody. b.l. thickness d turb = Re x > 5 ´ 10 5
(Re x )0 . 2
-1

Local friction coefficient Cfx = 0.0576× Re x Rex > 5 ´ 10 5, Pr > 0.5


5

1
hx × x
Local Nusselt number Nux = = 0.288× Re x0.8 × Pr 3 Rex > 5 ´ 10 5, Pr > 0.5
k
-1
Average Friction coefficient Cfa = 0.072× ReL5 5 ´ 105 < ReL < 107

0.455
Average Friction coefficient Cfa = 107 < ReL < 109
(log (Re L ))2 . 58

Flat Plate, mixed


boundary layer:

0.074 1742
Average Friction coefficient Cfa = 1
- 5 ´ 105 < ReL < 107
ReL
Re 5
L Rex, c = 5 ´ 10 5

Contd.

FORCED CONVECTION 469


Contd.

h ×L F 4 1 I
Average Nusselt number Nuavg =
k GH
= 0.036 ×ReL5 - 836 ×Pr 3 JK 0.6 < Pr < 60,

5 ´ 105 < ReL < 107

Cylinder in cross flow: 100 < Re < 107

LM F OP
4

IJ
1 1 5 5
h ×D 0.62 ×Re 2 ×Pr 3 Re
G 8

LM F 0.4 I OP MMN H K PP
Average Nusselt number Nucyl = = 0.3 + × 1+ Re.Pr > 0.2
Q
1
k 2 4 28200

MM1 + GH Pr JK PP
3

N Q
Cylinder in liquid metal Nucyl = 1.125×(Re × Pr)0.413 ...for 1 < Re.Pr < 100
cross flow

F 1 2
I Fm I
Flow across a sphere: GH
Nusph = 2 + 0.4 ×Re 2 + 0.06 ×Re 3 ×Pr 0 . 4 × JK GH m JK
w
a
For gases & liquids.

Comprehensive equation
of Whitaker. 3.5 < Re < 7.6.104
Average Nusselt number 0.71 < Pr < 380,
1 < m/m s < 3.2
Falling drop:
1 1
Average Nusselt no. Nuavg = 2 + 0.6× Re 2 ×Pr 3

Flow across Tube bank: Nua = 0.021× ReD0.84 × Pr 0.36 × (Pr/Prw )0.25 ...for In-line tubes,

Turbulent flow
F Pr I 0 . 25

(ReD > 2 ´ 105) Nua = 0.022× ReD0.84 × Pr 0.36 × GH Pr JK


w
...for staggered tubes, Pr >1 N > 20, and 0.7 < Pr
< 500, 1000 < ReD_max
< 2 ´ 10 6

Nua = 0.019× ReD0.84 ...for staggered tubes, Pr = 0.7

2
2 ×f ×Gmax ×N m F I 0 .14

Flow across Tube banks: Dp =


r
× w
mb GH JK Pa Gmax = r × umax

Pressure drop N = No. of transverse


rows

LM OP
M
MM
0.118 P ×Re
LM (S - D ) OP PP
- 0 .16
Friction factor in Eq. 9.118 f = 0.25 + 1. 08 D ...for staggered tubes.

MN N D Q PQ
T

LM FS I OP
0.08 × G J
Friction factor in Eq. 9.118
M
f = M0.044 +
HDK
L
PP ×Re - 0 .15
...for in-line tubes
MM LM (S - D ) OP 0 . 43 + 1.13 ×
D
SL PP D

MN PQ
T
N D Q
Contd.

470 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.

ha ×D p 1- e F 1 2
I 1
Flow through packed beds:
k
=
e GH
× 0.5 ×ReDp
2 + 0. 2 × Re 3
Dp ×Pr
3
JK 20 < ReDp < 10,000,

Heat transfer between 0.34 < e < 0.78.


See text
gas and packings for definition of ReDp
and e
1 1
ha ×D p 0. 8
Flow through packed beds: = 2.58× ReDp
3
×Pr 3 + 0.094× ReDp ×Pr 0 . 4 For particles like
k
cylinders,
Heat transfer between see text for definition
walls of bed and gas of ReDp
1 1
ha ×D p 0. 8
Flow through packed beds: = 0.203 × ReDp
3
×Pr 3 + 0.220× ReDp ×Pr 0 . 4 For particles like
k
spheres, 40 < ReDp < 2000
Heat transfer between see text for definition
walls of bed and gas of ReDp
Flow inside tubes: Lh_lam = 0.05× Re × D Re < 2300...laminar
Hydrodynamic and Lt_lam = 0.05× Re × Pr × D Re > 4000...turbulent
thermal entry lengths Lh_turb = Lt_turb = 10× D
Darcy – Weisbach
equation for pressure Dp f r ×u m2
= ×
drop L D 2

64
Friction factor f= Laminar flow in tubes
ReD
h ×D
Flow inside tubes: NuD = = 4.364 Pr > 0.6
k
Nusselt no. for fully
developed laminar flow,
constant wall heat flux

Flow inside tubes:


h ×D
Nusselt no. for fully NuD = = 3.66 Pr > 0.6
k
developed laminar flow,
constant wall temperature

Flow inside short tubes:

Nusselt no. for fully D FG IJ


developed velocity profile,
Nu avg = 3.66 +
0.0668
L
×Re ×Pr
H K ...Pr > 0.7 L/D < 60
laminar flow, constant wall
LMFG IJ OP
2
D 3

NH K
temperature 1 + 0.04 × ×Re ×Pr
L Q
Flow inside short tubes:

F I F I
1
Nusselt no. for fully
F I
1

GG Re × Pr JJ × G m J
3
developed velocity profile,
GG Re ×Pr JJ ×FG m IJ
0 .14 3
0 .14
laminar flow, constant wall
GH DL JK H m K
Nu avg = 1.86× £2
temperature..Sieder &
Tate relation.
s GH DL JK H m K s

Contd.

FORCED CONVECTION 471


Contd.

0.48 < Pr < 16,700


0.0044 < ( m / m s ) < 9.75

Flow inside short tubes:


Local Nusselt no. for fully FG D IJ ×Re ×Pr
developed velocity profile, HLK
0.023 ×
laminar flow, constant
Nu = 4.36 +
LF D I O
1 + 0.0012 × MG J ×Re ×Pr P
...Pr > 0.7
wall heat flux.
NH L K Q
Flow inside tubes:
Friction factor for smooth f = 0.316× Re –0.25 ...for 2 ´ 104 < Re < 8 < 104
pipes f = 0.184 × Re –0.2 ...for 104 < Re < 105.
f = (0.79×In (Re) – 1.64)–2 ...for 3000 < Re < 5 ´ 106
Flow inside tubes:
Friction factor for rough 1.325 Relative roughness,
f=
pipe LMln F e I + 5.74 OP 2 (e / D ) is known

MN GH 3.7D JK Re PQ09

h Nu D f
Reynold’s analogy St = = =
r ×C p ×u m Red ×Pr 8
2
f
Colburn analogy St × Pr 3 =
8
Flow inside tubes:
Turbulent flow: Nu = 0.023× Re0.8 × Pr n ...for 0.7 < Pr < 160, Re > 10,000 Dittus–Boelter equation
Nusselt number n = 0.4 when fluid is being heated, and 0.6 < Pr < 160
n = 0.3 when fluid is being cooled Re > 10,000 L/D > 10

Flow inside tubes:


Turbent flow: 1
Fm I 0 .14
Sieder - Tate eqn.
× GH m JK
b
Nusselt number, Nu = 0.027× Re 0.8
× Pr
3
0.7 < Pr < 16,700,
when there is s
6000 < Re < 107
property variation

FG f IJ ×Re ×Pr Fits the experimental


Flow inside tubes:
H 8K Fm I
×G
n
data better; n = 0.11 for
H m JK
Turbulent flow: Nu = b
Nusselt number Ff I
1.07 + 12.7 × G J ×(Pr
0.5 heating of fluids, n =
H 8K
0 . 67 s
- 1) 0.25 for cooling of fluids,
n = 0 for constant heat
flux, m b / m s = Ts /Tb ,
temperature in Kelvin

Flow of liquid metals


inside smooth pipes: Nu = 4.82 + 0.0185 × Pe 0.827 3600 < Re <9.05 ´ 105,
constant surface heat flux. 100 < Pe < 10,000

Flow of liquid metals Nu = 6.3 + (0.0167× Re0.85 ×Pr 0.93) Recent correlation
inside smooth pipes: which fits experimental
constant surface heat flux. data better.

Contd.

472 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.

Flow of liquid metals Nu = 5.0 + 0.025× Pe0.8 ...for Ts = constant, Pe > 100.
inside smooth pipes:
constant surface
temperature.

Helically coiled tubes: Nu a _ helical FG 21 IJ × FG D IJ D = diameter of tube


Turbulent flow:
Hausen' s relation
Nu a _ straight
= 1+
H Re
0 .14
K Hd K c
d c = diameter of helix

9.12 Summary
Convection is the mode of heat transfer with fluid motion. If the fluid motion is caused by density differences as
a result of temperature differences, then it is called ‘natural convection’; instead, if the fluid motion is imposed
due to a pump or fan, then it is called ‘forced convection’. Also, the flow may be ‘laminar’ or ‘turbulent’; in
laminar flow, the flow is ‘ordered’ and the different layers of fluid flow parallel to each other in an orderly
manner. In turbulent flow, the flow is ‘chaotic’ and the flow is highly disordered and there is ‘mixing’ between
different layers of fluid as a result of chunks of fluid (‘eddies’) moving between layers. Dimensionless number
that characterizes the flow as laminar or turbulent is the Reynolds number.
In this chapter, we studied the principles of forced convection and stated a few correlations for external flow
on flat plates, cylinders and spheres, and also for internal flow through circular and non-circular pipes.
Mathematical analysis of convection problem is complicated since the temperature profile has to be solved
in conjunction with the fluid flow relations. ‘Boundary layer concept’ simplifies this problem to some extent.
Boundary layer is a very thin, stagnant fluid layer that adheres to the wall surface wherein the velocity and
temperature gradients are significant. Thus, the flow field is considered to be made up of two regions, one ‘a
boundary layer region’ and the other, an ‘inviscid region’. Derivation of boundary layer equations and their
solution to the simple case of a flat plate was explained in some detail. Further simplification with the method of
integral equations was also demonstrated.
Central problem in convection heat transfer situation is to find out the heat transfer coefficient, ‘h’. Heat
transfer coefficient is generally represented in terms of the dimensionless Nusselt number, Nu. So, in the analysis,
our aim is to get a relation for Nusselt number. By ‘Dimensional Analysis’, it was shown that in forced convec-
tion, Nusselt number is expressed as function of Reynolds and Prandtl numbers.
We are also interested in the drag force between the fluid and the plate and the pressure drop occurs in the
pipe if a fluid is flowing through it. This is related to the shear stress at the walls, which in turn, is expressed in
terms of a ‘skin friction coefficient’ for the flat plate and a ‘friction factor’ for internal flow through a pipe. We
solve the momentum equation to get the shear stress and the friction coefficient, and by solving the energy
equation we get the temperature profile and thus the heat transfer coefficient.
There is a similarity in the governing equations of momentum and energy transfer. This leads to the idea of
‘analogy between momentum and heat transfer’ and we have extremely useful analogies such as Reynolds anal-
ogy and Colburn analogy. Particularly for rough tubes, an estimate of heat transfer coefficient is easily made just
by the knowledge of friction coefficient, with the help of these analogies.
Most of the convection correlations are empirical, deduced as result of large amount of experimental data.
Several empirical correlations for laminar as well as turbulent, forced convection, for many practically important
situations have been presented in this chapter.
In the next chapter, we study about heat transfer with natural convection.

Questions
1. Explain the difference between natural and forced convection in laminar and turbulent flow. [M.U.]
2. Write short notes on hydrodynamic and thermal boundary layers. What is the importance of these boundary
layers in heat transfer? [M.U.]
3. Explain the principle of dimensional analysis. What are its advantages and limitations?
[M.U.]
4. State Buckingham p-theorem. [M.U.]
5. Using dimensional analysis, derive an expression for heat transfer coefficient in forced convection in terms of
Nusselt number, Reynolds number and Prandtl numbers. [M.U.]

FORCED CONVECTION 473


6. Explain the physical significance of: (a) Reynolds number (b) Prandtl number, and (c) Nusselt number [M.U.]
7. Write a short note on applications of dimensional analysis. [M.U.]
8. Write a short note on Reynolds analogy between momentum and heat transfer, with reference to a flat plate.
[M.U.]
9. In flow across a cylinder, what is meant by friction drag and pressure drag? At what point on the cylinder is the
heat transfer maximum?
10. Show that for flow inside a circular tube, Reynolds number can be written as:
Re = 4.m/(p .D.m )
11. State the generally accepted values of critical Reynolds numbers at which the flow changes from laminar to
turbulent for: (a) a flat plate (b) flow across a circular cylinder (c) flow across a sphere, and (d) flow inside a
circular pipe.
12. Comment on the hydrodynamic and thermal entry lengths for laminar and turbulent flows for an oil inside a
circular pipe. How would they compare for a liquid metal?
13. What is the difference between friction factor and friction coefficient?
14. How is pressure drop in a tube related to the friction factor?
15. In a circular tube, where is the heat transfer coefficient higher, at the entry or exit? Why?
16. Does the roughness of tube surface affect the heat transfer in (a) laminar flow (b) turbulent flow. Explain your
answer.

Problems
1. Glycerine at 10°C flows over a flat plate, 6 m long, maintained at 30°C with a velocity of 1.5 m/s. Determine the
total drag force and the heat transfer rate over the entire plate per unit width. Properties of glycerine at 20°C are:
r = 1264 kg/m3, n = 1180 ´ 10– 6 m2/s, Pr = 12,500, k = 0.2861 W/ (mK) and Cp = 2387 J/(kgK).
2. Water at at 30°C is flowing with a velocity of 4 m/s along the length of a long, flat plate, 0.3 m wide, maintained
at 10°C.
(a) Calculate the following quantities at x = 0.3 m:
(i) boundary layer thickness (ii) local friction coefficient (iii) average friction coefficient (iv) local shear
stress due to friction (v) thickness of thermal boundary layer (vi) local convection heat transfer coefficient
(vii) average heat transfer coefficient (viii) rate of heat transfer from the plate between x = 0 and x = x, by
convection, and (ix) total drag force on the plate between x = 0 and x = 0.3 m
(b) Also, find out the value of xc (i.e. the distance along the length at which the flow turns turbulent, Rec = 5 ´
105 ).
Properties of water at a film temperature of 20°C are: r = 1000 kg/m3,
n = 1.006 ´ 10 –6 m2/s, Pr = 7.02, k = 0.5978 W/ (mK) and Cp = 4178 J/(kgK).
3. Consider water flowing at 30°C over a flat plate 1 m x 1m size, maintained at 10°C with a free stream velocity of
0.5 m/s. Plot the variation of local heat transfer coefficient along the length if heating starts from 0.25 m from
the leading edge.
4. Air at a pressure of 3 atm. and 200°C flows over a flat plate (1 m long ´ 0.3 m wide), at a velocity of 7 m/s. If the
plate is maintained at 40°C, find out the rate of heat removed continuously from the plate. [Hint: heat is re-
moved from both the surfaces of the plate. Properties k, m, Pr do not vary much with pressure, but, r varies as
per the Ideal gas law, viz. r = p/(R.T), temperature in Kelvin.]
Properties of air at 1 atm. and a film temperature of 120°C are: r = 0.898 kg/m3,
n = 25.45 ´ 10 –6 m2/s, Pr = 0.686, k = 0.03338 W/ (mK) and Cp = 1009 J/(kgK).
5. In problem 4, apply the Colburn analogy to estimate the drag force exerted on the plate.
6. Dry air at at atmospheric pressure and 30°C is flowing with a velocity of 2 m/s along the length of a flat plate,
(size: 1 m ´ 0.5 m), maintained at 90°C.
Using Blasius exact solution, calculate the the heat transfer rate from:
(a) the first half of the plate (b) full plate, and (c) next half of plate.
7. Air at 25°C and atmospheric pressure is flowing with a velocity of 2.5 m/s along the length of a flat plate,
maintained at 55°C. Calculate:
(i) hydrodynamic boundary layer thickness at 20 cm and 40 cm from the leading edge by the approximate
method (ii) mass entrainment rate between these two sections assuming a cubic velocity profile, and (iii) heat
transferred from the first 40 cm of the plate.
8. An air stream at 20°C and atmospheric pressure, flows with a velocity of 4 m/s over an electrically heated flat
plate (size: 0.6 m ´ 0.6 m), heater power being 1 kW. Calculate:

474 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i) the average temperature difference along the plate (ii) heat transfer coefficient, and (iii) temperature of the
plate at the trailing edge
9. Sodium–Potassium alloy (25% + 75 %), at 250°C , flows with a velocity of 0.5 m/s over a flat plate (size: 0.3 m ´
0.1 m), maintained at 550°C. Calculate:
(i) the hydrodynamic and thermal boundary layer thicknesses (ii) local and average value of friction coefficient
(iii) heat transfer coefficient, and (iv) total heat transfer rate
Properties of Na–k alloy at a film temerature of 400°C are: r = 775 kg/m3, n = 0.308 ´ 10 –6 m2/s, Pr = 0.0108, k
= 22.1 W/ (mK) and Cp = 1000.6 J/(kgK).
10. A thin plate of length 2 m and width 1.5 m is exposed to a flow of air parallel to its surface along the 2 m side.
The velocity and temperature of the free stream flow of air are 3 m/s and 20°C respectively. The plate surface
temperature is 90°C. Determine the lengthwise mean local heat transfer coefficient at the end of the plate and
the amount of heat transferred. Take the following properties of air at 20°C:
n = 15.06 ´ 10 –6 m2/s, Pr = 0.703, k = 0.0259 W/ (mK), and use the relation:
Nuavg = 0.664. Re L0.5. Pr 1/3. [M.U.]
11. A 10 cm diameter steam pipe, whose surface is at 90°C passes through an area where the wind is blowing across
the pipe at a velocity of 40 km/h at a temperature of 10°C and pressure of 1 atm. Determine the rate of heat loss
from the pipe per unit length.
12. A 6 mm diameter electrical cable carries a current of 60 Amp and its resistance is 0.002 ohm/metre. Determine
the surface temperature of the cable if air at a temperature of 10°C blows across the cable with a velocity of 50
km/h.
13. An incandescent bulb (60 W) can be considered as a sphere of 10 cm diameter. Only 10% of the energy supplied
is converted to light and the remaining 90% of the energy is converted to heat. If air at 20°C blows across the
bulb with a velocity of 2.5 m/s, determine the equilibrium temperature of the glass bulb.
14. A sphere suspended in an air stream is used as speed measuring device. A 12 mm diameter sphere, when
suspended in an air stream flowing at 40°C, maintains a surface temperature of 50°C, while dissipating an
electrical energy of 0.6 W. Calculate the air speed.
15. A 1.5 cm diameter ball bearing at a temperature of 100°C is cooled by passing water at a temperature of 15°C at
0.3 m/s over it. Calculate the value of average surface heat transfer coefficient between the ball bearing and
water.
16. In a packed bed heat exchanger, air is heated from 30°C to 370°C by passing it through a 10 cm diameter pipe,
packed with spheres of 6 mm diameter. The flow rate is 18 kg/h. Pipe surface temperature is maintained at
420°C. Determine the length of bed required. (Hint: In this case, heat is transferred between the gas and the
walls of the bed).
17. In a regenerator, (1 m dia ´ 2 m long), spherical rock fillings of diameter = 25 mm, are used to heat up air. Void
fraction of this bed is 40%. Initially, the rock fillings are at 25°C and the air is at 85°C, flowing in the axial
direction with a flow rate of 1.2 kg/s. Calculate the value of heat transfer coefficient (Hint: In this case, heat is
transferred between the gas and the spherical fillings).
18. A tube 15 mm ID is maintained at a constant temperature of 60°C. Water is flowing inside the tube at a rate of
10 g/s. Temperature of water at entry is 20°C and at a distance of 1 m from entry the temperature is 40°C.
Compute the average value of Nusselt number using the appropriate correlations.
19. Water at 20°C flows through a 15 mm ID, 4 m long tube with a velocity of 2 m/s. Tube wall is maintained at a
constant temperature of 90°C. What is the heat transfer coefficient and the total amount of heat transferred, if
the exit temperature of water is 60°C? Also, calculate the pressure drop.
20. If, in problem 15, three Globe valves are introduced in the pipe line, what will be the new pressure drop value?
21. Water at 20°C flows through a 15 mm ID, 4 m long tube with a velocity of 2 m/s. Tube wall is maintained at a
constant heat flux by electrical heating. What is the heat transfer coefficient and the total amount of heat trans-
ferred, and the temperature of tube wall at the exit, if the exit temperature of water is 60°C?
22. In a heat exchanger, water flows through a long 2.2 cm ID copper tube at a bulk velocity of 2 m/s and is heated
by steam condensing at 150°C on the outside of the tube. The water enters at 15°C and leaves at 60°C. Find the
heat transfer coefficient for water. Use the empirical relation: Nu = 0.023.Re 0.8.Pr 0.4. Physical properties of water
at the mean bulk temperature of 37.5°C are: r = 990 kg/m 3, m = 0.00069 kg/(ms), Pr = 0.0108, k = 0.63 W/(mK)
and Cp = 4160 J/(kgK). [M.U.]
23. A water heater consists of a thick walled tube of 20 mm ID and 40 mm OD, insulated on the outside surface.
Electrical heating within the wall provides a uniform heat generation rate of 5 ´ 105 W/m3. Water at a rate of
0.15 kg/s enters at 20°C and leaves at 70°C. Calculate the length of tube required. What is the local heat transfer
coefficient at the exit, if the inner wall surface temperature at exit is 80°C?

FORCED CONVECTION 475


24. Water is flowing through a tube of 6 mm ID at a rate of 4 kg/s. A constant heat flux of 250 W per metre length
is provided at the surface. If the water enters at 20°C and exits at 70°C, what is the length of tube required? Also,
what is the surface temperature at exit?
25. Liquid sodium is to be heated from 170°C to 230°C at a rate of 2 kg/s in a 2.5 cm diameter tube, heated electri-
cally on its surface. (Constant heat flux). Calculate the length of tube required if the wall temperature is not to
exceed 280°C. Properties of sodium at average bulk temperature of 200°C are: r = 903 kg/m3, n = 0.506 ´ 10 –6
m2/s, Pr = 0.0075, k = 81.41 W/(mK) and Cp = 1327.2 J/(kgK).
26. A square duct of 20 cm side carries cool air at 10°C over a length of 25 m. Average velocity at entrance is 1.5 m/
s. If the walls of duct are maintained at 30°C, determine the outlet temperature of air.
27. Water flows through a rough pipe of 40 mm ID and 3 m length. Relative roughness, (e/D) for pipe = 0.004. Inlet
temperature of water is 20°C and the inlet flow velocity is 1.5 m/s. Determine the outlet temperature and also
the pressure drop.
28. Consider a tube bank, made of tubes of 10 mm OD, in an in-line arrangement, longitudinal spacing and trans-
verse spacing being 15 mm and 17 mm respectively. Air is heated from 20°C to 40°C by pumping it through this
tube bank. Air approaches the tube bank with a velocity of 4 m/s, and the tube walls are maintained at a
constant temperature of 150°C. If there are 10 tube rows, what is the average heat transfer coefficient and the
pressure drop?
29. Water at 20°C flows across a tube bundle at a free stream velocity of 20 m/s. OD of the tubes is 8 cm. Longitu-
dinal and transverse spacings are 22.5 cm each. Tubes are in a staggered arrangement. If the tube surfaces are
maintained at 50°C, estimate the heat transfer coefficient.
30. Engine oil is to be cooled from 150°C to 90°C in an annulus of 15 mm ID and 30 mm OD. Flow velocity is 1 m/s.
Temperature of inside tube wall is maintained at 25°C. Determine the heat transfer coefficient and the length of
tube required. Properties of engine oil at a mean bulk temperature of 120°C are: r = 828 kg/m3, n = 12 ´ 10 –
6
m2/s, Pr = 175, k = 0.1349 W/ (mK) and Cp = 2307 J/(kgK).
31. Water is flowing at the rate of 20 kg/min. through a tube of inner diameter 2.5 cm. The surface of the tube is
maintained at 100°C. If the temperature of water increases from 25°C to 55°C, find the length of tube required.
Following empirical relation can be used:
Nu = 0.023.Re 0.8.Pr 0.4. Physical properties of water can be taken from the following table: [M.U.]

t, (deg. C) r, (kg/m3) Cp, (J/(kgK)) K ´ 102, (W/(mK)) m ´ 106, (kg/(ms))

40 992.2 4174 63.35 652


50 988.1 4178 64.74 550
60 983.2 4182 65.90 470
70 977.8 4187 66.72 405
80 971.8 4195 67.41 355

476 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

10
Natural (or Free)
Convection

10.1 Introduction
In the previous chapter, we studied heat transfer by forced convection, wherein fluid movement was caused by
an external agency such as a pump or fan. In this chapter, we shall study about heat transfer in ‘Natural or free
convection’; here, fluid movement is caused because of density differences in the fluid due to temperature differ-
ences, under the influence of gravity. Density differences cause a ‘buoyancy force’ which in turn, causes the fluid
circulation by ‘convection currents’. Buoyancy force is the upward force exerted by a fluid on a completely or
partially immersed body and is equal to the weight of the fluid displaced by the body. Obviously, fluid velocity in
natural convection is low as compared to that in forced convection, and as a result, the heat transfer coefficient is
also lower in the case of natural convection. Still, natural convection is one of the important modes of heat
transfer used in practice since there are no moving parts and as a result, there is an increased reliability. Natural
convection heat transfer is extensively used in the following areas of engineering:
(i) cooling of transformers, transmission lines and rectifiers
(ii) heating of houses by steam or electrical radiators
(iii) heat loss from steam pipe lines in power plants and heat gain in refrigerant pipe lines in air-conditioning
applications
(iv) cooling of reactor core in nuclear power plants
(v) cooling of electronic devices (chips, transistors, etc.) by finned heat sinks.

10.2 Physical Mechanism of Natural Convection


Consider the familiar example of a heated, vertical plate kept hanging in quiescent air. Let the temperature of the
heated surface be Ts and that of the surrounding air, Ta. A layer of air in the immediate vicinity of the plate will
get heated by conduction; density of this heated air layer decreases (since the total pressure of surroundings is
constant and p = rRair T for an ideal gas). As a result, the heated layer rises up and the cold air from the sur-
roundings moves in to take its place. This layer, in turn, gets heated up, moves up and is again replaced by cooler
air etc. Thus, convection currents are set up causing the heat to be carried away from the hot surface. This
situation is shown in Fig. 10.1 (a).
During the temperature induced flow, a boundary layer is set up along the length of the plate as shown.
With the x-axis taken along the vertical length of plate, and the y-axis perpendicular to it, the velocity and tem-
perature profiles are shown in the Fig. 10.1. As far as the velocity profile is concerned, at the plate surface, the
fluid velocity is zero due to ‘no slip’ condition; then, the velocity increases to a maximum value and then, drops
to zero at the outer edge of the boundary layer since the surrounding air is assumed to be quiescent. Note the
difference in this velocity profile as compared to that in the case of forced convection. The boundary layer is
laminar for some distance along the length, and then depending on the fluid properties and the driving tempera-
ture difference between the wall and the ambient, the boundary layer becomes turbulent.
x

y
Ts
Temperature profile

Ta
Velocity profile [P + (¶P/¶x).dx].dy

U=0 U=0
Boundary layer
tdx dx [t + (¶t/¶y).dy].dx
Ts Stationary fluid at Ta dy

Hot, vertical plate


P.dy
r.g.dx.dy

FIGURE 10.1(a) Velocity and temperature FIGURE 10.1(b) Differential control volume with
profiles in natural convection forces for natural convection

Analytical solution of natural convection heat transfer is a little more complicated since velocity field is
coupled to the temperature field because the flow is induced by temperature differences. Temperature field is
also coupled to the velocity field, i.e. we say that the velocity and temperature fields are ‘mutually coupled’.
Therefore, to get a solution, momentum and energy equations for the boundary layer have to be solved simulta-
neously. Solutions by the exact and approximate integral methods have been obtained for the simple cases, but
the predicted surface heat transfer coefficients are smaller than the experimentally measured values, because the
analysis do not take into account rate-increasing disturbances (remember: velocities are quite small in free con-
vection) present in actual equipments. Therefore, in handling natural convection problems, we rely mostly on
empirical relations derived as a result of large experimental work.

10.3 Dimensional Analysis of Natural Convection—Grashoff Number


Natural convection heat transfer is a good candidate for dimensional analysis since we can reliably list the pa-
rameters on which this phenomenon depends and the theoretical analysis is rather difficult and we have to
depend mostly on experimental work.
As we stated earlier, in natural convection, flow is induced by the density differences caused as a result of
temperature differences. In the gravitational field, the density differences induce a buoyancy force given by:
Fb = r fluid ×g×V body ...(10.1)
where rfluid = density of fluid
g = acceleration due to gravity
V body = volume of portion of body immersed in fluid
In the absence of other body forces (such as centrifugal, electromagnetic etc.),
Net vertical force acting on the body = weight of body – buoyancy force
i.e. Fnet = W – Fb
i.e. Fnet = r body ×g×V body – rfluid ×g×V body
i.e. Fnet = (r body – rfluid)×g×V body ...(10.2)
Now, the density differences can be related to the temperature differences by the temperature coefficient of
volumetric expansion, b, which is defined as:
b = (1/v).(¶v/¶T)p = –(1/r).(¶r/¶T)p ....(1/K) ...(10.3)
i.e. b = – (1/r).(Dr/DT) at constant P
i.e. Dr = – r.b.DT at constant P
For ideal gas, p = r.R.T and, b = 1/T, where T is expressed in Kelvin.

478 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 10.1 Parameters for Natural convection heat transfer

Sl. No. Parameter Symbol (Unit) Primary dimension


1 Significant length L c (m) L
2 Fluid density r (kg/m3) ML –3
3 Fluid viscosity m (kg/(m.s)) ML –1t–1
4 Temperature difference D T (deg.C) T
5 Coefficient of volume expansion b (1/K) T –1
6 Acceleration due to gravity g (m/s2) Lt –2
7 Th. conductivity of fluid k (W/(m.K)) MLt– 3 T–1
8 Heat transfer coefficient h (W/(m2.K)) Mt –3 T–1
9 Specific heat of fluid Cp (J/(kg.K)) L2 t –2 T –1

With this background, now let us list out the parameters (and their primary dimensions) on which the
phenomenon of Natural convection heat transfer depends, as shown in Table 10.1.
We see that there are 9 variables listed. Of these, the product b.g.DT represents buoyancy forces and is
considered as a single variable. Thus, we can say that there are 7 variables affecting the phenomenon and there
are 4 primary dimensions, viz. (M, L, T and t).
Then, from Buckingham theorem, Number of independent dimensionless groups that can be formed is
equal to 7 – 4 = 3.
Choosing Lc, r, m and k as the core group, we write:
p 1 = Lca × r b × m c × k d ×(g×b ×DT) ...(a)
p 2 = Lcp ×rq×mr ×k s ×Cp ...(b)
p 3 = Lcw ×rx ×m y ×k z ×h ...(c)
(i) Considering Eq. a:
p 1 = M 0 ×L0 ×T0 ×t 0 = La ×(M×L –3)b ×(M×L –1 ×t – 1)c×(M×L×t –3 ×T– 1)d ×(L×t –2)
Equating the coefficients of M, L, T and t on either side of above equation., we get:
M: 0 = b + c + d
L: 0 = a – 3 b – c + d + 1
T: 0 = –d
t: 0 = –c – 3 d – 2
Solving the above set of equations simultaneously, we get:
d = 0, c = –2, b = 2, a = 3
Therefore, p 1, the first dimensionless group is:

r 2 × ( g × b × D T ) × L3c ( g × b × D T ) × L3c
p1 = = = Gr
m2 n2
where Gr = Grashoff number
(ii) Considering Eq. b:
p 2 = M 0 ×L0 ×T0 ×t0 = Lp ×(M×L –3)q ×(M×L –1 ×t – 1)r ×(M×L×t–3×T –1)s×(L2×t –2 T –1)
Equating the coefficients of M, L, T and t on either side of above equation., we get:
M: 0 = q + r + s
L: 0 = p – 3 q – r + s + 2
T: 0 = – s – 1
t: 0 = –r – 3 s – 2
Solving the above set of equations simultaneously, we get:
s = –1, r = 1, q = 0, p = 0
Therefore, p 2, the second dimensionless group is:
m ×Cp
p2 = = Pr
k
where, Pr = Prandtl number

NATURAL (OR FREE) CONVECTION 479


(iii) Considering Eq. c:
p 3 = M 0 ×L0 ×T 0 ×t 0 = Lw ×(M×L –3)x ×(M×L –1 ×t –1)y ×(M×L×t –3 ×T– 1)z ×(M×t –3 ×T– 1)
Equating the coefficients of M, L, T and t on either side of above equation, we get:
M: 0 = x + y + z + 1
L: 0 = w – 3 x – y + z
T: 0 = –z – 1
t: 0 = –y – 3 z – 3
Solving the above set of equations simultaneously, we get:
z = –1, y = 0, x = 0, w = 1
Therefore, p 3, the third dimensionless group is:
h×L
p3 = = Nu
k
where, Nu = Nusselt number
Thus, for natural convection, we have:
Nu = f(Gr, Pr)
Of course, the exact form of equation with associated constants must be determined from experiments.
As we shall see later, in most of the empirical relations, product of Gr and Pr is taken together and the
relations are presented in the form:
Nu = C.Ram
where
Ra = (Gr.Pr) = Rayleigh number and ‘C’ and ‘m’ are constants determined from experiments.
By determining Nu, we determine h, the heat transfer coefficient in natural convection.
Then, the heat transfer rate for natural convection is given by Newton’s law of cooling, i.e.
Qconv = h.A.(Ts – Ta), W
Thus, from dimensional analysis, we have established that, in natural convection heat transfer problems,
dimensional groups of significance are: Grashoff number (Gr), Prandtl number (Pr) and Nusselt number (Nu).
Out of these, Gr plays the same role in natural convection as that of Reynolds number in forced convection.
Physical significance of Grashoff number is that it represents the ratio of buoyancy force to the viscous force
acting on the fluid, i.e.
Buoyancy forces g ×D r V g × b × D T ×V
Gr = = =
Viscous forces r ×n 2
n2
since Dr = r.b.DT.
So, we can write:

g × b × (Ts - Ta ) × L3c
Gr = ...(10.4)
n2
where,
g = acceleration due to gravity, m/s2
b = coefficient of volume expansion, 1/K (b = 1/T for ideal gases only, T in Kelvin)
Ts = temperature of the surface, deg. C
Ta = temperature of the fluid at sufficient distance from the surface, deg. C
Lc = characteristic length of the geometry, m, and
n = kinematic viscosity of fluid, m2/s
Product of Grashoff number and Prandtl number, i.e. Rayleigh number, Ra = Gr.Pr is the criterion to deter-
mine if the flow is laminar or turbulent, in natural convection. For example, in the case of heat transfer by natural
convection for vertical plates, for Ra > about 109, the flow is turbulent and for Ra < 109, the flow is laminar.

10.4 Governing Equations and Solution by Integral Method


As stated earlier, in solving natural convection heat transfer problems, we rely more on empirical relations than
on analytical relations. This is due to the fact that analytical relations are rather difficult to obtain since the

480 FUNDAMENTALS OF HEAT AND MASS TRANSFER


momentum and energy equations are ‘mutually’ coupled; also, analytical relations generally give lower values of
heat transfer coefficients as compared to empirical relations. Also, because of the very low velocities involved in
natural convection, it becomes difficult to take into account all factors in the analytical methods and the insertion
of measuring probes itself introduces disturbances in the flow fields.
So, we shall indicate the development of governing equations for the simple case of heat transfer by free
convection from a heated, vertical plate at a surface temperature Ts and the surrounding ambient at a tempera-
ture Ta, and very briefly give the outline of the solution by the approximate Integral method and give only the
final results.
In general, solution of the problem involves the simultaneous solution of equations of continuity, momen-
tum and energy.
Equation of continuity:
Considering an elemental volume in the two-dimensional boundary layer, with constant properties, the continu-
ity equation remains the same as derived for forced convection, i.e. Eq. 9.15. We rewrite it here as:
(¶u/¶x) + (¶v/¶y) = 0 ...(10.5)
Equation of momentum:
This is derived by applying Newton’s second law to the differential control volume. Fig. 10.1 (b) shows the
various forces acting on the control volume. Net force acting in the x-direction (= S Fx) must be equal to the rate
of change of momentum in that direction.
i.e. S Fx = r.{u.(¶u/¶x) + v.(¶u/¶y)}.dx.dy ...(Eq. A)
See Eq. a under section 9.7.2
As far as S Fx is concerned, compared to the case of forced convection, now there an additional force due to
gravity, acting in the downward direction i.e. opposed to the positive X-direction; so, we get
S Fx = – (¶p/¶x).dx.dy – r.g.dx.dy + m(¶ 2u/¶y2).dx.dy ...(Eq. B)
See Eq. b under section 9.7.2
Equating Eqs. A and B, we get:
r.{u.(¶u/¶x) + v.(¶u/¶y)} = –(¶p/¶x) – r.g + m(¶ 2u/¶y2) ...(10.6)
Since the pressure gradient in X-direction is due to change in elevation of plate, we write:
(¶p/¶x) =– ra.g
Therefore, Eq. 10.6 becomes:
r.{u.(¶u/¶x) + v.(¶u/¶y)} = g.(ra – r) + m(¶ 2u/¶y2) ...(10.7)
Now, the density difference (ra – r) may be related to the temperature difference as follows:
b = (1/V).(¶V/¶T)p = (1/V).{(V – V a)/(T – Ta)} = (ra – r) /{r.(T – Ta)}
So, we get:
r.{u.(¶u/¶x) + v.(¶u/¶y)} = g.r.b(T – Ta) + m(¶ 2u/¶ y2) ...(10.8)
Eq. 10.8 is the momentum equation for natural convection boundary layer; note that its solution requires a
knowledge of temperature distribution.
In general, volume coefficient of expansion, b for fluids has to be obtained from data tables; however, for
ideal gases, b = 1/T, where T is the absolute temperature in Kelvin.
Equation of energy:
Again, equation of energy remains the same as derived earlier for forced convection, equation 9.18. We rewrite it
here as:
u.(¶T/¶x) + v.(¶T/¶y) = a.(¶ 2T/¶y 2). ...(10.9)
While solving this problem by the approximate integral method, we make an assumption that the fluid is
incompressible except for the effect of variable density in the buoyancy force, since fluid motion is induced by
this variation. This is known as Boussinesq approximation. And, the flow is with laminar boundary layer, steady,
two-dimensional and with constant fluid properties. Further, since it is the temperature difference that induces
the flow in natural convection, both the hydrodynamic and thermal boundary layers are assumed to be identical,
i.e. thicknesses of both the boundary layers are assumed to be equal, or d = dt.
Integrating Eq. 10.8 over the boundary layer thickness, we get the integral momentum equation:
d
dx
F
GH z
0
d I
JK
r × u2 dy = t s +
z
0
d
r × g × b × (T - Ta ) dy

NATURAL (OR FREE) CONVECTION 481


i.e.
d
dx
F
GH z
0
d I
JK
r × u2 dy = – m×
F du I
GH dy JK y=0
+
z d

0
r × g × b × (T - Ta ) dy ...(10.10)

To solve this, we have to assume the velocity and temperature distributions which satisfy the boundary
conditions, just as we did in the case of forced convection.
For temperature distribution, the boundary conditions are:
T = Ts at y = 0
T = Ta at y = d
and, (¶T/¶y) = 0 at y = d
And the temperature distribution which satisfies these conditions is:

T - Ta y FG IJ 2

Ts - Ta
= 1-
d H K ...(10.11)

Boundary conditions for the velocity profile are:


u = 0 at y = 0
u = 0 at y = d
and, (¶u/¶y) = 0 at y = d
An additional condition from Eq. 10.8 is:
(¶ 2u/¶y2) =–g.b.(Ts – Ta)/n at y = 0 since both u and v are zero at the surface.
And the velocity profile which satisfies these conditions is:

u y y FG IJ 2

ux
= × 1-
d d H K ...(10.12)

Here, ux is a fictitious reference velocity, an arbitrary function of x, since there is no free stream velocity in
natural convection.
Maximum velocity and its position is determined by differentiating Eq. 10.12 w.r.t y and equating to zero.
The result is:
4
u max = ×ux at y = d/3 ...(10.12a)
27
And, the mean velocity at a section is obtained by integrating the velocity function over the boundary layer
thickness:

um =
1 d
z 1
× udy = ×
d 0 d z d

0
ux ×
FG y IJ ×FG 1 - y IJ
Hd K H d K
2
dy

1 27
i.e. um = ×ux = ×umax ...(10.12b)
12 48
Inserting Eqs. 10.11 and 10.12 in Eq. 10.10 and performing the mathematical operations, one gets:

1 d (ux2 ×d ) 1 n × ux
× = × g × b × (Ts - Ta ) ×d = ...(10.13)
105 dx 3 d
Similarly, integrating Eq. 10.9, we get the integral form of energy equation as follows:

d
dx
LM
N z
0
d
u ×(T - Ta ) dy = – a ×
OP
Q
F dT I
GH dy JK y=0
...(10.14)

Substituting the assumed velocity and temperature distributions in Eq. 10.14, final result is:
1 d (ux ×d ) (Ts - Ta )
×(Ts – Ta)× = 2×a × ...(10.15)
30 dx d
Assuming exponential functional variations for ux and d, i.e. ux = C 1.x 1/2 and d = C 2.x 1/4, we get the final
result for velocity function and boundary layer thickness in laminar flow as:

482 FUNDAMENTALS OF HEAT AND MASS TRANSFER


ux = 5.17×n×(Pr + 0.952) –0.5 ×
LM b × g ×(T - T ) OP
s a
0. 5
×x0.5
N n Q 2
...(10.16a)

and,

d 3.93× (0.952 + Pr )0 . 25
= ...(10.16b)
x Grx0. 25 × Pr 0. 5
Eq. 10.16b gives the variation of d along the height x of the plate. Grx is the Grashoff number.
Mass flow rate through the boundary:
Mass flow rate through a section for unit width of plate is given by:
ux r
m = um ×(d×1)×r = ×d×r = ×(d×ux) ...(10.17)
12 12
ux and d are obtained from Eq. 10.16 a & b.
Mass flow between two sections at x 1 and x 2 can be determined by the difference in values of m (as calcu-
lated from Eq. 10.17) between these two sections.
Total mass flow through the boundary is obtained by putting x 1 = 0 and x 2 = L. We get:

m total = 1.7×r ×n×


LM Gr L
OP 0. 25
...(10.18)
MN Pr ×(Pr + 0.952) PQ
2

Heat transfer coefficient is determined from


F dT I
qs = –k×A× GH dy JK s
= h×A×(Ts – Ta)

Using the temperature distribution given by Eq. 10.11, we get:


2×k
h=
d
h×x x
i.e. = Nux = 2 ×
k d
And the dimensionless heat transfer coefficient (i.e. Nusselts number) is:

0. 508 k 2 Pr 0. 5 × Grx0 . 25
Nux = (using eqn. 10.16(b))...(10.19)
(0.952 + Pr )0 . 25
Average heat transfer coefficient for the vertical plate is obtained by integrating over the height L:

h avg =
1 L
L 0
4
z
× hx dx

i.e. h avg = ×hL ...(10.20a)


3
0. 25
4 0.667 × Pr 0 . 5 × GrL
and, Nu avg = ×NuL = ...(10.20b)
3 ( 0.952 + Pr )0 . 25
Note that for forced convection over a flat plate, we had Nuavg = 2. NuL
Above equations are valid for laminar boundary layer flow only
For turbulent boundary layer flow, (Gr × Pr > 109), by following the integral method, we get:

F 2 I
GG
0. 565 × 1 + 0 494 × Pr 3 JJ
d turb
=
H K (Gr ×Pr = Ra = Rayleigh number)...(10.21)
8
x 0. 1
Gr × Pr 15
NATURAL (OR FREE) CONVECTION 483
and,

havg × L
LM 1.17
OP 0.4

= 0.0246 M PP
Pr × GrL
Nu avg = (for turbulent flow...(10.22))
k MMN 1 + 0.495 × Pr 2
3
PQ
In the above equation physical properties of fluid are taken at the average (film) temperature, i.e.
Ts + Ta
Tf =
2
An outline of the analytical procedure involved for the simple case of heat transfer by convection from a
heated vertical plate is presented above just to illustrate the fact that even for simple cases, analytical procedures
are rather involved. It is stated again, that this is due to the mutual coupling of momentum and energy equa-
tions.

10.5 Empirical Relations For Natural Convection


Over Surfaces and Enclosures
Free convection patterns from a few common geometries are shown in Fig. 10.2.

Cold plate

Hot Cold

(b) Hot surface facing up (c) Cold surface facing up

Hot Cold

(d) Hot surface facing down (e) Cold surface facing down
(a) Cold vertical surface
FIGURE 10.2 Free convection flow patterns

We shall present below empirical relations for natural convection from several types of surfaces and enclo-
sures of practical importance. While using the empirical relations, it is important to remember the conditions
under which these relations are valid. Observe that most of the relations are presented in the form: Nu = C.Ram,
where C and m are constants deduced from experiments. Nu is the Nusselt number (= h.Lc/k), Ra is the Rayleigh
number (= Gr.Pr); characteristic dimension Lc for vertical plates and cylinders is generally the plate (or cylinder)
height L or diameter D for a horizontal cylinder.
10.5.1 Vertical Plate at Constant Temperature Ts
Vertical plate is an important geometry since heat transfer from the walls of a furnace can be calculated by the
relations applicable to a vertical plate.
McAdams has suggested the following relations for fluids whose Prandtl number is close to unity, i.e. for air
and other gases, generally:
1
Nu = 0.59× Ra 4 ...104 < Ra < 109...(10.23)
1
and Nu = 0.13× Ra 3 ...109 < Ra < 1012...(10.24)
Eq. 10.23 is for laminar, boundary layer type, natural convection flow, while Eq. 10.24 is for turbulent,
boundary layer type, natural convection flows. Fluid properties are evaluated at film temperature Tf, already
defined.

484 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Churchill and Chu present following relations for the entire range of Ra and also valid for all Prandtl num-
bers from 0 to ¥.
For 0 < Ra < 109, 0 < Pr < ¥:
1
0.670 × Ra 4
Nu = 0.68 + (0 < Ra < 109...(10.25))
LM F 0.492 I OP
4
9 9

MM1 + GH Pr JK PP
16

N Q
For Ra > 109, 0.6 < Pr < ¥:
1
0.15× Ra 3
Nu = (Ra > 109...(10.26))
LM F 0.492 I OP
16
9 27

MM1 + GH Pr JK PP
16

N Q
For Ra > 109, 0 < Pr < 0.6:

LM OP 2

MM 1
PP
M
Nu = M0.825 +
0.387 × Ra 6 PP (Ra > 109...(10.27))
MM LM F 0.492 I 9 OP
8
27 PP
MM MM1 + GH Pr JK PP PP
16

N N Q Q
Eq. 10.25 is for fluids whose Pr is not too close to unity (or, to that of air). Eq. 10.26 is for high Prandtl no.
fluids, and Eq. 10.27 is for low Pr fluids i.e. for liquid metals.
In the above equations characteristic dimension for Nu and Ra is the height L of the plate; fluid properties
are evaluated at the film temperature Tf .
For inclined plates (inclined at an angle q to the vertical), vertical plate relations can be used by replacing g
by g.cos(q) for Ra < 109. Inclined length L is the characteristic dimension.
10.5.2 Vertical Cylinders At Constant Temperature Ts
A vertical cylinder can be treated as a vertical plate and the relations given above can be applied if the following
criterion is satisfied:
D 34
³ ...(10.28)
L 1
Ra 4
Height L of the cylinder is the characteristic dimension.
10.5.3 Vertical Plate With Constant Heat Flux
Equations of Churchill and Chu, 10.25 and 10.26 are valid, with the following modifications: (a) temperature of
the constant flux plate is considered at a point mid-way between top and bottom (b) constant 0.492 should be
changed to 0.437.
Alternative relations are given below for vertical and inclined plates for natural convection in water and air.
Here, a modified Grashoff number, Gr’ is defined:

g × b × qs × x 4
Gr ¢ = Gr×Nux = ...(10.29)
k ×n 2
NATURAL (OR FREE) CONVECTION 485
where qs is the wall heat flux in W/m2. Then the following two relations are recommended for local heat transfer
coefficients in laminar and turbulent ranges respectively:
Nux = 0.60×(Gr ¢×Pr)0.2 (105 < Grx¢ < 1011...(10.30))
and, Nux = 0.17×(Gr¢×Pr)0.25 (Grx¢ > 1011...(10.31))
And the average heat transfer coefficient in the laminar region is obtained by integration over the entire
height L of the plate as:
5
h= ×h L (for laminar...(10.32))
4
and, for turbulent region, hx is independent of x:
h = hL (for turbulent...(10.33))
Example 10.1. A hot plate 30 cm high and 1.2 m wide at 140°C is exposed to ambient air at 20°C. Using the approximate
solution, calculate the following:
(i) Maximum velocity at 12 cm from the leading edge of the plate (ii) boundary layer thickness at 12 cm from the
leading edge of plate (iii) local heat transfer coefficient at 12 cm from the leading edge of the plate (iv) average heat
transfer coefficient over the surface of the plate (v) total mass flow through the boundary (vi) total heat loss from the
plate, and (vii) temperature rise of air
Solution.
Data:
L := 0.3 m W := 1.2 m Ts := 140°C Ta := 20°C x := 0.12 m g := 9.81 m/s2
We need properties of air at film temperature Tf = (140 + 20)/2
Tf := 80°C
Properties of air at 80°C:
r := 1.00 kg/m3 n := 21.09 ´ 10 –6 m2/s Pr := 0.692 k := 0.03047 W/(mK) Cp := 1009 J/(kgK)
1
b := (coefficient of volume expansion...Note that temperature must be in Kelvin)
(Tf + 273)
i.e. b = 2.833 ´ 10 – 3 1/k
Grashoff number:
g × b × (Ts - Ta ) × x 3
At x = 0.12 m: Grx :=
n2
i.e. Grx = 1.296 ´ 10 7
g × b × (Ts - Ta ) × L3
At L = 0.3 m: GrL :=
n2
i.e. GrL = 2.024 ´ 10 8
(i) Maximum velocity
To calculate this, first we need the velocity function ux. We have from Eq. 10.16a:

ux := 5.17 ×n×(Pr + 0.952)–0.5 ×


LM b × g ×(T - T ) OP
s a
0.5
×x 0.5
N n Q 2
...(10.16a)

i.e. ux = 2.551 m/s (velocity function)


Therefore, maximum velocity is given by:
4
umax := × ux ...(10.12a)
27
i.e. umax = 0.378 m/s
(ii) Thickness of boundary layer:
We have:
d 3. 93 × ( 0. 952 + Pr ) 0 . 25
= ...(10.16b)
x Grx0 . 25 × Pr 0 . 5

3. 93 × ( 0. 952 + Pr ) 0 . 25
i.e. d := ×x
Grx0 . 25 × Pr 0 . 5
i.e. d = 0.0107 m = 10.7 mm.

486 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(iii) Heat transfer coefficient
We have:
0 . 508 × Pr 0 . 5 × Grx0 . 25
Nux := ...(10.19)
( 0.952 + Pr ) 0 . 25
i.e. Nu x = 22.39 (Check Nux = 2.x/d = 22.43...checks)
Nux × k
Therefore, hx :=
x
i.e. hx = 5.685 W/(m2K) (heat transfer coefficient)
Heat transfer coefficient at x = L:
NuL × k
hL =
L
0 .508 × Pr 0 . 5 × GrL0 . 25 k
i.e. hL := ×
(0 .952 + Pr ) 0 . 25 L
i.e. hL = 4.521 W/(m2 K) (heat transfer coefficient at x = L )
(iv) Average value of heat transfer coefficient:
4
havg :=
×hL ...(10.20a)
3
i.e. h avg = 6.028 W/(m2 K) (average heat transfer coefficient)
(v) Total mass flow rate through boundary layer:
We have, from Eq. 10.18

m total := 1.7 ×r×n ×


LM GrL OP 0 . 25

MN Pr 2
× ( Pr + 0. 952) PQ ...(10.18)

i.e. m total = 4.54 ´ 10 –3 kg/s.


(vi) Heat transfer from the plate:
Q= havg ×As ×(Ts – Ta), W (where As = surface area of both the surfaces of plate)
i.e. Q := havg ×(2× L×W) (Ts – Ta ), W
i.e. Q= 520.855 W (total heat transfer from plate.)
(vii) Temperature rise of air:
We have: Q= m total ×Cp × DT
Q
i.e. DT :=
mtotal × Cp
i.e. DT = 113.699 deg. C.
Example 10.2. A furnace door, 1.5 m high and 1 m wide, is insulated from inside and has an outer surface temperature
of 70°C. If the surrounding ambient air is at 30°C, calculate the steady state heat loss from the door.
Solution.
Data:
L := 1.5 m W := 1.0 m Ts := 70°C Ta := 30°C g := 9.81 m/s2
We need properties of air at film temperature Tf = (70 + 30)/2
Tf := 50°C (film temperature)
Properties of air at 50°C:
r := 1.093 kg/m 3 n := 17.95 ´ 10 –6 m2/s Pr := 0.698 k := 0.02826 W/(mK) Cp := 1005 J/kgK)
1
b := (coefficient of volume expansion...Note that temperature must be in Kelvin)
(Tf + 273)
i.e. b = 3.096 ´ 10 –3 1/K
Grashoff number:
g × b × (Ts - Ta ) × L3
At L = 1.5 m: GrL :=
n2
i.e. GrL = 1.273 ´ 1010

NATURAL (OR FREE) CONVECTION 487


Rayleigh number:
Ra := GrL ×Pr
i.e. Ra = 8.882 ´ 10 9
Then, applying Eq. 10.24, we get:
1
Nu := 0.13 ×Ra 3 (109 < Ra <1012...(10.24))
i.e. Nu = 269.227 (Nusselt number)
Nu × k
Therefore, h := W/(m2 K) (heat transfer coefficient)
L
i.e. h = 5.072 W/(m2 K) (heat transfer coefficient)
Heat loss:
Q = h×As ×(Ts – Ta), W (heat loss from outer surface)
i.e. Q := h×(L×W)×(Ts – Ta), W
i.e. Q = 304.334 W (heat loss)
Alternatively, we can apply Eq. 10.26:
1
0.15 × Ra 3
Nu := (Ra > 109...(10.26))
LM F 0.492 I OP
16
9 27

MM1 + GH Pr JK
16

PP
N Q
i.e. Nu = 217.746
Nu × k
Therefore, h := W/(m2 K) (heat transfer coefficient)
L
i.e. h = 4.102 W/(m2 K) (heat transfer coefficient)
And,
Q := h×(L×W)× (Ts – Ta) W
i.e. Q = 246.14 W (heat loss)
Difference between the two values of Q obtained is about 19%
Example 10.3. In a nuclear reactor core, parallel vertical plates, each 2.5 m high and 1.5 m wide, heat liquid Bismuth by
natural convection. Maximum temperature of the plates should not exceed 755°C and lowest allowable temperature of
Bismuth is 320°C. Calculate the maximum heat dissipation from both sides of each plate.
Solution.
Data:
L := 2.5 m W := 1.5 m Ts := 755°C Ta := 320°C g := 9.81 m/s2
We need properties of air at film temperature Tf = (755 + 320)/2
Tf := 537.5°C (film temperature)
Properties of Bismuth at 538°C:
r := 9739 kg/m3 n := 1.08 ´ 10 –7 m2/s Pr := 0.011 k := 15.58 W/(mK)
Cp := 154.5 J/(kgK) b := 0.126 ´ 10 –3 1/k
Note that we cannot put b = 1/(Tf + 273), since Bismuth is a liquid, and not ideal gas. Instead, we should read the
value of b from data tables.
Grashoff number:
g × b × (Ts - Ta ) × L3
At L = 2.5 m: GrL :=
n2
i.e. GrL = 7.203 ´ 1014
Rayleigh number:
Ra := GrL ×Pr
i.e. Ra = 7.923 ´ 1012
Then, applying Eq. 10.27 we get:

488 FUNDAMENTALS OF HEAT AND MASS TRANSFER


LM OP 2

MM 1
PP
Nu := M0 .825 + PP
0. 387 × Ra 6

MM LM F 0.492 I 9
OP
8
PP
(Ra > 109...(10.27))

MM
27

MM1 + GH Pr JK
16

MN PP PP
N Q Q
i.e. Nu = 834.346 (Nusselt number)
Nu × k
Therefore, h := W/(m2 K) (heat transfer coefficient)
L
i.e. h = 5.2 ´ 10 3 W/(m2K) (heat transfer coefficient)
Heat loss:
Q = h×2×As ×(Ts – Ta), W (heat loss from both surfaces of plate)
i.e. Q := h×2×(L ×W)×(Ts – Ta), W
i.e. Q = 1.696 ´ 10 7 W (heat loss.)
i.e. Q = 16.96 MW (heat loss from each plate. )
Example 10.4. A vertical steel plate, 0.4 m ´ 0.4 m in size and 3 mm thick, at an uniform temperature of 180°C, is exposed
to atmospheric air at 20°C. Find the approximate time required for the plate to cool to 30°C, if the heat transfer coeffi-
cient in natural convection for the vertical plate is given by: h = 1.42 ´ (DT/L)1/4. For steel, r = 7800 kg/m3, Cp = 473 J/
(kgK)
Solution.
Data:
L := 0.4 m W := 0.4 m t := 0.003 m r := 7800 kg/m3 Ta := 20°C Ts := 180°C g := 9.81 m/s2
2 2
A := L ×W m i.e. A := 0.16 m Cp := 473 J/(kgK)
At any instant, let the temperature of the plate be T. Then, we can write the heat balance:
Rate of decrease of enthalpy of the plate = rate of instantaneous heat transfer from plate by convection

i.e. – m×Cp dT = h×(2×A)×(T – Ta) ((a)...areas on both the sides of the plate lose heat by convection)
dt
where m is the mass of the plate
Put: q =(T – Ta)
dq dT
Then, =
dt dt
And Eq. a becomes:
- dq 2× h × A
= ×q ...(b)
dt m × Cp
Now, mass of plate: m := (L×W×t)×r kg
i.e. m = 3.744 kg

LM (T - T ) OP
1
a 4
Now, heat transfer coefficient: h = 1.42×
N L Q
1
i.e. h = 1.78556 q 4 , since q = (T – Ta)
Substituting in Eq. b:
5
- dq
= 3.22647 ´ 10 –4 × q 4 ...(c)
dt
Integrating Eq. c:
-1
4× q 4 = 3.22647 ´ 10 –4 ×t + C1 ...(d)
where C 1 is the integration constant.
To find C1, use the initial condition, i.e. at t = 0, q = 180 – 20 = 160°C:

NATURAL (OR FREE) CONVECTION 489


-1
i.e. C1 := 4.160 4
i.e. C 1 = 1.12468
Therefore Eq. d becomes
-1
q 4 = 8.06618 ´ 10–5 ×t + 0.28117 ...(e)
Eq. e gives the temperature of the plate at any time t.
Time required for the plate to reach 30°C:
i.e. T := 30°C
Therefore, q := T – Ta
i.e. q = 10°C
Then, from Eq. e:
-1
q 4 - 0 .28117
t :=
8 .06618 ´ 10 - 5
i.e. t = 3.486 ´ 10 3 s
i.e. t = 0.968 hrs.
Example 10.5. A vertical pipe, 15 cm OD, 1 m long, has a surface temperature of 90°C. If the surrounding air is at 30°C,
what is the rate of heat loss by free convection per metre length of pipe?
(b) If the pipe is inclined to the vertical at an angle of 30 deg. during installation, how does the heat loss/m change?
Solution.
Data:
L := 1.0 m D := 0.15 m Ta := 30°C Ts := 90°C g := 9.81 m/s2
Now, film temperature is (90 + 30)/2 = 60°C.
i.e. Tf := 60°C
Properties of air at 60°C:
n := 18.97 ´ 10 –6 m2/s Pr := 0.696 k := 0.02896 W/(mK)
1
b := (coefficient of volume expansion...Note that Temperature must be in Kelvin
(Tf + 273)
i.e. b = 3.003 ´ 10 –3 1/K
Grashoff number:
g × b × (Ts - Ta ) × L3
GrL :=
n2
i.e. GrL = 4.912 ´ 10 9
RaL := GrL ×Pr
Rayleight number:
i.e. RaL = 3.419 ´ 10 9
Now, to apply the vertical plate correlations for this case of a vertical cylinder, let us confirm the following condi-
tion:
D 34
³ ...(10.28)
L 1
RaL4
D 34
= 0.15 and, = 0.141
L 1
RaL4
Therefore, Eq. 10.28 is satisfied, and we can apply the vertical plate Eq. 10.24:
1
Nu := 0.13 × RaL3 (109 < Ra < 1012...(10.24))
i.e. Nu = 195.836 (Nusselt number)
k
Therefore, h := Nu ×
L
i.e. h = 5.671 W/(m2 K) (heat transfer coefficient)

490 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Heat loss/meter length of pipe:
Q := h×(p×D)×(Ts – Ta) W/m
i.e. Q = 160.356 W/m.
Alternatively:
We can use Eq. 10.27

LM OP 2

MM 1
PP
Nu := M0 .825 + PP
0. 387 × Ra 6

MM LM F 0.492 I 9
OP
8
PP
(ReL > 109...(10.27))

MM
27

MM1 + GH Pr JK
16

MN PP PP
N Q Q
i.e. Nu = 179.503 (compare with Nu = 195.836, obtained using Eq. 10.24)
k
Therefore, h := Nu ×
L
i.e. h = 5.198 W/(m2 K) (heat transfer coefficient)
and,
Q := h×(p×D)×(Ts – Ta) W/m
i.e. Q = 146.981 W/m (compare with 160.356 W/m obtained earlier.)
(b) When the pipe is inclined at 30 deg. to vertical:
q = 30 deg.
But, while using Mathad, arguments to trigonometric functions must be in radians.
p
i.e. q := 30× radians
180
i.e. q = 0.524 radians
1
We use Nu := 0.13×(RaL ×cos(q)) 3
i.e. Nu = 186.668 (Nusselt number)
k
Therefore, h := Nu ×
L
i.e. h = 5.406 W/(m2 K) (heat transfer coefficient)
Heat loss/metre length of pipe:
Q := h×(p×D)×(Ts – Ta) W/m
i.e. Q = 152.849 W/m.
10.5.4 Horizontal Plate at Constant Temperature Ts
Here, the characteristic length to be used in expressions for Nu and Gr is:
Lc = A/P
where, A is the surface area and P is the perimeter.
Property values are evaluated at film temperature, Tf.
(a) Upper surface of a hot plate (or, lower surface of a cold plate):
1
Nu = 0.54 Ra 4 (104 < Ra < 107...(10.34))
and,
1
Nu = 0.15 Ra 3 (107 < Ra < 1011...(10.35))
(b) Lower surface of a hot plate (or upper surface of a cold plate):
1
Nu = 0.27×Ra 4 (105 < Ra < 1011...(10.36))
Example 10.6. A hot, square plate, 50 cm ´ 50 cm, at 100°C is exposed to atmospheric air at 20°C. Find the heat loss from
both the surfaces of the plate:
(i) if the plate is kept vertical
(ii) if the plate is kept horizontal.

NATURAL (OR FREE) CONVECTION 491


Properties of air at mean temperature of 60°C are given below: r = 1.06 kg/m3,
k = 0.028 W/(mK), n = 18.97 ´ 10 –6 m2/s, Cp = 1.008 kJ/(kgK).
Following empirical relations can be used:
Case (i): Nu = 0.13 ´ (Gr.Pr)1/3
Case (ii): Nu = 0.71 ´ (Gr.Pr)1/4 for the upper surface, and
Nu = 0.35 x (Gr.Pr)1/4 for the lower surface. (M.U. 1995)
Solution.
Data:
L := 0.5 m W := 0.5 m Ts := 100°C Ta := 20°C g := 9.81 m/s2
We need properties of air at film temperature Tf = (100 + 20)/2
Tf := 60°C
Properties of air at 60°C
r := 1.06 kg/m 3 n := 18.97 ´ 10 –6 m2/s k := 0.028 W/(mK) Cp := 1008 J/(kgK) m := n×r
Cp × m 1
i.e. m = 2.011 ´ 10 –5 kg/m.s Pr := i.e Pr := 0.724 b :=
k Tf + 273
i.e. b = 3.003 ´ 10 –3 1/K
Case 1: Plate held vertical:
Now, the characteristic length is the vertical side, L
L3 × g × b ×(Ts - Ta )
Gr :=
n2
i.e. Gr = 8.816 ´ 10 8 (Grashoff number)
and, Ra := Gr×Pr
i.e. Ra = 5.926 ´ 10 8 (Rayleigh number)
Therefore,
1
Nu := 0.13×(Gr× Pr) 3
i.e. Nu = 109.194 (Nusselt number)
k × Nu
and, h :=
L
i.e. h = 6.115 W/(m2 K) (heat transfer coefficient)
Therefore, heat transferred:
Q := h×(2×L ×W)×(Ts – Ta) W (heat transfer from both surfaces)
i.e. Q = 244.594 W (heat transfer from both surfaces)
Case 2: Plate held horizontal:
Now, Characteristic length Lc = surface area of plate/Perimeter
L ×W
Lc :=
2 × (L + W )
i.e. Lc = 0.125 m
L3c × g × b × (Ts - Ta )
Then, Gr :=
n2
i.e. Gr = 1.279 ´ 10 7 (Grashoff number)
Therefore, Ra := Gr×Pr (Rayleigh number)
i.e. Ra = 9.259 ´ 10 6
For upper surface:
1
Nuupper := 0.71 ×Ra 4

i.e. Nu upper = 39.166 (Nusselt number)


Nuupper × k
and, hupper : = (heat transfer coefficient)
Lc
2
hupper = 8.773 W/(m K) (heat transfer coefficient for upper surface)
Qupper := hupper ×(L×W)×(Ts – Ta)
i.e. Qupper = 175.462 W (heat transfer from upper surface)

492 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For lower surface:
1
Nu lower := 0.35 ×Ra 4
i.e. Nu lower = 19.307 (Nusselt number)
Nulower × k
and, h lower :=
Lc
i.e. h lower = 4.325 W/(m2 K) (heat transfer coefficient for lower surface)
and, Q lower := hlower (L×W )×(Ts – Ta)
i.e. Qlower = 86.495 W (heat transfer from lower surface)
Total heat transferred:
Q tot := Q upper + Q lower
i.e. Qtot = 261.957 W (Total heat transferred.)

10.5.5 Horizontal Plate With Constant Heat Flux


Here, the characteristic length to be used in expressions for Nu and Gr is:
Lc = A/P
where, A is the surface area and P is the perimeter.
For a circle, Lc = 0.9D, and for rectangle, Lc = (L + W)/2.
All property values, except b , are evaluated at a temperature, Te, defined by:
Te = Ts – 0.25 (Ts – Ta) ...(10.37)
and, b is evaluated at Ta.
Ts is estimated from the basic relation:
havg (Ts – Ta) = qs ...(10.38)
(a) Upper surface of a hot plate (or lower surface of a cold plate):
1
Nu = 0.13×Ra 3 (Ra < 2 ´ 108...(10.39))
and,
1
Nu = 0.16×Ra 3 (2 ´ 10 8 < Ra < 1011...(10.40))
(b) For heated surface facing downward:
Nu = 0.58×Ra0.2 (106 < Ra < 1011...(10.41))
As in the case of vertical plates with constant heat flux, in this case also, iteration will be required while
solving problems.
Example 10.7. A horizontal metal plate, 0.5 m ´ 0.5 m, is exposed to sun and receives radiant energy at the rate 180
W/m2. If the heat transfer from the plate occurs to the surrounding air at 20°C by free convection only, find the steady
state temperature of the plate. Assume that the bottom of the plate is insulated.
Solution. The plate is subjected to constant heat flux. We do not know the surface temperature. So, we can assume either
the surface temperature or the heat transfer coefficient to start with, proceed with the calculations, repeat if necessary.
Data:
L := 0.5 m W := 0.5 m Qs := 180 W/m2 Ta := 20°C g := 9.81 m/s2
Let us assume the surface temperature to be 60°C. Ts := 60°C
This is consant flux condition. So, we need properties of air at Te:
Te := Ts – 0.25× (Ts – Ta)
i.e. Te = 50°C
Properties of air at 50°C:
1
n := 17.95 ´ 10 –6 m2/s k := 0.02826 W/(mK) Pr := 0.698 b := i.e. b = 3.413 ´ 10 –3 1/K
Ta + 273
Now, for horizontal plate, Characteristic length Lc = surface area of plate/Perimeter
L ×W
Lc :=
2 × (L + W )
i.e. Lc = 0.125 m

NATURAL (OR FREE) CONVECTION 493


L3c × g × b × (Ts - Ta )
Then, Gr :=
n2
i.e. Gr = 8.118 ´ 10 6 (Grashoff number)
Therefore, Ra := Gr×Pr (Rayleigh number)
i.e. Ra = 5.667 ´ 10 6
Therefore, for upper surface of horizontal plate losing heat, we have:
1
Nu := 0.13 ×Ra 3 (Ra < 2 ´ 10 8...(10.39))
i.e. Nu = 23.177 (Nusselt number)
k
and, h := Nu × (heat transfer coefficient)
Lc
i.e. h = 5.24 W/(m2 K) (heat transfer coefficient)
Therefore, equating the heat received by plate to the heat transfer from the plate by convection:
q s (L×W) = h×(L×W)×(Ts – Ta)
qs × (L ×W )
Therefore, Ts := + Ta
h × (L ×W )
i.e. Ts = 54.353°C
We had assumed Ts to be 60°C. So, let us repeat the calculations with Ts = 56°C
Ts := 56°C
We need properties of air at film temperature Te
Te := Ts – 0.25 ×(Ts – Ta)
i.e. Te = 47°C
Properties of air at 47°C
n := 17.7 ´ 10 –6 m 2/s (kinematic viscosity)
k := 0.0275 W/(mK) (thermal conductivity)
Pr := 0.71 (Prandtl number)
1
b :=
Ta + 273
i.e. b = 3.413 ´ 10 –3 1/K
L3c × g × b × (Ts - Ta )
Then, Gr :=
n2
i.e. Gr = 7.514 ´ 106 (Grashoff number)
Therefore, Ra := Gr×Pr (Rayleigh number)
i.e. Ra = 5.335 ´ 10 6
Therefore, for upper surface of horizontal plate losing heat, we have:
Nu := 0.13 ×Ra (Ra < 2 ´ 10 8...(10.39))
i.e. Nu = 22.716 (Nusselt number)
k
and, h := Nu × (heat transfer coefficient)
Lc
i.e. h = 4.997 W/(m2 K) (heat transfer coefficient)
Therefore, equating the heat received by plate to the heat transfer from the plate by convection:
Qs ×(L×W) = h×(L×W)×(Ts – Ta)
qs × (L ×W )
Therefore, Ts := + Ta
h × (L ×W )
i.e. Ts = 56.018°C
We had assumed Ts to be 56°C, whereas now, we got Ts = 56.018°C. This is in very good agreement.
Therefore, Ts = 56°C (steady state surface temperature of plate.)

10.5.6 Horizontal Cylinder At Constant Temperature


Here, D, diameter of the cylinder is the characteristic dimension.

494 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For heat transfer from (or to) a horizontal cylinder, Morgan recommends following correlation for fluids
with (0.69 < Pr < 7):
Nu = C×Ran ...(10.42)
where C and n are obtained from the following Table:

TABLE 10.2 Constants for use in Eq. 10.42

Ra C n
10–10–10 –2 0.675 0.058
10 –2–102 1.02 0.148
102–104 0.85 0.188
104–107 0.48 0.25
107–1012 0.125 0.333

Also, the following correlation of Churchill and Chu may be used for the complete range of Prandtl num-
bers: (0 £ Pr £ ¥) and for a wider range of Rayleigh numbers:

LM LM OP
1 OP 2

MM PP
6

MM PP
MM M PP PP
Nu = M0 .60 + 0 . 387 × M
Ra
MM MM L PP PP (10 - 5 < Ra < 10 12...(10.43))
OP
16

MM MM MM1 + FGH 0.Pr559 IJK


9 9

PP PP
PP
16

MN N MN Q Q PP
Q
And, only for the laminar range:
1
0. 518 × Ra 4
Nu = 0.36 + (10 -6 < Ra < 10 9...(10.44))
LM F 0.559 I OP
4
9 9

MM1 + GH Pr JK PP
16

N Q
Properties in the above equations are evaluated at film temperature, D is the characteristic dimension.
Churchill and Chu recommend that above two eqns. may be used for constant flux conditions too, with the
temperature Ts being half way up the cylinder at the 90 deg. angle from bottom.
For thin wires: (D = 0.2 mm to 1 mm): Rayleigh number is usually very small and a film type of flow
pattern is observed. Following correlation is used:
1
NuD = 1.18× ( RaD ) 8 (Ra < 500...(10.45))
Heat transfer from horizontal cylinders to liquid metals may be calculated from:
1
NuD = 0.53×(GrD ×Pr 2) 4 ...(10.46)
Example 10.8. A horizontal, steam pipe of 10 cm OD runs through a room where the ambient air is at 20°C. If the
outside surface of the pipe is at 180°C, and the emissivity of the surface is 0.9, find out the total heat loss per metre
length of pipe.
Solution. The pipe is horizontal and loses heat by natural convection as well as radiation. Diameter D is the characteris-
tic dimension to calculate Rayleigh number.
Data:
D := 0.1 m L := 1.0 m Ts := 180°C Ta := 20°C g := 9.81 m/s 2 e := 0.9 s := 5.67 ´ 10 – 8 W/(m2 K4)

NATURAL (OR FREE) CONVECTION 495


We need properties of air at film temperature Tf = (180 + 20)/2
Tf := 100°C (film temperature)
Properties of air at 100°C:
r := 0.946 kg/m 3 n := 23.02 ´ 10 – 6 m2/s k := 0.03127 W/(mK) Cp := 1011.3 J/(kgK) Pr := 0.704
1
b := i.e. b = 2.68 ´ 10 –3 1/K
Tf + 273

g × b × (Ts - Ta ) × D 3
Then, Gr :=
n2
i.e. Gr = 7.94 ´ 10 6 (Grashoff number)
Therefore, Ra := Gr×Pr (Rayleigh number)
i.e. Ra = 5.59 ´ 10 6
To find Nusselt number, we use Eq. 10.43:

LM LM OP
1
6
OP 2

MM MM PP PP
MM PP
Nu := 0 .60 + 0 . 387 × M PP
Ra
MM MM L OP
16
PP PP (10 –5 < Ra < 1012...(10.43))

MM MM1 + FGH 0.559 IJK


9 9

MM 16

PP PP PP
MN MN MN Pr Q Q PQ
i.e. Nu = 23.788 (Nusselt number)
k
And, h := Nu × (heat transfer coefficient)
D
2
i.e. h = 7.438 W/(m K) (heat transfer coefficient)
Therefore, heat loss by natural convection:
Qconv := h×(p×D×L)×(Ts – Ta) W/m
i.e. Qconv := 373.897 W/m
And, heat loss by radiation:
Remember that, here, the temperatures must be in Kelvin.
Qrad := e×(p×D×L)×s×[(Ts + 273)4 – (Ta + 273)4 ] W/m
i.e. Qrad = 556.947 W/m
Therefore, total heat loss from pipe surface:
Qtot := Qconv + Qrad W/m
i.e. Qtot = 930.844 W/m
Note that in this type of problems, radiation heat loss is quite comparable to the natural convection heat loss and
must, therefore, always be considered.
Alternatvely:
We can also use Eq. 10.42 to find out Nu, to determine the convection heat loss:
Nu = C×Ra n ...(10.42)
Where constants C and n for Ra = 5.59 ´ 10 6 are obtained from Table 10.2 as:
C := 0.48 n := 0.25
Therefore, Nu := C×Ran
i.e. Nu = 23.34 (Nusselt number...compare with Nu = 23.788 obtained earlier)
k
And, h := Nu × (heat transfer coefficient)
D
2
i.e. h = 7.298 W/(m K) (heat transfer coefficient...compare with h = 7.438 obtained earlier)
Therefore, heat loss by natural convection:
Qconv := h×(p×D×L)×(Ts – Ta) W/m
i.e. Qconv = 366.859 W/m (compares with 373.897 W/m, got earlier.)

496 FUNDAMENTALS OF HEAT AND MASS TRANSFER


And, total heat loss from pipe surface:
Qtot := Qconv + Qrad W/m
i.e. Qtot = 923.806 W/m.
Example 10.9. A tank contains water at 15°C. The water is heated by passing steam through a pipe placed in water. The
pipe is 60 cm long and 4 cm in diameter and its surface is maintained at 85°C. Find the heat loss from the pipe if:
(i) the pipe is kept horizontal
(ii) the pipe is kept vertical.
Following empirical relations may be used: Nu = C.(Gr.Pr)m, where
C = 0.53 and m = 0.25 when 104 < Gr.Pr < 109, and
C = 0.13 and m = 1/3 when Gr.Pr > 109. (M.U., 1998)
Following data may be used:
Properties of water at average temperature of 50°C are:
Tf := 50°C r := 988 kg/m3 n := 5.56 ´ 10 –7 m2/s Cp := 4178 J/(kgK)
k := 0.647 W/(mK) b := 5.1 ´ 10 –4 1/K
Other data:
L := 0.6 m D := 0.04 m Ts := 85°C Ta := 15°C g := 9.81 m/s2 m := n×r i.e. m = 5.493 ´ 10 –4 kg/(ms)
Cp × m
Pr := i.e. Pr := 3.547
k
Case 1: Pipe held horizontal:
Diameter D is the characteristic dimension to calculate Ra.
A := p×D×L m2 (surface area)
i.e. A = 0.075 m2 (surface area)
D 3 × g × b × (Ts -Ta )
And, Gr :=
n2
Gr = 7.25 ´ 107 (Grashoff number)
And, Ra := Gr×Pr
i.e. Ra = 2.572 ´ 10 8 (Rayleigh number)
Then, we have:
Nu := 0.53 ×Ra 0.25
i.e. Nu = 67.118 (Nusselt number)
Nu × k
and, h :=
D
i.e. h = 1.086 ´ 10 3 W/(m 2K) (heat transfer coefficient)
And heat transfer is given by:
Q horizl := h×A×(Ts – Ta)
i.e. Qhorizl = 5.73 ´ 10 3 W
Case 2: Pipe held vertical:
Now, the length L is the characteristic dimension to calculate Ra.
L3 × g × b ×(Ts - Ta )
We have: Gr :=
n2
i.e. Gr = 2.447 ´ 1011
and, Ra := Gr×Pr
i.e. Ra = 8.68 ´ 10 11
And, Nu := 0.13 ×Ra
i.e. Nu = 1.24 ´ 10 3 (Nusselt number)
Nu × k
and, h :=
L
i.e. h = 1.337 ´ 10 3 W/(m 2K) (heat transfer coefficient.)
And heat transfer is given by:
Qvert := h×A×(Ts – Ta)
i.e. Qvert = 7.058 ´ 10 3 W

NATURAL (OR FREE) CONVECTION 497


Example 10.10. A fine wire of 0.2 mm diameter is maintained at a constant temperature of 64°C by an electric current.
The wire is exposed to air at 1 bar and 10°C. Calculate the electric power necessary to maintain the wire temperature if
the length of wire is 1 m.
Solution.
Data:
L := 1.0 m D := 2 ´ 10 –4 m Ts := 64°C Ta := 10°C g := 9.81 m/s2
Properties of air at film temperature of (64 + 10)/2 = 37°C are:
Tf := 37°C r := 1.143 kg/m3 n := 16.7 ´ 10 –6 m2/s Cp := 1006 J/(kgK) k := 0.0268 W/(mK)
1
Pr := 0.711 b := 1/K i.e. b = 3.226 ´ 10– 3 1/K
Tf + 273
Diameter D is the characteristic dimension to calculate Ra.
A := p×D×L m2 (surface area)
i.e. A = 6.283 ´ 10 – 4 m2 (surface area)
D 3 × g × b × (Ts - Ta )
And, Gr :=
n2
Gr = 0.049018 (Grashoff number)
And, Ra := Gr×Pr
i.e. Ra = 0.035 (Rayleigh number)
Then, applying Eq. 10.45, we get:
1
Nu D := 1.18 × ( Ra) 8 (Ra < 500...(10.45))
i.e. NuD = 0.776 (Nusselt number)
k
and, h := NuD × (heat transfer coefficient)
D
i.e. h = 103.936 W/(m2 K) (heat transfer coefficient)
Therefore, rate of heat loss from wire;
Q := h×A×(Ts – Ta) W/m
i.e. Q = 3.526 W/m
i.e. Power required to maintain the surface temperature at 64°C = 3.526 W
Alternatively:
We can use Eq. 10.42:
Nu = C×Ran ...(10.42)
Where constants C and n are obtained from the Table, corresponding to Ra = 0.035:
C := 1.02 and n = 0.148
Then,
Nu := C×Ran
i.e. Nu = 0.621 (Nusselt number)
k
and, h := Nu × (heat transfer coefficient)
D
i.e. h = 83.168 W/(m2 K) (compare this with h = 103.936 obtained earlier.)
Therefore, rate of heat loss from wire:
Q := h×A×(Ts – Ta) W/m
i.e. Q = 2.822 W/m (compare this with Q = 3.526 W/m obtained earlier.)

10.5.7 Free Convection From Spheres


Sphere diameter D is the characteristic dimension. Yuge recommends following correlation for average Nusselt
number for free convection between a sphere and air. Properties are evaluated at the film temperature.
1
Nu = 2 + 0.43×(Ra) 4 (1 < Ra < 105, Pr = 1...(10.47))
For higher range of Ra:
1
Nu = 2 + 0.50 × (Ra) 4 (3 ´ 105 < Ra < 8 ´ 108...(10.48))

498 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Example 10.11. A sphere of 25 mm diameter, with its surface temperature at 100°C, is kept in still air at a temperature of
20°C. Determine the rate of convective heat loss.
Solution.
Data:
D := 25 ´ 10– 3 m Ts := 100°C Ta := 20°C g := 9.81 m/s2
Properties of air at film temperature of (100 + 20)/2 = 60°C are:
1
Tf := 60°C n := 18.97 ´ 10 –6 m2/s k := 0.02896 W/(mK) Pr := 0.696 b := 1/K
Tf + 273
i.e. b = 3.003 ´ 10 –3 1/K
Diameter D is the characteristic dimension to calculate Ra.
A := p×D 2 m2 (surface area of sphere)
i.e. A = 1.96 ´ 10 –3 m 2 (surface area)
D 3 × g × b × (Ts - Ta )
And, Gr :=
n2
Gr = 1.023 ´ 10 5 (Grashoff number)
And, Ra := Gr×Pr
i.e. Ra = 7.122 ´ 10 4 (Rayleigh number)
Then, using Eq. 10.47:
Nu := 2 + 0.43×(Ra)1/4 (1 < Re < 105, Pr = 1...(10.47))
i.e. Nu = 9.025 (Nusselt number)
k
and, h := Nu × (heat transfer coefficient)
D
i.e. h = 10.454 W/(m2K) (heat transfer coefficient)
Therefore, rate of heat loss from sphere:
Q := h×A×(Ts – Ta) W
i.e. Q = 1.642 W

10.5.8 Free Convection From Rectangular Blocks and Short Cylinders


Here, characteristic length L is defined as:
LH × LV
L= ...(10.49a)
LH + LV
where LH is the longer of the two horizontal dimensions and LV is the vertical dimension. Based on this character-
istic length, the heat transfer correlation is:
1
NuL = 0.55×(RaL) 4 (10 4 < Ra L < 109...(10.49))
For short cylinders (D = H):
Nu = 0.775×(Ra) 0.208 ...(10.50)
Example 10.12. A ceramic block is of 0.3 m ´ 0.2 m section and is 0.3 m in height. Surface temperature of the block is
380°C. If it is exposed to air at 20°C, determine the rate of convective heat loss.
Solution.
Data:
LH := 0.3 m LV := 0.3 m Ts := 380°C Ta := 20°C g := 9.81 m/s2
Properties of air at film temperature of (380 + 20)/2 = 200°C are:
1
Tf := 200°C n := 34.57 ´ 10 –6 m2/s k := 0.03781 W/(mK) Pr := 0.699 b := 1/K
Tf + 273
i.e. b = 2.114 ´ 10 –3 1/K
The characteristic dimension to calculate Ra, is:
LH × LV
L :=
LH + LV
i.e. L = 0.15 m
Also, A := 2×(0.3×0.3) + 2 ×(0.2×0.3) + 2 ×(0.3×0.2) m2 (surface area of the block)

NATURAL (OR FREE) CONVECTION 499


i.e. A = 0.42 m2 (surface area)
L × g × b ×(Ts - Ta )
3
And, Gr :=
n2
Gr = 2.109 ´ 10 7 (Grashoff number)
and, Ra := Gr×Pr
i.e. Ra = 1.474 ´ 10 7 (Rayleigh number)
Then using Eq. 10.49 we get:
1
Nu := 0.55 × ( Ra) 4 (104 < Ra < 109...(10.49)
i.e. Nu = 34.078 (Nusselt number)
k
and, h := Nu × (heat transfer coefficient)
L
i.e. h = 8.59 W/(m2 K) (heat transfer coefficient)

Free convection for various shapes


3.2
3
2.8
2.6
2.4
2.2
2
1.8
log(Nu)

1.6
Y
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0.2
0.4
5 4 3 2 1 0 1 2 3 4 5 6 7 8 9 10 11 12
X
log(Ra)

FIGURE Example 10.12 Graph of log(Ra) vs. log(Nu) for various shapes in free convection with various
fluids

Therefore, rate of heat loss from the block:


Q := h×A×(Ts – Ta) W
i.e. Q = 1.299 ´ 10 3 W
Alternatively:
Based on experimental data for vertical plates, vertical cylinders, horizontal cylinders, spheres and blocks to various
fluids such as air, water, alcohol and oil, King has drawn the following curve of log(Ra) vs. log(Nu). Here, fluid proper-
ties are evaluated at the film temperature and the characteristic dimension (L) to be taken to determine Ra and Nu are:
for a vertical plate L is the height, and for long, horizontal cylinder, L is the diameter, and for a short cylinder or
block, 1/L = (1/LV ) + (1/LH). For a sphere, radius is the characteristic dimension.
Above Fig. Example 10.12 is expected to give fair estimate of convection coefficient for objects other than horizontal
cylinder and plate. This figure can also be used for more common shapes when the Rayleigh number is outside the range
of the specific correlation for that shape.
Now, in this problem: Ra = 1.474 ´ 10 7
Therefore, log(Ra) = 7.168
Using this value in the x-axis of above Fig. Example 10.12 we get:
log(Nu) = 1.55
i.e. Nu = 101.55
i.e. Nu := 35.481

500 FUNDAMENTALS OF HEAT AND MASS TRANSFER


k
and, h := Nu × (heat transfer coefficient)
L
i.e. h = 8.944 W/(m2 K) (heat transfer coefficient)
Compare this value of h with h = 8.59 W/(m2 K), got earlier.
Therefore, rate of heat loss from the block:
Q := h×A×(Ts – Ta) W
i.e. Q = 1.352 ´ 10 3 W
Compare this value of Q with Q = 1299 W, obtained earlier.

10.5.9 Simplified Equations For Air


Since air is the common fluid in most of the free convection problems encountered in practice, it is useful to have
simplified relations for those situations:

TABLE 10.3 Simplified equations for free convection to air at atmospheric pressure (constant wall temp.)

Surface Laminar 10 4 < (Gr.Pr) < 109 Turbulent (Gr.Pr ) > 109

FG DT IJ
1
1
4
Vertical plate or cylinder h = 1.42×
HLK h = 1.31×(D T) 3

F DT IJ
1
1
h = 1.32× G
4
Horizontal cylinder
HDK h = 1.24×(D T ) 3

Horizontal plate:
FG DT IJ
1
1
4
Heated plate facing upward, or
cooled plate facing downward
h = 1.32×
HLK h = 1.52×(D T ) 3

FG DT IJ
1
Heated plate facing downward, 4

or cooled plate facing upward


h = 0.59×
HLK
where h = heat transfer coefficient, L = vertical or horizontal dimension,
D = diameter, and D T = Ts – Ta
k
Spheres h = [2 + 0.392 × Gr D1/4] × for 1 < GrD < 105
D

For pressures other than atmospheric, multiply the RHS of above expressions as below, where p is in bar:

F p I
1

GH 1.0132 JK
2
Laminar:

F p I
2

GH 1.0132 JK
3
Turbulent

10.5.10 Free Convection In Enclosed Spaces


Enclosed spaces may be formed by horizontal plates or vertical plates; also, enclosed spaces may be filled with
air or any other fluids. Typical example is a double-plane window, or a vacuum flask or a cryogenic container
involving concentric cylinders or spheres. Correlations for convection heat transfer for such situations are given
below.
Fig. 10.3 shows the horizontal and vertical enclosures and the nomenclature used. Here, the space between
the plates, ‘b’ is the characteristic dimension. Properties are evaluated at the average of two plate temperatures.
So, now, Grashoff number for the enclosure is defined as:

NATURAL (OR FREE) CONVECTION 501


b
T1 T2
T2

b Q L

T1
Q

FIGURE 10.3 (A) Free convection in a horizontal FIGURE 10.3 (B) Free convection in a vertical
Enclosure (T1 > T2) Enclosure (T1 > T2)

g × b ×(T1 - T2 ) × b 3
Grb = ...(10.51)
n2
and, the Rayleigh number for the enclosure is:

g × b ×(T1 - T2 ) × b 3
R ab = ...(10.52)
n ×a
For Horizontal enclosure:
For air:
Average Nusselt number (based on plate spacing ‘b’) is given by Jakob:
1
Nu = 0.195×Gr 4 (104 < Gr < 3.7 ´ 10 5...(10.53))
1
And, Nu = 0.068×Gr 3 (3.7 ´ 105 < Gr < 107...(10.54))
And, for Gr < 1700, we have Nu = 1.
For liquids (water, silicone oils and mercury), equation suggested by Globe and Dropkin:
1
Nu = 0.069×Ra 3 ×Pr0.074 (1.5 ´ 105 < Ra < 109...(10.55))
Here also, the space between the plates, ‘b’ is the characteristic dimension. Properties are evaluated at the
average of two plate temperatures.
For Vertical enclosure:
For Air:
For Gr (based on plate spacing ‘b’) < 1700, we have Nu = 1.
Jakob has given following correlations:
1
keff 0.18 × Gr 4
= Nu = (2 ´ 10 4 < Gr < 2 ´ 105...(10.56))
FG IJ
1
k
L 9
b H K
where keff = effective thermal conductivity
1
keff 0.065 ×Gr 4
and, = Nu = (2 ´ 105 < Gr < 10 7...(10.57))
FG IJ
k 1
L 9
b H K
Note that for above two relations, aspect ratio, L/b > 3.
If L/b < 3, each vertical surface is treated independently.

502 FUNDAMENTALS OF HEAT AND MASS TRANSFER


If the enclosed vertical layer contains fluids with Prandtl numbers between 3 and 30,000, following correla-
tion due to Emery and Chu may be used:
Nu = 1 (for Ra < 1000...(10.58))
1
0.28 × Ra 4
And, Nu = (for 1000 < Ra < 10 7...(10.59))
FG L IJ
1
4
H bK
Layer thickness ‘b’ is the characteristic dimension used in Nu and Ra.
Example 10.13. Air at 2 bar pressure is contained between two horizontal panels separated by a distance of 20 mm. The
lower panel is at a temperature of 70°C and the upper panel is at 30°C. Calculate the heat transfer rate by free convection
per sq. m. of the panel surface.
Solution.
Data:
L := 0.02 m T1 := 70°C T2 := 30°C g := 9.81 m/s 2 P := 2 ´ 10 5 Pa R := 287 J/kgK
We need properties of air at film temperature Tf = (70 + 30)/2
Tf := 50°C (average temperature)
Properties of air at 50°C:
Note that only density changes with pressure and m, k and Cp do not change much; however, n = m/r, and this
changes with pressure.
p
r := kg/m3 (density of air at 2 bar pressure)
R × (Tf + 273)
i.e. r = 2.157 kg/m3 (density of air at 2 bar pressure)
m := 19.57 ´ 10 –6 kg/ms (dynamic viscosity)
m
Therefore, n :=
r
i.e. n = 9.07 ´ 10 –6 m 2/s (kinematic viscosity at 2 bar pressure)
k := 0.02781 W/(mK) (thermal conductivity)
Pr := 0.709 (Prandtl number)
1
b :=
Tf + 273
i.e. b = 3.096 ´ 10 –3 1/K
g × b × (T1 - T2 ) × L3
Then, Gr :=
n2
i.e. Gr = 1.181 ´ 10 5 (Grashoff number)
Then, using Eq. 10.53, we get:
1
Nu := 0.195× Gr 4 (10 4 < Gr < 3.7 ´ 10 5...(10.53))
i.e. Nu = 3.615 (Nusselt number)
Then, heat flux across the gap is computed from:
q× L
Nu =
k × (T1 - T2 )
Nu × k × (T1 - T2 )
i.e. q := ...heat flux across the gap
L
2
i.e. q = 201.07 W/m ...heat flux across the gap
Example 10.14. Air gap between the two glass panels of a double-pane window (0.8 m wide ´ 1.5 m high) is 2 cms. If the
two glass surfaces are at 20°C and 0°C, determine the rate of heat transfer through the window.
Solution.
Data:
L := 1.5 m W := 0.8 m b := 0.02 m T1 := 20°C T2 := 0°C g := 9.81 m/s2

NATURAL (OR FREE) CONVECTION 503


We need properties of air at film temperature Tf = (20 + 0)/2
Tf := 10°C (average temperature)
Properties of air at 10°C:
1
n := 14.19 ´ 10 –6 m2/s k := 0.02487 W/(mK) Pr := 0.716 b := i.e. b = 3.534 ´ 10 –3 1/K
Tf + 273
Remember that here, the distance between panels ‘b’ is the characteristic dimension.
g × b × (T1 - T2 ) × b 3
Then, Gr :=
n2
i.e. Gr = 2.754 ´ 10 4 (Grashoff number)
L
And, = 75 > 3 (condition is satisfied.)
b
Then, using Eq. 10.56 we get:
1
0 .18 × Gr 4
Nu := (2 ´ 104 < Gr < 2 ´ 105...(10.56))
FG L IJ
1
9

H bK
i.e. Nu = 1.435 (Nusselt number)
Therefore, k eff := Nu×k
i.e. keff = 0.036 W/(mK) (effective thermal conductivity)
Then, heat flux across the gap is computed from:
q×b
Nu :=
k × (T1 - T2 )
Nu × k × (T1 - T2 )
i.e. q= (heat flux across the gap)
b
Defining ‘effective thermal conductivity’, we can also write the above relation as:
keff := Nu×k (effective thermal conductivity)
i.e. keff = 0.036 W/(mK) (effective thermal conductivity)
keff × (T1 - T2 )
and, q := (heat flux across the gap)
b
i.e. q = 35.696 W/m2 (heat flux across the gap)
Therefore, Q := q×(L×W) W (heat transfer rate)
i.e. Q = 42.835 W (heat transfer rate)

10.5.11 Free Convection In Inclined Spaces


This situation is encountered in flat plate solar collectors and double-glazed windows. Fig. 10.4 shows the no-
menclature for the relations given below. This configuration has been investigated for large aspect ra-
tios (L/d < 12) by Hollands et.al. Following equation correlated
d experimental data at tilt angles t less than 70 deg.:
T
OP
2
T1 F I LM 1708 ×(sin (1.8 ×t ))1. 6
GH
NuL = 1 + 1.44× 1 -
1708
RaL × cos(t ) JK MN
× 1-
RaL × cos(t ) PQ
L F L Ra ×cos(t ) O 1 I
+ GM - 1J ...(10.60)
GH N 5830 PQ
Q L 3
JK
t If the quantity in the first bracket and the last bracket is nega-
tive, then it must be set equal to zero.
FIGURE 10.4 Free convection in inclined, For tilt angles between 70 deg. and 90 deg. Catton recom-
enclosed space (T1 > T2) mends that the Nusselt number for a vertical enclosure (t = 90
deg.) be multiplied by (sin t)1/4
504 FUNDAMENTALS OF HEAT AND MASS TRANSFER
i.e.
NuL (t) = Nu L (t = 90)×(sin (t)) ...(10.61)
Example 10.15. In a solar flat plate collector, the plate is of size 1 m ´ 1 m and is at a temperature of 140°C. The glass
cover plate is at a distance of 8 cm from the collector surface and its temperature is 40°C. Space in between contains air
at 1 atm. If the collector plate is inclined to the horizontal at 20 deg., determine the heat transfer coefficient.
Solution.
Data:
L := 1.0 m W := 1.0 m b := 0.08 m T1 := 140°C T2 := 40°C t := 20 deg (angle of tilt (to horizontal)
But, while using Mathad, arguments for trigonometric functions must be in radians: So,
p
t := 20× (radians...angle of tilt (to horizontal))
180
2
g := 9.81 m/s (acceleration due to gravity)
We need properties of air at average temperature Tf = (140 + 40)/2
Tf := 90°C (average temperature)
Properties of air at 90°C:
n := 21.96 ´ 10 – 6 m2/s (kinematic viscosity)
k := 0.03059 W/(mK) (thermal conductivity)
Pr := 0.705 (Prandtl number)
1
b :=
Tf + 273
i.e. b = 2.755 ´ 10 –3 1/K
Remember that here, the height of panels ‘L’ is the characteristic dimension.
L
And, = 12.5 > 12 (condition is satisfied.)
b
g × b × (T1 - T2 ) × L3
Then, GrL :=
n2
i.e. GrL = 5.604 ´ 10 9 (Grashoff number)
and, RaL := GrL ×Pr (Rayleigh number)
i.e. RaL = 3.951 ´ 10 9 (Rayleigh number)
Then, using Eq. (10.60), we get:

L OP
OP + MMFG Ra × cos(t ) IJ
1

F 1708 I × LM1 - 1708 ×(sin(1.8×t ))


:= 1 + 1.44 × G 1 -
1. 6 3

- 1P
H Ra × cos(t ) JK MN PQ MMH 5830 K
L
Nu L
L Ra × cos (t )L PP ...(10.60)

N Q
F 1 - 1708 I = 1
We have: GH Ra × cos(t ) JK
L
(not negative)

LMF Ra ×cos (t ) I OP
1

MMGH 5830 JK - 1PP = 85.034


L 3
and, (not negative)
N Q
Note: If the above two terms are negative, then they must be set equal to zero.
Therefore, Nu L = 87.474 (Nusselt number)
NuL × k
i.e. h := ...W/(m2K) (heat transfer coefficient)
L
i.e. h = 2.676 W/(m2 K) (heat transfer coefficient.)

10.5.12 Natural Convection Inside Spherical Cavities


Diameter D is the characteristic dimension. Following relation is recommended:
D × havg
= C×(GrD ×Pr) n ...(10.62)
k

NATURAL (OR FREE) CONVECTION 505


where C and n are taken from table below:

GrD .Pr C n

104 – 109 0.59 1/4


109 – 1012 0.13 1/3

10.5.13 Natural Convection Inside Concentric To


Cylinders and Spheres
Free convection in enclosures formed between concentric cylinders Ti
and concentric spheres when the gap is filled with various fluids such
as air, water and oils have been correlated by Raithby and Hollands.
See Fig. 10.5. Here, Di and Do are the inside and outside diam-
eters of the long cylinders or spheres; Ti and To are the corresponding
temperatures. L is the length of long cylinders, and ‘b’ is the gap or
thickness of the enclosed fluid layer (i.e. b = [Do – Di]/2). Procedure is b
Di
b

to find out an effective thermal conductivity and then determine the


heat transfer as if by pure conduction, using this effective thermal Do

conductivity.
FIGURE 10.5 Free convection in an
Concentric cylindrical annuli: enclosure between long, concentric
Q 2×p × keff × (Ti - To ) cylinders and spheres (Ti > To)
L
=
ln o
FG IJ
D
...(10.63)

H K
Di

F I
1
1
keff Pr
GH JK
4
= 0.386× × Racc4 (100 < Racc < 107...(10.64))
k 0.861 + Pr
And,

F ln F D I I × Ra 4

GH GH D JK JK o
i
b
Racc = ...(10.65)
LM O 5

1 P
b ×M
MM D + D PPP
3 1
3 3

N i
5
Q 5
o

Concentric spherical annuli:


Di × Do
Q = p×k eff × ×(Ti – To) ...(10.66)
b

F I
1
1
keff Pr
GH JK
4
= 0.74× × Racs4 (102 < Racs < 104...(10.67))
k 0.861 + Pr
And,

b × Rab
Racs = ...(10.68)
LM O 5

1 P
D ×D × M PP
4 1
4
+
o
MM Di 7
D PQ
7

N i
5 5
o

506 FUNDAMENTALS OF HEAT AND MASS TRANSFER


In both Eqs. 10.65 and 10.68, Rayleigh number (Rab) is based on the thickness ‘b’ of the annular fluid layer.
Further, fluid properties are to be evaluated at the average of Ti and To, and Eqs. 10.64 and 10.67 are invalid if
(keff/k) found from them is less than unity. If (keff/k) is less than one, then the process is one of pure conduction
in the fluid and keff = k should be used.
Example 10.16. A sphere of 0.15 m diameter stores a brine at – 5°C and is insulated by enclosing it in another sphere of
0.2 m diameter and the intervening space contains air at 1 bar. The outside sphere is at 25°C. Estimate the convection
heat transfer rate.
Solution.
Data:
Di := 0.15 m Do := 0.2 m b := 0.025 m Ti := – 5°C To := 25°C g := 9.81 m/s2
We need properties of air at average temperature Tf = (25 – 5)/2
Tf := 10°C (average temperature)
Properties of air at 10°C:
1
n := 14.19 ´ 10 –6 m2/s k := 0.02487 W/(mK) Pr := 0.716 b := i.e. b = 3.534 ´ 10 –3 1/K
Tf + 273
Remember that here, the gap between spheres ‘b’ is the characteristic dimension.
g × b × (To - Ti ) × b 3
Then, GRb :=
n2
i.e. Grb = 8.07 ´ 10 4 (Grashoff number)
Therefore, Rab := Grb ×Pr
i.e. Ra b = 5.778 ´ 104 (Rayleigh number)
Now,

F I
1
1
k eff Pr
GH JK
4
= 0.74 × × Ra cs4 (10 < Racs < 106...(10.67))
k 0 .861 + Pr

b × Rab
where Racs := ...(10.68)
LM 1 O
1 P
5

D ×D × M
4 4
+ P
MN D D PQ
o i 7 7
5 5
i 0

i.e. Racs = 235.649


Therefore, from Eq. 10.67 we get:
k eff
= 2.38
k
i.e. keff := 2.38 ×k
i.e. k eff = 0.059 W/(mK) (effective thermal conductivity)
Therefore, rate of heat loss:
FG D × D IJ ×(T – T )
H b K
i o
Q := p×k eff × i o ...(10.66)

i.e. Q = –6.694 W.
Note: Negative sign indicates that heat flow is from outside to inside.
Example 10.17. A long tube of 0.1 m OD is maintained at 150°C. It is surrounded by a cylindrical radiation shield,
located concentrically, such that the air gap between the two cylinders is 10 mm. The shield is at a temperature of 30°C.
Estimate the convection heat transfer rate per metre length.
Solution.
Data:
Di := 0.1 m Do := 0.12 m b := 0.01 m Ti := 150°C To := 30°C g := 9.81 m/s2
We need properties of air at average temperature Tf = (150 + 30)/2
Tf := 90°C (average temperature)
Properties of air at 90°C:

NATURAL (OR FREE) CONVECTION 507


1
n := 21.96 ´ 10 –6 m2/s k := 0.03059 W/(mK) Pr := 0.705 b := i.e. b = 2.755 ´ 10 –3 1/K
Tf + 273
Remember that here, the gap between cylinders ‘b’ is the characteristic dimension.
g × b × (Ti - To ) × b 3
Then, Grb :=
n2
i.e. Grb = 6.725 ´ 10 3 (Grashoff number)
Therefore, Rab := Grb ×Pr
i.e. Ra b = 4.741 ´ 103 (Rayleigh number)
Now,

F I
1
1
k eff Pr
GH JK
4
= 0.386× × Racc4 (100 < Racc < 107...(10.64))
k 0 .861 + Pr
where,

F ln F D I I × Ra 4

GH GH D JK JK o

i
b

Racc := ...(10.65)
LM 1 1 OP 5

b ×M
MN D D PPQ
3
+ 3 3
5 5
i o

i.e. Racc = 213.597


Using this value of Ra cc in Eq. 10.64 we get:
k eff
= 1.209
k
i.e. k eff := 1.209× k
i.e. k eff = 0.037 W/(mK) (effective thermal conductivity)
Therefore, rate of heat loss per meter length:
Q 2× p × keff × (Ti - To )
L
=
Do FG IJ ...(10.63)
ln
Di H K
Q
i.e. = 152.943 W/m.
L

10.5.14 Natural Convection In Turbine Rotors, Rotating Cylinders, Disks and Spheres
Thermal analysis of shafting, flywheels, turbine blades, and other machine elements is of practical importance,
and this involves natural convection heat transfer from a rotating body to surrounding ambient.
Cooling of turbine blades:
Blade is cooled by drilling a blind hole from the root till near the tip of the blade and the coolant circulates
through this hole by centrifugal acceleration rm.w 2 where rm is the mean radius of the blade measured from shaft
centre and w is the angular velocity of the blade.
So, now, in Grashoff number, acceleration due to gravity term is replaced by centrifugal acceleration. There-
fore,

(rm ×w 2 ) × b × D T × L3
GrL =
n2
where L is the length of cooling passage.
In practice, Gr is always > 1012 and we use the following equation to find the heat transfer coeff. in fully
turbulent flow:

508 FUNDAMENTALS OF HEAT AND MASS TRANSFER


LM 1.17
OP 0. 4

h ×L
= 0.0246× M PP
a Pr × Gr L
Nua = ...(10.69)
k MMN 1 + 0.495 × Pr 2
3
PQ
Once average heat transfer coefficient is calculated, if d and L are the diameter and length of the hole respec-
tively, total heat transferred is calculated by applying the Newton’s law of cooling:
Q = ha ×(p×d×L)×(Ts – Ta) ...(10.70)
where Ts is the surface temperature of the hole and Ta is the coolant temperature.
Rotating cylinders:
Here, we define a peripheral-speed Reynolds number:

p × D2 ×w
Rew = ...(10.71)
n
At speeds greater than critical, (Rew > 8000 in air), following correlation is used for average Nusselt number
in natural convection from a rotating, horizontal cylinder, in air:
hc × D
NuD = = 0.11×(0.5×Rew2 + GrD ×Pr)0.35 ...(10.72)
k
Rotating disk:
At rotational Reynolds number w.D 2/n below about 10 6, boundary layer on the disk is laminar.
For laminar regime, average Nu for a disk rotating in air:

F I
1
h ×D w × D2 2
NuD = a
k
= 0.36×
n GH JK (for w. D 2/n < 10 6...(10.73))

For turbulent regime, local value of Nu at a radius r is given approximately by:

hc × r w ×r2 F I 0 .8
Nur =
k
= 0.0195×
n GH JK ...(10.74)

If there is laminar flow between r = 0 and r = rc, and turbulent flow between r = rc and r = r 0, average value
of Nusselt number is given by:

F I ×F r I F w ×r I × LM1 - F r I OP
1
2 2 0. 8 2. 6
h ×r w × ro2
GH n JK M GH r JK P
2
Nur = c o = 0.36× GH JK GH r JK
c + 0.015× o c
(for rc < ro...(10.75))
k n o
N Q o

Rotating sphere:
For Pr > 0.7, in laminar flow regime, (i.e. Rew = w.D2/n < 5 ´ 104), average Nusselt number is given by:
NuD = 0.43×Rew0.5 ×Pr 0.4 (Rew < 5 ´ 10 4...(10.76))
And,
NuD = 0.066×Rew0.67 ×Pr0.4 (5 ´ 104 < Rew < 7 ´ 105...(10.77))
Example 10.18. A turbine blade is cooled by free convection with water as coolant. The cooling passage is 8 mm in
diameter and 8 cm long. The blade velocity at a mean radius of 25 cm is 240 m/s. The hole surface temperature is at
230°C and cooling water temperature is 50°C. Find the average heat transfer coefficient and the rate of heat loss.
Solution.
Data:
D := 0.008 m L := 0.08 m rm := 0.25 m V := 240 m/s Ts := 230°C Ta := 50°C g := 9.81 m/s2
We need properties of water at average temperature Tf = (230 + 50)/2
Tf := 140°C ...average temperature
Properties of water at 140°C:
n := 0.2118 ´ 10 – 6 m2/s k := 0.6845 W/(mK) Pr := 1.23 b := 0.966 ´ 10 –3 1/K
Here, the length of hole ‘L’ is the characteristic dimension.

NATURAL (OR FREE) CONVECTION 509


Then,
( rm ×w 2 ) × b × D T × L3
GrL = (Grashoff number)
n2
In the above, (rm × w 2) is the centrifugal acceleration of water at the mean radius, rm.
Now, V = rm ×w
V
i.e. w=
rm

V2
And, rm ×w 2 = = 2.304 ´ 10 5 m/s2 (centrifugal acceleration of water.)
rm
Then, we get
( 2 . 304 ´ 10 5 ) × b × (Ts - Ta ) × L3
GrL := (Grashoff number)
n2
i.e. GrL = 4.572 ´ 1014 (Grashoff number)
Heat transfer coefficient
We use:

h ×L
LM Pr × Gr
1 . 17
OP 0.4

Nua = a
k
= 0.0246× MM L
2 PP ...(10.69)
N 1 + 0.495 × Pr 3
Q
i.e. Nu a := 1.655 ´ 10 4
Nua × k
i.e. ha := W/(m2K) (heat transfer coefficient)
L
i.e. ha = 1.416 ´ 10 5 W/(m2K) (heat transfer coefficient)
Heat transfer:
Q := ha × (p×D×L)×(Ts – Ta) ...(10.70)
i.e. Q = 5.125 ´ 10 4 W
Example 10.19. A 15 cm diameter steel shaft whose surface is at 120°C is allowed to cool while rotating about its own
horizontal axis at 3 r.p.m. in an environment of air at 20°C. Find the initial rate of heat loss.
Solution.
Data:
D := 0.15 m L := 1 m N := 3 r.p.m Ts := 120°C Ta := 20°C g := 9.81 m/s2
We need properties of air of average temperature Tf = (120 + 20)/2
Tf := 70°C (average temperature)
Properties of air at 70°C:
1
n := 19.9 ´ 10 –6 m2/s k := 0.02922 W/(mK) Pr := 0.707 b := 1/K i.e. b = 2.915 ´ 10 –3 1/K
Tf + 273
Now, rotation speed of the shaft is:
2 ×p × N
w := rad/s
60
i.e. w = 0.314 rad/s
Here, we have the peripheral—speed Reynolds number
p × D 2 ×w
Rew := ...(10.71)
n
i.e. Rew = 1.116 ´ 10 3
And,
g × b × (Ts - Ta ) × D 3
GrD := (Grashoff number)
n2
i.e. GrD = 2.437 ´ 10 7 (Grashoff number)
and, Ra := GrD ×Pr (Rayleigh number)

510 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Ra = 1.723 ´ 10 7 (Rayleigh number)
Then, from Eq. 10.72:
hc × D
NuD = = 0.11×(0.5×Rew2 + GrD ×Pr)0.35 ...(10.72)
k
i.e. Nu D := 37.796 (Nusselt number)
NuD × k
And, hc := W/(m2 K) (heat transfer coefficient)
D
i.e. hc = 7.363 W/(m2 K) (heat transfer coefficient)
Initial rate of heat loss:
Q := hc ×(p×D×L)×(Ts – Ta) W/m
i.e. Q = 346.957 W/m.
Example 10.20. A 20 cm diameter disk, being ground at 3000 r.p.m. has its surface at 70°C. Surrounding air is at 30°C.
Find the value of convection coefficient
Solution.
Data:
D := 0.2 m N := 3000 r.p.m Ts := 70°C Ta := 30°C g := 9.81 m/s 2
We need properties of air at average temperature Tf = (70 + 30)/2
Tf := 50°C (average temperature)
Properties of air at 50°C:
1
n := 17.92 ´ 10 –6 m2/s k := 0.02781 W/(mK) Pr := 0.709 b := 1/K i.e. b = 3.096 ´ 10 –3 1/K
Tf + 273
Now, rotational speed:
2 ×p × N
w :=
60
i.e. w = 314.159 rad/s
w ×D 2
Therefore, = 7.012 ´ 10 5 (this is less than 10 6...therefore, laminar.)
n
Then, applying Eq. 10.73 we get:

F I
1
h ×D w × D2
GH JK
2
NuD = a = 0.36× (for w.D 2/n < 106...(10.73))
k n
i.e. Nu D := 301.466 (Nusselt number)
NuD × k
And, ha := W/(m2 K) (heat transfer coefficient)
D
2
i.e. ha = 41.919 W/(m K) (heat transfer coefficient)
Example 10.21. A sphere, 0.1 m in diameter is rotating at 30 r.p.m. in a large container of Carbon dioxide at atmospheric
pressure. The sphere is at 180°C and the CO 2 is at 20°C. Estimate the rate of heat transfer.
Solution.
Data:
D := 0.1 m N := 30 r.p.m Ts := 180°C Ta := 20°C g := 9.81 m/s2
We need properties of CO 2 at average temperature Tf = (180 + 20)/2
Tf := 100°C (average temperature)
Properties of CO2 at 100°C:
1
n := 12.6 ´ 10 –6 m2/s k := 0.02279 W/(mK) Pr := 0.733 b := 1/k i.e. b = 2.681 ´ 10 –3 1/K
Tf + 273
Now, rotational speed:
2 ×p × N
w :=
60
i.e. w = 3.142 rad/s
w ×D 2
And, Rew :=
n
NATURAL (OR FREE) CONVECTION 511
i.e. Rew = 2.493 ´ 10 3 (this is less than 5 ´ 10 4...therefore, laminar.)
Then, average Nusselt number is given by:
Nu D := 0.43 ×Rew0.5 ×Pr 0.4 (Rew < 5 ´ 10 4...(10.76))
i.e. NuD = 18.963 (Nusselt number)
NuD × k
Then, h := W/(m2 K) (heat transfer coefficient)
D
i.e. h = 4.322 W/(m2 K) .(heat transfer coefficient)
Heat transfer rate:
Q := h×(p×D 2)×(Ts – Ta) W
i.e. Q = 21.723 W.

10.5.15 Natural Convection from Finned Surfaces


‘Heat sinks’ used in cooling of electronic devices have fins on their surfaces. Heat is transferred to the heat sink
from the electronic device by conduction and then the heat is dissipated to the ambient from the fins, mostly by
natural convection. Advantage of natural convection cooling is that there is no need to have an external moving
part (like a fan or pump) and therefore, there is increased reliability. Of course, there is a limitation to the amount
of heat that can be transferred, and if the heat to be dissipated is quite large, forced convection cooling may have
to be resorted to.
Fins increase the surface area for heat transfer. If the fins are very close to each other, we will have more
area, but the heat transfer coefficient will be low since too close a spacing of the fins impedes the flow of fluid by
convection. Instead, if the fins are far apart, total surface area will be less, but heat transfer coefficient will be
larger. Therefore, there is an optimum spacing for the fins, which maximizes the heat transfer by natural convec-
tion from a given base area of width W and height L.
Rectangular fins on a vertical surface:
See Fig. 10.6.
W H For a vertical heat sink with isothermal fins of thick-
ness ‘t’ much smaller than the fin spacing ‘S’, Bar–Cohen
Quiescent air, Ta and Rohsenow give the optimum fin spacing as:
L
L Sopt = 2.714× 1
...(10.78)
Ra 4
Ts where L is the fin length in vertical direction and it is the
characteristic dimension to calculate Ra.
Then, heat transfer coefficient for this case of optimum
spacing is given by:
k
t h = 1.31× ...(10.79)
S Sopt
and the rate of heat transfer by natural convection from the
fins is determined from;
FIGURE 10.6 Free convection from vertical heat
sink with fins Q = h×(2×n×L×H)×(Ts – Ta) ...(10.80)

where n = W/(S + t) = number of fins and Ts is the surface temperature of fins.


Rectangular fins on a horizontal surface:
See Fig. 10.7.
For rectangular fins on horizontal surfaces, fins facing upwards for Ts > Ta (or facing downward for Ts < Ta),
Jones and Smith give following correlation:
-1
LF 1500 I 2 OP
Nu = MG
2

MNH Ra JK
s
s
+ (0.081× Ras0 . 39 )- 2
PQ ...(10.81)

Above equation is valid over the range:

512 FUNDAMENTALS OF HEAT AND MASS TRANSFER


200 < Ras < 6 ´ 10 5, Pr = 0.71, 0.026 < H/W < 0.19, and S
0.016 < S/W < 0.20, with the following definitions: Quiescent air, Ta
3 t
q ×S g × b ×(Ts - Ta ) × S
Nus = and, Ras =
(Ts - Ta ) × k n ×a W
Example 10.22. A vertical heat sink, 0.3 m wide ´ 0.15 m high,
is provided with vertical, rectangular fins of 1 mm thickness.
Base and surface temperature of fins is 100°C and the sur-
rounding air is at 20°C. Determine the optimum fin spacing H
and the rate of heat transfer from the heat sink by natural con- Ts
vection.
Solution. FIGURE 10.7 Rectangular fins on a horizontal
Data: surface
W := 0.3 m L := 0.15 m H := 0.02 m
t := 0.001 m Ts := 100°C Ta := 20°C g := 9.81 m/s2
We need properties of air at average temperature T f = (100 + 20)/2
T f := 60°C (average temperature)

W = 0.3 m H = 20 mm

Quiescent air, Ta = 20°C

L = 0.15 m

Ts = 100°C

t = 1 mm

FIGURE Example 10.22 Free convection from vertical heat sink with fins

Properties of air at 60°C:


1
n := 18.97 ´ 10 – 6 m2/s k := 0.02896 W/(mK) Pr := 0.696 b := 1/K i.e. b = 3.003 ´ 10 –3 1/K
Tf + 273
Now, the characteristic length is the length of fins in vertical direction, i.e. L = 0.15 m.
Then,
g × b × (Ts - Ta ) × L3
GrL := (Grashoff number)
v2
i.e. GrL = 2.21 ´ 10 7 (Grashoff number)
And, Ra := GrL ×Pr
i.e. Ra = 1.538 ´ 10 7 (Rayleigh number)
Optimum fin spacing:
We use Eq. 10.78:
L
Sopt := 2.714 ...(10.78)
Ra
1
4

i.e. Sopt = 6.5 ´ 10 –3 m


i.e. Sopt = 6.5 mm

NATURAL (OR FREE) CONVECTION 513


No. of fins:
W
n := (no. of fins)
Sopt + t
i.e. n = 39.998 (say 40)
i.e. n := 40 (no. of fins)
Heat transfer coefficient:
From Eq. 10.79:
k
h := 1.31 × ...(10.79)
Sopt
i.e. h = 5.836 W/(m2K) (heat transfer coefficient)
Heat transfer rate:
From Eq. 10.80:
Q := h × (2 × n × L × H) (Ts – Ta) ...(10.80)
i.e. Q = 112.056 W.

10.6 Comprehensive Correlations from Russian Literature


Following correlations are from the text book by M. Mikheyev.
Free convection from different objects:
Free convection from different objects were investigated with various fluids such as air, hydrogen, carbon diox-
ide, water, aniline, glycerine, carbon tetrachloride, various oils etc. Objects studied included horizontal and ver-
tical wires, tubes, plates, and spheres of widely different sizes: wires and tubes from 0.015 to 245 mm in diameter,
spheres from 30 mm to 16 m in diameter, height of plates and tubes ranging from 0.25 to 6 m. Gas pressures were
varied from 0.03 to 70 ata.
While generalizing the data, reference dimension was diameter d for tubes and spheres, and height h for
plates. Properties of fluids were taken at film temperature, Tf = (Ts + Ta)/2. It is interesting to note that all the
data, when plotted with log(Gr.Pr) on the x-axis and log (Nu) on the y-axis, fall fairly well on one common curve.
So, the general relation is:
Nu f = C (Gr × Pr) n ...(10.82)
Values of C and n for different ranges of (Gr ×Pr) are taken from following Table:

(Gr × Pr ) f C n
1 x 10–3 – 5 x 102 1.18 1/8
5 x 102 – 2 x 107 0.54 1/4
2 x 107 – 1 x 1013 0.135 1/3

Note that with Ra = (Gr × Pr) < 1, Nu = 0.5 and remains constant, i.e. h = 0.5.k/d = heat transfer coefficient for
very low Rayleigh numbers. (e.g. for very thin wires).
Principal conclusions were: (a) Rayleigh number is the main dimensionless term to determine heat transfer
in fee convection (b) Shape of the body is of secondary importance in the process considered.
Eq. 10.82 is applicable to any fluid with Pr > 0.7 and for bodies of any shape and size. Same formula may be
used to calculate heat transfer from horizontal plates too. Then reference dimension is the smaller side of the
plate. Value of h determined from Eq. 10.82 must be increased by 30 % if the hot surface is facing upwards, and
decreased by 30 % if the heat losing surface faces downward.
For horizontal tubes, especially, following correlation is recommended for free convection with liquids and
gases:

F Pr I
×G
1
4

H Pr JK
1
a
Nua = 0.51× (Gr × Pr ) 4 ...(10.83)
w
Here, note that fluid properties are determined at free stream temperature Ta and the reference dimension is
the tube diameter, d.

514 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For air, Eq. 10.83 is given in the following simplified form:
1
Nua = 0.47× Gra4 ...(10.84)
Note that for horizontal tubes, Eq. 10.83 is to be preferred to Eq. 10.82.
Free convection in different enclosures:
Here, the concept of ‘equivalent thermal conductivity’ is used. It has the advantage that the heat transfer coeffi-
cient h need not be determined. In the following correlations, thickness of the enclosure d and the mean fluid
temperature (= [T1 + T 2]/2), are taken as the reference dimension and reference temperature respectively, irre-
spective of the shape of the enclosure. Passages considered are: horizontal passages, vertical passages, enclosures
within concentric cylinders and concentric spheres. Again, it is found that for all these enclosures, data fall well
within a single curve. Following are the correlations:
keff
=1 (for Ra < 1000...(10.85))
k
keff
= 0.105 × Ra0.3 (for 103 < Ra < 106...(10.86))
k
and,
keff
= 0.4 × Ra 0.2 (for 10 6 < Ra < 1010...(10.87))
k
In approximate calculations, Eqs. 10.86 and 10.87 may be replaced by the following single eqn. for the entire
range of Ra > 1000:
keff
= 0.18 × Ra0.25 (for Ra > 1000...(10.88))
k
If keff/k works out to be less than one, it means that Ra < 1000, and we should take keff = k.
Example 10.23. Work out Example 10.11 with formula from Russian literature: A sphere of 25 mm diameter, with its
surface temperature at 100°C, is kept in still air at a temperature of 20°C. Determine the rate of convective heat loss.
Solution.
Data:
D := 25 ´ 10 –3 m Ts := 100°C Ta := 20°C g := 9.81 m/s2
Properties of air at film temperature of (100 + 20)/2 = 60°C are:
1
Tf := 60°C n := 18.97 ´ 10 –6 m2/s k := 0.02896 W/(mK) Pr := 0.696 b := 1/K
Tf + 273
i.e. b = 3.003 ´ 10 – 3 1/K
Diameter D is the characteristic dimension to calculate Ra.
A := p × D 2 m2 (surface area of sphere)
i.e. A = 1.963 ´ 10 –3 m2 (surface area)
D 3 × g × b × (Ts - Ta )
And, Gr :=
v2
i.e. Gr = 1.023 ´ 10 5 (Grashoff number)
and, Ra := Gr × Pr
i.e. Ra = 7.122 ´ 10 4 (Rayleigh number)
Now, use Eq. 10.82:
Nuf = C × (Gr × Pr) n ...(10.82)
From the Table, for Ra = 7.122 ´ 10 4, we get:
1
C = 0.54 and, n =
4
Therefore, Nu f := C × Ra n (Nusselt number)
i.e. Nuf = 8.822 (Nusselt number)
Nu f × k
And, h := W/(m 2K) (heat transfer coefficient)
D
i.e. h = 10.219 W/(m2K) (heat transfer coefficient)

NATURAL (OR FREE) CONVECTION 515


Compare this value of h with h = 10.454 W/(m 2K) obtained earlier.
Therefore, rate of heat loss from sphere:
Q := h×A × (Ts – Ta) W
i.e. Q = 1.605 W.
Compare this value of Q with Q = 1.642 W, obtained earlier.
Example 10.24. Work out Example 10.16 with formula from Russian literature: A sphere of 0.15 m diameter stores a
brine at – 5°C and is insulated by enclosing it in another sphere of 0.2 m diameter and the intervening space contains air
at 1 bar. The outside sphere is at 25°C. Estimate the convection heat transfer rate.
Solution.
Data:
Di := 0.15 m Do := 0.2 m b := 0.025 m Ti := – 5°C To := 25°C g := 9.81 m/s2
We need properties of air at average temperature Tf = (25 – 5)/2
Tf := 10°C ...average temperature
Properties of air at 10°C:
1
n := 14.19 ´ 10 –6 m2/s k := 0.02487 W/(mK) Pr := 0.716 b := i.e. b = 3.534 ´ 10 –3 1/K
Tf + 273
Remember that here, the gap between spheres ‘b’ is the characteristic dimension.
g × b × (To - Ti ) × b 3
Then, Gr b :=
v2
i.e. Grb = 8.07 ´ 104 (Grashoff number)
Therefore, Rab := Grb × Pr
i.e. Ra b = 5.778 ´ 10 4 (Rayleigh number)
For this value of Ra, appropriate equation is Eq. 10.86 viz.
keff
= 0.105 ×Ra b0.3 (for 10 3 < Rab < 106...(10.86))
k
keff
i.e. = 2.817
k
i.e. keff := 2.817× k
i.e. keff = 0.07 W/(mK) (effective thermal conductivity)
Compare this value with keff = 0.059 W/(mK), obtained earlier.
Therefore, rate of heat loss:
FG D × D IJ ×(T – T )
H b K
i o
Q := p × k eff i o ...(10.66)

i.e. Q = – 7.923 W
Note: Negative sign indicates that heat flow is from outside to inside.
Compare this value of Q with Q = – 6.694 W, obtained earlier.

10.7 Combined Natural and Forced Convection


In many practical situations, natural and forced convection may occur together. At high velocities forced convec-
tion may be predominant, but at low velocities effect of natural convection also must be included. Further, natu-
ral and forced convection may occur in the same direction or they may act in opposite directions. We have the
following criteria to determine if the combined free and forced convection is to be considered.
GrL/(ReL2) << 0.1 (forced convection regime (negligible free convection)
GrL/(ReL2) >> 10 (free convection regime (negligible forced convection)
0.1 < GrL/(ReL2) » 10 (mixed convection regime (both free and forced convection are important)
In the mixed convection regime, following equation is used to calculate the Nusselt number:
Num = Nuforcedm ± Nufreem .....(10.89)
where first and second terms on RHS are Nusselt numbes for forced and free convection respectively. A value of
m = 3 is generally recommended. Positive or negative sign is taken if the free convection flow occurs in the same
or opposite direction to that of forced convection.
For the specific case of mixed convection for internal flow through a horizontal pipe, we have the follow-
ing correlations for average Nusselt number:

516 FUNDAMENTALS OF HEAT AND MASS TRANSFER


For laminar flow (ReD £ 2000): (correlation due to Brom and Gauwin):

Fm I
= 1.75 × G
0 .14
L
× MGz + 0.012 × F Gz ×Gr IK 4
OP 1
3

H m JK H
1 3
NuD b
MN PQ ...(10.90)
3
D
s
where m b and ms are viscosities of the fluid at the bulk mean temperature and surface temperature respectively,
and Gz is the Graetz number, given by:
FG D IJ = Graetz number.
Gz = ReD × Pr ×
H LK ...(10.91)

For turbulent flow : (correlation due to Metais and Eckert):

FG D IJ 0. 36
NuD = 4.69 ×Re D0.27× Pr0.21 × Gr D0.07
H LK ...(10.92)

Example 10.25. An un-insulated pipe of 50 mm OD, with a surface temperature of 50°C, runs through a plant room. An
exhaust fan creates a mild flow of air upwards across the pipe, with a velocity of 0.2 m/s. If the ambient temperature is
30°C, calculate the rate of heat loss by combined free and forced convection.
Solution.
Data:
D := 0.05 m L := 1 m V := 0.2 m/s Ts := 50°C Ta := 30°C g := 9.81 m/s2
We need properties of air at average temperature T f = (50 + 30)/2
Tf := 40°C (average temperature)
Properties of air at 40°C:
1
n := 16.96 ´ 10 – 6 m2/s k := 0.0271 W/(mK) Pr := 0.71 b := i.e. b = 3.195 ´ 10 – 3 1/K
Tf + 273
Remember that here, the diameter of pipe, D, is the characteristic dimension.
g × b × (Ts - Ta ) × D 3
Then, GrD :=
n2
i.e. GrD = 2.724 ´ 10 5 (Grashoff number)
Therefore, RaD := GrD × Pr
i.e. RaD = 1.934 ´ 10 5 (Rayleigh number)
Now, Reynolds number is given by:
D ×V
Re :=
n
i.e. Re = 589.623 (Reynold number)
Therefore,
Gr D
= 0.784 (this value is nearly equal to one. Therefore, flow is in mixed convection
Re 2 regime. i.e. both free and forced convection must be considered.)
Free convection Nusselt number:
From Eq. 10.43:

LM LM OP OP
1
6
2

MM M Ra PP PP
= M0 .60 + 0 . 387 × M
MM L F 0.559 I PP PP
D
Nufree (10 – 5 < RaD < 10 12...(10.43))
MM OP 16

MN MMN1 + GH Pr JK
9 9

PQ PP
16

MN PQ Q
i.e. Nufree = 9.261 (Free convection Nusselt number)
Forced convection Nusselt number:
Using Churchill and Burnstein correlation:

NATURAL (OR FREE) CONVECTION 517


LM F1 1
IJ 5
OP 4
5

LM F I OP MN GH
h×D 0 .62 × Re 2 × Pr 3 Re 8
Nu cyl =
k
= 0.3 +
2/3 1/4
× 1+
28200 K PQ ...(9.90)

G J
0. 4
NM H K PQ
1+
Pr

for 100 < Re < 10 7 and Re × Pr > 0.2


From Eq. 9.90 we get:
Nuforced = 12.927 (Forced convection Nusselt number)
We use Eq. 10.89 to determine the Nusselt number for mixed convection. Also, we use the ‘+’ sign, since both the
free convection from hot pipe and forced convection with air flowing from bottom of pipe upwards, are additive. We use
the value of exponent as 3.
i.e. Nu mixed3 = Nufree3 + Nu forced3 ...(10.89)
i.e. Numixed3 = 794.293 + 2160 = 2954
Therefore, Nu mixed := 14.348 (mixed Nusselt number)
Therefore, combined heat transfer coefficient:
Numixed k
h :=
D
i.e. h = 7.777 W/(m2K) (combined heat transfer coefficient)
Heat loss per meter length of pipe:
Q = h × (p × D × L) × (Ts – Ta) W/m
i.e. Q = 19.545 W/m.
Example 10.26 Air at 1 atm. and 20°C is forced through a 15 mm diameter tube at an average velocity of 20 cm/s. Tube
wall is maintained at a temperature of 100°C. The tube is 1 m long. Calculate the rate of heat transfer.
Solution.
Data:
D := 0.015 m L := 1.0 m V := 0.2 m/s Ts := 100°C Ta := 20°C g := 9.81 m/s2
Properties of air at bulk mean temperature of 20°C:
1
n := 15.06 ´ 10 – 6 m2/s mb := 18.14 ´ 10 – 6 kg/(ms) k := 0.02593 W/(mK) Pr := 0.703 b :=
Ta + 273
i.e. b = 3.413 ´ 10 – 3 1/K ms := 21.87 ´ 10 –6 kg/(ms) (dynamic viscosity of air at surface temperature of 100°C.)
Reynolds number:
V ×D
ReD := (Reynolds number)
n
i.e. ReD = 199.203 (Reynolds number)
Grashoff number:
g × b × (Ts - Ta ) × D 3
GrD := (Grashoff number)
n2
i.e. GrD = 3.986 ´ 10 4 (Grashoff number)
Therefore,
GrD
= 1.004
Re D2
Therefore, this is a case of mixed convection i.e. both free and forced convection are to be considered. And, Eq.
10.90 for average Nusselt number is applicable. i.e.
F m I × LGz + 0.012 ×eGz ×Gr j O
0 . 14 1

GH m JK MN PQ
4 3
b
NuD = 1.75
1 3
3
D ...(10.90)
s

Now, Graetz number is:


D
Gz := ReD × Pr × (Graetz number)
L
i.e. Gz = 2.101 (Graetz number)

518 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore,
F m I × LGz + 0.012 ×eGz ×Gr j O
0 . 14 1

GH m JK MN PQ
4 3
b 1
Nu D := 1.75 ×
3
3
D
s

i.e. NuD = 3.041 ...Average Nusselt number


NuD × k
Then, h := W/(m2K) ...heat transfer coefficient
D
i.e. h = 5.258 W/(m2K) ...heat transfer coefficient
Heat transfer rate/m length:
Q := h × (p × D × L) × (Ts – Ta) W/m
i.e. Q = 19.821 W/m.

10.7 Summary of Basic Equations for Natural Convection


Important correlations are summarized below:

Geometry Correlation
Heated, vertical plate: Integral method: Temperature distribution:

T - Ta y FG IJ 2

Ts - Ta
= 1-
d H K
Velocity distribution:

u y yFG IJ 2

ux
= × 1-
d d H K
Maximum velocity:
4
umax = × u x at y = d / 3.
27
Mean velocity:
1 27
um = × ux = × u max
12 48
Velocity function:

u x = 5.17 ×n ×(Pr + 0.952) –0.5×


LM b ×g ×(T s - Ta ) OP 0. 5
× x 0.5
N n 2
Q
Boundary layer thickness:

b
d 3.93 × 0.952 + Pr
=
g 0 . 25

x Grx0 . 25 ×Pr 0 . 5
Total mass flow through the boundary:

m total = 1.7r×n
LM GrL OP 0 . 25

N Pr 2
×(Pr + 0.952) Q
Average Nusselt number for laminar flow:
4 0.0667 ×Pr 0 . 5 ×GrL0 . 25
Nu avg = × Nu L =
3 (0.952 + Pr )0 . 25
Average Nusselt number for turbulent flow:

N uavg =
havg ×L
= 0.0246
LM Pr ×Gr
1.17
L
OP 0. 4

...for turbulent flow.


k MN 1 + 0.495 ×Pr 2
3
PQ
Contd.

NATURAL (OR FREE) CONVECTION 519


Contd.

Empirical relations:
Vertical plate, Ts = constant Height L is the characteristic length.
1
For air and other gases Nu = 0.59 × Ra 4 ...104 < Ra < 109
1
Nu = 0.13 × Ra 3 ...109 < Ra < 10 12

1
0.670 ×Ra 4
For all Prandl numbers:b 0< Pr < ¥: Nu = 0.68 + ...0 < Ra < 109
LM F 0.492 I OP
4
9 9

MMN1 + GH Pr JK
16

PPQ
1
0.15 ×Ra 3
0.6 < Pr < ¥: Nu = ...Ra > 109
LM F 0.492 I OP
16
(For high Prandtl No. fluids) 9 27

MMN1 + GH Pr JK
16

PPQ

LM OP 2

MM 1
PP
Nu = M0.825 + PP
0.387 ×Ra 6
0 < Pr < 0.6: (Entire range of Ra)
MM LM F 0.492 I 9 OP
8
PP
... Ra > 109

MM
27
(For low Prandtl No. fluids i.e. liquid metals)
MM1 + GH Pr JK
16

MN PP PPQ
N Q
Inclined plate, Inclined height L is the characteristic length.
inclined at an angle q to the vertical Use vertical plate equations as a first approximation.
Ts = constant Replace g by g × cos (q ).
Vertical cylinder Height L is the characteristic length.
Vertical cylinder can be treated as vertical plate, if the following
relation is satisfied:
D 34
³ 1
L
Ra 4
Vertical plate, qs = constant Eqs. 10.25 and 10.26 are still valid, with the modification that
constant 0.492 is changed to 0.437.
Alternatively:
A modified Grashoff number is defined:
g × b ×q s × x 4
Gr ¢ = Gr × Nux =
k ×v 2
And following two relations for local Nusselt no.:
Nux = 0.60 (Gr ¢× Pr )0.2 ...105 < Gr ¢x < 1011
Nux = 0.17 (Gr ¢× Pr) 0.25 ...Gr x > 1011
For Average Nu :
5
h= × hL ...for laminar
4
Contd.
520 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Contd.

h = hL ...for turbulent.
Horizontal plate, Ts = constant Characteristic Length: Lc = A / P
Upper surface of hot plate (or lower surface of cold plate):
1
Nu = 0.54 × Ra 4 ...104 < Ra < 107
1
Nu = 0.15 × Ra ...107 < Ra < 1011
3

Lower surface of hot plate (or upper surface of cold plate):


1
Nu = 0.27× Ra 4 ...105 < Ra < 1011

Horizontal plate, qs= constant Characteristic Length: Lc = A/P


All property values, except b, are evaluated at a temperature,
Te, defined by:
Te = Ts – 0.25 (Ts – Ta)
and, b is evaluated at Ta.
Upper surface of hot plate (or lower surface of cold plate):
1
Nu = 0.13 × Ra 3 ...Ra < 2 ´ 108
1
Nu = 0.16× Ra 3 ...2 ´ 10 8 < Ra < 1011
For heated surface facing downward:
Nu = 0.58 Ra0.2 ...106 < Ra < 1011

Horizontal cylinder, Ts= constant Diameter D is the characteristic length.


For air:
Nu = C × Ran
C and n from Table 10.2.
For (0 £ Pr £ ¥):

LM LM OP
1
6
OP 2

MM MM PP PP
MM M Ra PP PP
Nu =
MM0.60 + 0.387 × MM L OP
16
PP PP ...10 –5 < Ra < 10 12

MM MM1 + FGH 0.Pr559 IJK


9 9

MM 16

PPQ PPQ PP
MN NM MN PQ
And, only for laminar range:
1
0.518 ×Ra 4
Nu = 0.36 + ...10 –6 < Ra < 109
LM F 0.559 I OP
4
9 9

MM1 + GH Pr JK
16

PP
N Q
1
For thin wires: (D = 0.2 to 1 mm) NuD = 1.18 × (RaD ) 8 ...Ra < 500

1
From horizontal cylinders to liquid metals: NuD = 0.53 × (GrD ×Pr 2 ) 4

Contd.

NATURAL (OR FREE) CONVECTION 521


Contd.

Spheres: Diameter D is the characteristic length.


1
Nu = 2 + 0.43 × (Ra ) 4 ...1 < Ra < 105, Pr = 1
And, for higher range of Ra:
1
Nu = 2 + 0.50 × (Ra ) 4 ...3 ´ 105 < Ra < 8 ´ 108
Rectangular blocks: Ch. Lengh:
LH × LV
L=
LH + LV
1
NuL = 0.55 × (RaL ) 4 ...104 < RaL < 10 9
Short cylinders (D = H) Nu = 0.775(Ra) 0.208
Simplified equation for air: Refer to Table 10.3
Free convection in enclosed spaces: Space between the plates, ‘b ’ is the characteristic dimension.
g × b ×(T1 - T2 ) × b 3
Gr b =
n2
For Horizontal enclosure:
1
For air: Nu = 0.195 × G r 4 ...104 < Gr < 3.7 ´ 10 5
and,
1
and, Nu = 0.068 × G r 3 ...3.7 ´ 105 < Gr < 107
And, for Gr < 1700, we have Nu = 1.
1
For liquids (water, silicone oils and mercury): Nu = 0.069 × Ra 3 Pr 0.074 ...1.5 ´ 105 < Ra < 109
For Vertical enclosure:
For Air: For Gr (based on plate spacing ‘b’) < 1700, we have Nu = 1.
1
k eff 0.18 ×Gr 4
= Nu = ...2 ´ 104 < Gr < 2 ´ 105
FG IJ
1
k L 9
b H K
where keff = effective thermal conductivity.
and,
1
k eff 0.065 ×Gr 4
= Nu = ...2 ´ 105 < Gr < 10 7
FG IJ
1
k
L 9
b H K
Note that for above two relations, aspect ratio, L /b > 3.
If L/b < 3, each vertical surface is treated independently.
Nu = 1 ...for Ra < 1000
and,
1
For fluids with Pr = 3 and 30,000 0.28 ×Ra 4
Nu = ...for 1000 < Ra < 10 7
inside vertical enclosure:
FG IJ
1
L 4

H K
b

Contd.

522 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.

Free convection in inclined spaces: For (L /d < 12) and at tilt angles t less than 70 deg.:
F 1708 I LM
. × t ))1. 6
1708 ×(sin (18 OP
(Flat plate solar collectors and double
glazed windows)
Nu L = 1 + 1.44 1 - GH RaL × cos (t )
× 1- JK N
RaL × cos (t ) Q
LF Re × cos (t ) I OP
+ MG
1

MMH 5830 JK PP
L 3
-1
N Q
If the quantity in the first bracket and the last bracket is negative,
then it must be set equal to zero.
For tilt angles between 70 deg. and 90 deg. Catton recommends
that the Nusselt number for a vertical enclosure (t = 90 deg.) be
multiplied by (sin t )1/4, i.e.
1
NuL (t ) = NuL (t = 90) × (sin (t )) 4
Natural convection inside D × havg
spherical cavities: = C × (GrD × Pr ) n
k
For values of C and n, see table in text.

Concentric cylindrical annuli: ‘b’ is the gap or thickness of the enclosed fluid layer
(i.e. b = [Do – Di]/2).
Q 2× p × k eff (Ti - To )
L
=
F I
Do
ln GH JK
Di

F Pr IJ
1
1
k eff
= 0.386 × G
4
...100 < Racc < 107.
k H 0.861 + Pr K ×Racc4

and,

F ln F D I I ×Ra 4

GH GH D JK JK o

i
b

Ra =cc
LM 1 1 OP 5

b ×M
MN D D PPQ
3
+ 3 3
5 5
i o

Q = p ×k × G
F D ×D IJ × (T – T )
H b K
i o
Concentric spherical annuli: eff i o

F Pr IJ ×Ra
1
1
k
= 0.74 × G
4
...102 < Racs < 104
H 0.861 + Pr K
eff 4
cs
k
and,

b ×Rab
Racs =
LM 1 1
OP 5

D ×D 4
o
4
i × MM 7
+ 7 PP
ND i
5
D o5 Q
Contd.

NATURAL (OR FREE) CONVECTION 523


Contd.

Cooling of turbine blades:


(hole diameter D, hole (rm ×w 2 ) × b × D ×T ×L3
Gr L =
length, L) n2
Mostly, GrL > 1012, and we use:

h ×L
LM Pr ×Gr
1.17
OP 0.4

Nua = a
k
= 0.0246 × MM L
2 PP
N 1 + 0.495 ×Pr 3
Q
Total heat transferred:
Q = ha ×(p × d × L) × (Ts – Ta )
Rotating cylinders: Peripheral-speed Reynolds number:
p ×D 2 ×w
Re w =
n
For (Re w > 8000 in air): Average Nusselt number:

NuD =
hc ×D
k
d
= 0.11 0.5 ×Re w2 + GrD ×Pr i 0 . 35

Rotating disk: For laminar regime, average Nu for a disk rotating in air:

F I
1
ha ×D w ×D 2 2 w ×D 2
NuD =
k
= 0.36 ×
n GH JK ...for
n
< 10 6

For laminar flow between r = 0 and r = rc, and turbulent flow between
r = rc and r = r0, average value of Nusselt number is given by:

F I ×F r I F w ×r I × LM1 - F r I OP
1
2 0 .8 2. 6
2 2
h ×r w ×r 2
Nu r = c o = 0.36× GH o
JK GH r JK c
+ 0.015 × GH n JK M GH r JK P
o c
k n o
N o
Q
...for rc < ro

For turbulent regime, local value of Nu at a radius r is given by:

hc × r w ×r 2 F I 0 .8

Nur =
k
= 0.0195 ×
n GH JK
Rotating sphere: For Pr > 0.7, in laminar flow regime, (i.e. Re w = w . D 2/n < 5 ´ 104),
average Nusselt number is given by:
NuD = 0.43 Re w0 . 5 Pr 0.4 ...Re w < 5 ´ 104
and,
NuD = 0.066 Re w0 . 67 × Pr 0.4 ...5 ´ 10 4 < Rew < 7 ´ 10 5
Rectangular fins on a vertical surface: Optimum fin spacing:
L
See Fig. 10.6. S opt = 2.714 1
where L = fin length in vertical direction and is also
Ra 4 the characteristic length.

k
h = 1.31 ×
S opt
Rate of heat transfer:
Q = h × (2 × n × L × H ) × (Ts – Ta ) where n = no. of fins
Contd.

524 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.
Rectangular fins on a
horizontal surface: For fins facing upwards for Ts > Ta (or facing downward for Ts < Ta):
See Fig. 10.7

LF 1500 I OP
-1

= MG
2 2

Nus
MNH Ra JK
s
+ (0.081×Ra s0 . 39 )- 2
PQ
Above equation is valid over the range:
200 < Ras < 6 ´ 105, Pr = 0.71, 0.026 < H/W < 0.19, and 0.016 < S/W
< 0.20, with the following definitions:
q ×S
Nus =
(Ts - Ta ) × k
and,
g × b × (Ts - Ta ) ×S 3
Ras =
n ×a

Combined Natural and


GrL/(ReL2) << 0.1 ....forced convection regime (negligible free convection)
Forced convection
GrL/(ReL2) >> 10 ....free convection regime (negligible forced convection)
0.1 < GrL/(ReL2) » 10 ....mixed convection regime (both free and forced
convection are important)
In the mixed convection regime, following equation is used to calculate
the Nusselt number:
Num = Nuforcedm ± Nufreem
A value of m = 3 is generally recommended. Positive or negative sign is
taken if the free convection flow occurs in the same or opposite direc-
tion to that of forced convection.

10.8 Summary
In natural (or, free) convection, fluid flow is caused by density differences as a result of temperature differences.
Natural convection is, in fact, a preferred mode of heat transfer in many practical applications, since there is no
need for an external fan or pump to cause the flow, and is therefore more economical and reliable. Cooling of
electronic devices, transformers, motors, transmission lines, etc. are some of the common examples of applica-
tions of natural convection heat transfer.
In this chapter, first, an outline of the method of solution of the relevant conservation equations by the
approximate integral method was given. Solutions for the case of natural convection are move difficult as com-
pared to the case of forced convection since in the case of natural convection, the momentum and energy equa-
tions are ‘mutually coupled’ which means that they have to be solved simultaneously. Next, empirical relations
for several geometries and situations of practical importance were listed. Several examples have been worked
out, demonstrating the use of these correlations.

Questions
1. Explain the circumstances under which natural convection occurs. Differentiate between natural and forced
convection.
2. What is the criterion from laminar to turbulent flow in natural convection?
3. What is the physical significance of Grashoff number? Compare it with Reynolds number.
4. Use the principle of dimensional analysis to establish a relationship between Nusselt number, Grashoff number
and Prandtl number. [M.U.]
5. Why is the analytical solution of free convection problems more involved as compared to forced convection
problems?
6. State two important applications of heat transfer in an enclosure. What is meant by ‘aspect ratio’ of an enclo-
sure?

NATURAL (OR FREE) CONVECTION 525


7. What is a ‘heat sink’? Why are fins provided in a heat sink?
8. Explain why there is an ‘optimum’ spacing between fins in a heat sink.
9. What is the criterion to decide if the natural convection is negligible or not, in forced convection heat transfer?
10. How is the Nusselt number calculated in ‘mixed convection’ regime?

Problems
1. A hot plate 35 cm high and 1.1 m wide at 160°C is exposed to ambient air at 20°C. Using the approximate
solution, calculate the following:
(i) Maximum velocity at 10 cm from the leading edge of the plate (ii) boundary layer thickness at 10 cm from the
leading edge of plate (iii) local heat transfer coefficient at 10 cm from the leading edge of the plate (iv) average
heat transfer coefficient over the surface of the plate (v) total mass flow through the boundary (vi) total heat loss
from the plate, and (vii) temperature rise of air.
2. A hot plate 25 cm height and 100 cm width is exposed to atm. air at 20°C. The surface temperature of plate is
100°C. Find the heat loss from both the surfaces of the plate. If the height of the plate is changed to 50 cm, what
will be the change in heat loss? Following empirical relation may be used: Nu = 0.59(Gr ×Pr)1/4
Properties of air at average temperature are:
r =1.06 kg/m3; n =18.97 ´ 10 – 6 m2/s ; Cp =1004 J/kgK ; k = 0.029 W/mK [M.U.]
3. A hot plate 100 cm height and 25 cm width is exposed to atm. air at 25°C. The surface temperature of plate is
95°C. Find the heat loss from both the surfaces of the plate. If the height of the plate is reduced to 50 cm and the
width is increased to 40 cm, what will be the change in heat loss? Following empirical relation may be used:
Nu = C.(Gr . Pr)m where
C = 0.59 and m = 1/4 for (Gr ×Pr) < 109, and
C = 0.1 and m = 1/3 for (Gr × Pr) > 109
Properties of air at average temperature are:
r = 1.06 kg/m3; n =18.97 ´ 10 –6 m2/s ; Cp = 1004 J/kgK; k= 0.029 W/mK [M.U.]
4. A hot square plate 40 cm x 40 cm at 100°C is exposed to atmospheric air at 20°°C. Find the heat loses from both
surfaces of the plate if:
(a) the plate is held horizontal.
(b) the plate is held in vertical plane.
Properties of air at average temperature are: r =1.06 kg/m3; n =18.97 ´ 10 –6 m2/s
Cp = 1004 J/kgK; k = 2.89 ´ 10-2 W/mK;
Following empirical relations may be used to find average heat transfer coefficients:
Case(a): Nu = 0.13(Gr × Pr)1/4
Case(b): For lower surface Nu = 0.35(Gr × Pr)1/4
For upper surface Nu = 0.71(Gr × Pr)1/4 [M.U.]
5. A flat, vertical electrical heater is 0.5 m ´ 0.5 m in size and dissipates heat to still, ambient air at 20°C. Heat
generation rate is 1 kW/m2. Determine the average heat transfer coefficient and the average surface tempera-
ture.
6. A vertical steel plate, 0.5 m ´ 0.5 m in size and 3 mm thick, at an uniform temperature of 160°C, is exposed to
atmospheric air at 20°C. Find the approximate time required for the plate to cool to 30°C, if the heat transfer
coefficient in natural convection for the vertical plate is given by: h = 1.42 ´ (DT/L)1/4. For steel, r = 7800 kg/m3,
Cp = 473 J/(kgK).
7. A 4 cm diameter steel ball at 160°C loses heat only by free convection to ambient air at 20°C. Calculate the time
required for the temperature of the ball to reach 30°C. For steel, r = 7800 kg/m3, Cp = 473 J/(kgK).
8. (a) A vertical pipe, 7.5 cm OD, 1.8 m long, has a surface temperature of 90°C. If the surrounding air is at 30°C,
what is the rate of heat loss by free convection from this cylinder?
(b) If the pipe is inclined to the vertical at an angle of 30 deg. during installation, how does the heat loss/m
change?
9. A horizontal metal plate, 0.6 m x 0.6 m, is exposed to sun and receives radiant energy at the rate 170 W/m2. If
the heat transfer from the plate occurs to the surrounding air at 20°C by free convection only, find the steady
state temperature of the plate. Assume that the bottom of the plate is insulated.
10. A horizontal, steam pipe of 10 cm OD runs through a room where the ambient air is at 20°C. If the outside
surface of the pipe is at 160°C, and the emissivity of the surface is 0.85, find out the total heat loss per metre
length of pipe.
11. A horizontal pipe carrying steam passes through a large room and is exposed to air at 30°C. The outer diameter
of pipe is 20 cm. If the surface temperature of pipe is 200°C, find the loss of heat per metre length of the pipe by

526 FUNDAMENTALS OF HEAT AND MASS TRANSFER


convection and radiation. Assume emissivity of the pipe surface as 0.8. Properties of air given below:
k = 0.0331 W/mK; r = 0.8826 kg/m3; n =24.83 ´ 10 – 6; Cp = 1014 J/kgK.
Approximate empirical relation can be assumed as Nu = 0.53 (Ra)0.25 for (103 < Ra < 109). [M.U.]
12. A tank contains water at 20°C. The water is heated by passing steam through a pipe placed in water. The pipe is
3 m long and 5 cm in diameter and its surface is maintained at 100°C. Find the heat loss from the pipe if the pipe
is kept horizontal.
Following empirical relations may be used: Nu = C.(Gr.Pr)m, where
C = 0.53 and m = 0.25 when 104 < Gr.Pr < 109, and
C = 0.13 and m = 1/3 when Gr.Pr > 109.
13. A fine wire of 0.5 mm diameter is maintained at a constant temperature of 260°C by an electric current. The wire
is exposed to air at 1 bar and 20°C. Calculate the heat transfer coefficient and the current flowing through the
wire to maintain the wire temperature if the length of wire is 1 m. Electrical resistance of the wire is 8 ohms/m.
14. A spherical bulb of 5 cm diameter, with its surface temperature at 120°C, is exposed to still air at a temperature
of 20°C. Determine the rate of convective heat loss.
15. A metal block is of 6 cm x 9 cm section and is 15 cm in height. Surface temperature of the block is 80°C. If it is
exposed to air at 20°C, determine the rate of convective heat loss.
16. A circular disk heater of 3 cm diameter is maintained at a temperature of 60°C and is exposed to ambient air at
20°C. Calculate the free convection heat loss.
17. A short, solid vertical cylinder, 15 cm diameter and 15 cm high, is at 260°C and is exposed to air at 40°C.
Estimate the value of average surface coefficient of heat transfer for the entire outside surface.
18. Solve Problem 10.11 using simplified equations for air. Refer to Table 10.3 for appropriate equation.
19. Helium gas at 2 bar pressure is contained between two horizontal panels separated by a distance of 20 mm. The
lower panel is at a temperature of 80°C and the upper panel is at 20°C. Calculate the heat transfer rate by free
convection per sq. m. of the panel surface.
20. Air gap between the two glass panels of a double-pane window (0.8 m wide ´ 1.5 m high) is 2 cms. If the two
glass surfaces are at 25°C and – 5°C,
(a) determine the rate of heat transfer through the window.
(b) Verify the result with formula from Russian literature.
(c) Show graphically how the heat transfer coefficient varies as the gap spacing.
21. Two vertical plates of size 40 cm ´ 40 cm are separated by a space of 3 cm and the gap is filled with water. Plate
temperatures are 60°C and 20°C. What is the heat transfer rate? Verify the result with formula from Russian
literature.
22. In a solar flat plate collector, the plate is 1.5 m high and 2.5 m wide, and is at a temperature of 120°C. The glass
cover plate is at a distance of 3 cm from the collector surface and its temperature is 40°C. Space in between
contains air at 1 atm. If the collector plate is inclined to the horizontal at 30 deg., determine the heat transfer
coefficient and the rate of heat loss.
23. (a) Air at 1 bar fills the gap between two concentric spheres of 10 cm and 8 cm, respectively. Inner sphere is at
90°C and outer sphere is at 30°C. Calculate the free convection heat transfer across the gap. (b) Verify the result
with formula from Russian literature.
24. A long tube of 0.2 m OD is maintained at 130°C. It is surrounded by a cylindrical radiation shield, located
concentrically, such that the air gap between the two cylinders is 20 mm. The shield is at a temperature of 30°C.
Estimate the convection heat transfer rate per metre length. Verify the result with formula from Russian litera-
ture.
25. A turbine blade is cooled by free convection with water as coolant. The cooling passage is 8 mm in diameter and
7 cm long. The blade velocity at a mean radius of 20 cm is 210 m/s. The hole surface temperature is at 200°C and
cooling water temperature is 60°C. Find the average heat transfer coefficient and the rate of heat loss.
26. A 15 cm diameter steel shaft whose surface is at 150°C is allowed to cool while rotating about its own horizontal
axis at 3 r.p.m. in an environment of air at 30°C. Find the initial rate of heat loss.
27. A 2 m diameter disk rotates at 600 r.p.m. has its surface at 60°C. Surrounding air is at 20°C. Find the value of
convection coefficient and the rate of heat transfer from one side.
28. A sphere, 0.1 m in diameter is rotating at 30 r.p.m. in ambient air at 20°C. The sphere is at 160°C. Estimate the
rate of heat transfer.
29. Consider a vertical heat sink with fins as shown in Fig. Problem 29.
The vertical heat sink, 0.35 m wide x 0.15 m high, is provided with vertical, rectangular fins of 1 mm thickness
and 20 mm length. Base and surface temperature of fins is 80°C and the surrounding air is at 20°C. Determine
the optimum fin spacing and the rate of heat transfer from the heat sink by natural convection.

NATURAL (OR FREE) CONVECTION 527


W = 0.35 m H = 20 mm

Quiescent air, Ta = 20 C

L = 0.15 m

Ts = 80 C

t = 1 mm

FIGURE Problem 10.29 Free convection from vertical heat sink with fins

30. Water at 20°C with a velocity of 5 cm/s flows across a horizontal cylinder maintained at a temperature of 60°C.
Is the heat transfer by free convection significant? If so, calculate the rate of heat loss by combined free and
forced convection. What will be the situation if the fluid is air at atmospheric pressure?
31. Consider a 3 m long vertical plate at a temperature of 80°C, kept in still air at 20°C. What is the forced motion
velocity above which free convection heat transfer from the plate is negligible?
32. Atmospheric air flows through a 25 mm diameter horizontal tube at an average velocity of 25 cm/s. The tube is
maintained at 150°C and the bulk air temperature is 30°C. Estimate the heat transfer coefficient if the tube is 0.35
m long.

528 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

11
Boiling and
Condensation

11.1 Introduction
In the previous two chapters, we studied heat transfer in forced and free convection, i.e. heat transfer with fluid
motion, induced either by an external means or by density differences. In both the cases, fluid involved was
homogeneous and in single phase. But, there are many important practical cases which involve heat transfer with
a change of phase of the fluid, e.g. boiling where the liquid changes to vapour and condensation where the
vapour condenses into a liquid. Boiling and condensation are classified under convection since there is motion of
the fluid during heat transfer in these processes.
Some of the applications of boiling and condensation are:
(i) Evaporators and condensers of a vapour compression refrigerating system
(ii) Boilers and condensers of a steam power plant
(iii) Reboilers and condensers of distillation columns of cryogenic and petrochemical plants
(iv) Cooling of nuclear reactors and rocket motors
(v) Process heating and cooling, etc.
Unique features of boiling and condensation are:
(i) heat transfer, practically at a constant temperature, because of change of phase
(ii) latent heat and surface tension come into play in addition to buoyancy driven flow effects, resulting in
larger heat transfer rates and heat transfer coefficients compared to the usual free or forced convection
(iii) high heat transfer rates with small temperature difference.

11.2 Dimensionless Parameters in Boiling and Condensation


It is difficult to obtain governing equations for boiling and condensation by applying the usual conservation
laws. However, dimensional analysis has been successfully applied with the use of Buckingham p theorem. Heat
transfer coefficient in either boiling or condensation process may be reasonably assumed to depend on the tem-
perature difference DT between the surface temperature Ts and the saturation temperature Tsat of the fluid, body
force arising out of the density difference between the liquid and vapour phases {= g.( r l – r v )}, latent heat hfg,
surface tension s , a characteristic length L and the thermo-physical properties of the liquid or vapour; r , Cp , m
and k. Applying the Buckingham p theorem, it can be shown that the functional relationship between the various
dimensionless groups is given by:
LM O
h×L
MN c hL3 Cp × DT m × cp
= f r × g× rl - rv × 2 , , c h PP
, g × rl - r v ×
L2
...(11.1)
k m hf g k
Q
s
where, the dimensional groups are defined as follows:
h×L
NuL = = Nusselt number
k
Cp ×DT
Ja = = Jakob number
hf g

m ×Cp
Pr = = Prandtl number
k
L2
Bo = g ×(r l – r v) ×= Bond number.
s
The first term inside brackets in Eq. 11.1, i.e.
L3
r × g × ( r l – r v )×
m2
represents the effect of buoyancy induced fluid motion on heat transfer.
Jakob number (Ja) involves the ratio of maximum sensible energy absorbed to the latent heat absorbed.
Generally, Ja has a small numerical value.
Bond number (Bo) is the ratio of gravitational body force to the surface tension force.

11.3 Boiling Heat Transfer


11.3.1 Boiling and Evaporation
‘Boiling’ occurs at the solid–liquid interface when the solid surface is at a temperature Ts , sufficiently above the
saturation temperature Tsat of the liquid at that pressure. In contrast, ‘evaporation’ occurs at the liquid–vapour
interface when the vapour pressure above the liquid is less than the saturation pressure of the liquid at the given
temperature. Unique feature of the boiling phenomenon is the production of vapour bubbles at the solid–liquid
interface causing intense mixing.
11.3.2 Boiling Modes
Boiling is generally classified as ‘pool boiling’ and ‘flow boiling’.
In pool boiling, there is no bulk fluid flow, and any motion of the fluid is due to natural convection and the
movement of bubbles under buoyancy effects. Heating of a liquid by immersing a heating element in it is an
example of pool boiling. When boiling occurs while fluid is in motion under the influence of a pump, it is called
flow boiling. These two modes of boiling are further classified as ‘sub-cooled boiling’ and ‘saturated boiling’. In
sub-cooled boiling, main body of the liquid is at a temperature below the saturation temperature Tsat, while in
saturated boiling, main body of the liquid is at a temperature equal to Tsat. During initial stages of boiling, we
have the subcooled boiling where bubbles originate at the heating surface, move up due to buoyancy effects, and
dissolve in the cooler liquid since the body of the liquid is at a temperature lower than Tsat. As the body of the
liquid reaches the saturation temperature, bubbles start reaching up to the free surface of the liquid and we say
that bulk or saturated boiling is set in motion.
Since boiling is a form of convection heat transfer, boiling heat flux is given by Newton’s law of cooling, i.e.
q boiling = h ×(Ts – Tsat) = h×DTe W/m2 ...(11.2)
where, DTe = (Ts – Tsat) = excess temperature.
11.3.3 Origin and Growth of Bubbles
Boiling is associated with bubble formation at the heating surface. Bubbles are believed to originate at the small
cavities present on the heating surface by expansion of entrapped gas or vapour. These are known as ‘nucleation
sites’. Then, the bubbles grow in size, detach themselves from the surface when a critical size is reached. Bubble
growth and dynamics is closely related to the surface tension (s ) at the liquid – vapour interface, temperature
excess and the pressure. Surface tension signifies the wetting capability of the liquid; lower the surface tension,
higher the wetting capability.
Wetting capacity of a liquid signifies the contact angle q between the wall and the free surface. See Fig. 11.1.

530 FUNDAMENTALS OF HEAT AND MASS TRANSFER


q q

q
q

(i) (ii) (i) (ii)

(a) Contact angle q for (i) wetting (b) Shape of vapour bubble for (i) wetting
and (ii) non-wetting liquids and (ii) non-wetting liquids

FIGURE 11.1 Contact angle and shape of vapour bubbles for wetting and non-wetting liquids

Larger the angle q , poorer is the wetting capacity. A liquid is considered to wet a surface if q < 90 deg. This
is true for liquids like water (q = 50 deg.), kerosene (q = 26 deg.), ether (q = 16 deg.) etc. If q > 90 deg. liquid does
not wet the surface, eg. mercury (q = 137 deg.).
Based on theory of capillarity, following equation gives the separation diameter ‘d’ of a bubble in a quiet
liquid:
s
d = 6.386 × q × mm ...(11.3)
(r l - r v)
where, q is contact angle in deg., s is surface tension in N/m, r is density of liquid or vapour in kg/m3.
For water boiling on a metal surface, Eq. 11.3 gets the following form:

FG r - r IJ
l v
3
d = 2.65
H 1000 K mm., (11.4)

At atmospheric pressure, for water, r l = 957.9 kg/m3, r v = 0.5978 kg/m3, and we get
d = 2.482 mm. Note that separation diameter diminishes with increasing pressure.
Intense mixing caused by the separation of the bubbles from the surface and their movement through the
fluid results in high rate of heat transfer from the surface to the liquid. It is obvious that higher the number of
nucleation sites where the bubbles originate and higher the frequency with which the bubbles detach from the
surface, higher the heat transfer coefficient, h. Heat transfer coefficient is a function of excess temperature, DTe , as
discussed below:
11.3.4 Boiling Regimes and Boiling Curve
Nukiyama performed his pioneering experiments on boiling heat transfer in 1934. He used nichrome and plati-
num wires which were electrically heated while immersed in liquids. In general, four different boiling regimes
are observed depending upon the excess temperature (DT e) imposed, namely
(i) natural convection boiling (DTe upto about 5 deg.C)
(ii) nucleate boiling (DTe from 5 deg to about 30 deg.C)
(iii) transition boiling (DTe from 30 deg to about 120 deg.C), and
(iv) film boiling (DTe beyond 120 deg.C).
Fig. 11.2 shows a typical boiling curve for water at one atmosphere pressure. General shape of the boiling
curve is same for other fluids as well. In Fig. 11.2, boiling heat flux is plotted against the excess temperature.
Also, shape of the boiling curve is independent of the geometry of the heating surface, but depends on the fluid
pressure and the specific fluid–heating surface combination.
A brief explanation of the different boiling regimes is given below:
(i) Natural convection boiling This range is up to the point ‘A’ in Fig. 11.2. No bubbles are formed up to a small
excess temperature of about 5 deg. and the liquid is superheated, rises to the free surface and evaporates from the
surface. In this range, the free convection correlations derived in the previous chapter can be applied to make
heat transfer calculations.

BOILING AND CONDENSATION 531


Natural convection Nucleate Transition Film
boiling boiling boiling boiling

Bubbles Maximum
collapse (critical)
6 in the C heat flux, qm xa
10 liquid E

2
qboiling, W/m
5
10
B

4
10 D
A Bubbles
rise to the Leidenfrost point, qmin
free surface

3
10
1 ~5 10 ~ 30 100 ~ 120 1000
DTexcess = Ts – Tsat, °C

FIGURE 11.2 Typical boiling curve for water at one atmosphere pressure

(ii) Nucleate boiling Region between ‘A’ and ‘C’ is the nucleate boiling region. Starting from point ‘A’, as DTe
increases, bubbles start forming at nucleation sites at an increasing rate. Nucleate boiling region may classified
into two sub-regions:
(a) region A–B, where the isolated bubbles formed rise up, but do not reach the free surface and collapse in
the body of the liquid; space vacated by the bubbles formed at the surface as they move up, is filled by
fresh liquid, and the process is repeated. Movement of the bubbles through the body of the liquid causes
agitation which is responsible for increasing heat transfer in nucleate boiling.
(b) region B–C, where the bubbles form at a faster rate at a largely increased number of nucleation sites and
rise up in the liquid in almost continuous columns of vapour. These bubbles gush up in the liquid and
reach the free surface and then collapse. Heat flux in this region is very large due to this reason. Note that
after point B there is an inflection in the boiling curve; this is because of the fact that as excess tempera-
ture is increased, the heating surface gets almost covered with bubbles and the heat flux increases at a
lower rate as DTe increases, and reaches a maximum at point C. Heat flux at point C is called ‘critical’ or
‘maximum’ or ‘burnout’ heat flux, qmax. For water, qmax > 1 MW/m2.
It should be clear that from heat transfer point of view, nulcleate boiling regime is the most desirable range
to operate, since very high heat transfer rates are obtained with relatively small DTe .(under 30°C).
(iii) Transition boiling Region between ‘C’ and ‘D’ is the transition boiling region. In this range, as the excess
temperature increases, the heat flux decreases; this is due to the fact that now a major portion of the heater
surface is covered by the vapour film which has a smaller thermal conductivity as compared to that of the liquid,
and, therefore, acts as an insulation. Between points C and D, nucleate and film boiling occur partially or alter-
nately and is therefore called ‘unstable film boiling regime’. At point D, excess temperature is of the order of
120°C.
(iv) Film boiling This region is beyond the point D. As excess temperature is further increased, now a stable,
vapour blanket completely covers the heater surface. So, at point D, the heat flux reaches a minimum and this
point is known as ‘Leidenfrost point’, (in honour of Leidenfrost, who explained in 1756 that the water drops
dropped on a very hot surface ‘dance’ on a vapour film and boil away). Now, as the excess temperature is
increased further, heat transfer by radiation effect also comes into picture in addition to conduction through the
vapour film, and the heat flux increases as shown.
11.3.5 Burnout Phenomenon
In Fig. 11.2, a continuous boiling curve was shown. However, in practice, when Nukiyama conducted his experi-
ments with an electrically heated nichrome wire immersed in a pool of water, he observed that when a little
excess power was supplied to the nichrome wire after reaching point C, wire temperature suddenly increased
uncontrollably to the melting point of the wire (i.e.1500 K) and burnout occurred. When the experiment was

532 FUNDAMENTALS OF HEAT AND MASS TRANSFER


q
qmax Sudden jump in temperature
2
(W/m )

Bypassed part of boiling


curve

Sudden drop in qmin


temperature
DTe
(deg.C)

FIGURE 11.3 Actual boiling curve for water heated by a platinum wire

repeated with platinum wire (which has a higher melting point of 2045 K), it was possible to maintain heat flux
higher than qmax without a burnout. Now, when the power was gradually reduced after qmin was reached at point
D, there was a sudden drop in the excess temperature, landing into the nucleate boiling region. Note that the arm
C–D of the boiling curve cannot be obtained in the power controlled mode of heating, unless the power applied
is reduced suddenly when point C is reached. The phenomenon of ‘hysterisis effect’ explained above is shown in
Fig. 11.3.
As we go on supplying electrical energy to the heater, point C (Fig. 11.2), i.e. the point of critical or maxi-
mum heat flux is reached; now, if we try to go past this point by increasing the heater power, the fluid is not able
to accept this increased power as shown in Fig. 11.3, and as a result, the heater temperature increases. So, DTe
increases, and the fluid can accept even lesser energy at this increased DT, and the heater temperature further
increases, and so on. Thus, a steady state point E is reached in the boiling curve Fig. 11.2, which, unfortunately,
corresponds to a very high surface temperature, that the heater may even melt or ‘burnout’. Hence, the name
‘burnout heat flux’ for the heat flux at point C.
Knowledge of ‘burnout flux’ is very important from practical point of view (in electrically heated surfaces
and nuclear reactors), since any attempt to go past the point C of maximum heat flux will make the surface
temperature to jump suddenly to point E, causing a burnout. So, the aim should be to operate at a point as near
to the point C as possible, but never to go beyond it. In cryogenic applications, however, point E falls at tempera-
tures much lower than the melting point of materials concerned, and film boiling can be adopted without any
danger of a burnout.
11.3.6 Heat Transfer Correlations for Pool Boiling
There are different correlations for the different regimes of boiling discussed above. Most of these correlations are
empirical, since, as already mentioned, phenomenon of boiling is not easily amenable to theoretical analysis.
Natural convection boiling regime (i.e. up to an excess temperature of about 5 deg.C). In this regime, the corre-
lations already presented in the previous chapter on ‘Natural (or, Free) convection’ may be used.
Nucleate boiling regime (i.e. excess temperature varying from about 5 deg. up to about 30 deg.C). In this re-
gime, heat transfer depends on the number of nucleation sites, rate of vapour bubble formation, etc. It is thought
that much higher heat transfer rates obtained in this regime are due to the stirring and agitating effect caused by
the bubbles on the surrounding liquid. Further, experiments show that nucleate boiling heat flux is not very
much dependent on the geometry or orientation of the heater surface. Therefore, the correlation given below is
valid for flat plates, cylinders and other geometries.
Correlation proposed by Rohsenow in 1952, is the most widely used one, for heat flux in the nucleate boiling
regime:

L g ×( r - r ) OP × LM C OP
1 3
pL ×(Ts - Tsat )
×M L V 2
W/m2
q nucleate = m L × h fg
N s Q MN C sf × h f g × Pr Ln PQ ...(11.5)

BOILING AND CONDENSATION 533


where,
qnucleate = nucleate boiling heat flux, W/m2
m L = viscosity of liquid, kg/(m.s)
h f g = enthalpy of vaporisation, J/kg
g = gravitational acceleration, m/s2
r L = density of liquid, kg/m3
r V = density of vapour, kg/m3
C pL = Specific heat of liquid, J/(kg.C)
s = surface tension of liquid–vapour interface, N/m
Ts = surface temperature of heater, deg.C
Tsat = saturation temperature of fluid, deg.C
Csf = a constant depending upon the specific surface–fluid combination
Pr L = Prandtl number of liquid
n = 1 for water, and 1.7 for all other liquids.
Subscripts L and V refer to liquid and vapour, respectively.
Since water is one of the most common fluids used, it is useful to have its surface tension and the constant
Cs f for water–surface combination, readily available.
Surface tension and latent heat of water at a few temperatures are given in Table 11.1.

TABLE 11.1 s and hfg for water

Saturation temperature (Tsat), deg.C Surface tension (s ), N/m Latent heat (hfg ), kJ/kg
0 0.0755 2500.8
20 0.0729 2453.7
40 0.0695 2406.2
60 0.0661 2357.9
80 0.0627 2308.3
100 0.0589 2256.7
150 0.0487 2113.4
200 0.0378 1939.3
250 0.0261 1714.7
300 0.0143 1406.2
350 0.0036 916.1
374 0.0 0.0

A quick estimate of surface tension of water at a given temperature can be made using the following equa-
tion:
s = 0.0743 × (1 – 0.0026 T), N/m ...(11.6)
where, T is in deg.C
Experimentally determined values of constant Csf for a few liquid–surface combinations are given in Table
11.2:
Note that Rohsenow Eq. 11.5 is applicable for clean surfaces and for relatively smooth surfaces.
To calculate the heat flux in nucleate boiling, Collier recommends the following equation which is simpler to
use as compared to Eq. 11.5:

LM F P I 0.17
FPI 1. 2
F PI OP
10 3 . 33
qnucleate = 0.000481× DT e3.33 × P c2.3
r ×
MN1.8 ×GH P JK
cr
+ 4 ×G
H P JK
cr
+ 10 × G
H P JK
cr PQ W/m2 ...(11.7)

where, DTe is the excess temperature in deg.C, P is the operating pressure in atm., Pc r is the critical pressure in
atm.
Another correlation proposed by Mostinski to determine the heat transfer coefficient in nucleate boiling is:

534 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 11.2 Csf for a few liquid–surface combinations
Fluid and surface Csf
Water–copper:
Scored surface 0.0068
Polished surface 0.0130
Water–stainless steel:
Teflon coated surface 0.0058
Ground and polished surface 0.0068
Mechanically polished surface 0.0130
Chemically etched surface 0.0130
Water–brass 0.0060
Water–nickel 0.0060
Water–platinum 0.0130
n-pentane–copper:
Lapped surface 0.0049
Polished surface 0.0154
n-pentane–chromium 0.0150
Ethyl alcohol–chromium 0.0027
Benzene-chrommium 0.101

F PI 0 . 566
h nucleate = 0.00341×(Pcr)2.3 × (DTe ) 2.33× GH P JK
cr
W/(m2C) ...(11.8)

where, P and Pc r are the operating and critical pressure (bar), respectively, DTe is the excess temperature in deg.C.
Once ‘h’ is known, heat flux is calculated as: q = h.DTe .
Another useful correlation for heat transfer coefficient in nucleate boiling is from Russian literature:

LM Fr × hf g I ×FG r IJ OP
1 1
0 .75
kL × q0 .70
= M6.9 ´ 10 × G JK H s K PP
V 30 3
-3 L
kcal/(m2hC)
h nucleate
MN Hr - rV 7
...(11.9a)
L
Qm 0 . 45
L × C 60 ×(T
pL sat + 273) 0. 37

Note the units in the above equation: r in kg/m3, hfg in kcal/kg, s in kg/m, k in kcal/(mhrC), m in kgs/m2, Cp
in kcal/(kgC) and subscripts L and V refer to liquid and vapour, respectively.
When all terms are expressed in S.I. Units, as in Eq. 11.5, above equation becomes:

Fr I ×FG r IJ
1 1
× hf g 0. 75
kL × q 0. 70
= 0.07726 G JK H s K
V 30 3
L
hnucleate
HrL - rV
×
0. 45
7
m L × C pL
60 × (T + 273) 0 . 37
W/(m2C) ...(11.9b)

sat

Advantage of Eq. 11.9b is that heat transfer coefficient is presented as a function of physical properties of the
fluid only; therefore, it can be used to calculate ‘h’ for any fluid and at any pressure, if reliable data on physical
properties are available.
Based on Eq. 11.9, following calculation formulas are recommended specifically for water in nucleate boil-
ing, in the pressure range 0.2–100 ata:
hnucleate = 3.133 × q 0.7 × P 0.15 W/(m2C) (for water...(11.10a))
and,
hnucleate = 45.054 × DT 2.33 × P 0.5 W/(m2C) (for water...(11.10b))
In the above equations q is the heat flux in W/m2, P is the pressure in bar, and DT is the excess temperature
in deg.C.
Peak (or, maximum) heat flux in nucleate pool boiling:
During the design of heat transfer equipments (e.g. boiler tubes), it is extremely important to have an idea about
the peak heat flux, so that steps can be taken to avoid a burnout.
BOILING AND CONDENSATION 535
Leinhard and Dhir (1973) give the following correlation for peak heat flux in nucleate pool boiling:

L s × g ×(r OP
1

×M
4
L - rV )
qmax = Co × h f g ×rV W/m2 ...(11.11)
MN r 2
V PQ
where, Co = 0.149 for a large horizontal surface
and, Co= 0.116 for a large horizontal cylinder.
Unlike the nucleate boiling flux, peak heat flux depends on heater geometry and orientation.
Eq. 11.11 indicates that water will have larger peak heat flux than any other common liquids, because of its
large heat of vaporisation. Also, peak heat flux is a function of operating pressure, since the pressure affects the
boiling point, which in turn, affects the heat of vaporisation and surface tension. According to experimental data,
peak heat flux initially increases sharply as the pressure is increased, reaches a maximum, then decreases to zero
at critical pressure. This trend is shown clearly for water, in Fig. 11.4.
In Fig. 11.4, on X-axis, we have the ratio of P/Pcr and on Y-axis is plotted the ratio (qpeak, p/qpeak, 1), where
qpeak, p is the peak heat flux at the operating pressure P and qpeak, 1 is the peak heat flux at one atm. pressure. At
the maximum point in the curve, we have:
P/Pcr = 0.35 and qpeak, p/qpeak, 1 = 3.2. Remembering that for water, Pc r = 225 ata, we see that q peak = 4.652 x 106
W/m2 must occur at P = 80 ata.
In addition to peak heat flux, the excess temperature at peak heat flux is also important to determine if the
surface of the heater would reach the burnout point at a given peak heat flux. Experimental values of peak heat
flux and the corresponding excess temperature are given in Table 11.3 for a few fluids at 1 atm. pressure:
Another relation for peak heat flux on horizontal cylinders, which fits experimental data very well, is
presented by Sun and Lienhard :
amax
qmax F
= 0.89 + 2.27 exp - 3. 44 × R ¢ W/m2 e j ...for 0.15 < R¢ < 3.47...(11.12a)

Variation of peak heat flux for water


4
(Qpeak at P)/(Qpeak at 1 am.)

Yi
2

0
0 0.2 0.4 0.6 0.8 1
Xi
Ratio (P/Pcr)

FIGURE 11.4 Variation of peak heat flux with pressure for water

TABLE 11.3 Peak heat flux and critical temperature for a few liquids at one atm. pressure

Liquid H2O O2 N2 H2
2 5
Qpeak –1, (W/m ) 12 x 10 150 100 30
DTcr –1(deg.C) 25–30 11 11 2

536 FUNDAMENTALS OF HEAT AND MASS TRANSFER


amax
= 0.894 W/m2 (for R¢ > 3.47...(11.12b))
qmax F
where, R¢ is a dimensionless radius defined as:

L g ×(r - r ) OP
1

R¢ = R× M L V 2

N s Q
and, qmax F is the peak heat flux on an infinite horizontal plate, given as:
1
qmax F = 0.131× r V × h f g × [s × g × ( r L - r V )] 4 W/m2 (for infinite horizontal plate...(11.13))
Another useful correlation for peak heat flux in nucleate boiling is from Russian literature:
13 1 1
0.5
kL × ( r L - r V ) 24 ×[ rV × hf g ×(Tsat + 273)] 3 ×s 24
qmax = 1.7 × 10 4 × 5 1
kcal/(m2hr) ...(11.14a)
r 12 × C 6
L pL

Note the units in the above equation r in kg/m3, h fg in kcal/kg, s in kg/m, k in kcal/(m hr C), m in kgs/m2, Cp
in kcal/(kgC) and subscripts L and V refer to liquid and vapour, respectively.
When all terms are expressed in S.I. Units, as in Eq. 11.5, above equation becomes:
13 1 1
k 0 . 5 × ( r L - r V ) 24 ×[ rV × hf g ×(Tsat + 273)] 3 ×s 24
q max = 4152 × 5 1
W/m2 ...(11.14b)
r 12 × C 6
L pL

Advantage of Eq. 11.14b is that peak heat flux is presented as a function of physical properties of the fluid
only; therefore, it can be used to calculate ‘qmax’ for any fluid and at any pressure, if reliable data on physical
properties are available.
Minimum heat flux:
This occurs at point D in Fig. 11.2; minimum heat flux represents the lower limit of heat flux in film boiling. For
a large, horizontal plate, Zuber derived the following relation (modified by Berenson in 1961) for minimum heat
flux:

L s × g ×( r - r ) OP
1

×M
4
L V
q min = 0.09 × rV × h f g W/m2 ...(11.15)
MN (r + r ) PQ
L V
2

Film boiling:
Heat transfer coefficient in stable film boiling regime on a horizontal cylinder or sphere is predicted by Bromley’s
correlation:

L g × r ×( r - r ) × h ¢ × k OP
1
3
=C ×M
4
V L V fg V
hfilm
MN m ×(T - T )× L
o
V s sat PQ W/(m2C) ...(11.16)

where,
h¢f g = h f g + 0.4 × C pV × (Ts – Tsat)
Co = 0.62 and L = D (for a horizontal cylinder)
Co = 0.67 and L = D (for a sphere)
For a very large diameter tube (diameter D) or a horizontal surface, Eq. 11.16 is valid, with the following
value for Co (Westwater and Breen, 1962):
FG 0.69 × l IJ and, L = l
Co = 0.59 +
H D K
BOILING AND CONDENSATION 537
L s OP
1

l = 2×p × M
2
and,
N g ×( r - r ) Q L V
Note that for a horizontal surface, Co = 0.59, since D ® ¥
Vapour properties in Eq. 11.16 are evaluated at the mean film temperature,
Tf = (Ts + Tsat)/2
As stated earlier, during stable film boiling, at high temperatures (> 300°C), thermal radiation effects become
significant and Bromley suggested using an overall heat transfer coefficient given by:
h = h film + 0.75 × h rad ...(11.17)
and, h rad is given by:

hrad =
e
s × e × Ts4 - Tsat
4
j W/(m C)2
...(11.18)
(Ts - Tsat )
where, s = 5.67 ´ 10 –8 W/(m2K4) (Stefan–Boltzmann constant)
and, e is the emissivity of the heated surface.
Also, remember that in Eq. 11.18, the temperatures Ts and Tsat must be in Kelvin.
Heat flux in stable film boiling is easily calculated, once the heat transfer coefficient is determined, i.e.
qfilm = h × (Ts – Tsat) W/m2. ...(11.19)
11.3.7 Simplified Correlations for Boiling with Water
Since water is one of the most commonly used fluids in practice, it is useful to have some simplified correlations
for boiling water.
Jakob and Hawkins (1957) presented following simple relations for water boiling at atmospheric pressure on
submerged surfaces:
Heat transfer coefficients at pressures other than atmospheric may be calculated using the following empiri-
cal equation:

F pI 0.4
hp = ha × GH p JK
a
...(11.20)

where, hp = heat transfer coefficient at any pressure p,


ha = heat transfer coefficient at pressure pa (= 1 atm.) from Table 11.4.

TABLE 11.4 Simplified relations for boiling heat transfer coefficient for water at one atm. pressure

Type of surface Range of validity (kW/m 2) h (W/m 2K)


Horizontal: qs < 15.8 1040 ´ (DTe )1/3
15.8 < q s < 236 5.56 ´ (DTe )3

Vertical: qs < 3.15 539 ´ (DTe )1/7


3.15 < qs < 63.1 7.95 ´ (DTe )3

Example 11.1. Water at a pressure of one atm. is boiled in a polished copper pan, 300 mm diameter. If the surface
temperature of the pan is 110°C, (a) calculate the boiling heat flux and the heat transfer coefficient. What is the evapora-
tion rate of water? (b) compare the nucleate boiling flux with the maximum heat flux (c) compare the values of heat
transfer coefficient obtained from Rohsenow’s correlation with those obtained using Collier’s, Mostinski’s and Russian
correlations.
Solution.
Data:
Ts := 110°C Tsat := 100°C d := 0.3 m Cs f := 0.013 (from table 11.2)
Properties of water at Tsat = 100°C are:
r L = 958.4 kg/m3 r V = 0.5955 kg/m3 CPL = 4220 J/(kgK) mL = 279 ´ 10 – 6 kg/(ms) Pr L = 1.75

538 FUNDAMENTALS OF HEAT AND MASS TRANSFER


h f g = 2257 ´ 10 3 J/kg s = 58.9 ´ 10 – 3 N/m n = 1 (exponent n in eqn. 11.5) (exponent ‘n’ in eqn. 11.5)
2
g = 9.81 m/s
Since DT is 10 deg.C, it is reasonable to assume that correlation for nucleate boiling regime is applicable. Then, we
have, for heat flux:

LM g ×( r - r ) OP × LM C OP
1 3
L V 2 pL × (Ts - Tsat )
qnucleate := m L × h f g× W/m2
N s Q MN C sf × h f g × Pr
n
L PQ ...(11.5)

i.e. q nucleate = 1.396 ´ 10 5 W/m2


Heat transfer coefficient:
qnucleate
We have: h := W/(m2C)
(Ts - Tsat )
i.e. h = 1.396 ´ 10 4 W/(m2C)
Evaporation rate:
F p ×d I 2
Total amount of heat supplied: Q := qnucleate × GH 4 JK W

i.e. Q = 9.869 ´ 10 3 W
Q
Therefore, evaporation rate of water: m := kg/s
hf g
i.e. m = 4.373 ´ 10 – 3 kg/s
i.e. m = 15.742 kg/h
(b) Maximum heat flux:
We have, from Eq. 11.11 for a horizontal surface:

L s × g ×(r OP
1

×M
L - rV ) 4
qmax := 0.149× h f g× r V W/m2
MN r 2
V PQ ...(11.11)

i.e. qmax = 1.259 ´ 10 6 W/m2


Thus, actual heat flux is much smaller than the critical (max) heat flux.
Let also check the actual heat flux using Collier’s correlation:
We have: Pcr := 225 atm (critical pressure for water)
P := 1 atm (operating pressure)
DTe := Ts – Tsat
We have, from Collier’s correlation:

LM F P I 0 . 17
F PI 1. 2
F PI 10
OP 3 . 33

MN1.8× GH P JK + 4×G J + 10 × G J
PQ
3.33
qnucleate := 0.000481 × DT e P cr2.3 × W/m2
cr HP Kcr HP K cr
...(11.7)

i.e. qnucleate= 8.969 ´ 10 4 W/m2


Compare this value with 1.396 ´ 105, obtained using Rohsenow’s correlation.
Let us also check the actual heat flux using Mostinski’s correlation:
FPI 0 . 566

hnucleate := 0.00341× (Pc r)2.3 × (DTe)2.33 × GH P JK


cr
W/(m2C) ...(11.8)

i.e. hnucleate = 8.739 ´ 103 W/(m2C)


and, q nucleate := hnucleate × DTe
i.e. qnucleate = 8.739 ´ 10 4 W/m2.
Again, compare this value with 1.396 ´ 105, obtained using Rohsenow’s correlation, and 8.969 ´ 104, using Collier’s
correlation.
Also, from Russian literature:
hnucleate := 45.054 DTe2.33 × P0.5 W/(m2C) (for water (11.10b))
i.e. hnucleate = 9.632 ´ 103 W/(m2C)
and, q nucleate := h nucleate × DTe
i.e. qnucleate = 9.632 ´ 104 W/m2.

BOILING AND CONDENSATION 539


Thus, it may be noted that these empirical relations can give values that differ from each other considerably.
(c) comparing the values of h from different correlations:
h = 1.396 ´ 104 W/(m2C) ...Rohsenow’s correlation
h = 8969 W/(m2C) ...Collier’s correlation
h = 8739 W/(m2C) ...Mostinski’s correlation
h = 9632 W/(m2C) ...Russian correlation.
Example 11.2. A nickel wire, 1 mm diameter and 300 mm long, is submerged in a water bath open to atmosphere. What
is the value of current flowing through the wire that will cause burnout, if the applied voltage is 10 V?
Solution.
Data:
Tsat := 100°C R := 0.0005 m L := 0.3 m V := 10 V
Properties of water at Tsat = 100°C are:
r L := 958.4 kg/m3 r V := 0.5955 kg/m3 C p L := 4220 J/(kgK) mL := 279 ´ 10 – 6 kg/(ms) Pr L := 1.75
h f g := 2257 ´ 103 J/kg s := 58 ´ 10 – 3 N/m g := 9.81 m/s2
This is the case of a horizontal cylinder. So, let us use Eq. 11.12.
First, calculate the factor R’:

L g ×(r - r ) OP
1

R¢ := R× M L V 2

N s Q
i.e. R¢ = 0.2 (dimensionless radius < 3.47)
Therefore,
qmax
qmax F
e j
= 0.89 + 2.27×exp - 3. 44 R ¢ W/m2 ...for 0.15 < R¢ < 3.47 ...(11.12a)

and, q max F is the peak heat flux on an infinite horizontal plate, given as:
1
q max F := 0.131× r V × h f g s × g ×(r L - rV ) 4 W/m2 ...for infinite horizontal plate ...(11.13)
6 2
i.e. qmax F = 1.107 ´ 10 W/m (for infinite horizontal plate.)
From Eq. 11.12a we get:
qmax
= 1.378
qmax F
i.e. qmax := 1.378 × q max F
i.e. qmax = 1.525 ´ 106 W/m2 ...maximum heat flux for horizontal cylinder
This is the value of burnout flux.
If V is the voltage, I the current through the wire, we have:
V ×I
qmax =
2×p × R × L
qmax × 2 × p × R × L
i.e. I := A ...current through the wire
V
i.e. I = 143.724 A ...current through the wire.
Example 11.3. A horizontal, metal-clad heating element, 10 mm diameter and of surface emissivity 0.85, is submerged in
a water bath. Surface temperature of the heating element is 300°C. If the water is at atmospheric pressure, calculate the
power dissipation per unit length of the heater.
Solution.
Data:
Tsat := 100°C D := 0.01 m L := 1 m Ts := 300°C e := 0.85 rL := 958.4 kg/m3
3 2 –8 2 4
h f g := 2257 ´ 10 J/kg g := 9.81 m/s s := 5.67 ´ 10 W/(m K )
Since the excess temperature is (300 – 100) = 200°C, it is film boiling region. We need properties of vapour at the
mean film temperature of (300 + 100)/2 = 200°C.
Properies of vapour at 200°C:
kV := 0.0375 W/(mK) rV := 7.85 kg/m3 mV := 15.7 ´ 10 – 6 kg/(ms) CpV := 2910 J/(kgK)
h¢f g := h f g + 0.4 CpV × (Ts – Tsat ) i.e. h¢f g := 2.49 ´ 106 J/kg

540 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Then, for a horizontal cylinder, we have:

L g × r ×(r - r ) × h¢ × k OP
1

= C ×M
3 4
V L V fg V
hfilm W/(m2C)
MN m ×(T - T ) × L PQ
o
V s sat
...(11.16)

with Co = 0.62 and L = D...for a horizontal cylinder

L g × r × (r - r ) × h ¢ × k OP
1

:= 0.62× M
3 4
V L V fg V
hfilm W/(m2C)
i.e.
MN m ×(T - T ) × D
V s sat PQ
i.e. hfilm = 461.161 W/(m2C) (film boiling heat transfer coefficient)
Radiative heat transfer coefficient is given by:
s × e × [(Ts + 273) 4 - (Tsat + 273) 4 ]
hrad := W/(m2C) ...(11.18)
(Ts - Tsat )
i.e. h rad = 21.313 W/(m 2C)
Therefore, the ‘total’ heat transfer coefficient is given by:
h := h film + 0.75 × h rad ...(11.17)
i.e. h = 477.145 W/(m2C)
Therefore, power dissipation per unit length of heater:
Q := h × (p × D × L) × (Ts – Tsat) W/m
i.e. Q = 2.998 ´ 103 W/m.
Example 11.4. A large, horizontal plate is kept immersed in a water bath boiling at 1 atm, 100°C. Surface temperature of
the plate is 260°C. Calculate the heat transfer coefficient and the heat flux. Assume the emissivity of the surface as 0.9.
Solution.
Data:
Tsat := 100°C Ts := 260°C e := 0.9 rL := 958.4 kg/m3 h f g := 2257 ´ 10 3 J/kg
g := 9.81 m/s2 s := 58.9 ´ 10 – 3 N/m
Since the excess temperature is (260–100) = 160°C, it is film boiling region. We need properties of vapour at the
mean film temperature of (260 + 100)/2 = 180°C.
Properties of vapour at 180°C:
kV := 0.03268 W/(mK) rV := 5.16 kg/m3 m V := 15.1 ´ 10 – 6 kg/(ms) CpV := 2709 J/(kgK)
h ¢f g := h f g + 0.4 × CpV × (Ts – Tsat) i.e. h ¢f g := 2.43 ´ 106 J/kg
Then, we apply Eq. 11.16:

L g × r × (r - r ) × h ¢ × k OP
1

= C ×M
3 4
V L V fg V
hfilm W/(m2C)
MN m ×(T - T ) × L
o
V s sat PQ ...(11.16)

where, for a horizontal surface, we have:


L = l and,
FG 0.69 × l IJ
H
Co = 0. 59 +
D K where, D ® ¥

i.e. Co := 0.59

L s OP
1

l := 2 × p × M
2
and,
MN g ×( r - r ) PQ
L V

i.e. l = 0.016 m
Then, film boiling heat transfer coefficient:

L g × r × (r - r ) × h ¢ × k OP
1

:= C × M
3 4
V L V fg V
h film W/(m2C).
MN m ×(T - T )× l
o
V s sat PQ
i.e. h film = 337.783 W/(m2C)

BOILING AND CONDENSATION 541


And, radiative heat transfer coefficient:
s := 5.67 ´ 10 – 8 W/(m 2K4) (Stefan–Boltzmann constant)
s × e × [(Ts + 273) - (Tsat + 273) ]
4 4
h rad := W/(m2C) ...(11.18)
(Ts - Tsat )
i.e. h rad = 19.567 W/(m2C)
Note: Use absolute temperatures in Eq. 11.18 for radiative heat transfer.
Therefore, the ‘total’ heat transfer coefficient is given by:
h := h film + 0.75 × h rad ...(11.17)
i.e. h = 352.458 W/(m2C)
and, the heat flux:
q := h× (Ts – Tsat ) W/m2
i.e. q = 5.639 ´ 104 W/m2.
Example 11.5. Water is boiling at 8 atm. on the surface of a horizontal tube, whose wall temperature is maintained at
8°C above the boiling point of water. Calculate the nucleate boiling heat transfer coefficient
(b) What is the change in the value of heat transfer coefficient when (i) temperature difference is increased to 16°C at the
pressure of 8 atm., and (ii) pressure is raised to 16 atm. with DTe = 8°C.
Solution.
Data:
DTe := 8°C Pa := 1 atm P := 8 atm
Now, we use the following relation (assuming q > 15.8 kW/m2), from Table 11.4:
FPI 0.4

hp = ha × GH P JK
a

where, ha := 5.56 × (DT e ) 3 W/(m2K) (at atmospheric pressure)


Therefore, h a = 2.847 ´ 10 3 W/(m2K) (at atmospheric pressure)
and,
FPI 0.4

hp := ha × GH P JK
a

i.e. hp = 6.54 ´ 103 W/(m2K) (at 8 atm. Pressure)


i.e. hp = 6.54 kW/(m2K) (at 8 atm. pressure.)
(i) Now, when the DT is increased to 16°C, with the pressure remaining at 8 atm.:
DTe =16
F PI 0. 4

h p := 5.56 DTe3× GH P JK
a

i.e. hp = 5.232 ´ 10 4 W/(m2K)


i.e. hp = 52.32 kW/(m2K)
i.e. heat transfer coefficient increases by about 8 times as compared to the earlier value when DT was 8°C.
(ii) When the pressure is increased to 16 atm., with the DT remaining at 8°C.:
DTe := 8°C
P := 16 atm.
Again, we have:
FPI 0.4

hp := 5.56 × DTe × GH P JK
a

i.e. hp = 8.63 ´ 10 W/(m2K)3

i.e. hp = 8.63 kW/(m2K)


i.e. heat transfer coefficient increases by about 32% as compared to the original value at a pressure of 8 atm.

11.3.8 Flow Boiling


Flow boiling or boiling in forced convection, is important in the design of boiling nuclear reactors, in spacecrafts
and space power systems.

542 FUNDAMENTALS OF HEAT AND MASS TRANSFER


In flow boiling, a fluid is forced to move over a heated surface qboiling
while the phase change occurs. Therefore, combined effects of natu-
ral/forced convection and pool boiling come into play.
Flow boiling is classified as:
(i) External flow boiling, and
(ii) Internal flow boiling.
High velocity
(i) External flow boiling In external flow boiling, flow occurs over
the surface of a plate or cylinder; there are the flow regimes similar
to that in pool boiling, but due to the effect of flow velocity, both
the nucleate boiling heat flux and the critical heat flux get en-
hanced. See Fig. 11.5. For water in external flow boiling, critical Low velocity
heat flux value as high as 35 MW/m2 has been obtained (as com- Free convection
pared to the value of 1.3 MW/m2 in pool boiling at one atm.).
For cross flow over a cylinder of diameter D, Lienhard and DTe
Eichhorn have given following correlations, depending upon (deg.C)
whether the fluid velocity is ‘low’ or ‘high’.
FIGURE 11.5 Effect of flow velocity in
Criterion to determine if the velocity is low or high is:
external flow boiling
LMF 0.275 I F r I 1
OP
MMGH p JK ×GH r JK
qmax 2

r V × hf g × V
> L
+1 PP (low velocity)
N V
Q
and,
LMF 0.275 I F r I 1
OP
MMGH p JK ×GH r JK
qmax 2

r V × hf g × V
< L
+1 PP (high velocity)
N V
Q
Correlation for low velocity:

qmax
LM F I
1 OP
1 4
MM GH JK PP
3
= × 1+ ...(11.21)
r V × hf g × V p WeD
N Q
Correlation for high velocity:

FG r IJ FG r IJ
3 1
L 4 L 2
qmax
=
Hr K V
+
Hr K V
...(11.22)
r V × hf g × V 169 ×p 1
3
19.2 × p ×We D
Here, V is the fluid velocity and WeD is the Weber number, defined as the ratio of inertia forces to surface
tension forces, i.e.

rV V 2 × D
WeD = ...(11.23)
s
(ii) Internal flow boiling Internal forced convection boiling refers to flow inside a tube. This is more complicated
since, now, there is no free surface for the vapour to escape and results in two phase flow inside the tube. There
are different flow regimes occurring inside the tube depending upon the ‘quality’ of the fluid. (‘Quality’ is de-
fined as the ratio of mass of vapour to the total mass of fluid at a given location). This is illustrated in Fig. 11.6,
which also shows a qualitative graph of variation of heat transfer coefficient with local quality.
Consider a fluid, at a temperature below its boiling point, entering a vertical, heated tube. Progressive va-
porisation occurs along the length of the tube and the ‘quality’ increases. Up to a short distance from the inlet,
heat transfer coefficient for the single phase fluid may be predicted using the Dittus–Boelter equation.

BOILING AND CONDENSATION 543


(a) Bubble flow (b) Slug flow (c) Annular flow (d) Mist flow

Forced convection (liq) Annular mist Forced


h transition
2 convection (vap)
(W/(m C)
a, b c d
Annular Mist flow
flow

hc hc

0 Quality (%) 100

FIGURE 11.6 Flow regimes and heat transfer coefficient in forced convection flow in a vertical tube

Bubble–flow regime Soon, the bulk temperature reaches the saturation point, and bubbles are formed at the
nucleation sites on the wall and are carried into the main stream, as in nucleate boiling. This is known as the
‘bubble flow regime’ (see Fig. 11.6 (a)) and the heat transfer coefficient increases. Heat transfer coefficient in this
range can be predicted by superimposing the liquid–forced convection and nucleate pool boiling equations.
Slug–flow regime Further along the distance, vapour fraction increases and individual bubbles agglomerate
and slugs of vapour are formed. This regime is known as ‘slug flow regime’. See Fig. 11.6 (b). Fluid velocity
increases and since the slugs of vapour are compressible, flow oscillations may occur. Mass fraction of vapour in
this regime is around 1 %, but volume fraction of vapour may be even up to 50 %. In this regime also, heat
transfer coefficient may be calculated by superimposing the liquid–forced convection and nucleate pool boiling
equations. Heat transfer coefficient increases because of increased velocity.
Annular–flow regime As the fluid progresses further up the tube, quality increases due to further addition of
heat and vapour forms the core and a film of liquid flows on the inner wall surface. Vapour core travels at a
higher velocity than the liquid and vapours are formed primarily at the liquid–vapour interface and not at the
wall surface. Quality in this flow regime may be up to 25 %. See Fig. 11.6 (c).
Transition–flow regime Now, as the quality increases, there is a sudden drop in the value of heat transfer coef-
ficient. Heat flux at this point is known as ‘critical heat flux’. This is the point of dryout. This sudden drop
happens since the liquid film at the wall is now replaced by a vapour film, which has a poor thermal conductiv-
ity. There may be sharp increase in the wall temperature and even burnout may occur.
Mist–flow regime Now, the tube is fully occupied by the vapour, which may contain droplets of liquid. This is
known as mist–flow regime. See Fig. 11.6 (d). Heat transfer is from the wall to the vapour directly, and then from
the vapour, heat is transferred to the droplets of liquid contained in the vapour.
From the annular–flow regime onwards, prediction of heat transfer coefficient is a little difficult and uncer-
tain due to problems of two phase flow.
Correlations to find out heat transfer coefficient in nucleate flow boiling as well as in two–phase flow boiling
are presented below:
Correlations for nucleate flow boiling:
Rosenhow and Griffith (1955) have suggested that total heat flux be calculated by adding the nucleate pool
boiling flux (from Eq. 11.5) and the forced convection effect (from Dittus–Boelter equation with the coefficient
0.023 replaced by 0.019),

544 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e.
qtotal = qnucleate + qforced convection ...(11.24)
For forced convection flow inside vertical tubes, following correlation is recommended:
p

h = 2.54 × (DTe)3 × e 1. 551 W/(m2K) ...(11.25)


where, DTe = Ts – Tsat , and p = pressure in mega pascals.
Above equation is valid for the pressure range of 5 to 170 bar.
For horizontal tubes, McAdams et.al. suggest following relation for low pressure boiling water:
q = 2.253 × (DTe)3 .96 W/m2 ...for 0.2 < P < 0.7 MPa ...(11.25a)
For higher pressures, Levy recommends:
4
q = 283.2 × P 3 × (DTe ) 3 W/m2 ...for 0.7 < P < 14 MPa ...(11.25b)
Here, the pressure P is in mega-Pascals.
Correlations for two-phase, flow boiling:
Beyond the saturated nucleate boiling, there is annular–flow region, and Chen’s correlation (1966) has been
widely used for two-phase heat transfer calculations.
Here,
h = hc + h b
where, hc is the contribution due to annular region and hb is the contribution due to nucleate boiling region. We
have:

hc = 0.023 ×
LM G ×(1 - x)× D OP 0. 8
× Pr L ×
0. 4 kL
×F ...(11.26)
N m Q L D
and,

F 0 .79 0.45 0 . 49 I
hb = 0.00122 × GG kL × C pL × r L
JJ × DT 0.24
DPsat0.75× S ...(11.27)
Hs K
0. 5 0 . 29 0. 24 0 . 24 e
×m L ×h f g × r V

DPsat = change in vapour pressure corresponding to a temperature change of DTe.


(Ts - Tsat ) × hf g
D Psat =
Tsat ×( vv - vL )
‘v’ is the specific volume.
S.I. Units are used throughout.
Parameter F is calculated from:
1
F = 1.0 when < 0.1
Xtt

F 1 + 0.213I 0 .736
and, F = 2.35 × GH X tt
JK when
1
Xtt
> 0.1

F x I ×F r I ×F m I
0 .9 0. 5 0 .1

H 1 - x JK GH r JK GH m JK
= G
1 L V
where,
Xtt V L
Parameter S is given by:
S = (1 + 0.12 × ReTP1.14) –1 for ReTP < 32.5
S = (1 + 0.42 × ReTP0.78)–1 for 32.5 < ReTP < 70
S = 0.1 for Re TP > 70
where, Reynolds number ReTP is defined as:

BOILING AND CONDENSATION 545


G ×(1 - x ) × D 1. 25 - 4
×F
ReTP = ×10
mL
Chen’s correlation has been tested for several systems (including water) for pressures ranging from 0.5 to 35
atm and quality ‘x’ ranging from 1 to 71%.
More recent correlation For flow boiling, recent correlation is due to Klimenko (1988). This is for a liquid boil-
ing at a pressure P, in a pipe with a wall thermal conductivity of kw and cross-sectional area Ac. This correlation
is valid for both nucleate boiling and annular film boiling up to dry out.
First, we have to determine whether it is nucleate flow boiling regime or annular film boiling regime. This is
done by evaluating parameter f :

L Fr IO F r I
1
m × hf g
× M1 + x × G - 1J P × G
K PQ H r JK
L V 3

MN H r
F= ...(11.28)
q × Ac V L

Quality ‘x’ is defined as:


mV mV
X= = ...(11.29)
m ( mL + mV )
where, mV is the mass of vapour and m L is the mass of liquid, and m = (mV + mL)
Nucleate flow boiling (f < 1.6 ´ 104):
-1
Fk I 0.15
Nu = 7.4 ×10 – 3 (q¢ )0.6 × (P¢) 0.5 × Pr L3 × GH k JK
w
L
...(11.30)

where,
hb × Lc
Nu =
k

L s OP
1

L = M
2
c
N g ×(r - r ) Q
L V

q × Lc
q¢ =
h f g × r V × ×a L

P × Lc
P¢ =
s
F k I
aL = GH r ×C JK
p
L
Annular film boiling (f > 1.6 ´ 104):
1
F r I ×F k I
0.2 0 .09
Nu = 8.7 × 10 –2 × Re0.6 × Pr L6 × GH r JK GH k JK
V
L
w
L
...(11.31)

where,
r L ×V × Lc
Re =
mL

m LM F I OP
and, V=
Ac × r L MN
r
rV GH
× 1 + x× L - 1 JK PQ
All properties are evaluated at temperature Tsat.

546 FUNDAMENTALS OF HEAT AND MASS TRANSFER


The actual, effective heat transfer coefficient due to boiling and single phase forced convection is obtained
as:
1
e
h = hb3 + hc3 3 j ...(11.32)
Example 11.6. Water at 8 atm. flows inside a vertical tube of 2.5 cm diameter under flow boiling conditions. Tube wall
temperature is maintained at 8°C above the saturation temperature. Determine the heat transfer for one metre length of
tube.
Solution.
Data:
D := 0.025 m L := 1 m DTe := 8°C P := 8 atm i.e. P := 8.0 ´ 0.10132 MPa i.e. P := 0.811 MPa
We apply Eq. 11.25
P

h := 2.54 × (DTe)3 × e 1. 551 W/(m2K) ...(11.25)


where, P is in mega Pascals
Therefore, h = 2.193 ´ 10 3 W/(m2 K)
Then, heat transfer for 1 m length of tube:
Q := h × (p × D × L)× DTe W/m
i.e. Q := 1.378 ´ 10 3 W/m.
Example 11.7. A 50 mm diameter vertical evaporator tube (kw = 20 W/(mK)) carries 1 kg/s of steam at 14.55 bar at a
quality x = 0.2. The tube is subjected to a uniform heat flux of 106 W/m2. Identify the regime of flow boiling and
calculate the convective heat transfer coefficient and surface temperature of the tube.
(b) when the quality reaches 0.8, what is the boiling regime and how much is the boiling heat transfer coefficient?
Solution.
Data:
D := 0.05 m P := 14.55 ´ 105 N/m2 Tsat := 470 K(at 14.55 bar) hf g := 1951 ´ 103 J/kg (at 14.55 bar)
m := 1 kg/s x := 0.2 kw := 20 W/(mK) rL := 868.056 kg/m3 rV := 7.353 kg/m3 kL := 0.667 W/(mK)
CpL := 4480 J/(kgK) mL := 136 ´ 10 – 6 N.s/m2 Pr L := 0.92 s := 0.0385 N/m (surface tension)
q := 106 W/m2 g := 9.81 m/s2
` Cross-sectional area of tube:
p ×D 2
Ac :=
4
i.e. Ac = 1.963 ´ 10 – 3 m2 (cross-sectional area of tube)
First, find out the parameter F to determine if the flow boiling regime is nucleate or annular-film:
We have:

m × hf g LM Fr I OP F r I
1

GH r JK PQ GH r JK
3
L V
F := × 1 + x× -1 ×
q × Ac MN V L
...(11.28)

i.e. F = 4.944 ´ 10 3
Since F < 1.6 ´ 104, it is nucleate flow boiling regime.
Heat transfer coefficient:
We use Klimenko’s correlation to determine the boiling heat transfer coefficient h b .
From Eq. 11.30:
1
Fk I 0 . 15

GH k JK
- w
Nu = 7.4 × 10 – – 3 × (q ¢)0.6 × (P¢)0.5 × Pr L
3 × ...(11.30)
L

where,
hb × Lc
Nu =
kL

LM s OP
1
2
Lc := Lc = 2.135 ´ 10 – 3
MN g ×( r - r ) PQ
i.e.
L V

BOILING AND CONDENSATION 547


F k I
a L := GH r ×C JK
L
L

pL
i.e. aL = 1.715 ´ 10 – 7 m2/s (thermal diffusivity)

q × Lc
q ¢ := i.e. q¢ = 867.855
h f g × r V × ×a L

P × Lc
P ¢ := i.e. P = 8.07 ´ 10 4
s
Therefore,
-1
Fk I 0 . 15

Nu := 7.4 ´ 10 – 3 × (q¢)0.6× (P¢)0.5 × Pr L


3 × GH k JK
w

L
...(11.30)

i.e. Nu = 208.613 (Nusselt number)


Nu × kL
and, hb :=
Lc
i.e. hb = 6.516 ´ 10 4 W/(m2K) (boiling heat transfer coefficient)
Single-phase forced convection heat transfer coefficient hc:
We use Dittus–Boelter equation, namely,
Nu = 0.023 × Re L0.8 × Pr L0.4

Now, ReL :=
LM m×(1 - x) OP × D
N A Qm c L

i.e. ReL = 1.498 ´ 10 5 (Reynolds number for liquid flow)


Then, Nu := 0.023 Re L0.8 Pr L0.4
i.e. Nu = 307.352 (Nusselt number)
Nu × k L
Therefore, hc :=
D
i.e. hc = 4.1 ´ 10 3 W/(m2K) (single-phase convection heat transfer coefficient)
Therefore, total or effective heat transfer coefficient:
Total or effective heat transfer coefficient is given by Eq. 11.32:

d i
1

h := hb3 + hc3
3
...(11.32)
4 2
i.e. h = 6.517 ´ 10 W/(m K) (total heat transfer coefficient)
Tube surface temperature:
We have: q = h×DTe (heat flux)
q
Therefore, DTe := °C
h
i.e. DTe = 15.345°C
and, Ts := Tsat + DTe
i.e. Ts = 485.345 K
i.e. Ts = 212.345°C (tube surface temperature.)
(b) When quality, x = 0.8:
x := 0.8 (quality)
We have:

m × hf g L Fr
× M1 + x × G
IO F r I
- 1J P × G
1

K PQ H r JK
3
L V
F :=
q × Ac MN H r V L
...(11.28)

i.e. F = 1.917 ´ 10 4
Since F > 1.6 ´ 104, it is annular-film flow boiling regime.
In this regime, Klimenko’s correlation for boiling heat transfer coefficient is:

548 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F r I ×F k I 1 0.2 0 . 09

Nu = 8.7 ´ 10 –2 ×Re0.6× Pr L6 ×GH r JK GH k JK V

L
w

L
...(11.31)

m L F r - 1I OP i.e. V = 55.528 m/s


× M1 + x × G
MN H r JK PQ
L
where, V :=
Ac × r L V

r L × V × Lc
and, Re := i.e. Re = 7.568 ´ 105
mL
Therefore,
1
F r I ×F k I 0.2 0 . 09

Nu := 8.7 ´ 10 –2 × Re0.6 × Pr L6 × GH r JK GH k JK
V

L
w

i.e. Nu = 151.142 (Nusselt number)


Nu × kL
and, hb :=
Lc
i.e. hb = 4.721 ´ 104 W/(m2K) ...annular–film boiling heat transfer coefficient
Compare this value with hb = 6.516 ´ 104, obtained earlier for nucleate flow boiling. It is as it should be, since in
annular film flow boiling, the heat transfer coefficient is less than that in the nucleate flow boiling.
Example 11.8. In Example 11.7, if the tube surface is maintained at a constant temperature of 227°C, calculate the total
heat transfer coefficient and surface heat flux at the point where the quality is 0.2. Rest of the data are the same as in
Example 11.7.
Solution.
Data:
D := 0.05 m P := 14.55 × 10 5 N/m2 Tsat := 470 K (at 14.55 bar) hf g := 1951 ´ 103 J/kg Ts := 500 K
\ DTe := Ts – Tsat Ps := 26.4 bar, corresponding to 500 K i.e. DPsat := (26.4 – 14.55) ´ 10 5 N/m2 m := 1 kg/s
x := 0.2 (quality) rL := 868.056 kg/m3 rV := 7.353 kg/m3 kL := 0.667 W/(mK) CpL := 4480 J/(kgK)
mL := 136 ´ 10 – 6 Ns/m2 mV := 15.54 ´ 10 –6 Ns/m2 Pr L := 0.92 s := 0.0385 N/m g := 9.81 m/s2
Cross-sectional area of tube:
p ×D 2
Ac :=
4
i.e. Ac = 1.963 ´ 10 –3 m2
Now, let us use Chen’s correlation.
We have:
h = hc + hb
where, hc is the contribution of single-phase convection and hb is the contribution of boiling region.

hc = 0 .023 ×
LM G ×(1 - x) × D OP 0.8

× Pr L0 . 4 ×
kL
×F ...(11.26)
N m Q L D
and,
F k 0 . 79
× C 0pL. 45 × r 0L . 49 I × DT
hb = 0.00122 × GH s L
0. 5 0 . 24
× h fg × m 0L. 29 r V0 . 24
JK e
0 . 24
× DP 0sat. 75 ×S ...(11.27)

First, let us calculate 1/Xtt , so that factor F can be calculated:

F x I × F r I ×F m I
0. 9 0. 5 0.1
1
X tt
= GH 1 - x JK GH r JK GH m JK L

V
V

1
i.e. = 2.512
X tt
This is greater than 0.1. Therefore,
F1 I 0 . 736

F := 2.35 × GH X tt
+ 0. 213 JK when
1
X tt
> 0.1

BOILING AND CONDENSATION 549


i.e. F = 2.091
Also, G, the mass velocity is:
m
G :=
Ac
i.e. G = 509.296 kg/(sm2)
Then, we have:

hc := 0.023 ×
LM G ×(1 - x) × D OP 0 .8

× Pr 0L. 4 ×
kL
×F ...(11.26)
N m Q L D
i.e. hc = 8.573 ´ 103 W/(m2K) (convective heat transfer coefficient)
and, to calculate hb , we need to calculate the factor S, after finding out ReTP ,:
G ×(1 - x) × D 1.25
ReTP := ×F ´ 10 – 4
mL
i.e. ReTP = 37.665 9Two-phase Reynolds number)
Then, we have:
0.78 –1
S := (1 + 0.42 × Re TP ) for 32.5 < Re TP < 70
i.e. S = 0.123
Therefore,
F k 0L. 79 × C 0pL. 45 × r L0 . 49 I × DT
hb = 0.00122× GH s 0 .5
× m 0L . 29 × h f g 0 . 24 × r V0 . 24
JK e
0 . 24
× DP sat
0 . 75
×S ...(11.27)

i.e. hb = 1.385 ´ 10 4 W/(m 2 K) (boiling heat transfer coefficient)


Then, total heat transfer coefficient h:
h := hc + hb
i.e. h = 2.243 ´ 104 W/(m2K) (total heat transfer coefficient)
And, surface heat flux at that point where x = 0.2:
q := h ×DTe
q = 6.728 ´ 105 W/m2 (local heat flux.)

11.4 Condensation Heat Transfer


11.4.1 Introduction
Condensation heat transfer has important practical applications, e.g. in thermal power plants, low pressure ex-
haust steam from the steam turbine is condensed on the outside of water cooled tubes of a condenser; in vapour
compression refrigeration and air-conditioning systems, the refrigerant is condensed inside tubes of the con-
denser, in the condenser-reboiler of cryogenic distillation columns, etc.
Whenever a saturated vapour at a temperature Tsat is brought in contact with a surface maintained at tem-
perature Ts such that Ts is less than Tsat, vapours condense on the surface. Thus, in a way, condensation is the
‘reverse’ of boiling process. While condensing, naturally, the vapours will release the latent heat of vaporisation.
The vapours may condense on the surface in one of the two modes: ‘film-wise condensation’ or ‘drop-wise
condensation’.
In film-wise condensation, say, on a vertical surface, vapours condense on the surface and drip down
forming a continuous liquid film on the surface. Thickness of the condensate film increases as it travels down
towards the lower (or trailing) end of the plate. During the condensation process, latent heat of vaporisation is
released by the vapours. For further condensation to occur, the released latent heat has to be conducted through
this liquid film to the cooled surface at temperature Ts . However, the liquid film offers resistance to the flow of
heat and this resistance increases as the thickness of the film grows. Film-wise condensation occurs on surfaces
which tend to get ‘wetted’.
In drop-wise condensation, the vapours condense on the surface on drops, which drip down the surface. A
continuous film of liquid is not formed on the surface. Thus, more of the base area at temperature Ts is always
exposed to the vapours. Therefore, heat transfer rate is higher (up to ten times) in drop-wise condensation as
compared to the value in film-wise condensation. Generally, drop-wise condensation occurs on smooth surfaces
which do not get ‘wetted’.

550 FUNDAMENTALS OF HEAT AND MASS TRANSFER


While drop-wise condensation would appear to be the preferred y
mode, in practice, it is difficult to maintain this mode of condensation
since, with time, all surfaces tend to get wetted. x g
Attempts to achieve drop-wise condensation have been made ei-
Velocity profile
ther by coating the surface with some suitable material or by adding
some additives to the vapours; but, commercially, these techniques
have not yet become viable. Liquid-vapour
Interface
11.4.2 Film Condensation and Flow Regimes Ts
Consider film condensation of a vapour at saturation temperature Tsat Temp. profile
on the surface of a cooled vertical plate, maintained at a temperature
Ts (< Tsat.). See Fig. 11.7.
Vapour condenses on the top of the plate and flows down as a
film. Thickness of the film (d ) is zero at the top of the plate (i.e. at x =
0 in the coordinate system shown) and increases as we travel down
the plate (i.e. as x increases) due to additional condensation of va- Liquid Vapour, Tsat
pour. Initially, the liquid film flow is laminar; after some distance it
will become wavy and later, it may even turn turbulent. These differ-
ent flow regimes are identified according to a ‘film Reynolds
FIGURE 11.7 Film condensation on a
number’, defined as follows:
vertical plate
Dh × r L ×VL
Ref =
mL
4 × Ac × r L × VL 4 × r L ×VL ×d 4×m
= = = ...(11.32)
P × mL mL P × mL
where,
Dh = 4Ac/P = 4.d = hydraulic diameter of condensate
flow, m
P = wetted perimeter of condensate, m
Ac = P. d = area of cross section of flow at the lowest part of flow, m2
r L = density of liquid, kg/m3
m L = viscosity of liquid, kg/ms
V L = average velocity of condensate at the lowest part of flow, m/s
rLAcV L = m = mass flow rate of condensate at the lowest part of flow, kg/s.
For the common geometries of a vertical plate, vertical cylinder and a horizontal cylinder, hydraulic diam-
eter Dh is equal to 4 times the thickness of the condensate, d , at the location where the hydraulic diameter is to be
evaluated; this is clearly shown as follows:
For a vertical plate Let B be the breadth of the plate, perpendicular to paper. At any section along the vertical
height, let the thickness of the liquid film be d . Then,
Ac = B. d = area of cross section
P = B = wetted perimeter, and
Dh = 4.Ac/P = 4.d.
For a vertical cylinder Let D be the diameter of the cylinder. At any section along the vertical height, let the
thickness of the liquid film be d. Then,
Ac = p.D.d = area of cross section
P = p.D = wetted perimeter, and
Dh = 4.Ac/P = 4.d.
For a horizontal cylinder Let D be the diameter of the cylinder, and L the length. Along the length of the cylin-
der, let the thickness of the liquid film be d . Then,
Ac =2.L.d = area of cross section
P = 2.L = wetted perimeter, and
Dh = 4.Ac/P = 4.d.

BOILING AND CONDENSATION 551


Again, considering Eq. 11.32, for a vertical plate, wetted perimeter, P = B, the breadth; therefore, (m/P) is the
mass flow rate per unit breadth. If we denote (m/P) by m‘, we can write for the vertical plate:
4× m¢
Re f = ...(11.32a)
mL
Another point to be noted is regarding the latent heat of vaporisatiion (hfg ) released during condensation:
Vapour at a temperature of Tsat comes in contact with the plate at a temperature of Ts (< Tsat) and condenses.
However, the condensed liquid is, invariably, further sub-cooled to a temperature somewhere in between Ts and
Tsat, thus releasing some more heat. Rohsenow (1956) suggested that this subcooling of the liquid can be taken
into account by replacing h fg by a ‘modified latent heat of vaporisation’, h¢fg , defined as:
h¢f g = h fg + 0.68×C p L× (Tsat – Ts) ...(11.33)
where, C pL is the specific heat of liquid at the average film temperature.
Similarly, if a superheated vapour at a temperature, Tv , enters a condenser and condenses, the superheated
vapour has to be cooled to Tsat first, and then condensed at Tsat, and then sub-cooled to some temperature be-
tween Ts and Tsat. Then, modified latent heat of vaporisation is:
h¢f g = h fg + 0.68×C pL × (Tsat – Ts) + CpV × (Tv – Tsat) ...(11.34)
where, CpV is the specific heat of vapour at the average temperature of (TV + Tsat)/2.
Then, rate of heat transfer in condensation becomes:
Qconden = h × A× (Tsat – Ts) = m× h ¢f g ...(11.35)
where, A is the surface area on which condensation occurs.
Then, from Eq. 11.35 and 11.32, we can write:
4 ×Qconden 4 × A × h ×(Tsat - Ts )
Ref = = ...(11.36)
P × m L × h¢f g P × m L × h¢f g
When either Qconden or h is known, it is convenient to use Eq. 11.36 to determine Ref .
Now, different flow regimes are identified according to the value of Re f as follows:
Ref £ 30 (Liquid film is smooth and wave-free, i.e. fully laminar.)
450 < Ref < 1800 (Liquid film has ripples or waves and the flow is wavy—laminar.)
Ref > 1800 (Liquid film is fully turbulent.)
Heat transfer correlations vary depending upon the flow regime.
11.4.3 Nusselt’s Theory for Laminar Film Condensation on Vertical Plates
Nusselt developed his theory for laminar film condensation on vertical plates analytically in 1916.
Consider a vertical plate maintained at a temperature Ts and exposed to a saturated vapour at a temperature
of Tsat. (Ts < Tsat). See Fig. 11.8. Let the height of the plate be L and the breadth, ‘b’. Coordinate system is chosen
such that x-coordinate is to the downward direction, i.e. in the direction of flow of condensate and y-coordinate
is towards the right, as shown. Condensation occurs on the plate and the condensate moves down from top to
bottom. Thickness of condensate is zero at the top (i.e. at x = 0) and increases in the flow direction, due to
additional condensation of vapour. Since the liquid film offers resistance to the flow of heat from the vapour to
the cold surface, this also means that resistance to heat transfer is minimum at the top of the plate and the
resistance increases as one moves down in the flow direction.
Nusselt made the following simplifying assumptions in his analysis:
(i) Flow of liquid film is laminar
(ii) Inertia force in the film is negligible (i.e. negligible acceleration of the liquid in the film) compared to
viscosity and weight
(iii) Heat flow is mainly by conduction through the liquid film; convection in liquid as well as in vapour is
neglected.
(iv) Temperature is Ts at the liquid–plate interface and Tsat at the liquid–vapour interface and the temperature
gradient between them is linear
(v) Velocity of vapour is low, i.e. there is no viscous shear force at the liquid–vapour interface
(vi) Properties of the liquid are constant.

552 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Volume element
d-y

y dx

x g F2
F1
F3
x
d
F1 = weight = rLg(dy)(b.dx)
dx Velocity profile F2 = shear force = m(du/dy)(b.dx)
y
F3 = Buoyancy force = rVg(dy)(b.dx)
Temperature profile

Ts
Liquid vapour Interface

Liquid Vapour, Tsat

FIGURE 11.8 Film condensation on a vertical plate—Nusselt’s analysis

Velocity profile:
Considering a volume element as shown, apply Newton’s second law to get the velocity profile:
S F x = m × ax = 0
since the acceleration of the liquid film is negligible... by assumption.
So, making the force balance:
Fdown = Fup
i.e. Weight = Viscous shear force + Buoyancy force
du
rL × g×(d – y)×(b×dx) = mL × × (b× dx) + rV × g×(d – y)× (b× dx)
dy
du g × ( r L - r V ) ×(d - y )
i.e. =
dy mL
Integrating from y = 0 (i.e. the plate surface) to y = y, and remembering that at y = 0, u = 0, and at y = y, u =
u(y), we get:

F y2 I
GH
g × (r L - rV )× d × y -
2 JK
u(y) =
mL

LM
g × ( r L - rV ) ×d 2 y 1 y FG IJ OP
2
i.e. u (y) =
mL MN
× - ×
d 2 d H K PQ ...(11.37)

Eqn. 11.37 is the desired equation for velocity profile.


And, the mean flow velocity of the liquid at a section is given by:

z
d
1
um = × u dy
d
0

z LM FG IJ OP dy
d 2 2
1 g × ( r L - rV ) ×d y 1 y
i.e. um = ×
d
0
mL
× - ×
d 2 d MN H K PQ
BOILING AND CONDENSATION 553
g × ( r L - rV ) ×d 2
i.e. um = ...(11.38)
3×mL
Mass flow rate:
Mass flow rate of condensate through any x-position is given by:
Mass flow rate = density ´ area ´ mean velocity

g × ( r L - rV ) ×d 2
i.e. m = rL × (b × d ) ×
3×mL

r L ( r L - r V ) × g × b ×d 3
i.e. m= ...(11.39)
3× mL
Note that mass flow rate is a function of position (x), since the film thickness d is function of x.
As we proceed from position x to (x + dx), film thickness increases from d to (d + dd ), and there is additional
mass ‘dm’ condensed. This additional mass ‘dm’ condensed between x and (x + dx) is obtained by differentiating
Eq. 11.39 w.r.t. x (or d ):

dm =
LM r L ×(r L - r V )× g × b ×d 2
× dd
OP ...(11.40)
MN mL PQ
Heat flow rate:
While condensing ‘dm’ amount of liquid, certain amount of latent heat of vaporisation is released; this is equal to:
dQ = dm× h f g

i.e. dQ = h f g ×
LM r L ×( r L - r V ) × g × b ×d 2
× dd
OP ...(11.41)
MN mL PQ
But, as per the assumption, heat flow through the liquid film is by pure conduction, with linear temperature
gradient. Therefore, we can write:
kL ×(b × dx)
dQ = × (Tsat – Ts) ...(11.42)
d
From Eqs. 11.41 and 11.42:

h fg×
LM r L ×( r L
OP
- r V ) × g × b ×d 2 k × (b × bx)
× dd = L × (Tsat – Ts)
MN mL PQ d

kL × m L
i.e. d 3 × dd = ×(Tsat - Ts ) × dx (Tsat – Ts) dx
r L × ( r L - rV ) × g × hf g
Integrating the above equation with the boundary condition that d = 0 at x = 0, we get:

L 4 × k × m ×(T OP
1

d (x) = M sat - Ts ) × x
4
L L
MN r ×(r - r
L L V ) × g × hf g PQ ...(11.43)

Eq. 11.43 gives the liquid film thickness as a function of position x. Note that the film thickness increases as
the fourth root of the distance along the flow direction. Increase is rapid at the top end of the vertical plate and
slows down later.
Heat transfer coefficient:
For the heat flow through the liquid film, we have:
kL × ( b × d x )
dQ = × (Tsat – Ts)
d
Also, by Newton’s law of cooling:

554 FUNDAMENTALS OF HEAT AND MASS TRANSFER


dQ = hx × (b× dx) × (Tsat – Ts )
where, hx is the local heat transfer coefficient
From the above two relations, we get:
kL ×(b × bx)
× (Tsat – Ts) = hx × (b× dx) × (Tsat – Ts )
d
kL
i.e. hx = ...(11.44)
d
Note that at any position along the height, heat transfer coefficient is directly proportional to the thermal
conductivity kL of the condensate and inversely proportional to the thickness of the film.
Substituting the value of d from Eq. 11.43 in Eq. 11.44:

L r ×( r - r ) × k × g × h OP
1
3
= M
4
L L V L fg
hx
MN 4× m × L×(T - T )
L sat s PQ ...(11.45)

At x = L, i.e. at the lower end of the plate, local heat transfer coefficient is:

L r ×( r - r ) × k × g × h OP
1
3
= M
4
L L V L fg
hL
MN 4× m × L×(T - T )
L sat s PQ ...(11.46)

Obviously, rate of condensation heat transfer is higher at the upper end as compared to that at the lower
end.
Average value of heat transfer coefficient over the entire height of the plate is of interest to calculate the total
heat transfer rate. This is obtained by integrating Eq. 11.45 over the height L:

z
L
1
havg = × hx dx
L
0

4
We get: havg = × hL ...(11.47)
3
In the above, hL is the local heat transfer coefficient at x = L, i.e. at the lower end of the plate.
Substituting for h L from Eq. 11.46, we get:

4 Lr OP
1
3
= ×M
L × ( r L - r V ) × kL × g × hf g
4

3 M PQ
havg
N 4 × m L × L × (Tsat - Ts )

Lr OP
1
- rV ) × kL3 × g × hf g
= 0.943 × M
4
L ×( r L
i.e. havg
MN m L × L × (Tsat - Ts ) PQ ...(11.48)

Eq. 11.48 is Nusselt’s equation for average heat transfer coefficient for condensation on a vertical plate.
It is observed that in practice, experimental value of average heat transfer coefficient is about 20% higher
than that given by Nusselt’s Eq. 11.48. So, McAdams suggested to use a coefficient of 1.13 instead of 0.943 in Eq.
11.48.
Nusselts equation underpredicts the value of h, basically because:
(a) it does not take into account non-linear temperature profile in the liquid film, and
(b) it does not take into account the sub-cooling of the liquid film.
These effects can be accounted for by replacing hfg in Eq. 11.48 by h‘fg given by Eq. 11.33.
Then, we have, for average heat transfer coefficient for laminar film condensation on a vertical plate:

BOILING AND CONDENSATION 555


Lr OP
1
3
= 0.943 × M
L ×( r L - rV ) × kL × g × hf¢ g
4
havg
MN m L × L × (Tsat - Ts ) PQ W/(m2C) ...for 0 < Re f < 30 ...(11.49)

where, h¢ f g = h fg + 0.68 × CpL× (Tsat – Ts)


In the above equation all the liquid properties should be evaluated at the film temperature,
Tf = (Tsat + Ts)/2 and h fg and r V should be evaluated at Tsat.
It is desirable to get relations for heat transfer coefficient in terms of the film Reynolds number. Now, let us
define a dimensionless number called ‘Condensation number’, (Co) [or, ‘modified Nusselt number’] as follows:

L OP
1

Co = M
m 2L 3 havg
MN r L × ×( r L - r V ) × g PQ ×
kL
...(11.50a)

Since, rL >> rV , condensation number can be simplified as:

Fm I Fn I
1 1
havg 2 3 havg 2
×G
H r × g JK
3
Co = L = ×G J L
kL 2
L kL H gK ...(11.50b)

Then, Rohsenow (1985) has shown that above derived relation for heat transfer coefficient for condensation
on a vertical plate for the laminar regimes of condensate flow, can be re-cast as follows:
(a) Laminar flow, (Ref £ 30):
1
-
Co = 1.47× Re f 3 (laminar) (11.51)
(b) In the laminar–wavy region, (30 < Ref < 1800):
Kutatelazde recommends following correlation:
Re f
Co = 1. 22
(laminar–wavy) (11.52)
1.08 × Re f - 5. 2
(c) For turbulent region, (Ref > 1800):
Labuntsov recommends following correlation for turbulent film condensation:
Re f
Co = 0 .75 0. 5 (turbulent) (11.53)
8750 + 58 ×( Re f - 253) × Pr L-
Eqs. 11.51, 11.52 and 11.53 are depicted graphically in Fig. 11.9 below:
From Eq. 11.53, it may be observed that in the turbulent film condensation region, the condensation number
depends on liquid Prandtl number, PrL, too, in addition to the film Reynolds number, Ref. This Prandtl number
dependence of Co in turbulent film region is clearly shown in Fig. 11.9 for PrL = 1, 3, 5 and 10.
Above correlations for condensation on a vertical plate are applicable to condensation inside or outside
vertical tubes also, if the tube diameter is not too small.
Calculation formulas for all the three regions of film condensation on a vertical plate (or cylinder) are given
below:
Calculation formulas for laminar region, for vertcal plate are:

L OP 3

4× g M k
Ref_lam ×M L
PP
3 ×n M h
= 2
...(11.51a)
MN 3× 4 PQ
L avg

556 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Co vs. film-Re for vertical plate
1
Laminar Turbulent

Laminar-wavy
10

Condensation number
5

PrL = 1

0.1
3 4 5
10 100 1•10 1•10 1•10
Film Reynolds number

FIGURE 11.9 Condensation number vs. film Reynolds number for a vertical plate

where, havg is from Eq. (11.49)

F gI
1
-1
3
and, hvert_lam = 1.47 × kL × Re 3
f × GH n JK 2
L
...(11.51b)

Calculation formulas for laminar–wavy region, for vertcal plate are:

LM F gI OP
1 0 .82
3.7 × L × k ×(T - Ts )
= M 4.81 + PP
3
Ref_wavy L sat
×G J
MN m × h¢ Hn K 2
...(11.52a)

Q
L fg L

F gI
1
Re f _ wavy × kL 3
and, hvert_wavy= ×G J
1.08 × Re f _ wavy1. 22 - 5 . 2 Hn K 2
L
...(11.52b)

Calculation formulas for turbulent film region, for vertcal plate are:

LM 0.069 × L × k × Pr OP
4

F gI
1 3
0. 5
= M L ×(Tsat - Ts )
+ 253 P
GH n JK
L 3 0.5
Ref_turb × - 151× Pr L ...(11.53a)
MN m × h¢ L fg
2
L PQ
and,

F gI
1
Re f _ turb × kL 3
hvert_turb = ×G J
- 253) H n K
- 0.5 2
...(11.53b)
8750 + 58 × Pr L ×( Re f _ turb 0. 75 L

Chen et.al. (1987) have suggested the following universal correlation for both wavy and turbulent regions
(instead of Eqs. 11.52 and 11.53) for condensation on a vertical plate:

BOILING AND CONDENSATION 557


Fn I
1
1
havg 2
e j
3
×G J
- 0 . 44
Co = L = Re f + 5.82 ×10- 6 × Re 0f .8 × Pr 1L. 3 2 ...for Re f > 30
kL H gK ...(11.54)

11.4.4 Film Condensation on Inclined Plates, Vertical Tubes, Horizontal Tubes and
Spheres, and Horizontal Tube Banks
Inclined plates:
Eq. 11.49 for laminar condensation on vertical plates can also be used for inclined plates. If the plate is inclined at
an angle of q to the vertical, (q £ 60 deg.), replacing g by g.cos(q) in Eq. 11.49 gives satisfactory results for laminar
condensation on the upper surface of the inclined plate, i.e.

Lr OP
1
3
= 0.943 M
L × ( r L - r V ) × kL × g × cos (q )× hf¢ g
4
hinclined
MN m L × L ×(Tsat - Ts ) PQ W/(m2C) (for 0 < Ref < 30...(11.55))

We can also write:

c h
1
hinclined = hvert cos (q ) 4
(laminar...(11.56))
Vertical tubes:
Eq. 11.49 for laminar condensation on vertical plates can also be used to determine heat transfer coefficient
for laminar condensation on the outer or inner surface of a vertical tube, if the tube diameter is large compared to
the thickness of the liquid film, i.e.
if D >> d.
Horizontal tubes and spheres:
Horizontal tube–laminar film condensation:
For laminar film condensation on horizontal tubes and spheres, Nusselt type of analysis gives relations similar to
Eq. 11.49, except that L is replaced by diameter D and the value of the numerical constant is different. We get:

Lr OP
1
3
= 0.729 × M
L ×( r L - r V ) × kL × g × hf¢ g
4
hhoriz
MN m L × L ×(Tsat - Ts ) PQ W/(m2C) ...for 0 < Ref < 30 ...(11.57)

Horizontal tube–forced convection condensation:


Eq. 11.57 is for the case of a quiescent vapour condensing on a horizonal tube. However, for condensers
used in practice, a vapour may be forced through a condenser while being condensed. For the case of a cylinder
of diameter D exposed to cross flow of a vapour with a free stream velocity of U, following correlation due to
Shekriladze and Gomelauri (1966), may be applied:

LM L 1.69× g × h¢ × m × D O OP
1
1 2
1
hhoriz × D
MM1 + MM1 + U × k ×(T - T ) PP PP
fg L 2
= 0.64 × ReD2 × ...for ReD < 106 ...(11.57a)

N N Q
kL 2
L sat s
Q
r L ×U × D
where, ReD =
mL
Sphere–laminar film condensation:

Lr OP
1
- r V ) × kL3 × g × hf¢ g
= 0.826× M
4
L ×( r L
hsphere
MN m L × D ×(Tsat - Ts ) PQ W/(m2C) ...for 0 < Ref < 30 ...(11.58)

It is interesting to compare the laminar condensation on vertical and horizontal tubes. From Eqs. 11.49 and
11.57, we can write:

558 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG IJ F DI
1 1

= 1.294 × G J
hvert 0. 943 D 4 4
hhoriz
= ×
0.729 L H K H LK ...(11.59)

For hvert to be equal to hhoriz, we should have:


L = (1.294)4 × D
i.e. L = 2.8 × D
i.e. for L > 2.8.D, heat transfer coefficient will be higher for a horizontal tube. It is a fact that most of the tubes
used in practice have lengths such that L > 2.8.D. Therefore, tubes used in a steam condenser are generally
arranged in a horizontal orientation.
Horizontal tube banks: Correction factor
More than a single horizontal tube, an array of horizon- 1
tal tubes condensing a vapour on their outer surfaces is Staggered tubes
of practical interest. In a steam condenser, generally such In-line tubes
an array of tubes is used to condense low pressure steam
exiting from the steam turbine. These tubes are cooled
by circulating cold water through them. It is clear that
the vapour condensed on the outer surface of the tubes
at the top of the array will trickle down on to the tubes
0.4
below them; thus the thickness of the liquid film on the
1 20
surfaces of the tubes at the lower level will be larger, and Row number, n
as a result, the heat transfer coefficient will be lower for
the tubes at the lower level, i.e. the heat transfer coeffi- FIGURE 11.10 Correction factor for heat transfer
cient depends upon the number of the row, counting coefficient in different rows of a condenser
from top. Heat transfer coefficient for tube on the first
row is maximum, lower for second row, etc. Variation of heat transfer coefficient for condensation on the outside
of the tubes of different rows of a condenser are shown qualitatively in Fig. 11.10.
In Fig. 11.10, correction factor = h 1/hn , where
h1 = heat transfer coefficient for the first row, and
hn = heat transfer coefficient for the nth row.
Average heat transfer coefficient for film condensation on a vertical tier containing N tubes is obtained by
substituting (N.D) in place of D in Eq. 11.57 for a single horizontal tube, i.e.

L r ×(r - r )× k OP
1
3
= 0.729 × M
L × g × hf¢ g
4
L L V
h horiz_Ntubes
MN m ×(N × D)×(T
L sat - Ts ) PQ W/(m2C) ...for 0 < Ref < 30 ...(11.60)

Clearly, this is related to the value of heat transfer coefficient for a single horizontal tube as follows:
1
hhoriz_Ntubes = 1
× hhoriz_1 tube ...(11.61)
N4
Vertical Tier of N Horizontal tubes:
Chen (1961) has suggested the following modified form of Eq.11.57 for condensation on a vertical tube bank,
to take into account the condensation occurring on the sub-cooled film between two adjacent tubes:

L r ×(r - r )× k OP
1

= 0.725 × M
L L V
3
L × g × hf¢ g
4 LM CpL ×(Tsat - Ts )× ( N - 1) OP
hhoriz_Ntubes
MN m ×(N × D)×(T
L sat - Ts ) PQ × 1 + 0. 2 ×
MN hf g PQ ...(11.61)

Note that the second square brackets on the RHS, is a correction factor to Eq. 11.60; also note that term inside
this square bracket is h fg and not h¢fg. Eq. 11.61 is valid for:
CpL ×(Tsat - Ts )
£ 2 and, PrL ³ 1
hf g

BOILING AND CONDENSATION 559


and,
CpL ×(Tsat - Ts )
= Ja = Jacob number.
hf g

11.4.5 Effect of Vapour Velocity, Nature of Condensing Surface and


Non-condensable Gases
Effect of vapour velocity:
Above formulas are valid for stationary vapour or vapour moving at very low velocities (V < 10 m/s). At higher
velocities, there will be friction between the liquid film and moving vapour. For a vertical plate, if the vapour is
moving upwards, it will act to decrease the film velocity since the film is moving downwards and so, the film
thickness will increase and the heat transfer coefficient will decrease. When the frictional force overcomes the
gravity force, the film will get detached from the surface and the heat transfer coefficient will increase. Instead, if
the vapour moves downwards in the direction of motion of the liquid film, then the film velocity increases, film
thickness decreases, and as a result, the heat transfer coefficient increases. It is also observed that at low pres-
sures, effect of vapour velocity on heat transfer coefficient is small.
Effect of nature of condensing surface:
For a rough surface, or for a surface covered with an oxide film, the surface offers additional resistance to the
flow of the film, thus increasing the thickness of the liquid film. This results in reduction of heat transfer coeffi-
cient.
Effect of non-condensable gases:
If the vapours contain air or other non-condensable gases, heat transfer coefficient reduces drastically. This hap-
pens because, only the vapours condense on the surface, and the non-condensable gases form a ‘cloud’ near the
surface impeding the approach of vapours to the surface for further condensation. In connection with steam
condensers, practical data indicate that even a one per cent of air (by mass) in the vapour causes a reduction of
about 60% in the value of heat transfer coefficient. So, in industrial condensers, provision is made for continuous
venting of the non-condensables from the system.
11.4.6 Simplified Calculations for Water
With reference to industrially important steam condensers, water vapour is condensed on the outside of tubes
arranged mostly in a horizontal orientation. Then, for a horizontal tube, we have:

Lr OP
1
3
= 0.729 × M
L × ( r L - r V ) × kL × g × hf¢ g
4
hhoriz
MN m L × D ×(Tsat - Ts ) PQ W/(m2C) ...for 0 < Ref < 30 ...(11.57)

Then, without the correction for liquid sub-cooling, and for rL >> rV , we can write:
1
F r ×k I
1
2 3 4 ( g × hf g ) 4
= 0.729 × G
H m JK
L L
hhoriz × 1
W/(m2C) ...(11.62)
L
[D ×(Tsat - Ts )] 4

A×B
i.e. hhoriz = 0.729 × 1
W/(m2C) ...(1162a)
[D ×(Tsat - Ts )] 4
where,

F r ×k I
1
2 3 1
e j
4
A= G
H m JK
L L
and, B = g × h f g 4
L

For N horizontal tubes in a vertical tier, D in Eq. 11.62a is replaced by (N.D).


For vapours condensing on a vertical tube, again, Eq. 11.62a can be used, but with the modification that
numerical constant 0.729 is replaced by 1.13 and D is replaced by the height of the tube, L.

560 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Here, parameter A is evaluated at the mean film temperature {= (Tsat + Ts)/2} and parameter B is evaluated
at Tsat.
Values of A and B can be conveniently tabulated for water for different values of Tsat and Ts , respectively.
For the industrially important case of a steam condenser, Eq. 11.62a can be used to make a quick estimate of the
heat transfer coefficient; values of parameters A and B can be read from Table 11.5 and Table 11.6, respectively, or
from Fig. 11.10 and Fig. 11.11, respectively.
Example 11.9. Saturated steam at atmospheric pressure condenses on a vertical plate (size: 30 cm x 30 cm) maintained at
60°C. Determine heat transfer rate and the mass of steam condensed per hour.
(b) If the plate is tilted at an angle of 30 deg. to the vertical, what is the value of condensation rate?
Solution.
Data:
L := 0.3 m b := 0.3 m Tsat := 100°C Ts := 60°C and, Tf := 80°C (mean film temperature = (100 + 60)/2)
Properties of liquid at Tf = 80°C:
rL := 971.8 kg/m3 CpL := 4197 J/kg.C mL := 0.355 ´ 10 – 3 kg/(ms) kL := 0.67 W/(mC) g := 9.81 m/s2
Properties of saturation vapours at 100°C:
h f g := 2257 ´ 103 J/kg rV := 0.5978 kg/m3
Average heat transfer coefficient:
Let us assume laminar film condensation; we shall check this later.
Then, we have:
h¢ f g := hfg + 0.68×CpL×(Tsat – Ts) J/kg (modified heat of vaporisation)

TABLE 11.5 Values of parameter ‘A’ at different mean film temperatures for water

3
R| r ×k
A= S
2
L
3
L
U|V 1
4
Temp, deg.C r (liq), kg/m Cp (liq), J/kg.C k (liq), W/mC m , kg/m.s
|T m L |W
5 999.9 4205 0.571 0.001519 105.21
10 999.7 4194 0.58 0.001307 110.52
15 999 4186 0.589 0.001138 115.70
20 998 4182 0.598 0.001002 120.75
25 997 4180 0.607 0.000891 125.68
30 996 4178 0.615 0.000798 130.40
35 994 4178 0.623 0.00072 134.97
40 992.1 4179 0.631 0.000653 139.50
45 990.1 4180 0.637 0.000596 143.59
50 988.1 4181 0.644 0.000547 147.76
60 983.3 4185 0.654 0.000467 155.13
80 971.8 4197 0.67 0.000355 168.19
100 957.9 4217 0.679 0.000282 178.65
120 943.4 4244 0.683 0.000232 186.98
140 921.7 4286 0.683 0.000197 192.53
160 907.4 4340 0.68 0.00017 197.55
180 887.3 4410 0.673 0.00015 200.00
200 864.3 4500 0.663 0.000134 200.77
220 840.3 4610 0.65 0.000122 199.67
240 813.7 4760 0.632 0.000111 196.99
260 783.7 4970 0.609 0.000102 192.04
280 750.8 5280 0.581 0.000094 185.19
300 713.8 5750 0.548 0.000086 176.71
320 667.1 6540 0.509 0.000078 165.62
340 610.5 8240 0.469 0.00007 153.09
360 528.3 14690 0.427 0.00006 137.95

BOILING AND CONDENSATION 561


250

200

Parameter A
150
A
100

50

0
0 100 200 300 400
Mean film temperature (deg.C)
FIGURE 11.10 Parameter A for water

i.e. h¢ fg = 2.371 ´ 106 J/kg


and,

Lr OP
1

:= 0.943× M
L × ( r L - r v ) × kL3 × g × h ¢f g 4
havg W/(m2C)...for 0 < Ref < 30.
MN m L × L × (Tsat - Ts ) PQ ...(11.49)

i.e. h avg = 5.917 ´ 103 W/(m2C) (average heat transfer coefficient for the plate)
Heat transfer rate:
Q := havg × (L × b × (Tsat – Ts), W
i.e. Q = 2.13 ´ 104 W (heat transfer rate to the plate)
Condensation rate of steam:
Q
m := kg/s
h¢f g
i.e. m = 8.983 ´ 10 – 3 kg/s
i.e. m = 32.34 kg/h (condensation rate per hour.)
Now, check the assumption of laminar film condensation:
We have, film Reynolds number, given by:
4× m
Re f = ...(11.32)
P×mL
where, P = wetted perimeter = b, for vertical plate
4× m
Therefore, Re f :=
b× mL
i.e Ref = 337.401 (< 1800, therefore laminar.)
However, Re f is > 30; therefore, it is laminar–wavy region.
So, let us assume that condensation is laminar up to a particular distance L1 from top, and then it becomes wavy.
Let us find out the distance from top, where film becomes wavy.
For purely laminar film: Ref =30
Re f × b ×m L
i.e. m1 :=
4
i.e. m1 = 7.987 ´ 10 – 4 kg/s (mass flow rate at the end of laminar film)
Q
Now, m1 = 1 ...(a)
h¢f g
where, Q1 is the heat transfer to the plate up to the length L1

562 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 11.6 Values of parameter ‘B’ at different saturation temperatures for water
1
Tsat, deg.C Psat, kPa hfg, kJ/kg B = [9.81 ´ h f g ´ 103] 4
5 0.8721 2490 70.302
10 1.2276 2478 70.217
15 1.7051 2466 70.132
20 2.339 2454 70.046
25 3.169 2442 69.961
30 4.246 2431 69.882
35 5.628 2419 69.795
40 7.384 2407 69.709
45 9.593 2395 69.622
50 12.35 2383 69.534
60 19.94 2359 69.358
80 47.39 2309 68.988
100 101.33 2257 68.596
120 198.53 2203 68.182
140 361.3 2145 67.729
160 617.8 2083 67.234
180 1002.1 2015 66.679
200 1553.8 1941 66.058
220 2318 1859 65.349
240 3344 1767 64.525
260 4688 1663 63.554
280 6412 1544 62.385
300 8581 1405 60.931
320 11274 1239 59.045
340 14586 1028 56.353
360 18651 720 51.553
374.14 22090 0 0.000

80
70
60
Parameter B

50

40 B

30
20
10
0
0 100 200 300 400
Tsat (deg. C)

FIGURE 11.11 Parameter B for water

and, Q1 := havg × (L1× b)× (Tsat – Ts) ...(b)


3
i.e. Q1 = 5.255 ´ 104× L 4
1

BOILING AND CONDENSATION 563


Then, from Eq. a:
Q1
m1 =
h¢f g
3

–4 5. 255 ´ 10 4 × L14
i.e. 7.987 ´ 10 =
h f¢ g

F 7.987 ´ 10 × h¢ I
4
-4 3
:= G
H 5.255 ´ 10 JK
fg
i.e. L1 4

i.e. L1 = 0.012 m (distance from top, up to which film is laminar)


Therefore, length of plate over which film is wavy = L2 = 0.3 – L1
i.e. L2 = 0.288 m
Over the length L2, we use equation applicable to wavy film, namely,
Re f
Co = (laminar–wavy...(11.52))
1. 08 × Re 1f . 22 - 5. 2
where, by definition, Condensation number is:

Fm I
1
hwavy 2

H r × g JK
×G
3
L
Co = ...(11.50b)
kL 2
L

Let the rate of condensation over the total length of the plate be mtot:
Then, mtot = m1 + m2 where m2 = mass condensation over length L2
hwavy × ( L2 × b) × (Tsat - Ts )
i.e. mtot = m1 +
h¢f g
–4
i.e. mtot = 7.987 ´ 10 + hwavy (1.458 ´ 10 –6)
Now, film Reynolds number for wavy region:
4 × mtot 4
Ref_wavy = = × [7.987 ´ 10 –4 + hwavy × (1.458 ´ 10 –6)]
b× mL b× mL
i.e. Ref_wavy = 29.998 + 0.055 × hwavy (c)
Then, from Eq. 11.50b and 11.52 and Eq. c we get:

Fm I
1
hwavy 2 29.998 + 0.055 × hwavy
H r × g JK
×G
3
L
=
kL 2
L 1. 08 × ( 29.998 + 0 .055 × hwavy )1. 22 - 5 .2
Solve this for hwavy using solve block of Mathcad:
Start with a guess value for hwavy. Then, immediately after ‘Given,’ write the constraint equation Then, type the
command “Find (hwavy)” and it is calculated immediately:
hwavy = 1000 (guess value)
Given

Fm I
1
hwavy 2 29.998 + 0.055 × hwavy
GH r × g JK
3
L
× =
kL 2
L 1. 08 × ( 29.998 + 0 .055 × hwavy )1. 22 - 5 .2
Find (hwavy) = 6.934 ´ 10 3
i.e. hwavy := 6934 W/(m2C) (heat transfer coefficient in the wavy region)
Now, from Eq. c:
Ref_wavy := 29.998 + 0.055×hwavy ...(c)
i.e. Ref_wavy = 411.368 < 1800, and > 30 (therefore, within the wavy region.)
Therefore, use of Eq. 11.52 for wavy region is justified.

564 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Then, total rate of condensation:
mtot := 7.987 ´ 10 –4 + hwavy ×(1.458 ´ 10 –6 )
i.e. mtot = 0.011 kg/s
i.e. mtot × 3600 = 39.27 kg/h (total rate of mass condensation.)
Comments: Above procedure, wherein we took into account the small, initial length from top of the plate where the film
is laminar, is quite accurate. But, note that laminar film exists only for a distance of 0.012 m from top, whereas for the
remaining 0.288 m of height, the flow is laminar–wavy.
So, if we make an approximation that the flow is laminar–wavy for the entire height of the plate, we get:
Heat transferred: Q := hwavy ×(L× b) ×(Tsat – Ts)
i.e. Q = 2.496 ´ 10 4 W
Q
And, rate of condensation: mtot :=
h¢f g
i.e. mtot = 0.011 kg/s
Note that this value of mtot is practically the same as obtained earlier with exact analysis. Therefore, it appears that
calculations made assuming that the entire plate has laminar–wavy film, involve negligible error, with the advantage
that the calculations are much easier to make.
Assuming that the film is laminar–wavy over the entire length of plate, calculations would proceed as follows:
Using the calculation formulas for laminar–wavy film region:
Laminar–wavy Reynolds number:

LM 3.7 × L × k ×(T - Ts ) F gI
1
OP 0 . 82

×G J
3
Ref_wavy = M 4 .81 + L
PP
Hn K
sat
...(11.52a)
MN m × h¢ L fg
2
L
Q
mL
Now, since nL = we write:
rL

LM LM OP
1
3
OP 0 . 82

MM 3.7 × L × k ×(T -T ) M g PP PP
Ref_wavy := M 4.81 +
L sat s
×M PP
MM m × h¢ L fg MM FG m IJ 2
PP
PP
L

MN NH r K L Q Q
i.e. Ref_wavy = 398.911 (Laminar–wavy Reynolds number)
Heat transfer coefficient in laminar–wavy region:

F gI
1
Re f _wavy × kL
- 5. 2 GH n JK
3
We have: hvert_wavy := × ...(11.52b)
1. 08 × Re f _ wavy 1. 22 2
L

i.e. hvert_wavy = 6.982 ´ 103 W/(m2C) (heat transfer coefficient in the wavy region.)
Compare this value with the value of 6934 W/(m2C) obtained earlier with exact analysis.
Heat transferred:
Q := hvert_wavy ×(L×b)×(Tsat – Ts) W
i.e. Q := 2.514 ´ 10 4 W
and, rate of condensation:
Q
mcond :=
h¢f g
i.e. mcond = 0.011 kg/s
i.e. mcond = 38.162 kg/h.
(b) If the plate is tilted at 30 deg. to vertical:
We can determine h by replacing g by g.cos (q ) in equation for vertical plate.
Since h is already determined for vertical plate, we can use the relation:

BOILING AND CONDENSATION 565


1
hinclined = hvert × (cos (q )) 4 (laminar...(11.56))
Note: while using Mathcad, for trigonometric functions, q must be expressed in radians:
30 ×p
i.e. q := radians
180
i.e. q = 0.524 radians
Therefore,
1
h inclined := hvert_wavy × (cos (q )) 4
i.e. hinclined = 6.735 ´ 10 3 W/(m2C) (heat transfer coefficient for inclined plate.)
and,
Heat transferred:
Q := h inclined × (L× b)× (Tsat – Ts)) W
i.e. Q = 2.425 ´ 104 W
and, rate of condensation:
Q
mcond :=
h ¢f g
i.e. mcond = 0.01 kg/s
i.e. mcond = 36.814 kg/h.
i.e. for an inclined plate, rate of condensation is decreased by about 3.5% as compared to that on a vertical plate.
Example 11.10. Saturated steam at a temperature of 65°C condenses on a vertical surface at 55°C. Determine the thick-
ness of the condensate film at locations 0.2 m and 1.0 m from top. Also, calculate the condensate flow rate, local and
average heat transfer coefficients at these locations.
Properties of water at the mean temperature are: (M.U. 2002)
Data:
r L := 983.3 kg/m3 kL := 0.654 W/(mK) mL := 4.67 ´ 10 –4 kg/(ms) CpL := 485 J/(kgC)
h f g := 2346 ´ 103 J/kg Tsat := 65°C Ts := 55°C g := 9.81 m/s2 b := 1 m (width, assumed)
Solution. Thickness of condensate film:
We have, from Eq. 11.43

LM 4 × k × m ×(T OP
1

L L sat - Ts ) × x 4
d (x) :=
MN r ×(r - r
L L V ) × g × hf g PQ ...(11.43)

When rL >> rV we can write:

LM 4 × k × m OP
1

L × (Tsat - Ts ) × x 4
L
(thickness of condensate as a function of position)
d (x) :=
MN r L × g × hf g
2
PQ m

Therefore,
d (0.2) = 1.024 ´ 10 –4 m (thickness of condensate film at 0.2 m from top.)
and, d (1.0)= 1.531 ´ 10 – 4 m (thickness of condensate film at 1.0 m from top.)
Mass flow rate of condensate:
We have, from Eq. 11.39:
r L × ( r L - r V ) × g × b ×d 3
m= ...(11.39)
3×m L
Again, when r L >> rV we can write:
r L2 × g × b × [d ( x )]3
m(x) := kg/s (mass flow rate as function of position)
3×mL
Therefore,
m(0.2) = 7.262 ´ 10 – 3 kg/s (mass flow rate of condensate at 0.2 m from top)
and, m(1.0) = 0.024 kg/s (mass flow rate of condensate at 1.0 m from top)

566 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Heat transfer coefficients:
We can use Eq. 11.44 or 11.45. Since d is already calculated, it is easier to use Eq. 11.44:
kL
i.e. hx = ...(11.44)
d
Again, writing hx as a function of x, Eq. 11.44 is re-written as:
kL
hx (x) :=
d (x )
Then,
hx (0.2) = 6.389 ´ 103 W/(m2C) (local heat transfer coefficient at 0.2 m from top)
and, hx (1.0)= 4.272 ´ 103 W/(m2C) (local heat transfer coefficient at 1.0 m from top)
Further, average heat transfer coefficient between x = 0 and x = x is given by Eq. 11.47, i.e.
4
havg = × hx ...(11.47)
3
We re-write this as:
4
havg (x) := × hx (x) (average heat transfer coefficient between x = 0 and x = x )
3
Therefore,
havg (0.2) = 8.518 ´ 103 W/(m2C) (average heat transfer coefficient between x = 0 and x = 0.2)
and, havg (1.0) = 5.697 ´ 103 W/(m2C) (average heat transfer coefficient between x = 0 and x = 1.0 m)
McAdams has suggested that this calculated value of havg should be increased by 20% to account for liquid sub-
cooling, i.e.
havg (1.0) = 1.2 ´ 5697 = 6846.4 W/(m2C) (corrected average heat transfer coefficient between x = 0 and x = 1.0 )m.
Note: All the above calculations have assumed that the film is purely laminar during condensation.
(b) In addition, if the height of the plate is 1 m let us calculate the following:
(i) heat transfer rate to the plate,
(ii) maximum velocity of condensate at the trailing edge, and
(iii) also, draw the variation of d with distance from top.
L := 1 m (height of vertical plate)
Let us check if the condensation is of laminar type:
We should find film Reynolds number.
From Eq. 11.32 we have:
4× m
Re f = where P = wetted perimeter = b = 1 m
P×mL
m has already been calculated as 0.024 kg/s at a distance of 1 m from top.
4 × 0.024
Then, Re f :=
b× mL
i.e. Re f = 205.567 < 1800 (therefore, laminar.)
Note: However, note that Re f > 30; therefore, flow is really in the laminar–wavy region.
Assuming that the film is laminar–wavy over the entire length of plate, calculations would proceed as follows:
Using the calculation formulas for laminar–wavy film region:
We have:
h¢f g := hf g + 0.68 ×CpL× (Tsat – Ts) J/kg (modified heat of vaporisation)
i.e. h¢fg = 2.349 ´ 10 6 J/kg
Laminar–wavy Reynolds number:

LM 3.7 × L × k ×(T - Ts ) L g OP
1
OP 0 . 82

×M
3
Ref_wavy = M 4. 81 + L sat
PP ...(11.52a)
MN m × h¢
L fg N (n ) Q
L
2

Q
mL
Now, since n L = we write:
rL

BOILING AND CONDENSATION 567


LM LM OP
1
3
OP 0 . 82

MM 3.7 × L × k ×(T -T ) M g PP PP
:= M 4 .81 + ×M PP
L s
Re f_wavy sat

MM m × h¢ L fg MM FG m IJ 2
PP
PP
L

MN NH r K L Q Q
i.e. Ref_wavy = 235.355 (Laminar–wavy Reynolds number)
Heat transfer coefficient in laminar–wavy region:

F gI
1
Re f _wavy × kL
×G J
3
hvert_wavy :=
- 5. 2 H n K
We have: ...(11.52b)
1. 08 × Re f _ wavy 1. 22 2
L

i.e. hvert_wavy = 6.445 ´ 103 W/(m2C) (heat transfer coefficient in the wavy region.)
Note: Compare this value with the value of 6836.4 W/(m2C) obtained earlier for pure laminar film.
(i) Heat transfer rate:
Q := h vert_wavy × (L× b) × (Tsat – Ts) W
i.e. Q = 6.445 ´ 104 W (heat transfer rate to the plate.)
(ii) Maximum velocity of condensate at the bottom (i.e. trailing edge of plate):
Velocity is given by:

g ×(r L - rV )×d 2 y 1 y LM FG IJ OP 2
u (y) =
mL
× - ×
d 2 d MN H K PQ ...(11.37)

Now, the maximum velocity occurs when y = d .


Therefore, remembering that rL >> rV we can write for maximum velocity at the bottom of the plate, i.e. at x = L,
where d = d (L):
LM
g × r L × d ( L) 2 d 1 d FG IJ OP
2
umax =
mL
× - ×
MN
d 2 d H K PQ
g × r L × d (L ) 2
i.e. umax :=
2× mL
i.e. umax = 0.242 m/s (maximum velocity at the bottom of plate.)
(iii) Draw the variation of d with the distance from the top of the plate, x:
It is convenient to use Mathcad to draw this graph:
x := 0, 0.05, ..., 1.0 (define a range variable x varying from x = 0 to x = 1 m at an interval of 0.05 m.)
Then, to draw the graph, click on the graph pallete, choose x–y graph, fill up the place holder on x-axis by x, and
place holder on y-axis by [d (x). 104) Click anywhere outside the graph area, and the graph appears:
Verify that at x = 1 m, d is equal to 1.531 ´ 10 – 4, as already calculated. It is observed that film thickness is zero at x
= 0, i.e. the top edge and goes on increasing with increasing x, i.e. as one travels down the plate.
Example 11.11. Dry, saturated steam at atmospheric pressure condenses on a horizontal tube of diameter = 5 cm and
length L = 1 m; surface of the tube is maintained at 60°C. Determine heat transfer rate and the mass of steam condensed
per hour. Assume laminar film condensation.
(b) If the tube is vertical, what is the condensation rate?
Solution.
Data:
FG (100 + 60) IJ
L := 1.0 m D := 0.05 m Tsat := 100°C Ts := 60°C
H
\ Tf := 80°C film temp. =
2 K
Properties of liquid at Tf := 80°C:
rL := 971.8 kg/m3 CpL := 4197 J/kgC mL := 0.355 ´ 10 –3 kg/(ms) kL := 0.67 W/(mC) g := 9.81 m/s2
Properties of saturated vapour at 100°C:
h fg := 2257 ´ 103 J/kg (enthalpy of vaporisation)
rV := 0.5978 kg/m3 (density of vapour)

568 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Variation of film thickness with x
2
1.8
1.6

4
Film thickness (m) ´ 10
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Distance from top (m)
Film thickness

(a) For horizontal tube:


We shall make a quick estimate using Eq. 11.62a:
A×B
i.e. hhoriz = 0.729 × 1 W/(m2C) ...(11.62a)
[D × (Tsat - Ts )] 4
where,

F r ×k I
1
2 3 1
A= G
H m JK
4
L L
and, B = ( g × h f g ) 4 .
L

Now, A and B are to be evaluated at the mean film temperature (= 80°C) and at saturation temperature of steam (=
100°C), respectively. Value of A is read from Table 11.5 (or Fig. 11.10) and that of B from Table 11.6 (or Fig. 11.11). We
get:
A := 168.19
and, B := 68.596
Therefore,
A×B
hhoriz := 0.729 × 1 W/(m2C)
[D × (Tsat - Ts )] 4

i.e. hhoriz = 7.072 ´ 103 W/(m2C) (average heat transfer coefficient for horizontal tube)
Heat transfer rate:
Q := hhoriz × (p × D× L) × (Tsat – Ts) W
i.e. Q = 4.444 ´ 104 W (heat transfer rate to the tube.)
Condensation rate of steam:
Q
m := kg/s
hf g
i.e. m = 0.01969 kg/s
i.e. m = 70.879 kg/h (condensation rate per hour for horizontal tube.)
(b) For vertical tube:
We shall make an estimate of ‘h’ using Eq. 11.62a. Since that equation is for horizontal tubes, we should change the
numerical constant to 1.13, and D to L, to apply for vertical tubes:

BOILING AND CONDENSATION 569


A ×B
i.e. hvert = 1.13 × 1
W/(m2C) ...(11.62a)
[L × (Tsat - Ts )] 4
Values of A and B are already determined:
i.e. A := 168.19
and, B := 68.596
Therefore,
A×B
hvert := 1.13 × 1 W/(m2 C) ...(11.62a)
[L × (Tsat - Ts )] 4

i.e. hvert = 5.184 ´ 103 W/(m2C) (average heat transfer coefficient for vertical tube)
Heat transfer rate:
Q := hvert × (p × D× L)× (Tsat – Ts) W
i.e. Q = 3.257 ´ 10 4 W (heat transfer rate to vertical tube.)
Condensation rate of steam:
Q
m := kg/s
hf g
i.e. m = 0.01443 kg/s
i.e. m = 51.953 kg/h (condensation rate per hour for vertical tube)
Ratio of condensation rates:
mhoriz 70. 879
= = 1.364
mvert 51.953
i.e. amount of steam condensed on a horizontal tube is 36.4% more as compared to the amount condensed for a vertical
tube.
Note: In the above analysis, effect of liquid sub-cooling has been neglected; also, laminar film condensation is assumed.
Example 11.12. Dry, saturated steam at atmospheric pressure condenses on a vertical tube of diameter = 5 cm and length
L = 1.5 m; surface of the tube is maintained at 80°C. Determine the heat transfer rate and the mass of steam condensed
per hour.
Solution.
Data:
FG (100 + 80) IJ
L := 1.5 m D := 0.05 m Tsat := 100°C Ts := 80°C
H
\ Tf := 90°C mean film temp. =
2 K
Properties of liquid at Tf := 90°C
rL := 965.3 kg/m3 CpL := J/kgC mL := 0.315 ´ 10 –3 kg/(ms) kL := 0.675 W/(mC)
2
PrL := 1.96 g := 9.81 m/s
Properties of saturated vapour at 100°C:
h f g := 2257 ´ 103 J/kg (enthalpy of vapourization)
rV := 0.5978 kg/m3 (density of vapour)
To start with, we shall assume that the film thickness d << D, the tube diameter and also that the condensation is in
pure laminar region; we shall check these assumptions later.
For laminar film condensation on vertical surface, we have:

Lr OP
1

:= 0.943 × M
L × ( r L - r V ) × kL3 × g × h ¢f g 4
h avg W/(m2C) ...for 0 < Re f < 30
MN m L × L × (Tsat - Ts ) PQ ...(11.49)

where, h¢f g := h f g + 0.68 CpL (Tsat – Ts)


Therefore, h¢fg = 2.314 ´ 106 J/kg
Then,

Lr OP
1

:= 0.943× M
L × ( r L - r V ) × kL3 × g × hf¢ g 4
Now, havg W/(m2C)
MN m L × L × (Tsat - Ts ) PQ
570 FUNDAMENTALS OF HEAT AND MASS TRANSFER
i.e. havg = 4.83 ´ 10 3 W/(m2C)
Heat transfer rate:
Q := havg ×(p ×D×L) ×(Tsat – Ts) W
i.e. Q = 2.276 ´ 104 W (heat transfer rate to the tube.)
Condensation rate of steam:
Q
m := kg/s
h¢f g
i.e. m = 9.83452 ´ 10 – 3 kg/s
i.e. m = 35.404 kg/h (condensation rate per hour for horizontal tube.)
Now, let us check the assumptions:
Film thickness d :
We have, from Eq. 11.43:

LM 4 × k × m ×(T OP
1

L L sat - Ts ) × x 4
d (x) :=
MN r ×(r - r
L L V ) × g × hf g PQ ...(11.43)

Then, at x = L = 1.5 m: d (1.5) = 1.876 ´ 10 – 4 m (film thickness at x = 1.5 m)


i.e. film thickness at the bottom end of the tube = 0.1876 mm << 5 cm. Therefore, the condition that d << D is satisfied.
Next, let us check if film—Reynolds number is < 30:
Reynolds number:
We have:
4× m
Ref = ...(11.32)
P×mL
where, P = wetted perimeter, i.e.
P := p × D i.e. P = 0.157 m
4×m
Therefore, Re f :=
P ×mL
i.e. Re f = 795.028 (< 1800...therefore, laminar)
However, Ref > 30; so, it is in the laminar–wavy region. So, let us make calculations for laminar–wavy region:
Laminar–wavy region:
We have:
Calculation formulas for laminar–wavy region, for vertical plate/cylinder are:
mL
nL := kinematic viscosity of liquid
rL
i.e. nL = 3.263 ´ 10 – 7 m2/s

LM 3.7 × L × k ×(T - Ts ) F gI
1
OP 0 . 82

×G J
3
Re f_wavy := MM4.81 + m × h¢ L sat

Hn K 2 PP ...(11.52a)
N L fg L
Q
i.e. Ref_wavy = 1.016 ´ 103 (film Reynolds number for wavy region)

F gI
1
Re f _wavy × kL
GH n JK
3
and, hvert_wavy := × ...(11.52b)
1. 08 × Re f _ wavy 1. 22 - 5.2 2
L

i.e. hvert_wavy = 6.16 ´ 10 3 W/(m2C) (heat transfer coefficient for the wavy region.)
Assuming that film on the entire tube length is laminar–wavy (see comments in Example 11.9),
Heat transferred:
Q := hvert_wavy × (p × D× L)× (Tsat – Ts) W
i.e. Q = 2.903 ´ 10 4 W.
And, rate of condensation:
Q
mcond :=
h¢f g

BOILING AND CONDENSATION 571


i.e. mcond = 0.013 kg/s
i.e. mcond = 45.155 kg/h.
Note: Let us compare the value of heat transfer coefficient for the wavy region, obtained above using Eq. 11.52b with
that obtained using Chen’s correlation 11.54:
From Chen’s correlation:

Fv I
1
havg 2 1
×G J
3
L
Co = = ( Re f 0 . 04 + 5.82 ×10 - 6 × Re 0f . 8 × Pr 1L. 3) 2 ...for Re f > 30
kL H gK ...(11.54)

F gI
1
1

GH v JK
3
i.e. havg := k L × 2
× ( Re f _ wavy 0 . 04 + 5.82 ×10 - 6 × Re 0f _wavy
.8
× Pr 1L. 3) 2
L

i.e. havg := 6.89 ´ 10 3 W/(m2C) (heat transfer coefficient with Chen’s correlation.)
This value, when compared to the value of 6160 W/(m2C) obtained earlier, is about 11.9% higher.
Example 11.13. A steam condenser consists of a square array of 400 horizontal tubes, each 6 mm in diameter. The tubes
are exposed to exhaust steam arriving from the turbine at a pressure of 0.1 bar. If the tube surface temperature is main-
tained at a temperature of 25°C by circulating cold water through the tubes, determine the heat transfer coefficient and
the rate at which the steam is condensed per unit length of tubes for the entire array. Assume laminar film condensation
and that there are no condensable gases mixed with steam.
Solution.
Data:
N := 400 i.e. N := 20 L := 1.0 m D := 0.006 m Tsat := 45.8°C
Ts := 25°C \ Tf := 35.4°C (mean film temp.)
Properties of liquid at Tf = 35.4°C:
rL := 994.04 kg/m 3 CpL :=4178 J/kgC mL := 0.720 ´ 10 – 3 kg/(ms) kL := 0.623 W/(mC)
2
PrL := 4.83 g := 9.81 m/s
Properties of saturated vapour at 45.8°C:
h f g := 2393 ´ 103 J/kg (enthalpy of vapourization)
rV := 0.068 kg/m3 (density of vapour)
Then, for laminar film condensation on a vertical tier consisting of N horizontal tubes, we have

LM r ×( r - r ) × k OP
1
3
L L V L × g × hf¢ g 4
hhoriz_Ntubes = 0.729× W/(m2C) ...for 0 < Re f < 30
MN m ×( N × D) ×(T
L sat - Ts ) PQ ...(11.60)

where, h ¢fg := h f g + 0.68 × CpL × (Tsat – Ts)


i.e. h ¢f g = 2.452 ´ 106 J/kg

L r × (r - r ) × k × g × h ¢ OP
1

:= 0.729 × M
3 4
L L V L fg
hhoriz_Ntubes W/(m2C)
and,
MN m ×( N × D)×(T - T )
L sat s PQ
i.e. hhoriz_Ntubes = 5.482 ´ 103 W/(m2C) (average heat transfer coefficient)
Therefore,
Heat transfer rate/m length:
Q := hhoriz_Ntubes × (p × D × L) × (Tsat – Ts) W
i.e. Q = 2.149 ´ 103 W/m (heat transfer rate to each tube/metre length.)
Condensation rate of steam:
Q
m := kg/s
h¢f g
i.e. m := 8.76542 ´ 10 – 4 kg/s per metre length of each tube
Then, for the entire array of 400 tubes, total condensation rate per metre length is:
mtot := m × 400
i.e. mtot = 0.351 kg/s per meter length of array
i.e. mtot = 1262 kg/h (condensation rate per hour per metre length of array)

572 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Alternatively:
We an use Chen’s correlation for average heat transfer coefficient on a vertical tier of N horizontal tubes:

L r ×(r - r )× k OP × LM1 + 0.2 × C OP


1

= 0.725 × M
3
L L V L × g × hf¢ g 4
pL ×(Tsat - Ts ) × ( N - 1)
hhoriz_Ntubes
MN m ×( N × D) ×(TL sat - Ts ) PQ MN hf g PQ ...(11.61)

CpL × (Tsat - Ts )
for: £ 2 and, Pr L ³ 1
hf g
Let us check the conditions first:
CpL × (Tsat - Ts )
= 0.036 < 2
hf g
and, PrL = 4.83 > 1
Therefore, the conditions are satisfied.
Then, we have:

L r ×(r - r )× k OP LM OP
1

:= 0.725 M
3
L L V L × g × h f¢ g 4 CpL ×(Tsat - Ts ) × ( N - 1)
hhoriz_Ntubes × 1 + 0. 2 ×
MN m ×( N × D)×(T
L sat - Ts ) PQ MN hf g PQ
3 2
i.e. h horiz_Ntubes = 6.204 ´ 10 W/(m C) (average heat transfer coefficient.)
Compare this value with h = 5482 W/(m 2C), obtained earlier from Eq. 11.60. We see that Chen’s correlation gives
about 13.2% higher value of h.
Correspondingly, condensation rate also will be 13.2% higher:
i.e. mtot := 1262 × (1.132)
i.e. mtot = 1429 kg/h (condensation rate per hour per metre length of array.)

11.4.7 Film Condensation inside Horizontal Tubes


Condensers used in vapour compression refrigeration and air-conditioning systems have refrigerant vapours
condensing inside horizontal or vertical tubes. The situation is complicated to analyse and, obviously, heat trans-
fer depends on vapour velocity.
For low vapour velocities, Chato recommends following correlation for condensation inside horizontal
tubes, when the tube length is short or the rate of condensation is small:

L g × r ×(r - r )× k × Lh OPOP
1

= 0.555× M
3
3 4

MN m ×(T - T ) × D MN
L L V L
hinternal + × CpL ×(Tsat - Ts ) ...(11.63)
L sat s
fg
8 QPQ
F r × u ×D I < 35000
= G
H m JK
V× V
for Revap
V inlet

Reynolds number of vapours is evaluated at the inlet conditions, with the inside diameter of the tube as the
characteristic length.
If the rate of condensation is high, or the tube is long, correlation of Ackers et al. may be used (with the
understanding that error may be as high as 50%):
1
hinternal × D
= 0.026 ×Rem0.8 × Pr L3 ...(11.63a)
k

D
LM r F I OP
1

GH JK P
2
where, R em =
mL MM
× GL + GV × L
rV
PQ
N
and, G L and GV are liquid and vapour mass velocities (kg/(m2s)) calculated on the basis of internal cross section
of the tube (= p . D2/4). These correlations may be used for vapour Reynolds numbers > 20,000 and the liquid
Reynolds number > 5,000.
Example 11.14. Ammonia at 40°C is condensing inside a horizontal tube of 16 mm ID. Mass velocity of ammonia vapour
at inlet is 20 kg/(m2.s). Surface of the tube is maintained at a constant temperature of 20°C by circulating cold water.
Calculate the fraction of vapour that will condense if the tube is 0.5 m long.

BOILING AND CONDENSATION 573


Solution.
Data:
L := 0.5 m D := 0.016 m GV := 20 kg/(m2s) Tsat := 40°C Ts := 20°C \ Tf := 30°C (mean film temp.)
Properties of liquid at Tf = 30°C:
r L := 596.4 kg/m3 CpL := 4890 J/kgC mL := 2.081 ´ 10 – 4 kg/(ms) kL := 0.507 W/(mC) g := 9.81 m/s2
Properties of saturated vapour at 40°C:
h f g := 1098.8 ´ 10 3 J/kg rV := 12.029 kg/m3 mV := 1.0735 ´ 10 –5 kg/(ms)
Vapour Reynolds number:
GV
Vapour velocity: V V :=
rV
i.e. V V = 1.663 m/s (quite low)
GV × D
And, ReV :=
mV
i.e. ReV = 2.981 ´ 104 < 35,000 (Reynolds number)
Therefore, we can use the correlation of Chato for film condensation inside a tube:

LM g × r ×( r LM OPOP
1

L L - r V ) × kL3 3 4
hinternal := 0.555× × h f g + × CpL × (Tsat - Ts )
MN m ×(TL sat - Ts ) × D N 8 QPQ ...(11.63)

i.e. hinternal = 5.182 ´ 103 W/(m2C)


Then, rate of heat transfer:
Q := hinternal ×(p × D×L)×(Tsat – Ts) W
i.e. Q = 2.605 ´ 10 3 W
Rate of condensation:
Q
mcond = kg/s
h¢f g

3
where, h ¢ f g := h fg + × CpL × (Tsat – Ts)
8
i.e. h fg = 1.135 ´ 106 J/kg (modified latent heat)
Q
Then, mcond :=
h ¢f g
i.e. mcond = 2.294 ´ 10 – 3 kg/s
Rate of input of ammonia:
F p × D I kg/s
2
min := GV × GH 4 JK
i.e. min = 4.021 ´ 10 – 3 kg/s
Therefore, fraction of vapour condensed:
mcond
Fraction := × 100
min
i.e. Fraction = 57.042 %
i.e. 57.042% of the ammonia entering the tube is condensed over a length of 0.5 m.

11.4.8 Drop-wise Condensation


In drop-wise condensation, as already explained, heat transfer coefficient is higher (by as much as ten times) than
that in film condensation, due to exposure of larger surface area to the vapours. Drop-wise condensation is
achieved either by giving a coating of some material (such as teflon, gold, silver, rhodium, platinum, etc.) or by
adding some promoting chemicals (such as waxes or fatty acids) to the vapours. However, maintaining the
surface conditions favourable for drop-wise condensation is very difficult in practice, since these coatings lose
their effectiveness in course of time due to fouling, oxidation, etc. In addition, the cost consideration also comes
into picture.

574 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Condensation of steam on copper surfaces has been studied extensively, because of its industrial importance
(in Steam Power plants). P. Griffith recommends following simple correlations for drop-wise condensation of
steam on copper surfaces:
hdropwise = 51104 + 2044 × Tsat (for 22°C < Tsat < 100°C...(11.64a))
and, hdropwise = 255310 (for Tsat > 100°C...(1.64b))
It must be appreciated that in heat exchangers, it is the total thermal resistance to heat transfer that matters.
Heat transfer coefficients are inherently high in boiling and condensation; so, the associated thermal resistance [=
1/(h.A)], is quite low. So, the heat transfer is more likely to be affected by other resistances such as the film
resistance on the other side of the heat exchanger or the thermal resistance of the tube material. As such, it is not
necessary to have an extremely high accuracy in the determination of heat transfer coefficients in boiling and
condensation.

11.5 Summary
In this chapter, we studied two important phenomena that involve phase change, namely, boiling and condensa-
tion, which have great industrial importance.
When a liquid comes in contact with a heated surface maintained at a temperature Ts greater than the satu-
rated temperature Tsat of the liquid at the given pressure, boiling occurs. Boiling phenomenon may be classified
into two types: (a) pool boiling, and (b) flow boiling. In pool boiling, liquid remains stationary as in a ‘pool’ but
for the convection currents caused by temperature differences and the heater is submerged in the liquid. In flow
boiling, boiling occurs as the liquid flows past a heated surface or inside a heated tube.
In pool boiling, there are various boiling regimes depending upon the ‘excess temperature’, DTe (= Ts – Tsat).
Since it is difficult to get analytical heat transfer correlations for boiling phenomenon, practical correlations are of
empirical nature and such correlations for the different boiling regimes have been presented.
Flow boiling inside pipes is more complicated, because of the possibility of two-phase flow. Two of the
practically important correlations for two-phase flow boiling have been presented along with those for nucleate
flow boiling.
Condensation is of important relevance in steam condensers of thermal power plants. Condensation can be
considered as the ‘reverse’ of boiling phenomenon. Again, condensation may be of two types: (a) film condensa-
tion, and (b) drop-wise condensation.
Film condensation may be of laminar or turbulent type. Nusselt’s relation for laminar film condensation on
a vertical plate was derived. Several correlations for laminar film condensation on some other practically impor-
tant geometries were presented. Concept of ‘condensation number’ and its use in correlations for laminar and
turbulent film condensation correlations was explained. Some important data to quickly estimate the heat trans-
fer coefficients for condensation of steam in a condenser were also given. Finally, for drop-wise condensation,
simple correlations of Griffith for steam condensing in a drop-wise manner on copper surfaces were given.

Questions
1. Differentiate between:
(a) evaporation and boiling
(b) sub-cooled boiling and saturated boiling, and
(c) pool boiling and flow boiling.
2. With the aid of a neat sketch of a boiling curve for water (for pool boiling), explain the various regimes of
boiling. [M.U.]
3. Differentiate between nucleate boiling and film boiling. [M.U.]
4. Draw the boiling curve for pool boiling of water and show the ‘burnout point’. What is its significance?
5. What are the different flow regimes occurring in a vertical tube during flow boiling?
6. Mention a few industrial applications where boiling and condensation are important.
7. Differentiate between film condensation and drop-wise condensation. In which case is the heat transfer higher?
Why?
8. What is the mechanism of heat transfer in condensation? Using Nusselt’s theory, develop an expression for
average heat transfer coefficient in condensation over a length of a vertical plate. [M.U.]
9. What are the important assumptions in Nusselt’s theory of condensation on a vertical plate?
10. What is ‘modified latent heat of vaporisation’? What is the purpose of introducing this quantity?
11. Define ‘film—Reynolds number’. At what value of film Reynolds number does the condensate film flow become
turbulent?
BOILING AND CONDENSATION 575
12. Define ‘condensation number’. Draw a qualitative graph of film Reynolds number vs. condensation number,
and identify the laminar, wavy and turbulent film flow regions.
13. While calculating the film–Reynolds number, how is ‘wetted perimeter’ (P) defined?
14. Show that the value of hydraulic diameter for film condensation on a vertical plate, vertical cylinder and a
horizontal cylinder is given by: Dh = 4.d , where d is the film thickness.
15. Considering the film condensation on horizontal and vertical tubes, in which case is the heat transfer coefficient
higher? Explain.
16. How is the average heat transfer coefficient in a vertical tier of N horizontal tubes related to the heat transfer
coefficient for condensation on a single horizontal tube? In a given tier, where is the heat transfer maximum, for
the upper tube or the tubes at the lower levels? Why?
17. Write a short note on how the condensation heat transfer is affected by:
(a) nature of surface (smooth or rough)
(b) vapour velocity, and
(c) presence of non-condensable gases in the vapours.

Problems
1. Water is boiled at atmospheric pressure in a mechanically polished, stainless steel pan of 30 cm diameter. Bot-
tom surface of the pan is maintained at a temperature of 108°C. Calculate: (a) heat transfer coefficient (b) heat
transfer rate, and (c) rate of evaporation of water.
2. In Problem 1, (a) calculate the maximum heat flux and compare the nucleate boiling flux with the maximum
heat flux (b) compare the values of heat transfer coefficient obtained from Rohsenow’s correlation with those
obtained using Collier’s, Mostinski’s and Russian correlations.
3. In problem 1 if the stainless steel pan is Teflon coated, how do the heat transfer coefficient and rate of evapora-
tion change?
4. A nickel wire, 1.5 mm diameter and 400 mm long, is submerged in a water bath open to atmosphere. What is the
value of current flowing through the wire that will cause burnout, if the applied voltage is 10 V?
5. A horizontal, metal-clad heating element, 8 mm diameter, and of surface emissivity 0.9, is submerged in a water
bath. Surface temperature of the heating element is 340°C. If the water is at atmospheric pressure, calculate the
power dissipation per unit length of the heater.
6. A large, horizontal plate is kept immersed in a water bath boiling at 1 atm, 100°C. Surface temperature of the
plate is 300°C. Calculate the heat transfer coefficient and the heat flux. Assume the emissivity of the surface as
0.85.
7. Water is boiling at 7 atm. on the surface of a horizontal tube, whose wall temperature is maintained at 9°C
above the boiling point of water. Calculate the nucleate boiling heat transfer coefficient.
(b) What is the change in the value of heat transfer coefficient when (i) temperature difference is increased to
18°C at the pressure of 7 atm., and (ii) pressure is raised to 14 atm. with DTe = 9°C?
8. Water at 6 atm. flows inside a tube of 2.0 cm diameter under flow boiling conditions. Tube wall temperature is
maintained at 7°C above the saturation temperature. Determine the heat transfer for one metre length of tube.
9. A 50 mm diameter vertical evaporator tube (kw = 20 W/(mK)) carries 1 kg/s of steam at 11.71 bar (Psat = 460 K)
at a quality x = 0.2. The tube is subjected to a uniform heat flux of 1.1 ´ 106 W/m 2. Identify the regime of flow
boiling and calculate the convective heat transfer coefficient and surface temperature of the tube.
(b) When the quality reaches 0.8, what is the boiling regime and how much is the boiling heat transfer coeffi-
cient?
10. In Problem 9, if the tube surface is maintained at a constant temperature of 227°C, calculate the total heat
transfer coefficient and surface heat flux at the point where the quality is 0.2. Rest of the data are the same as
in Problem 9.
11. Saturated water at 1 atm. and velocity 1.5 m/s moves across a cylindrical heating element of 5 mm diameter.
Find out the maximum heating rate for the element (W/m). [Hint: use the correlation of Lienhard and Eichhorn,
Eq.11.21 or 11.22].
12. Saturated steam at atmospheric pressure condenses on a vertical plate (size: 30 cm x 30 cm) maintained at 80°C.
Determine heat transfer rate and the mass of steam condensed per hour. (b) If the plate is tilted at an angle of
30 deg. to the vertical, what is the value of condensation rate?
13. Saturated steam at a temperature of 75°C condenses on a vertical surface at 65°C. Determine the thickness of the
condensate film at locations 0.2 m and 1.0 m from top. Also, calculate the condensate flow rate, local and aver-
age heat transfer coefficients at these locations.
Properties of water at the mean temperature of 70°C are: rL = 977.5kg/m3, kL = 0.663 W/(mK), m L = 0.000404
kg/(ms), CpL = 4190 J/(kgC). And, hf g at saturation temperature of 75°C = 2321 kJ/kg. Take rV << rL.

576 FUNDAMENTALS OF HEAT AND MASS TRANSFER


14. In Problem 11.13, if the height of the plate is 1 m, calculate the following:
(i) heat transfer rate to the plate,
(ii) maximum velocity of condensate at the trailing edge, and
(iii) also, draw the variation of d with distance from top.
15. Dry, saturated steam at atmospheric pressure condenses on a horizontal tube of diameter = 3 cm and height L =
1 m; surface of the tube is maintained at 80°C. Estimate the heat transfer rate and the mass of steam condensed
per hour. Assume laminar film condensation.
(b) If the tube is vertical, what is the condensation rate?
[ Hint: Use Eq. 11.62a].
16. Dry, saturated steam at atmospheric pressure condenses on a vertical tube of diameter = 5 cm and length L = 1.5
m; surface of the tube is maintained at 60°C. Determine the heat transfer rate and the mass of steam condensed
per hour.
17. A steam condenser consists of a square array of 625 horizontal tubes, each 6 mm in diameter. The tubes are
exposed to exhaust steam arriving from the turbine at a pressure of 0.15 bar. If the tube surface temperature is
maintained at a temperature of 25°C by circulating cold water through the tubes, determine the heat transfer
coefficient and the rate at which the steam is condensed per unit length of tubes for the entire array. Assume
laminar film condensation and that there are no condensable gases mixed with steam.
18. A steam condenser consists of an array of horizontal tubes, each 2.0 cm in diameter and 1.5 m long. The tubes
are arranged in such a manner that each vertical tier has 10 tubes; tubes are exposed to saturated steam at 100°C.
If the tube surface temperature is maintained at a temperature of 80°C, determine the total number of tubes
required to get a condensation rate of 0.4 kg/s. Assume laminar film condensation and that there are no conden-
sable gases mixed with steam.
19. Ammonia at 40°C is condensing inside a horizontal tube of 25 mm ID. Mass velocity of ammonia vapour at inlet
is 10 kg/(m2s). Surface of the tube is maintained at a constant temperature of 20°C by circulating cold water.
Calculate the fraction of vapour that will condense if the tube is 0.5 m long.
20. In Problem 19, if ammonia is condensing on the outside surface of the tubes, what will be the heat transfer rate?

BOILING AND CONDENSATION 577


CHAPTER

12
Heat Exchangers

12.1 Introduction
‘Heat exchanger’ is one of the most commonly used process equipments in industry and research. Function of a
heat exchanger is to transfer energy; this transfer of energy may occur to a single fluid (as in the case of a boiler
where heat is transferred to water) or between two fluids that are at different temperatures (as in the case of an
automobile radiator where heat is transferred from hot water to air). In some cases, there are more than two
streams of fluid exchanging heat in a heat exchanger. Heat exchangers of several designs in a variety of sizes
varying from ‘miniature’ to ‘huge’ (with heat transfer areas of the order of 5000 to 10,000 sq. metres) have been
developed over the years.
Some typical examples of heat exchanger applications are:
(i) Thermal power plants (boilers, superheaters, steam condensers, etc.)
(ii) Refrigeration and air-conditioning (evaporators, condensers, coolers)
(iii) Automobile industry (radiators, all engine cooling and fuel cooling arrangements)
(iv) Chemical process industry (variety of heat exchangers between different types of fluids, in cumbustors
and reactors)
(v) Cryogenic industry (condenser–reboilers used in distillation columns, evaporators to produce gas from
cryogenic liquids, etc.)
(vi) Research (‘regenerators’ used in Stirling engines, special ceramic heat exchangers used in ultra-low tem-
perature devices, superconducting magnet systems, etc.).

12.2 Types of Heat Exchangers


Heat exchangers may be classified in several ways:
(i) according to heat exchange process
(ii) according to relative direction of flow of hot and cold fluids
(iii) according to constructional features, compactness, etc.
(iv) according to the state of the fluid in the heat exchanger.
(i) Classification according to heat exchange process Heat exchangers may be of ‘direct contact type’ or of
‘indirect contact type’. In direct contact type, two immiscible fluids come in direct contact with each other and
exchange heat, e.g. air and water exchanging heat in a cooling tower. Indirect contact type can be further classi-
fied as ‘recuperators’ and ‘regenerators’. Recuperators are most commonly used; here, the hot and cold fluids are
separated from each other by a solid wall and heat is transferred from one fluid to the other across this wall. In
regenerators, also called ‘periodic flow heat exchangers’, hot and cold streams alternately flow through a solid
matrix (made of solid particles or wire mesh screens); during the ‘hot blow’, the matrix stores the heat given up
by the hot stream and during ‘cold blow’, the stored heat is given up by the solid matrix to the cold stream.
Sometimes, the solid matrix is made to rotate across fluid passages arranged side by side, so that the heat ex-
change process is ‘continuous’.
(ii) Classification according to relative direction of hot and cold fluids If the hot and cold fluids flow parallel to
each other, it is known as ‘parallel flow’ heat exchanger; if the two fluids flow opposite to each other, it is of
Cold Temperature
fluid
Th1

Th2
Hot
fluid Tc2

Tc1

Length

FIGURE 12.1(a) Parallel flow heat exchanger

Cold
fluid Temperature
Th1

Hot Tc2 Th2


fluid
Tc1

Length

FIGURE 12.1(b) Counter-flow flow heat exchanger

‘counter-flow’ type. If the fluids flow perpendicular to each other, then, we have ‘cross flow’ type of heat ex-
changer. These three types of heat exchangers are shown schematically in Fig. 12.1.
Further, when a fluid is constrained to flow within a channel (such as a tube), the fluid is said to be ‘un-
mixed’; otherwise, it is ‘mixed’. In Fig. 12.1 (c), hot fluid is unmixed since it flows constrained within the tubes,
whereas the cold fluid is perfectly mixed as it flows through the heat exchanger. In Fig. 12.1 (d), both the cold
and hot fluids are constrained to flow within the tubes and therefore, both the fluids are unmixed.
(iii) Classification according to constructional features Basically, there are three types: (a) concentric tubes type
(b) shell and tube type, and (c) compact heat exchangers.
In concentric tubes type of heat exchanger, one tube is located inside another; one fluid flows through the
inside tube and the other fluid flows in the annular space between the tubes. Fluids may flow parallel to each
other as shown in Fig. 12.1 (a), or they may flow in opposite directions, as shown in Fig. 12.1 (b).
Shell and tube type of heat exchanger is very popular in industry because of its reliability and high heat
transfer effectiveness. Here, one of the fluids flows within a bundle of tubes placed within a shell. And, the other
fluid flows through the shell over the surfaces of the tubes. Suitable baffles are provided within the shell to make
the shell fluid change directions and provide good turbulence, so that heat transfer coefficient is increased.
Fig. 12.2 shows a schematic diagram of a typical shell and tube heat exchanger.
Fig. 12.2 is an example of two tube pass and one shell pass heat exchanger, i.e. flow passes through the tubes
twice in opposite directions, and shell fluid passes through the shell once. Other flow arrangements are also
used, such as: one shell pass + two, four or six tube passes; two shell passes and four, eight, twelve, etc. tube
passes.
Compact heat exchangers are special purpose heat exchangers which provide very high surface area per
cubic metre of volume, known as ‘area density’. According to usually accepted norms, a ‘compact heat ex-
changer’ has an area density of 700 m2/m3 or more. These are generally used for gases, since usually gas side
heat transfer coefficient is small and therefore, it is needed to provide larger areas. Compact heat exchangers are
of plate-fin type or tube-fin type. A typical example of a plate-fin type of compact heat exchanger is shown in Fig.
12.3.

HEAT EXCHANGERS 579


Cold fluid Cold fluid

Hot
Hot
fluid
fluid

FIGURE 12.1(c) Cross-flow heat exchanger, cold FIGURE 12.1(d) Cross-flow heat exchanger, both
fluid ‘mixed’, hot fluid ‘unmixed’ cold and hot fluids ‘unmixed’

Shell and shell-side nozzles


Channel cover
Tube sheet Tubes
Tube-side channel

Baffles
Tube sheet
Pass divider
Channel cover
Tube-side channel
and nozzles

FIGURE 12.2 Diagram of a typical (fixed tube sheet) shell-and-tube heat exchanger

Fluid 2 (iv) Classification according to state of the fluid In all


Matrix 2
the types of heat exchangers discussed above, both the
fluids changed their temperature along the length of
heat exchanger. But, this need not be the case always. A
heat exchanger may be used to condense a fluid in
which case the condensing fluid will be at a constant
temperature throughout the length of the heat ex-
changer, while the other (cold) fluid will increase in
temperature as it passes through the heat exchanger,
absorbing the latent heat of condensation released by
Fluid 1 the condensing fluid. Such a heat exchanger is called a
‘Condenser’. If, on the other hand, one of the fluids
Parting Sheet evaporates in a heat exchanger, tempera-ture of this
fluid will remain constant throughout the length of
heat exchanger, whereas the temperature of the other
Matrix 1 fluid, which supplies the latent heat of evaporation to
the evaporating fluid, goes on decreasing along the
length of the heat exchanger. Such a heat exchanger is
FIGURE 12.3 Section of a plate-fin called an ‘Evaporator’.
heat exchanger It is interesting to compare the surface area-to-vol-
ume ratios of different types of heat exchangers. See
Table 12.1:
580 FUNDAMENTALS OF HEAT AND MASS TRANSFER
TABLE 12.1 Surface area-to-volume ratios of different heat exchangers
Type of HX Hydraulic diameter (mm) Surface-area/Volume, (m2/m 3)
Plain tube, shell-and-tube 40 to 6 60 – 600
Plate heat exchangers 20 to 10 180 – 350
Strip fin and louvred fin
heat exchangers 10 to 0.5 350 – 7100
Automotive radiators 5 to 2.5 710 – 1500
Cryogenic-heat exchangers 3.7 to 1.7 1000 – 2500
Gas turbine rotary regenerators 1.2 to 0.5 3000 – 7100
Matrix types, wire screen, sphere 2.5 to 0.2 1500 – 18000
bed, corrugated sheets
Human lungs 0.2 to 0.15 18000 – 25000

In Table 12.1, hydraulic diameter of the flow passage is also given; note that smaller the hydraulic diameter,
larger is the ratio of surface area-to-volume. Note that the human lungs have the largest of all surface area-to-
volume ratios.

12.3 Overall Heat Transfer Coefficient


‘Overall heat transfer coefficient’, was first introduced in Chapter 4. So, the reader may please refer back to
Chapter 4 to refresh memory.
In most of the practical cases of heat exchangers, temperature of the hot fluid (Ta) and that of the cold fluid
(Tb) are known; then we would like to have the heat transfer given by a simple relation of the form
Q = U×A×(Ta – Tb) = U×A×DT ...(4.21)
where, Q is the heat transfer rate (W), A is the area of heat transfer perpendicular to the direction of heat transfer,
and (Ta – Tb) = DT is the overall temperature difference between the temperature of hot fluid (Ta) and that of the
cold fluid (Tb).
In a normally used recuperative type of heat exchanger, the hot and cold fluids are separated by a solid wall.
This may be a flat type of wall (as in the case of plate-fin type of heat exchangers), or, more often, a cylindrical
wall (as in the case of a tube-in-tube type of heat exchangers). See Fig. 12.4.
Recall from Chapter 4 that, in general, the overall heat transfer coefficient is related to the total thermal
resistance of the system, as follows:
1
U= W/(m2 C). ...(4.23)
A ×S Rth
Therefore, the task of finding the overall heat transfer coefficient reduces to finding out the total thermal
resistance of the system.
For plane wall:
Remember that for a plane wall, thermal resistance is L/(k.A), and convective resistance is 1/(h.A), and since
the resistances are in series, we get:

T1 T2 k

1 To, ho
To, ho
Q Q
Ti, hi
Ti, hi
k
ri
L ro
X
(a) Plane wall (b) Cylindrical wall

FIGURE 12.4 Heat exchanger walls


HEAT EXCHANGERS 581
1
U=
F 1 + L + 1 I
A ×G
H h × A k × A h × A JK
i o

1
i.e. U= W/(m2C). ...(12.1)
1 L 1
+ +
hi k ho
Now, if the thermal resistance of the wall is negligible compared to other resistances, we get:
1
U= W/(m2 C). ...(12.2)
1 1
+
hi ho
For cylindrical wall:
Remember that for a cylindrical wall, thermal resistance is:

Fr I
ln GH r JK
o
i
2 ×p × k × L
and, convective resistance is 1/(h.A) and the resistances are in series. However, the area to be considered has to
be specified since the inner surface area and the outer surface area of the cylinder are different. Now, we have,
the general relation for U:
1
U= W/(m2 C). ...(4.23)
A ×S Rth
1
i.e. U×A =
S Rth
We can also write:
1
Ui ×Ai = Uo×Ao = ...(12.3)
S Rth
Therefore, referred to outer surface area, U becomes:
1
Uo×Ao =
Fr I
ln G J o
...(12.4)

1
+ i HrK
+
1
hi × Ai 2 ×p × k × L ho × Ao
Now, for a cylindrical system, we have:
Ai = 2×p×ri×L
and, Ao = 2×p×ro ×L
1
Then, Uo =
Fr I
ln G J o
Ao
+ i H r K ×A o +
1
× Ao
hi × Ai 2 ×p × k × L ho × Ao

1
i.e. Uo =
1 ro F I FG IJ F I
r r
...(12.5)
×
hi ri GH JK H K GH JK
+ o × ln o +
k ri
1
ho
Similarly, referred to inner surface area, U becomes:

582 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1
Ui ×Ai =
Fr I
ln G J o
...(12.6)

1
+
Hr K
i
+
1
hi × Ai 2 ×p × k × L ho × Ao
and,
1
Ui =
Fr I
ln G J o
1
+
H r K ×A +
i
i
1
× Ai
hi 2 ×p × k × L ho × Ao

1
1 Fr I Fr I 1 FrI
i.e. Ui = ...(12.7)
+ G J × ln G J +
h H k K H r K h GH r JK
i × o i
i i o o
Again, if the thermal resistance of the wall is negligible compared to other resistances, (i.e. high value of
thermal conductivity, k), or, wall thickness of the tube is very small (i.e. (ri/ro) » 1), we get:
1
U= W/(m2 C) ...(12.8)
1 1
+
hi ho
Note that Eq. 12.8 is the same as Eq. 12.2. For many practical situations, this simple equation gives a quick
estimate of overall heat transfer coefficient, U. Observe from Eq. 12.2 or 12.8 that the value of U is controlled by
the smaller of the two heat transfer coefficients, hi and ho. Therefore, aim of the designer should be to focus on the
smaller of the two heat transfer coefficients and improve it, if possible. For example, in a gas-to-liquid heat
exchanger, heat transfer coefficient is generally smaller on the gas side, and, therefore, the gas side heat transfer
coefficient controls the final value of overall heat transfer coefficient. So, one tries to improve the heat transfer
coefficient on the gas side by providing fins on the gas side surface. If fins are provided on a particular surface,
then the total heat transfer area on that surface is:
A total = A fin + A unfinned ...(12.9)
where, A fin is the surface area of the fins and A unfinned is the area of the un-finned portion of the tube.
For short fins of a material of high thermal conductivity, since there is practically no temperature drop along
the length we can use the value of total area as given by Eq. 12.9 to calculate the convection resistance on the
finned surface. However, for long fins where there is a temperature drop along the length of fin, we should use
the total or effective area, given by:
Atotal = Aunfinned + h fin ×Afin ...(12.10)
where, h fin is the ‘fin efficiency’. Sometimes, an overall surface efficiency’ ho is used. h o is defined as:
ho ×A total = A unfinned + h fin ×A fin
i.e. h o tells us how much of the total surface area is really effective in transferring heat.
Then, since the effective surface area is also equal to the unfinned area plus the effective area of fin, we can
get an expression for overall surface efficiency as follows:
ho ×A total = (A total – A fin) + h fin ×Afin
Afin h ×A
i.e. ho = 1 – + fin fin
Atotal Atotal
Afin
i.e. ho = 1 – ×(1 – h fin). ...(12.11)
Atotal
Then, while determining U, we should use h o.Atotal for the finned surface, whether it is inner surface area,
outer surface area or both.
For example, if the outer surface of the tube is finned (which is usually the case), with a fin efficiency of h fin,
we write, neglecting the thermal resistance of tube material:

HEAT EXCHANGERS 583


1 1
Ui ×Ai = Uo ×Ao = = ...(12.12)
S Rth 1 1
+
hi × Ai ho × ( Aunfinned + h fin × Afin )outer_surface
Instead, if the total (i.e. unfinned + finned) surface area and the overall surface efficiency (ho) is given for the
outer surface, Eq. 12.12 can be written as:
1 1
Ui ×Ai = Uo ×Ao = = ...(12.13)
S Rth 1 1
+
hi × Ai ho × (h o × Atotal )outer_surface
Typical values of overall heat transfer coefficients are given in Table 12.2:
TABLE 12.2 Typical values of overall heat transfer coefficients
Type of HX U (W/(m 2C)
Water-to-water 850 – 1700
Water-to-oil 100 – 350
Water-to-gasoline or kerosene 300 – 1000
Feed water heaters 1000 – 8500
Steam-to-light fuel oil 200 – 400
Steam-to-heavy fuel oil 50 – 200
Steam condenser 1000 – 6000
Freon condenser (water cooled) 300 – 1000
Ammonia condenser (water cooled) 800 – 1400
Alcohol condenser (water cooled) 250 – 700
Gas-to-gas 10 – 40
Water-to-air in finned tubes 30 – 60 (based on
(water in tubes) water side surface area)
Steam-to-air in finned tubes 400 – 4000 (based on
(steam in tubes) steam side surface area)

Fouling factors Note that above analysis was for clean heat transfer surfaces. However, with passage of time,
the surfaces become ‘dirty’ because of scaling, deposits, corrosion, etc. This results in a reduction in heat transfer
coefficient since the scale offers a thermal resistance to heat transfer. Fouling may be categorized as follows:
(i) due to scaling or precipitation
(ii) due to deposits of finely divided particulates
(iii) due to chemical reaction
(iv) due to corrosion
(v) due to attachments of algae or other biological materials
(vi) due to crystallization on the surface by subcooling.
Effect of fouling is accounted for by a term called, ‘Fouling factor’, (or, ‘dirt factor’), defined as:
1 1
Rf = - m2 K/W ...(12.14)
U dirty Uclean
Rf is zero for a new heat exchanger. Rf for a fouled heat exchanger cannot be ‘calculated’ theoretically, but
has to be determined experimentally by finding out the heat transfer coefficients for a ‘clean’ heat exchanger and
a ‘dirty’ heat exchanger of identical design, operating under identical conditions.
While taking into account the effect of fouling, the ‘fouling resistance’ (= Rf/area) should be added to the
other thermal resistances. For example, for a tube, we can write:
1 1
Ui ×Ai = Uo ×Ao =
S Rth
=
FG r IJ
o
...(12.15)

1
+
R fi
+
ln
HrK
i
+
1
+
R fo
hi × Ai Ai 2 ×p × k × L ho × Ao Ao
584 FUNDAMENTALS OF HEAT AND MASS TRANSFER
where, Rfi and Rfo are the fouling factors for the inside and outside surfaces, respectively, and L is the length of
tube. From Eq. 12.15, Ui or Uo can easily be calculated.
Fouling factor depends on flow velocity and operating temperature; fouling increases with decreasing veloc-
ity and increasing temperature.
Based on experience, Tubular Exchanger Manufacturers’ Association (TEMA) have given suggested values
of fouling factors. Some of these values are given in Table 12.3:

TABLE 12.3 Fouling factors for industrial fluids (TEMA, 1988)

Fluid Rf (m2C/W)
LIQUIDS:
Fuel oil 0.00088
Quench oil 0.0007
Transformer oil 0.00018
Hydraulic fluid 0.000238
Molten salts 0.000119
Industrial organic heat transfer media 0.000119
Refrigerant liquids 0.00018
Caustic solutions 0.000476
Vegetable oils 0.000715
Gasoline, naptha, light distillates, kerosene 0.000238
Light gas oil 0.000476
Heavy gas oil 0.000715
GASES & VAPOURS:
Solvent vapours 0.000238
Acid gases 0.000238
Natural gas 0.000238
Air 0.000119 – 0.000238
Flue gases 0.000238 – 0.000715
Steam (sat., oil free) 0.000119 – 0.000357
WATER:
River water, sea water, distilled water, boiler feed water:
Below 50 deg.C 0.0001
Above 50 deg.C 0.0002

Example 12.1. Water at a mean temperature of Tm = 90°C and a mean velocity of um = 0.10 m/s flows inside a 2.5 cm ID,
thin-walled copper tube. Outer surface of the tube dissipates heat to atmospheric air at Ta = 20°C, by free convection.
Calculate the tube wall temperature, overall heat transfer coefficient and heat loss per metre length of tube. Use follow-
ing simplified expression for air to determine heat transfer coefficient by free convection:
FG T - T IJ 0 . 25

H D K
s a
h a = 1.32×

Solution.
Data:
Tm := 90°C um := 0.1 m/s D := 0.025 m Ta := 20°C
Properties of water at mean temperature of 90°C:
r := 965.3 kg/m3 k := 0.675 W/(mC) m := 0.315 ´ 10– 3 kg/(ms) Pr := 2.22
We need to calculate the heat transfer coefficients for the inner and outer surfaces:
For the water side (i.e. inner surface):
D × um × r
We have: Re := (Reynolds number)
m
i.e. Re = 7.66 ´ 10 3 > 4000 (therefore, turbulent)

HEAT EXCHANGERS 585


Using Ditus–Boelter equation to determine heat transfer coefficient for inside surface:
Nu := 0.023×Re 0.8 ×Pr 0.3
i.e. Nu = 37.417 (Nusselts number)
k
Therefore, h i := Nu ×
D
i.e. h i = 1.01 ´ 103 W/(m2 C) (inside surface heat transfer coefficient)
For the air side (i.e. outer surface):
Approximate, value of film temperature for air:
90 + 20
Tf :=
2
i.e. Tf = 55°C
Properties of air at film temperature of 55°C:
r := 1.076 kg/m 3 k := 0.0283 W/(mC) m := 1.99 ´ 10 –5 kg/(ms) Pr := 0.708
Then, free convection heat transfer coefficient for outer surface is given by:
FG T - T IJ 0 . 25

H D K
s a
ho = 1.32×

F T - 20 I 0 . 25

= 1.32× G
H 0.025 JK
s
i.e. ho

i.e. ho = 3.32 ×(Ts – 20)0.25 ...(a)


However, Ts is not known,
Applying overall energy balance, with Ai = Ao for thin-walled tube:
h i ×(Ti – Ts) = ho ×(Ts – Ta)
Substituting for hi and ho
1010 ×(90 – Ts) = [3.32×(Ts – 20)0.25]×(Ts – 20)
i.e. 1010×(90 – Ts) = [3.32 ×(Ts – 20)1.25]
This equation may now be solved for Ts by trial and error.
But, with Mathcad, it is easily solved using solve block. Start with a guess value of Ts, then, after typing ‘Given’
write down the constraint, and then, typing ‘Find (Ts)’ gives the value of Ts immediately:
Ts := 40°C (guess value)
Given
1010×(90 – Ts) = [3.32×(Ts – 20)1.25]
Find (Ts) := 89.342
i.e. Ts := 89.342°C (tube surface temperature)
Then, ho is calculated from Eq. a:
h o := 3.32 ×(Ts – 20)0.25 ...(a)
i.e. h o = 9.58 W/(m2 C) (outside surface heat transfer coefficient)
and, overall heat heat transfer coefficient, U:
1
U :=
1 1
+
hi ho
i.e. U = 9.49 W(m2C) (overall heat transfer coefficient)
Note that overall heat transfer coefficient is nearly equal to ho. As commented earlier, since hi >> ho, overall heat
transfer coefficient is conrolled by ho.
Heat loss per metre length of tube:
Q := U×(p×D×1)×(Ts – Ta) W/m
i.e. Q = 51.686 W/m.
Example 12.2. In Exaple 12.1, if we desire to increase the value of overall heat transfer coefficient U, the obvious choice
is to focus on the air-side, since the air side heat transfer coefficient is the lower of the inside and outside heat transfer
coefficients. Let us increase the area on the air side by providing 8 numbers of radial fins of rectangular cross section, 2
mm thick and 20 mm height. Material of fins is the same as that of the tube, i.e. copper (k = 380 W/(mK)).
Then, determine the overall heat transfer coefficient and the rate of heat transfer.
586 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Solution.
Data:
Tm := 90°C um := 0.1 m/s D := 0.025 m Ta := 20°C Ts := 89.342°C hi := 1010 W/(m2 C)
ho := 9.58 W/(m2C) L := 0.02 m (fin height) W := 1 m (fin width = length of cyl.)
t := 0.002 m (fin thickness) N := 8 (No. of fins) k := 380 W/(mK)
Ai := p×D×1 m2/metre i.e. A = 0.079 m2 /metre
Overall heat transfer coefficient Ui referred to the inside surface:
Neglecting the thermal resistance of tube wall, we write:
1 1
Ui ×Ai = = ...(a)
S Rth 1 1
+
hi × Ai ho × ( Aunfinned + h fin × Afins )
First term in the denominator in RHS is the thermal resistance due to film coefficient on the inside, and the second
term is the thermal resistance of the film coefficient on the outside.
Un-finned surface (or the base surface) on the outside is at the wall temperature and is fully effective for heat
transfer whereas the finned surface is not fully effective because of temperature drop along the length of fins; therefore,
effective area of fins is obtained by multiplying the total area of fins by the fin effectiveness, h fin.
Therefore, we need to find out the fin efficiency.
Fin efficiency:
For a rectangular fin with adiabatic tip, the fin efficiency is given by:
tan h ( m × L)
h fin = ...(b)
m×L

ho × P
where, m= 1/m (fin parameter)
k × Ac
P = 2×(W + t) (perimeter, W = width of fin = 1 m)
Ac = W×t (area of cross section of fin)
P 2× (W + t) 2
Then, = = (for t << W)
Ac W ×t t

2× ho
Therefore, m :=
k×t
i.e. m = 5.021 1/m
and, m×L = 0.1
Then, from Eq. b, we get:
tan h (m × L)
h fin :=
m× L
i.e. h fin = 0.997 (fin efficiency)
Areas:
Aunfinned := p × (D – N×t)×W
i.e. Aunfinned = 0.028 m2 (unfinned or prime (base) area)
Afins := N×(2 ×W×L) m2 (finned area of N fins (both upper and lower side of fins considered)
i.e. Afins = 0.32 m2
Now, Ai := p×D×W m2 (inside surface area per metre length)
i.e. Ai = 0.079 m2
Therefore, overall heat transfer coefficient Ui referred to the inside surface:
1 1
We have: Ui ×Ai = = ...(a)
S Rth 1 1
+
hi × Ai ho × ( Aunfinned + h fin × Afins )
i.e. Ui ×Ai = 3.192
3.192
and, Ui :=
Ai
i.e. Ui = 40.642 W/(m2 C) (overall heat transfer coefficient referred to inside area)

HEAT EXCHANGERS 587


Note: compare this to the earlier U value of 9.49 W/(m2C); there is great improvement in value of U by providing fins.
Heat loss per metre length of tube:
Q := Ui×(p×D×l)×(Ts – Ta) W/m
i.e. Q = 221.34 W/m.
Note: Compare this to the earlier Q value of 51.686 W/m; this substantial improvement in value of Q is the result of
providing fins.
Also, note that in the above analysis, we assumed that the outside heat transfer coefficient ho is the same for the un-
finned surface as well as for the finned surfaces.
Example 12.3. A shell and tube counter-flow heat exchanger uses copper tubes (k = 380 W/(mC)), 20 mm ID and 23 mm
OD. Inside and outside film coefficients are 5000 and 1500 W/(m2 C), respectively. Fouling factors on the inside and
outside may be taken as 0.0004 and 0.001 m2C/W respectively. Calculate the overall heat transfer coefficient based on: (i)
outside surface, and (ii) inside surface.
Solution.
Data:
Di := 0.020 m Do := 0.023 m L := 1 m k := 380 W/(mK) hi := 5000 W/(m 2C) ho := 1500 W/(m2 C)
2 2
Rfi := 0.0004 m C/W (fouling factor on inside surface) Rfo := 0.001 m C/W (fouling factor on outside surface)
Heat transfer areas:
Ao := p×Do ×L m2/metre (inside surface area)
i.e. Ao = 0.07226 m2/metre
and, Ai := p×Di ×L m2/metre (inside surface area)
i.e. Ai = 0.06283 m2/metre
Overall heat transfer coefficient:
We have:
1 1
Ui ×Ai = Uo ×Ao =
S Rth
=
F D IJ
ln G o
...(12.15)

1
+
R fi
+
HD K +
i 1
+
R fo
hi × Ai Ai 2 ×p × k × L h o × Ao Ao
In the denominator of RHS of Eq. 12.15 above, we have the various thermal resistances, as follows:
first term ® convective film resistance on the inside surface = 3.1831 ´ 10 –3 C/W
second term ® fouling resistance on the inside surface = 6.3662 ´ 10 –3 C/W
third term ® conductive resistance of the tube wall = 5.854 ´ 10 –5 C/W
fourth term ® convective film resistance on the outside surface = 9.2264 ´ 10 –3 C/W
fifth term ® fouling resistance on the outside surface = 0.01384 C/W.
Note the relative magnitude of fouling resistances, as compared to other resistances. As expected, conductive resist-
ance of the tube wall (of copper, which is a good conductor) is the smallest of all.
Calculating the RHS, we get:
1
FD I
ln G
= 30.606

H D JK +
o

1 R fi i 1 R fo
+ + +
hi × Ai Ai 2 ×p × k × L ho × Ao Ao
i.e. Ui ×Ai = Uo ×Ao = 30.606
30.606
Therefore, Ui :=
Ai
i.e. Ui = 487.11 W/(m2C) (Overall heat transfer based on inside surface.)
30.606
and, Uo :=
Ao
i.e. Uo = 423.574 W/(m2C) (Overall heat transfer coefficient based on outside surface.)
Comments:
‘Fouling’ affects the value of overall heat transfer coefficient and therefore, the size (or area) of the heat exchanger
adversely.
If the fouling resistances were not included, we should have obtained the following values for the overall heat
transfer coefficient:

588 FUNDAMENTALS OF HEAT AND MASS TRANSFER


RHS of Eq. 12.15, deleting the fouling resistances, will become:
1
F D IJ
ln G o
= 80.205

1
+
HD K +
i 1
hi × Ai 2 ×p × k × L ho × Ao
i.e. Ui ×Ai = Uo ×Ao = 80.205
80.205
Therefore, Ui :=
Ai
i.e. Ui = 1.277 ´ 10 3 W/(m 2C) (overall heat transfer coefficient based on inside surface.)
80.205
And, Uo :=
Ao
i.e. Uo = 1.11 ´ 10 3 W/(m2C) (Overall heat transfer coefficient based on outside surface.)
i.e. Uo and Ui with no fouling are about 2.62 times the corresponding values when fouling resistances are included.
Therefore, it is advisable to include the effect of fouling, if practicable, at the design stage.

12.4 The LMTD Method for Heat Exchanger Analysis


Basically, a complete design of a heat exchanger is a huge topic which involves an analysis of:
(i) Thermal aspects (i.e. temperatures of fluids at inlet/exit, rate of heat transfer, etc., off-design perform-
ance, etc.)
(ii) Hydrodynamic aspects (i.e. pressure drops in the flow channels)
(iii) Structural aspects (mechanical design and structural design).
However, here we shall consider only the thermal analysis aspects.
12.4.1 Parallel Flow Heat Exchanger
Consider a double pipe, parallel flow heat exchanger, in which a hot fluid and a cold fluid flow parallel to each
other, separated by a solid wall. Hot fluid enters at a temperature of Th1 and leaves the heat exchanger at a
temperature of Th2; cold fluid enters the heat exchanger at a temperature of Tc1 and leaves at a temperature of Tc2.
This situation is shown in Fig. 12.5.

Temperature
Cold
Th1
fluid dQ dTh

Th2 DT2
Hot
DT1
fluid
Tc2
dTc
dA
Tc1

Length

FIGURE 12.5 Parallel flow heat exchanger

We desire to get an expression for the rate of heat transfer in this heat exchanger in the following form:
Q = U×A×DTm ...(12.16)
where, U = overall heat transfer coefficient
A = area for heat transfer (should be the same area on which U is based), and
DTm = a mean temperature difference between the fluids.
Now, we make the following assumptions:
(i) U is considered as a constant throughout the length (or area) of the heat exchanger

HEAT EXCHANGERS 589


(ii) Properties of fluids (such as specific heat) are also considered to be constant with temperature
(iii) Heat exchange takes place only between the two fluids and there is no loss of heat to the surroundings,
i.e. perfect insulation of heat exchanger is assumed
(iv) Changes in potential and kinetic energy are negligible
(v) Temperatures of both the fluids remain constant (equal to their bulk temperatures) over a given cross
section of the heat exchanger.
Area ‘A’ is constant for a given heat exchanger. However, we see from Fig. 12.5 that the temperature of the
two fluids vary along the length (or area) of the heat exchanger, i.e. the temperature difference between the hot
and cold fluids is not a constant along the length of the heat exchanger, but varies along the length. Our aim is to
find out the appropriate ‘mean temperature difference (DTm)’ between the hot and cold fluids, so that Eq. 12.16
can be applied. We proceed as follows:
Consider an elemental area dA of the heat exchanger. Then, by applying the First law, we can write:
Heat given up by the hot fluid = heat received by the cold fluid.
i.e.
dQ = – mh ×Cph ×dT h = mc ×Cpc ×dTc ...(12.17)
Here, the temperature of hot fluid decreases as the length increases. So, a negative sign is put in front of
mh.Cph .dTh, so that the heat transferred is a positive quantity.
Now, dQ for the elemental area dA, can also be expressed as:
dQ = U×(Th – Tc)×dA ...(12.18)
Now, from Eq. 12.17, we have:
- dQ
dTh =
mh × C ph

dQ
and, dTc =
mc × Cpc
where, mh and mc are the mass flow rates, and Cph and Cpc are the specific heats of hot and cold fluids, respec-
tively.
Therefore,
F 1 I
dTh – dTc = d(Th – Tc) = – dQ× GH m ×Ch ph
+
1
mc × Cpc
JK ...(12.19)

Substituting for dQ from Eq. 12.18, we get:


F 1 + 1 I
d(Th – Tc) =– U×(Th – Tc)×dA× GH m × C m × C JK
h ph c pc

d (Th - Tc ) F 1 I
i.e.
(Th - Tc )
= – U×
1
m h × Cph
GH +
m ×C Kc
J ×dApc
...(12.20)

Integrating Eq. 12.20 between the inlet and exit of the heat exchanger (i.e. between conditions 1 and 2):
FT - Tc 2 I F 1 I
ln GH T
h2

h 1 - Tc 1
JK = – U×A× GH m ×C
h ph
+
1
mc × Cpc
JK ...(12.21)

Now, considering the total heat transfer rate for the entire heat exchanger, we have:
Q
m h ×Cph =
Th 1 - Th 2
and,
Q
mc×Cpc =
Tc 2 - Tc 1
Substituting in Eq. 12.21:
FT - Tc 2 I - U ×A
ln GH T
h2

h 1 - Tc 1
JK =
Q
×(Th1 – Th2 + Tc 2 – Tc1)

590 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(Th 2 - Tc 2 ) - (Th 1 - Tc 1)
Q = U×A×
i.e.
FT - Tc 2 I ...(12.22)
ln GH Th2

h 1 - Tc 1
JK
Now, comparing Eq. 12.22 and Eq. 12.16, we observe that:
(Th 2 - Tc 2 ) - (Th 1 - Tc 1)
DT m =
FT - Tc 2 I ...(12.23)
ln GH T h2

h 1 - Tc 1
JK
Since this mean temperature difference varies in a logarithmic manner, it is called ‘Logarithmic Mean Tem-
perature Difference’ or, simply LMTD.
So, we write:
(Th 2 - Tc 2 ) - (Th 1 - Tc 1)
LMTD =
FT - Tc 2 I ...(12.24)
ln GH T h2

h 1 - Tc 1
JK
Now, note that (Th2 – Tc2) is the temperature difference at the exit and (Th1 – Tc1) is the temperature differ-
ence at the inlet of the heat exchanger. If we denote the temperature differences at the inlet and exit of the heat
exchanger by DT 1 and DT 2, respectively, we can write:
D T2 - D T1 DT1 - DT2
LMTD =
F DT I =
F DT I ...(12.25)
ln G
H D T JK
2

1
ln GH DT JK
1

We can state Eq. 12.25 in words as follows: LMTD is equal to the ratio of the difference between the greater
and lower of the temperature differences at the two ends to the natural logarithm of the ratio between those
temperature differences.
Equation for LMTD is easily remembered as follows:
GTD - LTD
LMTD =
FG
GTD IJ ...(12.26)
ln
H
LTD K
where,
GTD = ‘greater (of the two) temperature difference’, and
LTD = ‘lower temperature difference’.
12.4.2 Counter-flow Heat Exchanger
Again, consider a double pipe, counter-flow heat exchanger, in which a hot fluid and a cold fluid flow in direc-
tions opposite to each other, separated by a solid wall. Hot fluid enters at a temperature of Th1 and leaves the heat
exchanger at a temperature of Th2; cold fluid enters the heat exchanger at a temperature of Tc1 and leaves at a
temperature of Tc2. This situation is shown in Fig. 12.6.
We desire to get an expression for the rate of heat transfer in this heat exchanger in the following form:
Q = U×A×DTm ...(12.16)
where,
U = overall heat transfer coefficient
A = area for heat transfer (should be the same area on which U is based), and
DTm = a mean temperature difference between the fluids.
We see from Fig. 12.6 that the temperatures of the two fluids vary along the length (or area) of the heat
exchanger, i.e. the temperature difference between the hot and cold fluids is not a constant along the length of the
heat exchanger, but varies along the length. Our aim is to find out the appropriate ‘mean temperature difference
(DTm)’ between the hot and cold fluids, so that Eq. 12.16 can be applied. We proceed as follows, with the same
assumptions as made for the analysis of parallel flow heat exchanger:
Consider an elemental area dA of the heat exchanger. Then, by applying the First law, we can write:
Heat given up by the hot fluid = heat received by the cold fluid.

HEAT EXCHANGERS 591


Temperature

Cold Th1
fluid dQ dTh

DT1
dTc
Hot
fluid
Tc2 Th2
DT2
dA Tc1

Length
FIGURE 12.6 Counter-flow heat exchanger

i.e.
dQ = –mh ×Cph ×dTh = –mc ×Cpc ×dTc ...(12.27)
Here, the temperatures of both hot and cold fluids decrease as the length increases. So, negative sign is put
in front of mh.Cph .dTh and mc.Cpc.dTc so that the heat transferred is a positive quantity.
Now, dQ for the elemental area dA, can also be expressed as:
dQ = U×(Th – Tc)×dA ...(12.28)
Now, from Eq. 12.27, we have:
- dQ
dTh =
mh × Cph

- dQ
and, dTc =
mc × Cpc
where, mh and mc are the mass flow rates, and Cph and Cpc are the specific heats of hot and cold fluids, respec-
tively.
Therefore,
F 1 I
dTh – dTc = d(Th – Tc) = – dQ× GH m ×C
h ph
-
1
mc × Cpc
JK ...(12.29)

Substituting for dQ from Eq. 12.28, we get:


F 1 - 1 I
d(Th – Tc) =–U×(Th – Tc)×dA× GH m ×C m ×C JK
h ph c pc

F 1 I
i.e.
d (Th - Tc )
(Th - Tc )
= –U×
1
GH
mh × Cph
-
m ×C K
c
J ×dApc
...(12.30)

Integrating Eq. 12.30 between the inlet and exit of the heat exchanger (i.e. between conditions 1 and 2):

FT - Tc 1 I= F 1 I
ln GH T
h2
h 1 Tc 2
- JK –U×A× GH m ×C
h ph
-
1
mc × Cpc
JK ...(12.31)

Now, considering the total heat transfer rate for the entire heat exchanger, we have:
Q
m h ×Cph =
Th 1 - Th 2

592 FUNDAMENTALS OF HEAT AND MASS TRANSFER


and,
Q
mc ×Cpc =
Tc 2 - Tc 1
Substituting in Eq. 12.31:
FT - Tc 1 I = - U × A ×(T
ln GH T
h2
h 1 - Tc 2
JK Q h1 – Th2 – Tc2 + Tc1)

(Th 2 - Tc 1) - (Th 1 - Tc 2 )
i.e. Q = U×A×
FT
ln G
- Tc 1 I ...(12.32)

HT
h2
h 1 - Tc 2
JK
Now, comparing Eq. 12.32 and Eq.12.16, we observe that:
(Th 2 - Tc 1) - (Th 1 - Tc 2 )
DT m =
FT - Tc 1 I ...(12.33)
ln GH Th2
h 1 Tc 2
- JK
Note that this mean temperature difference varies in a logarithmic manner; so, it is called ‘Logarithmic Mean
Temperature Difference’ or, simply LMTD.
So, we write:
(Th 2 - Tc 1) - (Th 1 - Tc 2 )
LMTD =
FT - Tc 1 I ...(12.34)
ln GH Th2
h 1 Tc 2
- JK
Now, note that (Th2 – Tc1) is the temperature difference at the exit and (Th1 – Tc2) is the temperature differ-
ence at the inlet of the heat exchanger. If we denote the temperature differences at the inlet and exit of the heat
exchanger by DT 1 and DT2, respectively, we can write:
D T2 - D T1 D T1 - D T2
LMTD =
F DT I
ln G
=
F DT I ...(12.35)

H D T JK ln G
H D T JK
2 1
1 2
Note that the LMTD expressions for the parallel flow and the counter-flow heat exchangers (i.e. Eqs. 12.25
and 12.35) are the same.
Again, equation for LMTD is easily remembered as follows:
GTD - LTD
LMTD =
FG
GTD IJ ...(12.35a)
ln
H
LTD K
where
GTD = ‘greater (of the two) temperature difference’, and
LTD = ‘lower temperature difference’
Comments:
(i) When DT1 = DT2: This is a special case, which can occur sometimes in the case of a counter-flow heat ex-
changer. Then, Eq. 12.35 reduces to a form 0/0, which is indeterminate. However, from physical considerations,
DT1 = DT2 means that the temperature difference between the hot and cold fluids is equal throughout the heat
exchanger. Therefore, obviously, the mean temperature difference between the two fluids is DT1 = DT2. (This can
be proved mathematically also, by applying L’Hospital’s rule).
(ii) LMTD for a counterflow heat exchanger is always greater than that for a parallel flow heat exchanger.
This means that to transfer the same amount of heat, counterflow unit will require a smaller heat transfer surface
as compared to a parallel flow unit. This is the reason why a counter-flow heat exchanger is usually preferred.
(iii) LMTD can easily be calculated when all the end temperatures of the fluids are known. Then, immedi-
ately, the heat transfer rate is determined from the Eq. 12.16, i.e. Q = U.A.(LMTD). Therefore, calculation of

HEAT EXCHANGERS 593


1.000 LMTD is an important step in the design of a heat ex-
changer. To facilitate quick calculation of LMTD, when
(LMTD/GTD) and (AMTD/GTD) 0.900
both the end temperatures are known, following graph
0.800 (Fig. 12.7) is provided. Here, (LMTD/GTD) is plotted
0.700 against the ratio (LTD/GTD), where GTD = greater of the
0.600 two end temperature differences, and LTD = lower of the
two end temperature differences. First, calculate the ratio
0.500
(LTD/GTD), and then, read (either from the graph or
0.400 Table 12.4) the value of LMTD/GTD. Next, multiply this
0.300 value by GTD to get LMTD.
0.200 In the same graph, the value of (AMTD/GTD) is also
plotted, for comparison. Here, AMTD is the arithmetic
0.100 mean temperature difference; AMTD = (DT1 + DT 2)/2. It
0.000 may be noted that for values beyond about DT 2/DT 1 = 0.7,
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
AMTD and LMTD are almost the same, i.e. when DT2/
LTD/GTD DT1 > 0.7, it would suffice to use AMTD (which is easier to
LMTD/GTD AMTD/GTD calculate) instead of LMTD. However, for lower values of
DT2/DT1, LMTD has to be used.
FIGURE 12.7 LMTD and AMTD for parallel and Graph for LMTD shown above is also represented in
counter-flow HX tabular form (for better accuracy) in Table 12.4,
(iv) One term occurring in the derivation of LMTD shown above, is the product of mass flow rate and the
specific heat of a fluid, i.e. C = m.Cp. Here, C is known as ‘heat capacity rate’ or, simply ‘capacity rate’ of that
particular fluid. Thus, the capacity rates for hot and cold fluids are:
Ch = mh ×Cph W/C ((12.36)...capacity rate for hot fluid)
Cc = mc ×Cpc W/C ((12.37)...capacity rate for cold fluid)
Then, the heat transfer rate is given by:
Q = Ch ×(Th1 – Th2) W ((12.38)...for hot fluid)
Q = Cc ×(Tc2 – Tc1) W ((12.38)...for cold fluid)
i.e. to transfer a given amount of heat, higher the heat capacity rate of a fluid, lower will be the temperature rise
(or fall) of that particular fluid.
If the heat capacity rates of both the hot and cold fluids are equal, then, the total temperature drop of the hot
fluid will be equal to the total temperature rise of the cold fluid. See Fig. 12.8 (a).
(v) When a fluid is condensing or boiling, its temperature is essentially constant, i.e. Th1 = Th2 for a condens-
ing fluid and Tc1 = Tc2 for a boiling liquid. In other words, DT for the condensing or boiling fluid is zero. But,
since a finite amount of heat is transferred, (= m.h fg.), we say that capacity rate of a condensing or boiling fluid
tends to infinity. Temperature profiles for fluids in a heat exchanger when one of the fluids is condensing or
boiling are shown in Fig. 12.8 (b) and (c), respectively. LMTD for both these cases is determined by the same
procedure as for the parallel or counter-flow heat exchangers, i.e.
D T2 - D T1 D T1 - D T2
LMTD =
F DT I
ln G
=
F DT I ...(12.35)

H D T JK GH D T JK
2 ln 1
1 2
Example 12.4. Furnace oil, flowing at a rate of 4000 kg/h, is heated from 10 to 20°C by hot water flowing at 75°C, with
a velocity of 0.8 m/s, through a copper pipe 2.15 cm OD, 1.88 cm ID. Oil flows through annulus between copper and
steel pipe of 3.35 cm OD and 3 cm ID. Find the length of counter-flow heat exchanger. Fluid properties are given.
Use Dittus–Boelter equation Nu = 0.023.Re 0.8.Pr0.4 .
Solution.
4000
Data: m c :=
3600
i.e. m c = 1.111 kg/s

Tc1 := 10°C Tc2 := 20°C Th 1 := 75°C V := 0.8 m/s kcu := 385 W/(mK)

594 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 12.4 LMTD/GTD for parallel flow and counter-flow HX
LMTD = Log mean temperature difference
GTD = Greater of two end temperature differences
LTD = Lower of two end temperature differences
LTD/GTD LMTD/GTD LTD/GTD LMTD/GTD LTD/GTD LMTD/GTD
0.01 0.215 0.36 0.626 0.71 0.847
0.02 0.251 0.37 0.634 0.72 0.852
0.03 0.277 0.38 0.641 0.73 0.858
0.04 0.298 0.39 0.648 0.74 0.863
0.05 0.317 0.4 0.655 0.75 0.869
0.06 0.334 0.41 0.662 0.76 0.875
0.07 0.350 0.42 0.669 0.77 0.880
0.08 0.364 0.43 0.675 0.78 0.885
0.09 0.378 0.44 0.682 0.79 0.891
0.1 0.391 0.45 0.689 0.8 0.896
0.11 0.403 0.46 0.695 0.81 0.902
0.12 0.415 0.47 0.702 0.82 0.907
0.13 0.426 0.48 0.708 0.83 0.912
0.14 0.437 0.49 0.715 0.84 0.918
0.15 0.448 0.5 0.721 0.85 0.923
0.16 0.458 0.51 0.728 0.86 0.928
0.17 0.468 0.52 0.734 0.87 0.933
0.18 0.478 0.53 0.740 0.88 0.939
0.19 0.488 0.54 0.747 0.89 0.944
0.2 0.497 0.55 0.753 0.9 0.949
0.21 0.506 0.56 0.759 0.91 0.954
0.22 0.515 0.57 0.765 0.92 0.959
0.23 0.524 0.58 0.771 0.93 0.965
0.24 0.533 0.59 0.777 0.94 0.970
0.25 0.541 0.6 0.783 0.95 0.975
0.26 0.549 0.61 0.789 0.96 0.980
0.27 0.558 0.62 0.795 0.97 0.985
0.28 0.566 0.63 0.801 0.98 0.990
0.29 0.574 0.64 0.807 0.99 0.995
0.3 0.581 0.65 0.812 1 1.000
0.31 0.589 0.66 0.818
0.32 0.597 0.67 0.824
0.33 0.604 0.68 0.830
0.34 0.612 0.69 0.835
0.35 0.619 0.7 0.841

Temperature

Condensing fluid
Hot fluid
Th
Ch
DT1 DT2
Q
DT DT1 Tc2
DT2
Cold fluid Cold fluid
Cc = C h Tc1

0 L 0 L
Length

FIGURE 12.8 (a) Both fluids have FIGURE 12.8 (b) One of the fluids condensing (C h Þ ¥)
same capacity rates
HEAT EXCHANGERS 595
Temperature Th1 = 75°C
Water, Ch
Th1
Hot fluid
DT1 Th2 = 52.07°C
Th2
DT1 Q
Tc DT2
Tc2 = 20°C
Boiling fluid, Tc Oil, Cc
Tc1 = 10°C

0 L Length 0 L
Figure 12.8 (c) One of the fluids FIGURE Example 12.4 Counter-flow
boiling (Cc Þ ¥) heat exchanger
Fluid properties:

Property Water Oil


Cp (kJ/kgK) 4.187 1.884
k (W/mK) 0.657 0.138
n (m2/s) 4.187 ´ 10 –7 7.43 ´ 10 –6
r (kg/m3 ) 982 854

Dh i := 1.88 ´ 10 –2 m (inside diameter of tube for hot fluid flow)


Dh o := 2.15 ´ 10 –2 m (outside diameter of tube for hot fluid flow)
Dci := 3 ´ 10 –2 m (inside diameter of tube for cold fluid flow)
Dco := 3.35 ´ 10 –2 m (outside diameter of tube for cold fluid flow)
Cph := 4187 J/(kgK) kh := 0.657 W/(mK) nh := 4.18 ´ 10 –7 m2/s rh := 982 kg/m3 cpc := 1884 J/(kgK)
kc := 0.138 W/(mK) nc := 7.43 ´ 10– 6 m2/s rc := 854 kg/m3
cph × (n h × r h ) cpc ×(n c × r c )
Prh := i.e. Prh = 2.616 Prc := i.e. Pr c = 86.626
kh kc
Total heat transferred:
Q := mc ×cpc ×(Tc 2 – Tc 1) W
i.e. Q = 2.093 ´ 104 W
Inside heat transfer coefficient:
Dhi2
Ai := p× m2 (cross-sectional area of inside tube)
4
i.e. Ai = 2.776 ´ 10 – 4 m2 (cross-sectional area of inside tube)
and, mh := Ai ×V ×r h
i.e. mh = 0.218 kg/s (mass flow rate of hot fluid)
Therefore,
mh
Gh :=
Ai
i.e. Gh = 785.6 kg/(sm2) (mass velocity of hot fluid)
Gh × Dhi
and, Reh := (Remember: m = P ×n)
r h ×u h
i.e. Re h = 3.598 ´ 104 (Reynolds number of hot fluid)
We have:
Nuh := 0.023×Reh0.8 ×Prh0.4 (Dittus–Boelter equation)
i.e. Nu h = 149.154 (Nusselts number)
Nuh × kh
and, h h :=
Dhi

596 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. hh = 5.212 ´ 10 3 W/(m 2C) (heat transfer coefficient for hot (inside) fluid)
Outside heat transfer coefficient:
Equivalent diameter of annulus: Deq = 4 ´ (area of cross section/wetted perimeter)
Dc i2 - Dho2
Acannulus := p×
4
i.e. Acannulus = 3.438 ´ 10 – 4 m2 (cross-sectional are of annulus)
and, P := p×(Dc 1 + Dh o)
i.e. P = 0.162 m (wetted perimeter of annulus)
Therefore,
4 ×Acannulus
Deq :=
P
Deq = 8.5 ´ 10 –3 m (equivalent diameter of annulus)
mc
Also, Gc :=
Acannulus
i.e. Gc = 3.232 ´ 10 3 kg/(sm2) (mass velocity of cold fluid)
Gc × Deq
and, Rec :=
n c × rc
i.e. Rec = 4.329 ´ 10 3 (Reynolds number of cold fluid)
From Dittus–Boelter’s equation
Nu c := 0.023× Rec0.8 × Prc0.4
i.e. Nuc = 111.152 (Nusselts number for cold fluid flow)
Therefore,
Nuc × kc
hc :=
Deq
i.e. hc = 1.805 ´ 10 3 W/(m 2C) (heat transfer coefficient for cold (outside) fluid)
Overall heat transfer coefficient U:
We have: U×A = 1/(Total thermal resistance)
F I Dho
Rt :=
1
hh × (p × Dhi × 1)
+
1
hc × (p × Dho × 1)
+
1
2 × p × kcu × 1
GH JK
× ln
Dhi
(Total thermal resistance)

i.e. R t = 0.012 K/W (per metre length)


Therefore,
1
Ui =
Rt × Ai

1
i.e. Ui :=
Rt × (p × Dhi × 1)
i.e. Ui = 1.471 ´ 10 3 W/(m 2K) (heat transfer coefficient referred to inside surface area)
1
and, Uo :=
Rt × (p × Dho × 1)
i.e. Uo = 1.287 ´ 10 3 W/(m 2K) (heat transfer coefficient referred to outside surface area)
Now, calculate LMTD:
Exit temperature of hot fluid:
Q
Th 2 := Th1 –
mh × c ph
i.e. Th2 = 52.074°C (exit temperature of hot fluid)
Therefore,
DT1 := Th 1 – Tc 2
or, DT 1 = 55°C (temperature difference between hot and cold streams at inlet of HX)
and, DT2 := Th 2 – Tc 1
i.e. DT 2 = 42.074°C (temperature difference between hot and cold streams at exit of HX)

HEAT EXCHANGERS 597


Therefore,
D T1 - D T2
LMTD :=
F DT I
ln GH D T JK
1

i.e. LMTD = 48.249°C (Log Mean Temperature Difference).


Alternatively:
We can calculate LMTD quickly by using graph of Fig. 12.7 or from Table 12.4: We have: DT2/DT1 = 42.074/55 = 0.765.
From the table, we read against DT2/DT1 = 0.765, a value of LMTD/DT1 = 0.8775. Then, LMTD = 0.8775 ´ 55 = 48.262°C.
Length of HX required:
Heat exchange area required:
Q
Area :=
U i LMTD
i.e. Area = 0.295 m2 (heat exchange area required)
Therefore,
Area
L :=
p ×Dhi
i.e. L = 4.993 m (length of HX, metres.)
Example 12.5. In a shell-and-tube heat exchanger, tubes are 4 m
Condensing fluid
Temperature long, 3.1 cm OD, 2.7 cm ID. Water is heated from 22°C to 45°C by
Th = 100°C
condensing steam at 100°C on the outside of tubes. Water flow
rate through the tubes is 10 kg/s. Heat transfer coefficient on
Q steam side is 5500 W/(m2 K) and on waterside, 850 W/(m2 K).
DT2 = 55°C Neglecting all other resistances, find the number of tubes.
DT1 = 78°C
Solution.
Tc2 = 45°C Data:
Tc 1 := 22°C Tc 2 := 45°C Th := 100°C
Tc1= 22°C Cold fluid hi := 850 W/(m2 K) ho := 5500 W/(m2 K)
di := 0.027 m do := 0.03 m L := 4 m
0 L Length
m := 10 kg/s cp := 4170 J/(kgK)
FIGURE Example 12.5 Heat exchanger with Overall heat transfer coefficient:
one of the fluids condensing (Ch Þ ¥) 1
We have: Uo ×Ao =
S Rth
1 1
i.e. Uo ×Ao = =
1 1 1 1
+ +
ho × Ao hi × Ai ho × p × do × L hi × p × di × L
Since Ai = p × di × L and Ao = p × do × L
1
i.e. Uo :=
F1+ d I
GH h h × d JK
o i
o

i.e. Uo = 671.588 W/(m2 K) (overall heat transfer coefficient referred to outside area)
To calculate LMTD:
Now, DT1 := Th – Tc1
i.e. DT1 = 78°C (temperature difference at inlet)
and, DT2 := Th – Tc2
i.e. DT2 = 55°C (temperature difference at exit)
D T1 - D T2
Also, LMTD :=
F DT I
ln GH D T JK
1

i.e. LMTD = 65.832°C (Log Mean Temperature Difference.)

598 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Alternatively:
We can calculate LMTD quickly by using graph of Fig. 12.7 or from Table 12.4. We have: DT2/DT1 = 55/78 = 0.705. From
the table, we read against DT2/DT1 = 0.705, value of LMTD/DT1 = 0.844. Then, LMTD = 0.844 ´ 78 = 65.832°C....same
value as obtained above by calculation.
Heat transfer:
Q total := m×cp ×(Tc2 – Tc1) W
Q total = 9.591 ´ 10 5 W
A := p×do ×L ; A = 0.377 (outside area of each tube, m2)
Number of tubes required:
Qtotal
Atotal := (Total heat transfer area required m2)
Uo × LMTD
i.e. Atotal = 21.693 (Total heat transfer area required m 2)
Atotal
Therefore, N :=
A
i.e. N = 57.543 (No. of tubes required)
i.e. Number of tubes required is, say, 58.
Example 12.6. In a double pipe counter-flow heat exchanger,
10,000 kg/h of oil (Cp = 2.095 kJ/kgK) is cooled from 80°C to
Th1 = 80°C
50°C by 8000 kg/h of water entering at 25°C. Determine the area Oil, Ch
of heat exchanger for an overall U = 300 W/(m2 K). Take Cp for
water as 4.18 kJ/kgK.
(M.U. 1997) DT1 Th2 = 50°C
Solution.
Data: DT2
10000 Tc2 = 43.795°C
mh := i.e. mh = 2.778 kg/s
3600 Water, Cc Tc1 = 25°C
8000
Cph := 2095 J/KgK mc := 0 L
3600
i.e. mc = 2.222 kg/s Cpc := 4180 J/kgK
FIGURE Example 12.6 Counter-flow
U := 300 W/(m 2K) Th1 := 80°C
heat exchanger
Th2 := 50°C Tc1 := 25°C
mh × C ph ×(Th 1 - Th 2 )
Therefore: Tc2 := Tc 1 + (from heat balance)
mc × Cpc
i.e. Tc2 = 43.795°C (exit temperature of cold fluid (water))
Total heat transfer:
Q := mh ×Cph ×(Th1 – Th2)
i.e. Q = 1.746 ´ 10 5 W (total heat transfer)
To calculate LMTD:
We have:
DT1 := Th1 – Tc2
i.e. DT 1 = 36.205°C (temperature difference at the inlet of HX)
and, DT2 := Th2 – Tc1
i.e. DT 2 = 25°C (temperature difference at the exit of HX)
Therefore,
D T1 - D T2
LMTD :=
F DT I
ln GH D T JK
1

i.e. LMTD = 30.258°C (Log Mean Temperature Difference.)


Alternatively:
We can calculate LMTD quickly by using graph of Fig. 12.7 or from Table 12.4. We have: DT2/DT1 = 25/36.205 = 0.691.
From the table, we read against DT2 /DT1 = 0.69, a value of LMTD/DT1 = 0.835. Then, LMTD = 0.835 ´ 36.205 =
30.231°C....almost the same value as obtained above by calculation.

HEAT EXCHANGERS 599


Area required:
Q
A :=
U ×LMTD
i.e. A = 19.233 m2 (area required.)

12.5 Correction Factors for Multi-pass and Cross-flow Heat Exchangers


LMTD relations derived above are applicable to parallel flow and counter-flow heat exchangers only. But, in
practice, cross-flow heat exchangers (e.g. automobile radiators) and shell-and-tube heat exchangers, with more
than one pass in shell side and/or tube side, are also used. In such cases, the flow situation is complex and the
analytic relations for mean temperature difference are very complicated. Then, first, LMTD is calculated as if for
a counter-flow heat exchanger with the inlet and exit temperatures for the two fluids as per the actual data, and
next, a ‘correction factor (F)’ is applied to the calculated LMTD to get the mean temperature difference between
the fluids. Now, heat transfer rate is calculated as:
Q = U×A×(F×LMTD) W ...(12.39)
where, A is the area of heat transfer, U is the overall heat transfer coefficient referred to that area, and F is the
correction factor. Note again that LMTD is calculated as if for a counter-flow heat exchanger, taking the inlet
and exit temperatures of the two fluids the same as for the actual heat exchanger.
Values of correction factor (F) for a few selected heat exchangers are given in graphical representation in Fig.
12.9. F varies from 0 to 1. In these graphs, correction factor F is plotted as function of two parameters, i.e. P and
R, defined as:

1.0 T1

0.9 t2
Correction factor F

t1
0.8
R = 4.0 3.0 2.0 1.5 1.0 0.8 0.6 0.4 0.2 T2
0.7

0.6 T1 - T2 t 2 - t1
R= P=
t 2 - t1 T1 - t 1
0.5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

FIGURE 12.9(a) One shell pass and 2,4,6, etc. (any multiple of 2), tube passes

1.0 T1

0.9 t2
Correction factor F

0.8
R = 4.0 3.0 2.0 1.5 1.0 0.8 0.6 0.4 0.2 t1
0.7
T1 - T2 T2
R=
0.6 t 2 - t1
t 2 - t1
P=
T1 - t 1
0.5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

FIGURE 12.9(b) Two shell passes and 4,8,12, etc. (any multiple of 4), tube passes

600 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1.0
T1

0.9
Correction factor F
0.8 t1 t2
R = 4.0 3.0 2.0 1.5 1.0 0.8 0.6 0.4 0.2

0.7
T1 - T2
R=
0.6 t 2 - t1 T2
t -t
P= 2 1
0.5 T1 - t 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

FIGURE 12.9(c) Single pass cross-flow with both fluids un-mixed

1.0
T1

0.9
Correction factor F

0.8
R = 4.0 3.0 2.0 1.5 1.0 0.8 0.6 0.4 0.2 t1 t2
0.7
T1 - T2
R=
0.6 t 2 - t1
t 2 - t1 T2
P=
T1 - t 1
0.5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

FIGURE 12.9(d) Single pass cross-flow with one fluid mixed and the other un-mixed

t 2 - t1
P= ...(12.40)
T1 - t1
T1 - T2 Ctube_side
R= = ...(12.41)
t2 - t1 Cshell_side
where, C is the capacity rate = m.Cp.
Also, for a shell-and-tube heat exchanger, T and t represent the temperatures of fluids flowing through the
shell and tube sides, respectively. And, subscripts 1 and 2 refer to the inlet and exit, respectively. It makes no
difference whether hot or cold fluid flows through the shell or the tube. Values of P vary from 0 to 1 and it is
equal to the ratio of the temperature change of the tube side fluid to the maximum temperature difference be-
tween the two fluids; thus, P represents the thermal effectiveness of the tube side fluid; values of R vary from 0
to ¥. When R = 0, it means that the fluid on the shell side is undergoing a phase change (i.e. boiling or conden-
sation, which occurs at a practically constant temperature, Tsat), and when R = ¥, the tube side fluid is undergo-
ing a phase change. Observe from the graphs that, when R = 0 or ¥, the correction factor F is equal to 1.
Therefore, for a condenser or boiler, F = 1, irrespective of the configuration of the heat exchanger.
Note: To apply the correction factor F from these graphs, it is necessary that the end temperatures of both the
fluids must be known.
Example 12.7. A one shell pass, two tube pass heat exchanger, with flow arrangement similar to that shown in Fig. 12.9
(a), has water flowing through the tubes and engine oil flowing on the shell side. Water flow rate is 1.2 kg/s and its
temperatures at inlet and exit are 25°C and 75°C, respectively. Engine oil enters at 110°C and leaves at 75°C. Overall U
= 300 W/(m2K). Take Cp for water as 4.18 kJ/(kgK) and calculate the heat transfer area required.

HEAT EXCHANGERS 601


Solution.
Data:
mc := 1.2 kg/s Cpc := 4180 J/kgK
Th1 = 110 C U := 300 W/(m2K) T1 := 110°C (hot fluid, inlet)
Oil, Ch
T2 := 75°C (hot fluid, exit)
DT1 t1 := 25°C (cold fluid, inlet)
Th2 = 75 C t2 := 75°C (cold fluid, exit)
Therefore, total heat load:
Tc2 = 75 C
DT2 Q := mc ×Cpc ×(t 2 – t1) W
i.e. Q = 2.508 ´ 10 5 W
Water, Cc
Since this is a multi-pass HX, LMTD must be calculated
Tc1 = 25 C as for a counter-flow HX, and, then a correction factor ap-
0 L plied from Fig. 12.9 (a):
To calculate LMTD:
FIGURE Example 12.7 Counter-flow DT1 := T1 – t 2
heat exchanger i.e. DT 1 = 35°C
and, DT 2 := T 2 – t1
i.e. DT2 = 50°C
Therefore,
D T1 - D T2
LMTD :=
F DT I
ln GH D T JK
1

i.e. LMTD = 42.055°C (Log Mean Temperature Difference)


Correction factor, F:
We have:
t2 - t1
P :=
T1 - t1
i.e. P = 0.588
T1 - T2
and, R :=
t2 - t1
i.e. R = 0.7
Then, from Fig. 12.9 (a):
F = 0.8 (correction factor)
And, the corrected temperature difference becomes:
DT := 0.8×LMTD
i.e. DT = 33.644°C (actual mean temperature)
Therefore, heat transfer area:
Q
A :=
U × DT
i.e. A = 24.848 m2 (heat transfer area required.)
Example 12.8. In a shell and tube HX, 50 kg/min of furnace oil is heated from 10 to 90°C. Steam at 120°C flows through
the shell and oil flows inside the tube. Tube size: 1.65 cm ID and 1.9 cm OD. Heat transfer coefficient on oil and steam
sides are: 85 and 7420 W/(m2 K), respectively. Find the number of passes and number of tubes in each pass if the length
of each tube is limited to 2.85 m. Velocity of oil is limited to 5 cm/s. Density and specific heat of oil are 900 kg/m3 and
1970 J/(kg.K), respectively. (M.U. 1994)
Solution.
Data:
Tc1 := 10°C Tc2 := 90°C Th := 120°C di := 0.0165 m do := 0.019 m L := 2.85 m hoil := 85 W/(m2 K)
h steam := 7420 W/(m2 K) V := 0.05 m/s roil := 900 kg/m3 Cpoil := 1970 J/(kgK)
50
moil := i.e. m oil = 0.833 kg/s
60

602 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Total heat transferred: Condensing fluid
Temperature
Q := m oil ×Cpoil ×(Tc2 – Tc1) Th = 120 C
i.e. Q = 1.313 ´ 105 W (total heat transferred)
Number of tubes required: DT2 = 30 C
Total cross-sectional area for flow: Q
Tc2 = 90 C
moil DT1 = 110 C
Acs := i.e. Acs = 0.019 m2
V ×r oil
Cross-sectional area of each tube: Cold fluid
d2 –4 2 Tc1= 10 C
A := p× i A = 2.138 ´ 10 m
4
0 L Length
Therefore,
Acs FIGURE Example 12.8 Heat exchanger with
N := N = 86.606 (number of tubes,
A say 87 from velocity consideration) one of the fluids condensing (Ch Þ ¥)

i.e. N =87
Overall U, based on outer surface area:
Total thermal resistance:
1 1
Rt := +
hoil × p × di × 1 hsteam × p × do × 1
i.e. Rt = 0.229 K/W
Now, Ao := p×do×1
i.e. Ao = 0.06 m2 (outside surface area of tube/metre length)
Then,
1
Uo :=
Ao × Rt
i.e. Uo = 73.089 W/(m2K) (overall heat transfer coefficient referred to outside area of tube)
D T1 - D T2 110 - 30
and, LMTD =
D T1 F
=
110 I FG IJ
ln
D T2 GHln
30 JK H K
i.e. LMTD = 61.572°C (Log Mean Temperature Difference)
Therefore, heat transfer area required:
Q
Aht :=
U o ×LMTD
i.e. Aht = 29.184 m2 (total heat transfer area required)
Now, Atube = p×do×1
i.e. Atube = 0.06 m2 (heat transfer area per metre length)
Therefore,
Aht
Length :=
N × Atube
Length = 5.62 m (length of tube required)
But, length is limited to 2.85 m. So, use 2 tube passes.
Then, it becomes a shell-and-tube HX with two tube passes. So, it appears at first sight that correction factor (F) has
to be obtained from Fig. 12.9; but, observe that one of the fluids is condensing. So, F = 1, irrespective of HX configura-
tion.
i.e. F = 1
Therefore, Aht remains same.
Then,
Aht
Length =
2× N × Atube
i.e. Length = 2.81 m (this is less than 2.85 m, So, OK.)

HEAT EXCHANGERS 603


12.6 The Effectiveness–NTU Method for Heat Exchanger Analysis
LMTD can readily be determined when all the four end temperatures are either given, or can easily be calculated.
Then, the area required, A (i.e. the size of the HX) is easily found out by applying the equation: Q = U.A.(LMTD).
In other words, LMTD method is very convenient to use for sizing problems, when all the end temperatures are
known. However, there are certain problems where only the inlet temperatures of both the fluids are specified,
along with the flow rates and the overall heat transfer coefficients, and the heat transfer rate and the exit tem-
peratures of the fluids are to be calculated. Solution of such rating problems by the LMTD method would require
tedious iterations. However, the Effectiveness–NTU method, developed by Kays and London in 1955, overcomes
this problem and makes the solution straight forward. Effectiveness–NTU method is also useful in solving heat
exchanger problems, where off-design conditions exist; i.e. for example, the heat exchanger might have been
designed for some particular flow rates of fluids; now, to find out what happens to the performance if flow rate
of one of the fluids is reduced to, say, 75 % of the design flow rate, and so on.
The effectiveness–NTU method is not an altogether new method; fundamental equations are the same as
used in the LMTD method, but the different variables are arranged rather differently.
Before we develop the Effectiveness–NTU relations for different types of heat exchangers, let us define a few
quantities:
Effectiveness of a heat exchanger (e ):
Q
e= ...(12.42)
Qmax
where,
Q = actual heat transferred in the heat exhanger
Qmax = maximum possible heat transfer in the heat exchanger
Now, actual heat transfer rate in a heat exchanger is given by:
Q = mh ×Cph ×(Th1 – Th2) = Ch ×(Th1 – Th2)
and,
Q = mc ×Cpc ×(Tc2 – Tc1) = Cc ×(Tc2 – Tc1)
where, C h = capacity rate of the hot fluid, and
Cc = capacity rate of the cold fluid
Now, Ch may be equal to Cc or less than Cc or greater than Cc.
If Ch < Cc, we designate Ch as Cmin;
Instead, if Ch > Cc, we designate Cc as Cmin.
And, in each case, capacity rate of the other fluid is designated as Cmax.
Capacity Ratio (C):
Capacity ratio is defined as:
Cmin
C= ...(12.43)
Cmax
Number of Transfer Units (NTU):
Number of Transfer Units (which is a dimensionless number), is defined as:
U ×A
NTU = ...(12.44)
Cmin
where, U is the overall heat transfer coefficient and A is the corresponding heat transfer area. For given value of
A and flow conditions, NTU is a measure of the area (i.e. size) of the heat exchanger. Larger the NTU, larger the
size of the heat exchanger.
Maximum possible heat transfer in a heat exchanger (Qmax):
Now, consider a heat exchanger where the hot fluid is cooled from a temperature of Th1 to Th2 and the cold fluid
heated from Tc1 to Tc2. So, the maximum temperature differential in the heat exchanger is (Th1 – Tc1). Now, if the
heat exchanger had an infinite area, the hot fluid will be cooled from Th1 to Tc1 or the cold fluid may be heated
from Tc1 to Th1. However, which fluid will experience the maximum temperature differential (Th1 – Tc1) will
depend upon which fluid has the minimum capacity rate.
If hot fluid has the minimum capacity rate, we can write:

604 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Qmax = Ch ×(Th1 – Tc1) (if Ch is minimum capacity rate, Cmin)
Instead, if cold fluid has the minimum capacity rate, we write:
Q max = Cc ×(Th1 – Tc1) (if Cc is minimum capacity rate, Cmin.)
Or, more generally, we write:
Q max = Cmin ×(Th1 – Tc1) ...(12.45)
These situations are represented graphically in Fig. 12.9:

Temperature
Temperature

Hot fluid Hot fluid


Th1 Ch = Cmin
Th1 = Tc2 Ch
Tc2
Th2
Cold fluid Th2 = Tc1
Cc = Cmin Cold fluid
Tc1
Cc

0 L 0 L Length
FIGURE 12.9(a) Cold fluid has FIGURE 12.9(b) Hot fluid has
minimum capacity rate minimum capacity rate

Therefore, we can write for effectiveness:


Ch ×(Th 1 - Th 2 )
Q Cc × (Tc 2 - Tc 1)
e= = = ...(12.46)
Qmax Cmin × (Th 1 - Tc 1) Cmin × (Th 1 - Tc 1)
Now, if hot fluid is the ‘minimum fluid’ (i.e. Ch < Cc), we get from Eq. 12.46:
(Th1 - Th 2 )
e= (for Ch < Cc...(12.47a).)
(Th 1 - Tc 1)
And, if cold fluid is the ‘minimum fluid’ (i.e. Cc < Ch), we get from Eq. 12.46:
(Tc 2 - Tc 1)
e= (for Cc < C h...(12.47b))
(Th 1 - Tc 1)
i.e. by suitably choosing the fluid, the effectiveness of a heat exchanger can be expressed as a ratio of tempera-
tures (or, as a temperature effectiveness).
If Ch = Cc, obviously, both the fluids will experience the maximum possible temperature differential, if the
heat exchanger had an infinite area.
Now, for any heat exchanger, effectiveness can be expressed as a function of the NTU and capacity ratio,
Cmin/Cmax, i.e.
F C I
GH
e = f NTU , min
Cmax JK ...(12.47c)

We shall derive below e-NTU relation for a parallel flow HX.


12.6.1 Effectiveness–NTU Relation for a Parallel-flow Heat Exchanger
Consider the parallel-flow heat exchanger shown in Fig. 12.5. Assumptions for this derivation remain the same as
for the LMTD method.
Continuing from Eq. 12.21:

FT - Tc 2 I F 1 I
ln GH T
h2
h 1 - Tc 1
JK = – U×A× GH m ×C
h ph
+
1
mc × Cpc
JK ...(12.21)

Now, out of the two fluids, one is the ‘minimum’ fluid and the other is the ‘maximum’ fluid. Whichever
may be the minimum fluid, we can write Eq. 12.21 as:
HEAT EXCHANGERS 605
FT - Tc 2 I -U ×A FC I
ln GH T h2
h 1 - Tc 1
JK =
Cmin GH
× 1 + min
Cmax JK ...(12.48)

Th 2 - Tc 2 LM C F I OP
i.e.
Th 1 - Tc 1
= exp - NTU × 1 + min
MNCmax GH JK PQ ...(12.49)

Now, substituting for Th2 and Tc2 from Eq. 12.46, we get:

LMT Cmin C OP LM OP
Q = exp LM- NTU ×FG 1 + C I OP
- e× ×(Th1 - Tc1 ) - e × min ×(Th1 - Tc1 ) + Tc 1
N h1
Ch Cc Q N min
JK PQ
Th1 - Tc1 MN H C max

FG 1 1I
+ J
(Th1 - Tc 1 ) - e × Cmin ×(Th1 - Tc1 ) ×
HC C K L F C
= exp M - NTU × G 1 +
I OP
i.e.
Th1 - Tc1
h c

MN H C
min
max
JK PQ
F1 1I L F C
= exp M - NTU × G 1 +
I OP
i.e. 1 – e×Cmin × GH C
h
+ J
C K c MN H C
min
max
JK PQ
Now, assuming Ch > Cc, i.e. cold fluid as the ‘minimum fluid’, we have: Cmin = Cc and Cmax = Ch.
Therefore,

F Cmin I LM F Cmin I OP
GH
1 – e× 1 +
Cmax JK = exp - NTU × 1 +
MN GH Cmax JK PQ
LM F Cmin I OP
1 - exp - NTU × 1 +
MN GH Cmax JK PQ
i.e. e= ...(12.50)
Cmin
1+
Cmax
Eq. 12.50 is the desired expression for effectiveness of a parallel flow heat exchanger.
Note that the same result would be obtained, if we assume the hot fluid as the ‘minimum’ fluid.
Eq. 12.50 is concisely expressed as:
1 - exp ( - N × (1 + C ))
e= ...(12.51)
1+ C
Cmin
where, N = NTU and, C=
Cmax
Special cases:
(i) For a condenser or boiler i.e. one of the fluids undergoes a phase change. Therefore, Cmax ® ¥ i.e. Capacity
ratio, C = 0. Then effectiveness relation (for all heat exchangers) reduces to:
e = 1 – exp(– NTU) ...(12.52)
(ii) When C = 1, i.e. Cmin = Cmax This is the case of a typical, gas turbine regenerator. In this case,
1 - exp( - 2 × NTU)
e= (for C = 1, parallel flow HX...(12.53))
2
12.6.2 Effectiveness–NTU Relation for a Counter-flow Heat Exchanger
Again, consider the counter-flow heat exchanger shown in Fig. 12.6. Assumptions for this derivation remain the
same as for the LMTD method.
Continuing from Eq. 12.31:

606 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FT - Tc 1 I = – U×A× FG 1 I
JK
ln GH T
h2
h1 - Tc 2
JK H m ×C h ph
-
1
mc × Cpc
...(12.31)

This can be written as:


FT - Tc 1 I = – U×A× F 1 I
ln GH T
h2
h1 - Tc 2
JK GH C h
-
1
Cc JK
Assuming hot fluid as the ‘minimum’ fluid,
C min = Ch and, Cmax = Cc
we have:
FT - Tc 1 I = – U×A× F 1 - 1 I
ln GH T
h2
h1 - Tc 2
JK GH C C JK
min max

Th 2 -T L F 1 - 1
= exp M - U × A × G
I OP
i.e.
Th1 -T
c1
c2 MN HC C min max
JK PQ
Th 2 -T L - U × A ×F 1 - C I OP
= exp M
MN C GH C JK PQ
c1 min
i.e.
Th1 -T c2 min max

Th 2 -T L F C I OP
= exp M - NTU × G 1 -
H C JK QP
c1 min
i.e. ...(12.54)
Th1 -T c2 NM max
Now, substituting for Th2 and Tc2 from Eq. 12.46, we get:

LMT Cmin OP
N h1 - e×
Ch
×(Th1 - Tc1 ) - Tc1
Q LM F Cmin I OP
LMT Cmin OP = exp - NTU × 1 -
MN GH Cmax JK PQ ...(12.55)
h1 - e × ×(Th1 - Tc1 ) - Tc1
N Cc Q
Now, put Cmin = Ch , C = Cmin/Cmax, and N = NTU, in Eq. 12.55:
(Th1 - Tc 1 ) ×(1 - e )
= exp (– N×(1 – C))
(Th1 - Tc1 ) × (1 - C × e )

1-e
i.e. = exp(– N×(1 – C))
1 - C ×e
i.e. 1 – e = exp(–N×(1 – C)) – C×e×exp(– N×(1 – C))
i.e. e×(1 – C×exp(– N×(1 – C))) = 1 – exp×(– N×(1 – C))
1 - exp ( - N × (1 - C ))
or, e= ...(12.56)
(1 - C × exp( - N ×(1 - C )))
Instead of assuming that the hot fluid is the minimum fluid, if we assume that the cold fluid is the ‘mini-
mum’ fluid, then also the same relation (namely, Eq. 12.56), will result.
Eq. 12.56 is the desired expression for the effectiveness of the counter-flow heat exchanger.
Special cases:
(i) For a condenser or boiler i.e. one of the fluids undergoes a phase change. Therefore, Cmax ® ¥. i.e.
Capacity ratio, C = 0. Then effectiveness relation (for all heat exchangers) reduces to:
e = 1 – exp (– NTU) ...(12.57)
(ii) When C = 1, i.e. Cmin = Cmax This is the case of a typical, gas turbine regenerator. In this case, relation for
e reduces to the indeterminate form, 0/0. Then, apply the L’Hospital’s rule to evaluate e. i.e. differentiate
the numerator and denominator w.r.t. C and taking the limit C ® 1, we get:

HEAT EXCHANGERS 607


NTU
e= -for C = 1., counter-flow HX...(12.58))
1 + NTU
Effectiveness–NTU relations and the corresponding graphical representations for several types of heat ex-
changers are given by Kays and London.
Table 12.5 gives the Effectiveness relations for a few types of heat exchangers; and Table 12.6 gives the NTU
relations:

TABLE 12.5 Effectiveness relations for heat exchangers


[N = NTU = U.A/Cmin, C = Cmin/Cmax]

Flow geometry Relation

1 - exp(- N × (1 + C ))
Double pipe: parallel-flow e=
1+ C

1 - exp(- N × (1 - C ))
Double pipe: counter-flow e=
(1 - C × exp(- N ×(1 - C )))

N
Counter-flow, C = 1 e=
1+ N

FG exp(- N ×C × n) - 1IJ where, n = N – 0.22


Cross-flow: (single pass) both fluids un-mixed e = 1 – exp
H C ×n K
FG 1 + C 1I
-1

Cross-flow: (single pass) both fluids mixed e=


H 1 - exp(- N ) 1 - exp(- N ×C ) - N JK
1
Cross-flow: (single pass) Cmax mixed, Cmin un-mixed e= ×[1 – exp[– C ×(1 – e – N)]]
C

Cross-flow: (single pass) Cmax un-mixed, Cmin mixed e = 1 – exp -


LM 1
× (1 - exp(- N ×C ))
OP
N C Q
Shell and tube:

LM L 1
OP OP -1

1 + exp M - N × (1 + C )
2 2

M
e = 2 × M1 + C + (1 + C )
1
MN PQ PP
L OP P
2 2
One shell pass, 2, 4, 6 tube passes ×
MM 1
1 - exp M - N × (1 + C )
PQ PQ
2 2

N MN
LM (1 - e ×C ) OP
p
n

-1
Multiple shell passes, 2n, 4n, 6n tube
passes (e P = effectiveness of each shell e=
MN (1 - e ) PQ
p

pass, n = number. of shell passes) LM (1 - e ×C ) OP


p
n

-C
NM (1 - e ) QPp

n ×e p
Special case for C = 1 e=
1 + (n - 1) × e p

All exchangers, with C = 0 (Condensers e = 1 – e –N


and Evaporators)

608 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 12.6 NTU relations for heat exchangers
[N = NTU = U.A/Cmin, C = Cmin/Cmax, e = effectiveness]
Flow geometry Relation
- ln(1 - (1 + C ) × e )
Double pipe: parallel-flow N=
1+ C

1 FG
e -1 IJ
Double pipe: counter-flow, for C = 1 N=
C -1
× ln
H
C ×e - 1 K
e
Counter-flow, C = 1 N=
1- e

FG 1 IJ
Cross-flow: Cmax mixed, Cmin un-mixed
H
N = - ln 1 +
C
× ln(1 - C × e )
K
-1
Cross-flow: Cmax un-mixed, Cmin mixed N= × ln(1 + C × ln(1 - e ))
C
Shell and tube:

LM 2 - 1 - C - (1 + C ) 1
2 2
OP
MM e PP
-1
2
One shell pass, 2, 4, 6 tube passes N = - (1 + C ) 2 ×ln 1

MN 2e - 1 - C + (1 + C )
2 2
PQ
All exchangers, with C = 0 (Condensers N = – ln(1 – e)
and Evaporators)

NTU–Effectiveness graphs:
NTU–Effectiveness relations are also represented in graphical form and these are quite instructive. However, it is
a bit difficult to read these graphs accurately; so, analytical relations may be used wherever possible.
NTU–Effectiveness relations for parallel-flow and counter-flow heat exchangers are shown graphically in
Fig. 12.10 and 12.11, respectively. In these figures, effectiveness values are plotted against NTU for different
values of capacity ratio, C.
For convenience and accuracy in reading, effectiveness values for the parallel flow and counter-flow heat
exchangers are given in Tabular form, in Table 12.7 and 12.8:

1.000
0.900
=0 .2
0.800 C C=0
C = 0.4
0.700
C = 0.6
Effectiveness

0.600 C = 0.8
C=1
0.500
0.400
0.300
0.200
0.100
0.000
0 1 2 3 4 5
NTU

FIGURE 12.10 NTU Vs. effectiveness for parallel-flow heat exchangers

HEAT EXCHANGERS 609


1.000
0
0.900 =
C
0.800
0.700 1
C= C=0

Effectiveness
0.600 C = 0.2
C
0.500 C = 0.4
C = 0.6
0.400 C = 0.8
0.300 C=1
0.200
0.100
0.000
0 1 2 3 4 5
NTU

FIGURE 12.11 NTU Vs. effectiveness for counter-flow heat exchangers

NTU–effectiveness graphs for some other types of heat exchangers, are given by Kays and London, and are
reproduced below:

100 100
¥ 0.25
0
= 0, 4
xa
=
/C m d 0.5
80 ni 5 ixe
Cm 0.2 0 80
nm
2
0.5 /C u 0.75
5 d 1.33
0.7 ixe
Effectiveness e, %
Effectiveness e, %

1.00 Cm 1
60 60

Cold fluid
40 40 Mixed
Hot
fluid fluid

20 20
Unmixed
fluid
0
0 1 2 3 4 5
1 2 3 4 5
Number of transfer units NTU = AU/Cmin
Number of transfer units NTU = AU/Cmin

FIGURE 12.12 Cross-flow heat exchanger FIGURE 12.13 Cross-flow heat exchanger with one
with both fluids un-mixed fluid mixed and the other un-mixed

Note: In Fig. 12.13, the dashed lines are for the case of Cmin un-mixed and Cmax mixed. And, the solid lines are for
the case of Cmin mixed and Cmax un-mixed.
From the NTU—Effectiveness graphs, following important points may be observed:
(i) For a given value of capacity ratio, C, the effectiveness increases with NTU. Value of effectiveness varies
from 0 to 1.
(ii) Initially, effectiveness increases rather rapidly as NTU increases (up to a value of NTU = about 1.5) and
then, slowly for larger values of NTU. Remember that NTU is a measure of the size (i.e. heat exchange
area, A) of the heat exchanger; so, we can conclude that increasing the size of the heat exchanger beyond
about NTU = 3, cannot be economically justified, since there will not be any corresponding increase in
effectiveness.
610 FUNDAMENTALS OF HEAT AND MASS TRANSFER
TABLE 12.7 NTU Vs. effectiveness for parallel-flow HX

NTU C= 0 C = 0.2 C = 0.4 C = 0.6 C = 0.8 C= 1


0.1 0.095 0.094 0.093 0.092 0.092 0.091
0.2 0.181 0.178 0.174 0.171 0.168 0.165
0.3 0.259 0.252 0.245 0.238 0.232 0.226
0.4 0.330 0.318 0.306 0.295 0.285 0.275
0.5 0.393 0.376 0.360 0.344 0.330 0.316
0.6 0.451 0.428 0.406 0.386 0.367 0.349
0.7 0.503 0.474 0.446 0.421 0.398 0.377
0.8 0.551 0.514 0.481 0.451 0.424 0.399
0.9 0.593 0.550 0.512 0.477 0.446 0.417
1 0.632 0.582 0.538 0.499 0.464 0.432
1.1 0.667 0.611 0.561 0.517 0.479 0.445
1.2 0.699 0.636 0.581 0.533 0.491 0.455
1.3 0.727 0.658 0.599 0.547 0.502 0.463
1.4 0.753 0.678 0.614 0.558 0.511 0.470
1.5 0.777 0.696 0.627 0.568 0.518 0.475
1.6 0.798 0.711 0.638 0.577 0.524 0.480
1.7 0.817 0.725 0.648 0.584 0.530 0.483
1.8 0.835 0.737 0.657 0.590 0.534 0.486
1.9 0.850 0.748 0.664 0.595 0.537 0.489
2 0.865 0.758 0.671 0.600 0.540 0.491
2.1 0.878 0.766 0.677 0.603 0.543 0.493
2.2 0.889 0.774 0.681 0.607 0.545 0.494
2.3 0.900 0.781 0.686 0.609 0.547 0.495
2.4 0.909 0.787 0.689 0.612 0.548 0.496
2.5 0.918 0.792 0.693 0.614 0.549 0.497
2.6 0.926 0.797 0.696 0.615 0.550 0.497
2.7 0.933 0.801 0.698 0.617 0.551 0.498
2.8 0.939 0.804 0.700 0.618 0.552 0.498
2.9 0.945 0.808 0.702 0.619 0.553 0.498
3 0.950 0.811 0.704 0.620 0.553 0.499
3.1 0.955 0.813 0.705 0.621 0.553 0.499
3.2 0.959 0.815 0.706 0.621 0.554 0.499
3.3 0.963 0.817 0.707 0.622 0.554 0.499
3.4 0.967 0.819 0.708 0.622 0.554 0.499
3.5 0.970 0.821 0.709 0.623 0.555 0.500
3.6 0.973 0.822 0.710 0.623 0.555 0.500
3.7 0.975 0.824 0.710 0.623 0.555 0.500
3.8 0.978 0.825 0.711 0.624 0.555 0.500
3.9 0.980 0.826 0.711 0.624 0.555 0.500
4 0.982 0.826 0.712 0.624 0.555 0.500
4.1 0.983 0.827 0.712 0.624 0.555 0.500
4.2 0.985 0.828 0.712 0.624 0.555 0.500
4.3 0.986 0.829 0.713 0.624 0.555 0.500
4.4 0.988 0.829 0.713 0.624 0.555 0.500
4.5 0.989 0.830 0.713 0.625 0.555 0.500
4.6 0.990 0.830 0.713 0.625 0.555 0.500
4.7 0.991 0.830 0.713 0.625 0.555 0.500
4.8 0.992 0.831 0.713 0.625 0.555 0.500
4.9 0.993 0.831 0.714 0.625 0.555 0.500
5 0.993 0.831 0.714 0.625 0.555 0.500

HEAT EXCHANGERS 611


TABLE 12.8 NTU Vs. effectiveness for counter-flow HX

NTU C= 0 C = 0.2 C = 0.4 C = 0.6 C = 0.8 C= 1


0.1 0.095 0.094 0.093 0.092 0.092 0.091
0.2 0.181 0.178 0.175 0.172 0.169 0.167
0.3 0.259 0.253 0.247 0.242 0.236 0.231
0.4 0.330 0.320 0.311 0.303 0.294 0.286
0.5 0.393 0.381 0.368 0.356 0.345 0.333
0.6 0.451 0.435 0.419 0.404 0.389 0.375
0.7 0.503 0.484 0.465 0.447 0.429 0.412
0.8 0.551 0.528 0.507 0.485 0.465 0.444
0.9 0.593 0.569 0.544 0.520 0.496 0.474
1 0.632 0.605 0.578 0.551 0.525 0.500
1.1 0.667 0.638 0.609 0.580 0.552 0.524
1.2 0.699 0.668 0.637 0.606 0.576 0.545
1.3 0.727 0.696 0.663 0.630 0.598 0.565
1.4 0.753 0.721 0.687 0.652 0.618 0.583
1.5 0.777 0.744 0.709 0.673 0.636 0.600
1.6 0.798 0.764 0.729 0.691 0.653 0.615
1.7 0.817 0.784 0.747 0.709 0.669 0.630
1.8 0.835 0.801 0.764 0.725 0.684 0.643
1.9 0.850 0.817 0.780 0.740 0.698 0.655
2 0.865 0.832 0.795 0.754 0.711 0.667
2.1 0.878 0.845 0.808 0.767 0.723 0.677
2.2 0.889 0.857 0.821 0.779 0.734 0.688
2.3 0.900 0.869 0.832 0.790 0.745 0.697
2.4 0.909 0.879 0.843 0.801 0.755 0.706
2.5 0.918 0.889 0.853 0.811 0.764 0.714
2.6 0.926 0.897 0.862 0.821 0.773 0.722
2.7 0.933 0.906 0.871 0.829 0.782 0.730
2.8 0.939 0.913 0.879 0.838 0.790 0.737
2.9 0.945 0.920 0.887 0.846 0.797 0.744
3 0.950 0.926 0.894 0.853 0.804 0.750
3.1 0.955 0.932 0.900 0.860 0.811 0.756
3.2 0.959 0.937 0.907 0.867 0.818 0.762
3.3 0.963 0.942 0.912 0.873 0.824 0.767
3.4 0.967 0.947 0.918 0.879 0.830 0.773
3.5 0.970 0.951 0.923 0.884 0.835 0.778
3.6 0.973 0.955 0.927 0.890 0.841 0.783
3.7 0.975 0.958 0.932 0.895 0.846 0.787
3.8 0.978 0.961 0.936 0.899 0.851 0.792
3.9 0.980 0.964 0.940 0.904 0.855 0.796
4 0.982 0.967 0.944 0.908 0.860 0.800
4.1 0.983 0.970 0.947 0.912 0.864 0.804
4.2 0.985 0.972 0.950 0.916 0.868 0.808
4.3 0.986 0.974 0.953 0.920 0.872 0.811
4.4 0.988 0.976 0.956 0.923 0.876 0.815
4.5 0.989 0.978 0.959 0.927 0.879 0.818
4.6 0.990 0.980 0.961 0.930 0.883 0.821
4.7 0.991 0.981 0.963 0.933 0.886 0.825
4.8 0.992 0.983 0.966 0.936 0.890 0.828
4.9 0.993 0.984 0.968 0.938 0.893 0.831
5 0.993 0.985 0.970 0.941 0.896 0.833

612 FUNDAMENTALS OF HEAT AND MASS TRANSFER


100 100
0
=

0
ax

=
/C m .25

x
25 a
80

0. Cm
in 0 80
Cm 0
0.5

in /
50
0. 5

m
C
5
Effectiveness e, %

7
0.7 1. 0

Effectiveness e, %
0
60 1.00 1.
60

Shell fluid Shell fluid


40 40

20 20
Tube fluid
Tube fluid
0 0
1 2 3 4 5
1 2 3 4 5
Number of transfer units NTU = AU/Cmin
Number of transfer units NTU = AU/Cmin

FIGURE 12.14 Shell and tube heat exchanger, with FIGURE 12.15 Shell and tube heat exchanger, with
one shell pass and 2, 4, 6 tube passes two shell passes and 4, 8, 12 tube passes

(iii) At a given value of NTU, effectiveness is maximum for C = 0, (i.e. for a condenser or evaporator), and
decreases as C increases.
(iv) For NTU less than about 0.3, effectiveness is independent of capacity ratio, C.
(v) For given NTU and C, a counter-flow heat exchanger has highest effectiveness and a parallel, flow heat
exchanger has the lowest effectiveness.
(vi) When C = 1 (i.e. capacity rates of both the fluids are equal, as in the case of a typical regenerator), maxi-
mum effectiveness of a parallel-flow heat exchanger is 50% only, whereas there is no such limitation for
a counter-flow HX. Therefore, for such applications, obviously, the counter-flow arrangement is pre-
ferred.
Example 12.9. Consider a heat exchanger for cooling oil which enters at 180°C, and cooling water enters at 25°C. Mass
flow rates of oil and water are: 2.5 and 1.2 kg/s, respectively. Area for heat transfer = 16 m2. Specific heat data for oil and
water and overall U are given: Cpoil = 1900 J/kgK; Cpwater = 4184 J/kgK; U = 285 W/m2 K. Calculate outlet temperatures
of oil and water for parallel and counter-flow HX. (M.U. 1995)
Solution. Here, the outlet temperatures of both the fluids are not known. Use of LMTD method would require an itera-
tive solution. i.e. to start with, assume outlet temperature of, say, hot fluid, Th2 and calculate the exit temperature of cold
fluid, Tc2 and then, the LMTD; then, calculate the heat transfer rate Q. From Q and capacity rates, recalculate Th2, and
compare this value with the initially assumed value; if they do not match, say, within 0.5 deg.C, repeat the iterative
cycle.
But, as will be shown below, Effectiveness–NTU method, offers a direct, straightforward solution:
Data:
mh := 2.5 kg/s mc := 1.2 kg/s Th1 := 180°C U := 285 W/(m2 K) A := 16 m2 Tc 1 := 25°C
Cph := 1900 J/(kgK) Cpc := 4184 J/(kgK)
Capacity rates:
Ch := mh ×Cph
i.e. Ch = 4.75 ´ 10 3 W/K
and, Cc := mc ×Cpc
i.e. Cc = 5.021 ´ 10 3 W/K
Therefore,
Cmin := Ch W/K (minimum capacity rate)
and, Cmax := Cc W/K (maximum capacity rate)
Therefore, Capacity ratio:
Cmin
C :=
Cmax
HEAT EXCHANGERS 613
i.e. C = 0.946 (capacity ratio)
Number of Transfer Units:
U ×A
NTU :=
Cmin
i.e. NTU = 0.96
Case (i): Parallel-flow HX:
For parallel-flow HX, we have the effectiveness relation:

LM F Cmin I OP
1 - exp - NTU × 1 +
MN GH Cmax JK PQ
e= ...(12.50)
Cmin
1+
Cmax

1 - exp(- NTU × (1 + C))


e :=
1+ C
i.e. e = 0.435 (effectiveness of parallel-flow HX)
Then, since hot fluid is the ‘minimum’ fluid, we have:
Th1 - Th 2
e=
Th1 - Tc1
i.e. Th2 := Th1 – e×(Th1 – Tc1)
i.e. Th2 = 112.65°C (exit temperature of hot fluid (oil))
and, Tc2 is obtained from heat balance:
Ch × (Th1 – Th2) = Cc ×(Tc2 – Tc1) (by heat balance)
Ch × (Th 1 - Th 2 )
i.e. Tc2 := Tc1 +
Cc
i.e. Tc2 = 88.718°C (exit temperature of cold fluid (water))
Case (ii): Counter-flow HX:
For counter-flow HX, we have:
1 - exp( - NTU ×(1 - C))
e := ...(12.56)
(1 - C × exp(- NTU × (1 - C)))
i.e. e = 0.496 (effectiveness of counter-flow HX)
Then, again, since hot fluid is the ‘minimum’ fluid, we have:
Th1 - Th 2
e=
Th1 - Tc1
i.e. Th2 := Th1 – e×(Th1 – Tc1)
i.e. Th2 = 103.074°C (exit temperature of hot fluid (oil).)
And, Tc2 is obtained from:
Ch × (Th1 - Th 2 )
Tc2 := Tc1 +
Cc
i.e. Tc2 = 97.777°C (exit temperature of cold fluid (water).)
Note: In this problem, it is difficult to read accurately the e values from the graphs, for the given values of NTU
and C. It is suggested that the analytical relations may be used to get accurate results.
(b) In the above Example, suppose that the flow rate of water is increased to 2 kg/s.
Calculate the new outlet temperatures of oil and water for parallel and counter-flow HX. Rest of the data remain the
same.
Now, the heat exchanger is operated at an off-design condition, i.e. the water flow is changed from 1.2 kg/s to 2
kg/s. Then, e–NTU method is convenient to use to find out the exit temperatures of both the fluids.
mc := 2 kg/s (mass flow rate of cold fluid (water))
Note that still, hot fluid is the ‘minimum’ fluid and NTU remains the same, but C changes:
Capacity rates:
Ch := mh ×Cph

614 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Ch = 4.75 ´ 10 3 W/K
and, Cc := mc ×Cpc
i.e. Cc = 8.368 ´ 10 3 W/K
Therefore,
Cmin := Ch W/K (minimum capacity rate)
and, Cmax := Cc W/K (maximum capacity rate)
Therefore, Capacity ratio:
Cmin
C :=
Cmax
i.e. C = 0.568 (capacity ratio)
U ×A
and, NTU :=
Cmin
i.e. NTU = 0.96
Case (i): Parallel-flow HX:
We have the effectiveness relation:
1 - exp(- NTU × (1 + C))
e :=
1+ C
i.e. e = 0.496 (effectiveness of parallel-flow HX)
Compare this e with e = 0.435 obtained earlier.
Then, since hot fluid is the ‘minimum’ fluid, we have:
Th1 - Th 2
e=
Th1 - Tc1
i.e. Th2 := Th1 – e×(Th1 – Tc1)
i.e. Th2 = 103.079°C (exit temperature of hot fluid (oil).)
And, Tc2 is obtained from heat balance:
Ch = (Th1 – Th2) = Cc ×(Tc 2 – Tc1) (by heat balance)
Ch × (Th1 - Th 2 )
i.e. Tc2 := Tc1 +
Cc
i.e. Tc2 = 68.664°C (exit temperature of cold fluid (water).)
Case (ii): Counter-flow HX:
For Counter-flow HX, we have:
1 - exp( - NTU ×(1 - C))
e := ...(12.56)
(1 - C × exp(- NTU × (1 - C)))
i.e. e = 0.543 (effectiveness of counter-flow HX)
Compare this e with e = 0.496 obtained earlier.
Now, again, since hot fluid is the ‘minimum’ fluid, we have:
i.e. Th2 := Th1 – e×(Th1 – Tc1)
i.e. Th2 = 95.779°C (exit temperature of hot fluid (oil).)
And, Tc2 is obtained from:
Ch × (Th1 - Th 2 )
Tc2 := Tc1 +
Cc
i.e. Tc2 = 72.807°C ...exit temperature of cold fluid (water).
Note that as a result of increasing the cold fluid (water) flow rate, the new exit temperature of both the hot and cold
fluids are lower, for both the parallel and counter-flow cases.
Example 12.10. A steam condenser, condensing at 70°C has to have a capacity of 100 kW. Water at 20°C is used and the
outlet water temperature is limited to 45°C. If the overall heat transfer coefficient is 3100 W/m2 K, determine the area
required.
(b) If the inlet water temperature is increased to 30°C, determine the increased flow rate of water to maintain the
same outlet temperature. (M.U. 1998)

HEAT EXCHANGERS 615


Solution. This problem can be solved by LMTD method, too. But, in part (b), since the heat exchanger is operated at an
off-design condition, we shall adopt the e–NTU method.
Data:
Q := 100 ´ 103 W Th := 70°C U := 3100 W/(m2 K) Tc1 := 20°C Tc2 := 45°C
Cph := 1900 J/(kgK) Cpc := 4180 J/(kgK)
Therefore, effectiveness:
Since steam is condensing, it is the ‘maximum’ fluid. So, we can write:
Tc 2 - Tc1
e :=
Th - Tc1
i.e. e = 0.5 (effectiveness)
Now, for a condenser, we have, from Table 12.6:
NTU := – ln(1 – e) (for a condenser)
i.e. NTU = 0.693
Q
and, Cmin :=
e × (Th - Tc1 )
i.e. Cmin =4 ´ 10 3 W/K
U ×A
But, NTU = (by definition of NTU)
Cmin
Therefore,
NTU ×Cmin
A :=
U
i.e. A = 0.894 m2 (area of heat transfer.)
Case (b): If Tc1 is increased to 30°C, and Tc2 maintained at 45°C, what is the increased flow rate?
Tc1 := 30°C (new inlet temperature of water)
Then,
Tc 2 - Tc1
e :=
Th - Tc1
i.e. e = 0.375 (new effectiveness)
Therefore, new NTU:
NTU := – ln(1 – e) (for a condenser)
i.e. NTU = 0.47 (new value of NTU for case (b))
Therefore,
U ×A
Cmin2 :=
NTU
i.e. Cmin2 = 5.899 ´ 10 3 W/K (new Cmin for case (b))
Compare this value with C min = 4000 obtained earlier.
Therefore, increased flow rate:
We have: Cpc := 4180 J/kgK (specific heat for water)
Therefore,
Cmin
m 1 :=
Cpc
i.e. m 1 = 0.957 kg/s (earlier flow rate)
Cmin 2
and, m 2 :=
Cpc
m 2 = 1.411 kg/s (new flow rate)
Also,
Cmin 2
F :=
Cmin
Or, F = 1.475 (increase of 47.5%.)

616 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Example 12.11. Hot oil at a temperature of 180°C enters a Oil, Th1 = 180 C
shell and tube HX and is cooled by water entering at 25°C. 0.4 kg/s 6 tube passes
There is one shell pass and 6 tube passes in the HX and the
overall heat transfer coefficient is 350 W/(m2 K). Tube is
thin-walled, 15 mm ID and length per pass is 5 m. Water
flow rate is 0.3 kg/s and oil flow rate is 0.4 kg/s. Deter-
mine the outlet temperatures of oil and water and also the Tc2
heat transfer rate in the HX. Given: specific heat of oil =
1900 J/(kgK) and specific heat of water = 4184 J/(kgK)
Solution. Since the exit temperatures of both the fluids are
not known, we shall use e – NTU method.
Water, Tc1 = 25 C
Data:
0.3 kg/s
mh := 0.4 kg/s mc := 0.3 kg/s Th1 := 180°C
U := 350 W/(m2 K) Tc1 := 25°C N:= 6
D := 0.015 m L := 5 m Cph := 1900 J/(kgK) Th2
Cpc := 4184 J/(kgK)
FIGURE Example 12.11 Shell and tube heat
Capacity rates:
exchanger with one shell pass and 6 tube passes
Ch := mh ×Cph
i.e. Ch = 760 W/K
and, Cc := mc ×Cpc
i.e. Cc := 1.255 ´ 10 3 W/K
Therefore, oil is the ‘minimum’ fluid.
i.e. Cmin := Ch W/K (minimum capacity rate)
and, Cmax := Cc W/K (maximum capacity rate)
Therefore, capacity ratio:
Cmin
C :=
Cmax
i.e. C := 0.605 (capacity ratio)
Number of Transfer Units:
A := N×p×D×L
i.e. A = 1.414 m2 (area of heat transfer)
U ×A
and, NTU :=
Cmin
i.e. NTU = 0.651 (Number of Transfer Units)
Effectiveness:
This is a shell and tube HX with one shell pass and 6 tube passes. So, its effectiveness can be determined for C =
0.605 and NTU = 0.651, from Fig. 12.14.
Since it is difficult to read from the graph accurately, let us calculate e from analytical relation given in Table 12.5:
N := NTU (notation in following equation)

LM LM
1 + exp - N × (1 + C 2 ) 2
1
OP OP -1

M
e := 2× M1 + C + (1 + C )
1
MN PQ PP
L OP P (one shell pass, 2, 4 , 6 tube passes)
2 2
×
MM 1 - exp M - N × (1 + C )
1

PQ PQ
2 2

N MN
i.e. e = 0.415 ...effectiveness
Actual heat transfer, Q:
We have: Q= e×Q max
where, Q max := Cmin×(Th 1 – Tc1) W (maximum possible heat transfer in the HX)
i.e. Qmax = 1.178 ´ 10 5 W
and, Q := e×Qmax
i.e. Q= 4.884 ´ 104 W (actual heat transfer in the HX.)

HEAT EXCHANGERS 617


Outlet temperatures of hot and cold fluids:
We have: Q = Ch ×(Th1 – Th2) = Cc ×(Tc2 – Tc1)
Therefore,
Q
Th2 := Th1 –
Ch
i.e. Th2 := 115.742°C (outlet temperature of hot fluid (i.e. oil))
Q
and, Tc2 := Tc1 +
Cc
i.e. Tc2 = 63.907°C (outlet temperature of cold fluid (i.e. water).)
Example 12.12. A feed water heater heats water entering at a temperature of 25°C, at a rate of 3 kg/s. Heating is due to
steam condensing at 117°C. When the feed water heater was new (i.e. ‘clean’ condition), the exit temperature of water
was 85°C. After prolonged operation, for the same flow rates and inlet conditions, it was observed that the outlet tem-
perature was 75°C. Determine the value of fouling factor. Given: area of heat exchange = 5.5 m2.
Solution. Fouling resistance, Rf is calculated from the relation:
1 1
Rf = - m2 K/W (fouling factor)
Udirty U clean
Also, since the steam is condensing, it is the ‘maximum’ fluid, and the water is the ‘minimum fluid.
Data:
mc := 3 kg/s Th := 117°C U := 350 W/(m2 K) Tc1 := 25°C
Tc2 := 85°C (exit temperature of cold fluid (water) for ‘clean’ HX) Cpc := 4180 J/(kgK) A := 5.5 m2
Capacity rates:
Cc := mc ×Cpc (Capacity rate of cold fluid (water))
i.e. Cc = 1.25 ´ 104 W/K (capacity rate of cold fluid (water))
Since this is a condenser, water is the ‘minimum’ fluid, and capacity rate of condensing steam is ¥.
i.e. Cmin := Cc
Therefore, capacity ratio:
Cmin
C=
Cmax
i.e. C =0 (for a condenser)
Effectiveness:
Remembering that water is the minimum fluid, effectiveness is given by:
Tc 2 - Tc1
e :=
Th - Tc1
i.e. e = 0.652 (effectiveness of the condenser)
and, from Table 12.6, NTU of the condenser is given by:
NTU := – ln(1 – e)
i.e. NTU = 1.056
But, by definition of NTU:
U ×A
NTU =
Cmin
NTU×Cmin
Therefore, U clean :=
A
i.e. Uclean = 2.408 ´ 103 W/(m2 K) (overall heat transfer coefficient for ‘clean’ HX.)
After prolonged operation:
Tc2 := 75°C (exit temperature of water for ‘dirty’ HX.)
Therefore, effectiveness of dirty HX:
Tc 2 - Tc1
e :=
Th - Tc1
i.e. e = 0.543 (effectiveness of ‘dirty’ HX.)

618 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore, NTU of condenser:
NTU := – ln(1 – e)
i.e. NTU = 0.784 (for ‘dirty’ HX)
NTU ×Cmin
Then, Udirty :=
A
Udirty = 1.788 ´ 103 W/(m2 K) (overall heat transfer coefficient for ‘dirty’ HX)
Therefore, Fouling factor:
1 1
Rf := - m2 K/W (fouling factor)
Udirty Uclean
i.e. Rf = 0.000144 m2 K/W (fouling factor.)
Example 12.13. Oil at 100°C (Cp = 3.6 kJ/kgK) flows at a rate of 30,000 kg/h and enters into a parallel-flow HX. Cooling
water (Cp = 4.2 kJ/kg.K) enters the HX at 10°C at the rate of 50,000 kg/h. The heat transfer area is 10 m2 and U = 1000
W/(m2K). Calculate the following: (i) outlet temperature of oil and water (ii) maximum possible outlet temperature of
water.
Solution. Exit temperature of both the fluids are not known; therefore, NTU method is to be used:
Data:
30000 50000
mh := kg/s i.e. mh = 8.333 mc := kg/s i.e. mc = 13.889 Th1 := 100°C U := 1000 W/(m2 K)
3600 3600
2
Tc1 := 10°C Cph := 3600 J/(KgK) Cpc := 4200 J/(kgK) A := 10 m
Capacity rates:
Ch := mh × Cph
i.e. Ch = 3 ´ 104 W/K
and, Cc := mc × Cpc
i.e. Cc = 5.833 ´ 104 W/K
Therefore, oil is the ‘minimum’ fluid.
i.e. Cmin := Ch W/K (minimum capacity rate)
and, Cmax := Cc W/K (maximum capacity rate)
Therefore, Capacity ratio:
C
C := min
Cmax
i.e. C = 0.514 (capacity ratio)
U×A
and, NTU :=
Cmin
i.e. NTU = 0.333 (Number of Transfer Units)
Effectiveness:
For parallel flow HX, we have:
1 - exp(- NTU × (1 + C))
e := ...(12.51)
1+ C

Temperature Temperature

Th1 = 100 C Hot fluid Th1 = 100 C

Th2
Th2 = Tc2
Tc2

Tc1 = 10 C Cold fluid Tc1 = 10 C

Length Length
(a) (b)

FIGURE Example 12.13 Parallel-flow heat exchanger

HEAT EXCHANGERS 619


i.e. e = 0.262 (effectiveness of parallel flow HX with NTU = 0.333 and C = 0.514.)
Note: We can use the graph of Fig. 12.10 or Table 12.7, but using the analytical relation is more accurate.
Outlet temperature of hot and cold fluids:
Th1 - Th 2
Since hot fluid is the minimum fluid, e =
Th1 - Tc1
Therefore,
Th2 := Th1 – e×(Th1 – Tc1)
i.e. Th2 = 76.443°C (outlet temperature of hot fluid (i.e. oil).)
And, from heat balance:
Ch × (Th1 – Th2) = Cc ×(Tc2 – Tc1)
Ch × (Th1 - Th 2 )
Or, Tc2 := Tc1 +
Cc
i.e. Tc2 = 22.115°C (outlet temperature of cold fluid (i.e. water).)
(b) Maximum possible outlet temperature of water:
For a very long parallel-flow HX, the outlet temperatures of hot and cold fluids would be the same:
i.e. Th2 = Tc2
Therefore, writing the heat balance:
Ch × (Th1 – Th2) = Cc ×(Tc2 – Tc1)
i.e. Ch × (Th1 – Tc2) = Cc ×(Tc2 – Tc1)
i.e. Tc2 ×(Cc + Ch) = Cc ×Tc1 + Ch ×Th1
Cc × Tc1 + Ch × Th1
i.e. Tc2 :=
Cc + C h
i.e. Tc2 = 40.566°C (maximum possible outlet temperature of water.)

12.7 The Operating-line/Equilibrium-line Method


NTU–e method can be represented graphically in another way.
Refer to Fig. 12.16. Here, the x-axis represents the cold fluid temperature and the y-axis, the hot fluid tem-
perature. Now, if we plot the entrance and exit temperatures of a heat exchanger on these axes, we see that the
operating range of the HX is represented by a single line; this line is called ‘the operating line’. On the same
graph, a line drawn at 45 deg. is called ‘the equilibrium line’. For equilibrium line, Th = Tc. Thermodynamically,
it is impossible for the operating line of a heat exchanger to drop below the equilibrium line, since, if it does, it
would mean a violation of the second law. Slope of the operating line for the counter-flow HX is: (Cc/Ch) = (Th1–
Th2)/(Tc2–Tc1). And, the slope of the operating line for the parallel-flow HX is: – (Cc/Ch), i.e. negative slope. For a
condenser, operating line is horizontal with Th = constant and (Cc/Ch) = 0, and for an evaporator, the operating
line is a vertical line with Tc = constant, and (Cc/Ch) = ¥.
Advantage of this method of representation is that the effectiveness of the heat exchanger can now be shown
geometrically as a ratio of two lengths. For example, for the counter-flow HX shown in Fig. 12.16 (a), we have: Cc
> Ch and the effectiveness is equal to d/D.
For constant specific heats of fluids, the operating line is a straight line. Variation in specific heats of fluids is
also shown easily in these graphs: As shown in Fig. 12.16 (d), if the operating line curves upwards, i.e. the slope
increases as the temperature increases, it means that Cp of cold fluid increases with temperature (or, the Cp of hot
fluid decreases with temperature). Similarly, if the operating line curves downwards, it means that Cp of cold
fluid decreases with temperature (or, the Cp of hot fluid increases with temperature).
Effectiveness of a parallel-flow heat exchanger:
Operating-line/Equilibrium line method can be used to determine the effectiveness of a heat exchanger. Let us
illustrate this briefly with reference to a parallel-flow HX:
Refer to Fig. 12.17. Line 1–2 is the operating line for the parallel-flow HX. We see from the figure that Cc < Ch ,
since the slope of the operating line = –Cc/Ch .
And, Capacity ratio, C = Cc/Ch .
From the Fig 12.17:
Th1 – Tc1 = D
Tc2 – Tc1 = a 1

620 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Operating line, counter-flow HX Operating line, parallel-flow HX
Th slope = Cc/Ch, Î = d/D Th slope = Cc/Ch Î = d/D

Th1 Equilibrium line, Th1 D


Th = Tc
D d Equilibrium line,
Th = Tc
Th2 Th2 d

Tc1 Tc2 Tc Tc1 Tc2 Tc


(a) (b)
Condenser, Cc/Ch = 0, Cpc decreases with temperature
Th = const. (or, Cph increases with temperature)
Th Th
Evaporator/boiler, 2
Equilibrium line, 2 Equilibrium line
Cc/Ch = ¥, Th = Tc
Th = Tc 1
Tc = const.
Cpc increases with temperature
Regenerator, Cc /Ch = 1
(or, Cph decreases with temperature)
1

Tc Tc
(c) (d)
FIGURE 12.16 Operating line and equilibrium lines for heat exchangers
a1 Operating line for counter-flow HX
Therefore, e=
D Th a1 slope = Cc/Ch, Î = d/D
For parallel-flow HX, we have: Equilibrium line,
1
I = – U×A× FG 1 I Th1
FT
Th = Tc
- Tc 2 D
ln GH T h2
JK
h1 - Tc1 H m ×C h ph
+
1
mc × Cpc
JK ...(12.21)
2 b2
Th2
F T - T I = – U×A× F 1
ln G
I= D
b1
i.e.
H T - T JK
h2
h1
c2
c1
GH C h
+
1
Cc JK d

-U ×A
×(1 + C) ...(A)
Cc
45
Now, from the Fig. 12.17 we have:
Th2 – Tc2 = b 1 Tc1 Tc2 Tc
Th1 – Tc1 = D
FIGURE 12.17 Parallel-flow heat exchanger
We also see from the Fig. 12.17:
b 1 = b 2 – (Th1 – Th2)
But, b 2 = D – a 1 (from the Fig. 12.17, since equilibrium line is at 45 deg. to horizontal.)
Therefore, b 1 = (D – a 1) – (Th1 – Th2)
Cc
i.e. b 1 = (D – a 1) – ×a 1
Ch

b1 a1 Cc a1
Therefore, =1– – ×
D D Ch D
HEAT EXCHANGERS 621
b1
i.e. = 1 – e – C×e
D
b1
i.e. = 1 – e×( 1 + C) ...(B)
D
Substituting in Eq. A:
b1
= exp(– NTU×(1 + C))
D
Using Eq. B:
1 – e×(1 + C) = exp(– NTU×(1 + C))
1 - exp (- NTU ×(1 + C ))
i.e. e= ...(C)
1+ C
Eq. C is the desired equation for the effectiveness of the parallel-flow HX.
This is the same as the equation derived earlier for parallel-flow HX, i.e. Eq. 12.51. While deriving Eq. C, it
was assumed that the cold fluid was the ‘minimum’ fluid; if we assume that the hot fluid is the minimum fluid,
then also, the same result would be obtained.

12.8 Compact Heat Exchangers


Heat exchangers with an area density greater than about 700 m2/m3 are classified as ‘compact heat exchangers’.
Generally, they are used for gases.
Compact heat exchangers are, typically, of three types:
(i) array of finned circular tubes
(ii) array of plate-fin matrix, and
(iii) array of finned flat-tube matrix.
Heat transfer and pressure drops for these compact heat exchangers are determined experimentally and are
supplied by manufacturers as their proprietary data.
As an example, a plate-fin type of heat exchanger matrix, manufactured by Marston–Excelsior Ltd., is shown
in Fig. 12.18. As shown in the Fig.12.18, a single element consists of two plates in between which is sandwiched
a corrugated sheet. The two edges are sealed. Dip brazing technique is used to build a complete heat exchanger
block from individual elements. Multi-flow configurations are possible, and the generally used corrugations
types are: plain (P), plain-perforated (R), serrated (S) and herringbone (H).
Table 12.9 gives the geometrical data for some typical corrugations.

Elements assembled to form


a cross-flow arrangement

Elements assembled to form


Exloded photograph of a single element
a contra-flow arrangement

FIGURE 12.18 Plate-fin heat exchangers for cryogenic service (Marston–Excelsior Ltd.)

622 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 12.9 Geometrical data for typical corrugations (Marston Excelsior Ltd.)

Type Height (mm) Thickness (mm) Pitch (fins/m) a (m2/m) A1 (m2/m) A2 (m2/m) Dh (mm)
P, R, S, H 3.8 0.20 470 0.00326 1.81 3.41 2.5
P, R 5.1 0.20 550 0.004328 1.776 5.376 2.42
P, R 5.1 0.20 1020 0.00387 1.584 9.984 1.34
P, R, S, H 6.35 0.30 470 0.005175 1.712 5.712 2.79
P, R, H 8.9 0.46 590 0.00615 1.46 9.96 2.16
P, R 8.9 0.61 240 0.00707 1.712 3.912 5.04

In the above table,


a = free flow area per metre width of corrugation
A 1 = (primary surface area per metre width) x (metre length of corrugation)
A 2 = (secondary surface area per metre width) x (metre length of corrugation)
Dh = hydraulic mean diameter
i.e.
4×a
Dh =
wetted perimeter
4×a
i.e. Dh = (for areas specified above.)
( A1 + A2 )
Kays and London have studied a large number of compact heat exchanger matrices and presented their
experimental results in the form of generalised graphs. Heat transfer data is plotted as St.Pr2/3 against Re, where,
St = Stanton number = h/(G.Cp), Pr = Prandtl number = m.Cp/k, and Re = G.Dh/m, G = mass velocity (= mass flow
rate/Area of cross section), kg/(sm2.)
In the same graphs, friction factor, f, is also plotted against Re.
As an example, heat transfer and friction factor characteristics for a particular tube-fin matrix are shown in
Fig.12.19.
Pressure drop in plate-fin heat exchangers:
Total pressure drop for the fluid flowing across the heat exchanger is given by:

G2 LM F A I OP
DP =
2× r i MN
r
ro GH
× ( Kc + 1 - s ) + 2 × i - 1 + f ×
2
JK
r
Amin r m
r
× i - (1 - Ke - s 2 ) × i
ro PQ
...(12.59)

0.060 1.00¢¢

0.040
0.402
0.125¢¢
0.030 0.866¢¢ d = 0.013 in. Tube outside diameter = 0.402 in.
Fin pitch = 8.0 per in.
f

0.020 Flow passage hydraulic diameter, 4rh = 0.01192 ft.


Fin thickness = 0.013 in.
Free-flow area/frontal area, s = 0.534
Heat transfer area/total volume, a = 179 ft2/ft3
0.010 Fin area/total area = 0.913
2/ 3

0.008 Note: Minimum free-flow area in spaces transverse to flow.


Ns1×Np1

0.006
–3
NR×10
0.004
0.4 0.6 0.8 1.0 2.0 3.0 4.0 6.0 8.0 10.0

FIGURE 12.19 Heat transfer and friction factor for plate-finned circular tube matrix (Trane Company)

HEAT EXCHANGERS 623


Amin minimum free flow area
where s= =
A fr frontal area

A 4×L total heat transfer area


= =
Amin Dh minimum free flow area
r× u × A fr r ×u
G= = = mass velocity ...kg/sm2
Amin s
Kc and Ke = flow contraction and expansion coefficients, respectively
ri and ro = density at inlet and exit, respectively
F I
1
rm
1 1
= × +
1
2 ri ro
GH
. JK
In Eq. 12.59, on the RHS, the first term inside the square brackets represents the entrance contraction effect,
second term—the flow acceleration, the third term—core friction and the fourth term—the exit expansion effect.
Core friction drop is generally 90% of the total pressure drop. For liquids, entrance and exit losses are negligible.
Values of Kc and Ke are given in Kays and London.
Pressure drop for finned-tube exchangers: Entrance and exit effects are included in the friction factor;
therefore, Kc = Ke = 0. Then, total pressure drop across the tube bank is:

G2 LM F I A ri OP
DP =
2× ri MN
r
roGH JK
× (1 + s 2 ) × i - 1 + f × ×
Amin r m PQ
...(12.60)

Here, first term on the RHS is the flow acceleration effect, and the second term is the core friction.
Fig. 12.20 shows a plate-fin exchanger for an ethylene plant and Fig. 12.21 shows another plate-fin exchanger
for an air liquefier.
Regenerators:
Regenerators are extensively used in blast furnace stoves, open hearth furnaces, coke manufacture, glass produc-
tion, for air pre-heating in power plants, in gas turbine systems and in cryogenic plants, in Stirling cycle air (or
helium) liquefiers, in cryogenic mini-coolers used for cooling infrared detectors, etc. In a regenerator, hot and
cold fluids flow alternately through the regenerator matrix. The matrix may be sand-lime bricks, metal packings,
wire screen mesh, lead balls, etc., depending upon application. During the ‘hot blow’ hot fluid flows through the

FIGURE 12.20 A heat exchanger assembly with associated pipe work for an ethylene plant
(Marston Excelsior Ltd.)

624 FUNDAMENTALS OF HEAT AND MASS TRANSFER


matrix and the matrix absorbs the heat from the fluid;
during the ‘cold blow’, cold fluid flows through the
matrix and the matrix gives up the absorbed heat to the
cold fluid, thus heating the fluid. Thus, suitable valving
is necessary to alternately switch the hot and cold fluids
through the regenerator. In a valved type of exchanger,
generally, two identical matrices are provided such that
when one matrix is being heated, the other is being
cooled. Alternately, regenerator may be of rotary type,
where a porous matrix is rotated around its axis cutting
the hot and cold fluid lines, thus transferring heat from
the hot to the cold fluid.
Analysis of a periodic flow HX is complicated since
the matrix and gas temperatures vary with both posi-
tion and time. A rough outline of the analysis is given
below:
Refer to Fig. 12.22, which shows a regenerator dia-
grammatically. Hot fluid flows through the matrix dur-
ing the hot blow and heats the matrix. Then, the flow is
switched to effect the cold blow and the cold fluid flows
FIGURE 12.21 Air liquefier for a 400
through the matrix and gets heated up. Thus, in effect,
Ton/day Oxygen plant using two, 763 mm ´ 763
heat is transferred from the hot fluid to the cold fluid.
mm blocks in parallel (Marston Excelsior Ltd.)
We are interested in the gas and matrix tempera-
tures at any location and at any time. These are obtained
t, tg Regenerator filling by writing an energy balance for an element of width
mc, in dx, shown in the Fig. 12.22. Following notations are
mh , in
x dx used:
Ms = Mass of solid (matrix filling) per unit length,
kg/m
M = Mass flow rate of gas, kg/s
Cps = specific heat of solid, J/(kgK)
Cpg = specific heat of gas, J/(kgK)
L
V = free volume per unit length
mh, out mc, out A = heat transfer area per unit length
r = density of gas, kg/m3
FIGURE 12.22 Periodic-flow heat exchanger L = length of matrix column
(Regenerator)
h = convective heat transfer coefficient between the
gas and the matrix
t = solid temperature at a given location x
tg = gas temperature at location x
Writing the heat balance:
Heat transferred by convection between the gas and the solid = Heat stored in the solid
i.e.
dt
h×A×(tg – t)×dx×dt = Ms ×Cps ×dx× ×dt ...(12.61)
dt
Now, the heat transferred by convection is also equal to the heat stored in the gas contained in length dx plus
the increase in the enthalpy of the gas as it passes through the element dx.
i.e.
dtg F dt ×dxI ×dt
h×A×(tg – t)×dx×dt = r×Cpg ×(V ×dx)×
dt
×dt + M×Cpg × GH dx JK
g
...(12.62)

HEAT EXCHANGERS 625


Above equations are simplified as:
¶t/¶t = {(h.A)/(Cps.Ms)}.(tg – t) ...(12.63)
¶tg/¶x + (r.V/M).¶tg/¶t = {(h.A)/(Cpg.M)}.(t – tg) ...(12.64)
In most of the practical situations, the term (r.V/M) is very small and is neglected. Then, making following
substitutions
h× A× x h× A
x= and h = ×t
Cpg × M Cps × Ms
the resulting equations are solved with the following boundary conditions:
t = to at h = 0 (initial solid temperature for all x)
and, tg = tgo at x = 0 (inlet gas temperature for all h)
The results are presented usually in graphical form and the nature of graphs is shown in Fig. 12.23 (a) and
(b).

(tg – tgo)/(to – tgo) (t – tgo)/(to – tgo)


1 1 h=1
h=0
3
3
7
7 10
20

0 0
0 x 10 0 x 10

(a) To find gas temperature (b) To find solid temperature

FIGURE 12.23 Gas and solid temperature charts for a regenerator

In these graphs, tgo is the initial temperature of the gas, and to is the initial temperature of the solid.
Fig. 12.23 (a) presents the dimensionless gas temperature at any location as a function of x and h and Fig.
12.23 (b) shows the dimensionless solid temperature as a function of x and h.
Effectiveness–NTU relations for regenerator:
Effectiveness of a regenerator is presented as a function of three dimensionless parameters, as follows:
F C Cr I
GH
e = f NTU mod , min ,
Cmax Cmin JK ...(12.65)

where, NTUmod = modified NTU, given by:

LM OP
1
×M
M 1 PP
MM FG 1 IJ + FG 1 IJ PP
NTUmod = ...(12.66)
Cmin

N H h× A K H h× A K Q
c h

and, matrix capacity rate is equal to matrix mass rate times the specific heat of the solid.
For the rotary type of regenerator,
FG Rev IJ ×(matrix mass)×C
Cr =
H sK ps W/K ...for rotary type regenerator...(12.67).

For the valved type of regenerator, total mass of both the identical matrices is used, multiplied by valve
cycles/s, where period is the interval between ‘valve–on-to–off–to–on’.
Kays and London have presented e–NTUmod graphs for different Cr/Cmin ratios (ranging from 1 to infinity),
for given Cmin/Cmax ratios (ranging from 0.5 to 1). Table 12.10 is a sample table showing e values for Cmin/Cmax =
1. Fig. 12.24 presents this table in graphical form.

626 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 12.10 Effectiveness of periodic flow HX (Cmin/Cmax = 1)
NTUmod Cr /Cmin = 1 Cr /Cmin = 1.5 Cr /Cmin = 2 Cr /Cmin = 5 Cr /Cmin = Infinity
0 0 0 0 0 0
0.5 0.322 0.328 0.33 0.333 0.333
1 0.467 0.485 0.491 0.499 0.5
1.5 0.548 0.576 0.586 0.598 0.6
2 0.601 0.636 0.649 0.664 0.667
2.5 0.639 0.679 0.694 0.711 0.714
3 0.667 0.712 0.728 0.746 0.75
3.5 0.69 0.738 0.755 0.774 0.778
4 0.709 0.759 0.776 0.796 0.8
4.5 0.724 0.776 0.794 0.814 0.818
5 0.738 0.791 0.809 0.829 0.833
5.5 0.749 0.803 0.821 0.842 0.846
6 0.759 0.814 0.832 0.853 0.857
6.5 0.768 0.824 0.842 0.862 0.867
7 0.776 0.833 0.85 0.87 0.875
7.5 0.784 0.84 0.858 0.878 0.882
8 0.79 0.847 0.865 0.884 0.889
8.5 0.796 0.854 0.871 0.89 0.895
9 0.802 0.859 0.876 0.895 0.9
9.5 0.807 0.864 0.881 0.9 0.905
10 0.811 0.869 0.886 0.904 0.909

0.9 Infinity
5
2

0.8 1.5
Effectiveness

Cr
0.7 =1
Cmin

0.6

0.5

0.4
0 1 2 3 4 5 6 7 8 9 10
Modified NTU
FIGURE 12.24 Effectiveness of a periodic-flow HX (regenerator) for Cmin/Cmax = 1

Higher NTUmod ranges (of the order of 100 or more) are generally applicable to regenerators used in cryo-
genic applications; in such cases, since the effectiveness approaches unity asymptotically, for better clarity,
graphs are plotted with (1 – e) vs. NTUmod. (See Appendix at the end of chapter).
To calculate NTUmod we need the heat transfer coefficients on the cold and hot fluid sides. We also need the
heat transfer area. Heat transfer characteristics in terms of Colburn j factor vs. Reynolds number, and friction
factor vs. Reynolds number are presented for many types of matrices by Kays and London. They also provide

HEAT EXCHANGERS 627


physical data such as hydraulic diameter, heat transfer area, porosity, etc., for those matrices. As an example,
heat transfer characteristics and friction factor data for a randomly stacked, wire screen matrix (used typically, in
gas turbine regenerators) are shown in Fig. 12.25 and Fig. 12.26, respectively.
Advantages of regenerators:
(i) High surface density, of the order of 3000 m2/m3 (for a 24 mesh screen matrix, typically used in gas
turbine regenerators), can be packed into a given volume
(ii) Tends to be ‘self-cleaning’ because of periodic flow reversals
(iii) Cheaper on per unit heat transfer area basis.
Disadvantages:
(i) Some mixing of hot and cold fluids is unavoidable
(ii) Sealing between the fluids presents some problem if the pressure differential is large.

0.5
0.6
0.4

0.2

0.1
2/3
×N Pr

0.8
0.06
Porosity
G×cp
h

0.04 0.832
0.817
0.02 0.766
0.725
0.675
0.01
0.602
0.08
0.06
0.04

0.02
4×rh×G
NR =
m
0.001
2 3 4 5
1 2 4 6 8 10 2 4 6 810 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

FIGURE 12.25 Colburn j factor vs. Reynolds number for a randomly sacked wire screen matrix

10
8
6

1
0.8
Porosity
0.6
0.832
f

0.766
0.4
0.725
0.3 0.602

0.2

NR = 4rh G/m
0.1
2 3 4 5
1 2 3 4 6 8 10 2 3 4 6 8 10 2 3 4 6 8 10 2 3 4 6 8 10 2 3 4 6 8 10

FIGURE 12.26 Friction factor vs. Reynolds number for a randomly stacked wire screen matrix

628 FUNDAMENTALS OF HEAT AND MASS TRANSFER


12.9 Hydro-mechanical Design of Heat Exchangers
So far, we studied thermal design aspects for a heat exchanger. But, from a practical point of view, the pressure
drop that occurs when the fluid passes through the heat exchangers and the pumping power required to effect
this flow, are also important. We shall only briefly mention about this aspect.
Obviously, flow of fluid through heat exchanger passages involves pressure drop. And, higher the viscosity
of the fluid, higher the pressure drop. Total pressure drop in a heat exchanger section is calculated by summing
up the following individual pressure drops:
(a) Pressure drops in straight passages and pipe bends, DPf
(b) Pressure drops due to ‘end effects’, i.e. due to flow contraction and expansion at the ends, DPe
(c) Pressure drops due to flow acceleration (in cases of gases in non-isothermal flow), DPa, and
(d) Pressure drop due to self draught (due to buoyant forces) as a result of change in elevation of flow
channels, DPs.
(a) Pressure drops in straight passages and bends These are determined by Darcy formula, as explained in the
chapter on convection.

L r ×V 2
DPf = fD × × N/m2 (Darcy formula,...(12.68))
D 2
In Eq. 12.68, r is density of fluid (kg/m3) and V is mean velocity of flow (m/s). The friction factor, fD is
determined depending on the Reynolds number, as explained in the chapter on Forced convection.
Effect of bends and valves in the flow lines is generally accounted for by adding an ‘equivalent length’ for
each bend or valve, to the straight length. Equivalent lengths of a few fittings are shown in Table 12.11:
(b) Pressure drops due to contractions and expansions DPe is a function of area ratio A 1/A 2, where A 1 is the
smaller area.

r ×V 2
DPe = f× N/m2 (where f is fcont or fexpn...(12.69))
2
where, V refers to velocity at smaller cross section.
Values of fcont and fexpn are given in Table 12.12; these are shown graphically in Fig. 12.27.
(c) Pressure drops due to flow acceleration This pressure drop, in a channel of constant crosssection, is equal to
twice the difference in velocity heads, i.e.
F r ×V 2
r 1 ×V12 I N/m
DPa = 2× GH 2
2 2
-
2 JK 2
...(12.70)

TABLE 12.11 Equivalent lengths of fittings


Fitting Le / D (for turbulent flow only)
45° Elbow 15
90° Elbow (standard radius) 31
90° Elbow (medium radius) 26
90° Elbow (long sweep) 20
90° Square elbow 65
180° Close return bend 75
Swing check valve, open 77
Tee (as eL, entering run) 65
Tee (as eL, entering branch) 90
Couplings, unions Negligible
Gate valve, open 7
Gate valve, 1/4 closed 40
Gate valve, 1/2 closed 190
Gate valve, 3/4 closed 840
Globe valve, open 340
Angle valve, open 170

HEAT EXCHANGERS 629


TABLE 12.12 Sudden contraction and expansion coefficients for a tube
A1 /A2 fcont fexpn
0 1 0.5
0.1 0.8 0.45
0.2 0.65 0.42
0.3 0.5 0.38
0.4 0.37 0.34
0.5 0.22 0.28
0.6 0.14 0.23
0.7 0.1 0.2
0.8 0.03 0.12
0.9 0.01 0.07
1 0 0

1 where, subscripts 1 and 2 refer to inlet and outlet,


respectively.
(d) Pressure drop due to self-draught If the height
0.8
of the vertical channel (flue) is ‘h’, ro = density of
cold fluid (say, air) and r = density of hot fluid (say,
fcont or, fexps

0.6 flue gas), then pressure drop due to self-draught is


fcont given by:
fexpn DP s = (ro – r)×g×h N/m2 ...(12.71)
0.4 where, ‘g’ is the acceleration due to gravity.
DPs is positive for the descending fluid and negative,
0.2 if the fluid is ascending through the channel. DPs is
zero if the heat exchanger is not exposed to ambient
air, but is connected in a closed system.
0 Then, the total pressure drop is given by the summa-
0 0.2 0.4 0.6 0.8 1
tion of all these pressure drops:
Area Ratio, A1/A2
DPt = DPf + DPe + DPa + DPs ...(12.72)
FIGURE 12.27 Sudden conraction and Power required to originate fluid flow:
expansion coefficients for a tube Once the total pressure drop in the system is deter-
mined, the power required to circulate the fluid
through the system is easily calculated:
Flow×D Pt M × D Pt
P= = ,W ...(12.73)
h r ×h
where, Flow = volumetric flow rate, m 3/s,
M = mass flow rate of fluid, kg/s
DPt = total pressure drop, N/m2, and
r = density of liquid or gas, kg/m3
h = efficiency of pump or fan.

12.10 Summary
Heat exchanger is one of the important pieces of process equipment, used extensively in research as well as
industrial applications. Heat exchangers may be of recuperative, regenerative or direct contact type.
In this chapter, we focussed on the thermal design aspects of heat exchangers. First, the method of calculat-
ing the overall heat transfer coefficient was explained. Inclusion of fouling resistance is an important aspect of
design and this was discussed next.
Calculation of logarithmic mean temperature difference, LMTD, between the two fluid streams exchanging
heat, is an important step in the design. Procedure of calculating the LMTD for parallel and counter-flow heat
exchangers was explained; for more complicated type of exchangers, such as cross flow or multi-pass shell-and-

630 FUNDAMENTALS OF HEAT AND MASS TRANSFER


tube heat exchangers, mean temperature difference is calculated by multiplying the LMTD of a counter-flow HX
by a correction factor. Correction factor graphs have been given for a few types of heat exchangers.
Problems in heat exchanger are mainly of two types: (i) design problems where one has to calculate the area
of the HX, and (ii) performance problems where one has to calculate the outlet temperatures of both the fluids,
given the inlet temperatures. LMTD approach is suitable for the first type of problems, whereas for the second
type of problems, e–NTU approach is recommended, since in this case LMTD approach would require a labori-
ous iterative procedure. e–NTU relations and graphs for a few important cases have been given. Further, operat-
ing-line/equilibrium-line method was also briefly explained.
Compact heat exchangers and regenerators are also used in a variety of applications. Brief mention has been
made about these; however, their design is rather more involved and use of proprietary technical information
from the suppliers’ catalogues will be required. Finally, calculation of pressure drops and the necessary pumping
power in a heat exchanger, has been explained.
Selection of heat exchangers for a particular application is a serious task for the engineer and the following
aspects must be borne in mind while selecting a heat exchanger:
(i) required heat transfer rate
(ii) necessary pumping power
(iii) type of heat exchanger most suitable, depending upon the process
(iv) materials of construction and fabrication and testing procedures, with due consideration to operating
temperatures and pressures
(v) size and weight, depending upon application
(vi) ease of maintenance and servicing
(vii) safety and reliability aspects, and
(viii) cost.

Questions
1. How are heat exchangers classified? Discuss briefly different types of heat exchangers. Why is counter-flow HX
better than parallel-flow HX? [M.U.]
2. Draw temperature vs. length profiles for: (i) Condenser (ii) Evaporator (iii) Counter-flow HX with Ch = Cc
[M.U.]
3. What is overall heat transfer coefficient? What is its importance? Derive an expression for overall heat transfer
coefficient for a tubular HX based on inner surface area. [M.U.]
4. Explain the terms: Fouling factor, Effectiveness, NTU and LMTD. [M.U.]
5. Write short notes on correction factor charts for cross-flow heat exchangers. [M.U.]
6. Starting from fundamentals, derive an expression for the mean temperature difference for counter-flow HX in
terms of inlet and outlet temperatures of hot and cold fluids. [M.U.]
7. Derive an expression for the LMTD of a parallel-flow HX. State clearly the assumptions. [M.U.]
8. Derive an expression for the effectiveness of a counter-flow HX when capacity rate of hot fluid is more than that
of cold fluid. Hence show that effectiveness of a condenser is given by:
e = 1 – exp(–NTU) ...[M.U.]
9. Starting from basics, derive an equation for the effectiveness of a parallel-flow HX in terms of NTU and capacity
ratio. Also, show that when capacity ratio is 1, effectiveness is given by:
e = (½).{1 – exp(– 2.NTU)} ...[M.U.]
10. Prove that for a counter-flow HX, when C min/Cmax = 1,
e = NTU/(1 + NTU). [M.U.]
11. Compare LMTD and e–NTU methods of solving heat exchanger problems.
12. Using the operating-line/equilibrium-line method, derive an expression for the effectiveness of a counter-flow
HX. Assume Cc > Ch.
13. Write a short note on compact heat exchangers and regenerators.

Problems
1. A copper pipe (k = 350 W/mK) of 17.5 mm ID and 20 mm OD conveys water and the oil flows through the
annular passage between this pipe and a steel pipe. On the water side, the film coefficient is 4600 W/(m 2K) and
fouling factor is 0.00034 m2 K/W. The corresponding values for the oil side are 1200 W/(m 2K) and 0.00086
m2 K/W. Calculate the overall heat transfer coefficient between the water and oil, based on outside surface area
of inner pipe.
HEAT EXCHANGERS 631
2. In a shell and tube counter-flow HX, water flows through a copper tube (20 mm ID, 23 mm OD), while oil flows
through the shell. Water enters at 20°C and comes out at 30°C while oil enters at 75°C and comes out at 60°C.
The water and oil side film coefficients. are: 4500 and 1250 W/(m2 K), respectively. Thermal conductivity of tube
wall is 355 W/(mK). Fouling factors on water and oil sides are: 0.0004 and 0.001 m2 K/W, respectively. If the
length of tube is 2.4 m, calculate the overall heat transfer coefficient and rate of heat transfer. [M.U.]
3. Saturated seam at 120°C is condensing on the outer surface of a single pass HX. The overall heat transfer coeffi-
cient is 1600 W/(m2 K). Determine the surface area of the HX required to heat 2000 kg/h of water from 20°C to
90°C. Also, determine the rate of condensation of steam in kg/h. Assume latent heat of steam to be 2195 kJ/kg.
[M.U.]
4. A HX is required to cool 55,000 kg/h of alcohol from 66°C to 40°C using 40000 kg/h of water entering at 5°C.
Calculate (i) the exit temperature of water (ii) heat transfer (iii) surface area required for: (a) parallel-flow type
(b) counter-flow type of HX.
Take overall heat transfer coefficient U = 580 W/(m2 K). Cp (alcohol) = 3760 J/(kgK) and Cp (water) = 4180
J/(kgK).
5. In a counter-flow double pipe HX, water flow rate is 1300 kg/h. and it enters at 15°C. It is heated by oil, Cp = 2
kJ/kgK; oil flow rate is 550 kg/h. Oil inlet temperature is 95°C. Overall U = 800 W/m2K. Surface area of HX:
1.34 m2. Table of NTU–e is given as follows:

Capacity ratio, R NTU Effectiveness


0.202 3 0.93
0.202 4 0.96

Find out e, NTU and outlet temperatures. [M.U.]


6. In a gas turbine installation, a counter-flow HX, has hot exhaust gas outlet at 330°C, and air outlet at 460°C. For
each element of HX, (dQ/(dT.dX) is uniform and is equal to C and C.L = 52.3 kW/K. Capacity rate for hot fluid
= 21.76 kW/K and for cold fluid = 19.04 kW/K. Temperature variation along the length is linear for both fluids.
Calculate temperatures at entry. [M.U.]
7. A one shell, 2-tube pass steam condenser, has 2000 tubes of 20 mm diameter, with cooling water entry at 20°C,
flow rate 3000 kg/s; U = 6890 W/m2 K. Total heat to be transferred, Q = 2.331 ´ 10 8 W. Steam condenses at 50°C.
Determine tube length per pass using NTU method. Given that at 0.6 and 0.64 effectiveness, NTU is 0.78 and
0.82. [M.U.]
8. A water pre-heater of ID:3.2 cm, OD:3.52 cm, is heated by steam at 180°C. Water flows through pipe at a
velocity of 1.2 m/s. ‘h’ on steam side:11,000 W/m2K; water is heated from 25°C to 95°C. k of pipe material: 59
W/mK. Properties of water at 60°C are given. Calculate the length required. Use appropriate empirical relation.
Data: m = 4.62 ´ 10– 4 kg/ms; k = 0.653 W/mK; Cp = 4200 J/kgK. [M.U.]
9. Consider a HX for cooling oil entering at 180°C, by water entering at 25°C; mass flow rates of oil and water are:
2.5 and 1.2 kg/s, respectively. Area: 16 m2. Specific heat data for oil and water and overall U are given:
Data: Cp of oil = 1900 J/kgK; Cp of water = 4184 J/kgK; U = 285 W/m2K.
Calculate outlet temperature of oil and water for parallel and counter-flow HX. [M.U.]
10. In a shell-and-tube HX, 50 kg/min of furnace oil is heated from 10°C to 90°C. Steam at 120°C flows through the
Shell and oil flows inside the tube. Tube size: 1.65 cm ID and 1.9 cm OD. Film coefficients on oil and steam sides
are: 85 and 7420 W/(m2K). Find the number of passes and number of tubes in each pass if the length of each
tube is limited to 2.85 m. Velocity of oil is limited to 5 cm/s. Density and specific heat of oil are 900 kg/m3 and
1970 J/(kgK), respectively. [M.U.]
11. Water at a rate of 4080 kg/h is heated from 35°C to 75°C by an oil of Cp = 1.9 kJ/(kgK). The HX is of counter-
flow, double pipe design. The oil enters at 110°C and leaves at 75°C. Determine: (i) mass flow rate of oil (ii) area
of HX necessary to handle this load, if overall heat transfer coefficient, U = 320 W/(m2 K). [M.U.]
12. A steam condenser, condensing at 70°C has to have a capacity of 100 kW. Water at 20°C is used and the outlet
water temperature is limited to 45°C. If the overall heat transfer coefficient is 3100 W/(m2K), determine the area
required. If the inlet water temperature is increased to 30°C, determine the increased flow rate of water to
maintain the same outlet temperature. [M.U.]
13. Water enters a counter-flow, double pipe HX at 38°C flowing at a rate of 0.76 kg/s. It is heated by oil (Cp = 1.88
kJ/kgK) flowing at a rate of 0.152 kg/s from an inlet temperature of 116°C. For an area of 1.3 m2 and an overall
heat transfer coefficient of 340 W/(m2K), determine the total heat transfer rate. Take Cp for water = 4.170 kJ/
kgK.
Given: expression for effectiveness of double pipe, counter-current HX:

632 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1 - exp(- NTU (1 - C ))
e= [M.U.]
1 - C × exp(- NTU (1 - C ))
14. A refrigerator is designed to cool 250 kg/h of hot liquid (Cp = 3350 J/kgK) at 120°C using a parallel-flow ar-
rangement. 1000 kg/h of cooling water is available at a temperature of 10°C. If overall U = 1160 W/(m2 K) and
the surface area of HX is 0.25 m 2, calculate the outlet temperatures of the fluids and also the effectiveness of the
HX. [M.U.]
15. A steam condenser, condensing at 100°C has a capacity of 150 kW. Water at 25°C is used and the outlet water
temperature is 35°C. If the overall heat transfer coefficient is 3000 W/(m 2K), determine the area required for the
HX. [M.U.]
16. A parallel-flow HX has hot and cold water streams running through it and has following data:
mh = 10 kg/min, mc = 25 kg/min, Cph = Cpc = 4.18 kJ/(kgK), Th1 = 70°C, Th2 = 50°C, Tc1 = 25°C. Individual heat
transfer coefficients on both sides are 60 W/(m2K), each. Calculate: (i) area of HX (ii) exit temperatures of hot
and cold fluids, if hot water flow rate is doubled. [M.U.]
17. Steam at atmospheric pressure enters the shell of a surface condenser in which water flows through a bundle of
tubes of diameter = 25 mm at a rate of 0.05 kg/s. The inlet and outlet temperatures of water are 15°C and 70°C,
respectively. Condensation of steam takes place on the outside surface of the tube. If U = 230 W/(m2K), using
NTU method, find: (i) effectiveness of the HX (ii) length of tube required, and (iii) rate of steam condensation.
[M.U.]
18. An economiser in a boiler has water flowing inside the pipes and hot gases on the outside, flowing across the
pipes. The flow rate of the gases is 2000 tonne/h and they are cooled from 390°C to 200°C and their specific heat
is 1005 J/(kgC). Water is heated under high pressure from 100°C to 220°C. Assuming an overall heat transfer
coefficient of 35 W/(m2C), determine the area required. The correction factor, F = 0.8. [M.U.]
19. Hot oil is being cooled from 200°C to 130°C in a parallel-flow HX by water entering a 25°C and exiting at 60°C.
Determine the outlet temperatures of both the streams if the HX is made counter-flow. [M.U.]
20. A shell and tube HX with two shell passes and 8 tube passes has ethyl alcohol (Cp = 2670 J/kgC) flowing inside
the tubes, and water (Cp = 4190 J/kgC) flows through the shell. Ethyl alcohol enters at 25°C and leaves at 75°C
with a flow rate of 2 kg/s whereas water enters at 95°C and leaves at 45°C. Overall heat transfer coefficient U =
850 W/(m2 K). Determine the surface area required for the HX.

Appendix
In this Appendix to Chapter 12, some more information on compact heat exchangers and regenerators is given.
Example A12.1. Air at 2 atm and 400 K flows at a rate of 5 kg/s,
across a finned circular tube matrix shown in Fig. 12.19. Dimen-
sions of the heat exchanger matrix are: 1 m (W) ´ 0.6 m (Deep) ´ 0.5 m
0.5 m (H), as shown in Fig. A12.1. Find: (a) the heat transfer coef-
ficient (b) the friction factor, and (c) ratio of core friction pressure Air, 2 atm, 400 K,
drop to the inlet pressure. 5 kg/s
Solution. 1.0 m
Data:
m = 5 kg/s ...mass flow rate 0.6 m
Afr = 0.5 m2 ...frontal area
FIGURE A12.1 Configuration of compact
L = 0.6 m ...length of flow
heat exchanger for Example A12.1
Physical properties of air at 2 atm and 400 K:
r := 0.883 ×2 kg/m3 i.e. r := 1.766 kg/m3 m := 2.29 ´ 10 –5 kg/ms Cp := 1013 J/kgK Pr := 0.703
From Fig. 12.19, we have:
Amin
s := 0.534 where, s =
A fr

0.01192
and, Dh := m
3. 28
i.e. Dh := 3.634 ´ 10 – 3 m (hydraulic diameter)
Then,
Mass velocity:
m
G= and, Amin = s×Afr
Amin

HEAT EXCHANGERS 633


m
i.e. G :=
s × A fr
i.e. G = 18.727 kg/sm2 (mass velocity)
Reynolds number:
G × Dh
Re :=
m
i.e. Re = 2.972 ´ 103 (Reynolds number)
Then, from Fig. 12.19, for Re = 2972, we get:
2
h
× Pr 3 = 0.0069
G × Cp
(a) And, heat transfer coefficient:
0 .0069 × G × Cp
h := 2
Pr 3
i.e. h = 165.557 W/(m2C) (heat transfer coefficient.)
(b) Friction factor:
From Fig. 12.19, for Re = 2972, we get:
f = 0.024 (friction factor)
(c) Pressure drop:
G2 A
DPf = f× × N/m2 (fricitional pressure drop)
2× r Amin

A 4×L
Now, =
Amin Dh
A
i.e. = 660.403
Amin
Therefore,
G2
DPf := f× ×660.403 N/m2 (core friction pressure drop)
2× r
i.e. DPf = 1.574 ´ 103 N/m2 (core friction pressure drop.)
D Pf 1574
And, = = 0.78%
P 2 × (1.0132 ´ 105 )
i.e. frictional pressure drop is 0.78% of the inlet pressure.
Fig. A12.2, A12.3, and A12.4 show data for three more compact heat exchanger matrices, from Kays and
London.
Fig. A12.5 shows data for crossed rod matrices, random stacking (d = 0.375 in.) used in regenerators (Kays
and London):
Fig. A12.6 gives data for an infinite, randomly stacked sphere matrix, with porosity varying from 0.37 to
0.39. (Kays and London)
Data of Fig. A12.6 is given in tabular form in Table A12.1.
Analysis of regenerators by NTU method:
As explained earlier, in a regenerator, the same space is alternately occupied by the hot and cold fluids; regenera-
tor matrix stores the ‘heat’ during the flow of hot fluid (i.e. during ‘hot blow’) and rejects this heat to the cold
fluid during the flow of cold fluid through the regenerator matrix (i.e. during ‘cold blow’). Temperatures of the
gas as well as of the matrix solid are functions of both position and time. After sufficiently long time, some sort
of steady state is reached, and the same temperature distribution is repeated in each cycle of operation.
An NTU analysis, similar to the one done for heat exchangers, proceeds as follows, with the assumption that
the same mass flow rate of gas is maintained during both the hot and cold blows. (Ref: Cryogenic Systems by
R.F. Barron).

634 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Tube outside diameter = 0.645 in.
0.070 Fin pitch = 8.7 per in.
B Fin thickness = 0.010 in.
0.060 Fin area/total area = 0.862
0.050 Flow passage hydraulic diameter, 4rh = 0.01797 0.03883 ft.
Free-flow area/frontal area, s = 0.443 0.628
A Heat transfer area/total volume, a = 98.7 2 3
0.040 65.7 ft /ft
Note: Minimum free-flow area is in spaces transverse to flow.
0.030

S 1.121"

0.020

0.645"
B
A 0.1149" 0.010

0.010 To scale for "A"


2/ 3
(h/Gcp)×Npr

0.008 Spacing S
A 1.232"
–3
0.006 NR ´ 10 B 1.848"

1.0 2.0 3.0 4.0 6.0 8.0 10.0

FIGURE A12.2 Heat transfer coefficient and friction factor data for finned circular tube matrix
(surface CF 8.7–5/8J)

Tube outside diameter = 0.645 in.


0.060 Fin pitch = 7.0 per in.
Flow passage hydraulic diameter, 4rh = 0.0219 ft
Fin thickness = 0.010 in.
Free-flow area/frontal area, s = 0.449
0.040 2 3
Heat transfer area/total volume, a = 82 ft /ft

0.030 Note: Minimum free-flow area is in spaces transverse to flow.

0.020
S 1.121"
f

0.645"
0.010 135 0.1149" 0.010
2/ 3

0.008
(h/Gcp)◊Npr

0.006
–3
NR ¥ 10

1.0 2.0 3.0 4.0 6.0 8.0 10.0 20 30

FIGURE A12.3 Heat transfer coefficient and friction factor data for finned circular tube matrix
(surface CF–7.0–5/8J)

Following notations are used:


ms = mass of solid (matrix filling) in regenerator, kg
m = mass flow rate of gas through regenerator, kg/s
cs = specific heat of solid, J/(kgK)
cp = specific heat of gas, J/(kgK)
L = length of regenerator, m
Ah = total heat transfer area of solid material in regenerator, m2
r = density of gas, kg/m3
T = temperature of gas at location x and time t
Ts = temperature of solid material at location x and time t

HEAT EXCHANGERS 635


Fin pitch = 9.1 per in.
0.100"
0.060 Flow passage hydraulic diameter, 4rh = 0.01380 ft
0.55" Fin metal thickness = 0.004 in., copper
0.050
Free-flow area/frontal area, s = 0.788

0.18"
0.79" 2 3
0.040 0.737" 1.110" Total heat transfer area/total volume, a = 224 ft /ft
Fin area/total area = 0.813
0.030

0.020

0.015
Best Interpretation
0.010
f

2/ 3
(h/Gcp)NPr

0.008

0.006
0.005 –3
NR ¥ 10
0.004

0.4 0.5 0.6 0.8 1.0 1.5 2.0 3.0 4.0 6.0 8.0 10.0

FIGURE A12.4 Heat transfer coefficient and friction factor data for finned flat tube matrix
(surface 9.–0.737-5)

1.00
0.80 Porosity p Transverse pitch Xt
I ¥ 0.832 4.675
0.60
II 0.817 4.292
III 0.766 3.356
0.40
IV 0.725 2.856
V 0.675 2.417
VI 0.602 1.974
f

0.20 VII 0.500 1.571

0.10
0.08
0.06

0.04
2/ 3
(h/Gcp)Npr

0.02

0.01
0.008
0.006 NR = (4rh G/m)
0.004
2 3 4 5
10 2 4 6 8 10 2 4 6 8 10 2 4 6 8 10

FIGURE A12.5 Crossed rod matrix, random stacking (rod diameter d = 0.375 in.)

hc = heat transfer coefficient between the solid material and gas, W/(m2/C).
Now, applying the First law to a differential element of gas:
heat transferred to or from the gas = change in enthalpy of gas in a length dx, as it flows through the regenerator
i.e.
hc ( T – Ts) (Ah/L) dx = – mc p (¶T/¶x) dx ...(1)
Here, change in energy stored in the gas within the differential element is neglected. (This is true for cryo-
genic regenerators.)
Again, applying the First law to a differential element of solid material, we have:
heat transferred to or from the solid material = change in energy stored within the solid material

636 FUNDAMENTALS OF HEAT AND MASS TRANSFER


8.0
6.0
5.0
4.0
3.0

2.0
f
0.10 1.0
0.08 0.8
0.06 0.6
0.05 0.5
0.04 0.4
0.03 0.3
2/ 3
Nst Npr

2/ 3 – 0. 3
0.02 Nst Npr = 0.23NR 0.2

0.01
0.008
0.006
2 3 4
2 3 4 5 6 8 10 2 3 4 5 6 8 10 2 3 4 5 6 8 10 2 3 4 5

FIGURE A12.6 Data for an infinite, randomly stacked sphere matrix with porosity varying from 0.37 to 0.39

TABLE A12.1 Heat transfer and friction data for sphere bed matrices (Kays and London)
(Random packing, p = 0.37 to 0.39)
Reynolds number, NR NSt.NPr2/3 f
50000 0.0089 0.30
20000 0.0118 0.34
10000 0.0144 0.37
5000 0.0178 0.41
2000 0.023 0.47
10000 0.029 0.52
500 0.0355 0.59
200 0.046 0.80
100 0.056 1.10
50 0.069 1.65
20 0.091 3.0
10 0.112 5.2

i.e. Gas temperature = T


hc ( T – Ts) (Ah/L) dx = ms (dx/L) cs (¶Ts/¶t) ...(2) Solid temperature = Ts Regenerator filling
Solving Eqs. 1 and 2, we get the partial differential mc , in mh , in
equation for the temperature of gas flowing through x
dx
the regenerator:
[¶2T/(¶x.¶t)] + [hc Ah/(m cp L)] (¶T/¶t)
+ [hc Ah/(ms cs)] (¶T/¶x) = 0 ...(3)
From Eq. (3), we observe that two important mc = mh = m
dimensionless quantities are involved in the analysis
of a regenerator, i.e.
Ntu = hc Ah/(m cp) = Number of heat transfer units,
and 0 L
mh, out mc, out
Fn = hc Ah/(ms cs f) = Frequency number, where f =
1/P = frequency of switching the hot and FIGURE A12.7 NTU analysis of a regenerator
cold stream, P = heating or cooling period.
f.t = dimensionless time.

HEAT EXCHANGERS 637


1.0 Fig. A12.8 Gas temperature distribution for a counter-flow
Counter-flow regenerator regenerator as a function of time. For the regenerator,
0.9
Ntu = 10 hc × Ah hc × Ah
1 Nlu = = 10 and Fn = = 1/p. During the heating
0.8 Fn = = 0.318 m ×c p ms × c s × f
p
ThL – Tco period (when the cold gas is flowing through the regenerator, use
0.7
ThL – T

(T – Tco)/(ThL – Tco) and x /L as coordinates; for the cooling period


0.6 (when the warm gas is flowing through the regenerator), use
(ThL – T )/(ThL – Tco) and (L – x)/L as coordinates. ThL is entering

0. 0.5 .25 0
0 =
0.5 temp. of warm stream (at x = L) and Tco is the entering temp. of
or

fr
cold gas stream (at x = 0).

1. 75 0
ThL – Tco

0.4
T – Tco

00
0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1.0
x L-x
or =
L L
FIGURE A12.8 Gas temperature distribution in a counter-flow regenerator
Eq. 3 has been solved numerically for temperature distribution of gas, by Hausen for steady state cyclic
operation of a regenerator. Results for a particular case of Ntu = 10 are shown in Fig. A 12.8, as an example.
Once the temperature distribution is known, actual energy transferred is calculated as:

Q actual =
zz
0
1
f

0
L
hc ×
Ah
L
× (T - Ts ) dx dt

Maximum possible heat transfer in the regenerator occurs when the gas is heated from Tco (i.e. temperature
...(4)

of cold gas entering at x = 0) to the temperature ThL (i.e. temperature of hot gas entering the regenerator at x = L).
We get:
m × c p × (ThL - Tco )
Qideal = ...(5)
f
Then, regenerator effectiveness e , is given by:
Qactual
e=
Qideal

i.e. e=
hc × Ah
m × cp
×
z z FGH
0
1 1

0
T - Ts I FG IJ
ThL - Tco
Hausen’s numerical solution for the effectiveness of a regenerator as a function of frequency number and
x
JK H K
d
L
d(f×t) ...(6)

number of transfer units, is shown in Fig. A 12.9.


It is clear from Fig. A12.9 that for a large effectiveness, we need a small frequency number (Fn) and a large
number of transfer units (Ntu), i.e. for a large effectiveness of a regenerator, the requirements are:
(i) large heat transfer coefficient, hc
(ii) small gas mass flow rate, m, or small gas capacity rate m.cp
(iii) large product of regenerator matrix mass and its specific heat, ms.cs
(iv) large frequency, f.
Example A12.2. Following data are given for a regenerator:
heat transfer coefficient = 640 W/(m2 K), heat transfer area per unit length = 100 m2/m, mass of matrix solid per unit
length = 8 kg/m, specific heat of matrix material = 800 J/(kgK), frequency of operation = 60 cpm (= 1 cycle/s), mass flow
rate of gas through regenerator = 0.013 kg/s and specific heat of gas = 5200 J/(kgK). Desired effectiveness of this regen-
erator, operating in a counter-flow mode, is 0.95. Determine the length of the regenerator required.
Solution.
Data:
Ah ms
hc := 640 W/(m2K) = 100 m2/m = 8 kg/m cs := 800 J/(kgK) f := 1 cycle/s m := 0.02 kg/s
L L
638 FUNDAMENTALS OF HEAT AND MASS TRANSFER
1.0
Counter-flow regenerator Fh = 0
0.9 10
20
0.8 30
0.7
40

mcp(ThL – Tco)
0.6 hc ◊ Ah
50 Fn =

Q
ms ◊ c s ◊ f
0.5

0.4

e= 0.3

0.2

0.1

0
0 10 20 30 40 50
h ◊A
Ntu = c h
m ◊c p

FIGURE A12.9 Effectiveness of a counter-flow regenerator


cp := 5200 J/(kgK) e := 0.95
Frequency number:

FG A IJ
H LK
h
hc ×
We have: Fn =
FG m IJ × c × f
H LK
s
s

640 ×100
i.e. Fn =
8 × 800 × 1
i.e. Fn =10 (frequency number)
Number of heat transfer units:
For a Fn = 10 and e = 0.95, get the value of Ntu from Fig. A 12.9:
We get: Ntu =45
hc × Ah
But, Ntu =
m × cp
Therefore, heat transfer area required:
N tu × m × c p
Ah :=
hc
i.e. Ah = 7.313 m2 (heat transfer area required)
Therefore, regenerator length required:
Ah
L=
Ah
L
7 . 313
i.e. L=
100
i.e. L = 0.073 m = 7.3 cm (length of regenerator required.)
Regenerator ineffectiveness (1 – e) vs. NTU graphs for cryogenic regenerators:
As stated in the text, cryogenic regenerators, generally, have large values of modified NTU (i.e. NTUmod), of the
order of 100 or more. It may be observed that in the usual e–NTU graphs, the value of e approaches unity
asymptotically; so, for cryogenic heat exchangers, it is more instructive and convenient to draw regenerator inef-

HEAT EXCHANGERS 639


20

10

Ineffectiveness, (1 – e), %
6
5
Cr/Cmin
4 •
5
3
3 2
1.5
1.0
2

1
10 20 30 40 50 60 80 100
NTUmod

FIGURE A12.10 Regenerator ineffectiveness as a function of Ntu0 and matrix capacity rate ratio (Cmin/Cmax = 1)

20

10

8
Ineffectiveness (I – e), %

Cc
6 Cmin
1.0
5
1.5
4
2
3
3

5
2

1
10 20 30 40 50 60 80 100
NTUmod

FIGURE A12.11 Regenerator ineffectiveness as a function of Ntu0 and matrix capacity rate ratio
(Cmin/Cmax = 0.95)

fectiveness (1 – e) against NTUmod in log–log coordinates. Two sample graphs, one for Cmin/Cmax = 1, and the
other for Cmin/Cmax = 0.95, are shown in Fig. A12.10 and Fig. A12.11, respectively. (Ref: Compact Heat Exchang-
ers by Kays and London).

640 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

13
Radiation

13.1 Introduction
So far, we considered heat transfer by conduction and convection. In these modes of heat transfer, there was
always a medium present for heat transfer to occur. However, radiation mode of heat transfer is radically differ-
ent in the sense that there is no need for a medium to be present for heat transfer to occur. Just as conduction and
convection heat transfers occur when there is a temperature gradient, net radiation heat transfer also occurs from
a higher temperature level to a lower temperature level. There are two theories concerning the radiation heat
transfer: one, classical electromagnetic wave theory of Maxwell, according to which energy is transferred during
radiation by electromagnetic waves, which travel as rays and follow the laws of optics; second, the ‘Quantum
theory’ of physics, according to which energy is radiated in the form of successive, discrete ‘quanta’ of energy,
called ‘photons’. Both the theories are useful to explain the radiation phenomenon and properties.
Radiation heat transfer is proportional to the fourth power of absolute temperature of the radiating surface.
Therefore, radiation becomes the predominant mode of heat transfer when the temperature of the body is high.
With this in mind, we can cite a few important applications of radiation heat transfer:
(i) industrial heating, such as in furnaces
(ii) industrial air-conditioning, where the effect of solar radiation has to be considered in calculating the heat
loads
(iii) jet engine or gas turbine combustors
(iv) industrial drying
(v) energy conversion with fossil fuel combustion, etc.
Following are some of the features of radiation:
(a) The electromagnetic magnetic waves are of all wavelengths, travelling at the velocity of light, i.e. c = 3 ´
10 10 cm/s
(b) Frequency (f) and wavelength (l) are connected by the relation: c = l.f, which means that higher the
frequency, lower the wavelength
(c) Smaller the wavelength, more powerful is the radiation, and also more damaging, e.g. X-rays and
Gamma rays.
A sketch of the electromagnetic spectrum is shown in Fig. 13.1. Different parts of the electromagnetic spec-
trum have wavelengths (l) as shown in Table 13.1.
In this chapter, we are interested in radiations, which on absorption, result in production of heat, i.e. ‘ther-
mal radiation’. It may be observed that thermal radiation falls in the wavelength range of 0.1 to 100 microns (Unit
of wavelength is 1 micron = 10–6 m, and 1 Angstrom = 10– 10 m), i.e. thermal radiation includes entire visible (i.e.
l = 0.4 to 0.8 microns) and infra-red and part of ultra-violet range. As a matter of interest, it may be stated that
most of the radiation from the sun (temperature: 5600°C approximately) is in the lower end of 0.1 to 0.4 microns
and, for comparison, radiation from an incandescent lamp is in the range of 1 to 10 microns.
While most of the solids and liquids emit radiation in a continuous spectrum, gases and vapours radiate
only in certain wavelength bands; therefore, they are known as ‘selective emitters’.
TABLE 13.1 Wavelengths of different types of radiation

Kind of radiation Wavelength, l


Cosmic rays up to 4 ´ 10 –7 mm
Gamma rays 4 ´ 10 –7 to 1.4 ´ 10 –4 mm
X-rays 1.0 ´ 10 –5 to 2 ´ 10 –2 mm
Ultra-violet 0.01 to 0.39 mm
Visible radiation 0.39 to 0.78 mm
Thermal radiation 0.1 to 100 mm
Infra-red 0.78 to 1000 mm
Microwave 0.8 to 1000 mm
Radio waves Beyond 1 m

Thermal

12 10 8 6 4 2 2 4
10 10 10 10 10 10 1 10 10

l, (m)

Cosmic X-rays Ultra- Visible Infrared Microwave Radio waves


rays violet

FIGURE 13.1 Electromagnetic spectrum

13.2 Properties and Definitions


Often, we use the term ‘spectral’; it means, dependence on wavelength.
And, value of a quantity at a given wave length is called ‘monochromatic value’.
Absorptivity, Reflectivity and Transmissivity:
In general, when radiant energy (Qo) is incident on a surface, part of it may be absorbed (Qa), part may be
reflected (Qr) and part may be transmitted (Qt) through the body. Then, obviously,
Qa + Qr + Qt = Qo
Qa Qr Qt
i.e. + + =1
Qo Qo Qo
i.e. a+r+t=1 ...(13.1)
where, a = absorptivity = fraction of incident radiation absorbed
r = reflectivity = fraction of incident radiation reflected
t = fraction of incident radiation transmitted.
Most of the solids and liquids are ‘opaque’, i.e. they do not transmit radiation, and t = 0; so, for most solids
and liquids: a + r = 1.
Gases reflect very little; so, for gases: a + t = 1.
If t = 1, entire radiation passes through the body; such a body is ‘transparent’ or ‘diathermaneous’.
If a = 1, the body absorbs all the incident radiation and such a body is called a ‘black body’.
If r = 1, all the incident radiation is reflected, and it is a perfectly ‘white body’.
In reality, there are no ‘perfectly’ black, white or transparent bodies.
However, some bodies are transparent to only waves of certain wavelength; for example, rock salt is trans-
parent to heat rays, but non-transparent to ultra-violet rays. And, window glass is transparent to visible light, but
almost non-transparent to ultra-violet and infra-red rays. Therefore, a space covered with glass (or plastic) enclo-

642 FUNDAMENTALS OF HEAT AND MASS TRANSFER


sure, allows solar radiation to pass through it and the objects inside the enclosure get heated up; the heated
objects radiate, but this radiation is in the higher wavelength range (infra-red) to which glass or plastic is opaque.
So, the heat gets ‘trapped’ inside the enclosure and the temperature inside the enclosure rises above that of
ambient. This is known as ‘Greenhouse effect’ and is used to keep the plants warm in cold weather. Another
example of Greenhouse effect is manifested in heating up the interior of a car to a temperature much above the
ambient temperature when the car is parked in hot sun, with all its windows closed.
Absorption and reflection of heat rays depend rather on the state of the surface than on the colour of the
surface. For example, snow has an absorptivity of 0.985 and is nearly ‘black’ for thermal radiation!
Absorptivity of a surface can be increased by applying coatings of dark paints; usually, lamp black is used
for this purpose.

q
Monochromatic
radiation emitted

Directional
distribution

FIGURE 13.2 Spectral and spatial energy distribution

Spectral and Spatial energy distribution:


Distribution of radiant energy is non-uniform with respect to both wavelength and direction, as shown in
Fig. 13.2.
Perfect black body A perfect black body does not exist in nature; however, a perfect black body can be approxi-
mated in the laboratory by having a sphere coated black on the inside; then, if there is a small hole on the wall of
the sphere, the radiation Q entering the hole goes through multiple reflections and after ‘n’ reflections, r n.Q is the
emergent energy flux. Obviously, the emerging flux tends to be zero when ‘n’ tends to infinity, i.e. the pin hole in
the sphere simulates a black body. This is known as ‘Hohlraum’. See Fig. 13.3.
Note the following points in connection with a black body:
(i) A black body absorbs all the incident radiation, of all wavelengths and from all directions
(ii) For a given temperature and wavelength, energy emitted by a black body is the maximum as compared
to any other body
(iii) Black body is a ‘diffuse emitter’, i.e. the radiation emitted by a black body is independent of direction
(iv) A black body does not reflect or transmit any of incident radiation
Reflection Reflection may be ‘specular’ (or mirror-like) or ‘diffuse’. See Fig. 13.4
In specular reflection, the angle of incidence is equal to the angle made by the reflected ray with the normal
to the surface. In case of diffuse reflection, magnitude of reflected energy in a given direction is proportional to
the cosine of the angle of that direction to the normal. ‘Roughness’ of the surface determines if the reflection is
specular or diffuse: if the ‘height’ of corrugations on the surface is much smaller than the wavelength of incident
radiation, the surface behaviour is specular; otherwise, it is diffuse.
Emissive power (E ) The ‘total (or hemispherical) emissive power’ is the total thermal energy radiated by a
surface per unit time and per unit area, over all the wavelengths and in all directions. Note, in particular, that
only the original, emitted energy is to be considered and the reflected energy is not to be included. Total emissive
power depends on the temperature, material and the surface condition.
Solid angle ‘Solid angle’ is defined as a region of a sphere, which is enclosed by a conical surface with the
vertex of the cone at the centre of the sphere. See Fig. 13.5.
If there is a source of radiation of a small area at the centre of the sphere O, then the radiation passes through
the area An on the surface of the sphere and we say that the area An subtends a solid angle w when viewed from

RADIATION 643
Sphere coated black on the centre of the sphere. Note that with this definition, An is always
inside surface normal to the radius of the sphere. Mathematically, solid angle is
expressed as:
An
w= (steradians (sr)...(13.2))
r2
However, in a practical case, the surface may not be part of a
Q sphere; but, if a plane area A intercepts the line of propagation of
radiation such that the normal to the surface makes an angle q with
the line of propagation, then we project the incident area normal to
the line of propagation, such that, the solid angle is now defined as:
A ×cos(q )
w= sr ...(13.3)
FIGURE 13.3 Simulation of a black r2
body in laboratory—‘Hohlraum’ Note that, A.cos(q) = An is the projected area of the incident
surface, normal to the line of propagation.
N N

Incident Reflected Incident

q q

(a) Specular (b) Diffuse

FIGURE 13.4 Specular and diffuse reflection

w An Intensity of radiation (Ib) Intensity of radiation for a black


body, Ib is defined as the energy radiated per unit time per
unit solid angle per unit area of the emitting surface pro-
2 jected normal to the line of view of the receiver from the
O w = An/r
radiating surface.
Mathematically, this is expressed as:
dQb
r Ib = W/(m2sr) ...(13.4)
( dA × cos(q )) × dw
FIGURE 13.5 Definition of solid angle
dEb
i.e. Ib = W/(m2sr) ...(13.5)
cos(q ) × dw
Note that Emissive power Eb of a black body refers to unit surface area whereas Intensity Ib of a black surface
refers to unit projected area.
Ibl is the intensity of black body radiation for radiation of a given wavelength l. And, Ib is the summation

z
over all the wavelengths, i.e.
¥
Ib = Ibl dl W/(m2sr) ...(13.6)
0
Consider a small, black surface dA emitting radiation all over a hemisphere above it. See Fig.13.6. Let a
radiation collector be located on the hemispherical surface at a zenith angle q to the normal to the surface and
azimuth angle j; further, let the collector subtend a solid angle dw when viewed from a point on the emitter.
Then, it will be observed that maximum amount of radiation is measured when the collector is vertically above
the emitter, normal to the emitter. In any direction q from the normal, rate of energy radiated is given by Lam-
bert’s cosine law: “A diffuse surface radiates energy such that the rate of energy radiated in a direction q from
the normal to the surface is proportional to the cosine of the angle q”, i.e.
Qq = Qn × cos(q)

644 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Emitted radiation
An
Qn
Radiation collector

Qq = Qn cos (q)
q
dw q
Black
emitter

dA

FIGURE 13.6 Lambert’s cosine law

In general, for real surfaces, intensity does not vary with f, but depends on q; however, with the intensity of
radiation as defined above, i.e. on basis of unit projected area, it can be shown that for a black surface, the
intensity is the same in all directions. Such a surface is known as ‘diffuse surface’.
For a diffuse, black surface, radiation intensity is independent of direction and such surfaces are also known
as ‘Lambertonian surfaces’.
Intensity can be thought of as brightness; looking down vertically along the normal, a viewer sees all of the
black surface dA at a particular level of brightness; and looking down along a line that makes an angle q with the
normal, the viewer will see only the projected area dA.cos(q), but at the same level of brightness.
Many real bodies, which are not diffuse, do not obey Lambert’s law and their radiation intensity changes
with the direction q; for example, for polished metals, the ‘brightness’ is a maximum not in the direction normal
to the surface, but at 60 to 80 deg. from the normal, and with further increase in q, the brightness drops abruptly
to zero. But for materials like corundum and copper oxide, the intensity (or brightness) is greater along the
normal than that in other directions.

13.3 Laws of Black Body Radiation


13.3.1 Planck’s Law for Spectral Distribution
Radiation energy emitted by a black surface depends on the wavelength, temperature of the surface and the
surface characteristics.
Planck’s distribution law relates to the spectral black body emissive power, Ebl defined as ‘the amount of
radiation energy emitted by a black body at an absolute temperature T per unit time, per unit surface area, per
unit wavelength about the wavelength l’. Units of Ebl are: W/(m2mm). The first subscript ‘b’ indicates black body
and the second subscript ‘l’ stands for given wavelength, or monochromatic. Planck derived his equation for Ebl
in 1901 in conjunction with his ‘quantum theory’.
Planck’s distribution law is expressed as:

C1 × l- 5
F C I
2
Ebl (l) = W/(m mm) ...(13.7)
exp G
H l ×T JK - 1
2

where, C 1 = 3.742 ´ 10 8 Wmm4/m2


and, C2 = 1.4387 ´ 104 mmK.
Plots of Ebl vs. l for a few different temperatures are shown in Fig. 13.7.
To plot the Planck’s distribution for a black body, using Mathcad, first, define Ebl as a function of T and l as
follows:

RADIATION 645
Spectral emissive power of a black body
9
1 ¥ 10 Locus of max. power,
1 ¥ 108 l×T = 2898 mm×K

Spectral emissive power, W/(m2micron)


T = 5800 K = Temp. of sun
1 ¥ 107
1 ¥ 106
T = 1000 K
1 ¥ 105
1 ¥ 104 T = 500 K
1 ¥ 103
100
10 T = 100 K
1
0.1
0.01
3
1 ¥ 10
4
1 ¥ 10
3
0.01 0.1 1 10 100 1 ¥ 10
Wavelength, microns
FIGURE 13.7 Planck’s distribution law for a black body

C1 × l- 5
F C I - 1 W/(m mm)
2
Ebl (l, T) := ...(13.7)
exp G
H l ×T JK
2

Then, define a range variable l varying from 0.01 mm o 1000 mm:


l := 0.01, 0.02, ... , 1000.
Now, select the x–y plot from the graph palette; on x-axis place holder, type l and on the y-axis place holder,
fill in Ebl (l,100), Ebl (l,500), Ebl (l,1000), and Ebl (l,5800). Click anywhere outside the graph region, and the
curves appear immediately.
This is an important graph that tells us quite a lot about the characteristics of black body radiation:
(i) At a given absolute temperature T, a black body emits radiation over all wavelengths, ranging from 0 to
¥.
(ii) Spectral emissive power curve varies continuously with wavelength.
(iii) At a given wavelength, as temperature increases, emissive power also increases.
(iv) At a given temperature, emissive power curve goes through a peak, and a major portion of the energy
radiated is concentrated around this peak wavelength l max.
(v) A significant part of the energy radiated by sun (considered as a black body at a temperature of 5800 K)
is in the visible region (l = 0.4 to 0.7 microns), whereas a major part of the energy radiated by earth at 300
K falls in the infra-red region.
(vi) As temperature increases, the peak of the curve shifts to the left, i.e. towards the shorter wavelengths.
(vii) Area under the curve between l and (l + dl) = Ebl .dl = radiant energy flux leaving the surface within the
range of wavelength l to (l + dl). Integrating over the entire range of wavelengths,

Eb =
z
0
¥
(El ) dl = s ×T4

Eb is the total emissive power (also known as ‘radiant energy flux density’) per unit area radiated from a
...(13.8)

black body, and s is the Stefan–Boltzmann constant = 5.67 ´ 10 –8W/(m 2 K4).


Corollaries of Planck’s law:
(a) For shorter wavelengths, (C 2/l.T) becomes very large, and exp(C2/l.T) >> 1. Then, Planck’s formula
(Eq. 13.7) reduces to:

646 FUNDAMENTALS OF HEAT AND MASS TRANSFER


C1 × l- 5
Ebl =
FC I ...(13.9)
exp GH l ×T JK
2

This equation is known as ‘Wein’s law’ and is accurate within 1 % for l.T < 3000 mmK.
(b) For longer wavelengths, the factor (C 2/l.T) becomes very small, and exp(C 2/l.T) can be expanded in a
series as follows:

F C I = 1 + C + 1 ×F C I
exp G
2

H l ×T JK l × T 2 ! GH l ×T JK
2 2 2
+ ...

F C I =1+ C
exp G
H l ×T JK
2 2
i.e. (approximate)
l ×T
and, Planck’s law becomes:

C1 × l- 5 C ×T
Ebl = = 1 4 ...(13.10)
C2 C2 ×l
1+ -1
l ×T
This is known as ‘Rayleigh–Jean’s law’ and is accurate within 1 % for ll. lT > 8 ´ 105 mmK. This law is
useful in analysing long wave radiations such as radio waves.
13.3.2 Wein’s Displacement Law
It is clear from Fig. 13.7 that the spectral distribution of emissive power of a black body at a given absolute
temperature goes through a maximum. To find out the value of l max, the wavelength at which this maximum
occurs, differentiate Planck’s equation w.r.t. l and equate to zero. We get:

C1 × l- 5
Ebl =
F C I -1
exp G
(from Eq. 13.7)

H l ×T JK
2

F exp F C I - 1I ×C ×(- 5)× l - C ×l ×F exp F C I I ×FG C IJ ×(- 1)× l


dEbl
GH GH l ×T JK JK2
1 GH GH l ×T JK JK H T K
-6
1
-5 2 2 -2

i.e. = =0
dl F exp F C I - 1I 2

GH GH l ×T JK JK 2

F F -C II – C
Simplifying, GH
5× 1 - exp GH l ×T JK JK l ×T
2 2 =0

Solving this transendental Eq. for C2/l.T by trial and error, we get:
C2
= 4.965
l ×T

C2 1. 4387 ´ 10 4
Therefore, l max ×T = =
4.965 4.965
i.e. l max ×T = 2898 mmK ...(13.11)
i.e. l max is inversely proportional to the absolute temperature T, and the maximum spectral intensity shifts to-
wards shorter wavelengths as the absolute temperature is increased.
Wein’s displacement law is stated as: “product of absolute temperature and wavelength at which emissive
power of a black body is a maximum, is constant”.

RADIATION 647
Value of maximum monochromatic emissive power of a black body at a given temperature is obtained by
substituting this value of l max T (= 2898 mmK) in Planck’s equation, i.e.

FG 0.002898 IJ -5

Ebl max =
C1 × l-5 3.742 ´ 10- 16 ×
H T K
F C I F 1. 4387 ´ 10 I
=
4
exp G
H l ×T JK - 1 H 2898 JK - 1
exp G
2

i.e. Ebl max = 1.287 ´ 10–5 ×T5 W/m 3. ...(13.12)


This is an important equation which tells us that the maximum monochromatic emissive power of a black
body varies as the fifth power of the absolute temperature of the body.
In practice, this law is applied to predict very high temperatures simply by measuring the wavelength of
radiation emitted.
Dividing monochromatic emissive power of a black body, Ebl, by its maximum emissive power at the same
temperature, Ebl , we get the dimensionless ratio:
max

C1 × l- 5
F C I -1
Ebl (T )
exp GH l ×T JK
2

= -5
Ebl max (T ) C ×l
F I -1
1 max
C
exp G
H l ×T JK
2
max

F I
F I G 5
exp( 4 .965) - 1 J
J
×G
-6
Ebl (T ) 2898 ´ 10
= G J
i.e.
Ebl max (T ) H l ×T K GG exp FG 0.01439 IJ - 1JJ (where, l is in microns, and T in Kelvin...(13.12a))

H H l ×T K K
Note that RHS of Eq. 13.12a is a function of lT only. Therefore, to determine the monochromatic emissive
power, Ebl, of a black body at any given temperature T and wavelength l, first find out (Ebl /Eblmax) from Eq.
13.12a, then evaluate Eblmax from Eq. 13.12, and then multiply them together.
13.3.3 Stefan–Boltzmann Law
Monochromatic emissive power of a black body is obtained from the Planck’s law. Then, the total emissive
power of a black body over the entire wavelength spectrum is obtained by integrating Ebl. Total emissive power
(or, hemispherical total emissive power) is denoted by Eb, and is given as:

Eb =
z0
¥
Ebl dl =
z
0
¥ C1 × l- 5

exp
F C I - 1 dl
GH l ×T JK
2

Performing the integration, we get:


Eb = s ×T 4 W/m2 ...(13.13)
where, s = 5.67 ´ 10 – 8 W/(m2 K4)
s is known as ‘Stefan–Boltzmann constant’.
Eq. 13.13 is the governing rate equation for radiation from a black body. Its significance lies in the fact that
just with a knowledge of the absolute temperature of a surface, one can calculate the total amount of energy
radiated in all directions over the entire wavelength range.
Net radiant energy exchange between two black bodies at temperatures T 1 and T 2 is, therefore, given by:

648 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Q net = s×(T14 – T24) W/m2. ...(13.14)
13.3.4 Radiation from a Wave Band
Often, it is required to know the amount of radiation emitted in a given wave band, i.e. in a wavelength interval
between l 1 and l 2. This is expressed as a fraction of the total emissive power and is written as Fl1–l2. Then, we
can write:

z l2
Ebl dl
z
z
l1 1 l2
Fl1 _ l2 = ¥
= 4
× Ebl dl
s ×T l1
Ebl dl
0

GH z z
Fl2 l1 I
i.e. Fl1 _ l2 =
s ×T
1
4
× Ebl dl -
0 0
Ebl dl JK
i.e. Fl1 _ l2 = F0 _ l2 – F0 _ l1. ...(13.15)
Above formula is not very convenient to use, since Ebl depends on absolute temperature T, and it is not
practicable to tabulate F0–l for each T. This difficulty is overcome by expressing F0 –l as follows:

F0 _ l =
z0
l
Ebl dl

s ×T 4
=
z0
l ×T
Ebl d ( l × T )

s ×T 5
i.e. F0 _ l = f(l×T) ...(13.16)
i.e. now, F0 _ l is expressed as a function of the product of wavelength and absolute temperature ( = l.T) only.
Values of F0 _ l vs. l.T are tabulated in Table 13.2 and plotted in Fig. 13.8.
Note that the units of product l.T is (micronKelvin).
Therefore,

i.e.
Fl1_l2 = Fl1× T _ l2 ×T =

Fl1_l2 = F0_ l2×T – F0 _ l1×T .


1
s
×
LM
N z
0
l 2 ×T Ebl
T 5
d (l × T ) -
z
0
l 1 ×T Ebl
T 5
d (l ×T )
OP
Q
...(13.17)
13.3.5 Relation between Radiation Intensity and Emissive Power
Consider a differential black emitter dA 1 radiating into a hemisphere of radius r, with the centre of the hemi-
sphere located at dA 1. To get a relation between the intensity of radiation and the emissive power, we first

1
0.9
0.8
0.7
0.6
l

0.5
F0

0.4
0.3
0.2
0.1
0
100 1000 10000 100000
l×T (micron K)
FIGURE 13.8 Fraction of black body radiation in the range (0–l ×T)

RADIATION 649
TABLE 13.2 Radiation functions
l T (mmK) F0 – l lT ( m mK) F0 – l
400 0 7000 0.8081
600 0 7200 0.8192
800 0.000016 7400 0.8295
1000 0.00032 7600 0.848
1200 0.00213 7800 0.848
1400 0.0078 8000 0.8563
1600 0.0197 8500 0.8746
1800 0.0393 9000 0.89
2000 0.0667 9500 0.9031
2200 0.1009 10000 0.9142
2400 0.1403 10500 0.9237
2600 0.1831 11000 0.9319
2800 0.2279 11500 0.9399
3000 0.2732 12000 0.9451
3200 0.3181 12500 0.9505
3400 0.3617 13000 0.9551
3600 0.4036 13500 0.9592
3800 0.4434 14000 0.9628
4000 0.4809 14500 0.9661
4200 0.516 15000 0.9689
4400 0.5488 16000 0.9738
4600 0.5793 17000 0.9776
4800 0.6075 18000 0.9808
5000 0.6337 19000 0.9834
5200 0.659 20000 0.9855
5400 0.6804 25000 0.9922
5600 0.701 30000 0.9953
5800 0.7201 35000 0.9969
6000 0.7378 40000 0.9979
6200 0.7541 45000 0.9985
6400 0.7962 50000 0.9989
6600 0.7832 75000 0.9997
6800 0.7961 100000 0.9999

calculate the rate of energy falling on a differential area dA 2 on the surface of the hemisphere using the definition
of intensity, then calculate the rate of energy falling on the whole of the hemisphere by integrating, and then
equate this amount to the rate of radiant energy issuing from the black surface dA 1.
Let the rate of radiant energy falling on dA 2 be dQ. Solid angle subtended by dA 2 at the centre of the sphere,
dw = dA 2/r 2. Projected area of dA 1 on a plane perpendicular to the line joining dA 1 and dA 2 = dA 1 .cos (q). Then,
by definition, intensity of radiation is the rate of energy emitted per unit projected area normal to the direction of
propagation, per unit solid angle, i.e.
dQ
Ib =
dA1 × cos(q ) × dw

dQ
i.e. Ib = ...(13.18)
dA2
dA1 × cos(q ) ×
r2
But, it is clear from Fig. 13.9 that differential area dA 2 is equal to:
dA 2 = (r×dq)×(r×sin(q)×df )
i.e. dA 2 = r 2 ×sin (q)×dq ×df ...(13.19)

650 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Then, from Eqs. 13.18 and 13.19, n
dQ = Ib ×dA 1 ×sin (q )×cos(q )×dq×df r.sin(q) dA2 = r2.sin(q).dq.df
Then, total rate of radiant energy falling on the
hemisphere, Q, is obtained by integrating this value of
dQ over the entire hemispherical surface. Noting that
r.dq
the whole of hemispherical surface is covered by tak-
ing q from 0 to (p/2) and, f from 0 to (2.p), we write: q

z z
Black
p
2 ×p emitter dq
2 r.sin(q).df
Q = Ib ×dA 1 × sin (q)×cos (q) dq df
q =0 f=0 dA1

z
p f df
i.e. Q = 2×p ×Ib ×dA 1 × 2 sin (q)×cos(q)dq
q=0

z
p FIGURE 13.9 Radiation from a differential area
i.e. Q = p×Ib ×dA 1 × 2 2×sin(q)×cos(q)dq dA1 to surrounding hemisphere

z
q =0
p
i.e. Q = p×Ib ×dA 1 × 2 sin(2×q)dq
q =0
i.e. Q = p×Ib ×dA 1 ...(13.20)
But, Q is also equal to: Eb ×dA 1
Therefore, Eb ×dA 1 = p×Ib ×dA1
or, Eb = p×Ib ...(13.21)
i.e. Total emissive power of a black (diffuse) surface is equal to p times the intensity of radiation.
This is an important relation, which will be used while calculating the view factors required to determine net
energy exchange between surfaces.
13.3.6 Emissivity, Real Surface and Grey Surface
As already stated, a ‘black body’ is an ideal, and it emits maximum amount radiation at a given temperature; a
black body also absorbs all the radiation incident on it. A perfect black body does not exist in practice, but this
concept is useful as a standard to compare radiation properties of different bodies.
Real surfaces always emit less radiation as compared to a black body.
Emissivity (e) of a surface is defined as ‘the ratio of radiation emitted by a surface to that emitted by a black
body at the same temperature’. Value of e varies between 0 and 1. For a black body, e = 1, and emissivity of a
surface is a measure of how closely that surface approaches a black body.
Emissivity of a surface is not a constant, but depends on nature of the surface, temperature, wavelength,
method of fabrication, etc. For example, oxide film on a metal surface increases its emissivity. Emissivity of alloys
is greater than that of pure metals. And, emissivity of semi-conductors is greater than 0.8 at 100 deg.C and goes
on decreasing with rising temperature. Dielectric materials have higher values of emissivity as compared to that
of pure metals, and in this case also, emissivity decreases with increasing temperature.
el refers to the emissivity at a given wavelength, l, and is known as spectral emissivity. When it is averaged
over all wavelengths, it is known as total emissivity.
Similarly, eq refers to emissivity in a given direction, q, where q is the angle made by the direction consid-
ered with the normal to the surface; this is known as directional emissivity. When eq is averaged over all direc-
tions, it is known as hemispherical emissivity. Thus, the total hemispherical emissivity (e) of a surface is the
average emissivity over all directions and all wavelengths and is expressed as:
E (T ) E (T )
e (T) = = ...(13.22)
Eb (T ) s ×T 4
where, E(T) is the emissive power of the real surface. Similarly, spectral emissivity is defined as:
El (T )
el (T) = ...(13.23)
Ebl (T )

RADIATION 651
TABLE 13.3 Emissivity values for a few surfaces at room temperature
Surface e
Aluminium:
Polished 0.03
Anodised 0.84
Foil 0.05
Copper:
Polished 0.03
Tarnished 0.75
Stainless Steel:
Polished 0.21
Dull 0.60
Concrete 0.88
White marble 0.95
Red brick 0.93
Asphalt 0.90
Black paint 0.97
Snow 0.97
Human skin 0.97

1.0 Emissivity values for a few surfaces at room tem-


Nonconductor perature are given in Table 13.3. More detailed listing is
given in Handbooks.
Generally, for simplification of calculations in radia-
eq
tion heat transfer, we make ‘grey’ and ‘diffuse’ approxi-
0.5 mations.
A surface is said to be grey if its properties are inde-
pendent of wavelength, and a surface is diffuse if its
Conductor properties are independent of direction.
A black body is perfectly diffuse, and, real bodies,
though not perfectly diffuse, come quite close to it. As an
0
example, a qualitative graph of directional emissivity, eq
0 15 30 45 60 75 90
with q, for electrical conductors and nonconductors, is
q given in Fig. 13.10 (q is measured from the normal to the
FIGURE 13.10 Variation of emissivity surface, and q = 0 means normal to the surface).
with direction It may be observed that for conductors, eq is nearly
constant for about q < 40 deg. and for nonconductors
(such as plastics), eq remains constant for q < 70 deg. Therefore, directional emissivity in the normal direction (i.e.
q = 0) is taken as true representative of hemispherical emissivity; further, in radiation analysis, generally, the
surfaces are assumed to be diffuse emitters.
Emissivities and emissive powers of black body, real surface and grey surfaces are compared in Fig. 13.11.
Grey surface approximation implies that e of grey surface is a constant, but less than that of a black surface
(= 1). In the above Fig. the grey surface curve is drawn such that areas under the emission curves of the real and

z
grey surfaces are equal. i.e.
¥
e (T)×s×T 4 = e l (T ) × Ebl (T ) dl
0
Therefore, average emissivity is given by:

e (T) =
z
0
¥
e l (T ) × Ebl (T ) dl

s ×T 4
...(13.24)

652 FUNDAMENTALS OF HEAT AND MASS TRANSFER


T = constant T = constant
el El
Black body, e = 1 Black body, Ebl
1
Grey surface
Grey surface, e = constant El = eEbl

e
Real surface
El = el Ebl

Real surface, el

0
0 l l

FIGURE 13.11 Emissivity and emissive power for black body, grey and real surfaces at a given temperature

Integrand on the RHS of the above equation has generally to be evaluated numerically. However, if the
wavelength spectrum can be divided into sufficient number of wave-bands and the emissivity can be assumed to
be constant (but different) in each band, then the integration can be performed quite easily.
For example, let the variation of spectral emissivity with wavelength be as follows:
e 1 = constant, 0 £ l £ l 1
e 2 = constant, l 1 £ l £ l 2
e 3 = constant, l 2 £ l £ ¥
Then, the average emissivity is calculated using Eq. 13.24 as follows:

e (T) =
z 0
l1
e 1 × Ebl (T ) dl

s ×T 4
+
zl1
l2
e 2 × Ebl (T ) dl

s ×T 4
+
z l2
¥
e 3 × Ebl (T ) dl

s ×T 4
...(13.25)

i.e. e (T) = e 1.F0–l1 (T) + e 2.Fl1– l2 (T) + e 3.Fl2–¥ (T) ...(13.26)


Factors F0–l1 (T), etc., can easily be determined using Table 13.2.
It should be clearly understood that emissivity values strongly depend on the surface conditions, oxidation,
roughness, cleanliness, type of finish, etc. So, there is always an element of uncertainty while using reported
values.
13.3.7 Kirchhoff ’s Law Black body, e = 1
Kirchhoff’s law establishes a relation between the total, hemispherical T
emissivity, e of a surface and the total, hemispherical absorptivity. This Eb
is a very useful equation in calculating the net radiant heat loss from E
surfaces.
Consider a small, grey body of area A, emissivity e and tempera- T
ture T be located inside an isothermal enclosure maintained at the same
temperature T. Since the enclosure (or, cavity) is isothermal, its behav-
iour can be taken as that of a black body, irrespective of its surface FIGURE 13.12 Kirchhoff’s Law
properties. Also, since the grey body inside the enclosure is small, it
does not affect the black body nature of the enclosure.
Now, radiation incident on the small body is equal to the radiation emitted by the black body at temperature
T, i.e. G = Eb (T) = s.T 4, per unit surface area. And, the radiation absorbed by the small body per unit surface area
= Gabs = a.G = a.s.T4. Further, radiation emitted by the small body per unit area of its surface = E = e.s.T 4.
Since both the small body and the enclosure are at the same temperature, T, they will be in thermal equilib-
rium and the net heat transfer rate to the small body must be equal to zero.
i.e. radiation emitted by the small body = radiation ab-sorbed by the small body,
i.e. A.e.s.T 4 = A.a.s.T 4

RADIATION 653
or,
e (T) = a(T) ...(13.27)
Eq. (13.27) represents Kirchhoff’s law. Kirchhoff’s law states that the “total hemispherical emissivity, e of a
grey surface at a temperature T is equal to its absorptivity, a for black body radiation from a source at the
same temperature T.”
Note the important restrictions on Eq. 13.27: one, incident radiation must be from a black body, and, second,
black body must be at the same temperature as that of the other body. However, for practical purposes, we
assume that the emissivity and absorptivity of a surface are equal, even when that surface is not in thermal
equilibrium with the surroundings, since absorptivity of most of the real surfaces is not very much sensitive to
temperature and wavelength.
Similar to Eq. 13.27, we can write for monochromatic radiation,
e l (T) = a l (T) ...(13.28)
Example 13.1. Incident radiation (G = 1577 W/m2) strikes an object. The amount of energy absorbed is 472 W/m2 and
the amount of energy transmitted is 78.8 W/m2. What is the value of reflectivity?
Solution.
Data:
G := 1577 W/m2 Qa := 472 W/m2 Qt := 78.8 W/m2
Let Q r be the reflected radiation.
Then, we have: Qr := G – Qa – Qt
i.e. Qr = 1.026 ´ 10 3 W/m2 (reflected radiation)
Therefore, reflectivity r is given by:
Qr
r :=
G
i.e. r = 0.651 (reflectivity.)
Example 13.2. A hole of area dA = 2 cm2 is opened on the surface of a large spherical cavity whose inside is maintained
at 1000 K. Calculate: (a) the radiation energy streaming through the hole in all directions into space, (b) the radiation
energy streaming per unit solid angle in a direction making a 60 deg. angle with the normal to the surface of the
opening.
Solution. See Fig. Ex. 13.2.
Data:
dA := 2 ´ 10 –4 m2 T := 1000 K s := 5.67 ´ 10– 8 W/(m2K4 ) (Stefan–Boltzann const.) q := 60 deg.
p
i.e. q := 60 × (radians)
180
(a) Radiation streaming out in all directions:
Since the spherical cavity can be considered as a black body, energy streaming out is given by Stefan–Boltzmann law:
i.e. Q := dA× s× T 4
or, Q = 11.34 W (radiation energy streaming through the hole.)
(b) Radiation streaming out through unit solid angle, in a direction making 60 deg. with normal:
Now, we have the relation: Eb = p× lb, where Eb = Emissive power, and lb = Intensity of radiation.

Radiation
streaming out in n
all directions
T = 1000K
q = 60°

dA

(a) (b)

FIGURE Example 13.2 Radiation streaming out from a hole on the surface of a sphere

654 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Eb
Therefore, Ib =
p
But, Eb = s×T 4 W/m2 (for a black body, by Stefan–Boltzmann law)
s ×T 4
i.e. Ib := W/(m2 sr) (intensity of radiation)
p
4 2
i.e. Ib = 1.805 ´ 10 W/(m sr)
and, radiation streaming out in that direction:
Q := Ib ×dA×cos (q) W (where q is in radians (Note: while using Mathcad, q must be in radians while calculating cos(q).))
i.e. Q = 1.805 W (radiaion through a solid angle of unity, in a direction of 60 deg. with normal to the surface.)
Example 13.3. It is observed that intensity of radiation is maximum in case of solar radiation at a wavelength of 0.49
microns. Assuming the sun as a black body, estimate its surface temperature and emissive power. Wein displacement
constant = 0.289 ´ 10 –2 mK.
Solution.
Data:
l max := 0.49 microns s := 5.67 ´ 10 –8 W/(m2 K4) (Stefan–Boltzmann constant)
By Wein’s displacement law, we have :
l max ×T = 2890 microns K
Therefore,
2890
T :=
l max
i.e. T = 5.898 ´ 10 3 K (surface temperature of sun)
Heat flux at the surface Eb:
Sun can be considered as a black body; then, from Stefan–Boltzmann law:
Eb := s× T 4
i.e. Eb = 6.86 ´ 10 7 W/m2 (heat flux at the surface of sun.)
Example 13.4. The temperature of a body of area 0.1 m2 is 900 K. Calculate the total rate of energy emission, intensity of
normal radiation in W/(m2sr), maximum monochromatic emissive power, and wavelength at which it occurs.
Solution.
Data:
T := 900 K A := 0.1 m2 s := 5.67 ´ 10 –8 W/(m2 K4) (Stefan–Boltzmann constant)
Total rate of energy emission:
Eb := s× T4 (from Stefan–Boltzmann law)
i.e. Eb = 3.72 ´ 104 W/m2 (total emissive power)
Therefore,
Q := Eb ×A
i.e. Q = 3.72 ´ 103 W (total energy emission from surface)
Intensity of normal radiation:
(Note that for a black (diffuse) surface, intensity is the same in all directions.)
Eb
I :=
p
i.e. I = 1.184 ´ 10 4 W/(m2sr)
Wavelength of maximum monochromatic emissive power:
From Wein’s displacement law, we have:
lm ×T = 2898 mmK
Therefore,
2898
l m := mm
T
i.e. lm = 3.22 mm (wave length for maximum monochromatic emissive power at 900 K.)
Maximum monochromatic emissive power:
We use Planck’s law, with values of constants, C1 and C 2:
C1 := 3.742 ´ 108 W/(mm4/m2 )
and, C2 := 1.4387 ´ 104 mmK

RADIATION 655
C1 × l-m5
Eblmax := (Planck’s law)
F C I -1
exp GH l ×T JK
m
2

i.e. Eblmax := 7.6 ´ 103 W/(m2.micron) (maximum monochrome emissive power.)


Alternatively:
We can directly apply Eq.13.12:
Eblm := 1.287 ´ 10 –5 ×T 5 W/m 3. ...(13.12)
i.e. Ebl m = 7.6 ´ 109 W/m3 = 7.6 ´ 10 3 W/(m2 mm) (same as obtained above.)
Example 13.5. Window glass transmits radiant energy in the wavelength range 0.4 mm to 2.5 mm. Determine the fraction
of total radiant energy which is transmittted, when the source temperature is: (a) 5800 K (i.e. sun’s surface temperature),
and (b) 300 K (i.e. room temperature).
Solution.
Data:
T 1 := 5800 K T2 := 300 K l1 := 0.4 mm l2 := 2.5 mm
s = 5.67 ´ 10 –8 W/(m2 K4 ) (Stefan–Boltzmann constant)
Case (a): Source temperature T 1 = 5800 K
We use Table 13.2 where radiation functions are tabulated against the product (l.T).
We have:
l 1 ×T 1 = 2.32 ´ 10 3 mm/K
l 2 ×T1 = 1.45 ´ 104 mm/K
Corresponding to these lT values, we get, from Table 13.2:
F0_l1 := 0.12454 (by interpolation)
and, F0_l2 := 0.9661
Therefore, fraction transmmitted is equal to:
0.9661 – 0.12454 = 0.842
i.e. 84.2% of the energy coming from the sun (at 5800 K) is transmitted through the window glass.
Case (b): Source temperature T2 = 300 K
Again, we use Table 13.2 where, radiation functions are tabulated against the product (l.T).
We have:
l 1 ×T 2 = 120 mmK
l2 × T2 = 750 mmK
Corresponding to these lT values, we get, from Table 13.2:
F 0_l2 = 0.0
and, F0_l2 = 1.2 ´ 10 –5 = almost zero,
i.e. practically no energy will be transmitted through the window glass in this wavelength range, if the source tempera-
ture = 300 K. In other words, glass is ‘opaque’ to radiation at 300 K in the wavelength range 0.4 mm to 2.5 mm.
As mentioned in text, this is the principle of ‘Greenhouse effect’, wherein radiation from a high temperature source
(i.e. sun) is allowed to pass through the glass into the enclosure of the greenhouse, while radiation at a relatively low
temperature from within the enclosure, is not allowed to es-
el cape out. This, in effect, causes an increase in the tempera-
ture of the space within the enclosure.
1 Example 13.6. Spectral emissivity of a particular surface at
0.8 800 K is approximated by a step function, as follows: e 1 = 0.1
for l = 0 to 2 mm, e 2 = 0.5 for l = 2 to 15 mm, and e 3 = 0.8 for
l = 15 to ¥. Calculate (i) the total (hemispherical) emissive
0.5 power, and (ii) total hemispherical emissivity, e over all
wavelengths.
Solution.
0.1 Data:
0 T := 800 K (temperature)
0 2 10 16 l, mm e1 := 0.1 (emissivity in wavelength range: 0 to 2 mm)
e2 := 0.5 (emissivity in wavelength range: 2 to 15 mm)
FIGURE Example 13.6 Spectral emissivity distribu- e3 := 0.8 (emissivity in wavelength range: 15 mm to ¥.)
tion against wavelength s := 5.67 ´ 10 – 8 W/(m2 K4 ) (Stefan–Boltzmann constant)

656 FUNDAMENTALS OF HEAT AND MASS TRANSFER


l1 := 2 mm
l2 := 15 mm
l3 := ¥
Planck’s law gives spectral emissive power of a black body.

z
For a non-grey surface considered in this problem, we can write:
¥
Total emissive power: E= e l (l )× Ebl dl
0

Variation of e l with l is specified in the problem.

z z z
Therefore, splitting the above integral into parts:
2 15 ¥
E= e l (l )× Ebl dl + e l (l )× Ebl dl + e l (l )× Ebl dl

z z z
0 2 15

2 15 ¥
i.e. E = 0.1 × Ebl dl + 0.5 × Ebl dl + 0.8 × Ebl dl
0 2 15

E
Then, e = = e 1 ×(F0_l1) + e 2 ×(Fl1_l2) + e 3 ×(Fl3_infinity)
Eb
i.e. e = e 1 ×(F0_l1) + e 2 ×(F0_l2 – F0_l1) + e 3 ×(F0_ infinity – Fl0_l2) ...(a)
Values of F 0–l1, etc., are obtained from Table 13.2.
We have
l1 × T = 1.6 ´ 10 3 (correspondingly, we get: F0_ l1 = 0.0197)
l2 × T = 1.2 ´ 104 (correspondingly, we get: F0_l2 = 0.9451)
l3 × T = ¥ (correspondingly, we get: F 0_l3 = 1)
Then, from Eq. a:
e := 0.0197 + 0.5×(0.9451 – 0.0197) + 0.8×(1 – 0.9451)
i.e. e = 0.516 (Total hemispherical emissivity over all wavelengths. )
And, total emissive power of this surface is given by:
E := e ×s×T 4 W/m2 (total emissive power)
i.e. E = 1.199 ´ 104 W/m2 (total emissive power.)

13.4 The View Factor and Radiation Energy Exchange between Black Bodies
So far, we studied the fundamental laws of radiation and radiative properties of surfaces. But, in practical situa-
tions, we are mostly interested in radiative heat exchange between surfaces. The radiative heat exchange may be
only between two surfaces, or from one or more surfaces in an enclosure. If the surfaces involved are ‘black’,
then, the problem is simplified since the radiation falling on a black surface is completely absorbed and none is
reflected; however, if the surfaces are ‘grey’, then the problem is slightly more complicated since one has to take
into account the multiple reflections from surfaces. In either case, the radiative heat exchange depends on:
(i) absolute temperatures of surfaces
(ii) radiative properties of surfaces, and
(iii) geometry and relative orientation of the surfaces involved.
Point (iii) mentioned above is obvious since, generally, in engineering problems, we assume the surfaces to
be ‘diffuse’, i.e. radiation is emitted in all possible directions, and all of the energy emitted by surface 1 may not
be intercepted by surface 2. This statement is quantified by what is known as ‘View factor’. View factor is also
known by other names such as: ‘configuration factor’, ‘shape factor’, ‘angle factor’, etc.
View factor is defined as the fraction of radiant energy leaving one surface which strikes a second surface directly.
Here, ‘directly’ means that reflection or re-radiated energy is not considered. View factor is denoted by F12, where
the first subscript, 1 stands for the emitting surface, and the second subscript, 2 stands for the receiving surface.
We have:
F12 = (Direct radiation from surface 1 incident on surface 2) divided by (Total radiation from emitting
surface 1).
We desire to develop a general relation for view factor between two surfaces.
Infinetisimal areas:
As a first step, consider differential areas dA 1 and dA 2 on two black surfaces A 1 and A 2 exchanging heat by
radiation only. See Fig. 13.13.
RADIATION 657
dA2 dA 1 and dA 2 are at a distance ‘r’ apart and the normals to
n A2
these areas make angles f1 and f2 with the line connecting
them, as shown. Then, using the definition of intensity, we
f2 can write:
Energy leaving dA1 and falling on dA 2 = dQ 12 =
n Black surface, e = 1 Intensity of black body dA 1 x projected area of dA1 on a plane
r perpendicular to line joining dA1 and dA2 x solid angle sub-
tended by dA 2 at dA1.
dA1 f1 FG dA × cos(f ) IJ
H r K
2 2
Black surface, e = 1 i.e. dQ 12 = Ib1 ×(dA 1 ×cos (f 1))× 2
...(13.29)
A1
Now, total energy radiated from dA 1 is given by:
dQ 1 = Eb1 ×dA 1
FIGURE 13.13 Areas and angles used in i.e. dQ 1 = (p ×Ib1)×dA 1 ...(13.30)
derivation of view factor relation Then, by definition, the view factor FdA1-dA2 is the ratio of
dQ 12 to dQ 1:
cos(f 1) × cos(f 2) × dA2
FdA1-dA2 = ...(13.31)
p ×r 2
Note that the view factor involves geometrical quantities only.
Eq. 13.31 gives the view factor between two infinetisimal areas. Such a situation is encountered even when
finite areas are involved, when the distance between these two areas ‘r’, is very large.
Infinitesimal to finite area: i.e. the emitter is very small and the receiving surface is of finite size. Here,
integration over the entire surface A 2 has to be considered.
Again, remembering the definition of view factor, and forming the ratio of dQ 12 to dQ 1:

FdA1–A2 =
z
A2
Ib 1 (cos(f 1) dA1 )cos(f 2 )
dA2
r2

p Ib1 dA1
Since both Ib1 and dA 1 are independent of integration, we can write:

FdA1–A2 =
z
A2
(cos(f 1) cos(f 2 ) dA2
p r2
...(13.32)

Practical situation of calculating view factors between infinitesimal to finite areas are encountered in the case
of a small thermocouple bead located inside a pipe or a small, spherical point source radiator located by the side
of a wall, etc.
Finite to finite area: once again, from the definition of view factor:

zz I b1 (cos(f 1) dA1 )cos(f 2 )


dA2

z
FA1–A2 = A1 A 2 r2
p Ib1 dA1
A1
For constant Ib1, above equation becomes:

FA1–A2 =
1
p A1 A1 A 2 zz
(cos(f 1) cos(f 2 )
r2
dA1 dA2 ...(13.33)

It is clear from Eqs. 13.31, 13.32 and 13.33 that the view factor depends only on the relative orientation (or
spatial relation) of the two bodies; it does not depend on the emissivities of the surfaces or the temperatures.
Further, also note that the surfaces are assumed to be isothermal and diffuse emitters.

zz
In general, we write Eq. 13.33 compactly as:
1 (cos(f 1) cos(f 2 )
F 12 = dA1 dA2 ...(13.34)
p A1 A1 A 2 r2
Here, F12 means ‘the view factor from surface 1 to surface 2’.
658 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Similarly, if we desire to get the view factor from surface 2 to surface 1, we simply interchange suffixes 1

zz
and 2:
1 (cos(f 2 ) cos(f 1 )
F 21 = dA 2 dA1 ...(13.35)
p A2 A 2 A1 r2
Note that in Eqs. 13.34 and 13.35, the double integrals differ only in the order of integration, and as such,
yield the same result. Then, multiplying Eq. 13.34 by A 1, and Eq. 13.35 by A 2, and equating the double integrals,
we get:
A 1 ×F12 = A 2 ×F21 ...(13.36)
Eq. 13.36 is known as ‘reciprocity theorem’ and is a very useful and important relation. It helps one to find
out one of the view factors when the other one is known. In practice, one of the view factors which is easier to
calculate is obtained first, and the other view factor is found out next, by using the reciprocity theorem.

zz
Note: It is easier to remember the view factor relation given in Eq. 13.34 as:
(cos(f 1) cos(f 2 )
A 1 × F12 = dA1 dA2 . ...(13.37)
A1 p r2
A2
Radiation energy exchange between black bodies:
As already mentioned, analysis of heat exchange between two black bodies is simpler since a black body absorbs
all the radiation impinging on it and none is reflected.
Consider two black surfaces A1 and A2 exchanging radiation energy with each other.
Then, rate of energy emitted by surface 1, which directly strikes surface 2 is given by:
Q 12 = A 1 ×F 12 ×Eb1 = A 1 ×F 12 ×s ×T14 ...(13.38)
This energy is completely absorbed by surface 2, since surface 2 is black.
Similarly, of energy emitted by surface 2, which directly strikes surface 1 is given by:
Q 21 = A 2 ×F21 ×Eb2 = A 2 ×F21 ×s ×T24 ...(13.39)
and, net radiation exchange between the two surfaces is:
Q net = A1 ×F 12 ×s ×T 41 – A 2 ×F 21 ×s ×T24
But, A 1 ×F12 = A 2 ×F21 by reciprocity theorem
Therefore,
Q net = A1 ×F12 ×s×(T14 – T24) = A2 ×F21 ×s×(T14 – T24), W. ...(13.40)

13.5 Properties of View Factor and View Factor Algebra


We shall enumerate the salient features of view factor. View factors of some geometries are easily calculated;
however, more often, calculation of view factors for more complex shapes is quite difficult. In many cases, the
complex shapes could be broken down into simpler shapes for whom the view factors are already known or
could easily be calculated. Then, with this knowledge, the view factor for the desired complex shape could be
calculated by remembering the definition of view factor as the fraction of energy emitted by surface 1, which is
intercepted directly by surface 2, and the interrelation between the various view factors. This is known as ‘view
factor algebra’.
Properties of view factor:
(i) The view factor depends only on the geometrics of bodies involved and not on their temperatures or
surface properties.
(ii) Between two surfaces that exchange energy by radiation, the mutual shape factors are governed by the
‘reciprocity relation’, namely, A 1.F 12 = A 2.F 21.
(iii) When a convex surface 1 is completely enclosed by another surface 2, it is clear from Fig. 13.14 (a) that all
of the radiant energy emitted by surface 1 is intercepted by the enclosing surface 2. Therefore, view factor
of surface 1 w.r.t. surface 2 is equal to unity. i.e. F12 = 1. And, the view factor of surface 2 w.r.t. surface 1
is then easily calculated by applying the reciprocity relation, i.e. A 1.1 = A 2.F 21, or, F 21 = A 1/A 2.
(iv) Radiation emitted from a flat surface never falls directly on that surface (see Fig. 13.14 (b)), i.e. view factor
of a flat surface w.r.t. itself is equal to zero, i.e. F11 = 0. This is valid for a convex surface too, as shown in
Fig. 13.14 (c).
(v) For a concave surface, it is clear from Fig. 13.14 (d) that F11 is not equal to zero since some fraction of
radiation emitted by a concave surface does fall on that surface directly.

RADIATION 659
2

F12 = 1
1 F21 = A1/A2
F11 = 0 F11 = 0 F11 ¹ 0

(a) Surface 1 completely (b) Flat surface (c) Convex surface (d) Concave surface
enclosed by surface 2
FIGURE 13.14 View factors for a few surfaces

(vi) If two, plane surfaces A 1 and A 2 are parallel to each other and separated by a short distance between
them, practically all the radiation issuing from surface 1 falls directly on surface 2, and vice-versa. There-
fore, F12 = F21 = 1.
(vii) When the radiating surface 1 is divided into, say, two sub-areas A 3 and A 4, as shown in Fig. 13.15 (a), we
have:
A 1 ×F 12 = A 3 ×F32 + A 4 ×F 42 ...(13.41)
Obviously, F12 ¹ F32 + F42.
(viii) Instead, if the receiving surface A 2 is sub-divided into parts A 3 and A 4 as shown in Fig. 13.15 (b), we
have:
A 1 ×F12 = A 1 ×F 13 + A 1 ×F14 ...(13.42)
i.e. F 12 = F 13 + F 14,
i.e. view factor from the emitting surface 1 to a sub-divided receiving surface is simply equal to the sum of the
individual shape factors from the surface 1 to the respective parts of the receiving surface. This is known as
‘Superposition rule’.

A3

F13 A4
A2

A1.F12 = A3.F32 + A4.F42 F14 Receiving


surface, A2
A3

A4 A1 A1.F12 = A1.F13 + A1.F14

A1 = A3 + A4

(a) Radiating surface A1 is subdivided into A3 and A4 (b) Receiving surface A2 is subdivided into A3 and A4

FIGURE 13.15 View factors for sub-divided surfaces

(ix) Symmetry rule If two (or more) surfaces are symmetrically located w.r.t. the radiating surface 1, then
the view factors from surface 1 to these symmetrically located surfaces are identical. A close inspection of
the geometry will reveal if there is any symmetry in a given problem.
(x) Summation rule Since radiation energy is emitted from a surface in all directions, invariably, we con-
sider the emitting surface to be part of an enclosure. Even if there is an opening, we consider the opening
as a surface with the radiative properties of that opening. Then, the conservation of energy principle
requires that sum of all the view factors from the surface 1 to all other surfaces forming the enclosure,
must be equal to 1. See Fig. 13.16, where the interior surface of a completely enclosed space is sub-
divided into n parts, each of area A 1, A 2, A 3, ... , A n.
Then,
F11 + F12 + ... + F1n = 1

660 FUNDAMENTALS OF HEAT AND MASS TRANSFER


F21 + F22 + ... + F2n = 1
Fn1 + Fn2 + ... + Fnn = 1 ...(13.43)
3 i.e.
n–1
n

n 2
åF ij = 1 i = 1, 2, 3, ... n ...(13.44)
j=1
n 1 (xi) In an enclosure of ‘n’ black surfaces, maintained at tempera-
S Fij = 1 i = 1, 2, 3,..., n tures T1, T2, ... ,Tn , net radiation from any surface, say, sur-
j=1 face 1, is given by summing up the net radiation heat
transfers from surface 1 to each of the other surfaces of the
FIGURE 13.16 View factor—summa- enclosure:
tion rule for radiation in an enclosure
Q1net = A 1 ×F12 ×s ×(T14 – T24) + A1 ×F13 ×s ×(T14 – T34)
+ A 1 ×F14 ×s×(T14 – T44) + ... + A 1 ×F1n ×s ×(T14 – Tn4) ...(13.45)
Note: Often, while solving radiation problems, determination of the view factor is the most difficult part. It
will be useful to keep in mind the definition of view factor, summation rule, reciprocity relation, superposition
rule and symmetry rule while attempting to find out the view factors.

13.6 Methods of Determining View Factors


While solving problems in radiation heat transfer, required numerical values of view factors may be obtained by
the following methods:
(i) By performing the necessary integrations in Eqs. 13.31, 13.32 or 13.33. However, except in very simple
cases, most of the time, the direct integration procedure is quite difficult.
(ii) Use of readily available analytical formulas or graphs prepared by researchers for the specific geometry
in question.
(iii) Use of view factor algebra in conjunction with definition of view factor, summation rule, reciprocity
relation, superposition rule and symmetry rule.
(iv) Experimental and graphical techniques.
13.6.1 By Direct Integration
We shall demonstrate direct integration procedure with two examples:
Example 13.7. Find out the view factor from an elemental disk dA1 to a much larger disk A 2 of radius R, located directly
above and parallel to the small disk at a vertical distance L from the small disk, as shown in Fig. Example 13.7.
Solution. Area dA1 is much smaller than area A 2; so, this is the case of finding out the view factor from a differential area

z
to a finite area. So, we shall apply Eq. 13.32, i.e.
cos(f 1) cos (f 2 ) dA 2
F dA1–A2 = ...(13.32)
p r2
A2

Now, on area A2, consider a differential area dA2 of radius x and width dx as shown. Angles f 1 and f 2, made by the
line ‘r’ connecting dA1 and dA2 with the two normals are equal, since the disks are parallel.
We have: r 2 = x2 + L2 , f 1 = f 2
L
cos (f1) =
r
L
and, cos (f1) = = cos (f2)
x 2 + L2
and, differential area, dA2 = 2 ×p×x×dx.

z
Then, from Eq. 13.32 we get:
R
2 × L2 × x
F 12 = dx
( x 2 + L2 ) 2

z
0

R
x
i.e. F 12 = 2×L2 × dx
0 ( x 2 + L2 ) 2

RADIATION 661
A2 A2
dx
dA2
x
R R

f2
L L 2 2
f1 r (L + R )
2a

dA1 dA1
(a) (b)

FIGURE Example 13.7 View factor from an elemental area dA1 to a larger area A2

Now, let: u = x2 + L2
Then, du = 2×x×dx

z z
Then, expression for above integral becomes:
2×x 1 -1 -1
dx = du = =
(x 2 + L2 )2 u2 u ( x2 + L2 )
Therefore, putting the limits for x from 0 to R:
F 1 1 I
F12 = L2 × GH R 2 -
+ L2 L2 JK
R2
i.e. F12 =
R + L2
2

We can also write:

F I 2

= G JJ
R
F12
GH L2 + R 2 K
i.e. F12 = sin2 (a) (where, 2a is the angle subtended by the area A 2 at dA1 as shown in Fig. Example 13.7.)
Example 13.8. Find out the view factors F12 and F21 between two square surfaces 1 and 2, oriented towards each other as
shown in Fig. Example 13.8. Plate 1 has an area of 0.08 m2 and plate 2 has an area of 0.05 m2.
Solution.
Since both the plane surfaces are small compared to the distance between them, they can be approximated as differential
areas, i.e. we can apply Eq. 13.31 to get the view factors:
cos (f 1) × cos(f 2) × dA2
FdA1_ dA2 = ...(13.31)
p ×r2
Data:
p
dA1 := 0.08 m2 dA 2 := 0.05 m2 r := 5 m f1 := 15 deg. i.e. f 1 := 15× radians f2 := 40 deg.
180
p
i.e. f 2 := 40× radians
180
Note: f1 and f 2 are expressed in radians, since Mathcad requires that the angles be in radians while evaluating trigono-
metric functions such as sin (f), cos(f), etc.
Therefore, cos (f1) = 0.966
and, cos (f2)= 0.766
Then, from Eq. 13.31, we get:

662 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Plate 1 cos (f 1) × cos(f 2 ) × dA2
F 12 =
p ×r2
i.e. F 12 = 4.7106 ´ 10– 4
f1 = 15°
i.e. view factor from surface 1 to surface 2 is 0.00047106.
Next, to determine the view factor from surface 2 to 1, we can conven-
iently use the reciprocity rule, i.e.
r = 5m A1 ×F 12 = A2 ×F 21
In the present case, we write:
dA1 ×F12 = dA2 ×F 21
f2 = 40°
dA1 × F12
i.e. F21 :=
dA2
Plate 2 i.e. F 21 = 7.537 ´ 10 –4
i.e. view factor from surface 2 to surface 1 is 0.0007537.
FIGURE Example 13.8 View factors 13.6.2 By Analytical Formulas and Graphs
between two small plates separated
As stated earlier, determination of view factors by direct integration is
by a large distance
rather involved because of the contour integrations to be performed
over the surfaces. However, several research workers have published
analytical relations and graphs for many of the commonly encountered geometries.
Figs. 13.17 and 13.18 show a few two-dimensional and three-dimensional geometries and Tables 13.4 and
13.5 give corresponding view factor relations.
Note: In Table 13.5, atan(x) means arctan(x) or tan – 1(x).

wi

w
i j
L
j i
a

wj w
(a) Parallel plates with midlines (b) Inclined parallel plates of equal
connected by perpendicular width and with a common edge
j

wj wj
wk
k j
i
i

wi wi
(c) Perpendicular plates with a (d) Three-sided enclosure
common edge

S
diameter, D

j
i

(e) Infinite plane and row of cylinders

FIGURE 13.17 Few two-dimensional geometries, infinitely long

RADIATION 663
rj

j
j j
L Z
L
ri i

Y i i Y X
X
(a) Aligned parallel rectangles (b) Coaxial parallel disks (c) Perpendicular rectangles
with a common edge
FIGURE 13.18 Few three-dimensional geometries

TABLE 13.4 View factors for a few two-dimensional geometries

Geometry View factor relation

wi wj
Parallel plates with midlines connected Wi = WJ =
by perpendicular (See Fig. 13.17,a) L L
1 1
(Wi + W j )2 + 4 2 - (W j - Wi )2 + 4 2
Fij =
2 ×Wi

Inclined parallel plates of equal width and FG 1 ×aIJ


with a common edge (See Fig. 13.17,b)
Fij = 1 – sin
H2 K
Perpendicular plates with a common LM LM F I OP OP 2
1
2
edge (See Fig. 13.17,c) Fij =
1
2
× 1+ MM
wj
wi
- 1+
wj
wi MN GH JK PQ PP
MN QP
Three-sided enclosure (See Fig. 13.17,d) w j + w j - wk
Fij =
2 ×w i

L FDI OP Fs -D I
1 1
Infinite plane and row of cylinders
= 1 – M1 - G J
2 2 2 2
D
GH D JK
2

MN H s K
(See Fig. 13.17,d) Fij + ×tan–1
PQ s 2

Table 13.5 gives view factor relations for three important three-dimensional geometries, often required in
practice. For example, view factors between aligned parallel rectangles will be useful to calculate heat transfer
between the floor and ceiling of a room or a furnace; view factors between coaxial parallel disks will be required
to calculate the heat transfer between the top and bottom of a cylindrical furnace, and the view factors between
perpendicular rectangles is necessary to calculate the fraction of energy entering a floor through a window on the
adjacent wall, or to determine the fraction of energy radiated from the door of a furnace to the floor outside, etc.
It may be observed from the view factor relations given in Table 13.5 that even for these simple cases, the
relations are rather complex and difficult to calculate. So, generally, view factors for these (and, many other)
geometries are presented in graphical form. It is convenient to use the graphs to determine the view factors
quickly, but with the sacrifice of a little accuracy. However, if a computer is available, it is suggested that the
analytical relations given in Tables 13.4 and 13.5 could be used for better accuracy.
View factor relation for aligned, parallel rectangles of Fig. 13.18a, is shown in graphical form in Fig. 13.19.
This graph is drawn with Mathcad. Here, Fij is plotted against X/L varying from 0.1 to about 30, for given values
of Y/L (with Y/L = 0.1, 0.2, 0.4, 0.6, 1.0, 2.0, 4.0 and 10.0).

664 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 13.5 View factors for a few three-dimensional geometries

Geometry View factor relation

Aligned parallel rectangles X Y


(See Fig. 13.18,a) XX = YY =
L L

2 LM (1 + XX 2 ) ×(1 + YY 2 ) OP
A=
p × XX ×YY
B = ln
MN 1 + XX 2 + YY 2 PQ
1 LM XX OP
C = XX ×(1 + YY 2 ) 2 ×atan MM 1 PP
N (1 + YY ) 2 2
Q
1
LM YY OP
D = YY ×(1 + X X 2 ) 2 ×atan M PP
MN (1 + XX ) 1
2 2
Q
E = XX ×a tan (XX) F = YY ×atan (YY)
F i j = A ×(B + C + D – E – F )

Coaxial parallel disks ri rj 1 + R 2j


(See Fig. 13.17,b)
Ri = Rj = S= 1+
L L Ri2

LM LM O
F I OP PP
1
2 2
Fi j =
1
MM
× S - S 2 - 4×
rj
MN GH JK P P
2
MN
ri
Q PQ
Perpendicular rectangles Z Y 1 FG 1 IJ
with a common edge (See
Fig. 13.18,c)
H=
X
W=
X
A=
p ×W
B = W ×atan
HW K
FG 1 IJ 1 LM 1 OP
C = H ×atan
HH K D = (H 2 + W 2 ) 2 ×atan MM P
N (H + W ) PQ
1
2 2 2

+H ) O
LM P × LMN (1H+ ×H(1 +) ×H(H ++WW )) OPQ
W2 H2
(1 + W 2 ) ×(1 + H 2 ) W 2 × (1 + W 2 2 2 2 2
E= ×
(1 + W 2 + H 2 ) (1 + W 2 ) × (W 2 N + H )Q 2 2 2 2

FG 1 IJ
H
F i j = A× B + C - D +
4
×ln(E )
K
Graph of view factor for coaxial parallel disks (of Fig. 13.18,b) is drawn using Mathcad and is shown in Fig.
13.20. Here, view factor Fij is plotted against L/ri for different values of rj/L.
And, graph of view factors for perpendicular rectangles with a common edge (of Fig. 13.18,c), drawn using
Mathcad, is shown in Fig. 13.21. Here, view factor Fij is plotted against Z/X for different values of Y/X.
Another practically important geometry is that of two concentric cylinders of finite length. View factors
associated with this geometry are shown in Fig. 13.22.
We shall illustrate the use of analytical relations for view factors given in Table 13.5 or the Figs. 13.19 to
13.21, with an example:
Example 13.9. Find out the net heat transferred between two circular disks 1 and 2, oriented one above the other, paral-
lel to each other on the same centre line, as shown in Fig. Example 13.9. Disk 1 has a radius of 0.5 m and is maintained
at 1000 K, and disk 2 has a radius of 0.6 m and is maintained at 600 K. Assume both the disks to be black surfaces.

RADIATION 665
View factor for parallel rectangles View factor for coaxial parallel disks
1 10.0 1
4.0 8
4
2.0
1.0
0.8 2
0.6 1.5
0.4
0.6
Fij

Fij
0.1 0.2 1.0

0.4
Y/L = 0.1
0.6

0.2
rj /L = 0.3

0.01
0.1 1 10 100 0.1 1 10
X/L L/ri
FIGURE 13.19 View factor for aligned, FIGURE 13.20 View factor for coaxial,
parallel rectangles (See Fig. 13.18a) parallel disks (See Fig. 13.18b)

View factor for perpendicular rectangles


0.5
2
Y/X = 0.0
0.05

0.4 0.1

0.2
0.3
Fij

0.6
1.0
0.2

2.0

0.1
4.0
10.0

0
0.1 1 10
z/x
FIGURE 13.21 View factors for coaxial, perpendicular rectangles with a common edge (See Fig. 13.18c)

Solution.
Data:
ri := 0.5 m rj := 0.6 m L := 1 m T1 := 1000 K T2 := 600 K s := 5.67 ´ 10– 8 W/m2K
2 2 2 2
A1 := p×ri i.e. A1 = 0.785 m A2 := p×rj i.e. A2 = 1.131 m
This is the case of heat transfer between two black surfaces. So, we use Eq. (13.40), i.e.
Q net = A1 × F12 ×s×(T 14 – T 24)
= A2 × F21 ×s×(T14 – T24), W ...(13.40)
So, the problem reduces to calculating the view factor F 12 or F 21. We can easily find out F 12 using Fig. 13.20. How-
ever, we can determine F 12 analytically more accurately with Mathcad using the view factor relation given in Table 13.5
for coaxial parallel disks.
We re-write the view factor relation given in Table 13.5 as follows, for ease of calculation with Mathcad:

666 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1.0
A2
L
r1
0.8
r2
A1 1.0
0.6 0.9
0.8
F2 ® 1 ¥ L/
= 0.7 r2
/ r2 =
0.4 L 0.6 4 ¥
2 F2 ® 2
1 0.5 2
5
0. 0.4
25 1
0.2 0. 0.3
1
0. 0.2 0.5
0.25
0.1
0 0
0.2 0.4 0.6 0.8 1.0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
r1/r2 r1/r2

FIGURE 13.22 View factors for two concentric cylinders of finite length: (a) outer cylinder to inner cylinder (b)
outer cylinder to itself (Source: Cengel, Yunus A. [1998]. Heat Transfer: A Practical Approach. Pub.: McGraw-Hill)

ri rj 1 + Rj2 T2 = 600 K rj = 0.6 m


Ri := Rj = S(Ri , Rj) = 1 +
L L Ri2

LM L O
F I OP PP
1

j MM
2
Rj 2
1
MM e
F12 , (Ri ×Rj) := × S Ri , R j - S ( Ri , R j ) 2 - 4 × GH JK P P ri = 0.5 m L=1m
2
MN N Ri
Q PQ T1 = 1000 K

(view factor for coaxial parallel disks)


Here, first, S is written as a function of Ri and Rj where, Ri = ri /
FIGURE Example 13.9 Coaxial
L and Rj = rj/L. Then, F12 is expressed as a function of Ri and Rj. Now,
parallel disks
F 12 is easily obtained for any values of Ri and Rj by simply writing
F12 ×(Ri , Rj) = .
Therefore, in this case, R i = 0.5
and, Rj = 0.6
We get: F12 (0.5, 0.6) = 0.232
Verify This result may be verified from Fig. 13.20 where, F 12 is plotted against L/rj for various values of rj/L. Now, for
our problem, L/ri = 1/0.5 = 2, and rj /L = 0.6/1 = 0.6. Then, from Fig. 13.20, we read F12 = 0.232, approximately,
i.e. F12 := 0.232
Therefore, net transfer between disks 1 and 2:
Qnet := A1 × F12×s×(T14 – T24) W (from Eq. 13.40)
i.e. Qnet = 8.992 ´ 10 3 W.

13.6.3 By Use of View Factor Algebra


Often, we have to find out view factors for geometries for which readily no analytical relations or graphs are
available. In such cases, sometimes, it may be possible to get the required view factor in terms of view factors of
already known geometries, by suitable manipulation using view factor algebra. For this purpose, we remember
the definition of view factor (as the fraction of energy emitted by surface 1 and directly falling on surface 2), and
invoke the summation rule, reciprocity relation, and inspection of geometry.
We shall illustrate this procedure with some important examples:
Example 13.10. Find out the net heat transferred between the areas A1 and A2 shown in Fig. Example 13.10. Area 1 is
maintained at 700 K, and area 2 is maintained at 400 K . Assume both the surfaces to be black.
Solution. This is the case of heat transfer between two black surfaces. So, we use Eq. 13.40 i.e.

RADIATION 667
Qnet = A 1 ×F12×s×(T14 – T24) = A2 ×F21 ×s×(T14 – T24), W ...(13.40)
So, the problem reduces to calculating the view factor F12 or F21. We
A1 (700K) A5 see that to calculate F12 for areas A1 and A2 as oriented in the Fig. Example
2m 13.10 we do not readily have an analytical relation or a graph. Let us de-
note the combined areas (A1 + A3) by A5 and (A2 + A4) by A6. Then, we see
A3
that A5 and A6 are perpendicular rectangles which have a common edge,
3m A6 and we have graphs or analytical relation for the view factor for such an
A4 orientation. Then, we resort to view factor algebra, as follows:
3m Remember that by definition, view factor F12 is the fraction of radiant
A2(400K)
energy emitted by surface 1 which falls directly on surface 2. Looking at
2m the Fig. Example 13.10 we can say that fraction of energy leaving A1 and
5m
falling on A2 is equal to the fraction falling on A6 minus the fraction falling
on A4.
FIGURE Example 13.10 Perpendicular
rectangles with a common edge i.e. F12 = F 16 – F14 (by definition of view factor)

A6 A
i.e. F12 = F 61 × – F41 × 4 (since by reciprocity relation, A1 × F16 = A6×F16, and A1 × F14 = A4 × F41.)
A1 A1

A6 A
i.e. F12 = ×(F65 – F63) – 4 ×(F45 – F43) (Eq. A ... using the definition of view factor, as done in first step above)
A1 A1
Now, observe that view factors F65, F63, F45 and F43 refer to perpendicular rectangles with a common edge, and can
be readily obtained from Fig. 13.21, or by analytical relation given in Table 13.5.
We re-write the view factor relation for perpendicular rectangles with a common edge, given in Table 13.5 as
follows, for ease of calculation with Mathcad:
Z Y 1 FG 1 IJ
H :=
X
W :=
X
A(W) :=
p ×W
B(W) = W×atan
HW K
FG 1 IJ 1
LM 1
OP
C(H) := H×atan
H HK D(H, W) := (H 2 + W 2) 2 ×atan MM 1 PP
N (H 2
+ W 2)2 Q
E(H, W) :=
LM
(1 + W 2 ) × (1 + H 2 ) W 2 × (1 + W 2 + H 2 )
×
OP
LM H ×(1 + H + W ) OP
W2

×
2 2 2
H2

(1 + W 2 + H 2 ) MN
(1 + W 2 ) × (W 2 + H 2 ) PQ
MN (1 + H ) ×( H + W ) PQ 2 2 2

F
F (H, W) := A (W)× G B (W ) + C ( H ) - D ( H , W ) + × ln (E ( H , W ))J
1 I
ij
H 4 K (Eq. B...view factor for coaxial
perpendicular rectangles with
a common edge)

To find F 65 :
X := 5 Y := 5 Z := 5 (w.r.t. Fig. 13.18 (c) and Fig. Example 13.12)
Z
H := i.e. H= 1
X
Y
W := i.e. W= 1
X
Therefore,
Fij (1, 1) = 0.2 (substituting in Eq. B)
i.e. F65 := 0.2 (view factor from area A 6 to A5)
Note: This value can be verified from Fig. 13.21 also.
To find F63 :
X := 5 Y := 5 Z = 3 (w.r.t. Fig. 13.18 (c) and Fig. Example 13.12)
Z
H := i.e. H = 0.6
X
Y
W := i.e. W= 1
X

668 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore,
Fij (0.6, 1) = 0.161 (substituting in Eq. B)
i.e. F63 := 0.161 (view factor from area A 6 to A3)
Note: This value also can be verified from Fig. 13.21.
To find F 45:
X := 5 Y := 3 Z := 5 (w.r.t. Fig. 13.18 (c) and Fig. Example 13.10)
Z
H := i.e. H= 1
X
Y
W := i.e. W = 0.6
X
Therefore,
Fij (1, 0.6) = 0.269 (substituting in Eq. B)
i.e. F45 := 0.269 (view factor from area A 4 to A5)
Note: This value also can be verified from Fig. 13.21.
To find F 43:
X := 5 Y := 3 Z := 3 (w.r.t. Fig. 13.18 (c) and Fig. Example 13.10)
Z
H := i.e. H = 0.6
X
Y
W := i.e. W = 0.6
X
Therefore,
Fij (0.6, 0.6) = 0.231 (substituting in Eq. B)
i.e. F 43 := 0.231 (view factor from area A 4 to A3)
Note: This value also can be verified from Fig. 13.21.
Areas:
From Fig. Example 13.10, we have:
A 1 := 10 A2 := 10 A3 := 15 A4 := 15
A 5 := 25 A6 := 25 m2
Then, from Eq. A:

F12 :=
LM A ×(F
6
- F63 ) -
A4 OP
× ( F45 - F43 )
NA1
65
A1 Q
i.e. F12 := 0.041 (view factor from A 1 to A2.)
Note: F 21 can be calculated, if required, by reciprocity relation, i.e. A1 .F12 = A2 .F21
Therefore, net heat transfer between surfaces 1 and 2:
Qnet = A1 ×F12× s × (T14 – T24) W (from Eq. 13.40)
Here, we have:
T1 := 700 K (temperature of surface A1)
T2 := 400 K (temperature of surface A2)
s := 5.67 ´ 10–8 W/m2K (Stefan–Boltzmann constant)
Therefore,
Qnet := A1×F12×s ×(T14 – T24)
i.e. Qnet = 4.926 ´ 103 W.
Example 13.11. Find out the relevant view factors for the geometries shown in Fig. Example 13.11:
(a) a long tube with cross section of an equilateral triangle
(b) a black body completely enclosed by another black body
(c) diagonal partition inside a long square duct
(d) sphere of diameter d inside a cubical box of sides, L = d
(e) hemispherical surface closed by a plane surface, and
(f) the end and surface of a circular cylinder whose length is equal to diameter.
Solution. General principle in solving these problems is to invoke: Summation rule, reciprocity theorem, inspection of
geometry for symmetry, and of course, remembering the definition of view factor:

RADIATION 669
1
2 3 L 2
2

1
3
1
(a) Equilateral triangle (b) A black body completely (c) Diagonal partition in
enclosed by another black body a long square duct

2 1
1 L=d r
2 2 L=d
1 3
(d) Sphere inside a (e) Hemispherical bowl (f) Cylinder with length
cubical box = diameter
FIGURE Example 13.11 Different geometries

(a) Long tube with cross section of equilateral triangle: See Fig. 13.11a.
For surface 1: F 11 + F 12 + F 13 = 1 (summation rule)
But, F11 = 0 (since surface 1 is flat and cannot ‘see’ itself.)
Therefore, F 12 + F 13 = 1
Now, by inspection of geometry, we find that surfaces 2 and 3 are located symmetrically w.r.t. surface 1, since it is
an equilateral triangle. Therefore, radiation from surface 1 is divided equally between surfaces 2 and 3.
i.e. F 12 = F13 = 0.5
Similarly, for surface 2, we write:
F 21 + F 23 = 1
i.e. F 23 = 1 – F 21
A1
But, F 21 = ×F 12 = F 12 (since A 1 = A2)
A2
Therefore, F23 = 1 – F 12 = 0.5
Similarly, for surface 3.
(b) Black body enclosed inside a black enclosure: See Fig. Ex. 13.11b
For surface 1: F 11 + F 12 = 1 (by summation rule)
and, A1 ×F 12 = A2 ×F 21 (by reciprocity)
A2
i.e. F12 = ×F21
A1
Now, F 11 = 1 – F 12
A
i.e. F 11 = 1 – 2 ×F21
A1
But, F21 = 1 (since all the energy radiated by surface 2 is directly intercepted by surface 1.)
A
Therefore, F 11 = 1– 2.
A1
(c) Diagonal partition within a long square duct: See Fig. Ex. 13.11c.
For surface 1: F11 + F 12 + F 13 = 1 (by summation rule)
But, F 11 = 0 (since surface 1 is flat and cannot ‘see’ itself.)
Therefore, F 12 + F 13 = 1
By symmetry: F 13 = F 12 = 0.5 (since radiation emitted by surface 1 is divided equally between surfaces 2 and 3)
A1
By reciprocity: F 21 = ×F 12
A2

670 FUNDAMENTALS OF HEAT AND MASS TRANSFER


2 ×L
i.e. F21 = ×0.5 = 0.71.
L
(d) Sphere inside a cubical box: See Fig. Ex. 13.11d
For surface 1: F 11 = 0 (since surface of the sphere is convex and cannot ‘see’ itself.)
And, F 11 + F 12 = 1 (by summation rule)
Therefore, F12 = 1
A1
By reciprocity F 21 = ×F 12
A2

p ×d2 p
i.e. F21 = = ...since L = d
6 ×d2 6
i.e. F 21 = 0.524.
(e) Hemispherical surface closed by a flat surface: See Fig. Ex. 13.11e.
For surface 1: F 11 + F 12 = 1 (summation rule)
Also, F 21 = 1 (since surface 2 is flat and cannot ‘see’ itself, and all radiation
emitted by surface 2 falls directly on the hemisperical surface 1.)
A2
By reciprocity: F12 = ×F21
A1

p ×r2
i.e. F 12 = ×1= 0.5
2 ×p × r 2
Therefore, F 11 = 1 – F 12 = 0.5.
(f) End and sides of a circular cylinder (L = d): See Fig. Ex. 13.11f.
From the Fig. note that the two end surfaces are denoted by 1 and 3 and the side surface is denoted by 2.
View factor F 13: Surfaces 1 and 3 can be considered as two concentric parallel disks. Therefore, F 13 can be found
out from Fig. 13.20 or by analytical relation given in Table 13.5. Let us use the analytical relation:
We have:
ri rj 1 + Rj2
Ri := Rj = S(Ri , Rj) = 1 +
L L Ri2

LM LM O
F I OP PP
1
2
Rj 2
1
MM MN
Fij (Ri, Rj) := × S ( Ri , R j ) - S ( Ri , R j ) 2 - 4 × GH JK P P (view factor for coaxial parallel disks)
2
MN
Ri
Q PQ
Now, for a cylinder with L = d:
Ri := 0.5 Rj = 0.5
Therefore, Fij (Ri , Rj) = 0.172
i.e. F 13 = 0.172 (view factor from surface 1 to surface 3)
For surface 1: F 11 + F12 + F 13 =1 (by summation rule)
But, F 11 = 0 (since surface 1 is flat and cannot ‘see’ itself.)
Therefore, F 12 := 1 – F 13
i.e. F 12 = 0.828
A
By reciprocity: F 21 = 1 ×F 12
A2

p ×d2
i.e. F 21 = 4 ×0.828
p × d× L

p × d2
i.e. F 21 = 4 2 × 0.828 (since L = d )
p ×d

RADIATION 671
0.828
i.e. F21 == 0.207
4
Also, by symmetry: F 32 = F12 = 0.828
and, F23 = F21 = 0.207.
Example 13.12. Find out the view factor (F11) of a cavity with respect to itself. Hence, find out the view factor F11 for the
following:
(a) a cylindrical cavity of diameter ‘d’ and depth ‘h’
(b) a conical cavity of diameter ‘d’ and depth ‘h’
(c) a hemispherical bowl of diameter ‘d’.
Solution. See Fig. Example 13.12

2 d
2

2 d
1
1 h h 2
1 1
2a

d
(a) Cavity (1) closed by a (b) Cylindrical cavity (c) Conical cavity (d) Hemispherical bowl
hypothetical flat surface (2)

FIGURE Example 13.12 View factors for cavities

View factor of a general cavity w.r.t. itself:


See Fig. Example 13.12a.
We desire to find F11. It is obvious from the Fig. Example 13.12 that part of the radiation emitted by the cavity
surface 1, falls on itself and therefore, F 11 exists.
Close the opening (or mouth) of the cavity by a hypothetical flat surface 2. Then, surfaces 1 and 2 together form an
enclosure. We can write:
For surface 1: F 11 + F 12 = 1 ((a)...by summation rule)
For surface 2: F 21 + F 22 = 1 (by summation rule)
But, F 22 = 0 (since surface 2 is flat and cannot ‘see’ itself.)
Therefore, F 21 = 1
Further, A1 ×F 12 = A2 ×F 21 (by reciprocity)
A2
i.e. F 12 =
A1
Now, F 11 = 1 – F 12 (from Eq. a)
A
i.e. F 11 = 1– 2 (Eq. b)
A1
Eq. b is an important result, since it gives the shape factor of any general cavity w.r.t. itself.
Now, this result will be applied to following specific cavities:
(a) F 11 for a cylindrical cavity of diameter ‘d’ and depth ‘h’: See Fig. Example 13.12b.
A
We have: F 11 = 1 – 2
A1

p ×d2
i.e. F 11 = 1 – 4 (Note that A 1 consists of the area of bottom circular surface and the cylindrical side surfaces)
p ×d2
+ p ×d× h
4

d
= 1–
d + 4× h

672 FUNDAMENTALS OF HEAT AND MASS TRANSFER


4×h
i.e. F 11 = .
4× h + d
(b) F 11 for a conical cavity of diameter ‘d’ and depth ‘h’: See Fig. Example 13.12c.
A2
We have: F 11 = 1 –
A1

p × d2
i.e. F11 = 1– 4 (where, L is the slant height of the cone)
p ×d ×L
2
d
i.e. F 11 = 1 –
2× L
i.e. F 11 = 1 – sin (a) (where, a is the half-vertex angle of the cone.)
Alternatively:
To get F 11 in terms of depth ‘h’, we write:
d
We have: F 11 = 1 –
2× L

d
i.e. F 11 = 1 –
d2
2 × h2 +
4
d
i.e. F 11 = 1 –
4 × h2 + d 2
(c) F 11 for a hemispherical bowl of diameter ‘d’: See Fig. Example 13.12d.
A2
We have: F 11 = 1 –
A1

p ×d2
i.e. F 11 = 1– 42
p ×d
2
1
i.e. F 11 = 1 –
2
i.e. F11 = 0.5.
This result means that for any hemispherical cavity, half of the radiation emitted by the surface 1 falls on itself; it
also means that the remaining half falls on the closing surface 2.

13.6.4 By Graphical Techniques dA2


In some cases, it is possible to get view factors for some geometries f2
by some simple graphical construction. Let us illustrate the princi- n1 dw1
ple of this method as follows: (See Fig. 13.23).
n2
dA 1 and dA 2 are two differential areas at a distance r as shown. f 1 dA3
and f 2 are the angles made by the normals to dA 1 and dA 2 with the f1
line connecting dA 1 and dA 2.
r=1 Hemisphere of
Now, let us make a graphical construction as follows: Con- radius = unity
struct a hemisphere of radius equal to unity with dA 1 as the centre,
and project the element dA 2 on the surface of this hemisphere; this dA1 dA4
projection is shown as dA3 in the Fig. 13.23 From geometry, we
know that dA3 = dA2 .cos (f 2)/r2. Next, project dA3 on the tangential FIGURE 13.23 Graphical determination
plane drawn through dA1, i.e. on the base of the hemisphere. This of view factor between two differential
projection is dA4 in the figure above. Again, dA4 is equal to dA3 areas dA1 and dA2
multiplied by the cosine of the angle f 1 formed between the two
projections. Thus, we have:
RADIATION 673
cos(f 1) × cos(f 2)
dA 4 = dA 2 ×
r2
Now, base of the hemisphere is a circle of unity radius, whose area is equal to p. Therefore, area dA4 divided
by the area of circle of unity radius is:
dA4 cos(f 1) × cos(f 2)
= dA 2 ×
p p ×r 2
Now, recollect that we have already proved the view factor between two differential areas dA 1 and dA2 to
be:
cos(f 1) × cos(f 2) × dA2
FdA1 _ dA2 = ...(13.31)
p ×r 2
i.e. from the above two expressions, it is clear that view factor from dA 1 to dA 2 is given as the ratio of two areas
viz. dA 4 to p, where dA 4 is the projected area of dA3 on the base of the hemisphere and dA3 is the projection of dA2
on the surface of the hemisphere of unity radius.
If the view factor is desired from a differential area dA1 to a
2 finite area A2 (instead of to a differential area dA2), the same proce-
dure is followed: area A2 is projected over the surface of the hemi-
L2
C sphere of unity radius, and the resulting area is further projected on
D
the base of the hemisphere, and then the view factor is computed as
L3 L5 L6 the ratio of this projected area to p.
L4
Many graphical and optical integrators have been developed to
L1 find out view factors between two surfaces, based on this principle.
A B
However, above method is difficult to apply for the case of
complex geometries. In such cases, experimental techniques have
1 been adopted with success, using scale models, the underlying prin-
ciple being: ‘view factors of geometrically similar systems are iden-
FIGURE 13.24 Crossed strings method tical’.
to determine view factor between two Hottel’s crossed strings method This is an extremely simple
infinitely long surfaces method to find out the view factors between infinitely long sur-
faces; generally, channels and ducts which have a constant cross
section and are very long can be modelled as two-dimensional and infinitely long. Consider Fig. 13.24:
A, B and C, D are the end points of two surfaces 1 and 2. These are connected by tightly stretched strings as
shown. Then, the view factor F12 between surfaces 1 and 2 is given by:
( L5 + L6 ) - ( L3 + L4 )
F 12 =
2 × L1
S ( Crossed strings) - S ( Uncrossed strings)
i.e. F 12 = ...(13.46)
2 ×(string on surface 1)
Note that this method can be applied even when the two surfaces 1 and 2 have a common edge (as in the
case of a triangle); then, the common edge is treated as an imaginary string of zero length. Also, note that sur-
faces 1 and 2 may be curved surfaces, but L 1 and L 2 are the straight lengths connecting the edges of the respective
surfaces.
Table 13.4 gives view factors for a few two-dimensional geometries.
Example 13.13. Find out the view factors F12 and F21 for two infinitely long, parallel planes whose centre lines are on the
same vertical line, as shown in Fig. Example 13.13. Plate 1 is 1 m wide, plate 2 is 0.5 m wide and they are spaced 0.6 m
apart.
Solution.
Data:
L 1 := 1 m L2 := 0.5 m S := 0.6 m
To apply crossed strings method, we calculate L 3, L 4 L5 and L 6:

674 FUNDAMENTALS OF HEAT AND MASS TRANSFER


FG L - L IJ 2
L2 = 0.5 m
2
L3 := S2 +
H 2 K
1 2
C D

i.e. L 3 = 0.65 m L3 L5 L6 L4
and, L4 := L 3 (by symmetry) S = 0.6 m

L5 := S 2 + 0.752
i.e. L 5 = 0.96 m A L1 = 1 m B
and, L6 := L 5 (by symmetry) 1
Now, we have:
S (Crossed strings) - S (Uncrossed strings) FIGURE Example 13.13 Crossed strings
F12 = ...(13.46) method to determine view factor between
2 × (string on surface 1)
two infinitely long surfaces

(L5 + L6 ) - (L3 + L4 )
i.e. F 12 :=
2 × L1
i.e. F 12 = 0.31 (view factor from surface 1 to surface 2)
(L5 + L6 ) - (L3 + L4 )
Similarly, F 21 :=
2 × L2
i.e. F 21 = 0.621 (view factor from surface 2 to surface 1.)
Alternatively:
We an use the ready formula given in Table 13.4,
1 1
[(Wi + Wj ) 2 + 4] 2 - [(Wj - Wi ) 2 + 4] 2
Fij = (parallel plates with midlines connected by perpendicular.)
2 ×Wi
In the above formula, notations are with reference to Fig. 13.17a. In the present case, according to the notation of
Fig. Example 13.13, i stands for plate 2 and j stands for plate 1, and the spacing L stands for S.
i.e. wi := L 2
wj := L1
L := S
wi
Wi := i.e. Wi = 0.833
L
wi
Wj := i.e. Wj = 1.667
L
1 1
[(Wi + Wj ) 2 + 4] 2 - [(Wj - Wi ) 2 + 4] 2
Then, F 21 :=
2 ×Wi
i.e. F 21 = 0.621 (view factor from surface 2 to surface 1...same as obtained earlier.)

13.7 Radiation Heat Exchange between Grey Surfaces


So far, we considered radiation heat exchange between black bodies. This was relatively simple since a black
body absorbs all the energy falling on it and none is reflected. However, in the case of grey bodies, absorptivity
is less than unity and the effect of multiple reflections has to be taken into account, and this makes the analysis
more complex.
Generally, there are three methods to deal with the problem of radiation heat exchange between grey bodies:
(i) The reflection method
(ii) The electrical network method, and
(iii) The absorption factor method.
Of these, the reflection method is applied to the simplest of cases, where the number of reflections between
the interacting surfaces is finite, or when the surfaces are black. The electrical network method is applied to cases
of moderate complexity where the number of reflections involved are infinite, but the number of surfaces in-
volved are not more than four or five; this method is very simple, since the standard techniques of solving

RADIATION 675
electrical networks are applied to solve the equivalent thermal networks. The absorption factor method is used to
solve radiation problems that can be graded as ‘difficult’; here, the resulting system of linear algebraic equations
have to be solved by the standard mathematical techniques (such as: matrix methods or using standard computer
library programs).
Whatever the method followed, following assumptions are made to simplify the solution:
(i) All the surfaces of the enclosure are opaque (t = 0), diffuse and grey
(ii) Radiative properties such as r, e and a are uniform and independent of direction and frequency
(iii) Irradiation and heat flux leaving each surface are uniform over the surface
(iv) Each surface of the enclosure is isothermal, and
(v) The enclosure is filled with a non-participating medium (such as vacuum or air).
In this book, we shall discuss only the ‘electrical network method’, since it is simple to apply and gives a
physical ‘feel’ of the problem. However, before we proceed with the discussion of electrical network method, we
shall study a special case of radiative heat transfer between small grey bodies.
13.7.1 Radiation Exchange between Small, Grey Surfaces
Let us consider radiative heat exchange between two small, grey bodies, 1 and 2. By ‘small’, we mean that their
size is very small compared to the distance between them. Let the emissivities of surfaces 1 and 2 be e 1 and e 2,
respectively, and their absorptivities be a 1 and a 2, respectively. Since the surfaces are grey (not black), surely we
have to consider the effect of multiple reflections; however, implication of ‘small’ body is that the portion of
radiation emitted by either body that is reflected by the other body is considered to be ‘lost’ in space and does not
return to the originating surface.
Then, we write:
Energy emitted by body 1 and incident on body 2 = F 12 .A 1 .e 1.s.T14
Of this energy, amount absorbed by body 2 = a 2 .F12.A 1.e 1 .s.T14
Therefore, energy transferred from body 1 to body 2:
Q 1 = e 1 ×e 2 ×A 1 ×F 12×s ×T14 (since a 2 = e 2 by Kirchhoffs law)
Similarly, energy transferred from body 2 to body 1 is:
Q 2 = e 1 ×e 2 ×A 2 ×F 21 ×s ×T 24
and, net radiant energy exchange between 1 and 2 is:
Q 12 = e 1 ×e 2 ×A 1 ×F 12 ×s×(T14 – T24) = e 1 ×e 2 ×A 2 ×F21 ×s×(T14 – T24) ...(13.47)
since A 1 ×F 12 = A2 ×F21 (by reciprocity)
The product, (e 1 .e 2) is known as ‘equivalent emissivity (e eq)’ for a system of two ‘small’ grey bodies.
13.7.2 The Electrical Network Method
This method, introduced by Oppenheim in the 1950s, is simple and direct; it emphasises on the physics of the
problem, and is easy to apply. Before we introduce this method, let us define two new quantities, namely irradia-
tion and radiosity: (See Fig. 13.25).
Irradiation, (G ) is the total radiation incident upon a surface per unit time, per unit area (W/m2).
Radiosity, (J ) is the total radiation leaving a surface, with no regard for its origin (i.e. reflected plus emitted
from the surface) per unit time, per unit area (W/m2).
Now, from Fig. 13.25, it is clear that total radiation leaving the surface (i.e. radiosity, J) is:
J = r×G + e×E b
For a grey, opaque (t = 0) surface, we have:
Radiosity, J r = (1 – a) = (1 – e) (from Kirchhoff’s law)
Therefore,
Incident: G Refleted: r.G Emitted: e.Eb J = (1 – e)×G + e×Eb
( J - e × Eb )
or, G=
(1 - e )
Now, net rate of radiation energy transfer from the surface is
Surface given by: (rate of radiation energy leaving the surface minus the
rate of radiation energy incident on the surface), i.e.
FIGURE 13.25 Irradiation
and radiosity Q
= J–G
A
676 FUNDAMENTALS OF HEAT AND MASS TRANSFER
Q F J - e ×E I
i.e.
A
=J– GH 1 - e JK b

Therefore,
F e × A I ×(E
Q= GH 1 - e JK b – J)

Eb - J
i.e. Q= , W. ...(13.48)
(1 - e )
A ×e
By analogy with Ohm’s law, we can think of Q in Eq. 13.48 as a current flowing through a potential differ-
ence (Eb– J), and the factor (1 – e)/A.e as the resistance. Now, this resistance is the resistance to the flow of radiant
heat due to the nature of the surface and is known as ‘surface resistance (R)’.
i.e.
(1 - e )
R= (surface resistance)
A ×e
Surface resistance for a surface i is shown schematically in Fig. 13.26a.

Ebj

Surface j
Rj

Qi Rij = 1/(Ai.Fij)
Ji
Ebi Ji Ri
Ri = (1 – ei)/(Ai.ei) Ebi
Surface i Surface i
(a) Surface resistance (Ri) (b) Space resistance (Rij)
FIGURE 13.26 Surface resistance and space resistance

From Eq. 13.48, we see that if Ebi > Ji, i.e. if the emissive power is greater than the radiosity, then Qi will be
positive, which means that the net heat transfer is from the surface i. On the other hand, if Ebi < Ji, i.e. if the
emissive power is less than the radiosity, then Qi will be negative, and this means that the net heat transfer is to
the surface i.
For a black body emissivity e = 1; so, the surface resistance is zero, and
Ji = Ebi (for a black body...(13.49))
Also, many surfaces in numerous applications are adiabatic, i.e. well insulated, and net heat transfer
through such a surface is zero, since in steady state, all the heat incident on such a surface is re-radiated. These
are known as re-radiating surfaces. Walls of a furnace is the familiar example of a re-radiating surface. Obvi-
ously, for a re-radiating surface, Qi = 0, and from Eq. 13.48 we get:
Ji = Ebi = s ×Ti4 (for a re-radiating surface...(13.50))
Note that the temperature of a re-radiating surface can be calculated from the above equation; further, note
that this temperature is independent of the emissivity of the surface.
Again, consider two diffuse, grey and opaque surfaces i and j, maintained at uniform temperatures Ti and Tj,
exchanging heat with each other. Then, remembering the definitions of radiosity and view factor, we can write
for the radiation leaving surface i that strikes surface j:
Q i = Ai ×Fij ×Ji

RADIATION 677
Similarly, for surface j, we have:
Qj = Aj ×Fji×Jj
Therefore, net heat interchange between surfaces i and j is:
Qij = Ai ×Fij ×Ji – Aj ×Fji ×Jj
i.e. Q ij = Ai ×Fij ×(Ji – Jj) W ...(13.51)
since Ai ×Fij = Aj×Fji (by reciprocity)
Ji - J j
i.e. Qij = W. ...(13.52)
1
( Ai × Fij )
Again, by analogy with Ohm’s law, we can write Eq. 13.52 as:
Ji - J j
Q ij = W
Rij

1
where, Rij = ...(13.53)
Ai × Fij
Rij is known as ‘space resistance’ and it represents the resistance to radiative heat flow between the radiosity
potentials of the two surfaces, due to their relative orientation and spacing.
Space resistance is illustrated in Fig. 13.26b. Note from Eq. 13.52 that if Ji > Jj, net heat transfer is from surface
i to surface j; otherwise, the net heat transfer is from surface j to surface i.
Thus, for each diffuse, grey, opaque surface, in radiant heat exchange with other surfaces of an enclosure,
there are two resistances, i.e. the surface resistance, Ri = (1 – e i)/(Ai.e i), and a space resistance, Rij = 1/(Ai.Fij).
For a N surface enclosure, net heat transfer from surface i should be equal to the sum of net heat transfers
from that surface to the remaining surfaces, i.e.
N N N
Ji - J j
Qi = å Qij = å Ai × Fij × ( J i - J j ) = å Rij
W ...(13.54)
j =1 j =1 j=1

N
Ebi - J i Ji - J j
i.e.
Ri
= å Rij
W ...(13.55)
j=1

where, R i is the surface resistance and Rij is the space resistance.


This situation is shown in Fig. 13.27.
As can be seen from Fig. 13.27 rate of radiation ‘current’ flow to surface i through its surface resistance must
be equal to the sum of all the radiation current flows from surface i to all other surfaces through the respective
space resistances.
In general, there are two types of radiation problems: first (and most common), when the surface tempera-
ture Ti, and therefore, the emissive power Ebi is known; and, the second type is when the net radiation heat
transfer at the surface i is known. Eq. 13.55 is useful in solv-
J1
ing the first type of problems, i.e. when the surface tempera-
Ri1 ture is known; instead, if the net heat transfer rate at the
Qi Ri2 J2 surface is the known quantity, Eq. 13.52 is the applicable
Ji equation. Essentially, the problem is to solve for the
Ri(N–1)
Ebi JN–1 radiosities J 1, J2, ..., Jn. As mentioned earlier, electrical net-
Ri work method is convenient to use if the number of surfaces
RiN in an enclosure is limited to about five; however, if the
number of surfaces is more than five, the direct approach is
Surface i JN
to apply Eq. 13.55 for each surface whose temperature is
known, and Eq. 13.52 for each surface at which the net heat
FIGURE 13.27 Radiation heat transfer from
transfer rate is known, and solve the resulting set of N linear,
surface i to other surfaces in a N-surface
algebraic equations for the N unknowns, namely, J 1, J 2, ... , Jn
enclosure

678 FUNDAMENTALS OF HEAT AND MASS TRANSFER


by standard mathematical methods. Once the radiosities Surface 2,
are known, Eq. 13.48 may be applied to determine either A2, e2, T2
the heat transfer rate or the temperature, as the case may
Q1
be.
R1 = (1 – e1)/A1.e1)
13.7.3 Radiation Heat Exchange in
R2 = (1 – e2)/A2.e2)
Two-zone Enclosures
Surface 1, R
12 = 1/(A .F )
1 12
Now, with the background of above discussion on the A1, e1, T1
surface resistance and space resistance in connection
with diffuse, grey, opaque surfaces, let us consider the Q1 Q12 Q2
radiant heat transfer in a two-zone enclosure. This sim- Eb1 J1 J2 Eb2
ply means that the two surfaces, together, make up the
enclosure and ‘see’ only themselves and nothing else. R1 R12 R2
Many, practically important geometries may be classi-
fied as two-zone enclosures, e.g. a small body enclosed FIGURE 13.28 Two-surface enclosure and its
by a large body, a pipe passing through a large room, radiation network
concentric spheres, concentric long cylinders, long, par-
allel plane surfaces, etc.
Fig. 13.28 shows a schematic of a typical two-zone enclosure and the associated radiation (or, thermal) net-
work.
Surfaces 1 and 2 forming the enclosure are diffuse, grey and opaque. Let their emissivities, temperatures and
areas be (e 1, T1, A1) and (e 2, T2, A2), respectively. The radiation network is shown in Fig. 13.28. Each surface has
one surface resistance associated with it and there is one space resistance between the two radiosity potentials,
and all the resistances are in series, as shown. The ‘heat current’ (Q 12) in this circuit is calculated by dividing the
‘total potential’ (Eb1 – Eb2) by the ‘total resistance’ (R 1 + R 12 + R 2). So, we write:
Eb1 - Eb 2
Q 12 = Q 1 = Q 2 = Q12
R1 + R12 + R2 Eb1 = J1 Eb2 = J2

Eb1 - Eb 2
i.e. Q 12 = R12 = 1/(A1.F12)
1 - e1 1 1 - e2
+ +
A1 × e 1 A1 × F12 A2 × e 2 FIGURE 13.29 Radiation network for
two black surfaces forming an enclosure
s ×(T14 - T24 )
i.e. Q 12 = W. ...(13.56)
1 - e1 1 1 - e2
+ +
A1 × e 1 A1 × F12 A2 × e 2
Eq. 13.56 is an important equation, which gives net rate of heat transfer between two grey, diffuse, opaque
surfaces which form an enclosure, i.e. which ‘see’ only each other and nothing else.
Now, let us consider a few special cases of two-surface enclosure. Basic radiation network for all these cases
is the same as given in Fig. 13.28 and the basic, governing equation is Eq. 13.56, which is modified depending
upon the case considered.
Case (i): Radiant heat exchange between two black surfaces:
For a black body, e = 1, and J = Eb, as explained earlier. i.e. surface resistance [= (1 – e)/(A.e)] of a black body is
zero. Then, the radiation network will consist of only a space resistance between the two radiosity potentials, as
shown in Fig. 13.29:
Then, from Eq. 13.56, we get:

s × (T14 - T24 )
Q 12 =
1
A1 × F12
i.e. Q 12 = A 1 ×F 12 ×s×(T14 – T24), W (for two black surfaces forming an enclosure...(13.57))
Next, we shall consider four cases of practical interest where the view factor between the inner surface 1 and
the outer surface 2 (i.e. F 12) is equal to 1, and also the net radiation from a grey cavity.

RADIATION 679
Case (ii): Radiant heat exchange for a small object in a large cavity:
See Fig. 13.30 (a). A practical example of a small object in a large cavity is the case of a steam pipe passing
through a large plant room.
For this case, we have:
A1
=0
A2
and, F 12 = 1
And, Eq. 13.56 becomes:
Q12 = A 1 ×s×e 1 ×(T14 – T24 ) (for small object in a large cavity...(13.58))
Case (iii): Radiant heat exchange between infinitely large parallel plates:
See Fig. 13.30 (b). In this case, A 1 = A 2 = A, say, and F12 = 1
Then, Eq. 13.56 becomes:

A ×s × (T14 - T24 )
Q 12 = (for infinitely large parallel plates...(13.59))
1 1
+ -1
e1 e 2
Case (iv): Radiant heat exchange between infinitely long concentric cylinders:
See Fig. 13.30 (c). In this case:
F 12 = 1
Then, Eq. 13.56 becomes:

A1 ×s ×(T14 - T24 )
Q 12 =
1 A1 FG IJ FG
1 IJ (for infinitely long concentric cylinders...(13.60))

e1
+
A2
×
H KH
e2
-1
K
where,
A1 r
= 1
A2 r2
Remember that A1 refers to the inner (or enclosed) surface.
Eq. 13.60 is known as ‘Christiansen’s equation’.
Case (v): Radiant heat exchange between concentric spheres:
See Fig. 13.30 (d). In this case:
F 12 = 1
Then, Eq. 13.56 becomes

A1 ×s ×(T14 - T24 )
Q 12 =
1 A1 FG IJ FG
1 IJ (for concentric spheres...(13.61))

e1
+
A2
×
H KH
e2
-1
K
where,

A1 Fr I 2

A2
= GH r JK
1
2
Remember, again, that A1 refers to the inner (or enclosed) surface.
Case (vi): Energy radiated from a grey cavity:
Consider a grey cavity as shown in Fig. 13.31. Let e 1, A1 and T1 be its emissivity, area and temperature (in
Kelvin), respectively. Now, energy will stream out of the cavity into the surrounding space through the opening
(or, mouth) of the cavity. Let the opening be covered by an imaginary surface A 2. Thus, it is a two-surface
enclosure. Now, since the cavity is very small compared to the space outside, practically all the energy emitted
by the cavity will be absorbed by space, and it is reasonable to assume that radiation coming to the cavity from
space is negligible, i.e. the space acts like a black body at a temperature of zero Kelvin as far as the cavity is

680 FUNDAMENTALS OF HEAT AND MASS TRANSFER


A2, T2, e2 A1, T1, e1

A1, T1, e1

A2, T2, e2

(a) Small object in a large cavity (b) Infinitely large parallel planes

r1 r2 r2

r1

(c) Infinitely long concentric cylinders (d) Concentric spheres

FIGURE 13.30 Few two-surface enclosures where F12 = 1

A2 Q12
Eb1 J1 J2 = Eb2 = 0

A1, T1, e1 R1 R12

R1 = (1 – e1)/(A1.e1)
R12 = 1/(A1.F12)
(a) Grey cavity (b) Radiation network

FIGURE 13.31 Radiation from a grey cavity

concerned. So, surface 2 is black at zero Kelvin for our analysis. Implication of this is that surface resistance of
surface 2 is zero, and radiosity of surface 2 is equal to its emissive power, which in turn, is equal to zero since the
temperature is zero Kelvin. So, the radiation network for this case will be as shown in Fig. 13.31:
Therefore, net energy radiated from the grey cavity is given by:
Eb 1 - Eb 2
Q 12 =
R1 + R12
Eb1 - 0
i.e. Q 12 = (since Eb2 = 0 at T 2 = zero Kelvin)
R1 + R12

s ×T14
i.e. Q 12 =
1 - e1 1
+
e 1 × A1 A1 × F12
Now, F11 + F12 = 1 (by summation rule)
i.e. F 12 = 1 – F11
Then, Q 12 becomes:

s × T14
Q 12 =
1 - e1 1
+
e 1 × A1 A1 ×(1 - F11 )

RADIATION 681
A1 × e 1 ×s × T14 ×(1 - F11)
i.e. Q 12 =
(1 - e 1 ) ×(1 - F11 ) + e 1

A1 × e 1 ×s ×T14 ×(1 - F11 )


i.e. Q 12 =
1 - e 1 - F11 + e 1 × F11 + e 1

i.e. Q 12 = A 1 ×e 1 ×s ×T14 ×
LM 1 - F OP W
11
(net radiation from grey...(13.62))
N 1 - (1 - e )× F Q
1 11

Eq. 13.62 is an important result, which gives net radiation from a


A2 grey cavity to surrounding space. If it is desired, for example, to calcu-
A1,e1,T1 late the net radiation from a blind hole drilled in a flange, then the
A1 << A2 relation to use is the Eq. 13.62.
Q12 F12 = 1 Example 13.14. A long pipe, 50 mm in diameter, passes through a room and
is exposed to air at 20 deg.C. Pipe surface temperature is 93 deg.C. Emissiv-
ity of the surface is 0.6. Calculate the net radiant heat loss per metre length
of pipe.
(M.U. 1991)
FIGURE Example 13.14 Two-surface
Solution. The pipe is enclosed by the room; so, it is two-surface enclosure
enclosure with A1 << A2 problem. Further, area of the pipe is very small, compared to the area of the
room. Therefore, this is a case of a small object surrounded by a large area,
and we have:
A1
= 0
A2
and, F 12 = 1
\ Q 12 = A1 ×s×e1 ×(T14 – T24) (for small object in a large cavity...(13.58))
Data:
d 1 := 0.05 m L := 1 m e1 := 0.6 T1 := 93 + 273 K T2 := 20 + 273 K
s := 5.67 ´ 10 –8 W/(m2 K) (Stefan–Boltzmann constant)
Now, A1 := p×d1×L
i.e. A1 = 0.157 m2 (surface area of the pipe per metre length)
Then, applying Eq. 13.58, we get:
Q12 := A1×s×e1×(T14 – T24)
i.e. Q12 = 56.507 W (net radiant heat loss from the pipe per metre length.)

T1 = 1073 K Example 13.15. Calculate the net radiant heat interchange per m2 for two large par-
e1 = 0.3 allel plates maintained at 800°C and 300°C. The emissivities of two plates are 0.3 and
0.6, respectively. (M.U. 1993)
Solution. The plates are parallel to each other, and are very large; so, it is a two-
T2 = 573 K surface enclosure problem, with two infinite parallel plates. We have:
Data:
e1 = 0.6
T 1 := 800 + 273 K T2 := 300 + 273 K e1 := 0.3 e2 := 0.6
s := 5.67 ´ 10 –8 W/(m2 K) (Stefan–Boltzmann constant)
FIGURE Example 13.15 Two A := 1 m2 (area of the surface)
infinitely large parallel plates For infinite parallel plates, we have the relation:

A ×s × (T14 - T24 )
Q 12 := W (for infinitely large parallel plates ...(13.59))
1 1
+ -1
e1 e2
i.e. Q 12 = 1.726 ´ 104 W/m2 (radiant heat transfer per m 2 of the plates.)
Example 13.16. A spherical liquid oxygen tank, 0.3 m in diameter is enclosed concentrically in a spherical container of
0.4 m diameter and the space in between is evacuated. The tank surface is at – 183°C and has an emissivity of 0.2. The
container surface is at 25°C and has an emissivity of 0.25. Determine the net radiant heat transfer rate. (M.U.)

682 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Solution. This is the case of two surface enclosure, with one sphere enclosed by another sphere. If 1 denotes inner
sphere, and 2 outer sphere, we have for view factors: F 11 = 0 and F 12 = 1
Data:
T1 := 183 + 273 K T2 := 25 + 273 K e1 := 0.2 e2 := 0.25 r1 := 0.15 m r2 := 0.2 m
s := 5.67 ´ 10 –8 W/(m2K) (Stefan–Boltzmann constant) A1 := 4×p× r12 i.e. A 1 = 0.283 m2 (area of inner surface)
A2 := 4 ×p× r22 i.e. A 2 = 0.503 m2 (area of outer surface)
Now, for the case of concentric spheres, we have:
A1 × s × (T14 - T24 )
Q 12 :=
A1 F IF I (for concentric spheres...(13.61))
1
e1
+
A2
×GH JK GH
1
e2
-1 JK
Therefore,
Q 12 = – 18.748 W (net radiant heat interchange between inner and outer spheres.)

T1 = 680 K Note: Negative sign indicates that heat flows from outer sphere to
e1 = 0.35 inner sphere; this is certainly so, since the inner sphere is at a lower
temperature than the outer sphere.
A1 = 4 m2
Example 13.17. A convex grey body having a surface area of 4 m2
T2 = 310 K has e 1 = 0.35 and T1 = 680 K. This is completely enclosed by a grey
e1 = 0.75 surface having an area of 36 m2, e 2 = 0.75 and T2 = 310 K. Find the
A2 = 36 m2 net rate of heat transfer Q 12 between the two surfaces.
(M.U. 1999)
Solution. This is the case of a two surface enclosure. Inner surface is
Eb1 Eb2 convex; so, view factor F11 = 0. Also, F12 = 1 since the inner body is
completely enclosed by the outer surface.
R1 R12 R2 The radiation network for this problem is shown in Fig. Exam-
ple 13.22 below:
Data:
FIGURE Example 13.17 Radiation net-work T 1 := 680 K T2 := 310 K e1 := 0.35 e2 := 0.75
for a convex grey body completely enclosed A1 := 4 m2 A2 := 36 m2
by another grey body s := 5.67 ´ 10 –8 W/(m2 K) (Stefan–Boltzmann constant)

F 12 := 1 (since all the heat radiation emitted by surface 1 is intercepted by surface 2.)
Eb1 := s×T14 i.e. Eb1 = 1.212 ´ 104 W/m2
Eb2 := s×T24 i.e. Eb2 = 523.636 W/m2
1 - e1
Now, R 1 :=
e 1 × A1
i.e. R 1 = 0.464 m–2 (surface resistance of inner surface)
1- e2
and, R2 :=
e 2 × A2
i.e. R 2 = 9.259 ´ 10 – 3 m– 2 (surface resistance of outer surface)
1
Also, R12 :=
A1 × F12
i.e. R 12 = 0.25 m– 2 (space resistance between inner and outer surface)
Therefore,
Rtot := R1 + R 12 + R 2
i.e. R tot = 0.724 m– 2 (total resistance between inner and outer surface)
Then, net rate of heat transfer between surfaces 1 and 2 is given by:
Eb1 - Eb 2
Q 12 :=
Rtot
i.e. Q 12 = 1.603 ´ 104 Watts.
Alternatively:
We can apply the direct formula for a two surface enclosure, for which F 11 = 0, F 12 = 1, i.e.

RADIATION 683
A1 × s × (T14 - T24 )
Q 12 := (for convex surface enclosed by another surface...(13.61))
1 A1FG IJ FG
1 IJ
e1
+
A2 H KH
×
e2
-1
K
i.e. Q 12 = 1.603 ´ 104 Watt.
Example 13.18. A hemispherical furnace of radius 1.0 m has a roof temperature of T1 = 800 K and emissivity e1 = 0.8. The
flat circular floor of the furnace has a temperature of T2 = 600 K and emissivity e2 = 0.5. Calculate the net radiant heat
exchange between the roof and the floor. (M.U. 1998)

T1 = 800 K Solution. This is a two-zone enclosure problem. Fig. Exam-


e1 = 0.8 ple 13.18 shows the radiation network for this problem. We
A1 = 6.28 m2 have:
Eb1 - Eb 2
R=1m Q 12 =
T2 = 600 K R1 + R12 + R2
e2 = 0.5
where, R 1 and R 2 are the two-surface resistances and R 12 is
A2 = 3.14 m2 the space resistance between the two radiosity potentials.
Eb1 Eb1 Data:
T 1 := 800 K T 2 := 600 K e 1 := 0.8
R1 R12 R2 e 2 := 0.5 R := 1 m
4 ×p × R2
A1 := m2 (area of hemispherical surface 1)
FIGURE Example 13.18 Radiation network for heat 2
transfer between a hemispherical furnace and its floor i.e. A 1 = 6.283 m 2 (area of surface 1)

A 2 := p×R2 m2 (area of surface 2)


i.e. A2 = 3.142 m2 (area of surface 2)
s := 5.67 ´ 10 –8 W/(m2K) (Stefan–Boltzmann constant)
View factors:
F 21 := 1 (since all the heat radiated by surface 2 is intercepted by hemispherical surface 1.)
Now, A1 ×F12 = A2×F 21 (by reciprocity)
A2 × F21
Therefore, F 12 :=
A1
i.e. F 12 = 0.5 (view factor from surface 1 to surface 2)
Resistances:
1 - e1
Now, R1 :=
e 1 × A1
i.e. R 1 = 0.0398 m–2 (surface resistance of inner surface)
1- e2
and, R2 :=
e 2 × A2
i.e. R 2 = 0.318 m–2 (surface resistance of outer surface)
1
Also, R12 :=
A1 × F12
i.e. R 12 = 0.318 m–2 (space resistance between inner and outer surface)
Therefore,
Rtot := R1 + R 12 + R 2
i.e. R tot = 0.676 m–2 (total resistance between inner and outer surface)
Also,
Eb1 := s×T14 Eb1 = 2.322 ´ 104 W/m2
Eb2 := s×T24 Eb2 = 7.348 ´ 103 W/m2
Then, net rate of heat transfer between surface 1 and 2 is given by:
Eb1 - Eb 2
Q 12 :=
Rtot

684 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Q 12 = 2.347 ´ 104 Watt.
Additionally:
If both the surface are black Now, both the surface resistances become zero, since e = 1 for both the black surfaces.
Then,
Eb1 - Eb 2
Q 12 :=
R12
i.e. Q 12 = 4.988 ´ 104 W (if both the surfaces are black.)
Example 13.19. Refer to Fig. Example 13.19. Three thin-walled,
d1 = 0.15 m, T1 = 80 K,
long, circular cylinders 1, 2 and 3, of diameters 15 cm, 25 cm and 35 3
cm, respectively, are arranged concentrically as shown. Tempera- e1 = 0.05
2 d2 = 0.25 m, T2 = ?
ture of cylinder 1 is 80 K and that of cylinder 3 is 300 K. 1
Emissivities of cylinders 1, 2 and 3 are 0.05, 0.1 and 0.2, respec- e2 = 0.1
tively. Assuming that there is vacuum inside the annular spaces, d3 = 0.35 m, T3 = 300 K,
determine the steady state temperature attained by cylinder 2. e3 = 0.2
Solution. This is the case of radiant heat transfer between long,
concentric cylinders.
Data:
T 1 := 80 K T3 := 300 K e1 := 0.05 e2 := 0.10 e3 :=
0.20 d1 := 0.15 m d2 := 0.25 m FIGURE Example 13.19 Three concentric
d 3 := 0.35 m s := 5.67 ´ 10–8 W/(m2K) cylinders
Let L be the length of cylinders. Then,
A1 p × d1 × L d
= = 1
A2 p × d2 × L d2
and,
A2 p × d2 × L d
= = 2
A3 p × d3 × L d3
In steady state, net radiant heat transfer between cylinders 1 and 2 must be equal to the net radiant heat transfer between
cylinder 2 and 3.
i.e. Q12 := Q23 ...(A)
We apply Eq. 13.60, namely,
A1 ×s × (T14 - T24 )
Q 12 =
1 FG IJ FG
A1 1 IJ (for infinitely long, concentric cylinders...(13.60))

e1
+
H KH
A2
×
e2
-1
K
Applying Eq. 13.60 in heat balance Eq. A:
A1 ×s × (T14 - T24 ) A2 ×s × (T24 - T34 )
F IF
A1 I =
F IF
A2 I ...(B)
1
e1
+ GH JK GH
A2
×
1
e2
-1 JK 1
e2
+ GH JK GH
A3
×
1
e3
-1 JK
In the above equation, T2 is the only unknown. Simplifying Eq. B:

A1 F IF I
A1 × (T14 - T24 )
1
e1
+
A2
×
1
e2
-1GH JK GH JK
A2 × (T24 - T34 )
=
A2 F IF I
1
e2
+
A3
×
1
e3
-1 GH JK GH JK
F IF I
d
(T14 T24 )
1
e1 GH JK GH JK F d I
+ 1 ×
d2
1
e2
-1
A1 d A d
I ×GH d JK
-
1 Fd I F 1
i.e. = 2
(since = 1 and, 2 = 2 )
(T24 - T34 ) A2 d2 A3 d3
+ G J ×G - 1J
1

Hd K He K
2
e 2 3 3

RADIATION 685
(T14 - T24 )
i.e. = 3.293
(T24 - T34 )
i.e. (T14 – T24) = 3.293×T24 – 3.293×T34
i.e. 4.293×T24 = T14 + 3.293 ×T34

FT I
1
4
+ 3. 293 × T34
:= G JK
4
T2 1
i.e.
H 4. 293
i.e. T 2 = 280.864 K (steady state temperature of cylinder 2.)
Alternatively:
We can get value of T 2 very easily by applying solve block of Mathcad.
Start with a guess value for T 2, and write the constraint, i.e. Eq. B immediately after ‘Given’ in the solve block; then,
typing ‘Find(T2) =’ gives the value of T2:
T2 := 200 K ...guess value )
Given
d1 ×s × (T14 - T24 ) d 2 ×s ×(T24 - T34 )
dFG IJ FG IJ =
d F IF I
1
e1 d2
1
H KH
+ 1 ×
e2
-1
K
1
e2 GH JK GH
+ 2 ×
d3
1
e3
-1 JK
Find(T2) = 280.862
i.e. T 2 = 280.862 K (same as obtained earlier.)
Note: While writing the constraint equation in the solve block above, we have substituted d 1/d 2 for A 1/A2,and d 2/d 3 for
A2/A3.
Once again, it is demonstrated that using solve block of Mathad, very much simplifies the solution, and reduces the
labour involved.
Example 13.20. A blind cylindrical hole of diameter 2 cm and length 3 cm is drilled
D = 2 cm into a metal slab having emissivity 0.6. If the metal slab is maintained at a tem-
e1 = 0.6 perature of 350°C, find the heat escaping out of the hole by radiation.
(M.U)
Solution. This is a problem on determining energy esaping from a grey cavity. We
use Eq. 13.62, i.e.
H = 3 cm
Q 12 = A 1 ×e 1 ×s×T14 ×
LM
1 - F11 OP
MN
1 - (1 - e 1 ) × F11
W
PQ
(net radiation from grey cavity)
Data:
D := 0.02 m H := 0.03 m e1 := 0.6 T1 := 350 + 273 K
s := 5.67 ´ 10–8 W/(m2 K) (Stefan-Boltzmann constant)
FIGURE Example 13.20 Grey,
Now, F11 for a cavity is already shown to be:
cylindrical cavity
A
F11 = 1 – 2 (where, A2 = area of closing surface, A 1 = area of the cavity surface )
A1
p ×D2
i.e. F11 = 1 – 4 (for cylindrical cavity of this problem)
p ×D2
+ p ×D× H
4

D 4×H
i.e. F11 = 1 – =
D + 4×H 4× H + D

4× H
Therefore, F 11 :=
4× H + D
i.e. F 11 = 0.857 (view factor of the cavity w.r.t. itself)
p ×D 2
and, A 1 := p×D×H +
4

686 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. A1 = 2.199 ´ 10–3 m2 (area of surface of cylindrical cavity)
Then, from Eq. 13.62:

Q 12 := A 1 ×e 1 ×s× T14 ×
LM 1 - F OP 11

MN 1 - (1 - e )× F PQ
1 11

i.e. Q 12 = 1.063 W (energy escaping from the cavity.)


Example 13.21. A hohlraum is to be constructed out of a thin copper Opening, A2
sphere of diameter = 15 cm. Its internal surface is highly oxidised. What
should be the area of a small opening to be made on the surface of the Sphere, D = 0.15 m
sphere, if the desired absorptivity is 0.95?
Solution.
Data:
A1, e1 = 0.95
D := 0.15 m (diameter of the sphere)
a 1 := 0.95 (absorptivity)
e 1 := a 1 (emissivity (= absorptivity), by Kirchhoff’s law )
The inside surface of the sphere must absorb 95% of the energy,
which means that 5% of the energy escapes out through the opening of FIGURE Example 13.21 Hole on the
area = A2, say. surface of a sphere—holhraum
Let Q 2 = energy escaping through the hole, and
Q1 = energy radiated from the spherical cavity
Then, we have:

Q 2 = A1 ×e 1 ×s×T14 ×
LM 1 - F OP 11
(from Eq. 13.62)
MN 1 - (1 - e )× F PQ
1 11

and, Q 1 = A1 ×e1 ×s×T14


Q2
By data, = 0.05
Q1

LM 1 - F OP = 0.05
11
i.e.
MN 1 - (1 - e )× F PQ
1 11

Solving,
1 – F 11 = 0.05 – (0.05)2 ×F 11
i.e. 1 – 0.05 = F11 ×(1 – 0.0025)
0. 95
i.e. F 11 :=
1 - 0.0025
i.e. F 11 = 0.952
But, we also have, for the view factor of cavity, w.r.t. itself:
A2
F11 = 1 – (where, A2 is the area of the opening, A1 is the area of surface of cavity)
A1
A2
i.e. F 11 = 1 – (where, As = total area of spherical surface)
As - A 2
Now, A s := p×D 2
i.e. As = 0.071 m2 (total area of spherical surface)
Therefore,
F A2 I = 0.952
GH
F 11 = 1 -
As - A2 JK
Solving,
As - A2 - A2
= 0.952
As - A2
i.e. As – 2 ×A2 = 0.952×As – 0.952×A2

RADIATION 687
i.e. A2 ×(2 – 0.952) = As ×(1 – 0.952)
i.e. 1.048×A2 = 3.393 ´ 10 –3
3. 393 ´ 10 - 3
i.e. A 2 :=
1.048
i.e. A2 = 3.238 ´ 10–3 m2 (area of the opening on the surface of sphere)
i.e. A2 = 32.38 cm2 (area of the opening on the surface of sphere.)

13.7.4 Radiation Heat Exchange in Three-zone Enclosures


Fig. 13.32 (a) shows an enclosure made of three opaque, diffuse, grey surfaces. Let the surfaces A 1, A 2, A 3 be
maintained at uniform temperatures of T 1, T 2 and T 3 , respectively. Also, let the emissivities be e 1, e 2 and e 3,
respectively. The radiation network for this system of three-surface enclosure is shown in Fig. 13.32 (b). While
drawing the radiation network, the principle to be followed is quite simple: first, draw the surface resistance
associated with each grey surface; then, connect the radiosity potentials between surfaces by the respective space
resistances.

R12
Q12
Eb1 J1 J2 Eb2
Q1 Q2
R1 R2
Surface 1, Surface 2, Q13 Q23 R1 = (1 – e1)/(A1.e1)
A1, e1, T1 A2, e2, T2 R2 = (1 – e2)/(A2.e2)
R13 J3 R23
R3 = (1 – e3)/(A3.e3)
R3
Surface 3, R12 = 1/(A1.F12)
A3, e3, T3 Eb3 R13 = 1/(A1.F13)
R23 = 1/(A2.F23)
Q3
(a) (b)

FIGURE 13.32 Three-surface enclosure and its radiation network

It is considered that the temperature of each surface is known, i.e. emissive power Eb for each surface is
known. Then, the problem reduces to determining the radiosities J1, J 2 and J3. This is done by applying
Kirchhoff’s law of dc circuits to each node: i.e. sum of the currents (or, rate of heat transfers) entering into each
node is zero. Doing this, we get the following three algebraic equations:
Eb1 - J1 J - J1 J - J1
Node J 1: + 2 + 3 =0 ...(13.63a)
R1 R12 R13
Eb2 - J 2 J - J2 J - J2
Node J 2: + 1 + 3 =0 ...(13.63b)
R2 R12 R23
Eb3 - J 3 J - J3 J - J3
Node J 3: + 1 + 2 =0 ...(13.63c)
R3 R13 R23
Solving these three equations simultaneously, we get J1, J2 and J3.
Remember to write each equation such that current flows into the node; then, the magnitudes of the
radiosities would adjust themselves when all the three equations are solved simultaneously. Once the
magnitudes of the radiosities are known, expressions for net heat flows between the surfaces are:
J1 - J 2 J - J2
Q 12 = = 1 ...(13.64a)
R12 1
A1 × F12

J1 - J 3 J - J3
Q 13 = = 1 ...(13.64b)
R13 1
A1 × F13
688 FUNDAMENTALS OF HEAT AND MASS TRANSFER
J - J3 J - J2
Q 23 = 2 = 1 ...(13.64c)
R23 1
A2 × F23
and, net heat flow from each surface is:
Eb1 - J1 E - J1
= b1
Q1 =
R1 F I ...(13.65a)
GH
1 - e1
A1 × e 1 JK
Eb2 - J 2 E - J2
= b2
Q2 =
R2 F I ...(13.65b)
GH
1- e2
A2 × e 2 JK
Eb3 - J3 E - J3
= b3
Q3 =
R3 F I ...(13.65c)
GH
1 - e3
A3 × e 3 JK
Eq. set 13.64 is a set of general equations for three diffuse, opaque, grey surfaces. However, these equations
will be modified depending upon any constraint that may be attached to any of the surfaces, i.e. say, if the
surface is black or re-radiating: Ji = Ebi = s.Ti4. And, Qi = 0 for a re-radiating surface. If Q i at any surface is
specified instead of the temperature (i.e. Ebi), then, (Ebi – Ji)/Ri is replaced by Qi.
We shall study a few such special cases of three-zone enclosures below:
Case (i): Two black surfaces connected to a third refractory surface:
This is a three-zone enclosure, with two of the surfaces being black and the third surface being a re-radiating,
insulated surface. Typical example is a furnace whose bottom is the ‘source’ and the top is the ‘sink’ and the two
surfaces are connected by a refractory wall which acts as a re-radiating surface. In effect, the source and sink
exchange heat through the re-radiating wall; however, in steady state, the re-radiating wall radiates as much heat
as it receives, which means that net heat exchange through the re-radiating wall (= Q) is zero, i.e. Eb = J for the re-
radiating wall. Therefore, once J (i.e. Eb) is calculated for the re-radiating surface, its steady state temperature can
easily be calculated from: Eb = s.T 4.

Eb1 = J1 R12 Eb2 = J2


Eb1 = J1 R12 Eb2 = J2
R1R R2R R12 = 1/(A1.F12)
R1R = 1/(A1.F1R)
R2R = 1/(A2.F2R) R1R JR R2R
EbR = JR
(a) (b)

FIGURE 13.33 Two black surfaces connected by a third re-radiating surface and its radiation network

Fig. 13.33 (a) shows the radiation network for this case. The radiation network is drawn very easily by
remembering the usual principles: for a black surface, the surface resistance is zero, i.e. Eb = J. For a re-radiating
surface too, Eb = J, as already explained; further, for a re-radiating surface, Q = 0. Between two given surfaces, the
radiosity potentials are connected by the respective space resistances, as shown. It may be observed that the
system reduces to a series–parallel circuit of resistances as shown in Fig. 13.33 (b).
So, we write, for the total resistance of the circuit, R tot :
1 1 1
= +
Rtot R12 ( R1R + R2 R )

and, Q 12 =
Eb1 - Eb 2
= (E b1 – Eb2)×
1
+
LM 1 OP
Rtot R12 N
( R1R + R2 R ) Q
RADIATION 689
LM OP
M
– T )× MA × F +
1 P
Q 12 = s×(T14
F 1 I PP
4
i.e. ...(13.66)
MM
2 1 12

N GH A × F 1 1R
+
1
A2 × F2 R JK PQ
Here, Q 12 is the net radiant heat transferred between surfaces 1 and 2. Similar expressions can be written for
heat transfer between surfaces 2 and 3 (= Q 23) and the heat transfer between surfaces 1 and 3 (= Q 13).
Case (ii): Two grey surfaces surrounded by a third re-radiating surface:
In this case, there are two grey surfaces, and the third surface is an insulated, re-radiating surface. As al-
ready explained, the re-radiating surface radiates as much energy as it receives; therefore, net radiant heat trans-
fer for that surface is zero, i.e.
Q3 = 0
Eb3 - J 3
i.e.
F 1- e I =0
GH A ×e JK
3
3
3
i.e. Eb3 = J3
i.e. once the radiosity of the re-radiating surface is known, its temperature can easily be calculated, since Eb3 =
s.T34. Further, note that T3 is independent of the emissivity of surface 3.
R12
Q12
Eb1 J1 J2 Eb2
Surface 1, Surface 2, Q1 Q2
A1, e1, T1 A2, e2, T2 R1 R
Q13 Q23 2
R1 = (1 – e1)/(A1.e1)
R13 R23 R2 = (1 – e2)/(A2.e2)
Re-radiating surface 3, J3 = Eb3
Q3 = 0 Q3 = 0 R12 = 1/(A1.F12)
R13 = 1/(A1.F13)
R23 = 1/(A2.F23)
(a) (b)

FIGURE 13.34 Two grey surfaces surrounded by a re-radiating surface


Now, the radiation network reduces to a simple series–parallel circuit of the relevant resistances.
Expression for heat flow rate is:
Eb1 - Eb 2
Q 1 = – Q2 =
Rtot
where, R tot is the resistance, given by:

LM OP
R =R + M
MM 1 + 1 PPP + R
1
tot 1 2

N R (R + R ) Q 12 13 23

R = G
F 1- e I + F 1- e I
H A ×e JK LM OP + GH A ×e JK
1 1 2
i.e. tot ...(13.67)
1 1 2 2

MM PP
MMA × F + 1
1 1 12 PP
MM F 1 + 1 IP
N GH A × F A × F 1 13 2 23
JK PQ
690 FUNDAMENTALS OF HEAT AND MASS TRANSFER
13.7.5 Radiation Heat Exchange in Eb2 = J2 R23 Eb3 = J3
Four-zone Enclosures R12 = 1/(A1.F12)
(a) When all the four surfaces are R24 R13 = 1/(A1.F13)
black Remembering the principles already ex-
plained, if the radiation network for an enclosure R12 R43 R14 = 1/(A1.F14)
comprising of four black surfaces is drawn, it will R13 R23 = 1/(A2.F23)
look as shown in Fig. 13.35.
R24 = 1/(A2.F24)
Expression for net radiant heat flow rate from
surface 1 is: R43 = 1/(A4.F43)
Eb1 = J1 R14 Eb4 = J4
J1 - J 2 J - J3 J - J4
Q1 = + 1 + 1
R12 R13 R14
FIGURE 13.35 Radiation network for an enclosure of
E - Eb 2 E - Eb 3 E - Eb 4 four black surfaces
i.e. Q 1 = b1 + b1 + b1
R12 R13 R14
...since Eb1 = J1, etc.
i.e. Q 1 = A 1 ×F 12 (E b1 – Eb2) + A 1 ×F 13 ×(Eb1 – Eb3)
+ A 1 ×F 14 ×(Eb1 – Eb4) ...(13.68)
Similar expressions can be written for the net heat flow from other three surfaces.
(b) When all the four surfaces are grey Now, for each surface, a surface resistance also has to be included, and
the radiation network for this system will be as shown in Fig. 13.36:
Expression for net radiant heat flow rate from surface 1 is:
Eb1 - J1 J - J2 J - J3 J - J4
Q1 = = 1 + 1 + 1
R1 R12 R13 R14

Eb1 - J1
i.e. Q1 =
F1- e I = A 1 ×F 12×(J1 – J2) + A 1 ×F 13 ×(J1 – J3) + A 1 ×F 14 ×(J1 – J4) ...(13.68)
GH A ×e JK
1
1
1
Similar expressions can be written for the net heat flow from other three surfaces.

Eb2 R2 J2 R23 J3 R3 Eb3

R24

R12 R43
R13

Eb1 R1 J1 R14 J4 R4 Eb4

R1 = (1 – e1)/(A1.e1) R12 = 1/(A1.F12) R23 = 1/(A2.F23)


R2 = (1 – e2)/(A2.e2) R13 = 1/(A1.F13) R24 = 1/(A2.F24)
R3 = (1 – e3)/(A3.e3) R14 = 1/(A1.F14) R43 = 1/(A4.F43)
R4 = (1 – e4)/(A4.e4)

FIGURE 13.36 Radiation network for an enclosure of four grey surfaces

Example 13.22. A long duct of equilateral triangular section, of side w = 0.75 m, shown in Fig. Example 13.22, has its
surface 1 at 700 K, surface 2 at 1000 K, and surface 3 is insulated. Further, surface 1 has an emissivity of 0.8 and surface
2 is black. Determine the rate at which energy must be supplied to surface 2 to maintain these operating conditions.
Solution. Since the duct is very long, the ‘end effects’ can be neglected. Therefore, this is a three zone enclosure, with
surface 1 being grey, surface 2 being black, and surface 3 being insulated (or, re-rediating).

RADIATION 691
Eb1 R1 J1 R12 J2 = Eb2
T2 = 1000 K, Insulated Q1 Q2 = Q1
Black
R13 R23 R1 = (1 e1)/(A1.e1)
R12 = 1/(A1.F12)
e1 = 0.8,
J3 = Eb3 R13 = 1/(A1.F13)
T1 = 700 K
Q3 = 0 R23 = 1/(A2.F23)

(a) (b)
FIGURE Example 13.22 One black surface, one insulated surface, and one grey surface forming an enclo-
sure, and its radiation network
Fig. Example 13.22 also shows the radiation network for this problem. This is drawn remembering the principles
already stated, i.e. (a) for a black surface, the surface resistance is zero, and Eb = J, (b) for an insulated (or re-radiating)
surface, Q = 0 and J = Eb, (c) for a grey surface, add a surface resistance, (1 – e)/(A.e), and (d) connect the radiosity
potentials by the respective space resistances (1/Ai ×Fij).
Data:
Let the length of the duct be 1 m
i.e. L := 1 m (length of duct) W := 0.75 m (side of equilateral triangle) T1 := 700 K e1 := 0.8
T2 := 1000 K s := 5.67 ´ 10 –8 W/(m2K) (Stefan–Boltzmann constant)
Now, we have, for view factors:
F 11 + F 12 + F 13 = 1 (by summation rule)
But, F 11 = 0 (since surface 1 is flat, and cannot ‘see’ itself.)
Then, F 12 + F 13 = 1
Further, by symmetry, F 12 = F13 for equilateral triangle.
Therefore, F 12 = 0.5
and, F 13 = 0.5
Similarly, F 23 = 0.5
Since surface 3 is re-radiating surface, net heat transfer for that surface Q 3 = 0.
Therefore, Q 1 = –Q 2
And, radiation network is a simple series–parallel network as shown in Fig. Example 13.22 (b) above. Then, Q 1 is
determined directly as:
Eb 1 - Eb 2
Q1 =
F1
R +G
I -1
W ...(a)

1
HR 12
+
1
R13 + R23 JK
Eb1 - Eb 2
i.e. Q1 =
F I -1

1 - e1 G
+ GA ×F +
1 JJ
A1 × e 1 GG 1 12 1 1
JJ
H +
A1 × F13 A2 × F23 K
Areas:
A 1 := W ×L i.e. A1 = 0.75 m2 (area of surface 1)
and, for equilateral triangle:
A 2 := A1
and, A 3 := A1
Resistances:
1 - e1
Surface resistance: R1 :=
A1 × e 1
i.e. R 1 = 0.333 m– 2
Space resistances:
1
R12 :=
A1 × F12

692 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. R 12 = 2.667 m– 2
1
and, R13 :=
A1 × F13
i.e. R 13 = 2.667 m– 2
1
and, R23 :=
A2 × F23
i.e. R 23 = 2.667 m–2
Therefore, from Eq. a:
s × (T14 - T24 )
Q 1 := W. (a)
F1 I -1

R1 + GH R
12
+
1
R13 + R23 JK
i.e. Q1 = –2.041 ´ 10 4 W
and, Q 2 = –Q 1 = 2.041 ´ 10 4 W (energy to be supplied to heated surface 2 per metre length. )
Example 13.23. Two co-axial cylinders of 0.4 m and 1 m diameter are 1 m long. The annular top and bottom surfaces are
well insulated and act as re-radiating surfaces. The inner surface is at 1000 K and has an emissivity of 0.6. The outer
surface is maintained at 400 K and its emissivity is 0.4.
(i) Determine the heat exchange between the surfaces
(ii) If the annular base surfaces are open to the surroundings at 300 K, determine the radiant heat exchange.
If the outer cylinder is surface 2, take F21= 0.25 and F22= 0.27. (M.U. Dec. 1998)

1.0 m diameter
Re-radiating surface 3,
0.4 m diameter R12
Q3 = 0
Q12
Surface 1, Eb1 J1 J2 Eb2
A1, e1, T1 Q1 Q2 = Q1
R1 R2
Q13 Q23
Surface 2, R1 = (1 e1)/(A1.e1)
A2, e2, T2 R13 R23
1.0 m R2 = (1 e2)/(A2.e2)
J3 = Eb3
Q3 = 0 R12 = 1/(A1.F12)
R13 = 1/(A1.F13)
R23 = 1/(A2.F23)

(a) (b)

FIGURE Example 13.23 Two gray surfaces surrounded by a re-radiating surface

Solution. See Fig. Example 13.23. Let the inner surface be denoted by 1, outer surface by 2, and the two annular surfaces
by 3. Then, surfaces 1, 2 and 3 form an enclosure. And, the rediation network will look as shown in the Fig. Example
13.23.
Data:
D 1 := 0.4 m D2 := 1 m L := 1 m T1 := 1000 K T2 := 400 K e1 := 0.6 e2 := 0.4
F21 := 0.25 F22 := 0.27
s := 5.67 ´ 10 –8 W/(m 2K) (Stefan–Boltzmann constant)
Areas:
A1 := p× D1 × L i.e. A1 = 1.257 m 2 (surface area of inner cylinder 1)
A2 := p× D2 × L i.e. A 2 = 3.142 m 2 (surface area of outer cylinder 2)
To find F 12:
F 11 = 0 (since surface 1 is convex, and does not ‘see’ itself.)
A2
Then, F 12 := ×F21 (by reciprocity)
A1
i.e. F 12 = 0.625 (view factor from surface 1 to surface 2)
Also, F11 + F12 + F 13 =1 (by summation rule)

RADIATION 693
i.e. F 13 := 1 – (F 11 + F 12)
i.e. F 13 = 0.375 (view factor from surface 1 to surface 3)
Also, F 21 + F22 + F 23 =1 (by summaion rule)
i.e. F 23 := 1 – (F 21 + F 22)
i.e. F 23 = 0.48 (view factor from surface 2 to surface 3)
Emissive powers:
Eb1 := s×T14 i.e. Eb1 = 5.67 ´ 104 W/m2 (Emissive power of surface 1)
E b2 := s×T24 i.e. Eb2 = 1.452 ´ 103 W/m2 .(Emissive power of surface 2)
Resistances:
1- e1
R 1 := i.e. R1 = 0.531 m–2 (surface resistance of inner cylinder 1)
A1 × e 1

1- e2
R 2 := i.e. R 2 = 0.477 m–2 (surface resistance of outer cylinder 2)
A2 × e 2
1
R 12 := i.e. R 12 = 1.273 m–2 (space resistance between surfaces 1 and 2)
A1 × F12
1
R 13 := i.e. R 13 = 2.122 m–2 (space resistance between surfaces 1 and 3)
A1 × F13
1
R 23 := i.e. R 23 = 0.663 m–2 (space resistance between surfaces 2 and 3)
A2 × F23
Case (i): When both the annular surfaces act as re-radiating surfaces:
Eb1 - Eb 2
We have: Q= W (heat exchange between the surfaces)
R1 + Reff + R2
The radiation network is as shown above. For the series–parallel network of resistances, we observe that R12 and
(R 13 + R23) are in parallel. Therefore, effective resistance Reff is given by:
1 1 1
= +
Reff R12 R13 + R23

F1 1 I -1

i.e. R eff := GH R
12
+
R13 + R23 JK
i.e. Reff = 0.874 m–2 (effective resistance)
Therefore,
Eb1 - Eb 2
Q := W (heat exchange between the surfaces)
R1 + Reff + R2
4
i.e. Q = 2.936 ´ 10 W (heat exchange between the surfaces.)
In addition, for case (i), if we wish to determine the temperature of re-radiating surface:
Apply the condition that for re-radiating surface, in steady state,
heat received by the surface = heat lost by the surface
J1 - J 3 J - J2
i.e. = 3
R13 R23
So, we have to determine J1 and J2.
Now, Q = Q 1 = –Q 2
Eb1 - J1
We have: Q1 = and, Q 1 := Q
R1
i.e. J1 := Eb1 – Q 1 ×R1
i.e. J1 = 4.112 ´ 104 W/m2
and,
Eb2 - J 2
Q2 = and, Q 2 := –Q1
R2

694 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. J2 := Eb2 – Q2 ×R 2
i.e. J2 = 1.547 ´ 104 W/m2
J1 - J 3 J - J2
Now, = 3 (for re-radiating surface)
R13 R23
Therefore, J3 ×(R13 + R 23) = J1 ×R 23 + J 2 ×R 13
J1 × R23 + J 2 × R13
i.e. J3 :=
R13 + R23
i.e. J3 = 2.158 ´ 104 W/m2
But, J3 = Eb3 = s×T34

FG J IJ
1
4
T3 :=
HsK
3
Therefore,

i.e. T 3 = 785.429 K (equilibrium temperature of re-radiating surface.)


Case (ii): When both the annular surfaces are open to surroundings at 300 K:
Now, Tsurr := 300 K (temperature of surroundings)
Further, E b3 = s× T 4surr
i.e. Eb3 = 459.27 W/m2 (emissive power of surface 3 (i.e. surroundings))
and, J3 := Eb3
To find J 1 and J 2 Apply Kirchhoff’s law to nodes J1 and J2 :
Eb1 - J1 J - J1 J - J1
At J 1: + 2 + 3 = 0 (a)
R1 R12 R13
Eb2 - J 2 J - J2 J - J2
At J 2: + 1 + 3 = 0 (b)
R2 R12 R23
To get values of J 1 and J2 solve Eqs. a and b simultaneously:
We shall use solve block of Mathcad to solve Eqs. a and b:
First, choose trial (or, guess) values for J 1 and J 2. Then, immediately after ‘Given’, write the constraints, i.e. Eqs. a
and b. Now, type ‘Find (J1, J2) = ’, and the result appears immediately:
J1 := 100 J2 := 100 (trial values)
Given
Eb1 - J1 J - J1 J - J1
+ 2 + 3 = 0
R1 R12 R13
Eb2 - J 2 J - J2 J - J2
+ 1 + 3 = 0
R2 R12 R23

Find (J 1, J2) =
LM3.591 ´ 10 OP
4

MN7.278 ´ 10 PQ
3

i.e. J1 := 3.591 ´ 104 W/m2


and, J2 := 7.278 ´ 103 W/m2
Therefore, heat lost by surface 1:
Eb1 - J1
Q 1 :=
R1
i.e. Q 1 = 3.919 ´ 104 W.
And, heat lost by surface 2:
Eb2 - J 2
Q 2 :=
R2
i.e. Q 2 = – 1.22 ´ 104 W.
Note that negative sign indicates that flow is into the surface.
Heat gained by surroundings:

RADIATION 695
J1 - J 3 J - J3
Qs = + 2
R13 R23
i.e. Qs = 2.699 ´ 104 W
Verify:
Heat gained by the surroundings must be equal to hea lost by the surfaces.
i.e. Qs = Q1 + Q2
Q 1 + Q 2 = 2.699 ´ 104 = Qs (verified.)
Example 13.24. Two parallel plates, 0.5 m ´ 1 m each, are spaced 0.5 m apart. The plates are at temperatures of 1000°C
and 500°C and their emissivities are 0.2 and 0.5, respectively. The plates are located in a large room, the walls of which
are at 27°C. The surfaces of the plates facing each other only exchange heat by radiation. Determine the rates of heat lost
by each plate and heat gain of the walls by radiation. Use radiation network for solution.
Assume shape factor between parallel plates: F12 = F21 = 0.285. (M.U. 1996)

R12
Q12
Surface 1, Eb1 J1 J2 Eb2
A1, e1 = 0.2, 1.0 m Q1 Q2 = Q1
T1 = 1000°C R1 R
Q13 Q23 2
1 0.5 m R1 = (1 e1)/(A1.e1)
R13 R23 R2 = (1 e2)/(A2.e2)
Surface 2, Eb3 = J3
R12 = 1/(A1.F12)
A2, e2 = 0.5,
2 0.5 m Surrounding room (Surface 3), R13 = 1/(A1.F13)
T2 = 500°C T3 = 27°C, A3 >> (A1, A2)
R23 = 1/(A2.F23)
1.0 m
(a) (b)

FIGURE Example 13.24 Two grey surfaces surrounded by a large room

Solution.
This is a three-zone enlosure, and the radiation network for this system is shown in Fig. Example 13.24 (b) above.
Since the area A3 of the room is very large, we can take the surface resistance of A3 as equal to zero.
1- e3
i.e. = 0
A3 × e 3
This means that Eb3 = J 3, i.e. a large room is equivalent to a black surface.
Data:
A1 := 0.5 m2 A2 := 0.5 m2 T 1 := 1000 + 273 K T2 := 500 + 273 K T3 := 27 + 273 K e1 := 0.2
e 2 := 0.5 F12 := 0.285 F21 := 0.285 s := 5.67 ´ 10–8 W/(m2K) (Stefan–Boltzmann constant)
Now, F 11 + F12 + F 13 =1 (by summation rule)
But, F 11 = 0 (since surface 1 is flat and can not ‘see’ itself.)
Therefore, F 12 + F 13 = 1
and, F 13 := 1 – F 12
i.e. F 13 = 0.715 (view factor of surface 1 w.r.t. surface 3)
Similarly, F 23 := 0.715 (view factor of surface 2 w.r.t. surface 3)
Resistances:
1 - e1
R 1 := i.e. R 1 = 8 m–2 (surface resistance of surface 1)
A1 × e 1
1- e2
R 2 := i.e. R 2 = 2 m–2 (surface resistance of surface 2)
A2 × e 2
1
R 12 := i.e. R 12 = 7.018 m–2 (space resistance between surfaces 1 and 2)
A1 × F12

696 FUNDAMENTALS OF HEAT AND MASS TRANSFER


1
R 13 := i.e. R 13 = 2.797 m–2 (space resistance between surfaces 1 and 3)
A1 × F13
1
R 23 := i.e. R 23 = 2.797 m–2 (space resistance between surfaces 2 and 3)
A2 × F23
Heat lost by each surface:
Eb1 - J1
Q1 = = heat lost by surface 1
R1

Eb2 - J 2
and, Q2 = = heat lost by surface 2
R2
And, heat gain by surface 3:
Q 3 = Q 13 + Q 23
J1 - J 3 J - J3
i.e. Q3 = + 2
R13 R23
Therefore, the problem reduces to calculating the radiosities, J1, J 2 and J3.
To calculate the radiosities J1 and J2, apply Kirchhoff’s law of electric circuits of nodes J1 and J2 :
Eb1 - J 1 J - J1 Eb3 - J 1
Node J1: + 2 + = 0 (a)
R1 R12 R13

J1 - J 2 E - J2 E - J2
Node J2: + b3 + b2 = 0 (b)
R12 R23 R2
Emissive powers:
Eb1 := s×T14 i.e. Eb1 = 1.489 ´ 105 W/m2 (for surface 1)
4
Eb2 := s×T2 i.e. Eb2 = 2.024 ´ 10 4 W/m2 (for surface 2)
E b3 := s×T34 i.e. Eb3 = 459.27 W/m2 (for surface 3)
Note that: J3 := Eb3 (for the large room.)
To get J 1 and J2, solve Eqs. a and b simultaneously. To do this, we shall use solve block of Mathcad.
First, choose trial (or, guess) values for J1 and J 2. Then, immediately after ‘Given’, write the constraints i.e. Eqs. a
and b. Now, type ‘Find (J1, J2) = ’, and the result appears immediately:
J1 := 100 J2 := 100 ...trial values
Given
Eb1 - J1 J - J1 E - J1
+ 2 + b3 = 0
R1 R12 R13
J1 - J 2 Eb3 - J 2 Eb2 - J 2
+ + = 0
R12 R23 R2

Find (J1, J2) =


LM3.3476 ´ 10 OP
4

MN1.5057 ´ 10 PQ
4

i.e. J1 := 3.3476 ´ 104 W/m2


and, J2 := 1.5057 ´ 104 W/m2
Therefore,
Heat lost by each surface:
Eb1 - J1
Q1 :=
R1
i.e. Q 1 = 1.443 ´ 104 W (= heat lost by surface 1.)
Eb2 - J 2
And, Q 2 :=
R2
i.e. Q 2 = 2.594 ´ 103 W (= heat lost by surface 2.)
Now, heat lost by both surfaces 1 and 2 is gained by the surroundings; so, heat gained by surroundings = Q 3 = Q 1
+ Q2

RADIATION 697
i.e. Q3 := Q 1 + Q 2
i.e. Q 3 = 1.702 ´ 104 W (= heat gained by surface 3.)
Verify:
FJ -J J2 - J3 I
We have: Q 3 := GH R
1

13
3
+
R23 JK
i.e. Q3 = 1.702 ´ 10 4 W (= heat gained by surface 3...verified.)

13.8 Radiation Shielding


In practice, quite often, one or more ‘radiation shields’ are used to reduce radiant heat transfer between two
given surfaces. Radiation shield is, simply a thin, high reflectivity surface placed in between the surfaces which
exchange heat between themselves. Radiation shields may be made of aluminium foils, copper foils, or
aluminised mylar sheets, etc. Radiation shields are extensively used in building industry to reduce radiant heat
loss from or to the walls; in cryogenic industry as ‘super-insulation’ (i.e. alternate layers of an insulator and
reflector, e.g. glass fibre mat + aluminium foils or, aluminised mylar sheets, about 25 numbers per inch) to reduce
the heat leakage into cryogenic vessels or cryostats, in space industry, again as ‘super-insulation’ to reduce heat
in-leaks, etc. Radiation shield does not participate in heat transfer, i.e. it does not add or remove heat from the
system as such, but reduces the heat transfer by interposing additional ‘resistance’ in the path of heat transfer.
We shall study how the heat transfer is reduced by the use of radiation shields, with reference to two infi-
nite, parallel plates, which exchange heat between themselves.
Fig. 13.37 (a) shows two large parallel plates, 1 and 2 exchanging heat between themselves; let their areas,
temperatures (in Kelvin) and emissivities be (A1, T1, e1) and (A2, T2, e2). Let a radiation shield 3, be placed be-
tween these plates. Plate 3 is thin and made of a material of high reflectivity. Let the emissivities of two sides of
the radiation shield be e 3–1 and e3–2 as shown. Radiation network for this system is shown in Fig. 13.37 (b). This
is drawn, as usual, remembering that each grey surface has a ‘surface resistance’ associated with it, and the two
radiosity potentials are connected by a ‘space resistance’.

Shield
Surface 1, e3 1, T3 e3 2, T3 Surface 2,
A1, e1, T1 A2, e2, T2
Q12 Q12 Q12

(a)
Eb1 J1 1/(A1.F13) 1/(A3.F32) J Eb2
2
Q1
(1 e1)/(A1.e1) (1 e2)/(A2.e2)
(1 e3,1)/(A3.e3,1) (1 e3,2)/(A3.e3,2)
(b)

FIGURE 13.37 Radiation shield between two parallel plates, and associated radiation network

When there is no shield, the radiation heat transfer between plates 1 and 2 is already shown to be:

A ×s × (T14 - T24 )
Q 12 = (for infinitely large parallel plates...(13.59))
1 1
+ -1
e1 e 2
With one shield placed between plates 1 and 2, the radiation network will be as shown in Fig. 13.17 (b)
above. Note that now all the relevant resistances are in series. Net heat transfer between plates 1 and 2 is given
as:
Q 12_one shield = (Eb1 – Eb2)/Rtot where, Rtot is the total resistance.

698 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e.
Eb1 - Eb 2
Q 12_one_shield =
1 - e1 1 1 - e 3 _1 1 - e 3_ 2 1 1 - e2
+ + + + +
A1 × e 1 A1 × F13 A3 × e 3 _ 1 A3 × e 3 _ 2 A3 × F32 A2 × e 2
(for two grey surfaces with one radiation shield placed in between....(13.70))
Now, for two large parallel plates, we note:
F 13 = F 32 = 1 and, A1 = A 2 = A 3 = A
Then, Eq. 13.70 simplifies to:

A ×s ×(T14 - T24 )
Q 12_one_shield =
FG 1 + 1 - 1IJ + F 1 I
- 1J
...(13.71)

H e e K GH e
1
+
1 2 3_1 e 3_ 2 K
Note that as compared to Eq. 13.59 for the case of no-shield, we have, with one shield, an additional term
appearing in the denominator of Eq. 13.71. Therefore, if there are N radiation shields, we have, for net radiation
heat transfer:

A ×s × (T14 - T24 )
Q12 N _ shields =
FG 1 + 1 - 1IJ + F 1 I F 1
- 1J + ... + G
I
- 1J
...(13.72)

H e e K GH e
1 1
+ +
1 2 3_1 e3_2 K He N _1 e N _2 K
If emissivities of all surfaces are equal, Eq. 13.72 becomes:

A ×s × (T14 - T24 ) 1
Q12 N _ shields =
FG 1 1
=
IJ
( N + 1)
× (Q12no _ shield ) ...(13.73)

H
( N + 1) × + - 1
e e K
Note this important result, which implies that, when all emissivities are equal, presence of one radiation shield
reduces the radiation heat transfer between the two surfaces to one-half, two radiation shields reduce the heat
transfer to one-third, 9 radiation shields reduce the heat transfer to one-tenth, etc.
For a more practical case of the two surfaces having emissivities of e 1 and e 2, and all shields having the same
emissivity of es, Eq. 13.72 becomes:

A ×s ×(T14 - T24 )
Q12 N _ shields =
FG 1 + 1 - 1IJ + N ×FG 2 - 1IJ ...(13.74)

He e K He K
1 2 s
To determine the equilibrium temperature of the radiation shield:
Once Q 12 is determined from Eq. 13.71, the temperature of the shield is easily found out by applying the condi-
tion that in steady state:
Q 12 = Q 13 = Q 32 ...(13.75)
We can use either of the conditions: Q12 = Q13 or Q12 = Q32.
Q13 or Q32 is determined by applying Eq. 13.59; i.e. we get:

A ×s × (T14 - T34 )
Q 12 = Q 13 = ...(13.76a)
1 1
+ -1
e1 e 3
or,

A ×s × (T34 - T24 )
Q 12 = Q 32 = ...(13.76b)
1 1
+ -1
e3 e2
In both the above equations, T3 is the only unknown, which can easily be determined.
RADIATION 699
For a cylindrical radiation shield placed in between two, long concentric cylinders:
Consider the case of radiation heat transfer between two long, concentric cylinders. The radiation heat transfer
between two long, concentric cylinders is already shown to be:

A ×s ×(T14 - T24 )
Q 12 =
1 FG IJ FG
A1 1 IJ ...for infinitely long concentric cylinders...(13.60)

e1
+
H KH
A2
×
e2
-1
K
where,
A1 r
= 1
A2 r2
Now, let a cylindrical radiation shield, 3, be placed in between the inner cylinder (1) and the outer cylinder
(2), as shown in Fig. 13.38.
The radiation network for this system is shown in Fig. 13.38 (b) and it is exactly the same as shown in Fig.
13.37 (b). And, the radiation heat transfer between cylinders 1 and 2, when the shield is present, is given by:
Eb1 - Eb 2
Q 12_one_shield =
1 - e1 1 1 - e 3 _1 1 - e 3_ 2 1 1 - e2
+ + + + +
A1 × e 1 A1 × F13 A3 × e 3 _ 1 A3 × e 3 _ 2 A3 × F32 A2 × e 2
(for two grey surfaces with one radiation shield placed in between...(13.70))

Surface 1,
A1, e1, T1
Surface 2,
A2, e2, T2
Shield 3,
e3, T3

r1
r2
r3
(a)
Eb1 J1 1/(A1.F13) 1/(A3.F32) J Eb2
2
Q1
(1 e1)/(A1.e1) (1 e2)/(A2.e2)
(1 e3,1)/(A3.e3,1) (1 e3,2)/(A3.e3,2)
(b)
FIGURE 13.38 Radiation shield between two concentric cylinders, and associated radiation network

Now, for the cylindrical system, we have:


F 13 = F 32 = 1
A 1 = 2×p×r1 ×L
A 2 = 2×p×r2 ×L
and, A 3 = 2×p×r3 ×L
Then, Eq. 13.70 reduces to:

A1 ×s × (T14 - T24 )
Q 12one_shield =
FG IJ FG
A1 A1 IJ F I F I
K GH JK GH JK
1 1 1 1
e1
+
H KH
A2
×
e2
-1 +
A3
× +
e 3_1 e 3_ 2
-1

(for concentric cylinders with one radiation shield...(13.77))


In Eq. 13.77, we have: (A1/A2) = (r1/r2), and (A1/A3) = (r1/r3).

700 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Note that as compared to the relation for two concentric cylinders with no shield (i.e. Eq. 13.60), an addi-
tional term appears in the denominator of Eq. 13.77 (i.e. the third term) when one radiation shield is introduced;
if there is a second radiation shield, say (4), then one more similar term will have to be added in the denominator
to take care of the resistance of that shield.
In this case, too, the equilibrium temperature of the shield is determined by applying the principle that, in
steady state,
Q 12 = Q 13 = Q 32.
For a spherical radiation shield placed in between two concentric spheres:
This case is also represented by Fig. 13.38 (a), where inner sphere 1 is enclosed by an outer sphere 2, and a
radiation shield 3, is placed in between. The radiation network for this system is shown in Fig. 13.38 (b).
When there is no radiation shield, radiation heat transfer between surfaces 1 and 2 is given by Eq. 13.61, i.e.

A1 ×s ×(T14 - T24 )
Q12 =
1 A1 FG IJ FG
1 IJ (for concentric spheres...(13.61))

e1
+
A2
×
H KH
e2
-1
K
where,

A1 Fr I 2

A2
= GH r JK
1
2
Again, when the radiation shield is present, the general relation for radiation heat transfer between surfaces
1 and 2 is Eq. 13.70. Remembering that for concentric spheres,
F 13 = F 32 = 1
A 1 = 4×p×r12
A 2 = 4×p×r22
and, A 3 = 4×p×r32
relation for radiant heat transfer between surfaces 1 and 2, is exactly as Eq. 13.77, i.e.

A1 ×s × (T14 - T24 )
Q 12one_shield =
FG IJ FG
A1 A1 IJ F I F I
K GH JK GH JK
1 1 1 1
e1
+
H KH
A2
×
e2
-1 +
A3
× +
e 3_1 e 3_ 2
-1

(for concentric spheres with one radiation shield...(13.78))


In Eq. 13.78, we have:

A1 Fr I 2
A1 Fr I 2

A2
= GH r JK
1
2
and,
A3
= GH r JK
1
3
In this case also, equilibrium temperature of the shield is determined by applying the principle that, in
steady state,
Q12 = Q13 = Q32.
Example 13.25. Two large parallel planes facing each other and having emissivities 0.3 and 0.5 are maintained at 827°C
and 527°C, respectively. Determine the rate at which heat is exchanged between the two surfaces by radiation. If a
radiation shield of emissivity 0.05 on both sides is placed parallel between the two surfaces, determine the percentage
reduction in the radiant heat exchange rate. (M.U., Jan. 2002)
Solution. This is the case of one radiation shield placed in between two parallel plates. See Fig. Example 13.25.
Data:
T 1 := 827 + 273 K T2 := 527 + 273 K e1 := 0.3 e2 := 0.5 e31 := 0.05 e32 := 0.05
s := 5.67 ´ 10–8 W/(m2 K) (Stefan–Boltzmann constant) A := 1 m2 (surface area of plates...assumed)
(a) Heat exchange between surfaces 1 and 2, when there is no shield:
A ×s × (T14 - T24 )
Q12 := W (for infinitely large parallel plates...(13.59))
1 1
+ -1
e1 e 2

RADIATION 701
Surface 1, Shield T3
e1 = 0.3, e31 = 0.05 e32 = 0.05 Surface 2,
T1 = 827°C Q12 Q12 Q12 e2 = 0.5,
T2 = 527°C

(a)

Eb1 J1 1/(A1.F13) 1/(A3.F32) J Eb2


2
Q1
(1 e1)/(A1.e1) (1 e2)/(A2.e2)
(1 e3,1)/(A3.e3,1) (1 e3,2)/(A3.e3,2)
(b)
FIGURE Example 13.25 Radiation shield between two parallel plates, and associated radiation network
i.e. Q 12 = 1.38 ´ 104 W/m2 (radiant heat transfer, without shield.)
Now, we have for radiant heat transfer between surfaces 1 and 2; when there is one shield in between 1 & 2:
A ×s × (T14 - T24 )
Q 12_one_shield :=
FG 1 + 1 - 1IJ + F 1 I ...(13.71)

H e e K GH e JK
1
+ -1
1 2 31 e 32
i.e. Q 12_one_shield = 1.38 ´ 103 W/m2 (radiant heat transfer, with one shield.)
Therefore, percentage reduction in heat transfer due to radiation shield:
Q12 - Q12 _one_shield (1. 38 ´ 10 4 - 1. 38 ´ 10 3 )
Reduction = ×100 = × 100
Q12 1. 38 ´ 10 4
i.e. Reduction = 90%.
In addittion, if we wish to find out equilibrium temperature of shield:
Let the equilibrium temperature of shield be T 3.
In steady state, we have:
Q 12_one_shield = Q 13 = Q 32
Q 12 is already calculated. Q 13 or Q 32 is calculated using Eq. 13.59.
Let us take: Q 12_one_shield = Q13
A ×s × (T14 - T34 )
i.e. Q12_one_shield =
1 1
+ -1
e 1 e 31

LM F 1 + 1 - 1I OP
1

×G
H e e JK P
4
Q12_one_shield
i.e. T3
M
:= MT 4
-
1
PP 31

MM
1
A ×s
N PQ
i.e. T3 = 979.537 K
or, T3 = 706.537°C (equilibrium temperature of shield.)
Verify: Use the equation: Q 12_one_shield = Q 32
We get, writing for Q 32:
A ×s × (T34 - T24 )
Q 12_one_shield =
1 1
+ -1
e 1 e 32

LM F 1 + 1 - 1I OP
1

×G
H e e JK P
4
Q12_one_shield
i.e. T3 :=
MMT 4
-
2 32
PP
MM
2
A ×s
N PQ
702 FUNDAMENTALS OF HEAT AND MASS TRANSFER
i.e. T 3 = 979.537 K (same result as obtained above.)
Example 13.26. Two very large parallel plates with emissivities 0.3 and 0.7 exchange heat. Find the percentage reduction
in heat transfer when two polished aluminium radiation shields (E = 0.4) are placed between them. (M.U., Dec. 2000)
Solution. This is the case of two radiation shields placed in between two parallel plates.
Data:
e1 := 0.3 e2 := 0.7 es := 0.4
Then, with no radiation shield, we have the radiant heat transfer:
A ×s × (T14 - T24 )
Q12 = W (for infinitely large parallel plates...(13.59))
1 1
+ -1
e1 e 2
and, with 2 radiation shields, the radiant heat transfer is:
A ×s × (T14 - T24 )
Q 12two_shields =
F 1 + 1 - 1I + 2 ×F 2 - 1I (from Eq. 13.74)
GH e e JK GH e JK
1 2 s

Therefore, dividing the above two equations, we have:


1 1
+ -1
Q12 e1 e 2
=
F 1 + 1 - 1I + 2 ×F 1 + 1 - 1I
two– shields

Q12 _no_shield
GH e e JK GH e e JK
1 2 s s

Q12
i.e. two– shields
= 0.32
Q12 _no_shield
i.e. by introducing 2 radiation shields, the heat transfer is reduced to 32% of that without the shields.
Example 13.27. The net radiation from the surface of two parallel plates maintained at temperatures T1 and T2 is to be
reduced by 79 times. Calculate the number of screens to be placed between two surfaces to achieve this reduction in heat
exchange, assuming the emissivity of screens as 0.05 and that of surfaces as 0.8. (M.U)
Solution. This problem is on parallel plates with more than one radiation shields.
Data:
e 1 := 0.8 e2 := 0.8 es1 := 0.05 es2 := 0.05 s := 5.67 ´ 10– 8 W/(m2 K) (Stefan–Boltzmann constant)
Let N be the number of screens required.
Then, with no radiation shield, we have the radiant heat transfer:
A ×s × (T14 - T24 )
Q 12 = W (for infinitely large parallel plates...(13.59))
1 1
+ -1
e1 e 2
and, with N radiation shields, the radiant heat transfer is:
A ×s × (T14 - T24 )
Q 12N_shields = (from Eq. 13.74)
FG 1 + 1 - 1IJ + N × FG 1 1 IJ
He e K He
1 2 s1
+
e s2
-1
K
1 1
+ -1
Q12 N_shields e1 e 2 1
Then, by data:
Q12
=
F 1 + 1 - 1I + N ×F 1 I =
79
...(a)
GH e e JK GH e
1 2 s1
+
1
e s2
-1 JK
Solving Eq. a, we get N, the number of screens required.
We get:

F 1 + 1 - 1I - F 1 + 1 - 1I
79 × GH e e JK GH e e JK
1 2 1 2
N :=
F 1 + 1 - 1I
GH e e JK s1 s2

RADIATION 703
i.e. N=3 (number of screens required to reduce heat loss by 79 times.)
Example 13.28. A 10 mm OD pipe carries a cryogenic fluid at 80 K. This pipe is encased by another pipe of 15 mm OD,
and the space between the pipes is evacuated. The outer pipe is at 280 K. Emissivities of inner and outer surfaces are 0.2
and 0.3, respectively. (a) Determine the radiant heat flow rate over a pipe length of 5 m. (b) If a radiation shield of
diameter 12 mm and emissivity 0.05 on both sides is placed between the pipes, determine the percentage reduction in
heat flow. (c) What is the equilibrium temperature of the shield?
Solution.

Cylinder 1,
e1 = 0.2, T1 = 80 K
Cylinder 2,
e2 = 0.3, T2 = 280 K
Shield 3, T3
e31 = 0.05, e32 = 0.05
r1 = 0.005 m
r3 = 0.006 m

r2 = 0.0075 m
(a)
Eb1 J1 1/(A1.F13) 1/(A3.F32) J Eb2
2
Q1
(1 e1)/(A1.e1) (1 e2)/(A2.e2)
(1 e3,1)/(A3.e3,1) (1 e3,2)/(A3.e3,2)
(b)

FIGURE Example 13.28 Radiation shield between two concentric cylinders, and associated radiation network

Data:
r 1 := 0.005 m r2 := 0.0075 m r3 := 0.006 m T1 := 80 K T2 := 280 K e1 := 0.2 e2 := 0.3
e 31 := 0.05 e32 := 0.05 s := 5.67 ´ 10 –8 W/(m2 K) L := 5 m
Surface areas for 5 m length:
A1 := 2×p×r 1 × L i.e. A1 = 0.157 m2 (surface area of inner pipe)
A2 := 2×p×r 2 × L i.e. A2 = 0.236 m2 (surface area of outer pipe)
A3 := 2×p×r 3× L i.e. A3 = 0.188 m2 (surface area of radiation shield)
(a) Heat transfer without the shield being present:
We have:
A1 ×s × (T14 - T24 )
Q 12 := (for infinitely long concentric cylinders...(13.60))
1 A1FG IJ FG
1 IJ
e1
+
A2 H KH
×
e2
-1
K
i.e. Q 12 = – 8.295 W
Note that negative sign indicates that heat flow is from outside to inner pipe.
(b) Heat transfer with one shield being present:
Now, we have, for heat transfer,
A1 ×s × (T14 - T24 )
Q 12_one_shield := (for concentric cylinders with one radiation shield...(13.77))
FG IJ FG
A1 A1 IJ F I F I
K GH JK GH JK
1 1 1 1
e1
+
A2H KH
×
e2
-1 +
A3
× +
e 31 e 32
-1

i.e. Q 12_one_shield = – 1.392 W (for concentric cylinders with one radiation shield.)
Again, note that negative sign indicates that heat flow is from outside to inner pipe.
Therefore percentage reduction in heat flow due to shield:
( 8. 295 - 1. 392)
Reduction := × 100
8. 295

704 FUNDAMENTALS OF HEAT AND MASS TRANSFER


i.e. Reduction = 83.219%.
(c) Equilibrium temperature of shield:
Let the equilibrium temperature of shield be T 3.
In steady state, we have:
Q 12_one_shield = Q 13 = Q 32
Q 12 with one shield is already calculated. Q 13 or Q 32 is calculated using Eq. 13.59.
Let us take: Q 12_one_shield = Q 13
A1 × s × (T14 - T34 )
Q 12_one_shield =
i.e.
A1 F IF I
1
e1
+
A3
× GH JK GH
1
e 31
-1 JK

LM L 1 F A I ×F 1 - 1I OP OP
1

×M + G
4

M Q12_one_shield
MN e H A JK GH e JK PQ PP
1

:= MT
1 3 31
i.e. T3 4
-
MM 1
A ×s 1 PP
MN PQ
i.e. T 3 = 239.639 K
or, T 3 = 33.361°C (equilibrium temperature of shield.)
Verify: Use the equation: Q 12_one_shield = Q 32
We get, writing for Q 32
A3 ×s × (T34 - T24 )
Q 12_one_shield =
1 FG IJ FG
A3 1 IJ
e 32
+
A2H KH
×
e2
-1
K
LM L 1 + F A I × F 1 - 1I OP OP
1

×M
4

MN e GH A JK GH e JK PQ PP
Q12_one_shield
MMT
3

32 2 2
i.e. T3 := 4
+
MM
2
A ×s 3 PP
N Q
i.e. T 3 = 239.639 K (same result as obtained above.)
Example 13.29. A spherical tank with diameter D1 = 40 cm, filled with a cryogenic fluid at T1 = 100 K, is placed inside a
spherical container of diameter D2 = 60 cm, maintained at T2 = 300 K. Emissivities of inner and outer tanks are e 1 = 0.10
and e 2 = 0.20, respectively.
(i) Find the rate of heat loss into the inner vessel by radiation
(ii) If a spherical radiation shield of diameter D3 = 50 cm, with an emissivity e 3 = 0.05 on both surfaces is placed
between the spheres, what is the new rate of heat loss? (M.U. Jan. 2002)
Solution. This is a problem on spherical radiation shield. See Fig. Example 13.29 for schematic and the associated radia-
tion network.
Data:
r 1 := 0.2 m r2 := 0.3 m r3 := 0.25 m T1 := 100 K T2 := 300 K e1 := 0.1 e2 := 0.2 e31 := 0.05
e 32 := 0.05 s := 5.67 ´ 10– 8 W/(m2 K)
Areas:
A1 := 4×p× r12 i.e. A1 = 0.503 m2
2
A2 := 4×p× r2 i.e. A2 = 1.131 m2
and, A3 := 4×p× r32 i.e. A3 = 0.785 m2
(a) When there is no radiation shield:
We have, for radiation heat transfer between two concentric spheres:
A1 × s × (T14 - T24 )
Q 12 := (for concentric spheres...(13.61))
1 A1 FG IJ FG
1 IJ
e1
+
A2
×
H KH
e2
-1
K
RADIATION 705
Sphere 1,
e1 = 0.1, T1 = 100 K
Sphere 2,
e2 = 0.2, T2 = 300 K
Shield 3, T3
e31 = 0.05, e32 = 0.05
r1 = 0.2 m
r3 = 0.25 m

r2 = 0.3 m
(a)
Eb1 J1 1/(A1.F13) 1/(A3.F32) J Eb2
2
Q1
(1 e1)/(A1.e1) (1 e2)/(A2.e2)
(1 e3,1)/(A3.e3,1) (1 e3,2)/(A3.e3,2)
(b)

FIGURE Example 13.29 Radiation shield between two concentric spheres, and associated radiation network

i.e. Q 12 = –19.359 W (radiation heat transfer when there is no shield.)


Note: Negative sign indicates that heat transfer is radially inwards, i.e. from outer sphere to inner sphere.
(b) When the radiation shield is present:
For radiation heat transfer between concentric spheres with a radiation shield placed in between, we can directly use the
Eq. 13.78. However, we shall work from fundamentals, and use the Eq. 13.70, written for the radiation network shown
above, and then verify the result from Eq. 13.78:
Now, we have from Eq. (13.70):
Eb1 - Eb 2
Q 12_one_shield =
1 - e1 1 1 - e 31 1 - e 32 1 1- e2
+ + + + +
A1 × e 1 A1 × F13 A3 × e 31 A3 × e 32 A3 × F32 A2 × e 2
(for two grey surfaces with one radiation shield placed in between...(13.70))
Emissive powers:
Eb1 := s×T14 i.e. Eb1 = 5.67 W/m2 (Emissive power of surface 1)
Eb2 := s×T24 i.e. Eb2 = 459.27 W/m2 (Emissive power of surface 2)
View factors:
F 13 := 1 (since all radiation emitted by surface 1 is intercepted by surface 3)
F 32 := 1 (since all radiation emitted by surface 3 is intercepted by surface 2)
Resistances:
1 - e1
R 1 := i.e. R 1 = 17.905 m–2 (surface resistance of surface 1)
A1 × e 1
1
R 13 := i.e. R 13 = 1.989 m–2 (space resistance between surfaces 1 and 3)
A1 × F13

1 - e 31
R3a := i.e. R 3a = 24.192 m–2 (surface resistance of 3, facing surface 1)
A3 × e 31
1 - e 32
R 3b := i.e. R3b = 24.192 m–2 (surface resistance of 3, facing surface 1)
A3 × e 32
1
R 32 := i.e. R 32 = 1.273 m–2 (space resistance between surfaces 3 and 2)
A3 × F32

1- e2
R 2 := i.e. R 2 = 3.537 m–2 (surface resistance of surface 2)
A2 × e 2

706 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Then, from Eq. 13.70:
Eb 1 - Eb 2
Q 12_one_shield :=
1 - e1 1 1 - e 31 1 - e 32 1 1- e2
+ + + + +
A1 × e 1 A1 × F13 A3 × e 31 A3 × e 32 A3 × F32 A2 × e 2
i.e. Q 12_one_shield = – 6.206 W (heat transfer between spheres, with radiation shield being present.)
Note: Negative sign indicates that heat transfer is radially inwards, i.e. from outer sphere to inner sphere.
Verify:
We have, from Eq. 13.78:
A1 × s ×(T14 - T24 )
Q 12one_shield = (for concentric spheres with one radiation shield...(13.78))
A1 FG IJ FG A1 IJ F I F I
K GH JK GH JK
1 1 1 1
e1
+
A2
×
H KH
e2
-1 +
A3
× +
e 31 e 32
-1

A1 Fr I 2
A1 Fr I 2

where,
A2
= GH r JK
1

2
and,
A3
= G J
Hr K
1

i.e. we get:
A1 ×s × (T14 - T24 )
Q 12_one_shield :=
F I ×F 1 - 1I + F r I × F 1
r
2 2
I
1
e1 GH JK GH e JK GH r JK GH e
+ 1
r2 2
1

3 31
+
1
e 32
-1 JK
i.e. Q 12_one_shield = –6.206 W (verified.)
In addition, if we wish to find out the equilibrium temperature of the shield:
(c) Equilibrium temperature of shield:
Let the equilibrium temperature of shield be T 3
In steady state, we have:
Q 12_one_shield = Q 13 = Q 32
Q 12 with one shield is already calculated. Q 13 or Q 32 is calculated using Eq. 13.61.
Let us take: Q 12_one_shield = Q 13
A1 × s × (T14 - T34 )
Q 12_one_shield =
i.e.
A1 F IF I
1
e1
+
A3
×GH JK GH
1
e 31
-1 JK

LM L 1 F A I ×F 1 - 1I OP OP
1

×M + G
4

M Q12_one_shield
MN e H A JK GH e JK PQ PP
1

:= MT
1 3 31
i.e. T3 4
-
MM1
A ×s 1 PP
MN PQ
i.e. T 3 = 264.919 K
or, T 3 = 8.081°C (equilibrium temperature of shield.)
Verify Use the equation: Q 12_one_shield = Q 32
We get, writing for Q 32
A3 ×s × (T34 - T24 )
Q 12_one_shield =
A3 F IF I
1
e 32
+
A2
×
1
GH JK GH
e2
-1 JK
LM L 1 + F A I × F 1 - 1I OP OP
1

×M
4

MN e GH A JK GH e JK PQ PP
Q12_one_shield
M
3

:= MT
32 2 2
i.e. T3 4
+
MM2
A ×s 3 PP
N Q
RADIATION 707
i.e. T 3 = 264.919 K (same result as obtained above.)

13.9 Radiation Error in Temperature Measurement


An important application of radiation shields is in reducing the radiation error in temperature measurement. To
explain this, consider the case of a hot fluid at a temperature Tf, flowing through a channel, whose walls are at a
temperature Tw. Let the convective heat transfer coefficient between the fluid and the thermometer bulb be h. To
measure the temperature of the fluid, a thermometer (or a thermocouple) is introduced into the stream, as shown
in Fig. 13.39 (a).

Tc Thermometer Tc
Thermometer
Tw Tw
Ts
Tf ,h Tf ,h
qconv
qrad
Tw Tw
(a) Thermometer without radiation shield (b) Thermometer with radiation shield

FIGURE 13.39 Radiation shielding of thermometers

Let the reading shown by the thermometer be Tc. This reading, however, does not represent the true tem-
perature of the fluid Tf, since the thermometer bulb will lose heat by radiation to the walls of the channel which
are at a lower temperature Tw (which is usually the case). So, in steady state, the thermometer bulb will gain heat
by convection from the flowing fluid and will lose heat by radiation to the walls, and as a result, the temperature
Tc shown by the thermometer will be some value in between Tf and Tw.
We wish to find out the true temperature of the fluid Tf, by knowing the thermometer reading Tc.
Making an energy balance on the thermometer bulb, in steady state, we have:
Without radiation shield:
q conv to the bulb = qrad from the bulb
i.e. h×Ac ×(Tf – Tc) = ec ×Ac ×s×(Tc4 – Tw4)

e c ×s ×(Tc4 - Tw4 )
i.e. Tf = Tc + ...(13.79)
h
where,
Ac = surface area of thermometer bulb,
ec = emissivity of thermometer bulb surface.
Eq. 13.79 gives the true temperature of the fluid Tf. Second term on the RHS of Eq. 13.79 represents the error
in temperature measurement due to radiation effect. It is clear that radiation error can be minimised by:
(i) having low value of e c , i.e. high reflectivity for the bulb surface
(ii) high value for convective heat transfer coefficient, h.
In practice, even if we start with a thermometer bulb surface of high reflectivity, soon, the emissivity value
rises to about 0.8 or 0.9 due to deposit formation, corrosion or erosion of the bulb surface, etc.
So, the most practical way to reduce the radiation error in temperature measurement is to provide a cylindri-
cal radiation shield around the thermometer bulb, as shown in Fig. 13.39 (b). Then, in steady state, the shield
temperature (Ts) will stabilise somewhere in between the fluid temperature Tf and the wall temperature Tw. Then,
in Eq. 13.79, Tw will be replaced by the effective shield temperature Ts.
Energy balance on the thermometer bulb:
Heat transferred to the bulb from the fluid by convection = Heat transferred from the bulb to the shield by
radiation,

s ×(Tc4 - Ts4 )
i.e. h ×Ac ×(Tf – Tc) =
F 1-e I + 1 + F1-e I ...(13.80)
GH A ×e JK A × F GH A ×e JK
c
c
c c cs s
s
s

708 FUNDAMENTALS OF HEAT AND MASS TRANSFER


In Eq. 13.80, Fcs = view factor of thermometer bulb w.r.t. the shield and is, generally equal to 1.
In the RHS of eqn. (13.80), first term in the denominator represents surface resistance of the bulb, second
term is the space resistance between the bulb and the shield and the third term is the surface resistance of the
shield.
Now, making an energy balance on the shield:
(heat transferred to shield from the fluid by convection + heat transferred to shield from bulb by radiation) = heat
transferred from shield to walls by radiation

s ×(Tc4 - Ts4 )
i.e. 2×As ×h×(Tf – Ts) +
F 1-e I + 1 + F1-e I = es ×As ×s×(Ts4 – Tw4) ...(13.81)
GH A ×e JK A × F GH A ×e JK
c
c
c c cs s
s
s

where, As = area of shield on one side


es = emissivity of shield surface
Ac = area of bulb surface
ec = emissivity of bulb surface
Fcs = view factor of bulb w.r.t. shield.
In the first term of the above equation factor 2 appears since convective heat transfer to the shield occurs on
both surfaces of the shield. Also, in writing the RHS, the inherent assumption is that:
Fsw = 1 (view factor between the shield and the walls)
and,
As
=0 (i.e. surface area of shield is negligible compared to the area of the channel walls.)
Aw
Solving Eqs. 13.80 and 13.81 simultaneously, we obtain the shield temperature Ts and the thermometer read-
ing Tc, (if Tf is known), or Tf (if Tc is known).
Example 13.30. Hot air is flowing in a duct whose walls are maintained at a temperature Tw = 450 K. A thermocouple
placed in the stream shows a reading of 650 K. If the emissivity of the thermocouple junction is ec = 0.8 and the convec-
tive heat transfer coefficient between the flowing air and the thermocouple is h = 85 W/(m2 C), find out the true tempera-
ture of the flowing stream.
(b) Now, if a radiation shield (e s = 0.3) is placed between the thermocouple and the walls, what will be new value
of Tc read by the thermocouple? And, how much is the temperature error? Take Ac/As = 1/5.
Solution. In case (a), there is no radiation shield and in case (b), the radiation shield is present. Both these cases are
shown in Fig. Example 13.30 (a) and (b).
Data:
Ac 1
Tw := 450 K Tc := 650 K h := 85 W/(m2 C) e c := 0.8 es := 0.3 =
As 5
s := 5.67 ´ 10– 8 W/(m2 K) (Stefan–Boltzmann constant.)
Case (a): When there is no radiation shield:
In steady state, making a heat balance on the thermocouple bead, we have:
q conv = q rad

Thermometer, Thermometer,
Tc = 650 K Tc = ?
Tw = 450 K Tw
Ts, es = 0.3
Tf , h = 85 W/(m2C) Tf ,h
ec = 0.8
qconv
qrad
Tw Tw
(a) Thermometer without radiation shield (b) Thermometer with radiation shield

FIGURE Example 13.30 Radiation shielding of thermometers

RADIATION 709
i.e. h×Ac ×(Tf – Tc) = s×ec ×Ac ×(Tc4 – Tw4)
e c × s ×(Tc4 - Tw4 )
i.e. T f := Tc + ...(13.79)
h
i.e. T f = 723.376 K (true temperature of air stream, when there is no radiation shield
Therefore, radiation error = Tf – Tc = 73.376 deg.
Case (b): When the radiation shield is present:
Making a heat balance on the thermocouple bead:
s × (Tc4 - Ts4 )
h×Ac ×(Tf – Tc) =
F 1- e I + 1 + F 1-e I ...(13.80)
GH A × e JK A × F GH A ×e JK
c
c

c c cs s
s

where, Fcs = 1 (view factor for thermocouple bead w.r.t. shield.)


Then, Eq. 13.80 becomes:
s × Ac × (Tc4 - Ts4 )
h×Ac ×(Tf – Tc) =
1 Ac 1 FG IJ FG IJ
ec
+
As
×
es H KH
-1
K
s × Ac × (Tc4 - Ts4 )
h×Ac ×(Tf – Tc) =
FG IJ FG IJ
i.e. (a)
1 1 1
ec
+
5
×
es H KH
-1
K
Next, making a heat balance on the shield:
s × (Tc4 - Ts4 )
2×As × h×(Tf – Ts) + = e s ×As ×s×(Ts4 – Tw4)
F 1- e I + 1 + F 1- e I ...(13.81)
GH A × e JK A × F GH A × e JK
c
c

c c cs s
s

where,
As = area of shield on one side
es = emissivity of shield surface
Ac = area of bulb surface
ec = emissivity of bulb surface
Fcs = view factor of bulb w.r.t. shield = 1
Fsw = 1 (view factor between the shield and the walls)
and,
As
= 0 (i.e. surface area of shield is negligible compared to the area of the channel walls)
Aw
Then, Eq. 13.81 becomes:
s × Ac × (Tc4 - Ts4 )
2×As × h×(Tf – Ts) +
Ac F IF I = e s ×As ×s×(Ts4 – Tw4)
1
ec
+
As
×
1
esGH JK GH
-1 JK
F A I ×h× (T – T ) + s × (Tc4 - Ts4 ) F A I ×s×(T
i.e. 2× GH A JK
s
f s
Ac F IF I = es × GH A JK
s
s
4
– Tw4)
c 1
ec
+
As
×
1
esGH JK GH
-1 JK c

s × (Tc4 - Ts4 ) Ac 1
10×h×(Tf – Ts) + = es ×(5)×s×(Ts4 – Tw4), since
IJ FG I
i.e. = (b)
1 1 FG
1
JK
As 5
ec
+
5
×
H
es
-1
KH
Now, Tf is already known, and solving Eqs. a and b simultaneously, we get Tc and Ts.
To do this, we use solve block of Mathcad. We start with trial values of Tc and Ts, and write the constraint Eqs. a
and b immediately after typing ‘Given’. Then, typing ‘Find’ (Tc, Ts) = gives immediately the values of Tc and Ts.

710 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Tc := 100 Ts := 100 (trial values of Tc and Ts)
Given
s × (Tc4 - Ts4 )
h ×(Tf – Tc) =
IJ FG IJ (a)
1 FG
1 1
ec
+
5H×
es KH
-1
K
s × (Tc4 - Ts4 )
10×h×(Tf – Ts) + = es ×(5)×s×(Ts4 – Tw4)
IJ FG IJ (b)
1 FG
1 1
ec
+
5H
KH
×
es
-1
K
L716.327 OP
Find (T , T ) = M
c s
N703.655 Q
i.e. Tc = 716.327 K (value of thermocouple reading, when the shield is present.)
And, Ts = 703.655 K (temperature of shield.)
Therefore, radiation error = Tf – Tc = 7.049 deg.
Note: When there is no radiation shield, the error in thermocouple reading is 73.376 deg. and when the radiation
shield is introduced, the radiation error is reduced to just 7.049 deg.

13.10 Radiation Heat Transfer Coefficient (hr)


Concept of “radiation heat transfer coefficient” is useful in solving problems where heat transfer occurs by both
convection and radiation. Typical examples of such a situation are: heat loss from a steam pipe passing through
a room, heat loss from hot combustion products passing through a duct, heat loss from the walls and door of a
furnace, etc.
Radiation heat transfer coefficient is defined in a manner analogous to convection heat transfer coefficient.
Consider hot gases at a temperature Tg flowing through a tube whose walls are at a temperature of Tw. Then,
recollect that the convective heat flux is given by:
q conv = h c ×(Tg – Tw)
where, hc = convective heat transfer coefficient.
In a similar manner, we write for radiant heat flux from the pipe:
q rad = hr ×(Tg – Tw)
where, hr = radiation heat transfer coefficient.
For the above case, hr is determined from:
hr ×(Tg – Tw) = e×s×(Tg4 – Tw4)
e ×s ×(Tg4 - Tw4 )
i.e. hr =
(Tg - Tw )
i.e. hr = e×s×(Tg2 + Tw2) (Tg + Tw) W/(m2C) ...(13.82)
Then, qtot = qconv + qrad, by a linear superposition of both heat fluxes.
= hc × (Tg – Tw) + hr × (Tg – Tw)
= (hc + hr) × (Tg – Tw) ...(13.83)
For any other configuration, we can determine hr if we know the expressin for radiant heat flux. For exam-
ple, for radiant heat transfer between two large parallel plates, we have:

Q s × (T14 - T24 )
= = hr ×(T1 – T2)
A 1 1
+ -1
e1 e2

s ×(T12 + T22 ) ×(T1 + T2 )


i.e. hr = W/(m2C) (for two parallel plates...(13.84))
1 1
+ -1
e1 e 2
Note that radiation heat transfer coefficient is a strong function of temperature, unlike the convective heat
transfer coefficient.
RADIATION 711
Gas layer 13.11 Radiation from Gases, Vapours and Flames
So far, we have dealt with radiation heat exchange between surfaces in an
Il(x) enclosure, with a non-participating medium, in between, i.e. the interven-
Il0 IlL ing gas neither absorbs nor scatters the radiation nor does it emit any radia-
tion; in other words, the intervening gas does not, in any way, affect the
radiant heat transfer between the surfaces. Such an assumption is valid for
x=0 x=L mono-atomic gases such as argon and helium, and for diatomic gases such
x as oxygen and nitrogen; these gases are extremely inert to thermal radia-
dx tion. However, the same thing is not true for poly-atomic gases such as
CO 2, H 2O (vap), NH 3 and hydrocarbon gases; these gases do absorb and
FIGURE 13.40 Absorption of emit radiation. Further, radiation from solids and liquids generally covers
monochromatic radiation the entire wavelength range whereas radiation from gases is over selected
in a gas layer wavelength ‘bands’. Also, note that radiation from solids is a ‘surface phe-
nomenon’ whereas that from gases is a ‘volumetric phenomenon’.
13.11.1 Volumetric Absorption and Emissivity
In gases, absorption of radiation depends upon the absorption coefficient k l (1/m) and thickness L of the gas
layer, in addition to the temperature Tg of the gas. Fig. 13.40 shows a monochromatic beam of intensity Il0
impinging on the gas layer at x = 0; its intensity decreases as a result of absorption and at x = L, let the intensity
be IlL.
We wish to develop a relation between the initial and final intensities: If Il is the intensity at any x, the
reduction in intensity occurring in an infinitesimal layer of thickness dx is given by:
dIl (x) = –kl ×Il (x)×dx
Separating the variables and integrating from 0 to L, i.e. over the entire thickness of gas layer, we get:

z
Il 0
I lL 1
I l ( x) z L
dI l ( x) = – kl 1dx
0
where, the absorption coefficient kl is assumed to be independent of x.
We get:
I lL
= exp (– kl ×L) ...(13.85)
Il 0
This is known as Beer’s Law.
i.e. the intensity of radiation decreases exponentially with thickness as it travels through the gas layer.
LHS of Eq. 13.85 is monochromatic transmissivity tl of the gas. Also, in general, gases do not reflect radia-
tion, i.e. their reflectivity is zero. Therefore, we write:
al + tl = 1
i.e. al = 1 – t l
i.e. al = 1 – exp(– kl ×L) (monochromatic absorptivity of gas...(13.86))
Then, from Kirchhoff’s law, since absorptivity is equal to emissivity, we have:
i.e. el = 1 – exp(– kl ×L) (spectral emissivity of gas...(13.87))
From Eq. 13.87, one can see that if gas layer thickness, L is very large,
al = el = 1
i.e. for very thick layers, radiation from the gas is equivalent to a black body radiation.
13.11.2 Gaseous Emission and Absorption
As mentioned earlier, gases are ‘selective’ absorbers and emitters, i.e. gases absorb or emit radiant energy only
within certain wavelength bands. Beyond these wavelength bands, these gases are transparent (or diathermic) to
thermal radiation. In thermal engineering, we are particularly interested in CO2 and H2O vapour, since these are
the main products of combustion of fuels.
Following wavelength bands are of importance for CO2 and H2O vapour:
For CO2:
Band 1: l = 2.40 to 3.80 microns
Band 2: l = 4.01 to 4.80 microns

712 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Band 3: l = 12.5 to 16.5 microns
For H2O vapour:
Band 1: l = 2.24 to 3.27 microns
Band 2: l = 4.80 to 8.5 microns
Band 3: l = 12.0 to 25 microns.
As discussed earlier, intensity of radiation decreases as it passes through a gas layer; this ‘attenuation’ in
intensity is proportional to the path length ‘L’ and the partial pressure ‘p’ of the gas (in a mixture of gases).
Emissive power of a gas is proportional to the gas temperature Tg and the product (p.L). From the experimental
data for CO 2 and H2O, following empirical relations for the emissive powers of CO2 and H2O have been sug-
gested:
1
FG T IJ 3.5
ECO 2 = 3.5×(p×L) 3 ×
H 100 K kcal/(m2hr.) ...(13.88)

×G
F T IJ 3.0
EH 2 O = 3.5×p0.8 ×L0.6
H 100 K kcal/(m2hr.) ...(1389)

where, p = partial pressure (atm) and L = layer thickness ( m).


(Note: 1 kcal/(m2hr) = 1.162 ´ 10–3 kW/m2).
For a diffuse surface, radiation is emitted in all directions. Therefore, path length L depends on direction and
shape of the body. For calculation purposes, a ‘mean path length of beam’ (L) is defined as follows:
FG V IJ m
L = 3.6×
H AK (mean path length...(13.90))

where, V is the volue of the body (m 3), and A is the surface area of enclosure (m2).
Emissivity of gases is a function of gas temperature Tg, total pressure p of the gas mixture, partial pressure pg
of the radiating gas and the mean path length, L.
Emissivity of water vapour (ew) in a mixture of
0.8 other gases which are non-radiating, at a total pressure
0.6
of 1 atm. are plotted, as a function of gas temperature
0.4 20 Tg and the product of partial pressure of water vapour
10
0.3 5 and the mean path length, (pw.L), in Fig. 13.41:
0.2 23 To determine emissivity of water vapour when the
1 total pressure is different from one atm., multiply the
0.6 value obtained from Fig. 13.41 by a correction factor
Emissivity, ew

0.1 0.4
(Cw), obtained from Fig. 13.42:
0.08
0.2 Similarly, Fig. 13.43 shows a plot of emissivity of
0.06
0.1 carbon dioxide gas in a mixture of other gases which
0.04 are non-radiating, at a total pressure of 1 atm. and the
0.06
0.03 0.04 Fig. 13.44 shows correction factor Cc for emissivity of
0.02 carbon dioxide, when the total pressure is other than 1
0.02 atm.
0.015
0.01 When water vapour and carbon dioxide appear to-
0.01 0.007
0.008 gether in a mixture of other non-radiating gases, total
pwL = 0.005 ft-atm gas emissivity (eg) is expressed as:
0.006
300 600 900 1200 1500 1800 2100 eg = e w + e c – De ...(13.91)
Gas temperature, Tg(K)
In Eq. 13.91, De is the correction factor, read from
FIGURE 13.41 Emissivity of water vapour in a Fig. 13.45. Note that total emissivity is less than the
mixture of other gases which are non-radiating, sum of the individual emissivities of water vapour and
at a total pressure of 1 atm. (Source: Incropera, Frank carbon dioxide because of mutual absorption of radia-
P. and David P. Dewitt [1998]. Fundamentals tion between these two gases.
of Heat and Mass Transfer. Pub.: John Wiley & Sons) Mean path length (L) to be used in Figs. 13.41 to
13.45, for various geometries, are given in Table 13.6:

RADIATION 713
1.8 pwL= 0 0.05ft-atm
0.25
1.6 0.50
1.0
1.4 2.5
5.0

Pressure correction, Cw
1.2 10.0

1.0

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1.0 1.2
(pw + p)/2 (atm)

FIGURE 13.42 Correction factor for emissivity of water vapour when the total pressure of mixture is other
than 1 atm. (Source: Incropera and Dewitt [1998]. op. cit.)

0.3
pcL = 4.0 ft-atm
0.2 2.0
1.0
0.8
0.4
0.1 0.2
0.08 0.1
0.06
0.06
0.04
Emissivity, ec

0.04 0.02
0.03
0.01
0.02 0.006
0.005
0.004
0.01 0.003
0.008
0.006 0.002
0.004
0.003 0.001
0.002

0.001
300 600 900 1200 1500 1800 2100
Gas temperature, Tg(K)
FIGURE 13.43 Emissivity of carbon dioxide in a mixture of other gases which are non-radiating, at a total
pressure of 1 atm. (Source: Incropera and Dewitt [1998]. op. cit.)

Once the emissivity (eg) of the gas mass in the given geometry is determined, we can proceed to find out the
radiant heat transfer from the gas mass to the surface of enclosure:
If the surface is black: Radiation emitted by the gas mass is completely absorbed by the black surface; black
surface also emits radiation which, in turn, is absorbed by the gas depending upon its absorptivity. Therefore, the
net radiant heat exchange between the gas mass at a temperature Tg and the surface at a temperature Ts is:
Q net = As ×s×(eg ×Tg4 – ag ×Ts4) ...(13.92)

714 FUNDAMENTALS OF HEAT AND MASS TRANSFER


2.0

1.5

Pressure correction, Cc
1.0
0.8 pcL = 2.5 ft-atm
0.6 1.0
0.5
0.5 0.25
0.4 0.12
0.05
0-0.02
0.3
0.05 0.08 0.1 0.2 0.3 0.5 0.8 1.0 2.0 3.0 5.0
p (atm)

FIGURE 13.44 Correction factor for emissivity of carbon dioxide when the total pressure of mixture is other
than 1 atm. (Source: Incropera and Dewitt [1998]. op. cit.)

0.07
Tg ³ 930°C
0.06 Tg = 125°C Tg = 540°C L(pw + pc) =
L(pw + pc) = L(pw + pc) =
Mixture correction, De

5 ft-atm
0.05 5 ft-atm 5 ft-atm 3
2
0.04 1.5
1.0
3 0.75
0.03 0.75 3
0.5 2
2
0.02 0.3 1.5 0.50
0.2 1.5 1.0
0.75
0.01 1.0
0.50 0.30
0.30
0 0.20 0.20
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
pw pw pw
pc + pw pc + pw pc + pw

FIGURE 13.45 Correction factor for mixtures of carbon dioxide and water vapour. (Source: Incropera and
Dewitt [year]. op. cit.)

The absorptivity, ag for water vapour and carbon dioxide is calculated as follows:
For water vapour:

FT I 0 . 45
Fp × L × Ts I
aw = C ×G J
HT K
w
g

s
GH
×ew Ts , w
Tg
JK ...(13.93)

For carbon dioxide:

FT I 0 . 65
F p × L ×Ts I
ac = Cc × GH T JK
g

s
GH
×ec Ts , c
Tg
JK ...(13.94)

Correction factors Cw and Cc are obtained from Figs. 13.42 and 13.44, respectively. Emissivities ew and ec are
obtained from Figs. 13.41 and 13.43, respectively, however, replacing Tg by Ts in the x-axis, and replacing (pw.L)
or (pc.L) by {pw.L.(Ts/Tg)} or {pc.L.(Ts/Tg)}, respectively.
When both water vapour and carbon dioxide are present in gas mixture, total gas absorptivity, (ag) is ob-
tained as:

RADIATION 715
TABLE 13.6 Mean beam lengths for various gas geometries

Geometry Characteristic dimension Mean path length (L)


Cylinder (height = diameter), Diameter, D 0.60 D
radiating to whole surface
Cylinder (height = diameter), Diameter, D 0.71 D
radiating to centre of base
Cylinder (height = 0.5 diameter), Diameter, D
radiating to:
(a) end 0.43 D
(b) side 0.46 D
(c) whole surface 0.45 D
Sphere, radiating to entire surface Diameter, D 0.65 D
Hemisphere, radiating to element Radius, R R
in centre of base
Cube, radiating to any face Edge, L 0.60 L
Two infinite planes Separation distance, L 1.8 L
Bank of tubes, diameter = D,
distance between surfaces of tubes = x:
(a) triangular arrangement, x = D (2.8) x
(b) triangular arrangement, x = 2D (3.8) x
(c) square arrangement, x = D (3.5) x
Arbitrary shape of volume, V Volume to area ratio, (V/A) 3.6(V/A)
(radiation to surface of area, A)

a g = aw + ac – Da ...(13.95)
where, Da = De is obtained from Fig. 13.45.
If the surface is grey: This is the most probable case, since with passage of time, enclosure walls will get dirty,
and the surface emissivity es becomes less than unity. However, effective emissivity of the surface es-eff in the
presence of gas mass is greater than es ; for es = 0.8 to 1.0, we have the approximate formula for e s-eff:
(e s + 1)
es_eff = ...(13.92)
2
Then, the net radiant heat exchange between the gas mass at a temperature Tg and the surface at a tempera-
ture Ts is given by:
Q net = es_eff ×As ×s×(eg ×Tg4 – ag ×Ts4) ...(13.93)
Radiation from flames Flame is produced during combustion (of a fuel). Radiation from flames occurs in fur-
naces, jet engine burners, etc. Flames may be luminous or non-luminous. Flames produced by household stoves
(burning kerosene or wood) are not luminous. Luminous flames have glowing particles of carbon, soot and
flying ash, and involve high temperatures. Radiation from the flame, obviously, depends on the emission of
particles contained in the flame, which in turn, depends on the kind of fuel burnt, mode of combustion, design of
the furnace, amount of air introduced, etc. Net radiation heat exchange between a flame and its enclosure is
given by:
Q net = s ×Af ×Ffw × ef × ew × (Tf4 – Tw4) ...(13.94)
where, Af is the area of the flame envelope, subscripts ‘f’ and ‘w’ refer to the flame and wall surface, respectively.
‘Effective flame temperature, Tf,’ (in Kelvin) is generally calculated as the geometric mean of the theoretical
temperature of combustion T1 and the temperature of combustion products, T2, at the furnace outlet.
1
i.e. e j
Tf = T12 ×T22 4 K ...(13.95)
Approximate values of flame emissivity (ef) for flames of different fuels are given in Table 13.7:
Example 13.31. A spherical chamber of 0.8 m diameter is filled with a gas mixture at 1 atm. and is at 1500 K. The gas
mixture contains 20% CO 2 by volume, and the rest of the mixture is non-radiating gases. Determine the emissivity of the
gas body.

716 FUNDAMENTALS OF HEAT AND MASS TRANSFER


TABLE 13.7 Flame emissivity (ef) for an infinitely thick layer

Kind of flame Flame emissivity, ef


Non-luminous gas flame (or, anthracite 0.40
in grate stoker combustion)
Luminous flame of pulverised anthracite 0.45
Luminous flame of lean coal 0.60
Luminous flame of coal with large volatile 0.70
content (brown coal, peat, etc., burned in
a layer or pulverised)
Luminous masut flame 0.85

(b) If the volume is filled to a pressure of 3 atm., but with the fraction of CO 2 still being 20%, what will be the value
of emissivity of gas body?
Solution. This is a spherical gas body. From Table 13.6, we see that for a spherical body, the mean path length of beam
is 0.65 D, where D is the diameter of sphere.
D := 0.8 m Tg := 1500 K L := 0.65 D i.e. L = 0.52 m p := 1 atm. pc := 0.2 atm.

Therefore, pc ×L = 0.104 m. atm.


i.e. pc ×L = 0.104 ´ 3.28 ft. atm.
i.e. pc ×L = 0.341 ft. atm.
Now, refer to Fig. 13.43. For Tg = 1500 K, and pc ×L = 0.341 ft. atm., we read:
ec = 0.09 (emissivity of carbon dioxide = emissivity of gas mixture)
(b) When the total pressure is 3 atm., with volume fraction of CO2 being 20%:
Now, L remains the same, but pc will be:
pc = 0.2 ´ 3 atm.
i.e. pc = 0.6 atm.
Then, pc ×L = 0.312 m. atm.
i.e. pc ×L = 0.312 ´ 3.28 ft. atm.
i.e. pc ×L = 1.023 ft. atm.
Now, refer to Fig. 13.43. For Tg = 1500 K, and pc ×L = 1.023 ft. atm., we read:
ec = 0.14 (emissivity of carbon dioxide.)
However, this value of emissivity is for a total pressure of 1 atm. In the present case, total pressure is 3 atm.
Therefore, obtained value of 0.14 has to be multiplied by a correction factor, read from Fig. 13.44. We get, from Fig.
13.44, for total pressure, p = 3 atm. And, pc ×L = 1.023 ft. atm.,
Cc = 1.35 (correction factor)
Therefore, emissivity of CO2 when the mixture pressure is 3 atm.:
ec = 0.14 ´ 1.35
i.e. ec = 0.189 (emissivity of CO2 when the mixture pressure is 3 atm.)

13.12 Solar and Atmospheric Radiation


We give a brief introduction to this fascinating topic because of its importance in the context of the ‘energy crisis’
and the resulting interest in ‘renewable energy sources’; further, this is a topic which affects our daily life.
Energy emitted by the sun is known as ‘solar energy’. Inexhaustible energy of sun is produced as a result of
nuclear ‘fusion’ reaction between two hydrogen atoms to form one atom of helium. ‘Atmospheric radiation’ is
the radiation emitted or reflected by the constituents of the atmosphere.
Sun is a spherical body of diameter D = 1.39 ´ 109 m and is located at a mean distance of L = 1.50 ´
1011 m from the earth. Even though the sun radiates an enormous amount of energy, only less than a billionth of
this energy reaches the earth’s surface. Solar radiation travels through the vacuum of space till it encounters
earth’s atmosphere. By conducting experiments with high altitude aircraft or balloons, and spacecrafts, scientists
have shown that average value of solar energy reaching the upper surface of earth’s atmosphere is about 1353
W/m2. This value is known as solar constant, Gs. The solar constant is the rate at which solar energy is incident
on a surface normal to the sun’s rays at the outer edge of the atmosphere when the earth is at its mean distance
from the sun. Since the earth moves in an elliptical orbit around the sun, this mean distance (L) varies with the

RADIATION 717
Sun, radius = r position of the earth and the value of Gs also varies; how-
Eb = s Tsun4 ever, the average value of Gs taken is 1353 W/m2. Con-
2
(4pr ).Eb stituents of the atmosphere absorb and/or scatter
radiations of different wavelengths contained in solar ra-
(4pL2).Ga diation. As a result, the amount of solar energy actually
reaching the earth’s surface is about 950 W/m2.
Outer edge of From the measured value of solar constant, we can
L Earth's easily determine the surface temperature of the sun. See
atmosphere
Fig. 13.46.
We use the condition that total energy radiated by
Earth
the sun (considered as a black body) must be equal to the
energy passing through the surface of a sphere whose ra-
FIGURE 13.46 Estimation of surface temperature dius is equal to the mean distance between the sun and
of sun when the solar constant is known the earth (= L), i.e.
(4×p×r 2)×s×Tsun4 = (4×p×L2)×Gs ...(13.96)
where, r = radius of the sun, and L = mean distance between the sun and earth. By this method, effective surface
temperature of the sun is determined to be 5762 K.
Solar energy incident on earth’s surface consists of two parts: direct solar radiation, GD (which reaches the
surface without any attenuation in the atmosphere) and diffuse solar radiation Gd (scattered radiation coming
uniformly from all directions). Then, total solar energy incident on a horizontal surface is:
Gsolar = GD ×cos (q) + Gd W/m2 ...(13.97)
where, q is the angle between the sun’s rays and the normal to the surface.
Constituents of the atmosphere absorb/scatter some of the solar radiation, as already mentioned; in addi-
tion, they also emit radiation. Main constituents contributing to this ‘atmospheric radiation’ are CO2 and H2O
molecules. Effective sky temperature, Tsky, is calculated assuming the atmosphere to be a blackbody, i.e.
G sky = s×Tsky4 W/m2 ...(13.98)
Value of Tsky varies from 230 K to 285 K, depending on the atmospheric conditions.
Sky radiation absorbed by a surface is:
E sky_absorbed = a×Gsky = a ×s×Tsky = e×s×Tsky4 W/m2 ...(13.99)
For a surface at temperature Ts, exposed to both solar and atmospheric radiation, net rate of heat transfer to
the surface is:
q net_rad = SEabsorbed – SEemitted
i.e. q net_rad = (as ×Gsolar + e×s×Tsky4) – e×s×Ts4
i.e. q net_rad = as ×Gsolar + e×s×(Tsky4 – Ts4) W/m2 ...(13.100)
Remember that incident solar energy coming from the sun originates at a very high temperature, and there-
fore, its spectral distribution is concentrated on short wavelength region; however, radiation emitted by the sur-
face is from a relatively low temperature, and its spectral distribution is concentrated at infra-red region. This
means that radiation properties (such as absorptivity and emissivity) for a surface are quite different for incident
and emitted radiations. Table 13.8 lists values of solar absorptivity, as and emissivity e (at 300 K) for some com-
mon materials. Obviously, solar collectors, widely used in solar energy applications, must be made of materials
having high a s and low e.

TABLE 13.8 Solar absorptivity (as) and emissivity (e) at room temperature for a few surface

Surface as e
Aluminium
Polished 0.09 0.03
Anodized 0.14 0.84
Foil 0.15 0.05
Copper
Polished 0.18 0.03
Tarmished 0.65 0.75
Contd.

718 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.
Stainless steel
Polished 0.37 0.60
Dull 0.50 0.21
Concrete 0.60 0.88
White marble 0.46 0.95
Red brick 0.63 0.93
Aspholt 0.90 0.90
Black paint 0.97 0.97
White paint 0.14 0.93
Snow 0.28 0.97
Human skin 0.62 0.97

13.13 Summary
Radiation heat transfer is unique as compared to other two modes of heat transfer, namely, conduction and
convection, in the sense that no medium is required for radiation heat transfer to occur. Radiation involves elec-
tromagnetic waves of all wavelengths, ranging from zero to infinity. All bodies at temperatures above zero Kel-
vin emit radiation; our interest in this chapter has been on ‘thermal radiation’, i.e. radiations in the wavelength
range of 0.1 to 100 microns.
After studying fundamental laws governing radiation heat transfer, we studied radiation properties of sur-
faces, such as absorptivity (a), emissivity (e) and transmissivity (t), since these properties affect the radiation heat
transfer.
Radiation heat transfer between surfaces is also dependent on the relative size and orientation of the sur-
faces. This is taken care of in calculations by introducing the concept of ‘view factor’. Analytical relations for
view factor are available only for simple geometries, and mostly, graphical solutions, available in heat transfer
hand-books, have to be referred to. Analytical relations and graphs for view factors for some of the commonly
required geometries have been given. ‘View factor algebra’ enables one to get view factors for some complicated
geometries, by ‘breaking down’ these geometries into simpler geometries for which values of view factors are
either already known or tabulated.
Next, radiation heat transfer between surfaces in two-surface and three-surface enclosures were considered,
using the radiation network method. This method greatly simplifies the analysis and gives a ‘physical feel’ of the
problem. Important practical examples of two-surface enclosure are: two infinite, parallel planes, long concentric
cylinders and concentric spheres. Furnaces with re-radiating (insulated) surfaces are examples of three-surface
enclosure.
‘Radiation shielding’ to reduce the radiation heat transfer between surfaces was studied next. Importance of
radiation shielding in reducing the radiation error in temperature measurement was studied.
Radiation has to be generally considered when the operating temperature level is high; as a rule, it will be
prudent to check its relevance in problems involving natural convection and forced convection at high tempera-
tures. Typical example is heat transfer from walls and doors of furnaces. In such problems, concept of ‘radiation
heat transfer coefficient’ simplifies the numerical calculations.
Finally, after giving a brief introduction to radiation heat transfer from gases, vapours and flames, we made
a mention of solar and atmospheric radiation, in view of its importance in the context of renewable energy
sources.

Questions
1. What is meant by ‘thermal radiation’? To which part of electromagnetic spectrum it belongs?
2. What is ‘visible light’? To which part of electromagnetic spectrum it belongs?
3. A local radio station broadcasts radio waves at a wavelength of 480 m. What is the frequency of those radio
waves?
4. Define: absorptivity, reflectivity and transmissivity. [M.U.]
5. Explain the following: (i) Black body and Grey body (ii) Specular reflector and Diffuse reflector (iii) Radiosity
and Irradiation. [M.U.]
6. State Planck’s law of monochromatic radiation. What is its significance? [M.U.]

RADIATION 719
7. State and explain Kirchhoff’s law of radiation. [M.U.]
8. State Wein’s law of displacement and prove that monochromatic emissive power of a black body is maximum
when lm.T = 2900 mmK. [M.U.]
9. What is intensity of radiation? Prove that total emissive power is p times the intensity of radiation. [M.U.]
10. Explain what is meant by ‘Greenhouse effect’.
11. What is meant by ‘view factor’? When is the view factor of a surface to itself equal to zero?
12. Write a short note on properties of view factor. [M.U.]
13. Explain ‘crossed-strings method’ of finding out view factors. When is it applicable?
14. Derive a general equation to find out the view factor of any cavity w.r.t. itself.
15. What is meant by ‘view factor algebra’? When is it resorted to?
16. Write a short note on ‘electrical network method’ to determine radiant heat exchange between grey surfaces.
17. What is a ‘radiation shield’? When is it used?
18. What is ‘radiation error’ in temperature measurement? Explain how radiation error can be reduced by the use of
radiation shields.
19. How is radiation from a gas mass different from radiation from a solid?
20. What is ‘mean path (or beam) length’?
21. How do you find out the emissivity of a gas mass containing carbon dioxide or/and water vapour, the mixture
pressure being one atmosphere?
22. What is ‘solar constant’? How is the effective surface temperature of sun determined when the value of solar
constant is known?
23. What is meant by ‘effective sky temperature’?
24. Why is solar absorptivity of a given surface quite different from its absorptivity for radiation from other sur-
rounding bodies?

Problems
1. A hole of area dA = 2 cm2 is opened on the surface of a large spherical cavity whose inside is maintained at 900
K. Calculate: (a) the radiation energy streaming through the hole in all directions into space, (b) the radiation
energy streaming per unit solid angle in a direction making a 45 deg. angle with the normal to the surface of the
opening.
2. The temperature of a body of area 0.1 m2 is 700 K. Calculate the total rate of energy emission, intensity of
normal radiation in W/(m2sr), maximum monochromatic emissive power, and wavelength at which it occurs.
3. Treating sun as a black body with a surface temperature of 5800 K, determine the rate at which infra-red radia-
tion (l = 0.76 – 100 mm) is emitted by the sun.
4. Filament of an incandescent light bulb is at 2800 K. Treating it as a black body, determine the fraction of the
radiant energy emitted by the filament that falls in the visible range. Also, find out at what wavelength is the
emission of radiation from the filament becomes maximum.
5. Window glass transmits radiant energy in the wavelength range 0.4 mm to 2.5 mm. Determine the rate of radiant
energy which is transmitted, through a glass window of size: 2 m ´ 2 m, when the black body source tempera-
ture is: (a) 5800 K (i.e. sun’s surface temperature), and (b) 1000 K .
6. Spectral emissivity of a particular surface at 900 K is approximated by a step function, as follows: e 1 = 0.3 for
l = 0 to 2 mm, e2 = 0.6 for l = 2 to 10 mm, and e 3 = 0.3 for l = 10 mm to ¥. Calculate (i) average emissivity of the
surface, and (ii) rate of radiation emission from the surface.
7. Two diffuse surfaces, a small disk of area A 1 and a large disk of area A2, are parallel to each other and directly
opposed, i.e. a line joining their centres is normal to both the surfaces. The large disk has a radius R and is
located at height L from the smaller disk. Obtain an expression for the view factor of small disk w.r.t. the large
disk. [M.U.]
8. Find out the net heat transferred between two circular disks 1 and 2, oriented one above the other, parallel to
each other on the same centre line as shown in Fig. 13.18. Disk 1 has a radius of 0.6 m and is maintained at 900
K, and disk 2 has a radius of 0.7 m and is maintained at 600 K . Assume both the disks to be black surfaces.
9. Find out the net heat transferred between two aligned parallel rectangles, as shown in Fig. 13.18. (X = 1 m, Y =
1.5 m and L = 1.5 m). Surface 1 is maintained at 600 K, and surface 2 is maintained at 1000 K . Assume both the
surfaces to be black surfaces.
10. Find out the net heat transferred between the areas A2 and A3 shown in Fig. Example 13.10 (See text for the
figure). Area 1 is maintained at 700 K, and area 2 is maintained at 400 K. Assume both the surfaces to be black.
11. Determine the view factor from the side surface to the base of a cylindrical enclosure whose height is twice its
diameter.

720 FUNDAMENTALS OF HEAT AND MASS TRANSFER


12. Determine the view factors from the base of a cube to each of the other five surfaces.
13. Find out the view factor from the dome of a hemispherical furnace to its circular base.
14. Find out the view factor (Fij ) between the plates i and j shown in Fig. Example 13.17a. Given: wi = 1 m, wj = 2 m
and L = 0.70.
15. A 0.3 m ´ 0.3 m ingot, 1.2 m in height, at a temperature of 1000 deg.C, is taken out of a furnace and rests on the
floor of a foundry room. Assuming that the surroundings are at a temperature of 30 deg.C, and the emissivity of
the surface of the ingot to be 0.8, calculate the net radiant heat loss from the ingot.
16. A spherical liquid oxygen tank, 0.3 m in diameter is enclosed concentrically in a spherical container of 0.4 m
diameter and the space in between is evacuated. The tank surface is at –183°C and has an emissivity = 0.2. The
container surface is at 25°C and has an emissivity = 0.25. Determine the net radiant heat transfer rate. [M.U.]
17. A hemispherical furnace of radius 1.6 m has a roof temperature of T1 = 900 K and emissivity 0.8. The flat circular
floor has a temperature of 500 K and emissivity of 0.5. Calculate the net radiant heat exchange between the roof
and floor. [M.U.]
18. Three thin-walled, long, circular cylinders 1, 2 and 3, of diameters 20 cm, 30 cm and 40 cm, respectively, are
arranged concentrically. Temperature of cylinder 1 is 100 K and that of cylinder 3 is 300 K. Emissivities of
cylinders 1, 2 and 3 are 0.05, 0.1 and 0.2, respectively. Assuming that there is vacuum inside the annular spaces,
determine the steady state temperature attained by cylinder 2.
19. A long pipe, 50 mm diameter passes through a room and is exposed to air at 20°C. The pipe surface temperature
is 93°C. Assuming that the emissivity of pipe surface is 0.6, calculate the radiation heat loss per metre length of
the pipe. [M.U.]
20. Calculate the net radiant heat interchange per square metre for two large parallel plates maintained at 800°C
and 300°C. The emissivities of two plates are 0.3 and 0.6, respectively.
21. Pipe carrying steam, OD = 20 cm, is exposed in a large room at 30°C. Pipe surface temperature = 400°C and
emissivity of pipe surface is 0.8. Calculate heat loss by pipe by radiation. What would be rate of loss of heat if
pipe is enclosed in a 40 cm diameter brick conduit of emissivity 0.91?
22. A blind cylindrical hole of diameter and length 3 cm is drilled into metal slab having emissivity 0.6. If the metal
slab is maintained at temp 350°C, find the rate of heat escaping out of the hole by radiation. [M.U.]
23. Calculate the radiation heat transfer from a hemispherical cavity if inside temperature is 800 K and its emissivity
is 0.6. Diameter of cavity is 500 mm.
24. A hohlraum is to be constructed out of a thin copper sphere of diameter = 20 cm. Its internal surface is highly
oxidised. What should be the area of a small opening to be made on the surface of the sphere, if the desired
absorptivity is 0.95?
25. A long duct of equilateral triangular section, of side w = 1.0 m, shown in Fig. Example 13.22, has its surface 1 at
600 K, surface 2 at 1100 K, and surface 3 is insulated. Further, surface 1 has an emissivity of 0.8 and surface 2 is
black. Determine the rate at which energy must be supplied to surface 2 to maintain these operating conditions.
26. Two co-axial cylinders of 0.5 m and 1 m diameter are 1.2 m long. The annular top and bottom surfaces are well
insulated and act as re-radiating surfaces. The inner surface is at 1100 K and has an emissivity of 0.6. The outer
surface is maintained at 500 K and its emissivity is 0.4.
(i) Determine the heat exchange between the surfaces
(ii) If the annular base surfaces are open to the surroundings at 300 K, determine the radiant heat exchange.
(Hint: If the outer cylinder is surface 2, first determine F21 and F22).
27. Two parallel plates, 0.5 m ´ 1 m each, are spaced 0.5 m apart. The plates are at temperatures of 900°C and 600°C
and their emissivities are 0.2 and 0.5, respectively. The plates are located in a large room, the walls of which are
at 25°C. The surfaces of the plates facing each other only exchange heat by radiation. Determine the rates of heat
lost by each plate and heat gain of the walls by radiation. Use radiation network for solution.
Assume shape factor between parallel plates: F12 = F21 = 0.285.
28. A furnace is of the shape of a frustrum of a cone. Diameters of top and bottom surfaces are 5 m and 3 m,
respectively, and the height is 3 m. Bottom surface is maintained at 1000°C and the top surface is at 600°C.
Emissivities of top and bottom surfaces are 0.8 and 0.9, respectively. Inclined side surface is refractory surface.
Find the radiation heat transfer from the bottom to the top surface and also the temperature of the inclined
surface.
29. Two very large parallel plates with emissivities 0.2 and 0.6 exchange heat. Find the percentage reduction in heat
transfer when two polished aluminium radiation shields (e = 0.3) are placed between them. Also, find the equi-
librium temperatures of the two shields.
30. Two large parallel planes facing each other and having emissivities 0.3 and 0.5 are maintained at 700°C and
500°C, respectively. Determine the rate at which heat is exchanged between the two surfaces by radiation. If a
radiation shield of emissivity 0.05 on both sides is placed parallel between the two surfaces, determine the
percentage reduction in the radiant heat exchange rate. What is the equilibrium temperature of the shield?
RADIATION 721
31. A spherical tank with diameter D1 = 30 cm filled with a cryogenic fluid at T1 = 90 K is placed inside a spherical
container of diameter D2 = 50 cm and is maintained at T2 = 300 K. Emissivities of inner and outer tanks are e1 =
0.10 and e 2 = 0.2, respectively. A spherical radiation shield of diameter D3 = 40 cm and having an emissivity e 3 =
0.05 on both surfaces is placed between the spheres. Calculate the rate of heat loss from the system by radiation.
Then, find the rate of evaporation of cryogenic liquid for hfg = 2.1 ´ 105 J/kg. What is the equilibrium tempera-
ture of the shield?
32. A double-walled flask may be considered as equivalent to two infinite parallel plates. The emissivities of walls
are 0.3 and 0.8, respectively. The space between the walls of the flask is evacuated. To reduce heat flow, a shield
of polished aluminium with emissivity equal to 0.04 (on both sides) is inserted between the two walls. Find the
percentage reduction in heat transfer. Also, find the equilibrium temperature of the shield. [M.U.]
33. The net radiation from the surface of two parallel plates maintained at temperatures T1 and T2 is to be reduced
to one-fifth. Calculate the number of screens to be placed between two surfaces to achieve this reduction in heat
exchange, assuming the emissivity of screens on both sides as 0.05 and that of surfaces as 0.2.
34. A 10 mm OD pipe carries a cryogenic fluid at 100 K. This pipe is encased by another pipe of 15 mm OD, and the
space between the pipes is evacuated. The outer pipe is at 300 K. Emissivities of inner and outer surfaces are 0.1
and 0.2, respectively. (a) Determine the radiant heat flow rate over a pipe length of 3 m. (b) If a radiation shield
of diameter 12 mm and emissivity 0.05 on both sides is placed between the pipes, determine the percentage
reduction in heat flow. (c) What is the equilibrium temperature of the shield?
35. Hot air is flowing in a duct whose walls are maintained at a temperature Tw = 500 K. A thermocouple placed in
the stream shows a reading of 800 K. If the emissivity of the thermocouple junction is e c = 0.8 and the convective
heat transfer coefficient between the flowing air and the thermocouple is h = 80 W/(m2 C), find out the true
temperature of the flowing stream. How much is the radiation error?
36. Hot air is flowing in a duct whose walls are maintained at a temperature Tw = 400 K. A thermocouple placed in
the stream shows a reading of 600 K. If the emissivity of the ther-mocouple junction is e c = 0.6 and the convec-
tive heat transfer coefficient between the flowing air and the thermocouple is h = 100 W/(m2 C), find out the true
temperature of the flowing stream. (b) Now, if a radiation shield (e s = 0.2) is placed between the thermocouple
and the walls, what will be new value of Tc read by the thermocouple? And, how much is the temperature
error? Take Ac/As = 0.1.
37. A spherical chamber of 1.5 m diameter is filled with a gas mixture at 1 atm. and is at 1200 K. The gas mixture
contains 18% CO2 by volume, and the rest of the mixture is non-radiating gases. Determine the emissivity of the
gas body.
(b) If the volume is filled to a pressure of 3 atm., but with the fraction of CO2 still being 18 %, what will be the
value of emissivity of gas body?
38. A cubical furnace of 2 m side, contains a gas mixture at 1500 K at a total pressure of 2 atm. The gas mixture
contains 15% of CO2 and 10% of H2 O by volume. If the furnace walls are at a temperature of 600 K, find out the
heat transferred by radiation from the gases to the walls. Assume that surfaces are black.

722 FUNDAMENTALS OF HEAT AND MASS TRANSFER


CHAPTER

14
Mass Transfer

14.1 Introduction
Mass transfer is an important topic with vast industrial applications in varied fields such as: mechanical, chemi-
cal and aerospace engineering, physics, chemistry, biology, etc. Few of the applications involving mass transfer
are:
(i) absorption and desorption (e.g. ammonia refrigeration systems)
(ii) solvent extraction
(iii) humidification (e.g. cooling towers and air-conditioning applications)
(iv) oxygenation of blood, food, etc.
(v) evaporation of petrol in internal combustion engines
(vi) neutron diffusion in nuclear reactors
(vii) distillation columns to separate components in a mixture.
Numerous every day applications such as dissolving of sugar in tea, drying of wood or clothes, evaporation
of water vapour into dry air, diffusion of smoke from a chimney into atmosphere, etc., are also examples of mass
diffusion.
Our aim in this introductory chapter on mass diffusion is, primarily, to show the similarity between the heat
transfer and mass transfer processes. For an in-depth study of this topic, one should consult specialised books in
the field.
Modes of mass transfer Mass transfer occurs whenever there is a concentration gradient between two fluids,
just as heat transfer occurs when there is a temperature gradient. Three modes of mass transfer may be distin-
guished:
(i) Molecular mass diffusion This occurs when mass transfer takes place in a fluid at rest, as a result of
concentration gradient and is analogous to diffusion heat transfer in conduction due to temperature gra-
dients.
(ii) Convective mass transfer This occurs when the fluid is in motion. Now, the effect of velocity field also
comes into picture. Mass transfer may be between a moving fluid and a surface or between two moving
fluids, which do not mix with each other. In fact, now, the mass transfer is by both by molecular diffu-
sion and convective motion of the fluid. This process is analogous to convective heat transfer process and
for low concentrations and low mass transfer rates, many of the equations for convective mass transfer
will be identical to those derived for convective heat transfer.
(iii) Mass transfer by change of phase Here, again, both convection and diffusion are involved. Boiling of
water in an open pan, evaporation of a cryogenic liquid from its container, diffusion of smoke from a
chimney, etc., are familiar examples.

14.2 Concentrations, Velocities and Fluxes


It is necessary to define a few terms:
14.2.1 Concentrations
Mass concentration (or mass density) Mass concentration or mass density rA of a species A in a mixture is
defined as mass of component A per unit volume of the mixture. It is expressed in kg/m3 units. Often, mass
concentration is also denoted by C; thus, CA = rA.
Molar concentration (or molar density) Molar concentration or molar density nA of a species A in a mixture is
defined as the number of moles of component A per unit volume of the mixture. It is expressed in kg moles/m3
units.
These two concentrations are related to each other as follows:
rA
nA = ...(14.1)
MA
where, MA = molecular weight of species A.
Mass fraction It is the ratio of mass concentration of species A to the total mass density of the mixture, i.e.
rA
wA = ...(14.2)
r
Mole fraction It is the ratio of molar density of species A to the total molar density of the mixture, i.e.
nA
yA = ...(14.3)
n
For a binary mixture of two components, A and B, we have, by definition:
rA + rB = r ...(14.4)
nA + nB = n ...(14.5)
wA + wB = 1 ...(14.6)
yA + yB = 1 ...(14.7)
Ideal gas mixtures At low pressures, a gas or gas mixture can be considered as an ideal gas. Familiar example
of such a case is the mixture of dry air and water vapour existing under atmospheric conditions. Then, by
Dalton’s law, total pressure of the mixture is equal to the sum of the partial pressures of each component, and is
given by:
P = S Pi ...(14.8)
Here, Pi is the partial pressure of component i and it is the pressure that would be exerted by the component
i if it alone occupied the whole mixture volume. Then, using the ideal gas relation (i.e. P.V = n.R u.T, where Ru =
universal gas constant = 8314 J/kg mole K), we can write:
N i × Ru × T
Pi V N
= = i = yi ...(14.9)
P N × Ru ×T N
V
i.e. pressure fraction is equal to the mole fraction.
14.2.2 Velocities
Mass diffusion may occur in a stationary medium or a moving medium. In a staionary medium, the components
in a mixture move because of the concentration gradients only and the velocity of each species is equal to the
‘diffusion velocity’ only. However, if the medium is also moving, then, the absolute velocity of a species is equal
to the sum of the bulk flow velocity and the diffusion velocity.
Remembering that, m = r.V.A, we can write for a mixture of two components A and B:
m = mA + mB
i.e. r×V mass ×A = rA ×V A ×A + rB ×V B ×A
Therefore,
r A ×VA + r B ×VB
V mass = = wA ×V A + wB ×V B ...(14.10)
r
Here, V mass is called the mass–average velocity of flow.
In the case of a stationary medium, mass–average velocity is equal to zero.

724 FUNDAMENTALS OF HEAT AND MASS TRANSFER


When there is no concentration gradient, velocity of all species is equal to the mass–average velocity of flow;
and when there is a concentration gradient, average velocity of each component is given by:
V A = V mass + V diff_A ...(14.11a)
and, V B = V mass + V diff_B ...(14.11b)
Similarly, molar average velocity is defined by:

V molar = nA ×VA + nB ×VB = yA ×V A + yB ×V B ...(14.12)


n
And, molar average velocity of each component is given by:
V A = V molar + V diff_A ...(14.13a)
and, V B = V molar + V diff_B ...(14.13b)
From Eqs. 14.11 and 14.13, we can write:
Mass diffusion velocities of A and B:
V diff_A = V A – V mass ...(14.14a)
V diff_B = V A – V mass ...(14.14b)
Molar diffusion velocities of A and B:
V diff_A = V A – V molar ...(14.15a)
V diff_B = V A – V molar ...(14.15b)
14.2.3 Fluxes
For species A: Absolute flux = rA.V A.
Bulk motion flux = rA.V mass.
Diffusion flux = mA/A = mass flow per unit time per unit area.
Absolute flux of a component is as seen by a stationary observer.
It is equal to diffusion flux + bulk motion flux
i.e.
mA
rA ×V A = + rA ×V mass
A
mA
i.e. Diffusion flux = = rA ×V A – rA ×V mass = rA ×(V A – V mass) ...(14.16)
A
Similarly, on molar basis:
Diffusion flux = nA ×(V A – V molar) ...(14.17)

14.3 Fick’s Law of Diffusion


Consider a chamber, containing a mixture of two gases B and C, divided into two volumes by a partition in the
middle, as shown in Fig. 14.1.
To start with, let the gas mixture in volume B be Partition
rich in species B, and the volume C be rich in species
C. Now, if the partition is removed, molecules of B
would diffuse to the right, i.e. in the direction of de- B C
creasing concentration of B, and the molecules of C
would diffuse to the left. Lower part of Fig. 14.1
shows the concentration profiles of B and C shortly
after the partition is removed. After sufficient time Concentration of CB CC Concentration of
has elapsed, equilibrium conditions would be species B species C
achieved, i.e. uniform concentrations of B and C
would be attained and there would be no more mass
diffusion. Fick’s law relates the mass flux by diffu-
sion to the concentration gradient. It is stated as: ‘Dif-
fusion mass flux of a species through a medium is x
proportional to the concentration gradient”, i.e.
FIGURE 14.1 Fick’s law of diffusion for a binary
m dC
j = B = –D × B kg/(sm2)
B BC ...(14.18) gas mixture
A dx
MASS TRANSFER 725
where,
jB = mB/A = mass flux, kg/(sm2)
A = area normal to the line of propagation of mass, m2
CB = concentration of species B which is diffusing, kg/m3
dCB/dx = concentration gradient for species B
DBC = diffusion coefficient or ‘diffusivity’ for the binary mixture of B and C, m2/s
The negative sign in Eq. 14.18 signifies the fact that diffusion takes place in the direction of decreasing
concentration, so that mass flux is a positive quantity.
Observe that unit of diffusivity is : m2/s.
Molar flux would be obtained by simply dividing jB by the molecular weight of species B, i.e.
jB
JB = kg moles(sm2) ...(14.19)
MB
Note the close similarity of Fick’s law as given by Eq. 14.18 to the Fourier’s law of heat conduction and to the
Newton’s law of viscosity, i.e.
Q dT d
q= = – k× = a× (r×Cp ×T) (Fouier’s law of heat conduction)
A dx dx
du d
t = m× = n× (r×u) (Newton’s law of viscosity)
dy dy
We can state that:
Fourier’s law ...describes transport of energy due to temperature gradient
Newton’s law...describes transport of momentum due to velocity gradient, and
Fick’s law...describes transport of mass due to concentration gradient.
Further, units of mass diffusivity (D), thermal diffusivity (a ), and kinematic viscosity (n ) are all same, i.e.
m2/s.
Now, while dealing with perfect gases, we can express the concentration gradient appearing in Fick’s Eq. 14.18,
in terms of partial pressures, as follows:
For species B:
pB = rB ×RB ×T (Perfact gas law)
R ×T
i.e. pB = r B × u (where, Ru = Universal gas constant = 8314 J/(kg mole K))
MB
In the above, pB = partial pressure of species B,
rB = density of species B,
MB = the molecular weight of species B,
and, T is the absolute temperature in Kelvin.
Remembering that density is nothing but concentration, i.e. rB= CB, we write, substituting for rB in Eq. 14.18:
mB dC
jB = = – DBC × B
A dx

mB F
d pB × MB I
i.e. jB =
A
= – DBC × GH
dx Ru × T JK
mB MB dpB
i.e. jB = = – DBC × × ...(14.20)
A Ru × T dx
Similarly, for species C, we can write:
mC MC dpC
jC = = – DCB× × ...(14.21)
A Ru ×T dx
Note that Eqs. 14.20 and 14.21 are valid for isothermal conditions only.
Following points must be noted well in connection with the Fick’s law equation 14.18:

726 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(i) This law is valid for mass diffusion, only due to concentration gradient. It is not applicable to the case of
mass diffusion occurring due to other reasons such as pressure gradients, temperature gradients or other
external forces.
(ii) Fick’s law, like Fourier’s law, is derived as a result of experimental observations; it is not derived from
first principles analytically.
(iii) Mass diffusion occurs in the direction of decreasing concentration, just as heat transfer occurs in the
direction of decreasing temperature.
(iv) Diffusion coefficient (D) depends upon pressure, temperature and the nature of the component con-
cerned; however, it may be assumed as constant for ideal gases and dilute liquids for a given range of
temperature and pressure.
Diffusion coefficient (D):
For gases/gas mixtures:
From kinetic theory of gases, it can be shown that at ordinary pressures, diffusion coefficient is independent
of mixture composition, but increases with temperature and decreases with pressure, i.e. for a binary gas mixture
of two components B and C, we have:
3
F1 I
1
T2
×G JK
1 2

I HM
D = 0.0043× + cm2/s ...(14.22)
F 1 1
2
B MC
p ×G
GH Vb3 + Vc3 JJ
K
t

where,
pt = total pressure (atm.)
T = absolute temperature (K)
MB, MC = molecular weights of gas species
V b, V c = molecular volumes of B and C at normal boiling points, cm3/gm. mole
Molecular weights and molecular volumes of a few gases are given in Table 14.1.
Data in Table 14.1 is useful to estimate the diffusion in binary gas mixtures.
From Eq. 14.22, it is clear that:

FG T IJ ×FG P IJ
3
D1 1 2 1
D2
=
HT K H P K
2 2
...(14.23)

i.e. if the diffusion coefficient at a certain temperature and pressure are known, then the diffusion coefficient at
any other pressure and temperature can be estimated using Eq. 14.23. Note that the temperatures must be ex-
pressed in Kelvin.
Table 14.2 gives values of Diffusion coefficient and Schmidt numbers for a few substances diffusing through
air at 25°C and 1 atm. Eq. 14.23 may be used to get values of diffusion coefficient at any other desired tempera-
ture and pressure.

TABLE 14.1 Molecular weights and molecular volumes for a few gases

Gas Molecular weight Molecular volume at normal


boiling point (cm3/gm.mole)
Air 29 29.89
Ammonia (NH3) 17 25.81
Carbon dioxide (CO2) 44 34.00
Carbon monoxide (CO) 28 30.71
Hydrogen (H2) 2 14.28
Nitrogen (N2) 28 31.20
Oxygen (O2) 32 25.63
Sulphur dioxide (SO2) 64 44.78

MASS TRANSFER 727


TABLE 14.2 Mass diffusivity and Schmidt number for a few gases and vapours diffusing through air at 25°C
and 1 atm.

Substance Mass diffusivity, D(m2/s) Schmidt number, Sc = n /D


–4
Ammonia 0.280 ´ 10 0.78
Carbon dioxide 0.164 ´ 10–4 0.94
Hydrogen 0.410 ´ 10–4 0.22
Oxygen 0.206 ´ 10–4 0.75
Water 0.256 ´ 10–4 0.60
Methanol 0.159 ´ 10–4 0.97
Ethyl alcohol 0.119 ´ 10–4 1.30
Acetic acid 0.133 ´ 10–4 1.16
Benzene 0.088 ´ 10–4 1.76
Toluene 0.084 ´ 10–4 1.84

For the practically important case of diffusion of water vapour in air, following formula has been proposed
by Marrero and Mason:
T 2. 072
DH 2 O_air = 1.87 ´ 10–10 × m2/s ...280 K < T < 450 K...(14.24)
P
where, P is the total pressure in atm. and T is the temperature in Kelvin.
Values of D at 1 atm., as calculated from Eq. 14.24, are given in Table 14.3:

TABLE 14.3 Diffusion coefficient for water vapour in air at 1 atm.

T (deg.C) D (m 2/s)
0 2.09E-05
5 2.17E-05
10 2.25E-05
15 2.33E-05
20 2.42E-05
25 2.5E-05
30 2.59E-05
35 2.68E-05
40 2.77E-05
45 2.86E-05
50 2.96E-05
55 3.05E-05
60 3.15E-05
65 3.25E-05
70 3.35E-05
75 3.45E-05
80 3.55E-05
85 3.66E-05
90 3.77E-05
95 3.88E-05
100 3.98E-05
105 4.1E-05
110 4.21E-05
115 4.32E-05
120 4.44E-05
125 4.56E-05
Contd.

728 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Contd.

130 4.68E-05
135 4.8E-05
140 4.92E-05
145 5.05E-05
150 5.17E-05

For steady state diffusion through a non-diffusing, multi-component mixture, an ‘effective diffusivity’ is de-
fined as:
1
D= ...(14.25)
yb yc y
+ + d
Dab Dac Dad
where, yb, yc, yd, ... = mole fractions of components on a free basis
Dab, Dac, Dad , ... = diffusivities of species A through B, C, D ...
For dilute liquids:
Following semi-empirical relation is suggested to estimate the diffusion coefficient of dilute liquids:
T
F= ...(14.26)
DAB × m B
where,
T = absolute temperature (K)
DAB = diffusivity of solute A through a solvent B, (m2/s)
mB = viscosity of solvent B (centipoises), and
F = a function of molal volume of solute A, Ks/cm2.centipoise...determined from charts
Table 14.4 gives mass diffusivity values of a few liquids at 20°C with water as solvent:
Observe that mass diffusivity in liquids is much lower than in gases; therefore, diffusion in liquids occurs at
a much slower rate than in gases.
For solids:
Diffusion in solids occurs still at a lower rate as compared to that in gases and liquids. This is evident from
the Table 14.5, which gives mass diffusivity values for a few substances diffusing through some solids.

14.4 General Differential Equation for Diffusion in Stationary Media


This equation for mass transfer of any species is derived in a manner analogous to the derivation of general
differential equation for conduction.

TABLE 14.4 Mass diffusivity of liquids at 20 deg.C, with water as solvent

Solute D, (m2/s) Schmidt number, Sc = n/D


–9
Oxygen 1.80 ´ 10 558
Carbon dioxide 1.77 ´ 10–9 559
Ammonia 1.76 ´ 10–9 570
Hydrogen 5.13 ´ 10–9 196
Chlorine 1.22 ´ 10–9 824
Hydrochloric acid 2.64 ´ 10–9 381
Sulphuric acid 1.73 ´ 10–9 580
Acetic acid 0.88 ´ 10–9 1140
Ethanol 1.00 ´ 10–9 1005
Urea 8.06 ´ 10–9 946
Glucose 0.60 ´ 10–9 -

MASS TRANSFER 729


TABLE 14.5 Diffusion through solids at 1 atm.

Substance A (Solute) Substance B (Solvent) T, (K) DAB,(m 2 /s)


Carbon dioxide Natural rubber 298 1.1 ´ 10–10
Nitrogen Natural rubber 298 1.5 ´ 10–10
Oxygen Natural rubber 298 2.1 ´ 10–10
Helium Pyrex 773 2.0 ´ 10–12
Helium Pyrex 293 4.5 ´ 10–15
Helium Silicon dioxide 298 4.0 ´ 10–14
Hydrogen Iron 298 2.6 ´ 10–13
Hydrogen Nickel 358 1.2 ´ 10–12
Hydrogen Nickel 438 1.0 ´ 10–11
Cadmium Copper 293 2.7 ´ 10–19
Zinc Copper 773 4.0 ´ 10–18
Zinc Copper 1273 5.0 ´ 10–13
Antimony Silver 293 3.5 ´ 10–25
Bismuth Lead 293 1.1 ´ 10–20
Mercury Lead 293 2.5 ´ 10–19
Copper Aluminium 773 4.0 ´ 10–14
Copper Aluminium 1273 1.0 ´ 10–10
Copper Iron (fcc) 773 5.0 ´ 10–15
Carbon Iron (fcc) 1273 3.0 ´ 10–11

Consider a differential control volume in a


y
given stationary medium (i.e. mass average veloc-
ity of the mixture is zero) as shown in Fig. 14.2.
dz Assumptions:
(i) Species B is diffusing through a solid or
jB dy { jB + jB(¶jB/¶x).dx} through a stationary fluid medium C
(ii) Mass transfer is by concentration differ-
dx ence only, and influence of other effects
such as pressure, temperature or other
forces is negligible
x
(iii) Diffusivity is constant in all directions
(Isotropic).
Consider the mass balance of species B:
z
Along the X-direction:
Mass influx at the left surface = (mb/A).dy.dz
FIGURE 14.2 General differential equation for mass
diffusion Mass efflux at the right surface = {(mb/A) +
(¶/¶x) (mb /A).dx}.dy.dz
Therefore, accumulation of species B in the elemental volume due to diffusion in X-direction =
(mb/A).dy.dz – {(mb/A) + (¶/¶x)(mb/A).dx}.dy.dz = –(¶/¶x)(mb/A).dx.dy.dz
Similarly, accumulation of species B due to diffusion in the Y and Z directions are:
Y-direction: –(¶/¶y)(mb/A).dx.dy.dz
Z-direction: –(¶/¶z)(mb/A).dx.dy.dz
Therefore, net accumulation of species B =
– {(¶/¶x)(mb/A) + (¶/¶y)(mb/A) + (¶/¶z)(mb/A) }.dx.dy.dz ...(i)
Let qb be the generation of mass species per unit volume (say, due to some chemical reaction). Then, total
mass of species generated =
qb.dx.dy.dz ...(ii)
Now, net effect of the mass accumulation due to diffusion and mass generation in the volume is an increase
in the mass concentration of species B; this results in a time rate of change of mass concentration in the control
volume, and is given by:
730 FUNDAMENTALS OF HEAT AND MASS TRANSFER
(¶Cb/¶t).dx.dy.dz ...(iii)
Then, writing the mass balance,
– {(¶/¶x )(mb/A) + (¶/¶y)(mb/A) + (¶/¶z)(mb/A)}.dx.dy.dz + qb .dx.dy.dz =
(¶Cb/¶t).dx.dy.dz
Cancelling (dx.dy.dz) throughout and using Fick’s law, i.e. (mb/A) = –D.(¶Cb/¶x),
(¶Cb/¶t) = (¶/¶x){D.(¶Cb/¶x)} + (¶/¶y){D.(¶Cb/¶y)}
+ (¶/¶z){D.(¶Cb/¶z)} + qb ...(14.27)
Eq. 14.27 is the general differential equation for mass transfer by diffusion in a solid or stationary medium,
with constant D and with internal mass generation.
With no internal mass generation (qb = 0), and constant D, we have:
(¶Cb/¶t) = D.{(¶ 2Cb/¶x2) + (¶ 2Cb/¶y 2) + (¶ 2Cb/¶z2)} = D.Ñ 2Cb ...(14.28)
To obtain the concentration gradient, and the mass diffusion rate, Eq. 14.28 has to be solved with the appro-
priate boundary conditions.
Eq. 14.28 can be written as:
Ñ 2Cb = (1/D).(¶Cb/¶t) ...(14.29)
Observe that Eq. 14.29 is similar in form to the heat conduction equation, i.e.
Ñ 2T = (1/a).(¶T/¶t) (differential equation of heat conduction )
Boundary conditions generally encountered in practice are:
1. Specified concentrations at the boundary: Cb = Cb0 at x = 0 and Cb = CbL at x = L
2. Impermeable surface at the boundary: ¶Cb/¶x = 0 at x = 0
3. Specified mass flux at the surface: jb =(mb/A) = –Dbc.(¶Cb/¶x) at x = 0
4. Specified mass transfer coefficient (convective) at the surface: jb = hm.(Cbs – Cba) where hm = convective
mass transfer coefficient, Cbs = concentration in the fluid adjacent to the surface, and Cba = bulk concentra-
tion in the fluid stream.

14.5 Steady State Diffusion in Common Geometries Concentration profile


Now, we shall derive relations for concentration profiles and mass transfer rates
in a few simple geometries such as a plain membrane, cylindrical shell and Cb1
spherical shell. (mb/A)

14.5.1 Steady State Diffusion Through a Plain Membrane Cb2


Consider a plain membrane whose thickness L is small compared to its other di-
mensions, i.e. mass diffusion through this membrane can be considered as one-
L
dimensional. See Fig. 14.3.
x
Then, for steady state (i.e. (¶Cb/¶t) = 0) and one-dimensional diffusion, dif-
ferential equation 14.29 reduces to: FIGURE 14.3 Diffusion
2 through a plain membrane
d Cb
=0
dx2
dCb
Integrating, = C1
dx
Integrating again, Cb = C1 ×x + C 2 ...(a)
BCs: Cb = Cb1 at x = 0, and Cb = Cb2 at x = L
Cb2 - Cb1
Therefore, C 2 = C b1 and, C1 =
L
Subst. in Eq. a:
x
Cb = (Cb2 – Cb1)× + Cb1 ...(14.30)
L
Eq. 14.30 gives the concentration profile, which is a straight line as shown in Fig. 14.3.
Now, mass transfer rate is obtained by applying the Fick’s law, i.e.

MASS TRANSFER 731


mb dCb
= – Db ×
A dx
mb
= – Db×
d LMx
(Cb 2 - Cb 1 ). + Cb1
OP
i.e.
A dx NL Q
mb FG
C - Cb1
= – Db × b 2
IJ
i.e.
A H
L K
mb D
i.e. = + b ×(Cb1 – Cb2) ...(14.31)
A L
Eq. 14.31 gives the mass flux through the membrane.
Alternatively:
Since Db is constant, we can directly integrate the Fick’s law equation between x = 0 and x = L, after separating
the variables:
We have, from Fick’s law:
mb dCb
= –Db ×
A dx

z
Separating the variables and integrating,
mb
A
×
z
0
L
dx = –Db ×
Cb 1
Cb 2
dCb (assuming Db = constant )

mb D
i.e. = b ×(Cb1 – Cb2) (same as Eq. 14.31)
A L
Cb1 - Cb 2 Concentration potential
i.e. mb = =
L Diffusion resistance
Db × A
Note that the above equation gives diffusion mass flow rate (kg/s), which can be expressed in a form analo-
gous to Ohm’s law, i.e. as a ratio of concentration potential to the diffusion resistance.
Therefore, diffusion resistance for a plane membrane is given by:
L
R membrane = s/m3 ...(14.32)
Db × A
Concentration profile is obtained by using the fact that in steady state, the mass diffusion rate through any
section in the membrane must be constant. Let at any x, the concentration be Cb. Then, from Eq. 14.31,
mb D
= b ×(Cb1 – Cb)
A x
Equating this with Eq. 14.31,
mb D D
= b ×(Cb1 – Cb) = b ×(Cb1 – Cb2)
A x L
(Cb1 - Cb ) (Cb1 - Cb2 )
i.e. =
x L
x
i.e. Cb = (Cb2 – Cb1)× + Cb1 (same as Eq. (14.30))
L
Note that diffusion mass transfer and conduction heat transfer are analogous.
14.5.2 Steady State Diffusion through a Cylindrical Shell
Consider a cylindrical shell of length L and inner and outer radii equal to r 1 and r 2, respectively, as shown in Fig.
14.4. Let the corresponding concentrations of species B at these radii be Cb1 and Cb2.

732 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Assumptions: dr
1. Steady state, one-dimensional diffusion in the r direction only Cb2
r mb, kg/s
2. Constant D Cb1
3. No internal mass generation
Now, consider an elemental cylindrical shell at any radius r, with
a thickness dr.
r1
We have, from Fick’s law:
r2
mb dCb Concentration profile
= – Db ×
A dx
For this system we write:
mb dCb FIGURE 14.4 Diffusion through a
= – Db ×
2×p × r × L dr cylindrical shell
Separating the variables, and integrating from r1 to r2 :

mb ×
z
FG 1IJ dr = –D ×2×p×L×
H rK
r1
r2

b z Cb 2

Cb 1
dCb (assuming Db = constant )

Fr I
m ×ln G J = 2×p ×D ×L×(C – Cb2)
Hr K
2
i.e. b b b1
1

Cb 1 - Cb 2
i.e. mb =
FG r IJ kg/s ...(14.33)
ln
Hr K
2

2 × p × Db × L

Cb 1 - Cb 2
i.e. mb = kg/s ...(14.34)
Rdiff_cyl

FG r IJ
Hr K
2
ln
where, R diff_cyl = 1
s/m3 = diffusion resistance of cylindrical shell
2 × p × Db × L
To get concentration distribution with radius:
Integrating from r1 to r, (and concentration from Cb1 to Cb), Eq. 14.33 becomes:
Cb 1 - Cb
mb =
FrI
ln G J
kg/s ...(14.35)

Hr K 1

2 × p × Db × L
But, in steady state, since mass flow rate is same through each section of the cylindrical shell, equating Eqs.
14.33 and 14.35, we get:
Cb 1 - Cb 2 Cb 1 - Cb
Fr I
ln G J
=
FG r IJ
Hr K ln
Hr K
2

1 1

2 × p × Db × L 2 × p × Db × L

FG r IJ
Cb - Cb 1
ln
Hr K
Fr I
1
i.e. = ...(14.36)
Cb 2 - Cb1
ln G J
Hr K
2

MASS TRANSFER 733


Eq. 14.36 gives the concentration distribution in the cylindrical shell as a function of radius, r. Note that the
concentration distribution is logarithmic. This profile is also shown in Fig. 14.4.
Now, bringing in the concept of ‘log mean area’ for a cylindrical system, just as we did in the case of one-
dimensional conduction through a cylindrical shell (see Chapter 4), we can write:
Db × (Cb1 - Cb 2 )
mb = ×Am ...(14.37)
Dx

2 × p × L × (r2 - r1 )
where, Dx = (r2 – r1) and, Am =
FG r IJ = log mean area.
ln
Hr K
2

14.5.3 Steady State Diffusion through a Spherical Shell


Consider a spherical shell of inner and outer radii equal to r1 and r2, respectively, as shown in Fig. 14.4. Let the
corresponding concentrations of species B at these radii be Cb1 and Cb2.
Assumptions:
1. Steady state, one-dimensional diffusion in the r direction only
2. Constant D
3. No internal mass generation
Now, consider an elemental spherical shell at any radius r, with a thickness dr.
We have, from Fick’s law:
mb dC
= –Db × b
A dx
For this system:
mb dC
= –Db × b
4 ×p × r 2 dr

z
Separating the variables and integrating from r1 to r2:

mb ×
z
FG 1 IJ dr = –D ×4×p×
r1
r2

Hr K 2 b
Cb 1
Cb 2
dCb (assuming Db = constant)

F 1 1I
m × G - J = 4×p ×D ×(C – Cb2)
i.e. b
Hr r K 1 2
b b1

Cb1 - Cb 2
i.e. mb =
FG 1 - 1 IJ kg/s ...(14.38)

Hr r K
1 2

4 × p × Db

Cb 1 - Cb 2
i.e. mb = ...(14.39)
Rdiff_sph

FG 1 - 1 IJ
where, R diff_sph =
Hr r K
1 2
s/m3 = diffusion resistance of spherical shell
4 × p × Db
To get concentration distribution with radius:
Integrating from r1 to r, (and concentration from Cb1 to Cb), Eq. 14.38 becomes:
Cb 1 - Cb
mb =
FG 1 - 1IJ kg/s ...(14.40)

H r rK
1

4 × p × Db

734 FUNDAMENTALS OF HEAT AND MASS TRANSFER


But, in steady state, since mass flow rate is same through each section of the cylindrical shell, equating Eqs.
14.38 and 14.40, we get:
Cb1 - Cb 2 Cb 1 - Cb
FG 1 - 1 IJ =
FG 1 - 1IJ
Hr r K
1 2 H r rK 1

4 × p × Db 4 × p × Db

1 1
-
Cb - Cb 1 r r1
i.e. = ...(14.41)
Cb 2 - Cb1 1 1
-
r2 r1
Eq. 14.41 gives the concentration distribution in the spherical shell as a function of radius, r. Note that the
concentration distribution is hyperbolic.
Now, bringing in the concept of ‘mean area’ for a spherical system, just as we did in the case of one-dimen-
sional conduction through a spherical shell (see Chapter 4), we can write:
Db × (Cb1 - Cb 2 )
mb = ×Am ...(14.42)
Dx
where, Dx = (r2 – r1) and, Am = 4×p×r1 ×r2 = mean area.
Solubility factor S:
Species concentration at the gas–solid interface is generally stated in terms of partial pressure of gas adjoining the
interface and a ‘solubility factor, S’. Then,
Cb = S×pa ...(14.43)
where, pa = partial pressure, and, S = solubility.
Solubility data for selected gas solid combinations are given in Tabl 14.6.

TABLE 14.6 Solubility of selected gases and solids (For gas i, S = Ci, solid side/Pi, gas side)

Gas Solid T, (K) S,(kmol /m3 bar)


O2 Rubber 298 0.00312
N2 Rubber 298 0.00156
CO2 Rubber 298 0.04015
He SiO2 293 0.00045
H2 Ni 358 0.00901
(Note: Permeability, P = S × DAB where DAB = diffusivity of gas in solid.)

Summary Table Formulas for one-dimensional, steady state diffusion in simple geometries are summarised in
Table 14.7 below:
Example 14.1. Air is contained in a vessel at a temperature of 20°C and pressure of 2 bar. Assuming the partial pressures
of O2 and N2 to be in ratios of 0.21 and 0.79, respectively, calculate: (i) Molar concentrations (ii) Mass concentrations (i.e.
densities), (iii) Mass fractions, and (iv) Molar fractions.
Solution.
Data:
pO 0.21
T := 20 + 273 K p := 2× 105 Pa 2
:= Ru := 8314 J/kg mole K MO 2 := 32 M N2 := 28
pN 2
0.79
(i) Molar concentrations
pO
nO2 = 2

Ru × T

0. 21× p
i.e. nO2 :=
Ru × T

MASS TRANSFER 735


TABLE 14.7 One-dimensional, steady state diffusion in simple geometries

Geometry Mass flow rate (kg /s) Rdiffusion (s /m3) Concentration distribution

C b1 - C b 2 L C b - C b1 x
Plane membrane mb = =
L Db × A Cb 2 - Cb 1 L
Db × A

Fr I FrI
Cb 1 - C b 2
ln GH r JK
2
C b - C b1
ln GH r JK
1

Fr I Fr I
1
Cylindrical shell mb = =
2 × p ×D b ×L Cb 2 - Cb 1
ln G J ln G J
Hr K Hr K
2 2

1 1
2 × p × Db × L

F 1 - 1I 1 1
C - Cb 2 GH r r JK C b - C b1
-
r r1
FG IJ
1 2
Spherical shell mb = b 1 =
1 1 4 × p ×D b Cb 2 - Cb 1 1 1
-
H
r1 r2
-
K r2 r1
4 × p ×Db

i.e. nO2 = 0.017 kg mole/m3 (molar concentration of Oxygen.)


Similarly,
pN
nN 2 = 2

Ru × T

79× p
i.e. nN 2 :=
Ru × T
i.e. nN 2 = 0.065 kg mole/m3 (molar concentration of Nitrogen.)
(ii) Mass densities
We have the relation between molar and mass densities:
rA
nA = ...(14.1)
MA
Therefore,
r O2 := MO2 × nO2
i.e. r O2 = 0.552 kg/m3 (mass density of Oxtygen.)
Similarly,
r N2 := MN2 × nN2
i.e. r N2 = 1.816 kg/m3 (mass density of Nitrogen.)
(iii) Mass fractions
Total mass density, r := r O2 + r N2
i.e. r = 2.368 kg/m3
Therefore,
rO
wO2 := 2

r
i.e. wO 2 = 0.233 (mass fraction of Oxygen.)
and,
rN
wN2 := 2

r
i.e. w N2 = 0.767 (mass fraction of Nitrogen.)

736 FUNDAMENTALS OF HEAT AND MASS TRANSFER


(iv) Molar fractions
Total molar concentration, n := nO2 + n N2
i.e. n = 0.082 kg mole/m3
Therefore,
nO
yO2 := 2

n
i.e. yO2 = 0.21 (mole fraction of Oxygen.)
and,
nN
y N2 := 2

n
i.e. y N2 = 0.79 (mole fraction of Nitrogen.)
Note that mole fractions should be equal to partial pressure fractions.
Example 14.2. Calculate the diffusion coefficient of ammonia (NH3) in air at 20°C and 1 atm. pressure. Then, calculate
the value of D for a pressure of 3 atm. and temperature of 57°C.
Solution. We shall use the empirical relation given by Eq. 14.22:

F1 I
3 1
T2 1
I GH M JK
2
D = 0.0043× × + cm2/s ...(14.22)
F
p ×GV
1 1
2
MC
JK
B
+ Vc
H
3 3
t b

where, pt = total pressure (atm.)


T = absolute temperature (K)
MB, MC = molecular weights of gas species
V b , V c = molecular volumes of B and C at normal boiling points, cm3/gm mole.
Let us denote B for NH3 and C for air. Getting the data for molecular weight and molecular volumes from Table
14.1, we have:
Data:
Pt := 1 atm. MB := 17 V b := 25.81 cm3/gm Mc := 29 V c := 29.89 cm3/gm T := 20 + 273 K
Then,

F1 I
3 1
T2 1
I × GH M JK
2
D := 0.0043 +
F 1 1
2
MC
GH JK
B
pt × Vb3 + Vc3

i.e. D = 0.179 cm 2/s (Diffusiviy of NH3 in air at 1 atm and 20°C.)


(ii) At 3 atm. And 57°C
We have the relation:

F T I ×F P I
3
D1
GH T JK GH P JK
2
1 1
= ...(14.23)
D2 2 2

Here, D1 := 0.179 m2/s


T1 := 20 + 273 K
P1 := 1 atm
P2 := 3 atm
T2 := 57 + 273 K
Therefore,
D1
D 2 :=
LMF T I F P I OP 3

MMGH T JK × GH P JK PP
2
1 1

N 2
Q 2

i.e. D2 = 0.642 cm 2/s (Diffusiviy of NH 3 in air at 3 atm and 57°C.)

MASS TRANSFER 737


Example 14.3. A steel, rectangular container having walls 15 mm thick, is used to store gaseous hydrogen at elevated
pressure. The molar concentrations of hydrogen in steel at the inside and outside surfaces are 1 kg mole/m3 and zero,
respectively. Assuming the diffusion coefficient for hydrogen in steel to be 0.25 ´ 10–12 m2/s, calculate the molar diffu-
sion flux for hydrogen through steel. (M.U. 1999)
Solution.
Data:
L := 0.015 m Db := 0.25 × 10–12 m2/s Cb1 := 1 kg mole/m3 Cb2 := 0.0 kg mole/m3
We have, for molar flux through a plane membrane:
Db
×(Cb1 – Cb2) kg mole/(m2s)
Jb :=
L
i.e. Jb = 16.667 × 10– 12 kg mole/m2s (Molal flux of H2 through steel.)
Example 14.4. Hydrogen gas at 2 bar, 25°C is flowing through a vulcanised rubber tube, 30 mm ID, 50 mm OD. Solubil-
ity of H2 in rubber is 0.053 m 3 of H 2 per atm. per m3 of rubber at 25°C. Diffusivity of H2 through rubber is 18 ´ 10– 11 m2/
s. Density of H2 is 0.0899 kg/m3 at 1 bar pres-sure and 0°C. Calculate percentage reduction in hydrogen loss if the rubber
pipe is covered by 2.5 mm thick steel tubing. Assume diffusivity of H2 through steel as 1.0 x 10-12 m2/s at 25°C. (M.U.)
Solution. See Fig. Ex. 14.4.

Steel tubing

Rubber tubing Mh1, kg/s Rubber tubing


Mh2, kg/s

Ch1 Ch2 = 0 Ch2 Ch2 = 0

r1
r2 r1
r2
r3
(a) only rubber tubing (b) with steel tubing over the rubber tubing

FIGURE Example 14.4 Diffusion through a cylindrical shell

Data:
P1 := 2 ´ 105 Pa L := 1 m r1 := 15 × 10–3 m r2 := 25 × 10–3 m r3 := r2 + 2.5 × 10–3 m T1 := 25 + 273 K
–11 2 –12 2
DH 2 _rubber
:= 18 × 10 m /s DH _rubber = 1 × 10
2
m /s
Solubility of hydrogen in rubber at 1 bar = 0.053 m3/bar/m3 of rubber
Therefore, solubility at 2 bar pressure: S = 2 × 0.053 m3/m3 of rubber
i.e. S := 0.106 m3/m3 of rubber
Gas constant R for Hydrogen:
Given that at 1 bar, 0°C, density is 0.0899 kg/m3
p
Therefore, RH = 2
r ×T

1.0 ´ 10 5
i.e. RH :=
2
0 .0899 ´ 273
i.e. RH = 4.075 × 103 J/kgK
2

Diffusion through rubber wall:


S × P1
Ch1 = r1 = kg/m3 of rubber (concentration on the inside surface)
RH × T1
2

i.e. Ch1 := 0.01746 kg/m3 of rubber


and, Ch2 := 0 kg/m3 of rubber (concentration on the outside surface)

738 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Therefore, resistance to diffusion through rubber:
Fr I
ln GH r JK
2

1
R diff_rubber := s/m3 (for cylindrical shell)
2 × p × DH 2 _rubber ×L
i.e. Rdiff_rubber = 4.517 × 108 s/m3.
Case (i): Only rubber tubing:
Ch 1 - Ch 2
Mass diffusion rate: mh1 :=
Rdiff_rubber
i.e. mh1 = 3.866 × 10– 11 kg/s.
Case (ii): With steel casing over rubber tubing:
Now, resistance to diffususion through steel:
Fr I
ln GH r JK3

2
Rdiff_steel := s/m3 (for cylindrical shell)
2 × p × DH 2 _steel ×L
i.e. R diff_steel = 1.517 ´ 1010 s/m3
Therefore,
Rtotal := Rdiff_rubber + R diff_steel
i.e. R total = 1.562 × 10 10 s/m3
Ch 1 - Ch 2
And, mh2 :=
Rtotal
i.e. mh2 = 1.118 × 10–12 kg/s (mass dffusion rate of H2 with steel tubing over the rubber turbing.)
Therefore, percentage reduction in mass diffusion because of steel turbing:
mh 1 - mh 2
Reduction := ×100
mh 1
i.e. Reduction = 97.109% (percentage reduction as a result of putting steel tubing over rubber tubing.)
Example 14.5. Hydrogen gas is stored at 358 K in a 4.0 m ID, 5 cm thick spheri-
Ni wall mdiff, kg/s
cal container made of Nickel. Molar concentration of hydrogen in Ni at the in-
ner surface is 0.09 kg mole/m3 and is equal to zero at the outer surface.
Determine the mass diffusion rate of hydrogen through the walls of the con-
tainer.
Ch1 Ch2 = 0
Solution. From Table 14.5, we have, for H 2–Ni at 358 K:
–12 2
DH2 _Nickel = 1.2 × 10 m /s
r1 := 2 m r2 := 2.05 m T1 := 358 K Cb1 := 0.09 kg mole/m3
Cb2 := 0 kg mole/m3 MH 2 := 2
r1
We have, the molar diffusion rate for a spherical shell:
r2
Cb 1 - Cb 2
Nb :=
F1 - 1I kg mole/s
GH r r JK
1 2 FIGURE Example 14.5 Diffusion
4 × p × DH through a spherical shell
2 _Nickel

i.e. Nb = 1.113 × 10 –10 kg mole/s


Therefore, mass diffusion rate:
mb := Nb × MH
2

i.e. mb = 2.226 × 10–10 kg/s (mass diffusion rate of hydrogen through Ni wall of spherical container.)
Example 14.6. Helium gas is stored at a pressure of 6 bar and 293 K in a 0.4 m ID, 3 mm thick spherical container made
of fused silica. Determine the rate of pressure drop due to diffusion. Given: D = 0.04 x 10–12 m2/s, and solubility of gas
at the solid surface on the inside is 18 ´ 10–9 kg/(m 3Pa).
Solution. Since the wall thickness is small as compared to the radius, we can approximate the shell as a flat plate.

MASS TRANSFER 739


8314
p := 6 ×10 5 Pa r := 0.2 m L := 0.003 m T := 293 K RHe := J(kgK) i.e. RHe = 2.079 ×103 J/kgK)
4
S := 18 × 10– 9 kg/(m3Pa) Dbc := 0.04 × 10–12 m2/s
p ×V
Mass of gas in the container: m =
RHe × T
Since pressure drop due to diffusion is very slow, temperature T can be considered as constant. Also, since the tank
is rigid, volume V can be considered as constant.
Therefore,
dm V dp
= × ...(a)
dt RHe × T dt
But, dm/dt must also be equal to the diffusion rate, i.e.
dm r - rb 2
= A×Dbc × b1 (for a flat plate (membrane)...(b))
dt L
where, A = area of cross section, rb2 = 0 at the outer surface
Therefore, equating Eqs. a and b,
V dp A × Dbc × r b1
× =
RHe × T dt L

dp R × T × A × Dbc × r b 1
i.e. = He ...(c)
dt V ×L
Now, for a sphere:
A := 4×p× r 2
4
V := ×p ×r3
3
A
and, =15
V
Now, r b1 = solubility x pressure
i.e. rb1 := S×p
Then, substituting in Eq. c, we get:
dp
= 1.315 ´ 10–6 Pa/s (rate of pressure drop.)
dt
Note: This is the initial rate of pressure drop. With time, pressure inside the container will fall; then, the rate of
pressure drop will also decrease.

14.6 Equimolal Counter-diffusion in Gases


Chamber B Chamber C Consider a binary mixture of species B and C. ‘Equimolal counter-diffu-
Nb Nc sion’ implies that in the diffusion process, each molecule of component B
is replaced by each molecule of component C and vice-versa. Refer to
pb, nb pc, nc Fig. 14.5, which shows two chambers B and C, both containing mixtures
of B and C, but at different concentrations, connected to each other by a
x
passage; component B diffuses from higher concentration to the lower
concentration and equimolal counter-diffusion occurs between B and C
Pressure
p t = p b + pc
and each molecule of B is replaced by each molecule of C and vice-versa.
Total pressure, pt of the system remains constant, and, pt = pb + pc.
pb1 pb(X) Distillation columns are good examples for equimolal counter-diffu-
pc2
sion. Venting of a gas to, say, atmosphere, also involves equimolal coun-
pc1 pc(X) pb2 ter-diffusion.
From Fick’s law, we have molar diffusion rates of species B and C:
Distance
(See eqn. 14.20)
FIGURE 14.5 Equimolal counter- mb A dpb
diffusion in a binary gas mixture Nb = = – Dbc × × kg mole/s ...(14.44a)
Mb Ru ×T dx

740 FUNDAMENTALS OF HEAT AND MASS TRANSFER


mc A dpc
and, Nc = = –Dcb × × kg mole/s ...(14.44b)
Mc Ru ×T dx
where, pb and pc are partial pressures of species B and C, respectively.
Note from the Fig. 14.5 that each of the components diffuses in the direction of its drop in concentration.
Now, from Dalton’s law, total pressure of the system is equal to the sum of the partial pressures of the
components of the mixture, i.e.
P t = pb + pc
Differentiating w.r.t. x,
dpt dpb dpc
= +
dx dx dx
However, under steady state conditions, total pressure of the system is constant, i.e.
dpt dpb - dpc
= 0 and, =
dx dx dx
Also, Nb and Nc are numerically equal, since both the species are diffusing in opposite directions but at a
constant rate, i.e.
Nb = –Nc
A dpb A dpc
i.e. –Dbc × × = Dcb × ×
Ru ×T dx Ru ×T dx
Substituting for dpc/dx:
A dpb A dpb
–Dbc × × = –Dcb × ×
Ru ×T dx Ru ×T dx
or, Dbc = Dcb = D. ...(14.45)
Note the important fact from Eq. 14.45 that in equimolal counter-diffusion, the diffusion coef-ficient of com-
ponent B in component C is equal to the diffusion coefficient of component C in component B. Value of D in a
binary mixture is calculated using the empirical Eq. 14.22, as already described.
For constant Dbc, Eq. 14.44 may be integrated between any two planes, to give:
mb A ( pb 2 - pb1 )
Nb = = –D× ×
Mb Ru ×T ( x2 - x1 )
mb A ( pb1 - pb 2 )
i.e. Nb = = D× × kg mole/s ...(14.46)
Mb Ru × T ( x 2 - x1 )
In Eq. 14.46, pb1 and pb2 are the partial pressures of component B at locations x 1 and x 2, respectively.
Example 14.7. A distillation column containing a mixture of Benzene and Toluene is at 101 kN/m2 pressure and tem-
perature = 100°C. The liquid and vapour phase contain 30 mol% and 45 mol% of Benzene. At 100°C, the vapour pressure
of Toluene is 70 kN/m2 and diffusivity is 5 ´ 10–6 m2/s. Work out the rate of interchange of Benzene and Toluene
between the liquid and vapour phases if resistance to mass transfer lies in a film 0.25 mm thick. Universal gas constant
= 8.314 kJ/kg mole K. (M.U. Dec. 1998)
Solution. This is the case of a distillation column, and equimolar counter-diffusion occurs in the column. Let subscripts:
t ® for Toluene and, b ® for Benzene.
Data:
ptot := 1.01× 10 5 N/m2 pvap_toluene := 7.0 × 10 4 N/m 2 yb_liq := 0.3 yb_vap := 0.45 A := 1 m2 (assumed)
–6 2
T := 373 K D := 5× 10 m /s Ru := 8314 J/kg mole K
Dx := 0.00025 m; (Note that Dx = (x 2 – x 1), distance through which diffusion occurs)
Now, we have for Toluene, mass diffusion rate:
D × A × Mt ( pt1 - pt 2 )
mt = × kg/s (from Eq. (14.46))
Ru × T ( x2 - x1 )
And, molal flux:
Nt mt D (pt1 - pt2 )
= = × kg mole/(sm2)
A Mt × A Ru × T (x2 - x1 )

MASS TRANSFER 741


Now, pt1 = partial pressure of Toluene at x 1, i.e. in liquid
x2
Vapour phase:
pt1 := (pvap_toluene)×(1 – yb_liq) ...By Raqult’s law
Dx i.e. pt1 = 4.9 ´ 104 N/m2
And, p t2 = partial pressure of Toluene at x 2, i.e. in vapour
phase:
x1 pt2 := (p tot)×(1 – yb_vap) ...By Dalton’s law
Liquid i.e. pt2 = 5.555 ´ 10 4 N/m2
Therefore, molal flux for Toluene:
Nt D (pt1 - pt2 ) D (pt1 - pt2 )
= × = ×
A Ru × T (x2 - x1 ) Ru × T Dx
Nt
FIGURE Example 14.7 Equimolal counter- i.e. = – 4.224 ´ 10 –5 kg mole/(sm2).
A
diffusion Negative sign indicates that Toluene diffuses from x 2 to x 1,
i.e. from vapour to liquid. Benzene will diffuse at the same rate but
in opposite direction. i.e. Nt = –Nb.
Example 14.8. A tank contains a mixture of CO2 and N2, in the mole proportions of 0.2 and 0.8 at 1 bar and 290 K. It is
connected by a duct of cross-sectional area 0.1 m2 to another tank containing a mixture of CO2 and N2 in the molal
proportions of 0.8 and 0.2. The duct is 0.5 m long. Determine the diffusion rates of CO2 and N2 in kg/s. Given: D = 0.16
´ 10 –4 m2/s for CO2/N2...at 293 K...from tables. (M.U. Dec. 1998)
Solution. This is the case of equimolar counter-diffusion between CO2 and N2.
Data:
P tot := 1.00 ×10 5 Pa p1CO := 0.2 ×105 Pa
2
p1N := 0.8 ´ 105 Pa
2
p2 CO := 0.8×105 Pa
2

p2 N := 0.2× 10 5 Pa
2
T := 290 K Dx := 0.5 m A := 0.1 m2 D := 16×10–6 m2/s
Ru := 8314 J/(kg mole K)
For equimolar counter-diffusion, we have:
Molar flow rate, from Eq. 14.46:
mb A ( p b1 - p b 2 )
2
0.1 m ...cross-sectional area Nb = = D× × kg mole/s ...(14.46)
Mb Ru × T ( x 2 - x1 )
For Nitrogen, in this case, we can write:
Chamber B Chamber C A ( p1N - p 2 N )
CO2(0.2) + CO2(0.8) + N nitrogen := D× × 2 2

Ru × T Dx
N2(0.8) NN NCO N2(0.2)
2 2
i.e. Nnitrogen = 7.963 ´ 10–8 kg mole/s (molar flow rate of N 2)
0.5 m and, mass flow rate of Nitrogen: = Nnitrogen ´ (Mol. wt. of Nitro-
x gen)
i.e. Nnitrogen ×28 = 2.23×10– 6 kg/s (mass flow rate of N2.)
Pressure For CO2, we can write:
pt = p1CO + p1N
2 2
A ( p1CO - p 2 CO )
p1N N CO := D× × 2 2

2 p2CO 2
Ru × T Dx
2
p2 N i.e. N CO = – 7.963 ´ 10–8 kg mole/s (molar flow rate of CO 2.)
p1CO 2
2

2
Note: Molar flow rate of CO2 is equal to that of N2, but in opposite
Distance direction, as indicated by negative sign.
And, mass flow rate of CO2: = N CO × (Mol. wt. of CO2)
FIGURE Example 14.8 Equimolal counter- 2

diffusion in a binary gas mixture i.e. N CO ×44 = – 3.504× 10–6 kg/s


2
(mass flow rate of CO2.)

Example 14.9. A spherical ball of ice, 1 cm diameter is suspended in still dry air at 1.013 bar. Calculate the initial rate of
evaporation at the surface.
Take D = 0.256 ´ 10–4 m2/s. At 0 deg.C, saturated vapour pressure = 0.0061 bar.
Solution.
Data: T := 273 K D := 0.256 ×10–4 m2/s Pt := 1.013 ×105 Pa r1 := 0.5 ´ 10–2 m

742 FUNDAMENTALS OF HEAT AND MASS TRANSFER


r2 := ¥ m Ru := 8314 J/kg mole K r2 = ¥
At 0 deg. C:pv 1 := 0.006 ×10 5 N.m2 (vapour pressure of water vapour)
Therefore, p a1 := pt – pvl
i.e. p a1 := 1.007 × 10 5 N/m2 (partial pressure of air)
At r = infinity: pv 2 := 0 (vapour pressure of water vapour)
Therefore, pa2 := pt – pv2
i.e. pa2 := 1.013 × 105 N/m2 (partial pressure of air)
For equimolar counter-diffusion in spherical coordinates
Nb - D dpb
= × kg mole/s m2 (molar flux)
A Ru × T dx
r1 = 5 mm
Nb - D dpb
i.e. = ×
4 ×p × r 2 Ru × T dx FIGURE Example 14.9
Equimolal counter-diffusion from

z zF FGH IIJK
Separating the variables and integrating,
2 2 a sphere to surroundings
- N b × Ru × T 1
dpv =– × dr
1 4 ×p × D 1 r2
- N b × Ru × T 1 1
i.e. (pv1 – pv2) =
4 ×p × D
× GH
-
r1 r2 JK
But, r 2 = ¥ and, pv2 = 0
Therefore,
4 ×p × D × pv × r1
N water :=
Ru × T
i.e. N water = 4.323 × 10–10 kg.mole/s (initial rate of evaporation of water)
We can also write:
Nwater ×18 = 7.781 × 10–9 kg/s. (since mol. wt. of water vap. = 18)
Example 14.10. A pipe carrying ammonia at 1 atm. is maintained at that
Air at 1 atm, 25°C
pressure by venting ammonia to atmosphere through a 5 mm ID, 5 m
long tube. Assuming both ammonia and atmospheric air to be at 25 NH3
deg.C, determine: (a) mass flow rate of ammonia diffusing into the at-
mosphere, and (b) mass flow rate of air that diffuses into the pipe line. x2
Take the diffusion coefficient of ammonia in air (or, air in ammonia) as:
D = 0.26 ´ 10–4 m2/s.
5 mm diameter
Solution. This is a case of equimolal counter-diffusion, where two large
reservoirs containing mixture of ideal gases at different concentrations,
L=5m
are connected to each other by a pipe.
Data:
P1_NH := 1.013× 105 Pa
3
(partial pressure of ammonia at section-1) x1
P2 _NH := 0 Pa (partial pressure of ammonia at section-2,
3
i.e. at atmosphere) NH3 at 1 atm, 25°C

x1 := 0 m x2 := 5 m d := 0.005 m L := 5 m T := 298 K
FIGURE Example 14.10 Equimolal
D := 0.26 ´ 10– 4 m2/s Ru := 8314 J/kg moleK
counter-diffusion of NH3 and Air
p × d2
A := i.e. A = 1.963 × 10–5 m2 (cross-sectional area of vent pipe)
4
Note that the pressure of ammonia at the bottom of vent pipe (x = 0) is 1 atm. and is equal to zero at the top (x = L)
For equi-molal diffusion, we have:
mb A ( pb1 - pb 2 )
Nb := = D× × kg mole/s. ...(14.46)
Mb Ru × T (x2 - x1 )
With the notations used above, we get:
A ( p1_ NH - p 2 _ NH )
N NH := D× × 3
kg mole/s.
3

3
Ru × T ( x 2 - x1 )

MASS TRANSFER 743


i.e. N NH = 4.175 × 10–12 kg mole/s.
3

Therefore, mass diffusion rate of ammonia into air:


MNH := 17
3
(molecular wt. of NH3)
and, mNH := N NH × MNH
3 3 3

i.e. mNH = 7.097 × 10– 11 kg/s


3
(mass diffusion rate of ammonia into atmosphere.)
And, mass diffusion rate of air into ammonia:
For equimolal diffusion, we have the molal diffusion rate of ammonia into air is equal to the molal diffusion rate of
air into ammonia.
i.e. N Air = N NH = (4.175 × 10–12) kg mole/s
3

Therefore, mass diffusion rate of air into ammonia:


M air := 29 (Molecular weight of air)
and, m Air = Nair × Mair
i.e. mAir = –1.211 × 10–10 kg/s (mass diffusion rate of air into ammonia.)
Note: negative sign simply indicates that mass diffusion of air is in the opposite direction, i.e. from atmosphere to
ammonia in the pipe.

14.7 Steady State Uni-directional Diffusion—Diffusion of Water Vapour


through Air
Let us consider a binary gas mixture wherein one gas diffuses through another gas which remains as a stagnant
layer; familiar examples of uni-directional diffusion through a stagnant gas layer are: absorption, humidification
and, the diffusion of water vapour through a layer of
Air air, when evaporation of water occurs, say in a well or
a test tube.
Consider the evaporation of water contained in a
2 X well and the subsequent diffusion of this water vapour
through the stagnant gas layer above the water. See
Fig. 14.6.
Assumptions:
pw pa
X (i) steady state and isothermal conditions exist
(ii) total pressure is constant
1 (iii) both air and water vapour behave like perfect
gases
Water pt = pw + pa
(iv) air has negligible solubility in waer, and
(v) air movement at the top is very slow, i.e. move-
ment of air is just sufficient to carry away the
FIGURE 14.6 Diffusion of water vapour through air evaporated water, but not large enough to
cause any change in the concentration profile of
air.
Water which evaporates at surface 1 (see the Fig. 14.6) diffuses through air standing above it; and, in steady
state, this upward movement of water vapour must be balanced by a downward diffusion of air so that concen-
tration at any location x, remains a constant.
Downward diffusion of air:
A × Ma dpa
ma = –D× × ...(14.47)
Ru ×T dx
Since air is not soluble in water, this downward movement of air will cause a bulk mass movement upward,
with a velocity just large enough to compensate for the diffusion of air downward.
Bulk mass transfer of air is equal to:
Ma
ma = –ra ×A×u = –pa× ×A×u ...(14.48)
Ru ×T
where, u = bulk velocity upward

744 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Equating Eqs. 14.47 and 14.48, we get:
A × Ma dpa Ma
D× × = pa × ×A×u
Ru ×T dx Ru ×T
D dpa
i.e. × u= ...(14.49)
pa dx
Now, the total mass transfer of water is equal to the upward mass diffusion of water plus water vapours
carried upward by bulk movement of air.
i.e.
A × Mw dpw
mwtot = –D× × + rw ×A×u ...(14.50)
Ru × T dx
A × Mw dpw p × Mw × A D dpa
i.e. mwtot = –D× × + w × × ...(14.51), using Eq. (14.49)
Ru × T dx Ru × T p a dx
In the above equations, pa = partial pressure of air and, pw = partial pressure of water vapour. Now, from
Dalton’s law of partial pressures, we have, the total pressure:
pt = pa + pw
dpt dpa dp
Therefore, = + w
dx dx dx
But, the total pressure of the system is constant; so,
dpt
=0
dx
dpa - dpw
Therefore, =
dx dx
Then, substituting in Eq. 14.51,
A × Mw dpw p × Mw × A D dpw
mwtot = –D× × – w × ×
Ru × T dx Ru × T pa dx

A × Mw dpwF I p
i.e. mwtot = –D× ×
Ru × T dx
GH JK
× 1+ w
pa

A × M dp F p + p I
R × T dx GH p JK
w w a w
i.e. mwtot = –D× × ×
u a

A × M dp F p I
mwtot w
×G w
J t
R × T dx H p - p K
i.e. = –D× × ...(14.52)
u t w
Eq. 14.52 is known as ‘Stefan’s law’ for an ideal gas diffusing through another, stationary, ideal gas compo-
nent in a binary gas mixture.
Integrating Stefan’s equation between planes x 1 and x 2,

mwtot ×
z
x1
x2
dx =
- D× A
Ru × T
×Mw ×pt ×
z pw 2

p w1
1
( pt - pw )
dpw

i.e. mwtot ×(x2 – x 1) =


D× A
Ru ×T
×Mw ×pt ×
z pw 2 1
p w1 ( pw - pt )
dpw

D× A F
p - pt I
i.e. mwtot ×(x2 – x 1) =
Ru ×T
×Mw ×pt × ln w 2 GH
pw1 - pt JK

MASS TRANSFER 745


Pressure
D× A F I
p - pw 2
i.e. mwtot ×(x2 – x 1) =
Ru ×T GH
×Mw ×pt × ln t JK
pt - pw1
pt = pw + pa
D× A
×M ×p × ln G
Fp I
H p JK
a2
pa2 i.e. mwtot ×(x2 – x 1) = w t
R ×Tu a1
pa1
Therefore,
pw1
D × A Mw × pt p F I ...kg/s
pw2
mwtot = ×
Ru T ( x2 x1 )
× -
× ln a 2
pa1 GH JK ...(14.53)

Now, let us define ‘Log Mean Partial pressure of Air (LMPA)’ as:
x1 Distance x2
p a 2 - p a1
FIGURE 14.7 Uni-directional LMPA =
Fp I
diffusion of water vapour in a station-
ary gas (i.e. air)
ln GH p JK
a2
a1

Fp I = p -p
i.e. ln GH p JK LMPA
a2
a1
a2 a1

Then, Eq. 14.53 can be written as:


D × A Mw × pt ( pa 2 - pa1 )
mwtot = × ×
Ru ×T ( x2 - x1 ) LMPA
D × A Mw × pt ( pw1 - pw 2 )
i.e. mwtot = × × kg/s. ...(14.54)
Ru ×T (x 2 - x1 ) LMPA
Note that, in Eq. 14.54, instead of LMPA we can use arithmetic mean pressure, [i.e. (pa1 + pa2)/2], if the
partial pressure of water vapour does not change much as compared to the total pressure of the mixture.
Let the partial pressure of water be pw at any plane x; then, integrating the Stefan’s equation between planes
x 1 and x, we get:

D× A F
p - pw I
mw ×(x – x 1) =
Ru ×T
×Mw ×pt × ln t GH
pt - pw1 JK
i.e. pw = pt – (pt – pw1)×exp
LM m w
×( x - x1 )×
Ru × T OP ...(14.55)
N p ×M
t w D×A Q
Eq. 14.55 gives the variation of partial pressure of water vapour with distance x along the tube.
And, for the stagnant gas, i.e. air:
pa = pt – pw

i.e. pa = (pt – pw1)×exp


LM m w
× (x - x1 ) ×
Ru × T OP ...(14.56)
N p ×M
t w D×A Q
Variation of pw and pa with x, are shown graphically in Fig. 14.7:
Example 14.11. Calculate the hourly loss of water from a well 6 m deep and cross-sectional area 5 m2. Temperature is 30
deg.C and pressure is 1.013 bar. Given: D = 0.256 cm2/s. Also, saturated pressure of water at 30°C = 0.042 bar.
Solution. This is a case of uni-directional diffusion of water vapour through a column of air. Therefore, Eq. 14.53 is
applicable. Refer Fig. 14.6.
Data:
Pwl := 0.042 ´ 105 Pa Pt = 1.013 ´ 105 Pa x 1 := 0 m x2 := 6 m A := 5 m2 T := 30 + 273 K
–4 2
D := 0.256 ´ 10 m /s Ru := 8314 J/kg moleK Mw := 18
Then, from Eq. 14.53, we have:

746 FUNDAMENTALS OF HEAT AND MASS TRANSFER


D × A M w × pt F I
p
mw = ×
Ru × T ( x2 - x1 ) GH JK
× ln a 2
pa1
...(14.53)

where, pa2 and pa1 are partial pressures of non-diffusing gas, i.e. air.
Now, pa1 := pt – pwl
i.e. pa1 = 9.71 ´ 104 Pa (partial pressure of air at bottom of well)
and, pa2 := pt – 0
i.e. pa2 = 1.013 ´ 105 Pa (partial pressure of air at top of well)
Therefore,
Diffusion rate of water vapour is given by:
D × A M w × pt p F I
mw := ×
Ru × T ( x2 - x1 )
× ln a 2
pa1 GH JK
i.e. mw = 6539 ´ 10–7 kg/s
and, mw ×3600 = 2.354 ´ 10–3 kg/hr (hourly loss of water.)
Example 14.12. In a Stefan tube experiment with carbon tetrachloride (CCl4), following data are noted: Diameter of tube
= 1 cm. Length of tube above liquid surface = 12 cm. Temperature maintained = 0°C. Pressure maintained = 76 cmHg.
Vapour pressure of CCl4 at 0°C = 33 mmHg. Evaporation of CCl4 = 0.037 g. Time of evaporation = 10 h. Estimate the
diffusion coefficient of CCl4 into air.
Solution. Stefan tube experiment is conducted to determine the diffusion coefficient of diffusing gas in a column of non-
diffusing gas, i.e. CCI4 in air, in this case.
Data:
d := 0.01 m Dx := 0.12 m T := 273 K pt := 760 mmHg pwl := 33 mmHg Pa1 := pt – pw1 mmHg
i.e. pa1 := 727 mmHg pa2 := pt – 0 mmHg i.e. pa2 = 760 mmHg Ru := 8314 J/kg moleK
0 .037 ´ 10 - 3
mw = kg/s i.e. mw = 1.028 ´ 10–9 kg/s (evap. rate of CCl 4) p := 1.013 ´ 105 Pa (total pressure)
10 ´ 3600
Mw := 154 (mol. wt. of CCl 4)
p × d2
A :=m2
4
i.e. A = 7.854 ´ 10–5 m (cross-sectional area of tube)
Now, we have for uni-directional gas diffusion:
Fp I
D × A × Mw × p × ln GH p JK
a2

a1
mw =
Ru × T × D x
Therefore,
mw × Ru × T × D x
D :=
Fp I
× p × ln G
m2/s (diffusivity of CCl4 in air)

H p JK
a2
A × Mw
a1

i.e. D = 5.147 ´ 10–6 m2/s (diffusivity of CCl4 in air.)

14.8 Steady-state Diffusion in Liquids


Now, let us study steady state equimolal counter-diffusion and uni-directional diffusion in a binary mixture of
liquids. While dealing with liquids, we generally write the equations in terms of molar concentrations.
14.8.1 Steady-state Equimolal Counter-diffusion in Liquids
Writing in terms of molar concentrations, the molar flux of component B is given by:
dCb
Jb = – D× ...(14.57)
dx
Integrating between x = x 1 and x = x 2 (with Cb varying from Cb1 to Cb2):
C - Cb 2
Jb = D× b1 ...(14.58)
x 2 - x1
MASS TRANSFER 747
where,
Cb1, Cb2 = molal concentrations at x 1 and x 2, in kg mole/m3,
D = mass diffusivity for liquid-to-liquid in binary liquid mixture, m2/s.
Jb = mass flux of species B, kg mole/(m2s), and
x = distance, m.
In case of liquids, mole fraction is defined as:
Ci
xi = = mole fraction of component i, where i = B or C
C
and, Cb + Cc= C = total molal concentration of the mixture, kg mole/m3
Then, in terms of mole fraction, Eq. 14.58 is expressed as:
xb1 - xb 2
Jb = D × C kg mole/(m2s). ...(14.59)
x2 - x1

14.8.2 Steady-state Uni-directional Diffusion in Liquids


This analysis is similar to that done in case of gases. However, now, Eq. 14.54 is written as a molal flux and,
instead of partial pressures used in case of gases, now molal concentrations are used for liquids.
Eq. 14.54 for gases may be re-written as:
mwtot D pt ( p - pw 2 )
= × × w1 kg mole/(m 2 s)
A × Mw Ru ×T (x 2 - x1 ) LMPA
In an analogous manner, for liquids, we write: (liquid B diffusing in stationary liquid C)
C C - Cb 2 C C - C c1
Jb = D × × b1 = D× × C2 kg mole/(m2s) ...(14.60)
( x2 - x1 ) Cc , ln ( x2 - x1 ) Cc , ln
where, Cc, ln = logarithmic mean concentration of component C, defined as:
Cc 2 - Cc 1
Cc, ln =
FG C IJ
c2
...(14.61)
ln
HC K
c1
and, note that C = Cb + Cc and Cb2 – Cb1 = Cc1 – Cc2
In terms of mole fractions, Eqs. 14.60 and 14.61 can be written as:
C X - X c1 C X - Xb 2
Jb = D × × c2 = D× × b1 kg mole/(m2s) ...(14.62)
( x2 - x1 ) X c , ln ( x 2 - x1 ) Xb , ln
where, Xc, ln = logarithmic mean mole fraction of component C, defined as:
X c 2 - Xc 1
Xc, ln =
FG X IJ
c2
...(14.63)
ln
HX K
c1
In the above equations,
C = total molal concentration of mixture, kg mole/m3,
Xi = Ci/C = mole fraction
D = mass diffusivity , m2/s,
Jb = mass flux of species B, kg mole/(m2s),
x = distance, m.

14.9 Transient Mass Diffusion in Semi-infinite, Stationary Medium


Now, consider a semi-infinite medium in which the component B is initially at a uniform concentration Cbi
throughout. Then, suddenly the surface at x = 0 is exposed to a concentration of B equal to Cbs and maintained at
that value for all t. We are interested to know the value of Cb at a given distance x from the surface at a given
time t.

748 FUNDAMENTALS OF HEAT AND MASS TRANSFER


We have already derived the general differential eqn. for mass diffusion, viz.
Ñ 2 Cb = (1/D).( ¶Cb /¶t) .... (14.29)
Then, transient, uni-directional mass diffusion equation can be written as:
dCb d 2 Cb
= D× ...(14.64)
dt dx 2
This equation has to be solved with the following boundary conditions:
(i) Cb = Cbi at t = 0, for all x
(ii) Cb = Cbs at x = 0, for all t
(iii) Cb = Cbi at x = ¥, for all t
Note immediately that Eq. 14.64 and the set of boundary conditions are similar to the case of transient one-
dimensional heat conduction encountered in Chapter 7. By analogy, now, we can write the solution for Eq. 14.64
also, looking at the solution in the case of heat conduction, i.e.
F x I
T ( x , t ) - T0
Ti - T0
= erf GH 2× a ×t JK ...(7.29)

where Ti = initial uniform temp. and To = temp. to which the suddenly exposed and then maintained for all
times.
So, we have for solution of Eq. 14.64:

Cb - Cbs C - Cb F I = erf (u)


Cbi - Cbs
= bs
Cbs - Cbi
= erf GH 2× x
J
D ×t K
...(14.65)

x
where, u =
2× D ×t
Mass flow rate through the boundary is given by:
mb dC - D ×(Cbi - Cbs ) - u2
= – D× b = ×e ...(14.66)
A dx p × D ×t

FG m IJ - D ×(Cbi - Cbs )
H AK
b
and, = ...(14.67)
x=0 p × D ×t
This solution is applicable, typically, in case of solid state diffusion in case-hardening of mild steel in a
carburising atmosphere.
Another quantity of interest in solid diffusion process is the ‘penetration depth’ (d diff), defined as the loca-
tion x where the tangent to the concentration profile at the surface (x = 0) intercepts the Cb = Cbi line (See Fig.14.8).
Then, penetration depth is obtained as:
Cbs - Cbi Cbs - Cbi
d diff =
LM- F dC I OP =
(Cbs - Cbi )
(differentiating Eq. 14.65)

MN GH dx JK PQ
b
p × D ×t
x=0

i.e. d diff = p × D ×t ...(14.68)


Example 14.13. A mild steel piece has uniform, initial carbon concentration of 0.2 % by mass. It is exposed to a
carburising atmosphere in a furnace, where the surface concentration is maintained at 1.3 %. Determine how long the
piece must be kept in the furnace for the concentration of carbon at a location 0.4 mm below the surface to reach 1 %.
Take D = 5 ´ 10 –10 m2/s.
Solution. This problem is of transient mass diffusion in a semi-infinite medium.
Data:
Cbi := 0.002 (initial concentration of carbon)
Cbs := 0.013 (surface concentration of carbon)
Cb := 0.01 (desired concentration at x = 0.4 mm)
x := 0.4 ´ 10 –3 m (depth of penetration)
D := 5 ´ 10 –10 m2/s (diffusion coefficient)

MASS TRANSFER 749


Semi-infinite medium

Cbs
Cb(X, t) Tangent to concentration gradient at x = 0

Cbi 0 x

ddiff

FIGURE 14.8 Transient diffusion in a semi-infinite medium-penetration depth

Then, for diffusion in a semi-infinite medium, we have:

Cbs - Cb F I
= erf
x
GG JJ ...(14.65)
Cbs - Cbi 2× D ×t H K
F I
Cbs - Cb w - wb
= bs GG 2× x
J
D ×t JK
i.e.
Cbs - Cbi wbs - wbi
= erf
H (A)

Now, from data:


Cbs - Cb
= 0.273
Cbs - Cbi
Then, from table of error functions, we read
erf (0.247) = 0.273
Then, from Eq. A:
x
= 0.247
2× D ×t

F x I 2

GH 2 × 0.247 JK
and, t :=
D
i.e. t := 1.311 × 103 s (time required)
i.e. t := 0.364 h (time required.)
Alternatively:
If we do not have the table of error functions, still the problem can be easily solved with the solve block of Mathcad;
Mathcad has built-in-error functions available.
We start with a guess value for t and write the constraint under ‘Given’, then, command ‘Find’ gives the value of t ,
as shown below:
t := 100 s (guess value)
Given

Cbs - Cb F I
= erf
x
GG JJ
Cbs - Cbi 2× D ×t H K
Find (t ) = 1.315 × 10 3
i.e. t = 1315 s (time of exposure required in the furnace)
t
i.e. = 0.365 h (time of exposure required in the furnace.)
3600
Note: Values of t obtained are practically the same, by both the methods.

750 FUNDAMENTALS OF HEAT AND MASS TRANSFER


14.10 Transient Mass Diffusion in Common Geometries
Transient mass diffusion analysis is important in hardening of mild steel by carburising, where the steel compo-
nent is packed in a carbonaceous material and kept in a furnace at high temperature for a desired length of time.
We have worked out an example of this type already. Transient analysis is also required in gem industry (to get
desired colour for valuable stones), in ‘doping’ of n-or p-type of semiconductor materials, in drying of coal,
timber, food, textiles, etc.
Transient mass diffusion and transient heat conduction are analogous if the solution is dilute. One-dimen-
sional transient heat conduction in plane walls, cylinders and spheres was discussed in Chapter 7 and Tabular
and Chart solutions were presented. These solutions are applicable to transient mass diffusion problems, too, if:
(i) the diffusion coefficient (D) is constant (corresponding to constant thermal diffusivity, in transient heat
conduction)
(ii) there are no homogeneous reactions occurring (corresponding to no heat generation), and
(iii) initial concentration of species B is constant throughout the medium (corresponding to uniform initial
temperature).
Quantities which are analogous in heat and mass transfer are tabulated in Table 14.8.
Then, for example, for a plane wall, we have the temperature distribution given by:
T ( x, t ) - Ta 2
Fo FG
= A1 × e - l 1 × × cos
l 1 ×x IJ
Plane wall: q (x, t ) =
Ti - Ta H L K (Fo > 0.2 ...(7.24a))

Now, the corresponding equation can be written for transient mass diffusion in a plane wall and the values
of constants A1 and l1 (as a function of Bi ) can be taken from the Table 7.1 given in Chapter 7. Similarly the chart
solutions given by Heisler’s charts can be applied for transient, one-dimensional mass diffusion problems.

14.11 Mass Transfer Coefficient


Mass transfer coefficient is defined in a manner analogous to convective heat transfer coefficient.
Remember that convective heat transfer coefficient ‘h’ is obtained from Newton’s law:
Q = h ×A× D T, W ...(14.69)
(a) Steady state diffusion of a fluid across a solid layer of thickness (x2 – x1):
Mass diffusion rate for diffusion through a solid layer is given by: (see Eq. 14.31)
D × A × (Cb1 - Cb 2 )
mb = kg/s ...(14.70)
x 2 - x1
Writing this in a manner analogous to Eq. 14.69:
D × A × (Cb1 - Cb 2 )
mb = = hmc × A × (Cb1 – C b2)
x 2 - x1
Therefore, hmc , the mass transfer coefficient based on concentration differences, can be written as:

TABLE 14.8 Analogous quantities in heat conduction and mass diffusion


Heat conduction Mass diffusion
T C, y, r or w
a Dbc

T (x , t ) - Ta w b (x , t ) - w ba
q (x, t ) = q mass =
Ti - Ta w bi - w ba
x x
z= z mass =
2× a ×t 2× D bc ×t

h ×L hmass ×L
Bi = Bi =
k mass
Dbc
a ×t Dbc ×t
Fo = Fo mass =
L2 L2

MASS TRANSFER 751


D
hmc = m/s ...(14.71)
( x2 - x1 )
Note the units of mass transfer coefficient: m/s.
(b) Steady state equimolal counter-diffusion:
In this case, from Eq. 14.46, we can write the mass diffusion rate as:
Mb ( pb1 - pb 2 )
mb = D × A × ×
Ru ×T ( x2 - x1 )
D Mb
i.e. mb = × × A× (pb1 – p b2)
( x 2 - x1 ) Ru ×T
Mb
i.e. mb = h m c × × A × (pb1 – pb2) kg/s ...(14.72)
Ru ×T
Now, if we define a mass transfer coefficient hmp , based on partial pressure differences, we rewrite Eq. 14.72
as:
Mb
mb = h m c × × A× (pb1 – pb2) = hmp × A× (pb1 – pb2)
Ru ×T
And, it is clear that:
Mb
hmp = h m c ×
Ru ×T
hmc
i.e. hmp = ...(14.73)
R ×T
i.e. mass transfer coefficient based on pressure difference is obtained by simply dividing the mass transfer coef-
ficient based on concentration differences by (R.T) where R = particular gas constant, and, T = temperature in
Kelvin.
(c) Diffusion of water vapour through a layer of stagnant air:
In this case, we have seen that the mass diffusion rate of water vapour is given by:

D × A Mw × pt F I p
mw = ×
Ru ×T ( x2 - x1 ) GH JK
× ln a2
p a1
...(14.53)

D× A M × p Fp -p I = h
× ln G
H p - p JK
w t t w2
i.e. mw = × m p × A× (pw1 – pw2) ...say
R ×T (x - x )
u 2 1 t w1
Then, for this case, the mass transfer coefficient based on pressure difference can be written as:

D × pt M F
p - pw 2 I
hmp = × w × ln t
( x2 - x1 ) ×( pw1 - pw 2 ) Ru × T GH
pt - pw1 JK ...(14.74)

And, for this case, the mass transfer coefficient based on concentration difference would be:

D × pt F p - pw 2 I
h m c = h mp ×(R×T) =
( x2 - x1 ) ×( pw1 - pw 2 ) GH
× ln t
pt - pw1 JK ...(14.75)

14.12 Convective Mass Transfer


So far, we considered molecular diffusion of a fluid through a solid, or between two fluids. In these cases, the
bulk velocities of the species were insignificant and only the diffusion velocities were of significance.
However, when the bulk velocities of species diffusing are significant, we have the convective mass transfer.
This is analogous to convective heat transfer, just as the molecular diffusion is analogous to conduction heat
transfer.

752 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Convective mass transfer involves the transportation of material across the boundary. Again, just as in the
case of convective heat transfer, convective mass transfer is of two types: (a) ‘Free (or natural) convective mass
transfer, and (b) ‘Forced’ convective mass transfer. In free convective mass transfer, buoyancy forces cause the
circulation due to density differences, whereas in forced convective mass transfer, an external agency such as a
fan or pump causes the circulation. Further, forced convective mass transfer may be of laminar or turbulent
types, depending on the Reynolds number.
Recall that for convective mass transfer, we have the Newton’s equation for heat flux:
Q
= h × DT
A
Similarly, the equation for mass flux in the case of convective mass transfer is:
mb
= hm × D C b ...(14.76)
A
where, h m = convective mass transfer coefficient, and DC b = concentration difference of species B.
Just as in the case of convective heat transfer, analytical treatment of convective mass transfer is compli-
cated, because of the effects of flow velocity, surface geometry, flow regime, flow type (i.e. external or internal
flow), composition and variation of fluid properties. Therefore, generally, empirical relations, obtained as a result
of experimentation, are resorted to.
For flow over a flat plate, we saw earlier that a velocity boundary layer develops; similarly, a concentration
boundary layer also develops and the equations for conservation of momentum, energy and concentration may
be written as follows:
Momentum: u.(¶ u/¶ x) + v.(¶ u/¶ y) = n .(¶ 2u/¶ y2)
Energy : u.(¶ T/¶ x) + v.(¶ T/¶ y) = a.(¶ 2T/¶ y2) , and
Concentration : u.(¶ C/¶ x) + v.(¶ C/¶ y) = D.(¶ 2 C/¶ y2),
where, C = concentration of component diffusing throug2h the boundary layer.
With reference to the velocity, temperature and concentration boundary layers, we have the following non-
dimensional numbers appearing in the empirical correlations:
Prandtl number (Pr ) Pr is defined as: Pr = n/a , and is the connecting link between the velocity and temperature
profiles. Pr = 1 indicates that these profiles are identical.
Schmidt number (Sc) Sc is defined as: Sc = n /D, and is the connecting link between the velocity and concentra-
tion profiles. Sc = 1 indicates that these profiles are identical.
Lewis number (Le) Le is defined as: Le = a /D, and is the connecting link between the temperature and concen-
tration profiles. Le = 1 indicates that these profiles are identical.
All the three boundary layers will coincide if Pr = Sc = Le.
Noting that the governing equations for momentum, energy and mass transfer are similar, it is reasonable to
guess that the empirical correlations for mass transfer coefficient will also be similar to the correlations for the
convective heat transfer, studied earlier. For heat transfer coefficient, we had the general correlation of the type:
h×L
Nu = = f (Re, Pr)
k
In an analogous manner, for mass transfer coefficient in convective mass transfer, we have:
hm × L
Sh = = f (Re, Sc) ...(14.77)
D
where, Sh is the ‘Sherwood number’, which represents a non-dimensional mass transfer coefficient.
Table 14.9 gives the Sherwood number relations for a few convective mass transfer situations, written by
analogy with Nusselt number relations for similar convective heat transfer situations.
Note that in Table 14.9, for mass transfer, Grashoff number is defined as follows:
FG D r IJ × L
3

Gr =
g × ( r a - r s ) × L3
=

HrK
r ×n 2 n2

MASS TRANSFER 753


TABLE 14.9 Sherwood number relations for convective mass transfer
Convective heat transfer Convective mass transfer
1. Forced convection over a flat plate:
Local heat transfer coefficient: Local mass transfer coefficient:
hm × x
Nux = 0.332 × Re x ×Pr 0 . 333 ...laminar flow (Re < 5 ´ 10 5 ) Sh x = = 0.332 × Re x × Sc 0.333
D
hm × x
Nux = 0.0298 × Rex0.8 × Pr 0.333 ...turbulent flow (Re > 5 ´ 10 5) Sh x = = 0.0298 × Re x0.8 × Sc 0.333
D
Average heat transfer coefficient: Average mass transfer coefficient:
1
hm ×L
Nua = 0.664 × ReL ×Pr 0 . 333 ...laminar flow Sh a = = 0.664 × Re L0.5× Sc 3 , Sc > 0.5
D
1
hm ×L
Nu a = 0.037× Re L0.8 × Pr 0.333 ...turbulent flow Sh a = = 0.037 × Re L0.8 × Sc 3 , Sc > 0.5
D
For mixed b.l. conditions with Re c = 5 ´ 10 5 : For mixed b.l. conditions with Rec = 5 ´ 105:
Nu a = (0.037× Re L0.8 – 870) × Pr 0.33 Sh a = (0.037 × Re L0.8 – 870) × Sc 0.33
2. Fully developed flow in smooth, circular pipes:
Laminar flow (Re < 2300): Sh = 3.66 ...(for uniform wall mass concen-
Nu = 3.66 ...for uniform wall temperature tration), and Sh = 4.36 for constant wall mass
Nu = 4.36 ...for uniform wall heat flux flux. Turbulent flow (2000<Re <35000):
Turbulent flow (Re > 10000): Gilliand’s relation:
Nu = 0.023 × Re 0.8 × Pr 0.4 0.7 < Pr < 160 Sh = 0.023 × Re 0.83× Sc 0.44× 0.6 < Sc < 2.5
3. Natural convection over surfaces:
(a) Vertical plate:
1 1
Nu = 0.59 × (Gr ×Pr ) 4 10 5 < Gr × Pr < 109 Sh = 0.59× (Gr ×Sc ) 4 105 < Gr × Sc < 10 9
1 1
Nu = 0.1× (Gr × Pr ) 3 109 < Gr × Pr < 1013 Sh = 0.1× (Gr ×Sc ) 3 109 < Gr × Sc < 10 13
(b) Upper surface of horizontal plate, Fluid near the surface is light,
(surface is hot, Ts > Ta ): (r s < r a ):
1 1
Nu = 0.54 × (Gr ×Pr ) 4 104 < Gr × Pr < 107 Sh = 0.54 × (Gr ×Sc ) 4 104 < Gr × Pr < 107
1 1
Nu = 0.15 (Gr × Pr ) 3 10 7 < Gr × Pr < 1011 Sh = 0.15 × (Gr ×Sc ) 3 107 < Gr × Pr < 10 11
(c) Lower surface of horizontal plate, Fluid near the surface is light,
(surface is hot, Ts > Ta ): (r s < r a ):
1 1
Nu = 0.27× (Gr ×Pr ) 4 105 < Gr × Pr < 1011 Sh = 0.27 × (Gr ×Sc ) 4 105 < Gr × Sc < 1011
4. For natural convection through a tube:
Steinberger and Treybol relation for mass transfer is given by:
Sh = 2 + 0.57 × (Gr × Sc ) 0.25 ...for Gr × Sc < 108
and, Sh = 2 + 0.025 × (Gr × Sc) 0.33 × Sc 0.245 ...for Gr × Sc > 108

14.13 Reynolds and Colburn Analogies for Mass Transfer


Reynolds and Colburn analogies for heat transfer can be extended to the case of mass transfer, to get a relation
between the mass transfer coefficient and the friction factor. Reynolds analogy for heat transfer in a of pipe flow
is written as:
Nu h Cf
= St = = ...(14.78)
Re × Pr r × C p × V 2

754 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Similarly, for mass transfer, we write:
Sh h Cf
= S tm = m = ...(14.79)
Re × Sc V 2
where, St m = h m /V is the Stanton number for mass transfer.
Remember that Reynolds analogy is valid only when Pr = Sc = 1.
When Pr (or, Sc) is different from unity, we have the Colburn analogy:
2
Cf
jH = St × ( Pr ) 3 = (0.5 < Pr < 50...for heat transfer...(14.80))
2
2
Cf
and, jM = Stm × (Sc) 3 = (0.6 < Sc < 3000...for mass transfer...(14.81))
2
where, jM is the Colburn factor for mass transfer.
Relations for Cf (= friction factor) are already given in Chapter on Forced Convection.
From Eqs. 14.80 and 14.81, we can write:

FG Sc IJ
2
St 3
Stm
=
H Pr K
FG IJ FG IJ
2 2
h Sc 3 a 3
i.e.
hm
= r × Cp ×
Pr H K
= r × C p×
D H K ...(14.82)

Now, non-dimensional number (a/D) = Le is known as ‘Lewis number’.


Therefore, we have:
2
h
= r × Cp × Le 3 ...(14.83)
hm
Above relation is useful in cases of simultaneous heat and mass transfer.
Air–water vapour mixtures are of special interest in air conditioning applications. For air–water vapour
mixtures, L e = 0.872 and L e 2/3 is nearly equal to unity. Therefore, for air–water vapour mixtures, the relation
between heat and mass transfer coefficients is conveniently expressed as:
h = r × Cp × h m (for air–water vapour mixture...(14.84))
Eq. 14.84 is known as Lewis relation and is normally used in air-conditioning applications.
Note: It should be remembered that the analogy between convection heat and mass transfer is valid only for low
mass flux conditions.
Example 14.14. Air at a temperature of 20°C, and RH of 40% flows over a water surface at a velocity of 1.5 m/s. Length
parallel to flow is 18 cm. Average surface temperature is 16°C. Estimate the amount of water evaporated per hr/m2 of
surface area. Partial pressure of water vapour at 20°C and 40% RH is 0.011 bar and at 16°C and saturated, the vapour
pressure is 0.017 bar. Viscosity and density of air are: 18.38 x 10-6 kg/ms and 1.22 kg/m3, respectively. Assume D = 0.256
cm2/s.
Solution. See Fig. Ex. 14.14.

Water surface, 16°C

Air, 20°C, 40% RH,


u = 1.5 m/s

L = 0.18 m

FIGURE Example 14.14 Convective mass transfer from surface of water

MASS TRANSFER 755


Data:
m
u := 1.5 m/s L := 0.18 m m := 18.38 × 10 –6 kg/(ms) r := 1.22 kg/m3 n := i.e. n = 1.507 × 10 –5 m2/s
r
Ru
D := 0.256 ´ 10– 4 m2/s Ru := 8314 J/kg mole K RH O := i.e. RH O = 461.889 J/kgK T1 := 16 + 273 K
2
18 2

2 5
T2 := 20 + 273 K A := 1 m pw1 := 0.017 × 10 Pa pw2 := 0.011 × 105 Pa
Reynolds number:
u× L
Re :=
n
i.e. Re = 1.792 ´ 104 (Less than 5 ´ 105...so, laminar)
Mass transfer coefficient:
For laminar flow over a surface, we have:
hmc × L
Sh = = 0.664 × Re 0.5 × Sc 0.33 (Sherwood number)
D
n
Now, Sc :=
D
i.e. Sc = 0.588
Therefore,
Sh := 0.664 × R e 0.5 × Sc 0.33
i.e. Sh = 74.623 (Sherwood number)
Sh × D
and, hmc := m/s (mass transfer coefficient based on concentration)
L
i.e. hmc = 0.011 m/s
Concentrations (i.e. densities):
pw1
Cw1 :=
RH O × T1
2

i.e. Cw1 = 0.013 kg/m3 (concentration (density) at temperature T1)


pw2
and, Cw2 :=
RH O × T2
2

i.e. Cw2 = 8.128 ´ 10 –3 kg/m3 (concentration (density) at temperature T2)


Therefore, mass diffusion of water is given by:
mw := h mc × A× (Cw1 – Cw2)
i.e. mw = 4.89 ´ 10 – 5 kg/s per m2 area
i.e. mw = 3600 = 0.176 kg/hr per m2.
Example 14.15. Air at a temperature of 25°C, and RH of 20% flows through a pipe of 20 mm ID with a velocity of 5 m/
s. The inside surface of the tube is constantly wetted with water such that a thin water film is maintained on the surface.
Determine the amount of water evaporated per m2 of surface area.
Given: n = 15.7 x 10-6 m2/s, Sc = 0.6, and D = 0.26 x 10-4 m2/s.
Solution. This is a case of convective mass transfer.
Data:
m := 5.0 m/s d := 0.02 m n := 15.7 × 10 –6 m2/s
r sat := 0.231 kg/m3 (density of water vapour at saturation at 25 deg.C)
Therefore, density of vapour in free stream at 25% RH:
r free_stream := 0.25 r sat
i.e. rfree_stream = 0.058 kg/m3 (density of water vapour at free stream conditions)
–4 2
D := 0.26 × 10 m /s Sc := 0.6 A := 1 m2
Reynolds number:
u× d
Re := (Reynolds number)
n
i.e. Re = 6.369 × 103 > 2300 (therefore, turbulent flow)

756 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Then, Sherwood number relation is:
Sh := 0.023 × Re 0.83 × Sc 0.44
i.e. Sh = 27.873
hmc × d
But, Sh =
D
where, h mc = mass transfer coefficient based on concentration (i.e. density) difference.
Therefore,
Sh × D
h mc :=
d
i.e. h mc = 0.036 m/s
Then, mass evaporaion rate:
m := h mc × A × (r sat – r free_stream ) kg/(sm 2 )
i.e. m = 6.278 ´ 10 – 3 kg/(sm2) (mass evaporation rate per m2 surface area.)
1
Note: One sq. m of surface area of the pipe is equivalent to a length of: = 15.915 metres.
p ×d
i.e. Mass evaporation rate per metre length of pipe is given by:
m
= 3.944 ´ 10 –4 kg/s per metre length.
15. 915
Example 14.16. In an experiment, when atmospheric air at Ta = 30 deg.C was blown past a wet bulb thermometer, the
wet bulb reading obtained was Tw = 20°C. What is the value of concentration of water vapour Ca in the free stream?
Also, determine the relative humidity (RH) of the free stream. (Note: RH is equal to the ratio of concentration Ca of water
vapour in free stream to the saturation concentration at the free stream temperature of 30°C, Csat.. Csat is obtained from
Steam Tables.)
Solution. This problem involves simultaneous heat and mass transfer.
Data:
Ta := 30°C (temperature of atmospheric air)
Tw := 20°C (wet bulb temperature)
In steady state, we can state the energy balance as:
heat transfer from air stream to wet cloth = latent heat of evaporation of water
i.e. h ×A× (Ta – Tw) = hm × A × (r w – r a)×h f g ...(a)
Remember Cw = rw
h
i.e. × (Ta – Tw) =( r w – r a) × h fg ...(b)
hm
But, the ratio, h/hm is given from Lewis relation, as:

FG IJ
2
2
h a 3

hm
= r × Cp × Le 3 = r × Cp ×
D H K
Then, Eq. b becomes:

FG a IJ
2
3
r × Cp ×
H DK × (Ta – Tw) =(r w – ra)× h f g ...(c)

30 + 20
Now, properties of air are evaluated at the film temperature T f =
2
i.e. properties at: Tf := 25°C
3
Air: r := 1.186 kg/m Cp := 1005 J/(kgC) a := 2.18 × 10 –5 m2/s
a
D := 0.26 × 10 –4 m2/s (from Tables) = 0.838 = Le = Lewis number
D
Water:
h fg := 2454.1 × 10 3 J/kg at 20°C
Saturated concentration of water vapour at Tw = 20°C is determined from:
pw × Mw 2339 × 18
rw = = = 0.01728
Ru × Tw 8314 × 293
MASS TRANSFER 757
Note that in the above, pw = saturation pressure = 2339 Pa, corresponding to Tw = 20°C, from Steam Tables.
i.e. rw = 0.01728 kg/m3 = Csat = saturation concentration (density) of water
vapour at wet bulb temperature of 20°C
Then, from Eq. c:

FaI
2

r ×C × G J
3
p
H DK × (Ta – Tw) = (r w – ra)× h f g ...(c)

FG a IJ
2
3

i.e. r a := r w –
r × Cp ×
H DK ×(Ta - Tw )

hf g
i.e. ra = 0.013 kg/m3 (concentration of water vapour in free stream)
Relative humidity:
Saturation concentration at Ta = 30°C: rsat := 0.0304 kg/m3 (from Steam Tables)
Therefore,
ra
RH :=
r sat
i.e. RH = 0.426 = 42.6% (relative humidity of free stream.)

14.14 Summary
Mass transfer is an important phenomenon with vast industrial applications.
Diffusion mass transfer occurs due to concentration difference and is similar to heat conduction. The gov-
erning law is the Fick’s law of diffusion, analogous to the Fourier’s law of conduction, i.e.
mB dCB
jB = = – DBC × kg/(sm2) ...(14.18)
A dx
DBC is the diffusion coefficient for species B in a mixture of B and C. Values of diffusion coefficients for gases,
liquids and solids was discussed.
Equimolal diffusion of gases in a binary mixture was studied. This is important in distillation process, and
in venting a gas from a pipe line to atmosphere.
Next, the diffusion of a gas in a stationary gas column was explained. This phenomenon has applications in
absorption and humidification. Evaporation of water vapour in a stationary column of air was studied, as an
example.
Transient diffusion was explained briefly. Equations for transient diffusion are written by analogy with
transient conduction.
Convective mass transfer involves transport of mass across the boundary and is affected by the flow field.
Analogous to ‘heat transfer coefficient’, a ‘mass transfer coefficient’ is also defined and the governing law is
similar to the Newton’s law of cooling.
Analogy between heat and mass transfer was explained and relations for convective mass transfer were
written for various geometries and flow conditions, by analogy with heat transfer relations under similar situa-
tions.
Finally, topic of simultaneous heat and mass transfer, which has important applications in the field of air
conditioning, was discussed.

Questions
1. State Fick’s law of mass transfer by diffusion and explain its analogy with Fourier’s law of conduction. [M.U.]
2. Define: (a) diffusion coefficient, (b) mass transfer coefficient.
3. Write a short note on diffusion coefficient in a binary mixture of: (a) gases (b) liquids, and (c) solids.
4. How does D depend on pressure and temperature in a binary gas mixture?
5. Derive a mass diffusion equation in general form in cartesian coordinates for mass diffusion in stationary me-
dium in the same lines as that of general heat conduction equation, in differential form.
Using the above general equation show that the governing differential equation for steady state diffusion
through a plane membrane reduces to the form:

758 FUNDAMENTALS OF HEAT AND MASS TRANSFER


d 2 CA
= 0
d x2
where, CA is the concentration of species A. [M.U.]
6. Define: Sherwood number, Schmidt number and Lewis number. [M.U.]
7. Derive a basic differential equation for equimolal counter-diffusion in gases and solve the same for constant
pressure situation to get the mass flux of species A as:
- D × MA (Pa 2 - Pa1 )
Mass flux = ×
Ru × T Dx
where, Pa1 and Pa2 are partial pressures of species A at x and (x + Dx) locations, MA is the molecular mass of
species A, T is the temperature of gases in Kelvin, and Ru is the Universal gas constant. [M.U.]
8. Prove that during isothermal evaporation, rate of mass transfer of water vapour into atmospheric air is given by:
D × P ×( Pa1 - Pa 2 )
NA = kg mole/(s m2)
Ru × T × D x × Pln
Notations in the above equation are as defined earlier. [M.U.]
9. Write a short note on the analogy between momentum, heat and mass transfer. [M.U.]
10. Explain in brief two main methods of mass transfer and bring out their differences. [M.U.]
11. Show that kinematic viscosity, thermal diffusivity and diffusion coefficient have the same units. [M.U.]
12. For convective mass transfer, name the non-dimensional number that plays the same role in mass transfer as
that of Prandtl number in heat transfer and write down an expression for the same. [M.U.]
13. Derive Stefan’s law (isothermal evaporation of water). [M.U.]
14. Derive an equation to determine the amount of mass transferred through a composite plane wall with one layer
of diffusivity D ab and another with diffusivity Dac , with concentrations Ca1 and Ca2 on either side of wall, wall
thicknesses Dx1 and D x2. [M.U.]

Problems
1. Air is contained in a vessel at a temperature of 20°C and pressure of 3 bar. Assuming the partial pressures of O2
and N2 to be in ratios of 0.21 and 0.79, respectively, calculate: (i) Molar concentrations (ii) Mass concentrations
(i.e. densities), (iii) Mass fractions, and (iv) Molar fractions.
2. Calculate the diffusion coefficient of CO2 in air at 20°C and 1 atm. pressure. Then, calculate the value of D for a
pressure of 3 atm. and temperature of 57°C.
3. A steel, rectangular container having walls 10 mm thick, is used to store gaseous hydrogen at elevated pressure.
The molar concentrations of hydrogen in steel at the inside and outside surfaces are 1.1 kg.mole/m3 and zero,
respectively. Assuming the diffusion coefficient for hydrogen in steel to be 0.25 ´ 10 – 12 m2/s, calculate the molar
diffusion flux for hydrogen through steel.
4. Hydrogen gas at 2 atm., 25°C is flowing through a rubber pipe, 25 mm ID, 50 mm OD. Solubility of H2 in rubber
is 0.053 cm3 of H 2 per cm3 of rubber at 1 atm. pressure. Diffusivity of H2 through rubber is 0.7 ´ 10 –4 m2/h. Find
the loss of hydrogen per metre length of pipe. [M.U.]
5. Hydrogen gas is stored at 358 K in a 3.0 m ID, 5 cm thick spherical container made of Nickel. Molar concentra-
tion of hydrogen in Ni at the inner surface is 0.12 kg.mole/m3 and is equal to zero at the outer surface. Deter-
mine the mass diffusion rate of hydrogen through the walls of the container. (Take D = 1.2 ´ 10 –12 m2/s)
6. Helium gas is stored at a pressure of 4 bar and 293 K in a 0.3 m ID, 3 mm thick spherical container made of
fused silica. Determine the rate of pressure drop due to diffusion. Given: D = 0.04 ´ 10 –12 m2/s, and solubility of
gas at the solid surface on the inside is 18 ´ 10 –9 kg/(m3 Pa).
7. In problem 6, if the container is a long cylinder of diameter 0.3 m, calculate mass of helium lost by diffusion per
metre length. Rest of the data are same.
8. A gas mixture consists of oxygen and nitrogen at 1 bar and 27°C. The oxygen content, by volume, at two planes
3 mm apart are 15 % and 30%, respectively. Calculate the rate of diffusion in kg mole/(sm2), if:
(i) nitrogen is non-diffusing
(ii) there is equimolar counter-diffusion of the two gases.
Take D = 0.181 cm2/s. [M.U.]
9. A tank contains a mixture of CO2 and N2, in the mole proportions of 0.3 and 0.7 at 1 bar and 290 K. It is
connected by a duct of cross-sectional area 0.1 m2 to another tank containing a mixture of CO2 and N2 in the
molal proportions of 0.7 and 0.3. The duct is 0.75 m long. Determine the diffusion rates of CO2 and N2 in kg/s.
Given: D = 0.16 ´ 10 – 4 m2/s for CO2/N2 at 293 K from tables.

MASS TRANSFER 759


10. A spherical ball of ice, 1.5 cm diameter is suspended in still dry air at 1.013 bar. Calculate the initial rate of
evaporation at the surface.
Take D = 0.256 ´ 10 –4 m2/s. At 0 deg.C, saturated vapour pressure = 0.0061 bar.
11. Each of two large vessels contains uniform mixture of nitrogen and carbon dioxide at 1 bar and 288.9 K. Vessel
1 contains 90 mole % of N2 and 10 mole % CO2, whereas vessel 2 contains 20 mole % N2 and 80 mole % CO2.
The two vessels are connected by a duct of 0.15 m ID and 1.22 m long. Determine the rate of transfer of N2
between the two vessels in kg/s, assuming steady state transfer. Mass diffusivity for N2–CO2 mixture at 1 bar
and 288.9 K may be taken as:
D = 0.16 ´ 10 –4 m2/sec. [M.U.]
12. Estimate the evaporation rate of water, which is available at the bottom of a well 2.5 m diameter and 5 m deep,
into dry atmospheric air at 25°C. The diffusion coefficient is 0.0925 m2/h and the atmospheric pressure is 1 bar.
The partial pressure of water at the water surface is 0.0312 bar. [M.U.]
13. A pan 20 mm diameter 20 mm deep, is filled with water to a level of 10 mm and is exposed to dry air at 40°C.
Calculate the time required for all water to evaporate. What will be the change in time required if the temp of air
is 30°C? D = 0.256 cm2/s? [M.U.]
14. Estimate the diffusion rate of water at 27°C in a test tube 20 mm diameter 5 cm deep, into dry air at same
temperature. Take D = 0.26 cm2/s. Saturated vapour pressure of water at 27°C = 0.035 bar. [M.U.]
15. Water at 20°C is spilled in a room. Thickness of water layer is 1 mm. Absolute humidity of air is 3 g of vapour
per kg of dry air. Calculate time required for complete evaporation of water spilled, if evaporation is by molecu-
lar diffusion through an air film of 5 mm thickness.
Atmospheric temperature and pressure are 1 bar and 20°C, respectively. Assume surface area of floor as 1 m2.
Take D = 0.26 ´ 10 –4 m2/s. [M.U.]
[Hint: Remember:
pw 2 m M
= w× a
p a2 ma M w
and, pw2 + pa2 = 1, where mw = mass of water vapour ma = mass of dry air, Ma and Mw are the molecular weighs
of air and water vapour, respectively.]
16. A mild steel piece has uniform, initial carbon concentration of 0.15% by mass. It is exposed to a carburising
atmosphere in a furnace, where the surface concentration is maintained at 1.2%. Determine how long the piece
must be kept in the furnace for the concentration of carbon at a location 0.4 mm below the surface to reach 1%.
Take D = 5 ´ 10 –10 m2/s.
17. Air at 1 atm. and 25°C, containing small quantities of iodine, flows with a velocity of 4.5 m/s inside a 5 cm
diameter tube. Determine the mass transfer coefficient for iodine transfer from the air stream to the surface.
Assume: D = 0.82 ´ 10 –5 m2/s; n = 15.5 ´ 10 –6 m2/s.
18. Air at a temperature of 21°C, and RH of 40% flows over a water surface at a velocity of 1.2 m/s. Length parallel
to flow is 15 cm. Average surface temperature is 15°C. Estimate the amount of water evaporated per hr/m2. of
surface area. Partial pressure of water vapour at 21°C and 40% RH is 0.011 bar and at 15°C and saturated, the
vapour pressure is 0.017 bar. Viscosity and density of air are: 18.38 x 10-6 kg/m.s and 1.22 kg/m3, respectively.
Assume D = 0.256 ´ 10 –4 m2/s, po = 1.013 bar, and
1
Sh = 0.023 × R e 0.8 × Sc 3 [M.U.]
19. Air at a temperature of 30°C, and RH of 15% flows through a pipe of 15 mm ID with a velocity of 5 m/s. The
inside surface of the tube is constantly wetted with water such that a thin water film is maintained on the
surface. Determine the amount of water evaporated per sq.m of surface area.
Given: n = 16 ´ 10 – 6 m2/s, Sc = 0.6, and D = 0.26 ´ 10 –4 m2/s.
20. Atmospheric air at Ta = 50 deg.C was blown past a wet bulb thermometer, the wet bulb reading obtained was Tw
= 30°C. What is the value of concentration of water vapour Ca in the free stream? Also, determine the relative
humidity (RH) of the free stream. (Note: RH is equal to the ratio of concentration Ca of water vapour in free
stream to the saturation concentration at the free stream temperature of 50°C, Csat . Csat is obtained from Steam
Tables.)

760 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Appendix

TABLE A-1 Conversion factors


Quantity Multiply by to get
Length m 3.2808 ft
ft 0.304804 m
Area m2 0.764 ft2
ft 2 1.308901 m2
Volume m3 35.34 ft3
ft 3 0.028297 m3
Mass kg 2.3046 lb
lb 0.433915 kg
Density kg/m 3 0.062428 lb/ft3
lb/ft3 16.01845 kg/m3
Temperature K 0.555556 R
R 1.8 K
Mass transfer coefficient m/s 11811 ft/h
ft/h 8.47E-05 m/s
Volume flow rate m3 /s 127130 ft3/h
ft3/h 7.87E-06 m3/s
Acceleration m/s 2 42520000 ft/h2
ft/h2 2.35E-08 m/s2
Energy J 0.000948 Btu
Btu 1054.997 J
Force N 0.22481 lbf
lbf 4.448201 N
Heat transfer rate W 3.4123 Btu/h
Btu/h 0.293057 W
Heat flux W/m2 0.3171 Btu/h ft2
Btu/h ft2 3.153579 W/m 2
Heat generation rate W/m3 0.09665 Btu/h ft3
Btu/h ft3 10.34661 W/m 3
Heat transfer coefficient W/m 2K 0.17612 Btu/h ft2 F
Btu/h ft2 F 5.677947 W/m2 K
Kinetic viscosity & diffusivity m2 /s 38750 ft2/h
ft2/h 2.58E-05 m2/s
Latent heat J/kg 0.00043 Btu/lbm
Btu/lbm 2325.852 J/kg
Contd.
Contd.

Mass flow rate kg/s 7936.6 lbm/h


lbm/h 0.000126 kg/s
Pressure and stress N/m2 0.020886 lbf/ft2
lbf/ft 2 47.87896 N/m2
Specific heat J/kgK 0.000239 Btu/lbmF
Btu/lbmF 4186.553 J/kgK
Thermal conductivity W/mK 0.57782 Btu/h ft F
Btu/h ft F 1.730643 W/mK
Thermal resistance K/W 0.5275 F/hBtu
F/h.Btu 1.895735 K/W
Dynamic viscosity kg/ms 2419.1 lbm/fth
lbm/fth 0.000413 kg/ms

762 FUNDAMENTALS OF HEAT AND MASS TRANSFER


Bibliography

Agrawal, Shyam K. Applied Thermosciences: Principles and Applications. New Delhi: Viva Books.
Arora, S. C., S. Domkundwar and A. V. Domkundwar, A Course in Heat and Mass Transfer. Delhi: Dhanpat Rai.
Becker, Martin. Heat Transfer: A Modern Approach. Plenum Press.
Bird, Stewart and Lightfoot, Transport Phenomena. John Wiley.
Cengel, Yunus A. Heat Transfer: A Practical Approach, McGraw Hill.
——— Introduction to Thermodynamics and Heat Transfer. McGraw Hill.
Chapman, Alan J. Heat Transfer, Macmillan.
Gebhart, Benjamin. Heat Transfer. New York: McGraw Hill.
Gupta, C. P. and Rajendra Prakash. Engineering Heat Transfer. Nem Chand & Bros.
Holman, J. P. Heat Transfer. McGraw Hill.
Incropera, Frank P. and David P. Dewitt, Fundamentals of Heat and Mass Transfer. John Wiley.
Isachenko, V., V. Osipova and A. Sukomel, Heat Transfer. Moscow: Mir Publishers.
Kays and London, Compact Heat Exchangers. McGraw Hill.
Kays, W. M. Convective Heat and Mass Transfer. McGraw Hill.
Kern, D. Q. Process Heat Transfer. McGraw Hill.
Kothandaraman, C. P. Fundamentals of Heat and Mass Transfer. New Delhi: New Age.
Kreith, Frank. Principles of Heat Transfer. International Text Book Company.
Kumar, D. S. Heat and Mass Transfer. S. K. Kataria & Sons.
Long, Christopher A. Essential Heat Transfer. Delhi: Pearson Education.
Madams, W. H. Heat Transmission. McGraw Hill.
Mikheyev, M. Fundamentals of Heat Transfer. Moscow: Peace Publishers.
Nag, P. K. Heat Transfer. New Delhi: Tata McGraw Hill.
Ozisik, M. N. Heat Transfer: A Basic Approach. McGraw Hill.
Pitts, Donald R. and Leighton E. Sisson. Heat Transfer (Schaum Outline Series). New York: McGraw Hill.
Rajput, R. K. Heat and Mass Transfer. New Delhi: S. Chand.
Rathore, M. M. Comprehensive Engineering Heat Transfer. New Delhi: Laxmi Publications.
Rogers, G. F. C. and Y. R. Mayers. Engineering Thermodynamics, Work and Heat Transfer. English Language Book
Society.
Sachdeva, R. C. Fundamentals of Engineering Heat and Mass Transfer. New Delhi: New Age.
Schenk, H. Heat Transfer Engineering. Longmans Green.
Sonntag, R. E., C. Borgnakke and G. J. van Wylen. Fundamentals of Thermodynamics. John Wiley.
Sucec, James. Heat Transfer. W.M.C. Brown.
Welty, James R., Robert E. Wilson and Charles E. Wicks. Fundamentals of Momentum, Heat and Mass Transfer. John
Wiley.
Index

A correction factors for multi-pass and cross-flow heat


analogy between momentum and heat transfer, 429 exchangers, 600
analysis with variable thermal conductivity, 200 counter-flow heat exchanger, 591
application of fin theory for error estimation, 261 criteria for lumped system analysis, 269
applications of heat transfer, 1 critical thickness of insulation, 101
cylinder with uniform internal heat generation, 166
cylindrical systems, 74
B
basic conduction relations, with heat generation, 215
basic equations for forced convection, 469
D
basic equations for natural convection, 519 dielectric heating, 204
boiling and condensation, 29 differential equations for the boundary layer, 390
boiling and evaporation, 530 dimensional analysis of natural convection, 478
boiling heat transfer, 530 dimensional analysis, 394
boiling modes, 530 dimensionless parameters in boiling and condensation,
529
boiling regimes and boiling curve, 531
drop-wise condensation, 574
boundary and initial conditions, 31
boundary layer equations, exact solutions of, 402
burnout phenomenon, 532 E
effect of variable thermal conductivity, 113
C electrical network method, 676
emissivity, real surface and grey surface, 651
combined heat transfer mechanism, 8
empirical relations for natural convection, 484
combined natural and forced convection, 516
equimolal counter-diffusion in gases, 740
compact heat exchangers, 622
composite cylinders, 79
composite spheres, 95 F
concentrations, velocities and fluxes, 723 Fick’s law of diffusion, 725
condensation heat transfer, 550 film condensation and flow regimes, 551
conduction with variable area, 66 film condensation inside horizontal tubes, 573
conduction, 3, 22 fin effectiveness, 255
conservation of energy equation for the boundary layer, fin efficiency, 250
392 fin formulae, 233
conservation of mass, 390 fin of finite length losing heat from its end by convection,
conservation of momentum equation, 391 229
convection boundary condition, 33 fin of finite length with insulated end, 226
convection, 5, 23 fin of finite length with specified temperature at its end,
convective mass transfer, 752 231
finite difference formulation from differential equations, I
330 infinitely long fin, 224
fins of non-uniform cross section, 248 insulation systems, 21
fins of uniform cross section, 222 interface boundary condition, 34
flow across a bank of tubes, 440
flow across bluff objects, 436
K
flow across cylinders and spheres, 432
Kirchhoff’s law, 653
flow across cylinders, spheres and other bluff shapes and
packed beds, 431
flow boiling, 542 L
flow inside tubes, 445 laws of black body radiation, 645
flow through packed beds, 436 LMTD method for heat exchanger analysis, 589
fluxes, 725 lumped system analysis, 267
forced convection, 382
Fourier’s law of heat conduction, 13 M
free convection from rectangular blocks and short mass transfer coefficient, 751
cylinders, 499 mass transfer, 11, 723
free convection from spheres, 498 methods of determining view factors, 661
free convection in enclosed spaces, 501 methods of solving a system of simultaneous, algebraic
free convection in inclined spaces, 504 equations, 344
fully developed laminar flow inside pipes of non-circular methods to determine convective heat transfer coefficient,
cross-sections, 453 393
fundamental laws of heat transfer, 2 mixed boundary condition, 276

G N
gaseous emission and absorption, 712 natural (or free) convection, 477
governing equations and solution by integral method, natural convection from finned surfaces, 512
480 natural convection in turbine rotors, rotating cylinders,
Grashoff number, 478 disks and spheres, 508
natural convection inside concentric cylinders and
H spheres, 506
heat exchangers, 78 natural convection inside spherical cavities, 505
heat exchangers, types of, 578 Newton’s law of cooling and heat transfer coefficient, 384
heat transfer considerations in a pipe, 448 NTU method for heat exchanger analysis, 604
heat transfer correlations for pool boiling, 533 NTU relation for a counter-flow heat exchanger, 606
heat transfer from extended surfaces, 221 NTU relation for a parallel-flow heat exchanger, 605
heat transfer in boiling and condensation, 10 numerical methods for transient heat conduction, 363
heat transfer in nuclear fuel rod with cladding, 212 numerical methods in heat conduction, 329
heat transfer in nuclear fuel rod without cladding, 208 Nusselt number, 384
heat transfer in transient conduction, 310 Nusselt’s theory for laminar film condensation on
heat transfer through a piston crown, 207 vertical plates, 552
heat transfer through composite slabs, 50
heat transfer, modes of, 2 O
Heisler and Grober charts, 284 one-term approximation solutions, 281
hollow cylinder with heat generation, 175 one-dimensional steady state conduction in cylindrical
hollow cylinder with variable thermal conductivity, 121 systems, 348
hollow sphere with variable thermal conductivity, 129 one-dimensional steady state conduction in spherical
horizontal cylinder at constant temperature, 494 systems, 352
horizontal plate at constant temperature, 491 one-dimensional steady state heat conduction in Carte-
horizontal plate with constant heat flux, 493 sian coordinates, 331
hydrodynamic and thermal boundary layers for flow in a one-dimensional steady state heat conduction with heat
tube, 445 generation, 147
hydro-mechanical design of heat exchangers, 629 one-dimensional steady state heat conduction, 47

INDEX 765
one-dimensional transient conduction in semi-infinite steady and unsteady heat transfer, 10
solids, 300 steady state diffusion in common geometries, 731
one-dimensional transient heat conduction in a plane steady state diffusion through a cylindrical shell, 732
wall, 365 steady state diffusion through a plain membrane, 731
operating-line/equilibrium-line method, 620 steady state diffusion through a spherical shell, 734
optimum (or economic) thickness of insulation, 109 steady state unidirectional diffusion—diffusion of water
origin and growth of bubbles, 530 vapour through air, 744
overall heat transfer coefficient for cylindrical system, 82 steady-state diffusion in liquids, 747
overall heat transfer coefficient for spherical system, 97 steady-state equimolal counter-diffusion in liquids, 747
overall heat transfer coefficient, 53, 581 steady-state unidirectional diffusion in liquids, 748
Stefan–Boltzmann law, 648
P
parallel flow heat exchanger, 589 T
performance of fins, 250 temperature distribution in transient conduction, 308
physical mechanism of forced convection, 382 thermal boundary layer, 388
physical mechanism of natural convection, 477 thermal conductivity of gases, 19
Planck’s law for spectral distribution, 645 thermal conductivity of liquids, 18
plane slab with uniform internal heat generation, 147–55 thermal conductivity of materials, 14
plane slab with variable thermal conductivity, 113 thermal conductivity of solids, 14
plane slab, 47 thermal contact resistance, 63
prescribed heat flux at the boundaries, 32 thermal diffusivity (a), 24
prescribed temperatures at the boundaries, 32 thermal resistance of a fin, 256
properties of view factor and view factor algebra, 659 thermal resistance, concept of, 22
thermodynamics and heat transfer, 1
R total surface efficiency, 257
radiation error in temperature measurement, 708 transient heat conduction in multi-dimensional systems,
radiation exchange between small, grey surfaces, 676 308
radiation from a wave band, 649 transient heat conduction, 266
radiation from gases, vapours and flames, 712 transient mass diffusion in common geometries, 751
radiation heat exchange between grey surfaces, 675 transient mass diffusion in semi-infinite, stationary
radiation heat exchange in four-zone enclosures, 691 medium, 748
radiation heat exchange in three-zone enclosures, 688 turbulent flow inside pipes, 454
radiation heat exchange in two-zone enclosures, 679 two-dimensional conduction, shape factor, 134
radiation heat transfer coefficient (hr), 711 two-dimensional steady state conduction in Cartesian
coordinates, 356
radiation shielding, 698
two-dimensional transient heat conduction, 372
radiation, 6, 23, 641
relation between radiation intensity and emissive power,
649 V
response time of a thermocouple, 271 velocities, 724
Reynolds and Colburn analogies for mass transfer, 754 velocity boundary layer, 385
velocity profile for fully developed, steady, laminar flow,
446
S
vertical cylinders at constant temperature, 485
simplified calculations for water, 560
vertical plate at constant temperature, 484
simplified correlations for boiling with water, 538
vertical plate with constant heat flux, 485
simplified equations for air, 501
view factor and radiation energy exchange between black
solar and atmospheric radiation, 717
bodies, 657
solid cylinder with internal heat generation, 167
volumetric absorption and emissivity, 712
solid sphere with internal heat generation, 197
von Karman integral equations, 408
solutions of boundary layer equations, 408
sphere with uniform internal heat generation, 197
spherical systems, 91 W
Wein’s displacement law, 647

766 INDEX

You might also like