Você está na página 1de 48

Kurdistan Iraqi Region

Ministry of Higher Education


University of Sulaimani
College of Science
Physics Department

Numerical Study of
Heat Flow in Material

Prepared by

Sakar M. Mhedin Hawdeng K. Muhamed


Soma S. Muhamed Amin

Supervised by

Dr. Omed Ghareb Abdullah

2009 - 2010

i
Acknowledgments

Praise to God that afford us to reach this work, we would like to thanks
our supervisor Dr. Omed Ghareb for his support and direction throughout
this work. We appreciate for Physics Department in College of Science at
University of Sulaimani, which giving us many information about this
work we want to express a great thanks to all lectures whom give us
information along our study, also we appreciation for those people that
helped us even by one word especially Soma’s brother Halwest, and deeps
gratitude to our colleagues for their encouraging.
Finally thanks to our Family for their support and encourages during
our study.

Sakar , Hawdeng and Soma

ii
Contents

Chapter One: Basic Concepts


1.1 Introduction.

1.2 Basics of Heat Transfer.

1.3 Heat Transfer Mechanisms.

1.4 Thermal Conductivity.

1.5 Umklapp and Normal Processes.

1.6 Heat Capacity.

1.7 Conduction Heat Transfer.

1.8 Convection Heat Transfer.

Chapter Two: Heat Flow Analyzing


2.1 Introduction.

2.2 Derivation of the Heat Equation.

2.3 Derivation of the Heat-Conduction Equation.

2.3 Derivation of the Heat-Conduction Equation.

2.4 Method of Analysis.

2.5 Analytical Solution.

2.6 Numerical Method of Analysis.

iii
Chapter three: Results and Discussion
3.1 Introduction.

3.2 Comparison of Analytical- and Numerical- Solution.

3.3 Temperature Distribution in Plate with Fixed Boundary

Temperatures.

3.4 Temperature Distribution in Plate with Convection

Environment Temperature.

3.5 Temperature Distribution in Plate with Composite

Materials.

3.6 Conclusions.

References

iv
Abstract

The determination of temperature distribution in a medium is the main


objective of a conduction analysis, that is, to know the temperature in the
medium as a function of space at steady state and as a function of time during
the transient state. Once this temperature distribution is known, the heat flux
at any point within the medium, or on its surface, may be computed from
Fourier’s law. Knowledge of the temperature distribution within a solid can
be used to determine the structural integrity via a determination of the thermal
stresses and distortion. The optimization of the thickness of an insulating
material and the compatibility of any special coatings or adhesives used on
the material can be studied by knowing the temperature distribution and the
appropriate heat transfer characteristics.
The general heat conduction equation can be solved either analytically or
numerically, given the proper set of boundary conditions, to yield the
temperature distribution in the computational domain. Each method has its
own pros and cons. Analytic methods revolve around finding an exact
solution for the differential equation by integration, Fourier or Laplace
transforms, etc. The drawback is that problems with irregular computational
domains and/or time varying physical properties are unsolvable by these
methods.
In this study the numerical solution of the heat flow for the temperature
distribution in two dimensional plates were investigated. The finite difference
method is extended, to enable calculation of temperature distribution and flow
in composite materials for different boundary conditions. The effect of
thermal convections and the ratio of thermal conductivity of composite
material are also investigated.
The results show that the finite difference method is simple and powerful
method to obtain a temperature distribution.

v
Chapter One
Basic Concepts

1.1 Introduction:
Heat transfer is the transition of thermal energy from a hotter object to a
cooler object ("object" in this sense designating a complex collection of
particles which is capable of storing energy in many different ways). When an
object or fluid is at a different temperature than its surroundings or another
object, transfer of thermal energy, also known as heat transfer, or heat
exchange, occurs in such a way that the body and the surroundings reach
thermal equilibrium; this means that they are at the same temperature. Heat
transfer always occurs from a higher-temperature object to a cooler-
temperature one as described by the second law of thermodynamics or the
Clausius statement. Where there is a temperature difference between objects
in proximity, heat transfer between them can never be stopped; it can only be
slowed.
Conduction is the transfer of heat by direct contact of particles of matter.
The transfer of energy could be primarily by elastic impact as in fluids or by
free electron diffusion as predominant in metals or phonon vibration as
predominant in insulators. In other words, heat is transferred by conduction
when adjacent atoms vibrate against one another, or as electrons move from
atom to atom. Conduction is greater in solids, where atoms are in constant
contact. In liquids (except liquid metals) and gases, the molecules are usually
further apart, giving a lower chance of molecules colliding and passing on
thermal energy.
Heat conduction is directly analogous to diffusion of particles into a fluid,
in the situation where there are no fluid currents. This type of heat diffusion
differs from mass diffusion in behaviour, only in as much as it can occur in
solids, whereas mass diffusion is mostly limited to fluids.

1
Metals (e.g. copper, platinum, gold, iron, etc.) are usually the best
conductors of thermal energy. This is due to the way that metals are
chemically bonded: metallic bonds (as opposed to covalent or ionic bonds)
have free-moving electrons which are able to transfer thermal energy rapidly
through the metal.
As density decreases so does conduction. Therefore, fluids (and especially
gases) are less conductive. This is due to the large distance between atoms in
a gas: fewer collisions between atoms means less conduction. Conductivity of
gases increases with temperature. Conductivity increases with increasing
pressure from vacuum up to a critical point that the density of the gas is such
that molecules of the gas may be expected to collide with each other before
they transfer heat from one surface to another. After this point in density,
conductivity increases only slightly with increasing pressure and density.

1.2 Basics of Heat Transfer:


In the simplest of terms, the discipline of heat transfer is concerned with
only two things: temperature, and the flow of heat. Temperature represents
the amount of thermal energy available, whereas heat flow represents the
movement of thermal energy from place to place.
On a microscopic scale, thermal energy is related to the kinetic energy of
molecules. The greater a material's temperature, the greater the thermal
agitation of its constituent molecules (manifested both in linear motion and
vibrational modes). It is natural for regions containing greater molecular
kinetic energy to pass this energy to regions with less kinetic energy.
Several material properties serve to modulate the heat tranfered between
two regions at differing temperatures. Examples include thermal
conductivities, specific heats, material densities, fluid velocities, fluid
viscosities, surface emissivities, and more. Taken together, these properties
serve to make the solution of many heat transfer problems an involved
process.

2
1.3 Heat Transfer Mechanisms:

Heat transfer mechanisms can be grouped into three broad categories:

1- Conduction:
Regions with greater molecular kinetic energy will pass their thermal
energy to regions with less molecular energy through direct molecular
collisions, a process known as conduction. In metals, a significant portion of
the transported thermal energy is also carried by conduction-band electrons.

2- Convection:
When heat conducts into a static fluid it leads to a local volumetric
expansion. As a result of gravity-induced pressure gradients, the expanded
fluid parcel becomes buoyant and displaces, thereby transporting heat by fluid
motion (i.e. convection) in addition to conduction. Such heat-induced fluid
motion in initially static fluids is known as free convection.

3- Radiation:
All materials radiate thermal energy in amounts determined by their
temperature, where the energy is carried by photons of light in the infrared
and visible portions of the electromagnetic spectrum. When temperatures are
uniform, the radiative flux between objects is in equilibrium and no net
thermal energy is exchanged. The balance is upset when temperatures are not
uniform, and thermal energy is transported from surfaces of higher to surfaces
of lower temperature.

3
1.4 Thermal Conductivity:
To quantify the ease with which a particular medium conducts, engineers
employ the thermal conductivity, also known as the conductivity constant or
conduction coefficient K. Thermal conductivity K is defined as "the quantity
of heat Q, transmitted in time (t) through a thickness (L), in a direction normal
to a surface of area (A), due to a temperature difference (ΔT)", see figure
(1.1).

Figure (1.1): One-Dimensional Heat Flow.

Thermal conductivity is a material property that is primarily dependent on


the medium's phase, temperature, density, and molecular bonding. The
thermal conductivity can be found using this equation:
𝐾𝐾 = 13 𝐶𝐶 𝑣𝑣 𝑙𝑙 (1.1)
where C is the specific heat, v is the velocity, and l is the mean free path
of the phonon. A phonon can be scattered by an electron, a defect (including a
boundary), and other phonons. Such scattering will shorten the mean free
path.

4
Three important mechanisms are to be considered:
(a) collision of a phonon with other phonons
(b) collision of a phonon with imperfections in the crystal
(c) collision of a phonon with the external boundaries of the crystal
The phonon-phonon scattering is due to the anharmonic interaction. At
high T the mean free path 𝑙𝑙 ∝ 1/𝑇𝑇 : number of phonons 𝑛𝑛 ∝T at high T
collision frequency ∝ 𝑛𝑛 ⇒ 𝑙𝑙 ∝ 1/𝑛𝑛.
Table (1.1) gives the values of thermal conductivity for a variety of
materials

Table (1.1): Thermal Conductivities of Selected Materials at Room


Temperature.
Material Thermal Conductivity, W/m K
Copper 401
Silver 429
Gold 317
Aluminum 237
Steel 60.5
Limestone 2.15
Bakelite 1.4
Water 0.613
Air 0.0263

1.5 Umklapp and normal processes:


Normal process: phonon collisions that conserve k-momentum (crystal
momentum) do not allow relaxation (N-processes) or Momentum is
conserved i.e. k final = k initial . This processes involve small momentum
exchange between phonons and have dominant contribution at low
temperature.
Umklapp processes: (U-processes) do allow relaxation and limit thermal
conductivity. This is because the medium is discrete, and k-values outside the
first Brillouin zone can be mapped back into the first Brillouin zone with a
reciprocal lattice vector, or (U-processes) is an anharmonic phonon-phonon

5
(or electron-phonon) scattering process, creating a phonon with a momentum
k-vector outside the first Brillouin zone. Umklapp scattering is one process
limiting the thermal conductivity in crystalline materials, the others being
phonon scattering on crystal defects and at the surface of the sample.
Figure (1.2) schematically shows the possible scattering processes of two
incoming phonons with wave-vectors (k-vectors) k 1 and k 2 (red) creating one
outgoing phonon with a wave vector k 3 (blue). As long as the sum of k 1 and
k 2 stay inside the first Brillouin zone (gray squares) k 3 is the sum of the
former two conserving phonon momentum. This process is called normal
scattering (N-process).

Figure (1.2): Normal process (N-process) and Umklapp process (U-process).

With increasing phonon momentum and thus wave vector of k 1 and k 2


their sum might point outside the Brillouin zone (k' 3 ). As shown in Figure
(1.2), k-vectors outside the first Brillouin zone are physically equivalent to
vectors inside it and can be mathematically transformed into each other by the
addition of a reciprocal lattice vector G. These processes are called Umklapp
scattering and change the total phonon momentum.
Umklapp scattering is the dominant process for thermal resistivity at high
temperatures for low defect crystals, as shown in figure (1.3).

6
Figure (1.3): Umklapp Scattering, Scattering that changes total crystal
momentum by a reciprocal lattice vector.

The thermal conductivity for an insulating crystal where the U-processes


are dominant has 1/T dependence.

𝑘𝑘�⃑̀ = 𝑘𝑘�⃑ + 𝑞𝑞⃑ 𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 (1.2)

𝑘𝑘�⃑̀ = 𝑘𝑘�⃑ + 𝑞𝑞⃑ + 𝐺𝐺⃑ 𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈𝑈 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 (1.3)


Suppose that two phonons of vectors q1 and q2 collide, and produce a
third phonon of vector q3.
Momentum conservation: q3 = q1 + q2
q3 may lie inside the Brillouin zone, or not. If it's inside → momentum of
the system before and after collision is the same. This is a normal process. It
has no effect at all on thermal resistivity, since it has no effect on the flow of
the phonon system as a whole.
If q3 lies outside the Brillouin zone, we reduce it to equivalent q4 inside
the first Brillouin zone: q3 = q4 + G
Momentum conservation: q1 + q2 = q4 + G
The difference in momentum is transferred to the center of mass of the
lattice. This type of process is is known as the umklapp process
•highly efficient in changing the momentum of the phonon
•responsible for phonon scattering at high temperatures

7
1.6 Heat capacity:
At high temperatures, the heat capacity tends to 3Nk where N is the
number of atoms in the material. This is called the law of Dulong and Petit. It
turns out that this can be understood classically in terms of the equipartition
theorem. Unfortunately, classical mechanics also predicts that the heat
capacity should be independent of temperature. Despite many heroic efforts,
no satisfactory classical explanation for the way in which the specific heat
varies with temperature has ever been found and Maxwell recognized this as a
very severe problem for classic mechanics. The classical theory for the
specific heat of solids does not explain the decrease of specific heat at low
temperatures. The physical models of the specific heat curves as given by
Einstein and subsequentlyby Debye employed the quantum theory and agreed
well with experiment.

Albert Einstein applied in these idea on the problem of heat capacities of


solids. Defining the dimensionless parameter 𝑥𝑥 = 𝜃𝜃𝐸𝐸 /𝑇𝑇 the form of the heat
capacity is universal – it does not depend on the specific matter.
𝐶𝐶𝑣𝑣 = 3𝑁𝑁𝑘𝑘𝐵𝐵 𝑥𝑥 2 𝑒𝑒 𝑥𝑥 ⁄(1 − 𝑒𝑒 𝑥𝑥 )2 (1.4)
This is a nice presentation of a scaling, i.e. an expression that depends on
a general parameter (usually dimensionless). It was Einstein who first showed
that quantum mechanics could resolve this outstanding problem and explain
the approximate temperature dependence of the heat capacity. Even though
we can argue from rather general principles that 𝐶𝐶𝑣𝑣 must tend to zero as T
becomes small, we shall see that the explanation for this arises from the
discreteness of the energy levels.
Despite the successes of Einstein's theory, it failed to predict in detail the
way in which the heat capacity actually approaches zero as T→ 0 In order to
examine the low-temperature region in more detail, it is 𝑇𝑇 2 preferable to plot
𝐶𝐶𝑉𝑉 ⁄𝑇𝑇 rather than T. It is found experimentally that if 𝑐𝑐𝑣𝑣 ⁄𝑇𝑇 is plotted as a
function of

8
𝐶𝐶𝑉𝑉
= 𝛼𝛼𝛼𝛼 + 𝛾𝛾 (1.5)
𝑇𝑇

𝐶𝐶𝑉𝑉=𝛼𝛼𝑇𝑇 3 +𝛾𝛾𝛾𝛾 (1.6)


For temperatures very large compared with the Einstein temperature
𝐶𝐶𝑉𝑉 ≈ 3𝑁𝑁𝑁𝑁 (1.7)
Thus the high temperature limit of Einstein’s equation gives the value of
Dulong and Petit.
The failure of their law becomes evident when we examine the low
temperature limit. For 𝜃𝜃𝐸𝐸 ⁄𝑇𝑇 ≫ 1,
𝜃𝜃 2 −𝜃𝜃 𝐸𝐸 ⁄𝑇𝑇
𝐶𝐶𝑉𝑉 = 3𝑁𝑁𝑁𝑁 � 𝐸𝐸 � 𝐸𝐸 𝑒𝑒 (1.8)
𝑇𝑇

As T approaches zero, 𝐶𝐶𝑉𝑉 also goes to zero, since the exponential decay
overpowers the growth of (𝜃𝜃𝐸𝐸 ⁄𝑇𝑇)2
Einstein’s theory also explains the low heat capacities of some elements at
moderately high temperatures. If an element has a large Einstein temperature,
the ratio 𝜃𝜃𝐸𝐸 ⁄𝑇𝑇 will be large even for temperatures well above absolute zero,
and will 𝐶𝐶𝑉𝑉 be small. This equation is known as Debye’s (𝑇𝑇 3 law).
12𝜋𝜋 4 𝑇𝑇 3
𝐶𝐶𝑣𝑣 = NK� � (1.9)
5 𝜃𝜃𝐷𝐷

It is valid when the temperature is lower than about 0.1 𝜃𝜃𝐷𝐷 , which means
for most substances about 10-20 K. The relation gives a better fit valid for all
monatomic solids. When the temperature is above the Debye temperature, the
heat capacity is to experimental data at very low temperatures than the
Einstein model, and is very nearly equal to the classical value 3NK Heat
capacity of Debye:
In the Debye approximation, it is given by
𝑇𝑇 3 𝜃𝜃 ⁄𝑇𝑇 𝑒𝑒 𝑥𝑥 𝑥𝑥 4
𝐶𝐶𝑣𝑣 (𝑇𝑇) = 9𝑅𝑅 � � ∫0 𝐷𝐷 (𝑒𝑒 𝑥𝑥 −1)2
𝑑𝑑𝑑𝑑 (1.10)
𝜃𝜃 𝐷𝐷

where 𝜃𝜃𝐷𝐷 is the Debye temperature of the solid, 𝑇𝑇 is the absolute


temperature, and 𝑅𝑅 is the gas constant. This Demonstration shows the
variation of the specific heat of solids with temperature of representative
solids according to the Debye theory.

9
Debye defined a value with dimension of temperature, 𝜃𝜃𝐷𝐷 = ℎ𝑣𝑣𝑚𝑚𝑚𝑚𝑚𝑚 /𝑘𝑘𝐵𝐵
(named later Debye temperature), which is typical to each solid. The Debye
heat capacity approaches the classical Dulong-Petit result at high
temperatures (𝑇𝑇 ≫ 𝜃𝜃𝐷𝐷 ). In terms of 𝑥𝑥 = 𝜃𝜃𝐷𝐷 /𝑇𝑇 the Debye heat capacity is a
universal function, though more complicated, including an integral.

Figure (1.3): Heat capacity of classic, Debye and Einstein.

Dulong-Petit: Classic approach treating atoms in a lattice as independent


particles. Einstein: Quantum model treating the lattice as a set of independent
Quantum harmonic oscillators. Debye: Quantum model treating the lattice as
a set of coupled Quantum harmonic oscillators.

1.7 Conduction Heat Transfer:


When a temperature gradient exists in a body, experience has shown that
there is an energy transfer from the high-temperature region to the low-
temperature region. The physical mechanism obeys the Fourier’s law of heat
conduction given by the following equation; Fourier made the empirical
observation that the heat flow is linearly proportional to the thermal gradient

10
𝑑𝑑𝑑𝑑
q= −𝑘𝑘 (1.11)
𝑑𝑑𝑑𝑑

where q is the heat flow, dT/dz is the thermal gradient with distance z and
the negative sign arises because heat flows down the thermal gradient. The
proportionality constant k is called thermal conductivity. Analysing the units
𝐽𝐽
of this equation shows that k has the units � �. It is important to note that k
𝑆𝑆𝑆𝑆𝑆𝑆

has the units of watt per meter per Celsius degree in a typical system of units
in which the heat flow is expressed in watt.
Considering the general case where the temperature may be changing with
time and heat sources may be present within the body, energy balance allows
to establish the basic equation which governs the transfer of heat in a solid,
using equation (1.11) as a starting point. This equation is called heat
conduction equation and is available at each point inside the solid:
1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑘𝑘
− +∆𝑇𝑇 + = 0 𝑤𝑤ℎ𝑒𝑒𝑒𝑒𝑒𝑒 𝑎𝑎 = (1.12)
𝑎𝑎 𝜕𝜕𝜕𝜕 𝐾𝐾 𝜌𝜌𝜌𝜌

The quantity a is called the thermal diffusivity of the material. The larger
the value of a, the faster heat will diffuse through the material, this may be
seen by examining the quantities which make up a. A high value of a could
result either from a high value of the thermal conductivity, which would
indicate a rapid energy-transfer rate, or from a low value of the thermal heat
capacity 𝑃𝑃𝑐𝑐 . A low value of the heat capacity would mean that less of the
energy moving through the material would be absorbed and used to raise the
temperature of the material.

1.8 Convection Heat Transfer:


Heat transfer occurring in a fluid system systematically involves the
process of convection heat transfer. The analytical treatment of that kind of
problem is very complex because of the heat and mass transfer where the
velocity field is coupled to the temperature field. Actually, it is a problem of

11
heat transported by fluid mass characterised by its heat capacity 𝑝𝑝𝑐𝑐 and its
velocity field which are both strongly temperature-dependent.
Fourier's second law is based on a simple energy balance. It simply states
that the change of the energy content H with time t corresponds to the change
of heat flow with distance z:
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
=− (1.13)
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑

Analysing the units of the equation its clear that heat content H is
𝐽𝐽
measured in: � �. The origin of the negative sign in equation (1.13) can be
𝑚𝑚 3

seen on the figure (1.4) If more heat flows into of a unit volume than flows
out of it, then the change of heat flow across the distance dz is negative and
the heat content inside it will rise.

Figure (1.4): Energy balance in a unity volume.

12
Chapter Two
Heat Flow Analyzing

2.1 Introduction:
Heat transfer engineers investigate the rate of transport of thermal energy
in engineering systems. Heat conduction is the primary thermal energy
transport mechanism in solid systems. Energy is transported through solid
materials when temperature gradients exist inside them. Energy moves from
atom to atom by phononic as well as electronic interactions. This process is
known as heat conduction. Let us consider a few selected examples of
practical situations involving conduction heat transfer processes.
A common processing method using in the metallurgical industry is
continuous casting. Steel, copper and aluminum are routinely produced using
this technology. The process is used to convert refined liquid metal into
solidified ingot. Liquid metal is poured into a chilled reciprocating mold on
one end and the (partially solidified) ingot is extracted on the other end. For
the metal to freeze heat must be conducted through the solidified shell in
contact with the mold and though the mold wall. Continuous casting
engineers are interested in controlling the rate of solidification in order to
avoid metallurgical defects or catastrophic breakouts. In the manufacture of
jet engines, turbine disks are mill annealed by heating and maintaining them
at a selected temperature and subsequent air cooling. The treatment leads to
optimal metallurgical structure and properties in the finished component. Heat
treating engineers are interested in controlling the rates of heating and cooling
during heat treatment in order to optimize the resulting component properties.
Another reason to undertake the study of conduction heat transfer is that
the fundamental notion involved (energy conservation) has important
analogues in other physical systems of considerable interest. Specifically, the
principle of mass conservation is used to formulate and solve problems in
diffusional mass transfer and the principles of conservation mass and

13
momentum are the foundation of fluid mechanics. Much of the intuition and
insight acquired from a study of conduction heat transfer can be utilized to
advantage when studying diffusion and fluid flow.
Furthermore, the mathematics of heat conduction has also important
applications in the study of Brownian motion, probability theory and in
financial investment theory. The mathematical formulation of heat conduction
problems is based on the principle of conservation of energy which is the
statement of the thermal energy balance inside a body containing temperature
gradients.

2.2 Derivation of the Heat Equation:


Consider a long thin bar of length one, of uniform material, and insulated
so that heat can enter or escape only at its ends. Let 𝑢𝑢(𝑥𝑥, 𝑡𝑡) be the temperature
at position 0 ≤ 𝑥𝑥 ≤ 1, and time 𝑡𝑡 ≥ 0. Consider a short stretch of bar between
𝑥𝑥 and 𝑥𝑥 + ℎ. It is a fact one can measure in the laboratory, that the rate 𝐻𝐻(𝑥𝑥)
at which heat flows from 𝑥𝑥 to 𝑥𝑥 + ℎ is proportional to the temperature
gradient (𝑢𝑢(𝑥𝑥, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥 + ℎ, 𝑡𝑡))/ℎ. (𝐻𝐻(𝑥𝑥) > 0, so heat flows from 𝑥𝑥 to 𝑥𝑥 +
ℎ, if 𝑢𝑢(𝑥𝑥, 𝑡𝑡) > 𝑢𝑢(𝑥𝑥 + ℎ, 𝑡𝑡)). 𝐻𝐻(𝑥𝑥 − ℎ) − 𝐻𝐻 (𝑥𝑥) is the rate at which heat flows
toward 𝑥𝑥, from both 𝑥𝑥 − ℎ and 𝑥𝑥 + ℎ, and so is the rate at which heat "builds
up" at 𝑥𝑥. Dividing (𝐻𝐻(𝑥𝑥 − ℎ) − 𝐻𝐻 (𝑥𝑥)) by ℎ yields the rate at which the "heat
density" builds up at 𝑥𝑥, which is in turn proportional to the rate at which the
temperature rises or falls at 𝑥𝑥. In other words:

𝑑𝑑𝑑𝑑(𝑥𝑥, 𝑡𝑡) (𝑢𝑢(𝑥𝑥 − ℎ, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥, 𝑡𝑡))/ℎ − (𝑢𝑢(𝑥𝑥, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥 + ℎ, 𝑡𝑡))/ℎ
= 𝐶𝐶 (2.1)
𝑑𝑑𝑑𝑑 ℎ

where 𝐶𝐶 is a constant of proportionality. For small ℎ,


𝑢𝑢(𝑥𝑥 − ℎ, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥, 𝑡𝑡) 𝑑𝑑𝑑𝑑(𝑥𝑥 − ℎ/2, 𝑡𝑡)
≈− (2.2)
ℎ 𝑑𝑑𝑑𝑑
And
𝑢𝑢(𝑥𝑥, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥 + ℎ, 𝑡𝑡) 𝑑𝑑𝑑𝑑(𝑥𝑥 + ℎ/2, 𝑡𝑡)
≈− (2.3)
ℎ 𝑑𝑑𝑑𝑑

14
and so:
(𝑢𝑢(𝑥𝑥 − ℎ, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥, 𝑡𝑡))/ℎ − (𝑢𝑢(𝑥𝑥, 𝑡𝑡) − 𝑢𝑢(𝑥𝑥 + ℎ, 𝑡𝑡))/ℎ 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑡𝑡)
≈ (2.4)
ℎ 𝑑𝑑𝑑𝑑 2
Thus in the limit as ℎ ⟶ 0, we get the heat equation:
𝑑𝑑𝑑𝑑(𝑥𝑥, 𝑡𝑡) 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑡𝑡)
= 𝐶𝐶 (2.5)
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 2

𝐶𝐶 is called the conductivity of the bar. To solve this equation, we still need
the boundary conditions, or the temperatures at the ends of the bar
(𝑢𝑢(0, 𝑡𝑡) 𝑎𝑎𝑎𝑎𝑎𝑎 𝑢𝑢(1, 𝑡𝑡) 𝑓𝑓𝑓𝑓𝑓𝑓 𝑎𝑎𝑎𝑎𝑎𝑎 𝑡𝑡 ≥ 0, and the initial conditions, or the
temperature of the bar at time 𝑡𝑡 = 0, (𝑢𝑢(𝑥𝑥, 0)𝑓𝑓𝑓𝑓𝑓𝑓 𝑎𝑎𝑎𝑎𝑎𝑎 0 ≤ 𝑥𝑥 ≤ 1).

If instead of a one-dimensional bar, we have a two-dimensional plate of


some shape, the heat equation becomes:
𝑑𝑑𝑑𝑑(𝑥𝑥, 𝑦𝑦, 𝑡𝑡) 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑦𝑦, 𝑡𝑡) 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑦𝑦, 𝑡𝑡)
= 𝐶𝐶 � + � (2.6)
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 2 𝑑𝑑𝑑𝑑 2
with boundary conditions consisting of temperature specified all along the
one-dimensional boundary of the plate, and the initial conditions consisting of
the temperature at every point in the plate at 𝑡𝑡 = 0.

Finally, if we have a complete three-dimensional object, the heat equation


becomes:
𝑑𝑑𝑑𝑑(𝑥𝑥, 𝑦𝑦, 𝑧𝑧, 𝑡𝑡) 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑦𝑦, 𝑧𝑧, 𝑡𝑡) 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑦𝑦, 𝑧𝑧, 𝑡𝑡) 𝑑𝑑2 𝑢𝑢(𝑥𝑥, 𝑦𝑦, 𝑧𝑧, 𝑡𝑡)
= 𝐶𝐶 � + + � (2.7)
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 2 𝑑𝑑𝑑𝑑 2 𝑑𝑑𝑑𝑑 2
With boundary conditions consisting of temperature all along the two-
dimensional boundary of the body, and initial conditions consisting of the
temperature throughout the body at 𝑡𝑡 = 0. (Note that the Laplacian is defined
both for functions of two and of three variables), (see Figure 2.1).

15
Figure(2.1): Heat flow boundary condition in one-, two-, three-dimensions.

2.3 Derivation of the heat-conduction equation:


Consider a small element of material in a solid body. The element has the
shape of a rectangular parallelepiped with its edges dx , dy and dz parallel,
respectively, to the x , y and z axes as shown in Figure (2.2). To obtain an
equation for the temperature distribution we write an energy balance for the
element semantically in the form.

Rate of Rate of heat generation Rate of heat Rate of change in


+ = +
Heat in-flow by Internal source out-flow internal energy

Or algebraically as:
�𝑞𝑞𝑥𝑥 + 𝑞𝑞𝑦𝑦 + 𝑞𝑞𝑧𝑧 � + 𝑞𝑞̇ (𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑)
𝜕𝜕𝜕𝜕
= �𝑞𝑞𝑥𝑥+𝑑𝑑𝑑𝑑 + 𝑞𝑞𝑦𝑦+𝑑𝑑𝑑𝑑 + 𝑞𝑞𝑧𝑧+𝑑𝑑𝑑𝑑 � + 𝑐𝑐𝑐𝑐(𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑) (2.8)
𝜕𝜕𝜕𝜕

16
Where the rate of heat generation per unit volume, (𝑞𝑞̇ ) and the
temperature, T, are in general function of the three coordinate x , y , z as well
as of time 𝜃𝜃.

Figure (2.2): Sketch illustrating nomenclature for the derivation of the


general heat-conduction equation in Cartesian coordinates.

The rate of heat conduction in to the element across the left face in the x
direction 𝑞𝑞𝑥𝑥 can acoording to Eq. (2.8) be written as:
𝜕𝜕𝜕𝜕
𝑞𝑞𝑥𝑥 = �−𝑘𝑘 � 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 (2.9)
𝜕𝜕𝜕𝜕
The temperature gradient is expressed as a partial derivate because T is
not only a function of x , but also of y , z and 𝜃𝜃 . The rate of heat-conduction
out of the element across the right face at x+dx, 𝑞𝑞𝑥𝑥+𝑑𝑑𝑑𝑑 is
𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
𝑞𝑞𝑥𝑥+𝑑𝑑𝑑𝑑 = ��−𝑘𝑘 �+ �−𝑘𝑘 � 𝑑𝑑𝑑𝑑� 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 (2.10)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
Subtracting the heat-flow rate out of the element from the heat-flow rate
in to the element yields.
𝜕𝜕𝜕𝜕
𝜕𝜕 �𝑘𝑘 �
𝑞𝑞𝑥𝑥 − 𝑞𝑞𝑥𝑥+𝑑𝑑𝑑𝑑 = 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 (2.11)
𝜕𝜕𝜕𝜕
And similarly for the y and z directions

17
𝜕𝜕𝜕𝜕
𝜕𝜕 �𝑘𝑘 �
𝜕𝜕𝜕𝜕
𝑞𝑞𝑦𝑦 − 𝑞𝑞𝑦𝑦+𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 (2.12)
𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕
𝜕𝜕 �𝑘𝑘 �
𝑞𝑞𝑧𝑧 − 𝑞𝑞𝑧𝑧+𝑑𝑑𝑑𝑑 = 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 (2.13)
𝜕𝜕𝜕𝜕
Substituting these relation in to the energy balance and dividing each
term by 𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 gives:
𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
�𝑘𝑘 � + �𝑘𝑘 � + �𝑘𝑘 � + 𝑞𝑞̇ = 𝑐𝑐𝑐𝑐 (2.14)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
If the system is homogeneous and the specific heat c and density, 𝛒𝛒 are
independent of temperature. If also k is assumed to be uniform, eq. (2.14) can
be written:
𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝑞𝑞̇ 1 𝜕𝜕𝜕𝜕
+ + + = (2.15)
𝜕𝜕𝑥𝑥 2 𝜕𝜕𝑦𝑦 2 𝜕𝜕𝑧𝑧 2 𝑘𝑘 𝑎𝑎 𝜕𝜕𝜕𝜕
Where the constant 𝑎𝑎 = 𝑘𝑘/𝑐𝑐𝑐𝑐 is called the thermal diffusivity and has the
unit, sq m∕sec, in the SI system. Equation (2.15) is known as the general heat-
conduction equation and governs the temperature distribution and the
conduction heat flow in a solid having uniform physical properties.
If the system contains no heat source, Eq. (2.15) reduces to the Fourier
equation:
𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 1 𝜕𝜕𝜕𝜕
+ + = (2.16)
𝜕𝜕𝑥𝑥 2 𝜕𝜕𝑦𝑦 2 𝜕𝜕𝑧𝑧 2 𝑎𝑎 𝜕𝜕𝜕𝜕
If the system is steady, but heat sources are present, Eq. (2.12) becomes
the Poisson equation:
𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝑞𝑞̇
+ + + =0 (2.17)
𝜕𝜕𝑥𝑥 2 𝜕𝜕𝑦𝑦 2 𝜕𝜕𝑧𝑧 2 𝑘𝑘
In the steady state the temperature distribution in a body free of heat
source must satisfy the Laplace equation
𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇
+ + =0 (2.18)
𝜕𝜕𝑥𝑥 2 𝜕𝜕𝑦𝑦 2 𝜕𝜕𝑧𝑧 2
For one-dimensional steady heat conduction, Eq. (2.18) becomes:

18
𝑑𝑑2 𝑇𝑇
=0 (2.19)
𝑑𝑑𝑑𝑑 2
𝑑𝑑𝑑𝑑
Which yield, after integration, = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐, for steady one-
𝑑𝑑𝑑𝑑

dimensional heat conduction.

Figure (2.3): Cylindrical coordinate system.

There are numerous problems in heat conduction which can be handled


more conveniently in a cylindrical or spherical coordinate system. The
general heat-conduction equation in the cylindrical coordinate system shown
in Figur(2.3) is:
𝜕𝜕 2 𝑇𝑇 1 𝜕𝜕𝜕𝜕 1 𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇 𝑞𝑞̇ 1 𝜕𝜕𝜕𝜕
+ + + + = (2.20)
𝜕𝜕𝑟𝑟 2 𝑟𝑟 𝜕𝜕𝜕𝜕 𝑟𝑟 2 𝜕𝜕𝜙𝜙 2 𝜕𝜕𝑧𝑧 2 𝑘𝑘 𝑎𝑎 𝜕𝜕𝜕𝜕
The derivation of this equation is left as an exercise. However, attention is
directed to the fact that the area through which heat flows in to the element in
the positive r direction is r d𝜙𝜙 dz and the area through which heat flows out of
the element in the r direction is (r+dr)d𝜙𝜙 dz. This change in area results in the
1 𝜕𝜕𝜕𝜕
term � � ( ) in Eq. (2.20).
𝑟𝑟 𝜕𝜕𝜕𝜕

In the spherical coordinate system shown in Figure(2.4), the general heat-


conduction equation becomes

19
1 𝜕𝜕 2 𝜕𝜕𝜕𝜕 1 𝜕𝜕 𝜕𝜕𝜕𝜕 1 𝜕𝜕 2 𝑇𝑇 𝑞𝑞̇ 1 𝜕𝜕𝜕𝜕
�𝑟𝑟 �+ 2 �sin 𝜙𝜙 � + 2 + = (2.21)
𝑟𝑟 2 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑟𝑟 sin 𝜙𝜙 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑟𝑟 sin 𝜙𝜙 2 𝜕𝜕𝜓𝜓 2 𝑘𝑘 𝑎𝑎 𝜕𝜕𝜕𝜕

Figure(2.4): Spherical coordinate system.

2.4 Method of Analysis:


In the preceding chapter we dealt with problems in which the able the
temperature and the heat flow can be treated as functions of a single variable.
Many practical problems fall into this category, but when the Boundaries of a
system are irregular or when the temperature along a Boundary is non
uniform a one-dimensional treatment may no longer be Satisfactory. In such
cases the temperature is a function of two and possibly even three
coordinates. The heat flow through a corner section where two or three walls
meet the heat conduction through the walls of a short hollow cylinder or the
heat loss from a buried pipe are typical Examples of this class of problems.
In this chapter we shall consider some methods for analyzing conduction
in two- and three-dimensional systems. The emphasis will be placed on two-
dimensional problems because they are less cumbersome to solve yet they
illustrate the basic methods of analysis for three- dimensional systems.
Heat conduction in two- and three-dimensional systems can be treated by
analytical graphical analogical and numerical methods. A complete treatment

20
of analytical solutions requires a prior knowledge of Fourier series Bessel
functions legendre polynomials Laplace transform methods and complex
variable theory.
We shall consider only the analytical solution of one relatively simple
problem to illustrate the analytical method of approach. Emphasis will be
placed on numerical method which are suitable for solution by computer.

2.5 Analytical solution:


The objective of any heat transfer analysis is to predict either the rate of
heat flow or the temperature distribution. In a two-dimensional system
without heat sources, the equation governing the temperature distribution in
the steady state is:
𝜕𝜕 2 T 𝜕𝜕 2 T
+ =0 (2.22)
∂x 2 ∂y 2
If the thermal conductivity is uniform. The solution of Eq. (2.22) will give
T(x, y), the temperature as a function of the two space-coordinates x and y.
The rate of heat flow per unit area in the x and y directions, respectively, can
then be obtained from Fourier's law:
𝜕𝜕𝜕𝜕
(𝑞𝑞 ⁄𝐴𝐴)x = −𝑘𝑘 (2.23)
𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕
(𝑞𝑞 ⁄𝐴𝐴)y = −𝑘𝑘 (2.24)
𝜕𝜕𝜕𝜕
It should be noted that whereas the temperature is a scalar, the heat flux
depends on the temperature gradient and is therefore a vector. The total rate of
heat flow at a given point x, y, is the resultant of 𝑞𝑞𝑥𝑥 and 𝑞𝑞𝑦𝑦 at that point at is
directed perpendicular to the isotherm as shown in Figure(2.5). Thus, if the
temperature distribution in a system is known, the rate of heat flow can easily
be calculated. Usually, heat-transfer analyses concentrate therefore on
determining the temperature field.
An analytical solution of a heat-conduction problem must satisfy the heat-
conduction equation as well as the boundary conditions specified by the

21
physical conditions of the particular problem. The classical approach to an
exact solution of the Fourier equation is the separation-of-variables technique.
We shall illustrate this approach by applying it:

𝒒𝒒 = 𝒒𝒒𝒙𝒙 + 𝒒𝒒𝐲𝐲
𝒒𝒒𝒚𝒚

𝒒𝒒𝒙𝒙

T(x,y)

Figure(2.5): sketch showing heat flow in two dimension Isotherm.

To a relatively simple problem, consider a thin rectangular plate, free of


heat sources and insulated at the top and bottom surface (Figure 2.6). Since
𝜕𝜕𝜕𝜕⁄𝜕𝜕𝜕𝜕 is negligible the temperature is a function of x and y only.
If the thermal conductivity is uniform, the temperature distribution must
satisfy the equation
𝜕𝜕 2 𝑇𝑇 𝜕𝜕 2 𝑇𝑇
+ =0 (2.25)
∂x 2 ∂y 2
Equation (3.16) is a linear and homogeneous partial-differential equation
which can be integrated by assuming a product solution for T (x, y) of the
form
𝑇𝑇 = 𝑋𝑋 𝑌𝑌 (2.26)
Where X= X(x), a function of x only, and Y= Y(y), a function of y alone.
Substituting Eq. (2.26) in Eq. (2.25) yields:
1 𝑑𝑑2 𝑋𝑋 1 𝑑𝑑2 𝑌𝑌
− = (2.27)
𝑋𝑋 𝑑𝑑𝑑𝑑 2 𝑌𝑌 𝑑𝑑𝑑𝑑 2
The variables are now separated. The left-hand side a function of x only,
while the right-hand side is a function of y alone. Since neither side can

22
change as x and y vary, both must be equal to a constant, say 𝜆𝜆2 we have,
therefore, the two total-differential equations:
𝑑𝑑2 𝑋𝑋
+ 𝜆𝜆2 𝑋𝑋 = 0 (2.28)
𝑑𝑑𝑑𝑑 2
𝑑𝑑2 𝑌𝑌
2
− 𝜆𝜆2 𝑌𝑌 = 0 (2.29)
𝑑𝑑𝑦𝑦

y
𝝅𝝅𝝅𝝅
𝑻𝑻 = 𝑻𝑻𝒎𝒎 𝒔𝒔𝒔𝒔𝒔𝒔 � 𝑳𝑳 �
b

T=0 T=0

0 T=0 L x

Figure(2.6):Rectangular adiabatic plate.

The general solution to Eq. (2.28) is:


𝑋𝑋 = 𝐴𝐴 cosλ 𝑥𝑥 + 𝐵𝐵 sinλ 𝑥𝑥 (2.30)
The general solution to Eq. (2.29) is:
𝑌𝑌 = 𝐶𝐶𝑒𝑒 −𝜆𝜆𝜆𝜆 + 𝐷𝐷𝑒𝑒 𝜆𝜆𝜆𝜆 (2.31)
and, therefore, from Eq.(2.26):
𝑇𝑇 = 𝑋𝑋𝑋𝑋 = (𝐴𝐴 cosλ 𝑥𝑥 + 𝐵𝐵 sinλ 𝑥𝑥) �𝐶𝐶𝑒𝑒 −𝜆𝜆𝜆𝜆 + 𝐷𝐷𝑒𝑒 𝜆𝜆𝜆𝜆 � (2.32)
Where A , B , C, and D are constants to be evaluated from the boundary
conditions. As shown in figure (2.5), the boundary conditions to be satisfied
are:
1. T= 0 at y= 0
2. T= 0 at x= 0
3. T= 0 at x= 𝐿𝐿
𝜋𝜋𝜋𝜋
4. T=𝑇𝑇𝑚𝑚 𝑠𝑠𝑠𝑠𝑠𝑠 � � at y=b
𝐿𝐿

23
Substituting these conditions into Eq. (2.32) for T we get from the first
condition:
(𝐴𝐴 cos 𝜆𝜆𝜆𝜆 + 𝐵𝐵 sin 𝜆𝜆𝜆𝜆) (𝐶𝐶 + 𝐷𝐷) = 0 (2.33)
From the second condition:
𝐴𝐴�𝐶𝐶𝑒𝑒 −𝜆𝜆𝜆𝜆 + 𝐷𝐷𝑒𝑒 𝜆𝜆𝜆𝜆 � = 0 (2.34)
And from the third condition:
(𝐴𝐴 cos 𝜆𝜆𝜆𝜆 + 𝐵𝐵 sin𝜆𝜆 𝐿𝐿) (𝐶𝐶𝑒𝑒 −𝜆𝜆𝜆𝜆 + 𝐷𝐷𝑒𝑒 𝜆𝜆𝜆𝜆 ) = 0 (2.35)
The first condition can be satisfied only if 𝐶𝐶 = −𝐷𝐷, and the second if
𝐴𝐴 = 0. Using these results in the third condition, we obtain:
(𝐵𝐵 sin 𝐿𝐿)𝐶𝐶�𝑒𝑒 −𝜆𝜆𝜆𝜆 − 𝑒𝑒 𝜆𝜆𝜆𝜆 � = 2𝐵𝐵𝐵𝐵 sin 𝜆𝜆𝜆𝜆 sinh 𝜆𝜆𝜆𝜆 = 0 (2.36)
To satisfy this condition, sin 𝐿𝐿 must be zero or λ= 𝑛𝑛𝑛𝑛⁄𝐿𝐿, where n= 1, 2,3,
etc. There exists therefore a different solution for each integer n and each
solution has a separate integration constant C n . summing these solution, we
get:

𝑛𝑛𝑛𝑛𝑛𝑛 𝑛𝑛𝑛𝑛𝑛𝑛
𝑇𝑇 = � Cn sin sinh (2.37)
𝐿𝐿 𝐿𝐿
1

The last boundary condition demands that, at 𝑦𝑦 = 𝑏𝑏,



𝑛𝑛𝑛𝑛𝑛𝑛 𝑛𝑛𝑛𝑛𝑛𝑛 𝜋𝜋𝜋𝜋
𝑇𝑇 = � Cn sin sinh = 𝑇𝑇𝑚𝑚 sin � � (2.38)
𝐿𝐿 𝐿𝐿 𝐿𝐿
1

So that only the first term in the series solution with


𝐶𝐶1 = 𝑇𝑇𝑚𝑚 ⁄sinh(𝜋𝜋𝜋𝜋⁄𝐿𝐿) is needed. The solution therefore becomes:

sinh(𝜋𝜋𝜋𝜋⁄𝐿𝐿) 𝜋𝜋𝜋𝜋
𝑇𝑇(𝑥𝑥, 𝑦𝑦) = 𝑇𝑇𝑚𝑚 sin � � (2.39)
sinh(𝜋𝜋𝜋𝜋⁄𝐿𝐿) 𝐿𝐿
The corresponding temperature filed is shown in Figure (2.7) the solid
lines are isotherms and the dotted lines are heat-flow lines. It should be noted
that lines indicating the direction of heat flow are perpendicular to the
isotherms.

24
When the boundary conditions are not as simple as in the illustrative
problem, the solution is obtained in the form of an infinite series. For
example, if the temperature at the edge 𝑦𝑦 = 𝑏𝑏 is a function of x, say

Isotherm
Heat-Flow lines

Figure(2.7): Isotherm Heat-Flow lines in the rectangular adiabatic


plate shown in Figure (2.6).

𝑇𝑇(𝑥𝑥, 𝑏𝑏) = 𝐹𝐹 (𝑥𝑥), then the solution, is the infinite series:


∞ 𝐿𝐿
2 sinh(𝑛𝑛𝑛𝑛⁄𝐿𝐿) 𝜋𝜋𝜋𝜋 𝑛𝑛𝑛𝑛
𝑇𝑇 = � sin 𝑥𝑥 � 𝐹𝐹 (𝑥𝑥) sin 𝑥𝑥𝑥𝑥𝑥𝑥 (2.40)
𝐿𝐿 sinhnπ(𝑏𝑏⁄𝐿𝐿) 𝐿𝐿 𝐿𝐿
𝑛𝑛=1 0

which is quite laborious to evaluate quantitatively the separation-of-


variables method can be extended to three-dimensional cases by assuming
𝑇𝑇 = 𝑋𝑋𝑋𝑋𝑋𝑋 substituting this expression for T in Eq. (2.18), separating the
variables, and integrating the resulting total-differential equations subject to
the given boundary conditions.

Analytical solution are useful, but there are few practical problems with
geometries and boundary conditions which can be solved analytically; and
even when a solution has been obtained, it is often too complicated to justify
the time and effort required to evaluate it quantitatively.

25
2.6 Numerical method of analysis:
An immense number of analytical solutions for conduction heat-transfer
problems have been accumulated in literature over the past 100 years. Even
so, in many practical situations the geometry or boundary conditions are such
that an analytical solution has not been obtained at all, or if the solution has
been developed, it involves such a complex series solution that numerical
evaluation becomes exceedingly difficult. For such situation the most fruitful
approach to the problem is one based on finite-difference techniques, the
basic principles of which we shall outline in this section.
Consider a two-dimensional body that is to be divided into equal
increment in both the 𝑥𝑥 and 𝑦𝑦 directions, as shown in figure (2.8). The nodal
points are designated as shown, the m locations indicating the 𝑥𝑥 increment
and the 𝑛𝑛 locations indicating the y increment. We wish to establish the
temperatures at any of these nodal points within the body, using Equation
(2.22) as a governing condition. Finite difference are used to approximate
differential increment in the temperature and space coordinates; and the
smaller we choose these finite increments, the more closely the true
temperature distribution will be approximated.

m, n+1

△ 𝒚𝒚
m-1, n m, n m+1, n

△ 𝒚𝒚

m, n-1

Figure (2.8): sketch illustrating nomenclature used in two-dimensional


numerical analysis of heat conduction.

26
The temperature gradients may be written as follows:
𝜕𝜕𝜕𝜕 𝑇𝑇𝑚𝑚 +1,𝑛𝑛 − 𝑇𝑇𝑚𝑚 ,𝑛𝑛
� ≈ (2.41)
𝜕𝜕𝜕𝜕 𝑚𝑚 +1/2,𝑛𝑛 ∆𝑥𝑥
𝜕𝜕𝜕𝜕 𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇𝑚𝑚 −1,𝑛𝑛
� ≈ (2.42)
𝜕𝜕𝜕𝜕 𝑚𝑚 −1/2,𝑛𝑛 ∆𝑥𝑥
𝜕𝜕𝜕𝜕 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 − 𝑇𝑇𝑚𝑚 ,𝑛𝑛
� ≈ (2.43)
𝜕𝜕𝜕𝜕 𝑚𝑚 ,𝑛𝑛+1/2 ∆𝑦𝑦
𝜕𝜕𝜕𝜕 𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1
� ≈ (2.44)
𝜕𝜕𝜕𝜕 𝑚𝑚 ,𝑛𝑛−1/2 ∆𝑦𝑦
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
� − �
2
𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑚𝑚 +1/2,𝑛𝑛 𝜕𝜕𝜕𝜕 𝑚𝑚 −1/2,𝑛𝑛
� ≈
𝜕𝜕𝑥𝑥 2 𝑚𝑚 ,𝑛𝑛 ∆𝑥𝑥
𝑇𝑇𝑚𝑚 +1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 − 2𝑇𝑇𝑚𝑚 ,𝑛𝑛
= (2.45)
(∆𝑥𝑥)2
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
� − �
2
𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑚𝑚 ,𝑛𝑛+1/2 𝜕𝜕𝜕𝜕 𝑚𝑚 ,𝑛𝑛−1/2
� ≈
𝜕𝜕𝑦𝑦 2 𝑚𝑚 ,𝑛𝑛 ∆𝑦𝑦
𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 − 2𝑇𝑇𝑚𝑚 ,𝑛𝑛
= (2.46)
(∆𝑦𝑦)2

Thus the finite-difference approximation for equation (2.22) becomes:

𝑇𝑇𝑚𝑚 +1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 − 2𝑇𝑇𝑚𝑚 ,𝑛𝑛 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 − 2𝑇𝑇𝑚𝑚 ,𝑛𝑛
+ =0 (2.47)
(∆𝑥𝑥)2 (∆𝑦𝑦)2
If ∆𝑥𝑥 = ∆𝑦𝑦, then
𝑇𝑇𝑚𝑚 +1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 − 4𝑇𝑇𝑚𝑚 ,𝑛𝑛 = 0 (2.48)

Since we are considering the case of constant thermal conductivity, the


heat flows may all be expressed in terms of temperature differentials,
Equation (2.48) states very simply that the net heat flow into any node is zero
at steady-state conditions. In effect, the numerical finite-difference approach
replaces the continuous temperature distribution by fictitious heat-conducting
rods connected between small nodal points which do not generate heat.

27
We can also devise a finite-difference scheme to take heat generation into
𝑞𝑞̇
account. We merely add the term �𝑘𝑘 into the general equation and obtain

𝑇𝑇𝑚𝑚 +1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 − 2𝑇𝑇𝑚𝑚 ,𝑛𝑛 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 − 2𝑇𝑇𝑚𝑚 ,𝑛𝑛 𝑞𝑞̇
+ + = 0 (2.49)
(∆𝑥𝑥)2 (∆𝑦𝑦)2 𝑘𝑘

Then for a square grid in which ∆𝑥𝑥 = ∆𝑦𝑦,

𝑞𝑞̇ (∆𝑥𝑥)2
𝑇𝑇𝑚𝑚 +1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 + − 4𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 = 0 (2.50)
𝑘𝑘

To utilize the numerical method, Equation (2.48) must be written for each
node within the material and the resultant system of equations solved for the
temperature at the various nodes.

When the solid is exposed to some convection boundary condition, the


temperatures at the surface must be computed differently from the method
given above. Consider the boundary shown in figure (2.9). The energy
balance on node (𝑚𝑚, 𝑛𝑛) is:

𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 ∆𝑥𝑥 𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 ∆𝑥𝑥 𝑇𝑇𝑚𝑚 ,𝑛𝑛 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1
−𝑘𝑘∆𝑦𝑦 − 𝑘𝑘 − 𝑘𝑘
∆𝑥𝑥 2 ∆𝑦𝑦 2 ∆𝑦𝑦
= ℎ∆𝑦𝑦�𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇∞ � (2.51)

If ∆𝑥𝑥 = ∆𝑦𝑦 , the boundary temperature is expressed in the equation

ℎ∆𝑥𝑥 ℎ∆𝑥𝑥 1
𝑇𝑇𝑚𝑚 ,𝑛𝑛 � + 2� − 𝑇𝑇∞ − �2𝑇𝑇𝑚𝑚 −1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛+1 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 � = 0 (2.52)
𝑘𝑘 𝑘𝑘 2

28
m, n+1

𝑇𝑇∞

m-1, n m, n △ 𝒚𝒚

△ 𝒚𝒚
q

m, n-1
∆𝒙𝒙
𝟐𝟐
surface
∆𝑥𝑥

Figure (2.9): Nomenclature for nodal equation with convective boundary


condition.

An equation of this type must be written for each node along the surface
shown in figure (2.9). So when a convection boundary condition is present, an
equation like (2.52) is based at the boundary and an equation (2.48) is used
for the interior points.
Equation (2.52) applies to a plane surface exposed to a convection
boundary condition. It will not apply for other situations, such as an insulate
wall or a corner exposed to a convection boundary condition. Consider the
corner section shown in figure (2.9). The energy balance for the corner
section is
∆𝑦𝑦 𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇𝑚𝑚 −1,𝑛𝑛 ∆𝑥𝑥 𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1
−𝑘𝑘 − 𝑘𝑘
2 ∆𝑥𝑥 2 ∆𝑦𝑦
∆𝑥𝑥 ∆𝑦𝑦
=ℎ �𝑇𝑇𝑚𝑚 ,𝑛𝑛 − 𝑇𝑇∞ � + ℎ �𝑇𝑇 − 𝑇𝑇∞ � (2.53)
2 2 𝑚𝑚 ,𝑛𝑛

If ∆𝑥𝑥 = ∆𝑦𝑦 ,
ℎ∆𝑥𝑥 ℎ∆𝑥𝑥
2𝑇𝑇𝑚𝑚 ,𝑛𝑛 � + 1� − 2 𝑇𝑇 − �𝑇𝑇𝑚𝑚 −1,𝑛𝑛 + 𝑇𝑇𝑚𝑚 ,𝑛𝑛−1 � = 0 (2.54)
𝑘𝑘 𝑘𝑘 ∞

29
Other boundary condition may be treated in a similar fashion, and a
convenient summary of nodal equation is given below for different
geometrical and boundary situation. Situation 𝑓𝑓 and g are of particular interest
since they provide the calculation equations which may be employed with
curved boundaries, while still using uniform increments in ∆x and ∆y.

m-1, n m, n

∆𝑦𝑦
2

∆𝑦𝑦 𝑇𝑇∞

m, n-1
m-1, n-1
∆𝑥𝑥
2

∆𝑥𝑥

Figure(2.9): Nomenclature for nodal equation with convection at a corner


section.

(a) Interior node:

m, n+1

△ 𝒚𝒚

m-1, n m, n m+1, n

△ 𝒚𝒚

m, n-1

△ 𝒙𝒙 △ 𝒙𝒙

0 = 𝑻𝑻𝒎𝒎+𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 − 𝟒𝟒𝑻𝑻𝒎𝒎,𝒏𝒏


𝑻𝑻𝒎𝒎,𝒏𝒏 = (𝑻𝑻𝒎𝒎+𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 )/𝟒𝟒

30
(b) Convection boundary node:

m, n+1

△ 𝒚𝒚

m-1, n 𝒉𝒉, 𝑻𝑻∞


m, n

△ 𝒚𝒚

m, n-1

△ 𝒙𝒙

𝒉𝒉∆𝒙𝒙 𝟏𝟏 𝒉𝒉∆𝒙𝒙
𝟎𝟎 = 𝑻𝑻∞ − �𝟐𝟐𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 � − 𝑻𝑻𝒎𝒎,𝒏𝒏 � + 𝟐𝟐�
𝒌𝒌 𝟐𝟐 𝒌𝒌
(𝑻𝑻 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏)
𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝒎𝒎,𝒏𝒏+𝟏𝟏 � + 𝑩𝑩𝒊𝒊 𝑻𝑻∞
𝟐𝟐
𝑻𝑻𝒎𝒎,𝒏𝒏 =
𝟐𝟐 + 𝑩𝑩𝒊𝒊
𝒉𝒉∆𝒙𝒙
𝑩𝑩𝒊𝒊 =
𝒌𝒌
(c) Exterior corner with convection boundary:
𝒉𝒉, 𝑻𝑻∞
m-1, n m, n

△ 𝒚𝒚

m, n-1

△ 𝒙𝒙

𝒉𝒉∆𝒙𝒙 𝒉𝒉∆𝒙𝒙
𝟎𝟎 = 𝟐𝟐 𝑻𝑻∞ − �𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 �−𝟐𝟐𝑻𝑻𝒎𝒎,𝒏𝒏 � + 𝟏𝟏�
𝒌𝒌 𝒌𝒌
(𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏)
� + 𝑩𝑩𝒊𝒊 𝑻𝑻∞
𝟐𝟐
𝑻𝑻𝒎𝒎,𝒏𝒏 =
𝟏𝟏 + 𝑩𝑩𝒊𝒊
𝒉𝒉∆𝒙𝒙
𝑩𝑩𝒊𝒊 =
𝒌𝒌

31
(d) Interior corner with convection boundary:

m, n+1

m-1, n m, n m, n+1

△ 𝒚𝒚 𝒉𝒉, 𝑻𝑻∞

m, n-1

△ 𝒙𝒙

𝒉𝒉∆𝒙𝒙 𝒉𝒉∆𝒙𝒙
𝟎𝟎 = 𝟐𝟐 𝑻𝑻∞ + 𝟐𝟐𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎+𝟏𝟏,𝐧𝐧 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 + 𝟐𝟐(𝟑𝟑 + )𝑻𝑻𝒎𝒎,𝒏𝒏
𝒌𝒌 𝒌𝒌
(𝑻𝑻 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏)
𝑩𝑩𝒊𝒊 𝑻𝑻∞ + 𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 + 𝒎𝒎+𝟏𝟏,𝒏𝒏 �
𝟐𝟐
𝑻𝑻𝒎𝒎,𝒏𝒏 =
𝟑𝟑 + 𝑩𝑩𝒊𝒊
𝒉𝒉∆𝒙𝒙
𝑩𝑩𝒊𝒊 =
𝒌𝒌
(e) Insulated boundary:

m, n+1
Insulated

m-1, n m, n

△ 𝒚𝒚

△ 𝒙𝒙

𝟎𝟎 = 𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 + 𝟐𝟐𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 − 𝟒𝟒𝑻𝑻𝒎𝒎,𝒏𝒏


𝑻𝑻𝒎𝒎,𝒏𝒏 = (𝑻𝑻𝒎𝒎,𝒏𝒏+𝟏𝟏 + 𝑻𝑻𝒎𝒎,𝒏𝒏−𝟏𝟏 + 𝟐𝟐𝑻𝑻𝒎𝒎−𝟏𝟏,𝒏𝒏 )/𝟒𝟒

32
Chapter Three
Results and Discussion

3.1 Introduction:
Heat transfer is a process by which internal energy from one
substance transfers to another substance. Thermodynamics is the study of
heat transfer and the changes that result from it. An understanding of heat
transfer is crucial to analyzing a thermodynamic process, such as those
that take place in heat engines and heat pumps.
Numerical methods aim at finding a solution to the equation at only a
discrete set of points by replacing the differentials in the equation by
differences. These methods, however, can deal with complicated solution
domains and time varying physical properties. Examples of these methods
include the Finite Difference Method, the Finite Volume Method, and the
Finite Element Method. We will consider the Finite Difference Method
because of its simplicity and capability.
Microsoft Excel is a very versatile spreadsheet software utility. In this
chapter we describe some of the features that may be applicable to
numerical solution of heat-transfer for steady-state of two-dimensional
conducting problems. All the investigated samples in this chapter are
taking to be square, and solved for the case of square grids with ∆𝑥𝑥 = ∆𝑦𝑦.
The solution of the equations is accomplished by a built-in iterative-
solution feature of Excel. The goal here is to provide useful information
about the temperature distribution in a two dimensional coordinate, and
investigating the effect of different parameters on this temperature
distribution.

33
3.2 Comparison of Analytical- and Numerical- Solution:
The analytical solution of temperature distribution of to-dimensional
system without heat sources which described in section (2.5), was
compared with the numerical solution for the same boundary conditions.
The square sample of length 𝑙𝑙 = 𝑏𝑏 = 100 𝑐𝑐𝑐𝑐, was assumed with
maximum temperature 𝑡𝑡𝑚𝑚 = 100𝑜𝑜 𝐶𝐶. These values were set in the
equation (2.39) to obtain analytical solution, the numerical solution shows
the same results as shown in Fig (3.1).
This is indicates that the numerical method was a power full method
to perform temperature distribution of two dimensional plate for different
boundary conditions.

100
90
90
80
80

70
70

60
60

50
50

40
40

30
30

20
20

10
10

0
10 20 30 40 50 60 70 80 90 100

Figure(3.1): Analytical and numerical solution of the temperature


distribution of the problem in the section (2.5), (a) Counter plot, (b) Three
dimensional mesh.

3.3 Temperature Distribution in Plate with Fixed Boundary


Temperatures:
The two-dimensionally plate system is very easily analyzed
numerically using an Excel set up for the nodal equations. We choose
𝑇𝑇1 = 20 𝑜𝑜 𝐶𝐶 as the constant boundary temperature along the top and sides
of plate, and 𝑇𝑇2 = 100 𝑜𝑜 𝐶𝐶 as the constant tempature along the bottom of
the plate. The sketch of this problem is shown in figure (3.2).

34
T = 100℃

T=20℃ T=20℃

T=20℃

Figure (3.2): Sketch of the fixed boundary temperature problem.

The constant temperatures inserted at the boundaries cells. All the


other nodes are interior nodes and have the same form of nodal equation,
writing the equation of interior points in one of the interior cells, and then
drag-copied to the remaining temperature cells. The iterative solution
procedure is executed up to some thousands to obtain the steady state
solution.
Graphical display of the result can be presented simply as plan and
elevation views of the temperature profiles as shown in figure (3.3.a), or
as a three dimensional charts coupled with the plan and elevation views,
as shown in figure (3.3.b).

Figure(3.3): Temperature distribution of the fixed boundary problem, (a)


Counter plot, (b) Three dimensional mesh.

35
3.4 Temperature Distribution in Plate with convection environment
temperature:
The two-dimensionally plate system of convection environment
temperature 𝑇𝑇∞ = 20 𝑜𝑜 𝐶𝐶 along the top and sides of plate, with base
temperature sat at 𝑇𝑇 = 100 𝑜𝑜 𝐶𝐶. The sketch of this problem is shown in
figure (3.4). The nodal equations are described in section (2.6) are used,
with the respective values of 𝐵𝐵𝑖𝑖 = ℎ ∆𝑥𝑥/𝑘𝑘. The results of the calculations
shown in figure (3.5) for different values of the parameter 𝐵𝐵𝑖𝑖
(0.05 , 0.5 , 5.0). The 3D displays illustrate the following behavior of the
system:
1- Smaller values of 𝐵𝐵𝑖𝑖 (small convection, large conduction) result in
more uniform temperature profiles across the thickness of the
sample.
2- Smaller values of 𝐵𝐵𝑖𝑖 also result in less rapid decrease in
temperature along the length of the sample.
3- Large values of 𝐵𝐵𝑖𝑖 (high convection, low conduction) cause a rapid
drop in temperature along the length of the sample.

𝑇𝑇∞ =20℃

𝑇𝑇∞ =20℃ 𝑇𝑇∞ =20℃

T = 100℃

Figure (3.4): Sketch of the convention boundary temperature problem.

36
100

90
90

80 80

70
70

60
60
50

50

(a)
40

30 40

20
30

10

20
10 20 30 40 50 60 70 80 90 100

(b)

(c)

Figure(3.5): The two- and three-dimensional temperature distribution, for


different thermal convection (a) Bi = 0.05, (b) Bi = 0.5, (c) Bi = 5.0.

Now consider the existing of a hollow in the plate, in this case there
are eight convection boundaries at the outer and inner of the sample. The
process was proceeded for the same boundary conditions as previous, the
three dimensional mesh and plat contours view of the result are shown in
figure (3.6).

37
100
90
90

80
80

70
70

60
60

50
50

(a)
40
40

30
30

20
20

10
10

0
10 20 30 40 50 60 70 80 90 100

100
90
90

80
80

70
70

60
60

50
50

(b)
40
40

30
30

20
20

10
10

0
10 20 30 40 50 60 70 80 90 100

100
90
90

80
80

70
70

60
60

50
50

40
40
(c)
30
30

20
20

10
10

0
10 20 30 40 50 60 70 80 90 100

Figure(3.6): Temperature distribution of the convection boundary


problem, for different thermal convection (a) 𝐵𝐵𝑖𝑖 = 0.05, (b) 𝐵𝐵𝑖𝑖 = 0.5, (c)
𝐵𝐵𝑖𝑖 = 5.0.

38
3.5 Temperature Distribution in Plate with composite Materials:
Consider the composite solid plate shown in figure (3.7), in this case
there are four interface surfaces and four corners separating two materials
with different thermal conductivities, therefore the thermal conductivity
ratios 𝐾𝐾 = 𝑘𝑘𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 /𝑘𝑘𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 must be considered.
The solutions for two extremes in thermal conductivity rations are
presented
1. 𝐾𝐾 = 0.01, interior material good conductor, exterior material good
insulator. Sample materials: copper and fiberglass.
2. 𝐾𝐾 = 1000, interior material good insulator, exterior material good
conductor.
There are too many nodes to display all of the numerical solutions so
we have given three dimensional mesh charts of the temperature
distributions to illustrate the effect.
The solution for case 1 is given in figure (3.8.a) and indicates what
one would expect; strong temperature gradients through the exterior
insulating material, and very small temperature gradients in the interior
material. Note the “flat top” of the high conductivity interior material that
results from the small temperature gradients.
The solution for case 2 is displayed in figure (3.8.b) and illustrates
just the opposite of case 1; strong temperature gradients for the interior
material and relatively shallow gradients for the exterior material.
Plan and elevation views have been included for all there cases to
further illustrate the temperature gradients effects.

39
𝑇𝑇∞ =20℃

𝑇𝑇∞ =20℃ 𝑇𝑇∞ =20℃

interior material
Exterior material
T = 100℃

Figure (3.7): Sketch of the fixed boundary temperature problem.

100

90
90

80
80

70 70

60
60

50
50

40
40 (a)
30

30
20

20
10

10 20 30 40 50 60 70 80 90 100

100

90
90

80
80

70 70

60 60

50
50

40
40
(b)
30

30
20

20
10

10 20 30 40 50 60 70 80 90 100

Figure(3.8): Temperature distribution of the composite, for different ratio


of thermal conductivity (a) 𝐾𝐾 = 0.01, (b) 𝐾𝐾 = 1000.

Now consider the existing of a hollow in the inner plate as shown in


figure (3.9), in this case there are eight convection boundaries at the outer
and inner of the sample, as well as four boundary composite surfaces. The

40
process was proceeded for the same boundary conditions as previous, the
three dimensional mesh and plat contours view of the result are shown in
figure (3.10).
𝑇𝑇∞ =20℃

𝑇𝑇∞ =20℃ 𝑇𝑇∞ =20℃

interior material
Exterior material
T = 100℃

Figure (3.9): Sketch of the composite material with a hollow in the inner
plate.

100
90
90

80
80

70
70

60
60

50
50

(a)
40
40

30
30

20
20

10
10

0
10 20 30 40 50 60 70 80 90 100

100
90
90

80
80

70
70

60
60

50
50

40

(b)
40

30
30

20
20

10
10

0
10 20 30 40 50 60 70 80 90 100

Figure(3.10): Temperature distribution of the composite with a hollow in


the inner plate, the thermal conductivity (a) 𝐾𝐾 = 0.01, (b) 𝐾𝐾 = 1000.

41
3.6 Conclusions:
The objective of this work was to show the capability of numerical
techniques to investigate the heat flow in solids with different shapes and
boundary conditions; therefore the numerical method of the heat transfer
analysis is performed to calculate the temperature distribution of a two
dimensional plate for different boundary conditions.
The general equation of the temperature distribution has been first
derived for the case of unsteady state with the present of heat sources,
which is a nonlinear and inhomogeneous partial differential equation. In
general it is not a straightforward to solve this equation. In this work a
relatively simple case of a thin rectangular plate in the absence of heat
source with uniform thermal conductivity has been first addressed and
studied analytically. In this case the temperature distribution is reduced to
the Laplace equation. In general there are a few practical problems with
geometries and boundary conditions which can be solved analytically,
while there are many problems that cannot be solved analytically, so in
this case a numerical study of the problems is beneficial. For that purpose,
the problem of heat distribution for the same geometry that has been used
in the analytical case is studied numerically via the finite different
techniques.
The influence of fixed boundary temperature, convenient temperature
are investigated, while in the case of composite material the influence of
the ratio of the thermal conductivity of two material are considered, as
well as the effect of the presented of the hollow in the plate are also
illustrated.
Comparison of the temperature distribution results of the finite
difference method with the analysis method shows an excelling
agreement.

42
References
[1] J. P. Holman, “Heat Transfer”, Ninth Edition, McGraw-Hill Companies,
Inc., (2002).

[2] Steven C. Chapra, and Raymond P. Canale, “Numerical Methods for


Engineers”, Fifth Edition, McGraw-Hill Companies, Inc., (2006).

[3] B. Gebhart, Heat Conduction and Mass Diffusion, McGraw-Hill, New


York, (1993).

[4] K. Stüwe , “Principles of Heat Flow Modeling”, Intrepid Geophysics,


(2008).

[5] John H. Lienhard IV, and John H. Lienhard V, "A Heat Transfer
Textbook", Third Edition, John H. Lienhard IV and John H. Lienhard V,
(2005)

[6] Kreith, Frank, "Principles of Heat Transfer", 3rd Edition, Intext Press, Inc.,
New York, ISBN 0-7002-2422-X.

[7] http://home.comcast.net/~szemengtan/StatisticalMechanics/HeatCapacity
OfSolids.pdf

[8] http://web.njit.edu/~sirenko/Phys-446/Lecture5-SSP-2007.pdf

[9] http://www.cmmp.ucl.ac.uk/~ahh/teaching/3C25/Lecture11s.pdf

[10] http://ruelle.phys.unsw.edu.au/~gary/Site/PHYS3020_files/SM3_6.pdf

[11] http://www.enertron-inc.com/enertron-resources/PDF/Brief-Introducti on-


to-Heat-Transfer.pdf

[12] http://nptel.iitm.ac.in/courses/Webcourse-contents/IISc-BANG/Heat%20
and%20Mass%20Transfer/pdf/M1/Student_Slides_M1.pdf

[13] http://www.wlv.com/products/databook/ch1_1.pdf

[14] http://www.ewp.rpi.edu/hartford/~ernesto/C_Su2003/MMHCD/Notes/
Notes_pdf/s01.pdf

43

Você também pode gostar