0% found this document useful (0 votes)
68 views47 pages

11 Chapter 2

This chapter reviews concepts related to heat and mass transfer in nanofluid boundary layer flow. It begins by presenting the basic equations of fluid motion, heat, and mass transfer. It then discusses fundamentals of nanofluids and their applications. Finally, it reviews previous related studies. The chapter goes on to define key fluid properties like density, pressure, and viscosity. It also discusses fluid statics and dynamics, and analytical and computational solution methods in fluid mechanics.

Uploaded by

solomon
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
68 views47 pages

11 Chapter 2

This chapter reviews concepts related to heat and mass transfer in nanofluid boundary layer flow. It begins by presenting the basic equations of fluid motion, heat, and mass transfer. It then discusses fundamentals of nanofluids and their applications. Finally, it reviews previous related studies. The chapter goes on to define key fluid properties like density, pressure, and viscosity. It also discusses fluid statics and dynamics, and analytical and computational solution methods in fluid mechanics.

Uploaded by

solomon
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 47

Chapter 2

Literature Review

This chapter is devoted to the discussion and presentation of various concepts that
are important to understand the heat and mass transfer in the boundary layer flow of
nanofluids. We start this chapter with a review of the fundamental concepts of fluid
mechanics that form the framework for heat and mass transfer in the nanofluid flow.
We first present the basic equation of motion, heat and mass transfer. We then present
the fundamentals of nanofluids and its applications. Finally, the review of the previous
related studies is presented.

2.1 Fundamental Concepts and Definitions

Fluid is a substance that continually deforms (flows) under an applied shear stress.
Fluids are a subset of the phases of matter and include liquids, gases, plasmas and, to
some extent, plastic solids. Liquids form a free surface (that is, a surface not created
by the container) while gases do not. The gross properties of solids, liquids and gases
are directly related to their molecular structure and to the nature of the forces between
the molecules. The distinction between solids and fluid is not a sharp one, since there
are many materials which in some respects behave like a solid and in the other respects
like a fluid. The distinction between liquids and gases is much less fundamental, so far
as dynamical studies are concerned. For reasons related to the nature of intermolecular
forces, most substances can exist in either of two stable phases which exhibit the property
of fluidity, or easy deformability. The density of a substance in the liquid phase is
normally much larger than that in the gaseous phase, but this is not in itself a significant
basis for distinction since it leads mainly to a difference in the magnitudes of forces
required to produce given magnitudes of acceleration rather than to a difference in the
types of motion. The most important difference between the mechanical properties of

8
Chapter 2. Literature Review 9

liquids and gases lies in their bulk elasticity, that is, in their compressibility. Gases
can be compressed much more readily than liquids, and as a consequence any motion
involving appreciable variations in pressure will be accompanied by much larger changes
in specific volume in the case of a gas than in the case of a liquid.

Fluid mechanics is concerned with understanding, predicting, and controlling the be-
havior of a fluid. Since we live in a dense gas atmosphere on a planet mostly covered by
liquid, a rudimentary grasp of fluid mechanics is part of everyday life. For an engineer,
fluid mechanics is an important field of the applied sciences with many practical and
exciting applications. If you examine municipal water, sewage, and electrical systems,
you will notice a heavy dependence on fluid machinery. Pumps and steam turbines are
obvious components of these systems, as are the valves and piping found in your home,
under your city streets, in the Alaska oil pipeline, and in the natural gas pipelines that
crisscross the country. Moreover, aircraft, automobiles, ships, spacecraft, and virtually
all other vehicles involve interactions with fluid of one type or another, both externally
and internally, within an engine or as part of a hydraulic control system. Learning
more about fluid mechanics also allows us to better understand our bodies and many
interesting features of our environment. The heart and lungs, for example, are wonder-
fully designed pumps that operate intermittently rather than steadily as most man-made
pumps do. Yet the heart moves blood efficiently through the branching network of arter-
ies, capillaries, and veins, and the lungs cycle air quite effectively through the branching
pulmonary passages, thereby keeping the cells of our bodies alive and functioning. Many
other sophisticated fluid handling devices are found throughout the biological world in
living creatures of all types, sizes, and degree of complexity. The environment is another
source of complex and interesting fluid mechanics problems. These range from the pre-
diction of weather, hurricanes, and tornadoes to the spread and control of air and water
pollution. Add to this list the flow of rivers and streams, the movement of groundwa-
ter, the jet stream and great ocean currents, and the tidal flows in estuaries. The lava
flows of volcanoes and the movements of molten rock within the earth also lie within
the domain of fluid mechanics. Looking beyond Earth, stellar processes and interstellar
events are striking examples of fluids in motion on a grand scale. Knowledge of fluid
mechanics is also the key to understanding and sometimes controlling other interesting,
if not vital phenomena, such as the curving flight of a tennis, golf, or soccer ball, and
the many different pitches in baseball.

The field of fluid mechanics has historically been divided into two branches, fluid statics
and fluid dynamics. Fluid statics, or hydrostatics, is concerned with the behavior of a
fluid at rest or nearly so. Fluid dynamics involves the study of a fluid in motion.
Chapter 2. Literature Review 10

Modern engineering science is rooted in the ability to create and solve mathematical
models of physical systems. Fluid mechanics is often considered as a challenging subject,
primarily because the underlying mathematical model appears to be complex and diffi-
cult to apply. The governing equation of fluid statics, called the hydrostatic equation,
is actually relatively simple and may always be solved to find the pressure distribution
in the fluid. On the other hand, the governing equation of fluid dynamics, called the
Navier-Stokes equation, would never be described as simple. The inherent difficulties
of fluid mechanics have been recognized for centuries, yet engineers have demonstrated
great ingenuity in developing a number of different approaches to solving specific fluid
flow problems. The common theme is to simplify the mathematical or experimental
model used to describe the flow without sacrificing the relevant physical phenomena.
The art of fluid mechanics, however, is developed primarily through experience, both
your own and that of others. This art consists in knowing when it is safe to neglect the
effects of physical phenomena that are judged to have little impact on the flow. Once
we decide to neglect certain physical phenomena, we drop the corresponding terms in
the governing equations, thereby decreasing the difficulty in obtaining a solution. Such
fundamental topics as boundary layer theory, the Bernoulli equation, potential flow, and
even fluid statics can be considered to be part of the art of fluid mechanics in this sense.
Learning about these historical approximations and others like them is essential in fluid
mechanics, and engineers working with fluids should continue to seek to attain it.

Once the analysis of a fluid mechanics problem has been cast in the form of an ap-
propriate mathematical model, a solution method must be chosen. For example, one
might employ an analytical solution method that results in a representation of the flow
variables as functions of space and time. An analytical solution is a highly compact and
useful form of solution that should always be acquired if possible. Be aware, however,
that an analytical solution of the governing equations of fluid dynamics is usually not
possible. Complex engineering geometries and a natural tendency for fluid flows to be-
come unstable ensure that analytical solutions will remain elusive. Nevertheless, it is
wise to consult the engineering literature to determine what has been accomplished in
treating the same or related flow problems. If you find that an approximate analytical
solution to a problem of current interest is available, you may be able to use it as the
starting point for your analysis. Today, an engineer will increasingly choose to employ
computational methods to solve the equations of fluid motion. These methods include
finite difference, finite element, finite volume, and other computational approaches in
which digital computers are used to supply numerical solutions of approximate versions
of the governing equations. These solutions are discrete, meaning that the flow variables
are known only at specific spatial locations in the flow field. Computational tools of all
Chapter 2. Literature Review 11

kinds, ranging from commercially available computational fluid dynamic codes to visu-
alization packages and symbolic mathematics codes, are among the most important aids
in the modern practice of fluid mechanics. The one motivation in writing this thesis is
to integrate some of the modern computational aids into fluid mechanics. The computer
programming languages C, C++, Mathematica, Maple, Matlab, and others like them
are superb aids in learning fluid mechanics. It is recommended using them to simplify
calculations and to visualize the mathematics.

2.1.1 Some basic properties of the fluid

i. Density, specific weight, specific volume


The density of a fluid is defined as the mass per unit volume. The symbol most
often used for density is ρ (the lower case Greek letter rho). Mathematically, the
density ρ at a point P may be defined as

δm
ρ = limδv→0 δv

Where δv is the volume element around P and δm is the mass of the fluid within
δv.
The specific weight γ of a fluid is defined as the weight per unit volume. Thus
γ = ρg, where g is the acceleration due to gravity.
The specific volume of a fluid is defined as the volume per unit mass and is clearly
the reciprocal of the density.

ii. Pressure
When a fluid is contained in a vessel, it exerts a force at each point of the inner
side of the vessel. Such a force per unit area is known as pressure. Mathematically,
the pressure p at a point P may be defined as

δF
p = limδS→0 δS

Where δS is an elementary area around P and δF is the normal force due to fluid
on δS.

iii. Temperature
Suppose two bodies of diffferent heat content are brought into contact while iso-
lated from all other bodies. Then some thermal energy will move from one body
into the other body. The body from where the thermal energy moves is said to be
at a higher temperature while the body into which the energy flows is said to be
at a lower temperature. When two bodies are in thermal equilibrium then they
are said to have a common property, known as temperature T.
Chapter 2. Literature Review 12

iv. Thermal conductivity


The well-known Fourier heat conduction law states that the conductive heat flow
per unit area( or heat flux) qn is proportional to the temperature decrease per
unit distance in a direction normal to the area through which the heat is flowing.
Thus, mathematically

qn ∝ − ∂T ∂T
∂n so that qn = −k ∂n

where k is said to be the thermal conductivity

v. Specific heat
The specific heat C is of a fluid is defined as the amount of heat required to raise
the temperature of a unit mass of the fluid by one degree. Thus

∂Q
C= ∂T

where δQ is the amount of heat added to raise the temperature by δT . The value of
the specific heat depends on two well-known process- the constant volume process
and the constant pressure process. The specific heats of the above process are
denoted and defined as
 
∂Q
Specific heat at constant volume=Cv =
∂T v
∂Q
Specific heat at constant pressure=Cp = ∂T p
Cp
Ratio of these two specific heats is denoted by γ. Thus γ = Cv

vi. Incompressible and Compressible fluids


Gasses are compressible and their density changes readily with temperature and
pressure. Liquids, on the other hand, are rather difficult to compress and for most
problems we can treat them incompressible. Only in such situations as sound
propagation in liquids does one need to consider their compressibility.

The density of a fluid is a thermodynamic property which depends on the state of


the fluid. The density ρ can be expressed as a function of pressure and temperature.
Such a relation is known as an equation of state. For an ideal gas the equation of
state may be expressed as

P = ρRT

where R is the characteristic gas constant. The constant R has different values for
various gases and its units has the form

energy J
R= =
massXtemperature kg − K
Chapter 2. Literature Review 13

we also have R = Cp − Cv
where Cp is specific heat at constant pressure and Cv is specific heat at constant
volume.

vii. Compressibility and Bulk modulus


The compressibility of a fluid is expressed by the quantity bulk modulus K, which
is defined as the ratio of volumetric stress to volumetric strain. If a small increase
in pressure dp causes a change dv of the specific volume v, then by definition

dp dp dp 1 dv dρ
K= = =ρ as v ∝ =⇒ =−
dv/v dρ/ρ dρ ρ v ρ

The coefficient of compressibility β is defined as β = 1/K.


We now proceed to establish relationship between the bulk modulus and the local
pressure for a perfect gas for two different types of compression.

Relationship for isothermal process. We have

p/ρ = Const. = C1 (2.1)

From the equation of state for a perfect gas,

p = ρRT (2.2)

From (2.2),
dp/dρ = RT = p/ρ (2.3)

By definition, K = ρ(dp/dρ) = ρX(p/ρ) = p , by (2.3)

Relationship for isentropic process. We have

p/ργ = Const. = C2 (2.4)

Where γ = Cp /Cv ; Cv being specific heat at constant volume and Cp being specific
heat at constant pressure.
From (2.2) and (2.4), dp/dρ = RT = γργ−1 C2
By definition, K = ρ(dp/dρ) = γργ C2 = γp, by (2.4)

viii. Viscosity ( or internal friction)


The viscosity of a fluid is a measure of its resistance to gradual deformation by
shear stress or tensile stress. For liquids, it corresponds to the informal notion of
Chapter 2. Literature Review 14

Figure 2.1: Couette flow

”thickness”. For example, honey has a higher viscosity than water [21].

Viscosity is due to friction between neighboring parcels of the fluid that are mov-
ing at different velocities. When fluid is forced through a tube, the fluid generally
moves faster near the axis and very slowly near the walls, therefore some stress
(such as a pressure difference between the two ends of the tube) is needed to over-
come the friction between layers and keep the fluid moving. For the same velocity
pattern, the stress required is proportional to the fluid’s viscosity. A liquid’s vis-
cosity depends on the size and shape of its particles and the attractions between
the particles.

Dynamic (shear) viscosity


The dynamic (shear) viscosity of a fluid expresses its resistance to shearing flows,
where adjacent layers move parallel to each other with different speeds. It can be
defined through the idealized situation known as a Couette flow, where a layer of
fluid is trapped between two horizontal plates, one fixed and one moving horizon-
tally at constant speed u. (The plates are assumed to be very large, so that one
need not consider what happens near their edges). See Fig. 2.1

If the speed of the top plate is small enough, the fluid particles will move parallel
to it, and their speed will vary linearly from zero at the bottom to u at the top.
Each layer of fluid will move faster than the one just below it, and friction between
them will give rise to a force resisting their relative motion. In particular, the fluid
will apply on the top plate a force in the direction opposite to its motion, and an
equal but opposite one to the bottom plate. An external force is therefore required
in order to keep the top plate moving at constant speed. The magnitude F of this
Chapter 2. Literature Review 15

force is found to be proportional to the speed u and the area A of each plate, and
inversely proportional to their separation y. That is,

u
F = µA (2.5)
y

The proportionality factor µ in this formula is the viscosity (specifically, the dy-
namic viscosity) of the fluid. The ratio u/y is called the rate of shear deformation
or shear velocity, and is the derivative of the fluid speed in the direction perpen-
dicular to the plates. Isaac Newton expressed the viscous forces by the differential
equation
∂u
τ =µ (2.6)
∂y
where τ = F/A and ∂u/∂y is the local shear velocity. This formula assumes that
the flow is moving along parallel lines and the y axis, perpendicular to the flow,
points in the direction of maximum shear velocity. This equation can be used
where the velocity does not vary linearly with y, such as in fluid flowing through
a pipe. Equation (2.6) is known as Newtons law of viscosity.

Kinematic viscosity
The kinematic viscosity is the ratio of the dynamic viscosity µ to the density of
the fluid ρ. It is usually denoted by the Greek letter ν. It is a convenient concept
when analyzing the Reynolds number, that expresses the ratio of the inertial forces
to the viscous forces:
ρuL uL
Re = = , (2.7)
µ ν

where L is a typical length scale in the system.


Bulk viscosity
When a compressible fluid is compressed or expanded evenly, without shear, it may
still exhibit a form of internal friction that resists its flow. These forces are related
to the rate of compression or expansion by a factor s, called the volume viscosity,
bulk viscosity or second viscosity. The bulk viscosity is important only when the
fluid is being rapidly compressed or expanded, such as in sound and shock waves.
Bulk viscosity explains the loss of energy in those waves, as described by Stokes’
law of sound attenuation.

Viscosity tensor
In general, the stresses within a flow can be attributed partly to the deformation of
the material from some rest state (elastic stress), and partly to the rate of change
of the deformation over time (viscous stress). In a fluid, by definition, the elastic
Chapter 2. Literature Review 16

stress includes only the hydrostatic pressure. In very general terms, the fluid’s
viscosity is the relation between the strain rate and the viscous stress. In the
Newtonian fluid model, the relationship is by definition a linear map, described by
a viscosity tensor that, multiplied by the strain rate tensor (which is the gradient
of the flow’s velocity), gives the viscous stress tensor. The viscosity tensor has nine
independent degrees of freedom in general. For isotropic Newtonian fluids, these
can be reduced to two independent parameters. The most usual decomposition
yields the stress viscosity µ and the bulk viscosity σ.

2.1.1.1 Classification of fluids on the basis of density and viscosity

Fluids can be classified on the basis of density and viscosity as

a. Compressible and Incompressible fluids


b. Viscous and Inviscid fluids
c. Newtonian and Non-Newtonian fluids
d. Real and Ideal fluids

a. Compressible and Incompressible fluids


Compressible flow describes the behavior of fluids that experience significant vari-
ations in density. For flows in which the density does not vary significantly, the
analysis of the behavior of such flows may be simplified greatly by assuming a
constant density. This is an idealization, which leads to the theory of incompress-
ible flow. However, in the many cases dealing with gases (especially at higher
velocities) and those cases dealing with liquids with large pressure changes, the
significant variations in density can occur, and the flow should be analyzed as a
compressible flow if accurate results are to be obtained. The incompressible fluids
have constant density while the compressible fluids have variable density.

b.Viscous ( or real) and Inviscid (non-viscous,frictionless, perfect or


ideal) fluids
An infinitesimal fluid element is acted upon by two types of forces, namely, body
forces and surface forces. The body force is a type of force which is proportional
to the mass (or possibly the volume) of the body on which it acts while the surface
force is one which acts on a surface element and is proportional to the surface area.
Suppose that the fluid element be enclosed by the surface S. Let P be an arbitrary
point of S and let dS be the surface element around P. Then the surface force on
dS is, in general, not in the direction of normal at P to dS. Hence the force may
be resolved into components, one normal and the other tangential to the area dS.
Chapter 2. Literature Review 17

The normal force per unit area is sid to be the normal stress or pressure while the
tangential force per unit area is said to be the shearing stress.

A fluid is said to be viscous when the normal as well as shearing stress exist. On
the other hand, a fluid is said to be inviscid when it does not exert any shearing
stress, whether at rest or in motion. Clearly the pressure exerted by an inviscid
fluid on any surface is always along the normal to the surface at that point. Due
to shearing stress a viscous fluid produces resistance to the body moving through
it as well as between the particles of the fluid itself. Water and air are treated
inviscid fluids where as syrup and heavy oil are treated as viscous.

c. Newtonian and Non-Newtonian fluids


The fluids that obeys Newtonian law of viscosity (2.6) is known as Newtonian
fluid, for example, air and water. Viscous fluids such as tar and polymers do not
obey Newton’s law of of viscosity and the relation between stress and rate of shear
strain is non-linear Such fluids are known as non-Newtonian fluids.

d. Real and Ideal fluids


The concept of an ideal fluid is based on theoretical consideration because all real
fluids exhibit viscous property. From relation (2.6), it follows that τ vanishes either
for µ = 0 or for ∂u/∂y = 0. Hence, we see that a real fluid with small viscosity
and small velocity gradient can be regarded as a frictionless. For an ideal fluid,
w shall take shear stress as zero and assume that, at the surface of contact with
solid, and ideal fluid can have relative velocity in the tangential direction although
normal velocity must be zero at the surface of contact.

ix. The Mach number, subsoinc and supersonic flows


In compressible flows there is a great distinction between flow involving velocities
less than that of sound (subsonic flow) and flow involving velocities greater than
that of sound (supersonic flow).

The Mach number M is defined as the ratio of of the fluid speed to the local speed
V
of sound. Thus, M = c , where V is the fluid velocity and c the local speed of
sound.
Chapter 2. Literature Review 18

2.1.1.2 Classification of fluids on the basis of Mach number

The Mach number gives us an important information about the type of compressible
flow. The compressible flows can be classified on the basis of the Mach number as:

• Subsonic (M < 1)

• Sonic (M = 1)

• Supersonic (1 < M < 6)

• Hypersonic (M > 6)

• Transonic (M < 1aswellasM > 1)

2.1.2 Some important types of flows

I. Laminar (streamline) and turbulent flows


A flow, in which each fluid particle traces out a definite curve and the curves traced
out by any two different fluid particles do not intersect, is said to be laminar. On
the other hand, a flow, in which each fluid particle does not trace out a definite
curve and the curves traced out by fluid particles intersect, is said to be turbulent.

II. Steady and unsteady flows


A flow, in which properties and conditions ( P, say) associated with the motion of
the fluid are independent of the tie so that the flow pattern remains unchanged
with the time, is said to be steady. Mathematically, we may write ∂P/∂t = 0.
Her P may be velocity, density, pressure, temperature etc. On the other hand,
a flow, in which properties and conditions associated with the motion of the the
fluid depend on the time so that the flow pattern varies with time, is said to be
unsteady

III. Uniform and non-uniform flows


A flow, in which the fluid particles possess equal velocities at each section of the
channel or pipe is called uniform. On the other hand, a flow, in which the fluid
particles possess different velocities at each section of the channel or pipe is called
non-uniform. These terms are usually used in connection with flow in channels.
Chapter 2. Literature Review 19

IV. Rotational and Irrotational flows


A flow, in which the fluid particles go on rotating about their own axes, while
flowing, is said to be rotational. On the other hand, a flow in which the fluid
particles do not rotate about their own axes, while flowing, is said to be irrotational.

V. Barotropic flow

The flow is said to be barotropic when the pressure is a function of the density
[22].

2.2 Heat and Mass Transfer

2.2.1 Heat Transfer

The concept of energy is used in thermodynamics to specify the state of a system. It is


a well-known fact that energy is neither created nor destroyed but only changed from
one form to another. The science of thermodynamics deals with the relation between
heat and other forms of energy, but the science of heat transfer is concerned with the
analysis of the rate of heat transfer taking place in a system. The energy transfer by
heat flow cannot be measured directly, but the concept has physical meaning because it
is related to the measurable quantity called temperature. It has long been established
by observations that when there is temperature difference in a system, heat flows from
the region of high temperature to that of low temperature. Since heat flow takes place
whenever there is a temperature gradient in a system, a knowledge of the temperature
distribution in a system is essential in heat transfer studies. Once the temperature dis-
tribution is known, a quantity of practical interest, the heat flux, which is the amount of
heat transfer per unit area per unit time, is readily determined from the law relating the
heat flux to the temperature gradient. The problem of determining temperature distri-
bution and heat flow is of interest in many branches of science and engineering. In the
design of heat exchangers such as boilers, condensers, radiators, etc., for example, heat
transfer analysis is essential for sizing such equipment. In the design of nuclear-reactor
cores, a thorough heat transfer analysis of fuel elements is important for proper sizing
of fuel elements to prevent burnout. In aerospace technology, the temperature distri-
bution and heat transfer problems are crucial because of weight limitations and safety
considerations. In heating and air conditioning applications for buildings, a proper heat
transfer analysis is necessary to estimate the amount of insulation needed to prevent
excessive heat losses or gains. In the studies of heat transfer, it is customary to consider
three distinct modes of heat transfer: conduction, convection, and radiation. In reality,
Chapter 2. Literature Review 20

temperature distribution in a medium is controlled by the combined effects of these three


modes of heat transfer; therefore, it is not actually possible to isolate entirely one mode
from interactions with the other modes. However, for simplicity in the analysis, one
can consider, for example, conduction separately whenever heat transfer by convection
and radiation is negligible. With this qualification, we present below a brief qualitative
description of these three distinct modes of heat transfer.

2.2.1.1 Conduction

Conduction is the mode of heat transfer in which energy exchange takes place from the
region of high temperature to that of low temperature by the kinetic motion or direct
impact of molecules, as in the case of fluid at rest, and by the drift of electrons, as in
the case of metals. In a solid which is a good electric conductor, a large number of free
electrons move about in the lattice; hence materials that are good electric conductors
are generally good heat conductors (i.e., copper, silver, etc.).
The empirical law of heat conduction based on experimental observations originates-from
Biot but is generally named after the French mathematical physicist Joseph Fourier (2.4)
who used it in his analytic theory of heat. This law 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. For heat flow in the x
direction, for example, the Fourier law is given as

dT
Q̇x = −kA W (2.8)
dx
or
Qx dT
qx = = −k W/m2
A dx

where Q̇x is the rate of heat flow through area A in the positive x direction and qx is called
the heat flux in the positive x direction. The proportionality constant k is called the
thermal conductivity of the material and is a positive quantity. If temperature decreases
in the positive x direction, then dT /dx is negative; hence qx (or Qx ) becomes a positive
quantity because of the presence of the negative sign in Eq. (2.4). Therefore, the minus
sign is included in Eq. (2.4) to ensure that qx (or Qx ) is a positive quantity when the heat
flow is in the positive x direction. Conversely, when the right-hand side of Eq. (2.4) is
negative, the heat flow is in the negative x direction. The thermal conductivity k in Eq.
2.4 must have the dimensions W/(mo C)orJ/(mso C) if the equations are dimensionally
correct. There is a wide difference in the range of thermal conductivities of various
engineering materials. Between gases and highly conducting metals, such as copper or
Chapter 2. Literature Review 21

silver, k varies by a factor of about 104 . The highest value is for highly conducting
pure metals, and the lowest value is for gases and vapors, excluding the evacuated
insulating systems. The nonmetallic solids and liquids have thermal conductivities that
lie between them. Metallic single crystals are exceptions, which may have very high
thermal conductivities; for example, with copper crystals, values of 8000W/(mo C) and
even higher are possible.
Thermal conductivity also varies with temperature. This variation, for some materials
over certain temperature ranges, is smalt enough to be neglected; but for many cases
the variation of k with temperature is significant. Especially at very low temperatures
k varies rapidly with temperature; for example, the thermal conductivities of copper,
aluminum, or silver reach values 50 to 100 times those that occur at room temperature.

2.2.1.2 Convection

When fluid flows over a solid body or inside a channel while temperatures of the fluid
and the solid surface are different, heat transfer between the fluid and the solid surface
takes place as a consequence of the motion of fluid relative to the surface; this mechanism
of heat transfer is called convection. If the fluid motion is artificially induced, say with
a pump or a fan that forces the fluid flow over the surface, the heat transfer is said to
be by forced convection. If the fluid motion is set up by buoyancy effects resulting from
density difference caused by temperature difference in the fluid, the heat transfer is said
to be by free (or natural ) convection. For example, a hot plate vertically suspended in
stagnant cool air causes a motion in the air layer adjacent to the plate surface because
the temperature gradient in the air gives rise to a density gradient, which in turn sets up
the air motion. As the temperature field in the fluid is influenced by the fluid motion, the
determination of temperature distribution and of heat transfer in convection for most
practical situations is a complicated matter. In engineering applications, to simplify the
heat transfer calculations between a hot surface at Tw and a cold fluid flowing over it at
a bulk temperature Tf , a heat transfer coefficient h is defined as

q = h(Tw − Tf ) (2.9)

where q is the heat flux (in watts per square meter) from the hot wall to the cold fluid.
Alternatively, for heat transfer from the hot fluid to the cold wall, Eq. 2.9 is written as

q = h(Tf − Tw ) (2.10)

where q represents the heat flux from the hot fluid to the cold wall. Historically, the
form given by Eq. (2.9) was first used as a law of cooling as heat is removed from a
Chapter 2. Literature Review 22

body to a liquid flowing over it, and it is generally referred to as ”Newton’s law of
cooling.” If the heat flux in Eqs. (2.9) and (2.10) is given in watts per square meter and
the temperatures are in degrees Celsius (or kelvins), then the heat transfer coefficient
h in Eqs. (2.9) and (2.10) must have the dimensions W/(m2o C) if the equations are
dimensionally correct.
The heat transfer coefficient h varies with the type of flow (i.e., laminar or turbulent),
the geometry of the body and flow passage area, the physical properties of the fluid, the
average temperature, and the position along the surface of the body. It also depends on
whether the mechanism of heat transfer is by forced convection (i.e., the fluid motion
is caused by a pump or a blower) or by natural convection (i.e., the fluid motion is
caused by the buoyancy). When h varies with the position along the surface of the
body, for convenience in engineering applications, its average value hm over the surface
is considered instead of its local value h. Equations (2.9) and (2.10) are also applicable
for such cases by merely replacing h by hm ; then q represents the average value of the
heat flux over the region considered.
The heat transfer coefficient can be determined analytically for flow over bodies having
a simple geometry such as a flat plate or flow inside a circular tube. For flow over bodies
having complex configurations, the experimental approach is used to determine h. There
is a wide difference in the range of the values of the heat transfer coefficient for various
applications.

2.2.1.3 Radiation

All bodies continuously emit energy because of their temperature, and the energy thus
emitted is called thermal radiation. The radiation energy emitted by a body is trans-
mitted in the space in the form of electromagnetic waves according to Maxwell’s classic
electromagnetic wave theory or in the form of discrete photons according to Planck’s hy-
pothesis. Both concepts have been utilized in the investigation of radiative-heat transfer.
The emission or absorption of radiation energy by a body is a bulk process; that is, radia-
tion originating from the interior of the body is emitted through the surface. Conversely,
radiation incident on the surface of a body penetrates to the depths of the medium where
it is attenuated. When a large proportion of the incident radiation is attenuated within
a very short distance from the surface, we may speak of radiation as being absorbed
or emitted by the surface. For example, thermal radiation incident on a metal surface
is attenuated within a distance of a few angstroms from the surface; hence metals are
opaque to thermal radiation.
Chapter 2. Literature Review 23

The solar radiation incident on a body of water is gradually attenuated by water as the
beam penetrates to the depths of water. Similarly, the solar radiation incident on a sheet
of glass is partially absorbed and partially reflected, and the remaining is transmitted.
Therefore, water and glass- are considered semitransparent to the solar radiation.

It is only in a vacuum that radiation propagates with no attenuation at all. Also the
atmospheric air contained in a room is considered transparent to thermal radiation for
all practical purposes, because the attenuation of radiation by air is insignificant unless
the air layer is several kilometers thick. However, gases such as carbon dioxide, carbon
monoxide, water vapor, and ammonia absorb thermal radiation over certain wavelength
bands; therefore they are semitransparent to thermal radiation. It is apparent from
the previous discussion that a body at a temperature T emits radiation owing to its
temperature; also a body absorbs radiation incident on it. Here we briefly discuss the
emission and absorption of radiation by a body.

Emission of Radiation The maximum radiation flux emitted by a body at temperature


T is given by the Stefan-Boltzmann law

Eb = σT 4 W/m2 (2.11)

where T is the absolute temperature in kelvins, σ is the Stefan-Boltzmann constant


[σ = 5.6697x10−8 W/(m2 K 4 )], and Eb is called the blackbody emissive power.

Only an ideal radiator or the so-called blackbody can emit radiation flux according to
Eq. (2.11). The radiation flux emitted by a real body at an absolute temperature T is
always less than that of the blackbody emissive power Eb , it is given by

q = Eb = σT 4 (2.12)

where the emissivity  lies between zero and unity; for all real bodies it is always less
than unity.

Absorption of Radiation If a radiation flux qinc is incident on a blackbody, it is


completely absorbed by the blackbody. However, if the radiation flux qinc is incident
on a real body, then the energy absorbed qabs by the body is given by

qabs = αqinc (2.13)

where the absorptivity a lies between zero and unity; for all real bodies it is always less
than unity.
Chapter 2. Literature Review 24

The absorptivity α of a body is generally different from its emissivity . However, in


many practical applications, to simplify the analysis, αis assumed to equal .

Radiation Exchange When two bodies at different temperatures ”see” each other,
heat is exchanged between them by radiation. If the intervening medium is filled with a
substance such as air which is transparent to radiation, the radiation emitted from one
body travels through the intervening medium with no attenuation and reaches the other
body, and vice versa. Then the hot body experiences a net heat loss, and the cold body
a net heat gain, as a result of the radiation heat exchange.

Combined heat transfer mechanism So far we have considered the heat transfer
mechanism, conduction, convection, and radiation separately. In many practical sit-
uations heat transfer from a surface takes place simultaneously by convection to the
ambient air and by radiation to the surroundings. Consider a small plate of area A and
emissivity ε that is maintained at Tw and exchanges energy by convection with a fluid
at T∞ with a heat transfer coefficient hc and by radiation with the surroundings at Ts .
The heat loss per unit area of the plate, by the combined mechanism of convection and
radiation, is given by
qw = hc (Tw − T ∞) + σ(Tw 4 − Ts 4 ) (2.14)

If |Tw − Ts |  Tw the second term can be linearized. We obtain

qw = hc (Tw − T∞ ) + hr (Tw − Ts ) (2.15)

Where hr = 4εσTw 3 .

2.2.2 Mass Transfer

Distinction should be made between mass transfer and the bulk fluid motion (or fluid
flow) that occurs on a macroscopic level as a fluid is transported from one location
to another. Mass transfer requires the presence of two regions at different chemical
compositions, and mass transfer refers to the movement of a chemical species from a
high concentration region toward a lower concentration one relative to the other chemical
species present in the medium. The primary driving force for fluid flow is the pressure
difference, whereas for mass transfer it is the concentration difference. Therefore, we do
not speak of mass transfer in a homogeneous medium.

It is a common observation that whenever there is an imbalance of a commodity in a


medium, nature tends to redistribute it until a ”balance” or ”equality” is established.
This tendency is often referred to as the driving force, which is the mechanism behind
Chapter 2. Literature Review 25

many naturally occurring transport phenomena. If we define the amount of a commod-


ity per unit volume as the concentration of that commodity, we can say that the flow
of a commodity is always in the direction of decreasing concentration; that is, from the
region of high concentration to the region of low concentration. The commodity simply
creeps away during redistribution, and thus the flow is a diffusion process. The rate of
flow of the commodity is proportional to the concentration gradient dC/dx, which is the
change in the concentration C per unit length in the flow direction x, and the area A
normal to flow direction and is expressed as

Flow rate ∝ (Normal area)(concentration gradient)

dC
Q̇ = −kdif f A (2.16)
Dx
Here the proportionality constant kdif f is the diffusion coefficient of the medium, which
is a measure of how fast a commodity diffuses in the medium, and the negative sign
is to make the flow in the positive direction a positive quantity (note that dC/dx is a
negative quantity since concentration decreases in the flow direction).

The mechanisms of heat and mass transfer are analogous to each other, and thus we can
develop an understanding of mass transfer in a short time with little effort by simply
drawing parallels between heat and mass transfer. Temperature
The driving force for heat transfer is the temperature difference. In contrast, the driving
force for mass transfer is the concentration difference. Therefore, both heat and mass
are transferred from the more concentrated regions to the less concentrated ones. If
there is no temperature difference between two regions, then there is no heat transfer.
Likewise, if there is no difference between the concentrations of a species at different
parts of a medium, there will be no mass transfer.

Heat is transferred by conduction, convection, and radiation. Mass, however, is trans-


ferred by conduction (called diffusion) and convection only, and there is no such thing
as ”mass radiation”.

2.2.2.1 Diffusion

Similar to the Fourier’s law of heat conduction (2.4), the rate of mass diffusion ṁdif f
of a chemical species A in a stationary medium in the direction x is proportional to
the concentration gradient dC/dx in that direction and is expressed by Fick’s law of
diffusion by
dC
ṁdif f = −DA (2.17)
dx
Chapter 2. Literature Review 26

where D is the diffusion coefficient (or mass diffusivity) of the species in the mixture
and C is the concentration of the species in the mixture.

2.2.2.2 Convection

You will recall that heat convection is the heat transfer mechanism that involves both
heat conduction (molecular diffusion) and bulk fluid motion. Fluid motion enhances
heat transfer considerably by removing the heated fluid near the surface and replacing
it by the cooler fluid further away. In the limiting case of no bulk fluid motion, convec-
tion reduces to conduction. Likewise, mass convection (or convective mass transfer) is
the mass transfer mechanism between a surface and a moving fluid that involves both
mass diffusion and bulk fluid motion. Fluid motion also enhances mass transfer con-
siderably by removing the high concentration fluid near the surface and replacing it by
the lower concentration fluid further away. In mass convection, we define a concentra-
tion boundary layer in an analogous manner to the thermal boundary layer and define
new dimensionless numbers that are counterparts of the Nusselt and Prandtl numbers.
The rate of heat convection for external flow was expressed conveniently by Similar to
Newton’s law of cooling (2.10), the rate of mass convection can be expressed as

ṁconv = hmass As (Cs − C∞ ) (2.18)

where hmass is the mass transfer coefficient, As is the surface area, and Cs − C∞ is a
suitable concentration difference across the concentration boundary layer.

Convection is the concerted, collective movement of groups or aggregates of molecules


within fluids (e.g., liquids, gases) and rheids, either through advection or through dif-
fusion or as a combination of both of them. Convection of mass cannot take place in
solids, since neither bulk current flows nor significant diffusion can take place in solids.
Diffusion of heat can take place in solids, but that is called heat conduction. Convection
can be demonstrated by placing a heat source (e.g. a Bunsen burner) at the side of a
glass full of a liquid, and observing the changes in temperature in the glass caused by
the warmer fluid moving into cooler areas.

Convective heat transfer is one of the major modes of heat transfer, and convection is also
a major mode of mass transfer in fluids. Convective heat and mass transfer take place
both by diffusion - the random Brownian motion of individual particles in the fluid
- and by advection, in which matter or heat is transported by the larger-scale motion
of currents in the fluid. In the context of heat and mass transfer, the term ”convection”
is used to refer to the sum of advective and diffusive transfer. In common use the
term ”convection” may refer loosely to heat transfer by convection, as opposed to mass
Chapter 2. Literature Review 27

transfer by convection, or the convection process in general. Sometimes ”convection”


is even used to refer specifically to ”free heat convection” (natural heat convection) as
opposed to forced heat convection. However, in mechanics the correct use of the word
is the general sense, and different types of convection should be properly qualified for
clarity. Convection can be qualified in terms of being natural, forced, gravitational,
granular, or thermomagnetic.

2.3 Basic Equations

2.3.1 Equation of motion

Our understanding of the physical significance of various parameters in influencing the


friction factor and the drag coefficient can be improved if we have some knowledge of the
equations governing the fluid motion. We present here the equations of continuity and
momentum for the two-dimensional, steady motion of constant-property, incompressible
Newtonian fluid in the two-dimensional, rectangular coordinate system for the x and
y variables. We also illustrate the simplification of these equations for simpler flow
conditions.

2.3.1.1 Continuity Equation

The continuity equation is essentially the equation for the conservation of mass which
states that fluid mass can be neither created nor destroyed; it is derived by a mass
balance on the fluid entering and leaving a volume element taken in the flow field.
Consider a differential volume element ∆x∆yl about a point (x, y) in the flow field, as
illustrated in Fig. (2.2). The equation for the conservation of mass for two-dimensional
steady flow may be stated as:

   
N et rate of mass f low net rate of mass f low
   
 entering volume element  +  entering volume element  = 0 (2.19)
   
in x direction in y direction

Let u = u(x, y) and v = v(x, y) be the velocity components in the flow in the x and
y directions, resp.. If Mx = ρu∆y.l is the mass flow rate into the element in the
x direction through the surface at x, then Mx + (∂Mx /∂x) ∆x is the mass flow rate
leaving the element in the x direction through the surface at x + ∆x. The net rate of
mass flow into the element in the x direction is the difference between the entering and
Chapter 2. Literature Review 28

Figure 2.2: Nomenclature for the derivation of the continuity equation.

leaving flow rates, given by


!
N et rate of mass f low entering ∂Mx ∂(ρu)
=− ∆x = − ∆x∆yl (2.20)
element in x direction ∂x ∂x

Similarly, for the y direction, we write


!
N et rate of mass f low entering ∂My ∂(ρv)
=− ∆x = − ∆x∆yl (2.21)
element in y direction ∂y ∂y

Substituting Eqs. (2.20) and (2.21) in (2.19), we obtain

∂(ρu) ∂(ρv)
+ =0 (2.22)
∂x ∂y
When density ρ is treated as a constant, Eq. (2.22) simplifies to

∂u ∂v (2.23)
∂x + ∂y =0

Equation (2.23) is the continuity equation in rectangular coordinates for the steady,
two-dimensional flow of an incompressible fluid.

2.3.1.2 Momentum Equations

The momentum equations are derived from Newton’s second law of motion, which states
that mass times the acceleration in a given direction is equal to the external forces acting
on the body in the same direction. The external forces acting on a volume element in a
Chapter 2. Literature Review 29

flow field are considered to consist of the body forces and the surface forces. The body
forces may result from such effects as the gravitational, electric, and magnetic fields
acting on the body of the fluid, and the surface forces result from the stresses acting on
the surface of the volume element. With this consideration Newton’s second law may
be stated for flow in direction i as
     
acceleration body f orces surf ace f orce
     
(M ass)  in  =  acting in
 
+
  acting in 
 (2.24)
direction i direction i direction i

for a three-dimensional flow, for example, in the rectangular coordinate system i = x,


y, and z; hence Eq. (2.24) provides three independent momentum equations. In this
analysis, we consider two-dimensional, steady, incompressible, constant property flow
in the (x, y) rectangular coordinate system. Therefore, for i = x and y, Eq. (2.24)
will provide two momentum equations, one for the x direction and the other for the y
direction.
Let u = u(x, y) and v = v(x, y) be the velocity components in the x and y directions,
respectively. We consider a volume element ∆x∆yl about a point (x, y) in the flow field.
Various terms in Eq. (2.24) are determined as follows:
First, if ρ is the density of the fluid, the mass term in this equation is given by

(M ass) = (∆x∆yl)ρ (2.25)

Second, Eq. (2.24) contains a term called acceleration. Commonly, acceleration implies
time rate of change of velocity, but for the steady flow considered here, there is an
acceleration associated with the convective motion of fluid in other directions, because
we have a two-dimensional flow. Consider, for example, motion of the fluid in the
x direction. If u = u(x, y) is the velocity component in the x direction, there is an
acceleration of the fluid in the x direction associated with the motion of the fluid in
other directions, given by
!
Acceleration ∂u ∂u
=u +v (2.26)
in x direction ∂x ∂y

The derivation of Eq. (2.26) is as follows :


Consider u = u(x, y) for the two-dimensional steady flow. The total derivative of u is

∂u ∂u
du = dx + dy
∂x ∂y
Chapter 2. Literature Review 30

Dividing both sides by dt, we have

du ∂u dx ∂u dy ∂u ∂u
= + =u +v
dt ∂x dt ∂y dt ∂x ∂y

Which is the same as Eq. (2.26). Similarly, one obtains Eq. (2.27) by considering the
derivative of v = v(x, y).
Similarly, if v = v(x, y) is the velocity component in the y direction, the acceleration of
the fluid in the y direction associated with the motion of the fluid in the other directions
is given by !
Acceleration ∂v ∂v
=u +v (2.27)
in y direction ∂x ∂y

Third, Eq. (2.24) contains a term called body forces acting on the fluid. Let Fx and
Fy be the body forces acting per unit volume of the fluid in the x and y directions,
respectively (that is, ρg denotes the gravitational force acting per unit volume). Then
!
Body f orces acting
= Fx (∆x∆yl) (2.28)
in x direction

!
Body f orces acting
= Fy (∆x∆yl) (2.29)
in y direction

Fourth, Eq. (2.24) contains a term called the surface forces acting on the fluid. The
surface forces acting per unit area are called stresses. When the stress acts normal to
the surface, it is called the normal stress: when it acts along the surface, it is called the
shear stress.
Figure 2.3 shows various stresses acting on the surfaces of a differential volume element.
In this figure σx and σy denote the normal stresses in the x and y directions, respectively.
The shear stresses are denoted by τxy and τyx , where the first subscript indicates the axis
to which the surface is perpendicular and the second subscript indicates the direction of
the shear stress. Thus, τxy is the shear stress acting on the surface ∆yl (i.e., the surface
perpendicular to the x axis) at x in the direction y. Then the net normal surface force
acting on the element in the positive x direction is (∂/∂y)(σx ∆yl)∆x, and the net shear
force acting on the element in the positive x direction is (∂/∂y)(τyx ∆xl)∆y. Hence, the
net surface forces acting on the element in the positive x direction becomes
!  
N et surf ace f orces acting ∂σx ∂τyx
= + (∆x∆yl) (2.30)
in x direction ∂x ∂y
Chapter 2. Literature Review 31

Similarly, the net surface force acting in the y direction is


!  
N et surf ace f orces acting ∂σy ∂τxy
= + (∆x∆yl) (2.31)
in y direction ∂y ∂x

Figure 2.3: Nomenclature for the various stresses acting on the surfaces of the volume
element.

When Eqs. (2.25) to (2.31) are introduced into Eq. (2.24) and ∆x∆y terms are canceled,
the x momentum and the y momentum equations, respectively, become
 
∂u ∂u ∂σx ∂τyx
xM omentum : ρ u +v = Fx + + (2.32)
∂x ∂y ∂x ∂y
 
∂v ∂v ∂σy ∂τxy
yM omentum : ρ u +v = Fy + + (2.33)
∂x ∂y ∂y ∂x
The final stage in the analysis involves the determination of the expressions for vari-
ous stresses appearing in these equations. Such relations depend on the type of fluid
considered, and a discussion of this matter can be found in several references. For the
two-dimensional, incompressible, constant-property flow and the newtonian fluid con-
sidered here, various stresses are related to the velocity components by
 
∂u ∂v
τxy = τyx = µ +
∂y ∂x
∂u
σx = −P + 2µ
∂x
∂v
σy = −P + 2µ (2.34)
∂y

where P is the pressure and ıis the viscosity in the flow field.
When Eqs (2.34) are introduced into Eqs. (2.32) and (2.33), after some manipulation,
one obtains
Chapter 2. Literature Review 32

 2
∂ u ∂2u
  
∂u ∂u ∂P
xM omentum : ρ u +v = Fx − +µ + 2 (2.35)
∂x ∂y ∂x ∂x2 ∂y
 2
∂2v
  
∂v ∂v ∂P ∂ v
yM omentum : ρ u +v = Fy − +µ + (2.36)
∂x ∂y ∂y ∂x2 ∂y 2

   
∂u ∂u ∂P ∂2u ∂2u
xM omentum : ρ u ∂x +v ∂y = Fx − ∂x +µ ∂x2
+ ∂y 2
(2.37)
   
∂v ∂v ∂P ∂2v ∂2v
yM omentum : ρ u ∂x +v ∂y = Fy − ∂y +µ ∂x2
+ ∂y 2

where Fx and Fy are the body forces per unit volume acting in the x and y directions,
respectively. Equations (2.35) and (2.36) are called, respectively, the x and y momentum
equations for the steady, two-dimensional flow of an incompressible, constant-property,
newtonian fluid in the rectangular coordinate system or The Navier-Stokes equations
of motion.
The physical significance of the various terms in Eqs. (6-55) is as follows: The terms on
the left-hand side represent the inertia forces, the first term on the right-hand side is the
body force, the second term is the pressure force, and the last term in the parentheses is
the viscous forces acting on the fluid element.
If the body forces Fx and Fy are known, the continuity equation (2.23) and the two
momentum equations (2.35) and (2.36) provide three independent equations for the de-
termination of the three unknown quantities u, v, and P for the steady, two-dimensional
flow of an incompressible fluid. The analytical solution of these equations is extremely
difficult except for very simple situations.

2.3.2 Equation of Energy

The temperature distribution in the flow field is governed by the energy equation, which
can be derived by writing an energy balance according to the first law of thermodynamics
for a differential volume element in the flow field. If radiation is absent and there are
no distributed energy sources in the fluid, the energy balance on a differential volume
Chapter 2. Literature Review 33

element may be stated as


     
  Rate of energy Rate of energy Rate of
Rate of energy      
  
+ input due to   input due to   increase of 
imput due to + =
     
  
 work done by   work done by   energy in 

conduction      
body f orces surf ace f orces element
(2.38)
To derive the energy equation, each term in this expression should be evaluated. Here
we consider the energy equation in the rectangular coordinate system for steady, two-
dimensional (x, y) flow of an incompressible, constant-property, newtonian fluid. Let
∆x∆yl be the differential volume element about a point (x, y) in the flow field. Various
terms in Eq. (2.38) are evaluated now.

First, the heat addition into the element ∆x∆yl by conduction occurs in the x and y
directions. Referring to the nomenclature shown in Fig.2.4, we write
 
Rate of energy    
 
=− ∂QX ∂Qy ∂qX ∂qy
 input due to ∆x + ∆y =− + ∆x∆yl
  ∂x ∂y ∂x ∂y
conduction
∂2T ∂2T
 
=k + ∆x∆yl (2.39)
∂x2 ∂y 2

∂T ∂T
since qx = −k ∂x and qy = −k ∂y Second, if Fx and Fy are the body forces acting
per unit volume of the element and u and y are the velocity components in the x and y
directions, respectively, the energy input into the volume element ∆x∆yl resulting from
the increase in potential energy becomes
 
Rate of energy
 
 imput by  = (uFx + vFy ) ∆x∆yl (2.40)
 
bodyf orces

Third, the rate of energy input to the volume element ∆x∆yl due to surface stresses
consists of the contributions from the stresses σx , σy , τyx , and τxy . By referring to the
illustration and nomenclature in Fig. 2.5, the energy input due to the normal stress σx
is given by
  
∂ ∂
−uσx + uσx + (uσx )∆x ∆yl = ∆x∆yl (uσx )
∂x ∂x

and, due to the normal stress σy , is given by


  
∂ ∂
−vσy + vσy + (vσy )∆y ∆xl = ∆x∆yl (vσy )
∂y ∂y
Chapter 2. Literature Review 34

Figure 2.4: Nomenclature for heat addition by conduction.

Similarly, the energy input due to the stresses τyx , and τxy are given, respectively, by
 
∂ ∂
−uτyx + uτyx + (uτyx )∆y ∆xl = ∆x∆yl (uτyx )
∂y ∂y
 
∂ ∂
−vτxy + vτxy + (vτxy )∆x ∆yl = ∆x∆yl (vτxy )
∂x ∂x

Figure 2.5: Nomenclature for frictional work done by the surface forces.

The total rate of energy input into the element due to the stresses is obtained by summing
the above four quantities:
 
Rate of ener.  
 = ∂ (uσx ) + ∂ (vσy ) + ∂ (uτyx ) + ∂ (vτxy ) ∆x∆yl(2.41)
 
 input by
  ∂x ∂y ∂y ∂x
surf. stress

Fourth the energy contained in the volume element is considered to consist of the specific
Chapter 2. Literature Review 35

internal energy e per unit mass and the kinetic energy 12 (u2 + v 2 ) per unit mass of the
fluid. Then the energy content of the volume element ∆x∆yl becomes

1
ρ[e + (u2 + v 2 )]∆x∆yl
2

The rate of increase of this energy is obtained by taking its total derivative, that is,

 
Rate of increase  
 = ρ ∂e + 1 D (u2 + v 2 ) ∆x∆yl
 
 of energy of (2.42)
  ∂t 2 Dt
element

where the total derivative D/Dt for the two-dimensional, steady flow considered here is
defined as
D ∂ ∂
=u +v (2.43)
Dt ∂x ∂y
Finally, Eqs. (2.39) to (2.43) are introduced to the energy balance equation (2.38),
and the resulting expression is simplified by combining it with the momentum equa-
tions (2.35) and (2.36) and introducing the definition of various stress terms given by
Eqs. (2.34). After quite lengthy manipulations, the energy equation in the rectan-
gular coordinate system for steady, two-dimensional (x, y) flow of an incompressible,
constant-property, Newtonian fluid is determined as

   2 
∂2T
ρcp u ∂T
∂x + v ∂T
∂y = k ∂ T
∂x2
+ ∂y 2
+ µΦ

where the viscous-energy-dissipation function Φ is definedas (2.44)


  2   2
∂u 2 ∂v ∂v ∂u

Φ=2 ∂x + ∂y + ∂x + ∂y

The physical significance of various terms in Eq. (7.3) is as follows: The left hand side
represents the net energy transfer due to mass transfer; on the right hand side the terms
in parentheses represent conductive heat transfer; and the last term on the right-hand
side is the viscous-energy dissipation in the fluid due to internal fluid friction.
For most engineering applications, the flow velocities are moderate; hence the viscous-
energy dissipation term can be neglected. Then, the energy equation (7.3) simplifies
to
Chapter 2. Literature Review 36

 
1 ∂2T ∂2T
α u ∂T ∂T
∂x + v ∂y = ∂x2
+ ∂y 2
(2.45)

k
where α = ρcp . For the case of no flow (that is, u = v = 0), the energy equation
simplifies to the two-dimensional, steady-state heat conduction equation with no heat
generation.

2.3.3 Mass Conservation

The center piece of convective mass transfer analysis is the principle of mass conservation
in (or continuity through) the control volume sketched in Fig. 2.6. The net flow of
constituent i into the control volume is equal to the rate of accumulation of constituent
i inside the control volume,

 
∂ρi ∂(ρi ui )
∆x∆y = ρi ui ∆y − ρi ui + ∆x ∆y + ρi vi
∂t ∂x
 
∂(ρi vi )
− ρi vi + ∆y ∆x + m000 i ∆x∆y (2.46)
∂y

Figure 2.6: Conservation of the mass of component i during the flow of a mixture

where, ρi is the number of kilograms of constituent i per cubic meter found locally in
the mixture at point (x, y). The velocity components (ui , vi ) account for the motion of
constituent i relative to the control volume. The use of the (ui , vi ) notation at this stage
should not be taken as a suggestion that a motion with such velocity components can
actually be seen or measured. These velocity components do have a physical meaning:
Chapter 2. Literature Review 37

The group ρi ui represents the net mass flux of constituent i (measured in kg/s m2 ) in the
x direction. Concluding eq. (2.46) is the term containing m000
i , which is the volumetric
rate of constituent i generation (the units of m000 3
i are kg/s m ). This last term must be
taken into account in reactive flows that generate constituent i locally, as a product of
reaction. If constituent i is consumed by the reaction, the species generation ratem000
i is
negative. The mass conservation statement (2.46) reduces to

∂ρi ∂ ∂
+ (ρi ui ) + (ρi vi ) = m000
i (2.47)
∂t ∂x ∂y

In the absence of constituent generation (m000


i = 0) for steady flows has the same form as
the mass conservation statement for mixture flow, eq. (2.23). Superimposing the flows,
that is, summing eq. (2.47) over i letting m000
i , we obtain

∂ρ ∂ X ∂ X
+ ρi ui + ρi vi = 0 (2.48)
∂t ∂x ∂y

A term-by-term comparison of eqs. (2.47) and (2.48) brings us to a very important


concept in mass convection the concept of mass-averaged velocity components (u, v),

1X 1X
u= ρi ui , v= ρ i vi (2.49)
ρ ρ

The velocity difference (ui −u) is the diffusion velocity of constituent i in the x direction.
The product ρi (ui − u) is the flow rate of constituent i per unit area in the x direction
relative to the bulk motion of the mixture; a shorter name for this quantity is diffusion
flux, jx,i . Combining the diffusion flux definitions

jx,i = ρi (ui − u)
jy,i = ρi (vi − v) (2.50)

with the mass continuity equation for constituent i [eq. (2.47)] yields

∂ρi ∂ ∂jx,i ∂jy,i


+ (ρi v) = − − + m000
i (2.51)
∂t ∂x ∂x ∂y

Reverting now to the concentration notation [Ci = ρi ] and assuming that the mixture
flow may be treated as one with ρ = constant, the conservation of constituent i requires
that
∂Ci ∂Ci ∂Ci ∂jx,i ∂jy,i
+u +v =− − + m000
i (2.52)
∂t ∂x ∂y ∂x ∂y
Chapter 2. Literature Review 38

Making use of Fick’s law, taking the steady flow with m000
i = 0 and dropping i, the mass
conservation equation in two dimensional flow is

∂C ∂C ∂2C ∂2C
u +v = D( 2 + ) (2.53)
∂x ∂y ∂x ∂y 2

2.4 Boundary layer flow

2.4.1 Boundary Layer Theory

The boundary layer was first defined by Ludwig Prandtl in a paper presented on Au-
gust 12, 1904 at the third International Congress of Mathematicians in Heidelberg,
Germany. For convenience, Consider laminar two dimensional flow of small viscosity
(large Rynolds’s number) over a fixed semi-infinite plate. It is observed that, unlike
an ideal (no-viscous) fluid flow, the fluid does not slide over the plate, but ”stick” to
it. Since the plate is at rest, the fluid in contact with it will also be at rest. As we
move outwards along the norma, the velocity of the fluid will gradually increase and at
a distance far from the plate the full stream velocity U is attained. Strictly speaking
this is approached asymptotically. However, it will be assumed that the transition from
zero velocity at the plate to the full magnitude U takes place with in a thin layer of fluid
in contact with the plate. This is known as the boundary layer. Thus, boundary layer
is the layer of fluid in the immediate vicinity of a bounding surface where the effects
of viscosity are significant. It simplifies the equations of fluid flow by dividing the flow
field into two areas: one inside the boundary layer, dominated by viscosity and creating
the majority of drag experienced by the boundary body; and one outside the boundary
layer, where viscosity can be neglected without significant effects on the solution. This
allows a closed-form solution for the flow in both areas, a significant simplification of the
full Navier-Stokes equations. The majority of the heat transfer to and from a body
also takes place within the boundary layer, again allowing the equations to be simpli-
fied in the flow field outside the boundary layer. The pressure distribution throughout
the boundary layer in the direction normal to the surface (such as an airfoil) remains
constant throughout the boundary layer, and is the same as on the surface itself.

The velocity boundary layer thickness is normally defined as the distance from the
solid body at which the viscous flow velocity is 99% of the free stream velocity (the
surface velocity of an inviscid flow). Displacement Thickness is an alternative definition
stating that the boundary layer represents a deficit in mass flow compared to inviscid
flow with slip at the wall. It is the distance by which the wall would have to be displaced
in the inviscid case to give the same total mass flow as the viscous case. The no-slip
Chapter 2. Literature Review 39

condition requires the flow velocity at the surface of a solid object be zero and the fluid
temperature be equal to the temperature of the surface. The flow velocity will then
increase rapidly within the boundary layer, governed by the boundary layer equations,
mentioned in the next section.
The thermal boundary layer thickness is similarly the distance from the body at
which the temperature is 99% of the temperature found from an inviscid solution. The
concentration boundary layer thickness is the distance from the body at which the
concentration is 99% of the concentration found from an inviscid solution.

The ratio of the velocity and thermal boundary layer thicknesses is governed by the
Prandtl number (Pr). If the Prandtl number is 1, the two boundary layers are the
same thickness. If the Prandtl number is greater than 1, the thermal boundary layer is
thinner than the velocity boundary layer. If the Prandtl number is less than 1, which is
the case for air at standard conditions, the thermal boundary layer is thicker than the
velocity boundary layer. In heat transfer problems, the Prandtl number controls the
relative thickness of the momentum and thermal boundary layers. When Pr is small, it
means that the heat diffuses very quickly compared to the velocity (momentum). This
means that for liquid metals the thickness of the thermal boundary layer is much bigger
than the velocity boundary layer.

The mass transfer analog of the Prandtl number is the Schmidt number. Schmidt
number (Sc) is a dimensionless number defined as the ratio of momentum diffusivity
(viscosity) and mass diffusivity, and is used to characterize fluid flows in which there are
simultaneous momentum and mass diffusion convection processes. It was named after
the German engineer Ernst Heinrich Wilhelm Schmidt (1892-1975). Lewis number
(Le) is a dimensionless number defined as the ratio of thermal diffusivity to mass diffu-
sivity. It is used to characterize fluid flows where there is simultaneous heat and mass
transfer by convection. It is named after Warren K. Lewis (1882-1975).
The Prandtl number (Pr), Schmidt number and Lewis number are defined as [23, 24]

ν viscous dif f usion rate ν viscous dif f usion rate


Pr = = , Sc = = ,
α thermal dif f usion rate D mass dif f usion rate
α thermal dif f usion rate
Le = =
D mass dif f usion rate

where ν is the kinematic viscosity, α is the thermal diffusivity and D is the mass diffu-
sivity.
Chapter 2. Literature Review 40

2.4.2 Boundary Layer Equations

The deduction of the boundary layer equations was one of the most important advances
in fluid dynamics (Anderson, 2005). Using an order of magnitude analysis, the well-
known governing Navier-Stokes equations of viscous fluid flow can be greatly simplified
within the boundary layer. Notably, the characteristic of the partial differential equa-
tions (PDE) becomes parabolic, rather than the elliptical form of the full Navier-Stokes
equations. This greatly simplifies the solution of the equations. By making the boundary
layer approximation, the flow is divided into an inviscid portion (which is easy to solve
by a number of methods) and the boundary layer(See fig. 2.7), which is governed by an
easier to solve PDE. The continuity, Navier-Stokes equations, energy and concentration
equations for a two-dimensional steady incompressible flow in Cartesian coordinates are
given by (refer Eqs. (2.23), (2.37), (2.45), (2.53))

∂u ∂v
+ = 0 (2.54)
∂x ∂y
 2
∂2u
  
∂u ∂u ∂P ∂ u
ρ u +v = − +µ + (2.55)
∂x ∂y ∂x ∂x2 ∂y 2
 2
∂2v
  
∂v ∂v ∂P ∂ v
ρ u +v = − +µ + (2.56)
∂x ∂y ∂y ∂x2 ∂y 2
∂2T ∂2T
 
1 ∂T ∂T
u +v = + (2.57)
α ∂x ∂y ∂x2 ∂y 2
∂C ∂C ∂2C ∂2C
u +v = D( 2 + ) (2.58)
∂x ∂y ∂x ∂y 2

The approximation states that, for a sufficiently high Reynolds number the flow over
a surface can be divided into an outer region of inviscid flow unaffected by viscosity
(the majority of the flow), and a region close to the surface where viscosity is important
(the boundary layer). Let u and v be streamwise and transverse (wall normal) velocities
respectively inside the boundary layer. Using scale analysis(x ∼ L, y ∼ δ, u ∼ U∞
∂p
and δ  L), asymptotic analysis(v  u) and incompressibility of the fluid( ρ1 ∂y = 0) it
can be shown that the above equations of motion reduce within the boundary layer to
become

∂u ∂u ∂2u
u +v = ν (2.59)
∂x ∂y ∂y 2
∂T ∂T ∂2T
u +v = α 2 (2.60)
∂x ∂y ∂y
∂C ∂C ∂2C
u +v = D 2. (2.61)
∂x ∂y ∂y

These are called the boundary layer equations for momentum, energy and concentration,
respectively. These approximations are used in a variety of practical flow problems of
Chapter 2. Literature Review 41

Figure 2.7: Velocity and temperature boundary layers near a plate parallel to a
uniform flow.

scientific and engineering interest.

2.5 Nanofluids

2.5.1 Fundamentals of Nanofluids

Heat transfer is one of the most important processes in many industrial and consumer
products. The inherently poor thermal conductivity of conventional fluids puts a fun-
damental limit on heat transfer. Therefore, for more than a century since Maxwell
(1973), scientists and engineers have made great efforts to break this fundamental limit
by dispersing millimeter- or micrometer-sized particles in liquids. Maxwell [3] presented
a theory for effective conductivity of slurries. However, major problems such as sedi-
mentation, erosion, and high pressure drop prevented the usual microparticle slurries
to be used as heat transfer fluids. In 1993, Masuda et al. [4] worked on the thermal
conductivity and viscosity of suspensions of Al2 O3 , SiO2 , and T iO2 ultrafine particles
and published a paper written in Japanese. In addition to this, in 1993, Arnold Grimm,
an employee of R.-S. Automatis in Mannheim, Germany obtained a patent related to
improved thermal conductivity of a fluid containing dispersed solid particles [5]. He dis-
persed Al particles measuring 80nm to 1µm into a fluid. He claimed a 100% increase in
the thermal conductivity of the fluid for loadings of 0.5 to 10vol%. The serious problem
with these suspensions was rapid settling of the Al particles, presumably because in
Chapter 2. Literature Review 42

his study the particle size was large. However, Choi et al.[6] were able to make stable
nanofluids with no dispersants at all. Nanofluids were found to be very stable, devoid
of such problems, due to the small size of the particles and the small volume fraction
of the particles needed for heat transfer enhancement. Nanofluids, in which nano-sized
particles (typically less than 100nanometers) are suspended in liquids, has emerged as a
potential candidate for the design of heat transfer fluids. A study by a group at Argonne
National Laboratory in U.S, showed that these fluids enhance thermal conductivity of
the base liquid enormously, which is beyond the explanation of theories on suspensions.

Nanotechnology is the creation of functional materials, devices, and systems by control-


ling matter at the nanoscale level, and the exploitation of their novel properties and
phenomena that emerge at that scale. Cooling is indispensable for maintaining the
desired performance and reliability of a wide variety of products, such as computers,
power electronics, car engines, and high-powered lasers or x-rays. With the unprece-
dented increase in heat loads and heat fluxes caused by more power and/or smaller
feature sizes for these products, cooling is one of the top technical challenges facing
high-tech industries such as microelectronics, transportation, manufacturing, metrol-
ogy, and defense. Nanofluid technology offers a great potential for further development
of high-performance, compact, cost-effective liquid cooling systems.

There exists different ways of enhancing heat transfer such as changing flow geometry,
boundary conditions, or by enhancing thermal conductivity of the fluid. The conven-
tional way to enhance heat transfer in thermal systems is to increase the heat transfer
surface area of cooling devices and the flow velocity or to disperse solid particles in heat
transfer fluids. However a new approach to enhancing heat transfer to meet the cooling
challenge is necessary because of the increasing need for more efficient heat transfer
fluids in many industries, such as the electronics, photonics, transportation, and energy
supply industries. Nanofluids have pioneered in overcoming these problems by stably
suspending in fluids nanometer-sized particles instead of millimeter- or micrometer-sized
particles. Compared with microparticles, nanoparticles stay suspended much longer and
possess a much higher surface area. The surface/volume ratio of nanoparticles is 1000
times larger than that of microparticles. The high surface area of nanoparticles enhances
the heat conduction of nanofluids since heat transfer occurs on the surface of the parti-
cle. Researchers in nanofluids exploit the unique properties of these tiny nanoparticles
to develop stable and high-thermal-conductivity heat transfer fluids. Stable suspension
of small quantities of tiny particles makes conventional heat transfer fluids cool faster
and thermal management systems smaller and lighter.
Chapter 2. Literature Review 43

Nanofluids are a new type of heat transfer fluid engineered by uniform and stable sus-
pension of nanometer-sized particles into liquids. Most nanofluids are very dilute sus-
pensions of nanoparticles in liquids and contain a very small quantity, preferably less
than 1% by volume, of nanoparticles. The average size of nanoparticles used in nanoflu-
ids may vary from 1 to 100nm (pref erably < 10nm). Because nanoparticles are so
small, they remain in suspension almost indefinitely and dramatically reduce erosion
and clogging compared with the suspension of larger particles. Also, their larger surface
area may improve heat transfer.

Modern fabrication technology provides great opportunities to process materials actively


at nanometer scales. Fabrication of nanoparticles can be classified into two broad cat-
egories: physical processes and chemical processes (Kimoto et al. [25]; Granqvist and
Buhrman [26]; Gleiter [27]). Typical physical methods include inert-gas condensation
(IGC), developed by Granqvist and Buhrman (1976), and mechanical grinding. Chem-
ical methods include chemical vapor deposition (CVD), chemical precipitation, micro
emulsions, thermal spray, and spray pyrolysis. Nanoparticles used in nanofluids have
been made of various materials, such as oxide ceramics (Al2 O3 , CuO), nitride ceram-
ics (AlN , SiN ), carbide ceramics (SiC, T iC), metals (Cu, Ag, Au), semiconductors
(T iO2 , SiC), carbon nanotubes, and composite materials such as alloyed nanoparticles
Al70 Cu30 or nanoparticle core-polymer shell composites, and many types of liquids, such
as water, ethylene glycol, oil, polymer solutions, bio-fluids and other common fluids have
been used as host liquids in nanofluids.

Many experimental studies have shown that the thermal conductivity of nanofluids de-
pends on several parameters; nanoparticle material, particle volume fraction, spatial
distribution, particle size, particle shape, base-fluid type, temperature, and pH value.
According to Lee et al. [28] and Eastman et al. [29], metallic nanoparticles enhance the
thermal and electrical conductivity of the base fluid as well as the overall heat transfer
rate compared to non-metallic ones. They also found that nanofluids’ heat transfer rate
increases with increase in nanoparticle volume fraction. This unique property makes
nanofluids applicable as a means of heat removal in numerous thermal-fluid microsys-
tems such as micro heat pipes, microchannel heat sinks, microreactors among others. As
proposed by Buongiorno [30], the convective transport process in nanofluids is mainly
attributed to Brownian diffusion and thermophoresis, which necessitates their inclusion
in the conservation equations for mass and energy. The concept of convective trans-
port in nanofluids taking into consideration these physical mechanisms has been widely
investigated.
Chapter 2. Literature Review 44

2.5.2 Future Research on Nanofluids

Despite recent advances in the field of nanofluids, the mysteries of nanofluids are un-
solved, presenting new opportunities and challenges for thermal scientists and engineers.
Nanofluid research could lead to a major breakthrough in developing next-generation
coolants for numerous engineering applications. Better ability to manage thermal prop-
erties translates into greater energy efficiency, smaller and lighter thermal systems, lower
operating costs, and a cleaner environment.

Many industries have a strong need for improved fluids that can transfer heat more effi-
ciently. Nanofluids transfer heat more efficiently than do conventional fluids. Therefore,
when used to improve the design and performance of thermal management systems,
nanofluids offer several benefits, including improved reliability, reduction in cooling sys-
tem size, decreased pumping-power needs, increased energy and fuel efficiencies, and
lower pollution. Thus, nanofluids can have a significant impact in cooling a number of
high-heat-flux devices and systems used in consumer, industrial, and defense industries.

Future research on nanofluids can be classified in two broad categories: basic research,
and applied research including development and demonstration. The goal of future basic
research on nanofluids is to gain a fundamental understanding of the static and dynamic
mechanisms of enhanced heat transfer in nanofluids. At present, understanding the
fundamental mechanisms of the enhanced thermal conductivity of nanofluids remains
a key challenge in nanofluid research. The three main categories of new mechanisms
proposed for enhanced thermal conductivity of nanofluids are conduction, nanoscale
convection, and near-field radiation. Although these mechanisms have been proposed
the validity of most of them remains a subject of debate, and there is no agreement in
the nanofluids community about their use.

Applied researches have shown that a number of nanofluids provide extremely desir-
able thermal properties, such as higher thermal conductivity, convection heat transfer
coefficients compared to their base liquids without dispersed nanoparticles. These key
thermal features of nanofluids, together with excellent nanoparticle suspension stability,
would open the door to a wide range of engineering applications, such as engine cool-
ing and microelectronics cooling, and biomedical applications, such as cancer therapy.
Nanofluid research presents us with very promising opportunities for applications, but
there are still a number of technical issues on the road to commercialization.
Chapter 2. Literature Review 45

2.5.3 Applications of Nanofluids

The applied research in nanofluids find most of their applications in thermal management
of industries and consumer products. Some of the applications are the following

I. Cooling applications
Crystal silicon mirror cooling : Crystal Silicon Mirror Cooling One of the first
applications of research in the field of nanofluids is for developing an advanced cool-
ing technology to cool crystal silicon mirrors used in high-intensity x-ray sources
[31]. Because an x-ray beam creates tremendous heat as it bounces off a mir-
ror, cooling rates of 2000 to 3000 W/cm2 should be achievable with the advanced
cooling technology.

Electronics cooling : Chien et al. [32] were probably the first to show experi-
mentally that the thermal performance of heat pipes can be enhanced by nearly a
factor of 2 when nanofluids are used. They used water-based nanofluids containing
17 − nmgold nanoparticles as the working fluid in a disk-shaped miniature heat
pipe (DMHP). They measured the thermal resistance of the DMHP with both
nanofluids and deionized (DI) water. The results show that the thermal resistance
of a DMHP is reduced significantly (40%) when nanofluids are used instead of
DI water. Tsai et al. [33] used gold nanofluids as the working fluid for a con-
ventional meshed circular heat pipe. Kang et al. [12] measured the temperature
distribution and thermal resistance of a conventional grooved circular heat pipe
with water-based nanofluids containing a tiny amount (1to50ppm) of 35-nm silver
nanoparticles. Ma et al. [34, 35] were first to develop an ultrahigh-performance
chip cooling device called the nanofluid oscillating heat pipe (OHP). They esti-
mated that an OHP with water-based nanofluids containing Al2 O3 nanoparticles
has the ability to remove heat in excess of 1000W/cm2 , so they proposed the novel
concept of combined nanofluids and OHPs for breakthrough chip cooling. Chien
and Huang [36] numerically investigated silion microchannel heat sink (MCHS)
performance using nanofluids as the coolant. Koo and Kleinstreuer [37] simu-
lated and analyzed steady laminar flow of nanofluids in microchannels. Chein
and Chuang [38] investigated MCHS performance analytically and experimentally
using nanofluids as the coolant. Palm et al. [39] investigated the heat transfer
enhancement capabilities of nanofluids inside typical radial flow impingement jet
cooling systems. Zhou [40] investigated the heat transfer characteristics of copper
nanofluids with acoustic cavitation bubbles.

Vehicle cooling : Nanoparticles can be dispersed not only in coolants and engine
oils, but in transmission fluids, gear oils, and other fluids and lubricants. These
Chapter 2. Literature Review 46

nanofluids may provide better overall thermal management and better lubrication.
Tzeng et al. [41] were probably the first to apply nanofluid research in cooling a
real-world automatic power transmission system. They dispersed CuO and Al2 O3
nanoparticles into automatic transmission oil to investigate the optimum possible
compositions of a nanofluid for higher heat transfer performance.

Transformer cooling : The power generation industry is interested in trans-


former cooling application of nanofluids for reducing transformer size and weight.
The ever-growing demand for greater electricity production will require upgrades
of most transformers at some point in the near future at a potential cost of mil-
lions of dollars in hardware retrofits. If the heat transfer capability of existing
transformers can be increased, many of the upgrades may not be necessary. Xuan
and Li [42] and Yu et al. [43] have demonstrated that the heat transfer properties
of transformer oils can be improved by using nanoparticle additives.

Space and nuclear systems cooling : You et al. [44] and Vassallo et al. [45]
have discovered the unprecedented phenomenon that nanofluids can double or
triple the CHF in pool boiling. Kim et al. [46] found that the high surface wet-
tability caused by nanoparticle deposition can explain this remarkable thermal
properties of nanofluids. The work is important in developing realistic predictive
models of the CHF in nanofluids. The ability to greatly increase the CHF, the
upper heat flux limit in nucleate boiling systems, is of paramount practical impor-
tance to ultrahigh-heat-flux devices that use nucleate boiling, such as high-power
lasers and nuclear reactor components. Therefore, nanofluids have opened up ex-
citing possibilities for raising chip power in electronic devices or simplifying cooling
requirements for space applications. Most of all, leading nuclear researchers are
very much interested in the use of nanofluids with dramatically increased CHF val-
ues because it could enable very safe operation of commercial or military nuclear
reactors.

Defense Applications: A number of military devices and systems, such as high


powered military electronics, military vehicle components, radars, and lasers, re-
quire high-heat-flux cooling, to the level of thousands of W/cm2 . At this level,
cooling with conventional heat transfer fluids is difficult. Some specific examples
of potential military applications include power electronics and directed-energy
weapons cooling. Since directed-energy weapons involve heat sources with high
heat fluxes (> 500 to 1000W/cm2 ), cooling of the direct-energy weapon and as-
sociated power electronics is critical and is further complicated by the limited
capability of current heat transfer fluids. Nanofluids also provide advanced cool-
ing technology for military vehicles, submarines, and high-power laser diodes. It
appears that nanofluid research for defense applications considers multifunctional
Chapter 2. Literature Review 47

nanofluids with added thermal energy storage or energy harvesting through chem-
ical reactions.

II. Tribological Applications


Nanofluid technology can help develop better oils and lubricants. Recent nanofluid
activity involves the use of nanoparticles in lubricants to enhance tribological
properties of lubricants, such as load-carrying capacity and antiwear and friction-
reducing properties between moving mechanical components. In lubrication appli-
cation it has been reported that surface-modified nanoparticles stably dispersed in
mineral oils are very effective in reducing wear and enhancing load-carrying capac-
ity [47]. Li et al. [48] performed experiments on lubricant nanofluids containing
IrO2 and ZrO2 nanopartcles. The results showed that nanoparticles decrease
friction remarkably on the surface of 100C6 steel.

III. Biomedical Applications


Nanofluids was originally developed primarily for thermal management applica-
tions such as engine, microelectronics, and photonics. However, nanofluids can
be formulated for a variety of other uses for faster cooling. Nanofluids are now
being developed for medical applications, including cancer therapy. Traditional
cancer treatment methods have significant side effects. Iron-based nanoparticles
can be used as delivery vehicles for drugs or radiation without damaging nearby
healthy tissue by guiding the particles up the bloodstream to a tumor with mag-
nets. Nanofluids could also be used for safer surgery by cooling around the sur-
gical region, thereby enhancing a patient’s chance of survival and reducing the
risk of organ damage. In contrast to cooling, nanofluids could be used to pro-
duce higher temperatures around tumors, to kill cancerous cells without affecting
nearby healthy cells [49].

IV. Other Potential Applications


Other possible areas for the application of nanofluids technology include cooling a
new class of super powerful and small computers and other electronic devices for
use in military systems, airplanes, or spacecraft as well as for large-scale cooling.
In the future, nanofluids could be used to maintain a high temperature gradi-
ent in thermoelectrics that would convert waste heat to useful electrical energy.
Novel projected applications of nanofluids include sensors and diagnostics that
instantly detect chemical warfare agents in water or water- or food borne con-
tamination; biomedical applications such as cooling medical devices, detecting
unhealthy substances in the blood, cancer treatment, or drug delivery; and devel-
opment of advanced technologies such as advanced vapor compression refrigeration
Chapter 2. Literature Review 48

systems. These are just a few of the almost endless variety of nanofluids appli-
cations. Therefore, nanofluids will be increasingly important for high-value-added
niche applications as well as for high-volume applications.

2.6 Important conditions

2.6.1 Magnetohydrodynamics

Magnetohydrodynamics is the study of the macroscopic interactions of electrically-


conducting fluids with a magnetic field. Examples of such fluids include plasmas, liquid
metals, and salt water or electrolytes. The word magnetohydrodynamics (MHD) is
derived from magneto- meaning magnetic field, hydro- meaning liquid, and -dynamics
meaning movement. The fundamental concept behind MHD is that magnetic fields can
induce currents in a moving conductive fluid, which in turn creates forces on the fluid
and also changes the magnetic field itself. The set of equations which describe MHD
are a combination of the Navier-Stokes equations of fluid dynamics and Maxwell’s equa-
tions of electromagnetism. These differential equations have to be solved simultaneously,
either analytically or numerically.

The ideal MHD equations consist of the continuity equation, the Cauchy momentum
equation, Ampere’s Law neglecting displacement current, and a temperature evolution
equation. As with any fluid description to a kinetic system, a closure approximation
must be applied to highest moment of the particle distribution equation. This is often
accomplished with approximations to the heat flux through a condition of adiabaticity
or isothermality.

In the following, B is the magnetic field, E is the electric field, v is the bulk plasma
velocity, J is the current density, ρ is the mass density, p is the plasma pressure, and t
is time. The continuity equation is

∂ρ
+ ∇ · (ρv) = 0. (2.62)
∂t

The momentum equation is


 

ρ + v · ∇ v = J × B − ∇p. (2.63)
∂t

The Lorentz force term J × B can be expanded to give

B2
 
(B · ∇) B
J×B= −∇ , (2.64)
µ0 2µ0
Chapter 2. Literature Review 49

where the first term on the right hand side is the magnetic tension force and the second
term is the magnetic pressure force. The ideal Ohm’s law for a plasma is given by

E + v × B = 0. (2.65)

Faraday’s law is
∂B
= −∇ × E. (2.66)
∂t
The low-frequency Ampere’s law neglects displacement current and is given by

µ0 J = ∇ × B. (2.67)

The magnetic divergence constraint is

∇ · B = 0. (2.68)

The energy equation is given by


 
d p
= 0, (2.69)
dt ργ

where γ = 5/3 is the ratio of specific heats for an adiabatic equation of state. This energy
equation is, of course, only applicable in the absence of shocks or heat conduction as it
assumes that the entropy of a fluid element does not change. The principle of magnetic
field is applied in the medical field in the form of a device called Magnetic Resonance
Imaging (MRI). Now MRI is widely used for diagnosis of brain, vascular diseases and
all the body. It is known that several physiological fluids and also fluids in engineering
systems possess electrically conducting properties.

Many experimental and theoretical studies on conventional electrically conducting flu-


ids indicate that magnetic field markedly changes their transport and heat transfer
characteristics. Jafar et al. [50] studied the MHD Flow and Heat Transfer Over stretch-
ing/shrinking sheets with external magnetic field, viscous dissipation and Joule Effects.
Hamada and Pop [51] carried out Magnetic field effects on free convection flow of a
nanofluid past a vertical semi-infinite flat plate. Beg et al.[52] investigated the Nu-
merical study of free convection magnetohydrodynamic heat and mass transfer from a
stretching surface to a saturated porous medium with Soret and Dufour effects. Anuar
Ishak [53] studied Unsteady MHD flow and heat transfer over a stretching plate. Fang
et al. [54, 55] studied the Slip MHD viscous flow over a stretching sheet and over a
permeable shrinking sheet, respectively. Makinde [56] investigated Entropy analysis for
MHD boundary layer flow and heat transfer over a flat plate with a convective surface
boundary condition.
Chapter 2. Literature Review 50

2.6.2 Slip Boundary Condition

The description of the fluid flow in confined geometry requires specification of the bound-
ary condition for the fluid velocity at the solid wall. Usually the fluid is assumed to be
immobile at the boundary. Although this assumption is successful in describing fluid
flow on macroscopic length scales, it needs a revision for the microscopic scales due to
possible slip of the fluid relatively to the wall. The existence of liquid slip at the solid
surfaces was established in many experiments on the pressure driven flow in narrow cap-
illaries and drainage of thin liquid films in the surface force apparatus. The most popular
Navier model relates the fluid slip velocity to the interfacial shear rate by introducing
the slip length, which is assumed to be rate independent. The slip length is defined as
a distance from the boundary where the linearly extrapolated fluid velocity profile van-
ishes. Typical values of the slip length inferred from the experiments on fluids confined
between smooth hydrophobic surfaces is of the order of ten nanometers. Despite the
large amount of experimental data on the slip length, the underlying molecular mecha-
nisms leading to slip are still poorly understood because it is very difficult to resolve the
fluid velocity profile in the region near the liquid/solid interface at these length scales.

The slip flow problem of laminar boundary layer is of considerable practical interest.
Microchannels which are at the forefront of today’s turbomachinery technologies, are
widely being considered for cooling of electronic devices, micro heat exchanger systems,
etc. If the characteristic size of the flow system is small or the flow pressure is very
low, slip flow happens [57]. Wang [58] reported that the partial slip between the fluid
and the moving surface may occur in situations that the fluid is particulate such as
emulsions, suspensions, foams and polymer solutions. Wang [59], Andersson [60] and
Ariel [61] employed a partial slip boundary condition to study the flow of a pure fluid
over a stretching sheet. The effects of partial slip on the steady flow of an incompressible,
electrically conducting third grade fluid past a stretching sheet has been examined by
Sahoo and Do [62]. Hayat et al. [63] analyzed the effect of the slip boundary condition on
the magneto hydrodynamic flow and heat transfer over a stretching sheet. The influence
of partial slip, thermal radiation and temperature dependent fluid properties on the
hydro-magnetic fluid flow and heat transfer over a flat plate with heat generation has
been analyzed by Das [64].

2.6.3 Porous medium

A porous medium is a matter which contains a number pores (small holes or voids)
distributed throughout the matter. The skeletal portion of the matter is often called
the ”matrix” or ”frame”. The pores are typically filled with a fluid (liquid or gas).
Chapter 2. Literature Review 51

A porous medium is most often characterized by its porosity. Other properties of the
medium (e.g., permeability, tensile strength, electrical conductivity) can sometimes be
derived from the respective properties of its constituents (solid matrix and fluid) and
the media porosity and pores structure, but such a derivation is usually complex.

Flows through porous medium occur in filtration of fluids and seepage of water in river
beds. Movement of underground water and oils are some important examples of flows
through porous medium. An oil reservoir mostly contains of sedimentary formation
such as limestone and sandstone in which oil is entrapped. Another example of flow
through porous medium is the seepage under a dam which is very important. There are
examples of natural porous medium such as beach sand, rye bread, wood, filter, loaf of
bread, animal fur, biological tissue (bones, skin) human lung, gall bladder and bile with
stones, building materials (sand, cement, plaster board, brick), in petroleum production
engineering and in many other processes as well . The type of porous materials we can
think of is virtually limitless. Several works have been published by using the generalized
Darcy’s law.

2.6.4 Thermophoresis and Brownian motion

Thermophoresis, or thermodiffusion, is a phenomenon observed in mixtures of


mobile particles where the different particle types exhibit different responses to the force
of a temperature gradient. Thermophoresis occurs because of kinetic theory in which
high energy molecules in a warmer region of liquid impinge on the molecules with greater
momentum than molecules from a cold region. This leads to a migration of particles in
the direction opposite the temperature gradient, from warmer areas to cooler areas.

Brownian motion Brownian motion or pedesis is the random motion of particles sus-
pended in a fluid (a liquid or a gas) resulting from their collision with the quick atoms or
molecules in the gas or liquid. It is one of the key heat transfer mechanisms in nanoflu-
ids. Brownian motion only exists when the particles in the fluid are extremely small,
and as the size of the particles gets larger, Brownian motion effects diminish.

The effective thermal conductivity of nanofluids depends not only on the nanostructures
of the suspensions but also on the dynamics of nanoparticles in liquids. Nanofluids are
dynamic systems, so the motion of nanoparticles and the interactions between dancing
nanoparticles or between dancing nanoparticles and liquid molecules should be consid-
ered to develop more realistic models. Interestingly, the Brownian motion of nanoparti-
cles was considered as a most probable mechanism. The studies of Wang et al. [65] and
Keblinski et al. [66] clearly showed that Brownian motion is not a significant contributor
to heat conduction based on the results of a time-scale study. However, it is important
Chapter 2. Literature Review 52

to understand that the heat transfer mechanism that Wang et al. [65] and Keblinski
et al. [66] explored is heat conduction through particle-particle collisions caused by the
Brownian motion of nanoparticles. Despite the work of Wang et al. [65] and Keblinski et
al. [66], a few investigators did not drop the idea that Brownian motion of nanoparticles
is a most probable mechanism. In fact, one of the key concepts used in most dynamic
models is that nanoparticle motion is essential to enhanced energy transport in nanoflu-
ids. This is to address one of the most important thermal phenomena in nanofluids: the
strongly temperature-dependent thermal conductivity of nanofluids.

2.7 Review of previous study

Although increases in effective thermal conductivity are important in improving the heat
transfer behavior of fluids, a number of other variables also play key roles. For example,
the heat transfer coefficient for forced convection in tubes depends on many physical
quantities related to the fluid or the geometry of the system through which the fluid
is flowing. These quantities include intrinsic properties of the fluid such as its thermal
conductivity, specific heat, density, and viscosity, along with extrinsic system parameters
such as tube diameter and length and average fluid velocity. Therefore, it is essential
to measure the heat transfer performance of nanofluids directly under flow conditions.
Experimentalists have shown that nanofluids have not only better heat conductivity but
also greater convective heat transfer capability than that of base fluids. Experiments
show unexpectedly that the heat transfer coefficients of nanofluids are much better than
expected from enhanced thermal conductivity alone in both laminar and turbulent flow.
However, for natural convection, nanofluids have lower heat transfer than that of base
fluids. Putra et al. [67] were first to study natural convection in nanofluids. Using
water with 130 − nm Al2 O3 and 90 − nm CuO particles, they showed that the natural
convective heat transfer is lower in nanofluids than in pure water and that this decrease
in natural convection heat transfer coefficient increases with particle concentration.

Experimental investigations have demonstrated a remarkable heat transfer enhancement


when using nanofluids in forced convection: a 40% increase in turbulent convection
heat transfer with the addition of 2.0vol% of Cu nanoparticles in water and roughly a
twofold increase in laminar convection heat transfer by the addition of 1.1vol% CNTs
in water (Xuan and Li [68], Faulkner et al.[69]). The enhancement of heat transfer
coefficient measured is much higher than that of predictions based on enhanced effective
thermal conductivity of nanofluids alone. Such dramatic enhancement of convective
heat transfer has inspired several investigators to propose new mechanisms of enhanced
convection heat transfer coefficient under both laminar and turbulent flow. In the flow
Chapter 2. Literature Review 53

of a nanofluid, thermal dispersion, particle migration, and Brownian diffusion may be


some mechanisms of enhanced convection in nanofluids. Xuan and Roetzel [70] were
first to employ the concept of thermal dispersion for modeling enhanced convection in
nanofluids. This concept adds a fictitious conductivity called the thermal dispersion
coefficient to the effective thermal conductivity of nanofluids by assuming that there is
velocity slip between nanoparticle and liquid and that the nanoparticles induce a velocity
and temperature perturbation.

Buongiorno [71] considered seven slip mechanisms that can produce a relative velocity
between the nanoparticles and the base fluid: inertia, Brownian diffusion, thermophore-
sis, diffusiophoresis, Magnus effect, fluid drainage, and gravity. Of all of these mecha-
nisms, only Brownian diffusion and thermophoresis were found to be important. Buon-
giorno’s analysis consisted of a two-component equilibrium model for mass, momentum,
and heat transport in nanofluids. He found that a non-dimensional analysis of the equa-
tions implied that energy transfer by nanoparticle dispersion is negligible, and cannot
explain the abnormal heat transfer coefficient increase. Buongiorno suggested that the
boundary layer has different properties because of the effect of temperature and ther-
mophoresis. The viscosity may be decreasing in the boundary layer, which would lead
to heat transfer enhancement. Later, Buongiorno [71] developed an analytical model for
convective transport in nanofluids in which Brownian motion and thermophoresis effect
were considered.

Sakiadis [72] studied the boundary layer behavior for the sheet moving with a constant
velocity in a viscous fluid. The analytical solution for steady stretching of the surface was
given by Crane [2]. After this pioneering work, various aspects of the problem including
magneto hydrodynamic flows [73], non-Newtonian fluids [74], nanofluids [75–77], flows
with chemical reactions [78], and also different hydrodynamic and thermal boundary
conditions have been investigated.

Some of the researchers considered different idealized thermal boundary conditions for
the sheet surface. Gupta and Gupta [79] analyzed heat and mass transfer over a stretch-
ing sheet with constant surface temperature. Different thermal boundary conditions
such as power-law surface temperature and power-law surface heat flux were discussed
by Fang [80]. Cortell [81] investigated the viscous flow and heat transfer over a stretching
sheet prescribed power law surface temperature. However, when the sheet is prescribed
to a convective fluid from below, the consideration of constant or variable tempera-
ture/heat flux is not a realistic boundary condition in many engineering applications
of stretching sheet. In this case, the convective boundary condition is a more realistic
thermal boundary condition. Recently, number of researchers examined the convective
boundary condition. Aziz [82] considered the classical problem of hydrodynamic and
Chapter 2. Literature Review 54

thermal boundary layers over a flat plate in a uniform stream of fluid. Aziz obtained a
similarity solution for laminar thermal boundary layer over a flat plate with a convective
surface boundary condition. Later, Magyari [83] introduced an analytical approach for
heat equation which has been implemented in the work of Aziz [82]. Hamad et al. [84]
investigated the heat and mass transfer of boundary layer stagnation-point flow over a
stretching sheet in a porous medium saturated by a nanofluid using a lie group analy-
sis. Ishak [85] obtained a similarity solution of flow and heat transfer over a permeable
surface with a convective boundary condition.

The forced convection of a uniform stream flow over a flat surface with a convective
surface boundary condition has been theoretically analyzed by Merkin and Pop [86].
Yao et al. [87] studied heat transfer in the stretching sheet problem with convective
boundary condition, and they obtained an exact solution in the form of an incomplete
Gamma function. They reported that the convective boundary condition results in
temperature slip at the wall, which it is greatly affected by the Prandtl number and the
wall stretching parameters. They found that the temperature profiles are quite different
from the prescribed wall temperature cases.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy