You are on page 1of 115

Contents

Heat transfer: Introduction. 4

Chapter 1. Structure of materials. Heat transfer in solids. 5


Solid state, material structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Fourier’s law, thermal conductivity, differential equation . . . . . . . . . . . . . . . . 5
Differential HT equation. Initial and boundary conditions. . . . . . . . . . . . . . . . 7
Differential HT equations in other coordinate systems . . . . . . . . . . . . . . . . . . 10
Solid object and an infinite fluid medium . . . . . . . . . . . . . . . . . . . . . . . . . 12
Thermal resistance: analogy heat transfer - electricity. . . . . . . . . . . . . . . . . . 14
Characteristic values in non-steady heat conductivity . . . . . . . . . . . . . . . . . . 15

Chapter 2. Heat transfer in fluids. 17


Conduction, convection, advection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Differential equation for convective heat tranfer . . . . . . . . . . . . . . . . . . . . . 18
Differential equation for fluid flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Boundary conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Example: Couette Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
External flow. Boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
External flow. Flat plane in parallel flow . . . . . . . . . . . . . . . . . . . . . . . . . 26
External flow: some additional remarks . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Free convection. Vertical surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
Free convection: Vertical surface - nondimensional analysis . . . . . . . . . . . . . . . 34
Free convection: Vertical surface - similarity solution . . . . . . . . . . . . . . . . . . 35

Chapter 3. Radiation. 38
Types of electromagnetic waves. And what is thermal radiation . . . . . . . . . . . . 38
Specific features of radiative energy transfer . . . . . . . . . . . . . . . . . . . . . . . 41
Emissive power. Solid angle. Intensity . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Diffuse surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

Chapter 4.Blackbody. 46
Blackbody . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Planck’s distribution. Rayleigh-Jeans and Wein’s approximations . . . . . . . . . . . 47
Total emisive power of a blackbody. Fractions . . . . . . . . . . . . . . . . . . . . . . 51

Chapter 5. Opaque materials. Emissivity of real surfaces 54


Spectral directional emissivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Total directional emissivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Spectral hemispherical emissivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

1
Total hemispherical emissivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Solutions to the examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

Chapter 6. Opaque materials. Absorptivity of real surfaces 65


Incident radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
Spectral directional absorptivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Kirchhoff’s Law for directional spectral properties . . . . . . . . . . . . . . . . . . . . 67
Total directional absorptivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
Kirchhoff’s Law for directional total properties. Gray surfaces . . . . . . . . . . . . . 68
Spectral hemispherical absorptivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Kirchhoff’s Law for hemispherical spectral properties. Diffuse spectral surfaces . . . . 70
Total hemispherical absorptivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
Kirchhoff’s Law for hemispherical total properties. . . . . . . . . . . . . . . . . . . . . 72

Chapter 7. Opaque materials: reflectivity. Remarks on transparent materials:


transmissivity 73
Spectral and total bidirectional reflectivity . . . . . . . . . . . . . . . . . . . . . . . . 73
Directional reflectivities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Hemispherical reflectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Radiosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Special remark 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Spectral and total transmissivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Special remark 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

Chapter 8. Radiative heat exchange between black isothermal surfaces 82


Differential view factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
View factors between a differential surface and a finite one . . . . . . . . . . . . . . . 83
View factors for finite surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Algebra of view factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

Chapter 9. Radiative heat exchange in a black enclosure 86

Chapter 10. RHT in an Enclosure Composed of Diffuse-Gray Surfaces 88


Resulting heat fluxes and temperatures in a diffuse-gray enclosure . . . . . . . . . . . 90
Graphical representation of RHT in a diffuse-gray enclosure . . . . . . . . . . . . . . 91
Continuation of graphical representation of RHT in a diffuse-gray enclosure . . . . . . 93
Network representation of radiation shields . . . . . . . . . . . . . . . . . . . . . . . . 96

Chapter 11. Solar radiation. Environmental radiation. 100


If Sun is a black body? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Scattering in the atmosphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

2
Emission from the Earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

Chapter 12. Heat exchange with absorption and emission in the gas volume 106
Absorptivity and emissivity in a volume . . . . . . . . . . . . . . . . . . . . . . . . . 106
Kirchhoff’s law for the absorption and emission in the gas. Emissivity of a gas . . . . 107
Average path length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Gray gas model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Nomograms and empirical formulae . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Radiating gas in a mixture with non-radiating gases . . . . . . . . . . . . . . . . 112
Mixture of radiating gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Radiating gas near a surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

3
Heat transfer: Introduction.
The subject of the course is a heat transfer. But what is the heat transfer?

Definition 1. Heat transfer is a process of the transmitting (transfer) the thermal energy in
space.

At least 3 questions can be asked to this definition:


q.1. What is thermal energy?
q.2. Why (where) is it transmitting?
q.3. How can it be transmitted?

The answers are the following:


a.1. Particles of an object (or of a media) posses kinetic energy. Temperature of the object
represents the kinetic energy of particles comprising the body.
a.2. The energy is transmitting due to interactions of the particles – (*it should be clarified
what ”particle” means).
a.3. The modes (types) of heat transfer in a media depend on the physical properties and the
state of the media.

There are three modes of heat transfer:


– conduction, which exists in a media (does not exist in a vacuum), both solid and fluid (liquid
and gases). Conduction is provided by interaction of moving microparticles which comprise a
solid object or a fluid.
– convection, which exists in moving fluids (liquid and gases). Here the energy is transferred
due to interactions of agglomerates (microvolumes) of atoms and/or molecules which can move
for some distance.
– radiation, the heat is transferred by waves (or by photons). This mode of the heat transfer
differs significantly from the previous two since it does not require any media between the ob-
jects which exchange with their energies.

In first lectures we will recall fundamentals of conduction and thermal convection. The most
part of the course is devoted to the radiative heat transfer.

4
Chapter 1. Structure of materials. Heat transfer in solids.
To recall fundamentals of the heat transfer we should refer to the materials structure.

Solid state, material structure


Every material consists of atoms and every atom has a nucleus and electrons which ”moves
somehow around” the nucleus. The structure of atoms is given in a Periodic table (Mendeleev
table). Let us consider a silicon, Si. In its nucleus it has 14 protons and 14 neutrons, there are
14 electrons around its nucleus, as shown in fig.1.

Figure 1: Structure of Si

The radius of a stable nucleus can be estimated as R = r0 A1/3 , where r0 = 1.25f m = 1.25 ·
10−15 m) and A = Z + N , Z and N are the number of protons and neutrons, respectively.
Then, the radius of Si nucleus is about Rn (Si) = 1.25f m · 281/3 ≈ 3.8f m = 3.8 · 10−15 m.
The radius of the Si atom can be estimated using the data given in a periodic table: atomic
volume is about 2.05cm3 /mol, a mole contains 6.02 · 1023 atoms, so the radius of an atom is
Ra (Si) = (2.05/6.02 · 1023 )1/3 ≈ 1.54 · 10−8 cm = 1.54 · 10−10 m. The size of the nucleus is
negligibly small compared to that the atom, i.e. the atom is mostly ”empty”.
Electrons ”move near their own” atom in certain areas which are called ”orbitals”. In metals
and also in semiconductors electrons from external orbitals can leave their own atom and
move from one atom to another.

Fourier’s law, thermal conductivity, differential equation


Inside metals the energy is transferred from point to point by mechanical interaction of particles
(mostly electrons) and this mode of energy transfer is named conduction. The heat flux
characterises how fast the energy is transferred in space.

Definition 2. The density of the heat flux ~q is the energy transfered through an isothermal surface
per unit of time per unit of the surface.

This value is a vector which is perpendicular to the isothermal surface:

~q = e~x · qx + e~y · qy + e~z · qz (1)


W J
[~q] = 2 =
m s · m2

5
An empirical Fourier law relates the temperature distribution and the heat flux.
Fourier’s Law: The density of a conductive heat flux is proportional to the negative tem-
perature gradient:
~q = −κ(x, y, z)∇T (2)

Here −∇T = −grad(T ) that in cartesian system can be written as

∂T ∂T ∂T
−∇T = −e~x · − e~y · − e~z ·
∂x ∂y ∂z

The κ(x, y, z) in eq.2 presents the thermal conductivity which is generally a tensor but in
many practical cases it is considered as a scalar, which however may depend on temperature.
W J
Dimension of the thermal conductivity is [κ] = = .
m·K s·m·K
In the fig.2 the values of thermal conductivity for some materials are given and the fig.3 provides
a qualitative idea about thermal conductivity of various media.

Figure 2: Values of the thermal conductivity for some materials.

Figure 3: Qualitative relation of thermal conductivity for various materials at normal temper-
ature and pressure

6
Note:
1. In metals and semiconductors the energy transfer is occurred due to the motion of the
electrons from the highest orbitals which move freely from atom to atom.
2. In solid dielectric materials the energy transfer is occurred via oscillation of the crystalline
cells of materials (see fig.4) which is known as ”phonon conductivity”.

a b

Figure 4: Crystal structure. (a) Unit (elementary) body center cubic (bcc) cell. (b) Crystalline
cell.

3. For the solids the thermal conductivity of metals (Ag, Cu, Pb, Al, Steel, etc) is higher than
the sonductivity of insulator. The exception is the diamond, which is one of crystallographic
modification of the carbon.
4. In gases and liquids the particles which provide the heat transfer are molecules.
5. The Fourier’s law is remarkable since it is valid for all solids, whether they are metals or
dielectrics, and also for liquids and gases.

Differential HT equation. Initial and boundary conditions.


Fourier’s law is not enough for the solution of the thermal problems. An equation which would
describe the conservation of the total energy is required. Derivation of the equation for
heat transfer is based on the First law of thermodynamics.

First law of thermodynamics: A change in the internal energy of a system dU is equal to


the difference between the amount of heat supplied to the system dQ, and the amount of work
performed by the system on its surroundings dW .

dU = dQ − dW (3)

The thermodynamics states that the variation of the internal energy during the period of time
dt of a material with a constant mass m and a constant volume dv is related to the variation
of the temperature T in this volume:
∂T
dU = mCp dT = ρ · dv · Cp dt (4)
∂t

7
where Cp is a specific heat (J/(kg · K)), ρ is a density (kg/m3 ), dv is the volume of the system,
(m3 ).
The work of the system (or over the system) with a constant volume during the period of time
dt related to the internal heat sources with a volume density qv0 :

dW = qv0 · dt · dv (5)

The amount of heat supplied to the system dQ during the time dt is the thermal heat flux ~q
~ which surrounds the volume dv, i.e.
through the surface ds
~ = −(∇
dQ = −~q · dsdt ~ · ~q)dv · dt (6)

Bringing together Eqs.4, 5 and 6, we obtain:


∂T ~ · ~q)dv · dt + q 0 · dv · dt
ρCp dvdt = −(∇ v
∂t
Re-writing the previous expression, we obtain a ”classical” form of the equation for the con-
ductive energy transfer:
∂T 
~ κ∇T~

ρCp =∇ + qv0 (7)
∂t
If the physical properties of a material (ρ, Cp , κ) are constant, Eq.7 can be re-written in a form

∂T
= a∆T + qv00 (8)
∂t
Where a is the thermal diffusivity and laplasian in cartesian system is written as
 2
∆= ∇ ~

∂2 ∂2 ∂2
∆= + +
∂x2 ∂y 2 ∂z 2
The eq.7 allows us to solve the problem of heat transfer if we know material properties (Cp , ρ, κ),
distribution of internal heat sources qv0 (x, y, z), and also if we have boundary conditions and,
if the problem is non-stationary, initial conditions are given.
As an illustration, imagine that we have to calculate the temperature distribution in a block
and let us supposed that a block is attached to a surface and is surrounded by the air as shown
in fig.5.

Figure 5: Illustration to the boundary condition

For the surface 6 where the block is attached to the motherboard we can use boundary condi-
tions of the first or of the second type.

8
1. Dirichlet or first-type boundary condition:
this condition specifies the temperature at the boundary of the object (i.e. the temperature of
the motherboard is known for every time):

T (S, t) = TS (t)

2. Neumann or second-type boundary condition:


this conditions specifies a density of the heat flux at the boundary of the object for every
moment of time:
∂T
−κ = ~q(x, y, z)S · ~n
∂n S
In the example with a processor this condition can be used, for example, if the surface is insu-
lated (which is not generally the case) or if the surface is cooled with a known heat flux.

Let us consider surfaces from 1 to 5. If temperature calculations does not include the temper-
ature distribution in the air, the appropriate boundary conditions are
3. Third-type boundary conditions(Robin boundary conditions):
A widely used form for this type of the boundary condition is

∂T
−κ = h · (T − Text )
∂n S
where h(W/(m2 · K) is a coefficient of the convective heat transfer and Text is a temperature of
the air at a long distance from the surface.
In fact, the coefficient of the convective heat transfer includes all the uncertainties related to
the convective heat transfer in the air. Convection is considered in the next chapter.
Note that a more general form for the third-type boundary conditions is a combination of the
first and second conditions:
 
∂T
A·T −B·κ = FS (x, y, z)
∂n S
and this formulation is better for the example under consideration because the block loses the
energy via the surfaces 1 − 5 also by radiation.

How to solve Eq.7 for a domain under consideration ?


1. Analytically
If the geometry, boundary and initial conditions and distribution of internal heat sources are
not complicated, it is possible to do analytically, using methods of Mathematical Physics.
Good numbers of examples are given in the book by H.S. Carslaw and J.C.Jaeger, ”Conduction
of heat in solids” (several editions, the 2nd was in 1959). In this case the solution generally has
a form of a series.
2. Numerically
Practically for any configuration, for any conditions the Eq.7 can be solved numerically.

9
Example
Let a temperature distribution in a wall of a thickness of 1m and the area of 10m2 at a certain
time instant is given as
T (x) = a + bx + cx2 (9)

with T (x) given in Celsius degree and x is in meters, with a = 9000 C, b = −3000 C/m,
c = −500 C/m2 . A uniform heat generation q 0 = 1000W/m3 exists inside the wall whose
properties are ρ = 1600kg/m3 , κ = 40W/(m · K), and Cp = 4kJ/(kg · K)
Questions:
1. Make a scheme (drawing) for the given problem and provide its mathematical description
2. Determine:
2.1 Rate of the heat entering the wall at x = 0 and leaving the volume at x = L, L = 1m
2.2 Rate at which the energy is stored in the volume
2.3 Rate at which temperature changes at x = 0, x = 0.25 and x = 0.5m

Differential HT equations in other coordinate systems


Another coordinate system, non-Cartesian, can be chosen for the solution of the problems be-
cause of some reasons. The most important is that the geometrical description of the object
under consideration is easier in another system. Example: spherical shell is easier to present
in spherical coordinate system, an ellipsoidal particle in an elliptic system etc.

Yet, the presentation of differential operator ∇ is different in these cases and it accounts for
the variation of the curvature of the spatial coordinates. Also, a differential volume in each
case will be presented in a different manner as shown in figures and equations below.

Geometry, nabla operator and Heat transfer equation in cylindrical coordinate


system

Figure 6: Cylindrical coordinate system

10
Nabla operator in cylindrical coordinate systems:

~ r,ϕ,z = ~er ∂ + ~eϕ 1 ∂ + ~ez ∂


∇ (10)
∂r r ∂ϕ ∂z
where ~er , ~eϕ and ~ez are unit vectors along the coordinate axes r, ϕ and z, respectively (fig.6).

~ 2r,ϕ,z = ~er ∂ + ~eϕ 1 ∂ + ~ez ∂


∇ (11)
∂r r ∂ϕ ∂z
Elementary volume in cylindrical coordinate systems (fig.6).

dvr,ϕ,z = rdϕ dz dr (12)

Accounting for this, the Differential Heat Transfer Equation is presented in the cylindrical
coordinate system as follows:
     
∂T 1 ∂ ∂T 1 ∂ ∂T ∂ ∂T
Cp ρ = κr + 2 κ + κ + q0 (13)
∂t r ∂r ∂r r ∂ϕ ∂ϕ ∂z ∂z

Geometry, nabla operator and Heat transfer equation in spherical coordinate sys-
tem

Figure 7: Cylindrical coordinate system

Nabla operator in the spherical coordinate systems

~ r,ϕ,θ = ~er ∂ + ~eϕ 1




+ ~eθ
1 ∂
(14)
∂r r sin θ ∂ϕ r ∂θ
where ~er , ~eϕ and ~eθ are unit vectors along the coordinate axes r, ϕ and θ, respectively
Elementary volume in the spherical coordinate systems (fig.7)

dvr,ϕ,θ = r sin θ dϕ rdθ dr (15)

Accounting for this, the Differential Heat Transfer Equation is presented in the spherical
coordinate system as follows:
     
∂T 1 ∂ 2 ∂T 1 ∂ ∂T 1 ∂ ∂T
Cp ρ = 2 κr + 2 2 κ + 2 κ sin θ + q0 (16)
∂t r ∂r ∂r r sin θ ∂ϕ ∂ϕ r sin θ ∂θ ∂θ

11
Boundary condition for a solid object in an infinite fluid medium
Before going into details of convective heat transfer let us consider a solid object inside a liquid
whose velocity and temperature are fixed at infinity (fig.8)

Figure 8: Flow around a solid object

The questions are:

1. if in this situation the boundary conditions for the differential heat transfer equation inside
the object can be written as a constant temperature, equal to that of the liquid?
2. If not, what is a condition to use?

The answer to the first question is ”no”, because it is not possible to assure that the
layers of the fluid adjacent to the object have the temperature T∞ , they will be heated or cooled
depending on the temperature of the object, an example of the temperature distribution in a
wall and inside the liquid surrounding it is given in a figure above, fig.8. Suppose that the given
distribution is steady and is conditioned by internal heat sources in the solid object.
Note that despite the convective flow in the liquid (that will be discussed later) the boundary
condition for the heat flux in the liquid are still the same because of viscosity in the liquid
which creates a thin layer near the object where the liquid does not move:

qs,R = qf,R
qs,L = qf,L
dT
qs,R = −κs |s,R thermal flux in the solid at the right boundary
dx
dT
qs,L = −κs |s,L thermal flux in the solid at the left boundary
dx
dT
qf,R = −κf |f,R thermal flux in the liquid at the right boundary
dx
dT
qf,L = −κf |f,L thermal flux in the liquid at the left boundary
dx
(17)

Here the indices s and f are for the solid and fluid, respectively and L,R ar for the right and
the left boundaries.

12
Note that for both sides is is assumed that the direction of the vector ~q is the same as
direction of the x − axis.

Figure 9: Flow around a solid object

According to the presented illustration, the temperature gradient inside the liquid near the
wall near the the right boundary can be approximated as:
dT T∞ − TS,R
qf,R = −κf |f,R = −κf (18)
dx δR
where δR > 0 is a distance at which the tangent line to the temperature curve in the liquid
Tf (y) constructed at the left boundary of the object cross the level T∞ .
The value of δR is not known a priori (without rigorous calculation).
This condition can be rewritten as:

qf,R = h(TS,R − T∞ ) (19)


κL
hR = (20)
δR
(21)

The coefficient hL is termed ”convective heat exchange coefficient.” Note that according to the
choice of the axis, the q is directed out from the object. The same condition can be written on
the left boundary:
dT TS,R − T∞
qf,L = −κf |L,R = −κf
dx δL
qf,L = hR (T∞ − TS,L )
κf
hR =
δL
(22)

13
In this case the qf,L is directed toward the object, consequently, the sign in the temperature
difference has changed.
Attention! In all (to my knowledge) numerical softwares the following agree-
ment exists: the flux is considered to be positive if its direction is the same as
the direction of the normal vector defining a surface.
In the condition written as eq.19, the coefficient h contains uncertainties related to the definition
of the distance δ which, in particular, depends on the velocity of the forced flow. Indeed, if the
velocity is higher, then the fluid with the temperature of the flow T∞ arrives faster, therefore,
the layer of the heated liquid will be thinner fig.10.

Figure 10: Flow around a solid object

Note that in the case of the object of a complex shape, the distance δ can vary along the surface
of the object although the imposed flow is fixed. In this case a local coefficient h(~r) and an
average coefficient h can be introduced.

Thermal resistance: analogy heat transfer - electricity.


From previous consideration, it is seen that generally the heat flux in a steady case of conduc-
tiviy between two surfaces can be presented in a form:
κ
q= (Ts,1 − Ts,2 ) (23)
L
and in a case of a surface with convective heat exchange:

q = h(Ts,1 − T∞ ) (24)

An analogy between the heat transfer and the electricity can be made, assuming that q
corresponds ot the electrical current I, the temperature difference to the potential and, conse-

14
quently, having as a resistancesuch that I = U/R, i.e. q = ∆T /RT

1 1
RT = and RT =
κ/L h
This concept becomes really interesting in consideration of more complex objects.
Example
A composite wall shown in fig.11 is subjected to a flow of the hot fluid with T∞,1 on its left side
and to the cold fluid with T∞,2 on the right side. The thermal conductivity and thicknesses of
all layers are known, the two convective heat exchange coefficients are given. Find a heat flux
in the system.
Solution is given in the class

Figure 11: Thermal resistance in a composite wall

Note: some systems can be presented with parallel thermal resistance, similar to the elec-
trical circuits.

Characteristic values in non-steady heat conductivity


Let us remind a problem considered on the exercises: a cube 0.5cm × 0.5cm × 0.5cm with
the initial temperature T0 = 350K is placed inside a liquid having a temperature Tf = 273K.
The convective heat exchange coefficient between the cube and the liquid in this case is h =
1W/(m2 · K). Suppose that the temperature inside the cube is uniform during its cooling. Find
the temperature of the cube at t = 5s after beginning of the cooling.

15
The mathematical statement for the described problem is the following:
∂T ~ ∇T ~ )
ρCp = ∇(κ
∂t
T |t=0 = T0
dT
−κ |S = h(T − Tf )
dx
(25)

If in the last equation related to the boundary condition a derivative is replaced with its linear
form ∆T /δ where ∆T is a temperature drop insider the object over its characteristic size δ,
then:

∆T = − (T − Tf )
κ
the non-dimensional complex on the right side of the equation is termed Biot
number:
hδ conduction thermal resistance
Bi = =
κ convection thermal resistance
The solution of the given problem is presented in a form

T = Tf + (T0 − Tf )e−hS/(V ρCp )t = Tf + (T0 − Tf )e−t/τ

The complex in the exponential term shows how fast the temperature of the object will change
in time, it is a time constant which can be seen as follows:
V ρCp 1
τ= = V ρCp = (thermal capacitance) · (convective thermal resistance)
hS h

Diffusion time and diffusion length

Introducing some characteristic time related to the system τ and a characteristic distance L, it
is possible to write the conductive heat transfer equation for a media with constant properties
in the form
∆T ∆T
∼a 2 here ∆ is a simple difference!
τ L
Then, the following values can be derived:
1. Diffusion time τd = L2 /a that can be interpreted as:
· time which is required for that a thermal perturbation propagated at the distance L
· time which is required for a thermal perturbation of a size L to be smoothed


2. Diffusion length Ld = aτ that can be interpreted as:
· a distance at which a thermal perturbation will propagate during the time tau
· a size of a thermal perturbation which will be smoothed during the time tau

16
Chapter 2. Heat transfer in fluids.
To remind:

ˆ Solids → crystalline structure.

ˆ The heat is transferred via interaction of individual particles → motion of electrons (in
conductors) AND/OR oscillation of atoms/molecules (in dielectrics)

ˆ This process is called (thermal) conduction → Fourier’s law → thermal conductivity

ˆ Differential conductivity equation and boundary conditions

Conduction, convection, advection.


With increase of the temperature of a solid body, oscillations of the particles in crystalline cell
increase. At some temperature the crystalline structure starts to destroy and the solid phase
tranforms into liquid phase. It happens at the melting temperature of the material. For the
silicon the melting temperature is about 1414◦ C. With further heating the liquid transforms
into gas at the boiling point. For the silicon the boiling temperature is about 3265◦ C

In the liquid state the atoms and molecules do not construct a cell and they can move for
long distances. However, in the liquid state they are still close and groups of them (agglom-
erate) move in the same direction. They say that in a liquid a short-range order is preserved
(Fig.12b). With further increase of the temperature this short-range order destroyes as well.
Distances between the molecules increase even more and transition from the liquid to the gas
occurs (Fig.12c).

Figure 12: Transition (a)solid -(b)liquid -(c)gas.

What kinds of interactions between the particles exist in fluids? Firstly, in fluids, the micro-
particles (atoms, molecules and electron in the liquid metals) interact ”individually”. Thus,
the conductive mode of heat transfer persists in fluids (it always exists). Secondly, the

17
microvolumes (agglomeration of molecules) of the fluids can move and transfer the energy, see
fig.13.

Figure 13: Interaction of macrovolumes and molecules in fluids.

Definition 3. Convective heat transfer (often referred to as ”convection”), is transfer of the


energy in space by the movement of microvolumes of fluids.

Note, that sometimes the term ”advection” is used in relation to the heat transfer in fluid.
The difference between ”advection” and ”convection” is that
– advection is the transfer of heat in space only by the movement of fluids
– convection unifies advection and conduction.

Differential equation for convective heat tranfer


The first law of thermodynamics (Eq.3) is still valid, whether the fluid or solid is considered.
Equations for 4 and 5 are true as well. Expression for dQ in a form of density heat fluxes is
still OK but the energy may be transferred both by conduction and convection:

~q = ~qcond + ~qconv

Conductive flow is defined by Fourier’s law given by Eq.2 and convective flow is

qconv = V~ · (ρCp T ) (26)

Combining equations 3, 4, 5, 2 and 26, and assuming that physical properties of the material
ρ, Cp , κ are constant, we obtain:

∂T
ρCp + ρCp V~ · ∇T
~ = κ∆T + qv0 (27)
∂t
Note that in some cases (see an example below) a heat can be generated because of the
”friction” inside the liquid, so-called ”viscous dissipation”. This term can be added to the right
part of te energy equation as will be considered below.

18
This differential equation (eq.27) contains velocity of the motion of the liquid. Therefore,
to solve a general problem of convective heat transfer another equation for the liquid flow is
required. For derivation of the differential equation for fluid flow (the momentum equation)
the second Newton’s law is applied to the elementary volume of the liquid. The derivation
of this equation is beyond the scope of our course. Here we will present basic ideas for its
derivavtion and we will use the equation obtained in a course of fluid mechanics.

Differential equation for fluid flow – momentum equation


Newton’s second law of motion The acceleration a of a body is parallel and directly pro-
portional to the net force F~ = and inversely proportional to the mass m, i.e., F~ = m~a.

Forces acting on the element of fluid can be divided into bulk forces and surface forces.

ˆ Bulk forces are forces which act on all the particles of the volume. Examples are: gravity
force, centrifugal force, electromagnetic forces. Below we will take into account only the
gravity force.

ˆ Surface forces act from the fluid which surrounds the elementary volume and are imposed
to the boundary surfaces of the volume. The pressure and the shear stress are surface
forces

Therefore, three forces act on the elementary volume of the fluid: (i) the gravity force, (ii) the
resultant force of pressure and (iii) the resultant force of the shear stress. This is expressed via
the following equation:

∂ V~  
ρ + ρ V~ , ∇ V~ = ρ~g − ∇P + µ∇2 V~ (28)
∂t
here P is the pressure, µ is a dynamic viscosity. Note that

∇2 V~ = ∇(∇ · V~ ) − ∇ × (∇ × V~ ) (29)

and in Cartesian coordinates:

∇2 V~ = (∇2 Vx , ∇2 Vy , ∇2 Vz ) (30)

In eq.28 the right part is the sum of forces acting on a differential volume, whereas the left
part is the result of their action which leads to the acceleration of the volume of fluid and to
its deformation. Note that eq.28 can be decomposed into 3 equations, each correspond to a
component of the velocity.

In that way,
∗ the equation of momentum, eq.28,

19
∗ the equation of convective heat transfer, eq.27,
∗ the continuity equation
∂ρ ~
+ ∇ · (ρV~ ) = 0 (31)
∂t
compose the system of equations to solve for a general convective heat transfer
problem

Boundary conditions
Conditions at the wall:

ˆ Momentum and continuity equations: no-slip conditions, that is velocities equal to that
of the wall

ˆ Convective heat transfer equation: same conditions as discussed earlier for conductive
heat transfer.

Example: Couette Flow


Couette flow is a case of a parallel flow in which fluid is moving only in one direction. Couette
flow is a stationary flow between two infinite plates one of which is steady and another moves
with a velocity U in its plane, the distance between plates is L. In the present case suppose
that each plate has fixed temperature T0 and TL , respectively. The illustration to the problem
is given at fig.14, let us define heat transfer and mass transfer in the channel for U = 10m/s,
TL = 300 C, T0 = 100 C, L = 3mm. As a liquid, let us consider two cases: oil and water, whose
properties are given below.

Figure 14: Illustration to the problem of Couette flow with heat transfer

20
First, let us write down the system of governing equations for two dimensions:
∂u ∂v
+ =0 (32)
∂x ∂y
 2
∂ u ∂ 2u

∂u ∂u ∂p
ρu + ρv =− +µ + (33)
∂x ∂y ∂x ∂x2 ∂y 2
 2
∂ v ∂ 2v

∂v ∂v ∂p
ρu + ρv =− +µ + (34)
∂x ∂y ∂y ∂x2 ∂y 2
 2
∂ 2T

∂T ∂T ∂ T
ρcp u + ρcp v =κ + + µΦ (35)
∂x ∂y ∂x2 ∂y 2
 2
∂ C ∂ 2C

∂C ∂C
ρu + ρv =D + (36)
∂x ∂y ∂x2 ∂y 2

Here µΦ is a viscous dissipation that for incompressible fluid is presented as:


" 2  2  2 #
∂u ∂v ∂u ∂v
µΦ = µ + +2 +2 (37)
∂y ∂x ∂x ∂y

Boundary conditions:

u|y=0 = 0 u|y=L = U
T |y=0 = T0 T |y=L = TL
C|y=0 = C0 C|y=L = CL
(38)

In the given case v ≡ 0, so, no equation for the v component of the fluid. In the equation
for the x-component of the velocity ∂u/∂x = 0 , the terms which are left:

∂p ∂ 2 u
− + =0
∂x ∂y 2
(39)

The motion of the liquid happens due to the viscosity and the the motion of the plate, but not
because of the pressure difference. Then, the only term which is left and the solution is:

∂ 2u
=0
∂y 2
u(y) = C1 y + C2

Accounting for the boundary conditions:


y
u(y) = U
L
(40)

21
The energy equation is simplified as well as viscous dissipation term:

∂ 2T
κ + µΦ = 0
∂y 2
 2
∂u
µΦ = µ
∂y
2
∂ T ∂u
2
= −µ
∂y ∂y
(41)

Taking the derivative of the velocity, we can solve the equation for the energy taking into
account its boundary conditions:
 2
∂ 2T U
2
= −µ
∂y L
2 2
µU y
T =− + C3 y + C4
2Lκ L2
TL − T0 µU 2
C4 = T0 C3 = + (42)
L 2κL
Finally,

µU 2 y  y 2
 
y
T (y) = T0 + (Tl − T0 ) + − (43)
L 2κ L L

The solution for the concentration can be written immediately:


y
C(y) = C0 + (Cl − C0 ) (44)
L
Let us consider the case of an oil, the water and the air, properties are given below:
oil water air
Thermal conductivity κ, W/(m · K) 0.145 0.6 0.03
dynamic viscosity µ, P a · s 0.8 0.006 20.8 · 10−6

External flow. Boundary layer


Illustration to the problem of an external flow around a solid object is given in fig.??. For the
flow in the boundary layer the following simplifications are generally made: the thickness of the
boundary layer δ is very small, so, the dervatives of all variables over direction perpendicular
to the surface within the boundary layer are much larger than derivatives along the surface:
∂Ψ ∂Ψ
>> , Ψ can be one of those: u, v, T, C
∂y ∂x
Also, the velocity along the surface is much higher than the velocity perpendicular to it: u >> v.
∂u ∂v
>> (45)
∂y ∂y

22
Figure 15: A turbine blade in an air stream

Taking this into account, find that equations given for two dimensions eq.32-eq.36 can be
simplified up to the following system:
∂u ∂v
+ =0
∂x ∂y
∂u ∂u ∂p ∂ 2u
u +v =− +ν 2
∂x ∂y ρ∂x ∂y
∂p
=0
∂y
∂T ∂T ∂ 2T
u +v =a 2
∂x ∂y ∂y
∂C ∂C ∂ 2C
u +v =D 2
∂x ∂y ∂y
Let us introduce non-dimensional variables of the form:
x y u v
x∗ = y∗ = u∗ = v∗ =
L L U∞ U∞
and
p T − T0 C − C0
p∗ = 2
T∗ = C∗ =
ρU∞ T∞ − T0 C∞ − C0
The the given equations can be rewritten:
∂u∗ ∂v∗
+ =0
∂y∗ ∂y∗
∂u∗ ∂u∗ dp∗ ν ∂ 2u
u∗ +v∗ =− +
∂x∗ ∂y∗ dx∗ U∞ L ∂y 2
∂T ∗ ∂T ∗ a ∂ 2T ∗
u∗ +v∗ =
∂x∗ ∂y∗ U∞ L ∂y∗2
∂C∗ ∂C∗ D ∂ 2 C∗
u∗ +v∗ =
∂x∗ ∂y∗ U∞ L ∂y∗2
Let us introduce for non-dimensional complexes the following numbers:

23
U∞ L
Reynolds: Re ≡ (46)
ν
ν
Prandtl: Pr ≡ (47)
a
ν
Schmidt: Sc ≡ (48)
D
(49)

That finally gives:


∂u∗ ∂v∗
+ =0
∂y∗ ∂y∗
∂u∗ ∂u∗ dp∗ 1 ∂ 2u
u∗ +v∗ =− +
∂x∗ ∂y∗ dx∗ Re ∂y 2
∂T ∗ ∂T ∗ 1 1 ∂ 2T ∗
u∗ +v∗ =
∂x∗ ∂y∗ Re P r ∂y∗2
∂C∗ ∂C∗ 1 1 ∂ 2 C∗
u∗ +v∗ =
∂x∗ ∂y∗ Re Sc ∂y∗2
With the boundary conditions:

u ∗ |y∗=0 = 0 u ∗ |y∗←∞ ← 1
v ∗ |y∗=0 = 0
T ∗ |y∗=0 = 0 T ∗ |y∗←∞ ← 1
C ∗ |y∗=0 = 0 C ∗ |y∗←∞ ← 1

Important: although physical properties affect the motion of the liquid and the heat transfer,
only their combinations will define the character of the flow. In particular, it is clear that
the non-dimensional velocity is a function of the coordinates, of the pressure gradient and of
the Reynolds number:

u∗ = f1 (x∗, y∗, dp ∗ /dx∗, Re)

Similarly:

T ∗ = f2 (x∗, y∗, dp ∗ /dx∗, Re, P r)


C∗ = f3 (x∗, y∗, dp ∗ /dx∗, Re, Sc)

Furthermore, as it can be discussed in a previous chapter, the heat flux between the surface
and the fluis is expressed as:

dT
q = −kf on the other hand q = h(T0 − T∞ )
dy y=0

24
consequently, a local convective heat coefficient can be defined as

kf dT /dy|y=0
h=−
T0 − T∞
dT T∞ − T0 dT ∗
=
dy L dy∗ y∗=0

kf dT ∗
h=
L dy∗ y∗=0

Using this, define a non-dimensional heat exchange coefficient which is also termed Nusselt
number:

hL dT ∗
Nu ≡ = (50)
kf dy∗ y∗=0

In a similar way, accounting for the fact that mass transfer in the liquid can be written in the
same way as for the heat:

dC
qm = −D on the other hand qm = hm (C0 − C∞ )
dy y=0

D dC∗
hm =
L dy∗ y∗=0

we can define a dimensionless parameters for the mass transfer termed Sherwood number:

hm L dC∗
Sh ≡ = (51)
D dy∗ y∗=0

It is clear that for a given (defined) geometry which is involved in a heat and mass transfer
(this means that for all cases dp ∗ /dx∗ is the same) with various liquids, one has:

N u = f4 (x∗, Re, P r)
Sh = f5 (x∗, Re, Sc)

Furthermore, the averaged heat and mass transfer coefficient can be obtained by integration
over a surface of the body, i.e., over x∗. consequently, the averaged dimensionless parameters
are:
hL
Nu ≡ = f6 (Re, P r)
a
hm L
Sh ≡ = f7 (Re, P r)
a
Often the average values of N u and Sh related to an object with a characteristic size L can
be presented in a form

N uL = CRem
L Pr
n

ShL = CRem
L Sc
n

25
Example:
A turbine blade is subjected to the air flow as indicated in fig.16. The experimental test shows
that in this case the density of the heat flux from the air to the turbine blade is q = 95kW/m2 .
To maintain the turbine in a steady state with the surface temperature Ts = 8000 C, the heat
should be removed and this is done with a coolant circulating inside the blade. Determine the

Figure 16: A turbine blade in an air stream

heat flux at the same dimensionless location for a similar turbine blade having a chord length
of L = 80mm (two times larger) if the airflow has the same temperature but its velocity is
decreased by two.

Flat plane in parallel flow


In a case of a steady, incompressible laminar flow with constant fluid properties, negligible
viscous dissipation and for dp/dx = 0 the equations for the boundary layer take the form:
∂u ∂v
+ =0
∂x ∂y
∂u ∂u ∂ 2u
u +v =ν 2
∂x ∂y ∂y
∂T ∂T ∂ 2T
u +v =a 2
∂x ∂y ∂y
∂C ∂C ∂ 2C
u +v =D 2
∂x ∂y ∂y
Blasius proposed a method for the solution of these equation with replacing of variables. It is

based on an idea that the thickness of the dynamic boundary layer increases as δ ∼ x with
the distance from the edge of the plate. Consequently, it can be supposed that the velocity
within the boundary layer is ”udapted” to its growth and behaves similarly to itslef, or, in
other words
u y 
=f
U∞ δ
Actually, using a dimensional analysis it can be shown that
 1/2
νx
δ= (52)
U∞

26
Then, first, introduce a new coordinate taken as:
r
y U∞
η= =y (53)
δ(x) νx

Second, the stream function is used instead of two velocities. The stream function ψ is defined
in a way such that a relation between it and the velocities is:
∂ψ ∂ψ
u≡ and v≡−
∂y ∂x
With use of the stream function the continuity equation is automatically satisfied, so, it is not
needed any more.
Third, we introduce a new variable based on a stream function, which is supposed to be
dependent only on η:
ψ ψ
f (η) = =√
U∞ δ(x) U∞ νx
ψ = U∞ δ(x)f (η) (54)

Then using the chain rule for taking derivatives, obtain:


∂ψ ∂ψ ∂η ∂f 1 ∂f
u≡ = = U∞ δ(x) = U∞
∂y ∂η ∂y ∂η δ(x) ∂η
and r  
∂ψ ∂ 1 νU∞ df
v≡− = (U∞ δ(x)f (η)) = η −f
∂x ∂x 2 x dη

Taking the necessary derivatives, obtain further:

∂u U∞ d2 f
=− η
∂x 2x dη 2
r
∂u U∞ d2 f
= U∞
∂y νx dη 2
2
∂ u U∞ d3 f
2
= η
∂x2 νx dη 3
Taking obtained derivatives and putting them into the momentum equation, find a third-
order ordinary homogeneous differential equation (Blausius’ equation) with appropriate bound-
ary conditions:

d3 f d2 f
+ f =0 (55)
dη 3 dη 2

df
= f (0) = 0
dη η=0

df
=1

η=1

27
Figure 17: Solution for a Blasius equation: hydrodynamic laminar boundary layer over a flat
plate, from Incropera, De Witt

This equation can be integrated numerically and tabulated. The table of solution is given in
fig.17. It is reasonable to accept the ”edge” of the boundary layer at a coordinate η (or y) at
which u/U∞ = 0.99. From the table at fig.17 we can find that it occurs at η = 5. Then,
5 5x
δ=p =√
U∞ /νx Rex

Note that now the energy equation can be solved: introducing the non-dimensional temperature
T ∗ = (T − Ts )/(Ts − T∞ ) and assuming T ∗ = T ∗ (η), obtain:

d2 T ∗ P r dT ∗
+ f =0 (56)
dη 2 2 dη
T ∗ (0) = 0
T ∗ (∞) = 1

Similarly, for the mass transfer equation it can be obtained :

d2 C∗ Sc dC∗
+ f =0 (57)
dη 2 2 dη
C ∗ (0) = 0
C ∗ (∞) = 1

These equations also can be integrated numerically for various values of Re and P r.

28
Important to note:

1. The eq.55, eq.56 and eq57 are similar, moreover, they are identical for P r = 1 and Sc = 1.
This means that if P r ≈ Sc ≈ 1, then the hydrodynamic, thermal and concentration boundary
layer will behave similarly. More generally:
δ
1 ≈ P r1/3 (58)
δT
2. It is demonstrated that for P r > 0.6 the non-dimensional gradients of temperature and
concentration are
dT ∗
= 0.332P r1/3
dη η=0

dC∗
= 0.332Sc1/3
dη η=0

Then, the convective heat exchange can be found as well as N u number:


q 1 −kf dT
hx = =
Ts − T∞ Ts − T∞ dy
hx x 1/2
N ux ≡ = 0.332Rex P r1/3 (59)
ν
In the same manner:
q 1 −DdC
hm,x = =
Cs − C∞ Cs − C∞ dy
hm,x x 1/2
Shx ≡ = 0.332Rex Sc1/3 (60)
ν
Then the values of the convective coefficient averaged over a finite distance can be found
that gives:
hx = 2hx
hm,x = 2hm,x
1/2
N ux = 0.664Rex P r1/3
1/2
Shx = 0.664Rex Sc1/3
3. For small P r numbers the solution given above is not applicable. According to 58, the
thermal boundary layer will be thinner than the hydrodynamic one and for the estimations a
uniform velocity equal to U∞ can be taken within the thermal boundary layer. An appropriate
relation is:
1/2
N ux = 0.565P ex
1/2
Shx = 0.565P em,x
where P e = Rex P r and P em,x = Rex Sc
4. Finally, a general correlation is proposed for all P r numbers:
1/2
0.3387Rex P r1/3
N ux = 1/4
(61)
[1 + (0.0468/P r)2/3 ]

29
External flow: some additional remarks
Laminar to turbulence transition

The formulae obtained in the previous section are valid for the laminar flow over the plate.
Yet, it is known that a laminar flow may transform into a turbulent flow if a critical Reynolds
number is achieved (fig.18). The value of the critical Reynolds number lies between 105 to
3 × 106 depending on various factors, so, a representative value Rec = 5 × 105 is often taken.

Figure 18: Transition from laminar to turbulent flow inear the flat plate. From F.P.Incropera,
D.P.DeWitt, Fundamentals of heat and mass transfer, 4th edition

A nice video about the turbulent flow can be seen at https://www.youtube.com/watch?v=


e1TbkLIDWys.

When flow becomes turbulent, its character changes, in particular all its characteristics (u, v, T )
oscillate because of multiple vortices existing everywhere in the liquid. These vortices provide
additioinal transport of the momentum, heat and mass and to take this into account the Navier-
Stokes equations were ”updated”. Also, the flow character near the wall changes and it can
be shown that for turbulent flows δ ≈ δT ≈ δm . Proposed relations for local values of the
non-dimensional convective heat and mass transfer coeficients are:

4/5
N ux = 0.0296Rex P r1/3 , 0.6 < P r < 60 (62)
4/5
Shx = 0.0296Rex Sc1/3 , 0.6 < Sc < 60 (63)

30
External flow across a steady cylinder

An example of the external flow across a cylinder is shown in fig.19. It is seen that the distance
between the streamline closest to the cylinder and the surface of the latter increases alongthe
flow and wakes (vortices) appear behinf the cylinder. A more detailed scheme is shown in fig.21.

Figure 19: Transition from laminar to turbulent flow inear the flat plate. From a movie at
https: // www. youtube. com/ watch? v= TOUylg7Eyec

Note that the “forced” velocity (upstream velocity) V is different from the velocity around the
cylinder (free stream velocity) u∞ The flow is stopped at the stagnation point and, according to
Bernulli law, the pressure is the highest at this point, from this point the fluid starts to accelerate
and a boundary layer starts to develop. Further it moves, the pressure gradient along the surface
decreases till zero and becames negative. At this point a boundary layer is separated from the
surface and the wake (region of recirculating flow - cf.https://en.wikipedia.org/wiki/Wake)
appears downstream. Yet, appearance of wake is different from the transition from a laminar
flow to the turbulent!

Figure 20: Transition from laminar to turbulent flow inear the flat plate. From F.P.Incropera,
D.P.DeWitt, Fundamentals of heat and mass transfer, 4th edition

For the circular cylinder the characteristic length is the diameter, and the Reynolds number is

31
defined with the upstream velocity V as
VD
ReD ≡ (64)
ν
Since the momentum of fluid in a turbulent boundary layer is larger than in the laminar
boundary layer, it is reasonable to expect transition to delay the occurrence of separation. If
ReD ≤ 2 × 105 , the boundary layer remains laminar, and separation occurs at θ = 800 , if
ReD ≥ 2 × 105 , boundary layer transition occurs, and separation is delayed to θ = 1400 .
Various correlation are proposed for the non-dimensional heat transfer coefficient, i.e. Nus-
selt number, most of which are of the form

N uD = CRem
DP r
1/3
(65)

Churchill and Bernstein have proposed a single comprehensive equation that covers the
entire range of ReD for which data are available, as well as a wide range of P r. The equation
is recommended for all ReD and P r ≥ 0.2 and has the form:

Figure 21: Churchill and Bernstein general equation for a Nusselt number for a flow across
the cylinder, From F.P.Incropera, D.P.DeWitt, Fundamentals of heat and mass transfer, 4th
edition

32
Free convection. Vertical surface
In a previous chapter a case of the forced convection was considered, where the flow is defined
by an imposed velocity and does not depend on temperature distribution in the system. Yet,
there is a force in the momentum equation 28:
∂ V~  
ρ ~
+ ρ V , ∇ V~ = ρ~g − ∇p + µ∇2 V~
∂t
Let us consider a motion of the liquid near a vertical plate, the temperature of the plate is
higher than that of the liquid: Tp > Tf , the gravity vector is directed downward

Figure 22: Illustration to the formation of the boundary layer near a vertical surface due to
natural convection From F.P.Incropera, D.P.DeWitt, Fundamentals of heat and mass transfer,
4th edition

In this situation the fluid near the plate is moving up due to buoyancy force (Archimede
force) and it is quiescent (immovable) far from the plate. Also, a temperature in the fluid will
change from the value equal to the temperature of the plate to the temperature of the fluid.
In other words, a kinematic boundary layer and a thermal boundary layer will appear near
the plate. Moreover, in this situation temperature distribution affects the motion of the fluid.
Indeed, the governing equations within the boundary layer are:

∂u ∂v
+ =0
∂x ∂y
∂u ∂u 1 ∂p ∂ 2u
u +v =− −g+ν 2 (66)
∂x ∂y ρ ∂x ∂y
∂p
=0
∂y
∂T ∂T ∂ 2T
u +v =a 2
∂x ∂y ∂y
Outside the boundary layer the velocity is zero and x momentum equations are reduced to
1 ∂p
− =g (67)
ρ∞ ∂x

33
Since the pressure does not change with the distance from the plate (derivative on y is zero), the
pressure gradient remains similar within the boundary layer and outside the boundary layer,
i.e. eq.67 can be put in the momentum equation:
∂u ∂u g ∂ 2u
u +v = (ρ∞ − ρ) + ν 2
∂x ∂y ρ ∂y
In the first term the density is variable since it depends on temperature. This dependence can
be introduced with the use of the volumetric thermal expansion coefficient:
1 dρ
β=− (68)
ρ dT
The derivative of the density over the temperature can be approximated as
1 ρ∞ − ρ
β≈−
ρ T∞ − T
(ρ∞ − ρ) ≈ −ρβ(T∞ − T )

with this approximation, the complete system of equation for the flow is:

∂u ∂v
+ =0
∂x ∂y
∂u ∂u ∂ 2u
u +v = gβ(T − T∞ ) + ν 2 (69)
∂x ∂y ∂y
2
∂T ∂T ∂ T
u +v =a 2
∂x ∂y ∂y
Note that variation of the density was taken into acount only in the term related to the force,
but not in other terms. This approach is named Boussinesque approximation.

Remark: for the gases which can rather often be considered as ideal, the thermal expansion
coefficient is obtained directly from the equation of the state ρ = p/RT . Then:
 
1 ∂ρ 1
β=− = (70)
ρ ∂T p T

Free convection: Vertical surface - nondimensional analysis


A non-dimensional analysis of the problem can be done using the an approach similar to the
one used in the case of the forced flow. First, introduce the non-dimensional coordinate and
variables:
x y
x∗ = y∗ =
L L
u v T − T∞
u∗ = v∗ = T∗ =
u0 u0 T0 − T∞
Note that since the initial statement problem does not contain any velocities, u0 is an arbitrary
velocity of the upstream flow and L is a characteristic size (length of the plate)

34
Then x- momentum and energy equation take the following form in the non-dimensional coor-
dinates:
∂u∗ ∂u∗ gβ(Ts − T∞ ) 1 ∂ 2u
u∗ +v∗ = T ∗+
∂x∗ ∂y∗ u20 ReL ∂y 2
∂T ∗ ∂T ∗ 1 ∂ 2T ∗
u∗ +v∗ =
∂x∗ ∂y∗ ReL P r ∂y∗2
It is convenient to introduce a ratio of the buoyancy and viscous force which is termed Grashof
number GrL :
gβ(Ts − T∞ )L3
GrL = (71)
ν2
Then the equations take the form
∂u∗ ∂u∗ GrL 1 ∂ 2u
u∗ +v∗ = T ∗+
∂x∗ ∂y∗ Re2L ReL ∂y 2
∂T ∗ ∂T ∗ 1 ∂ 2T ∗
u∗ +v∗ =
∂x∗ ∂y∗ ReL P r ∂y∗2
From these equations it can be expected that
1.The non-dimensional velocity (and the kinematic boundary layer) will be the function of the
numbers ReL and GrL
2.The non-dimensional temperature gradient, i.e.N u number, will be the function of the ve-
locities and the P r number, that is, of numbers P r, ReL and GrL . But this will be the case
only if both types of convection exist in the system: a forced convection characterized by the
ReL (with the imposed velocity U∞ ) and a free convection characterized by the GrL , and if
GrL /ReL ≈ 1. In other cases:
· The forced convection is absent or weak, then
GrL /ReL >> 1 => N u = f (GrL , P r)
· The forced convection prevails over the free convection:
GrL /ReL << 1 => N u = f (ReL , P r)

Free convection: Vertical surface - similarity solution


A system of equation describing the problem is given above. It should be subjected to the
boundary conditions:

uy=0 = vy=0 = 0 Ty=0 = Ts


uy→∞ → 0 Ty→∞ → T∞

Now, introduce a new variable:


 1/4
y Grx
η≡ (72)
x 4

35
and introduce a stream function ψ(x, y) related to the u and v velocities similar to the case of
the forced convection. Then, introduce a new function f (η):
"  1/4 #
Grx
ψ(x, y) ≡ f (η) 4ν (73)
4

Using u = ∂ψ/∂y, obtain:


 1/4  1/4
∂ψ ∂η Grx 0 1 Grx
u= = 4ν f (η)
∂η ∂y 4 x 4
in a similar way a velocity v = −∂ψ/∂x can be obtained and all required derivatives. Finally,
a system of equations for a functions f (η) and a non-dimensional temperature T ∗ can be
obtained:

f 000 + 3f f 00 − 2(f 0 )2 + T ∗ = 0
T ∗00 +3P rf T ∗0 = 0

with the boundary conditions:


0
fη=0 = fη=0 =0 T ∗η=0 = 1
0
fη→∞ →0 T ∗η→∞ → 0

Since temperature enters into the equation for f (η) and temperature depends on P r, the
function f (η) also depends on P r. So, solutions can be obtained for various values of the P r
number. Some examples are given below in fig.23.
Now, when the temperature distribution is known, we can obtain the convective heat transfer
coefficient, which will be (obviously) a function of the P r number:
q
q = h(Ts − T∞ ) => h =
Ts − T∞
hx qx
Nu = =
κ κ(Ts − T∞ )
On the other hand, by definition: q = −κ(dT /dy)y=0 and using non-dimensional variable:
 1/4
dT κ Grx dT ∗
q = −κ = (Ts − T∞ )
dy y=0 x 4 dη η=0
That gives for N u:
 1/4  1/4
hx Grx dT ∗ Grx
Nu = =− = g(P r)
κ 4 dη η=0 4
dT ∗
The dependence of the non-dimensional derivative on P r is easily observed in fig.23b.

From numerical simulations an interpolation formula was found for g(P r):

0.75P r1/2
g(P r) = 0 ≤ Pr ≤ ∞
(0.609 + 1.221P r1/2 + 1.238P r)1/4

36
Figure 23: Laminar, free convection boundary layer conditions on an isothermal, vertical sur-
face. (a) Velocity profiles. (b) Temperature profiles. From F.P.Incropera, D.P.DeWitt, Fun-
damentals of heat and mass transfer, 4th edition

Free convection: Vertical surface - transition to turbulent flow


Similar to the case forced convection, it can happen that the flow will change its regime from
the laminar to turbulent (fig.24).
This transition is defined by a Rayleigh number which is a product of the Grashof and
Prandtl numbers:

Ra = GrP r (74)

For a vertical flat plate the critical Rayleigh number is:


gβ(Ts − T∞ )x3
Rax,c = Grx,c P r = ≈ 109 (75)
νa
Example
Consider a 0.25-m-long vertical plate that is at 70C. The plate is suspended in air that is at
25C. Estimate the boundary layer thickness at the trailing edge of the plate if the air is quies-
cent. How does this thickness compare with that which would exist if the air were flowing over
the plate at a free stream velocity of 5 m/s?

37
Figure 24: Transition from laminar to turbulent free convective flow near a heated vertical
plate From F.P.Incropera, D.P.DeWitt, Fundamentals of heat and mass transfer, 4th edition

Hint: use the results presented in fig.23 to answer the question related to the free convection.

38
Chapter 3. Radiation - Introduction.
To remind:

ˆ Solids → mechanical interaction of individual particles → conduction

ˆ Fluids:
mechanical interaction of individual particles → conduction
+
mechanical interaction of microvolumes (agglomeration of molecules) → advection.
conduction + advection = convection

Conclusion: conductive and convective modes of heat transfer require presence of media
(presence of particle). Here we start to learn radiation, a process of energy transfer which can
occur without any media.

Types of electromagnetic waves. And what is thermal radiation


Definition 4. Thermal radiation is the transfer of internal energy by electromagnetic waves.

Classically, we can consider that electromagnetic waves are originated by a charged particles
wich moves with acceleration, for examples, by electrons.

The electromagnetic wave can be considered as related to each other oscillations of electrical
and magnetic field in time and space. Examples of a linearly polarized el.-m. wave is given in
Fig.25

Figure 25: A linearly polarized el.-m. wave.

http://en.wikipedia.org/wiki/Electromagnetic_radiation
Electromagnetic radiation is a transverse wave meaning that the oscillations of the waves are
perpendicular to the direction of its propagation and to the energy transfer. The electromag-
netic wave propagates (moves) with a velocity v and is characterised by a wavelength λ or by
a frequency f . The frequency and the wavelength are inversely proportional and related to the

39
velocity v:

v = fλ (76)
[f ] = Hz = s−1
[λ] = [L] = m, µm = 10−6 m, nm = 10−9 m

The maximal velocity the el.-m. wave has in vaccum and it is equal to the speed (velocity) of
light c = 299792458m/s or (3 · 108 m/s).

Definition 5. Inside a material the velocity of the wave is less by n times: n = c/v, n is called
”index of refraction” or ”refractive index” of material.

Definition 6. The range of all possible wavelength (or all possible frequencies) is called ”spec-
trum”.

As you see in Fig.26, according to the wavelength (or to the frequencies) the electromagnetic
waves are divided into: radio waves, microwaves, infrared radiation, visible light, ultraviolet
radiation, X-rays, gamma rays.

Figure 26: Ranges for different el.-m. waves.

http://commons.wikimedia.org/wiki/File:EM_Spectrum_Properties_edit.svg
! IMPORTANT ! Usual indication for the thermal radiation is that the electromagnetic
waves which transfer the thermal energy cover a small part of ultraviolet spectrum, a visible
light and an infrared part of the spectrum as it is shown in fig.27 and 28

40
Figure 27: ” ”

41
Figure 28: Place of the thermal radiation in spectrum.

In fact, any electromagnetic radiation can heat a material when it is absorbed. To prove,
several phenomena can be mentioned, e.g.:

* Microwaves heat the food in the microwaves ovens (λ = 120 − 320mm). It occurs due
to interaction of el.-m. waves with polar molecules which typically can be found in liquids.

Some remarks about the term ”microwaves” (e.g. ”microwave oven”) Microwaves are
radio waves with wavelengths ranging from as long as one meter to as short as one millimeter,
or equivalently, with frequencies between 300 MHz (0.3 GHz) and 300 GHz.
The prefix ”micro-” in ”microwave” is not meant to suggest a wavelength in the micrometer
range. It indicates that microwaves are ”small” compared to waves used in typical radio broad-
casting, in that they have shorter wavelengths.

* Also, radio waves interact with electrons in antennas and make them move and produce an
electromagnetic current in a wire. This current can heat the antenne.

However, if a ”usual object” has a temperature different from the absolute zero, it emits the
most intensively electromagnetic waves in a spectrum range which is defined as spectrum of
thermal radiation. This is conditioned by the energetical levels (which are discrete!) available
for particles (e.g. electrons) which make up the body.

Specific features of radiative energy transfer


Generally, an object can emit the electromagnetic waves in the whole spectrum (fig.29a). This
is the first ”feature” of the thermal radiative heat transfer. The second ”feature” is that
these waves can be emitted by the object with various intensity in different directions (fig.29b).
That is in a general case the radiative heat flux can depend:

42
Figure 29: Illustration to spectral and directional dependence of radiative energy transfer: a)
Liquid iron radiates in visible spectrum but also in infrared; b) and c) show examples of different
directional radiation

1. on coordinates (x, y, z)

2. on the wavelength λ (or, the same, on the frequency of radiation, f )

3. and on the direction which is defined by a polar angle θ and azimuatl angle φ

Emissive power. Solid angle. Intensity


We start to study the radiative heat transfer from radiation of surfaces. Let Q be the rate
at which energy leaves the surface A in all directions over the whole spectrum. Dimension of
[Q] = J/s = W , Watt.

A value which is similar to the density of the heat flux ~q which was defined in the first lecture
is called emissive power.

Definition 7. A total hemispherical emissive power E∩ is the energy which leaves a unit surface
dA in all direction over the whole spectrum per unit of time.

Figure 30: Radiation from a surface – for definition of the emissive power

dQ W
E∩ = [E∩ ] = (77)
dA m2
Let dQλ (λ) be the rate at which energy leaves the surface A in all directions within the spectral
range [λ, λ + dλ] (around the wavelength λ). Dimension of [dQλ (λ)] = J/(s · µm) = W/µm.
R∞
Evidently: Q = 0 dQλ (λ)dλ.

43
Definition 8. A spectral hemispherical emissive power dE∩,λ (λ) is the energy which leaves a
unit surface dA in all direction within the spectral range [λ, λ + dλ] per unit of time.

d2 Qλ (λ) W
E∩,λ (λ) = [E∩,λ ] = (78)
dAdλ m2
· µm
Also, we can consider waves which leave the elementary surface dA in a certain direction. A
direction in space is defined by the two angles, polar angle θ and an azimuthal angle φ, the
last can be chosen arbitrarily, but the polar angle is always calculated from the normal to the
surface. Similarly to ”small” (elementary) surface or volume, a small spatial angle (or solid
angle) is considered in space.

Definition 9. The value of a spatial angle dω is defined as a ratio of the element dAsp which
the angle cuts on a sphere (see Fig.31) to the square of the radius of the sphere. The measure
of the solid angle is steradian, sr.

2
dω = dAsp /rsp (79)
dAsp = rsinθdθrdφ
dω = sinθdθdφ (80)

Figure 31: (a) For definition of a solid (spatial) angle. (b) Illustration to ”directional” radiation

Definition 10. Intensity I(θ, φ) is the radiative energy flux in a solid angle (θ − dθ, θ +
dθ; φ − dφ, φ + dφ) per unit area normal to the direction (θ, φ).

44
Intensity is a radiative flux of the thermal energy related to the area of a surface
which is perpendicular toward a direction of interest, fig.32.
d2 Q0θ,φ
Iθ,φ = (81)
dA cos θdω

Figure 32: Illustration to definition of intensity

Definition 11. Consequently, the spectral intensity Iλ (θ, φ) is the radiative energy flux in
a solid angle (θ − dθ, θ + dθ; φ − dφ, φ + dφ) per unit area normal to the direction (θ, φ) within
a spectral range [λ, λ + dλ].

d3 Qλ,θ,φ
Iλ,θ,φ (λ, θ, φ) = (82)
dAcosθdωdλ
Similarly to the emissive power, the relation of the intensity and spectral intensity is via the
integral:
Z ∞
Iθ,φ = Iλ,θ,φ (θ, φ)dλ (83)
0

The dimension of the intensity is J/(m2 · sr) and of the spectral intensity J/(m2 · m · sr).

Diffuse surface
Note that
d2 Q
=
dAdλZ
= dωIλ,θ,φ (θ, φ)cosθ =

Z π/2 Z 2π
= dθ dφIλ,θ,φ (θ, φ)cosθsinθ =
0 0

= E∩,λ (84)

45
Definition 12. A surface dA is called diffuse if intensity of the emitted radiation is independent
on direction of this radiation.

Using Eq.84, we obtain the relation between the hemispherical emissive power and inten-
sity which is emitted by a surface dA for spectral and total values:

Eed,∩,λ = πIedλ (85)


Eed,∩ = πIed (86)

46
Chapter 4.Blackbody.
To remind:
• Radiative heat transfer: energy is transferred by electromagnetic waves
• Two features of RHT: dependence on wavelength, dependence on direction
• Spectrum. Solid (spatial) anlge
• Characteristics of radiation:
– Total hemispherical emissive power, spectral hemispherical emissive power
– Intensity, spectral intensity
• Diffuse emitters
In fact, in addition to its two specific features (spectral and direction dependence), radiative
heat transfer has a third feature that radiative properties of materials are defined with respect
to an ideal object which does not exist in reality and which is called ”blackbody”. To remind:
a body can emit, abosrb and reflect radiation.

Blackbody
Definition 13. A blackbody is defined as an ideal body which absorbs (internally) all the inci-
dent radiation regardless of wavelength and direction (whatever is wavelength or direction).

From this definition follow the properties of a black-body:

1. The blackbody is an ideal emitter – no body emits more than a blackbody being at the
same temperature

2. The blackbody is a perfect emitter at every wavelength, i.e. (at a fixed temperature) it
emits maximal energy at every wavelength

3. The blackbody is a perfect emitter at every direction, i.e. it emits maximal energy in
any direction

4. The radiation which fills a cavity with black walls is isotropic

The properties (1)-(4) of a blackbody can be proved based on the definition of a black-
body, as you can read, for example, in a book ”Thermal radiation heat transfer”, vol.1,
R.Siegel,J.R.Howell (1st ed.1968).

Despite that the blackbody is an ideal object which does not exist in the nature, there are some
models of it. The most known ”approximation to a blackbody” is a cavity with a small opening
(a hole), as shown in fig.33. When radiation falls into the hole, it undergoes multiple reflections
and absorptions. If the opening is small, very little of incident radiation will escape back, i.e.
the energy will be completely absorbed that corresponds to the definition of the black body.
This model can be used also as a source of a ”black radiation” which can be obtained by

47
heating up the cavity. The copper walls (or another material witha high thermal conductiv-
ity), can provide equal temperature distribution over the walls, i.e. a radiation of the same
temperature will be emitted by any part of the wall.

Figure 33: A model of a blackbody, from Thermal radiation heat transfer, vol.1,
R.Siegel,J.R.Howell, 1968

There are several modern versions of black bodies and black surfaces, as you can find in litera-
ture or in the internet. A super-black coating based on chemically etching of a nickel-phosphorus
alloy was created in 2002 by Richard Brown and his colleagues at the UK’s National Physical
Laboratory (NPL) in Teddington. In Fig.34 a surface of this material is shown.

Later, in 2009, a team of Japanese scientists created a material whose absorption is even closer
to the ideal black body. This material is based on vertically aligned single-walled carbon
nanotubes shown in fig.35.
This material absorbs between 98% and 99% of the incoming light in the spectral range from
the ultra-violet to the far-infrared regions. The photo as well as microscopic images of this
material is taken from the open article in Proc Natl Acad Sci U S A. 2009 April 14;
106(15): 60446047 . An abbreviation SWNT means ”single-walled carbon nanotubes”.

Planck’s distribution. Rayleigh-Jeans and Wein’s approximations


The properties of the blackbody is known, its model exists. The question is how to calculate
the magnitude of the energy which blackbody with a temperature T emits at each wavelength
λ, or within a range of the spectrum? The relation which allows one to do this, was obtained by

48
Figure 34: SEM (scanning electronic microscope) image of NiP black surfaces with different ini-
tial concentration of phosphorus R.J.C.Brown, P.J.Brewer, M.J.T. Milton, J. Mater. Chem.,
2002, 12, 27492754

Max Planck. It is necessary to stress that this formula cannot be obtained from purely thermal
(thermodynamic) consideration. The knowledge from quantum physics is required.

For a blackbody the spectral distribution of hemispherical emissive power in a vacuum is given
as a function of the absolute temperature T and the wavelength λ by:
2πhC02
E∩,λ,b (λ, T ) =     (87)
hC0
λ5 exp −1
λkT
h = 6.6266 · 10−34 J · s Planck’s constant
J
k = 1.381 · 10−23 Boltzmann constant
K
m
C0 = 299792458 ≈ 3 · 108 speed of light
s
If a blackbody is surrounded by a media with the index of refraction n, the Eq.87 takes the
form:
2πhC02
E∩,λ,b (λ, T ) =     (88)
hC 0
n2 λ5 exp −1
nλkT
For most engineering problems where surrounding media is the air or a gas the eq.87 (without
n) is applicable.

Note that eq.87 is often given in a form


2πC1
E∩,λ,b (λ, T ) =     (89)
C 2
λ5 exp −1
λT
2 8 W · µm4
C1 = hC0 = 3.742 · 10 − − 1st radiation constant
m2
C2 = hC0 /k = 1.439 · 104 µmK − − 2nd radiation constant

49
Figure 35: Microscopic structure of SWNT forest. (A) SWNT forest grown on an 8-in silicon
wafer. (B) SEM image of SWNT forest vertically standing on a silicon substrate. (Scale bar,
0.5 mm.) (C) SEM image showing top surface of SWNT forest. (Scale bar, 0.5µm.) (D) SEM
image showing side surface of SWNT forest. (Scale bar, 5µm.)

Before we plot the graph Eλ,b (λ, T ), let us investigate characteristics of this function.
Firstly, With temperature increase the Eλ,b (λ, T ) increase as well.
Secondly, what if λT >> hC0 /k? Then (let x = hC0 /(λkT )):

x2
ex = 1 + x + + ...
2
2πhC02
E∩,λ,b (λ, T )λT →∞ = "  2 #
hC0 hC0
λ5 1+ + + ... − 1
λkT λkT
2πC0
E∩,λ,b (λ, T )λT →∞ = kT (90)
λ4
The eq.90 is known as Rayleigh-Jeans distribution which is valid for long wavelength (or
for a very high temperatures). It is seen that at a constant temperature Eλ,b (λ, T ) decreases
when λ increases.

Thirdly, what if λT → 0? Then:


 
C2
exp >> 1
λT
2πhC02
E∩,λ,b (λ, T ) =   (91)
hC 0
λ5 exp
λkT

The eq.91 is known as Wien’s approximation (do not confuse with Wien’s displacement law
considered below). It is seen that Eλ,b (λ, T )λ→0 → 0

50
Since Eλ,b (λ, T ) tends to zero when its argument λ approaches either to 0 or to ∞, but it is
not zero for other λ, it can be expected that this functions has a maximal value within this
range. To obtain this value, one has to differentiate the initial eq.87 over λ and equate to zero.
As a result, a transcendental equation will be obtained, solution of which gives the following
relation:

λmax T = C3 (92)
C3 = 2898µm · K

Eq.92 is called Wien’s displacement law. From it follows that if temperature of a black body
increases (T2 > T1 ) then the maximum of the spectral emissive power will occur at another
wavelength: λ2 < λ1 , i.e. – with growth of the temperature, the wavelength at which the
blakbody emits the maximal energy displaces toward shorter values.
Knowing the fundamental properties of the Planck’s distribution, it is possible to plot a distri-
bution of the spectral emissive power over the wavelength, which is given in fig.36.

Figure 36: Spectral emissive power of a blackbody according to Planck’s law

Problem 1. The radiation from the Sun and from the Earth can be considered similar to
radiation of a blackbody. At what wavelengths has the Sun and the Earth their maximal emissive
power? The TSun = 5777K, TEarth = 298K

To solve the Problem 1, the Wein’s law is applied. Consequently, for the Sun λmax,Sun ≈ 0.5µm
(corresponds to green color), for the Earth λmax,Sun ≈ 10µm (infrared).

51
For info: wavelength range for visible light
violet: 380 ÷ 450nm = 0.38 ÷ 0.45µm yellow: 570 ÷ 590nm = 0.57 ÷ 0.59µm
blue: 450 ÷ 475nm = 0.450 ÷ 0.475µm orange: 590 ÷ 620nm = 0.59 ÷ 0.62µm
cyan: 475 ÷ 495nm = 0.475 ÷ 0.495µm red: 620 ÷ 750nm = 0.62 ÷ 0.75µm
green: 495 ÷ 570nm = 0.495 ÷ 0.570µm

Total emisive power of a blackbody. Fractions


Knowing the spectral emissive power, the total value can be obtained by integration over the
whole range of the wavelengths:
Z 0
E∩,b = E∩λ,b (λ, T ) · dλ = . . . = σT 4 (93)

W
σ = 5.67 · 10−8
m2
· K4
here σ is a Stephan-Boltzmann constant. To remind: from geometrical point of view, the inte-
gral gives the area under the curve.

For some reasons it is often required to know the emissive power emitted by a unit surface dA
within a wavelength band (wavelength range) [λ1 , λ2 ]. How to find it? To perform the inten-
gration similar to one in eq.93 but from [λ1 , λ2 ]. This can be done either NUMERICALLY
or with use of tabulated values which called ”fractions”.

Definition 14. Fraction at a certain temperature T , F0→λ1 ,T , is defined for a wavelength band
which lies between 0 and λ1 a ratio of a hemispherical emissive power of a blakbody in this band
to the total hemispherical emissive power of a blackbody at this temperature:

R λ1 R λ1
0
E∩,λ,b (λ, T )dλ E∩,λ,b (λ, T )dλ
F0→λ1 ,T = = 0 (94)
E∩,b (T ) σT 4
1 λ1 T 1 1 λ1 T 2πhC02
Z Z
F0→λ1 ,T = E ∩,λ,b (λ, T )d(λT ) =  hC0
  d(λT ) (95)
σ 0 T5 σ 0 (λT )5 exp kλT −1

As you can see from the eq.94, the faction F0→λ1 ,T depends only on the product λT , that
gives the possibility to tabulate this value. The table given below is taken from the book
”Fundamentals of heat and mass transfer”, 6th edition, by Incropera, DeWitt, Bergman, Lavine.

Problem 2. Black surface is at uniform temperature 2000K.


a. What is the total emissive power of the surface?
b. What is the wavelength below which 10% of emitting radiation is concentrated?
c. What is the wavelength above which 10% of emitting radiation is concentrated?
See the scheme for the problem below in fig.37.

52
Figure 37: Illustration to the problem 2 (from the book ”Fundamentals of heat and mass
transfer”, 6th edition, by Incropera, DeWitt, Bergman, Lavine)

For a: E∩,b (T ) = σT 4 = 5.67 · 10−8 · 20004 = 5.67 · 204 = 907.22kW/m2 For b: F0→λ1 ,T = 0.1.
From the Table: λ1 T = 2195µmK. Therefore, λ1 = 2195/2000 ≈ 1.097µm
For c: F0→λ2 ,T = 0.9. From the Table: λ2 T = 9382µmK. Therefore, λ2 = 9382/2000 ≈
4.691µm

53
54
Chapter 5. Opaque materials. Emissivity of real surfaces
To remind:
• In radiative heat transfer (RHT) we deal with intensity and emissive power.
• These values can be spectral or total that is related to the wavelength of electromagnetic
waves.
• In RHT a ”reference body” – blackbody – exists
• The properties of the black body were derived with the help of a quantum physics and are
well-known
• The spectral hemispherical emissive power of a blackbody is given by Planck’s distribution
• The total hemispherical emissive power of a blackbody is given by Stephan-Boltzmann law
• Balckbody is a diffuse emitter.
• Diffuse emitter is a body (a surface) intensity of which is independent of direction.

The ideal behaviour of a blackbody serves as a standart. The ability of a particular body (or
surface) to emit or to absorb radiation is first of all related to its atomic structure. But it also
depends on the quality of the surface (if we speak about a solid body).
In the next lectures we will discuss opaque materials, i.e. materials (body or surfaces) which
do not allow electromagnetic waves (within the spectral region which coreponds to the thermal
radiation - see lecture 2) to propagate. We will follow the notation proposed in the book
”Thermal radiation heat transfer”, vol.1, R.Siegel, J.R.Howell,
Notation:

ˆ dQ0 means the flux of energy emitting within a small (differential ) solid angle

ˆ dQλ means the flux of energy emitting within a wavelength range [λ, λ + dλ]

ˆ d2 Q0λ means the flux of energy emitting in a unit direction within a wavelength range
[λ, λ + dλ]

ˆ d3 Q0λ means the flux of energy emitting in a unit direction within a wavelength range
[λ, λ + dλ] per unity of the surface (by a differential area).

Spectral directional emissivity


It follows from the definition of the blackbody, that any real body must emit less energy than
a blackbody. It is logical to introduce a coefficient which would describe a relation between the
ability to emit radiation at a temperature T for a real body and for a blackbody, i.e.

energy emitting by a real surf ace at temperature TA


(emissivity) = (96)
energy emitting by a blackbody at temperature TA
Consequently, a blackbody would have the emissivity equal to 1 for all wavelength.
In general case the amount of energy emitting by a surface may vary with wavelength and

55
direction. Therefore, the emisivity has to be dependent on these values as well. In fig.38
a qualitative illustration of a directional spectral emissivity is given. Here an intensity of
emitting power of a real surface for various angles is given in comparison with an intensity of
a blackbody.

Figure 38: qualitative illustration to the definition of a directional spectral emissivity.

Definition 15. Spectral directional emissivity 0λ (λ, θ, φ, TA ) is a ”fundamental value”


which is defined as the ratio of the emissive ability of the real surface at a temperature TA into
a solid angle directed at (θ, φ) at the wavelength λ to the emissive ability of the blackbody at a
temperature TA at the same wavelength λ within a differential solid angle at the same direction.
The ”emisive ability” to be understood by the energy heat flux:

d3 Q0λ,e (λ, θ, φ, TA )
0λ (λ, θ, φ, TA ) = 3 0 (97)
d Qλ,b (λ, θ, φ, TA )

Recalling the definition of intensity given by eq.82, we can write:

Iλ,e (λ, θ, φ, TA )cosθdω(θ, φ)dA


0λ (λ, θ, φ, TA ) =
Iλ,b (λ, TA )cosθdω(θ, φ)dA
Iλ,e (λ, θ, φ, TA )
0λ (λ, θ, φ, TA ) = (98)
Iλ,b (λ, TA )

Example 1. At 600 from the normal, a surface heated to 833K has a directional spectral
emissivity of 0.70 at a wavelength of 5µm. What is the spectral intensity in this direction?

Solution for the example is given at the end of the chapter.


Spectral directional emissivity is a property which is often unknown for a given material.
However under certain conditions radiative properties of materials can be estimated using the
elecromagnetic theory. For large wavelengths (after the visible spectrum) the electromag-
netic theory predicts different behaviour of 0 (θ, φ, T ) for dielectric and metals (see fig.39).
For dielectrics (Fig.40) the spectral emissivity remains constant until the angle from the
normal direction to the surface is about 700 , then 0λ (λ, θ, φ, T ) decreases rapidly to zero. There
is very few emitted radiation from dielectrics for θ > 700 .

56
Figure 39: Behaviour of the (spectral or total) directional emissivity for metals (conductors)
and dielectrics (non-conductors)

Figure 40: Behaviour of the total directional emissivity for dielectrics (non-conductors), illus-
tration from the book ”Radiative Heat Transfer”, 2nd edition M.F.Modest, page 63

For metals the electromagnetic theory also predicts constant behaviour in the range between
00 and 700 (generally, with a rather small value of emissivity). But in the range between 700 and
900 the 0λ (λ, θ, φ, T ) firstly increases sharply and then drops to 0. For example, this behaviour
was obtained theoretically for a pure and smooth surface made of Platinum and was confirmed
by the measurements (see fig.41).

Figure 41: Spectral directional emissivity of a smooth Pt surface at λ = 2µm for temperatures
300K – 5, 687K – ◦ and 1400K – , the straight line shows theoretical values, from ”Thermal
radiation heat transfer, vol.1” by Robert Siegel and John R. Howell.

57
Similar dependence of 0λ (λ, θ, φ, T ) on the direction is shown in Fig.42 for Nickel (Ni), Chromium
(Cr), Manganese (Mn) and Aluminium (Al) (without the drop of the emissivity to the zero).

Figure 42: Behaviour of the total directional emissivity for metals, from ”Radiative Heat Trans-
fer”, 2nd edition M.F.Modest, page 63

Important remark: For radiation of shorter wavelength (visibile spectrum) the dependence of
0λ (λ, θ, φ, T ) on direction may be quite different! As an example, in fig.43 a directional spectral
emissivity of pure titanium is shown, the example is taken from the book ”Thermal radiation
heat transfer, vol.1” by Robert Siegel and John R. Howell. Here for the wavelength ≥ 4µm
directional emissivity is constant for 0 < θ ≤ 450 , then it increases and then drops to zero. But
for λ = 1.62µm, 1.0µm and 0.43µm the directional emissivity does not vary with angle!

Figure 43: Spectral directional emissivity of pure titanium, from ”Thermal radiation heat
transfer”, vol.1, R.Siegel, J.R.Howell

Another example which shows comparison between the values of normal (perpendicular to the
emitting surface) spectral emissivities obtained theoretically and experimentally is given in the
table 44. Note, that values for Mg differ in 4 times.

58
Figure 44: Normal spectral emissivity 0λ,n (λ, T ) of various metals in visible and inrared spec-
trum (n and η - are spectral refractive and absorption indice, respectively. From ”Thermal
radiation heat transfer”, vol.1, R.Siegel, J.R.Howell

Total directional emissivity


If for a certain material a spectral directional emissivity is known, the intensity of the radiation
which emits this surface can be calculated using the eq.98. However, for most of materials only
averaged values of emissivity are known and they are discussed below.

Definition 16. Total directional emissivity 0 (θ, φ, TA ) is defined as the ratio of the emis-
sive ability of the real surface at a temperature TA into a solid angle directed at (θ, φ) at all
wavelength to the emissive ability of the blackbody at a temperature TA within a differential solid
angle at the same direction at all wavelength. The ”emisive ability” to be understood by the
energy heat flux:

d2 Q0e (θ, φ, TA )
0 (θ, φ, TA ) = (99)
d2 Q0b (θ, φ, TA )
Recalling that the total heat flux is an integral over the whole spectrum of the spectral heat
flux, the eq.99 can be re-written:
R∞ 3 0
0
d Qλ,e (λ, θ, φ, TA )dλ
 (θ, φ, TA ) = R0 ∞ 3 0
0
d Qb (λ, θ, φ, TA )dλ

As in previous definition, we use intensity to re-write this equation:


R∞
0 Iλ,e (λ, θ, φ, TA ) cos θdω(θ, φ)dAdλ
 (θ, φ, TA ) = 0 R ∞
0
Iλ,b (λ, TA ) cos θdω(θ, φ)dAdλ
R∞ R∞
Iλ,e (λ, θ, φ, T A )dλ Iλ,e (λ, θ, φ, TA )dλ
0 (θ, φ, TA ) = 0 R ∞ = 0
0
Iλ,b (λ, TA )dλ σTA4 /π

59
Further, we can use spectral directional emissivity that has been defined above to continue:
R∞
0 0
0λ (λ, θ, φ, TA )Iλ,b (λ, TA )dλ
 (θ, φ, TA ) = (100)
σTA4 /π
The last eq.100 shows that the total directional emissivity is an average (over the wavelength)
of the spectral directional emissivity an that the Planck’s distribution Iλ,b /(σTA4 ) plays the role
of the density probability function.

Note: If a total directional emissivity of a surface is known, than the total intensity Ie (θ, φ, T )
of the radiation emitting by the surface in a direction (θ, φ) may be calculated using eq.100.

Remark: Most of dependecies discussed above for directional spectral values are applicable to
the total directional emissivity.

Consequently, if a dependence of the spectral directional emissivity on the wavelengths is known,


a total directional emissivity can be calculated, let us consider another example.

Example 2. At TA = 556K the directional spectral emissivity 0λ (λ, θ, φ, TA ) of a surface can
be approxmated by 0.8 in the range λ = 0 ÷ 5µm and 0.4 for λ > 5µm. What is the value of
0 (θ, φ, TA )?

Solution for the example is given at the end of the chapter.

Spectral hemispherical emissivity


Definition 17. Spectral hemispherical emissivity ∩,λ (λ, TA ) is defined as the ratio of the
emissive ability of a real surface at a temperature TA into a hemisphere at a wavelength λ to the
emissive ability of the blackbody at a temperature TA into a hemisphere at the same wavelength.
The ”emisive ability” to be understood by the energy heat flux:

d2 Qλ,e (λ, TA )
∩,λ (λ, TA ) = (101)
d2 Qλ,b (λ, TA )
Recalling that the hemispherical spectral heat flux is an integral over the hemisphere of the
directional spectral heat flux, obtain:
R 3 0
d Qλ,e (λ, θ, φ, TA )dω
∩,λ (λ, TA ) = R∩ 3 0

d Qb (λ, θ, φ, TA )dω
Using intensity as above, we can re-write:
R R
I (λ, θ, φ, TA ) cos θdAdω
∩ R λ,e
Iλ,e (λ, θ, φ, TA ) cos θdω
∩,λ (λ, TA ) = = ∩ R

I λ,b (λ, TA ) cos θdAdω Iλ,b (λ, TA ) ∩
cos θdω
R 2π R π/2
0
dφ 0
dθIλ,e (λ, θ, φ, TA ) cos θ sin θ
∩,λ (λ, TA ) =
π · Iλ,b (λ, TA )

60
Further, Iλ,b (λ, TA ) can be introduced into the inegral since it does not depend on the angle
Z 2π Z π/2
Iλ,e (λ, θ, φ, TA ) cos θ sin θ
∩,λ (λ, TA ) = dφ dθ
0 0 π · Iλ,b (λ, TA )

where we can use definition of a spectral directional emissivity:


Z 2π Z π/2
1
∩,λ (λ, TA ) = dφ dθ 0λ (λ, θ, φ, TA ) cos θ sin θ (102)
π 0 0

Note, that the integration can be performed on φ


Z π/2
∩,λ (λ, TA ) = 2 · dθ 0λ (λ, θ, φ, TA ) cos θ sin θ (103)
0

The eq.102 demonstrates that the hemispherical directional emissivity is a spectral directional
emissivity averaged over a hemisphere. A distribution of the radiative energy of a black body
over space (cos θdω = cos θ sin θdθdφ) plays the role of the density probability function.

Illustration to the hemispherical spectral emissivity is given in fig.45. Using notation given in
this figure, it can be written that ∩,λ1 = b/a

Figure 45: Illustration to a hemispherical spectral emissivity, from ”Fundamentals of heat and
mass transfer”, 6th edition, by Incropera, DeWitt, Bergman, Lavine

Remark: Both, for metals and dielectrics, hemispherical spectral emissivities does not differ
much from the normal value of the emissivity. This is explained by the fact that they varies
strongly only in a very narrow region of the angles.
Generally, for dielectrics
∩,λ (λ, T )
0.95 ≤ 0 ≤ 1.0 (104)
n,λ (λ, θ = 0, T )
And for metals
∩,λ (λ, T )
1.0 ≤ 0
≤ 1.3 (105)
n,λ (λ, θ = 0, T )

61
Spectral emissivities for different materials

In this section some illustrations to spectral behaviour of emissivities of different materials are
given. Since there is almost no difference if a hemispherical emissivity or emissivity in normal
direction is taken (eq.104, eq.105) for most of materials the value of the 0n,λ (λ, θ = 0, T ) is
presented.

For metals in the visible and infrared spectral range the emissivity decreases if the wavelength
increases (Fig46). The curve 5 for copper in fig.46 is an exception.

Figure 46: Normal spectral emissivities. a. Spectral nomal emissivity of polished Ni, curves
1-4 are experimental data from diffferent sources: 1-T=294K, 2-T=1200K, 3- T=1272K, 4-
T=294K, another work; curves 5 and 6 are calculated: 5-T=1100K, 6-T=294K. b: Variation
with wavelength of normal spectral emissivity for polished metals, 1 - Mo at 1110K, 2- Fe at
1317K, 3 - Pt at 1217K, 4 - Ni at 1200K, 5 - Cu at 1242K. From ”Thermal radiation heat
transfer”, vol.1, R.Siegel, J.R.Howel

For dielectrics there are very few measurements which cannot be summarized briefly. This
is explained by the fact the dielectric surfaces cannot be smoothed (polished) as metallic one
and the quality (roughness) of the surface affects the emissivity.

Commercial and non-commercial ”black surfaces”.


Normal spectral emissivities for a SWNT forest and NiP alloy (presented in the previous lecture)
and some other commercialle available ”black surfaces” are given in the fig.47
The authors of the article ”...report that among all known materials, a forest of vertically aligned
single-walled carbon nanotubes (SWNTs) behaves most similarly to a black body. Specifically,
from optical studies, we revealed that a SWNT forest possesses a nearly constant and near-
unity emissivity (absorptivity) of 0.980.99 across a wide spectral range from UV (200 nm) to

62
Figure 47: Normal spectral emissivity of SWNT forest (red line) and commercially available
black surfaces that are denoted as NiP alloy (blue), black coating (black), and black paint
(yellow)

far infrared (200µm). We speculate that this important black body behavior originates from
the homogeneous sparseness and alignment of the SWNTs within the forest.”(Kohei Mizuno,
Juntaro Ishii, Hideo Kishida, Yuhei Hayamizu, Satoshi Yasuda, Don N. Futaba, Motoo Yumura,
and Kenji Hata, PNAS, vol.106, April 14 2009)

Total hemispherical emissivity


Finally, as it follows from a previous consideration a total hemispherical value represents a
directional spectral emissivity averaged over the whole spectrum and over the whole hemisphere.
The averaging over a spectrum is done with a Planck’s distribution taken as an probability
distribution function and the averaging over a hemisphere is done with an expression of an
elementary solid angle.

Definition 18. Total hemispherical emissivity (TA ) is a ratio of the emissive ability of
the real surface at a temperature TA into a hemisphere over the whole spectrum to the emissive
ability of the blackbody at a temperature TA into a hemisphere over the whole spectrum. The
”emisive ability” to be understood by the energy heat flux:

dQe (TA )
(TA ) = (106)
dQb (TA )
The eq.106 can be re-written using the definition of intensity and the ”fundamental value” of
a spectral directional emissivity:
R∞ R
dλ dω · Iλ,e (λ, θ, φ, TA ) cos θdA
∩ (TA ) = 0 R ∞ ∩ R
0
dλ ∩ dω · Ib (λ, TA ) cos θdA
R∞ R
0
dλ ∩ dω · 0λ (λ, θ, φ, TA )Iλ,b (λ, TA ) cos θ
∩ (TA ) = (107)
σTA4

63
The eq.107 can be written in different ways, for example:
Z ∞ Z
1
∩ (TA ) = dλIλ,b (λ, TA ) dω · 0λ (λ, θ, φ, TA ) cos θ
σTA4 0 ∩

Using relations obtained above the value of a total hemispherical emissivity ∩ (TA ) can be
expressed via hemispherical spectral emissivity ∩,λ (λ, TA ) or directional total emissivity
0 (θ, φ, TA ).

Dependence of emissivities on temperature

Generally, for metals emissivity of metals increases with increasing temperature (fig.48).

Figure 48: Temperature dependence of total hemispherical emissivity for different material: 1
- Graphite, 2 - Magnesium oxide, 3 - polished inconel, 4 - Magnesium, 5 - Tunsten (W), 6 -
polished gold. From ”Thermal radiation heat transfer”, vol.1, R.Siegel, J.R.Howel

For dielectrics the emissivity may either increase or decrease with increasing temperature. It
depends on the material, see fig49 and curve 2 in fig.48.

Figure 49: Temperature dependence of total normal emissivities 0n of selected materials

64
Emissivities for various materials
• Generally, emissivities are small for metallic polished surfaces.
• Emissivity of rough surfaces are higher, than for polished.
• Oxide layer may increase eissivity.

Solutions to the examples


Solution to the example 1:
From the table on page 29, for a blackbody at λTA = 5 · 833 = 4165µm · K,
Iλ,b (λ, TA )/(σT 5 ) ≈ 0.540391 · 10−4 (µm · K · sr)−1 .
Then: Iλ,b (λ, TA ) = 0.540391 · 10−4 · 5.67 · 10−8 · (8.33)4 (100)4 · 833 = 1.2289 · 103 .
noindent Let us look at dimensions:
(µm · K · sr)−1 · W/(m2 · K 4 ) · K 5 = µm−1 · sr−1 · W/m2
Then, the intensity from the surface is
0λ,e (λ, θ, φ, TA ) · Iλ,b (λ, TA ) = 8.6 · 102 µm−1 · sr−1 · W/m2 = 8.6 · 108 W/(m3 · sr)

Solution to the example 2:


Since the spectral directional emissivity is known, we start with the eq.100 which can be re-
written as follows:
R5 0 R∞ 0
0  (λ, θ, φ, TA )Iλ,b (λ, θ, φ, TA )dλ
0 λ 5
λ (λ, θ, φ, TA )Iλ,b (λ, θ, φ, TA )dλ
 (θ, φ, TA ) = + (108)
σTA4 /π σTA4 /π

Further, R5 R∞
0 0
Iλ,b (λ, θ, φ, TA )dλ 5
Iλ,b (λ, θ, φ, TA )dλ
 (θ, φ, TA ) = 0.8 + 0.4 (109)
σTA4 /π σTA4 /π
Recalling the previous lecture:

0 (θ, φ, TA ) = 0.8F (0 → 5µm)+0.4·F (5µm → ∞) = 0.8F (0 → 5µm)+0.4 [1 − F (0µm → 5µm)]


(110)

λ · TA = 2830µm · K (111)

From the table at page 29, F (2830µm · K) = 0.227, consequently:

0 (θ, φ, TA ) = 0.8 · 0.227 + 0.4 · (1 − 0.227) = 0.490 (112)

65
Chapter 6. Opaque materials. Absorptivity of real sur-
faces
To remind:
• In RHT a ”reference body” – blackbody – exists
• To define the ability of emit radiative energy for real surface a coefficient named ”emissivity”
is used
• The emissivity can be directional or hemispherical with respect to the space
• The emissivity can be spectral or total with respect to the wavelengths (spectrum)
The definition of the blackbody is that it absorbs all the possible radiation incident on it re-
gardless the direction or the wavelength of radition. Obviously, any real surface cannot absorb
more radiation than a blackbody, but it will absorb less radiation than a blackbody. In order
to characterise the ability of a real surface to absorb the energy a coefficient termed ”absorp-
tivity” is used.

Note that absorption is related to the energy which is incident onto the surface (or the object)
which is discussed below.

Incident radiation
In fig.50a a surface dAe (this surface is a part of a hemisphere, it is not a black surface) emits
the energy. A part of this energy reaches a surface dA.

a b

Figure 50: Incident radiation. a - In terms of energy emitting by the surface dAe , b - In terms
of energy intercepting by the surface dA.

This part of the energy is defined by a solid angle dωe and by the radiative properties of the
surface dAe that can be written as follows:

d3 Q0λ,e (λ, θ, φ, TAe ) = Iλ,e (λ, θ, φ, TAe )dωe · dλdAe (113)


dA cos θ
dωe = (114)
R2

66
On the other hand, this is the energy incident on the surface dA, which we will denote as
d3 Q0λ,e = d3 Q0λ,i , here the index ”i” in d3 Q0λ,i means Incident:

d3 Q0λ,i (λ, θ, φ, TAe ) = Iλ,e (λ, θ, φ, TAe ) cos θdωe dλ · dAe

Note that
dA cos θ dAe
dωe · dAe = 2
· dAe = dA cos θ 2 = dA cos θ · dω
R R
Therefore,
d3 Q0λ,i = Iλ,e (λ, θ, φ, TAe ) cos θ · dωdλdA (115)

Finally, intensity with which the surface dAe emits the radiation can be considered as intensity
of the radiation incident on the surface dA

d3 Q0λ,i = Iλ,i (λ, θ, φ, Ti ) cos θ · dωdλdA (116)

Similarly to the definition of the intensity of the energy which a surface emits, the intensity
of the incident radation can be defined as follows:

Definition 19. Spectral intensity of the incident radiation Iλ,i (λ, θ, φ) is the energy
incident from a unit solid angle dω(θ, φ) per unit time, per unit wavelength interval per area
normal to the direction of the solid angle.

Definition 20. Total intensity of the incident radiation Ii (θ, φ) is the energy incident
from a unit solid angle dω(θ, φ) per unit time, per area normal to the direction of the solid
angle.

Important! Note that it in a general case it is not possible to indicate a temperature


to which the intensity of the incident radiation corresponds.

Similary to the spectral hemispherical emissive power E∩,λ the following value can be defined:

Definition 21. Hemispherical spectral irradiation G∩,λ is the energy incident onto a unit
surface dA from all direction per unit time, per unit wavelength interval.

Z
G∩,λ = Iλ,i (λ, θ, φ) cos θ · dω

Z 2π Z π/2
G∩,λ = dφ dθIλ,i (λ, θ, φ) cos θ sin θ (117)
0 0
W
[G∩,λ ] =
m2 · µm
Definition 22. Hemispherical total irradiation G∩ is the energy incident onto a unit
surface dA from all direction in a whole spectrum.

67
Z ∞
G∩ = G∩,λ (λ)dλ
0
Z ∞ Z 2π Z π/2
G∩ = dφ dθIλ,i (λ, θ, φ) cos θ sin θdλ (118)
0 0 0
W
[G∩ ] =
m2

Spectral directional absorptivity


An opaque body (or an opaque surface) which is subjected to the radiation can absorb or not
the incident energy. A fraction of the energy which is absorbed can be defined with absorption
coefficient (absorptivity):

energy f lux incident onto a real surf ace and absorbed by it


absorptivity = (119)
energy f lux incident onto a real surf ace
Similarly to the emisivity, absorptivity has to be dependent on directions and wavelengths.
The fig.50b is a good illustration for directional absorptivities, the first of which is a spectral
directional absorptivity

Definition 23. The fraction of spectral energy incident within the unit solid angle dω(θ, φ)
which is absorbed by the unit surface is defined as the directional spectral absorptivity a0λ (λ, θ, φ).

d3 Q0λ,a (λ, θ, φ)
a0λ (λ, θ, φ) = (120)
d3 Q0λ,i (λ, θ, φ)
The eq.120 can be re-written in the form of intensity of incident radiation:

d3 Q0λ,a (λ, θ, φ)
a0λ (λ, θ, φ) = (121)
Iλ,i (λ, θ, φ) cos θdωdλdA

Absorptivity depends on the temperature of the surface which is subjected to the radiation
(Ta ).

Kirchhoff ’s Law for directional spectral properties


There are multiple formulation of the Kirchchoff’s law in radiation, the one is given here.
Kirchhoff ’s Law For a body of any arbitrary material, emitting and absorbing thermal
electromagnetic radiation at every wavelength in thermodynamic equilibrium, the ratio of its
emissive power to its dimensionless coefficient of absorption is equal to a universal function
only of radiative wavelength and temperature, the perfect black-body emissive power.

68
d3 Q0λ,e (λ, θ, φ, TA )/dA
= d3 Q0λ,b (λ, θ, φ, TA )/dA
a0λ (λ, θ, φ, TA )
0λ (λ, θ, φ, TA ) · d3 Q0λ,b (λ, θ, φ, TA )/dA
= d3 Q0λ,b (λ, θ, φ, TA )/dA
a0λ (λ, θ, φ, TA )
That is:
0λ (λ, θ, φ, TA ) = a0λ (λ, θ, φ, TA ) (122)
Remark: since dependence of spectral directional emissivity on the temperature is rather weak,
eq.122 is often used even if the temperature of the surface TA is not equal to the ”effective
temperature” of the incident radiation, i.e. if the surface is not in thermal equilibrium with the
environment.

Total directional absorptivity


Definition 24. The directional total absorptivity is the ratio of the energy including all wave-
lengths that is absorbed from a given direction to the energy incident from that direction.

R∞ 3 0
0 d2 Q0a (θ, φ, Ta ) 0
d Qλ,a (λ, θ, φ)
a (θ, φ) = = ∞ (123)
d2 Q0i (θ, φ)
R
0
d3 Q0λ,i (λ, θ, φ)
Similarly to emissivities, a relation between a ”fundamental” value a0λ (λ, θ, φ) and all averaged
absorptivities can be obtained:
R∞
a0 (λ, θ, φ)d3 Q0λ,i (λ, θ, φ)dλ
a0 (θ, φ) = 0
R∞
0
d3 Q0λ,i (λ, θ, φ)dλ
R∞
a0 (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdωdλ
a0 (θ, φ) = 0 R∞
0
Iλ,i (λ, θ, φ) cos θdωdλ
R∞ 0
0 a (λ, θ, φ)Iλ,i (λ, θ, φ)dλ
a (θ, φ) = 0 R ∞ (124)
0
Iλ,i (λ, θ, φ)dλ
With use of Kirchhoff’s law the equation 124 can be re-written in terms of spectral directional
emissivity: R∞ 0
 (λ, θ, φ)Iλ,i (λ, θ, φ)dλ
a (θ, φ, Ta ) = 0 λR ∞
0
(125)
0
Iλ,i (λ, θ, φ)dλ
What is a relation between a0 (θ, φ) and 0 (θ, φ)?

Kirchhoff ’s Law for directional total properties. Gray surfaces


To remind: relation between 0 (λ, θ, φ) and 0 (θ, φ) is given in Chapter 5, eq.100, which is
repeated here: R∞ 0
 (λ, θ, φ, TA )Iλ,b (λ, θ, φ, TA )dλ
0 (θ, φ, TA ) = 0 λ 100
σTA4 /π

69
There are 2 situations when a0 (θ, φ) = 0 (θ, φ)
• The first case: if intensity of the incident radiation can be presented as a product of two
functions, one of which depends on angles but does not depend on wavelengths C(θ, φ) and
another part is proportional (with a coefficient V independent of wavelength) to the intensity
of the blackbody, with the same temperature TA i.e: Iλ,i (λ, θ, φ, TA ) = C(θ, φ) · V · Iλ,b (λ, TA ).
Then:
R∞ 0
 (λ, θ, φ, TA )C(θ, φ) · V · Iλ,b (λ, TA )dλ
a (θ, φ) = 0 λ R ∞
0

0
C(θ, φ) · C · Iλ,b (λ, TA )dλ
R∞ 0
 (λ, θ, φ, TA )Iλ,b (λ, TA )dλ
a0 (θ, φ) = 0 λ R ∞ = 0 (θ, φ, TA ) (126)
0
Iλ,b (λ, TA )dλ

Note that if ”effective temperature” of the incident radiation is not equal to the temperature
of the surface, total directional emissivity and total directional absorptivity are not equal!
• The second case: If directional spectral emissivity does not depend on wavelength, then
directional spectral absorptivity does not depend on wavelength as well as 0λ (θ, φ) and eq.43
becomes: R∞
0 0 Iλ (λ, θ, φ)dλ
0
a (θ, φ) = λ (λ, θ, φ) R ∞ = 0 (θ, φ)
0
Iλ,i (λ, θ, φ)dλ

Definition 25. A surface which behaves like this, i.e. a0λ (λ, θ, φ) = 0λ (λ, θ, φ) = a0 (θ, φ) =
0 (θ, φ) is termed a directional gray surface.

”Gray” here indicates that the surface is ”almost black”: absorptivity is constant for all wave-
length but is not unity.

Spectral hemispherical absorptivity


Definition 26. The hemispherical spectral absorptivity is the fraction of the spectral energy that
is absorbed from the spectral energy incident from all directions over a surrounding hemisphere
(see fig.51): a∩,λ (λ) = d2 Qλ,a /d2 Qλ,i .

Figure 51: Hemispherical irradiation incident onto a surface dA.

70
The incident energy on dA from all directions of the hemisphere (fig.51) is given by the integral
Z
2
d Qλ,i (λ) = Iλ,i (λ, θ, φ) cos θdωdAdλ

The amount of the incident energy which is absorbed by the surface dA is
Z
2
d Qλ,a (λ) = a0λ,i (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdωdAdλ

The ratio of these quantities gives:


a0λ (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdω
R

a∩,λ (λ) = R (127)
I (λ, θ, φ) cos θdω
∩ λ,i

or by using Kirchhoff’s law for directional spectral properties:


R 0
 (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdω
a∩,λ (λ) = ∩ λ R (128)
I (λ, θ, φ) cos θdω
∩ λ,i

Kirchhoff ’s Law for hemispherical spectral properties. Diffuse spec-


tral surfaces
To remind: a hemispherical spectral emissivity is given in the Chapter 5 102 and is repeated
here:
1 2π
Z Z π/2
∩,λ (λ, TA ) = dφ dθ 0λ (λ, θ, φ, TA ) cos θ sin θ 102
π 0 0
The question is: when a∩,λ (λ) = ∩,λ (λ, TA )?
Looking at the eq.128, two cases can be revealed:
• The first case:
if Iλ,i (λ, θ, φ) is independent of (θ, φ) that is, if the incident spectral intensity is uniform over all
directions (diffuse irradiation of the surface). If this is the case, the eq.128 can be re-written:
Iλ,i (λ) ∩ 0λ (λ, θ, φ, TA ) cos θdω
R
a∩,λ (λ) = R = ∩,λ (λ, TA ) (129)
Iλ,i (λ) ∩ cos θdω
• The second case:
If the directional spectral properties are independent of angle, then:

0λ (λ, θ, φ, TA ) = 0λ (λ, TA ) = a0λ (λ, θ, φ) = a0λ (λ) (130)

for any angular variation of incident intensity. A surface for which the eq.130 is true is termed
a diffuse spectral surface.

Total hemispherical absorptivity


Definition 27. The hemispherical total absorptivity represents the fraction of energy absorbed
that is incident from all directions of the enclosing hemisphere and for all wavelengths (see also
fig.51).

71
The total incident energy (per unit of time) dQ∩,i that is intercepted by a surface element dA is
determined by integrating equation 116 over all λ and all angles (θ, φ) of the hemisphere which
results in: Z ∞Z
dQ∩,i = Iλ,i (λ, θ, φ) cos θdωdλdA
0 ∩
The total amount of energy absorbed is equal to
Z ∞Z
dQ∩,a = a0λ (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdωdλdA
0 ∩

Then, R∞R
a0 (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdωdλ
a∩ = 0 R ∞λ R

(131)
0
I (λ, θ, φ) cos θdωdλ
∩ λ,i
or using Kirchhoff’s law for directional spectral values:
R∞R 0
 (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdωdλ
a∩ = 0 ∩R ∞λ R (132)
0
I (λ, θ, φ) cos θdωdλ
∩ λ,i

Total hemispherical absorptivity can be expresses via two others averaged absorp-
tivities:
from eq.124, it follows that
Z ∞ Z ∞
0 0
a (λ, θ, φ)Iλ,i (λ, θ, φ)dλ = a (θ, φ) Iλ,i (λ, θ, φ)dλ
0 0

then the hemispherical total absorptivity can be written in terms of directional total ab-
sorptivity: R R ∞
Iλ,i (λ, θ, φ)dλ a0 (θ, φ) cos θdω

∩ 0
a∩ = R∞R
0
I (λ, θ, φ) cos θdωdλ
∩ λ,i
or: R 0
a (θ, φ)Ii (θ, φ) cos θdω
a∩ = ∩ R (133)
I (θ, φ) cos θdω
∩ i
from eq.127, it follows that
Z Z
0
aλ (λ, θ, φ)Iλ,i (λ, θ, φ) cos θdω = a∩,λ (λ) Iλ,i (λ, θ, φ) cos θdω
∩ ∩

then the hemispherical total absorptivity can be written in terms of spectral hemispherical
absorptivity: R∞ R 
a (λ) ∩ Iλ,i (λ, θ, φ) cos θdω dλ
0 R ∩,λ
a∩ = ∞ R 
0 ∩
Iλ,i (λ, θ, φ) cos θdω dλ
or, using the definition 21 of irradiation:
R∞
a∩,λ (λ)Gλ (λ)dλ
a∩ = 0 R λ (134)
0
G(λ)dλ

72
Kirchhoff ’s Law for hemispherical total properties.
To remind, the total hemispherical emissivity is given in the Chapter 5 by eq.107 which is
repeated here:
R∞ R
dλ ∩ dω · 0λ (λ, θ, φ, TA )Iλ,b (λ, θ, φ, TA ) cos θdA
∩ (TA ) = 0 R∞ R eq.107
0
dλ ∩
dω · Ib (λ, TA ) cos θdA

By comparison of the eq.132 and eq.107 it is seen that in a general case, the a∩ is equal to ∩
only if the intensity of the incident radiation can be presented in a form Iλ,i (λ, θ, φ) = C ·Iλ,b (λ),
i.e. if the incident radiation arrives from an object ”similar to a black body”.
However, there are others possible combinations. The simplest case is, obviously, when the
surfaces are diffuse-gray, i.e. when directional spectral properties do not depend on direction
or wavelength, in this case

a0λ = a0 = aλ = a∩ = 0λ = 0 = λ = ∩ (135)

Example 3. A diffuse firewall made of brick with a temperature Tw = 500K has the spectral
emissivity shown in fig.52a and is exposed to a bed of coals at 2000K (see a scheme of the
process in fig.52b. a.Determine the total hemispherical emissivity and emissive power of the
brick firewall. b.Determine the total absorptivity of the wall to the radiation from emission
by the coal (suppose that the emission from the coal proportional to that of the blackbody at
2000K).

Figure 52: a.Directional spectral emissivity of a brick wall; b.Scheme of the problem

73
Chapter 7. Opaque materials: reflectivity. Remarks on
transparent materials: transmissivity
To remind:
• Emissivity
• Absorptivity
• Emissivity and absorptivity of a certain surface (body) are linked by the Kirchhoff’s law
• Kirchhoff’s law is true for spectral directional values for a surface in the ”thermal equilibrium
with incident energy”
• Kirchhoff’s law is true for averaged values under certain conditions (see lectures)

Spectral and total bidirectional reflectivity


The reflectivity is a property that determines the fraction of the incident radiation reflected by
a surface.

Similarly to the emissivity and absorptivity, the reflectivity can be related to a specific wave-
length or to the whole spectrum, i.e. it can be spectral or total. With respect to the space (to
directions) the situation is more complicated since the incident energy is characterised by a cer-
tain distribution over directions, i.e. d3 Q0λ,i (λ, θi , φi ) but also a reflected energy is characterised
by (another) distribution over directions, i.e. d3 Q0λ,r (λ, θr , φr , θi , φi ), see fig.53. Therefore, the
reflectivity depends on two combinations of angles, in other words it is a bidirectional prop-
erty.

Figure 53: Incident and reflected radiation – illustration to bidirectional reflectivity. Note,
generally θr 6= θi and φi 6= φr

74
REMARK: There are various definitions for the bidirectional spectral reflectivity. The one is
taken from the book ”Thermal radiation heat tranfer”, vol.1, R.Siegel,J.R.Howell (1st ed.1968)
and is given below

Definition 28. Bidirectional spectral reflectivity is a ratio expressing the contribution


0
that Iλ,i (λ, θi , φi ) cos θi dωi makes to the reflected spectral intensity Iλ,r (λ, θr , φr ) in the (θr , φr )
direction:

0
Iλ,r (λ, θr , φr , θi , φi )
ρ00λ (λ, θr , φr , θi , φi )
= (136)
Iλ,i (λ, θi , φi ) cos θi dωi
Then, for the illustration in fig.53, the energy which leaves the surface dAe and then is reflected
from the surface dA toward the surface dAf is
0
d3 Qλ,r (λ, θi , φi , θr , φr ) = Iλ,r (λ, θr , φr , θi , φ) cos θr dωr · dA · dλ (137)
d3 Qλ,r (λ, θi , φi , θr , φr ) = ρ00λ (λ, θr , φr , θi , φi )Iλ,i cos θi dωi cos θr dωr · dA · dλ

Note that definition 28 presents a ratio of a reflected intensity divided by the energy
which an infinitesimal surface dA got from the direction (θi , φi ).
It can be demonstrated (see Seigel, Howell, ”Thermal radiation heat transfer”) that for the
bidirectional reflectivity defined with eq.136 the reciprocity law is fulfilled, i.e.

ρ00λ (λ, θr , φr , θi , φi ) = ρ00λ (λ, θi , φi , θr , φr ) (138)

Directional reflectivities
Directional-hemispherical reflectivity defines how much of the radiant energy incident
from one direction will be reflected into all directions (fig.54,left).

Hemispherical-directional reflectivity defines reflected intensity into one direction result-


ing from incident radiation coming from all directions, see (fig.54, right).

Definition 29. The spectral directional-hemispherical reflectivity is defined as the spec-


tral energy reflected into all solid angles divided by the incident energy from one direction (θ, φ).

d3 Q0λ,r (λ)
ρ0λ (λ, θi , φi )= 3 0
d Qλ,i (λ, θi , φi )
R 0
Iλ,r (λ, θr , φr , θi , φi ) cos θr dωr
ρλ (λ, θi , φi ) = ∩
0
(139)
Iλ,i (λ, θi , φi ) cos θi dωi
If the intensity of the reflected radiation does not depend on the angle of reflection, i.e. the
energy is reflected diffusely, then it can be taken out of the integral and we obtain
0
Iλ,r,d (λ, θi , φi )
ρ0λ,d (λ, θi , φi ) = (140)
1
· Iλ,i (λ, θi , φi ) cos θi dωi
π

75
Figure 54: Illustration to directional reflectivities. On the left: directional - hemispherical
reflectivity; On the right: hemispherical - directional reflectivity.

The problem is that along with the definition 29, you can find other definitions for spectral
directional-hemispherical reflectivity,as, for example, one given below:

Definition 30. The spectral directional-hemispherical reflectivity of a surface is defined


as the fraction of the spectral intensity incident in the direction (θ, φ) which is reflected by the
surface:

Iλ,r (λ, θi , φi )
ρ0λ (λ, θi , φi ) = (141)
Iλ,i (λ, θi , φi )
REMARK 1. The definition 30 with eq.141 in fact is not good because it is not clear what is
Iλ,r (λ, θi , φi ), note, that we usually define intensities via the energy.
REMARK 2. Even if we suppose that Iλ,r (λ, θi , φi ) in the definition 30 with eq.141 is fo a
diffuse surface, there will be a difference between two definitions in π/(cos θi dωi ) !
Nevertheless, to solve various problems we will use the definition 30 with eq.141.
As it is indicated above, another directional spectral reflectivity is spectral the hemispherical-
directional spectral reflectivity.

Definition 31. The spectral hemispherical-directional spectral reflectivity is the reflected spec-
tral intensity in the (θr , φr ) direction divided by the integrated average incident spectral intensity.

From eq.136, the intencity of radiation incident from the direction (θi , φi ) and reflected into the
direction (θr , φr ) is given as
0
Iλ,r (λ, θr , φr , θi , φi ) = ρ00λ (λ, θr , φr , θi , φi ) · Iλ,i (λ, θi , φi ) cos θi dωi (142)

Then, the reflected intensity into the (θr , φr ) direction is found by integrating equation 142 over
all incident directions:
Z
Iλ,r (λ, θr , φr ) = ρ00λ (λ, θr , φr , θi , φi ) · Iλ,i (λ, θi , φi ) cos θi dωi
∩,i

76
Then, the spectral hemispherical-directional reflectivity is:

ρ00 (λ, θr , φr , θi , φi ) · Iλ,i (λ, θi , φi ) cos θi dωi


R
0 ∩,i λ
ρλ (λ, θr , φr ) = 1
R (143)
I (λ, θi , φi ) cos θi dωi
π ∩,i λ,i

Hemispherical reflectivity
Finally, for the hemispherical reflectivity (spectral and total) which is schematically shown in
fig.55 the definition is unique.

Figure 55: Illustration to the hemispherical reflectivity.

Definition 32. Spectral hemispherical reflectivity is the ratio of the amount of spectral energy
d2 Qλ,r that is reflected by the surface dA into all directions to the amount of incident spectral
energy d2 Qλ,i that is intercepted by the surface dA from all directions.

Z
2
d Qλ,i = dλdA Iλ,i (λ, θi , φi ) cos θi dωi = Gλ (λ)
∩,i
Gλ,r (λ)
ρλ = (144)
G (λ)
R λ 0
ρ (λ, θi , φi )Iλ,i (λ, θi , φi ) cos θi dωi
∩,i λ
ρλ = R (145)
I (λ, θi , φi ) cos θi dωi
∩,i λ,i

Consequently, total hemispherical reflectivity is given by the following equations:


Gr
ρλ = (146)
RG ∞
0
ρλ (λ)Gλ,r (λ)dλ
ρλ = R∞ (147)
0
Gλ dλ

77
Radiosity
The radiation which leaves the surface consists of the radiation emitting by the surface and
reflecting by it. That is the spectral energy which leaves the surface dA in a direction (θ, φ) is
usually given as

d3 Qλ,tot (λ, θ, φ) = (Iλ,r (λ, θ, φ) + Iλ,e (λ, θ, φ)) cos θdω(θ, φ)dλdA (148)

For radiation which leaves the surface in all direction a term ”radiosity” is used. The spectral
radiosity is:
Z Z
3
Jλ (λ) = d Qλ,tot (λ, θ, φ)) = Iλ,r+e (λ, θ, φ) cos θdω(θ, φ)dλdA (149)
∩ ∩

The total radiosity is:


Z ∞ Z ∞ Z
0
J= Jλ (λ)dλ = dλ Iλ,r+e (λ, θ, φ) cos θdω(θ, φ)dA (150)
0 0 ∩

Special remark 1
If a surface is opaque, i.e. there is no transmission of the radiation, then the radiative energy
incident on the surface is either absorbed, or reflected, i.e.

d3 Q0λ,i (λ, θi , φi ) = d3 Q0λ,r (λ, θi , φi ) + d3 Q0λ,a (λ, θi , φi ) (151)


d Q0λ,r (λ, θi , φi )
3
d Q0λ,a (λ, θi , φi )
3
1= +
d3 Q0λ,i (λ, θi , φi ) d3 Q0λ,i (λ, θi , φi )
1 = ρ0λ (λ, θi , φi ) + a0λ (λ, θi , φi ) (152)

Using a Kirchhoff law for the spectral directional values, we can write:

1 = ρ0λ,a (λ, θi , φi ) + 0λ,a (λ, θi , φi , Ta ) (153)

1. It is possible to integrate the eq.151 over the spectrum that will lead to

1 = ρ0 (θi , φi ) + a0 (θi , φi , Ta ) (154)

0
If the surface is gray or if the itensity of the incident radation Iλ,i (λ, θ, φ) = C(θ, φ) · Iλ,i (λ, Ta )
a Kirchhoff law for total directional properties can be used:

1 = ρ0 (θi , φi ) + 0 (θi , φi , Ta ) (155)

2. If the incident radiation arrives from all direction, the eq.151 can be inegrated over the
hemisphere that will lead to

1 = ρ∩,λ (λ) + a∩,λ (λ) (156)

78
If the surface is diffuse or if the itensity of the incident radation does not depend on direction,
a Kirchhoff law for spectral hemispherical properties can be used:

1 = ρ∩,λ (λ) + ∩,λ (λ, Ta ) (157)

3. Finally, if the incident radiation arrives from all direction and the whole spectrum should
be taken into account, the eq.151 can be inegrated over the hemisphere and over the spectrul
that will lead to

1 = ρ + a∩ (158)

If the surface is diffuse and gray or if the intensity of the incident radiation is proportional to
the one from the black surface, a Kirchhoff law for total hemispherical properties can be used:

1 = ρ(Ta ) + ∩ (Ta ) (159)

In fig.56 the normal relfectivity and absorptivity are plotted for different materials.

Figure 56: Spectral normal reflectivity and absorptivity of selected materials, from the book
”Fundamentals of heat and mass transfer”, 6th edition, by Incropera, DeWitt, Bergman, Lavine.

Spectral and total transmissivity


The transmissivity is not a characteristic of the surface, but of a layer of a material.
This is a property that determines the fraction of the incident radiation propagated through
the layer. If the layer is not very thick the following simplified parameters may be introduced
in order to characterize the ability of the object to let the radiation pass through it.

79
Definition 33. Hemispherical spectral transmissivity is defined as a ratio of the energy at a
certain wavelength propagated through the surface (thin layer) to the energy incident onto the
surface (layer) from all directions at this wavelength.

d2 Q∩,λ,tr (λ)
τλ = (160)
d2 Q∩,λ,i (λ)

Spectral transmissivities τλ for some materials are presented in Fig.57. Note, that for each
material its thickness is indicated.

Figure 57: Spectral transmissivities τλ of selected materials, from the book ”Fundamentals of
heat and mass transfer”, 6th edition, by Incropera, DeWitt, Bergman, Lavine.

Definition 34. Hemispherical total transmissivity is defined as a ratio of the energy over the
whole spectrum propagated through the surface (thin layer) to the energy incident onto the
surface (layer) from all directions over the whole spectrum.

dQ∩,tr
τ=
dQ
R ∞ ∩,i
τλ (λ)d2 Q∩,λ,tr dλ
τ = 0 R∞ 2 (161)
0
d Q∩,λ,i

Note that glass or water transparent in the visible spectrum are opaque at longer wavelengths.
The transmissivity of plastic (Tedlar) in the infrared region is higher than the transmissivity
of the glass.

80
Example 4. The cover galss of on a flat-plate solar collector has a low iron content and its
spectral transmissivity may be approximated by the distribution presented in the fig.58. What is
the total transmissivity of the cover glass to solar radiation? Note that the spectral distribution
of solar irradiation is proportional to that of the blackbody emission at 5800K.

To answer the question, use the eq.161 using the fact that d2 Q∩,λ,i = f (T ) ∗ Eb,λ (T ).

Figure 58: Spectral transmissivities τλ (λ) of the cover glass for the example 4

Special remark 2
It should be noted that if the surface under consideration is not opaque than the relations pre-
sented in the section Special remarks 1 should be updated because the transmitted radiation
has to be taken into account, see fig.59.

Figure 59: Energy balance for a thin semitransparent layer, from the book ”Fundamentals of
heat and mass transfer”, 6th edition, by Incropera, DeWitt, Bergman, Lavine.

With definitions given above we cannot update the eq.154 for spectral directional balance
since there is no notion for the spectral directional transmissivity. But we can re-write eqs.156
and 158 in the forms, respectively:

81
1 = ρ∩,λ (λ) + a∩,λ (λ) + τ∩,λ (λ) (162)
1 = ρ∩ + a∩ + τ∩ (163)

82
Chapter 8. Radiative heat exchange between black isother-
mal surfaces
Differential view factors
Consider two differential isothermal surfaces as shown in fig.60. The surfaces dAi and dAj are
isothermal at temperatures Ti and Tj , respectively, arbitrarily oriented, and have their normals
at angles θi and θj to the line of length R joining them.

Figure 60: Radiative interchange between two differential area elements.

The total energy per unit time leaving dAi and incident upon dAj is d2 QdAi→dAj = Ii cos θi dAi dωij ,
where dωij , is the solid angle subtended by dAj when viewed from dAi : dωij = dAj cos θj /R2 .
So,
cos θi dAi cos θj dAj
d2 QdAi→dAj = Ii (164)
R2
A similar expression can be written for the total energy per unit time leaving dAj and incident
upon dAi :
cos θi dAi cos θj dAj
d2 QdAj→dAi = Ij
R2
To make the next step, suppose that the surfaces dAi and dAj are black. Then intensities
Ii = σTi4 /π and Ij = σTj4 /π

Definition 35. A fraction of energy which leaves the black surface element dAi and arrives at
black element dAj is defined as the differential geometric configuration factor or view factor
dFdAi−dAj .

According to the definition, differential view factor dFdAi−dAj is

d2 QdAi→dAj d2 QdAi→dAj
dFdAi−dAj = =
dQe σTi4 dAi
cos θi cos θj dAj
dFdAi−dAj = (165)
πR2

83
Note that the differential view factor dFdAj−dAi is
cos θi cos θj dAi
dFdAj−dAi = (166)
πR2
Comparing equations 165 and 166 we obtain the first relation of the algebra of view factors
or view factors reciprocity law:
dAi · dFdAi−dAj = dAj · dFdAj−dAi (167)
If no other surfaces are included in the system (no other sources of radiation), the resutling
radiative flux for the surface dAi is
cos θi dAi cos θj dAj
d2 QdAi,res = d2 QdAi→dAj − d2 QdAj→dAi = (Ii − Ij )
R2
d2 QdAi,res = σ(Ti4 − Tj4 )dFdAi−dAj dAi (168)
Note that the resulting flux for the second surface is
cos θi dAi cos θj dAj
d2 QdAj,res = (Ij − Ii )
R2
d2 QdAj,res = σ(Tj4 − Ti4 )dFdAj−dAi dAj

View factors between a differential surface and a finite one


It is possible to consider the radiative heat exchange between an elementary surface dAi and a
finite surface Aj as it is shown in fig.61.
As in previous case, we can consider a small (differential) part dAj of the surface Aj . Then

Figure 61: Radiative interchange between a differential element and a finite area.

the energy which leaves the surface dAi and arrives to the subsurface dAj will be given by the
eq.164. To obtain the energy which leaves the surface dAi and arrives to the surface Aj the
integration over the latter should be performed:
Z
cos θi dAi cos θj dAj
dQdAi→Aj = Ii
Aj R2
Z
cos θi cos θj dAj
dQdAi→Aj = σTi4 dAi (169)
Aj R2

84
Similarly to the definition 35, the definition for a view factor between a differential element dAi
and a finite surface Aj can be obtained:
dQdAi→Aj
dFdAi−Aj =
σT 4 dAi
Z i
cos θi cos θj dAj
dFdAi−Aj = (170)
Aj πR2

The second view factor is


dQAj→dAi
dFAj−dAi = R
Aj
σTj4 dAj
Z
dAi cos θi θj dAj
dFAj−dAi = (171)
Aj Aj πR2

Comparison fo the eqs.170 and 171 shows that the reciprocity law is valid:

dAi · dFdAi−Aj = Aj · dFAj−dAi (172)

View factors for finite surfaces


Similar consideration for finite surfaces (as shown in fig.62) gives view factors in this case as:

Figure 62: Radiative interchange between two finite surfaces.

Z Z
1 cos θi cos θj
FAi−Aj = dAi dAj (173)
Ai Ai Aj πR2
Z Z
1 cos θi cos θj
FAj−Ai = dAi dAj (174)
Aj Ai Aj πR2
(175)

Evidently, the reciprocity law is valid in this case as well:

Ai · FAi−Aj = Aj · FAj−Ai (176)

85
Algebra of view factors
Algebra of view factors includes (1)definitions, (2) reciprocity laws and (3) a law of additivity
wich is illustrated with a fig.63

Figure 63: Radiative interchange between two finite surfaces.

It follows from the definition of the view factor that fot the surface shown in fig.63

FA1−A2 = FA1−A21 + FA1−A22 (177)

Example 5. Let for the fig.63 the FA1−A2 and FA1−A22 are known. Find FA21−A1

86
Chapter 9. Radiative heat exchange in a black enclosure
In fig.64 a cross-section of an enclosure is shown. The heat balance equation for a surface k

Figure 64: Enclosure composed of N black isothermal surfaces (shown in cross section for
simplicity).

can be written in a form of the energy which leaves the surface k and the energy which arrives
from all other surfaces:
N
X
Qk,res = σ · Tk4 · Ak − σ · Tj4 · Fj−k · Aj (178)
j=1

Using a reciprocal relation Aj Fj−k = Ak Fk−j and a sumation rule N


P
j=1 Fkj = 1 ∀ k, the eq.178
can be re-written in a form of the energy exchange between the surface k with every other
surface j:
N
X
Tk4 − Tj4 · Fk−j

Qk,res = σ · Ak (179)
j=1

In an enclosure which consists of N surfaces there are Q1 , . . . , QN resulting heat fluxes and
T1 , . . . , TN temperatures, that gives 2N variables. The problem of the heat transfer lies in find-
ing of either the temperature or the heat flux for a surfaces if other heat fluxes and temperatures
are known. Let us look for example at the next problem:

Example 6. A row of regularly spaced, cylindrical heating elements is used to maintain an


insulated furnace wall 1 at Tw1 = 500K. The opposite wall, wall 2, is at uniform temperature
of Tw2 = 300K. The insulated wall experiences convection with air at Tair = 450K and a
convection coefficient of 200W/(m2 · K). Assuming the walls and elements are black, estimate
the required operating temperature for the elements Tel . See illustration to the problem at fig.65.

The system presented in fig.65 can be considered as an enclosure which consists of 3 ”effective
surfaces” characterised with appropriate view factors Fw1−el , Fw1−w2 , Fel−w1 , Fel−w2 , Fw2−el ,

87
Figure 65: Illustration to the example 6. Cylindrical elements heat the lower surface which is
insulated.

Fw2−w1 . For each of these three surfaces the heat balance equation can be written and these
surfaces has a temperature Tw1 , Tw2 or Tel , respectively. That is there are 3 resulting heat
fluxes and 3 temperatures in the system, i.e. 6 variables. Three of them are known: Tw1 , Tw2
and Qw1 = 0. Therefore, all others can be found from 3 equations of the heat balance. Since
the question if the problem is related only to the temperature of the heating elements, it is
convinient to write the heat balance for the wall 1.
Example 7. An enclosure of triangular cross section is made up of three plane plates each of
finite width and infinite length (thus forming an infinitely long triangular prism). The surfaces
are maintained at temperatures T1 , T2 , and T3 , respectively. Determine the amount of energy
that must be supplied to each surface per unit time in order to maintain these temperatures.
Note that this amount of energy is also the net radiative loss from each surface.
Equation 179 is written for each surface as:

Q1,res = σ · A1 T14 − T24 · F1−2 + σ · A1 T14 − T34 · F1−3


 

Q2,res = σ · A2 T24 − T14 · F2−1 + σ · A2 T24 − T34 · F2−3


 

Q3,res = σ · A3 T34 − T14 · F2−1 + σ · A3 T34 − T24 · F3−2


 

It can be verified that the total heat flux in the enclosure is

Q1,res + Q2,res + Q3,res = 0

Example 8. The enclosure of example 7 has two of its sides maintained at temperatures T1
and T2 , respectively. The third side is an insulated (adiabatic) surface, Q3 = 0. Determine Q1 ,
Q2 , and T3 .
Again equation 179 can be written for each surface as

Q1,res = σ · A1 T14 − T24 · F1−2 + σ · A1 T14 − T34 · F1−3


 

Q2,res = σ · A2 T24 − T14 · F2−1 + σ · A2 T24 − T34 · F2−3


 

0 = σ · A3 T34 − T14 · F2−1 + σ · A3 T34 − T24 · F3−2


 

From the last equation the unknown temperature T3 can be found. After that the heat fluxes
Q1 , Q2 can be calculated.

88
Chapter 10. RHT in an Enclosure Composed of Diffuse-
Gray Surfaces
A black surface is a too strong assumption for a real surface. In practice, the enclosure shown
in fig.64 can have non-black internal surfaces. The next approximation which provides rather
simple solution is an enclosure composed of diffuse-gray surfaces.

Because a gray surface is not a perfect absorber, a part of the energy incident on a surface
is reflected. The total energy which leaves a gray surface k, Qout,k consists of the emitting
energy Qe,k and a reflected energy Qr,k . With regard to the reflected energy, two assumptions
are made:
1 – the reflected energy is diffuse, that is, the reflected intensity at each position on the bound-
ary is uniform for all directions
2 – the reflected energy is uniform over each surface of the enclosure.

To remind: By definition, when a surface is diffuse-gray, the directional spectral emissivity


and absorptivity do not depend on either angle or wavelength. As a result of this definition,
the hemispherical total absorptivity and emissivity are equal and depend only on TA , that is,
∩ = a∩ .
Moreover, since the reflected energy is also distributed uniformly everywhere, a hemispher-
ical diffuse reflectivity ρ∩ can be considered and

ρ∩ = 1 − a∩ = 1 − ∩ (180)

Also, the view factors F which were derived for black surfaces can be used for the present
enclosure theory.

REMARK. Most of the problems encountered in practice are at steady state. However, the
radiative heat balances considered here are not limited to steady-state conditions.

Radiative heat balance equations can be written in different ways, however, the starting point
is the same: resulting heat fux which is a difference between the energy which leaves a surface
and which arrives to it:

Qres,k = Qout,k − Qinc,k = Ak · qout,k − Ak · qinc,k (181)

where Qout,k is the energy which leaves a gray surface k and Qinc,k is the energy which is incident
on the surface k. Then:

Qout,k = Qe,k + Qr,k = Qe,k + ρk · Qinc,k = k · Qeb,k + (1 − k ) · Qinc,k


Ak · qout,k = Ak · k · qeb,k + Ak · (1 − k ) · qinc,k

89
and accounting for Planck’s relation:

Ak · qout,k = Ak · k · σTk4 + Ak · (1 − k ) · qinc,k (182)

Equations 181 and 182 contain too many unknowns: N · Qres , N · Qout , N · Qinc and N · T ,
the system of equations can be simplified. But there are different ways to simplify it. Let us
consider the first way which was proposed by Poljak.
What is the energy flux incident on the surface k? From the one hand using the eq.182:

qout,k − k · σTk4
qinc,k = (183)
(1 − k )

On the other hand,

Ak · qinc,k = A1 · qout,1 F1−k + A2 · qout,2 F2−k + . . .


+ Aj · qout,j Fj−k + . . . + Ak · qout,k Fk−k + . . . + AN · qout,N FN −k (184)

Using reciprocal relations, the eq.184 can be re-written:

Ak · qinc,k = Ak · qout,1 Fk−1 + Ak · qout,2 Fk−2 + . . .


+ Ak · qout,j Fk−j + . . . + Ak · qout,k Fk−k + . . . + Ak · qout,N Fk−N
N
X
qinc,k = Fk−j qout,j (185)
j=1

Therefore, from eq.182 and 185 it follows that


N
X
qout,k = k · σTk4 + (1.0 − k ) Fk−j qout,j (186)
j=1

Coming back to the eq.181, the resulting heat flux can be written as:
N
!
X
Qres,k = Qout,k − Qinc,k = Ak qout,k − Fk−j qout,j (187)
j=1

or, taking into account eq.183 as

qout,k − k · σTk4
 
Qres,k = Ak qout,k −
1 − k
(1 − k )qout,k − qout,k + σk Tk4
Qres,k = Ak ·
1 − k
k 4

Qres,k = Ak · σTk − qout,k (188)
1 − k
The equations 187 and 188 is a required system of equations written for non-black surfaces
(!). In that way for each surface there are 2 equations with 3 unknowns: Qres,k , qout,k and Tk ,
that are 3N variables. If N variables are known, the system of equations can be resolved.

90
Example 9. As an example, let us consider 2 infinite surfaces characterised with 1 and 2 and
having the temperatures T1 and T2 . Find the resulting heat fluxes for these surfaces.

There are 6 variables in the equations, 2 are known + 2 equations for each of 2 surfaces, i.e.
4 equations and 2 temperatures are known. Therefore, the system of equations can be solved.
Let us write down the system. The solution is not given here. Please see your own
notes for details.

Example 10. Derive an expression for the radiation exchange between two concentric diffuse-
gray spheres at uniform temperatures as shown in figure 66

Figure 66: Radiative interchange between two gray spheres

The solution is not given here. Please see your own notes for details.

Resulting heat fluxes and temperatures in a diffuse-gray enclosure


Note that in fact in the system consisting of eq.187 and 188 the heat flux qout,k is not important.
The system can be re-written without this heat flux. The eq.188 is used in order to find qout,k
and subsitute the result in eq.187. As a result, one can obtain a general equation for the surface
k:
Q1 1 − 1 Q2 1 − 2
− Fk−1 − Fk−2 − ...
A1 1 A2 2
 
Qk 1 1 − k QN 1 − N
+ − Fk−k − ... − Fk−N =
Ak k k AN N
= −Fk−1 σT14 − Fk−2 σT24 − . . . + 1 − Fk−k σTk4 − . . . − Fk−N σTN4

(189)

This equation can be re-written in a compact form using the Kronecker delta defined as:

1 if i = j
δij = (190)
0 if i 6= j

91
Then the eq.189 becomes:
N   N
X δkj 1 − j Qj X
− Fk−j = (δkj − Fk−j ) σTj4 (191)
j=1
j j Aj j=1

Graphical representation of RHT in a diffuse-gray enclosure


Generally, using the system of equations 187 and 188 it is possible to solve the problem of the
radiative heat exchange for any enclosure for N variables ( resulting heat fluxes or temperatures)
if N other are known. Sometimes utilisation of eq.189 is easier. The problem which arises in
analytical solution (without computers) that it is necessary to solve a system of 2N equations
(i.e. for 3 surfaces it gives 6 equations) that is not always easy. Therefore, in some publications
a graphical representation of a RHT in diffuse-gray enclosure is provided.
To use this representation and to make our equations shorter we will us following notations,
some of which you already know:
·Gk - irradiation of the surface k, i.e. the radiation (energy) incident onto surface from ev-
erywhere per unit surface per second;
·Ek,b = σTk4 - hemispherical emissive power power of a black surface k;
·Jk = Ak · qout,k = Ak · (k σTk4 + ρk · qinc,k ) - radiosity, i.e the radiation (energy) which leaves
the unit surface per unit of time.

The graphical representation of the energy incident onto the surface k (or i) and the energy
leaving the surface k (or i) is given in fig.67(b,c).
Then the eq.188 for a surface k can be re-written in another way:

(σTi4 − qout,i ) (Ei,b − Ji )


Qres,i = = (192)
(1 − i ) / (i Ai ) (1 − i ) / (i Ai )

To be consistent with the figures, the index k used in eq.181–188 is changed to the index i.
The eq.192 provides a convinient interpretation for the net radiative heat transfer from a sur-
face. This transfer (see fig.192) is associated with a driving potential (Eb,i − Ji ) and a surfaces
radiative resistance of the form (1 − i )/(i Ai ). That is a surface is like a ”sandwich” with a
”resistance inside”. Hence, if the emissive power of a black surface i at temperature Ti exceeds
radiosity of the surface i, there is a net (resulting) radiation heat transfer from the surface; if
the inverse is true, the net (resulting) radiation transfer is to the surface.

REMARK A If i = 1 then there is no resistance and ”potentials” Ji and Eb,i are equal!
REMARK B Sometimes one of the surfaces can be very large relative to the other surfaces
(e.g. small surfaces in a large room). In this case it is possible to say that Ai → ∞, thereforen
the surface resistance (1 − i )/(i Ai ) → 0. Situation is similar to a black surface (considered in
the remark A) when i = 1. Hence, Ji = Eb,i and a surface which is large relative to all other

92
Figure 67: Radiation exchange in an enclosure consisting of diffuse gray surfaces. (a) –
Schematic of the enclosure; (b) Radiative balance for a surface i with irradiation and ra-
diosity; (c) Radiative balance using emitted, incident and reflected energy (d)Network element
representing the net radiation transfer from a surface according to eq.192.

surfaces under consideration can be treated as if it were a blackbody.


Note that the eq.187 can be also re-written in terms of radiosities:
N N
X X Ji − Jj
Qres,i = Ai (Fi−j qout,i − Fi−j qout,j ) = (193)
j=1 j=1
(Ai Fi−j )−1

Finally, if we combine eqs.192 and 197, we obtain:


N
Ei,b − Ji X (Ji − Jj )
= (194)
(1 − i ) /i Ai j=1
(Ai Fij )−1

The eq.197 and eq.198 are presented in the fig.68 This figure (as well as equations) represents
a radiation balance for the radiosity node associated with the surface i. The rate of radiation
transfer (flow of the current) to the node i through its surface resistance (1 − i ) /(i Ai ) must
equal the rate of radiation transfer (currents flow) from i to all other surfaces through the
corresponding geometrical resistances Ai Fij .

REMARK. Note that the network can include parallel curcuit and/or series circuits,
similar to electrical circuits.

93
Figure 68: Radiation exchange for the node i
.

Continuation of graphical representation of RHT in a diffuse-gray


enclosure
In eq.192 a misprint was made. Please, correct eq.192 in the following way:
Ei,b − Ji
Qi = (195)
(1 − i ) /i Ai

To use the eq.195 for the solution of the heat exchange problem, the knowledge of the radiosity
Ji is necessary. It can be defined by consideration of the heat exchnage between the surface
i with all other surfaces. It can be written that irradiation Gi of the surface i is a sum of
radiosities J from all surfaces taken with appropriate view factors Fj−i :
N
X
Ai Gi = Fji Aj Jj
j=1

Using the reciprocity relation:


N
X
Ai Gi = Fij Ai Jj
j=1

94
Using eq.?? Qi = Ai Ji − Ai Gi , i.e. Ai Gi = Ai Ji − Qi :
N
X
Ai Ji − Qi = Fij Ai Jj
j=1
N
!
X
Qi = Ai Ji − Fij Jj
j=1

PN
From the summation rule j=1 Fij = 1, the last equation can be written in the form:

N N
! N N
X X X X
Qi = Ai Fij Ji − Fij Jj = Ai Fij (Ji − Jj ) = qij (196)
j=1 j=1 j=1 j=1

This result equates the net radiative flux Qi from the surface i to the sum of radiative heat
exchange fluxes qij between the surface i with all other surfaces. It can be rewritten in the
following form:
N
X (Ji − Jj )
Qi = −1 (197)
j=1
(A i F ij )
It is preferable to use the equation of the radiative heat exchange between a par-
ticluar surface i with other surfaces in the form of eq.197 if the net heat flux Qi is
known for the surface.
It is not always the case and the temperature Ti of the surface i is known. Then, making a
combination of the eq.195 and the eq.196, obtain:
N
Ei,b − Ji X (Ji − Jj )
= (198)
(1 − i ) /i Ai j=1
(Ai Fij )−1

where Ei,b is known since the Ti is known.


This figure (as well as equations) represents a radiation balance for the radiosity node as-
sociated with the surface i. The rate of radiation transfer (flow of the current) to the node i
through its surface resistance (1 − i ) /(i Ai ) must equal the rate of radiation transfer (currents
flow) from i to all other surfaces through the corresponding geometrical resistances Ai Fij .

95
This figure (as well as equations) represents a radiation balance for the radiosity node asso-
ciated with the surface i. The rate of radiation transfer (flow of the current) to the node i
through its surface resistance (1 − i ) /(i Ai ) must equal the rate of radiation transfer (currents
flow) from i to all other surfaces through the corresponding geometrical resistances Ai Fij .

Example 11. Let us consider an eclosure that involves only 2 surfaces as it is shown in fig.69a.
Note that this enclosure is similar to the case of 2 infinite plates (that was considered last time).

The enclosure can be presented using network approach as it is shown in fig.69b. This network
represents two radiative nodes (marked as ”surface 1” and ”surface 2”) with a resistance
A1 F12 = A2 F21 between them. Note, that every node is not a simple ”point” but it looks like a
sandwich (see also fig.67), which consists of an ”internal rezistance” (1 − i ) /(i Ai ), of driving
potential Eb,i − Ji and of resulting ”current” Qi .

Figure 69: illustration to the example 11. (a)A two-surfaces enclosure and (b) network repre-
sentation of the RHT in enclosure.

However, despite this complicated structure it can be seen that the driving potential between
the two surfaces is U12 = Eb,1 − Eb,2 and the total resistance of the network is the sum of all
resistances connected in series:
(1 − 1 ) 1 (1 − 2 )
Rtot = + + (199)
1 A1 A1 F12 2 A2

96
Therefore, the resulting ”current”, i.e. the resulting flux Q12 is
U12
I12 = (200)
Rtot
Eb,1 − Eb,2
Qres,1 =
(1 − 1 ) 1 (1 − 2 )
+ +
1 A1 A1 F12 2 A2
4 4
σ (T1 − T2 )
Qres,1 = (201)
(1 − 1 ) 1 (1 − 2 )
+ +
1 A1 A1 F12 2 A2
Note that result given by eq.201 is consistent with results obtained during the last lecture for
the two infinite surfaces.

Network representation of radiation shields


Radiation shields constructed from low emissivity (high reflectivity) materials can be used to
reduce the net radiation transfer between two surfaces.

Example 12. For example, consider placing a radiation shield (surface 3) between the two
large parallel planes as shown in fig.70.

Figure 70: Illustration to the example 12

Similarly to the case of two-surfaces enclosure, this configuration can be presented using network
approach as it is shown in fig.70 Using the same idea of driving potential, current and resistance
(eq.200), and taking into account that Fij = 1 ∀ i, j obtain:

Qres,12 σ (T14 − T24 )


qres,12 = = (202)
A 1 (1 − 3,1 ) (1 − 3,2 ) 1
+ + +
1 3,1 3,2 2

97
Figure 71: Network representation for the example 12

If the temperatures of the surfaces T1 and T2 are known, the eq.202 provides the values of
the resulting heat flux between the surfaces 1 and 2 but this flux is equal (up to sign) to all
resulting heat fluxes in the system:

qres,12 = qres,13 = qres,32 = −qres,21 = −qres,23 (203)

Taking the last equation into account the temperature of the shield T3 can be found.

Example 13. In manufacturing the special coating on a curved solar absorber surface of area
A2 = 15m2 is cured by exposing it to an infrared heater of width W=1m. The absorber and the
heater are each of length L = 10m and are separated by a distance of H = 1m. There is no
other heat exchange processes in the system.
The heater is at T1 = 1000K and has an emissivity of 1 = 0.9, while the absorber is at
T2 = 600K and has an emissivity of 1 = 0.5. The system is in a large room whose walls are
at T3 = 300K.
What is the net rate of heat transfer to the absorber surface? The scheme of the problem is
given in fig.72

Figure 72: Illustration to the example 13

To solve the problem, the following assumptions must be made:


1. Steady state conditions
2. Surfaces of absorber and heater are diffuse and gray
3. The wall of the room can be considered as black

98
The radiation network is constructed by first identifying nodes associated with the radiosi-
ties of each surface as shown in step 1, fig.73. Then each radiosity node is connected to each of
the other radiosity nodes through the appropriate space resistance as shown in step 2, fig.73.
Difficulty: the surface of the room wall is not known. However, it can be replaced with the use
of the reciprocity law: A3 F31 = A1 F13 and A3 F32 = A2 F23 .
The last step is to connect the ”internal resitances” and ”driving potentials” to each of the
node as it is shown in fig.73, step 3.

Figure 73: Network representation for the example 13

Note that for the surface 3 3 = 1, therefore, J3 = Eb,3 = σT34 .


Summation of all ”currents” (i.e. heat fluxes) in the node 1 gives zero:
Eb,1 − J1 J1 − J2 J1 − J3
= + (204)
(1 − 1 )/1 A1 1/(A1 F12 ) 1/(A1 F13 )
Similarly for the node 2:
Eb,2 − J2 J2 − J1 J2 − J3
= + (205)
(1 − 2 )/2 A2 1/(A2 F21 ) 1/(A2 F23 )
Form the equations 204 and 205 radiosities J1 and J2 can be found, the only difficulty is to
define all view factors in the system. The way of the solution is as follows:
1. To simplify the problem, the surface of the absorber A2 can be replaced with an ”effective’
surface A02 as shown in fig.74. Then, the view factor F12 = F120
2. The view factor F120 can be found from the figures (or schemes) that you have for view
factors
3. From the summation rules F11 + F120 + F13 = 1 and F11 = 0, i.e. F13 = 1 − F120
4. The view factor F23 is required. In order to find it first, we use a reciprocal law: A2 F23 =
A20 F20 3 , and from the symmetry F20 3 = F13 . Therefore, F23 = A20 F20 3 /A2
With known F23 , F12 = F120 and F13 the system of equations 204 and 205 an be solved and the
final answer can be obtained.

99
Figure 74: Illustration to the example 13

100
Chapter 11. Solar radiation. Environmental radiation.
Solar radiation is essential for the life on Earth. The Sun is the star at the center of the Solar
System. It consists of hot plasma, i.e. ionized gas at high temperature. These gases are mostly
hydrogen (75%) and helium while the rest (about 1.7%) are heavier elements includin oxygen,
carbon, neon, iron and some others.

The Sun is almost spherical with a diameter of about 1392684km (109 times that of Earth) and
its mass is ≈ 1.99 · 1030 kg that is about 330000 times the mass of Earth. The mass of the Sun
accounts for about 99.86% of the total mass of the Solar System. A mean distance between the
Sun and the Earth is 1.496 · 108 km. That is the solid angle at which the Sun is seen from the
Earth is
πDS2 πRS2
ΩS = = (206)
4 · RS2 E RS2 E
3.14 · (1.39 · 109 )2
ΩS = 11 2
= 0.67 · 10−4 sr
4 · (1.496 · 10 )

Since the solid angle is very small, is is generally assumed that solar radiation comes from a
single direction, i.e., that all the light beams are parallel.

Example 14. The energy which is emitted by the Sun is generated by nuclear fusion of hydrogen
nuclei into helium. The surface temperature of the Sun is approximately 5778K. Let us assume
that the Sun is a black body and calculate the total solar heat flux incident on the collector (per
unit area) which is mounted on a satellite orbiting Earth and directed at the sun (i.e., normal
to the sun’s rays) as shown in fig.75a.

Figure 75: Solar radiation incident onto a solar collector in outer space

The total heat rate leaving the sun is Qsol = 4·πRS2 ·Eb (TS ). If we construct a sphere around the
Sun in that way that a satellite is on its surface, then the area of this sphere is Ssat = 4πRS2 E .
Note, that a geostationary orbit for satellites is at 36000 km which is negligible compared with

101
the distance to the Sun. Then, the heat flux which leaves the Sun and pass through a unit area
of this sphere is

Qsol 4 · πRS2 · Eb (TS ) Eb (TS ) πRS2 Eb (TS )


qse = = 2
= 2
= ΩS
Ssat 4πRSE π RSE π
σTS4 W
qse = ΩS = 1.8057 · (57.78)4 · 6.77 · 10−5 = 1348 2 (207)
3.14 m
Note that if the collector is not perpendicular to the solar rays, but parallel to the Earth’s surface
(as in fig.75a) the total flux GS,0 (extraterrestrial solar irradiation) may be calculated as

GS,0 = qse · f · cos(θ) (208)

where f is a small correction factor which account for the eccentricity pf the arth’s orbit about
the Sun 0.97 ≤ f ≤ 1.03.
The calculated value qse ≈ 1340W/m2 is called solar constant. The most accurate measure-
ments and spectral integration of extraterrestrial solar irradiation data have been made during
the past twenty years by Earth-orbiting satellites (F.Modest ”Radiative transfer”) resulting in
(1367 ± 2)W/m2 .

If Sun is a black body?


As it is seen in fig.76 the spectrum of extraterrestrial radiation approximates black body radi-
ation with temperature between T = 5800K and T = 5770K.

Figure 76: Spectrum of solar irradiation

102
Scattering in the atmosphere
Solar radiation is attenuated significantly as it penetrates through the atmosphere. The change
is due to absorption and scattering of the radiation by the gases in the atmosphere. In the
UV (ultraviolet) spectrum ozone O3 absorbs almost all radiation with λ < 0.3µm and some
part of radiatio in 0.3µm < λ < 0.4µm. In the visible spectrum some absorption occurs due
to O3 and O2 and in the near and far IR (infrared) region absorption is dominated by water
vapor.
Atmospheric scattering consists (mainly) of two kinds shown below

Figure 77: Rayleigh and Mie scattering

Rayleigh scattering

Rayleigh (molecular) scattering by very small gas molecules occurs when the ratio of the effective
molecule diameter D to the wavelength of radiation λ is much less then unity: πD/λ << 1.
This condition provides nearly uniform scattering of radiation in all directions (fig.77a), hence,
about a half of the scattered radiation is redirected to space while the remaining portion strikes
the earth’s surface. After Rayleigh scattering the scattered radiation continues to propagate in
all directions.
Rayleigh scattering is more effective with shorter waves an therefore it provides a blue color of
the sky in daytime.

Mie scattering

Mie scattering occurs due to larger dust and aerosol particles in the atmosphere when πD/λ ≈ 1
and direction of the radiation changes only a little as shown in fig.77b.
The cumulative effect of the scqattering processes of the solar radiation is shown in fig.78.
The radiation which arrives to the observer (surface) consists of a part of direct radiation (or
lightly scattered) and of scattered radiation which is often supposed to be diffuse. The diffuse
contribution may vary from 10% of the total solar radiation on a clear day to nearly 100% on
a totally cloudy day.
Scattering may also happen on particles with a characteristic size significantly large than a
wavelength in a visible range of the spectrum. This leads for various optical phenomena as,

103
Figure 78: Cumulative scattering effect

for example, rainbow. A brief descritpion of rainbow’s formation is given in the end of the
chapter.

Emission from the Earth


Radiation from the Sun is concentrated in short wavelength range of the spectrum. The long
wavelength form of environmental radiation include emission from the Earth’s surface as well as
from atmospheric elemnts. The emissive power from the Earth’s surface may be approximated
as
4
Eearth = εσTearth (209)

where ε ≈ 0.97 and Tearth ≈ 250 − 320K. That is the peak of the emission from the Earth’s
surfaces is about 10µm and most of radiation in concentrated within the range 4µm to 40µm.
As it was mentioned above in that region the most absorption in the atmosphere occurs due to
CO2 and H2 O molecules which then re-emit this energy. This emission does not similar to
that of the blackbody but its contribution to irradiation of the earth’s surface can be estimated
as
4
Gatm = σTsky (210)

where Tsky is an ”effective sky temperature” which varies from 230K to 280K.
REMARK 2 Absorption by CO2 and H2 O of radiation emitted by the Earth and its re-
emission leads to w well-known ”greenhouse effect”.
REMARK 2 If a problem includes a solar radiation incident onto a surface (on the Earth)
and emission from that surface in most cases the surface cannot be considered as a gray one!

Example 15. A flat-plate solar collector has a selective abosrber surface of εa = 0.1 and
solar absorptivity aS = 0.95. The absorber temperature is Ta = 1200 C, the colar irradiation is
750W/m2 , the effective sky temperature is −100 C and the air temperature Tair = 300 C. Take

104
the heat transfer convective coefficient h = 1W/m2 (see fig.79).
Question What is useful heat removal rate (W/m2 ) from the collector for these conditions?
What is the efficiency of the collector?

Figure 79: Cumulative scattering effect

The heat balance equation per unit area is:

aS · Gs + asky · Gsky − qconv − Ee − qu = 0 (211)

where Ee is emission from the surface and qu is a useful heat removal rate. The emission ”from
the sky” is
4
Gsky = σTsky (212)

The solar absorption coefficent aS is given, but the absorption coefficient asky should not be
taken equal to aS because the temperature of the surface differs significantly from the ”effective
temperature” of the solar radiation. According to Kirgchoff law, asky ≈ εa !
Combining everything,

4
qu = aS · Gs + εa σ · Tsky − h(Ta − Tair ) − εa σTa4 = 516W/m2 (213)

Efficiency of the collector is


qu 516
η= = = 0.69 (214)
Gs 750

Seasons change

Season variation of temperature occurs due to the inclination of the Earth’s axis with respect
to its rotational orbit around the Sun (fig.80

105
Figure 80: Inclination of the Earth’s axis which leads to the change of the seasons

106
Chapter 12. Heat exchange with absorption and emission
in the gas volume
Attenuation of the radiation in a participating medium
As it was discussed starting from the beginning of the course, intensity is the main chrac-
teristics of the energy flux in the radiative heat transfer. We will consider a radiation at a
wavelength λ from an elementary surface dF into a volume filled with a participating media
whose index of refraction is 1 (i.e. this is a gas)(fig.81). Let us find how intensity of the radia-
tion varies along the path of the ray beacuse of absorption and scattering on the gas molecules.
First, we assume that the temperature of the gas is low, i.e. emission of the gas is negligible.
Let s is the path length, I0 (λ) is a known intensity at a certain s0 .

Figure 81: Emission into a gaseous media

Intensity decrease because of radiation propagation through a layer of the thickness dS is


proportional to the intensity and to dS:

dIλ = −Kλ · Iλ dS (215)

This equation is written with the assumption that no intensity is scattered from the radiation
field into the direction of S. The coefficient Kλ is called the extinction coeficient of the material.
The extinction coefficient is a physical property of the material and has the units of reciprocal
length. For gases, it is a function of the temperature T, pressure P, composition of the material
(specified here in terms of the concentration Ci of the i component), and the wavelength of
the incident radiation so that Kλ = Kλ (λ, T, P, Ci ). Since it is directly proportional to the
pressure, it can be written also as Kλ = κλ · p where p is the (partial) pressure of the gas.

107
Integrating equation 215 over a length S of the ray’s path, obtain
Z Iλ (L) Z L
dIλ
=− Kλ dS ∗
Iλ (0) Iλ 0
  Z L
Iλ (L)
ln =− Kλ dS ∗
Iλ (0) 0
 Z L 

Iλ (L) = Iλ (0) exp − Kλ dS (216)
0

Equation 216 is known as Bouguer’s law. It shows that the intensity of monochromatic radiation
along a path is attenuated exponentially while passing through an absorbing-scattering medium.
The eq.216 can be re-written as eq.220 known as Beer’s equation:

Iλ (S) = Iλ (0)e−κλ (L) (217)

where the κλ (L) is called optical thickness of the gas and in case that Kλ does not vary with
the path, κλ (L) = Kλ · L, L is a characteristic size of the volume.
If the optical thickness of the gas is high, attenuation of the radiation because of its propagation
through the gas will be significant. In case of a small opticla thickness, it is possible to neglect
the presence of the gas in the volume.

The coefficient of the radiation attenuation (extinction coefficient) Kλ is equal to the sum of
two coefficients: absorption coefficient aλ (λ, T, P ) and scattering coefficient σλ (λ, T, P ):

Kλ = aλ (λ, T, P ) + σλ (λ, T, P ) (218)

In further consideration we shall neglect the scattering. Then, using Beer’s (or Bouger’s)
equation, an absorptivity for a volume of gas of thickness L can be obtained as:
Iλ,L − Iλ,0
ag,λ = − = 1 − e−κλ (L) = 1 − e−aλ L (219)
Iλ,0
and transmissivity for a volume of gas of thickness L is given as
Iλ,L
τg,λ = = e−κλ (L) = e−aλ L (220)
Iλ,0
It should be noted that absoprtion in gases is (highly) selective, i.e. it is big in some
spectral zones, ”spectral bands” around certain wavelengths. The most important gases in
engineering practice are water vapor and carbon dioxide since they are product of combustion
and other chemical reactions. In fig.82 and 83 spectral absorptivities are shown for water vapor
and and CO2

Kirchhoff ’s law for the absorption and emission in the gas. Emissivity
of a gas
Similar to the radiative heat properties related to the surfaces, it is of high interest to relate
absorption and emission properties of the gas. To do this, let us conisder geometry presented in

108
Figure 82: Spectral absorptivity of water vapor for layers a,b,c,d,e of different thickness, from
V.P.Isachenko, V.A.Osipova, A.S.Sukomel, ”Heat transfer”

Figure 83: Spectral absorptivity of CO2 for a layer of thickness 50mm (1), 20mm (2), 63mm
(3) and 1m (4), from V.P.Isachenko, V.A.Osipova, A.S.Sukomel, ”Heat transfer”

fig.81. Assume that the black surface dF and the gas have the same temperature. The surface
emits the energy, the gas absorbs it and re-emits. Since we assume the thermal equilibrium,
the resulting heat fluxes in the system should be zero. Therefore, :
the gas above the surface emits toward the black surface dF the same amount of energy as it
aborbs.
Absorbed radiation along each direction within a hemisphere is given as

Ia,λ = ag,λ I0,λ,b (λ, T )

where I0,λ,b (λ, T ) is intensity of the radiation of the surface dF . On the other hand, this is the
energy emitted by the gas which can be expressed in terms of emissivity and radiation of the

109
black body:
Ie,λ = g,λ I0,λ,b (λ, T )

Comparing two last equations, find that

g,λ = ag,λ = 1 − e−aλ L (221)

The eq.221 is valid for any direction in space. If the distance L does not vary with direction or
if we can use an averaged value L̄, it is possible to integrate these relations over a hemisphere
and to introduce a hemispherical emissive power at a surface of a gas volume from a hemisphere
of a radius L filled with a gas:

Eg,λ = g,λ πI0,λ,b (λ, T ) = g,λ E0,λ,b (λ, T )

where I0,λ,b = E0,λ,b (λ, T )/π is defined via Planck’s law.

Average path length

In practice the volume of gas can have a shape different from a hemispherical one, i.e. the path
length for a radation varies with direction. However, it can be averaged and in the table below
mean path lenghs are given for various geometries.

Figure 84: Mean path lengths for various gas geometries

Gray gas model


Definitions given above were obtain for a certain wavelength λ, i.e. for a monochromatic rada-
tion. This approach is natural for gases because their absorption and emission are characterised
by separate spectral bands (i.e. it is not continuous as for liquids and solids). However, the
practical approach requires simpler description and gray gas model is used where a total
emissivity is used:
Eg (T ) = g Eg,b (T ) = g σT 4

110
Total emissivity g , similar to extinction coefficient, depends on T, C, P etc. Let us consider as
an example a gray gas model for a carbon dioxide at T = 1400K. At this temperature main
absorption bands are ( see also fig.85 below), note that these data are a little different from
those given above):
2.56 − 2.88µm
4.15 − 4.76µm
9 − 20µm
Then, emissive power can be calculated within each bands of the spectrum assuming that the
coefficient Kλ is constant within this band:
Z λi,2
∆Ei,g = i,g (λ, L)Eg,b,λ (λ, T )
λi,1
Z 2.86µm
−aλ,1 L
∆E1,g = (1 − e ) σTg4 (222)
2.56µm
Z 4.76µm
∆E2,g = (1 − e−aλ,2 L ) σTg4 (223)
4.15µm
Z 20µm
∆E3,g = (1 − e−aλ,3 L ) σTg4 (224)
9µm

Figure 85: Line 1: Spectrum of emission for a black body at T=1400K; line 2: spectrum of
emission for CO2 , with L=3m; line3 : spectrum of emission for gray gas approximation of CO2

Note that spectral emissivity for the gas depends not only on the absorption coeffcient but also
on the path length of the radiation inside the gase volume:

i,g = 1 − e−aλ,i L

This means that at a certain given L the emissive power of radiation in one spectral band can
be equal to that of the black body because the coefficient aλ is high for this band whereas in
other spectral bands emissivity is below unity as it is shown in fig.85.
The total emissivity of the gas can be calculated via a standard definition of emissivity as
∆E1,g + ∆E2,g + ∆E3,g
g (T ) =
σT 4
111
For given data for CO2 , obtain

g (T ) = (1 − e−aλ,1 L )F2.56→2.56 (λ, T ) + (1 − e−aλ,2 L )F4.15→4.76 (λ, T )+


(1 − e−aλ,3 L )F9→20 (λ, T ) (225)

where notation in term of fraction Fλ,1→λ,2 is used for the emission within a spectral band (not
from the zero!). Now remember that Kλ,i = kλ,i · p. Then equation 225 shows that the total
emissivity of the gas depends on the size of the gas volume, on the absorption coefficient aλ
(we neglect the scattering), on the partial pressure of the gas. When Kλ,i L = kλ,i · pL >> 1
the exponents go to zero and the total emissivity of the gas has the maximal value which is for
the case under consideration, for T = 1400K CO2 = 0.2
Dependence of the total emissivity on temperature is manifested in two ways. First, the den-
sity of the gas decreases with temperature, therefore, the extinction coefficient decreases that
decreases Kλ . Second, the width of spectral bands increases if temperature decreases that in-
creases the value of fractions involved in calculation of the total emissivity of the gas, so there
is competition between two effects. In fig.86 the dependence of maximal possible values of total
emissivities for water vapor and carbon dioxide is shown (i.e. in this case kλ,i · pL >> 1). Note
that presented value for CO2 is sligntly different from the one just calculated (0.2).

Figure 86: Dependence of maximal emissivities for H2 O and CO2 on temperature

Experimental data on emissive power of CO2 and H2 O may be interpolated with formulae
 3.5
0.33 T
ECO2 = 3.5(pc L)
100
 3
0.8 0.6 T
EH2 O = 3.5pw L
100

This generalization shows that the emissive power of a gas may differ significantly from the
Stephan-Boltzman law, nevertheless, the emissive power is still proportional to the temperature.

112
Nomograms and empirical formulae
Radiating gas in a mixture with non-radiating gases

Nomograms are constructed using various measurements of emissivities (and absorptivities) of


gases under different conditions. In particular, they present variation of the total emissivity of
a gas volume (hemisphere for a given case) dependent on the temperature of the gas Tg , the
total and the partial pressure of the gas p and pg , and the raidus of the hemisphere L. Two
nomograms are shown in fig.87a,b for the total pressure p = 1atm.

A B

Figure 87: A:Total emissivity of water vapor in mixture with nonradiating gases at 1-atm total
pressure and of hemispherical shape; B:Total emissivity of CO2 in mixture with nonradiating
gases at 1-atm total pressure and of hemispherical shape. From Incropera, DeWitte, Bergman,
Lavine, ”Fundamentals of heat and mass transfer”.

To evaluate the emissivity under total pressure different from 1 atm the emissivities given in
fig.87 should be multiplied by the correction factor Cw and CC which are given in fig.88

Mixture of radiating gases

Nomograms given above are apply when water vapor or carbon dioxide appear separately in the
mixture with other gases that are nonradiating. For the gaseous mixture of H2 O and CO2
the total emissivity is given as a sum of the total emissivities of these gases with a subtraction
of a correction factor:
m = w + c − ∆gm (226)

113
A

Figure 88: Correction factors for total pressure different from 1 atm for water vapor H2 O (A)
and for carbon dioxide CO2 (B). From Incropera, DeWitte, Bergman, Lavine, ”Fundamentals
of heat and mass transfer”.

Correction ∆gm is presented in graphical form in the fig.90 and takes into account the mutual
absorption between two gases.

Figure 89: Correction factor for the total emissivity of water vapor and carbon dioxide mixture
given by eq.226. From Incropera, DeWitte, Bergman, Lavine, ”Fundamentals of heat and mass
transfer”.

Note that in the nomograms the value p · L is given in f t · atm, 1f t = 30.48cm ≈ 0.3m.

Instead of nomograms, formulae developed by various reseracher (Hottel et al) can be used.

114
Radiating gas near a surface
Consider heat fluxes in a system consisting of a black surface with temperature Ts and a gas
volume with temperature Tg near this surface. The rate of the radiant heat transfer to a surface
due to the emission is given as
Qg,e = Ag σTg4
Radiation emitted by the surface σTs4 is absorbed by the gas with a coefficient ag . Therefore,
the resulting heat flux for the gas is

Qg,res = Aσ(g Tg4 − ag Ts4 ) (227)

However the question rises about the absorption coefficient ag since the radiant flux in gas
depends both on the gas temperature and surface temperature. Recommended formula for
evaluation of absorptivity in water vapor and carbon dioxide are
 0.45
Tg Ts
aw = C w · w (Ts , pw · L · ) (228)
Ts Tg
 0.65  
Tg Ts
ac = C c · c Ts , pc · L · (229)
Ts Tg
where w and c are avaluated from fig.87a and b, respectively, accoutning for that Tg is replaced
with Ts and pw L is replaced with pw LTs /Tg (pc L is replaced with pc LTs /Tg ). In the presence
of both gases the total gas absorptivity is expressed as

ag = aw + ac − ∆am (230)

and ∆am = ∆m taken from the fig.90.

Example
A gas turbine combustion chamber may be considered as a long tube with of 0.4 m diameter.
The combustion gas is at T = 10000 C and pressure of 1 atm, the temperature of the chamber
surface is 5000 C. If the combustion gas contains CO2 and water vapor, each with a mole frac-
tion of 0.15, what is the net radiative flux between the gas and the chamber surface (which can
be approximated as black body)?

Figure 90: Illustration to the example.

115

You might also like