You are on page 1of 41

TOPIC 4.

Introduction to Thermal Energy 1

Topic 4

Introduction to Thermal Energy

Contents

4.2 TEMPERATURE................................................................................................................... 4
4.2.1 Thermal Equilibrium.............................................................................................6
4.2.2 Zeroth Law of Thermodynamics...........................................................................6
4.2.3 Temperature Measurement..................................................................................7
4.3 MODES OF HEAT TRANSFER...............................................................................................8
4.3.1 Heat Conduction...................................................................................................8
4.3.2 Heat Convection...................................................................................................9
4.3.3 Thermal Radiation..............................................................................................10
4.4 HEAT TRANSFER BY CONDUCTION.....................................................................................11
4.4.1 Conduction through a Single Element................................................................11
4.4.2 Heat Conduction through Two Elements in Series.............................................13
4.4.3 Total Conductance of Series Network of Resistances........................................17
4.5 HEAT TRANSFER BY CONVECTION.....................................................................................19
4.5.1 Heat Convection through a Single Element........................................................20
4.5.2 Typical Heat Transfer Correlations.....................................................................22
4.5.3 Using a Heat Transfer Correlations....................................................................23
4.5.4 Many Resistances in Series...............................................................................24
4.5.5 Case Study – Jacketed Vessel...........................................................................25
4.6 HEAT TRANSFER BY RADIATION.........................................................................................29
4.6.1 Stefan-Boltzmann Law.......................................................................................30
4.7 ENERGY BALANCES ON REACTING SYSTEMS......................................................................32
4.7.1 Extent of Reaction..............................................................................................32
4.7.2 Hess’s Law.........................................................................................................34
4.7.3 Enthalpy of Formation........................................................................................35
4.7.4 Enthalpy of Combustion.....................................................................................36
4.7.5 Internal Energy Change of Reaction...................................................................37
4.8 TUTORIAL QUESTIONS TOPIC 4...........................................................................................39
4.9 Bibliography....................................................................................................................41

Prerequisite knowledge

 Engineering mathematics.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 2

 Engineering chemistry.
 Mass and energy balances for both closed and open systems.
 Detailed understanding of the ideal gas law.
 Thermodynamic principles and definitions.
 Detailed understanding of both the steady-flow energy equation and
the non-flow energy equation.

Learning objectives
After studying this Process Industries C topic you should be able to:
 Write down and defined thermal equilibrium and the Zeroth Law of
Thermodynamics
 Convert between common empirical and absolute temperature scales.
 Differentiate between conduction, convection and radiative modes of
heat transfer.
 Draw resistance network diagram for conduction and convection
resistances in series.
 Differentiate between overall resistance and overall conductance.
 Solve heat transfer problems using resistance network and overall
conductance approach.
 Outline main facts concerning radiative heat transfer.
 Carry out energy balances on reacting systems and review basic
definitions.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 3

4.1 Introduction

When a system changes state it is said to undergo a “process”. Some


common processes are as follows:
 Isobaric: constant pressure process.
 Isothermal: constant temperature process.
 Isochoric: constant volume process.
 Isenthalpic: constant enthalpy process.
 Isentropic constant entropy process.
 Adiabatic a process involving no heat transfer with surroundings.

Energy balances are crucial in Chemical Engineering since they help to define
what utilities are needed, the energy cost and the size of the heating or cooling
equipment that will be needed.

Large energy changes are associated with the following: separation of fine
mixtures, as in distillation; phase change, as in evaporation, condensation or
solidification; chemical reactions, both endothermic and exothermic.

Smaller energy changes are associated with: pumping and transportation of


fluids; separation of coarsely mixed materials (sand/water); heating or cooling
without phase change.

The non-flow and steady-flow energy equations are both expressions of the
First Law of Thermodynamics, which encapsulates the well-known energy
conservation principle.

Heat and work are both forms of energy. They are energy exchange terms
and not energy storage terms. The First Law is about balancing “quantities”. In
this regard 1 (kJ) of work is the same “quantity” of energy as 1 (kJ) of heat.

The Second Law of Thermodynamics is about grading energy in terms of


“quality”:
 Work may be converted completely into heat, but heat cannot be
completely converted into work; therefore, 1 (kJ) of work is a higher
“quality” (higher grade of energy) than 1 (kJ) of heat.
 A fixed quantity of heat at a higher temperature is a higher grade more
useful form of energy (higher quality) than when it is at a lower
temperature.

Although the quantity of energy is always conserved, energy quality for any
real process tends to decline:
 Work naturally tends to become degraded into heat.
 Heat naturally tends to flow from a higher to a lower temperature.
 Therefore, any device or process that works against this natural
tendency – i.e. heat engines, turbines, compressors, refrigerators, heat
pumps, etc. must have their efficiency limited by the Second Law.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 4

4.2 Temperature

An empirical temperature scale is one where arbitrary numbers are placed


against a pair of reproducible reference states – see empirical Celsius and
Fahrenheit scales below:

Reproducible Celsius Celsius Fahrenheit Fahrenheit


Reference Scale “Degrees” Scale “Degrees”
State
Boiling Point 100oC 212oF
of Pure Water
100
180
Freezing Point 0oC 32oF
degre
degre
of Pure Water

Between these fixed points, the empirical Celsius scale has 100 “degrees”,
while the empirical Fahrenheit scale has 180 “degrees”:
 Therefore, each oC is equivalent to 180/100 = 1.8 oF.
 There is an offset of 32 degrees between the scales.
 Using TC for the Celsius scale and TF for the Fahrenheit scale the
conversion between the two scales is given by

TF  (1.8  TC )  32 ………………………………..…(4.1)

An absolute scale places the zero point at the absolute zero of temperature.
By international agreement in 1956 the Triple Point of pure water was chosen
as the reproducible reference state for the Kelvin scale

The Triple Point pressure is 0.006112 bar and by international agreement the
triple point temperature was assigned a precise value of 273.16 K.

This value was chosen so that there was approximately 100 K between the
“ice point” and the “steam point” of pure water, which meant that one degree
on the Celsius scale is the same as one degree on the Kelvin scale.
The Celsius scale TC is now defined in terms of the absolute Kelvin scale
TK as follows:

TC  TK  273.15 ……………………………………(4.2)

Notice that the Celsius scale is only shifted with respect to the Kelvin scale
and the size of a degree is the same in both cases – thus, temperature
difference is the same for both scales and (W/m oC) is the same as (W/m K).
The Triple Point is now the reproducible reference state and since its value
was fixed at 273.16 K (exactly), it temperature on the Celsius scale is given by

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 5

TCTP  273.16  273.15  0.01o C

Thus, the Triple Point temperature of water is either 273.16 K or 0.01oC.


Likewise the temperature of ice/water or the “ice point” is still 0 oC, so that its
temperature on the Kelvin scale is given by equation (3.2)

TK ICE  0  273.15  273.15 K

The temperature of steam/water or the “steam point” is still 100 oC, so that its
temperature on the Kelvin scale is again given by equation (3.2)

TK STEAM  100  273.15  373.15 K

Absolute zero on the Celsius scale is also given by equation (3.2), knowing
that absolute zero must correspond to 0 K:

TC Abs zero  0  273.15  -273.15 o C

On the Fahrenheit scale absolute zero can be readily be found

TF Abs zero  1.8    273.15   32  -459.67 o F

The Fahrenheit absolute temperature scale is the “Rankine” scale, so that


absolute zero also corresponds to zero rankine.

TR  TF  459.67 ………………………………….(4.3)

In summary the various fixed reference states are listed below:

Reference Celsius Kelvin Fahrenheit Rankine


State Scale Scale Scale Scale
TC (oC) TK (K) TF (oF) TR (R)
Boiling Point of 100 373.15 212 671.67
Pure Water
Triple Point 0.01 273.16 32.018 491.688
Pure Water
Freezing Point 0 273.15 32 491.67
of Pure Water
“Absolute Zero” -273.15 0 -459.67 0

4.2.1 Thermal Equilibrium

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 6

A system that is in equilibrium implies that there are no driving forces across
the boundary (between system and surroundings). This in turn implies
complete uniformity of all thermodynamic properties throughout the system.

Thermal equilibrium implies no temperature difference across the boundary,


which means complete uniformity of temperature throughout the system; under
these conditions there can be no heat transfer

Thus if hot body “A” is brought in contact with cold body “B”, heat will flow from
“A” to “B” until they are in thermal equilibrium. Once at equilibrium, heat
transfer stops at which point both bodies must have the same temperature.

No matter what “A” and “B” are made off, if they are in thermal equilibrium and
if a thermometer is brought in contact with each body, then the thermometer
must read the same value. However, the sense of touch is more subjective:
 Take it that body “A” and “B” are in thermal equilibrium with each other,
but “A” is metallic while “B” is expanded polystyrene.
 They must both be at the same temperature – assume that this
temperature is 0oC.
 When touched, body “A” will feel colder than body “B”, this is simply
because the metal has a much higher thermal conductivity and
conducts heat away from the hand faster – making it “feel” cooler.

4.2.2 Zeroth Law of Thermodynamics

This can be stated quite simply:

“If body A and body B are each equal in temperature to a third body C, then
body A and body B must be equal in temperature to each other”.

This is how thermometers work:


 The thermometer is brought in contact with a body of unknown
temperature.
 Time must be allowed for the thermometer to reach thermal
equilibrium.
 Once in thermal equilibrium the thermometer will read the correct
temperature so long as it is “calibrated” properly.

When measuring air temperature on a hot sunny day a thermometer will often
be hotter than the air temperature. The thermometer is picking up heat directly
from the sun by a process known as “thermal radiation”.

This is why air temperature is usually measured in the shade. Under these
circumstances, direct exchange of solar radiation cannot heat the thermometer
above the prevailing air temperature to give a false reading.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 7

4.2.3 Temperature Measurement

There are a wide variety of thermometers on the market. One of the oldest
types is the “expansion” thermometer.

When an expansion thermometer is brought in contact with a hot substance,


the fluid inside the thermometer (mercury or alcohol) will expand out of the
bulb and rise in the capillary tube – thus indicating the temperature.

However, the expansion is quite small so that the liquid volume in the bulb
must be large in relation to the capillary tube volume. Water is not used since
its expansion is non-linear – having a minimum volume at 4oC.

Thermocouples are based on the “thermo-electric” effect. When two different


metals are joined to form a “junction” a small voltage is generated which is
temperature dependent and this may be used to measure temperature.

Resistance thermometers work on the principle that the resistance of a wire


(usually platinum) depends on temperature. Hence, by measuring the
resistance of the thermometer bulb, temperature can be measured.

Thermocouples and resistance thermometers are widely used throughout


chemical plants. The electrical signals are conditioned and transmitted from
any point in the plant to some remote centralised control room.

Two thermometers, using different fluids, may read the same at two fixed
calibration points, but there may be small differences between them at other
points on the scale.

Pyrometers do not require any physical contact with the material whose
temperature is being measured.
 All bodies emit thermal radiation and the wavelength of the emitted
radiation depends on the temperature and the surface characteristics
of the body.
 The pyrometer measures the infra-red wavelengths being emitted.
 The operator adjusts the instrument to account for the surface
characteristics of the body.
 The reading then reflects the temperature of the object.

Pyrometers are good for measuring very hot objects that are dangerous to
approach or would burn out other measuring devices.

Pyrometers are also very useful for measuring the temperature of moving
objects where direct physical contact would be difficult or impossible.

In the next section the different modes of heat transfer will be discussed and
these fall into one of three categories: heat transfer by conduction; heat
transfer by convection; heat transfer by radiation.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 8

4.3 Modes of Heat Transfer

4.3.1 Heat Conduction

In the case of a solid, its atoms or molecules are in fixed positions relative to
each other – they do, however, vibrate around these fixed positions and
heating a solid causes them to vibrate faster.

If one side of a block is hotter than the other, the atoms (or molecules) on the
hotter side will have a higher energy and will be vibrating faster than the atoms
or molecules on the other side of the block.

Since the atoms are continuously bumping into each other, this increased rate
of vibration is passed from atom to atom. In this way, thermal energy “flows” by
conduction from the hot side of the block to the cold side.

Heat transfer by conduction may also occur in fluids. If the fluid is flowing
under laminar flow conditions, then heat transfer will occur by conduction.

Also, when a hot vapour condenses onto the outside of a cold pipe, a thin
laminar film of condensate forms on the outside of the pipe. Heat is transferred
by conduction through this laminar film to the cold surface.

One dimensional steady-state heat conduction law is known as Fourier’s Law


and is given by

   kA dT ………………………………………….…(4.4)
Q
dx

Where,
 
Q The rate of heat transfer (W)
k  The thermal conductivity of the solid or fluid (W/m K)
A  The heat transfer area, perpendicular to heat flow (m2)
dT / dx  The temperature gradient (K/m)

The minus sign is needed because the temperature gradient is negative in the
direction of heat flow ( T final  Tinital ).

In the case of metals conduction is greatly enhanced by the movement of “free


electrons”. Thus good electrical conductors are also good heat conductors.

The thermal conductivity of gases is much lower because the atoms or


molecules are further apart. The same applies to solids that contain pockets of
gas, such as polystyrene – these solids make very good thermal insulators.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 9

4.3.2 Heat Convection

On the other hand, the molecules (or atoms) of a fluid do not simply vibrate
about fixed positions:
 They are free to move both macroscopically, as fluid eddies, and
microscopically (within a fluid eddy), by diffusion.
 This mechanism transmits energy as heat either into or out of a moving
body of fluid.
 This mode of heat transfer is called convection.

There are two types of convection:


 Forced convection where fluid movement is mechanically induced by a
pump or a fan.
 Free convection where fluid movement is induced naturally by fluid
buoyancy effects – hot fluid rises and is replaced by cold fluid which
sinks.

Steady-state heat convection law is given by the expression below:

  hAT
Q ………………………………………….…(4.5)

Where,
 
Q The rate of heat transfer (W)
h  The fluid heat transfer coefficient (W/m2 K)
A  The heat transfer area, perpendicular to heat flow direction (m2)
T  The temperature difference in the direction of heat flow (K)

If hot water is circulated by a pump through a cooler tube, then heat will be
transferred by forced convection from the body of the hot water to the cooler
metal (the more turbulence within the fluid, the faster the rate of heat transfer).

If water is heated on a stove, then “free convection cells” will form within the
body of fluid; these are caused by hotter/less dense water rising and
colder/more dense water falling – thus, heat is transferred by free convection.

Consider a car radiator: heat is transferred by forced convection, from hot


water within the radiator; by conduction, through the radiator metal wall; and
finally by forced convection to the cold surroundings.

Heat transfer coefficients are found experimentally and many correlations have
been developed and published – broadly speaking, these correlations depend
on the following:
 Whether heat transfer is by forced convection or natural convection.
 Whether the flow regime is laminar or turbulent.
(Whenever hot vapours condense on to cold tubes, or whenever liquids
are boiling, then special correlations are required.)

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 10

4.3.3 Thermal Radiation

Thermal radiation involves electromagnetic transfer of heat from a hot body to


a cold body. The heat does not need to be transferred through any substance,
indeed it may even be transferred through a complete vacuum.

All bodies emit thermal radiation, but this is usually invisible (infra-red region of
the electromagnetic spectrum). However, if a body is hot enough, then it will
emit visible light – the wavelength depends on its temperature.

The rate of thermal radiation from an “ideal blackbody” is given by the Stefan-
Boltzmann Law, as follows:

  AT 4 ………………………………………….…(4.6)
Q

Where,
 
Q The rate of heat transfer (W)
 The Stefan-Boltzmann constant (5.67x10-8 W/m2 K4)
A  The heat transfer area (m2)
T  The absolute temperature (K)

Corrections are usually required to equation (4.6) to deal with the following
complications:
 “Non-black” behaviour – no body is a “perfect” emitter so that some
adjustment is needed to equation (4.6).
 A body not only emits thermal radiation, but receives thermal radiation
from other bodies – thus, equation (4.6) is modified to allow for a net
energy exchange between two bodies having different temperatures.
 Since thermal radiation is a directed form of energy transfer, the “view”
that different bodies have of each other is important – thus, equation
(4.6) is modified to include these "View Factors".

In the case of a car radiator, not only is heat transferred through the radiator
successively by forced convection, conduction and then forced convection to
the surroundings, but also it is radiated away from the hot metal to the cold
surroundings.

If the area is brought over to the LHS of equation (4.4), (4.5) or (4.6), so that
 .
rate of heat transfer is divided by area, then this is called the “heat flux” Q f

With heat transfer it is easier to start with heat transfer through flat (plane)
surfaces. Heat transfer through pipes is complicated by the fact that the curved
area, perpendicular to the direction of flow, continually changes.

Heat transfer is made more complex, not only by the different modes of heat
transfer, but also by whether or not these modes occur in series or in parallel.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 11

4.4 Heat Transfer by Conduction

Conduction can occur through a single element or, more usually, a number of
elements in series. Consider a single element first.

4.4.1 Conduction through a Single Element

Temperatures on either side of the single flat element are fixed at T1 (hotter
face) and T2 (cooler face) respectively:

Heat Transfer
Area A1 (m2)

Rate of
Heat Transfer
Q (W)

Thermal
Conductivity
k1 (W/m K)

 The heat transfer rate is


TemperatureT1
steady with respect to
(K)
time.
 Thus, the temperature
change from the inlet
T2 face to the outlet face is
uniform

Thickness
0 x1 (m)

Fourier’s one dimensional steady-state heat conduction law is given by


equation (4.4)

   k A dT
Q 1 1
dx

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 12

Now separate the variables to get

x1 T2

Q  dx   k1 A1  dT
0 T1

Integrating and then re-arranging this expression leads to

k A
Q  1 1  T1  T2  ……………………………………..(4.7)
x1

Finally, put this equation into the form shown below

Q  T  T2 
 1
A1 x1 / k1

This expression takes the general form

Driving Force
Rate 
Resistance

Where,
  The heat flux or the “rate” (W/m2)
Q f

 T1  T2   The temperature difference or “driving force” (K)

And, the heat transfer resistance R1 (K m2/W) of a “single conductive


element” is given by

x1
R1  ………………………………………………..(4.8)
k1

The resistance concept is an excellent way of visualising heat transfer


problems – heat transfer is now analogous to Ohm’s Law for electricity:
 Voltage drop is analogous to temperature drop.
 Flow of electrical current is analogous to heat flux.
 Electrical resistance is analogous to thermal resistance.

Through equation (4.8), the higher the thermal conductivity k (W/m K) then
the lower the resistance R .

Double glazing works on the principle that two layers of glass, separated by a
thin layer of some low conductivity gas, has a higher resistance to heat
transfer than a single pane of glass on its own.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 13

The higher the thermal resistance of a window, then the lower will be the rate
of heat loss per unit temperature drop across it. Triple glazing has an even
higher thermal resistance (and lower heat loss) than double glazing.

Thermal conductivity is a measure of how many joules per second of heat can
be conducted through a block of material, with unit thickness and unit area,
subject to unit temperature difference:
 Solids are better thermal conductors than liquids.
 Liquids are better thermal conductors than gases.
 Water is a better thermal conductor than most liquids.
 Gas molecules at the same temperature have the same kinetic energy,
so lighter gases move faster and are better thermal conductors than
heavier gases.

Thermal conductivities of some common materials are as follows:

Material Thermal Conductivity


k (W/m K)
Copper 377
Mild Steel 45
Stainless Steel 16
Glass 1.09
Water 0.62
Benzene 0.16
Hydrogen 0.17
Air 0.024

4.4.2 Heat Conduction through Two Elements in Series

Consider heat conduction through two elements in series (both with same area
perpendicular to the direction heat flow):

Element Element
“1”: “2”:

Rate of Heat
Transfer Q
 (W)

Thermal Thermal
Conductivity Conductivity
k1 (W/m K) k 2 (W/m K)
Thickness T1 T2 T3 Thickness
x1 (m) x 2 (m)

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 14

Each of the elements, numbered as shown, has its own thermal conductivity
and thickness. The areas of each element are the same and the temperature
of each face are numbered “1” to “3”, going from left to right.

Apply equation (4.7) to the first element, Re-arrange the equation and drop the
subscript on the area term. The temperature drop across the first element is

x1 
(T1  T2 )  Q
k1 A

Then apply equation (4.7) to find the temperature drop across the second
element as follows:

x2 
(T2  T3 )  Q
k2 A

Now add both equations together. The heat transfer rate Q is the same
through each element (steady-state assumption). The result is the overall
temperature drop across both elements, as follows:

x1  x
(T1  T3 )  Q  2 Q
k1 A k2 A

x x  Q
 (T1  T3 )   1  2 
 k1 k 2  A

T  T 
 Q f  1 3
 x1 x2 
  
 k1 k 2 

This expression again takes the general form

Driving Force
Rate 
Resistance
Where,
  The heat flux or the “rate” (W/m2)
Q f

 T1  T3   The overall temperature difference or “driving force” (K)

One of the advantages of this approach is that the unknown intermediate


temperature has now been eliminated in favour of the overall temperature
drop. In real problems it is this overall temperature drop that is usually known.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 15

The total heat transfer resistance RT of both elements (the combined


resistance) is given by

x1 x1
RT   …………………………………..………(4.9)
k1 k1

The total heat transfer resistance RT of both elements together is simply the
sum of the resistances of each element, so that

RT  R1  R2 …………………………………………(4.10)

Just as electrical resistances in series are additive, so thermal resistances in


series are also additive; this is what makes the resistance-analogy so
valuable.

The result may be extended to any number of thermal resistances in series as


follows:

RT   Ri ………………………………………….(4.11)
i

Example: 4.4.1

Problem:

The inside of a metal furnace wall, thermal conductivity 45 W/m K and


thickness 5 cm, is 150oC. If there is 2.5 cm thickness of lagging, with thermal
conductivity 0.05 W/m K and, if the outside lagging temperature is 30 oC, then
calculate the individual resistances of each element, the overall resistance of
both elements, the heat flux through the overall system and the temperature of
the outside furnace metal wall (the same as the inner lagging temperature).

Assume steady-state operation throughout.

Solution:

The individual resistances are given by equation (4.8) – first for element “1”

x1 0.05
R1    0.001 (K m2/W)
k1 45

Then for element “2” – the lagging

x2 0.025
R2    0.5 (K m2/W)
k2 0.05

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 16

The total resistance for the combined system– the metal and the lagging

RT  R1  R2

 RT  0.5  0.001  0.501 (K m2/W)

The heat flux Q f (W/m2) through the combined system is given by

T  T 
Q f  1 3
 x1 x 2 
  
 k1 k 2 

150  30
 Q f   239.5 (W/m2)
0.501

Notice that this calculation could be represented by the following “resistance


network”, where temperature is analogous to voltage, thermal resistance to
electrical resistance and heat flux to current flow – see below:

This greatly helps to visualise the problem. The intermediate temperature T2


is found in exactly the same way as if it were a missing intermediate voltage.
Just as the voltage drop across R1 would be

V1  i R1

So, the temperature drop across this first element is simply

T1  Q
 R  239.5  0.001
f 1

 T1  0.24 oC

 T1  T2  0.24 oC

T2  149.76 oC
…………………………………………………
Notice how small the temperature drop is across the metal (very small
resistance) and how large it is across the lagging (much larger resistance).

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 17

4.4.3 Total Conductance of Series Network of Resistances

The “ U -value” is the total “conductance” of an electric circuit. The total


conductance is simply the reciprocal of the total resistance, so that

1
U 
RT

This can be expressed the other way around and, for a system of thermal
resistances in series, this becomes

1
  Ri ……………………………………………(4.12)
U i

Generalising the “two resistances in series” system, to any number of


resistances in series leads to

TT
Q f 
 Ri …………………………………….……(4.13)
i

Where,
  The heat flux through each element in series (Wm2)
Q f

TT  The total temperature difference across all elements in series


(K)
R i
i  RT The total resistance across entire series network (K m2/W)

Equation (4.13) may be expressed in terms of the overall “ U -value” simply by


applying equation (4.12), as follows:

  UT ………………………………………..…(4.14)
Q f T

Eliminate the heat flux Q f  Q / A to obtain

  UAT ……………………………………….…(4.15)
Q T

Where, Q  (W) is the constant rate of heat flow through the series network of
resistances.

Equation (4.15) is one of the most well-known heat transfer expressions and is
used widely throughout industry to design all types of heat exchangers. The “
U -value” is always quoted when rating the thermal performance of doors,
windows, walls, roofs, floor slabs, etc.
Example: 4.4.2

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 18

Problem:

Using the same information as the previous problem, calculate the U -value
of the system of two resistances in series and find the rate of heat loss, to the
surroundings, through 10 m2 of lagging.

Solution:

The total resistance for the combined system – the metal and the lagging was
found to be

RT  R1  R2

 RT  0.5  0.001  0.501 (K m2/W)

This may be converted to a U -value using equation (4.12)

1
  Ri
U i

1
U   2 W/m2 K
0.501

Then the total rate of heat loss through 10 m 2 of lagging area is given by
equation (4.15)

  UAT
Q T

Q
  2  10  (150  30)

Q kW
  2 .4
…………………………………………………

Each square metre of lagging loses about 0.24 kW or approximately 240 W,


which is the same as 240 J every second. This represents a large heat loss
and cost per annum, which is why the performance of lagging is so important.

When solving heat transfer problems the resistance network approach is used
to find the total thermal resistance. Once this is known it is converted into a
U -value, using equation (4.12), then equation (4.15) may be solved.

In many heat exchanger design problems it is the heat transfer area A (m2)
that is needed. This area determines the size of the heat exchanger.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 19

4.5 Heat Transfer by Convection

Many problems involve a complex network of resistances in series, however


very few problems involve heat conduction only; most involve a combination of
heat conduction and heat convection resistances.

In order to find the heat convection resistance, start with the steady-state heat
convection law that was presented earlier

  h A T
Q i i i

Where,
 
Q The rate of heat transfer through a single convection element (W)
hi  The fluid heat transfer coefficient through an element (W/m2 K)
Ai  The area of the element, perpendicular to heat flow direction (m2)
Ti  The temperature drop across the single convective element (K)

Putting this into the same form as Ohm’s Law leads to

Q Ti

Ai 1 / hi

In terms of heat flux Q f  Q / A this becomes

Ti
Q f 
1 / hi

This is now in the same form as Ohm’s Law for electricity

Vi
i
Ri

Where,
i  The current flow through resistance element (A)
V  Voltage drop across a single element (V)
R  Electrical resistance of the element (  )

By comparing the last two expressions the thermal resistance of a convective


element is clearly

1
Ri  ……………………………………………….(4.16)
hi

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 20

4.5.1 Heat Convection through a Single Element

Consider a flat wall with fluid flowing parallel to the wall; consider also that the
fluid is being cooled by a fluid on the other side of the wall (not shown):

Wall Fluid
The fluid flow is parallel to the
wall, as shown:
Flow  The heat flow direction is
Past perpendicular to fluid flow.
Wall  Since fluid is being cooled,
heat flow is from fluid to the
wall.
Heat
Flow
The temperature profile through
Temperature Tb the fluid, at any point along the
(K) flow path, is shown to the left:
 The temperature varies from
the undisturbed “bulk”
T
temperature b .
Tw
 To the “wall” temperature Tw

Distance from
T1
Wall

At steady-state, the heat convection law was shown earlier to be

  h A T
Q i i i

Where, Ti  (Tb  Tw ) is the temperature difference across the “convective


element” (the difference between the bulk and wall temperatures). The thermal
resistance of this single convective element is then given by

1
Ri 
hi

Where, hi is the “film” heat transfer coefficient (W/m 2 K) and Ri is the


equivalent resistance of this element. Using this approach, convective and
conductive resistances in series may be combined.

Despite the fact that fluid flow may be turbulent, there is a thin sub-layer (see
dotted red line above) immediately adjacent to the solid wall where flow is
entirely laminar – this is caused by the “no-slip” condition.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 21

Within the laminar sub-layer heat is transferred by a “conduction-only”


mechanism which tends to be slow. In order to maintain the rate of heat
transfer through this layer a large temperature drop is needed.

Outside the laminar sub-layer


Heat is the turbulent core where
Flow heat may be transferred by:
Temperature Tb  Macroscopic eddy mixing,
(K) see circular patterns.
 Within eddies heat is also
transferred by conduction.
Tw  However, the eddy mixing
greatly enhances the rate of
heat transfer.
Distance from
Wall

Thus, to maintain the rate of heat transfer from the bulk of the fluid to the edge
of the sub-layer only a small T is needed – see temperature profile above.

The same rate of heat transfer occurs through both the sub-layer and the
turbulent core but, in order to maintain this rate of heat transfer through the
sub-layer, a much larger T is required – see temperature profile.

This is why convective heat transfer coefficients are referred to as “film” heat
transfer coefficients – most of the temperature drop actually occurs across this
thin laminar sub-layer (it controls the overall rate of heat transfer).

Since the sub-layer controls the overall rate of heat transfer, it is fairly obvious
that the thinner the film:
 The faster will be the overall rate of heat transfer from the bulk of the
fluid to the wall.
 The larger will be the film heat transfer coefficient.
 The smaller will be the convective resistance of this element.

Thus, higher Reynolds Numbers will cause more turbulence, reduce the
thickness of the sub-layer and increase the heat transfer coefficient. It follows
that heat transfer coefficient must depend on the Reynolds Number

hi  f (Re)

However, heat transfer coefficients are not correlated in this way – they are
“embedded” inside another dimensionless group called the Nusselt Number
“Nu”, so that the actual correlation takes the form

Nu  f (Re)

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 22

4.5.2 Typical Heat Transfer Correlations

In fact it turns out to be a little more complicated. For forced convection


through a circular pipe, under certain specific conditions, the heat transfer
correlation (in full) becomes

Nu  0.023(Re) 0.8 (Pr) 0.4 ………………………………..(A)

Where, “Nu” is the Nusselt Number, defined as follows:

hd
Nu  Nusselt Number 
k

Where,
h  The fluid heat transfer coefficient through an element (W/m2 K)
d  The inside diameter of the circular pipe (m)
k  The fluid thermal conductivity (W/m K)

And “Re” is the Reynolds Number, defined as follows:

du
Re  Reynolds Number 

Where,
d  The inside diameter of the circular pipe (m)
u  The average fluid velocity through the pipe (m/s)
  The fluid density (kg/m3)
  The fluid viscosity (N s/m2)

And “Pr” is the Prandtl Number, defined as follows:

CP 
Pr  Prandtl Number 
k

Where,
C P  The fluid specific heat capacity (kJ/kg K)
  The fluid viscosity (N s/m2)
k  The fluid thermal conductivity (W/m K)

Dimensionless correlations have proved very effective when correlating


complex phenomena. Notice that only three groups (Nu, Re and Pr) are
needed to correlate heat transfer between eight different variables.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 23

4.5.3 Using a Heat Transfer Correlations

A typical heat transfer correlation for convective heat transfer through circular
pipes was given by equation (A):

Nu  0.023(Re) 0.8 (Pr) 0.4

These correlations have been developed experimentally by many workers


using many different fluids – the correlations represent a “best-fit” through the
experimental data and may be used for most fluids.

Each “Number” is a dimensionless group and the procedure for finding the film
heat transfer coefficient is usually as follows:
 Select a correlation – this step is crucial and depends on the geometry,
the mode of heat transfer, as well as many other factors. Equation (A)
is an example of one such correlation.
 Calculate the “Re” and the “Pr” from their definitions.
 Use the correlation to find the “Nu”.
 Given the definition of the Nusselt Number “Nu”, back-calculate the film
heat transfer coefficient.

The next steps are then to define the resistance network, calculate all
remaining convective or conductive resistances and then find the total
resistance of the overall series network.

This overall resistance is then converted into a U -value and an equation,


such as equation (4.15), is solved – often it is the area A that is needed.

A very wide range of heat transfer coefficients are possible depending on heat
transfer mechanism, fluid properties, geometry, etc.

Transfer Mechanism Fluid Typical Film


Coefficient
h (W/m2 K)
No Phase Change: Water 1700 - 11000
Organic solvents 350 - 3000
Gases 20 - 300
Air 5 - 25
Condensation: Steam 6000 - 17000
Boiling: Water 2000 - 12000

Condensation and boiling are special cases involving phase change – each
have higher film coefficients than water alone.

In the case of boiling, bubbles are formed and these greatly increase the
degree of turbulence and the rate of which heat is transferred away from the
hot surface into the bulk of the boiling fluid.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 24

In the case of condensation, the condensing vapour moves rapidly to the cold
surface due to bulk flow (there is a large change in volume as the vapour
condenses and essentially the vapour is “pumped” to the cold surface).

On reaching the cold surface the vapour condenses into a thin film of colder
condensate, releasing all its latent heat very close to the cold surface itself.

Heat must then be transferred by conduction through this thin condensate film;
it is flowing slowly and is laminar). However, the film is often so thin (due to
continual drainage) that the heat transfer coefficients are large.

Boiling and condensation come with specialised correlations and will be dealt
with more extensively in later courses – this subject area is well-covered in
standard textbooks (Coulson and Richardson, Vol. 1, 1999).

4.5.4 Many Resistances in Series

In many heat transfer systems there can be up to five resistances in series,


each with the same heat flux passing through them, as follows:

In many cases the resistances are typically:


 Inner convective resistance – “inner side” of tube/vessel wall.
 Inner conductive dirt/scale resistance – “inner side” of tube/vessel.
 Conductive resistance through tube/vessel wall.
 Outer conductive dirt/scale resistance – “outer side” of tube/vessel.
 Outer convective resistance – “outer side” of tube/vessel wall.

It is usual that only T1 and T6 are known, which is why the overall
resistance RT and the U -value are needed – using these overall values
eliminates all the unknown intermediate temperatures.

The greater part of solving heat exchanger problems lies in calculating all
these resistances and then combining them into an overall U -value.

Allowances must also be made for the fact that dirt and scale accumulates on
surfaces over time – thus, their resistances change with time.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 25

4.5.5 Case Study – Jacketed Vessel

Consider a more realistic heat transfer situation:

Agitator

Steam
IN

Condensate
OUT

Heat is transferred from the steam jacket (located on the outside of the vessel)
into the agitated liquid. There are now five resistances in series:
 Steam-side film resistance (convective resistance R1 ).
 Steam-side dirt factor resistance (conductive resistance R2 ).
 Vessel wall resistance (conductive resistance R3 ).
 Liquid-side dirt factor resistance (conductive resistance R4 ).
 Liquid-side film resistance (convective resistance R5 ).

All these resistances exist between the bulk of the steam in the outer jacket,
through the jacket wall to the bulk of the liquid within the agitated vessel. The
only known temperatures are T1 & T6 - see network below:

In heat transfer problems the “jacket-side” is often labelled “outer” and the
“process-side” is often labelled “inner”. Notice that the heat flux is the same
through each thermal resistance – the heat must pass through each in turn.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 26

Just as in electrical circuits, the total thermal resistance is found by adding the
individual resistances together (this applies to resistances in series):

RT   Ri
i

 RT  R1  R2  R3  R4  R5

1 x dS xV x dl 1
 RT     
hS k dS kV k dl hl

The thickness of the dirt factors are never specified so they are just left as
Rdo (outer dirt resistance) and Rdi (inner dirt resistance), leading to

1 x 1
RT   Rdo  V  Rdi 
hS kV hl

Tidy up the notation to outer film coefficient ho (the service fluid in the jacket
need not always be steam) and inner film coefficient hi (within the vessel)
and then drop the “vessel” subscripts to get

1 x 1
RT   Rdo   Rdi 
ho k hi

Changing from overall resistance RT to overall U -value, leads to

1 1 x 1
  Rdo   Rdi  ……………………….(4.17)
U ho k hi

Where,
ho  The outer film coefficient for steam (W/m2 K)
hi  The inner film coefficient for liquid (W/m2 K)
Rdo  The outer dirt factor for steam-side (K m2/W)
Rdi  The inner dirt factor for liquid-side (K m2/W)
x  The thickness of the vessel wall (m)
k  The thermal conductivity of the vessel wall (W/m K)

Once the U -value is calculated, then equation (4.15) may be solved

  UAT
Q T

Example: 4.5

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 27

Problem:

A jacketed vessel similar to the one illustrated previously has a wall thickness
of 2 cm and an overall jacket heating area of 5 m 2. If the vessel contents are
heated by steam at 2 bar ( T  120.3 oC), determine the following:
a) The overall heat transfer coefficient.
b) The rate of heat transfer into the vessel when the vessel contents are at
20oC.
c) The steam flow needed to supply this amount of heat.
d) The temperature of the dirty wall inside the vessel.

Use the following information:

hs  ho  8000 (W/m2 K)
hl  hi  300 (W/m2 K)
Rdo  5x10-4 (K m2/W), expressed as a coefficient this is 2000 (W/m2 K)
Rdi  1x10-3 (K m2/W), expressed as a coefficient this is 1000 (W/m2 K)
k  45 (W/m K)

Solution:

For part a: solve equation (4.17) for the Overall heat transfer coefficient as
follows

1 1 x 1
  Rdo   Rdi 
U ho k hi

1 1 0.02 1
   (5  10  4 )   (1  10 3 ) 
U 8000 45 300

U  185 W/m2 K

The dirt factors could be expressed as equivalent heat transfer coefficients;


however, the overall heat transfer coefficient is the same either way:

1 1 1 x 1 1
    
U ho hdo k hdi hi

1 1 1 0.02 1 1
     
U 8000 2000 45 1000 300

U  185 W/m2 K
For part b: use equation (4.15) as follows

  UAT
Q T

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 28

  185  5  120.3  20


Q

Q
  92.8 kW

For part c: for steam at 2 bar pressure h fg  2201.9 kJ/kg (steam tables)

 m
Q  S h fg

Q 92.8
m
S  
h fg 2201.9

m
 S  0.042 kg/s

For part d: examine the resistance network to see the easiest way of finding
the intermediate temperature. Since T6 is known and since the heat flux Q 
f

and R5 are easily found, the easiest approach is to find T5 across R5 :

Treat the temperature drop just like a voltage drop (resistance network)


  Q  92,800  18560 W/m2
Q Finding an
f
A 5 intermediate
temperature may be
1 1 important:
R5    0.00333 K m2/W
hi 300  Will material in
contact with
 T5  Q
  R  0.00333  18560  61.8 oC surface boil or
f 5
degrade?
 Will material in
T5  T6  T5  20  61.9  81.8 oC
…………………………………………………

4.6 Heat Transfer by Radiation

All bodies emit heat by thermal radiation. Thermal radiation is part of the
continuous electromagnetic spectrum and is defined as that electromagnetic
radiation between wavelengths 0.1 μm and 100 μm.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 29

Radiant energy emissions are propagated electromagnetically with a


maximum velocity of 3 x 108 m/s in a vacuum. Conduction and convection
require some physical medium – radiation can occur in a complete vacuum.

Usually thermal radiation is not visible (except when using thermal imaging
techniques), but if the temperature of a body is hot enough it will emit part of
its radiant energy as visible light.

When sufficiently hot the object will first glow cherry red in colour but, if its
temperature is high enough, it will appear white – so-called “white hot”.

If an object absorbs thermal radiation it will heat up. In the Middle East cars
become very hot in direct sunlight, which is why most drivers in hot countries
try to park in the shade.

An object may absorb, reflect or transmit thermal radiation. For instance, many
gases tend to both absorb and reflect very little thermal radiation – mostly they
just transmit thermal radiation (the exceptions are “greenhouse” gases).

Solids, on the other hand (except for glass), tend to reflect or absorb thermal
radiation, since they transmit little or no radiation they are “opaque”. The way a
solid behaves is determined by its surface characteristics.

A perfectly black object, called a “perfect blackbody”, will absorb 100% of all
incident radiation. Matt black objects behave very like “perfect blackbodies”;
however, no body is really perfectly black.

Black objects are good absorbers of thermal radiation. They are also good
emitters of thermal radiation for the following reasons:
 Whenever a photon of thermal radiation is incident to a surface, an
electron on the surface will absorb that photon.
 This causes the electron to rise briefly to a higher energy level.
 When the electron falls back to its original energy level thermal energy
is once again released.
 Since one process is the reverse of the other, it follows that good
absorbers make good emitters.
 Black objects make better absorbers (and emitters) than non-black
objects, simply because fewer photons are reflected.
 The emitted radiant energy has a different spectrum of wavelengths
when compared to the incident radiation – the emitted spectrum
depends on the body’s temperature and it surface characteristics.

4.6.1 Stefan-Boltzmann Law

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 30

White and silver surfaces are very good reflectors of thermal radiation, thus
they are both poor absorbers and poor emitters; in order to retain body heat
when a person is hypothermic, they are often wrapped in insulated silver foil.

Atoms inside a sold surface also emit thermal radiation and this is part of the
heat conduction mechanism.

The rate of thermal radiation from an “ideal blackbody” is given by the Stefan-
Boltzmann Law, as follows:

Eb  T 4 ………………………………………….…(4.18)

Where,
Eb  The blackbody emissive power (W/m2)
  The Stefan-Boltzmann constant (5.67x10-8 W/m2 K4)
T  The absolute temperature (K)

The emissive power of an object is a flux of radiant energy given out by a body
and for a perfect blackbody it depends only on the absolute temperature T
 (W) per unit area A (m2).
(K). The flux is the rate of heat emission Q

No surface is truly an “ideal blackbody”, so that equation (4.18) must be


modified to allow for real non-black behaviour as follows:

E  T 4 ………………………………………….…(4.19)

Where,  is the dimensionless quantity called the “emissivity” and is defined


as the ratio of the emissive power of a non-black surface E to the emissive
power of a perfect black body Eb at the same temperature T .

E
 ………………………………………..…..….(4.20)
Eb

(It is necessary to emphasise the fact that both E & Eb , in the above
expression, must be at the same temperature T ).

Example: 4.6.1

Problem:

Calculate the radiant heat flux from a heating panel, emissivity 0.85, when its
temperature is 50oC and again when its temperature is increased to 150oC.

Solution:

Use equation (4.19) at 50oC:

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 31

E  0.85  5.67  10 8  3234

 E  524.6 W/m2

Then use equation (4.19) again at 150oC:

E  0.85  5.67  10 8  4234

 E  1543 W/m2

…………………………………………………

Example: 4.6.2

Problem:

Calculate the net radiant heat flux from a heating panel, emissivity 0.85, when
its temperature is 50oC and it is completely surrounded by a room at 20oC.

Assume that the surfaces of the room behave as a perfect black body.

Solution:

In this case body “1” (hot object) is emitting thermal radiation, but it is also
receiving thermal radiation from the room “2” (cold object), thus the net heat
flux leaving surface “1” q
 1 (W/m2) is amount leaving less the amount arriving:


q 1   T14  T24 

 q 1  0.85  5.67  10 8  3234  2934 
 q 1  169 W/m2

…………………………………………………

A more detailed treatment of thermal radiation must take into account the
number of surfaces, the “view” that each surface has of each other and the
emissivity of each surface – this is where the resistance approach is helpful.

Thermal radiation is not usually considered in heat exchanger problems unless


the surfaces are at a high temperature, or the overall rate of heat transfer is
slow (such as natural convection), or both.

4.7 Energy Balances on Reacting Systems

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 32

The following are required to properly define the enthalpy of reaction: a


balanced chemical reaction; the T & P of reactants and products must be
fixed; the state of aggregation of each species must be specified.

Standard conditions are usually taken to be a temperature of 298 K (25oC)


and a pressure of 1 atm. When evaluated under these conditions the enthalpy
of reaction is denoted H R .

The superscript  , often ignored, simply denotes that the reaction is carried
out under standard pressure P  . Often the subscript “ R ” is replaced by the
standard temperature – the enthalpy of reaction then becomes simply H 298 .

Doubling the stiochiometric coefficients of a balanced chemical equation does


not change the units of H R (doubling is simply multiplication by a constant);
however, it does double the value of H R .

4.7.1 Extent of Reaction

The extent of reaction is defined as the amount of any species reacting


divided by the stoichiometric coefficient of that species:
 Remember reactants will be consumed – thus, amounts reacted will be
negative as will their stoichiometric coefficients.
 Products will be produced – thus, amounts reacted will be positive as
will their stoichiometric coefficients.
 This is why the extent of reaction  is always positive.

For a batch or non-flow reactor the extent of reaction is defined as follows:

ni  nio
 …………………………………………..(4.21)
i

Where,
  Extent of reaction (mol)
n i  Moles of species i at the end of the batch (mol)
n io  Moles of species i at the beginning of the batch (mol)
 i  The stoichiometric coefficient of species i (dimensionless)

For equation (4.21), it does not matter which reacting species is chosen to
calculate the extent of reaction, this is because the extent of reaction is the
same for all species taking part in the reaction.
For a continuous, or steady-flow, reactor the extent of reaction is defined in
much the same way:

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 33

n  n io
  i ………………………………………….(4.22)
i
Where,
  Extent of reaction (mol/s)
n i  Mole flow of species i leaving the reactor (mol/s)
n io  Mole flow of species i entering the reactor (mol/s)
 i  The stoichiometric coefficient of species i (dimensionless)

The extent of reaction is a key variable for solving material balances around a
chemical reactor. The units for the extent of reaction come from the numerator
of either equation (4.21) or (4.22) – the denominator is dimensionless.

Example: 4.7.1

Problem:

Take the complete combustion of propane in oxygen in a steady flow


combustion chamber, as follows:

C3H8(g) + 5 O2(g) → 3 CO2(g) + 4 H2O(g) H R  2220 kJ/mol

If 2.4 kmol/s of CO2 is produced in the product stream and if there is no CO 2 in


the inlet stream, then calculate the rate of enthalpy change H  between the
inlet and the outlet assuming that both streams are at 25oC.

Solution:

Since the flow of CO 2 entering and leaving the reactor are known, then
equation (4.22) may be used to find the extent of reaction  (mol/s)

n  n
  i io
i

2.4  0
   0.8 kmol/s
3
And,
H  H R  0.8  103    2220 kJ/s

 H  1.776  10 6 kJ/s


…………………………………………………

4.7.2 Hess’s Law

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 34

Hess’s Law can be used to calculate any unknown enthalpy of reaction from
known enthalpies of reaction. The same approach can be used for enthalpies
of formation (see next).

The basis of Hess’s Law is that enthalpy is a state function and, so long as the
initial and final states are the same, all possible calculation paths between
these two states must yield the same enthalpy change:
 If the stoichiometric coefficients are doubled or halved then the same
operation must be carried out on the enthalpy of reaction.
 If the reaction is reversed the sign of the enthalpy of reaction must also
be reversed.
 If different reactions are added or subtracted, then the enthalpies of
reaction must also be added or subtracted and, if the same species
appears on both sides to the equation, simply cancel it out.

Example: 4.7.2

Problem:

Take the following two reactions:

C(s) + O2(g) → CO2(g) H R1  394 kJ/mol

CO(g) + 1/2 O2(g) → CO2(g) H R 2  283 kJ/mol

Use Hess’s Law to find the enthalpy of reaction H R 3 for the following:

C(s) + 1/2 O2(g) → CO(g) H R 3

Solution:

Reverse reaction “2”, call it H R 4 and add it to reaction “1” as follows:

CO2(g) → CO(g) + 1/2 O2(g) H R 4  283 kJ/mol

C(s) + O2(g) → CO2(g) H R1  394 kJ/mol

Adding the two together and cancelling like components leads to

C(s) + 1/2 O2(g) → CO(g) H R 3  111 kJ/mol (-394+283 kJ/mol)


…………………………………………………
The same algebraic manipulations that are carried out on the balanced
chemical reactions are also carried out on the enthalpies of reaction.

4.7.3 Enthalpy of Formation

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 35


The standard enthalpy of formation H f is the enthalpy change associated
with forming one mole of a compound from its constituent elements:
 The elements should be as they commonly occur in nature, e.g. Cl 2
rather than Cl (for sulphur use S, not S8).
 Pure elements, such as H2, C, O2, N2 are defined as having zero
enthalpy of formation.
 “Standard” refers to a standard temperature and pressure of 298 K and
1 atm, respectively – the superscript “  ” is sometimes omitted.
 Extensive enthalpy of formation tables are to be found in various books
(Perry and Green, 2008), (Felder and Rousseau, 2008) and
(Himmelblau and Riggs, 2013).

If the standard enthalpies of formation of all species taking part in a reaction


are known, then the standard enthalpy of reaction may be found from the
expression below:

H R   i H fi ……………………………..…….(4.23)
i

Where,
 i  The stoichiometric coefficient of species “ i ” (dimensionless)
H fi  The standard enthalpy of formation of species “ i ” (kJ/mol)
H R  The standard enthalpy of reaction at temperature TR
(kJ/mol)

Stoichiometric coefficients have a sign and this must be properly applied to


equation (4.23) and any other algebraic expression:
  i is positive for product species
  i is negative for reactant species
  i is zero for inert (non-reacting) species

Hess’s Law is really the basis of equation (4.23); thus, enabling the standard
enthalpy of reaction to be found from standard enthalpies of formation.

For instance write down the standard formation reaction for benzene liquid and
look up the standard enthalpy of formation of benzene:

6 C(s) + 3 H2(g) → C6H6(l) H f  48.66 kJ/mol

The enthalpy of formation of gaseous benzene is H f  82.93 kJ/mol . The


formation of benzene is endothermic (requiring energy). When it is formed as a
liquid it will release energy (as it condenses) and will, therefore, be less
endothermic.

4.7.4 Enthalpy of Combustion

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 36

An enthalpy of combustion is simply an enthalpy of reaction where one mole


of a combustible compound is completely oxidised with oxygen – the basic
assumptions are as follows:
 Reactants and products are all at standard conditions, that is a
temperature of 25oC and a pressure of 1 atm.
 All carbon is oxidised to CO2(g).
 All hydrogen is oxidised to H2O (l).
 All sulfur is oxidised to SO2(g).
 All nitrogen forms N2(g).

The example below is an ideal case – more realistic scenarios will be covered
in later modules.

Example: 4.7.3

Problem:

Calculate the enthalpy of combustion of propane in oxygen using the standard


enthalpies of formation of each species as follows:
H fCO 2  393.5 kJ/mol
H fH 2O  285.84 kJ/mol
H fC 3 H 8  103.8 kJ/mol

Solution:

The proper combustion reaction for propane is given by

C3H8(g) + 5 O2(g)→ 3 CO2(g) + 4 H2O(l) H C

Apply equation (4.23) and then expand to all species (enthalpies of formation
of elements are zero and stoichiometric coefficients have a sign).

H R   i H fi
i

H R   CO 2 H fCO 2   H 2O H fH 2 O   C 3 H 8 H fC 3 H 8

H R  (3)  393.5  (4)  285.84  (1)  103.8

H C  2220 kJ/mol
…………………………………………………

4.7.5 Internal Energy Change of Reaction

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 37

The enthalpy change of reaction is used to solve energy balances around non-
flow, constant pressure batch reactors, or steady-flow continuous reactors.

In order to solve energy balances around constant volume batch reactors, the
internal energy change of reaction is needed. Any property change of reaction,
may be defined as (Smith, Van Ness and Abbott, 2005)

M R   i mi …………………………………….(4.24)
i
Where,
 i  The stoichiometric coefficient of species “ i ” (dimensionless)
M R  The standard property change of reaction at TR
m i  The standard property of species “ i ”

The “standard” superscript symbol  refers to the fixed pressure of 1 atm and
the temperature TR is often taken to be 298 K – at least for the purposes of
tabulating enthalpy change of reaction data.

In reality the standard property change of reaction may be at any temperature


(so long as products and reactants are at the same temperature), but the
pressure is always fixed at 1 atm.

Writing equation (4.24) for standard enthalpy of reaction

H R   i hi
i

But from the definition of enthalpy h  u  Pv , so that this becomes


H R   i u i  Pvi 
i

For reactant and product species that are ideal gases Pv  RT , so that


H R   i u i  RT 
i

Expanding out the brackets on the RHS yields

H R   i u i   i RT
i i

Compare the first term on the RHS with the general definition of any property
change of reaction – see equation (4.24) above.
The first term on the RHS is evidently the internal energy change of reaction

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 38

 U R   i u i
i

Substitute this into the former equation leads to an expression that allows
U R to be calculated from H R , as follows:

U R  H R   i RT ……………………..……..(4.25)
i

The last term on the RHS only applies to gaseous species. For any liquid or
solid species the specific volume will be negligible compared to gaseous
species, thus these solid or liquid species may be ignored in (4.25).

If all the reaction species are only solid or liquid, it follows that

U R  H R

Example: 4.7.4

Problem:

Calculate the relationship between the internal energy change of reaction and
the enthalpy change of reaction for the following reaction:

C3H8(g) + 5 O2(g)→ 3 CO2(g) + 4 H2O(g)

Solution:

Apply equation (4.25) to this reaction – all species are gaseous

U R  H R   4  3  5  1 RT

U R  H R  RT

Take care when solving the above expression – all terms must have same
units.

Internal energy and enthalpy change of reaction terms are usually expressed
in units of (kJ/mol); therefore, “ RT ” must also be expressed in the same
units.

…………………………………………………

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 39

4.8 Tutorial Questions Topic 4

1. The wall of a fermenter was constructed of solid stainless steel with


thermal conductivity 16 W/m K, thickness 5 mm and with surface area 40
m2.

If the temperature of fermentation is 25 oC while the outside atmospheric


temperature is 12°C, calculate the heat transfer rate from the fermenter
to the outside atmosphere.

2. Consider a glass window of 1 m x 1 m, with a constant inside window


pane temperature of 30°C. If the constant inside air temperature of the
room is 20°C (air-conditioned room), then calculate the following:
a) The rate of convective heat gain from the inside of the glass
window to the room ( hair = 10.45 W/m2 K).
b) The rate of radiative heat gain, in one direction, from the window
pane to the room (not the exchange between the window pane and
the room). Take it that glass has an emissivity  = 0.92.

3. Consider the following combustion system

100 mol/s C3H8


gas
At 25oC O2, N2, CO2 and
H2O gas
At 1000oC
Furnace

600 mol/s O2 and


2256 mol/s N2 gas
At 300oC

C3H8(g) + 5 O2(g)→ 3 CO2(g) + 4 H2O(l) H R  2200 kJ/mol

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 40

Assuming complete combustion of all propane and, given the above,


determine the following:
a) Complete the mass balance.
b) Set reference states for specific enthalpy calculations.
c) Prepare and inlet-outlet table for enthalpies, including molar
flowrates for all components (use data from Specific Enthalpy
Table below).
d) Find out all enthalpies at the given conditions
e) Calculate the rate of enthalpy change for the reactor and the
amount of heat transferred from the reaction system.

Use Molar enthalpy tables from standard course textbook (Felder and
Rousseau, 2008).

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1


TOPIC 4. Introduction to Thermal Energy 41

4.9 Bibliography

Felder, Richard M. and Rousseau, Ronald W. 2008. Elementary Principles of


Chemical Processes. 3rd ed. India: Wiley

Himmelblau, David M. and Riggs, James B . 2013. Basic Principles and


Calculations in Chemical Engineering. 8th ed. London: Pearson.

Harker, J.H., Backhurst, J.R., Coulson, J.M. and Richardson, J.F. 1999.
Coulson and Richardson’s Chemical Engineering, Volume 1, Fluid Flow, Heat
Transfer and Mass Transfer. 6th ed. Oxford: Elsevier Butterworth-Heinemann.

Smith, J.M. Van Ness, H.C., Abbott, M.M. 2005. Introduction to Chemical
Engineering Thermodynamics. 7th ed. New York: McGraw-Hill.

Perry, Robert H. and Green, Don W. 2008. Perry’s Chemical Engineers’


Handbook. 8th ed. New York: McGraw-Hill.

©HERIOT-WATT UNIVERSITY…………………………………………………………..June 2014 v1

You might also like