Download Internal Energy Work Heat

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Insulated glazing wikipedia , lookup

Thermal expansion wikipedia , lookup

R-value (insulation) wikipedia , lookup

Temperature wikipedia , lookup

Chemical thermodynamics wikipedia , lookup

Conservation of energy wikipedia , lookup

Thermal conductivity wikipedia , lookup

Heat capacity wikipedia , lookup

Equipartition theorem wikipedia , lookup

Heat transfer wikipedia , lookup

Thermal radiation wikipedia , lookup

First law of thermodynamics wikipedia , lookup

Heat equation wikipedia , lookup

Heat wikipedia , lookup

Thermodynamic system wikipedia , lookup

Van der Waals equation wikipedia , lookup

Second law of thermodynamics wikipedia , lookup

Internal energy wikipedia , lookup

Equation of state wikipedia , lookup

Thermal conduction wikipedia , lookup

Thermodynamic temperature wikipedia , lookup

Heat transfer physics wikipedia , lookup

Adiabatic process wikipedia , lookup

History of thermodynamics wikipedia , lookup

Otto cycle wikipedia , lookup

Transcript
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3. WORK, INTERNAL ENERGY, HEAT &
THE FIRST LAW OF THERMODYNAMICS.
Work
dW  PdV
1st LAW
Clausius
(1850)
Heat
dQ  TdS
dU  dQ  dW
Internal Energy
U
3
3
PV  NK BT
2
2
3.1 Work.
One of the most important ideas in thermodynamics is the concept of work, of a system’s
capability to perform mechanical work in a system transition between equilibrium states.
Continuing with use of the fluid/gas system as the paradigm to stand for general systems
the concept of work can be explored with the aid, once again, of the piston now acting as a
38
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
moving wall with which the contained gas particles may exchange energy, something that
has already been looked at.
Consider the piston system sketched below.
P, V
A
-F
F= PA
x
dx
x
The force acting on the piston due to the pressure P of the gas is
F = PA
And this acts to the right and in mechanical equilibrium (the piston not moving) an equal
but opposite external force –F must act to the left to maintain the position of the piston. If
the position of the piston changes under the external force exerted by the environment
from x to x - dx, thus changing the volume of the gas from
Vi  xA to Vf  x  dx A with dV  Vf  Vi   Adx
The infinitesimal work done on the gas by the external force is
dW  Fdx  PAdx  PdV
The infinitesimal work done on a fluid during an infinitesimal change of volume, dV is then;
đW =-PdV
Note that because in this case (compression) dV is negative the value of –PdV is positive.
39
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
This is an example of a very important sign convention;
work done on a system is positive
To compress the gas work was done ON it by the surroundings.
Reversible processes
It is important to note that in defining the work in this way it is essential to ensure that the
pressure remains well defined at all times and this requires that the process must be
performed quasi statically, that is, it must proceed such that at any instant the system
is in an equilibrium state. This further means that at any instant the change could be
reversed and the system returned to its original state (along with the surroundings in as
much as they are affected by the system). Such a quasi-static process would need to be
performed infinitely slowly and is more usually called a reversible process.
To be reversible there cannot be any loss or dissipation in the system and therefore the
piston must be frictionless. In actuality a reversible process can only be approximately
achieved but the concept is extremely important in enabling the theoretical description of
the thermodynamic process.
In the equation, đW = -PdV , where an infinitesimal amount of work, đW has been
performed (in this case on the system) the bar through the đ indicates the fact that the
amount of work for a given volume change depends on the path chosen ie how the change
dV is made. The change from an initial equilibrium state with volume Vi to a final
equilibrium state with volume Vf can be made in an infinity of ways and any one of them
can be performed quasi-statically and reversibly but requiring different amounts of work to
be performed.
NB The basic difference between đW and dV is simply stated as;
dV represents a change of state from one equilibrium state to another and its value is
path independent.
whereas
đW represents a process and its value is path dependent
dV is said to be a perfect differential whereas đW is an imperfect differential.
40
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
It is generally the case that an infinitesimal change in a function of state (U, T, V, P…) is
represented by a perfect differential while an imperfect differential represents how that
change was carried out (a process or flow). NB W is NOT a function of state!
If a large change is made in the volume it is necessary to find the net effect through
integration and care has to be taken in evaluating any integral in accounting for how the
large changes were made, ie. the path followed. The volume change in going from A to B
is easily evaluated
V  Vf  Vi 
Vf
 dV
Vi
The work performed has to be more carefully evaluated because in the integral
Vf
  Pd V
W 
Vi
In the integrand the value of P will also change and its value at each value of V (or
equilibrium state) will depend on the path taken.
Furthermore there is no “initial work” Wi or “final work” Wf the concept is meaningless as it
is a process that is being considered. There will be an amount of work done during the
process represented by the integral but caution is required here as depending on the path
chosen to get from i to f the value of P inside the integral will change in different ways and
needs to be included as part of the integral.
A
P
A
P
PB=PA/2
C1
C3
C2
C3
O
B
V
VA
41
VB=2VA
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
The PV diagram shown above, includes the start and end point of a process taking an
ideal gas from equilibrium point A (PA , VA) to the equilibrium point B (PA/2, 2VA). There are
three processes (pathways) indicated and each is examined in turn to find the work done
in going from A to B.
Example 1.
C1 shows the direct route from A to B (on the diagram anyway). To find the work the first
requirement is to know how P can be written in terms of V in order to evaluate the integral
of PdV. In general the equation of the straight line joining (x1 , y1 ) to (x2, y2) may be written
as
y = y1 + [(y2 - y1) / (x2 - x1)]·(x - x1)
Using this the equation of the line C1 is
 PA  P 
A
P
P
3
2
P  PA  
(V  VA )  PA  A (V  VA )  PA  A V
 2V  V 
2VA
2
2VA
A
A


Therefore the integral may be written
VB
W 

VB
3
 PdV   PA
2
VA

VA
VB
P
dV  A
2VA

P
3
VdV   PA VB  VA   A
2
2VA
VA
V 2 V 2 
 B
A 


2


Knowing VB = 2VA this can be tidied up to obtain
3
3
W   PAVA   nRT
4
4
Example 2
C2 shows the isotherm. The equation of state for the ideal gas may be used to find P as a
function of V. As the temperature is constant along C2,
The integral is then
42
P
nRT
V
Thermal & Kinetic Physics: Lecture Notes
VB
W 

©
Kevin Donovan
VB
 PdV  nRT
VA

dV
V
V

 nRT lnV VB  nRT ln B

V
A
V
A

VA
W  nRT ln 2
Example 3
The path C3 is in fact an isochore (V = VA = constant) followed by an isobar (P = PA/2 =
constant). This can be broken into two integrals and
VA
W 

VA
VB
P
 PdV  A
2
 dV
VA
The first integral is zero as V = 0 and the second is easily evaluated
P
P V
nRT
W   A VB  VA    A A  
2
2
2
In summary
C1
W  1.5nRT
C2
W  nRT ln 2  0.69nRT
C3
W  0.5nRT
Going from A to B necessitated the system doing work on the surroundings (all processes
have negative values for the work). This was of necessity the case as the gas expanded!
The work expended by the system was different for each path. Being the least for C3 and
the greatest for C1.
43
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
VB
Another view of the work can be seen from consideration of the integral
  PdV and the
VA
PV diagram.
Isochore
P
Isotherm
Isobar
V
VB
The integral   PdV is seen to be the negative of the area under the line of the path
VA
between the path and the P = 0 volume axis and looking at the above pressure volume
indicator graph it is easy to see the results obtained earlier for C1 and C3 by geometric
area evaluation. This is not so easily done for path C 2 the isotherm. To get the sign correct
it is necessary to take note of the direction in which the path is traversed. If going from left
to right the sign convention, positive for work done on a system, the area is taken as
negative. Proceeding from right to left or high to low volumes the area is positive as the
gas is compressed and work is therefore done on the system.
Frequently in this module cyclic processes will be of interest and importance and these
may be represented on the PV diagram and the work carried out by or done on the system
can be calculated over a cycle where the system at an equilibrium state A is caused to
pass through a series of points equilibrium points before returning to A.
44
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
Example 4.
Isotherm
P
3
Isochore
Isobar
2
2P1
P1
1
V1/2
V
V1
An ideal gas is compressed isothermally from V1 to V1/2 on the path 1  2. It is then
expanded isobarically at 2P1 back to its original volume on path 2  3 before finally being
returned isochorically to its original equilibrium state on the path 3  1
What is the work done by/on the gas?.
We need to evaluate and add three integrals, one for each path.
Path 1  2 is an integral evaluated in an earlier example
V2
W1 2 

V1
V2
 PdV  nRT

dV
V
V
 nRT lnV V2  nRT ln 1   nRT ln 2
1
V
 V2 
V1
Path 2  3 is isobaric and P is constant (again this was done earlier)
V
W2  3  P2 V1  V2   P2 1  P1V1  nRT
2
And path 3  1 is isochoric, V = 0 and therefore no work is done
45
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
W3 1  0
The total work done in the closed cycle is therefore
Wtot  W12  W23  W3 1  nRT ln 2  nRT  0
Wtot  nRT (0.69  1)  0.31nRT
Note the signs of W in each of the arms of the cycle recalling that positive means work is
done on the gas. The net work is negative meaning that overall in the cycle work is carried
out by the gas.
If the cycle is now reversed;
W13  0
V
W3  2  P2 V2  V1   P2 1  P1V1  nRT
2
V1
W2 1 

V1
 PdV  nRT
V2

dV
V
V
 nRT lnV V1  nRT ln 1   nRT ln 2
2
V
 V2 
V2
This time the net work is
WTot  W13  W3 2  W21  0  nRT  0.69nRT
WTot  nRT 1 0.69  0.31nRT
and is positive, ie. the system has work done on it by the surroundings.
In fact, when the work extracted in one direction is equal to the work done in the other
direction. The cycle is said to be reversible.
46
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3.2 Internal Energy.
Internal energy changes during the cycle may now be dealt with having looked at the work.
Previously it was found for an ideal monatomic gas (no rotation or vibration) that the
internal energy was given by the expression U 
3
PV
2
This may be used to find the changes in U as the cycle is traversed, first in the clockwise
sense from 1  2  3
U1 2  U 2  U1 
3
P2V2  P1V1   3 nRT2  nRT1   0
2
2
Because 1 and 2 lie on an isotherm and T1 = T2
For the next stage
U 2 3  U 3  U 2 
3
P2V1  P2V2   3 2P1V1  P1V1   3 2nRT1  nRT1   3 nRT1
2
2
2
2
And for the final stage
U 3 1  U1  U 3 
3
P1V1  P2V1   3 P1V1  2P1V1    3 nRT1
2
2
2
The net change in internal energy around the cycle is then
U1 2  U 2 3  U 3 1  0 
3
3
nRT1  nRT1  0
2
2
In performing this analysis two important things have been learned about the changes in
internal energy after undergoing processes of a certain type;
i)
Around a closed loop there is no change in internal energy. This is because
there is no net temperature change and therefore the molecules cannot be going
slower or faster otherwise continuous looping would lead to infinite or zero velocities.
More succinctly, there is no change in U because it is a STATE VARIABLE and its
47
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
value is representative of the particular equilibrium state. Unlike W which is not a
state variable and is not therefore associated with any particular equilibrium state but
rather with the process of going from one state to another.
ii)
Also note that equilibrium states on an isotherm all have the same internal
energy and U12  0 along the isotherm.
It is obviously true that U is zero if the loop is traversed in the opposite sense (try it) for
the same reason, U is a state variable. In fact it is possible to be completely general,
whatever the closed loop from A back to A there will be no change in U because it is a
state variable.
P
CU
2
1
CL
V
V1
The above arbitrary closed cycle shows the general situation
V2
U   dU  0
C
For W we can find the work done after one cycle by finding the area of the upper and
lower curve and subtracting the lower curve, work done by the system, (going from left to
right with volume increasing) from the upper curve, work done on the system, (going from
48
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
right to left with volume decreasing) to obtain the net work done on the system. It is the
area enclosed by the circuit.
If the circuit is traversed in the other sense then the upper and lower curves would have
their signs altered and work done on the system would be minus the area enclosed.
W = +Area Enclosed (counter clockwise)
W = -Area Enclosed (clockwise)
Around a closed cycle;
i)
The work done depends on direction traversed and on the particular cycle
chosen. It is path dependent and dW is an imperfect differential.
ii)
The internal energy is unaltered independent of the closed cycle. U is a state
variable and dU a perfect differential.
49
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3.3 Heat.
In the eighteenth century the popular view of heat was that it had the nature of a fluid and
would flow from hot bodies to cold bodies. This fluid, caloric, could neither be created or
destroyed and every body contained a certain amount of caloric depending on its
temperature. This view allowed explanations of many phenomena and was almost
universal although a few notable scientists including Boyle and Hooke held a contrary
position that heat was an expression of some form of motion, the so called kinetic theory of
heat. Indeed the phrase heat flow persists to this day and a Calorie is an alternative unit of
energy much favoured still by chemists.
Benjamin Thompson aka Count Rumford (1753 – 1814), became the
most well known opponent to the caloric viewpoint due to his experiments
on frictional heat developed during the boring of cannons in the Munich
arsenal which was published in 1798. In this work he used a blunt borer
and with the heat produced was able to boil large quantities of water
implying, from the caloric viewpoint, a limitless supply of the fluid. These
experiments along with other later experiments finished off the idea of heat as an actual
physical fluid. It was the careful quantitative work of Joule that finally put paid to the idea of
caloric and enabled the kinetic theory of heat to be placed on its pedestal.
James Prescott Joule (1818 – 1889)
Demonstrated the relationship between heat and mechanical work.
In so doing the foundations of the law of conservation of energy were
finally established almost 150 years after Leibnitz first proposed it in his
theory of vis-viva.
Adiabatic Work In 1843 Joule demonstrated in a classic set of carefully carried through
experiments that if he enclosed a system within adiabatic walls (insulating walls allowing
no heat transfer into or out of the system) then by doing a known amount of work on the
system, that starting from a particular equilibrium state the final end state depended only
on the work done and not on how the system was taken from the initial to the final state.
eg. stirring with paddles
or
heating with a known current and resistance
50
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
i
i
He showed that the temperature rise T  U = W the adiabatic work done on the system.
In fact this is the empirical definition of the internal energy
WAdiabatic  Ufinal  Uinitial
3.4 Heat and the First Law.
The case of a thermally/adiabatically isolated system is a special one and the observations
noted previously have important implications when followed through.
When the system is not adiabatically isolated then it is the case that
U  W
This is of great consequence since at first sight it implies non-conservation of energy! By
now the conservation of energy had taken its place in the pantheon of physical principles
as an inviolable requirement and this position could only be reversed at great cost to the
existing understanding of the universe. It was now an item of faith that the amount of
energy in the universe was a constant and unchanging, only the form in which it was
51
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
present altered, it’s actual quantity was God given, from the start of the universe and could
not be added to or subtracted from by any mechanism. In this respect it had been a late
comer to the group of conserved quantities which included momentum and angular
momentum (Newton), again fixed quantities which were present at the creation and
continued unaltered, and mass (Lavoisier). To maintain this inviolability of energy
conservation in the face of the evidence of careful thermodynamic measurements where
heat flows were allowed to occur it was recognised that in thermodynamic systems the
concept of heat flow, Q was essential and that Q was just another form of energy.
Heat flow like work however was a form of energy in transit. No body could be said to hold
a quantity of work or heat, the two only existed as transient forms while energy changes
from one form to another. Thus the First Law of Thermodynamics was born. The
change in internal energy after some process is carried through is given by
U  Q  W
W is the work done on the system in that process and Q is the heat transferred to
the system by that process.
A simple and early statement of the first law is that the change in the internal energy of a
closed system is equal to the amount of heat absorbed by the system, minus the
amount of work done by the system on its surroundings. Given the sign convention that
we are using this statement corresponds to the preceding formulation of the first law in the
form of an equation.
The first comprehensive statement of this most important of laws was made by Clausius in
1850 some seven years after Joule’s experiments.
Note that Q like W is not a state variable and should be formally written with the
strikethrough across the  or d. dQ and dW are both path dependent and imperfect
differentials unlike dV, dP , dT and dU.
The sign convention must be upheld when using the first law.
i)
Earlier the convention that W was positive was established when the system
has work done on it and negative when it is the system that does work.
52
Thermal & Kinetic Physics: Lecture Notes
ii)
©
Kevin Donovan
Similarly the convention for heat flow is that Q is positive when it flows into
the system and negative when it flows from the system.
The first law may also be written in infinitesimal form as
dU  dQ  dW
In the infinitesimal form the first law holds for all processes reversible and irreversible.
When the process is reversible the expression(s) for dW may be used to cast the first law
in the form
Fluid
dU  dQ  PdV
(Reversible)
The heat flow must be path dependent as the work is path dependent and the first
law must hold for any path.
B
Q

dQ
A
Returning to the previous example to look at heat flow around a closed cycle using the first
law.
53
Thermal & Kinetic Physics: Lecture Notes
Isobar
2
P
Kevin Donovan
3
Isochore
2P1
©
Isotherm
P1
1
V1/2
V
V1
V
Q1 2  U1 2  W1 2  0  nRT1 ln 1  nRT1 ln 2
V2
This is negative on going from 1  2 and means that heat flows out from the system.
Q2 3  U 2 3  W2 3 
3
5
5
P1V1  ( P1V1)  P1V1   nRT1
2
2
2
This is positive and on going from 2  3 heat flows into the gas.
3
3
Q3 1  U 3 1  W3 1   P1V1  0   nRT1
2
2
This is negative and therefore on going from 3  1 heat flows out of the gas.
The net heat flow around the cycle is
Qnet  Q12  Q2 3  Q3 1  nRT ln 2 
54
5
3
nRT1  nRT1  nRT1(1  ln 2)
2
2
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
This is positive and therefore there is net heat flow into the gas if this closed cycle is
traversed. Previously the net work was found to be;
Wnet  nRT (ln 2  1)  0
The net work was done by the gas and is equal to the net flow of heat into the gas. They
balance one another leaving U = 0 as required after traversing a closed cycle and in
order to satisfy the first law of thermodynamics.
A Test for Perfect Differentials
The differences between perfect and imperfect differentials have been mentioned
frequently with respect to state variables such as internal energy which are examples of
the former and work and heat, which are associated with particular processes therefore
depending on the path taken from initial to final state and are examples of the latter. An
important test for a perfect differential may now be established as follows;
Suppose z = z(x, y) is a function of state which depends on two independent variables, x
and y.
Then
 z 
 z 
dz    dx    dy
 x  y
 y  x
And hence
dz  a( x, y )dx  b( x, y )dy
For dz to be a perfect differential requires by definition that
 2z
 2z

yx xy
implying
a( x, y ) b( x, y )

y
x
55
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
Given a differential in the form;
dz  a( x, y )dx  b( x, y )dy
Such that
a b

y x
Then the differential is perfect and can be integrated to give z independent of the path, and
z is a function of state.
If
a b

y x
Then the differential is imperfect and the integral of dz depends on the path. z is not a
function of state.
Example.
U
An ideal monatomic gas,
i)
3
PV
2
3
 U 
 U 
dU  
 dP  
 dV  VdP  PdV 
2
 P V
 V  P
3
a V
2
b
3
P
2
a 3 b 3
 

V 2 P 2
U is a perfect differential.
ii)
dW  PdV
First law
dQ  dU  dW 
3
VdP  PdV   PdV  3 VdP  5 PdV
2
2
2
56
Thermal & Kinetic Physics: Lecture Notes
3
a V
2
b
©
Kevin Donovan
5
P
2
a 3 b 5
 

V 2 P 2
dQ is not a perfect differential
Applications of the First Law may now be examined.
i)
Heat Capacities.
There are many simple applications of the first law for example it may be used to
define heat capacity.
If an amount of heat Q is introduced to a finite system there will be an increase in
temperature, T and the heat capacity is defined as
QRe v dQRe v

dT
T 0 T
C  limit
(Reversible change)
It is defined when the change in temperature is a reversible one, that is to say at
each intermediate temperature the new state is an equilibrium state obeying the state
equation. In this case the input or output of heat will raise or lower the temperature by
the same amount.
The specific/molar heat capacity is the heat capacity per unit mass/mole,
c
1 dQRe v
C

m dT
m
cm 
1 dQRe v C

n dT
n
The value of C depends on the type of process in which the heat was transferred.
a)
Constant Volume Heat capacity
QRe v
T 0 T V  const
CV  limit
57
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
From the first law
U  Q 
 PdV  Q
(V = const, dV = 0)
This allows for a measure of the specific heat capacity for a gas at constant volume
that involves only state functions;
U
3
 U 

  nR
2
T  0 T V  const  T V
CV  limit
b)
Constant Pressure Heat Capacity
Q
T  0 T P  const
CP  limit
From the first law
U  Q 
 PdV  Q  (PV )
so for the heat
Q  U  (PV )  (U  PV )
A new state function, Enthalpy, H is defined
H  U  PV
The infinitesimal change in H may be found as follows;
dH = dU + PdV + VdP
And for isobaric processes where dP = 0 the heat change is given by;
dQ  dU  PdV  dH
Therefore a useful definition of the specific heat capacity for a gas at constant
pressure that involves only state functions is
58
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
Q
H
 H 
CP  limit Re v  limit


T
T  0 T P  const  T  P
U
H  U  PV 
3
3
PV  nRT
2
2
3
5
5
PV  PV  PV  nRT
2
2
2
3
 U 
CV  
  nR
 T V 2
5
 H 
CP  
  nR
 T  P 2
And the relationship between CP and CV is then
CP  CV  nR
It should be noted that all of the above only applies to a monatomic gas, (no
potential, rotational or vibrational contributions to U) where U 
3
3
PV  nRT .
2
2
Taking a closer look at ENTHALPY the newly defined state function
H  U  PV 
3
5
5
PV  PV  PV  nRT
2
2
2
dH  dU  PdV  VdP
In situations where the pressure is constant, and these situations arise frequently in liquids
or solutions which are often at atmospheric pressure;
dH  dU  PdV  dQ
59
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
And the natural variables of the state function, H, in an isobaric situation are U (or T )
and V where by natural variables it is meant those variables which when held constant
lead to no further change occurring in that state function ie. for the state function enthalpy
if U and V are constant, dU = dV = 0 and the enthalpy is constant. A further example is for
the internal energy of an ideal gas where U 
3
3
PV and dU  PdV  VdP  .
2
2
Clearly it is V and P that are the natural variables of U as when dV = dP =0 then
dU = 0.
If there is heat flow during a chemical reaction it is equal to the enthalpy of that reaction
and the enthalpy is widely used in physical chemistry and is tabulated for common
reactions.
In the definition of heat capacity the need for a reversible heat flow process was stressed.
This is similar to reversible work previously encountered as it means the heat flow is
induced by infinitesimal temperature differences rather than by finite difference in
temperature, ie. it occurs quasi statically.
An example of irreversible heat flow is the dropping of a hot metal (metals conduct heat
well and therefore the temperature of the metal will be uniform from inside to out and
remain so) into a bowl of water at room temperature. The metal will rapidly transfer heat to
the water in an irreversible process passing through non-equilibrium intermediate states.
There is no way to take the metal back to its original (hot) state by infinitesimal changes to
the environment.
An example of reversible heat flow is to place a finite system eg. the bowl of water in
contact with a heat reservoir at a slightly different temperature (higher or lower). This could
raise/lower the temperature of the water by dT before iterating with a reservoir at a slightly
higher/lower temperature than the previous one. In this way the water could be taken from
one equilibrium state to another at a higher/lower temperature in incremental steps,
reversibly.
At this point the HEAT RESERVOIR which is a very important and useful concept in
thermodynamics, may be defined. Recall that, as stated earlier, it is wrong to think of an
object as containing heat, where heat like work is the result of a process and is correctly
termed heat flow. The heat reservoir is a body of such large heat capacity that a flow of
heat Q to or from the system of interest, whilst changing the temperature of that system
60
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
will have no effect in changing the temperature of the reservoir. Using a series of heat
reservoirs at different temperatures heat may be added to or removed from a system
reversibly at any specified temperature.
ii)
Adiabatic Process for an Ideal Gas.
Beginning with the analysis of an ideal gas undergoing a reversible adiabatic change
(expansion/compression) and starting with the first law,
dU  dQ  dW  dQ  PdV
In an adiabatic process dQ = 0 by definition.
whence
dU  PdV
For an ideal monatomic gas it has already been established that U 
And therefore
dU 
3
PdV  VdP   PdV
2
Re-arranging
5
3
PdV  VdP  0
2
2
and
5 dV
3 dP

2 V
2 P
5 dV
dP

3 V
P
By integration
5
lnV   ln P  const
3
Rewritten using
n ln x  ln( x n )
 5 
ln P  lnV 3   const


Combining the logs
61
3
PV
2
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
5
PV 3  PV   PV V  1  nRTV  1  const
We have then demonstrated that at all points on the adiabatic PV curve (where in this case
T does vary) are linked through the relationship;
5
PV 3  PV   const
and
TV  1  const
that remain true.
NB the constants in the two equations are not the same!
This problem could have been approached in a different fashion;
At constant volume
U  U(T )
Meaning that
 U 
dU  
 dT  CV dT  PdV
 T V
CV dT  
CV
(1st Law with dQ = 0)
nRT
dV
V
dT
dV
 nR
T
V
Integrating and using the relationship CP  CV  nR
CV lnT  nR lnV  const  (CV  CP ) lnV  const
C

lnT   P  1 lnV  const
 CV

62
Thermal & Kinetic Physics: Lecture Notes
CP
TV
CV 1
©
Kevin Donovan
 TV  1  const
With the important exponent  given by

CP
CV
From a previous result;
3
5
 U 
 H 
CV  
  nR and CP  
  nR
 T V 2
 T  P 2

The exponent  
CP 5

CV 3
5
is thus placed on a more physical basis. It should be noted that the
3
equation of state U 
3
PV is only true for monatomic gases and therefore the exponent,
2
, will be different for non-monatomic gases.
3.5 Diatomic Gases.
All
of
the
foregoing
was
related
to
monatomic
gases
where
the
relations
U
3
3
PV  nRT were found by considering the translational momentum exchange of a
2
2
molecule at a container wall and how this contributed to the pressure. Turning now to the
case of diatomic gases in order to be able to compare and contrast their behaviour with
that of the monatomic gas.
Previously it was shown that for the monatomic gas, where there is only translational
energy to be considered;
U N
Note that
1
mv x2
2
 1
1 2
1
1
mv
 N  mv x2 
mv y2 
mv z2
2
2
2
 2
 3
  PV
 2
represents the energy available in one of the 3 degrees of freedom
available to a spherically symmetric atom which comprises a monatomic gas.
63
Thermal & Kinetic Physics: Lecture Notes
 1
1
1
U  N  mv x2 
mv y2 
mv z2
2
2
 2
n
N
NA
and
©
Kevin Donovan
 3
3
  PV  nRT
2
 2
R
 kB
NA
 1
1
1
U  N  mv x2 
mv y2 
mv z2
2
2
 2
 3
  Nk BT
 2
We note here without formal proof the equipartition theorem, that it is the case that for
each degree of freedom available to the atom there is
1
k BT average thermal energy. Ie.
2
1
1
mv x2  k BT
2
2
1
1
With s degrees of freedom U  s NkBT  s nRT
2
2
With this reminder it is now possible to ask what happens if more degrees of freedom are
allowed or in other words if more independent or orthogonal ways in which a constituent
part of a system may hold energy are included.
Examples of this are diatomic gases in two special circumstances;
i)
A gas composed of rigid diatomic molecules where by rigid it is meant that
there is a constant internuclear separation.
z
y
x
64
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
The molecule shown as a dumbbell is now able to rotate about the x, y and z axes with
consequent rotational kinetic energies;
1
I z 2
2
1
I x 2
2
1
I y 2
2
respectively where I is the moment of inertia.
With the molecule aligned along the z axis and the atoms treated as point like objects the
moment of inertia Iz << Ix = Iy and there are effectively two new degrees of freedom added
to each molecule.
For each of these
1
1
I x 2  k BT
2
2
and
1
1
I y 2  k BT
2
2
The internal energy of a collection of N of these rigid diatomic molecules is
U  N  ETrans  E Rot
  N 
1
1
1
1
1
mv x2 
mv y2 
mv z2  I x 2  I y 2
2
2
2
2
 2



The
number of degrees of freedom is now s = 5 and the equations obtained previously for a
monatomic gas where s = 3 are now modified;
U s
1
1
1
5
NkBT  s nRT  s PV  PV
2
2
2
2
CV is now given by;
1
5
 U 
CV  
  s nR  nR
2
2
 T V
And for H = U + PV
H  U  PV 
5
7
7
PV  PV  PV  nRT
2
2
2
So for CP
s2
7
 H 
CP  
nR  nR
 
2
2
 T  P
and
65
Thermal & Kinetic Physics: Lecture Notes

©
Kevin Donovan
CP s  2 7


CV
s
5
For an adiabatic process T and V are always related by
CP
TV
ii)
CV
1
2
 TV 5  TV  1  const
A gas composed of vibrating diatomic molecules where the condition of
rigidity were to be relaxed and the diatomic molecule was allowed to vibrate.
z
y
x
The vibration occurs along the axis joining the two atoms and this motion has potential
energy and kinetic energy associated with it, adding a further two degrees of freedom
and a further k BT per molecule to the internal energy of the gas which is now
U
7
7
7
Nk BT  nRT  PV
2
2
2
The number of degrees of freedom is now s = 7
And the changes to CV , H and CP that follow this change in s and in internal energy are
now;
1
7
 U 
CV  
  s nR  nR
2
2
 T V
and
66
Thermal & Kinetic Physics: Lecture Notes
H  U  PV 
©
Kevin Donovan
7
s2
9
PV  PV 
PV  nRT
2
2
2
with
`
s2
9
 H 
CP  
nR  nR
 
2
2
 T  P
and

CP s  2 9


CV
s
7
and finally for an adiabatic process on a gas of non rigid diatomic molecules
CP
TV
CV
1
2
 TV 7  TV  1  const
67
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3.5 Heat capacity of classical ideal gases
Equation of state PV = nRT
Monatomic Gas
3
nRT
2
U
CV 
H
5
nRT
2
U
3
nR
2
H
5
nR
2

U
7
nRT
2
CV 
7
nRT
2
H
CP 
CP 5

CV 3
Vibrating Diatomic Gas
5
nR
2
CV 
5
nRT
2
CP 

Rigid Diatomic Gas
7
nR
2
9
nRT
2
CP 
CP 7

CV 5

7
nR
2
9
nR
2
CP 9

CV 7
Internal energy due to
Internal energy due to
Internal energy due to
3 translational DoF .
3 translational DoF
3 translational DoF
s=3
plus 2 rotational DoF.
plus 2 Rotational DoF
s=5
plus 2 Vibrational DoF.
s=7
For more complex gases there may be even more degrees of freedom and the
measurement of
CP
CV
may provide valuable insights into the make up of the gas
molecules.
It should be remarked upon here that classically all is fine with the foregoing and there are
nice relationships between specific heat capacities and the numbers of degrees of
freedom. However if the behaviour of the heat capacity of an actual diatomic gas is
examined it would be found that the heat capacity would depend on temperature, a fact
that is not reflected in the preceding classical model. To explain this quantum mechanics
must be appealed to where it is known that the energies of rotation and vibration are in fact
quantised;
For vibrational states;
1

EVib   n  
2

68
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
Where  is the frequency of oscillation of the simple harmonic oscillator



With  the spring constant (related to bond strength) and  
m1m2
is the reduced
m1  m2
mass and n the vibrational quantum number or occupancy of the vibrational level. The
quantised vibration behaves as a boson and therefore it is not restricted by Pauli’s
exclusion principle and any state (vibrational frequency) may have multiple occupancy, the
quantum equivalent of an increased classical amplitude of oscillation.
For rotational states;
E Rot 
J (J  1) 2
2I
where J is the rotational quantum number and I the moment of inertia.
For both of these quantised energies, if the unit of quantisation is larger than
1
k BT , the
2
process is effectively unavailable to the molecule and the degree of freedom therefore
unavailable also. This means that as temperature is lowered the extra processes are
“frozen” out and the internal energy is dropped in an almost stepwise fashion causing the
specific heats to drop “suddenly”. A non rigid diatomic molecule will first lose the two
degrees of freedom due to vibration as temperature is lowered before losing the two
rotational degrees of freedom. There is always the zero point vibrational energy
1
 , but
2
this does not contribute to the heat capacity as it doesn’t contribute to a rise in internal
energy as heat is introduced into the system. ie. it is always there irrespective.
69
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3.6 Adiabatic Processes.
Adiabatic processes, as seen, are described by the equation
PV   const
P
const
P
nRT
V
V
cf Isothermal processes
PV  nRT  const
For adiabatic changes the rate of change of pressure with volume or slope of the adiabatic
line on a P V diagram is found;
dP  const  PV   P



dV
V
V  1
V  1
cf Isothermal changes where;
dP  nRT  PV
P



dV
V
V2
V2
Adiabatic
Change
P
Isothermal
Change
V
It is to be noted that the magnitude of the local slope in a PV diagram is always greater for
adiabatic change than for isothermal change,
 P
P

V Adiabatic
V
70
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
Looking now at actual gases to see how well the foregoing discussion describes them;
Gas

CV
CP
Oxygen, O2
0.658
0.981
1.49
Nitrogen, N2
0.743
1.04
1.40
Hydrogen, H2
10.2
14.3
1.402
Carbon Monoxide, CO
0.744
1.04
1.398
Water, H2O
1.4
1.86
1.33
Carbon Dioxide, CO2
0.657
0.846
1.287
Sulphur DioxideSO2
0.47
0.60
1.276
Ammonia, NH3
1.66
2.15
1.295
Benzene, C6H6
0.67
0.775
1.157
Table of specific heat capacities for several monatomic, diatomic, triatomic and
more complex gases. The values are taken at high temperature where all the
degrees of freedom are allowed to play a role, ie. kBT > Equantum
71
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
70
CO2
H2O
NO2
60
CP(kJ/m/K)
50
O2
H2
CO
40
30
N
O
H
20
10
0
1000
2000
3000
4000
5000
T (K)
Graph showing variation of CP with temperature for a variety of monatomic,
diatomic and triatomic gases.
72
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3.7 The Van Der Waals Gas.
Having looked at other types of “ideal” gas, (rigid diatomic and non rigid diatomic) and their
behavior’s, attention may now be turned to real gases. The most well known of these is the
van der Waals gas, where the modified equation of state discovered empirically by Dutch
physicist Johannes van der Waals (1837 – 1923)
was found to describe well the P-V behavior of real gases with the ideal equation of state,
PV = nRT , being replaced by
2 

 P  n a V  nb   nRT

V 2 

In this equation of state designed to describe real gases with potential interactions the
pressure,, P, has a modification;
PP
n 2a
V2
Atoms/molecules in the bulk of the gas have no net force imposed upon them by the
presence of the intermolecular force as they are pulled equally from all sides but the
atoms/molecules adjacent to the walls will have a net force pulling on them into the bulk of
 N 1

 a
the gas/fluid and therefore e reduction in P. The term
2
V
 NA V 
n 2a
2
represents a
decrease on the pressure by an amount proportional to the square of the number density
of particles in the system. If it is added to the actual pressure P the adjusted pressure is
what would have been exerted by an ideal gas and allows the ideal gas equation of state
to be applied.
73
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
In the second term, the volume V is reduced by an amount b per mole in order to account
for the finite size of the atoms themselves and their collisions with each other. The
atoms/molecules have a size assigned to them described by the van der Waals radius
rvdW . Recalling that an atom is largely empty space this radius is determined by the
extent of the interatomic potential. If the centres of two molecules approach within a
distance 2rvdW they are deemed to have collided and each molecule has an “excluded”
volume assigned to it
Vex 


3
1 4
4 3 
  2rvdW  4   rvdW
.
2 3
3

It should be noted that the excluded volume per molecule is 4x the volume of the
molecule.
The total excluded volume is
4 3
VTot ex  4N  rvdW
3
and the actual volume of the container is effectively reduced by this amount and
V  V  VTot ex  V  nb in the van der Waals equation allowing the ideal gas equation
thus modified to be used. b may be found by experimental determinastion of the P-V
diagram and then rvdW
4 3
N
VTot ex  4N  rvdW
 nb 
b
3
NA
1
 3 1  3

rvdW   b
 16 N A 
This has been the most successful modification made to the ideal gas equation of state
and is still widely used.
The internal energy of the van der Waals gas.
When T = 0 and kinetic energy is by definition zero, in an ideal gas the pressure would go
to zero as well as it is the direct result of the exchange of molecular momenta with the
walls in which the gas is contained. However, the vdW equation of state reduces to
74
Thermal & Kinetic Physics: Lecture Notes
P
©
n 2a
Kevin Donovan
(T = 0)
V2
Where it is seen that the pressure is finite and negative at zero temperature and is
proportional to the square of the density of molecules, and this is physically the result of
the net attraction of surface molecules towards those in the bulk.
Beginning with an extreme dilution whose energy is taken as the reference energy level
and compressing this gas the work required to achieve the compression may be simply
calculated,
V
V

W   PdV  n 2a



dV /
V /2
integrating,
V
n 2a
W  n a

V
V/ 
2
1
ie. after doing this work at T = 0 the gas has some energy which is taken to be the stored
molecular potential energy at a volume V, not available to ideal gases! This means that the
total internal energy of a vdW gas has both, kinetic and potential energy contributions
3
n 2a
UvdW  nRT 
2
V
This is for monatomic VdW (real) gases with three translational degrees of freedom.
NB, from the above expression for UVdW and the equation for the heat capacity at constant
volume
 U
CV (vdW )   vdW
 T
75
3

  nR
V 2
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
is unchanged from that of a real gas. This is of course as it should be as the increase in
heat flow should only affect the kinetic part of the internal energy and not the potential
energy part.
It is to be noted in passing that the internal energy is now a function of 2 state variables
unlike previously;
UvdW  U(T ,V )
UIdeal  U(T )
The internal energy of the vdW gas may be recast using the equation of state to obtain it in
terms of pressure and volume;
U vdW 
2
3n 
n 2 a  V
 n a
P

b



2 
V

V 2  n
NB the factor n has been taken outside of the second bracket on the RHS.
Now expanding the brackets[;
UvdW 
3n  PV
na n 2 ab  n 2a
 Pb 



2  n
V
V 2  V
Simplify
3
1 n 2 a 3 n 3 ab
UvdW  P V  nb  

2
2 V
2 V2
and
UvdW  U(P,V )
76
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
The Enthalpy of a van der Waals gas
The enthalpy is defined in the usual way,
HVdW  UVdW  PV
The terms in the equation of state may be multiplied out and the equation rewritten as
PV  nRT 
 n 2a 
n 2a
(nb)
 nbP  
 V2 
V


The enthalpy may then be written as;
HVdW 
5
n 2a
n 3 ab
nRT  2
 nbP 
2
V
V2
The adiabatic law for van der Waals gases.
With the equation of state of the Van der Waals gas ;
2 

 P  n a V  nb   nRT

V 2 

it is sensible to ask if an equation describing an adiabatic process may be found,
5
equivalent to that which has already found for an ideal gas, PV 3  PV   const (where
5/3 indicates that it is a monatomic gas that is being considered).
Previously the internal energy of the van der Waals gas, UvdW 
3
n 2a
was
nRT 
2
V
obtained. To find the adiabatic law the first law is used as follows;
dU  dQ  dW  0  PdV
Finding dU from the equation for UVdW , and P from a rearranged equation of state with
pressure as the subject
77
Thermal & Kinetic Physics: Lecture Notes
P
©
Kevin Donovan
nRT
n 2a

V  nb V 2
and using these in the first law
3
n 2a
 nRT
n 2a
dU  nRdT 
dV  PdV 
dV 
dV
2
V  nb
V2
V2
3
 nRT
nRdT 
dV
2
V  nb
3 dT
dV

2 T
V  nb
By integrating the above
3
lnT fi  ln(V  nb)fi
2
Re-arranging
T
ln f
 Ti
3
 2
V  nb 
   ln f


Vi  nb 
Therefore re arranging
3
ln
Tf 2 Vf  nb 
3

Ti 2 Vi  nb

 const
And finally
3
T 2 V  nb   const
is the adiabatic rule for a van der Waals gas.
78
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
3.8 Free expansion and Joule Kelvin effect.
i)
Joule Free Expansion.
Consider now another irreversible process in order to see its consequences for ideal and
real gases. This is the Joule free expansion illustrated below, where a gas is allowed to
expand without constraint into a vacuum.
V
V
2V
Vacuum
Gas
Gas
The free expansion is represented in the above diagrams where initially a gas is contained
in one half of a container, with adiabatic and rigid external walls, by a separating internal
wall. These constraints (adiabatic and rigid respectively) mean that;
W = 0
and that
Q = 0.
If the internal wall is suddenly removed/broken the gas is free to expand into the other half
of the container.
Applying the first law to this process is straight forward and
U  Q  W  0
This means that for an ideal gas, where U ~ T , there is no change in temperature and
for an isothermal process as V doubles the pressure halves. It is an irreversible process
clearly as it can’t run backwards. This is approximately what was found for air by Joule but
in fact there is a very slight cooling. The molecules on the right hand side are on average
further apart and therefore their potential attraction drops meaning that their potential
energy goes up (recall that the attractive potential is a negative quantity). If the potential
increases then the kinetic must decrease in order that the total, U = 0. This is the origin of
the slight cooling. It is an effect that can be expected for real gases where U = U(T, V) but
not for an ideal gas where U = U(T). However that the effect is small shows that the ideal
approximation is a good one.
79
Thermal & Kinetic Physics: Lecture Notes
ii)
©
Kevin Donovan
Joule Kelvin Effect.
The Joule Kelvin effect is a way to effect an expansion accompanied by cooling that is
more effective than the free expansion and used in the liquefaction of gases, also known
as the throttling effect.
Pi +dP
Pi , Ti , Vi
Pf
Porous Plug
Pf , Tf , Vf
Pf - dP
Pi
Beginning with the situation depicted in the top figure where a porous plug divides a
chamber with adiabatic walls into two with the porous plug allowing high pressure on the
left hand side and lower pressure on the right hand side (a nozzle or constriction between
the two may also be used). There is a piston either side, on the left hand side it is
withdrawn and there is gas in the left hand side at initial pressure and temperature, Pi and
Ti The pressure is slightly greater on the outside of the piston forcing it slowly in and
pushing the gas through the porous plug into the right hand chamber where the pressure,
Pf , is lower. The right hand piston slowly withdraws (the pressure is slightly lower on the
outside of this piston to allow this) to accommodate the arriving gas at the final pressure
and temperature, Pf and Tf . Because Pf is lower than Pi the volume Vf must be greater
than Vi , an expansion has occurred. This is an isobaric process where the pressure
remains unchanged on either side of the plug and the work done on the gas in forcing it
through the plug is;
80
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
0

WLHS   Pi dV  PiVi
Vi
The work done by the gas expanding into the right hand side is
Vf

WRHS   Pf dV  Pf Vf
0
We can apply the first law recalling that the walls are adiabatic and therefore Q = 0
U  Uf  U i  Q  W  PiVi  Pf Vf
Or re-expressed as
U f  Pf Vf  H i  U i  Pi Vi  H f
and therefore this is an isenthalpic process with H  0 . The pressure and volume
have changed but the enthalpy has remained unchanged.
In general there will be a temperature change in the process and either heating or cooling
can occur.
Pf
 T 
dT  
 dP   JT dP
 P  H
T 

Pi
 T 

 dP 
 P  H
Pf

JT dP
Pi
or
 T
dT  
 V
Vf

 dV   J dV
H
T 

Vi
 T

 V

 dV 
H
Vf
  dV
J
Vi
Where  JT is called the Joule Kelvin or Joule Thomson coefficient (Lord Kelvin and
Thomson are of course the same person!) and  J is the Joule coefficient. An everyday
example of this effect is the cooling of a propellant gas as it is released from a high
pressure container into a room at ambient (low) pressure as with a deodorant spray.
To proceed further the enthalpy of real gases must used for the Van der Waals gas as
found earlier;
HVdW 
5
n 2a
n 3 ab
nRT  2
 nbP 
2
V
V2
81
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
 T 
The partial differential 
 is required to get the Joule Thomson coefficient  JT ’
 P  H
Since T and P are the state variables of interest the enthalpy is required in terms of these
two variables. The above expression is in terms of T, V and P and it is necessary to
substitute for V and to simplify the expression as follows;
1. The van der Waals correction terms involving a or b are small first order corrections
and therefore the final term on the RHS involving the product ab is of second order in
smallness and may be dropped,
HVdW 
2. It is necessary to replace
5
n 2a
nRT  2
 nbP
2
V
a
in favour of P and T removing V. We do this by
V
approximating to an ideal gas for the third term on the RHS;
a aP
aP


V PV nRT
Thus
HVdW 
5
2naP
nRT 
 nbP
2
RT
With reference to the throttling process it has been demonstrated that while P changes the
 H 
enthalpy remains constant and therefore 
  0 . It is then possible to obtain the Joule
 P 
 T 
Thomson coefficient 
 by differentiating both sides of the equation wrt P.
 P  H
5
2na 2naP  T 
 H 
 T 


  0  nR 
  nb 


2
RT
 P  H
 P  H
RT 2  P  H
 T 
Now solving for 
   JT
 P  H
2a
b
 T 
RT
 JT  


 P  H 5 R  2aP
2
RT 2
82
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
All of the terms in the denominator on the RHS have positive values and it is therefore
positive, however the numerator may take either sign dependant on the temperature and
so then may  JT .
For T 
2a
coefficient is positive, ie. for low enough T the coefficient  JT > 0 and a drop
Rb
in pressure produces a drop in temperature whereas if T 
2a
a drop in pressure will
Rb
cause an increase in temperature.
The different behaviors arise because of the two effects of molecular interactions in a real
gas. This can be seen by considering the molecular interaction via eg. the Lennard-Jones
potential.
12
6
 r

 rmin  
min
U LJ r    
  2
 
r  
 r 




U(r)
rmin
r
The proximity of neighboring molecules as r (or V) becomes smaller from the right) causes
a lowering of potential energy and as their proximity increases still further collisions
(“touching” of molecules) causes a rise in potential energy.
i)
At low temperatures where there are low collision rates between gas molecules the
effect of increasing the intermolecular distance (as described for the Joule free
expansion) increases the molecule-molecule potential interaction as represented by
the rising part of the potential interaction curve between the minimum and infinity thus
increasing the potential energy part of the internal energy. For the internal energy to
83
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
remain constant the kinetic part of the internal energy must decrease and this is the
origin of the decrease in temperature.
ii)
Collisions between molecules, as represented by the rising part of the potential
interaction curve between the minimum and the origin results in a temporary
conversion of kinetic to potential energy. At high temperatures where the molecular
velocity and therefore collision rate is high this may be a large factor and an increase
in volume resulting in a reduction in the collision rate will consequently result in a
reduction of mean potential energy and therefore an increase in kinetic energy with
resultant rise in temperature.
TInv 
2a
is known as the inversion temperature and divides the two regimes.
Rb
This process is used in a continuous flow process (rather than the single shot process
schematically described earlier) called the Linde liquefaction process.
For good measure the Joule coefficient may be found in similar fashion;
Starting with the van der Waals enthalpy;
5
n 2a
n 3 ab
HVdW  nRT  nRT  2
 nbP 
2
V
V2
 T
The partial differential 
 V

 is required to get the Joule coefficient  J ’
H
Since T and V are now the state variables of interest the enthalpy is required in terms of
these two variables. The above expression is in terms of T, V and P and it is necessary to
eliminate P and simplify as follows;
1. As previously the final term on the RHS involving the product ab is of second order in
smallness and may be dropped,
HVdW 
5
n 2a
nRT  2
 nbP
2
V
84
Thermal & Kinetic Physics: Lecture Notes
©
Kevin Donovan
2. It is necessary to replace P in favour of V and T. We do this again, by approximating to
an ideal gas for the third term on the RHS;
P
nRT
V
Thus
HVdW 
5
na 2 n 2 bRT
nRT  2

2
V
V
 T
It is now possible to obtain the Joule coefficient 
 V

 by differentiating both sides of the
H
equation wrt V.
5  T 
n 2a n 2 bR  T 
n 2 bRT
 H 

0

nR

2








2  V  H
V  V  H
 V  H
V2
V2
 T 
Now solving for 
  J
 V  H
n 2 bR  n 2 2a  bRT 
 T   5

  nR 

V 
 V  H  2
V2
n 2 2a  bRT 
 T 

  J 
 5 nb 
 V  H
V 2 nR  

2 V 
All of the terms in the denominator on the RHS have positive values and it is therefore
positive, however the numerator may take either sign dependant on the temperature and
so then may  J . For bRT  2a or T 
2a
the coefficient is positive and an increase in V
bR
(expansion) will result in an increase in T. The inversion temperature TInv 
2a
is as
Rb
previously found for  JT
The expansion of a gas causing a drop in temperature is frequently used in refrigeration
processes and requires a negative Joule coefficient and therefore operation of the system
above the inversion temperature of the particular gas used.
85