You are on page 1of 8

Lecture 6: The Ideal Gas Partition Function

6.1 Recommended textbook chapters for this section


• Chandler - Ch. 4 intro, 4.6, 4.7

6.2 Topics in this lecture


• Factorization of the partition function

• The ideal gas partition function

6.3 Announcements
• Pset 2 due Thursday

6.4 Factorization of the partition function


In the last lecture, we derived expressions for generalized partition functions and related connections
to macroscopic thermodynamic potentials. We specifically paid attention to four key ensembles:
the microcanonical ensemble (our starting point for all further derivations), the canonical ensem-
ble, the grand canonical ensemble, and the isothermal-isobaric ensemble. These are the four most
common ensembles used in statistical mechanics. To date, we have considered partition functions
describing all states for an entire system of interest. However, often times we will be able to divide
the system into multiple independent subsystems (indeed, we did this for the polymer adsorption
example), perhaps even independent particles, suggesting that we may want a relationship between
the properties of an entire system and the properties of a subsystem. Two subsystems are indepen-
dent if the microstate of one subsystem does not depend on the microstate of another - this requires
that the subsystems non-interacting so that positions and energies are uncorrelated. We will thus
introduce a method for relating a system partition function to subsystem partition functions.
Consider a system for which we can break the energy into two parts (i.e., into the energies of
two subsystems, for example, or two different sources of energy) such that we can write in general
E = (1) + (2) . The superscripts in this case indicate the energies of subsystems 1 and 2, and we
partition the total system energy between them; this is similar to the partitioning of energy between
the system of interest and the bulk reservoir that we used to derive the canonical partition function
(1) (2)
in Lecture 3. The energy of a single configuration, j, of the combined system is then Ej = n +m ,
where the label j identifies a particular configuration of the combined system and n and m denote
specific configurations of the two subsystems. We assume that n and m are independent of each

1
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

other since the subsystems themselves are independent. Rather than enumerating all possible
microstates by summing over all j, we can thus instead enumerate all possible microstates by
summing over all possible values of n and m independently. The corresponding canonical partition
function for the combined system is then:

X
Z= e−βEj (6.1)
j
h i
(1) (2)
−β n +m
XX
= e (6.2)
n m
XX (1) (2)
= e−βn e−βm (6.3)
n m

If the indices n and m are independent of each other, we can now further write:

! !
X (1) X (2)
−βn −βm
Z= e e (6.4)
n m
(1)
≡Z Z (2) (6.5)

Expressing the sum over all possible indices as a product of sums over single indices is the
key to this method - because E can be decomposed into the sum of two independent values,
the product of the two summations yields all possible values of E. You can demonstrate this
manually by assuming a two-state energy system per subsystem, for example, and counting the
possible products. Note that this factorization is only possible because the two subsystem energies,
(1) and (2) are independent and uncorrelated - if either energy were a function of the other (i.e.
if n and m were not independent), then the same factorization would not be possible. Since
each of the independent sums takes the same form as the canonical partition function, we use
Z (1) to refer to the partition function for subsystem 1, etc. Again, note that we write these as
sums over particular microstates of the system and microstates of the subsystem, so there could
be many possible microstates with the same decomposition of energies but with different particle
configurations.
Extending this result, if we break the system into N independent, uncorrelated subsystems then
the partition function can be written as:

Z = Z (1) Z (2) . . . Z (N ) (6.6)


In principle, each of the N partition functions for the subsystems could be distinct, because
each subsystem could represent a collection of particles giving rise to distinct possible energies.
If, however, we have a system of completely identical, independent subsystems (e.g. each
subsystem could represent a single particle), then the partition function for each subsystem is
identical and the partition function of the combined system becomes:

Z = (Z (1) )N (6.7)
N
=z (6.8)

To simplify this notation, we will define Z (1) ≡ z as the single-subsystem partition function. This
factorization dramatically simplifies our calculation of a system partition function - for example, if

2
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

we had a system with 100,000 independent molecules (e.g., ideal gas molecules), and each molecule
can exist in three different microstates, then the total number of states in the entire system is
3100,000 , which is much too large to calculate. Eq. 6.8 instead says that we only need an expression
for the three-state partition function z which is trivial. Again, we assume that each subsystem
is independent and identical - so this implies that each subsystem has exactly the same partition
function associated with it (same possible energies, same degeneracy for each energy, etc.).
In the preceding definition we have applied a distinct label to each of the N different single-
subsystem partition functions, implying that each molecule is distinguishable. As we discussed in
the derivation of the microcanonical ensemble, however, in most systems of independent particles
the particles are instead indistinguishable. Identifying distinguishable vs. indistinguishable
subsystems can be confusing, but the idea here is that particles/subsystems are distinguishable if
we can uniquely identify the subsystem if it were randomly selected. For example, each monomer in
a polymer is distinguishable because we can identify exactly where in the polymer chain it is (i.e. its
if the first, second, third, etc. monomer), whereas each molecule in an ideal gas is indistinguishable.
Similarly, each site on a lattice is distinguishable, but if particles are able to move between lattice
points then the particles are indistinguishable.
If each subsystem of a combined system is indistinguishable, then for a given value of E we can
distribute values of (1) , (2) , (3) , . . . to each subsystem in many ways that all yield the same value
of E, but these different distributions all lead to sets of microstates that are indistinguishable, and
thus each set of configurations should only be counted once in the system partition function. For
example, assume a two particle system where each particle can have a discrete energy arbitrarily
labeled as 1, 2, 3, or 4. Treating each particle as a subsystem and assigning an energy of 0 to
particle 1 and 4 to particle 2 is equivalent to assigning an energy of 0 to particle 2 and 4 to particle
1; hence this state should only be counted once in the partition function. Note that there can still
be many different ways of assigning energies to achieve the same total system energy of 4 - we could
have assigned the values of 0/4, 1/3, or 2/2, and these would all yield unique states. We just do
not want to overcount the redundant states 4/0 and 3/1 in this example.

To avoid overcounting states associated with indistinguishable subsystems, we recognize from


the discussion of the microcanonical ensemble that there are N ! possible ways to assign the energies
i , j , etc. to the N subsystems. That is, we choose one of the N subsystems first and assign it
a value i , then we choose one of the N − 1 subsystems second and assign it a value j , etc., such

3
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

that there N ! different ways of assigning a given set of i , j , . . . to N subsystems, and each of these
would have the same energy, so for this particular value of E we have counted N ! states when we
should have counted only 1. If we imagine counting our partition function by summing over energy
levels, then the degeneracy of each energy level is overcounted by this amount. More
specifically, each possible way of assigning these states is the same for indistinguishable subsystems,
so we overcount the degeneracy of each energy level in the combined system partition function by
a factor of N !.
Dividing by this factor of N ! gives a final expression for the partition function of a system of
N independent, indistinguishable subsystems as:

zN
Z= (6.9)
N!
X
z= e−βi (6.10)
i

Finally, we can note that  itself may be divided into a series of individual components - for
example, we can imagine a single particle as having a component of its energy related to its kinetic
energy or its potential energy. Writing  = a + b + c . . . allows us to immediately write for the
system partition function:

(za zb zc . . . )N
Z= (6.11)
N!
where za are single-subsystem partition functions containing only the energetic term of the sub-
division. The only important point to note here is that typically the different energetic components
of the single-subsystem energy are distinguishable (i.e. kinetic energy and potential energy) and
thus the factorial in the denominator does not change.
To summarize, we thus obtain:

X X
Z= e−βEj = Ω(N, V, Eν )e−βEν Single system with N particles (6.12)
j ν
X
Z = zN → z = e−βj N independent, distinguishable subsystems (6.13)
j

zN X
Z= →z= e−βj N independent, indistinguishable subsystems (6.14)
N!
j

(za zb zc . . . )N
Z= Eq. 6.14 with multiple energetic components per subsystem (6.15)
N!

6.5 The ideal gas partition function


We will now use these previous relations to now derive the partition function for a monatomic
ideal gas with N particles as a exemplary case of a many-body non-interacting system and show
that we can recover results expected from classical thermodynamics. The basic steps for identifying
thermodynamic properties of the system will be the following:

• Determine a single-particle energy and corresponding single-particle partition function

4
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

• Write a partition function for the entire system using the single-particle partition function

• Derive thermodynamic quantities via derivatives of the partition function

We will start by performing this derivation in the canonical ensemble. We assume a volume of
monatomic gas contained with a P cubic box of length L, such that V = L3 . Each particle has an
associated partition function z = j e−βj . Because the system is ideal and each molecule is only a
single atom (and hence there are no degrees of freedom associated with molecular vibrations, etc),
the energy of each particle is given from the solution to the quantum “particle in a box”, which is a
generalization of a result stated in Lecture 1 (note that we will not derive where this energy comes
from as it is outside the scope of this class). A quantum particle in a box has a discrete spectrum
of energies corresponding to the possible positions (or, more accurately, probability distributions
associated with specific positions, which are equivalent to the square of the wavefunction). The
energy is given as:

h2 (n2x + n2y + n2z )


ν = , (6.16)
8mL2
nx , ny , nz = 1, 2, 3 . . . (6.17)

Here, h is the Planck constant, m is the mass of the particle, and nx , ny , and nz are three
unitless quantum numbers that specify a particular quantum state in a 3D box. Recall that in
quantum mechanics energies are discretized, such that the three quantum numbers above specify
discrete values of the energy, ν , and also contribute to the degeneracy, ω(ν ) (written in lower
case to indicate a single-particle degeneracy). Since particle positions are associated with different
energies, the degeneracy of each energy level is specified only by the number of possible ways of
assigning quantum numbers to yield that energy.
To proceed, we would like to calculate the single particle partition function. We write the
partition function by summing over all possble values of the energy, ν :

X
z= w(ν )e−βν (6.18)
ν

The prefactor w(ν ) is the degeneracy of a given energy level, ν . We can start by computing
an expression for the degeneracy by picturing a three-dimensional space of quantum numbers in
a coordinate system defined by nx , ny , and nz , where each point in this space that has positive
values of all three numbers is a single possible molecular quantum state (i.e., one quantum state
per unit volume) corresponding to some energy value ν . The equation

8mL2 ν
n2x + n2y + n2z = = R2 (6.19)
h2
then defines the equation for a sphere of radius R2 in the nx ny nz space, where the expression for
R is from the rearrangement of Eq. 6.16.

5
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

We can quantify the number of possible microstates as the volume of the sphere for which all
3
three quantum numbers are positive, or πR
6 . We can write this phase-space volume as:

πR3
Θ(ν ) = (6.20)
6
π 8mν 3/2
 
= V (6.21)
6 h2

Note that V = L3 is the volume of the physical system, while R3 defines the volume of the
phase space. What we really want is an expression for the degeneracy of a particular energy level ν
- this can then be approximated as the number of states with an energy between ν and ν + ∆ν :

ω(ν ) = Θ(ν + ∆ν ) − Θ(ν ) (6.22)


∆Θν
= ∆ν (6.23)
∆ν
We now assume the difference in energy ∆ν is extremely small, such that ∆ν → dν . This
approximation then gives us the volume of an extremely thin shell of phase space, and therefore
is a reasonable approximation for the number of states on the surface of our sphere. We can thus
approximate the degeneracy as:

dΘν
ω(ν ) ≈ dν (6.24)
dν
!
π 8mν 3/2
 
d
= V dν (6.25)
dν 6 h2
π 8m 3/2 1/2
 
= V ν dν (6.26)
4 h2

Substituting this expression into our partition function gives:


π 8m 3/2 1/2 −βν
X  
z= V ν e dν (6.27)
ν
4 h2

π 8m 3/2 X 1/2 −βν
 
= V ν e dν (6.28)
4 h2 ν

6
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

6.6 Ideal gas partition function in the classical limit


The previous expression has now assumed that the quantum numbers are effectively continuous
in space. To simplify this expression further, we further assume that the sum over all energies
can be converted to an integral because the differences in the energy between states, ∆, is much
smaller than kB T , such that the energy states are also effectively continuous (which is consistent
with the assumption that quantum numbers are continuous). Converting a sum to an integral is
equivalent to what we did when discussing Stirling’s approximation - we can think of the sum over
energies as performing rectangular integration of the energy vs. ν, and as the difference between
consecutive terms in the sum becomes small, rectangular integration can be well-approximated using
a continuous integral. The assumption that energies are continuous is reasonable in a macroscopic
system where incrementing the quantum numbers by one leads to a very small change in the energy
(≈ 10−9 kB T ). This assumption is referred to as the classical limit, and indeed, writing a partition
function in terms of an integral over phase space is a hallmark of classical statistical mechanics
(to date, everything we have done is in the quantum limit by writing energy levels as discrete). In
the classical limit the partition function becomes:

π 8m 3/2
  Z ∞
z= V ν1/2 e−βν dν (6.29)
4 h2 0
We
R ∞can1/2analytically solve this expression by defining u = βν and recognizing the standard inte-
−u √
gral 0 u e du = π/2. Simplifying yields our final expression for the single-particle partition
function for the ideal gas:

 3/2 Z ∞
π 8m
z= V ν1/2 e−βν dν (6.30)
4 h2 0
 3/2 Z ∞
π 8m
= V (kB T )3/2 u1/2 e−u du (6.31)
4 h2 0
!
π 8mkB T 3/2 π 1/2
 
= V (6.32)
4 h2 2
2πmkB T 3/2
 
= V (6.33)
h2
Finally, we write the full partition function of the system, remembering that the particles are
indistinguishable, as:

zN
Z= (6.34)
N!
1 2πmkB T 3N/2 N
 
= V (6.35)
N! h2
 N
1 V
= (6.36)
N ! λ3
In the last line, we define:
s
h2
λ= (6.37)
2πmkB T

7
University of Wisconsin-Madison Lecture 6
CBE 710, Fall 2019 - Prof. R. C. Van Lehn September 24, 2019

where λ is the thermal de Broglie wavelength which has units of length and is a function of T ;
this simplifies our notation and defines a characteristic length scale for treating a gas classically - if
V  λ3 , which we will generally assume to be the case, then the classical limit is reasonable. With
this partition function, we can proceed to derive thermodynamic relationships as will be discussed
in the next lecture.

You might also like