Click here to view G.U. B.Sc. 3rd yr. examination questions on statistical thermodynamics & data analysis

A Brief Introduction to Statistical Thermodynamics
A Web-Book by Rituraj Kalita, Dept. of Chemistry, Cotton College, Guwahati-781001 (Assam, India)
Preface   Ch. 1   Ch. 2   Ch. 3   Ch. 4   Ch. 5   Ch. 6   Ch. 7   Bibliog.
     Background topics/ vocabulary           General topics           Advanced (avoidable) topics
© 2006. Copyright reserved. The book or any portion of it can't be reproduced/ re-published/ circulated.

Ch. 3: Molecular Partition Function for Gases
and their Thermodynamic Functions

 

3.1 Towards the Evaluation of the Molecular Partition Function q for Gases:

The molecular energy e discussed earlier is actually the complete (overall) molecular energy that is expressible (approximately) as a sum of several specific-mode molecular energy components such as translational, electronic etc. energies. Because of this fact, the (complete or overall) molecular p. f. (denoted as q) may be expressed as a product (we'll show that soon) of several such specific-mode p. f. factors (qtr, qel etc.). The basic idea is that a molecule, say an O2 molecule, at the same time, undergoes several types of molecular motions such as translational, electronic, intra-nuclear, rotational and vibrational, with at least some approximate degree of independence among these motions (talking in irritating quantum-mechanical jargon, we say that the molecular Hamiltonian h may be separated into specific-mode Hamiltonians, at least as a reasonable approximation), and so such specific-mode (translational, electronic etc.) energies and corresponding specific-mode energy-states (e.g., a translational molecular state, an electronic molecular state) can be thought of. When you've recently discussed the electronic motion and the associated electronic energy of H-atom in your quantum mechanics course (and derived that electronic energy to be of the form –Z2E0/n2), you were made to overlook the fact that the H-atom may also have other (i.e., non-electronic) form of energy, particularly the translational energy (the translational energy is associated with translational motion, i.e., with the movement of the atom/molecule as a whole from one place to another). At this point, I would also like to recall an interesting incident of 'ragging' I faced as a fresher student then, during the mid-eighties, at the college where I now teach: I was asked to move my head, hand, feet, fingers and toes. I could do each of that, naturally, but then I was asked to move all these body-parts together! This came out to be totally impossible for me; but, strangely, the gas molecules always keep undergoing all the specific-mode motions simultaneously. An O2 molecule moving from one place to another (i.e., undergoing translation) is simultaneously also undergoing haphazard electron-motions, a 'to and fro' vibration in the internuclear bond and a revolving rotation of the nuclear framework!

For the non-monatomic molecules (i.e., those having two or more nuclei within it, e.g., O2, NO, NH3) the molecular energy is approximated (via Born-Oppenheimer Approximation and Rigid-Rotor-Harmonic-Oscillator Approximation etc.) to be a sum (as e = etr + eel + enuc + evib + erot) of a translational energy etr, an electronic energy eel, an (intra-)nuclear energy enuc, a vibrational energy evib and a rotational energy erot, each of these specific-mode energies being associated with a corresponding specific-mode energy state (so, the overall molecular state f is obtained as a product of specific-mode states as: f = ftrfelfnucfvibfrot). Thus, a complete (overall) molecular state for such a molecule is a combination of a translational state, an electronic state, an (intra-)nuclear state, a vibrational state and a rotational state. For system of such gas molecules, q (expressed as a sum over states) is given by q = Ss exp{-es /(kT)}, and so we have:
q = Ss exp{-e /(kT)}= Ss exp{-(
etr+eel+enuc+evib+erot)/(kT)}
= Ss exp{-
etr/(kT)}.exp{-eel/(kT)}.exp{-enuc/(kT)}.exp{-evib/(kT)}.exp{-erot/(kT)}
As an overall state is composed of a combination of the specific-mode states, so the sum over all the overall molecular states is same as the sum over all translational states, all electronic states, all (intra-)nuclear states, all vibrational states and all rotational states. So:
q = Str-states Sel-states Snuc-states Svib-states Srot-states exp{-
etr/(kT)}.exp{-eel/(kT)}. 
 exp{-
enuc/(kT)}. exp{-evib/(kT)}. exp{-erot/(kT)}
As the specific-mode states are independent of one another, so the above five summations could be performed independently of one another. This means
q = [Str-states exp{-
etr/(kT)}]. [Sel-states exp{-eel/(kT)}]. [Snuc-states exp{-enuc/(kT)}].
 [Svib-states exp{-
evib/(kT)}]. [Srot-states exp{-erot/(kT)}]
The sums within the square-brackets in this expression are called respectively as translational p. f. qtr, electronic p. f. qel, (intra-)nuclear p. f. qnuc, vibrational p. f. qvib, rotational p.f. qrot.
So we have
q = qtr qel qnuc qvib qrot
Notes: (i) A complete (overall) molecular energy-level for such a molecule is similarly a combination of a translational level, an electronic level, an (inra-)nuclear level, a vibrational level and a rotational level, with the degeneracy g of the overall level equaling the product of the degeneracies of the specific-mode energy levels, i.e., g = gtr gel gnuc gvib grot. The specific-mode degeneracies get multiplied thus, because every translational state (within the overall level) may coexist with every electronic state (therein), and so on. (ii) The intra-nuclear (molecular) state (clearly different from the roto-vibrational inter-nuclear state) is popularly also called just as the nuclear state, so we're now calling it '(intra-)nuclear state'. Soon, we'll call it just as nuclear state, and the corresponding energy as nuclear energy!

For the monatomic molecules (i.e., those having only one nuclei within it, e.g., He, Ne, Ar, H) the molecular (i.e., atomic) energy is approximated to be a sum (as e = etr + eel + enuc) of a translational energy etr, an electronic energy eel and an (intra-)nuclear energy enuc, and so for them we get q = qtr qel qnuc   (for them, rotational and vibrational motions do not exist!).

While evaluating the partition functions, it is also important to note the convention in which the zero-values of the molecular energies are expressed. Just as the gravitational potential energy (expressed as Vgrav = mgh) of a piece of chalk in the teacher's hand may be measured with a zero-value reference (i.e., h = 0, Vgrav = 0) to the classroom ground, to the nearby street, or to the world-wide sea-level reference, the energy of a molecule also may be expressed with several different zero-value references. However, the widely popular convention is that any specific-mode (i.e., translational etc.) energy of a molecule is considered to be zero at the ground energy level (of that mode). In contrast to this popular convention (which is analogous to the classroom-ground reference for the piece of chalk), another convention particularly useful for treating chemical reactions and equilibriums (i.e., molecule-interconversions) considers instead the (overall) ground energy level for all neutral gaseous atoms to be at zero energy (see Ch. 4,  this convention is obviously analogous to the world-wide sea-level reference one). In a few coming sections, we'll try to find simple expressions for all the specific-mode p. f. 's, using the above popular convention.
 

3.2 Derivation of the Expression for the Translational Partition Function (qtr):

From a quantum mechanics outlook, the translational energy of a gas molecule enclosed within a system is nothing but the energy of a particle in a 3-dimensional box model (with the mass of the particle equaling the mass m of the molecule). [Why? This is because the energy of the particle in the 3-D box is clearly by virtue of only its translational motion (i.e., motion from one place to another).] Similarly, every energy-state of the particle in a 3-D box problem implies a translational energy state of the gas molecule. For simplicity of derivation, we choose a cubical 3-D box (of equal sides, say L).
Note: Mass m of a gas molecule equals M/NA, where M is the molar mass. Thus, the mass of an ordinary (16O2) oxygen molecule (with M = 0.032 kg mol–1) is 5.314 x 10–26 kg.

From quantum mechanics, we know that the translational energy etr of the molecule in such a cubical box is etr = (nx2 + ny2 + nz2) h2 / (8 m L2), where (nx, ny, nz) are all positive integers ranging from 1, 2, 3 onwards. Furthermore, a specific set of (nx, ny, nz) [say the set (nx = 235, ny  = 641, nz = 3726)] implies a particular translational state.
So we have, qtr = Σtr-st exp [–etr / (k T)] = Σtr-st exp [– (nx2 + ny2 + nz2) h2 / (8 m L2 k T)]  
where tr-st implies translational states, and the above sum is made over all translational states.
Note: The above defining expression for translational energy, etr = (nx2 + ny2 + nz2) h2 / (8 m L2) doesn't exactly obey the aforesaid popular convention, as per which the ground level (with nx = ny = nz = 1) should has energy zero. However, as h2 / (8 m L2) is extremely small, this inconsistency is negligible.

As every possible (positive integral) set of (nx, ny, nz) implies one translational state, and every translational state is indicated by one possible set of (nx, ny, nz), so the above sum over all translational states is identical with the similar sum over all the positive integral values of the three numbers (nx, ny, nz). This means (in expressions below, nx,ny,nz varies from 1 to +a):
qtr = Σnx,ny,nz exp [– (nx2 + ny2 + nz2) h2 / (8mL2 kT)] 
= Σnx,ny,nz  [expnx2h2/(8mL2 kT)}]. [expny2h2/(8mL2 kT)}]. [expnz2h2/(8mL2 kT)]

As each of the above three exponential functions within the square brackets involve only one out of the three variables (nx, ny, nz), so each of the summations over each of the variables (nx, ny, nz) can also be performed independently. This gives qtr as a product of three sums as:
qtr = [nx=1Σ+a exp{– nx2h2 / (8mL2kT)}]. [ny=1Σ+a exp{– ny2h2 / (8mL2 kT)}].
                 [nz=1Σ+a exp{– nz2h2 / (8mL2 kT)}]

Expanding the three sums within the square brackets above, it becomes obvious that all the three sums are actually identical, differing only in the name (nx, ny, nz) of the dummy variable. So we now have a simple expression: qtr = [nx=1Σ+a exp{– nx2h2 / (8mL2kT)}]3
To evaluate the sum nx=1Σ+a exp{– nx2h2 / (8mL2kT)}, we note that for any macroscopic system (with L at least 0.5 mm, say) the quantity a = h2/(8mL2kT) is extremely small, and so it can be shown that the difference between two consecutive terms (
say, the two terms with nx = 342 and nx = 343) in the series is much smaller than the two terms themselves. This is the condition for approximating a sum by an integral, and so the sum nx=1Σ+a exp{– nx2 a} can be replaced by the integral I = nx=1+a exp{– nx2 a} dnx. Further, as this integral value is quite large, so it makes little error if the lower limit is replaced by nx = 0 instead of nx = 1. So we have:  I =  nx=1+a exp(– nx2 a) dnx = nx=0+a exp(– nx2 a) dnx = 0+a exp(– a s2) ds. 
Note: The sum [nx=1Σ+a exp{– nx2 a}] = [nx=1Σ+a exp{– nx2 a} Δnx], where Δnx = 1. So wasn't there a Δnx hidden in the sum (which got converted to dnx)? (If not interesting, don't bother about this math-insight!)

We know that value of the standard integral 0+a exp(–as2)ds = (1/2) (p/a)1/2. So we have:
qtr = [0+a exp(– a s2) ds]3 = [(1/2)(p/a)1/2]3 = (1/8)(p/a)3/2  = (1/8) [(8pmL2kT)/h2]3/2
 
= 81/2[(pmkT)3/2 / h3] L3 = 23/2[(pmkT)3/2 / h3] L3 = [(2pmkT)3/2 / h3] L3
As L3 is nothing but the volume V of the system, so we have qtr = [(2pmkT)3/2 / h3] V
Though derived for the system within a cubical-box shaped enclosure only, this is the generally valid expression for qtr. So we have:  qtr = (2pmkT)3/2 V / h3 
Notes: (i) Making a substitution v = a s2, we could get I = (1/2) a–1/2 v=0+a v1/2–1 exp(–v) dv   (please verify!). The integral 0+a vn–1 exp(–v) dv   (where n is a positive real number) is known in mathematics as the gamma function, G (n). We know that G (2) = G (1) = 1 and G (1/2) = p1/2. [For n > 1, this function obeys the recursion relation G (n) = (n–1) G (n–1) which gives, for n = 2, 3, 4 etc. natural-number values of n, G (n) = (n–1)!]. So we have,   I = (1/2) a–1/2 w=0+a w1/2–1 exp(–w) dw = (1/2) a–1/2 p1/2  = (1/2) (p/a)1/2.
(ii) Take care to express V in m3 (SI) or cm3 (CGS) unit (not dm3), and m, k & h also in the SI or CGS unit. (iii) In a problem, the volume V might have to be found from the P-T data, using the ideal gas equation. Thus, for 2 moles of He gas (obviously here m = {(4.002/6.022) x 10–26} kg at P = 1 atm and T = 298 K, V = (2 mol) x (0.08206 atm L K–1 mol–1) x (298 K) / (1 atm) = 48.91 L = 4.891 x 10–2 m3. So, qtr = (2 x 3.142 x 6.646 x 10–27 kg x 1.381 x 10–23 J K–1 x 298 K)3/2. (4.891 x 10–2 m3) / (6.626 x 10–34 J s)3  = 3.788 x 1029. To perform this sort of complicated calculations, some software-package calculators such as the free package Assam-Calcu are the most suitable tools. (iv) The expression for qtr may also be written as qtr = V/Λwhere Λ = h / (2pmkT)1/2 is called the thermal de Broglie wavelength for the gas system (Λ obviously has the dimension of length: in the example in (iii), its value is 5.054 x 10–11 m = 0.5054 A0). [Please check!]
 

3.3 Derivation of the Expression for the Electronic Partition Function (qel):

The electronic energy levels eel and electronic degeneracies gel for any molecule may be obtained (from quantum mechanics) via solution of its molecular electronic Schrφdinger equation. Having obtained these energy levels and degeneracies, the electronic partition function qel is expressed (and calculated) simply from the following, basic, 'sum over levels' form:
  qel = gel,0 exp{– eel,0 /(kT)} + gel,1 exp{– eel,1 /(kT)} + gel,2 exp{– eel,2 /(kT)} + .........
(where eel,0 , eel,1 , eel,2 are respectively the ground, 1st-excited, 2nd-excited energy-level energies and gel,0 , gel,1 , gel,2 are respectively their corresponding degeneracies.)

Using the above-mentioned popular convention, the eel are expressed with the ground-level electronic energy as zero (i.e., eel,0 = 0), so the actual expression looks as follows:
  qel = gel,0 + gel,1 exp{– eel,1 /(kT)} + gel,2 exp{– eel,2 /(kT)} + .........

In most cases, there arises a great simplification to this expression. For most of the stable gas molecules (e.g., Ne, Ar, H2, N2, O2, HCl, HI, CO etc.) the excited electronic energy levels lie quite high (i.e., compared to the ground one), so that at usual, not-too-high temperatures (say, 300 K or even 1000 K), the exp{–eel /(kT)} values for any of the excited levels is practically zero (as eel >> kT), and so the excited-level contribution-terms to qel are negligible. So in such cases qel usually equals almost gel,0 (the degeneracy of the ground electronic level):     qel = gel,0
(An exception to this generalization is NO, for which eel,1 = 0.015 eV only: however, here also eel,2 , eel,3 etc. lie quite high.) Furthermore, for many of those stable gases (e.g., Ne, Ar, H2, N2, HCl, HI etc.) the ground electronic level (generally a singlet 1S+ one) is non-degenerate (i.e., gel,0 = 1) so that qel equals simply 1 (a notable exception is O2, for which gel,0 = 3 – this level is a triplet 3S-g one). The excited levels lie similarly high even for the atomic alkali metal gases (e.g. Na, Cs etc. atomic gases) and for atomic hydrogen (i.e., atomic H), though for them gel,0 = 2 (a doublet term): so for them even at 1000-2000 K temperatures, qel equals practically 2. Similarly, also for electron gas [e– (g), available within ionized gases] qel = 2.
Note: For NO at 300 K, eel,0 = 0, eel,1 = 0.015 eV and gel,0 = gel,1 = 2 while the 2nd excited and still upper levels hardly contribute (as for them eel >> kT), so we get:
qel = 2 + 2 exp{–(0.015 x 1.609 x 10–19 J) /(1.381 x 10–23 J K–1 x 300 K)} = 3.117
 

3.4 Derivation of the Expression for the Nuclear Partition Function (qnuc):

The (intra-)nuclear energy levels for a molecule are always such that the upper (i.e., the excited) nuclear levels lie extremely above the ground level, so that at any temperature of chemical interest the excited nuclear levels are hardly occupied, and thus the contributions of the excited nuclear levels to the nuclear p. f. is negligible. Considering the universally accepted convention of the ground level nuclear energy being taken as zero, we so get qnuc = gnuc,0 (where gnuc,0 is the degeneracy of the ground nuclear level).
Note: For an atom (i.e., monatomic molecule) with nuclear-spin number I (I = ½ for H, 1 for 14N), gnuc,0 = 2I+1. For a diatomic molecule with atoms a & b, gnuc,0 = (2Ia + 1)(2Ib + 1).

However we have an even more simplifying, but nevertheless useful*, expression for qnuc. We note that the ground nuclear degeneracy gnuc,0 of an atom never actually changes in any physical or chemical process (but simply combining into their product during formation of a molecule). So in any calculation about change of a thermodynamic property for any physical or chemical process (say that of entropy change for a reaction, ΔSreac) it makes no difference to the calculation-result if nuclear p. f. of all the atoms and molecules are taken simply as 1. This idea gives rise to the very popular, simplifying convention among chemists, according to which qnuc = 1 (for any atom/ molecule/ ion). 
* Notes:
(1) This simplifying convention is obviously not useful for calculating the absolute value of entropy (the so-called third-law entropy) of a system (in contrast to calculating an entropy-change). (2) With this simplifying convention, the relation q = qtr qel qnuc qvib qrot (for non-monatomic molecules) obviously becomes simply q = qtr qel qvib qrot.
 

3.5 Derivation of the Expression for the Vibrational Partition Function (qvib):

At this point it should be repeated that qvib and qrot exists only for non-monatomic molecules. Let us now evaluate qvib for the simplest case of diatomic molecules, using the rigid-rotor harmonic-oscillator model. According to this simple model (please note that this model is rather an approximate one), a diatomic molecule (e.g., CO) has a characteristic frequency of vibration ν, equaling* (2p)–1(K/m)½. From quantum mechanics of harmonic oscillators, we further know that the vibrational energy levels evib are given by (v + ½)hν (where v = 0, 1, 2, ... etc.), each of these levels being non-degenerate (i.e., gvib = 1). However, considering the popular convention of the ground (v = 0) vibrational level having the zero energy, the vibrational energy levels are to be re-assigned as:
evib = vhν    (with v = 0, 1, 2, 3, ...... etc.)
* Note: To clarify further, let us note that here K is the (oscillator) force constant [equaling the second derivative (d2U/dR2)Re of molecular electronic energy U w.r.t. internuclear distance R at R = Re, the equilibrium internuclear distance] and m is the reduced mass equaling (mamb)/(ma+mb), where ma & mb are the masses of the constituent atoms a & b respectively. Each of ν, K and m is a molecular constant for a given (diatomic) molecule (i.e., say, for the CO molecule). Frequency ν is sometimes also expressed in terms of wavenumber ύ (ύ = 1/λ, λ being the wavelength, ύ being in cm–1 or m–1 unit).

So the vibrational partition function is qvib = v=0Σa gvib,v exp {– vhν /( kT )}
  = v=0Σa [exp {– hν /( kT )}]v 
       [as degeneracy gvib,v = 1]
  = 1 + y + y2 + y3 + ......           [where we take y = exp {– hν /( kT )}]
  = 1/(1 – y)           [using the well-known resulting expression for this series for |y| < 1]
  = 1/ [1 – exp{– hν /( kT )}]           [putting the value of y]
Here ( hν / k ) is another characteristic molecular constant with the dimension of temperature. This quantity is called the characteristic vibrational temperature, denoted as θvib

Thus we get the following simple expression for vibrational p. f. of diatomic molecules:
   qvib         =  1/ [1 – exp{– hν /( kT )}]         =  1/ [1 – exp{– θvib / T}]

For example, let us consider H2 gas at 2000 K temperature. If it is given that here the force constant K is 554.4 N m–1 [we may easily get the reduced mass m for H2 as:
 
m = {(1.6678 x 10–27 x 1.6678 x 10–27)/(1.6678 x 10–27 + 1.6678 x 10–27)} kg
= 8.369 x 10–28 kg, please check it yourself], then ν = (2p)–1(K/m)½ = 1.295 x 1014 s–1 and  θvib = hν / k = 6215 K.
Thus we have qvib  =  1/ [1 – exp{– 6215 K / 2000 K}] = 1.0468.

For polyatomic molecules (atomicity a > 2), there exists several characteristic frequencies of vibration νi and several corresponding characteristic vibrational temperatures θvib,i corresponding to the several normal modes of vibration i [the number of these normal modes equal (3a–5) for linear molecules and (3a–6) for non-linear molecules, where the atomicity a is the number of atoms in the molecule]. The vibrational partition function here is a product of the factors from each of the normal modes. So for polyatomic molecules we get:
  qvib         =  1/ Πi [1 – exp{– hνi /( kT )}]         =  1/ Πi [1 – exp{– θvib,i / T}]
Note: We see that for most molecules, qvib values are at most marginally more than 1 (in other cases practically equal to 1), and so is generally immaterial in contributing to q and to thermodynamic functions.
 

3.6 Derivation of the Expression for the Rotational Partition Function (qrot):

Let us start with evaluation of qrot for the simplest case of heteronuclear diatomic molecules (e.g., HCl, CO, HI, and even HD or O16O18 – but not simple H2 which is homonuclear). Considering the aforesaid rigid-rotor harmonic-oscillator model, we get that the rotational energy levels are:
    erot = J(J+1) h2 / (8π2I )    with degeneracy grot = (2 J+1)   where J = 0, 1, 2, 3, .....
As the above expression obeys the criterion that the ground rotational level (J = 0 level) has zero energy, so it is already consistent with the aforesaid popular convention (i.e., it can be safely used to find qrot).
Note: Here I is the moment of inertia (a characteristic molecular constant) of the diatomic molecule equal to m Re2 , where m is the aforesaid reduced mass (mamb)/(ma+mb) and Re is the equilibrium internuclear distance (popularly known as the bond-length).

So the rotational partition function is qrot = J=0Σa (2J+1) exp {– J(J+1)h2 / (2I kT )}
Here h2/(2Ik) is another characteristic molecular constant with the dimension of temperature. This quantity is called the characteristic rotational temperature, denoted as θrot. Also, h/(2Ic) is another molecular constant with the dimension of wavenumber, known as the spectroscopic rotational constant, B [so that  h2/(2I) = Bhc and  θrot = Bhc/k]. Thus we get:   qrot = J=0Σa (2J+1) exp {– J(J+1)θrot / T }

There are several approximate approaches to get simpler expressions for this sum. The simplest approach, valid for rot / T) << 1 (i.e., for T >> θrot) uses the idea that under the condition T >> θrot, the above sum could be transformed to the integral I stated below:
I = J=0+a (2J+1) exp {– J(J+1)θrot / T } dJ. Using the transformation v = J(J+1)θrot / T    we get dv = (2J+1)(θrot / T ) dJ  or    (2J+1) dJ = (T/θrot) dv.    As at J = 0, v = 0 and at
J = + a, v = +a, we finally get: qrot = (T/θrot) v=0+a exp(–v) dv = (T/θrot)    [as, clearly, 0+a exp(–v) dv = 1].  So, we get the following simplest, most used expression for this qrot:
 
qrot    =  T/θrot    =  (2I kT / h2)  
  =  kT/(Bhc)      [valid for T >> θrot]

Above expression is better than 1% accurate for T > 33 θrot. For T not so extremely larger than θrot but rot / T) still smaller than 0.7, the four-term Mulholland expression
[qrot = (T/
θrot) {1 + (1/3).(θrot/T) + (1/15).(θrot/T)2 + (4/315).rot/T)3}] is good enough.
However for rot / T) 0.7, the following direct 4-term summation is sufficiently accurate:
  qrot = 1 + 3
exp (– 2 θrot / T ) + 5 exp (– 6 θrot / T ) + 7 exp (– 12 θrot / T )
(
Please check this expression: i.e., do find out wherefrom the above numbers 3, 2, 5, 6, 7 & 12 appear!)

It will be shown [from a detailed discussion of the symmetry-interaction of their intra-nuclear and rotational states] that for the homonuclear diatomic molecules (e.g., H2, O2, N2) qrot is: 
 
qrot    =  T/(2θrot)    =  2I kT / (2h2)     =  kT/(2Bhc)      [valid for T >> θrot]

Introducing the symmetry factor s the value of which is 2 for homonuclear diatomic molecules and 1 for heteronuclear ones, we get the following general relation for T >> θrot
 
qrot    =  T/(sθrot)    =  2I kT / (sh2)     =  kT/(sBhc)      [valid for T >> θrot]

The above relations are valid even for linear polyatomic (i.e., with atomicity a > 2) molecules (for CO2, s is 2, not 1 – can you tell why?). For non-linear polyatomic molecules, the analogue of the above relation (obviously valid only for some sufficiently large values of T) is:
qrot   = (p0.5/s).{83 π6 (Ixx Iyy Izz) k3 T3 / h6}0.5   =  (p0.5/s).{T 3 / (θrot-x θrot-y θrot-z )}0.5 
(here Ixx, Iyy and Izz are the 3 moments of inertia along the 3 perpendicular directions, and
s = 3 for NH3, 12 for CH4 etc. Guess defining expressions for θrot-x , θrot-y and θrot-z yourself!)

As a simple example, it is known that for ordinary O2 (i.e., for O2 containing only the O16 isotope) θrot = 2.07 K. So at 298 K, the condition T >> θrot is obeyed, and we get qrot = T/(sθrot) where s = 2 (as ordinary O2 is homonuclear). So, here, qrot = 71.98. We may also get that for O2, the spectroscopic rotational constant B = kθrot/(hc) [from relation θrot = Bhc/k] is obviously 143.9 m–1 = 1.439 cm–1. Also, using h2/(2Ik) = θrot we get that here the moment of inertia I = mRe2 equals I = h2/(2kθrot) = 1.945 x 10–46 kg m2, so that (we know that for O2, m is obviously 32*32/(32+32) a.m.u. = 2.657 x 10–26 kg) the bond-length
Re = √( I/
m) = 0.8556 x 10–10 m = 0.8556 A0. Thus, from knowledge of θrot or of B, even the bond-length value of a diatomic molecule may be calculated (and vice-versa)!

To calculate or to compare populations in two rotational levels (irrespective of the molecular occupation of other-mode levels) the (mode-specific form of the) Boltzmann distribution law (or its corollary about ratio of populations) comes handy. Thus for HCl gas for which
θ
rot
= 15.02 K, the ratio of populations at J = 1 (the 1st excited) and J = 3 (the 3rd excited) rotational levels at a temperature T = 300 K are (see section 2.2 for the background theory):
  population-ratio  NJ=3/NJ=1 = (gJ=3/gJ=1). exp{– (εJ=3 – εJ=1)/(kT)}
   = {(2x3+1)/(2x1+1)}. exp{– (
3x4 θrot/ T – 1x2 θrot / T)}
   = {7/3}. exp{– (
12 θrot/T – 2 θrot / T)}
   = (7/3). exp(–
10 θrot / T ) = (7/3). exp (– 150.2 K / 300 K) = 2.333 x 0.6061 = 1.414

Note: Have you noted that a upper rotational level is here more (1.414 times more) occupied than a lower one! This can happen because the degeneracy of the upper level is more (i.e., gJ=3 is 7 while gJ=1 is 3).


3.7 Expression
s for the Thermodynamic Functions (U, P and S) for Gases:

Let us now find what sort of expressions for the thermodynamic functions U, P & S of gases arise from the above knowledge about qtr, qel, qnuc, qvib, and qrot (as q = qtrqelqnuc for monatomic gases and q = qtrqelqnucqvibqrot  for non-monatomic gases). In doing this, we'll first find generalized expressions for them in terms of the specific-mode partition functions and then apply them for specific examples of the commonly occurring gases. As q is a function of V & T, and the said thermodynamic functions are expressed in terms of N, V, T & q, so this procedure ultimately gives us these said thermodynamic quantities as just functions of N, V & T (while N, V & T are treated as independent variables). 

3.7.1 Expression for energy U (with corollary-expressions for molar heat capacities) 
For non-monatomic gases, q =
qtrqelqnucqvibqrot and so
ln q = ln qtr + ln qel + ln qnuc + ln qvib + ln qrot. Using U = NkT2(∂ lnq /∂T)V, we get
U = NkT2(∂ lnqtr/∂T)V + NkT2(d lnqel/dT) + NkT2(d lnqnuc/dT) + NkT2(d lnqvib/dT)
      + NkT2(d lnqrot/dT)             
[as qel, qnuc, qvib, and qrot
are independent of V, so the criterion about constancy of V is omitted in these four corresponding differential expressions]
So, naturally, the translational, electronic, nuclear, vibrational and rotational contributions to U are defined as:
(i) translational energy Utr = NkT2(∂ lnqtr/∂T)V    (ii) electronic energy
Uel = NkT2 d lnqel/dT     (iii) nuclear energy Unuc = NkT2 d lnqnuc/dT       (iv) vibrational energy Uvib = NkT2 d lnqvib/dT       (v) rotational energy Urot = NkT2 d lnqrot/dT
(where, obviously, U = Utr+Uel+Unuc+Uvib+Urot)

As qtr = [(2pmkT)3/2 / h3] V, so     ln qtr = ln [V(2pmk)3/2/ h3] + ln T3/2   
[where (2pmk)3/2/ h3 is obviously a constant]
So (at constant V),   
(∂ lnqtr/∂T)V  = 0 + d ln T3/2 /dT = (3/2) d ln T /dT = (3/2).(1/T)
so that    Utr = NkT2 (3/2).(1/T) = (3/2) NkT    i.e.,
Utr = (3/2) NkT

As generally qel = gel,0 (i.e., a constant), so, generally, Uel (= NkT2 d lnqel/dT) = 0.
Similarly using either qnuc = gnuc,0 or the conventional qnuc = 1, we (always) get
Unuc = 0
[
This zero value of Unuc is consistent with the convention of avoiding Unuc via use of q = qtrqelqvibqrot]

For usual diatomic gases, applying relation qvib = 1/ {1 – exp(– θvib / T )}, we get
   ln qvib  =  – ln {1 – exp(– θvib / T )}, and so    [remember that
Uvib = NkT2 d lnqvib/dT]
  
d lnqvib/dT  =  – [1/ {1 – exp(– θvib / T )}]. {– exp(– θvib / T )}. (– θvib). (– 1 / T2 )
                     
( θvib / T2 ). [exp(– θvib / T ) / {1 – exp(– θvib / T )}]  
                     
( θvib / T2 ). [1 / {exp( θvib / T ) – 1}]            
So for
such diatomic gases Uvib = Nkθvib /{exp( θvib/ T ) – 1}       (≈ 0  for T << θvib)
For polyatomic gases also,
Uvib ≈ 0 at a not-so-high temperature T.

For common diatomic (or linear-molecule) gases with T >> θrot, qrot = T/(sθrot) so that 
   ln qrot = lnT
– ln(sθrot)
This gives
d lnqrot/dT = 1/ T ,  ln (sθrot) being a constant. So,  Urot = NkT. However, for non-linear gases T >> θrot, we get  qrot = (constant). T3/2 which gives Urot = (3/2) NkT
So we get:
For linear-molecule gases Urot = NkT  but for non-linear ones Urot = (3/2) NkT

For monatomic gases, q = qtrqelqnuc and  U = Utr+Uel+Unuc. [Here Uvib & Urot (as well as qvib & qrot) simply do not exist.]

So for Uel ≈ 0 & Uvib ≈ 0 (Uvib ≈ 0 for not-so-high T), U ≈ (5/2) NkT for linear-molecule non-monatomic gases but U ≈ 3 NkT for non-linear ones. However, for monatomic gases with Uel ≈ 0, U ≈ (3/2) NkT. Talking in a per-mole basis, we thus see that the molar energy Umol ≈ (3/2) RT for monatomic gases, Umol ≈ (5/2) RT for linear-molecule non-monatomic gases but Umol ≈ 3 RT for non-linear ones (provided Uel ≈ 0 & Uvib ≈ 0).    [To get the per-mole expressions, remember the fact that NAk = R , where NA is the Avogadro number.]

The molar heat capacity at constant volume, Cv,mol is nothing but (dUmol/dT)V. So we may easily get that Cv,mol ≈ 1.5 R for monatomic gases, Cv,mol ≈ 2.5 R for linear-molecule non-monatomic gases but Cv,mol ≈ 3 R for non-linear ones (provided Uel ≈ 0 & Uvib ≈ 0). The values of the molar heat capacity at constant pressure, Cp,mol and the ratio γ may be obviously obtained from the relations    Cp,mol = Cv,mol + R    &      γ = Cp,mol / Cv,mol.
Note: Please try to show that for monatomic gases with Uel ≈ 0,
γ = 1.67

3.7.2 Expression for pressure P (with corollary-expression for enthalpy H)
We see above that qtr = (2pmkT)3/2 V/ h3 is obviously of the form V. ftr (T) while each of
qel, qnuc, qvib and qrot are either constants or functions of T only, so q is a function of only T & V in the specific form V. f (T) where f (T) is a function of T only (as mentioned earlier, in Sec 2.3). This gives that ln q = ln V + ln f (T). As P = NkT (d lnq /dV)T, we get:
   P = NkT (1/V + 0) = NkT/V. So,  i.e., for all gases we get
P = NkT / V  or  PV = NkT.

It can be easily realized that P is wholly translational in the exact sense (unlike U or S) i.e., P = Ptr –– or in other words there's no non-translational contribution to P (i.e., Pel, Pnuc, Pvib, Prot = 0). This may be understood as follows: Using q = qtrqelqnucqvibqrot, P =
NkT (d lnq /dV)T = NkT (d lnqtr /dV)T +NkT (d lnqel /dV)T + NkT (d lnqnuc /dV)T
+ NkT (d lnqvib /dV)T + NkT (d lnqrot /dV)T. we have only Ptr = NkT (d lnqtr /dV)T = NkT / V while the other four terms above are zero. So P = Ptr.

The relation PV = NkT found above immediately gives enthalpy H (= U + PV) = U + NkT (with, obviously, molar enthalpy Hmol = Umol + RT).

This relation PV = NkT which is identical with the experimentally observed ideal gas relation PV = nRT = NkT  [as n = N/NA so that Nk = nNAk = nR] also means the constant k found in the defining relation q = Sl gl exp {-el /(kT)} for q is nothing other than the Boltzmann constant k (= 1.381 x 10–23 J K–1) found in the ideal gas relation PV = NkT.

3.7.3 Expression for entropy S (with corollary-expression for Helmholtz free energy A)
Starting with the relation q =
qtrqelqnucqvibqrot for non-monatomic gases and using
S = Nk ln(q/N) + U/T
+ Nk ,  we get: 
  S = Nk ln qtr + Nk ln qel + Nk ln qnuc + Nk ln qvib + Nk ln qrot – Nk ln N
        + (
Utr+Uel+Unuc+Uvib+Urot) / T + Nk
In this expression, we see that if only the translation motion would have existed, i.e., if q would have equaled just qtr, then S would have been (Nk ln qtr – Nk ln N +
Utr / T + Nk). [Here Utr/T = 1.5 Nk] So we call this quantity as Str, the translational contribution to entropy (i.e., translational entropy). Definition of the other four contributions are rather obvious:
(i) electronic entropy Sel = Nk ln qel + Uel / T     (ii) nuclear entropy
Snuc =
Nk ln qnuc + Unuc / T       (iii) vibrational entropy Svib = Nk ln qvib + Uvib / T     
(iv) rotational entropy Srot =
Nk ln qrot + Urot / T
[where, obviously, S = Str+Sel+Snuc+Svib+Srot
with Str = Nk ln (qtr / N) +  Utr / T  +  Nk]

The above general expression for Str attracts some special interest, a simplified per-mole form of it being called by a special name. To arrive at that per-mole relation, we take [noting that Utr / T = (3/2) Nk ]
Str = Nk ln (qtr / N) + (3/2) Nk + Nk =
Nk ln (qtr / N) + (5/2) Nk
For one mole of the gas, N = NA (with NAk = R) and transl.-entropy is Str,mol. So we have:
Str,mol
= 2.5 R + R ln (qtr / NA). Putting qtr = (2pmkT)3/2 V/ h3 where m = M/NA we get (here M is the molar mass of the gas):    Str,mol = R [2.5 + ln {(2pMkT)3/2 V/ (NA5/2 h3)}] 
This means,
Str,mol = R [2.5 + ln {(2pk)3/2 / (NA5/2 h3)} + 1.5 ln M + ln V + 1.5 ln T]. This is the Sackur-Tetrode equation, where ln {(2pk)3/2 / (NA5/2 h3)} is a constant (its value in SI unit being the number 16.016). 

The Sackur-Tetrode equation could be written also in an alternative form, involving the pressure P instead of the volume V (via use of PV = NkT), and involving the relative molecular mass Mr instead of the molar mass M. To get this form, we note that for one mole V = NAkT/P and that m = Mr mo (where mo is the atomic mass unit - a.m.u., equal to 1.6606 x 10–27 kg, i.e., 1/12 -th of the mass of a C-12 atom). Thus the relation 
Str,mol = R [2.5 + ln {
(2pmkT)3/2 V/ (NA h3)}] gives:
      Str,mol = R [2.5 + ln {
(2pMr mo kT)3/2 NAkT/ (PNA h3)}] 
           = R [2.5 + ln {
(2pmo)3/2 ( kT)5/2 / h3} – ln P + ln Mr]. Putting P = Po (P/Po) where Po is the standard reference value of 1 atm (i.e., 1.013 x 105 N m–2), we get ln P = ln Po + ln (P/Po). So, Str,mol = R [2.5 + ln {(2pmo)3/2 ( kT)5/2 / h3} – ln Po – ln (P/Po) + ln Mr
        = R [2.5 + ln {
(2pmo)3/2  k5/2 / (Poh3)} + (5/2) ln T – ln (P/Po) + ln Mr]
  or,   
Str,mol = R [2.5 + ln {(2pmo)3/2  k5/2 / (Poh3)} + 1.5 ln M – ln (P/Po) + 2.5 ln T ]
Here also ln {
(2pmo)3/2  k5/2 / (Poh3)} is a constant (its value in SI unit being –3.6639).

From Sel = Nk ln qel + Uel / T  we find that for the usual case of qel = gel,0 (along with
Uel = 0), the electronic entropy Sel = Nk ln gel,0 (so, if gel,0 = 1, Sel = 0). On the per-mole basis, Sel,mol = R ln gel,0.
Similarly, as always Unuc = 0, so Snuc = Nk gnuc,0 in the exact sense; however, if we are interested only in the entropy change in the usual physical/ chemical processes, the convention qnuc = 1 giving simply Snuc = 0 is perfectly good enough.

Svib and Srot (as you must have understood by now) arise only for non-monatomic gases. Using the diatomic-gas Uvib expression  Uvib = Nk θvib [1 / {exp( θvib / T ) – 1}], we get that for diatomic gases Svib = Nk ln [1/{1 – exp(– ζvib)}] + Nk ζvib. [1 / {exp(ζvib) – 1}] where ζvib = θvib / T. However for the usual values of temperature with T << θvib , ζvib is very large, giving Svib ≈ 0 – so there's hardly much need to calculate Svib. Similarly for the polyatomic gases also, in the usual cases, Svib ≈ 0.
The rotational entropy Srot, however, evokes much more interest because of its non-negligible value.
For diatomic and other linear molecules that obey qrot = T/(sθrot) – and thus Urot = NkT, we get: Srot = Nk ln{ T/(sθrot) } + Nk = Nk [ ln{ T/(sθrot) } + 1], the molar value being Srot,mol = R.[ln{ T/(sθrot) } + 1]. It can be easily shown that (try it yourself) that for non-linear molecules at sufficiently enough temperatures, 
Srot = Nk [ ln qrot
+ 1.5].

The mode-specific contribution to A are defined with the help of those for U and S. Thus 
A
tr = Utr – TStr = – NkT ln (qtr / N) – NkT. The other contributions to A are as follows:
 
Ael = – NkT ln qel, Anuc = – NkT ln qnuc, Avib = – NkT ln qvib, Arot = – NkT ln qrot.

Let us now calculate the molar translational and the molar rotational entropy of HCl at 298.15 K temperature and 1 atm pressure, given that for HCl, θrot = 15.3 K (you may note that for this case Sel and Svib are practically zero). Using PV (= NkT) = nRT, we get
V = (1mol x 0.08206 L atm K–1mol–1 x 298.15 K/ 1atm) = 24.466 L = 2.4466 x 10–2 m3
Using the Sackur-Tetrode equation, the molar
translational entropy is (M = 36.5 x 10–3 kg):
   Str,mol = R [2.5 + ln {(2pk)3/2 / (NA5/2 h3)} + 1.5 ln M + ln V + 1.5 ln T]
       = R [2.5 + 16.016 + 1.5 ln
(36.5 x 10–3) + ln (2.4466 x 10–2) + 1.5 {ln (298.15)}]
  = 8.314 J K–1mol–1. [2.5 + 16.016 – 4.9657
– 3.7105 + 8.5464]  = 152.9 J K–1mol–1
Similarly, using Srot,mol = R [ln qrot + 1] = R [ln{ T/(sθrot) }
+ 1]   for diatomic gases, we get (for HCl being heteronuclear, s = 1):
   Srot,mol = R. [ln (298.15 K / 15.3 K) + 1]
               =
8.314 J K–1mol–1. [ln 19.487 + 1] =  33.00 J K–1mol–1
Note: Please cross-check the above value (~153 J K–1mol–1) of Str,mol using the alternative Sackur-Tetrode form [in doing so use Mr = 36.5,  (P/Po) = 1 and T = 298.15]. 


3.8 Statistical-Thermodynamic Calculation of Entropy of Gases in Contrast to its Calorimetric Measurement -- The Concept of Residual Entropy of Solids at 0 K
:

Similar to the Sackur-Tetrode calculation of molar translational entropy, the absolute (actual) entropy per mole of a (pure) nearly-ideal gas may be calculated (Subsection 3.7.3 above) by calculating its various components, taking care to include the (generally ignored) nuclear component (~ R.gnuc,0). On the other hand, the difference DSmol = Smol(T)–Smol(0) between the molar entropy Smol(T) of a pure gas at the given temperature T (Kelvin), and the entropy Smol(0) of its pure crystalline solid form at 0 K temperature (0 K in only a practical sense, as exactly absolute zero temperature can't be attained), may be determined calorimetrically via measurements of the molar heat capacity (at constant pressure) Cp at different temperatures in its various (solid, liquid, gas etc.) phases, and of the molar enthalpies DHmol of phase transitions: the molar entropy change DSmol is a sum of several integrals of (Cp/T).dT, throughout the complicated transition of the solid at 0 K into the nearly-ideal gas at T K, plus the familiar phase-transition contributions (DHmol/Tph-tr) to this entropy. However, the third law of thermodynamics states that the entropy of a pure perfectly crystalline solid at absolute zero temperature is zero, implying thereby that the aforesaid Smol(0) = 0 so that DSmol = Smol(T), and thus meaning that the aforesaid calorimetric determination of DSmol also determines the molar entropy Smol(T) of the gas at the given temperature. Thus there is a provision for comparison of the gas-phase entropy predicted by statistical thermodynamics on one hand, and that determined by calorimetric measurements on the other.  

However for some substances, such as carbon monoxide, it is seen that the statistical-thermodynamically predicted and the calorimetrically measured values of the gas-phase entropy do not exactly match. In such cases, the discrepancy that arises invariably involves a somewhat lower value for the calorimetrically measured entropy. The obvious explanation for this is that the entropy of the crystalline solid phase of that substance at the absolute zero temperature {i.e., Smol(0) above} is not exactly zero (thus apparently going against the third law). Such a non-zero, positive entropy at the absolute zero temperature for the crystalline solid phase of a substance is called its residual entropy. Actually, the residual entropy arises because of a disorder or randomness of the molecular arrangement within the crystal lattice of some substances, even at temperatures approaching absolute zero, and implying that such crystals are not exactly perfect ones. Thus the third law of thermodynamics is not actually being disobeyed in such cases!

As an example, carbon monoxide is found to crystallize with the CO molecules linearly oriented in its crystal lattice. But during crystallization, many CO molecules were found to align themselves in the direction exactly opposite to that for the other CO molecules (figure 3.1). 

 Figure 3.1: The random linear alignments of molecules in crystalline CO

Thus there is a disorder or randomness in the crystal lattice -- as there is freedom about which of the CO molecules would be oppositely oriented. A theoretical estimate about the residual entropy of the crystal in such situations concerns the thermodynamic probability or the number of distinct ways W in which the crystal structure may be attained, involving the Boltzmann equation S = k.lnW

As an example, for CO a reasonable assumption would be to consider half of the N molecules of the crystal to be aligned in an opposite way. Assuming this, the number of ways in which the crystal structure may be attained equals W = N!/{(N/2)!.(N/2)!}, i.e., the number of ways in which the N/2 number of oppositely aligned molecules may be chosen out of the total N molecules. [It may be noted here that as the molecules in the crystal are localized in space they are distinguishable, unlike the molecules in gaseous carbon monoxide, and so we get this value for W.] 
Thus here we have, S = k.lnW = k ln [N!/{(N/2)!.(N/2)!}]
        = k [ln(N!) – ln{(N/2)!} – ln{(N/2)!}]  = k [ln(N!) – 2ln{(N/2)!}]
Applying Stirling's approximation to ln(N!) & ln{(N/2)!}, we get:
     S = k [N lnN – N – 2{(N/2) ln(N/2) – (N/2)}
        = k [N lnN – N – N ln(N/2) + 2(N/2)]
        = k [N lnN – N lnN + N ln2] = Nk ln2
Thus the molar residual entropy here is NAk ln2 = R ln2 = 11.758 J K–1mol–1. This predicted estimate for residual entropy of CO is of the same order to its reported experimental values.


3.9 Looking back towards a Verification for the Relation b = 1/(kT)
:

In Section 2.2 we assumed that the Lagrange undetermined multiplier b (in the Boltzmann distribution law and in the definition of q) equals 1/(kT) [where k is the Boltzmann constant and T is the thermodynamic temperature]. Let us now verify this relation, as we are now in a position to do so!

Let us begin with the relation q = Sl gl exp (-bel) from Section 2.2, and differentiate it w.r.t. b at constant volume V. This gives:
(dq/d
b)V-Sl gl el exp (-bel),  or,   Sl gl el exp (-bel)  =  - (dq/db)V
Using
Nl
= (N/q) gl exp(-bel) from Section 2.2, we get:
U
=  Sl Nl e = (N/q) Sl gl el exp (-bel ) = - (N/q) (dq/db)V -N (d ln q /db)V  
So, the translational system-energy is definable as:   Utr = -
N
(d ln qtr /db)V.
Avoiding the replacement
b = 1/(kT) in the definition of qtr, we may easily derive that
qtr = (2pm/
b)3/2 V/ h3 , so that    ln qtr = ln {(2pm)3/2 V/ h3} - (3/2) ln b. This gives:
 
(d ln qtr /db)V = 0 - (3/2).(1/b) = -(3/2).(1/b).  This gives:
 
Utr = -N.{-(3/2).(1/b)} = (3/2). (N/b)

As we know (from experimental knowledge and from kinetic theory of gases etc.) that the translational (kinetic) energy for gases is Utr = (3/2).NkT, so obviously,
(3/2)(N/
b) = (3/2)NkT  implying  b = 1/(kT).

Note: Using b = 1/(kT) we arrived at the approximate Boltzmann equation S = k lnW*  (where k is the Boltzmann constant). As  lnW* ≈ lnW, we get that the constant k in the Boltzmann equation S = k lnW is also nothing other than the Boltzmann constant!

 
3.10 Looking back at the rotational p. f. for
homonuclear diatomic molecules:

In any discussion about the rotational p. f. of homonuclear diatomic molecules, we need to note that all the rotational states of such molecules can't coexist with all the nuclear states, because of certain symmetry consideration of the complete molecular wavefunction with respect to exchange of the pair of identical nuclei therein. Just as any many-electron electronic wavefunction (of an atom or molecule) must be antisymmetric with respect to exchange of any pair of electrons therein (electrons being a set of identical Fermions), the complete molecular wavefunction of a molecule (e.g., CH4) with two or more identical nuclei (e.g., the four 1H nuclei in CH4) must also either be antisymmetric (for Fermion kind of nuclei, e.g., for 1H as in CH4) or be symmetric (for Boson kind of nuclei, e.g., for 14N in N2 or 16O in O2) with respect to exchange of any pair of identical nuclei therein. Now, the complete molecular wavefunction f of a molecule is a product of five specific-mode wavefunctions: a translational, an electronic, a nuclear, a rotational and a vibrational wavefunction, i.e., f = ftr fel fnuc frot fvib. Out of these, each of ftr, fel & fvib are always symmetric with respect to exchange of a pair of identical nuclei, and so these three never determines the overall symmetry of f. It is the rotational wavefunction frot and nuclear wavefunction fnuc that are, with respect to exchange of a pair of identical nuclei in the molecule, symmetric in some cases and antisymmetric in other ones: thus frot is symmetric for even-J states (i.e., for J = 0, 2, 4, 6 etc.) and anti-symmetric for odd-J states (i.e., for J = 1, 3, 5 etc.); whereas fnuc is antisymmetric if the nuclear-spin function part of fnuc is of antisymmetric type [e.g., of the form w(a,b) = {an(a) bn(b) – an(b) bn(a)} for H2 that satisfies w(b,a) = –w(a,b), this w(a,b) being built from the two one-nuclear spin functions an (corresponding to mI = ½) & bn (for mI = –½) found in the 1H nuclei], otherwise fnuc is symmetric. Furthermore, we may note that a pair of identical nuclei are Fermions if the nuclear-spin number I is half-integral (e.g., 1H with I = ½) and Bosons if the nuclear-spin number I is integral (e.g., 14N with I = 1, or 16O with I = 0). 

Thus, for homonuclear diatomic molecules with half-integral nuclear-spin number I (e.g., H2) the nuclei are Fermions, and so the complete molecular wavefunction f must be antisymmetric. Thus there are two kinds of H2 molecules: in the ortho variety (~75% for H2 molecules are found to be ortho-hydrogen at room-temperature) the nuclear wavefunction fnuc is symmetric, coexisting with only the odd-J rotational states so that frot is antisymmetric, making f (= ftr fel fnuc frot fvib) antisymmetric. On the other hand, in the para variety (at room-temperature, the rest ~25% of H2 molecules are found to be para-hydrogen) the nuclear wavefunction fnuc is anti-symmetric, coexisting with only the even-J rotational states so that frot is symmetric, again making f antisymmetric as a whole. Now there are three symmetric nuclear- spin functions w(a,b) for H2: w(a,b) = an(a) an(b), bn(a) bn(b) & {an(a) bn(b) + an(b) bn(a)}, in contrast to the only one antisymmetric w(a,b) mentioned above, and so there are 3 symmetric nuclear ground-level wavefunctions in H2 in contrast to only 1 antisymmetric one. Thus we get that the combined rotational-nuclear partition function for the homonuclear diatomic molecule H2 is:
qrot-nuc =  
1{1 + 5 exp (– 6 θrot / T ) + 9 exp (– 20 θrot / T ) + ..... }
             + 3{
3 exp (– 2 θrot / T ) + 7 exp (– 12 θrot / T ) + 11 exp (– 30 θrot / T ) + ....}
dividing this by nuclear p. f. expression qnuc = (2I + 1)2 = 4 for the homonuclear diatomic molecule H2 to get qrot (as qrot-nuc must, obviously, equal qrot qnuc), we finally get: 
qrot =  (1/4)
{1 + 5 exp (– 6 θrot / T ) + 9 exp (– 20 θrot / T ) + ..... }
         + (3/4){
3 exp (– 2 θrot / T ) + 7 exp (– 12 θrot / T ) + 11 exp (– 30 θrot / T ) + ....}
where the upper curly-bracketed sum is over the even values (0, 2, 4 etc.) of J, whereas the lower curly-bracketed sum is over the odd values (1, 3, 5 etc.) of J
(please verify that the coefficients 1, 5, 9 etc. and the exponents 0, 6, 20 etc. are indeed so). In general for homonuclear diatomic gases consisting of Fermion-nuclei we have 
qrot =  {I/(2I+1)}
{1 + 5 exp (– 6 θrot / T ) + 9 exp (– 20 θrot / T ) + ..... } + 
 {(I+1)/(2I+1)}{
3 exp (– 2 θrot / T ) + 7 exp (– 12 θrot / T ) + 11 exp (– 30 θrot / T ) + ...}
Now for temperatures T >> θrot both the curly-bracketed sum are naturally equal, and as their total sum (i.e., the unrestricted-J sum giving qrot in the heteronuclear case) was earlier shown to be (T/θrot) as in Section 3.6, so each curly-bracketed sum above is (T/θrot)/2, and so qrot here equals [(1/4)(T/ θrot)/2 + (3/4)(T/ θrot)/2] i.e., (T/ θrot)/2. This is the basis behind the formula qrot =  T/(2θrot) for homonuclear diatomic molecules, valid under the condition T >> θrot, and already mentioned in Section 3.6 earlier.

The ortho and para forms of a homonuclear diatomic molecule are rather distinct and can even be separated. The ratio of the numbers of ortho- and para- molecules are naturally given by their respective contributions to the combined rotational-nuclear p.f. and thus to the complete molecular partition function. So for H2 gas at equilibrium at a temperature T, the ratio ro/p = Northo/Npara  of ortho-hydrogen and para-hydrogen molecules is: 
Northo/Npara = 3{
3 exp (– 2 θrot / T ) + 7 exp (– 12 θrot / T ) + 11 exp (– 30 θrot / T ) + ....} / {1 + 5 exp (– 6 θrot / T ) + 9 exp (– 20 θrot / T ) + 13 exp (– 42 θrot / T ) + ..... }
For H2, θrot is rather high (85.3 K) so that at the room temperatures the condition T >> θrot is not exactly valid, and so for H2 gas at room temperature, ro/p only approximately equals its high temperature limit 3.0. At very low temperatures, ro/p for H2 approaches 0.
Problem: Show that at 300 K, ro/p for H2 is 2.990, but at 20 K it is only 0.002! Also show that even at 300 K, qrot for H2 is 1.934, a lot different from the T/(2θrot) value 1.758! 

On the other hand for homonuclear diatomic molecules with integral nuclear-spin number I (e.g., 14N2, 16O2) the nuclei are Bosons, and so the complete molecular wavefunction f must be symmetric. If, however, I = 0 (e.g., 16O2) then only one spin state (symmetric) is there, so that there are only ortho variety existing, with only the even-J rotational states allowed. But for integral yet non-zero I (e.g., for 14N2 with I = 1) there are both symmetric and anti-symmetric spin states. So for such molecules even-J rotational states coexist with symmetric nuclear states giving the ortho form (67% of N2 at room-temperature are ortho-nitrogen), while odd-J rotational states coexist with anti-symmetric nuclear states giving the para form (the rest 33%). At any temperature T, qrot and equilibrium ro/p are given here by the relations:
qrot =  {(I+1)/(2I+1)}
{1 + 5 exp (– 6 θrot / T ) + 9 exp (– 20 θrot / T ) + ..... } + 
 {I/(2I+1)}{
3 exp (– 2 θrot / T ) + 7 exp (– 12 θrot / T ) + 11 exp (– 30 θrot / T ) + ...}
and
ro/p = {(I+1)/I} {
1 + 5 exp (– 6 θrot / T ) + 9 exp (– 20 θrot / T ) + 13 exp (– 42 θrot / T ) + ..... } / {3 exp (– 2 θrot / T ) + 7 exp (– 12 θrot / T ) + 11 exp (– 30 θrot / T ) + ....}