CHEM2410 FinalRevision
CHEM2410 FinalRevision
Physical Chemistry I
REVISION: Chapters 5 ,6 ,12
Course Instructor:
Prof. Jonathan Halpert
CHEM 2410: Chapter 5
CHAPTER OUTLINE
Thermodynamics of mixtures Phase Diagrams of Binary Systems
1. Partial Molar Quantities 6. Vapor Pressure Diagrams
2. Thermodynamics of Mixing 7. Temp/Composition Diagrams
3. Chemical Potential of Liquids 8. Liquid-Liquid Phase Diagrams
9. Liquid-Solid Phase Diagrams
Properties of Solutions
4. Liquid Mixtures Activities
5. Colligative Properties 10. Solvent Activity
11. Solute Activity
12. Activity of Regular Solutions
13. Activities of Ions in Solution
Required Reading:
“Atkin’s Physical Chemistry”, Atkins, de Paula & Steeler 11th Ed. 2018: Chapter 5
Simple Mixtures
State Properties
Partial Molar Volume
State Properties
V V
dV = dnA + dnB dV = VA dnA + VB dnB
nA p ,T ,nB nB p ,T ,nA
nA nB
only true if both integration steps are at the same
By integrating V = dV = VA dnA + VB dnB = VA nA + VB nB time so that nA/nB (and thus dVA/dVB) is constant
0 0 (as we will prove later on)…the result is obviously
true for the marginal values of VA and VB
Partial Molar Gibbs Energy = μn
State Properties
such that for a two component mixture of A and B, G total changes by dG:
G G
dG = dnA + dnB dG = A dnA + B dnB
nA p ,T ,nB nB p ,T ,nA
nA nB
G
J = n2
J p ,T ,n '
n
n1
G G
dG = dnA + dnB
nA p ,T ,nB nB p ,T ,nA
dG = A dnA + B dnB G = A nA + B nB n1 n2
• the slope of line G (of the system) vs nA gives the chemical potential at each amount of A (or xA = nA/n)
• the chemical potential is strongly dependent on the relative amount, xA in a mixture
Gibbs-Duhem Equation
State Properties
Gibbs-Duhem Equation (p,T constant)
We derived that for a binary systems G = A nA + B nB so for an infinitesimal change in G, dG (p,T const.)
however dG p ,T = A dnA + B dnB from our earlier derivation (which was valid!)
since G is a state function, these two functions must be equal, which implies that
The Gibbs-Duhem Equation, suggests that changes in the chemical potential of any one species relies on
changes in the chemical potential of the other species
nB
nA d A = −nB d B d A = − d B see addendum for example
nA this reasoning can be extended to dV = − nB dV
A B
any partial molar quantity, e.g. VJ nA
Gibbs Energy of Mixing: Gases
State Properties
Example: Gibbs energy of mixing
A container is divided into two equal compartments. One contains 3.0 mol H2(g), the other contains 1.0 mol N2(g),
both at 25 °C. Calculate the Gibbs energy of mixing when the partition is removed, assume perfect behavior.
initial pressure of N2 side is p, pressure of H2 side is 3p (Vi,Ti equal)
final partial pressure of N2 is 0.5p and of H2 is then 1.5p
Gi = nA ( A + RT ln p ) + nB ( B + RT ln p )
A B
G f = nA ( A + RT ln p A ) + nB ( B + RT ln pB )
G f = ( 3.0 mol ) H 2 + RT ln (1.5 p / bar ) + (1.0 mol ) N2 + RT ln ( 0.5 p / bar )
change
vapor pressure exists b.c. the entropy is increased by having molecules in the vapor phase, in a mixture the
entropy of the liquid phase is higher, so less entropic gain to having vapor…thus pA is lower than pA*
Chemical Potential of Mixing: Solutes
State Properties
Chemical Potential of Solutes in Solvents
A solute is much more dilute, surrounded by solvent molecules and is thus in a very
different environment:
In cases where the solvent molecules are very similar to the solute molecules, the local
environment is similar and the solute may obey Raoult’s Law
ideal-dilute solutions
mixture that obeys Raoult’s Law for the solvent, and Henry’s Law for the solute
Thermodynamics of Mixing: Liquids
State Properties
Gibbs Energy, Entropy, Enthalpy of Solutions
To find the Gibbs energy of mixing (just like for gases) vapor and liquid are at
equilibrium, so u(l)=u(g)
Gi = nA *A + nB *B from Raoult’s Law, So we calculate Gmix for
since xA=pA/p*
( ) ( )
Gvapour to get Gliquid
G f = nA *A + RT ln x A + nB *B + RT ln xB
mix G = nA RT ln x A + nB RT ln xB
mix G = nRT ( x A ln x A + xB ln xB ) Gibbs Energy of mixing to form an ideal solution
(very similar to gas mixing)
The entropy for mixing ideal solutions (same steps as gas mixing)
mix S = −nR ( x A ln x A + xB ln xB )
the change in volume comes from:
The enthalpy for mixing ideal solutions is thus:
mix H = 0 G mix G
= V so V = =0
p T ,nA ,nB p T ,nA ,nB
mix
Implications for Real Gibbs Energies
Colligative properties: properties that depend on the amount of the solute species present but not its identity.
Colligative properties arise from changes in the chemical potential of the solvent due to entropic effects.
(which is why molecular identity is not important, just the amounts of other species)
These properties are explored here for dilute solutions of A and B such that nA>>nB and thus xA>>xB
Common features of colligative properties
A (l ) = *A (l ) + RT ln x A
for solvent A, the liquid has a reduced μ due to RT ln xA<0 but the solid is
*A ( g ) = *A (l ) + RT ln x A
unaffected (assumption): thus the gas phase becomes lower in G at a higher T
With the presence of the solute B, the Tb* (at 1 atm) is raised to Tb= Tb* + ∆Tb where
R (Tb* )
2
Tb = KxB and K = xB is the mole fraction of the solute in the solvent
vap H
For practical use (like with other empirical expressions), the equation is:
*A ( s) = *A (l ) + RT ln x A
By similar arguments, the liquid potential μ(l) is lower so G(l) is still lower than
G(s) even at smaller T…thus the freezing point, where G(l)=G(s), decreases
With the presence of the solute B, the Tf* (at 1 atm) is lowered to Tf= Tf* – ∆Tf where
R (T f * )
2
For practical use (like with Henry’s law and other empirical expressions), the equation is
fus H 1 1
ln xB = −
R T f T
=fusH/(RT*)
Osmosis
From the Greek “to push”, osmosis is the spontaneous passage of pure solvent into a solution separated from
it by semi-permeable membrane that is permeable to solvent but not to the solute
The derivation requires the chemical potential be equal on each side of the
membrane at a pressure (p+Π) at which solvent no longer passes through
x A p*A
yA = and yB = 1 − y A
x A ( p A − pB ) + pB
* * *
pB* p*A
in addition, we can p= to get the vapor pressure from the fraction of A in
rearrange this into y A ( pB* − p*A ) + pB* the vapor (instead of xA in the liquid)
(several steps, algebra!)
Mixed Phase Diagrams: Binary Systems
State Properties
Counting molecules, finding fractions
We can combine the two diagrams to form a single mixed phase
diagram (for a closed system, e.g. in a piston chamber) vs p (at r.t.)
Lα Lβ
Temperature vs. Composition: Liquid-Vapor
State Properties
Temperature Composition Diagrams
p = p*A + pB*
the fraction recovered is effectively the this is useful for heat sensitive, water insoluble compounds as
partial pressure ratio they can be collected at a lower temperature
Partially Miscible Liquids
State Properties
• For points in the two-phase regime, the tie-line is used The upper critical point is the T at which both
to find the relative quantities of the phases by the components are miscible at all compositions
lever rule nαLα = nβLβ
Low temperature miscibility
• Some species, (trimethylamine), form stable complexes at low T, that break apart at high T
• This causes a lower critical temperature (Tlc) above which two phases can form at some compositions
• Other species exhibit both a Tlc and Tuc: first forming stable compounds, later high T makes miscibility spontaneous
Distillation of Partially Miscible Liquids
State Properties
In a eutectic system:
• The line b3 to a3 is called a peritectic line, at which K Position of the product is noticeable
suffuses into the solid Na and reacts to form Na2K due to the 2:1 ratio needed to form it
Multiple Products
So rather than change the form of the expression for chemical potential (based on an ideal solvent), we can
instead alter the ideal ln(xA) to be the real ratio ln(p/p*)
This ratio is called the activity (aA) the real partial pressure ratio, substituted as an effective mole fraction
(instead of the actual mole fraction xA)
This is similar to our use of fugacity, the idealized version of the real pressure, such that equations for
perfect gases worked… by translating real p to f, we could then plug f into the Gibbs equations
In this same way, for any mole fraction, we can determine the activity and plug that into the chemical
potential equation
“gamma”, the activity coefficient, a property of the real solvent
pA
A = A* + RT ln ( a A ) where aA or a A = A x A so A = A* + RT ln ( xA ) + RT ln ( A )
p*A
If the solvent is ideal, then aA = xA and γ = 1 ln(ab)=ln(a) + ln(b)
Real Solutes: Activity
Unlike the ideal solvent (which is pure), the ideal solute is dilute so γ→1 as xB→0
Implications of the Margules Equations
a A = A x A = x A exp (1 − x A ) pA = p x = p x e
2 * *
A A A A A
• positive values of ξ, (endothermic mixing, unfavorable interactions) give pA higher than ideal
• negative value of ξ, (exothermic mixing, favorable interactions) give pA lower than ideal
(1− x A )
2
• plots become linear as x→1, and coincide with Raoult’s Law since e → 1 as (1-x)→0
p
• as x→0 equation becomes A = p *
x
A A e K
which is Henry’s Law with A = x A e
Activities of Ions
State Properties
Activities of Ions
If the chemical potential of the cation M+ is μ+ and that of the anion is μ– then
Gmideal = + + −
J ideal = J + RT ln xJ or J real = J + RT ln aJ
Gmreal = Gmideal + RT ln ( + − ) where all the deviations from ideality are captured in the last term
Mean Activity Coefficients
Previously Gmreal = Gmideal + RT ln ( + − ) but it is difficult to assign separate activities to the individual ions
Instead we can find the geometric mean of the activities of the two species and find the mean activity coefficient:
= ( + − ) Gmreal = Gmideal + 2 RT ln ( )
12
so that for an ionic species M+X-
We can generalize this to MpXq which creates p moles of cations and q moles of anions per mole of MX
but if we take the geometric mean to get a mean activity coefficient (general expression)
= ( + − )
p q 1 ( p+q)
then i = i ideal + RT ln for each ion, i
then finally
real
G
m = p+ + q− but how to get γ± ? u geom. = ( x y
p
)
q 1 ( p+q)
Debye-Hückel Limiting Law
Debye-Huckel Theory: the long range of Coulombic (charge-charge) interactions between ions
means that this is likely the dominant factor in the non-ideality of solutions
Charged ions attract each other, so each ion is likely to have a nearby cloud of oppositely charged
ions called the ionic atmosphere: the energy (and thus μ) of the ion is lower due to this interaction
Debye-Huckel
model
The stabilization of ions due to local electrostatic interactions explains why chemists use dilute
solutions to achieve precipitation of ions from solution (less stabilization)
At very low concentration the activity can be derived from the Debye-Huckel Limiting Law: molality of ion, i
a constant related to dielectric of the solvent
1 2 bi
log = − A z+ z− I 1 2 A = 0.509 for aq. solution at 25C I= i b
2 i
z
signed total charge +/- ionic strength of the solution signed charge on ion, i
Called the “limiting law” because the activity coefficient becomes more accurate as the solution is more dilute
CHEM 2410: Chapter 6
CHAPTER OUTLINE
Spontaneous Chemical Reactions Equilibrium Electrochemistry
1. The Gibbs Energy Minimum 5. Half-rxns and electrodes
2. The Description of Equilibrium 6. Varieties of Cells
7. The Cell potential
The Response of Equilibria to Conditions 8. Standard electrode potential
3. Changes in Pressure 9. Applications of Standard Potentials
4. Changes in Temperature
Required Reading:
“Atkin’s Physical Chemistry”, Atkins, de Paula & Steeler 11th Ed. 2018: Chapter 6
Extent of Reaction (ξ)
State Properties
Equilibrium and Extent of Reaction
The direction of spontaneity for RTP (room temp. and pressure) experiments (decreasing Gibbs) tends toward
equilibrium (as long as there are no kinetic barriers): the point at which macroscale changes in the system stop
• At the equilibrium point, dG = 0 → represents the minimum of the Gibbs energy curve for that system
• A and B are interconverting, but at equilibrium this happens at the same rate
The extent of the reaction: a variable ξ (“xi” or “ksai”, same Greek letter but not the enthalpy variable from Ch.
5) which serves as a metric of how far the reaction goes, with units of quantity (think “moles of the reaction”)
• for a small change in the reaction dξ, a small amount of A (reactant) –dnA becomes B (product) by +dnB
• for a noticeable change Δξ the amount of A, nA,0 → nA,0 – Δξ and B goes to nB,0 → nB,0 + Δξ
Reaction Gibbs Energy (ΔrG)
State Properties
Reaction Gibbs Energy (ΔrG)
With this in mind, we can write an expression for the Gibbs reaction energy at any point during the reaction
G
dG = J dnJ = J v J d = J v J d so then rG = = vJ J
all species, J all species, J all species, J p ,T J
We can then write the chemical potentials as a function of the activities of each species
r G = v J J + RT v J ln aJ = r G + RT v J ln aJ = r G + RT ln aJ v J ( ) a ln x = ln xa
J J J J
vJ
and since ln a + ln b +… = ln (a*b*…) then we can write r G = r G + RT ln aJ
for any reaction
J
Q, the reaction quotient
r G = r G + RT ln aJ v J from the last slide, so we can write
J
( a ) (a )
vp vq
...(all products)
Q = aJ =
vJ p q
( am ) ( an ) ...(all reactants)
vm vn
J
vJ
furthermore, K can then be defined as K = aJ but only when given their values at equilibrium (eq.)
J eq.
K (dimensionless) in terms of stoichiometric activities is called a thermodynamic equilibrium constant.
Using Activity to Calculate
State Properties
Equilibrium Constants, K
List of activities for various phases
K is independent of…
Catalysts: may change the rate, but not the energy of the products vs. reactants
Pressure via Inert Gas: relative compositions constant (total composition changes w.r.t. to total moles of gas)
Pressure via Compression (ΔV~1/Δp): K is independent of pressure but compositions may change significantly
K is dependent on…
Temperature: affects standard Gibbs Energy, and T appears in the equation below → must affect ln K and thus K
− r G = + RT ln K where K = aJ v J
J eq.
In the case of a compression, the change in volume does effect each partial pressure
( pB )
2
ptot must follow the gas law pV=nRT
A 2B K=
( pA ) p o (before reaching a new equilibrium)
• As the reaction progresses ni is converted to neq which is the endpoint of the reaction
• it is often useful to consider αn = ni-neq = Δn the amount by which the reactant changes during the
reaction (neq ≡ nf) and is closely related to Δξ
this is very useful for examining reactions where the final ratio of products to reactants is non-trivial
in particular for reactions where the flask is loaded with reactants, and a noticeable amount of reactant
will be consumed while products form (constrained by the initial mass and the equilibrium constant)
van ‘t Hoff Equation: ΔrH from ln K
State Properties
Measuring Standard Reaction Enthalpy
Use data below to show the temperature variation of the equilibrium constant of the reaction
so plot data as (–ln K) vs (1/T) and then get linear fit y=mx+b and m=∆H/R
Measuring ln K from ΔrH
State Properties
Value of K at different temperatures
1 r H
d ln K = 2
dT
R T
K2 T2
1 r H 1 T 2 r H r H T2
dT r H 1 1
K 1 d ln K = T1 R T 2 dT = R T1 T 2 dT R T 1 T 2 = R − − −
T2 T1
r H 1 1 r H 1 1
ln K 2 − ln K1 = − − ln K 2 = ln K1 − −
R T2 T1 R T2 T1
Summary of CHEM2410 to this point
The master equation that combines the 0th, 1st, 2nd, 3rd Laws of Thermodynamics and
chemical equilibrium constant:
where
0th Law of thermodynamics: T Chapt. 1
1st Law of thermodynamics: H (U) Chapt. 2
2nd & 3rd Laws of Thermodynamics: S Chapt. 3
Combine 0th, 1st, 2nd, and 3rd Laws : G (A) Chapt. 4-5
Chemical equilibrium constant : K (rGꝋ) Chapt. 6
Basic Concepts for Batteries
State Properties
Now we have all the parts in place to evaluate cell potentials under standard conditions…
New Slide
of Electrochemical Energy
(Cell Potential + Nernst Equations)
Electrical Work
Maximum non-expansion work (additional or extra) a system can do is ∆G = wadd, max which is the amount of
additional work done during a reversible process
change in total charges
In electrochemistry wadd=we and in physics dwe = dQ
electrostatic potential
dG
rG = so r G d = dG since dG = dwe and then r G d = dwe
d p ,T
if the reaction advances by dξ then the number of electrons moving is v∙dξ and the total charge is -veNA∙dξ
which if F=eNA then this becomes -vF∙dξ where F=96485 C/mol
r G = −vFEcell
r G = −vFEcell and r G = r G + RT ln Q
r G RT
−vFE = r G + RT ln Q E=− − ln Q
vF vF
G
if we define a standard cell potential E = − r related to standard reaction Gibbs, ∆rG
vF
RT
E = E − ln Q Nernst Equation: relates composition to cell potential (via standard cell potential)
vF
The Nernst Equation: some common examples
Metal/metal ion
Gas/ion
Examples of
common half-cell
potentials and _____
how to use their 1
activities and v to
find E from Eθ Redox
Insoluble salt/metal __
1
Measuring Thermodynamic Variables from
State Properties
If the standard formation energy ∆fG(H+,aq) = 0 then the ∆fG of other ions can be determined
dG dE r S
similarly if = −S then we can derive =
dT p dT vF
dE
From this we can also derive r H = r G + T r S = −vF Ecell − T cell
dT
Thus finding cell potentials gives us information on K, ∆rG, ∆rS, and ∆rH
CHEM 2410: Chapter 12
CHAPTER OUTLINE
Distribution of States
1. The Boltzmann Distribution Derived Functions
2. Molecular Partition Functions 5. Internal Energy and Entropy
6. Thermodynamic Functions
Energy and Entropy in Molecules
3. Molecular Energies
4. Canonical Ensemble
Required Reading:
“Atkin’s Physical Chemistry”, Atkins, de Paula & Steeler 11th Ed. 2018: Chapter 12
Statistical Thermodynamics (Ch.12)
State Properties
Distribution of States
Although we can measure the system energy, E, we cannot say for certain how it is arranged
Collisions between molecules cause constant changes in the energy for each particular molecule
However we can report the average population of states: for a large system this average is very consistent
even as individual molecules are themselves changing their energy
We make an assumption, the principle of a priori probability: all possibilities of equal energy are equally
likely to be populated…that is the nature of the state (vib, rot, trans etc.) is irrelevant in this assumption
Configurations and Probability Weighting
If all configurations with are equally probable (later we will constrain them to E), then the probability is the
number of configurations of that type divided by the total number of configurations
For a system with energy E, the system consists of states ε0, ε1, ε2… ε∞ where ε0=0 is the lowest energy state (i.e.
the ground state)
If we add up all the molecules and their energy states we get E the total energy…however since we set ε0=0 we
may need to add a constant U(0) to get U the internal energy (e.g. zero point energy for oscillators)
For a system with N molecules: N0 will be in state ε0 ; N1 will be in state ε1 ; N2 will be in state ε2 ;….etc.
N = Ni E = i Ni
i i
N is conserved E is conserved
State Configurations
x! is “x factorial” where
We can write configurations as populations in each state: x! = 1×2×3×4….(x-2) ×(n-1) ×n
population of state #3
{N, 0, 0, 0,….} all ground state with ε0
{N-1,1,0,0,….} one in state 1, the rest in the ground state (state 0)
{N-2, 1, 1, 0, 0, ….} one in state 2, one in state 1, the rest in the ground state….etc…
There are more ways to create the latter configurations (i.e. by swapping molecules between the ground state)
{N, 0, 0, 0,….} {N-1, 1, 0, 0,….} {N-2, 1, 1, 0,….} {N0, N1, N2, N3,….}
N!
1 way to obtain it N ways to obtain it ½N(N-1) ways to obtain it N 0 ! N1 ! N 2 ! N 3 !... ways to obtain it
N!
In the last example, the general case, we can define W as the weight W=
(# distinguishable permutations) N 0 ! N1 ! N 2 ! N 3 !...
Molecular Partition Functions
N i e − i N i gi e − i
= e − i
=q = i =q
g e − i
N q i N q i
the Boltzmann Equation for states gives the probability of a particular state i, which has an energy of εi and a
partition function which is a sum of the other states I
the Boltzmann Equation for energy levels gives the probability of an energy level, which is the state probability
times a simple degeneracy factor gi (since P ~ ε/T only) and partition function, a sum over all energy levels i
Types of Partition Functions
State Properties
Types of Partition Functions
Partition functions vary only in expressions for εi, which is determined by the energy levels for a particular type of
state (translational, rotational, vibrational, electronic, spin, etc.)
Example: if the energy levels are evenly spaced then for state n, εn = nε = [0, ε, 2ε, 3ε, 4ε…] with no degeneracy
q = gi e − i = e − ( 0) + e − ( ) + e − ( 2 ) + e − (3 ) + e − ( 4 ) + ....
i
+ (e ) + (e ) + (e )
− − 2 − 3 − 4
q = 1+ e + ....
1 1
1 + x + x 2 + x 3 + x 4 + .... = for x 1 q
1− x 1 − e −
an infinite ladder
Fractional population in a 2-state system
2-state system: a system consisting of 2 states, state 0 with ε0=0, and state 1 with ε1=ε
q = e − i = e − 0 + e − 1 = 1 + e −
i
1
1 + e − for i = 0
N i e − i e − i
= = =
N q 1 + e − e −
for i = 1
1 + e −
Translational Contribution: x,y,z
The same analysis holds for all directions (x,y,z) such that: n ,n
x y , nz
= nx + n y + nz
And then (as for any independent energy states we can separate summands):
( )= − − −
− nx ,n y ,nz − TX + YT + ZT − − −
q = = =
T T T T T T
T
e e e e e e e e
X Y Z X Y Z
nx , n y , nz X ,Y , Z X Y Z X Y Z
volume, V=XYZ
X Y Z V h
so that qT = = 3 = β=1/kT
( 2 mkT )
12
k= 1.380×10−23 J∙K-1
Boltzmann constant
If q is the total number of accessible states, then qT/N is the number of translational states per molecule
So for a large value, V/NΛ >> 1.
Since V/N is the volume per molecule, and d=(V/N)1/3 is avg distance between molecules
so d>> Λ for this to be true…
The distance between molecules must be much larger than the thermal wavelength for the integration to be valid
The Rotational Contribution: linear rotor
As we saw earlier, the rotational energy levels for a linear rotor are: J = hcBJ ( J + 1) for integers J 0
speed of light in a vac. rotational constant (in cm-1)
c= 299 792 458 m / s “characteristic frequency”
where no adjustment is needed since E(0)=0, and the degeneracy is g J = ( 2 J + 1)
this means that q = gJ e
R
J
− J
= ( 2 J + 1) e − hcBJ ( J +1) (for linear molecules except A2 symmetrical)
J =0 J =0
discussed later
for many molecules, if kT is larger than the separation between energy levels then:
qJR = ( 2 J + 1) e − hcBJ ( J +1) ( 2 J + 1) e − hcBJ ( J +1) dJ
J =0 0
12
1 kT
32 12
qR = q =
hcB hc A
R
2
hc AB
constants Rotational constant terms
In the case of a non-symmetric top (with frequencies A,B,C), similar arguments lead to
kT
32 12
q =
R
hc ABC
Rotational constant terms (one for each of three unequal rotational axes)
Characteristic Rotational Temperature
In the derivation, the major assumption (integration instead of summation) is valid when changes for e-βε from J
to J+1 are small (e.g. lots of populated states), so as to be nearly continuous
Determining when that is valid, requires introducing a characteristic rotational temperature: at a “high”
temperature T>>θR these (continuous) equations are valid
One further complication to rotational levels, is that rotations which produce indistinguishable states (in
highly symmetric molecules) are forbidden:
An example is 1H2 where the nuclear spin states are either paired or parallel….
in this case J must be even (para) or odd (ortho) for the total molecular wavefunction to be non-zero
There are four possible spin combinations, one para and three ortho, so the average (per molecule) partition
function will be
1 kT
32 12
qR = non-linear
hc ABC
1 3
q R
J , avg = ( 2 J + 1) e − hcBJ ( J +1) + ( 2 J + 1) e − hcBJ ( J +1)
4 J =even 4 J =odd 1 kT
qJR = linear
which approximates to (since term under summation is identical) under high T hcB
1
qJR,avg =
2 J = all
( 2 J + 1) e − hcBJ ( J +1)
so the term ½=1/σ where σ=2 is the symmetry number
for a symmetric diatomic (linear) molecule
Vibrational Contribution to Partition Function
For a harmonic oscillator, any given normal mode will have an infinite ladder of states so the partition function
is the same as that given earlier:
1
q = e − v
V
1 − e − hcv
v
v =0
for a polyatomic molecule, there are 3N-6 normal modes (3N-5 for linear molecule) all of which are treated
independently such that:
qvV1 ,v2 ,v3 ... = qvV1 qvV2 qvV3 qvV4 .... By the same “separation” arguments made previously
Characteristic Vibrational Temperture
Many common modes have a high vibrational frequency v, such that hcv 1 at low temperatures
Some modes may have a low vibrational frequency such that hcv 1 , and in this case further
approximations can be made since e-x=1+x+x2/2+x3/3… for x <<1 (also vibrational state is active at r.t.)
1
q = e − v
V
q V 1
1
1 − e − hcv 1 − (1 − hcv...) hcv
v v
v =0
kT
so for low frequency modes: qvV
hcv
hcv
so the vibrational characteristic temperature V
V
such that approximations are valid for T
k
Electronic Contribution to Partition Function
Electronic energy differences (ΔE) from the ground state are very large so for most cases qE=1
Exceptions occur for degenerate ground states, where qE≈gE×1= gE such as alkali metals which have spin
degeneracy of 2 and thus qE=2
Another example is when atoms/molecules have degenerate ground and low lying degenerate excited states
E.g. NO is a molecule with degenerate ground states 1 2 (unpaired spin) and low energy degenerate excited
2
states the 2 3 2 (also unpaired spin) at an energy ∆E=ε above the ground state (εgnd=0 so εexc= ε)
q E = e − i = 2 + 2e −
E
Average Energy
State Properties
If the total energy is E, we can then describe the average energy per molecule (average energy state) as E/N
E 1 N N i e − i
E = i Ni = = i Ni = i i where =
via the
i N N i N N q Boltmann Eq.
i
(if εgs=0)
i e − 1
= = i e − i
i
so then
q q i at constant volume, e.g. no change in state energies with change in T
i
if we notice that i e − = −i
d
(
d − i
e ) then we can make a substitution Average energy ONLY in
terms of partition function
1
= −
d e − i
( ) = − 1 d 1 dq d ln q d ln q
q i d q d
e − i
=−
q d
=−
d
= gs −
i d V
constant for assumption about εgs
Ensembles
State Properties
Canonical Ensembles
The microcanonical system (1 member, N molecules) is the focus of our study to this point (isolated, same V,E,N)
It is important to remember that ensembles are imaginary replications of the system, each version may be slightly
different with variations based on probability of configurations (the number of members can be infinite)
So to recap:
Ñ is the number of members (replicas of system, not N),
Ẽ is total energy of all members,
Ŵ is the weight of the configuration (Ñ1, Ñ2, Ñ3….) with (E1, E2, E3….)
a canonical ensemble
by the same arguments that gave us the weight of configurations for a single system of states
N! N i e − Ei
where Q = e − Ei
Q is the canonical
W= =
N 0 ! N1 ! N 2 ! N 3 !.... N Q partition function
i
Energy Density of States
The canonical distribution function gives the probability of occurrence of members in a single state i of energy Ei
In a large ensemble it is likely that there are many members with energy near (but not quite equal to) Ei, we can
define the energy density of states as the number of states in an energy range divided by the width of the range
dΩ(E)/dE
= #/J or #/cm-3/J
Working with the Canonical Ensemble
State Properties
Independent Molecules Revisited
In any system, the total energy of N molecules in state i will be Ei=εi(1)+εi(2)+εi(3)… where εi(1) is the energy of
molecule #1 in configuration state i, etc.
Q = e − Ei = e
− i (1) + i ( 2 ) + i ( 3).... i ( N )
The canonical partition function is then
i i
Rather than cycle through each configuration state i, for independent molecules we can rewrite the index as
cycling through all the states of each molecule (i.e. each molecule’s partition function independently evaluated)
− J − J − J − J
Q = e = ( ) q = e − i
N
e e .... e q remember that
j 1 j 2 j 3 j N i
this is true as long as each molecule is distinguishable from the others so that Q = qN
qN
if they are indistinguishable, then Q must be divided by indistinguishable permutations (there are N!) so Q =
N!
same molecule/atom, able to switch positions at random, like in a gas
Partition Functions and Cases
Partition functions thus far:
qN
Molecules are indistinguishable and independent Q=
e.g. perfect gas N!
Molecules are interacting (indist. or dist.) use Q (but we did not derive any expression)
e.g. imperfect gas, liquids, some solids
Mean Energy of a Canonical Ensemble
q assumed that molecules were independent, non-interacting and isolated (i.e. N single molecular partition functions)
Q accepts they may be interacting (and is thus more likely to be true for real gases and condensed phases)
Total energy is Ẽ and Ñ is the number of members, so average energy per system is <E>= Ẽ/ Ñ
NNi i ee−−EEi i
Pi = == Q == e− Ei
where Q
where
NN QQ ii
so it follows that the average can be calculated in the same manner as <ε> (average energy per molecule)
probability of the system having energy Ei this is the average energy of a system at
thermal equilibrium with the surroundings
HW Questions
State Properties