Indroduction to Statistical Mechanics: Building Up to the Bulk
Shaun Williams, PhD
Atomic Structure
Atomic Structure
A concept key to interpreting the second law of thermodynamics is the limitation by quantum mechanics on the number of states of our system
The number of quantum states of any real system is a finite number
This is an important difference between our classical understanding and our modern understanding
Quantization of Properties
At molecular sizes, particles behave like waves
To fulfill the mathematical rules of waves, parameters, such as energy and momentum, are only allowed certain quantized values
For a moving particle particle, this quantization becomes measurable when the distance traveled is much greater than its de Broglie wavelength
$$ \lambda_{dB} = \frac{h}{mv} $$
In this equation, called the de Broglie wavelength equation
\(h\) is Planck’s constant, \( 6.626 \times 10^{-34}\,\mathrm{J}\cdot \mathrm{s}\)
\(m\) is the particles mass
\(v\) is the particles velocity
Quantizer Energy
In a classical system, energy is a continuous function
In a quantum system, energy comes only in discrete values
Problem with the quantum nature
When dealing with waves, the particles/waves cannot have their position pinned down to a single location
For any parameter, \(A\left(\vec{r}\right)\), the average value of this parameter is given by the average value theorem
$$ \expect{A} = \int_{\text{all space}} \mathcal{P}_\vec{r}\left(\vec{r}\right) A\left(\vec{r}\right) d\vec{r} $$
Where \(\mathcal{P}_\vec{r}\left(\vec{r}\right)\) is the probability distribution
Particle in a Three-Dimensional Box
Consider a 3-D box with lengths of \(a\), \(b\), and \(c\) along the \(x\), \(y\), and \(z\) axes, respectively
Possible energy values for a particle in the box is
$$ \begin{align}
\varepsilon_{n_x,n_y,n_z} &= \frac{\pi^2 \hbar^2}{2m}\left( \frac{n_x^2}{a^2}+\frac{n_y^2}{b^2}+\frac{n_z^2}{c^2} \right) \\
&\approx \frac{\pi^2\hbar^2}{2m(abc)^\bfrac{2}{3}}\left( n_x^2+n_y^2+n_z^2 \right) = \varepsilon_0n^2
\end{align} $$
Energy Levels of a Three-Dimensional Box
States of Such a Particle
A particle which has the lowest values of \(n_x\), \(n_y\) and \(n_z\) is said to be in its ground state
Any other state is said to an excited state
When two or more quantum states share the same energy, we say they are degenerate
(2,1,1), (1,2,1), and (1,1,2)
Correspondence principle – quantum mechanics approaches classical mechanics as we increase the particle’s kinetic energy, mass, or domain
Use of Quantum Mechanics
Due to correspondence principle:
Use QM when \(\lambda_{dB}\) is comparable to the domain size
Use classical when \(\lambda_{dB}\) is much smaller
Entropy allows us to indirectly measure the density of quantum states, \(W(\varepsilon)\)
The density of states tells how many quantum states near energy \(\varepsilon\) there are per unit energy
Quantum States of Atoms
The same analysis can be used on one-electron atoms yielding
$$ E_n = -\frac{Z^2m_ee^4}{2\left(4\pi \epsilon_0\right)^2n^2\hbar^2} \equiv -\frac{Z^2}{2n^2}E_h $$
\(E_h\) is the energy unit Hartree
Hartrees are convenient units for calculations involving electrons in atoms
QM accurately predicts the properties of individual molecules
QM will work as the system becomes bigger
It is no longer practical to analyze a mole of particles with QM, simply to complicated
With so many molecules, we typically don’t care how a single one of them behaves
Macroscopic Properties
We seek macroscopic properties of the system
Two group of macroscopic properties:
Extensive – obtained by summing together contributions from all the molecules in the system
Intensive – obtained by averaging the contributions from all the molecules in the system
Parameters
Extensive Parameters
\(N\), the total number of molecules in the system or the total number of moles, \(n\)
\(V\), the total space occupied
\(E\), the sum of the translational, rotational, vibrational, and electron energies
Intensive Parameters
\(P\), the pressure, which is the average force per unit area exerted by the molecules on their surroundings
\(\rho\), the number density, which is the average number of molecules per unit volume
Related to mass density
$$ \rho_m=\rho\frac{\mathcal{M}}{\mathcal{N}_A} $$
Statistical Mechanics
Statistical mechanics used probability theory to describe the macroscopic groups of molecules
Based on common characteristics of individual molecules
Statistical mechanics extends the results of QM to the behavior of bulk quantities of substances
The things we want to study are perfect statistical samples
Common Assumptions in Statistical Mechanics
Chemically identical molecules share the same physics
Macroscopic variables are continuous variables
Measured properties reflect the ensemble average
If we fix some values (e.g. volume and moles)
Ensemble - Set of all quantum states, same values
Microstate – a unique quantum state
\(\Omega\) – number of microstates in the ensemble
Ergodic Hypothesis
If we fix \(E\), \(N\), and \(V\)
Let the system find its own pressure
The result will be the average of \(P\) over all microstates
The ensemble average
12-Member Ensemble
Here are 12 distinct microstates
All have the same properties
Let the particle exert a pressure of p0 on the walls next to it
Half the top walls have two particles next to it and half have one
\(\expect{p_{top}} = 1.5p_0\)
The Microcanonical Ensemble
Approaching a problem, we must decide which parameters are allowed to vary
More flexible is more realistic but more difficult to solve
If we fix all the extensive variables, \(E\), \(V\), and \(N\) we have our first ensemble – the microcanonical ensemble
Properties of Microcanonical Ensemble
If we fix all extensive variables then intensive variables that are ratios of extensive variables must also be fixed
Density, \(\rho=\frac{N}{V}\)
What can change?
Microscopic properties that we’re not bothering to measure
Some intensive properties (such as pressure) that are not ratios of extensive variables
Advantages of Microcanonical Ensemble
Energy and mass are rigorously conserved
There system must be completely isolated from the rest of the universe
Constant volume means that no mechanical force pushes on the surroundings
Isolation greatly simplifies the system
Entropy
Boltzmann Entropy
The Boltzmann entropy is given by
$$ S\equiv k_\mathrm{B}\ln \Omega $$
Boltzmann constant is \(k_\mathrm{B}=1.381\times 10^{-23}\,\mathrm{J}\,\mathrm{K}^{-1}\)
This provides a different definition of the relation between heat and temperature
This will be our rigorous definition of entropy since it works under and circumstances
The entropy counts the total number of distinct microstates of the system
Entropy and Microstates
\(\Omega\) versus \(g\)
For a microcanonical ensemble
\(\Omega\) is the total number of microstates that have the same energy
Under these conditions \(\Omega\) is the same as \(g\)
\(\Omega\) is the total number of microstates in an ensemble
\(g\) is the degeneracy of quantum states in some particular individual particle
Size of \(\Omega\)
\(\Omega\) is enormous
However, \(\Omega\) is always finite
The number of arrangements of system is countable
In a classical system, \(\Omega\) would be infinite due to the continuous nature of classical mechanics
Quantization makes \(\Omega\) finite
Two Non-Interacting Subsystems
The degeneracy is the product of degeneracies
The entropy is the sum of the entropies
$$ \begin{align}
\Omega_{\chem{A+B}} &= \Omega_\chem{A}\Omega_\chem{B} \\
\ln \Omega_{\chem{A+B}} &= \ln \left(\Omega_\chem{A}\Omega_\chem{B}\right) = \ln \Omega_\chem{A}+\ln \Omega_\chem{B} \\
S_\chem{A+B} &= S_\chem{A}+S_\chem{B}
\end{align} $$
Gibbs Definition of Entropy
Gibbs developed a general equation for estimating the entropy in any ensemble using probability rather than \(\Omega\)
Derivation:
The total number of ways of arranging the labels on \(N\) molecules is \(N!\)
There are a total of \(N_i!\) ways of rearranging the labels
Example of Calculating \(\Omega\)
The Ensemble Size
These definitions mean that we get
$$ \Omega=\frac{N!}{N_1!N_2!N_3!\dots N_k!} = \frac{N!}{\prod_{i=1}^kN_i!} $$
Example: 6 particles with 4 possible states with 3 in state 1, 2 in state 2, 1 in state 3 and 0 in state 4
$$ \Omega = \frac{6!}{(3!)(2!)(1!)(0!)}=60 $$
Using this definition of the size of the ensemble we find
$$ \begin{align}
S &= k_\mathrm{B} \ln \Omega = k_\mathrm{B} \ln \left( \frac{N!}{\prod_{i=1}^kN_i!} \right) \\
&= k_\mathrm{B} \left[ \ln N! - \ln \prod_{i=1}^k N_i! \right]
\end{align} $$
Now we can use Stirling’s approximation: \( \ln N! \approx N\ln N\) to yield
$$ \begin{align}
S &= k_\mathrm{B} \left[ N\ln N - N - \left( \sum_{i=1}^k N_i \ln N_i - \sum_{i=1}^k N_i \right) \right] \\
&= k_\mathrm{B} \left[ \sum_{i=1}^k N_i \ln N - N - \sum_{i=1}^k N_i \ln N_i +N \right] \\
&= k_\mathrm{B} \sum_{i=1}^k \left[ N_i \ln N - N_i \ln N_i \right] \\
&= k_\mathrm{B} \sum_{i=1}^k N_i \left[ \ln N - \ln N_i \right]
\end{align} $$
Rewriting Entropy in Terms of Probability
$$ S = k_\mathrm{B} \sum_{i=1}^k N_i \left[ \ln N - \ln N_i \right] $$
We can write this in terms of probabilities \(\mathcal{P}(i)=\frac{N_i}{N}\)
$$ \begin{align}
S &= Nk_\mathrm{B} \sum_{i=1}^k \frac{N_i}{N} \ln \frac{N}{N_i} \\
&= Nk_\mathrm{B} \sum_{i=1}^k \mathcal{P}(i) \ln \frac{1}{\mathcal{P}(i)} \\
&= -Nk_\mathrm{B} \sum_{i=1}^k \mathcal{P}(i) \ln \mathcal{P}(i)
\end{align} $$
Temperature and the Partition Function
The Ideal Gas
A collection of particles
Don’t interact with each other at all
Elastically bounce off walls
We want to use statistical mechanics to find the pressure exerted by an ideal gas on the walls of the container as a function of the macroscopic variables
The ensemble size of the universe is
$$ \Omega_T\left(E_T\right) = \sum_E \Omega\left(E\right) \Omega_r \left(E_r\right) $$
\(\mathcal{P}(i)\) and Number of Microstates
Choose a single microstate, \(i\), with energy \(E_i\)
If there are 9 distinct microstates with energy the reservoir must be changing
Thus we can determine the system number from the reservoir number (\(E_r=E_T-E\))
The probability of the system being in state \(i\) is
$$ \mathcal{P}(i)=\frac{\Omega_r\left(E_r\right)}{\Omega_T\left(E_T\right)} $$
Ensemble Average
A Problem
The problem is we don’t want to have to measure the reservoir instead of the system
First we will rewrite this in terms of entropy (Taylor series expansion)
$$ S_r \left( E_r \right) = S_r \left( E_T-E_i \right) = S_r \left( E_T \right) - \left. \left( \frac{\partial S_r}{\partial E_r} \right)_{V,N} \right|_{E_T} E_i + \cdots $$
For a fixed total energy we know that \(dE_T = d\left(E_T-E\right)=-dE\)
We also know that \(\Omega_T\left(E_T\right)=\Omega_r\left(E_r\right)\Omega\left(E\right)\)
Continuing to Solve the Problem
We also know that
$$ \begin{align}
dS_r &= k_\mathrm{B} d \ln \Omega_r \left(E_r\right) = k_\mathrm{B} d\left[ \ln \left( \frac{\Omega_T\left(E_T\right)}{\Omega\left(E\right)} \right) \right] \\
&= k_\mathrm{B} d \left[ \ln \Omega_T \left(E_T\right) - \ln \Omega\left(E\right) \right] = -k_\mathrm{B} d\ln \Omega\left(E\right) \\
&= -dS
\end{align} $$
Combining these equations we find that
$$ \left. \left( \frac{\partial S_r\left(E_r\right)}{\partial E_r} \right) \right|_{E_T} = \left. \left( \frac{\partial S(E)}{\partial E} \right) \right|_{E_T} $$
Rewriting the Entropy Equation
Using these previous equation we find
$$ S_r\left(E_r\right) = S_r\left(E_T-E_i\right) = S_r\left(E_T\right) - \left. \left( \frac{\partial S(E)}{\partial E} \right) \right|_{E_T} E_i $$
This introduces temperature
$$ T \equiv \left( \frac{\partial E}{\partial S} \right)_{V,N} $$
Using temperature in the entropy equation
$$ S_r\left(E_r\right) = S_r\left(E_T-E_i\right)=S_r\left(E_T\right)-\frac{E_i}{T} $$
Solving the Issues
We can now go back to \(\Omega\)
$$ k_\mathrm{B} \ln \left[ \Omega_r\left(E_r\right) \right] = k_\mathrm{B} \ln \left[ \Omega_r\left(E_T\right) \right] -\frac{E_i}{T} $$
Solving for the number of reservoir states we get
$$ \Omega_r\left(E_r\right)=\Omega_r\left(E_T\right)e^{-\bfrac{E_i}{k_\mathrm{B}T}} $$
Temperature
A New Ensemble
In typical experiments, the energy is not fixed
The system and the surroundings/reservoir can exchange energy
If the reservoir is big enough, its properties are constant, so we fix \(T\)
So a canonical ensemble has \(T\), \(V\), and \(N\) fixed and \(E\) as a variable
More Equations, Now Within the Canonical Ensemble
Combining some previous equations
$$ \mathcal{P}(i)=\frac{\Omega_r\left(E_T\right)}{\Omega_T\left(E_T\right)} e^{-\bfrac{E_i}{k_\mathrm{B}T}} $$
The \(\Omega\) are constants
We can eliminate them by requiring that the probability be normalized
$$ \sum_{i=1}^\infty \mathrm{P}(i) = \frac{\Omega_r\left(E_T\right)}{\Omega_T\left(E_T\right)} \sum_{i=1}^\infty e^{-\bfrac{E_i}{k_\mathrm{B}T}} = 1 $$
The likelihood of getting the system into a particular state \(i\) drops exponentially with respect to the ratio of the microstate energy \(E_i\) to the thermal energy \(k_\mathrm{B}T\)
The temperature establishes how likely the system is to be found at any particular energy.
Distribution in a 3D Box
Finding Another Likelihood
Let’s find the likelihood of finding the system t any particular energy E
$$ \mathcal{P}(E)=\Omega(E)\mathcal{P}(i) = \frac{\Omega(E) e^{-\bfrac{E_i}{k_\mathrm{B}T}}}{Q(T)} $$
In all these equation the beauty of the canonical ensemble is that is allows us to ignore how \(\Omega\) depends on \(V\) and \(N\)
Assertion
The probability extends to the probability of a particle within our system having some particular energy \(\varepsilon\)
Say we have a particular quantum state \(i\) and degeneracy \(g\)
$$ \begin{align}
\mathcal{P}(\varepsilon) &= \frac{g(E) e^{-\bfrac{\varepsilon}{k_\mathrm{B}T}}}{q(T)} \\
q(T) &= \sum_{\varepsilon=0}^\infty g(E) e^{-\bfrac{\varepsilon}{k_\mathrm{B}T}}
\end{align} $$
Rotational Partition Functions
Check of Zero Energy
If we measure all energy relative to \(E_0\) then
$$ \begin{align}
\mathcal{P}(E) &= \frac{ge^{-\bfrac{\left(E-E_0\right)}{k_\mathrm{B}T}}}{\sum_E ge^{-\bfrac{\left(E-E_0\right)}{k_\mathrm{B}T}}} \\
&= \left( \frac{ge^{-\bfrac{E}{k_\mathrm{B}T}}}{\sum_E ge^{-\bfrac{E}{k_\mathrm{B}T}}} \right) \left( \frac{e^{-\bfrac{E_0}{k_\mathrm{B}T}}}{e^{-\bfrac{E_0}{k_\mathrm{B}T}}} \right) = \frac{ge^{-\bfrac{E}{k_\mathrm{B}T}}}{\sum_E ge^{-\bfrac{E}{k_\mathrm{B}T}}}
\end{align} $$
Example 2.1
At 298 K, calculate the ratio of the number of \(\chem{NH_3}\) molecules in the excited state to the number in the ground state, where the excited state is (a) the \(0^-\) state of the inversion, which lies \(0.79\,\mathrm{cm}^{-1}\) above the \(0^+\) ground state; and (b) the \(1^+\) state, which lies \(932.43\,\mathrm{cm}^{-1}\) above the \(0^+\). (The wavenumber unit, \(\mathrm{cm}^{-1}\), is conventionally used by spectroscopist as an energy unit, based on the relation between the transition energy in the experiment and the reciprocal wavelength of the photon that induces the transition: \(E_{photon}=\frac{hc}{\lambda}\). Because the energy is inversely proportional to the wavelength, energy is given in units of \(\frac{1}{\text{distance}}\).)
Example 2.2
At a temperature of 1000 K, how many vibrational states of \(\chem{H_2}\) are populated by at least 1% of the molecules, given the vibrational constant \(\omega_e=4395\,\mathrm{cm}^{-1}\) and the vibrational energy (relative to the ground state)of approximately \(E_{vib}=\omega_e\nu\)?
Example 2.3
You discover a molecular system having the energy levels and degeneracies
$$ \varepsilon = c\left(n-1\right)^6;\;g=n;\;\text{for }n=1,2,3,\dots\text{ and }k_\mathrm{B}T=400c $$
Evaluate the partition function and calculate \(\mathcal{P}(\varepsilon)\) for each of the four lowest energy levels.
The Ideal Gas Law
The Original Problem - Pressure of an Ideal Gas
Moving A Wall
We momentarily free one wall
It reduces its forces therefore its pressure on the gas, \(P_{min}\)
The gas pushes the wall out infinitesimally, \(ds\)
The energy spent moving the wall is \(dE=-F\,dS = -P_{min} A\,dS = -P_{min} \,dV\)
The Pressure Equation
If we hold \(N\) and \(S\) constant then
$$ P_{min} = -\left( \frac{\partial E}{\partial V}\right)_{S,N} $$
Mostly we will deal with the case where \(P\cong P_{min}\)
$$ \text{if}\,P=P_{min}:P=-\left(\frac{\partial E}{\partial V}\right)_{S,N} $$
Cyclic Rule for Partial Derivatives
The cyclic rule for partial derivatives is
$$ \left( \frac{\partial X}{\partial Y}\right)_Z = -\left( \frac{\partial X}{\partial Z}\right)_Y\left(\frac{\partial Z}{\partial Y}\right)_X $$
This allows us to write
$$ P=-\left(\frac{\partial E}{\partial V}\right)_{S,N} = \left(\frac{\partial E}{\partial S}\right)_{V,N} \left(\frac{\partial S}{\partial V}\right)_{E,N} $$
Another equation for temperature
$$ T=\left(\frac{\partial E}{\partial S}\right)_{V,N} $$
Determining Entropy Change with Volume
Let the translational states of the gas correspond to all the possible ways the gas particles can be arranged inside our box, regardless of what energy the gas has
The box has a volume \(V\)
Break up the volume into units \(V_0\), each able to hold a single particle
We have \(M\) number of units so \(V=MV_0\)
More On Our Gas
The particles are indistinguishable
This reduces the number of distinct states
If we had two particles the number of distinct states would be \(\bfrac{M\left(M-1\right)}{2}\)
Extending this to \(N\) particles
$$ \frac{M(M-1)(M-2)\dots (M-N)}{N!} $$
For a gas \(M \gg N\)
Even More on Our Gas
Introduce a constant \(A\) to absorb any other parameter’s effects, such as energy
$$ \Omega = \lim_{M\gg N} A\frac{1}{N!} M(M-1)(M_2)\dots (M_N)=A\frac{1}{N!}M^N $$
Because \(V=MV_0\)
$$ \Omega\frac{1}{N!}\left(\frac{V}{V_0}\right)^N = A\frac{1}{V_0^NN!} V^N $$