We are going to calculate the total entropy of a gas
Assume: constant energy and volume
Assume: gas is dilute – ignore intermolecular forces
Our gas molecules will be able to rotate and vibrate
We will skip the derivation however and show the results
Ensemble Size of the 3D Box System
Using geometry and calculus we can arrive at an expression of the ensemble size of gas in a 3D box
$$ \Omega = \frac{1}{N!\left[\left(\frac{3N}{2}\right)-1\right]!}\left(\frac{\pi}{4}\right)\left(\frac{2\pi mEV^\bfrac{2}{3}}{h^2}\right)^{\left(\frac{3N}{2}\right)-1} $$
This means that the size of the ensemble becomes huge as \(m\), \(E\), and \(V\) approach bulk.
Example Ensemble Sizes
Entropy
We can use Boltzmann’s equation to determine the entropy based on our ensemble size
$$ \begin{align}
S &= k_\mathrm{B}\ln \Omega \\
&= k_\mathrm{B} \ln \left[ \frac{1}{N!\left[\frac{3N}{2}-1\right]!} \frac{\pi}{4} \left( \frac{2\pi mEV^\bfrac{2}{3}}{h^2} \right)^{\frac{3N}{2}-1} \right] \\
&= k_\mathrm{B}\left[ \begin{split} & -\ln N! - \ln \left(\frac{3N}{2}-1\right)! + \ln \left(\frac{\pi}{4}\right) \\ & + \left(\frac{3N}{2}-1\right) \ln \left(\frac{2\pi mEV^\bfrac{2}{3}}{h^2}\right) \end{split} \right]
\end{align} $$
Using Stirling's approximation
$$ S = k_\mathrm{B}\left\{ \begin{split} & -N\ln N + N - \left(\frac{3N}{2}-1\right) \ln \left(\frac{3N}{2}-1\right) + \left(\frac{3N}{2}-1\right) \\ & + \ln \left(\frac{\pi}{4}\right) + \left(\frac{3N}{2}-1\right) \ln \left(\frac{2\pi mEV^\bfrac{2}{3}}{h^2}\right) \end{split} \right\} $$
Simplifying with a fixed, large number of moles
$$ \left(\frac{\partial \Omega}{\partial V}\right)_{E,N} \approx \frac{1}{N!\left[\left(\frac{3N}{2}\right)-1\right]!} \left(\frac{2\pi mE}{4\pi^2\hbar^2}\right)^{\frac{3N}{2}-1} \frac{\pi}{4V^\bfrac{2}{3}}NV^{N-1} $$
Increasing Volume
What happens when our volume increases?
If our volume increases by \(\frac{dV}{V}=10^{-10}\)
$$ \frac{d\Omega}{\Omega} = N\frac{dV}{V} = \left(6.02\times 10^{23}\right)\left(10^{-10}\right)=6.02\times 10^{13} $$
So, the chance our system will be found in a larger volume than its initial volume is immensely greater
The number of microstates has increased immensely
Meaning
Say that the sample changes quantum states each time two molecules collide, \(\sim 10^{30}\) times per sec.
The probability at each of these collisions that the system will remain at the original volume is one in \(6.02\times 10^{13}\).
This means that the sample will expand immediately
When possible, systems always move to a greater number of quantum states (greater entropy)
Second Law of Thermodynamics
The second law of thermodynamics can be stated as:
For any process in a system totally isolated from its surroundings, the macroscopic entropy measured over the entire system \(S_T\) can never decrease: \(\Delta S_T \ge 0\)
Qualifiers
The entropy can decrease in a process, but only in a local region or (for a microscopic system) over a short time
If we allow the system to communicate with its surroundings, the entropy change in the system could decrease, but it would have to be matched by an increase in the entropy of the surroundings
Second Law and Temperature
Let’s look at the system in thermal contact with a reservoir (from chapter 2)
System had fixed volume and number of particles
It could exchange energy with the surrounding reservoir as long and \(E_T\) was conserved.
Let’s look at the effects of the second law on this system.
As energy is transferred into the system from the reservoir
Energy of the system rises
\(\Omega(E)\) rises (this is the system)
Energy of the reservoir falls
\(\Omega_r(E_r)\) falls (this is the reservoir)
As long as there are no additional interactions then \(\Omega_T(E_T)=\Omega(E)\Omega_r(E_r)\)
The Take-Away
Based on that equation we know that if either ensemble size gets too small then the total ensemble size will decrease
This means that the configuration with the most available microstates is somewhere in the middle
System and Reservoir Microstates
Finding the Largest Ensemble State
We can find this maximum total by first keeping V and n constant for both the reservoir and the system and setting derivative equal to zero
$$ \begin{align}
\left. \left(\frac{\partial \ln \Omega_T}{\partial E}\right)_{V,n}\right|_\chem{max} &= 0 = \left. \left(\frac{\partial \ln \left(\Omega \Omega_r\right)}{\partial E}\right)_{V,n} \right|_\chem{max} \\
&= \left. \left(\frac{\partial \left(\ln \Omega+\ln \Omega_r\right)}{\partial E}\right)_{V,n} \right|_\chem{max} \\
&= \left. \left(\frac{\partial \ln \Omega}{\partial E}\right)_{V,n} \right|_\chem{max} + \left. \left(\frac{\partial \ln \Omega_r}{\partial E}\right)_{V,n} \right|_\chem{max} \\
\end{align} $$
Because \(E+E_r=E_T\) is constant then \(dE=-dE_r\)
$$ 0 = \left. \left(\frac{\partial \ln \Omega}{\partial E}\right)_{V,n} \right|_\chem{max} - \left. \left(\frac{\partial \ln \Omega_r}{\partial E_r}\right)_{V,n} \right|_\chem{max} $$
For an expansion \(V_f \gt V_i\) therefore \(\Delta S\) is positive
\(\Delta S\) is negative for compression
Notes
It is important to remember that energy, entropy, temperature, volume, pressure, and number of moles are all state function
This means is does not matter how we go between initial and final states
It also means that \(\Delta E\), \(\Delta S\), \(\Delta T\), \(\Delta V\), \(\Delta P\), and \(\Delta n\) also do not depend on the path taken
\(\Delta S_T\) does depend on path
Why Does \(\Delta S_T\) Depend on Path?
Consider \(\Delta E_\chem{rev}\) for a reversible process and \(\Delta E_\chem{irr}\) for an irreversible process
$$ q_\chem{rev} + w_\chem{rev} = \Delta E_\chem{rev} = \Delta E_\chem{irr} = q_\chem{irr}+w_\chem{irr} $$
As we have shown previously for the isothermal expansion, the reversible process is the one that did the most work, so for any process
$$ 0 \gt w_\chem{irr} \gt w_\chem{rev},\,q_\chem{irr}\lt q_\chem{irr} $$
For the reversible process \(\dbar q_\chem{rev} = T\, dS\)
This gives the Clausius inequality
$$ \dbar q \le T\,dS $$
The equality only holds for the reversible process
Thermodynamic Potentials
The thermodynamic potentials can also give the direction of spontaneous process on certain conditions
We rarely measure entropy directly
We don’t often want to hold \(E\) and \(V\) constant
If we keep \(T\) and \(P\) constant, the Gibbs energy is useful
Gibbs Energy
From our Gibbs energy definition in chapter 7
$$ \begin{align}
G &= E-TS+PV \\
dE &= \dbar q + \dbar w \\
\left(dG\right)_{T,P} &= \left( dE-T\,dS-S\,dT+P\,dV + V\,dP \right)_{T,P} \\
&= \dbar q+\dbar w - T\,dS + P\,dV
\end{align} $$
So what?
From the Clausius inequality we know \(\dbar q \le T\,dS\)
We know the maximum work is done in a reversible process so \(\dbar w = \dbar w_\chem{rev} \ge -P\,dV\)
So, if we only consider \(PV\) work then \(\left(dG\right)_{T,P} \le 0\)
Again, the equality hold only for reversible processes
\(A\) process is spontaneous under constant \(T\) and \(P\) that will minimize \(G\)
\(G\) is at its minimum value at equilibrium
In Benchtop Chemistry
The direction in which the reaction will move spontaneously is the one that leads to a negative \((\Delta G)_T\), where
$$ \begin{align}
(\Delta G)_T &= \Delta (H-TS)_T \\
&= (\Delta H)_T - (T\Delta S + S\Delta T)_T \\
&= (\Delta H)_T - T(\Delta S)_T
\end{align} $$
If temperature is understood to be fixed then
$$ \Delta G = \Delta H - T\Delta S $$
The change in enthalpy is equal to the heat flowing into the surrounding bath which can be related to \(\Delta S_r\)
$$ \Delta H = q = -q_r = -T\Delta S_r $$
So using this in our equations we have
$$ (\Delta G)_{T,P} = -T\Delta S_r -T\Delta S=-T(\Delta S_r+\Delta S) $$
Ideal Mixing
The Setup
We have two containers each holding an ideal gas at the same temperature and pressure
Container \(\chem{A}\) holds \(n_\chem{A}\) moles of gas
Container \(\chem{B}\) holds \(n_\chem{B}\) moles of gas
The containers are connected through a tube with a valve
What happens to the entropy and other parameters when the value is opened?
Intuition
Intuitively we know that this mixing will be spontaneous
This means that the entropy must increase
How can this be true when the density of gas throughout the containers is never changing.
We know that expanding each gas individually increases its entropy
$$ \begin{align}
dS &= \frac{P}{T}dV=\frac{nR}{V}dV \\
\Delta S &= nR\ln \left(\frac{V_f}{V_i}\right) \\
\Delta S_\chem{A} &= n_\chem{A}R\ln \left(\frac{V_\chem{A}+V_\chem{B}}{V_\chem{A}}\right) \\
\Delta S_\chem{B} &= n_\chem{B}R\ln \left(\frac{V_\chem{A}+V_\chem{B}}{V_\chem{B}}\right)
\end{align} $$
Volume to Moles
Because we are at constant temperature and pressure, we can rewrite our equation in terms of mole fractions
$$ \begin{align}
\frac{V_\chem{A}}{V_\chem{A}+V_\chem{B}} &= \frac{n_\chem{A}}{n_\chem{A}+n_\chem{B}} = X_\chem{A} \\
\frac{V_\chem{B}}{V_\chem{A}+V_\chem{B}} &= \frac{n_\chem{B}}{n_\chem{A}+n_\chem{B}} = X_\chem{B}
\end{align} $$
Note that the first term in each are identical so they will cancel out when we take the difference
$$ \begin{align}
\Delta S_\chem{mix} =& S_f - S_i \\
=& (N_\chem{A}+N_\chem{B})k_\mathrm{B} \ln (V_\chem{A}+V_\chem{B}) - \left[ (N_\chem{A}+N_\chem{B})k_\mathrm{B} \ln (V_\chem{A}+V_\chem{B}) \right] \\
& \vdots \\
=& -R\left(n_\chem{A}\ln X_\chem{A}+n_\chem{B}\ln X_\chem{B}\right)
\end{align} $$
Result
So we got the same result that we get classically.
Now we have shown that as long as the A-B interaction is essentially the same as the A-A and B-B interactions it doesn’t matter if we are dealing with gases or liquids, monatomics or polyatomics.
The number of distinct translational states increases by a factor of \( \frac{\left(V_\chem{A}+V_\chem{B}\right)^{N_\chem{A}+N_\chem{B}}}{V_\chem{A}^{N_\chem{A}}V_\chem{B}^{N_\chem{B}}} \)
Ultimate Results
Even with no work done and no net heat flow, system will evolve so as to find more microstates of the ensemble
The driving force behind mixing has nothing to do with the mixing itself
Each substance is taking advantage of the opportunity to distribute its particles over a greater volume thereby increasing its own entropy.
We can calculate the enthalpy change of mixing the same way (remember that \(T\) is constant)
$$ \begin{align}
\Delta G &= \Delta H - T\Delta S\\
\Delta H_\chem{mix} &= \Delta_\chem{mix}G + T\Delta_\chem{mix}S \\
&= \left\{ \begin{split} & RT\left(n_\chem{A}\ln X_\chem{A}+n_\chem{B}\ln X_\chem{B}\right) \\ & + T\left[-R\left(n_\chem{A}\ln X_\chem{A}+n_\chem{B}\ln X_\chem{B}\right)\right] \end{split} \right\} \\
&= 0
\end{align} $$
So, because the process takes place at constant pressure, \(\Delta H=q\), and no heat is flowing, so
$$ \Delta H =0 $$
The Third Law of Thermodynamics
The Third Law of Thermodynamics
Up to this point we have basically limited ourselves to calculating changes in entropy
We need a reference value for entropy
This reference is given in the third law of thermodynamics
The entropy of a substance approaches a minimum as \(T\) approaches zero
We also \(S=0\) at \(T=0\)
Problem
There is an experimental problem with this law
At absolute zero, the most stable form of a substance is a crystal that puts each particle in the most stable, reproducible place.
At very cold temps, however, the atoms don’t have enough energy to cross the energy barrier separating crystalline phases
Crystal domains form - local ordered regions
Meaning
Cooling the substance traps it in any of several low-energy state
Not necessarily the lowest energy state
This means there is residual entropy that cannot be removed from the system
This means there are still several ensemble states possible
Thus \(S=k_\mathrm{B} \ln \Omega \ne 0\), but it is close to zero
Getting to Absolute Zero
It is thermodynamically impossible to reach absolute zero
Near absolute zero it also because quantum mechanically impossible to reach absolute zero
Getting there would involve remove all molecular motion
We can get to less that \(1\,\mathrm{K}\) using boiling liquid helium as a coolant
The Curie Temperature
Magnetic material have a temperature at with the thermal motion of the atom (called phonons) exceed the force keeping the spins aligned (magetism).
The temperature this occurs at is called the Curie temperature
Iron has a Curie temperature of \(1043\,\mathrm{K}\)
Many substances become magnetic at some low temperature
Adiabatic Demagnetization; Step 1
Liquid helium is used to cool a crystal to near \(1\,\mathrm{K}\) as it is magnetized by a high-field external solenoid magnet.
Adiabatic Demagnetization; Step 2
The helium is replaced by vacuum once the crystal is fully magnetized.
Adiabatic Demagnetization; Step 3
The external field is turned off.
Adiabatic Demagnetization; Step 4
The spins in the crystal equilibrate with the phonons.
The net angular momentum of the spins is conserved by a current induced in the solenoid, which does work and further lowers the temperature of the crystal.
Final temperatures below \(10^{-6}\,\mathrm{K}\) can be achieved using this technique.
Adiabatic Demagnetization: Effect on Temperature
The Debye Theory
At constant pressure
$$ S=\int_0^T \frac{C_P(T')}{T'}dT' $$
The Debye theory for heat capacities predicts a low temperature heat capacity
$$ C_V \approx \frac{12\pi^4 Nk_\mathrm{B}^4 T^3}{5\omega_\mathrm{D}^3} $$
The Debye Theory Continued
For crystals, the heat capacity at constant pressure is nearly equal to \(C_V\)