A single thermodynamic state is a specific set of macroscopic parameters
Because a thermodynamic state specifies all the parameters, the parameters are called state functions
Speed is not a state function because that parameter does not describe the entire system
An equation that expresses one state function in terms of the others is an equation of state
Equations of State
Equations like \(P=-\left(\frac{\partial E}{\partial V}\right)_{S,N}\) and \(\left(\frac{\partial E}{\partial S}\right)_{V,N}\equiv T\) give the state function
$$ PV=nRT $$
First Law of Thermodynamics
From our previous discussions we have seen that
$$ \dbar q \le T\,dS $$
The first law of thermodynamics states that
$$ \Delta E \equiv E_f - E_i =q+w $$
If we look at infinitesimal changes in energy
$$ dE=\dbar q + \dbar w $$
Inexact Differential, \(\dbar\)
The inexact differential is used to remind us that \(q\) and \(w\) are not state functions
They are instead defined by the exact process
\(q\) and \(w\) do not have a value at any individual point
To evaluate \(q\) or \(w\), we cannot in principle integrate \(\dbar q\) or \(\dbar w\) directly
We must replace inexact with exact to integrate
Approximations and Assumptions
Equilibrium
We will measure the energy of a state where there is no net change in any macroscopic parameter over time
These are equilibrium states
Basic thermodynamics has difficulty dealing with non-equilibrium states
Types of Systems
Closed system – the number and type of molecules is fixed
Pure – molecules all have the same chemical structure
Homogeneous – all chemical components in the same phase of matter
Heterogeneous – more than one phase of matter present
Definitions of Terms Used for Thermodynamics
System
Meaning
Open
May exchange mass and energy with surroundings
Closed
All \(n_i\) fixed (no mass exchange and no reactions)
Pure
One chemical component (implied unless stated otherwise)
Isolated
No mass or energy exchange with surroundings
Homogeneous
One phase of matter
Heterogeneous
More than one phase of matter
Container
Meaning
Adiabatic
Walls prevent heat flow
Diathermal
Walls allow heat flow
Processes
Meaning
Quasistatic
System and surroundings always in equilibrium
Reversible
System and surroundings at constant total entropy (\(\Delta S_T=0\))
Adiabatic
No heat flow (\(q=0\))
Isentropic
System at constant entropy (\(\Delta S=0\))
Isothermal
System at constant temperature (\(\Delta T=0\))
Isobaric
system at constant pressure (\(\Delta P=0\))
Isochoric
System at constant volume (\(\Delta V=0\))
Isoenergetic
System at constant energy (\(\Delta E=0\))
The Standard State
The standard state is typically labeled with a subscript \(\ominus\)
Chemical potential: \(\mu\equiv \left(\frac{\partial E}{\partial n}\right)_{S,V}\)
The Fundamental Equation
The fundamental equation is \(E=E(S,V,n_1,\dots,n_k)\)
Using this and our previous derivative equation and the slope rule for derivatives we come to
$$ \begin{align}
dE =& \left(\frac{\partial E}{\partial S}\right)_{V,n_1,\dots,n_k} dS + \left(\frac{\partial E}{\partial V}\right)_{S,n_1,\dots,n_k} dV \\
&+ \left(\frac{\partial E}{\partial n_1}\right)_{S,V,n_2,\dots,n_k} dn_1 + \cdots + \left(\frac{\partial E}{\partial n_k}\right)_{S,V,n_1,\dots,n_{k-1}} dn_k \\
=& T\,dS - P\,dV + \mu_1\,dn_1 + \dots + \mu_k dn_k
\end{align} $$
Simplifying
Our differential equation is most convenient when a number of the derivatives on the RHS can be set to zero
For a closed system, the moles numbers do not change so \(dn_i = 0\)
This results in closed systems obeying the equation
$$ \text{closed system: }dE=T\,dS - P\,dV $$
Legendre Transformation
We would prefer equations in terms of \(T\) and \(P\) rather than \(S\)
Let’s say the energy only depends on two extensive variables such that
$$ \left(\frac{\partial E}{\partial X_1}\right)_{X_2} = Y_1 \;\; \left(\frac{\partial E}{\partial X_2}\right)_{X_1}=Y_2 $$
The partial Legendre transform of energy with respect to \(X_1\) switches the functional dependence to \(Y_1\)
$$ Z(Y_1,X_2)=E(X_1,X_2)-X_1Y_1 $$
$$ \left(\frac{\partial E}{\partial Y_1}\right)_{X_2} = -X_1 \;\; \left(\frac{\partial E}{\partial X_2}\right)_{Y_1}=Y_2 $$
Doing some differential math we arrive at
$$ dZ=-X_1\,dY_1 + Y_2\, dX_2 $$
Legendre Transforms - Thermodynamic Potentials
The enthalpy, \(H\), transform of \(E\) with respect to \(V\)
$$ H(S,P,n_1,\dots,n_k) = E+PV $$
The Helmholtz free energy, \(F\), transform of \(E\) with respect to \(S\)
$$ F(T,V,n_1,\dots,n_k) = E-TS $$
The Gibbs free energy, \(G\), transform of \(E\) with respect to \(V\) and \(S\)
$$ G(T,P,n_1,\dots,n_k) = E-TS+PV = H-TS $$
Derivatives From the Thermodynamic Potentials
From the thermodynamic potentials we can define some useful derivative equations
$$ \begin{align}
dE &= T\, dS - P\, dV + \mu_1\, dn_1 + \cdots + \mu_k\, dn_k \\
dH &= T\, dS + V\, dP + \mu_1\, dn_1 + \cdots + \mu_k\, dn_k \\
dF &= -S\, dT - P\, dV + \mu_1\, dn_1 + \cdots + \mu_k\, dn_k \\
dG &= -S\, dT + V\, dP + \mu_1\, dn_1 + \cdots + \mu_k\, dn_k
\end{align} $$
Closed System Example
In a closed system in contact with a temperature and pressure reservoir (fixes \(T\) and \(P\)) the entropy change is
$$ \begin{align}
T\,dS &= dE+P\,dV \\
dS &= \frac{1}{T} \left(dE+P\,dV\right) \\
dS &= \frac{dH}{T}
\end{align} $$
The order of when you take the two derivatives so these expressions are the same
$$ \begin{align}
\left[ \frac{\partial}{\partial V} \left( \frac{\partial E}{\partial S} \right)_{V,n} \right]_{S,n} &= \left[ \frac{\partial}{\partial S} \left( \frac{\partial E}{\partial V} \right)_{S,n} \right]_{V,n} \\
\left( \frac{\partial T}{\partial V} \right)_{S,n} &= -\left( \frac{\partial P}{\partial S} \right)_{V,n}
\end{align} $$
We know that
$$ \begin{align}
q_\chem{rot} &= \frac{k_\mathrm{B}T}{B} \;\; q_\chem{vib} = \left[ 1-e^{-\frac{\omega_e}{k_\mathrm{B}T}}\right]^{-1} \\
Q(T) &= q_\chem{elec}(T)q_\chem{vib}(T)q_\chem{rot}(T)Q_\chem{trans}(T)
\end{align} $$
The internal energy can be defined in terms of the partition function, as we have learned
$$ E=k_\mathrm{B}T^2\left( \frac{\partial \ln Q(T)}{\partial T}\right)_{V,n} $$
Heat capacity quantifies the amount of energy we must add to a sample to change its temperature by a certain amount
Effectively it is the flow of heat per unit change in temperature \(C(T)\equiv \frac{\dbar q}{dT}\)
Molar heat capacity is an intensive rather than extensive property
$$ C_m(T) \equiv \frac{1}{n} C(T) = \frac{1}{n} \frac{\dbar q}{dT} $$
Measuring Heat Capacities
We can measure the heat capacity under various conditions
When measured under constant pressure conditions we get \(C_P\)
When measured under constant volume conditions we get \(C_V\)
\(C_P\) is the most commonly tabulated form of heat capacity.
Different Types of Work
More on Heat Capacities
\(C\) measures an incremental change in one form of transferred energy
Incremental change in energy of a closed system can be written as \(dE=T\,dS-P\,dV = \dbar q + \dbar e\)
Changes in volume are usually associated with work
The most straightforward work is the mechanical energy expended or absorbed as a sample expands or contracts, called \(PV\) work
\(PV\) Work
If only \(PV\) work, then the incremental \(\dbar w\) is the change in energy due to the incremental expansion or contraction \(\dbar w = -P_\chem{min}\,dV\)
For reversible process, we change the volume so slowly that the pressure on both sides of the boundary are always equal to within our ability to measure the difference, therefore \(\dbar w = -P\,dV\)
$$ \dbar q_\chem{rev} = dE-\dbar w_\chem{rev} = T\,dS-P\,dV-(-P\,dV) = T\,dS $$
We know that \(E=\frac{1}{2}N_{ep}NK_\mathrm{B}T=\frac{1}{2}N_{ep}nRT\)
\(N_{ep}\) is the number of equipartition degrees of freedom in translation, rotation, and vibration
Using our new \(C_V\) equation we can find that
$$ C_V=\left(\frac{\partial E}{\partial T}\right)_V=\frac{1}{2}N_{ep}nR\left(\frac{\partial T}{\partial T}\right)_V = \frac{1}{2}N_{ep}nR $$
For monatomic gas, \(N_{ep}=3\) so \(C_V=\frac{3}{2}nR\)
Example 7.1
You have a constant-volume sample with a known quantity of mercury gas, a thermometer, and a calorimeter that allows you to add or remove measured amounts of energy from the sample. How could you use this apparatus to determine whether gas-phase mercury was in monatomic or diatomic form, that is, \(\chem{Hg(g)}\) or \(\chem{Hg_2(g)}\)? What numerical results would you look for in your measurements?
Constant Pressure Heat Capacity
We can begin with enthalpy
$$ dH=T\,dS + V\,dP + \mu\,dn $$
For a closed sample at constant pressure, we set \(dP\) and \(dn\) to zero
Recalling that \(H=E+PV\) we find that
$$ dH=dE+P\,dV + V\,dP = \dbar q + \dbar w + P\,dV + V\,dP $$
So at constant pressure if \(dw=-P\,dV\), \(dH=\dbar q\)
$$ C_P \equiv \left(\frac{\partial q_\chem{rev}}{\partial T}\right)_P = \left(\frac{\partial H}{\partial T}\right)_P $$
Heating at Constant Pressure vs Constant Volume
How Are \(C_P\) and \(C_V\) Related?
Start with enthalpy \(H=E+PV\)
Now we expand our expression for \(C_P\)
$$ C_P=\left(\frac{\partial (E+PV)}{\partial T}\right)_P=\left(\frac{\partial E}{\partial T}\right)_P + P\left(\frac{\partial V}{\partial T}\right)_P $$
We can use a calculus identity
$$ \begin{align}
\left(\frac{\partial X}{\partial Y}\right)_Z &= \left(\frac{\partial X}{\partial Y}\right)_W + \left(\frac{\partial X}{\partial W}\right)_Y \left(\frac{\partial W}{\partial Y}\right)_Z \\
\left(\frac{\partial E}{\partial T}\right)_P &= \left(\frac{\partial E}{\partial T}\right)_V+\left(\frac{\partial E}{\partial V}\right)_T\left(\frac{\partial V}{\partial T}\right)_P
\end{align} $$
How are \(C_P\) and \(C_V\) Related?
We can now rewrite our equation
$$ C_P = \left(\frac{\partial E}{\partial T}\right)_V+\left(\frac{\partial E}{\partial V}\right)_T\left(\frac{\partial V}{\partial T}\right)_P+P\left(\frac{\partial V}{\partial T}\right)_P $$
The first term is \(C_V\)
The second term is called the internal pressure
Using our definition of the coefficient of thermal expansion
$$ \begin{align}
C_P &= C_V + V\alpha \left[ \left(\frac{\partial E}{\partial T}\right)_T+P\right] \\
C_{Pm} &= C_{Vm} + V_m \alpha \left[ \left(\frac{\partial E}{\partial T}\right)_T+P\right]
\end{align} $$
Ideal Gas
The coefficient of thermal expansion of an ideal gas is
$$ \alpha=\frac{1}{V}\left(\frac{\partial V}{\partial T}\right)_P = \frac{1}{V}\left[\frac{\partial}{\partial V} \left(\frac{nRT}{P}\right)\right]_P = \frac{nR}{PV} = \frac{1}{T} $$
The ideal gas constant pressure heat capacity is therefore \(C_P=C_V+V\alpha P=C_V+nR\)
Example 7.2
Use the definition of the coefficient of thermal expansion to find the relationship between \(C_P\) and \(C_V\) at the temperature at maximum density (TMD) for any substance.
Heat Capacities and Real Gases
The general relationship between the heat and the temperature change is
$$ q=\int_{T_1}^{T_2}\dbar q = \int_{T_1}^{T_2}\frac{\dbar q}{dT}dT=\int_{T_1}^{T_2}C(T)\,dT $$
If we assume that the heat capacity does not depend on temperature
$$ q=\int_{T_1}^{T_2} C\,dT=C\int_{T_1}^{T_2}dT=C\Delta T $$
Heat Capacities of Several Substances
More on Heat Capacities of Real Gases
Heat capacities for gas-phase atoms tend to lie very close to the equipartition value
The more atoms in the molecule, the more the equipartition heat capacity overestimates the experimental heat capacity
Heat capacities for diatomics and triatomics are more accurately predicted by full equipartition when the atoms are heavier
A lower limit to the heat capacity can therefore be estimated by neglecting all the vibrations
Example 7.3
The molar heat capacity at constant pressure of \(\chem{CCl_4}\) is \(83.30\,\mathrm{J\,K^{-1}\,mol^{-1}}\) at \(298\,\mathrm{K}\). If we heated a \(1.00\,\mathrm{mol}\) sample of \(\chem{CCl_4}\) gas with \(1.00\,\mathrm{J}\), starting near \(298\,\mathrm{K}\), roughly what percent of the heat would go into rotation, and what percent into vibration? Assume that the temperature change is small enough that the heat capacity may be treated as constant.
For a Linear Diatomic Molecule
After a lot of math we can arrive at an equation for the constant pressure heat capacity in terms of vibrational frequency
The vibrational constant of \(\chem{{}^{35}Cl_2}\) is \(560.5\,\mathrm{cm^{-1}}\). Estimate the value of \(C_{Pm}\) for \(\chem{{}^{35}Cl_2}\) at \(298.15\,\mathrm{K}\).
Idealized Solid
Einstein Solid
Albert Einstein derived the heat capacity for an idealized crystal, called the Einstein solid
He treated each particle as an independent harmonic oscillator
Einstein postulated that he could find the value ε by assuming some single effective vibrational constant \(\omega_E\)
$$ q_\chem{vib}(T,V) = \frac{1}{1-e^{-\frac{\omega_E}{k_\mathrm{B}T}}} $$
The Einstein Heat Capacity
If we assume that each atom vibrates as an independent harmonic oscillator
$$ E_m = 3\mathcal{N}_\mathrm{A} \expect{\varepsilon}_\chem{vib} = \frac{3\mathcal{N}_\mathrm{A}\omega_E}{e^{\frac{\omega_E}{k_\mathrm{B}T}}-1} $$
From this we can derive the constant volume molar heat capacity
$$ C_{Vm} = \frac{3\mathcal{N}_\mathrm{A}\omega_e^2e^{\frac{\omega_e}{k_\mathrm{B}T}}}{k_\mathrm{B}T^2\left(e^{\frac{\omega_e}{k_\mathrm{B}T}}-1\right)^2} $$
Success of the Einstein Solid
At high temperatures, the heat capacity predicted by the Einstein solid is usually close to experimental
When the solid is cooled, the individual stretches no longer can be excited and the heat capacity becomes radically different
At cold temperature, the vibrations are dominated by low-frequency, collective motions of the atoms called phonon modes
Phonon Modes in Crystals
Debye Heat Capacity
The Debye heat capacity improves on the Einstein solid by taking collective motions into account
Debye argued that the density of quantum states \(W\) for these oscillations goes as (\(A\) is a normalization constant)
$$ W(\omega)\,d\omega = AV\omega^2\,d\omega $$
$$ A=\frac{9N}{\omega_D^2V} $$
The Debye Heat Capacity Equation
Without doing all the calculus, we can find that
$$ \begin{align}
E_m &\approx \frac{3\pi^4 \mathcal{N}_\mathrm{A} k_\mathrm{B}^4T^4}{5\omega_D^3} \\
C_{Vm} &= \left( \frac{\partial E_m}{\partial T}\right)_V \approx \frac{12\pi^4\mathcal{N}_\mathrm{A}k_\mathrm{B}^4T^3}{5\omega_D^3}
\end{align} $$
Comparison of Heat Capacities
Specific Heat Capacities
It is often easier to measure the mass of liquids and solids compared to gases
This means that we can put the heat capacities of solids and liquids on a per gram basis rather than a per mole basis
Heat capacities per gram are called specific heat capacities and designated with a lower case c
Heat Capacities of Selected Solids
Solid
\(\mathcal{M} \\ (\mathrm{g\,mol^{-1}})\)
\(C_{Pm} \\ (\mathrm{J\,K^{-1}\,mol^{-1}})\)
\(c_{P} \\ (\mathrm{J\,K^{-1}\,g^{-1}})\)
Solid
\(c_{P} \\ (\mathrm{J\,K^{-1}\,g^{-1}})\)
Lithium
6.94
24.8
0.516
Hematite
Beryllium
9.01
16.4
0.113
Pyrex glass
0.84
Carbon (graphite)
12.01
8.53
0.059
Gypsum
1.09
Carbon (diamond)
12.01
6.11
0.043
Loose wool
1.26
Sodium
22.99
28.2
1.23
Paper
1.34
Silicon
28.09
20.0
0.75
White pine
2.5
Sulfur
32.07
22.6
0.71
Paraffin wax
2.9
Arsenic
74.92
24.6
0.33
Tantalum
180.9
25.4
0.14
Uranium
238.0
27.7
0.12
Heat Capacities of Monatomic Liquids
We can define the heat capacities in terms of the pair correlation function, \(\mathcal{G}(R)\)
$$ \begin{align}
C_{Vm} &= \frac{3R}{2} + 2\pi \mathcal{N}_\mathrm{A}^2\left(\frac{\rho_m}{\mathcal{M}}\right)\int_0^\infty u(R) \left( \frac{\partial \mathcal{G}(R)}{\partial T}\right)_V R^2\,dR \\
&\approx \frac{3R}{2}+\left(\frac{2\pi \mathcal{N}_\mathrm{A}^2}{k_\mathrm{B}T}\right)\left(\frac{\rho_m}{\mathcal{M}}\right)\left\{ \frac{304\varepsilon^2R_\chem{LJ}^3}{315}\right\} \\
&\approx \frac{3R}{2} + \left(2\pi \mathcal{N}_\mathrm{A}k_\mathrm{B}R_\chem{LJ}^3\right)\left(\frac{\rho_m}{\mathcal{M}}\right) \\
&\approx \frac{3R}{2}+\left(\frac{3Rb\rho_m}{\mathcal{M}}\right)=3R\left(\frac{1}{2}+\frac{b\rho_m}{\mathcal{M}}\right)
\end{align} $$