You are on page 1of 19

Introduction 1.

Foundations of thermodynamics
• Thermodynamics: phenomenological description of 1.1. Fundamental thermodynamical
equilibrium bulk properties of matter in terms of only concepts
a few “state variables” and thermodynamical laws.
System : macroscopic entity under consideration.
• Statistical physics: microscopic foundation of
thermodynamics Environment : world outside of the system (infinite).

• ∼ 1023 degrees of freedom → 2–3 state variables! Open system : can exchange matter and heat with the
environment.
• “Everything should be made as simple as possible,
but no simpler” (A. Einstein) Closed system : can exchange heat with the
environment while keeping the number of particles
fixed.
Summary of contents: Isolated system : can exchange neither matter nor heat
with the environment. Can (possibly) still do work
• Review of thermodynamics
by e.g. expanding.
• Thermodynamical potentials
Thermodynamical equilibrium:
• Phase space and probability • No macroscopic changes.
• Quantum mechanical ensembles • Uniquely described by (a few) external variables of
• Equilibrium ensembles state.

• Ideal fluids • System forgets its past: no memory effects, no


hysteresis.
• Bosonic systems
• Often the term global equilibrium is used, as opposed
• Fermionic systems to local equilibrium, which is not full equilibrium at
all (next page)!
• Interacting systems
Nonequilibrium:
• Phase transitions and critical phenomena
• Generally much more complicated than equilibrium
state.
• Simplest case: isolated systems each in an
equilibrium state.
• In a local thermodynamical equilibrium small regions
are locally in equilibrium, but neighbour regions in
different equilibria ⇒ particles, heat etc. will flow.
Example: fluid (water) with non-homogeneous
temperature.
• Stronger nonequilibrium systems usually relax to a
local equilibrium.

Degrees of freedom (d.o.f.) is the number of quantities


needed for the exact description of the microscopic
state.
Example: classical ideal gas with N particles: 3N
coordinates (x, y, z), 3N momenta (px , py , pz ).
State variables are parameters characterizing the
macroscopic thermodynamical state. These are all
extensive or intensive:
Extensive variable: change value when the size
(spatial volume and the number of degrees of
freedom) is changed: volume V , particle number
N , internal
R energy U , entropy S, total magnetic
moment d3 r M.

1
Intensive variable: independent of the size of the
system, and can be determined for every 1 = 2
semimicroscopical volume element: e.g.
temperature T , pressure p, chemical potential µ,
magnetic field H, ratios of extensive variables
like ρ = N/V , s = S/N, . . .. I I
Conjugated variables: A and B appear in pairs in dp = dU = · · · = 0.
expressions for the differential of the energy (or 1→2 1→2

more generally, some state variable), i.e. in (B) The total change of an exact differential is
forms ±A dB or ±B dA; one is always extensive independent on the path of integration:
and the other intensive. 1
b
Example: pressure p and volume V ; change in
internal energy U when V is changed
(adiabatically, at constant S) is dU = −pdV . a
2
Process is a change in the state. Z Z
Reversible process: advances via states dU = 0, dU −
a b
infinitesimally close to equilibrium,
so that we can write
quasistatically (“slow process”). The direction of
a reversible process can be reversed, obtaining Z 2
the initial state (for system + environment!) U (2) = U (1) + dU
1
Isothermal process : T constant.
Isobaric process : p constant.
Exact differentials
Isochoric process : V constant. Let us denote by dF¯ a differential which is not necessarily
Isentropic or adiabatic process: S constant. exact (i.e. integrals can depend on the path). Assuming
Irreversible process is a sudden or spontaneous it depends on 2 variables x, y, the differential
change during which the system is far from
equilibrium. In the intermediate steps global dF
¯ = F1 (x, y) dx + F2 (x, y) dy
state variables (p, T , . . .) are usually not well
is exact differential if
defined.
Cyclic process consists of cycles which take the ∂F1 ∂F2
= .
system every time to its initial state. ∂y ∂x
∂F (x,y)
Then ∃F (x, y) so that F1 (x, y) = ∂x and
1.2. State variables and exact F2 (x, y) = ∂F∂y
(x,y)
and
differentials
2
Let us suppose that, for example, the state of the system
Z
can be uniquely described by state variables T , V ja N . dF
¯ = F (2) − F (1)
1
Other state variables are then their unique functions:
is independent on the path, and integrable. In this case
p = p(T, V, N ) (x, F1 ) and (y, F2 ) are pairs of conjugated variables with
U = U (T, V, N ) respect to F .
Examples: are the following differentials exact?
S = S(T, V, N ) . . .
d̄F = y dx + x dy
By applying differential calculus, the differential of p, for
example, is dF
¯ = x dx + x dy
     
∂p ∂p ∂p All physical state variables are exact differentials! This
dp = dT + dV + dN
∂T V,N ∂V T,N ∂N T,V will enable us to derive various identities between state
.. variables.
.
Integrating factor
The differentials of state variables, If dF
¯ = F1 dx + F2 dy is not exact, there exists an
dp, dT , dV , . . ., are exact differentials. These have the integrating factor λ(x, y) so that in the neighbourhood of
following properties the point (x, y)
(A) Their total change evaluated over a closed path
vanishes: λ¯
dF = λF1 dx + λF2 dy = df

2
is an exact differential. λ and f are state variables. 1.3. Equations of state
Example: find λ for the differential Encodes (some of the) physical properties of the
equilibrium system. Usually these relate “mechanical”
dF
¯ = x dx + x dy . readily observable variables, like p, T , N , V ; not
“internal” variables like S, internal energy U etc. A
typical example: pressure of some gas as a function of T
Legendre transformations and density ρ.
Legendre transformations can be used to make changes in Some examples:
the set of the indepependent state variables. For example,
let us look at the function f (x, y) of two variables. We
Classical ideal gas
denote
∂f (x, y)
z = fy = pV = N kB T
∂y
and define the function where N = number of molecules
T = absolute temperature
g = f − yfy = f − yz. kB = 1.3807 · 10−23 J/K = Boltzmann constant. Chemists
use often the form
(Note: z, y is a congjugated pair with respect to f !) Now
pV = nRT
dg = df − y dz − z dy = fx dx + fy dy − y dz − z dy
n = N/N0 = number of moles
= fx dx − y dz.
R = kB N0 = 8.315J/K mol
Thus we can take x and z as independent variables of the = gas constant
function g, i.e. g = g(x, z). Obviously
N0 = 6.0221 · 1023 = Avogadro’s number.
∂g(x, z)
y=− . If the gas is composed of m different species of molecules
∂z
the equation of state is still
Corresponding to the Legendre transformation f → g
there is the inverse transformation g → f pV = N kB T,

f = g − zgz = g + yz. where now


m
X
N= Ni
i=1
Often needed identities
and
Let F = F (x, y), x = x(y, z), y = y(x, z) and z = z(x, y). X
If we want to give F in terms of (x, z), we can write p= pi , pi = Ni kB T /V,
i
F (x, y) = F (x, y(x, z)). where pi is the partial pressure of the i:th component
Applying differential rules we obtain identities
Virial expansion of real gases
When the interactions between gas molecules are taken
       
∂F ∂F ∂F ∂y
= + into account, the ideal gas law receives corrections which
∂x z ∂x y ∂y x ∂x z
are suppressed by powers of density ρ = N/V :
     
∂F ∂F ∂y
= p = kB T ρ + ρ2 B2 (T ) + ρ3 B3 (T ) + · · ·
 
∂z x ∂y x ∂z x

One can show that


  Here Bn is the n:th virial coefficient.
∂x 1
= 
∂y z
∂y Van der Waals equation
∂x
z The molecules of real gases interact
     
∂x ∂y ∂z • repulsively at short distances; every particle needs at
= −1 >
∂y z ∂z x ∂x y least the volume b ⇒ V ∼ N b.
and
 

∂F
 • attractively (potential ∼ (r/r0 )6 ) at large distances
∂x ∂y due to the induced dipole momenta. The pressure
= z .
∂y z
∂F decreases when two particles are separated by the
∂x z
attraction distance. The probability of this is
∝ (N/V )2 .

3
We improve the ideal gas state equation where

p′ V ′ = N kB T P = electric polarization
= atomic total dipole momentum/volume
so that
D = electric flux density
V ′
= V − Nb ǫ0 = 8.8542 · 10−12 As/Vm
p = p′ − aρ2 = true pressure. = vacuum permeability.

then In homogenous dielectric material one has


2
(p + aρ )(V − N b) = N kB T.  
b
P= a+ E,
T

Solid substances where a and b are almost constant and a, b ≥ 0.


The thermal expansion coefficient
  Curie’s law
1 ∂V When a piece of paramagnetic material is in magnetic
αp =
V ∂T p,N field H we write

and the isothermal compressibility B = µ0 (H + M),

where
 
1 ∂V
κT = −
V ∂p T,N
M = magnetic polarization
of solid materials are very small, so the Taylor series = atomic total magnetic moment/volume
B = magnetic flux density
V = V0 (1 + αp T − κT p)
µ0 = 4π · 10−7 Vs/Am = vacuum permeability.
is a good approximation.
Polarization obeys roughly Curie’s law
Typically
ρC
κT ≈ 10−10 /Pa M= H,
T
αp ≈ 10−4 /K.
where ρ is the number density of paramagnetic atoms and
C an experimental constant related to the individual
atom.
Stretched wire Note: Use as a thermometer: measure the quantity
Tension [N/m2 ] M/H.

σ = E(T )(L − L0 )/L0 ,

where L0 is the length of the wire when σ = 0 and E(T )


is the temperature dependent elasticity coefficient.

Surface tension

 n
t
σ = σ0 1 − ′
t
t = temperature ◦ C
t′ and n experimental constants,
< <
1∼ n∼2
σ0 = surface tension when t = 0◦ C.

Electric polarization
When a piece of material is in an external electric field E,
we define
D = ǫ0 E + P,

4
Laws of thermodynamics If the system can exchange heat and particles and do
Thermodynamics is based upon 4 laws (these can be work, the energy conservation law gives the change of the
derived from statistical physics, but in thermodynamics internal energy
these are considered “fundamental”).
dU = d̄Q − d̄W + µdN,
1.4. 0th law of thermodynamics
If each of two bodies is separately in thermal equilibrium where µ is the chemical potential. More generally,
with a third body then they are also in thermal X
dU = d̄Q − f · dX + µi dNi .
equilibrium with each other. ⇒ there exists a property
i
called temperature and thermometer which can be used to
measure it. U is a state variable, i.e. dU is exact.

1.5. Work Cyclic process


Before we discuss the 1st law, let us introduce the In a cyclic process the system returns to the original state.
concept of Work. Work is exchange of such ”noble”
H
Now dU = 0, so ∆W = ∆Q (no change in thermal
energy (as opposed to exchange of heat or matter) that energy). Now ∆Q+ is the heat absorbed by the system
can be completely transformed to some other noble form during one cycle and −∆Q− (> 0) is the heat released.
of energy; e.g. mechanical and electromagnetic energy.
Sign convention: work ∆W is the work done by the
p
system to its environment. Example: pV T system 6
∆Q+

'- $
∆W = p ∆V.

Note: dW¯ is not an exact differential: the work done by 
the system is not a function of the final state of the -∆W =Area =
H
p dV = in pV T -system
system (need to know the history!). Instead
1 & Q%
Q
dW
¯ = dV s∆Q−
Q
p
-
is exact, i.e, 1/p is the integrating factor for work.
V
Example:

dW
¯ = p dV − σA dL − E · dP − H · dM. The total change of heat is
In general X ∆Q = ∆W = ∆Q+ + ∆Q− ,
dW
¯ = fi dXi = f · dX,
i The efficiency of a heat engine (∆W > 0) is work/(heat
where fi is a component of a generalized force and Xi a taken):
component of a generalized displacement.
∆W ∆Q+ + ∆Q− |∆Q− |
η= = = 1 − .
1.6. 1st law of thermodynamics ∆Q+ ∆Q+ |∆Q+ |
Total energy is conserved
In addition to work a system can exchange heat (thermal
energy) or chemical energy, associated with the exchange 1.7. 2nd law of thermodynamics
of matter, with its environment. Thermal energy is 2nd law can be stated in various (historical) ways:
related to the energy of the thermal stochastic motion of
(a) Heat flows spontaneously from high temperatures to
microscopic particles.
low temperatures.
The total energy of a system is called internal energy
U. (b) Heat cannot be transferred from a cooler heat
Sign conventions: reservoir to a warmer one without other changes.

' $
(c) In a cyclic process it is not possible to convert all
∆Q Environment
 heat taken from the hotter heat reservoir into work.


+
 (d) It is not possible to reverse the evolution of a system
towards thermodynamical equilibrium without
System -∆W
converting work to heat.
Q
k
Q
Q (e) The change of the total entropy of the system and its
Q
& %chemical energy
Q µ ∆N environment is positive and can be zero only in
reversible processes.

5
(f) Of all the engines working between the temperatures We define the efficiency as
T1 and T2 the Carnot engine has the highest
efficiency. ∆W ∆Q1
η= =1− .
∆Q2 ∆Q2
We consider the infinitesimal process
Because the processes are reversible the cycle C can be
y reversed and C works as a heat pump. Let us consider
6
two Carnot cycles A and B, for which

' t
U (2) ∆WA = ∆WB = ∆W.
C 2
U (1) t CW
d̄Q
A is an enegine and B a heat pump. The efficiences are
1 correspondingly
∆W ∆W
ηA = and ηB = .
∆QA ∆QB
-
x T 2
D Q D Q
A
W B

Now A B T 2> T 1
d̄Q = dU + d̄W = dU + f · dX, D Q A -W D Q B -W
so there exists an integrating factor 1/T so that T 1

1 Let us suppose that


dQ
¯ = dS
T
ηA > ηB ,
is exact. The state variable S is entropy and T turns out
to be temperature (on an absolute scale) The second law so that ∆QB > ∆QA or ∆QB − ∆QA > 0. The heat
(e) can now be written as would transfer from the cooler reservoir to the warmer
one without any other changes, which is in contradiction
dStot with the second law (form b). So we must have
≥ 0,
dt
ηA ≤ ηB .
where Stot is the entropy of the system + environment.
For the entropy of the system only we have By running the engines backwards one can show that
1
dS ≥ d̄Q, ηB ≤ ηA ,
T
where the equality holds only for reversible processes. so that ηA = ηB , i.e. all Carnot engines have the same
The entropy of the system can decrease, but the total efficiency.
entropy always increases (or stays constant). Similarly, it follows that the Carnot engine has the
For reversible processes the first law can be rewritten as highest efficiency among all engines (also irreversible)
dU = dQ
¯ − d̄W + µ dN = T dS − p dV + µ dN. working between given temperatures: assume that engine
A in previous figure is some other engine than Carnot.
Then the argument above implies that ηA ≤ ηB . If A is
1.8. Carnot cycle not reversible, efficiency is not necessarily the same while
Illustrates the concept of entropy. The Carnot engine C running the system backwards and the inequality remains
consists of reversible processes in force; if it is reversible, reversing the process gives
ηA = ηB .
a) isothermal T2 ∆Q2 > 0 Note: The efficiency does not depend on the realization
b) adiabatic T2 → T1 ∆Q = 0 of the cycle (e.g. the working substance). Only
c) isothermal T1 ∆Q1 > 0 reversibility is essential! ⇒ The efficiency depends only
d) adiabatic T1 → T2 ∆Q = 0 on the temperatures of the heat reservoirs.
Now ∆U = 0, so ∆W = ∆Q2 − ∆Q1 (wrong sign here, Defining temperature scale via Carnot engines: Let us
for simplicity). consider Carnot’s cycle between temperatures T3 and T1 .
Now
a D Q 2
η = 1 − f (T3 , T1 ),
d
b where we have the identity
c
∆Q1
D Q 1 f (T3 , T1 ) = .
∆Q3

6
T 3 D Q
is exact and the entropy S and the temperature T are
3
D W D Q 3 state variables. Because the Carnot cycle has the highest
2 3
D Q D W
T
2 efficiency, a cycle containing irreversible processes satisfies
D Q
1 3
2
D Q
2
D W 1 2 1
D Q ∆Q1 T1
T
1 ηirr = 1 − < ηCarnot = 1 −
1 ∆Q2 T2
Here or
∆Q2 ∆Q1
∆Q2 − < 0.
f (T3 , T2 ) = T2 T1
∆Q3 Thus for an arbitrary cycle we have
∆Q1
f (T2 , T1 ) = I
dQ
¯
∆Q2 ≤ 0, (∗)
∆Q1 T
f (T3 , T1 ) =
∆Q3 where the equality holds only for reversible processes.
so Because entropy S is a state variable, it cannot depend
f (T3 , T1 ) = f (T3 , T2 )f (T2 , T1 ). on the integration path. Thus, for an arbitrary process
1 → 2 the change of the entropy can be obtained from the
The simplest solution is formula
dQ
Z Z
¯
∆S = dS = .
T1 rev rev T
f (T2 , T1 ) = .
irr 2
T2

We define the absolute temperature so that

T1
η =1− . 1
T2
re v
This (theoretical) definition was first used by Kelvin, and According to the formula (∗) we have
it gives us our familiar absolute temperature scale, up to Z
dQ
Z
dQ
¯ ¯
a scale factor. − < 0,
irr T rev T
(This is by no means unique; one could also use
f (t2 , t1 ) = [t2 /t1 ]a , with any a 6= 0, this would just give us or
dQ
Z
a different temperature scale t = const. × T 1/a .) ¯
∆S > .
The Carnot cycle satisfies irr T
This is usually written as
dQ
I
¯
= 0,
T dQ
¯
dS ≥ ,
T
since, during the isothermal part a,
and the equality is valid only for reversible processes. For
dQ ∆Q2
Z
¯
= an isolated system (∆Q = 0, e.g. system + environment!)
a T T2 we have
∆S ≥ 0.
and part c
dQ ∆Q1 ∆Q2
Z
¯ Thus, for an irreversible process,
=− =− .
c T T1 T2
∆Qirr ≤ ∆Qrev = T ∆S,
This is valid also for an arbitrary reversible cycle
∆Wirr = ∆Qirr − ∆U ≤ ∆Wrev = f · dX
C
where “rev” is a hypothetical reversible process
C i connecting the initial and final states of “irr”-trajectory.
Note the sign conventions: ∆W work done by the system;
∆Q heat absorbed by the system. ∆Wirr ≤ ∆Wrev means
less work done by “irr” (∆W > 0) or more work done to
because it (∆W < 0).
dQ dQ
I I
¯ X ¯
= = 0.
C T i Ci T 1.9. 3rd law of thermodynamics
Nernst’s law (1906):
So
dQ
¯
dS = lim S = 0.
T T →0

7
A less strong form can be stated as: 2. Thermodynamic potentials
When the maximum temperature occuring in the process
from a state a to a state b approaches zero, also the 2.1. Fundamental equation
entropy change ∆Sa→b → 0. According to the first law (in a T V N -system, generalizes
Note: There are systems whose entropy at low easily), for reversible processes
temperatures is larger than true equilibria would allow.
This is due to very slow relaxation time dU = T dS − p dV + µ dN . (∗)

S, V and N can be considered to be natural variables of


the internal energy U , i.e. U = U (S, V, N ). Furthermore,
from the law (∗) one can read the relations
 
∂U
= T
∂S V,N
 
∂U
= −p (∗∗)
∂V S,N
 
∂U
= µ.
∂N S,V

Scaling law of extensive variables: all extensive variables


must be linear functions of system size V (and each
other). Now U , S, V and N are extensive so we have
U (λS, λV, λN ) = λU (S, V, N ) ∀λ. (∗ ∗ ∗)
Taking a derivative of (∗ ∗ ∗) wrt. λ, we obtain the Euler
equation for homogenous functions
     
∂U ∂U ∂U
U =S +V +N .
∂S V,N ∂V S,N ∂N S,V

Substituting the partial derivatives in (∗∗) this takes the


form
U = T S − pV + µN
or
1
S= (U + pV − µN ).
T
This is called the fundamental equation.

2.2. Internal energy and Maxwell


relations
Because  
∂U
T =
∂S V,N

and  
∂U
p=− ,
∂V S,N
so
∂T ∂ ∂U ∂ ∂U ∂p
= = =− .
∂V ∂V ∂S ∂S ∂V ∂S
Similar relations can be derived also for other partial
derivatives of U and we get so called Maxwell’s relations
   
∂T ∂p
= −
∂V S,N ∂S V,N
   
∂T ∂µ
=
∂N S,V ∂S V,N
   
∂p ∂µ
= − .
∂N S,V ∂V S,N

8
In an irreversible process is called enthalpy.
Now
T ∆S > ∆Q = ∆U + ∆W,
dH = dU + p dV + V dp
so
∆U < T ∆S − p ∆V + µ ∆N. = T dS − p dV + µ dN + p dV + V dp
If S, V and N stay constant in the process then the or
internal energy decreases. Thus we can deduce that dH = T dS + V dp + µ dN.
In an equilibrium with given S, V and N the internal
energy is at the minimum. From this we can read the partial derivatives
We consider a reversible process in an isolated system 
∂H

(∆Q = 0) T =
∂S p,N
D L  
∂H
p 1
p 2 V =
F ∂p S,N
V 1 V 2
 
∂H
µ = .
∂N S,V
e q u ilib riu m p o s itio n
We partition ∆W into the components Corresponding Maxwell relations are
 
work due to the
   
Z ∂T ∂V
p dV = change of the total  =
∂p S,N ∂S p,N
volume ( = 0)    
∂T ∂µ
work done by the =
" #
∆Wfree = . ∂N S,p ∂S p,N
gas against the    
force F ∂V ∂µ
= .
Now ∂N S,p ∂p S,N

∆Wfree = ∆W1 + ∆W2 = p1 ∆V1 + p2 ∆V2 In an irreversible process one has


= (p1 − p2 )∆V1 = (p1 − p2 )A ∆L
∆Q = ∆U + ∆W − µ ∆N < T ∆S.
= −F ∆L.
Now ∆U = ∆(H − pV ), so that
According to the first law we have
Z ∆H < T ∆S + V ∆p + µ ∆N.
∆U = ∆Q − ∆W = ∆Q − p dV − ∆Wfree
We see that
= ∆Q − ∆Wfree . In a process where S, p and N are constant spontaneous
Because now ∆Q = 0, we have changes lead to the minimum of H, i.e. in an equilibrium
of a (S, p, N )-system the enthalpy is at the minimum.
∆U = −∆Wfree = F ∆L, The enthalpy is a suitable potential for an isolated system
in a pressure bath (p is constant). Let us look at an
i.e. when the variables S, V and N are kept constant the isolated system in a pressure bath. Now
change of the internal energy is completely exhangeable
with the work. ∆U is then called free energy and U dH = dU + d(pV )
thermodynamic potential. Note: If there are
irreversible processes in an isolated system (V and N and
constants) then dU = d̄Q − d̄W + µ dN.
∆Wfree ≤ −∆U. Again we partition the work into two components:
If the system does no work, ∆U ≤ 0, i.e. the system
tends to minimize its internal energy. dW
¯ = p dV + d̄Wfree .

2.3. Enthalpy Now


Using the Legendre transform dH = dQ
¯ + V dp − d̄Wfree + µ dN
  and for a finite process
∂U
U →H =U −V = U + pV
∂V S,N
Z Z Z
∆H ≤ T dS + V dp − ∆Wfree + µ dN.
We move from the variables (S, V, N ) to the variables
(S, p, N ). The quantity When (S, p, N ) is constant one has

H = U + pV ∆H ≤ −∆Wfree

9
i.e. ∆Wfree is the minimum work required for the change and the partial derivative relation
∆H.   ∂T

Note: Another name of enthalpy is heat function (in ∂V ∂S p
= ∂T 
constant pressure). ∂S p ∂V p

we can write
Joule-Thomson phenomenon  
Flow of gas through a porous wall (“choke”): T T ∂V
dT = dS + dp.
D Q = 0 Cp Cp ∂T p

p 1
Substituting into this the differential dS in constant
V 1
enthalpy (∗) we get so called Joule-Thomson coefficients
" #
  
∂T T ∂V V
c h o k e ∂p H
=
Cp ∂T p T
− .
p1 and p2 are constant (in time), p1 > p2 and the process
irreversible. When a differential amount of matter passes Defining the heat expansion coefficient αp so that
through the choke the work done by the system is
 
1 ∂V
αp = ,
V ∂T p
dW
¯ = p2 dV2 + p1 dV1 .
we can rewrite the Joule-Thomson coefficient as
V1 V2 
∂T

V
Initial state Vinit 0 = (T αp − 1).
∂p H Cp
Final state 0 Vfinal
The work done by the system is We see that when the pressure decreases the gas
Z • cools down, if T αp > 1.
∆W = dW ¯ = p2 Vfinal − p1 Vinit .
• warms up, if T αp < 1.
   
According to the first law we have For ideal gases ∂T∂p H = 0 holds. For real gases
∂T
∂p H
is below the inversion temperature positive, so the gas
∆U = Ufinal − Uinit = ∆Q − ∆W = −∆W,
cools down.
so that 2.4. Free energy
Uinit + p1 Vinit = Ufinal + p2 Vfinal . The Legendre transform
Thus in this process the enthalpy H = U + pV is 
∂U

constant, i.e. the process is isenthalpic, U →F =U −S
∂S V,N

∆H = Hfinal − Hinitial = 0. or
F = U − TS
We consider now a reversible isenthalpic (and dN = 0) defines the (Helmholtz) free energy.
process init→final. Here Now
dF = −S dT − p dV + µ dN,
dH = T dS + V dp = 0,
so the natural variables of F are T , V and N . We can
so read the partial derivateves
V
dS = − dp. (∗)
 
T ∂F
S = −
∂T V,N
Now T = T (S, p), so that  
∂F

∂T
 
∂T
 p = −
dT = dS + dp. ∂V T,N
∂S p ∂p S  
∂F
µ = .
On the other hand ∂N T,V

∂T

T From these we obtain the Maxwell relations
= , 
∂S
 
∂p

∂S p Cp =
∂V T,N ∂T V,N
where Cp is the isobaric heat capacity (see 
∂S
 
∂µ

thermodynamical responses). Using the Maxwell relation = −
∂N T,V ∂T V,N
       
∂T ∂V ∂p ∂µ
= = − .
∂p S ∂S p ∂N T,V ∂V T,N

10
In an irreversible change we have 2.6. Grand potential
The Legendre transform
∆F < −S ∆T − p ∆V + µ ∆N,    
∂U ∂U
i.e. when the variables T , V and N are constant the U →Ω=U −S −N
∂S V,N ∂N S,V
system drifts to the minimum of the free energy.
Correspondingly defines the grand potential
∆Wfree ≤ −∆F,
Ω = U − T S − µN.
when (T, V, N ) is constant.
Free energy is suitable for systems where the exchange of
Its differential is
heat is allowed; i.e. the control variables are T and V
(typically at constant N ). Very useful quantity in physics! dΩ = −S dT − p dV − N dµ,

2.5. Gibbs free energy so the natural variables are T , p andµ.


The Legendre transformation The partial derivatives are now
   
∂U ∂U
 
∂Ω
U →G=U −S −V S = −
∂S V,N ∂V S,N ∂T p,µ
 
defines the Gibbs function or the Gibbs free energy ∂Ω
p = −
∂V T,µ
G = U − T S + pV. 
∂Ω

N = − .
∂µ T,V
Its differential is
We get the Maxwell relations
dG = −S dT + V dp + µ dN,
   
∂S ∂p
so the natural variables are T , p and N . For the partial =
∂V T,µ ∂T V,µ
derivatives we can read the expressions    
∂S ∂N

∂G
 =
S = − ∂µ T,V ∂T V,µ
∂T p,N    
∂p ∂N

∂G
 = .
V = ∂µ T,V ∂V T,µ
∂p T,N

∂G
 In an irreversible process
µ = .
∂N T,p ∆Ω < −S ∆T − p ∆V − N ∆µ,
From these we obtain the Maxwell relations holds, i.e. when the variables T , V andµ are kept
constant the system moves to the minimum of Ω.
   
∂S ∂V
= − Correspondingly
∂p T,N ∂T p,N
   
∂S ∂µ ∆Wfree ≤ −∆Ω,
= −
∂N T,p ∂T p,N
    when (T, V, µ) is constant.
∂V ∂µ
= . The grand potential is suitable for systems that are
∂N T,p ∂p T,N allowed to exchange heat and particles.
In an irreversible process
Bath
∆G < −S ∆T + V ∆p + µ ∆N, A bath is an equilibrium system, much larger than the
system under consideration, which can exchange given
holds, i.e. when the variables T , p and N stay constant extensive property with our system.
the system drifts to the minimum of G. Pressure bath
Correspondingly F
D V
∆Wfree ≤ −∆G,

when (T, p, N ) is constant. The exchanged property is the volume or a corresponding


The Gibbs function is suitable for systems which are generalized displacement; for example magnetization in a
allowed to exchange mechanical energy and heat in heat- magnetic field.
and pressure baths. Heat bath

11
T The heat capacity C is defined so that
T
D Q
D S ∆Q ∆S
C= =T .
∆T ∆T
Particle bath
D S m Keeping volume and N constant, we define
m
D N
 
∂S
CV = T .
∂T V,N
Baths can also be combined; for example a suitable
potential for a pressure and heat bath is the Gibbs Now, according to the first law at constant N ,V
function G.
dU = T dS − p dV + µ dN = T dS.
2.7. Thermodynamic response functions
Response functions are thermodynamic quantities most Using S = −(∂F/∂T )V,N , we can write
accessible to experiment. They give us information about
∂2F
   
how a specific state variable changes as other independent ∂U
CV = = −T
state variables are changed. They can be classfied as ∂T V,N ∂T 2 V,N
mechanical (compressbility, susceptibility) and thermal
(heat capacity) responses.

1) Heat expansion coefficient 5) Isobaric heat capacity


 

1 ∂V
 ∂S
αp = Cp = T
V ∂T p,N ∂T p,N

or   Because
1 ∂ρ dH = T dS + V dp + µ dN,
αp = − ,
ρ ∂T p,N and using S = −(∂G/∂T )p,N , one can write
where ρ = N/V .
∂2G
   
∂H
Cp = = −T .
2) Isothermal compressibility ∂T p,N ∂T 2 p,N
   
1 ∂V 1 ∂ρ
κT = − =
V ∂p T,N ρ ∂p T,N Relating response functions
Now
3) Adiabatic compressibility
   
∂S ∂S (V (p, T ), T )
=
    ∂T p ∂T p
1 ∂V 1 ∂ρ
κS = − = . 
∂S
 
∂S
 
∂V

V ∂p S,N ρ ∂p S,N = +
∂T V ∂V T ∂T p
The velocity of sound depends on the adiabatic
compressibility like and (a Maxwell relation)
r
1
   
cS = , ∂S ∂p
mρκS = ,
∂V T ∂T V
where m the particle mass. so    
One can show that ∂p ∂V
Cp = CV + T .
∂T V ∂T p
α2p
κT = κS + V T . Since
Cp 
∂p
 
∂T
 
∂V

= −1
∂T V ∂V p ∂p T

4) Isochoric heat capacity or


∂V
  
∂p ∂T αp
Heat capacity C is a measure of the amount of heat = − p = ,
needed to raise the temperature of a system by a given ∂T V
∂V κT
∂p
T
amount.
so
In a reversible process we have
α2p
Cp = CV + V T .
∆Q = T ∆S. κT

12
Thus, Cp > CV . 2.9. Stability conditions of matter
In a steady equilibrium the entropy has the true
2.8. Thermodynamical equilibrium state maximum so that small variations can only reduce the
According to the 2nd law, the entropy of an isolated entropy.
equilibrium system must be at maximum. Thus, any local We again consider an isolated system divided into parts,
fluctuation must cause the entropy to decrease; if it were and denote the equilibrium values common for all
not so, the system could move to a new higher entropy fictitious parts by the symbols T , p and {µj } and the
state, which cannot happen in an equilibrium system by equilibrium values of other variables by the superscript 0 .
definition. In order to study the maximality of the entropy, the
We divide the system into fictitious parts: entropy of the partial system α, Sα , needs to be expanded
into second order in {∆Uα , ∆Vα , ∆Njα }. We write Sα
close to an equilibrium as the Taylor series
a p , T ,
a a
V a , ... 0
Sα (Uα , Vα , {Njα }) = Sα0 (Uα0 , Vα0 , {Njα })
 0  0
∂S ∂S
D U = D V = D N i= 0 + ∆Uα + ∆Vα
∂U V,N ∂V U,N
Extensive variables satisfy X  ∂S 0
+ ∆Njα
∂Nj U,V
X
S = Sα j
α (  0  0
X 1 ∂Sα ∂Sα
V = Vα + ∆ ∆Uα + ∆ ∆Vα
2 ∂Uα V,N ∂Vα U,N
α
)
X  ∂Sα 0
X
U = Uα
+ ∆ ∆Njα
X
α
j
∂Njα U,V
Nj = Njα . +···.
α

Here ∆Uα = Uα − Uα0 and correspondingly for other


Since each element is in equlibrium the state variables are
quantities. The variations of partial derivatives stand for
defined in each element, e.g.
 0
Sα = Sα (Uα , Vα , {Njα }) ∂Sα
∆ =
∂Uα V,N
and  2 0 "   #0
1 pα µjα ∂ S ∂ ∂S
∆Sα = ∆Uα + ∆Vα − ∆Njα . ∆Uα + ∆Vα
Tα Tα Tα ∂U 2 V,N ∂V ∂U V,N
U,N
Let us assume that the system is composed of two parts: "
X ∂  ∂S 
#0
α ∈ {A, B}. Then + ∆Njα
j
∂Nj ∂U V,N
U,V
∆UB = −∆UA , ∆VB = −∆VA and ∆NjB = −∆NjA
and similarly for other partial derivatives.
so The
P 1st order differentials drop out, because
α ∆Uα = 0. Thus,
X
∆S = ∆Sα
α ∆Sα =
   
1 1 pA pB (  0  0
= − ∆UA + − ∆VA 1 ∂Sα ∂Sα
TA TB TA TB ∆ ∆Uα + ∆ ∆Vα
X  µjA  2 ∂Uα V,N ∂Vα U,N
µjB
− − ∆NjA . X  ∂Sα 0
)
T A T B
j + ∆ ∆Njα .
j
∂Njα U,V
In an equilibrium S is at its maximum, so ∆S = 0 and
This can be rewritten as
TA = TB
pA = pB ∆Sα =
(    
µjA = µjB . 1 1 pα
∆ ∆Uα + ∆ ∆Vα
2 Tα Tα
This is valid also when the system consists of several )
phases. X  µjα 
− ∆ ∆Njα .
j

13
Using the first law we get 3. Applications of thermodynamics
(
∆S =
1 X
−∆Tα ∆Sα + ∆pα ∆Vα
3.1. Classical ideal gas
2T α For a full description of the thermodynamics of a system
) we need to know both the equation of state and some

X
∆µjα ∆Njα . thermodynamic potential. EOS gives us mechanical
j
response functions, but for thermal response functions we
need also some potential.
This can be further written as From the ideal gas equation of state
(
1 X CV 1 pV = N kB T
∆S = − (∆Tα )2 + [(∆Vα )2Nα ]
2T α T κT V we obtain directly the mechanical response functions
 0 )
∂µ
 
2 1 ∂V N kB 1
+ (∆Nα ) , αp = = =
∂N p,T V ∂T p,N Vp T
 
1 ∂V N kB T 1
where κT = − = = .
V ∂p T,N V p2 p
 0  0
∂V ∂V
(∆Vα )Nα = ∆Tα + ∆pα . Thermal response functions cannot be derived from the
∂T N,p ∂p N,T equation of state. Empirically it has been observed
Since ∆S ≤ 0, we must have 1
CV = f kB N.
2
∂µ
CV ≥ 0, κT ≥ 0, ≥ 0. Here 12 f kB is the specific heat capacity/molecule and f is
∂N
the number of degrees of freedom of the molecule.
The condition CV ≥ 0 is a condition for thermal stability: Atoms/molecule f translations rotations
if a small excess of heat energy is added to a volume 1 3 3 0
element of fluid, the temperature of the volume element 2 5 3 2
must increase. polyatomic 6 3 3
The condition κT ≥ 0 is a condition for mechanical For real gases f = f (T, p), which is different from ideal
stability: if a volume of a small fluid element (fixed N ) gas because of internal degrees of freedom (vibrations),
increases, the pressure must go down, so that the larger interactions between the molecules and quantum
pressure from the environment halts and reverses the mechanical effects.
growth.
∂µ Entropy
Likewise ∂N ≥ 0 is a condition for chemical stability.
Entropy can be obtained from thermal and mechanical
Recall that
response functions by integrating along the trajectory
V
   2 
∂S ∂ F
CV = T = −T >0
∂T V,N ∂T 2 V,N V S
S 0
Thus, F is a concave function of T . V 0

T
Similarly,
T 0 T
   2 
1 ∂p ∂ F The differential is
= −V =V
κT ∂V T,N ∂V 2 T,N    
∂S ∂S
dS = dT + dV
and F is a convex function of V . ∂T V ∂V T
 
1 ∂p
= CV dT + dV,
T ∂T V

since according to Maxwell relations


   
∂S ∂p
= .
∂V T ∂T V
Integrating we get
Z V T
CV N kB
Z
S = S0 + + dV dT
T0 T V0 V
T V
= S0 + CV ln + N kB ln
T0 V0

14
or " f /2 # In the process V1 → V2 and ∆Q = ∆W = 0, so ∆U = 0.
T V Process is irreversible.
S = S0 + N kB ln .
T0 V0 a) Ideal gas
Note: A contradiction with the 3rd law: S → −∞, when Now
1
T → 0. 3rd law relies on the quantum nature of real U= f kB T N,
2
matter!
so T1 = T2 , because U1 = U2 . The cange in the entropy is
thus
Internal energy V2
We substitute into the first law (N const.) ∆S = N kB ln .
V1
dU = T dS − p dV
b) Real gas equation of state
the differential The internal energy and the number of particles are
    constant:
∂S ∂S
dS = dT + dV,
   
∂U ∂U
∂T V ∂V T dU = dT + dV = 0 .
∂T V ∂V T
and get ∂T

The Joule coefficient ∂V U,N
characterizes the behaviour
of the gas during free expansion (cf. Joule-Thompson
   
∂S
dU = CV dT + T − p dV. coefficient):
∂V T
∂U
  
According to Maxwell relations and to the equation of ∂T ∂V T
= − ∂U 
state we have ∂V U,N ∂T V
 

∂S
 
∂p

N kB p 1 αp
= = = , = p−T .
∂V T ∂T V V T CV κT

so Note that this is differential form (∂V ); with a finite


dU = CV dT process one must integrate over differential changes.

and 3.3. Mixing entropy


1 Conside different gases A and B, separated by a partition:
U = U0 + CV (T − T0 ) = U0 + f kB N (T − T0 ).
2 A B
If we choose U0 = CV T0 , we get for the internal energy T A T B
p A p B
1
U= f N kB T.
2
We assume that initially pA = pB = p and TA = TB = T .
Now The partition is removed and the gases mix. The process
α2p
Cp = CV + V T is irreversible is adiabatic so ∆Q = 0.
κT In a mixture of ideal gases every component satisfies the
or   state equation
1
Cp = N kB f +1 pj V = Nj kB T.
2
or The concentration of the component j is
Cp = γCV , Nj pj
xj = = ,
where γ is the adiabatic constant N p
where the total pressure p is
γ = Cp /CV = (f + 2)/f. X
p= pj .
j
3.2. Free expansion of gas Method 1:
Each constituent gas expands to volume V . Since
V 1 D Q = D W = 0 pA = pB and TA = TB , we have Vj = V xj . The change in
the entropy is (see the free expansion of a gas)
X V
∆S = Nj kB ln
j
Vj

15
or X Example:
∆Smix = −N kB xj ln xj .
j j A B C D
Mj H2 S O2 H2 O SO2
Now ∆Smix ≥ 0, since 0 ≤ xj ≤ 1.
νj −2 −3 2 2
Method 2:
For a process taking place in constant pressure and We define the degree of reaction ξ so that
temperature the Gibbs function is the suitable potential:
dNj = νj dξ.
G = U − T S + pV
1 When ξ increments by one, one reaction of the reaction
= f kB T N − T S + pV = · · ·
2 formula from left to right takes place.
= N kB T [φ(T ) + ln p] = N µ(p, T ), Convention: When ξ = 0 the reaction is as far left as it
can be. Then
where ξ ≥ 0.
µ0 f
φ(T ) = − ξ − ( + 1) ln T. Let us assume that p and T remain constant during the
kB T 2
reaction. Then a suitable potential is the Gibbs function
Before mixing
X
X G= µj Nj .
G(b) = Nj kB T [φj (T ) + ln p] j
j
Its differential is
and after mixing X X
X dG = µj dNj = dξ νj µj .
G(a) = Nj kB T [φj (T ) + ln pj ], j j
j
We define
so the change in the Gibbs function is  
∂G X
X pj ∆r G ≡ = νj µj .
∆G(mix) = G(a) − G(b) = Nj kB T ln ∂ξ p,T
j
p j

∆r G is thus the change in the Gibbs function per one


X
= Nj kB T ln xj .
j
reaction (often called the affinity).
Since (p, T ) is constant G has a minimum at equilibrium.
Because   The equilibrium condition is thus
∂G
S=− ,
νj µeq
X
∂T P,{Nj } ∆r Geq = j = 0.
j
we get for the mixing entropy
X In a nonequilibrium dG/dt < 0, so if ∆r G > 0 we must
∆Smix = S(a) − S(b) = − Nj kB ln xj . have dξ/dt < 0, i.e. the reaction proceeds to left and vice
j
versa. Let us assume that the components obey the ideal
Gibbs’ paradox: If A ≡ B, i.e. the gases are identical no gas equation of state. Then
changes take place in the process. However, according to
µj = kB T [φj (T ) + ln p + ln xj ],
the former discussion, ∆S > 0. The reason is that in
classical ensemble the particles are distinguishable, and where pj = xj p is the partial pressure of component j and
mixing of A and B really happens. In quantum mechanics
this apparent contradiction is removed by employing µ0j 1
quantum statitics of identical particles. φ j (T ) = − ηj − (1 + fj ) ln T.
kB T 2

3.4. Chemical reactions So


Consider for example the chemical reaction X  P Y
ν

∆r G = kB T νj φj (T ) + kB T ln p νj xj j .

2 H2 S + 3 O2 ← 2 H2 O + 2 SO2 . j

In general the chemical reaction formula is written as The equilibrium condition can now be written as
P
− νj
X Y ν
0= νj Mj . xj j = p j K(T ),
j j

Here νj ∈ I are the stochiometric coefficients and Mj where P


− νj φj (T )
stand for the molecular species. K(T ) = e j

16
is the equilibrium constant of the reaction, which depends an intensive quantity it depends only on relative
only on T . The equilibrium condition is historically called fractions, so
the law of mass action.
For the reaction above µjα = µjα (p, T, x1α , . . . , xH−1,α ),

x2C x2D and the conditions (∗) take the form


= pK(T ).
x2A x3C
µ1A (p, T, x1A , . . . , xH−1,A ) =
The heat of reaction is the change of heat energy ∆r Q per µ1B (p, T, x1B , . . . , xH−1,B )
one reaction to right. A reaction is ..
.
• Endothermic, if ∆r Q > 0 i.e. the reaction takes heat.
µHA (p, T, x1A , . . . , xH−1,A ) =
• Exothermic, if ∆r Q < 0 i.e. the reaction releases µHB (p, T, x1B , . . . , xH−1,B ).
heat.
There are
We write ∆r G as
X • M = (H − 1)F + 2 variables,
∆r G = −kB T ln K(T ) + kB T νj ln pxj .
• Y = H(F − 1) equations.
j
The simultaneous equations can have a solution only if
Now
M ≥ Y or
∆Q = ∆U + ∆W = ∆U + p ∆V = ∆(U + pV ) F ≤ H + 2.
= ∆H, This condition is know as the Gibbs phase rule.
For pure matter the equilibrium condition
since ∆p = 0. When the total amount matter is constant
µA (p, T ) = µB (p, T )
dG = −S dT + V dp
defines in the (p, T )-plane a coexistence curve. If the
holds in a reversible process and phase B is in equilibrium with the phase C we get
    another curve
G 1 G G S V µB (p, T ) = µC (p, T ).
d = dG − 2 dT = − 2
+ dT + dp
T T T T T T
The phases A, B can C can be simultaneously in
H V equilibrium in a crossing point, so called triple point, of
= − 2 dT + dp,
T T these curves.
because G = H − T S. We see that
   3.6. Phase transitions
2 ∂ G In a phase transition the chemical potential
H = −T .
∂T T p,N
 
∂G
µ=
Now   ∂N p,T
∂ ∆r G d
= −kB ln K(T ), is continuous. Instead
∂T T dT
 
so that ∂G
d S=−
∆r H = kB T 2 ln K(T ). ∂T p
dT
This expression is known as the heat of reaction. and  
∂G
V =
3.5. Phase equilibrium ∂p T
In a system consisting of several phases the equilibrium are not necessarily continuous.
conditions for each pair (A and B) of phases are A transition is of first order, if the first order derivatives
(S, V ) of G are discontinuous, and of second order, if the
TA = TB = T 1st order derivatives are continuous but 2nd order
pA = pB = p discontinuous. Otherwise the transition is continuous.
µjA = µjB , j = 1, . . . , H, (∗) T
m 1= m 2
where H is the number of particle species in the system. 1
Let us assume that the number of phases is F , so for each 2
species there are F − 1 independent conditions (∗). Now
µiα = µiα (p, T, {Njα }). Because the chemical potential is p

17
In a first order transition from a phase 1 to a phase 2 so that
 (2)  (1) −S1 dT + V1 dp = −S2 dT + V2 dp
∂G ∂G
∆S = − + or on the curve
∂T p ∂T p
(2)  (1) dp S2 − S1 ∆S T −1 ∆H

∂G ∂G = = =
∆V = − . dT V2 − V1 ∆V ∆V
∂p T ∂p T
and we end up with the Clausius-Clapeyron equation
When we cross a coexistence curve p and T stay constant,  
so dp 1 ∆H
= .
dT coex T ∆V
∆Q = T ∆S = ∆U + p ∆V = ∆(U + pV )
= ∆H. Here ∆H = H2 − H1 and ∆V = V2 − V1 .

∆Q is called the phase transition heat or the latent heat. Examples


Note: First order transitions have non-zero latent heat a) Vapour pressure curve
but not the higher order ones. p fu s io n c u rv e
Typical phase diagram of solid-liquid-gas -system: lines c ritic a l
p o in t
are 1st order transitions, critical point is 2nd order. C
(triple point is 1st order). s o lid flu id
s u b lim a tio n v a p o r
Water: pT = 610Pa, TT = 0.01◦C; T
c u rv e v a p o r p re ssu re
pc = 22 MPa, Tc = 374.15◦C. trip le p o in t c u rv e
T
p
We consider the transition
solid
critical liquid → vapour.
point
liquid
Assuming ideal gas we have
triple
point N kB T
∆V = Vv = ,
gas
T p
because Vl(iquid) ≪ Vv(apor) , and
p
 
dp ∆Hlv p
critical = .
solid point dT coex N kB T 2
liquid const T isotherm
If the vapourization heat (the latent heat) ∆Hlv is
gas roughly constant on the vapour pressure curve we can
triple
integrate
point V p = p0 e−∆Hlv /N kB T .
coexistence (this assumption is not true near crit. point!)
Because of phase coexistence, phase diagrams are simplest b) Fusion curve
~ . . .).
in “force”-type coordinates (T, p, µ, H, Now
∆Vls = Vl(iquid) − Vs(olid)
3.7. Phase coexistence
c o e x is te n c e
can be positive or negative (for example H2 O).
T
c u rv e
According to the Clausius-Clapeyron equation

1
dp ∆Hls
=
2 dT T ∆Vls

p we have
dp
On the coexistence curve dT > 0, if ∆Vls > 0 1)

dp
G1 (p, T, N ) = G2 (p, T, N ) dT < 0, if ∆Vls < 0 2) .

p d p p d p
< 0
and
> 0
dG = −S dT + V dp
1 ) s o lid d T 2 ) d T
flu id flu id
when the number of particles N is constant. Along the s o lid
curve

G1 (p + dp, T + dT, N ) = G2 (p + dp, T + dT, N ), T T

18
p
is o th e rm
We see that when the pressure is increased in constant
temperature the system
B I A
1) drifts ”deeper” into the solid phase,
II
2) can go from the solid phase to the liquid phase.
V
c) Sublimation curve In many equations of state the phase transition happens
Now when there is apparent instability dp/dV > 0 (for
dH = T dS + V dp = Cp dT + V (1 − T αp ) dp, example, van der Waals). In this case, we can use
Maxwell’s construction: The points A and B have to be
because S = S(p, T ) and using Maxwell relations and chosen so that the area I = area II.
definitions of thermodynamic response functions
     
∂S ∂S ∂V Cp
dS = dp + dT = − dp + dT.
∂p T ∂T p ∂T p T
The vapour pressure is small so dp ≈ 0, and
Z T
0
Hs = H s + Cps dT solid phase
0
Z T
Hv = Hv0 + Cpv dT vapour (gas).
0

Let us suppose that the vapour satisfies the ideal gas


state equation. Then
N kB T N kB T
∆Vvs = − Vs ≈ ,
p p
so
dp ∆Hvs p ∆Hvs
= ≈ ,
dT T ∆Vvs N kB T 2
where ∆Hvs = Hs − Hv . For a monatomic ideal gas
Cp = 25 kB N , and
0 Z RT s ′
p ∆Hvs 5 1 0 Cp dT
ln =− + ln T − dT +constant.
p0 N kB T 2 kB N T2
0
Here ∆Hvs is the sublimation heat at vanishing
temperature and pressure.

Coexistence range
p is o th e rm s
C
u n d e rc o o le d

B A
V
o v e rh e a te d

Matter is mechanically stable only if dVdp < 0. Thus the


range of stability lies outside of the points A and B.
Overheated liquid and undercooled vapour are metastable
(supercooling, -heating).
According to the Gibbs-Duheim relation (consider dG!)
S V
dµ = − dT + dp
N N
we have along isotherms
V
dp.
dµ =
N
Thus, when the phases A and B are in equilibrium,
Z B
V
µA − µB = dp = 0.
A N

19

You might also like