If there exists a relation f(x,y,z)=0 between 3 variables, one can write: x=x(y,z), y=y(x,z) and z=z(x,y). The total differential dz of z is than given by:
dz=(∂z∂x)ydx+(∂z∂y)xdy
By writing this also for dx and dy it can be obtained that
(∂x∂y)z⋅(∂y∂z)x⋅(∂z∂x)y=−1
Because dz is a total differential ∮dz=0.
A homogeneous function of degree m obeys: εmF(x,y,z)=F(εx,εy,εz). For such a function Euler’s theorem applies:
mF(x,y,z)=x∂F∂x+y∂F∂y+z∂F∂z
For an ideal gas it follows that: γp=1/T, κT=1/p and βV=−1/V.
For an ideal gas holds: Cmp−CmV=R. Further, if the temperature is high enough to thermalize all internal rotational and vibrational degrees of freedom: CV=12sR. Hence Cp=12(s+2)R. From their ratio it now follows that γ=(2+s)/s. For a lower T one needs only to consider the thermalized degrees of freedom. For a Van der Waals gas: CmV=12sR+ap/RT2.
In general holds:
Cp−CV=T(∂p∂T)V⋅(∂V∂T)p=−T(∂V∂T)2p(∂p∂V)T≥0
Because (∂p/∂V)T is always <0, the following is always valid: Cp≥CV. If the coefficient of expansion is 0, Cp=CV, and this is true also at T=0K.
The zeroth law states that heat flows from higher to lower temperatures. The first law is the conservation of energy. For a closed system: Q=ΔU+W, where Q is the total added heat, W the work done and ΔU the difference in the internal energy. In differential form this becomes: đQ=dU+đW, where đ means that the it is not a differential of a state function. For a quasi-static process: đW=pdV. So for a reversible process: đQ=dU+pdV.
For an open (flowing) system the first law is: Q=ΔH+Wi+ΔEkin+ΔEpot. One can extract an amount of work Wt from the system or add Wt=−Wi to the system.
The second law states: for a closed system there exists an additive quantity S, called the entropy, the differential of which has the following property:
dS≥đQT
If the only processes occurring are reversible: dS=đQrev/T. So, the entropy difference after a reversible process is:
S2−S1=2∫1đQrevT
So, for a reversible cycle: ∮đQrevT=0.
For an irreversible cycle: ∮đQirrT<0.
The third law of thermodynamics is (Nernst's law):
lim
From this it can be concluded that the thermal heat capacity \rightarrow0 if T\rightarrow0, so absolute zero temperature cannot be reached by cooling through a finite number of steps.
The state functions and their differentials are:
Internal energy: | U | dU=TdS-pdV |
Enthalpy: | H=U+pV | dH=TdS+Vdp |
Free energy: | F=U-TS | dF=-SdT-pdV |
Gibbs free energy: | G=H-TS | dG=-SdT+Vdp |
From this one can derive Maxwell’s relations:
\left(\frac{\partial T}{\partial V}\right)_{S}=-\left(\frac{\partial p}{\partial S}\right)_{V}~,~~\left(\frac{\partial T}{\partial p}\right)_{S}=\left(\frac{\partial V}{\partial S}\right)_{p}~,~~ \left(\frac{\partial p}{\partial T}\right)_{V}=\left(\frac{\partial S}{\partial V}\right)_{T}~,~~\left(\frac{\partial V}{\partial T}\right)_{p}=-\left(\frac{\partial S}{\partial p}\right)_{T}
From the total differential and the definitions of C_V and C_p it can be derived that:
TdS=C_VdT+T\left(\frac{\partial p}{\partial T}\right)_{V}dV~~\mbox{and}~~TdS=C_pdT-T\left(\frac{\partial V}{\partial T}\right)_{p}dp
For an ideal gas:
S_m=C_V\ln\left(\frac{T}{T_0}\right)+R\ln\left(\frac{V}{V_0}\right)+S_{m0}~~\mbox{and}~~ S_m=C_p\ln\left(\frac{T}{T_0}\right)-R\ln\left(\frac{p}{p_0}\right)+S_{m0}'
Helmholtz’ equations are:
\left(\frac{\partial U}{\partial V}\right)_{T}=T\left(\frac{\partial p}{\partial T}\right)_{V}-p~~,~~\left(\frac{\partial H}{\partial p}\right)_{T}=V-T\left(\frac{\partial V}{\partial T}\right)_{p}
For a macroscopic surface: đW_{\rm rev}=-\gamma dA, with \gamma the surface tension. From this follows:
\gamma=\left(\frac{\partial U}{\partial A}\right)_{S}=\left(\frac{\partial F}{\partial A}\right)_{T}
The efficiency \eta of a process is given by: \displaystyle\eta=\frac{\mbox{Work done}}{\mbox{Heat added}}
The Cold factor \xi of a cooling down process is given by: \displaystyle\xi=\frac{\mbox{Cold delivered}}{\mbox{Work added}}
For adiabatic processes: W=U_1-U_2. For reversible adiabatic processes Poisson’s equation holds. With \gamma=C_p/C_V one gets that pV^\gamma=constant. Also: TV^{\gamma-1}=constant and T^\gamma p^{1-\gamma}=constant. Adiabats are steeper on a p-V diagram than isotherms because \gamma>1.
Here: H_2-H_1=\int_1^2 C_pdT. For a reversible isobaric process: H_2-H_1=Q_{\rm rev}.
This is also called the Joule-Kelvin effect and is an adiabatic expansion of a gas through a porous material or a small opening. Here H is a conserved quantity, and dS>0. In general this is accompanied with a change in temperature. The quantity which is important here is the throttle coefficient:
\alpha_H=\left(\frac{\partial T}{\partial p}\right)_{H}=\frac{1}{C_p}\left[T\left(\frac{\partial V}{\partial T}\right)_{p}-V\right]
The inversion temperature is the temperature where an adiabatically expanding gas keeps the same temperature. If T>T_{\rm i} the gas heats up, if T<T_{\rm i} the gas cools down. T_{\rm i}=2T_{\rm B}, with for T_{\rm B}: [\partial(pV)/\partial p]_T=0. The throttle process is, for example, applied in refrigerators.
The system undergoes a reversible cycle with 2 isothemics and 2 adiabats:
The efficiency for a Carnot cycle is:
\eta=1-\frac{|Q_2|}{|Q_1|}=1-\frac{T_2}{T_1}:=\eta_{\rm C}
The Carnot efficiency \eta_{\rm C} is the maximal efficiency at which a heat machine can operate. If the process is applied in reverse order and the system performs a work -W the cold factor is given by:
\xi=\frac{|Q_2|}{W}=\frac{|Q_2|}{|Q_1|-|Q_2|}=\frac{T_2}{T_1-T_2}
The Stirling cycle
Stirling’s cycle consists of 2 isotherms and 2 isochorics. The efficiency in the ideal case is the same as for a Carnot cycle.
Consider a system that changes from state 1 into state 2, with the temperature and pressure of the surroundings given by T_0 and p_0. The maximum work which can be obtained from this change is, when all processes are reversible:
The minimal work needed to attain a certain state is: W_{\rm min}=-W_{\rm max}.
Phase transitions are isothermic and isobaric, so dG=0. When the phases are indicated by \alpha, \beta and \gamma: G_m^\alpha=G_m^\beta and
\Delta S_m=S_m^\alpha - S_m^\beta=\frac{r_{\beta\alpha}}{T_0}
where r_{\beta\alpha} is the transition heat of phase \beta to phase \alpha and T_0 is the transition temperature. The following holds: r_{\beta\alpha}=r_{\alpha\beta} and r_{\beta\alpha}=r_{\gamma\alpha}-r_{\gamma\beta}. Further
S_m=\left(\frac{\partial G_m}{\partial T}\right)_{p}
so G has a kink in the transition point and the derivative is discontinuous. In a two phase system Clapeyron’s equation is valid:
\frac{dp}{dT}=\frac{S_m^\alpha-S_m^\beta}{V_m^\alpha-V_m^\beta}= \frac{r_{\beta\alpha}}{(V_m^\alpha-V_m^\beta)T}
For an ideal gas one finds for the vapor line at some distance from the critical point:
p=p_0{\rm e}^{-r_{\beta\alpha/RT}}
There exist also phase transitions with r_{\beta\alpha}=0. For those there will occur only be a discontinuity in the second derivatives of G_m. These second-order transitions appear at organization phenomena.
A phase-change of the 3rd order, so with e.g. [\partial^3 G_m/\partial T^3]_p non continuous arises e.g. when ferromagnetic iron changes to the paramagnetic state.
When the number of particles within a system changes this number becomes a third quantity of state. Because addition of matter usually takes place at constant p and T, G is the relevant quantity. If a system has many components this becomes:
dG=-SdT+Vdp+\sum_i\mu_idn_i where \displaystyle\mu=\left(\frac{\partial G}{\partial n_i}\right)_{p,T,n_j}
is called the thermodynamic potential. This is a partial quantity. For V:
V=\sum_{i=1}^c n_i\left(\frac{\partial V}{\partial n_i}\right)_{n_j,p,T}:=\sum_{i=1}^c n_i V_i
where V_i is the partial volume of component i. The following holds:
\begin{aligned} V_m&=&\sum_i x_i V_i\\ 0&=&\sum_i x_i dV_i\end{aligned}
where x_i=n_i/n is the molar fraction of component i. The molar volume of a mixture of two components can be a concave line in a V-x_2 diagram: the mixing leads to a contraction of the volume
The thermodynamic potentials are not independent in a multiple-phase system. It can be derived that \sum\limits_i n_i d\mu_i=-SdT+Vdp, this gives at constant p and T: \sum\limits_i x_i d\mu_i=0 (Gibbs-Duhmen).
Each component has as many \mu’s as there are phases. The number of free parameters in a system with c components and p different phases is given by f=c+2-p which is called Gibb's phase rule. .
For a mixture of n components (the index ^0 is the value for the pure component):
U_{\rm mixture}=\sum_i n_i U^0_i~~,~~H_{\rm mixture}=\sum_i n_i H^0_i~~,~~ S_{\rm mixture}=n\sum_i x_i S^0_i+\Delta S_{\rm mix}
where for ideal gases: \Delta S_{\rm mix}=-nR\sum\limits_i x_i\ln(x_i).
For the thermodynamic potentials: \mu_i=\mu_i^0+RT\ln(x_i)<\mu_i^0. A mixture of two liquids is rarely ideal: this is usually only the case for chemically related components or isotopes. In spite of this Raoult’s law holds for the vapour pressure for many binary mixtures: p_i=x_ip^0_i=y_ip. Here x_i is the fraction of the ith component in liquid phase and y_i the fraction of the ith component in gas phase.
A solution for one component in a second gives rise to an increase in the boiling point \Delta T_{\rm k} and a decrease of the freezing point \Delta T_{\rm s}. For x_2\ll1:
\Delta T_{\rm k}=\frac{RT_{\rm k}^2}{r_{\beta\alpha}}x_2~~,~~ \Delta T_{\rm s}=-\frac{RT_{\rm s}^2}{r_{\gamma\beta}}x_2
with r_{\beta\alpha} the heat of evaporation and r_{\gamma\beta}<0 the melting heat. For the osmotic pressure \Pi of a solution: \Pi V_{m1}^0=x_2RT.
These are called collegative properties
When a system evolves towards equilibrium the only changes that are possible are those for which: (dS)_{U,V}\geq0 or (dU)_{S,V}\leq0 or (dH)_{S,p}\leq0 or (dF)_{T,V}\leq0 or (dG)_{T,p}\leq0. In equilibrium for each component: \mu_i^\alpha=\mu_i^\beta=\mu_i^\gamma.
The number of possibilities P to distribute N particles on n possible energy levels, each with a g-fold degeneracy is called the thermodynamic probability and is given by:
P=N!\prod_i\frac{g_i^{n_i}}{n_i!}
The most probable distribution, that with the maximum value for P, is the equilibrium state. When Stirling’s equation, \ln(n!)\approx n\ln(n)-n is used, one finds for a discrete system the Maxwell-Boltzmann distribution. The occupation numbers in equilibrium are then given by:
n_i=\frac{N}{Z}g_i\exp\left(-\frac{W_i}{kT}\right)
The state sum Z is a normalization constant, given by: Z=\sum\limits_ig_i\exp(-W_i/kT). For an ideal gas:
Z=\frac{V(2\pi mkT)^{3/2}}{h^3}
The entropy can then be defined as: S=k\;ln\left ( P \right ) . For a system in thermodynamic equilibrium this becomes:
S=\frac{U}{T}+kN\ln\left(\frac{Z}{N}\right)+kN\approx\frac{U}{T}+k\ln\left(\frac{Z^N}{N!}\right)
For an ideal gas, with U=\frac{3}{2}kT then: \displaystyle S=kN+kN\ln\left(\frac{V(2\pi mkT)^{3/2}}{Nh^3}\right)
Thermodynamics can be applied to other systems than gases and liquids. To do this the term đW=pdV has to be replaced with the correct work term, like đW_{\rm rev}=-Fdl for the stretching of a wire, đW_{\rm rev}=-\gamma dA for the expansion of a soap bubble or đW_{\rm rev}=-BdM for a magnetic system.
A rotating, non-charged black hole has a temperature of T=\hbar c/8\pi km. It has an entropy S=Akc^3/4\hbar\kappa with A the area of its event horizon. For a Schwarzschild black hole A is given by A=16\pi m^2. Hawkings area theorem states that dA/dt\geq0.
Hence, the lifetime of a black hole \sim m^3.