Isothermal–isobaric ensemble

Ensemble of states at constant pressure
Statistical mechanics
Particle statistics
Thermodynamic ensembles
Models
  • Debye
  • Einstein
  • Ising
  • Potts
  • v
  • t
  • e

The isothermal–isobaric ensemble (constant temperature and constant pressure ensemble) is a statistical mechanical ensemble that maintains constant temperature T {\displaystyle T\,} and constant pressure P {\displaystyle P\,} applied. It is also called the N p T {\displaystyle NpT} -ensemble, where the number of particles N {\displaystyle N\,} is also kept as a constant. This ensemble plays an important role in chemistry as chemical reactions are usually carried out under constant pressure condition.[1] The NPT ensemble is also useful for measuring the equation of state of model systems whose virial expansion for pressure cannot be evaluated, or systems near first-order phase transitions.[2]

In the ensemble, the probability of a microstate i {\displaystyle i} is Z 1 e β ( E ( i ) + p V ( i ) ) {\displaystyle Z^{-1}e^{-\beta (E(i)+pV(i))}} , where Z {\displaystyle Z} is the partition function, E ( i ) {\displaystyle E(i)} is the internal energy of the system in microstate i {\displaystyle i} , and V ( i ) {\displaystyle V(i)} is the volume of the system in microstate i {\displaystyle i} .

The probability of a macrostate is Z 1 e β ( E + p V T S ) = Z 1 e β G {\displaystyle Z^{-1}e^{-\beta (E+pV-TS)}=Z^{-1}e^{-\beta G}} , where G {\displaystyle G} is the Gibbs free energy.

Derivation of key properties

The partition function for the N p T {\displaystyle NpT} -ensemble can be derived from statistical mechanics by beginning with a system of N {\displaystyle N} identical atoms described by a Hamiltonian of the form p 2 / 2 m + U ( r n ) {\displaystyle \mathbf {p} ^{2}/2m+U(\mathbf {r} ^{n})} and contained within a box of volume V = L 3 {\displaystyle V=L^{3}} . This system is described by the partition function of the canonical ensemble in 3 dimensions:

Z s y s ( N , V , T ) = 1 Λ 3 N N ! 0 L . . . 0 L d r N exp ( β U ( r N ) ) {\displaystyle Z^{sys}(N,V,T)={\frac {1}{\Lambda ^{3N}N!}}\int _{0}^{L}...\int _{0}^{L}d\mathbf {r} ^{N}\exp(-\beta U(\mathbf {r} ^{N}))} ,

where Λ = h 2 β / ( 2 π m ) {\displaystyle \Lambda ={\sqrt {h^{2}\beta /(2\pi m)}}} , the thermal de Broglie wavelength ( β = 1 / k B T {\displaystyle \beta =1/k_{B}T\,} and k B {\displaystyle k_{B}\,} is the Boltzmann constant), and the factor 1 / N ! {\displaystyle 1/N!} (which accounts for indistinguishability of particles) both ensure normalization of entropy in the quasi-classical limit.[2] It is convenient to adopt a new set of coordinates defined by L s i = r i {\displaystyle L\mathbf {s} _{i}=\mathbf {r} _{i}} such that the partition function becomes

Z s y s ( N , V , T ) = V N Λ 3 N N ! 0 1 . . . 0 1 d s N exp ( β U ( s N ) ) {\displaystyle Z^{sys}(N,V,T)={\frac {V^{N}}{\Lambda ^{3N}N!}}\int _{0}^{1}...\int _{0}^{1}d\mathbf {s} ^{N}\exp(-\beta U(\mathbf {s} ^{N}))} .

If this system is then brought into contact with a bath of volume V 0 {\displaystyle V_{0}} at constant temperature and pressure containing an ideal gas with total particle number M {\displaystyle M} such that M N N {\displaystyle M-N\gg N} , the partition function of the whole system is simply the product of the partition functions of the subsystems:

Z s y s + b a t h ( N , V , T ) = V N ( V 0 V ) M N Λ 3 M N ! ( M N ) ! d s M N d s N exp ( β U ( s N ) ) {\displaystyle Z^{sys+bath}(N,V,T)={\frac {V^{N}(V_{0}-V)^{M-N}}{\Lambda ^{3M}N!(M-N)!}}\int d\mathbf {s} ^{M-N}\int d\mathbf {s} ^{N}\exp(-\beta U(\mathbf {s} ^{N}))} .
The system (volume V {\displaystyle V} ) is immersed in a much larger bath of constant temperature, and closed off such that particle number remains fixed. The system is separated from the bath by a piston that is free to move, such that its volume can change.

The integral over the s M N {\displaystyle \mathbf {s} ^{M-N}} coordinates is simply 1 {\displaystyle 1} . In the limit that V 0 {\displaystyle V_{0}\rightarrow \infty } , M {\displaystyle M\rightarrow \infty } while ( M N ) / V 0 = ρ {\displaystyle (M-N)/V_{0}=\rho } stays constant, a change in volume of the system under study will not change the pressure p {\displaystyle p} of the whole system. Taking V / V 0 0 {\displaystyle V/V_{0}\rightarrow 0} allows for the approximation ( V 0 V ) M N = V 0 M N ( 1 V / V 0 ) M N V 0 M N exp ( ( M N ) V / V 0 ) {\displaystyle (V_{0}-V)^{M-N}=V_{0}^{M-N}(1-V/V_{0})^{M-N}\approx V_{0}^{M-N}\exp(-(M-N)V/V_{0})} . For an ideal gas, ( M N ) / V 0 = ρ = β P {\displaystyle (M-N)/V_{0}=\rho =\beta P} gives a relationship between density and pressure. Substituting this into the above expression for the partition function, multiplying by a factor β P {\displaystyle \beta P} (see below for justification for this step), and integrating over the volume V then gives

Δ s y s + b a t h ( N , P , T ) = β P V 0 M N Λ 3 M N ! ( M N ) ! d V V N exp ( β P V ) d s N exp ( β U ( s ) ) {\displaystyle \Delta ^{sys+bath}(N,P,T)={\frac {\beta PV_{0}^{M-N}}{\Lambda ^{3M}N!(M-N)!}}\int dVV^{N}\exp({-\beta PV})\int d\mathbf {s} ^{N}\exp(-\beta U(\mathbf {s} ))} .

The partition function for the bath is simply Δ b a t h = V 0 M N / [ ( M N ) ! Λ 3 ( M N ) {\displaystyle \Delta ^{bath}=V_{0}^{M-N}/[(M-N)!\Lambda ^{3(M-N)}} . Separating this term out of the overall expression gives the partition function for the N p T {\displaystyle NpT} -ensemble:

Δ s y s ( N , P , T ) = β P Λ 3 N N ! d V V N exp ( β P V ) d s N exp ( β U ( s ) ) {\displaystyle \Delta ^{sys}(N,P,T)={\frac {\beta P}{\Lambda ^{3N}N!}}\int dVV^{N}\exp(-\beta PV)\int d\mathbf {s} ^{N}\exp(-\beta U(\mathbf {s} ))} .

Using the above definition of Z s y s ( N , V , T ) {\displaystyle Z^{sys}(N,V,T)} , the partition function can be rewritten as

Δ s y s ( N , P , T ) = β P d V exp ( β P V ) Z s y s ( N , V , T ) {\displaystyle \Delta ^{sys}(N,P,T)=\beta P\int dV\exp(-\beta PV)Z^{sys}(N,V,T)} ,

which can be written more generally as a weighted sum over the partition function for the canonical ensemble

Δ ( N , P , T ) = Z ( N , V , T ) exp ( β P V ) C d V . {\displaystyle \Delta (N,P,T)=\int Z(N,V,T)\exp(-\beta PV)CdV.\,\;}

The quantity C {\displaystyle C} is simply some constant with units of inverse volume, which is necessary to make the integral dimensionless. In this case, C = β P {\displaystyle C=\beta P} , but in general it can take on multiple values. The ambiguity in its choice stems from the fact that volume is not a quantity that can be counted (unlike e.g. the number of particles), and so there is no “natural metric” for the final volume integration performed in the above derivation.[2] This problem has been addressed in multiple ways by various authors,[3][4] leading to values for C with the same units of inverse volume. The differences vanish (i.e. the choice of C {\displaystyle C} becomes arbitrary) in the thermodynamic limit, where the number of particles goes to infinity.[5]

The N p T {\displaystyle NpT} -ensemble can also be viewed as a special case of the Gibbs canonical ensemble, in which the macrostates of the system are defined according to external temperature T {\displaystyle T} and external forces acting on the system J {\displaystyle \mathbf {J} } . Consider such a system containing N {\displaystyle N} particles. The Hamiltonian of the system is then given by H J x {\displaystyle {\mathcal {H}}-\mathbf {J} \cdot \mathbf {x} } where H {\displaystyle {\mathcal {H}}} is the system's Hamiltonian in the absence of external forces and x {\displaystyle \mathbf {x} } are the conjugate variables of J {\displaystyle \mathbf {J} } . The microstates μ {\displaystyle \mu } of the system then occur with probability defined by [6]

p ( μ , x ) = exp [ β H ( μ ) + β J x ] / Z {\displaystyle p(\mu ,\mathbf {x} )=\exp[-\beta {\mathcal {H}}(\mu )+\beta \mathbf {J} \cdot \mathbf {x} ]/{\mathcal {Z}}}

where the normalization factor Z {\displaystyle {\mathcal {Z}}} is defined by

Z ( N , J , T ) = μ , x exp [ β J x β H ( μ ) ] {\displaystyle {\mathcal {Z}}(N,\mathbf {J} ,T)=\sum _{\mu ,\mathbf {x} }\exp[\beta \mathbf {J} \cdot \mathbf {x} -\beta {\mathcal {H}}(\mu )]} .

This distribution is called generalized Boltzmann distribution by some authors.[7]

The N p T {\displaystyle NpT} -ensemble can be found by taking J = P {\displaystyle \mathbf {J} =-P} and x = V {\displaystyle \mathbf {x} =V} . Then the normalization factor becomes

Z ( N , J , T ) = μ , { r i } V exp [ β P V β ( p 2 / 2 m + U ( r N ) ) ] {\displaystyle {\mathcal {Z}}(N,\mathbf {J} ,T)=\sum _{\mu ,\{\mathbf {r} _{i}\}\in V}\exp[-\beta PV-\beta (\mathbf {p} ^{2}/2m+U(\mathbf {r} ^{N}))]} ,

where the Hamiltonian has been written in terms of the particle momenta p i {\displaystyle \mathbf {p} _{i}} and positions r i {\displaystyle \mathbf {r} _{i}} . This sum can be taken to an integral over both V {\displaystyle V} and the microstates μ {\displaystyle \mu } . The measure for the latter integral is the standard measure of phase space for identical particles: d Γ N = 1 h 3 N ! i = 1 N d 3 p i d 3 r i {\displaystyle {\textrm {d}}\Gamma _{N}={\frac {1}{h^{3}N!}}\prod _{i=1}^{N}d^{3}\mathbf {p} _{i}d^{3}\mathbf {r} _{i}} .[6] The integral over exp ( β p 2 / 2 m ) {\displaystyle \exp(-\beta \mathbf {p} ^{2}/2m)} term is a Gaussian integral, and can be evaluated explicitly as

i = 1 N d 3 p i h 3 exp [ β i = 1 N p i 2 2 m ] = 1 Λ 3 N {\displaystyle \int \prod _{i=1}^{N}{\frac {d^{3}\mathbf {p} _{i}}{h^{3}}}\exp {\bigg [}-\beta \sum _{i=1}^{N}{\frac {p_{i}^{2}}{2m}}{\bigg ]}={\frac {1}{\Lambda ^{3N}}}} .

Inserting this result into Z ( N , P , T ) {\displaystyle {\mathcal {Z}}(N,P,T)} gives a familiar expression:

Z ( N , P , T ) = 1 Λ 3 N N ! d V exp ( β P V ) d r N exp ( β U ( r ) ) = d V exp ( β P V ) Z ( N , V , T ) {\displaystyle {\mathcal {Z}}(N,P,T)={\frac {1}{\Lambda ^{3N}N!}}\int dV\exp(-\beta PV)\int d\mathbf {r} ^{N}\exp(-\beta U(\mathbf {r} ))=\int dV\exp(-\beta PV)Z(N,V,T)} .[6]

This is almost the partition function for the N p T {\displaystyle NpT} -ensemble, but it has units of volume, an unavoidable consequence of taking the above sum over volumes into an integral. Restoring the constant C {\displaystyle C} yields the proper result for Δ ( N , P , T ) {\displaystyle \Delta (N,P,T)} .

From the preceding analysis it is clear that the characteristic state function of this ensemble is the Gibbs free energy,

G ( N , P , T ) = k B T ln Δ ( N , P , T ) {\displaystyle G(N,P,T)=-k_{B}T\ln \Delta (N,P,T)\;\,}

This thermodynamic potential is related to the Helmholtz free energy (logarithm of the canonical partition function), F {\displaystyle F\,} , in the following way:[1]

G = F + P V . {\displaystyle G=F+PV.\;\,}

Applications

  • Constant-pressure simulations are useful for determining the equation of state of a pure system. Monte Carlo simulations using the N p T {\displaystyle NpT} -ensemble are particularly useful for determining the equation of state of fluids at pressures of around 1 atm, where they can achieve accurate results with much less computational time than other ensembles.[2]
  • Zero-pressure N p T {\displaystyle NpT} -ensemble simulations provide a quick way of estimating vapor-liquid coexistence curves in mixed-phase systems.[2]
  • N p T {\displaystyle NpT} -ensemble Monte Carlo simulations have been applied to study the excess properties[8] and equations of state [9] of various models of fluid mixtures.
  • The N p T {\displaystyle NpT} -ensemble is also useful in molecular dynamics simulations, e.g. to model the behavior of water at ambient conditions.[10]

References

  1. ^ a b Dill, Ken A.; Bromberg, Sarina; Stigter, Dirk (2003). Molecular Driving Forces. New York: Garland Science.
  2. ^ a b c d e Frenkel, Daan.; Smit, Berend (2002). Understanding Molecular Simluation. New York: Academic Press.
  3. ^ Attard, Phil (1995). "On the density of volume states in the isobaric ensemble". Journal of Chemical Physics. 103 (24): 9884–9885. Bibcode:1995JChPh.103.9884A. doi:10.1063/1.469956.
  4. ^ Koper, Ger J. M.; Reiss, Howard (1996). "Length Scale for the Constant Pressure Ensemble: Application to Small Systems and Relation to Einstein Fluctuation Theory". Journal of Physical Chemistry. 100 (1): 422–432. doi:10.1021/jp951819f.
  5. ^ Hill, Terrence (1987). Statistical Mechanics: Principles and Selected Applications. New York: Dover.
  6. ^ a b c Kardar, Mehran (2007). Statistical Physics of Particles. New York: Cambridge University Press.
  7. ^ Gao, Xiang; Gallicchio, Emilio; Roitberg, Adrian (2019). "The generalized Boltzmann distribution is the only distribution in which the Gibbs-Shannon entropy equals the thermodynamic entropy". The Journal of Chemical Physics. 151 (3): 034113. arXiv:1903.02121. Bibcode:2019JChPh.151c4113G. doi:10.1063/1.5111333. PMID 31325924. S2CID 118981017.
  8. ^ McDonald, I. R. (1972). " N p T {\displaystyle NpT} -ensemble Monte Carlo calculations for binary liquid mixtures". Molecular Physics. 23 (1): 41–58. Bibcode:1972MolPh..23...41M. doi:10.1080/00268977200100031.
  9. ^ Wood, W. W. (1970). " N p T {\displaystyle NpT} -Ensemble Monte Carlo Calculations for the Hard Disk Fluid". Journal of Chemical Physics. 52 (2): 729–741. Bibcode:1970JChPh..52..729W. doi:10.1063/1.1673047.
  10. ^ Schmidt, Jochen; VandeVondele, Joost; Kuo, I. F. William; Sebastiani, Daniel; Siepmann, J. Ilja; Hutter, Jürg; Mundy, Christopher J. (2009). "Isobaric-Isothermal Molecular Dynamics Simulations Utilizing Density Functional Theory:An Assessment of the Structure and Density of Water at Near-Ambient Conditions". Journal of Physical Chemistry B. 113 (35): 11959–11964. doi:10.1021/jp901990u. OSTI 980890. PMID 19663399.