<<

Statistical of classical States and ensembles

• A microstate of a statistical is specied by the complete about the states of all microscopic degrees of freedom of the system. It is impossible in practice to measure or mathematically express the enormous amount of information contained in a microstate. However, this concept is essential for the theoretical formulation of .

• A macrostate of a statistical system represents the macroscopic properties of the system, and may be specied by quantities such as T , p, V , internal E, magnetization M, etc. Descriptions of non-equilibrium are also of interest, using macroscopic quantities such as and dependent temperature T (r, t), pressure p(r, t), current density of particle ow j(r, t), etc. • Statistical mechanics studies the of macroscopic systems with many degrees of freedom starting from the microscopic point of view. The observable macroscopic behavior of the system is derived by applying the principles of to the dynamics of all microscopic degrees of freedom. In other words, the basic goal is to obtain equations for the evolution of the macrostate variables from the known fundamental dynamics and statistics of microstates.

• Microstates of classical particles (in a uid or ) are typically specied by the instantaneous positions ri and momenta pi (or vi = pi/mi) of all particles. One represents a microstate of N particles by a 6N-dimensional vector: Ω = (r1, p1, r2, p2, ··· , rN , pN )  In many cases one needs to generalize this description. For example, particles may have internal degrees of freedom (vibrational and rotational modes of molecules) that need to be included in Ω. If one wants to describe a system of magnets that are free to rotate but not change their location in , one may want to use , where is the magnetic of the th Ω = (µ1, µ2, ··· , µN ) µi i magnet. If the number of particles N can change, then it must be included in the the denition of Ω (whose dimensionality is then variable).

• Discrete classical systems have microstates Ω = (s1, s2, ··· , sN ), where si is some discrete variable that th indicates the state of i component. For example, two-state variables si ∈ {0, 1} would be appropriate for describing a system of many switches in a .

• The vector space spanned by all possible microstate vectors Ω is called . • Statistical ensemble and microstate : The of a classical macroscopic system is a Ω(t) in phase space. One could in principle determine Ω(t) from the initial state Ω(0) by solving the appropriate equations of for all of the system's degrees of freedom. This is, however, futile because the system of equations is too large and there is no hope to measure the initial state. Instead, we set up a statistical description of dynamics by dening a statistical ensemble and the probability of microstates.

 Imagine preparing N → ∞ copies of the system through some realistic process in which one controls a small number of the systems' properties. For example, one may , tem- perature and the number of particles, and make those properties the same in each copy of the system. The obtained collection of systems is called ensemble. Since one cannot control the many microscopic degrees of freedom, each member of the ensemble will generally have a dierent mi- crostate consistent with the controlled parameters. The phase space points Ω corresponding to the ensemble microstates should span the entire phase space more or less evenly.  For continuum systems of particles, consider a small phase space volume:

N Y 3 3 dΩ = d rid pi i=1

24 centered at the point Ω in phase space. Count the number dN (Ω) of ensemble microstates within dΩ. The probability density of a microstate Ω is: 1 dN (Ω) f(Ω) = lim N →∞ N dΩ f(Ω) is a PDF that depends on many random scalar variables. If we realistically prepare a single macroscopic system, then f(Ω)dΩ is the probability that the system will be within the phase space volume dΩ of the microstate Ω. As any PDF, f(Ω) is normalized:

dΩ f(Ω) = 1 ˆ

 For discrete systems whose microstates are Ω = (s1, s2, ··· , sN ), one can directly with normalized microstate p(Ω): X p(Ω) = 1 Ω Equilibrium

• The state of equilibrium is a macrostate that does not change in time. This implies that macroscopic state variables do not change in time. The microstate, however, may change in time as long as it remains consistent with the given static macrostate.

• Postulate about the equilibrium : The probability distribution of en- semble microstates must adequately reect our knowledge about the system's condition. If we knew

exactly all details about the system's microstate Ω0, then such knowledge would be captured by a perfectly sharp and deterministic distribution:   1 , Ω = Ω0 f(Ω) = δ(Ω−Ω0) for continuum systems ; p(Ω) = for discrete systems 0 , Ω 6= Ω0

where δ(Ω − Ω0) is the dened by: Y δ(Ω − Ω0) = δ(xi − x0i)δ(yi − y0i)δ(zi − z0i)δ(pxi − p0xi)δ(pyi − p0yi)δ(pzi − p0zi) i

∞  0 , x 6= 0  δ(x) = , dx δ(x) = 1 ∞ , x = 0 ˆ −∞

The above distribution is completely biased toward a single microstate Ω0. However, in reality we never have so much information about a macroscopic system. Information that we have in an equilibrium is given by state variables such as temperature T , pressure p, µ, total energy H, volume V , the number of particles N, etc. Since we have no knowledge about anything else, we should not arbitrarily introduce it into the probability distribution. In other words, the probability distribution should depend only on the quantities we know, and be completely unbiased toward any quantities we do not know - microstate in particular:   f(Ω) = f T, p, µ . . . ; H(Ω),V (Ω),N(Ω) ... for continuum systems in equilibrium   p(Ω) = p T, p, µ . . . ; H(Ω),V (Ω),N(Ω) ... for discrete systems in equilibrium

 We anticipated that extensive state variables (like total energy H) depend explicitly on the mi- crostate Ω, but required that the PDF f or probability p depend on Ω only implicitly through macroscopic state variables. Knowing nothing about microstates, all microstates with the same energy, etc. appear equally likely.

25  The postulated equilibrium probability distribution must be tested in experiments. In practice, all predictions of this postulate match the experiments because there are so many microstates corresponding to the same macrostate. The detailed time evolution from one microstate to another looks completely chaotic. One indeed cannot tell anything about which particular microstate realizes the observed macrostate at any instant of observation (i.e. there is no signicant bias).    The implicit dependence of the equilibrium PDF on total energy, f(Ω) = f H(Ω) , can be rigorously derived from the . We will show this later.  We used the notation H for total energy to emphasize its dierence from E. The former is a random state variable H(Ω) that depends on the microstate, while the latter is the average value E = hHi. Also, the total energy expressed as a function H(Ω) of the system's state is called Hamiltonian in mechanics.

represents perfectly isolated systems in equilibrium. The total energy E of an is conserved, so that the PDF or probability is:

     1 ,H(Ω) = E  f Ω = Cδ H(Ω)−E continuum ; p(Ω) = C0δ ≡ C0 discrete H(Ω),H 0 ,H(Ω) 6= E

0 where C,C are normalization constants, δ(x) is the Dirac delta function, and δx,y is Kronecker symbol (the discrete analogue of Dirac function). The system's volume and number of particles are also xed (but we emphasize only the PDF's dependence on energy).

represents the systems in equilibrium at given temperature T . The canonical PDF or probability is:   f H(Ω) = Ce−βH(Ω) continuum ; p(Ω) = C0e−βH(Ω) discrete

where C,C0 are normalization constants, and β is a temperature-dependent factor with units of inverse energy (we will show later that β = 1/kBT ). This PDF is known as . The system exchanges energy with its environment (so it is not conserved), but keeps its volume and particle number xed.

 One can derive the canonical PDF in the following manner. Consider two systems A and B in contact and thermal equilibrium with each other. Regard A as a member of a canonical ensemble EA(T ) and B as a member of a separate but equivalent canonical ensemble EB(T ). There is a unique functional form of the PDF f for all canonical ensembles. It is parametrized by temperature T and cannot explicitly depend on the microstate (since we have equilibrium). It can, however, depend on a few extensive microstate functions such as system's total energy H, volume V , number of particles N, etc. Consequently, the PDFs for nding the systems A and B in particular microstates are f(T ; HA,VA,NA ... ) and f(T ; HB,VB,NB ... ) respectively. Now consider the joint system A + B. Since energy can be exchanged between its parts, A + B is another equilibrium system belonging to the same type of canonical ensemble EA+B(T ). The PDF of A+B is therefore f(T ; HA +HB,VA +VB,NA +NB ... ). However, the degrees of freedom that belong to A are completely independent from the degrees of freedom that belong to B (particles are not exchanged between A and B). The properties of probability then ensure:

f(T ; HA + HB,VA + VB,NA + NB ... )dΩAdΩB =

f(T ; HA,VA,NA ... )dΩA × f(T ; HB,VB,NB ... )dΩB Cancel the phase space volume factors dΩ and take the :

log f(T ; HA + HB,VA + VB,NA + NB ... ) = log f(T ; HA,VA,NA ... ) + log f(T ; HB,VB,NB ... ) The most general functional form that satises this equation is:

log f(T ; H,V,N... ) = −βH − γV − νN − · · ·

26 where the coecients β, γ, ν are parameters that characterize the established equilibrium in which energy, , particles, etc. can be exchanged with the environment. Specically, the need to introduce the parameter β can be taken as evidence that temperature exists as a state variable; β is some function of temperature, xed only by the precise denition of temperature. If we consume γV + νN + ··· into the normalization constant C, then:

f(T ; H,V,N... ) = C(T ; V,N... ) × e−βH

represents the systems in equilibrium at given temperature T and in contact with a particle reservoir at chemical potential µ (the systems exchange both energy and particles with their environment). The grand canonical PDF or probability is:       −β H(Ω)−µN −β H(Ω)−µN f H(Ω),N = Ce continuum ; p(Ω) = C0e discrete

where C,C0 are normalization constants, N is the number of particles, and β is a temperature- dependent factor with units of inverse energy.

 The proof is analogous to that for the canonical distribution, but the dependence on N is kept outside of the normalization constant: f = C × e−βH−νN . Then we may write ν = −β0µ, where 0 β0 has the same units as β, in order to obtain the correct units for ω: f = Ce−βH+β µN . Lastly, we identify β0 = β, so the exponent becomes proportional to E = H − µN. This is justied by the fact that chemical potential µ is of a particle that has been taken from the particle reservoir into the system. It is a real potential energy because it can be returned to the reservoir at any time by returning the particle. Therefore, we may regard E as the true total energy of the system, which includes the potential energy with respect to the reservoir. The system with energy E is still in thermal equilibrium, so it should have a canonical PDF (β0 = β).

• Density and number of states: The equilibrium probability must be normalized over the entire phase space:

  X   dΩ f H(Ω) = 1 continuum ; p H(Ω) = 1 discrete ˆ Ω Given that equilibrium probability distributions depend only on energy, we can reorganize the normal- ization conditions as /sums over energy. The weight function of energy in these and similar integrals/sums is the density/number of states.

 In continuum systems, we substitute the identity

    f H(Ω) = dH δ H − H(Ω) f(H) ˆ in the normalization condition:     dΩ f H(Ω) = dH dΩ δ H(Ω) − H f(H) = dH ω(H)f(H) = 1 ˆ ˆ ˆ ˆ   ω(H) = dΩ δ H(Ω) − H ˆ The function ω(H) is the density of microstates at energy H; ω(H)dH is the phase space volume occupied by the microstates whose energy lies in the interval (H,H + dH).  In discrete systems, we substitute the identity

  X p H(Ω) = δH,H(Ω) p(H) H

27 in the normalization condition: X   X X X p H(Ω) = δH(Ω),H p(H) = Ω(H)p(H) = 1 Ω H Ω H X Ω(H) = δH(Ω),H Ω The function Ω(H) is the number of microstates whose energy is H.

• Partition function: The normalization of the canonical ensemble PDF yields: 1 dH ω(H)f(H) = C dH ω(H)e−βH = 1 ⇒ C = ,Z = dH ω(H)e−βH ˆ ˆ Z ˆ

where ω(H) is the density of states, and Z is the so called partition function. Exactly the same form of the normalization constant C0 is obtained for discrete canonical distributions: X X 1 X Ω(H)p(H) = C0 Ω(H)e−βH = 1 ⇒ C0 = ,Z = Ω(H)e−βH Z H H H  Partition function depends on various state variables, which are then constrained and related to the internal energy by the normalization condition.

Single and multi-particle probability distributions

• For the sake of generality, let us represent the microstate vector Ω = (ω1, ω2 ... ωN ) in terms of the state vectors ωi of individual degrees of freedom. If the system is a uid or solid of N particles, we would use ωi = (ri, pi), but more generally these could be other continuous or discrete quantities.  Referring to a single degree of freedom as particle, we dene the single-particle PDF or probability as:

0 0 0 X 0 f1(ωi) = dΩ f(Ω )δ(ω − ω ) (continuous) ; p1(ωi) = p(Ω ) δω0 ω (discrete) ˆ i i i i Ω0

using the microstate PDF f(Ω) or probability p(Ω). Note that we used Dirac function δ(x − y) or Kronecker symbol δxy (equal to 1 only if x = y, otherwise zero) to constrain the integra- 0 th tion/summation to only those microstates Ω whose i particle's state is xed to ωi. The func- tions f1 and p1 capture the statistics of a single particle. If all particles are identical, these functions do not depend on the particle index i, e.g. f1(ωi) ≡ f1(ω).  Similarly, we dene the two-particle PDF or probability:

f (ω , ω ) = dΩ0 f(Ω0)δ(ω0 − ω )δ(ω0 − ω ) (continuous) 2 i j ˆ i i j j X 0 p (ω , ω ) = p(Ω ) δ 0 δ 0 (discrete) 2 i j ωiωi ωj ωj Ω0 and so on. Such multi-particle functions contain information about correlations between particles. Specically,

g2(ωi, ωj) = f2(ωi, ωj) − f1(ωi)f1(ωj) is a two-body that equals zero only if the particles are uncorrelated (according to the fundamental property of probability for independent random variables). Assuming that all particles are identical, the multi-particle distribution and correlation functions depend on the 0 coordinates of unspecied particles, as in f2(ω, ω ).

28 • As long as the system has a macroscopically large number N of , and we are only interested in few-particle n  N (rather than many-particle n ∼ N) PDFs fn, we can obtain the PDFs by averaging over the degrees of freedom of a single system. In other words,

0 1 dN(ω) 0 1 dN(ω, ω ) f1(ω) = lim , f2(ω, ω ) = lim ··· N→∞ N dω N→∞ N 2 dωdω0 where dN(ω) is the number of particles (in a single system) whose coordinates are within a small volume dω from the state ω, and dN(ω, ω0) is the number of particle pairs whose coordinates are within a small volume dωdω0 from the joint state (ω, ω0).  Proof: Consider the original single-particle PDF and substitute in its denition the dening for the objective microstate PDF f(Ω):

0 0 0 0 0 0 1 dN (Ω ) 1 dN (ω) f 1(ω) = dΩ f(Ω )δ(ωi − ω) = dΩ δ(ωi − ω) × lim = lim ˆ ˆ N →∞ N dΩ0 N →∞ N dω

When the number dN (Ω0) of ensemble members in the microstate Ω0 is integrated over the constrained phase space volume, one gets the number dN (ω) of ensemble members that contain a particle in the given state ω. Since dN (ω) is obtained by a single counting experiment on the entire ensemble, it can be regarded as a binomially-distributed random variable. Its mean hdN (ω)i is, then, equal to N P (ω), where P (ω) is the probability that a single ensemble member will have a particle in the state ω. We can alternatively extract P (ω) from a single system by applying the denition of objective probability on the ensemble of particles in a given single system of interest: dN(ω) P (ω) = lim N→∞ N Therefore:   0 1 dN (ω) 1 dN (ω) 1 N P (ω) dN(ω) f1(ω) = lim → = = lim = f1(ω) N →∞ N dω N dω N dω N→∞ Ndω The equality indicated by the arrow can be regarded as a consequence of central limit theorem: the random quantity dN /N is an average whose distribution converges to the Gaussian distribution with the same mean and vanishing variance (of the order of 1/N ). Using analogous arguments, one can prove the equivalence between any multi-particle PDFs 0 , and also generalize to fn = fn discrete cases.

Fluctuations in equilibrium

• Instantaneous total energy H of a macroscopic system is a random variable (except in the microcanon- ical ensemble). The standard deviation σH of H is negligible in comparison to its mean E = hHi. In other words, the statistical uctuations of H are negligible, so all ensembles that describe the same macrostate (given by E) are equivalent.

 Proof: A particle energy i is a random variable with some nite mean  and standard deviation σ (particles are equivalent even if they are correlated by interactions). The total energy H is the sum of N individual particle i, e.g.:

N X p2 1 X H =  ,  = i + U(r ) + V (r − r ) i i 2m i 2 j i i=1 j6=i

Since all particles are equivalent, the average of H is:

* N + N X X E = hHi = i = hii = N i=1 i=1

29 and the variance of H is:

* N !2+ * N + 2 2 X X 2 X σH = h(H − E) i = (i − ) = (i − ) + (i − )(j − ) i=1 i=1 i6=j 2 X = Nσ + h(i − )(j − )i i6=j Any correlations between particles must be local if interactions have short range (Coulomb in- teractions in plasmas can be included because screening makes them eectively short-ranged). In other words, a particle may be correlated with other particles in its vicinity, but not with the macroscopic majority of particles far away. Therefore, we may express the last term on the right-hand-side in 2 as: σH

near i far from i X X X X X h(i − )(j − )i = h(i − )(j − )i + h(i − )(j − )i i6=j i j6=i i j far from i X X = Nφ + h(i − )ih(j − )i i j = Nφ

We merely separated the sum over the second particle j into the contribution from particles that are near the rst particle i and the contribution from far-away particles. The ones that are near i can interact with it and contribute the amount Nφ proportional the total number N of particles i (the value of φ is not important). The far-away particles are uncorrelated with i, so we can use the property of independent random variables that the average of their product equals

the product of their averages. Then, by denition of  = hii, the average of i −  vanishes, and we conclude that the full variance of H is proportional to the number of particles:

2 2 σH = N(σ + φ) It can be now clearly seen that the standard deviation of is much smaller than its mean, √ H σH = const × N  E = const × N in the N → ∞ limit.

 The theorem trivially applies to the microcanonical ensemble, where σH ≡ 0.  It can be shown using the same argument that all central moments h(H − E)ni in systems with short-range correlations are proportional to . Therefore, all measures of uctuations √ N pn h(H − E)ni ∝ n N are small in comparison to the mean E ∝ N.  The same theorem applies also to the uctuations of other extensive quantities, such as the number of particles N.

• Few body PDF's fn (for N  n) are the same in all types of equilibrium ensembles (microcanonical, canonical, etc.) that correspond to the same macrostate. In this sense (much stricter than previously), all ensembles are equivalent.

 Proof: Consider N identical particles and focus on f1(ω):

N 1 X f (ω) = dΩ0 f(Ω0)δ(ω0 − ω) = dΩ0 f(Ω0) δ(ω0 − ω) 1 ˆ 1 ˆ N i i=1

The microstate PDF f(Ω0) ≡ f(H) depends only on the energy H(Ω0) in equilibrium, and must produce the mean E = hH(Ω0)i xed by the macrostate specication. The ensemble type deter- mines the system's boundary condition (interaction with the environment) in a given macrostate.

30 Thus, only the uctuations of H (the variance and higher-order ) may depend on the type of ensemble. Having this in mind, it is useful to rewrite f1 as an over energy:

N N 1 X 1   X f (ω) = dΩ0 f(Ω0) δ(ω0 − ω) = dH f(H) × dΩ0 δ H(Ω0) − H δ(ω0 − ω) 1 N ˆ i N ˆ ˆ i i=1 i=1 This looks like the average hX(H, ω)i of the random quantity

N   X X(H, ω) = dΩ0 δ H(Ω0) − H δ(ω0 − ω) ˆ i i=1 that depends on random total energy H and the given single-particle state ω. By construction, X(H, ω) is proportional to the average number of particles at ω in the microstates with total energy H. This is a local property of the system's dynamics (mathematically encoded in H(Ω), the Hamiltonian) and the density of states, but expressed as a function of the macroscopic total energy. Small changes of H have very little eect on X; only small changes of energy per particle  = H/N do aect X: dkX dkX 1 dkX ∼ O(N 0) ⇒ = dk dHk N k dk

We can exploit this feature in the expansion of f1(ω) = hX(H, ω)i in terms of central moments h(H − E)ki: ∞ X 1 dkX(E, ω) 1 d2X(E, ω) hX(H, ω)i = h(H − E)ki = X(E, ω) + Var(H) + ··· k! dEk 2 dE2 k=0 This expansion follows from the Taylor expansion of hX(H, ω)i about H = E under the ensemble average. We proved before that Var(H) ∝ N (all central moments are proportional to N). Using dkX/dEk ∝ 1/N k we nd: 1 1 f (ω) = hX(H, ω)i = X(E, ω) + C N + C N + ··· −−−−→N→∞ X(E, ω) 1 2 N 2 3 N 3

Therefore, f1(ω) equals the variable X evaluated at the mean energy H = E in the N → ∞ limit. This way of calculating f1 is identical to the microcanonical ensemble calculation, where total energy is xed. In other ensembles, energy uctuations produce corrections (the terms involving

constants C2,C3 ... ), but all of them vanish).

 The proof for other multi-particle PDFs fn proceeds in the same steps. It breaks down, however, when n becomes comparable to N (because the assumption about the changes of the generalized X with E fails).

Entropy

• The statistical formula for consistent with laws of is:   X   S = −k dΩ f(Ω) log f(Ω)∆Ω (continuum) ; S = −k p(Ω) log p(Ω) (discrete) B ˆ B Ω In continuum systems, f(Ω)dΩ → f(Ω)∆Ω inside the logarithm is microstate probability p(Ω) in analogy with discrete systems. Then, ∆Ω is the elementary phase space volume that cannot contain more than one microstate.

 Proof: Entropy S is dened in thermodynamics as an independent state variable related to the transfer of in a reversible change from one equilibrium state A to another B:

B dQ S − S = B A ˆ T A

31 If the system is unable to do any work, then, by , the amount of transferred heat equals the change of internal energy:

B dE SB − SA = ˆ T dW =0 A We will use this relationship and 3rd law of thermodynamics to deduce the statistical denition of entropy. Without loss of generality, we will work with a discrete ensemble and seek an expression of the form: X S = hSi = p(Ω)S(Ω) Ω which properly extracts entropy as a state variable (ensemble average) from the distribution of some physical quantity S(Ω). We ought to nd S(Ω) that is fundamentally independent of microstate energy H(Ω), because entropy is an independent state variable. Consider a process in which the internal energy of the system is changed at xed temperature and without any ability for the system to do work (an example is Joules' experiment with a freely expanding that has nothing to push and do work on). Based on this process and the thermodynamic denition of entropy, we nd: 1 E S − S = (E − E ) ⇒ S = const + B A T B A T This is a type of an : S and E are independent at the microscopic level, but related as above only in equilibrium under the conditions we xed. Let us express internal energy E as a desired ensemble average: 1 X S = const + p(Ω)H(Ω) T Ω and then try to eliminate the reference to the microstate energy in the sum: 1 X   1 X   S = const − p(Ω) log e−βH(Ω) = const − p(Ω) log Zp(Ω) βT βT Ω Ω 1 X   = const0 − p(Ω) log p(Ω) βT Ω We temporarily used the canonical probability distribution 1 X p(Ω) = e−βH(Ω) ,Z = e−βH(Ω) Z Ω and the fact that partition function Z is a constant whose logarithm we may absorb into const0. However, any probability distribution xed by the macrostate (i.e. any type of ensemble) is acceptable, because all ensembles are statistically equivalent. We would like to identify: 1   S(Ω) → − log p(Ω) βT but this is not adequate because we need a truly microscopic denition of S that makes no reference to macrostate quantities such as temperature. The problem is eliminated only if βT is constant. In other words, 1 β ∝ kBT is fundamentally required by the units of β (βE is dimensionless) and the fact that entropy is an independent state variable. We will later show that β = 1/kBT by studying the statistical mechanics of an - this is a matter of how exactly temperature is dened. Hence, the meaning of the physical quantity S(Ω) that microscopically relates to entropy is derived from the fraction of occurrences of the microstate Ω in the ensemble of microstates prepared from the same macrostate.

32  We will now prove that the above constant should be zero using 3rd law of thermodynamics. Canonical distribution (which makes a reference to temperature) shows that the system is certainly

in its lowest energy state H = H0 at T = 0:

1 1 T →0 1 p(Ω) = e−βH = e−H/kB T −−−→ e−H/kB T P e−βH P e−H/kB T e−H0/kB T

−(H−H0)/kB T T →0 = e −−−→ δH(Ω),H0

because the partition function in the denominator becomes dominated by the smallest energy H0 in the spectrum, i.e. Z ≈e−H0/kB T . All natural systems have a unique lowest energy state, or a few lowest energy states. If the lowest energy state is unique, then : Ω0 p(Ω) → δH(Ω),H0 = δΩ,Ω0   const0 X T →0 const0 X S = − kB p(Ω) log p(Ω) −−−→ − kB δΩ,Ω0 log(δΩ,Ω0 ) Ω Ω 0 0 = const − kB × 1 × log(1) = const → 0

Since entropy vanishes at T = 0 according to the 3rd law of thermodynamics, we nd that const0 = 0. If, more generally, there are a few minimum energy states, then entropy dened this way will be nite at T = 0, but of the order one (hence negligible), instead of being proportional to the number of particles N as an extensive quantity.

• In discrete microcanonical ensemble at total energy E, entropy equals:   S = kB log Ω(E)

where Ω(E) is the number of microstates at that energy.

 Proof: The microcanonical probability distribution specied by internal energy E is:

1 X p(Ω) = δ , Ω(E) = δ Ω(E) H(Ω),E H(Ω),E Ω

where Ω(E) is the number of microstates at energy E. Hence, entropy in a discrete microcanonical ensemble becomes:

0 X   S = const − kB p(Ω) log p(Ω) Ω   kB X δH(Ω),E = const0 − δ log Ω(E) H(Ω),E Ω(E) Ω   kB X 1 = const0 − δ log Ω(E) H(Ω),E Ω(E) Ω k  1  = const0 − B × Ω(E) log Ω(E) Ω(E) 0   = const + kB log Ω(E)

Note that the Kronecker symbol inside the logarithm (2nd line) has no eect because the only time it matters is when the same Kronecker symbol outside of the logarithm is equal to one. Then, the logarithm is essentially independent of Ω and can be pulled out of the sum. The isolated sum over microstates is just the number of microstates Ω(E), which is canceled by the normalization of the probability. The nal result is that entropy is the logarithm of the number of microstates at given energy E, up to a constant. 3rd law of thermodynamics says that S = 0 at zero temperature.

33 • Having the statistical denition of entropy, it is possible to microscopically derive the 2nd law of ther- modynamics. Any system constrained to be in a given macrostate still has many accessible microstates. Out of equilibrium, there could be some bias in the statistical distribution of microstates. For example, one could prepare a gas by putting all of its particles in a small volume at a corner of a large box. Among all existing microstates for the gas in the box, only those with particles conned to the corner are initially likely. The gas will, however, quickly spread through the entire box. In this process, the entropy   S = kB log Ω(E) grows because the number of available microstates of the gas with conserved energy E grows with the volume that the gas occupies. As the system evolves from a less to a more likely (i.e. chaotic) condition, its entropy increases - which is one way to state 2nd law of thermodynamics.

• The laws of nature are invariant under time-reversal. For any possible process, the one unfolding in exactly the opposite order is also possible according to the equation of motion. So, if one drops an egg on the oor and the egg breaks, it is possible in principle to simultaneously impart on every atom exactly the opposite from the one it had at the instant of fall - the pieces of the broken egg would y upward, reassemble and land back to one's hand. Such a reversed process is never observed, even though it is possible. This is a reection of the 2nd law of thermodynamics, now interpreted statistically. Every macroscopic system always evolves to more probable and more featureless states. This is the most likely scenario, and when the number of degrees of freedom is huge, the probability of this is huge. The reversed process is just statistically extremely unlikely. In that sense, there is an arrow of time, the causal passage of time from past to future.

• Entropy dened by X   S ∝ − p(Ω) log p(Ω) Ω is a mathematical concept, an abstract property of a probability distribution. It expresses the overall of measurement outcomes, or the amount of information we lack about the described statistical system. We could have postulated this formula based solely on its mathematical meaning, and then derived all equilibrium distributions from the requirement that entropy (uncertainty) be maximized. The following statements have rigorous mathematical proofs:

 Among all possible ensembles of a statistical system (not just equilibrium ones) with xed energy, the one with maximum entropy is microcanonical.  Among all possible ensembles of a statistical system with xed temperature, the one with maxi- mum entropy is canonical.  Among all probability distributions dened on a set of possible measurement outcomes, the one with maximum entropy is uniform.  Among all probability distributions dened on x ∈ (0, ∞) with xed mean µ = hxi, the one with maximum entropy is exponential.  Among all probability distributions dened on x ∈ (−∞, ∞) with xed mean µ = hxi and variance σ2 = Var(x), the one with maximum entropy is Gaussian.

34