• Keine Ergebnisse gefunden

Digression: Statistical ensembles and potentials

2.3 Generalisations of AdS/CFT

2.3.1 Digression: Statistical ensembles and potentials

We want to model strongly coupled quantum systems at finite temperature and density. Thereby, we automatically enter the regime of thermodynamics.

This section is devoted to remind us of the fundamental concepts of statistical mechanics, so that we are able to generalise these to holographic dualities. To begin with, we recapitulate the most important thermodynamic potentials which follow from the microcanonical, the canonical, and the grand canonical ensemble on a classical level. Then we proceed by explaining how to deal with quantum systems in the same manner. This will be useful to explain the notion of entanglement entropy and the related topic of why event horizons are always associated with a temperature.

In themicrocanonical ensemble we consider the total energy of the system fixed. The microcanonical partition function is then defined as the number of all microstates of total energy E by

Zmicro = Tr(1), (2.49)

where Tr defines the sum over all accessible states. The crucial assumption in the microcanonical ensemble is that we assign the same probability p = 1/Zmicro to all states accessible at the fixed energy E. Now, we are able to define the entropy of the system as

S =hlog(Zmicro)i=hlog(1/p)i= hlog(p)i= Tr (p log(p)) . (2.50) Consider the configuration of the classical system to be a random variable over the probability distribution set by the uniform probabilityp= 1/Zmicro. The entropy then defines the average surprise per sampling3, if we draw a random state from the set of possible states of total energyE, and is indeed maximised if we use the uniform probability. The entropy of the system is the thermodynamic potential of the microcanonical ensemble, meaning that we can use it to derive macroscopic information about the system. For example, the statistical definition of the inverse temperature of the system is given by deriving the entropy with respect to the energy,

⌘ @S

@E . (2.51)

Likewise, the pressure p is defined by deriving w.r.t. the volume V, and chemical potentials µk are defined by deriving w.r.t. fixed particle numbers Nk where k labels di↵erent sorts of particles. To summarise, the entropy S(E, V, Nk) is the thermodynamic potential of the microcanonical ensemble, in which the independent variables are the energy E, the volumeV and the

3To obtain the connection between expected surprise and entropy, we can think about a system in which one of the states has probabilityp0= 1 and, due to normalisation, all other states obeypi= 0. Then, the entropy is zero due to (2.50), which is intuitive, since we know the outcome of making observations in advance.

particle numbers Nk. Other ensembles are defined by performing Legendre transformations on one or more of the conjugated variables (E, ), (V, p) and (Nk, µk).

The canonical ensemble follows from the microcanonical ensemble by a Legendre transformation of (E, ), i.e. instead of the energy E, we use the temperature T or, equivalently, its inverse as independent variable. The new thermodynamic potential is the (Helmholtz) free energy F which is de-fined by

F( , V, Nk) = E( ) S(E( ), V, Nk). (2.52) Physically, this requires us to couple the system of interest to an infinite heat bath. Under the assumption that changes in our system are negligible for the temperature in the heat bath, one can derive that the partition function of the system becomes

Zcan = Tr exp ( E) . (2.53)

Hence, the contribution of states to the partition function is weighted by exp( E). Unlike fixing the energy, fixing the temperature in an experi-mental environment is much more convenient, which is why the canonical ensemble is more useful for applications.

For the same reason and especially in condensed matter applications, it is unfeasible to keep the particle number fixed. The intrinsic quantity asso-ciated to the particle number is the chemical potential. The grand canon-ical ensemble follows from the Legendre transformation of (Nk, µk), which exchanges the particle numbers and chemical potentials as independent vari-ables. Once again, this need us to couple our system to an infinite particle reservoir for each particle sort labelled by k. The thermodynamic potential is given by the Landau free energy⌦, which is defined as

⌦( , V, µk) =F( , V, Nkk)) µkNkk). (2.54) In order to fix the associated chemical potentials µk instead of the particle numbers, we change to the grand canonical ensemble by defining its partition functionZgrand as

Zgrand= Tr exp E X

k

µkNk

!!

. (2.55)

If we want to proceed to quantum systems, we need to initialise some notation. First of all, quantum mechanics deals with states | i 2H which are vectors in some Hilbert spaceH. A state is said to bepure if it is a vector

in H. We can compute expectation values of observablesO in the state | i by

hOi =h |O| i , (2.56)

where h | is the co-vector to | i w.r.t. the scalar product defined over H. Given a state | i, we can define a projection operator ⇢ ⌘ | i h |, which is also called thestate/density matrix/operator, where all four possible notations appear equivalently in the literature. Any expectation values can now be rewritten in the form

hOi = Tr (⇢ O) , (2.57)

where the trace Tr sums over an arbitrary normalised orthogonal set of basis vectors of the Hilbert space, e.g. eigenstates of the Hamilton operator. We can define amixed state by

⇢=X

i

pi | ii h i| , (2.58)

where| ii 2H and the probabilities of each statepi are normalised,P

ipi = 1. A mixed state is a genuine statistical mixture of pure states and, in general, there is no way to define a pure state | i whose state operator matches the one of the mixed state. The state operator always satisfies Tr⇢ = 1. If and only if the density matrix is the projection operator of a pure state, it also satisfies ⇢2 =⇢ or, equivalently, Tr(⇢2) = 1.

In the context of thermodynamics, the state matrix ⇢ replaces any oc-currences of the probability distribution p in the discussion about classical statistical mechanics. We always have a Hamilton operator ˆH defining the dynamics of the quantum theory. In equilibrium, any operators and states are supposed to be stationary. This means that, in particular, the state operator ⇢ commutes with the Hamilton operator,

@t⇢⌘[⇢,H] = 0ˆ , (2.59) which means that both operators can be diagonalised simultaneously in terms of energy eigenstates ˆH|ii=Ei|ii.

The microcanonical ensemble is then given by defining the partition func-tion as in (2.49), but the trace being over all energy eigenstates |ii with eigenvalues Ei =E. The microcanonical partition function Zmicro then just turns out to be the multiplicity of the energy level E.4 The von Neumann

4 Obviously, there are problems in this definition, because the energy levels of a quan-tum system are typically discrete and hence the derivatives w.r.t. the energy are probably not well-defined. The technical solution to this is to define an energy width E in which the states may be, which also explains the terminology for thedensity operator.

entropy associated with a given the state operator reads

S(⇢) = Tr (⇢log⇢), (2.60) where log⇢ is well-defined due to the state matrix being diagonalisable.

In the canonical ensemble, similar to the probability of a configuration of energy E in the classical context, the state matrix is now given by

⇢= 1

Zcane Hˆ , (2.61)

whereZcan is the canonical partition function Zcan = Tr⇣

e Hˆ

, (2.62)

which normalises the state operator such that Tr⇢= 1.

Finally, in the same fashion, the grand canonical ensemble is defined by its partition function

Zgrand = Tr exp Hˆ X

k

µkk

!!

, (2.63)

where we promoted the particle numbers Nk to conserved charge operators Qˆk.5 The state operator is then likewise given by

⇢= 1 Zgrand

e (Hˆ PkµkQˆk), (2.64) which concludes our recapitulation of both classical and quantum statistical mechanics.