* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download Chapter 12 Quantum gases
Quantum computing wikipedia , lookup
Ensemble interpretation wikipedia , lookup
Quantum chromodynamics wikipedia , lookup
Bohr–Einstein debates wikipedia , lookup
Scalar field theory wikipedia , lookup
Density matrix wikipedia , lookup
Many-worlds interpretation wikipedia , lookup
Renormalization wikipedia , lookup
Measurement in quantum mechanics wikipedia , lookup
Renormalization group wikipedia , lookup
Quantum machine learning wikipedia , lookup
Spin (physics) wikipedia , lookup
Probability amplitude wikipedia , lookup
Quantum group wikipedia , lookup
Coherent states wikipedia , lookup
Bell's theorem wikipedia , lookup
Quantum key distribution wikipedia , lookup
Copenhagen interpretation wikipedia , lookup
Matter wave wikipedia , lookup
Wave function wikipedia , lookup
Quantum field theory wikipedia , lookup
Quantum entanglement wikipedia , lookup
Path integral formulation wikipedia , lookup
EPR paradox wikipedia , lookup
Interpretations of quantum mechanics wikipedia , lookup
Second quantization wikipedia , lookup
Double-slit experiment wikipedia , lookup
History of quantum field theory wikipedia , lookup
Wave–particle duality wikipedia , lookup
Particle in a box wikipedia , lookup
Quantum teleportation wikipedia , lookup
Hidden variable theory wikipedia , lookup
Atomic theory wikipedia , lookup
Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup
Quantum state wikipedia , lookup
Relativistic quantum mechanics wikipedia , lookup
Symmetry in quantum mechanics wikipedia , lookup
Elementary particle wikipedia , lookup
Chapter 12 Quantum gases In classical statistical mechanics, we dealt with an ideal gas which was a good approximation for a real gas in the highly diluted limit. An important difference between classical and quantum mechanical many-body systems lies in the distinguishable character of their constituent particles. The inherent indistinguishability in quantum mechanics gives rise to phenomena and concepts that are not present in classical physics. In this chapter, we will present the ideal quantum gases, the Fermi gas and the Bose gas, as well as some practical examples of these systems. 12.1 Identical particles We consider a system of N non-interacting particles: Ĥ = N X Ĥ (i) , i=1 with (i) Ĥ (i) ϕ(i) αi = εα i ϕ α i . (12.1) E (i) In eq. (12.1), ϕαi are the one-particle wavefunctions, with the subindex αi denoting a complete set of quantum numbers such as, for instance, (n, l, ml , ms ), (kx , ky , kz , ms ), etc. In the case of N distinguishable particles, we can enumerate them so that the wavefunction of the N -particle system is given as a product of the one-particle states: (2) (N ) ϕα . . . ϕα . |ϕN i ≡ |ϕα1 , . . . , ϕαN i ≡ ϕ(1) α1 2 N If the particles are indistinguishable, then the principle of indistinguishability imposes certain symmetry properties on the N -particle state, i.e., that ⋆ the interchange of two particles results in, at most, the change of the sign of the N -particle wavefunction. 153 154 CHAPTER 12. QUANTUM GASES Therefore, the N -particle wavefunction has to be properly symmetrized: E (±) ≡ |ϕα1 , . . . , ϕαN i(±) ϕN (2) (N ) 1 X ϕα . . . ϕ α ) . (±)p P(ϕ(1) ≡ α1 2 N N! P p: number of pair interchanges (transpositions) performed by the permutation P. E (+) or antisymThe state of a system of identical particles can be either symmetrical ϕN E (−) metrical ϕN . These two types of states are orthogonal to one another; they belong to (+) (−) different Hilbert spaces HN and HN . Pauli spin-statistics theorem Wolfgang Pauli showed that, for a system of indistinguishable particle of certain sort, the spin of the particles and the nature of the system’s wavefunction (symmetrical or antisymmetrical) are related to one another. E (+) (+) HN : Hilbert space of symmetrical states ϕN of a system that consists of identical particles with integer spin (s = 0, 1, 2, 3, . . .) called bosons . Examples: photons (s = 1), phonons (s = 1), magnons (s = 1), α-particles (s = 0). E (−) (−) HN : Hilbert space of antisymmetrical states ϕN of a system that consists of identical particles with half-integer spin (s = 12 , 23 , . . .) called fermions . Examples: electrons (s = 1/2), protons (s = 1/2), neutrons (s = 1/2). The N -particle wavefunction of a fermionic minant: (1) E ϕ α1 E 1 (−) .. ϕN = N! . E ϕ(1) αN system can be represented by a Slater deter E (2) ... ϕ α1 .. . E (2) ... ϕαN E (N ) ϕ α1 .. . E (N ) ϕαN . E (−) If two raws or two columns of the determinant are identical then ϕN = 0. This is a consequence of the Pauli principle: ”two fermions can never coincide in all of their quantum numbers” 155 12.2. FOCK SPACE 12.2 Fock space A very different way of representing the N -particle wavefunction for a system of fermions or bosons is by considering the occupation numbers nαi . E (±) nαi : number of times that the one-particle state |ϕαi i is present in ϕN . In terms of nαi , the N -particle state is written as X (p) (p+1) (±) (1) (2) p (±) P ϕα1 ϕα1 . . . ϕαi ϕαi |N ; nα1 . . . nαi . . .i ≡ c± ... , {z } | {z } | P n α1 n αi with N = X n αi and i c± = p N! 1 Q i n αi ! . (12.2) c± is calculated from the normalization for the orthogonal Fock states: Y (±) δnαi ñαi . N ; . . . nαi . . . |Ñ ; . . . ñαi . . . (±) = δN Ñ i Note that nαi = 0 or 1 for fermions, nαi = 0, 1, 2, . . . for bosons. (±) The Fock states constitute a complete orthonormal basis in HN . Second quantization Having extended our description of a quantum system by considering the Fock space, we can introduce now the formalism of second quantization. We define the creation (annihilation) operators a†αi and aαi as |ϕαi i = a†αi |0i , aαi |ϕαi i = |0i , aαi |0i = 0, |0i: vacuum state. The operator a†αi (aαi ) creates (annihilates) the one-particle state |ϕαi i. By considering the corresponding normalization factors (QM II), these operators are defined in terms of the N -particle wavefunction as follows: 156 CHAPTER 12. QUANTUM GASES Bosons: p nαr + 1 |N + 1; . . . nαr + 1 . . .i(+) , √ nαr |N − 1; . . . nαr − 1 . . .i(+) , = a†αr |N ; . . . nαr . . .i(+) = aαr |N ; . . . nαr . . .i(+) nαr = 0, 1, 2, . . . . Fermions: a†αr |N ; . . . nαr . . .i(−) = (−1)Nr δnαr ,0 |N + 1; . . . nαr + 1 . . .i(−) , aαr |N ; . . . nαr . . .i(−) = (−1)Nr δnαr ,1 |N − 1; . . . nαr − 1 . . .i(−) , nαr = 0, 1; Nr = r−1 X n αj . j=1 The Fock states can be represented by the creation operators as n j a†αj Y p |N ; . . . nαr . . .i(±) = (±1)Nj |0i . nj ! j In this representation, we get rid of the symmetrization/antisymmetrization aspect of the wavefunction. The information about the symmetry of the wavefunction is now contained in the commutation relations of the creation/annihilation operators: • for bosons, • for fermions, [aαr , aαs ] = a†αr , a†αs = 0, aαr , a†αs = δrs ; {aαr , aαs } = {a†αr , a†αs } = 0, {aαr , a†αs } = δrs . Above, [ , ] denotes a commutator and { , } denotes an anticommutator. Usually, one uses a and a† for denoting fermionic creation/annihilation operators and b and b† for bosonic ones. Quantum mechanical operators describing observables can also be represented in terms of creation and annihilation operators. - Occupation number operator: n̂αr = a†αr aαr . 12.3. PARTITION FUNCTION OF THE IDEAL QUANTUM GASES 157 - Particle number operator : N̂ = X n̂αr = r X a†αr aαr . r - For an ideal quantum gas of N independent particles, the Hamiltonian operator is given by N X (i) Ĥ = Ĥ1 , i=1 with (i) Ĥ1 = 1 2 p̂ + V (r̂i ). 2m i V (r̂i ) describes the possible interaction of particle i with an external potential V (electric field, magnetic field, etc). In second quantization, Ĥ can be written as X X Ĥ = εr a†αr aαr = εr n̂αr , r r with εr δrs = hεr | Ĥ1 |εs i . The Fock states are the eigenstates of Ĥ: Ĥ |N ; . . . nαr . . .i 12.3 (±) = X εr n α r r ! |N ; . . . nαr . . .i(±) . Partition function of the ideal quantum gases The most adequate ensemble to deal with quantum gases is the grand canonical ensemble: Z(T, V, µ) = Tr e−β(Ĥ−µN̂ ) , (12.3) Ĥ: Hamilton operator, N̂ : particle number operator. In order to evaluate the trace, it is very useful to consider the Fock states, which are eigenstates of both the particle number operator and the Hamilton operator. This allows us to rewrite (12.3) as Z (±) (T, V, µ) = ∞ X N =0 = ∞ X N =0 P P nr =N X r e−β P r nr (εr −µ) {nr } nr =N X Y r {nr } r e−βnr (εr −µ) , (12.4) 158 CHAPTER 12. QUANTUM GASES where Z (+) (T, V, µ) is the partition function for an ideal gas of bosons and Z (−) (T, V, µ) is the partition function for an ideal gas of fermions, and nr was used instead of nαr for brevity. The sum {nr } in eq. (12.4) runs over all combinations of occupation numbers for a given P total number of particles N . In fact, since the summation ∞ frees the constraint of N =0 P nr =N X r , {nr } one can alternatively write the double sum ∞ X N =0 as XX n1 P nr =N X r (12.5) {nr } ... n2 X ..., (12.6) nr P where the sums nr are independent. Now, using the equivalence between the summations (12.5) and (12.6), we calculate the partition function: ! ! ! X X X e−βnr (εr −µ) e−βn1 (ε1 −µ) e−βn2 (ε2 −µ) . . . Z (±) (T, V, µ) = n2 n1 = Y X r nr e−βnr (εr −µ) ! . nr (12.7) Eq. (12.7) can be further simplified if one makes use of the statistical properties of particles in a given system. Thus, • for a system of bosons , nr runs over all non-negative integers so that in the parenthesis of eq. (12.7) we have a geometrical series: Y 1 (+) Z (T, V, µ) = , 1 − e−β(εr −µ) r • for a system of fermions , nr = 0, 1 and therefore we only have two terms in eq. (12.7): Y Z (−) (T, V, µ) = 1 + e−β(εr −µ) . r Out of Z (±) one can obtain all the thermodynamical properties of the ideal quantum gases. 12.4. THERMODYNAMICAL PROPERTIES OF IDEAL QUANTUM GASES 12.4 159 Thermodynamical properties of ideal quantum gases (i) Grand canonical potential : Ω(+) (T, V, µ) = −kB T ln Z (+) (T, V, µ) X = kB T ln 1 − e−β(εr −µ) ; r Ω(−) (T, V, µ) = −kB T ln Z (−) (T, V, µ) X = −kB T ln 1 + e−β(εr −µ) . r ⋆ Note that the volume dependence of Ω(±) (T, V, µ) is in the one-particle energies εr . Thus, for instance, if the particles are in a potential well with infinitely high walls, then, accordingly, ǫr will have only certain allowed values. (ii) Expectation value of the particle number operator N̂ : X 1 1 ∂ ln Z (+) (T, V, µ) = ; β(ǫr −µ) − 1 β ∂µ e r X 1 ∂ 1 (−) = ln Z (T, V, µ) = . β(εr −µ) + 1 β ∂µ e r < N̂ >(+) = < N̂ >(−) Out of these equations, we can obtain the chemical potential µ = µ(T, V, < N̂ >(±) ). (iii) Thermal equation of state for the ideal quantum gas: P V = kB T ln Z (±) (T, V, µ). (iv) Internal energy for bosons: ∂ ln Z (+) (T, V, µ) + µ < N̂ >(+) ∂β X εr − µ + µ < N̂ >(+) = β(εr −µ) − 1 e r X εr = β(ε −µ) r e −1 r U (+) = − and for fermions: U (−) = X r εr . eβ(εr −µ) + 1 (12.8) 160 CHAPTER 12. QUANTUM GASES 12.5 Particle distribution function for bosons and fermions We define the average occupation number < n̂r >(±) of the rth one-particle state, P < n̂r >(±) nl =N X lX P 1 [−β l nl (εl −µ)] = n e r Z (±) (T, V, µ) N {nl } = − 1 ∂ ln Z (±) (T, V, µ) , β ∂εr as the distribution function of particles in the state r. This quantity is especially important for understanding quantum gases. For bosons and fermions we have, correspondingly, → Bose-Einstein distribution function: 1 < n̂r >(+) = eβ(εr −µ) −1 = f (+) (εr ) and → Fermi-Dirac distribution function: < n̂r >(−) = 1 eβ(εr −µ) +1 = f (−) (εr ). Properties: (1) For fermions, µ can have any value and 0 ≤ < n̂r >(−) ≤ 1 (Pauli principle). (2) For bosons, µ must be smaller than the smallest one-particle energy ε0 , otherwise the occupation number < n̂0 > would be negative and diverge for ε0 = µ. For T → 0, this would mean that all occupation numbers are zero and therefore the total number of bosons N → 0. This would be fine for the case that we have no bosons. But what happens when N is given and nonzero? ⋆ In order to treat the limit T → 0, we make an assumption that in this limit µ → ε0 in a way that the lowest one-particle state is macroscopically occupied, i.e., < n̂0 >(+) (T = 0) = N ⇒ Bose-Einstein condensation. 12.5. PARTICLE DISTRIBUTION FUNCTION FOR BOSONS AND FERMIONS 161 (3) Making use of < nr >(±) , we can calculate the average total number of particles and the internal energy as < N̂ >(±) = X < nr >(±) , r U (±) = X εr < n̂r >(±) . r (4) For large one-particle energies, i.e., for εr − µ >> kB T , both Bose-Einstein and Fermi-Dirac distribution functions converge to the Maxwell-Boltzmann distribution function: (εr − µ >> kB T ), < n̂r >(±) ∼ e−βεr which implies that in the classical limit there is no difference between bosons and fermions. The meaning of this limit will be explicitly discussed in the next chapter Graphical representation of the distribution functions We define x = β(εr − µ) and plot < nr > vs. x for the three distribution functions (Fig. 12.1). Figure 12.1: Fermi-Dirac (F-D), Bose-Einstein (B-E) and Maxwell-Boltzmann (M-B) distribution functions vs. x. From Fig. 12.1, the limit x >> 1, corresponding to εr − µ >> kB T , where all the distribution functions coincide, is the classical limit. For x → 0, the distribution function for bosons diverges in order to avoid the situation when µ > ε, as mentioned previously. In the next graphic Fig. 12.2, we explicitly show the distribution function for fermions. For T = 0, all states up to εr = µ are occupied (with one particle each (Pauli principle)) and all states with εr > µ are empty. For T = 0, the chemical potential coincides with the Fermi energy: µ = εF . For T > 0, fermions get statistically excited to higher energy states and their distribution function is as shown in Fig. 12.2. Exercise: find the particle fluctuations for bosons, fermions and classical particles. 162 CHAPTER 12. QUANTUM GASES Figure 12.2: Fermionic distribution function vs. temperature T . The relative particle fluctuations with respect to the average occupation number is given by: σn2 r 1 < n̂2r > − < n̂r >2 = = eβ(εr −µ) = z −1 eβεr = −c, 2 2 < nr > < nr > < nr > where z is the fugacity and the constant c is cM −B = 0 cB−E = −1 cF −D = +1 for classical particles (Maxwell-Boltzmann), for bosons (Bose-Einstein), for fermions (Fermi-Dirac). The particle fluctuations for bosons are larger than those for classical particles while for fermions they are smaller. The reason for that is that for fermions the change of a particle state is hindered by the Pauli principle while it is enhanced for bosons!