Você está na página 1de 41

COMMONWEALTH OF AUSTRALIA Copyright Regulations 1969 WARNING This material has been reproduced and communicated to you by or on behalf

of the University of Queensland pursuant to Part VB of the Copyright Act 1968 (the Act). The material in this communication may be subject to copyright under the Act. Any further reproduction or communication of this material by you may be the subject of copyright protection under the Act. Do not remove this notice.

Chapter 9 Degenerate Fermi Gases


Key Concepts
Examples. Fermi energy. Fermi temperature. Degeneracy condition (T TF ).

Reading
Schroeder, Section 7.3. Kittel & Kroemer, Chapter 7. Ashcroft & Mermin, Solid State Physics, Ch. 2. Kittel, Introduction to Solid State Physics, Ch. 6.

9.1

What is a degenerate Fermi gas?

A degenerate Fermi gas is a system of fermions where quantum statistics has a signicant effect on the macroscopic properties of the system. Examples include: Conduction (free) electrons in a metal Liquid 3 He Neutrons in a neutron star

100

Electrons in a white dwarf star Protons and neutrons in an atomic nucleus We assume the temperature is sufciently low that vQ = h 2mkB T
3

V . N

This is the opposite condition to that required for the validity of Boltzmann statistics. Here, we are considering the case where the fermions are packed very closely together so that quantum statistics are important. As a rst step, we take T = 0.

9.2

Zero temperature

nFD()

=F

Figure 9.1: Fermi-Dirac distribution at T = 0. All single particle states with < are occupied, and all with > are unoccupied mathematically, we obtain this by letting T 0 in the Fermi-Dirac distribution. Physically, it makes sense because at absolute zero we expect all particles to be in their lowest possible states. But because we are looking at fermions, we cant have more than one fermion per state (two if we consider spin). Therefore the particles ll up all the lowest energy states in order. The Fermi energy is dened by F (T = 0). N.B. The chemical potential varies with temperature we will show this in a moment. (9.1)

101

The Fermi energy gives the cut off energy for the distribution at T = 0. We can see that the above denition makes sense because = U N
S,V

U . N

(9.2)

At T = 0, if we add one particle (N = 1) then it must go in with an energy F (because that is the next available level) so the energy of the system goes up by U = F . Thus U = F . N The Fermi energy is determined by the total number of particles in the system the more particles, the more states are needed to place them all, and so the higher the Fermi energy. (T = 0) = We now calculate F for a system of N free electrons conned to a cube of volume V = L3 . We rst nd the allowed energy levels. For a one dimensional box, the allowed wavelengths and corresponding momenta are n = 2L , n pn = h hn , = n 2L

as we have already shown previously, and this extends to the three components of momentum in three dimensions. In three dimensions, the allowed energies are p2 , 2m 1 2 (p + p2 + p2 ), = y z 2m x h2 = (n2 + n2 + n2 ). (9.3) y z 8mL2 x In n-space, a constant energy surface is a sphere. The number of states with energy less than F is the volume of a sphere with corresponding radius nmax given by (p) = h2 2 n , (9.4) 8mL2 max Because all particles have energy less than F by denition, this must be N , and so we can nd nmax and hence F in terms of N . F = N = 2 volume of 1 of a sphere, 8

1 4 = 2 n3 , 8 3 max 3 n , (9.5) = 3 max where the factor of two takes spin into account, and the factor of 1/8 is because we only include nx , ny , nz > 0. The Fermi energy is then F h2 h2 2 nmax = = 8mL2 8m 102 3N V
2/3

(9.6)

(compare with the calculation of the Debye wavevector). Note that F only depends on the density N/V , meaning it is intensive if we double the size of the system, F does not change. We now reconsider the requirement for degeneracy V /N vQ . This is equivalent to kB T F . We can also dene the Fermi temperature by TF so that the degeneracy condition becomes T TF . (9.9) F , kB (9.8) (9.7)

In a typical elemental metal, n = N/V 1022 1023 cm3 , with F 110 eV. Hence, TF 104 105 K, and so most metals at room temperature are highly degenerate and it is often sufcient to consider them to be at T = 0!

Figure 9.2: Free electron Fermi gas parameters for metals at room temperature. Taken from Kittel, Introduction to Solid State Physics.

103

9.3

Internal energy at zero temperature

We can now calculate the internal energy per particle of a Fermi gas. We would expect that it would be approximately F /2, since the particles have all energies between 0 and F . U = 2
nx ,ny ,nz nmax

(n) 2 dn n
2 0
2

d3 n(n),
2

= 2
0 nmax

sin d
0

d(n),

=
0

(n)n2 dn,

nmax h2 = n4 dn, 8mL2 0 3 h2 5 n = N F . = 2 max 40mL 5 So the average energy per particle is 3F /5.

(9.10)

9.4

Degeneracy pressure

The Fermi gas has pressure even at T = 0 P = U V ,


S,N

= V =

3 h2 N 5 8m

3N

2 3

V 3

2N F 2U = . (9.11) 5V 3V This last equality is the same as for a classical ideal gas and is required by Newtons laws. We can also write 2 (9.12) P V = N kB TF . 5 This pressure is due to the exclusion principle, which prevents two fermions being in the same state. When the particles are degenerate (or close to it) they resist being pushed closer together. This degeneracy pressure prevents the gravitational collapse of white dwarf and neutron stars (see tutorial problem).

9.5

Bulk modulus

The bulk modulus is the change in pressure over the fractional change in volume a high bulk modulus means the material is difcult to compress, while an easily compressible material has a

104

Metal Li Na K Rb Cs Cu Ag Al

Free electron B Measured B (1010 dynes cm2 ) (1010 dynes cm2 ) 23.9 11.5 9.23 6.42 3.19 2.81 2.28 1.92 1.54 1.43 63.8 134.3 34.5 99.9 228 76.0

Table 9.1: Bulk moduli for some typical metals. The free electron value is that for a free electron gas at the observed density of the metal. small B. Mathematically, B = V where is the compressibility. For a Fermi gas,
5 2 F V 3 P = N = V 3 . 5 V V 2

P V

=
T

1 ,

(9.13)

So then,

2 F 2N 5 = kB TF . (9.14) B= P = N 3 3 V 3V Compare this to an ideal gas where B = P = N kB T /V (check this!). Electrons in a (room temperature) metal have a much higher bulk modulus (because usually TF T ) and are therefore less compressible than an ideal gas of similar density. Metals are composed of ions and electrons, and while the crystal of ions is hard to compress, the compressibility of metals is dominated by that due to the degeneracy pressure of the electrons.

105

Chapter 10 Degenerate Fermi gases at low temperatures


Key Concepts
T TF , but T > 0. Density of states. Temperature dependence of chemical potential (T ). Sommerfeld expansion. Heat capacity CV (T ) T . Pauli paramagnetism. Heavy fermions.

Reading
Schroeder, Section 7.3. Kittel & Kroemer, Ch. 7. Ashcroft & Mermin, Solid State Physics, Ch. 2, p. 661. Kittel, Introduction to Solid State Physics, Ch. 6.

10.1

Small non-zero temperatures

We now consider the case 0 < T F /kB = TF . This applies to metals at room temperature. Recall that the Fermi-Dirac distribution function looks like Figure 10.1, also see Figure 10.2.

106

I nFD()

T=0 KBT

Figure 10.1: Fermi-Dirac distribution at low temperatures. Now a fraction of the electrons, of order N kB T /F have kinetic energy kB T greater than F . Thus we expect the internal energy at low temperatures to be U (T ) U (T = 0) + aN U T kB T F kB T,

where a is a number of order one. The heat capacity at low temperatures will then be CV (T ) = a
2 N kB T, F

which is consistent with the third law of thermodynamics. We now derive this result in a more systematic manner.

10.2

Density of states

This is an important concept with very wide-ranging applications in quantum physics. We have already come across the density of states in treating both phonons and black body radiation, without actually identifying it. The density of states, denoted g() is the number of single particle states per unit energy the number of states near energy . If we let N (E) be the total number of single particle states with energy less than E then
E

N (E) =

d g()

It follows that g(E) = dN (E) , dE (10.1)

107

Figure 10.2: Fermi-Dirac distribution at various temperature for TF = 50000 K. The results apply to an ideal gas in three dimensions with a constant total number of particles. You can read off the chemical potential for each of these distributions from the graph the energy at which f = 0.5. Figure taken from Kittel, Introduction to Solid State Physics.

that is, the number of states between E and E + dE is g(E) dE. If {} denotes all the single-particle energy states, we can also write g() =

( )

(10.2)

where (x) is the Dirac delta function, dened such that (x) = for x = 0, 0 otherwise.

but with the special property that the function is normalised to one

dx (x) = 1.

It is also the derivative of the step function (x) which is dened to be (x) = 1 for x 0, 0 otherwise.

108

It follows that
E

N (E) =
E

d g(), d
E

= =

( ),

d ( ), (10.3)

(E ).

which is the total number of states with energy less than E, consistent with the original denition. For a system of N fermions at T = 0, F is dened by N = = =

no. of energy levels with F

(F ) g() d

(10.4)

The total energy at T = 0 is


F

U (T = 0) =

(F ) =

d g().

(10.5)

The above expressions generalise naturally to nite temperature: N (T ) =


nF D ( ) d g()nF D (),

= and U (T ) =

(10.6)

nF D ( ) d g()nF D ().

(10.7)

In words, to nd U we sum over all energies, adding the number of states with each energy times the (average) number of particles in each of those states. The Fermi-Dirac distribution at T = 0 is a backwards step function with the jump down at F , and so the general expressions above are consistent with the zero temperature limit.

10.2.1 Density of states in a three dimensional cube method 1


We now nd the density of states for a Fermi gas conned to a three dimensional cube, subject to the boundary condition that the wave function vanishes on the surface of the cube.

109

Previously, we showed that the allowed values of the energy of a single state particle (fermion or boson) are (p) = for nx , ny , nz = 1, 2, . . .. Then N (E) = total no. of states with energy E, 1 = volume of of the sphere in n-space, 8 1 4 3 = n , 8 3 0 where E = h2 n2 /8mL2 . This is similar to what we did for the T = 0 case, except that we are 0 no longer saying that all these states are lled, i.e. N = N . (We are considering particles with no spin.) We thus have N (E) = 6 8mL2 E h2 2m h2
3/2

p2 h2 = (n2 + n2 + n2 ), y z 2m 8mL2 x

,
3/2

4 3 L = 3 And so

E 3/2 .

(10.8)

3/2 2m dN (E) 3 = 2L g(E) = E. (10.9) dE h2 For fermions with spin-half we multiply this by two, because there are twice as many states.

10.2.2 Density of states in a three dimensional cube method 2


Lets repeat the derivation using the denition g() =

( ),

where the sum in this case is over all n = (nx , ny , nz ), and h2 (n2 + n2 + n2 ). y z 2 8mL x We convert the sum to an integral (in polar co-ordinates) and if the energy spacings are small, this is a good approximation. (n) = g() =

h2 |n|2 , 8mL2
/2 /2

(10.10) d h2 2 n , 8mL2 (10.11) (10.12)

=
0

dn n 2
0

2 0

sin d
0

dn n2

h2 2 n . 8mL2

110

When working with this denition of the density of states, the next step is to make a substitution in the delta function. Here we set h2 2 x= n, 8mL2 so n = 2L n dn = 8L Thus, g() = 4L
3 2 3

2m h2 2m h2

1/2

x, 1 dx . 2 x

3/2

2m h2 2m h2

3/2 0 3/2

dx x( x),

= 2L3

which agrees with the previous result. We also dene g( 0) = 0 there are no states with energy less than zero in this case. Such states can correspond to bound states such as occur in potential wells, etc.

10.3

Temperature dependence of the chemical potential

For a general density of states g(), we said before that

N =

d g()nF D (), d g()

1 e()/kB T +1

(10.13)

If N is xed, then Eq. (10.13) determines as a function of temperature as T increases, must decrease. To see why, note that nF D has some symmetry about = , in the sense that the probability of a state at energy + being occupied is the same as the probability of a state at being unoccupied. The total number of particles in the system is determined by Eq. (10.6). Lets consider rst a constant density of states g() g. If the number of particles in the system is conserved, then at non-zero temperatures, the area under the graph should still be N . Because of the symmetry of nF D about , the missing area below from the T = 0 distribution must be equal to the additional area above , and so the total area is indeed still N .

111

g (E)

g (E)

g(E)nFD(E) area = N area = N

F= (T = 0)

F= (T = 0)

Figure 10.3: Non-constant density of states Now, however, consider the case where the density of states is not constant, and increases with , as shown in Fig. 10.3. Now, the area under the graph to the right of will be scaled more than the empty area above the line to the left, due to the higher density of states on the right. Therefore, the area of the graph is greater than N , and it would appear the number of particles has increased! This problem is resolved if, as the temperature increases, decreases slightly, so that the peak of nF D moves to the left, and isnt scaled as much by g(), reducing the overall size.

Figure 10.4: Temperature dependence of for an ideal Fermi gas in a box. Figure taken from Schroeder.

112

-dnFD /d

~5k BT

Figure 10.5: The derivative of the Fermi-Dirac distribution is negligible everywhere except within a few kB T of .

10.4

The Sommerfeld expansion

We now explicitly nd (T ) when kB T F = (T = 0). From

N=

d g() F D (), n

(10.14)

we rst integrate by parts to get


N = N () F D () n

d N ()

d F D n (). d

(10.15)

As , N () 0 and nF D () 1. Also, as +, nF D () 0, so the boundary term is zero. For kB T F , nF D / is sharply peaked at = . We can see this mathematically by d F D n d 1 = , d d e()/kB T + 1 1 ex = . kB T (ex + 1)2

(10.16)

where x = ( )/kB T . We now change our integration variables in Eq. (10.14) from to x

N=

dx N ( + xkB T )

ex . (ex + 1)2

(10.17)

We perform a Taylor expansion of N () about = . This is reasonable, because the total number of states varies slowly on the energy scale of kB T when T is small. 1 N ( + xkB T ) = N () + N ()xkB T + N ()(xkB T )2 + . . . 2 113

However, by denition, N () = so that 1 N = N ()I0 + kB T g()I1 + (kB T )2 g ()I2 , 2 where In = I0 I1 I2 Thus, we nd that N = N () + Now, we let (T ) = F a(kB T )2 + . . . (10.20) [remembering that F (T = 0)] and substitute in this expression and work consistently to second order in T : N = N (F ) a(kB T )2 g(F ) + 2 g (F )(kB T )2 . 6 (10.21) ex , (ex + 1)2 1 d = 1, dx = x+1 dx e = 0, since the integral is antisymmetric about x = 0 2 . (See Appendix B5 in Schroeder) = 3 dx xn 2 g ()(kB T )2 + . . . 6

N ()

= g( = ),
=

(10.18)

(10.19)

But by denition of the Fermi energy, N (F ) = N there must be enough states with energy less than F to contain all the particles. Thus, we must have a= and so to second order in kB T /F , 2 g (F ) (kB T )2 , 6 g(F ) For a three dimensional system we have g() so (T ) = F g (F ) 1 = g(F ) 2F and so we do have an expansion on powers of kB T /F . (10.23) 2 g (F ) , 6 g(F ) (10.22)

114

10.5

Heat capacity

We now evaluate the internal energy

U (T ) =

d g() F D (), n

(10.24)

for kB T F . The calculation is similar in spirit to that for nding (T ). Let h() be dened by

h() =

dE [E g(E)],

(10.25)

noting that h(F ) = U (T = 0). Integrating U (T ) by parts gives

U (T ) =

d [g()] nF D (), d

d d

dE [Eg(E)] nF D (),

= h() F D ()| n

d h()

d nF D () . d

(10.26)

Again, because nF D () 0 as and h() 0 as , the boundary term is zero. Proceeding as before,

U (T ) =

dx h( + xkB T )

ex . (ex + 1)2

(10.27)

We can expand h as a Taylor series 1 h( + xkB T ) = h() + xkB T h () + (xkB T )2 h () + . . . 2 and substituting this into the integral gives U (T ) = h() + Then, we previously showed that (T ) = F 2 g (F ) (kB T )2 , 6 g(F ) (10.30) 2 (kB T )2 h () + . . . 6 (10.29) (10.28)

so we can Taylor expand h() around F , giving h() = h(F ) h (F ) From the denition of h(), h (F ) = F g(F ), h (F ) = g(F ) + F g (F ). (10.32) (10.33) 2 g (F ) (kB T )2 . 6 g(F ) (10.31)

115

Collecting terms, we have 2 2 F g (F )(kB T )2 + (g(F ) + F g (F )) (kB T )2 . 6 6 The second and fourth terms cancel exactly, leaving U (T ) = U (T = 0) U (T ) U (T = 0) = It then follows immediately that CV (T ) = where U T = T,
V

(10.34)

2 g(F )(kB T )2 6

(10.35)

(10.36)

2 2 g(F )kB . 3 For non-interacting electrons in three dimensions, the Fermi energy is = h2 F = 8m and the density of states per unit volume is
3 (8m) 2 F , 3 2h which can be rewritten in terms of the Fermi energy as

(10.37)

3N V

2 3

(10.38)

g(F ) =

(10.39)

g(F ) =

3N/V . 2F

(10.40)

Notice that g(F ) is completely determined by the conduction electron density N/V . We can also write m kF , (10.41) g(F ) = ( )2 where kF is the Fermi wave number. Note that for xed density, g(F ) scales with m. If g(F ) is larger than for the free (non-interacting) electron model, we say the electrons have an effective mass given by obs m = me . (10.42) free

10.6

Low temperature heat capacity of metals

For the total heat capacity of metals, we have now calculated that at low temperatures we have CV (T ) = T + T 3 . electrons phonons

Thus, plotting CV (T )/T against T 2 should be a straight line with gradient and intercept at T = 0 of .

116

Figure 10.6: Low-temperature measurements of the heat capacities per mole of copper, silver and gold. Figure taken from Schroeder.

10.7

Magnetic susceptibility of metals Pauli paramagnetism

In a magnetic eld B, the energy of the electrons becomes = + B B, = B B, (10.43) (10.44)

due to Zeeman splitting, where is the energy in the absence of a magnetic eld. (The arrow subscript represents whether the electron spin is parallel or anti-parallel to the magnetic eld. Remember the intrinsic magnetic moment of elecrons is in the opposite direction to its spin as it is negatively charged.) If we neglect the orbital motion of the electron, the density of states is 1 g () = g( B B), (10.45) 2 1 g () = g( + B B), (10.46) 2 i.e. the density of states of the spin up electrons (for example) is half the density of states in no magnetic eld at an energy B B less than . The total number density of electrons of spin = (, ) is n = d g () F D (), n (10.47)

117

Figure 10.7: Experimental values of electronic heat capacity. Taken from Kittel, Introduction to Solid State Physics.

where nF D () = The chemical potential (T ) is determined by n = n + n , where n = d g () F D (). n (10.49) 1 e() +1 . (10.48)

g() varies on an energy scale of F so since B B F , to a good approximation we have g () = 1 g( B B), 2 1 1 = g() B B g (). 2 2

(10.50)

and therefore g () + g () = 1 1 g() B Bg () + 2 2 = g(). 1 1 g() + B Bg () , 2 2 (10.51)

118

Thus, n = n + n , = = d g () F D () + n d g() F D (). n d g () F D (), n (10.52)

which is the same as the zero-eld equation. Hence, we can neglect the B eld dependence of and from Eq. 10.23, we have (T ) = F 1 + O The magnetisation is M = B (n n ),

kB T F

(10.53)

= B = 2 B B

d nF D () [g () g ()] , d nF D ()g (). (10.54)

Integrating by parts, and using the fact that nF D () 0 as and g() 0 for 0

M = 2 B B

d g()

nF D ()

For kB T F , the function nF D ()/ is sharply peaked at = F and so to leading order = in (kB T /F )2

d g()

nF D ()

= g(F )

nF D ()

, (10.55)

= g(F ). The magnetic susceptibility is given by = M 3N 2 B = 2 g(F ) = . B B 2V kB TF

(10.56)

Note that > 0 i.e. this is paramagnetism. Working to second order in (kB T /F )2 we nd (T ) = 2 g(F ) 1 B 2 12 kB T F
2

(10.57)

Note that this weak temperature dependence is in contrast to that for immobile spin-half particles which obey Curies law 1 (T ) . T

119

Figure 10.8: Temperature dependence of the magnetic susceptibility of metals. Figure taken from Kittel, Introduction to Solid State Physics.

120

Metal rs /a0 Li 3.25 Na 3.93 K 4.86 Rb 5.20 Cs 5.62

Theory 106 0.80 0.66 0.53 0.50 0.46

Measured 106 2.0 1.1 0.8 0.8 0.8

Table 10.1: Comparison of free electron and measured Pauli susceptibilities. The discrepancies between theory and experiments are mainly due to electron-electron interactions.

Figure 10.9: Temperature dependence of magnetic susceptibility of liquid 3 He. (T ) const for T TF 1 K. (T ) C/T for T > TF (Curies law). Taken from A. L. Thomson et al., Phys. Rev. 128, 509 (1962).

121

Figure 10.10: Temperature dependence of the heat capacity of liquid 3 He. CV (T ) T for T TF 1 K. m increases with pressure. Taken from D. S. Greywall, Phys. Rev. B 27, 2747 (1983).

122

10.8

Heavy fermion metals

To conclude this chapter we briey mention heavy fermion metals. These are systems for which the interaction between neighbouring electrons so strong that electrons cannot be considered independently of one another. They were discovered in the late 1970s, and some of them are superconductors. Examples include UPt3 , UBe13 , CeCu2 Si2 , LiV2 O4 , . . . . Some typical parameters for these materials are m 1 J/mol K2 , 1001000. me e (T ) is also enhanced by a factor of 1001000 compared to elemental metals.

123

Chapter 11 Degenerate Bose Gases


Key Concepts
Macroscopic occupation of the ground state Bose-Einstein condensation (BEC) Critical temperature for BEC Superuid 4 He Two uid model of superuidity Ultra-cold atomic gases

Reading
Schroeder, Section 7.6 Kittel & Kroemer, Chapter 7. Anglin and Ketterle, Nature 416, 211 (2002). Nobel prize article 2001, linked from course webpage. Bose-Einstein Condensation in Dilute Gases, Pethick and Smith, QC175.47.B65 P48 2002. (advanced textbook on BEC.)

124

11.1

Macroscopic occupation of the ground state

Earlier in the course we learnt that there are two types of particles in nature bosons and fermions. The difference between the two is that while Pauli exclusion principle applies to identical fermions, any number of identical bosons can co-exist in the same quantum state. We have already derived the Bose-Einstein distribution for the mean occupation number of bosons occupying a single quantum state of energy s nBE (s ) = 1 e(s )/kB T 1 .

The chemical potential is something of a mysterious quantity, but (in my opinion) the easiest way to deal with it is to treat it as a normalisation constant for an isolated system. Because we know that physically the mean occupation number of all quantum states in a system must be positive, this means that we must have < 0 , the lowest energy state in the system. However, can be arbitrarily close to 0 if we set = 0 , then nBE (0 ) = 1 e/kB T 1 1 kB T = . 1 + /kB T 1

Thus if is really small, then we can have a huge number of atoms in the lowest energy level. Obviously there must be some constraint on , because in a real system there are only a nite number of particles.

11.2

More on the density of states

In this chapter we will consider a system of N atomic bosons of mass m conned to a volume V = L3 (an innite box potential.) However, in real experiments trapping potentials are usually not spatially homogeneous, and it is worthwhile to be able to calculate the density of states for these systems for more general treatments of BEC. For systems where the trapping potential varies slowly on the length scale of the de Broglie wavelength T , we can calculate the density of states for a three dimensional system as g() 1 h3 d3 rd3 p [ H(r, p)],

where H = p2 /2m + U (r). In your assignment you will show that g() = 2 2m h2
3/2

d3 r
V

U (r).

For the innite box, the spatial integral simply gives the volume, and leaves us with the same result from earlier 3/2 2m g() = 2V . h2

125

11.3

Bose-Einstein condensation (BEC)

Consider N bosons of mass m conned to a volume V = L3 . At zero temperature all of the N atoms will be in the ground state which has an energy of h2 3h2 2 2 2 0 = . (11.1) 1 +1 +1 = 8mL2 8mL2 At very low temperatures it is possible to nd a macroscopic number N0 of the N bosons still in the ground state. The value of N0 is given by the Bose-Einstein distribution function . (11.2) 1 By taking the Taylor expansion of the denominator, we can nd that the chemical potential as a function of temperature is given by N0 = e(0 )/kB T kB T (T ) 0 as T 0. = N0 Hence for low temperatures (T ) is very close to the energy of the ground state. The chemical potential can be determined by calculating the total number of bosons N=
s

(11.3)

nBE (s ) =
s

1 e(s )/kB T 1

(11.4)

We can approximate this sum by an integral as long as kB T is signicantly larger than the energy level spacing. However, if there are a large number of bosons in the ground state, there will be a spike in the function nBE (). This is not accounted for correctly in the integral. In fact, this approximation may be bad for several of the lowest energy levels. However, only the ground state will be wildly incorrect, and so we treat N0 separately from the integral. Thus we can write

N N0 +

g()
0

d e/kB T 1

(11.5) (11.6)

N0 (T ) + Nex (T ),

where g() is the density of states and Nex (T ) is the number of bosons in all the excited states. For now, we will assume that (T ) = 0 when there is a large occupation of the ground state. We can do this because the energy of the ground state approaches zero as the volume becomes large; any change in will have negligible effect on the occupation of the excited states when N0 is large. We now have 2 Nex (T ) = 2 = 2m h2
3/2

V
0 3/2

e/kB T

2mkB T h2 126

d 1 xdx x1 e

(11.7)

where we have made the substitution x = /kB T . This integral can be tackled by expanding the fraction as an innite series (see assignment question). The general result is

dx
0

xn1 = (n)(n), ex 1

(11.8)

where the gamma function (n) interpolates the factorial function (i.e. for integer n, (n) = (n 1)!) and the zeta function (n) is dened as (n) = For the 3D box, we have n = 3/2, so (3/2) = Nex (T ) = 2.612 where vQ = 3 is the quantum volume. T We can see from the equation for Nex that as T increases, the value of Nex also increases. We dene the critical temperature Tc as the temperature at which the number of excited particles Nex is equal to the total number of particles Nex (Tc ) = N. (11.10) 1 . mn m=1 /2 and (3/2) = 2.612. Thus
3/2

2mkB T h2

V = 2.612

V , vQ

(11.9)

So, as the temperature approaches the critical temperature the number of particles in the ground state N0 (T ) approaches zero. Since the initial assumption that (T ) = 0 was based on the fact that a large number of particles were in the ground state, for temperatures greater than the critical temperature the assumption that (T ) = 0 is no longer valid. A Bose-Einstein condensate forms when the temperature is below the critical temperature, which can be written as kB Tc = 0.527 h2 2m N V
2/3

(11.11)

The fraction of particles in the ground state of a system of bosons is given by N0 (T ) = N 1 (T /Tc )3/2 T < Tc , 0 T Tc . (11.12)

Note that Bose-Einstein condensation occurs approximately when the quantum volume vQ is equal to or greater than the volume per particle V /N this is a simple physical picture.

11.4

Why does it happen?

From the maths we have seen that BEC does occur, but this doesnt give much understanding of why it occurs. The answer to this question was hinted at in the tutorial on quantum statistics.

127

For a gas of noninteracting distinguishable particles, we can treat them all separately using Boltzmann statistics. A single particle has a reasonable chance of being found in any single particle state whose energy is of order kB T . The available number of such single particle states is quite large usually, and is of order of the single particle partition function Z1 . The probability of being in the ground state is thus 1/Z1 , and as this applies to each particle individually, only a miniscule fraction will ever be found in the ground state away from T = 0. From the perspective of the system, each total system state has its own probability. The conguration with all particles in the ground state will have a Boltzman factor of 1, whereas the Boltzmann factor of a typical system state at temperature T with U N kB T will be eN kB T /kB T = eN . This seems rather small however, we must remember to multiply it by the number of different N arrangements of the N particles. For Z1 single particle states, the number of arrangements is Z1 , and this factor overwhelms the eN . Now think about the case of identical bosons. When there are more available states than particles (Z1 N ) then the situation is similar to that of distinguishable particles. However, when we have Z1 < N , then there are far fewer arrangements of identical bosons as compared to distinguishable particles. Thus the probability of the system being in its ground state is much higher than for distinguishable particles. The number of system states Ns is roughly the number of ways of arranging N indistinguishable particles among Z1 single particle states which is mathematically the same as distributing N units of energy among Z1 oscillators. It is given by Ns N + Z1 1 N (eZ1 /N )N (eN/Z1 )Z1 when Z1 N, when Z1 N. (11.13)

If Z1 N then (eZ1 /N )N eN , then the probability of the system being in an excited state is very close to one. However, if Z1 N , then the total number of states is not very large, and the eN term becomes more signicant making the probability of the system being in an excited state very low, thus allowing a Bose-Einstein condensate to form. It is worth comparing the transition temperature with the energy of the ground state of the system kB Tc = 0.527 0 = h2 2m N V
2/3

= N 2/3

0.527h2 , 2mL2

3h2 . 8mL2 Both are of order h2 /mL2 , however the BEC temperature is enhanced by a factor of N 2/3 .

11.5

Important points

11.5.1 Thermodynamic quantities


The behaviour of the condensate occupation, the chemical potential, and the heat capacity of the Bose gas in the region of Tc is shown in Fig. 11.1 for the case of the innite box potential. All

128

thermodynamic quantities can be calculated relatively easily, but we have to resort to numerical solution and we will not do so in the course. However, the same general principles as used earlier apply. An important point to note is the peak in the heat capacity at Tc . This is relevant in the case of superuid helium (see later). The condensate fraction of the Bose gas does not contribute to the heat capacity or the entropy of the system. The heat capacity behaves as CV T 3/2 below Tc , and reduces to 3N kB /2 well above Tc .

11.5.2 BEC in other potentials


BEC does not occur in every Bose gas system, and it is worth looking at the maths to understand why. A BEC forms when we have 0 so that the ground state contains a macroscopic number of particles. This will only happen if the integral for the number of excited particles Eq. (11.7) reaches a limit as 0. From the result given in Eq. (11.8), if the density of states is given by g() n1 , then we must have n > 1 so that the zeta function (n) = 1 , mn m=1

will converge. If n 1, the result for the number of excited state particles diverges, and so our approximation that 0 was incorrect. This means that the excited states can accomodate all particles at any temperature, and there is no BEC. For a 2D homogeneous system, we have g() constant, and so there is no BEC.

11.5.3 Massless bosons


We have already considered ideal gases of bosons both photons and phonons. However, we do not get BEC in these systems. The reason is that there is no restriction on the particle number phonons and photons can be created and destroyed at will, and so = 0 always. It is the requirement that particles be conserved that results in the formation of a BEC.

11.5.4 Examples of BEC


The phenomenon of BEC and macroscopic quantum phenomena are intricately linked, and there are three real-world examples of where Bose-Einstein statistics are important. The rst is in superconductors, where electrons can form bound Cooper pairs below a transition temperature. They can be thought of as composite bosons forming a BEC, and can ow without resistance through the conductor. The second two examples are superuid helium and trapped atomic gases, and we will consider these in more detail in the following sections.

129

N N0 Nex

(a)
0 0.5 1 1.5 2

T / Tc
0 0.2

/ kB T

0.4 0.6 0.8

(b)
1 0 0.5 1 1.5 2

T / Tc
2

1.5

CV / N kB

0.5

(c)
0 0 0.5 1 1.5 2

T / Tc
Figure 11.1: Thermodynamics of a Bose gas in the region of the critical temperature for BoseEinstein condensation. (a) The behaviour of the ground state occupation. (b) The chemical potential as a function of temperature. (c) The behaviour of the heat capacity. Note the peak at Tc , and CV 3/2N kB for T Tc . 130

Figure 11.2: The specic heats of (a) 4 He and (b) 3 He showing the anomalies associated with their respective phase transitions. Note that the temperature scales in (a) and (b) are quite different, being a factor of 103 lower for 3 He.

11.6

Superuid helium

Helium-4 is a bosonic atom that liquies at a temperature of 4.2 K at atmospheric pressure. However, it undergoes another phase transition at Tc = 2.17 K to a superuid phase (called Helium-II) that has essentially zero viscosity. This transition occurs at the lambda-point, so called because the heat capacity diverges at Tc , and a plot of CV versus T is similar to the Greek letter see Fig. 11.2. The phenomenon was discovered by Kapitsa, Allen and Misener in 1938. In 1962 Landau received a Nobel prize for his theory of superuidity and in 1978 Kapitsa also received a Nobel prize for his work in superuidity. The superuid state involves macroscopic quantum coherence, similar to that responsible for superconductivity. The theory of BEC in an ideal gas of atoms was rst published by Einstein in 1924 (before fermions were known about), but was pretty much ignored for several years. In fact, it was even suggested

131

Figure 11.3: Phase diagrams of 4 He (left) and 3 He (right). Neither diagram is to scale, but qualitative relations between the diagrams are shown correctly. Not shown are the three different solid phases (crystal structures), or the superuid phases of 3 He below 3 mK.

by Uhlenbeck that the prediction of a phase transition was incorrect (and there is evidence that Einstein accepted this criticism!) However, when superuidity was discovered Fritz London suggested that Bose condensation could be at the heart of the physics. In fact, using the ideal gas formula for the ideal gas BEC transition temperature gives a good estimate for 4 He. This idea was further strengthened when it was found there was no similiar superuid transition in fermionic 3 He (see Fig. 11.3). The existence of a BEC in superuid 4 He was argued about for many years, but now it is accepted that this is the case. [N.B. there is a superuid transition in 3 He, but at milliKelvin temperatures. Here the fermions form Cooper pairs, similar to electrons in a superconductor. Nobel prizes were awarded for this work in 1996 (Lee, Osheroff, and Richardson, experiment) and 2003 (Leggett, theory).] There is one important fact about superuid 4 He it is necessary to consider the interactions between the atoms to explain superuidity, and so it is necessary to go beyond the ideal gas model. The interactions are very strong, and the system is difcult to treat theoretically. However, the weakly-interacting Bose gas turns out to be a reasonable qualitative model, and much work was done on this in the 1950s and 1960s in the context of superuid 4 He.

11.6.1 Two uid model of superuidity


For Bose-Einstein condensates we have N = The total number of bosons, = N0 (T ) + Ne (T ), = Number in the ground state + number in excited states. It turns out that a very good model for superuid helium is to treat it as a mixture of two components, superuid and normal uid, with the proportion of each component depending on the

132

Property Viscosity Thermal conductivity Entropy

Superuid 0 0

Normal uid non-zero nite nonzero

Table 11.1: Properties of superuid and normal uid components temperature. In terms of density we write = s (T ) + n (T ). (11.14)

However, it is not the case that the superuid component directly corresponds to atoms in the ground state of the system. In fact, by using neutron scattering methods it has been determined that even in the absence of a normal component only about 10% of the atoms are in the condensate state. This is due to the strong interactions between the atoms this results in large depletion of the condensate.

11.6.2 Physical properties


Superuid 4 He does not boil, i.e. no bubbles form when it is heated. This is because thermal conduction within the 4 He occurs via the superuid component, which has innite thermal conductivity. This means that if one part of the uid is heated, the heat is transferred throughout the uid immediately. Thus it is impossible to set up a temperature gradient within the uid (which is required to form bubbles.) Heat transfer within a superuid occurs not by conduction, but by the movement of the normal component. If a heat source is placed somewhere within a superuid, and a heat sink placed elsewhere within the superuid, the normal uid component moves from the source towards the sink, and the superuid component moves in the opposite direction. This motion occurs under the constraint that the total density remains constant throughout the superuid. This fact results in some remarkable properties. Table 11.1 contains a comparison of some of the properties of the superuid and normal uid components.

11.6.3 Superleaks
A superleak is a ow of the superuid component through a highly non-porous material that the superuid can ow through, but the normal uid component cannot pass through due to viscosity. If in the system shown in Fig. 11.4 we have T1 > T2 and P1 = P2 , then for a normal connection thermal equilibrium is reached by the conduction of heat. Thus the uid carries entropy from 1 to 2. For a superleak this is NOT possible because only the superuid can pass from 1 to 2 and the superuid carries no entropy. Suppose T1 were to slowly increase. For thermal and diffusive

133

T2 ,P2 T1 ,P1 Superleak 1 powder superfluid


Figure 11.4: Representation of a superleak. The superuid can pass through the powder due to its vanishing viscosity, however the normal component cannot.

equilibrium to hold it is required that 1 = 2 . Using the thermodynamic identities G = N = U T S + P V, and T dS = dU + P dV dN, we get the equation N d = SdT + V dP From the conditions for equilibrium we know that 1 2 = 0 = S(T1 T2 ) + V (P1 P2 ), and therefore P1 P2 = This leads to several interesting phenomena. S (T1 T2 ). V (11.17) (11.16) (11.15)

11.6.4 Thermomechanical effect


It is possible to establish a pressure difference between two superuids connected by a superleak by creating a temperature difference. The capillary in Fig. 11.5 contains a superleak. The heater in the main section raises T1 above T2 , and this causes P1 to become greater than P2 . The result is that the superuid is forced through the capillary.

134

Figure 11.5: A fountain based on the thermomechanical effect

11.6.5 Mechanocaloric Effect


Similarly, creating a pressure difference between two systems connected by a superleak will cause a temperature difference to arise. This is in contrast to a normal uid, for which a pressure difference causes a mass ow. The magnitude of this effect can be found as follows V T = . P S If we let P = gz, then this gives T g = , z S (11.18)

where is the mass per mole and S is the entropy per mole. For superuid 4 He at around T = 1.3 K we have T 1 mK cm1 . z Thus a temperature difference of only 1 mK will produce a height difference of 1 cm! The quantitative details of this effect have been experimentally veried.

135

11.6.6 Other phenomena


Superuid helium displays a variety of other interesting phenomena. These include Second sound temperature waves rather than pressure waves, caused by the normal and superuid fractions moving out of phase. Vortices and quantised circulation.

11.6.7 Video
If we have time, part of the video Superuid helium by J.F. Allen, University of St. Andrews will be shown. Jack Allens obituary follows at the end of this chapter.

11.7

Ultra-cold Bose gases

In 1995 at the University of Colorado in Boulder, a team of physicists lead by Eric Cornell and Carl Wieman suceeded in creating the rst ever Bose-Einstein condensate in a dilute gas of 87 Rb atoms. The condensate they formed consisted of only 2000 atoms at a temperature of about 100 nK. This temperature was achieved by rst laser cooling in a vacuum to microKelvin temperatures, before transferring the atoms to a magnetic trap. The walls of the trap were then lowered, allowing evaporative cooling to occur. In this process the hottest atoms in the velocity distribution escape, and the remaining atoms cool via collisions and rethermalisation (the process by which the atoms return to equilibrium). At these temperatures you would normally expect 87 Rb to be a solid! However, if they are kept sufciently dilute only two particle collisions will ever occur, and these must be elastic. Three particles are required to collide such that two form a molecule, and the third takes away the excess energy. Therefore the gas will be metastable however this requires it to be 100,000 times less dense than air! The important feature of BEC in atomic gases is that the interactions are much, much weaker than in superuid helium, and therefore the system is much closer to the ideal Bose gas that Einstein rst imagined. In fact, quantities such as Tc are not much shifted from the values you will calculate in your assignment. This also means that it is feasible to treat these systems theoretically (although it is still not easy!) A wide variety of interesting phenomena have been observed in gaseous Bose-Einstein condensates, including vortices, vortex lattices, ultra-slow light (17 m s1 ), molecular BEC, four-wave mixing, entangled atomic beams, BECs in periodic lattices, and so on. One of the most talked about possibilities is to make atom lasers from BECs, consisting of a continuous stream of atoms sharing the same quantum mechanical phase. The Nobel Prize was awarded to researchers in this eld in 1997 (laser cooling) and 2001 (BEC), and it remains a very active area today. There is

136

Figure 11.6: Shadow images of a degenerate Bose gas of 87 Rb, taken at UQ in February 2004. From left to right: above Tc , slightly below Tc , well below Tc . There are about 30,000 atoms in the picture on the right.

a large theory team at UQ (the Centre of Excellence for Quantum-Atom Optics), as well as an experimental BEC group.

11.8

Ultra-cold Fermi gases

Even more recently, interest has exploded in the eld of degenerate Fermi gases. In a similar manner to Bose gases, they can be cooled to very low temperatures. However, it is more difcult with fermions, as the collisional process between identical fermions that is required for evaporative cooling turns off at very low temperatures. The solution is either to use a Bose gas refrigerator, or to cool and trap two different spin states simultaneously. Currently people are working on trying to observe superuidity in two-component degenerate Fermi gases, in the hope that this will teach us about the nature of high temperature superconductivity. In particular there is interest in what is known as the BEC-BCS crossover the transition from having a BCS-like state of weakly paired fermions forming the superuid to having strongly bound fermions making up a bosonic molecule and BEC. This is also a topic of research at UQ. It is predicted that in the next few years a Nobel prize is likely to be awarded in this area as well.

137

news and views


Obituary

John Frank (Jack) Allen (19082001)


In late 1937, two young physicists at the Royal Society Mond Laboratory in Cambridge, UK, found that liquid helium could flow through very small capillaries with essentially zero viscosity below a temperature of 2.17 kelvin. While they were preparing a note for publication, they heard of another paper, just submitted by P. L. Kapitza in Moscow, reporting similar results. Both papers appeared side by side on 8 January 1938 in Nature. Kapitza and the young physicists Jack Allen and A. D. Misener had discovered the mysterious phenomenon of superfluidity in liquid 4 He. It was the start of a golden period in low-temperature physics. Allen, the last survivor of this heroic time, died of a stroke on 22 April 2001, aged 92. Most of Allens greatest discoveries were made at the outset of his career (in 1938 he was essentially a postdoc, working without a supervisor, which suited him fine). Between 1937 and 1939 he and his associates at Cambridge produced a stream of papers on superfluidity in liquid 4 He, this output coming to an abrupt end with the start of the Second World War. In 1946, with normal life resuming, Allen organized the first international lowtemperature meeting at the University of Cambridge, UK. This was the birth of the major triennial event in the lowtemperature physics community; the twenty-third meeting will be held in Hiroshima next year. In 1947, Allen was appointed professor of natural philosophy and head of physics at the University of St Andrews in Scotland, and two years later he became a Fellow of the Royal Society. But although he stayed active in the world of lowtemperature physics, revitalizing research at St Andrews and helping to run many conferences and workshops, his own research was never again central. He retired from St Andrews in 1978. Jack Allen was born in Winnipeg, Canada, in 1908 (the year 4He was first liquefied in Leiden, in the Netherlands). He received his BA from the University of Manitoba, where his father was head of the physics department. In 1929, he went to study at the University of Toronto where James F. McLennan had built up a strong physics department. In 1923 this had been only the second laboratory in the world to liquefy helium, and when Allen arrived he immersed himself in the new world of cryogenics. He received his PhD in 1933, having already written ten papers on superconductivity. Late in 1935 he went
436

Co-discoverer of superfluidity

to Cambridge to work with Kapitza, only to find that Kapitza was being detained in Moscow, and was in the process of setting up what was to become the famous Institute of Physical Problems. In Kapitzas absence, Allen effectively took over the low-temperature work, although officially the director was J. D. Cockroft. In 1935, Don Misener, a graduate student at Toronto, had carried out the first experimental study of the viscosity of liquid 4He. By then it was known that liquid helium underwent some sort of phase transition at 2.17 K, as there were abrupt changes in various thermodynamic properties. Miseners work suggested that the viscosity decreased substantially when the liquid passed through this transition. Misener joined Allen at Cambridge in 1937 to do his PhD, and the two set out to study the phenomenon by examining flow in thin capillaries. Miseners 1935 experimental work had also attracted the notice of Kapitza in Moscow. Both groups reported their independent discovery of superfluid flow in 1938, with Kapitza being the first to coin the term superfluid. It is puzzling and unfortunate that when Kapitza finally received a well-deserved Nobel prize in physics in 1978, the citation concerning superfluidity made no reference to the work of Allen and Misener. Allen quickly found other dramatic manifestations of superfluidity, all of which involved the counterflow of the normal and superfluid components or the clamping of the normal component with
2001 Macmillan Magazines Ltd

fine powder. But after 1945 the Moscow group under Kapitza (helped by L. D. Landau, who developed a complete theory of the two-fluid behaviour of superfluid helium in 1941) dominated further research on quantum liquids. The study of superfluid 4He increasingly involved microscopic theories and new experimental probes such as neutron scattering, none of which interested Allen. Allen was the last of a generation of independent-minded classical physicists who delighted in explaining the visible world. He prized his own ability, and that of glass-blowers and technical people, to build experimental apparatus. Allen was as proud of his invention of the O-ring vacuum seal as anything else he did. It should be no surprise that his greatest work on superfluidity in liquid 4He involved phenomena that could be seen. Indeed, it is most fitting that Allen discovered the famous fountain effect in 1938 with the help of a pocket flashlight. Over a ten-year period Allen made a movie of the various two-fluid phenomena exhibited by liquid 4He. The photography of these effects was a real challenge, because liquid 4He is essentially transparent. This unique colour movie (the fifth edition was completed in 1982) is one of Allens great legacies to physics. Allen had a commanding presence and a dry sense of humour. He strongly identified with the physics of an earlier day, and I imagine that he would have enjoyed talking with classical physicists such as Lord Rayleigh, Michael Faraday and Daniel Bernoulli more than with Werner Heisenberg and Erwin Schrdinger. Superfluidity is a dramatic visible manifestation of quantum mechanics, being the result of BoseEinstein condensation in which a macroscopic number of 4He atoms occupy the same, single-particle quantum state. It is paradoxical that the phenomenon was first observed by Allen: a great physicist who wasnt much interested in atoms. Walking the old stone streets of St Andrews, one quickly notices the elegant metal historical plaques on many buildings, commemorating famous people who have been associated with this historic town and its university over the centuries. Almost every plaque connects its subject to physics not surprising, because Jack Allen was the motivating force behind the sign committee. I very much hope that the town sees fit to so Allan Griffin honour Allen himself.
Allan Griffin is in the Department of Physics, University of Toronto, Toronto, Ontario, M5S 1A7, Canada. e-mail: griffin@physics.utoronto.ca

138

NATURE | VOL 411 | 24 MAY 2001 | www.nature.com

UNIV. OF ST ANDREWS

Você também pode gostar