0% found this document useful (0 votes)
3 views

Chapter 2

Chapter 2 provides an overview of the experimental system used to study atomic gases, focusing on neutral alkali atoms and their properties. It discusses the significance of achieving low temperatures and densities for observing quantum phenomena like Bose-Einstein condensation, as well as the role of magnetic and optical trapping techniques. The chapter also covers the atomic properties, including the distinction between bosons and fermions, the Zeeman effect, and the concept of polarizability in the context of trapping and imaging atoms.

Uploaded by

Ramesh Mani
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
3 views

Chapter 2

Chapter 2 provides an overview of the experimental system used to study atomic gases, focusing on neutral alkali atoms and their properties. It discusses the significance of achieving low temperatures and densities for observing quantum phenomena like Bose-Einstein condensation, as well as the role of magnetic and optical trapping techniques. The chapter also covers the atomic properties, including the distinction between bosons and fermions, the Zeeman effect, and the concept of polarizability in the context of trapping and imaging atoms.

Uploaded by

Ramesh Mani
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 14

Chapter 2

The experimental system

In this first lecture, we are going to try and give a lightning introduction to the most important
aspects of the atomic gases that are actually realized in experiments. The goal is to provide
justifications for the kinds of models that we will be considering in the following lectures, and
to gain some feel for when experimental reality is likely to intrude!
We will be concerned with the properties of gases of neutral alkali atoms. The number
of atoms in typical experiments range from 104 to 107 : often N → ∞ will be a good enough
approximation. The atoms are confined in a trapping potential of magnetic or optical origin,
with peak densities at the centre of the trap ranging from 1013 cm−3 to 1015 cm−3 .
As we just discussed in the previous chapter, the observation of quantum phenomena
like Bose-Einstein condensation requires a phase-space density of order one, or nλ3dB ∼ 1.
The above densities then correspond to temperatures

~2 n2/3
T ∼ ∼ 100nK − few µK
mk

At these temperatures the atoms move at speeds of ∼ kT /m ∼ 1 cm s−1 , which should be


p

compared with around 500 m s−1 for molecules in this room, and ∼ 106 m s−1 for electrons
in a metal at zero temperature. Achieving the regime nλdB ∼ 1, through sufficient cooling is
the principle experimental advance that gave birth to this new field of physics.
It should be noted that such low densities of atoms1 are in fact a necessity. We are
dealing with systems whose equilibrium state is a solid (that is, a lump of Sodium, Rubidium,
etc.). The first stage in the formation of a solid would be the combination of pairs of atoms
into diatomic molecules, but this process is hardly possible without the involvement of a third
atom to carry away the excess energy. The rate per atom of such three-body processes is
10−29 − 10−30 cm6 s−1 , leading to a lifetime of several seconds to several minutes2 . These
relatively long timescales suggest that working with equilibrium concepts may be a useful
first approximation.
1
Compare 1019 cm−3 for the number density of air molecules at ground level, and ∼ 1022 cm−3 for atomic
densities in liquids and solids.
2
This three-body rate is reduced by the appearance of Bose-Einstein condensation, further enhancing the
sample lifetime.

3
2.1 Atomic properties
2.1.1 Boson or Fermion?
Since the alkali elements have odd atomic number Z, we readily see that alkali atoms with
odd mass number are bosons, and those with even mass number are fermions.
Alkali atoms have a single valence electron in an nS state, so have electronic spin
J = S = 1/2. Thus bosonic and fermionic alkalis have half integer and integer nuclear spin
respectively. We list the following experimental ‘star players’

Bosons Nuclear spin, I


87 Rb 3/2
23 Na 3/2
7 Li 3/2
Fermions
6 Li 1
40 K 4

2.1.2 The Zeeman effect


The Zeeman effect plays a crucial role in the trapping of atoms. Assuming that we deal with
atoms in a state of zero orbital angular momentum, the effect of magnetic field is described
by the Hamiltonian
HZ = AI · J + gµB B · J, (2.1)
where I and J are the nuclear and electronic angular momenta respectively, and the first term
originates from the hyperfine interaction. We can ignore the Zeeman effect of the nuclear
spin, as the nuclear magneton is approximately me /mp ∼ 1/2000 of the Bohr magneton.
We will consider this case I = 3/2, exemplified by 87 Rb, 23 Na, and 7 Li. Solving the
Hamiltonian (2.1) gives energy levels labeled by the conserved quantum numbers F
(F = I + J) and mF , the component parallel to the field. In the future, when we speak of
an atomic ”species”, we will normally mean one of these hyperfine-Zeeman states. The
zero-field splitting between the F = 2 and 1 states is 2A, and we can define a crossover
scale Bbf ≡ |A|/µB beyond which the complexities of the hyperfine coupling become
unimportant. The full field dependence of the energy levels is3

mF F E(B)
2 2 A (1 + B/Bhf )
1 2, 1 ±A[1 + B/Bhf + (B/Bhf )2 ]1/2
0 2, 1 ±A[1 + (B/Bhf )2 ]1/2
−1 2, 1 ±A[1 − B/Bhf + (B/Bhf )2 ]1/2
−2 2 A (1 − B/Bhf )
These are plotted in Fig. 2.1. Many experiments are done in the regime B  Bhf , so a
linear expansion of the above energies suffices.
 
1
E(B) = ± A + µB mF B , (2.2)
2
3
We shunt all the energy levels up by A/4 for convenience

4
Figure 2.1: Magnetic field dependence of atomic states of an atom with J = 1/2, I = 3/2.

with plus (minus) for the upper (lower) multiplet. Since magnetic traps have a local minimum
in the field, it is the ”low field seekers” with positive gradient that can be trapped. In the
present case these are F = 2, mF = 2, 1, 0, and F = 1, mF = −1.
We will see that in general collisions between atoms can convert low field seekers to high
field seekers that are then lost from the trap. Two states are however of special experimental
importance in being immune to this process. They are the doubly polarized state with F =
I + 1/2, mF = F and the maximally stretched state with F = I − 1/2, mF = −F .

2.1.3 Excited states and polarizabilty


A second approach to trapping atoms is to use the potential that they feel in the presence of
the electric fields created by a laser. The field polarizes the atoms, giving them an electric
dipole moment that in turn interacts with the field.
Let us start by considering the effect of a static electric field. Second-order perturbation
theory gives us an expression for the polarizability, defined through the quadratic energy shift
in the presence of an electric field ∆E = −αE 2 /2
X |hn|d · ε̂|0i|2 X
α=2 , d ≡ −e rj . (2.3)
n
En − E0
j

We leave in the unit direction vector of the electric field ε̂ in order to avoid the tensorial
structure of α. It is convenient to write this result in terms of the dimensionless oscillator
strengths
2me (En − E0 )
fn0 = |hn|d · ε̂|0i|2 .
e2 ~2
These satisfy the f-sum rule (or Thomas-Reiche-Kuhn sum rule)
X
fn0 = Z (2.4)
n

5
H, di , di (no sum).
  
Problem 1 Prove this by considering

We have
e2 X fn0
α= 2 , (2.5)
me n ωn0
with ωn0 = (En − E0 ) /~.
The response to external fields is mostly determined by the ”fundamental” transition of
the valence electron nS → nP (a doublet due to spin-orbit coupling). The wavelength of this
transition is in the range 500 − 700 nm. We can use these facts to estimate the polarizability.
First we assume that the valence electron states are well-approximated by those of a single
electron moving in a coulomb potential due to the nucleus and core electrons. Thus their
oscillator strengths satisfy Eq. (2.4) with Z = 1. Next we neglect all but the nS → nP
transition, giving
1
α∼ .
(∆E)2
The result should be understood in atomic units with α measured in units of a30 , and energies
in e2 /a0 ∼ 27.2 eV.
The case of oscillating fields can be treated by considering the ground state to ground
state amplitude in second order perturbation theory in the field E(t) = Eω e−iωt + E−ω eiωt , with
E−ω = Eω∗ 4
Z t
1
h0|0it = 1− 2 dt1 dt2 h0|T d(t1 ) · Eω d(t2 ) · Eω∗ |0ie−iω(t1 −t2 ) + {ω → −ω}
2~ 0
 i
i X 1 i h
= 1+ 2 t− e−i(ωn0 +ω)t − 1 |h0|d · Eω |ni|2
~ n ωn0 + ω ωn0 + ω
+{ω → −ω}. (2.6)

After time averaging the imaginary linear in t term can be thought of as a shift in the energy
of the ground state, leading to a phase factor e−i∆Et/~ ∼ 1 − i∆Et/~, equal to

1 X |h0|d · Eω |ni|2
∆E = − + {ω → −ω}.
~ n ωn0 + ω

This gives us the (real part of the) dynamical polarizability through ∆E = −α0 (ω)hE 2 (t)i/2
X 2 (En − E0 ) |hn|d · ε̂|0i|2
α0 (ω) =
n (En − E0 )2 − (~ω)2
e2X fn0
= 2 , (2.7)
me n ωn0 − ω 2

which generalizes the static results Eq. (2.3) and Eq. (2.5). Note that in contrast to the
static case, where an attractive force is always felt towards regions of high field intensity,
4
Field-theoretic types may prefer to think of the self-energy of the atom at second order in the ”dressing”
electric field

6
the dynamical polarizability can be of either sign. In particular, when the polarizability is
dominated by a single transition we have
|hn|d · ε̂|0i|2
α0 (ω) ∼ , (2.8)
~ (ωn0 − ω)
and the sign from positive to negative as we go from ω < ωn0 (red detuning) to ω > ωn0 (blue
detuning).
Finally, inclusion of a finite excited state lifetime Γe tends to broaden this behaviour. The
effect can be included by putting an imaginary part in the denominator of Eq. (2.8) to obtain
the complex polarizability
|hn|d · ε̂|0i|2
α(ω) ∼ . (2.9)
~ (ωn0 − ω − iΓe )
In particular the imaginary part of this expression can be thought of as giving the rate of
transitions out of the ground state (an ‘imaginary part to the energy’ Im ∆E = −α00 (ω)hE(t)2 i)
2 1
Γg = − Im ∆E = α00 (ω)hE(t)2 i (2.10)
~ ~
Excited state lifetimes are of order 10 ns.

Problem 2 Consider a harmonically confined electron. The equation of motion for the dipole
moment d = −er is
e2
d̈ + ω02 d = E
me
Show that the expression for the classical polarizability is
e2 1
α(ω) = 2
me ω0 − ω 2
This correpsonds to Eq. (2.5) with one oscillator strength equal to one at ω0 . Verify that this
is case in a quantum mechanical calculation.

2.2 Trapping and imaging


In the previous section we have discussed the two important features of the alkali elements
that make them such a versatile experimental system. The spin of the valence electron
allows magnetic trapping, while the wavelength of the nS → nP transition means that lasers
can be used for trapping and cooling. Cooling is discussed in some detail in Ref. [1]. Here
we discuss briefly trapping and imaging, which are probably more important for the theorist
seeking to understand experimental papers.

2.2.1 Magnetic traps


In free space B cannot attain a maximum, so magnetic traps work by creating a field mini-
mum in which the low field seeking states are confined. Most produce an axially symmetric
magnetic field of the form
1 1
|B(r, z, φ)| = B0 + αr2 + βz 2 . (2.11)
2 2

7
Provided that the atoms move slowly, we can make an adiabatic approximation and assume
that the atoms remain in the instantaneous hyperfine-Zeeman state appropriate to their cur-
rent position, even though the direction of B(r) may change. This is fine as long as |B(r)|
does not become too small, which in practice means various strategies are required to plug
such ‘holes’ where unwanted transitions between species can occur. Details of the various
types of magnetic traps can be found in Ref. [1].
The relationship between (2.11) and the potential experienced by the atom is then very
simple in the linear regime described by Eq. (2.2). There is one subtlety that makes it-
self known only occasionally. The effective quantum Hamiltonian of an atom of a particular
species in the adiabatic approximation does not simply involve a conservative potential due
to the magnetic field, but in general includes a gauge potential whose origin is the Berry
phase accumulated by a varying direction of B(r). This potential is expressed in terms of
the instantaneous hyperfine-Zeeman state |α(r)i as

Aα (r) = ihα(r)|∇|α(r)i (2.12)

Problem 3 The magnetic field of a quadrupole trap is

B = Bz ẑ + B 0 (xx̂ − yŷ)

If Bz  B 0 , we can think of the hyperfine states as being eigenstates of mF , the component


in the z-direction. Inversion of Bz would carry atoms adiabatically from mF to −mF , as the
direction of B rotates through π. The rotation axis, however, depends on where they are
in the trap, being about an axis parallel to yx̂ + xŷ. Show that the effect of this rotation
is to multiply the wavefunction by the position-dependent phase factor e−2imF φ , where φ
is the azimuthal angle. This technique has been used to create vortices in Bose-Einstein
condensates [2].

2.2.2 Optical traps


Optical traps work by focusing a laser to create a field maximum. If the laser is red-detuned
the result is a potential minimum for the atoms, as discussed in Section 2.1.3. This type of
trap is useful when we are interested in interactions that depend on species or Feshbach
resonances, which are tuned by magnetic field. In such circumstances we don’t want the
Zeeman energy confusing things.
In this context, it is useful to consider if the optical potential really is independent of
species. Any species-dependence requires spin-orbit coupling to be taken into account in
the excited state. In the nP state spin-orbit causing a splitting into nP1/2 and nP3/2 cor-
responding to total electron angular momentum of 1/2 and 3/2 respectively. Since these
states contain Kramers doublets due to time-reversal symmetry, the dipole matrix elements
between these states and the ground state are independent of the initial spin state of the
electron for linearly polarized light. Thus any effect will arise from the hyperfine splitting,
which alters the detuning from a particular transition. If the detuning is much larger than
this splitting, as it usually is, the effect is negligible. For circularly polarized light the matrix
elements can differ, and the greatest effect is achieved by tuning between nP1/2 and nP3/2 .

8
Optical traps are usually made using far-detuned light with Γe /(ωn0 − ω) ∼ 10−7 − 10−6 .
Although the trapping potential is inversely proportional to the detuning, the ground state
lifetime Eq. (2.10) goes like the inverse square. Note that the absorption of even one photon
would be a disaster for a sample, heating it far from degeneracy.

2.2.3 Imaging
As we saw, the scale of the dependence of optical properties on frequency is set by the
excited state lifetime Γ and is typically a few MHz. This is much less than the zero field
hyperfine splitting on the GHz scale. Thus distinguishing optically between different F values
presents no problem. For the sublevels within a multiplet, the dependence of transitions upon
polarization can be exploited to further enhance resolution. Usually it is possible to image
the individual species in a gas separately.
The images that one most frequently sees in experimental papers are simple absorption
images, in which light is absorbed by the gas, creating real transitions and heating the gas .
Such measurements are therefore of a one-shot character in that they destroy the sample.
The second kind of imaging is dispersive (phase-contrast) imaging that relies on diffrac-
tion. Many images may be taken with this technique without much heating. The creation of
such images can be regarded as instantaneous.

2.3 Interactions
For the most part, interactions between alkali atoms in the parameter ranges of experimental
interest are highly amenable to theoretical analysis. The effective range of the potential is
always small compared to other length scales, and normally we are also in the dilute gas
limit na3s  1 where as is the s-wave scattering length. These happy circumstances go some
way to explaining the popularity of these systems with theorists.

2.3.1 Effective interaction between like species


Working within the standard Born-Oppenheimer approximation we consider the interatomic
potential describing the interaction between two alkali atoms. In general this potential is
strongly dependent on the spin state of the valence electrons. The singlet state has a far
deeper minimum, of the order ∼ 5000K, than the triplet state. This is because the two
electrons in the singlet state can share an orbital and form a covalent bond. By contrast, the
minimum in the triplet potential results from an interplay between the −C6 /r6 van der Waals
attraction at large distances, and a hardcore repulsion at short distances, see Fig. 2.2.
Two atoms in the same hyperfine state clearly have a symmetric electron spin wavefunc-
tion, so the triplet potential is the appropriate one5 . The van der Waals potential defines a
1/4
length r0 ≡ 2mr C6 /~2 , where mr is the reduced mass. This scale, the typical extent
of the last bound state in the potential, is of order 50Å, much smaller than the de Broglie
wavelength. This allows us to ignore all but s-wave scattering.
5
In the next section we will consider interactions between different species, which in general will involve both
singlet and triplet potentials. As explained at the start of this chapter, however, the strongly bound molecular
states are slow to form so still have little effect, and singlet and triplet scattering lengths can be roughly equal.

9
Figure 2.2: Sketch of the interatomic potentials for two alkali atoms with valence electrons in
singlet and triplet states.

On general grounds, we expect all features in the scattering amplitude to be at the high
energy scale ∼ ~2 /2mr r02 . Then the low energy scattering of interest is described simply
by the s-wave scattering length, defined through the form of the wavefunction in the relative
displacement of two atoms
sin [k (r − as )]
ψ(r) = const. . (2.13)
r
The scattering length will prove to be a basic parameter of all theories of the alkali gases. For
all but the heaviest atoms theoretical calculation of scattering lengths is extremely difficult.
They can, however, be reliably measured using photoassociative spectroscopy, see Ref. [1].
It is of the same order as the scale r0 introduced above.
as normally enters into our theoretical considerations through the notion of the pseudopo-
tential. The idea is that provided that all other scales in the problem are much larger than as
– in particular this requires the smallness of the gas parameter na3s – the expectation value
of the interaction energy has the form6

1 4πas ~2 X Y
Z
hHint i = drk δ̃(rij )|Ψ({r})|2 . (2.14)
2 m
i6=j k

δ̃ (rij ) denotes a delta-function smeared on a scale much larger than as but much smaller
than all other scales. The integral in Eq. (2.14) is just the (spatially averaged) probability
density for two particles to approach each other, and the interaction energy is just propor-
tional to this quantity summed over all pairs. It is natural to expect that a pairwise form holds
in the dilute limit and the quantity summed over in then the energy of a pair in vacuo. To be
6
This is a slight abuse of notation: the origin of this effective potential is both kinetic and potential in general:
see below

10
absolutely clear about the origin of Eq. (2.14), I present an argument leading to it in more
detail7 .

• At low densities interaction energy (meaning the expectation value of the interaction
hamiltonian) should have a pairwise form.
• Wavefunctions we will consider correspond to energy per particle Etyp much less than
~2 /2ma2s . It’s reasonable that such wavefunctions satisfy the ‘boundary condition’

Ψ({r}) ∼ A(1 − as /rij ),


p
at ~/ 2mEtyp  rij  as .
• In the case of a single pair, such a wavefunction has energy
4πas ~2 2
|A| , (2.15)
m
which is found by solving the Schrodinger equation in a spherical box of size R  as ,
then finding the normalization constant for this wavefunction. (don’t forget the reduced
mass!). Eq. (2.15) is in fact the O(as /R3 ) part of the energy, with the leading term being
π 2 ~2 /4mR2 . Since it represents the leading as dependence, however, it is reasonable
to call this the ’interaction energy’. Note, however, that the energy can range from
being all kinetic for a hard sphere potential, to all potential for a very ‘soft’ potential. In
any case, it arises from a scale ∼ as and should be affected by long distance changes
in the wavefunction only through A.
• Finally, |A|2 corresponds to the value of the integral in Eq. (2.14).

Several other ways of writing the pseudopotential are in common usage. One is to dis-
pense with the slightly cumbersome form of Eq. (2.14) and write the interaction Hamiltonian
as a delta-function potential.
4πas ~2
U (r) = δ(r), (2.16)
m
or in second quantized notation
1 4πas ~2
Z
Hint = dr φ† (r)φ† (r)φ(r)φ(r). (2.17)
2 m
Obviously this requires careful handling in light of the above. Another way of approaching
these difficulties is to ask how we can define a δ-function potential U0 δ(r) in three dimen-
sions. The scattering amplitude in general satisfies the integral equation
dq U (k0 , q)F (k, q)
Z
0 0 1
F (k, k ) = −U (k, k ) − (2.18)
2 (2π)3 (q) − (k) − i0
where (k) = k 2 /2m. The δ-function potential can then be taken to be the limit of the (non-
translationally invariant!) separable potential

U (k, k0 ) = U0 g(k)g(k0 )
7
Alternative versions may be found in Ref. [3] and Ref. [4]. The interesting feature of Leggett’s argument is
that it seems to apply even when na3s  1 as long as nr03  1, e.g. near a Feshbach resonance.

11
where g(k) is equal to unity at small k, but falls to zero at some cut-off scale. Then the integral
equation is solved trivially for the low energy scattering amplitude F (k, k0 ) → −4π~2 as /m
Z
m 1 dq g(k)
= − .
4πas ~ 2 U0 (2π)3 2k
which is compatible with Eq. (2.16), apart from the (divergent) second term. This is just the
momentum space version of our real space difficulties.
Finally, Eq. (2.16) is sometimes written

4πas ~2
U (r) = δ(r)∂r r.
m
This serves to remove the 1/r piece from the boundary condition Ψ({r}) ∼ A(1 − as /rij )
when the expectation value is taken.
All of these complexities aside, by far the most important thing we will do with the pseu-
dopotential is use it to find the interaction energy for trial wavefunctions that have no addi-
tional correlations built in between particles (the ‘Gross-Pitaevskii’ approximation). Then the
summand in Eq. (2.14) is n/N , and we find the energy per particle

2πnas ~2
E/N = .
m
The remarkable thing is that, although we used the diluteness condition to derive it, the
pseudopotential can be used to compute the next order in the ground state energy of a
1/2
system of bosons as an expansion in na3s

2πnas ~2  1/2 
E/N = 1 + α na3s + ... , (2.19)
m

where α = 128/15 π.

2.3.2 Interaction between species


The natural generalization of the pseudopotential Eq. (2.17) to several species is

1 X 4πaαβγδ ~2
Z
Hint = dr φ†α (r)φ†β (r)φγ (r)φδ (r). (2.20)
2 m
αβγδ

Bose (Fermi) statistics allows us to take aαβγδ = (−)aβαγδ = (−)aαβδγ 8 . Where B  Bhf and
the hyperfine splitting is large enough to rule out scattering to other values of F , rotational
invariance allow us to write, for the states within a given hyperfine multiplet
1
aαβγδ = [a1 δαγ δβδ + a2 Fαγ · Fβδ ± {α ↔ β}] ,
2
where F is the total spin operator within the multiplet. Even when scattering between mul-
tiplets is possible, the total angular momentum and projection mF = mF 1 + mF 2 of the two
particles is conserved.
8
The definition of these quantities is again from the scattering state, where the incoming and outgoing waves
are now in the internal states γ1 δ2 ± δ1 γ2 and α1 β2 ± β1 α2 respectively

12
Figure 2.3: A Feshbach resonance is caused by hybridization of a closed channel bound
state with the open channel.

We are now in a position to explain the importance of the low-field seeking doubly polar-
ized state with F = I + 1/2, mF = F and the maximally stretched state with F = I − 1/2,
mF = −F , introduced in Section 2.1.2. Atoms in the doubly polarized state can only scat-
ter into the same state, as no other states have larger mF . Two atoms in the maximally
stretched state could scatter so that one ends up with mF = −F − 1, but these states lie in
the F = I + 1/2 multiplet. Since we are concerned with positive splittings A on the scale
100 mK - 1 K, collisions might not be expected to depopulate the maximally stretched state.
Transitions to other states within the F = I − 1/2 multiplet do however occur due to the
magnetic dipole interactions, see Section 2.3.4, but these are typically much less frequent.

2.3.3 The Feshbach resonance


If the centre-of-mass energy of two scattering atoms is close to the energy of a bound state,
the scattering amplitude can be strongly modified. This phenomenon is called a Feshbach
resonance. In recent years its exploitation by experimentalists has made the strength of
the interaction between atoms a continuously tunable experimental parameter, something
unthinkable in conventional condensed matter systems.
The idea is illustrated in Fig. 2.3. The curves that we drew in Fig. 2.2 are really just
representative of the many that we could draw for the interatomic potentials corresponding
to different hyperfine states of the two atoms. As we just explained, the relatively large
hyperfine splitting makes it impossible for either of the atoms to scatter at low energy into
a higher multiplet (the ‘closed channel’ of the diagram). Their wavefunctions will, however,
in general hybridize with bound or nearly bound states in these closed channels due the
presence of exchange interactions.
The simplest model for this kind of scattering is the two-channel model, that accounts for

13
one nearby bound state in one closed channel
X  q  g X
bq a†1,q+p a†−1,−p + h.c
X
H= p a†s,p as,p + + ε0 b†q bq + √ (2.21)
p,s q
2 V p,q

Here bq annhilates a molecule (bound state of two atoms in the closed channel), and as,p
annihilates an atom in the open channel. The energy of the closed channel bound state is
ε0 . We have introduced the two species s = ±1 so that we can discuss either bosons or
fermions: the atoms are distinguishable in either case 9 .

Problem 4 Find the scattering amplitude for two atoms in the model Eq. (2.21).
Solution For two atoms the wavefunction in the centre-of-mass frame has the form
" #
† † †
X
|ψi = βb0 + αp a1,p a−1,−p |0i.
p

Substituting into Eq. (2.21) gives the equations


g
2p αp + √ β = Eαp
V
g X
ε0 β + √ αp = Eβ.
V p

Eliminating β
g2 X
2p αp + αp0 = Eαp . (2.22)
V (E − ε0 ) 0
p

We look for a scattering state of the form

4π~2 f (E)
αp (E) = δp,p0 + ,
m 2p − E − i0

where p0 is the wavevector of the incoming wave with 2p0 = E. Substituting into Eq. (2.22)
gives
g2
 Z 
m dp f (E)
f (E) + + =0
(E − ε0 ) 4π~2 (2π)3 2p − E − i0
(c.f. Eq. (2.18)). In order to tame the singular behaviour of the integral, we need to shift the
detuning parameter ε0 by the infinite constant
Z
2 dp 1
ε0 → ε0 + g ,
(2π)3 2p
to give
g2
 Z  
m dp 1 1
f (E) + + f (E) − = 0.
(E − ε0 ) 4π~2 (2π)3 2p − E − i0 2p
9
The same model can be applied to bosons of one species – the result for the scattering amplitude below
should be doubled – while fermions of the same species have no s-wave scattering. This exemplifies the general
relation for the scattering amplitude of identical bosons or fermions f (θ) ± f (π − θ) in the s-wave case, when
f (θ) = const

14
Taking real and imaginary parts now yields
" √ #
g2 m m3/2 E
Re f (E) + − Im f (E) = 0
(E − ε0 ) 4π~2 4π

g2 m3/2 E
Im f (E) + Re f (E) = 0.
(E − ε0 ) 4π

The final result for the scattering amplitude is


~γ 1
f (E) = − √ √ (2.23)
m E − ε0 + iγ E

with γ = g 2 m3/2 /4π.

The pole in f (E) is lies at real energies for ε0 < γ 2 /4, passing through zero when ε0 = 0.
When the pole lies at negative values of energy, its position corresponds to the bound state
energy (modified by coupling). When the pole is no longer at negative energy we refer to a
virtual state10 . The scattering length f (0) = −a is
~γ 1
a = −√ ,
m ε0

and displays the divergence characteristic of a Feshbach resonance as we pass from positive
to negative detuning, signaling the occurrence of a bound state11 . In a sense the model we
introduced can now be discarded. The form of the scattering amplitude Eq. (2.23) is in fact
the most general one allowed at low energies [5]: the model was just a convenient physical
realization where this two-parameter asymptotic description turns out to be exact. The γ
parameter characterizes the width of the resonance. If we were only interested in energies
very low compared to γ 2 , a single parameter description in terms of the scattering length
only would suffice.
Feshbach resonances have been found in a variety of alkali atoms. Those in the fermions
6 Li and 40 K have been exploited to great effect recently to probe the BCS-BEC crossover that

we will discuss later. The consensus view is that these resonances are broad in the above
sense, so that a single parameter description is possible. Tuning through the resonance
is achieved by varying an applied field, as the atom and molecule generally have different
magnetic moments.
The divergence of the gas parameter na3s implied by the approach to a Feshbach res-
onance suggests that sample lifetime will be dramatically reduced, as three-body collisions
leading to the formation of diatomic molecules become more frequent. One should bear in
mind, however, that such processes are a function of statistics. In a fermionic system a Fes-
hbach resonance for scattering between two species can occur in the s-wave channel, but
of any three particles scattering in this way, two will be of the same species. The formation
of a molecule of size r0 is then suppressed by some power of r0 q, for q a typical wavevector.
This power turns out to be about 3.33, so that even when as > 3000Å, the molecule lifetime
can be > 100 ms.
10
Although the pole is at real positive energies up to for 0 < ε0 < γ 2 /4, there is no singularity in the scattering
amplitude as the pole is not on the physical sheet, see Ref. [5].
11
A background scattering length is normally added to this to include the effect of non-resonant scattering.

15
2.3.4 Dipolar interactions
Finally, there is a dipole-dipole interaction between the valence electron spins of two atoms

µ0 (2µB )2
Umd = [S1 · S2 − 3 (S1 · r̂) (S2 · r̂)] . (2.24)
4πr3
This interaction only conserves total angular momentum (orbital plus spin), so can lead to
a decay from the doubly polarized or maximally stretched states. It is possible to show,
however, that the rate for such processes is slow and in general does not limit the lifetime in
experiments on alkali atoms [1].
Since the creation of a Bose-Einstein condensate of Chromium atoms last year, the mag-
netic dipole interaction has returned to prominence. Chromium has a dipole moment of 6µB ,
six times larger than the alkalis, so the dipole interaction is 36 times stronger. Recent theo-
retical work has focussed on understanding the consequent properties of the condensate.

16

You might also like