Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Report of the internship at IISc

Vivishek Sudhir

18th June – 20th July, 2010


Résumé

Name: Vivishek Sudhir

Date of Birth: 28th June, 1988

Academic Qualification: B. Tech. (Electrical & Electronics)

Current Affiliation: Graduate Student


Center for Quantum Control
Department of Physics
Imperial College, London

Permanent Address: ‘Krishnangie’, Puthur Road


Koppam Junction
Palakkad 678001, Kerala

E-mail: vivishek.sudhir@gmail.com

Division: Atomic & Optical Physics Lab


Department of Physics
Indian Institute of Science, Bangalore

Supervisor: Prof. Vasant Natarjan

1
Acknowledgement

At the outset, I would like to thank Prof. Vasant Natarajan for giving me this opportunity to work
in his group. Without his welcoming attitude, I would not have learned so much nor would I have
experienced such a deep passion for experimental physics. I would also like to identify the strong
support that he has extended to me and for patiently listening to all my chatter.

Without the friendly support of the students in the lab, I would not have got to speed with much of
the stringent demands of experimentation. In particular, I would like to identify the helping hands of
Kanhaiya, Ketan, Alok and Ranjitha.

Last, but never the least, I would like to thank Raghuveer for all the help he has done, from procuring
the copper plates and screws to organizing the wonderful trip to Hoggenekal.

Thank you all!

Vivishek Sudhir

2
Contents

Resume 1

Acknowledgment 2

1 Zeeman tuned Doppler cooling 4


1.1 Theory of Doppler cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.1 Zeeman tuning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Design and test of Zeeman slower . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Axial magnetic field and current coils . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Adaptive Zeeman tuning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 Study of Hyperfine structure 12


2.1 Theory of hyperfine interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.1 First and second order hyperfine corrections . . . . . . . . . . . . . . . . . . . . 13

Bibliography 15

3
Chapter 1

Zeeman tuned Doppler cooling

Cooling of few-body atomic/molecular systems has become an essential tool in modern atomic physics,
particularly in the field of condensate physics. In this context, “cooling” refers to the attainment of a
few-body ensemble which is as close to its theoretical ground state as possible. This means that the
ensemble necessarily has its thermal rotational and vibrational modes frozen. Thus, the individual
entities of the ensemble are slowed down to a velocity as close to rest as possible (remember that the
average velocity and temperature of an ensemble are related, hence the usage of the term “cooling”).
For systems which possess an internal structure constituted by bound states, the preferred method
of slowing is by photon recoil. In this method, a laser beam of suitable energy is focussed against the
oncoming matter beam; the bound particles are excited by photon absorption and in the subsequent
spontaneous decay of the excited state, a photon is emitted, which provides a recoil momentum in
the direction opposite to the beam velocity, thus slowing it down.
One of the most efficient and advantageous cooling method is the so called Doppler cooling tech-
nique. Using this method, one can achieve temperatures as low as a few hundred micro Kelvins; it
is theoretically impossible to cool below this Doppler limit. Other methods like Sisyphus cooling or
evaporative cooling are used on a Doppler cooled system to take it to lower temperatures.

1.1 Theory of Doppler cooling


In most experiments in atomic physics, two distinct levels of an atom are chosen to be resonant with
a laser beam shooting photon in the direction opposing the direction of motion of the atomic beam.
Then the photon recoil force slows the atoms as they move in the optical “drag” of the radiation.
To understand Doppler cooling, it is necessary to first understand how a two-level atom interacts
with the radiation field. Let ρ be the density matrix of the two-level atom, i.e.,
  ∗
c c c c∗
 
ρee ρeg
ρ= = e e∗ e g∗ ,
ρge ρgg cg ce cg cg

so that we have the two implicit relations, ρee + ρgg = 1 and ρ∗eg = ρge ; thus, the density matrix
has only three independent real parameters. In the interaction picture, the Hamiltonian for an atom
interacting with a classical radiation field is 1 ,

H = −eE · r, (1.1)

where, E = E0 êeikz−ωf t is the electric field.


When an atom and a radiation field interact with each other, there is a possibility that the excited
atomic state may spontaneously decay or that the ground state may spontaneously get excited; these
1
as pointed out by C. Cohen-Tannoudji etal. [1], there are some serious covariance issues in the form of the interaction
Hamiltonian (1.1) when the radiation field is also quantized, but for the present semi-classical treatment, this is well
justified [2]

4
effects are included in the dynamical equations by explicitly adding a term Γρ into the Liouville
equation for the density matrix, which then reads,
dρ 1
= [H, ρ] + Γρ.
dt i~
Written out in full, the Liouville equation takes the form,

iΩ∗ /2
    
ρgg 0 γ −iΩ/2 ρgg
d  ρ 0 −γ iΩ/2 −iΩ ∗ /2
 ee  =    ρee  ,
    
∗ ∗ (1.2)
dt ρ̄ge
   −iΩ /2 iΩ /2 − (iδ + γ/2) 0   ρ̄ge 
ρ̄eg iΩ/2 −iΩ/2 0 − (−iδ + γ/2) ρ̄eg

where ρ̄ge = ρge e−iδt , Ω = − eE~ 0 he|r|gi is the Rabi frequency, δ = ωf − ωa is the detuning and γ is the
spontaneous transition rate. In arriving at the above form of the dynamical equation, we have made
use of the dipole and rotating wave approximations [3].
Introducing new real variables, u = ρee , v = <e(ρge ) and w = =m(ρge ), the above equations can
be recast into the form,
      
u −γ 0 Ω u 0
d   
v = 0 −γ/2 −δ   v  +  0  . (1.3)
dt
w −Ω δ −γ/2 w Ω/2

These are the optical Bloch equations.


In steady state, the upper level population ρee is given by,
Ω2 /4
ρee = ,
δ 2 + Ω2 /2 + γ 2 /4
so that the scattering force is given by,
Ω2 /4
F = ~kγ , (1.4)
δ 2 + Ω2 /2 + γ 2 /4
which we have obtained by multiplying the photon momentum by the upper level decay rate and
the upper level population. It is easily noted that this photon recoil force has its maximum value of
~kγ/2 when the detuning δ = 0 i.e., for maximum recoil force, the atom and light should be in prefect
resonance.
But now, since the atom is moving through the light beam, the frequency of the laser as seen by the
atom will be Doppler shifted by an amount −kV , where V (t) is the velocity of the atom. So, if, initially
the atom and laser are in resonance, as it slows down by photon recoil, it moves below resonance and
the subsequent number of photons it absorbs will fall down gradually and further cooling becomes
impossible. In an older technique, initially, the laser is tuned slightly below the atomic transition so
that as the atom starts to experience the Doppler effect, the laser comes in resonance and cooling can
be effectively affected.

1.1.1 Zeeman tuning


To circumvent the problem of the laser falling out of resonance due to Doppler shifting, one uses
the technique of Zeeman tuning. Zeeman tuning is a far more effective method of maintaining zero
detuning than, say red shifting the laser initially, which still only gives a brief window of perfect
resonance.
The idea here is to apply a magnetic field on the atom in such a fashion that the variation of the
detuning due to Doppler shift is exactly balanced by the Zeeman shift of the atomic levels induced by
this magnetic field. The objective is to establish a magnetic field such that the Zeeman shift of the
atomic level nullifies the Doppler shift, thus maintaining perfect resonance throughout the stopping
distance.

5
If L is the desired stopping distance, then the velocity profile of the atom across this distance is
given by,
 z 1/2
V (z) = V0 1 − ,
L
where we have assumed constant deceleration. For the Zeeman effect to exactly counter the Doppler
shift, the condition is,
µB B(z)
ωa + = ωf + kV (z).
~
The above two equations give the required magnetic field profile viz.,
 z 1/2
B(z) = B0 + Bi 1 − , (1.5)
L
where µB B0 = −~δ and µB Bi = ~kV0 .

1.2 Design and test of Zeeman slower


The particular atom that we aim to cool here is Ytterbium, a lanthanide with atomic number 70.
The ultimate aim of trapping Yb is to perform sensitive studies on its electron dipole moments and
for that, one requires an isotope whose nucleus is a fermion. Thus, we choose the isotope with mass
number 171, which has a nuclear spin of 21 . The ground state electronic configuration of this isotope
is [Xe]4f14 6s0 . Some of the excited energy states useful for cooling and trapping is shown in Fig. 1.1.

Figure 1.1: Low-lying electronic energy levels of Yb171


70

The transition 1 P1 → 1 S0 is the one used for Zeeman tuned Doppler cooling. However, this transition
is not radiatively closed, so that the population in 1 S0 decays to the 3 P0 , 3 P1 and 3 P2 states. The 3 P1
population decays back to the 1 S0 state via the emission of a 556nm photon, closing the decay cycle.
But the states 3 P0 and 3 P2 are meta-stable, so that they do not decay back and so the population
excited into them are lost forever in the context of laser cooling. Thus, the relative stability of these
states limit the time available for laser cooling, so that there is a maximum distance within which the
atom has to be cooled or it cannot be cooled any further.
The fact that the thermal Yb source gives off Yb atoms at a most-probable velocity of 350m/s
and that we require to slow them down to about 20m/s most-probable and the fact that our Zeeman
slower had to be within 40cm length, fixed the required magnetic field constants B0 and Bi .

1.2.1 Axial magnetic field and current coils


Now the required magnetic field profile on the axis of the Zeeman slower flight tube is given by (1.5),
which is a semi-parabola opening in the −z direction. The way to achieve this profile is by using a
large number of current carrying coils wound over the flight tube, and energized by passing current
over it. Consider such a coil confined in the xy−plane as shown in Fig. 1.2.

6
Figure 1.2: Current carrying loop in xy−plane [4]

Using standard spherical coordinates, the current density J has components only in the φ direction,
given by,
 δ (r0 − a)
Jφ = I sin θ0 δ cos θ0 ,
a
so that the vectorial current density is given by,

J = −Jφ sin φ0 i + Jφ cos φ0 j.

Then by the Biot-Savart law,


J(r0 ) 3 0
Z
µ0
A(r) = d r,
4π |r − r0 |
is the magnetic vector potential. For our problem, since there is an axial symmetry, only the Aφ com-
ponent survives, and is given by the integral (after the preliminary integration over delta functions),

cos φ0 dφ0
Z
µ0 Ia
Aφ (r, θ) = .
4π 0 (a2 + r2 − 2ar sin θ cos φ0 )1/2

This integral can be evaluated in closed form using elliptic functions, to give the result,

(2 − k 2 )K(k) − 2E(k)
 
µ0 4Ia
Aφ (r, θ) = √ , (1.6)
4π a2 + r2 + 2ar sin θ k2

where K(x) and E(x) are complete elliptic integrals of the first kind and the argument k is given by,

4ar sin θ
k2 = .
a2 + r2 + 2ar sin θ
In an actual experiment, the atomic beam travels, neglecting the beam width itself, through the
axis (θ → 0, r → z) and it is the axial magnetic field Bz (z) that needs to fit the profile (1.5).
Evaluating the axial component of the magnetic field using (1.6), we get,

µ0 Ia2
Bz (z) = . (1.7)
2 (z 2 + a2 )3/2

The hypothesis is that the profile (1.5) can be obtained by winding current coils carrying the same
current, but geometrically arranged in layers arranged one on top of the other, with varying number
of turns in each layer.

7
We assume that the coil arrangement consists of N layers with the nth layer consisting of a
maximum of Mn coil turns. We also assume that the conductor has a radius r and carries a uniform
steady current I throughout all the turns in all the layers. Then, the magnetic field produced by the
mth turn of coil in the nth layer is given by,
2
µ0 I n + 21
Bmn (z) = 2 i3/2 ,
2r h 2
n + 12 + z − m − 12
so that the magnetic field produced by the totality of all the coil turns in all the layers is given by,
N Mn 2
µ0 I X X n + 12
Bcoil (z) =  i3/2 . (1.8)
2r h
1 2
 1 2
n=1 m=1 n+ 2 + z−m− 2
Now the “error” in the axial field profile is given by the function,
e(z) = |B(z) − Bcoil (z)|2 ,
where B(z) is the theoretically required profile from (1.5). The problem of finding the required coil
configuration can now be formulated as the mathematical problem of minimizing e(z) with respect to
the variables {N, M1 , M2 ...MN }, which is well-posed and well known. It can be solved in a straight
forward manner by employing the MMSE (minimum mean square error) algorithm, which is well
studied in other contexts [5].
We designed the coil configuration in two separate segments, the reverse and the forward. This was
necessitated because the field strengths we required were on the higher side, upto a maximum of 300
Gauss. The reverse windings were made with a 0.9mm thick cooled copper wire and the current used
was 60A. The forward part was wound using a high-capacity cooled copper wire of 1.2mm diameter
and a current of 127A was passed through it. The exact coil configurations for both the windings are
given in Table 1.2.1.

Coil layer Reverse coil turns Forward coil turns


Layer 1 9 turns 4 turns
Layer 2 9 turns 4 turns
Layer 3 14 turns 4 turns
Layer 3 14 turns 4 turns
Layer 4 2 turns 4 turns
Layer 5 5 turns 4 turns
Layer 6 5 turns 4 turns
Layer 7 5 turns 6 turns
Layer 8 5 turns –

Table 1.1: Coil configuration of the forward and reverse segments of the Zeeman tuner

The Zeeman tuner was built acording to the above specifications and the axial magnetic field
was measured before the equipment was deployed on the actual experimental setup. Fig. 1.3 shows
the experimentally measured axial magnetic field strength and its near perfect concordance with the
theoretically desired profile.
The main laser system is the one used to effect cooling by coupling the 1 P1 → 1 S0 tansition. A
SpectraPhys MilleniaXs laser operating at 532nm is fed to a Coherent 899-21 Ti:Sapphire ring laser
to produce a continuously tunable output from 700 − 800nm. The Ti:Sa is frequency stabilized by
locking it to a Febry-Perot reference cavity at 797.6nm. This is then frequency doubled using a Laser
Analytic System - WaveTrain doubler. The final output is 398.8nm at just above 120mW.
The final RF magneto-optical trapping cavity consists of two pairs of coils in anti-Helmholtz
configuration, made of hollow copper tubes through which chilled water is circulated to keep them
cool. The complete arrangement, commonly called the Ioffe-Pritchard MOT is shown in Fig. 1.4.

8
Figure 1.3: Theoretically desired and experimentally obtained axial magnetic field strength inside
Zeeman tuner

Figure 1.4: Schematic of a Ioffe-Pritchard trap configuration

1.3 Adaptive Zeeman tuning


The analysis of §1.1 assumes that the deceleration of the atom via photon recoil is constant i.e., that
the recoil force is a constant over time. This assumption has its origins in the expression (1.4), which
was calculated assuming that the upper level population has reached steady state. It is of course true
that after a brief transient, the population reaches a steady state value; but it is a fact that during the
transient period, the laser is significantly off resonance since this is the time when the atomic cloud
has the highest velocity and hence experiences the maximum Doppler shift. Thus, during this period,
the atom does not really undergo much slowing and so the initial segment of the Zeeman slower is
simply a flight tube for all practical purposes.
The developments presented in this section, aims at rectifying this ineffectiveness and forms one of
the novel pieces of work done during the internship period. It is quite obvious that when the atom is
moving through a Doppler shifted laser field, the optical Bloch equations (1.3), are to be supplemented
with another equation for the linear velocity of the ensemble, a freedom of degree which is very much
coupled with the internal atomic degrees of freedom.
To obatin this final equation, one notes that the detuning δ is a time varying parameter, unlike

9
what it is made to look in (1.3); in particular, we have

dδ d
= (ωf − ωa − khvi)
dt dt
dhvi
= −k
dt
k ∂H
= h i,
m ∂z
where in going to the last line, we have used the Ehrenfest Theorem. Using the dipole and rotating
wave approximations, one can express the above equation after taking the necessary expectation values
as,
∂Ω∗
 
dδ ~k ∂Ω
= ρ̄ge + ρ̄eg .
dt m ∂z ∂z
Using the fact that the Rabi frequency Ω is real, the above equation is equally well an equation for
the linear velocity V (t) in terms of the atomic coherences. Thus, we arrive at the full coupled optical
Bloch equations,
      
u −γ 0 Ω 0 u 0
d v =  0
  −γ/2 kV − δ0 0 v +  0 ,
   
(1.9)
dt w
   −Ω δ0 − kV −γ/2 0   w   Ω/2
V 0 2s 0 0 V 0
~ ∂Ω
where s ≡ m ∂z . It is to be noted that the above equations form a set of nonlinear coupled ODE’s
and is thus not possible to find closed form analytical solutions for them; it becomes necessary to
solve them numerically.
Instead of doing that, a realtime simulator was programmed using Simulink to simulate the dy-
namics predicted by the coupled optical Bloch equations (1.9). The state diagram for the equations
in Simulink is shown in Fig. 1.5.

Figure 1.5: Simulink state diagram for the coupled optical Bloch equation simulator

10
The system was simulated for a realistic scenario of atoms starting out at 100m/s, but being
controlled by a cooling laser of very little power, namely of the order of a few pico-watts. In the
standard Zeeman tuned Doppler cooling situation of previous sections, the claculations predict that
the atoms would be cooled to the Doppler limit in about 20secs. But in this case, where even the
initial transient region is utilized for resonant cooling, the results of the simulation Fig. 1.6 show that
the atoms are cooled to the Doppler limit in about 4secs.

Figure 1.6: Results of a typical simulation of the coupled optical Bloch equations (1.9) in Simulink.
The top three plots show the variables u(t), v(t), w(t), while the bottom most plot shows V (t), the
atomic velocity.

Thus, it is seen that utilization of the initial transient dynamics can markedly improve the result
of laser cooling, especially for the case if light atoms [6]. The other use is for the development of a
realtime automatic Zeeman tuner for a Doppler cooling system. In this case, instead of creating a
space varying magnetic profile for Zeeman tuning, one can use realtime simulations of the coupled
optical Blocj equations (1.9), to arrive at the time dependent magnetic field required for perfect
resonant coupling of the laser and atom. Once this is known, it is easily possible to use a “gated”
current to develop the required field. This technique would eliminate the need for a cumbersome coil
system for the Zeeman tuner. These are developments for the future.

11
Chapter 2

Study of Hyperfine structure

In the typical study of the atomic spectra, the nucleus is considered a point entity, which does not
exert any other influence on the electron(s) other than the Coulombic central force; this state of
affairs is modeled, in the relativistic case, by a Hamiltonian which is the sum of the Dirac-Coulomb
Hamiltonian and the configuration interaction term,
!  
X X Ze2 X Ze2
HDC = cα · pi + βmc2 −  + , (2.1)
ri |ri − rj |
i i j

where, the α, β are anti-commuting Dirac matrices and the sums are over the electrons. In the
above expression, the last double sum represents the configuration interaction (CI) terms, while the
remaining terms represent the Dirac-Coulomb terms.
The above Hamiltonian has undergone several revisions [7], each bringing in a new term to account
for various fine structure effects, including the fine splitting, the Lamb shift and the Zeeman-Stark
shifts. All these effects have been due to the presence of electromagnetic fields that affect the electron
structure; with the former two originating from vacuum polarization while the later originating from
externally applied fields.
On the other hand, the hyperfine splitting is a fine structure effect caused due to the finite size of
the nucleus. In particular, it is caused by the overlap of the electronic wavefunctions with that of the
nucleus. As the spherically symmetric component of the electron-nucleus interaction is already taken
into account in (2.1), hyperfine splitting is mainly caused by direction dependent coupling between
the electrons and the nucleus.

2.1 Theory of hyperfine interaction


The fact that the nucleus has a finite size means that its charge distribution is may exert a direction
dependent force on the electronic cloud; of course, this is typified by the multipole moment expansion
of the nuclear electromagnetic potentials [8],
!
X 1 X
φ(r) = (−1)µ Ck,µ (r̂)Qk,−µ
rk+1 µ
k
r ! (2.2)
X 1 k+1 X µ
A(r) = −i (−1) Ck,µ (r̂)Mk,−µ ,
rk+1 k µ
k

where, Ck,µ and Ck,µ are the normalized scalar and vector spherical harmonic respectively; while,
Qk,−µ and Mk,−µ are the components of the nuclear electric and magnetic 2k –pole moment operators
respectively.
In this abstract model, we still approximate the nucleus by a collection of multipoles, rather than
elucidate its exact potential surfaces using a nuclear structure calculation.

12
The hyperfine interaction Hamiltonian is given by,
X
Hhf s (r) = e (α · A(ri ) − φ(ri )) , (2.3)
i

where the sum is over the electron coordinates. The above interaction Hamiltonian is the minimally
coupled Dirac Hamiltonian with the nuclear momentum set to zero, the Born-Oppenheimer approxi-
mation, which basically assumes that the nucleus is massive enough to be considered stationary.
Substituting the multipole expansions (2.2) in the hyperfine Hamiltonian (2.3), gives,
X
Hhf s = (−1)µ Tk,µ Tk,µ , (2.4)
k,µ

where the k th rank spherical tensor Tk,µ acts on the electronic states and Tk,µ acts on the nuclear
space. We make the identification Tk,µ ≡ Mk,µ for odd k and Tk,µ ≡ Qk,µ for even k. Explicitly, the
electronic tensors are given by, X
Tk,µ = tk,µ (ri ),
i

where,
1
(
− rk+1 Ck,µ (r̂) even k
tk,µ (r) = i
q
k+1
(2.5)
− rk+1 k α · Ck,µ (r̂) odd k

2.1.1 First and second order hyperfine corrections


In the presence of multipolar fields typical of the hyperfine interaction, the total electronic angular
momentum J is no longer a conserved quantity; it is the “total” angular momentum F = I + J that
is conserved. Here, I is the nuclear angular momentum. Thus, the proper basis set of states are the
ones formed by taking the tensor product of the electronic and nuclear states, viz.,
X F MF
|γIJF MF i = CJM J ;IMI
|γJMJ i ⊗ |IMI i, (2.6)
MJ ,MI

where the expansion coefficients are the usual Clebsch–Gordan coefficients and γ represents the set
of all the other electronic quantum numbers. For fixed values of I and J, we have the selection rule,
|J − I| ≤ F ≤ J + I.
Using perturbation theory, the first order corrections in the electronic energies due to the hyperfine
interaction can be shown to be given by the diagonal elements of Hhf s in the coupled basis, i.e.,
(1)
WF = hγIJF MF |Hhf s |γIJF MF i. (2.7)

Similarly, the second order correction is given by,

(2)
X |hγ 0 IJ 0 F MF |Hhf s |γIJF MF i|2
WF = , (2.8)
Eγ 0 J 0 − EγJ
γ0J 0

where the sum excludes the case γ 0 = γ, J 0 = J.


Before we actually set out to evaluate the above expression, it is useful to introduce a set of c-
number parameters that characterize the hyperfine interaction and which have a direct bearing with
experimental data. These hyperfine structure constants, are usually denoted by A, B, C, D... [9] and
are defined in terms of the diagonal products of the “stretched” matrix elements of the hyperfine
spherical tensors T and T . For any tensor operator Ok,µ , its “stretched” matrix element is defined as
the expectation value of its zeroth component in the stretched state |I, MI = Ii, i.e.,

hOk iI ≡ hI, MI = I|Ok,µ=0 |I, MI = Ii.

13
In particular, the magnetic dipole, electric quadrupole, magnetic octupole and electric hexadecapole
moments of the nucleus are defined in terms of these c-number stretched matrix elements via,

µ = hM1 iI
Q = 2hQ2 iI
Ω = −hM3 iI
Π = hQ4 iI .

Now, the hyperfine structure constants A, B, C, D are defined as [9],


1 µ
A= hT1 iI hT1 iJ = hT1 iJ
IJ IJ
B = 4 hT2 iI hT2 iJ = 2Q hT2 iJ
C = hT3 iI hT3 iJ = −Ω hT3 iJ
D = hT4 iI hT4 iJ = Π hT4 iJ .

Now, the first order correction (2.7) can be expressed in the form [10],
(1)
X
WF = Xk (I, J, F ) hTk iI hTk iJ , (2.9)
k

where the coefficients Xk are given by,


 
F J I
(−1)I+J+F
k I J
Xk (I, J, F ) =   .
J k J I k I
−J 0 J −I 0 I
   
j1 j2 j3 j1 j2 j3
In the above expression, stands for the Wigner 3–j symbol while stands
m1 m2 m3 j4 j5 j6
for the Wigner 6–j symbol. The second order correction (2.8), can also be explicitly expressed as [11],

(2)
X (−1)J−J 0 X  I J F

I J F

WF = hI|Tk |IihI|Tl |IihγJ|Tk |γ 0 J 0 ihγ 0 J 0 |Tl |γJi.
0 0
EγJ − Eγ 0 J 0 J0 I k J0 I l
γJ k,l
(2.10)
Before the expressions (2.9) (2.10) can be used in an actual calculation, it is useful to know
the explicit expressions for the electronic spherical tensor operator’s expectation over the relativistic
electronic Dirac wavefunctions. In the “low” energy regime (where QED effects are negligible) of
relativistic atomic theory, the typical electronic wavefunction is given by the bispinor,
 
1 iPnκ (r)Ωκm (r̂)
|nκmi = ,
r Qnκ (r)Ω−κm (r̂)

where P, Q are the radial parts of the wavefunctions obtained as eigenstates of the Hamiltonian (2.1)
and Ω are spherical spinors [8], which are eigenstates of the operator K = 1 − σ · L with eigenvalue
κ = (l − j)(2j + 1). Then we have the following important result [10],
R ∞ dr
−hκ0 |Ck |κi  0 rk+1
(
 R(Pn κ Pnκ + Qn κ Qnκ ) even k (electric)
0 0 0 0
0 0
hn κ |tk |nκi = 0 κ0 +κ ∞ dr
hκ |Ck |−κi k 0 rk+1 (Pn κ Qnκ + Qn κ Pnκ )
0 0 0 0 odd k (magnetic),

where the selection rules: |j − j 0 | ≤ k ≤ j + j 0 and l + l0 even, are to be observed.

14
Bibliography

[1] C. Cohen-Tannoudji, J. D. Roc, and G. Grynberg, Photons and Atoms (John-Wiley and Sons,
1997).

[2] C. Cohen-Tannoudji, in Fundamental Systems in Quantum Optics, edited by J. Dalibard, J. M.


Raimond, and J. Zinn-Justin, Ecole de Physique des Houches (Elsevier Science, 1990), vol. Session
LIII.

[3] H. J. Metcalf and P. van der Straaten, Laser cooling and trapping (Springer, 2002).

[4] J. D. Jackson, Classical Electrodynamics (John-Wiley and Sons, 2009), 3rd ed.

[5] S. Haykin, Adaptive Signal Processing (Tata Mc-Graw Hill, 2004).

[6] G. B. A. etal (ALPHA Collaboration), Phys. Rev. Lett. 105, 013003 (2010).

[7] B. Brandsen and C. Joachain, Physics of Atoms and Molecules (John-Wiley and Sons, 1990).

[8] W. R. Johnson, Lectures on Atomic Physics (Springer, 2007).

[9] W. Dankworth, J. Ferch, and H. Gebauer, Z. Phys. 267, 229 (1974).

[10] K. Beloy, A. Derevianko, and W. R. Johnson, Phys. Rev. A 77, 012512 (2008).

[11] K. Beloy and A. Derevianko, Phys. Rev. A 78, 032519 (2008).

15

You might also like